Suyama, Takashi L.: Electronic Theses and Dissertations UC San Diego
Suyama, Takashi L.: Electronic Theses and Dissertations UC San Diego
UC San Diego
Peer Reviewed
Title:
Organic synthesis as an effective approach to chemical, pharmaceutical, and biosynthetic
investigations of natural products
Author:
Suyama, Takashi L.
Acceptance Date:
01-01-2009
Series:
UC San Diego Electronic Theses and Dissertations
Degree:
Ph. D., UC San Diego
Permalink:
http://www.escholarship.org/uc/item/2pn8j589
Local Identifier:
b6296689
Abstract:
The partnership between natural products and synthetic organic chemistry date back to the
origin of organic chemistry itself. While natural products became a major driving force for the
development of novel organic reactions and synthetic strategies, organic synthesis has contributed
in many ways to the elucidation and confirmation of structure, pharmaceutical development,
and biosynthetic studies of natural products. Due to the recent advances in both of these two
disciplines, there are new opportunities and issues surrounding natural products that organic
synthesis can be applied to, and such studies comprise this dissertation. Chapter I introduces the
background information and rationale for the dissertation research, which is based on the history
of organic chemistry and newly emerging research topics in natural products chemistry. Chapter II
describes the isolation, characterization, and toxicological evaluation of polybrominated diphenyl
ethers from a mixed assemblage of a marine red alga and cyanobacteria. Chapter III describes
the isolation, characterization, and biological evaluation of a novel vinylchloride-containing fatty
acid, credneric acid, from a Lyngbya sp. Chapter IV describes the conception of a method for
determining the absolute stereochemistry of natural products lacking proper functional groups for
derivatization. This methodology was applied to a cyclopropane-containing fatty acid and resulted
in a conflict with newly published literature, which is discussed in detail. Chapter V describes
the development of an expedient and efficient total synthesis of a novel alkaloid, epiquinamide,
isolated from the skin of a rainforest frog, Epipedobates tricolor. This work contributed to clarifying
the identity of a potent and selective nicotinic receptor agonistic activity observed in the extract of
E. tricolor. Chapter VI describes the development of a total synthesis of a marine cyanobacterial
metabolite, somocystinamide A, which has been shown to be a potent inhibitor of endothelial
cell proliferation and angiogenesis. The synthesis was later modified for scalability to meet the
Copyright Information:
Doctor of Philosophy
in
Oceanography
by
Takashi L.Suyama
Committee in charge:
2009
Copyright
Chair
2009
iii
DEDICATION
Christ of Nazareth. For, “we are hard-pressed on every side, yet not crushed; we are
perplexed, but not in despair; persecuted, but not forsaken; struck down, but not
destroyed— always carrying about in the body the dying of the Lord Jesus, that the life of
Jesus also may be manifested in our body.... Therefore we do not lose heart.... For the
things which are seen are temporary, but the things which are not seen are eternal
(2Corinthians 4:8-10, 16, 18, NKJV; emphasis mine).” Indeed, it is because of Him that
even I have persevered through this doctoral program. The credit for all that follows is
iv
EPIGRAPH
My Hope is Built.
v
TABLE OF CONTENTS
Dedication................................................................................................... iv
Epigraph....................................................................................................... v
Table of Contents......................................................................................... vi
Acknowledgments....................................................................................... xx
Abstract........................................................................................................ xxvi
Chapter I: Introduction................................................................................ 1
vi
Chapter VIII: Conclusions........................................................................... 368
Appendix..................................................................................................... 375
vii
LIST OF ABBREVIATIONS
CD circular dichroism
CoA coenzyme A
EC effective concentration
EDC 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide
EtOH ethanol
DMAP 4-dimethylamino pyridine
EI electron impact
ESI electron spray ionization
FAB fast atom bombardment
FLIPR fluorescence laser plate reader
GCMS gas chromatography mass spectrometry
HMBC heteronuclear multiple bond correlation
HMG 3-hydroxy-3-methyl-glutaryl
HMQC heteronuclear multiple quantum coherence
HPLC high performance liquid chromatography
HR high resolution
HSQC heteronuclear single quantum coherence
INADEQUATE incredible natural abundance double quantum transfer
experiment
i-PrOH isopropanol
IR infrared spectroscopy
MeOH methanol
MS mass spectrometry
MsO methanesulfonate
NCS N-chlorosuccinimide
NMR nuclear magnetic resonance
viii
NOE nuclear Overhauser effect
OR optical rotation
rt room temperature
THF tetrahydrofuran
UV ultra violet
ix
LIST OF ILLUSTRATION AND FIGURES
Figure I.1: The originally assigned structure (13) of mururin C and its
revised structure (12).................................................................................................... 11
Figure I.2: Types of errors in natural product structure revisions................................. 30
Figure I.3: Structure determination methods used in 2005-2008 for the erroneous
cases.............................................................................................................................. 30
Figure I.4: Structure revision methods used in 2005-2008........................................... 30
Figure I.5: Discrimination of possible structures by HMBC........................................ 31
Figure I.6: HMBC correlations observed and recorded for the labdane diterpenoid... 31
Figure I.7: All new small chemical entities that were approved for clinical use,
01/1981-06/2006, by source......................................................................................... 33
Figure I.8: Drug Development Process......................................................................... 35
Figure II.1: OH-PBDE congeners isolated from the red alga/cyanobacteria assembly
(1 and 2) and an alternative regioisomer of 1 (3)......................................................... 63
Figure II.2: X-ray crystal structure of 1........................................................................ 68
Figure II.3: Ca2+ modulation assay in mouse neocortical neurons for compounds
1-3................................................................................................................................. 72
Figure II.4: Photographs of the red algal voucher sample............................................ 77
Figure II.5: Photograph of the red algal voucher sample............................................. 78
Figure II.6: Isolation scheme for 1 and 3...................................................................... 85
Figure II.7: 1H NMR Spectrum of 1 in CDCl3 ............................................................. 86
Figure II.8: 13C NMR Spectrum of 1 in CDCl3 ............................................................ 87
Figure II.9: 1H-1H COSY Spectrum of 1 in CDCl3 ...................................................... 88
Figure II.10: 1H-13C HSQC Spectrum of 1 in CDCl3 ................................................... 89
Figure II.11: 1H-13C HMBC Spectrum of 1 in CDCl3 .................................................. 90
Figure II.12: 1H NMR Spectrum of 1 in CDCl3 ........................................................... 91
Figure II.13: 13C NMR Spectrum of 3 in CDCl3 .......................................................... 92
Figure II.14: 1H-1H COSY Spectrum of 3 in CDCl3 .................................................... 93
Figure II.15: 1H-13C HSQC Spectrum of 3 in CDCl3 ................................................... 94
Figure II.16: 1H-13C HMBC Spectrum of 3 in CDCl3 .................................................. 95
Figure II.17: 1H NMR Spectrum of 2 in CDCl3 ........................................................... 96
x
Figure II.18: 1H-1H COSY Spectrum of 2 in CDCl3 .................................................... 97
Figure III.1: Jamaicamides A-C.................................................................................... 104
Figure III.2: Malyngamides A-X.................................................................................. 104
Figure III.3: Satellite Image of Lyngbya Collection Site.............................................. 106
Figure III.4: Zoom-up of the Satelite Image of the Lyngbya Collection Site............... 106
Figure III.5: "Green Lyngbya" collected in Papua New Guinea................................... 107
Figure III.6: 1H NMR Spectrum of Credneric Acid (38).............................................. 108
Figure III.7: Credneric Acid and Credneramide........................................................... 112
Figure III.8: N-Acyl Homoserine Lactones (AHLs)..................................................... 112
Figure III.9: Credneric Acid (38) 13C NMR Spectrum in CDCl3.................................. 116
Figure III.10: Credneric Acid (38) 1H-1H COSY Spectrum in CDCl3.......................... 117
Figure III.11: Credneric Acid (38) 1H-13C HSQC Spectrum in CDCl3......................... 118
Figure III.12: Credneric Acid (38) 1H-13C HMBC Spectrum in CDCl3........................ 119
Figure III.13: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 2.73 ppm in
CDCl3............................................................................................................................ 120
Figure III.14: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 5.79 ppm in
CDCl3............................................................................................................................ 121
Figure III.15: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 2.16 ppm in
CDCl3............................................................................................................................ 122
Figure III.16: IR spectrum of 38................................................................................... 123
Figure IV.1: Proposed transition state of alkylation of the enolate of 5 guided by
oxazolidinone
auxiliary........................................................................................................................ 132
Figure IV.2: Dihedral angles predicted from the 1H-1H coupling constants and
Altona equation (A) and select ROESY correlations for 7 and 8 (B)........................... 135
Figure IV.3: History of the absolute stereochemical assignment of cyclopropane
derivative 11.................................................................................................................. 136
Figure IV.4: 1H and 13C NMR assignments and an NOE correlation of 16 in CDCl3 138
Figure IV.5: X-ray crystal structure of 15..................................................................... 139
Figure IV.6: Corrected stereochemistry of 1-m and 1-c............................................... 139
Figure IV.7: Synthetic Introduction of a Stereochemical Reference (SISTER)
method conceptual figure............................................................................................. 141
Figure IV.8: 1H NMR spectrum of 7 in C6D6................................................................ 146
Figure IV.9: 1H NMR spectrum of 7 in CDCl3............................................................. 147
Figure IV.10: 13C NMR spectrum of 7 in CDCl3.......................................................... 148
Figure IV.11: 1H-1H Homonuclear decoupling experiment for 7 in C6D6 – Part 1....... 149
Figure IV.12: 1H-1H Homonuclear decoupling for 7 in C6D6 – Part 2.......................... 150
xi
Figure IV.13: 1H-1H Homonuclear decoupling experiment for 7 in C6D6 – Part 3....... 151
Figure IV.14: 1H-1H COSY spectrum of 7 in C6D6....................................................... 152
Figure IV.15: ROESY spectrum of 7 in C6D6............................................................... 153
Figure IV.16: LR FAB MS spectrum of 7..................................................................... 154
Figure IV.17: IR spectrum of 7..................................................................................... 155
Figure IV.18: 1H NMR spectrum of 8 in C6D6.............................................................. 156
Figure IV.19: 1H NMR spectrum of 8 in CDCl3........................................................... 157
Figure IV.20: 13C NMR spectrum of 8 in CDCl3.......................................................... 158
Figure IV.21: 1H-1H Homonuclear decoupling experiment for 8 in C6D6 – Part 1....... 159
Figure IV.22: 1H-1H Homonuclear decoupling experiment for 8 in C6D6 – Part 2....... 160
Figure IV.23: 1H-1H Homonulcear decoupling experiment for 8 in C6D6 – Part 3 161
Figure IV.24: 1H-1H COSY spectrum of 8 in C6D6....................................................... 162
Figure IV.25: ROESY spectrum of 8 in C6D6............................................................... 163
Figure IV.26: IR spectrum of 8..................................................................................... 164
Figure IV.27: 1H NMR spectrum of 9 in CDCl3........................................................... 165
Figure IV.28: IR spectrum of 9..................................................................................... 166
Figure IV.29: Geometry optimized 3D structures of 7 (above) and 8 (below)............. 167
Figure IV.30: HPLC chromatogram for 4 and 7........................................................... 168
Figure IV.31: HPLC chromatogram for 5 and 8........................................................... 169
Figure IV.32: UV spectra of 7 and 8............................................................................. 170
Figure V.1: Frog skin alkaloids of various core skeletons............................................ 174
Figure V.2: Skin of Epipedobates tricolor contains epibatidine and epiquinamide...... 175
Figure V.3: X-ray crystal structure of epi-epiquinamide (15)....................................... 178
Figure V.4: 1H NMR Spectrum of 11 in CDCl3............................................................ 200
Figure V.5: 13C NMR Spectrum of 11 in CDCl3........................................................... 201
Figure V.6: 1H NMR Spectrum of 12 in CDCl3............................................................ 202
Figure V.7: 13C NMR Spectrum of 12 in CDCl3........................................................... 203
Figure V.8: 1H NMR Spectrum of 13 in CDCl3............................................................ 204
Figure V.9: 13C NMR Spectrum of 13 in CDCl3........................................................... 205
Figure V.10: 1H NMR Spectrum of 14 in CDCl3.......................................................... 206
Figure V.11: 13C NMR Spectrum of 14 in CDCl3......................................................... 207
Figure V.12: 1H NMR Spectrum of 15 in MeOH-d4..................................................... 208
Figure V.13: 13C NMR Spectrum of 15 in MeOH-d4.................................................... 209
Figure V.14: 1H NMR Spctrum of 16 in CDCl3............................................................ 210
xii
Figure V.15: 13C NMR Spectrum of 16 in CDCl3......................................................... 211
Figure V.16: EI-GCMS Chromatogram and Mass Spectrum of 16.............................. 212
Figure V.17: 1H NMR spectrum 29 in CDCl3............................................................... 213
Figure V.18: 13C NMR spectrum of 29 in CDCl3.......................................................... 214
Figure V.19: 1H NMR spectrum of 30 in CDCl3........................................................... 215
Figure V.20: 13C NMR spectrum of 30 in CDCl3.......................................................... 216
Figure V.21: 1H NMR spectrum of 28 in CDCl3........................................................... 217
Figure V.22: 13C NMR spectrum of 28 in CDCl3.......................................................... 218
Figure V.23: 1H NMR spectrum of 31 in CDCl3........................................................... 219
Figure V.24: 13C NMR spectrum of 31 in CDCl3.......................................................... 220
Figure V.25: 1H NMR spectrum of 27 in CDCl3........................................................... 221
Figure V.26: 13C NMR spectrum of 27 in CDCl3........................................................ 222
Figure V.27: 1H NMR spectrum of 32 in CDCl3........................................................... 223
Figure V.28: 1H NMR NMR spectrum of (+)-9 in CDCl3............................................ 224
Figure V.29: 1H NMR spectrum of (+)-9 in CDCl3...................................................... 225
Figure V.30: 13C NMR spectrum of (+)-9 in CDCl3...................................................... 226
Figure V.31: 1H NMR spectrum of (-)-9 in CDCl3....................................................... 227
Figure V. 32: 13C NMR spectrum of (-)-9 in CDCl3..................................................... 228
Figure Vl.1: Antiangiogenic activity of 1 in zebrafish................................................. 236
Figure VI.2: Impact of somocystinamide A (1) on proliferation of paired
neuroblastoma cells lacking caspase 8 expression (triangles) or expressing caspase
8 (squares)..................................................................................................................... 238
Figure VI.3: Extrinsic pathway via activation of caspase-8 leading to apoptosis........ 238
Figure VI.4: Intrinsic pathway via activation of caspase-9 leading to apoptosis......... 239
Figure VI.5: Most stable conformer of somocystinamide A found through molecular
modeling....................................................................................................................... 240
Figure VI.6: Proposed biosynthesis of somocystinamide A (1)................................... 243
Figure VI.7: Enamide containing natural products....................................................... 244
Figure VI.8: Examples of disulfide-containing natural products................................. 247
Figure VI.9: Model compounds used in the study of enamide formation.................... 255
Figure VI.10: Reaction setup for enamide formation using molecular sieves and
glass wool in a Soxhlet extractor.................................................................................. 259
Figure VI.11: Solubility of somocystinamide A (1) in various media.......................... 261
Figure VI.12: 1H NMR spectrum of 22 in CDCl3......................................................... 282
Figure VI.13: 13C NMR spectrum of 22 in CDCl3........................................................ 283
xiii
Figure VI.14: 1H NMR spectrum of 2 in CDCl3........................................................... 284
Figure VI.15: 13C NMR spectrum of 2 in CDCl3.......................................................... 285
Figure VI.16: 1H NMR spectrum of 44 in CDCl3......................................................... 286
Figure VI.17: 13C NMR spectrum of 44 in CDCl3........................................................ 287
Figure VI.18: 1H NMR spectrum of 51 in CDCl3......................................................... 288
Figure VI.19: 13C NMR spectrum of 51 in CDCl3........................................................ 289
Figure VI.20: 1H NMR spectrum of 52 in CDCl3......................................................... 290
Figure VI.21: 13C NMR spectrum of 52 in CDCl3........................................................ 291
Figure VI.22: 1H NMR spectrum of 52-cis in CDCl3................................................... 292
Figure VI.23: 13C NMR spectrum of 52-cis in CDCl3.................................................. 293
Figure VI.24: 1H NMR spectrum of 56 in CDCl3......................................................... 294
Figure VI.25: 13C NMR spectrum of 56 in CDCl3........................................................ 295
Figure VI.26: COSY spectrum of 56 in CDCl3............................................................. 296
Figure VI.27: HMBC spectrum of 56 in CDCl3........................................................... 297
Figure VI.28: 1H NMR spectrum of 57 in CDCl3......................................................... 298
Figure VI.29: 13C NMR spectrum of 57 in CDCl3........................................................ 299
Figure VI.30: 1H NMR spectrum of 3 in 1:1 CDCl3/CD3OD....................................... 300
Figure VI.31: 13C NMR spectrum of 3 in 1:1 CDCl3/CD3OD...................................... 301
Figure VI.32: 1H NMR spectrum of 1 in CDCl3........................................................... 302
Figure VI.33: 13C NMR spectrum of 1 in CDCl3.......................................................... 303
Figure VI.34: 1H NMR spectrum of 70 in CDCl3........................................................ 304
Figure VI.35: 13C NMR spectrum of 70 in CDCl3........................................................ 305
Figure VI.36: 1H NMR spectrum of 71 in CDCl3......................................................... 306
Figure VI.37: 13C NMR spectrum of 71 in CDCl3........................................................ 307
Figure VI.38: 1H NMR spectrum of 73 in CDCl3......................................................... 308
Figure VI.39: 13C NMR spectrum of 73 in CDCl3........................................................ 309
Figure VII.1: Light absorption spectrum of scytonemin (1) (taken from ref. 4).......... 318
Figure VII.2: Select HMBC, COSY, and ROESY correlations of scytonemin
monomer (22) in DMF-d7............................................................................................. 323
Figure VII.3: Scytonemin monomer (22): geometry optimized computer model........ 324
Figure VII.4: 1H NMR spectrum of 22 in DMF-d7....................................................... 325
Figure VII.5: pH-dependent color change of monomer 22.......................................... 328
Figure VII.6: Alternative structures for scytonemin..................................................... 335
Figure VII.7: 1H NMR spectrum of 21 in CDCl3.......................................................... 346
Figure VII.8: 13C NMR spectrum of 21 in CDCl3......................................................... 347
xiv
Figure VII.9: 1H NMR spectrum of 18 in CDCl3......................................................... 348
Figure VII.10: 13C NMR spectrum of 22 in DMF-d7.................................................... 349
Figure VII.11: COSY NMR spectrum of 22 in DMF-d7.............................................. 350
Figure VII.12: HSQC NMR spectrum of 22 in DMF-d7.............................................. 351
Figure VII.13: HMBC NMR spectrum of 22 in DMF-d7............................................. 352
Figure VII.14: NOESY NMR spectrum of 22 in DMF-d7............................................ 353
Figure VII.15: HRMS (EI) spectrum of 22................................................................... 354
Figure VII.16: 1H NMR spectrum of 29 in CDCl3........................................................ 355
Figure VII.17: 13C NMR spectrum of 29 in CDCl3....................................................... 356
Figure VII.18: COSY spectrum of 29 in CDCl3........................................................... 357
Figure VII.19: HSQC spectrum of 29 in CDCl3........................................................... 358
Figure VII.20: HMBC spectrum of 29 in CDCl3.......................................................... 359
Figure VII.21: ROESY spectrum of 29 in CDCl3......................................................... 360
Figure VII.22: HRMS (ESI-FT) spectrum of 29 (showing [M+Na])........................... 361
Figure VII.23: 1H NMR spectrum of 34 in CDCl3........................................................ 362
Figure VII.24: 13C NMR spectrum of 34 in CDCl3....................................................... 363
Figure VII.25: 1H NMR spectrum of 37 in CDCl3........................................................ 364
Figure VII.26: 13C NMR spectrum of 37 in CDCl3....................................................... 365
Figure VIII.1: Natural products isolated during the dissertation research.................... 371
Figure VIII.2: Natural products that were synthetically investigated during the
disseration research....................................................................................................... 372
xv
LIST OF SCHEMES
xvi
Scheme V.10: Chelation-controlled hydride reduction of ketone (30)....................... 183
Scheme VI.1: Enamide............................................................................................... 242
Scheme VI.2: Instability of enamide as seen in the facile decomposition of
somocystinamide A..................................................................................................... 243
Scheme VI.3: Enamide synthesis on model systems through acylation of imine....... 246
Scheme VI.4: Different approaches to enamide functional group.............................. 247
Scheme VI.5: Approaches to install disulfide of the epidithiapiperazinedione core.. 248
Scheme VI.6: Total synthesis of glyotoxin involving sensitive installation of
disulfide...................................................................................................................... 248
Scheme VI.7: Synthesis of psammaplin A.................................................................. 249
Scheme VI.8: Retrosynthetic Analysis of Somocystinamide A.................................. 249
Scheme VI.9: Synthesis of vinyl iodide for Suzuki coupling (Scheme VI.10).......... 250
Scheme VI.10: Suzuki coupling with varying coupling partners............................... 250
Scheme VI.11: Ruthenium catalyzed cross metathesis approach............................... 251
Scheme VI.12: Initially planned final steps to Somocystinamide (1)......................... 254
Scheme VI.13: Possible mechanism of enamide formation....................................... 255
Scheme VI.14: Copper-mediated vinylation of a primary amide model compound
(65) and attempted chemoselective methylation......................................................... 257
Scheme VI.15: Final steps to somocystinamide A (1)................................................ 258
Scheme VI.16: Revised synthesis of the bis-methylamide 3...................................... 263
Scheme VII.1: Fragmentation of the reduced form of scytonemin (2) through
ozonolysis in its structure elucidation......................................................................... 319
Scheme VII.2: Biosynthetic gene cluster for scytonemin in Nostoc punctiforme...... 319
Scheme VII.3: Chemical synthesis of a putative biosynthetic precursor of
scytonemin (1), soraphinol A (S-8)............................................................................. 320
Scheme VII.4: Proposed mechanism for the biosynthesis of scytonemin (1) by
Balsksus and Walsh..................................................................................................... 321
Scheme VII.5: Initially conceived synthesis of scytonemin monomer 11.................. 322
Scheme VII.6: Predicted mechanism for regioselectivity of aldol condensation
between ketone 18 and aldehyde 19........................................................................... 322
Scheme VII.7: Synthesis of scytonemin monomer..................................................... 323
Scheme VII.8: Absence of tautomerization and resonance characteristics of
scytonemin monomer (22).......................................................................................... 324
Scheme VII.9: Proposed mcehanism for dimerization of oxidized monomer 25 to
reduced scytonemin 2................................................................................................. 328
Scheme VII.10: Synthesis of methylated monomer 29............................................... 331
Scheme VII.11: Possible role of glycosylation in the biosynthesis of scytonemin.... 332
xvii
Scheme VIl.12: Attempts to synthesize 30 and/or a mimic thereof (e.g. 33)............. 333
Scheme VII.13: Attempts to perform Rubottom oxidation via doubly silylation of
22................................................................................................................................. 334
Scheme VII.14: Newly proposed route for the synthesis of 2.................................... 336
Scheme VII.15: Highly chemo-selective DDQ oxidation of (E)-16-epinormacusine
B (45).......................................................................................................................... 337
Scheme VII.16: Chemo-selective oxidation of benzylic ether by DDQ in a neat
mixture........................................................................................................................ 337
Scheme VII.17: Completion of total synthesis of scytonemin albeit low yield
(yield to be determined).............................................................................................. 338
xviii
LIST OF TABLES
xix
ACKNOWLEDGMENTS
I would like to first thank my PhD advisor, Prof. William Gerwick, who has given
me great opportunities in his laboratory. His enthusiasm and mentorship were invaluable
during my graduate program. Bill has also been a great person to work for, and he has
been greatly patient with me through the hard and slow times that are associated with
as well. I am very grateful for the many hours that he has spent editing, reviewing, and
critiquing my manuscripts for this disseration and other publications. I would not like to
It was very helpful to discuss research issues with other members of my PhD
Bradley Moore, and William Fenical for taking their time out to be my advisors. I thank
Professor Fenical for allowing me to work in his laboratory while our new labs at SIO
I am very thankful for all the postdoctoral researchers, graduate students, and staff
that have worked with me over the years both at OSU and UCSD. I particularly enjoyed
and benefited from bouncing various ideas off Keith Schwartz (Dr. White's lab at OSU)
and my officemate, Karla Malloy. I acknowledge the technical assistance from Dr.
Anthony Mrse, Dr. Yongxuan Su, Dr. Harry Gross, Dr. Arnold Rhinegold and his staff,
Josh Wingerd, and Tara Byrum at UCSD and Roger Kohnert at OSU. I am especially
indebted to Professor Kerry McPhail at OSU, who helped me in the early years of my
scientific career.
xx
I must not forget to thank Professor David A. Horne at OSU, who first introduced
me to the joy of synthetic organic and bioorganic chemistry and taught me invaluable
skills in his laboratory. Going back further in time, I am thankful for my high school
biology teacher, Mr. Kudara, who inspired me to pursue a career in science back when all
I ever thought about was improving my batting average and hitting a home run some day
On a more personal note, I am thankful for the ministry by Dr. Kent Hovind and
his associates, who have inspired me to appreciate having a doctoral degree in science for
the Kingdom of our God. I am also very grateful for the church family I have in
Oceanside, CA and in Albany, OR. I cannot replace their fellowship, support, and prayers
with anything else. I am extremely grateful for Pastor Ron Reed, who has been a great
friend and a mentor despite the long distance between us. Knowing that he prays to our
I am deeply thankful for my parents and other family members for their support.
Thank you! I am especially grateful for my mother, who came to the Lord a few years
To my dear wife, Janet, I cannot thank God enough for you. Indeed, “whoso
findeth a wife findeth a good thing, and obtaineth favour of the LORD (Proverbs 18:22,
KJV).” You have been a great gift from the Lord. Your support, love, and friendship have
been a great source of strength. Thank you, honey. I look forward to serving our Lord
together with you in the years to come. I also thank you for your help with the tedious
xxi
Lastly, but most definitely not the least, I owe everything to my greatest mentor,
hero, and friend, the Lord Jesus Christ. Lord, You have taken my place and sacrificed
Your life for me, a wretched sinner who was a miserable atheist until six years ago, so
that I may be forgiven and have an eternal life with You. Knowing this fact has been an
infinite source of confidence and strength in all the situations I have faced in the last six
years. Any difficult situation seemed so small compared to the hope You have given me
through Your resurrection. With indescribable gratitude, all I can say is “Now unto Him
[Jesus] that is able to keep you from falling, and to present you faultless before the
presence of His glory with exceeding joy, to the only wise God our Savior, be glory and
majesty, dominion and power, both now and ever. Amen (Jude 24-25).” Hallelujah!
journal. The dissertation author was the primary investigator and author of the research
which forms the basis of this chapter. A coauthor will be William H. Gerwick.
The text of II, in whole, is the manuscript that has been submitted to Toxicon as it
will appear: Takashi L. Suyama, Zhengyu Cao, Thomas F. Murray, and William H.
dissertation author was the primary investigator and author of the research which forms
academic journal. The dissertation author was the primary investigator and author of the
xxii
research which forms the basis of this chapter. Coauthors will include Karla L. Malloy
academic journal. The dissertation author was the primary investigator and author of the
research which forms the basis of this chapter. Coauthors will include Kerry L. Mcphail,
the primary investigator and author of the research which forms the basis of this
chapter.
author was the primary investigator and author of the research which forms the
basis of this chapter. A coauthor for the second publication will be William H.
Gerwick.
The text of VII, in part, will be used for two manuscripts to be submitted
to two different academic journals. The dissertation author was the primary
investigator and author of the research which forms the basis of this chapter.
xxiii
VITA
PUBLICATIONS
xxiv
Wingerd; Teatulohi Matainaho; William H. Gerwick, Journal of Natural
Products 2008, 71, 1099-1103.
•“Stereospecific Total Synthesis of Somocystinamide A” Takashi L. Suyama;
William H. Gerwick, Organic Letters 2008, 10, 4449-4452.
•“Lipoproteins, lipopetides and analogs, and methods for making and using
them.” William Gerwick; Wolfgang Wrasidlo; Dennis Carson; Dwayne Stupack;
Takashi Suyama. PCT Int. Appl. 2008, 119pp.
•“Ichthyotoxic Brominated Diphenyl Ethers from a Mixed Assemblage of a Red
Alga and Cyanobacterium: Structure Clarification and Biological Properties.”
Takashi L. Suyama; Zhengyu Cao; Thomas F. Murray; William H. Gerwick.
Toxicon, 2009, Manuscript in review.
•“Vinylchloride-Containing Fatty acid and its Neurotoxic Phenethylamide
Derivative from a Papua New Guinea Collection of Lyngbya sp.” Takashi L.
Suyama; Karla L. Malloy; Zhengyu Cao; Thomas F. Murray; William H.
Gerwick. J. Nat. Prod. 2009, Manuscript in preparation.
•“Probing the Enzymatic Potential of Scy1263, a Hypothetical Protein from the
Scytonemin Biosynthetic Gene Cluster in Nostoc punctiforme ATCC 29133”
Carla M. Sorrels; Takashi L. Suyama; William H. Gerwick. Manuscript in
preparation.
•“Expedient Synthesis of α,α-dimethyl-β-hydroxy Carbonyl Scaffolds via Evans'
Aldol Reaction with a Tertiary Enolate” Joshawna K. Nunnery; Takashi L.
Suyama; William H. Gerwick. Manuscript in preparation.
•“Recent Structural Revisions of Natural Products” Takashi L. Suyama; William
H. Gerwick. Manuscript in preparation.
•“Insights into the Biosynthesis of Scytonemin through Synthetic Investigations”
Takashi L. Suyama; William H. Gerwick. Manuscript in preparation.
•“Improved Synthesis of Somocystinamide A” Takashi L. Suyama; William H.
Gerwick. Manuscript in Preparation.
FIELDS OF STUDY
xxv
ABSTRACT OF THE DISSERTATION
by
Takashi L. Suyama
The partnership between natural products and synthetic organic chemistry date
back to the origin of organic chemistry itself. While natural products became a major
driving force for the development of novel organic reactions and synthetic strategies,
organic synthesis has contributed in many ways to the elucidation and confirmation of
to the recent advances in both of these two disciplines, there are new opportunities and
xxvi
issues surrounding natural products that organic synthesis can be applied to, and such
Chapter I introduces the background information and rationale for the dissertation
research, which is based on the history of organic chemistry and newly emerging research
of a marine red alga and cyanobacteria. Chapter III describes the isolation,
credneric acid, from a Lyngbya sp. Chapter IV describes the conception of a method for
fatty acid and resulted in a conflict with newly published literature, which is discussed in
detail. Chapter V describes the development of an expedient and efficient total synthesis
Epipedobates tricolor. This work contributed to clarifying the identity of a potent and
endothelial cell proliferation and angiogenesis. The synthesis was later modified for
scalability to meet the need for further pharmaceutical investigations of this important
xxvii
intriguing insights into the biosynthesis of the natural product were obtained. Finally,
Chapter VIII provides the conclusions drawn from this dissertation research.
xxviii
Chapter I
Introduction
1
2
Abstract
partnership. While natural products became the major driving force for development of
novel organic reactions and synthetic strategies, organic synthesis has contributed in
and biosynthetic studies of natural products. Historical and modern examples of these
“I can no longer, as it were, hold back my chemical urine; and I have to let
out that I can make urea without needing a kidney, or even of an animal,
whether of man or dog: the ammonium salt of cyanic acid (cyansäures
Ammoniak) is urea.”
Friedrich Wöhler (1828)1
With the advent of chemical nomenclature and the acceptance of the atomic
theory, chemistry emerged as a scientific discipline during the late eighteenth century. 2
study of organic compounds did not begin until Wöhler synthesized urea from
organic compounds could only be made by “vital forces” in living organisms and that
they obeyed laws other than the physical laws observed for inorganic compounds.2 The
new field of organic chemistry flourished due to the developing confidence that organic
compounds could be studied logically and systematically in the same way inorganic
compounds were being studied. The backbone of this breakthrough was the structural
theory, which explained that the diversity and chemical properties of organic compounds
are due to the way the constituting atoms are connected to one another. The introduction
of this theory is regarded as “of the same importance in the history of science as the
development of the two laws of thermodynamics around 1850, the quantum theory and
the theory of general relativity after 1900, and the explanation of the molecular basis of
This diagram utilizes the principles of Venn diagram. Where there is an overlap between different
fields, strong interdisciplinary influence is implied. The order of overlaps is meant to roughly depict
the order of emergence of each discipline chronologically.
5
Due to scientists' constant interest in nature and health, many organic chemists of
compounds, isolated from living organisms.5 In time, two sub-disciplines were formed to
essential for normal growth, sustaining of life, and reproduction of the organism.
Examples of primary metabolites include proteins, carbohydrates, fats, and nucleic acids.
Unlike primary metabolites, the absence of secondary metabolites would not result in
mankind for various purposes. They have been used as medicines for pain relief and
healing of wounds, as poisons for hunting and warfare, as agents for capital punishment,
flavor of food, as perfume to obscure the odor of dead bodies, and in religious worship
and anointment (Table I-1).5,6 A great many of these natural products are still used today
and often for the same purposes.7 The difference is that the full chemical identities of the
natural products are now mostly known through advances in chemical sciences.
The rise of natural product chemistry not only stimulated the growth of biology,
but it also had a great impact on the art and science of organic synthesis (Illustration I.1).
The complex and elegant structures of many naturally occurring compounds became the
favored targets of total synthesis. In its infancy, the field of organic total synthesis
haemin (1)9 and equilenin (2),10 that would be considered complex even by today's
In the classical era, many of the natural product chemists had their primary
training in synthetic organic chemistry. Often times, the only possible means of structure
method is echitamine (8), a tumoricidal alkaloid isolated from the bark of Alstonia
glyoxylic reagent (Scheme I.1). Treatment with aqueous base and acid revealed a methyl
ester and imine-like moiety, respectively. Much structural information around the olefin
8
was obtained by the observed reduction with palladium and the resulting fragments from
ozonolysis. Acetylation added two acetate units, thereby showing that there are two
hydroxyl groups. Further chemical analysis of a similar nature, the molecular formula
derived by mass spectrometry, and the compound's chemical properties finally led to
structure 7 (Scheme I.2).14 Even though 7 was very similar to the correct structure 8, the
true identity of this intriguing alkaloid was not discovered until X-ray crystallography
was applied in 1962. Since the initial isolation of this natural product,15 some 85 years
spectroscopic structural analysis of natural products appeared and consequently, the field
of natural product chemistry matured.12,16 Today, in many cases, the processes of natural
Subsequently, the focus of natural product chemists has shifted from merely describing
molecules.17 However, the age old partner of natural products chemistry, chemical
synthesis, still plays a major role today in confirmation of the chemical structure of the
natural product, particularly its stereochemistry, and in providing a larger quantity of the
and mururin C (12). In the first case, the issue concerned stereochemical assignments,
while an incorrect planar structure was assigned to the latter compound. Such revisions
were required for 30 natural products in 2008 alone.18 It is safe to say that the art of
structure elucidation has not achieved perfection and that the partnership between natural
products chemistry and synthetic organic chemistry is still very much in order.
by Cimino and coworkers.19 Originally, structure 10 was proposed based on NMR and
coworkers suggested that the correct structure is most likely 9.20 A biogenetic rationale
for structure 9 was inspired by Cimino and Faulkner, who noted that the extensive
conrotatory product, 9, whereas the corresponding thermal reaction would yield the
Indeed, Moses and coworkers synthesized 11, which was converted to 9 upon
exposure to sunlight.23 Furthermore, this synthetic material, 9, had the same NMR and
failed.23 These results not only confirm that 9 is the correct structure for this natural
product, but also suggests that the actual biosynthetic product produced by Tridachia may
revision as well as insights into its biosynthesis that could not have been easily obtained
Figure I.1: The originally assigned structure (13) of mururin C and its
revised structure (12)
Although the majority of structural revisions of natural products have been due to
interesting to consider the less frequent cases of incorrectly assigned planar structures.
Such was the case with mururin C (12), a natural product isolated from the bark of
product based on HR-MS and 1D and 2D NMR analyses.24 In the course of the synthesis
12
of 13, however, it was found that allenes analogous to 13, when formed in situ, were
intercepted readily by nucleophiles in the reaction mixture and the allene species could
not be isolated.25 This observation led to the rationale that if 13 was formed at all in the
course of biosynthesis, the hydroxyl group of the phenol would attack the allenic carbon
and form benzofuran, thereby converting to structure 12. Therefore it is highly unlikely
that 13 was the correct structure. Moreover, comparison of predicted chemical shifts for
12 and 13 revealed that 12 is very likely the correct structure. In this work, synthetic
studies of the natural product eloquently revealed the erroneous feature of the originally
In 2005, Nicolaou and Snyder published a fairly comprehensive and inspiring paper
on natural product structural revisions that were made possible by total synthesis in the
period of 1990 to 2004.16 In that paper, the authors articulate the amazement, in contrast
to their expectation, of discovering the large number and profoundness of structural mis-
assignments made in the recent years. In Table I.2, all of the structural revisions 26 of
The year in which the initial structure was published is in parentheses. Only the structure
determination methods used for the portion of the structure that is erroneous are mentioned in this
table. Predict.: predictions based on molecular modeling. Mol. Mod.: use of geometry optimized
structure.
Total
NOE
Synthesis28
Calafianin27 (2000)
Total
NOESY
Synthesis30
Tribulusterine29 (1999)
Total
Absence of NOE
Synthesis32
Ayapin derivative31 (1999)
1
H & 13C NMR
Total
HMBC
Synthesis34
NOE
Isoaurostatin33 (2001)
1
H & 13C NMR
Total
HMQC
Synthesis36
HMBC
Chloroaurone35 (2001)
14
NOESY Total
HMBC Synthesis38
Feigrisolide A37 (2000)
NOESY Total
HMBC Synthesis38
Feigrisolide B37 (2000)
HMBC Total
NOESY Synthesis40
Derivatization Total
NMR Synthesis42
Agardhilactone41 (1996)
EIMS Total
Fragmentation Synthesis44
Pyrinodemin A43 (2001)
CD
Total
Comparison Synthesis46
NMR
CD
Total
Comparison Synthesis46
NMR
J-based Total
ROESY Synthesis48
X-ray
HMBC
NOE50
Aspernigrin A49 (2004)
1 NMR
H NMR
Comparison52
1 NMR
H NMR
Comparison52
Derivatization
Total
J-based
Synthesis54
Mosher
Amphidinolide W53 (2002)
Comparison
NOE Total
55 Synthesis56
Batzelladine F (1997) DCI-MS
Fragmentation
Total
NOE
Synthesis58
Brevenal57 (2005)
Total
NOE
Synthesis60
Trihydroxycadinane59 (1998)
Total
Comparison
Synthesis62
Aspergione B61 (2003)
Semi-
synthesis
1
H & 13C NMR
NMR
Botcinic acid63 (1993)
Comparison64
(Botcinolide)
17
NMR NMR
Comparison Comparison65
Homobotcinic acid65 (1996)
(Homobotcinolide)
NMR NMR
Comparison Comparison65
Botcinic acid methyl ester66
(1996) (4-O-methylbotcinolide)
EI-MS
HMBC 13
C Prediction
NOE NMR
Peribysin C67 (2004)
comparison68
EI-MS
HMBC 13
C Prediction
NOE NMR
Peribysin C62 (2004)
comparison61
Chemical
ROESY Reactivity
CD predict.70
Pyranonigrin69 (2004)
FAB-MS
HMBC72
Degradation
Shatavarin I71 (1987)
18
EI-MS
Degradation HMBC57
OR predict
Shatavarin IV58 (1987)
HMBC
1D & 2D NMR NMR
Comparison74
Fusapyrone73 (1994)
HMBC
1D & 2D NMR NMR
Comparison74
Deoxyfusapyrone73 (1994)
HMQC
COSY Total
Comparison Synthesis76
Solandelactone A75 (1996)
J-based
Comparison Total
J-based Synthesis76
75
Solandelactone B (1996)
Comparison Total
J-based Synthesis76
75
Solandelactone C (1996)
19
Comparison Total
J-based Synthesis76
75
Solandelactone D (1996)
Comparison Total
J-based Synthesis76
Solandelactone E75 (1996)
Comparison Total
J-based Synthesis76
Solandelactone F75 (1996)
Comparison Total
J-based Synthesis76
75
Solandelactone G (1996)
Comparison Total
J-based Synthesis76
75
Solandelactone H (1996)
NOESY
NOESY78
NMR Comprison
Barlerin77 (2004)
20
NOESY
NOESY78
NMR Comprison
Penstemoside77 (2004)
NOESY
NOESY78
NMR Comprison
Dehydropenstemoside77 (2004)
HR-MS
NMR
NOESY
Comparison78
HMBC
Harpagoside79 (2003)
Degradation
GCMS NOESY78
OR
80
Agnuside (2004)
21
NMR
NOESY78
Comparison
3,4-hydroxybenzoylaucubin80
(2004)
COSY
1
H NMR
HMBC
CD
X-ray82
Camellenodiol81 (1981)
1
H NMR COSY
CD HMBC82
Camelledionol81 (1981)
ROESY
ROESY
Mol. Mod.
J-based
J-based84
Palau'amine83 (1993)
Total
D2O Shift86
Synthesis87
Intricatin85 (1989)
22
Total
Synthesis89
Degradation
First partial
1 D & 2D NMR
synthesis had
an error.90
Kedarcidin88 (1993)
J-based Total
NOESY Synthesis92
Total
NOE
Synthesis96
Biyouyanagin95 (2005)
Total
Synthesis98
Aciphyllene97 (1983)
23
Mosher
Total
J-based
Synthesis100
NOE
Palmerolide A99 (2006)
EI-MS
1 Total
H, 13C NMR
Synthesis102
Benz[g]isochromene-dione101 HETCOR
(1995)
Total
NOE
Synthesis104
Laurentristich-4-ol103 (2005)
1 Total
H, 13C NMR
Synthesis106
Biomimetic
Semi-
synthesis108
Xuxuarine Eα
107
Rzedowskia bistriterpenoid α
(1996)
24
Biomimetic
Semi-
synthesis108
Xuxuarine Eα
25-acetoxy-3a-hydroxyolean-12-
en-28-oic acid111 (2004)
ROESY Total
NOE Synthesis114
Total
J-based
Synthesis116
Diversifolide115 (1999)
1D & 2D NMR
IR Total
HR EI-MS Synthesis118
NOE
Labdane diterpenoid117 (2006)
Total
NOE
Synthesis120
Netamine E119 (2006)
Total
NOE
Synthesis120
J-based
Mosher Total
Derivatization Synthesis122
Tyroscherin121 (2004)
Optical Rotation
26
Ozonolysis
Combinatorial
Derivatization synthesis of
fragments124
Chiral HPLC
Amphidinol 3 Fragment123
Derivatization Total
Chiral GCMS Synthesis126
HMBC
ROESY 13
C Predict.
X-ray128
Spiroleucettadine127 (2004)
Mol. Mod.
13
NMR C Predict.130
Comparison
Obtusallene V129 (2000)
27
Mol. Mod.
13
NMR C Predict.130
Comparison
Obtusallene VI129 (2000)
Mol. Mod.
13
NMR C Predict.130
Comparison
Obtusallene VII129 (2000)
INEPT2-
INADEQUATE HMBC
HMBC- X-ray132
Circumdatin A131 (1999) INADEQUATE
INEPT2-
INADEQUATE HMBC
HMBC- X-ray132
INADEQUATE
Circumdatin B131 (1999)
28
HR-MS
X-ray134
1
H NMR
Bromoditerpene133 (1987)
NOE
NMR NOE136
Comparison
Hiyodorilactone B135 (1978)
ESI-MS
X-ray138
Rhizopodin137 (1993)
NMR
Comparison
ESI HR-MS
Semi-
HMBC
synthesis
Bioinformatics
ESI HR-
Microcyclamide 139
(2008) MS140
HMBC
13
Biosynthetic C Predict.
Prediction UV Predict.142
Hassanane141 (1996) Mol. Mod.
29
Comparison
NOESY with synthetic
ESI HR-MS model
compounds144
Zamamistatin 143
(2001) See next entry
Comparison with IR
synthetic model NMR and OR
compounds Comparison145
Zamamistatin144 (2001)
30
60%
Stereochemical
Constitutional
Constitutional &
Stereochemical
15%
25%
24% 26%
NOE
Comparison
J-based
Deriv atization
NMR
CD/OR
2% Model
4% Chromatography
MS
18% 4%
6%
12% 6%
Figure I.3: Structure determination methods used in 2005-2008 for the erroneous cases
54%
Total sy nthesis
Other sy nthetic
X-ray
Others
8%
8% 29%
years (1990-2008). Of the 83 structure revisions, 40% had constitutional errors (Figure
I.2). This statistic comes as a surprise in this century when NMR tools, such as HMBC, to
connect elucidated substructures are readily available (Figure I.3). However, a close
examination of these structures reveals a subtle problem. Unless all theoretically possible
HMBC correlations are observed, some structures are indistinguishable from their
The HMBC correlations necessary for distinction between the two possible structures are depicted
as blue and orange arrows. If the blue correlation was only present, structure 14 would be correct.
Likewise, if the orange correlation was only present, then structure 15 would be correct.
Figure I.6: HMBC correlations observed and recorded for the labdane diterpenoid108
The HMBC correlations necessary for discrimination of structures 14 and 15, as noted in Figure
I.5, were not recorded (or observed) for this natural product.
32
(Figure I.4).
one case where total synthesis clarified the true identity of the bioactive metabolite. In
2003, Daly and coworkers isolated a novel quinolizidine alkaloid, epiquinamide (16)
was believed to be a subtype selective nicotinic receptor agonist and it attracted much
attention from the synthetic community. By 2006, there had been several asymmetric
evidence surfaced that the originally observed activity might have been due to a minor
contamination of the sample with epibatidine.148 Awareness of this error was made
possible by the synthetic supply of 16 free of any other bioactive natural product
products advance further, the need for supplying pure samples of natural products by total
Successful attempts have been made to present the case that natural products are a
great resource for drug development.149 Indeed, among all of the drugs that were
approved between 1981 and 2006 for clinical use, 63% had their ultimate origin in
33
natural products (Figure I.7).93 This recent success of natural products did not emerge
from a vacuum, but it has been nurtured by the partnership with synthetic chemistry since
the conception of organic chemistry. In the earliest era of organic chemistry, it was rarely
and other fields.4,5 Even today, it is difficult to find an academic publication in organic
synthesis that does not appeal to medical or ecological interests for a justification of the
research.
indicates that many are neurologically relevant compounds, such as caffeine, nicotine,
codeine, and morphine (Table I.1).5 It is disappointing, then, that the pool of clinically
this earliest era of natural product chemistry. In the last 25 years, only two neurologically
relevant natural products, and an additional two compounds derived from natural
Figure I.7: All new small chemical entities that were approved for clinical use,
01/1981-06/2006, by source.93
N: Natural products. ND: Derived from a natural product and is usually a semisynthetic modification.
S: Totally synthetic drug, often found by random screening/modification of an existing agent. S*:
Made by total synthesis, but the pharmacophore is/was from a natural product. NM: Natural product
mimic.
34
products, were approved for clinical use.93 In contrast, 21 natural products and 99
compounds derived from natural products were approved for treatments of cancer and
infectious diseases during the same period.93 To respond to this challenge, current efforts
Despite more anti-cancer agents being approved for human treatment than any
other pharmaceutical classes, there is still a great need for new drugs to treat cancer. For
pancreatic cancer.152 Even for the types of cancer that are known to be treatable by anti-
cancer agents, severe side-effects and ineffectiveness towards metastasis still plague
patients. With cancer being the second most common cause of death in the USA (23.1%
of all deaths in 2004),153 the search for new and more effective anti-cancer agents is still
warranted indeed. Though outside the scope of this thesis, new anti-microbial agents are
also urgently needed. In recent decades, infectious diseases have emerged as a substantial
In the past few years, the re-isolation of known compounds has become a
terrestrial plants and microbes have been the major sources of natural products, and
researchers have exhaustively studied those organisms.5 Therefore, the current trend is to
investigate organisms that have not been previously studied in detail. For example,
investigation of poisonous frogs155 and marine organisms91 has enjoyed great productivity
in recent years. Between 1968 and 1998, over 400 alkaloids of over 20 structural classes,
most of which are neurologically relevant, were isolated from amphibians alone.87
Considering the difficulty of obtaining extract samples from amphibian skin, this is a
35
metabolites, a total of 812 new compounds were isolated in 2005 alone (716 for 2004),
With such an outstanding rate of discovery, one would expect to see more natural
products being approved for clinical use. However, this is not the case. One cause for this
high potential are never tested in vivo in animals, which often requires more than
hundreds of milligrams of material. Often times, the major barrier in drug development is
this multi-hundred milligram (MHM) quantity, due to the frequent inability to produce
may show interest in carrying out industrial-scale production of the compound for further
testing (Figure I.8). Otherwise, due to financial difficulties, it is virtually impossible for
Note that “identification of compound” in this figure refers to both lead compound discovery and
medicinal chemistry (analog synthesis and evaluation) processes.
36
If the natural product is from a bacterium and/or fungus, culturing of the source
organism is often possible at a scale large enough for isolation of the compound to meet
the demands of in vivo assays. For most natural products isolated from animals, such as
organizations would generally tend to avoid high risk projects, such as a compound at the
in vivo testing stage, academic, non-profit institutions, and benevolent small businesses
As Paul Wender has articulated, an ideal synthesis is “one in which the target
operation that proceeds quickly and in quantitative yield.”157 Employing these principles
for achieving an efficient synthesis while applying creativity and utilizing the vast
Inspiring news hit the synthetic community in 2004 when Novartis disclosed their
Discodermolide, 17
37
synthetic routes developed by both Smith's and Patterson's laboratories (Scheme I.4),159,160
Novartis pushed the scalability limit of total synthesis. Smith and coworkers devised a
highly convergent and elegant synthesis, in which all three of the units (19, 20, 21)
required for the discodermolide skeleton are derived from a single common precursor
(22) with elaborate stereochemistry.83 However, the preparation for the last coupling in
Smith's synthesis involved a tranformation to a phosphonium salt that required 12.8 kbar
of pressure, which would be too dangerous for an industrial scale synthesis. In the
strategy developed by Patterson and coworkers, the last coupling proceeded with good
38
diastereoselectivity (21:4) and with excellent yield (87%), and it did not involve any
"Some 3,000 kg of the sponge--a quantity that probably does not exist--would
have been needed to deliver 60 g of (+)-discodermolide. The large-scale total
synthesis of such a complex natural product in such quantities was a first for
Novartis and probably the entire pharmaceutical industry,“
Stuart J. Mickel, Novartis161
Another obstacle to the development of drugs from natural products is the lack of
against HUVEC and in vivo anti-angiogenic effect in zebrafish at 30 nM) was only
Somocystinamide, 23
39
products. Yet, the discovery and development of these compounds as drugs will depend
on collaborative efforts of chemists and biologists to identify bioactivities that are not yet
known.
there is usually still work to be done by synthetic chemists. In the example of 23, the
disulfide and enamide are of concerns with regards to human toxicity and compound
stability. These functionalities may need to be replaced in order for the drug to be usable
in clinical situations. Indeed, in Figure I.5, only 6% of the newly approved small
molecule drugs are unmodified natural products, whereas 28% are modified natural
supply issue, access to analogs, and detailed insights of the chemical properties of the
intellectually satisfying problem for chemists to solve and can be useful for engineering
of the biosynthetic pathway and in vitro application of the enzyme in chemical synthesis.
intermediate and then analyzing the product of the enzymatic reaction that utilizes the
intermediate.
Brevianamides and their related compounds are a family of natural products that
are biosynthetically derived from tryptophan, proline, and one or two isoprene unit(s).164
Many metabolites of this family share a core structure that appears to be derived from an
fungi Penicilium sp. and Aspergillus sp., possess a “Diels-Alderase,” has been an exciting
topic in the field of bioorganic chemistry. 78,165 Both brevianamide A (33) and
30, which was prepared through chemical synthesis.166 When a synthetic precursor similar
any of the isolated natural products was the major product and an isomer that is
analogous to 34 was the minor product (isomeric ratio = 2:1).167 This result gives
credence to the hypothesis that the Diels-Alder reaction in Schem I.5 is indeed enzyme-
catalyzed and that 33 and 34 are not artifacts, but natural products.
synthesis of a natural product in a way that mimics the proposed biosynthetic pathway,
one can gain direct chemical insight into the biochemical reactions of interest, as seen in
the example above with brevianamides. Because a biomimetic synthesis relies upon the
or an elaborate intramolecular reaction are often involved. 168 A recent and elegant
isolated from a marine tunicate, Dendrodoa grossularia.169 The dimeric nature of the
molecule is not immediately apparent in the structure (35), but is clearly revealed in the
whose total synthesis was reported simultaneously by Porco and Baldwin in 2003
(Scheme I.7).170,171 Here the inherent propensity of the monomer (42) for dimerization is
reaction to produce the natural product (41). Without this biomimetic step, the core
Presumably, the exact same reaction can happen in nature with or without any assistance
from an enzyme, which brings up the question about whether compounds like 41 are truly
this case, given that 42 requires an extremely high concentration (neat) for smooth
example of a dimeric metabolite that is now known to require enzymatic assistance for
putative monomer 45 was prepared and subjected to various oxidizing conditions that
could exist in nature. However, it was found that 45 does not dimerize under such
Scheme I.8: Scytonemin monomer does not dimerize under naturally occurring conditions
Three separate areas in which natural products chemistry has benefited from
organic synthesis were discussed above; namely, structure elucidation and confirmation,
natural products have been the major driving force for the advancement of organic
synthesis. Given this rich history, direct application of the power of organic synthesis to
natural products should improve productivity in all three areas mentioned above. As an
added significant benefit, new chemical transformations and synthetic strategies can be
ichthyotoxic polybrominated diphenyl ethers from a mixed assemblage of a red alga and
cyanobacteria. Coauthors contributed the following: Z. Cao conducted the [Ca2+]i assays;
T. F. Murray oversaw the [Ca2+]i assay work and contributed to the corresponding portion
44
of the article. The text of II, in whole, is the manuscript that has been submitted to
Toxicon as it will appear: Takashi L. Suyama, Zhengyu Cao, Thomas F. Murray, and
Properties. The dissertation author was the primary investigator and author of the
Chapter III describes the discovery of a novel fatty acid containing a vinylchloride
functionality, credneric acid, from a marine cyanobacterium Lyngbya sp. The text of III,
dissertation author was the primary investigator and author of the research which forms
contributed the following: K. L. McPhail assisted with the purification of the natural
product and acquisition and interpretation of the NMR data. The text of IV, in part, will
Cyanobacterial Lipid. The dissertation author was the primary investigator and author of
a rainforest frog alkaloid epiquinamide and its enantiomer. The text of V, in part, is
45
4541-4543. The dissertation author was the primary investigator and author of
author was the primary investigator and author of the research which forms the
UV-screening metabolite, scytonemin. The text of VII, in part, will be used for a
the primary investigator and author of the research which forms the basis of this
chapter.
The appendix describes my additional research activities that could not constitute
4 Tarbell, D. S.; Tarbell, T. Essays on the history of organic chemistry in the United States,
1875-1955. 1986, K & S Press, Nashville, TN. pp 5.
7 Indeed, the bioactivities of some of these compounds are still being discovered. For example,
some of the health benefits of garlic are now attributed to the diallyl sulfide compounds found
in it (Table I.1 and ref. 8).
8 Xiao, D.; Li, M.; Herman-Antosiewicz, A.; Antosiewicz, J.; Xiao, H.; Lew, K. L.; Zeng, Y.;
Marynowski, S. W.; Singh, S. V. Diallyl Trisulfide Inhibits Angiogenic Features of Human
Umbilical Vein Endothelial Cells by Causing Akt Inactivation and Down-Regulation of VEGF
and VEGF-R2. Nutr. Canc. 2006, 55, 94-107.
9 Fischer, H.; Zeile, K. Synthese des Hämatoporphyrins, Protoporphyrins und Hämins. Justus
Liebigs Ann. Chem. 1929, 468, 98-116.
10 Bachmann, W. E.; Cole, W.; Wilds, A. L. The total synthesis of sex hormone equilenin. J. Am.
Chem. Soc. 1939, 61, 974.
13 Saraswathi, V.; Ramamoorthy, N.; Subramaniam, S.; Mathuram, V.; Gunasekaran, P.;
Govindasamy, S. Inhibition of glycolysis and respiration of sarcoma-180 cells by echitamine
chloride. Chemotherapy 1998, 44, 198-205.
16 Nicolaou, K. C.; Snyder, S. A. Chasing Molecules That Were Never There: Misassigned
Natural Products and the Role of Chemical Synthesis in Modern Structure Elucidation.
Angew. Chem. Int. Ed. 2005, 44, 1012-1044.
47
18 SciFinder Scholar search for “structure revision” and other related keywords
19 Gavagnin, M.; Mollo, E.; Cimino, G.; Ortea, J. A new γ-dihydropyrone-propionate from the
caribbean sacoglossan Tridachia crispata. Tetrahedron Lett. 1996, 37, 4259–4262.
20 (a) Jeffery, D. W.; Perkins, M. V.; White, J. M. Synthesis of an Analogue of the Marine
Polypropionate Tridachiahydropyrone. Org. Lett. 2005, 7, 407–409. (b) Jeffery, D. W.;
Perkins, M. V.; White, J. M. Synthesis of the Putative Structure of Tridachiahydropyrone. Org.
Lett. 2005, 7, 1581–1584.
21 Ireland, C.; Faulkner, D. J. The metabolites of the marine molluscs tridachiella diomedea and
tridachia crispata. Tetrahedron 1981, 37, 233–240, Suppl. 1.
23 Sharma, P.; Griffiths, N.; Moses, J. E. Biomimetic synthesis and structural revision of (±)-
tridachiahydropyrone. Org. Lett. 2008, 10, 4025-4027.
24 Takashima, J.; Asano, S.; Ohsaki, A. Tennen Yuki Kagobutsu Toronkai Koen Yoshishu 2000,
42, 487.
25 Hu, G.; Liu, K.; Williams, L. J. The brosimum allene: a structural revision. Org. Lett. 2008,
10, 5493-5496.
26 Publications that merely claim that the original structures are incorrect, but do not give
revised structures are omitted.
28 (a) Ogamino, T.; Nishiyama, S. Synthesis and structural revision of calafianin, a member of
the spiroisoxazole family isolated from the marine sponge, Aplysina gerardogreeni.
Tetrahedron Lett. 2005, 46, 1083–1086. (b) Ogamino, T.; Obata, R.; Tomoda, H.; Nishiyama,
S. Total Synthesis, Structural Revision, and Biological Evaluation of Calafianin, a Marine
Spiroisoxazoline from the Sponge, Aplysina gerardogreeni. Bull. Chem. Soc. Jpn. 2006, 79,
134-139.
29 Wu, T.; Shi, L.; Kuo, S. Alkaloids and other constituents from Tribulus terrestris.
Phytochemistry 1999, 50, 1411-1415.
30 Bremner, J.; Sengpracha, W.; Southwell, I.; Bourke, C.; Skelton, B.; White, A. The alkaloids
of Tribulus terrestris: a revised structure for the alkaloid tribulusterine. Acta Horticulturae
2005, 677, 11-17.
48
31 Baetas, A. C. S.; Arruda, M. S. P.; Mu¨ller, A. H.; Arruda, A. C. J. Braz. Chem. Soc. 1999, 10,
181–183.
32 Maesa, D.; Vervischa, S.; Debenedettib, S.; Davioc, C.; Mangelinckxa, S.; Giubellinaa, N.; De
Kimpe, N. Synthesis and structural revision of naturally occurring ayapin derivatives.
Tetrahedron 2005, 61, 2505-2511.
33 Suzuki, K.; Yahara, S.; Maehata, K.; Uyeda, M. Isoaurostatin, a Novel Topoisomerase
Inhibitor Produced by Thermomonospora alba. J. Nat. Prod. 2001, 64, 204–207.
35 Atta-Ur-Rahman; Choudhary, M. I.; Hayat, S.; Khan, A. M.; Ahmed, A. Two New Aurones
from Marine Brown Alga Spatoglossum variabile. Chem. Pharm. Bull. 2001, 49, 105–107.
37 Tang, Y.-Q.; Sattler, I.; Thiericke, R.; Grabley, S.; Feng, X.-Z. Feigrisolides A, B, C and D,
New Lactones with Antibacterial Activities from Streptomyces griseus. J. Antibiot. 2000, 53,
934-943.
38 Alvarez-Bercedo, P.; Murga, J.; Carda, M.; Marco, J. A. Stereoselective Synthesis of the
Published Structure of Feigrisolide A. Structural Revision of Feigrisolides A and B. Org.
Chem. 2006, 71, 5766–5769.
39 Tang, Y.-Q.; Sattler, I.; Thiericke, R.; Grabley, S.; Feng, X.-Z. Feigrisolides A, B, C and D,
New Lactones with Antibacterial Activities from Streptomyces griseus. J. Antibiot. 2000, 53,
934-943.
40 Kim, W. H.; Jung, J. H.; Lee, E. Feigrisolide C: Structural Revision and Synthesis. J. Org.
Chem. 2005, 70, 8190-8192.
42 Miyaoka, H.; Hara, Y.; Shinohara, I.; Kurokawa, T.; Yamada, Y. Synthesis and structural
revision of marine eicosanoid agardhilactone. Tetrahedron Lett. 2005, 46, 7945–7949.
43 Snider, B. B.; Shi, B. Synthesis of pyrinodemins A and B. Assignment of the double bond
position of pyrinodemin A. Tetrahedron Lett. 2001, 42, 1639–1642.
44 Ishiyama, H.; Tsuda, M.; Endo, T.; Kobayashi, J. Asymmetric Synthesis of Double Bond
Isomers of the Structure Proposed for Pyrinodemin A and Indication of Its Structural Revision.
Molecules 2005, 10, 312-316.
49
45 Nakamura, H.; Wu, H.; Ohizumi, Y.; Hirata, Y. Agelasine-A, -B, -C and -D, novel bicyclic
diterpenoids with a 9-methyladeninium unit possessing inhibitory effects on na,K-atpase from
the okinawa sea sponge Agelas sp. Tetrahedron Lett. 1984, 25, 2989-2992.
46 Marcos, I. S.; Garcia, N.; Sexmero, M. J.; Basabe, P.; Diez, D.; Urones, J. G. Synthesis of (+)-
agelasine C. A structural revision. Tetrahedron 2005, 61, 11672-11678.
47 Rao, M. R.; Faulkner, D. J. Clavosolides A and B, Dimeric Macrolides from the Philippines
Sponge Myriastra clavosa. J. Nat. Prod. 2002, 65, 386-388.
48 Son, J. B.; Kim, S. N.; Kim, N. Y.; Lee, D. H. Total Synthesis, Structural Revision, and
Absolute Configuration of (+)-Clavosolide A. Org. Lett. 2006, 8, 661-664.
49 Hiort, J.; Maksimenka, K.; Reichert, M.; Perovi-Ottstadt, S.; Lin, W. H.; Wray, V.; Steube, K.;
Schaumann, K.; Weber, H.; Proksch, P.; Ebel, R.; Mller, W. E. G.; Bringmann, G. New
Natural Products from the Sponge-Derived Fungus Aspergillus niger. J. Nat. Prod. 2004, 67,
1532-1543.
50 Ye, Y. H.; Zhu, H. L.; Song, Y. C.; Liu, J. Y.; Tan, R. X. Structural Revision of Aspernigrin
A, Reisolated from Cladosporium herbarum IFB-E002. J. Nat. Prod. 2005, 68, 1106-1108.
52 Zhang, W.; Guo, Y. W.; Gavagnin, M.; Mollo, E.; Cimino, G. Suberoretisteroids A-E, five
new uncommon polyoxygenated steroid 24-ketals from the Hainan gorgonian Suberogorgia
reticulata. Helv. Chim. Acta 2005, 88, 87-94.
53 Shimbo, K.; Tsuda, M.; Izui, N.; Kobayashi, J. Amphidinolide W, a New 12-Membered
Macrolide from Dinoflagellate Amphidinium sp. J. Org. Chem. 2002, 67, 1020.
55 Patil, A. D.; Freyer, A. J.; Taylor, P. B.; Carte´, B.; Zuber, G.; Johnson, R. K.; Faulkner, D. J.
Batzelladines F−I, Novel Alkaloids from the Sponge Batzella sp.: Inducers of p56lck-CD4
Dissociation. J. Org. Chem. 1997, 62, 1814-1819.
56 Cohen, F.; Overman, L. E. Evolution of a Strategy for the Synthesis of Structurally Complex
Batzelladine Alkaloids. Enantioselective Total Synthesis of the Proposed Structure of
Batzelladine F and Structural Revision. J. Am. Chem. Soc. 2006, 128, 2594-2603.
57 Bourdelais, A. J.; Jacocks, H. M.; Wright, J. L. C.; Bigwarfe, P. M., Jr.; Baden, D. G. A New
Polyether Ladder Compound Produced by the Dinoflagellate Karenia brevis. J. Nat. Prod.
2005, 68, 2-6.
50
58 Fuwa, H.; Ebine, M.; Bourdelais, A. J.; Baden, D. G.; Sasaki, M. Total Synthesis, Structure
Revision, and Absolute Configuration of (-)-Brevenal. J. Am. Chem. Soc. 2006, 128,
16989-16999.
59 Ahmed, A. A.; Mahmoud, A. A. Jasonol, a rare tricyclic eudesmane sesquiterpene and six
other new sesquiterpenoids from Jasonia candicans. Tetrahedron 1998, 54, 8141-52.
60 Fang, L.; Bi, F.; Zhang, C.; Zheng, G.; Li, Y. Total synthesis of 4a,5a,10b-trihydroxycadinane
and its C4-isomer: structural revision of a natural sesquiterpenoid. Synlett 2006, 16,
2655-2657.
61 Lin, W.; Brauers, G.; Ebel, R.; Wray, V.; Berg, A.; Sudarsono; Proksch, P. Novel Chromone
Derivatives from the Fungus Aspergillus versicolor Isolated from the Marine Sponge
Xestospongia exigua. J. Nat. Prod. 2003, 66, 57–61.
62 Saito, F.; Kuramochi, K.; Nakazaki, A.; Mizushina, Y.; Sugawara, F.; Kobayashi, S. Synthesis
and Absolute Configuration of (+)-Pseudodeflectusin: Structural Revision of Aspergione B.
Eur. J. Org. Chem. 2006 21, 4796–4799.
63 Cutler, H. G.; Jacyno J. M.; Harwood J. S.; Dulik, D.; Goodrich, P. D.; Roberts, R. G.
Botcinolide: A Biologically Active Natural Product from Botrytis cinerea. Biosci. Biotech.
Biochem. 1993, 57, 1980-1982.
64 Tani, H.; Koshino, H.; Sakuno, E.; Cutler, H. G.; Nakajima, H. Botcinins E and F and
Botcinolide from Botrytis cinerea and structural revision of botcinolides. J. Nat. Prod. 2006,
69, 722–725.
65 Cutler, H. G.; Parker, S. R.; Ross, S. A.; Crumley, F. G.; Schreiner, P. R. A Biologically Active
Natural Homolog of Botcinolide from Botrytis cinerea. Biosci. Biotech. Biochem. 1996, 60,
656-658.
66 Collado, I. G.; Aleu, J.; Herna´ndez-Gala´n, R.; Hanson, J. R. Some metabolites of Botrytis
cinerea related to botcinolide. Phytochemistry 1996, 42, 1621-1624.
67 Yamada, T.; Iritani, M.; Minomura, K.; Kawai, K.; Numata, A. Peribysins A–D, potent cell-
adhesion inhibitors from a sea hare-derived culture of Periconia species. Org. Biomol. Chem.
2004, 2, 2131–2135.
68 Koshino, H.; Satoh, H.; Yamada, T.; Esumi, Y. Structural revision of peribysins C and D.
Tetrahedron Lett. 2006, 47, 4623-26.
69 Hiort, J.; Maksimenka, K.; Reichert, M.; Perovic-Ottstadt, S.; Lin, W. H.; Wray, V.; Steube,
K.; Schaumann, K.; Weber, H.; Proksch, P.; Ebel, R.; Mu¨ller, W. E. G.; Bringmann, G. New
Natural Products from the Sponge-Derived Fungus Aspergillus niger. J. Nat. Prod. 2004, 67,
1532-1543.
51
70 Schlingmann, G.; Taniguchi, T.; He, H.; Bigelis, R.; Yang, H. Y.; Koehn, F. E.; Carter, G. T.;
Berova, N. Reassessing the structure of pyranonigrin. J. Nat. Prod. 2007, 70, 1180-1187.
71 (a) Ravikumar, P. R.; Soman, R.; Chetty, G. L.; Pandey, R. C.; Sukh, D. Chemistry of
Ayurvedic Crude Drugs: Part VI-(Shatavari-1): Structure of Shatavarin-IV. Ind. J. Chem. 1987,
26B, 1012–1017. (b) Joshi, J.; Sukh, D. Chemistry of Ayurvedic Crude Drugs: Part III-
Shatavari-2: Structure Elucidation of Bioactive Shatavarin-I & Other Glycosides. Ind. J.
Chem. 1988, 27B, 12–16.
72 Hayesa, P. Y.; Jahidina, A. H.; Lehmannb, R.; Penmanb, K.; Kitchinga, W.; De Voss, J. J.
Structural revision of shatavarins I and IV, the major components from the roots of Asparagus
racemosus. Tetrahedron Lett. 2006, 47, 6965-6969.
73 Evidente, A.; Conti, L.; Altomare, C.; Bottalico, A.; Sindona, G.; Segre, A. L.; Logrieco, A.
Fusapyrone and deoxyfusapyrone, two antifungal a -pyrones from Fusarium semitectum. Nat.
Toxins 1994, 2, 4–13
74 Hiramatsu, F.; Miyajima, T.; Murayama, T.; Takahashi, K.; Koseki, T.; Shiono, Y. Isolation
and Structure Elucidation of Neofusapyrone from a Marine-derived Fusarium species, and
Structural Revision of Fusapyrone and Deoxyfusapyrone. The Journal of Antibiotics 2006 59,
704–709.
75 Seo, Y.; Cho, K. W.; Rho, J.-R.; Shin, J.; Kwon, B.-M.; Bok, S.-H.; Song, J.-I.
Solandelactones A-I, Lactonized Cyclopropyl Oxylipins Isolated from the Hydroid Solanderia
secunda. Tetrahedron 1996, 52, 10583-96.
77 Delazar, A.; Byres, M.; Gibbons, S.; Kumarasamy, Y.; Modarresi, M.; Nahar, L.; Shoeb, M.;
Sarker, S. D. Iridoid Glycosides from Eremostachys glabra. J. Nat. Prod. 2004, 67,
1584-1587.
78 Jensen, S. R.; Çalış, I.; Gotfredsen, C. H.; Søtofte, I. Structural Revision of Some Recently
Published Iridoid Glucosides. J. Nat. Prod. 2007, 70, 29–32.
80 Ahmad, I.; Afza, N.; Anis, I.; Malik, A.; Fatima, I.; ul-Haq, A.; Tareen, R. B. Iridoid
glycosides and a benzofuran type sesquiterpene from Buddleja crispa. Heterocycles 2004, 63,
1875-1881.
81 Itokawa, H.; Nakajima, H.; Ikuta, A.; Iitaka, Y. Two triterpenes from the flowers of Camellia
japonica. Phytochemistry. 1981, 20, 2539-2542.
52
82 Yoshikawa, M.; Morikawa, T.; Asao, Y.; Fujiwara, E.; Nakamura, S.; Matsuda, H. Medicinal
flowers. XV. The structures of noroleanane- and oleanane-type triterpene oligoglycosides with
gastroprotective and platelet aggregation activities from flower buds of Camellia japonica.
Chem. Pharmaceut. Bull. 2007, 55, 606-612.
84 (a) Koeck, M.; Grube, A.; Seiple, I. B.; Baran, P. S. The pursuit of Palau'amine. Angew.
Chem. Int. Ed. 2007, 46, 6586-6594. (b) Grube, A.; Köck, M. Structural Assignment of
Tetrabromostyloguanidine: Does the Relative Configuration of the Palau'amines Need
Revision? Angew. Chem. Int. Ed. 2007, 46, 2320 – 2324.
85 Wall, M. E.; Wani, M. C.; Manikumar, H. T.; Taylor, H.; McGivney, R. Plant Antimutagens, 6.
Intricatin and Intricatinol, New Antimutagenic Homoisoflavonoids from Hoffmanosseggia
intricata. J. Nat. Prod. 1989, 52, 774-778.
87 Siddaiah, V.; Maheswara, M.; Rao, C. V.; Venkateswarlub, S.; Subbarajub, G. V. Synthesis,
structural revision, and antioxidant activities of antimutagenic homoisoflavonoids from
Hoffmanosseggia intricata. Bioorg. Med. Chem. Lett. 2007, 17, 1288–1290.
88 (a) Leet, J. E.; Schroeder, D. R.; Hofstead, S. J.; Golik, J.; Colson, K. L.; Huang, S.; Klohr, S.
E.; Doyle, T. W.; Matson, J. A. Kedarcidin, a new chromoprotein antitumor antibiotic:
structure elucidation of kedarcidin chromophore. J. Am. Chem. Soc. 1992, 114, 7946-7948. (b)
Leet, J. E.; Schroeder, D. R.; Langley, D. R.; Colson, K. L.; Huang, S.; Klohr, S. E.; Lee, M.
S.; Golik, J.; Hofstead, S.J.; Doyle, T. W.; Matson, J. A. Chemistry and structure elucidation of
the kedarcidin chromophore. J. Am. Chem. Soc. 1993, 115, 8432-8443.
89 Ren, F.; Hogan, P. C.; Anderson, A. J.; Myers, A. G. Kedarcidin chromophore: Synthesis of
Its Proposed Structure and Evidence for a Stereochemical Revision. J. Am. Chem. Soc. 2007,
129, 5381–5383.
90 Kawata, S.; Ashizawa, S.; Hirama, M. Synthetic Study of Kedarcidin Chromophore: Revised
Structure. J. Am. Chem. Soc. 1997, 119, 12012–12013.
91 McPhail, K. L.; Gerwick, W. H. Three New Malyngamides from a Papua New Guinea
Collection of the Marine Cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2003, 66, 132.
92 Li, Y.; Feng, J.-P.; Wang, W.-H.; Chen, J.; Cao, X.-P. Total Synthesis and Correct Absolute
Configuration of Malyngamide U. J. Org. Chem. 2007, 72, 2344-2350.
95 Tanaka, N.; Okasaka, M.; Ishimaru, Y.; Takaishi, Y.; Sato, M.; Okamoto, M.; Oshikawa, T.;
Ahmed, S. U.; Consentino, L. M.; Lee, K.-H. Biyouyanagin A, an Anti-HIV Agent from
Hypericum chinense L. var. salicifolium. Org. Lett. 2005, 7, 2997-2999.
96 Nicolaou, K. C.; Sarlah, D.; Shaw, D. M. Total synthesis and revised structure of
biyouyanagin A. Angew. Chem. Int. Ed. 2007, 46, 4708-4711.
97 Nii, H.; Furukawa, K.; Iwakiri, M.; Kubota, T. Nippon Nogeikagaku Kaishi 1983, 57, 733–
741; Chem. Abstr. 1983, 99, 200329.
98 Blay, G.; Garcia, B.; Molina, E.; Pedro, J. R. Synthesis of (+)-pechueloic acid and (+)-
aciphyllene. Revision of the structure of (+)-aciphyllene. Tetrahedron 2007, 63, 9621-9626.
100(a) Nicolaou, K. C.; Guduru, R.; Sun, Y.-P.; Banerji, B.; Chen, D. Y.-K. Total synthesis of the
originally proposed and revised structures of palmerolide A. Angew. Chem. Int. Ed. 2007, 46,
5896-5900. (b) Nicolaou, K. C.; Sun, Y.-P.; Guduru, R.; Banerji, B.; Chen, D. Y.-K. Total
Synthesis of the Originally Proposed and Revised Structures of Palmerolide A and Isomers
Thereof. J. Am. Chem. Soc. 2008, 130, 3633-3644.
101 Solis, P. N.; Lang’at, C.; Gupta, M. P.; Kirby, G. C.; Warhurst, D. C.; Phillipson, J. D. Bio-
active Compounds from Psychotria camponutans. Planta Med. 1995, 61, 62-65.
102 Jacobs, J.; Claessens, S.; De Kimpe, N. First straightforward synthesis of 1-hydroxy-3,4-
dihydro-1H-benz[g]isochromene-5,10-dione and structure revision of a bioactive
benz[g]isochromene-5,10-dione from Psychotria camponutans. Tetrahedron 2008, 64,
412-418.
103 Sun, J.; Shi, D.; Ma, M.; Li, S.; Wang, S.; Han, L.; Yang, Y.; Fan, X.; Shi, J.; He, L.
Sesquiterpenes from the Red Alga Laurencia tristicha. J. Nat. Prod. 2005, 68, 915-919.
104 Chen, P.; Wang, J.; Liu, K.; Li, C. Synthesis and Structural Revision of (±)-Laurentristich-4-
ol. J. Org. Chem. 2008, 73, 339-341.
106 Harkat, H.; Blanc, A.; Weibel, J.-M.; Pale, P. Versatile and Expeditious Synthesis of Aurones
via AuI-Catalyzed Cyclization. J. Org. Chem. 2008, 73, 1620-1623.
54
107 Gunatilaka, A. A. L. In: Herz, W.;Kirby, G. W.; Moore, R. E.; Steglich, W.; Tamm,C.
Progress in the Chemistry of Organic Natural Products vol.67, Springer-Verlag/Wien, New
York 1996, pp. 1–123.
108 Jacobsen, N. E.; Wijeratne, E. M. K.; Corsino, J.; Furlan, M.; Bolzani, V. S.; Gunatilaka, A.
A. L. Biomimetic synthesis of xuxuarines Ea and Eb: Structure revision of Rzedowskia
bistriterpenoids. Bioorg. Med. Chem. 2008, 16, 1884-1889.
109 Bastida, J.; Codina, C.; Peeters, P.; Rubiralta, M.; Orozco, M.; Luque, F. J.; Chharbra, S. C.
Alkaloids from Crinum kirkii. Phytochemistry 1995, 40, 1291-1293.
110 Biechy, A.; Hachisu, S.; Quiclet-Sire, B.; Ricard, L.; Zard, S. Z. The total synthesis of (±)-
fortucine and a revision of the structure of kirkine. Angew. Chem. Int. Ed. 2008, 47,
1436-1438.
111 Sakai, K.; Fukuda, Y.; Matsunaga, S.; Tanaka, R.; Yamort, T. New Cytotoxic Oleanane-Type
Triterpenoids from the Cones of Liquidamber styraciflua. J. Nat. Prod. 2004, 67, 1088-1093.
112 Sun, H.; Fang, W.-S.; Hu, C. An efficient semi-synthesis and structure revision of a cytotoxic
triterpenoid 25-acetoxy-3a-hydroxyolean-12-en-28-oic acid from Liquidamber styraciflua. J.
Asian Nat. Prod. Res. 2008, 10, 271-276.
113 Zampella, A.; D'Auria, M.V.; Minale, L.; Debitus, C.; Roussakis, C. Callipeltoside A: A
Cytotoxic Aminodeoxy Sugar-Containing Macrolide of a New Type from the Marine
Lithistida Sponge Callipelta sp. J. Am. Chem. Soc. 1996, 118, 11085-11088.
114 Carpenter, J.; Northrup, A. B.; Chung, D.; Wiener, J. J. M.; Kim, S.-G.; MacMillan, D. W. C.
Total synthesis and structural revision of callipeltoside C. Angew. Chem. Int. Ed. 2008, 47,
3568-3572.
115 Kuo, Y.-H.; Lin, B.-Y. A New Dinorxanthane and Chromone from the Root of Tithonia
diversifolia. Chem. Pharm. Bull. 1999, 47, 428-429.
116 Matsuo, K.; Yokoe, H.; Shishido, K.; Shindo, M. Synthesis of diversifolide and structure
revision. Tetrahedron Lett. 2008, 49, 4279-4281.
117 Akiyama, K.; Kikuzaki, H.; Aoki, T.; Okuda, A.; Lajis, N. H.; Nakatani, N. Terpenoids and a
Diarylheptanoid from Zingiber ottensii. J. Nat. Prod. 2006, 69, 1637-1640.
118 Boukouvalas, J.; Wang, J.-X. Structure Revision and Synthesis of a Novel Labdane
Diterpenoid from Zingiber ottensii. Org. Lett. 2008, 10, 3397-3399.
119 Sorek, H.; Rudi, A.; Gueta, S.; Reyes, F.; Martin, M. J.; Aknin, M.; Gaydou, E.; Vacelet, J.;
Kashman, Y. Netamines A–G: seven new tricyclic guanidine alkaloids from the marine sponge
Biemna laboutei. Tetrahedron 2006, 62, 8838–8843.
55
120 Yu, M.; Pochapsky, S. S.; Snider, B. B. Synthesis of 7-Epineoptilocaulin, Mirabilin B, and
Isoptilocaulin. A Unified Biosynthetic Proposal for the Ptilocaulin and Batzelladine Alkaloids.
Synthesis and Structure Revision of Netamines E and G. J. Org. Chem. 2008, 73, 9065-9074.
121 Hayakawa, Y.; Yamashita, T.; Mori, T.; Nagai, K.; Shin-ya, K.; Watanabe, H. J. Structure of
Tyroscherin, an Antitumor Antibiotic against IGF-1-dependent Cells from Pseudallescheria
sp. Antibiot. 2004, 57, 634–638.
122 Katsuta, R.; Shibata, C.; Ishigami, K.; Watanabe, H.; Kitahara, T. Synthesis and structure
revision of tyroscherin, a growth inhibitor of IGF-1-dependent tumor cells. Tetrahedron Lett.
2008, 49, 7042-7045.
123 Murata, M.; Matsuoka, S.; Matsumori, N.; Paul, G. K.; Tachibana, K. Absolute Configuration
of Amphidinol 3, the First Complete Structure Determination from Amphidinol Homologues:
Application of a New Configuration Analysis Based on Carbon−Hydrogen Spin-Coupling
Constants. J. Am. Chem. Soc. 1999, 121, 870–871.
124 Oishi, T.; Kanemoto, M.; Swasono, R.; Matsumori, N.; Murata, M. Synthesis of the 1,5-
Polyol System Based on Cross Metathesis: Structure Revision of Amphidinol 3. Org. Lett.
2008, 10, 5203-5206.
125 Mueller, D.; Krick, A.; Kehraus, S.; Mehner, C.; Hart, M.; Kuepper, F. C.; Saxena, K.; Prinz,
H.; Schwalbe, H.; Janning, P.; Waldmann, H.; Koenig, G. M. Brunsvicamides A-C: sponge-
related cyanobacterial peptides with Mycobacterium tuberculosis protein tyrosine phosphatase
inhibitory activity. J. Med. Chem. 2006, 49, 4871-4878.
126 Walther, T.; Arndt, H.-D.; Waldmann, H. Solid-support based total synthesis and
stereochemical correction of brunsvicamide A. Org. Lett. 2008, 10, 3199-3202.
127 Ralifo, P.; Crews, P. J. A New Structural Theme in the Imidazole-Containing Alkaloids from
a Calcareous Leucetta Sponge. Org. Chem. 2004, 69, 9025–9029.
128 White, K. N.; Amagata, T.; Oliver, A. G.; Tenney, K.; Wenzel, P. J.; Crew, P. Structure
Revision of Spiroleucettadine, a Sponge Alkaloid with a Bicyclic Core Meager in H-Atoms J.
Org. Chem. 2008, 73, 8719–8722.
129 Guella, G.; Mancini, I.; Öztunç, A.; Pietra, F. Conformational Bias in Macrocyclic Ethers and
Observation of High Solvolytic Reactivity at a Masked Furfuryl (=2-Furylmethyl) C-Atom.
Helv. Chim. Acta 2000, 83, 336–348.
130 Braddock, D. C.; Rzepa, H. S. Structural Reassignment of Obtusallenes V, VI, and VII by
GIAO-Based Density Functional Prediction. J. Nat. Prod. 2008, 71, 728-730.
131 Rahbæk, L.; Breinholt, J.; Frisvad, J. C.; Christophersen, C. Circumdatin A, B, and C: Three
New Benzodiazepine Alkaloids Isolated from a Culture of the Fungus Aspergillus ochraceus.
J. Org. Chem. 1999, 64, 1689-1692.
56
132 Ookura, R.; Kito, K.; Ooi, T.; Namikoshi, M.; Kusumi, T. Structure Revision of Circumdatins
A and B, Benzodiazepine Alkaloids Produced by Marine Fungus Aspergillus ostianus, by X-
ray Crystallography. J. Org. Chem. 2008, 73, 4245-4247.
133 Cafieri, F.; De Napoli, L.; Fattorusso, E.; Santacroce, C. Diterpenes from the red alga
Sphaerococcus coronopifolius. Phytochemistry 1987, 26, 471–473.
134 Smyrniotopoulos, V.; Quesada, A.; Vagias, C.; Moreau, D.; Roussakis, C.; Roussis, V.
Cytotoxic bromoditerpenes from the red alga Sphaerococcus coronopifolius. Tetrahedron
2008, 64, 5184-5190.
135 Takahashi, T.; Eto, H.; Ichimura, T.; Murae, T. Hiyodorilactones A and B, new tumor
inhibitory germacradienolides from Eupatorium sachalinense makino. Chem. Lett. 1978,
1345-1348.
136 Tori, M.; Morishita, N.; Hirota, N.; Saito, Y.; Nakashima, K.; Sono, M.; Tanaka, M.;
Utagawa, A.; Hirota, H. Sesquiterpenoids isolated from Eupatorium glehnii. Isolation of
guaiaglehnin A, structure revision of hiyodorilactone B, and genetic comparison. Chem.
Pharmaceut. Bull. 2008, 56, 677-681.
137 Sasse, F.; Steinmetz, H.; Höfle, G.; Reichenbach, H. Rhizopodin, a new compound from
Myxococcus stipitatus (myxobacteria) causes formation of rhizopodia-like structures in animal
cell cultures. J. Antibiot. 1993, 46, 741–748.
138 Jansen, R.; Steinmetz, H.; Sasse, F.; Schubert, W.-D.; Hagelueken, G.; Albrecht, S. C.;
Mueller, R. Isolation and structure revision of the actin-binding macrolide rhizopodin from
Myxococcus stipitatus (Myxobacteria). Tetrahedron Lett. 2008, 49, 5796-5799.
139 Ziemert, N.; Ishida, K.; Quillardet, P.; Bouchier, C.; Hertweck, C.; Tandeau de Marsac, N.;
Dittmann, E. Microcyclamide Biosynthesis in Two Strains of Microcystis aeruginosa: from
Structure to Genes and Vice Versa. Appl. EnViron. Microbiol. 2008, 74, 1791-1797.
140 Portmann, C.; Blom, J. F.; Kaiser, M.; Brun, R.; Juttner, F.; Gademann, K. Isolation of
aerucyclamides C and D and structure revision of microcyclamide 7806A: Heterocyclic
ribosomal peptides from Microcystis aeruginosa PCC 7806 and their antiparasite evaluation.
J. Nat. Prod. 2008, 71, 1891-1896.
141 Luis, J. G.; Lahlou, E. H.; Andre´s, L. S. Hassananes: C23 terpenoids with a new type of
skeleton from Salvia apiana Jeps. Tetrahedron 1996, 52, 12309-12312.
142 Yang, J.; Huang, S.-X.; Zhao, Q.-S. Structure Revision of Hassanane with Use of Quantum
Mechanical 13C NMR Chemical Shifts and UV-Vis Absorption Spectra. J. Phys. Chem. 2008,
112, 12132-12139.
143 Takada, N.; Watanabe, R.; Suenaga, K.; Yamada, K.; Ueda, K.; Kita, M. Uemura,
Zamamistatin, a significant antibacterial bromotyrosine derivative, from the Okinawan sponge
Pseudoceratina purpurea. Tetrahedron Lett. 2001, 42, 5265-5267.
57
144 Hayakawa, I.; Teruya, T.; Kigoshi, H. Revised structure of zamamistatin. Tetrahedron Lett.
2006, 47, 155-158.
145 Kita, M.; Tsunematsu, Y.; Hayakawa, I.; Kigoshi, H. Structure of zamamistatin - a correction.
Tetrahedron Lett. 2008, 49, 5383-5384.
146 Fitch, R. W.; Garraffo, H. M.; Spande, T. F.; Yeh, H. J. C.; Daly, J. W. Bioassay-guided
isolation of epiquinamide, a novel quinolizidine alkaloid and nicotinic agonist from an
Ecuadoran poison frog, Epipedobates tricolor. J. Nat. Prod. 2003, 66, 1345-1350.
147 (a) Wijdeven, M. A.; Botman, P. N. M.; Wijtmans, R.; Schoemaker, H. E.; Rutjes, F. P. J. T.;
Blaauw, R. H. Total Synthesis of (+)- Epiquinamide. Org. Lett. 2005, 7, 4005-4007. (b)
Suyama, T. L.; Gerwick, W. H. Practical Total Syntheses of Epiquinamide Enantiomers. Org.
Lett. 2006, 8, 4541-4543. (c) Wijdeven, M. A.; Wijtmans, R.; van den Berg, R. J. F.; Noorduin,
W.; Schoemaker, H. E.; Sonke, T.; van Delft, F. L.; Blaauw, R. H.; Fitch, R. W.; Spande, T. F.;
Daly, J. W.; Rutjes, F. P. J. T. N,N-Acetals as N-Acyliminium Ion Precursors: Synthesis and
Absolute Stereochemistry of Epiquinamide. Org. Lett. 2008, 10, 4001-4003. (d) Ghosh, S.;
Shashidhar, J. Total synthesis of (+)- epiquinamide from -mannitol. Tetrahedron Lett. 2009,
50, 1177-1179. (e) Huang, P.-Q.; Guo, Z.-Q.; Ruan, Y.-P. A Versatile Approach for the
Asymmetric Syntheses of (1R,9aR)- Epiquinamide and (1R,9aR)-Homopumiliotoxin 223G.
Org. Lett. 2006, 8, 1435-1438. (f) Voituriez, A.; Ferreira, F.; Perez-Luna, A.; Chemla, F.
Asymmetric synthesis of (-)-1-hydroxyquinolizidinone, a common intermediate for the
syntheses of (-)-homopumiliotoxin 223G and (-)- epiquinamide. Org. Lett. 2007, 9,
4705-4708. (g) Tong, S.; Barker, D. A concise synthesis of (±) and a total synthesis of (+)-
epiquinamide. Tetrahedron Lett. 2006, 47, 5017-5020. (h) Kanakubo, A.; Gray, D.; Innocent,
N.; Wonnacott, S.; Gallagher, T. The synthesis and nicotinic binding activity of (± )-
epiquinamide and (± )-C(1)-epiepiquinamide. Bioorg. Med. Chem. Lett. 2006, 16, 4648-4651.
148 Fitch, R. W.; Sturgeon, G. D.; Patel, S. R.; Spande, T. F.; Garraffo, H. M.; Daly, J. W.;
Blaauw, R. H. Epiquinamide : A Poison That Wasn't from a Frog That Was. J. Nat. Prod. 2009,
72, 243-247.
149 Newman, D. J.; Cragg, G. M. Natural products as sources of new drugs over the last 25
years. J. Nat. Prod. 2007, 70, 461-477.
150 Füllbeck, M.; Michalsky, E.; Dunkel, M.; Preissner, R. Natural products: sources and
databases. Nat. Prod. Rep. 2006, 23, 347-356.
151 Hill, R. A. Annu. Rep. Prog. Chem. Sect. B, 2003, 99, 183-207.
152 Ghaneh, P.; Costello, E. J P Neoptolemos Biology and management of pancreatic cancer. J.
Postgrad. Med. 2008, 84, 478-497.
153 Heron, M. Deaths: Leading Causes for 2004 National Vital Statistics Reports 2007, 56, 1-96.
154 Sarker, S. D.; Latif, Z.; Gray, A. I. Natural Products Isolation, 2006, Humana Press Inc.
58
155 Daly, J. W.; Spande, T. F.; Garraffo, H. M. Alkaloids from Amphibian Skin: A Tabulation of
Over Eight-Hundred Compounds. J. Nat. Prod. 2005, 68, 1556-1575.
157 Wender. P. A. Introduction: Frontiers in Organic Synthesis. Chem. Rev. 1996, 96, 1-2.
158 Mickel, S. J.; Sedelmeier, G. H.; Niederer, D.; Daeffler, R.; Osmani, A.; Schreiner, K.;
Seeger-Weibel, M.; Bérod, B.; Schaer, K.; Chen, R. G.; Chen, W.; Jagoe, C. T.; Kinder, F. R.,
Jr.; Loo, M.; Prasad, K.; Repi, O.; Shieh, W.; Wang, R.; Waykole, L.; Xu, D. D.; Xue, S.
Large-Scale Synthesis of the Anti-Cancer Marine Natural Product (+)-Discodermolide. Part 1:
Synthetic Strategy and Preparation of a Common Precursor. Org. Process Res. Dev. 2004, 8,
92-100.
159 Smith, A.B.; Beauchamp, T. J.; LaMarche, M. J.; Kaufman, M. D.; Qiu, Y.; Arimoto, H.;
Jones, D. R.; Kobayashi, K. Evolution of a Gram-Scale Synthesis of (+)-Discodermolide. J.
Am. Chem. Soc. 2000, 122, 8654-8664.
160 Patterson, I.; Florence, G. J.; Gerlach, K.; Scott, J. P.; Sereinig, N. A Practical Synthesis of
(+)-Discodermolide and Analogues: Fragment Union by Complex Aldol Reactions. J. Am.
Chem. Soc. 2001, 123, 9535-9544.
161 Freemantle, M. Scaled-up synthesis of discodermolide. Science & Technology 2004, 82,
33-35.
162 Nogle, L. M.; Gerwick, W. H. Somocystinamide A, a Novel Cytotoxic Disulfide Dimer from
a Fijian Marine Cyanobacterial Mixed Assemblage. Org. Lett. 2002, 4, 1095-1098.
163 Wrasidlo, W.; Mielgo, A.; Torres, V. A.; Barbero, S.; Stoletov, K.; Suyama, T. L.; Klemke, R.
L.; Gerwick, W. H.; Carson, D. A.; Stupack, D. G. The marine lipopeptide somocystinamide A
triggers apoptosis via caspase 8. Proc. Natl. Acad. Sci. 2008, 105, 2313-2318.
165 Williams, R. M. Total synthesis and biosynthesis of the paraherquamides: an intriguing story
of the biological Diels-Alder construction. Chem. Pharm. Bull. 2002, 50, 711-740.
168 For example, see Paul G. Bulger,a Sharan K. Bagalb and Rodolfo Marquez, Recent advances
in biomimetic natural product synthesis, Nat. Prod. Rep., 2008, 25, 254–297 .
59
169 Miyake, F. Y.; Yakushijin, K.; Horne, D. A. Biomimetic synthesis of grossularines-1. Angew.
Chem. Int. Ed. 2005, 44, 3280 –3282 .
170 Lei, X.; Johnson, R. P.; Porco Jr., J. A. Total Synthesis of the Ubiquitin-Activating Enzyme
Inhibitor (+)-Panepophenanthrin. Angew. Chem. Int. Ed. 2003, 42, 3913-3917.
171 Moses, J. E.; Commeiras, L.; Baldwin, J. E.; Adlington, R. M. Total Synthesis of
Panepophenanthrin. Org. Lett. 2003, 5, 2987–2988.
172 Proteau, P. J.; Gerwick, W. H.; Garcia-Pichel, F.; Castenholz, R. The structure of scytonemin,
an ultraviolet sunscreen pigment from the sheaths of cyanobacteria . Experientia 1993, 49,
825-829.
Ichthyotoxic brominated diphenyl ethers from a mixed assemblage of a red alga and
60
61
Abstract
Primary fractions from the extract of a tropical red alga mixed with filamentous
cyanobacteria, collected from Papua New Guinea, were active in a neurotoxicity assay.
Bioassay guided isolation led to two natural products (1, 2) with relatively potent calcium
ion influx properties. The more prevalent of the neurotoxic compounds (1) was
literature, but reported with different NMR properties. To clarify this anomalous result,
we synthesized a candidate isomeric polybrominated diphenyl ether (3), but this clearly
had different NMR shifts than the reported compound. We conclude that the original
compound, giving rise to the observed anomalous NMR shifts. The second and more
minor natural product (2) isolated in this study was a more highly brominated species.
All three compounds showed a low micromolar ability to increase intracellular calcium
and accumulate in marine organisms at higher trophic level (mammals, fish, birds), these
II.1 Introduction
trophic level animals, including sperm whales,1,2 sea gulls,3 seals,2 polar bears,4 as well as
in humans.5,6 Some PBDEs are industrially produced in large quantities for use as flame
retardants, and therefore their presence in animals has generally been assumed to be of
found in many of these same higher animals,3,4,8 and recent studies have shown that the
OH-PBDEs found in large marine-associated animals may be of mixed origins with some
deriving from natural sources and the others being derivatives of anthropogenic PBDEs.
concerns have been raised over the occurrence of PBDEs in higher animals because
chlorinated aromatic compounds have been observed with PBDEs, such as aryl
hydrocarbon-receptor agonist and antagonist activities, thyroid toxicity, and effects on the
immune system.15 Neonatal exposure to PBDEs has been shown to cause neurotoxicity in
adult animals.14 OH-PBDEs have also been shown to disrupt thyroid hormone
thyroid hormones.4 Recently, BDE-47 and 6-OH-BDE-47 have been shown to trigger an
neuronal cells within a few minutes.16,17 These findings indicate that PBDEs and OH-
PBDEs have the potential to acutely disrupt normal neuronal communication in animals.
Our program has been systematically screening marine algae and cyanobacteria
environmental impacts these substances may impose, and partially to detect agents with
trigger an increase in cytosolic Ca2+ levels in neuronal cells, we have employed a FLIPR-
based fluorescence assay which detects Ca2+ modulation in mouse neocortical neurons.
mixed assemblage of marine cyanobacteria and a red alga. In the course of this work, we
have also corrected the assignment of a 13C NMR signal for 1 in the literature which
All solvents used were of HPLC grade from EMD and were used without
purification. TLC grade silica gel from Sigma-Aldrich was used for vacuum liquid
chromatography (VLC). Flash chromatography was performed using EMD silica gel
horse serum and soybean trypsin inhibitor were obtained from Atlanta Biologicals.
fluorescent dye Fluo-3, and pluronic acid F-127 were obtained from Invitrogen
Corporation.
Algal material was collected at Grabo reef in Papua New Guinea at 12-18 m depth
by SCUBA. The algae was identified in the field as Vidalia sp., but subsequent
A total of 2 L of the red algal and cyanobacterial assemblage (dry weight 128 g)
was extracted with CH2Cl2 / MeOH (2:1). Upon removal of the solvents in vacuo, a
portion (4.0 g) of the crude extract (4.6 g) was fractionated into nine fractions using silica
gel vacuum liquid chromatography (hexanes to EtOAc to MeOH), one of which (308 mg,
eluted with 2:3 EtOAc/hexanes) possessed Ca2+ modulating activity in mouse neocortical
neurons. A portion (306 mg) of the active fraction was further fractionated by flash
column chromatography22 on silica gel (hexanes/Et2O 1:9 to 1:8) to obtain seven sub-
65
fractions, of which two (2 mg and 77 mg) were associated with Ca 2+ modulation activity.
The larger sub-fraction was further purified by HPLC using a Jupiter 10μ C18 300A
column (250 x 10 mm) with a gradient solvent system (2.5 mL/min, 3:2 MeCN/H 2O to
4:1 over 30 min, then to MeCN over 10 min) to obtain 15 mg (0.38% of extract) of
compound 1 (Rt = 52 min) and ~0.5 mg (0.01% of extract) of compound 2 (Rt = 55 min),
Swiss-Webster mice. Briefly, pregnant mice were euthanized by CO2 asphyxiation and
their neocortices were collected. Isolated neocortices were then removed of their
meninges, minced by trituration using a Pasteur pipette, and treated with trypsin for 25
minutes at 37ºC. The cells were then dissociated by two successive trituration and
sedimentation steps in soybean trypsin inhibitor and DNase containing isolation buffer,
with 2 mM L-glutamine, 10% fetal bovine serum, 10% horse serum, 100 IU/mL
penicillin and 0.10 mg/mL streptomycin, pH 7.4. Cells were plated onto poly-L-lysine-
coated 96-well clear-bottomed black-well plate at a density of 1.5 x 105 cells/well. Plates
were incubated at 37ºC with 5% CO2 and 95% humidity. To prevent proliferation of
nonneuronal cells, cytosine arabinoside (10 μM) was added to the culture medium on day
2 after plating. The culture media was changed both on days 5, 8 and 11 using a serum-
free growth medium containing Neurobasal Medium supplemented with B-27, 100 I.U./
were used between 11-13 days in vitro (DIV) for Ca2+ influx assay. All animal use
protocols were approved by the Institutional Animal Care and Use Committee (IACUC).
Neocortical neurons cultured in 96-well plate were removed their medium and
replaced with dye loading buffer (100 μL/well) containing 4 μM Fluo-3 and 0.04%
Pluronic F-127 in Lock’s buffer (in mM: 8.6 Hepes, 5.6 KCl, 154 NaCl, 5.6 Glucose, 1.0
MgCl2, 2.3 CaCl2, 0.0001 glycine, pH 7.4). After 1 h incubation in dye loading buffer,
cells were washed five times with Locke’s buffer using automated cell washer (Bioteck
instrument Inc., VT, USA) leaving a final volume of 150 μL in each well. The plate was
Sunnyvale, CA, USA) chamber. Cells were excited by the 488-nm line of argon laser and
Ca2+ bound fluo-3 emission in the 500–600 nm range was recorded with the charge
coupled device (CCD) camera. Fluorescence readings were taken once every 9 s for 3
min to establish the baseline and then 50 µL of test compound containing solution (4x)
was added to each well from the compound plate, yielding a final volume of 200 μL/well.
The cells were exposed to the testing compounds for another 42 min.
All NMR data were collected in CDCl3 using Varian Inova 300 MHz and 500
MHz instruments. 1H and 13C NMR signals were referenced to the solvent peaks at 7.26
ppm and 77.0 ppm, respectively. High resolution mass spectra were recorded on a
Thermo-Finnigan MAT900XL spectrometer. A crystal was grown from a CH2Cl2 solution
of 1 by diffusional exchange with hexanes for X-ray crystallographic analysis on a
Bruker APEX CCD detector.
was added to the water. Both treatments and controls were run in duplicate. Adult male
zebra fish of approximately 4 cm length were randomly selected from aquaria and
individually placed in the Erlenmeyer flasks. Time of death was determined by observing
a lack of gill movement and no visible responses to swirling of the flask for one full
minute.
Fractions from the crude lipid extract of a Papua New Guinea mixed collection of
a red marine alga, tentatively identified as Leptofauchea sp. and a marine cyanobacterium
sequentially used flash column chromatography and HPLC, and yielded one more
compound 1 and indicated a molecular weight of 1160 (obs. [M -] m/z 1159). Subsequent
analysis by 13C NMR showed 4 protonated and 8 quarternary 13C signals, indicating that
the compound was either a symmetric dimer having a molecular weight of 1160 or that it
was a monomer with a molecular weight of 580 that easily dimerizes in the mass
spectrometer. High resolution EIMS clarified that it was a monomer with an exact mass
of m/z 575.6211, yielding a molecular formula of C12H5O2Br5 (calc. 575.6201). 1H and 13C
NMR, COSY, HSQC, and HMBC data readily identified two benzenoid rings connected
to one another by an ether linkage. By 13C NMR chemical shifts and consideration of the
molecular formula, one of the rings possessed 2 bromines while the other had 3 bromines
and a hydroxyl group. Further analysis of the HMBC spectrum revealed the A ring to be a
could not be made confidently based on NMR data alone. Further, the 13C NMR data for
69
Table II.1: Comparison of NMR data for two isolates of 1 and 3 (CDCl3)
1 did not match that reported for any known OH-PBDE. Because 1 deposited well-
to give an excellent quality crystal structure (Figure II.2). Comparison of this structure
with the literature for naturally occurring OH-PBDEs showed it to be identical to one
In comparing our 13C NMR data to those reported in the literature for compound
1, a significant discrepancy was noted in the chemical shift for C3 (Table II.1). Only the
Bowden data are included in this table because the only other 13C NMR data available
were recorded in DMSO and not in CDCl3,24 but they were similar to our data. One
hypothesis for this anomalous NMR data was that previous workers had mis-assigned
structure 1 to their material, and that it was in fact the data for another structure. Of the
59 possible regioisomers of 1, 50 could be eliminated based on the fact that the A ring of
Bowden’s isolation of 1 had three protons and the HMBC and COSY correlations were
Bowden. Specifically, the lone proton on the B ring showed HMBC correlations to 4
carbons,25 and thus, the only carbon for which there was not an HMBC correlation should
be para to the carbon bearing this proton. Based on the predicted chemical shifts of the
oxygen bearing carbons, six other isomers could be disregarded, leaving only a single
occurring compound having also been isolated from the marine sponge Dysidea
herbacea.11
71
13
Because the C NMR data for 3 in CDCl3 has not been published,11 this
the 13C NMR shifts of synthetic 3 with those reported by Bowden revealed that they were
13
considerably different (Table II.1). Subsequently, a close inspection of the C NMR
spectrum from the Bowden material revealed a small peak at 121.1 ppm (B. F. Bowden,
personal communication), which corresponds well with our data for C-3 in compound 1.
Because C-3 is para to the only protonated carbon in ring B, this 13C atom is predicted to
relax slowly such that its detection may require longer delay times.27 Moreover, during
our purification of 1-3, it was difficult to remove minor contaminants that had similar
NMR signatures to these OH-PBDEs. In this regard, it is possible that the 113.6 ppm
signal reported for C3 by Bowden et al. was actually the quickly relaxing (e.g.
characterized that eluted very closely to 1 by RP-HPLC and showed Ca2+ modulating
low resolution negative ion ESI-LCMS (obs. [M-H]- m/z 658.74). The isotopic pattern
72
observed for the molecular ion was consistent with the presence of 6 bromine atoms.
Although NMR characterization of this compound was challenging due to the small
quantity isolated (~0.5 mg) and the presence of chromatographically similar impurities,
1
H NMR and 1H-1H COSY analysis revealed a three proton spin system consisting of a
vicinally coupled pair (7.29 ppm and 6.39 ppm) and a third (7.79 ppm) which was meta
coupled to δ 7.29. These shifts and coupling pattern matched those reported for synthetic
phenol (2).
Figure II.3: Ca2+ modulation assay in mouse neocortical neurons for compounds 1-3.
Compounds 1-3 were tested in the Ca2+ modulation assay in mouse neocortical
neurons at various concentrations and their EC50 values were determined to be 16.4 μM
73
(9.8 μM- 27.2 μM, 95% Confidence Intervals), 27.2 μM (12.3 μM- 59.8 μM, 95%
Confidence Intervals) and 42.4 μM (23.2 μM- 77.5 μM, 95% Confidence Intervals),
respectively (Figure II.3). In light of the finding that 1 may act as an environmental
neurotoxin,16 its toxicity to zebrafish was also evaluated. At 5 μg/mL (8.6 μM) media
concentration, the fish immediately attempted to escape out of the container and its
apparent rate of respiration visibly increased; such behaviors were not observed for the
control group. Within 5 minutes, the treated fish appeared to have difficulty maintaining
their proper orientation, and within 20 minutes, both died. Similar effects were seen at 1
μg/mL (1.7 μM), but it took 50 min for the fish to expire. At 0.5 μg/mL (0.86 μM), the
fish appeared to be unaffected for the duration of the 3 hour assay. Regioisomer 3 had the
same effect but was less potent (8.6 μM, death at 25 min; no effect at 1.7 μM), suggesting
Both the long term and the short term neurotoxic effects of polychlorinated
biphenyls (PCBs) have been studied extensively. Notably, PCBs disrupt the intracellular
Ca2+ homeostasis, which is critical for a variety of cell functions, including release of
neurotransmitters.14 As for the longer term effects, PCBs can modulate muscarinic and
nicotinic receptors and also disrupt neurological development.14 Similarly, PBDEs have
been shown to have detrimental effects to learning and memory functions in mice. These
behavioral effects are likely related to the observation that PBDEs reduced the densities
These findings are significant in light of recent reports that OH-PBDEs have
of the dry weight of the organism is composed of OH-PBDEs and visible crystals of these
compounds have been observed in the sponge tissue.13 In addition to sponges, tunicates,
and blue-green algae, we report here that a red alga and/or cyanobacterium living in
association with the red alga as a source of OH-PBDEs.29 Since most of the world's
seaweeds are red algae30 and many invertebrates feed on these algae, the presence of OH-
PBDEs in marine invertebrates may actually reflect the natural occurrence and
lipophilicity.7 OH-PBDEs are expected to be reasonably lipophilic and yet more water
soluble than PBDEs. Therefore, they may be even more prone to bioaccumulation due to
The possibility raised by these findings is that many OH-PBDEs are of natural
regard, it is possible that human populations have been exposed to these materials
through seafood consumption for a lengthy period and that this exposure may have had
historical as well as current health implications. PCBs and brominated flame retardants
other than PBDEs are suspected to be responsible for abnormal behavior phenomena such
events. Thus, it is important to both characterize the origin and occurrence of OH-PBDEs
75
Acknowledgment
The text of II, in whole, has been submitted for publication to Toxicon at the time
of writing. The dissertation author was the primary author and directed and carried out
acknowledged. We thank L. Tan and D. Sherman for help with collection of the algal
13
material, B. F. Bowden for providing copies of their C NMR spectra of 1, and L.
Gerwick for providing zebrafish. We further thank the government of Papua New Guinea
for scientific collection permits and funding from NIH Grant 053398 is acknowledged.
76
Appendix
A colorless block 0.12 x 0.08 x 0.08 mm in size was mounted on a Cryoloop with
Paratone oil. Data were collected in a nitrogen gas stream at 100(2) K using phi and
omega scans. Crystal-to-detector distance was 60 mm and exposure time was 5 seconds
per frame using a scan width of 0.5°. Data collection was 98.7% complete to 67.00° in θ.
A total of 8115 reflections were collected covering the indices, -9<=h<=9, -10<=k<=10,
0.0280. Indexing and unit cell refinement indicated a primitive, monoclinic lattice. The
space group was found to be P2(1)/c (No. 14). The data were integrated using the Bruker
SAINT software program and scaled using the SADABS software program. Solution by
with the proposed structure. All non-hydrogen atoms were refined anisotropically by
full-matrix least-squares (SHELXL-97). All hydrogen atoms were placed using a riding
model. Their positions were constrained relative to their parent atom using the
Table II.3: Atomic coordinates ( x 104) and equivalent isotropic displacement parameters (Å2x
103) for 1. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor.
_____________________________________________________________________________
x y z U(eq)
_____________________________________________________________________________
C(1) 1442(6) 2778(6) 4601(2) 17(1)
C(2) 1873(6) 4251(6) 4435(2) 18(1)
C(3) 2477(6) 5284(6) 4901(2) 17(1)
C(4) 2615(6) 4782(6) 5531(2) 18(1)
C(5) 2167(7) 3329(6) 5701(2) 19(1)
C(6) 1584(6) 2306(6) 5237(2) 18(1)
C(7) 1886(7) 904(6) 3823(2) 19(1)
C(8) 3595(7) 721(6) 3994(2) 19(1)
C(9) 4599(7) -253(6) 3651(2) 19(1)
C(10) 3873(7) -1037(6) 3133(2) 21(1)
C(11) 2171(7) -826(6) 2941(2) 21(1)
C(12) 1183(6) 146(6) 3285(2) 19(1)
O(1) 756(4) 1782(4) 4149(2) 21(1)
O(2) 1203(5) 875(4) 5391(2) 30(1)
Br(1) 1606(1) 4788(1) 3572(1) 23(1)
Br(2) 3061(1) 7263(1) 4676(1) 23(1)
Br(3) 3451(1) 6077(1) 6189(1) 22(1)
Br(4) 5202(1) -2438(1) 2681(1) 24(1)
Br(5) -1156(1) 448(1) 3038(1) 31(1)
_____________________________________________________________________________
82
Table II.5: Anisotropic displacement parameters (Å2x 103)for 1. The anisotropic displacement
factor exponent takes the form: -22[ h2a*2U11 + ... + 2 h k a* b* U12 ]
_____________________________________________________________________________
U11 U22 U33 U23 U13 U12
_____________________________________________________________________________
C(1) 16(2) 20(2) 14(2) -8(2) 1(2) -3(2)
C(2) 17(2) 26(3) 10(2) 1(2) 2(2) 0(2)
C(3) 16(2) 19(2) 15(2) 0(2) 1(2) 0(2)
C(4) 15(2) 24(2) 15(2) -3(2) -2(2) -1(2)
C(5) 23(2) 22(3) 12(2) 1(2) 0(2) 1(2)
C(6) 16(2) 19(2) 18(3) -1(2) 2(2) 1(2)
C(7) 23(2) 22(2) 12(2) 1(2) 4(2) -5(2)
C(8) 25(3) 18(2) 13(2) -1(2) -2(2) -5(2)
C(9) 23(2) 20(2) 14(2) 2(2) -2(2) 0(2)
C(10) 27(3) 23(3) 13(2) 0(2) 5(2) -3(2)
C(11) 30(3) 21(2) 12(2) -2(2) 0(2) -5(2)
C(12) 17(2) 29(3) 12(2) 1(2) 0(2) -9(2)
O(1) 17(2) 30(2) 16(2) -10(2) 1(1) -4(2)
O(2) 31(2) 22(2) 37(2) 2(2) 7(2) -4(2)
Br(1) 22(1) 34(1) 12(1) 4(1) -1(1) -2(1)
Br(2) 24(1) 20(1) 25(1) 5(1) -1(1) -4(1)
Br(3) 25(1) 24(1) 17(1) -6(1) -2(1) -5(1)
Br(4) 28(1) 26(1) 18(1) -4(1) -1(1) 6(1)
Br(5) 16(1) 53(1) 23(1) -18(1) -1(1) -3(1)
_____________________________________________________________________________
84
Table II.6: Hydrogen coordinates ( x 104) and isotropic displacement parameters (Å2x 103)
for 1.
_____________________________________________________________________________
x y z U(eq)
_____________________________________________________________________________
2 Vos, J. G.; Becher, G.; van den Berg, M.; de Boer, J.; Leonards, P. E. G. Brominated flame
retardants and endocrine disruption. Pure Appl. Chem. 2003, 75, 2039-2046.
3 Verreault, J.; Gebbink, W. A.; Gauthier, L. T.; Gabrielsen, G. W.; Letcher, R. J. Brominated
flame retardants in glaucous gulls from the Norwegian Arctic: more than just an issue of
polybrominated diphenyl ethers. Environ. Sci. Technol. 2007, 41, 4925-4931.
4 Kelly, B. C.; Ikonomou, M. G.; Blair, J. D.; Gobas, F. A. P. C. Hydroxylated and methoxylated
polybrominated diphenyl ethers in a Canadian arctic marine food web. Environ. Sci. Technol.
2008, 42, 7069-7077.
5 Betts, K. S. Science news: rapidly rising PBDE levels in North America. Environ. Sci.
Technol. 2002, 36, 50A-2A.
6 Athanasiadou, M.; Cuadra, S. N.; Marsh, G.; Bergman, Å.; Jakobsson, K. Polybrominated
diphenyl ethers (PBDEs) and hydroxylated PBDE metabolites in young humans from
Managua, Nicaragua. Environ. Health Perspect. 2008, 116, 400-408.
8 McKinney, M. A.; Cesh, L. S.; Elliott, J. E.; Williams, T. D.; Garcelon, D. K.; Letcher, R. J.
Brominated flame retardants and halogenated phenolic compounds in North American West
Coast bald eaglet (Haliaeetus leucocephalus) plasma. Environ. Sci. Technol. 2006, 40,
6275-6281.
9 Malmberg, T.; Athanasiadou, M.; Marsh, G.; Brandt, I.; Bergman, Å. Identification of
hydroxylated polybrominated diphenyl ether metabolites in blood plasma from
polybrominated diphenyl ether exposed rats. Environ. Sci. Technol. 2005, 39, 5342-5348.
11 Carte, B.; Faulkner, D. J. Polybrominated diphenyl ethers from Dysidea herbacea, Dysidea
chlorea and Phyllospongia foliascens. Tetrahedron 1981, 37, 2335-2339.
12 Carte, B.; Kernan, M. R.; Barrabee, E. B.; Faulkner, D. J.; Matsumoto, G. K.; Clardy, J.
Metabolites of the nudibranch Chromodoris funerea and the singlet oxygen oxidation products
of furodysin and furodysinin. J. Org. Chem. 1986, 51, 3528-3532.
16 Dingemans, M. M. L.; de Groot, A.; van Kleef, R. G. D. M.; Bergman, Å.; van den Berg, M.;
Henk P.M. Vijverberg, H. P. M.; Westerink , R. H. S. Hydroxylation increases the neurotoxic
potential of BDE-47 to affect exocytosis and calcium homeostasis in PC12 cells. Environ.
Health Perspect. 2008, 116, 637-643.
17 Dingemans, M. M. L.; Ramakers, G. M. J.; Gardoni, F.; van Kleef, R. G. D. M.; Bergman, Å;
Di Luca, M.; van den Berg, M.; Westerink, R. H. S.; Vijverberg, H. P. M. Neonatal exposure to
brominated flame retardant BDE-47 reduces long-term potentiation and postsynaptic protein
levels in mouse hippocampus. Environ. Health Perspect. 2007, 115, 865–870.
19 Li, W. I.; Berman, F. W.; Okino, T.; Yokokawa, F.; Shioiri, T.; Gerwick, W. H.; Murray, T. F.,
2001. Antillatoxin is a novel marine cyanobacterial toxin that potently activates voltage-gated
sodium channels. Proc. Natl. Acad. Sci. 2001, 98, 7599-7604.
20 Rogers, K. L.; Fong, W. F.; Redburn, J.; Griffiths, L. R. Fluorescence detection of plant
extracts that affect neuronal voltage-gated Ca2+ channels. Eur J Pharm Sci. 2002, 15, 321-330.
22 Still, W. C.; Khan, M.; Mitra, A. Rapid chromatographic technique for preparative separations
with moderate resolution. J. Org. Chem. 1978, 43, 2923-2925.
24 Fu, X.; Schmitz, F. J. New brominated diphenyl ether from an unindentified species of
Dysidea sponge. 13C NMR data for some brominated diphenyl ethers. J. Nat. Prod. 1996, 59,
1102-1103.
25 Bowden, B. F.; Towerzey, L.; Junk, P. C. A new brominated diphenyl ether from the marine
sponge Dysidea herbacea. Aust. J. Chem. 2000, 53, 299-301.
27 Crews, P.; Rodríguez, J.; Jaspars, M. Organic Structure Analysis, Oxford, 1998, pp. 169-181.
28 Viberg, H.; Fredriksson, A.; Eriksson, P. Neonatal exposure to polybrominated diphenyl ether
(PBDE 153) disrupts spontaneous behaviour, impairs learning and memory, and decreases
hippocampal cholinergic receptors in adult mice. Toxicol. Appl. Pharmacol. 2003, 192,
95-106.
29 Malmvärn, A.; Marsh, G.; Kautsky, L.; Athanasiadou, M.; Bergman, Å.; Asplund, L.
Hydroxylated and methoxylated brominated diphenyl ethers in the red algae Ceramium
tenuicorne and blue mussels from the Baltic Sea. Environ. Sci. Technol. 2005, 39, 2990-2997.
31 Symons, R. K.; Burniston, D.; Jaber, R.; Piro, N.; Trout, M.; Yates, A.; Gales, R.; Terauds, A.;
Pemberton, D.; Robertson, D. Southern hemisphere cetaceans. A study of the POPs
PCDDs/PCDFs and dioxin-like PCBs in stranded animals from the Tasmanian coast.
Organohalogen Compounds 2003, 62, 257-260.
32 Law, R. J.; Allchin, C. R.; deBoer, J.; Covaci, A.; Herzke, D.; Lepom, P.; Morris, S.;
Tronczynski, J.; de Wit, C. A. Levels and trends of brominated flame retardants in the
European environment. Chemosphere 2006, 64, 187–208.
Chapter III
Lyngbya sp.
101
102
Abstract
A novel fatty acid containing a vinyl chloride moiety was discovered through
sp. collected in Papua New Guinea. While no bioactivity has yet been found to be
associated with this natural product, its biogenesis is of significance in light of other
III.1 Introduction
It has been exceptionally productive to search for novel natural products in marine
malyngamides A-X have been found from marine cyanobacteria and nudibranchs (Figure
III.1, III.2),2,3 and nudibranchs probably obtain these compounds from feeding on
cyanobacteria.4 All of these natural products have a common biosynthetic theme in that
they possess a PKS-derived long lipid chain and an NRPS-derived peptidic portion.5
The apparent rule for the construction of malyngamides is that if two acetate units
and a glycine unit are used for the “head group,” then the final product will contain a
linear “head group” that is attached to either a pyrrolone or a pyrrolidinone with the
exception of malyngamide P (19). Alternatively, if three acetate units and a glycine unit
are used, then a cyclohexane with various oxidation states will be the biosynthetic
product for the “head group.” Moreover, the vinylchloride unit resides on the C-1 of the
glycine unit, suggesting that the sp2 carbon bearing the chloride is not derived from S-
adenosyl-methionine (SAM), but from the C-2 carbon of another acetate unit.6 In HMG-
CoA biosynthesis, an intermediate to the 5-carbon unit of terpenes, the same strategy is
employed, where an aldol reaction between an acetoacetyl CoA and an acetyl CoA is
The two different biosynthetic pathways leading to 33 and 37 utilize the same chemistry until the
chlorinated intermediate 31, after which the corresponding decarboxylases distinguish the
pathways (32 for jamaicamide A; 34 for curacin A).
has been shown to be derived from the C-2 of acetate.2 Recently, the biosynthesis of
jamaicamides was found to be closely related to that of curacins, where the two different
pathways share a common intermediate (31) as shown in Scheme III.1.8 Herein described
is a novel natural product isolated from Lyngbya that also possesses a vinylchloride
group.
Tropical areas are biologically very diverse environments in which many species
compete for space and nutritional resources. Therefore, organisms in such environments
106
The satellite image was obtained through Google Earth™. The site of collection is indicated by
the red circle.
Figure III.4: Zoom-up of the Satelite Image of the Lyngbya Collection Site
The satellite image was obtained through Google Earth™. The site of collection is indicated by
the red circle.
107
are expected to possess a rich collection of secondary metabolites suited for competition.
In order to acquire such metabolites, a sample of Lyngbya sp. was collected in Papua
New Guinea (PNG). The cyanobacterial material was collected at a site between the
northeast coast of New Britain Island and the southwest coast of Credner Island by
snorkeling (S 4º 15.981'; E 152º 19.735'; Figure III.3 and III.4) at the depth of 1 m.9 The
alga had the typical hair-like morphology of Lyngbya, but had a greenish color unlike
Since water-soluble organic natural products are very difficult to handle, they are
not as well studied as other organic compounds. Therefore, it is a better likelihood to find
a novel compound in aqueous soluble materials. One of the aims of this study was to
develop a methodology to purify natural products from the aqueous extract. The
cyanobacterial material (WG-1560) was extracted exhaustively with CH2Cl2 and MeOH
(2:1) for the isolation of lipid contents.10 The remaining aqueous layers and the EtOH
used for preservation of the alga were combined and the volatiles were removed under
vacuum. The residues were resuspended in EtOH and the insolubles were filtered off
108
(Scheme III.2). The filtrate was then subjected to two steps of rough normal phase
separable by TLC was subjected to flash column chromatography with silica gel as the
stationary phase and a gradient mixture of MeOH and CH2Cl2 as the mobile phase to
obtain several fractions. One of the fractions contained a fairly pure compound and it was
further purified by semi-prep reverse phase HPLC to obtain 4 mg of a colorless oil. The
compound was named “credneric acid,” after the nearby island where the cyanobacterial
The purified material was first analyzed by 1H NMR (Figure III.6), which
revealed that the compound was a short chain fatty acid. However, one unusual singlet
signal was observed at 5.79 ppm. Further analysis by 13C NMR, COSY, and HSQC (Table
III.1) led to substructures as shown in Scheme III.3. Finally, the two and three bond
the three substructures (Table III.1). The unknown “R” group was determined to be a
chlorine atom based on the high resolution mass spectroscopy analysis (m/z 217.0990
expected for C11H18O2Cl [M+H]; observed m/z 217.0995) and the NMR chemical shift
(1H 5.79 ppm; 13C 113.1). The stereochemical assignment of the two olefins were made
by the coupling constant between the two vinyl protons (5.42 ppm and 5.49 ppm; J =
15.4 Hz) and 1D ROESY correlations of key protons on and around the olefins (Scheme
III.4).
110
Table III.1: 1H and 13C NMR, COSY, HSQC, and HMBC data for credneric acid (38)
The HSQC data are reflected in the assignment of the 1H NMR signals to their respective 13C
NMR signals.
13 1
Position C H COSY HMBC
While the gene cluster responsible for the biosynthesis of jamaicamides has been
identified already,2 the biosynthetic process for this halogenation step is only speculative
and largely based on other related natural products.7 Since there is remarkable
resemblance between credneric acid (38) and the middle portion of the jamaicamides, it
insights into the halogenation process for the jamaicamides and the malyngamides. For
halogenating enzyme would be easier to synthesize for credneric acid than for the
jamaicamides.
Credneric acid was tested in H-460 cytotoxicity and Na+ channel activation and
blocking assays, but it was found to be inactive (data not shown). It is possible that the
compound is not able to cross cell membranes due to its polar carboxylic acid group.
112
Alternatively, it is possible that this natural product lacks functional groups necessary for
bioactivity. After the discovery of 38, a related natural product, credneramide (39, Figure
III.7), was isolated from the CH2Cl2 / MeOH extract of the same collection of Lyngbya in
our laboratory.11 This amide derivative of 38 was found to be only moderately active in
anti-parasitic assays against malaria, chagas, and leishmania and Na+ channel blocking
assay in Neuro-2a cells. It may be that 38 is merely a biosynthetic precursor for amides,
such as 39, and it is not produced as a secondary metabolite with its own biological or
ecological significance.
foundational biosynthetic units. For example, malyngamides all have lyngbic acid in
A), and lyngbyamides B-D have the same cyclopropane fatty acid in common (Scheme
III.5).12 All of these amides have some structural resemblance to bacterial quorum sensing
molecules such as AHLs (N-acyl homoserine lactones; Figure III.8).13 This resemblance
may be related to the biological roles that these amide natural products have in nature.
other compounds, such as 39. Given the precedence with lyngbyamides and
malyngamides, discovery of other credneric acid derivatives are expected in the future.
III.6 Conclusion
characterized from a PNG collection of Lyngbya. Despite its inactivity in the few
bioassays tested thus far, this natural product may play an important role as a template for
synthesizing bioactive amide derivatives. Credneric acid may also be useful for the study
General
114
Unless noted otherwise, all materials were purchased from commercially available
sources and were used without further purification. Flash chromatorgraphy14 and vacuum
liquid chromatography were performed using EM Science silica gel (230-400 mesh).
TLC was performed using EM Science pre-coated silica gel plates (Merck 60 F254). IR
spectra were recorded on a Nicolet Magna-IR 550. 1H NMR and 13C NMR spectra were
referenced to the residual solvent peaks at 7.26 ppm and 77.0 ppm, respectively. High
The cyanobacterial material (41 g, dry weight) was extracted exhaustively with
CH2Cl2 and MeOH (2:1) for the isolation of lipid contents. The remaining aqueous layers
and the EtOH used for preservation of the alga were combined and the volatiles were
removed under vacuum. The residues were resuspended in EtOH and the insolubles were
filtered off (Scheme III.1). By using vacuum, the filtrate was pulled through a pad of
water-treated silica gel with differing solvents (vacuum liquid chromatography, VLC).
Most of the organic content was eluted with EtOH, but not with other solvents (Scheme
III.1). Upon concentration under vacuum, the residues were subjected to solid phase
extraction (SPE) with silica gel. By eluting with different solvents (1:1 CH 2Cl2/EtOAc;
EtAcO; 1:9 MeOH/CH2Cl2; 3:7 MeOH/CH2Cl2), four fractions (4 mg, 422 mg, 1.08 g,
and 1.09 g, respectively) were obtained (Scheme III.1). Only the third fraction (1:9
MeOH/CH2Cl2 elution) had a TLC profile that appeared suitable for normal phase
115
separation of compounds. The material in the third SPE fraction was subjected to flash
column chromatography with silica gel as the stationary phase and a gradient mixture of
MeOH and CH2Cl2 as the mobile phase to obtain several fractions. One of the fractions
contained a fairly pure compound (7 mg) and it was further purified by semi-prep reverse
phase HPLC (250 x10.00 mm Synergi 4μ Hydro-RP 80A column, MeCN/H 2O gradient)
to obtain 4 mg of a colorless oil. IR (film on KBr) 3428 (br), 2965, 2930, 2873, 1711,
1632, 1431, 1410, 1284, 1252, 1212, 1158, 968, 795 cm-1. 1H NMR (500 MHz, CDCl3) δ
5.79 (s, 1H), 5.49 (dt, J = 15.4, 6.2 Hz, 1H), 5.42 (dt, J = 15.4, 6.4 Hz, 1H), 2.73 (d, 6.2
Hz, 2H), 2.44 (t, J = 7.4 Hz, 2H), 2.35 (dt, J = 6.4, 7.0 Hz, 2H), 2.16 (t, J = 7.7 Hz, 2H),
1.44 (h, J = 7.5 Hz, 2H), 0.92 (t, J = 7.3 Hz, 3H). 13C NMR (75 MHz, CDCl3) δ 178.9,
141.4, 130.5, 128.1, 113.1, 37.9, 33.7, 32.2, 27.4, 20.2, 13.9. HRMS (EI): calcd for
Figure III.13: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 2.73 ppm in CDCl3
121
Figure III.14: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 5.79 ppm in CDCl3
122
Figure III.15: Credneric Acid (38) 1D ROESY Spectrum: Irradiated at 2.16 ppm in CDCl3
123
1 Gerwick, W. H.; Tan, L. T.; Sitachitta, N. The Alkaloids; Cordell, G. A., Ed.; Academic Press;
San Diego, CA 2001; Vol. 57, pp 75-184.
2 Edwards, D. J.; Marquez, B. L.; Nogle, L. M.; McPhail, K.; Goeger, D. E.; Roberts, M. A.;
Gerwick , W. H. Structure and biosynthesis of the jamaicamides, new mixed polyketide-
peptide neurotoxins from the marine cyanobacterium Lyngbya majuscula. Chemistry &
Biology 2004, 11, 817–833.
3 (a) Appleton, D. R.; Sewell, M. A.; Berridge, M. V.; Copp, B. R. A new biologically active
malyngamide from a New Zealand collection of the sea hare Bursatella leachii. J. Nat. Prod.
2002, 65, 630-631. (b) Kan, Y.; Sakamoto, B.; Fujita, T.; Nagai, H. New malyngamides from
the Hawaiian cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2000, 63, 1599-1602. (c)
Gallimore, W. A.; Scheuer, P. J. Malyngamides O and P from the sea hare Stylocheilus
longicauda. J. Nat. Prod. 2000, 63, 1422-1424. (d) Milligan, K. E.; Marquez, B.; Williamson,
R. T.; Davies-Coleman, M.; Gerwick, W. H. Two New Malyngamides from a Madagascan
Lyngbya majuscula. J. Nat. Prod. 2000, 63, 965-968. (e) Kan, Y.; Fujita, T.; Nagai, H.;
Sakamoto, B.; Hokama, Y. Malyngamides M and N from the Hawaiian red alga Gracilaria
coronopifolia. J. Nat. Prod. 1998, 61, 152-155. (f) Wu, M.; Milligan, K. E.; Gerwick, W. H.
Three new malyngamides from the marine cyanobacterium Lyngbya majuscula. Tetrahedron
1997, 53, 15983-15990. (g) Todd, J. S.; Gerwick, W. H. Malyngamide I from the tropical
marine cyanobacterium Lyngbya majuscula and the probable structure revision of
stylocheilamide. Tetrahedron Lett. 1995, 36, 7837-40. (h) Praud, A.; Valls, R.; Piovetti, L.;
Banaigs, B. Malyngamide G: structure of a new chlorinated amide from a blue-green alga
epiphytic on Cystoseira crinita. Tetrahedron Lett. 1993, 34, 5437-40. (i) Gerwick, W. H.;
Reyes, S.; Alvarado, B. Two malyngamides from the Caribbean cyanobacterium Lyngbya
majuscula. Phytochem. 1987, 26, 1701-4. (j) Ainslie, R. D.; Barchi, J. J., Jr.; Kuniyoshi, M.;
Moore, R. E.; Mynderse, J. S. Structure of malyngamide C. J. Org. Chem. 1985, 50, 2859-62.
(k) Cardellina, J. H., II; Marner, F. J.; Moore, R. E. Malyngamide A, a novel chlorinated
metabolite of the marine cyanophyte Lyngbya majuscula. J. Am. Chem. Soc. 1979, 101,
240-2. (l) Cardellina, J. H., II; Dalietos, D.; Marner, F. J.; Mynderse, J. S.; Moore, R. E. (-)-
trans-7(S)-Methoxytetradec-4-enoic acid and related amides from the marine cyanophyte
Lyngbya majuscula. Phytochem. 1978, 17, 2091-5.
4 Garson M. J. Marine mollusks from Australia and New Zealand: chemical and ecological
studies. Prog. Mol. Subcell. Biol. 2006, 43, 159-74.
6 Dewick, P. M. Medicinal natural products; a biosynthetic approach, 2nd Ed. John Wiley &
Sons Ltd. 2001, pp 169-170.
7 Vaillancourt, F. H.; Yeh, E.; Vosburg, D. A.; Garneau-Tsodikova, S.; Walsh, C. T. Nature’s
inventory of halogenation catalysts: oxidative strategies predominate. Chem. Rev. 2006, 106,
3364–3378.
125
8 Gu, L.; Wang, B.; Kulkarni, A.; Geders, T. W.; Grindberg, R. V.; Gerwick, L.; Håkansson, K.;
Wipf, P.; Smith, J. L.; Gerwick, W. H.; Sherman, D. H. Metamorphic enzyme assembly in
polyketide diversification. Submitted for peer review.
10 The extracted alga was numbered as “WG-1560” for tracking purposes in the lab.
13 Uroz, S.; Dessaux, Y.; Oger, P. Quorum sensing and quorum quenching: the Yin and Yang of
bacterial communication. ChemBioChem 2009, 10, 205-216.
14 Still, W. C.; Khan, M.; Mitra, A. Rapid chromatographic technique for preparative separations
with moderate resolution. J. Org. Chem. 1978, 43, 2923-2925.
Chapter IV
stereochemical reference
126
127
Abstract
absolute stereochemistry of an enatiomeric pair of novel fatty acids isolated from two
subsequent report of a total synthesis of a related natural product was in conflict with the
results of this approach. There may have been an error in the NMR analysis of the
cyclopropane fatty acid derivative or in the optical rotation measurement of the fatty
acids themselves.
128
IV.1 Introduction
these derivatives was significant quantities of the parent fatty acid (1). Subsequently, we
isolated what appeared to be the identical fatty acid, based on 1H, 13C NMR, and FAB-MS
data, from the same genus of cyanobacterium collected in Curaçao. However, the optical
rotation values of these two samples were found to be opposite one another ([α]D25 +19.3˚
for the Madagascar sample 1-m and [α]D25 –9.8° for the Curaçao sample 1-c), a finding
purification, and measurement of the rotation value for these derivatives.2 The
Madagascar-derived 4-bromobenzyl ester gave an [α]D25 +8.8° whereas the Curaçao fatty
The finding of two enantiomeric forms of 1 from two different collections of the
each in order to gain insight into the biological properties of the two series as well as the
129
fatty acids had been reported at the outset of this project. Initial attempts to crystallize the
were inspired to develop a new strategy for their stereochemical determination. The use
of a Mosher ester was discounted because this technique is limited to the configurational
was realized that the carboxyl functionality in 1 provided an opportunity for the addition
introduce a chiral appendage at C-2. We then envisioned that this latter appendage could
be related to chiral centers present in the natural product using a variety of NMR
techniques (nOe and J values). To our knowledge, this strategy which interfaces the
power of chirally-induced chemical synthesis with NMR spectroscopy has never been
Both enantiomers of the cyclopropyl fatty acid (1) were isolated from the
(3), was condensed with each acid to produce imide derivatives 4 and 5 (Scheme IV.1).
auxiliary and the Madagascar-derived cyclopropyl fatty acid (1-m) to produce imide
derivative 6. Each of these imides was methylated under standard conditions to produce
130
α-methyl derivatives 7 (from 4), 8 (from 5) and 9 (from 6). For all three synthetic
sequences, only a single stereoisomer was observed as the final reaction product (7-9).
The 1H NMR of 7 and 8 showed distinctive chemical shifts and splitting patterns
for the cyclopropane ring protons, confirming that they were indeed diasteriomeric to one
other, and thus 1-m and 1-c were indeed enantiomers. However, the 1H NMR spectrum of
8 was identical to that of 9, indicating that these derivatives were enantiomeric. This was
substantiated by measuring similar rotations of opposite sign for 8 (+48°) and 9 (-25°).
and R, respectively) and the cyclopropane ring (+ and -), the stereochemistry of the α-
methyl carbon in these two derivatives must be opposite to each other as well. Because 7
and 9 possess different relative stereochemistries between the cyclopropyl ring and the α-
methyl group, as evidenced by their different NMR spectra, we conclude that the
and not the cyclopropyl ring stereochemistry. The overriding influence of the auxiliary
during the conversion of 6 to 9 was predicted based on the rationale that the long alkyl
chain would be as far away from ionic lithium-chelated six-membered ring as possible
during the alkylation transition state, thus overcoming any mismatching effects during
131
this reaction (Figure IV.1). Hence, the stereochemistry of the α-methyl group in 7 and 8
relate the stereochemistry of the α-methyl group to that of the cyclopropyl bridgehead
stereocenters in derivatives 7 and 8. First, all proton chemical shifts were assigned for all
three derivatives (7-9) using 1H-1H COSY data. The values of the key 1H-1H coupling
decoupled 1H NMR, and X-WIN NMR spin simulation program. These were in good
Using the Karplus-type equation proposed by Altona,4 the dihedral angles between
the key bonds were estimated (Figure IV.2). Alternatively, the Karplus-type equation
proposed by Barfield and Smith,5 which takes into consideration the bond angles, did not
yield reasonable dihedral angles (Table IV.1). Despite the fact that the key coupling
constants are all similar to each other (around 7 Hz), a qualitative trend was seen in the
calculated dihedral angles. For example, the dihedral angle predicted from the ROESY
analyses for H3α-C3-C1′-H1′ is larger (60°) than that for H3β-C3-C1′-H1′ and the coupling
constant is indeed larger for H3β-H1′ than for H3α-H1′ (Figure IV.2). Detailed analysis of
derivative 7 and 17 for derivative 8. The relative stereochemistry predicted between the
α-methyl center and cyclopropyl ring centers by these two approaches were in excellent
agreement (Figure IV.2). Finally, the energy minimized structures of both 7 and 8 were
optimization (Hartree-Fock) calculations and compared with those obtained from the
Altona calculations and the ROESY data.6 All three lines of evidence converge to
133
establish compound 7 as having 1R, 2R, 2´S stereochemistry and compound 8 to have 1S,
Table IV.1: Dihedral angles calculated or simulated by different methods for 7 and 8
Dihedral angles from Spartan were calculated with semi-empirical PM3; those from ROESY
were estimated manually using a molecular model kit; those from Altona & Barfield equations
were calculated from experimental coupling constants.
134
However, as this work was concluding, Baird and coworkers reported their total
cyclopropyl fatty acid to be 1R, 2R (Scheme IV.2).7 This result directly contradicted our
work because (+)-1 was assigned 1S, 2S in our hands whereas (+)-1, one of the
optical rotation measured by us or Baid et al. had the wrong sign, 3) The stereochemical
assignment of the chiral starting material that Baird et al. used was erroneous. It should
be noted that the only other asymmetric synthesis of grenadamide was reported by Taylor
and coworkers who simply correlated their absolute stereochemical assignments of their
synthetic products with those of Baird's.9 Therefore, Baird's work is the only source of
mentioned above was evaluated through a literature search. The history of absolute
remarkably complex (Figure IV.3). It is interesting that the stereoconfiguration was never
is unclear how the original compound's absolute configuration was assigned in 1927
(Figure IV.3) because the first absolute stereochemical assignments (tartaric acid) were
Figure IV.2: Dihedral angles predicted from the 1H-1H coupling constants and Altona
equation (A) and select ROESY correlations for 7 and 8 (B).
Here lies the weakness of the modern scientific research; so much of our work
relies heavily on the previous work, but often without any verification due to time
constraints. It is therefore admirable that so many natural product chemists have re-
visited the structures of natural products that were previously characterized and have
136
means.12 Synthetic chemists, by default, synthesize the reported structures and are able to
discover errors in the structure assignment as an additional benefit of the synthetic work.
However, re-visiting a natural product's structure is not an activity that natural product
used by Baird et al. Thus, the synthesis of 11 was performed according to literature.13 X-
ray crystallography was considered the best choice of stereochemical assignment due to
IV.3). Careful 1D and 2D NMR analysis of 15 and 16 indicated that the C-2 position had
correlation was observed between H-1 and H-4, which suggested that the cyclopropane
ring had a trans relative stereochemistry (Figure IV.4). Considering that α-epimerization
of 14 was possible during chromium oxidation to the carboxylic acid (Scheme IV.3), the
verify or discount this assignment, a suitable crystal of 15 was obtained for X-ray
crystallographic analysis. To our grief, the crystal structure revealed that the
Figure IV.4: 1H and 13C NMR assignments and an NOE correlation of 16 in CDCl3
The structure is shown as trans due to the NOE correlation observed between H-1 and H-4.
However, the relative sterechemistry was later shown to be cis based on an X-ray crystal structure
of 15 (Figure IV.5). The 1H and 13C assignments were made through the use of COSY, HSQC, and
HMBC spectra in conjunction with the 1H and 13C NMR spectra. The 1H chemical shifts are
shown in red and the 13C chemical shifts are shown in blue.
cyclopropane ring retained its cis configuration (Figure IV.5). This result unequivocally
assigns the absolute configurations of 1-m to be 1S, 2S (Figure IV.6), thereby re-
The final explanation to be evaluated for reconciling our work and Baird's work
was to ascertain the validity of optical rotation values obtained for 1-m and 1-c. Of the
many structure determination tools utilized by natural product chemists today, the
contamination, and depending on the condition of the light source, different values can be
obtained for the same sample. Moreover, unlike spectra obtained from NMR or MS,
optical rotation values usually do not leave substantial records of their measurement. The
optical rotation value obtained by C. Nannini ([α]D –7.6° for 1-c)1 was consistent with
139
this work ([α]D –9.8° for 1-c). Therefore, it is unlikely that any human error, such as
IV.4 Conclusions
A pair of enantiomeric natural products were isolated from the same genus
collected from different locations. We developed a new concept for determining absolute
stereochemistry of natural products (i.e. the SISTER method, Figure IV.7). However, X-
ray crystallographic investigation indicated that either our ROESY and J-based analysis
was incorrect or one of the optical rotation values was wrong. It may require a re-
collection of the algal material to resolve the confusion that has plagued this project.
However, as shown in Figure IV.3, the ultimate origin of the absolute stereochemical
absolute configurations of 11 were never fully determined, but that its arbitrary
assignments happened to be the correct ones. After all, there is always a fifty-fifty chance
General
and a Nicolet 510 spectrophotometer, respectively. NMR spectra were recorded either at a
1
H resonance frequency of 600.04 MHz (Bruker DRX 600) or 400.13 MHz (Bruker DPX
400). The Bruker DRX 600 was equipped with a Bruker Q-Switch TXI probe and the
DPX 400 was equipped with a Bruker BBO probe. All chemical shifts are reported
relative to residual CHCl3 (δH 7.27, δC 77.23) or C6HD5 (δH 7.16) as internal standard.
measured with a Perkin-Elmer model 243 polarimeter. HPLC was performed using
Waters 515 pumps and a Waters 996 photodiode array spectrophotometer. TLC grade
(10-40 mesh) silica gel was used for vacuum chromatography, and Merck aluminum-
backed TLC sheets (Silica gel 60 F254) were used for TLC. All solvents were purchased
as HPLC grade. THF and triethylamine were freshly distilled before they were used
fatty acids (1-m and 1-c), and the amide derivatives (7, 8, and 9) were dried with a
141
vacuum pump before use in synthetic reactions. Other reagents were used without any
purification unless specified. All reagents, except for the natural fatty acids 1-m and 1-c
and the solvents were purchased from Sigma-Aldrich. All reactions were carried out
under argon.
acetonitrile/CH2Cl2 (1:9). The same procedure was used to purify 1-c from a Curaçao
collection of this cyanobacterium. A total of 42.7 mg of crude 1-m and 23.3 mg of crude
1-c were isolated. Due to the inherently poor chromatographic properties of these fatty
acids, these crude materials with relatively low purity were used for subsequent reactions
oxazolidin-2′′′-one (5). In 300 μL of THF were dissolved 23.3 mg crude 1-c which was
cooled to -78°C. To this solution were added pivaloyl chloride (17 μL, 0.14 mmol) and
triethylamine (16 μL, 0.11 mmol). The reaction mixture was stirred and allowed to warm
(58.7 mg, 0.331 mmol) was dissolved in 300 μL of THF and cooled to -78°C. To this
solution was added 185 μL of 2.0 M n-butyllithium solution in cyclohexane (0.37 mmol)
and the reaction mixture was stirred for 20 minutes. The lithiated oxazolidinone solution
was transferred to the solution of the mixed anhydride at -78°C. This reaction was stirred
and allowed to warm up to RT over 24 hours, and then quenched with 5 mL of aqueous 1
M NH4Cl solution and the aqueous layer extracted with CH2Cl2 (3 × 5 mL). The organic
layer was subsequently washed with 5 mL of H2O and dried with MgSO4. The
concentrated mixture was analyzed by TLC (MeOH/CH2Cl2, 1:19) which showed that no
starting material (2) was present, thus indicating a yield >95%. The resultant mixture was
purified by silica gel flash chromatography with CH2Cl2 as eluent to yield 36.2 mg of a
pale-yellow oil. This was used for subsequent reactions without further purification. 1H
NMR (CDCl3; 400MHz) δ 0.23 (2H, m, 3′-CH2), 0.49 (2H, m, 1′-CH and 2′-CH), 0.88
(3H, t, J = 6.9 Hz, 7′′-CH3), 1.26 (12H, m, alkyl chain), 1.60 (2H, m, 3-CH 2), 2.75 (1H,
dd, J = 13.2 Hz, 9.7 Hz, Bn-CH2), 3.02 (2H, m, 5′′′-CH2), 3.31 (1H, dd, J = 13.3 Hz, 3.0
Hz, Bn-CH2), 4.18 (2H, m, 2-CH), 4.67 (1H, m, 4′′′-CH), 7.19-7.36 (5H, m, phenyl); MS
(FAB) 372.1 (M+1). Likewise, semi-purified compound 4 (111.4 mg, crude weight) was
400 μL of THF and the solution was cooled to -78°C. Lithium hexamethyldisilazide (95
μL, 1.0 M solution in THF, 95 μmol) was added dropwise and the mixture was stirred for
20 minutes. Excess methyl iodide (55 μL) was added dropwise at -78°C and the reaction
mixture was allowed to warm to RT over 6 hours. The reaction was then quenched with 4
mL of NH4Cl aqueous solution and extracted with CH2Cl2 (3 ×5 mL). The combined
organic layers were washed with 2 mL of H2O and then dried with MgSO4. The resulting
mixture was concentrated under reduced pressure and passed through a short silica gel
column followed by a C18 column (eluents were CH2Cl2 and MeOH, respectively). After
evaporation of solvents under reduced pressure, 27.7 mg of crude material was recovered,
and this was purified by HPLC using a reverse phase semi-preparative column (YMC,
ODS-AQ, 250 × 10 mm I.D., 5 μm, 120 Å; H2O/MeOH, 0.7:99.3). The major compound
in the reaction mixture was the by-product, pivaloyl oxazolidinone; however, 2.1 mg of 7
were isolated. Based on the HPLC peak integration of 4 and 7 in the reaction mixture, the
yield of 7 was estimated to be 70%. Only one isomer of the methylated product was
observed. 8 and 9 were synthesized in the same manner and only one diastereomer was
observed in each synthesis (50% and 60% yields, resepectively). Colorless oil; UV λ max
255 nm; IR υ 2923cm-1, 1776 cm-1, 1697cm-1; 1H NMR (C6D6; 400MHz) δ 0.27 (2H, m,
3′-CH2), 0.58 (1H, m, 1′-CH), 0.62 (1H, m, 2′-CH), 0.91 (3H, t, J = 6.8 Hz, 7′′-CH3), 1.29
(12H, m, alkyl chain), 1.34 (3H, d, J = 7.0 Hz, 2-CCH3), 1.46 (1H, ddd, J = 13.6 Hz, 6.5
Hz, 6.4 Hz, pro-S-3-CH2), 1.84 (1H, ddd, J = 13.6 Hz, 7.2 Hz, 7.2 Hz, pro-R-3-CH2), 2.29
(1H, dd, J = 13.4 Hz, 9.3 Hz, Bn-CH2), 2.97 (1H, dd, J = 13.4 Hz, 3.0 Hz, Bn-CH 2), 3.20
144
(1H, dd, 5′′′-CH2), 3.46 (1H, dd, J = 9.1 Hz, 2.6 Hz, 5′′′-CH2), 4.14 (1H, ddt, J = 7.2 Hz,
13
7.0 Hz, 6.4 Hz, 2-CH), 4.25 (1H, m, 4′′′-CH), 6.90-7.10 (5H, m, phenyl); C NMR
(CDCl3; 100 MHz) δ 12.4, 14.3, 16.7, 17.5, 18.9, 22.9, 29.6, 29.7, 29.8, 32.1, 34.4, 38.1,
38.3, 38.4, 55.6, 66.2, 127.6, 129.2, 129.7, 135.6, 153.3, 177.5. [α]D28 = + 48° (c = 1.2,
CHCl3).
oxazolidin-2′′′-one (8). Colorless oil; UV λ max 257 nm, IR υ 2918cm-1, 1776 cm-1,
1692cm-1 ; 1H NMR (C6D6; 400MHz) δ 0.23 (1H, m, 3′-CH2), 0.39 (1H, m, 3′-CH2), 0.45
(1H, m, 2′-CH), 0.59 (1H, m, 1′-CH), 0.92 (3H, t, J = 6.9Hz, 7′′-CH3), 1.29 (12H, m,
alkyl chain), 1.33 (3H, d, J = 7.0 Hz, 2-CCH3), 1.43 (1H, ddd, J = 13.7 Hz, 6.9 Hz, 6.0
Hz, pro-S-3-CH2), 1.84 (1H, ddd, J = 13.7 Hz, 7.3 Hz, 7.1 Hz, pro-R-3-CH2), 2.30 (1H,
dd, J = 13.4 Hz, 9.7 Hz, Bn-CH2), 2.98 (1H, dd, J = 13.5 Hz, 2.7 Hz, Bn-CH 2), 3.22 (1H,
dd, 5′′′-CH2), 3.47 (1H, dd, J = 9.0 Hz, 2.5 Hz, 5′′′-CH2), 4.16 (1H, ddq, 7.3 Hz, 7.0 Hz,
6.0 Hz, 2-CH), 4.23 (1H, m, 4′′′-CH), (5H, m, phenyl); 13C NMR (CDCl3, 100MHz) δ
11.9, 14.3, 16.7, 17.4, 19.3, 22.9, 29.6, 29.7, 29.8, 32.1, 34.1, 38.1, 38.2, 55.7, 66.2,
127.5, 129.1, 129.7, 135.6, 153.3, 177.6. [α]D28 = + 42° (c = 2.5, CHCl3).
(C6D6; 400MHz) δ 0.23 (1H, m, 3′-CH2), 0.39 (1H, m, 3′-CH2), 0.45 (1H, m, 2′-CH), 0.59
(1H, m, 1′-CH), 0.92 (3H, t, J = 6.9Hz, 7′′-CH3), 1.29 (12H, m, alkyl chain), 1.33 (3H, d,
145
J = 7.0 Hz, 2´-CCH3), 1.43 (1H, ddd, J = 13.7 Hz, 6.9 Hz, 6.0 Hz, pro-S-3´-CH2), 1.84
(1H, ddd, J = 13.7 Hz, 7.3 Hz, 7.1 Hz, pro-R-3´-CH2), 2.30 (1H, dd, J = 13.4 Hz, 9.7 Hz,
Bn-CH2), 2.98 (1H, dd, J = 13.5 Hz, 2.7 Hz, Bn-CH2), 3.22 (1H, dd, 5´´-CH2), 3.47 (1H,
dd, J = 9.0 Hz, 2.5 Hz, 5´´-CH2), 4.16 (1H, ddq, 7.3 Hz, 7.0 Hz, 6.0 Hz, 2´-CH), 4.23
the OSU Biochemistry and Biophysics NMR facility for use of the Bruker DRX600
NMR instrument (V. Hsu), the OSU Mass Spectrometry Facility (J. Moore).
146
2 Complete purification of the fatty acids themselves were not possible by normal means.
3 (a) Teng, Rong-Wei; Shen, Ping; Wang, De-Zu; Yang, Chong-Ren., Bopuxue Zazhi, 2002,
19(2), 203-223. (b) Harada, Nobuyuki; Watanabe, Masataka; Kosaka, Masashi; Kuwahara,
Shunsuke., Yuki Gosei Kagaku Kyokaishi, 2001, 59(10), 985-995. (c) Kusumi, Takenori;
Ohtani, Ikuko I., Biology-Chemistry Interface, 1999, 103-137
6 See the experimental section for the geometry optimized 3D structures of 7 and 8.
8 Coxon, G. D.; Al-Dulayymi, J. R.; Baird, M. S.; Knobl, S.; Robertsa, E.; Minnikin , D. E. The
synthesis of (11R,12S)-lactobacillic acid and its enantiomer. Tetrahedron Asymm. 2003, 14,
1211-1222.
9 Avery, T. D.; Culbert, J. A.; Taylor, D. K. The first total synthesis of natural grenadamide.
Org. Biomol. Chem. 2006, 4, 323-330.
10 Alternative interpretations of the ROESY correlations were considered. However, the original
interpretations were more consistent with the molecular modeling results.
13 Grandjean, D.; Pale, P.; Chuche, J. Enzymatic hydrolysis of cyclopropanes. Total synthesis of
optically pure dictyopterenes A and C. Tetrahedron 1991, 47, 1215-1230.
Chapter V
172
173
Abstract
due to its apparent selectivity for a nicotinic acetylcholine receptor subtype. In order to
high overall yields and high stereoselectivity from readily commercially available
materials. However, it was found that none of the stereoisomers of epiquinamide was
responsible for the observed bioactivity. Through this error, the importance of total
V.1 Introduction
rich source of bioactive and structurally complex alkaloids (Figure V.1).1,2 Typically,
these alkaloids have pyrrolidine (1), piperidine (2), pyrrolizidine (3), indole (4),
secondary metabolites can have a variety of neurological activities both in vitro and in
vivo.3 Many of the frog skin natural products have been traced back to the amphibian's
dietary sources, such as insects and other arthropods.1 However, the ultimate source may
175
be of microbial origin in some cases.4 This is an ecological theme parallel to that seen
Epipedobates tricolor
Photograph from reference 3
Figure V.2: Skin of Epipedobates tricolor contains epibatidine and epiquinamide
(Figure V.2) in 1992 by Daly and co-workers and has become one of the cornerstone
Epiquinamide (9) was isolated along with 8 from the same frog species in 20037 and has
attracted much attention due to its potent and selective agonistic activities against β2
them are known to be sufficiently selective for subtypes of the nicotinic acetylcholine
tissues of the human body, and it is conceivable that by selectively targeting a certain
tissue, the possibility of side effects could be reduced.9 However, the major limiting
factor in testing this compound was its low availability from nature (240 μg from the
176
Scheme V.2: Applying the synthetic logic of quinolizidine biosynthesis to the synthesis of
epiquinamide skeleton
skins of 183 frogs).7 To meet this need and to explore new chemistry for the construction
of the novel quinolizidine skeleton, a number of research groups have undertaken the
total synthesis of 9.10 To date, there have been four11,12,13,14 enantiospecific total syntheses
there have been two racemic syntheses of 917,18 and two syntheses of the epimer of 9
177
(15).12,19 In the literature on epiquinamide, the optical rotation value of the natural product
was never reported due to its low availability.7 Access to both enantiomers was therefore
of great importance.
appendages that could be cyclized to give the second ring. The initial synthetic strategy
was formed around the reductive amination reaction to furnish the bicyclic tertiary amine
(Schemes IV.1 through IV.3). This idea was inspired by the observation that there had
exists in the literature that suggested that such reductive amination using H2 would give
trans stereochemistry via chelation control22 (Scheme V.2; H2 reaction) while cis
stereochemistry would be achieved if hydride was used20 via a Felkin-Anh type transition
state (Scheme V.2; hydride reaction).23 This latter stereochemical outcome was expected24
178
for the stereoselective reduction of 2 based on the assumption that the protecting group
would not chelate to the boron atom of NaBH3CN under acidic conditions.
(Scheme V.3).25 Being inspired by nature's strategy, it was realized that utilization of
ornithine could result in retention of the chiral amine group and stereochemical guidance
179
by the same group (Scheme V.4). This strategy eliminates the need for introducing
chirality at any point in the synthesis. In order to maximize the flexibility of the
route that employs the exact chemistry used in the biosynthesis. For example, Koomen
lupinamine (Scheme V.5).26 If the same strictly biosynthetic approach was applied to
naturally occurring chiral molecule and using it in a biomimetic manner, the synthesis
excellent de (13). The relative stereochemistry of this product was confirmed by an X-ray
crystal structure of the Cbz-protected quinolizidine after the second cyclization (Figure
V.3). This result may be explained by an exchange of ligands on the boron atom; i.e., the
chelation model. If the oxygen atom of the Boc group becomes one of the ligands, the
tethered cyanoborohydride would transfer its hydride from the Si face (when R and R' are
intermediate were devised (Schemes IV.6 and IV.7). One route features addition of vinyl
borylate to the acyliminium ion27 (from 13) and the other route centers around a hetero
Diels-Alder reaction28 with the acyliminum ion (from 22) as the dienophile. Despite some
180
of the promising precedents in the literature, these approaches proved unsuccessful due to
give the anti configuration.29,30 Using a similar synthetic logic as applied to the first
attempted synthesis (Scheme V.1), the first piperidine ring can be synthesized via
cyclization through SN2 displacement as shown in Scheme V.7. The substrate for the SN2
reaction can be easily derived from a corresponding amino alcohol, such as 30, which
available ornithine derivative 10, which was first converted smoothly to the Weinreb
amide 29 using a common coupling condition (Scheme V.8).31,32 After simple washings,
the product was fairly pure and required no further purification. Upon treatment with an
182
V.10).30,34 By 1H and 13
C NMR, none of the other diastereomer was observed in the
alcohol proceeded smoothly with an excellent yield, and again the product was
crystalline. Therefore, the synthesis of the SN2 substrate 31 was accomplished very
shown in Scheme V.8. Removal of the Boc group in TFA/CH 2Cl2 followed by
acetonitrile yielded the diallyl piperidine 27 in good yield.36,37,38,39,40 The synthesis of the
using the Grubbs second-generation catalyst.41,42 It is noteworthy that this was one of the
However, the purification of this rather polar RCM product presented difficulties.45 The
problem was overcome when Cho and Kim’s method was employed with a small
was observed when this process was separated into two steps as shown in Scheme V.8 (9
steps overall).11 By 1H and 13C NMR, a single isomer was observed. This material had the
identical 1H and 13C NMR and IR spectra and HR mass (EI) to those of the authentic
material.7 The reproducible overall yields of (+)-9 beginning from 10 were 38% and 28%
184
for the longer and shorter sequences, respectively. The specific optical rotation value also
matched that of the material from the previous total synthesis ((+)-9 [α]22D = +24° (c
0.10, CHCl3)).50 Only three chromatographic purification steps were required throughout
the synthesis. The synthesis of the other enantiomer was performed in the same manner
was accomplished in two weeks, demonstrating the practicality of this synthetic route
9 (12-15 steps),11-17 efficiency of this synthesis is outstanding. This is largely due to the
With both of the enantiomers of epiquinamide in hand, the next goal was to
dichroism spectrum of the authentic natural product was ever reported.7 In the absence of
access to the natural sample for comparison in chiral chromatography with the two
compare the biological activity of the two isomers. If one was as active as the reported
natural product, but not the other stereoisomer, then it can be concluded that the former
However, neither of the compounds were active in the in-house Na+ channel
activation and blocking activity assays using neuroblastoma cells (Neuro 2a),52 the
cytotoxicity assays using human lung cancer cells (H460), or the brine shrimp assay.
185
was inactive in their competitive binding assays using [3H]epibatidine.53 With all of these
results combined, it was considered highly likely that the active compound in the original
sample from the frogs was not epiquinamide, but some other exceptionally bioactive
contaminant.
Two years later, Rutjes and coworkers compared their synthetic standards of both
enantiomers of epiquinamide with the authentic sample and determined that the natural
and coworkers reported GCMS evidence that the observed bioactivity of epiquinamide
product; the total synthesis helps to confirm the identity of the bioactive metabolite.
“While these data are somewhat embarrassing to report, they are not
entirely unusual in the isolation of natural products. Further, the results are
a reminder that the presence of multiple active substances in an extract (in
this case compounded by small amounts of sample available) can
sometimes lead to erroneous results from cross-contamination that are not
always evident a priori. This underscores the importance of obtaining
synthetic material whenever possible to corroborate both structure and
pharmacology. In this case, we were able to produce a correct structure,
but were misled by contamination in the pharmacology. Often with more
complex structures, the converse occurs. Thus,it is important that the
pharmacognosist work in collaboration with the synthetic/medicinal
chemist to produce data that are unequivocal.”
Fitch et al, 200919
V.6 Conclusion
unfortunate realization that the structure 9 does not possess the reported bioactivity
further confirms the contemporary need for total synthesis of novel natural products.
General
Unless noted otherwise, all materials were purchased from commercially available
sources and were used without further purification. Anhydrous dichloromethane and
diethyl ether were purchased from VWR. All reactions were carried out under dry argon
atmosphere unless otherwise noted. Flash chromatorgraphy was performed using Fisher
silica gel (230-400 mesh). TLC was performed using EM Science pre-coated silica gel
device and are uncorrected. Optical rotations were measured on a Rudolph Research
Autopol III polarimeter and a Jasco P1010 polarimeter. IR spectra were recorded on a
Nicolet Magna-IR 550. For 11-16, 1H NMR and 13C NMR spectra were recorded on a
Bruker spectrometer (300 MHz and 75 MHz, respectively). For 27-32 and 9, 1H NMR
and 13C NMR spectra were recorded on a Varian Inova spectrometer (500MHz) or on a
recorded in CDCl3 were referenced to the residual solvent peaks at 7.26 ppm and 77.0
MAT900XL spectrometer.
187
(S)-methyl 12,12-dimethyl-3,10-dioxo-1-phenyl-2,11-dioxa-4,9-diazatridecane-5-
carboxylate (11). To αN,δN-protected ornithine 10 (718 mg, 1.96 mmol) in CH2Cl2 (10
mL) at 0 ºC were added anhydrous MeOH (300 μL, 7.4 mmol), DMAP (43 mg, 0.35
mmol), EDC·HCl (403 mg, 2.10 mmol), and Et 3N (280 μL, 2.0 mmol) successively. After
the reaction mixture was warmed to rt over night under stirring, the volatiles were
removed under vacuum. The residues were resuspended in Et2O. The solution was
washed with 0.5 M HCl, 1M NaHCO3, and brine successively. Upon drying over
Na2SO4, the solution was passed through a silica gel plug (1:1 EtOAc / Hex) and the
solvents were removed under vacuum. Colorless gummy substance (11; 703 mg, 1.85
mmol, 94%) was obtained, and it was pure enough for the subsequent reactions. TLC Rf
= 0.43 (1:1 EtOAc / Hex). 1H and 13C NMR spectra: see Figures V.4 and 5.
mg, 3.68 mmol) in THF (7 mL) at −78 ºC was added 1.0 M vinyl magnesium bromide
solution in THF (50 mL) over ~7 min. The solution of 11 (3.14 g, 8.25 mmol) in THF (30
mL) was cannulated in over 5 min. The reaction mixture was gradually warmed to rt over
188
night under stirring. The mixture turned black. After cooling to 0 ºC, the reaction was
quenched by addition of 1M HCl (200 mL) and the mixture was stirred for 15 min. Then
the mixture was extracted with Et2O thrice. The combined organic layer was washed with
1M NaHCO3, water, and brine successively. Upon drying over Na2SO4, the solution was
concentrated under vacuum and the residues were subjected to flash column
chromatography to obtain off-white solid (12; 3.056 g, 7.56 mmol, 92%). TLC Rf = 0.57
(1:1 EtOAc / Hex). 1H and 13C NMR spectra: see Figures V.6 and 7.
ketone 12 (1.72 g, 4.27 mmol) in CH2Cl2 (25 mL) at rt was added TFA (25 mL). The
solution was stirred at rt for 1 hr and the volatiles were removed under vacuum. The oily
residues were dissolved in i-PrOH (40 mL) and the solution was cooled to 0 ºC. NaOH
solution in EtOH was added to adjust the pH to ~3. NaBH3CN (1.38 g, 22.0 mmol) was
added in two portions and the solution was warmed to rt over night under stirring. Then
1M HCl (30 mL) was added to quench the reaction. Evolution of H 2 gas was observed.
The volatiles were removed under vacuum. To the residues were added saturated
NaHCO3 and the aqueous layer was extracted with CH2Cl2 twice and CHCl3 twice. The
combined organic layer was dried over Na2SO4. Upon removal of the solvents under
vacuum, the residues were subjected to flash column chromatography using silica gel
189
(0.5% Et3N buffer, 1:49 MeOH / CH2Cl2 to 1:4 MeOH / CH2Cl2) to obtain the product
(13; 325 mg, 1.13 mmol, 26% over two steps). 1H and 13C NMR spectra: see Figures V.8
and 9.
details. 1H and 13C NMR spectra: see pages Figures V.10 and 11.
amino alcohol 14 (56 mg, 0.183 mmol) in CH2Cl2 were added CBr4 (91 mg, 0.27 mmol)
and PPh3 (120 mg, 0.46 mmol) successively at 0 ºC under stirring. The solution was
warmed to rt quickly after 30 min, and it was kept at rt for 2.5 hrs. Then Et3N (64 μL,
0.46 mmol) was added, which made the solution turn yellow. The solution was left
stirring at rt for over night. Upon quenching the reaction with 1M HCl (30 mL), the
organic phase was washed with 1M HCl (20 mL x 3). The combined aqueous phase was
basified with solid NaOH, and it was extracted with CH2Cl2 (30 mL x 4). Upon drying
over K2CO3 and removal of the solvent under vacuum, the residues were subjected to
190
flash column chromatography using silica gel (1:9 MeOH / CH2Cl2). However, the
product was still contaminated with phosphine compounds and it was further purified by
semi-prep HPLC using a Synergi 4μ Hydro-RP 80A (250 x 10 mm) column (1.2 mL /
min H2O to 1.5 mL / min 5:3 MeCN / H 2O) with acetic acid buffer (0.1%). The product
(15) had a retention time of 6.02 min (31 mg, 0.11 mmol, 59%). 1H and 13C NMR spectra:
9-epi-epiquinamide (16). No experimental details. 1H and 13C NMR spectra: see Figures
(2S)-2-benzyloxycarbonylamino-5-tert-butyloxycarbonylamino-(N-methoxy-N-
hydrochloride (2.180 g, 22.35 mmol), Et3N (3.78 mL, 26.90 mmol), (3-dimethylamino-
pyridine (240 mg, 1.96 mmol) sequentially. The resulting solution was left stirring to
warm up to room temperature over night (18 hours). The volatiles were removed on a
191
rotary evaporator and the residue was suspended in ethyl acetate (200 mL). The insoluble
solids were filtered off and rinsed with EtAcO. The filtrate was sequentially washed with
0.2 M HCl, water, 1.0 M NaHCO3, water, and brine and was dried over MgSO4. After
filtration using Celite®, the solvents were removed under vacuum and a colorless glass
(29) was obtained and was used without purification for the next step. (7.564 g, 18.47
mmol, 93%). TLC Rf = 0.30 EtAcO/hexane (3:2). [α]D24 –6.3º (c 1.20, CHCl3). IR (film)
3332, 3080, 3024, 2975, 2936, 1709, 1654, 1530, 1450, 1398, 1367, 1250, 1160 cm -1. 1H
NMR (500 MHz, CDCl3) δ 7.26-7.36 (m, 5H), 5.53 (d, J = 9.0 Hz, 1H), 5.08 (m, 2H),
4.73 (m, 1H), 4.57 (m, 1H), 3.77 (s, 3H), 3.20 (s, 3H), 3.12 (m, 2H), 1.70-1.80 (m, 2H),
1.49-1.61 (m, 2H), 1.41 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 172.3, 156.1, 155.8, 136.2,
128.4, 128.0, 127.9, 79.0, 66.8, 61.5, 50.6, 40.0, 30.0, 28.3, 25.8. HRMS (EI): calcd for
(5S)-2-benzyloxycarbonylamino-5-tert-butyloxycarbonylamino-4-oxo-1-octene (30).
To the solution of allylmagnesium bromide in Et2O (1.0 M, 100 mL, 100 mmol) at -78 ºC
was added the solution of the Weinreb amide 29 (9.060 g, 22.13 mmol) in Et2O (135 mL)
dropwise over 1 hour while vigorously stirring. The resulting slurry mixture was warmed
to room temperature and stirred for 30 min. Upon cooling to -10 ºC, aqueous HCl
solution (1.0 M, 125 mL) was added carefully and the mixture was stirred for 5 min at
-10 ºC. After partitioning, the aqueous layer was extracted with Et 2O (30 mL × 3) and the
192
combined organic layer was washed sequentially with cold 0.1 M HCl, cold water, cold
1.0 M NaHCO3, and brine, and then was dried over MgSO4. Upon filtration through
Celite®, the solvents were removed under vacuum. The residue was dissolved in a
mixture of Et2O and CH2Cl2, and hexane was added and the product was re-crystallized
(7.138 g, white needles). The filtrate was concentrated under vacuum and was purified by
flash chromatography (EtAcO / hexane 1:9 to 1:4). The separation of the double bond-
isomerized products from the desired product was difficult. The product (30) was a white
solid (total 7,420 g, 19.00 mmol, 86%). TLC Rf = 0.46 EtAcO / hexane (2:3). Mp 85~86
ºC. [α]D24 +46º (c 0.34, CHCl3). IR (film) 3332, 3080, 3025, 2975, 2936, 1693, 1639,
1525, 1452, 1250, 1168 cm-1. 1H NMR (500MHz, CDCl3) δ 7.29-7.37 (m, 5H), 5.82-5.96
(m, 1H), 5.55 (d, J = 7.2 Hz, 1H), 5.23-5.19 (m, J =11.0 Hz, 1H), 5.13-5.09 (m, J = 11.0
Hz, 1H), 5.09 (s, 2H), 4.57 (bs, 2H), 4.44-4.50 (m, 1H), 3.28 (d, J = 6.6 Hz, 2H),
3.09-3.17 (m, 2H), 1.86-1.97 (m, 2H), 1.46-1.59 (m, 2H), 1.43 (s, 9H). 13C NMR (75
MHz, CDCl3) δ 206.5, 156.0,156.0, 136.2, 129.5, 128.5, 128.2, 128.1, 119.5, 79.3, 67.0,
59.1, 59.0, 44.5, 39.9, 28.6, 28.4, 25.8. HRMS (EI): calcd for C21H31N2O5 (M + H):
(4R,5S)-5-benzyloxycarbonylamino-8-tert-butyloxycarbonylamino-4-hydroxy-1-
octene (28). To the suspension of LiAl(Ot-Bu)3H (10.25 g, 40.31 mmol) in EtOH (70
mL) stirring at -78 ºC was added ketone 30 (3.952 g, 10.12 mmol) in EtOH (70 mL)
193
dropwise over 30 min. After stirring for 5 hours at -70 to -78º C, 10% citric acid (100
mL) was added, and the mixture was stirred for 20 min. This mixture was extracted with
CH2Cl2 four times, and the combined organics were washed with water, saturated
NaHCO3, brine and dried over Na2SO4. Upon concentration under vacuum, the mixture
was dissolved in minimal amount of CH2Cl2 / Et2O (1:2) and small amount of hexane was
added. After slow cooling to -20 ºC, crystals were washed with Et2O / hexane (1:1) and
then dried under vacuum (3.516 g). The filtrate containing the product was purified by
flash chromatography (EtAcO / hexane 1:9 to 1:1) after concentration under vacuum (28;
total 3.703g, 9.435 mmol, 93%). TLC Rf = 0.33 EtAcO / hexane (1:1). Mp 106 ºC. [α]D24
+22º (c 0.37, CHCl3). IR (film) 3348, 3317, 3068, 2944, 1689, 1538, 1452, 1260, 1160.
1
H NMR (300 MHz, CDCl3) δ 7.26-7.38 (m, 5H), 5.74-5.88 (m, 1H), 5.16 (m, 2H),
5.05-5.12 (m, 2H), 5.09 (s, 2H), 4.57 (m, 1H), 3.67 (m, 2H), 3.11 (m, 2H), 2.11-2.32 (m,
2H), 1.51-1.70 (m, 2H), 1.28-1.51 (m, 11H). 13C NMR (300 MHz, CDCl3) δ 156.6, 156.1,
136.4, 134.4, 128.5, 128.1, 128.1, 118.3, 79.2, 73.2, 66.8, 55.0, 40.3, 38.3, 28.4, 26.8,
26.2. HRMS (EI): calcd for C21H33N2O5 (M + H): 393.2384, found 393.2375.
(4R,5S)-5-benzyloxycarbonylamino-8-tert-butyloxycarbonylamino-4-
10.91 mmol) in CH2Cl2 (320 mL) stirring at -15 ºC was added Et3N (3.13 mL, 22.3
mmol). Methanesulfonyl chloride (1.27 mL, 16.4 mmol) was added dropwise, and then 4-
194
dimethylaminopyridine (702 mg, 5.75 mmol) was added. The solution was left stirring at
temperature ranging from -15 ºC to 0 ºC for 4 hours. Cold water (100 mL) was then
added rapidly at 0 ºC. The mixture was extracted with CH2Cl2 three times and the
combined organics were washed with cold 1.0 M HCl, cold water, cold NaHCO3, and
dried over Na2SO4. Upon removal of the solvents under vacuum, the product was
recrystallized from CH2Cl2 / Et2O / hexane system as white crystals (4.838 g, 10.28
mmol). The residual product in the filtrate was purified by flash chromatography
(EtAcO / hexane 1:9 to 2:3) and white solid (31) was obtained (total 4.978 g, 10.58
mmol, 97 % yield). TLC Rf = 0.70 EtAcO / hexane 1:1. Mp 67ºC. [α] D24 –27º (c 0.50,
CHCl3). IR (film) 3402, 2983, 2944, 1689, 1639, 1525, 1452, 1365, 1339, 1254, 1173,
913 cm-1. 1H NMR (500 MHz, CDCl3) δ 7.26-7.36 (m, 5H), 5.74-5.85 (m, 1H), 5.25 (m,
1H), 5.18 (d, J = 9.5 Hz, 1H), 5.16 (m, 1H), 5.10 (s, 2H), 4.83 (m, 1H), 4.58 (m, 1H),
3.88 (m, 1H), 3.11 (m, 2H), 2.98 (s, 3H), 2.46-2.54 (m, 1H), 2.37-45 (m, 1H), 1.54-1.66
(m, 2H), 1.27-1.49 (m, 11). 13C NMR (75 MHz, CDCl3) δ 156.1, 156.0, 136.4, 132.3,
128.4, 128.0, 127.9, 119.2, 83.5, 79.2, 66.8, 53.0, 39.9, 38.5, 36.1, 28.3, 26.7, 25.4.
the mesylate 31 (2.734 g, 5.809 mmol) in CH2Cl2 (43 mL) at 0ºC was added
195
trifluoroacetic acid (11 mL) dropwise over 5 min while stirring. Immediately the reaction
mixture was warmed to room temperature and was stirred for 45 min. Upon cooling the
reaction mixture to 0ºC again and diluting the mixture with CH 2Cl2 (100 mL), 2M K2CO3
solution (100 mL) was added carefully. This mixture was partitioned and the aqueous
phase was extracted with CH2Cl2 (50 mL × 4). The combined organic phase was dried
over anhydrous K2CO3 and was filtered through celite. The solvents were removed on a
rotary evaporator. The residues were then dissolved in CH3CN (450 mL), and K2CO3
(4.20 g, 30.4 mmol) was added in two portions over 3 hours. After stirring the mixture for
24 hours, the mixture was gradually heated up to 70ºC over 1.5 hours. The consumption
of the primary amine was confirmed on TLC, and the mixture was filtered through celite
and was concentrated in vacuum to ca. 90 mL. After the addition of K 2CO3 (1.78g, 12.9
mmol), the mixture was stirred for 30 min at room temperature. Then allylbromide
(800μL, 9.19 mmol) and additional K2CO3 (0.98 g, 7.1 mmol) were added and the
mixture was left stirring at room temperature for 30 hours. Additional allylbromide
(200μL, 2.30 mmol) and K2CO3 (0.54 g, 3.9 mmol) was added and the mixture was
stirred at room temperature for 4 hours. Upon filtration of the mixture, the solvents were
removed on a rotary evaporator. To the residues were added Et 2O (200 mL) and saturated
NaHCO3 (150 mL). After partitioning the phases, the aqueous phase was extracted with
Et2O (50 mL × 4). The combined organic phases were dried over Na2SO4, and filtered
through Celite®. After the removal of the solvents under vacuum, the residue was purified
by flash chromatography (EtAcO / hexane, 1:19 to 1:4), yielding a colorless oil (27;
1.382 g, 4.395 mmol, 76 % over 2 steps). TLC Rf = 0.45 EtAcO / hexane 1:1. [α] D24 =
+16º (c 0.23, CHCl3). IR (film) 3427, 3078, 3029, 2942, 2864, 2811,1720, 1640, 1500,
196
1456, 1341, 1214, 919 cm-1. 1H NMR (300 MHz, CDCl3) δ 7.28-7.38 (m, 5H), 5.86-5.71
(m, 2H), 5.48 (m, 1H), 5.16-4.96 (m, 7H), 3.84 (m, 1H), 3.29 (m, 1H), 2.92 (m, 1H), 2.80
(m, 1H), 2.35-2.46 (m, 2H), 2.15-2.03 (m, 2H), 1.84 (m, 1H), 1.64 (m, 1H), 1.47 (m, 1H),
1.39 (m, 1H). 13C NMR (75 MHz, CDCl3) δ 155.5, 136.7, 134.9, 134.4, 128.3, 127.9,
127.9, 117.6, 117.2, 66.3, 63.0, 56.3, 52.6, 48.1, 33.9, 29.5, 20.8. Both 1H and 13C NMR
spectra contain rotamer signals.55 HRMS (EI): calcd for C19H26N2O2 (M): 314.1989,
found 314.1994.
of the 1,2-diallyl piperazine 27 (1.131 g, 3.597 mmol) in CH2Cl2 (360 mL) was added
Grubb’s 2nd generation catalyst (269 mg, 0.360 mmol) in two portions over 2 hours. The
mixture was brought to reflux slowly and the temperature was reduced to room
temperature after 2 hours for the second addition of the catalyst. Then the mixture was
refluxed again for 5 hours (while monitoring by TLC) and cooled again to room
temperature. Silica gel (3.0 g) was added and the mixture was stirred for 10 min. The
solids were filtered through Celite® and rinsed with CH2Cl2 thoroughly. To the filtrate was
added activated charcoal (ca. 20 g) and the mixture was left stirring at room temperature
over night. Upon filtration through Celite®, the filtrate was concentrated on a rotary
evaporator. The residue was purified by flash chromatography on silica gel twice to yield
197
transparent brown oil still contaminated with a miniscule amount of the ruthenium
catalyst (32; 855 mg, 2.986 mmol, ). TLC Rf = 0.24 CH2Cl2 / MeOH 49:1. [α]D24 –40º (c
0.33 CHCl3). IR (film) 3431, 3055, 2960, 2807, 2758, 1718, 1644, 1500, 1337, 1217,
1095 cm-1. 1H NMR (500 MHz, CDCl3) δ 7.28-7.35 (m, 5H), 5.68-5.71 (m, 1H), 5.57 (m,
1H), 5.10 (s, 2H), 3.77 (dd, J = 9.2, 2.5 Hz, 1H), 3.19 (m, J = 16.4 Hz, 1H), 2.89 (dd, J =
11.2, 2.2 Hz, 1H), 2.68 (m, J = 16.4 Hz, 1H), 2.35 (m, J = 8.3 Hz, 1H), 2.19-2.26 (m,
1H), 2.03 (td, J = 12.2, 3.1 Hz, 1H), 1.93 (m, 1H), 1.90 (m, 1H), 1.75 (m, J = 12.9, 3.8
Hz, 1H), 1.54 (m, J = 14.6 Hz, 1H), 1.49 (dt, J = 13.3, 3.7 Hz, 1H). 13C NMR (75 MHz,
CDCl3) δ 156.2, 136.7, 128.4, 127.9, 127.8, 124.2, 123.4, 66.4, 58.9, 56.1, 54.3, 49.1,
29.8, 27.8, 20.3. Both 1H and 13C NMR spectra contain rotamer signals.54 HRMS (EI):
mg, 0.698 mmol) in EtOH (3.7 mL) was added Pd/C (10% wt, 180 mg) carefully. Argon
was evacuated and H2 was bubbled into the mixture for 5 min. Then the reaction mixture
was left stirring under H2 atmosphere over night using balloon pressure. Acetic anhydride
(170 μL, 1.80 mmol) was added in three portions over 1 hour. The reaction mixture was
left stirring again over night and then was filtered through Celite® and the filter case was
rinsed with ethanol. Upon concentration, the residue was purified by flash
yield pale yellow solid (9; 87 mg, 0.44 mmol, 63 % yield). Rf = 0.39 CH2Cl2 / MeOH /
28% NH4OH(aq) 90:9:1. Mp 130ºC. [α]D24 +24º (c 0.10, CDCl3). IR (film) 3427, 2949,
2862, 2814, 2766, 1637, 1540, 1376, 1295, 1211 cm-1. 1H NMR (300 MHz, CDCl3) δ
6.24 (bs, 1H), 3.92 (m, J = 8.8, 2.0 Hz, 1H), 2.78 (m, 2H), 2.01 (s, 3H), 1.94 (m, 3H),
1.85 (m, J = 13.4 Hz, 1H), 1.72 (m, J = 13.2 Hz, 2H), 1.19-1.63 (m, 7H). 13C NMR (75
MHz, CDCl3) δ 169.6, 64.4, 56.7, 56.6, 48.0, 29.5, 28.9, 25.4, 23.9, 23.4, 20.4. HRMS
1.01 mmol) in EtOH (3.0 mL) was added Pd/C (10% wt, 100 mg) carefully. Argon was
evacuated and H2 was bubbled into the mixture for 5 min. Then the reaction mixture was
left stirring under H2 atmosphere over night using balloon pressure. The reaction mixture
was left stirring again over night and then was filtered through Celite ® and the filter cake
was rinsed with ethanol. The solvent was removed under vacuum, and the residues were
dissolved in dioxane (10 mL). While stirring, aqueous solution of NaOH (1.0 M, 10 mL)
was added slowly at room temperature. After 2 hours, the reaction mixture was extracted
with CH2Cl2 (20 mL × 4). The organic layer was dried over Na2SO4 and was concentrated
under vacuum. The residue was purified by flash chromatography on silica gel (CH2Cl2 /
MeOH / 28% NH4OH(aq) 99:0.9:0.1 to 90:9:1) to yield pale yellow solid (9; 173 mg,
(-)-epiquinamide. TLC Rf = 0.39 CH2Cl2 / MeOH / 28% NH4OH(aq) 90:9:1. [α]D24 –22º
(c 0.13, CHCl3). . 1H and 13C NMR: identical to (+)-9. HRMS (EI): calcd for C11H21N2O1
NMR Spectra
13
Figure V.32:34:
Figure
Figure 33:C13C
NMR
1H spectrum
NMR
NMR of (-)-9
spectrum
spectrum of in CDCl3
of (-)-9
(+)-9
229
2 Daly, J. W.; Spande, T. F.; Garraffo, H. M. Alkaloids from amphibian skin: a tabulation of over
eight-hundred compounds. J. Nat. Prod. 2005, 68, 1556-1575.
3 Daly, J. W. Nicotinic agonists, antagonists, and modulators from natural sources. Cell. Mol.
Neurobiol. 2005, 25, 513-552.
4 Daly, J. W. Marine toxins and nonmarine toxins: convergence or symbiotic organisms? J. Nat.
Prod. 2004, 67, 1211-1215.
5 Simmons, T. L.; Coates, R. C.; Clark, B. R.; Engene, N.; Gonzalez, D.; Esquenazi, E.;
Dorrestein, P. C.; Gerwick, W. H. Biosynthetic origin of natural products isolated from marine
microorganism–invertebrate assemblages. Proc. Natl. Acad. Sci. 2008, 105, 4587–4594.
6 Spande, T. F.; Garraffo, H. M.; Edwards, M. W.; Yeh, H. J. C.; Pannell, L.; Daly, J. W.
Epibatidine: a novel (chloropyridyl)azabicycloheptane with potent analgesic activity from an
Ecuadoran poison frog. J. Am. Chem. Soc. 1992, 114, 3475-3478.
7 Fitch, R. W.; Garraffo, H. M.; Spande, T. F.; Yeh, H. J. C.; Daly, J. W. Bioassay-guided
isolation of epiquinamide, a novel quinolizidine alkaloid and nicotinic agonist from an
Ecuadoran poison frog, Epipedobates tricolor. J. Nat. Prod. 2003, 66, 1345-1350.
8 Baker, D. D.; Alvi, K. A. Small-molecule natural products: new structures, new activities.
Curr. Opin. Biotechnol. 2004, 15, 576-583.
9 Gotti, C.; Riganti, L.; Vailati, S.; Clementi, F. Brain Neuronal Nicotinic Receptors as New
Targets for Drug Discovery. Curr. Pharm.Design 2006, 12, 407-428.
11 Wijdeven, M. A.; Botman, P. N. M.; Wijtmans, R.; Schoemaker, H. E.; Rutjes, F. P. J. T.;
Blaauw, R. H. Total Synthesis of (+)- Epiquinamide. Org. Lett. 2005, 7, 4005-4007.
13 Wijdeven, M. A.; Wijtmans, R.; van den Berg, R. J. F.; Noorduin, W.; Schoemaker, H. E.;
Sonke, T.; van Delft, F. L.; Blaauw, R. H.; Fitch, R. W.; Spande, T. F.; Daly, J. W.; Rutjes, F. P.
J. T. N,N-Acetals as N-Acyliminium Ion Precursors: Synthesis and Absolute Stereochemistry
of Epiquinamide. Org. Lett. 2008, 10, 4001-4003.
14 Ghosh, S.; Shashidhar, J. Total synthesis of (+)- epiquinamide from -mannitol. Tetrahedron
Lett. 2009, 50, 1177-1179.
230
15 Huang, P.-Q.; Guo, Z.-Q.; Ruan, Y.-P. A Versatile Approach for the Asymmetric Syntheses of
(1R,9aR)- Epiquinamide and (1R,9aR)-Homopumiliotoxin 223G. Org. Lett. 2006, 8,
1435-1438.
16 Voituriez, A.; Ferreira, F.; Perez-Luna, A.; Chemla, F. Asymmetric synthesis of (-)-1-
hydroxyquinolizidinone, a common intermediate for the syntheses of (-)-homopumiliotoxin
223G and (-)- epiquinamide. Org. Lett. 2007, 9, 4705-4708.
17 Tong, S.; Barker, D. A concise synthesis of (±) and a total synthesis of (+)- epiquinamide.
Tetrahedron Lett. 2006, 47, 5017-5020.
18 Kanakubo, A.; Gray, D.; Innocent, N.; Wonnacott, S.; Gallagher, T. The synthesis and nicotinic
binding activity of (± )- epiquinamide and (± )-C(1)-epiepiquinamide. Bioorg. Med. Chem.
Lett. 2006, 16, 4648-4651.
19 Fitch, R. W.; Sturgeon, G. D.; Patel, S. R.; Spande, T. F.; Garraffo, H. M.; Daly, J. W.; Blaauw,
R. H. Epiquinamide : A Poison That Wasn't from a Frog That Was. J. Nat. Prod. 2009, 72,
243-247.
20 Pal, K.; Behnke, M. L.; Tong, L. A general stereocontrolled synthesis of cis-2,3 disubstituted
pyrrolidines and piperidines. Tetrahedron Lett. 1993, 34, 6205-6208.
23 (a) Anh, N. T.; Eisenstein, O. Nouv. J. Chim. 1977, 1, 61-70. (b) Lodge, E. P.; Heathcock, C.
H. Acyclic stereoselection. 40. Steric effects, as well as .sigma.*-orbital energies, are
important in diastereoface differentiation in additions to chiral aldehydes. J. Am. Chem. Soc.
1987, 109, 3353-3361.
24 The corresponding transition states were modeled at semi-empirical (PM3) level and the
relative energy difference between the two transition states leading to the two stereoisomers
was calculated. Based on this result, Si-face addition of hydride was predicted (data not
shown).
27 Batey, R. A.; MacKay, D. B.; Santhakumar, V. Alkenyl and aryl boronates – mild
nucleophiles for the stereoselective formation of functionalized N-heterocycles. J. Am. Chem.
Soc. 1999, 121, 5075-5076.
29 Guarna, A.; Occhiato, E. G.; Machetti, F.; Scarpi, D. A Concise route to 19-nor-10-azasteroids,
a new class of steroid 5α-reductase Inhibitors. Synthesis of (+)-19-nor-10-azatestosterone and
(+)-17β-(acetyloxy)-(5β)-10-azaestr-1-en-3-one. J. Org. Chem. 1998, 4111-4115.
30 So, R. C.; Ndonye, R.; Izmirian, D. P.; Richardson, S. K.; Guerrera, R. L.; Howell, A. R.
Straightforward synthesis of sphinganines via a serine-derived Weinreb amide. J. Org. Chem.
2004, 69, 3233-3235.
35 The bridgehead proton of 9 has a chemical shift of δ 3.91 ppm, and that of 15 has a shift of δ
3.95 ppm in MeOH-d4. Otherwise, the two compounds have exactly the same carbon
framework and molecular weight by NMR and LRMS.
36 Abe, H.; Aoyagi, S.; Kibayashi, C. Total synthesis of the tricyclic marine alkaloids (-)-
lepadiformine, (+)-cylindricine C, and (-)-fasicularin via a common intermediate formed by
formic acid-induced intramolecular conjugate azaspirocyclization. J. Am. Chem. Soc. 2005,
127, 1473-1480.
39 Sundberg, R. J.; Amat, M.; Fernando, A. M. Analogs of the iboga alkaloids. Synthesis and
reactions of (.+-.)-15-oxo-20-deethylcoronaridine derivatives. J. Org. Chem. 1987, 52,
3151-3159.
40 Kinderman, S. S.; Wekking, M. M. T.; van Maarseveen, J. H.; Schoemaker, H. E.; Hiemstra,
H.; Rutjes, F. P. J. T. A stereodivergent approach to substituted 4-hydroxypiperidines. J. Org.
232
41 Scholl, M.; Ding, S.; Lee, C. W.; Grubbs, R. H. Synthesis and activity of a new generation of
ruthenium-based olefin metathesis catalysts coordinated with 1,3-dimesityl-4,5-
dihydroimidazol-2-ylidene ligands. Org. Lett. 1999, 1, 953-956.
42 Nicolaou, K. C.; Bulger, P. G.; Sarlah, D. Metathesis reactions in organic synthesis. Angew.
Chem. Int. Ed. 2005, 44, 4490-4527.
43 Hong, S. H.; Grubbs, R. H. Highly active water-soluble olefin metathesis catalyst. J. Am.
Chem. Soc. 2006, 128, 3508-3509.
46 Published methods for the removal of the ruthenium catalyst are better suited for nonpolar
compounds. For examples, see: (a) Maynard, H. D.; Grubbs, R. H. Purification technique for
the removal of ruthenium from olefin metathesis reaction products. Tetrahedron Lett. 1999,
40, 4137-4140. (b) Haack, K. L.; Ahn, Y. M.; Georg, G. I. A convenient method to remove
ruthenium byproducts from olefin metathesis reactions using polymer-bound
triphenylphosphine oxide (TPPO). Mol. Diversity 2005, 9, 301-303.
47 Amir H. Hoveyda, Dennis G. Gillingham, Joshua J. Van Veldhuizen, Osamu Kataoka, Steven
B. Garber, Jason S. Kingsbury and Joseph P. A. Harrity Ru complexes bearing bidentate
carbenes: from innocent curiosity to uniquely effective catalysts for olefin metathesis. Org.
Biolmol. Chem. 2004, 2, 8-23.
48 Michrowska, A; Grela, K. Quest for the ideal olefin metathesis catalyst. Pure Appl. Chem.
2008, 80, 31–43.
49 Shinada, T.; Hayashi, K.; Yoshida, Y.; Ohfune, Y. Squaric acid derivatives prevent the removal
of N-Cbz and N-Fmoc groups under catalytic hydrogenation reaction. Synlett 2000, 10,
1506-1508.
50 The literature value of (+)-9 was [α]20D = +28° (c 0.23, CHCl3) (see ref 11) and [α]22D = +26.2°
(c 0.23, CHCl3) (see ref 17).
51 The literature value of (-)-9 was [α]20D = - 25° (c 0.26, CHCl3) (see ref 15).
52 LePage, K. T.; Dickey, R. W.; Gerwick, W. H.; Jester, E. L.; Murray, T. F. On the use of
neuro-2a neuroblastoma cells versus intact neurons in primary culture for neurotoxicity
studies. Critical Reviews in Neurobiology 2005, 17, 27-50.
53 Kanakubo, A.; Gray, D.; Innocent, N.; Wonnacott, S.; Gallagher, T. The synthesis and nicotinic
binding activity of (±)-epiquinamide and (±)-C(1)-epiepiquinamide. Bioorg. Med. Chem. Lett.
233
54 Moraczewski, A. L.; Banazsynski, L. A.; From, A. M.; White, C. E.; Smith, B. D. Using
hydrogen bonding to control carbamate C−N rotamer equilibria. J. Org. Chem. 1998, 63,
7258-7262.
Chapter VI
234
235
Abstract
endothelial cells. Due to its promising bioactivity profile, total synthesis of this natural
enamide and disulfide. After an initial successful route was developed, the total synthesis
was revised to be suitable for larger-scale synthesis to provide for in vivo evaluations and
analog production.
236
VI.1 Introduction
mixed assemblage of Schizothrix and Lyngbya majuscula collected in Somo Somo, Fiji.3
The initial report by Nogle and Gerwick cited its activity against Neuro-2a neuroblastoma
human umbilical vein endothelial cells (HUVEC) with an IC50 of 500 fM.4 HUVEC cells
are primary endothelial cells derived from human tissues and are a good model for
in the tissues surrounding a tumor. Growing tumors need an even greater supply of
nutrients from the blood than normal tissue. Therefore, a compound that inhibits the
Vascular endothelial growth factor (VEGF) is a key growth factor that is involved
particular area indicates active formation of blood vessels. The use of transgenic
zebrafish embryos, which are transparent, allows direct observation of a drug's effects on
angiogensis through the visualization of green fluorescent protein (GFP) that is placed
under the control of a VEGF promotor.6 In such a study by our collaborators at the
Moores Cancer Center, somocystinamide A (1) was found to significantly reduce the
level of GFP, and thus indicate a reduced level of VEGF activity present in zebrafish
embryos even at low drug concentrations (<0.1 μM, the media concentration).4 Equally
significant, the zebrafish were not adversely affected by the compound even at 1.6 μM,4
apoptosis is the most effective mechanism. This is because certain forms of apoptotic
contact with the corresponding receptor on the cell surface.7 Therefore, when extremely
110
IC50
90 WG-144 (NB-7 cell) 8.369e-007
% Cell proliferation
WG-144(NB-7-c-8 cell) 1.589e-008
70
50
30
10
-10
-12.5 -10.0 -7.5 -5.0 -2.5 0.0
Log Drug Concentration
Cell Membrane
Cell membrane
Nucleus
caused via an apoptotic mechanism. A study conducted using two neuroblastoma-7 cell
lines, one with caspase 8 expression and the other without it, lends further evidence to the
hypothesis that somocystinamide A is an apoptosis inducer.8 The former cell line was
found to be 50-fold more susceptible to the effects of 1 than the latter (Figure VI.2).4
FADD proteins at the death receptor and subsequent self-proteolysis of the pro-domain, it
can activate executioner caspases (caspases-3,6,7), which will exercise their destructive
proteolytic ability and cause apoptosis (Figure VI.3).7 Because the initial recruitment of
240
through the caspase-9 pathway (Figure VI.4), but to a lesser extent.4 Moreover, when 1
was reduced to the corresponding monomeric thiol, its bioactivity was lost, suggesting
that either the dimeric nature or the disulfide functionality of 1 is necessary for the
hydrogen bonding between the acetamide groups, which would hold the molecule in a
pair-of-forceps fashion (Figure VI.5). Given that the enamide functionality appears to be
proteins are brought together through the virtue of the aforementioned conformation.
observed activity was simply due to the lipophilic nature of the natural product, an analog
of 1 that lacks any functionalization in the middle of the chain was tested (2).9 This
analog (2) was found to be completely inactive. Likewise, the hydrolysis product of 1,
bis-methylamide 3 was also inactive.10 Therefore, the enamide as well as some or all of
the other functional groups present in 1 seem to be necessary for induction of apoptosis.
liposomes would be a potentially viable mechanism for delivery of the drug. The
endogenous protein, angiostatin,12 has a similar bioactivity profile, and it has advanced to
phase II clinical trial as of 2008 for treatment of lung cancer. 13 Given its bioactivity
of the compound changes more drastically compared to the alkylamide analog. One of the
meant a vinyl amide herein, possesses conjugation of the carbonyl double bond and the
alkenyl group. This conjugated system absorbs UV at 230-240 nm,14 which can be a
useful physical property when the presence of an enamide compound is sought by HPLC
techniques. Easily accessible enamides will almost always have a proton at R3 position
(Scheme VI.1). The proton is also very diagnostic for an enamide moiety in that it gives
Chemically, enamides can be significantly more unstable than amides due to the
possible tautomerism to become an acyl imine or an acyl iminium ion that is further
NMR tube due to the residual moisture and acid (HCl) present in CDCl3 (Scheme VI.2).3
Given this precedence, it is possible that some natural products, as we know them, are
will be discussed later, this distinction becomes significant for devising a strategy to
synthesis have been for secondary enamides, and the same approaches have been found
As discussed earlier, the enamide is a labile functional group and therefore care
isolated in recent years (Figure VI.7). All of these natural products appear to have been
assembled through hybrid PKS-NRPS systems (except for 10, a peptide derivative),17
where the bulk of the carbon skeletons are constructed through PKS catalytic activities
and amino acid residues are embedded via NRPS catalytic activities to provide the
(1) was proposed as shown in Figure VI.6.18 Most of these natural products have been
categorized into secondary enamides (6-10) and tertiary enamides (11, 12, 1). The latter
group generally has a methyl group on the enamide nitrogen that is most likely derived
from S-methy-adenosylmethionine.19,20
Salicylihalamide A (7) is strongly cytotoxic against NCI melanoma cell line (GI50 = 7
nM)22 and is a very potent inhibitor of vacuolar H+-ATPases (IC50 < 1 nM),23 whose
possesses cytotoxicity against human cancer cell lines and inhibitory activity against
vacuolar H+-ATP ases.25 The peptidic compounds terpeptin (8) and chondriamide A (10)
have cell-cycle inhibition activity and cytotoxicity against colon cancer cells (LOVO
246
cells), respectively.26,27 Ulapualides, isolated from sponges and nudibranchs, have been
leukemia cells.28 Interestingly, almost all of the recently isolated enamide natural products
are of marine origin. These newly discovered and bioactive enamide natural products
heterocycle and asymmetric synthesis, where enamides are used as synthons.29,30 There is
one review on the preparation of this labile, yet crucial functional group.15 In Scheme
VI.3, four major types of approaches for synthesizing enamides are summarized.
In devising a strategy for the synthesis of 1, the stabilities of both the disulfide
and the enamide became of concern. Since the installment of the enamide was envisioned
to be performed at a very late stage, if not at the last step, it was desirable to assess the
corresponding carboxylic acid with the fully furnished imine (Scheme VI.3).31 This
approach was successful with simple substrates bearing carboxylic acids and no other
Furthermore, the presence of the disulfide group in 1 requires great care and
epidithiapiperazine-dione natural products (such as 24 in Figure VI.8) has met with much
difficulty in the installation of the disulfide (Scheme VI.6).32,33 To date, only one complete
total synthesis of a compound of this class has been reported.34 Another case in point is
psammaplin A (25); in all three of the published syntheses of 25, the sulfur atoms were
introduced as a disulfide in the final step so as to avoid side reactions (Scheme VI.7).35
the natural product led to the proposed synthetic approach shown in red in Scheme VI.8.
Suzuki coupling is well known as a robust reaction to connect a vinyl or aryl halide to a
vinyl or aryl borylate.36 The reaction is somewhat less reliable if the boron is on an sp 3
Scheme VI.9: Synthesis of vinyl iodide for Suzuki coupling (Scheme VI.10)
carbon.36,37 Thus, besides the enamide synthesis, other steps of concern in the proposed
synthesis were the Takai vinyl iodide synthesis38 and the Suzuki coupling. The Takai
reaction was tested on a commercially available aldehyde derived from proline, which is
structurally very similar to the actual substrate except for the absence of sulfur. This
251
reaction proceeded with good yield (70%). The necessary vinyl halide for the Suzuki
coupling could be synthesized via the Takai protocol from the corresponding aldehyde. 38
Finally late-stage dimerization at the sulfur atom via air oxidation and installation of the
whose stereochemistry translates into the final product. The cysteine was then protected
aldehyde 43 via a Swern reaction,39,40 the aldehyde was treated with chromium (II)
chloride and iodoform (Takai protocol).38 The Takai reaction seemed to give a good yield
by direct TLC analysis of the reaction mixture. However, after work-up and
chromatography, only a minute amount of the desired product was recovered, possibly
due to the presence of excess iodoform (elutes closely with the product) and
decomposition of the vinyl iodide. Brief efforts at optimizing this reaction were not
successful.
252
The vinyl iodide 44 was tested for the Suzuki coupling reaction and only a trace
amount of the desired product was formed. The difficulty of this reaction was attributed
to the combination of functional group incompatibility and, possibly, the excess 9-BBN
that could have reduced the palladium-vinyl complex to the terminal alkene.41 To
examine the latter possibility, the same Suzuki coupling reaction was carried out with a
trifluoroboronate substrate, which is a solid that can be isolated.42 However, this approach
give the correct chain length for the natural product because the terminal sp 2 carbon
would be cleaved off in the process. The metathesis reaction was carried out using 48 and
the vinyl thiazolidine 51, which was synthesized from the aldehyde 43 via a simple
Wittig reaction.44 Different Ruthenium catalysts and various amounts of each substrate
and the catalyst were screened to find an optimum condition for this coupling reaction
(Table VI.1).
The production of the dimeric ester 53 seemed much faster than the desired
product by TLC analysis. This observation seemingly implied that the dimerization of
the ester happens first and then the dimeric ester is used in the actual metathesis.
Therefore it was hypothesized that the yield could be improved by utilizing pre-prepared
dimeric ester for the metathesis, which would presumably reduce the work load for the
catalyst. However, this hypothesis was discounted because the use of 12 as one of the
253
coupling partners did not improve the yield and apparently opened another reaction
pathway for the starting material (entry 5, Table VI.1). In any case, the second generation
Hoveyda catalyst was found to be the best catalyst for this purpose.45
(a) Concentration of 9 (M); (b) Equivalents of 48 or 53 with respect to 51; (c) Isolated yields.
Yields in parentheses are based on recovered starting material; (d) 3.2 g scale (ca. 10 times more
than entries 1~6); (e) 5.5 g scale. The E/Z ratio was 18:1.
254
With a reliable coupling protocol in hand, a new synthetic route was devised, as
shown in Scheme VI.12. The deprotection of the cysteine residue was effected by
reduction of the benzylic carbon with sodium in liquid ammonia in excellent yield (54).13
Upon methylation with diazomethane, the resulting thiol (55) was treated with TFA and
subsequently with acetic anhydride and triethylamine to replace the Boc protecting group
with acetate, which results in acylation of the thiol group as well (56). Both the removal
of the unwanted acetate on the thiol and hydrolysis of the methyl ester were achieved in
one step by treatment with lithium hydroxide in aqueous THF, which also causes the
oxidative dimerization of the thiol if performed in air (somocystinoic acid, 57).46 This air
oxidation only occurred with small scale reactions (<100 mg). Attempts to effect the
dimerization with conventional means, such as treatment with I2, did not give good
yields.
The method developed from the model study was used to effect the acylation with
the corresponding imine. Various conditions were investigated to couple the in situ
generated imine 58 to the corresponding di-acyl chloride 60, whose formation was
verified by the reaction with methylamine to produce 3 (Scheme VI.13, Table VI.2).
Model compounds (63-65) were also tested in this investigation for enamide formation
(Figure VI.9, Table VI.2). In most cases, the starting material decomposed while in some
cases a trace amount of 1 was observed (Table VI.2). This result was curious because, in
addition to the model study, there are reports of synthesis of simple enamides via
acylation of the corresponding acid chloride with imine.47 It is possible that the putative
nature of the substrate (Scheme VI.13).48 In support of this hypothesis, the tautomer 59
of imines, direct addition of amides to alkynes, condensation of aldehyde with amide, and
the olefination of amides (Scheme VI.3). Most of these protocols suffer from low yields
and/or lack of stereocontrol on the double bond geometry. The most recent collective
amides with vinyl halide in the presence of copper (I) iodide.50 At the time when there
this noble method guaranteed both aspects of reaction specificity. Since the
regiospecificity and stereospecificity are solely dictated by the position of the halide on
the olefin, this methodology could be universally applied to any enamide systems as long
as all the functional groups are tolerated. In the past decade, Buchwald and coworkers
have improved this reaction through identification of a more suitable ligand for the
as expected from the literature,16a direct vinylation of acyclic secondary amide was not
possible. While it was possible to vinylate a primary amide model compound, the
required chemoselective methylation of the enamide moiety was not successful (Scheme
VI.14).
relatively low activiation energy3 inspired us to carry out the opposite reaction, namely
condensation of the aldehyde 551 with 3 (Scheme VI.15). Because of the need to remove
water during the course of reaction, the use of a Soxhlet extraction apparatus was more
Figure VI.10: Reaction setup for enamide formation using molecular sieves and glass
wool in a Soxhlet extractor
convenient for small scale reactions and it also allowed for use of solvents heavier than
water (Figure VI.10).52 Observing that the putative intermediate 21 did not yield 1, we
hypothesized that the E1 path way was not viable (Scheme VI.13). In support of this
hypothesis, use of a more polar solvent, THF, had an adverse effect in comparison to 1,2-
dichloroethane, a more non-polar solvent.53 The best result was obtained when TsOH was
used as the catalyst, which gave a 47% yield (Scheme VI.15).54 Thus, a total synthesis of
1 was completed. This synthesis is fairly robust for all steps but the last two such that >1
g of 57 was prepared. However, scale-up for the production of 3 proved difficult due to
activity was investigated in brine shrimp. While 1 was not lethal to the brine shrimp even
significantly impaired even at 1.3 μM. A much decreased quantity of intestinal contents
were found in the treated group. Thus, 1 seems to possess biological properties not yet
understood.
somocystinamide A (1) possessed anti-proliferative activity, but to a lesser extent than the
natural sample. Initially there was concern about the possibility that 1 was not the
this decreased activity, a hypothesis was formed and examined; synthetic 1 is purer than
natural 1 and the natural sample contained contaminants that assisted in solubilizing 1,
whereas the synthetic sample had poor solubility. While much better than 3, the solubility
of 1 is not satisfactory for many purposes. Indeed, in the brine shrimp assays, the samples
assays, serial dilutions are made from the highest concentration solution. Therefore, if
any portion of 1 precipitates in the highest concentration solution, all of the wells
Because all the in vitro bioassays are done in aqueous environments, the solubility
compound was detected when only organic solvents were used (the first two entries in
Figure VI.9). However, when a solution of 1 was diluted with H2O, the measured
dodecamide, DMEM) eliminated the solubility of 1. Given these results, any future
262
bioassay of 1 must be carried out with care. It is recommended that the maximum amount
reprotection maneuver.
readily dissolved in solvents such as EtOAc, the workup and purification would be
simply a matter of aqueous acid/base washings for this amine coupling step.
3.The final step of introducing the enamide moiety has not been optimized beyond 41%
yield.
In order to remedy all three of these issues, an efficient alternative method must
be first developed for the production of 3, which would enable optimization studies on
the finals step from 3 to 1. When investigating the enamide formation reaction with
simpler substrates, such as undecanoic acid, it was observed that the resulting tertiary
enamide had excellent solubility in common solvents such as CH2Cl2 and MeOH.
However, the dimeric equivalent of this compound, 2, had extremely poor solubility in
these same solvents. Suspecting that the insolubility of bis-amide 3 was at least partially
due to its dimeric nature and the methyl amide functionality, a synthetic route for 3 was
devised in which the intermediates remain as monomers until the enamide formation step
This scheme shortened the synthesis by two steps. However, the cross metathesis
step suffered low yields (~40%) under the same condition as the previous synthesis. After
insoluble solid. In the previous reaction (51 to 52), the dimeric ester by-product was able
to be recycled in the reaction (Table IV.1, entry 5), which was probably a large
contributing factor to the excellent yield (82%) despite the faster formation of 53 than 52.
Therefore, it was hypothesized that if a solvent system that solubilize 74 was utilized, the
yield would increase due to recycling of 74. When 9:1 CH2Cl2 / MeOH was used, the
yield decreased and most of the starting material (51) was left intact. When CHCl3 was
used, the cross metathesis proceeded smoothly with excellent yield (81-84%). It was
264
noted, however, that the purification of the product (71) was more troublesome than 52
revised synthesis, the use of a potentially dangerous and toxic reagent, diazomethane was
also eliminated.56 The hydrolysis/oxidation step from 73 to 3 only required washing of the
precipitated product, 3, thereby taking advantage of its insolubility. However, the yield
was somewhat less than ideal (65%). Purification methods such as washing of a solid is
well-suited for future adaptation to industrial scale production should the compound
become a pharmaceutical product.57 All of these steps were done at fairly large scales and
>1 g of 73 and >500 mg of 3 were obtained, the latter of which would not have been
reasonable without this revision of the synthesis. Finally, the penultimate natural product
3 (301 mg) could be converted to somocystinamide A (1, 119 mg, 33% yield) through the
scale synthesis58 required for future in vivo investigations. The overall yield of this
revised synthetic route was 11% over 7 steps from the known aldehyde 43, whereas the
overall yield of the earlier synthesis was 9% over 10 steps from 43. Despite a small
difference in the overall yield, the smaller number of steps required for the revised route
VI.8 Conclusions
the thiol, and enamide formation through condensation of aldehyde with amide. This
synthesis provided material for further bioassays and helped to identify the solubility
problem of this natural product. Furthermore, the synthesis was successfully revised to
suit the future needs for synthesis of analogs and a large supply of the compound for in
vivo evaluations.
266
and were used without further purification. Anhydrous benzene was purchased from
EMD. Tetrahydrofuran (THF) was distilled from sodium / benzophenone. Et3N and
CH2Cl2 were distilled from CaH. Ac2O was distilled from quinone. Dimethyl sulfoxide
(DMSO), oxalyl chloride, and trifluoroacetic acid (TFA) were distilled without desiccant.
All reactions were carried out under dry argon atmosphere unless otherwise noted.
Reaction temperatures herein recorded are external temperatures unless otherwise noted.
Flash chromatorgraphy was performed using EMD silica gel (230-400 mesh) in all
cases.59 Thin layer chromatographic (TLC) analysis was performed using EM Science
pre-coated silica gel plates (Merck 60 F254). Melting points were determined on an
Electrothermal Mel-Temp device and are uncorrected. Optical rotations were measured
NMR and 13C NMR spectra were recorded on a Varian Inova spectrometer (500 MHz and
125 MHz, respectively) or on a Varian Inova spectrometer (300 MHz and 75 MHz,
respectively). 1H and 13
C spectra recorded in CDCl3 (or in CD3OD/CDCl3) were
referenced to the residual solvent peaks at 7.26 ppm (CHCl3) and 77.0 ppm (CDCl3),
MAT900XL spectrometer.
267
43
The first step was done according to reference 59.60 The rest of the steps to the aldehyde
Synthesis of enamides 22 and 2. To a solution of the aldehyde 5 (97 μL, 0.98 mmol) in
CH2Cl2 (5 mL) was added a 2M solution of MeNH2 in THF (0.49 mL, 0.98 mmol) and
MgSO4 (710 mg) at 0˚C. Immediately, the mixture was warmed to rt and left stirring for
1.5 hrs. Then neat acetylchloride (70 μL, 0.98 mmol) was added. After 10 min, pyridine
(95 μL, 1.18 mmol) was added, and the mixture was left stirring for 3 hrs at rt. The
reaction mixture was filtered and was poured into a cold saturated solution of NaHCO3.
Upon phase separation, the organic layer was washed with saturated NaHCO 3 and was
dried over Na2SO4. The volatiles were removed in vacuo and a brown oil was obtained,
which was mostly pure by TLC and 1H NMR (96 mg, ~70% yield). In order to obtain an
analytically pure sample, the brown oil was subjected to flash column chromatography
using silica gel (1:6 EtOAc / Hexanes to 1:5 EtOAc / Hexanes). A colorless oil (51 mg,
0.37 mmol, 37%) was obtained. TLC Rf = 0.41 (3:7 EtOAc/Hexanes). The bis-enamide 2
was synthesized in the same manner except that two equivalents of 5 and MeNH2 were
(5.189 g, 42.22 mmol) was flame-dried in a flask under vacuum. Upon cooling to rt, THF
(31 mL) was added and the suspension was stirred at rt for 25 min. Then the mixture was
cooled to 0°C. A solution of the aldehyde 43 (1.972 g, 6.693 mmol) and CHI3 (5.270 g,
13.38 mmol) in THF (30 mL) was cannulated dropwise into the CrCl 2 suspension over 20
min. An additional amount of THF (10 mL) was used to rinse the flask containing 43 and
CHI3, which was cannulated again into the CrCl2 suspension. At this point, TLC analysis
showed that there was little starting material left, and that a significant amount of the
desired product had formed. This mixture was taken out of the ice bath immediately and
was stirred at rt for 4 hrs. The mixture was diluted with Et2O (30 mL), and it was poured
into brine, which was extracted with Et2O (20 mL x 2). The organic layer was washed
with brine, and was dried over Na2SO4. Upon removal of the solvents, the residues were
mg of the product (44) as a pale yellow solid (4.4% yield). TLC Rf = 0.30 (1:4 EtOAc /
hexanes). 1H and 13C NMR spectra: see Figures VI.15 and 16.
of MePPh3·Br (13.37 g, 37.43 mmol) in THF (150 mL) at –78°C (dry ice bath) was added
1.0 M NaHMDS (43 mL, 43 mmol) dropwise under vigorous stirring, which resulted in a
yellow solution. The dry ice bath was removed and the solution stirred for 45 min at rt.
270
The solution was cooled again to –78°C and a solution of the aldehyde (8.302 g, 28.30
mmol, synthesized according to literature2) in THF (100 mL) was added dropwise. Then
the solution was warmed to rt over night under stirring. Consumption of the aldehyde was
confirmed by TLC analysis. Then 1.0 M aqueous NH4Cl (100 mL) and Et2O (100 mL)
were added successively. The aqueous phase was extracted with Et2O (100 mL x 2). The
combined organic phase was washed with brine, dried over Na2SO4, and was filtered
through a plug of silica gel. Upon evaporation of solvents, the residue was purified by
flash column chromatography (3:97 to 1:9 Et2O / hexane) and a white solid (6.04 g, 20.7
mmol) was obtained (74% yield). Mp 44-45 ºC. TLC Rf 0.39 Et2O / hexane (1:4). [α]D23
+136.2º (c 6.50, CHCl3). IR (film, KBr) νmax 2977, 1698, 1367, 1164 cm-1. 1H NMR (300
MHz, CDCl3) δ 7.22-7.39 (m, 5H), 6.10 (s, 1H), 6.03 (m, 1H), 5.37, (d, 17.1 Hz, 1H),
5.25 (dt, J=10.2, 0.9 Hz, 1H), 4.83 (m, 1H), 3.27 (dd, J=11.7, 6.6 Hz, 1H), 2.85 (dd,
J=11.7, 5.1 Hz, 1H), 1.33 (s, 9H). 13C NMR (75 MHz, CDCl3) δ 153.7, 141.6, 137.8,
128.2, 127.5, 126.3, 117.2, 80.7, 66.3, 64.5, 36.1, 28.2. HRMS (EI): calcd for
(2R,4R)-3-tert-butoxycarbonyl-4-((E)-11-methoxy-11-oxoundec-1-enyl)-2-
mmol) in CH2Cl2 (95 mL) was added methyl undecenoate (9.2 mL, 41 mmol) at rt. The
271
solution was then subjected to vacuum-sonication-Ar introduction cycle three times for
degassing. Then second generation Hoveyda-Grubbs catalyst (592 mg, 0.944 mmol) was
addedin one portion. After 20 min of stirring at rt, the solution was refluxed for 16 hrs.
Then the solvents were removed under vacuum and the residue was subjected to flash
column chromatography (hexane to 1:9 Et2O / hexane) to obtain colorless oil (7.174 g,
15.54 mmol, 82% yield). Along with the desired product was isolated the cis isomer
(TLC Rf=0.52 Et2O / hexane 3:7, 539 mg, 1.17 mmol, 6% yield cis). TLC Rf 0.47 Et2O /
hexane (3:7). [α]D23 –81.7º (c 1.85, CHCl3). IR (film, KBr) νmax 2978, 2928, 2853, 1740,
1696, 1367, 1166 cm-1. 1H NMR (300 MHz, CDCl3) δ 7.21-7.38 (m, 5H), 6.08 (s, 1H),
5.78 (dt, J = 15.3 Hz, 6.6 Hz, 1H), 5.63 (dd, J = 15.3 Hz, 7.6 Hz, 1H), 4.81 (ddd, J = 7.6,
6.5, 4.6 Hz, 1H), 3.65 (s, 3H), 3.24 (dd, J=11.7, 6.5 Hz, 1H), 2.79 (dd, J = 11.7, 4.6 Hz,
1H), 2.29 (t, J = 7.6 Hz, 2H), 2.06 (q, J = 6.8 Hz, 2H), 1.61 (m, 2H), 1.28-1.41 (m, 20H).
13
C NMR (75 MHz, CDCl3) δ 174.2, 153.5, 141.8, 133.9, 129.4, 128.1, 127.3, 126.2,
80.4, 66.0, 63.8, 51.3, 36.5, 34.0, 32.1, 29.2, 29.1, 29.1, 29.0, 29.0, 28.2, 24.9. HRMS
52-cis. 1H NMR (500 MHz, CDCl3) δ 7.40 – 7.16 (m, 5H), 6.13 (s, 1H), 5.63 (t, J = 9.7
Hz, 1H), 5.58 – 5.45 (m, 1H), 5.13 (d, J = 6.1 Hz, 1H), 3.65 (d, J = 1.7 Hz, 3H), 3.20 (dd,
J = 11.6, 6.4 Hz, 1H), 2.72 (dd, J = 11.6, 5.5 Hz, 1H), 2.29 (t, J = 7.5 Hz, 2H), 2.25 –
2.10 (m, 2H), 1.60 (dd, J = 13.6, 6.8 Hz, 2H), 1.51 – 1.14 (m, 20H). 13C NMR (125 MHz,
CDCl3) δ 174.15, 153.59, 141.92, 132.08, 129.72, 128.11, 127.30, 126.01, 80.46, 66.05,
59.14, 51.33, 36.75, 33.99, 29.58, 29.24, 29.19, 29.11, 29.02, 28.17, 27.60, 24.84.
272
(E)-11-((2R,4R)-3-(tert-butoxycarbonyl)-2-phenylthiazolidin-4-yl)undec-10-enoic
acid (52-OH). The solution of the methyl ester 52 (11.40 g, 24.70 mmol) in THF (240
mL) was cooled to 0°C. A solution of LiOH·H2O (10.07 g, 240.2 mmol) in H2O (120 mL)
was added in one portion. Then the reaction mixture was stirred and warmed gradually
over night to rt. Upon seeing some intact starting material on TLC, MeOH (25 mL) was
added at rt. 6 hrs later, the solution was cooled to 0°C and Et2O (200 mL) and 1M
NaHSO4 solution (300 mL) were added, after which the pH was ca 2~3. Upon phase
separation, the aqueous layer was extracted with Et2O (150 mL x 3). The combined
organic layer was washed with brine (250 mL), dried over MgSO4, and was filtered.
Upon removal of the solvents, a gummy oil (11.27 g) was obtained, which was used in
the next step without further purification. TLC Rf 0.70 EtAcO / hexane (3:2). [α]D23
+69.5º (c 4.4, CHCl3). IR (film, KBr) νmax 2975, 2928, 2855, 1701, 1369, 1164 cm-1. 1H
NMR (300MHz, CDCl3) δ 7.21-7.38 (m, 5H), 6.08 (s, 1H), 5.78 (dt, J = 15.3, 6.6 Hz,
1H), 5.63 (dd, J = 15.3, 7.6 Hz, 1H), 4.82 (ddd, J = 7.6, 6.4, 4.7 Hz, 1H), 3.24 (dd, J =
12.2, 6.4 Hz, 1H), 2.80 (dd, J = 12.2, 4.7 Hz, 1H), 2.34 (t, J = 7.5 Hz, 2H), 2.07 (q, J =
6.7 Hz, 2H), 1.63 (m, 2H), 1.30-1.43 (m, 19H). 13C NMR δ 179.5, 153.6, 141.8, 134.0,
129.4, 128.1, 127.4, 126.3, 80.5, 66.1, 63.9, 36.6, 34.2, 32.2, 29.21, 29.15, 29.04, 28.99
(3 C's), 28.2, 24.6. HRMS (EI): calcd for C25H37O4N1S1: 447.2438, found 447.2441.
273
thiazolidine 52-OH (11.145 g from the previous step) in THF (10 mL) was added NH3(l)
(350 mL) that was distilled over Na. To this mixture was added cubic pieces (5~7 mm) of
Na under vigorous stirring until dark blue color persisted. The mixture was kept refluxing
at rt and Na pieces were added as necessary to keep the color for a total of 2 hrs. Then
NH4Cl was carefully added until the color disappeared. NH3 was removed under gentle
The carboxylic acid (8.030 g, from the previous step) was dissolved in Et 2O (50
mL) and was cooled to 0°C. Ethereal solution of CH2N2 (0.14 M, 170 mL, 24 mmol) was
added dropwise over 30 min. Then N2 was bubbled in for 45 min to remove excess
CH2N2. Upon removal of the solvent in vacuo, white crystals were obtained (8.74 g).
These crystals were dissolved in CH2Cl2 (200 mL) and was cooled to 0°C. Then
TFA (50 mL) was added dropwise over 30 min. Upon further stirring at 0°C for 30 min,
the volatiles were removed in vacuo to yield orange oil. Residual TFA was removed
under hi-vacuum for 1 hr. This TFA salt was dissolved in CH 2Cl2 (220 mL) and the
solution was cooled to 0°C. Then Et3N (30.6 mL, 220 mmol) and Ac2O (10.4 mL, 110
mmol) were added slowly under stirring. The solution was let warm to rt over night under
stirring. Then the volatiles were removed in vacuo. The residues were dissolved in Et2O
(200 mL) and 0.5 M HCl solution (200 mL) was added. Upon phase separation, the
aqueous phase was extracted with Et2O (200 mL x 2). The combined organic phase was
274
washed with H2O (150 mL), saturated NaHCO3 solution (100 mL), H2O (100 mL), and
brine (100 mL) successively. After flash column chromatography (15:85 EtAcO / hexane
to 1:1 EtAcO / hexane), the desired product was obtained as a white solid (~4.2 g, 12
mmol, 50% yield over 5 steps). The disulfide methyl ester was also isolated as a colorless
oil (173 mg, 0.275 mmol, 1% yield). Mp 70-71 ºC. TLC Rf 0.38 EtAcO / hexane (3:2).
[α]D23 –13.2º (c 6.10, CHCl3). IR (film, KBr) νmax 3278, 3081, 2985, 2926, 2852, 1737,
1691, 1647, 1547, 1434, 1373, 1186, 1139, 968, 630 cm-1. 1H NMR (500MHz, CDCl3) δ
5.97 (d, J = 8.1 Hz, 1H), 5.57 (dtd, J = 15.5, 6.6, 1.2 Hz, 1H), 5.29 (ddt, J = 15.5, 6.4, 1.2
Hz, 1H), 4.50 (quin, J = 6.8 Hz, 1H), 3.60 (s, 3H), 3.01 (d, J = 6.6 Hz, 2H), 2.28 (s, 3H),
2.24 (t, J = 7.7 Hz, 2H), 1.94 (q, J = 7.8 Hz, 2H), 1.90 (s, 3H), 1.55 (quin, J = 7.4 Hz,
13
2H), 1.21-1.30 (m, 12H). C NMR (125MHz, CDCl3) δ 195.9, 174.1, 169.3, 133.1,
127.1, 51.3, 50.9, 33.9, 33.4, 32.0, 30.4, 29.0, 29.0, 28.9, 28.8, 28.8, 24.7, 23.1. HRMS
Somocystinoic acid (57). To the solution of the methyl ester (487 mg, 1.36 mmol) in
THF (40 mL) was added 1M NaOH solution (20 mL) at rt under stirring. O2 was
continuously bubbled into the mixture while stirring at rt until the consumption of the
275
thiol was observed by TLC (20 hrs). The reaction mixture was diluted with Et2O (30 mL)
and 1M KHSO4 solution (40 mL) was added at 0°C. Upon phase separation, the aqueous
layer was extracted with Et2O (30 mL) and EtAcO (30 mL x 2). The combined organic
layer was dried over MgSO4 and was filtered. Removal of the solvents in vacuo yielded
pale yellow oil (458 mg), which was used without further purification for the next step.
TLC Rf 0.41 AcOH / MeCN / EtAcO (1:9:90). [α]D23 +41.6º (c 0.66, CHCl3). IR (film,
KBr) νmax 3272(br), 3079(br), 2927, 2855, 1712, 1648, 1550, 1374, 969 cm -1. 1H NMR
(300 MHz, CDCl3) δ 6.23 (d, J = 8.1 Hz, 2H), 5.66 (dtd, J = 15.5, 6.8, 0.7 Hz, 2H), 5.43
(dd, J = 15.5, 6.2 Hz, 2H), 4.70 (quin, J = 6.5 Hz, 2H), 3.06 (dd, J = 13.5, 6.0 Hz, 2H),
2.85 (dd, J = 13.5, 6.2 Hz, 2H), 2.35 (t, J = 7.1 Hz, 4H), 2.04 (m, 4H), 2.03, (s, 6H), 1.64
(quin, J = 7.4 Hz, 4H), 1.26-1.43 (m, 20H). 13C NMR (75 MHz, CDCl3) δ 178.7, 170.1,
133.7, 127.6, 50.8, 44.5, 33.9, 32.0, 28.7, 28.7, 28.6, 28.5, 28.3, 24.6, 23.3. HRMS
(72 mg, 0.1198 mmol) in CH2Cl2 (8 mL) was added MeNH3Cl (24 mg, 0.36 mmol), Et3N
(84 µL, 0.60 mmol), and EDC·HCl (50 mg, 0.26 mmol), and DMAP (5 mg) successively
276
at 0°C under stirring. The reaction was immediately warmed to rt and was stirred at rt for
5 hrs. After 3 hrs into the reaction, white precipitates appeared. The solvents were
removed under vacuum. The residues were taken up in CH2Cl2 / MeOH (2:1) and the
mixture was washed with 0.5 M HCl solution and saturated NaHCO3 solution and was
dried over Na2SO4. Upon removal of the solvents, sufficiently pure material for the next
step was obtained as a white solid (52 mg, 0.083 mmol, 69% yield over 2 steps). Mp
154-155 ºC. TLC Rf 0.36 MeOH / CH2Cl2 (1:9). [α]D23 –17.8º (c 1.4, CHCl3/MeOH 1:1).
IR (film, KBr) νmax 3306, 3276, 2918, 2849, 1642, 1540 cm-1. 1H NMR (300 MHz,
CD3OD/CDCl3 1:1, rotamers) δ 7.55 (d, J =8.3 Hz, 2H), 7.25 (bs, 2H), 5.31 (dt, J = 15.5,
6.7 Hz, 2H), 5.06 (dd, J = 15.5, 6.6 Hz, 2H), 4,24 (m, 2H), 2.53 (d, J = 7.1 Hz, 4H), 2.38
& 2.39 (s, 6H), 1.83 (t, J = 7.8 Hz, 4H), 1.68 (q, J = 6.5 Hz, 4H), 1.64 (s, 6H), 1.25 (quin,
J = 7.4 Hz, 4H), 0.95-1.05 (m, 20H). 13C NMR (75 MHz, CD3OD/CDCl3 1:1, rotamers)
δ 175.2, 175.1, 170.7, 170.6, 132.7, 127.2, 50.1, 43.3, 35.6, 35.5, 31.6, 28.6, 28.5, 28.5,
28.4, 28.3, 25.2, 25.1, 21.7,. 21.6. HRMS (EI): calcd for C32H59O4N4S2 (M+H): 627.3972,
found 627.3973.
277
Somocystinamide A (1). The suspension of the methyl amide (42 mg, 0.0670 mmol) in
introduction cycle. Then 4-pentenal (100 µL, 1.01 mmol, synthesized according to
literature)51 and TsOH·H2O (3 mg, 0.015 mmol) were added successively at rt. The
reaction mixture was refluxed for 14 hours under stirring in a Soxhlet apparatus that is
equipped with glass wool and molecular sieves 4Å so as to remove residual H 2O. Then
another batch of 4-pentenal (100 μL, 1.01 mmo) and TsOH·H 2O (3 mg, 0.015 mmol)
were added. After 1.5 hours, 4-pentenal (100 μL, 1.01 mmol) was added again. Upon
further refluxing and stirring for 1.5 hours, the reaction mixture was cooled to 0°C, and
Et3N (100 µL) was added. Most of the solvent was removed in vacuo, but ca. 1 mL of
solvent was retained, which was directly subjected to flash column chromatography
(EtAcO to 1:1:8 MeOH / CH2Cl2 / EtAcO) to obtain somocystinamide A (21 mg, 0.0277
mmol, 41% yield) as an off-white amorphous solid. Mp 93-95 ºC. TLC Rf 0.56 MeOH /
CH2Cl2 / EtAcO (1:1:8). [α]D22 +19.1º (c 0.59, CHCl3). IR (film, KBr) νmax 3286(br), 2926,
2854, 1645, 1541, 1395, 1289, 1087, 968 cm-1. 1H NMR (300 MHz, CDCl3, rotamers) δ
7.35 (d, J = 14.4 Hz 0.6H), 6.67 (dt, J = 13.7, 1.3 Hz, 1.4H), 6.28 (d, J = 7.8 Hz, 2H),
5.84 (m, 2H), 5.65 (dtd, J = 15.5, 6.7, 1.1 Hz, 2H), 5.43 (dd, J = 15.5, 6.2 Hz, 2H),
4.91-5.11 (m, 4H), 4.67 (quin, J = 6.5 Hz, 2H), 3.07 and 3.08 (s, 6H), 3.06 (m, 2H), 2.82
(m, 6H), 2.42 (m, 4H), 2.03 (m, 4H), 2.01 (s, 6H), 1.64 (quin, J = 7.5 Hz, 4H), 1.25-1.38
(m, 20H). 13C NMR (75 MHz, CDCl3, rotamers) δ 171.7, 169.1, 137.4, 137.0, 133.7,
129.2, 128.1, 127.6, 115.4, 115.0, 109.0, 108.4, 50.8, 44.7, 34.5, 34.5, 34.4, 33.8, 32.3,
32.2, 29.7, 29.7, 29.3, 29.2, 29.0, 28.9, 25.0, 23.4. UV (MeOH) λ max 234 nm (ε 3.2×105).
mmol) in CH2Cl2 (350 mL) was added MeNH2·HCl (18.45 g, 273.3 mmol) at rt. This
mixture was cooled to –15ºC and Et3N (57.0 mL, 408 mmol) was added. Then EDC·HCl
(27.29 g, 142.4 mmol) and DMAP (238 mg, 1.95 mmol) were added successively under
vigorous stirring. The mixture was very cloud at this point. The mixture was gradually
warmed to rt, and was stirred over night. The mixture appeared milky. Upon removal of
the volatiles in vacuo, the solid residues were triturated with EtOAc (500 mL). This
mixture was filtered, and the solids were rinsed with EtOAc. Then the EtOAc solution
was poured into 1M HCl (250 mL), and the phases were separated. The organic layer was
washed with H2O (150 mL), saturated NaHCO3 (150 mL), H2O (100 mL), and brine (100
mL) successively. After the solution was dried over Na2SO4 and the solvent was removed,
a white solid (70) resulted, which was sufficiently pure for the subsequent step. 1H NMR
(300 MHz, CDCl3) δ 5.80 (ddt, J = 16.9, 10.2, 6.7 Hz, 1H), 5.50 (s, 1H), 5.05 – 4.86 (m,
2H), 2.80 (d, J = 4.9 Hz, 3H), 2.15 (t, J = 7.6 Hz, 2H), 2.02 (dd, J = 14.1, 6.9 Hz, 2H),
1.70 – 1.53 (m, 3H), 1.45 – 1.18 (m, 15H). 13C NMR (75 MHz, CDCl3) δ 173.80, 139.15,
(2R,4R)-tert-butyl 4-((E)-11-(methylamino)-11-oxoundec-1-enyl)-2-phenyl
thiazolidine -3-carboxylate (71). To the thiazolidine (51) (2.063 g, 7.079 mmol) and the
methylamide (70) (2.270 g, 11.50 mmol) was added CHCl3 (45 mL) that was passed
through basic alumina (activity I). Then the reaction mixture was degassed in vacuum
with ultrasonic assistance three times. Grubbs-Hoveyda II catalyst (95 mg, 0.152 mmol)
was added. After stirring at rt for 30 min, the reaction mixture was refluxed (bath
temperature = 80ºC). Over the course of 18 hrs, additional amounts of 70 (1.932 g, 9.791
mmol) and Grubbs-Hoveyda II catalyst (354 mg, 0.564 mmol, 5.8 mol%) were added in
four portions. Then the mixture was cooled to rt, concentrated under vacuum, and
subjected to silica gel flash column chromatography (1:9 EtOAc / Hex to 2:9:9 acetone /
EtOAc / Hex). Some of the colored impurities were not completely removed. A dark-
brown thick oil (2.752 g, 5.974 mmol, 84%) was obtained. TLC Rf = 0.47 (2:9:9 MeOH/
EtOAc/hexanes). [α]22D +58.6º (c 3.6, CHCl3). IR (film, KBr) νmax 3298(br), 2927, 2854,
1696, 1650, 1554, 1368, 1163, 1109, 697 cm-1. 1H NMR (400 MHz, CDCl3) δ 7.37 – 7.19
(m, 5H), 6.05 (s, 1H), 5.83 – 5.70 (m, 1H), 5.61 (dd, J = 15.3, 7.4 Hz, 1H), 4.79 (s, 1H),
3.22 (dd, J = 11.7, 6.5 Hz, 1H), 2.81 – 2.75 (m, 1H), 2.75 (d, J = 4.9 Hz, 3H), 2.17 – 2.08
(m, 1H), 2.08 – 1.99 (m, 2H), 1.64 – 1.53 (m, 2H), 1.41 – 1.20 (m, 21H). 13C NMR (100
MHz, CDCl3) δ 173.8, 153.5, 141.7, 134.0, 129.3, 128.1, 127.4, 126.2, 80.4, 66.0, 63.8,
36.6, 36.5, 32.1, 29.2, 29.2, 29.1, 29.0, 28.9, 28.1, 26.1, 25.7. HRMS (ESI): calcd for
Liquid NH3 (150 mL) was distilled from Na onto a solution of 71 (3.098 g, 6.725 mmol)
in THF (10 mL). Then Na pieces (~ 5 mm cubes) were added to keep the color of the
solution dark blue for 2 hrs under reflux (at rt) and vigorous stirring. Solid NH 4Cl was
added carefully until the blue color disappeared. NH3 was evaporated under a stream of
N2. Then 1M NaHSO4 (300 mL) was added, which was extracted with EtOAc (200 mL x
3). The organic layer was washed with brine and dried over Na2SO4. Upon removal of the
solvent in vacuo, amorphous pale brown solid (2.531 g) resulted, which was used in the
next step without purification. This solid (2.531 g) was dissolved in CH2Cl2 (75 mL) and
freshly distilled TFA (18 mL) was added at 0ºC under stirring. After the solution was kept
at 0ºC under stirring for 30 min, the volatiles were removed in vacuo. The residues were
placed under high vacuum for 2 hours, which resulted in a thick orange oil. This oil was
dissolved in CH2Cl2 (100 mL) and the solution was cooled to 0ºC. Et3N (15.8 mL, 113.2
mmol) and acetic anhydride (5.3 mL, 56.6 mmol) were added successively at 0ºC under
stirring. The solution was gradually allowed to warm to rt. After 16 hrs of stirring at rt,
the volatiles were removed in vacuo. The residues were treated with 1M NaHSO4 (250
mL), which was extracted with EtOAc (200 mL x 3). The combined organic layer was
washed with brine (150 mL) and dried over Na2SO4. Most of the solvent was removed
under vacuum, and the crude solution was let stand in a -20ºC freezer over night. Pale
brown solids were precipitated and filtered, which was rinsed with hexanes. Upon drying
281
under high vacuum, 73 was obtained as a pale brown solid (1.112 g). The filtrate
contained some of the product, and it was purified by flash column chromatography (1:1
0.90, CHCl3). IR (film, KBr) νmax 3314, 3286, 3089(br), 2925, 2853, 1693, 1646, 1553,
1412, 1373, 1133, 968 cm-1. 1H NMR (300 MHz, CDCl3) δ 6.03 (d, J = 8.3 Hz, 1H), 5.90
(s, 1H), 5.59 (ddd, J = 7.9, 7.3, 4.0 Hz, 1H), 5.43 – 5.21 (m, 1H), 4.52 (p, J = 6.3 Hz,
1H), 3.04 (d, J = 6.6 Hz, 2H), 2.76 (d, J = 4.8 Hz, 3H), 2.32 (s, 3H), 2.13 (t, J = 7.6 Hz,
2H), 2.00 – 1.95 (m, 2H), 1.93 (s, 3H), 1.72 – 1.42 (m, 2H), 1.26 (m, 10H). 13C NMR (75
MHz, CDCl3) δ 196.2, 173.9, 169.5, 133.1, 127.8, 58.1, 51.0, 36.5, 33.5, 32.0, 30.5, 29.1,
29.0, 29.0, 28.7, 26.2, 25.6, 23.2. HRMS (ESI): calcd for C18H32N2O3SNa [M+Na+]:
70
70
2 Gerwick, W. H.; Tan, L. T.; Sitachitta, N. The Alkaloids; Cordell, G. A., Ed.; Academic Press;
San Diego, CA 2001; Vol. 57, pp 75-184.
4 Wrasidlo, W.; Mielgo, A.; Torres, V. A.; Barbero, S.; Stoletov, K.; Suyama, T. L.; Klemke, R.
L.; Gerwick, W. H.; Carson, D. A.; Stupack, D. G. The marine lipopeptide somocystinamide A
triggers apoptosis via caspase 8. Proc. Natl. Acad. Sci. 2008, 105, 2313-2318.
5 Park, H.-J.; Zhang, Y.; Georgescu, S. P.; Johnson, K. L.; Kong, D.; Galper, J. B. Human
umbilical vein endothelial cells and human dermal microvascular endothelial cells offer new
insights into the relationship between lipid metabolism and angiogenesis. Stem Cell Reviews
and Reports 2006, 2, 93-101.
6 Goishi, K.; Klagsbrun, M. Vascular endothelial growth factor and its receptors in embryonic
zebrafish blood vessel development. Current Topics in Developmental Biology 2004, 62,
127-152.
8 There is some evidence that 1 also induces intrinsic apoptosis, which involves caspase-9, in
addition to the extrinsic apoptotic pathway involving caspase-8 (see ref. 4).
10 This compound was obtained through unintentional hydrolysis of 1 (see Scheme VI.2).
14 Crews, P.; Rodrǐguez, J.; Jaspars, M. Optical and chiroptical techniques: Ultraviolet
spectroscopy. In Organic Structure Analysis; Oxford University Press: Oxford, 1998, pp
349-373.
311
15 Tracey, M. R.; Hsung, R. P.; Antoline, J.; Kurtz, K. C. M.; Shen, L.; Slafer, B. W.; Zhang, Y.
Product class 4: N-Arylalkanamides, ynamides, enamides, dienamides, and allenamides.
Science of Synthesis 2005, 21, 387-475.
16 (a) Jiang, L.; Job, G. E.; Klapars, A.; Buchwald, S. L. Copper-Catalyzed Coupling of Amides
and Carbamates with Vinyl Halides . Org. Lett. 2003, 5, 3667–3669. (b) Dehli, J. R.; Legros,
J.; Bolm, C. Synthesis of enamines, enol ethers and related compounds by cross-coupling
reactions . Chem. Commun. 2005, 973–986. (c) Han, C.; Shen, R.; Su, S.; Porco, J. A., Jr.
Copper-Mediated Synthesis of N-Acyl Vinylogous Carbamic Acids and Derivatives: Synthesis
of the Antibiotic CJ-15,801. Org. Lett. 2004, 6, 27–30.
17 (a) Ramaswamy, A. V.; Flatt, P. M.; Edwards, D. J.; Simmons, T. L.; Han, B.; Gerwick, W. H.
The secondary metabolites and biosynthetic gene clusters of marine cyanobacteria.
Applications in biotechnology. Frontiers Mar. Biotech. 2006, 175-224. (b) Moore, Bradley
S. Biosynthesis of marine natural products: macroorganisms (Part B). Nat. Prod. Rep, 2006,
23, 615-629.
18 Chang, Z.; Sitachitta, N.; Rossi, J. V.; Roberts, M. A.; Flatt, P. M.; Jia, J.; Sherman, D, H.;
Gerwick, W. H. Biosynthetic Pathway and Gene Cluster Analysis of Curacin A, an Antitubulin
Natural Product from the Tropical Marine Cyanobacterium Lyngbya majuscula. J. Nat. Prod.
2004, 67, 1356-1367.
19 Cantoni, G. L. The Nature of the Active Methyl Donor Formed Enzymatically from L-
Methionine and Adenosinetriphosphate. J. Am. Chem. Soc. 1952, 74, 2942-2943.
21 Rukachaisirikul, V.; Sommart, U.; Phongpaichit, S.; Sakayaroj, J.; Kirtikara, K. Metabolites
from the endophytic fungus Phomopsis sp. PSU-D15. Phytochem. 2008, 69, 783-787.
22 (a) Erickson, K. L.; Beutler, J. A.; Cardellina, J. H., II; Boyd, M. R. Salicylihalamide A and B,
novel cytotoxic macrolides from the marine sponge Haliclona sp. J. Org. Chem. 1997, 62,
8188–8192. (b) Erickson, K. L.; Beutler, J. A.; Cardellina, J. H., II; Boyd, M. R.
Salicylihalamides A and B, Novel Cytotoxic Macrolides from the Marine Sponge Haliclona
sp. [Erratum]. J. Org. Chem. 2001, 66, 1532.
23 Boyd, M. R.; Farina, C.; Belfiore, P.; Gagliardi, S.; Kim, J. W.; Hayakawa, Y.; Beutler, J. A.;
McKee, T. C.; Bowman, B. J.; Bowman, E. J. Discovery of a novel antitumor benzolactone
enamide class that selectively inhibits mammalian vacuolar-type (H+)-ATPases. J.
Pharmacol. Exp. Ther. 2001, 297, 114–120.
24 Sun-Wada, G.-H.; Wada, Y.; Futai, M. Diverse and essential roles of mammalian vacuolar-type
proton pump ATPase: toward the physiological understanding of inside acidic compartments.
Biochim. Biophys. Acta 2004, 1658, 106–114.
312
25 Huss, M.; Wieczorek, H.; Inhibitors of V-ATPases: old and new players. J. Exp. Biol. 2009,
212, 341-346.
27 Jorge A. Palermo, Patricia Blanch Flower and Alicia M. Seldes. Chondriamides A and B, New
Indolic Metabolites Red Alga Chondria sp. Tetrahedron Lett. 1992, 33, 3097-3100.
28 Pattenden, G.; Ashweek, N. J.; Baker-Glenn, C. A. G.; Kempson, J.; Walker, G. M.; Yee , J. G.
K. Total synthesis of (−)-ulapualide A, a novel tris-oxazole macrolide from marine
nudibranchs, based on some biosynthesis speculation . Org. Biolmol. Chem. 2008, 6,
1478-1497.
29 Su, S.; Kakeya, H.; Osada, H.; Porco, J. A. Jr. Synthesis and cell cycle inhibition of the peptide
enamide natural products terpeptin and the aspergillamides . Tetrahedron 2003, 59,
8931-8946.
33 Middleton, M. D.; Peppers, B. P.; Diver, S. T. Studies directed toward the synthesis of the
scabrosins: validation of a tandem enyne metathesis approach . Tetrahedron 2006, 62,
10528-10540.
313
34 Fukuyama, T.; Nakatsuka, S.; Kishi, Y. Total synthesis of gliotoxin, dehydrogliotoxin, and
hyalodendrin. Tetrahedron 1981, 37, 2045-2078.
35 (a) Nicolaou, K. C.; Hughes, R.; Pfefferkorn, J. A.; Barluenga, S.; Roecker, A. J.
Combinatorial Synthesis through Disulfide Exchange: Discovery of Potent Psammaplin A
Type Antibacterial Agents Active against Methicillin-Resistant Staphylococcus aureus
(MRSA) . Chem. Eur. J. 2001, 7, 4280-4295. (b) Hoshino, O.; Murakata, M.; Yamada, K. An
improved synthesis of psammaplin A. Bioorg. Med. Chem. Lett. 1992, 2, 1561-1562. (c)
Hoshino, O.; Murakata, M.; Yamada, K. A convenient synthesis of a bromotyrosine derived
metabolite, psammaplin A, from Psammaplysilla sp. Bioorg. Med. Chem. Lett. 1992, 12,
1561-1562.
38 a) Takai, K.; Nitta, K.; Utimoto, K. Simple and selective method for aldehydes (RCHO)
.fwdarw. (E)-haloalkenes (RCH:CHX) conversion by means of a haloform-chromous chloride
system. J. Am. Chem. Soc. 1986, 108, 7408-7410. b) Concellon, J. M.; Bernad, P. L.; Mejica,
C. Synthesis of enantiopure (S)-(E)-1-haloalk-1-ene-3-amines with total or very high
diastereoselectivity by halomethylenation of α-amino aldehydes promoted by CrCl2.
Tetrahedron Lett. 2005, 46, 569-571.
40 Mancuso, A. J.; Huang, S. L.; Swern, D. Oxidation of long-chain and related alcohols to
carbonyls by dimethyl sulfoxide "activated" by oxalyl chloride. J. Org. Chem. 1978, 43,
2480-2482.
41 There is evidence that 1 also induces intrinsic apoptosis in addition to the extrinsic apoptotic
pathway involving caspase-8. However, initiation of the extrinsic pathway, which is more
sensitive, appears to be the predominant mechanism of action.
43 Nicolaou, K. C.; Bulger, P. G.; Sarlah, D. Metathesis reactions in total synthesis. Angew.
Chem. Int. Ed. 2005, 44, 4490-4527.
45 Garber, S. B.; Kingsbury, J. S.; Gray, B. L.; Hoveyda, A. H. Efficient and Recyclable
Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122,
314
8168-8179.
46 I took the liberty to name this compound (57) somocystinoic acid because it is a new
compound.
47 (a) Cook, G. R.; Stille, J. R. Stereochemical consequences of the lewis acid-promoted 3-aza-
cope rearrangement of N-alkyl-N-allyl enamines. Tetrahedron 1994, 50, 4105–4124. (b)
Gangadasu, B.; Narender, P.; Kumar, S. B.; Ravinder, M.; Rao, B. A.; Ramesh, C.; Raju, B. C.;
Rao, V. J. Facile and selective synthesis of chloronicotinaldehydes by the Vilsmeier reaction.
Tetrahedron 2006, 62, 8398–8403. (c) Cook, G. R.; Barta, N. S.; Stille, J. R. Lewis acid-
promoted 3-aza-Cope rearrangement of N-alkyl-N-allyl enamines. J. Org. Chem. 1992, 57,
461–467. (d) Couture, A.; Dubiez, R.; Lablache-Combier, A. Secondary enamide and
thioenamide photochemistry. A new spiroannulation method. J. Org. Chem. 1984, 49, 714–
717. (e) Pattenden, G.; Ashweek, N. J.; Baker-Glenn, C. A. G.; Kempson, J.; Walker, G. M.;
Yee, J. G. K. Total synthesis of (-)-ulapualide A, a novel tris-oxazole macrolide from marine
nudibranchs, based on some biosynthesis speculation. Org. Biomol. Chem. 2008, 6, 1478–
1497. (f) Morales, O.; Seide, W.; Watson, S. E. Synthesis of 7,6-fused bicyclic lactams for use
as beta-turn mimics. Synth. Commun. 2006, 36, 965–973.
48 The acetamide, disulfide, or already formed enamide of the other half of the dimeric
intermediate may act as a nucleophile to attack the iminium ion. For examples of nucleohilic
additions to iminium ions, see: Moonen, K.; Stevens, C. V. Synthesis 2005, 20, 3603–3612.
50 Ogawa, T.; Kiji, T.; Hayami, K.; Suzuki, H. Stereospecific one-pot synthesis of enamides and
enimides by the copper iodide promoted vinylic substitution. Chem. Lett. 1991, 1443-1446.
51 4-pentenal (5) was synthesized according to the following paper: Griffith, G. A.; Percy, J. M.;
Pintat, S.; Smith, C. A.; Spencer, N.; Uneyama, E. Towards novel difluorinated sugar
mimetics; syntheses and conformational analyses of highly-functionalised difluorinated
cyclooctenones. Org. Biomol. Chem. 2005, 3, 2701–2712.
52 Instead of a thimble, glass wool and molecular sieves were used to capture any moisture
distilled as an azeotrope.
53 (a) Saunders, W. H. Jr. Acc. Chem. Res. 1976, 9, 19-25. (b) Carey, F. A.; Sundberg, R. J.
Advanced Organic Chemistry Part A 4th Ed, Kluwer Academic / Pleunum Publishers; New
York, 2000, 378-383.
54 Since the publication of this synthesis, the condition for enamide formation was slightly
modified (higher concentration and more equivalents of the aldehyd) to increase the yield
from 41 to 47%.
55 If this is the case, it would not be the first time. See Chapter V for the bioactivity of the
synthetic sample of epiquinamide.
315
57 Lee, S. Process development : fine chemicals from grams to kilograms. Oxford ; New York :
Oxford University Press, 1995.
58 Unfortunately, the 200 mg mark was not met for technically being a multi-hundred-milligram
(MHM) scale synthesis. See Chapter I of this dissertation for the discussion on MHM
synthesis.
59 Still, W. C.; Khan, M.; Mitra, A. Rapid chromatographic technique for preparative separations
with moderate resolution. J. Org. Chem. 1978, 43, 2923-2925.
60 Chavan, S. P.; Chittiboyina, A. G.; Ramakrishna, G.; Tejwani, R. B.; Ravindranathan, T.;
Kamat, S. K.; Rai, B.; Sivadasan, L.; Balakrishnan, K.; Rmalingam, S.; Deshpande, V. H. An
unusual stereochemical outcome of radical cyclization: synthesis of (C)-biotin . Tetrahedron,
2005, 61, 9273-9280.
Chapter VII
316
317
Abstract
to gain insights into the biogenesis of the unique structure and to produce a chelation-free
dimerize this monomer and explanations for their failure are described herein. As a result,
VII.1 Introduction
and UVC (190-280 nm) radiations, pigmented metabolites, including scytonemin (1),2 are
brown indole alkaloid is particularly well suited for absorbing UV light in the blue
Figure VII.1: Light absorption spectrum of scytonemin (1) (taken from ref. 4)
reduced form (2) as shown in Scheme VII.1.2 Almost 150 years after the initial report of
the compound,6 the structure of 1 was finally elucidated in 1993 by Gerwick and
319
coworkers through NMR and MS analysis of 1, as well as those of the ozonolysis product
Due to the unique carbon skeleton and ecological significance of this metabolite,
there is a considerable interest in its biosynthesis.8 The first study in this regard by
Scheme VII.1: Fragmentation of the reduced form of scytonemin (2) through ozonolysis in its
structure elucidation (Gerwick et al. 1993)
Scheme VII.3: Proposed mechanism for the biosynthesis of scytonemin (1) by Balsksus and
Walsh5
gene cluster in 2007 (Scheme VII.2).5,8 The involvement of tryptophan and chorismate in
the biosynthesis were implicated by the presence of genes NpR1261-1266 and by the
presence of genes NpR1260, 1267, and 1269, respectively. Subsequently, Balskus and
Scheme VII.3 is based on the observed activity of the ThDP (thiamine diphosphate)-
of 7 is expected to be necessary due to the steric hindrance by the carboxylic acid in the
Balskus and Walsh were able to synthesize α-hydroxy ketone S-8 without spontaneous
cyclization through an independent route, and this material was used as a synthetic
VII.4).5
As can be seen in Scheme VII.3, there are many steps remain to be determined in
the biosynthesis of scytonemin. For example, despite genetic evidence, the intermediacy
322
Dissection “a” represents the initially conceived approach. Dissection “b” represents the revised
approach. While the biomimetic approach of “a” is attractive, the starting material 16 would
require many steps to prepare.
Scheme VII.6: Predicted mechanism for regioselectivity of aldol condensation between ketone
18 and aldehyde 19
of the monomer 11 has not been established. Furthermore, it is unclear from the gene
this question is to synthesize the putative monomer precursor. If the monomer does not
dimerize spontaneously, synthetic attempts to dimerize it could provide insight into the
chemistry that must occur for this important step in the biosynthesis of scytonemin (1).
323
The product of the condensation of 18 and 19 was exclusively in the keto form (22) and not in
the enol form (11) as predicted by others. Therefore, this product is given a new compound
number (22).
Figure VII.2: Select HMBC, COSY, and ROESY correlations of scytonemin monomer
(22) in DMF-d7
1
H NMR shifts are shown in red and 13C NMR shifts are shown in blue.
324
activities.11 However, due to its strong metal-binding properties (especially for zinc), a
pure sample of 1 could not be obtained for further biological assays. Therefore, a metal-
At the outset, a synthetic route that involves a putative biosynthetic precursor was
VII.3. However, due to the fairly large number of steps 16 would require to synthesize,
another approach was conceived through the use of an alternative intermediate 17 instead
of 16 (Scheme VII.5). In this strategy (dissection b), aldol condensation of ketone 18 with
aldehyde 19 was envisioned. The required regioselectivity was expected based on the
326
(Scheme VII.6).
synthesis of scytonemin monomer began with indole-2-acetic acid (20), which was
ketone 21. After the solvent was changed from diethyl ether (literature12) to THF, the
yield increased to 81% from 50~60%, presumably due to the improved solubility of the
starting material. The rhodium-catalyzed decomposition of the diazo group was sluggish
despite the literature precedence,12 and the product (18) was somewhat unstable during
purification. This material (18) slowly converts to a pink-purple compound that further
converts to an insoluble blue material upon standing in solution. The identity of these
extractor equipped with a plug of glass wool and 3 Å molecular sieves for removal of
(Scheme VII.6) and the other possible regioisomer (22) and over-condensed product (23)
could not be found in the reaction mixture. The major product's color was bright yellow
in a dilute solution and bright orange when dried. This solid had good solubility in DMF
and DMSO, moderate solubility in pyridine and MeOH, and somewhat poor solubility in
CH2Cl2 and EtOAc, and was insoluble in diethyl ether, hexane, and water. This material's
structure was fully elucidated by NMR and HR-MS analyses (Figure VII.2). It should be
noted that it was difficult to remove all of the unreacted 4-hydroxybenzaldehyde (19),
327
and complete purification was only possible by RP-HPLC or by another round of flash
column chromatography.
HMBC and ROESY correlations. The ROESY correlation between the indole NH proton
and meta-proton on the phenol indicated that the geometry of the exocyclic double bond
is E (Figure VII.2). The 1H NMR spectra in both MeOH-d4 and DMF-d7 revealed that the
ketone was exclusively in the keto form (22) and not at all in the enol form (11), as has
been suggested by others.5 This finding is supported by the calculated energy difference
between the keto and enol forms (24 kcal/mol).14 This is possibly due to both the anti-
aromaticity of the cyclopentadiene in the enol form (11) and the electron-donating ability
of the phenol and electron-withdrawing ability of the ketone (Scheme VII.8). Due to the
NMR equivalence of the two pairs of protons on the phenol moiety (H-14, H-14' and
H-15, H-15'; see Scheme VII.9 for numbering), the phenyl ring is expected to be
perpendicular to the tricyclic system, and not in direct conjugation with the rest of the
molecule (Figure VII.3). Finally, the color of a solution of 22 could be changed to bright
red when treated with a base (Figure VII.4). Upon neutralization, the color of the solution
changed back to bright yellow and a TLC analysis showed that 22 was intact.
Now with the monomer 22 available, oxidative dimerization was all that was
necessary for the total synthesis of scytonemin (1). Once the methylene carbon in 22
becomes oxidized to a methine (25), tautomerization should be able to render that carbon
328
both electrophilic and nucleophilic (Scheme VII.9). This dual property of 25 may
Note that the conversion of 2 to 1 is known to occur spontaneously under ambient conditions.2
oxidative conditions, such as O2, UV light, and the combination thereof to oxidize the
methylene carbon selectively. However, even after prolonged exposure to aerial O2 and
UV light (1 week), the monomer 22 remained intact (entry 1 in Table VII.1). Therefore, it
enzymatic reaction or the oxidation of the methylene carbon occurs earlier in the
biosynthesis of 1.
(22) could be synthesized and fed into scytonemin-producing organisms, such as Nostoc
punctiforme, to look for incorporation of the labeled monomer. The synthesis of 22 was
repeated with [1-13C] 4-hydroxybenzaldehyde15 for the feeding study (see Scheme
intermediate and that oxidation of the methylene carbon does not happen earlier in the
biosynthesis. If incorporation does not occur, nothing can be firmly concluded due to the
Given the knowledge that the monomer 22 does not spontaneously oxidize under
ambient conditions, various oxidizing reagents were tested for selective oxidation of 22
as shown in Table VII.1. These reaction conditions included two-electron oxidants, such
conditions yielded scytonemin (1) or its reduced form (2) in amounts detectable by
LCMS or 1H NMR except for entry 13, but because the detected amount was so little, the
amount of 22 was still intact, which presumably meant that oxidative decomposition of 1
(if it was formed at all) was much faster than oxidative dimerization of 22.
In light of these failed reactions, it was hypothesized that at least some of the
failed oxidation reactions were due to chemoselective oxidation of the phenol and not the
cyclopentenone of 22. This hypothesis seemed reasonable due to the electron-rich nature
Note that the yield could be increased to ~70% if a significant excess of 28 was used.
However, the chromatographic purification became increasingly difficult due to the
unreacted 28, which almost completely coeluted with 29.
332
In the gene cluster for the biosynthesis of 1 (Scheme VII.2), there are a few genes
whose roles are not clearly understood in light of the proposed biosynthesis (Scheme
involved in the dimerization step. One hypothesis has been brought forth in P. Proteau's
thesis as follows. The enzyme from NpR1270 is hypothesized to glycosylate the ketone
oxygen, converting the keto form to the enol form, and the glycosylated enol is
transported across the cell membrane to be transported into the extracellular sheaths,
In 30, resonance forms with a negative charge on the cyclopentadiene are preferred due to
creation of aromaticity, which renders the ring susceptible to oxidation due to its rich electron
density. Once oxidized to 31, dimerization could occur in the same manner as in Scheme VII.9.
333
where the electron-rich enol α-carbon is oxidized and coupled to another monomer
was attempted to form 33 (Scheme VII.12). However, the only product isolated from the
the reactive phenolic hydroxy group in 34, glycosylation of the corresponding enolate
was attempted. It is noteworthy that there have not been rigorous synthetic studies on the
butoxide resulted in the loss of the acetyl group and reversion to monomer 22.
a silyl enol ether of the ketone in 36, it was treated with LiHMDS and TBDMSOTf. A
new product was formed in this reaction, but due to its unstable nature, it could not be
334
isolated for full characterization. Therefore, this crude material was treated with m-CPBA
to oxidize the α carbon. However, the resulting product showed intact methylene protons
scytonemin 1 was questioned. While the HMBC correlations noted by Gerwick et al. are
fairly conclusive, they leave some room for three additional possibilities (Figure VII.6).20
same manner as the conversion of 1 to 2 would also result in the formation of nostodione
(3). However, there is no other way to evaluate these structures unless they are actually
synthesized and their spectral data are compared with the authentic compound, which
assume the enol form to enrich the electron density on C-5, presumably due to the
realizable. To eliminate the anti-aromaticity problem and to be able to convert C-5 to sp2
hybridization, a carbon other than C-5 in the cyclopentenone ring could be converted to
an sp3 carbon, thereby breaking the conjugation of the system. For example, the reason
for the ketone 18 being able to undergo an aldol condensation via acid-catalyzed
enolilzation is likely owing to the non-conjugated nature of the enol form of the
cyclopentenone ring. The most feasible site of oxidation level change is the carbonyl
carbon in the cyclopentenone ring. Thus, the synthesis of 22 was modified for the
Neither 44a or 44b suffers from anti-aromaticity. However, an important resonance form of 25
suffers from anti-aromaticity (Scheme VII.9), and it may not be possible to synthesize 25 at all. If
this route is successful, credence to the intermediacy of desilylated version of 44 will be
provided.
reasonable to propose that 10 is the true monomeric precursor, and not 22, provided that
and/or 10 are expected to provide evidence for or against the alternative intermediacy of
transformation may be possible for 42, which would allow access to glycosylated
337
Scheme VII.16: Completion of total synthesis of scytonemin albeit very low yield
monomer 30. Testing for incorporation of 13C-labeled 30 as well as the desilylated version
amounts of 1 and 2. The presence of these two compounds was detected by RP-HPLC
coupled with ESI-MS, in which the retention time and mass spectra of 1 and 2 were
matched with the crude reaction products. However, the integration of the peaks
corresponding to 1 and 2 were too small to be certain that they were not due to some
well known in the literature (Scheme VII.15).21 Therefore, this reaction was revisited to
used for the oxidative dimerization studies were contaminated with significant amounts
of 19. In order to avoid any possible side reactions involving 19, such as reduction of the
oxidant by 19, fine-tuned flash column chromatography conditions (1:19 to 1:4 THF /
hexanes) were developed to completely purify 22. Moreover, being inspired by Li and
this DDQ oxidation (Scheme VII.17). In this attempt, distinct chromatographic peaks
corresponding to 1 and 2 were observed after flash column purification. Semi-prep HPLC
purification of the reaction products is under way at the time of this writing. Due to the
low yield, however, it is still thought that the alternative intermediacy of 43 is possible. It
is possible that an unnatural reaction pathway with a high activation energy is operating
VII.5 Conclusions
of this alkaloid, a few important insights into its biosynthesis were obtained. The
biosynthetic precursor proposed by Walsh (11) was found to be a structure that does not
exist because the equilibrium between 11 (enol form) and 22 (keto form) lies exclusively
towards 22. Through various failed attempts to oxidatively dimerize 22 (although a trace
amount of 1 and 2 appeared to have been produced under certain conditions), this newly
appear to be incompatible with the oxidation required for conversion to the natural
by an enzyme with unusual chemical properties. An alternative conclusion from this work
is that a compound other than 22 is the biosynthetic precursor. In order to further evaluate
feeding study. The intermediacy of an alternative precursor, such as 44, was proposed and
it is being evaluated through chemical synthesis. Finally a small yield of 1 and 2 were
obtained after revisiting the DDQ oxidation of 22. Further studies are required to improve
General
340
Unless noted otherwise, all materials were purchased from commercial sources
and were used without further purification. Anhydrous benzene, 1,4-dioxane, and 1,2-
benzophenone. Pyridine, Et3N, and CH2Cl2 were distilled from CaH. Ac2O was distilled
from quinone. CHCl3 was passed through a column of basic alumina activity I. All
reactions were carried out under dry argon atmosphere unless otherwise noted. Reaction
temperatures herein recorded are external temperatures unless otherwise noted. Flash
chromatorgraphy was performed using EMD silica gel (230-400 mesh) in all cases.23
Thin layer chromatographic (TLC) analysis was performed using EM Science pre-coated
silica gel plates (Merck 60 F254). Melting points were determined on an Electrothermal
Mel-Temp device and are uncorrected. Optical rotations were measured on a Jasco
P-2000 polarimeter. IR spectra were recorded on a Nicolet Magna-IR 550. NMR spectra
were recorded on a Varian Inova spectrometer (500 MHz and 125 MHz, respectively for
1
H and 13C), Varian Mercury spectrometer (400 MHz and 100 MHz, respectively for 1H
and 13C), Varian Mercury spectrometer (300 MHz and 75 MHz, rspectively for 1H and
13
C), or Varian Inova spectrometer (300 MHz and 75 MHz, respectively for 1H and 13C).
1
H and 13C spectra recorded in CDCl3 were referenced to the residual solvent peaks at
7.26 ppm (CHCl3) and 77.0 ppm (CDCl3), respectively. 1H and 13C spectra recorded in
DMF-d7 were referenced to the residual solvent peaks at 8.03 ppm (CHNO(CH3)2) and
163.15 ppm (CDCl3), respectively. High resolution mass spectra were recorded on a
to the literature procedure12 with minor modifications, such as the solvent and
an excess amount of benzoic acid and back titration with NaOH in the presence of phenol
red. To a solution of indole-3-acetic acid (2.500 g, 14.27 mmol) in THF (100 mL) at 0ºC
were added ethyl chloroformate (1.42 mL, 14.9 mmol) and N-methyl morpholine (1.64
mL, 14.9 mmol) successively under stirring. After 15 min of stirring at 0ºC, the solution
of CH2N2 was added in one portion (225 mL, 54.0 mmol). This solution was allowed to
gradually warm to rt over night. A pale brown-orange solution with black tar-like residues
resulted. Most of the volatiles were evaporated under a stream of N2. The residues that
basic alumina (activity III) as the stationary phase (1:9 EtOAc / hexanes to 1:1) to obtain
an orange oil (2.31 g, 11.6 mmol, 81% yield). TLC Rf = 0.39 (1:1 EtOAc / hexanes). 1H
NMR (400 MHz, CDCl3) δ 8.64 (bs, 1H), 7.57 (d, J = 7.8 Hz, 1H), 7.36 (d, J = 8.1 Hz,
1H), 7.28 – 7.19 (m, 1H), 7.19 – 7.12 (m, 1H), 7.04 (d, J = 2.4 Hz, 1H), 5.19 (s, 1H), 3.77
(s, 2H). 13C NMR (101 MHz, CDCl3) δ 194.5, 136.2, 126.9, 123.6, 122.2, 119.7, 118.5,
according to according to the literature procedure.12 However, the yield was consistently
lower than reported (20~40%). TLC Rf = 0.37 (3:7 THF / hexanes). 1H NMR (300 MHz,
CDCl3) δ 8.16 (s, 1H), 7.50 (d, J = 7.6 Hz, 1H), 7.39 (dd, J = 4.7, 3.4 Hz, 1H), 7.25 –
7.12 (m, 2H), 3.54 (s, 2H), 3.53 (d, J = 1.0 Hz, 2H).
best overall yield was obtained when a crude mixture or 18 was used in this reaction. To a
solution of the diazoketone 21 (277 mg, 1.39 mmol) in CH2Cl2 at rt was added Rh2(OAc)4
(6 mg, 0.014 mmol). Vigorous evolution of gas was observed. After 2 hrs, no starting
material was detected by TLC analysis. A reddish orange solution resulted. The reaction
mixture was filtered through a plug of silica gel, which was rinsed with 1:1
EtOAc/hexanes (100 mL). Upon removal of the solvents in vacuo, 155 mg of crude
material was obtained (a mixture of white and orange solids). To a solution of this
343
19 (144 mg, 1.18 mmol), toluenesulfonic acid monohydrate (16 mg, 0.080 mmol), and
molecular sieves 3 Å (155 mg) at rt. This mixture was brought to reflux in a flask
equipped with a condenser. After 14 hrs of refluxing, additional amounts of 19 (27 mg,
0.22 mmol) and molecular sieves 3 Å (60 mg) were added. After 4 hrs, the reaction
mixture was filtered through a plug of silica gel, which was rinsed with EtOAc (200 mL).
Upon concentration under vacuum, this solution was subjected to flash column
chromatography (1:19 THF/hexanes to 1:4) to obtain a bright orange solid (58 mg, 0.211
mmol, 15% yield from 21 over 2 steps). Mp >210ºC. IR (film, KBr) νmax 3375(br), 2963,
2921, 1712, 1599, 1483, 1441, 1274, 1231, 1165, 1123, 745 cm -1. 1H NMR (300 MHz,
DMF-d7) δ 11.11 (s, 1H), 10.22 (s, 1H), 7.70 (d, J = 8.5 Hz, 2H), 7.60 (d, J = 8.1 Hz, 1H),
7.57 (d, J = 8.0 Hz, 1H), 7.23 (ddd, J = 8.2, 7.2, 1.3 Hz, 1H), 7.12 (ddd, J = 7.9, 4.5, 0.6
Hz, 1H), 7.06 (s, 1H), 6.97 (dd, J = 6.1, 2.4 Hz, 2H), 3.56 (s, 2H). 13C NMR (75 MHz,
DMF-d7) δ 205.3, 160.2, 141.6, 141.4, 131.7, 127.6, 127.3, 125.4, 125.2, 124.5, 121.2,
120.9, 120.4, 117.3, 113.9, 36.9. HRMS (EI): calcd for C18H13NO2[M+]: 275.0941, found
compound was synthesized in the same manner as 22. Yellow-green solid. 1H NMR (300
MHz, CDCl3) δ 8.21 (s, 1H), 7.60 (d, J = 8.8 Hz, 2H), 7.55 (d, J = 8.0 Hz, 1H), 7.35 (d, J
= 8.1 Hz, 1H), 7.25 (t, J = 7.2 Hz, 1H), 7.17 (s, 1H), 7.16 (dd, J = 7.4 Hz, 1H), 7.04 (d, J
= 8.7 Hz, 2H), 3.90 (s, 3H), 3.55 (s, 2H). 13C NMR (75 MHz, CDCl3) δ 204.6, 160.3,
140.2, 138.8, 129.6, 128.7, 127.8, 124.3, 124.3, 120.9, 120.3, 119.8, 114.8, 111.8, 55.5,
344
36.5. HRMS (EI): calcd for C19H15NO2Na [M+Na]: 312.0995, found 312.0998 (Δ 1.0
ppm).
(E)-4-((2-oxo-1,2-dihydrocyclopenta[b]indol-3(4H)-ylidene)methyl)phenyl acetate
(34). To a solution of 22 (21 mg, 0.076 mmol) in pyridine (200 μL) was added Ac2O (200
μL) at rt. This solution was left at rt over night. Most of the volatiles were removed in
vacuo, and the residues (dark orange oil) was subjected to flash column chromatography
yield). TLC Rf = 0.64 (1:2 acetone/hexanes). 1H NMR (400 MHz, CDCl3) δ 8.29 (s, 1H),
7.64 (d, J = 8.7 Hz, 2H), 7.55 (d, J = 7.9 Hz, 1H), 7.36 (d, J = 8.2 Hz, 1H), 7.27 (m, 1H),
7.23 (d, J = 8.6 Hz, 2H), 7.16 (m, 1H), 7.14 (s, 1H), 3.56 (s, 2H), 2.36 (s, 3H). 13C NMR
(101 MHz, CDCl3) δ 204.3, 169.5, 150.8, 139.6, 139.1, 134.0, 131.9, 129.1, 124.7, 124.0,
122.9, 122.5, 121.5, 121.0, 119.9, 112.0, 36.4, 21.2. LRMS (ESI, negative): 316 [M-H].
(E)-3-(4-(trimethylsilyloxy)benzylidene)-3,4-dihydrocyclopenta[b]indol-2(1H)-one
(37). To a solution of 22 (40 mg, 0.145 mmol) in THF (1.1 mL) at -78ºC were added 2,6-
lutidine (42 μL, 0.426 mmol) and TMSOTf (52 μL, 0.29 mmol) dropwise under stirring.
After stirring for 5 min at -78ºC, the solution (orange-yellow color) was warmed
immediately to rt. After stirring at rt for 3 hrs, saturated NaHCO3 solution (10 mL) was
added to quench, which was extracted with EtOAc (20 mL x 3). The combined organic
layer was washed with H2O (15 mL) and brine (25 mL) successively, and was dried over
Na2SO4. This product readily decomposed to 22 on a silica gel TLC plate. Upon removal
(0.098 mmol, 67% yield). 1H NMR (300 MHz, CDCl3) δ 8.26 (s, 1H), 7.55 (d, J = 8.3 Hz,
2H), 7.35 (d, J = 8.0 Hz, 1H), 7.26 (t, J = 7.2 Hz, 1H), 7.17 (t, J = 7.2 Hz, 1H), 7.16 (s,
1H), 6.99 (d, J = 8.6 Hz, 2H), 3.53 (s, 2H), 0.36 (s, 9H). 13C NMR (75 MHz, CDCl3) δ
204.6, 156.2, 140.1, 138.9, 129.6, 127.9, 125.5, 124.4, 124.3, 124.3, 120.9, 120.8, 119.8,
95
90
85
80
75
70
65
60
Relative Abundance
55
50
45
40
35
30
25
276.0987
20
280.9824
15
10
268.9824
274.0876 277.1047
5
272.0715 273.0780
267.9909 269.9863 271.0820
274.9935
276.0002 277.0111 278.1091 278.9859 279.9927 281.9864
268 269 270 271 272 273 274 275 276 277 278 279 280 281 282
m /z
95
90
85
80
75
70
65
60
Relative Abundance
55
50
45
40
35
30
25
20 313.1032
15
10
5
310.5966 311.2948 313.2053 314.1069 314.5144 315.1631 315.5026
311.5897 312.3627 316.7042
0
310.5 311.0 311.5 312.0 312.5 313.0 313.5 314.0 314.5 315.0 315.5 316.0 316.5
m/z
1 Garcia-Pitchel, F.; Sherry, N. D.; Castenholz, R. W. Evidence for an ultraviolet sunscreen role
of the extracellular pigment scytonemin in the terrestrial cyanobacterium Chlorogloeopsis sp.
Photochem. Photobiol. 1992, 56, 17-23.
2 Proteau, P. J.; Gerwick, W. H.; Garcia-Pitchel, F.; Castenholz, R. The structure of scytonemin,
an ultraviolet sunscreen pigment from the sheaths of cyanobacteria. Experientia 1993, 49,
825-829.
4 Sinha, R. P.; Häder, D. UV-protectants in cyanobacteria. Plant Sci. 2008, 174, 278-289.
5 Balskus, E. P.; Walsh, C. T. Investigating the initial steps in the biosynthesis of cyanobacterial
sunscreen scytonemin. J. Am. Chem. Soc. 2008, 130, 15260-15261.
7 Kobayashi, A.; Kajiyama, S.; Inawaka, K.; Kanzaki, H.; Kawazu, K. Nostodione A, a novel
mitotic spindle poison from a blue-green alga Nostoc commune. Z. Naturforsch. 1994, 49c,
464–470.
9 Matthews, C. K.; van Holde, K. E.; Ahern, K. G. Biochemistry, 3rd Ed. 2000, Addison Wesley
Longman, Inc.
11 (a) Stevenson, C. S.; Capper, E. A.; Roshak, A. K.; Marquez, B.; Grace, K.; Gerwick, W. H.;
Marshall, L. A. Inflamm. Res. 2002, 51, 112–114. (b) Stevenson, C. S.; Capper, E. A.; Roshak,
A. K.; Marquez, B.; Eichman, C.; Jackson, J. R.; Mattern, M.; Gerwick, W. H.; Jacobs, R. S.;
Marshall, L. A. J. Pharmacol. Exp. Ther. 2002, 303, 858–866.
13 This system worked well in the synthesis of somocystinamide A (see Chapter VI).
18 Proteau, P. J. Oxylipins from temperate marine algae and a photoprotective sheath pigment
from blue-green algae. 1993, M.S. Thesis.
20 All of these structures (38-40) require at least one four-bond HMBC correlation, which is not
unheard of in a polycyclic aromatic system like scytonemin.
21 Yu, J.; Wang, T.; Wearing, X. Z.; Ma, J.; Cook, J. M. Enantiospecific Total Synthesis of (-)-
(E)16-Epiaffinisine, (+)-(E)16-Epinormacusine B, and (+)-Dehydro-16-epiaffinisine as well as
the Stereocontrolled Total Synthesis of Alkaloid G . J. Org. Chem. 2003, 68, 5852-5859.
23 Still, W. C.; Khan, M.; Mitra, A. Rapid chromatographic technique for preparative separations
with moderate resolution. J. Org. Chem. 1978, 43, 2923-2925.
24 Leonard, J.; Lygo, B.; Procter, G. Advanced Practical Organic Chemistry, 2nd Ed. 1998,
Stanley Thomas Ltd.
Chapter VIII
Conclusions
368
369
Abstract
VIII.1 Thesis
Natural products chemistry has benefited from organic synthesis in the areas of
study of natural products. Simultaneously, natural products have been a major driving
force for the advancement of organic synthesis. Given this rich history, direct application
improve productivity in all three areas mentioned above. As a significant added benefit,
new chemical transformations and synthetic strategies may be discovered in the process.
In order to evaluate the thesis that contributions from organic synthesis enhance
the productivity of natural products chemistry, various studies were conducted, which
included applying the techniques learned through the mastery of organic synthesis to
natural product isolation and structure elucidation. The structures of the natural products
that were isolated in this endeavor are summarized below in Figure VIII.1. A chemist
the chance of isolating natural products of interest with a good yield and purity.
Moreover, a synthetic chemist has seen a large number of NMR spectra of many different
classes of compounds, which is an experience that can aid in the structure elucidation
process.
371
many natural products remain undiscovered due to their polarity and solubility in water.
Skills nurtured through organic synthesis were utilized to isolate credneric acid (1), a
metabolite found in the aqueous extract of a Lyngbya sp. from Papua New Guinea. In the
The literature 13C NMR values for 2 contradicted our findings, and this discrepancy was
natural products that do not have a functional group appropriate for traditional
contribution was made to resolve the confusion about the identity of the potent nicotinic
Unfortunately, 5 was found to be inactive in many bioassays tested thus far. However,
through the concept of using a biomimetic starting material (an ornithine derivative), a
very expedient and practical synthetic route to 5 was developed. Had 5 been a bioactive
molecule, this synthesis should have been able to provide sufficient quantities of the
presence of an aqueous base). Finally, a revision of the synthesis was successfully made
to improve the scalability of the first synthesis, thereby providing a sufficient quantity of
7 for in vivo studies. This latter synthesis should enable future investigations of
Figure VIII.2: Natural products that were synthetically investigated during the disseration
research
putative monomeric precursor (9) of this natural product was accomplished in one step
from a known indole derivative. Access to the monomer 9 allowed experiments to show
that dimerization of 9 cannot occur spontaneously under ambient conditions and that such
chemical and biosynthetic insights were gained through attempts to dimerize 9, which
was successful at a small yield albeit a very low yield. The synthesis of 9 and its
3.5
2.5
1.5
0.5
-0.5
210 260 310 360 410 460 510 560
λmax : 255 nm (ε = 3.8 m2/mol); 289 nm (ε = 2.9 m2/mol); 401 nm (ε = 6.1 m2/mol)
Taken in 99:1 MeCN/THF (0.1 mg/ mL)
375
376
The experimental results for the feeding study are summarized in Scheme 1.
Labeled scytonemin's identity was confirmed through its retention time, which coincided
with that of an authentic sample of scytonemin.
It is apparent that monomer 22 can be converted to 1 by a scytonemin-producing
cyanobacterium. Moreover, two different hypotheses can be postulated based on the
absence of singly labeled scytonemin (Figure 1). It is possible that presence of 22
suppressed any production of scytonemin as well as its monomer in the organism through
a feedback mechanism. In this hypothesis, the observation of the unlabeled scytonemin is
explained by small basal production of 1 prior to the UV radiation treatment and the
Surprisingly, singly labeled scytonemin was not observed at all. However, almost equal amounts
of unlabeled and doubly labeled scytonemin were observed. Tolypothrix distorta is a
cyanobacterum that is known to be able to sustain moderate UV radiation through producing
scytonemin.
Figure 1: ESI mass spectra of control and 13C-labeled monomer treatment scytonemin
The peak 543.3 corresponds to scytonemin ([M-H]) while the 545.3 peak corresponds to doubly
labeled scytonemin ([M-H]).
possible that the dimerization process only occurs in the sheath. If so, the labeled
monomer that penetrated into the sheath would be converted only to doubly labeled
scytonemin due to the suppressed production of scytonemin monomer in the cell.
378
The above scheme summarizes the experiments done on N2024 so far. The 13C and 1H
NMR spectra of N-Bz-O-Ac-N2024 were consistent with the expected structure. The
decision was made to protect the free amines as amides because of the expected side
reaction of the amine nitrogens attacking the ester(s) in the hydrolysis step (thereby
forming lactams).
The attempt to perform ethenolysis by olefin metathesis was not successful even after
prolonged reaction time. The 1H NMR spectrum of the resulting material revealed olefin
peaks, indicating that there was no reaction. However, oxidative ozonolysis seemed to
have proceeded well based on the crude 1H NMR taken on the product, in which the
olefin peaks had disappeared. Subsequently, barium hydroxide mediated hydrolysis was
carried out. After work-up, there were a few different compounds in the mixture and upon
HPLC purification, only one of them was isolated in sufficient quantity for NMR
analysis. Despite the 1.7 mm probe, it was not possible to obtain 1H NMR spectrum of
good quality for this material. This disappointing result may be due to two factors; 1)
insufficient starting material, 2) the final product's poor solubility. The latter issue may
have been worsened by the presence of an additional carboxylic acid moiety created by
the oxidative ozonolysis step. For the next time, doubly reductive ozonolysis may be for
performed, where the resulting ketone and aldehyde are reduced to alcohol. One
complication would be the introduction of a new stereo center, the secondary alcohol.
Another approach that could be taken either independently from the above chemistry or
in conjunction is to utilize the newly installed carbon-sensitive probe for 13C NMR
experiments. With this probe, experiments like INADEQUATE and 13C-13C COSY can be
performed without the need for 13C enrichment, thereby clarifying 13C peaks through
decoupling.
379
Pier Lucio Anelli, Fernando Montanari, and Silvio Quici Organic Syntheses, 1993, 8,
367-372.
For the MeOH run, 46.4 mg of (S)-2-methylbutanal was dissolved in 1.5 mL of MeOH.
Immediately the mixture's optical rotation was measured at 1 minute intervals.
40
35
30
25
20
15
10
0
0 20 40 60 80 100 120 140 160 180 200