You are on page 1of 5

pubs.acs.

org/JACS Communication

Dearomative Synthetic Entry into the Altemicidin Alkaloids


Claire S. Harmange Magnani and Thomas J. Maimone*
Cite This: J. Am. Chem. Soc. 2021, 143, 7935−7939 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Altemicidin and related Streptomyces-derived monoterpene alkaloids possess dense, highly polar azaindane cores as
well as potent cytotoxic and tRNA synthetase inhibitory properties. The congested α-amino acid motif decorating their presumed
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

iridoid-like core structure has proven to be both a synthetic challenge and a biosynthetic mystery to date. Herein, we report a
Downloaded via IISER THIRUVANANTHAPURAM on January 16, 2023 at 12:09:38 (UTC).

distinct, abiotic strategy to these alkaloids resulting in a concise synthesis of altemicidin from simple chemical feedstocks. Key
chemical findings include the exploitation of a dearomative pyridinium addition and dipolar cycloaddition sequence to
stereospecifically install the quaternary amine moiety, and a chemoselective molybdenum-mediated double reduction to establish the
fully functionalized azaindane nucleus with minimal redox manipulations.

I n a screening effort to identify novel acaricidal and


insecticidal natural products, Takahashi and co-workers
isolated (−)-altemicidin (1) from Streptomyces sioyaensis in
203207 (2) and SB-203208 (3) while searching for novel
microbial tRNA synthetase inhibitors.5,6 Preliminary medicinal
chemistry work involving modification of natural 1 supports
the speculation that the sulfonamide side chains of 2 and 3
1989 (Figure 1A).1−4 Following this report, researchers at
may mimic natural aminoacyl adenosine monophosphate
SmithKline Beecham identified the related alkaloids SB- (AMP) substrates.7,8 This is notable given the importance of
tRNA synthetase inhibitors as antibiotic and antimalarial drug
leads.9 The cytotoxic mechanism of the action of 1, which lacks
the presumed tRNA synthetase-anchoring isoleucyl side chain,
has not been vetted, but we are intrigued by the possibility that
1 serves as an AMP (4) mimic owing to its complementary size
and electronic distribution (Figure 1A). Notably, nucleotide
mimics have touched many areas of modern drug discovery
from antivirals to cancer immunotherapeutics.10
To date, the enzymatic origins of 1 remain unverified with
the exception of the sulfonamide side chain, which Abe has
shown originates from cysteine (Figure 1B).11 While an
iridoid-like monoterpene is a likely source of the 10-carbon
core of 1, we were compelled to use simple feedstock
chemicals in a net reductive dearomative assembly of 1,12
rather than a nature-inspired oxidative approach (Figure 1B).13
We recognized the latent tetrahydropyridine harbored in
altemicidin and envisioned this motif arising from the
hypothetical annulation shown involving nicotinamide and an
oxime derivative. In contrast to fully synthetic routes to 1 and
2, which proceed in approximately 30 steps,14,15 this strategy
could potentially introduce differentiated polar functionality
rapidly, construct the synthetically challenging α-tertiary amine
stereocenter,16 and reduce reliance on protecting group
chemistry and redox manipulations. Herein, we describe a
short route to (±)-1 guided by this blueprint.

Received: April 21, 2021


Published: May 21, 2021

Figure 1. Altemicidin monoterpene alkaloids. (A) Members and


electrostatic potential surfaces. (B) Potential biosynthesis of 1 and
retrosynthetic plan.

© 2021 American Chemical Society https://doi.org/10.1021/jacs.1c04147


7935 J. Am. Chem. Soc. 2021, 143, 7935−7939
Journal of the American Chemical Society pubs.acs.org/JACS Communication

Scheme 1. Total Synthesis of Altemicidin (1)a

a
Reagents and conditions: (a) 5 (1.0 equiv), 6 (1.5 equiv), phenyl chloroformate (1.3 equiv), TMSOTf (0.10 equiv), MeCN, 23 °C, 3 h, 77%; (b)
Pd(PPh3)4 (1 mol %), NDMBA (0.5 equiv), DCM, 23 °C, 16 h, 92%; (c) NaBH4 (4.0 equiv), 2% MeOH/THF, −50 °C, 8 h, 65% (1.7:1 dr); (d)
trifluorotoluene, BHT (5 mg/mL), MWI, 130 °C, 8 h then add LiBr (5.0 equiv), DBU (1.0 equiv), 3 Å mol. sieves, MeOH, 0 °C, 3 h, 45%; (e)
Ac2O (0.2 M), 65 °C, 89%, 12 h; (f) NaH (90%, 20 equiv), MeI (40 equiv), 3 Å mol. sieves, DCM/DMF (4:1), −30 to −10 °C, 2 h, 98%; (g)
Mo(CO)6 (1.05 equiv), MeCN, 87 °C, 4 h then add 11, NaBH3CN (10 equiv), 6 h, 82%; (h) Pd(OAc)2 (10 mol %), acetaldoxime (10 equiv), 1,4-
dioxane, 105 °C, 12 h, 82%; (i) AcCl (1.5 M in EtOH), 100 °C (pressurized tube), 1.5 h, 63%; (j) 2-sulfamoyl acetic acid (1.1 equiv), DCC (1.1
equiv), DMAP (1.1 equiv), DMF, 23 °C, 12 h, 73%; (k) LiI (12 equiv), pyridine, 130 °C, 9 h, 67%. TMSOTf = trimethylsilyl
trifluoromethanesulfonate, TBS = tert-butyldimethylsilyl, NDMBA = 1,3-dimethylbarbituric acid, DBU = 1,8-diazabicyclo[5.4.0]undec-7-ene, MWI
= microwave irradiation, BHT = 2,6-di-tert-butyl-4-methylphenol, DCC = N,N′-dicyclohexylcarbodiimide, DMAP = 4-dimethylaminopyridine.

In our initial synthetic studies, we attempted to add various oxime π-bond failed and led us to the realization that a dipolar
keto-oxime enolates to N-methylpyridinium salts with the cycloaddition strategy was required. Efforts to generate a 1,3-
hopes of trapping the unstable dihydropyridine adduct in a [3 dipole from the allyl oxime through reported palladium-
+ 2] cycloaddition, but these reactions proved fruitless.17 After mediated O → N allyl migration were appealing but
extensive experimentation, we found that phenyl chloroformate unsuccessful.23 The allyl group could, however, be removed
activation of commercially available and inexpensive nicotino- (cat. Pd(PPh3)4, NDMBA) to cleanly afford free oxime 7 in a
nitrile (5) allowed for addition of silylenol ether 6 (prepared high yield (92%). Gratifyingly, microwave heating of this
from commercially available ethyl-2-(hydroxyimino)-3-oxobu- compound induced the requisite [3 + 2] dipolar cycloaddition
tanoate) in 77% on large scales (Scheme 1).18 In addition to in 30% isolated yield. Improving the yield of this trans-
the identification of 6 as a suitable nucleophile for this process, formation was challenging, because 7 had a propensity to
and the optimal chloroformate activator (PhOCOCl), the use revert to nicotinonitrile at elevated temperatures. Additionally,
of catalytic TMSOTf as a coactivator and polar solvent the general instability of the tricyclic product hampered efforts
acetonitrile were key to engendering greater yields and to reduce the ketone. We reasoned that reduction of the
selectivity for addition at the desired C-4 position over those ketone prior to the cycloaddition could potentially alleviate
of C-6.19,20 Notably, even slight variations to these coupling these problems, and that the incorporation of an additional sp3
conditions were highly detrimental: lithium, potassium, center might provide a more optimal transition state for the
sodium, and titanium enolates afforded no product, and dipolar cycloaddition. Low-temperature sodium borohydride
enolates containing larger oxime protecting groups generated reduction of 7 produced alcohol 8 as a mixture of alcohol
only trace amounts of product.21,22 epimers (1.7:1 dr) in 65% yield.24 Compared to 7, the mixture
Our efforts were next directed toward forging the key α- of 8/epi-8 was indeed a superior cyclization substrate,
tertiary amine stereocenter. Attempts to utilize the dihydropyr- undergoing the thermal [3 + 2] cyclization in over 80%
idine as a nucleophilic enamine in an addition reaction to the combined yield. Microwave heating and the addition of the
7936 https://doi.org/10.1021/jacs.1c04147
J. Am. Chem. Soc. 2021, 143, 7935−7939
Journal of the American Chemical Society pubs.acs.org/JACS Communication

radical scavenger (BHT) were essential to this process. Table 1. Studies on the Chemoselective Reduction of the
Interestingly, through careful chromatography, we isolated an Tricyclic Isoxazolidine
isomer of 7 in 5% yield after the palladium-mediated
deallylation, which we suspect is the (E)-oxime geometric
isomer. Upon reduction (NaBH4) of this material and workup
with aqueous acid, we found that the material had directly
cyclized to the tricyclic product suggesting that the depicted
(Z)-oxime isomer is carried through the initial steps of the
synthesis as the major product and that oxime isomerization is
a significant component of the activation barrier for the
thermal cycloaddition process. Given the clean reaction profile
for the cycloaddition of 8, we were able to telescope this
process with phenyl carbamate cleavage by simply adding
methanol and LiBr-DBU.25 From this one-pot transformation,
diastereomerically pure tricyclic isoxazolidine 9, whose
structure was confirmed by X-ray crystallography, could be
isolated in 45% yield. Of note, we have easily generated
multigram quantities of 9 through this process.
While our route to the carbocyclic core of altemicidin
proceeded efficiently (i.e., 4 steps from 5 and 6), we were
cognizant that managing the correct oxidation states and
orchestrating the reactivity of the several polar functional
groups would be critical to realizing an overall concise route to
1. Particularly, direct and chemoselective reduction of the C−
O and N−O bonds in isoxazolidine 9 stood as a primary
hurdle. While reduction efforts began with tricycle 9, we
quickly realized that incorporation of an electron-withdrawing latter. Despite the low conversion, we noted that NaBH3CN
acetyl group was required to activate the N−O bond (vide cleanly intercepted the transient imine without concomitant
infra). Accordingly, acylated tricycle 10 was cleanly generated formation of 20. A report that NaBH3CN reacts with
by heating 9 in neat acetic anhydride at 65 °C (Scheme 1). Mo(CO)6 illuminated a potential problem,27 which was
Other activating groups such as carbamates, trifluoro- and remedied by preforming the active Mo(CO) 3(MeCN)3
trichloroacetates, and even a protected form of the natural complex in situ before introducing 10 and NaBH3CN (entry
sulfonamide side chain were ineffective in downstream 8). Under these conditions, 19 was delivered in 78% yield on
chemistry. Compound 10 could be methylated on nitrogen half-gram scales. While earlier attempts to achieve this
(NaH, MeI) producing intermediate 11 whose structure was reduction with the corresponding N-methyl derivative 11
also verified crystallographically. using the phenylsilane reductant were met with single N−O
With both 10 and 11 accessible, we proceeded to examine bond reduction, we were gratified to observe that 11
the reduction of these compounds (Table 1). Canonical undergoes the desired double reductions under the Mo(0)/
reducing and hydrogenation conditions were first examined, NaBH3CN conditions to give 12 in 82% yield as an
and representative examples are shown. In methanol, samarium inconsequential mixture of O-acylated and deacylated products
diiodide reduction of 10 gave only N−O reduction product 21. (entry 9).
When water was added as a cosolvent (entry 2), extensive To complete the total synthesis, the nitrile moiety was
reduction of the cyclopentane ring occurred (see 22). carefully converted to its corresponding amide using Pd(OAc)2
Hydrogenolysis conditions (H2, Pd/C) completely reduced and acetaldoxime in refluxing dioxane (Scheme 1).28 Extensive
the amino acrylonitrile to methylpiperidine 23 leaving the N− experimentation revealed that the N-acetyl group of 13 could
O bond untouched (entry 3). In contrast, molybdenum only be removed under strongly acidic conditions at elevated
hexacarbonyl in wet acetonitrile at reflux produced dihydro- temperatures (AcCl, EtOH, 100 °C).29 Following this
pyridine 20 (entry 4), indicating that N−O reduction was transformation, the sulfonamide side chain could be attached
accompanied by loss of water and tautomerization rather than via a DCC/DMAP-mediated coupling with acid 15 to provide
further reduction. With the hypothesis that the addition of a ethyl altemicidin (16). With 1 within reach, we found 16 to be
hydride source might intercept the presumed imine robustly resistant to saponification conditions. Despite the
intermediate, phenylsilane, which has reported orthogonality challenges associated with dealkylating an ethyl group,
to Mo0, was included.26 Gratifyingly, these conditions (entry subjecting 16 to lithium iodide in refluxing pyridine provided
5) favored formation of tetrahydropyridine 19 over dihy- the free carboxylic acid. We believe that this challenging ethyl
dropyridine 20. Resubjecting 20 to the reaction conditions did dealkylation is facilitated by the formation of isolable oxazoline
not result in conversion to the desired product, suggesting that intermediates (see 17 and 18) which sterically disencumber
20 is not an intermediate under these conditions. Increasing the ethyl ester to permit dealkylation. Hydration of oxazoline
phenylsilane equivalents improved the formation of 19 relative 18 occurs during concentration of the aqueous, reverse-phase
to 20, but yields remained below 50%. Reductants such as column fractions at elevated temperatures. Overall, altemicidin
sodium triacetoxyborohydride and sodium cyanoborohydride is generated in 67% yield from 16 and in 11 steps from 5 and
were examined and showed some desired reactivity (entries 6 6.
and 7); however, these reactions suffered from significant Our synthesis of 1 leverages a dearomative approach to
production of 20 in the former and poor conversion in the install the core heteroatoms convergently and rapidly and is
7937 https://doi.org/10.1021/jacs.1c04147
J. Am. Chem. Soc. 2021, 143, 7935−7939
Journal of the American Chemical Society


pubs.acs.org/JACS Communication

strategically distinct to past syntheses of these compact, polar REFERENCES


alkaloids. A chemo- and diastereoselective dipolar cyclo- (1) Takahashi, A.; Kurasawa, S.; Ikeda, D.; Okami, Y.; Takeuchi, T.
addition forges the challenging α-tertiary amine stereocenter, Altemicidin, A New Acaricidal and Antitumor Substance I.
which is ultimately converted into altemicidin through a Taxonomy, Fermentation, Isolation and Physico-Chemical and
carefully orchestrated reduction/functionalization sequence. As Biological Properties. J. Antibiot. 1989, 42, 1556.
a result of the robust chemistry described herein, ample (2) Takahashi, A.; Ikeda, D.; Nakamura, H.; Naganawa, H.;
opportunity exists to make deep-seated changes to the cores of Kurasawa, S.; Okami, Y.; Takeuchi, T.; IITAKA, Y. Altemicidin, A
these biologically active natural products as well as procure New Acaricidal and Antitumor Substance II. Structural Determi-
sufficient quantities for further target identification and nation. J. Antibiot. 1989, 42, 1562.
mechanism of action studies. (3) Takahashi, A.; Naganawa, H.; Ikeda, D.; Okami, Y. Structure


Determination of Altemicidin by NMR Spectroscopic Analysis.
Tetrahedron 1991, 47 (22), 3621.
ASSOCIATED CONTENT (4) Isaac, B. G.; Ayer, S. W.; Stonard, R. J. Microorganisms: A
* Supporting Information
sı Remarkable Source of Diverse Chemical Structures for Herbicide
The Supporting Information is available free of charge at Discovery in Secondary-Metabolite Biosynthesis and Metabolism;
Petroski, R. J., McCormick, S. P., Eds.; Springer US: Boston, MA,
https://pubs.acs.org/doi/10.1021/jacs.1c04147. 1992; p 157.
Experimental procedures and spectroscopic data for all (5) Stefanska, A. L.; Cassels, R.; Ready, S. J.; Warr, S. R. SB-2037307
compounds (PDF) and SB-203208, Two Novel Isoleucyl TRNA Synthetase Inhibitors
from a Streptomyces Sp. I. Fermentation, Isolation and Properties. J.
Accession Codes Antibiot. 2000, 53 (4), 357.
CCDC 2078779−2078780 contain the supplementary crys- (6) Houge-Frydrych, C. S. V.; Gilpin, M. L.; Skett, P. W.; Tyler, J.
W. SB-203207 and SB-203208, Two Novel Isoleucyl TRNA
tallographic data for this paper. These data can be obtained Synthetase Inhibitors from a Streptomyces Sp. J. Antibiot. 2000, 53
free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by (4), 364.
emailing data_request@ccdc.cam.ac.uk, or by contacting The (7) Banwell, M. G.; Crasto, C. F.; Easton, C. J.; Karoli, T.; March, D.
Cambridge Crystallographic Data Centre, 12 Union Road, R.; Nairn, M. R.; O’Hanlon, P. J.; Oldham, M. D.; Willis, A. C.; Yue,
Cambridge CB2 1EZ, UK; fax: +44 1223 336033. W. Multi-Component Assembly of the Bicyclic Core Associated with


the tRNA Synthetase Inhibitors SB-203207 and SB-203208.
AUTHOR INFORMATION Application to the Synthesis of Biologically Active Analogues. Chem.
Commun. 2001, 21, 2210.
Corresponding Author (8) For an insightful review on inhibitors of adenylate-forming
Thomas J. Maimone − Department of Chemistry, University enzymes, see: Lux, M. C.; Standke, L. C.; Tan, D. S. Targeting
of CaliforniaBerkeley, Berkeley, California 94720, United Adenylate-Forming Enzymes with Designed Sulfonyladenosine
States; orcid.org/0000-0001-5823-692X; Inhibitors. J. Antibiot. 2019, 72, 325.
Email: maimone@berkeley.edu (9) For examples, see: (a) Hurdle, J. G.; O’Neill, A. J.; Chopra, I.
Prospects for Aminoacyl-tRNA Synthetase Inhibitors as New
Author Antimicrobial Agents. Antimicrob. Agents Chemother. 2005, 49, 4821.
Claire S. Harmange Magnani − Department of Chemistry, (b) Pham, J. S.; Dawson, K. L.; Jackson, K. E.; Lim, E. E.; Pasaje, C. F.
A.; Turner, K. E. C.; Ralph, S. A. Aminoacyl-tRNA synthetases as drug
University of CaliforniaBerkeley, Berkeley, California
targets in eukaryotic parasites. Int. J. Parasitol.: Drugs Drug Resist.
94720, United States; orcid.org/0000-0003-0334-2535 2014, 4, 1.
Complete contact information is available at: (10) (a) Eastman, R. T.; Roth, J. S.; Brimacombe, K. R.; Simeonov,
https://pubs.acs.org/10.1021/jacs.1c04147 A.; Shen, M.; Patnaik, S.; Hall, M. D. Remdesivir: A Review of Its
Discovery and Development Leading to Emergency Use Author-
Notes ization for Treatment of COVID-19. ACS Cent. Sci. 2020, 6, 672.
(b) Berger, G.; Marloye, M.; Lawler, S. E. Pharmacological
The authors declare no competing financial interest.


Modulation of the STING Pathway for Cancer Immunotherapy.
Trends Mol. Med. 2019, 25, 412.
ACKNOWLEDGMENTS (11) Hu, Z.; Awakawa, T.; Ma, Z.; Abe, I. Aminoacyl Sulfonamide
Financial support was provided by the NIH (R01GM136945). Assembly in SB-203208 Biosynthesis. Nat. Commun. 2019, 10, 184.
(12) Huck, C. J.; Sarlah, D. Shaping Molecular Landscapes: Recent
C.S.H.M. thanks the NSF (DGE-1106400) for predoctoral Advances, Opportunities, and Challenges in Dearomatization. Chem.
fellowship support. We acknowledge José R. Hernández- 2020, 6, 1589.
Meléndez for contributions to the optimization of the (13) For recent examples of this strategy from our group, see:
pyridinium addition reaction and thank the UC-Berkeley (a) Hung, K.; Condakes, M. L.; Morikawa, T.; Maimone, T. J.
Amgen Scholars Program for their support. We are Oxidative Entry into the Illicium Sesquiterpenes: Enantiospecific
appreciative for spectroscopic assistance from Dr. Hasan Synthesis of (+)-Pseudoanisatin. J. Am. Chem. Soc. 2016, 138, 16616.
Celik under NIH grant S10OD024998 and Dr. Jeff Pelton (b) Condakes, M. L.; Hung, K.; Harwood, S. J.; Maimone, T. J. Total
under NIH grant GM68933. Dr. Nicholas Settineri is Synthesis of (−)-Majucin and (−)-Jiadifenoxolane A, Complex
acknowledged for X-ray crystallographic analysis wherein Majucin-Type Illicium Sesquiterpenes. J. Am. Chem. Soc. 2017, 139,
support from NIH Shared Instrument Grant (S10- 17783. (c) Hung, K.; Condakes, M. L.; Novaes, L. F. T.; Harwood, S.
J.; Morikawa, T.; Yang, Z.; Maimone, T. J. Development of a Terpene
RR027172) is also acknowledged. We also thank Dr. Ulla
Feedstock-based Oxidative Synthetic Approach to the Illicium
Andersen for mass spectrometry analysis and Prof. Wenjun Sesquiterpenes. J. Am. Chem. Soc. 2019, 141, 3083.
Zhang and Dr. Zhijuan Hu for HPLC assistance. Finally, we (14) For total syntheses, see: (a) Kende, A. S.; Liu, K.; Jos Brands,
are grateful to Prof. Ikuro Abe (University of Tokyo) for K. M. Total Synthesis of (−)-Altemicidin: A Novel Exploitation of the
providing an authentic sample of altemicidin and helpful Potier-Polonovski Rearrangement. J. Am. Chem. Soc. 1995, 117,
discussions. 10597. (b) Hirooka, Y.; Ikeuchi, K.; Kawamoto, Y.; Akao, Y.; Furuta,

7938 https://doi.org/10.1021/jacs.1c04147
J. Am. Chem. Soc. 2021, 143, 7935−7939
Journal of the American Chemical Society pubs.acs.org/JACS Communication

T.; Asakawa, T.; Inai, M.; Wakimoto, T.; Fukuyama, T.; Kan, T. (27) Liu, F.-C.; Sheu, Y.-C.; She, J.-J.; Chang, Y.-C.; Hong, F.-E.;
Enantioselective Synthesis of SB-203207. Org. Lett. 2014, 16, 1646. Lee, G.-H.; Peng, S.-M. Syntheses and Characterizations of Group 6
(15) For approaches toward these natural products, see: (a) Kan, T.; Metal Cyanotrihydroborate Complexes. J. Organomet. Chem. 2004,
Kawamoto, Y.; Asakawa, T.; Furuta, T.; Fukuyama, T. Synthetic 689, 544.
Studies on Altemicidin: Stereocontrolled Construction of the Core (28) Chan, W. C.; Koide, K. Total Synthesis of the Reported
Framework. Org. Lett. 2008, 10, 169. (b) Nasirova, D. K.; et al. Structure of Stresgenin B Enabled by the Diastereoselective
Rearragnement of 2-azanorbornenes to tetrahdrocyclopenta[c]- Cyanation of an Oxocarbenium. Org. Lett. 2018, 20, 7798.
pyridines under the action of activated alkynes − A short pathway (29) The remaining material (33%) in this reaction goes on to form
for construction of the altemicidin core. Tetrahedron Lett. 2017, 58, an unusual β-lactam product (SI-6) further discussed in the
4384. (c) Hayakawa, I.; Nagayasu, A.; Sakakura, A. Toward the Supporting Information. See also: Nisole, C.; Uriac, P.; Huet, J.;
Synthesis of SB-203207: Construction of Four Contiguous Nitrogen- Toupet, L. Reactivity of 4-Aminoazetidin-2-ones: Obtention of gem-
Containing Stereogenic Centers. J. Org. Chem. 2019, 84, 15614. difunctional derivatives by N-1−C-4 Cleavage. J. Chem. Res. (S) 1991,
(16) For relevant reviews on the synthesis of α-tertiary amine- 204.
containing alkaloids, see: (a) Ohfune, Y.; Shinada, T. Enantio- and
Diastereoselective Construction of α,α-Disubstituted α-Amino Acids
for the Synthesis of Biologically Active Compounds. Eur. J. Org. Chem.
2005, 2005, 5127. (b) Hager, A.; Vrielink, N.; Hager, D.; Lefranc, J.;
Trauner, D. Synthetic approaches towards alkaloids bearing α-tertiary
amines. Nat. Prod. Rep. 2016, 33, 491.
(17) Bull, J. A.; Mousseau, J. J.; Pelletier, G.; Charette, A. B.
Synthesis of Pyridine and Dihydropyridine Derivatives by Regio- and
Stereoselective Addition to N-Activated Pyridines. Chem. Rev. 2012,
112, 2642.
(18) Quirion, J. C.; Leclerc, E.; Jubault, P. Product Subclass 18:1,2-
Dihydropyridines, 1,4-Dihydropyridines, and Derivatives. In Science of
Synthesis; Molander, G. A., Ed.; Georg Thieme Verlag: Stuttgart,
2000; p 601.
(19) For selected uses of silyl triflates to promote pyridinium
addition reactions, see: (a) Pabel, J.; Hoesl, C. E.; Maurus, M.; Ege,
M.; Wanner, K. Generation of N-Acylpyridinium Ions from Pivaloyl
Chloride and Pyridine Derivatives by Means of Silyl Triflates. J. Org.
Chem. 2000, 65, 9272. (b) Hoesl, C. E.; Maurus, M.; Pabel, J.;
Polborn, K.; Wanner, K. Generation of chiral N-acylpyridinium ions
by means of silyltriflates and their diastereoselective trapping
reactions: formation of N-acyldihydropyridines and N-acyldihydro-
pyridones. Tetrahedron 2002, 58, 6757. (c) Yamaguchi, R.; Hatano,
B.; Nakayasu, T.; Kozima, S. Triflate Ion-promoted Addition
Reactions of Allylsilane to Quinolines and Isoquinolines Acylated
by Chloroformate Esters. Tetrahedron Lett. 1997, 38, 403. (d) Ref
17.
(20) While this manuscript was under review, Obradors and List
reported an elegant, high-yielding, and C-4 selective silyl triflate-
catalyzed addition reaction of silyl enol ethers and pyridines, see:
Obradors, C.; List, B. Azine Activation via Silylium Catalysis. J. Am.
Chem. Soc. 2021, 143, 6817.
(21) Comins, D. L.; Brown, J. D. Regioselective Addition of
Titanium Enolates to 1-Acylpyridinium Salts. A Convenient Synthesis
of 4-(2-Oxoalkyl)Pyridines. Tetrahedron Lett. 1984, 25, 3297.
(22) While the chiral auxiliary approach reported by Yamada and
Morita appears well-suited for preparing 7 in enantioenriched form,
we were unsuccessful in employing silyl enol ether 6 in this chemistry.
6 did not react under the reported conditions, and added TMSOTf
caused degradation of the chiral auxiliary. For details, see: Yamada, S.;
Morita, C. Face-Selective Addition to a Cation-Complex of a
Pyridinium Salt: Synthesis of Chiral 1,4-dihydropyridines. J. Am.
Chem. Soc. 2002, 124, 8184.
(23) Grigg, R.; Markandu, J. Palladium (II) Catalysed Tandem
[2,3]-Sigmatropic Shift-1,3-Dipolar Cycloaddition Processes in Oxime
O-Allyl Ethers. Tetrahedron Lett. 1991, 32, 279.
(24) Attempts to produce optically active 8 from racemic 7 via
divergent asymmetric reduction or kinetic resolution have not proven
to be successful to date.
(25) Scebach, D.; Thaler, A.; Blaser, D.; Ko, S. Y. Trans-
esterifications with 1,8-Diazabicyclo[5.4.0]Undec-7-Ene/Lithium
Bromide (DBU/LiBr) − Also Applicable to Cleavage of Peptides
from Resins in Merrifleld Syntheses. Helv. Chim. Acta 1991, 74, 1102.
(26) Perez, D.; Greenspoon, N.; Keinan, E. Molybdenum(0)-
Catalyzed Reductive Dehalogenation of.Alpha.-Halo Ketones with
Phenylsilane. J. Org. Chem. 1987, 52, 5570.

7939 https://doi.org/10.1021/jacs.1c04147
J. Am. Chem. Soc. 2021, 143, 7935−7939

You might also like