Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 May 28.
Published in final edited form as: Angew Chem Int Ed Engl. 2018 Apr 26;57(22):6605–6609. doi: 10.1002/anie.201800596

Formation of Aminocyclopentadienes from Silyldihydropyridines: Ring Contractions Driven by Anion Stabilization

Morgan M Walker [a], Shuming Chen [b], Brandon Q Mercado [a], K N Houk [b],, Jonathan A Ellman [a],
PMCID: PMC6040663  NIHMSID: NIHMS979963  PMID: 29570926

Abstract

Highly functionalized aminocyclopentadienes are formed by rearrangement of anions generated from readily prepared 6-silyl-1,2-dihydropyridines by desilylation with fluoride. The scope and generality of the reaction are defined, and diverse transformations are performed on the highly functionalized products. A mechanism and driving force for rearrangement is identified from experiment and DFT calculations.

Keywords: rearrangement, amines, ring contraction, density functional calculations, reaction mechanisms

A new contract(ion)

Densely functionalized aminocyclopentadienes are efficiently accessed by anionic rearrangement of 6-silyl-1,2-dihydropyridines. DFT calculations have been performed to elucidate the mechanism of this transformation.

graphic file with name nihms979963u1.jpg


We previously reported the Rh(I)-catalyzed C–H bond addition/electrocyclization cascade reaction of readily available α,β-unsaturated imines 1 with trimethylsilyl alkynes, providing efficient access to diverse 6-trimethylsilyl-substituted 1,2-dihydropyridines (DHPs) 3 (Scheme 1).[1] Subsequent protonation/desilylation of the silyl-substituted DHPs 3 generated unstabilized ylides 4 that were converted to pharmaceutically relevant tetrahydropyridines 5 and tropanes 6.

Scheme 1.

Scheme 1

Reactions of silyl-substituted DHPs 3.

Upon investigation of the desilylation of DHPs 3 under basic conditions using tetrabutylammonium fluoride hydrate, we were surprised to obtain aminocyclopentadienes 8. This transformation provides rapid entry to densely substituted derivatives that would be difficult to access by other methods.[2] Notably, aminocyclopentanes are present in many natural products[3] and pharmaceuticals such as the recently approved drugs Ticagrelor[4] and Peramivir.[5] We now report on the scope and limitations of this new transformation and demonstrate that products 8 are versatile intermediates for further elaboration. We have determined a mechanism for this transformation from experiment and computation.

A variety of conditions were explored for mediating desilylative rearrangement of silyl DHP 3 (see Table S1). Bu4NF hydrate proved to be the optimal reagent for effecting this transformation, but the presence of water was also necessary. The optimized conditions were next applied to a number of silyl-substituted DHPs 3 (Table 1). We first evaluated DHPs with an ester group at the R5 position to provide aminocyclopentadienes 8a–8e. DHPs 3a with R1, R2 and R3 = Me and 3b with R1 and R2 = Me and R3 = H provided products 8a and 8b, respectively, in high yields. The bicyclic product 8c was similarly obtained in a high yield. Moreover, the aminocyclopentadiene product 8d demonstrates efficient access to densely substituted products with differential substitution at all five of the R1 to R5 sites. Replacement of the methyl ester (8a) with the more sterically encumbered tert-butyl ester (8e) resulted in minimal reduction in yield.

Table 1.

Substrate scope.[a]

graphic file with name nihms979963f8.jpg
[a]

Conditions: 3 (1 equiv), Bu4NF hydrate (1.2 equiv) in THF (0.1 M). Reactions generally proceeded much more rapidly than 18 h, but for evaluating scope were allowed to continue for longer times. Isolated yields after silica gel chromatography.

[b]

X-ray structure shown with anisotropic displacement ellipsoids at the 50% probability level. The pircrylsulfonate counterion and hydrogen atoms omitted for clarity.

[c]

Conditions: Bu4NF (2.0 equiv), 72 h.

Different R1-nitrogen substituents could be incorporated successfully into the aminocyclopentadiene product, including benzyl (8f), alkyl (8g and 8h), and even aryl (8i) N-substituents. The structure of the rearrangement product was rigorously confirmed to be the aminocyclopentadiene motif by X-ray structural analysis of the amine salt of 8i.[6] As demonstrated with aminocyclopentadienes 8f to 8l, an aryl group could be incorporated at the R5 position instead of the ester group, including trifluoromethylphenyl (8f–8i), phenyl (8j), 2-fluorophenyl (8k), and 3-chlorophenyl (8l) groups.

To enhance synthetic utility, the reaction sequence from imine 1 was carried out without isolation of DHPs 3 (Table 2). Previously prepared aminocyclopentadienes 8a and 8f were obtained in 52% and 67% overall yields from 1a, respectively. The additional aminocyclopentadienes, 8m displaying different alkyl substituents from R2 to R4 and bicyclic 8n, further establish that these densely functionalized products can be rapidly prepared from readily available TMS alkynes and imines.

Table 2.

Direct synthesis of aminocyclopentadienes 8 from imine 1.[a]

graphic file with name nihms979963f9.jpg
[a]

Conditions: 1st step: alkyne (2 equiv), from 1.25 to 15 mol % of [RhCl(coe)2]2 with equimolar amounts of 4-Et2PPhNMe2 and from 2 to 48 h depending on DHP 3 synthesized (see Supporting Information). 2nd step: Bu4NF hydrate in THF (0.1 M). Isolated overall yield for sequence based on 1 after silica gel chromatography.

While not all substitution patterns for DHPs 3 underwent clean rearrangement to aminocyclopentadienes 8, the side products obtained for these derivatives provided very useful information for ascertaining a mechanism for this new transformation (vide infra). For DHP 3o with an R5 methyl group, no reaction occurred under the standard conditions, and upon heating to 50 °C, while the rearrangement product 8o was not observed, 1,4-dihydropyridine 9o was obtained in moderate yield (Scheme 2).[7] In contrast, DHP 3p with the electron-rich 4-methoxyphenyl group at the R5 position, gave a mixture of rearrangement product 8p and 1,4-dihydropyridine 9p (entry 2).

Scheme 2.

Scheme 2

DHPs 3 that provide the 1,4-dihydropyridines. Conditions: 3 (1 equiv), Bu4NF hydrate in THF (0.1 M). [a] 18 h at 50 °C. [b] 72 h at rt. Yields determined by 1H NMR spectroscopy with 2,6-dimethoxytoluene as the external standard.

The N-benzyl DHP 3q with an R5 phenyl group also resulted in an unexpected outcome. The aminocyclopentadiene 8q was produced in very low yield with the pyridine byproduct 10 obtained in 52% isolated yield instead (eq 1). Extrusion of the tolyl anion from the N-benzyl substituent likely occurred due to the considerable resonance stabilization achieved by aromatization after generating the anion upon silyl cleavage. In support of this hypothesis, pyridine 10 was not detected for the reaction of DHP 3j, which possesses an N-cyclohexyl instead of an N-benzyl group (see 8j, Table 1). Extrusion of the more basic cyclohexyl anion would be less favorable in this case.

graphic file with name nihms979963e1.jpg (1)

Synthetic transformations were performed upon the aminocyclopentadiene products 8 to demonstrate their utility as intermediates for further elaboration. Diels-Alder reaction of aminocyclopentadiene 8f with N-phenyl maleimide provided the norbornene adduct 11 as a single isomer with the facial and endo selectivity rigorously determined by X-ray crystallography (eq 2).[8,2b]

graphic file with name nihms979963e2.jpg (2)

More extensive elaboration of the fused bicyclic aminocyclopentadiene 8c was also carried out (Scheme 3). Diels-Alder cycloaddition with N-phenyl maleimide and methyl acrylate gave the polycyclic adducts 12 and 13, respectively, in good yields and with very high regio- and diastereoselectivity. Dihydroxylation with catalytic osmium under acidic conditions occurred exclusively at only one of the two double bonds and proceeded with moderate diastereoselectivity to provide 14.[9] Selective reduction of the ester to the corresponding alcohol 15 was accomplished in high yield with DIBAL. Finally, straightforward acetylation of the secondary amine gave amide 16.

Scheme 3.

Scheme 3

Elaboration of aminocyclopentadiene 8c. Reaction conditions: a) N-phenylmaleimide, CH2Cl2; b) methyl acrylate; c) OsO4, NMO, citric acid, tBuOH/H2O; d) DIBAL, CH2Cl2; e) (CH3CO)2O, NEt3, CH2Cl2.

The most closely related transformation is that reported by Fusi and Adamo for ring contraction of highly electron-deficient N-trifluoroacetyl-3,5-dicyano-1,4-dihydropyridines (Scheme 4).[10,11]

Scheme 4.

Scheme 4

Previously reported transformation by Fusi and Adamo.

We investigated the mechanism of the ring-contraction of 3 to aminocyclopentadiene 8 with DFT calculations. The mechanism in Pathway A (Scheme 5), analogous to that proposed by Fusi and Adamo, is consistent with the experimental observation that an electron-withdrawing R5 substituent is necessary to obtain 8 in a good yield. After desilylation of dihydropyridine 3 to yield anion 17, 6π disrotatory electrocylization occurs via TS1 to afford aziridine intermediate 18, with its formal negative charge stabilized by the electron-withdrawing R5 substituent. Aziridine intermediate 18 opens regioselectively in the presence of a proton source to furnish 8. An alternative mechanism, Pathway B, where the C6–N1 bond cleaves heterolytically to yield a carbene intermediate 19 that undergoes 6π-electrocyclization to provide 8 was also considered. We calculated the energy profiles for each pathway with the results shown in Table 3. DFT computations were performed with B3LYP geometry optimizations and M06-2X energy evaluations[12] in implicit CPCM solvent.[13] Computational details and discussion on the possible TMS elimination pathways to give 18 are given in the Supporting Information.[14]

Scheme 5.

Scheme 5

Possible mechanistic pathways from DHP 3 to aminocyclopentadiene 8.

Table 3.

Computed free energies (kcal/mol) at the M06-2X/6-311G++(d,p)-CPCM (THF)// B3LYP/6-31G(d) level of theory of transition states and intermediates along Pathways A and B relative to desilylated DHP 17 (R1 to R4 = Me).

R5 Structure ΔG Structure ΔG
4-CF3-Ph TS1a 5.9 18a −8.2
Ph TS1b 7.9 18b −2.8
CO2Me TS1c 1.2 18c −14.7
Me TS1d 15.3 18d 11.3
4-CF3-Ph TS2a 48.2 19a 18.6

The relative energies of the two pathways A and B were determined with a trifluoromethylphenyl group at the R5 position and with the N-R1 substituent modeled with a methyl group to simplify the calculations (Table 3). The ΔG of the aziridine-forming transition state TS1a was calculated to be 5.9 kcal/mol, indicating the feasibility of Pathway A. In contrast, the ΔG of the carbene-forming transition state TS2a was calculated to be 48.2 kcal/mol, clearly excluding Pathway B from consideration.

For electron-withdrawing substituents at the R5 position the ring-contracting step along Pathway A was calculated to have free energies of activation TS1a–c ranging from 1.2 kcal/mol (R5 = CO2Me) to 7.9 kcal/mol (R5 = Ph), with the length of the forming C2–C6 bond ranging from 1.92 Å (R5 = Ph) to 2.01 Å (R5 = CO2Me). With a methyl group as the R5 substituent, the free energy of activation TS1d is much higher at 15.3 kcal/mol,[15] and the forming C2–C6 bond length of 1.79 Å indicates a considerably later transition state. The formation of 18 was found to be exergonic for all electron-withdrawing R5 substituents, with ΔG for 17a–c to 18a–c ranging from −14.7 kcal/mol for R5 = CO2Me to −2.8 kcal/mol for R5 = Ph. In contrast, for 17d with R5 = Me, the formation of 18d is endergonic with a ΔG of +11.3 kcal/mol. These results indicate that both TS1 and 18 are strongly stabilized by an electron-withdrawing R5 substituent. For 17a with R5 = 4-CF3-Ph, the free energy of activation is 2.0 kcal/mol lower and the reaction is >5 kcal/mol more exergonic than for R5 = Ph, showing that for R5 groups with similar steric profiles, stronger electron-withdrawing ability provides better stabilization.

The calculated energies are consistent with the experimental results. In particular, when 3o with R5 = Me was submitted to the reaction conditions, the aminocyclopentadiene product 8o was not observed and DHP 9o was instead obtained (Scheme 2). This result is in agreement with high endergonicity of 17d to 18d, with direct protonation of 17d occurring to give 9o rather than rearrangement.

The energy profiles of the two regiodivergent pathways, C and D, of opening up 18a are shown in Figure 1, with Pathway C leading to the experimentally observed aminocyclopentadiene isomer 8a. Spontaneous opening of the aziridine ring in 18a would result in a highly basic amide anion, which is calculated to be unstable. This result indicates that a proton source is required for the ring-opening, consistent with the observation that anhydrous fluoride sources result in greatly diminished yields (Table S1). Computed geometries of TS4 involve mainly C–N bond breaking with very little proton transfer. At the level of theory used for geometry optimizations, hydrogen-bonded amide intermediates 20 were located, but these local minima disappear at the higher level of theory used for the energetics (Figure 1), suggesting that the C–N breaking and proton transfer are likely energetically concerted but highly asynchronous. While the energetics are not expected to be accurate, these results do provide information on how hydrogen bonding and proton transfer influence the reaction barrier. At both levels of theory, the calculations show the C–N breaking to be rate-determining, with TS4a being lower than TS4b by 2.2 kcal/mol. This result is in excellent agreement with the experimentally observed product, isomer 8a.

Figure 1.

Figure 1

Energy profiles (in kcal/mol) of Pathways C and D for the regiodivergent aziridine ring-opening of 18a to yield aminocyclopentadiene isomers at the M06-2X/6-311G++(d,p)-CPCM (THF)// B3LYP/6-31G(d)-CPCM (THF) level of theory. Distances denoted in Angstroms.

We sought to provide direct experimental support for the aziridine intermediate 18 based upon the observation that silyl DHP 3p provided a mixture of 8p and 9p albeit in low yield (Scheme 2). We hypothesized that deprotonation of the secondary amine in 8p might result in the aziridine intermediate, which upon protonation could provide a mixture of 8p and 9p. One caveat is that rearrangement of 3 requires that Bu4NF be employed as the hydrate while deprotonation of 8p requires a strong base under anhydrous conditions. Perhaps for this reason, we did not detect the formation of 1,4-DHP 9p upon deprotonation under a variety of conditions. However, deprotonation with BuLi and 12-crown-4[16] at −40 °C followed by protonation with acetic acid gave the aminocyclopentadiene isomer 21p along with 8p, providing significant support for aziridine intermediate 18p (Scheme 6).

Scheme 6.

Scheme 6

Support for aziridine intermediate. [a] Yields determined by 1H NMR spectroscopy with 2,6-dimethoxytoluene as the external standard.

In conclusion, we have discovered a new type of ring-contraction of 6-azacyclohexadienyl anions to obtain densely functionalized aminocyclopentadienes 8 from readily accessible silyl DHPs 3. The scope and limitations of the reaction were defined with many different types of products 8 obtained in good to excellent yields. The mechanism of this transformation was elucidated by DFT computations. A variety of regio- and stereoselective transformations of aminocyclopentadienes demonstrate the versatility of these products as synthetic intermediates.

Supplementary Material

Supplemental Information

Acknowledgments

This work was supported by NIH Grant R35GM122473 (to J.A.E.) and NSF Grant CHE-1059084 (to K.N.H.). We gratefully acknowledge Eric Paulson (Yale) for NMR assignment of 21p and Dr. James Phelan (Yale) for helpful discussions. Density functional theory (DFT) calculations were performed on the Hoffman2 cluster at the University of California, Los Angeles, and the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by the National Science Foundation (Grant OCI-1053575).

Footnotes

Supporting information for this article is given via a link at the end of the document

References

  • 1.a) Chen S, Bacauanu V, Knecht T, Mercado BQ, Bergman RG, Ellman JA. J Am Chem Soc. 2016;138:12664. doi: 10.1021/jacs.6b08355. [DOI] [PMC free article] [PubMed] [Google Scholar]; b) Ischay MA, Takase MK, Bergman RG, Ellman JA. J Am Chem Soc. 2013;135:2478. doi: 10.1021/ja312311k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.For alternative methods to prepare aminocyclopentadienes, see: Wu YT, Flynn B, Schirmer H, Funke F, Müller S, Labahn T, Nötzel M, de Meijere A. Eur J Org Chem. 2004:724.Macaulay JB, Fallis AG. J Am Chem Soc. 1990;112:1136.
  • 3.For leading references and reviews on select natural products that incorporate tertiary carbinamine-based aminocyclopentanes, see: Pactamycin Malinowski JT, Sharpe RJ, Johnson JS. Science. 2013;340:180. doi: 10.1126/science.1234756.Halochlorine and pinnaic acids Liu R, Gutierrez O, Tantillo DJ, Aubé J. J Am Chem Soc. 2012;134:6528. doi: 10.1021/ja300369c.Clive DLJ, Yu M, Wang J, Yeh VSC, Kang S. Chem Rev. 2005;105:4483. doi: 10.1021/cr0406209.Himandrine and himgaline Movassaghi M, Tjandra M, Qi J. J Am Chem Soc. 2009;131:9648. doi: 10.1021/ja903790y.Evans DA, Adams DJ. J Am Chem Soc. 2007;129:1048. doi: 10.1021/ja0684996.Cylindricines Pandey G, Janakiram V. Chem Eur J. 2015;21:13120. doi: 10.1002/chem.201501859.Weinreb SM. Chem Rev. 2006;106:2531. doi: 10.1021/cr050069v.Cephalotaxines Abdelkafi H, Nay B. Nat Prod Rep. 2012;29:845. doi: 10.1039/c2np20037f.Notoamide B Artman GD, Grubbs AW, Williams RM. J Am Chem Soc. 2007;129:6336. doi: 10.1021/ja070259i.
  • 4.Wiviott SD, Steg PG. Lancet. 2015;386:292. doi: 10.1016/S0140-6736(15)60213-6. [DOI] [PubMed] [Google Scholar]
  • 5.McLaughlin MM, Skoglund EW, Ison MG. Expert Opin Pharmacother. 2015;16:1889. doi: 10.1517/14656566.2015.1066336. [DOI] [PubMed] [Google Scholar]
  • 6.Synthetic details, characterization, and molecular structures obtained by X-ray structural analysis are described in the Supporting Information. CCDC 1816814 (8i), 1817118 (11) and 1817123 (14) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre www.ccdc.cam.ac.uk/data_request/cif.
  • 7.Protodesilylation of certain types of 6-silyl-1,2-DHPs provide desilylated DHPs, see: Fallon BJ, Garsi J-B, Derat E, Amatore M, Aubert C, Petit M. ACS Catal. 2015;5:7493.
  • 8.Kim S, Ciufolini MA. Org Lett. 2011;13:3274. doi: 10.1021/ol201236j. [DOI] [PubMed] [Google Scholar]
  • 9.Dupau P, Epple R, Thomas AA, Fokin VV, Sharpless KB. Adv Synth Catal. 2002;344:421. [Google Scholar]
  • 10.Fusi S, Ponticelli F, Ventura A, Adamo MFA. Tetrahedron Lett. 2008;49:5820. [Google Scholar]
  • 11.Azirdinyl cyclopentene derivatives have been prepared by irradiation of N-alkyl pyridiniums in protic solvents. For this photochemical process, an azirdinyl cyclopentallyl cation is proposed as an intermediate: Penkett CS, Simpson ID. Tetrahedron Lett. 2001;42:1179.Penkett CS, Simpson ID. Tetrahedron. 1999;55:6183.Ling R, Mariano PS. J Org Chem. 1998;63:6072. doi: 10.1021/jo980855i.Ling R, Yoshida M, Mariano PS. J Org Chem. 1996;61:4439. doi: 10.1021/jo960316i.Kaplan L, Pavlik JW, Wilzbach KE. J Am Chem Soc. 1972;94:3283.
  • 12.Zhao Y, Truhlar DG. Theor Chem Acc. 2008;120:215. [Google Scholar]
  • 13.Barone V, Cossi M. J Phys Chem A. 1998;102:1995. [Google Scholar]
  • 14.Frisch MJ, et al. Gaussian 09. Gaussian Inc; Wallingford, CT: 2013. [Google Scholar]
  • 15.For a comparable activation barrier found in the analogous theoretical rearrangement of iodabenzene, see: Rawashdeh AM, Parambil PC, Zeng T, Hoffmann R. Am Chem Soc J. 2017;139:7124. doi: 10.1021/jacs.7b03388.
  • 16.Aggarwal VK, Wang MF, Zaparucha A. J Chem Soc, Chem Commun. 1994:87. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental Information

RESOURCES