Next Article in Journal
N,N-Bis(3β-acetoxypregn-5(6)-en-20-on-16α-yl)hydroxylamine
Previous Article in Journal
1,5-(4,4'-Dipyridyl)naphthalene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Syn-Ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate HCl Salt

Department of Chemistry, Durham University, South Road, Durham, DH1 3LE, UK
*
Author to whom correspondence should be addressed.
Molbank 2015, 2015(1), M846; https://doi.org/10.3390/M846
Submission received: 12 December 2014 / Revised: 24 February 2015 / Accepted: 26 February 2015 / Published: 3 March 2015

Abstract

:
This short note describes a one-step synthesis of the title compound from commercially available starting materials and reports its full spectroscopic characterization data.

Graphical Abstract

Introduction

The pharmaceutical and agrochemical industries are under constant pressure to streamline their development cycles to provide new chemical entities required to target human diseases and increase agricultural output [1]. Despite recent breakthroughs in the areas of biologics and protein engineering simple heterocyclic structures still remain principle targets offering rapid modification points, high atom to biological activity ratios, and beneficial pharmacological profiles [2,3,4]. Consequently many synthesis endeavors are focused on accessing new heterocyclic derivatives. Recently, we [5,6] and others [7] have reviewed the available literature focusing on top-selling drug molecules and were able to confirm and analyse the prevalence of small molecular entities containing five- and six-membered heterocyclic rings. Amongst the heteroaromatic scaffolds indoles, imidazoles and pyridines were found to be most common, whereas pyrrolidines, piperazines and piperidines are most frequently encountered in saturated heterocycles. Consequently, it comes with no surprise that chemists in both industry and academia continue to seek new entries into highly versatile heterocyclic building blocks bearing specific functionalities (i.e., halides, esters, amines, etc.) or distinct hydrogen bonding networks. Over the past ten years one key area of our research has been to devise modern, often flow-based routes towards such interesting heterocyclic architectures [8,9,10,11].
In this short note we wish to report on the efficient synthesis of syn-ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate hydrochloride salt, which represents an interesting fused tricyclic system containing several hydrogen bond acceptors and donors.

Experimental Section

To a 100 mL round bottom-flask containing 2-chloro-6-methoxyquinoline-3-carboxaldehyde (1, 2.21 g, 10 mmol) and glycine ethyl ester hydrochloride (2, 1.39 g, 10 mmol) in acetonitrile (20 mL) was added triethylamine (1.4 mL, 10 mmol) in one portion. The resulting suspension was stirred at room temperature for 20 h. After evaporation of the solvent the resulting residue was redissolved in ethyl acetate and extracted with water (3 × 25 mL). The organic layer was dried over sodium sulfate, filtered and evaporated to yield a waxy solid. This material was recrystallised from dichloromethane to furnish a pale yellow amorphous solid (3, 1.95 g, 69% yield, Scheme 1).
Initial analysis of the reaction product by LC-MS and high resolution LC-MS had revealed the molecular formula to be C15H16N2O4. Furthermore IR spectroscopy indicated the presence of a carbonyl stretch (1724.5 cm−1) as well as an amine (3346.8 and 3276.0 cm−1) and a hydroxyl group (2619.2 cm−1) consistent with the tricyclic structure proposed for 3. In addition elemental analysis was consistent with the proposed HCl salt of the parent compound, and further analysis of the result suggest the presence of a hemihydrate species of 3.
As the reaction product was isolated as a single diastereoisomer (1H NMR) further 2-dimensional NMR experiments were used to determine its relative configuration. Crucially, it was found that both methine protons on the pyrrolidine moiety show a correlation in the COSY spectrum with the coupling constant being measured as J = 2.8 Hz, which is indicative of a long-range coupling. NOESY spectral data furthermore revealed a through space coupling of these two methine protons (see Figure 1 and SI for details) indicating the presence of the cis-diastereoisomer. By creating an energy-minimised 3D model (ChemDraw 3D, PerkinElmer, Waltham, MA, USA) of the free base form of cis-3 the distance between these two protons was estimated to be ~3.3 Å.
It is believed that this conformationally stable diastereoisomer is due to the presence of the hydrochloride salt.

Spectroscopic Data

1H-NMR (600 MHz, DMSO-d6) δ 8.38 (s, 1H), 7.82 (d, J = 9.1 Hz, 1H), 7.46 (d, J = 2.8 Hz, 1H), 7.40 (dd, J = 9.2, 2.8 Hz, 1H), 5.95 (s, 1H), 5.34 (d, J = 2.8 Hz, 1H), 4.14 (m, 2H), 3.88 (s, 3H), 3.68 (d, J = 2.8 Hz, 1H), 1.19 (t, J = 7.1 Hz, 3H).
13C-NMR (151 MHz, DMSO-d6) δ 173.8 (C), 158.2 (C), 145.4 (C), 142.6 (C), 137.7 (CH), 134.8 (C), 129.3 (CH), 128.6 (C), 123.2 (CH), 106.4 (CH), 70.7 (CH), 60.9 (CH2), 58.1 (CH), 56.1 (CH3), 14.6 (CH3).
IR (neat, HCl-salt) ν 3346.8 (w), 3276.0 (w), 2619.2 (broad), 1724.5 (s), 1624.4 (m), 1499.3 (m), 1340.9 (m), 1218.3 (s), 1157.4 (s), 1015.8 (s), 876.5 (m), 823.9 (s), 537.9 (m) cm−1.
LC-MS (ESI) m/z = 289.1 (M+H). HR-MS (ESI) calculated for C15H17N2O4 289.1188, found 289.1201 (M+H, Δ = 4.5 ppm).
Elemental analysis: calculated for C15H16N2O4.HCl.0.5H2O C: 53.98%, H: 5.44%, N: 8.39%; measured C: 53.75%, H: 5.18%, N: 8.45%.
Melting range: ~135–140 °C (HCl-salt; CH2Cl2, decomposition).

Supplementary materials

Supplementary File 1Supplementary File 2Supplementary File 3Supplementary File 4

Acknowledgments

Both authors gratefully acknowledge financial support through the Royal Society.

Author Contributions

Both authors contributed equally to the research outlined in this paper and the writing of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scannel, J.W.; Blanckley, A.; Boldon, H.; Warrington, B. Diagnosing the decline in pharmaceutical R&D efficiency. Nat. Rev. Drug Disc. 2012, 11, 191–200. [Google Scholar]
  2. Li, J.J. (Ed.) Heterocyclic Chemistry in Drug Discovery; Wiley: Hoboken, NJ, USA, 2013; ISBN 978-1-118-14890-7.
  3. Silverman, R.B. The Organic Chemistry of Drug Design and Drug Action; Elsevier Academic Press: Burlington, MA, USA, 2004; ISBN 978-0126437324. [Google Scholar]
  4. Wermuth, C.G. (Ed.) The Practice of Medicinal Chemistry, 3rd ed.; Elsevier Academic Press: Waltham, MA, USA, 2008; ISBN 978-0-12-374194-3.
  5. Baumann, M.; Baxendale, I.R.; Ley, S.V.; Nikbin, N. An overview of the key routes to the best selling 5-membered ring heterocyclic pharmaceuticals. Beilstein J. Org. Chem. 2011, 7, 442–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Baumann, M.; Baxendale, I.R. An overview of the synthetic routes to the best selling drugs containing 6-membered heterocycles. Beilstein J. Org. Chem. 2013, 9, 2265–2319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Vitaku, E.; Smith, D.T.; Njardarson, J.T. Analysis of the structural diversity, substitution patterns, and frequency of nitrogen heterocycles among U.S. FDA approved pharmaceuticals. J. Med. Chem. 2014, 57, 10257–10274. [Google Scholar] [CrossRef] [PubMed]
  8. Baumann, M.; Baxendale, I.R.; Ley, S.V.; Smith, C.D.; Tranmer, G.K. Fully automated continuous flow synthesis of 4,5-disubstituted oxazoles. Org. Lett. 2006, 8, 5231–5234. [Google Scholar] [CrossRef] [PubMed]
  9. Baumann, M.; Baxendale, I.R.; Ley, S.V. Synthesis of 3-Nitropyrrolidines via dipolar cycloaddition reactions using a modular flow reactor. Synlett 2010, 5, 749–752. [Google Scholar] [CrossRef]
  10. Baumann, M.; Baxendale, I.R.; Kirschning, A.; Ley, S.V.; Wegner, J. Synthesis of highly substituted nitropyrrolidines, nitropyrrolizines and nitropyrroles via multicomponent-multistep sequences within a flow reactor. Heterocycles 2010, 82, 1297–1316. [Google Scholar] [CrossRef]
  11. Baumann, M.; Baxendale, I.R.; Kuratli, C.; Ley, S.V.; Martin, R.E.; Schneider, J. Synthesis of a drug-like focused library of trisubstituted pyrrolidines using integrated flow chemistry and batch methods. ACS Comb. Sci. 2011, 13, 405–415. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of syn-ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate hydrochloride salt 3.
Scheme 1. Synthesis of syn-ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate hydrochloride salt 3.
Molbank 2015 m846 sch001
Figure 1. Assignment of the relative configuration of 3 based on 2D NMR data and a modelled 3D structure (free base form).
Figure 1. Assignment of the relative configuration of 3 based on 2D NMR data and a modelled 3D structure (free base form).
Molbank 2015 m846 g001

Share and Cite

MDPI and ACS Style

Baumann, M.; Baxendale, I.R. Syn-Ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate HCl Salt. Molbank 2015, 2015, M846. https://doi.org/10.3390/M846

AMA Style

Baumann M, Baxendale IR. Syn-Ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate HCl Salt. Molbank. 2015; 2015(1):M846. https://doi.org/10.3390/M846

Chicago/Turabian Style

Baumann, Marcus, and Ian R. Baxendale. 2015. "Syn-Ethyl 1-hydroxy-7-methoxy-2,3-dihydro-1H-pyrrolo[3,4-b]quinolone-3-carboxylate HCl Salt" Molbank 2015, no. 1: M846. https://doi.org/10.3390/M846

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop