Next Article in Journal
Paenidigyamycin G: 1-Acetyl-2,4-dimethyl-3-phenethyl-1H-imidazol-3-ium
Next Article in Special Issue
(E)-(1-(4-Ethoxycarbonylphenyl)-5-(3,4-dimethoxyphenyl)-3-(3,4-dimethoxystyryl)-2-pyrazoline: Synthesis, Characterization, DNA-Interaction, and Evaluation of Activity Against Drug-Resistant Cell Lines
Previous Article in Journal
1,1,4,7-Tetramethyldecahydro-1H-cyclopropa[e]azulen-7-ol from the Stembark Chisocheton pentandrus
Previous Article in Special Issue
Synthesis of 2-Cyanopyrimidines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

2,2′-((1,4-Dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid)

1
Department of Chemical Technology of Drugs, Faculty of Pharmacy, Medical University of Gdańsk, Al. Gen. J. Hallera 107, Gdańsk 80-416, Poland
2
Department of Organic Chemistry, Faculty of Pharmacy, Medical University of Gdańsk, Al. Gen. J. Hallera 107, Gdańsk 80-416, Poland
*
Author to whom correspondence should be addressed.
Molbank 2019, 2019(4), M1093; https://doi.org/10.3390/M1093
Submission received: 27 October 2019 / Revised: 19 November 2019 / Accepted: 19 November 2019 / Published: 21 November 2019

Abstract

:
The title compound, 2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid) was synthesized from isoxazolo[3,4-b]quinolin-3(1H)-one and dimethyl acetylenedicarboxylate (DMAD) via a double aza-Michael addition followed by [1,3]-H shifts. The product was characterized by infrared and nuclear magnetic resonance spectroscopy, as well as elemental analysis and high-resolution mass spectrometry (HRMS). The proposed reaction mechanism was rationalized by density functional theory (DFT) calculations.

Graphical Abstract

1. Introduction

Isoxazol-3(2H)-one and isoxazol-5(2H)-one derivatives (Figure 1) represent an extensive class of heterocyclic ring systems found in natural products and building blocks employed in medicinal chemistry. They may be treated as useful tools in organic synthesis since they are small and easy to functionalize molecules that can be utilized to design novel bioactive compounds. It has been proven that these synthetic products exhibit antibacterial [1,2,3,4,5,6,7,8], antifungal [7,8,9,10,11,12,13,14,15,16], antitubercular [3], anticancer [3,6,17,18], antileucemic [5], antinflammatory [19,20,21], antiviral [22], anticonvulsant [1], antioxidant [2,7], and antiandrogenic [23,24] properties. They may act as inhibitors of p38 MAP kinases [25], protein kinase C [26], and protein-tyrosine phosphatase 1B, which consequently cause antiobesity effect [27,28]. They also find applications as GABAA receptor ligands [29] and glutamate receptor agonists [30,31,32], therefore they can affect learning and memory processes. Despite the molecular mechanism of antifungal action of isoxazol-5(2H)-ones such as TAN-950A [16] and drazoxolon [8,15] has not been established, it should be underlined that structurally similar antimicrobial oxazolidones such as posizolid, tedizolid, radezolid and linezolid or cycloserine serve as inhibitors of protein synthesis that prevent binding of N-formylmethionyl-tRNA to the ribosome. [33].
Recently, our research group reported antibacterial [34] and antifungal [35] N-substituted derivatives of 4,6-dimethylisoxazolo[3,4-b]pyridin-3(1H)-one 1, which were obtained in N1-alkylation, N1-acylation, and N1-sulfonation reactions. Moreover, we have implemented compound 1 and its benzo analogue, isoxazolo[3,4-b]quinolone-3(1H)-one 2, to tandem Mannich—electrophilic amination reactions with formaldehyde and (fluoro)quinolones as secondary amines to obtain hybrid quinolone-based quaternary ammonium compounds with proved antibacterial and antibiofilm activities along with enhanced hydrophilic properties [36,37]. Furthermore, our studies aimed at reactivity exploration of isoxazolones proved that under alkaline conditions the said heterocyclic ring system (2) undergoes N1-alkylation followed by bimolecular base-catalyzed acyl-oxygen cleavage (BAC2) and O-alkylation to yield N,O-dialkyl hydroxylamines (Scheme 1) [38]. Finally, we found that compounds 1 and 2 react with α,β-acetylenic carbonyl compounds (Michael acceptors) to give N-vinyl isoxazolones, which transform into N-vinylhydroxylamines by means of BAC2 cleavage of C-O bond. The later react with an excess of Michael acceptors to yield N,O-divinylhydroxylamines (enamines) that undergo [3,3]-sigmatropic rearrangement to give Paal–Knorr intermediates (Scheme 1) [39]. These findings prompted us to examine the aza-Michael addition reaction of isoxazolone 2 to double activated electron-deficient acetylene, i.e., dimethyl acetylenedicarboxylate (DMAD).
Herein, we describe a facile approach to 2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid) 3 along with its characterization by experimental methods such as 1H-NMR, IR, HRMS, and elemental analysis, as well as theoretical DFT calculations.

2. Results and discussion

2.1. Chemistry

We have performed the reaction of isoxazolone 2 with DMAD in anhydrous methanol at room temperature. Triethylamine was used as a base to deprotonate the acidic isoxazolone ring and hence form an ambident nucleophile capable to serve as a Michael donor. We reasoned that in the case of the reaction between compound 2 and double activated acetylene derivatives such as DMAD, dual addition to the Michael acceptor would occur (product A, Scheme 2). Unexpectedly, the isolated compound proved alternate structure as evidenced by 1H-NMR spectrum (Supplementary Materials). The analysis of the spectrum revealed a lack of the aliphatic signal of the methine CH groups as depicted in structure A. Instead, two acidic protons were observed as broad singlet at δ = 13.85. Presumably, the initial step of the reaction sequence involved nucleophilic attacks of isoxazolones on the α,β-unsaturated carbonyl compound. The initially formed intermediate A undergoes base-promoted [1,3]-hydrogen shifts with isoxazolone rings cleavage to yield 2,3-diiminosuccinate derivative 3. The elucidated structure was supported by the results of the high-resolution mass spectrometry and elemental analysis.

2.2. DFT Calculations

Quantum-chemical calculations were carried out to rationalize the formation of product 3. The results obtained with density functional calculations (DFT) revealed that compound 3 is more favorable than the double Michael addition product A, both in gas phase and in the solvent (Table 1). Hence, the data obtained with use of B3LYP, ωB97XD, and APFD methods indicate that isomer 3 is 36.8 to 51.4 kcal/mol more stable than the intermediate A. Albeit the predictions are to some extent sensitive to variation of the functional type applied, the results unequivocally prove that the base-promoted [1,3] proton shifts that comprise isoxazolone ring opening transformations are intensely exothermic.

3. Materials and Methods

3.1. General Methods

All reagents and solvents were purchased from commercial sources (Acros Organics, Geel, Belgium; Alfa-Aesar, Haverhill, MA, USA; or Sigma-Aldrich, Saint Louis, MO, USA) and used without further purification. Isoxazolo[3,4-b]quinolin-3(1H)-one 2 was obtained according to the known procedure [40,41]. Analytical TLC was performed on silica gel Merck 60 F254 plates (0.25 mm) with UV light visualization. Melting point was determined on an X-4 melting point apparatus with a microscope and was uncorrected. The IR spectrum was recorded on a Thermo Scientific Nicolet 380 FT-IR spectrometer. The 1H NMR spectrum was registered on a Varian Unity Plus 500 MHz spectrometer. 1H NMR data were internally referenced to DMSO-d6 (2.50 ppm). The ESI-MS spectra were recorded on Shimadzu single quadrupole LCMS 2010 eV mass spectrometer (Shimadzu, Kyoto, Japan). The HRMS spectra were obtained using Agilent LC/MS Q-TOF 6550 mass spectrometer (Agilent Technologies, Santa Clara, CA, USA). Elemental analysis was performed with Elementar Vario El Cube CHNS (Elementar Analysensysteme GmbH, Langenselbold, Germany).

3.2. Synthesis of 2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid) (3)

Isoxazolo[3,4-b]quinolin-3(1H)-one 2 (0.186 g, 1 mmol), dimethyl acetylenedicarboxylate (0.123 mL, 1 mmol), and triethylamine (0.139 mL, 1 mmol) were dissolved in 10 mL of anhydrous methanol. The reaction was stirred at room temperature and the reaction progress was monitored by TLC (chloroform). After 12 h, the reaction mixture was concentrated to 5 mL under reduced pressure. Upon cooling, the precipitated solid was filtered off and washed with diethyl ether (3 × 3 mL). The product was obtained as a yellow solid. Yield 0.162 g (63%); mp 240 °C (with decomposition); IR (KBr) νmax 2952, 1747, 1736, 1627, 1610, 1574, 1557, 1454, 1204, 1133, 1064, 787, 759 cm–1; 1H-NMR (500 MHz, DMSO-D6) δ 3.76 (s, 6H, OCH3), 7.35 (t, J = 7.8 Hz, 2H, CH), 7.41 (d, J = 7.8 Hz, 2H, CH), 7.79 (t, J = 7.8 Hz, 2H, CH), 7.99 (d, J = 7.8 Hz, 2H, CH), 8.71 (s, 2H, CH), 13.65 (bs, 2H, COOH); ESI-MS positive ionization m/z 515 [M + 1]+, negative ionization m/z 513 [M − 1]; HRMS m/z 515.1194 [M + 1]+ (calcd for C26H19N4O8+, 515.1197); anal. C 59.61, H 3.59, N 10.74%, calcd for C26H18N4O8·0.5H2O, C 59.66, H 3.66, N 10.70%.

3.3. DFT Calculations

All calculations have been performed with the Gaussian 16 [42] using standard algorithms and thresholds. The hybrid Becke-3–Lee–Yang–Parr functional (B3LYP) [43], long-range-corrected hybrid functional ωB97XD [44], as well as Austin–Frisch–Petersson hybrid density functional with dispersion (APFD) [45] were utilized. The bulk solvent effects were taken into account for the DFT calculations by means of polarizable continuum model (IEF-PCM) [46]. The 6-31G+(d) standard basis set has been used in the course of this study. The geometry optimizations for the studied molecules were carried out in their ground states with the inclusion of solvent effects. Vibrational analyses were used to verify that the optimized structures correspond to local minima on the energy surface. Gibbs energies including zero-point corrections, temperature corrections, and vibrational energies were computed for standard conditions (T = 298.15 K, P = 1.0 atm) using the harmonic oscillator approximation.

4. Conclusions

In summary, we have shown that isoxazolone 2 in the presence of triethylamine reacts with highly electrophilic DMAD via double aza-Michael addition followed by [1,3]-H shifts to give 2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid) 3 in good yield. The structure of compound 3 was confirmed by spectroscopic characterization. Finally, compound 3 was proven to be significantly more stable than intermediate A as evidenced by DFT calculations.

Supplementary Materials

The following are available online. 1H-NMR spectrum of compound 3 and HRMS Qualitative Compound Report.

Author Contributions

Conceptualization, J.F. and J.S.; Synthesis, J.F.; Molecule characterization, J.F. and P.W.; NMR data analysis, J.F.; DFT calculation and visualization, K.G. and J.S.; Data evaluation, J.F., K.G., P.W., and J.S.; Writing—original draft preparation, J.F.; Writing—review and editing, J.S.; Supervision and project administration, J.S. All authors read and approved the final manuscript.

Funding

This research was funded by the Funds for Statutory Activity of the Medical University of Gdansk (ST-020038/07).

Acknowledgments

The NMR spectrum was carried out at The Intercollegiate Nuclear Magnetic Resonance Laboratory, Gdańsk University of Technology, Poland. The elemental analysis and HRMS experiments were performed at The Physics-Chemistry Workshops, Faculty of Chemistry, University of Gdańsk, Poland. We thank Prof. Franciszek Sączewski, at the Medical University of Gdańsk, Poland, for his consultation on reaction mechanism.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Ergenc, N.; Capan, G.; Otuk, G. New 4-arylhydrazono-5(4H)-isoxazolone derivatives as possible antibacterial and anticonvulsant agents. Pharmazie 1993, 48, 780–782. [Google Scholar]
  2. Mazimba, O.; Wale, K.; Loeto, D.; Kwape, T. Antioxidant and antimicrobial studies on fused-ring pyrazolones and isoxazolones. Bioorg. Med. Chem. 2014, 22, 6564–6569. [Google Scholar] [CrossRef]
  3. Chande, M.S.; Verma, R.S.; Barve, P.A.; Khanwelkar, R.R.; Vaidya, R.B.; Ajaikumar, K.B. Facile synthesis of active antitubercular, cytotoxic and antibacterial agents: A Michael addition approach. Eur. J. Med. Chem. 2005, 40, 1143–1148. [Google Scholar] [CrossRef]
  4. Yu, M.; Wang, J.; Tang, K.; Shi, X.; Wang, S.; Zhu, W.-M.; Zhang, X.-H. Purification and characterization of antibacterial compounds of Pseudoalteromonas flavipulchra JG1. Microbiology 2012, 158, 835–842. [Google Scholar] [CrossRef]
  5. Wierenga, W.; Evans, B.R.; Zurenko, G.E. Benzisoxazolones: Antimicrobial and antileukemic activity. J. Med. Chem. 1984, 27, 1212–1215. [Google Scholar] [CrossRef]
  6. Saidachary, G.; Veera Prasad, K.; Divya, D.; Singh, A.; Ramesh, U.; Sridhar, B.; China Raju, B. Convenient one-pot synthesis, anti-mycobacterial and anticancer activities of novel benzoxepinoisoxazolones and pyrazolones. Eur. J. Med. Chem. 2014, 76, 460–469. [Google Scholar] [CrossRef]
  7. Padmavathi, V.; Venkata Subbaiah, D.R.C.; Mahesh, K.; Radha Lakshmi, T. Synthesis and Bioassay of Amino-pyrazolone, Amino-isoxazolone and Amino-pyrimidinone Derivatives. Chem. Pharm. Bull. 2007, 55, 1704–1709. [Google Scholar] [CrossRef]
  8. Tanaka, K.; Matsuo, K.; Nakanishi, A.; Jo, M.; Shiota, H.; Yamaguchi, M.; Yoshino, S.; Kawaguchi, K. Syntheses and Antimicrobial Activities of Five-Membered Heterocycles Having a Phenylazo Substituent. Chem. Pharm. Bull. 1984, 32, 3291–3298. [Google Scholar] [CrossRef]
  9. Braunholtz, J.T.; Freeman, P.F.H. Fungicidal Isoxazolones. GB1074803, 5 July 1967. [Google Scholar]
  10. Das, N.P.; Mishra, P.K.; Sahu, S. Fungicidal activity of some substituted 5-isoxazolones. Acta Cienc. Indica Chem. 2011, 37, 239–243. [Google Scholar]
  11. Tabarki, M.A.; Besbes, R. Regioselective ring opening of β-phenylglycidate and aziridine-2-carboxylates with N-alkylhydroxylamines: Synthesis of isoxazolidinones. Tetrahedron 2014, 70, 1060–1064. [Google Scholar] [CrossRef]
  12. Heintz-Buschart, A.; Eickhoff, H.; Hohn, E.; Bilitewski, U. Identification of inhibitors of yeast-to-hyphae transition in Candida albicans by a reporter screening assay. J. Biotechnol. 2013, 164, 137–142. [Google Scholar] [CrossRef]
  13. Miyake, T.; Yagasaki, Y.; Kagabu, S. Potential new fungicides: N-acyl-5-methyl-3(2H)-isoxazolone derivatives. J. Pest. Sci. 2012, 37, 89–94. [Google Scholar] [CrossRef]
  14. Kömürcü, S.G.; Rollas, S.; Yilmaz, N.; Cevikbas, A. Synthesis of 3-methyl-4-[(2,4-dihydro-4-substituted-3H-1,2,4-triazole-3-thione-5-yl)phenylhydrazono]-5-isoxazolone and evaluation of their antimicrobial activities. Drug Metabol. Drug Interact. 1995, 12, 161–169. [Google Scholar] [CrossRef]
  15. Summers, L.A. Agricultural fungicides. V. Preparation of 3-Methyl-4-(3′-pyridylhydrazono)-isoxazol-5-one and related compounds. Aust. J. Chem. 1973, 26, 2723–2725. [Google Scholar] [CrossRef]
  16. Cox, M.; Jahangri, S.; Perkins, M.V.; Prager, R.H. Some Synthetic Approaches to Glutamate AMPA Receptor Agonists Based on Isoxazolones. Aust. J. Chem. 2004, 57, 685–688. [Google Scholar] [CrossRef]
  17. Panathur, N.; Gokhale, N.; Dalimba, U.; Koushik, P.V.; Yogeeswari, P.; Sriram, D. New indole–isoxazolone derivatives: Synthesis, characterisation and in vitro SIRT1 inhibition studies. Bioorg. Med. Chem. Lett. 2015, 25, 2768–2772. [Google Scholar] [CrossRef]
  18. Rollas, S.; Kokyan, Ş.; Koçyiǧit-Kaymakçioǧlu, B.; Özbaş-Turan, S.; Akbuǧa, J. Synthesis and evaluation of cytotoxic activities of some substituted isoxazolone derivatives. Marmara Pharm. J. 2011, 15, 94–99. [Google Scholar] [CrossRef]
  19. Vergelli, C.; Schepetkin, I.A.; Crocetti, L.; Iacovone, A.; Giovannoni, M.P.; Guerrini, G.; Khlebnikov, A.I.; Ciattini, S.; Ciciani, G.; Quinn, M.T. Isoxazol-5(2H)-one: A new scaffold for potent human neutrophil elastase (HNE) inhibitors. J. Enzyme Inhib. Med. 2017, 32, 821–831. [Google Scholar] [CrossRef]
  20. Hou, X.Q.; Gao, Y.W.; Yang, S.T.; Wang, C.Y.; Ma, Z.Y.; Xia, X.Z. Role of macrophage migration inhibitory factor in influenza H5N1 virus pneumonia. Acta Virol. 2009, 53, 225–231. [Google Scholar] [CrossRef]
  21. Laughlin, S.K.; Clark, M.P.; Djung, J.F.; Golebiowski, A.; Brugel, T.A.; Sabat, M.; Bookland, R.G.; Laufersweiler, M.J.; VanRens, J.C.; Townes, J.A.; et al. The development of new isoxazolone based inhibitors of tumor necrosis factor-alpha (TNF-α) production. Bioorg. Med. Chem. Lett. 2005, 15, 2399–2403. [Google Scholar] [CrossRef]
  22. Deng, B.-L.; Hartman, T.L.; Buckheit, R.W., Jr.; Pannecouque, C.; De Clercq, E.; Cushman, M. Replacement of the Metabolically Labile Methyl Esters in the Alkenyldiarylmethane Series of Non-Nucleoside Reverse Transcriptase Inhibitors with Isoxazolone, Isoxazole, Oxazolone, or Cyano Substituents. J. Med. Chem. 2006, 49, 5316–5323. [Google Scholar] [CrossRef]
  23. Ishioka, T.; Tanatani, A.; Nagasawa, K.; Hashimoto, Y. Anti-Androgens with full antagonistic activity toward human prostate tumor LNCaP cells with mutated androgen receptor. Bioorg. Med. Chem. Lett. 2003, 13, 2655–2658. [Google Scholar] [CrossRef]
  24. Ishioka, T.; Kubo, A.; Koiso, Y.; Nagasawa, K.; Itai, A.; Hashimoto, Y. Novel Non-Steroidal/Non-Anilide Type Androgen Antagonists with an Isoxazolone Moiety. Bioorg. Med. Chem. 2002, 10, 1555–1566. [Google Scholar] [CrossRef]
  25. Laufer, S.A.; Margutti, S. Isoxazolone Based Inhibitors of p38 MAP Kinases. J. Med. Chem. 2008, 51, 2580–2584. [Google Scholar] [CrossRef]
  26. Demers, J.P.; Hageman, W.E.; Johnson, S.G.; Klaubert, D.H.; Look, R.A.; Moore, J.B. Selective inhibitors of protein kinase C in a model of graft-vs-host disease. Bioorg. Med. Chem. Lett 1994, 4, 2451–2456. [Google Scholar] [CrossRef]
  27. Kafle, B.; Aher, N.G.; Khadka, D.; Park, H.; Cho, H. Isoxazol-5(4H)one Derivatives as PTP1B Inhibitors Showing an Anti-Obesity Effect. Chem. Asian J. 2011, 6, 2073–2079. [Google Scholar] [CrossRef]
  28. Kafle, B.; Cho, H. Isoxazolone Derivatives as Potent Inhibitors of PTP1B. Bull. Korean Chem. Soc. 2012, 33, 275–277. [Google Scholar] [CrossRef]
  29. Frolund, B.; Kristiansen, U.; Brehm, L.; Hansen, A.B.; Krogsgaard-Larsen, P.; Falch Partial, E. Partial GABAA Receptor Agonists. Synthesis and in Vitro Pharmacology of a Series of Nonannulated Analogs of 4,5,6,7-Tetrahydroisoxazolo[4,5-c]pyridin-3-ol. J. Med. Chem. 1995, 38, 3287–3296. [Google Scholar] [CrossRef]
  30. Tamura, N.; Matsushita, Y.; Kishimoto, S.; Itoh, K. Synthesis and Biological Activity of (S)-2-Amino-3-(2, 5-dihydro-5-oxo-4-isoxazolyl)propanoic Acid (TAN-950 A) Derivatives. Chem. Pharm. Bull. 1991, 39, 1199–1212. [Google Scholar] [CrossRef]
  31. Tamura, N.; Itoh, K.; Iwama, T. Synthesis and Glutamate-Agonistic Acitivity of (S)-2-Amino-3-(2, 5-dihydro-5-oxo-3-isoxazolyl)-propanoic Acid Derivatives. Chem. Pharm. Bull. 1992, 40, 381–386. [Google Scholar] [CrossRef]
  32. Toshi, I.; Yasuo, N.; Norikazu, T.; Setsuo, H.; Akinobu, N. A novel glutamate agonist, TAN-950 A, isolated from streptomycetes. Eur. J. Pharmacol. 1991, 197, 187–192. [Google Scholar] [CrossRef]
  33. Shinabarger, D. Mechanism of action of the oxazolidinone antibacterial agents. Exp. Opin. Investig. Drugs 1999, 8, 1195–1202. [Google Scholar] [CrossRef] [PubMed]
  34. Sączewski, J.; Jalińska, A.; Kędzia, A. New derivatives of 4,6-dimethylisoxazolo[3,4-b] pyridin-3(1H)-one: Synthesis, tautomerism, electronic structure and antibacterial activity. Heterocycl. Commun 2014, 20, 215–223. [Google Scholar] [CrossRef]
  35. Sączewski, J.; Fedorowicz, J.; Kędzia, A.; Ziółkowska-Klinkosz, M.; Jalińska, A. Synthesis and Antifungal Activity of Some 4,6-Dimethylisoxazolo[3,4-b]pyridin-3(1H)-one Derivatives. Med. Chem. 2016, 12, 640–646. [Google Scholar] [CrossRef]
  36. Fedorowicz, J.; Sączewski, J.; Konopacka, A.; Waleron, K.; Lejnowski, D.; Ciura, K.; Tomašič, T.; Skok, Ž.; Savijoki, K.; Morawska, M.; et al. Synthesis and biological evaluation of hybrid quinolone-based quaternary ammonium antibacterial agents. Eur. J. Med. Chem. 2019, 179, 576–590. [Google Scholar] [CrossRef]
  37. Ciura, K.; Fedorowicz, J.; Andrić, F.; Greber, K.E.; Gurgielewicz, A.; Sawicki, W.; Sączewski, J. Lipophilicity Determination of Quaternary (Fluoro)Quinolones by Chromatographic and Theoretical Approaches. Int. J. Mol. Sci. 2019, 20, 5288. [Google Scholar] [CrossRef] [Green Version]
  38. Sączewski, J.; Fedorowicz, J.; Korcz, M.; Sączewski, F.; Wicher, B.; Gdaniec, M.; Konopacka, A. Experimental and theoretical studies on the tautomerism and reactivity of isoxazolo[3,4-b]quinolin-3(1H)-ones. Tetrahedron 2015, 71, 8975–8984. [Google Scholar] [CrossRef]
  39. Sączewski, J.; Fedorowicz, J.; Gdaniec, M.; Wiśniewska, P.; Sieniawska, E.; Drażba, Z.; Rzewnicka, J.; Balewski, Ł. The Elusive Paal–Knorr Intermediates in the Trofimov Synthesis of Pyrroles: Experimental and Theoretical Studies. J. Org. Chem. 2017, 82, 9737–9743. [Google Scholar] [CrossRef]
  40. Sączewski, J.; Hinc, K.; Obuchowski, M.; Gdaniec, M. The Tandem Mannich–Electrophilic Amination Reaction: A Versatile Platform for Fluorescent Probing and Labeling. Chem. Eur. J. 2013, 19, 11531–11535. [Google Scholar] [CrossRef]
  41. Fedorowicz, J.; Sączewski, J.; Drażba, Z.; Wiśniewska, P.; Gdaniec, M.; Wicher, B.; Suwiński, G.; Jalińska, A. Synthesis and fluorescence of dihydro-[1,2,4]triazolo[4,3-a]pyridin-2-iumcarboxylates: An experimental and TD-DFT comparative study. Dyes Pigments 2019, 161, 347–359. [Google Scholar] [CrossRef]
  42. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  43. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Chai, J.-D.; Head-Gordon, M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys. 2008, 128, 084106. [Google Scholar] [CrossRef] [PubMed]
  45. Austin, A.; Petersson, G.A.; Frisch, M.J.; Dobek, F.J.; Scalmani, G.; Throssell, K. A density functional with spherical atom dispersion terms. J. Chem. Theor. Comput. 2012, 8, 4989–5007. [Google Scholar] [CrossRef] [PubMed]
  46. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
Figure 1. The structure of isoxazolone derivatives.
Figure 1. The structure of isoxazolone derivatives.
Molbank 2019 m1093 g001
Scheme 1. Base-promoted reactivity of isoxazolones.
Scheme 1. Base-promoted reactivity of isoxazolones.
Molbank 2019 m1093 sch001
Scheme 2. The synthesis of 3 (2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid).
Scheme 2. The synthesis of 3 (2,2′-((1,4-dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid).
Molbank 2019 m1093 sch002
Table 1. Relative electronic energies (ΔE), and Gibbs free energies (ΔG) for isomers A and 3 calculated using B3LYP, ωB97XD, and APFD density functionals and 6-31G+(d) basis set in vacuum and with PCM (MeOH) solvation model.
Table 1. Relative electronic energies (ΔE), and Gibbs free energies (ΔG) for isomers A and 3 calculated using B3LYP, ωB97XD, and APFD density functionals and 6-31G+(d) basis set in vacuum and with PCM (MeOH) solvation model.
Relative Energy Molbank 2019 m1093 i001 Molbank 2019 m1093 i002
VacuumA3
ΔE (kcal/mol)
B3LYP0−46.5
ωB97XD0−40.5
APFD0−36.8
ΔG (kcal/mol)
B3LYP0−50.9
ωB97XD0−45.7
APFD0−40.6
MeOHA3
ΔE (kcal/mol)
B3LYP0−47.8
ωB97XD0−42.2
APFD0−38.4
ΔG (kcal/mol)
B3LYP0−51.4
ωB97XD0−47.4
APFD0−41.1

Share and Cite

MDPI and ACS Style

Fedorowicz, J.; Gzella, K.; Wiśniewska, P.; Sączewski, J. 2,2′-((1,4-Dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid). Molbank 2019, 2019, M1093. https://doi.org/10.3390/M1093

AMA Style

Fedorowicz J, Gzella K, Wiśniewska P, Sączewski J. 2,2′-((1,4-Dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid). Molbank. 2019; 2019(4):M1093. https://doi.org/10.3390/M1093

Chicago/Turabian Style

Fedorowicz, Joanna, Karol Gzella, Paulina Wiśniewska, and Jarosław Sączewski. 2019. "2,2′-((1,4-Dimethoxy-1,4-dioxobutane-2,3-diylidene)bis(azanylylidene))bis(quinoline-3-carboxylic acid)" Molbank 2019, no. 4: M1093. https://doi.org/10.3390/M1093

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