Next Article in Journal
Design, Synthesis and Bioactive Evaluation of Oxime Derivatives of Dehydrocholic Acid as Anti-Hepatitis B Virus Agents
Previous Article in Journal
Tamoxifen and the PI3K Inhibitor: LY294002 Synergistically Induce Apoptosis and Cell Cycle Arrest in Breast Cancer MCF-7 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Spectrophotometric Analysis of 1-Azafluorenone Derivatives

by
Nicholas H. Angello
,
Robert E. Wiley
,
Christopher J. Abelt
* and
Jonathan R. Scheerer
*
Department of Chemistry, The College of William & Mary, P.O. Box 8795, Williamsburg, VA 23187, USA
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(15), 3358; https://doi.org/10.3390/molecules25153358
Submission received: 10 June 2020 / Revised: 20 July 2020 / Accepted: 21 July 2020 / Published: 24 July 2020
(This article belongs to the Section Photochemistry)

Abstract

:
A new extension for the ‘one pot’ construction of diverse 1-azafluorene derivatives featuring a Diels–Alder/retro-Diels–Alder cycloaddition is reported. Conditions were also determined for oxidation to the derived azafluorenones. The spectrophotometric analysis of five different azafluorenones were performed. Moderate fluorescence was observed with azafluorenone derivatives that bear an imbedded pyridone motif; whereas those bearing substituted pyridines do not fluoresce.

Graphical Abstract

1. Introduction

The spectrophotometric properties of fluorenone have been studied for over a century, and the molecule remains of current interest to a diverse community of physical and organic chemists, photochemists and material scientists [1]. The fundamental photophysical properties for the parent 9-fluorenone (1) continue to be refined [2] and the resulting enhanced understanding has enabled various applications such as incorporating fluorenone-based molecules for resistive memory devices [3]. Recent advances in modern organic photochemistry have returned to fluorenone as well defined molecule that can in some cases provide superior results as compared to other photosensitizers [4,5]. Several analogs of fluorenone have been prepared with heteroatom substitutions on or within the core. The inclusion of more basic functionalities permits the tuning of the donor–acceptor properties. In particular, 3-aminofluorenone derivatives (e.g., 2) have been extensively studied and show both an enhanced quantum yield and strong emissive properties stemming from the intramolecular charge transfer excited state [6]. The emission of these derivatives is strongly quenched by hydrogen-bonding interactions with solvents. The mechanism (in plane/out of plane) and stoichiometry of the interactions with the emissive excited state are currently under investigation [7].
Pyridine analogs of fluorenone have been known for some time as natural products [8]. Renewed isolation efforts have revealed nearly a dozen substituted 4-azafluorenone derivatives related to onychine (3), the simplest member of the family [9,10]. Due to potent biological activities, a number of synthetic azafluorenone derivatives have been prepared and show promising therapeutic potential against several diseases and cancer [11,12]. The analysis of the photophysical properties of azafluorenone derivatives have been more limited; 3-azafluorenone 4 has been spectroscopically characterized [13,14] and 1,8-diazafluorenone has been advanced in the field of forensic analysis for the fluorescent illumination of finger print residue [15]. Since the pyridine motif can impart favorable aqueous solubility properties, azafluorenones containing this structural feature have recently emerged as a new class of biocompatible imaging reagents. The 1-aza- and 2-azafluorenone derivatives 5 and 6 are illustrated as representative examples (Figure 1) [16,17].
We recently developed a new synthetic strategy for the construction of 1-azafluorene derivatives 8 and 9 in one reaction vessel from 2-alkynylbenzaldehyde derivatives (e.g., 7) [18,19]. This manuscript both advances the related methods for the rapid construction of diverse 1-azafluorene structures and determines the photophysical properties of a series of derived azafluoreneones represented by structures 10 and 11. As such, the products of both the synthetic effort and spectroscopic characterization are described. We were particularly interested in comparing the properties of 1-azafluorenone derivatives bearing either the pyridine structure motif in 11 and the pyridone functionality present in 10. The pyridone 10 bears an in-plane N–H bond in close proximity to the carbonyl and we wondered what effect such an orientation would have on the optical properties.

2. Results and Discussion

2.1. Synthesis

In our previous research, we discovered a domino reaction pathway starting from either the diketopiperazine or oxazinone precursors 12 or 13 that enabled a rapid assembly of azafluorene derivatives [18,19]. Although the reaction conditions can vary somewhat for different precursors, both substrates undergo a cascade reaction sequence comprised of an aldol condensation, alkene isomerization, Diels–Alder cycloaddition, and cycloreversion steps (Reactions (1) and (2)). The foundational discovery highlighted in Reaction (1) affords pyridone-containing structures 8 after a base-promoted extrusion of the lactimide bridge. Starting with the lactim-lactone precursor 13, the derived intermediate [2.2.2]-bicycloadduct is not isolable under the reaction conditions (PhMe, 110 °C) and undergoes the cycloreversion and extrusion of carbon dioxide to reveal the 2-methoxy-1-azafluorene product 9a (Reaction (2)). The reactivity and scope for these reaction sequences have been disclosed in initial communications.
Molecules 25 03358 i001
We now wish to report that a modified and analogous reaction sequence with phenyl substituted dihydrooxazinone 14 was achieved using very mild reaction conditions (Reaction (3)). We determined that the entire sequence to give azafluorene product 9b—the aldol, isomerization, [4+2], and retro-[4+2]—could be accomplished with a non-hygroscopic weak fluoride base (TBAT, tetrabutylammonium difluorotriphenylsilicate) at ambient temperatures and in reliable yields (50–70%). To the best of our knowledge, this is the lowest temperature observed for any cycloaddition/cycloreversion sequence with an oxazinone derivative. In practice, due to the propensity for degradation or polymerization on storage, dihydrooxazinone 14 used in this sequence was prepared in situ via Staudinger reductive cyclization from the derived azide precursor (see Supplementary Materials). Azafluorene prepared in this manner gave more consistent results, on average a 70% yield over the complete operation. Overall, the three related domino reaction sequences highlighted in Reactions (1)–(3) enable the rapid preparation of diverse azafluorene derivatives.
Molecules 25 03358 i002
We selected a series of azafluorenes to advance through oxidation to the derived fluorenone products for spectroscopic analysis. In our hands, we found that the fluorenone product 11b could be most conveniently accessed by treating a DMF solution of the tricyclic fluorene 9b with Cs2CO3 and air (Scheme 1). This oxidation proved general and another four products were prepared in an analogous fashion. Thus, five 1-azafluorenone substrates (10a, 10b, 11a11c) were advanced for spectrophotometric analysis. These 1-azafluoreneone derivatives were selected in order to possess a representative mix of both pyridine and pyridone structural motifs. Additionally, varied substitutions at the C2 position, including carbon (Ph, 11b), oxygen (OMe, 11a; O, 10a and 10b) and halogen (Cl, 11c), were included for substrate diversity.

2.2. Optical Properties

The electronic absorption spectra of 10a, 10b, and 11a11c are shown in Figure 2. All have the expected strong, short-wavelength UV bands below 300 nm. Methoxyazafluorenone 11a and the related phenyl derivative 11b have moderate bands in the medium UV (300–330 nm) and very weak bands around 400 nm, giving the solutions of these compounds a yellow color. The substituted azafluorenones, especially the phenyl derivative, have larger molar absorptivities than the pyridones. The pyridone derivatives 10a and 10b show three bands spaced between 300 and 480 nm of similar strength.
The pyridone derivatives 10a and 10b showed moderate fluorescence in ethanol, while the fluorenones 11a11c were essentially non-fluorescent (Figure 3). Quantum yields are 0.04 ± 0.01 for 10a; 0.07 ± 0.01 for 10b.

2.3. Theoretical Calculations

The electronic structures of the ground (DFT, B3YLP, 6-311G+(2d,p)) and the first excited (TD-DFT, B3LYP, 6-31G+(2d,p)) [20] states for 10a, 10b, and 11a11c were calculated with Gaussian 16 [21]. The molecular orbitals for the unsubstituted pyridone 10a and the phenylazafluorenone 11b are representative of the two groups. The frontier molecular orbitals, calculated for the ground state structures, are shown in Figure 4. For both molecule sets, the HOMO and LUMO orbitals are π-systems, whereas the HOMO-1 orbital is an n-type orbital. Electron density in this orbital includes contributions from the carbonyl oxygen and the pyridine nitrogen. In 11c and 11a, the chlorine and methoxy also contribute. For the pyridone set, the HOMO-3 is another n-orbital that is centered on the pyridone carbonyl. The calculations predict that strong, short-wavelength (<280 nm) absorption bands were π→π*, whereas the longer wavelength bands have a partial to mostly n→π* character (Table 1).

2.4. Discussion of Fluorescence

The azafluorenones 11a11c show no appreciable fluorescence, whereas the pyridones 10a and 10b fluoresce abstemiously. Biczόk and coworkers found that rapid intersystem crossing from the lowest singlet excited state (S1) to lower energy triplet levels reduced the fluorescence quantum yields of 3-azafluorenone derivatives to the order of 10−2 [13,14]. For 1-azafluorenones, 11a11c, the lowest energy singlet state (S1) has n→π* character, while the lowest energy triplet state (T1) has only π→π* character. These triplet states lie at 11.9, 14.8 and 14.9 kcal/mol lower in energy than the initial S1 states.
As a result, the typical spin-orbit coupling mechanism may be responsible for rapid intersystem crossing and explain the relative lack of fluorescence.
In contrast to the azafluorenones, the pyridones 10a and 10b show moderate fluorescence. Here, the π,π* triplet states lie at 37.1 and 36.6 kcal/mol lower in energy than the n,π* singlet states. The large energy gap may impede the intersystem crossing. The nature of the fluorescent state for the pyridones is not definite. The structures of the pyridones allow for the possibility of the excited-state intramolecular proton transfer (ESIPT). Proton transfer to the amide carbonyl would give an iminol structure that is similar to 11a (Figure 5, right). Since 11a does not show fluorescence, we conclude that this enol structure is not likely to be the source of the pyridone emission. Proton transfer to the fluorenone carbonyl gives the enol structure (Figure 5, left). This structure has an extended, cross-conjugated π-system. For the emission, the calculations predict bands at 504 nm for 10a and 452 nm for 10b. However, both of these transitions have some n,π* character and zero oscillator strength. For the enol structures, the longest wavelength emission bands are calculated to be at 483 and 489 nm, respectively. These bands are pure π,π* and have non-zero oscillator strengths (both 0.0004). The possibility of ESIPT in the pyridone derivatives remains under investigation.

3. Materials and Methods

All reactions were carried out under an atmosphere of nitrogen in flame-dried or oven-dried glassware with magnetic stirring unless otherwise indicated. Acetonitrile, THF, toluene, and Et2O were degassed with argon and purified by passage through a column of molecular sieves and a bed of activated alumina [22]. Dichloromethane was distilled from CaH2 prior to use. All reagents were used as received unless otherwise noted. Flash column chromatography was performed using SiliCycle siliaflash P60 silica gel (230–400 mesh) [23]. Analytical thin layer chromatography was performed on SiliCycle 60Å glass plates. Visualization was accomplished with UV light, anisaldehyde, ceric ammonium molybdate, potassium permanganate, or ninhydrin, followed by heating. Infrared spectra were recorded using a Digilab FTS 7000 FTIR spectrophotometer. 1H-NMR spectra were recorded on a Varian Mercury 400 (400 MHz) spectrometer and are reported in ppm using solvent as an internal standard (CDCl3 at 7.26 ppm) or tetramethylsilane (0.00 ppm). Proton-decoupled 13C-NMR spectra were recorded on a Mercury 400 (100 MHz) spectrometer and are reported in ppm using solvent as an internal standard (CDCl3 at 77.00 ppm). All compounds were judged to be homogeneous (>95% purity) by 1H and 13C NMR spectroscopy, unless otherwise noted. Mass spectra data analysis was obtained through positive electrospray ionization (w/ NaCl) on a Bruker 12 Tesla APEX–Qe FTICR-MS with an Apollo II ion source. Absorption and fluorescence data were obtained using a fiber optic system with an Ocean Optics Maya CCD detector, a miniature deuterium/tungsten lamp (uv/vis) and a 365 nm LED light source (fluorescence). Relative quantum yields were determined using anthracene as the reference (Φ = 0.30). The emission intensity data was treated as follows: (1) the electronic noise was subtracted, (2) the wavelength values were converted to wavenumbers, (3) the corresponding net intensity values were multiplied by λ2/(λmax)2 to account for the effect of the abscissa-scale transformation and (4) the resulting intensity values were divided by the spectral response of the Hamamatsu S10420 CCD. Electronic structure calculations were conducted using Gaussian 16. Ground state geometries were optimized using the DFT B3YLP method employing the 6-311G+(2d,p) basis set. Excited states were optimized using the TD-DFT B3LYP method employing the 6-31G+(2d,p) basis set.

4. Conclusions

We successfully prepared several 1-azafluorenones and related pyridones using an aldol/isomerization/Diels–Alder cycloaddition/cycloreversion cascade reaction sequence followed by fluorene oxidation. The photophysical properties of these compounds were characterized and modeled through computations. The pyridones fluoresce moderately, but the azafluorenones bearing the pyridine motif do not.

Supplementary Materials

The following are available online: the experimental procedures, characterization data, 1H and the 13C-NMR spectra for all new compounds.

Author Contributions

J.R.S. and C.J.A. conceived and designed the experiments, analyzed the data, and prepared the manuscript. N.H.A. and R.E.W. performed experiments and analyzed the data. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the National Institutes of Health (R15GM107702 to J.R.S.). Additional support was also provided by the Camille and Henry Dreyfus Foundation (Henry Dreyfus Teacher− Scholar Award to J.R.S.).

Acknowledgments

The authors acknowledge William & Mary Research Computing for providing computational resources and/or technical support that have contributed to the results reported within this paper. URL: https://www.wm.edu/it/rc.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Herkstroeter, W.G.; Hammond, G.S.; Lamola, A.A. Mechaisms of Photochemical Reactions in Solution. 28. Values of Triplet Excitation energies of Selected Sensitizers. J. Am. Chem. Soc. 1964, 86, 4537–4540. [Google Scholar] [CrossRef]
  2. Chang, C.W.; Solling, T.I.; Diau, E.W.G. Revisiting the photophysics of 9-fluorenone: Ultrafast time-resolved fluorescence and theoretical studies. Chem. Phys. Lett. 2017, 686, 218–222. [Google Scholar] [CrossRef]
  3. Jia, M.; Li, Y.; Huang, M.; Kim, K.J.; Liu, Y.; Cao, S. Fluorenone-based molecules for resistive memory devices: Tuning memory behavior by adjusting end groups. Synth. Met. 2020, 226, 116431. [Google Scholar] [CrossRef]
  4. Xia, J.B.; Zhu, C.; Chen, C. Visible Light-Promoted Metal-Free C-H Activation: Diarylketone-Catalyzed Selective Benzylic Mono-and Difluorination. J. Am. Chem. Soc. 2013, 135, 17494–17500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Pitts, C.R.; Bloom, M.S.; Bume, D.D.; Zhang, Q.A.; Lectka, T. Unstrained C-C bond activation and directed fluorination through photocatalytically-generated radical cations. Chem. Sci. 2015, 6, 5225–5229. [Google Scholar] [CrossRef] [Green Version]
  6. Biczόk, L.; Berces, T.; Inoue, H. Effects of molecular structure and hydrogen bonding on the radiationless deactivation of singlet excited fluorenone derivatives. J. Phys. Chem. 1999, 103, 3837–3842. [Google Scholar] [CrossRef]
  7. Alty, I.G.; Abelt, C.J. Stereoelectronics of the Hydrogen-Bond-Induced Fluorescence Quenching of 3-Aminofluorenones with Alcohols. J. Phys. Chem. A 2017, 121, 5110–5115. [Google Scholar] [CrossRef]
  8. De Almeida, M.E.L.; Braz-F, R.; Von Bulow, M.V.; Gottlieb, O.R.; Maia, J.G.S. Onychine, an alkaloid from Onychopetalum amazonicum. Phytochemistry 1976, 15, 1186–1187. [Google Scholar] [CrossRef]
  9. Pumsalid, K.; Thaisuchat, H.; Loetchutinat, C.; Nuntasaen, N.; Meepowpan, P.; Pompimon, W. A New Azafluorenone from the Roots of Polyalthia cerasoides and its Biological Activity. Nat. Prod. Commun. 2010, 5, 1931–1934. [Google Scholar] [CrossRef] [Green Version]
  10. Mueller, D.; Davis, R.A.; Duffy, S.; Avery, V.M.; Camp, D.; Quinn, R.J. Antimalarial Activity of Azafluorenone Alkaloids from the Australian Tree Mitrephora diversifolia. J. Nat. Prod. 2009, 72, 1538–1540. [Google Scholar] [CrossRef]
  11. Shi, Y.B.; Gao, S.H. Recent advances of synthesis of fluorenone and fluorene containing natural products. Tetrahedron 2016, 72, 1717–1735. [Google Scholar] [CrossRef]
  12. Ye, F.; Tran, C.; Jullien, L.; Le Saux, T.; Haddad, M.; Michelet, V.; Ratovelomanana-Vidal, V. Synthesis of Fluorescent Azafluorenones and Derivatives via a Ruthenium-Catalyzed 2+2+2 Cycloaddition. Org. Lett. 2018, 20, 4950–4953. [Google Scholar] [CrossRef] [PubMed]
  13. Biczόk, L.; Cser, A.; Nagy, K. Substituent and solvent effects on the photophysical properties of 3-azafluorenone derivatives. J. Photochem. Photobiol. A Chem. 2001, 146, 59–62. [Google Scholar] [CrossRef]
  14. Biczόk, L. Photophysical properties of 3-azafluorenone. React. Kinet. Catal. Lett. 1997, 61, 57–62. [Google Scholar] [CrossRef]
  15. Grigg, R.; Mongkolaussavaratana, T.; Pounds, C.A.; Sivagnanam, S. 1,8-Diazafluorenone and related-compounds-a new reagent for the detection of alpha-amino-acids and latent fingerprints. Tetrahedron Lett. 1990, 31, 7215–7218. [Google Scholar] [CrossRef]
  16. Gao, M.; Su, H.F.; Lin, Y.H.; Ling, X.; Li, S.W.; Qin, A.J.; Tang, B.Z. Photoactivatable aggregation-induced emission probes for lipid droplets-specific live cell imaging. Chem. Sci. 2017, 8, 1763–1768. [Google Scholar] [CrossRef] [Green Version]
  17. Sharma, A.; Umar, S.; Kar, P.; Singh, K.; Sachdev, M.; Goel, A. A new type of biocompatible fluorescent probe AFN for fixed and live cell imaging of intracellular lipid droplets. Analyst 2016, 141, 137–143. [Google Scholar] [CrossRef]
  18. Angello, N.H.; Wiley, R.E.; Elmore, T.G.; Perry, R.S.; Scheerer, J.R. Domino Reaction Sequence for the Synthesis of 2.2.2 Diazabicycloalkenes and Base-Promoted Cycloreversion to 2-Pyridone Alkaloids. Org. Lett. 2018, 20, 5203–5207. [Google Scholar] [CrossRef]
  19. Williamson, J.B.; Smith, E.R.; Scheerer, J.R. A Merged Aldol Condensation, Alkene Isomerization, Cycloaddition/Cycloreversion Sequence Employing Oxazinone Intermediates for the Synthesis of Substituted Pyridines. Synlett 2017, 28, 1170–1172. [Google Scholar] [CrossRef] [Green Version]
  20. Foresman, J.B.; Frish, A.E. Exploring Chemistry with Electronic Structure Methods, 3rd ed.; Gaussian Inc.: Pittsburg, PA, USA, 2015. [Google Scholar]
  21. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; Version 1.1; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  22. Pangborn, A.B.; Giardello, M.A.; Grubbs, R.H.; Rosen, R.K.; Timmers, F.J. Safe and Convenient Procedure for Solvent Purification. Organometallic 1996, 15, 1518–1520. [Google Scholar] [CrossRef]
  23. Still, W.C.; Kahn, M.; Mitra, A. Rapid chromatographic technique for preparative separations with moderate resolution. J. Org. Chem. 1978, 43, 2923–2925. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Fluorenone derivatives and the proposed synthesis and evaluation of 1-azafluorenones.
Figure 1. Fluorenone derivatives and the proposed synthesis and evaluation of 1-azafluorenones.
Molecules 25 03358 g001
Scheme 1. Preparation of the diverse 1-azafluorenones for spectroscopic characterization.
Scheme 1. Preparation of the diverse 1-azafluorenones for spectroscopic characterization.
Molecules 25 03358 sch001
Figure 2. Absorption and long wavelength expansion (inset) of pyridones 10a,b in ethanol and azafluorenones 11ac in EtOAc. 11c (Cl), ——; 11b (Ph), — . . —; 11a (MeO), .; pyridone 10a (H), — —; 10b (F), – – –.
Figure 2. Absorption and long wavelength expansion (inset) of pyridones 10a,b in ethanol and azafluorenones 11ac in EtOAc. 11c (Cl), ——; 11b (Ph), — . . —; 11a (MeO), .; pyridone 10a (H), — —; 10b (F), – – –.
Molecules 25 03358 g002
Figure 3. Normalized fluorescence of pyridinones 10a and 10b in ethanol (10a (H), — —; 10b (F), – – –).
Figure 3. Normalized fluorescence of pyridinones 10a and 10b in ethanol (10a (H), — —; 10b (F), – – –).
Molecules 25 03358 g003
Figure 4. Frontier molecular orbitals (top, LUMO; middle, HOMO; bottom, HOMO-1) for pyridone 10a (left) and phenylazafluorenone 11b (right).
Figure 4. Frontier molecular orbitals (top, LUMO; middle, HOMO; bottom, HOMO-1) for pyridone 10a (left) and phenylazafluorenone 11b (right).
Molecules 25 03358 g004
Figure 5. Tautomers of 10a.
Figure 5. Tautomers of 10a.
Molecules 25 03358 g005
Table 1. The calculated absorption bands (λmax (nm), oscillator strength, % n→π* character) for 10 and 11.
Table 1. The calculated absorption bands (λmax (nm), oscillator strength, % n→π* character) for 10 and 11.
Compound
11c (Cl)λmax277286301306
f × 104448944123
% n010010058
11b (Ph)λmax288306308316322327
f × 104499521594233219907
% n00045052
11a (MeO)λmax279279303306427
f × 1044040364711
% n01001000100
10a (H)λmax295302324349403
f × 1043065743400
% n000100100
10b (F)λmax291306322347399
f × 1043042100619400
% n044416436

Share and Cite

MDPI and ACS Style

Angello, N.H.; Wiley, R.E.; Abelt, C.J.; Scheerer, J.R. Synthesis and Spectrophotometric Analysis of 1-Azafluorenone Derivatives. Molecules 2020, 25, 3358. https://doi.org/10.3390/molecules25153358

AMA Style

Angello NH, Wiley RE, Abelt CJ, Scheerer JR. Synthesis and Spectrophotometric Analysis of 1-Azafluorenone Derivatives. Molecules. 2020; 25(15):3358. https://doi.org/10.3390/molecules25153358

Chicago/Turabian Style

Angello, Nicholas H., Robert E. Wiley, Christopher J. Abelt, and Jonathan R. Scheerer. 2020. "Synthesis and Spectrophotometric Analysis of 1-Azafluorenone Derivatives" Molecules 25, no. 15: 3358. https://doi.org/10.3390/molecules25153358

Article Metrics

Back to TopTop