Next Article in Journal
Robust Synthesis of Ciprofloxacin-Capped Metallic Nanoparticles and Their Urease Inhibitory Assay
Next Article in Special Issue
Tandem Catalysis of an Aldol-‘Click’ Reaction System within a Molecular Hydrogel
Previous Article in Journal
The Chemical Composition of Achillea wilhelmsii C. Koch and Its Desirable Effects on Hyperglycemia, Inflammatory Mediators and Hypercholesterolemia as Risk Factors for Cardiometabolic Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Ethano-Bridged Diazapolycenes as Potential Precursors for Diazapolycenes and Their Properties

Department of Pharmacy, College of Pharmacy, Yeungnam University, Gyeongsan 38541, Korea
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(4), 407; https://doi.org/10.3390/molecules21040407
Submission received: 4 February 2016 / Revised: 15 March 2016 / Accepted: 19 March 2016 / Published: 25 March 2016
(This article belongs to the Special Issue New Molecular Materials)

Abstract

:
A series of ethanodiazapolycenes were prepared in 87%–89% yields by Friedländer reactions of three o-aminoarenecarbaldehydes with bicyclo[2.2.2]octane-2,5-dione and their spectral, thermal, and structural properties were studied. Subsequent attempts to convert them to diazapolycenes have proved unsuccessful.

Graphical Abstract

1. Introduction

The polycenes, represented by pentacene (1a) have long been of interest in the area of semiconductors with potential applications in organic thin-film transistors (OTHFTs) [1,2,3], and organic light-emitting diodes [4]. In order to overcome the major drawbacks of pentacene such as its insolubility at room temperature and rapid degradation [5], continuous efforts have been pursued to prepare soluble precursors that can be thermally converted to pentacene [6], the introduction of substituent(s) on the basic skeleton [7,8,9], and substitution of the benzene moiety by isosteric aromatics such as thiophene (compound 2) [10], and pyridine (compounds 3 [11] and 4 ([12]) (Figure 1).
Theoretically, introduction of a nitrogen atom significantly decreases the frontier molecular orbital energy and hence improves the stability in air and to light [13]. This has led to the synthesis of large numbers of N-heteroacenes with pyrazine units in recent years [14,15]. The 1-azapentacenes with bis(triisopropylsilyl)ethynyl groups such as 3 in fact showed potential for high-performance organic semiconductors with superior structure-activity characteristics that the corresponding pentacene derivative, 6,13-bis[(triisopropylsilyl)ethynyl]pentacene [16]. Recently, we reported a synthesis and the properties of 5-azapentacene (1b, Figure 1) [17], which might open a new vista for the studies on azapolycenes. As part of our ongoing studies on azaaromatics [17,18,19], we describe herein a preparation of ethano-bridged diazapolycenes as dimeric diazapolycenes and our attempts at the preparation of the corresponding diazapolycenes.

2. Results and Discussions

2.1. Synthesis

Synthesis of the compounds was straightforward as shown in Scheme 1. The Friedländer reactions of o-aminoarenecarbaldehydes 5 with diketone 6 afforded the corresponding ethano-bridged diazapolycenes 7 in 87%–89% yields. The prerequisite starting 2-aminobenzaldehyde (1a) [20], 3-aminonaphthalene-2-carbaldehyde (5b) [21], 3-aminoanthracene-2-carbaldehyde (5c) [17], and bicyclo[2.2.2]octane-2,5-dione (6) [22] were prepared employing the corresponding previously reported methods.

2.2. Attempts to Convert 7 to the Corresponding Diazapolycenes 8

Catalytic dehydrogenation and/or dealkylation [23,24,25,26] and vacuum pyrolysis [27,28,29,30,31] of cyclohexenes have been commonly employed to build up benzene rings, in which the former proceeds via elimination of alkyl substituent(s) and the latter via retro-Diels-Alder reaction. However, all attempts, including catalytic dehydrogenation (10% Pd/C in nitrobenzene at 200 °C for 8 h) and vacuum pyrolysis (0.02 mmHg, 600 °C, 45–60 min) of the compounds 7 leading to the corresponding diazapolycenes 8 have been as yet unsuccessful (Scheme 2). It should be noted that vacuum pyrolysis of 7a,b resulted in sublimation of the compounds which remain unchanged, while 7c decomposed.

2.3. Spectroscopic Properties

1H-NMR spectral data of selected protons are summarized in Table 1. All the proton resonances were assigned by comparison with previously reported data for related compounds such as quinoline [32], benzo[g]quinoline [33], and 5-azapentacene [17] and by double quantum H-H COSY experiments. The resonances of H4 and the proton at the peri-position in the most of the polypyridines and polyquinolines, are characteristic and have been used as a diagnostic probe for understanding the structural information. The resonances of H4 were shifted downfield by 0.73 ppm for 7b compared to 7a, and 0.40 ppm for 7c compared to 7b. These values are comparable to those of related compounds such as quinoline, benzo[g]quinoline, and 5-azapentacene, reflecting increased delocalization of π-orbitals in the aromatic systems as the number of aromatic rings increases. Similar trends were also observed for the proton at the peri-position. It should be noted that the ethane bridge protons of 7a resonated at δ 2.03 as a singlet, while those of 7b and 7c were multiplets in the range of δ 2.25–2.10 and δ 2.27–2.23, respectively, reflecting the fact these bridges are rigid at room temperature on the NMR time scale.
UV absorption spectra of 7 were obtained in EtOH (1.0 × 10−6 M) (Figure 2) and the absorption maxima and extinction coefficients are summarized in Table 2. The absorption bands in the 245–303 and 331–400 nm regions correspond quite closely to the π-π* absorptions. The absorption maxima for the more linear compounds (7a vs. 7b, 7b vs. 7c) appear shifted towards longer wavelengths (see Figure 2). These data are consistent with an electronic transition state in which the energy of the receptor π* orbital is lowered by the increasing delocalization which would be found for the more conjugated and linear systems. The electronic spectrum of the most linear compound 7c showed π-π* absorption bands at 303 and 325 nm which are red-shifted compared to the parent compound 7a by 58 and 80 nm, with little effect on the intensity of the absorption compared to 7a. It should be noted that benzoannulation led an additional absorption band in the region of 350–408 nm, of which 7c showed the most intense and bathochromic shifted absorption due to the additional benzene ring.
The photoluminescence (PL) of the compounds was studied in EtOH (1 × 10−5 mol/L) at room temperature (Figure 2) and are summarized in Table 2. Excitation of the absorbances at 245, 280, and 303 nm for 7a, 7b, and 7c, respectively, showed emissions at 400, 465, and 500 nm. The observed emission wavelength is somewhat dependent on the nature of the conjugated system: The most extended compound 7c showed a green light emission at 500 nm while 7a showed a purple light emission at 400 nm. The quantum yields were determined by a previously reported method [34,35] employing quinine sulfate as a standard to give values of 0.38, 0.54, and 0.58 for 7a, 7b, and 7c, respectively.

2.4. Structural and Thermal Properties

The crystallinity of the compounds prepared was analyzed by X-ray diffraction (XRD) and the corresponding X-ray diffractograms are shown in Figure 3. The diffractograms of compounds 7a and 7b showed distinctive peaks, indicating their crystalline nature. Compounds 7a, of which the X-ray crystal structure has been reported [36], and 7b have more crystalline character comparing to 7c. The crystallite sizes of 7a and 7b were calculated by employing Scherrer’s equation, Dp = 0.93λ/Lcosθ [37], where Dp is the averaged particle size of the crystallites, λ is the incident wavelength (1.54056 Å), θ is the Bragg angle and L is the diffracted full width at half maximum (in radians) caused by crystallization to give 6.89 and 10.54 nm for 7a and 7b, respectively.
The thermal behaviors of the compounds were analyzed by differential scanning calorimetry (DSC). All the compounds showed a single sharp endothermic peak at the melting transition temperature (Tm) of 287.3, 397.35, and 409.78 °C, respectively. However only compound 7a showed sharp exothermic peaks as a crystallization temperature (Tc) at 202.74 °C (data not shown). It should be noted that none of the compounds showed a glass transition temperature (Tg).

3. Experimental Section

3.1. General Information

Melting points were determined using a Fischer-Jones melting points apparatus (Fischer Scientific, Grand Island, NY, USA) and are not corrected. UV spectra were recorded on a V550 spectrophotometer (JASCO, Oklahoma City, OK, USA). NMR spectra were obtained using a Bruker-250 spectrometer (Fällanden, Switzerland) or VNS600 FT-NMR (Varian, Australia) operating at 250 MHz or 600 MHz for 1H-NMR and 62.5 MHz or 150 MHz for 13C-NMR and are reported as parts per million (ppm) from the internal standard TMS. Chemicals and solvents were commercial reagent grade and used without further purification. Electrospray ionization (ESI) mass spectrometry (MS) experiments were performed on a LCQ advantage-trap mass spectrometer (Thermo Finnigan, San Jose, CA, USA). Elemental analyses were taken on a Hewlett-Packard Model 185B elemental analyzer (Hewlett Packard, Littleton, MA, USA). XRD analysis was performed by X-ray diffractometry (MPD for bulk, PANalytical, Wesybrough, MA, USA) with nickel-filtered CuKα radiation (30 kV, 30 mA) at 2θ angles from 10° to 90°, a scan speed of 10°/min and a time constant of 1 s. Thermal behaviors of the compounds were analyzed using differential scanning calorimeter (DSC Q200, TA Instrument, Wilminton, NJ, USA) with 1~2 mg of sample sealed in alumina in the range of 40–385 °C increasing temperature in a rate of 10 °C/min. An empty pan was used as a reference, and the DSC baseline, temperature, and enthalpy were calibrated. High and low resolution FAB-MS data were obtained on a JMS-700 instrument (JEOL, Peabody, MA, USA).

3.2. Synthesis and Characterization of the Products

6,13-Dihydro-6,13-ethano-5,12-diazapentacene (7a)

To a mixture of 2-aminobenzaldehyde (5a, 0.59 g, 4.9 mmol) and bicyclo[2.2.2]octane-2,5-dione (6, 0.30 g, 2.18 mmol) in EtOH (15 mL) was added 2 M KOH in EtOH (1 mL). The resulting mixture was refluxed for 8 h to form a precipitate which was collected and washed with EtOH to give pale yellow needles (0.60 g, 88%): mp 297–298 °C (lit. [36] mp 284–285 °C). Spectral data were identical to those reported previously [36]. 1H-NMR (CDCl3, 250 MHz) δ 8.33 (s, 2H, H7 and H14), 8.01 (d, J = 8.3 Hz, 2H, H4 and H11), 7.95 (d, J = 7.9 Hz, 2H, H1 and H8), 7.73 (td, J = 8.3, 1.0 Hz, H3 and H10), 7.60 (td, J = 8.0, 1.0 Hz, H2 and H9), 4.80 (s, 2H, H6 and H13), 2.03 (s, 4H). 13C-NMR (62.5 MHz, CDCl3) δ 25.5, 45.8, 126.3, 127.3, 128.0, 128.6, 129.5, 130.3, 134.2, 146.3, 163.4.

7,16-Dihydro-7,16-ethanobenzo[b]benzo[6,7]quinolino[3,2-i]acridine (7,16-dihydro-7,16-ethano-6,15-diaza-heptacene, 7b)

To a mixture of 3-aminonaphthalene-2-carbaldehyde (5b, 280 mg, 1.63 mmol) and bicyclo[2.2.2]octane-2,5-dione (6, 90 mg, 0.65 mmol) in EtOH (10 mL) was added 2 M KOH in EtOH (1 mL). The resulting reaction mixture was refluxed for 8 h. Evaporation of the solvent provided pale yellow crystalline solids (236 mg, 89%), which was purified by flash column chromatography on silica eluting with EtOAc to give the desired product as pale yellow needles: mp 397.35 °C (DSC). 1H-NMR (CDCl3, 250 MHz) δ 8.60 (s, 2H, H5 and H14), 8.33 (s, 2H, H8 and H17), 8.19 (s, 2H, H9 and H18), 8.03 (overlapped dd, 4H, J = 7.8, 2.8 Hz, H1, H4, H10, H13), 7.50–7.46 (m, 4H, H2, H3, H11, H12), 4.77 (s, 2H, H7 and H16), 2.25–2.10 (m, 4H). 13C-NMR (CDCl3, 62.5 MHz) δ 163.9, 143.6, 133.8, 133.1, 131.8, 130.6, 128.6, 128.2, 126.7, 126.6, 126.4, 126.3, 126.1, 46.7, 25.5. MS (ESI) calcd for C30H20N2 [M + H]+ 409, found 409. HR-FAB-MS (m/z): [M+] calcd for C30H20N2, 408.16265; found, 408.1629. Anal. calcd for C30H20N2 C, 88.21; H, 4.94; N, 6.86. Found C, 88.39; H, 4.93.

8,19-Dihydro-8,19-ethanonaphtho[2,3-b]naphtho[2′,3′:6,7]quinolino[3,2-i]acridine (8,19-dihydro-8,19-ethano-7,18-diazanonacene, 7c)

A mixture of 3-aminoanthracene-2-carbaldehyde (5c, 360 mg, 1.63 mmol, 2.5 equiv), bicyclo[2.2.2]octane-2,5-dione (6, 90 mg, 0.65 mmol), and 2 M KOH in EtOH (1 mL) in EtOH (10 mL) was refluxed for 8 h. Work-up as described above for 7a afforded a pale crystalline solid (291 mg, 88%) as a desired product which was flash-chromatographed on silica eluting with EtOAc. The later fractions afforded yellow solid: mp 409.78 °C (DSC) (dec). 1H-NMR (CDCl3, 250 MHz) δ 8.85 (s, 2H, H6 and H17), 8.73 (s, 2H, H9 and H20), 8.68 (s, 2H, H10 and H21), 8.59 (s, 2H, H5 and H16), 8.21 (s, 2H, H11 and H22), 8.03 (dd, 2H, J = 7.8, 1.5 Hz, H4, H14), 8.00 (dd, 2H, J = 7.8, 1.5 Hz, H1, H12), 7.45–7.41 (m, 4H, H2, H3, H13, H14), 4.76 (s, 2H, H8 and H18), 2.27–2.23 (m, 4H). A 13C-NMR spectrum could not be recorded due to the low solubility in common NMR solvents. MS (ESI) calcd for C38H25N2 [M + H]+ 509, found 509. HR-FAB-MS (m/z): [M+] calcd for C38H24N2, 508.19395; found, 508.1641. Anal. calcd for C38H24N2-1.5H2O C, 85.21; H, 5.08; N, 5.23. Found C, 85.26; H, 5.07; N, 5.21.

3.3. Attempted Synthesis of Diazapolycenes 8

Method A: A mixture of 7 (1.0 mmol) and 10% Pd/C (100 mg) in nitrobenzene (10 mL) was heated 200 °C for 8 h. After cooling the reaction mixture to room temperature, reaction mixture was filtered through Celite®. Work up as usual afforded a pale yellow solid, which turned out the starting material based on TLC as well as 1H-NMR.
Method B: The vacuum pyrolysis (0.02 mmHg, 600 °C, 45–60 min) of the compound 7 (1.0 mmol) did not lead to the corresponding diazapolycenes 8 but instead sublimation of the product occurred for 7a,b, while 7c was decomposed at 409 °C without sublimation.

3.4. Measurement of Quantum Yield (Φ)

The quantum yield of the compounds prepared was measured by using a quinine sulfate solution (in 0.1 M H2SO4, literature quantum yield 0.577 at 360 nm [34,35]) as the standard and calculated with the equation of Φ = ΦR × (I/IR) × (ODR/OD) × (n2/nR2), where Φ is the quantum yield, I is the measured integrated emission intensity, n is the refractive index, OD is the optical density and the subscript R refers to the quinine sulfate.

4. Conclusions

The Friedländer reactions of three o-aminoarenecarbaldehydes, namely 2-aminobenzaldehyde, 3-aminonaphthalene-2-carbaldehyde and 3-aminoanthracene-2-carbaldehyde, with bicyclo[2.2.2]-octane-2,5-dione yielded a series of ethanodiazapolycenes as dimeric azapolycenes in 87%–89% yields and their spectral, thermal, and structural properties were studied. All attempts to further convert the products into diazapolycenes by retro-Diels-Alder reaction or vacuum pyrolysis were not successful.

Acknowledgments

Financial support from Yeungnam University (213A380193) is gratefully appreciated.

Author Contributions

MK performed synthesis, characterization, gathering spectral data of the compounds and participated in manuscript writing. YJ planned the experiments, supervised, and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brown, A.R.; Pomp, A.; Hart, C.M.; de Leeuw, D.M. Logic gates made from polymer transistors and their use in ring oscillators. Science 1995, 270, 972–974. [Google Scholar] [CrossRef]
  2. Dimitrakopoulos, C.D.; Malenfant, P.R.L. Organic thin film transistors for large area electronics. Adv. Mater. 2002, 14, 99–117. [Google Scholar] [CrossRef]
  3. Kleemann, H.; Gutierrez, R.; Avdoshenko, S.; Cuniberti, G.; Leo, K.; Luessem, B. Reverse breakdown behavior in organic pin-diodes comprising C60 and pentacene: Experiment and theory. Org. Electron. 2013, 14, 193–199. [Google Scholar] [CrossRef]
  4. Wolak, M.A.; Jang, B.; Palilis, L.C.; Kafafi, Z.H. Functionalized pentacene derivatives for use as red emitters in organic light-emitting diodes. J. Phys. Chem. B 2004, 108, 5492–5499. [Google Scholar] [CrossRef]
  5. Etienne, A.; Beauvois, C. Photooxidation of 6,13-diphenylpentacene and 5,7,12,14-tetraphenylpentacene. Compt. Rend. 1954, 239, 64–66. [Google Scholar]
  6. Afzali, A.; Kagan, C.R.; Traub, G.P. N-Sulfinylcarbamate-pentacene adduct: A novel pentacene precursor soluble in alcohols. Synth. Met. 2005, 155, 490–494. [Google Scholar] [CrossRef]
  7. Allen, C.F.H.; Bell, A. Action of Grignard reagents on certain pentacenequinones, 6,13-diphenylpentacene. J. Am. Chem. Soc. 1942, 64, 1253–1260. [Google Scholar] [CrossRef]
  8. Kaur, I.; Jia, W.J.; Kopreski, R.P.; Selvaprasah, S.; Dokmeci, M.R.; Pramanik, C.; McGruer, N.E.; Miller, G.P. Substituent effects in pentacenes: Gaining control over HOMO–LUMO gaps and photooxidative resistances. J. Am. Chem. Soc. 2008, 130, 16274–16286. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, J.; Pawle, R.H.; Haas, T.E.; Thomas, S.W., III. Combining electronic and steric effects for highly stable unsymmetric pentacenes. Chem. Eur. J. 2014, 20, 5880–5884. [Google Scholar] [CrossRef] [PubMed]
  10. Kraft, U.; Anthony, J.E.; Ripaud, E.; Loth, M.A.; Weber, E.; Klauk, H. Low-voltage organic transistors based on tetraceno[2,3-b]thiophene: Contact resistance and air stability. Chem. Mater. 2015, 27, 998–1004. [Google Scholar] [CrossRef]
  11. Song, C.-L.; Ma, C.-B.; Yang, F.; Zeng, W.-J.; Zhang, H.-L.; Gong, X. Synthesis of tetrachloro-azapentacene as an ambipolar organic semiconductor with high and balanced carrier mobilities. Org. Lett. 2011, 13, 2880–2883. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, Y.-Y.; Song, C.-L.; Zeng, W.-J.; Zhou, K.-G.; Shi, Z.-F.; Ma, C.-B.; Yang, F.; Zhang, H.-L.; Gong, X. High and balanced hole and electron mobilities from ambipolar thin-film transistors based on nitrogen-containing oligoacences. J. Am. Chem. Soc. 2010, 132, 16349–16351. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, Y.; Shen, L.; Li, X. Effects of heteroatoms of tetracene and pentacene derivatives on their stability and singlet fission. J. Phys. Chem. A 2014, 118, 5700–5708. [Google Scholar] [CrossRef] [PubMed]
  14. Bunz, U.H.F. The larger linear N-heterocycles. Acc. Chem. Res. 2015, 48, 1676–1687. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, J.; Wang, C.; Long, G.; Aratani, N.; Yamadab, H.; Zhang, Q. Fusing N-heterocene analogues into one “kinked” molecule with slipped two-dimensional ladder-like packing. Chem. Sci. 2016, 7, 1309–1313. [Google Scholar] [CrossRef]
  16. Anthony, J.E.; Brooks, J.S.; Eaton, D.L.; Parkin, S.R. Functionalized pentacene: Improved electronic properties from control of solid-state order. J. Am. Chem. Soc. 2001, 123, 9482–9483. [Google Scholar] [CrossRef] [PubMed]
  17. Jahng, Y.; Karim, M. Synthesis of 5-azapentacene by Friedländer reaction and its properties. Tetrahedron 2016, 72, 199–204. [Google Scholar] [CrossRef]
  18. Karim, M.; Jahng, Y. Unusual product distribution from Friedländer reaction of di and triacetylbenzenes with 3-aminonaphthalene-2-carbaldehyde and properties of new benzo[g]quinoline-derived aza-aromatics. Molecules 2014, 19, 12842–12851. [Google Scholar] [CrossRef] [PubMed]
  19. Jahng, Y.; Rahman, A.F.M.M. Synthesis and properties of 2,2′-di(heteroaryl)-9,9′-spirobifluorenes. Bull. Chem. Soc. Jpn. 2010, 83, 672–677. [Google Scholar] [CrossRef]
  20. Smith, L.I.; Opie, J.W. o-Aminobenzaldehyde. Org. Synth. Coll. 1955, 3, 56–58. [Google Scholar]
  21. Taffarel, E.; Chirayil, S.; Thummel, R.P. Synthesis and properties of ligands based on benzo[g]quinoline. J. Org. Chem. 1994, 59, 823–828. [Google Scholar] [CrossRef]
  22. Funel, J.-A.; Schmidt, G.; Abele, S. Design and scale-up of Diels-Alder reactions for the practical synthesis of 5-phenylbicyclo[2.2.2]oct-5-en-2-one. Org. Proc. Res. Dev. 2011, 15, 1420–1427. [Google Scholar] [CrossRef]
  23. House, H.O. Modern Synthetic Reactions, 2nd ed.; W.A. Benjamin, Inc.: Readings, MA, USA, 1972; pp. 34–44. [Google Scholar]
  24. Kutz, W.M.; Corson, B.B. Dealkylation of dialkylbenzenes: Use of superatmospheric pressure and alumina-silica catalyst in presence of benzene. Ind. Eng. Chem. 1946, 38, 761–764. [Google Scholar] [CrossRef]
  25. Linstead, R.P.; Thomas, S.L.S. Dehydrogenation. II. Elimination and migration of methyl groups from quaternary carbon atoms during catalytic dehydrogenation. J. Chem. Soc. 1940, 1127–1134. [Google Scholar] [CrossRef]
  26. Csicsery, S.M.; Pines, H. Alumina: Catalyst and support. XVIII. Aromatization of 2,2-dimethyl-3-methyl-C14-pentane over “non-acidic” chromia-alumina catalyst. Contribution to the mechanism of aromatization. J. Am. Chem. Soc. 1962, 84, 3939–3944. [Google Scholar] [CrossRef]
  27. McNab, H. Chemistry without reagents: Synthetic applications of flash vacuum pyrolysis. Aldrichim. Acta 2004, 37, 19–26. [Google Scholar] [CrossRef]
  28. Rickborn, B. The retro-Diels-Alder reaction, Part I: C-C dienophiles. Org. React. 1998, 52, 1–393. [Google Scholar]
  29. Trahanovsky, W.S.; Swenson, K.E. Flash vacuum pyrolysis of 2,3-dialkyltetralins. J. Org. Chem. 1981, 46, 2984–2985. [Google Scholar] [CrossRef]
  30. Alder, K.; Rickert, H.F. Zur Kenntnis der Dien-synthese. I. Über eine Methode der direkten Unterscheidung cyclischer Penta- und Hexa-diene. Ann. Chem. Pharm. 1936, 524, 180–189. [Google Scholar] [CrossRef]
  31. Cadogan, J.I.G.; Hickson, C.L.; McNab, H. Short contact time reactions of large organic free radicals. Tetrahedron 1986, 42, 2135–2165. [Google Scholar] [CrossRef]
  32. Pretsch, E.; Clerc, T.; Siebel, J.; Simon, W. Tables of Spectral Data for Structure Determination of Organic Compounds 13C-NMR 1H-NMR IR MS UV/VIS, 2nd ed.; Springer-Verlag: Berlin, Geramny, 1999; p. H335. [Google Scholar]
  33. Krapcho, A.P.; Gilmor, T.P. General preparative route to benzo[g]quinolines (1-Azaanthracenes). J. Heterocycl. Chem. 1999, 36, 445–452. [Google Scholar] [CrossRef]
  34. Lakowicz, J.P. Principles of Fluorescence Spectroscopy, 3rd ed.; Springer-Verlag: Berlin, Geramny, 2006; pp. 54–55. [Google Scholar]
  35. Eastman, J.W. Quantitative spectrofluorimetry: The fluorescence quantum yield of quinine sulfate. Photochem. Photobiol. 1967, 6, 55–72. [Google Scholar] [CrossRef]
  36. Alshahateet, S.F.; Rahman, A.; Noman, M.M.; Bishop, R.; Craig, D.C.; Scudder, M.L. Interlocking molecular grid lattices involving weak assembly forces. CrystEngComm 2002, 4, 585–590. [Google Scholar] [CrossRef]
  37. Patterson, A. The Scherrer formula for X-ray particle size determination. Phys. Rev. 1939, 10, 978–982. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 7 are available from the authors.
Figure 1. Pentacene and its isosteres.
Figure 1. Pentacene and its isosteres.
Molecules 21 00407 g001
Scheme 1. Synthesis of ethano-bridged diazapolycenes.
Scheme 1. Synthesis of ethano-bridged diazapolycenes.
Molecules 21 00407 g004
Scheme 2. Synthesis of diazapolycenes.
Scheme 2. Synthesis of diazapolycenes.
Molecules 21 00407 g005
Figure 2. UV absorption and photoluminescence (PL) spectra of 7a, 7b, and 7c in EtOH (1 × 10−6 M) at 298 K.
Figure 2. UV absorption and photoluminescence (PL) spectra of 7a, 7b, and 7c in EtOH (1 × 10−6 M) at 298 K.
Molecules 21 00407 g002
Figure 3. X-ray diffractograms of compounds 7a, 7b, and 7c in powder state.
Figure 3. X-ray diffractograms of compounds 7a, 7b, and 7c in powder state.
Molecules 21 00407 g003
Table 1. 1H-NMR chemical shifts and coupling constants of selected protons.
Table 1. 1H-NMR chemical shifts and coupling constants of selected protons.
CompoundsH4H at Peri-Position-CH2CH2-
Quinoline (a)8.00 (3J = 8.3, 4J = 1.1 Hz)8.05 (3J = 8.2, 4J = 1.6 Hz)
Benzo[g]quinoline8.58 (3J = 8.5, 4J = 1 Hz) (b)
8.31 (3J = 8.5 Hz) (c)
8.76 (s) (b) 8.70 (s) (c)
Naphtho[2,3-g]quinoline (d)8.50 (3J = 7.5, 4J = 1.5 Hz)8.87 (s)
5-Azapentacene (e)9.12 (s)9.26 (s)
7a7.60 (s)8.01 (3J = 8.3, 4J = 1.2 Hz)2.03 (s)
7b8.33 (s)8.59 (s)2.25–2.10 (m)
7c8.73 (s)8.85 (s)2.27–2.23 (m)
(a) Ref. [32]; (b) Taken from 250 MHz 1H-NMR in CDCl3; (c) Ref [33]; (d) Taken from the ChemDraw® computer program from CambridgeSoft; (e) Ref [17].
Table 2. UV absorption and FL data of compounds 7a, 7b, and 7c.
Table 2. UV absorption and FL data of compounds 7a, 7b, and 7c.
Compd.λ [nm (logε)]λexcitλemitQuantum Yield (a)
7a245 (5.49) 331 (5.18)2454000.38
7b256 (5.20, sh) 280 (5.46) 297 (5.20, sh) 350 (4.61) 370 (4.81)2804650.54
7c257 (5.39) 303 (5.46) 325 (5.41) 408 (4.85)3035000.58
(a) Values were calculated from the equation given in Ref. [34,35] using quinine sulfate as a reference.

Share and Cite

MDPI and ACS Style

Karim, M.; Jahng, Y. Synthesis of Ethano-Bridged Diazapolycenes as Potential Precursors for Diazapolycenes and Their Properties. Molecules 2016, 21, 407. https://doi.org/10.3390/molecules21040407

AMA Style

Karim M, Jahng Y. Synthesis of Ethano-Bridged Diazapolycenes as Potential Precursors for Diazapolycenes and Their Properties. Molecules. 2016; 21(4):407. https://doi.org/10.3390/molecules21040407

Chicago/Turabian Style

Karim, Moinul, and Yurngdong Jahng. 2016. "Synthesis of Ethano-Bridged Diazapolycenes as Potential Precursors for Diazapolycenes and Their Properties" Molecules 21, no. 4: 407. https://doi.org/10.3390/molecules21040407

Article Metrics

Back to TopTop