Next Article in Journal
Reduction of Diphenyl Diselenide and Analogs by Mammalian Thioredoxin Reductase Is Independent of Their Gluthathione Peroxidase-Like Activity: A Possible Novel Pathway for Their Antioxidant Activity
Previous Article in Journal
Free Radical Scavenging Activity and Characterization of Sesquiterpenoids in Four Species of Curcuma Using a TLC Bioautography Assay and GC-MS Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis of a Highly Luminescent Three-Dimensional Pyrene Dye Based on the Spirobifluorene Skeleton

Department of Organic and Polymeric Materials, Graduate School of Science and Engineering, Tokyo Institute of Technology, Japan
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(11), 7582-7592; https://doi.org/10.3390/molecules15117582
Submission received: 24 August 2010 / Revised: 25 October 2010 / Accepted: 27 October 2010 / Published: 28 October 2010

Abstract

:
We have synthesized a highly luminescent (log ε > 5.0, Φ > 0.9) pyrene dye based on a spirobifluorene skeleton [2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi[9H-fluorene; 4-PySBF]. The use of spirobifluorene prevents fluorescence quenching by intramolecular energy transfer and/or electron transfer among the chromophores in the excited state. The emission spectra of 4-PySBF exhibited a red shift of 20 nm in comparison to a model compound [9,9′-dioctyl-2,7-bis(7-tert-butyl-1-pyrenyl)-9H-fluorene; 2-PyF], but its UV-Vis spectrum remained unchanged.

Graphical Abstract

Introduction

Recently, blue emitting dyes with high efficiency (Φ ≈ 1.0, log ε > 4.5) [1] have been attracting considerable attention because of their applicability to molecular electronic materials such as organic field effect transistors (OFETs) and organic light-emitting diodes (OLEDs). Many studies have been conducted on strategies for synthesizing a highly efficient dye. For example, rod- [1,2] and star-shaped [3,4,5,6] hydrocarbons that expand π-conjugation to two dimensions can be used to control the emission color of a dye with high efficiency.
Spirobifluorene derivatives are considered the most promising candidates for organic optoelectronics [7,8,9,10,11]. The rigidity of spiro-compounds affords them high thermal stability. In addition, their solubility is higher than that of corresponding compounds without a spiro moiety, because their perpendicular conformations that are based on the spiro-linkage efficiently suppress intermolecular interactions between π-systems. Recently, several studies were conducted on introducing dyes into a molecule for π-conjugation in three dimensions. Examples of such studies include those on the introduction of dyes into the spirobifluorene skeleton [12,13,14,15,16,17,18,19]. Because of the high thermal stability and unique characteristics of the spirobifluorene skeleton, it has been adopted in various materials [20,21,22].
In this paper, we report the synthesis and characterization of a new pyrene dye that is based on the spirobifluorene skeleton. Pyrene derivatives are highly absorptive [23,24,25,26]; further, the introduction of pyrene substituents into spirobifluorene derivatives is expected to improve the fluorescence quantum yield and thermal stability of the derivatives because the substituents are highly emissive, bulky, and rigid.

Results and Discussion

A spirobifluorene dye, 2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (4-PySBF), and a model compound, 9,9′-dioctyl-2,7-bis(7-tert-butyl-1-pyrenyl)-9H-fluorene (2-PyF), were synthesized in 37% and 24% yields, respectively, by palladium-catalyzed Suzuki-Miyaura coupling reactions (Scheme 1). The precursors, 7-tert-butylpyrene-1-boronic acid pinacol ester (1) and 1-bromo-7-tert-butylpyrene (3), were prepared from pyrene according to a previously reported method [27]. The chemical structures of 4-PySBF and 2-PyF were confirmed by 1H-NMR, 13C-NMR, and FT-IR spectroscopy. Each compound was found to be soluble in various organic solvents such as chloroform, tetrahydrofuran (THF), dichloromethane, and toluene. It should be noted that the tert-butyl group is necessary for the successful synthesis of 4-PySBF; we found that without tert-butyl groups, the solubility is quite poor, making purification and structural analysis by techniques such as 1H-NMR and 13C-NMR rather difficult.
Scheme 1. Synthesis of spirobifluorene dye and model compound.
Scheme 1. Synthesis of spirobifluorene dye and model compound.
Molecules 15 07582 g007
We measured the UV-Vis spectra, fluorescence spectra, fluorescence quantum yields, and fluorescence lifetimes, and lifetimes of 4-PySBF and 2-PyF in CH2Cl2 solution. The UV-Vis spectra (Figure 2, left side) show that the absorption maxima (λmax) of 4-PySBF and 2-PyF are at 360 nm and 359 nm, with molar absorption (λmax) coefficients ε of 131,000 and 71,000 M−1 cm−1, respectively. The finding that the ε value of 4-PySBF is about twice that of 2-PyF is in accordance with the fact that 4-PySBF has two chromophores per molecule whereas 2-PyF has but one.
The observed fluorescence spectra (Figure 2, right side) show that 4-PySBF and 2-PyF exhibit intense blue emissions (λem) at 441 nm (Φ = 0.92) and 421 nm (Φ = 0.90), respectively, in CH2Cl2 solution (λex = 360 nm; c = 1.0 × 10−6 M). The fluorescence spectra also reveal that the fluorescence intensity of 4-PySBF is about twice that of 2-PyF, again because 4-PySBF has two chromophores per molecule. A 20-nm red-shift between 4-PySBF and 2-PyF was observed. It is noted that we observed a spiroconjugation (a large red shift in the fluorescence spectrum and no red shift in the absorption spectrum in this 4-PySBF and 2-PyF pair). We hypothesize that this observed red-shift could be due to conjugation of 4-PySBF via the spiro carbon [28,29,30,31,32] in the excited state [33]. Another possibility is a through-space interaction between upper and lower pyrene chromophores.
Figure 1. Absorption and emission spectra of 2-PyF and 4-PySBF in CH2Cl2; c = 10 μM, λex = 360 nm.
Figure 1. Absorption and emission spectra of 2-PyF and 4-PySBF in CH2Cl2; c = 10 μM, λex = 360 nm.
Molecules 15 07582 g001
Table 1. Photochemical data for compounds 4-PySBF and 6.
Table 1. Photochemical data for compounds 4-PySBF and 6.
dyeλab[nm]ε (×105) [l/mol・cm]log ελem[nm]Φτ [ns]kf (×108) [S-1]kisc + knr (×107) [S-1]
4-PySBF3601.315.124410.921.526.15.3
2-PyF3590.714.854210.901.884.85.3
Measurement of the photophysical properties of both compounds at the same concentration confirm that 4-PySBF had twice the number of chromophores in solution as compared to 2-PyF. Next, we measured the fluorescence spectra of both compounds under the same absorption conditions. We then found that the solutions of 4-PySBF and 2-PyF had the same number of chromophores. The fluorescence spectra (Figure 3) revealed that both 4-PySBF and 2-PyF had approximately the same fluorescence intensity. This result shows that the spirobifluorene derivatization doubled the performance of the chromophores by preventing intramolecular energy transfer and/or electron transfer, although the pyrene chromophores are placed close together. The compound 4-PySBF was found to have ideal luminescence properties (log ε > 5.0, Φ > 0.9).
Figure 2. Emission spectra of 2-PyF and 4-PySBF in CH2Cl2; Abs = 0.1, λex = 360 nm.
Figure 2. Emission spectra of 2-PyF and 4-PySBF in CH2Cl2; Abs = 0.1, λex = 360 nm.
Molecules 15 07582 g002
Fluorescence quantum yield can be expressed as Φf = kfτs, or Φf = kf/(kf + kisc + knr + kq[O2]), where kf, kisc, knr, and kq[O2] denote the rate constants for fluorescence radiation, intersystem crossing, nonradiative decay, and fluorescence quenching by oxygen, respectively, and τs denotes the lifetime (s = singlet) [14]. Table 1 summarizes the values of kf and kisc + knr for 4-PySBF and 2-PyF, calculated from the experimental values of Φf and τs under degassed conditions, that is, under the assumption that kq[O2] = 0. As knr for aromatic hydrocarbons is known to be negligible [34], kisc should be the dominant component in kisc+knr [35,36,37,38]. Under aerated conditions, kq[O2] is estimated to be nearly diffusion controlled (up to 3.89 × 1010 m−1 s−1 in hexane) [34,39], and [O2] is less than 2.0 × 10−3 M in usual solvents. These data imply that the maximum possible value of kq[O2] is 8.0 × 107 s−1. As can be seen from Table 1, the kf values for 4-PySBF and 2-PyF are much larger than the kq[O2] values. This situation is most probably responsible for the observed large values of Φf for both compounds, even in the presence of oxygen.

Conclusions

In summary, we have synthesized a highly luminescent spirobifluorene dye (log ε > 5.0, Φ > 0.9) with tetrapyrene chromophores (4-PySBF) and measured its photophysical properties. In comparison to a model compound, 2-PyF, the emission spectrum of 4-PySBF exhibited a red shift of 20 nm, but its UV-Vis spectrum remained unchanged. It is not easy to find a satisfactory explanation for this observed red-shift at the present stage. However, in our next study, we intend to calculate the electron states of these compounds in the ground and excited states by the molecular orbital (MO) method and examine this conjugation system in greater detail. Further, we intend to apply 4-PySBF in a previously developed cholesteric liquid crystal laser (distributed feedback laser) as the laser dye in the future.

Experimental

Instruments

All the 1H- and 13C-NMR spectra were recorded on a 400 MHz JEOL LMN-EX400 instrument with tetramethylsilane (TMS) as the internal standard. FT-IR spectra were recorded on a JASCO FT-IR 469 plus spectrometer. Melting points were obtained by a Stuart Scientific Melting Point Apparatus SMP3. MS spectra (FAB) were obtained by JEOL JMS700 mass spectrometer. UV-Vis spectra were recorded with a Beckman Coulter DU800 UV-Vis Spectrophotometer. Fluorescence spectra were recorded on a JASCO FP-6500 Spectrofluorometer. Quantum Yields were measured by a Hamamatsu Photonics C9920-02 Absolute PL Quantum Yield Measurement system. Fluorescence lifetimes were measured using a Hamamatsu Photonics OB 920 Fluorescence Lifetime Spectrometer.

Materials

Unless otherwise noted, all reagents, chemicals and solvents were obtained from commercial sources and used without further purification. Pyrene, 2-chloro-2-methylpropane, Pd(PPh3)4, 4,4,5,5-tetramethyl-1,3,2-dioxaborolane were obtained from TCI. 9,9-Dioctyl-fluorene-2,7-bis(trimethylene-borate) (4) and PdCl2(PPh3)2, were obtained from Aldrich. 2,2′,7,7′-Tetrabromospirobifluorene (2) was a gift of JFE Chemical Coporation (Japan).
2,2′,7,7′-tetrakis(7-tert-Butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (3, 4-PySBF)
Under an argon atomosphere, 2,2′,7,7′-tetrabromospirobifluorene (0.25 g, 0.4 mmol), 7-tert-butylpyrene-1-boronic acid pinacol ester (0.77 g, 2.0 mmol) and cesium carbonate (5.21 g, 16 mmol) were mixed together with Pd(PPh3)4 (100 mg, 0.1 mmol) and degassed THF (30 mL). The mixture was refluxed for 24 h. After cooling to room temperature, the resulting mixture was extracted with chloroform. The organic extract was washed sequentially with water and brine and then dried over MgSO4. After removal of the solvent, the residue was purified by column chromatography (chloroform:hexane = 1:5) and preparative HPLC (CHCl3). After removal of the solvent, the product was recrystallized from cyclohexane to afford 4-PySBF as a yellow powder in 37% yield. 1H-NMR (400 MHz, THF-d8): δ 8.23–7.36 (m, Ar-H, 44H), 1.53 (s, tert-butyl-H, 36H) ppm.; 13C-NMR (100 MHz, CDCl3): δ 149.3, 149.1, 140.9, 140.7, 137.4, 131.3, 130.7, 130.5, 130.3, 128.4, 128.2, 128.1, 127.7, 127.5, 127.3, 127.2, 126.5, 125.0, 124.9, 124.4, 123.1, 122.4, 122.0, 120.2 (aromatic C), 66.3 (spiro-C), 35.2 (Ar-C (CH3)3), 31.9 (Ar-C (CH3)3) ppm; FT-IR (KBr): 3042, 2961, 2901, 2867, 1594, 1457, 1227 cm−1; mp 286–288 °C; HR-MS (FAB+) Calcd. for C105H80 [M] 1341.6294, found [M+] 1341.6290.
9,9′-Dioctyl-2,7-bis (7-tert-butyl-1-pyrenyl)-9H-fluorene (6, 2-PyF)
Under an argon atomosphere, 2-bromo-7-tert-butylpyrene (0.8 g, 2.4 mmol) and 9,9-dioctyl-fluorene-2,7-bis(trimethyleneborate) (0.28 g 0.5 mmol) were mixed together with Pd(PPh3)4 (60 mg, 0.05 mmol), degassed toluene (20 mL), one drop of Aliquat 336, and 1M aqueous sodium carbonate solution (2 mL). The resulting mixture was refluxed for 24 h. After cooling to room temperature, the mixture was extracted with chloroform. The organic extract was washed sequentially with water and brine and then dried over MgSO4. After removal of the solvent, the residue was purified by column chromatography using chloroform/hexane (5:1, v/v) following HPLC (CHCl3). After subsequent recrystallization in ethanol, 2-PyF was obtained as a yellow powder in 24% yield. 1H-NMR (400 MHz, CDCl3): δ 8.27–7.97 (m, Ar-H, 18H), 7.69–7.67 (m, Ar-H, 4H), 2.12–2.08 (s, Ar-CH-Ar, 2H), 1.60 (s, tert-butyl-H, 18H), 1.25–1.19 (m, alkyl-H, 20H), 0.96 (m, alkyl-H, 4H), 0.82 (t, J = 6.84, alkyl-H, 6H) ppm; 13C-NMR (100 MHz, CDCl3): δ 151.2, 149.1, 140.1, 140.0, 138.1, 131.4, 130.9, 130.4, 129.4, 128.4, 127.6, 127.5, 127.3, 125.4, 125.3, 125.0, 124.5, 123.2, 122.4, 122.0, 119.8 (aromatic C), 55.3 (spiro-C), 40.4 (alkyl-C), 35.2 (Ar-C (CH3)3), 31.9 (Ar-C (CH3)3), 30.1, 29.7, 29.4, 29.3, 24.2, 22.7, 14.2 (alkyl-C) ppm; FT-IR (KBr): 3050, 2953, 2925, 2853, 1608, 1481, 1457, 1310 cm−1; mp 206–208 °C; HR-MS (FAB+) calcd. for C69H74 [M] 902.5791, found [M+] 902.5786.

Acknowledgements

We thank JFE Chemical for the kind gift of some fluorene compounds. This work was supported by the Industrial Technology Research and Development Grant (09C46622) from NEDO of Japan.

Appendix

1H-NMR and 13C NMR spectra of 3 and 6
Figure S1. 1H-NMR spectrum of 2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (4-PySBF, 400 MHz, THF-d8).
Figure S1. 1H-NMR spectrum of 2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (4-PySBF, 400 MHz, THF-d8).
Molecules 15 07582 g003
Figure S2. 13C-NMR spectrum of 2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (4-PySBF, 100 MHz, CDCl3).
Figure S2. 13C-NMR spectrum of 2,2′,7,7′-tetrakis(7-tert-butyl-1-pyrenyl)-9,9′-spirobi [9H-fluorene] (4-PySBF, 100 MHz, CDCl3).
Molecules 15 07582 g004
Figure S3. 1H-NMR spectrum of 9,9′-dioctyl-2,7-bis (7-tert-butyl-1-pyrenyl)-9H-fluorene (2-PyF, 400 MHz, CDCl3).
Figure S3. 1H-NMR spectrum of 9,9′-dioctyl-2,7-bis (7-tert-butyl-1-pyrenyl)-9H-fluorene (2-PyF, 400 MHz, CDCl3).
Molecules 15 07582 g005
Figure S4. 13C-NMR spectrum of 9,9′-dioctyl-2,7-bis (7-tert-butyl-1-pyrenyl)-9H-fluorene (2-PyF, 100 MHz, CDCl3).
Figure S4. 13C-NMR spectrum of 9,9′-dioctyl-2,7-bis (7-tert-butyl-1-pyrenyl)-9H-fluorene (2-PyF, 100 MHz, CDCl3).
Molecules 15 07582 g006
  • Samples Availability: Available from GK.

References and Notes

  1. Yamaguchi, Y.; Tanaka, T.; Kobayashi, S.; Wakamiya, T.; Matsubara, Y.; Yoshida, Z. Light-Emitting Efficiency Tuning of Rod-Shaped π-Conjugated Systems by Donor and Acceptor Groups. J. Am. Chem. Soc. 2005, 127, 9332–9333. [Google Scholar]
  2. Yamaguchi, Y.; Ochi, T.; Wakamiya, T.; Matsubara, Y.; Yoshida, Z. New Fluorophores with Rod-Shaped Polycyano π-Conjugated Structures: Synthesis and Photophysical Properties. Org. Lett. 2006, 8, 717–720. [Google Scholar] [CrossRef]
  3. Yamaguchi, Y.; Matsubara, Y.; Ochi, T.; Wakamiya, T.; Yoshida, Z. How the π-Conjugation Length Affects the Fluorescence Emission Efficiency. J. Am. Chem. Soc. 2008, 130, 13867–13869. [Google Scholar]
  4. Kanibolotsky, A.L.; Berridge, R.; Skabara, P.J.; Perepichka, I.F.; Bradley, D.D.C.; Koeberg, M. Synthesis and Properties of Monodisperse Oligofluorene-Functionalized Truxenes: Highly Fluorescent Star-Shaped Architectures. J. Am. Chem. Soc. 2004, 126, 13695–13702. [Google Scholar]
  5. Leventis, N.; Rawashdeh, A.M.M.; Elder, I.A.; Yang, J.; Dass, A.; Sotiriou-Leventis, C. Synthesis and Characterization of Ru(II) Tris(1,10-phenanthroline)-Electron Acceptor Dyads Incorporating the 4-Benzoyl-N-methylpyridinium Cation or N-Benzyl-N‘-methyl Viologen. Improving the Dynamic Range, Sensitivity, and Response Time of Sol−Gel-Based Optical Oxygen Sensors. Chem. Mater. 2004, 16, 1493–1506. [Google Scholar] [CrossRef]
  6. Barltrop, J.A.; Coyle, J.D. Principles of Photochemistry; Wiley: New York, NY, USA, 1987; p. 68. [Google Scholar]
  7. Sagari, T.P.I.; Spehr, T.; Siebert, A.; Fuhmann-Lieker, T.; Salbeck, J. Spiro Compounds for Organic Optoelectronics. Chem. Rev. 2007, 107, 1011–1065. [Google Scholar]
  8. Cocherel, N.; Poriel, C.; Vignau, L.; Bergamini, J.; Rault-Berthelot, J. A New Blue Emitter for Nondoped Organic Light Emitting Diode Applications. Org. Lett. 2010, 12, 452–455. [Google Scholar] [CrossRef]
  9. Thirion, D.; Poriel, C.; Barriere, F.; Metivier, R.; Jeannin, O.; Rault-Berthelot, J. Tuning the Optical Properties of Aryl-Substituted Dispirofluorene-indenofluorene Isomers through IntramolecularExcimer formation. Org. Lett. 2009, 11, 4794–4797. [Google Scholar]
  10. Luo, J.; Zhou, Y.; Niu, Z.Q.; Zhou, Q.F.; Ma, Y.; Pei, J. Three-Dimensional Architectures for Highly Stable Pure Blue Emission. J. Am. Chem. Soc. 2007, 129, 113414–113415. [Google Scholar]
  11. Pudzich, R.; Fuhrmann-Lieker, T.; Salbeck, J. Spiro Compounds for Organic Electroluminescence and Related Applications. Adv. Polym. Sci. 2006, 199, 83–142. [Google Scholar] [CrossRef]
  12. Kim, S.Y.; Lee, M.; Boo, B.H. Second molecular hyperpolarizability of 2,2′-diamino-7,7′-dinitro-9,9′-spirobifluorene: An experimental study on third-order nonlinear optical properties of a spiroconjugateddimer. J. Chem. Phys. 1998, 109, 2593–2595. [Google Scholar] [CrossRef]
  13. Poriel, C.; Ferrand, Y.; Maux, P.L.; Rault-Berthelot, J.; Simonneaux, G. Organic Cross-Linked Electropolymers as Supported Oxidation Catalysts: Poly((tetrakis(9,9‘spirobifluorenyl)porphyrin)-manganese) Films. Inorg.Chem. 2004, 43, 5086–5095. [Google Scholar] [CrossRef]
  14. Poriel, C.; Ferrand, Y.; Maux, P.L.; Paul, C.; Rault-Berthelot, J.; Simonneaux, G. Poly(ruthenium carbonyl spirobifluorenylporphyrin): A new polymer used as a catalytic device for carbene transfer. Chem. Commun. 2003, 2308–2309. [Google Scholar]
  15. Wong, K.T.; Ku, S.Y.; Cheng, Y.M.; Lin, X.Y.; Hung, Y.Y.; Pu, S.C.; Chou, P.T.; Lee, G.H.; Peng, S.M. Synthesis, Structures, and Photoinduced Electron Transfer Reaction in the 9,9‘-Spirobifluorene-Bridged Bipolar Systems. J. Org. Chem. 2006, 71, 456–465. [Google Scholar] [CrossRef]
  16. Chien, Y.Y.; Wong, K.T.; Chou, P.T.; Cheng, Y.M. yntheses and spectroscopic studies of spirobifluorene-bridged bipolar systems; photoinduced electron transfer reactions. Chem. Commun. 2002, 2874–2875. [Google Scholar]
  17. Sartin, M.M.; Shu, C.; Bard, A.J. Electrogenerated Chemiluminescence of a Spirobifluorene-Linked Bisanthracene: A Possible Simultaneous, Two-Electron Transfer. J. Am. Chem. Soc. 2008, 130, 5354–5360. [Google Scholar]
  18. Kowada, T.; Matsuyama, Y.; Ohe, K. Synthesis, Characterization, and Photoluminescence of Thiophene-Containing Spiro Compounds. Synlett 2008, 12, 1902–1906. [Google Scholar]
  19. Seto, R.; Sato, T.; Kojima, T.; Hosokawa, K.; Koyama, Y.; Takata, T. 9,9′-Spirobifluorene-Containing Polycarbonates: Transparent Polymers with High Refractive Index and Low Birefringence. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 3658–3667. [Google Scholar] [CrossRef]
  20. Lei, T.; Luo, J.; Wang, L.; Ma, Y.; Wang, J.; Cao, Y.; Pei, J. Highly stable blue light-emitting materials with a three-dimensional architecture: Improvement of charge injection and electroluminescence performance. New J. Chem. 2010, 34, 699–707. [Google Scholar] [CrossRef]
  21. Liao, Y.L., Lin; Wong, K.T.; Hou, T.H.; Hung, W.Y. A Novel Ambipolar Spirobifluorene Derivative that Behaves as an Efficient Blue-Light Emitter in Organic Light-Emitting Diodes. Org. Lett. 2007, 9, 4511–4514. [Google Scholar] [CrossRef]
  22. Cocherel, N.; Poriel, C.; Rault-Berthelot, J.; Barriere, F.; Audebrand, N.; Slawin, A.M.Z.; Vignau, L. New 3π-2Spiro Ladder-Type Phenylene Materials: Synthesis, Physicochemical Properties and Applications in OLEDs. Chem. Eur. J. 2008, 14, 11328–11342. [Google Scholar] [CrossRef]
  23. Shimizu, H.; Fujimoto, K.; Furusho, M.; Maeda, H.; Nanai, Y.; Mizuno, K.; Inouye, M. Highly Emissive -Conjugated AlkynylpyreneOligomers: Their Synthesis and Photophysical Properties. J. Org. Chem. 2007, 72, 1530–1533. [Google Scholar]
  24. Oyamada, T.; Akiyama, S.; Yahiro, M.; Saigou, M.; Shiro, M.; Sasabe, H.; Adachi, C. Unusual photoluminescence characteristics of tetraphenylpyrene (TPPy) in various aggregated morphologies. Chem. Phys. Lett. 2006, 421, 295–299. [Google Scholar] [CrossRef]
  25. Uchimura, M.; Watanabe, Y.; Araoka, F.; Watanabe, J.; Takezoe, H.; Konishi, G. Development of Laser Dyes to Realize Low Threshold in Dye-doped Cholesteric Liquid Crystal Lasers. Adv. Mater. 2010, 22, 4473–4478. [Google Scholar] [CrossRef]
  26. Watanabe, Y.; Uchimura, M.; Araoka, F.; Konishi, G.; Watanabe, J.; Takezoe, H. Extremely Low Threshold in a Pyrene-doped Distributed Feedback Cholesteric Liquid Crystal Laser. Appl. Phys. Express 2009, 2, 102501. [Google Scholar] [CrossRef]
  27. Figueira-Duarte, T.M.; Simon, S.C.; Wagner, M.; Druzhinin, S.I.; Zachariasse, K.A.; Müllen, K. Polypyrene Dendrimers. Angew. Chem. Int. Ed. 2008, 47, 10175–10178. [Google Scholar] [CrossRef]
  28. Schweig, A.; Weidner, U.; Hellwinkel, D.; Krapp, W. Spiroconjugation. Angew. Chem. Int. Ed. 1973, 12, 310–311. [Google Scholar] [CrossRef]
  29. Cocherel, N.; Poriel, C.; Jeannin, O.; Yassin, A.; Rault-Berthelot, J. The synthesis, physicochemical properties and anodic polymerization of a novel ladder pentaphenylene. Dyes Pigm. 2009, 83, 339–347. [Google Scholar] [CrossRef]
  30. Poriel, C.; Liang, J.J.; Rault-Berthelot, J.; Barriere, F.; Cocherel, N.; Slawin, A.M.Z.; Horhant, D.; Virboul, M.; Alcaraz, G.; Audebrand, N.; Vignau, L.; Huby, N.; Wantz, G.; Hirsch, L. Dispirofluorene-Indenofluorene Derivatives as New Building Blocks for Blue Organic Electroluminesent Devices and Electroactive Polymers. Chem. Eur. J. 2007, 13, 10055–10069. [Google Scholar]
  31. Tao, S.; Peng, Z.; Zhang, X.; Wang, P.; Lee, C.S.; Lee, S.T. Highly Efficient Non-Doped Blue Organic Light-Emitting Diodes Based on Fluorene Derivatives with High Thermal Stability. Adv. Funct. Mater. 2005, 15, 1716–1721. [Google Scholar] [CrossRef]
  32. Johansson, N.; dos Santos, D.A.; Guo, S.; Cornil, J.; Fahlman, M.; Salbeck, J.; Schenk, H.; Arwin, H.; Bredas, J.L.; Salanek, W.R. Electronic Structure and Optical Properties of Electroluminescent Spiro-type Molecules. J. Chem. Phys. 1997, 107, 2542–2549. [Google Scholar] [CrossRef]
  33. This phenomenon is thought to be caused by intramolecular overlapping of the electron clouds of the 2,7-dipyrenylfluorene units. Few researchers have successfully observed distinct and strong spiroconjugation in the excited state [29,30]. However, in general, observed spiroconjugations were weak [31]. For example, in ref. [31], the authors reported that a compound in which two pyrenes were introduced into the spirobifluorene skeleton (2,7-dipyrene-9,9′-spirobifluorene, SDPF) showed a red-shift of 4 nm (in the ground state) and 1 nm (in the excited state) in comparison with a model compound, 2,7-dipyrene-9,9′-dimethyl fluorene (DPF). Futhermore in ref. [32], the authors compared UV-Vis spectra between a spiro-compound and an appropriate model compound by using MO calculations. However, to the best of our knowledge, no comparison between a 2,7-diarylfluorene and an appropriate spiro-compound (2,2’,7,7’-tetraarylspirobifluorene) has been carried out yet to investigate the spiroconjugation by experimental methods. This pyrene-fluorene-conjugated compound (4-PySBF) exhibited extremely large red-shift of fluorescence (20 nm). This phenomenon is very interesting. We intend to calculate the electron states of these compounds in the ground and excited states by the MO method and discuss this conjugation system in greater detail.
  34. Maeda, H.; Maeda, T.; Mizuno, K.; Fujimoto, K.; Shimizu, H.; Inouye, M. Alkynylpyrenesas Improved Pyrene-Based Biomolecular Probes with the Advantages of High Fluorescence Quantum Yields and Long Absorption/Emission Wavelengths. Chem. Eur. J. 2006, 12, 824–831. [Google Scholar] [CrossRef]
  35. Abdel-Shafi, A.A.; Wilkinson, F. Charge Transfer Effects on the Efficiency of Singlet Oxygen Production Following Oxygen Quenching of Excited Singlet and Triplet States of Aromatic Hydrocarbons in Acetonitrile. J. Phys. Chem. A. 2000, 104, 5747–5757. [Google Scholar] [CrossRef]
  36. Kyushin, S.; Ikarugi, M.; Goto, M.; Hiratsuka, H.; Matsumoto, H. Synthesis and Electronic Properties of 9,10-Disilylanthracenes. Organometallics 1996, 15, 1067–1070. [Google Scholar]
  37. Usui, Y.; Shimizu, N.; Mori, S. Mechanism on the Efficient Formation of Singlet Oxygen by Energy Transfer from Excited Singlet and Triplet States of Aromatic Hydrocarbons. Bull. Chem. Soc. Jpn. 1992, 65, 897–902. [Google Scholar] [CrossRef]
  38. McLean, A.J.; McGarvey, D.J.; Truscott, T.G.; Lambert, C.R.; Land, E.J. Effect of Oxygen-enhanced Intersystem Crossing on the observartion efficiency of Formation of Singlet Oxygen. J. Chem. Soc. Faraday Trans. 1990, 86, 3075–3080. [Google Scholar] [CrossRef]
  39. Pyrene has a kisc value of 1.2 × 106 S−1 in cyclohexane, see Kikuchi, K. A new method for determining the efficiency of enhanced intersystem crossing in fluorescence quenching by molecular oxygen. Chem. Phys. Lett. 1991, 183, 103–106. [CrossRef]

Share and Cite

MDPI and ACS Style

Sumi, K.; Konishi, G.-i. Synthesis of a Highly Luminescent Three-Dimensional Pyrene Dye Based on the Spirobifluorene Skeleton. Molecules 2010, 15, 7582-7592. https://doi.org/10.3390/molecules15117582

AMA Style

Sumi K, Konishi G-i. Synthesis of a Highly Luminescent Three-Dimensional Pyrene Dye Based on the Spirobifluorene Skeleton. Molecules. 2010; 15(11):7582-7592. https://doi.org/10.3390/molecules15117582

Chicago/Turabian Style

Sumi, Kentaro, and Gen-ichi Konishi. 2010. "Synthesis of a Highly Luminescent Three-Dimensional Pyrene Dye Based on the Spirobifluorene Skeleton" Molecules 15, no. 11: 7582-7592. https://doi.org/10.3390/molecules15117582

Article Metrics

Back to TopTop