A Radical Anion of Structurally Constrained Triphenylborane

Jul 10, 2013 - Soichiro Nakatsuka , Hajime Gotoh , Keisuke Kinoshita , Nobuhiro .... Tomokatsu Kushida , Ayumi Shuto , Masafumi Yoshio , Takashi Kato ...
2 downloads 0 Views 881KB Size
Communication pubs.acs.org/Organometallics

A Radical Anion of Structurally Constrained Triphenylborane Tomokatsu Kushida and Shigehiro Yamaguchi* Institute of Transformative Bio-Molecules (WPI-ITbM) and Department of Chemistry, Graduate School of Science, Nagoya University, Furo, Chikusa, Nagoya 464-8602, Japan S Supporting Information *

ABSTRACT: Chemical reduction of the structurally constrained triphenylborane 1 with K produced a radical anion. EPR analysis demonstrated the delocalization of the unpaired electron spin density over the entire π skeleton to a greater extent than is the case for a nonconstrained Ph3B radical anion. DFT calculations indicated that the coplanarization of the benzene rings with the boron plane and the shortening of the B−C bond lengths by the structural constraint are responsible for the spin delocalization. The geometry optimization also suggested a bowl-shaped conformation as a viable local minimum structure in addition to the planar conformation. X-ray crystal structure analysis indeed revealed the nonplanar structure of the radical anion 1•−. series of more π-extended planarized boranes has also been synthesized as a model system of the boron-doped graphene.12 The planarized triphenylborane was also a useful precursor to produce a borataanthracene, which showed an intense fluorescence.13 In addition, compound 1 itself showed intriguing fluorescence properties with a significantly redshifted emission despite its constrained structure.11 These results suggest the great potential of this skeleton as a building unit for more elaborated triarylborane-based materials, including electron-transporting materials. This consideration prompted us to study the reduced state of the planarized triphenylborane. We now report the chemical reduction of the planarized triphenylborane 1 to produce its radical anion. The structure of the product was successfully determined by an Xray crystallographic analysis. We also investigated the electronic structure, especially the distribution of the unpaired electron spin density, of the radical anion by an EPR analysis as well as quantum chemical calculations. We revealed that the structural constraint in the planar fashion not only effectively delocalizes the spin density over the molecule but also allows the structural deformation to a nonplanar bowl-shaped conformation upon one-electron reduction. We previously reported the cyclic voltammetry of the planarized triphenylborane 1, which showed a reversible redox wave at a half-wave potential of −2.59 V (vs Fc/Fc+) in THF with n-Bu4N+PF6− as a supporting electrolyte.11 The reversibility indicates the formation of a stable radical anion under the measurement conditions. We therefore conducted the chemical reduction of 1 to isolate the radical anion species.

riarylboranes with extended π conjugation have emerged as an promising class of electronic materials because of their characteristic electron-accepting ability due to conjugation through a vacant p orbital of the boron atom.1 Their application as electron-transporting materials for organic light-emitting diodes (OLEDs) has particularly attracted increasing attention.2 For such uses, the investigation of the one-electron-reduced species should provide crucial information, because they are assumed to be the active species for the electron transportation in the devices. Furthermore, triarylborane radical anions are also of interest as the isoelectronic congener of the triarylmethyl radicals, which represent one of the stable organic radicals.3 From these points of view, the reduction of the triarylboranes with alkali metals to form the corresponding radical anions has been studied for a long time.4 The electrochemical reduction of triarylboranes has also been investigated.5 The electron paramagnetic resonance (EPR) studies of the one-electron-reduced triarylboranes revealed the distribution of the unpaired electron spin density, which is mainly localized on the boron center.6−9 However, the structural information for the radical anions is still quite limited. X-ray crystallographic analysis of the triarylborane radical anion has been achieved only for trimesitylborane, namely [Li(12-crown-4)2]+[Mes3B]•− (Mes = 2,4,6-trimethylphenyl).10 Therefore, accumulation of more abundant knowledge on the structure−property relationship for the radical anions of the triarylboranes is required for propelling the applied chemistry of the triarylborane-based materials. Recently, we have reported the synthesis of the planarized triphenylborane 1, in which three benzene rings are fixed in a coplanar fashion with three methylene bridges.11 Thanks to the structural constraint, the compound and its derivatives exhibit outstanding stability not only to water and oxygen but also to silica gel and amines, despite the absence of the steric protection of the boron atom. On the basis of this finding, a

T

© 2013 American Chemical Society

Special Issue: Applications of Electrophilic Main Group Organometallic Molecules Received: June 15, 2013 Published: July 10, 2013 6654

dx.doi.org/10.1021/om400560u | Organometallics 2013, 32, 6654−6657

Organometallics

Communication

The hyperfine coupling constants for the peripheral hydrogen atoms of 1·K were comparable to those of a neutral carbon congener 2, in which the a(p-1H) and a(m-1H) values were reported to be 3.13 and 0.95 G, respectively (Figure 1c).3c These results demonstrate that the unpaired electron spin is similarly distributed on these methylene-bridged planarized structures regardless of the central atoms. In other words, the planarized framework in 1·K is beneficial to effective delocalization of the unpaired electron spin density over the entire molecule. We conducted theoretical calculations at the UB3PW91/631+G(d) level of theory,14 to clarify the origin of the greater extent of spin delocalization in the structurally constrained system 1·K. The structural optimization gave the planar conformation p-1•− as an energy-minimum structure. We first compared the spin density on the boron atom between the optimized structures of p-1•− and [Ph3B]•−. The calculated spin density on the boron center in p-1•− (0.502) was smaller than that of [Ph3B]•− (0.574) (Figure 2a,b). Since the spin

The reaction of 1 with potassium metal in THF resulted in an immediate color change of the solution from colorless to deep blue. Recrystallization from the reaction mixture by slow diffusion of a hexane solution of [2.2.2]cryptand gave deep blue crystals of a potassium salt of the radical anion 1·K in 62% yield (Scheme 1). The compound 1·K was thermally stable in the Scheme 1. Chemical Reduction of 1

solid state. It did not decompose up to 200 °C, at which point the deep blue color of the crystals turned to reddish purple and then black. The EPR spectrum of the radical anion 1·K was measured in THF at 195 K (Figure 1a), which showed hyperfine splitting,

Figure 2. Calculated unpaired electron spin densities on the boron atom ρ(B) in (a) the optimized geometry of p-1•−, (b) the optimized geometry of [Ph3B]•−, and model geometries of [Ph3B]•− with (c) fixed dihedral angles or (d) shortened B−C bonds (UB3PW91/631+G(d)). Figure 1. (a) EPR spectrum of 1·K in THF at 195 K together with its simulated spectrum and hyperfine coupling constants of (b) the planarized triphenylborane radical anion 1·K and (c) the planarized triphenylmethyl radical 2.

density ρ is correlated to the hyperfine coupling constants a in the EPR spectrum with McConnell’s equation a = Q·ρ, where Q is the proportional constant, 15 the calculated spin distributions are in good agreement with experimental results that 1·K has a smaller value of a(11B) than does [Ph3B]•−. There are two structural differences between the planarized radical anion p-1•− and [Ph3B]•−. One is the dihedral angles of the phenyl groups to the plane of the central boron moiety, which are 0° in p-1•−, while those in [Ph3B]•− are twisted by 29.4°. The other difference is the B−C bond lengths. We previously reported that in the neutral state the constrained 1 has shorter B−C bond lengths (1.519(2)−1.520(2) Å)11 by about 0.05−0.07 Å in comparison to those in Ph3B (1.571(3)− 1.589(5) Å).16 In the radical anion species p-1•−, the B−C bond length (1.516 Å) is also shorter than that (1.572 Å) in [Ph3B]•−. The extent of the difference (0.056 Å) is comparable to that in the neutral species. The B−C bond length tends to be slightly shortened by the one-electron reduction. This is reasonable in light of the shape of the LUMO for triphenylboranes, which consists of a bonding interaction

and the g value was 2.0039. The simulation analysis successfully reproduced the spectrum. The hyperfine coupling constants for the boron atom and the peripheral hydrogen atoms were estimated as shown in Figure 1b. The most notable feature is the hyperfine coupling constant a(11B) for the boron of 6.2 G, which is smaller than those of the radical anions of the conventional triarylboranes, such as [Ph3B]•− (7.84 G)7 and [Mes3B]•− (9.87 G).8 This result suggested less localized character of the spin density on the boron center in the planarized radical anion 1·K. Furthermore, the hyperfine coupling constant for the peripheral hydrogen atoms (a(p-1H) = 3.1 G and a(m-1H) = 0.8 G) were larger than those of [Ph3B]•−, in which a(p-1H) and a(m-1H) are 2.73 and 0.670 G, respectively.7 These comparisons clearly indicated that the unpaired electron spin is more delocalized over the three benzene rings in the planarized skeleton. A comparison with isoelectronic triphenylmethyl radicals is also worthy of note. 6655

dx.doi.org/10.1021/om400560u | Organometallics 2013, 32, 6654−6657

Organometallics

Communication

between the p orbital of the boron atom and the π* orbitals of the benzene moieties. To elucidate the impacts of these structural differences on the spin delocalization, we constructed several structurally constrained model geometries for [Ph3B]•− and assessed their spin distribution based on the single point calculations. When the dihedral angles of the phenyl groups to the boron plane were fixed to 0° while the B−C bond lengths of 1.572 Å were maintained, the spin density on the boron atom decreased to 0.538 (Figure 2c). On the other hand, when the B−C bond lengths were shortened to 1.520 Å while the propeller structure was maintained with a dihedral angle of 29.4°, this deformation also lowered the spin density to 0.536 (Figure 2d). Namely, these two different structural deformations result in a comparable degree of decrease in the spin density on the boron center with respect to each other. This result suggests that both the coplanarization of the triphenylborane skeleton and the shortening of the B−C bond lengths almost equally contribute to the effective spin delocalization in 1•−. Notably, the structural optimization of the radical anion 1•− at the UB3PW91/6-31+G(d) level also gave another local minimum structure in addition to the planar p-1•− geometry, which has the bowl-shaped conformation b-1•− (Figure 3). The

Figure 4. (a) ORTEP drawing of 1·K (50% probability for thermal ellipsoids). THF and hydrogen atoms are omitted for clarity. (b) Side view of the anion part of 1·K. Selected bond lengths (Å) and angles (deg): B(1)−C(1), 1.517(4); B(1)−C(10), 1.523(4); B(1)−C(16), 1.521(4); C(1)−B(1)−C(10), 118.8(3); C(1)−B(1)−C(16), 117.5(3); C(10)−B(1)−C(16), 118.0(3).

one benzene ring largely deviates from the central boron moiety (Figure 4b). The dihedral angle between the deviated benzene ring and the plane that consists of the boron atom and the three ipso carbon atoms (C1, C10, and C16) is 23.7°, while those for the other two benzene rings are 14.6 and 10.5°, respectively. As for the geometry around the boron atom, the B−C bond lengths are 1.517(4), 1.523(4), and 1.521(4) Å for the B1−C1, B1−C10, and B1−C16 bonds, respectively, which are slightly shorter than the calculated values (1.531 Å) and comparable to those in the crystal structure of the neutral 1 (1.519(2)−1.520(2) Å). 11 The boron atom is slightly pyramidalized. The sum of the three C−B−C angles is 354.3°. These structural parameters imply that the triphenylborane framework constrained with the three methylene bridges is still flexible enough to form a bowl-shaped structure upon one-electron reduction. According to the structural optimization at the B3PW91/631+G(d) level, the neutral 1 has a planar conformation, which is consistent with its X-ray crystal structure.11 The bowl-shaped structure is not a local minimum on the potential energy surface in the ground state, and therefore the plane-to-bowl structural change is not likely to occur in the ground state of the neutral 1. In contrast, the increase in the electron density on the boron center by the one-electron reduction renders the bowl-shaped structure a local minimum on the energy surface. This result is interesting, because it is opposite to the general idea of the structural constraint in a planar fashion, which is an effective strategy for making the skeleton rigid. Key for producing this discrepancy are the compressed B−C bonds in the neutral 1. The structural constraint with the three methylene bridges in the triphenylborane framework forces the B−C bonds to be shortened to a significant extent. The change in the electronic configuration of the boron atom in such a confined situation results in inducing a significant structural deformation through the elongation of the B−C bonds. This structural change may be also relevant to the structural deformation of 1 in the excited state, which showed a

Figure 3. Optimized structures of the radical anions p-1•− and b-1•− and the energy difference ΔE0 calculated at the UB3PW91/6-31+G(d) level of theory.

b-1•− conformation was estimated to be energetically less stable than the planar conformation p-1•− by only +0.70 kcal/mol. This rather small ΔE0 value implies that p-1•− and b-1•− likely convert into each other at ambient temperature. Although we also investigated the viability of the pyramidalized conformation for the nonconstrained [Ph3B]•−, we could not find such a geometry as a local minimum in the geometry optimization. Thus, the bowl-shaped conformation is unique to the structurally constrained triphenylborane framework. The B−C bond length in b-1•− is 1.531 Å, which is slightly elongated than that (1.516 Å) in the p-1•− conformation. In the b-1•− conformation, the sum of the C−B−C angles is 352.9°. The viability of the bowl-shaped conformation b-1•− has been demonstrated by an X-ray crystal structure analysis of 1·K with [2.2.2]cryptand.17 As shown in Figure 4, it was revealed that the radical anion 1•− does not have a planar but rather a shallow bowl-shaped structure. This is in contrast to the planar geometry around the boron center reported for [Li(12-crown4)2]+[Mes3B]•−.10 Whereas pyramidalized geometries were reported for amine- or phosphine-coordinated boryl radicals,18 a nonplanar structure is unprecedented for triarylborane radical anions. The structure is not a symmetrical bowl shape, where 6656

dx.doi.org/10.1021/om400560u | Organometallics 2013, 32, 6654−6657

Organometallics

Communication

Chem. Soc. 1971, 93, 2215. (c) Neugebauer, F. A.; Hellwinkel, D.; Aulmich, G. Tetrahedron Lett. 1978, 19, 4871. (4) (a) Krause, E.; Polack, H. Ber. Dtsch. Chem. Ges. 1926, 59, 777. (b) Krause, E.; Polack, H. Ber. Dtsch. Chem. Ges. 1928, 61, 271. (c) Bent, H. E.; Dorfman, M. J. Am. Chem. Soc. 1935, 57, 1259. (d) Chu, T. L. J. Am. Chem. Soc. 1953, 75, 1730. (e) Chu, T. L.; Weismann, T. J. J. Am. Chem. Soc. 1956, 78, 23. (f) Chu, T. L.; Weismann, T. J. J. Am. Chem. Soc. 1956, 78, 3610. (g) Brown, H. C.; Dodson, V. H. J. Am. Chem. Soc. 1957, 79, 2302. (h) Moeller, C. W.; Wilmarth, W. K. J. Am. Chem. Soc. 1959, 81, 2638. (5) DuPont, T. J.; Mills, J. L. J. Am. Chem. Soc. 1975, 97, 6375. (6) Weissman, S. I.; van Willigen, H. J. Am. Chem. Soc. 1965, 87, 2285. (7) Leffler, J. E.; Watts, G. B.; Tanigaki, T.; Dolan, E.; Miller, D. S. J. Am. Chem. Soc. 1970, 92, 6825. (8) Griffin, R. G.; van Willigen, H. J. Chem. Phys. 1972, 57, 86. (9) Kwaan, R. J.; Harlan, C. J.; Norton, J. R. Organometallics 2001, 20, 3818. (10) Olmstead, M. M.; Power, P. P. J. Am. Chem. Soc. 1986, 108, 4235. (11) Zhou, Z.; Wakamiya, A.; Kushida, T.; Yamaguchi, S. J. Am. Chem. Soc. 2012, 134, 4529. (12) (a) Saito, S.; Matsuo, K.; Yamaguchi, S. J. Am. Chem. Soc. 2012, 134, 9130. (b) Dou, C.; Saito, S.; Matsuo, K.; Hisaki, I.; Yamaguchi, S. Angew. Chem., Int. Ed. 2012, 51, 12206. (13) (a) Kushida, T.; Zhou, Z.; Wakamiya, A.; Yamaguchi, S. Chem. Commun. 2012, 48, 10717. (b) Kushida, T.; Yamaguchi, S. Angew. Chem., Int. Ed. 2013, DOI: 10.1002/anie.201303830. (14) The theoretical calculations were performed with the Gaussian 09 program. (15) (a) McConnell, H. M. J. Chem. Phys. 1956, 24, 764. (b) McConnell, H. M.; Chesnut, D. B. J. Chem. Phys. 1958, 28, 107. (16) Zettler, F.; Hausen, H. D.; Hess, H. J. Organomet. Chem. 1974, 72, 157. (17) In the packing structure, while C−H···π interactions between the CH2 moieties of [2.2.2]cryptand and the benzene moieties or the boron atom of 1•− are observed (Figure S1, Supporting Information), no significant interaction between the radical anion 1•− and the potassium ion is detected, in which the shortest B···K distance is 6.50 Å. (18) (a) Paul, V.; Roberts, B. P.; Robinson, C. A. S. J. Chem. Res., Synop. 1988, 264. (b) Kirwan, J. N.; Roberts, B. P. J. Chem. Soc., Perkin Trans. 2 1989, 539.

significant red-shifted emission despite the structural constraint. A detailed study of the structural change in the excited state will be reported elsewhere. In summary, we have succeeded in the synthesis and characterization of the planarized triphenylborane radical anion 1·K. The EPR analysis of 1·K demonstrated that the unpaired electron spin is more delocalized over the three benzene rings in comparison to the nonconstrained [Ph3B]•−. The theoretical calculations predicted that the bowl-shaped geometry is a viable conformation for 1•− in addition to the planar conformation. The small energy difference between the planar and bowlshaped conformations suggested that the plane-to-bowl conversion likely takes place at room temperature. Actually, the X-ray crystal structure analysis revealed that 1·K adopts a nonplanar bowl-shaped conformation. These structural features are unique for the structurally constrained triphenylborane skeleton. These findings demonstrated the impacts of the structural modifications on the intrinsic characters of the boron-centered radical species and should provide important information on the application of the planarized triphenylborane skeleton as a building unit for more extended π-conjugated materials.



ASSOCIATED CONTENT

S Supporting Information *

Text, a figure, tables, and a CIF file giving experimental procedures, results of theoretical calculations, and crystallographic data for 1·K. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail for S.Y.: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by CREST, JST, and a Grant-in-Aid for Scientific Research on Innovative Areas (Stimuli-responsive Chemical Species) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan. T.K. thanks the JSPS Research Fellowship for Young Scientists. We greatly thank Prof. K. Awaga, Prof. M. M. Matsushita, Dr. H. Yoshikawa, Dr. Y. Shuku, and Mr. K. Matsuura (Nagoya University) for measurements and valuable discussions on EPR spectroscopy.



REFERENCES

(1) (a) Entwistle, C. D.; Marder, T. B. Angew. Chem., Int. Ed. 2002, 41, 2927. (b) Entwistle, C. D.; Marder, T. B. Chem. Mater. 2004, 16, 4574. (c) Jäkle, F. Coord. Chem. Rev. 2006, 250, 1107. (d) Yamaguchi, S.; Wakamiya, A. Pure Appl. Chem. 2006, 78, 1413. (e) Jäkle, F. Chem. Rev. 2010, 110, 3985. (2) (a) Noda, T.; Shirota, Y. J. Am. Chem. Soc. 1998, 120, 9714. (b) Noda, T.; Ogawa, H.; Shirota, Y. Adv. Mater. 1999, 11, 283. (c) Shirota, Y.; Kinoshita, M.; Noda, T.; Okumoto, K.; Ohara, T. J. Am. Chem. Soc. 2000, 122, 11021. (d) Noda, T.; Shirota, Y. J. Lumin. 2000, 87−89, 1168. (e) Shirota, Y. J. Mater. Chem. 2000, 10, 1. (f) Doi, H.; Kinoshita, M.; Okumoto, K.; Shirota, Y. Chem. Mater. 2003, 15, 1080. (g) Tanaka, D.; Takeda, T.; Chia, T.; Watanabe, S.; Kido, J. Chem. Lett. 2007, 36, 262. (h) Sun, C.; Hudson, Z. M.; Helander, M. G.; Lu, Z.-H.; Wang, S. Organometallics 2011, 30, 5552. (3) (a) Gomberg, M. J. Am. Chem. Soc. 1900, 22, 757. (b) Ballester, M.; Riera-Figueras, J.; Castañer, J.; Badfa, C.; Monsó, J. M. J. Am. 6657

dx.doi.org/10.1021/om400560u | Organometallics 2013, 32, 6654−6657