Improved Dynamic Properties of Charge-Transfer-Type

Sep 29, 2015 - Synopsis. Vacuum co-sublimation gave a charge-transfer complex, (coronene)2F4TCNQ, consisting only of DDA-type alternating π-columns. ...
1 downloads 17 Views 3MB Size
Article pubs.acs.org/crystal

Improved Dynamic Properties of Charge-Transfer-Type Supramolecular Rotor Composed of Coronene and F4TCNQ Yukihiro Yoshida,*,† Yoshihide Kumagai,‡ Motohiro Mizuno,‡ Kazuhide Isomura,§ Yuto Nakamura,§ Hideo Kishida,§ and Gunzi Saito†,∥ †

Faculty of Agriculture, Meijo University, Shiogamaguchi 1-501, Tempaku-ku, Nagoya 468-8502, Japan Department of Chemistry, Graduate School of Natural Science & Technology, Kanazawa University, Kakuma, Kanazawa, Ishikawa 920-1192, Japan § Department of Applied Physics, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan ∥ Toyota Physical and Chemical Research Institute, 41-1 Yokomichi, Nagakute, Aichi 480-1192, Japan ‡

S Supporting Information *

ABSTRACT: A charge-transfer (CT)-type supramolecular rotor, (coronene)2F4TCNQ, was obtained by vacuum cosublimation. The CT complex has an unprecedented crystal structure consisting only of DDA-type alternating π-columns of coronene (D) and electron acceptor (A), and coronene molecules with a pronounced undulating structure form no C−H···F hydrogen bonds with adjacent F4TCNQ molecules in the side-by-side direction. These structural features are in contrast with those reported for (coronene)F4TCNQ that has DA-type alternating π-columns, which was obtained by diffusion in dichloromethane/pentane. Coronene molecules in the present complex undergo an in-plane rotation in the gigahertz region at 233 K, which is 3 orders of magnitude faster than that in (coronene)F4TCNQ. Furthermore, the activation energy for the rotation was found to be about half of the value reported for (coronene)F4TCNQ. These results clearly demonstrate that dynamic properties of the assemblies (CT complexes) can be varied by changing the assembly method (crystallization method), even when the parts (molecules) used as rotator and stator components are the same.



INTRODUCTION Supramolecular rotors that emulate the structures of macroscopic rotary devices can be designed for specific purposes by preselecting diverse combinations of the component molecules as rotators and stators in the assemblies.1−4 Formation of the assemblies is certainly driven by intermolecular (noncovalent) attractions, such as Coulomb, hydrogen-bonding, and van der Waals (including π−π) interactions, which can be used to tune the molecular arrangements and that play a crucial role in the stability of the assemblies. In most cases, such a supramolecular approach is free from any complicated organic synthesis to create the covalent bonds between the rotator and stator moieties. Thus far, highly symmetric planar molecules such as cyclopentadienyls, 1,5−7 porphyrins, 1,2,8−10 and crown ethers3,4,11,12 have been used as rotators for supramolecular rotors. We have recently focused on a polycyclic aromatic hydrocarbon (PAH), coronene (Figure 1a), which is the smallest D6h-symmetric PAH molecule and is a model molecule for fragmented graphene, because of its promising low rotational barrier.13−17 We reported dynamic properties of various neutral charge-transfer (CT) complexes combining the coronene molecule as an electron donor (D) with 7,7,8,8tetracyanoquinodimethane (TCNQ) derivatives as electron © XXXX American Chemical Society

Figure 1. Molecular structures and atomic numbering of (a) coronene and (b) F4TCNQ in 2 at 100 K (gray, C; white, H; blue, N; lime green, F). Thermal ellipsoids are shown at 50% probability.

acceptors (A).18,19 Among them, a 1:1 CT complex formed with F4TCNQ (Figure 1b), (coronene)(F4TCNQ) (1), which was prepared by room-temperature diffusion in a dichloromethane/pentane mixed solution, exhibited the slowest inplane molecular rotation, mainly due to the multiple side-byReceived: August 8, 2015 Revised: September 22, 2015

A

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

side C−H···N and C−H···F hydrogen bonds as well as the faceto-face π−π interactions between adjacent coronene and F4TCNQ molecules.19 In this study, we obtained another CT complex composed of coronene and F4TCNQ, (coronene)2F4TCNQ (2), by vacuum co-sublimation of coronene and F4TCNQ. Whereas the molecular packing of the coronenebased CT complexes is limited to either a DA-type alternating π-column (Figure 2a) or a DDA-type alternating one flanked by

Figure 2. Molecular arrangements of CT complexes composed of coronene and TCNQ analogues, where coronene and TCNQ analogues appear in green and red, respectively. (a) DA-type alternating column in 1,19 (b) DDA-type alternating column flanked by another coronene in (coronene)3TCNQ,18 and (c) genuine DDAtype alternating column in 2.

Figure 3. Infrared transmission spectra of 1 (green)18 and 2 (blue) together with monoanionic F4TCNQ (red), neutral F4TCNQ (orange), and neutral coronene (pink), measured in dispersed KBr pellets.

coronene (1.23 V vs saturated calomel electrode (SCE)) compared to the E11/2(A) value of F4TCNQ (0.60 V vs SCE). An electronic absorption spectrum measured using the KBr pellet showed a CT band at 9.9 × 103 cm−1 in addition to the intramolecular transition bands of the component molecules above 20 × 103 cm−1 (Figure S2, Supporting Information). Complex 2 crystallizes in a monoclinic lattice P21/n (Table S1, Supporting Information) and undergoes no structural phase transition in the whole measured temperature range (100−298 K). Hereafter, we use the crystallographic data at 100 K unless otherwise stated. One coronene molecule and a half of the F4TCNQ structure are crystallographically independent. Theoretical calculations predicted that the coronene monocation undergoes an in-plane Jahn−Teller distortion caused by the splitting of the degenerate of the e2u HOMO level.21,22 Figure 4a displays the deviation of each rim carbon atom from the average of the distances of the atoms from the molecular centroid in 2 (Δdi = di − ∑di/12, where di is the distance from carbon atom with a number i to the molecular centroid). The root-mean-square deviation of Δdi (dRMS = (∑Δdi2/12)1/2) was estimated to be 2.52 × 10−3 Å, which is comparable to 1.73 × 10−3 Å in 119 and 1.70 × 10−3 Å in pristine coronene crystal, but it is much lower than 1.25 × 10−2 Å of a D2h-symmetric coronene monocation optimized by density functional theory (DFT) calculations (Figure 4b). The nearly isotropic structure of coronene molecules in 2 is firm evidence of the absence of a Jahn−Teller distortion, namely, the neutral ground state. As for the charge state of the F4TCNQ molecule, the ratio of bond lengths, c/(b + d) (see Figure 1b), is a diagnostic parameter for determining δ in the crystal, in view of the structural similarity between TCNQ and F4TCNQ.23,24 The δ value in 2, which was estimated to be −0.15(5) on the basis of a standard line through pristine F4TCNQ solid (δ = 0)25 and monoanionic F4TCNQ salts (δ = −1),26−28 is in good agreement with that expected from the vibrational spectra described above. We note that the X-ray data at 298 K yield a δ value of −0.02(12), which is identical within the experimental error to that at 100 K.

another coronene (Figure 2b), complex 2 has an unprecedented crystal structure consisting only of DDA-type alternating π-columns (Figure 2c). In contrast to 1, each coronene molecule has a pronounced undulating structure and is free of C−H···F hydrogen bonds with adjacent F4TCNQ molecules; the latter feature, in particular, could modify the dynamic behavior toward a reduced rotational barrier. Indeed, solid-state 1H NMR measurements indicate in-plane rotation of component coronene molecules in the gigahertz region at 233 K, which is 3 orders of magnitude faster than that in 1. The drastically different dynamic properties in the assemblies formed with the same components will create many opportunities for selecting synthesis conditions as well as for selecting and designing components for supramolecular rotors.



RESULTS AND DISCUSSION Dark brown single crystals of 2 (Figure S1, Supporting Information) were prepared by co-sublimation of purified coronene and F4TCNQ, which were ground together and heated in a borosilicate glass tube sealed under vacuum (see Experimental Section for details). The present result suggests that complex 1 is a kinetic product, whereas complex 2 is a thermodynamic one. Here, we note that complex 2 with less crystallinity can also be obtained by diffusion in tetrahydrofuran solution and can be completely transformed into the dark green crystals of 1 and pale yellow crystals of pristine coronene by exposure to benzene. As shown in Figure 3, the b1u νCC mode of 2 (1591 cm−1) is apparently shifted to a lower frequency by 9 cm−1 compared to that of neutral F4TCNQ (1600 cm−1). Assuming that the charge (δ) of F4TCNQ is linearly correlated with the νCC shift, namely, δ = −(ωobs − ω0)/(ω1 − ω0), where ωobs, ω0, and ω1 are the νCC mode frequencies of the CT complex, neutral F4TCNQ, and monoanionic F4TCNQ, respectively,20 the δ value of 2 was estimated to be about −0.15, which is comparable to the value obtained for 1.18,19 Also, the neutral ground state is in accordance with the larger E11/2(D) value of B

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 4. Deviations from the average value of the distances of 12 rim carbon atoms from the molecular centroid (Δd) on a cobweb chart for (a) coronene in 2 and (b) coronene monocation with a structure optimized by the DFT method (D2h symmetry). (c) Out-of-plane deviations from a mean plane containing 24 carbon atoms (Δh) on a bar graph. The red and blue spheres (left panel) and the bars (right upper panel) indicate atoms with out-of-plane deviations greater than 0.04 Å perpendicularly out of and into the plane of the paper, respectively. The numbers correspond to those of the carbon atoms in Figure 1a. The right lower panel is the corresponding bar graph for 1.19

Figure 5. (a) Crystal structure of 2 viewed along the b axis. (b) Overlap pattern of coronene molecules within the dimer in 2. Short atomic contacts viewed along the molecular side-by-side directions are shown in (c) and (d), where the green dotted lines show short C···C contacts and the arrows show the directions of the deviation from the mean plane.

(hRMS = 3.51 × 10−2 Å18), possibly associated with the undulating structure in 2. In a previous paper, we reported that the dihedral angle between adjacent coronene and TCNQ analogues within the DA-type alternating π-column exerts an effect on the dynamic properties of the CT complexes such as rotational rate and activation energy.19 Although it is difficult to assess precisely the dihedral angle in 2 due to the undulating structure of coronene, it is apparent that the overlap integral between the HOMO of coronene and the LUMO of F4TCNQ within the column in 2 (1.46 × 10−4) is smaller than that in 1 (2.70 × 10−4). The π-conjugated planes of neighboring columns along the c axis, which are connected through a C−H···N (C18···N2, 3.39 Å; sum of van der Waals radii, 3.25 Å30) hydrogen bond between coronene and F4TCNQ, are nonparallel with a dihedral angle of about 33°. Along the b direction, each coronene molecule connects with another coronene molecule through a C−H···H−C (C4···C4, 3.96 Å) dihydrogen bond31,32 and with a F4TCNQ molecule through a C−H···N (C15···N2, 3.46 Å) hydrogen bond, resulting in a DDA-type infinite array along the [130] direction at z = 0 and 1 and along the [13̅ 0] direction at z = 0.5 (Figure 6a). Particularly when focusing on one coronene molecule, we found that each molecule interacts with only two neighboring F4TCNQ molecules through sideby-side C−H···N hydrogen bonds (Figure 6b). This is strikingly different from 1, which involves coronene molecules interacting with six F4TCNQ molecules through side-by-side C−H···N and C−H···F hydrogen bonds (Figure 6c).19 Earlier studies by Akutagawa et al., which describe the crystal polymorphs of Ni(dmit)2-based (dmit, 1,3-dithiole-2-thione4,5-dithiolate) ionic salts including supramolecular rotators of o-aminoanilinium33 and trans-cyclohexane-1,4-diammonium34

It is notable that the coronene molecules in 2 have an unprecedented undulating structure (right upper panel in Figure 4c), with a root-mean-square deviation from a mean plane containing 24 carbon atoms (hRMS = (∑Δhi2/12)1/2, where Δhi is the out-of-plane deviation of the rim carbon atom with a number i from the mean plane) of 5.29 × 10−2 Å. This value is more than 2 times larger than those in 1 (2.05× 10−2 Å;19 right lower panel in Figure 4c) and pristine coronene crystal (1.37× 10−2 Å). Although DFT calculations predicted that a coronene molecule with C2h symmetry shows an out-ofplane distortion,29 the distortion in 2 is 2 orders of magnitude greater than that in the theoretically optimized structure. It is thus possible that the intermolecular interactions cause the pronounced out-of-plane distortion of the coronene molecules in 2. Two coronene molecules sandwich a F4TCNQ molecule to construct a DDA-type alternating π-column along the a axis (Figure 5a). The coronene dimer, which is located on an inversion center, has a ring-over-atom-type overlap (Figure 5b), as in the case of graphite and (coronene)3TCNQ, with the molecular arrangement shown in Figure 2b.18 As shown in Figure 5c,d, four carbon atoms (blue in Figure 4c; i = 1, 9, 10, and 18) deviate from the least-squares plane to form a face-toface intradimer π−π interaction (C1···C9, 3.32 Å; sum of van der Waals radii, 3.40 Å30), whereas four other carbon atoms (red in Figure 4c; i = 4, 6, 13, and 15) deviate in the direction opposite that of the blue carbon atoms. The overlap integral between the HOMOs of coronene molecules within the dimer (1.35 × 10−4) is apparently smaller than that in (coronene)3TCNQ (2.48 × 10−4) with fairly flat coronene molecules C

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 7. (a) Temperature dependence of the solid-state 1H NMR spectra of 2. (b) Spin−lattice relaxation time (T1) of 2 plotted as a function of the reciprocal temperature. For the red line, see the text.

Figure 7b displays the temperature dependence of T1 of 2 at ν = 294.988 MHz. The T1 value can be described well by the Bloembergen−Purcell−Pound (BPP) relaxation35 involving a single motion above 177 K, and the T1 minimum occurs at 233 K, where ωoτc = 2πντc = 0.616 (ωo, observed frequency, τc, correlation time). Although the motional mode is unclear, the rotational rate (krot) at 233 K can be determined as 1.5 GHz using the relation krot = 1/(2τc) for the assumed 2-fold flipping model. We note the krot value is much higher than those of (coronene)3TCNQ (ca. 200 MHz at 233 K),18 1 (1 MHz at 233 K),19 and pristine coronene crystal (3 MHz at 233 K).19 Coronene molecules in 2, although slower than those in an isotropic three-dimensional ionic complex, (coronene)3Mo6Cl14, with no face-to-face π−π interactions (ca. 300 MHz for slower species and ca. 2.6 GHz for faster species at 233 K),36 were found to be one of the fastest rotators in the coronene-based CT complexes with alternating columns. This confirms the importance of molecular arrangement or intermolecular interactions, i.e., C−H···N and C−H···F hydrogen bonds with adjacent F4TCNQ molecules along the side-byside direction and π−π interactions with adjacent coronene and F4TCNQ molecules along the stacking direction, in determining dynamic properties. Since, in general, the hydrogenbonding interaction is stronger than the π−π interaction,37 the absence of C−H···F hydrogen bonds in 2, in contrast to 1, may be a primary cause of the improved dynamic properties. The estimated activation energy (Ea) of 2 (14 kJ mol−1) is significantly lower than those of 1 (30 kJ mol−119) and pristine coronene crystal (25 kJ mol−138 or 27 kJ mol−119). The reason must be the same as that given above for the high rotational rate. It is worth mentioning that investigation of the arrangement patterns of electron-donating/accepting rotator molecules suitable for reducing the rotational barrier would encourage the rational design of CT-type supramolecular rotors.1−4,13−19,36,39 In particular, the present results strongly suggest the possibility of controlling dynamic properties over a wide range by intermolecular interactions. By extrapolating the Arrhenius law to higher temperatures, it is likely that the molecular rotation of coronene in 2 should occur with a frequency as high as ca. 6.5 GHz at room temperature.

Figure 6. (a) Formation of DDA-type infinite chain along the [130] direction (z = 0 and 1) and the [1̅30] direction (z = 0.5) in 2, where green dotted lines show short atomic contacts. Short atomic contacts around one coronene molecule in (b) 2 and (c) 1,19 where green and orange dotted lines show C−H···N and C−H···F hydrogen bonds, respectively.

cations as showing different frequency-dependent dielectric constants indicative of molecular motion, prompted us to investigate the dynamic properties of 2. We therefore conducted variable-temperature solid-state 1H NMR studies of 2 because the line shape and spin−lattice relaxation time (T1) are very sensitive to molecular motion. The solid-state 1H NMR spectra of polycrystalline 2 exhibit a single broad signal over the whole measured temperature range (113−353 K), with a width of approximately 9.2 kHz at 113 K and 7.0 kHz at 353 K (Figure 7a). Since the crystallographic study revealed that the coronene molecule in 2 undergoes a pronounced in-plane thermal fluctuation at room temperature (Figure S3, Supporting Information), it appears that the motion is mainly originated from the in-plane rotation of coronene molecules, as in other coronene-based CT complexes.18,19 The relative arrangement of coronene molecules in the dimer shown in Figure 5c,d suggests a 180° rotational (2-fold flipping) motion. The 60° rotational (6-fold flipping) motion, which has often been found for coronene-based CT complexes,18,19 may occur when the coronene dimer shows a collective motion. The absence of band broadening at low temperatures due to the freezing of molecular motion (Figure 7a) indicates that the coronene molecules in 2 rotate even at 113 K.



CONCLUSIONS Vacuum co-sublimation of coronene and F4TCNQ afforded a new neutral CT complex with 2:1 stoichiometry. The crystal is the first coronene-based CT complex that comprises only a DDA-type alternating π-column. In the crystal, each coronene D

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

were taken from references.43,44 The calculations were performed with the program developed by Mori et al.45

molecule has a pronounced undulating structure that might lead to the reduced CT interactions along the stacking direction. Because of the significantly weakened intermolecular interactions, especially along the side-by-side direction, compared to those in the reported 1:1 CT complex composed of the same components, coronene molecules in the present complex undergo an in-plane molecular motion with a significantly low rotational barrier; the activation energy for the rotation is lowered to about half the value for the 1:1 CT complex. The rotational rate lies in the gigahertz region at 233 K, which is 3 orders of magnitude faster than that in the 1:1 complex. Of particular importance is that the two CT complexes with different stoichiometries can be obtained separately by changing the synthesis conditions. Since the ability to improve dynamic properties by changing the stoichiometry is a distinctive characteristic of supramolecular assemblies, the present results suggest the possibility of developing a new type of supramolecular rotors and controlling a wide range of dynamic properties. Studies along this line are now in progress.





ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.5b01138. Photographs of single crystals of 1 and 2 (Figure S1), electronic absorption spectra of 1 and 2 (Figure S2), thermal fluctuation of coronene molecule in 2 (Figure S3), and crystallographic data of 2 (Table S1) (PDF). X-ray crystallographic data for 2 (CIF).



AUTHOR INFORMATION

Corresponding Author

*Phone: +81-52-838-2552. Fax: +81-52-833-7200. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



EXPERIMENTAL SECTION

ACKNOWLEDGMENTS The authors thank Hiromi Hayama for experimental support. This work was supported by JSPS KAKENHI grant nos. 23225005, 25288041, and 26110512.

General Details. Tetrahydrofuran (Aldrich, > 99.9%) was used as received, and benzene (Wako Chemical, > 99.5%) was distilled prior to use. Coronene (Tokyo Kasei, 98%) was recrystallized several times from dry benzene, and F4TCNQ was synthesized according to a literature procedure40 and purified by sublimation. Infrared transmission spectra were measured on KBr pellets using a Shimadzu IRPrestage-21 spectrophotometer (380−7800 cm−1). UV−vis−NIR spectra were measured on KBr pellets using a Shimadzu UV-3100 spectrophotometer (3.8−42 × 103 cm−1). Preparation of (Coronene)2F4TCNQ (2). Recrystallized coronene (0.06 mmol) and sublimated F4TCNQ (0.07 mmol) were ground together in an agate mortar and transferred into a 12 mm diameter borosilicate glass tube. After sealing under vacuum (5 × 10−5 mbar), the 13 cm long tube was placed in a two-zone furnace with a temperature gradient from 214 °C (sample side) to 131 °C (opposite side); these temperatures were measured using a thermocouple in contact with the inside wall of the furnace. After heating for 6 days, dark brown rod crystals of 2 with a typical length of 0.5 mm were grown in the center of the tube. X-ray Structural Analysis. A single crystal of 2 was mounted to the end of the glass capillary using a minimal amount of adhesive and was cooled to 100(2) K by a stream of cooled nitrogen gas. Data collections were carried out on a CCD-type diffractometer (Bruker SMART APEX II) with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). The crystal structure was solved by direct methods (SIR200441) and refined by a full-matrix least-squares method on F2 using SHELXL.42 All non-hydrogen atoms were anisotropically refined. The positional parameters of hydrogen atoms were calculated under fixed C−H bond lengths of 0.93 Å with sp2 configuration of the parent carbon atoms. Crystallographic data were deposited with the Cambridge Crystallographic Data Centre: deposition number CCDC 1025088 for 2 at 100 K. Solid-State 1H NMR Measurements. The 1H NMR spectra and T1 were measured using a JEOL ECA-300 spectrometer at 294.988 MHz. The NMR spectra were observed using a solid echo sequence (90°)x − τ − (90°)y − τ − tacq, where τ and tacq are the intervals of echo and acquisition time, respectively. The 90° pulse width and τ were 2.0 and 40 μs, respectively. 1H NMR T1 was measured by the inversion recovery and saturation recovery methods. Measurements of temperature-dependent 1H NMR spectra and T1 were performed during the heating process. Calculation of Overlap Integrals. The overlap integrals between the HOMOs of coronene molecules and between the HOMO of coronene and the LUMO of F4TCNQ were calculated through the extended Hückel method. Parameters for Slater-type atomic orbitals



REFERENCES

(1) Kottas, G. S.; Clarke, L. I.; Horinek, D.; Michl, J. Chem. Rev. 2005, 105, 1281−1376. (2) Shinkai, S.; Takeuchi, M. Bull. Chem. Soc. Jpn. 2005, 78, 40−51. (3) Akutagawa, T.; Nakamura, T. Dalton Trans. 2008, 6335−6345. (4) Vogelsberg, C. S.; Garcia-Garibay, M. A. Chem. Soc. Rev. 2012, 41, 1892−1910. (5) Brydges, S.; Harrington, L. E.; McGlinchey, M. J. Coord. Chem. Rev. 2002, 233−234, 75−105. (6) Kinbara, K.; Aida, T. Chem. Rev. 2005, 105, 1377−1400. (7) Fukino, T.; Joo, H.; Hisada, Y.; Obana, M.; Yamagishi, H.; Hikima, T.; Takata, M.; Fujita, N.; Aida, T. Science 2014, 344, 499− 504. (8) Tashiro, K.; Konishi, K.; Aida, T. J. Am. Chem. Soc. 2000, 122, 7921−7926. (9) Ogi, S.; Ikeda, T.; Wakabayashi, R.; Shinkai, S.; Takeuchi, M. Chem. - Eur. J. 2010, 16, 8285−8290. (10) Écija, D.; Auwärter, W.; Vijayaraghavan, S.; Seufert, K.; Bischoff, F.; Tashiro, K.; Barth, J. V. Angew. Chem., Int. Ed. 2011, 50, 3872− 3877. (11) Ballardini, R.; Balzani, V.; Credi, A.; Gandolfi, M. T.; Venturi, M. Acc. Chem. Res. 2001, 34, 445−455. (12) Akutagawa, T.; Koshinaka, H.; Sato, D.; Takeda, S.; Noro, S.; Takahashi, H.; Kumai, R.; Tokura, Y.; Nakamura, T. Nat. Mater. 2009, 8, 342−347. (13) Boyd, R. K.; Fyfe, C. A.; Wright, D. A. J. Phys. Chem. Solids 1974, 35, 1355−1365. (14) Gavezzotti, A.; Desiraju, G. R. Acta Crystallogr., Sect. B: Struct. Sci. 1988, 44, 427−434. (15) Feng, X.; Marcon, V.; Pisula, W.; Hansen, M. R.; Kirkpatrick, J.; Grozema, F.; Andrienko, D.; Kremer, K.; Müllen, K. Nat. Mater. 2009, 8, 421−426. (16) Karlen, S. D.; Reyes, H.; Taylor, R. E.; Khan, S. I.; Hawthorne, M. F.; Garcia-Garibay, M. A. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 14973−14977. (17) Wheeler, S. E. CrystEngComm 2012, 14, 6140−6145. (18) Yoshida, Y.; Shimizu, Y.; Yajima, T.; Maruta, G.; Takeda, S.; Nakano, Y.; Hiramatsu, T.; Kageyama, H.; Yamochi, H.; Saito, G. Chem. - Eur. J. 2013, 19, 12313−12324. E

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

(19) Yoshida, Y.; Kumagai, Y.; Mizuno, M.; Saito, G. Cryst. Growth Des. 2015, 15, 1389−1394. (20) Horiuchi, S.; Kumai, R.; Okimoto, Y.; Tokura, Y. Chem. Phys. 2006, 325, 78−91. (21) Hoijtink, G. J. Mol. Phys. 1959, 2, 85−95. (22) Kato, T.; Yoshizawa, K.; Yamabe, T. J. Chem. Phys. 1999, 110, 249−255. (23) Kistenmacher, T. J.; Emge, T. J.; Bloch, A. N.; Cowan, D. O. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1982, 38, 1193−1199. (24) Mahns, B.; Kataeva, O.; Islamov, D.; Hampel, S.; Steckel, F.; Hess, C.; Knupfer, M.; Büchner, B.; Himcinschi, C.; Hahn, T.; Renger, R.; Kortus, J. Cryst. Growth Des. 2014, 14, 1338−1346. (25) Emge, T. J.; Maxfield, M.; Cowan, D. O.; Kistenmacher, T. J. Mol. Cryst. Liq. Cryst. 1981, 65, 161−178. (26) Metzger, R. M.; Heimer, N. E.; Gundel, D.; Sixl, H.; Harms, R. H.; Keller, H. J.; Nöthe, D.; Wehe, D. J. Chem. Phys. 1982, 77, 6203− 6214. (27) Mochida, T.; Yamazaki, S.; Suzuki, S.; Shimizu, S.; Mori, H. Bull. Chem. Soc. Jpn. 2003, 76, 2321−2328. (28) Mochida, T.; Akasaka, T.; Funasako, Y.; Nishio, Y.; Mori, H. Cryst. Growth Des. 2013, 13, 4460−4468. (29) Todorov, P. D.; Jenneskens, L. W.; van Lenthe, J. H. J. Chem. Phys. 2010, 132, 034504. (30) Bondi, A. J. Phys. Chem. 1964, 68, 441−451. (31) Custelcean, R.; Jackson, J. E. Chem. Rev. 2001, 101, 1963−1980. (32) Matta, C. F.; Hernández-Trujillo, J.; Tang, T.-H.; Bader, R. F. W. Chem. - Eur. J. 2003, 9, 1940−1951. (33) Akutagawa, T.; Koshinaka, H.; Ye, Q.; Noro, S.; Kawamata, J.; Yamaki, H.; Nakamura, T. Chem. - Asian J. 2010, 5, 520−529. (34) Ye, Q.; Akutagawa, T.; Hoshino, N.; Kikuchi, T.; Noro, S.; Xiong, R.-G.; Nakamura, T. Cryst. Growth Des. 2011, 11, 4175−4182. (35) Bloembergen, N.; Purcell, E. M.; Pound, R. V. Phys. Rev. 1948, 73, 679−712. (36) Yoshida, Y.; Maesato, M.; Kumagai, Y.; Mizuno, M.; Isomura, K.; Kishida, H.; Izumi, M.; Kubozono, Y.; Otsuka, A.; Yamochi, H.; Saito, G.; Kirakci, K.; Cordier, S.; Perrin, C. Eur. J. Inorg. Chem. 2014, 2014, 3871−3878. (37) Robertazzi, A.; Krull, F.; Knapp, E.-W.; Gamez, P. CrystEngComm 2011, 13, 3293−3300 and references therein. (38) Fyfe, C. A.; Dunell, B. A.; Rimeester, J. Can. J. Chem. 1971, 49, 3332−3335. (39) Jose, D.; Datta, A. J. Phys. Chem. Lett. 2010, 1, 1363−1366. (40) Wheland, R. C.; Martin, E. L. J. Org. Chem. 1975, 40, 3101− 3109. (41) Burla, M. C.; Caliandro, R.; Camalli, M.; Carrozzini, B.; Cascarano, G. L.; De Caro, L.; Giacovazzo, C.; Polidori, G.; Spagna, R. J. Appl. Crystallogr. 2005, 38, 381−388. (42) Sheldrick, G. M. SHELXL-97, Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. (43) Chen, M. M. L.; Hoffmann, R. J. Am. Chem. Soc. 1976, 98, 1647−1653. (44) Summerville, R. H.; Hoffmann, R. J. Am. Chem. Soc. 1976, 98, 7240−7254. (45) Mori, T.; Kobayashi, A.; Sasaki, Y.; Kobayashi, H.; Saito, G.; Inokuchi, H. Bull. Chem. Soc. Jpn. 1984, 57, 627−633. Overlap integrals calculated using the program are free from intermolecular interactions through the excited states such as Förster-type resonance energy or Dexter-type electron transfer processes.

F

DOI: 10.1021/acs.cgd.5b01138 Cryst. Growth Des. XXXX, XXX, XXX−XXX