Next Article in Journal
Structure-Properties Correlation of Cross-Linked Penicillin G Acylase Crystals
Next Article in Special Issue
Blue-Shifting Mechanofluorochromic Luminescent Behavior of Polymer Composite Films Using Gelable Mechanoresponsive Compound
Previous Article in Journal
Displacements in the Cationic Motif of Nonstoichiometric Fluorite Phases Ba1−xRxF2+x as a Result of the Formation of {Ba8[R6F68–69]} Clusters: III. Defect Cluster Structure of the Nonstoichiometric Phase Ba0.69La0.31F2.31 and Its Dependence on Heat Treatment
Previous Article in Special Issue
Electrorheological Effect of Gold Nanoparticles Coated with Fluorescent Mesogenic Groups Dispersed in Nematic Liquid Crystal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Fluoroalkyl-Substituent in Bistolane-Based Photoluminescent Liquid Crystals on Their Physical Behavior

1
Faculty of Molecular Chemistry and Engineering, Kyoto Institute of Technology, Matsugasaki, Sakyo-ku, Kyoto 606-8585, Japan
2
School of Chemistry, University of St. Andrews, North Haugh, St. Andrews KY16 9ST, UK
3
Department of Quantum Beam Science, Graduate School of Science and Engineering, Ibaraki University, 4-12-1 Naka-narusawa, Hitachi, Ibaraki 316-8511, Japan
4
Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(4), 450; https://doi.org/10.3390/cryst11040450
Submission received: 1 April 2021 / Revised: 14 April 2021 / Accepted: 17 April 2021 / Published: 20 April 2021

Abstract

:
Photoluminescent liquid crystals (PLLCs) have attracted significant attention owing to their broad applicability in thermosensing and PL switching. Extensive efforts have been made to develop bistolane-based PLLCs containing flexible units at both molecular terminals, and it has been revealed that their PL behavior can switch with the phase transition between the crystalline and LC phases. Although slight modulation of the flexible unit structure dramatically alters the LC and PL behaviors, few studies into the modification of the flexible units have been conducted. With the aim of achieving dynamic changes in their physical behaviors, we developed a family of bistolane derivatives containing a simple alkyl or a fluoroalkyl flexible chain and carried out a detailed systematic evaluation of their physical behaviors. Bistolanes containing a simple alkyl chain showed a nematic LC phase, whereas switching the flexible chain in the bistolane to a fluoroalkyl moiety significantly altered the LC phase to generate a smectic phase. The fluoroalkyl-containing bistolanes displayed a stronger deep blue PL than their corresponding non-fluorinated counterparts, even in the crystalline phase, which was attributed to the construction of rigid molecular aggregates through intermolecular F···H and F···F interactions to suppress non-radiative deactivation.

Graphical Abstract

1. Introduction

The fluorine atom possesses numerous unique characteristics [1,2,3,4], including the largest electronegativity of all elements (4.0 on the Pauling scale), the second smallest atomic size (van der Waals radius, rvdw = 147 pm) next to hydrogen (rvdw = 120 pm), and the dissociation energy of the C–F bond (105.3 kJ mol−1) is higher than those of the C–H (98.8 kJ mol−1) and C–C (83.1 kJ mol−1) bonds. Owing to these unique characteristics, the incorporation of fluorine atoms into organic molecules results in dramatic augmentations or alterations in their physical properties, in addition to introducing new functionalities [1,2,3,4]. Over the past few decades, significant attention has been paid to the development of fluorinated functional materials, worldwide. Consequently, several fluorinated materials have been developed for application in the pharmaceutical [5,6,7], agrochemical [8,9,10], and industrial fields, including organic dyes, liquid crystals, and photoluminescence molecules (Figure 1a) [11,12,13].
Previously, our group developed an efficient synthetic methodology for a wide variety of organofluorine compounds using readily available fluorinated substances [14,15] and carried out detailed explorations of fluorine-containing organic materials, such as negative dielectric liquid-crystalline (LC) molecules (A) [16] and efficient solid-state photoluminescence (PL) molecules (B) (Figure 1b) [17,18]. From the latter studies on fluorinated PL molecules, an interesting finding was that polyfluorinated bistolanes (C) exhibited both LC and PL properties [19,20,21,22,23] and were, thus, promising stimulus-responsive PL switching molecules [24,25]. As a result of intensive investigations into fluorinated PL molecules, it was revealed that fluorine atoms in these structures play an essential role in determining the electron density distribution and in the construction of rigid molecular aggregates via F···H hydrogen bonding interactions.
To understand the effect of the fluoroalkyl-type flexible unit on the LC and PL characteristics of bistolane-based photoluminescent LCs (PLLCs) (see Figure 1c), we designed a family of bistolane derivatives 1 bearing an alkyl chain (e.g., methyl or cyclohexyl) and bistolane derivatives 2 bearing a fluoroalkyl unit (e.g., trifluoromethyl, 2,3,5,6-tetrafluorocyclohexyl or undecafluorocyclohexyl). In this article, we discuss the results of our evaluation of the phase transitions and photophysical behaviors of these derivatives and describe in detail the effects of the fluoroalkyl substituents on the physical properties of the bistolane derivatives.

2. Materials and Methods

2.1. General

All chemicals were of reagent grade and were purified in the usual manner prior to use when necessary. Column chromatography was conducted using silica gel (FUJIFILM Wako Pure Chemical Corporation, Wako-gel® 60N, 38–100 µm; Osaka, Japan), while thin-layer chromatography (TLC) was performed on silica gel TLC plates (Merck, Silica gel 60F254; New Jersey, NJ, USA).
High-resolution mass spectra (HRMS) were recorded on a JMS700MS spectrometer (JEOL, Tokyo, Japan) using the fast atom bombardment (FAB) method. Infrared (IR) spectra were recorded using the KBr method with an FTIR-4100 type A spectrometer (JASCO, Tokyo, Japan). All the optical spectra are reported in terms of the wavenumber (cm−1). 1H NMR (400 MHz) and 13C NMR (100 MHz) spectra were obtained using an AVANCE III 400 NMR spectrometer (Bruker, Rheinstetten, Germany) in chloroform-d (CDCl3) solution, and the chemical shifts are reported in parts per million (ppm) based on the residual protons in the NMR solvent. 19F NMR (376 MHz) spectra were obtained using an AVANCE III 400 NMR spectrometer (Bruker, Rheinstetten, Germany) in CDCl3 solution with CFCl3 (δF = 0 ppm) as an internal standard.

2.2. Materials

Bistolanes 1a and 1b bearing an alkyl chain and 2a2c bearing a fluoroalkyl unit were synthesized via a Sonogashira cross-coupling reaction using 4-substituted halobenzene and 4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenylacetylene, both of which are readily accessible from the corresponding commercially-available chemicals (Figure 2).

2.3. Typical Synthetic Procedure for Preparation of the Bistolanes (as an Example, 2c)

In a two-necked round-bottomed flask, equipped with a magnetic stirrer bar, were placed Cl2Pd(PPh3)2 (Sigma-Aldrich; 22 mg, 0.031 mmol), PPh3 (Fujifilm-Wako Pure Chemicals; 10 mg, 0.038 mmol), 4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenylacetylene (76 mg, 0.33 mmol), 1-(undecafluorocyclohexyl)-4-iodobenzene (191 mg, 0.39 mmol), CuI (Fujifilm-Wako Pure Chemicals; 6.0 mg, 0.032 mmol), and Et3N (Fujifilm-Wako Pure Chemicals; 1.5 mL). The mixture was heated to 60 °C and continuously stirred for 19 h at the same temperature. After 19 h, the precipitate formed in the reaction mixture was removed by atmospheric filtration, the filtrate was added to an aqueous saturated NH4Cl solution (Fujifilm-Wako Pure Chemicals; 10 mL), and the organic component was extracted with ethyl acetate (EtOAc, 3 × 15 mL). The combined organic layer was then dried over anhydrous Na2SO4 and separated by filtration. The filtrate was evaporated under reduced pressure using a rotary evaporator and the residue was subjected to silica-gel column chromatography (eluent: hexane/EtOAc = 30/1), followed by recrystallization from CH2Cl2/MeOH (v/v = 1/1) to afford 1-undecafluorocyclohexyl-4-[2-[4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]benzene (2c) as colorless crystals in 40% yield (87 mg, 0.13 mmol).

2.3.1. 1-Undecafluorocyclohexyl-4-[2-[4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]benzene (2c)

Yield: 40%; m.p.: 163 °C determined by DSC; IR (KBr) νmax 2960, 2933, 2861, 2213, 1597, 1521, 1249, 1195, 1025, 968, 866 cm−1; 1H NMR (CDCl3) δ 0.91 (3H, t, J = 6.8 Hz), 1.31–1.39 (4H, m), 1.42–1.53 (2H, m), 1.79 (2H, quin., J = 6.4 Hz), 3.98 (2H, t, J = 6.4 Hz), 6.88 (2H, d, J = 8.8 Hz), 7.46 (2H, d, J = 8.8 Hz), 7.51 (3H, s), 7.67 (4H, s); 13C NMR (CDCl3) δ 14.0, 22.6, 25.7, 29.1, 29.7, 31.6, 68.1, 87.6, 89.3, 91.9, 104.2–114.0 (m, 2C), 114.6, 114.7, 121.8, 123.7, 123.9, 124.2, 126.8, 127.0 (t, J = 4.4 Hz), 127.2 (t, J = 3.6 Hz), 131.4, 131.6, 131.9, 133.1, 159.5; 19F NMR (CDCl3): δ −118.4 to −119.5 (dm, J = 295.5 Hz, 2F), −121.9 to −123.0 (dm, J = 275.5 Hz, 2F), −123.0 to −124.1 (dm, J = 286.1 Hz, 1F), −132.7 to −133.8 (dm, J = 298.5 Hz, 2F), −138.4 to −139.6 (dt, J = 282.0, 13.6 Hz, 2F), −141.6 to −142.7 (dm, J = 286.1 Hz, 1F), −181.0 to −181.4 (m, 1F); HRMS (FAB+) m/z 658.1720 (calcd for C34H25F11O, 658.1730).

2.3.2. 4-[2-[4-[2-(4-n-Hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]toluene (1a)

Yield: 33%; m.p.: 154 °C determined by DSC; IR (KBr) νmax 2952, 2921, 2848, 1604, 1519, 1466, 1284, 1242, 1180, 839, 813 cm1; 1H NMR (CDCl3) δ 0.91 (3H, t, J = 7.2 Hz), 1.31–1.39 (4H, m), 1.40–1.52 (2H, m), 1.75–1.83 (2H, m), 2.37 (3H, s), 3.97 (2H, t, J = 6.4 Hz), 6.87 (2H, d, J = 8.8 Hz), 7.16 (2H, d, J = 8.4 Hz), 7.42 (2H, d, J = 8.4 Hz), 7.45 (2H, d, J = 8.8 Hz), 7.475 (4H, s); 13C NMR (CDCl3) δ 14.0, 21.5, 22.6, 25.7, 29.2, 31.5, 68.1, 87.8, 88.6, 91.26, 91.34, 114.6, 114.9, 120.0, 122.9, 123.3, 129.1, 131.3, 131.4, 131.5, 133.0, 138.5, 159.4; HRMS (FAB+) m/z 392.2148 (calcd for C29H28O, 392.2140).

2.3.3. 1-Cyclohexyl-4-[2-[4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]benzene (1b)

Yield: 53%; m.p.: 164 °C determined by DSC; IR (KBr) νmax 2926, 2852, 2215, 1604, 1596, 1518, 1246, 1103, 827 cm−1; 1H NMR (CDCl3) δ 0.91 (3H, t, J = 7.2 Hz), 1.21–1.30 (1H, m), 1.30–1.52 (10H, m)‚ 1.70–1.92 (8H, m), 3.97 (2H, t, J = 6.4 Hz), 6.87 (2H, d, J = 8.8 Hz), 7.19 (2H, d, J = 8.4 Hz), 7.449 (2H, d, J = 8.4 Hz), 7.454 (2H, d, J = 8.8 Hz), 7.47 (4H, s); 13C NMR (CDCl3) δ 14.0, 22.6, 25.7, 26.1, 26.8, 29.1, 31.5, 34.2, 44.5, 68.1, 87.8, 88.5, 91.3, 114.5, 114.8, 120.3, 122.9, 123.2, 126.9, 129.6, 131.3, 131.4, 131.6, 133.0, 148.6, 159.3; HRMS (FAB+) m/z 460.2760 (calcd for C34H36O, 460.2766).

2.3.4. 4-[2-[4-[2-(4-n-Hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]benzotrifluoride (2a)

Yield: 43%; m.p.: 130 °C determined by DSC; IR (KBr) νmax 2962, 2937, 2859, 1607, 1594, 1510, 1332, 1286, 1251, 1127, 1067, 1024, 837 cm−1; 1H NMR (CDCl3) δ 0.91 (3H, t, J = 6.8 Hz), 1.32–1.38 (4H, m), 1.40–1.52 (2H, m), 1.73–1.86 (2H, m), 3.98 (2H, t, J = 6.8 Hz), 6.88 (2H, d, J = 8.8 Hz), 7.46 (2H, d, J = 8.8 Hz), 7.50 (4H, s), 7.62 (4H, d, J = 1.6 Hz); 13C NMR (CDCl3) δ 14.0, 22.6, 25.7, 29.2, 31.6, 68.1, 87.6, 89.5, 91.5, 91.9, 114.6, 114.7, 121.2 (q, J = 271.3 Hz), 121.8, 124.2, 125.3 (q, J = 3.7 Hz), 126.9, 130.0 (q, J = 33.0 Hz), 131.4, 131.6, 131.8, 133.1, 159.5; 19F NMR (CDCl3): δ −63.30 (s, 3F); HRMS (FAB+) m/z 446.1861 (calcd for C29H25F3O, 446.1858).

2.3.5. 1-(2,3,5,6-Tetrafluorocyclohexyl)-4-[2-[4-[2-(4-n-hexyloxyphenyl)ethyn-1-yl]phenyl]ethyn-1-yl]benzene (2b)

Yield: 64%; m.p.: 175 °C determined by DSC; IR (KBr) νmax 2934, 2868, 2213, 1604, 1519, 1247, 1107, 1016, 973, 837 cm−1; 1H NMR (CDCl3) δ 0.91 (3H, t, J = 6.8 Hz), 1.31–1.39 (4H, m), 1.41–1.52 (2H, m), 1.79 (2H, quin., J = 6.8 Hz), 2.43–2.56 (1H, m), 2.62 (1H, t, J = 36.4 Hz), 2.70–2.86 (1H, m), 3.97 (2H, t, J = 6.4 Hz), 4.51–4.77 (2H, m), 4.87–5.18 (2H, m), 6.87 (2H, d, J = 8.8 Hz), 7.46 (2H, d, J = 8.4 Hz), 7.491 (6H, brs), 7.54 (2H, d, J = 8.4 Hz); 13C NMR (CDCl3) δ 14.0, 22.6, 25.7, 29.1, 31.6, 43.6–44.2 (m, 1C), 68.1, 85.6–90.0 (2C), 87.7, 89.9, 90.4, 114.6, 114.8, 122.4, 123.1, 123.7, 129.3, 131.3, 132.0, 133.1, 135.7, 159.4, three carbons cannot be detected due to low solubility in NMR solvent; 19F NMR (CDCl3): δ −190.6 to −190.9 (m, 2F), −210.1 to −210.6 (m, 2F); HRMS (FAB+) m/z 532.2402 (calcd for C34H32F4O, 532.2389).

2.4. X-ray Crystallographic Analysis

Single crystals of 2a and 2c were obtained by recrystallization from a mixed solvent system of CH2Cl2/MeOH (v/v = 1/1). Each single crystal was mounted on a glass fiber and X-ray diffraction patterns were recorded on a XtaLabMini diffractometer equipped with a VariMax Mo optical system (λ = 0.71073 Å) and a Pilatus P200 detector (Rigaku, Tokyo, Japan). The reflection data were processed using the CryslAlisPro (ver. 1.171.38.46; Rigaku Oxford Diffraction, 2015). The structures were solved by a directed method (SHELXT-2014/5) and SHELXL-2014/7 programs [26]. The crystallographic data were deposited in the Cambridge Crystallographic Data Centre (CCDC) database (CCDC 2,065,068 for 2a and 2,065,069 for 2c). These data can be obtained free of charge from the CCDC via www.ccdc.cam.ac.uk/data_request/cif (accessed on 19 April 2021).

2.5. Powder X-ray Diffraction (PXRD) Measurements

The liquid crystal structures were evaluated using an FR-E X-ray diffractometer attached to an R-axis IV two-dimensional (2D) detector (Rigaku, Tokyo, Japan). For the purpose of these measurements, 0.3-mm collimated Cu Kα radiation (λ = 1.54187 Å) was used as the X-ray beam, and the camera length was set to 300 mm. The powder sample was loaded into a thin-walled glass capillary tube for XRD analysis (ϕ 1.0–2.5 mm, Hilgenberg GmbH), and the sample was annealed up to an isotropic temperature under vacuum. The glass capillary was placed on a ceramic heater attached to the FR-E sample holder. The exposure time of the X-ray beam was 5–60 min.

2.6. Computations

All computations were performed using the Gaussian 16 software package (Revision B.01) [27]. Geometry optimizations were carried out using the hybrid meta-GGA functional (M06-2X) [28,29] and the 6-31+G(d) basis set with an implicit solvation model (conductor-like polarizable continuum model; CPCM [30,31,32]) for CH2Cl2. The vertical excitation energies and dipole moments of the optimized structures were calculated using the time-dependent self-consistent field approximation [33,34] at the same level of theory.

2.7. Thermal Measurements

The phase transition behavior was observed by polarizing optical microscopy using a BX53 microscope (Olympus Corporation, Tokyo, Japan), equipped with a heating and cooling stage (Linkam Scientific Instruments, 10,002 L, Surrey, UK). The thermodynamic behavior was determined using differential scanning calorimetry (DSC, SHIMADZU DSC-60 Plus, Kyoto, Japan) at heating and cooling rates of 5.0 °C min−1 under a N2 atmosphere.

2.8. Photophysical Measurements

The UV–Vis absorption spectra were recorded using a V-500 absorption spectrometer (JASCO, Tokyo, Japan). The PL spectra of the solution and crystal forms were acquired using an FP-6600 fluorescence spectrometer (JASCO, Tokyo, Japan). The absolute quantum yields in both the solution and crystalline phases were measured using the Quantaurus-QY measurement system C11347-01 (Hamamatsu Photonics, Hamamatsu, Japan).

3. Results and Discussion

3.1. Theoretical Assessment

Our study was initiated with a theoretical assessment based on quantum chemical calculations using the Gaussian 16 software to discuss the effects of the fluoroalkyl units on the physical properties of the various compounds. Table 1 lists the molecular dipole moments (i.e., μ|| along the major molecular axis and μ along the minor molecular axis), the highest occupied molecular orbital (HOMO) and lowest unoccupied MO (LUMO) energies, the theoretical transition probability and distribution, and the theoretical absorption wavelength (λcalcd) for each compound. Figure 3 illustrates the MOs alongside the corresponding orbital energies.
Bistolanes 1a and 1b, bearing an alkyl chain, were found to possess small dipole moments for both the major and minor molecular axes, µ|| and µ, respectively, in both the major and minor molecular axes in the ground state. Comparing the µ|| and µ values for 1a and 1b, the µ|| values for perfluoroalkyl-containing bistolanes 2a and 2c increased approximately six-fold along the major molecular axis. Interestingly, in the case of bistolane 2b, bearing an all-cis-2,3,5,6-tetrafluorocyclohexyl unit, ~2.5-fold higher values of both µ|| and μ were obtained because of the 1,3-diaxial arrangement of the C–F bonds stemming from the all-cis configuration [35,36]. The dramatic changes in the dipole moments may, therefore, induce a significant alteration in the molecular aggregates, associated with changes in the intermolecular interactions; a significant change in the phase transition behavior during the crystal ⇄ LC and LC ⇄ isotropic liquid phase transformations would also be expected.
Based on our calculations, 1a and 1b, both of which bear alkyl chains, exhibited broad HOMOs and LUMOs that covered the entire π-conjugated structure. In contrast, the partially or fully fluoroalkyl-containing bistolanes 2a2c showed a molecular orbital distribution wherein the orbital lobes were primarily localized over the electron-rich aromatic ring attached to the n-hexyloxy chain in the HOMO, while in the LUMO, the orbital lobes were primarily localized over the electron-deficient aromatic ring bearing a fluoroalkyl unit. 1a and 1b exhibited high energy levels of −6.88 eV and −6.87 eV for the HOMO, respectively, and −1.27 eV for both LUMOs of 1a and 1bE = 5.60–5.61 eV). Meanwhile, 2a and 2c with the perfluoroalkyl substituent exhibited reduced HOMO and LUMO energy levels, i.e., −7.00 eV and −1.52 eV, respectively, together with a smaller ΔE value (5.48 eV). Notably, 2b, with a partially fluorinated alkyl fragment, had intermediate HOMO and LUMO energy levels between 1b and 2c: −6.93 eV for the HOMO and −1.35 eV for the LUMO. Consequently, it can be anticipated that switching the flexible unit from an alkyl chain to the corresponding fluoroalkyl unit could also result in significant changes in the photophysical behavior, such as the ultraviolet-visible (UV–Vis) absorption and the PL behavior.

3.2. Synthesis and Crystal Structure Determination

Based on the theoretical assessments, bistolanes 1 and 2 were synthesized via a Pd(0)-catalyzed Sonogashira cross-coupling reaction of 4-[2-(4-n-hexyloxyphneyl)ethyn-1-yl]phenylacetylene with the appropriate iodobenzene or bromobenzene bearing an alkyl or fluoroalkyl substituent at the para-position (Figure 2). All compounds were successfully obtained in 33–64% isolated yields after purification by column chromatography and subsequent recrystallization. Identification of the crystalline compounds was achieved by spectroscopic analyses, and the obtained products were found to be sufficiently pure to allow evaluation of their phase transitions and PL behaviors.
Among the crystalline compounds obtained, trifluoromethyl (CF3)-containing 2a and undecafluorocyclohexyl (cyclo-C6F11)-substituted 2c furnished single crystals suitable for X-ray crystallographic analysis. Figure 4 shows the molecular structures and packing structures of these two crystals.
As indicated, the CF3-containing 2a formed colorless rod-shaped single crystals of a monoclinic crystal system in the Cc space group, and the unit cell contained four molecular units. For 2a, a rigid packing structure was constructed that included several intermolecular interactions (Figure 4a, left). Considering the short interatomic distances, as shown in Figure 4a (right), C–F···H hydrogen bonding (259.5 pm), O···H hydrogen bonding (268.4 pm), C–F···π interactions (316.4 pm) [37,38,39], C–H···π interactions (282.4 and 284.0 pm), and van der Waals interactions (Calkyl–H···H–Calkyl = 235.3 pm and Caryl–H···H–Calkyl = 237.7 pm) all appeared to be present; in each of these cases, the interatomic distances were less than the sum of rvdW, for the two atoms (i.e., C: 170 pm, H: 120 pm, and F: 149 pm) [40]. In contrast, the cyclo-C6F11-substituted 2c formed colorless block-shaped crystals of a monoclinic crystal system in the P21 space group, and four molecular units of 2c were contained in a unit cell (Figure 4b, left). As shown in Figure 4b (right), rigid packing structures were formed through several intermolecular interactions, including C–F···H hydrogen bonding (242.8 and 254.7 pm), C–F···F interactions (289.4, 291.1, 293.9, and 294.0 pm) [41,42], Caryl–H···π interactions (282.2, 283.0, and 289.4 pm), Calkyl–H···π interactions (283.8 pm), and van der Waals interactions (Caryl–H···H—Calkyl = 238.8 pm). It is notable that 2a produced layered structures with a head-to-tail molecular alignment through C–F···H hydrogen bonding interactions. Although 2c also formed layered structures with a head-to-head molecular alignment, in this case, these were attributed to C–F···F interactions resulting from the fluorophilic effect of the highly fluorinated alkyl moieties [42,43].

3.3. Phase Transition Behaviors

The phase transition behaviors of crystalline bistolane derivatives 1a and 1b and 2a2c were then evaluated using polarizing optical microscopy (POM), differential scanning calorimetry (DSC), and variable temperature powder X-ray diffraction (VT-PXRD). Figure 5 shows the phase sequences and POM images observed for the LC phase during the 2nd heating and cooling processes.
During the POM observations, a bright-viewing field with good fluidity was found for all the bistolanes (i.e., 1a, 1b, and 2a2c) over both the heating and cooling processes. The POM images of 1a and 1b in the LC phase display the four-brush Schlieren texture, which is a typical image of a nematic (N) LC phase. In addition, the VT-PXRD measurements of 1a and 1b at 180 °C revealed a hollow pattern at ~20° (2θ) with no sharp diffraction signals. Based on the POM and VT-PXRD measurements, the observed LC phases for 1a and 1b were identified as the nematic (N) phase with an orientational order. In contrast to the alkyl-substituted 1a and 1b, CF3-substituted 2a and cyclo-C6F11-substituted 2c displayed distinct POM textures, namely fan-shaped arrangements, which are typical textures for smectic (Sm) phases with both orientational order and positional order in the LC phase. In the case of 2a, which bears a CF3-unit, VT-PXRD measurements at 150° showed three sharp peaks, identified as the (110), (200), and (210) mirror indexes, in the wide-angle region. This indicates that 2a forms an in-plane herringbone structure (Figure S23, Supporting Information). Consequently, the LC phase observed between 130 and 202 °C was identified as a smectic E (SmE) phase [44]. Furthermore, the PXRD pattern at 210 °C showed only one sharp diffraction peak, assigned to the (110) plane, which corresponds to a hexagonal structure; the LC phase observed in the range of 202–214 °C was identified as the smectic B (SmB) phase. Additionally, the LC phase displayed between 214 and 236 °C was identified as the smectic A (SmA) phase, which produced a single sharp diffraction peak in the small-angle region. In contrast, for 2c, which bears a cyclo-C6H11 unit, a single sharp signal was detected in the small-angle region at 180 °C, which indicates that 2c displayed only the SmA phase in the LC form (Figure S25 in SI). Although 2b, which bears an all-cis-2,3,5,6-tetrafluorocyclohexyl (cis-cyclo-C6H7F4) unit, contained four fluorine atoms in its flexible unit, only the N phase was observed as an LC phase, analogous to the case of 1, which bears an alkyl chain. Based on these observations, the dipole moment µ|| along the major molecular axis was considered to be important for controlling the LC phase. More specifically, the incorporation of fluorine atoms to increase µ|| induces the formation of a higher-order LC phase with both orientational and positional order, similar to the Sm phase, whereas bistolanes with smaller µ|| values form an LC phase with only an orientational order, such as the N phase.

3.4. Photophysical Behavior

Subsequently, we focused on the photophysical behaviors of bistolanes 1a and 1b, which bear an alkyl chain, and 2a2c, which bear a fluoroalkyl unit. Figure 6a shows the UV–Vis absorption (1.5 × 10−5 mol L−1) and PL spectra (1.0 × 10−6 mol L−1) in CH2Cl2, and the obtained photophysical data are summarized in Table 2.
For the UV–Vis absorption measurements carried out in CH2Cl2, a single UV–Vis absorption band (λabs) was observed for all bistolanes with a maximum absorption wavelength at ~328–332 nm, together with a shoulder peak at ~348–350 nm (Figure 6a, Table 2). Based on the theoretical assessment (λcalcd = 330–334 nm, Table 1), for all compounds, the major absorption band was concluded to originate from the HOMO→LUMO transition. The slight red-shift of the λabs values of 2a2c, compared to those of 1a and 1b, were likely caused by the narrower HOMO–LUMO energy gap (ΔE) of the former, which is induced by the electron-withdrawing character of the fluoroalkyl unit.
Upon irradiation of the CH2Cl2 solutions of the bistolanes with UV light of an energy equal to λabs, all compounds were found to exhibit strong PL (ΦPL = 1.00), with a single PL band of maximum PL wavelength (λPL) at ~377–405 nm (Figure 6a, Table 2). In this case, the incorporation of fluorine atoms into the flexible unit caused a gradual red-shift in λPL by 20–23 nm for 2a and 2c, with respect to the values for 1a and 1b, although 2b, which contained four fluorine atoms, did not exhibit any such shift in λPL. Using theoretical assessments, it was concluded that the red-shift in λPL is associated with the orbital gap (ΔE) between the HOMO and LUMO; full substitution of the flexible unit with fluorine atoms, as in the cases of 2a and 2c, caused a 0.12 eV decrease in ΔE compared to that of their non-fluorinated counterparts, i.e., 1a and 1b, resulting in a slight red-shift in λPL. However, the partial incorporation of fluorine atoms into the flexible unit, as in the case of 2b, had no significant impact, and this was attributed to the less effective electron withdrawing capability of the flexible unit along the major molecular axis. As shown by the Commission Internationale de l’Eclailage (CIE) color diagram in Figure 6b, all the bistolane analogues produced a deep blue PL color in CH2Cl2.
To gain additional insight into the effect of fluorine atoms on the absorption and PL behaviors, the influence of the solvent was examined for CH3-substituted 1a and CF3-substituted 2a as a representative comparison (Figure 6c). It was found that the absorption behaviors of 1a and 2a were consistent, without showing any dependence on the solvent polarity, i.e., λabs = 324–328 nm for 1a and 325–333 nm for 2a (Figures S28 and S29 in SI). By contrast, the PL behavior was found to change upon varying the solvent polarity. More specifically, a solution of 1a in the less polar hexane (ET(30) = 31.0) [45] exhibited sharp PL signals at ~353 nm, along with a vibrational structure at ~368 nm. Upon increasing the solvent polarity, the PL band was gradually shifted to the long-wavelength region, reaching up to 388 nm when DMF (ET(30) = 43.2) or MeCN (ET(30) = 45.6) was employed as the solvent. CF3-substituted 2a also exhibited a sharp band at ~361 nm in hexane, together with vibrational structure at ~377 nm. In the case of 2a, the λPL reached up to 420 nm when the polar solvents DMF (ET(30) = 45.6) or MeCN (ET(30) = 45.6) were used. Upon comparison of the PL behaviors of 1a and 2a in various solvents, a longer-wavelength shift in the PL was observed in the case of CF3-substituted 2a because 2a is likely to be susceptible to stabilization via solvation in polar solvents due to its large dipole moment, µ||. To elucidate the correlation between solvent polarity and PL behavior, Lippert–Mataga plots were created with a solvent polarity parameter (Δf) on the horizontal axis and the Stokes’ shift (Δν = νabs − νPL) on the vertical axis (Figure 6d) [46,47]. Linear relationships between Δν and Δf were found for 1a and 2a, and these are expressed in Equations (1) and (2), respectively:
Δν = 8404.5 Δf + 2342,
Δν = 16,559 Δf + 1859,
It is known that the slope of these equations is strongly correlated with the difference in the dipole moments between the excited and ground states [44]: Δμ = µeµg, where μe and μg represent the dipole moments in the excited and ground states, respectively. As a result, the value of Δμ for 2a was larger than that for 1a (i.e., 7.44 D for 2a and 6.33 D for 1a), and this was attributed to the fact that compared to 1a, the CF3-substituted 2a is likely to exhibit a larger polarization in the excited state. Therefore, the incorporation of a fluoroalkyl substituent resulted in a strong intermolecular charge transfer transition-based PL due to the strong electron-withdrawing character of the fluoroalkyl unit.
It is of significant interest that bistolanes 1a and 1b (bearing an alkyl chain) and 2a2c (bearing a fluoroalkyl unit) exhibited PL even in the crystalline state, despite the fact that the majority of π-conjugated luminophores have been reported to undergo immediate quenching when molecular aggregates are formed [48,49]. Figure 7 shows the PL spectra and photographic images depicting the PL behaviors of 1a, 1b, and 2a2c in the crystalline state, along with the corresponding CIE color diagram. The corresponding photophysical data are listed in Table 3.
As indicated in Figure 7a, compounds 1a and 1b exhibited two PL bands with λPL values of ~404 and 422 nm, while bands at ~418, 417, and 421 nm, were observed for 2a2c, respectively, in addition to shoulder peaks at 434 nm for 2a and 2c, and at 401 nm for 2b. In comparison with the results obtained for the alkyl-containing 1a and 1b, the λPL values for 2a2c were slightly red-shifted by 10–15 nm. In addition, λPL was significantly red-shifted by ~30 nm, in comparison with the λPL obtained in the solution states, which is attributable to the stabilization of the excited states via intermolecular interactions in the aggregated states.
It is noteworthy that bistolanes 1a and 1b exhibited high PL efficiencies (ΦPL) of 0.57 and 0.79, respectively, even in the crystalline state, although the values of ΦPL in the aggregated states were drastically reduced compared with those in the solution states, due to the acceleration of non-radiative deactivation via intermolecular interactions [48,49]. As mentioned above, 2b, which contains a partially fluorinated alkyl unit, afforded a similar result for the PL measurements, even in the crystalline state, whereas bistolanes 2a and 2c, which bear fully fluorinated alkyl units, exhibited enhanced ΦPL values. As discussed above, 2a and 2c formed rigid molecular aggregates through multiple intermolecular interactions, which effectively weakened the molecular motions (e.g., vibration and rotation). This, in turn, suppressed non-radiative deactivation, and produced a significant enhancement in ΦPL. In the other compounds, it is anticipated that non-radiative deactivation cannot be sufficiently suppressed because there are fewer intermolecular interactions, although the aggregated structures of non-fluorinated 1a and 1b and partially fluorinated 2b remain unclear. Accordingly, the incorporation of a fluoroalkyl unit into the bistolane scaffold can achieve rigid packing in the molecular aggregates, leading to a significant enhancement in the ΦPL, even in the crystalline state.
Notably, the bistolane derivatives bearing a non-fluorinated or fluorinated alkyl chain were also found to be emissive in their LC phase that was formed after a thermal transition from the crystalline to the LC phase. The PL intensity, however, was gradually retarded because of the increasing non-radiative deactivation via micro-Brownian motion under the thermal conditions. The photoluminescent liquid crystal (PLLC) characteristics could be significantly improved if alignment of the LC on a film or substance could be controlled using a photoalignment technique, such as a mechanical rubbing process [50] or a rubbing-free process using photo-cross-linkable polymers [51,52] and photoreactive polyimide polymers [53,54].

4. Conclusions

To elucidate the effects of fluorine atoms in bistolane-based photoluminescent liquid crystals (PLLCs) on their liquid-crystalline (LC) and photoluminescent (PL) behaviors, bistolanes bearing an alkyl chain or a fluoroalkyl unit were synthesized, and their physical behaviors were systematically evaluated. All bistolanes exhibited LC behavior; the non-fluorinated alkyl-containing bistolanes mainly displayed a nematic LC phase, whereas their fluoroalkyl-containing counterparts exhibited smectic LC phases that possessed both an orientational order and a positional order. The large difference in the LC behavior can be attributed to two factors, namely the dipole moment along a major molecular axis, which is induced by the electron-withdrawing fluoroalkyl units, and the construction of head-to-tail or head-to-head ordered structures in molecular aggregates through F···H or F···F intermolecular interactions. In the photophysical measurements, interestingly, all bistolanes exhibited an intense PL behavior both in the dilute solution and in the crystalline phases. The incorporation of fluorine atoms into the flexible unit reduced not only the HOMO and LUMO energy levels, but also the energy gap between the two, thereby generating a slight red-shift of the maximum PL wavelength. Although the PL efficiency in the crystalline state normally decreases via non-radiative deactivation induced by intermolecular energy transfer, it was notable that the bistolanes containing fluoroalkyl units exhibited enhanced PL efficiencies due to the effective suppression of non-radiative deactivation. This was attributed to the formation of rigid molecular packing structures through multiple intermolecular interactions involving F···H and F···F interactions, among others. Consequently, the incorporation of fluorine atoms into the flexible units of the bistolane-based PLLCs likely contributed to the generation of a higher-ordered smectic LC phase in addition to blue PL in both the solution and aggregated states, with a high PL efficiency. These results will pave the way for the development of high-performance functional materials by incorporating fluorine atoms into organic molecules. Further fluorine-containing PLLCs will be reported by our research group in the near future.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11040450/s1, Figures S1–S13: NMR spectra, Figures S14 and S15: Crystal structure, Figures S16–S20: Optimized structures and their HOMO and LUMO distributions, Figures S21–25: DSC thermograms, Figure S26: POM image, Figures S27–S29: UV–Vis and PL spectra, Figure S30: Lippert–Mataga plots, Figure S31: Excitation and PL spectra, Table S1: Crystallographic data, Tables S2–S11: Cartesian coordinates for calculated structure, Tables S12–S16: phase transition behavior, Tables S17 and S18: photophysical data in a different solvent.

Author Contributions

Conceptualization, S.Y.; methodology, S.Y.; validation, S.Y. and M.M.; investigation, S.Y., M.M., Y.W., T.K. (Tsutomu Konno), Q.Z., D.O., M.N., T.A., H.F., T.K. (Toshio Kubota), and M.H.; resources, S.Y., M.M., Y.W., Q.Z., D.O.; data curation, S.Y., M.M., Y.W., T.K. (Tsutomu Konno); writing—original draft preparation, S.Y.; writing—review and editing, S.Y., M.M., Y.W., T.K. (Tsutomu Konno), Q.Z., D.O., M.N., T.A., H.F., T.K. (Toshio Kubota), and M.H.; visualization, S.Y.; supervision, S.Y.; project administration, S.Y.; funding acquisition, S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI through a Grant-in-Aid for Scientific Research (C), grant number JP18K05262.

Data Availability Statement

Data is contained within the article and supplementary material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Caron, S. Where does the fluorine come from? A review on the challenges associated with the synthesis of organofluorine compounds. Org. Process Res. Dev. 2020, 24, 470–480. [Google Scholar] [CrossRef]
  2. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of organofluorine compounds to pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef] [PubMed]
  3. Ogawa, Y.; Tokunaga, E.; Kobayashi, O.; Hirai, K.; Shibata, N. Current contributions of organofluorine compounds to the agrochemical industry. iScience 2020, 23, 101467. [Google Scholar] [CrossRef] [PubMed]
  4. Kirsch, P. Modern Fluoroorganic Chemistry: Synthesis, Reactivity, Applications, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2013; pp. 7–21. [Google Scholar]
  5. Behrang, N.; Fischbach, F.; Kipp, M. Mechanism of siponimod: Anti-inflammatory and neuroprotective mode of action. Cells 2019, 8, 24. [Google Scholar] [CrossRef] [Green Version]
  6. Mei, H.; Han, J.; Fustero, S.; Medio-Simon, M.; Sedgwick, D.M.; Santi, C.; Ruzziconi, R.; Soloshonok, V.A. Fluorine-containing drugs approved by the FDA in 2018. Chem. Eur. J. 2019, 25, 11797–11819. [Google Scholar] [CrossRef]
  7. Böhm, H.-J.; Banner, D.; Bendels, S.; Kansy, M.; Kuhn, B.; Müller, K.; Obst-Sander, U.; Stahl, M. Fluorine in medicinal chemistry. ChemBioChem 2004, 5, 637–643. [Google Scholar] [CrossRef]
  8. Lowe, P.T.; O’Hagan, D. A role for fluorine in flavours, fragrances and pheromones. J. Fluor. Chem. 2020, 230, 109420. [Google Scholar] [CrossRef]
  9. Li, H.; Gao, M.; Chen, Y.; Wang, Y.; Zhu, X.; Yang, G. Discovery of pyrazine-carboxamide-diphenyl-ethers as novel succinate dehydrogenase inhibitors via fragment recombination. J. Agric. Food Chem. 2020, 68, 14001–14008. [Google Scholar] [CrossRef]
  10. Giornal, F.; Pazenok, S.; Rodefeld, L.; Lui, N.; Vors, J.-P.; Leroux, F.R. Synthesis of diversely fluorinated pyrazoles as novel active agrochemical ingredients. J. Fluor. Chem. 2013, 152, 2–11. [Google Scholar] [CrossRef]
  11. Fan, Q.; Méndez-Romero, U.; Guo, X.; Wang, E.; Zhang, M.; Li, Y. Fluorinated photovoltaic materials for high-performance organic solar cells. Chem. Asian J. 2019, 14, 3085–3095. [Google Scholar] [CrossRef]
  12. Funabiki, K.; Saito, Y.; Kikuchi, T.; Yagi, K.; Kubota, Y.; Inuzuka, T.; Miwa, Y.; Yoshida, M.; Sakurada, O.; Kutsumizu, S.J. Aromatic Fluorine-Induced One-Pot Synthesis of Ring-Perfluorinated Trimethine Cyanine Dye and Its Remarkable Fluorescence Properties. J. Org. Chem. 2019, 84, 4372–4380. [Google Scholar] [CrossRef]
  13. Hird, M. Fluorinated liquid crystals—Properties and applications. Chem. Soc. Rev. 2007, 36, 2070–2095. [Google Scholar] [CrossRef]
  14. Konno, T. Trifluoromethylated internal alkynes: Versatile building blocks for the preparation of various fluorine-containing molecules. Synlett 2014, 25, 1350–1370. [Google Scholar] [CrossRef]
  15. Konno, T. Trifluoromethyl (CF3) group insertion methods in stereoselective synthesis. In Stereoselective Synthesis of Drugs and Natural Products; Andrunshko, V., Andrunshko, N., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2013; pp. 769–806. [Google Scholar]
  16. Yamada, S.; Konno, T. New types of negative dielectric anisotropy liquid crystals: Synthesis of CF2CF2-carbocyclic mesogens and an evaluation of their liquid crystal characteristics. In Chemical Elements (Fluorine, Rhodium and Rubidium): Properties, Synthesis and Applications; Huff, A., Ed.; Nova Science Publishers: New York, NY, USA, 2018; pp. 53–92. [Google Scholar]
  17. Yamada, S.; Mitsuda, A.; Miyano, K.; Tanaka, T.; Morita, M.; Agou, T.; Kubota, T.; Konno, T. Development of novel solid-state light-emitting materials based on pentafluorinated tolane fluorophores. ACS Omega 2018, 3, 9105–9113. [Google Scholar] [CrossRef]
  18. Morita, M.; Yamada, S.; Konno, T. Fluorine-induced emission enhancement of tolanes via formation of tight molecular aggregates. New J. Chem. 2020, 44, 6704–6708. [Google Scholar] [CrossRef]
  19. Yamada, S.; Mitsuda, A.; Adachi, K.; Hara, M.; Konno, T. Development of light-emitting liquid-crystalline polymers with a pentafluorinated bistolane-based luminophore. New J. Chem. 2020, 44, 5684–5691. [Google Scholar] [CrossRef]
  20. Morita, M.; Yamada, S.; Agou, T.; Kubota, T.; Konno, T. Luminescence tuning of fluorinated bistolanes via electronic or aggregated-structure control. Appl. Sci. 2019, 9, 1905. [Google Scholar] [CrossRef] [Green Version]
  21. Yamada, S.; Morita, M.; Agou, T.; Kubota, T.; Ichikawa, T.; Konno, T. Thermoresponsive luminescence properties of polyfluorinated bistolane-type light-emitting liquid crystals. Org. Biomol. Chem. 2018, 16, 5609–5617. [Google Scholar] [CrossRef]
  22. Yamada, S.; Morita, M.; Konno, T. Multi-color photoluminescence induced by electron-density distribution of fluorinated bistolane derivatives. J. Fluorine Chem. 2017, 202, 54–64. [Google Scholar] [CrossRef]
  23. Yamada, S.; Miyano, K.; Konno, T.; Agou, T.; Kubota, T.; Hosokai, T. Fluorine-containing bistolanes as light-emitting liquid crystalline molecules. Org. Biomol. Chem. 2017, 15, 5949–5958. [Google Scholar] [CrossRef]
  24. Nishikawa, M.; Kume, S.; Nishihara, H. Stimuli-responsive pyrimidine ring rotation in copper complexes for switching their physical properties. Phys. Chem. Chem. Phys. 2013, 15, 10549–10565. [Google Scholar] [CrossRef]
  25. Kobayashi, A.; Kato, M. Stimuli-responsive luminescent copper(I) complexes for intelligent emissive devices. Chem. Lett. 2017, 46, 154–162. [Google Scholar] [CrossRef]
  26. Sheldrick, G.M. Crystal structure refinement with SHELEX. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  27. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  28. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  29. Chan, B.; Gilbert, A.T.B.; Gill, P.M.W.; Radom, L. Performance of density functional theory procedures for the calculation of proton-exchange barriers: Unusual behavior of M06-type functionals. J. Chem. Theory Comput. 2014, 10, 3777–3783. [Google Scholar] [CrossRef]
  30. Andzelm, J.; Kölmel, C.; Klamt, A. Incorporation of solvent effects into density functional calculations of molecular energies and geometries. J. Chem. Phys. 1995, 103, 9312–9320. [Google Scholar] [CrossRef]
  31. Barone, V.; Cossi, M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  32. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef]
  33. Grad, J.; Yan, Y.J.; Mukamel, S. Time-dependent self-consistent field approximation for semiclassical dynamics using Gaussian wavepackets in phase space. Chem. Phys. Lett. 1987, 134, 291–295. [Google Scholar] [CrossRef]
  34. Makri, N. Time-dependent self-consistent field approximation with explicit two-body correlations. Chem. Phys. Lett. 1990, 169, 541–548. [Google Scholar] [CrossRef]
  35. Durie, A.J.; Slawin, A.M.Z.; Lebl, T.; Kirsch, P.; O’Hagan, D. Fluorocyclohexenes: Synthesis and structure of all-syn-1,2,4,5-tetrafluorocyclohexane. Chem. Commun. 2012, 48, 9643–9645. [Google Scholar] [CrossRef] [PubMed]
  36. Durie, A.J.; Fujiwara, T.; Cormanich, R.; Bühl, M.; Slawin, A.M.Z.; O’Hagan, D. Synthesis and elaboration of all-cis-1,2,4,5-tetrafluoro-3-phenylcyclohexane: A polar cyclohexane motif. Chem. Eur. J. 2014, 20, 6259–6263. [Google Scholar] [CrossRef] [PubMed]
  37. Kawahara, S.; Tsutzuki, S.; Uchimaru, T. Theoretical study of the C–F/π interaction: Attractive interaction between fluorinated alkane and an electron-deficient π-system. J. Phys. Chem. A 2004, 108, 6744–6749. [Google Scholar] [CrossRef]
  38. Li, P.; Maier, J.M.; Vik, E.C.; Yehl, C.J.; Dial, B.E.; Rickher, A.E.; Smith, M.D.; Pellechia, P.J.; Shimizu, K.D. Stabilizing fluorine–π interactions. Angew. Chem. Int. Ed. 2017, 56, 7209–7212. [Google Scholar] [CrossRef]
  39. Mooibroek, T.J.; Gamez, P.; Reedijk, J. Lone pair–π interactions: A new supramolecular bond? CrystEngComm 2008, 10, 1501–1515. [Google Scholar] [CrossRef]
  40. Bondi, A. van der Waals volumes and radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  41. Bauzá, A.; Frontera, A. Electrostatically enhanced F···F interactions through hydrogen bonding, halogen bonding and metal coordination: An ab initio study. Phys. Chem. Chem. Phys. 2016, 18, 20381–20388. [Google Scholar] [CrossRef] [Green Version]
  42. Beker, R.J.; Colavita, P.E.; Murphy, D.M.; Platts, J.A.; Wallis, J.D. Fluorine-fluorine interactions in the solid state: An experimental and theoretical study. J. Phys. Chem. A 2012, 116, 1435–1444. [Google Scholar] [CrossRef]
  43. Takezawa, H.; Murase, T.; Resnati, G.; Metrangolo, P.; Fujita, M. Recognition of polyfluorinated compounds through self-aggregation in a cavity. J. Am. Chem. Soc. 2014, 136, 1786–1788. [Google Scholar] [CrossRef]
  44. Sugisawa, S.; Tabe, Y. Induced smectic phases of stoichiometric liquid crystal mixtures. Soft Matter. 2016, 12, 3103–3109. [Google Scholar] [CrossRef]
  45. Reichardt, C. Solvatochromic dyes as solvent polarity indicators. Chem. Rev. 1994, 94, 2319–2358. [Google Scholar] [CrossRef]
  46. Mataga, N.; Kaifu, Y.; Koizumi, M. The solvent effect on fluorescence spectrum, change of solute-solvent interaction during the lifetime of excited solute molecule. Bull. Chem. Soc. Jpn. 1955, 28, 690–691. [Google Scholar] [CrossRef] [Green Version]
  47. Mataga, N.; Kaifu, Y.; Koizumi, M. Solvent effects upon fluorescence spectra and the dipole moments of excited molecules. Bull. Chem. Soc. Jpn. 1956, 29, 465–470. [Google Scholar] [CrossRef] [Green Version]
  48. Birks, J.B. Photophysics of Aromatic Molecules; Birks, J.B., Ed.; Wiley: London, UK, 1970. [Google Scholar]
  49. Thomas, S.W.; Joly, G.D.; Swager, T.M. Chemical sensors based on amplifying fluorescent conjugated polymer. Chem. Rev. 2007, 107, 1339–1386. [Google Scholar] [CrossRef]
  50. Yaroshchuk, O.; Reznikov, Y. Photoalignment of liquid crystals: Basics and current trends. J. Mater. Chem. 2012, 22, 286–300. [Google Scholar] [CrossRef]
  51. Kawatsuki, N.; Goto, K.; Yamamoto, T. Photoinduced anisotropy and photoalignment of nematic liquid crystals by a novel polymer liqud crystal with a coumarin-containing side group. Liq. Cryst. 2001, 28, 1171–1176. [Google Scholar] [CrossRef]
  52. Kawatsuki, N.; Kuwabara, M.; Matsuura, Y.; Sasaki, T.; Ono, H. Photoalignment control of liquid crystals on photo-cross-linkable polymer liquid crystal film. J. Photopolym. Sci. Technol. 2005, 18, 17–22. [Google Scholar]
  53. Li, X.D.; Zhong, Z.X.; Lee, S.H.; Ghang, G.; Lee, M.H. Liquid crystal photoalignment using soluble photosensitive polyimide. Jpn. J. Appl. Phys. 2006, 45, 906. [Google Scholar] [CrossRef]
  54. Zhong, Z.X.; Li, X.; Lee, S.H.; Lee, M.H. Liquid crystal photoalignment material based on chloromethylated polyimide. Appl. Phys. Lett. 2004, 85, 2520. [Google Scholar] [CrossRef]
Figure 1. (a) Examples of fluorine-containing organic materials reported by other groups, and (b) by our group; (c) chemical structures of bistolanes 1 (bearing an alkyl chain) and 2 (bearing a fluoroalkyl unit).
Figure 1. (a) Examples of fluorine-containing organic materials reported by other groups, and (b) by our group; (c) chemical structures of bistolanes 1 (bearing an alkyl chain) and 2 (bearing a fluoroalkyl unit).
Crystals 11 00450 g001
Figure 2. Synthetic procedures for the preparation of the bistolanes 1a, 1b, and 2a2c used in this study.
Figure 2. Synthetic procedures for the preparation of the bistolanes 1a, 1b, and 2a2c used in this study.
Crystals 11 00450 g002
Figure 3. Molecular orbitals and orbital energies for bistolanes 1a and 1b (bearing an alkyl chain), and 2a2c (bearing a fluoroalkyl unit). In the MO diagrams, hydrogen atoms are omitted for clarity from the molecular structures.
Figure 3. Molecular orbitals and orbital energies for bistolanes 1a and 1b (bearing an alkyl chain), and 2a2c (bearing a fluoroalkyl unit). In the MO diagrams, hydrogen atoms are omitted for clarity from the molecular structures.
Crystals 11 00450 g003
Figure 4. Crystal structures, packing structures, and intermolecular interactions between neighboring molecules for (a) 2a and (b) 2c.
Figure 4. Crystal structures, packing structures, and intermolecular interactions between neighboring molecules for (a) 2a and (b) 2c.
Crystals 11 00450 g004
Figure 5. Phase transition behaviors of the bistolane derivatives during the 2nd heating [H] and cooling [C] processes (scan rate: 5.0 °C min−1). POM texture images of the LC phases are also shown.
Figure 5. Phase transition behaviors of the bistolane derivatives during the 2nd heating [H] and cooling [C] processes (scan rate: 5.0 °C min−1). POM texture images of the LC phases are also shown.
Crystals 11 00450 g005
Figure 6. (a) UV–Vis and PL spectra of 1a, 1b, and 2a2c measured in CH2Cl2 (inset: photographic images of the PL observed under UV irradiation at λex = 365 nm); (b) Commission Internationale de l’Eclailage (CIE) diagram indicating the PL color; (c) PL spectra of 1a and 2a in various solvents. The values in parentheses indicate the ET(30) values, which are quantitative indicators of the solvent polarity; (d) Lippert–Mataga plot calculated from the Stokes’ shift (Δν = νabsνPL) and a solvent polarity parameter (Δf).
Figure 6. (a) UV–Vis and PL spectra of 1a, 1b, and 2a2c measured in CH2Cl2 (inset: photographic images of the PL observed under UV irradiation at λex = 365 nm); (b) Commission Internationale de l’Eclailage (CIE) diagram indicating the PL color; (c) PL spectra of 1a and 2a in various solvents. The values in parentheses indicate the ET(30) values, which are quantitative indicators of the solvent polarity; (d) Lippert–Mataga plot calculated from the Stokes’ shift (Δν = νabsνPL) and a solvent polarity parameter (Δf).
Crystals 11 00450 g006
Figure 7. (a) PL spectra of 1a, 1b, and 2a2c in the crystalline form (inset: photographic images showing PL under UV irradiation at λex = 365 nm); (b) CIE color diagram obtained from the corresponding PL spectra.
Figure 7. (a) PL spectra of 1a, 1b, and 2a2c in the crystalline form (inset: photographic images showing PL under UV irradiation at λex = 365 nm); (b) CIE color diagram obtained from the corresponding PL spectra.
Crystals 11 00450 g007
Table 1. Theoretical data obtained from quantum chemical calculations 1.
Table 1. Theoretical data obtained from quantum chemical calculations 1.
Compoundµ||/µ (D) 2HOMO/LUMO (eV)Theoretical Transition Probability (%)λcalcd (nm)
1a0.96/1.79−6.88/−1.27HOMO→LUMO (92%)330
1b0.96/1.79−6.87/−1.27HOMO→LUMO (92%)331
2a6.19/1.49−6.99/−1.50HOMO→LUMO (87%)333
2b2.44/4.17−6.93/−1.35HOMO→LUMO (91%)331
2c6.57/1.54−7.00/−1.52HOMO→LUMO (87%)334
1 The ground state geometry and the TD-SCF approximation were obtained at the M06-2X/6-31+G(d) level of theory with CPCM for CH2Cl2. 2 µ|| and µ are the dipole moments along the major and minor molecular axes, respectively.
Table 2. Photophysical data for 1a, 1b, and 2a2c recorded in various solvents 1.
Table 2. Photophysical data for 1a, 1b, and 2a2c recorded in various solvents 1.
CompoundSolvent [ET(30)] 2λabs (nm) [ε, 103 L mol−1 cm−1]λex (nm)λPL (nm)ΦPL 3
1aCH2Cl2 [40.7]328 [80.4], 348 [sh, 55.6]3303771.00
Hexane [31.0]324 [20.4], 344 [sh, 12.9]324353, 368, 3800.30
CHCl3 [39.1]328 [57.2], 348 [sh, 37.3]330366, 3800.66
DMF [43.2]328 [33.7], 348 [sh, 23.2]3303880.78
MeCN [45.6]324 [38.1], 344 [sh, 26.8]3243880.68
1bCH2Cl2 [40.7]330 [63.0], 350 [sh, 43.0]3303861.00
2aCH2Cl2 [40.7]331 [50.9], 350 [sh, 37.0]3304001.00
Hexane [31.0]330 [2.6], 349 [sh, 1.4]330361, 3770.20
CHCl3 [39.1]333 [23.4], 350 [sh, 19.0]3303880.86
DMF [43.2]333 [23.4], 350 [sh, 19.0]3304200.83
MeCN [45.6]325 [5.3], 346 [sh, 3.7]3254200.71
2bCH2Cl2 [40.7]329 [61.0], 350 [sh, 42.4]3303861.00
2cCH2Cl2 [40.7]332 [56.7], 350 [sh, 41.7]3304051.00
1 Concentration: 1.0 × 10−5 mol L−1 for the UV–Vis absorption measurements and 1.0 × 10−6 mol L−1 for the PL measurements. 2 The ET(30) value is a quantitative indicator of the solvent polarity. 3 Measured using an integrating sphere; sh: shoulder peak.
Table 3. Photophysical data for 1a, 1b, and 2a2c in the crystalline form.
Table 3. Photophysical data for 1a, 1b, and 2a2c in the crystalline form.
CompoundλPL1ΦPL 2
1a404, 4220.57
1b405, 4210.79
2a418, 434(sh)1.00
2b401(sh), 4170.71
2c421, 434(sh)0.82
1 Excitation wavelength of incident light: 300 nm; sh: shoulder peak. 2 Measured using an integrating sphere.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yamada, S.; Wang, Y.; Morita, M.; Zhang, Q.; O’Hagan, D.; Nagata, M.; Agou, T.; Fukumoto, H.; Kubota, T.; Hara, M.; et al. Effect of Fluoroalkyl-Substituent in Bistolane-Based Photoluminescent Liquid Crystals on Their Physical Behavior. Crystals 2021, 11, 450. https://doi.org/10.3390/cryst11040450

AMA Style

Yamada S, Wang Y, Morita M, Zhang Q, O’Hagan D, Nagata M, Agou T, Fukumoto H, Kubota T, Hara M, et al. Effect of Fluoroalkyl-Substituent in Bistolane-Based Photoluminescent Liquid Crystals on Their Physical Behavior. Crystals. 2021; 11(4):450. https://doi.org/10.3390/cryst11040450

Chicago/Turabian Style

Yamada, Shigeyuki, Yizhou Wang, Masato Morita, Qingzhi Zhang, David O’Hagan, Masakazu Nagata, Tomohiro Agou, Hiroki Fukumoto, Toshio Kubota, Mitsuo Hara, and et al. 2021. "Effect of Fluoroalkyl-Substituent in Bistolane-Based Photoluminescent Liquid Crystals on Their Physical Behavior" Crystals 11, no. 4: 450. https://doi.org/10.3390/cryst11040450

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop