Next Article in Journal
Characterization of the Microstructure Evolution in IF-Steel and AA6016 during Plane-Strain Tension and Simple Shear
Previous Article in Journal
Comparative Study on Corrosion Protection of Reinforcing Steel by Using Amino Alcohol and Lithium Nitrite Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Liquid Crystalline Esters of Dibenzophenazines

by
Kevin John Anthony Bozek
1,
Kuan I Ho
1,
Tom Saint-Martin
1,2,
Panos Argyropoulos
1 and
Vance E. Williams
1,*
1
Department of Chemistry, Simon Fraser University, Burnaby, BC V5A 1S6, Canada
2
Department of Chemistry, Université Pierre & Marie Curie, Paris 75252, France
*
Author to whom correspondence should be addressed.
Materials 2015, 8(1), 270-284; https://doi.org/10.3390/ma8010270
Submission received: 31 October 2014 / Accepted: 4 January 2015 / Published: 14 January 2015

Abstract

:
A series of esters of 2,3,6,7-tetrakis(hexyloxy)dibenzo[a,c]phenazine-11-carboxylic acid was prepared in order to probe the effects of the ester groups on the liquid crystalline behavior. These compounds exhibit columnar hexagonal phases over broad temperature ranges. Variations in chain length, branching, terminal groups, and the presence of cyclic groups were found to modify transition temperatures without substantially destabilizing the mesophase range.

Graphical Abstract

1. Introduction

Discotic liquid crystals (DLCs) have been the focus of intense scrutiny over the past several years, owing in large part to their usefulness as organic semiconductors with high charge carrier mobilities, a lack of grain boundaries, and a predisposition to uniformly align. Such properties make DLCs attractive candidate materials for light emitting diodes, photovoltaic devices, and field effect transistors [1,2,3,4,5,6]. Successful exploitation of discotic liquid crystals requires balancing molecular electronic properties with the requirements of self-assembly. In principle, this can be accomplished by creating series of molecules in which phase behavior can be fine-tuned without substantially altering the electronic properties of the building blocks. Ideally, such families would also be constructed from readily available starting materials via reliable and facile synthetic approaches.
With these criteria in mind, we undertook the synthesis of a series of dibenzophenazine esters 26. Because esters can be formed from the reaction of an appropriate acid (e.g., 1) with any of number of readily available alcohols, discotic esters, as a family, hold the promise for considerable structural diversity. In the present work, we explore the impact of varying the R group with linear (2), branched (3) cyclic (45) and hydroxy-terminated substituents (6). Because communication across the ester group is limited, these structural changes are expected to have only a minimal impact on the electronic properties of the discotic core while allowing for the fine-tuning of the LC phase range. Surprisingly, despite these potential advantages, discotic esters have been largely neglected in the literature [7,8,9,10]. Dibenzophenazine esters in particular were targeted because of our previous observation that the methyl ester 2a possesses a columnar hexagonal phase over an extremely broad temperature range (111–201 °C) [11]. Because of the breadth of this phase, we posited that replacing the methyl group would lead to changes in phase stability without completely destroying the liquid crystallinity.

2. Results and Discussion

The synthesis of esters 2bf and 36 is outlined in Scheme 1. The linear alkyl esters 2bf and hydroxy-terminated esters 6a and 6c were prepared according to the method previously reported for 6b [12]. The acid 1 was first converted to its acid chloride using oxalyl chloride in the presence of dimethylformamide; subsequent treatment of this acid chloride with an excess of the appropriate alcohol afforded the target esters in moderate to good yields (33%–85%) after final purification.
Scheme 1. Reagents and conditions: (a) oxalyl chloride, dimethylformamide, dichloromethane, room temperature (RT), 3 h; ROH, pyridine, dichloromethane, RT, 20 h; (b) 4-dimethylaminopyridine (DMAP), 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDCI), dichloromethane, corresponding alcohol, RT, 24 h.
Scheme 1. Reagents and conditions: (a) oxalyl chloride, dimethylformamide, dichloromethane, room temperature (RT), 3 h; ROH, pyridine, dichloromethane, RT, 20 h; (b) 4-dimethylaminopyridine (DMAP), 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDCI), dichloromethane, corresponding alcohol, RT, 24 h.
Materials 08 00270 g008
Attempts to prepare the branched isopropyl analog 3 via this route were unsuccessful; rather than isolating this ester, the dimethylamide 7 was formed, presumably from the reaction of residual dimethylamine present in the DMF. Although not specifically targeted, this amide was isolated in sufficient yields to allow us to examine its phase behavior. In order to access the branched ester 3, we adopted an alternative synthetic approach in which the acid was activated with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDCI)/4-dimethylaminopyridine (DMAP), followed by reaction with isopropanol. This improved synthetic route was also used to prepare the cyclic esters, 4 and 5, which were isolated in 70% and 33% yields, respectively.
The electronic properties of all compounds were examined by UV-visible absorption spectroscopy in dilute chloroform solutions. The spectra of all esters are similar, with the λmax of the longest wavelength absorption band falling in the range of 419–422 nm; this small variation is unsurprising, since the electronic structure of the core is not expected to be sensitive to the nature of the functional group on the ester. A second, unresolved peak is also observed at slightly shorter wavelengths (λmax ~ 405 nm). Both peaks are slightly blue-shifted to λmax = 414 nm and 395 nm in the spectrum of the amide 7 (Figure 1), presumably because of differing electron-withdrawing ability of this substituent [13].
Figure 1. UV-visible absorption spectra of 2b (blue), 4 (black) and 7 (red) in CHCl3.
Figure 1. UV-visible absorption spectra of 2b (blue), 4 (black) and 7 (red) in CHCl3.
Materials 08 00270 g001
The optical properties of the esters were further probed by examining the absorption and emission spectra of the methyl ester, 2a, in solution and when cast as a thin film (Figure 2). The ester 2a exhibits strong luminescence in both the solution and the solid state, although we have not measured quantum yields at this time. The solid state and solution absorption spectra of this compound are very similar to one another, although the two long wavelength absorption bands noted above are sharper and more resolved in the thin film. Much larger changes are observed between the emission spectra. In solution, 2a exhibits a single broad peak centered at 600 nm; this peak is blue-shifted by approximately 40 nm in the solid state. The excitation spectra (not shown) of both samples closely match the absorption profiles. The solution emission and absorption spectra of 2a closely resemble those previously reported by Szydłowska and coworkers for the carboxylic acid 1 [14].
Figure 2. UV-visible absorption and emission spectra of 2a in solution (10−5 M CHCl3) and as thin films. Emission spectra were obtained using an excitation wavelength of 420 nm.
Figure 2. UV-visible absorption and emission spectra of 2a in solution (10−5 M CHCl3) and as thin films. Emission spectra were obtained using an excitation wavelength of 420 nm.
Materials 08 00270 g002
The liquid crystal properties of compounds 27 were studied by polarized optical microscopy (POM), differential scanning calorimetry (DSC), and variable temperature X-ray diffraction (XRD) experiments, the results of which are summarized in Table 1 and Table 2. DSC endotherms of all compounds exhibited at least two phase transitions upon first heating: a large enthalpy transition at low temperature followed by a smaller enthalpy transition at higher temperatures, consistent with crystalline-to-liquid crystalline melting and liquid crystalline (LC)-to-isotropic clearing transitions, respectively. The DSC cooling curves for all compounds show the isotropic-to-LC transition with minimal hysteresis. Notably, the LC-solid transitions are absent in all cases, indicating that the mesophases supercool from their LC phases (Figure 3). Second heating curves obtained immediately following cooling runs generally either lack a melting peak, or exhibit an exothermic peak associated with cold-crystallization of the sample. In contrast, cooled DSC samples that have been allowed to sit for extended periods do exhibit melting peaks on heating, providing evidence that samples crystallize over time.
Examination of 27 by microscopy confirms that all compounds form liquid crystalline phases on both heating and cooling. Samples cooled slowly into their LC phase from the isotropic state exhibit dendritic textures (Figure 4) that are typical of columnar hexagonal (Colh) phases. These samples show a strong tendency to homeotropically align between untreated glass substrates when slowly cooled from their isotropic liquids, which is a common feature of dibenzophenazine discotic mesogens [15,16]. The liquid crystalline textures of all samples are retained down to room temperature, well below their melting points, consistent with the DSC observation of supercooling. Most of the compounds were observed to crystallize at room temperature over the course of approximately 30 min. A notable exception was the phenyl ester 4, which crystallizes over a period of weeks rather than minutes (Figure 5). It is not clear at this time why this compound crystallizes at a much slower rate than the other esters studied, but does suggest that variations in the ester group may provide a practical means to kinetically suppress crystallization.
Table 1. Phase behavior of compounds 2–7. a Transition temperatures and enthalpies were determined by differential scanning calorimetry (DSC) (scan rate = 10 °C/min) on second heating run (see text); b Cr = crystal, Colh = columnar hexagonal, I = isotropic; c see reference 11; d see reference 12.
Table 1. Phase behavior of compounds 2–7. a Transition temperatures and enthalpies were determined by differential scanning calorimetry (DSC) (scan rate = 10 °C/min) on second heating run (see text); b Cr = crystal, Colh = columnar hexagonal, I = isotropic; c see reference 11; d see reference 12.
CompoundPhaseT/°C (ΔH/J g−1) aPhaseT/°C (ΔH/J g−1) aPhase b
2a cCr111.3 (39.1)
Materials 08 00270 i001
Colh200.6 (4.7)
Materials 08 00270 i002
196.5 (4.8)
I
2bCr64.3 (45.6)
Materials 08 00270 i001
Colh210.8 (8.0)
Materials 08 00270 i002
209.5 (11.1)
I
2cCr86.8 (40.1)
Materials 08 00270 i001
Colh212.1 (7.9)
Materials 08 00270 i002
210.2 (9.8)
I
2dCr92.5 (42.0)
Materials 08 00270 i001
Colh209.6 (8.3)
Materials 08 00270 i002
208.2 (8.6)
I
2eCr97.5 (68.0)
Materials 08 00270 i001
Colh202.9 (7.7)
Materials 08 00270 i002
202.0 (7.7)
I
2fCr92.7 (61.8)
Materials 08 00270 i001
Colh191.3 (4.6)
Materials 08 00270 i002
190.9 (6.3)
I
3Cr78.6 (32.7)
Materials 08 00270 i001
Colh218.8 (9.6)
Materials 08 00270 i002
217.1 (10.6)
I
4Cr124.7 (38.7)
Materials 08 00270 i001
Colh220.5 (4.1)
Materials 08 00270 i002
218.6 (5.8)
I
5Cr84.7 (18.5)
Materials 08 00270 i001
Colh202.5 (5.6)
Materials 08 00270 i002
199.1 (7.3)
I
6aCr99.7 (15.0)
Materials 08 00270 i001
Colh198.7 (6.6)
Materials 08 00270 i002
197.6 (6.5)
I
6b dCr130.1 (71.8)
Materials 08 00270 i001
Colh202.0 (9.4)
Materials 08 00270 i002
201.1 (9.3)
I
6cCr96.5 (51.2)
Materials 08 00270 i001
Colh188.8 (7.9)
Materials 08 00270 i002
187.8 (8.8)
I
7Cr122.9 (45.5)
Materials 08 00270 i001
Colh180.3 (2.9)
Materials 08 00270 i002
180.1 (3.1)
I
Table 2. X-ray diffraction (XRD) data and lattice constants for compounds 27.
Table 2. X-ray diffraction (XRD) data and lattice constants for compounds 27.
CompoundTemperature (°C)d-spacings (Å)Miller Index (hkl)Phase (Lattice Constants)
2a15017.3100Colh
4.3Alkyl Halo(a = 20.0 Å)
3.5π–π stacking
2b17917.7100Colh
5Alkyl Halo(a = 20.5 Å)
3.8π–π stacking
2c17717.4100Colh
4.9Alkyl Halo(a = 20.1 Å)
2d17817.7100Colh
4.9Alkyl Halo(a = 20.5 Å)
2e17917.9100Colh
4.7Alkyl Halo(a = 20.7 Å)
3.6π–π stacking
2f15518.5100Colh
4.6Alkyl Halo(a = 21.4 Å)
318517.7100Colh
10.3110(a = 20.4 Å)
5.4Alkyl Halo
417517.7100Colh
4.5Alkyl Halo(a = 20.4 Å)
515518.1100Colh
10.6110(a = 20.9 Å)
6.6Alkyl Halo
6a16017.3100Colh
5.1Alkyl Halo(a = 20.0 Å)
3.6π–π stacking
6b17017.5100Colh
4.6Alkyl Halo(a = 20.2 Å)
3.5π–π stacking
6c15718.3100Colh
5.1Alkyl Halo(a = 21.1 Å)
716917.1100Colh
4.7Alkyl Halo(a = 19.8 Å)
Figure 3. Differential scanning calorimetry (DSC) endotherm and exotherm for first heating and cooling cycle of 3 (heating/cooling rate = 10 °C/min).
Figure 3. Differential scanning calorimetry (DSC) endotherm and exotherm for first heating and cooling cycle of 3 (heating/cooling rate = 10 °C/min).
Materials 08 00270 g003
Figure 4. Representative polarized optical micrographs of columnar phases formed upon cooling from the isotropic phase of (a) 2b; (b) 4; (c) 6c and (d) 7 (200× magnification). Images (a)–(c) viewed employing a 530 nm quarter wavelength retardation plate in order to improve viewing contrast.
Figure 4. Representative polarized optical micrographs of columnar phases formed upon cooling from the isotropic phase of (a) 2b; (b) 4; (c) 6c and (d) 7 (200× magnification). Images (a)–(c) viewed employing a 530 nm quarter wavelength retardation plate in order to improve viewing contrast.
Materials 08 00270 g004
Figure 5. Polarized optical microscopy (POM) micrographs of compound 4 (a) at 175 °C; (b) at room temperature after 7 days and (c) at room temperature after 27 days.
Figure 5. Polarized optical microscopy (POM) micrographs of compound 4 (a) at 175 °C; (b) at room temperature after 7 days and (c) at room temperature after 27 days.
Materials 08 00270 g005
XRD experiments are consistent with the assignment of Colh phases being formed for all mesogens. X-ray diffraction patterns of the liquid crystalline phases of all compounds exhibit one intense peak in the low angle region, which was identified as the (100) peak of a hexagonal lattice. A second broad alkyl halo peak at distances of approximately 4.5 Å is also observed in all cases. Peaks associated with the π–π stacking distances at approximately 3.5 Å were also observed in the cases of compounds 2a, 2b, 2e, 6a and 6b. Changes in the ester group have only a small impact on the intercolumnar spacing (Table 2).
The effects of chain length on the LC phase ranges of the linear esters 2af are shown graphically in Figure 6. Increasing chain length is accompanied by a general tendency towards lower clearing temperatures and higher melting temperatures, resulting in an overall narrowing of the LC range, from a maximum 137 °C range for 2b (64–211 °C) to a 98 °C range for 2f (93–191 °C). This effect parallels the typical trend for discotic mesogens when one or more of the peripheral side chains are varied [17,18]. In the present case, the narrowing of the phase range is much less pronounced than when all of the side chains of a discotic are lengthened. The methyl ester 2a represents an anomaly within this series, with a melting temperature that is almost 50 °C higher than that of the ethyl ester 2b; as a consequence, this molecule has the narrowest LC phase range in this family. Thus, extending the ester chain length beyond a single carbon has the beneficial effect of shifting the LC phase range closer to room temperature.
Figure 6. Plot of melting (Tm, black circles) and clearing (Tc, red squares) temperatures of linear alkyl esters 2af, as a function of ester chain length, where n = #carbons in alkyl chain of ester (i.e., 2a, n = 1; 2b, n = 2, etc.).
Figure 6. Plot of melting (Tm, black circles) and clearing (Tc, red squares) temperatures of linear alkyl esters 2af, as a function of ester chain length, where n = #carbons in alkyl chain of ester (i.e., 2a, n = 1; 2b, n = 2, etc.).
Materials 08 00270 g006
The introduction of branched chains into discotic mesogens is often associated with significant changes in melting and clearing temperatures. However, the nature of these changes is far from consistent; both elevated and depressed transition temperatures are found to accompany branching [19,20,21]. It was therefore of interest to compare the differences between the isopropyl derivative 3 with the linear esters in series 2. Of particular note, the clearing temperature of the branched ester 3 (219 °C) is appreciably higher than that of its linear isomer 2c (Tc = 212 °C), indicating that branching leads to a stabilization of the LC phase. In contrast, the melting temperature of the linear ester 2c (87 °C) is higher than that of 3 (79 °C). This may reflect either the greater stability of the LC phase of 3 (which would lead to a depression of the melting point) or, as seems more likely, the destabilization of its crystalline solid phase. More generally, the branched-chain derivative 3 exhibits one of the lowest melting temperatures and highest clearing points of any of the esters studied in this work, with only the ethyl compound 2b having a broader LC phase range. This stabilization of the LC phase may result from the isopropyl group being able to fill space more effectively, allowing for tighter packing and stronger interactions [22].
Replacing the linear and branched alkyl groups in 2 and 3 with a cyclohexyl group in 5 led to only modest changes in phase behavior. Despite the bulkiness of this group, 5 both melts and clears at temperatures that are similar to the members of series 2 and 3, and it appears that the cyclohexyl group neither strongly stabilizes nor destabilizes the LC phase. In marked contrast, both transition temperatures of the phenyl ester 4 are substantially elevated, with melting and clearing temperatures that are 40 °C and 18 °C higher, respectively, than those of the structurally similar derivative 5. Indeed, 4 has the highest Tm and Tc values of any of the compounds prepared for this study, denoting that both the crystalline and liquid crystalline phases are stabilized by the phenyl substituent, likely via π–π interactions.
We had anticipated that incorporating chains terminated with hydroxyl groups (6ac) would lead to a stabilization of the liquid crystalline phases, as is often the case for mesogens capable of hydrogen bonding [11,23,24]. Surprisingly, replacing the terminal hydrogen of an alkyl chain with a hydroxyl group, rather than increasing the LC-isotropic transition temperature, actually leads to a slight decrease in Tc (Figure 7). Thus, it appears that hydrogen bonding does little to stabilize the LC phases in the present systems, and may in fact destabilize these phases. Similar trends have also been observed for triphenylene and penta-ethynylbenzene discotic mesogens bearing a single hydroxy-terminated chain [25,26].
Figure 7. Phase ranges of compounds 2b, 2c, 2f and 6ac; Cr = crystalline solid, Col = columnar hexagonal LC, I = isotropic liquid.
Figure 7. Phase ranges of compounds 2b, 2c, 2f and 6ac; Cr = crystalline solid, Col = columnar hexagonal LC, I = isotropic liquid.
Materials 08 00270 g007
Finally, the amide derivative 7 also forms a columnar hexagonal phase, albeit with a clearing temperature (180 °C) that is substantially lower than any of the esters studied. We ascribe this depression of the isotropization temperature to differences in the electronic properties between esters and amides, with amides being less electron-withdrawing (Hammett σp = 0.37) than esters (σp = 0.45) [13]. This observation is consistent with our earlier finding of a strong correlation between clearing temperatures of discotic mesogens and the electron-withdrawing ability substituents directly attached to the dibenzophenazine core [27]. It is notable that changing the functional group from an ester to an amide has a much greater tendency to destabilize the columnar phase than changing the nature of the ester. Thus, electronic effects appear to trump steric considerations, at least over the range of functionalities examined.

3. Experimental Section

All solvents employed were reagent grade and were used as obtained from suppliers unless otherwise noted. Oxalyl chloride was purchased from Aldrich (St. Louis, MO, USA). The 1,3-propanediol was purchased from ICN biomedical supplies (Santa Ana, CA, USA). Pyridine was purchased from Anachemia (Richmond, Canada) and was dried over KOH pellets. The DMAP and EDCI were purchased from TCI (Portland, OR, USA). Column chromatography was performed using silica gel 60 (230–400 mesh) purchased from Silicyle Inc. (Quebec City, Canada). The carboxylic acid derivative, 1 [11] and esters 2a [11] and 6b [12] were prepared according to previously published methods.
The 400 MHz 1H NMR spectra were obtained using a Bruker AMX-400 400MHz NMR spectrometer (Billerica, MA, USA). 500 MHz 1H NMR spectra and 125 MHz 13C NMR spectra were obtained using a Varian AS500 Unity Inova 500 MHz spectrometer (Billerica, MA, USA). The 600 MHz 1H NMR and 150 MHz 13C NMR spectra were obtained using a Bruker Avance 600 MHz spectrometer equipped with a TCI cryoprobe. Phase transition temperatures and enthalpies were investigated using differential scanning calorimetry on a TA Instruments DSC Q2000 (New Castle, DE, USA) equipped with a TA Instruments Refrigerated Cooling System 90, heating and cooling at a rate of 10 °C min−1. Texture analysis was carried out using optical polarizing microscopy on an Olympus BX50 microscope (Richmond Hill, Canada) with crossed polarizers using a Linkam LTS350 heating stage (Tadworth, UK). X-ray scattering experiments were conducted using a Rigaku R-Axis Rapid diffractometer (The Woodlands, TX, USA) equipped with a temperature controller [28]. X-ray samples were prepared by first heating the compound to its isotropic phase, then loading it by capillary action into a glass capillary.
Infrared spectroscopy (IR) spectra were obtained using a Perkin-Elmer Spectrum One FT-IR with an attenuated total reflectance (ATR) attachment. High resolution mass spectroscopy was acquired on a BrukermicrOTOF-Q II using electrospray ionization positive at the University of Notre Dame. UV-Vis spectra were obtained using a Varian Cary 300 Bio UV-Visible spectrometer (Santa Clara, CA, USA) and fluorescence measurements were obtained using a PTI QuantaMaster Fluorescence Spectrometer (London, ON, Canada). Thin films used for UV-visible and emissions studies of 2a were prepared by spin coating (1700 rpm) a 10−3 M chloroform solution onto a microscope cover slip using a Laurell WS-400A-6NPP/LITE spin coater (North Wales, PA, USA).

3.1. Synthesis Method 1 (Oxalyl Chloride Method)

In a round bottom flask (100 mL) equipped with a magnetic stir bar, the carboxylic acid (1 equivalent) was suspended in dry DCM (50 mL) and the mixture was set to stir. To this was added DMF (0.1 equivalent) and oxalyl chloride (0.1 mL 2 M solution, 1.1 equivalent). The resulting suspension was allowed to stir for 2.5 h at room temperature. At this time the solvent was removed under vacuum. The solid was then resuspended in dry DCM (25 mL) and a mixture of alcohol (10 mL, 30 equivalent) and pyridine (10 mL, 30 equivalent) in dry DCM (40 mL) was added and the reaction was allowed to stir overnight at room temperature. The resulting suspension was filtered and washed with water (2 × 50 mL), HCl (10%, 50 mL), and brine (50 mL). The organic layers were combined, dried over MgSO4, and the solvent was removed under vacuum affording the target compound.

3.2. Synthesis Method 2 (EDCI/DMAP Method)

In a round bottom flask (100 mL) equipped with a magnetic stir bar, the carboxylic acid (1 equivalent), the alcohol (1.1 equivalent), and DMAP (3 equivalent) were dissolved in dry DCM (25 mL) and the mixture was set to stir and cooled to 0 °C in an ice water bath. To this was added EDCI (2 equivalent). The resulting mixture was allowed to stir for 24 h. The resulting solution was washed with water (2 × 50 mL), HCl (10%, 50 mL), and brine (50 mL). The organic layers were combined, dried over MgSO4, and the solvent was removed under vacuum to afford the target compound.
Compound 2b was prepared according to Method 1 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 36% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 14H), 1.42 (m, 15H), 1.59 (m, 10H), 1.98 (m, 8H), 4.26 (t, 4H, J = 6.6 Hz), 4.33 (dd, 4H, J = 12.1, 6.5 Hz), 4.52 (q, 2H, J = 7.1 Hz), 7.66 (m, 2H), 8.34 (m, 2H), 8.73 (m, 2H), 9.0 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 170.80, 152.08, 152.01, 149.49, 142.65, 142.60, 141.63, 140.66, 136.35, 129.68, 127.78, 127.73, 126.85, 126.72, 123.62, 123.58, 108.88, 108.79, 106.45, 106.43, 69.64, 69.19, 69.15, 39.81, 35.54, 31.67, 29.71, 29.32, 29.25, 25.85, 25.82, 22.66, 14.07, 14.05; UV (λmax(log ε), CHCl3): 288(4.58), 339(4.47), 419(4.44); IR (neat): 1723 cm−1 (ν: C=O); High resolution mass spectrometry (HRMS) (Electrospray ionization (ESI))+ : m/z calcd for C47H65N2O6+: 753.4838, found: 753.4837.
Compound 2c was prepared according to Method 1 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 49% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 16H), 1.42 (m, 15H), 1.59 (m, 10H), 1.98 (m, 8H), 4.29 (t, 4H, J = 6.5 Hz), 4.36 (dd, 4H, J = 12.3, 6.4 Hz), 4.42 (t, 2H, J = 6.6 Hz), 7.74 (s, 2H), 8.37 (m, 2H), 8.81 (d, 2H, J = 4.1 Hz), 9.04 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV (λmax(log ε), CHCl3): 288(4.85), 339(4.75), 419(4.41); IR (neat): 1726 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C48H67N2O6+: 767.4994, found: 767.4994.
Compound 2d was prepared according to Method 1 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 36% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 18H), 1.42 (m, 15H), 1.59 (m, 10H), 1.98 (m, 8H), 4.29 (t, 4H, J = 6.5 Hz), 4.36 (dd, 4H, J = 12.3, 6.4 Hz), 4.46 (t, 2H, J = 6.6 Hz), 7.74 (s, 2H), 8.35 (m, 2H), 8.81 (d, 2H, J = 4.1 Hz), 9.04 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 288(4.83), 339(4.72), 419(4.31); IR (neat): 1726 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C49H69N2O6+: 781.515, found: 781.5147.
Compound 2e was prepared according to Method 1 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 37% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 20H), 1.42 (m, 15H), 1.59 (m, 10H), 1.98 (m, 8H), 4.29 (t, 4H, J = 6.5 Hz), 4.36 (dd, 4H, J = 11.9, 6.6 Hz), 4.45 (t, 2H, J = 6.7 Hz), 7.73 (s, 2H), 8.37 (m, 2H), 8.81 (d, 2H, J = 4.6 Hz), 9.04 (d, 1H, J = 1.3 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 288(4.72), 339(4.61), 419(4.28); FT-IR (neat): 1722 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C50H71N2O6+: 795.5307, found: 795.5307.
Compound 2f was prepared according to Method 1 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 33% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 22H), 1.42 (m, 15H), 1.59 (m, 10H), 1.98 (m, 8H), 4.27 (t, 4H, J = 6.6 Hz), 4.34 (dd, 4H, J = 11.8, 6.5 Hz), 4.45 (t, 2H, J = 6.7 Hz), 7.68 (s, 2H), 8.35 (m, 2H), 8.75 (d, 2H, J = 5.3 Hz), 9.01 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 288(4.92), 339(4.81), 419(4.48); FT-IR (neat): 1722 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C51H73N2O6+: 809.5463, found: 809.5465.
Compound 7 was prepared according to Method 1 and purified by column chromatography using 1% methanol in dichloromethane as eluent (Rf = 0.08) to afford orange crystals in 40% yield. 1H NMR (400 MHz, CDCl3): δ 0.94 (m, 11H), 1.25 (s, 3H), 1.42 (m, 15H), 1.60 (m, 10H), 1.97 (m, 8H), 3.14 (s, 3H), 3.23 (s, 3H), 4.29 (m, 8H), 7.67 (s, 1H), 7.68 (s, 1H), 7.84 (dd, 1H, J = 8.6, 1.9 Hz), 8.31 (m, 2H), 8.70 (s, 1H), 8.74 (s, 1H) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 286(4.70), 336(4.53), 414(4.16); FT-IR (neat): 1608 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C48H69N3O5+: 752.4997, found: 752.4997.
Compound 6a was prepared according to Method 1 and purified by recrystallization from acetonitrile to afford red crystals in 85% yield. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 13H), 1.42 (m, 18H), 1.61 (m, 7H), 1.99 (m, 8H), 4.08 (m, 4H), 4.31 (m, 8H), 4.61 (m, 4H), 7.71 (s, 2H), 8.36 (m, 2H), 8.76 (s, 2H), 9.05 (d, 1H, J = 1.3 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 170.80, 152.08, 152.01, 149.49, 142.65, 142.60, 141.63, 140.66, 136.35, 129.68, 127.78, 127.73, 126.85, 126.72, 123.62, 123.58, 108.88, 108.79, 106.45, 106.43, 69.64, 69.19, 69.15, 39.81, 35.54, 31.67, 29.71, 29.32, 29.25, 25.85, 25.82, 22.66, 14.07, 14.05; FT-IR (neat): 1726 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C47H65N2O7+: 769.4786, found: 769.4786.
Compound 6c was prepared according to Method 1 and purified by recrystallization from acetonitrile to afford red crystals in 78% yield. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 13H), 1.42 (m, 17H), 1.61 (m, 12H), 1.96 (m, 2H), 2.17 (9H), 3.70 (t, 2H, J = 6.5 Hz), 4.33 (m, 8H), 4.47 (t, 2H, J = 6.6 Hz), 7.73 (s, 2H), 8.37 (m, 2H), 8.80 (s, 1H), 8.81 (2, 1H), 9.04 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 170.80, 152.08, 152.01, 149.49, 142.65, 142.60, 141.63, 140.66, 136.35, 129.68, 127.78, 127.73, 126.85, 126.72, 123.62, 123.58, 108.88, 108.79, 106.45, 106.43, 69.64, 69.19, 69.15, 39.81, 35.54, 31.67, 29.71, 29.32, 29.25, 25.85, 25.82, 22.66, 14.07, 14.05; FT-IR (neat): 1722 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C51H73N2O7+: 825.5413, found: 825.5412.
Compound 3 was prepared according to Method 2 and purified by column chromatography using dichloromethane as eluent (Rf = 0.43) to afford orange crystals in 27% yield and then further purified via recrystallization from hexanes. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 11H), 1.42 (m, 15H), 1.47 (s, 3H), 1.49 (s, 3H), 1.59 (m, 10H), 1.98 (m, 8H), 4.29 (t, 4H, J = 6.5 Hz), 4.36 (dd, 4H, J = 12.3, 6.4 Hz), 5.39 (m, 1H), 7.74 (s, 2H), 8.37 (m, 2H), 8.81 (d, 2H, J = 4.1 Hz), 9.04 (d, 1H, J = 1.6 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; FT-IR (neat): 1723 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C48H67N2O6+: 767.4994, found: 767.4994.
Compound 4 was prepared according to Method 2 and purified by recrystallization from acetonitrile to afford red crystals in 70% yield. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 13H), 1.43 (m, 15H), 1.61 (m, 10H), 1.98 (m, 8H), 4.32 (m 8H), 7.34 (m, 3H), 7.49 (m, 2H), 7.72 (s, 2H), 8.47 (m, 2H), 8.80 (s, 2H), 9.23 (d, 1H, J = 1.8 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 290(4.75), 342(4.66), 422(4.33); FT-IR (neat): 1738 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C51H65N2O6+: 801.4838, found: 801.4837.
Compound 5 was prepared according to Method 2 and purified by recrystallization from acetonitrile to afford red crystals in 33% yield. 1H NMR (400 MHz, CDCl3): δ 0.95 (m, 11H), 1.43 (m, 16H), 1.60 (m, 14H), 1.98 (m, 13H), 4.32 (m 8H), 5.32 (m, 1H), 7.71 (s, 2H), 8.37 (m, 2H), 8.79 (s, 2H), 9.02 (d, 1H, J = 1.5 Hz) ppm; 13C NMR (151 MHz, CDCl3): δ 166.29, 152.32, 152.06, 149.56, 149.46, 143.38, 143.22, 142.77, 140.58, 132.03, 130.41, 129.27, 128.24, 127.19, 126.68, 123.57, 123.46, 108.99, 108.79, 106.45, 106.32, 69.66, 69.59, 69.18, 69.16, 61.45, 31.67, 29.33, 29.31, 29.25, 25.85, 25.82, 22.67, 22.66, 14.45, 14.07, 14.05; UV-vis (λmax(log ε), CHCl3): 288(4.68), 339(4.57), 419(4.24); FT-IR (neat): 1722 cm−1 (ν: C=O); HRMS (ESI+): m/z calcd for C51H71N2O6+: 807.5307, found: 807.5307.

4. Conclusions

In the current work, we have demonstrated that dibenzophenazine esters are a versatile class of compounds that form columnar hexagonal phases over extremely broad temperature ranges. These systems tolerate considerable variation in structure without the loss of liquid crystallinity and without significantly altering the electronic structure of the core. Such features may make these compounds promising candidates as organic semiconductors, where fine-tuning self-assembly independently of the electronic properties is an important criterion. We have further observed that replacing the ester functionality with an amide group substantially lowers the clearing temperature, underscoring the importance of electronic considerations in the design of new mesogens. A common feature of all molecules examined in this study was their tendency to supercool. While most compounds crystallized quickly, the phenyl ester 4 retained liquid crystalline ordering for prolonged periods at room temperature, pointing the way to the design of stable glass-forming mesogens. Such studies are currently being undertaken in our laboratory.

Acknowledgments

The authors would like to acknowledge the Natural Sciences and Engineering Research Council of Canada for funding this research. Tom Saint-Martin gratefully acknowledges Université Pierre & Marie Curie (UPMC) and The Solidarity Fund and Student Initiatives Deleopment UPMC for a financial support.

Author Contributions

This project was overseen by Vance Williams. Kevin Bozek synthesized all compounds and carried out XRD, DSC and polarized microscopy on all samples. Solution UV-visible spectra were measured by Kuan I Ho and Panos Argyropoulos. Tom Saint-Martin prepared thin films of 2b and carried out all UV-visible and emission experiments on this compound. This manuscript was prepared by Kevin Bozek and Vance Williams.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaafarani, B.R. Discotic liquid crystals for opto-electronic applications. Chem. Mater. 2011, 23, 378–396. [Google Scholar] [CrossRef]
  2. Laschat, S.; Baro, A.; Steinke, N.; Giesselmann, F.; Hägele, C.; Scalia, G.; Judele, R.; Kapatsina, E.; Sauer, S.; Schreivogel, A.; et al. Discotic liquid crystals: From tailor-made synthesis to plastic electronics. Angew. Chem. Int. Ed. 2007, 46, 4832–4887. [Google Scholar] [CrossRef]
  3. Imrie, C.T.; Henderson, P.A. Liquid crystal dimers and higher oligomers: Between monomers and polymers. Chem. Soc. Rev. 2007, 36, 2096–2124. [Google Scholar] [CrossRef] [PubMed]
  4. Bushby, R.J.; Kawata, K. Liquid crystals that affected the world: Discotic liquid crystals. Liq. Cryst. 2011, 38, 1415–1426. [Google Scholar] [CrossRef]
  5. Kumar, S. Self-organization of disc-like molecules: Chemical aspects. Chem. Soc. Rev. 2006, 35, 83–109. [Google Scholar] [CrossRef] [PubMed]
  6. Kumar, S. Triphenylene-based discotic liquid crystal dimers, oligomers and polymers. Liq. Cryst. 2005, 32, 1089–1113. [Google Scholar] [CrossRef]
  7. Keuker-Baumann, S.; Bock, H.; Sala, F.D.; Benning, S.A.; Haßheider, T.; Frauenheim, T.; Kitzerow, H.-S. Absorption and luminescence spectra of electroluminescent liquid crystals with triphenylene, pyrene and perylene units. Liq. Cryst. 2001, 28, 1105–1113. [Google Scholar] [CrossRef]
  8. Hassheider, T.; Benning, S.A.; Kitzerow, H.-S.; Achard, M.-F.; Bock, H. Color-tuned electroluminescence from columnar liquid crystalline alkyl arenecarboxylates. Angew. Chem. Int. Ed. 2001, 40, 2060–2063. [Google Scholar] [CrossRef]
  9. Benning, S.; Kitzerow, H.-S.; Bock, H.; Achard, M.-F. Fluorescent columnar liquid crystalline 3,4,9,10-tetra-(n-alkoxycarbonyl)-perylenes. Liq. Cryst. 2000, 27, 901–906. [Google Scholar] [CrossRef]
  10. Ito, S.; Wehmeier, M.; Brand, J.D.; Kübel, C.; Epsch, R.; Rabe, J.P.; Müllen, K. Synthesis and self-assembly of functionalized hexa-peri-hexabenzocoronenes. Chem. Eur. J. 2000, 6, 4327–4342. [Google Scholar] [CrossRef] [PubMed]
  11. Foster, E.J.; Lavigueur, C.; Ke, Y.-C.; Williams, V.E. Self-assembly of hydrogen-bonded molecules: Discotic and elliptical mesogens. J. Mater. Chem. 2005, 15, 4062–4068. [Google Scholar] [CrossRef]
  12. Bozek, K.J.A.; Williams, V.E. Folded discotic dimers. Soft Matter 2014, 10, 5749–5754. [Google Scholar] [CrossRef] [PubMed]
  13. Hansch, C.; Leo, A.; Unger, S.H.; Kim, K.H.; Nikaitani, D.; Lien, E.J. Aromatic substituent constants for structure-activity correlations. J. Med. Chem. 1973, 16, 1207–1216. [Google Scholar] [CrossRef] [PubMed]
  14. Szydłowska, J.; Krzyczkowska, P.; Salamończyk, M.; Górecka, E.; Pociecha, D.; Maranowski, B.; Krówczyński, A. Gelling and fluorescent mesogens of quinoxaline analogs. J. Mater. Chem. C 2013, 1, 6883–6889. [Google Scholar] [CrossRef]
  15. Ong, C.W.; Hwang, J.-Y.; Tzeng, M.-C.; Liao, S.-C.; Hsu, H.-F.; Chang, T.-H. Dibenzo[a,c]phenazine with six-long alkoxy chains to probe optimization of mesogenic behavior. J. Mater. Chem. 2007, 17, 1785–1790. [Google Scholar] [CrossRef]
  16. Ong, C.W.; Hwang, C.-Y.; Liao, S.-C.; Pan, C.-H.; Chang, T.-H. Indent and bulge-discogens for controlling the columnar mesophase. J. Mater. Chem. 2009, 19, 5149–5154. [Google Scholar] [CrossRef]
  17. Kumar, S. Chemistry of Discotic Liquid Crystals: From Monomers to Polymers; CRC Press: Boca Raton, FL, USA, 2010. [Google Scholar]
  18. Lavigueur, C.; Foster, E.J.; Williams, V.E. Self-assembly of discotic mesogens in solution and in liquid crystalline phases: Effects of substituent position and hydrogen bonding. J. Am. Chem. Soc. 2008, 130, 11791–11800. [Google Scholar] [CrossRef] [PubMed]
  19. Collard, D.M.; Lillya, C.P. Control of thermal phase behavior of disklike molecules by modification of side-chain structure. J. Am. Chem. Soc. 1989, 111, 1829–1830. [Google Scholar] [CrossRef]
  20. Collard, D.M.; Lillya, C.P. Structure–property relationships for the thermal phase behavior of discotic liquid crystals: The effect of branching and unsaturation in the side chains of disklike molecules. J. Am. Chem. Soc. 1991, 113, 8577–8583. [Google Scholar] [CrossRef]
  21. Varshney, S.K.; Prasad, V.; Takezoe, H. Room-temperature discotic cholesteric and nematic phases: influence of 3,7-dimethyloctane peripheral chain on the molecular self-assembly of radial polyalkynylbenzene. Liq. Cryst. 2011, 38, 53–60. [Google Scholar] [CrossRef]
  22. Stackhouse, P.J.; Hird, M. Influence of branched chains on the mesomorphic properties of symmetrical and unsymmetrical triphenylene discotic liquid crystals. Liq. Cryst. 2008, 35, 597–607. [Google Scholar] [CrossRef]
  23. Paleos, C.M.; Tsiourvas, D. Supramolecular hydrogen-bonded liquid crystals. Liq. Cryst. 2001, 28, 1127–1161. [Google Scholar] [CrossRef]
  24. Wang, B.Q.; Zhao, K.Q.; Hu, P.; Yu, W.H.; Gao, C.Y.; Shimizu, Y. Tuning hydrogen-bonding with amide groups for stable columnar mesophases of triphenylene discogens. Mol. Cryst. Liq. Cryst. 2007, 479, 135/[1173]–150/[1188]. [Google Scholar] [CrossRef]
  25. Boden, N.; Bushby, R.J.; Lu, Z.B. A rational synthesis of polyacrylates with discogenic side groups. Liq. Cryst. 1998, 25, 47–58. [Google Scholar] [CrossRef]
  26. Grafe, A.; Janietz, D.; Frese, T.; Wendorff, J.H. Star-shaped discotic oligomesogens based on radial pentakisphenylethynylbenzene moieties. Chem. Mater. 2005, 17, 4979–4984. [Google Scholar] [CrossRef]
  27. Foster, E.J.; Jones, R.B.; Lavigueur, C.; Williams, V.E. Structural factors controlling the self-assembly of columnar liquid crystals. J. Am. Chem. Soc. 2006, 128, 8569–8574. [Google Scholar] [CrossRef] [PubMed]
  28. Lavigueur, C.; Foster, E.J.; Williams, V.E. A simple and inexpensive capillary furnace for variable-temperature X-ray diffraction. J. Appl. Crystallogr. 2008, 41, 214–216. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Bozek, K.J.A.; Ho, K.I.; Saint-Martin, T.; Argyropoulos, P.; Williams, V.E. Liquid Crystalline Esters of Dibenzophenazines. Materials 2015, 8, 270-284. https://doi.org/10.3390/ma8010270

AMA Style

Bozek KJA, Ho KI, Saint-Martin T, Argyropoulos P, Williams VE. Liquid Crystalline Esters of Dibenzophenazines. Materials. 2015; 8(1):270-284. https://doi.org/10.3390/ma8010270

Chicago/Turabian Style

Bozek, Kevin John Anthony, Kuan I Ho, Tom Saint-Martin, Panos Argyropoulos, and Vance E. Williams. 2015. "Liquid Crystalline Esters of Dibenzophenazines" Materials 8, no. 1: 270-284. https://doi.org/10.3390/ma8010270

Article Metrics

Back to TopTop