Next Article in Journal
Experimental Evaluation of Cement Mortars with End-of-Life Tyres Exposed to Different Surface Treatments
Previous Article in Journal
Enhancement of Optical Activity and Properties of Barium Titanium Oxides to Be Active in Sunlight through Using Hollandite Phase Instead of Perovskite Phase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Thermal and Optical Characterizations of New Lateral Organic Systems

by
Laila Ahmed. Al-Mutabagani
1,
Latifah A. Alshabanah
1,
Sobhi M. Gomha
2,3 and
Hoda A. Ahmed
2,4,*
1
Department of Chemistry, College of Science, Princess Nourah bint Abdulrahman University, Riyadh 11671, Saudi Arabia
2
Department of Chemistry, Faculty of Science, Cairo University, Cairo 12613, Egypt
3
Chemistry Department, Faculty of Science, Islamic University in Almadinah Almonawara, Almadinah Almonawara 42351, Saudi Arabia
4
Chemistry Department, College of Sciences, Yanbu, Taibah University, Yanbu 30799, Saudi Arabia
*
Author to whom correspondence should be addressed.
Submission received: 17 April 2021 / Revised: 3 May 2021 / Accepted: 13 May 2021 / Published: 15 May 2021

Abstract

:
New laterally OCH3-substituted optical organic Schiff base/ester series, namely 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates, were prepared and characterized with different thermal, mesomorphic, and photoactive techniques. The prepared group constitutes five homologues that differ from each other in the number of carbons in the terminal alkoxy chain (n), which varies between n = 6, 8, 10, 12, and 16 carbons. The laterally protruded methoxy group is attached to the central benzene ring that makes an angle of 120° with the molecular long axis. Molecular structures of all newly prepared homologues were fully elucidated via FT-IR, 1H and 13C NMR spectroscopy. Mesomorphic transitions were determined via differential scanning calorimetry (DSC) and the phases identified by polarized optical microscopy (POM). Independent of the length of the terminal alkoxy chain attached to phenyl ester ring, only a monomorphic nematic (N) phase was observed for all the synthesized compounds. A comparative study was made between the present lateral methoxy-substituted homologues and their corresponding laterally-neat analogues. The results revealed that, depending on the length of the alkoxy chain and the presence or absence of the lateral methoxy group, different mesophases with different thermal stability and temperature ranges were observed. Finally, UV-vis spectra showed that the present nematogenic series possess photoactive properties that are of importance for many applications.

1. Introduction

Optical and mesomorphic characteristics of organic compounds are mainly dependent on their geometries, in which slight modification in the structure is associated with a considerable change in their mesophase behavior. Large numbers of calamitic thermotropic liquid crystalline materials, with rigid cores containing two or more aromatic rings and terminal flexible chains, have been prepared [1,2,3]. Many of these investigations were based on the Schiff bases linkage [4,5,6,7]. It is worth mentioning that the first room-temperature nematogenic material, namely 4-methoxybenzylidene-4′-butylaniline (MBBA), was investigated by Kelker et al. [8].
The incorporation of a high polar, small size lateral group to the main architecture of the molecule influences the physical or thermal properties of the forming mesomorphic material, such as melting temperature, phase transition temperature, morphology, dielectric anisotropy and dipole moment [9,10,11,12,13]. In such cases, the intermolecular separation increases by the addition of the lateral substituent, which broadens the molecular core, leading to the depression of the lateral-interactions [14,15,16]. Furthermore, as the molecule width increases, the stability of both the smectic and nematic phases will be reduced [17]. The small size of the lateral substituent enables its incorporation within the mesomorphic geometrics without being sterically disrupted and, consequently, their mesophases are still observed. It was reported that the mesophase type, stability and range were affected by the protrusion of a lateral-fused ring in the central core of mesogenic part of the molecule [18]. In another work [19], disruption of the mesophase was observed in a laterally nitro substituted series due to the increase of the molecular width, which affected its lateral intermolecular interactions. In addition, the high dipole moment and steric hindrance of the NO2 group disrupted their mesomorphic properties.
On the other hand, the length of the terminal flexible chain plays a significant role in mesomorphic behavior, enhancing the twist-bend nematic and heliconical phases [20]. It has also been reported that the lateral or terminal polar group processes an impact effect on the mesomorphic properties of a large number of Schiff bases/ester systems [21].
The growing interest of optical organic materials in the last few decades is because of their photo-physical properties [22] that lead to industrial applications, such as laser dyes and fluorescent brighteners [23]. Due to their reasonable thermal stabilities, they have been widely used in photonic and electronic applications, such as solar cells, charge-transfer agents and non-linear optical (NLO) materials. Most of them are based on chromophores and constitute one of the highest types of fluorescence sensors [24,25,26], which accounts for its exponentially increased application, such as fluorescent probes [27,28]; they are also broadly used in organic light-emitting diodes [29]. Moreover, the thermal property causes the nematogenic materials to flow and their optical responses, activated by small temperature changes, can be used in sensing and photonics applications [30]. It was found that this effect had essential ramifications for the nematic phases as the optical properties of thermally induced geometrical changes [30].
Recently, in further structure–property relationship studies of calamitic liquid crystals, the molecular biaxiality of the rode-like molecules was increased by introducing a lateral methoxy (–OCH3) group within the molecular skeleton. From this point of view, the first goal of our present work is to synthesize new geometrical structure of laterally substituted three rings Schiff base derivatives, bearing terminal alkoxy chains with different lengths (n), namely, 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates (Figure 1). The lateral OCH3 group is attached to the middle ring. The second goal is to investigate their optical, mesomorphic, and photophysical behaviors as well as to study the impact of the terminal flexible alkoxy chain on the mesomorphic properties of prepared compounds. Finally, a comparison will be made between the present laterally substituted derivatives and their previously reported three rings Schiff base analogues in order to investigate the effect of the introduction of lateral substituent on their mesophase behavior.

2. Experimental

Synthesis

Schiff base and hydrazone derivatives are well known as valuable intermediates in the synthesis of many organic compounds that are used in many applications [31,32,33,34,35,36,37,38,39]. A series of new laterally methoxy Schiff base derivatives 3 and In were formed as the following shown in Scheme 1:
Synthetic method is attached in the supplementary data. The melting points and IR data were determined for the prepared imine 3: Yield: 90.1%; mp 101.1 °C, FTIR (ύ, cm−1): 3400 (OH), 2928, 2832 (CH2 stretching), 1733 (C=O), 1609 (C=N), 1566 (C=C), 1466 (C–OAsym), 1251 (C–OSym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.85–0.90 (t, 3H, CH3(CH2)3CH2CH2O–), 1.29–1.31 (m, 6H, CH3(CH2)3CH2CH2O–), 1.71–1.78 (q, 2H, CH3(CH2)3CH2CH2O–), 3.72 (s, 1H, OH), 3.93 (s, 3H, OCH3), 4.08 (t, 2H, CH3(CH2)3CH2CH2O–), 6.79 (d, 2H, Ar–H), 6.89 (d, 2H, Ar–H), 7.01 (d, 2H, Ar–H), 7.24 (s, 1H, Ar–H), 7.38–7.40 (m, 3H, Ar–H), 7.84 (d, 2H, Ar–H), 9.76 (s, 1H, CH=N). Anal. Calcd. for C20H25NO3 (327.42): C, 73.37; H, 7.70; N, 4.28. Found: C, 73.21; H, 7.59; N, 4.13%.

General Procedure for Synthesis of 4-(4-(hexyloxyphenyl)iminomethyl)-3-Methoxyphenyl 4-Alkoxybenzoates

Synthetic method is attached in the supplementary data. The physical data of products In are listed below:
4-(4-(Hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-hexyloxybenzoate (I6).
Yield: 88.2%; mp 114–116 °C, FTIR (ύ, cm−1): 2929, 2869 (CH2 stretching), 1735 (C=O), 1612 (C=N), 1573 (C=C), 1458 (C–OAsym), 1260 (C–OSym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.86–0.97 (m, 6H, 2xCH3(CH2)3CH2CH2O–), 1.23–1.33 (m, 12H, 2xCH3(CH2)3CH2CH2O–), 1.78–180 (m, 4H, 2xCH3(CH2)3CH2CH2O–), 3.88 (s, 3H, OCH3), 4.01–4.05 (m, 4H, 2xCH3(CH2)3CH2CH2O–), 6.81 (s, 1H, Ar–H), 6.94–6.97 (m, 3H, Ar–H), 7.31–7.33 (d, 2H, Ar–H), 7.51 (d, 2H, Ar–H), 8.12–8.14 (m, 3H, Ar–H), 9.95 (s, 1H, CH=N). 13C-NMR (100 MHz, CDCl3): δ/ppm: 14.0 (2CH3), 18.3, 22.5, 22.6, 25.6, 29.0, 29.6, 31.5, 31.9 (8CH2), 56.1 (OCH3), 68.3, 68.4 (2CH2–O), 109.5, 114.4, 114.5, 117.5, 123.6, 129.3, 132.4, 132.5, 133.8, 141.2, 143.2, 145.0, 145.4, 152.3, 157.4 (Ar-C and C=N), 164.1 (C=O). Anal. Calcd. for C33H41NO5 (531.68): C, 74.55; H, 7.77; N, 2.63. Found: C, 74.46; H, 7.63; N, 2.51%.
4-(4-(Hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-octyloxybenzoate (I8).
Yield: 84.0%; mp 102–104 °C, FTIR (ύ, cm−1): 2921, 2871 (CH2 stretching), 1731 (C=O), 1603 (C=N), 1583 (C=C), 1463 (C–OAsym), 1262 (C–OSym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.86–0.91 (m, 6H, CH3(CH2)3CH2CH2O– and CH3(CH2)5CH2CH2O–), 1.23–1.39 (m, 16H, CH3(CH2)3CH2CH2O– and CH3(CH2)5CH2CH2O–), 1.78–1.82 (m, 4H, CH3(CH2)3CH2CH2O– and CH3(CH2)5CH2CH2O–), 3.88 (s, 3H, OCH3), 4.01–4.05 (m, 4H, CH3(CH2)3CH2CH2O– and CH3(CH2)5CH2CH2O–), 6.79–6.96 (m, 4H, Ar–H), 7.31–7.35 (d, 2H, Ar–H), 7.48–7.51 (d, 2H, Ar–H), 8.09–8.13 (m, 3H, Ar–H), 9.95 (s, 1H, CH=N) ppm. Anal. Calcd. for C35H45NO5 (559.74): C, 75.10; H, 8.10; N, 2.50. Found: C, 75.25; H, 8.03; N, 2.38%.
4-(4-(Hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-decyloxybenzoate (I10).
Yield: 87.5%; mp 1099–101 °C, FTIR (ύ, cm−1): 2936, 2882 (CH2 stretching), 1728 (C=O), 1609 (C=N), 1575 (C=C), 1455 (C–OAsym), 1254 (C–OSym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.84–0.90 (m, 6H, CH3(CH2)3CH2CH2O– and CH3(CH2)7CH2CH2O–), 1.23–1.45 (m, 20H, CH3(CH2)3CH2CH2O– and CH3(CH2)7CH2CH2O–), 1.78–1.80 (m, 4H, CH3(CH2)3CH2CH2O– and CH3(CH2)7CH2CH2O–), 3.87 (s, 3H, OCH3), 3.88–3.91 (t, 2H, CH3(CH2)7CH2CH2O–), 4.01–4.04 (t, 2H, CH3(CH2)3CH2CH2O–), 6.78–6.98 (m, 4H, Ar–H), 7.31–7.33(d, 2H, Ar–H), 7.49–7.51 (d, 2H, Ar–H), 8.10–8.14 (m, 3H, Ar–H), 9.95 (s, 1H, CH = N) ppm. 13C-NMR (100 MHz, CDCl3): δ/ppm: 14.02 (CH3), 22.1, 22.6, 26.1, 28.9, 29.3, 31.2, 31.6 (CH2), 56.7 (OCH3), 68.6 (CH2–O), 114.5, 118.1, 121.3, 124.9, 128.1, 132.3, 133.6, 134.0, 135.9, 137.7, 140.8, 146.3 (Ar–C), 150.7 (C=N), 152.4, 157.6 (Ar–C), 164.0 (C=O) ppm. Anal. Calcd. for C37H49NO5 (587.79): C, 75.60; H, 8.40; N, 2.38. Found: C, 75.49; H, 8.31; N, 2.27%.
4-(4-(Hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-dodecyloxybenzoate (I12).
Yield: 89.2%; mp 95–97 °C, FTIR (ύ, cm−1): 2928, 2871 (CH2 stretching), 1727 (C=O), 1610 (C = N), 1573 (C=C), 1462 (C–OAsym), 1464 (C–OAsym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.69–0.76 (m, 6H, CH3(CH2)3CH2CH2O– and CH3(CH2)9CH2CH2O–), 1.23–1.45 (m, 24H, CH3(CH2)3CH2CH2O– and CH3(CH2)9CH2CH2O–), 1.56–1.67 (m, 4H, CH3(CH2)3CH2CH2O– and CH3(CH2)9CH2CH2O–), 3.72 (s, 3H, OCH3), 3.74–3.78 (t, 2H, CH3(CH2)9CH2CH2O–), 3.86–3.93 (t, 2H, CH3(CH2)3CH2CH2O–), 6.69–6.83 (m, 4H, Ar–H), 7.11–7.37 (m, 4H, Ar–H), 7.95–7.98 (m, 3H, Ar–H), 9.82 (s, 1H, CH=N) ppm. 13C-NMR (100 MHz, CDCl3): δ/ppm: 13.9 (2CH3), 22.1, 22.2, 25.2, 28.6, 28.9, 29.1, 29.2, 31.1, 31.5 (14 CH2), 55.7 (OCH3), 67.9 (2CH2–O), 110.6, 113.9, 114.8, 114.9, 120.3, 123.1, 124.2, 124.3, 129.6, 132.0, 132.1, 145.0, 151.9, 157.2, 158.5 (Ar–C and C=N), 163.5 (C=O) ppm. Anal. Calcd. for C39H53NO5 (615.84): C, 76.06; H, 8.67; N, 2.27. Found: C, 76.00; H, 8.57; N, 2.14%.
4-(4-(Hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-hexadecyloxybenzoate (I16).
Yield: 90.1%; mp 90–92 °C, FTIR (ύ, cm−1): 2923, 2882 (CH2 stretching), 1729 (C=O), 1612 (C=N), 1580 (C=C), 1464 (C–OAsym), 1251 (C–OSym). 1H-NMR (400 MHz, CDCl3): δ/ppm: 0.84–0.90 (m, 6H, CH3(CH2)3CH2CH2O– and CH3(CH2)7CH2CH2O–), 1.24–1.43 (m, 32H, CH3(CH2)3CH2CH2O– and CH3(CH2)13CH2CH2O–), 1.71–1.84 (m, 4H, -CH3(CH2)3CH2CH2O– and CH3(CH2)13CH2CH2O–), 3.87 (s, 3H, OCH3), 3.88–3.90 (t, 2H, CH3(CH2)13CH2CH2O–), 4.00–4.03 (t, 2H, CH3(CH2)3CH2CH2O–), 6.94–6.97 (m, 4H, Ar–H), 7.30–7.32 (d, 2H, Ar–H), 7.49–7.51 (d, 2H, Ar–H), 8.11–8.14 (m, 3H, Ar–H), 9.95 (s, 1H, CH=N) ppm. Anal. Calcd. for C43H61NO5 (671.95): C, 76.86; H, 9.15; N, 2.08. Found: C, 76.93; H, 9.01; N, 1.95%.

3. Results and Discussion

3.1. Optical and Mesophase Studies

The mesomorphic and optical characterizations for the synthesized laterally methoxy-substituted homologous, In, were investigated by differential scanning calorimetry (DSC) and their textures identified by polarized optical microscopy (POM). DSC thermograms of designed compounds (In), through heating and cooling cycles, are shown in Figure 2. It is clearly shown that upon heating, all compounds (Figure 2a) showed two endotherms intrinsic peaks ascribed to the crystal–to–nematic and nematic–to–isotropic liquid transitions. During the cooling cycle, all derivatives were shown to possess only the nematic mesophase (Figure 2b). The POM investigations revealed images which confirm the threads/schlieren nematic mesophase (Figure 3). This means that these materials exhibited the monomorphic enantiotropic property. Transition temperatures, enthalpies, normalized entropies and the nematic temperature ranges, recorded by DSC measurements upon heating, were collected in Table 1. Figure 4 depicts the graphical plot of the transition temperatures of all designed compounds in order to evaluate the effect of the terminal alkoxy chain length (n) on the mesomorphic and optical behaviors of the prepared series, In.
Data illustrated in Table 1 and Figure 4 indicate that the melting transitions (TCr–N) of all of the prepared homologues, independent of the length of the terminal chain, exhibit irregular trends. In general, the melting point is related with the polarizability of the compound as well as the packing of crystalline molecules and other factors. Data also show that all the members of the homologous series are enantiotropic with suitable nematic thermal stability and nematic temperature range. As usual, the stability of the N mesophase decreased with increases to the length of terminal chain (n), [40,41]. That descending trend of the nematic thermal stability with the increase of the flexible-chain length is ascribed to the dilution of the rigid mesogenic core. TN–I transitions were showed to decrease from 149.8 to 119.2 °C as n increased from 6 to 16 carbons. Moreover, the homologue with the shortest chain (I6) exhibited the highest N temperature range (ΔTN = 52.9 °C), while the N phase range for the longest chain (I16) showed the lowest value (ΔTN = 23.9 °C). The decline of ΔTN is not regular but obeys the order of I6 > I12 > I8 > I10 > I16. It was known that the rigidity, aspect ratio, molecular geometry, polarity and polarizability of the molecule were considered essential factors that are responsible for the thermal stability of the observed mesophases. The different contribution extents of these factors lead to differences in the mesophase property. Generally, the thermal stability of a phase of a given liquid crystalline material is increased by influences of the polarity or polarizability of the molecule mesogenic core that is actually affected by the polarity of the attached group, which would consequently affect the polarity of the whole molecule. Furthermore, the kind and stability of the formed phase are mainly dependent upon the dipole moment of the mesogenic group of the molecule which differs according to the position of the polar substituent too [42].
The normalized transition entropy changes, ΔSN–I/R, of the synthesized lateral methoxy derivatives (In) are collected in Table 1. The results revealed that the value of the entropy change was related, independently, to the length of the terminal chains (n). The small values observed for the entropy change can be attributed to the decrease of the length–to–breadth ratio resulting from their lower anisotropy in terms of their molecular geometry and the increment of their molecular biaxiality [42,43]. Moreover, due to the nematic nature of the mesophase, this exhibits the lowest order mesophase.

3.2. Thermal Properties Studies

Thermal stability of the present homologous set (In) was measured by thermogravimetric analysis (TG). A representative example of TG curve, and its corresponding derivative (DTG) of the compounds I6, is depicted in Figure 5. Parameters of thermal degradation, such as Tonset (the start thermal degradation temperature), Tmax (the temperature at which the thermal degradation rate is maximum), and m% (the mass percentage loss), were estimated. As can be seen from Figure 5, the decomposition occurs through two degradation steps (the second step is not seen in (DTG)) depending on the molecular structure. The first step occurs in the temperature range of ≈250–300 °C and starts at 280 °C with the maximum degradation rate (Tmax) at 300 °C, indicating that the material has high thermal stability. The second step occurs between 330 and 400 °C, with the maximum degradation rate of ≈350 °C. For all prepared homologues series, the first main step occurs with maximum degradation rates (Tmax) at 350, 353, 358, 361 and 368 °C for I6, I8, I10, I12 and I16, respectively, indicating that the thermal stability of materials increases with n (See Figures S1–S4 in supplementary data). Results revealed that the investigated compounds exhibited high thermal stabilities up to 350–368 °C, which covers the temperature window of mesophase transition detected thermally and extends over this transition too.

3.3. Effect of the Lateral Methoxy Group on the Mesomorphic Behaviors

Because of its small size, the incorporation of the lateral methoxy group into the mesogenic part of the molecule was accompanied by a slight steric effect. Moreover, since this is an electron donating group, it affected the intermolecular dispersion interactions. In order to analyze the effect of the special orientation and location of the lateral OCH3 group in the middle ring of the molecule on the mesomorphic and thermal stability of the compounds, a comparison was established between the present synthesized lateral methoxy derivatives (In) and the previously documented laterally neat homologues series IIn [44]; graphically represented in Figure 6a for the mesomorphic stability, and their temperature ranges in Figure 6b. As can be seen from Figure 6a, the type and thermal stability of the formed mesophase varied according to the impacted molecular dipole moment and polarizability of the mesogenic moiety, which were dependent upon the orientation and position of the lateral methoxy group. Moreover, the addition of the lateral OCH3 group into the ortho-position with respect to the azomethine linkage of In, was shown to disrupt the smectic A molecular packing and give only N mesophase. The mesomeric interactions of the lateral methoxy group with the carbonyl ester C=O moiety may be prohibited to the extent that the present compounds (In) possess lower thermal stability than the laterally neat homologue (IIn). The mesophase thermal stability was dependent on the enhanced dipole of the mesogenic portion of the molecule, which was mainly dependent on the position of the lateral OCH3 group. In addition to the molecular geometry of the molecule, which varied according to the location of the attached substituent, it also had high effects on the molecular thermal stability. For the nematic range (Figure 6b) the laterally neat series, IIn, showed higher N temperature ranges than the laterally methoxy-substituted analogues, In, in the short terminal chain lengths (n = 6, 8), while for longer terminals (n ≥ 10) the present laterally substituted series showed wider N temperature ranges than the previously investigated laterally neat analogues. The comparison revealed that the mesophase type and stability, as well as the temperature range, depended on the protrusion of the lateral OCH3 group, which was incorporated within the mesogenic portion of the molecule. Furthermore, it was found that the introducing of the methoxy group into the π-system induced a bathochromic shift in the absorption spectrum of the molecule and improved the light absorption characteristics [45].

3.4. Photoactive Studies

UV-vis absorption spectra were measured for the present lateral methoxy-substituted derivatives (In) in dichloromethane solution (C = 1.1 × 10−3 mol L−1) and the results are represented graphically in Figure 7. As can be seen from Figure 7, the length of terminal alkoxy chain enhanced the maximum absorbance of the members of homologous series, In. Figure 7 also showed that the absorption of light within the wavelength range 380–800 nm, and its strong absorption maxima at ~442–453 nm due to the terminal alkoxy carbon chain (n), can be attributed to the π–π* transition of the chromophore within the molecule. Moreover, the intensity of the peak and absorption bands is dependent on the geometry of the molecule that absorbs the light at a given wavelength. The absorption spectra of I16 showed the maximum bands at 453 nm, which is also due to the π–π electronic transitions from the highest occupied molecular orbitals (HOMO) to the lowest unoccupied molecular orbitals (LUMO) [46,47,48,49]. The present nematogenic materials are promising for use in photoactive organic compounds for many applications.

4. Conclusions

We have documented the synthesis of new optical lateral-methoxy Schiff base/ester homologues using different thermal, optical and mesomorphic techniques. Molecular structures were elucidated by FT-IR and NMR spectroscopy. Mesomorphic, optical and thermal behaviors were characterized using DSC, POM, TGA as well as UV-vis spectroscopy. Results revealed that all members of the homologues series exhibited only the monomorphic N phase irrespective of the terminal alkoxy-chain length. On the other hand, their thermal stabilities were shown to be dependent on the length of the terminal chain. A comparative study was made between the present group of homologues and their corresponding laterally-neat Schiff base/ester analogues. The study revealed that the type and thermal mesophase stability were influenced by the incorporated lateral OCH3 group within the mesogenic portion of the molecule. Finally, the proper choice for new optical organic materials, which possess nematic temperature range, good thermal stability and are affected by UV-light, are promising additional factors for industrial applications.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/cryst11050551/s1. Figure S1: TG (a) and DTG (b) curves of the compound I8. Figure S2: TG (a) and DTG (b) curves of the compound I10. Figure S3: TG (a) and DTG (b) curves of the compound I12. Figure S4: TG (a) and DTG (b) curves of the compound I16.

Author Contributions

Conceptualization, L.A.A.-M. and H.A.A.; methodology, L.A.A.-M., L.A.A. and H.A.A.; validation, L.A.A.-M., L.A.A. and S.M.G.; formal analysis, L.A.A.-M. and H.A.A.; investigation, L.A.A.-M. and H.A.A.; resources, L.A.A. and S.M.G.; data curation, H.A.A.; writing—original draft preparation, H.A.A., L.A.A.-M. and S.M.G.; writing—review and editing, L.A.A.-M., L.A.A., H.A.A., S.M.G.; visualization, H.A.A.; supervision, L.A.A.-M. and H.A.A.; project administration, L.A.A.-M. and L.A.A.; funding acquisition, L.A.A.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research project was funded by the Deanship of Scientific Research, Princess Nourah bint Abdulrahman University, through the Program of Research Project Funding After Publication, grant No (PRFA-P-42-8).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This research project was funded by the Deanship of Scientific Research, Princess Nourah bint Abdulrahman University, through the Program of Research Project Funding After Publication, grant No (PRFA-P-42-8).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kelker, H.; Scheurle, B. A Liquid-crystalline(Nematic) Phase with a Particularly Low Solidification Point. Angew. Chem. Int. Ed. 1969, 8, 884–885. [Google Scholar] [CrossRef]
  2. Kato, T.; Uchida, J.; Ichikawa, T.; Sakamoto, T. Functional Liquid Crystals towards the Next Generation of Materials. Angew. Chem. Int. Ed. 2018, 57, 4355–4371. [Google Scholar] [CrossRef]
  3. Sharma, V.S.; Patel, R.B. Design and investigation of calamatic liquid crystals: Schiff base (‒CH˭N), chalcone (‒CO‒CH˭CH‒), and ester (‒COO‒) linkage group contain rigid rod shape with various terminal parts. Mol. Cryst. Liq. Cryst. 2017, 643, 141–158. [Google Scholar] [CrossRef]
  4. Ahmed, N.H.S.; Saad, G.R.; Ahmed, H.A.; Hagar, M. New wide-stability four-ring azo/ester/Schiff base liquid crystals: Synthesis, mesomorphic, photophysical, and DFT approaches. RSC Adv. 2020, 10, 9643–9656. [Google Scholar] [CrossRef] [Green Version]
  5. Al-Mutabagani, L.; Alshabanah, L.A.; Ahmed, H.; Abu Al-Ola, K.; Hagar, M. New Rod-like H-bonded Assembly Systems: Mesomorphic and Geometrical Aspects. Crystals 2020, 10, 795. [Google Scholar] [CrossRef]
  6. Ahmed, H.A.; Hagar, M.; Aljuhani, A. Mesophase behavior of new linear supramolecular hydrogen-bonding complexes. RSC Adv. 2018, 8, 34937–34946. [Google Scholar] [CrossRef] [Green Version]
  7. Ahmed, H.A.; Mansour, E.; Hagar, M. Mesomorphic study and DFT simulation of calamitic Schiff base liquid crystals with electronically different terminal groups and their binary mixtures. Liq. Cryst. 2020, 47, 2292–2304. [Google Scholar] [CrossRef]
  8. Kelker, H.; Scheurle, B. Eine flüssig-kristalline (nematische) Phase mit besonders niedrigem Erstarrungspunkt. Angew. Chem. 1969, 81, 903–904. [Google Scholar] [CrossRef]
  9. Saccone, M.; Kuntze, K.; Ahmed, Z.; Siiskonen, A.; Giese, M.; Priimagi, A. ortho-Fluorination of azophenols increases the mesophase stability of photoresponsive hydrogen-bonded liquid crystals. J. Mater. Chem. C 2018, 6, 9958–9963. [Google Scholar] [CrossRef]
  10. Jessy, P.; Radha, S.; Patel, N. Morphological, optical and dielectric behavior of chiral nematic liquid crystal mixture: Study on effect of different amount of chirality. J. Mol. Liq. 2018, 255, 215–223. [Google Scholar] [CrossRef]
  11. Mishra, R.; Hazarika, J.; Hazarika, A.; Gogoi, B.; Dubey, R.; Bhattacharjee, D.; Singh, K.N.; Alapati, P.R. Dielectric properties of a strongly polar nematic liquid crystal compound doped with gold nanoparticles. Liq. Cryst. 2018, 45, 1661–1671. [Google Scholar] [CrossRef]
  12. Zaki, A.A. Optical measurements of phase transitions in difluorophenylazophenyl benzoate thermotropic liquid crystal with specific orientated fluorine atoms. Phase Transit. 2018, 92, 135–148. [Google Scholar] [CrossRef]
  13. Zaki, A.A.; Ahmed, H.; Hagar, M. Impact of fluorine orientation on the optical properties of difluorophenylazophenyl benzoates liquid crystal. Mater. Chem. Phys. 2018, 216, 316–324. [Google Scholar] [CrossRef]
  14. Naoum, M.M.; Saad, G.R.; Nessim, R.I.; Abdel-Aziz, T.A.; Seliger, H. Effect of molecular structure on the phase behaviour of some liquid crystalline compounds and their binary mixtures II. 4-Hexadecyloxyphenyl arylates and aryl 4-hexadecyloxy benzoates. Liq. Cryst. 1997, 23, 789–795. [Google Scholar] [CrossRef]
  15. Saad, G.R.; Nessim, R.I. Effect of molecular structure on the phase behaviour of some liquid crystalline compounds and their binary mixtures VI[1]. The effect of molecular length. Liq. Cryst. 1999, 26, 629–636. [Google Scholar] [CrossRef]
  16. Naoum, M.; §, S.M.; Ahmed, H. Lateral protrusion and mesophase behaviour in pure and mixed states of model compounds of the type 4-(4′-substituted phenylazo)-2-(or 3-)methyl phenyl-4’-alkoxy benzoates. Liq. Cryst. 2010, 37, 1245–1257. [Google Scholar] [CrossRef]
  17. Luckhurst, G.; Gray, G.W. The molecular Physics of Liquid Crystals; Academic Press: Cambridge, MA, USA, 1979. [Google Scholar]
  18. Ahmed, H.A.; El-Atawy, M.A. Synthesis, mesomorphic and geometrical approaches of new non-symmetrical system based on central naphthalene moiety. Liq. Cryst. 2021. [Google Scholar] [CrossRef]
  19. El-Atawy, M.A.; Naoum, M.M.; Al-Zahrani, S.A.; Ahmed, H.A. New Nitro-Laterally Substituted Azomethine Deriva-tives; Synthesis, Mesomorphic and Computational Characterizations. Molecules 2021, 26, 1927. [Google Scholar] [CrossRef]
  20. Takezoe, H.; Takanishi, Y. Bent-Core Liquid Crystals: Their Mysterious and Attractive World. Jpn. J. Appl. Phys. 2006, 45, 597–625. [Google Scholar] [CrossRef]
  21. Ahmed, H.A.; Hagar, M.; El-Sayed, T.H.; Alnoman, R.B. Schiff base/ester liquid crystals with different lateral substituents: Mesophase behaviour and DFT calculations. Liq. Cryst. 2019, 46, 1–11. [Google Scholar] [CrossRef]
  22. Naveen, K.; Udayabhanu, A.; Mahadevan, K.M.; Nagaraju, G. Solvent free and green synthesis of efficient solvochrom-ism based coumarin moieties for quick visualization of LFPs and OLEDs applications. J. Mol. Struct. 2021, 1223, 129208. [Google Scholar]
  23. Ghasem, B.; Shahram, M.; Saeed, S. A novel method for the synthesis of coumarin laser dyes derived from 3-(1H-benzoimidazol-2-yl) coumarin-2-one under microwave irradiation. Arab. J. Chem. 2014, 7, 972. [Google Scholar]
  24. Aulsebrook, M.L.; Graham, B.; Grace, M.R.; Tuck, K.L. Coumarin-based fluorescent sensors for zinc (II) and hypo-chlorite. Supramol. Chem. 2015, 27, 798. [Google Scholar] [CrossRef]
  25. Huang, X.; Dong, Y.; Huang, Q.; Cheng, Y. Hydrogen bond induced fluorescence recovery of coumarin-based sensor system. Tetrahedron Lett. 2013, 54, 3822–3825. [Google Scholar] [CrossRef] [Green Version]
  26. Khan, R.I.; Pitchumani, K. beta-Cyclodextrin included coumarin derivatives as selective fluorescent sensors for Cu2+ ions in HeLa cells. Royal Soc. Chem. Adv. 2016, 6, 20269. [Google Scholar]
  27. Breul, A.M.; Hager, M.D.; Schubert, U.S. Fluorescent monomers as building blocks for dye labeled polymers: Synthesis and application in energy conversion, biolabeling and sensors. Chem. Soc. Rev. 2013, 42, 5366–5407. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, X.; Feng, H.; Yan, M.; Guo, H.; Yang, F. The novel rufigallol-based liquid crystals with cholesterol units: Synthesis, mesomorphic and photophysical properties. Liq. Cryst. 2019, 46, 787–796. [Google Scholar] [CrossRef]
  29. Lin, L.; Qin, W.; Cheng, B.; Guo, H.; Yang, F. The influence of multiple alkyl chains on mesomorphic and photophysical properties of diphenylacrylonitrile liquid crystals. Liq. Cryst. 2019, 47, 967–976. [Google Scholar] [CrossRef]
  30. Young-Ki, K.; Bohdan, S.; Oleg, D. Lavrentovich. Molecular reorientation of a nematic liquid crystal by thermal ex-pansion. Nat. Commun. 2012, 3, 1113. [Google Scholar]
  31. Sayed, R.; Abd El-lateef, H.M.; Gomha, S.M. L-Proline catalyzed green synthesis and anticancer evaluation of novel bi-oactive benzil bis-hydrazones under grinding technique. Green Chem. Lett. Rev. 2021, 14, 179–188. [Google Scholar] [CrossRef]
  32. Gomha, S.M.; A Abdelhady, H.; Hassain, D.Z.; Abdelmonsef, A.H.; El-Naggar, M.; Elaasser, M.M.; Mahmoud, H.K. Thiazole-Based Thiosemicarbazones: Synthesis, Cytotoxicity Evaluation and Molecular Docking Study. Drug Des. Dev. Ther. 2021, ume 15, 659–677. [Google Scholar] [CrossRef]
  33. Gomha, S.M.; Muhammad, Z.A.; Abdel-Aziz, H.M.; Matar, I.K.; El-Sayed, A.A. Green synthesis, molecular docking and anticancer activity of novel 1,4-dihydropyridine-3,5-Dicarbohydrazones under grind-stone chemistry. Green Chem. Lett. Rev. 2020, 13, 6–17. [Google Scholar] [CrossRef]
  34. Gomha, S.M.; Edrees, M.M.; Muhammad, Z.A.; Kheder, N.A.; Melha, S.A.-; Saad, A.M. Synthesis, Characterization, and Antimicrobial Evaluation of Some New 1,4-Dihydropyridines-1,2,4-Triazole Hybrid Compounds. Polycycl. Aromat. Compd. 2020, 1–13. [Google Scholar] [CrossRef]
  35. Gomha, S.M.; Riyadh, S.M. Synthesis under Microwave Irradiation of [1,2,4]Triazolo[3,4-b] [1,3,4]thiadiazoles and Other Diazoles Bearing Indole Moieties and Their Antimicrobial Evaluation. Molecules 2011, 16, 8244–8256. [Google Scholar] [CrossRef] [PubMed]
  36. Abu-Melha, S.; Edrees, M.M.; Riyadh, S.M.; Abdelaziz, M.R.; Elfiky, A.A.; Gomha, S.M. Clean grinding technique: A facile synthesis and in silico antiviral activity of hydrazones, pyrazoles, and pyrazines bearing thiazole moiety against SARS-CoV-2 main protease (Mpro). Molecules 2020, 25, 4565. [Google Scholar] [CrossRef] [PubMed]
  37. Sayed, A.R.; Gomha, S.M.; A Taher, E.; A Muhammad, Z.; El-Seedi, H.R.; Gaber, H.M.; Ahmed, M.M. One-Pot Synthesis of Novel Thiazoles as Potential Anti-Cancer Agents. Drug Des. Dev. Ther. 2020, ume 14, 1363–1375. [Google Scholar] [CrossRef] [Green Version]
  38. Ouf, S.A.; Gomha, S.M.; Ewies, M.M.; Ouf, A.S.; Sharawy, I.A.A. Antidermatophytic activity of some newly synthe-sized arylhydrazonothiazoles conjugated with monoclonal antibody. Sci. Rep. 2020, 10, 20863. [Google Scholar] [CrossRef]
  39. Rashdan, H.R.M.; Abdelmonsef, A.H.; Shehadi, I.A.; Gomha, S.M.; Soliman, A.M.M.; Mahmoud, H.K. Synthesis, mo-lecular docking screening and anti-proliferative potency evaluation of some new imidazo[2,1-b]thiazole linked thiadi-azole conjugates. Molecules 2020, 25, 4997. [Google Scholar] [CrossRef] [PubMed]
  40. Imrie, C.T.; Taylor, L. The preparation and properties of low molar mass liquid crystals possessing lateral alkyl chains. Liq. Cryst. 1989, 6, 1–10. [Google Scholar] [CrossRef]
  41. Imrie, C.T. Non-symmetric liquid crystal dimers: How to make molecules intercalate. Liq. Cryst. 2006, 33, 1449–1485. [Google Scholar] [CrossRef]
  42. Attard, G.; Imrie, C.; Karasz, F. Low molar mass liquidcrystalline glasses: Preparation and properties of the. al-pha.-(4-cyanobiphenyl-4ʹ-oxy)-. omega.-(1-pyrenimin ebenzylidene-4ʹ-oxy) alkanes. Chem. Mater. 1992, 4, 1246–1253. [Google Scholar] [CrossRef]
  43. Attard, G.S.; Imrie, C.T. Liquid-crystalline and glass-forming dimers derived from 1-aminopyrene. Liq. Cryst. 1992, 11, 785–789. [Google Scholar] [CrossRef]
  44. Ahmed, H.A.; Hagar, M.; Saad, G.R. Impact of the proportionation of dialkoxy chain length on the mesophase be-haviour of Schiff base/ester liquid crystals; experimental and theoretical study. Liq. Cryst. 2019, 46, 1611–1620. [Google Scholar] [CrossRef]
  45. Avhad, K.; Jadhav, M.; Patil, D.; Chowdhury, T.H.; Islam, A.; Bedja, I.; Sekar, N. Rhodanine-3-acetic acid containing D-π-A push-pull chromophores: Effect of methoxy group on the performance of dye-sensitized solar cells. Org. Electron. 2019, 65, 386–393. [Google Scholar] [CrossRef]
  46. Yang, M.; Liu, Z.; Li, X.; Yuan, Y.; Zhang, H. Influence of flexible spacer length on self-organization behaviors and photophysical properties of hemiphasmidic liquid crystalline polymers containing cyanostilbene. Eur. Polym. J. 2020, 123, 109459. [Google Scholar] [CrossRef]
  47. Ahmed, H.A.; Khushaim, M.S. Nematic Phase Induced from Symmetrical Supramolecular H-Bonded Systems Based on Flexible Acid Core. Cryst. 2020, 10, 801. [Google Scholar] [CrossRef]
  48. Ahmed, H.; Khushaim, M.S. Nematogenic Laterally Substituted Supramolecular H-Bonded Complexes Based on Flexible Core. Cryst. 2020, 10, 878. [Google Scholar] [CrossRef]
  49. Al-Mutabagani, L.A.; Alshabanah, L.A.; Naoum, M.M.; Hagar, M.; Ahmed, H.A. Experimental and Computational Approaches of Newly Polymorphic Supramolecular H-Bonded Liquid Crystal Complexes. Front. Chem. 2020, 8, 930. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates, In.
Figure 1. Chemical structure of 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates, In.
Crystals 11 00551 g001
Scheme 1. Synthesis of 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates, In.
Scheme 1. Synthesis of 4-(4-(hexyloxyphenyl)iminomethyl)-3-methoxyphenyl 4-alkoxybenzoates, In.
Crystals 11 00551 sch001
Figure 2. The recorded DSC thermograms of series In at a rate of ± 10 °C/min (a) the second heating and (b) cooling scans.
Figure 2. The recorded DSC thermograms of series In at a rate of ± 10 °C/min (a) the second heating and (b) cooling scans.
Crystals 11 00551 g002
Figure 3. Textures observed upon heating under POM of nematic mesophase for (a) compound I6 at 141.0 °C and (b) compound I16 at 116.0 °C.
Figure 3. Textures observed upon heating under POM of nematic mesophase for (a) compound I6 at 141.0 °C and (b) compound I16 at 116.0 °C.
Crystals 11 00551 g003
Figure 4. Impact of terminal chain length (n) on the mesomorphic behavior of the present homologous In series.
Figure 4. Impact of terminal chain length (n) on the mesomorphic behavior of the present homologous In series.
Crystals 11 00551 g004
Figure 5. TG (a) and DTG (b) curves of the compound I6.
Figure 5. TG (a) and DTG (b) curves of the compound I6.
Crystals 11 00551 g005
Figure 6. Comparison of (a) Tc and (b) ΔTc of homologues series In and IIn.
Figure 6. Comparison of (a) Tc and (b) ΔTc of homologues series In and IIn.
Crystals 11 00551 g006
Figure 7. UV-vis spectra of present series, In, in CH2Cl2 at 25 °C.
Figure 7. UV-vis spectra of present series, In, in CH2Cl2 at 25 °C.
Crystals 11 00551 g007
Table 1. Mesophase transition temperatures (°C), enthalpy ΔH, kJ/mole, and normalized entropy of transition ΔS/R for compounds In.
Table 1. Mesophase transition temperatures (°C), enthalpy ΔH, kJ/mole, and normalized entropy of transition ΔS/R for compounds In.
Comp.TCr–NΔHCr–NTN–IΔHN–IΔTNΔSN–I/R
I696.944.9149.81.552.90.4
I896.351.5138.32.842.00.8
I1089.949.3129.12.239.20.7
I1273.844.3122.11.048.30.3
I1695.354.7119.23.023.90.9
Cr-N = solid to the nematic phase; N–I = nematic to the isotropic liquid phase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Al-Mutabagani, L.A.; Alshabanah, L.A.; Gomha, S.M.; Ahmed, H.A. Synthesis, Thermal and Optical Characterizations of New Lateral Organic Systems. Crystals 2021, 11, 551. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050551

AMA Style

Al-Mutabagani LA, Alshabanah LA, Gomha SM, Ahmed HA. Synthesis, Thermal and Optical Characterizations of New Lateral Organic Systems. Crystals. 2021; 11(5):551. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050551

Chicago/Turabian Style

Al-Mutabagani, Laila Ahmed., Latifah A. Alshabanah, Sobhi M. Gomha, and Hoda A. Ahmed. 2021. "Synthesis, Thermal and Optical Characterizations of New Lateral Organic Systems" Crystals 11, no. 5: 551. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11050551

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