Next Article in Journal
Analysis of Floquet Waves in Periodic Multilayered Isotropic Media with the Method of Reverberation-Ray Matrix
Previous Article in Journal
Elastic Metagratings with Simultaneous Modulation of Reflected and Transmitted Waves
Previous Article in Special Issue
Self-Localized Liquid Crystal Micro-Droplet Arrays on Chemically Patterned Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

2-Pyridinyl-Terminated Iminobenzoate: Type and Orientation of Mesogenic Core Effect, Geometrical DFT Investigation

1
Department of Chemistry, College of Science, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Department of Chemistry, Faculty of Science, Alexandria University, Alexandria 21321, Egypt
3
Smart-Health Initiative (SHI) and Red Sea Research Center (RSRC), Division of Biological and Environmental Sciences and Engineering (BESE), King Abdullah University of Science and Technology (KAUST), P.O. Box 4700, Thuwal 23955-6900, Saudi Arabia
4
Core Labs, King Abdullah University of Science and Technology, P.O. Box 4700, Thuwal 23955-6900, Saudi Arabia
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(7), 902; https://doi.org/10.3390/cryst12070902
Submission received: 24 May 2022 / Revised: 17 June 2022 / Accepted: 17 June 2022 / Published: 24 June 2022
(This article belongs to the Special Issue Self-Assembly in Liquid Crystalline Materials)

Abstract

:
A new liquid crystal series of pyridin-2-yl 4-[4-(alkylphenyl)iminomethyl]benzoate was synthesized and characterized for their mesomorphic behavior. These compounds contain Schiff base and carboxylate ester mesogenic cores, in addition to terminal alkyl chains with a different number of carbons. The structures were confirmed via FT-IR, and 1H NMR spectroscopy. The phase transitions were studied by differential thermal analysis (DSC) and the mesophase types were identified by polarized optical microscopy (POM). A comparative study was performed between the synthesized compounds and previously reported compounds. Density functional theory (DFT) calculations were included in the study to compute the dipole moment and the polarizability, as well as the frontier molecular orbitals and the charge distribution mapping, which impact the terminal and lateral interactions of the compounds. The theoretical results were discussed to confirm the experimental data and explain the mesomorphic behavior of the compounds. Finally, the energy gap, global softness, and chemical hardness were calculated to determine the suitability of the liquid crystalline compounds to be employed in applications.

1. Introduction

Liquid crystal (LC) represents a phase formed in certain compounds that intermediates solid and liquid states of matter. Therefore, the compounds that can form LC have two melting points: the one at which the LC is formed, which appears as a cloudy liquid, and the other one at which the clear liquid state is formed [1]. Liquid crystals are also known as mesophases. This phase is characterized by having unique properties, as it possesses properties from both solids and liquids. It can flow as a liquid but with certain crystalline physical properties [2], such as optical, electrical, and magnetic properties, in addition to the molecule’s arrangement in spatial directions [3]. In 1888, the botanist Friedrich Reinitzer worked on the cholesteryl benzoate, which is a crystal cholesterol derivative extracted from carrot; he found that it has two melting points. He explained that the cholesteryl benzoate crystals lose their rigidity and turn into a milky fluid at the first melting point, which is 145.5 °C, and this fluid finally turns into a clear liquid at the second melting point, which is 178.5 °C. He also discovered different optical properties as violet and blue colors appeared and then disappeared while cooling these crystals. This strange phase is then called liquid crystal [4].
LCs are divided into three classes, which are lyotropic, thermotropic, and polymeric. Lyotropic LCs are formed depending on the change in concentration of material in a certain solvent. One of the most commonly observed systems is the LC formed by water and amphiphilic molecules, forming micelles, such as soap, detergents, and lipids [5]. Polymeric LCs are high-molecular-weight liquid crystals with a degree of flexibility. They are classified according to the molecular structure arrangement of mesogenic monomers [2]. Thermotropic LCs are formed depending on the temperature change. This class is widely observed in televisions, laptops, and mobile phone displays. Thermotropic LCs have many phases, depending on the molecular arrangement and symmetry, which are divided into smectic, nematic, and cholesteric [3]. They are usually organic compounds similar to cholesteryl nonanoate and MBBA [6]. LCs that are either rod-like or disk-like in shape can undergo a nematic phase. In this phase, molecules are ordered in a long-range orientation parallel to each other. They are aligned in a certain direction, defined according to the medium director (the principal axis). Moreover, they lose their positional order upon melting [3,6]. There is another similar phase to the nematic, which is the cholesteric phase. It also has a long-range orientation, with no positional order, unlike the nematic phase. The director of the cholesteric phase varies periodically in solution [3,7]. On the other hand, the smectic phase is stratified, and these well-defined layers can slide over each other, unlike the nematic and cholesteric phases. The smectic phase has no long-range orientation but preserves the positional order, and according to this order, the smectic phase has many types. Smectic A is a phase through which a molecule is positioned perpendicular to the other layers. Smectic B is a phase that exhibits a hexagonal crystalline order within the layers. Smectic C is the reverse of smectic A as the molecules are not positioned perpendicularly [3].
Most thermotropic liquid crystals are mainly composed of an aromatic part that consists of two benzene rings or more and an aliphatic part that is flexible alkyl chains. The rings of the aromatic part are connected by linkage groups such as a Schiff base (-CH=N-). Therefore, it provides a stepped core structure and preserves the molecular linearity and develops high stability for the compound that enables mesophase formation [8,9]. Compounds containing Schiff bases have many applications; they act as systems with high thermal stability, corrosion inhibitors, and have some biological activity. In addition, they have been proven to have photo-efficiency, with the wavelength according to their structure [9,10]. Schiff base LCs have gained attention since the discovery of 4-methoxybenzylidene-4′-butylaniline, which is characterized by being in the nematic phase at room temperature [11,12]. Schiff base LCs that have low molecular weight or exhibit a twist-bend nematic phase have been synthesized and investigated [12]. Compounds containing ester group linking between two benzene rings obtain the characters of double bonds. The mesophase becomes more stable when the mutual conjugation increases between the ester carbonyl and the substituent [10]. Several studies have used LC compounds containing a Schiff base ester mesogenic core because of their remarkable properties and wide temperature range [8]. Schiff base ester LCs containing two or three aromatic rings were noticed to have interesting optical behavior. Moreover, LCs containing only one mesogenic core differ in behavior from other LCs having two or more mesogenic cores [13]. Furthermore, as the terminal substituent length increases, the molecules tend to arrange in a parallel orientation [8]. Moreover, to create new thermotropic LCs having phase transitions, it is important to select a proper mesogenic moiety, terminal groups, and flexible wings. Therefore, to design a new material, computational analysis is considered an excellent and helpful tool to do so. LCs must be investigated for their properties, such as MO energies, the frontier molecular orbitals’ energy difference, and the geometry of molecules [10].
Therefore, the aim of the current study was the synthesis of new liquid crystalline compounds A, B, and C, which were prepared from the reaction between 4-formylbenzoic acid and 4-alkylaniline. Their structure was confirmed with IR and NMR spectroscopy, and their mesomorphic behavior was inspected through POM and DSC. Then, a comparative study was performed of the newly synthesized compounds with previously reported ones, D [14], E [15], F [16], G, H [17], and I [18], to investigate the effect of the terminal chain—whether an alkyl chain or alkoxy chain—the effect of the mesogenic core—whether Schiff base/ester or azo/ester—and whether the aromatic ring is a benzene ring or heterocyclic ring. Figure 1 illustrates the structures of these compounds.

2. Materials and Methods

2.1. Materials

In this study, 4-Formylbenzonitrile (95%) and N,N′-dicyclohexylcarbodiimide (DCC) (99%) were purchased from Sigma Aldrich (Shanghai, China); 4-octyl aniline (97%) and 4-dodecylaniline (97%) were purchased from Sigma Aldrich (Verona, WI, USA); 4-tetradecylaniline (97%) was obtained from Sigma Aldrich (Delhi, India); 2-hydroxy pyridine (98%), 4-dimethylaminopyridine (DMAP) (98%), dichloromethane ( 99.8), HCl (38%), and ethanol (95%) were purchased from Sigma Aldrich (Homburg, Germany).

2.2. Synthesis

The three compounds were prepared according to the following Scheme 1.

2.2.1. Synthesis of 4-Formylbenzoic Acid

First, 2 g of 4-formylbenzonitrile and 10 mL of hydrochloric acid in water were refluxed for four hours. Later, this mixture was left to cool and then the product (4-formylbenzoic acid) was separated by filtration.

2.2.2. Synthesis of 4(((4-Alkylphenyl)imino)methyl)benzoic Acid (X)

Equimolar amounts of 4-formylbenzoic acid (75 mg, 0.5 mmole) and 4-alkylaniline (0.5 mmole) in 15 mL ethanol were refluxed for four hours. Afterward, the mixture was left to cool and then the product was separated by filtration.

2.2.3. Synthesis of Pyridin-2-yl 4(((4-alkylphenyl)imino)methyl)benzoate (A, B, and C)

First, 4(((4-alkylphenyl)imino)methyl)benzoic acid (1 mmole) and 2-hydroxy pyridine (1.05 mmole) were dissolved in dichloromethane (10 mL). Then, 4 mmole of DCC and traces of DMAP (act as a catalyst) were added to the mixture. This mixture was left and stirred for 72 h at normal room temperature. After stirring, the reaction by-product, dicyclohexylurea, was filtered out. Then, the filtrate was evaporated and the obtained product was recrystallized from ethanol. The progress of the reaction was checked using thin-layer chromatography (TLC) on silica gel 60 F254 E-Merck (layer thickness is equal to 0.2 mm) plates with hexane-ethyl acetate (1:1 v/v) mobile phase. The spots were observed under a UV lamp at λ = 254 nm.
Pyridin-2-yl 4-[4-(octylphenyl)iminomethyl]benzoate (A)
Yield: 67%, FTIR (ύ, cm−1): 2946–2849 (CH2 stretching), 1733 (C=O), 1605 (C=N), 1593 (C=C), 1460 (C-OAsym), 1271 (C-OSym). 1H NMR (400 MHz, CDCl3): δ/ppm: 0.90 (t, 3 H, J = 6.0 Hz, CH3(CH2)6CH2), 1.31–1.65 (m, 12 H, CH3(CH2)6CH2), 2.63 (t, 2 H, J = 6.0 Hz, CH3(CH2)6CH2), 7.19 (d, 2 H, J = 8 Hz, Ar-H), 7.22 (d, 2 H, J = 8.0 Hz, Ar-H), 7.37 (d, 2 H, J = 8.0 Hz, Ar-H), 7.51 (dd, 1 H, J1 = 8.0, J2 = 4.0 Hz, Py-H), 8.02 (d, 2 H, J = 8.0 Hz, Ar-H), 8.48–8.52 (m, 2 H, Py-H), 8.90 (dd, 1 H, J1 = 8.0, J2 = 4.0 Hz, Py-H), 9.44 (s, 1 H, CH=N). 1H NMR (100 MHz, CDCl3): δ/ppm: 14.12, 22.68, 24.94, 29.53, 29.61, 31.43, 31.78, 33.89, 35.52, 120.83, 121.98, 123.54, 125.37, 129.14, 130.05, 134.54, 137.66, 141.18, 149.31, 151.30, 152.65, 154.19, 158.07, 163.60.
Pyridin-2-yl 4-[4-(dodecylphenyl)iminomethyl]benzoate (B)
Yield: 83%, FTIR (ύ, cm−1): 2946–2849 (CH2 stretching), 1730 (C=O), 1601 (C=N), 1591 (C=C), 1462 (C-OAsym), 1269 (C-OSym). 1H NMR (400 MHz, CDCl3): δ/ppm: 0.89 (t, 3 H, J = 6.0 Hz, CH3(CH2)10CH2), 1.29–1.67 (m, 20 H, CH3(CH2)10CH2), 2.62 (t, 2 H, J = 6.0 Hz, CH3(CH2)10CH2), 7.21 (d, 2 H, J = 8 Hz, Ar-H), 7.23 (d, 2 H, J = 8.0 Hz, Ar-H), 7.38 (d, 2 H, J = 8.0 Hz, Ar-H), 7.51 (dd, 1 H, J1 = 8.0, J2 = 4.0 Hz, Py-H), 8.01 (d, 2 H, J = 8.0 Hz, Ar-H), 8.48–8.52 (m, 2 H, Py-H), 8.90 (dd, 1 H, J1 = 8.0, J2 = 4 Hz, Py-H), 9.44 (s, 1 H, CH=N). 1H NMR (100 MHz, CDCl3): δ/ppm: 14.12, 22.71, 24.93, 25.61, 29.54, 29.77, 31.56, 31.93, 33.94, 35.52, 120.80, 121.99, 123.53, 125.38, 129.14, 130.03, 134.53, 137.68, 141.19, 149.31, 151.30, 152.62, 154.18, 158.06, 163.59.
Pyridin-2-yl 4-[4-(tetradecylphenyl)iminomethyl]benzoate (C)
Yield: 85%, FTIR (ύ, cm−1): 2946–2849 (CH2 stretching), 1732 (C=O), 1601 (C=N), 1592 (C=C), 1462 (C-OAsym), 1269 (C-OSym). 1H NMR (400 MHz, CDCl3): δ/ppm: 0.90 (t, 3 H, J = 6.0 Hz, CH3(CH2)12CH2), 1.28–1.67 (m, 24 H, CH3(CH2)12CH2), 2.6 (t, 2 H, J = 6.0 Hz, CH3(CH2)12CH2), 7.20 (d, 2 H, J = 8 Hz, Ar-H), 7.23 (d, 2 H, J = 8.0 Hz, Ar-H), 7.37 (d, 2 H, J = 8.0 Hz, Ar-H), 7.51 (dd, 1 H, J1 = 8.0, J2 = 4.0 Hz, Py-H), 8.01 (d, 2 H, J = 8.0 Hz, Ar-H), 8.48–8.52 (m, 2 H, Py-H), 8.90 (dd, 1 H, J1 = 8.0, J2 = 4.0 Hz, Py-H), 9.44 (s, 1 H, CH=N). 1H NMR (100 MHz, CDCl3): δ/ppm: 14.12, 22.69, 24.93, 25.61, 29.30, 29.54, 29.66, 29.78, 31.54, 31.93, 33.94, 35.52, 120.80, 121.99, 123.53, 125.39, 129.15, 130.02, 134.53, 137.66, 141.19, 149.31, 151.31, 152.62, 154.19, 158.07, 163.59. The 1HNMR and C13 NMR of Pyridin-2-yl 4-[4-(tetradecylphenyl)iminomethyl]benzoate (C) are shown in Figures S1 and S2 (supplementary materials).
The phase transitions in the materials were determined via differential scanning calorimetry (DSC), DSC-60A, Shimadzu, Kyoto, Japan. Here, 2–3 mg specimens were encased in aluminum pans and heated or cooled in a dry nitrogen environment. The temperature was measured at a rate of 10.0 °C/min. Under an inert nitrogen gas atmosphere, samples were heated from room temperature to 200 °C and then cooled back to room temperature at the same rate. The endothermic peak minima of enthalpy in the heating curves were used to calculate the phase transition temperatures. Temperature monitoring accuracy was better than 1.0 °C.
A polarized optical microscope (POM, Wild, Germany) coupled to a Mettler FP82HT hot stage was used to check transition temperatures and identify phases for the prepared compounds.

3. Results and Discussion

3.1. Mesomorphic Behavior

The phase transitions of all synthesized compounds were studied by differential thermal analysis (DSC), and the mesophase types were recognized by observing their textures through polarized optical microscopy (POM). DSC was used to measure the phase transition temperatures (°C), their enthalpy of transition ΔH (kJ/mol), and their normalized entropy of transition ΔS/R. The results that were derived from DSC measurements and verified by POM are given in Table 1. A representative example demonstrating the DSC thermogram of heating/cooling processes for compound A is shown in Figure 2. The results of the POM showed enantiotropic nematic mesophases for all prepared compounds. The existence of sharp peaks in the DSC thermogram indicates the phase transition between the crystal phase and the liquid crystalline phase. Moreover, the small peaks signify the transition from the liquid crystalline phase to the isotropic liquid. The observed mesophase textures for compounds B and C are provided as examples in Figure 3.
A comparative study of the prepared compounds with the reportedly structurally similar compounds E, F, G, H, and I was carried out to explain the effect of the mesogenic core and its orientation on the mesomorphic behavior. It is clear from Figure 4 that compound D is non-mesomorphic. It is worth noticing that the mesophase ranges of compounds A (15.92 °C), B (13.9 °C), and C (13.28 °C) are decreased by increasing the length of the terminal alkyl chain. This could be attributed to alkyl group disorder. The longer alkyl group may result in insufficient space between the layers for the terminal chains to accommodate, causing the layers to be parted. This could reduce the overlap of aromatic rings and the interactions between distinct mesogenic groups, which are essential for the development of the nematic mesophase [19]. On the other hand, the absence of the mesophase for compound D could be due to the presence of the electronegative nitrogen heteroatom in the middle ring of the compound. Conversely, its analogue E compound, which has the nitrogen heteroatom in the terminal outer ring, exhibits a smectic A mesophase with a wide range of thermal stability (25.4 °C). It could be explained in terms of the higher degree of charge separation in compound E compared with its analogue D. The presence of the terminal pyridyl ring affords a higher packing pattern compared to that of the compound D to show a smectic mesophase. It is also observed that compounds F, H, and I possess nematic phases and their mesophase ranges are increased in order H (14.17 °C) < F (45.6 °C) < I (46 °C). Moreover, compound G has a high smectic A mesomorphic range (20 °C). On the other hand, it is worth noting the change in the nematic mesophase range by comparing all compounds discussed. For example, F and I have a broad nematic range, while A, B, and C have a short nematic mesophase range. It is clear that longitudinal diffusion is not favored for long alkyl chains.

3.2. DFT Calculations

3.2.1. Geometrical Structure

We utilized theoretical calculations to complement the experimental findings regarding the investigated compounds. The calculations were executed by Gaussian 09 Revision D.01 software [20] and performed using the DFT/B3LYP method at basis set 6–311G (d, p). The geometrical structures of compounds were subjected to an optimization process to develop a new lowest-energy geometrical structure, which is referred to as the convergence, by minimizing the energy of conformations regarding all geometrical parameters, without exerting any molecular symmetry constraints. All of the optimized compounds were stable due to lacking the imaginary frequency. The structures were drawn by GaussView 5.0.8 [21] and the optimized structures were captured by Chemcraft 1.8 [22]. Additionally, the optimized structures were subjected to a frequency process in the same level of theory and basis set to calculate some significant parameters. The optimized molecular structures are shown in Figure 5 and the twist angles of compounds’ rings are tabulated in Table 2. It can be noticed that the newly synthesized compounds A, B, and C are non-coplanar. Moreover, the length of the terminal alkyl chain has an insignificant effect on the twist angles of the rings of the compound, but it is worth mentioning that the D and E compounds are coplanar, which can be attributed to the presence of the azo group (-N=N-), which enhances the planarity of the compounds. This could be one of the factors responsible for the formation of the smectic A mesophase in the E compound, as the planarity increases the degree of molecular packing, which creates a more ordered smectic mesophase. Moreover, the presence of an alkoxy terminal chain in compound H may slightly enhance the planarity compared to the alkyl chain in the A compound.
The aspect ratio is another structural parameter that may explain the behavior of the mesophases as it can indicate the collision diameter of the compounds, so the intermolecular interaction increases when raising the aspect ratio values. Thus, we correlate in Table 3 and Figure 6 the mesomorphic parameters of the investigated compounds with aspect ratio values and the ratio of the width to the height.
It is observed that the nematic mesophase ranges of the A, B, and C compounds are decreased upon increasing the aspect ratio. However, compound E has the highest aspect ratio and also possesses the more-ordered smectic A mesophase with a wide range (25.4 °C). The high aspect ratio could increase the space filling of the liquid crystal compounds, consequently enhancing terminal and lateral interactions, thus developing a smectic mesophase. On the other hand, the mesophase ranges decrease for the F, G, H, and I compounds upon the elevation of the aspect ratio. This finding may be explained in terms of the combined impact of various factors, such as the polarizability and dipole moment, that could contribute together to show the developed mesophase.
It is well known that the increment in the dipole moment has a major effect on mesomorphic behavior [23]. It may be rationalized in terms of the degree of molecular packing as it could permit π–π stacking interactions and other particular interactions, such as quadrupolar. Consequently, these may affect the mesophase stability and mesophase range of the liquid crystalline compounds. Thus, the dipole moments were estimated and results were correlated to the mesomorphic parameters of the investigated compounds, and they are tabulated in Table 4 and graphically represented in Figure 7. It is worth noticing that the increase in the dipole moment may cause an increment in the temperature at which the nematic phase is formed for the A, B, and C compounds. Compound A (TN = 134.16 °C) has a higher dipole moment (1.52 Debye) than compound C (TN = 133.60 °C), which has a dipole moment value equal to 1.43 Debye. Furthermore, compound B (TN = 132.55 °C) has the lowest dipole moment value (1.39 Debye) among the newly synthesized compounds. Through the same approach, it was noticed that the polar alkoxy terminal chain increases the dipole moment value to be higher than that of the non-polar alkyl chain. This slight change in the dipole moment may influence the competitive terminal and lateral intermolecular interactions of the compounds. Additionally, the formation of the smectic mesophase with a wide range (ΔTSmA = 25.4 °C) in compound E might be caused by the high dipole moment (6.43 Debye), which allows a high degree of molecular ordering. Furthermore, the dipole moment increment resulted in an increment in the mesophase range for the F, G, H, and I compounds. These findings can be explained in terms of the high degree of molecular packing in the lattice, which may enhance the terminal intermolecular interactions.
The polarizability of the molecules could be highly influenced by the dimension and the electronic nature of the mesogenic core. Moreover, an increase in the polarizability of the compounds’ mesogenic cores may influence the mesophase ranges and stabilities. Thus, the polarizability of the investigated compounds was also estimated, and the values are included in Table 4 and illustrated in Figure 8. It is obvious that the polarizability increments with the chain length, as seen in the A, B, and C compounds. Moreover, compound E, which possesses a smectic A mesophase, has the highest polarizability (475.63 Bohr). This can be attributed to the presence of the mesogenic core (-N=N-) and the nitrogen atom in the terminal ring, as well as the large size of the compound, which facilitates the polarizability. The relatively low polarizability of the D compound compared to E can be attributed to the presence of a pyridine ring in the center of the compound. Moreover, compound H has slightly higher polarizability than F, which may be rationalized by the orientation of the mesogenic core (COO) and the position of the nitrogen atom in the heterocycle ring. Additionally, the change in the position of the mesogenic core (-CH=N-) and (COO) slightly affects the polarizability, as seen in compounds G and I. These findings accumulate together with other factors, such as dipole moment, to elucidate the mesomorphic behavior of the compounds.

3.2.2. Molecular Electrostatic Potential (MEP)

The electron density at the atomic sites of the studied compounds has a significant impact on the dipole moment, polarizability, and electronic structure. Moreover, the molecular electrostatic potential (MEP) is a useful tool to analyze the electron density distribution in the molecule and is considered an effective method for predicting intermolecular interactions. Thus, the charge distribution map of the compounds was computed by the same method, using the same basis set, according to molecular electrostatic potential (MEP) and is illustrated in Figure 9. The most negatively charged sites are colored in red, whereas the least negatively charged are colored in blue. It is obvious that the negatively charged sites are those with high numbers of electronegative nitrogen and oxygen atoms in the mesogenic cores, such as Schiff base portion (-CH=N-), azo group (-N=N-), and carboxylate ester (COO). On the other hand, the terminal chain is the least negatively charged site. It is noticed that the length of the terminal alkyl chain may have an insignificant effect on the charge distribution of the aromatic ring moieties, as seen in the A, B, and C compounds. Moreover, the charge map distribution is affected by the orientation and position of the mesogenic core, which could influence the type and the temperature range of the mesophase by varying the competitive interaction between terminal interaction and lateral interaction. The distribution of terminal charge for the F and I compounds permits the head-to-tail interaction to develop the nematic mesophase, whereas the high charge separation for the E compound enhances the side–side interaction, resulting in the formation of the more ordered smectic mesophase. In the same context, the accumulated negative charge in the center of the D compound reduces the lateral in addition to the terminal interactions, which could explain the non-mesomorphic behavior.

3.2.3. Frontier Molecular Orbitals (FMOs)

The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) are the two frontier molecular orbitals (FMOs). The HOMO is an electron donor, whereas the LUMO is an electron acceptor. The energy gap is the difference in energy between them, which may indicate the electron transition capability. Furthermore, it is inversely proportional to the molecules’ reactivity. The energies of the compounds’ HOMO and LUMO orbitals were computed using the same method and the same basis set. The results are tabulated in Table 5 and the isodensity surface plots for the ground state FMOs are illustrated in Figure 10. One can conclude from the results that the polar alkoxy terminal chain significantly lowers the energy gap of FMOs compared to the non-polar alkyl chain, although it does not influence the location of the FMOs’ electron densities. Additionally, the aromatic rings have most of the electron densities of the sites that form the HOMOs and the LUMOs. In the same approach, the coplanarity of compound E caused by the azo group enhances the conjugation of the aromatic rings, thus decreasing the energy gap of FMOs. Furthermore, the energy gap is greatly affected by the position and the orientation of the mesogenic cores (-CH=N- and COO), as seen in the G and I compounds, which could be attributed to the high degree of conjugation in the G compound. On the other hand, the changing of the terminal chain length has no effect on the energy gap value, as seen in the A, B, and C compounds.
The energy gap values are important to calculate the global softness and chemical hardness of the investigated compounds. The global softness (S), which is equal to 1/(ELUMO-EHOMO), can predict the sensitivity of the compound for the photoelectric behavior [24]. The photoelectric sensitivity of the compound improves as its global softness increases. The chemical hardness (η), which is equal to (ELUMO−EHOMO)/2, indicates the resistance of the charge transfer. These parameters can determine the suitability of the liquid crystalline compounds to be employed in many applications as transistors. Some molecules that contain alkyl chains were functionalized to give smectic phases, which are optimized for applications as field-effect transistors [1]. In addition, LCs have been used in the manufacture of optical photovoltaic cells. The mechanism of OPV cell operation to exchange light energy into electrical energy is designed as follows: charge generation occurs between two different organic semiconductors, followed by separation and movement toward two opposite electrodes [25]. Recently, LCs served as active signal amplifiers in sensing and biosensing applications due to their high sensitivity properties, rapid response, and easy manufacturing. Thermotropic LC biosensors depend on the molecular interactions that lead to adsorption and desorption at the interface, and due to the optical anisotropy of LCs, the output optical signals will vary according to the samples analyzed [26].

4. Conclusions

Three new liquid crystalline compounds were synthesized and characterized, namely pyridin-2-yl 4-[4-(alkylphenyl)iminomethyl]benzoate. Each of the three compounds has an enantiotropic nematic mesophase. The enhancement of the nematic phase may be attributed to the alkyl group disorder. The longer alkyl group may result in insufficient space between the layers for the terminal chains to accommodate, causing the layers to be parted, which is essential in nematic phase development. A comparative study of the prepared compounds with their reported structurally similar compounds was carried out to explain the effect of the mesogenic core and its orientation on the mesomorphic behavior. The results reveal that the presence of the azo mesogenic core (-N=N-) might enhance the planarity of the compounds. This could be responsible for the formation of the smectic A mesophase in the E compound as the planarity could increase the degree of molecular packing, which creates a more ordered smectic mesophase. Moreover, the increment in the temperature at which the nematic phase was formed for the A, B, and C compounds may have caused the dipole moment to be increased. Additionally, the polar alkoxy terminal chain increases the dipole moment value to be higher than that of the non-polar alkyl chain. With the same approach, the dipole moment increment resulted in an increment in the mesophase range for the F, G, H, and I compounds. Meanwhile, the polarizability increases with the chain length, as seen in the A, B, and C compounds. Moreover, the relatively low polarizability of the D compound compared to E can be attributed to the presence of a pyridine ring in the center of the compound. In addition, it was found using the molecular electrostatic potential that the D compound has non-mesomorphic behavior due to the accumulated negative charge in the center of the compound, which reduces the lateral in addition to the terminal interactions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12070902/s1, Figure S1: 1HNMR of Pyridin-2-yl 4-[4-(tetradecylphenyl)iminomethyl]benzoate (C); Figure S2: C13 NMR of Pyridin-2-yl 4-[4-(tetradecylphenyl)iminomethyl]benzoate (C).

Author Contributions

Conceptualization, M.H.; Data curation, S.N., M.J., A.-H.E. and M.H.; Formal analysis, S.N., M.G., A.-H.E. and M.H.; Investigation, A.-H.E. and M.H.; Methodology, S.N., M.G. and M.E.; Project administration, M.J.; Resources, M.J.; Software, M.H. and J.Y.A.-H.; Writing—original draft, S.N., M.G., M.H. and J.Y.A.-H.; Writing—review and editing, S.N., M.G., J.Y.A.-H. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research project has been funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R24), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Samples of the compounds are available from the authors.

Acknowledgments

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R24), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ashikur Rahman Rabbi, J.A.F. Preparation, Characterization and Applications of Liquid Crystals: A Review. J. Appl. Chem. 2020, 13, 43–54. [Google Scholar]
  2. Khoo, I.-C. Liquid Crystals; John Wiley & Sons: Hoboken, NJ, USA, 2022. [Google Scholar]
  3. Andrienko, D. Introduction to liquid crystals. J. Mol. Liq. 2018, 267, 520–541. [Google Scholar] [CrossRef]
  4. Mitov, M. Liquid-Crystal Science from 1888 to 1922: Building a Revolution. ChemPhysChem 2014, 15, 1245–1250. [Google Scholar] [CrossRef] [PubMed]
  5. Ashok, C.K.; Ramesh, D.R.; Ola, M.M.; Chaudhari, V.A. Liquid crystals: A review. Int. J. All Res. Writ. 2019, 1, 119–129. [Google Scholar]
  6. Priestly, E. Introduction to Liquid Crystals; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  7. Chester, A.N.; Martellucci, S. Phase Transitions in Liquid Crystals; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013; Volume 290. [Google Scholar]
  8. Jber, N.R.; Shukur, M.M.; Najaf, A.A. Schiff Base liquid Crystals with Terminal Alkoxy Group Synthesis and Thermotropic Properties. Al-Nahrain J. Sci. 2014, 17, 64–72. [Google Scholar] [CrossRef]
  9. Hagar, M.; Ahmed, H.; Aouad, M. Mesomorphic and DFT diversity of Schiff base derivatives bearing protruded methoxy groups. Liq. Cryst. 2020, 47, 2222–2233. [Google Scholar] [CrossRef]
  10. Nafee, S.S.; Hagar, M.; Ahmed, H.A.; El-Shishtawy, R.M.; Raffah, B.M. The synthesis of new thermal stable schiff base/ester liquid crystals: A computational, mesomorphic, and optical study. Molecules 2019, 24, 3032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Ha, S.-T.; Yeap, G.-Y.; Boey, P.-L. Synthesis and liquid crystalline properties of new Schiff bases N-[4-(4-nalkanoyloxybenzoyloxy) benzylidene]-4-cyano-, 4-hydroxy-, 4-thio-and 4-nitroanilines. Aust. J. Basic Appl. Sci. 2009, 3, 3417–3422. [Google Scholar]
  12. Hagar, M.; Ahmed, H.; Saad, G. Mesophase stability of new Schiff base ester liquid crystals with different polar substituents. Liq. Cryst. 2018, 45, 1324–1332. [Google Scholar] [CrossRef]
  13. Ahmed, N.H.; 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]
  14. Hagar, M.; Ahmed, H.; Saad, G. New calamitic thermotropic liquid crystals of 2-hydroxypyridine ester mesogenic core: Mesophase behaviour and DFT calculations. Liq. Cryst. 2020, 47, 114–124. [Google Scholar] [CrossRef]
  15. Ahmed, H.A.; Hagar, M.; Alhaddad, O.A.; Zaki, A.A. Optical and geometrical characterizations of non-linear supramolecular liquid crystal complexes. Crystals 2020, 10, 701. [Google Scholar] [CrossRef]
  16. Al-Mutabagani, L.A.; Alshabanah, L.A.; Ahmed, H.A.; Alalawy, H.H.; Al Alwani, M.H. Synthesis, mesomorphic and computational characterizations of nematogenic schiff base derivatives in pure and mixed state. Molecules 2021, 26, 2038. [Google Scholar] [CrossRef] [PubMed]
  17. Nada, S.; Hagar, M.; Farahat, O.; Hasanein, A.A.; Emwas, A.-H.; Sharfalddin, A.A.; Jaremko, M.; Zakaria, M.A. Three Rings Schiff Base Ester Liquid Crystals: Experimental and Computational Approaches of Mesogenic Core Orientation Effect, Heterocycle Impact. Molecules 2022, 27, 2304. [Google Scholar] [CrossRef] [PubMed]
  18. Ahmed, H.; Hagar, M.; Alhaddad, O. Mesomorphic and geometrical orientation study of the relative position of fluorine atom in some thermotropic liquid crystal systems. Liq. Cryst. 2020, 47, 404–413. [Google Scholar] [CrossRef]
  19. Paterson, D.A.; Crawford, C.A.; Pociecha, D.; Walker, R.; Storey, J.M.; Gorecka, E.; Imrie, C.T. The role of a terminal chain in promoting the twist-bend nematic phase: The synthesis and characterisation of the 1-(4-cyanobiphenyl-4′-yl)-6-(4-alkyloxyanilinebenzylidene-4′-oxy) hexanes. Liq. Cryst. 2018, 45, 2341–2351. [Google Scholar] [CrossRef]
  20. Frisch, M.; Trucks, G.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. Gaussian 09, Revision d. 01; Gaussian. Inc.: Wallingford, CT, USA, 2009; p. 201. [Google Scholar]
  21. Dennington, R.; Keith, T.; Millam, J. GaussView, Version 5; Semichem, Inc.: Shawnee, KS, USA, 2009. [Google Scholar]
  22. Adrienko, G.A. Chemcraft 1.8 (build445). Available online: https://www.chemcraftprog.com/download.html (accessed on 24 April 2022).
  23. Ribera, D.; Serra, A.; Mantecón, A. Dimeric liquid-crystalline epoxyimine monomers: Influence of dipolar moments on mesomorphic behavior and the formation of liquid-crystalline thermosets. J. Polym. Sci. Part A Polym. Chem. 2003, 41, 1465–1477. [Google Scholar] [CrossRef]
  24. Al-Hossainy, A.F.; Zoromba, M.S. Doped-poly (para-nitroaniline-co-aniline): Synthesis, semiconductor characteristics, density, functional theory and photoelectric properties. J. Alloys Compd. 2019, 789, 670–683. [Google Scholar] [CrossRef]
  25. Kumar, M.; Kumar, S. Liquid crystals in photovoltaics: A new generation of organic photovoltaics. Polym. J. 2017, 49, 85–111. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, Z.; Xu, T.; Noel, A.; Chen, Y.-C.; Liu, T. Applications of liquid crystals in biosensing. Soft Matter 2021, 17, 4675–4702. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The investigated compounds A, B, C, D, E, F, G, H, and I.
Figure 1. The investigated compounds A, B, C, D, E, F, G, H, and I.
Crystals 12 00902 g001
Scheme 1. Synthesis of pyridin-2-yl 4(((4-alkylphenyl)imino)methyl)benzoate (A, B, and C).
Scheme 1. Synthesis of pyridin-2-yl 4(((4-alkylphenyl)imino)methyl)benzoate (A, B, and C).
Crystals 12 00902 sch001
Figure 2. Differential scanning calorimetry (DSC) thermogram of compound A during heating and cooling cycles, at 10.0 °C/min.
Figure 2. Differential scanning calorimetry (DSC) thermogram of compound A during heating and cooling cycles, at 10.0 °C/min.
Crystals 12 00902 g002
Figure 3. Nematic texture under polarized optical microscopy (POM) upon heating (a) for B at 123 °C and (b) for C at 130 °C, with heating rate of 10.0 °C/min.
Figure 3. Nematic texture under polarized optical microscopy (POM) upon heating (a) for B at 123 °C and (b) for C at 130 °C, with heating rate of 10.0 °C/min.
Crystals 12 00902 g003
Figure 4. Graphical representations of (a) mesophase ranges of all investigated compounds A, B, C, D, E, F, G, H, and I; (b) mesophase temperatures in terms of the compounds.
Figure 4. Graphical representations of (a) mesophase ranges of all investigated compounds A, B, C, D, E, F, G, H, and I; (b) mesophase temperatures in terms of the compounds.
Crystals 12 00902 g004aCrystals 12 00902 g004b
Figure 5. The estimated molecular geometry of A, B, C, D, E, F, G, H, and I compounds.
Figure 5. The estimated molecular geometry of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g005
Figure 6. Graph of mesophase range versus aspect ratio of A, B, C, D, E, F, G, H, and I compounds.
Figure 6. Graph of mesophase range versus aspect ratio of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g006
Figure 7. Graph of mesophase range versus total dipole moment of A, B, C, D, E, F, G, H, and I compounds.
Figure 7. Graph of mesophase range versus total dipole moment of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g007
Figure 8. Graph of mesophase range versus polarizability of A, B, C, D, E, F, G, H, and I compounds.
Figure 8. Graph of mesophase range versus polarizability of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g008
Figure 9. The estimated molecular electrostatic potentials (MEP) of A, B, C, D, E, F, G, H, and I compounds.
Figure 9. The estimated molecular electrostatic potentials (MEP) of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g009
Figure 10. Graphical representation of frontier molecular orbitals and energy gaps of A, B, C, D, E, F, G, H, and I compounds.
Figure 10. Graphical representation of frontier molecular orbitals and energy gaps of A, B, C, D, E, F, G, H, and I compounds.
Crystals 12 00902 g010
Table 1. Phase transition temperatures (°C) at 10.0 °C/min, enthalpy of transition ΔH (kJ/mol), entropy of transition ΔS (J/mol.K), and normalized entropy of transition of compounds A, B, C, D, E, F, G, H, and I; all values have been calculated for the heating transition.
Table 1. Phase transition temperatures (°C) at 10.0 °C/min, enthalpy of transition ΔH (kJ/mol), entropy of transition ΔS (J/mol.K), and normalized entropy of transition of compounds A, B, C, D, E, F, G, H, and I; all values have been calculated for the heating transition.
CompoundsTCr-SmATCr-NTCr-ITSmA-ITN-IΔHCr-NΔSCr-NΔSCr-N/RΔHCr-SmAΔSCr-SmAΔSCr-SmA/RΔHCr-IΔSCr-IΔSCr-I/RΔHN-IΔSN-IΔSN-I/RΔHSmA-IΔSSmA-IΔSSmA-I/R
A-118.24--134.1629.775.99.1------0.300.740.09---
B-118.65--132.5538.297.511.7------0.411.010.12---
C-120.32--133.6044.7113.613.7------0.501.230.15---
D--124.3--------35.1788.4910.64------
E96.1--121.5----23.563.647.65------1.64.050.49
F-128.6--174.243.40108.0312.99------1.804.020.48---
G67.3--87.3----28.3883.3610.02------1.925.332.65
H-116.42--130.5950.48129.5815.58------2.035.030.60---
I-110.8--156.826.1568.118.19------0.882.050.25---
Cr-SmA = Crystal to smectic A transition; Cr-N = Crystal to nematic transition; Cr-I = Crystal to isotropic liquid transition; SmA-I = Smectic A to isotropic liquid transition; N-I = Nematic to isotropic liquid transition.
Table 2. The estimated twist angles for A, B, C, D, E, F, G, H, and I compounds.
Table 2. The estimated twist angles for A, B, C, D, E, F, G, H, and I compounds.
CompoundsθI-IIθI-IIIθII-III
A27.2811.2037.04
B11.8626.1737.57
C15.1522.1636.80
D0.060.060.01
E0.010.040.05
F6.2522.5128.70
G48.9179.9731.42
H26.416.6330.57
I40.1557.0716.96
θI-II = Twist angle between the planes of I ring and II ring; θI-III = Twist angle between the planes of I ring and III ring; θII-III = Twist angle between the planes of II ring and III ring.
Table 3. Mesomorphic parameters for heating cycle (°C), the calculated dimensions (Ǻ), and aspect ratios of A, B, C, D, E, F, G, H, and I compounds.
Table 3. Mesomorphic parameters for heating cycle (°C), the calculated dimensions (Ǻ), and aspect ratios of A, B, C, D, E, F, G, H, and I compounds.
CompoundsCrTSmATNΔTCΔTSmAΔTNDimensionsAspect Ratio (L/D)
Length (L)Width (D)
A118.24-134.1615.92-15.9229.627.553.92
B118.65-132.5513.9-13.934.588.424.11
C120.32-133.6013.28-13.2837.168.454.40
D124.3-----31.327.933.95
E96.1121.5-25.425.4-41.707.895.29
F128.6-174.245.6-45.631.158.063.86
G67.387.3-2020-28.787.273.96
H116.42-130.5914.17-14.1731.357.484.19
I110.8-156.846-4629.057.473.89
Table 4. Mesomorphic parameters for heating cycle (°C), polarizability (Bohr), and dipole moment (Debye) of A, B, C, D, E, F, G, H, and I compounds.
Table 4. Mesomorphic parameters for heating cycle (°C), polarizability (Bohr), and dipole moment (Debye) of A, B, C, D, E, F, G, H, and I compounds.
CompoundsTSmATNΔTCΔTSmAΔTNPolarizabilityDipole Moment
(x)(y)(z)Total
A-134.1615.92-15.92376.160.16−1.45−0.431.52
B-132.5513.9-13.9426.560.12−0.59−1.251.39
C-133.6013.28-13.28451.29−0.120.79−1.181.43
D-----380.01−1.810.760.001.96
E121.5-25.425.4-475.63−5.87−2.630.006.43
F-174.245.6-45.6377.903.283.270.114.64
G87.3-2020-368.62−2.87−1.860.513.45
H-130.5914.17-14.17391.56−1.89−0.14−1.312.30
I-156.846-46352.823.820.460.333.87
Table 5. The frontier molecular orbital energies (eV), global softness S, and chemical hardness η (eV) of A, B, C, D, E, F, G, H, and I compounds.
Table 5. The frontier molecular orbital energies (eV), global softness S, and chemical hardness η (eV) of A, B, C, D, E, F, G, H, and I compounds.
CompoundsELUMOEHOMOΔESη
A−2.38−6.153.770.271.89
B−2.39−6.153.760.271.88
C−2.39−6.153.760.271.88
D−2.75−6.343.590.281.80
E−2.75−6.263.510.281.76
F−2.33−5.783.450.291.73
G−2.28−5.813.530.281.77
H−2.30−5.803.500.291.75
I−1.97−6.074.100.242.05
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-Humaidi, J.Y.; Nada, S.; Gerges, M.; Ehab, M.; Jaremko, M.; Emwas, A.-H.; Hagar, M. 2-Pyridinyl-Terminated Iminobenzoate: Type and Orientation of Mesogenic Core Effect, Geometrical DFT Investigation. Crystals 2022, 12, 902. https://doi.org/10.3390/cryst12070902

AMA Style

Al-Humaidi JY, Nada S, Gerges M, Ehab M, Jaremko M, Emwas A-H, Hagar M. 2-Pyridinyl-Terminated Iminobenzoate: Type and Orientation of Mesogenic Core Effect, Geometrical DFT Investigation. Crystals. 2022; 12(7):902. https://doi.org/10.3390/cryst12070902

Chicago/Turabian Style

Al-Humaidi, Jehan Y., Shady Nada, Mariam Gerges, Marwa Ehab, Mariusz Jaremko, Abdul-Hamid Emwas, and Mohamed Hagar. 2022. "2-Pyridinyl-Terminated Iminobenzoate: Type and Orientation of Mesogenic Core Effect, Geometrical DFT Investigation" Crystals 12, no. 7: 902. https://doi.org/10.3390/cryst12070902

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