Next Article in Journal
Analysis of the Physiognomy of Unique Sets in the Maximum Number of Repetitions Strategy—The Case of One-Arm Scott Machine Seated Bicep Curls
Next Article in Special Issue
β-Phosphonated Glycine Pendant Groups Grafted on Styrene-6.7% Divinylbenzene Copolymers: Synthesis and Their Application as Photocatalysts
Previous Article in Journal
Performance Assessment of Using Docker for Selected MPI Applications in a Parallel Environment Based on Commodity Hardware
Previous Article in Special Issue
Influence of Operating Parameters on Photocatalytic Oxidation of 2,4-Dichlorofenol in Aqueous Solution by TiO2/Stainless Steel Photocatalytic Membrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pentacoordinated Liquid Crystalline Zn(II) Complex Organized in Smectic Mesophase: Synthesis, Structural and Electrochemical Properties

1
“Coriolan Drăgulescu” Institute of Chemistry, Romanian Academy, 24 Mihai Viteazu Bvd., 300223 Timisoara, Romania
2
Institut de Physique et Chimie des Matériaux de Strasbourg (IPCMS), UMR7504, CNRS-Université de Strasbourg, 67034 Strasbourg, France
3
Department of Applied Chemistry and Engineering of Inorganic Compounds and Environment, Politehnica University of Timisoara, Blv. Vasile Parvan No. 6, 300223 Timisoara, Romania
4
National Institute of Research and Development for Industrial Ecology (INCD ECOIND), Timisoara Branch, 300431 Timisoara, Romania
5
Department of Environmental Engineering and Management, “Gheorghe Asachi” Technical University of Iasi, 700050 Iasi, Romania
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2022, 12(16), 8306; https://doi.org/10.3390/app12168306
Submission received: 22 July 2022 / Revised: 10 August 2022 / Accepted: 16 August 2022 / Published: 19 August 2022
(This article belongs to the Special Issue New Trends in Functional and Multifunctional Advanced Materials)

Abstract

:

Featured Application

The study was conducted as fundamental research aiming new advanced materials characterized envisaging applications of non-enzymatic electrochemical sensing.

Abstract

The synthesis and structural characterization of a new liquid crystalline coordination complex based on pentacoordinated Zn(II) metal centre with the coordination fulfilled by the tridentate chelating N^N^N 2,2′;6′,2″-terpyridine ligand and two monoanionic gallates decorated with several long alkyl chains is described. The mesomorphic properties were accurately investigated by small- and wide-angle X-ray scattering studies. Despite the bulky coordination around the metal centre, the complex self-organizes into a smectic phase and, based on the structural and geometrical parameters, a model for the supramolecular organization in the liquid crystalline phase is proposed. Electrochemical investigations showed the importance of the molecular structure of the coordination complex in enhancing its aqueous sensing capacities: the bulky organic ligands form an organic shell separating the metal centres and favouring the redox system through their reduction followed by stripping.

1. Introduction

Coordination complexes based on d-block metal centres are considered as promising molecular materials for bringing important advancements in applicative fields, such as opto-electronics, photonics, data storage, catalysis, biomedicine, etc., due to unique features imparted by the different nature of the molecular fragments, which, when combined together, contribute to the collective properties of the final material [1,2,3,4,5,6,7,8,9,10]. The redox, magnetic, optical, catalytic, electrical or electrochemical properties of the d-block metal ion centre can cooperatively interact with those of the organic ligands (redox, optic, biologic properties) and/or counterions (solubility in polar solvents, anion coordination) in the case of ionic species, to enhance or modify the overall properties of the architectural ensemble [11,12,13,14,15,16,17,18,19]. Indeed, the properties of a d-block-based metal coordination complex are more than the sum of the individual component’s properties, and in many cases, synergistic unique physicochemical properties may be engineered this way [4,5,20], while the supramolecular arrangement is important due to the sensitivity of the coordination complex to the molecular surroundings, particularly for emission properties [21,22]. Thus, different soft or crystalline supramolecular architectures built up with the same “d-block metal containing building block” through non-covalent interactions can have varied photophysical properties, and small changes in their molecular structure may induce a huge modification of the materials response to external stimuli [5,15,16,21,22].
Zn(II)-based metallomesogens were mainly designed for application in electro-optics, due to the possibility of obtaining luminescent polarized emission [23,24,25,26,27,28]. Recently, a series of pentacoordinated Zn(II) coordination complexes (Zn_1–4 in Figure 1) self-assembling into 2D hexagonal columnar and 3D mesophases was reported [27,28]. The symmetry of the mesophases was somehow expected taking into account the bulky coordination sphere around the Zn(II) ion fulfilled by a tridentate chelating N^N^N terpyridine (tpy) ligand differently substituted in apical position and two monoanionic gallate-based ligands. Indeed, the majority of tetra-, penta- and hexa-coordinated Zn(II) derivatives displayed columnar mesophases [24,25,26,27,28,29,30,31,32,33,34]; however, some smectic phases [23,35,36,37,38] or lamello-columnar phases were also reported [39].
The monoanionic ligands decorated with several long alkyl chains were responsible for the induction of liquid crystalline properties, while the terpyridine-based substituents (−C6H4R, Figure 1) influenced the supramolecular arrangement by the modification of the interface between antagonistic aromatic and alkyl chains segments and were the determining factor for the photophysical properties: the electron-donating -NEt2 group originated an ILCT excited state and Zn_2 reaching a record luminescence quantum yield in solution (95%).
In electrochemical sensing, the presence of a d-block metal centre brings redox properties, while the organic ligands may act as a separator of the metal centres avoiding close contacts and implicit agglomeration; the separation of the redox active centres may yield micro or even nano-electrode arrays/assemblies, increasing the sensing performances, especially for bioanalysis devices [40,41,42]. The electrocatalytic activities of various forms of zinc oxides have been studied for multiple applications, e.g., the oxygen reduction (ORR) reaction that play an important role in renewable energy technologies [43,44], water electrooxidation for the generation of hydrogen peroxide [45] and for electrochemical sensing of various biologically active compounds [46,47]. Additionally, metallomesogens based on the redox active Cu(I) metal centre showed potential in improving the electrocatalytic performances of carbon-based materials for detection of bioanalytes [48,49]. Hence, our aim was to study the electrochemical behaviour of related Zn(II) coordination complex as a function of the molecular structure. In order to reduce the workload in the synthesis of complexes, we employed the non-substituted 2,2′;6′,2″-terpyridine to obtain Zn_tpy (Scheme 1) with a reduced molecular structure and followed the changes in the supramolecular arrangement in condensed states. The electrochemical properties of complex Zn_tpy were investigated and discussed as a function of molecular structure.

2. Materials and Methods

The starting materials commercially available were used as received. The synthesis of the silver(I) salt of 3,4,5-tridodecyloxybenzoate was performed as reported in the literature [50]. 1H and 13C NMR experiments were recorded on Bruker Avance III HD–500 MHz spectrometer in CD2Cl2 or DMSO-d6, using tetramethylsilane (TMS) as internal standard. Elemental analyses (CHN) were performed with a Flash 2000 microanalyzer from ThermoFisher Scientific. The percentage of Zn(II) was determined using a SensAA flame atomic absorption spectrometer (GBC Scientific Equipment, Australia) equipped with a zinc hollow cathode lamp (detection limit: 0.4–1.5 mg/L, wavelength: 213.9 nm, slit width: 0.5 nm). The flame used was an air–acetylene (oxidizing) mixture. Two determinations were made and the average absorbance value was further used.
Mesophases optical textures of complex Zn_tpy were observed with an Olympus BX53M polarizing microscope (POM) equipped with a Linkam hot stage. Images of the various phases were recorded using an Olympus UC90 camera. Enthalpies and transition temperatures were recorded using a Q1000 apparatus from TA Instruments. The apparatus was calibrated with indium; three heating/cooling cycles were performed on each sample, with a heating and cooling rate of 5 °C/min. Thermal gravimetric analysis (TGA) was carried out using a Q50 instrument from TA Instruments at a scanning rate of 5 °C/min and with air as purge gas.
S/WAXS patterns were obtained with a transmission Guinier-like geometry. A linear focalized monochromatic Cu Kα1 beam (λ = 1.5405 Å) was obtained using a sealed-tube generator (600 W) equipped with a bent quartz monochromator. In all cases, samples were filled in home-made sealed cells of 1 mm path. The patterns were recorded with a curved Inel CPS120 counter gas-filled detector linked to a data acquisition computer (periodicities up to 90 Å) and on image plates scanned by Amersham Typhoon IP with 25 µm resolution (periodicities up to 120 Å). The sample temperature was controlled within ±0.01 °C, and exposure times were varied from 2 to 6 h.

2.1. Synthesis

Complex Zn_p. A solution of ZnCl2 (0.292 g, 2.143 mmol) in 10 mL of MeOH was added dropwise to a solution of ligand 2,2′;6′,2″-terpyridine (0.500 g, 2.143 mmol) in 20 mL MeOH. During the addition of the Zn(II) salt, a pale-yellow precipitate formed. The reaction mixture was stirred for 2 h at r. t. Then, the yellowish precipitate was filtered off, washed with MeOH and dried under vacuum to yield 0.785 g (2.122 mmol, η = 99%) of the desired compound. FT-IR (KBr, cm−1): 1594–1448 (υC=C, υC=N); 1H-NMR (DMSO-d6, δ/ppm): 8.82 (overlapped peaks, 4H), 8.60 (overlapped peaks, 3H), 8.34 (t, 3J = 7.8 Hz, 2H), 7.86 (t, 3J = 6.4 Hz, 2H). Anal. Calcd. for C15H11Cl2N3Zn (369.56 g·mol−1): C, 48.75; H, 3.00; N, 11.37. Found: C, 48.58; H, 3.06; N, 11.03%.
Complex Zn_tpy. A solution of Ag(I) salt of 3,4,5-tridodecyloxybenzoate (0.698 g, 0.893 mmol) in 25 mL CHCl3 was added dropwise to a suspension of Zn_p (0.150 g, 0.4054 mmol) in 40 mL of CHCl3. The reaction mixture was stirred at room temperature for 6 h. After the removal of AgCl by filtration, the mother liquor was evaporated under reduced pressure. The residue was dissolved in Et2O and acetone was added. The precipitate formed was filtered out, and the pure product was obtained from the mother solution by repeated recrystallization from acetone and hexane/EtOH (0.315 g, η = 47%). FT-IR (KBr, cm−1): 2924 (νas,(-CH2−)), 2854 (νs,(-CH2−)), 1623 (νas,(COO)), 1567–1428 (υC=C, υC=N) 1359 (νs(COO)); 1H-NMR (CD2Cl2, δ/ppm): 8.87 (d, 3J = 5.0 Hz, 2H), 8.20 (overlapped peaks, 5H), 7.96 (t, 3J = 7.8 Hz, 2H), 7.49 (t, 3J = 7.6 Hz, 4H), 7.09 (s, 2H), 3.83 (overlapped peaks, 12H), 1.77–1.26 (overlapped peaks, 120 H), 0.79 (t, 3J = 6.8 Hz, 18H). Anal. Calcd. for C101H165N3O10Zn (1646.79 g·mol−1): C, 73.66; H, 10.10; N, 2.55. Found: C, 73.46; H, 10.05; N, 2.64%. AAS: Zn% calcd.: 3.97, found: 3.91.

2.2. Electrochemical Characterization

All electrochemical experiments were carried out in 0.1 M NaOH solutions by cyclic voltammetry (CV) technique in a three-electrode cell connected with an Autolab potentiostat/galvanostat PGSTAT 302 (EcoChemie, The Netherlands) controlled with GPES 4.9 software. The working electrode was a glassy carbon (GC) provided by Metrohm (Switzerland) with the disk diameter of 5.0 mm. The counter electrode was a platinum plate electrode with the geometrical surface area of 1.0 cm2. The electrode potentials were recorded against saturated calomel reference (SCE) electrode. GC working electrode was modified with Zn_tpy and named Zn_tpy/GC electrode; its electrochemical behaviour was studied by comparison with simple GC electrode in 0.1 M NaOH supporting electrolyte. The GC electrode was mechanically cleaned by polishing with alumina, followed by rinsing with distilled water and drying. The modification was done by simple film casting through cleaned GC electrode immersion within a non-aqueous suspension containing 5 mg Zn_tpy /mL acetonitrile. After immersion, the electrode surface was dried at room temperature. All measurements were performed at room temperature (25  ±  1 °C).

3. Results and Discussion

3.1. Synthesis and Characterization

For the synthesis of the complex Zn_tpy, we employed the synthetic strategy previously reported, that requires first the synthesis of the dichloro Zn(II) precursor (Zn_p in Scheme 1), followed by its reaction with the Ag(I) salt of 3,4,5-tridodecyloxybenzoate [27,28]. The complex Zn_tpy was obtained by repeated crystallizations as a yellowish powder and was dried in vacuum for several hours. Structural characterizations were performed by FT-IR and 1H NMR spectroscopies, while the purity was confirmed by elemental analysis and AAS spectroscopy (Supplementary Materials, Figure S1).

3.2. Thermal and Mesomorphic Properties

The thermal stability was determined by thermogravimetric analysis (TGA) (Supplementary materials, Figure S2), while the transition temperatures and the enthalpy changes associated were investigated by differential scanning calorimetry (DSC) (Supplementary materials, Figure S3). The data are presented in Table 1.
For Zn_tpy, weight-loss from degradation becomes significant above ca. 170 °C, with a 5% weight-loss at 230 °C (at a rate of 5 °C min−1). Polarized optical microscopy investigations (POM) showed that, on cooling from the isotropic liquid, the complex self-assembles into a liquid crystalline mesophase (Figure 2a) below 114 °C, as indicated by the growth of a fan-shaped, fluid but non-characteristic, optical texture. The final identification of the type and symmetry of the mesophase was achieved by small- and wide-angle X-ray scattering studies (S/WAXS–Figure 2b, Supplementary materials, Figure S4).
The SWAXS patterns are typical of a liquid crystalline smectic phase, with the overlapping broad scattering signal arising from liquid-like lateral distances between undifferentiated molecular segments, i.e., molten chains, hch, and aromatic parts, har, in the wide-angle region (hch + har), and with two sharp small-angle reflections corresponding to the first two lamellar orders (h00) (100, very strong, and 200, weak), emerging from the nanosegregated structure formed by the antagonistic aromatic and aliphatic segments (Figure 2b); an additional broad signal, corresponding to short-range periodicity from the arrangement of the metallic-tpy moiety within aromatic layers, Dar, was also observed at ca. 7.5–8.0 Å.
Due to the specific amphiphilic molecular structure of the zinc complex, with the six alkyl chains radiating in the same direction, whereas the tpy moiety is deprived of chains, the complexes are probably self-organized pairwise in a more or less head-to-head fashion (Figure 3), with the aromatic and metallic moieties confined into a median sublayer (Dar), whose periodic alternation with molten chain sublayers generates a SmA-like structure with a periodicity dlam of ca. 29.3 Å at 85 °C (ca. 28.9 Å at 100 °C, Table 2).
The geometry of the Zn complexes can be deduced from the single-crystal structures of molecules, including the same complex moiety. In this way, structures of several Zinc bis(enzoate)-phenylterpyridine derivatives are available (CSD-BETWOO, CSD-BETWUU, CSD-BETXAB, CSD-LEXWES) [52] and show irregular trigonal bipyramidal geometries, in which both benzoate ligands are nearly orthogonal to the terpyridine ring plane.
In Zn_tpy, benzoate ligands consist of gallate units substituted by three long dodecyl chains, whose rejection into adjacent molten aliphatic layers generates a smectic A-like organization (Figure 3). The formation of such lamellar structure implies that the molecular packing within the aromatic layer matches the high overall cross-sectional area of the six pending chains (6×σch ≈ 135 Å2). This condition is realized through a monolayer configuration alternating terpyridine rings and gallate segments in the layer plane. This arrangement, which is also found in CSD-BETWOO and otherwise explains the in-plane periodicity Dar, offers a high molecular cross-section (SZnAr = 87 Å2 in CSD-BETWOO and SZnAr ≈ 95 Å2 in Zn_tpy, Table 2). Above all, it allows sharing alkyl substituents between both adjacent layers, and thus elevates the molecular area to Amol ≈ 190 Å2, which even exceeds the requirement for alkyl chains (qch). This discrepancy is, however, easily compensated by a higher degree of the folding of the chains, to the cost of more diffuse interfaces and of a lower efficient nanosegregation between antagonistic moieties. This effect should contribute to the narrowness of the smectic range and to the difficulty to detect the lamellar higher order reflections.
The reduction of the molecular structure beside easing the synthesis has a significant influence on the supramolecular organisation in the mesophase: the non-substituted coordination complex described herein Zn_tpy, arranges into smectic phases while the apical substitution of the tpy ligand direct the self-assembly into room temperature hexagonal mesophases, as previously reported for Zn_1-4 [27,28] (Figure 1). Indeed, depending on the functionality of the tpy substituent, the complexes self-assemble either into 3D hexagonal mesophase (Mhex) or in 2D hexagonal lattice (Colh) (See Figure 4).

3.3. Electrochemistry

Studies regarding the electrochemical behaviour of functional materials, such as metallomesogens, are mandatory for any type of electrochemical application, such as electrochemical sensing. CV technique is a valuable tool to understand the processes that occur at the electrode surface containing a liquid crystal. However, the electrode composition cannot only contain the metallomesogen because it is not able to ensure the electrical conductivity suitable for electrochemical applications, but it can be used as modifier for the carbon-based electrode to enhance its electrochemical properties. In this study, the GC electrode was used as a substrate and Zn_tpy as a modifier, resulting in a Zn_tpy/GC electrode. In order to evaluate the importance of the molecular structure on the electrochemical performance of the modified GC electrode, the electrode was characterized in comparison with a GC electrode modified with the dichloro Zn(II) precursor Zn_p (Scheme 1), which lacks alkyl chains yielding electrode Zn_p/GC, and with a simple unmodified GC electrode. The electrochemical behaviour was studied in 0.1 M NaOH and attempted to reach zinc oxide species, which exhibited enhanced electrocatalytic activity for the electrooxidation processes.
The comparative cyclic voltammograms of the GC electrode, Zn_p/GC electrode and Zn_tpy/GC electrode in 0.1 M NaOH are presented in Figure 5.
CV shape recorded at GC electrode is changed significantly by the presence of Zn_tpy onto the GC electrode vs. the Zn_p presence onto the electrode that influences negligibly the negative cathodic range in comparison with GC. This means that no electrochemical behaviour is manifested by Zn_p to be considered as redox systems for further involvement within oxidation and reduction electrode processes. First of all, to understand the electrochemical behaviour of Zn_tpy, it is important to identify all processes that occurred onto GC electrode in 0.1 M NaOH. By forward scanning, starting from −1.50 V vs. SCE, an anodic peak (AI) is evidenced at the potential value of about −0.25 V vs. SCE, and at the potential value of about +1.00 V vs. SCE, the oxygen evolution reaction (OER) occurred with a low shoulder type anodic peak at about +1.25 V, which is characterised by the diffusion-controlled oxygen evolution in alkaline medium of carbon-based electrodes. The first anodic peak should be attributed to the oxidation of the easily oxidizable functional groups of the GC surface [53]. By reverse scanning, three cathodic peaks (CI-CIII, Figure 5) appeared on the GC electrode. CI recorded at the potential value of +0.26 V vs. SCE is usually attributed to the reduction of H2O and/or dissolved O2 that is well-known as the oxygen reduction reaction (ORR). However, in accordance with the literature data [54], this peak can be attributed to the reduction of difficult oxidized surface functional groups (e.g., edge or basal hydroxyl sites). Considering the high anodic potentials corresponding to OER, the produced oxygen can be divided into inner (adsorbed or trapped onto electrode surface) and dissolved outer oxygen. The appearance of CII and CIII cathodic peaks at −0.50 V and −1.00 V vs. SCE are ascribed to the reduction of the inner and dissolved outer O2, which could occur by 4-electron pathway (reaction 1) and two successive 2-electron pathways, according to the reactions (reactions 2 and 3):
O2 + 2H2O + 4e → 4HO
O2 + H2O + 2e → HO2 + HO
HO2 + H2O + 2e → 3HO
Moreover, under the potential range of CIII, the hydrogen evolution reaction (HER) occurred onto GC electrode, which is partially overlapping with ORR.
All cathodic and anodic peaks were influenced by the presence of Zn_tpy. Thus, a slight polarization effect towards the OER is noticed and the shoulder recorded at +1.25 V/SCE is higher due to zinc oxide dissolution overlapping with O2 evolution. Considering the literature data related to zinc deposition and zinc oxide formation [55,56,57], during cathodic direction scanning, starting with −0.760 V, metallic zinc is deposited, which can continue at more negative potentials than −1.5 V vs. SCE. Moreover, under this negative potential range of ORR, ZnO can be formed based on EC (electrochemical/chemical) mechanism by coupling the zinc ions with peroxide and superoxide ions, which are generated from oxygen reduction reaction, according with reactions (4 and 5):
Zn2+ + O22− → ZnO2 → ZnO + ½º2
Zn2+ + 2º2 → ZnO4 → ZnO + 3/2º2
These processes are proved by the CII cathodic peak increasing, which is related with the anodic peak from +1.25 V/SCE, considering both the higher concentrations of Zn2+ and ORR products.
Another alternative considers the conversion of metallic zinc deposit into ZnO through peroxide ions generated by ORR, based on reaction (6):
Zn + ½O22− → ZnO + 1e
This process occurred at a more negative potential, at which zinc ions are reduced to metallic deposits that can be oxidized via peroxide generated from ORR, which is evidenced by CIII cathodic peak increasing. During anodic scanning, the first anodic peak (A0) appeared only in the presence of Zn_tpy at about −0.55 V vs. SCE, being a characteristic feature of the stripping of Zn deposited onto a GC electrode. At more positive scanning, the AI anodic peak increased due to the dissolution of metallic zinc and zinc oxide. A new anodic peak is evidenced at the potential value of about +0.65 V vs. SCE (A II), which can be attributed to the zinc oxide dissolution with the zincate formation (reaction (7)), which is controlled by the HO anions diffusion at the electrode surface.
ZnO + H2O + 2HO → Zn(OH)42−
To further elucidate several mechanistic aspects, the influence of the scan rate on CV shape recorded at Zn_tpy/GC electrode was studied. The family of CVs recorded for the scan rate, ranged between 0.01 to 0.20 Vꞏs−1, are presented in Figure 6. When the scan rate is ranged from 0.01 to 0.10 Vꞏs−1, the usual behaviour characteristics to the diffusion-controlled processes is observed for all cathodic and anodic peaks. However, the negative and low current of the A0 anodic peak was noticed, which did not vary linearly with the scan rate that informed us about the slow rate and the electron transfers-controlled process at the beginning of zinc stripping from zinc deposited onto the GC electrode. Moreover, the currents corresponding to the CIII cathodic peak linearly increased with the scan rate, which means the diffusion-controlled overall process of metallic zinc deposition and more overlapped cathodic processes, including hydrogen evolution. The linear dependences of the cathodic and anodic peaks currents with the square root of the scan rates are presented in Figure 7a, and all the ORR and zinc and zinc oxide stripping at more positive potentials are diffusion-controlled.
Very slight shifting of the oxidation and reduction peak potentials is found (Figure 7b), being practically independent of the voltammetry timescale, which informed us about the steady-state cathodic and anodic processes on the Zn_tpy/GC electrode.

4. Conclusions

A new metallomesogen based on a pentacoordinated Zn(II) metal centre was synthesized and characterized. Despite its bulky coordination, the coordination complex arranged into smectic layers starting from 89 °C. On cooling, the smectic phase is retained at room temperature and found to co-exist with the crystalline phase. An arrangement in the smectic phase is proposed based on X-ray scattering studies at both small- and wide-angles.
The electrochemical behaviour of the amphiphilic complex Zn_tpy, as a modifier of the GC electrode, was studied in comparison with its precursor Zn_p that lacks alkyl chains. The amphiphilic structure of the metallomesogen confers special properties for the complex regarding its electrochemical behaviour. The anodic and cathodic peaks identified for Zn_tpy/GC electrode in 0.1 M NaOH supporting electrolyte using cyclic voltammetry revealed the presence of redox systems for the whole potential range from −1.5 to +1.5 V vs. SCE. Within the negative potential range, the metallic zinc deposition and zinc oxide formation occurred during cathodic scanning at a potential value more negative than −0.55 V vs. SCE and zinc stripping was identified at a higher potential value. At a more positive potential range, higher than +0.65 V vs. SCE, the zinc oxide dissolution with the zincate formation controlled by the HO anions diffusion at the electrode surface occurred, which is overlapped at about +1.25 V vs. SCE with oxygen evolution reaction. This behaviour makes this complex suitable to be used in electrode modification/functionalization to confer it electrocatalytic activity vs. its dichloro precursor Zn_p that did not exhibit redox systems. The differences may derive from the different molecular structure of the two complexes; while in the case of Zn_p there is more likely a closer vicinity of the metal centres and their unavailability for involvement in the reduction and oxidation electrode processes, in the case of Zn_tpy, the bulky gallate ligands form an organic shell that keeps the metal centres more or less isolated, not permitting close contacts, which make them available for Zn reduction followed by its stripping. This can be a strategy to build performant sensing materials with nanoseparated electrochemically active centres.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app12168306/s1, Figure S1: 1H NMR spectra of complex Zn_tpy; Figure S2: TGA trace of complex Zn_tpy; Figure S3: DSC trace of complex Zn_tpy; Figure S4: S/WAXS pattern of complex Zn_tpy at 100 °C.

Author Contributions

Conceptualization, B.D., F.M. and E.I.S.; methodology, E.I.S.; investigation, B.H., A.A.A., E.P., S.M., S.I., S.N., C.C.; resources, E.I.S.; data curation, B.D., F.M. and E.I.S.; writing—original draft B.D., F.M. and E.I.S. and E.I.S.; writing—review and editing, B.D., F.M. and E.I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant of the Romanian Ministry of Education and Research, CNCS-UEFISCDI, project number PN-III-P4-ID-PCE-2020-1958, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the Romanian Academy, Program 4 and B.D. and B.H. thank CNRS and University of Strasbourg for support. This work was supported by a grant of the Ministry of Research, Innovation and Digitization, CNCS/CCCDI–UEFISCDI, project number PN-III-P4-ID-PCE-2020-1958, within PNCDI III.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Housecroft, C.E.; Constable, E.C. Solar energy conversion using first row d-block metal coordination compound sensitizers and redox mediators. Chem. Sci. 2022, 13, 1225–1262. [Google Scholar] [CrossRef]
  2. Bozic-Weber, B.; Constable, E.C.; Housecroft, C.E. Light harvesting with Earth abundant d-block metals: Development of sensitizers in dye-sensitized solar cells (DSCs). Coord. Chem. Rev. 2013, 257, 3089–3106. [Google Scholar] [CrossRef]
  3. Howarth, A.J.; Majewski, M.B.; Wolf, M.O. Photophysical properties and applications of coordination complexes incorporating pyrene. Coord. Chem. Rev. 2015, 282, 139–149. [Google Scholar] [CrossRef]
  4. Greenfield, T.J.; Julve, M.; Doyle, R.P. Exploring the biological, catalytic, and magnetic properties of transition metal coordination complexes incorporating pyrophosphate. Coord. Chem. Rev. 2019, 384, 37–64. [Google Scholar] [CrossRef]
  5. Pan, M.; Liao, W.-M.; Yin, S.-Y.; Sun, S.-S.; Su, C.-Y. Single-Phase White-Light-Emitting and Photoluminescent Color-Tuning Coordination Assemblies. Chem. Rev. 2018, 118, 8889–8935. [Google Scholar] [CrossRef]
  6. Kumar, N.; Bhalla, V.; Kumar, M. Beyond zinc coordination: Bioimaging applications of Zn(II)-complexes. Coord. Chem. Rev. 2021, 427, 213550. [Google Scholar] [CrossRef]
  7. Gupta, D.; Sathiyendiran, M. Rhenium-Carbonyl-Based Supramolecular Coordination Complexes: Synthesis, Structure and Properties. Chem. Sel. 2018, 3, 7439–7458. [Google Scholar] [CrossRef]
  8. Cargnelutti, R.; Schumacher, R.F.; Belladona, A.L.; Kazmierczak, J.C. Coordination chemistry and synthetic approaches of pyridyl-selenium ligands: A decade update. Coord. Chem. Rev. 2021, 426, 213537. [Google Scholar] [CrossRef]
  9. Marchetti, F.; Pettinari, C.; Di Nicola, C.; Tombesi, A.; Pettinari, R. Coordination chemistry of pyrazolone-based ligands and applications of their metal complexes. Coord. Chem. Rev. 2019, 401, 213069. [Google Scholar] [CrossRef]
  10. Malinowski, J.; Zych, D.; Jacewicz, D.; Gawdzik, B.; Drzezdzon, J. Application of Coordination Compounds with Transition Metal Ions in the Chemical Industry—A Review. Int. J. Mol. Sci. 2020, 21, 5443. [Google Scholar] [CrossRef]
  11. Chalkley, M.J.; Drover, M.W.; Peters, J.C. Catalytic N2-to-NH3 (or -N2H4) Conversion by Well-Defined Molecular Coordination Complexes. Chem. Rev. 2020, 120, 5582–5636. [Google Scholar] [CrossRef] [PubMed]
  12. Chylewska, A.; Biedulska, M.; Sumczynski, P.; Makowski, M. Metallopharmaceuticals in Therapy—A New Horizon for Scientific Research. Curr. Med. Chem. 2018, 25, 1729–1791. [Google Scholar] [CrossRef]
  13. Pervaiz, M.; Sadiq, S.; Sadiq, A.; Younas, U.; Ashraf, A.; Saeed, Z.; Zuber, M.; Adnan, A. Azo-Schiff base derivatives of transition metal complexes as antimicrobial agents. Coord. Chem. Rev. 2021, 447, 214128. [Google Scholar] [CrossRef]
  14. Doistau, B.; Jimenez, J.R.; Piguet, C. Beyond Chiral Organic (p-block) Chromophores for Circularly Polarized Luminescence: The Success of d-Block and f-Block Chiral Complexes. Front. Chem. 2020, 8, 555. [Google Scholar] [CrossRef] [PubMed]
  15. Maiuolo, L.; Calandra, P.; Lombardo, D.; Szerb, E.I. Amphiphiles-metals interactions for applications in modern technologies: Recent developments and future perspectives. Rev. Roum. Chim. 2020, 65, 647–671. [Google Scholar] [CrossRef]
  16. Cretu, C.; Maiuolo, L.; Lombardo, D.; Szerb, E.I.; Calandra, P. Luminescent Supramolecular Nano- or Microstructures Formed in Aqueous Media by Amphiphile-Noble Metal Complexes. J. Nanomater. 2020, 2020, 5395048. [Google Scholar] [CrossRef]
  17. Pellei, M.; Del Bello, F.; Porchia, M.; Santini, C. Zinc coordination complexes as anticancer agents. Coord. Chem. Rev. 2021, 445, 214088. [Google Scholar] [CrossRef]
  18. Yang, D.; Zhao, J.; Yang, X.-J.; Wu, B. Anion-coordination-directed self-assemblies. Org. Chem. Front. 2018, 5, 662–690. [Google Scholar] [CrossRef]
  19. Kameo, H.; Nakazawa, H. Recent Developments in the Coordination Chemistry of Multidentate Ligands Featuring a Boron Moiety. Chem. Asian J. 2013, 8, 1720–1734. [Google Scholar] [CrossRef]
  20. Zhang, X.; Hou, Y.; Xiao, X.; Chen, X.; Hu, M.; Geng, X.; Wang, Z.; Zhao, J. Recent development of the transition metal complexes showing strong absorption of visible light and long-lived triplet excited state: From molecular structure design to photophysical properties and applications. Coord. Chem. Rev. 2020, 417, 21337. [Google Scholar] [CrossRef]
  21. Dey, N.; Haynes, C.J.E. Supramolecular Coordination Complexes as Optical Biosensors. Chem. Plus Chem. 2021, 86, 418–433. [Google Scholar] [CrossRef] [PubMed]
  22. Cuerva, C.; Cano, M.; Lodeiro, C. Advanced Functional Luminescent Metallomesogens: The Key Role of the Metal Center. Chem. Rev. 2021, 121, 12966–13010. [Google Scholar] [CrossRef]
  23. Pastor, M.J.; Cuerva, C.; Fernández-Lodeiro, A.; Lodeiro, C.; Campo, J.A.; Cano, M. Designing Zn(II) complexes as a support of bifunctional liquid crystal and luminescent materials. Dyes Pigm. 2018, 149, 37–50. [Google Scholar] [CrossRef]
  24. Chakraborty, S.; Mondal, P.; Prasad, S.K.; Rao, D.S.S.; Bhattacharjee, C.R. Zinc(II)-salphen complexes bearing long alkoxy side arms: Synthesis, solvent dependent aggregation, and spacer group substituent effect on mesomorphism and photophysical property. J. Molec. Liq. 2017, 246, 290–301. [Google Scholar] [CrossRef]
  25. Chakraborty, S.; Mondal, P.; Prasad, S.K.; Rao, D.S.S.; Bhattacharjee, C.R. Induction of Mesomorphism through Supramolecular Assembly in Metal Coordination Compounds of “salphen”-Type Schiff Bases: Photoluminescence and Solvatochromism. Eur. J. Inorg. Chem. 2016, 28, 4604–4614. [Google Scholar] [CrossRef]
  26. Kumar, N.S.S.; Shafikov, M.Z.; Whitwood, A.C.; Donnio, B.; Karadakov, P.B.; Kozhevnikov, V.N.; Bruce, D.W. Mesomorphism and Photophysics of Some Metallomesogens Based on Hexasubstituted 2,2′:6′, 2″-Terpyridines. Chem. Eur. J. 2016, 22, 8215–8233. [Google Scholar] [CrossRef] [PubMed]
  27. Andelescu, A.A.; Heinrich, B.; Spirache, M.A.; Voirin, E.; La Deda, M.; Di Maio, G.; Szerb, E.I.; Donnio, B.; Costisor, O. Playing with PtII and ZnII Coordination to Obtain Luminescent Metallomesogens. Chem. Eur. J. 2020, 26, 4850–4860. [Google Scholar] [CrossRef]
  28. La Deda, M.; Di Maio, G.; Candreva, A.; Heinrich, B.; Andelescu, A.A.; Popa, E.; Voirin, E.; Badea, V.; Amati, M.; Costisor, O.; et al. Very intense polarized emission in self-assembled room temperature metallomesogens based on Zn(II) coordination complexes: An experimental and computational study. J. Mater. Chem. C 2022, 10, 115–125. [Google Scholar] [CrossRef]
  29. Cavero, E.; Uriel, S.; Romero, P.; Serrano, J.L.; Giménez, R. Tetrahedral Zinc Complexes with Liquid Crystalline and Luminescent Properties: Interplay Between Nonconventional Molecular Shapes and Supramolecular Mesomorphic Order. J. Am. Chem. Soc. 2007, 129, 11608–11618. [Google Scholar] [CrossRef]
  30. Pucci, D.; Crispini, A.; Ghedini, M.; La Deda, M.; Liguori, P.F.; Pettinari, C.; Szerb, E.I. ‘‘Green light’’ for Zn(II) mesogens. RSC Adv. 2012, 2, 9071–9078. [Google Scholar] [CrossRef]
  31. Keerthiga, R.; Kaliyappan, T.; Kannan, P. Studies on Twist Bent Core Zinc (II) Methacrylate Supramolecular Columnar Hexagonal Phase Mesogen derived from azobenzene moiety and its Photoluminescent Behaviours. Inorg. Chem. Comm. 2019, 99, 26–30. [Google Scholar] [CrossRef]
  32. Morale, F.; Date, R.W.; Guillon, D.; Bruce, D.W.; Finn, R.L.; Wilson, C.; Blake, A.J.; Schröder, M.; Donnio, B. Columnar mesomorphism from hemi-disklike metallomesogens derived from 2,6-bis[3′,4′,5′-tri(alkoxy)phenyliminomethyl]pyridines (L): Crystal and molecular structures of [M(L)Cl2 (M = Mn, Ni, Zn). Chem. Eur. J. 2003, 9, 2484–2501. [Google Scholar] [CrossRef] [PubMed]
  33. Bhattacharjee, C.R.; Das, G.; Mondal, P.; Prasad, S.K.; Rao, D.S.S. Novel Green Light Emitting Nondiscoid Liquid Crystalline Zinc(II) Schiff-Base Complexes. Eur. J. Inorg. Chem. 2011, 2011, 1418–1424. [Google Scholar] [CrossRef]
  34. Bhattacharjee, C.R.; Chakraborty, S.; Das, G.; Mondal, P. Emissive ‘zinc(II)-salphen’ core: Building block for columnar liquid crystals. Liq. Cryst. 2012, 39, 1435–1442. [Google Scholar] [CrossRef]
  35. Cuerva, C.; Ovejero, P.; Campo, J.A.; Cano, M. Tetrahedral and octahedral metallomesogenic Zn(II) complexes supported by pyridine-functionalised pyrazole ligands. New J. Chem. 2014, 38, 511–517. [Google Scholar] [CrossRef]
  36. Pucci, D.; Aiello, I.; Bellusci, A.; Crispini, A.; Ghedini, M.; La Deda, M. Coordination Induction of Nonlinear Molecular Shape in Mesomorphic and Luminescent ZnII Complexes Based on Salen-Like Frameworks. Eur. J. Inorg. Chem. 2009, 2009, 4274–4281. [Google Scholar] [CrossRef]
  37. Ames, K.A.; Collinson, S.R.; Blake, A.J.; Wilson, C.; Love, J.B.; Bruce, D.W.; Donnio, B.; Guillon, D.; Schröder, M. Design of Neutral Metallomesogens from 5,5-Dimethyldipyrromethane: Metal Ion Mediated Control of Folding and Hairpin Structures. Eur. J. Inorg. Chem. 2008, 2008, 5056–5066. [Google Scholar] [CrossRef]
  38. Giménez, R.; Manrique, A.B.; Uriel, S.; Barberá, J.; Serrano, J.L. Mesomorphism of a tetrahedral zinc complex. Chem. Commun. 2004, 18, 2064–2065. [Google Scholar] [CrossRef]
  39. Chakraborty, S.; Bhattacharjee, C.R.; Mondal, P.; Prasad, S.K.; Rao, D.S.S. Synthesis and aggregation behaviour of luminescent mesomorphic zinc(II) complexes with ‘salen’ type asymmetric Schiff base ligand. Dalton Trans. 2015, 44, 7477–7488. [Google Scholar] [CrossRef]
  40. Liu, Y.; Li, X.; Chen, J.; Yuan, C. Micro/Nano Electrode Array Sensors: Advances in Fabrication and Emerging Applications in Bioanalysis. Front. Chem. 2020, 8, 573865. [Google Scholar] [CrossRef]
  41. Hannah, S.; Damion, E.B.; Corrigan, K. Developments in Micro and Nanoscale Sensors for Biomedical Sensing. Curr. Opin. Electrochem. 2020, 23, 7–15. [Google Scholar] [CrossRef]
  42. Karimian, N.; Ugo, P. Recent advances in sensing and biosensing with arrays of nanoelectrodes. Curr. Opin. Electrochem. 2019, 16, 106–116. [Google Scholar] [CrossRef]
  43. Yu, J.; Huang, T.; Jiang, Z.; Sun, M.; Tang, C. Synthesis and Characterizations of Zinc Oxide on Reduced Graphene Oxide for High Performance Electrocatalytic Reduction of Oxygen. Molecules 2018, 23, 3227. [Google Scholar] [CrossRef] [PubMed]
  44. Sofianos, M.V.; Lee, J.; Silvester, D.S.; Samanta, P.K.; Paskevicius, M.; English, N.J.; Buckley, C.E. Diverse morphologies of zinc oxide nanoparticles and their electrocatalytic performance in hydrogen production. J. Energy Chem. 2020, 56, 162–170. [Google Scholar] [CrossRef]
  45. Kelly, S.R.; Shi, X.; Back, S.; Vallez, L.; Park, S.Y.; Siahrostami, S.; Zheng, X.; Norskov, J.K. ZnO as an Active and Selective Catalyst for Electrochemical Water Oxidation to Hydrogen Peroxide. ACS Catal. 2019, 9, 4593–4599. [Google Scholar] [CrossRef]
  46. Dayakar, T.; Rao, K.V.; Bikshalu, K.; Rajendar, V.; Park, S.H. Novel synthesis and structural analysis of zinc oxide nanoparticles for the non-enzymatic glucose biosensor. Mater. Sci. Eng. C 2017, 75, 1472–1479. [Google Scholar] [CrossRef]
  47. Ramya, B.; Priya, P.G. Rapid phytochemical microwave-assisted synthesis of zinc oxide nano flakes with excellent electrocatalytic activity for non-enzymatic electrochemical sensing of uric acid. J. Mater. Sci. Mater. Electron. 2021, 32, 21406–21424. [Google Scholar] [CrossRef]
  48. Motoc, S.; Cretu, C.; Costisor, O.; Baciu, A.; Manea, F.; Szerb, E.I. Cu(I) Coordination Complex Precursor for Randomized CuOx Microarray Loaded on Carbon Nanofiber with Excellent Electrocatalytic Performance for Electrochemical Glucose Detection. Sensors 2019, 9, 5353. [Google Scholar] [CrossRef]
  49. Motoc, S.; Schinteie, B.; Pop, A.; Negrea, S.; Cretu, C.; Szerb, E.I.; Manea, F. Graphene Quantum Dots and Cu(I) Liquid Crystal for Advanced Electrochemical Detection of Doxorubicine in Aqueous Solutions. Nanomaterials 2021, 11, 2788. [Google Scholar] [CrossRef]
  50. Szerb, E.I.; Pucci, D.; Crispini, A.; La Deda, M. Soft Luminescent Materials Based on Ag(I) Coordination Complexes. Mol. Cryst. Liq. Cryst. 2013, 573, 34–45. [Google Scholar] [CrossRef]
  51. Pucci, D.; Barberio, G.; Bellusci, A.; Crispini, A.; Donnio, B.; Giorgini, L.; Ghedini, M.; La Deda, M.; Szerb, E.I. Silver Coordination Complexes as Room-Temperature Multifunctional Materials. Chem. Eur. J. 2006, 12, 6738–6747. [Google Scholar] [CrossRef] [PubMed]
  52. The Cambridge Structural Database; Ma, Z.; Lu, W.; Liang, B.; Pombeiro, A.J.L. Synthesis, Characterization, Photoluminescent and Thermal Properties of Zinc(II) 4′-phenyl-terpyridine Compounds. New J. Chem. 2013, 37, 1529–1537. [Google Scholar] [CrossRef]
  53. Soliman, A.B.; Abdel-Samad, H.S.; Abdel-Rehim, S.S.; Hassan, H.H. Surface functionality and electrochemical investigations of a graphitic electrode as a candidate for alkaline energy conversion and storage devices. Sci. Rep. 2016, 6, 22056. [Google Scholar] [CrossRef] [PubMed]
  54. Elder, J.P. The Electrochemical Behavior of Zinc in Alkaline Media. J. Electrochem. Soc. 1969, 116, 757. [Google Scholar] [CrossRef]
  55. Dumitru, C.-S.; Cojocaru, A.; Anicai, L.; Cotarta, A.; Visan, T. Electrodeposition of zinc oxide films from choline chloride based ionic liquid media containing zinc and nitrate ions. UPB Sci. Bull. Ser. B 2016, 78, 59. [Google Scholar]
  56. Anwar, S.; Zhang, Y.; Khan, F. Electrochemical behaviour and analysis of Zn and Zn–Ni alloy anti-corrosive coatings deposited from citrate baths. RSC Adv. 2018, 8, 28861–28873. [Google Scholar] [CrossRef]
  57. Renuka, R.; Srinivasan, L.; Ramamurthy, S.; Veluchamy, A.; Venkatakrishnan, N. Cyclic voltammetric study of zinc and zinc oxide electrodes in 5.3 M KOH. J. Appl. Electrochem. 2001, 31, 655–661. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of the Zn(II) coordination complexes previously reported [27,28].
Figure 1. Chemical structure of the Zn(II) coordination complexes previously reported [27,28].
Applsci 12 08306 g001
Scheme 1. Synthesis of Zn_tpy: (i) ZnCl2, MeOH, r.t., 2 h; (ii) CHCl3, r.t., 6 h.
Scheme 1. Synthesis of Zn_tpy: (i) ZnCl2, MeOH, r.t., 2 h; (ii) CHCl3, r.t., 6 h.
Applsci 12 08306 sch001
Figure 2. (a) POM texture of Zn_tpy on cooling at 110 °C, magnification 20×; SWAXS patterns of Zn_tpy; (b) at 85 °C in the mesophase and (c) at room temperature in the pasty pseudo-crystalline state, showing coexistence of the signals of the crystal phase and smectic A phase.
Figure 2. (a) POM texture of Zn_tpy on cooling at 110 °C, magnification 20×; SWAXS patterns of Zn_tpy; (b) at 85 °C in the mesophase and (c) at room temperature in the pasty pseudo-crystalline state, showing coexistence of the signals of the crystal phase and smectic A phase.
Applsci 12 08306 g002
Figure 3. Idealized supramolecular representation of the Zn_tpy complexes in the SmA phase through self-assembly of aromatic and metallic moieties into monolayers (various possibilities of head-to-head arrangements are shown, with, in some cases, the possibility of larger interdigitation) alternating with molten chains (aliphatic chains continuum is shown in pale blue); dlam and Amol are the lamellar periodicity and molecular area, and Dar is the average separation between [Zn_tpy] moieties, Dar ~ √(Amol/π) (see Table 2).
Figure 3. Idealized supramolecular representation of the Zn_tpy complexes in the SmA phase through self-assembly of aromatic and metallic moieties into monolayers (various possibilities of head-to-head arrangements are shown, with, in some cases, the possibility of larger interdigitation) alternating with molten chains (aliphatic chains continuum is shown in pale blue); dlam and Amol are the lamellar periodicity and molecular area, and Dar is the average separation between [Zn_tpy] moieties, Dar ~ √(Amol/π) (see Table 2).
Applsci 12 08306 g003
Figure 4. Mesomorphic behaviour of Zn(II) coordination complexes: Cr—crystalline phase; SmA—smectic A mesophase; Mhex—three-dimensional mesophase with a two-dimensional hexagonal lattice symmetry, Colh: two-dimensional columnar hexagonal mesophase. DSC data on the second cooling cycle.
Figure 4. Mesomorphic behaviour of Zn(II) coordination complexes: Cr—crystalline phase; SmA—smectic A mesophase; Mhex—three-dimensional mesophase with a two-dimensional hexagonal lattice symmetry, Colh: two-dimensional columnar hexagonal mesophase. DSC data on the second cooling cycle.
Applsci 12 08306 g004
Figure 5. Cyclic voltammograms recorded in 0.1 M NaOH supporting electrolyte at a scan rate of 0.05 V⸱s−1 with GC electrode (black line), Zn_p/GC (red line) and Zn_tpy/GC electrode (blue line) within the potential range of −1.5 to +1.5 V/SCE; Inset: details of CVs, including the anodic and cathodic peaks marking.
Figure 5. Cyclic voltammograms recorded in 0.1 M NaOH supporting electrolyte at a scan rate of 0.05 V⸱s−1 with GC electrode (black line), Zn_p/GC (red line) and Zn_tpy/GC electrode (blue line) within the potential range of −1.5 to +1.5 V/SCE; Inset: details of CVs, including the anodic and cathodic peaks marking.
Applsci 12 08306 g005
Figure 6. CVs recorded at Zn_tpy/GC electrode in 0.1 M NaOH supporting electrolyte at various scan rates of: 0.01; 0.02; 0.03; 0.05; 0.06; 0.1; 0.2 Vꞏs−1 (curves 1–8) within potential range from −1.5 V to +1.5 V vs. SCE.
Figure 6. CVs recorded at Zn_tpy/GC electrode in 0.1 M NaOH supporting electrolyte at various scan rates of: 0.01; 0.02; 0.03; 0.05; 0.06; 0.1; 0.2 Vꞏs−1 (curves 1–8) within potential range from −1.5 V to +1.5 V vs. SCE.
Applsci 12 08306 g006
Figure 7. (a) Linear dependence of the peaks currents vs. square root of the scan rate; (b) Dependence of the peak potentials vs. the logarithm of the scan rate.
Figure 7. (a) Linear dependence of the peaks currents vs. square root of the scan rate; (b) Dependence of the peak potentials vs. the logarithm of the scan rate.
Applsci 12 08306 g007
Table 1. Thermal and mesomorphic data for Zn_tpy.
Table 1. Thermal and mesomorphic data for Zn_tpy.
Tdec5% [a] [°C]Mesophases [b], Transition Temperatures (°C) and Enthalpies
[ΔH (kJ·mol−1)] [c]
245Cr 89 [12.0] SmA 109 [2.5] Isso/Isso 114 [2.6] SmA 73 [10.2] Sm + Cr
[a] Decomposition temperature at 5% weight-loss (from TGA); [b] SmA: smectic A mesophase; [c] DSC data on second heating and cooling cycles.
Table 2. Geometrical mesophases parameters.
Table 2. Geometrical mesophases parameters.
Sample
Phase, T
Vmol
ρ
dlamNmolL
Amol
NarL
SZnAr
NchL
Sch
qch
Zn_tpy
SmA, 85
27,530.9929.312121.41
1889431.3
Zn_tpy
SmA, 100
278128.892121.43
0.98 1929632.1
T, temperature of the measurement (°C); Vmol, calculated molecular volume (Å3); ρ, density (g/cm3); dlam, lamellar periodicity (Å); NmolL, number of molecular layers per lamella; Amol = NmolL × Vmol/d, molecular area (Å2); NarL, number of aromatic layers per lamella (NarL = 1 for monolayer arrangement); SZnAr = (NarL/NmolL) × Amol, surface per Zinc complex (Å2); NchL, number of chain layers per lamella (NchL = 2 for non-interdigitated chains); Sch = (NchL/NmolL) × Amol/nch, surface per chain (Å2), nch = 6 being the number of peripheral chains per molecule; qch = Schch, chain packing ratio, σch = (26.5616 + 0.02023T)/1.27 (Å2) being the cross-sectional area of a molten chain [51].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Andelescu, A.A.; Ilies, S.; Cretu, C.; Popa, E.; Marinescu, S.; Heinrich, B.; Manea, F.; Negrea, S.; Donnio, B.; Szerb, E.I. Pentacoordinated Liquid Crystalline Zn(II) Complex Organized in Smectic Mesophase: Synthesis, Structural and Electrochemical Properties. Appl. Sci. 2022, 12, 8306. https://doi.org/10.3390/app12168306

AMA Style

Andelescu AA, Ilies S, Cretu C, Popa E, Marinescu S, Heinrich B, Manea F, Negrea S, Donnio B, Szerb EI. Pentacoordinated Liquid Crystalline Zn(II) Complex Organized in Smectic Mesophase: Synthesis, Structural and Electrochemical Properties. Applied Sciences. 2022; 12(16):8306. https://doi.org/10.3390/app12168306

Chicago/Turabian Style

Andelescu, Adelina A., Sorina Ilies (b. Motoc), Carmen Cretu, Evelyn Popa, Sorin Marinescu, Benoît Heinrich, Florica Manea, Sorina Negrea, Bertrand Donnio, and Elisabeta I. Szerb. 2022. "Pentacoordinated Liquid Crystalline Zn(II) Complex Organized in Smectic Mesophase: Synthesis, Structural and Electrochemical Properties" Applied Sciences 12, no. 16: 8306. https://doi.org/10.3390/app12168306

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