Next Article in Journal
Jacobi and Lyapunov Stability Analysis of Circular Geodesics around a Spherically Symmetric Dilaton Black Hole
Previous Article in Journal
Bonding of Dissimilar Metals in the Interlayer Region in Al-Based Composites: Molecular Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent Effects and Metal Ion Recognition in Several Azulenyl-Vinyl-Oxazolones

by
Mihaela Homocianu
1,
Anton Airinei
1,*,
Ovidiu-Teodor Matica
2,
Mihaela Cristea
3 and
Eleonora-Mihaela Ungureanu
2,*
1
Petru Poni Institute of Macromolecular Chemistry, Grigore Ghica Voda Alley, 41A, 700487 Iasi, Romania
2
Doctoral School of Chemical Engineering and Biotechnologies, Faculty of Chemical Engineering and Biotechnologies, University “Politehnica” of Bucharest, Gheorghe Polizu 1-7, Sector 1, 011061 Bucharest, Romania
3
“C.D. Nenitzescu” Institute of Organic and Supramolecular Chemistry, Romanian Academy, 060023 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Symmetry 2023, 15(2), 327; https://doi.org/10.3390/sym15020327
Submission received: 30 December 2022 / Revised: 16 January 2023 / Accepted: 20 January 2023 / Published: 24 January 2023
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
The spectral properties of several azulene-oxazolone derivatives containing a phenyloxazolone moiety linked to a substituted azulene ring via a C=C double bond were studied in different solvents of varying polarity. The solvatochromism and the ability of azulene-oxazolone derivatives to recognize heavy metal ions were investigated. In order to estimate the contribution of the non-specific and specific solute–solvent interactions, multiple linear regression analysis using Kamlet–Taft, Catalan and Laurence parameters was applied. These azulene derivatives demonstrate positive solvatochromism. The methyl and isopropyl substituents at the seven-membered azulene ring determine the highest red shifts of the absorption maxima of these azulenyl-vinyl-oxazolones. According to Catalan and Laurence models, the solvent polarizability is a more significant parameter in describing the solvatochromic properties of the azulene-oxazolone derivatives. The azulene-oxazolone compounds under study showed a good response to heavy metal cations (Cd2+, Hg2+, Cu2+ and Pb2+).

1. Introduction

Due to their unusual physicochemical properties (photophysical, electrochemical and photoelectrochemical properties), azulene derivatives present a broad range of applications in the research field of functional materials, including optical materials, electrochromic materials, conducting polymers, molecular switches, organic solar cells, molecular devices, and so on [1,2,3,4,5]. Furthermore, certain azulene derivatives exhibit antimicrobial, antifungal, anti-inflammatory and anticancer activity, and are utilized in photodynamic therapy [2,6,7].
The structure of azulene, a bicyclic nonalternant aromatic hydrocarbon (Figure 1a), consists of an electron-deficient seven-membered ring (tropylium cation) fused with an electron rich five-membered ring (cyclopentadienyl anion). Its polarity is reflected by the presence of a relatively large dipole moment (1.08 D) (Figure 1b). Even though both skeleton isomers, azulene and naphthalene, have the same composition (C8H10) and degree of unsaturation (five double bonds and two fused rings), azulene has distinct absorption bands and electronic transitions compared to its naphthalene isomer (Figure 2a). By varying the nature of the electronic substituents grafted on the azulene moiety in the positions where the electron density is the highest, the wavelength of the light absorbed can be finely tuned [8] and significant changes of the optical and photophysical properties have been observed [9,10,11].
Azulene exhibits a weak absorption band in the visible range which causes the blue color of this compound [9]. Due to a small energy gap between the ground state (S0) and the first excited state (S1), this weak absorption corresponds to a transition from the ground state (S0) to the first excited state (S1) (S0 → S1 transition), making azulene become the smallest isolable organic compound [12,13].
The literature data show that while electron-withdrawing groups (EWG) present at the C-1 and C-3 positions increase the S0–S1 gap by stabilizing the HOMO and LUMO+1 orbitals, having no effect on the LUMO orbital, electron-donating groups (EDG) at these positions destabilize the HOMO and LUMO+1 orbitals, reducing the S0–S1 gap [14]. Additionally, azulene presents a strong absorption band related to a transition from S0 to the second excited state (S2). In addition to these absorption bands, a third strong band showing an S0 → S3 transition was observed for azulene in the ultraviolet range [1,15,16,17]. In this way, the absorption bands of azulenes can be connected with four distinct spectral regions: λ > 500 nm (S0 → S1 transition), 350 nm < λ < 500 nm (S0 → S2 transition), λ = 250–350 nm (S0 → S3 transition), and λ > 250 nm (S0 → S4 transition) [16,17].
Azulene is one of the few chromophores that present a strong emission from the S2 excited singlet state to S0 at room temperature, distinctly from the great majority of aromatic compounds which present the S1 → S0 emission. This anomalous situation can be due to the low-lying first excited state S1 and a large energy gap between S1 and S2 levels which contribute to the decrease of the radiationless transition rate between S2 and S1 [1,17,18]. Additionally, the anomalous emission from the S2 state is determined by the fact that the energy gap between S2 and S1 states is over 10,000 cm−1 (cf. 14,000 cm−1 in azulene), while a mixture of S2 and S1 emissions occurs if the S2 → S1 energy gap has values between 9000 and 10,000 cm−1 [19]. If the gap is below 9000 cm−1, the internal conversion rate between S1 and S2 is higher than the relaxation rate from S2 and the emission takes place from S1 [1,3,16]. The nature of azulene fluorescence is highly dependent on substitution at the azulene moiety, which affects whether the emission is dominant from the S2 or S1 singlet excited state, except for azulene aldehydes whose emission from the S3 and S2 states depends on the nature of the hydrogen bonds present in the system [20,21].
Due to the ability of solvents to modify the physicochemical properties of molecules, the investigation of the solvent influence on the spectral characteristics of azulene derivatives can provide information about the inter- and intramolecular interactions with the solvent and the excited state energy. The solvent influence is associated with a change in polarizability, polarity, dielectric constant and viscosity of the surrounding microenvironment. The dependence of the electronic absorption spectra, mainly the position and intensity of the absorption bands, on the solvent polarity can be described by the interaction of the chromophoric groups with solvent through non-specific (dipolarity/polarizability) and specific (H-bonding, acceptor–donor) interactions [22,23,24,25]. The solvent effects were studied using different solvatochromic parameters and polarity scales, such as the empirical solvent polarity parameter ET [26], the Catalan solvent scale, the Kamlet–Taft scale or the Laurence solvent scale, which were applied to investigate the nature of solute–solvent interactions for a given compound [24,25,26,27,28,29,30,31,32].
In this paper, a detailed investigation of the effects of solvents and substituents on the electronic absorption spectra of some azulenyl-vinyl-oxazolones (Figure 3) was performed. Each investigated compound can act as a ligand for heavy metal (HM) ions. Table S1 lists the codes and names of the investigated compounds. They contain the phenyloxazolone moiety attached to the substituted azulene by a C=C double bond. The azulene moiety is either substituted with electron-donating alkyl groups (4,6,8-Me3, 5-iPr-3,8-Me2) or unsubstituted, whereas the phenyloxazolone moiety is either substituted with the electron-withdrawing nitro group or unsubstituted. This polarized structure ensures the increased degree of asymmetry of these azulene compounds by increasing the dipole moment as compared to unsubstituted 4-(azulen-1-ylmethylene)-2-phenyloxazol-5(4H)-one.
The spectral response of the samples under study was discussed for solvents of different polarities. The contribution of the non-specific and specific solute–solvent interactions for these azulene derivatives was analyzed using the Kamlet–Taft, Catalan and Laurence solvent scales.

2. Materials and Methods

The spectral studies were performed in 1,4-dioxane, tetrachloromethane (CCl4), toluene, chloroform (CLF), ethyl acetate (EtAc), dichloromethane (DCM), dichloroethane (DCE), acetone, methanol, acetonitrile (ACN), dimethylformamide (DMF), and dimethyl sulfoxide (DMSO). All the solvents were of spectroscopic grade and were commercially acquired from Sigma-Aldrich. The azulene derivatives (Table S1) were prepared following a previously reported synthetic method based on the condensation of azulene-1-carbaldehydes with hippuric acids by Erlenmeyer–Plöchl synthesis [33] (Figure 3). Copper sulfate pentahydrate, cadmium sulfate hydrate, lead(II) acetate trihydrate and mercury(II) chloride (Sigma-Aldrich) were utilized as received. A SPECORD 210 Plus spectrometer (Analytik Jena, Germany) was used to record the electronic absorption spectra in solution. Quartz cells with an optical path length of 10 mm were used for measurements.
The spectral properties of azulene derivatives were investigated by titration experiments monitoring the intensity of the absorption bands at longer wavelengths upon the addition Cd2+, Cu2+, Pb2+ and Hg2+ cations. Stock solutions of each metal cation were freshly prepared before each measurement at a concentration of 10−3 mol/L in acetonitrile. To perform the absorption titration experiments, a solution (2.5 mL) of azulene derivative in acetonitrile was introduced into a quartz cuvette and then increasing volumes of each metal ion (1–1500 μL) were added. After mixing, the electronic absorption spectra were recorded.

3. Results and Discussion

To investigate the solvent effect on the electronic absorption spectra of azulene derivatives, the measurements were performed in solvents of varying polarities. The UV–Vis absorption spectra of the studied azulenes consist of distinct absorption ranges located around 400–600 nm (S0 → S2 transition), 300–400 nm (S0 → S3 transition) and 300–220 nm (S0 → S4 transition). The transition S0 → S1 appearing at wavelengths over 600 nm has a very low intensity, and it is not evident in the spectra (Figure 4 and Figure S1). The absorption bands in the range 300–600 nm can be attributed to a ππ* transition being localized on the whole system.
When the solvent polarity changes from dioxane (ε = 2.22) to dimethyl sulfoxide (DMSO) (ε = 47.24), the maximum of the absorption band at 405 nm for O1 shifts to longer wavelengths of about 492 nm (Table 1). The introduction of the electron donating methyl groups in the 4,6,8 positions and the methyl groups in the 3,8 positions and an isopropyl group in 5 position, leads to a greater red shift, namely 758 cm−1 and 887 cm−1, for O2 and O3, respectively, compared to O1. The substitution of the nitro electron-withdrawing group at the para position of the phenyl ring of oxazolone induces a lower red shift for compounds O5 (588 cm−1) and O6 (429 cm−1) due to the higher asymmetry degree in these compounds determined by the nitro group. In this way, by changing the position of the substituents on the azulene ring, the azulene compounds that present colors covering the visible range to a great extent (Figure 4 and Figure S1) can be obtained, giving the possibility of tuning the color of azulene derivatives. The absorption bands corresponding to the S0 → S3 transition are of lower intensity and are poorly defined (Figure 4). Although, most azulene compounds exhibit strong fluorescence, the azulene derivatives under study do not show fluorescence.
To gain an understanding of the solvatochromic behavior of the azulene derivatives, the Lippert–Mataga and Bakhshiev solvent polarity functions were utilized in a simple approximation, according to the relations (1) and (2) [24,26,34,35].
νmax = ν0 + 2(μe−μg)2/hca3FLM(n,ε) + const
νmax = ν0 + 2(μe−μg)2/hca3FB(n,ε) + const
In (1) and (2), νmax denotes the absorption maximum wavenumber; ν0 is the spectral position of the absorption maximum wavenumber in the gas phase; n is the refractive index; ε is the dielectric constant of the solvent; μg and μe represent the ground and excited state dipole moments, respectively; h is the Planck constant; c is the light velocity in vacuum; a is the Onsager cavity radius; and FLM(n,ε) and FB(n,ε) are the solvent polarity functions given by expressions (3) and (4), respectively.
FLM(n,ε) = (ε − 1)/(2ε + 1)−(n2 − 1)/(2n2 + 1)
FB(n,ε) = {(ε − 1)/(ε + 2)−(n2 − 1)/(n2 + 2)}(2n2 + 1)/(n2 + 2)
The plots FLM(n,ε) and FB(n,ε) as a function of νmax do not show a linear trend, but present a high dispersion of the data (Figures S2 and S3), which denotes that the specific interactions must be taken into consideration for understanding the solvent effects.
In recent years, many multiparameter solvent polarity scales have been developed to quantitatively describe solvent–solute interactions, resulting in a good linear correlation between spectral data and solvent polarity parameters. Linear solvation energy relationship analysis was applied using the Kamlet–Taft [29], Catalan [30] and Laurence [31] models in order to explore the solvent influence on the electron absorption spectra of the investigated azulene derivatives. The linear solvation energy relationships that describe the above-mentioned empirical solvent scales can be formulated through Equations (5)–(7), respectively [28,29,30,31,32,36].
ν max = ν max , 0 + a α α + b β β + c π π
In (5), νmax represents the wavenumber in the absorption maximum; νmax,0 is the value of the wavenumber in the gas phase; π* is a measure of the dipolarity/polarizability of the solvent; α and β denote the solvent hydrogen bond donor acceptor (HBD) and solvent hydrogen bond acceptor basicity (HBA), respectively; and aα, bβ and cπ are regression coefficients.
ν max = ν max , 0 + a SA SA + b SB SB + c SP SP + d SdP SdP
In (6), SA, SB, SP and SdP are parameters that refer to acidity, basicity, polarizability and dipolarity of the solvents, respectively, and aSA, bSB, cSP and dSdP are regression coefficients.
ν max = ν max , 0 + a DI DI + b ES ES + c α 1 α 1 + d β 1 β 1
In (7), DI describes the dispersion and induction interactions, ES denotes the electrostatic interactions between permanent dipoles of the solute and the solvent, α1 and β1 are HBD and HBA abilities of the solvents, and aDI, bES, cα1 and dβ1 represent the corresponding regression coefficients.
The Kamlet–Taft (α, β and π*), Catalan (SA, SB, SP and SdP) and Laurence (D1, ES, α1 and β1) solvatochromic parameters used in multilinear regression analysis are given in Table 2 [24,29,30,37]. The solvent-independent correlation coefficients from Equations (5)–(7) are estimated using multiple linear regression analysis, and they reflect the relative contributions of each solvent factor on the spectral shift in νmax.
The results of the multiple regression analysis using Equations (5)–(7) are illustrated in Table 3. The multilinear analysis of the absorption data according to the Kamlet–Taft solvatochromic model shows a poor correlation, evidenced by the low values of R2 for all azulene derivatives (Table 3). This weak correlation can be due to the fact that in the Kamlet–Taft equation, the non-specific effects are included in a single parameter π*. In the Catalan model, these effects are split, and a very good fit was obtained (R2 = 0.94 for O1). It is clear from Table 3 that the Catalan model gives better correlations than the Laurence model, as seen from the slightly higher values of R2. The percentage contributions of Catalan and Laurence solvent parameters on the absorption spectral shifts of the azulene derivatives are given in Table 4. The multilinear fit of νmax using SP, SdP, SA and SB parameters (Equation (6)) shows that the solvent polarizability parameter (SP) dominates the solvatochromism with a lower contribution of solvent dipolarity and acidity. The solvent polarizability has a major effect on the absorption energy of all studied azulenes with percentage contributions of 73.89%, 62.48%, 71.65%, 67.62% and 65.67%, respectively (Table 4). In this way, the specific interactions (α + β) have around 20% contributions to the solvent–solute interactions. According to the Catalan scale analysis, a good correlation was obtained for O2 (R2 = 0.92) with the largest impact of solvent polarizability (71.65%), but compared to O5 and O6, the correlation is less good (R2 = 0.86 and 0.74, respectively), with a higher contribution (30.28%) of the specific interactions for O6. This shows that the introduction of the nitro group in the p-position of the phenyl ring of the oxazolone determines a reduction of the non-specific interactions at the position of absorption maxima.
The multiple linear regressions in Figure 5, Figures S4 and S5 reveal a good correlation between the experimental and calculated data using Equation (6). In the fitting relations to the absorption spectral data (Table 3), the coefficients a, c and d have a negative sign for all azulene derivatives, which suggests a red shift of the absorption maxima with the increase of solvent polarizability and dipolarity. At the same time, the negative coefficients associated with SP, SdP and SA confirm the stabilization of the excited state compared to the ground state [38].
From the multilinear correlations using the Laurence solvent scale (Equation (7)), similar results were obtained. The fitting equations relating to the absorption data reveal that the non-specific interactions, given by the dispersion and induction contributions (DI ~75%) and the electrostatic interactions (ES ~15%), are the most important factors responsible for the red shift of the absorption band, while the specific interactions have a low influence for all compounds (Table 4).
Given the outstanding binding characteristics of azulene compounds with metal ions, they can be used for heavy metal ion detection [39,40]. The metal ion recognition behavior of the azulene derivatives was evaluated upon their titration by successive increments of metal ion, and the changes in the electronic absorption spectra were monitored. The evolution of the UV–Vis spectra of O2 and O5 derivatives with the four metal ions at different concentrations is depicted in Figure 6, Figure 7, Figures S6 and S7. The addition of each cation caused the progressive decrease in intensity of the absorption bands of azulene derivatives (Figure 6, Figure 7, Figures S6 and S7), with the exception of the absorption band in the 250–270 nm range where, during titration, an increase in absorbance was noticed due to the absorption of metal salts [41] (Figure 7b inset). A good linear relationship (R2 = 0.99) was obtained for the plot of absorbance of the longer wavelength band versus the metal ion concentration for O2 during titration with Cd2+ ions over the range of 2.71 × 10−5 to 5.35 × 10−4 M Cd2+ content (Figure 8). The decrease in intensity of the absorption band at longer wavelengths with adding increasing contents of metal ions was seen to be 35.5% for Pb2+, 32.6% for Cu2+, 30.4% for Cd2+ and 32.7% for Hg2+, taking into account the ratio Af/Ao, where Ao is the absorbance before the addition of metal ion and Af is the absorbance corresponding to the constant value after the addition of metal ion.

4. Conclusions

The aim of this work was to investigate the spectral characteristics of a series of structurally different azulene-oxazolone derivatives in various solvents and their spectroscopic absorption behavior in the presence of Cd2+, Hg2+, Cu2+ and Pb2+ ions. Each investigated compound contains a phenyloxazolone moiety linked to a substituted azulene ring through a C=C double bond. The azulene moiety substituted with electron-donating alkyl groups and the phenyloxazolone moiety substituted with an electron-withdrawing nitro group provides an increased degree of asymmetry of these azulene compounds compared to unsubstituted 4-(azulen-1-ylmethylene)-2-phenyloxazol-5(4H)-one. The solvent-dependent characteristics of some azulene-oxazolone derivatives were analyzed using Kamlet–Taft, Catalan and Laurence solvent scales. Catalan and Laurence multiparametric solvatochromic analysis showed that the solvent polarizability parameters (SP + SdP, α1 + β1) have a greater impact (over 80%) on the spectral shifts of the absorption maxima of the investigated azulenyl-vinyl-oxazolones. However, the specific interactions are quite important in the solvatochromic behavior. The solvatochromic response also depends on the position of the substituents from the azulenic ring. These results might be useful for the design of new solvatochromic compounds and to understand the nature of the solvent–solute interactions given by this polarized structure. In the presence of Cd2+, Hg2+, Cu2+ and Pb2+ metal ions, the main absorption band intensity decreases by about 30% and exhibits a linear dependence on the concentration of metal ions, showing a good recognition ability of these ions at 10−4 M.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym15020327/s1, Basic properties for ligands and characterization by elemental analysis, UV–Vis, 1H NMR, 13C NMR, IR, MS; Lippert–Mataga equation; Bakhshiev equation; Figure S1: UV–Vis absorption spectra of azulene derivative in solvents of different polarities for (a) O3 and (b) O6; Figure S2: Lippert–Mataga solvent polarity function (FLM) plots for (a) O1 and (b) O6; Figure S3: Bakhshiev solvent polarity function plots for (a) O1 and (b) O6; Figure S4: Linear correlation plots using Equation (6) for (a) O2 and (b) O5; Figure S5: Linear correlation plots using Equation (7) for (a) O1 and (b) O2; Figure S6: Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of (a) Pb2+ for O2 and (b) Cu2+ for O5; Figure S7: Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of Hg2+ (a) for O2 and (b) for O5.

Author Contributions

Conceptualization, E.-M.U. and A.A.; methodology, E.-M.U.; validation, E.-M.U. and A.A.; formal analysis, M.H., O.-T.M. and M.C.; investigation, M.H. and O.-T.M.; resources, E.-M.U.; data curation, A.A. and M.C.; writing—original draft preparation, A.A. and M.H.; writing—review and editing, E.-M.U. and A.A.; visualization, A.A.; supervision, E.-M.U.; project administration, E.-M.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xin, H.; Gao, X. Application of azulene in constructing organic optoelectronic materials: New tricks for an old dog. ChemPlusChem 2017, 82, 945–956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ito, A.; Amaki, T.; Ishii, A.; Fukuda, K.; Yamasaki, R.; Okamoto, I. Conformational analysis of N-aryl-N-(2-azulenyl)acetamides. Tetrahedron Lett. 2018, 59, 3994–3998. [Google Scholar] [CrossRef]
  3. Dong, J.X.; Zhang, H.L. Azulene-based organic functional molecules for optoelectronics. Chin. Chem. Lett. 2016, 27, 1097–1104. [Google Scholar] [CrossRef]
  4. Tao, T.; Fan, Y.; Zhao, J.; Yu, J.; Chen, M.; Huang, W. Reversible alteration of spectral properties for azulene decorated multiphenyl-ethylenes by simple acid base and redox processes. Dye. Pigment. 2019, 164, 346–354. [Google Scholar] [CrossRef]
  5. Yang, C.C.; Ma, J.Y.; Su, X.; Zheng, X.L.; Chen, H.; He, Y.Y.; Tian, W.Q.; Li, W.Q.; Yang, L. High performance nonlinear materials with simple aromatic hydrocarbons. Flatchem 2022, 33, 100362. [Google Scholar] [CrossRef]
  6. Damrongrungruang, T.; Rattanayatikul, S.; Sontikan, N.; Wutirak, B.; Teerakapong, A.; Kaewrawang, A. Effect of different irradiation modes of azulene-mediated photodynamic therapy on singlet oxygen and PGE2 formation. Photochem. Photobiol. 2021, 97, 427–434. [Google Scholar] [CrossRef]
  7. Leino, T.O.; Sieger, P.; Yli-Kauhaluoma, J.; Walen, E.A.A.; Kley, T. The azulene scaffold from a medicinal chemist’s perspective: Physicochemical and in vitro parameters relevant for drug discovery. Eur. J. Med. Chem. 2022, 237, 114379. [Google Scholar] [CrossRef]
  8. Liu, R.S.H.; Muthyala, R.S.; Wang, X.S.; Asato, A.E.; Wang, P.; Ye, C. Correlation of substituent effects and energy levels of the two lowest excited states of the azulenic chromophore. Org. Lett. 2000, 2, 269–271. [Google Scholar] [CrossRef]
  9. Murfin, L.C.; Lewis, S.E. Azulene—A bright core for sensing and imaging. Molecules 2021, 26, 353. [Google Scholar] [CrossRef]
  10. Shevyakov, S.V.; Li, H.; Muthyala, R.; Asato, A.E.; Croney, J.C.; Jameson, D.M.; Liu, R.S.H. Orbital control of the color and excited state properties of formylated and fluorinated derivatives of azulene. J. Phys. Chem. A 2003, 107, 3295–3299. [Google Scholar] [CrossRef]
  11. Razus, A.C. Azulene moiety as electron reservoir positively charged systems; A short survey. Symmetry 2021, 13, 526. [Google Scholar] [CrossRef]
  12. Lash, T.D. Out of the blue! Azuliporphyrins and related carbaporphyrinoid systems. Acc. Chem. Res. 2016, 49, 471–482. [Google Scholar] [CrossRef] [PubMed]
  13. Dias, J.R. Electronic and structural properties of biazulene, terazulene and polyazulenes isomers. J. Phys. Org. Chem. 2007, 20, 395–409. [Google Scholar] [CrossRef]
  14. Liu, R.S.H. Colorful azulene and its equally colorful derivatives. J. Chem. Educ. 2002, 79, 183–185. [Google Scholar] [CrossRef]
  15. Ruth, A.A.; Kim, E.K.; Hese, A. The So→S1 cavity ring-down absorption spectrum of jet-cooled azulene dependence of internal conversion on the excess energy. Phys. Chem. Chem. Phys. 1999, 1, 5121–5128. [Google Scholar] [CrossRef]
  16. Koch, M.; Blacque, O.; Venkatesan, K. Impact of 2,6-connectivity in azulene: Optical properties and stimuli responsive behavior. J. Mater. Chem. C 2013, 1, 7400–7408. [Google Scholar] [CrossRef] [Green Version]
  17. Foggi, P.; Neuwahl, F.V.R.; Moroni, L.; Salvi, P.R. S1→Sn and S2→Sn absorption of azulene: Femtosecond transient spectra and excited state calculation. J. Phys. Chem. A 2003, 107, 1689–1696. [Google Scholar] [CrossRef]
  18. Okamoto, M.; Hirayama, S.; Steer, R.P. A reinterpretation of the unsusual barochromism of azulene. Can. J. Chem. 2007, 85, 432–437. [Google Scholar] [CrossRef]
  19. Itoh, T. Fluorescence and phosphorescence from higher excited states of organic molecules. Chem. Rev. 2012, 112, 4541–4568. [Google Scholar] [CrossRef]
  20. Zhou, Y.Y.; Baryshnikov, G.; Li, X.P.; Zhu, M.J.; Ågren, H.; Zhu, L.L. Anti-Kasha’s rule emissive switching induced by intermolecular H-bonding. Chem. Mater. 2018, 30, 8008–8016. [Google Scholar] [CrossRef]
  21. Gong, Y.; Zhou, Y.; Yue, B.; Wu, B.; Sun, R.; Qu, S.; Zhu, L. Multiwavelength anti-Kasha’s rule emission on self-assembly of azulene-functionalized persulfurated arene. J. Phys. Chem. C 2019, 123, 22511–22518. [Google Scholar] [CrossRef]
  22. Homocianu, M.; Airinei, A.; Dorohoi, D.O. Solvent effects on the electronic absorption and fluorescence spectra. J. Adv. Res. Phys. 2011, 2, 011105. [Google Scholar]
  23. Airinei, A.; Isac, D.I.; Homocianu, M.; Cojocaru, C.; Hulubei, C. Solvatochromic analysis and DFT computational study of an azomaleimide derivative. J. Mol. Liq. 2017, 240, 476–485. [Google Scholar] [CrossRef]
  24. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  25. Bakhshiev, N.G. Spectroscopy of Intermolecular Interactions; Nauka: Saint Petersburg, Russia, 1972. (In Russian) [Google Scholar]
  26. Bakhshiev, N.G.; Gubaryan, S.K.; Dobretsov, G.E.; Kirillova, A.Y.; Svetlichnyi, V.Y. Solvatochromism and solvatofluorochromism of the intramolecular charge transfer band of 4-dimethylaminochalcone in the electronic spectra of its solutions. Opt. Spectrosc. 2006, 100, 700–708. [Google Scholar] [CrossRef]
  27. Reichardt, C. Solvatochromic dyes as solvent polarity indicators. Chem. Rev. 1994, 94, 2319–2358. [Google Scholar] [CrossRef]
  28. Kamlet, M.J.; Abboud, J.L.; Taft, R.W. The solvatochromic comparison method. 6. The π* scale of solvent polarities. J. Am. Chem. Soc. 1977, 90, 6027–6038. [Google Scholar] [CrossRef]
  29. Kamlet, M.J.; Abboud, J.L.; Abraham, M.H.; Taft, R. Linear solvation energy relationships. 23. A comprehensive collection of the solvatochromic parameters π*, α and β, and some methods for simplifying the generalized solvatochromic method. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  30. Catalan, J. Toward a generalized treatment of the solvent effect on the empirical scales: Dipolarity (SdP, a new scale), polarizability (SP), acidity (SA) and basicity (SB) of the medium. J. Phys. Chem. B 2009, 113, 5951–5960. [Google Scholar] [CrossRef]
  31. Catalan, J.; Hopf, H. Empirical components of solute-solvent interactions: The polarizability solvent scale. Eur. J. Org. Chem. 2004, 2004, 4694–4702. [Google Scholar] [CrossRef]
  32. Laurence, C.; Legros, J.; Chantzis, A.; Phanchat, A.; Jacquemin, D. A database of dispersion-induction DI, electrostatic ES, and hydrogen bonding α1 and β1 solvent-parameters and some applications to the multiparameter correlation analysis of solvent effects. J. Phys. Chem. B 2015, 119, 3174–3184. [Google Scholar] [CrossRef]
  33. Cristea, M.; Birzan, L.; Dumitrascu, F.; Enache, C.; Tecuceanu, V.; Hanganu, A.; Draghici, C.; Deleanu, C.; Nicolescu, A.; Maganu, M.; et al. 1-Vinylazulenes with oxazolonic ring-potential ligands for metal ion detectors; Synthesis and product properties. Symmetry 2021, 13, 1209. [Google Scholar] [CrossRef]
  34. Lippert, E. Dipolmoment und Electronenstruktur von angeregten Molekulen. Z. Für Nat. A Phys. Sci. 1955, 10, 541–545. [Google Scholar]
  35. Mataga, N.; Kaifu, Y.; Koizumi, M. Solvent effects upon fluorescence spectra and dipolmoments of excited states. Bull. Chem. Soc. Jpn. 1956, 29, 465–470. [Google Scholar] [CrossRef] [Green Version]
  36. Kamlet, M.J.; Abboud, J.L.M.; Taft, R.W. An examination of linear solvation energy relationships. Progr. Phys. Org. Chem. 1981, 13, 485–630. [Google Scholar]
  37. Marcus, Y. The properties of organic liquids that are relevant to their use as solvating solvents. Chem. Soc. Rev. 1993, 22, 409–416. [Google Scholar] [CrossRef]
  38. Rauf, M.A.; Hisaindee, S. Studies on solvatochromic behavior of dyes using spectral techniques. J. Mol. Struct. 2013, 1042, 45–56. [Google Scholar] [CrossRef]
  39. Birzan, L.; Cristea, M.; Draghici, C.C.; Tecuceanu, V.; Maganu, M.; Hanganu, A.; Arnold, G.L.; Ungureanu, E.M.; Razus, A.C. 1-Vinylazulenes–potential host molecules in ligands for metal ion detectors. Tetrahedron 2016, 72, 2316–2326. [Google Scholar] [CrossRef]
  40. Arnold, G.L.; Lazar, I.G.; Ungureanu, E.M.; Buica, G.O.; Birzan, L. New azulene modified electrodes for heavy metal ions recognition. Bulg. Chem. Commun. 2017, 49, 205–210. [Google Scholar]
  41. Muravev, A.; Yakupov, A.; Gerasimova, T.; Islamov, D.; Lazarenko, V.; Shokurov, A.; Ovsyannikov, A.; Dorovaovskii, P.; Zubavichus, Y.; Naummkin, A.; et al. Thiacalixarenes with sulfur functionalized at lower rim: Heavy metal ion binding in solution and 2D-confined space. Int. J. Mol. Sci. 2022, 23, 2341. [Google Scholar] [CrossRef]
Figure 1. (a) Alternant nature of naphthalene and nonalternant nature of azulene; (b) azulene and its polarized resonance structure.
Figure 1. (a) Alternant nature of naphthalene and nonalternant nature of azulene; (b) azulene and its polarized resonance structure.
Symmetry 15 00327 g001
Figure 2. (a) Unique fluorescence observed for naphthalene and azulene. (b) Probability of locating the electron (squares of coefficients of each wave−function) in the HOMO, LUMO, and LUMO+1 of azulene (left) and naphthalene (right).
Figure 2. (a) Unique fluorescence observed for naphthalene and azulene. (b) Probability of locating the electron (squares of coefficients of each wave−function) in the HOMO, LUMO, and LUMO+1 of azulene (left) and naphthalene (right).
Symmetry 15 00327 g002
Figure 3. Chemical structure of investigated azulenyl-vinyl-oxazolones: R1 = H; 4,6,8-Me3; 5-iPr-3,8-Me2; R2 = H; NO2.
Figure 3. Chemical structure of investigated azulenyl-vinyl-oxazolones: R1 = H; 4,6,8-Me3; 5-iPr-3,8-Me2; R2 = H; NO2.
Symmetry 15 00327 g003
Figure 4. UV–Vis absorption spectra in different solvents: (a) O3, (b) O6.
Figure 4. UV–Vis absorption spectra in different solvents: (a) O3, (b) O6.
Symmetry 15 00327 g004
Figure 5. Linear relationship between experimental and calculated absorption maxima for (a) O3 using Equation (6) and (b) O3 using Equation (7).
Figure 5. Linear relationship between experimental and calculated absorption maxima for (a) O3 using Equation (6) and (b) O3 using Equation (7).
Symmetry 15 00327 g005
Figure 6. Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of Cd2+ (a) for O2 and (b) for O5.
Figure 6. Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of Cd2+ (a) for O2 and (b) for O5.
Symmetry 15 00327 g006
Figure 7. Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of (a) Cu2+ for O2 and (b) Pb2+ for O5. The inset shows the enlarged view of the 220–270 nm range.
Figure 7. Changes in the UV–Vis absorption spectra in acetonitrile solution upon the addition of increasing amounts of (a) Cu2+ for O2 and (b) Pb2+ for O5. The inset shows the enlarged view of the 220–270 nm range.
Symmetry 15 00327 g007
Figure 8. Plot of the absorption change of O2 as a function of Cd2+ content.
Figure 8. Plot of the absorption change of O2 as a function of Cd2+ content.
Symmetry 15 00327 g008
Table 1. Spectroscopic data of the investigated azulenyl-vinyl-oxazolones in different solvents (νmax, cm−1).
Table 1. Spectroscopic data of the investigated azulenyl-vinyl-oxazolones in different solvents (νmax, cm−1).
SolventAzulene Derivative
O1O2O3O5O6
1,4-Dioxane21,50020,96020,08020,08020,080
CCl421,45920,79020,00020,00020,000
Toluene21,32120,70319,92019,92019,920
CLF21,14120,44919,45519,45519,455
EtAc21,55120,87620,16120,16120,161
DCM21,32120,53319,49319,49319,493
DCE21,18620,57619,84119,84119,841
Acetone21,45921,32119,72319,72319,723
Methanol21,32120,49119,56919,56919,569
ACN21,45920,87619,72319,72319,723
DMF21,18620,36619,45519,45519,455
DMSO21,00820,20219,19319,19319,193
Table 2. Empirical Kamlet–Taft (KAT) [29], Catalan [30,31] and Laurence [32] solvent parameters used in the multiple regression analysis.
Table 2. Empirical Kamlet–Taft (KAT) [29], Catalan [30,31] and Laurence [32] solvent parameters used in the multiple regression analysis.
MethodKATCatalanLaurence
SolventαβπSASBSPSdPDIESα1β1
1,4-Dioxane0.00.370.550.000 0.4440.737 0.3120.770.360.000.44
CCl40.00.00.280.000 0.0440.768 0.0000.820.100.000.00
Toluene0.00.110.540.0000.1280.7820.2840.860.200.000.15
CLF0.440.00.580.0470.0710.7830.6140.800.400.200.00
EtAc0.00.450.540.0000.5250.6740.5350.710.510.000.52
DCM0.300.00.820.040 0.1780.761 0.7690.780.600.100.00
DCE0.00.00.8070.0300.1260.7710.7420.800.740.000.00
Acetone0.080.480.710.0000.4750.6510.9070.690.780.040.49
Methanol0.930.620.600.6050.5450.6080.9040.640.841.000.54
ACN0.190.310.750.0440.2860.6450.9740.670.840.230.37
DMF0.00.690.870.0310.6130.7590.9770.780.870.000.69
DMSO0.00.761.00.0720.6470.8301.0000.841.000.000.71
Table 3. Coefficients obtained by multivariate linear regression analysis of the investigated azulene derivatives using Kamlet–Taft (KAT), Catalan and Laurence solvent parameters.
Table 3. Coefficients obtained by multivariate linear regression analysis of the investigated azulene derivatives using Kamlet–Taft (KAT), Catalan and Laurence solvent parameters.
Fitting MethodLigandy0 × 103Correlation CoefficientsR2
KAT aαbβcπ*
O121.70 ± 1.84−86.73 ± 1.7556.72 ± 2.17−583.31 ± 3.100.435
O220.60 ± 2.29−379.22 ± 2.18200.96 ± 2.70−1368.62 ± 3.860.743
O321.18 ± 2.52−228.32 ± 2.40−12.44 ± 2.98−790.21 ± 4.250.498
O518.86 ± 2.50376.51 ± 2.31156.34 ± 2.49212.37 ± 3.830.401
O618.09 ± 2.77110.16 ± 2.63429.86 ± 3.27−233.94 ± 4.670.252
Catalan aSAbSBcSPdSdP
O123.39 ± 2.55−502.30 ± 1.3285.15 ± 1.00−2535.06 ± 3.33−308.2 ± 68.260.943
O222.70 ± 2.7 −822.99 ± 1.42400.93 ± 1.07−3447.95 ± 3.58−846.20 ± 0.730.980
O323.48 ± 4.16 −883.34 ± 2.1644.75 ± 1.63−3435.86 ± 5.44−431.24 ± 1.110.928
O521.35 ± 5.05 −428.37 ± 2.62415.66 ± 1.98−2774.90 ± 6.60−484.54 ± 1.350.861
O622.14 ± 13.70−1377.65 ± 7.11141.29 ± 5.38−5461.76 ± 17.89−336.09 ± 3.660.743
Laurence aDIbEScα1dβ1
O123.75 ± 4.14−2774.5 ± 5.04−507.2 ± 1.25−305.15 ± 1.14103.52 ± 1.300.900
O323.03 ± 5.14−3535.5 ± 6.25−1322.9 ± 1.5−423.99 ± 1.42588.61 ± 1.610.955
O223.92 ± 6.83−3701.7 ± 830−754.3 ± 2.06−498.44 ± 1.89137.89 ± 2.140.871
O521.61 ± 7.58−2869.8 ± 9.21−746.93 ± 2.29−222.79 ± 2.09480.85 ± 2.370.792
O620.28 ± 8.71−2796.9 ± 10.58−279.95 ± 2.63−175.33 ± 2.40447.12 ± 2.730.742
Table 4. Percentage contributions of the Catalan and Laurence solvent parameters to the solvatochromism of the investigated azulenes.
Table 4. Percentage contributions of the Catalan and Laurence solvent parameters to the solvatochromism of the investigated azulenes.
Fitting MethodLigandRelative Contribution of Correlation ParametersSpec.*a (%)Non-Spec.*b (%)
Catalan SA (%)SB (%)SP (%)SdP (%)
O114.642.4873.898.9817.1282.87
O214.917.2662.4815.3322.1877.81
O318.420.9371.658.9919.3580.64
O510.4310.1267.6211.8020.5679.43
O616.5613.7265.674.0430.2869.71
Laurence DI (%)ES (%)α1 (%)β1 (%)
O175.1813.748.262.8011.0788.92
O260.2122.537.2210.0217.2482.75
O372.6914.819.782.7012.4987.50
O566.4217.285.1511.1216.2883.71
O675.607.564.7312.0816.8283.17
*a specific interaction; *b non-specific interaction.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Homocianu, M.; Airinei, A.; Matica, O.-T.; Cristea, M.; Ungureanu, E.-M. Solvent Effects and Metal Ion Recognition in Several Azulenyl-Vinyl-Oxazolones. Symmetry 2023, 15, 327. https://doi.org/10.3390/sym15020327

AMA Style

Homocianu M, Airinei A, Matica O-T, Cristea M, Ungureanu E-M. Solvent Effects and Metal Ion Recognition in Several Azulenyl-Vinyl-Oxazolones. Symmetry. 2023; 15(2):327. https://doi.org/10.3390/sym15020327

Chicago/Turabian Style

Homocianu, Mihaela, Anton Airinei, Ovidiu-Teodor Matica, Mihaela Cristea, and Eleonora-Mihaela Ungureanu. 2023. "Solvent Effects and Metal Ion Recognition in Several Azulenyl-Vinyl-Oxazolones" Symmetry 15, no. 2: 327. https://doi.org/10.3390/sym15020327

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