Next Article in Journal
Hetero-Disubstituted Sugarcane Bagasse as an Efficient Bioadsorbent for Cationic Dyes
Previous Article in Journal
Lingonberry Leaves Modify Rumen Protozoa Population, Carbohydrate Digestion, and Morphology of Gastrointestinal Tract in Sheep: A Preliminary Study
Previous Article in Special Issue
Catalyst-Driven Improvements in Conventional Methods for Imine-Linked Covalent Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvatochromic and Computational Study of Three Benzo-[f]-Quinolinium Methylids with Photoinduced Charge Transfer

by
Mihaela Iuliana Avadanei
1,*,
Ovidiu Gabriel Avadanei
2 and
Dana Ortansa Dorohoi
2,*
1
Petru Poni Institute of Macromolecular Chemistry, 41A Gr. Ghica Voda Alley, 700487 Iasi, Romania
2
Faculty of Physics, Alexandru Ioan Cuza University, 11 Carol I Blvd, 700506 Iasi, Romania
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(15), 3162; https://doi.org/10.3390/molecules30153162
Submission received: 18 June 2025 / Revised: 23 July 2025 / Accepted: 25 July 2025 / Published: 29 July 2025
(This article belongs to the Special Issue Feature Papers in Applied Chemistry: 4th Edition)

Abstract

The solvatochromic properties of 48 solvents of three benzo-[f]-quinolinium methylids (BfQs) were analyzed within the theories of the variational model and Abe’s model of the liquid. The electro-optical properties of BfQs in the first excited state were determined based on the charge transfer process that occurs from the ylid carbon to the nitrogen atom. The dipole moments and the polarizabilities in the first excited state were calculated according to the two models. The quantum chemical calculations helped in understanding the relationship between the molecular structure and absorption properties of the ground state. It is concluded that several key parameters modulate the strength of the charge transfer and they work in synergy, and the most important are as follows: (i) isomerism around the single polar bond, and (ii) the properties of the solvent. The link between geometrical conformation and the zwitterionic character make the studied BfQs very sensitive chromophores for sensors and optical switching devices.

Graphical Abstract

1. Introduction

The nitrogen-containing fused heterocycles are the most abundant building block in medicinal chemistry, being used in the development of anticancer [1,2] and antifungal drugs [3,4,5]. The benzo-[f]-quinolines (BfQs) and their [3+2] cycloadducts stand as a versatile class of precursors for the pharmacological industry, in organic chemistry [1,2,3,4,5,6], and in obtaining dyes and fluorescent probes [7,8]. BfQs have two interesting structural features: a large delocalized electronic distribution, due to the polyheterocyclic structure, and a permanent dipole, due to the separated electron acceptor and donor moieties. Like all ylids, the 1,2-dipolar structure is achieved through the positive charge located on the nitrogen atom and the negative charge on the carbanion [9,10]. In addition, the electron withdrawing character and the hyperpolarizability of the quaternary amine endow BfQs with strong non-linear optical properties [11,12], which are now desired in photonics and information technologies.
The unusual photophysical properties of the charge-separated compounds such as BfQs stem from the low conjugation between the electron donor and acceptor units. The zwitterionic character is enforced in ground state and provokes pronounced negative solvatochromism. The molecular polarization is strongly dependent on the electric and coordinating properties of the solvent. In the solvation shell that surrounds the solute molecule, the solvent creates an internal electric field that influences the electronic states of the analyzed molecule [13,14]. The influence of the solvent on the electronic absorption spectra can be analyzed based on theoretical models [15,16,17] or by using empirical relations established between the wavenumber in the maximum of the absorption band and some solvent parameters [18,19,20,21,22,23]. The empirical dependences have the drawbacks of their limited applicability to similar chemical structures only.
In this work, we have addressed the negative solvatochromism of three BfQ derivatives by using a quasi-empirical procedure. The chemical structure of the benzo-f-quinolinium methylides allowed us to differentiate the solvatochromic effects according to differences in the short substituent. The compounds are as follows: benzo-[f]-quinolinium acetyl benzoyl methylid (I1), benzo-[f]-quinolinium carbethoxy benzoyl methylid (I2), and benzo-[f]-quinolinium dibenzoyl methylid (I3) [24,25] with structural features from Scheme 1. Some of the optical properties of I1 were described in a previous study [25]. Because of their zwitterionic structure, I1I3 are able to distinguish between apolar and polar and/or protic solvents in terms of the hypsochromic or bathochromic behavior of their charge transfer absorption band.
In our study, the universal interactions are described by theoretically established functions of the refractive index and dielectric permittivity [26,27,28] and the specific interactions expressed by the empirical Kamlet-Abboud-Taft parameters [18,19,20,21,22,23]. Accordingly, the relation that shows the influence of the solvent on the electronic spectrum of the active molecule is as follows:
ν c a l c . = ν 0 + C 1 f ( ε ) + C 2 f ( n ) + C 3 β + C 4 α
where Cj, j = 14, are the correlation coefficients; and f ( ε ) = ( ε 1 ) / ( ε + 2 ) and f ( n ) = ( n 2 1 ) / ( n 2 + 2 ) are the functions of dielectric permittivity and of the refractive index [26]. The second and third terms in Equation (1) were theoretically established. The α and β parameters describe the ability of the solvents to participate in specific interactions with the solute molecule by donating or accepting protons, respectively. The correlation coefficients depend on the electro-optical parameters of the solute. Equation (1) gives information about the ground state of the studied molecule, which generally is non-fluorescent. Details about the descriptors of the excited state of a solvated and non-fluorescent molecule can then be obtained by using various theories, like those of MacRae [17] or Abe [15,16]. We will apply the variational model [29,30,31], which is based on MacRae’s theory [17], in order to determine the electro-optical parameters in the excited state of I1, I2 and I3 ylids, according to the observed solvatochromism. In the second stage, we will use Abe’s theory of liquid [16,17] for determining the dipole moment and the polarizabilities in the excited state. The structure–optical property correlation was finally investigated by means of quantum chemical calculations with the DFT and TD-DFT methods.

2. Results and Discussion

2.1. Experimental Solvatochromism of I1, I2 and I3

The absorption spectra of I1, I2 and I3 have been recorded in a series of solvents with different characteristics in order to test their responses. As shown in Figure 1 for representative spectra in alcohols, I1, I2 and I3 have several electronic absorption bands. Two intense ππ* transitions are observed in the UV region around 270 nm and 350 nm, and a shallow one above 235 nm. Then n → π* transition in the Vis region is of a low intensity and is generated from an intramolecular charge transfer (ICT) [9,10,25,30,31,32], which makes it heavily dependent on the properties of the solvent, either electric or coordinative. The spectra of I1, I2, and I3 in alcohol in comparison with those in acidified alcohol show the disappearance of the ICT band when protonation of the nitrogen atom blocks the electron transfer. The exemplary spectra in the visible region for low alcohols vs. solvents of low polarity in Figure 1b,d,f evidence the large blue shift in the ICT maximum. Its negative solvatochromism points to the stabilization of the molecule in apolar and aprotic solvents and a redistribution of the electronic cloud around the dipolar bond in polar solvents. The maximum of the ICT band for I1, I2 and I3 as a function of the solvent parameters for all solvents analyzed is listed in Table 1. As presented in Figure 2, the graphical representation of the ε value ν (cm−1) of the ICT maximum as a function of f(ε) for each BfQ splits the plot into two clear regions: the apolar/aprotic solvents, where the universal interactions dominate, and the polar/protic solvents, influenced by specific solvent–solute interactions.
In Figure 2, the experimental points corresponding to aprotic solvents are near the curve situated at smaller wavenumbers, and the representative points for protic solvents are situated near the curve situated at higher wavenumbers. The distance between the two curves estimates the energy of the quasi-chemical interactions in the protic solvents compared to the aprotic ones with the same macroscopic parameters. The separation of the points in Figure 2a–c is due to specific interactions of ylids with protic solvents shifting the ICT visible absorption band through the high wavelengths.
When both universal and specific interactions act in solutions, the Kamlet–Abboud–Taft solvent empirical parameters π * , α, and β are used to express the shift in the electronic absorption band by multilinear equations of the following type:
ν c a l c .   = ν 0 + P π * + A α + B   β  
where ν 0 is the value of the wavenumber in the maximum of the electronic band in a vacuum, P is the magnitude of the universal interactions, and A and B measure the strength of the specific interactions in which the solvent donates or receives protons from solute, respectively. The sign of the correlation coefficients shows the sense of the spectral shift.
Based on the data from Table 1, the results of the multiparametric regression are exhaustively listed in Table S1 and are briefly given through the following expressions:
ν(I1) = 21,079.48 + 922.02·π* + 1728.03·α + 622.60·β (R = 0.92)
ν(I2) = 20,359.59 + 887.83·π* + 2260.64·α + 1677.93·β (R = 0.92),
ν(I3) = 19,872.56 + 1220.10·π* + 2660.40·α + 1656.91·β (R = 0.95)
Equations (3)–(5) show that in diluted solutions of the studied BFQs, the universal interactions are dominant in non-polar and non-protic solvents, while in protic solvents the specific interactions with receiving protons by the ylids become very important (Table 2).
The excited state dipole of I1, I2 and I3 was estimated based on both the results obtained by solvatochromic analysis of the spectral shifts in 48 solvents and the electro-optical parameters determined by quantum chemical calculations. The dependence of the ν value (cm−1) on the solvent parameters is given by Equation (1). By using the multilinear regression on Equation (1) and data from Table 1, the following dependencies were obtained:
ν c a l c .   ( I 1 ) = 20,018 + 1145 f ( ε ) +   3978   f ( n ) + 226   β +   1888   α     ( R   =   0.919 )
ν c a l c .   ( I 2 ) = 19,982 + 1183   f ( ε ) + 1555   f ( n ) + 767   β + 2281   α     ( R   = 0.928 )
ν c a l c .   ( I 3 ) = 19,288 + 1524 f ( ε ) + 1983   f ( n ) + 907   β + 3032   α   ( R = 0.970 )
The variational method uses the coefficients C1 and C2, whose expressions contain the solvent descriptors, according to the following relations [33,34,35]:
C 1 r 3 = 2 µ g ( µ g µ e c o s φ ) + 3 k T ( α g α e )
( C 1 + C 2 )   r 3 = µ g 2 µ e 2 + 1.5   I u   I v I u + I v ( α g α e )
The C1 and C2 coefficients are expressed in erg; r is the molecular radius (cm); μg and μe are the dipole moments in the ground and in excited state (Debye); αg and αe are the molecular polarizabilities (cm3); and Iu and Iv are the ionization potentials of the solute (u) and solvent (v) (in erg) (1 eV = 8066 cm−1; k = 1.38041 × 10−16 erg). The electro-optical parameters of I1, I2 and I3 calculated in vacuum (RCAM-B3LYP6-31G(d,p)/def2SV level) are listed in Table 3.
The variational model is based on the hypothesis of McRae [17,29,30,31] that the molecular polarizability could be considered constant in the process of light absorption. By using Equations (9) and (10) for C1 and C2, a relationship between the dipole moment in the excited state, μe, and the angle between μe and μg (noted φ) can be determined. One obtains solutions of this relation as a function of the φ angle. By varying the angle and by calculating the polarizability for the obtained value of the dipole moment in the excited state, we can estimate that the electronic transition takes place when the polarizability in the excited state becomes equal to that in the ground state. By using the molecular descriptors from Table 3 and Equations (9) and (10), the variational model applied to I1, I2 and I3 resulted in the relations listed in Table 4 for the dipole moments in the excited state.
With the values obtained by the multilinear regression on Equation (1), the dipole moment and the polarizability of the spectrally active molecule (I1, I2 and I3) in the excited state can be calculated. These values are close to the ground state, so in the limits of the measurement precision and of conditions in which the C1 and C2 parameters are determined, one can suppose that McRae’s hypothesis is accomplished in condition φ  0 . Due to the chemical structure of the carbanion, we consider that the electronic transition upon light excitation occurs, in a first approximation, parallel to the ylid bond.
The variational model allows us to estimate numerically the values of μe, if the angle φ between the dipoles in the ground and the excited state is varied and the dipole moment in the excited state is calculated for each value of φ. The method that considers the electronic transition in the absorption process taking place without any change in the molecular polarizability of the spectrally active molecule can be applied here only with some approximations. This is because the free term in the expression of the molecular polarizability vs. the squared excited dipole moment is smaller than that obtained by DFT calculations for the ground state.
Therefore, in the paradigm of the variational model and by using the relations from Table 4, we have determined the values of the dipole moments and polarizability of I1, I2 and I3 in the excited state for a number of values of the φ angle. The calculated values are given in Table 5, Table 6 and Table 7.
As it results from relation (5), the correlation coefficient C contains a term dependent only on the solute dipole moments and a term dependent on the solute polarizabilities in the two electronic states responsible for visible transition in absorption process. If one defines ratio R by relation (11), its variation with the angle φ can offer information about the visible absorption process. Equations (9) and (10) can be rewritten as follows:
R = C 1 µ C 1 α = 2 µ g ( µ g µ e c o s φ ) 1.5   I u   I v I u +   I v ( α g α e )
The C1μ and C1α and the R ratio can be calculated for each individual value of the φ angle and the corresponding values are listed in Tables S2–S4. The graphical representations of R vs. φ for I1, I2, and I3 are presented in Figure 3.
The sign of R gives information about the electronic transitions. Figure 3 shows that R is a continuous function with a discontinuity near φ ≈ 80°. One can consider that the visible absorption can induce a charge redistribution determining such a high value of the angle between the dipole moments of solutes in the two electronic states participating in this process.

2.2. Abe’s Model of the Liquid

The results obtained above according to the variational model will be compared to those determined in the limits of the liquid theory developed by T. Abe [15,16]. In this model, the linear relationship between the values of the squared dipole moments in the ground and in the excited state is given by Equation (12):
µ e 2 ( u ) µ g 2 ( u ) + α e ( u ) A = B
A   ( erg ) = µ g 2 ( v ) + 3 2 α g ( v ) I g ( v ) [ I g ( u ) h c ν s ] I g ( v ) + I g ( u ) h c ν s 2 3 k T µ g 2 ( v ( p ) ) + α ( v )
B   ( erg · cm 3 ) = ν 0 ν s C + { µ g 2 ( v ) + 3 2 α g ( v ) I g ( v ) I g ( u ) I g ( v ) + I g ( u ) }   α g ( u ) 2 3 k T µ g 2 ( v ( p ) ) + α ( v )
C = 16 π 3 N A 2 9 ( ρ v M v ) 2 3 { [ ( M u ρ u ) 1 3 + ( M v ρ v ) 1 3 ] 4 + [ ( M u ρ u ) 1 3 + 3 ( M v ρ v ) 1 3 ] 4 + [ ( M u ρ u ) 1 3 + 5 ( M v ρ v ) 1 3 ] 4 + }
The A, B, and C parameters in relations (13) and (14) are functions of the molecular descriptors of the studied molecules, that is I1, I2 and I3, and of the solvent parameters (Table S5). The other notations are designating the following: M—molar mass (g/mol); ρ—density (g/cm3); T—absolute temperature (K); N A —Avogadro’s number; and k—gaseous constant. The values of A, B, and C for I1, I2 and I3 in the 48 solvents used in this work are listed in Tables S6 and S7. The graphical representation of B vs. A allows for the determination of the polarizability of the excited state from the slope, while the intercept represents the difference µ e 2 ( u ) µ g 2 ( u ) . Accordingly, Figure 4 shows the B vs. A plot for I1, I2, and I3.
In Figure 4, two different curves are seen for each of the analyzed BfQs: one with a positive slope that follows the regular trend for most of ICT compounds, and the other with a steep negative slope, which is highly uncommon. The linear increase in B occurs in apolar and non-H-bonding solvents, from which the dipole moment in the excited state (Equation (12)) is determined as being lower than in the ground state (Table 8). In polar and coordinating solvents, the calculated values of the dipole moment in the excited state are higher than in the ground state.
The calculated values for the excited state dipole moments and polarizabilities of the studied BfQ molecules were obtained based on the computed corresponding parameters for the ground state of the isolated molecules in a vacuum. For this reason, the obtained values could be affected by solvent influence on the molecular parameters of the ground state. The specific interactions between ylids and the protic solvents, for example, increase the ground state dipole moments of ylids. The data obtained based on the Abe model are affected by the hypotheses in which this model is developed. So, one considers that the ground and excited dipole moments are collinear and also that the specific solute–solvent interactions are neglected. Nevertheless, the decrease in the dipole moment by excitation in the absorption process, common to the studied ylids, is evidenced both in the data computed based on the variational model and also in the Abe model.

2.3. Quantum Chemical Investigations of I1, I2 and I3

The structural features of the studied benzo-[f]-quinoline methylids were investigated at the RCAM-B3LYP6-31G(d,p)/def2SV level of theory. The geometry in the relaxed ground state in vacuo for I1, I2, and I3 is shown in Figure 5. Irrespective of the substituent on the ylidic carbon, the molecules are generally non-planar, with the two arms twisted with respect to each other and with the plane of the fused heterocycles. The twisting degree is variable and the ground energy does not differ too much from one rotamer to another. The charge distribution on the donor moiety is slightly different according to the kind of short substituent on the ylidic carbon. From carbethoxy to benzoate as substitution groups, the Mulliken charge on the carbanion decreased from I1 (−0.824 e) to I2 (−0.734 e) and increased again in I3 (−0.788 e). The magnitude of the dipole moment follows the same trend, but overall, the dipole moment in the ground state is around 4.5 D on average. The electrostatic potential (Figure 5b) does not differ from one BfQ to another, having the nucleophilic activity concentrated on the carbanion oxygen atoms. In the ground state, the charge is accumulated on the carbanion, as HOMO orbitals show (Figure 5d), and becomes displaced on the fused heterocycles in LUMO. The intramolecular charge transfer in I1, I2 and I3 occurs from HOMO to LUMO, which is in agreement with previous studies on cycloimmonium ylids [29,30,31,32,33,34].
Insights into the electronic absorption properties of I1, I2 and I3 were further obtained by TD-DFT methods with the same functional and basis set. The calculations were made for a vacuum and two representative solvents, cyclohexane (CHX) and water. The TD-DFT analysis revealed that the spectra profile and especially the intensity of the ICT band are completely dependent on the conformation of the molecule. The free rotation around the single bond between the heterocycle and ylid carbon, and the many degrees of freedom at the two substituents, lead to a bunch of conformations that affect the energetics of the molecule to a certain degree. The most mobile group is the short substituent, and its orientation relative to the heterocycle is the decisive factor in the absorption process. The orientation of the benzenoid substituent is always orthogonal on the heterocycle, due to steric hindrance, so it does not greatly affect the intensity of the ICT process, but rather the vibronic structure of the intense ππ* band around 250 nm.
The influence of the molecular orientations is strong for I1 and I2. Table 9 lists the values of the dipole moments and Mulliken charges in the ground state for the planar and twisted conformations for I1I3. From linear to twisted, the dipole moment in the ground state increases and the electric charge on the N atom decreases. Also, from CHX to water, the dipole moment becomes higher and the Mulliken charge on the nitrogen atom becomes smaller. In very close connection with these parameters, the profiles of the calculated UV-Vis spectra differ more in terms of their molecular conformation and less in regard to the type of the solvent. Figure 6a,c illustrates the UV-Vis spectra of I1, I2 and I3 for the planar and the orthogonal orientation of the short substituent at the ylide carbon atom relative to the plane of the polyheterocycle, and the oscillator strength of the electronic transitions, f. Overall, the calculated profile is similar to the experimental spectra, although the band around 344–366 nm of all BFQs was not captured by DFT calculations, whatever the method used. The ICT wavelength maxima were slightly overestimated in water in comparison with the experimental values, and were underestimated in cyclohexane. Yet, the ICT is at its maximum when the short substituent at the ylid carbon is oriented almost in plane with the fused heterocycle. When the substituent is twisted out of the plane of the molecule, the ICT band is weak and the maximum is bathochromically shifted towards 500 nm and beyond for vacuum and non-polar solvents, and hypsochromically shifted in protic solvents. For I3, the two substituents pose a strong hindrance with respect to the heterocycle and they are generally parallel to each other and perpendicular to the plane of the heterocycle. Therefore, there are no great differences in molecular conformations nor in the absorption properties between them. Figure 6c presents the representative UV-VIS spectra of I3 as an average for the electronic absorption. One can state that the orientation of substituents can be somewhat predicted by analyzing the profile of the UV-Vis spectrum, the intensity, and the position of the ICT band. We conclude that the experimental spectra, where the ICT band is of low intensity, consists of overlapping contributions from molecules with twisted conformations at the polar bond and with many rotamers within the short substituent.
The lowest excited state, S1, is basically described by a HOMO–LUMO transition of low oscillator strength in non-polar cyclohexane, and by extension, in a vacuum, and of high oscillator strength in polar solvents. In water, the calculated electronic spectra are higher in intensity but with a weaker ICT band, which suggests a redistribution of the electronic cloud by interaction with the water dipoles. In conjunction with the hole–electron distribution analysis presented in Table 10 and Table 11, the HOMO–LUMO transition arises from a charge transfer from the carbanion to the fused heterocycle, a transition that occurs in whatever medium is analyzed. One can see that the carbonyl groups also do participate in transition, as the hole has a nodal plane along the oxygen atoms. The electron distribution is localized mainly near the nitrogen, where most of the charge is transferred after excitation. We observe that the transferred charge is distributed on half of the heterocycle. Accordingly, the charge difference upon electronic transition follows the profile of the hole–electron localization, where the electronic density of the ylid carbon decreases and further increases in the heterocycle.
The comparison of the calculated energy of the frontier molecular orbitals between cyclohexane and water, shown in Figure 6d, confirms the negative solvatochromism by lowering the HOMO energy for any of BfQs and enlarging the HOMO–LUMO gap. There is an expected difference in the energy of HOMO and LUMO between the planar and the twisted conformations. The HOMO and LUMO levels are lower for non-planar molecules and with a higher energy gap than those of planar ones.
The TD-DFT analysis of the S1/ICT state, presented in Table 10 for I1 as a representative example, shows that the electronic charge is reversed as compared to the ground S0 state, as the nitrogen atom carries a negative charge after electronic excitation. The same situation appears for I2 and I3 and is not presented here. The charge transfer occurs irrespective of the conformation, but is strongly favored when the carbonyl group in the substituent arm aligns with the aromatic structure in the heterocycle. It leads to a higher negative charge on nitrogen in S1/ICT than in the twisted molecule, according to the values of Q [-N-] in Table 10. But the values of the dipole moments listed in Table 10 for I1 indicate that the excited planar molecule is less polar than the excited twisted molecule. The dipole moment in S1/ICT is twice the magnitude in water as compared to cyclohexane.
Table 11 contains the analysis of the electronic excitation for I2 and I3 in planar conformation in order to focus on the intense ICT process. As was explained above, the charge transfer for planar and twisted I2 is similar to that in I1. For both I2 and I3, there is a net separation in the centers for electronic charge between donor and acceptor units. As in I1, the transferred electronic charge is restricted to half of the heterocycle, which is near the nitrogen.
In contrast with other cycloimmonnium ylids, whose hole and electron distribution have changed according to the type of solvent [33,34], the hole of the BfQ remains mainly located on the carbanion and the electrons transfer to nitrogen and the nearest aromatics. The ICT character of the S0 → S1 transition is confirmed for any solvent and does not seem to be affected, even in strongly coordinating solvents like water. Summing up the above observations, the negative solvatochromism of zwitterionic benzo-[f]-quinolinium methylids is given by three key factors: (i) alkyl vs. aryl units as terminal groups in the short substituent, because the less voluminous and virtually linear alkyl chain will allow the free rotation around the polar bond without the steric hindrance imposed by bulk aromatic rings; (ii) the conformation around the polar bond, which is mandatory for the ylids with the acetyl and the carbethoxy substituents, and the electron transfer is maximal where the carbonyl group in these substituents is in the vicinity of the polyheterocycle; and iii) the polarity and coordinating properties of the solvent.

3. Materials and Methods

The synthesis, structural characterization, and analysis of the electro-optical properties of I1, I2, and I3 followed the established procedure [24]. Elemental analysis was performed on an Elemental Exeter Analytical CE 440 apparatus (Exeter Analytical Inc., Coventry, UK). The 1H-NMR and 13C-NMR measurements were made on a high-resolution liquid NMR 400 MHz spectrometer Bruker Neo-1 (Bruker BioSpin GmbH, Rheinstetten, Germany) for a direct detection probe, with 5 mm Quadra nuclei probes, or QNPs (four nuclei, 1H/13C/19F/29Si). The FTIR spectra were recorded as KBr pellets with a Bruker Vertex 70 spectrometer (Bruker Optics GmbH, Ettlingen, Germany) between 4000 and 400 cm−1 at a 2 cm−1 resolution.
Benzo-[f]-quinolinium-acetyl-benzoyl methylid (I1). Orange powder. Anal. calcd for: C23H17NO2: 81.42% C; 5.01% H; 4.13% N. Found: 81.83% C; 5.39% H; 4.61% N. M. p. 136 °C. 1H-NMR (DMSO-d6, 400 MHz), δ (ppm): 2.89 (s, 3H, CH3); 7.22–7.24 (d, 2H, CHAr); 7.39–7.40 (d, 1H, CHAr); 7.45–7.47 (d, 2H, CHAr); 7.57–7.59 (d, 2H, CHAr); 7.65–7.67 (d, 1H, CHAr); 7.88–7.91 (d, 2H, CHAr); 8.12–8.14 (d, 1H, CHAr); 8.37–8.40 (d, 1H, CHAr); 8.83–8.85 (d, 1H, CHAr); 8.99–9.02 (d, 1H, CHAr). 13C-NMR (DMSO-d6, 400 MHz), δ (ppm): 25.86 (CH3); 125.3; 126.7; 127.2; 127.5; 127.9; 128.2; 128.5; 128.9; 129.1; 130.1; 130.6; 130.9; 131.1; 134.4; 134.7; 134.9; 135.4; 140.8; 142.6 (Ar); 139.1 (C-); 181.6; 193.4 (C=O). FTIR (KBr) νmax, cm−1: 3015w (νCHarom), 2919w (νCH), 2838w (νCH), 1720s (νC=O), 1703m (νC=O), 1620m (νC=C), 1595s (νC=C), 1545s (νC=C), 1505s (νC=C), 1460m (δCH), 1404m, 1396m (δCH3), 1232m, 862m, 803m (δCCarom), 730m (δCC/(δCH).
Benzo-[f]-quinolinium-carbethoxy-benzoyl methylid (I2). Orange powder. Anal. calcd. for: C24H19NO3: 78.05% C; 5.15% H; 3.79% N. Found: 78.44% C; 5.50% H; 4.26% N. M. p. 142–143 °C. 1H-NMR (DMSO-d6, 400 MHz), δ (ppm): 1.35–1.37 (m, 3H, CH3); 4.52–4.54 (m, 2H, CH2); 7.19–7.21 (d, 2H, CHAr); 7.31–7.33 (d, 1H, CHAr); 7.43–7.46 (d, 2H, CHAr); 7.67–7.69 (d, 2H, CHAr); 7.75–7.77 (d, 1H, CHAr); 7.84–7.85 (d, 2H, CHAr); 8.17–8.19 (d, 1H, CHAr); 8.42–8.44 (d, 1H, CHAr); 8.86–8.89 (d, 1H, CHAr); 9.00–9.03 (d, 1H, CHAr). 13C-NMR (DMSO-d6, 400 MHz), δ (ppm): 26.11 (CH3); 61.55 (CH2); 125.8; 127.1; 127.5; 127.7; 128.1; 128.6; 128.9; 129.2; 129.4; 130.3; 130.9; 131.6; 131.9; 134.6; 135.2; 135.7; 135.8; 141.3; 142.9 (Ar); 184.3; 195.1 (C=O); 139.9 (C). FTIR (KBr) νmax, cm−1: 3025w (νCHarom), 2998w (νCHarom), 2931w (νCH), 2858w (νCH), 1725s (νC=O), 1705m (νC=O), 1615m (νC=C), 1603m (νC=C), 1598s (νC=C), 1560s (νC=C), 1505s (νC=C), 1453m (δCH), 1420s (νC=C), 1383m, 1378s (δCH3), 1327m, 797m (δCC), 737m (δCC/(δCH).
Benzo-[f]-quinolinium-dibenzoyl methylid (I3). Orange crystals (recrystallization from methanol). Anal. calcd for: C28H19NO2: 83.79% C; 4.74% H; 3.49% N. Found: 84.15% C; 5.11% H; 3.86% N. M. p. 230–231 °C. 1H-NMR (DMSO-d6, 400 MHz), δ (ppm): 7.25–7.27 (d, 2H, CHAr); 7.34–7.37 (d, 1H, CHAr); 7.42–7.44 (d, 4H, CHAr); 7.51–7.53 (d, 4H, CHAr); 7.60–7.62 (d, 2H, CHAr); 7.83–7.85 (d, 2H, CHAr); 8.22–8.24 (d, 1H, CHAr); 8.46–8.48 (d, 1H, CHAr); 8.79–8.81 (d, 1H, CHAr); 9.11–9.13 (d, 1H, CHAr). 13C-NMR (DMSO-d6, 400 MHz), δ (ppm): 126.6; 126.9; 127.4; 127.6; 127.8; 128.3; 128.5; 128.7; 128.8; 128.9; 129.9; 130.3; 130.5; 130.8; 131.3; 131.6; 131.7; 131.9; 134.6; 134.8; 135.6; 135.9; 136.2; 142.8; 143.3 (Ar); 194.7; 194.9 (C=O); 138.8 (C). FTIR (KBr) νmax, cm−1: 3049w (νCHarom), 3021w (νCHarom), 2932w (νCH), 2854w (νCH), 1748s (νC=O), 1735m (νC=O), 1627m (νC=C), 1603m (νC=C), 1598s (νC=C), 1556s (νC=C), 1498s (νC=C), 1462m (δCH), 1327m (δC=C), 874m (δCHarom), 797m (δCCarom), 735m (δCC/(δCH), 650m.
The UV-Vis absorption spectra were recorded with an Analytic Jena Specord Plus-5 spectrophotometer (Jena GmbH, Germany), in suprasil quartz cuvettes of a 10 mm pathlength. The solvents used were of spectral grade and from Merck (Merck KGaA, Darmstadt, Germany), Sigma Aldrich (Saint-Louis, MO, USA) and Chemopar (Iasi, Romania). The solvent parameters, which are microscopic (α and β) and macroscopic (ε and n), were obtained from previous studies [10,25,29,30,31,32,33,34] and from the literature [17,18,19,20,21,22,23].
The electronic properties of the studied BfQs were investigated in the DFT and TD-DFT methods with the Gaussian 09W version A.01 software [35]. The relaxed geometries were calculated in vacuo and in the presence of an implicit solvent method without symmetry constraints, firstly with the PM3MM method, and refined with the long-range corrected functional RCAM-B3LYP with the 6-31G(d,p) basis set and def2SVfitting model. The effects of electron correlation and long-range correction on the optical properties were taken into account [36]. The calculations were made for cyclohexane and water as the solvents of choice. The minimum on the potential energy surface in every case was verified by calculating the harmonical vibrational frequencies. The electronic excitation was analyzed by using the Multifwn 3.8 software [37].

4. Conclusions

The values of the dipole moments of the studied BfQs in the S1/ICT state, estimated by two methods, the variational model and the Abe model, are smaller in comparison to those calculated for the ground state, proving the decrease in the solvation energy in the ICT process due to light absorption. The differences between the obtained values of the excited state dipole moments in the two models are due to the hypotheses in which these methods were developed. The studied molecule BfQs interact with the protic solvents by means of hydrogen bonds, which increase the moment dipole in the ground state and change the energetics of the excited state. The small twist angle favored the ICT transition. The structure of the short substituent influenced the energy levels of the molecules, as the carbethoxy substituent raised the HOMO level in comparison to the shorter acetyl chain and compressed the energy gap. The HOMO–LUMO gap progressively increased from non-polar to polar solvents. The clear negative solvatochromism and the electro-optical properties lead to interesting applications of I1, I2 AND I3 as a sensor for harmful reagents and gases. Further studies of non-linear optical characteristics may recommend BfQs as useful in optical switching technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30153162/s1, Tables S1–S7: Multiparametric regression in the Kamlet–Taft approach. Calculated values of C, C, and the R ratio. The Abe parameters (A, B, and C). Molecular descriptors of the solvents.

Author Contributions

Conceptualization, D.O.D. and M.I.A.; methodology, D.O.D.; software, O.G.A. and M.I.A.; validation, D.O.D. and M.I.A.; formal analysis, D.O.D., M.I.A. and O.G.A.; investigation, D.O.D.; resources, D.O.D.; data curation, D.O.D.; writing—original draft preparation, D.O.D. and M.I.A.; writing—review and editing, D.O.D. and M.I.A.; visualization, D.O.D. and M.I.A.; supervision, D.O.D. 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 are available by request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ledade, P.V.; Lambat, T.L.; Gunjate, J.K.; Chopra, P.K.P.G.; Bhute, A.V.; Lanjewar, M.R.; Kadu, P.M.; Dongre, U.J.; Mahmood, S.H. Nitrogen-containing fused heterocycles: Organic synthesis and applications as potential anticancer agents. Curr. Org. Chem. 2023, 27, 206–222. [Google Scholar] [CrossRef]
  2. Zbancioc, G.; Mangalagiu, I.I.; Moldoveanu, C. The effective synthesis of new benzoquinoline derivatives as small molecules with anticancer activity. Pharmaceuticals 2024, 17, 52. [Google Scholar] [CrossRef] [PubMed]
  3. Antoci, V.; Oniciuc, L.; Amariucai–Mantu, D.; Moldoveanu, C.; Mangalagiu, V.; Amarandei, A.M.; Lungu, C.N.; Dunca, S.; Mangalagiu, I.I.; Zbancioc, G. Benzoquinoline derivatives: A straightforward and efficient route to antibacterial and antifungal agents. Pharmaceuticals 2021, 14, 335. [Google Scholar] [CrossRef] [PubMed]
  4. Boateng, C.A.; Zhu, X.Y.; Jacob, M.R.; Khan, S.I.; Walker, L.A.; Ablordeppey, S.Y. Optimization of 3 (phenylthio)quinolinium compounds against opportunistic fungal pathogens. Eur. J. Med. Chem. 2011, 46, 1789–1797. [Google Scholar] [CrossRef]
  5. Boateng, C.A.; Eyunni, S.V.K.; Zhu, X.Y.; Etukala, J.R.; Bricker, B.A.; Ashfaq, M.K.; Jacob, M.R.; Khan, S.I.; Walker, L.A.; Ablordeppey, S.Y. Benzothieno [3,2-b]Quinolinium and 3(phenylthio)quinolinium compounds: Synthesis and evaluation against opportunistic fungal pathogens. Bioorg. Med. Chem. 2011, 19, 458–470. [Google Scholar] [CrossRef]
  6. Yamamoto, K.; Torigoe, K.; Kuriyama, M.; Demizu, Y.; Onomura, O. (3+2) Cycloaddition of Heteroaromatic N-Ylides with Sulfenes. Org. Lett. 2024, 26, 798–803. [Google Scholar] [CrossRef]
  7. Liu, F.-T.; Zhai, S.-M.; Gao, D.-F.; Yang, S.-H.; Zhao, B.-X.; Lin, Z.-M. A Highly Sensitive Ratiometric Fluorescent Probe for Detecting HSO3−/SO32− and Viscosity Change Based on FRET/TICT Mechanism. Anal. Chim. Acta 2024, 1305, 342588. [Google Scholar] [CrossRef]
  8. Wang, M.-Q.; Zhang, Y.; Zeng, X.-Y.; Yang, H.; Yang, C.; Fu, R.-Y.; Li, H.-J. A Benzo(f)Quinolinium fused chromophore-based fluorescent probe for selective detection of c-Myc G-quadruplex DNA with a red emission and a large Stockes shift. Dye. Pigment. 2019, 168, 334–340. [Google Scholar] [CrossRef]
  9. Zugravescu, I.; Petrovanu, M. N-Ylid Chemistry; McGraw Hill: New York, NY, USA; Academic Press: London, UK, 1976. [Google Scholar]
  10. Dorohoi, D.; Partenie, H. The Spectroscopy of the cycloimmonium ylides. J. Mol. Struct. 1993, 293, 129–132. [Google Scholar] [CrossRef]
  11. Teran, N.B.; He, G.S.; Baev, A.; Shi, Y.; Swihart, M.T.; Prasad, P.N.; Marks, T.J.; Reynolds, J.R. Twisted Thiophene-Based Chromophores with Enhanced Intramolecular Charge Transfer for Cooperative Amplification of Third-Order Optical Nonlinearity. J. Am. Chem. Soc. 2016, 138, 6975–6984. [Google Scholar] [CrossRef]
  12. Zelichenok, A.; Burtman, V.; Zenou, N.; Yitzchaik, S.; Di Bella, S.; Meshulam, G.; Kotler, Z. Quinolinium-Derived Acentric Crystals for Second-Order NLO Applications with Transparency in the Blue. J. Phys. Chem. B 1999, 103, 8702–8705. [Google Scholar] [CrossRef]
  13. Onsager, L. Electric Moments of Molecules in Liquids. J. Am. Chem. Soc. 1936, 58, 1486–1493. [Google Scholar] [CrossRef]
  14. Wang, X.; Dao, R.; Yao, J.; Peng, D.; Li, H. Modification of the Onsager Reaction Field and Its Application on Spectral Parameters. ChemPhysChem 2017, 18, 763–771. [Google Scholar] [CrossRef]
  15. Abe, T. Theory of solvent effects on molecular electronic spectra. Frequency shifts. Bull. Chem. Soc. Jpn. 1965, 38, 1314–1318. [Google Scholar] [CrossRef]
  16. Abe, T. The Dipole Moments and Polarizabilities in the Excited States of Four Benzene Derivatives from Spectral Solvent Shifts. Bull. Chem. Soc. Jpn. 1967, 40, 1571–1574. [Google Scholar] [CrossRef]
  17. McRae, E.G. Theory of Solvent Effects on Molecular Electronic Spectra. Frequency Shifts. J. Phys. Chem. 1957, 61, 562–572. [Google Scholar] [CrossRef]
  18. Kawski, A. On the estimation of excited-state dipole moments from solvatochromic shifts of absorption and fluorescence spectra. Z. Naturforsch. 2002, 57, 255–262. [Google Scholar] [CrossRef]
  19. Kamlet, M.J.; Abboud, J.I.; Taft, R.W. The solvatochromic comparison method 6. The π* scale of solvent polarities. J. Am. Chem. Soc. 1977, 99, 6027–6038. [Google Scholar] [CrossRef]
  20. Kamlet, M.J.; Abboud, J.L.M.; Abraham, M.H.; Taft, R.W. Linear solvation energy relationships. 23. A comprehensive collection of the solvatochromic parameters π*, α, and β, and some methods for simplifying the generalized solvatochromic equation. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  21. Kamlet, M.J.; Abboud, J.I.; Taft, R.W. An examination of linear solvation energy relationships. Prog. Phys. Org. Chem. 1981, 13, 485–630. [Google Scholar] [CrossRef]
  22. Reichardt, C. Solvents and Solvent effects in Organic Chemistry, 3rd ed.; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  23. Catalán, J.; Hopf, H. Empirical Treatment of the Inductive and Dispersive Components of Solute−Solvent Interactions: The Solvent Polarizability (SP) Scale. Eur. J. Org. Chem. 2004, 2004, 4694–4702. [Google Scholar] [CrossRef]
  24. Van Do, N.; Rucinschi, E.; Druta, I.; Zugravescu, I. Recherches sur les benzo(f)quinoléinium-ylures. I. Synthèse de benzo(f)quinoléinium-ylures. Bull. Inst. Pol. Iasi 1977, 23, 53–58. [Google Scholar]
  25. Dorohoi, D.O.; Partenie, D.H.; Calugaru, A.C. Specific and Universal Interactions in Benzo-[f]-Quinolinium Acetyl-Benzoyl Methylid (BQABM) Solutions; Excited State Dipole Moment of BQABM. Spectrochim. Acta A 2019, 213, 184–191. [Google Scholar] [CrossRef] [PubMed]
  26. Hilliard, I.J.; Foulk, D.S.; Gold, H.S.; Rechisteiner, C.F. Effects of solute-solvent interactions on electronic spectra-a predictive analysis. Anal. Chim. Acta 1981, 133, 319–327. [Google Scholar] [CrossRef]
  27. Gulseven Sidir, Y.; Sidir, I.; Beker, H.; Tasal, E. UV spectral changes for some azo-compounds in the presence of different solvents. J. Mol. Liquids 2011, 162, 148–154. [Google Scholar] [CrossRef]
  28. Sidir, I.; Gulseven Sidir, Y.; Demiray, F.; Berker, H. Estimation of ground and excited states dipole moments of α-hydroxy-phenyl hydrazone derivatives. J. Mol. Liquids 2014, 197, 386–394. [Google Scholar] [CrossRef]
  29. Morosanu, A.C.; Gritco-Todirascu, A.; Creanga, D.E.; Dorohoi, D.O. Computational and solvatochromic study of pyridinium-acetyl-benzoyl-methylid (PABM). Spectrochim. Acta A 2018, 189, 307–315. [Google Scholar] [CrossRef]
  30. Dorohoi, D.O.; Dimitriu, D.G.; Dimitriu, M.; Closca, V. Specific interactions in N-ylid solutions, studied by nuclear magnetic resonance and electronic absorption spectropscopy. J. Mol. Struct. 2013, 1044, 79–86. [Google Scholar] [CrossRef]
  31. Dorohoi, D.O.; Gosav, S.; Morosanu, A.C.; Dimitriu, D.G.; Apreotesei, G.; Gosav, S. Molecular descriptors-spectral property relations for characterization molecular interactions in binary and ternary solutions. Symmetry 2023, 15, 2075. [Google Scholar] [CrossRef]
  32. Dorohoi, D.O.; Partenie, D.H.; Chiran, L.M.; Anton, C. About the electronic absorption spectra (EAS) and electronic diffusive spectra (EDS) of some pyridazinium ylids. J. Chim. Phys. Phys.-Chim. Biol. 1994, 91, 419–431. [Google Scholar] [CrossRef]
  33. Avădănei, M.I.; Griţco-Todiraşcu, A.; Dorohoi, D.O. Negative solvatochromism of the intramolecular charge transfer band in two structurally related pyridazinium Ylids. Symmetry 2024, 16, 1531. [Google Scholar] [CrossRef]
  34. Avadanei, M.I.; Dorohoi, D.O. Comparative Study of Two Spectral Methods for Estimating the Excited State Dipole Moment of Non-Fluorescent Molecules. Molecules 2024, 29, 3358. [Google Scholar] [CrossRef]
  35. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision, A.02; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  36. Pant, D.; Sitha, S. Roles of bridges on electronic, linear and nonlinear optical properties: A computational study on zwitterions with N-methyl pyridinium and p-dicyanomethanide phenylene. Comput. Theor. Chem. 2023, 1229, 114308. [Google Scholar] [CrossRef]
  37. Liu, Z.; Lu, T.; Chen, Q. An sp-hybridized all-carboatomic ring, cyclo [18] carbon: Electronic structure, electronic spectrum, and optical nonlinearity. Carbon 2020, 165, 461–467. [Google Scholar] [CrossRef]
Scheme 1. Chemical structure of the studied benzo-[f]-quinoline methylids.
Scheme 1. Chemical structure of the studied benzo-[f]-quinoline methylids.
Molecules 30 03162 sch001
Figure 1. Exemplary electronic absorption spectra are as follows: (a) the UV-VIS region of I1 in ethanol and acidified ethanol; (b) the ICT band of I1 as a function of solvents of different polarity; (c) the UV-VIS region of I2 in ethanol and acidified ethanol; (d) the ICT band of I2 as a function of solvents of different polarity; (e) the UV-VIS region of I3 in methanol and acidified methanol; (f) the ICT band of I3 as a function of solvents of different polarity. The blue arrow indicates the main ICT band.
Figure 1. Exemplary electronic absorption spectra are as follows: (a) the UV-VIS region of I1 in ethanol and acidified ethanol; (b) the ICT band of I1 as a function of solvents of different polarity; (c) the UV-VIS region of I2 in ethanol and acidified ethanol; (d) the ICT band of I2 as a function of solvents of different polarity; (e) the UV-VIS region of I3 in methanol and acidified methanol; (f) the ICT band of I3 as a function of solvents of different polarity. The blue arrow indicates the main ICT band.
Molecules 30 03162 g001aMolecules 30 03162 g001b
Figure 2. Plots of the maximum of the ICT band as a function of the dielectric permittivity of the solvent for (a) I1; (b) I2; (c) I3. The dotted lines are guides for the eye. The numbers on the experimental data points are the solvents’ numbering in Table 1.
Figure 2. Plots of the maximum of the ICT band as a function of the dielectric permittivity of the solvent for (a) I1; (b) I2; (c) I3. The dotted lines are guides for the eye. The numbers on the experimental data points are the solvents’ numbering in Table 1.
Molecules 30 03162 g002
Figure 3. Graphical representation of the R values, calculated according to Equation (11), as a function of the angle between the dipole moments in the ground and in the excited state for (a) I1; (b) I2; (c) I3. The dotted lines are guides for the eye.
Figure 3. Graphical representation of the R values, calculated according to Equation (11), as a function of the angle between the dipole moments in the ground and in the excited state for (a) I1; (b) I2; (c) I3. The dotted lines are guides for the eye.
Molecules 30 03162 g003
Figure 4. Graphical representation of the Abe parameters B vs. A for (a) I1; (b) I2, (c) I3. The experimental data are numbered according to the solvents’ numbering in Table 1. The dotted lines are the linear fit.
Figure 4. Graphical representation of the Abe parameters B vs. A for (a) I1; (b) I2, (c) I3. The experimental data are numbered according to the solvents’ numbering in Table 1. The dotted lines are the linear fit.
Molecules 30 03162 g004
Figure 5. (a) Optimized molecular structures of I1, I2, and I3 in ground state in vacuum. (b) Electrostatic potential surfaces of I1, I2, and I3; Q represents Mulliken charges for acceptor [-N] and donor [-C-] sites. (c) Values of dipole moments in ground state. (d) Frontier molecular orbitals HOMO and LUMO. Calculations: RCAM-B3LYP6-31G(d,p)/def2SV. Color code: positive lobes - orange; negative lobes - blue (HOMO) and red (LUMO).
Figure 5. (a) Optimized molecular structures of I1, I2, and I3 in ground state in vacuum. (b) Electrostatic potential surfaces of I1, I2, and I3; Q represents Mulliken charges for acceptor [-N] and donor [-C-] sites. (c) Values of dipole moments in ground state. (d) Frontier molecular orbitals HOMO and LUMO. Calculations: RCAM-B3LYP6-31G(d,p)/def2SV. Color code: positive lobes - orange; negative lobes - blue (HOMO) and red (LUMO).
Molecules 30 03162 g005
Figure 6. TD-DFT analysis of the electronic transitions in a vacuum as compared to the implicit solvents cyclohexane and water, and the oscillator strength (f) for (a) I1; (b) I2; (c) I3. (d) Energy levels of the frontier molecular orbitals of I1, I2 and I3 in the three considered media.
Figure 6. TD-DFT analysis of the electronic transitions in a vacuum as compared to the implicit solvents cyclohexane and water, and the oscillator strength (f) for (a) I1; (b) I2; (c) I3. (d) Energy levels of the frontier molecular orbitals of I1, I2 and I3 in the three considered media.
Molecules 30 03162 g006
Table 1. Solvent parameters and the position of the ICT band of I1, I2 and I3.
Table 1. Solvent parameters and the position of the ICT band of I1, I2 and I3.
No.εf(ε)f(n)π*βαν (nm)
I1
ν (nm)
I2
ν (nm)
I3
1n-Heptane 0.2270.24−0.0800476477500
2n-Hexane0.2410.229−0.0400476477496
3Dioxane0.2860.2510.550.370473475486
4Carbon tetrachloride 0.2920.2740.280.10472476497
5Benzene0.2990.2950.590.10454485488
6p-Xylene0.2990.291 0.130452487490
71,3,5-Trimethylbenzene0.3020.2930.410.130453488487
8o-Xylene0.3020.2920.410.110459491495
9Toluene0.3480.2970.540.110459494492
10Propanoic acid0.4340.239 0.451.12411414406
11Trichloroethylene0.4480.2820.480.050463481485
12Methoxybenzene0.5240.3020.730.320456472478
131,2-Dibromoethane0.5380.313 00458476478
14Chloroform0.5520.2670.690.10.2465476465
15Butyl acetate0.5770.240.460.450462474463
163-Methylbutyl acetate0.5890.2410.710.070465468466
17Chlorobenzene0.6050.3070.710.070453467482
18Ethyl acetate0.6250.2280.550.450454467479
19Methyl acetate0.6550.2210.60.420459465469
20Dichloromethane0.7270.2560.820.10.2459458474
21Octan-1-ol0.7450.2580.40.810.77438426435
221,2-Dichloroethane0.7520.2640.810.10459465472
23Pyridine0.790.2990.870.640455457468
24Hexan-1-ol0.8390.2520.040.840.8436437428
25Phenylmethanol0.8040.3110.980.520.6435416416
26Cyclohexanol0.8240.2760.450.840.66435417422
27Pentan-1-ol0.8260.2410.40.860.84432415418
28Butan-1-ol0.8330.2420.470.840.84425416418
291-Phenylethanone0.8330.3120.90.490.04455462466
30Butan-2-one0.850.232 0.480.06452454462
31Cyclohexanone0.850.2690.760.530454454466
322-Methylpropan-1-ol0.8520.2390.40.840.69426414422
33Propan-2-ol0.8520.2340.480.840.76423420417
34Propan-1-ol0.8660.2390.520.90.84425412411.
35Acetone0.8680.2220.620.480.08450456466
36Propane-1,2-diol0.8820.259 0.80.58421414414
374-Hydroxy-4-methylpentan-2-one0.8830.2530.650.550425416420
38Pentane-2,4-dione0.8920.269 0.60451453462
39Ethanol0.8950.2210.860.750.86424414413
40Methanol0.9090.2030.60.660.98416413407
41Propane-1,3-diol0.9150.263 0.850.77411414406
42Acetonitrile0.9210.2190.750.40.19449451450
43Dimethylformamide0.9240.2580.880.690444451454
44Ethane-1,2-diol0.930.2590.90.520.9410409402
45Dimethylsulfoxide0.9460.28110.760446449450
46Propane-1,2,3-triol0.9480.28 1.140.7408405400
47Water0.9640.2061.090.471.17405404402
48Formamide0.9730.2710.970.480.71420415416
Table 2. Percentage contributions of the solvatochromic parameters in the Kamlet–Taft approach.
Table 2. Percentage contributions of the solvatochromic parameters in the Kamlet–Taft approach.
CompoundPπ* (%)Pα (%)Pβ (%)
I128.1752.8019.02
I218.3946.8334.76
I322.0348.0429.92
Table 3. The electro-optical parameters of isolated I1, I2, and I3 molecules, determined by the DFT method.
Table 3. The electro-optical parameters of isolated I1, I2, and I3 molecules, determined by the DFT method.
ParameterI1I2I3
E HOMO (eV)−5.21−5.21−5.23
E LUMO (eV)−2.36−2.53−2.39
µ g   ( D ) 4.564.3194.708
α g   ( A · 3 ) 41.07745.1234.50
A (Å2)306.67328.43360.88
V (Å3)313.87339.65368.71
r33)28.9429.8628.79
Table 4. Equations for determining the dipole moment in the excited state of I1, I2 and I3 molecules by the variational model.
Table 4. Equations for determining the dipole moment in the excited state of I1, I2 and I3 molecules by the variational model.
I1 C 1   r 3 = 6.5830 0.01506 µ e 2 9.12 µ e c o s φ + 35.1347 = 0
( C 1 + C 2 )   r 3 = 29.4539
α e = 40.0125 0.1229 µ e 2 µ 1 , 2 e = 9.12 c o s φ ± 9.12 c o s φ 2.1165 2 0.03012
I2 C 1   r 3 = 11.1686 0.01506 µ e 2 8.64 µ e c o s φ + 26.3688 = 0
( C 1 + C 2 )   r 3 = 20.3918
α e = 44.9074 0.1229 µ e 2 µ 1 , 2 e = 8.64 c o s φ ± ( 8.64 c o s φ ) 2 1.5885 2 0.03012
I3 C 1   r 3 = 8.7163 0.015 µ e 2 9.416 µ e c o s φ + 35.6459 = 0
( C 1 + C 2 )   r 3 = 20.0579
α e = 34.766 0.1225 µ e 2 µ 1 , 2 e = 8.7163 c o s φ ± ( 9.416 c o s φ ) 2 2.1388 2 0.03014
Table 5. The dipole moments and polarizabilities of I1 in the first excited state, calculated in the limits of the variational model.
Table 5. The dipole moments and polarizabilities of I1 in the first excited state, calculated in the limits of the variational model.
φ (degree) µ e   ( D ) α e   ( A 3 )φ (degree) µ e   ( D ) α e   ( A 3 )
03.8838.167618.18−0.61
204.1337.927720.09−9.59
405.0836.847822.58−22.65
607.9132.327926.07−43.54
7516.656.018031.81−84.31
Table 6. The dipole moments and polarizabilities of I2 in the first excited state, calculated in the limits of the variational model.
Table 6. The dipole moments and polarizabilities of I2 in the first excited state, calculated in the limits of the variational model.
φ (degree) µ e   ( D ) α e   ( A 3 )φ (degree) µ e   ( D ) α e   ( A 3 )
03.0743.7578.818.930.87
203.2743.6078.919.19−0.33
403.9443.007919.45−1.60
606.2440.128021.99−14.54
7512.9224.4181.28.67−56.09
78.51.7218.2181.535.46−109.5
Table 7. The dipole moments and polarizabilities of I3 in the first excited state, calculated in the limits of the variational model.
Table 7. The dipole moments and polarizabilities of I3 in the first excited state, calculated in the limits of the variational model.
φ (degree) µ e   ( D ) α e   ( A 3 )φ (degree) µ e   ( D ) α e   ( A 3 )
03.8132.9977
204.0632.757821.87−23.85
404.9931.717925.10−42.42
607.7627.388030.13−76.44
7516.252.418143.24−194.29
Table 8. The calculated values of the dipole moment and polarizabilities in the excited state of I1, I2 and I3, according to Abe’s model of the liquid.
Table 8. The calculated values of the dipole moment and polarizabilities in the excited state of I1, I2 and I3, according to Abe’s model of the liquid.
BfQSolvent Equation µ e   ( D ) α e   ( c m 1 )
I1Aprotic B   =   ( 18.64   ±   3.59 ) +   ( 25.98   ±   0.94 )A1.4725.98
Protic B   =   ( 1.86   ±   4.80 )       ( 82.27   ±   4.50 )A4.76−82.27
I2Aprotic B   =   ( 23.28   ±   4.76 )   +   ( 46.03   ±   1.39 )A4.0246.03
Protic B   =   ( 4.01   ±   3.26 )     ( 56.60   ±   3.23 )A6.59−56.60
I3Aprotic B   =   ( 17.99   ±   5.60 )   +   ( 27.00   ±   1.26 )A4.9227.00
Protic B   =   ( 0.63   ±   5.26 )     ( 79.28   ±   5.17 )A4.77−79.28
Table 9. Calculated dipole moments and Mulliken charges on the donor and acceptor atoms for the ground state of I1, I2 and I3 in vacuum, cyclohexane, and water for the two extreme rotamers.
Table 9. Calculated dipole moments and Mulliken charges on the donor and acceptor atoms for the ground state of I1, I2 and I3 in vacuum, cyclohexane, and water for the two extreme rotamers.
BfQSolventDipole Moment, (D)Dipole Moment, (D)Mulliken Charges (e)Mulliken Charges (e)
PlanarNon-PlanarPlanarNon-Planar
I1Gas4.2564.742Q[-N-] = 0.852
Q[-C-] = −0.824
Q[-N-] = 0.861
Q[-C-] = −0.769
CHX4.4704.76Q[-N-] = 0.898
Q[-C-] = −0.781
Q[-N-] = 0.833
Q[-C-] = −0.809
Water6.0555.968Q[-N-] = 0.880
Q[-C-] = −0.800
Q[-N-] = 0.793
Q[-C-] = −0.770
I2Vacuum4.2095.050Q[-N-] = 0.901
Q[-C-] = −0.731
Q[-N-] = 0.865
Q[-C-] = −0.786
CHx4.326.094Q[-N-] = 0.892
Q[-C-] = −0.718
Q[-N-] = 0.837
Q[-C-] = −0.789
Water6.2837.189Q[-N-] = 0.865
Q[-C-] = −0.762
Q[-N-] = 0.796
Q[-C-] = −0.776
I3Vacuum4.526Q[-N-] = 0.852
Q[-C-] = −0.788
CHx4.825Q[-N-] = 0.842
Q[-C-] = −0.812
Water6.494Q[-N-] = 0.821
Q[-C-] = −0.775
Table 10. Characteristics of I1 in the S1/ICT state in two different conformations, the dipole moment, the hole (blue) – electron (green) distribution during excitation, and the charge density difference between states for the S0 → S1/ICT transition (blue and green parts correspond to a decrease and an increase, respectively, in the electronic density after electronic excitation).
Table 10. Characteristics of I1 in the S1/ICT state in two different conformations, the dipole moment, the hole (blue) – electron (green) distribution during excitation, and the charge density difference between states for the S0 → S1/ICT transition (blue and green parts correspond to a decrease and an increase, respectively, in the electronic density after electronic excitation).
In VacuumPlanar Conformation Non-Planar Conformation
ConformationMolecules 30 03162 i001Q[-N-] = − 0.594
Q[-C-] = 0.082
μe = 2.27 D
Molecules 30 03162 i002Q[-N-] = − 0.565
Q[-C-] = 0.048
μe = 3.53 D
Hole–electron distributionMolecules 30 03162 i003Molecules 30 03162 i004
Charge density differenceMolecules 30 03162 i005Molecules 30 03162 i006
In CHXPlanar ConformationNon-Planar Conformation
ConformationMolecules 30 03162 i007Q[-N-] = −0.576
Q[-C-] = 0.101
μe = 2.77 D
Molecules 30 03162 i008Q[-N-] = −0.565
Q[-C-] = 0.080
μe = 4.87 D
Hole–electron distributionMolecules 30 03162 i009Molecules 30 03162 i010
Charge density differenceMolecules 30 03162 i011Molecules 30 03162 i012
In WaterPlanar ConformationNon-Planar Conformation
ConformationMolecules 30 03162 i013Q[-N-] = −0.576
Q[-C-] = 0.125
μe = 4.15 D
Molecules 30 03162 i014Q[-N-] = −0.565
Q[-C-] = 0.048
μe = 6.09 D
Hole–electron distributionMolecules 30 03162 i015Molecules 30 03162 i016
Charge density differenceMolecules 30 03162 i017Molecules 30 03162 i018
Table 11. The hole (blue) – electron (green) distribution during excitation and the charge density difference between states for the S0 → S1 transition (blue and green parts correspond to a decrease and an increase, respectively, in the electronic density after electronic excitation).
Table 11. The hole (blue) – electron (green) distribution during excitation and the charge density difference between states for the S0 → S1 transition (blue and green parts correspond to a decrease and an increase, respectively, in the electronic density after electronic excitation).
BfQSolvent Hole–Electron DistributionCharge Density Difference
I2vacuumMolecules 30 03162 i019Molecules 30 03162 i020
cyclohexaneMolecules 30 03162 i021Molecules 30 03162 i022
waterMolecules 30 03162 i023Molecules 30 03162 i024
I3vacuumMolecules 30 03162 i025Molecules 30 03162 i026
cyclohexaneMolecules 30 03162 i027Molecules 30 03162 i028
waterMolecules 30 03162 i029Molecules 30 03162 i030
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

Avadanei, M.I.; Avadanei, O.G.; Dorohoi, D.O. Solvatochromic and Computational Study of Three Benzo-[f]-Quinolinium Methylids with Photoinduced Charge Transfer. Molecules 2025, 30, 3162. https://doi.org/10.3390/molecules30153162

AMA Style

Avadanei MI, Avadanei OG, Dorohoi DO. Solvatochromic and Computational Study of Three Benzo-[f]-Quinolinium Methylids with Photoinduced Charge Transfer. Molecules. 2025; 30(15):3162. https://doi.org/10.3390/molecules30153162

Chicago/Turabian Style

Avadanei, Mihaela Iuliana, Ovidiu Gabriel Avadanei, and Dana Ortansa Dorohoi. 2025. "Solvatochromic and Computational Study of Three Benzo-[f]-Quinolinium Methylids with Photoinduced Charge Transfer" Molecules 30, no. 15: 3162. https://doi.org/10.3390/molecules30153162

APA Style

Avadanei, M. I., Avadanei, O. G., & Dorohoi, D. O. (2025). Solvatochromic and Computational Study of Three Benzo-[f]-Quinolinium Methylids with Photoinduced Charge Transfer. Molecules, 30(15), 3162. https://doi.org/10.3390/molecules30153162

Article Metrics

Back to TopTop