Next Article in Journal
On Some Symmetries of Quadratic Systems
Next Article in Special Issue
Symmetry/Asymmetry of the NHN Hydrogen Bond in Protonated 1,8-Bis(dimethylamino)naphthalene
Previous Article in Journal
Bounds on Relativistic Deformed Kinematics from the Physics of the Universe Transparency
Previous Article in Special Issue
First-Principles Calculations of High-Pressure Physical Properties of Ti0.5Ta0.5 Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational and Spectral Means for Characterizing the Intermolecular Interactions in Solutions and for Estimating Excited State Dipole Moment of Solute

by
Dana Ortansa Dorohoi
1,*,
Dorina Emilia Creanga
1,
Dan Gheorghe Dimitriu
1,
Ana Cezarina Morosanu
1,
Antonina Gritco-Todirascu
1,
Gabriel Grigore Mariciuc
2,
Nicoleta Puica Melniciuc
3,
Elena Ardelean
3 and
Corina Cheptea
4
1
Faculty of Physics, Alexandru Ioan Cuza University, 11 Carol I Blvd., RO-700506 Iasi, Romania
2
Faculty of General Medicine, Gr. T. Popa University of Medicine and Pharmacy, 16 Universității Street, RO-700115 Iasi, Romania
3
Faculty of Orthodox Theology, Alexandru Ioan Cuza University, 6 Closca Street, RO-700066 Iasi, Romania
4
Faculty of Biomedical Sciences, Gr. T. Popa University of Medicine and Pharmacy, 9-13 M. Kogalniceanu Street, RO-700454 Iasi, Romania
*
Author to whom correspondence should be addressed.
Symmetry 2020, 12(8), 1299; https://doi.org/10.3390/sym12081299
Submission received: 8 July 2020 / Revised: 21 July 2020 / Accepted: 29 July 2020 / Published: 4 August 2020
(This article belongs to the Special Issue Quantum Chemistry)

Abstract

:
The results obtained both in quantum chemical computation and in solvatochromic study of pyridinium di-carbethoxy methylid (PCCM) are correlated in order to estimate the electric dipole moment in the excited state of this molecule. This estimation is made by a variational method in the hypothesis that the molecular polarizability does not change in time of the absorption process. Ternary solutions of PCCM in protic binary solvents are used here, both establishing the contribution of each type of interaction to the spectral shift and to characterize the composition of the first solvation shell of PCCM. Results are compared with those obtained before for other binary solvents. The difference between the interaction energies in molecular pairs of PCCM-active solvent and PCCM-less active solvent was also estimated based on the cell statistical model of the ternary solutions.

Graphical Abstract

1. Introduction

Knowledge about the electro-optical parameters of molecules in different electronic states obtained by experimental means is of the greatest importance for the researchers working in quantum chemistry field, either for comparison with theoretical data or for overtaking some steps of calculations based on them.
The electronic spectra (absorption or fluorescence) involve the valence electronic cloud very sensitive to the external fields. In solutions, the solute (spectrally active) molecules are under the action of the internal electrostatic field created by the solvent molecules surrounding them (named by Onsager internal reactive field [1]). The strength of the interactions between the solute and a given solvent depends both on the internal reactive field magnitude and on the electro-optical parameters of the spectrally active molecule.
The molecular parameters are usually different in the two electronic states between which the absorption transition takes place. Consequently, the difference of the solvation energies of the solute (in the electronic states responsible for the electronic band appearance) determines the spectral shift [2,3,4] in solvent related to the vacuum state (vaporous state of the solute).
Information on the intermolecular interactions strength in liquids can be obtained from the spectral shifts of the electronic bands. Thus, the spectrally active molecules can be used as probes in testing the internal field in different liquids.
Based on the existent theories regarding the solvent influence on the electronic bands, one can establish the nature and the strength of the intermolecular interactions in different liquids, or one can estimate some electro-optic parameters in the electronic states of the spectrally active molecules [5,6,7,8,9,10] (in the limits in which these theories were developed). This kind of information is useful both for quantum mechanics and for developing the liquid theory, which, in the actual stage, cannot generalize in mathematical treatment all phenomena occurring in this state of mater. The insurmountable difficulty for the liquid theories is the gentle equilibrium between the strength of the internal interactions and the thermal motion energy which generates a continuous order–disorder in the molecular arrangement. For these reasons, the study of the electronic spectra in liquids with various physicochemical properties offers more information about the molecular structure of liquids, compared with FTIR or Raman spectra, which emphasize only the local modification in the structural features of the studied molecules.
Pyridinium di-carbethoxy methylid (PCCM) was chosen as the spectrally active molecule in this study.
Pyridinium ylids are zwitterionic compounds in which the positively charged heterocycle (pyridine) is covalently bonded to a negative carbanion bearing a positive charge [11]. Pyridinium ylids [11,12,13] are important in chemistry, especially because they are precursors in obtaining new organic compounds. Pyridinium ylids are suitable building blocks [14] from the synthesis of indolizines, cyclopropanes, 2(1H)-pyridin-(ethio)-ones, cyclopropanes, 2,3-dihydrofuranes, mono-oligopyridines, pyranonaphthoquinones, nitrones, or azepines. Pyridinium ylids are nucleophilic molecules with long lifetime used in 1,3 dipolar cycloaddition chemistry [15,16,17]. In cycloaddition of pyridinium ylids with alkene made in aqueous buffer at room temperature [17], the highest yield was obtained for electron withdrawing groups attached both to ylids and alkenes.
The carbanion centers of the ylid-partial structures can serve as electron donators in the reversible color changes [18] for the photochromatic Merrifield resins.
Pyridinium ylids are used in some applications based on polymers [19,20,21]. The UV exposure of pyridinium ylid accompanied by a change in polarity was applied in photo resistors. Polymers containing pyridinium ylids were prepared [19] as water soluble MID UV resist materials. Pyridinium ylids and their salts are used to obtain photochromic materials, such as photosensitive resins [20], due to their reversible color change in which the carbanion lone electron pair is implied. Some novel negative photoresists were obtained [20] based on the photo isomerization of water- soluble polymeric pyridinium ylid.
Polymers containing pyridinium ylid moieties were subjected to actinic irradiation in order to convert the pyridinium ylid moieties to water-insoluble N-acyl diazepine polymers [21], used for the water-proofing or hydrophobization of surfaces, as well as for the production of photoresists, printing plates, and printed circuit boards.
In material science, pyridinium ylids were also used for obtaining thin conducting layers [22,23].
As all cycloimmonium ylids [14,15,16,17], pyridinium ylids show a visible electronic absorption band due to an intramolecular charge transfer from the carbanion toward the heterocycle. This band is very sensitive to the solvent nature. Due to this spectral property, pyridinium ylids can be used as acid–basic indicators [11,24].
Pyridinium ylids were used in physics as probes to characterize the strengths of the internal electric field in solutions [11,24].
Both the quantum mechanical and solvatochromic studies can give interesting details regarding the structural features in the electronic state responsible for the pyridinium ylid visible band appearance [24,25,26,27,28].
Pyridinium di-carbethoxy methylid (PCCM) [11] is an amphiphilic compound with the positive charge located on the nitrogen atom belonging to pyridine heterocycle and the negative charge distributed on the carbanion double substituted by the carbethoxy groups (Figure 1). The symmetrically substituted carbanion increases the stability of PCCM [29,30,31].
The aim of this study was to use the spectral data obtained for PCCM binary and ternary solutions in order to obtain information about the excited state dipole moment of ylid and the nature and the strength of the intermolecular interactions in diluted ylid solutions. The difference between the potential energies in molecular pairs achieved between the ylid and the hydroxyl solvent molecules in ternary solutions is also estimated in this study.
The inhomogeneity of the first solvation shell of ylid in ternary solutions was also evidenced, by estimating the molar fractions of the two solvents in the first solvation shell of PCCM. The results can be compared with those obtained for a pyridinium ylid with non-symmetrical carbanion, pyridinium acetyl benzoyl methylid (PABM), or for benzo-[f]-quinolinium acetyl benzoyl methylid (BQABM) published before [28,32].

2. Experimental and Computational Details

The studied substance was prepared as it is described in Reference [11], in the Alexandru Ioan Cuza University labs. The purity of the obtained substance was checked by spectral (NMR and IR) and chemical (elemental) means.
The spectrally grade solvents were achieved from Merck Company and Fluke (Sigma-Aldrich) and used without any purification.
The solvent parameters were taken from https://www.stenutz.eu./chem.solv26.php.
The binary solvents used in this study were achieved in molar fractions from double-distilled water, ethanol, and methanol.
The visible electronic absorption band was recorded at a Specord UV–Vis Carl Zeiss Jena spectrophotometer.
The quantum mechanical analysis was made by Spartan’14 program, using EDF2/6–31G* density functional model [33].
Statistical analysis was made by using Origin 9.

3. Theoretical Notions

The optimized geometry (the arrangement of molecular atoms for which the molecular energy is minim) and some energetic, electro-optical, and QSAR parameters of the PCCM molecule in electronic ground state were established in computational estimations with Spartan’14 program.
Some of the data obtained in quantum mechanical analysis are used here for estimating, in a variational method, the excited state dipole moment of PCCM based on solvatochromic results. In this method, a relation is established between the excited dipole moment and the polarizability of PCCM, and then the excited state dipole moment is determined in the McRae [2] and Kawschi [7,34] hypothesis that the molecular polarizability does not change during of electronic transition responsible for the visible band appearance.
The solvent influence on the visible band of PCCM was explained both by using empirical parameters introduced by Kamlet Taft [35] and also by using Relation (2), derived by the theory of diluted solutions developed by McRae [2], Bakhshiev [3], Abe [4], and others, in the hypothesis that only long-range (universal or non-specific) interactions determine the spectral shifts. In this case, the influence of the specific interactions is taken into consideration by two empirically introduced terms in Relations (1) and (2).
v ¯ ( cm 1 ) = C 1 + C 2 π * + C 3 α + C 4 β ,
v ¯ ( cm 1 ) = v 0 + C 1 f ( ε ) + C 2 f ( n ) + C 3 α + C 4 β ,
with
f ( ε ) = ε 1 ε + 2   and   f ( n ) = n 2 1 n 2 + 2
The significance of the parameters in (1)–(3) is the following: v ¯ ( cm 1 ) is the wavenumber in the maximum of the intramolecular charge transfer (ICT) band of PCCM; ε and n are the solvent electric permittivity and refractive index; π * is the solvent polarity/polarizability indicator; and α (HBD) and β (HBA) are the solvent abilities to donate or to accept protons in hydrogen bonds [27,36]. The empirically introduced solvent parameters [35] π * , α, and β are named Kamlet Taft parameters and are usually used in solvatochromic studies [37,38].
The correlation coefficients, C1 and C2, from Relation (2) depend, as indicated by Relations (4) and (5), on molecular dipole moments, µ; polarizability, α; ionization potentials, I; and temperature, T, at which the experiments are made. The indices g and e refer to ground and excited electronic states, and u and v refer to the solute and solvent molecules, respectively.
C 1 = 2 μ g ( μ g μ e cos φ ) h c a 3 + 3 k T α g α e a 3 ,
C 2 = μ g 2 μ e 2 h c a 3 2 μ g ( μ g μ e cos φ ) h c a 3 3 k T α g α e a 3 + 3 2 α g α e a 3 I u I v I u + I v .
In (4) and (5), a is the molecular radius, approximated (in the hypothesis that the molecule is a spherical entity) by using Relation (6), and φ is the angle between the dipole moments.
a = 3 V A
The volume, V, and the area, A, from (6) were obtained here by quantum mechanical computations (see Table 1).
The correlation coefficients, C1 and C2, resulting from the solvatochromic analysis, can be used for estimating both the excited state dipole moment and the angle between the electric dipole moments in the electronic state responsible for the electronic band appearance when some electro-optical parameters of the studied molecule are obtained by other methods (for example, by computational methods). The results obtained in solvatochromic study and the molecular parameters for the PCCM ground electronic state established by Spartan’14 are used here to estimate the dipole moment in the excited state of this molecule.
For the molecules showing electronic spectra both in absorption and fluorescence [3,8,34,39], the correlation coefficients, C1 and C2, are usually transformed in an equations system by introducing molecular parameters of the ground electronic state obtained by quantum mechanical analysis. The excited state dipole moment, the excited state polarizability, and also the angle between the solute dipole moments in the two states of electronic transition are estimated in the limits of approximation in which the theory of liquids was developed [2,3,4,5].
There are molecules showing only absorption spectra. In their case, the values of the coefficients, C1 and C2, and the electro-optical parameters theoretically obtained could offer information about the excited state dipole moment by using a variational method [40] applied with good results for other molecules [10,38].
Some relations between the excited state dipole moment, the polarizability, and angle between the dipole moments in the electronic states of the transition are established in this method. Based on these relations, the excited state and polarizability can be computed at each angle (considered as being variable) between the dipole moments. The absorption transition is supposed to be achieved when the excited state polarizability of the solute molecule equalizes the ground state polarizability, in accordance with McRae’s [2] or Kawski’s [7,34] approximations.
The majority of the publications [40,41,42,43,44] takes into consideration the ground electronic state parameters of the spectrally active molecules computed in quantum mechanics for isolated molecules, in their vaporous state, while in solutions, their parameters can drastically be changed by interactions. This approximation could also influence the value of the excited state dipole moment estimated by using computational and solvatochromic results.
Ternary solutions were used both to enlarge the physical parameters of the liquids used in solvatochromic study and to characterize the local structure of the diluted solutions of PCCM.
Additionally, the statistic cell model of ternary solutions [45,46,47,48,49] applied to PCCM permitted us to estimate the difference between the interaction energies in molecular pairs of the types: ylid-active solvent (1) and ylid-inactive solvent (2). This is a very important result because there are very few methods for estimating the intermolecular energy for molecular pairs in solutions.
The average statistic weight of the most active (from the interactions point of view) solvent in the first solvation shell of the solute molecule, noted by p 1 , must be defined in the statistical cell model [50,51] of the diluted ternary solutions achieved in binary homogeneous liquids, because the composition of the first solvation shell of the spectrally active molecule can differ from the composition of the liquid mixture in the bulk solution. This parameter is the average percent of the active solvent molecules in the first solvation shell of the solute molecule, which can differ by the molar concentration of this solvent in the bulk solution, noted by x 1 . Relations of the type p 1 + p 2 = 1 and x 1 + x 2 = 1 are true for the two solvents of ternary solution.
Between these parameters, a linear relation of Type (7) was established [28,32] with the slope near the unity and the cut dependent on the difference w 2 w 1 between the interaction energies in molecular pairs ylid-active solvent (indexed by 1) and ylid-less active solvent (indexed by 2).
ln p 1 p 2 = ln x 1 x 2 + w 2 w 1 k T
Relation (7) offers the possibility to estimate the difference w 2 w 1 , which is practically impossible to be determined by other methods, except the spectral ones.
The average statistic weight of the active solvent (indexed by 1), p 1 , in the first solute solvation shell can be calculated by using the values of the wavenumbers in the maximum of the ICT band of the spectrally active molecule determined in the binary solutions realized in the two solvents and in the ternary solution, indexed with 1, 2, and t, respectively:
p 1 = v t v 2 v 1 v 2   and   p 2 = v 1 v t v 1 v 2
The excess function is as follows:
δ 1 = p 1 x 1
Reference [50] defines an indicator of the ternary solution non-homogeneity. When δ 1 < 0 , the molecules of the inactive solvent (indexed by 2) are predominant in the first solvation shell of the solute. When δ 1 > 0 , the molecules of the active solvent are dominant in the solute first solvation shell.

4. Results and Discussion

4.1. Computational Results

PCCM is a pyridinium ylid with symmetrically substituted carbanion having planar structure at the carbanion level. The ylid carbanion is sp2 hybridized, so the two covalent simple bonds between it and the carbethoxy groups lie in the plane of the heterocycle. The lone electron pair of the carbanion is perpendicular on this plane, facilitating the heterocycle π electrons’ conjugation with the electrons belonging to C=O groups of carbethoxy substituents [11].
The Spartan evaluations were made in vacuum, in water, and in ethanol. As one can see from Figure 1, the symmetry of the molecule is not changed in the two hydroxyl solvents, compared with the vacuum structure.
The ground state electric dipole moment of the studied molecule keeps its orientation, parallel to the ylid bond both in water and in ethanol, but changes its absolute value, as it results from the data of Table 1. In the same time, the molecular polarizability does not support significant change when PCCM is passed from vaporous state in solution with water or ethanol (see Table 1).
A significant increase of dipole moment takes place when a PCCM molecule is solvated in water or ethanol, hydroxyl solvents in which hydrogen bond are possible between the ylid carbanion electron lone pair or the electrons of C=O groups and their hydroxyl groups. As seen from Table 1, three places are possible for a hydrogen bond addition to PCCM (three HBA counts are possible for PCCM).
The changes in the EHOMO and ELUMO values listed in Table 1 at passing from vacuum to solutions in water or in ethanol are important because they show us that, in these solvents, the visible band of PCCM will be shifted to blue, which is related to its vaporous state.
The changes in area and volume of PCCM influence the dimensions of the solvation shell of this molecule. The deviation of the molecular structure at sphericity demonstrates that the theories of liquids can be applied only in some approximations, because these theories consider the spectrally active molecules as being spherical entities surrounded by solvent spherical layers.
Let us observe, in Figure 2, some change in the electronic charges near the molecular atoms of PCCM in water or ethanol, compared with vaporous state. The charges located near the atoms component of ylid bond or that al the level of C=O groups of the ylid carbanion are important for the interactions of this molecule with the hydroxyl solvent molecules, as this will result from the solvatochromic analysis.
The computational study offers some information about the electron transitions in the visible absorption process, as it results from Figure 3a–c, in which the aspects of the HOMO and LUMO orbitals are illustrated.
The similitude in the orbital aspects in hydroxyl solvents can be observed in Figure 3b,c. All three figures show that, by the visible photon absorption, the heterocycle will be enriched in electron density, thus demonstrating that intramolecular charge transfer (ICT) process takes place from the carbanion and its substituents toward the ylid heterocycle.
An electrostatic potential map (for elucidation the molecular charges distribution), local ionization potential map, and |LUMO| map of PCCM (anticipating the electrophilic and nucleophilic reactivity, respectively) in vaporous phase and in the two hydroxyl solvents are illustrated in Figure 4. Colors in these figures have the signification [28,32] specified below.
The electrostatic map corresponds to the electronic reactivity, providing a measure of the overall size and shape of the molecule. The red color corresponds to high negative potentials, while the blue color indicates high positive potentials.
The electron removing energy as a function of its location is shown by the local ionization potential map. The local ionization potentials are always positive. The blue color indicates large ionization potentials, while the red color corresponds to small ionization potentials.
The |LUMO| map displays the absolute value of LUMO onto the electron density surface. The red color indicates small absolute values of LUMO, and the blue color shows large absolute values of it.
The places of nucleophilic attack (located near carbanion and near the two oxygens of C=O groups of carbethoxy substituents) can be observed in Figure 4a–c for PCCM in vaporous phase, or in solutions in water and ethanol, respectively.
The electrostatic potential, local ionization potential and |LUMO| maps keep similar aspects for the vaporous state and also in solutions of PCCM achieved with water and ethanol.
Important for our study are the orientation of the electric dipole moment of PCCM and the aspects of the HOMO and LUMO orbitals, giving information about the sense of the electronic transition in the visible light process and also about the strength of the intermolecular interactions of PCCM in solutions.
The values of the volume and molecular surface area (Table 1) were used to estimate the molecular radius in expressions of the correlation coefficients, C1 and C2, in Relations (4) and (5).

4.2. Spectral Results

PCCM shows a UV region with an intense band attributed to π-π* transitions and a visible band of small intensity, sensitive to the solvent nature, which disappears in acid media, attributed to n-π* transition realized by electronic charge transfer from the carbanion towards the ylid heterocycle (see Figure 3 and Figure 5).
The electronic absorption spectrum in ternary solution obtained by adding sulphuric acid in the PCCM solution in ethanol is represented in Figure 5 by a dashed line. The protonated ylid shows spectral changes both in UV and in visible range, as it results from Figure 5. The visible band of PCCM disappears in the acid media in which the lone electron pair of the ylid carbanion is blocked by protonation [11,48].
The spectral data regarding the binary solutions of PCCM are given in Table 2. The solvent parameters are also listed in Table 2. The computed values of the wavenumbers in Table 2 were made by using Relations (10) and (11).
From the data of Table 2, we see the spectral shifts to blue of the visible electronic absorption band both in the polar and in the protic solvents, compared to the non-polar and aprotic ones. Based on the disappearance in acid solutions and on its small intensity, the visible electronic absorption band of PCCM has been attributed to intramolecular charge transfer (ICT) from the carbanion toward the heterocycle [11,24]. This attribution is sustained by the aspects of HOMO and LUMO orbitals (see Figure 3) and by other solvatochromic and computational studies made on similar structures [28,32].
As seen in Table 2, the visible ICT band is very sensitive to the solvent nature and can help us to establish the nature and the strength of interactions between the PCCM molecule and the solvents in diluted solutions.
In order to analyze the solvent effects on the visible electronic absorption band of PCCM, relations of Types (1) and (2) were used.
Based on Kamlet Taft parameters [5,35] of the used solvents (Table 2), the correlation coefficients obtained for Relation (1) in the statistical analysis are listed below.
v c a l c .   = ( 22059 ± 118 ) + ( 1044 ± 156 ) π * + ( 2006 ± 118 ) α + ( 791 ± 171 ) β
Adj. R-square = 0.948, SD = 247 cm−1 and
v c a l c .   = ( 1152 ± 807 ) + ( 0.951 ± 0.034 ) v e x p , with: Adj. R-square = 0.950, SD = 235 cm−1
The dependence between the experimental wavenumbers in the ICT band maximum and those computed by Relation (10) values is illustrated in Figure 6.
The Kamlet Taft solvent parameters were empirically introduced, and they cannot be bonded with the electro-optical parameters of the spectrally active molecules.
For this reason, Relation (2) was also applied to the spectral data given in Table 2, because the correlation coefficients, C1 and C2, are expressed as function on the molecular parameters of the spectrally active molecule [28,32] and offer the possibility to characterize the excited state of this molecule, as it results from Relations (3) and (4). The correlation coefficients obtained based on Relation (2) and the data from Table 2 are listed below. The results of statistical analysis of spectral data based on Relation (11) are illustrated in Figure 7, in which the computed values of the ICT wavenumbers are plotted vs. those experimentally obtained.
v c a l c .   = ( 22026 ± 419 ) + ( 1642 ± 254 ) f ( ε ) + ( 261 ± 1412 ) f ( n ) + ( 1695 ± 139 ) α + ( 135 ± 213 ) β
Adj. R-square = 0.943, SD = 253 cm−1 and
v c a l c .   = ( 1232 ± 843 ) + ( 0.948 ± 0.036 ) v e x p , with: Adj. R-square = 0.947, SD = 237 cm−1.
Having in view the high values of the errors affecting the correlations coefficients which multiply the solvent parameters f ( n ) and β, these parameters were eliminated from statistics and the results are the following:
v c a l c .   = ( 21942 ± 131 ) + ( 1731 ± 212 ) f ( ε ) + ( 1735 ± 122 ) α
Adj. R-square = 0.945, SD = 248 cm−1 and
v c a l c .   = ( 1247 ± 848 ) + ( 0.947 ± 0.036 ) v e x p , with: Adj. R-square = 0.946, SD = 238 cm−1.
Relation (12) describes, in a good manner, the dependence of the wavenumber in the maximum of the ICT band of PCCM on the solvent parameters in the binary solutions. The dependence, based on Relation (12), of the computed wavenumbers vs. those experimentally determined for the PCCM binary solutions is illustrated in Figure 7.
The solvatochromic analysis was also made for ternary solutions realized in binary hydroxyl solvents water (1) + ethanol (2) and for water (1) + methanol (2) with the Kamlet Taft solvent parameters from [51]. The solvent parameters and the wavenumbers in the maximum of ICT visible band of PCCM are listed in Table 3 and Table 4. The data from Table 3 and Table 4 demonstrate that δ 1 (excess function of the active solvent) is positive, that the first solvation shell of PCCM in ternary solutions is enriched in water molecules (water was considered as being the active solvent from the interactions point of view).
The dependence of Type (1) was observed for the ternary PCCM solutions in both pairs of binary solvents. The correlation coefficients obtained using Relation (1) and the data from Table 3 and Table 4 for the two types of solutions are given in Relations (13) and (14). The dependence between the computed, v calc (cm−1) and the experimental wavenumber, v exp (cm−1), and for PCCM in water (1) + ethanol (2) and in water (1) + methanol (2) is illustrated in Figure 8 and Figure 9.
The presence of some aberrant points for high water content in Figure 8 could be explained [52] by the increased ability of water to make some cages around the hydrocarbon part of superior alcohols. This ability increases with the length of hydrocarbon chain from the alcohol molecules, and it was emphasized for water + ethanol mixtures [51].
Results or ternary solution PCCM in water (1) + ethanol (2):
v c a l c .   = ( 20153 ± 533 ) + ( 2174 ± 165 ) π * + ( 3451 ± 367 ) α + ( 862 ± 145 ) β
Adj. R-square = 0.967
v c a l c .   = ( 736 ± 886 ) + ( 0.971 ± 0.035 ) v e x p , with: Adj. R-square = 0.950, SD = 235 cm−1
Results for solution PCCM in water (1) + methanol (2):
v c a l c .   = ( 23772 ± 817 ) + ( 827 ± 62 ) π * + ( 961 ± 148 ) α + ( 194 ± 71 ) β
Adj. R-square = 0.992
v c a l c .   = ( 180 ± 445 ) + ( 0.993 ± 0.018 ) v e x p . , with: Adj. R-square = 0.993, SD = 100 cm−1
The contribution of each type of intermolecular interaction in PCCM diluted solutions can be estimated based on the data from Table 2, Table 3 and Table 4 and the values of the correlation coefficients established in statistical analysis. As it results from the linear dependences expressed in Figure 6, Figure 7 and Figure 8, in PCCM, diluted binary and ternary solutions act both universal and specific interactions.
The universal interactions of the orientation–induction type are dominant in aprotic solvents, while, in the hydroxyl solvents, a great contribution to the spectral shift of ICT visible band of PCCM is given by specific interactions between the ylid and the hydroxyl groups of these solvents (see Table 5 and Table 6).
The dependence of the type ln p 1 p 2 vs. x 1 x 2 , where p1 and x1 are the statistic average weight of the active solvent (considered by us as being water) in the first solvation shell of PCCM and the molar fraction of the same solvent in the bulk solution, respectively, for PCCM ternary solutions is illustrated in Figure 10 and Figure 11. The index 2 shows the same parameters for the second solvent, considered as being a few active from the interactions point of view.
From Figure 10 and Figure 11, it results a linear dependence between the points corresponding to both types of the binary hydroxyl solvents, proving the applicability of the statistic cell model of ternary solutions to the studied mixtures.
The difference between the interaction energies in the molecular pairs of the types PCCM–water (1) and PCCM–ethanol (2) and, respectively, PCCM–Water (1) and PCCM–Methanol (2) was estimated based on the significance of the cut at origin of the corresponding lines (see Figure 10 and Figure 11).
The difference w 2 w 1 , determined using the cut of line (7) for the ternary solution PCCM + water (1) + ethanol (2), has the value w 2 w 1 = ( 0.574 ± 0.013 ) 10 20 J , and for solution PCCM + water (1) + methanol (2) has the value w 2 w 1 = ( 0.475 ± 0.010 ) 10 20 J . These values demonstrate (by their positive signs) that the interaction energy of ylid molecule with a water molecule is higher in modulus compared with those between the ylid and one alcohol molecule (either ethanol or methanol).
The difference w 2 w 1 is higher for the case of ternary solutions in water + ethanol, proving that the interaction energy in molecular pairs: ylid–ethanol is smaller compared with that corresponding to pairs: ylid–methanol.
A study of the protic solvent influence on the visible ICT band of pyridinium di-carbethoxy methylid and of iso-quinolinium di-carbethoxy methylid was also accomplished in Reference [27]. The difference w 2 w 1 was estimated for the binary solvents containing one protic and one aprotic solvent. The w 2 w 1 values obtained in Reference [27] for pyridinium di-carbethoxy methylid were 0.289 10 20 J for the binary solvent octanol (1) + dichloroethane (2) and 0.190 10 20 J for the binary solvent methanol (1) + benzene (2).
The higher values obtained in this paper for binary solvents containing water and ethanol or methanol, compared with those obtained in Reference [27], could be explained by the action of the orientation–induction forces on the complexes formed by ylid with hydroxyl molecule, having in view the high value of the water electric permittivity. A study of the orientation–induction interactions in ternary solutions pyridinium ylids + water + ethanol has been made in Reference [53,54] and emphasized the great contribution of the orientation interactions on the ylid complexes realized by hydrogen bonds.
The experimental data regarding the interactions between two molecules are important for the researchers who study the interactions in liquids, because there are few methods in which these types of data can be obtained (any of the devices introduced in liquids for measurements disturb the internal equilibrium in this state of matter). Only spectrally active molecules are used as probes able to estimate, by spectra modifications, the strength of the reactive internal field in liquids.
Information about the excited states of the spectrally active molecules, especially about their excited electric dipole moment and its orientation relative to the ground state dipole moment are important for quantum chemistry and can serve to develop the calculations in the molecular excited states.
The dipole moment of PCCM in its first excited state (corresponding to the ICT process caused by the visible photon absorption) is estimated here by using the variational method proposed in Reference [28,32,41] and the data of PCCM in the ground state obtained in quantum chemical estimation (Table 1).
With the values of the correlation coefficients, C1 and C2, obtained in (11) based on Relations (3) and (4) and using the computed by Spartan’14 values of the ground state electric dipole moment and polarizability of PCCM, the variational method proposed in Reference [41] permitted us to estimate the excited state dipole moment of PCCM and the angle between the dipole moments in the electronic state between which the absorption transition takes place, in the approximation that the molecular polarizability does not vary in the electronic transition time [2,34].
Let us apply this method for the PCCM by using the data from Table 1, the values of the coefficients, C1 and C2, from Relation (11), and the molecular radius computed by using the surface area and volume in Formula (6).
The estimation is made in some approximations, having in view the ovality of PCCM molecule obtained by Spartan (see Table 1) and the hypothesis of the cell model regarding the solute and solvation shell sphericity. Using Definitions (4) and (5) and the values of the coefficients, C1 and C2, one obtains the following:
α e = 64.3783 0.1168 µ e 2
0.0143 µ e 2   12.56   µ e cos φ + 72.1007 = 0
Equations (15) and (16) where obtained by using the values for the sum C 1 + C 2 and C 1 regression coefficients from Relation (11) and the ground state dipole moment of PCCM computed in water (Table 1).
Equation (16) has real solutions in the hypothesis that its discriminator is a positive quantity. A limitation for the angle, φ, between the dipole moments results in this condition.
φ [ 0 ,   80.69 ) ( 99.31 , 180 ] °
The excited state dipole moment, obtained as the real solution of Equation (16), is used for estimating the excited state polarizability of PCCM based on Relation (15).
In variational method, the angle between the molecular dipole moments in the electronic states participating to the absorption process is varied until the polarizability in the excited state of the spectrally active molecule equalizes the ground state polarizability. In conformity with the supposition of McRae [2], the electronic transition takes place for the corresponding value of the angle, φ. The data are listed in Table 7.
In Table 7, only some of the estimations are listed, because for φ = 8.5°, the value of the excited state polarizability equalizes the ground state polarizability, and one can appreciate that the electronic transition with ICT takes place at this angle. The results obtained when the dipole moments triangle (Figure 12) was solved are µ e = 5.84 D , Δµ = 1D, φ = 8.5°, for initial values µ g = 6.28 D , computed by Spartan for PCCM in water; I u = 5.58   eV and I v = 9.96   eV (for cyclohexane); and C1 = 1642 cm−1 and C2 = −261 cm−1. In Figure 12, the angle between the ground state dipole moment and the vector Δµ is 59.72 degrees, and the angle between the excited state dipole moment and the vector Δµ is 111.78 degrees.
In the case when only f ( ε ) and α influence the wavenumber in the maximum of the ICT band (see Relation (12’)), the correlation coefficients are C1 = 1731 cm−1 and C2 = 0 cm−1. A system of equations of the Types (18) and (19) is obtained in this case.
α e = 64.2213 0.1168 µ e 2
0.0143 µ e 2 12.56   µ e cos φ + 71.7802 = 0
Equations (18) and (19) have the solutions as it follows.
The real solutions are obtained for φ [ 0 ,   80.71 ) ( 99.28 , 180 ] °, but the condition that the excited state polarizability is the same as in the ground electronic state imposes the following solutions: φ = 0°, µ e = 5.75 D, and Δµ = 0.53D. The parallelism between the ground and excited dipole moments of the molecule in the time of ICT can be admitted with good precision, having in view the approximation in which the theory regarding the solvent influence on the wavenumbers of the electronic transitions in the absorption process has been developed.
In the case of symmetrically substituted carbanion, the angle between the dipole moments process is small, and the excited state dipole moment is smaller than the ground state dipole moment. This fact determines the hypsochromic shift of the ICT band in polar solvents, because the solvation energy of the orientation induction type is higher in the ground state of PCCM molecule, compared with that in its excited state.

5. Conclusions

The experimental data obtained in solvatochromic study and quantum mechanical parameters computed by Spartan can be used in order to characterize the excited electronic state of a spectrally active molecule without fluorescence spectrum.
The excited state dipole moment of PCCM in the excited state of the ICT transition, estimated here by the variational method, is smaller than the ground state dipole moment, and it makes a small angle with this.
The long-range (universal) interactions of the dipolar types of PCCM with the solvents are stronger in the ground electronic state, as compared with the excited state.
The universal interactions are predominant in PCCM solutions achieved in aprotic solvents, while in the protic ones, the specific interactions of the hydrogen types in which the ylid molecules accept protons significantly influence the position of the visible ICT band of ylid.
A part of experimental data obtained in this research could be used by researchers in the quantum chemistry field for developing quantum chemical computations for the excited state parameters of the ylid molecules in different liquids.

Author Contributions

Spectral studies in binary and ternary solutions, D.O.D., D.E.C., D.G.D., and G.G.M.; computational studies of PCCM molecule in vacuum, in water and in ethanol, by SPARTAN 14, A.C.M., A.G.-T., and C.C.; chemical reactions and analyses, N.P.M. and E.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Onsager, L. Electric moments of molecules in liquids. J. Am. Chem. Soc. 1936, 58, 1486–1493. [Google Scholar] [CrossRef]
  2. McRae, E.G. Theory of solvent influence on the molecular electronic spectra. Frequency shifts. J. Phys. Chem. 1957, 61, 562–572. [Google Scholar] [CrossRef]
  3. Bakhshiev, N.G. Spectroscopy of Intermolecular Interactions; Nauka: St. Petersburg, Russian, 1972. (In Russian) [Google Scholar]
  4. Abe, T. Theory of solvent influence on the molecular electronic spectra. Frequency shifts. Bull. Chem. Soc. Jpn. 1965, 38, 1314–1318. [Google Scholar] [CrossRef] [Green Version]
  5. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  6. Dorohoi, D.O.; Holban, V. Intermolecular interactions in some pyridazinium ylids solutions. J. Mol. Struct. 1993, 293, 133–136. [Google Scholar] [CrossRef]
  7. Kawski, A. Ground and excited state dipole moments of 6-propionyl-2-(dimethyl amino)-naphthalene determined by solvatochromic shifts. Z. Naturforsch 1999, 54, 379–381. [Google Scholar] [CrossRef]
  8. Zakerhamidi, M.S.; Moghadam, M.; Ghanadzadeh, A.; Hosseini, S. Anisotropic and isotropic solvent effects on the dipole moment and photophysical properties of rhodamine dyes. J. Lum. 2012, 132, 931–937. [Google Scholar] [CrossRef]
  9. Stanculescu, E.R.; Stroia, L.; Dorohoi, D.O. Solvatochromism Used in Determining Some Electro-Optical Parameters. Proceedings SPIE nr. 8001, 2011, nr. Art. 80012S, Published in 2011. Available online: https://www.spiedigitallibrary.org/conference-proceedings-of-spie/8001/80012S/Solvatochromism-used-in-determining-some-molecular-electro-optical-parameters/10.1117/12.891949.short (accessed on 20 June 2020).
  10. Ivan, L.M.; Dimitriu, D.G.; Gritco-Todirascu, A.; Morosanu, A.C.; Dorohoi, D.O.; Cheptea, C. Excited state dipole moments of two pyridazinium p-nitro-phenacylids estimated from solvatochromic study. Spectrosc. Lett. 2020, 53, 1–11. [Google Scholar] [CrossRef]
  11. Zugravescu, I.; Petrovanu, M. N-Ylid Chemistry; Academic Press: New York, NY, USA, 1976. [Google Scholar]
  12. Beak, P.; Reitz, D.B. Dipole stabilized carbanions: Novel and useful intermediates. Chem. Rev. 1978, 78, 275–316. [Google Scholar] [CrossRef]
  13. Radzhabov, M.R.; Tsyganov, D.V.; Pivina, T.S.; Krayushkin, M.M.; Seeboth, N.; Ivanov, S.; Frederique-Salit, A. Syntheses of pyridinium ylids and simulation of their 1, 3 dipolar Cycloaddition mechanism. Mendeleev Commun. 2017, 27, 503–505. [Google Scholar] [CrossRef]
  14. Jacobs, J.; van Hende, E.; Claessens, S.; De Kimpe, N. Pyridinium ylids in heterocyclic synthesis. Curr. Org. Chem. 2011, 15, 1340–1362. [Google Scholar] [CrossRef]
  15. Pawda, A.; Austin, D.J.; Percedo, L.; Lin, Z. Cycloaddition reactions of pyridinium and related azomethine ylides. J. Org. Chem. 1993, 58, 1144–1150. [Google Scholar] [CrossRef]
  16. Allgäuer, D.S.; Mayer, P.; Mayr, H. Nucleophilicity parameters of pyridinium ylides and their use in mechanistic analyses. J. Am. Chem. Soc. 2013, 135, 15216–15224. [Google Scholar] [CrossRef] [PubMed]
  17. Bonte, S.; Ghenea, I.O.; Dinica, R.; Baussanne, I.; Demeunynck, M. Investigation of the pyridinium ylide-alkene Cycloaddition as a fluorogenic coupling reaction. Molecules 2016, 21, 332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Albrecht, M.; Yulikov, M.; Kohn, T.; Jeschke, G. Pyridinium salts and ylides as partial structures of photoresponsive Merrifield resins. J. Mat. Chem. 2010, 20, 3025–3034. [Google Scholar] [CrossRef]
  19. Schwalm, R.; Bottcher, A.; Koch, H. MID UV resistant materials containing pyridinium ylides. Proc. SPIE 1988. [Google Scholar] [CrossRef]
  20. Streith, J. The photochemistry of N-iminopyridinium ylides in retrospect. From a simple concept to same applications. CHIMIA 1991, 45, 65–76. [Google Scholar]
  21. Taylor, L.D.; Karl, M.; Haubs, J. Polymeric Pyridinium Ylide and Products from Same; European Patent Office: Munich, Germany, 1985.
  22. Leontie, L.; Olariu, I.; Rusu, I.G. On the charge transport in some new carbanion disubstituted in thin films. Mater. Chem. Phys. 2003, 80, 506–511. [Google Scholar] [CrossRef]
  23. Leontie, L.; Roman, M.; Branza, F.; Podaru, C.; Rusu, I.G. Electrical and optical properties of some new synthetized ylids in thin films. Synth. Mat. 2003, 138, 157–163. [Google Scholar] [CrossRef]
  24. Dorohoi, D.O. Electronic spectroscopy of N-ylids. J. Mol. Struct. 2004, 704, 31–43. [Google Scholar] [CrossRef]
  25. Pop, V.; Dorohoi, D.O.; Holban, V. Molecular interactions in binary solutions of 4-aminophthalimide and 3p-cumyl-pyridazinium acetyl benzoyl methylid. Spectrochim. Acta A 1994, 50, 2281–2289. [Google Scholar] [CrossRef]
  26. Dorohoi, D.O.; Partenie, H.D. The Spectroscopy of N-ylids. J. Mol. Struct. 1993, 293, 129–132. [Google Scholar] [CrossRef]
  27. Gheorghies, C.; Gheorghies, L.V.; Dorohoi, D.O. Solvent influence on some complexes realized by Hydrogen bond. J. Mol. Struct. 2008, 887, 122–127. [Google Scholar] [CrossRef]
  28. 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] [PubMed]
  29. Jorgenson, C.K. Symbiotic ligands, hard and soft central atoms. Inorg. Chem. 1964, 3, 1201–1202. [Google Scholar] [CrossRef]
  30. Pearson, R.G.; Songstad, J. Application of the principle of Hard and Soft acids and bases to Organic Chemistry. J. Am. Chem. Soc. 1967, 89, 1827–1836. [Google Scholar] [CrossRef]
  31. Surpateanu, G.; Dorohoi, D.O. Determination of the group electronegativity. An. Univ. Al. I. Cuza Sect. Ib Fizica 1977, 23, 99–102. [Google Scholar]
  32. 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]
  33. Spartan’14 for Windows, Macintosh and Linux, Tutorial and User’s Guide, January 10, 2014 Modeling. 2014. Available online: http://downloads.wavefun.com/Spartan14Manual.pdf (accessed on 3 June 2020).
  34. Kawschi, A. On the estimation of excited state dipole moments from solvatochromic shifts of absorption and fluorescence spectra. Z. Nat. 2002, 57, 255–262. [Google Scholar]
  35. Kamlet, M.J.; Abboud, J.I.; Abraham, M.; Taft, R.W. Linear solvation energy relationships 23. A comprehensive collection of the solvatochromic parameters π*, α,β and some methods for simplifying the generalized solvatochromic equation. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  36. Nadejde, C.; Creanga, D.E.; Humelnicu, I.; Dorohoi, D.O. Study of intermolecular interactions in rifampicin ternary solutions. J. Mol. Liq. 2009, 150, 51–55. [Google Scholar] [CrossRef]
  37. Morosanu, A.C.; Benchea, A.C.; Babusca, D.; Dimitriu, D.G.; Dorohoi, D.O. Quantum mechanical and solvatochromic characterization of Quercetin. Anal. Lett. 2017, 50, 2725–2739. [Google Scholar] [CrossRef]
  38. Babusca, D.; Morosanu, A.C.; Dorohoi, D.O. Spectral study of specific interactions between zwitterionic compounds and protic solvents. Spectrochim. Acta A 2017, 172, 58–64. [Google Scholar] [CrossRef] [PubMed]
  39. Dorohoi, D.O.; Surpateanu, G.; Gheorghies, L.V. A spectral method for determination some molecular parameters of cycloimmonium ylids. Balk. Phys. Lett. 1998, 6, 198–203. [Google Scholar]
  40. Dorohoi, D.O. Excited state molecular parameters determined by spectral means. Ukr. J. Phys. 2018, 63, 701–707. [Google Scholar] [CrossRef] [Green Version]
  41. Pantil, S.K.; Wari, M.N.; Yohannan Panicker, C.; Inamdar, S.R. Solvatochromic study of coumarin 545 in alcohols for the determination of ground and excited state dipole moments. Int. J. Adv. Res. 2013, 1, 616–626. [Google Scholar]
  42. Thipperudrappa, J. Analysis of solvatochromism of a biologically active ketocyanine due using different solvent polarity scales and estimation of dipole moments. Int. J. Life Sci. Pharma Res. 2014, 4, 1–11. [Google Scholar]
  43. Zakerhamidi, M.S.; Golghasemi Sorkhabi, S.H.; Shamkhali, S. Polar and low polar solvents media effect on dipole moments of some diazo-Sudan dyes. Spectrochim. Acta A 2014, 127, 340–348. [Google Scholar] [CrossRef]
  44. Gahlaut, R.; Tewari, N.; Bridhkoti, J.P.; Koshi, H.C.; Pant, S. Determination of ground and excited state dipole moments of some naphthols using solvatochromic shift method. J. Mol. Liq. 2011, 163, 141–146. [Google Scholar] [CrossRef]
  45. Pop, V.; Dorohoi, D.O.; Delibas, M. Considerations on the statistic model of the intermolecular interactions in ternary solutions. An. Stiin. Univ. Al. I. Cuza Iasi 1987, 33, 78–86. [Google Scholar]
  46. Dorohoi, D.O.; Pop, V. Spectral shifts of the electronic absorption spectra of some cycloimmonium ylids in ternary solutions. An. Stiint. Univ. Al. I. Cuza Iasi Ib Fizica 1987, 33, 78–85. [Google Scholar]
  47. Puica-Melniciuc, N.; Avadanei, M.I.; Caprosu, M.; Dorohoi, D.O. Interaction energy in pairs of phthalazinium-dibenzoyl methylid (PDBM)–protic solvent molecules estimated in the limits of ternary solution model. Spectrochim. Acta A 2012, 96, 271–277. [Google Scholar]
  48. Dorohoi, D.O.; Avadanei, M.; Postolache, M. Characterization of the solvation spheres of some dipolar spectrally active molecules in binary solvents. Optoelectron. Adv. Mater. 2008, 2, 51–55. [Google Scholar]
  49. Avadanei, M.; Dorohoi, D.O. Interaction energy in pairs of pyridazinium ylid-solvent molecules estimated by spectral means within the cell ternary solution model. Ukr. J. Phys. 2012, 57, 118–122. [Google Scholar]
  50. Sasirekha, V.; Vanelle, P.; Terme, T.; Ramakrishnan, V. Solvatochromism and preferential solvation of 1, 4-dihydroxy-2, 3-dimethyl-9, 10-anthraquinone by UV-vis absorption and laser-induced fluorescence measurements. Spectrochim. Acta A 2008, 71, 766–772. [Google Scholar] [CrossRef] [PubMed]
  51. Buhvestov, U.; Rived, F.; Ràfols, C.; Bosch, E.; Rosés, M. Solute-solvent and solvent-solvent interactions in binary solvent mixtures. Part 7. Comparison of the enhancement of the water structure in alcohol-water mixtures measured by solvatochromic indicators. J. Phys. Org. Chem. 1998, 11, 185–192. [Google Scholar] [CrossRef]
  52. Rosés, M.; Buhvestov, U.; Ráfols, C.; Rived, F.; Bosch, E. Solute-solvent and solvent-solvent interactions in binary solvent mixtures. Part.6. A quantitative measurement of the enhancement of water structure in 2-methylpropan-2-ol-water and propan-2-ol-water mixtures by solvatochromic indicators. J. Chem. Soc. Perkin Trans. 1997, 2, 1341–1348. [Google Scholar] [CrossRef]
  53. Dulcescu, M.M.; Stan, C.; Dorohoi, D.O. Spectral investigations of the influence of the hydroxyl solvents on the intermolecular interactions in some pyridinium ylid ternary solution. Rev. Chim. Buchar. 2010, 61, 1219–1222. [Google Scholar]
  54. Dulcescu, M.M.; Stan, C.; Dorohoi, D.O. Spectral study of intermolecular interactions in water-ethanol solutions of some carbanion di-substituted pyridinium ylids. Rev. Roum. Chim. 2010, 55, 403–408. [Google Scholar]
Figure 1. Optimized structures of pyridinium di-carbethoxy methylid (PCCM) in (a) vacuum, (b) water, and (c) ethanol.
Figure 1. Optimized structures of pyridinium di-carbethoxy methylid (PCCM) in (a) vacuum, (b) water, and (c) ethanol.
Symmetry 12 01299 g001
Figure 2. Atomic charges near the atoms of PCCM in (a) vaporous state and in solutions with (b) water and (c) ethanol.
Figure 2. Atomic charges near the atoms of PCCM in (a) vaporous state and in solutions with (b) water and (c) ethanol.
Symmetry 12 01299 g002
Figure 3. HOMO and LUMO orbitals of PCCM in (a) vaporous state and in solutions with (b) water and (c) ethanol.
Figure 3. HOMO and LUMO orbitals of PCCM in (a) vaporous state and in solutions with (b) water and (c) ethanol.
Symmetry 12 01299 g003
Figure 4. Electrostatic potential map, local ionization potential map, and |LUMO| map of PCCM in (a) vaporous phase, (b) water, and (c) ethanol.
Figure 4. Electrostatic potential map, local ionization potential map, and |LUMO| map of PCCM in (a) vaporous phase, (b) water, and (c) ethanol.
Symmetry 12 01299 g004
Figure 5. Electronic absorption spectrum of PCCM in ethanol and in ethanol with addition of H2SO4. The UV spectrum was recorded by using 0.2 cm quartz cells, and the visible range by using 5 cm glass cells.
Figure 5. Electronic absorption spectrum of PCCM in ethanol and in ethanol with addition of H2SO4. The UV spectrum was recorded by using 0.2 cm quartz cells, and the visible range by using 5 cm glass cells.
Symmetry 12 01299 g005
Figure 6. Computed wavenumber with Relation (10) and data from Table 2, v calc (cm−1), vs. experimental wavenumber, v exp (cm−1).
Figure 6. Computed wavenumber with Relation (10) and data from Table 2, v calc (cm−1), vs. experimental wavenumber, v exp (cm−1).
Symmetry 12 01299 g006
Figure 7. Computed, v calc (cm−1), (with Relation (12′) and data from Table 2) vs. experimental wavenumber, v exp (cm−1).
Figure 7. Computed, v calc (cm−1), (with Relation (12′) and data from Table 2) vs. experimental wavenumber, v exp (cm−1).
Symmetry 12 01299 g007
Figure 8. Computed wavenumber, v calc (cm−1) (with Relation (13) and data from Table 3), vs. experimental wavenumber, v exp (cm−1) for ternary solution PCCM + water + ethanol.
Figure 8. Computed wavenumber, v calc (cm−1) (with Relation (13) and data from Table 3), vs. experimental wavenumber, v exp (cm−1) for ternary solution PCCM + water + ethanol.
Symmetry 12 01299 g008
Figure 9. Computed wavenumber, v calc (cm−1), (with Relation (14) and data from Table 4), vs. experimental wavenumber, v exp (cm−1) for ternary solution PCCM + water + methanol.
Figure 9. Computed wavenumber, v calc (cm−1), (with Relation (14) and data from Table 4), vs. experimental wavenumber, v exp (cm−1) for ternary solution PCCM + water + methanol.
Symmetry 12 01299 g009
Figure 10. ln p 1 p 2 vs. ln x 1 x 2 for ternary for ternary solution PCCM + water (1) + ethanol (2).
Figure 10. ln p 1 p 2 vs. ln x 1 x 2 for ternary for ternary solution PCCM + water (1) + ethanol (2).
Symmetry 12 01299 g010
Figure 11. ln p 1 p 2 vs. ln x 1 x 2 for ternary solution PCCM + water (1) + methanol (2).
Figure 11. ln p 1 p 2 vs. ln x 1 x 2 for ternary solution PCCM + water (1) + methanol (2).
Symmetry 12 01299 g011
Figure 12. Triangle of the dipole moments for visible ICT transition of PCCM.
Figure 12. Triangle of the dipole moments for visible ICT transition of PCCM.
Symmetry 12 01299 g012
Table 1. Computed (by Spartan’14) parameters of PCCM for isolated molecule, in water and ethanol solutions.
Table 1. Computed (by Spartan’14) parameters of PCCM for isolated molecule, in water and ethanol solutions.
Molecule Properties
Density Functional EDF2, 6–31 G*,
in Vacuum
Density Functional EDF2, 6–31 G*,
in Water
Density Functional EDF2, 6–31 G*,
in Ethanol
FormulaC12H15NO4C12H15NO4C12H15NO4
Weight237.255 amu237.255 amu237.255 amu
Energy−821.401842 au−821.422833 au−821.428045 au
Energy (aq)−821.414893 au--
Solvation E−34.27 kJ/mol--
EHOMO−5.12 eV−5.58 eV−5.47 eV
ELUMO−1.93 eV−1.86 eV−1.83 eV
Dipole Moment2.87 Debye6.28 Debye6.15 Debye
Tautomers000
Conformers363636
QSAR (Quantitative Structure–Activity Relationship)
Area270.44 Å2274.40 Å2274.30 Å2
Volume243.96 Å3245.24 Å3245.15 Å3
PSA35.050 Å237.812 Å237.737 Å2
Ovality1.43 1.451.45
Polarizability60.41 Å360.39 Å360.40 Å3
HBD Count000
HBA Count333
Table 2. Solvent parameters and wavenumber in the maximum of the visible intramolecular charge transfer (ICT) band of PCCM in binary solutions (experimental, Exp.; computed using Relation (1), Calc. (1), and computed using Relation (2), Calc. (2).
Table 2. Solvent parameters and wavenumber in the maximum of the visible intramolecular charge transfer (ICT) band of PCCM in binary solutions (experimental, Exp.; computed using Relation (1), Calc. (1), and computed using Relation (2), Calc. (2).
No.Solventf(ε)f(n) π * βα v ¯ (cm−1)
Exp.Calc. (1′)Calc. (2′′)
1n-Heptane0.2270.240−0.080.000.0022,08021,97522,337
2Dioxane0.2860.2510.550.370.0022,90022,92522,480
3Carbon tetrachloride0.2920.2740.280.100.0021,91022,43022,448
4Benzene0.2990.2950.590.100.0022,55022,75322,454
5p-Xylene0.2990.2910.400.130.0022,52022,60222,459
61,3,5 Trimethylbenzene0.3020.2930.410.130.0022,41022,58922,463
7o-Xylene0.3020.2920.410.110.0022,45022,57422,461
8Toluene0.3480.2970.540.110.0022,72022,70922,535
9Trichloroethylene 0.4480.2820.530.050.0022,50022,65122,695
10Anisole0.5240.3020.730.320.0023,04023,07422,851
111,2-Dibromoethane0.5380.3130.750.000.0022,74022,84122,828
12Chloroform0.5520.2670.690.100.2023,28023,25923,216
13n-Butyl acetate0.5770.2400.460.450.0023,02022,89522,972
14Iso-amyl acetate0.5890.2410.710.070.0023,21022,85522,940
15Chlorobenzene0.6050.3070.710.070.0022,95022,85522,949
16Ethyl acetate0.6250.2280.550.450.0023,30022,98923,054
17Methyl acetate0.6550.2210.600.420.0023,40023,01722,101
18Dichloromethane0.7270.2560.820.100.2023,42023,39522,506
19N-Octyl alcohol0.7450.2580.400.810.7724,56024,66224,597
201,2-Dichloroethane0.7520.2640.810.100.0023,09022,98323,206
21Pyridine0.7900.2990.870.640.0023,10023,47323,332
22N-hexyl alcohol0.8390.2520.040.840.8024,52024,37024,807
23Benzyl alcohol0.8040.3110.980.520.6024,77024,69724,352
24Cyclohexanol0.8240.2760.450.840.6624,60024,51724,340
25N-Amyl alcohol0.8260.2410.400.860.8424,35024,84224,860
26N-Butyl alcohol0.8330.2420.470.840.8424,55024,89924,868
27Acetophenone0.8330.3120.900.490.0423,37023,46623,446
28Iso-Butyl alcohol0.8520.2390.400.840.6924,70024,52524,646
29Iso propyl alcohol0.8520.2340.480.840.7625,00024,74924,646
30N-Propyl alcohol0.8660.2390.520.900.8424,95024,99924,931
31Acetone0.8680.2220.620.480.0823,45023,24623,594
32Benzo nitrile0.8890.3080.900.370.0023,00023,29123,456
33Ethanol0.8950.2210.860.750.8624,97025,27424,997
34Methanol0.9090.2030.600.660.9825,23025,17325,216
35Acetonitrile0.9210.2190.750.400.1923,75025,53923,857
36DMF0.9240.2580.880.690.0023,65023,52323,560
37Ethylene glycol0.9300.2590.900.520.9025,56023,51525,081
38DMSO0.9460.2811.000.760.0023,72023,70423,609
391,2,3-Propanetriol0.9480.2800.620.511.2125,62025,53625,630
40Water0.9640.2061.090.471.1725,43025,91525,600
41Formamide0.9730.2710.970.480.7124,19024,87524,821
Table 3. Binary solvent water (1) + ethanol (2): molar composition (x1) and Kamlet Taft solvent parameters (π*, α, β); wavenumbers, ν (cm−1), in the maximum of the PCCM visible ICT band and p 1 .
Table 3. Binary solvent water (1) + ethanol (2): molar composition (x1) and Kamlet Taft solvent parameters (π*, α, β); wavenumbers, ν (cm−1), in the maximum of the PCCM visible ICT band and p 1 .
No. x 1 π * αΒ ln x 1 x 2 PCCM
ν (cm−1) p 1 ln p 1 p 2
10.0000.510.830.98-24,970--
20.0500.540.830.97−2.9425,0250.119−2.00
30.1000.570.840.96−2.2025,0850.249−1.10
40.1500.600.830.94−1.7425,1260.338−0.67
50.2000.630.830.93−1.3925,1700.435−0.26
60.2500.650.830.93−1.1025,2100.5220.09
70.3000.680.820.92−0.8525,2450.5990.40
80.3500.700.810.91−0.6225,2750.6630.68
90.4000.730.800.91−0.4125,3150.7501.10
100.4500.750.790.89−0.2025,3350.7921.34
110.5000.770.790.900.0025,3600.8481.72
1205500.800.780.890.2025,3700.8701.90
130.6000.820.770.890.4125,3800.8912.10
140.6500.850.770.890.6225,3850.9022.22
150.7000.900.740.880.8525,3900.9132.35
160.7500.940.710.861.1025,4000.9352.67
170.8001.000.670.871.3925,4080.9522.99
180.8251.030.660.871.5525,4100.9573.10
190.8501.060.640.901.7425,4130.9633.26
200.8751.090.610.921.9525,4150.9673.38
210.9001.110.590.972.2025,4200.9783.79
220.9251.120.561.032.5125,4220.9834.06
230.9501.130.541.112.9425,4250.9894.50
240.9751.130.521.183.6625,4280.9965.52
251.0001.130.501.26-25,430--
Table 4. Binary solvent water (1) + methanol (2): solvent molar composition (x1) and Kamlet Taft solvent parameters (π*, α, β); wavenumbers ν (cm−1) in the maximum of the PCCM visible ICT band and p 1 .
Table 4. Binary solvent water (1) + methanol (2): solvent molar composition (x1) and Kamlet Taft solvent parameters (π*, α, β); wavenumbers ν (cm−1) in the maximum of the PCCM visible ICT band and p 1 .
No. x 1 π * αΒ ln x 1 x 2 PCCM
ν (cm−1) p 1 ln p 1 p 2
10.0000.580.741.14-25,190--
20.0500.610.741.13−2.9425,2010.046−3.03
30.1000.640.741.12−2.2025,2250.146−1.77
40.1500.660.741.10−1.7425,2500.249−1.10
50.2000.700.741.09−1.3925,2700.334−0.69
60.2500.730.741.07−1.1025,2850.397−0.42
70.3000.760.741.06−0.8525,3130.5120.05
80.3500.780.721.04−0.6225,3300.5840.34
90.4000.820.721.04−0.4125,3500.6660.69
100.4500.850.711.02−0.2025,3650.7290.99
110.5000.880.701.030.0025,3750.7701.21
1205500.910.681.020.2025,3850.8131.47
130.6000.950.661.010.4125,3900.8331.61
140.6500.980.651.010.6225,4010.8791.98
150.7001.010.631.010.8525,4080.9082.29
160.7501.040.611.021.1025,4150.9382.72
170.8001.060.591.061.3925,4200.9583.13
180.8251.080.581.071.5525,4220.9673.38
190.8501.090.561.091.7425,4240.9753.66
200.8751.100.551.121.9525,4250.9793.84
210.9001.110.541.132.2025,4260.9834.06
220.9251.110.521.172.5125,4270.9884.41
230.9501.120.521.192.9425,4280.9924.82
240.9751.120.511.223.6625,430--
251.0001.140.491.23-25,430--
Table 5. Contribution of each type of intermolecular interactions (in binary solutions) to the spectral shifts of PCCM visible ICT band.
Table 5. Contribution of each type of intermolecular interactions (in binary solutions) to the spectral shifts of PCCM visible ICT band.
No.Solvent C 1 f ( ε )
(cm−1)
p 1
(%)
C 2 f ( n )
(cm−1)
p 2
(%)
C 3 α
(cm−1)
p 3
(%)
C 4 β
(cm−1)
p 4
(%)
1n-Heptane 372.785.662.614.40.00.00.00.0
2Dioxane469.680.365.511.20.00.049.88.5
3Carbon tetrachloride 479.485.071.512.70.00.013.52.4
4Benzene490.984.576.913.20.00.013.52.3
5p-Xylene490.984.075.913.00.00.017.53.0
61,3,5-Trimethylbenzene495.984.176.413.00.00.017.53.0
7o-Xylene495.984.576.113.00.00.014.82.5
8Toluene571.486.177.511.70.00.014.82.2
9Trichloroethylene735.690.273.59.00.00.06.70.8
10Anisole860.487.678.88.00.00.043.14.4
111,2-Dibromoethane883.491.581.68.50.00.00.00.0
12Chloroform906.368.269.65.2339.025.513.51.0
13n-Butyl acetate947.488.562.65.80.00.060.65.7
14Iso-amyl acetate967.193.062.86.00.00.09.40.9
15Chlorobenzene993.491.780.17.40.00.09.40.9
16Ethyl acetate1026.289.559.55.20.00.060.65.3
17Methyl acetate1075.590.457.64.80.00.056.54.8
18Dichloromethane1193.774.066.84.1339.021.013.50.8
19N-Octyl alcohol1223.245.267.32.51305.248.3109.04.0
201,2-Dichloroethane1234.793.868.85.20.00.013.51.0
21Pyridine1297.188.878.05.30.00.086.15.9
22N-hexyl alcohol1377.647.365.72.31356.046.6113.03.9
23Benzyl alcohol1320.153.181.13.31017.040.970.02.8
24Cyclohexanol1353.050.972.02.71118.742.1113.04.3
25N-Amyl alcohol1356.245.862.82.11423.948.1115.73.9
26N-Butyl alcohol1367.746.163.12.11423.948.0113.03.8
27Acetophenone1367.786.481.45.167.84.365.94.2
28Iso-Butyl alcohol1398.951.062.32.31169.642.6113.04.1
29Iso propyl alcohol1398.948.961.02.11288.245.0113.04.0
30N-Propyl alcohol1421.946.962.32.11423.947.0121.14.0
31Acetone1425.284.757.93.4135.68.164.63.8
32Benzo nitrile1459.791.880.35.10.00.049.83.1
33Ethanol1469.547.657.61.91457.847.2100.93.3
34Methanol1492.545.352.91.61661.250.488.82.7
35Acetonitrile1512.277.757.12.9322.116.653.82.8
36DMF1517.190.567.34.00.00.092.95.5
37Ethylene glycol1527.047.967.52.11525.647.870.02.2
38DMSO1553.389.873.34.20.00.0102.35.9
391,2,3-Propanetriol1556.541.573.01.92051.054.768.61.8
40Water1582.843.053.71.51983.253.863.21.7
41Formamide1597.654.470.72.41203.541.064.62.2
Table 6. Contribution (in percents) of each type of intermolecular interaction to the spectral shift in PCCM ternary solutions.
Table 6. Contribution (in percents) of each type of intermolecular interaction to the spectral shift in PCCM ternary solutions.
No.x1Water + EthanolWater + Methanol
Δ ν ¯
(cm−1)
C1π*
%
C2α
%
C3β
%
Δ ν ¯
(cm−1)
C1π*
%
C2α
%
C3β
%
10.000481723.059.517.5141733.950.315.8
20.050487224.058.817.2142835.549.915.6
30.100493225.058.516.5145235.549.215.3
40.150497326.157.616.3147737.048.314.7
50.200501727.057.016.0149738.747.314.0
60.250505728.056.615.4151239.947.213.9
70.300509227.855.615.6154040.846.213.0
80.350512229.755.015.3155741.445.513.1
90.400516230.854.015.2157743.043.913.1
100.450518231.653.614.8159244.242.813.0
110.500520732.352.814.9160245.442.012.6
120550521733.751.614.7161246.740.812.5
130.600522734.550.814.7161748.639.212.2
140.650523235.450.814.8162849.838.212.0
150.700523737.447.814.8163551.037.012.00
160.750524739.246.714.1164252.435.612.0
170.800525541.544.114.4164753.134.412.5
180.825525742.643.214.2164954.033.812.2
190.850526043.842.014.2165154.632.612.8
200.875526245.040.015.0165255.032.013.0
210.900526745.838.715.5165355.531.413.1
220.925526946.436.716.9165455.530.613.9
230.950527246.635.418.0165556.030.014.0
240.975527546.634.019.4165755.929.714.4
251.000527723.059.517.5346.632.720.7
Table 7. Excited state dipole moment and polarizability for different angles, φ.
Table 7. Excited state dipole moment and polarizability for different angles, φ.
No.Φ (Degree)µe (D)αe (cm3)
105.77860.48
255.80160.45
385.83760.40
48.55.84260.39
58.65.84560.388
695.85370.37

Share and Cite

MDPI and ACS Style

Dorohoi, D.O.; Creanga, D.E.; Dimitriu, D.G.; Morosanu, A.C.; Gritco-Todirascu, A.; Mariciuc, G.G.; Melniciuc, N.P.; Ardelean, E.; Cheptea, C. Computational and Spectral Means for Characterizing the Intermolecular Interactions in Solutions and for Estimating Excited State Dipole Moment of Solute. Symmetry 2020, 12, 1299. https://doi.org/10.3390/sym12081299

AMA Style

Dorohoi DO, Creanga DE, Dimitriu DG, Morosanu AC, Gritco-Todirascu A, Mariciuc GG, Melniciuc NP, Ardelean E, Cheptea C. Computational and Spectral Means for Characterizing the Intermolecular Interactions in Solutions and for Estimating Excited State Dipole Moment of Solute. Symmetry. 2020; 12(8):1299. https://doi.org/10.3390/sym12081299

Chicago/Turabian Style

Dorohoi, Dana Ortansa, Dorina Emilia Creanga, Dan Gheorghe Dimitriu, Ana Cezarina Morosanu, Antonina Gritco-Todirascu, Gabriel Grigore Mariciuc, Nicoleta Puica Melniciuc, Elena Ardelean, and Corina Cheptea. 2020. "Computational and Spectral Means for Characterizing the Intermolecular Interactions in Solutions and for Estimating Excited State Dipole Moment of Solute" Symmetry 12, no. 8: 1299. https://doi.org/10.3390/sym12081299

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