Vitória M. R. Vasconcelosa,
Bruna B. Postacchini
b,
Hélcio S. dos Santosc,
Francisco F. M. Cajazeirasc,
Valder N. Freire
d,
Clodomiro Alves Juniorf,
Cláudia Pessoae,
Roner F. da Costa
f,
Igor F. Vasconcelos
a and
Eveline M. Bezerra
*f
aPrograma de Pós-Graduação em Engenharia e Ciência de Materiais, Universidade Federal do Ceará (UFC), CEP 60440-554, Fortaleza, CE, Brazil. E-mail: vitoriavasconcelos@alu.ufc.br; ifvasco@ufc.br
bDepartamento de Física, Universidade Federal de Ouro Preto (UFOP), CEP 35400-000, Ouro Preto, MG, Brazil. E-mail: bruna@ufop.edu.br
cDepartamento de Química, Universidade Estadual Vale do Acaraú (UVA), CEP 62040-370, Sobral, CE, Brazil. E-mail: helciodossantos@gmail.com; ferdinandocajazeiras@gmail.com
dDepartamento de Física, Universidade Federal do Ceará (UFC), CEP 60440-900, Fortaleza, CE, Brazil. E-mail: valder@fisica.ufc.br
ePrograma de Pós-Graduação em Farmacologia, Universidade Federal do Ceará (UFC), CEP 60430-275, Fortaleza, CE, Brazil. E-mail: cpessoa@ufc.br
fPrograma de Pós-Graduação em Ciência e Engenharia de Materiais, Universidade Federal Rural do Semi-Árido (UFERSA), CEP 59625-900, Mossoró, RN, Brazil. E-mail: roner.costa@ufersa.edu.br; clodomiro.jr@ufersa.edu.br; eveline.bezerra@ufersa.edu.br
First published on 24th January 2025
Chalcones demonstrate significant absorption in the near ultraviolet-visible spectrum, making them valuable for applications such as solar cells, light-emitting diodes, and nonlinear optics. This study investigates four dibenzalacetone derivatives (DBAd), DBA, DBC, DEP, and DMA, examining the impact of electron-donating and electron-withdrawing groups and conjugation elongation on their electronic structure in solvents of varying polarities. Using the Polarizable Continuum Model (PCM) and time-dependent density functional theory (TD-DFT), we characterized the excited states of these compounds. Our results reveal a consistent red-shift in the absorption spectrum, with electron-donating groups like ethoxy inducing a more pronounced red-shift than chlorine. Extending conjugation in DMA further shifted the absorption band to lower energy. Solvatochromism influenced the absorption intensities, underscoring the importance of evaluating parameters beyond λmax. Although our methodologies provided a satisfactory correlation between theoretical and experimental data, they also indicate the need for further theoretical models to accurately capture solute–solvent interactions and describe charge-separated states. The results indicated that dibenzalacetone derivatives have potential as alternative materials for development of organic solar cells.
Among these materials, chalcones are generally non-luminescent due to quenching processes caused by intramolecular twisting motions and isomerization of the α, β-unsaturated ketone fragment.9 Howeverm, their derivatives exhibit a wide range of absorption energies in the ultraviolet–visible light spectrum and extensive π-electron delocalization. This results from the reaction of a diketone with two equivalents of an aromatic aldehyde or, alternatively, a dialdehyde reacting with two equivalents of an aromatic ketone.10–14 In this case, the “chalcone” term is used to designate the core molecular fragment of a variety of relevant natural compounds that have two aromatic rings in conjugation with an electrophilic α,β-unsaturated carbonyl system, denoted as 1,3-diaryl-2-propen-1-one.15 In addition, they comprise a class of compounds that are efficiently synthesized with high yields and isomeric selectivity.16
For this reason, chalcones and their derivatives have been extensively studied for their biological properties as anti-inflammatory agents,17 antibacterial agents,18 anticancer agents,19 antituberculosis agents,20 and antileishmanial agents.21 Moreover, chalcones and their associated compounds have gained recognition as π-conjugated organic dyes with significant potential in various photovoltaic applications, such as solar cells, photodetectors, and light-emitting diodes.22 Additionally, they present applications in nonlinear optics, optical limiting, and electrochemical sensing.23,24 Also, they can be employed as precursors in synthesizing advanced polymer materials25 and act as effective corrosion inhibitors for carbon steel.26
However, solvation also plays a significant role in the electronic properties of chalcone derivatives. Specifically, polar solvents can enhance the red-shift effect by interacting with electron donor and acceptor groups, reducing the energy of the excited state.27,28 Due to the above, we were motivated to investigate four dibenzalacetone derivatives (DBAd): (i) DBA [molecular formula C17H14O, (1E,4E)-1,5-diphenylpenta-1,4-dien-3-one, PubChem CID 640180];29 (ii) DBC [molecular formula C17H12Cl2O, (1E,4E)-1,5-bis(4-chlorophenyl)penta-1,4-dien-3-one, PubChem CID 5378584];30 (iii) DEP [molecular formula C21H22O3, (1E,4E)-1,5-bis(4-ethoxyphenyl)penta-1,4-dien-3-one, PubChem CID 668155];31 and (iv) DMA [molecular formula C21H18O, (1E,3E,6E,8E)-1,9-diphenylnono-1,3,6,8-tetraen-5-one, PubChem CID 5378259].32 Their chemical structures are presented in Fig. 1. Although these compounds are known and their crystal structures have been solved, to the best of our knowledge, there has not yet been a complete characterization of the low-energy excited states in solvents with varying polarities.
![]() | ||
Fig. 1 Two-dimensional chemical structures of the four dibenzalacetone derivatives (DBAd): DBA (C17H14O), DBC (C17H12Cl2O), DEP (C21H22O3), and DMA (C21H18O). Carbons and hydrogens are not explicitly shown. Non-carbon and non-hydrogen atoms are colored by atom type (oxygen in red, chloride in green). The figure was drawn using ChemCraft33 program (graphical software for visualization of quantum chemistry computations https://www.chemcraftprog.com). |
In the present study, we address experimentally and computationally the impact of the modification of dibenzalacetone derivatives (DBAd), substituted with electron-donating and electron-withdrawing groups in the aromatic rings and with elongation of the conjugated chain, in a series of solvents with increasing order of dielectric constants. We used the Polarizable Continuum Model (PCM) with the variant of the integral equation formalism of the polarizable continuum model (IEFPCM) for density functional theory (DFT)34–36 and time-dependent density functional theory (TD-DFT)37 to determine the electronic structure of DBAd in their ground and excited states. This research provides insights into the molecular architecture and solvatochromism that drive the functionality of these compounds in advanced technological applications.
Initially, stock solutions with concentrations on the order of 10−3 mol L−1 were prepared for each dibenzalacetone derivative – as detailed in Table S1 of the ESI.† These solutions were homogenized in an ultrasonic bath for 5 minutes. Optical absorption spectra measurements of each DBAd with each of the three solvents were taken for five distinct concentrations in the range from an initial low concentration (LC, 0.5 × 10−5 mol L−1) to high concentration (HC, 3.5 × 10−5 mol L−1), obtained from the stock solution. According to the Beer–Lambert law, the molar absorption coefficient of DBAd was calculated via linear regression analysis of the experimental data in the graph of maximum absorption as a function of concentration. The result is expressed as 1 × 105 L mol−1 cm−1. The y-axis of the graph containing the optical absorption spectra, which are usually acquired in arbitrary units, was corrected to L mol−1 cm−1; see obtained coefficients in Table S2 of the ESI.† Fig. 2 summarizes the experimental and computational protocol.
![]() | ||
Fig. 2 The experimental and computational protocol employed to obtain the UV-vis spectra and electronic properties of the four dibenzalacetone derivatives (DBAd) – DBA, DBC, DEP, and DMA – (presented in Fig. 1) in the toluene (TOL), dichloromethane (DCM), and acetonitrile (ACN) solvents. The experimental protocol (left side) is pink, and the theoretical protocol (right) is purple. Solvents (bottom) are represented in blue, and methods that used solvents are shaded in light blue. |
Geometry optimizations with no constraints were performed using the gradient method, with the final maximum force below 7 × 10−5 Ha Bohr−1 and the RMS force less than 1 × 10−5 Ha Bohr−1, producing geometries accurate to 0.001 Å. The harmonic vibrational frequencies of the ground state were obtained using analytical derivatives within the harmonic approximation in the Gaussian 09 (ref. 45) program at the same theory level (M06-2X/6-311+G(d,p)) used for geometry optimizations. These frequencies (ω) do not include anharmonic corrections due to the lack of scaling factors for this specific combination of functional and basis set in the Computational Chemistry Comparison and Benchmark Database (CCCBDB).46 All optimized structures for each dibenzalacetone derivative (DBAd) were considered absolute minima due to the absence of an imaginary mode in the vibrational analysis calculations. Finally, for the four DBAd as the optimized geometry fundamental state, we use the time-dependent density functional theory (TD-DFT) method based on the Runge–Gross theory and the time-dependent Kohn–Sham formulation47–49 to determine the 50 lowest-energy singlet–singlet vertical electronic transitions using the same functional and base optimization parameters, also with the implicit solvent method (IEFPCM) using the same three solvents in the geometry optimization calculations (TOL, DCM, and ACN) chosen to evaluate the influence of solvatochromism on the electronic properties of the four DBAd. Finally, the theoretical UV-vis absorption spectra were obtained using Gaussian functions centered on the vertical excitation energy (σ = 0.283 eV and the half width at half height (HWHH) = 0.333 eV or 2685.830 cm−1). The band’s intensity is proportional to the strength of the oscillator, which is related to the probability of the transition between states.
The M06-2X functional43 is known for its precision in handling noncovalent interactions, long-range charge transfer, and electronic excitations. The 6-311+G(d,p) basis set was chosen for its ability to accurately model electronic and optical properties, and it complements M06-2X, providing a good correlation between theoretical results, such as optimized geometries and electronic transitions, and experimental observations.
![]() | ||
Fig. 3 Optimized dibenzalacetone derivatives (DBAd) using the DFT method with functional M06-2X and 6-311+G(d,p) basis sets. Superposition image of ball and stick representations of all four DBAd (top), followed by representations of DBA (C17H14O), DBC (C17H12Cl2O), DEP (C21H22O3), and DMA (C21H18O) with atom labels. The carbons of each of the DBAd are colored differently (DBA in pink balls, DBC in purple balls, DEP in cyan balls, and DMA in orange balls), while non-carbon atoms are colored by atom type (oxygen in red, chloride in green, and hydrogen in light grey). The atom labels referring to hydrogen atoms were suppressed. The figure was drawn using PyMOL51 (PyMOL Molecular Graphics System; http://www.pymol.org). |
Furthermore, calculations show that the bond lengths and angles are slightly perturbed due to the electronic effects of the substituents.11 This difference arises because the DFT/PCM calculations considered a single molecule was considered in a continuous dielectric medium, whereas the experimental measurements accounted for closely packed molecules in a condensed phase.53,54 Additionally, the four dibenzalacetone derivatives (DBAd) adopt a quasi-planar s-trans configuration across the enone fragment O1C9–C10
C11, causing lower steric repulsion between the aryl and carbonyl groups and favoring π-extended conjugation throughout the molecule.55,56 The crystallographic parameters for the DBAd have been deposited at the Cambridge Crystallographic Data Centre (CCDC). They can be consulted using the reference number 187856 for DBA,57 608330 for DBC,58 1015509 for DEP,59 and 1563359 for DMA.60
Experimental frequencya (cm−1) | Theoretical frequencyb (cm−1) | |||||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
DBAd | DBA | DBC | DEP | DMA | DBA | DBC | DEP | DMA | ||||||||
Ac/solventd | Solid | Solid | Solid | Solid | TOL | DCM | ACN | TOL | DCM | ACN | TOL | DCM | ACN | TOL | DCM | ACN |
a Frequencies were determined using a second-derivative-based method.b ω harmonic vibrational frequencies calculated using the M06-2X/6-311+G(d,p) DFT method.c Assignment.d TOL, toluene (κTOL = 2.37); DCM, dichloromethane (κDCM = 8.93); ACN, acetonitrile (κACN = 35.67); where κ is the dielectric constant for the solvent.e ν: stretching modes. | ||||||||||||||||
ν(C–H)sp2e | 3375 | 3036 | 3036 | 3028 | 3200 | 3206 | 3206 | 3054 | 3053 | 3056 | 3153 | 3147 | 3146 | 3172 | 3174 | 3190 |
ν(C–H)sp3 | — | — | 2883 | — | — | — | — | — | — | — | 3054 | 3053 | 3059 | — | — | — |
ν(C![]() |
1690 | 1650 | 1640 | 1654 | 1747 | 1720 | 1714 | 1750 | 1729 | 1719 | 1684 | 1685 | 1680 | 1688 | 1692 | 1685 |
ν(C![]() |
1450 | 1587 | 1568 | 1446 | 1500 | 1494 | 1493 | 1542 | 1541 | 1541 | 1561 | 1560 | 1560 | 1538 | 1536 | 1536 |
ν(C–C)e | 1448 | 1323 | 1255 | 1357 | 1317 | 1321 | 1323 | 1319 | 1321 | 1323 | 1298 | 1295 | 1291 | 1312 | 1312 | 1312 |
ν(C–O)e | — | — | 1167 | — | — | — | — | — | — | — | 1198 | 1196 | 1200 | — | — | — |
ν(CCl)e | — | 565 | — | — | — | — | — | 446 | 448 | 447 | — | — | — | — | — | — |
In α,β-unsaturated ketones, the strong coupling between the CC bond and the adjacent carbonyl group causes the delocalization of the π electrons due to the conjugation effect.61 This delocalization causes the C
O and C
C bonds to exhibit characteristics that are intermediate between those of single and double bonds (and decreases the localized electron density at a specific bond), thereby lowering their force constants and resulting in decreased frequency vibrations of the carbonyl and double bonds.62 Therefore, the formation of chalcones can be strongly indicated by the presence of bands with values between 1650 to 1780 cm−1 for C
O, 1550 to 1650 cm−1 for C
C, and 800 to 1200 cm−1 for C–C in the IR spectra.
In the theoretical results, the IR frequency of the CO group agrees with typical values for α,β-unsaturated ketones.63 The values calculated in TOL, DCM, and ACN are, respectively, 1747, 1720, and 1714 cm−1 for DBA, 1750, 1729, and 1719 cm−1 for DBC, 1684, 1685, and 1680 cm−1 for DEP, and 1688, 1692, and 1685 cm−1 for DMA (see Table 1), subtly shifted to lower frequencies due to the increased conjugation effect along the enone fragment (C
C–C
O) and the nature of the substituents on the aromatic rings. That is, electron-donating substituents such as the ethoxy group in the DEP dibenzalacetone derivative and excellent conjugation such as in DMA increase electronic delocalization by the conjugation effect, reducing the force constant of the C
O bond and lowering the vibrational frequency.64
In contrast, electronegative substituents that withdraw electrons through inductive effects, such as chlorine atoms in DBC, exhibit the opposite behavior to DEP and DMA. Although they exert an electron-donating effect through their lone pairs (also known as a positive mesomeric effect), the electron-withdrawing inductive effect predominates, leading to a subtle increase in vibrational frequency due to the increase in electronic coupling. Moreover, polarized solvents stabilize the carbonyl dipole moment, further influencing these frequencies.65 In general, for the four dibenzalacetone derivatives (DBAd), the carbonyl stretching frequencies subtly decrease as the solvent polarity increases (TOL > DCM > ACN), consistent with stabilizing the carbonyl dipole through dipole–dipole interactions – see Table 1.
The CC stretch typically occurs at 1650 cm−1, but conjugation moves it to lower frequencies and increases the intensity.65,66 In addition, the presence of the electron donor and acceptor also changes the vibration of the C
C aliphatic and aromatic groups. In the calculated results for DEP and DBC, the presence of substituents in the para-position of the aromatic rings, in addition to shifting the IR to frequencies of approximately 1540 and 1560 cm−1, respectively, increased the absorption intensity of C
C compared to DMA and DBA due to conjugation and inductive effects that modify the force constants, despite the inertial mass of the ethoxy group and chlorine atoms. Furthermore, in the ground state, the substitutions in DEP and DBC appear to remove the electron density of the aromatic rings instead of interacting by resonance,66,67 which could justify the frequency values being slightly higher than 1495 cm−1 for DBA and 1536 cm−1 for para-DMA.
In general, the C–H stretching peaks for the sp2 carbon appear at values greater than 3100 cm−1 for the four dibenzalacetone derivatives (DBAd) in the results calculated by DFT/PCM. However, since the C–H stretching for both aliphatic and aromatic chain alkenes appears in the same range,66 it is not easy to use the C–H bands to differentiate between both types of contribution. For simplicity, we do not distinguish the contributions of the aliphatic and aromatic chains to the C–H vibrations. In the DEP exception, the calculated C–H stretching peaks for the sp2 carbon in TOL, DCM, and ACN occur at 3153, 3147, and 3146 cm−1, respectively, and the C–H peaks for the sp3 carbon atoms appear below these values (at 3054, 3054, and 3059 cm−1, respectively). C–H out-of-plane bending bands occur in the range of 650 to 1000 cm−1, but these will not be assigned here.
The vibrations for C–C, C–O and C–Cl occur in the range of 600 to 1550 cm−1,68 so that the greater the inertial mass, the lower the vibrational frequency. In the theoretical results, the observed vibrational frequencies for DBA demonstrate vibrations around 800 to 1200 cm−1, corresponding to C–C stretching modes and C–H out-of-plane bending. For DBC, in addition to the C–C stretching modes, vibrations around 440 to 780 cm−1 indicate the presence of C–Cl stretching. For DEP, the presence of C–O stretching is evident in the range of 1000 to 1300 cm−1 along with C–C stretching modes, and the results for DMA are similar to those for DBA.
The experimental IR spectra of dibenzalacetone derivatives show similarities (see Fig. 4). However, we initially noticed that the DBA derivative exhibited an extensive band around 3400 cm−1 associated with forming a hydrated precipitate or adsorbed water.11 A strong CO stretching mode appeared at 1690, 1650, 1640, and 1654 cm−1 for DBA, DBC, DEP, and DMA, respectively, associated with enone (O1
C9–C10
C11) stretching. The C
C stretching modes of the aromatic rings and the aliphatic chain appeared at 1604 cm−1 for DBA, 1591 cm−1 for DBC, 1600 cm−1 for DEP, and 1600 cm−1 for DMA. Below 1500 cm−1, the modes were mainly attributed to C–C stretching, C–C
C and O–C
C bending, and out-of-plane torsions.11,66 The region between 3000 and 3200 cm−1 was attributed to stretching of the C–H bond of the sp2 carbon of the aromatic rings and aliphatic chain. Furthermore, for DEP, we attribute the mode at 2880 cm−1 to the stretching of C–H sp3 due to ethoxy groups.
The DFT/PCM-calculated vibrational frequencies (unscaled) were systematically overestimated relative to the experimental values. The reason for this disagreement between calculated and observed vibrational wavenumbers is that the calculations were made for a free-solvated molecule,69 while experiments were performed for a crystalline conformation of the solid sample. The reason is also partly due to the anharmonicity and the approximate nature of the quantum mechanical methods.70
Initially, the optical absorption spectra were collected in arbitrary units by default by the equipment. However, to standardize the interpretation with the theoretical results, the absorbance in arbitrary units was converted to molar absorption coefficient values. We used the linear regression technique on the absorbance values at the maximum wavelength (λmax) for different concentrations of each DBAd in TOL, DCM, and ACN, according to the Beer–Lambert law.73,74 The regression plots are shown in Fig. 6(a) for DBA, Fig. 6(c) for DBC, Fig. 6(e) for DEP, and Fig. 6(g) for DMA. The TD-DFT calculations were performed as function of the oscillator strength, vertical excitation energies (E in eV), and probability of an electronic transition between orbital types (%); see Fig. 6(b) for DBA, Fig. 6(d) for DBC, Fig. 6(f) for DEP, and Fig. 6(h) for DMA. Table 2 shows the first six main electronic transitions of the DBAd in the employed solvents, calculated using the TD-DFT/PCM/M06-2X/6-311+G(d,p) methodology.
![]() | ||
Fig. 6 Determination of molar absorption coefficients for DBA (a), DBC (c), DEP (e) and DMA (g) in toluene (κTOL = 2.37), dichloromethane (κDMC = 8.93) and acetonitrile (κACN = 35.67) by linear regression analysis of the experimental data of maximum absorption as a function of molar concentration, according to the Beer–Lambert law. The results on the y-axis, initially described in arbitrary units, were corrected to mole per L per cm−1 to standardize the interpretation with the theoretical results and can be found in Table S2 in the ESI.† Theoretical oscillator strengths as a function of photon energy (in eV) for DBA (b), DBC (d), DEP (f) and DMA (h) in TOL, DCM, and ACN. |
DBAd | ES | Toluene (κTOLa = 2.37) | Dichloromethane (κDCM = 8.93) | Acetonitrile (κACN = 35.67) | |||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
E (eV) | O Strc | Electronic transition | (%)d | E (eV) | O Strc | Electronic transition | (%)d | E (eV) | O Strc | Electronic transition | (%)d | ||
a κ is the dielectric constant for the solvent.b H = HOMO, L = LUMO.c O Str, oscillator strength.d Probability. | |||||||||||||
DBA | 1 | 3.61 | 0.006 | H−4 → Lb | 76.3 | 3.68 | 0.007 | H−4 → L | 77.1 | 3.71 | 0.007 | H−4 → L | 77.3 |
2 | 4.27 | 0.839 | H → L | 87.5 | 4.22 | 0.824 | H → L | 88.8 | 4.21 | 0.810 | H → L | 89.3 | |
3 | 4.59 | 0.490 | H−1 → L | 87.2 | 4.58 | 0.474 | H−1 → L | 88.1 | 4.58 | 0.454 | H−1 → L | 88.6 | |
4 | 4.98 | 0.014 | H−3 → L | 27.5 | 4.96 | 0.019 | H−2 → L | 32.9 | 4.95 | 0.019 | H−3 → L | 26.8 | |
5 | 4.98 | 0.011 | H−2 → L | 27.4 | 4.96 | 0.012 | H−3 → L | 33.1 | 4.95 | 0.013 | H−2 → L | 26.6 | |
6 | 5.30 | 0.077 | H−1 → L+1 | 79.3 | 5.30 | 0.084 | H−1 → L+1 | 81.4 | 5.31 | 0.086 | H−1 → L+1 | 81.5 | |
DBC | 1 | 3.60 | 0.007 | H−4 → L | 74.0 | 3.67 | 0.008 | H−4 → L | 74.7 | 3.69 | 0.008 | H−4 → L | 74.8 |
2 | 4.18 | 1.014 | H → L | 85.0 | 4.14 | 0.992 | H → L | 86.4 | 4.13 | 0.975 | H → L | 87.0 | |
3 | 4.51 | 0.549 | H−1 → L | 83.8 | 4.50 | 0.534 | H−1 → L | 85.1 | 4.51 | 0.514 | H−1 → L | 85.8 | |
4 | 4.96 | 0.000 | H−2 → L | 19.4 | 4.95 | 0.002 | H−2 → L | 26.3 | 4.95 | 0.003 | H−2 → L | 27.5 | |
5 | 4.96 | 0.001 | H → L+3 | 21.1 | 4.96 | 0.002 | H−3 → L | 26.8 | 4.96 | 0.002 | H−3 → L | 28.1 | |
6 | 5.21 | 0.100 | H−1 → L+1 | 70.4 | 5.23 | 0.110 | H−1 → L+1 | 72.5 | 5.23 | 0.035 | H → L+1 | 71.1 | |
DEP | 1 | 3.67 | 0.011 | H−4 → L | 72.7 | 3.74 | 0.016 | H−4 → L | 73.5 | 3.76 | 0.011 | H−4 → L | 74.0 |
2 | 3.97 | 1.065 | H → L | 85.8 | 3.90 | 1.053 | H → L | 87.2 | 3.91 | 1.048 | H → L | 87.2 | |
3 | 4.32 | 0.451 | H−1 → L | 82.7 | 4.30 | 0.449 | H−1 → L | 82.0 | 4.28 | 0.424 | H−1 → L | 85.5 | |
4 | 4.85 | 0.013 | H → L+1 | 23.7 | 4.83 | 0.005 | H → L+1 | 35.5 | 4.85 | 0.002 | H → L+1 | 30.7 | |
5 | 4.86 | 0.015 | H → L+3 | 24.6 | 4.85 | 0.018 | H → L+2 | 27.7 | 4.86 | 0.017 | H → L+2 | 27.2 | |
6 | 5.12 | 0.073 | H−1 → L+1 | 30.1 | 5.09 | 0.038 | H → L+1 | 46.3 | 5.11 | 0.055 | H → L+1 | 41.2 | |
DMA | 1 | 3.53 | 0.002 | H−4 → L | 63.5 | 3.61 | 0.003 | H−4 → L | 57.4 | 3.64 | 0.034 | H−4 → L | 54.0 |
2 | 3.72 | 1.397 | H−1 → L | 77.6 | 3.68 | 1.376 | H → L | 79.6 | 3.65 | 1.336 | H → L | 75.7 | |
3 | 3.98 | 0.418 | H → L | 81.1 | 3.96 | 0.413 | H−1 → L | 82.4 | 3.95 | 0.402 | H−1 → L | 79.4 | |
4 | 4.76 | 0.078 | H → L+1 | 70.7 | 4.74 | 0.049 | H → L+1 | 48.1 | 4.71 | 0.032 | H → L+1 | 54.4 | |
5 | 4.77 | 0.014 | H−1 → L+1 | 71.7 | 4.75 | 0.063 | H−1 → L+1 | 47.5 | 4.72 | 0.088 | H−1 → L+1 | 53.4 | |
6 | 4.93 | 0.011 | H−3 → L | 23.6 | 4.92 | 0.011 | H−3 → L | 24.5 | 4.90 | 0.009 | H−3 → L | 40.1 |
According to the Franck–Condon principle, the maximum absorption peaks in a UV-vis spectrum correspond to a vertical excitation from the ground (S0) to excited (Sn) electronic states.75,76 In the theoretical spectra, three absorption bands located at 170 to 187 nm, 192 to 214 nm, and 284 to 336 nm were observed for the four dibenzalacetone derivatives (DBAd) in the employed solvents, relating to π–π* and n–π* electronic transitions originating from the aromatic rings and α,β-unsaturated ketone fragment. In DMA, an additional absorption band with weak intensity was observed at about 260, 261, and 263 nm (approximately 4.77, 4.75, and 4.71 eV) for TOL, DCM, and ACN, respectively, due to increased oscillator strength in this region. This band corresponds to the HOMO−1 to LUMO+1 and HOMO to LUMO+1 electronic transitions; see Table 2.
However, due to the inherent limitations of experimental measurements, constrained by the equipment setup, source limitations, optical components, atmospheric absorption, quartz cuvette absorption, and solvent absorption, only the lowest energy band was reported in the DBAd experimental UV-vis spectra with the cutoff point being above 285 nm for toluene, 235 nm for dichloromethane, and 190 nm for acetonitrile; see Fig. 5. In some cases, it is possible to observe two experimental bands. The main bands characterized by absorption at a maximum wavelength (λexpmax) of 327.0, 326.5, and 322.5 nm for DBA, 331.0, 332.0, and 326.0 nm for DBC, 361.5, 364.0, and 359.0 nm for DEP, and 369.0, 370.5, and 369.0 nm for DMA in the solvents TOL, DCM, and ACN, respectively, are in good approximation to the calculated λthemax values. These values are shown in Table 3.
DBAd | Solvent | Experimentala | Theoreticalb | Error% | |||||
---|---|---|---|---|---|---|---|---|---|
λexpmax (nm) | εexpmax (105 mol L−1 cm−1) | Eexpoptg (eV) | λtheomax (nm) | εtheomax (105 mol L−1 cm−1) | Etheopth (eV) | O Strf | |||
a Experimental peak maxima and molar absorption coefficients determined using a second-derivative-based method.b Theoretical values calculated using M06-2X/6-311+G(d,p).c TOL, toluene (κTOL = 2.37).d DCM, dichloromethane (κDCM = 8.93).e ACN, acetonitrile (κACN = 35.67); where κ is the dielectric constant for the solvent.f O Str, oscillator strength.g Experimental optical gap (Eexpopt) determined using the Tauc plot method,77 which involves plotting (absorption coefficient × photon energy)n in (eV cm−1)n against photon energy (in eV) to extrapolate the linear portion to the energy axis. Adopted n = 2 considering indirect allowed transitions. Data expressed with one decimal place due to the resolution of the spectrophotometer.h Theoretical optical gap (Etheopt) obtained from the second excited state (S2) and selected due to its higher oscillator strength and HOMO → LUMO transition probability, which indicates allowed electronic transitions;78 ↓ increases the value in this direction; ↑ decreases the value in this direction. | |||||||||
DBA | TOLc | 327.0↑ | 0.471↓ | 3.50↑ | 290.69↓ | 0.463↑ | 4.27↑ | 0.839↑ | 12 |
DCMd | 326.5↑ | 0.659↓ | 3.49↑ | 294.07↓ | 0.441↑ | 4.22↑ | 0.824↑ | 17 | |
ACNe | 322.5↑ | 0.744↓ | 3.54↓ | 294.70↓ | 0.425↑ | 4.21↑ | 0.810↑ | 18 | |
DBC | TOLc | 331.0↓ | 0.324↓ | 3.42↑ | 296.89↓ | 0.544↑ | 4.18↑ | 1.014↑ | 12 |
DCMd | 332.0↓ | 0.327↓ | 3.42↑ | 299.33↓ | 0.520↑ | 4.14↑ | 0.992↑ | 11 | |
ACNe | 326.0↑ | 0.444↓ | 3.48↓ | 299.68↓ | 0.501↑ | 4.13↑ | 0.975↑ | 9 | |
DEP | TOLc | 361.5↓ | 0.405↓ | 3.15↑ | 311.98↓ | 0.535↑ | 3.97↑ | 1.065↑ | 16 |
DMCd | 364.0↓ | 0.483↓ | 3.10↑ | 318.19↓ | 0.511↑ | 3.90↑ | 1.053↑ | 14 | |
ACNe | 359.0↑ | 0.596↓ | 3.15↓ | 317.47↑ | 0.512↓ | 3.91↓ | 1.048↑ | 13 | |
DMA | TOLc | 369.0↓ | 0.409↓ | 3.07↑ | 332.99↓ | 0.689↑ | 3.72↑ | 1.397↑ | 11 |
DMCd | 370.5↓ | 0.463↓ | 3.05↑ | 336.90↓ | 0.671↑ | 3.68↑ | 1.376↑ | 10 | |
ACNe | 369.0↑ | 0.675↓ | 3.07↓ | 339.67↓ | 0.656↑ | 3.65↑ | 1.336↑ | 8 |
In good agreement with λexpmax, the TD-DFT calculations predict the occurrence of a higher probability electronic transition (allowed) at 290.69, 294.07, and 294.70 nm (4.26, 4.22, and 4.21 eV) for DBA, 296.89, 299.33, and 299.68 nm (4.18, 4.14, and 4.13 eV) for DBC, 311.98, 318.19, and 317.47 nm (3.97, 3.90, and 3.90 eV) for DEP, and 322.99, 336.90, and 339.67 nm (3.72, 3.68, and 3.65 eV) for DMA in TOL, DCM, and ACN, respectively, due to the highest calculated value of the oscillator strength in this vertical transition. These absorption maxima correspond mainly to the electronic excitation from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO); this transition between frontier orbitals corresponds to the energy of the second excited state, as seen in Table 2. Fig. 7–10 illustrate the symmetry of the molecular orbitals involved in the main electronic transitions of the first three lowest-energy excited states that compose the band centered at λthemax.
For qualitative analysis of the experimental (Eexpopt) and theoretical (Etheopt) optical gap values, we used the Tauc method77 to determine Eexpopt and the energy of the second excited state (S2) to quantify Etheopt, because the optical gap corresponds to the energy of the lowest electronic transition that is accessed through the absorption of a photon.78 Here, we consider that the first excited state (S1) for DBAd may be optically forbidden due to the symmetry of the orbitals, justifying the low values of oscillator strength; see excited state (S1) for DBAd in Fig. 7–10. In general, Eexpopt decreases in the order DBA > DBC > DEP > DMA and from TOL to DCM, while acetonitrile increases the optical gap energy. Similarly, Etheopt decreases from DBA to DMA (with DEP in ACN as an exception) and decreases with solvent polarity. These occur because in α,β-unsaturated ketones, the π–π* and n–π* electronic transitions are shifted to longer wavelengths due to the conjugation effect causing increased proximity of the electronic levels of the chromophores that form molecular orbitals with lower electronic excitation energy.66,79
In addition, auxochromes with unpaired electrons also cause shifts in the absorption bands to longer wavelengths since the unbonded electrons enhance the delocalization due to the conjugation effect with the π system.80 So, the more n electrons interacting with unsaturated bonds, the greater the bathochromic shifts.81 Therefore, the presence of a para-position substitution group at the aromatic rings justifies shifting the π–π* and n–π* transitions in compounds DBC and DEP to a lower energy region (a longer wavelength) relative to non-substituted DBA. Despite the electron-withdrawing inductive effect of chlorine atoms – chlorine atoms withdraw electrons by induction but can donate one electron by conjugation when in aromatic rings82 – when substituted in aromatic rings, they weaken the bond order of CC and C
O due to increased electronic delocalization.
In DEP, in addition to the conjugation effect by unpaired electron-donation from the oxygen at the ethoxy group, the overlap of the C–H bonding orbitals of the alkyl group with the π system results in an extension of the conjugation. This type of interaction is often called hyperconjugation;83 see Fig. 9, which shows the electron density arising from the alkyl group. In DMA, the increase in the α,δ-unsaturated chain promoted the largest red-shift due to the increased interaction between the π orbitals that promoted the increase in electronic delocalization. Therefore, the red-shift of λexpmax and λthemax in DMA > DEP > DBC > DBA is justified by the donor–acceptor–donor structural modification in DBAd. Fig. 11 shows the experimental and calculated red-shift by the TD-DFT/PCM/M06-2X/6-311+G(d,p) methodology.
The excitation of n electrons in the π-extended chromophore renders the excited atom partially electron-deficient, while the π system acquires an electron in the π* antibonding orbital.84 This state causes charge separation in the molecule, and is called the charge-transfer excited state,85 which solvents also influence. In general, a very low red-shift observed in the λthemax of the DBAd (with the exception of DEP) upon going from TOL to DCM and DCM to ACN solvents suggests a low increase in the molecular dipole moment in the excited state with solvent polarity. However, the acetonitrile solvent practically promoted a hypsochromic shift in λexpmax by obliterating the absorption peak due to the dipole–dipole interaction with the solute molecules (or due to the effect not being captured by the theoretical methodology); see Fig. 5 and Table 3 for comparison of the variation in λexpmax and λthemax values.
In summary, DBC induced a red-shift in the optical absorption spectra, with λexpmax increasing by approximately 4.00, 5.50, and 3.50 nm compared to DBA in the solvents toluene (TOL), dichloromethane (DCM), and acetonitrile (ACN), respectively. In contrast, DEP and DMA induced the largest shifts of 34.50, 37.50, and 36.50 nm for DEP, and 42.00, 44.00, and 46.50 nm for DMA compared to DBA. In the theoretical results, the red-shift of λthemax in the absorption spectra corresponded to 6.20, 5.26, and 4.98 nm for DBC compared to DBA in the solvents TOL, DCM, and ACN, respectively. For DEP, the shifts were 21.29, 24.12, and 22.77 nm compared to DBA, while for DMA they were 42.30, 42.83, and 44.97 nm compared to DBA; see comparative values of λexpmax and λtheomax in Table 3. The similar trend observed in the theoretical and experimental results corroborates the accuracy and sensitivity of the study.
Our results demonstrate a satisfactory match between λexpmax and λtheomax for the dibenzalacetone derivatives (DBAd), with a theoretical–experimental difference that is between 8% and 18% (see Table 3). DMA in acetonitrile (ACN) showed the highest accuracy with an 8% discrepancy, indicating good predictive reliability of the theoretical models employed. In contrast, DBA in ACN presented the most significant discrepancy at 18%. These findings underscore the impact of molecular structure and solvatochromism on the absorption characteristics of these compounds. Although the methodologies employed produced a satisfactory correlation, they also demonstrate the need to evaluate other theoretical models.
According to Koopmans’ theorem,86 we calculated the global chemical reactivity descriptors from the theoretical results obtained using the DFT/TD-DFT/PCM/M06-2X/6-311+G(d,p) methodology. The results in Table S5 of the ESI† show that the exciton binding energy decreases with increasing π-conjugation and solvent polarity. The trend observed for the derivatives follows the order: DMA < DEP < DBC < DBA, with EB values in toluene (TOL), dichloromethane (DCM), and acetonitrile (ACN) of 2.29, 2.15, and 2.07 eV for DMA, 2.41, 2.24, and 2.21 eV for DEP, 2.71, 2.60, and 2.56 eV for DBC, and 2.86, 2.73, and 2.69 eV for DBA.
Reducing the exciton binding energy by modifying the donor–acceptor–donor structure in dibenzalacetone derivatives (DBAd) tends to improve the exciton dissociation efficiency in charge-transfer significantly states,87 especially in interactions with electron-acceptor materials such as fullerene and its derivatives. This results in an effective minimization of energy losses in high-efficiency organic devices. Therefore, the reported results suggest that DBAd have promising potential for both biological applications and optoelectronic devices, significantly expanding their field of utilization.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra07256a |
This journal is © The Royal Society of Chemistry 2025 |