Next Article in Journal
Application of Pharmacokinetic Prediction Platforms in the Design of Optimized Anti-Cancer Drugs
Next Article in Special Issue
Discovery of Novel Coumarin Derivatives as Potential Dual Inhibitors against α-Glucosidase and α-Amylase for the Management of Post-Prandial Hyperglycemia via Molecular Modelling Approaches
Previous Article in Journal
Fluorescent Immunoassay with a Copper Polymer as the Signal Label for Catalytic Oxidation of O-Phenylenediamine
Previous Article in Special Issue
Effects of Coumarinyl Schiff Bases against Phytopathogenic Fungi, the Soil-Beneficial Bacteria and Entomopathogenic Nematodes: Deeper Insight into the Mechanism of Action
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Optical Characterization in Solution and Solid-State, and DFT Calculations of 3-Acetyl and 3-(1′-(2′-Phenylhydrazono)ethyl)-coumarin-(7)-substituted Derivatives

by
Cesar A. Villa-Martínez
1,
Nancy E. Magaña-Vergara
2,*,
Mario Rodríguez
3,
Juan P. Mojica-Sánchez
4,
Ángel A. Ramos-Organillo
1,
Joaquín Barroso-Flores
5,6,
Itzia I. Padilla-Martínez
7 and
Francisco J. Martínez-Martínez
1,*
1
Facultad de Ciencias Químicas, Universidad de Colima, Km 9 Carretera Coquimatlán-Colima, Coquimatlán 28400, Mexico
2
CONACyT, Facultad de Ciencias Químicas, Universidad de Colima, Km 9 Carretera Coquimatlán-Colima, Coquimatlán 28400, Mexico
3
Centro de Investigaciones en Óptica A. P. 1-948, León 37000, Mexico
4
Tecnológico Nacional de Mexico, Instituto Tecnológico José Mario Molina Pasquel y Henríquez Campus Tamazula de Gordiano, Carretera Tamazula-Santa Rosa No. 329, Tamazula de Gordiano 49650, Mexico
5
Centro Conjunto de Investigación en Química Sustentable UAEM-UNAM, Unidad San Cayetano, Carretera Toluca-Atlacomulco Km.14.5, Toluca de Lerdo 50200, Mexico
6
Instituto de Química, Universidad Nacional Autónoma de Mexico, Circuito Exterior, Ciudad Universitaria, Ciudad de Mexico 04510, Mexico
7
Laboratorio de Química Supramolecular y Nanociencias, UPIBI, Instituto Politécnico Nacional, Av. Acueducto s/n Barrio la Laguna Ticomán, Ciudad de Mexico 07340, Mexico
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(12), 3677; https://doi.org/10.3390/molecules27123677
Submission received: 13 April 2022 / Revised: 1 June 2022 / Accepted: 2 June 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Coumarin and Its Derivatives II)

Abstract

:
Intramolecular charge transfer (ICT) effects are responsible for the photoluminescent properties of coumarins. Hence, optical properties with different applications can be obtained by ICT modulation. Herein, four 3-acetyl-2H-chromen-2-ones (1ad) and their corresponding fluorescent hybrids 3- (phenylhydrazone)-chromen-2-ones (2ad) were synthesized in 74–65% yields. The UV-Vis data were in the 295–428 nm range. The emission depends on the substituent in position C-7 bearing electron-donating groups. Compounds 1bd showed good optical properties due to the D-π-A structural arrangement. In compounds 2ad, there is a quenching effect of fluorescence in solution. However, in the solid, an increase is shown due to an aggregation-induced emission (AIE) effect given by the rotational restraints and stacking in the crystal. Computational calculations of the HOMO-LUMO orbitals indicate high absorbance and emission values of the molecules, and gap values represent the bathochromic effect and the electronic efficiency of the compounds. Compounds 1ad and 2ad are good candidates for optical applications, such as OLEDs, organic solar cells, or fluorescence markers.

1. Introduction

Intramolecular charge transfer (ICT), as well as linear and nonlinear optical properties [1,2,3,4,5], are associated with organic molecules having an extended electronic π-backbone. This molecular characteristic is generally modulated by the insertion of electron-donor or -acceptor groups into the electronic π-system, to favor the ICT [6]. The typical electronic architectures that can be presented, are donor-acceptor-donor (D-A-D), donor-π-acceptor (D-π-A), or donor-π-acceptor-π-donor (D-π-A-π-D), known as push-pull type compounds. Photoluminescent compounds and materials with fluorescent and phosphorescent properties play a fundamental role in many areas, such as biology, pharmacology, chemistry, biomedicine, and optoelectronics [7,8,9,10,11,12]. For example, coumarin compounds coupled to phenylhydrazine have been reported as fluorescent probes for visualizing living tissues [13]. Photoluminescent organic compounds derived from coumarins have been applied in two-photon absorption phenomena, optical data storage, energy limitation, and the manufacture of lasers [5]. Coumarins can be obtained in good yields [9] through synthetic techniques, such as Knoevenagel, Pechmann, and Wittig reactions. Therefore, the photophysical properties of coumarins can be easily modified by substituting in positions 3, 6, 7, and 8 to obtain hybrid materials [9,14,15], which exhibit high quantum yields, important changes in Stokes shifts, and good absorption and emission profiles [16]. On the other hand, fluorescent hybrid compounds derived from hydrazines and coumarins have shown synergic effects in their fluorescent properties [17,18]. These compounds have been applied as on-off sensors [19,20], photoluminescent dyes [21], anticonvulsants [22], hydrogels [23], and antimicrobials [24].
In this work, we describe the synthesis, structural characterization, and photoluminescence proprieties of small compounds derived from 3-acetyl-2H-chromene-2-one (1ad) coupled with phenylhydrazine, forming the hybrid fluorescent compounds 3-(1′-(2′-phenylhydrazono) ethyl)-2H-2-chromen-2-one (2ad). The optical characterization of 2ad was performed in solution and in the solid-state, and the results contrasted with ab initio theoretical calculations. The UV-vis spectra of coumarins 1a,c and 2b have been reported in solution but not in the solid-state [25,26].

2. Experimental

2.1. Materials and Methods

Salicilaldehyde, 4-methoxy-salicylaldehyde, 4-hydroxy-salicylaldehyde, 4-diethylamino-salicylaldehyde, ethyl acetoacetate, piperidine, glacial acetic acid, ethanol, and phenylhydrazine were purchased from Aldrich (Toluca, Mexico) and used as received. The ACS grade solvents were purchased from CTR (Jalisco, Mexico) and used without further purification. Melting points were measured on an Electrothermal Mel-Temp 1201D apparatus. IR spectra were collected using Varian 3100 FT-IR EXCALIBUR series spectrophotometer. 1H and 13C NMR spectra were recorded in a BRUKER Ultrashield plus 400 MHz in DMSO-d6 solutions using (CH3)4Si as an internal reference compound; chemical shifts (δ) are in ppm and coupling constants (nJH-H) in Hz. Mass analysis was performed by electro spray ionization in a high-resolution mass spectrometer Bruker micOTQF- QTOF instrument (Bruker Daltonik GmbH, Bremen, Germany). Absorbances were obtained with a Perkin Elmer Lambda 900 UV/Vis/IR spectrophotometer; emissions were obtained in an Edinburan F55 fluorometer, employing the stock solutions of 1 × 10−3 M of compounds 1ad and 2ad. The fluorescence spectra were recorded under 400 nm excitation (except in the compound 2d was in 372 nm). The quantum efficiencies were obtained using an integration sphere of direct photon counting, relating the absorbed and emitted photons. The optical data in the solid state were obtained similarly.

2.2. General Methods of Synthesis and Characterization

3-Acetyl-2H-chromene-2-one (1ad) were prepared following reported procedures [15,27] with modifications. 4-Substituted-salicylaldehydes, ethyl acetoacetate, and piperidine were dissolved in 15 mL of ethanol and placed under reflux at 78 °C for 24 h. After cooling, the solvent was removed by vacuum filtration, and the resulting solid was recrystallized from cold ethanol.
3-Acetyl-2-chromen-2-one (1a) [25]. Obtained from salicilaldehyde (0.42 mL, 4.1 mmol), ethyl acetoacetate (0.51 mL, 4.1 mmol), and piperidine (17.4 μL, 3 drops). White solid, yield = 0.62 g (80%), mp = 94–96 °C. λ Abs (nm) = 297 (THF), 296 (acetonitrile). IR (cm−1): ν (C-H) 2978, ν (C-H aromatic) 3047, ν (C=C) 1674–1450, ν (C=O) 1681, ρ (C-H f.p) 771. 1H-NMR (DMSO-d6), δ (ppm): 8.67 (1H, s, H-4), 7.95 (1H, d, 3J 8.3, H-5), 7.74 (1H, dd, 3J 7.7, 7.3, H-7), 7.48 (1H, d, 3J 7.7, H-8), 7.44 (1H, d, 3J 8.3, 7.3, H-6), 2.51 (3H, s, H-12). 13C-NMR (DMSO-d6), δ (ppm): 195.5 (C-11), 158.8 (C-2), 155.0 (C-9), 147.4 (C-4), 134.9 (C-7), 131.2 (C-5), 125.4 (C-6), 124.9 (C-3), 118.6 (C-10), 116.5 (C-8), 30.4 (C-12).
3-Acetyl-7-diethylamino-2-chromen-2-one (1b) [28,29]. Obtained from 4-diethylamino-salicylaldehyde (0.500 g, 2.6 mmol), ethyl acetoacetate (0.32 mL, 2.6 mmol) and piperidine (17.4 μL, 3 drops). Yellow solid, yield = 0.37 g (59%), mp = 156–158 °C. λ Abs (nm) = 426 (THF), 429 (acetonitrile), 424 (toluene), λ Em (nm) = 466 (THF), 477 (acetonitrile), 461 (toluene). IR (cm−1): ν (C-H) 2924, ν (C-H aromatic) 2962, ν (C=C) 1608–1411, ν (C=O) 1658, ν (C-N) 1188, ρ (C-H f.p) 759. 1H-NMR (DMSO-d6), δ (ppm): 8.4 (1H, s, H-4), 7.65 (1H, d, 3J 9.0, H-5), 6.78 (1H, dd, 3J 9.0, 4J 2.4, H6), 6.56 (1H, d, 4J 2.3, H-8), 3.9 (3H, s, H-12), 3.49 (4H, q, 3J 7.0, N (CH2CH3)2.), 1.14 (6H, t, 3J 7.0, N (CH2CH3)2.). 13C-NMR (DMSO-d6), δ (ppm): 194.6 (C-11), 160.3 (C-2), 158.7 (C-9), 153.4 (C-7), 148.0 (C-4), 132.8 (C-5), 115.5 (C-3), 110.6 (C-6), 107.9 (C-10), 96.3 (C-8), 44.8 (N (CH2CH3)2.), 12.8 N (CH2CH3)2.
3-Acetyl-7-hydroxy-2-chromen-2-one (1c) [30]. Obtained from 4-hydroxy-salicylaldehyde (0.500 g, 3.62 mmol), ethyl acetoacetate (0.42 mL, 3.6 mmol) and piperidine (17.4 μL, 3 drops). Green solid, yield = 0.68 g (89%), mp = 240–242 °C. λ Abs (nm) = 359 (THF), 356 (acetonitrile), λ Em (nm) = 460 (THF), 453 (acetonitrile). IR (cm−1): ν (C-H) 2927; ν (C-H aromatic) 3032, ν (C=C) 1678–1450, ν (C=O) 1678; ν (O-H) 3194, ρ (C-H f.p) 763. 1H-NMR (DMSO-d6), δ (ppm): 8.57 (1H, s, H-4), 7.65 (1H, d, 3J 8.6, H-5), 6.86 (1H, dd, 3J 8.6, 4J 2.2, H-6), 6.73 (1H, d, 4J 2.2, H-8), 2.63 (3H, s, H-12). 13C-NMR (DMSO-d6), δ (ppm): 195.1 (C-11), 165.1 (C-2), 159.5 (C-9), 157.8 (C-7), 148.2 (C-4), 133.1 (C-5), 119.4 (C-3), 114.8 (C-6), 111.1 (C-10), 102.2 (C-8), 30.4 (C-12).
3-Acetyl-7-methoxy -2-chromen-2-one (1d) [31]. Obtained from 4-methoxy-salicylaldehyde (0.500 g, 3.28 mmol), ethyl acetoacetate (0.41 mL, 3.3 mmol) and piperidine (17.4 μL, 3 drops). Yellow solid, yield = 0.63 g (87%), mp = 172–174 °C. λ Abs (nm) = 356 (THF), 354 (acetonitrile). IR (cm−1): ν (C-H) 2978, ν (C-H aromatic) 3032, ν (C=C) 1666–1462, ν (C=O) 1670, ν (C-O-C) 1126, ρ (C-H f.p) 767. 1H-NMR (DMSO-d6), δ (ppm): 8.62 (1H, s, H-4); 7.86 (1H, d, 3J 8.6, H-5), 7.02 (1H, dd, 3J 8.6, H-6), 7.04 (1H, s, H-8), 2.56 (3H, s, H-12), 3.9 (3H, s, OMe). 13C-NMR (DMSO-d6), δ (ppm): 195.1 (C-11), 165.3 (C-2), 159.3 (C-9), 157.5 (C-7), 148.0 (C-4), 132.6 (C-5), 120.8 (C-3), 113.9 (C-6), 112.2 (C-10), 100.7 (C-8), 56.7 (OMe), 30.4 (C-12).
Compounds 2a,c were prepared according to previous reports [31]. Compounds 2b,d have not been reported yet. The corresponding 4-substituted compound 1ad, phenylhydrazine, and glacial acetic acid were added to 10 mL of ethanol and refluxed at 78 °C for 4 h. The solvent was removed by gravity filtration, and the solid was washed with ethanol, recrystallized from ethanol at room temperature, and then characterized.
3-(1-(2-Phenylhydrazone) ethyl) 2-chromen-2-one (2a) [32]. Obtained from 1a (0.300 g, 1.07 mmol), phenylhydrazine (0.105 mL, 1.07 mmol) and glacial acetic acid (13.6 μL, 3 drops). Orange solid, yield = 0.39 g (74%), mp = 184–186 °C. λ Abs (nm) = 277 (THF), 275 (acetonitrile). IR (cm−1): ν (C-H) 2978, ν (C-H aromatic) 3032, ν (C=C) 1597–1489, ν (NH) 3294, ν C=N (1597), ρ (C-H f.p) 732. 1H-NMR (DMSO-d6), δ (ppm): 9.46 (NH, s), 8.20 (1H, s, H-4), 7.84 (1H, d, 3J 7.8, H-5), 7.60) (1H, dd, 3J 7.9, 7.4, H-7), 7.41 (1H, d, 3J 8.0, H-8), 7.36 (1H, dd, 3J 7.9, 7.5, H-6), 7.21–7.27 (4H, o, m, -Ph), 6.79 (1H, p, 3J 7.0, -Ph). 2.23 (3H, s, H-12), 13C-NMR (DMSO-d6), δ (ppm): 160.4 (C-2), 153.4 (C-9), 146.0 (C-11), 139.6 (C-4), 139.1 (ipso, -Ph), 132.0 (C-7), 129.3 (C-5), 129.2 (m, -Ph), 127.8 (C-3), 125.0 (C-6), 119.8 (C-10), 119.7 (p, -Ph), 116.3 (C-8), 113.5 (o, -Ph), 15.5 (C-12).
7-(Diethylamino)-3-(1-(2-phenylhydrazone)ethyl)2-chromen-2-one (2b). Obtained from 1b (0.300 g, 1.15 mmol), phenylhydrazine (0.113 mL, 1.15 mmol) and glacial acetic acid (13.6 μL, 3 drops). Orange solid, yield = 0.32 g (65%), mp = 236–238 °C. λ Abs (nm)= 411 (THF), 413 (acetonitrile), 408 (toluene), λ Em (nm) = 479 (THF), 494 (acetonitrile), 469 (toluene). IR (cm−1): ν (C-H) 2900; ν (C-H aromatic) 2966; ν (C=C) 1589–1408, ν (NH) 3302, ν C=N (1589), ρ (C-H f.p) 748. 1H-NMR (DMSO-d6), δ (ppm): 9.23 (NH, s), 8.01 (1H, s, H-4), 7.56 (1H, d, 3J 8.9, H-5), 7.22-7-20 (4H, o, m, -Ph), 6.73 (1H, p, -Ph), 6.71 (1H, d, 3J 8.9, H-6), 6.55 (1H, s, H-8), 2.20 (3H, s, H-12), 3.45 (4H, q, 3J 6.9, (N (CH2CH3)2), 1.14 (6H, t, 3J 7.0, N (CH2CH3)2), 13C-NMR (DMSO-d6), δ (ppm): 160.9 (C-2), 156.4 (C-9), 150.9 (C-7), 146.4 (C-11), 140.4 (ipso, -Ph), 140.6 (C-4), 130.2 (C-5), 129.2 (m, -Ph), 120.1 (C-3), 119.3 (p, -Ph), 113.4 (m, -Ph), 109.7 (C-6), 108.6 (C-10), 96.6 (C-8), 44.5 (N (CH2CH3)2), 15.7 (C-12), 12.8 (N (CH2CH3)2). MS-MS (ESI) m/z = 350.1866 [M + H]+ (experimental), 350.1869 (calculated).
7-(Hydroxy)-3-(1-(2-phenylhydrazone)ethyl)-2-chromen-2-one (2c) [26]. Obtained from 1c (0.300 g, 1.46 mmol), phenylhydrazine (0.144 mL, 1.46 mmol) and glacial acetic acid (13.6 μL, 3 drops). Yellow solid, yield = 0.36 g (65%), mp = 228–230 °C. λ Abs (nm) = 278 (THF), 276 (acetonitrile), λ Em (nm) = 465 (THF). IR (cm−1): ν (C-H aromatic) 3032; ν (C=C) 1593–1442, ν (NH) 3325, ν (O-H) 3221 ν (C=N) 1597, ρ (C-H f.p) 748. 1H-NMR (DMSO-d6), δ (ppm): 10.62 (1H, s, OH) 9.31 (NH, s), 8.09 (1H, s, H-4), 7.66 (1H, d, 3J 8.4, H-5), 6.81 (1H, d, 3J 8.4, H-6), 6.74 (1H, s, H-8), 7.22 (4H, o, p, -Ph), 6.77 (1H, p, -Ph), 2.20 (3H, s, H-12). 13C-NMR (DMSO-d6), δ (ppm): 161.6 (C-7), 160.4 (C-2), 155.5 (C-9), 146.0 (C-11), 140.4 (C-4), 139.7 (ipso, -Ph), 130.6 (C-5), 129.2 (m, -Ph), 123.3 (C-3), 119.6 (p, -Ph), 113.9 (C-6), 113.4 (o, -Ph), 112.2 (C-10), 102.2 (C-8), 15.6 (C-12).
7-(Methoxy)-3-(1-(2-phenylhydrazone)ethyl)-2-chromen-2-one (2d). Obtained from 1d (0.300 g, 1.48 mmol), phenylhydrazine (0.148 mL, 1.46 mmol) and glacial acetic acid (13.6 μL, 3 drops). Yellow solid, yield = 0.41 g (65%), mp = 228–230 °C. λ Abs (nm) = 278 (THF), 275 (acetonitrile), λ Em (nm) = 467 (THF). IR (cm−1): ν (C-H) 2997, ν (C-H aromatic) 3020, ν (C=C) 1612–1496, ν (NH) 3309, ν (C=N) 1608, ρ (C-H f.p) 736. 1H-NMR (DMSO-d6), δ (ppm): 9.36 (NH, s), 8.15 (1H, s, H-4), 7.77 (1H, d, 3J 8.65 H-5), 7.23–7.22 (4H, o, m, -Ph); 7.0 (1H, s, H-8), 6.98 (1H, d, 3J 8.6, H-6), 6.79 (1H, p, -Ph), 3.88 (3H, s, OMe), 2.21 (3H, s, H-12), 13C-NMR (DMSO-d6), δ (ppm): 162.8 (C-2), 160.3 (C-2), 155.4 (C-9), 146.1 (C-11), 140.1 (C-4), 139.5 (ipso, -Ph), 130.3 (C-5), 129.3 (m, -Ph), 124.3 (C-3), 119.6 (p,-Ph), 113.5 (C-6), 113.3 (o, -Ph), 113.3 (C-6), 100.7 (C-8), 56.4 (OMe), 15.5 (C-12). MS-MS (ESI) m/z = 309.1235 [M + H]+ (experimental), 309.1239 (calculated).

2.3. Determination of Quantum Yield

Fluorescence quantum yield (Φ) for coumarins (1ad) and hydrazones (2ad) were measured through the integration sphere direct excitation method, with “Direct Excitation” measurements, which record the scatter and the emission of the sample being directly excited by the radiation from the excitation monochromator only. The absolute fluorescence quantum yield (Φ) is the ratio of the number of photons emitted to the number of photons absorbed, according to Equation (1).
ϕ = N e m N a b s
The absolute fluorescence quantum yield (ΦDexc) was calculated with the direct excitation method according to Equation (2).
ϕ D e x c = E B E A S A S B
where E denotes the emission region of sample and solvent, S the excitation scatter region of sample and solvent, and A, B the experimental setup and integral of the scans of samples and solvent.
For calculating the integrals, the selection of the integral regions, and the final calculation of ΦDexc, we used the quantum yield wizard supplied with the equipment software [33].

2.4. Computational Details

Geometry optimizations were carried out using the framework of density functional theory implemented in ORCA software [34]. For the exchange and correlation, the PBE0 hybrid functional [35,36] was combined with a complete electronic basis set aug-cc-pVDZ [37]. A vibrational frequency analysis was effectuated to verify minima energy states; only positive values were found in all cases. The HOMO-LUMO gap energies were calculated, and frontier orbitals were plotted using Chemcraft [38]. Excited states were determined using TD-DFT [39,40,41] under the same level of theory PBE0/aug-cc-pVDZ [42], and solvent effects were included using a conductor-like polarizable model (CPCM) [43,44] considering THF, acetonitrile, and toluene.

3. Results and Discussion

3.1. Synthesis and Structural Characterization

Coumarins 1ad were synthesized by the Knoevenagel [28] condensation of ethyl acetoacetate with the corresponding 4-substituted salicylaldehyde (Scheme 1), with yields higher than 80%. Afterward, the functionalization of 3-acetyl-coumarin derivatives 1ad with phenylhydrazine was performed through acid catalysis with glacial acetic acid to obtain compounds 2ad (Scheme 1). The experimental yields are in the 65–74% range, according to precise reports about the synthesis of the same compounds [2,29,32,45].
The H-4 signal in the 1H NMR spectra indicated the formation of compounds 1ad (Figures S1–S8), appearing in the 8.40–8.67 ppm range [31]. The functionalization with phenylhydrazine in compounds 2ad, generated an anisotropic effect perpendicularly extended from the phenyl group of the hydrazone fragment to the pyran ring, causing the protection of H-4 and shifting the signal to lower frequencies in the 8.00–8.15 ppm range. The NH proton in the phenylhydrazone was shifted to high frequencies appearing in the 9.31–9.46 ppm range; whereas the 13C-NMR signal was shielded from 194.6–195.5 ppm, for compounds 1ad, to the range of 146.0–146.4 ppm in compounds 2ad (Figures S9–S16). The IR spectra of compounds 1ad showed the C=O band in the 1658–1681 cm−1 range. Meanwhile, compounds 2ad showed the characteristic hydrazone C=N and N-H bands in the 1589–1608 cm−1 and 3294–3325 cm−1 ranges, respectively. Selected 1H and 13C NMR values and IR wavenumbers are listed in Table 1.

3.2. UV-vis Absorption and Photoluminescence Spectra

The UV-vis and fluorescence spectroscopic data of compounds 1ad and 2ad, measured in tetrahydrofuran and acetonitrile solvents at room temperature, are listed in Table 2. Their scarce solubility in solvents, such as ethanol, DMF, and water, limited the tests in these solvents. The absorption and emission spectra in solution are shown in Figure 1 and Figure 2, respectively. The maximum absorbance bands are attributed to π → π* and n → π* electronic transitions. The absorption band in 1bd showed a red-shift due to the insertion of electron-donor substituents in the C-7 position. The largest red-shift in 1b is attributed to the strong electron-donor character of the diethylamino group [4,9]. Similar behavior is conserved in phenylhydrazone derivatives, wherein 2b shows the largest bathochromic effect in the absorption band. The molar absorptivity depends on the solvent used, obtaining higher values with acetonitrile than with THF. In contrast to the full palette observed in the absorption, compounds 1a,d and 2a showed no emission band. The maximum wavelength of the coumarin emission band is around 453–494 nm. The increase of emission is related to the nature of the substituent in C-7 following the order: N (CH2CH3)2 > OH > OCH3 > H, in agreement with the strength of the substituent electron-donor properties [4,9]. Compound 1b showed the largest quantum yield (ΦF) with values above 80%.
On the other hand, a brief comparison between the photophysical properties of coumarins 1ad with the corresponding values of hydrazones 2ad allowed us to conclude that coumarin functionalization in the C-3 position with hydrazine caused fluorescence quenching with null fluorescence yields. This result is explained by the diminished electron acceptor capabilities of the resulting hydrazone functionality.
Compounds 1b,d, and 2c showed the best photophysical properties in the solid-state. Their fluorescence spectra are shown in Figure 3, and their photophysical properties are listed in Table 3. The emission λmax of coumarin 1b was red-shifted to 595 nm in comparison with the solution (477 nm), and unexpectedly, compound 1d turned emissive, showing a band at 540 nm, although with low ΦF values (<6.6%) in both cases. In contrast, hydrazone 2c has emissive properties compared to the solution: emission at 582 nm and ΦF value of 52.6. These results revealed that the intermolecular interactions in the solid-state favor the emissive relaxation mechanism from an excited state. The X-ray molecular structures of compounds 1a,b,d are known [46,47,48,49]. Hence, a brief analysis of their photoluminescent properties in relation with their molecular structure in the solid state seems appropriate. The powder X-ray diffractogram of 1a,b,d was obtained and compared to the single crystal reported ones, shown in Figure S17. In the solid, the free rotation of the pendant –OMe and 3-Ac groups is restricted due to intermolecular hydrogen bonds, leading to a coplanar system and increasing the emissive properties of 1d. Therefore, the solid-state exerts an aggregation-induced emission (AIE) effect [46] on 1d. In the solid, the molecules of 1d are π-stacked along the (100) direction, at a distance of 3.509 Å between the benzo fused (Bz) and lactone (L) rings centroids (Cg (Bz)-Cg (L)) [47], in agreement with medium strength π-D-π-A interactions arranged in the head-to-tail fashion [50]. It is worth mentioning that the Bz ring acts as the donor and the L ring as the acceptor. In contrast, coumarin 1a crystallizes in two forms, triclinic (form A) and monoclinic (form B). Form A (herein obtained) is a head-to-head stacked, while form B generates head-to-tail π-stacked dimers with long Cg (Bz)-Cg (L) distances of 4.77 Å and 3.98 Å, respectively [48]. On the contrary, an aggregation-caused quenching (ACQ) effect is observed on 1b. In the crystal lattice, the molecules of 1b are arranged in dimers through antiparallel type carbonyl–carbonyl interactions [51] at Cg (Bz)-Cg (L) distance of 6.38 Å and the amino-ethyl groups pointing to opposite directions [49]. The free rotation of the pendant –NEt2 and 3-Ac groups in 1b is restricted as in 1d, but with the contrary effect. This brief analysis supports that the ΦF in the solid-state depends on the crystal lattice arrangement. Coumarin molecules that show head-to-tail π-staking arrangement with a Cg(Bz)-Cg(L) distance shorter than 3.5 Å seem to favor large ΦF values. The large value of ΦF of hydrazone 2c in the solid-state is also explained as an AIE effect. Intermolecular hydrogen bonding, probably between the –OH and the C=N functionality, should favor an entire head-to-tail π-staking arrangement in the crystal lattice with the appropriate Cg (Bz)-Cg (L) distance, as mentioned earlier.
Compounds 15 have been used as fluorescent dyes for cellular imaging (1), indicators of bathochromic effects with red emissions (2), chemosensors for anions (3), for the study of nonlinear optical properties (4), and observation of highly fluorescent compounds depending on the substituents (5). Their photophysical properties are listed in Table 4. 7-Diethylamino substituted coumarin 1b showed better photophysical properties (ΦF) in solution than compounds 15, previously reported and shown in Figure 4 [6,7,10,16,52]. Then, compound 1b in solution and 2c in the solid-state could be evaluated in similar applications, since they have even better optical properties, and their synthesis is simpler and with higher chemical yields than the reported compounds 17.

3.3. Theoretical Calculations

The theoretical data of the molecular orbitals allowed us to explain the spectroscopic behavior and the electronic effects in the molecules. The HOMO (H), LUMO (L), H-L gap, and Egop energy values are listed in Table 5. The H-L gap energies are similar among them. Nevertheless, the smallest value corresponds to compound 1b, whose λmax is displaced towards the red. Likewise, the H-L gap represents the excited state electronic transition [6,38] from S0 to S1. Compounds 2ad show low H-L gap values similar to coumarins 1ad. The theoretical values, listed in Table 5, agree with the experimental optical values.
The optical gap values (Egop) for coumarins rank from 2.90 to 4.18 eV, and they are smaller than the theoretical data (3.84–4.47 eV); where compound 1b shows the smallest Egop value, indicating good ICT capability. Egop values of compounds 2ad show low energy values (3.03–4.52 eV) like coumarins. These results indicate that the absorbance and emission red-shifts are due to the good electronic efficiency in their D-A structure. The theoretical and experimental H-L gap values are in agreement with the experimental optical properties and increase their possibility to function as semiconductor compounds [55]. The experimental values are comparable with the data for well-known luminophores 3, 6, and 7 listed in Table 6 and whose structures are shown in Figure 5.
The data of compounds 1bd showed HOMO-LUMO interactions (Table S1) in a higher percentage of contribution energy due to the ED groups generating n → π* transitions, in contrast to compound 1a, which causes HOMO-1-LUMO π → π* transitions, in agreement with the experimental data.
Compounds 2ad showed π → π* and n → π*transitions represented by the HOMO-LUMO, HOMO+1-LUMO-1 orbitals (supplementary material Table S2), indicating a D-A behavior of the molecules which is consistent with the experimental data.
Figure 5 shows the FMOs (frontier molecular orbitals) of compounds 1ad. The structures show a planar geometry through the benzopyran ring. The HOMO-LUMO and HOMO-1-LUMO electronic distributions indicate the presence of π → π* and n → π* transitions, confirming the excitation of the FMOs from S0 to S1.
In compounds 2ad, the phenylhydrazine group acts as an ED group. In this case, it represents the energy of the HOMO, and often, the whole part of the coumarin where the substituent in the benzopyrone ring acts as EA group, represented by the energy of the LUMO (Figure 6). The previous result agrees with the experimental findings, suggesting that the electronic cloud of the ED moiety moves towards the EA, which is from the phenylhydrazine to the coumarin. The hydrazo-coumarins 2ad showed low quantum yields and unobservable emissions because the electronic cloud does not include the ED group which is interacting with the solvents.
Compound 2c shows different behavior in the FMOs, and its electron density is both as HOMO and LUMO in the hydrazone group, generating only n → π* type transitions, which agrees with its behavior in the solid-state, since the solvent generates the quenching of the fluorescence in solution.

4. Conclusions

Molecules 1ad and 2ad can be good candidates for semiconductor compounds suitable to be used in several optical applications. The photophysical properties of coumarins 1ad and their hydrazone derivatives 2ad were compared. The bandgap indicates the efficient ICT, which confirms the D-A type structure. Compounds 1b and 2c presented the best optical properties. Compound 1b showed absorbances at λ > 420 nm and emission at λ > 460 nm in solution and quantum yields >80%. Compound 2c also showed good emission values in the solid-state due to the push-pull system, supported by the π-π interactions. Compound 1d turned emissive in the solid-state due to the π-π- interactions in the crystal network. This work also confirms that small coumarins derivatives possess photoluminescent properties which can be tuned in both solution and in the solid-state.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27123677/s1, Figures S1–S16, 1H and 13C NMR spectra; Figure S17, Powder X-ray diffractogram of 1a,b,d; Table S1, HOMO-LUMO interactions of compounds 1bd, Table S2, HOMO-LUMO interactions of compounds 2bd.

Author Contributions

Conceptualization, F.J.M.-M., M.R. and N.E.M.-V.; Methodology, C.A.V.-M.; Software, J.P.M.-S. and J.B.-F.; Investigation, I.I.P.-M. and Á.A.R.-O.; Writing, C.A.V.-M., F.J.M.-M. and I.I.P.-M. All authors have read and agreed to the published version of the manuscript.

Funding

Partially supported by Universidad de Colima and Consejo Nacional de Ciencia y Tecnología (CONACYT, grants, 316892 and 255354) and SIP-IPN.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Centro de Investigaciones en Óptica de León Gto. Mexico. The computational calculations were supported by the high-performance computing system of PIDi-UTEM (SCC-PIDi-UTEM-CONICYT-FONDEQUIP-EQM180180).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 1ad and 2ad are available from the authors.

References

  1. Rodríguez-Romero, J.; Aparicio-Ixta, L.; Rodríguez, M.; Ramos-Ortíz, G.; Maldonado, J.L.; Jiménez-Sánchez, A.; Farfán, N.; Santillan, R. Synthesis, chemical-optical characterization and solvent interaction effect of novel fluorene-chromophores with D-A-D structure. Dyes Pigments 2013, 98, 31–41. [Google Scholar] [CrossRef]
  2. Kamath, P.R.; Sunil, D.; Ajees, A.A.; Pai, K.S.R.; Das, S. Some new indole-coumarin hybrids; synthesis, anticancer and Bcl-2 docking studies. Bioorg. Chem. 2015, 63, 101–109. [Google Scholar] [CrossRef]
  3. Srinivasa, H.T.; Palakshamurthy, B.S.; Abdulkarim, T.M. Ethyl 7-hydroxycoumarin-3-carboxylate derivatives: Synthesis, characterization and effect of liquid crystal properties. J. Mol. Struct. 2018, 1155, 513–519. [Google Scholar] [CrossRef]
  4. Akchurin, I.O.; Yakhutina, A.I.; Bochkov, A.Y.; Solovjova, N.P.; Medvedev, M.G.; Traven, V.F. Novel push-pull fluorescent dyes-7-(diethylamino)furo-and thieno [3,2-c] derivatives: Structure, electronic spectra and TD-DFT study. J. Mol. Struct. 2018, 1160, 216–221. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Shi, Z.; Li, Y. Photoexcitation mechanism of new D-π-A coumarin derivatives in linear and nonlinear optical processes. J. Mater. Sci. 2016, 27, 7132–7140. [Google Scholar] [CrossRef]
  6. García, S.; Vázquez, J.L.; Renteria, M.; Aguilar, G.I.G.; Delgado, F.; Trejo, D.M.; Garcia, R.M.A.; Alvarado, M.E.; Vazquez, M.A. Synthesis and experimental computational characterization of nonlinear optical properties of triazacyclopentafluorene-coumarin derivatives. Opt. Mater. 2016, 62, 231–239. [Google Scholar] [CrossRef]
  7. Chen, J.; Liu, W.; Ma, J.; Xu, H.; Tang, X.; Fan, Z.; Wang, P. Synthesis and properties of fluorescence dyes: Tetracyclic pyrazolo [3,4-b]pyridine-based coumarin chromophores with intramolecular charge transfer character. J. Org. Chem. 2012, 77, 3475–3482. [Google Scholar] [CrossRef]
  8. Bassolino, G.; Nancoz, C.; Thiel, Z.; Bois, E.; Vauthey, E.; Rivera, F.P. Photolabile coumarins with improved efficiency through azetidinyl substitution. Chem. Sci. 2018, 9, 387–391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Aksungur, T.; Aydiner, B.; Seferolu, N.; Ozkutuk, M.; Arslan, L.; Reis, Y.; Acik, L.; Seferoglu, Z. Coumarin-indole conjugate donor-acceptor system: Synthesis photophysical properties, anion sensing ability, theoretical and biological activity studies of two coumarin-indole based push-pull dyes. J. Mol. Struct. 2017, 1147, 364–379. [Google Scholar] [CrossRef]
  10. Aydiner, B.; Yalcin, E.; Korkmaz, V.; Seferoglu, Z. Efficient one-pot three-component method for the synthesis of highly fluorescent coumarin-containing 3,5-disubstituted-2,6-dicyanoaniline derivatives under microwave irradiation. Synth. Commun. 2017, 47, 2174–2188. [Google Scholar] [CrossRef]
  11. Sinha, S.; Gaur, P.; Dev, S.; Mukhopadhyay, S.; Mukherjee, T.; Ghosh, S. Hydrazine response molecular material: Optical signaling and mushroom cell staining. Sens Actuators B Chem. 2015, 221, 418–426. [Google Scholar] [CrossRef]
  12. Huang, Y.; Zhang, Y.; Yuan, Y.; Wenwen, C. Organogelators based on iodo 1,2,3-triazole functionalized with coumarin: Properties and gelator solvent interaction. Tetrahedron. 2015, 71, 2124–2133. [Google Scholar] [CrossRef]
  13. Li, T.; Wang, L.; Lin, S.; Xu, X.; Liu, M.; Shen, S.; Yan, Z.; Mo, R. Rational design and bioimaging applications of highly specific “Turn-On” fluorescent probe for hypochlorite. Bioconjugate Chem. 2018, 29, 2838–2845. [Google Scholar] [CrossRef] [PubMed]
  14. Minoru, Y.; Yuma, H.; Hideki, O.; Fumito, T. Photochemical synthesis and photophysical properties of coumarins bearing extended polyaromatic rings studied by emission and transient absorption measurements. Photochem. Photobiol. Sci. 2017, 16, 555–563. [Google Scholar] [CrossRef]
  15. Arcos, R.R.; Maldonado, D.M.; Ordóñez, H.J.; Romero, A.M.; Farfán, N.; del Pilar Carreón-Castro, M. 3-Substituted-7-(diethylamino)coumarins as molecular scaffolds for the bottom-up self-assembly of solids with extensive π-stacking. J. Mol. Struct. 2017, 1130, 914–921. [Google Scholar] [CrossRef]
  16. Tathe, A.B.; Gupta, V.D.; Sekar, N. Synthesis and combined experimental and computational investigations on spectroscopic and photophysical properties of red emitting 3-styryl coumarins. Dyes Pigments 2015, 119, 49–55. [Google Scholar] [CrossRef]
  17. Liangliang, L.; Yanjun, W.; Lin, W.; Aihua, G.; Feilong, H.; Chi, Z. A fluorescent probe for hypochlorite based on the modulation of the unique rotation of the N-N single bond in acetohydrazine. Chem. Commun. 2015, 51, 10435–10438. [Google Scholar] [CrossRef]
  18. Wu, Q.; Ma, L.; Xu, Y.; Cao, D.; Guan, R.; Liu, Z.; Yu, X. Two Coumarin formhydrazide compounds as chemosensors for copper ions. Inorg. Chem. Commun. 2016, 69, 7–9. [Google Scholar] [CrossRef]
  19. Essaidi, Z.; Krupka, O.; Iliopoulos, K.; Champigny, E.; Sahraoui, B.; Sallé, M.; Gindre, D. Synthesis and functionalization of coumarin containing copolymers for second order optical nonlinearities. Opt. Mater. 2013, 35, 576–581. [Google Scholar] [CrossRef] [Green Version]
  20. Mukherjee, K.; Chio, T.I.; Gu, H.; Banerjee, A.; Sorrentino, A.M.; Sackett, D.L.; Bane, S.L. Benzocoumarin hydrazine: A large stoke shift fluorogenic sensor for detecting carbonyls in isolated biomolecules and in live cells. ACS Sens. 2017, 2, 128–134. [Google Scholar] [CrossRef] [PubMed]
  21. Dai, X.; Zhao, Y.W.; Zhi, F.D.; Jun, Y.M.; Bao, X. Z A simple but effective near infrared ratiometric fluorescent probe for hydrazine and its application in bioimaging. Sens. Actuator B-Chem. 2016, 232, 369–374. [Google Scholar] [CrossRef]
  22. Dehestani, L.; Ahangar, N.; Hashemi, S.M.; Irannejad, H.; Masihi, P.H.; Shakiba, A.; Emami, S. Desing, synthesis, in vivo and in silico evaluation of phenacyl triazole hidrazones as new anticonvulsant agents. Bioorg. Chem. 2018, 78, 119–129. [Google Scholar] [CrossRef]
  23. Iturmendi, A.; Theis, S.; Maderegger, D.; Monkowius, U.; Teasdale, I. Coumarin caged polyphosphazenes with a visible light driven on demand degradation. Macromol. Rapid Commun. 2018, 39, 1800377–1800383. [Google Scholar] [CrossRef] [PubMed]
  24. Soni, J.N.; Soman, S.S. Reactions of coumarin-3-carboxylate, its crystallographic study and antimicrobial activity. Pharma Chem. 2014, 6, 396–403. [Google Scholar]
  25. Ming, K.L.; Jing, L.; Ying, Z.; Xiao, Y.X.; Xin, S.; Zhe, Y.T.; Zheng, H.; Xiao, X.L. Synthesis, photoluminescent behaviors, and antibacterial activities of 3-acetylcoumarins. Res. Chem. Intermed. 2015, 41, 3289–3296. [Google Scholar] [CrossRef]
  26. Liu, M.; Xu, Z.; Song, Y.; Li, H.; Xian, C. A novel Coumarin Based chemosensor for colorimetric detection of Ag(I) ion and fluorogenic sensing of Ce(III) ion. J. Lumines. 2018, 17, 31882–311991. [Google Scholar] [CrossRef]
  27. Chimenti, F.; Secci, D.; Bolasco, A.; Chimenti, P.; Granese, A.; Befani, O.; Turini, P.; Alcaro, S.; Ortuso, F. Inhibition of monoamine oxidases by coumarin-3-acyl derivatives: Biological activity and computational study. Bioorg. Med. Chem. Lett. 2004, 14, 3697–3703. [Google Scholar] [CrossRef] [PubMed]
  28. Valizadeh, H.; Shockravi, A.; Gholipur, H. Microwave assisted synthesis of coumarins via potassium carbonate catalyzed Knoevenagel condensation in 1-n-butyl-3-methylimidazolium bromide ionic liquid. J. Heterocycl. Chem. 2007, 44, 867–871. [Google Scholar] [CrossRef]
  29. Grover, J.; Kumar, V.; Sobhia, M.E.; Jachak, S.M. Synthesis, biological evaluation and docking analysis of 3-methyl-1-phenylchromeno [4,3-c]pyrazol-4(1H)-ones as potential cyclooxygenase-2 (COX-2) inhibitors. Bioorg. Med. Chem. Lett. 2014, 24, 4638–4642. [Google Scholar] [CrossRef]
  30. Dai, X.; Qing, H.W.; Peng, C.W.; Jie, T.; Yu, X.; Sheng, Q.W.; Jun, Y.M.; Bao, X.Z. A simple and effective coumarin based fluorescent probe for cysteine. Biosens. Bioelectron. 2014, 59, 35–39. [Google Scholar] [CrossRef] [PubMed]
  31. Kumar, J.A.; Saidachary, G.; Mallesham, G.; Sridhar, B.; Jain, N.; Kalivendi, S.V.; Jayathirtha, R.J.; Raju, C.B. Synthesis, anticancer activity and photophysical properties of novel substituted 2-oxo-2H-cromenylpyrazolecarboxylates. Eur. J. Med. Chem. 2013, 65, 389–402. [Google Scholar] [CrossRef]
  32. Padilla-Martinez, I.; Flores, L.I.Y.; García, B.E.V.; Gonzalez, J.; Cruz, A.; Martinez-Martinez, F.J. X-Ray supramolecular structure, NMR spectroscopy and synthesis of 3-methyl-1-phenyl-1H-cromeno [3,4-c]pyrazol-4-ones formed by the unexpected cyclization of 3-[1-(phenyl-hydrazono)ethyl]-cromen-2-ones. Molecules 2011, 16, 915–932. [Google Scholar] [CrossRef] [Green Version]
  33. Reference Guide Edinburgh Instruments FLS980; Integrating sphere for measurements of fluorescence quantum yields and spectral reflectance; Edinburgh Instruments Ltd.: Livingston, UK, 2016.
  34. Neese, F. The ORCA program system. Wiley Interdisciplinary Reviews: Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  36. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  37. Kendall, R.A.; Dunning, T.H., Jr.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  38. Zhurko, G.A. Chemcraft-Graphical Program for Visualization of Quantum Chemistry Computations. Available online: https://chemcraftprog.com (accessed on 12 April 2022).
  39. Laurent, A.D.; Jacquemin, D. TD-DFT benchmarks: A review. Int. J. Quantum. Chem. 2013, 113, 2019–2039. [Google Scholar] [CrossRef]
  40. Jacquemin, D.; Wathelet, V.; Perpete, E.A.; Adamo, C. Extensive TD-DFT benchmark: Singlet-excited states of organic molecules. J. Chem. Theory Comput. 2009, 5, 2420–2435. [Google Scholar] [CrossRef]
  41. Charaf-Eddin, A.; Planchat, A.; Mennucci, B.; Adamo, C.; Jacquemin, D. Choosing a functional for computing absorption and fluorescence band shapes with TD-DFT. J. Chem. Theory Comput. 2013, 9, 2749–2760. [Google Scholar] [CrossRef] [PubMed]
  42. Taouali, W.; Casida, M.E.; Znaidia, S.; Alimi, K. Rational design of (DA) copolymers towards high efficiency organic solar cells: DFT and TD-DFT study. J. Mol. Graph. 2019, 89, 139–146. [Google Scholar] [CrossRef] [PubMed]
  43. Takano, Y.; Houk, K.N. Benchmarking the conductor-like polarizable continuum model (CPCM) for aqueous solvation free energies of neutral and ionic organic molecules. J. Chem. Theory Comput. 2005, 1, 70–77. [Google Scholar] [CrossRef]
  44. Raj, R.K.; Gunasekaran, S.; Gnanasambandan, T. Combined spectroscopic and DFT studies on 6-bromo-4-chloro-3-formyl coumarin. Spectrochim. Acta Part A: Mol. Biomol. Spectrosc. 2015, 139, 505–514. [Google Scholar] [CrossRef]
  45. Grover, J.; Roy, K.S.; Jachak, S.M. Potassium carbonate mediated efficient and convenient synthesis of 3-methyl-1-phenylchromeno [4,3-c]pyrazol-4(1H)-ones. Synth. Commun. 2014, 44, 1914–1923. [Google Scholar] [CrossRef]
  46. Luo, J.; Xie, Z.; Lam, J.W.Y.; Cheng, L.; Chen, H.; Qiu, C.; Kwok, H.S.; Zhan, X.; Liu, Y.; Zhu, D.; et al. Aggregation-induced emission of 1-methyl-1,2,3,4,5-pentaphenylsilole. Chem. Commun. 2001, 18, 1740–1741. [Google Scholar] [CrossRef] [PubMed]
  47. Han, H.-M.; Lu, C.-R.; Zhang, Y.; Zhang, D.-C. 3-Acetyl-7-methoxycoumarin. Acta Crystallogr. Sect. E 2005, 61, o1864. [Google Scholar] [CrossRef]
  48. Parthapratim, M.; Tayur, N.; Guru, R. Topological Analysis of Charge Density Distribution in Concomitant Polymorphs of 3-Acetylcoumarin, A Case of Packing Polymorphism. Cryst. Growth Des. 2006, 6, 708–718. [Google Scholar]
  49. Christopher, G.; Hamaker, C.S. McCully, 3-Acetyl-7-(diethyl amino)coumarin. Acta Crystallogr. Sect. E 2006, 62, o2072–o2074. [Google Scholar] [CrossRef]
  50. García-Báez, E.V.; Martínez-Martínez, F.J.; Höpfl, H.; Padilla-Martínez, I.I. π-Stacking Interactions and CH·X (X = O, aryl) Hydrogen Bonding as Directing Features of the Supramolecular Self-Association in 3-Carboxy and 3-Amido Coumarin Derivatives. Cryst. Growth Des. 2003, 3, 35–45. [Google Scholar] [CrossRef]
  51. Cabrera-Pérez, L.C.; García-Báez, E.V.; Franco-Hernández, M.O.; Martínez-Martínez, F.J.; Padilla-Martínez, I.I. Carbonyl–carbonyl interactions and amide π-stacking as the directing motifs of the supramolecular assembly of ethyl N-(2-acetylphenyl)-oxalamate in a synperiplanar conformation. Acta Crystallogr. Sect. C 2015, C71, 381–385. [Google Scholar] [CrossRef]
  52. Yanar, U.; Babur, B.; Pekyilmaz, D.; Yahaya, I.; Aydiner, B.; Dede, Y.; Seferoglu, Z. A fluorescent coumarin thiophene hybrid as a ratiometric chemosensor for anions: Synthesis, photophysics, anions sensing and orbital interactions. J. Mol. Struct. 2016, 1108, 269–277. [Google Scholar] [CrossRef]
  53. Shu, X.; Li, J.; Li, M.-K. Crystal structure photoluminescent and theoretical studies of 3-acetyl-8-methoxy-coumarine derivatives. Asian J. Chem. 2014, 26, 8427–8430. [Google Scholar] [CrossRef]
  54. Garcias-Morales, C.; Romero-Borja, D.; Maldonado, J.-L.; Roa, A.E.; Rodríguez, M.; García-Merinos, J.P.; Ariza-Castolo, A. Small Molecules derived from Thieno [3,4-c] pyrrole-4,6-dione (TPD) and their use in solution processed organic solar cells. Molecules 2017, 22, 1607–1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Costa, J.C.S.; Taveira, R.J.S.; Lima, C.F.R.A.C.; Mendes, A.; Santos, L.M.N.B.F. Optical band gaps of organic semiconductor materials. Opt. Mater. 2016, 58, 51–60. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of 3-acetyl-2H-chromen-2-one (1ad) and 3-(1-(2-phenylhydrazone)ethyl)-2-chromen-2-one (2ad).
Scheme 1. Synthesis of 3-acetyl-2H-chromen-2-one (1ad) and 3-(1-(2-phenylhydrazone)ethyl)-2-chromen-2-one (2ad).
Molecules 27 03677 sch001
Figure 1. Absorption UV-vis spectra of compounds 1ad and 2ad (c = 1 × 10−4 M, at 25 °C) in (a) THF, and (b) acetonitrile.
Figure 1. Absorption UV-vis spectra of compounds 1ad and 2ad (c = 1 × 10−4 M, at 25 °C) in (a) THF, and (b) acetonitrile.
Molecules 27 03677 g001
Figure 2. Fluorescence spectra of compounds 1b,c, 2bd in THF and 1b,c, 2b in acetonitrile solutions (c = 1 × 10−4 M, at 25 °C). λex (Em) = λmax (Abs). Compound 2c,d do not have emission.
Figure 2. Fluorescence spectra of compounds 1b,c, 2bd in THF and 1b,c, 2b in acetonitrile solutions (c = 1 × 10−4 M, at 25 °C). λex (Em) = λmax (Abs). Compound 2c,d do not have emission.
Molecules 27 03677 g002
Figure 3. Fluorescence spectra in the solid-state of compounds 1b,d, and 2c (at 25 °C).
Figure 3. Fluorescence spectra in the solid-state of compounds 1b,d, and 2c (at 25 °C).
Molecules 27 03677 g003
Figure 4. Reported coumarin compounds with photophysical properties in solution, used for comparison purposes [6,7,10,16,52,53,54].
Figure 4. Reported coumarin compounds with photophysical properties in solution, used for comparison purposes [6,7,10,16,52,53,54].
Molecules 27 03677 g004
Figure 5. Frontier molecular orbitals corresponding to the optimized structures of compounds 1ad.
Figure 5. Frontier molecular orbitals corresponding to the optimized structures of compounds 1ad.
Molecules 27 03677 g005
Figure 6. Frontier molecular orbitals corresponding to the optimized structures of compounds 2ad.
Figure 6. Frontier molecular orbitals corresponding to the optimized structures of compounds 2ad.
Molecules 27 03677 g006
Table 1. Selected 1H, 13C NMR (δ = ppm, DMSO-d6) and IR (ν = cm−1) spectroscopic data for compounds 1ad and 2ad.
Table 1. Selected 1H, 13C NMR (δ = ppm, DMSO-d6) and IR (ν = cm−1) spectroscopic data for compounds 1ad and 2ad.
Comp.δ 1Hδ 13Cν/cm−1Comp.δ 1Hδ 13Cν/cm−1ν/cm−1
H-4C-11C=O H-4N-HC=NN-HC=N
1a8.67195.516812a8.009.46146.032941597
1b8.40194.616582b8.019.23146.433021589
1c8.57195.116782c8.099.31146.033251597
1d8.62195.116702d8.159.36146.133091608
Table 2. Experimental photophysical properties in a solution of compounds 1ad and 2ad.
Table 2. Experimental photophysical properties in a solution of compounds 1ad and 2ad.
CompoundSolventλmaxAbsε (M−1 cm−1)λmax EmStokes Shift
(cm−1)
ΦF (%)
1aTHF
Acetonitrile
297
296
100,870
138,563
NO
NO
NO
NO
NO
NO
1bTHF
Acetonitrile
426
429
42,277
45,357
466
477
2014
2345
6
80
1cTHF
Acetonitrile
359
356
194,444
196,477
460
453
6116
6014
>1
>1
1dTHF
Acetonitrile
356
354
238,403
244,724
NO
NO
NO
NO
NO
NO
2aTHF
Acetonitrile
277
275
21,480
268,504
NO
NO
NO
NO
NO
NO
2bTHF
Acetonitrile
411
413
31,247
39,880
479
494
3454
3970
>1
2
2cTHF
Acetonitrile
278
276
16,525
235,090
465
NO
14465
NO
>1
NO
2dTHF
Acetonitrile
278
275
17,588
237,623
467
NO
14557
NO
>1
NO
THF = Tetrahydrofuran, Quantum Yield (ΦF). λex(Em) = λmax(Abs). NO Not Observed.
Table 3. Photophysical properties of compounds 1b,d, and 2c in the solid-state.
Table 3. Photophysical properties of compounds 1b,d, and 2c in the solid-state.
Compoundλmax Em (nm)ΦF (%)
1b595<1
1d5406.6
2c58252.6
Table 4. Reported photophysical properties of 7-diethylamino-coumarins, used for comparison purposes.
Table 4. Reported photophysical properties of 7-diethylamino-coumarins, used for comparison purposes.
StructureAbsorbance (nm)Emission (nm) ΦF (%)References
1379501 (479)71 (42)[7]
2510585>1[16]
3406482>1[52]
4432490NR[6]
5436544>1[10]
6316458NR[53]
7527551NR[54]
NR = Not reported; (Emission in solid-state).
Table 5. HOMO-LUMO (H-L) gap values of compounds 1ad and 2ad. Data are calculated at aug-cc-pVDZ/PBE0 level of theory. Experimental Egop values are also shown.
Table 5. HOMO-LUMO (H-L) gap values of compounds 1ad and 2ad. Data are calculated at aug-cc-pVDZ/PBE0 level of theory. Experimental Egop values are also shown.
CompoundHOMO (eV)LUMO (eV)H-L Gap (eV)Egop (eV) *
1a−7.28−2.804.474.18
1b−6.29−2.443.842.90
1c−6.93−2.604.323.50
1d−6.81−2.534.283.56
2a−5.34−2.512.834.52
2b−4.93−2.052.883.03
2c−5.57−2.573.003.28
2d−5.18−2.292.883.34
* Obtained from equation Egop = 1243/λ (values obtained from highest wavelengths in Table 2).
Table 6. Comparison of reported theoretical calculations with compounds 1ad y 2ad.
Table 6. Comparison of reported theoretical calculations with compounds 1ad y 2ad.
CompoundsHOMO (eV)LUMO (eV)Gap (eV)References
3−5.68 to −1.83−2.15 to 0.38NR[52]
6−6.21 to −6.35−2.39 to −2.513.69 to 3.95[53]
7−5.17 to −5.34−2.10 to −3.191.98 to 2.47[54]
NR = Not reported.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Villa-Martínez, C.A.; Magaña-Vergara, N.E.; Rodríguez, M.; Mojica-Sánchez, J.P.; Ramos-Organillo, Á.A.; Barroso-Flores, J.; Padilla-Martínez, I.I.; Martínez-Martínez, F.J. Synthesis, Optical Characterization in Solution and Solid-State, and DFT Calculations of 3-Acetyl and 3-(1′-(2′-Phenylhydrazono)ethyl)-coumarin-(7)-substituted Derivatives. Molecules 2022, 27, 3677. https://doi.org/10.3390/molecules27123677

AMA Style

Villa-Martínez CA, Magaña-Vergara NE, Rodríguez M, Mojica-Sánchez JP, Ramos-Organillo ÁA, Barroso-Flores J, Padilla-Martínez II, Martínez-Martínez FJ. Synthesis, Optical Characterization in Solution and Solid-State, and DFT Calculations of 3-Acetyl and 3-(1′-(2′-Phenylhydrazono)ethyl)-coumarin-(7)-substituted Derivatives. Molecules. 2022; 27(12):3677. https://doi.org/10.3390/molecules27123677

Chicago/Turabian Style

Villa-Martínez, Cesar A., Nancy E. Magaña-Vergara, Mario Rodríguez, Juan P. Mojica-Sánchez, Ángel A. Ramos-Organillo, Joaquín Barroso-Flores, Itzia I. Padilla-Martínez, and Francisco J. Martínez-Martínez. 2022. "Synthesis, Optical Characterization in Solution and Solid-State, and DFT Calculations of 3-Acetyl and 3-(1′-(2′-Phenylhydrazono)ethyl)-coumarin-(7)-substituted Derivatives" Molecules 27, no. 12: 3677. https://doi.org/10.3390/molecules27123677

Article Metrics

Back to TopTop