Research Article | | Peer-Reviewed

Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases

Received: 23 December 2024     Accepted: 9 January 2025     Published: 24 January 2025
Views:       Downloads:
Abstract

The first objective of the study is based on experimental characterization of the studied compound. The synthetization process of C11H8O4 (I) involved the O-acetylation of 6-hydroxycoumarin with acetic anhydride, utilizing diethyl ether as a solvent and pyridine as a base. The obtained structure was characterized by both spectroscopic analyses such as ESI-MS, FT-IR, 1H and 13C NMR analysis and by single-crystal X-ray diffraction studies. In the latter case, we employed direct methods to solve the structure of (I) and subsequently refined to a final R value of 0.054 for 1896 independent reflections. In the structure, C—H•••O hydrogen bonds connect the molecules into R22 (8) dimers, which are linked together by C—H•••O interactions, forming layers parallel to the bc crystallographic plane. Similarly, the crystal structure is sustained by π–π interactions between neighboring rings, with inter-centroid distances lower than 3.8 Å. The second objective of the study is to use theoretical calculation methods to analyze the effect of solvent polarity on the energy gap of the boundary molecular orbitals and the overall chemical reactivity of coumarin-6-yl acetate in order to provide a better understanding of stability and reactivity. A series of density functional theory computations were achieved with B3LYP/6-311++G(d,p) basis set in both gas and solvent phases. In addition to the dipole moment, the natural bond orbital charge distribution was estimated in toluene, tetrahydrofuran (THF) and benzene solvents. The calculations were conducted utilizing the Gaussian 09 software, and the outcomes exhibited that the solvents have an influence on the optimized parameters. Furthermore, dual and local reactivity indices as Fukui functions from the natural bond orbital (NBO) charges were estimated in order to have a better comprehension of the electrophilic and nucleophilic regions, as well as the chemical activity of (I). The obtained dipole moment in the gas phase is 6.03 Debye and those in the presence of the solvents are 7.89, 6.87, 7.51 and 6.83 Debye for water, toluene, THF and benzene, respectively. Additionally, the global chemical reactivity parameters exhibit variation contingent on the molecular compound and polarity of the solvents, making this an important consideration in the selection of appropriate solvents for a given chemical reaction. The studied compound shows higher stability in the benzene solvent evidenced by an EHOMO-ELUMO energy gap of 9.48 eV, while its low stability is observed in the gas phase with an EHOMO-ELUMO energy gap of 6.64 eV.

Published in Science Journal of Chemistry (Volume 13, Issue 1)
DOI 10.11648/j.sjc.20251301.12
Page(s) 11-32
Creative Commons

This is an Open Access article, distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution and reproduction in any medium or format, provided the original work is properly cited.

Copyright

Copyright © The Author(s), 2025. Published by Science Publishing Group

Keywords

Coumarin Ester, Crystal Structure, Spectroscopic Analysis, Quantum Chemical Calculations, Fukui Functions

1. Introduction
Coumarin is a sweet-smelling substance naturally present in many plants, such as cinnamon and tonka bean. Its derivatives have been identified in various plants that are frequently utilized as licorice flavorings. These plants include, but are not limited to, fennel, aniseed, and licorice root . Coumarin has been utilized as a flavoring agent in the food and cosmetics industry for a considerable duration. Despite its continued employment in the cosmetics industry, its utilization has been discontinued in the food industry due to the potential for toxic and deleterious effects on the liver . However, low exposure to this naturally occurring compound is expected and should not present a health risk. The Canadian Food Inspection Agency (CFIA) considered it important to examine coumarin levels in common products such as ground cinnamon, cinnamon-containing products and licorice-flavored products to ensure that they are safe for consumption. The Acceptable Daily Intake (ADI) for coumarin is set at 0.1 mg by the European Food Safety Authority (EFSA) in 2004 . Furthermore, coumarin derivatives remain a significant area of interest for organic and medicinal chemists, with a wide range of applications in fragrances, pharmaceuticals, and agrochemicals. They are essential scaffolds for numerous therapeutic compounds, including those with antimicrobial , antioxidant, and anti-inflammatory properties . In a similar vein, certain noncentrosymmetric coumarin derivatives find application in the realm of cutting-edge research in domains such as optical communications, optical computing, dynamic imaging, and data storage. These derivatives offer a plethora of advantageous characteristics, including the phenomenon of photoswitching. Extensive research has been conducted to identify novel materials with enhanced nonlinear optical (NLO) properties. This research has been conducted either experimentally or guided by theoretical calculations, with the objective of synthesizing more effective photon-manipulating materials .
In the present work, the first objective is to elucidate the synthesized structure through both spectroscopic analyses such as ESI-MS, FT-IR, 1H and 13C NMR analysis, and single-crystal X-ray diffraction (XRD) studies . The 3D structure, as determined by X-ray analysis, has been subjected to further analysis through the Multipurpose Crystallographic Tool PLATON . The second objective is to investigate frontier molecular orbital (FMO) energies and derive global reactivity indices in both gas and solvent phases, underpinned by density functional theory (DFT) implemented in Gaussian09 . Furthermore, natural bond orbital (NBO) charges and Fukui Functions (FFs) were calculated to determine the atomic charge distribution for the purpose of identifying electrophilic and nucleophilic areas of the Coumarin-6-yl acetate using the B3LYP/6-311++G(d,p) basis set of Gaussian or the Perdew-Wang (PWC) functional and DND (Double-numerical + d-DNP) basis set of the DMol3 module implemented in Materials Studio software .
2. Experimental and Theoretical Methods
2.1. Synthesis
The reaction is an O-acetylation of 6-hydroxycoumarin with acetic anhydride in the presence of diethyl ether as solvent and pyridine as base. In this reaction, we exploited the HSAB theory which recommends that in acylation with anhydride or aliphatic acyl groups RCO+, which are known to be soft acids, best results are obtained by using soft bases like pyridine .
Figure 1. Synthesis of the title compound. Reagents and conditions: Pyridine, Diethyl ether, room temperature, 3 h.
The synthesis of the studied compound is analogous to the method described by Abou et al. (2021) . A 100 mL round-necked flask equipped with a water condenser was employed in the successive steps of the reaction. The following reagents were introduced successively: dried diethyl ether (25 mL), acetic anhydride (0.65 mL; 6.17 mmol), and dried pyridine (2.35 mL; 4.7 molar equivalents). While stirring vigorously, 6-hydroxycoumarin (1 g; 6.17 mmol) was added in small portions over 30 minutes. The reaction mixture was left to stir at room temperature for 3 hours. The mixture was then poured in a separating funnel containing 40 mL of chloroform and washed with diluted hydrochloric acid solution until the pH was 2–3. The organic phase was extracted, washed with water to neutrality, dried over magnesium sulfate (MgSO4) and the solvent removed. The resulting precipitate was filtered off with suction, washed with hexane and recrystallized from chloroform to obtain colorless needle-like crystals of the title compound: yield 72%; M.p. 371-373 K.
2.2. Spectroscopic Spectra Collection
The spectra were collected using the equipment used by abou et al. (2021) . The protocols are the same and are described as follows.
2.2.1. Electrospray Ionization Mass Spectrum
The analyses were conducted on a 3200 QTRAP spectrometer (Applied Biosystems SCIEX) furnished with a pneumatically abetted air pressure ionization (API) source for ESI-MS+ experiment. The sample in solution was ionized under the following conditions: electrospray tension (ISV): 5500 V; orifice tension (OR): 20 V; nebulizing gas pressure (air): 10 psi. The mass spectra (Figures 2 and 3) were attained with a quadrupole analyzer.
2.2.2. ATR-FTIR Spectrum
The infrared spectrum (Figure 4) was analyzed using a Bruker IFS 66/S Fourier Transform Infrared (FTIR) spectrometer, which was operated by the OPUS 6.5 software and employed the attenuated total reflectance (ATR) technique with a germanium tip. The absorption bands in the range 4000-400 cm-1 are expressed in wavenumber ῡ (cm-1): resolution 1 cm-1, 300 scans.
2.2.3. NMR Spectra
1H and 13C-NMR spectra (Figures 5 and 6) were performed on a Bruker AMX-400 spectrometer at 300 and 100 MHz respectively, utilizing TMS as internal standard (chemical shifts in δ ppm, coupling constants J in Hz) and deuterated chloroform (CDCl3) as a solvent.
The 13C spectrum was gained from an APT (Attached Proton Test) experiment.
Figure 2. Numbering of carbon atoms used in spectra analysis.
2.3. Crystal Structure Analysis
Diffraction intensities for Coumarin-6-yl Acetate were measured on Rigaku Oxford Diffraction SuperNova, Dual, Cu at zero, AtlasS2 diffractometer utilizing a mirror monochromator and Cu Kα radiation (λ = 1.54184 A) at 298 K. The structure was determined by direct methods utilizing SIR 2014 incorporated in the WinGX program suite. The refinement of the resolved structure was made by full-matrix least squares method on the positional and anisotropic temperature parameters of the non-hydrogen atoms, using 137 crystallographic parameters, with SHELXL2014 program . All H atoms were placed in calculated positions with [C—H = 0.93 Å (aromatic), 0.96 Å (methyl)] and refined using a riding model approximation with Uiso(H) constrained to 1.2 times Ueq(C-aromatic) or 1.5 times Ueq(C-methyl) of the respective parent atom. Data collection, cell refinement and data reduction are by CrysAlis PRO . The universal crystallographic tool PLATON was utilized to analyze the structure and present the results. Specifics information of the data acquisition conditions and the parameters of the refinement process are presented in Table 1.
2.4. Theoretical Computational Procedures
The optimized structure of the coumarin ester was performed via the density functional theory (DFT) implemented in the Gaussian 09W software package developed by Frisch and coworkers, using the job type Opt+Freq and restricted exchange correlation functional (RB3LYP) as calculation method with the 6-311++G(d,p) basis set in the ground state due to its accurate computational results on geometric and energetic parameters . Subsequent calculations are conducted using the M062X method with the 6-311++G(d,p) basis to estimate the electronic properties of this compound, such as ionization potential (I), electronic affinity (A), lowest unoccupied molecular orbital (LUMO), highest occupied molecular orbital (HOMO) and energy gap (ΔEg). In addition, Koopman's theorem for closed-shell molecules is used to compute the various global chemical reactivity descriptors in diverse types of solvents. Fukui functions, natural bond orbital (NBO) charges, and thermodynamic parameters can be calculated and discussed . All the output files originating from the calculations were visualized by the Gaussian View 06 program and the module DMol3 implemented in Materials Studio software.
3. Results and Discussion
3.1. Spectra Analysis
3.1.1. Interpretation of Electrospray Ionization Mass Spectrum
As illustrated in Figures 3 and 4, the peak observed at m/z 205 due to the pseudo-molecular ion [M+H]+ is consistent with the molecular weight of 204 g.mol-1, which is in accordance with the chemical formula C11H8O4.
ESI-MS m/z 205 ([M+H]+)
ESI-MS/MS m/z (%): 205 (MH+, 25), 163.2 (100), 135.2 (4.5), 119.0 (1.8 weak), 107.2 (3.6), 91.2 (1 weak), Figure 3.
3.1.2. Infrared Spectrum
For the studied compound, the FTIR spectrum showed absorption bands at 3224.1 cm-1 (C-H, aromatic), 2982.7 cm-1 (C-H, aliphatic), 1741.4 cm-1 and 1672.4 cm-1 for the two carbonyls, 1241.4 cm−1 (COC, lactone), and 1086.2 cm-1 (COC, ester). C=C signals were in the range of 1448.3 cm−1 to 1639.7 cm−1, (Figure 5) .
3.1.3. 1H-NMR Spectrum
The analysis (chemical shifts and coupling constants) of the 1H NMR spectrum (Figure 6) highlighted five spots, four of which were in the range 6-8.5 ppm and due to aromatic hydrogens. The three equivalent methyl protons are clearly visible at 2.3 ppm.
1H-NMR (CDCl3, 400 MHz, δ ppm): 7.6 (d, 1H, J = 9.6 Hz, H-4); 7.4 (m, 1H, H-5); 7.3 (m, 2H, H-7 and H-8); 6.4 (d, 1H, J = 9.6 Hz, H-3); 2.3 (s, 3H, H-12).
3.1.4. 13C (APT)-NMR Spectrum
In NMR studies, an APT sequence is employed to detect attached protons: CH3 and CH signals are positive, whilst CH2 and quaternary carbons signals are negative.
As illustrated in Figure 7, the APT spectrum of the molecule consists of 11 signals, in line with expectations. Six positive peaks were observed, suggesting the presence of five aromatic tertiary carbons and the shielded primary carbon of a methyl group. In contrast, five peaks were inverted, indicating quaternary carbons (C-2, C-6, C-9, C-10 and C-11).
13C (APT)-NMR (CDCl3, 100 MHz, δ ppm): 169.5 (C-11), 160.6 (C-2), 151.8 (C-6), 146.9 (C-9), 142.9 (C-4), 125.5 (C-8), 120.3 (C-7), 119.4 (C-10), 118.1 (C-5), 117.7 (C-3), 21.2 (C-12).
3.1.5. Heteronuclear Single-Quantum Correlation (HSQC) NMR Spectrum
As illustrated in Figure 8, the HSQC-NMR spectrum displayed six peaks demonstrating a strong correlation between the primary carbon C-12 and the methyl protons, as well as between each tertiary carbon (C-3, C-4, C-5, C-7, C-8) and the proton directly attached via 1JC-H scalar coupling. The spots obtained through this analysis confirm the previously identified signals in the 1H and 13C (APT)-NMR spectra.
3.1.6. Conclusion of Spectra Analysis
The results of the spectrometric analysis, when superimposed, corroborate the molecule illustrated in Figure 2. Other investigative methods, including X-ray and theoretical calculations, have been employed to validate this conclusion .
Figure 3. Electrospray ionization mass spectrum of the studied sample.
Figure 4. MS/MS spectrum of the protonated molecular ion peak (MH+) at m/z 205.
Figure 5. Experimental ATR-FTIR Spectrum.
Figure 6. Experimental 1H-NMR Spectrum: CDCl3, 300 MHz.
Figure 7. Experimental 13C (APT)-NMR Spectrum: CDCl3, 100 MHz.
Figure 8. Experimental HSQC Spectrum: CDCl3, 1H-NMR 300 MHz; 3C (APT)-NMR 100 MHz.
3.2. Structural Description
3.2.1. Structural Commentary
The structural configuration of the title compound (Figure 9) is characterized by a planar conformation of the coumarin ring system, as evidenced by the Puckering analysis parameter tau (τ=0.7°), which is less than 5°, the maximum value considered indicative of an effective planar conformation .
A thorough analysis of the bond lengths indicates a slightly uneven distribution around the pyrone ring. The C2-C3 bond length (1.340(2) Å) and the C1-C2 bond length (1.451(2) Å) exhibit a shorter and longer deviation, respectively, from the expected Car-Car bond length. This finding suggests that the electron density is less concentrated in the C2-C3 bond of the pyran-2-one ring, which results in the formation of a double bond, as reported in other coumarin ester derivatives .
Figure 9. An ORTEP view of the title compound (I) with the atomic numbering scheme. Displacement ellipsoids are shown at the 50% probability level.
Table 1. Crystal data and details of the structure determination.

chemical formula

C11H8O4

Theta range for data collection [°]

7.1 - 76.2

Formula weight

204.17

Crystal size [mm3]

0.28× 0.28 × 0.04

Temperature [K]

298

Index ranges

-4 ≤ h ≤4; -46 ≤ k ≤ 47; -8 ≤ l ≤ 8

Wavelength λ [Å]

1.54184

Reflections collected

8017

Crystal system

Monoclinic

Absorption coefficient [mm-1]

0.96

Space group

P21/n

Theta full [°]

67.684

Unit cell dimensions

F(000)

424

a [Å]

3.9039(2)

Refinement method

Full-matrix least squares on F2

b [Å]

37.5379(12)

Data/restraints/parameters

1712/0/ 136

c [Å]

6.4621(3)

Goodness of fit

1.03

α [°]

90

Final R indices [F2 > 2.0 σ(F2)]

R1 = 0.055, wR1= 0.164

β [°]

103.726(4)

Density calculated [g.cm-3]

1.265

γ [°]

90

Independent reflections

1896

Volume [Å3]

919.94(7)

Rint

0.032

Z

4

R indices (all data)

0.0574

Crystal description

plate

Δρmax, Δρmin (e Å−3)

0.28, -0.17

crystal color

Colorless

(Δ/σ)max

< 0.001

Diffractometer

SuperNova, Dual, Cu at zero, AtlasS2

Absorption correction

multi-scan;

CrysAlisPro 1.171.42.79a (Rigaku Oxford Diffraction, 2022)

Empirical absorption correction using spherical harmonics, implemented in SCALE3 ABSPACK scaling algorithm.

3.2.2. Supramolecular Features
The crystal structure reveals the generation of R228 dimeric units via C—H···O interactions. These dimers are linked together by non-classical hydrogen bonds through C—H···O, forming chains in the bc plane (Figure 10). Similarly, aromatic π-π stacking interactions have been observed between adjacent coumarin and pyrone rings, with centroid-centroid distances measuring less than 3.8 Å. This distance is considered to be the maximum threshold for effective π-π interaction . These aromatic interactions are present and link the dimers (Figure 11), (Table 3). The collective contribution of these molecular interactions is instrumental in ensuring the stable assembly of three-dimensional crystals.
For convenience, we have provided a summary of the perpendicular distances of Cg(I) on the J ring and the distances between Cg(I) and the perpendicular projection of Cg(J) on the I ring (slip) in Table 3.
Table 2. Hydrogen-bond geometry (Å, °).

D—H…A

D—H

H…A

D…A

D—H…A

C2—H5…O2i

0.93

2.51

3.397 (2)

160

C11—H11C…O4ii

0.96

2.58

3.378 (2)

141

Symmetry code: (i) 2−x, −y, −z; (ii) -1/2+x, 1/2-y, 1/2+z.
Figure 10. A view of the crystal packing, showing C—H···O hydrogen bonds linking molecules into R228 dimeric units and their propagation into the bc plane.
Figure 11. π···π stacking interactions in the crystal packing.
Table 3. Analysis of short ring interactions (Å). Cg1 and Cg3 are the centroids of the pyrone and the coumarin rings, respectively. The distances between the centroid of ring I and its perpendicular projection on ring J, as well as the distances between the centroid of ring I and the perpendicular projection of ring J on ring I (slippage), are reported.

Cg(I)

Cg(J)

Symmetry Cg(J)

Cg(I)…Cg(J)

CgI_Perp

CgJ_Perp

Slippage

Cg1

Cg3

1+x, y, z

3.7192(8)

-3.4080 (6)

3.4147 (5)

1.474

Cg2

Cg3

-1+x, y, z

3.7226(8)

3.4061(6)

-3.4004(5)

1.515

Table 4. Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2).

Atom

x

y

z

Ueq

O1

0.5293 (3)

0.08275 (3)

0.03878 (15)

0.0436 (3)

O2

0.6298 (4)

0.03526 (3)

−0.13667 (18)

0.0603 (4)

O3

0.7013 (3)

0.79975 (19)

0.76450 (16)

0.0434 (3)

O4

0.61611 (4)

0.21027 (3)

0.6250 (2)

0.0677 (4)

C1

0.6875 (4)

0.04991 (4)

0.0338 (2)

0.0429 (4)

C2

0.9057 (4)

0.03680 (4)

0.2330 (2)

0.0446 (4)

C3

0.9557 (4)

0.05577 (4)

0.4134 (2)

0.0415 (4)

C4

0.7911 (4)

0.09011 (3)

0.4138 (2)

0.0348 (3)

C5

0.5812 (4)

0.10251 (3)

0.2228 (2)

0.0360 (3)

C6

0.4105 (4)

0.13520 (4)

0.2078 (2)

0.0420 (4)

C7

0.4547 (4)

0.15620 (4)

0.3872 (2)

0.0414 (4)

C8

0.6686 (4)

0.14433 (3)

0.5779 (2)

0.0367 (3)

C9

0.8354 (4)

0.11183 (4)

0.5949 (2)

0.0376 (3)

C10

0.8350 (4)

0.19810 (4)

0.7676 (2)

0.0424 (4)

C11

0.8240 (5)

0.21674 (5)

0.9701 (3)

0.0575 (5)

Table 5. Experimental and DFT/ RB3LYP/6-311++G(d,p) calculated bond lengths in (Å) for compound (I).

Bond

X-Ray

Calc.(gas)

Bond

X-Ray

Calc.(gas)

Bond

X-Ray

Calc.(gas)

O1—C5

1.375 (2)

1.366

C5—C4

1.390 (2)

1.404

C8—C9

1.375 (2)

1.380

O1—C1

1.383 (2)

1.391

C4—C9

1.403 (2)

1.406

C9—C4

1.403 (2)

1.406

O3—C10

1.358 (2)

1.374

C3—C2

1.340 (2)

1.350

O2—C1

1.204 (2)

1.207

O4—C10

1.192 (2)

1.202

C4—C3

1.441 (2)

1.440

C10—C11

1.494 (2)

1.502

O3—C8

1.407 (2)

1.398

C6—C7

1.379 (2)

1.387

C1—C2

1.451 (2)

1.456

C5—C6

1.389 (2)

1.394

C7—C8

1.388 (2)

1.397

Table 6. Experimental and DFT/ RB3LYP/6-311++G(d,p) calculated bond angles (°) for compound (I).

Bond angle

X-Ray

Calc.(Gas)

Bond angle

X-Ray

Calc.(Gas)

Bond angle

X-Ray

Calc.(Gas)

C5—O1—C1

121.72 (11)

122.85

O1—C5—C4

121.53 (12)

121.37

C2—C3—C4

120.19 (13)

120.71

C10—O3—C8

117.99 (11)

119.20

C6—C5—C4

121.99 (13)

121.10

O2—C1—O1

116.16 (14)

115.87

C8—C9—C4

119.07 (12)

119.82

C6—C7—C8

119.46 (13)

119.75

O2—C1—C2

126.91 (14)

126.22

C5—C4—C9

118.47 (13)

118.72

C3—C2—C1

121.76 (14)

121.76

O1—C1—C2

116.93 (13)

115.87

C5—C4—C3

117.86 (12)

117.44

C9—C8—C7

122.06 (13)

121.12

O4—C10—O3

123.35 (14)

123.52

C9—C4—C3

123.67 (12)

123.83

C9—C8—O3

117.82 (12)

117.67

O4—C10—C11

125.86 (15)

125.72

C7—C6—C5

118.95 (13)

119.49

C7—C8—O3

119.99 (13)

121.09

O3—C10—C11

110.79 (13)

110.76

O1—C5—C6

116.48 (12)

117.53

Table 7. Experimental and DFT/ RB3LYP/6-311++G(d,p) calculated dihedral angles (°) for compounds (I).

Dihedral angles

X-Ray

Calc.(Gas)

Dihedral angles

X-Ray (I)

Calc.(Gas)

C8—C9—C4—C5

0.5 (2)

0.17

C6—C7—C8—C9

−1.0 (2)

-0.23

C8—C9—C4—C3

179.43 (13)

-179.89

C6—C7—C8—O3

−176.70 (13)

-176.10

C1—O1—C5—C6

179.51 (13)

-179.97

C10—O3—C8—C9

123.21 (15)

120.51

C1—O1—C5—C4

−1.0 (2)

-0.090

C10—O3—C8—C7

−60.88 (19)

-63.48

C7—C6—C5—O1

−179.67 (13)

179.82

C1—C2—C3—C4

0.1 (2)

-0.08

C7—C6—C5—C4

0.9 (2)

-0.06

C5—C4—C3—C2

0.1 (2)

0.05

C9—C4—C5—O1

179.35 (12)

179.98

C9—C4—C3—C2

−178.86 (14)

-179.89

C3—C4—C5—O1

0.3 (2)

0.04

C5—O1—C1—O2

−178.57 (14)

-179.99

C9—C4—C5—C6

−1.2 (2)

-0.15

C5—O1—C1—C2

1.2 (2)

0.05

C3—C4—C5—C6

179.74 (13)

179.91

C3—C2—C1—O2

178.99 (17)

-179.91

C5—C6—C7—C8

0.2 (2)

0.25

C3—C2—C1—O1

−0.8 (2)

0.04

C4—C9—C8—C7

0.6 (2)

0.02

C8—O3—C10—O4

−4.4 (2)

-0.41

C4—C9—C8—O3

176.43 (12)

176.03

C8—O3—C10—C11

175.44 (13)

179.69

3.3. Theoretical Calculations
In this study, we will compare the geometrical parameters obtained from theoretical calculations to those originating from x-ray crystallography to check their agreement before estimating the chemical reactivity descriptors.
3.3.1. Comparison of Geometrical Parameters
The geometrical parameters of (I) obtained from theoretical calculations are then confronted with those yielded from the X-ray crystallographic study. The comparison demonstrates a high level of consistency between bond lengths and bond angles, as evidenced by root mean square deviations (RMSD) of 0.01 Å for bond lengths and 0.67° for bond angles (Tables 5 and 6). Additionally, analysis of the calculated torsion angles confirms the planarity of the coumarin moiety (Table 7).
The above comparisons demonstrate that the theoretical calculations are in good alignment with the crystallographic predictions based on solid-state structure. Therefore, the calculated structure can be utilized to compute the chemical properties of (I) in solid state.
3.3.2. Molecular Electrostatic Potential (MEP)
Electrostatic interactions in a range of chemical systems have been extensively described by the MEP . The force exerted on a positive test charge (a proton) positioned above the charge cloud produced by the molecule's electrons and nuclei at any given time is known as the molecular electrostatic potential. Equation (1) was utilized to compute the values for the current system as previously mentioned .
Vr= AZARA-r-ρr'd3r'r'-r (1)
Where ZA denotes the charge of nucleus A situated in RA, whilst ρ(r') signifies the electron density function of the molecule and r' serves as the dummy integration variable.
Figure 12 shows the color visualizations of the calculations' outcomes in the gaseous state of (I) and in various solvents. Blue coloration indicates higher positive potential, which is advantageous for nucleophilic attack, while red coloration indicates higher negative potential, which is advantageous for electrophilic attack. The figure shows two possible locations for electrophilic attack on the compound in gaseous and solvent media. The negative regions are concentrated on the oxygen atoms O2 and O4, with maximum values of -0.0574 a.u. (gas), -0.0721 a.u. (water), -0.0692 a.u. (THF), -0.0640 a.u. (benzene), and -0.0651 a.u. (toluene). The presence of the intermolecular C2—H5⋅⋅⋅O2 [2−x, −y, −z] and C11—H11C⋅⋅⋅O4 [−1/2+x, 1/2−y, 1/2+z] hydrogen bonds is thus confirmed. As for the remaining atoms in the coumarin nucleus, their light blue environment suggests that they constitute electropositive weakening areas.
Figure 12. MEP map (in atomic units) calculated using DFT/ RM062X /6-311++G(d,p).
3.3.3. Reactivity Descriptors
(i). Global Descriptors
A number of theoretical descriptors, which are related to the conceptual DFT, are utilized in order to determine chemical reactivity forecasts. These include electronegativity (χ), which indicates a molecule's ability to retain its electrons; global softness (σ), which indicates a system's resistance to changes in its electron number; and the overall electrophilicity index (𝝎), which indicates a molecule's electrophilic power. The least vacant molecular orbital energy (ELUMO) is also of significance in this regard, as it describes the sensitivity of the molecule to nucleophilic attack, whilst the highest occupied molecular orbital energy (EHOMO) is important in terms of its description of the molecule's sensitivity to electrophilic attack. The following equations are utilized to calculate these parameters :
Eg=ELUMO-EHOMO(2)
I=-EHOMO (3)
A=-ELUMO (4)
χ=-μ=-ELUMO+EHOMO2 (5)
η=ELUMO - EHOMO2 (6)
ω=μ22η (7)
σ=12η (8)
Table 8 presents the results of the calculated energetic parameters, which demonstrate that compound (I) in these mixtures are stables. However, compound (I) in the benzene solvent is the most stable with an energy gap value of 9.48 eV, whereas (I) in the gas phase remains the least stable with the lowest energy gap value of 6.64 eV . This is also backed up by the fact that Gaussian did not identify any negative frequencies in its frequency calculation.
Table 8. Global reactivity descriptors calculated with the DFT/ M062X/6-311++G(d,p) method.

Gas

Water

THF

Benzene

Toluene

ELUMO (eV)

-1.48

-1.51

0.74

0.75

-1.51

EHOMO (eV)

-8.12

-8.24

-8.72

-8.73

-8.20

I (eV)

8.12

8.24

8.72

8.73

8.20

A (eV)

1.48

1.51

-0.74

-0.75

1.51

𝜒 (eV)

4.80

4.88

3.99

3.99

4.86

(eV)

-4.80

-4.88

-3.99

-3.99

-4.86

𝜂 (eV)

3.32

3.37

4.73

4.74

3.35

(eV-1)

0.151

0.149

0.106

0.105

0.149

ω (eV)

3.470

3.531

1.683

1.679

3.523

ΔEg (eV)

6.64

6.73

9.46

9.48

6.69

For convenience, Table 8 also shows the overall chemical reactivity indices. From these results, it can be seen that compound (I) in gaseous phase has the lowest overall hardness value (η = 3.32eV). It can therefore be considered as the softest of the series. In contrast, compound (I) in water gives the highest value of electronegativity (χ = 4.88 eV) and electrophilicity index (ω = 3.531 eV), indicating that it is more electron accepting than the Coumarin-6-yl acetate derivative in other solvents.
Figure 13. Calculated HOMO and LUMO orbital distributions and energy levels for the molecule (I) in solvents.
(ii). Local Descriptors and Dual Descriptors
Various indices have been used to distinguish between the reactive behaviors of the atoms that make up a molecule. These are local and dual descriptors of reactivity at the local level. In this case, the Fukui function fk+,fk- , the local softness σk+,σk-, the local electrophilic power ωk+,ωk-, and the dual descriptors have been used to elucidate the electrophilic and nucleophilic selectivity of the molecule. It is essential to note that the Fukui function fk+ quantifies reactivity when the molecule is subjected to nucleophilic attack, whereas the Fukui function fk- elucidates the electrophilic reactivity of a specific site. The uppermost Fukui function value is attributed to the greatest active location. The condensed indices σk+ and ωk+ indicate a site's capacity to gain electron density through nucleophilic attack, while the descriptors σk- and ωk- reflect a site's ability to produce electron density through electrophilic attack. In terms of the dual descriptor, it is an effective tool for forecasting efficiency and identifying issues related to regioselectivity. In fact, a positive dual descriptor indicates a position that is probably to accept electron density, making it further electrophilic. Contrarywise, a negative double descriptor suggests a location capable of generating electron density, reinforcing its nucleophilic character. A site with a dual descriptor value near zero indicates an equal ability to accept and provide electronic density. The local descriptor values are calculated using the following equations :
fk+=qkN+1-qkN (9)
fk-=qkN-qkN-1(10)
σk+=σfk+ (11)
σk-=σfk- (12)
ωk+ =ωfk+(13)
ωk- =ωfk- (14)
ηk+=ηfk+ (15)
ηk-=ηfk- (16)
Where:
qkN is the electron population of the atom k in the neutral molecule.
qkN+1 is the electron population of the atom k in the anionic molecule.
qkN-1 is the electron population of the atom k in the cationic molecule.
The values of the dual descriptors are obtained through the subsequent equations .
f=fk+-fk- (17)
σ=σk+-σk- (18)
ω=ωk+-ωk- (19)
All the results are summarized in tables 9-12 below
Table 9. NBO charges for (I) computed with DFT/ RM062X/6-311++G(d,p) method in gas and solvent phases.

Atoms

(I) in gas phase

(I) in water

(I) in tetrahydrofuran (THF)

q(N)

q(N-1)

q(N+1)

q(N)

q(N-1)

q(N+1)

q(N)

q(N-1)

q(N+1)

O1

-0.53336

-0.46999

-0.31181

-0.53639

-0.20537

-0.32012

-0.53563

-0.47618

-0.31842

O2

-0.56065

-0.39709

-0.38437

-0.63017

-0.13050

-0.41971

-0.61698

-0.45751

-0.41318

O3

-0.58162

-0.55533

-0.29229

-0.57929

-0.27335

-0.29221

-0.58658

-0.57051

-0.29603

O4

-0.57973

-0.55064

-0.29714

-0.62658

-0.30323

-0.31862

-0.61039

-0.59029

-0.31042

C1

0.77962

0.74965

0.34921

0.79262

0.34976

0.32145

0.79033

0.76789

0.32745

C2

-0.30309

-0.18428

-0.34109

-0.31474

0.04652

-0.31964

-0.31245

-0.16755

-0.32442

C3

-0.11439

-0.12996

-0.25830

-0.09152

-0.09878

-0.29738

-0.09644

-0.10713

-0.29016

C4

-0.14627

-0.04456

-0.06660

-0.14306

0.08937

-0.05036

-0.14343

-0.02250

-0.05277

C5

0.35491

0.47528

0.12460

0.35158

0.36138

0.10857

0.35223

0.48672

0.11184

C6

-0.22437

-0.21621

-0.13673

-0.22777

-0.12955

-0.12445

-0.22677

-0.22181

-0.12674

C7

-0.19454

-0.12251

-0.26101

-0.19893

0.00942

-0.24761

-0.19814

-0.10453

-0.25041

C8

0.28574

0.43492

0.16657

0.28051

0.30544

0.16885

0.28301

0.41843

0.16947

C9

-0.19124

-0.21069

-0.21340

-0.18776

-0.14883

-0.21011

-0.18824

-0.20544

-0.21095

C10

0.83414

0.83262

0.41574

0.85488

0.43029

0.42756

0.85026

0.85070

0.42499

C11

-0.68336

-0.68734

-0.33992

-0.68700

-0.34308

-0.34353

-0.68512

-0.68437

-0.34249

Table 9. Continued.

Atoms

(I) in benzene

(I) in toluene

q(N)

q(N-1)

q(N+1)

q(N)

q(N-1)

q(N+1)

O1

-0.53436

-0.19076

-0.31548

-0.60952

-0.19175

-0.58218

O2

-0.59137

-0.10870

-0.40012

-0.69016

-0.09898

-0.58218

O3

-0.58792

-0.27062

-0.29643

-0.65029

-0.27263

-0.58575

O4

-0.59214

-0.28372

-0.30178

-0.69248

-0.29096

-0.61261

C1

0.78555

0.33982

0.33826

0.92877

0.33413

0.74432

C2

-0.30818

0.02414

-0.33297

-0.34388

0.03413

-0.45571

C3

-0.10521

-0.10344

-0.27590

-0.03853

-0.10174

-0.28178

C4

-0.14464

0.07342

-0.05837

-0.17075

0.07665

-0.12881

C5

0.35367

0.35840

0.11716

0.40352

0.35440

0.31055

C6

-0.22525

-0.11956

-0.13078

-0.22442

-0.12234

-0.25998

C7

-0.19600

-0.00791

-0.25553

-0.16227

-0.00285

-0.32170

C8

0.28538

0.32426

0.16892

0.30069

0.31554

0.28041

C9

-0.18950

0.32426

-0.21160

-0.15429

-0.15038

-0.27529

C10

0.84209

0.42322

0.42038

0.97617

0.42385

0.83905

C11

-0.68346

-0.34198

-0.34129

-0.61559

-0.34330

-0.68410

Table 10. Reactivity descriptors calculated with natural charges for compound (I).

Atoms

Local descriptors

Dual descriptors

f-

f+

σ-

σ+

η-

η+

ω-

ω+

f

σ

Δω

O1

-0.06337

0.22155

-0.00957

0.03345

-0.21039

0.73555

-0.21989

0.76878

0.28492

0.04302

0.98867

O2

-0.16356

0.17628

-0.02470

0.02662

-0.54302

0.58525

-0.56755

0.61169

0.33984

0.05132

1.17924

O3

-0.02629

0.28933

-0.00397

0.04369

-0.08728

0.96058

-0.09123

1.00398

0.31562

0.04766

1.09520

O4

-0.02909

0.28259

-0.00439

0.04267

-0.09658

0.93820

-0.10094

0.98059

0.31168

0.04706

1.08153

C1

0.02997

-0.43041

0.00453

-0.06499

0.09950

-1.42896

0.10400

-1.49352

-0.46038

-0.06952

-1.59752

C2

-0.11881

-0.038

-0.01794

-0.00574

-0.39445

-0.12616

-0.41227

-0.13186

0.08081

0.01220

0.28041

C3

0.01557

-0.14391

0.00235

-0.02173

0.05169

-0.47778

0.05403

-0.49937

-0.15948

-0.02408

-0.55340

C4

-0.10171

0.07967

-0.01536

0.01203

-0.33768

0.26450

-0.35293

0.27645

0.18138

0.02739

0.62939

C5

-0.12037

-0.23031

-0.01818

-0.03478

-0.39963

-0.76463

-0.41768

-0.79918

-0.10994

-0.01660

-0.38149

C6

-0.00816

0.08764

-0.00123

0.01323

-0.02709

0.29096

-0.02832

0.30411

0.0958

0.01447

0.33243

C7

-0.07203

-0.06647

-0.01088

-0.01004

-0.23914

-0.22068

-0.24994

-0.23065

0.00556

0.00084

0.01929

C8

-0.14918

-0.11917

-0.02253

-0.01799

-0.49528

-0.39564

-0.51765

-0.41352

0.03001

0.00453

0.10413

C9

0.01945

-0.02216

0.00294

-0.00335

0.06457

-0.07357

0.06749

-0.07690

-0.04161

-0.00628

-0.14439

C10

0.00152

-0.4184

0.00023

-0.06318

0.00505

-1.38909

0.00527

-1.45185

-0.41992

-0.06341

-1.45712

C11

0.00398

0.34344

0.00060

0.05186

0.01321

1.14022

0.01382

1.19172

0.33946

0.05126

1.17790

Table 11. Reactivity descriptors calculated with natural charges for compound (I).in water.

Atoms

Local descriptors

Dual descriptors

f-

f+

σ-

σ+

η-

η+

ω-

ω+

f

σ

Δω

O1

-0.33102

0.21627

-0.04932

0.03222

-1.11554

0.72883

-1.16883

0.76365

0.54729

0.08155

1.93248

O2

-0.49967

0.21046

-0.07445

0.03136

-1.68389

0.70925

-1.76433

0.74313

0.71013

0.10581

2.50747

O3

-0.30594

0.28708

-0.04559

0.04277

-1.03102

0.96746

-1.08027

1.01368

0.59302

0.08836

2.09395

O4

-0.32335

0.30796

-0.04818

0.04589

-1.08969

1.03783

-1.14175

1.08741

0.63131

0.09407

2.22916

C1

0.44286

-0.47117

0.06599

-0.07020

1.49244

-1.58784

1.56374

-1.66370

-0.91403

-0.13619

-3.22744

C2

-0.36126

-0.0049

-0.05383

-0.00073

-1.21745

-0.01651

-1.27561

-0.01730

0.35636

0.05310

1.25831

C3

0.00726

-0.20586

0.00108

-0.03067

0.02447

-0.69375

0.02564

-0.72689

-0.21312

-0.03175

-0.75253

C4

-0.23243

0.0927

-0.03463

0.01381

-0.78329

0.31240

-0.82071

0.32732

0.32513

0.04844

1.14803

C5

-0.0098

-0.24301

-0.00146

-0.03621

-0.03303

-0.81894

-0.03460

-0.85807

-0.23321

-0.03475

-0.82346

C6

-0.09822

0.10332

-0.01463

0.01539

-0.33100

0.34819

-0.34681

0.36482

0.20154

0.03003

0.71164

C7

-0.20835

-0.04868

-0.03104

-0.00725

-0.70214

-0.16405

-0.73568

-0.17189

0.15967

0.02379

0.56379

C8

-0.02493

-0.11166

-0.00371

-0.01664

-0.08401

-0.37629

-0.08803

-0.39427

-0.08673

-0.01292

-0.30624

C9

-0.03893

-0.02235

-0.00580

-0.00333

-0.13119

-0.07532

-0.13746

-0.07892

0.01658

0.00247

0.05854

C10

0.42459

-0.42732

0.06326

-0.06367

1.43087

-1.44007

1.49923

-1.50887

-0.85191

-0.12693

-3.00809

C11

-0.34392

0.34347

-0.05124

0.05118

-1.15901

1.15749

-1.21438

1.21279

0.68739

0.10242

2.42717

Table 12. Reactivity descriptors calculated with natural charges for compound (I).in toluene.

Atoms

Local descriptors

Dual descriptors

f-

f+

σ-

σ+

η-

η+

ω-

ω+

f

σ

Δω

O1

-0.41777

0.02734

-0.06225

0.00407

-1.39953

0.09159

-1.47180

0.09632

0.44511

0.06632

1.56812

O2

-0.59118

0.10798

-0.08809

0.01609

-1.98045

0.36173

-2.08273

0.38041

0.69916

0.10417

2.46314

O3

-0.37766

0.06454

-0.05627

0.00962

-1.26516

0.21621

-1.33050

0.22737

0.4422

0.06589

1.55787

O4

-0.40152

0.07987

-0.05983

0.01190

-1.34509

0.26756

-1.41455

0.28138

0.48139

0.07173

1.69594

C1

0.59464

-0.18445

0.08860

-0.02748

1.99204

-0.61791

2.09492

-0.64982

-0.77909

-0.11608

-2.74473

C2

-0.37801

-0.11183

-0.05632

-0.01666

-1.26633

-0.37463

-1.33173

-0.39398

0.26618

0.03966

0.93775

C3

0.06321

-0.24325

0.00942

-0.03624

0.21175

-0.81489

0.22269

-0.85697

-0.30646

-0.04566

-1.07966

C4

-0.2474

0.04194

-0.03686

0.00625

-0.82879

0.14050

-0.87159

0.14775

0.28934

0.04311

1.01934

C5

0.04912

-0.09297

0.00732

-0.01385

0.16455

-0.31145

0.17305

-0.32753

-0.14209

-0.02117

-0.50058

C6

-0.10208

-0.03556

-0.01521

-0.00530

-0.34197

-0.11913

-0.35963

-0.12528

0.06652

0.00991

0.23435

C7

-0.15942

-0.15943

-0.02375

-0.02376

-0.53406

-0.53409

-0.56164

-0.56167

-1E-05

0.00000

-0.00004

C8

-0.01485

-0.02028

-0.00221

-0.00302

-0.04975

-0.06794

-0.05232

-0.07145

-0.00543

-0.00081

-0.01913

C9

-0.00391

-0.121

-0.00058

-0.01803

-0.01310

-0.40535

-0.01377

-0.42628

-0.11709

-0.01745

-0.41251

C10

0.55232

-0.13712

0.08230

-0.02043

1.85027

-0.45935

1.94582

-0.48307

-0.68944

-0.10273

-2.42890

C11

-0.27229

-0.06851

-0.04057

-0.01021

-0.91217

-0.22951

-0.95928

-0.24136

0.20378

0.03036

0.71792

The local and dual descriptors of compound (I), calculated at the M062X /6-311++G(d,p) level, indicate that the carbon atom C1 of compound (I) in water, in a sp2 hybridization state, is the preferred site for electrophile attack, with a Δω value of -3.22744. Based on the aforementioned level of calculation, a nucleophilic attack is predicted to occur on the O2 atom with a value of Δω = 2.50747. These local and dual descriptors results refine those obtained with Fukui functions, thanks to their precision .
3.3.4. Thermodynamic Parameters
The thermodynamic properties or parameters value (ΔS, ΔH and ΔG) provides insight into the spontaneity, randomness, and exothermic or endothermic nature of a process. To achieve these parameters, we need to know the entropy, enthalpy and free energy of the different compounds (reactants and products) of the synthesis reaction. For this reason, we have computed these standard thermodynamic quantities or parameters namely entropy S, heat capacity Cp, the enthalpy H and the free energy G from 25 K to 1000 K in steps of 25 K with the Perdew-Wang (PWC) functional and DND (Double-numerical + d-DNP) basis set of the DMol3 module implemented in Materials Studio software. The results are summarized in table 13 and plotted in figure 14.
Table 13. PWC/DND calculated of standard thermodynamic quantities (S(Cal.K-1.mol-1), Cp(Cal.K-1.mol-1), H((kcal.mol-1), G(kcal.mol-1)) for compound (I).

T(K)

S

Cp

H

G

25

55.757

11.433

103.617

102.223

50

65.087

15.67

103.959

100.705

75

72.048

18.801

104.391

98.987

100

77.864

21.798

104.898

97.112

125

83.064

24.956

105.482

95.099

150

87.907

28.308

106.148

92.962

175

92.533

31.822

106.899

90.706

200

97.018

35.452

107.74

88.336

225

101.407

39.149

108..672

85.856

250

105..724

42.868

109.697

83.266

275

109.984

46.563

110.815

80.57

298.15

113.882

49.932

111.932

77.768

300

114.192

50.198

112.025

77.768

325

118.35

53.741

113.324

74.861

350

122.459

57.169

114.711

71.85

375

126.516

60.463

116.182

68.738

400

130.52

63.613

117.733

65.525

425

134.467

66.614

119.361

62.213

450

138.356

69.464

121.063

58.802

475

142.185

72.165

122.833

55.295

500

145.952

74.721

124.67

51.693

525

149.657

77.138

126.568

47.998

550

153.299

79.422

128.525

44.211

575

156.877

81.582

130.538

40.334

600

160.393

83.624

132.603

36.368

625

163.846

85.556

134.718

32.315

650

167.238

87.385

136.88

28.176

675

170.568

89.119

139.087

23.953

700

173.839

90.763

141.336

19.648

725

177.052

92.323

143.624

15.262

750

180.207

93.806

145.951

10.796

775

183.306

95.216

148.314

6.252

800

186.35

96.557

150.711

1.631

825

189.341

97.835

153.141

-3.065

850

192.28

99.053

155.603

-7.836

875

195.168

100.215

158.094

-12.679

900

198.007

101.324

160.613

-17.594

925

200.798

102.383

163.159

-22.579

950

203.542

103.395

165.732

-27.633

975

206.24

104.362

168.329

-32.755

1000

208.894

105.288

170.949

-37.945

Figure 14. Thermodynamic property curves with entropy in cal/mol, heat capacity in cal/mol.K, enthalpy and free energy in kcal/mol.
4. Conclusion
In this paper, spectroscopic and X-ray crystallographic methods were used to solve and refine the molecular structure of the title compound (I). The intermolecular interactions present in the structure, the geometrical parameters as well as the conformations were analyzed using the versatile crystallographic tool Platon . The resulting bond lengths, bond angles, and torsion angles were confronted to the corresponding calculated data. The comparatively analysis showed no substantial differences between the experimental and theoretical structures. Furthermore, we conducted an investigation into the molecular electrostatic potential and HOMO-LUMO analysis for Coumarin-6-yl acetate utilizing DFT/ M062X /6-311++G(d,p) calculations. The MEP maps highlighted the nucleophilic sites around the oxygen atoms O2 and O4 and are materialized by the zones marked in red color (negative potential zone), while positive potential sites were found in proximity to the hydrogen atoms. This information provides insight into the regions where intra- and intermolecular interactions could potentially be observed. Furthermore, the reactivity descriptors of (I) were estimated via PWC/DND methods. The overall descriptors show that compound (I) in the benzene solvent demonstrates optimal stability, exhibiting an energy gap value of 9.48 eV. In contrast, compound (I) in the gaseous phase exhibits the lowest energy gap value (ΔEg = 6.64 eV), indicating that it is the least stable among the five mixtures of (I) with the solvents. In addition, the local and dual descriptors of compound (I) indicate that the carbon atom C1 of compound (I) in water is the optimal site for electrophile attack, with a Δω value of -3.22744. Based on the aforementioned level of calculation, it is predicted that a nucleophilic attack will occur on the O2 atom with a value of Δω = 2.50747 that seem to the best prevision because of the best accuracy of the method . These findings, derived from the electrophilic or nucleophilic attack zone approach, will facilitate precise regulation of chemical reactions, thereby enabling the realization of desired molecular structures. The standard thermodynamic quantities or parameters namely entropy S, heat capacity Cp, the enthalpy H and the free energy G have also been performed from 25 K to 1000 K in steps of 25 K. The obtained results at 298.15K and 1 Pa are S = 113.882 Cal.K-1.mol-1, Cp = 49.932 Cal.K-1.mol-1, H = 111.932 kcal.mol-1 and G = 77.768 kcal.mol-1.
Abbreviations

ESI-MS

ElectroSpray Ionization Mass Spectrometry

FT-IR

Fourier Transform Infrared Spectroscopy

1H NMR

Proton Nuclear Magnetic Resonance

13C NMR

Carbon-13 (C13) Nuclear Magnetic Resonance

B3LYP

Becke, 3-parameter, Lee–Yang–Parr

THF

TetraHydroFuran

NBO

Natural Bond Orbital

HOMO

Highest Occupied Molecular Orbital

LUMO

Lowest unoccupied molecular orbital

CFIA

Canadian Food Inspection Agency

ADI

Acceptable Daily Intake

EFSA

European Food Safety Authority

NLO

Non-Linear Optical

XRD

X-ray Diffraction

3D

Three-Dimensional

FMO

Frontier Molecular Orbitals

FF

Fukui Function

PWC

Perdew-Wang

DND

Double-numerical + d

TMS

Tetramethylsilane

DFT

Density Functional Theory

HSQC

Heteronuclear Single-Quantum Correlation

ATR

Attenuated Total Reflectance

APT

Attached Proton Test

CDCl3

Deuterated Chloroform

RMSD

Root Mean Square Deviations

MEP

Molecular Electrostatic Potential

Acknowledgments
The authors thank the Spectropole Service of the Federation of Chemical Sciences (Aix-Marseille University, France) for conducting the complete analysis.
Author Contributions
Honoré Kouadio Yao: Funding acquisition, Validation, Writing – original draft
Zakaria Koulabiga: Validation, Synthesis
Akoun Abou: Conceptualization, Investigation, Methodology, Supervision, Validation, Writing – review & editing
Abdoulaye Djandé: Conceptualization, Project administration, Supervision, Validation, Writing – review & editing
Stéphane Coussan: Data curation, Formal Analysis, Project administration, Resources, Software, Supervision, Validation, Visualization
Olivier Ouari: Data curation, Formal Analysis, Project administration, Resources, Software, Validation, Visualization
Conflicts of Interest
The authors declare no conflicts of interest.
References
[1] Matos, M. J., Santana, L., Uriarte, E., Abreu, O. A., Molina, E., & Yordi, E. G. (2015). Coumarins — An Important Class of Phytochemicals. Phytochemicals - Isolation, Characterisation and Role in Human Health. (pp. 113-140). Royaune-Uni: IntechOpen.
[2] Zeng, L., Zhang, R.-Y., Meng, T., Lou, Z.-C. (1990). Determination of nine flavonoids and coumarins in licorice root by high-performance liquid chromatography. J. Chromatogr. A, 513, 247-254.
[3] Shojaii, A., Fard, M. H., 2012. Review of pharmacological properties and chemical constituents of Pimpella anisum. ISRN Pharmaceutics, 2012, 510795.
[4] Abraham, K., Wöhrlin, F., Lindtner, O., Heinemeyer, G., Lampen, A., 2010. Toxicology and risk assessment of coumarin: Focus on human data. Mol. Nutr. Food Res, 54(2), 228-239.
[5] Lake, B. G. (1999). Coumarin metabolism, toxicity and carcinogenicity: Relevance for human risk assessment. Food and Chemical Toxicology, 37(4), 423-453.
[6] Coumarin in flavourings and other food ingredients with flavouring properties. Scientific opinion of the panel on food additives, flavourings, processing aids and materials in contact with food (AFC). (2008). EFSA Journal, 793, 1-15.
[7] Basanagouda M, Kulkarni M V, Sharma D, Gupta V K, Pranesha P, Sandhyarani P and Rasal V P., (2009). J. Chem. Sci. 121, 485–495.
[8] Vuković N, Sukdolak S, Solujić S and Niciforović N, (2010). An efficient synthesis and antioxidant properties of novel imino and amino derivatives of 4-hydroxy coumarins. Arch. Pharm. Res. 33; 5–15.
[9] Emmanuel-Giota A A, Fylaktakidou K C, Litinas K E, Nicolaides D N and Hadjipavlou-Litina D J,. (2001). Synthesis and biological evaluation of several 3-(coumarin-4-yl) tetrahydroisoxazole and 3-(coumarin-4-yl) dihydropyrazole derivatives. Heterocycl. Chem.; 38: 717–722.
[10] D. R. Kanis, M. A. Ratner, T. J. Marks, (1994). Design and construction of molecular assemblies with large second-order optical nonlinearities. Quantum chemical aspects. Chem. Rev. 94, 195–242.
[11] D. J. Williams, (1992). Non-linear optical properties of organic materials. Thin Solid Films 216, 117–122
[12] P. J. Mendes, J. P. Prates Ramalho, A. J. E. Candeias, M. P. Robalo, M. H. Garcia, (2005). Density functional theory calculations on η 5 - monocyclopentadienylnitrilecobalt complexes concerning their second-order nonlinear optical properties. J. Mol. Struct. (Theochem.) 729 109–113
[13] N. M. F. S. A. Cerqueira, A. M. F. Oliveira-Campos, P. J. Coelho, L. H. Melo de Carvalho, A. Samat, R. Guglielmetti, (2005). Synthesis of Photochromic Dyes Based on Annulated Coumarin Systems. Helv. Chim. Acta 85, 442–450. ttps://doi.org/10.1002/1522-2675(200202)85:2<442::AID-HLCA442>3.0.CO;2-9
[14] S. P. G. Costa, J. Griffiths, G. Kirsch, A. M. F. Oliveira-Campos. Synthesis of thieno[2,3-d]thiazole derived dyes with potential application in nonlinear optics. Ann. Quım. Int. Ed. (1998), 94, 186–188.
[15] M. M. M. Raposo, A. M. R. C. Sousa, A. M. C. Fonseca, G. Kirsch, (2005). Thienylpyrrole Azo Dyes: Synthesis, Solvatochromic and Electrochemical Properties. Tetrahedron 61, 8249–8256.
[16] N. Arumugam, A. I. Almansour, R. S. Kumar, V. S. Krishna, D. Sriram, N. Dege. (2021). Stereoselective synthesis and discovery of novel spirooxindolopyrrolidine engrafted indandione heterocyclic hybrids as antimycobacterial agents. Bioorg. Chem., 110, 104798.
[17] Spek, A. L., 2009. Structure validation in chemical crystallography. Acta Cryst., D65, 148–155.
[18] Frisch M. J., Trucks G. W., Schlegel H. B., Scuseria G. E., Robb M. A., Cheeseman J. R., et al.; (2013) GAUSSIAN09. Gaussian, Inc., Wallingford, CT, USA.
[19] B. Delley, (1990). An All-Electron Numerical Method for Solving the Local Density Functional for Polyatomic Molecules. J. Chem. Phys. 92, 508-517.
[20] B. Delley, (2000). From Molecules to Solids with DMol3. J. Chem. Phys. 113, 7756-7764.
[21] Tse-Lok, H., (1975). Hard soft acids bases (HSAB) principle and organic chemistry. Chem. Rev. 75(1), 1-20.
[22] R. G. Pearson, (1963). Hard and Soft Acids and Bases. J. Amer. Chem. Soc, 85(22), 3533-3539.
[23] Abou A, Sosso S., Kouassi A. F., Zoueu T. J., Djandé A., Ouari O. (2021) Synthesis, Characterization, Crystal Structure and Quantum Chemical Calculations of 2-oxo-2H-chromen-3-yl Acetate. Sci. J. Chem., 9(2), 29-44
[24] Rigaku OD (2015). CrysAlis PRO. Rigaku Oxford Diffraction, Yarnton, England.
[25] Burla M C, Caliandro R, Carrozzini B, Cascarano G L, Cuocci C, Giacovazzo C, Mallamo M, Mazzone A and Polidori G, (2015). Crystal structure determination and refinement via SIR2014. J. Appl. Cryst. 48, 306–309.
[26] Farrugia L J., (2012). WinGX and ORTEP for Windows: An Update. J. Appl. Cryst. 45, 849–854.
[27] Sheldrick G M., (2015). Crystal Structure Refinement with SHELXL. Acta Cryst. C71, 3–8.
[28] Martinez-Araya J. I., (2014). Why is the dual descriptor a more accurate local reactivity descriptor than Fukui functions? J. Math. Chem. 53(2), 451 - 465.
[29] Canevet, D., Salle, M., Zhang, G., Zhang, D., and Zhu, D., (2009). Tetrathiafulvalene (TTF) derivatives: key building-blocks for switchable processes. Chem. Comm. 17, 2245-2269.
[30] Vektariene, A., Vektaris, G., and Svoboda, J., (2009). A theoretical approach to the nucleophilic behavior of benzofused thieno [3,2-b] furans using DFT and HF based reactivity descriptors, Arkivoc: Online Journal of Organic Chemistry. ARKIVOC. vii, 311-329.
[31] Islam, M. J., Kumer, A., Sarker, N., Paul, S., and Zannat, A., (2019). The prediction and theoretical study for chemical reactivity, thermophysical and biological activity of morpholinium nitrate and nitrite ionic liquid crystals: A DFT study. Adv. J. Chem. A 2(4), 316-326.
[32] E. AlShamaileh, (2014). DFT Study of Monochlorinated Pyrene Compounds. Comput. Chem. 2, 43-49.
[33] Flores-Holguín, N., Frau, J., and Glossman-Mitnik, D., (2019). Chemical reactivity and bioactivity properties of the Phallotoxin family of fungal peptides based on Conceptual Peptidology and DFT study. Heliyon, 5(8), e02335.
[34] Abbaz, T., Bendjeddou, A., and Villemin, D., (2018). Molecular structure, HOMO, LUMO, MEP, natural bond orbital analysis of benzo and anthraquinodimethane derivatives. Pharmaceutical and Biological Evaluations, 5(2), 27-39.
[35] El_Kalai F., Çınar E. B., Lai C. H., Said Daoui S., Chelfi T., Allali M., Dege N., Karrouchi K., Benchat N., (2020). Synthesis, spectroscopy, crystal structure, TGA/DTA study, DFT and molecular docking investigations of (E)-4-(4-methylbenzyl) -6-styrylpyridazin-3(2H)-one, J. Mol. Struct., 129435
[36] Agrahari B., Layek S., Ganguly R, Dege N., Pathak D. D. (2019). Synthesis, characterization and single crystal X-ray studies of pincer type Ni(II)-Schiff base complexes: Application in synthesis of 2-substituted benzimidazoles. J. Organomet. Chem., 890, 13-20.
[37] Abou A., Yoda J., Djandé A., Coussan S. and Zoueu T J. (2018). Crystal structure of 2-oxo-2H-chromen-7-yl 4-fluorobenzoate. Acta Cryst. E74, 761–765.
[38] Abou A., Sosso S., Kouassi A F., Zoueu T J., Djandé A., Ouari O. (2021). Synthesis, Characterization, Crystal Structure and Quantum Chemical Calculations of 2-oxo-2H-chromen-3-yl Acetate. Sci. J. Chem. 9(2), 29-44
[39] Koulabiga Z., Yao K H., Abou A., Djandé A., Giorgi M., Coussan S. (2024). Synthesis, Characterization, Hirshfeld Surface Analysis and Quantum Chemical Calculations of 2-oxo-2H- Chromen-6-yl 4-Methoxybenzoate. Am. J. Org. Chem., 12(1): 1-19.
[40] Janiak J, J. (2000). A critical account on π–π stacking in metal complexes with aromatic nitrogen-containing ligands. Chem. Soc. Dalton Trans. 3885–3896.
[41] D. L. Beveridge; R. Lavery Theoretical Biochemistry and Molecular Biophysics: DNA. Proteins, Adenine Press, 1990.
[42] Politzer P. and J. S. Murray J. S.; (2002). The fundamental nature and role of the electrostatic potential in atoms and molecules. Theor. Chem. Acc. 108(3), 134–142.
[43] Pearson, R. G. (1986). Absolute electronegativity and hardness correlated with molecular orbital theory. Proc. Natl. Acad. Sci. U.S.A. Nov;. 83(22): 8440-8441.
[44] Pearson, R. G. (1963) Hard and Soft Acids and Bases. J. Am. Chem. Soc. 85, 3533-3539.
[45] Yang, W., Parr, R. G. and Pucci, R. (1984). Electron Density, Kohn-Sham Frontier Orbitals, and Fukui Functions. J. Chem. Phys. 81, 2862-2863.
[46] Yang, W. and Mortier, W. (1986) The Use of Global and Local Molecular Parameters for the Analysis of the Gas-Phase Basicity of Amines. J. Am. Chem. Soc. 108, 5708-5711.
[47] Yang, W. and Parr, R. G. (1985). Chemistry. Hardness, Softness, and the Fukui Function in the Electronic Theory of Metals and Catalysis. Proceedings of the National Academy of Sciences of the United States of America, 82, 6723-6726.
[48] Chattaraj, P. K., Maiti, B. and Sarkar, U. (2003) Philicity: A Unified Treatment of Chemical Reactivity and Selectivity. J. Phys. Chem. A 107, 4973-4975.
[49] Fuentealba, P. and Contreras, R. (2002). Fukui Function in Chemistry. In: Sen, K. D., Ed., Reviews in Modern Quantum Chemistry: A Celebration of the Contributions of Robert G Parr, World Scientific, River Edge, 1013-1052.
[50] Padmanabhan, J., Parthasarathi, R., Elango, M., Subramanian, V., Krishnamoorthy, B. S., Gutierrez-Oliva, S., Toro-Labbe, A., Roy, D. R. and Chattaraj, P. K. (2007) Multiphilic Descriptor for Chemical Reactivity and Selectivity. J. Phys. Chem. 111, 9130-9138.
[51] Morell, C., Grand, A. and Toro-Labbé, A. (2005). New Dual Descriptor for Chemical Reactivity. J. Phys. Chem. 109, 205-212.
[52] Morell, C., Grand, A. and Toro-Labbé, A. (2006) Theoretical Support for Using the Delta f(r). Descriptor Chem. Phys. Lett. 425, 342-346.
Cite This Article
  • APA Style

    Yao, H. K., Koulabiga, Z., Abou, A., Djandé, A., Coussan, S., et al. (2025). Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases. Science Journal of Chemistry, 13(1), 11-32. https://doi.org/10.11648/j.sjc.20251301.12

    Copy | Download

    ACS Style

    Yao, H. K.; Koulabiga, Z.; Abou, A.; Djandé, A.; Coussan, S., et al. Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases. Sci. J. Chem. 2025, 13(1), 11-32. doi: 10.11648/j.sjc.20251301.12

    Copy | Download

    AMA Style

    Yao HK, Koulabiga Z, Abou A, Djandé A, Coussan S, et al. Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases. Sci J Chem. 2025;13(1):11-32. doi: 10.11648/j.sjc.20251301.12

    Copy | Download

  • @article{10.11648/j.sjc.20251301.12,
      author = {Honoré Kouadio Yao and Zakaria Koulabiga and Akoun Abou and Abdoulaye Djandé and Stéphane Coussan and Olivier Ouari},
      title = {Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases},
      journal = {Science Journal of Chemistry},
      volume = {13},
      number = {1},
      pages = {11-32},
      doi = {10.11648/j.sjc.20251301.12},
      url = {https://doi.org/10.11648/j.sjc.20251301.12},
      eprint = {https://article.sciencepublishinggroup.com/pdf/10.11648.j.sjc.20251301.12},
      abstract = {The first objective of the study is based on experimental characterization of the studied compound. The synthetization process of C11H8O4 (I) involved the O-acetylation of 6-hydroxycoumarin with acetic anhydride, utilizing diethyl ether as a solvent and pyridine as a base. The obtained structure was characterized by both spectroscopic analyses such as ESI-MS, FT-IR, 1H and 13C NMR analysis and by single-crystal X-ray diffraction studies. In the latter case, we employed direct methods to solve the structure of (I) and subsequently refined to a final R value of 0.054 for 1896 independent reflections. In the structure, C—H•••O hydrogen bonds connect the molecules into R22 (8) dimers, which are linked together by C—H•••O interactions, forming layers parallel to the bc crystallographic plane. Similarly, the crystal structure is sustained by π–π interactions between neighboring rings, with inter-centroid distances lower than 3.8 Å. The second objective of the study is to use theoretical calculation methods to analyze the effect of solvent polarity on the energy gap of the boundary molecular orbitals and the overall chemical reactivity of coumarin-6-yl acetate in order to provide a better understanding of stability and reactivity. A series of density functional theory computations were achieved with B3LYP/6-311++G(d,p) basis set in both gas and solvent phases. In addition to the dipole moment, the natural bond orbital charge distribution was estimated in toluene, tetrahydrofuran (THF) and benzene solvents. The calculations were conducted utilizing the Gaussian 09 software, and the outcomes exhibited that the solvents have an influence on the optimized parameters. Furthermore, dual and local reactivity indices as Fukui functions from the natural bond orbital (NBO) charges were estimated in order to have a better comprehension of the electrophilic and nucleophilic regions, as well as the chemical activity of (I). The obtained dipole moment in the gas phase is 6.03 Debye and those in the presence of the solvents are 7.89, 6.87, 7.51 and 6.83 Debye for water, toluene, THF and benzene, respectively. Additionally, the global chemical reactivity parameters exhibit variation contingent on the molecular compound and polarity of the solvents, making this an important consideration in the selection of appropriate solvents for a given chemical reaction. The studied compound shows higher stability in the benzene solvent evidenced by an EHOMO-ELUMO energy gap of 9.48 eV, while its low stability is observed in the gas phase with an EHOMO-ELUMO energy gap of 6.64 eV.},
     year = {2025}
    }
    

    Copy | Download

  • TY  - JOUR
    T1  - Synthesis, Experimental Characterizations and Theoretical Study of the Chemical Reactivity of Coumarin-6-yl Acetate in Gas and Solvent Phases
    AU  - Honoré Kouadio Yao
    AU  - Zakaria Koulabiga
    AU  - Akoun Abou
    AU  - Abdoulaye Djandé
    AU  - Stéphane Coussan
    AU  - Olivier Ouari
    Y1  - 2025/01/24
    PY  - 2025
    N1  - https://doi.org/10.11648/j.sjc.20251301.12
    DO  - 10.11648/j.sjc.20251301.12
    T2  - Science Journal of Chemistry
    JF  - Science Journal of Chemistry
    JO  - Science Journal of Chemistry
    SP  - 11
    EP  - 32
    PB  - Science Publishing Group
    SN  - 2330-099X
    UR  - https://doi.org/10.11648/j.sjc.20251301.12
    AB  - The first objective of the study is based on experimental characterization of the studied compound. The synthetization process of C11H8O4 (I) involved the O-acetylation of 6-hydroxycoumarin with acetic anhydride, utilizing diethyl ether as a solvent and pyridine as a base. The obtained structure was characterized by both spectroscopic analyses such as ESI-MS, FT-IR, 1H and 13C NMR analysis and by single-crystal X-ray diffraction studies. In the latter case, we employed direct methods to solve the structure of (I) and subsequently refined to a final R value of 0.054 for 1896 independent reflections. In the structure, C—H•••O hydrogen bonds connect the molecules into R22 (8) dimers, which are linked together by C—H•••O interactions, forming layers parallel to the bc crystallographic plane. Similarly, the crystal structure is sustained by π–π interactions between neighboring rings, with inter-centroid distances lower than 3.8 Å. The second objective of the study is to use theoretical calculation methods to analyze the effect of solvent polarity on the energy gap of the boundary molecular orbitals and the overall chemical reactivity of coumarin-6-yl acetate in order to provide a better understanding of stability and reactivity. A series of density functional theory computations were achieved with B3LYP/6-311++G(d,p) basis set in both gas and solvent phases. In addition to the dipole moment, the natural bond orbital charge distribution was estimated in toluene, tetrahydrofuran (THF) and benzene solvents. The calculations were conducted utilizing the Gaussian 09 software, and the outcomes exhibited that the solvents have an influence on the optimized parameters. Furthermore, dual and local reactivity indices as Fukui functions from the natural bond orbital (NBO) charges were estimated in order to have a better comprehension of the electrophilic and nucleophilic regions, as well as the chemical activity of (I). The obtained dipole moment in the gas phase is 6.03 Debye and those in the presence of the solvents are 7.89, 6.87, 7.51 and 6.83 Debye for water, toluene, THF and benzene, respectively. Additionally, the global chemical reactivity parameters exhibit variation contingent on the molecular compound and polarity of the solvents, making this an important consideration in the selection of appropriate solvents for a given chemical reaction. The studied compound shows higher stability in the benzene solvent evidenced by an EHOMO-ELUMO energy gap of 9.48 eV, while its low stability is observed in the gas phase with an EHOMO-ELUMO energy gap of 6.64 eV.
    VL  - 13
    IS  - 1
    ER  - 

    Copy | Download

Author Information