Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Aug 16;8(34):31548–31566. doi: 10.1021/acsomega.3c04922

Spectroscopic, Biological, and Topological Insights on Lemonol as a Potential Anticancer Agent

A Ram Kumar , S Selvaraj ‡,*, Mohammad Azam §, GP Sheeja Mol ∥,, N Kanagathara , Mahboob Alam #, P Jayaprakash
PMCID: PMC10468887  PMID: 37663516

Abstract

graphic file with name ao3c04922_0019.jpg

A monoterpene alcohol known as lemonol was investigated experimentally as well as theoretically in order to gain insights into its geometrical structure, vibrational frequencies, solvent effects on electronic properties, molecular electrostatic potential, Mulliken atomic charge distribution, natural bond orbital, and Nonlinear Optical properties. The frontier molecular orbital energy gap values of 5.9084 eV (gas), 5.9261 eV (ethanol), 5.9185 eV (chloroform), 5.9253 eV (acetone), and 5.9176 eV (diethyl ether) were predicted, and it shows the kinetic stability and chemical reactivity of lemonol. Topological studies were conducted using Multiwfn software to understand the binding sites and weak interactions in lemonol. The antiproliferative effect of lemonol against the breast cancer cell line Michigan Cancer Foundation (MCF-7) was determined by 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide assay, while nuclear damage, condensation, and reactive oxygen species generation were identified using acridine orange/ethidium bromide, propidium iodide, and dichlorodihydrofluorescein diacetate staining. The theoretical and experimental findings are highly correlated, confirming the structure, and the results of in vitro studies suggest that lemonol acts as a potent inhibitor against the human breast cancer cell line MCF-7, highlighting its strong antiproliferative activity.

1. Introduction

Lemonol, chemically known as 3,7-dimethylocta-2,6-dien-1-ol with a molecular formula of C10H18O and a molecular mass of 154.25 g/mol, is a terpene alcohol and is structurally related to limonene, linalool, and menthol. There are two double bonds in the eight-carbon chain; the first is at position 2, while the second occurs at position 6. The hydroxyl (−OH) group is located at position 1 at the terminus of the chain.1,2 Lemonol also called geraniol is a colorless liquid oil that is readily soluble in organic solvents including ethanol, chloroform, and diethyl ether and is marginally dissolved in water.3 In addition, the indispensable oils of fragrant herbs such as Cinnamomum tenuipilum Kosterm and Verbena officinalis contain lemonol naturally and are also found in oils of rose, palmarosa, lime, ginger, and orange.4,5 A genetically modified Escherichia coli strain was used to produce lemonol by knocking out the yjgB gene, which is responsible for geraniol endogenous dehydrogenation and increasing the expression of geraniol synthase activity. Also, through metabolic engineering in E. coli, the enzyme YjgB has been found to be responsible for the process of producing 2-phenylethanol (2-PE) and 2-phenylethyacetate (2-PEAc). Lemonol is also chemically synthesized by the reduction of trans methyl geranate and hydrogenation of alpha-pinene.68 Numerous researchers have reported the biological and pharmacological benefits of lemonol including its antibacterial, antiproliferative, antioxidant, and anti-inflammatory properties.911 Due to the increased toxicity of N,N-diethyl-3-methylbenzamide (DEET) is a standard insect repellent, and lemonol is utilized as a natural insect repellent.12

Cancer is a group of disorders defined by the uncontrollable growth and spread of abnormal cells to neighboring tissues. According to a recent investigation by the International Agency for Research on Cancer (IARC) and the World Health Organization (WHO), cancer is thought to be the primary cause of one in seven fatalities worldwide. Over time, cancer has moved to less urbanized countries, which now account for about 57% of new cases and 65% of deaths worldwide.13 Breast cancer is most common in women aged 40–55. Several studies have shown that postmenopausal women with a higher Body Mass Index (BMI) are more likely to acquire breast cancer. Breast cancer may be caused by genetic, hormonal, behavioral, environmental, and dietary factors. Oral contraceptives and hormone replacement therapy for postmenopausal women may increase the risk. Early breast cancer surgery may remove the tumor mass in the breast or regional lymph nodes. Chemotherapy is the standard short-term treatment for cancer although it kills both malignant and normal cells. Most anticancer drugs work only at higher doses. Naturally occurring chemicals that diminish clinical protest and are less toxic may prevent breast cancer in patients and healthy people.

In view of the above-mentioned research works, a close examination of the scientific literature reveals that a detailed theoretical and experimental analysis of the vibrational assignments, electronic properties, topological properties, and biological activities of lemonol has yet to be reported. In response to this knowledge gap, this present investigation lies in the spectroscopic, electronic, and structural properties of lemonol by employing Density Functional Theory (DFT) to conduct quantum chemical computational calculations, and the results were compared with experimental findings of the Fourier transform infrared (FT-IR), Fourier transform-Raman (FT-Raman), and ultraviolet–visible (UV–vis) analytical techniques. The topological analyses including the Electron Localization Function (ELF), Localized Orbital Locator (LOL), Reduced Density Gradient (RDG), Non-Covalent Interaction (NCI), and Quantum Theory of Atoms in Molecule (QTAIM) have been carried out to distinguish about the distribution of lone pair electrons, bonding orbitals, hydrogen bonding, van der Waals interactions, weak interactions, critical points of the electron density, and chemical reactivity of lemonol. Furthermore, a more profound understanding of the physiochemical properties and in vitro antiproliferative activity of lemonol against human breast cancer cell line Michigan Cancer Foundation-7 (MCF-7) was analyzed.

2. Materials and Methods

2.1. Sample and Experimental Details

The pure sample of lemonol in liquid form was acquired from M/s. Sigma-Aldrich Co., USA, is used as such for experimental measurements. The FT-IR spectrum was recorded in the range 4000–400 cm–1 using a 0.5 cm–1 resolution PerkinElmer Spectrum Two FT-IR/ATR Spectrometer. The FT-Raman spectrum was recorded in the range 4000–400 cm–1 using a 2 cm–1 resolution Bruker RFS 27 stand-alone FT-Raman spectrometer system. The UV–vis spectrum was recorded in the range of 400–200 nm using a PerkinElmer-Lambda 35 UV Winlab V6.0 Spectrometer.

2.2. Computational Details

All of the quantum chemical calculations were done without any kind of geometric limitation using the DFT/B3LYP level of theory1416 that was included in the Gaussian 09 W program package17 along with the 6-311++G(d,p) basis set. The vibrational frequencies, chemical shifts, electronic transitions, and molecular geometry optimizations were all determined. Vibrational Energy Distribution Analysis (VEDA 4) was used to do the calculations for the Potential Energy Distribution (PED) studies.18 The vibrational assignments, chemical shifts, and electronic transitions were determined by combining the results of the GaussView 619 and the Chemcraft program,20 which provides a visual representation and a high degree of accuracy. In addition, topological parameters like Localized Orbital Locator (LOL), Electron Localization Function (ELF), and Reduced Density Gradient (RDG) analysis were carried out using Multiwfn software.21 All the correlation graphs were analyzed using Origin software.22

2.3. Cell Line and Culture

Michigan Cancer Foundation-7 (MCF-7) cells were purchased from the National Centre for Cell Sciences (NCCS), Pune, and cultured in Dulbecco’s Modified Eagle Medium (DMEM) with 10% fetal bovine serum (FBS) and maintained at 37 °C, 5% CO2, and 95% humidity in a carbon dioxide (CO2) incubator to avoid contaminations and maintain the cells under standard growth conditions.

2.4. Cell Proliferative Assay

The cytotoxic activity of lemonol on cell viability was obtained using a 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay.23,24 In this study, the dimethyl sulfoxide (DMSO) solvent was used to dissolve the crystal formazans because it has high solubility, high stability, and low toxicity by comparison with other organic solvents. The cells were cultured in 96-well plates and maintained for 24 h with a complete medium. Then, the cells were treated with lemonol (5, 10, 20, 30, 40, 80, 100, 200, 300, 400, and 500 μM) and incubated for 24 and 48 h in a CO2 incubator along with blank and control. After the incubation, the medium was withdrawn, and each well plate received 10 μL of MTT stock solution and was then incubated with the cells for 4 h. After that, the MTT solution was discarded and 100 μL of DMSO was added to each well and read at 570 nm.

2.5. Propidium Iodide (PI) Nucleic Acid Staining

3000 cells per well were used to culture the cells on 12-well plates, and they were left to incubate for a whole night. After, the cells were treated with lemonol (300 μM) for 24 and 48 h. Afterward, the cells were rinsed with phosphate buffer saline (PBS), and the treated cells were stained with PI for 5 min. In the end, the intensity of the fluorescence was measured using a fluorescence microscope.

2.6. 2′,7′-Dichlorodihydrofluorescein Diacetate (DCFHDA) Assay

Using the oxidatively sensitive DCFHDA staining technique, the formation of intracellular reactive oxygen species (ROS) was assessed. At a density of 3 × 103 cells/well, the cells were grown on 12-well plates with sterile coverslips. The plates were incubated with DMEM medium overnight to reach confluence. Then the cells were healed with the half-maximal inhibitory concentration (IC50) value of lemonol for 24 and 48 h, respectively. The culture medium was taken out after incubation, and the wells were gently rinsed with PBS buffer. The culture plates were then maintained in the dark for 20 min with the addition of a nonfluorescent probe (2,7-dichlorofluorescein-diacetate), before being examined under a fluorescence microscope with a green filter at a magnification of 10×.

2.7. Acridine Orange/Ethidium Bromide (Ao/EtBr) Staining

In the culture plates with coverslips, the cells were cultured at a population of 3 × 103 cells/well in 12-well plates. The plates were incubated with DMEM medium overnight to reach confluence. Then the cells were treated with a half-maximal inhibitory concentration (IC50) value of lemonol for 24 and 48 h. The medium was taken out after incubation, and the wells were gently rinsed with PBS buffer. After 10 min of Ao/EtBr staining, the cells were immediately imaged using a fluorescence microscope.

2.8. Statistical Data Analysis

Using one-way analysis of variance (ANOVA) and least significant difference (LSD), Statistical Package for Social Sciences (SPSS/10) software25 was used to calculate all the data statistically. The quantitative results are presented as mean ± standard deviation (SD) and measured as statistically significant if P 0.05.

3. Results and Discussion

3.1. Optimized Structural Parameters

The optimized structure of lemonol with atomic numbering is shown in Figure 1, and the bond lengths and angles are presented in Table 1. These values are compared with the experimental values of structurally similar derivatives of lemonol downloaded from the Cambridge Crystallographic Data Centre (CCDC), as reported in the literature.26 In the optimized structure of lemonol, there are 4 types of bond lengths such as carbon–carbon (C–C), carbon–hydrogen (C–H), oxygen–hydrogen (O–H), and oxygen–carbon (O–C) as well as 6 types of bond angles carbon–carbon–carbon (C–C–C), carbon–oxygen-hydrogen (C–O–H), oxygen–carbon-hydrogen (O–C–H), carbon–carbon–hydrogen (C–C–H), oxygen–carbon-carbon (O–C–C), and hydrogen–carbon-hydrogen (H–C–H) being identified. The optimized bond length of O1–C11 is observed as 1.428 Å and calculated as 1.439 Å, while the bond angle of O1–C11-C8 is observed as 107.9° and theoretically estimated as 113.9 °. The bond length of O1–H29 is observed as 0.840 Å and calculated as 0.962 Å. In addition, optimized bond angles of C–O–H and O–C–H were theoretically estimated in the ranges of 107.94 and 104.6–109.7 and experimentally observed as 109.5 and 110.1, respectively. In comparison to other C–C bonds in the structure, the bond length of C4–C8 is 1.342 Å (DFT), 1.325 Å (experimental) and that of C5–C6 is 1.339 Å (DFT), 1.276 Å (experimental), and these are shorter because of the influence of the methyl group connected to carbon. Additionally, the optimized theoretical bond angles of C–C–C were estimated to vary from 112.1 to 128.6 and experimentally observed to range from 110.4 to 128.5.

Figure 1.

Figure 1

Optimized molecular structure of lemonol.

Table 1. Experimental and Theoretical Geometrical Parameters of Lemonol.

Bond lengths (Å) B3LYP Experimentala Bond lengths (Å) B3LYP Experimentala
O1–C11 1.439 1.428 C7–H17 1.095 0.981
O1–H29 0.962 0.840 C7–H18 1.095 0.980
C2–C3 1.549 1.543 C7–H19 1.089 0.979
C2–C4 1.512 1.511 C8–C11 1.507 1.503
C2–H12 1.098 0.988 C8–H20 1.088 0.950
C2–H13 1.093 0.991 C9–H21 1.096 0.980
C3–C5 1.503 1.557 C9–H22 1.091 0.981
C3–H14 1.092 0.990 C9–H23 1.096 0.980
C3–H15 1.096 0.990 C10-H24 1.096 0.980
C4–C7 1.507 1.488 C10-H25 1.089 0.980
C4–C8 1.342 1.325 C10-H26 1.096 0.980
C5–C6 1.339 1.276 C11–H27 1.096 0.990
C5–H16 1.090 0.950 C11–H28 1.092 0.990
C6–C9 1.509 1.481 C6–C10 1.508 1.502
Bond angles (°) B3LYP Experimentala Bond angles (°) B3LYP Experimentala
C11–O1-H29 107.9 109.5 C4–C8-H20 117.9 115.8
O1–C11-C8 113.9 107.9 C6–C5-H16 117.0 115.7
O1–C11-H27 109.7 110.1 C5–C6-C9 120.7 121.6
O1–C11-H28 104.6 110.1 C5–C6-C10 124.9 124.1
C3–C2-C4 114.2 114.6 C9–C6-C10 114.2 114.4
C3–C2-H12 108.3 108.6 C6–C9-H21 110.9 109.5
C3–C2-H13 108.7 108.5 C6–C9-H22 111.8 109.4
C2–C3-C5 112.1 110.4 C6–C9-H23 110.9 109.4
C2–C3-H14 108.9 108.1 C6–C10-H24 110.4 109.5
C2–C3-H15 108.9 109.5 C6–C10-H25 113.4 109.4
C4–C2-H12 108.9 108.7 C6–C10-H26 110.4 109.5
C4–C2-H13 109.6 108.6 H17–C7-H18 106.7 109.4
C2–C4-C7 116.2 116.0 H17–C7-H19 108.8 109.6
C2–C4-C8 120.5 119.3 H18–C7-H19 107.0 109.5
H12–C2-H13 106.6 107.6 C11–C8-H20 116.0 115.9
C5–C3-H14 111.5 109.6 C8–C11-H27 109.4 110.1
C5–C3-H15 108.6 109.6 C8–C11-H28 111.6 110.2
C3–C5-C6 128.6 128.5 H21–C9-H22 108.1 109.4
C3–C5-H16 114.3 115.8 H21–C9-H23 106.5 109.5
H14–C3-H15 106.3 108.1 H22–C9-H23 108.1 109.5
C7–C4-C8 123.2 124.7 H24–C10-H25 107.8 109.5
C4–C7-H17 110.4 109.4 H24–C10-H26 106.4 109.6
C4–C7-H18 110.8 109.5 H25–C10-H26 107.9 109.4
C4–C7-H19 112.6 109.5 H27–C11-H28 107.1 108.3
C4–C8-C11 126.0 128.3      
a

Taken from ref (26).

In contrast, the bond angles of C–C–H and H–C–H are theoretically predicted in the range of 108.3–117.9° and 106.3–108.9° and experimentally observed as 108.5–115.8° and 107.6–109.6°, respectively. The optimized bond lengths of aliphatic C–H in methyl and methylene groups were calculated in the range of 1.089–1.096 and 1.092–1.098 Å, and experimental results show that they fall in the range of 0.979–0.981 and 0.988–0.991 Å. According to the findings, theoretical values were discovered to be marginally greater than practical values because optimization was carried out in an isolated setting in the gaseous state, but the experimental findings were impacted by the crystal environment. Figure 2 displays the correlation graphs between the optimal calculated and observed bond length and bond angle values. Bond lengths and bond angles have linear coefficient values (R2) of 0.99465 and 0.92554, respectively. Nevertheless, there is good agreement between the calculated and observed data in the literature.26

Figure 2.

Figure 2

Correlation graph for bond lengths and bond angles of lemonol.

3.2. Vibrational Properties

Lemonol contains 29 atoms and has 81 fundamental vibrational modes based on the (3N-6) degrees of freedom expected to be C1 symmetry. The theoretical vibrational frequencies were determined using the DFT/B3LYP levels with the 6-311++G(d,p) basis set in the gas phase. The vibrational spectra of lemonol revealed six different bands of stretching and bending vibrations, including CH, OH, CC, CH3, CH2, and CO. The theoretical along with experimental wavenumbers of lemonol were presented in Table 2, and the theoretical and experimental FT-IR and FT-Raman spectra were shown in Figures 3 and 4. The computed vibrational wavenumbers are scaled with 0.96 for all frequencies.27 The correlation graph between theoretical and experimental wavenumbers is displayed in Figure 5, with a linear coefficient value (R2) of 0.99403, and the optimized calculated values exhibit good agreement with the experimental values.

Table 2. Experimental FT-IR and FT-Raman and Theoretical Wave Numbers of Lemonol.

  Experimental wavenumbers (cm–1)
Theoretical wave numbersb (cm–1)
 
Modes FT-IR FT-Raman Unscaled Scaled IIR SARaman Vibrational assignmentsa(≥10% PED)
1   3400 3825 3672 19.79 65.77 υ OH(100)
2 3326   3125 2999 27.67 61.17 υ CH(46), υasCH3(52)
3     3124 2998 19.06 87.68 υ CH(88)
4     3120 2994 32.50 42.35 υ CH(42), υasCH3(55)
5     3106 2981 49.81 111.83 υ CH(91)
6 2967   3096 2972 6.30 33.37 υ CH(16), υasCH3(71)
7     3077 2953 19.81 23.74 υ CH(18), υas CH2(60)
8     3074 2951 16.20 94.16 υas CH2(53), υ CH(46)
9     3062 2940 18.46 79.46 υas CH2 in CH3(97)
10     3058 2936 16.97 34.56 υas CH2(69), υ CH(24)
11 2917   3048 2926 36.52 170.02 υas CH2 in CH3(86)
12   2913 3044 2922 8.42 47.95 υs CH2 in CH3(86)
13     3020 2899 27.46 297.26 υs CH3(27), υs CH2(54)
14     3013 2892 16.82 199.08 υs CH3(64), υs CH2(32)
15     3011 2891 56.60 242.67 υs CH3(49), υs CH2(29)
16     3007 2887 79.43 138.43 υs CH2 (93)
17     3004 2883 32.26 121.80 υs CH3 (75)
18 2856, 2108 2731, 2260 2998 2877 20.89 90.88 υs CH2 (88)
19 1668 1956, 1846, 1668 1727 1658 2.12 96.04 υ C=C(71), δas CH3(11), ω CH2(10)
20     1702 1634 33.21 116.27 υ C=C (72), δasCH3(10), δ CH(11)
21     1520 1459 4.64 8.93 χ CH2(25), δas CH3(56)
22 1442 1442 1502 1442 7.20 18.36 χ CH2(10), δas CH3(44), β CH(10)
23     1496 1436 8.39 23.32 χ CH2(11), δas CH3(12), υ CC(30)
24     1492 1432 12.94 0.66 δas CH3(39), β CH(38)
25     1489 1429 7.03 3.70 δas CH3(32), χ CH2(15)
26     1485 1425 8.56 13.36 δas CH3(48), χ CH2(30)
27     1481 1422 3.54 7.78 χ CH2(35), δas CH3(16)
28     1473 1414 0.88 16.15 δas CH3 (25), χ CH2(43)
29     1471 1412 3.83 8.68 δas CH3(11), χ CH2 (47)
30 1378 1378 1421 1363 2.72 19.21 ω CH3 (38), ω CH(23)
31     1419 1362 11.99 9.92 β CH(12), δs CH3(15), υ CC(35)
32     1412 1355 49.85 10.40 ω CH2(10), δ OH(20), υ CC(43)
33     1411 1353 6.36 5.39 β CH(18), υ CC(43)
34   1327 1385 1328 2.07 8.21 β CH (18), δs CH3(17), ω CH2(23)
35     1381 1324 8.75 9.61 β CH(23), δs CH3(18), δ OH(12), τ CH2(22)
36     1358 1303 3.57 44.23 ω CH2(33), δ OH(42), υ CC(11)
37   1280 1346 1291 2.86 8.64 δ OH(22), τ CH2(11), υ CC(10)
38     1310 1258 0.48 16.21 τ CH2(80)
39 1235 1231 1283 1231 0.78 2.61 ω CH2(60)
40     1246 1195 11.75 7.92 τ CH2(22), δ OH(10)
41 1182   1239 1189 4.94 2.17 δ OH(12), τ CH2(11)
42     1188 1140 10.63 5.61 δ OH(11), τ CH2(28)
43     1172 1125 1.28 7.07 τ CH2(33)
44 1094   1130 1085 12.90 9.71 ω CH2 (34)
45     1102 1057 1.61 1.41 β CH (42)
46     1100 1056 5.16 4.84 δ OH(20), τ CH2 (36)
47     1067 1023 4.84 9.56 β CH(48)
48 995 999 1047 1004 8.34 3.58 τ CH2(11), δ OH(10), υ CO(18)
49     1028 986 0.51 0.92 γ CH(28)
50     1015 973 3.51 1.61 γ CH (27)
51     1005 963 7.97 36.13 δas CH3(58)
52     990 950 90.34 4.97 δ OH (55)
53     973 934 1.86 1.14 δas CH3(10), τ CH2 (12)
54     964 925 33.84 4.82 δ OH(10), δas CH3(25)
55     959 921 0.72 1.25 δas CH3(45)
56 830   863 828 4.90 3.22 δas CH3(33), γ CH(18)
57     856 822 7.32 0.64 γ CH (22)
58 779 781 811 779 1.03 9.32 δs CH3(56)
59 744   789 756 7.90 4.27 δs CH3(45)
60     755 725 1.72 1.20 ρ CH2(44)
61 592   631 605 13.32 0.64 δ OH(10), δas CH3(15), γ CH(11), τ CH2(12)
62     565 541 7.78 1.20 δ OH (12), γCH(15), δas CH3(18)
63     505 485 0.89 2.39 δas CH3(38)
64     481 462 2.22 1.16 δas CH3(38)
65     434 417 12.64 1.57 δ OH(18), δas CH3(11)
66     400 383 0.21 0.76 δas CH3(18), υ CC(26)
67     389 373 2.45 2.06 δs CH3(18), τ CH2(11)
68     314 300 14.68 4.35 δ OH(18), δ CH(22)
69     307 294 40.72 0.73 δ OH(12), δ CH(18), τ CH2(22)
70     290 277 28.35 0.42 δ OH (34), ρ CH2(30)
71     258 247 66.66 0.45 δ OH(30), ρ CH2(34)
72     226 217 10.76 1.31 δas CH3(20), δ OH(20), ρ CH2(18)
73     198 190 1.64 0.46 δas CH3(40), δ OH(18)
74     151 145 0.75 0.28 δas CH3(15), δ OH(10)
75     138 132 0.67 0.27 τ CH3(63)
76     106 102 0.03 0.56 τ CH3(43)
77     70 67 0.13 0.68 τ CH3(66), δ OH(13)
78     60 58 0.20 0.72 δ OH(10), ρ CH2(11), δas CH3(12),δ CH(11)
79     52 49 1.20 1.34 δ OH(11), ρ CH2(18), δas CH3(11),δ CH(12)
80     32 30 0.72 1.76 δ OH(12), ρ CH2(10), δs CH3(10), δ CH(12)
81     30 28 0.10 1.53 δ OH(22), ρ CH2(18), δs CH3(11)
a

υs - symmetric stretching; υas - asymmetric stretching; δ - bending/deformation; β - in-plane bending; γ - out-of-plane bending; χ - scissoring; ω - wagging; τ - twisting; ρ - rocking.

b

Theoretical wavenumber scaling factor: 0.96 for all vibrations.

Figure 3.

Figure 3

Theoretical and experimental FT-IR vibrational spectra of lemonol.

Figure 4.

Figure 4

Theoretical and experimental FT-Raman vibrational spectra of lemonol.

Figure 5.

Figure 5

Correlation between all experimental and theoretical wavenumbers.

3.2.1. Hydroxyl Group Vibrations

When compared to other groups, the hydroxyl group vibrations are more susceptible to the environment and offer the three common vibrations of stretching in-plane and out-of-plane deformation. The O–H stretching modes are generally expected in the range of 3550–3700 cm–1.28 The O–H stretching vibration is expected to occur at 3672 cm–1 theoretically from the current investigation, and a single band was seen in the matching experimental FT-Raman spectra with PED of 100% at 3400 cm–1. The out-of-plane and in-plane deformation vibrations are expected at 710–517 and 1420–1330 cm–1,29 respectively. Experimentally, a medium band is observed at 1378 cm–1 (FT-IR) and a weak intense band is observed at 1378 cm–1 (FT-Raman) in the spectrum, while the corresponding theoretical bands are simulated at 1363 and 1324 cm–1. The theoretical assignments of the out-of-plane deformation vibrations at 605 and 541 cm–1 are in good agreement with the observed FT-IR spectrum at 592 cm–1.

3.2.2. CH Vibrations

C–H stretching modes are anticipated in the range of 3000–3100 cm–1, as this range is the defining region for the detection of C–H vibrations.30,31 From this study, CH vibrations are anticipated to be at 2999, 2998, 2994, 2981, 2972, 2953, 2951, and 2936 cm–1, with their matching experimental peaks in the FT-IR spectrum at 3326 and 2967 cm–1. The typical ranges for the in-plane and out-of-plane deformation modes are 1450 to 1000 cm–1 and 1000 to 750 cm–1.32 A theoretical simulation of the C–H in-plane deformation modes falls in the range 1442–1023 cm–1, as observed experimentally ranging from 1442 to 1378 cm–1 (FT-IR) and 1442–1327 cm–1 (FT-Raman). The DFT approach assigned the out-of-plane deformation vibrations to wavelengths of 986, 973, 828, and 822 cm–1, while the equivalent experimental wavenumber is found in the FT-IR spectrum at 830 cm–1.

3.2.3. CC Vibrations

The frequency ranges of the double-bonded (C=C) and single-bonded (C–C) stretching vibrations are typically in the ranges of 1280–1380 and 1400–1625 cm–1.33 In this instance, the C=C bands appeared at 1436, 1362, 1355, 1353, and 1291 cm–1 theoretically; however, the experimental Raman spectrum at 1280 cm–1 has been ascribed to these bands. The peaks at 1668 cm–1 (FT-IR) and 1668 cm–1 (FT-Raman) correspond to the predicted values at 1658 and 1634 cm–1 and are ascribed to the C–C stretching modes.

3.2.4. Methyl Group Vibrations

The symmetric and asymmetric stretching vibrations of methyl groups are typically observed between 2870 and 2980 cm–1.34,35 Theoretically, the CH3 symmetric stretching occurs at 2922, 2899, 2892, 2891, and 2883 cm–1, and experimentally, the corresponding mode is found in the FT-Raman spectrum at 2913 cm–1. Similarly, the asymmetric stretching occurs at 2999, 2994, and 2972 cm–1, and the corresponding experimental values are found at 3326 and 2967 cm–1 in the FT-IR spectrum. The predicted ranges for the asymmetric and symmetric deformation vibrations are 1465–1440 and 1390–1370 cm–1.36 In contrast to what is experimentally observed, the CH3 asymmetric bending is anticipated to occur between 1668 and 1442 cm–1 in the FT-IR and FT-Raman spectrum and the corresponding simulated values at 1658–1412 cm–1. The bands for CH3 symmetric deformation seen in this work at 1362, 1328, and 1324 cm–1 exhibited good harmony with the experimental spectrum at 1327 cm–1 in the FT-Raman spectrum.

3.2.5. Methylene Group Vibrations

Six fundamental frequencies are displayed by the methylene group, including symmetric and asymmetric, two in-plane deformations (rocking and scissoring), and two out-of-plane deformations (wagging and twisting). In the ranges of 3000–2900 and 3100–3000 cm–1, the symmetric and asymmetric stretching vibrations of CH2 are observed in the previous report.37 In the molecular structure of lemonol, the CH2 symmetric stretching vibrations are theoretically assigned at 2899, 2892, 2891, and 2887 cm–1, experimentally observed at 2913 cm–1 in the FT-Raman spectrum. The DFT technique predicts the asymmetric stretching vibrations at 2953, 2951, 2940, 2936, and 2926 cm–1, and the equivalent experimental band was noticed at 2917 cm–1 in the FT-IR spectrum. The CH2 deformation vibrations were anticipated in the range below 1500 cm–1. The CH2 scissoring frequencies were observed at 1459, 1442, 1436, 1429, 1425, 1422, 1414, and 1412 cm–1 as well as the experimental frequencies at 1442 cm–1 in both FT-IR and FT-Raman spectra. While the CH2 wagging vibrations were experimentally measured at 1235, 1094 cm–1 in the FT-IR spectrum and at 1327, 1231 cm–1 in the FT-Raman spectrum, the corresponding simulated vibrations are at 1355, 1328, 1303, 1231, and 1085 cm–1. The CH2 twisting vibrations are measured and assigned to wavenumbers 1280, 999 cm–1 in the FT-Raman spectrum and to 1182, 995 cm–1 in the FT-IR spectrum and simulated in the range of 1324–1004 cm–1. Similarly, the calculated CH2 rocking vibrations were at 725, 277, 247, and 217 cm–1.

3.2.6. CO Vibrations

In general, the frequency range of CO stretching vibrations is observed as 1260–1000 cm–1.38,39 The CO stretching vibration of lemonol is calculated at 1004 cm–1, and the corresponding experimental band is found as a strong intensity peak at 995 cm–1 (FT-IR) and a weak intensity peak at 999 cm–1 (FT-Raman), which are satisfactorily correlated with the theoretical value.

3.3. Electronic Properties

When the physical and chemical activities of molecules are described, the frontier molecular orbital (FMO) energies play an intriguing role. By using the time-dependent (TD) DFT approach with the B3LYP/6-311++G (d,p) basis set, the theoretically determined oscillator strengths (f), wavelength (λ), and excitation energy (E) of lemonol have been simulated in the gas phase and different solvents (ethanol, chloroform, acetone, and diethyl ether) which are compared to experimental findings. Figure 6 shows the theoretical and experimental spectra, and Table 3 lists the theoretical and experimental values. In the present work, a strong peak was observed at 210 nm experimentally, which shows good agreement with the theoretical spectra at 217 nm (gas), 214 nm (ethanol, acetone), and 215 nm (chloroform, diethyl ether). The second peak observed as a medium band at 278 nm has a good correlation with the calculated spectrum at 232 nm (gas, chloroform) and 231 nm (ethanol, acetone, diethyl ether). The deviation between the calculated and experimental spectra is due to a specific solvent–solute interaction. The data shown in Figure 7 and Table 4 represent the FMO plots of lemonol; the positive phases are shown in red, and the negative phases are shown in green. The charge transfer and interactions that occur within the molecules were revealed by the FMO energy gap, which was found to be 5.9084 eV (gas), 5.9261 eV (ethanol), 5.9185 eV (chloroform), 5.9253 eV (acetone), and 5.9176 eV (diethyl ether). In addition, the energy gap of lemonol was determined as 5.186 eV experimentally with the help of Tauc’s plot as shown in Figure 8, which matches well with the theoretical energy gap values calculated in the gas phase.

Figure 6.

Figure 6

(a) Simulated electronic spectra of lemonol in different solvents. (b) Experimental electronic spectrum of lemonol in ethanol.

Table 3. Experimental and Calculated Wavelengths (λ), Excitation Energies (E), Absorbance Values (A), and Oscillator Strengths (f) of Lemonol.

    TDDFT/B3LYP/6-311++G(d,p)
Experimental(Ethanol)
Gas
Ethanol
Chloroform
Acetone
Diethyl ether
λ (nm) Abs. λ (nm) E (eV) f λ (nm) E (eV) f λ (nm) E (eV) f λ (nm) E (eV) f λ (nm) E (eV) f
278 0.298 232 5.33 0.0841 231 5.35 0.1355 232 5.34 0.1365 231 5.35 0.1351 231 5.34 0.1299
    229 5.41 0.0199 226 5.47 0.0019 227 5.45 0.0041 226 5.47 0.0019 227 5.45 0.0044
210 2.286 217 5.70 0.0101 214 5.77 0.0227 215 5.75 0.0207 214 5.77 0.0225 215 5.75 0.0194

Figure 7.

Figure 7

FMO plots of lemonol.

Table 4. FMO Energies and Energy Gap of Lemonol.

  TDDFT/B3LYP/6-311++G(d,p)
Parameters(eV) Gas Ethanol Chloroform Acetone Diethyl ether
HOMO energy –6.4542 –6.4368 –6.4319 –6.4363 –6.4316
LUMO energy –0.5458 –0.5107 –0.5134 –0.5110 –0.5140
HOMO–LUMO energy gap 5.9084 5.9261 5.9185 5.9253 5.9176
HOMO–1 energy –6.9372 –6.9489 –6.9402 –6.4363 –6.9394
LUMO+1 energy –0.3322 –0.2277 –0.2487 –0.2288 –0.2514
(HOMO–1)-(LUMO+1) energy gap 6.6050 6.7212 6.6915 6.2075 6.6880

Figure 8.

Figure 8

Tauc’s plot of lemonol.

When substituents within the molecular structure of lemonol can donate or accept electrons, the electronic properties are influenced, resulting in a smaller band gap. The electron-donating groups facilitate electron donation to the system, which stabilizes the energy level of the highest occupied molecular orbital (HOMO). In contrast, the electron-accepting groups can withdraw electrons from the system, thereby lowering the lowest unoccupied molecular orbital (LUMO) energy levels. While the HOMO–1 and LUMO+1 indicate the donor and acceptor energy levels, which lie one energy state lower and higher than their respective levels, the nature of HOMO and LUMO is revealed through band theory, which interprets HOMO as the conduction band and LUMO as the valence band.4042 The combination of these substituents results in an effective decrease in the band gap of lemonol. Furthermore, the conjugation present in lemonol can also contribute to a smaller band gap by introducing additional conjugated π systems such as double bonds. Lemonol in various phases is characterized by a smaller band gap due to the extended conjugation, which allows electrons to disperse over a larger spatial region. In line with this observation, the minimum energy gap of lemonol was observed in the gas phase at 5.9084 eV exploring the low kinetic stability and high chemical reactivity. Furthermore, the electronic descriptors such as ionization potential (IA), electron affinity (EA), chemical hardness (η), chemical softness (S), chemical potential (μ), electronegativity (χ), and global electrophilicity (ω) are presented in Table 5. It is important to note that the values of chemical hardness (η) and softness (S) are two important characteristics to determine the chemical stability and polarizability of molecules. In the present work, the maximum and minimum chemical hardness were found at 2.9630 (ethanol) and 2.9542 (gas) exploring that the lemonol has more stability and less reactivity in the solvent phase and less stability and more reactivity in the gas phase, respectively.

Table 5. Calculated Quantum Chemical and Physicochemical Parameters of the Lemonol.

Parameters (eV) Formula Gas Ethanol Chloroform Acetone Diethyl ether
EHOMO   –6.4542 –6.4368 –6.4319 –6.4363 –6.4316
ELUMO   –0.5458 –0.5107 –0.5134 –0.5110 –0.5140
Energy gap ΔE (EHOMOELUMO) 5.9084 5.9261 5.9185 5.9253 5.9176
Ionization potential (IP) EHOMO 6.4542 6.4368 6.4319 6.4363 6.4316
Electron affinity (EA) ELUMO 0.5458 0.5107 0.5134 0.5110 0.5140
Electronegativity (χ) –1/2 (ELUMO + EHOMO) 3.50 3.4737 3.4726 3.4736 3.4728
Chemical hardness (η) 1/2 (ELUMOEHOMO) 2.9542 2.9630 2.9592 2.9626 2.9588
Chemical softness (S) 1/2η 0.08462 0.08437 0.08448 0.08438 0.08436
Chemical potential (μ) –χ –3.50 –3.4737 –3.4726 –3.4736 –3.4728
Global electrophilicity (ω) μ2/2η 1.0366 1.0180 1.0186 1.0180 1.0174
Maximum electronic charge (ΔNmax) –(μ/η) 0.5923 0.5861 0.5867 0.5862 0.5868

3.4. Natural Bond Orbital (NBO) Properties

The chemical nature of orbitals, hyperconjugative relations, and hydrogen bonding can be accomplished with the help of natural bond orbital analysis. To review the donor and acceptor interactions of lemonol, the second-order Fock matrix is used. Table 6 includes the electron density, donor and acceptor stabilization energies, and localized NBO. The hyperconjugation relationships between electron donors and acceptors within the molecule become more intense as E(2) increases. In the lemonol compound, the strongest interactions are associated with electron donation from C5–H16 to the acceptor C6–C10 with a stabilizing energy of 8.24 kJ mol–1 with a transition of σ–σ*. Similarly, significant delocalization was also found at electron donation from C8–H20 to the acceptor C4–C7 with a stabilizing energy of 7.95 kJ mol–1. Moreover, the electron is donated from lone pair donor O1 to the acceptor C11–H27 with a stabilizing energy of 6.98 kJ mol–1 with a transition of L(2)-σ*. The simulated stabilizing energy of charge transfer from donor C4–C8 to acceptor O1–C11 is 6.68 kJ mol–1 with a π–σ* transition. In the lemonol, the maximum contribution was made by Cieplak interaction from occupied σ orbital to unoccupied σ* including C2–C4, C2–H13, C3–C5, C6–C9, C7–H19, and C8–C11 donated its σ electron to σ* antibonding orbitals C8–C11, C4–C7, C6–C9, C3–C5, C2–C4, and C2–C4 with stabilizing energies of 4.09, 4.82, 4.14, 4.69, 4.29, and 4.41 kJ mol–1.

Table 6. Second-Order Perturbation Theory of the Fork Matrix in NBO Analysis of Lemonol Based on the B3LYP/6-311++G(d,p) Basis Set.

Donor (i) Type of bond Acceptor (j) Type of bond Type of transition E(2)a(kJ/mol) E(j) – E(i)b(a.u.) F(i,j)c(a.u.)
C2–C4 σ C8–C11 σ* σ–σ* 4.09 1.04 0.058
C2–H13 σ C4–C7 σ* σ–σ* 4.82 0.91 0.059
C3–C5 σ C6–C9 σ* σ–σ* 4.14 1.05 0.059
O1 L(2) C11–H27 σ* L(2)-σ* 6.98 0.68 0.062
C4–C8 π O1–C11 σ* π–σ* 6.68 0.54 0.054
C5–H16 σ C6–C10 σ* σ–σ* 8.24 0.92 0.078
C6–C9 σ C3–C5 σ* σ–σ* 4.69 1.06 0.063
C7–H18 σ C4–C8 π* σ–π* 4.81 0.56 0.041
C7–H19 σ C2–C4 σ* σ–σ* 4.29 0.93 0.057
C8–C11 σ C2–C4 σ* σ–σ* 4.41 1.06 0.061
C8–H20 σ C4–C7 σ* σ–σ* 7.95 0.93 0.077
a

E(2) – Mean energy of hyper-conjugative interactions (stabilization energy).

b

E(j) – E(i) – The energy difference between the donor (i) and acceptor (j) natural bonding orbitals.

c

F(i,j) – Fock matrix element between i and j natural bonding orbitals.

3.5. Nonlinear Optical (NLO) Properties

Several studies on novel compounds with nonlinear characteristics have attracted a lot of attention in recent years. To find NLO property molecules on an inexpensive basis and to ascertain the molecule’s first-order hyperpolarizability tensor, computational research is performed. The effectiveness of the electron connection between the donor and acceptor groups determines these polarizabilities and hyperpolarizabilities. In the present case, the NLO properties of dipole moment, anisotropy polarizability, and first-order hyperpolarizability of lemonol are presented in Table 7. The dipole moment is a first-order tensor that describes the distribution of electric charge in a molecule expressed in the Debye unit (1 Debye = 3.336 × 10–30 Coulomb-meter). The dipole moment along the X axis is found to be 1.8978 Debye, which is high when compared to the Y (0.1214 Debye) and Z (0.7879 Debye) axes. The overall dipole moment is calculated to be 2.0584 Debye, which indicates a moderate to relatively strong dipole moment. The ability of the material to be polarized is measured by the second-order polarizability tensor. αxx, αyy, and αzz represent interactions between the X, Y, and Z components of the electric field and induced polarization. The corresponding polarizability values are calculated to be −80.2609, 66.8832, and −73.0347 e.s.u., respectively. The negative values indicate that the induced polarization is opposite in direction to the applied electric field. In this case, the total polarizability is calculated to be −73.39 e.s.u., which is 19.05 times greater than that of urea. Each component of the third-order hyperpolarizability tensor obtained from theoretical calculations provides information about the directional dependence and magnitude of the nonlinear response of a material. Lemonol exhibited the highest NLO property in the βxxx direction with a positive value of 51.4818 × 10–31 e.s.u. The next strongest interaction is in βzxx with 24.8102 e.s.u. followed by βyzz with 5.1512 e.s.u. Similarly, the component βxxy = −8.8482 e.s.u. represents the interaction between the x-component of the applied electric field and the y-component of the induced polarization. The negative value indicates that there is an opposite phase or orientation in this interaction. The computed other parameters are given in Table 7. The value of first-order hyperpolarizability was found to be 46.74 e.s.u., which is 12.534 times greater than the value of urea, making it an excellent choice for future research on nonlinear optics.4346

Table 7. Electric Dipole Moment μ(D), the Average Polarizability αtot (×10–24 e.s.u), and First-Order Hyperpolarizability βtot (×10–31 e.s.u) Calculated at B3LYP/6-311++G(d,p) for Lemonol.

Parameters B3LYP/6-311++G(d,p) Parameters B3LYP/6-311++G(d,p)
μx 1.8978 βxxx 51.4818
μy 0.1214 βxxy –8.8482
μz 0.7879 βxyy –9.2997
μ(D) 2.0584 βyyy 8.9232
αxx –80.2609 βzxx 24.8102
αxy 2.7516 βxyz –6.8277
αyy –66.8832 βzyy –1.6729
αxz –4.3588 βxzz –2.5006
αyz 1.1042 βyzz 5.1512
αzz –73.0347 βzzz –3.6737
αtotal(e.s.u) –73.39 βtotal(e.s.u) 46.74

3.6. Mulliken Charges

The application of quantum mechanical calculations to molecular systems relies profoundly on the accurate determination of the atomic charges. The Mulliken charge can be thought of as a quantitative representation of how the electronic structure changes in response to atomic displacement.47,48Table 8 shows the calculated Mulliken charges of lemonol as well as graphically represented in Figure 9. The results of lemonol illustrate that all the hydrogen atoms have positive potential; especially, the hydrogen atom H29 has a higher positive charge (0.2289e) than all other hydrogen atoms due to the interaction of oxygen atom O1, which is evident for the electronegativity of the oxygen atom. The carbon atoms C4 (0.8782e) and C6 (0.4313e) show positive potential which is attached to methyl groups, whereas other carbon atoms C2, C3, C5, C7, C8, and C11 show negative potential; especially, the C11 atom exhibits more negativity due to the attachment of oxygen atom. On the whole, carbon atom C4 exhibits a more positive charge of (0.8782e), and carbon atom C7 exhibits a more negative charge of (−0.7306e), which is also confirmed with molecular electrostatic potential (MESP) analysis.

Table 8. Mulliken Population Analysis of Lemonol.

Atoms Charges (e) Atoms Charges (e)
O1 –0.231876 H16 0.166465
C2 –0.585991 H17 0.15039
C3 –0.289292 H18 0.172824
C4 0.878218 H19 0.172086
C5 –0.015905 H20 0.162222
C6 0.431364 H21 0.142858
C7 –0.730697 H22 0.150826
C8 –0.232234 H23 0.138774
C9 –0.633561 H24 0.13852
C10 –0.621358 H25 0.139948
C11 –0.707736 H26 0.140575
H12 0.135283 H27 0.152511
H13 0.158713 H28 0.165559
H14 0.09728 H29 0.228919
H15 0.125315    

Figure 9.

Figure 9

Mulliken atomic charge distribution of lemonol.

3.7. Molecular Electrostatic Potential Surface Analysis

By using different colors on the map, the molecular electrostatic potential (MESP) surface is utilized to examine the hydrogen bonding and electron-rich and -poor regions in the molecular structure of lemonol.49 The mapped color bands illustrated in Figure 10 correspond to specific regions, which are predicted between variations ranging from −4.974 × 10–2 to 4.974 × 10–2 e.s.u. for lemonol. From the surface map, the electron-poor regions, which serve as binding sites for electrophilic reactive species, are depicted as white color around the hydrogen atoms, while the electron-rich region, acting as binding sites for nucleophilic reactive species, is depicted as red color around the electronegative oxygen atom in the hydroxyl group of lemonol. The obtained results show an excellent correlation with the Mulliken population analysis.

Figure 10.

Figure 10

Total electron density (left) and the contour map (right) with the molecular electrostatic potential surface of lemonol.

3.8. Electron Localization Function (ELF) and Localized Orbital Locator (LOL)

Figures 11 and 12 describe the two-dimensional description of the electron localization function (ELF) and the localized orbital locator (LOL). The ELF is a tool used in computational chemistry to analyze the electronic structure of molecules. The ELF is a function that measures the degree of electron localization in a molecule. It provides a quantitative measure of the degree to which electrons are localized in a given nucleus. The ELF is useful in the study of chemical bonding, particularly in the analysis of bonding in solid-state materials. It can be used to distinguish between covalent, metallic, and ionic bonding and to identify regions of high electron density where chemical reactions are likely to occur. The electron localization function is a valuable tool in computational chemistry for understanding the electronic structure of molecules and materials. It provides a quantitative measure of the degree of electron localization, which can be used to analyze chemical bonding and predict reactivity.50 The LOL is a function that measures the degree of localization of an orbital in a molecule or material. It provides a quantitative measure of the extent to which a given orbital is localized around a given nucleus or group of nuclei. The hydrogen atoms of the lemonol compound with single electrons exhibit the highest Pauli repulsion, which is shown by the red zone, and the carbon and oxygen atoms exhibit the highest Pauli repulsion, which are shown by the blue region. Red spots in the figure represent locations with high LOL values and may be attributed to covalent regions, whereas blue areas can be interpreted as areas where there is electron depletion among the valence and inner shell.51 The C5 and C6 atoms show the active region of the lemonol compound.

Figure 11.

Figure 11

ELF projection and contour map of lemonol.

Figure 12.

Figure 12

LOL projection and contour map of lemonol.

3.9. RDG (Reduced Density Gradient) Analysis and NCI (Noncovalent Interaction) Analysis

RDG stands for reduced density gradient analysis. It is a tool used in computational chemistry to analyze the electronic structures of molecules and determine their reactivity. The reduced density gradient (RDG) is a function that measures the magnitude of the electron density and its gradient. High values of the RDG correspond to regions of high electron density, which are associated with chemical reactivity. RDG analysis is useful in the study of various chemical processes, such as nucleophilic and electrophilic reactions, hydrogen bonding, and charge transfer. In summary, RDG analysis is a useful tool in computational chemistry for studying molecular electronic structure and reactivity. It provides a quantitative measure of the electron density and its gradient, which can be used to identify reactive regions and predict chemical reactivity. Figure 13 demonstrates how the noncovalent interaction (NCI) approach may be used to examine all kinds of interactions, especially in low-density areas or regions.

Figure 13.

Figure 13

NCI of lemonol.

The three-dimensional isosurfaces of the lemonol were plotted in Figure 14, which presents a color code that differentiates the various types of interactions. The color blue signifies a hydrogen bond, the color green signifies the van der Waals interaction, and the color red signifies the strong repulsive force, also known as the steric effect in rings and cages. In strong attractive force ρ > 0, λ2 < 0, in van der Waals interaction ρ > 0, λ2 = 0, and strong repulsive force or steric effect in ring and cage which shows ρ > 0, λ2 > 0. The van der Waals interactions are located at −0.010 to 0.010 a.u. The strong attractive force of the lemonol compound is exhibited at −0.050 to −0.010 a.u., and the corresponding strong repulsive force or steric effect in the ring and cage is localized at 0.010 to 0.050 a.u.

Figure 14.

Figure 14

RDG analysis of lemonol.

3.10. QTAIM Analysis

The QTAIM (Quantum Theory of Atoms in Molecules) analysis is a tool used in computational chemistry to analyze the electronic structure of molecules. The QTAIM analysis is a method that allows the study of the topology of the electron density in a molecule and interaction within the molecule including hydrogen and noncovalent interactions.5255 It provides information about the nature of the chemical bonds, the distribution of the electronic charge, and the geometry of the molecule. The definition of the critical points (CPs) of the electron density, which are locations where the gradient of the electron density is zero, serves as the foundation for the QTAIM analysis.

The CPs are classified into four types: nuclear, bond, ring, and cage critical points. The QTAIM analysis allows the outcome of diverse properties of the electron density, such as the electron density at the CPs, Laplacian, and the total energy of the compound.56,57 Also, the analysis provides information about the electron density distribution and the nature of the chemical bonds, such as the bond path and the bond critical point (BCP). The BCPs are used to classify chemical bonds into covalent, ionic, and hydrogen bonds. The quantum theory of atoms in lemonol is plotted in Figure 15, and the parameters are presented in Table 9. The primary covalent connections between the atoms of the molecule also include a summary of the kinetic energy (G), potential energy (V), total electron density (H), ellipticity, and interaction energies (in kcal/mol) of the principal molecular interactions. The MOD(V/G) values are greater than 1, which means that all of the interactions are covalent.58,59

Figure 15.

Figure 15

QTAIM analysis of lemonol.

Table 9. Quantum Theory of Atom in Molecules Bond Critical Points (BSP) and Its Parameters.

Interaction δ(r)a.u. V2(r)a.u. H(r)a.u. G(r)a.u. V(r)a.u. λ1a.u. λ2a.u. λ3a.u. ELF LOL Ellipticity of electron density
BCP1 0.0110 0.0524 0.00280 0.0102 –0.00748 –0.00938 0.0110 0.0507 0.02259 0.1390 –1.847
BCP2 0.4330 –0.00248 –0.0623 –0.00286 –0.0626 –0.0879 –0.0877 –0.0727 0.9983 0.9614 0.0024

3.11. Cell Proliferative Assay

Lemonol’s ability to inhibit the proliferation of the human breast cancer cell line MCF-7 was examined using the MTT assay; with increasing lemonol concentrations, cell death ensues as illustrated in Figure 16. For 24 and 48 h, the cells were exposed to lemonol at various doses ranging from 5 to 500 μM. The solvent DMSO-treated cells served as a control. From the results, lemonol exhibits potential cytotoxicity against MCF-7 cells with an IC50 value of 300 μM concentration; meanwhile, further experiments were carried out using the IC50 value of the lemonol.

Figure 16.

Figure 16

Effect of lemonol on the cell viability in cultured MCF-7 cells. Cultured MCF-7 cells were treated with lemonol at different concentrations (5, 10, 20, 40, 80, 100, 200, 300, 400, 500 μM) for 24 and 48 h.

3.12. Propidium Iodide Nucleic Acid Staining

The nuclear staining of propidium iodide was used to examine the changes in the nucleus caused by lemonol in MCF-7 cells. Nuclear condensation and nuclear fragmentation are biochemical features of programmed cell death or apoptosis. In comparison with the control, lemonol-treated MCF-7 cells show more intense fluorescence due to condensation and fragmentation of nuclear deoxyribonucleic acid (DNA) shown in Figure 17A.

Figure 17.

Figure 17

(A) Fluorescent microscope images of PI staining of MCF-7 cells controlled and treated with 300 μM of lemonol, blue arrow shows PI-stained cells. (B) DCFHDA Staining of MCF-7 cells, blue arrow shows the generation of ROS (Reactive oxygen species).

3.13. 2′,7′-Dichlorodihydrofluorescein Diacetate (DCFHDA) Assay

To determine the mechanism of the lemonol antiproliferative effect and whether reactive oxygen species-mediated programmed cell death apoptosis occurs, the reactive oxygen species generation was examined using DCFHDA in fluorescence microscopy. The lemonol was treated with a 300 μM concentration for 24 and 48 h, as shown in Figure 17B. MCF-7 cells emit more fluorescence intensity compared with the control MCF-7 cells. This reactive oxygen species generation analysis revealed that lemonol at 300 μM induced increased reactive oxygen species generation in MCF-7 cells after 24 and 48 h of treatment.

3.14. Acridine Orange/Ethidium Bromide (Ao/EtBr) Staining

To detect the morphological evidence of apoptosis, AO/EtBr staining was carried out. The MCF-7 cells were treated with 300 μM concentration for 24 and 48 h as shown in Figure 18. Acridine orange stains live and dead cell components, including the cytoplasm and the nucleus. Ethidium bromide stains the nucleus of the damaged cells. AO/EtBr staining revealed that lemonol-treated MCF-7 cells were yellow and red due to nuclear damage, confirming apoptosis. However, no EtBr was detected in the control cell.

Figure 18.

Figure 18

Fluorescent microscope images of AO/EtBr dual staining blue arrows indicate DNA damage, as evidenced by nuclear staining with ethidium bromide.

4. Conclusions

High-level quantum chemical calculations were employed to determine the structural and spectroscopic characteristics of lemonol, which is a significant class of monoterpenoid. The experimental and theoretical bond lengths and bond angles of lemonol are correlated well with linear coefficient values (R2) of 0.99465 and 0.92554, respectively. Due to the influence of electron donors and acceptors as well as conjugation systems, the minimum energy gap value found theoretically in the gas phase is 5.9084 eV. With the help of Tauc’s plot, the energy gap was determined as 5.186 eV from the experimental UV–vis spectrum. In the NBO, the strongest interactions associated with electron donation were observed from C5–H16 to the acceptor C6–C10 with a stabilizing energy of 8.24 kJmol–1 with a transition of σ–σ*. The reactive sites and distribution of charges and noncovalent bonds of lemonol were found and confirmed by topological, MESP, and Mulliken charge distribution studies. Furthermore, the polarizability and hyperpolarizability values of lemonol are 19.05 and 12.53 times greater than urea, respectively. The in vitro MTT assay revealed that lemonol shows potent antiproliferative activity against the human breast cancer cell line MCF-7 cells with an IC50 value of 300 μM, which resulted in DNA damage, inhibition of DNA synthesis, generation of increased intracellular reactive oxygen species, and activation of the programmed cell death mechanism apoptosis. In conclusion, lemonol exhibits promising characteristics that make it a highly favorable candidate for a wide range of applications to contribute significantly to various scientific and pharmacological fields.

Acknowledgments

The authors acknowledge the financial support through the Researchers Supporting Project number (RSP2023R147), King Saud University, Riyadh, Saudi Arabia.

The authors declare no competing financial interest.

References

  1. Crespo R.; Rodenak-Kladniew B. E.; Castro M. A.; Soberon M. V.; Lavarias S. M. Induction of oxidative stress as a possible mechanism by which geraniol affects the proliferation of human A549 and HepG2 tumor cells. Chemico-Biological Interactions 2020, 320, 109029 10.1016/j.cbi.2020.109029. [DOI] [PubMed] [Google Scholar]
  2. Lira M. H. P. D.; Andrade Junior F. P. D.; Moraes G. F. Q.; Macena G. D. S.; Pereira F. D. O.; Lima I. O. Antimicrobial activity of geraniol: an integrative review. Journal of Essential Oil Research 2020, 32 (3), 187–197. 10.1080/10412905.2020.1745697. [DOI] [Google Scholar]
  3. Hosseini S. M.; Hejazian L. B.; Amani R.; SiahchehrehBadeli N. Geraniol attenuates oxidative stress, bioaccumulation, serological and histopathological changes during aluminum chloride-hepatopancreatic toxicity in male Wistar rats. Environmental Science and Pollution Research 2020, 27, 20076–20089. 10.1007/s11356-020-08128-1. [DOI] [PubMed] [Google Scholar]
  4. Lei Y.; Fu P.; Jun X.; Cheng P. Pharmacological properties of geraniol–a review. Planta Medica 2019, 85, 48–55. 10.1055/a-0750-6907. [DOI] [PubMed] [Google Scholar]
  5. Wu Y.; Wang Z.; Fu X.; Lin Z.; Yu K. Geraniol-mediated osteoarthritis improvement by down-regulating PI3K/Akt/NF-κB and MAPK signals: In vivo and in vitro studies. International Immunopharmacology 2020, 86, 106713 10.1016/j.intimp.2020.106713. [DOI] [PubMed] [Google Scholar]
  6. Burrell J. W. K.; Garwood R. F.; Jackman L. M.; Oskay E.; Weedon B. C. L. Carotenoids and related compounds. Part XIV. Stereochemistry and synthesis of geraniol, nerol, farnesol, and phytol. Journal of the Chemical Society C: Organic 1966, 2144–2154. 10.1039/j39660002144. [DOI] [Google Scholar]
  7. Sell C. S.Terpenoids. In Kirk-Othmer Encyclopedia of Chemical Technology ;Wiley, 2006. 10.1002/0471238961.2005181602120504.a01.pub2 [DOI] [Google Scholar]
  8. Zhou J.; Wang C.; Yoon S. H.; Jang H. J.; Choi E. S.; Kim S. W. Engineering Escherichia coli for selective geraniol production with minimized endogenous dehydrogenation. J. Biotechnol. 2014, 169, 42–50. 10.1016/j.jbiotec.2013.11.009. [DOI] [PubMed] [Google Scholar]
  9. de Carvalho K. I. M.; Bonamin F.; dos Santos R. C.; Perico L. L.; Beserra F. P.; de Sousa D. P.; Filho J. M. B.; da Rocha L. R. M.; Hiruma-Lima C. A. Geraniol a flavoring agent with multifunctional effects in protecting the gastric and duodenal mucosa. Naunyn-Schmiedeberg's Arch. Pharmacol. 2014, 387 (4), 355–365. 10.1007/s00210-013-0947-z. [DOI] [PubMed] [Google Scholar]
  10. Inouye S.; Takizawa T.; Yamaguchi H. Antibacterial activity of essential oils and their major constituents against respiratory tract pathogens by gaseous contact. J. Antimicrob. Chemother. 2001, 47 (5), 565–573. 10.1093/jac/47.5.565. [DOI] [PubMed] [Google Scholar]
  11. Su Y. W.; Chao S. H.; Lee M. H.; Ou T. Y.; Tsai Y. C. Inhibitory effects of citronellol and geraniol on nitric oxide and prostaglandin E2 production in macrophages. Planta Medica 2010, 76 (15), 1666–1671. 10.1055/s-0030-1249947. [DOI] [PubMed] [Google Scholar]
  12. Muller G. C.; Junnila A.; Butler J.; Kravchenko V. D.; Revay E. E.; Weiss R. W.; Schlein Y. Efficacy of the botanical repellents geraniol, linalool, and citronella against mosquitoes. Journal of Vector Ecology 2009, 34 (1), 2–8. 10.1111/j.1948-7134.2009.00002.x. [DOI] [PubMed] [Google Scholar]
  13. Torre L. A.; Bray F.; Siegel R. L.; Ferlay J.; Lortet-Tieulent J.; Jemal A. 2015. Global cancer statistics. CA: A Cancer Journal for Clinicians 2015, 65 (2), 87–108. 10.3322/caac.21262. [DOI] [PubMed] [Google Scholar]
  14. Kohn W.; Sham L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140 (4A), A1133–A1138. 10.1103/PhysRev.140.A1133. [DOI] [Google Scholar]
  15. Becke A. D. Density functional thermo chemistry – III: The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
  16. Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
  17. Frisch M. J.; et al. Gaussian 09 ;Gaussian, Inc.: Wallingford, CT, 2009.
  18. Jamróz M. H.Vibrational Energy Distribution Analysis VEDA 4 ;Institute of Nuclear Chemistry and Technology: Warsaw, Poland, 2004. –2010.
  19. Dennington R.; Keith T. A.; Millam J. M.. GaussView , Ver. 6.1; Semichem Inc.: Shawnee Mission, KS, 2016.
  20. Zhurko G.A.; Zhurko D.A.. Chemcraft Program , Version 1.6 (Build 315); 2009. Online at http://www.chemcraftprog.com.
  21. Lu T.; Chen F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33 (5), 580–592. 10.1002/jcc.22885. [DOI] [PubMed] [Google Scholar]
  22. Mitchell K. E.Origin 6.0 ;American Association for the Advancement of Science, 2000.
  23. Mosmann T. Rapid colorimetric assay for cellular growth and survival: application to proliferation and cytotoxicity assays. Journal of Immunological Methods 1983, 65 (1–2), 55–63. 10.1016/0022-1759(83)90303-4. [DOI] [PubMed] [Google Scholar]
  24. Perumal N.; Perumal M.; Kannan A.; Subramani K.; Halagowder D.; Sivasithamparam N. Morin impedes Yap nuclear translocation and fosters apoptosis through suppression of Wnt/β-catenin and NF-κB signaling in Mst1 overexpressed HepG2 cells. Exp. Cell Res. 2017, 355 (2), 124–141. 10.1016/j.yexcr.2017.03.062. [DOI] [PubMed] [Google Scholar]
  25. SPSS Inc . SPSS Base 10.0 for Windows User’s Guide ;SPSS Inc.: Chicago IL, 1999. [Google Scholar]
  26. Zhang S. Y.; Yang C. X.; Shi W.; Yan X. P.; Cheng P.; Wojtas L.; Zaworotko M. J. A chiral metal-organic material that enables enantiomeric identification and purification. Chem. 2017, 3 (2), 281–289. 10.1016/j.chempr.2017.07.004. [DOI] [Google Scholar]
  27. Thirunavukkarasu M.; Balaji G.; Muthu S.; Raajaraman B. R.; Ramesh P. Computational spectroscopic investigations on structural validation with IR and Raman experimental evidence, projection of ultraviolet-visible excitations, natural bond orbital interpretations, and molecular docking studies under the biological investigation on N-Benzyloxycarbonyl-L-Aspartic acid 1-Benzyl ester. Chemical Data Collections 2021, 31, 100622 10.1016/j.cdc.2020.100622. [DOI] [Google Scholar]
  28. Selvaraj S.; Rajkumar P.; Kesavan M.; Mohanraj K.; Gunasekaran S.; Kumaresan S. A combined experimental and theoretical study on 4–hydroxy carbazole by FT-IR, FT-Raman, NMR, UV–visible and quantum chemical investigations. Chemical Data Collections 2018, 17–18, 302–311. 10.1016/j.cdc.2018.10.003. [DOI] [Google Scholar]
  29. Thirunavukkarasu K.; Rajkumar P.; Selvaraj S.; Suganya R.; Kesavan M.; Gunasekaran S.; Kumaresan S. Vibrational (FT-IR and FT-Raman), electronic (UV-Vis), NMR (1H and 13C) spectra and molecular docking analyses of anticancer molecule 4-hydroxy-3-methoxycinnamaldehyde. J. Mol. Struct. 2018, 1173, 307–320. 10.1016/j.molstruc.2018.07.003. [DOI] [Google Scholar]
  30. Rajkumar P.; Selvaraj S.; Anthoniammal P.; Ram Kumar A.; Kasthuri K.; Kumaresan S. Structural (monomer and dimer), spectroscopic (FT-IR, FT-Raman, UV-Vis and NMR) and solvent effect (polar and nonpolar) studies of 2-methoxy-4-vinyl phenol. Chemical Physics Impact 2023, 7, 100257 10.1016/j.chphi.2023.100257. [DOI] [Google Scholar]
  31. Selvaraj S.; Rajkumar P.; Kesavan M.; Thirunavukkarasu K.; Gunasekaran S.; Devi N. S.; Kumaresan S. Spectroscopic and structural investigations on modafinil by FT-IR, FT-Raman, NMR, UV-Vis and DFT methods. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2020, 224, 117449 10.1016/j.saa.2019.117449. [DOI] [PubMed] [Google Scholar]
  32. Selvaraj S.; Ram Kumar A.; Ahilan T.; Kesavan M.; Goncagul S.; Rajkumar P.; Mani M.; Gunasekaran S.; Kumaresan S. Experimental and Theoretical Spectroscopic Studies of the Electronic Structure of 2-Ethyl-2-phenylmalonamide. Physical Chemistry Research 2022, 10 (3), 333–344. 10.22036/PCR.2021.304087.1966. [DOI] [Google Scholar]
  33. Silverstein R. M.; Bassler G. C. Spectrometric identification of organic compounds. J. Chem. Educ. 1962, 39 (11), 546. 10.1021/ed039p546. [DOI] [Google Scholar]
  34. Ram Kumar A.; Kanagathara N.; Selvaraj S. Spectroscopic, Structural and Molecular Docking Studies on N, N-Dimethyl-2-[6-methyl-2-(4-methylphenyl) Imidazo [1, 2-a] pyridin-3-yl] Acetamide. Physical Chemistry Research 2024, 12 (1), 95–107. 10.22036/PCR.2023.387911.2306. [DOI] [Google Scholar]
  35. Kanagathara N.; Thirunavukkarasu M.; Selvaraj S.; Ram Kumar A.; Marchewka M. K.; Janczak J. Structural elucidation, solvent (polar and non-polar) effect on electronic characterization, non-covalent charge interaction nature, topology and pharmacological studies on NLO active l-argininium methanesufonate. J. Mol. Liq. 2023, 385, 122315 10.1016/j.molliq.2023.122315. [DOI] [Google Scholar]
  36. Premkumar S.; Jawahar A.; Mathavan T.; Kumara Dhas M.; Sathe V.G.; Milton Franklin Benial A. DFT calculation and vibrational spectroscopic studies of 2-(tert-butoxycarbonyl (Boc)-amino)-5-bromopyridin. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2014, 129, 74–83. 10.1016/j.saa.2014.02.147. [DOI] [PubMed] [Google Scholar]
  37. Gunasekaran S.; Arun Balaji R.; Kumeresan S.; Anand G.; Srinivasan S. Experimental and theoretical investigations of spectroscopic properties of N-acetyl-5-methoxytryptamine. Canadian J. Anal. Sci. Spectroscopy 2008, 53, 149–162. [Google Scholar]
  38. Selvaraj S.; Rajkumar P.; Kesavan M.; Gunasekaran S.; Kumaresan S. Experimental and theoretical analyzes on structural and spectroscopic properties of monomer and dimeric form of (S)-Piperidine-2-Carboxylic acid: An attempt on medicinal plant. Vib. Spectrosc. 2019, 100, 30–39. 10.1016/j.vibspec.2018.10.008. [DOI] [Google Scholar]
  39. Selvaraj S.; Rajkumar P.; Kesavan M.; Gunasekaran S.; Kumaresan S.; Rajasekar R.; Renuga Devi T.S. Spectroscopic and quantum chemical investigations on structural isomers of dihydroxybenzene. J. Mol. Struct. 2019, 1196, 291–305. 10.1016/j.molstruc.2019.06.075. [DOI] [Google Scholar]
  40. Khan M. U.; Mehboob M. Y.; Hussain R.; Afzal Z.; Khalid M.; Adnan M. Designing spirobifullerene core based three-dimensional cross shape acceptor materials with promising photovoltaic properties for high-efficiency organic solar cells. Int. J. Quantum Chem. 2020, 120 (22), e26377 10.1002/qua.26377. [DOI] [Google Scholar]
  41. Khalid M.; Khan M. U.; Razia E. T.; Shafiq Z.; Alam M. M.; Imran M.; Akram M. S. Exploration of efficient electron acceptors for organic solar cells: rational design of indacenodithiophene based non-fullerene compounds. Sci. Rep. 2021, 11 (1), 19931. 10.1038/s41598-021-99254-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Devi P.; Fatma S.; Parveen H.; Bishnoi A.; Singh R. Spectroscopic analysis, first order hyperpolarizability, NBO, HOMO and LUMO analysis of 5-oxo-1-phenylpyrrolidine-3-carboxylic acid: Experimental and theoretical approach. Indian J. Pure Applied Phys. 2018, 56 (10), 814–829. [Google Scholar]
  43. Kanagathara N.; Senthilkumar K.; Sabari V.; Ragavendran V.; Elangovan S. Structural and Vibrational Investigation of Benzil-(1, 2-Diphenylethane-1, 2-Dione): Experimental and Theoretical Studies. J. Chem. 2022, 2022, 1–11. 10.1155/2022/5968496. [DOI] [Google Scholar]
  44. Senthilkumar K.; Kanagathara N.; Ragavendran V.; Natarajan V.; Marchewka M. K. Quantum chemical computational studies of 1, 3-diammonium propylarsenate: a semi organic crystalline salt. Inorganic and Nano-Metal Chemistry 2022, 52 (10), 1352–1363. 10.1080/24701556.2021.1963282. [DOI] [Google Scholar]
  45. Khalid M.; Khan M. U.; Shafiq I.; Hussain R.; Mahmood K.; Hussain A.; Jawaria R.; Hussain A.; Imran M.; Assiri M. A.; Ali A.; ur Rehman M. F.; Sun K.; Li Y. NLO potential exploration for D−π–A heterocyclic organic compounds by incorporation of various π-linkers and acceptor units. Arabian Journal of Chemistry 2021, 14 (8), 103295 10.1016/j.arabjc.2021.103295. [DOI] [Google Scholar]
  46. Khalid M.; Shafiq I.; Umm-e-Hani; Mahmood K.; Hussain R.; ur Rehman M. F.; Assiri M. A.; Imran M.; Akram M. S. Effect of different end-capped donor moieties on non-fullerenes based non-covalently fused-ring derivatives for achieving high-performance NLO properties. Sci. Rep. 2023, 13 (1), 1395. 10.1038/s41598-023-28118-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Kumar A. R.; Selvaraj S.; Jayaprakash K.S.; Gunasekaran S.; Kumaresan S.; Devanathan J.; Selvam K.A.; Ramadass L.; Mani M.; Rajkumar P. Multi-spectroscopic (FT-IR, FT-Raman, 1H NMR and 13C NMR) investigations on syringaldehyde. J. Mol. Struct. 2021, 1229, 129490 10.1016/j.molstruc.2020.129490. [DOI] [Google Scholar]
  48. Selvaraj S.; Rajkumar P.; Kesavan M.; Gunasekaran S.; Kumaresan S. Experimental and theoretical investigations on spectroscopic properties of tropicamide. J. Mol. Struct. 2018, 1173, 52–62. 10.1016/j.molstruc.2018.06.097. [DOI] [Google Scholar]
  49. Selvaraj S.; Ram Kumar A.; Ahilan T.; Kesavan M.; Gunasekaran S.; Kumaresan S. Multi spectroscopic and computational investigations on the electronic structure of oxyclozanide. Journal of the Indian Chemical Society 2022, 99 (10), 100676 10.1016/j.jics.2022.100676. [DOI] [Google Scholar]
  50. Becke A. D.; Edgecombe K. E. A simple measure of electron localization in atomic and molecular systems. J. Chem. Phys. 1990, 92 (9), 5397–5403. 10.1063/1.458517. [DOI] [Google Scholar]
  51. Rizwana B F.; Muthu S.; Prasana J. C.; Abraham C. S.; Raja M. Spectroscopic (FT-IR, FT-Raman) investigation, topology (ESP, ELF, LOL) analyses, charge transfer excitation and molecular docking (dengue, HCV) studies on ribavirin, Chemical Data. Chemical Data Collections 2018, 17, 236–250. 10.1016/j.cdc.2018.09.003. [DOI] [Google Scholar]
  52. Bader R.F.W.Atoms in Molecule- A Quantum Theory ;Oxford University Press: New York, 1990. [Google Scholar]
  53. Ali A.; Khalid M.; Ashfaq M.; Malik A. N.; Tahir M. N.; Assiri M. A.; Imran M.; de AlcantaraMorais S. F.; Braga A. A. C. Preparation, QTAIM and Single-Crystal Exploration of the Pyrimethamine-Based Co-Crystal Salts with Substituted Benzoic Acids. Chem. Select 2022, 7 (17), e202200349 10.1002/slct.202200349. [DOI] [Google Scholar]
  54. Khalid M.; Ali A.; Abid S.; Tahir M. N.; Khan M. U.; Ashfaq M.; Imran M.; Ahmad A. Facile Ultrasound-Based Synthesis, SC-XRD, DFT Exploration of the Substituted Acyl-Hydrazones: An Experimental and Theoretical Slant towards Supramolecular Chemistry. Chemistry Select 2020, 5 (47), 14844–14856. 10.1002/slct.202003589. [DOI] [Google Scholar]
  55. Tariq S.; Khalid M.; Raza A. R.; Rubab S. L.; Morais S. F. d. A.; Khan M. U.; Tahir M. N.; Braga A. A. C. Experimental and computational investigations of new indole derivatives: A combined spectroscopic, SC-XRD, DFT/TD-DFT and QTAIM analysis. J. Mol. Struct. 2020, 1207, 127803 10.1016/j.molstruc.2020.127803. [DOI] [Google Scholar]
  56. Bader R. F. Atoms in molecules. Accounts of chemical research 1985, 18 (1), 9–15. 10.1021/ar00109a003. [DOI] [Google Scholar]
  57. Bader R. F. W. A Quantum Theory of Molecular Structure and Its Applications. Chem. Rev. 1991, 91 (5), 893–928. 10.1021/cr00005a013. [DOI] [Google Scholar]
  58. Raajaraman B. R.; Sheela N. R.; Muthu S. Investigation on 1-Acetyl-4- (4hydroxyphenyl) piperazine an antifungal drug by spectroscopic, quantum chemical computations and molecular docking studies. J. Mol. Struct. 2018, 1173, 583–595. 10.1016/j.molstruc.2018.07.030. [DOI] [Google Scholar]
  59. KUMAR P S. V.; RAGHAVENDRA V; SUBRAMANIAN V Bader’s theory of atoms in molecules (AIM) and its application to chemical bonding. J. Chem. Sci. 2016, 128 (10), 1527–1536. 10.1007/s12039-016-1172-3. [DOI] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES