Next Article in Journal
In Vitro Selective Growth-Inhibitory Activities of Phytochemicals, Synthetic Phytochemical Analogs, and Antibiotics against Diarrheagenic/Probiotic Bacteria and Cancer/Normal Intestinal Cells
Next Article in Special Issue
The Role of Adaptogens in Prophylaxis and Treatment of Viral Respiratory Infections
Previous Article in Journal
CCL-11 or Eotaxin-1: An Immune Marker for Ageing and Accelerated Ageing in Neuro-Psychiatric Disorders
Previous Article in Special Issue
Characterization of Resveratrol, Oxyresveratrol, Piceatannol and Roflumilast as Modulators of Phosphodiesterase Activity. Study of Yeast Lifespan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biological Evaluation, DFT Calculations and Molecular Docking Studies on the Antidepressant and Cytotoxicity Activities of Cycas pectinata Buch.-Ham. Compounds

1
Department of Pharmacy, International Islamic University Chittagong, Kumira, Chittagong 4318, Bangladesh
2
Department of Theoretical and Computational Chemistry, University of Dhaka, Dhaka 1000, Bangladesh
3
Department of Pharmacy, BGC Trust University Bangladesh, Chittagong 4381, Bangladesh
4
Department of Agricultural Sciences, University of Naples Federico II, 80055 Portici, Italy
5
Department of Biochemistry and Molecular Biology, University of Chittagong, Chittagong 4331, Bangladesh
6
Nutrition and Bromatology Group, Department of Analytical and Food Chemistry, Faculty of Food Science and Technology, University of Vigo—Ourense Campus, E32004 Ourense, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceuticals 2020, 13(9), 232; https://0-doi-org.brum.beds.ac.uk/10.3390/ph13090232
Submission received: 9 August 2020 / Revised: 29 August 2020 / Accepted: 31 August 2020 / Published: 3 September 2020
(This article belongs to the Special Issue Medicinal Plants 2020)

Abstract

:
Cycas pectinata Buch.-Ham. is commonly used in folk medicine against various disorders. The present study investigated the antidepressant and cytotoxicity activity of methanol extract of C. pectinata (MECP) along with quantitative phytochemical analysis by GC-MS method. Here, the GC-MS study of MECP presented 41 compounds, among which most were fatty acids, esters, terpenoids and oximes. The antidepressant activity was assessed by the forced swimming test (FST) and tail suspension test (TST) models. In contrast, MECP (200 and 400 mg/kg) exhibited a significant and dose-dependent manner reduction in immobility comparable with fluoxetine (10 mg/kg) and phenelzine (20 mg/kg). MECP showed a weak toxicity level in the brine shrimp lethality bioassay (ED50: 358.65 µg/mL) comparable to the standard drug vincristine sulfate (ED50: 2.39 µg/mL). Three compounds from the GC-MS study were subjected to density functional theory (DFT) calculations, where only cyclopentadecanone oxime showed positive and negative active binding sites. Cyclopentadecanone oxime also showed a good binding interaction in suppressing depression disorders by blocking monoamine oxidase and serotonin receptors with better pharmacokinetic and toxicological properties. Overall, the MECP exhibited a significant antidepressant activity with moderate toxicity, which required further advance studies to identify the mechanism.

1. Introduction

Depression is a condition characterized by a lowering of the mood and dislike for movement that may distress an individual’s thoughts, conduct, emotions, and comfort [1]. Depressive behavior is additionally connected with suicide, which ranges from 10 and 20 million each year [2,3]. According to the World Health Organization (WHO) report, around 450 million people have a mental disorder, which may rise to 15% by 2020 [4]. In addition, the physical changes additionally happen in extreme, vital, or melancholia or melancholic depression. These comprise sleep deprivation or hypersomnia, modified eating disorders, anorexia and weight reduction and several endocrine dysfunctions with alterations in body temperature. Depressive behavior is the feature of some psychiatric disorders, which may also be caused by somewhat normal life situations; for example, deprivation of sleep, sicknesses, or an adverse effect of drugs and clinical treatments. Patients with major depressive behavior have several symptoms that may reflect in the brain, monoamine synapses or neurotransmitters, explicitly norepinephrine, serotonin, and dopamine [5,6,7].
There are several antidepressant drugs available to treat depression, but the rate of success of first-line therapy for depression [e.g., selective serotonin reuptake inhibitors (SSRIs) and serotonin-norepinephrine reuptake inhibitors (SNRIs)] is low due to several limitations (adverse effects, lower response and the onset of action, etc.), which have been mentioned in several reviews [8]. Thus, it is imperative that new antidepressant drugs demonstrate improvement of these drawbacks. Several phytochemicals (alkaloids, flavonoids, sterol, terpenes) were reported to have an antidepressant effect [9]. Oximes (R1R2C = NOH) are chemicals containing nitrogen produced by organisms in all kingdoms of life [10]. In recent years, oxime derivatives were reported to have several pharmacological activities: cytotoxicity, antibiotic effect, anticonvulsant, antimicrobial, cardiac dysrhythmia, antinociceptive activities [11,12,13,14].
Presently monoamine oxidase A (MAO-A) is useful in the treatment of depression disorders, because MAO metabolizes serotonin or 5-hydroxytryptamine (5-HT) in the central nervous system (CNS) [15]. SSRIs are effective in depression, but due to their limitations, the evaluation of new bioactive substances is a major target for the researchers [16]. Oxime derivatives are reported to have an antidepressant effect [17,18,19], whereas chalcone oxime ethers are reported to have potent inhibitory activity against MAO-B [20]. Our present study design aimed at the evaluation of the biological activity along with a computational study (DFT, molecular docking, ADME/T), where the MAO-A and serotonin receptor are used as a molecular targets for oxime derivatives in depression disorders.
Cycas pectinata Buch.-Ham. (Family: Cycadaceae), commonly known as moniraj or nagmoni, belongs to the genus Cycas [21]. This plant has traditionally been useful for hair growth, curing stomach aches, and curing ulcers [22,23]. Various ethnopharmacological uses in different treatment aspects are documented for Cycas species. Cycas revoluta Thunb. was used for inflammation, vomiting and tonic conditions [24], while Cycas circinalis L. is used for healing wounds and swollen glands. Cycas rumphii Miq. male pollen and cones are reported to have strong narcotic effects [25]. Like in a previous C. pectinata study, a number of fatty acid methyl esters along with other compounds have been reported for C. revolute [26], whereas 16 different bioactive compounds have been reported for C. circinalis [27]. In our previous study, several secondary metabolites from the methanol extract of C. pectinata exhibited the following pharmacological activities, including antioxidant, anti-inflammatory, thrombolytic, anxiolytic, sedative, antinociceptive and antidiarrheal properties [22]. In the present study we report the antidepressant activity along with the cytotoxicity activity of C. pectinate to find a potential lead compound from C. pectinata in alleviating depression disorders by blocking monoamine oxidase (MAOs) and serotonin receptors. To explain this possible mechanism of action of compounds isolated from C. pectinata, we also performed a quantum chemical analysis (DFT calculations) with molecular docking, and ADME/T studies to reveal the potential target(s) for inhibition of the human MAO and serotonin receptors.

2. Results and Discussion

2.1. Qualitative and Quantitative Phytochemical Analysis with Acute Toxicity Study

Phytochemical analysis is useful to evaluate the therapeutic and physiological activities of a plant extract. A qualitative phytochemical screen is performed to determine the presence or absence of secondary plant metabolites. The investigation showed positive results for carbohydrates, alkaloids, phenol, proteins, flavonoids, and saponins (data not shown), which was similar to our previous study that reported similar results [22]. The phytochemical analysis of C. pectinata leaves showed the presence of several phytochemicals. Glycosides are a group of compounds with drug-likeness and numerous studies have suggested that they are a fruitful source of potential drugs. Flavonoids are reported to have anti-inflammatory and anti-cancer activity, whereas tannins possess anti-inflammatory and anti-microbial activity [28]. Phenolic compounds are also present, which possess various physiological functions like anti-aging, anti-inflammation, anti-apoptosis, anti-carcinogenic, inhibition of angiogenesis and enhancement of endothelial function [29].
A total of 66 compounds were identified in the GC-MS analysis, whereas 25 compounds were reported by Tareq et al. [22]. In addition, 41 other compounds are presented in Table 1 and Figure S1, most of which were esters, organic compounds and alcohols. The most abundant compounds along with their retention times were (E)-2-decen-1-ol (20.360), chloroacetic acid 4-pentadecyl ester (20.360), glycerol 1-palmitate (20.009), octadecanoic acid 2-hydroxy-1,3-propanediyl ester (20.009), hexadecanoic acid 2-hydroxy-1-(hydroxymethyl)ethyl ester (20.009), docosanoic acid docosyl ester (19.440), cyclopentadecanone oxime (19.440), and 1-O-(16-hydroxyhexadecyl)-d-mannitol (19.440). These compounds isolated from MECP could help develop a new drug for depression and cancer diseases. The antidepressant activity was evaluated in Swiss albino mice, which required a prior toxicity study. Before starting the experiments and the acute toxicity study of MECP at 400–2000 mg/kg dose was conducted in Swiss albino mice. The methanol extract of C. pectinata leaves was determined to be safe. There was no change of behavioral rush or mortality, morbidity in 8 h observation period of 400, 600, 800, 1000, 2000 mg/kg of MECP doses which were similar to the previous study [22].

2.2. Antidepressant Activity

Anxiety and depression are mental conditions that may recur and are generally undiagnosed and untreated. Physical problems might join these mental conditions, and patients frequently present in medical care centers with physical problems instead of mental situations or problems [30]. Though several antidepressant drugs available, but the rate of success is falling day by day (e.g., SSRIs and SNRIs) [8,31]. Thus, the phytochemical study is a topic of interest for the researcher to evaluate a lead compound to treat depression. Several phytochemicals (alkaloids, flavonoids, sterol, terpenes) are reported to have antidepressant effects [9]. Additionally, a few medicinal plants such as M. angolensis [32], N. sativa [33] R. rosea [34] are reported to have bilateral anxiolytic and antidepressant effects. In our previous study, MECP showed decreased locomotor activity with a significant anxiolytic activity and also a strong binding affinity against the human serotonin receptor (PDV: 5I6X) suggested by the interacted compounds [22]. Here, the antidepressant activity of MECP was evaluated by a tail suspension test (TST) and forced swimming test (FST), which are the most promising models to assess antidepressant activity.
Moreover, TST is proposed to have a higher pharmacological sensitivity as compared to FST. In these models, the immobility time or the period of time when animals stopped struggling was calculated, where the antidepressant agents reduce the immobility time in rodents [35]. Fluoxetine is a SSRI very widely used as an antidepressant agent, which basically works by inhibiting access of serotonin transporter protein into presynaptic serotonin neurons by inhibiting the transporter protein and also has mild activity at the 5-hydroxytryptamine 2A (5HT2A) and 5-hydroxytryptamine 2C (5HT2C) receptors [36]. Also, phenelzine was used in this study is a monoamine oxidase inhibitor (MAOI) that acts by inhibiting MAO activity and afterward raises the neural concentration of neurotransmitters, thus increasing monoamine secretion in the synaptic cleft and alleviating depression [37].
In our study, both doses of MECP showed significant antidepressant activity, leading to a possibility that MECP may act in the presynaptic serotonin neurons by inhibiting serotonin transporter protein and by inhibiting the activity of MAO. In TST, 200 and 400 mg/kg dose exhibited 46.26% and 51.52% inhibition in immobility, whereas the standard drugs fluoxetine and phenelzine exhibited 55.06% and 39.40%, respectively. Additionally, the FST showed significant (p < 0.001) immobility, whereas the 200 and 400 mg/kg exhibited 28.51% and 32.55% inhibition of immobility. The results are presented in Figure 1. The presence of alkaloids and saponins in MECP may be a possible reason for this antidepressant activity as well as the presence of oxime derivatives [17,18,19,38].

2.3. Cytotoxicity Activity

The evaluation of the bioactivity of plant products by the brine shrimp lethality bioassay is an effective, safe and economical method. A good correlation is found in the brine shrimp lethality bioassay with solid human tumors for cytotoxic and pesticidal activity, which is useful for the discovery of active antitumor agents and natural pesticides [39]. This method is also used as a pre-screening test for antitumor research. Generally, the higher the ED50, the lower the toxicity of the extract is and vice versa [40]. In our study, the ED50 of the test samples was calculated using a concentration against the viability of the nauplii. Vincristine sulfate demonstrated the viability of nauplii when the concentration gradually decreased from 10 μg/mL (zero viability) to 0.125 μg/mL (90% viability). MECP has an ED50 of 358.65 µg/mL, which is weakly toxic, whereas the standard drug vincristine sulfate exhibited 2.39 µg/mL (highly toxic). The results are presented in Figure 2. This moderate toxicity level of MECP may be due to the presence of oxime derivatives, which reported to have cytotoxicity and antitumor activity [11,41].

2.4. In Silico Study

2.4.1. Molecular Geometry

The stable configurations of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime; and trans-2-dodecen-1-ol trifluoroacetate obtained from the conformational analysis which has been used for reactivity analysis are shown in Figure 3 with the numbering of atoms. From the structural point of view, these three compounds belongs to the C1 point group symmetry group and hence all the calculated frequencies transform to the same A symmetry species.
The total energies of the three compounds calculated by the B3LYP method are −1845.68068, 718.77081 and −997.04879 Hartree, respectively (Table 2). Among the three compounds trans-2-dodecen-1-ol trifluoroacetate showed a higher dipole moment value. Dipole moments tell us about the charge separation in a molecule. The larger the difference in electronegativity of bonded atoms, the larger the dipole moment [42]. Among the three isolated compound, 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] has higher polarizability value. Generally, polarizability increases as the volume occupied by electrons increases. In atoms, this occurs because larger atoms have more loosely held electrons than smaller atoms with tightly bound electrons [43].

2.4.2. Charges and MESP Calculations

The atomic charges (Mulliken and NBO) play an important role in molecular polarizability, dipole moment, electronic structure, molecular reactivity and a lot of related properties of molecular systems. The charge distributions over the atoms suggest the formation of donor and acceptor pairs involving the charge transferring the molecule. The charges on the atoms of the present 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime]; cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate; were calculated by Mulliken population analysis [44] and NBO charges [45] using B3LYP method with 6-31G+ (d,p) basis set, the tabular representation of the results are presented in Tables S1–S3.
For 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] it can be easily seen that the highest positive Mulliken charge value of 1.710 a.u was accommodated on the Si16 atom that is attached to the pyrrole ring, while in NBO charges the highest positive value was 1.905 a.u on the Si30 atom which connect with O atom. Also, the Mulliken charge with the highest negative value of (−0.862~−0.877) a.u was on the methyl group C atom wherein NBO charges provide the highest negative value of (−1.219~−1.225) a.u on the methyl group C atom. Due to the electron-withdrawing nature of the methyl group, its C atom is pulling electrons towards it.
As for cyclopentadecanone oxime, it showed the highest positive Mulliken charge value of 0.428 a.u accommodated on the H16 atom, which is bonded to the O atom. The highest negative Mulliken charge value of −0.638 a.u belongs to the O15 atom which is attached to the N atom. The natural atomic charges value is in excellent agreement with the highest positive and negative Mulliken charge values for the same atom of the molecule. From the table it can be easily seen regarding the Mulliken charge values for trans-2-dodecen-1-ol trifluoroacetate, the highest positive value of 0.769 a.u was accommodated on the C1 atom which is bond with CF3. This natural atomic charge value also agreed with the obtained result for the same carbon. It shows the highest positive value was 0.983 a.u. In Mulliken charges, the highest negative value of −0.480 a.u is accommodated on the O3 atom. This natural atomic charge value does not also agree with the obtained result for the same atom, whereas it shows the highest negative value of −0.524 a.u is accommodated on O42 atom at molecule. From the charges calculation the highest positive and negative value of Mulliken and NBO charge of atoms sometimes did not agreed with each other due to the two methods used.
The molecular electrostatic potential (MESP) surface [46] from Figure 4 illustrates the molecules’ charge distributions three-dimensionally. This map allows us to visualize variably charged regions of a molecule. The knowledge of the charge distributions can be used to determine how molecules may interact with one another and it is also used to determine the nature of their chemical bonds [47]. The MESP map was checked out by theoretical calculations using the B3LYP/6-31G+ (d,p) level. Molecular electrostatic potential shows the electronic density and is useful in recognizing sites for electrophilic attack, nucleophilic reactions, and hydrogen bonding interactions. Different colors represent the different values of the electrostatic potential at the surface. The negative areas (red, orange and yellow color) of MESP were related to electrophilic reactivity, the positive areas (blue color) ones to nucleophilic reactivity and green color are neutral regions. This figure also provides a visual representation of the chemically active sites and the comparative reactivity of atoms. The computed 3D plot of MESP for the title compounds is depicted in the figure, based on the electron density at different points on the molecule. However, potential values of the three isolated compounds ranges from −6.383 × 10−2 a.u (deepest red) to +6.383 × 10−2 a.u (deepest blue), −5.902 × 10−2 a.u (deepest orange) to +5.902 × 10−2 a.u (deepest blue), −4.638 × 10−2 a.u (deepest red) to +4.638 × 10−2 a.u (deepest blue) respectively.
According to the MESP map in the figure for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate; the negative regions are associated with the O15, O15, and O42 atoms, respectively. Therefore these atom positions are suitable sites for electrophilic attack. Alternatively, only cyclopentadecanone oxime showed a positive region associated with the H16 atom that indicates that this atom can be the most probably involved in nucleophilic processes. Here, 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime]; and trans-2-dodecen-1-ol trifluoroacetate; didn’t show any nucleophilic attack sites. The presence of positive and negative binding sites in cyclopentadecanone oxime may result in good interactions with proteins in biological systems.

2.4.3. FMOs and Global Descriptors

The frontier molecular orbitals, HOMO and LUMO, are the most important orbitals in a molecule. They play an important role in the optical and electric properties, as well as in quantum chemistry and the UV–Vis spectra [48]. The highest occupied molecular orbital (HOMO), represents the ionization potential of the molecule and lowest occupied molecular orbital (LUMO), corresponding electron affinity value is called the frontier molecular orbitals (FMOs) showed in Figure 5 were calculated at the B3LYP/6-31G+ (d,p) level for the three isolated compounds.
These orbitals determine the way how the molecule interacts with other species and give information about the reactivity/stability of specific regions of the molecule. The energy of HOMO characterizes the electron-donating ability of a molecule, while LUMO energy determines the ability to accept an electron. Therefore, higher values of EHOMO indicate a better tendency towards the donation of an electron. From Figure 5, trans-2-dodecen-1-ol trifluoroacetate is the better molecule which has the ability to accept electrons while the energy value of HOMO (EHOMO = −7.39524 eV) that allows it to be the best electron donor molecule. The energy gap between the HOMO and LUMO is very important in determining a molecule’s chemical reactivity. A high value of the energy gap indicates that the molecule shows high chemical stability; indicates a hard molecule, while a small HOMO-LUMO gap means small excitation energies to the manifold of excited states, and action as a soft molecule. Among three isolated compounds, 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] shows the lowest energy gap indicating it is more reactive than the two other compounds.
Using Koopmans’ theorem [49,50] (I) and (A) values can be correlated with the frontier orbitals by the relation: I = −EHOMO and A = −ELUMO. Ionization potential (I) is defined as the amount of energy needed to remove an electron from a molecule. High ionization energy indicates high stability, chemical inertness and small ionization energy indicating high reactivity of the atoms and molecules. Trans-2-Dodecen-1-ol, trifluoroacetate has the lowest ionization potential value (I = 7.39524 eV), which indicates that it is the best electron donor. The electronic affinity (A) is defined as the energy released when an electron is added to a neutral molecule. A molecule with high (A) values tends to accept electrons easily. From Table 3 it is clear that 5-chloro-1-(trimethylsilyl)-1H-Indole-2,3-dione 3-[O-(trimethylsilyl)oxime] is the most reactive. The global chemical reactivity descriptors such as chemical potential (µ), electronegativity (χ), hardness (η), softness (S), and electrophilicity index (ω) which were calculated from the HOMO and LUMO energies were obtained at the level of theory B3LYP/6-31G+ (d,p) and are incorporated in Table 3.
According to these parameters, the chemical reactivity varies with the structural configuration of the molecules. Global reactivity descriptors such as chemical potential denote as (µ = -χ), the absolute electronegativity (χ) is given by the relation (χ = (IP + EA)/2), global hardness and global softness (S) are defined as (η = (ELUMO − EHOMO)/2) and (S = 1/2η), the electrophilicity (ω) can be calculated using the electronic chemical potential and the chemical hardness (ω = µ2/2η) [51,52,53,54,55]. Hardness (η) and softness (S) are useful concepts for understanding the behavior of chemical systems. A hard molecule has a large energy gap and a soft molecule has a small energy gap [56]. Therefore, soft molecules will be more polarizable than hard molecules. From the established theoretical calculations cyclopentadecanone oxime has the highest hardness value (η = 3.16210 eV), which indicates that it is the hardest molecule. 5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] has the highest softness (S = 0.53083eV), so it is the softest molecule. The chemical potential µ (eV) measures the escaping tendency of an electron and it can be associated with the molecular electronegativity [57] then, as µ becomes more negative, it is more difficult to lose an electron but easier to gain one. As shown in Table 3, trans-2-dodecen-1-ol trifluoroacetate is the least stable and the most reactive among all the compounds. Electronegativity (χ), represents the ability of molecules to attract electrons. The (χ) values displayed in Table 3 show that cyclopentadecanone oxime; has the higher electronegativity (4.47546 eV) value compared to all the other molecules. Electrophilicity (ω), that gives an idea of the stabilization energy when the system gets saturated by electrons, which come from the external environment. This reactivity information shows if a molecule is capable of donating charges. A good, more reactive nucleophile is characterized by a lower value of (ω), while higher values indicate the presence of a good electrophile. Our results indicate that cyclopentadecanone oxime has lower values of (ω), so that compound is a good nucleophile, whereas trans-2-dodecen-1-ol, trifluoroacetate is a good electrophile.

2.4.4. Vibrational Spectral Analysis

The vibrational band assignments had been performed based on the normal coordinate analysis. Internal coordinates of three isolated compounds had constructed according to Pulay’s recommendations [58]. The calculated wavenumbers were selectively scaled according to the scaled quantum mechanical (SQM) method recommended by Rauhut and Pulay [59] using a scale factor with the root mean square (RMS) wavenumber error, which is in the reasonable limit for proper assignment. The observed FT-IR and simulated theoretical spectra calculated at the B3LYP/6-31G+ (d, p) basis set are shown in Figure S2. The calculated wavenumbers and their assignments are also presented in Table 4. The detailed analyses of vibrational wavenumbers for various functional groups are discussed below.

2.4.5. Hydroxyl (O–H) Group Vibrations

Bands due to O–H stretching are of medium to strong intensity in the infrared spectrum, although it may be broad. For solids, liquids and concentrated solutions a broad band of less intensity is normally observed [60]. The very weak FT-IR band at 3696 cm−1 is assigned to the O–H stretching vibrations. Normally free O–H stretching vibrations appeared around 3600 cm−1 for phenols [61]. The observed broad intense IR band for cyclopentadecanone oxime at corresponds to O–H stretching mode, which is calculated at 3884 cm−1.

2.4.6. C-H Vibrations

Aromatic compounds commonly exhibit multiple weak bands in the 3100–3000 cm−1 region due to aromatic C-H stretching vibrations and also in-plane bending vibrations generally lie in the range 1000–1300 cm−1 [62]. The bands appearing at (3081 ~ 3099) cm−1, (3100 ~ 3200) cm−1, (3023 ~ 3069) cm−1 for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate; respectively in the FT-IR spectrum are assigned to C–H ring stretching vibrations. In the present study, the C–H in-plane bending vibrations of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] and trans-2-dodecen-1-ol trifluoro-acetate is identified at 1137 cm−1 and 989 cm−1 at the B3LYP methods are assigned to C–H in-plane bending vibrations.

2.4.7. Methylene (H-C-H) group Vibrations

Methyl groups are generally referred to as electron donating substituents in an aromatic ring system. Whenever a methyl group is present in a compound, it gives rise to asymmetric and symmetric stretching vibrations [63]. The asymmetric stretch is usually at a higher wavenumber than the symmetric stretch. The asymmetric stretching vibrations of CH3 are expected in the 2925–3000 cm−1 region and the symmetric CH3 stretching vibrations in the 2905–2940 cm−1 range [64,65]. The predicted asymmetric and symmetric stretching vibrations for CH3 are at (2927~3017) cm−1, (3051~3093) cm−1, (2920~3001) cm−1 and (2916 ~ 2919) cm−1, (3007~3019) cm−1, (2805~2917) cm−1 for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate, respectively. Furthermore, the observed peaks at (1464~1459) cm−1, (1262 cm−1~1267) cm−1, 1237 cm−1 can be assigned to the scissoring, twisting and wagging modes of CH3 and CH2 groups in aliphatic chains, respectively [66]. The predicted scissoring, wagging and twisting vibrations for CH2 are at (1459~1499) cm−1, (1346~1373) cm−1, (1246~1287) cm−1 and (1407~1506) cm−1, (1324~1373) cm−1, (1287~1312) cm−1 1 for cyclopentadecanone oxime; and trans-2-dodecen-1-ol trifluoroacetate, respectively.

2.4.8. C-N Vibrations

The identification of C = N vibrations is a difficult task since mixing of vibrations is possible in this region. Silverstein et al. [67] assigned the C = N stretching absorption in the 1690–1640 cm−1 range for aromatic amines. The present work shows that the theoretically computed value of C = N stretching vibrations band observed at 1593 cm−1 and 1759 cm−1 in the FT-IR spectrum for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] and cyclopentadecanone oxime, respectively.

2.4.9. C=C Vibrations

The phenyl ring CC stretching vibrations are generally observed between 1600–1400 cm−1 [68], in which the bands between 1600–1500 cm−1 are assigned to C=C stretching and the rest to C-C stretching, even though no such distinction is present within the ring. In the present study, the bands observed at (1555~1579) cm−1 and (1680~1689) cm−1 are assigned to C=C for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] and trans-2-dodecen-1-ol trifluoroacetate, respectively.

2.4.10. Carbonyl (C=O) Group Vibration

The C=O stretching vibrations give rise to the characteristic bands in IR spectra, and the intensity of these bands can increase owing to the conjugation or formation of hydrogen bonds. The C=O stretching of ketones is expected in the region 1760–1730 cm−1 [69]. C=O stretching mode is not an independent vibrational mode because of it coupled with the vibrations of adjacent groups. The FT-IR band with strong intensity at 1717 cm−1 alone was assigned to the carbonyl stretching mode of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime].

2.4.11. NMR Analysis

After the optimization of molecular geometry of the three isolated compounds the 1H and 13C nuclear magnetic resonance (NMR) chemical shift values calculated at the B3LYP/6-31G+ (d,p) level in chloroform solvents by comparing their observed values in CDCl3 solvent with respect to TMS as an internal reference [70]. The theoretically calculated 1H- and 13C-NMR chemical shift values are presented in Table 5 and Table 6. The theoretically determined 1H-and 13C-NMR spectra are shown in Figures S3 and S4, respectively.
The 1H atoms chemical shift values of 1H-Indole-2,3-dione, 5-chloro-1-(trimethylsilyl)-, 3-[O-(trimethylsilyl)oxime] are divided into two ranges; the first range is approximately 0~6.5 ppm, the second range is around 0~−0.956 ppm. The first group is due to the H atoms in the benzene ring and methyl group. The second group is due to the H atoms in the methyl group attached to Si atoms. Also, 1H atoms chemical shift values of cyclopentadecanone, oxime divide into two ranges; the first one is around 0~4.5 ppm, the second one is greater than 4.5 ppm. The first group is due to the H atoms in the cyclic alkyl chain and that atoms have slightly positive charges. The highest chemical shift was found for H16 atom which associated with the O atom. Lastly, the chemical shift values of 1H atoms for trans-2-Dodecen-1-ol, trifluoroacetate are divided into two ranges; the first one is around 0~5 ppm, the second one is greater than 5 ppm. Frist range chemical shift values determined those H atoms in the alkyl chain and showed a slightly positive charge. The highest chemical shifts were found for H35, H36 atoms which associated with the C atom nearly O atom. Due to different chemical atmospheres at various regions around the H atoms for 1H-Indole-2,3-dione, 5-chloro-1-(trimethylsilyl)-, 3-[O-(trimethylsilyl)oxime]; cyclopentadecanone, oxime; and trans-2-Dodecen-1-ol, trifluoroacetate the chemical shift inequality had originated.
The calculated 13C chemical shift values of for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] are in the −15~143 ppm range. This range is divided into two parts; the first range is greater than 100 ppm for C7, C8, C1, C4, C5, C2, C3 atoms which are located in the benzene ring. The second range is less than 100 ppm for C6, C17, C33, C19, C18, C32, C31 atoms that alkyl chain carbons attached to silicon atoms. Also, the 13C chemical shift values of cyclopentadecanone oxime are in the 11~147 ppm range. This range divided into two part; firstly 146.55 ppm was found for O bonded C12 atom and below 100 ppm corresponds to the C11, C10, C8, C36, C1, C13, C3, C4, C31, C5, C6, C7, C2, C9 atoms in the cyclic alkyl chain. Finally, the trans-2-dodecen-1-ol trifluoroacetate chemical shift values are found in the 3~152 ppm range. The highest chemical shift values were found at 151.74, 129.32, 121.59 and 105.65 ppm for the C2, C6, C1, and C5 atoms that are bonded with a highly negative charge O atom. Besides, less than 100 ppm values are found for the C4, C7, C13, C12, C11, C10, C9, C8, C14 and C15 atoms that are located in the straight alkyl chain.

2.4.12. Molecular Docking Study

Computer-aided drug design (CADD) plays a significant role in developing new drugs. There are mainly two types of drug design methods available, namely: structure-based and ligand-based drug design [71]. In our previous study, we used ligand-based interactions to select a lead compound with sedative activity, which exhibited a significant binding affinity towards the human serotonin receptor (PDB: 5I6X) [22]. Here, the MAO receptor is used because MAO-A is generally targeted to treat depression and anxiety, whereas MAO-B useful for Alzheimer’s disease (AD) and Parkinson’s disease [72]. As oxime derivatives were reported to have antidepressant effects [17,18,19], in our present study, human monoamine oxidase A (PDB: 2Z5X) was used for a molecular docking study of antidepressant activity. The antidepressant activity is presented in Table 7. In the present study, cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate showed the highest and lowermost binding affinity against human monoamine oxidase A (PDB: 2Z5X), with docking scores of −4.333 kcal/mol and −3.155 kcal/mol, respectively. 5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] did not show any interaction, whereas the standard drug phenelzine showed −5.324 kcal/mol binding affinity. Cyclopentadecanone oxime interacted with the monoamine oxidase A (PDB: 2Z5X) by one π-alkyl interaction to Phe 112. The interaction of the compounds is presented in Figure S5.
Human serotonin receptor (PDB: 5I6X) used also for the molecular docking study, where cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate exhibited the highest and lowermost binding affinity, with docking scores of −6.537 kcal/mol and −2.387 kcal/mol, respectively The standard drug fluoxetine shows a −9.07 kcal/mol interaction. Cyclopentadecanone oxime interacted with the human serotonin receptor (PDB: 5I6X) by one H-bond to Asp 98 and one alkyl interaction to Ile 172. The interaction of the compounds is presented in Figure S6.
The molecular docking study of cytotoxicity activity was performed against the human estrogen receptor (PDB ID: 1ERR) and epidermal growth factor receptor tyrosine kinase (PDB ID: 1M17). Cyclopentadecanone oxime gave a −7.685 kcal/mol and −4.59 kcal/mol binding interaction against the human estrogen receptor (PDB ID: 1ERR) and epidermal growth factor receptor tyrosine kinase (PDB ID: 1M17), whereas the standard drug vincristine sulfate exhibited −3.896 kcal/mol and −3.85 kcal/mol interactions, respectively. Cyclopentadecanone oxime interacted with the human estrogen receptor (PDB: 1ERR) through one H-bond to Glu 353, one π-alkyl interaction to Phe 404 and an alkyl-interaction to Leu 346. In addition cyclopentadecanone oxime interacted with the epidermal growth factor receptor tyrosine kinase (PDB ID: 1M17) through one H-bond to Met 769, and two alkyl-interactions to Val 702 and Leu 820. The interactions of the compounds are presented in Figures S7 and S8.

2.4.13. ADME/T and Toxicological Properties Analysis

ADME properties and drug toxicity are important in preventing the early introduction of drugs into the commercial market. From a business point of view, it is necessary to remove the poor pharmacokinetic profile compounds, which reduces the cost of the drug development stage. As a result, over the previous decade, ADME/T screening has been applied in the early drug discovery phase [73]. The selected isolated compounds from the methanol extract of C. pectinata were subjected to the ADME/T profiling by following Lipinski’s (Rule of Five) [74] and Veber’s rules [75]. The three compounds 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl) oxime]; cyclopentadecanone oxime; and trans-2-dodecen-1-ol trifluoroacetate satisfy Lipinski’s Rule of Five, whereas trans-2-dodecen-1-ol trifluoroacetate violated Veber’s rules (Table 8).
In the toxicological study, 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] did not exhibit any risk of toxicity, whereas cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate showed risks of Ames toxicity and carcinogenic effect, respectively (Table 9). Aside from these effects, all three compounds can be considered as lead compounds with antidepressant and cytotoxicity activity.

3. Materials and Methods

3.1. Chemicals

Fluoxetine (Square Pharmaceuticals Ltd., Dhaka, Bangladesh), phenelzine (Ranbaxy Laboratories, Haryana, India), Tween-80 (Sigma Aldrich Co., St. Louis, MO, USA), and vincristine sulfate (2 mg/vial) (Beacon Pharmaceuticals Ltd. Dhaka, Bangladesh) were purchased from a local trader. All other chemicals were analytical grade.

3.2. Plant Materials and Preparation of Crude Extract

The details of the C. pectinata leaves (MECP) plant material were described in our earlier study [22]. The freshly collected leaves were ground into a coarse powder using a grinder (NOWAKE, Hokuto, Japan). The maceration of powder and methanol solvent was followed in a 1:4 ratio, with filtration by Whatman filter paper (#1) after seven days. The filtration was followed by evaporation in a water bath (40 °C) to obtain a crude extract. The crude extract was kept under refrigeration at 4 °C until further use.

3.3. Experimental Animals

The average weight of 25–35 g of six-seven weeks old Swiss albino mice of both sexes was obtained from the animal house of Department of Pharmacy, International Islamic University of Chittagong (IIUC), Chittagong, Bangladesh. The animals were adapted with the laboratory condition (room temperature 25 ± 2 °C, relative humidity 55–60%) by supplying food pellets and water. For the use of the experiment, all the animals were adapted for 14 days with laboratory conditions. The study was approved by the Institutional Animal Ethical Committee, Department of Pharmacy, International Islamic University Chittagong, Bangladesh, according to governmental guidelines under the reference (Pharm/p&d/138/13-′19,22/12/2019) [76].

3.4. GC-MS (Gas Chromatography-Mass Spectroscopy) Analysis of MECP

The detailed gas chromatography-mass spectroscopy (GC-MS) analysis of the methanol extract of C. pectinata leaves (MECP) were described in the earlier study of Tareq et al. [22].

3.5. Acute Toxicity Study

The acute oral toxicity of methanol extract of C. pectinata was determined by the OECD (2002) guidelines No. 423 method [77]. Mice were divided into six groups, where each group contained five animals. The first group received 1% Tween-80 in normal saline. The other groups were received 400, 600, 800, 1000, 2000 mg/kg of MECP dose. Then all the animals were observed for 8 h to detect early symptoms such as behavioral changes or mortality, morbidity and later for 3 days.

3.6. Phytochemical Screening

In the preliminary phytochemical screening of freshly prepared methanol leaves crude extract was qualitatively tested for the determination of carbohydrates, alkaloids, glycosides, tannins, terpenoids, flavonoids, and saponins [78,79].

3.7. Antidepressant Activity

3.7.1. Experimental Design for Anti-Depressant Activity

The antidepressant activity of the extract evaluated by the tail suspension test and forced swimming test. The mice were divided into four groups (n = 5). Administration of extract/control to the animals was followed after 60 min prior to study [80,81]:
  • Group I: Negative control received 1% Tween-80 (10 mL/kg, b.w.) orally
  • Group II: Positive control phenelzine received 20 mg/kg b.w. I.P.
  • Group III: Positive control fluoxetine received 10 mg/kg b.w. I.P.
  • Group IV: Received MECP 200 mg/kg b.w. orally
  • Group V: Received MECP 400 mg/kg b.w. orally

3.7.2. Tail Suspension Test (TST)

The antidepressant activity of MECP was executed by the method described by Steru et al. [80]. The treatment was followed as described in Section 3.7.1. After 60 min of treatment, each mouse was suspended by using adhesive tape at the tip of the tail over the rim of a box. Then the immobility time was recorded from the 6 min suspended period, whereas the first 2 min for initial adjustment and last 4 min for immobility time:
Inhibition   ( % ) = A B A   × 100
where, A = immobile time in the control group; B = immobile time in the test group.

3.7.3. Forced Swimming Test (FST)

The antidepressant activity of MECP was evaluated by the forced swimming test, as described by Porsolt et al., [81]. A glass box (25 × 15 × 25 cm3) filled to 15 cm with water (25 ± 2 °C) was utilized as a test apparatus for swimming. The treatment was followed as described in Section 3.7.1. After 60 min of treatment, each mouse was forced to swim in the apparatus. The immobility time was calculated from the 6 min swimming period. When the mice stopped struggling and remained suspended in water was considered as the immobility time and the period is recorded.
Inhibition   ( % ) = A B A   × 100
where, A = immobile time in the control group; B = immobile time in the test group.

3.8. Brine Shrimp Lethality Bioassay

The brine shrimp lethality bioassay was followed to evaluate the cytotoxicity of methanol extract of C. Pectinata leaves by Meyer et al. [39]. In 1000 mL distilled water, 38 g NaCl was dissolved to prepare artificial seawater. NaOH was added to maintain the pH at 8.0. Then serially diluted concentrations of 50, 100, 200, 400, 600 and 800 μg/mL were obtained. Vincristine sulfate used as a positive control as the preceding method in a serial concentration dilution 0.125, 0.25, 0.5, 1, 5 and 10 μg/mL. Then ten matured live shrimp were placed in all test tubes at room temperature (25 ± 1 °C) and after 24 h, each test tube was assessed, and the number of alive nauplii was counted and recorded.
% of viability = (Nl/N0) ×100
where, N0 = Number of nauplii taken; Nl = Number of nauplii alive

3.9. In Silico Study

3.9.1. Quantum Chemical Analysis

Quantum chemical analysis was performed with the Gaussian 09 software package [82] via the Gauss view 6.0.10 [83] molecular visualization program on a Pentium IV/3.02Hz personal computer. The selected isolated compounds 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl) oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate were fully optimized at the level of density functional theory (DFT) using the B3LYP with the 6-31G+ (d,p) basis set. The minima of the potential energy hypersurfaces were considered to be the stationary points and confirmed from the absence of any imaginary frequency. Electronic properties, such as HOMO-LUMO energies, molecular electrostatic potential (MESP) were calculated using the B3LYP method, based on the optimized structure in the gas phase. Furthermore, Mulliken and natural bond orbital (NBO) charges and global reactivity descriptors of the proposed compounds were analyzed. Calculated vibrational frequencies were multiplied by a suitable scaling number (0.964) [84] to better match experimental frequencies. Besides, the 1H and 13C nuclear magnetic resonance (NMR) chemical shift [85] (with respect to a TMS reference and chloroform solution) of the proposed compounds were also carried out by GIAO method in same method and level of basis set.

3.9.2. Molecular Docking Study

The optimized structure of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl) oxime], cyclopentadecanone oxime; and trans-2-dodecen-1-ol trifluoroacetate were subjected to a molecular docking study according to Sastry et al. as briefly explained in Adnan et al. [86,87]. The proteins used for the docking study were retrieved from the Protein data bank (https://www.rcsb.org/structure/): human monoamine oxidase A (PDB ID: 2Z5X), human serotonin (PDB ID: 5I6X), human estrogen receptor (PDB ID: 1ERR), and epidermal growth factor receptor tyrosine kinase (PDB ID: 1M17) [88]. The molecular docking study was performed using Schrödinger (Maestro v11.1).

3.9.3. ADME/T and Toxicological Properties Analysis

The optimized structures of 5-chloro-1-(trimethylsilyl)-1H-Indole-2,3-dione 3-[O-(trimethylsilyl) oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate were subjected to ADME/T following the rules of Lipinski (Rule of Five) [74] and Veber [75]. In addition, the toxicological properties were analyzed by the admetSAR (http://lmmd.ecust.edu.cn/admetsar2/). The ADME/T analysis was evaluated by SwissADME (http://www.swissadme.ch/) [89].

3.10. Statistical Analysis

The values are shown as mean ± standard error mean (SEM). * p < 0.001 statistical significance was calculated by one-way ANOVA (Dunnett’s test) using the GraphPad Prism (version 8.4.) software (San Diego, CA, USA).

4. Conclusions

This study reports that methanol leaves extract of C. pectinata could be a potential source of compounds with antidepressant and cytotoxicity activity due to the presence of secondary metabolites. In addition, the computational study of the oxime derivatives by DFT and molecular docking study unveiled better binding interaction against the MAO and serotonin receptor with good pharmacokinetic and toxicological properties. Further advanced studies are recommended to identify the mechanism of action of C. pectinata.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1424-8247/13/9/232/s1, Table S1: Mulliken atomic charges and NBO charges at different atoms in gas phase of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] computed by B3LYP/methods with 6-31G+ (d,p) basis set. Table S2: Mulliken atomic charges and NBO charges at different atoms in gas phase of cyclopentadecanone oxime computed by B3LYP/methods with 6-31G+ (d,p) basis set. Table S3: Mulliken atomic charges and NBO charges at different atoms in gas phase of trans-2-dodecen-1-ol trifluoroacetate computed by B3LYP/methods with 6-31G+ (d,p) basis set. Figure S1: Total ionic chromatogram (TIC) of MECP by GC-MS. Figure S2: Fourier transform Infrared- (FT-IR) spectrum of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively in the wavenumber range 4000 – 0 cm-1. Figure S3: Calculated 1H-NMR isotropic chemical shift spectrum of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively in chloroform solvent. Figure S4: Calculated 13C NMR isotropic chemical shift spectrum of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively in chloroform solvent. Figure S5: 3D and 2D interactions of cyclopentadecanone oxime (A); trans-2-dodecen-1-ol trifluoroacetate (B); phenelzine (C), against the human monoamine oxidase A (PDB: 2Z5X) for antidepressant activity; Figure S6: 3D and 2D interactions of cyclopentadecanone oxime (A); trans-2-dodecen-1-ol trifluoroacetate (B); fluoxetine (C), against the human serotonin receptor (PDB: 5I6X) for antidepressant activity; Figure S7: 3D and 2D interactions of cyclopentadecanone oxime (A); trans-2-dodecen-1-ol trifluoroacetate (B); vincristine sulfate (C), against the human estrogen receptor (PDB ID: 1ERR) for cytotoxicity activity; Figure S8: 3D and 2D interactions of cyclopentadecanone oxime (A); trans-2-dodecen-1-ol trifluoroacetate (B); vincristine sulfate (C), against the epidermal growth factor receptor tyrosine kinase (PDB ID: 1M17) for cytotoxicity activity.

Author Contributions

J.R., and A.M.T.; conceptualization, planning, designing, investigation, data analysis, software, manuscript writing, M.M.H.; designing, manuscript writing, data analysis, S.A.S.; planning, designing, investigation, data analysis, DFT calculation, M.N.I.; designing, investigation, data analysis, A.B.M.N.U., M.H.; and M.S.N. data curation, data analysis, in silico investigation, A.S.M.A.R., T.B.E., M.H.A., and J.S.-G.; monitoring, visualization, supervision, A.S.M.A.R., R.C. and J.S.-G., together with writing—review & editing, correspondence. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Center for Research and Publication (CRP) grant (IRG 180111), International Islamic University Chittagong (IIUC).

Acknowledgments

This research was supported by the Department of Pharmacy, International Islamic University Chittagong-4318, Bangladesh. The authors also express gratitude to Department of Theoretical and Computational Chemistry, University of Dhaka, Dhaka-1000, Bangladesh for logistics support and providing the Gaussian 09 software package.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MECPmethanol extract of C. pectinata leaves
GC-MSGas Chromatography-Mass Spectroscopy
IPintraperitoneal
b.w.body weight
MAOmonoamine oxidase
NMRnuclear magnetic resonance
DFTdensity functional theory
ADME/Tabsorption, distribution, metabolism, excretion, and toxicity
PDBprotein data bank
SEMstandard error mean
ANOVAone-way analysis of variance
FMOsfrontier molecular orbitals
HOMOhighest occupied molecular orbital
LUMOlowest unoccupied molecular orbital
MESPmolecular electrostatic potential
NBOnatural bond orbital

References

  1. Rosenbaum, D.; Hagen, K.; Deppermann, S.; Kroczek, A.M.; Haeussinger, F.B.; Heinzel, S.; Berg, D.; Fallgatter, A.J.; Metzger, F.G.; Ehlis, A.-C. State-dependent altered connectivity in late-life depression: A functional near-infrared spectroscopy study. Neurobiol. Aging 2016, 39, 57–68. [Google Scholar] [CrossRef] [PubMed]
  2. Calvó-Perxas, L.; Vilalta-Franch, J.; Turró-Garriga, O.; López-Pousa, S.; Garre-Olmo, J. Gender differences in depression and pain: A two year follow-up study of the Survey of Health, Ageing and Retirement in Europe. J. Affect. Disord. 2016, 193, 157–164. [Google Scholar] [CrossRef] [PubMed]
  3. Gadassi, R.; Mor, N. Confusing acceptance and mere politeness: Depression and sensitivity to Duchenne smiles. J. Behav. Ther. Exp. Psychiatry 2016, 50, 8–14. [Google Scholar] [CrossRef] [PubMed]
  4. Ridout, K.K.; Ridout, S.J.; Price, L.H.; Sen, S.; Tyrka, A.R. Depression and telomere length: A meta-analysis. J. Affect. Disord. 2016, 191, 237–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Perviz, S.; Khan, H.; Pervaiz, A. Plant Alkaloids as an Emerging Therapeutic Alternative for the Treatment of Depression. Front. Pharmacol. 2016, 7, 28. [Google Scholar] [CrossRef] [Green Version]
  6. Gold, P.W.; Goodwin, F.K.; Chrousos, G.P. Clinical and Biochemical Manifestations of Depression. N. Engl. J. Med. 1988, 319, 413–420. [Google Scholar] [CrossRef]
  7. Tondo, L.; Isacsson, G.; Baldessarini, R.J. Suicidal behaviour in bipolar disorder: Risk and prevention. CNS Drugs 2003, 17, 491–511. [Google Scholar] [CrossRef]
  8. Alexander, R.C.; Preskorn, S. Clinical pharmacology in the development of new antidepressants: The challenges. Curr. Opin. Pharmacol. 2014, 14, 6–10. [Google Scholar] [CrossRef]
  9. García-Ríos, R.I.; Mora-Pérez, A.; Ramos-Molina, A.R.; Soria-Fregozo, C. Neuropharmacology of Secondary Metabolites from Plants with Anxiolytic and Antidepressant Properties. In Behavioral Pharmacology-From Basic to Clinical Research; IntechOpen: London, UK, 2020. [Google Scholar]
  10. Sørensen, M.; Neilson, E.H.J.; Møller, B.L. Oximes: Unrecognized Chameleons in General and Specialized Plant Metabolism. Mol. Plant 2018, 11, 95–117. [Google Scholar] [CrossRef] [Green Version]
  11. Hertiani, T.; Edrada-Ebel, R.; Ortlepp, S.; van Soest, R.W.M.; de Voogd, N.J.; Wray, V.; Hentschel, U.; Kozytska, S.; Müller, W.E.G.; Proksch, P. From anti-fouling to biofilm inhibition: New cytotoxic secondary metabolites from two Indonesian Agelas sponges. Bioorg. Med. Chem. 2010, 18, 1297–1311. [Google Scholar] [CrossRef]
  12. Wei, K.; Wang, G.-Q.; Bai, X.; Niu, Y.-F.; Chen, H.-P.; Wen, C.-N.; Li, Z.-H.; Dong, Z.-J.; Zuo, Z.-L.; Xiong, W.-Y.; et al. Structure-Based Optimization and Biological Evaluation of Pancreatic Lipase Inhibitors as Novel Potential Antiobesity Agents. Nat. Prod. Bioprospect. 2015, 5, 129–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Karakurt, A.; Dalkara, S.; Özalp, M.; Özbey, S.; Kendi, E.; Stables, J.P. Synthesis of some 1-(2-naphthyl)-2-(imidazole-1-yl) ethanone oxime and oxime ether derivatives and their anticonvulsant and antimicrobial activities. Eur. J. Med. Chem. 2001, 36, 421–433. [Google Scholar] [CrossRef]
  14. Schenone, S.; Bruno, O.; Ranise, A.; Bondavalli, F.; Filippelli, W.; Falcone, G.; Rinaldi, B. O-[2-Hydroxy-3-(dialkylamino) propyl]ethers of (+)-1,7,7-trimethyl bicyclo[2.2.1]heptan-2-one oxime (camphor oxime) with analgesic and antiarrhythmic activities. IL Farmaco 2000, 55, 495–498. [Google Scholar] [CrossRef]
  15. Sivaraman, D.; Vignesh, G.; Selvaraj, R.; Dare, B.J. Identification of potential monoamine oxidase inhibitor from herbal source for the treatment of major depressive disorder: An in-silico screening approach. Der. Pharma. Chem. 2015, 7, 224–234. [Google Scholar]
  16. Rudorfer, M.V.; Potter, W.Z. Antidepressants. Drugs 1989, 37, 713–738. [Google Scholar] [CrossRef]
  17. Bozdağ, O.; Gümüşel, B.; Demirdamar, R.; Büyükbingöl, E.; Rolland, Y.; Ertan, R. Synthesis of some novel oxime ether derivatives and their activity in the ‘behavioral despair test’. Eur. J. Med. Chem. 1998, 33, 133–141. [Google Scholar] [CrossRef]
  18. Davrinche, C.; Nguyen-Tri-Xuong, E.; El Hamad, Y.; Reynaud, P.; Rinjard, P.; Tran, G. Amide-oximes et hydroximates benzodioxaniques: Synthèse de nouveaux composés et étude en neuropsycho-pharmacologie. Eur. J. Med. Chem. 1992, 27, 765–778. [Google Scholar] [CrossRef]
  19. Ertan, R.; BozdaĞ, O.Y.A.; KesİCİ, B.; Palaska, E.; Ertan, M. Studies on the synthesis and antidepressant activity of some new oxime-ether derivatives. Acta Pharm. Sci. 1998, 40, 131–135. [Google Scholar]
  20. Oh, J.M.; Rangarajan, T.M.; Chaudhary, R.; Singh, R.P.; Singh, M.; Singh, R.P.; Tondo, A.R.; Gambacorta, N.; Nicolotti, O.; Mathew, B. Novel Class of Chalcone Oxime Ethers as Potent Monoamine Oxidase-B and Acetylcholinesterase Inhibitors. Molecules 2020, 25, 2356. [Google Scholar] [CrossRef]
  21. Islam, M.; Rahman, M.; Hossain, G. Floristic composition and phytodiversity status of Sitakunda Ecopark, Chittagong, Bangladesh. Jahangirnagar Univ. J. Biol. Sci. 2016, 5, 29. [Google Scholar] [CrossRef]
  22. Rakib, A.; Ahmed, S.; Islam, M.A.; Uddin, M.M.N.; Paul, A.; Chy, M.N.U.; Emran, T.B.; Seidel, V. Pharmacological studies on the antinociceptive, anxiolytic and antidepressant activity of Tinospora crispa. Phytotherapy Res. 2020. [Google Scholar] [CrossRef] [PubMed]
  23. Tareq, A.M.; Farhad, S.; Neshar Uddin, A.B.M.; Hoque, M.; Nasrin, M.S.; Uddin, M.M.R.; Hasan, M.; Sultana, A.; Munira, M.S.; Lyzu, C.; et al. Chemical profiles, pharmacological properties, and in silico studies provide new insights on Cycas Pectinata. Heliyon 2020, 6, e04061. [Google Scholar] [CrossRef] [PubMed]
  24. Moawad, A.; Hetta, M.; Zjawiony, J.K.; Jacob, M.R.; Hifnawy, M.; Marais, J.P.J.; Ferreira, D. Phytochemical investigation of Cycas circinalis and Cycas revoluta leaflets: Moderately active antibacterial biflavonoids. Planta Med. 2010, 76, 796–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nair, J.J.; van Staden, J. Isolation and quantification of the toxic methylazoxymethanol glycoside macrozamin in selected South African cycad species. S. Afr. J. Bot. 2012, 82, 108–112. [Google Scholar] [CrossRef] [Green Version]
  26. Negm, W.; Ibrahim, A.R.; Aboelsauod, K.; Ragab, A.; Attia, G.I. GC-MS Analysis of Petroleum Ether Extract and Volatiles of Cycas revoluta Thunb Growing in Egypt. Inventi Rapid Planta Act. 2016, 2016, 1–5. [Google Scholar]
  27. Kumar, S.B.; Kumar, V.J. GC-MS Analysis of Bioactive Constituents from Cycas circinalis L. and Ionidium suffruticosum Ging. Int. J. Pharm. Sci. Rev. Res. 2014, 28, 197–201. [Google Scholar]
  28. Ben, I.O.; Woode, E.; Abotsi, W.K.M.; Boakye-Gyasi, E. Preliminary phytochemical screening and in vitro antioxidant prop-erties of Trichilia monadelpha (Thonn.) JJ De Wilde (Meliaceae). J. Med. Biomed. Sci. 2013, 2, 6–15. [Google Scholar]
  29. Bell, W.R. Evaluation of Thrombolytic Agents. Drugs 1997, 54, 11–17. [Google Scholar] [CrossRef]
  30. Schenone, S.; Bruno, O.; Ranise, A.; Bondavalli, F.; Filippelli, W.; Falcone, G.; Rinaldi, B. Treating Depression and Anxiety in Primary Care. Prim. Care Companion J. Clin. Psychiatry 2008, 10, 145–152. [Google Scholar] [CrossRef]
  31. Adnan, M.; Chy, M.N.; Kamal, A.T.M.M.; Chowdhury, K.A.; Rahman, M.A.; Reza, A.S.M.A.; Moniruzzaman, M.; Rony, S.R.; Nasrin, M.S.; Azad, M.O.; et al. Intervention in Neuropsychiatric Disorders by Suppressing Inflammatory and Oxidative Stress Signal and Exploration of In Silico Studies for Potential Lead Compounds from Holigarna caustica (Dennst.) Oken leaves. Biomolecules 2020, 10, 561. [Google Scholar] [CrossRef] [Green Version]
  32. Benneh, C.K.; Biney, R.P.; Adongo, D.W.; Mante, P.K.; Ampadu, F.A.; Tandoh, A.; Jato, J.; Woode, E. Anxiolytic and Antidepressant Effects of Maerua angolensis DC. Stem Bark Extract in Mice. Depress. Res. Treat. 2018, 2018, 1537371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Beheshti, F.; Khazaei, M.; Hosseini, M. Neuropharmacological effects of Nigella sativa. Avicenna J. Phytomed. 2016, 6, 104–116. [Google Scholar] [PubMed]
  34. Baldwin, D.S.; Polkinghorn, C. Evidence-based pharmacotherapy of Generalized Anxiety Disorder. Int. J. Neuropsychopharmacol. 2005, 8, 293–302. [Google Scholar] [CrossRef] [Green Version]
  35. Cryan, J.F.; Mombereau, C.; Vassout, A. The tail suspension test as a model for assessing antidepressant activity: Review of pharmacological and genetic studies in mice. Neurosci. Biobehav. Rev. 2005, 29, 571–625. [Google Scholar] [CrossRef] [PubMed]
  36. El Refaey, H. Fluoxetine. In xPharm: The Comprehensive Pharmacology Reference; Elsevier: Amsterdam, The Netherlands, 2007; pp. 1–9. [Google Scholar] [CrossRef]
  37. Brigitta, B. Pathophysiology of depression and mechanisms of treatment. Dialogues Clin. Neurosci. 2002, 4, 7–20. [Google Scholar]
  38. Abdulrasheed, M.; Ibrahim, I.H.; Mubarak, M.A.; Umar, F.A. Comparison of antimicrobial activity of seed oil of garlic and Moringa oleifera against some food-borne microorganisms. Bayero J. Pure Appl. Sci. 2015, 8, 196–201. [Google Scholar] [CrossRef]
  39. Meyer, B.N.; Ferrigni, N.R.; Putnam, J.E.; Jacobsen, L.B.; Nichols, D.E.J.; McLaughlin, J.L. Brine shrimp: A convenient general bioassay for active plant constituents. Planta Med. 1982, 45, 31–34. [Google Scholar] [CrossRef]
  40. Parvez, M.; Mosaddik, A. Evaluation of Brine Shrimp Cytotoxicity of Mango Peel and Flesh after Formalin Treatment. Int. J. Innov. Pharm. Sci. Res. 2016, 4, 900–908. [Google Scholar] [CrossRef]
  41. Soga, S.; Sharma, S.V.; Shiotsu, Y.; Shimizu, M.; Tahara, H.; Yamaguchi, K.; Ikuina, Y.; Murakata, C.; Tamaoki, T.; Kurebayashi, J.; et al. Stereospecific antitumor activity of radicicol oxime derivatives. Cancer Chemoth. Pharm. 2001, 48, 435–445. [Google Scholar] [CrossRef]
  42. Minkin, V.I.; Osipov, O.A.; Zhdanov, Y.A. Dipole Moments in Organic Chemistry; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar] [CrossRef]
  43. Baldin, A.M. Polarizability of nucleons. Nucl. Phys. 1960, 18, 310–317. [Google Scholar] [CrossRef]
  44. Gómez-Jeria, J.S. An empirical way to correct some drawbacks of mulliken population analysis. J. Chil. Chem. Soc. 2009, 54, 482–485. [Google Scholar] [CrossRef] [Green Version]
  45. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  46. Sheikhi, M.; Sheikh, D. Quantum chemical investigations on phenyl-7, 8-dihydro-[1,3]-dioxolo [4,5-g] quinolin-6 (5h)-one. Rev. Roum. Chim. 2014, 59, 761–767. [Google Scholar]
  47. Choi, C.H.; Kertesz, M. Bond length alternation and aromaticity in large annulenes. J. Chem. Phys. 1998, 108, 6681–6688. [Google Scholar] [CrossRef]
  48. Fleming, I. Frontier Orbitals and Organic Chemical Reactions; Wiley: Hoboken, NJ, USA, 1977. [Google Scholar]
  49. Koopmans, T. Ordering of wave functions and eigenenergies to the individual electrons of an atom. Physica 1933, 1, 104–113. [Google Scholar] [CrossRef]
  50. Phillips, J.C. Generalized Koopmans’ Theorem. Phys. Rev. 1961, 123, 420–424. [Google Scholar] [CrossRef]
  51. Flippin, L.A.; Gallagher, D.W.; Jalali-Araghi, K. A convenient method for the reduction of ozonides to alcohols with borane-dimethyl sulfide complex. J. Org. Chem. 1989, 54, 1430–1432. [Google Scholar] [CrossRef]
  52. Parr, R.G.; Szentpály, L.v.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  53. Chattaraj, P.K.; Giri, S. Stability, Reactivity, and Aromaticity of Compounds of a Multivalent Superatom. J. Phys. Chem. A 2007, 111, 11116–11121. [Google Scholar] [CrossRef] [Green Version]
  54. Padmanabhan, J.; Parthasarathi, R.; Subramanian, V.; Chattaraj, P.K. Electrophilicity-Based Charge Transfer Descriptor. J. Phys. Chem. A 2007, 111, 1358–1361. [Google Scholar] [CrossRef]
  55. Ayers, P.W.; Parr, R.G. Variational Principles for Describing Chemical Reactions:  The Fukui Function and Chemical Hardness Revisited. J. Am. Chem. Soc. 2000, 122, 2010–2018. [Google Scholar] [CrossRef]
  56. Obi-Egbedi, N.O.; Obot, I.B.; El-Khaiary, M.I.; Umoren, S.A.; Ebenso, E.E. Computational simulation and statistical analysis on the relationship between corrosion inhibition efficiency and molecular structure of some phenanthroline derivatives on mild steel surface. Int. J. Electrochem. Sci. 2011, 6, e5675. [Google Scholar]
  57. Parr, R.G.; Donnelly, R.A.; Levy, M.; Palke, W.E. Electronegativity: The density functional viewpoint. J. Chem. Phys. 1978, 68, 3801–3807. [Google Scholar] [CrossRef]
  58. Pulay, P.; Fogarasi, G.; Pang, F.; Boggs, J.E. Systematic ab initio gradient calculation of molecular geometries, force constants, and dipole moment derivatives. J. Am. Chem. Soc. 1979, 101, 2550–2560. [Google Scholar] [CrossRef]
  59. Rauhut, G.; Pulay, P. Transferable Scaling Factors for Density Functional Derived Vibrational Force Fields. J. Phys. Chem. 1995, 99, 3093–3100. [Google Scholar] [CrossRef]
  60. Rajesh, P.; Gunasekaran, S.; Gnanasambandan, T.; Seshadri, S. Experimental and theoretical study of ornidazole. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 153, 496–504. [Google Scholar] [CrossRef]
  61. Lampert, H.; Mikenda, W.; Karpfen, A. Molecular Geometries and Vibrational Spectra of Phenol, Benzaldehyde, and Salicylaldehyde:  Experimental versus Quantum Chemical Data. J. Phys. Chem. A 1997, 101, 2254–2263. [Google Scholar] [CrossRef]
  62. Puviarasan, N.; Arjunan, V.; Mohan, S. FT-IR and FT-Raman studies on 3-aminophthalhydrazide and N-aminophthalimide. Turk. J. Chem. 2002, 26, 323–334. [Google Scholar]
  63. Govindarajan, M.; Ganasan, K.; Periandy, S.; Karabacak, M. Experimental (FT-IR and FT-Raman), electronic structure and DFT studies on 1-methoxynaphthalene. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2011, 79, 646–653. [Google Scholar] [CrossRef]
  64. Colthup, N. Introduction to Infrared and Raman Spectroscopy; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  65. Roeges, N.P.G.; Baas, J.M.A. A Guide to the Complete Interpretation of Infrared Spectra of Organic Structures; Wiley: New York, NY, USA, 1994. [Google Scholar]
  66. Laszlo, V.; Endre, K.; Jozsef, B.; Boris, D.; Antal, S. Diethylamino and Pyrrolidino Lower Alkyl Esters of 3,5-dimethoxy-4-butoxy and Amyloxy Benzoic Acids. US3228961A, 11 January 1966. [Google Scholar]
  67. Silverstein, R.; Bassler, G.; Morrill, T. Spectrometric Identification of Organic Compounds; John Wiley and Sons: New York, NY, USA, 1981. [Google Scholar]
  68. Krishnakumar, V.; Manohar, S.; Nagalakshmi, R. Crystal growth and characterization of N-hydroxyphthalimide (C8H5NO3) crystal. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2008, 71, 110–115. [Google Scholar] [CrossRef]
  69. Smith, B.C. Infrared Spectral Interpretation: A Systematic Approach; CRC Press: Boca Raton, FL, USA, 1998. [Google Scholar] [CrossRef]
  70. Abbas, A.; Gökce, H.; Bahçeli, S. Spectroscopic (vibrational, NMR and UV–vis.) and quantum chemical investigations on 4-hexyloxy-3-methoxybenzaldehyde. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2016, 152, 596–607. [Google Scholar] [CrossRef] [PubMed]
  71. Baig, M.H.; Ahmad, K.; Roy, S.; Ashraf, J.M.; Adil, M.; Siddiqui, M.H.; Khan, S.; Kamal, M.A.; Provazník, I.; Choi, I. Computer Aided Drug Design: Success and Limitations. Curr. Pharm. Des. 2016, 22, 572–581. [Google Scholar] [CrossRef] [PubMed]
  72. Youdim, M.B.; Edmondson, D.; Tipton, K.F. The therapeutic potential of monoamine oxidase inhibitors. Nat. Rev. Neurosci. 2006, 7, 295–309. [Google Scholar] [CrossRef] [PubMed]
  73. Lin, J.; Sahakian, D.C.; de Morais, S.M.; Xu, J.J.; Polzer, R.J.; Winter, S.M. The role of absorption, distribution, metabolism, excretion and toxicity in drug discovery. Curr. Top. Med. Chem. 2003, 3, 1125–1154. [Google Scholar] [CrossRef] [PubMed]
  74. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 1997, 23, 3–25. [Google Scholar] [CrossRef]
  75. Veber, D.F.; Johnson, S.R.; Cheng, H.-Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular Properties That Influence the Oral Bioavailability of Drug Candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef]
  76. Zimmermann, M. Ethical guidelines for investigations of experimental pain in conscious animals. Pain 1983, 16, 109–110. [Google Scholar] [CrossRef]
  77. Mielke, H.; Strickland, J.; Jacobs, M.N.; Mehta, J.M. Biometrical evaluation of the performance of the revised OECD Test Guideline 402 for assessing acute dermal toxicity. Regul. Toxicol. Pharmacol. 2017, 89, 26–39. [Google Scholar] [CrossRef]
  78. Hossain, M.S.; Reza, A.; Rahaman, M.M.; Nasrin, M.S.; Rahat, M.R.U.; Islam, M.R.; Uddin, M.J.; Rahman, M.A. Evaluation of morning glory (Jacquemontia tamnifolia (L.) Griseb) leaves for antioxidant, antinociceptive, anticoagulant and cytotoxic activities. J. Basic Clin. Physiol. Pharmacol. 2018, 29, 291–299. [Google Scholar] [CrossRef]
  79. Evans, W.C. Trease and Evans Pharmacognosy, International Edition E-Book; Elsevier: Amsterdam, The Netherlands, 2009. [Google Scholar]
  80. Steru, L.; Chermat, R.; Thierry, B.; Simon, P. The tail suspension test: A new method for screening antidepressants in mice. Psychopharmacology 1985, 85, 367–370. [Google Scholar] [CrossRef]
  81. Porsolt, R.D.; Bertin, A.; Jalfre, M. Behavioral despair in mice: A primary screening test for antidepressants. Arch. Int. Pharmacodyn. Ther. 1977, 229, 327–336. [Google Scholar] [PubMed]
  82. Frisch, M.; Trucks, G.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.; et al. Gaussian 09; Revision d. 01; Gaussian Inc.: Wallingford CT, USA, 2009; p. 201. [Google Scholar]
  83. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6.0; Semichem Inc.: Shawnee, KS, USA, 2016; p. 16. [Google Scholar]
  84. Fekete, Z.A.; Hoffmann, E.A.; Körtvélyesi, T.; Penke, B. Harmonic vibrational frequency scaling factors for the new NDDO Hamiltonians: RM1 and PM6. Mol. Phys. 2007, 105, 2597–2605. [Google Scholar] [CrossRef]
  85. Ramazani, A.; Sheikhi, M.; Yahyaei, H. Molecular Structure, NMR, FMO, MEP and NBO Analysis of Ethyl-(Z)-3-phenyl-2-(5-phenyl-2H-1,2,3,4-tetraazol-2-yl)-2-propenoate Based on HF and DFT Calculations. Chem. Methodol. 2017, 1, 28–48. [Google Scholar] [CrossRef] [Green Version]
  86. Sastry, G.M.; Adzhigirey, M.; Day, T.; Annabhimoju, R.; Sherman, W. Protein and ligand preparation: Parameters, protocols, and influence on virtual screening enrichments. J. Comput. Aided Mol. Des. 2013, 27, 221–234. [Google Scholar] [CrossRef]
  87. Adnan, M.; Chy, M.N.; Kamal, A.T.M.M.; Chowdhury, M.R.; Islam, M.S.; Hossain, M.A.; Tareq, A.M.; Bhuiyan, M.I.; Uddin, M.N.; Tahamina, A.; et al. Unveiling Pharmacological Responses and Potential Targets Insights of Identified Bioactive Constituents of Cuscuta reflexa Roxb. Leaves through In Vivo and In Silico Approaches. Pharmaceuticals 2020, 13, 50. [Google Scholar] [CrossRef] [Green Version]
  88. Berman, H.M.; Battistuz, T.; Bhat, T.N.; Bluhm, W.F.; Bourne, P.E.; Burkhardt, K.; Feng, Z.; Gilliland, G.L.; Iype, L.; Jain, S.; et al. The Protein Data Bank. Acta Crystallogr. Sect. D Biol. Crystallogr. 2002, 58, 899–907. [Google Scholar] [CrossRef]
  89. Natarajan, A.; Sugumar, S.; Bitragunta, S.; Balasubramanyan, N. Molecular docking studies of (4Z,12Z)-cyclopentadeca-4, 12-dienone from Grewia hirsuta with some targets related to type 2 diabetes. BMC Complement. Altern. Med. 2015, 15, 73. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Antidepressant effect of methanol extract of C. pectinata leaves (MECP), fluoxetine and phenelzine in tail suspension test (A) and forced swimming test (B). The values are shown as mean ± standard error of the mean (SEM). * p < 0.001 statistically significant compared with the control by Dunnett’s test (n = 5).
Figure 1. Antidepressant effect of methanol extract of C. pectinata leaves (MECP), fluoxetine and phenelzine in tail suspension test (A) and forced swimming test (B). The values are shown as mean ± standard error of the mean (SEM). * p < 0.001 statistically significant compared with the control by Dunnett’s test (n = 5).
Pharmaceuticals 13 00232 g001
Figure 2. Percentage of mortality of brine shrimp lethality bioassay of methanol extract of C. pectinata leaves (MECP) and standard drug vincristine sulfate (VCS) at different concentrations.
Figure 2. Percentage of mortality of brine shrimp lethality bioassay of methanol extract of C. pectinata leaves (MECP) and standard drug vincristine sulfate (VCS) at different concentrations.
Pharmaceuticals 13 00232 g002
Figure 3. Optimized geometric structures of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] (A); cyclopentadecanone oxime (B) and trans-2-dodecen-1-ol trifluoroacetate (C).
Figure 3. Optimized geometric structures of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] (A); cyclopentadecanone oxime (B) and trans-2-dodecen-1-ol trifluoroacetate (C).
Pharmaceuticals 13 00232 g003
Figure 4. Calculated 3D surface mapped of electrostatic potential for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively in (a.u), the electron density isosurface being 0.0004 (a.u).
Figure 4. Calculated 3D surface mapped of electrostatic potential for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively in (a.u), the electron density isosurface being 0.0004 (a.u).
Pharmaceuticals 13 00232 g004
Figure 5. HOMO-LUMO plot 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively, by B3LYP/6-31G+ (d,p) level of theory.
Figure 5. HOMO-LUMO plot 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime] (A); cyclopentadecanone oxime (B); trans-2-dodecen-1-ol trifluoroacetate (C), respectively, by B3LYP/6-31G+ (d,p) level of theory.
Pharmaceuticals 13 00232 g005
Table 1. Quantitative compounds identified from methanol extract of C. pectinata by GC-MS analysis.
Table 1. Quantitative compounds identified from methanol extract of C. pectinata by GC-MS analysis.
Sl. No.RTCompound Namem/zAreaPA (%)Molecular FormulaMW (g/mol)Class
15.8811H-Indole-2,3-dione, 5-chloro-1-(trimethylsilyl)-, 3-[O-(trimethylsilyl)oxime]73.008515493.80C14H21ClN2O2Si2340.95Oxime
211.6403-Octyn-2-ol44.00259270.12C8H14O126.2Fatty alcohol
311.6402-Cyclohexen-1-one, 3-(3-hydroxybutyl)-2,4,4-trimethyl-44.00259270.12C13H22O2210.31Ketone
411.640Bioallethrin44.00259270.12C19H26O3302.4Pyrethroid
511.6403-Nonyn-2-ol44.00259270.12C9H16O140.22Secondary alcohol
612.5161-Octadecyne43.002095360.94C18H34250.5Hydrocarbon
712.516Z-2-Dodecenol43.002095360.94C12H24O184.32Fatty alcohol
812.516Phytol, acetate43.002095360.94C22H42O2338.6Diterpene
912.5155-Nonadecen-1-ol81.001226080.55C19H38O282.5Alcohols
1012.5152-Tridecyne81.001226080.55C13H24180.33Alkyne
1112.5169-Eicosyne43.001871410.84C20H38278.5Alkyne
1212.516Dodecanal43.001871410.84CH3(CH2)10CHO184.32Aldehyde
1312.516trans-2-Dodecen-1-ol, trifluoroacetate43.001871410.84C14H23F3O2280.33Ester
1413.450Tridecanoic acid, 12-methyl-, methyl ester74.004174741.86C15H30O2242.4Fatty acid
1513.450Eicosanoic acid, methyl ester74.004174741.86C21H42O2326.6FAME
1613.450Octadecanoic acid, 17-methyl-, methyl ester74.004174741.86C20H40O2312.5FAME
1715.17013-Tetradece-11-yn-1-ol67.00479050.21C14H24O208.34Alcohol
1815.1709,12-Octadecadienoic acid, methyl ester, (E,E)-67.00479050.21C19H34O2294.5FAME
1915.339Cyclopropaneoctanoic acid, 2-[[2-[(2-ethyl- cyclopropyl)methyl]cyclopropyl]methyl]-, methyl ester55.00703170.31C22H38O2334.5Fatty acid
2015.3393-Tetradecyn-1-ol55.00703170.31C14H26O210.36Alkyne
2115.3397-Hexadecenoic acid, methyl ester, (Z)-55.00703170.31C17H32O2268.4Fatty acid
2215.339Ethyl iso-allocholate55.00703170.31C26H44O5436.6Steroid
2315.337Isophytol, acetate71.001729500.77C22H42O2338.6Diterpene
2415.337E-2-Tetradecen-1-ol71.001729500.77C14H28O212.37Alkyne
2515.478Tetradecanoic acid, 12-methyl-, methyl ester, (S)-74.001073330.96C16H32O2316.5FAME
2615.478Heptacosanoic acid, methyl ester74.001073330.48C28H56O2424.7Fatty acid
2715.478Cyclopentanetridecanoic acid, methyl ester74.001073330.48C19H36O2296.5Fatty acid
2816.199Dodecanoic acid, 2-(acetyloxy)-1-[(acetyloxy)methyl]ethyl ester73.001110540.49C19H34O6338.5Ester
2916.199 and 5.881Phloroglucitol73.00111054 and 8515490.49 and 3.80C6H12O3132.16Alcohol
3016.604Octadecanal, 2-bromo-44.00113550.05C18H35BrO347.4Aldehyde
3116.604 and 12.515Undecanal81.00 and 44.00122608 and 113550.05 and 0.55C10H21CHO170.29Aldehyde
3217.623Octasiloxane, 1,1,3,3,5,5,7,7,9,9,11,11,13,13,15,15-hexadecamethyl-73.00774480.35C16H50O7Si8577.2Volatile organic compound
3317.623Dodecanoic acid, 2,3-bis(acetyloxy)propyl ester73.00774480.35C19H34O6358.5Ester
3419.440D-Mannitol, 1-O-(16-hydroxyhexadecyl)-73.00772150.34C22H46O7422.6Alcohol
3519.440Cyclopentadecanone, oxime73.00772150.34C15H29NO239.4Oxime
3619.440Docosanoic acid, docosyl ester73.00772150.34C44H88O2649.2Emollient
3720.009Hexadecanoic acid, 2-hydroxy-1-(hydroxymethyl)ethyl ester44.00496050.22C19H38O4330.5Fatty acid glycerol ester
3820.009Octadecanoic acid, 2-hydroxy-1,3-propanediyl ester44.00496050.22C39H76O5625.0Monoalkyl ester
3920.009Glycerol 1-palmitate44.00496050.22C19H38O4330.5Fatty acid
4020.360Chloroacetic acid, 4-pentadecyl ester44.00248940.11C17H33ClO2304.9Ester
4120.3602-Decen-1-ol, (E)-44.00248940.11C10H20O156.26Fatty acid
RT: Retention Time; m/z: m stands for mass and z stands for the charge number of ions, PA: Peak Area, MW: Molecular weight; FAME: fatty acid methyl ester.
Table 2. Optimized energies of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime]; cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate with dipole moment and polarizability.
Table 2. Optimized energies of 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)-oxime]; cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate with dipole moment and polarizability.
CompoundsEnergy (a.u)Dipole Moment (Debye)Polarizability (a.u)
5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime] −1845.680681.367261.403
Cyclopentadecanone oxime −718.770810.712162.046
trans-2-Dodecen-1-ol trifluoroacetate −997.048794.968159.682
Table 3. Global reactivity descriptors values in the gas phase.
Table 3. Global reactivity descriptors values in the gas phase.
Global Reactivity Descriptors5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethyl-silyl)oxime]Cyclopentadecanone Oximetrans-2-Dodecen-1-ol Trifluoroacetate
Ionisation potential (I) eV6.176716.597407.39524
Electron affinity (A) eV2.409020.273201.55567
Chemical hardness (η)1.883853.162102.91979
Softness (S)0.530830.316250.34249
Chemical potential (μ)−4.29287−3.43530−4.47546
Electronegativity (χ)4.292873.435304.47546
Electrophilicity index (ώ)9.214345.9006410.01485
Table 4. Calculated scaled infra-red (IR) frequencies (cm−1) for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate, respectively by DFT B3LYP/6-31+G (d,p) method (atom positions numbered as in the table).
Table 4. Calculated scaled infra-red (IR) frequencies (cm−1) for 5-chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime], cyclopentadecanone oxime and trans-2-dodecen-1-ol trifluoroacetate, respectively by DFT B3LYP/6-31+G (d,p) method (atom positions numbered as in the table).
MD5-Chloro-1-(trimethyl-silyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime]MDCyclopentadecanone OximeMDtrans-2-Dodecen-1-ol Trifluoroacetate
--υ(O15-H16)3884--
υ(C- H)3081~3099υ(C- H)3100~3200υ(C- H)3023~3069
υAsy(H-C-H)2927~3017υAsy(H-C-H)3051~3093υAsy(H-C-H)2920~3001
υSy(H-C-H)2916~2919υSy(H-C-H)3007~3019υSy(H-C-H)2805~2917
--δS(H-C-H)3023 ~ 3037--
υ(C=C)Aro1555~1579--υ(C=C)1680~1689
υ(C4=Cl12)695--υ(C-C)1283
--δS(H-C-H)1459~1499δS(H-C-H)1407~1506
--δW(H-C-H)1346~1373δW(H-C-H)1324~1373
--δT(H-C-H)1246~1287δT(H-C-H)1287~1312
δW(C-H)Aro916--δ(C-H)989
δ(C-H)Aro1137----
υ(C7=O15)1717----
υ(C18=N14)1593υ(C12=N14)1759--
υ(O29-N14)1035/979υ(O15-N14)891--
--δS(C12-N14-O15)524--
----υ(C-F)1123~1169
Calculated values were corrected by multiplying the frequency factor, f = 0.964. MD = Mode of Vibration, υ = Stretching, υSy = Symmetric Stretching, υAsy = Asymmetric Stretching, δ = Bending, δS = Scissoring, δW = Wagging, δR = Rocking, δT = Twisting, FS = Scaled frequency, Aro = Aromatic.
Table 5. Calculated 1H-NMR isotropic chemical shift (TMS and chloroform solution) by the DFT/B3LYP/6-31G+ (d,p) method (atom positions are numbered in the table).
Table 5. Calculated 1H-NMR isotropic chemical shift (TMS and chloroform solution) by the DFT/B3LYP/6-31G+ (d,p) method (atom positions are numbered in the table).
Compound (Chemical Shift-ppm)
Proton
No.
5-Chloro-1-(trimethylsilyl)-
1H-indole-2,3-dione
3-[O- (trimethylsilyl)oxime]
Proton
No.
Cyclopentadecanone OximeProton
No.
trans-2-Dodecen-1-ol, Trifluoroacetate
9-H6.36416-H4.64335-H5.567
10-H6.12918-H2.07436-H5.183
11-H5.94340-H1.80938-H4.266
22-H0.09717-H1.22137-H3.952
21-H0.07839-H1.12633-H1.459
27-H−0.27628-H1.11934-H1.319
25-H−0.28119-H1.06831-H0.699
39-H−0.36443-H0.97130-H0.645
34-H−0.38746-H0.73619-H0.611
40-H−0.41230-H0.70520-H0.607
36-H−0.42024-H0.67328-H0.600
24-H−0.62127-H0.63724-H0.596
28-H−0.63020-H0.58726-H0.594
42-H−0.74325-H0.57423-H0.593
37-H−0.75934-H0.57027-H0.592
23-H−0.84829-H0.52422-H0.582
26-H−0.85137-H0.52029-H0.579
35-H−0.90441-H0.50821-H0.579
20-H−0.91833-H0.50525-H0.579
41-H−0.94838-H0.50532-H0.503
38-H−0.95621-H0.47816-H0.245
35-H0.39518-H0.131
22-H0.35417-H0.124
44-H0.316
42-H0.298
23-H0.256
32-H0.252
45-H0.248
26-H0.181
Table 6. Calculated 13C-NMR isotropic chemical shift (TMS and chloroform solution) by the DFT/B3LYP/6-31G+ (d,p) method (atom positions numbered in the table).
Table 6. Calculated 13C-NMR isotropic chemical shift (TMS and chloroform solution) by the DFT/B3LYP/6-31G+ (d,p) method (atom positions numbered in the table).
Compound (Chemical Shift-ppm)
Carbon No.5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime]Carbon No.Cyclopentadecanone OximeCarbon No.trans-2-Dodecen-1-ol Trifluoroacetate
7-C142.31112-C146.4552-C151.741
8-C133.86911-C21.7606-C129.323
1-C128.58910-C16.8761-C121.591
4-C120.9758-C15.6035-C105.659
5-C113.28436-C15.4994-C63.818
2-C108.0361-C15.3877-C24.997
3-C103.79313-C15.32013-C23.385
6-C96.6373-C14.51112-C22.139
17-C−12.5274-C14.09411-C22.046
33-C−13.72431-C13.77110-C21.930
19-C−13.8955-C13.4449-C21.861
18-C−13.9146-C13.0248-C21.102
32-C−13.9187-C12.40714-C14.600
31-C−14.1102-C11.63515-C3.962
9-C11.045
Table 7. Docking scores of the identified compounds from methanol extract of C. pectinata leaves.
Table 7. Docking scores of the identified compounds from methanol extract of C. pectinata leaves.
CompoundsDocking Score (kcal/mol)
2Z5X5I6X1ERR1M17
5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethyl- silyl)oxime]
Cyclopentadecanone oxime−4.333−6.537−7.685−4.59
trans-2-Dodecen-1-ol trifluoroacetate−3.155−2.387−1.857−2.674
Standard drugs (Phenelzine/Fluoxetine/Vincristine sulfate)−5.324−9.07−3.896−3.85
Table 8. ADME/T properties of the selected compounds in MECP by SwissADME.
Table 8. ADME/T properties of the selected compounds in MECP by SwissADME.
CompoundsLipinski RulesLipinski ViolationsVeber Rules
MWHBAHBDLog PnRBTPSA
5-Chloro-1-(trimethylsilyl)-1H-indole-2,3-dione 3-[O-(trimethylsilyl)oxime]340.95302.380341.90
Cyclopentadecanone oxime239.40213.550032.59
trans-2-Dodecen-1-ol trifluoroacetate280.33503.9601226.30
MW, Molecular weight (<500 g/mol); HBA, Hydrogen bond acceptor (<10); HBD, Hydrogen bond donor (<5); Log P, Lipophilicity (≤5); nRB: number of rotatable bond (≤10); TPSA: topological polar surface area (≤140 Ų).
Table 9. Toxicological properties of the selected compounds in MECP.
Table 9. Toxicological properties of the selected compounds in MECP.
ParametersCompounds
5-Chloro-1-(trimethylsilyl)-1H-Indole-
2,3-dione 3-[O-(trimethylsilyl)oxime]
Cyclopentadecanone Oximetrans-2-Dodecen-1-ol Trifluoroacetate
Ames toxicity NATATNAT
Carcinogens NCNCC
Acute oral toxicity IIIIIIIII
Rat acute toxicity 2.68492.12032.6831
NAT, Non-Ames toxic; AT, Ames toxic; NC, Non-carcinogenic; C, carcinogenic; NR, Non-required. Category-I means (LD50 ≤ 50 mg/kg) and Category-III (500 mg/kg > LD50 < 5000 mg/kg).

Share and Cite

MDPI and ACS Style

Rahman, J.; Tareq, A.M.; Hossain, M.M.; Sakib, S.A.; Islam, M.N.; Ali, M.H.; Uddin, A.B.M.N.; Hoque, M.; Nasrin, M.S.; Emran, T.B.; et al. Biological Evaluation, DFT Calculations and Molecular Docking Studies on the Antidepressant and Cytotoxicity Activities of Cycas pectinata Buch.-Ham. Compounds. Pharmaceuticals 2020, 13, 232. https://0-doi-org.brum.beds.ac.uk/10.3390/ph13090232

AMA Style

Rahman J, Tareq AM, Hossain MM, Sakib SA, Islam MN, Ali MH, Uddin ABMN, Hoque M, Nasrin MS, Emran TB, et al. Biological Evaluation, DFT Calculations and Molecular Docking Studies on the Antidepressant and Cytotoxicity Activities of Cycas pectinata Buch.-Ham. Compounds. Pharmaceuticals. 2020; 13(9):232. https://0-doi-org.brum.beds.ac.uk/10.3390/ph13090232

Chicago/Turabian Style

Rahman, Jinnat, Abu Montakim Tareq, Md. Mohotasin Hossain, Shahenur Alam Sakib, Mohammad Nazmul Islam, Md. Hazrat Ali, A. B. M. Neshar Uddin, Muminul Hoque, Mst. Samima Nasrin, Talha Bin Emran, and et al. 2020. "Biological Evaluation, DFT Calculations and Molecular Docking Studies on the Antidepressant and Cytotoxicity Activities of Cycas pectinata Buch.-Ham. Compounds" Pharmaceuticals 13, no. 9: 232. https://0-doi-org.brum.beds.ac.uk/10.3390/ph13090232

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop