Next Article in Journal
Microalgal and Nitrogen-Fixing Bacterial Consortia: From Interaction to Biotechnological Potential
Next Article in Special Issue
Variation in the Chemical Composition of Endemic Specimens of Hedychium coronarium J. Koenig from the Amazon and In Silico Investigation of the ADME/Tox Properties of the Major Compounds
Previous Article in Journal
Cloning and Functional Analysis of the VfRR17 Gene from tung tree (Vernicia fordii)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chamaecyparis lawsoniana Leaf Essential Oil as a Potential Anticancer Agent: Experimental and Computational Studies

1
Department of Pharmacognosy, Faculty of Pharmacy, Zagazig University, Zagazig 44519, Egypt
2
Department of Pharmacognosy, Collage of Pharmacy, King Saud University, Ryiadh 11451, Saudi Arabia
3
Department of Medicinal Chemistry, Faculty of Pharmacy, Zagazig University, Zagazig 44519, Egypt
4
Department of Chemistry, School of Computer, Mathematical and Natural Sciences, Morgan State University, Baltimore, MD 21251, USA
*
Authors to whom correspondence should be addressed.
Plants 2023, 12(13), 2475; https://doi.org/10.3390/plants12132475
Submission received: 14 May 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Advances in Essential Oils from Medicinal Plants)

Abstract

:
Cancer remains one of the leading causes of death worldwide, affected by several factors including oxidative stress; and although conventional synthetic medicines have been used to treat cancer, they often result in various side effects. Consequently, there is a growing need for newer, safer and more effective alternatives, such as natural plant products. Essential oils (EOs) are one such alternative, offering a wide range of bioactivities, including antibacterial, antiviral, antioxidant, and anticancer properties. Accordingly, the objective of the present study was to investigate the chemical composition, as well as the antioxidant and anticancer properties of the leaf essential oil of Chamaecyparis lawsoniana (CLLEO) belonging to the Cupressaceae family. Totally, 59 constituents were identified by gas chromatography-mass spectrometry (GC-MS) analysis. cis-Abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene were the major components. The in vitro cytotoxicity study against human breast (MCF-7), colon (HCT-116), lung (A-549), hepatocellular (HepG-2) carcinoma cells using MTT assay indicated a promising cytotoxic activity against all the tested cancer cells, particularly HepG-2, with significant selectivity indices. CLLEO exhibited weak antioxidant activity according to the DPPH, ABTS and FRAP assays. In silico docking of these constituents against the epidermal growth factor receptor (EGFR), the myeloid cell leukemia-1 (Mcl-1) and caspase-8 using Molecular Operating Environment (MOE) software demonstrated good binding affinities of the components with the active site of these targets. These findings suggested using CLLEO, or its individual components, as a potentially viable therapeutic option for managing cancerous conditions.

1. Introduction

Last decades, medicinal and aromatic plants played fundamental roles in the area of therapeutics worldwide [1]. Among several secondary metabolites produced from such plants, essential oils (EOs) gained a great interest owing to their complex chemical framework besides their efficacy in several traditional healing approaches [2]. The biological activities of EOs as well as their possible mechanisms of action and pharmacological targets have been reported by several preclinical studies demonstrating their antimicrobial, anticancer, anti-inflammatory, antioxidant, antidiabetic, and spasmolytic activities in several in vitro and in vivo models [2,3,4,5,6].
Recently, species of the family Cupressaceae are considered as a valuable source for various bioactive natural products including EOs [7,8,9,10,11,12]. Various studies have explored the cytotoxic properties of EOs from different species in this family, revealing their potential as effective drugs for treating cancer [12,13,14,15,16,17,18,19]. Chamaecyparis lawsoniana (A. Murray) Parl. (Syn. Cupressus lawsoniana A. Murray) is a member of the family Cupressaceae which is also named ginger-pine, Lawson’s cypress or Port-Orford-Cedar. It is a huge tree with 43–55 m height that is native to North America with a limited distribution to the coastal forests of Northern California and Southwestern Oregon in the USA [20,21]. Regarding the chemical composition of its EO, few reports from Belgium [22], Iran [23], Spain [21], Greece [24], and Belgrade [25] have revealed the chemical constitution of C. lawsoniana EOs. In addition, the literature survey indicated that only one report has concerned with its biological activities revealing its antibacterial and antifungal effects [25,26,27]. Nevertheless, there are no reports about the other possible biological activities including antioxidant capacity and cytotoxic potential of CLLEO.
Cancer is one of the most serious diseases around the world and is exponentially increasing with recently recognized lifestyles. Globally, it was recorded as the primary cause of death since in 2020 about 19.3 million new cancer cases were reported besides 9.9 million cancer fatalities [28]. One of the principal issues with cancer cells is their capability to evade apoptosis via unidentified mutations, leading to accumulation of cells, that in consequence migrate to several body parts [29]. Therefore, drugs which affect cancer cells without influencing normal cells, through restoring the apoptosis mechanisms in the cancer ones and that able to avoid multidrug resistance are considered as effective anticancer drugs [30]. Moreover, the toxicity of the existing chemotherapeutic agents for substantial limitations in their use. Thus, the innovation of safe drug candidates is considered as an important challenge [31]. In this context, natural products have formed an essential role as anticancer drugs and chemotherapeutic agents for about half a century, e.g., vinblastine, vincristine, paclitaxel, camptothecin and doxorubicin [32]. Hence, the discovery of new natural products with anticancer features has unique concern for medical care purposes.
Oxidative stress emerges from the imbalance between the generated free radicals particularly reactive oxygen species (ROS) and the endogenous antioxidant defense systems. High levels of ROS and peroxides have been revealed to be associated with the pathogenesis of various disorders such as cancer, aging, diabetes, neurodegenerative and cardiovascular diseases [33,34].
With the growing focus on the potential anticancer effects of essential oils (EOs) derived from various plants and herbs, this study aimed to assess the chemical composition of essential oil extracted from C. lawsoniana fresh leaves, cultivated in Egypt, using GC-MS analysis. Moreover, it validated its antioxidant capacity and tested its effectiveness on different types of cancer cell lines including human breast (MCF-7), colon (HCT-116), lung (A-549) and hepatocellular (HepG-2) carcinoma cells. Furthermore, investigating the conceivable mechanisms that contribute to its cytotoxic activity was accomplished via in silico molecular docking of the CLLEO main components against different targets included in cancer progression.

2. Results

2.1. Chemical Composition of CLLEO

The hydrodistillation of C. lawsoniana fresh leaves resulted in the isolation of a transparent, yellow coloured EO with an average yield of 0.75 ± 0.05% v/w (from three independent extractions). The GC/MS analysis of CLLEO led to the identification of 59 components, comprising 92.77% of the total composition. The corresponding chemical names of these components and their area percentages are represented in Table 1. Figure 1 displays the GC/MS chromatogram of CLLEO, and the chemical structure of its main constituents are revealed in Figure 2.

2.2. Antioxidant Activity

The antioxidant scavenging capacity and reducing antioxidant power of CLLEO were evaluated using DPPH, ABTS and FRAP assays and the results are illustrated in Table 2. In comparison with ascorbic acid, the CLLEO revealed lower radical scavenging potentials against DPPH and ABTS radicals and lower ferric reducing ability, demonstrating relatively weak antioxidant activity.

2.3. In Vitro Cytotoxic Potential and Selectivity of CLLEO

In comparison with cisplatin, the cytotoxic effect of CLLEO versus MCF-7, HCT-116, A-549, HepG-2 carcinoma cell lines as well as MRC-5 normal cell line was explored in a concentration dependent manner using MTT assay (Table 3, Figure 3). CLLEO exhibited a pronounced cytotoxic action towards all the tested cancer cell lines. The HepG-2 carcinoma cell line was more sensitive to CLLEO (IC50 = 15.34 µg/mL) compared with MCF-7, HCT-116 and A-549 carcinoma cell lines (IC50 = 23.74, 28.27 and 25.79 µg/mL, respectively). Moreover, CLLEO revealed a selectivity against the four cell lines (SI > 3) in relative to MRC-5 as shown in Table 3.

2.4. In Silico Molecular Docking Study

To gain insights into binding affinity of CLLEO with epidermal growth factor receptor (EGFR) tyrosine kinase (PDB: 1M17) [45], induced myeloid leukemia cell differentiation (Mcl-1) (PDB: 2NLA) [46] and caspase-8 (PDB: 1F9E) [47]; molecular docking was performed.

2.4.1. Docking with Epidermal Growth Factor Receptor (EGFR)

Docking studies of cis-abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene on epidermal growth factor receptor (EGFR) tyrosine kinase (PDB: 1M17) revealed that the components reached the binding site of the enzyme. Docked components exhibited a good binding affinity as the docking energy score ranged from −5.7955 to −4.9882 Kcal/mol. cis-Abienol and trans-ferruginol showed H-bond interaction with the acidic Asp831 residue. trans-Ferruginol exhibited further stabilization by formation of arene-H interaction with Val702 residue. α-Cadinol showed H-bond interaction with Met769 (Table 4).

2.4.2. Docking with Induced Myeloid Leukemia Cell Differentiation (Mcl-1)

Docking studies of cis-abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene on Mcl-1 (PDB: 2NLA) indicated that the components reached the binding site of the enzyme. Docking energy scores of the docked components ranged from −5.3337 to −4.1318 Kcal/mol. Both cis-abienol and trans-ferruginol exhibited H-bond interaction with acidic Asp218 residue. α-Cadinol showed H-bond interaction with Gly192 residue (Table 5).

2.4.3. Docking with Caspase-8

Docking studies of cis-abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene on caspase-8 (PDB: 1F9E) exhibited good binding affinity within the binding site of the enzyme as the docking energy scores ranged from −5.4981 to −4.2315 Kcal/mol. cis-Abienol showed two H-bond interactions with Asp239 and Tyr244. trans-Ferruginol and α-cadinol formed H-bond interactions with Cys285 and Asp239, respectively (Table 6).

3. Discussion

The essential oil obtained from the fresh leaves of C. lawsoniana cultivated in Egypt was subjected, for the first time, to GC/MS analysis to profile its chemical composition. A total of 59 compounds representing 92.77% of the oil were identified. The major identified compound was the bicyclic diterpene alcohol cis-abienol (23.43%), followed by the tricyclic diterpene alcohol trans-ferruginol (14.31%). Both major compounds belong to the class of oxygenated diterpenes, represented by an overall contribution of 37.74%. Three other components were also detected in resonibally high amounts including the bicyclic sesquiterpene alcohol α-cadinol (8.84%), the bicyclic sesquiterpene hydrocarbon δ-muurolene (8.57%), and the bicyclic monoterpene hydrocarbon α-pinene (6.20%). CLLEO has also been characterized by significant amounts of γ-amorphene (4.10%), δ-cadinene (3.71%), 6-epi-β-cubebene (3.16%), sandaracopimara-8(14),15-diene (2.05%), abietatriene (1.89%), epi-α-cadinol (1.74%), pimara-7,15-dien-3-one (1.65%), 1,10-di-epi-cubenol (1.52%), germacrene D (1.32%), trans-totarol (1.27%) and terpinen-4-ol (1.20%), while the rest of the identified compounds were found in amounts less than 1%. To our knowledge, there are no studies revealed the chemical composition of the essential oil extracted from C. lawsoniana leaves cultivated in Egypt. However, in 1986, Karawya et al. [48] reported that the young twigs essential oil of cultivated specimens from Egypt consisted of limonene (15.8%), gscarene (14.1%), α-terpinene (11.2%), and sabinene (10.1%). Moreover, few reports about the chemical composition of C. lawsoniana essential oils that were obtained from other sources have been reported. In this regard, De Pooter et al. [22] carried out a study to investigate the chemical composition of C. lawsoniana essential oil obtained from leaves cultivated in Belgium. They found that limonene was the main constituent (45–60%) of this essential oil. In a similar work, terpinen-4-ol (22.0%), sabinene (21.0%), camphor (7.8%), citronellol (7.3%), γ-terpinene (7.0%) were identified as dominent components of the aerial parts essential oil extracted from the Iranian C. lawsoniana [23]. Additionally, Palá-Paúl et al. [21] reported limonene as the only major component (77.7%) of the essential oils isolated from the young stems and leaves of C. lawsoniana cultivated in Spain. Giatropoulos et al. [24] reported that Greek C. lawsoniana leaf essential oil composed mainly of 18.5% limonene, 17.1% beyerenne, 15.9% oplopanonyl acetate, and 9.7% methyl myrtenate. More recently, Nikolić et al. analyzed the essential oil obtained from the fresh leaves of C. lawsoniana grown in Belgrade [25], demonstrating that the principal constituents of this essential oil were limonene (16.7%), oplopanonyl acetate (14.5%), beyerene (10.1%), and 13-epi-dolabradiene (6.7%). These previous studies demonstrated that the investigated CLLEO had a significant difference in chemical composition and content percentages than that obtained from the same species in different countries. This fluctuation is mainly due to differences in location, climate, harvest period, age of the plant age, distillation method, and type of distillation apparatus used, plant parts used for oil isolation, the environmental conditions and method of analysis [49].
Since a single assay is limited in providing insights into the antioxidant capability of the oil, the CLLEO was subjected to scrutiny using three distinct assays including DPPH, ABTS, and FRAP assays to evaluate its prospective antioxidant activity. The essential oil exhibited weak radical scavenging activity towards DPPH and ABTS radicals as well as low ferric reducing power compared to standard ascorbic acid. These results align with prior research on the essential oil extracted from Chamaecyparis formosensis wood, which had a moderate to weak capacity for scavenging DPPH radicals [50]. Moreover, the present findings of the DPPH assay are corroborated by the insignificant anti-DPPH potential exhibited by the essential oil derived from the leaves and fruits of Chamaecyparis obtuse [51,52]. Conversely, the essential oils extracted from C. obtusa fruits demonstrated a remarkable ability to combat ABTS radicals and a moderate ability to reduce ferric ions [51]. Furthermore, it has been noted that C. lawsoniana bark extract demonstrated extraordinary antioxidative properties [53]. The fluctuation in the outcomes could be attributed to the intricate chemical nature of essential oils, which has the potential to yield diverse findings. Consequently, the decreased efficiency of CLLEO could be linked to the lack of phenolic components that contribute to the antioxidative capability [54].
On the other hand, the investigated CLLEO demonstrated a remarked cytotoxic activity against the cancer cell lines under examination. It exhibited a selective potent growth inhibitory activity to HepG-2 cells followed by MCF-7, then A-549 and lastly HCT-116. In addition, it showed a CC50 value of 95.17 µg/mL on the normal MRC-5 cells, revealing its selectivity to the carcinoma cell lines, especially towards HepG-2 cell (SI = 6.20). Regarding to the guidelines of the American National Cancer Institute (NCI) that considers IC50 values ≤ 30.0 μg/mL as considerably significant for active crude extract or essential oil [55,56], CLLEO could be deemed as a promising anticancer drug candidate. This is considered as the first report about the screening of CLLEO on various cancer cell lines. Regarding to genus Chamaecyparis, the review of literature showed that several studies have been conducted to uncover the cytotoxic capabilities of various Chamaecyparis species on either the same cancer cells as those used in the current study or on differing cell lines. Remarkably, the findings from these studies were in accordance with the present results. In this sense, C. obtusa leaf methanolic extract demonstrated an antiproliferative activity against HCT116 cells at a concentration of 1.25 μg/mL [13]. In addition, the methanolic extract of branches and leaves of C. obtusa var. breviramea f. crippsii exhibited notable cytotoxic effects against A549, BGC-823, Du145 and MDA-MB-231 with IC50 values of 0.94, 1.07, 0.95 and 0.96 μg/mL, respectively. Moreover, various chemical components isolated from the latter species have exhibited cytotoxic effects towards various cell lines. In particular, quercetin demonstrated activity against BGC-823, Hela, and A549 cell lines [14], while 13-epi-toruolsol showed efficacy against BGC-823 and Hela, and 3-epitriptobenzene B displayed activity against BGC-823, Hela, and A549 cancer cell lines [57].
The possible mechanisms involved in the cell death as well as the suggested molecular targets included in the noticed anticancer potential was investigated by accomplishing molecular docking studies for the CLLEO main components (cis-abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene) using Molecular Operating Environment (MOE) software to explore their binding affinity with the targeted active sites of EGFR, Mcl-1, and caspase-8 proteins and to emphasize its cytotoxic effects on cancer cells.
The epidermal growth factor receptor (EGFR), also named as ErbB1/HER1, is a transmembrane tyrosine kinase receptor protein manifested on the epithelial, mesenchymal, and neurogenic tissues. Upregulation of EGFR is commonly observed in cancer pathogenesis, including metastatic colorectal cancer, non-small-cell lung cancer, head and neck cancer, glioblastoma, breast, and pancreatic cancers. It has a key role in signaling pathways that trigger cell proliferation and apoptosis inhibition. EGFR targeting by anticancer agents suppresses the signal transduction pathways necessary to regulate the growth and proliferation of cancer cells in addition to resistance to cell death [58,59]. The molecular docking of major components of CLLEO on EGFR revealed that the components demonstrated a good binding revealed by the docking scores ranging from −5.7955 to −4.9882 Kcal/mol. Moreover, H-bond interaction with the acidic Asp831 residue was displayed by cis-abienol and trans-ferruginol. Additional stabilization was shown by trans-ferruginol through the formation of arene-H interaction with Val702 residue. α-Cadinol, as well, revealed H-bond interaction with Met769.
The programmed cell death, which is known as apoptosis, is a regulatory mechanism that has a key role in the management of cell proliferation in normal physiological processes as well as the pathological conditions. It enhances the removal of unnecessary damaged cells to keep the healthy balance between cell survival and cell death which is missed in case of cancer since the cells be removed are unable to receive the death signals due to a problem in any step of the apoptosis process [60,61]. Therefore, apoptosis could act as a main target for anticancer research. Apoptosis is induced through two main pathways; the intrinsic mitochondrial apoptotic pathway which is initiated by intracellular signals such as DNA damage, oxidative stress, ischemia, radiotherapy, or chemotherapy and triggers the activation of B-cell lymphoma-2 (Bcl-2) family proteins, and the extrinsic death receptor apoptotic pathway which is activated in response to extracellular signals emitted by other cells [60,61,62].
A potent anti-apoptotic protein from Bcl-2 family is the myeloid cell leukemia-1 (Mcl-1) which has a vital role in cellular apoptosis regulation. Mcl-1 overexpression was reported in various types of human cancers and was responsible for the resistance against several anticancer drugs. Additionally, its downregulation was proved to promote the induction of apoptosis and improve the sensitivity toward anticancer drugs [62,63,64]. On the other hand, caspase-8 is one of cysteine proteases involved in the induction of extrinsic apoptosis where it acts as initiator for the apoptosis signal propagation via direct cleavage of downstream effector caspases such as caspase-3. Furthermore, it has a vital role in several cellular processes included in cancer development and progression [65,66]. Consequently, Mcl-1 and Caspase-8 constitute appealing molecular targets for the development of new anticancer drugs. CLLEO main components were docked on Mcl-1 and showed docking energy scores ranged from −5.3337 to −4.1318 Kcal/mol. Both cis-abienol and trans-ferruginol revealed H-bond interaction with acidic Asp218 residue while α-cadinol interacted with Gly192 residue via a H-bond. Moreover, docking of such components on caspase-8 exhibited a good binding affinity within the binding site of the enzyme where the docking energy scores ranged from −5.4981 to −4.2315 Kcal/mol. Two H-bond interactions with Asp239 and Tyr244 were displayed by cis-abienol. Also, trans-ferruginol and α-cadinol created H-bond interactions with Cys285 and Asp239, respectively. Overall, the docked components of CLLEO showed promising binding affinities with the targeted proteins; EGFR, Mcl-1 and caspase-8. The docking results were correlated with the cytotoxic activity of CLLEO, indicating a potential anticancer effect. Therefore, using of each component in this EO, particularly the major ones or the whole EO as anticancer treatment is highly recommended.
In brief, CLLEO could be considered as a promising phytotherapeutic drug candidate in treatment of cancer diseases that acts through the inhibition of cell proliferation and the induction of apoptosis by targeting active sites of the key proteins, EGFR, Mcl-1, and caspase-8. Clinical trials in humans are crucial to guarantee the role of C. lawsoniana leaf EO intake as a cancer treatment.

4. Material and Methods

4.1. Chemicals

The chemicals utilized in this study including 2,2-Diphenyl-1-1-picrylhydrazyl (DPPH), 2,2′-azinobis (3-ethylbenzo-thiazoline-6-sulfonic acid) (ABTS), ascorbic acid, 3-(4,5-dimethythiazol-2-yl)-2,5-diphenyl tetrazolium bromide (MTT), dimethyl sulfoxide (DMSO), and cisplatin, were procured from Sigma (St. Louis, MO, USA).

4.2. Plant Material

Chamaecyparis lawsoniana (A. Murray) Parl. fresh leaves (Figure 4a,b) were gathered in February 2023 from El-Orman Botanical Garden, Giza, Egypt. Taxonomical verification of the plant species was carried out by Eng. Therese Labib, Plant Taxonomy Consultant at the Ministry of Agriculture and Ex-director of El-Orman Botanical Garden, Giza, Egypt. A voucher specimen, with the code ZU-Ph-Cog-0301, was preserved at the Herbarium of the Pharmacognosy Department, Faculty of Pharmacy, Zagazig University.

4.3. Essential Oil Extraction

The fresh leaves of C. lawsoniana (200 g) were exposed to hydrodistillation by the Clevenger-type apparatus under atmospheric pressure at about 100 °C for 5 h. The obtained essential oil was dried over anhydrous sodium sulphate and kept in brown-colored vials at 4 °C until further chemical and biological analyses. Three extractions were performed, and the average yield was determined.

4.4. Gas Chromatography–Mass Spectrometry (GC-MS) Analysis

CLLEO was analysed using Shimadzu GCMS-QP2010 (Kyoto, Japan) fitted with Rtx-5MS fused bonded column (30 m × 0.25 mm i.d. × 0.25 µm film thickness; Restek, PA, USA) equipped with a split–splitless injector. The column temperature was initially kept at 45 °C for 2 min (isothermal) and raised to 300 °C at a rate of 5 °C/min then held at 300 °C for 5 min (isothermal). The injection temperature was 250 °C. Helium carrier gas was used at a flow rate of 1.41 mL/min. The mass spectrometer was scanned over the 35 to 500 m/z with an ionizing voltage of 70 eV; a filament emission current of 60 mA and an ion source temperature 200 °C. A 1 µL of the diluted sample (1% v/v) was injected with split mode (split ratio 1:15). The identification of CLLEO components was based on the comparison of their retention indices (RI) and mass spectra (MS) to that reported in Adams library [35], NIST 11 Mass Spectral Library (NIST11/2011/EPA/NIH), Wiley library database 10th edition and the literature data [36,37,38,39,40,41,42,43,44]. The retention indices were assigned in relation to those of a homologous set of standard n-alkanes (C8–C28) injected under the same conditions. The identified compounds and their percentages are listed in Table 1.

4.5. Antioxidant Activity Assays

The antioxidant activity of CLLEO was assessed by three methods with different mechanisms; the 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging assay according to Elkomy et al. [67], 2,2′-azinobis(3-ethylbenzo-thiazoline-6-sulfonic acid) radical cation (ABTS●+) scavenging assay according to Ling et al. [68] and ferric reducing antioxidant power (FRAP) assay according to Elaasser et al. [69]. The experiments were performed in triplicate and average values were considered. Ascorbic acid was used as a reference compound.

4.6. In Vitro Anticancer Assay

4.6.1. Cell Line Propagation

The cytotoxic effects of CLLEO were evaluated in vitro against human breast (MCF-7), colon (HCT-116), lung (A-549), hepatocellular (HepG-2) carcinoma cells and a normal human Lung fibroblast cell line (MRC-5) obtained from the American Type Culture Collection (ATCC, Rockville, MD, USA). RPMI-1640 medium supplemented with 50 µg/mL gentamycin and 10% inactivated fetal calf serum was employed for cells propagation. The cells were kept in a humidified atmosphere with 5% CO2 at 37 °C and were sub-cultured two to three times a week.

4.6.2. Cytotoxicity Evaluation Using Viability Assay

MCF-7, HCT-116, A-549, HepG-2 and MRC-5 cell lines were plated into Corning® 96-well tissue culture plates with a cell population density of 5 × 104 cell/well, then incubated for 24 h. A fresh medium containing different concentrations of CLLEO (from 0.25 to 500 µg/mL) was then added. For each 96 well plate, a control was conducted using six vehicle controls containing 0.5% DMSO. The plates were maintained in a humidified incubator with 5% CO2 at 37 °C for 24 h. Control cells were incubated without the tested sample. Positive control containing cisplatin was also tested as reference for comparison. After the incubation period, the counts of viable cells were valued by the microplate 3-(4,5-dimethythiazole-2yl)-2,5-diphenyl-tetrazolium bromide (MTT) assay [70]. In brief, the mixture of media and sample was removed and replaced with 100 µL of fresh culture RPMI 1640 medium without phenol red then 10 µL of the 12 mM MTT stock solution (5 mg of MTT in 1 mL of PBS) was added to each well including the untreated controls. After 4 h of incubation, removal of an 85 µL aliquot of the media from the wells was done, and followed by the addition of a 50 µL of DMSO to each well then thoroughly mixed with the pipette and incubated at 37 °C for 10 min. The optical density was then measured at 590 nm with the microplate reader (SunRise, TECAN, Inc., Morrisville, NC, USA) to define the number of viable cells. All experiments were held in triplicate and the percentage of viability was calculated as:
The percentage of viability = (ODt/ODc) × 100%
where ODt is the mean optical density of wells treated with the tested sample and ODc is the mean optical density of untreated cells. The relation between surviving cells and drug concentration was plotted to obtain the survival curve of each cell line after treatment with the tested sample. The 50% inhibitory concentration (IC50, the concentration required to trigger toxic effects in 50% of intact cells), was determined from graphic plots of the dose response curve for each concentration using Graphpad Prism software version 9.4.1 (San Diego, CA, USA). For the normal cell line (MRC-5) it was expressed as CC50 (50% cytotoxic concentration).

4.6.3. Statistical Analysis

The results were expressed as the mean values ± standard deviation (SD) of three investigates. Statistica software version 8.0 (StatSoft Inc., Tulsa, OK, USA) was utilized for the statistical analysis. The Tukey’s test was used with a significance level of 5%.

4.6.4. Calculation of Selectivity Index (SI)

To evaluate the efficacy and degree of selectivity of CLLEO, SI value was calculated based on the following formula: SI = [CC50 of a normal cell line (MRC-5)/IC50 of cancer cell line]. High SI value (>3) suggests a drug with high selectivity against cancer cell lines [71,72].

4.7. In Silico Molecular Docking Study

Molecular docking studies of the following components: cis-abienol, trans-ferruginol, α-cadinol, δ-muurolene and α-pinene were performed to evaluate their binding affinity with the targeted active sites of EGFR, Mcl-1, and caspase-8 proteins. Molecular Operating Environment MOE version 2019.0102 software (Chemical Computing Group, Montreal, Canada) [73] was utilized for the docking studies. Protein and ligand structures were prepared as previously described [74].
The crystal structures of EGFR (PDB: 1M17) [45], Mcl-1 (PDB: 2NLA) [46] and caspase-8 (PDB: 1F9E) [47] were retrieved from the Protein Data Bank (http://www.rcsb.org, accessed on 24 March 2023) [75]. The structures of EGFR, MCl-1 and caspase-8 were prepared through the MOE QuickPrep tool.
cis-Abienol, trans-ferruginol, α-cadinol, δ-muurolene, and α-pinene were drawn through the Chemdraw® (PerkinElmer Informatics, Inc., Buckinghamshire, UK), then transferred to the MOE using smiles canonical. The energy of the components was minimized with root mean square (RMS) gradient 0.1 kcal/mol and finally preparing a database file.

Author Contributions

Conceptualization, E.F., R.O., S.S.E., S.P., A.M.E.-S., M.M.E.-D. and N.T.; Methodology, E.F., S.S.E. and N.T.; Software, S.S.E.; Investigation, E.F., R.O., S.S.E., S.P. and N.T. Validation, E.F., A.M.E.-S., M.M.E.-D. and N.T.; Formal analysis, E.F., R.O., S.S.E., S.P. and N.T.; Resources, R.O., S.P., A.M.E.-S. and M.M.E.-D.; Data curation, E.F., R.O., S.S.E. and N.T.; Writing—original draft preparation E.F., S.S.E. and N.T.; Writing—review and editing, R.O, S.P., A.M.E.-S. and M.M.E.-D.; Supervision, A.M.E.-S. and M.M.E.-D.; Project administration, E.F., A.M.E.-S., M.M.E.-D. and N.T.; Funding acquisition, R.O and S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia Project number (IFKSUOR3–135).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data and materials used are available in the manuscript.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work through the project no. (IFKSUOR3–135).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maiko, I.; Shinichiro, H.; Lyle, E.C. Role of Medicinal and Aromatic Plants: Past, Present, and Future. In Pharmacognosy; Shagufta, P., Areej, A.-T., Eds.; IntechOpen: Rijeka, Croatia, 2019; Chapter 2. [Google Scholar]
  2. Sharifi-Rad, J.; Sureda, A.; Tenore, G.C.; Daglia, M.; Sharifi-Rad, M.; Valussi, M.; Tundis, R.; Sharifi-Rad, M.; Loizzo, M.R.; Ademiluyi, A.O. Biological activities of essential oils: From plant chemoecology to traditional healing systems. Molecules 2017, 22, 70. [Google Scholar] [CrossRef]
  3. Lang, G.; Buchbauer, G. A review on recent research results (2008–2010) on essential oils as antimicrobials and antifungals. A review. Flavour Fragr. J. 2012, 27, 13–39. [Google Scholar] [CrossRef]
  4. Camele, I.; Gruľová, D.; Elshafie, H.S. Chemical composition and antimicrobial properties of Mentha × piperita cv.‘Kristinka’essential oil. Plants 2021, 10, 1567. [Google Scholar] [CrossRef] [PubMed]
  5. Abd-ElGawad, A.M.; Bonanomi, G.; Al-Rashed, S.A.; Elshamy, A.I. Persicaria lapathifolia essential oil: Chemical constituents, antioxidant activity, and allelopathic effect on the weed Echinochloa colona. Plants 2021, 10, 1798. [Google Scholar] [CrossRef]
  6. Mohamed, M.E.; Mohafez, O.M.; Khalil, H.E.; Alhaider, I.A. Essential oil from myrtle leaves growing in the eastern part of Saudi Arabia: Components, anti-inflammatory and cytotoxic activities. J. Essent. Oil Bear. Plants 2019, 22, 877–892. [Google Scholar] [CrossRef]
  7. Bharti, K.; Sharma, M.; Vyas, G.K.; Sharma, S. A Review on Phytochemical Pharmacological and Biological Activities of Thuja Occidentalis. Asian J. Pharm. Res. Dev. 2022, 10, 111–115. [Google Scholar] [CrossRef]
  8. Mazari, K.; Bendimerad, N.; Bekhechi, C.; Fernandez, X. Chemical composition and antimicrobial activity of essential oils isolated from Algerian Juniperus phoenicea L. and Cupressus sempervirens L. J. Med. Plants Res. 2010, 4, 959–964. [Google Scholar]
  9. Ben Nouri, A.; Dhifi, W.; Bellili, S.; Ghazghazi, H.; Aouadhi, C.; Chérif, A.; Hammami, M.; Mnif, W. Chemical composition, antioxidant potential, and antibacterial activity of essential oil cones of Tunisian Cupressus sempervirens. J. Chem. 2015, 2015, 538929. [Google Scholar] [CrossRef] [Green Version]
  10. Galovičová, L.; Čmiková, N.; Schwarzová, M.; Vukic, M.D.; Vukovic, N.L.; Kowalczewski, P.Ł.; Bakay, L.; Kluz, M.I.; Puchalski, C.; Obradovic, A.D. Biological Activity of Cupressus sempervirens Essential Oil. Plants 2023, 12, 1097. [Google Scholar] [CrossRef]
  11. Boufares, K.; Hassani, A.; Alem, A.S. Essential oil composition and antimicrobial activities of some Cupressaceae species from Algeria against two phytopathogenic microorganisms. J. Crop. Prot. 2019, 8, 223–234. [Google Scholar]
  12. Bajes, H.R.; Oran, S.A.; Bustanji, Y.K. Chemical composition and antiproliferative and antioxidant activities of essential oil from Juniperus phoenicea L. cupressaceae. Res. J. Pharm. Technol. 2022, 15, 153–159. [Google Scholar] [CrossRef]
  13. Kim, H.-Y.; Lee, S.-G.; Oh, T.-J.; Lim, S.R.; Kim, S.-H.; Lee, H.J.; Kim, Y.-S.; Choi, H.-K. Antiproliferative and apoptotic activity of Chamaecyparis obtusa leaf extract against the HCT116 human colorectal cancer cell line and investigation of the bioactive compound by gas chromatography-mass spectrometry-based metabolomics. Molecules 2015, 20, 18066–18082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhang, Y.-M.; Xu, J.; Xiao, L.; Zeng, G.-Z.; Sun, Z.-H.; Tan, N.-H. A New Phenolic Glycoside from Chamaecyparis obtusa var. breviramea f. crippsii. Molecules 2013, 18, 1255–1261. [Google Scholar] [CrossRef] [PubMed]
  15. Khanavi, M.; Enayati, A.; Shams Ardekani, M.; Akbarzadeh, T.; Karimpour Razkenari, E.; Eftekhari, M. Cytotoxic activity of Juniperus excelsa M. Bieb. Leaves Essent. Oil Breast Cancer Cell Lines. Res. J. Pharm. 2019, 6, 1–7. [Google Scholar]
  16. Jlizi, S.; Lahmar, A.; Zardi-Bergaoui, A.; Ascrizzi, R.; Flamini, G.; Harrath, A.H.; Chekir-Ghedira, L.; Ben Jannet, H. Chemical composition and cytotoxic activity of the fractionated trunk bark essential oil from Tetraclinis articulata (Vahl) Mast. growing in Tunisia. Molecules 2021, 26, 1110. [Google Scholar] [CrossRef]
  17. Elsharkawy, E.; Aljohar, H.; Donia, A. Comparative study of antioxidant and anticancer activity of Thuja orientalis growing in Egypt and Saudi Arabia. Br. J. Pharm. Res. 2017, 15, 1–9. [Google Scholar] [CrossRef]
  18. Elshafie, H.; Caputo, L.; De Martino, L.; Gruľová, D.; Zheljazkov, V.; De Feo, V.; Camele, I. Biological investigations of essential oils extracted from three Juniperus species and evaluation of their antimicrobial, antioxidant and cytotoxic activities. J. Appl. Microbiol. 2020, 129, 1261–1271. [Google Scholar] [CrossRef]
  19. Fayed, S.A. Chemical composition, antioxidant, anticancer properties and toxicity evaluation of leaf essential oil of Cupressus sempervirens. Not. Bot. Horti Agrobot. Cluj-Napoca 2015, 43, 320–326. [Google Scholar] [CrossRef] [Green Version]
  20. Rathi, A.; Baburaj, D.; Sundaram, E.; Kumar, S.; Khurana, A.; Manchanda, R. Pharmacognostic study of Chamaecyparis lawsoniana (Murr.) Parl.: A drug used in Homoeopathy. Indian J. Res. Homoeopath. 2015, 9, 20–24. [Google Scholar] [CrossRef]
  21. Palá-Paúl, J.; Usano-Alemany, J.; Granda, E.; Soria, A.-C. Antifungal and antibacterial activity of the essential oil of Chamaecyparis lawsoniana from Spain. Nat. Prod. Commun. 2012, 7, 1934578X1200701036. [Google Scholar] [CrossRef] [Green Version]
  22. De Pooter, H.; Vermeesch, J.; De Buyck, L.; Huang, Q.-L.; Schamp, N.; De Bruyn, A. Volatile Fractions of Chamaecyparis lawsoniana (A. Murray) Parl, and Chamaecyparis pisifera (Sieb. et Zucc.) Endl. var. filifera. Identification of Oplopanonyl Acetate. J. Essent. Oil Res. 1991, 3, 1–6. [Google Scholar] [CrossRef]
  23. Emami, S.A.; Massoomi, H.; Moghadam, M.S.; Asili, J. Identification of volatile oil components from aerial parts of Chamaecyparis lawsoniana by GC-MS and 13C-NMR methods. J. Essent. Oil Bear. Plants 2009, 12, 661–665. [Google Scholar] [CrossRef]
  24. Giatropoulos, A.; Pitarokili, D.; Papaioannou, F.; Papachristos, D.P.; Koliopoulos, G.; Emmanouel, N.; Tzakou, O.; Michaelakis, A. Essential oil composition, adult repellency and larvicidal activity of eight Cupressaceae species from Greece against Aedes albopictus (Diptera: Culicidae). Parasitol. Res. 2013, 112, 1113–1123. [Google Scholar] [CrossRef] [PubMed]
  25. Nikolić, B.M.; Milanović, S.D.; Milenković, I.L.; Todosijević, M.M.; Đorđević, I.Ž.; Brkić, M.Z.; Mitić, Z.S.; Marin, P.D.; Tešević, V.V. Bioactivity of Chamaecyparis lawsoniana (A. Murray) Parl. and Thuja plicata Donn ex D. Don essential oils on Lymantria dispar (Linnaeus, 1758) (Lepidoptera: Erebidae) larvae and Phytophthora de Bary 1876 root pathogens. Ind. Crop. Prod. 2022, 178, 114550. [Google Scholar] [CrossRef]
  26. Smith, E.C.; Williamson, E.M.; Wareham, N.; Kaatz, G.W.; Gibbons, S. Antibacterials and modulators of bacterial resistance from the immature cones of Chamaecyparis lawsoniana. Phytochemistry 2007, 68, 210–217. [Google Scholar] [CrossRef]
  27. Zazharskyi, V.; Davydenko, P.; Kulishenko, O.; Borovik, I.; Kabar, A.; Brygadyrenko, V. Antibacterial and fungicidal effect of ethanol extracts from Juniperus sabina, Chamaecyparis lawsoniana, Pseudotsuga menziesii and Cephalotaxus harringtonia. Regul. Mech. Biosyst. 2020, 11, 105–109. [Google Scholar] [CrossRef]
  28. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  29. Fernald, K.; Kurokawa, M. Evading apoptosis in cancer. Trends Cell Biol. 2013, 23, 620–633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Kumar, B.; Singh, S.; Skvortsova, I.; Kumar, V. Promising targets in anti-cancer drug development: Recent updates. Curr. Med. Chem. 2017, 24, 4729–4752. [Google Scholar] [CrossRef]
  31. Abdullaev, F. Pharmacology and therapeutics in the new millennium. In New Delhi: Narosa Puplising House. Plant-Derived Agents against Cancer; Springer: Berlin/Heidelberg, Germany, 2001; pp. 345–354. [Google Scholar]
  32. Sarker, S.D.; Nahar, L.; Miron, A.; Guo, M. Anticancer natural products. In Annual Reports in Medicinal Chemistry; Elsevier: Amsterdam, The Netherlands, 2020; Volume 55, pp. 45–75. [Google Scholar]
  33. Pham-Huy, L.A.; He, H.; Pham-Huy, C. Free radicals, antioxidants in disease and health. Int. J. Biomed. Sci. IJBS 2008, 4, 89. [Google Scholar]
  34. Mani, S. Production of reactive oxygen species and its implication in human diseases. In Free Radicals in Human Health and Disease; Springer: Berlin/Heidelberg, Germany, 2015; pp. 3–15. [Google Scholar]
  35. Adams, R.P. Identification of Essential Oil Components by Gas Chromatography/Mass Spectrometry, 5th ed.; Texensis Publishing: Gruver, TX, USA, 2017. [Google Scholar]
  36. Kasali, A.A.; Ekundayo, O.; Paul, C.; König, W.A. epi-Cubebanes from Solidago canadensis. Phytochemistry 2002, 59, 805–810. [Google Scholar] [CrossRef]
  37. Sacchetti, G.; Medici, A.; Maietti, S.; Radice, M.; Muzzoli, M.; Manfredini, S.; Braccioli, E.; Bruni, R. Composition and functional properties of the essential oil of Amazonian basil, Ocimum micranthum Willd., Labiatae in comparison with commercial essential oils. J. Agric. Food Chem. 2004, 52, 3486–3491. [Google Scholar] [CrossRef]
  38. Asuming, W.A.; Beauchamp, P.S.; Descalzo, J.T.; Dev, B.C.; Dev, V.; Frost, S.; Ma, C.W. Essential oil composition of four Lomatium Raf. species and their chemotaxonomy. Biochem. Syst. Ecol. 2005, 33, 17–26. [Google Scholar] [CrossRef]
  39. Yaşar, A.; Üçüncü, O.; Güleç, C.; İnceer, H.; Ayaz, S.; Yayl, N. GC-MS Analysis of Chloroform Extracts in Flowers, Stems, and Roots of Tripleurospermum callosum. Pharm. Biol. 2005, 43, 108–112. [Google Scholar] [CrossRef] [Green Version]
  40. Gooré, S.G.; Ouattara, Z.A.; Yapi, A.T.; Békro, Y.-A.; Bighelli, A.; Paoli, M.; Tomi, F. Chemical composition of the leaf oil of Artabotrys jollyanus from Côte d’Ivoire. Rev. Bras. Farmacogn. 2017, 27, 414–418. [Google Scholar] [CrossRef]
  41. Duquesnoy, E.; Marongiu, B.; Castola, V.; Piras, A.; Porcedda, S.; Casanova, J. Combined analysis by GC (RI), GC-MS and 13C NMR of the supercritical fluid extract of Abies alba twigs. Nat. Prod. Commun. 2010, 5, 1934578X1000501235. [Google Scholar] [CrossRef] [Green Version]
  42. Isidorov, V.A.; Lech, P.; Żółciak, A.; Rusak, M.; Szczepaniak, L. Gas chromatographic–mass spectrometric investigation of metabolites from the needles and roots of pine seedlings at early stages of pathogenic fungi Armillaria ostoyae attack. Trees 2008, 22, 531–542. [Google Scholar] [CrossRef]
  43. Stashenko, E.E.; Cervantes, M.; Combariza, Y.; Fuentes, H.; Martínez, J.R. HRGC/FID and HRGC/MSD analysis of the secondary metabolites obtained by different extraction methods from Lepechinia schiedeana, and in vitro evaluation of its antioxidant activity. J. High Resolut. Chromatogr. 1999, 22, 343–349. [Google Scholar] [CrossRef]
  44. Veličković, D.T.; Ranđelović, N.V.; Ristić, M.S.; Veličković, A.S.; Šmelcerović, A.A. Chemical constituents and antimicrobial activity of the ethanol extracts obtained from the flower, leaf and stem of Salvia officinalis L. J. Serb. Chem. Soc. 2003, 68, 17–24. [Google Scholar] [CrossRef]
  45. Stamos, J.; Sliwkowski, M.X.; Eigenbrot, C. Structure of the epidermal growth factor receptor kinase domain alone and in complex with a 4-anilinoquinazoline inhibitor. J. Biol. Chem. 2002, 277, 46265–46272. [Google Scholar] [CrossRef] [Green Version]
  46. Czabotar, P.E.; Lee, E.F.; van Delft, M.F.; Day, C.L.; Smith, B.J.; Huang, D.C.; Fairlie, W.D.; Hinds, M.G.; Colman, P.M. Structural insights into the degradation of Mcl-1 induced by BH3 domains. Proc. Natl. Acad. Sci. USA 2007, 104, 6217–6222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Blanchard, H.; Donepudi, M.; Tschopp, M.; Kodandapani, L.; Wu, J.C.; Grütter, M.G. Caspase-8 specificity probed at subsite S4: Crystal structure of the caspase-8-Z-DEVD-cho complex. J. Mol. Biol. 2000, 302, 9–16. [Google Scholar] [CrossRef]
  48. Karawya, M.S. Volatile oil of Chamaecyparis lawsoniana Parl. Egypt. J. Pharm. Sci. 1986, 27, 341–346. [Google Scholar]
  49. Barbosa, L.C.A.; Filomeno, C.A.; Teixeira, R.R. Chemical variability and biological activities of Eucalyptus spp. essential oils. Molecules 2016, 21, 1671. [Google Scholar] [CrossRef] [Green Version]
  50. Yoewono, J.R.; Purwanto, R.O.; Woolley, C.L.; Chang, C.; Shih, W. Antioxidant activities and oral toxicity studies of Chamaecyparis formosensis and Cymbopogon nardus essential oils. Int. J. Adv. Sci. Res. Manag 2016, 1, 47–58. [Google Scholar]
  51. Eltayeb, L.M.; Yagi, S.; Mohamed, H.M.; Zengin, G.; Shariati, M.A.; Rebezov, M.; Uba, A.I.; Lorenzo, J.M. Essential Oils Composition and Biological Activity of Chamaecyparis obtusa, Chrysopogon nigritanus and Lavandula coronopifolia Grown Wild in Sudan. Molecules 2023, 28, 1005. [Google Scholar] [CrossRef]
  52. Kim, S.; Lee, S.; Hong, C.; Gwak, K.; Yeo, H.; Lee, J.; Choi, I. Whitening and antioxidant activities of essential oils from Cryptomeria japonica and Chamaecyparis obtusa. Mokchae Konghak = J. Korean Wood Sci. Technol. 2011, 39, 291–302. [Google Scholar] [CrossRef] [Green Version]
  53. Gao, H.; Shupe, T.F.; Hse, C.Y.; Eberhardt, T.L. Antioxidant activity of extracts from the bark of Chamaecyparis lawsoniana (A. Murray) Parl. De Gruyter 2006, 60, 459–462. [Google Scholar] [CrossRef]
  54. Miguel, M.G. Antioxidant and anti-inflammatory activities of essential oils: A short review. Molecules 2010, 15, 9252–9287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Pérez-González, C.; Pérez-Ramos, J.; Méndez-Cuesta, C.A.; Serrano-Vega, R.; Martell-Mendoza, M.; Pérez-Gutiérrez, S. Cytotoxic activity of essential oils of some species from Lamiaceae family. In Cytotoxicity: Definition, Identification, and Cytotoxic Compounds; Istifli, E.S., Ila, H.B., Eds.; IntechOpen: Rijeka, Croatia, 2019; pp. 29–43. [Google Scholar]
  56. Capello, T.M.; Martins, E.G.; de Farias, C.F.; Figueiredo, C.R.; Matsuo, A.L.; Passero, L.F.D.; Oliveira-Silva, D.; Sartorelli, P.; Lago, J.H.G. Chemical composition and in vitro cytotoxic and antileishmanial activities of extract and essential oil from leaves of Piper cernuum. Nat. Prod. Commun. 2015, 10, 1934578X1501000217. [Google Scholar] [CrossRef] [Green Version]
  57. Xu, J.; Zeng, G.-Z.; Liu, Y.-M.; Chen, K.-L.; Sun, Z.-H.; Zhang, Y.-M.; Tan, N.-H. Sesquiterpenoids and diterpenes from Chamaecyparis obtusa var. breviramea f. crippsii. Z. Nat. B 2014, 69, 362–368. [Google Scholar] [CrossRef]
  58. Wee, P.; Wang, Z. Epidermal growth factor receptor cell proliferation signaling pathways. Cancers 2017, 9, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ongko, J.; Setiawan, J.V.; Feronytha, A.G.; Juliana, A.; Effraim, A.; Wahjudi, M.; Antonius, Y. In-silico screening of inhibitor on protein epidermal growth factor receptor (EGFR). In Proceedings of the IOP Conference Series: Earth and Environmental Science, Qingdao, China, 16–18 July 2022; p. 012075. [Google Scholar]
  60. Mady, A. Development and Characterization of Novel Mcl-1 Inhibitors for Treatment of Cancer. Ph.D. Thesis, University of Michigan, Ann Arbor, MI, USA, 2016. [Google Scholar]
  61. Wong, R.S. Apoptosis in cancer: From pathogenesis to treatment. J. Exp. Clin. Cancer Res. 2011, 30, 87. [Google Scholar] [CrossRef] [Green Version]
  62. Negi, A.; Murphy, P.V. Development of Mcl-1 inhibitors for cancer therapy. Eur. J. Med. Chem. 2021, 210, 113038. [Google Scholar] [CrossRef]
  63. Wei, S.-H.; Dong, K.; Lin, F.; Wang, X.; Li, B.; Shen, J.-j.; Zhang, Q.; Wang, R.; Zhang, H.-Z. Inducing apoptosis and enhancing chemosensitivity to gemcitabine via RNA interference targeting Mcl-1 gene in pancreatic carcinoma cell. Cancer Chemother. Pharmacol. 2008, 62, 1055–1064. [Google Scholar] [CrossRef] [PubMed]
  64. Guoan, X.; Hanning, W.; Kaiyun, C.; Hao, L. Adenovirus-mediated siRNA targeting Mcl-1 gene increases radiosensitivity of pancreatic carcinoma cells in vitro and in vivo. Surgery 2010, 147, 553–561. [Google Scholar] [CrossRef]
  65. Mnich, K.; Koryga, I.; Pakos-Zebrucka, K.; Thomas, M.; Logue, S.E.; Eriksson, L.A.; Gorman, A.M.; Samali, A. The stressosome, a caspase-8-activating signalling complex assembled in response to cell stress in an ATG5-mediated manner. J. Cell. Mol. Med. 2021, 25, 8809–8820. [Google Scholar] [CrossRef]
  66. Fulda, S.; Debatin, K.-M. Extrinsic versus intrinsic apoptosis pathways in anticancer chemotherapy. Oncogene 2006, 25, 4798–4811. [Google Scholar] [CrossRef] [Green Version]
  67. Elkomy, N.M.; El-Shaibany, A.; Elnagar, G.M.; Abdelkhalek, A.S.; Al-Mahbashi, H.; Elaasser, M.M.; Raweh, S.M.; Aldiyarbi, M.A.; Raslan, A.E. Evaluation of acute oral toxicity, anti-diabetic and antioxidant effects of Aloe vera flowers extract. J. Ethnopharmacol. 2023, 309, 116310. [Google Scholar] [CrossRef]
  68. Ling, L.T.; Yap, S.-A.; Radhakrishnan, A.K.; Subramaniam, T.; Cheng, H.M.; Palanisamy, U.D. Standardised Mangifera indica extract is an ideal antioxidant. Food Chem. 2009, 113, 1154–1159. [Google Scholar] [CrossRef]
  69. Elaasser, M.; Morsi, M.; Galal, S.; Abd El-Rahman, M.; Katry, M. Antioxidant, anti-inflammatory and cytotoxic activities of the unsaponifiable fraction of extra virgin olive oil. Grasas Aceites 2020, 71, e386. [Google Scholar] [CrossRef]
  70. Gomha, S.M.; Riyadh, S.M.; Mahmmoud, E.A.; Elaasser, M.M. Synthesis and anticancer activity of arylazothiazoles and 1,3,4-thiadiazoles using chitosan-grafted-poly(4-vinylpyridine) as a novel copolymer basic catalyst. Chem. Heterocycl. Compd. 2015, 51, 1030–1038. [Google Scholar] [CrossRef]
  71. Yousefbeyk, F.; Dabirian, S.; Ghanbarzadeh, S.; Eghbali Koohi, D.; Yazdizadeh, P.; Ghasemi, S. Green synthesis of silver nanoparticles from Stachys byzantina K. Koch: Characterization, antioxidant, antibacterial, and cytotoxic activity. Part. Sci. Technol. 2022, 40, 219–232. [Google Scholar] [CrossRef]
  72. Mansour, K.A.; Elbermawi, A.; Al-Karmalawy, A.A.; Lahloub, M.-F.; El-Neketi, M. Cytotoxic effects of extracts obtained from plants of the Oleaceae family: Bio-guided isolation and molecular docking of new secoiridoids from Jasminum humile. Pharm. Biol. 2022, 60, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  73. Chemical Computing Group. Molecular Operating Environment; Chemical Computing Group Inc.: Montreal, QC, Canada, 2008. [Google Scholar]
  74. Mohamed, M.E.; Tawfeek, N.; Elbaramawi, S.S.; Elbatreek, M.H.; Fikry, E. Agathis robusta Bark Extract Protects from Renal Ischemia-Reperfusion Injury: Phytochemical, In Silico and In Vivo Studies. Pharmaceuticals 2022, 15, 1270. [Google Scholar] [CrossRef] [PubMed]
  75. RCSB Protein Data Bank (PDB). Available online: https://www.rcsb.org/ (accessed on 24 March 2023).
Figure 1. GC/MS chromatogram of Chamaecyparis lawsoniana leaf essential oil. Numbers in red are related to Table 1.
Figure 1. GC/MS chromatogram of Chamaecyparis lawsoniana leaf essential oil. Numbers in red are related to Table 1.
Plants 12 02475 g001
Figure 2. Chemical structures of the main components present in Chamaecyparis lawsoniana leaf essential oil.
Figure 2. Chemical structures of the main components present in Chamaecyparis lawsoniana leaf essential oil.
Plants 12 02475 g002
Figure 3. In vitro cytotxicity of Chamaecyparis lawsoniana leaf essential oil and cisplatin against: (a) human breast (MCF-7), (b) colon (HCT-116), (c) lung (A-549) and (d) hepatocellular (HepG-2) carcinoma cell lines.
Figure 3. In vitro cytotxicity of Chamaecyparis lawsoniana leaf essential oil and cisplatin against: (a) human breast (MCF-7), (b) colon (HCT-116), (c) lung (A-549) and (d) hepatocellular (HepG-2) carcinoma cell lines.
Plants 12 02475 g003
Figure 4. Chamaecyparis lawsoniana (A. Murray) Parl. (a) Whole tree (b) Fresh leaves.
Figure 4. Chamaecyparis lawsoniana (A. Murray) Parl. (a) Whole tree (b) Fresh leaves.
Plants 12 02475 g004
Table 1. Chemical composition of Chamaecyparis lawsoniana leaf essential oil.
Table 1. Chemical composition of Chamaecyparis lawsoniana leaf essential oil.
Peak Compound NameChemical ClassRIExp. aRILit. bArea%Identification c
1α-ThujeneBicyclic monoterpene hydrocarbon9229240.13MS, RI
2α-PineneBicyclic monoterpene hydrocarbon 9319326.20MS, RI
3SabineneBicyclic monoterpene hydrocarbon9689690.13MS, RI
4β-PineneBicyclic monoterpene hydrocarbon9719740.06MS, RI
5 MyrceneAcyclic monoterpene hydrocarbon9879880.24MS, RI
6 α-TerpineneMonocyclic monoterpene hydrocarbon101210140.15MS, RI
7 p-CymeneAromatic monoterpene hydrocarbon102010200.03MS, RI
8 LimoneneMonocyclic monoterpene hydrocarbon102410240.14MS, RI
9γ-TerpineneMonocyclic monoterpene hydrocarbon105510540.36MS, RI
10 TerpinoleneMonocyclic monoterpene hydrocarbon108410860.15MS, RI
11 trans-Sabinene hydrateMonocyclic monoterpene alcohol111710980.02MS
12 Terpinen-4-olMonocyclic monoterpene alcohol117611741.20MS, RI
13 α-TerpineolMonocyclic monoterpene alcohol118811860.21MS, RI
14 Bornyl acetateBicyclic monoterpene ester128212840.02MS, RI
15 α-Terpinyl acetateMonocyclic monoterpene ester134613460.12MS, RI
16α-IonolMonocyclic sesquiterpene alcohol137613760.30MS, RI
17 β-ElemeneMonocyclic sesquiterpene hydrocarbon139013890.12MS, RI
18 β-CaryophylleneBicyclic sesquiterpene hydrocarbon141814170.15MS, RI
19 6-epi-β-CubebeneBicyclic sesquiterpene hydrocarbon14481449 3.16MS, RI
20 α-HumuleneMonocyclic sesquiterpene hydrocarbon145414520.14MS, RI
21 δ-Muurolene (cis-muurola-4(14)-diene)Bicyclic sesquiterpene hydrocarbon146714678.57MS, RI
22 γ-MuuroleneBicyclic sesquiterpene hydrocarbon147714780.09MS, RI
23 Germacrene DBicyclic sesquiterpene hydrocarbon148314801.32MS, RI
24 β-SelineneBicyclic sesquiterpene hydrocarbon148714890.03MS, RI
25 cis-Cadina-1,4-dieneBicyclic sesquiterpene hydrocarbon149414950.17MS, RI
26 γ-AmorpheneBicyclic sesquiterpene hydrocarbon150014954.10MS, RI
27 γ-CadineneBicyclic sesquiterpene hydrocarbon151415130.08MS, RI
28 δ-CadineneBicyclic sesquiterpene hydrocarbon152415223.71MS, RI
29 α-CadineneBicyclic sesquiterpene hydrocarbon153715370.05MS, RI
30 α-CalacoreneAromatic bicyclic sesquiterpene hydrocarbon154315440.03MS, RI
31 ElemolMonocyclic sesquiterpene alcohol155015480.59MS, RI
32 trans-NerolidolAcyclic sesquiterpene alcohol156115610.04MS, RI
33 Caryophyllene oxideBicyclic sesquiterpene oxide158415820.19MS, RI
34 Humulene epoxide IIMonocyclic sesquiterpene epoxide161116080.04MS, RI
35 1,10-di-epi-CubenolBicyclic sesquiterpene alcohol161716181.52MS, RI
36 JunenolBicyclic sesquiterpene alcohol162116180.04MS, RI
37 γ-EudesmolBicyclic sesquiterpene alcohol163416300.14MS, RI
38 epi-α-Cadinol (tau-cadinol)Bicyclic sesquiterpene alcohol164416381.74MS, RI
39 CubenolBicyclic sesquiterpene alcohol165016450.21MS, RI
40 β-EudesmolBicyclic sesquiterpene alcohol165416490.22MS, RI
41 α-CadinolBicyclic sesquiterpene alcohol16611660 8.84MS, RI
42 Germacra-4(15),5,10(14)-trien-1-α-olMonocyclic sesquiterpene alcohol169016850.44MS, RI
43 cis-14-nor-Muurol-5-en-4-oneBicyclic sesquiterpene ketone169216880.15MS, RI
44 6-Isopropenyl-4,8a-dimethyl-1,2,3,5,6,7,8,
8a-octahydronaphthalene-2-ol
Bicyclic sesquiterpene alcohol171017140.08MS, RI
45 7-HydroxycalameneneBicyclic sesquiterpene alcohol179818030.07MS, RI
46 Isopimara-9(11),15-dieneTricyclic diterpene hydrocarbon191519050.10MS, RI
47 PimaradieneTricyclic diterpene hydrocarbon194919480.04MS, RI
48 Sandaracopimara-8(14),15-diene
(13-isopimaradiene)
Tricyclic diterpene hydrocarbon196919682.05MS, RI
49 Kaur-15-eneTetracyclic diterpene hydrocarbon199719970.43MS, RI
50 13-epi-Manool oxideTricyclic diterpene oxide201920090.04MS, RI
51 AbietatrieneTricyclic diterpene hydrocarbon206120551.89MS, RI
52 AbietadieneTricyclic diterpene hydrocarbon208720870.07MS, RI
53 cis-AbienolBicyclic diterpene alcohol21202112 23.43MS, RI
54 SandaracopimarinalTricyclic diterpene aldehyde219521840.83MS
55 Pimara-7,15-dien-3-oneTricyclic diterpene ketone222522271.65MS, RI
56 SandaracopimarinolTricyclic diterpene alcohol228222690.73MS
57 trans-TotarolTricyclic diterpene alcohol229122901.27MS, RI
58 trans-FerruginolTricyclic diterpene alcohol2325233114.31MS, RI
59 cis-FerruginolTricyclic diterpene alcohol23372340 0.44MS, RI
Total identified 92.77
Monoterpenes hydrocarbons 7.59
Oxygenated monoterpenes 1.57
Sesquiterpene hydrocarbons 21.72
Oxygenated sesquiterpenes 14.61
Diterpene hydrocarbons 4.58
Oxygenated diterpenes 42.70
Compounds are ranked in the order of their elution on Rtx-5MS GC column. a Retention index defined experimentally on Rtx-5MS column relative to C8–C28 n-alkanes. b Reported retention index. c Identification was established by matching of the mass spectral (MS) data and retention index (RI) values of compounds from Adams library [35], NIST 11 Mass Spectral Library, Wiley Registry of Mass Spectral Data 10th edition, and the literature [36,37,38,39,40,41,42,43,44].
Table 2. Antioxidant activity of Chamaecyparis lawsoniana leaf essential oil.
Table 2. Antioxidant activity of Chamaecyparis lawsoniana leaf essential oil.
IC50 ± SD (µg/mL)
DPPHABTSFRAP
CLLEO116.91 ± 5.7373.02 ± 4.06218.64 ± 8.41
Ascorbic acid10.22 ± 0.5610.66 ± 0.8920.89± 1.25
Table 3. Cytotoxicity of Chamaecyparis lawsoniana leaf essential oil and cisplatin against some cell lines. CC50 (50% cytotoxic concentration) and IC50 (50% inhibitory concentration) values are expressed by mean ± standard deviation; SI (selectivity index) = CC50 value of normal cell/IC50 value of cancer cell.
Table 3. Cytotoxicity of Chamaecyparis lawsoniana leaf essential oil and cisplatin against some cell lines. CC50 (50% cytotoxic concentration) and IC50 (50% inhibitory concentration) values are expressed by mean ± standard deviation; SI (selectivity index) = CC50 value of normal cell/IC50 value of cancer cell.
MRC-5MCF-7HCT-116A-549HepG-2
CC50IC50SIIC50SIIC50SIIC50SI
CLLEO95.17 ± 3.7123.74 ± 1.724.0128.27 ± 2.133.3725.79 ± 1.953.6915.34 ± 0.966.20
Cisplatin19.43 ± 3.665.69 ± 0.373.412.51 ± 0.677.747.51 ± 0.822.593.68 ± 0.245.28
Table 4. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on EGFR.
Table 4. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on EGFR.
ComponentS Score Kcal/mol3D Protein-Component Interactions
cis-Abienol −5.5184Plants 12 02475 i001
trans-Ferruginol −5.7955Plants 12 02475 i002
α-Cadinol −5.1630Plants 12 02475 i003
δ-Muurolene −5.7461Plants 12 02475 i004
α-Pinene −4.9882Plants 12 02475 i005
Table 5. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on Mcl-1.
Table 5. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on Mcl-1.
ComponentS Score Kcal/mol3D Protein-Component Interactions
cis-Abienol −5.3337Plants 12 02475 i006
trans-Ferruginol −5.2220Plants 12 02475 i007
α-Cadinol −4.5189Plants 12 02475 i008
δ-Muurolene −4.4583Plants 12 02475 i009
α-Pinene −4.1318Plants 12 02475 i010
Table 6. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on caspase-8.
Table 6. Docking details of the selected components of Chamaecyparis lawsoniana leaf essential oil on caspase-8.
ComponentS Score Kcal/mol3D Protein-Component Interactions
cis-Abienol−5.3517Plants 12 02475 i011
trans-Ferruginol −5.4981 Plants 12 02475 i012
α-Cadinol −5.0823 Plants 12 02475 i013
δ-Muurolene −4.7582 Plants 12 02475 i014
α-Pinene −4.2315 Plants 12 02475 i015
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fikry, E.; Orfali, R.; Elbaramawi, S.S.; Perveen, S.; El-Shafae, A.M.; El-Domiaty, M.M.; Tawfeek, N. Chamaecyparis lawsoniana Leaf Essential Oil as a Potential Anticancer Agent: Experimental and Computational Studies. Plants 2023, 12, 2475. https://doi.org/10.3390/plants12132475

AMA Style

Fikry E, Orfali R, Elbaramawi SS, Perveen S, El-Shafae AM, El-Domiaty MM, Tawfeek N. Chamaecyparis lawsoniana Leaf Essential Oil as a Potential Anticancer Agent: Experimental and Computational Studies. Plants. 2023; 12(13):2475. https://doi.org/10.3390/plants12132475

Chicago/Turabian Style

Fikry, Eman, Raha Orfali, Samar S. Elbaramawi, Shagufta Perveen, Azza M. El-Shafae, Maher M. El-Domiaty, and Nora Tawfeek. 2023. "Chamaecyparis lawsoniana Leaf Essential Oil as a Potential Anticancer Agent: Experimental and Computational Studies" Plants 12, no. 13: 2475. https://doi.org/10.3390/plants12132475

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