Next Article in Journal
Resolvins, Protectins, and Maresins: DHA-Derived Specialized Pro-Resolving Mediators, Biosynthetic Pathways, Synthetic Approaches, and Their Role in Inflammation
Next Article in Special Issue
Analysis of Fatty Acids, Amino Acids and Volatile Profile of Apple By-Products by Gas Chromatography-Mass Spectrometry
Previous Article in Journal
Chemotaxonomic Classification of Peucedanum japonicum and Its Chemical Correlation with Peucedanum praeruptorum, Angelica decursiva, and Saposhnikovia divaricata by Liquid Chromatography Combined with Chemometrics
Previous Article in Special Issue
Supercritical CO2 Extraction of Narcissus poeticus L. Flowers for the Isolation of Volatile Fragrance Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical Composition, Biological Activities and In Silico Analysis of Essential Oils of Three Endemic Prangos Species from Turkey

by
Gokhan Zengin
1,
Mohamad Fawzi Mahomoodally
2,
Evren Yıldıztugay
3,
Sharmeen Jugreet
2,
Shafi Ullah Khan
4,
Stefano Dall’Acqua
5,
Adriano Mollica
6,
Abdelhakim Bouyahya
7 and
Domenico Montesano
8,*
1
Department of Biology, Science Faculty, Selcuk University, Konya 42130, Turkey
2
Department of Health Sciences, Faculty of Medicine and Health Sciences, University of Mauritius, Réduit 80837, Mauritius
3
Deparment of Biotechnology, Science Faculty, Selcuk University, Konya 42130, Turkey
4
Department of Pharmacy, Abasyn University, Peshawar 25000, Pakistan
5
Department of Pharmaceutical and Pharmacological Sciences, University of Padova, Via Marzolo 5, 35131 Padova, Italy
6
Department of Pharmacy, University “G. d’Annunzio” of Chieti-Pescara, 66100 Chieti, Italy
7
Laboratory of Human Pathologies Biology, Department of Biology, Faculty of Sciences, Mohammed V University in Rabat, Rabat 10106, Morocco
8
Department of Pharmacy, University of Naples Federico II, Via D. Montesano 49, 80131 Naples, Italy
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(5), 1676; https://doi.org/10.3390/molecules27051676
Submission received: 11 February 2022 / Revised: 28 February 2022 / Accepted: 1 March 2022 / Published: 3 March 2022
(This article belongs to the Special Issue Featured Papers on Bioactive Flavour and Fragrance Compounds 2022)

Abstract

:
In this study, the essential oils (EOs) obtained from three endemic Prangos species from Turkey (P. heyniae, P. meliocarpoides var. meliocarpoides, and P. uechtritzii) were studied for their chemical composition and biological activities. β-Bisabolenal (12.2%) and caryophyllene oxide (7.9%) were the principal components of P. heyniae EO, while P. meliocarpoides EO contained sabinene (16.7%) and p-cymene (13.2%), and P. uechtritzii EO contained p-cymene (24.6%) and caryophyllene oxide (19.6%), as the most abundant components. With regard to their antioxidant activity, all the EOs were found to possess free radical scavenging potential demonstrated in both DPPH and ABTS assays (0.43–1.74 mg TE/g and 24.18–92.99 mg TE/g, respectively). Additionally, while no inhibitory activity was displayed by P. meliocarpoides and P. uechtritzii EOs against both cholinesterases (acetyl- and butyryl-cholinesterases). Moreover, all the EOs were found to act as inhibitors of tyrosinase (46.34–69.56 mg KAE/g). Molecular docking revealed elemol and α-bisabolol to have the most effective binding affinity with tyrosinase and amylase. Altogether, this study unveiled some interesting biological activities of these EOs, especially as natural antioxidants and tyrosinase inhibitors and hence offers stimulating prospects of them in the development of anti-hyperpigmentation topical formulations.

1. Introduction

Essential oils (EOs) are recognized for their exceptional medicinal value and are considered among the most attractive and potent plant-derived products. Eos, also referred to as ethereal oils, are volatile and odorous oils present in only 10% of the plant kingdom and are stored in plants in special brittle secretory structures, for instance glands, secretory hairs, secretory ducts, secretory cavities or resin ducts. EOs have been used as perfumes, flavors in foods and beverage ingredients, or to heal both the body and mind since ages and even today they continue to be of paramount importance [1].
In fact, many studies have focused on the pharmacological and cosmeceutical potentials of EOs such as antioxidant, antimicrobial, anticancer, anti-inflammatory, antiaging and antimelanogenic amongst many others [2,3,4]. Accordingly, growing evidences of the health benefits of these natural essences have prompted researchers to further investigate EOs and their individual components. Additionally, their mechanisms of action have been elucidated with respect to their biological activities [5,6,7,8], which so far are showing promising prospects.
The genus Prangos (Apiaceae) consists of 45 worldwide species and the genus has a wide distribution area ranging from Portugal to Tibet [9]. The genus is represented by 19 taxa, including 19 species in the flora of Turkey, of which 10 of them are endemic [10]. The members of this genus have been widely used in traditional medicine and are greatly valued as spices and medicinal plants in Asia, especially in Iran, Turkey, and Iraq. The above-ground parts, the roots as well as the EOs of different species of this genus have both internal and external applications. The most popular indications of these plants are the alleviation of gastrointestinal symptoms, but various other uses have also been reported [9]. For instance, they are used as carminative, tonic, and anthelmintic agents, to heal scars and in the treatment of external bleeding, gastric or digestive disorders, wounds, and leuckoplakia. Moreover, Prangos species are known to act as stimulants, aphrodisiacs and natural fertilizers [9,11]. Based on the ethnobotanical uses of the members of the Prangos genus, several phytochemical studies have been performed and the presence of different groups of bioactive compounds including coumarins [12,13,14], essential oils [15,16], flavonoids [17,18] and phenolic acids [19] have been reported.
P. uechtritzii Boiss & Hausskn is a rigid and perennial plant. The leaves have long lobes and the flowers are yellow. The plant prefers rock limestone slopes and roadsides. It is generally distributed in the southeastern region of Turkey [20]. It is known as “deli çakşır” in local area of Turkey and has reputed aphrodisiac properties. In addition, the aerial parts of P. uechtritzii have been used to treat hemorrohoids in Anatolian folk medicine [14]. This plant’s antioxidant and antimicrobial effects as well as its chemical composition have been examined in earlier studies [14,21,22]. P. meliocarpoides Boiss. var. meliocarpoides is a perennial plant that ranges in height from 15 to 30 cm. It has yellow and glabrous flowers. It is distributed in the Inner Anatolia region of Turkey up to an altitude of 2000 m [20]. In earlier studies, the antioxidant properties of fruit extracts of the plant have been reported by several authors [21,23]. P. heyniae H. Duman & M.F. Watson is a perennial plant that reaches a height of 80 cm. Its flowers are yellow and glabrous. It is an endemic plant to the flora of Turkey and is distributed in the Central Anatolia region of Turkey, especially the calcareous slopes of Konya [24]. In recent studies, several coumarins and volatile compounds have been isolated from this plant, depending on the plant parts used [12,15].
In recent years, the number of studies reporting experimental data on the biological effects of Prangos species have increased considerably [17,22,25,26,27] and significant information has been gathered on the therapeutic properties of different species. Nevertheless, there are still a few species that have remained largely ignored with regard to certain aspects of their biological potentials, thereby reducing the possibility of their exploitation as phytomedicines. Therefore, in this study, the chemical composition, antioxidant and enzyme inhibitory effect of EOs extracted from three endemic Prangos species from Turkey (P. heyniae, P. meliocarpoides var. meliocarpoides, and P. uechtritzi.) were investigated and the molecular docking technique was applied to elucidate the binding interactions of selected EOs’ components with select enzymes.

2. Results and Discussion

2.1. Essential Oil Composition

The chemical composition of the EOs were analyzed using the gas-chromatography/mass spectrometry (GC/MS) and gas-chromatography-flame ionization detector (GC/FID). A total of 41 components was detected in P. heyniae EO (0.1–12.2%), while 40 components were found in P. meliocarpoides var. meliocarpoides EO (0.1–16.7%). On the other hand, only 30 components were identified in P. uechtritzii EO (0.1–24.6%) (Table 1). Ten compounds (α-pinene, β-pinene, sabinene, myrcene, limonene, p-cymene, α-copaene, γ-muurolene, caryophyllene oxide and spathulenol) were found to be common to all three EOs, although they varied in their percentages. As summarized in Table 2, β-bisabolenal (12.2%), caryophyllene oxide (7.9%), germacrene D (7.8%), elemol (7.4%) and α-humulene (6.7%) were the principal components of P. heyniae EO, while P. meliocarpoides var. meliocarpoides EO contained sabinene (16.7%), p-cymene (13.2%), bornyl acetate (11.8%), α-pinene (6.2%), p-cymen-8-ol (6.1%) as the major components. Moreover, p-cymene (24.6%), caryophyllene oxide (19.6%), 7-epi-1,2-dehydrosesquicineole (12.6%), limonene (3.2%) and α-bisabolol (3.2%) were present as the chief components of P. uechtritzii EO, accounting for 63.2% of the identified compounds (Table 2).
Previous studies have also analyzed the EOs of Prangos species under investigation herein. For instance, the bisabolene ether 7-epi-1,2-dehydrosesquicineole was also found to be present in the hydrodistilled fruit EO of P. uechtritzii as the major component (13.44%) in one study [28]. In another study, the EO of air-dried fruits of P. uechtritzii from East Anatolian region of Turkey was revealed to contain α-pinene (40.82%), nonene (17.03%), β-phellandrene (11.14%), δ-3-carene (7.39%), and p-cymene (4.90%) as principal components [22]. Furthermore, hydrodistilled EOs obtained from P. heyniae collected from four locations in Turkey were also studied for their chemical composition. The EOs were found to be rich in sesquiterpenes, germacrene D (10.3–12.1%), β-bisabolene (14.4%), kessane (26.9%), germacrene B (8.2%), elemol (3.4–46.9%), β-bisabolenal (1.4–70.7%), β-bisabolenol (8.4%) and an eudesmane type sesquiterpene (16.1%) which was later revealed to be 3,7(11)-eudesmadien-2-one [29]. Although, similar EO components were found to be present as previously reported in the EOs of the same species, they were found to vary. Indeed, EOs can vary greatly in their chemical composition both in qualitative and quantitative terms as they can be influenced by several factors including seasonal variations, plant organ, degree of maturity of the plant, geographic origin, and extraction method, among others [30]. The EOs derived from other Prangos species such as P. pabularia Lindl., P. peucedanifolia Fenzl., P. ferulacea, P. platychlaena and P. pabularia Lindl. have also been subject of investigations by other researchers [17,22,31,32].

2.2. Antioxidant Activity

Oxidative stress has been identified as the root cause of the development and evolution of several diseases. Supplementation of exogenous antioxidants or increasing endogenous antioxidant defenses of the body is a promising way of fighting the undesirable effects of reactive oxygen species (ROS) induced oxidative damage [33]. The plant kingdom is a rich source of health-promoting compounds especially as natural antioxidants [34]. Several studies have accentuated on the high antioxidant capacity of plants, their derivatives such as EOs and isolated compounds in the recent years, thus highlighting their usefulness in pharmaceutical, cosmetics, food and beverage industries, especially as some synthetic antioxidants such as BHA and BHT are now suspected to be potentially harmful to human health [35,36,37,38,39].
In the present investigation, the EOs extracted from all the three Prangos species were found to possess free radical scavenging potential in both DPPH and ABTS assays (0.43–1.74 mg TE/g and 24.18–92.99 mg TE/g). While P. uechtritzii EO showed the highest scavenging activity in DPPH assay, P. heyniae EO demonstrated the most significant scavenging activity in ABTS assay. Reducing potential was also noted by all EOs in CUPRAC and FRAP assays (103.15–113.43 mg TE/g and 47.98–61.20 mg TE/g, respectively) (Table 3). In FRAP assay also, P. heyniae EO displayed the highest activity while in CUPRAC assay P. meliocarpoides var. meliocarpoides EO showed higher reducing activity compared the other EOs. Metal chelating activity was demonstrated as well in the range of 28.66–30.94 mg EDTAE/g. Total antioxidant capacity was revealed by the phosphomolybdenum assay and the order of the potency of the EOs were as follows: P. meliocarpoides var. meliocarpoides > P. heyniae > P. uechtritzii (15.64–24.37 mmol TE/g) (Table 3).
Interestingly, several previous studies have also reported other members of the Prangos genus to exhibit strong antioxidant abilities [40,41]. In another study, the antioxidant activities of the water and methanol extracts obtained from the root, herb, and fruits of P. ferulacea, including the three species studied herein (P. heyniae, P. meliocarpoides var. meliocarpoides, and P. uechtritzii) from Konya Province (Turkey) were also evaluated using DPPH and thiobarbituric acid assays [21].

2.3. Enzyme Inhibitory Effects

Enzyme inhibitors play a significant role in the drug discovery process. An understanding of diseases at the molecular level has revealed the root cause of many to be the dysfunction, overexpression, or hyperactivation of enzymes. This hyperactivation or overexpression of enzymes can be treated by using suitable enzyme inhibitors. These efforts have provided several enzyme inhibitors in the clinic, including some from a natural origin [42]. Hence, in this present study, an attempt was made to assess the inhibitory effects of the EOs against some enzymes of clinical interest, notably cholinesterases, tyrosinase, amylase and glucosidase.
Cholinesterases are a group of serine hydrolases that split the neurotransmitter acetylcholine (ACh) and terminate its action. While acetylcholinesterase (AChE) plays the key role in ending cholinergic neurotransmission Butyrylcholinesterase (BChE) is a nonspecific cholinesterase enzyme that hydrolyzes choline-based esters. BChE plays a critical role in maintaining normal cholinergic function like AChE through hydrolyzing ACh [43]. Thus, cholinesterase inhibitors are useful substances that help to interfere with the break-down of ACh and prolong its action [44].
In this study, while no inhibitory activity was demonstrated by P. meliocarpoides var. meliocarpoides and P. uechtritzii EOs against both cholinesterases (AChE and BChE), while P. heyniae EO displayed only anti-BChE activity (9.85 mg GALAE/g) (Table 4).
The inhibition of α-glucosidase and α-amylase, enzymes involved in the digestion of carbohydrates, can significantly diminish the post-prandial surge of blood glucose and consequently can be an important strategy in the management of blood glucose level in type 2 diabetic and borderline patients. Presently, there is renewed interest in plant-based medicines and functional foods modulating physiological effects in the prevention and cure of diabetes and obesity. The plant kingdom is a wide field to search for natural effective oral hypoglycaemic agents that have minor or no side effects [45]. In the present study, the EOs were found to inhibit amylase (0.09–0.61 mmol ACAE/g) only, although the activity was not prominent, whereas they showed no activity against glucosidase (Table 4).
Tyrosinase plays a vital role because it is the critical enzyme and restriction enzyme in the course of melanin composition. Pigment spots and melanoma are markedly increased by cumulative tyrosinase activity and quantity. Consequently, tyrosinase inhibitors have received broad consideration owing to their use as hypopigmented agents in recent years [46].
Herein, all the EOs were found to act as tyrosinase inhibitors with inhibitory activity ranging from 46.34 to 69.56 mg KAE/g (Table 4). The anti-tyrosinase potency of the EOs was obtained in the following order: P. meliocarpoides var. meliocarpoides > P. heyniae > P. uechtritzii.
Another similar study was conducted but with an EO obtained from a different Prangos species, P. gaubae whereby the EO showed AChE (2.97 mg GEs/g oil), BChE (3.30 mg GEs/g oil), α-amylase (1.35 mmol ACEs/g oil), α-glucosidase (38.84 mmol ACEs/g oil), tyrosinase (29.24 mg KAEs/g oil) and lipase (1.59 mmol OEs/g oil) inhibitory activities. Additionally, strong antioxidant effects were observed in antiradical (DPPH and ABTS), reducing power (CUPRAC and FRAP), total antioxidant, as well as metal chelating assays [41]. On the other hand, in a recent study [40], methanol extracts of P. ferulacea (131.94 mg kojic acid (KAE) equivalent/g extract) and P. peucedanifolia (4.97 mmol acarbose equivalent (ACAE)/g extract) were reported to be potent inhibitors of tyrosinase and α-glucosidase, respectively.

Multivariate Analysis

To gain more insights between the detected compounds and biological activities of the tested essential oils, we performed a PLS analysis. The results are given in Figure 1. R2X and Q2 values are indicator the quality of PLS parameters and the values were 0.87 and 0.97, respectively. Apparently, good connections were established between the identified compounds and biological activities. For example, p-cymene (compound 16) made the main contribution to DPPH scavenging ability and this fact was confirmed by several researchers in previous studies [47,48]. As another example, β-elemene (compound 28) was very close to the ability of phosphomolybdenum (PBD) and this compound has already been described as an agent for antioxidant therapy [49]. In metal chelating ability (MCA), α-copaene (compound 20) was a major contributor. β-Pinene was very close in cupric reducing ability. Principal component analysis (PCA) was also performed to determine similarities or differences between the tested Prangos species based on their biological activities. The tested species were different and two components (PC1: 53% and PC2: 33.9%) were obtained in PCA (Figure 2). PC1 was mainly contributed by ABTS, BChE, FRAP and CUPRAC, while PBD, tyrosinase, DPPH and amylase were the main players in PC2. The obtained results could be useful for further application by using the tested species in future studies.

2.4. Molecular Docking

Molecular modelling has been used for the predicting the ligand–target affinity as well as interaction [50,51]. In this study, molecular docking investigation of 12 components (selected from five or the five most abundant components of each EO) from the three EOs of PH (P. heyniae), PM (P. meliocarpoides var. meliocarpoides) and PU (P. uechtritzii) against tyrosinase and amylase were investigated. The reason for selecting these two targeted enzymes was that these three EOs have demonstrated better inhibitory activity against these two enzymes. These EO components were selected to decipher the inhibition pattern against tyrosinase and amylase enzyme. Detailed docking scores of all selected compounds against tyrosinase and amylase are shown in Table 5 and details of the binding interactions of best docked pose of α-bisabolol and elemol against tyrosinase and amylase are reported in Table 6. Analysis of docking score revealed elemol and α-bisabolol to be the most effective in binding with tyrosinase and amylase based on their ChemGauss scores of −8.10 and −8.94, respectively. Detailed 3-D binding interactions of α-bisabolol against tyrosinase and elemol against amylase are depicted in Figure 3.

3. Materials and Methods

3.1. Plant Materials

Prangos species were collected in the city of Konya. The location information is given in Table 7 below. The plants’ identity were confirmed by one of co-authors (Evren Yıldıztugay) and voucher specimens have been deposited in Selcuk University. The aerial parts of the plant samples were dried under shade conditions for 10 days at room temperature. The plant samples were then powdered using a laboratory mill and the powdered plant samples were stored in the dark at room temperature.

3.2. Essential Oil Extraction and GC-MS Analysis

The essential oil was obtained by using the hydro-distillation technique. One hundred g dried plant samples were distilled in a Clevenger-type apparatus for 5 h. The essential oil was dried over sodium sulphate (anhydrous) and then the obtained essential oils were stored in an amber vial at +4 °C until analysis.
The obtained essential oil was characterized by gas chromatography-flame ionization detector (GC-FID) and gas chromatography-mass spectrophotometry (GC-MS) techniques. GC-MS analysis was performed by using a 5975 GC-MS system (Agilent, city, state abbreviation if USA, country) coupled to an Agilent 7890 A GC. To separate chemical components, a HP-Innowax column (60 m × 0.25 mm, 0.25 μm film thickness) was used. Other analytical parameters were reported in our earlier paper [52]. All analytical details are given in the Supplementary Materials.
The retention index (RI) calculated by co-injection with reference to a homologous series of n-alkanes (C8-C30) under identical experimental circumstances was used to identify the components. By comparing their mass spectra to those from the NIST 05 and Wiley 8th edition libraries, as well as comparing their RIs to literature values, we were able to make more accurate identifications.

3.3. Determination of Antioxidant and Enzyme Inhibitory Effects

The antioxidant activity of the essential oils tested in this study was determined using a variety of assays [53]. The assays used were 1,1-diphenyl-2-picrylhydrazyl (DPPH) and 2,2′-azino-bis(3-ethylbenzothiazoline)-6-sulfonic acid (ABTS) radical scavenging capacity (CUPRAC), ferric ion reducing antioxidant power (FRAP), metal chelating ability (MCA), and phosphomolybdenum assay (PDA). The data for the DPPH, ABTS, CUPRAC, and FRAP assays were expressed in mg Trolox equivalents (TE)/g essential oils whereas the data for MCA and PDA were expressed in mg EDTA equivalents (EDTAE)/g essential oils and mmol TE/g essential oils, respectively. Previously, we provided the experimental components for the acetylcholinesterase, butyrylcholinesterase, tyrosinase, α-amylase, and α-glucosidase assays. In cholinesterase assays, galanthamine was used as a positive control, and data were expressed as mg galanthamine equivalents (GALAE)/g essential oils. In the tyrosinase inhibitory assay, kojic acid was used as a standard inhibitor, and the results were expressed as mg kojic acid equivalents (KAE)/g essential oils [53,54]. In the anti-diabetic assays, acarbose was chosen as an inhibitor of both amylase and glucosidase, and the results are expressed as mmol acarbose equivalents (ACAE)/g essential oils. The experimental procedures of the assays are given in Supplementary Materials. The assays were performed in triplicate, and ANOVA (Tukey’s test) was used to determine the differences in the essential oils (p < 0.05).

3.4. Multivariate Analysis

A partial least squared (PLS) regression was performed for the relation between bioactive components and biological activities of the tested essential oils. A principal component analysis (PCA) was also performed to detect differences between the tested species. The multivariate analysis was done with the SIMCA 14.0 Software (Umetrics, Umeå, Sweden).

3.5. Molecular Docking

A molecular docking investigation of the 12 most abundant components from the three essential oils of Prangos heyniae, Prangos meliocarpoides var. meliocarpoides and Prangos uechtritzii against tyrosinase and amylase were carried out. A list of these compounds is summarized in Table 1.
The 2D structures of these selected compounds were retrieved from PubChem and then imported into Discovery studio client for making the 3D structures. OMEGA tool of OpenEye software was used to generate molecular structures of these compounds with their optimized stereoisomers, ring conformations, tautomers, and ionization states to make broad structural and chemical diversity from a single input compound. The default settings of OMEGA were used for the generation of multiconformers of each compound, as the generation of conformer is a prerequisite for subsequent molecular docking.
Three-dimensional X-rays crystallographic structures of the target enzymes tyrosinase and amylase were retrieved from the Enzyme Data Bank utilizing the PDB ID: 2Y9X and 4W93, respectively [55]. To prepare the enzyme structures, the Discovery Studio Client software was utilized, which eliminates the heteroatoms and water molecules, inserts hydrogens and missing residues and assigns charges. The binding site of each enzyme was located based on the co-crystalized ligand within each targeted enzyme. Prior to docking of compounds, optimization of docking calculations was performed by re-docking the co-crystal ligand within the active site of the respective enzyme. Binding site coordinates was obtained using a co-crystal ligand in the binding site of tyrosinase and amylase. In case of tyrosinase, active site was adjusted upon the coordinate X = −10.021, Y = −28.82, and Z = −43.596111 in XYZ dimensions. While in case of amylase, coordinate was selected using X = −9.63, Y = 4.34 and Z = −23.10 dimensions. In both cases of redocking co-crystal ligands within the target protein, the RMSD was found to be less than 2 Å which shows the reliability of the docking protocol [50].
After optimization of the docking protocol, molecular docking of all selected compounds (as listed in Table 1) were performed using the FRED tool in OEDOCKING of the OPENEYE Software. For each compound, ten poses were generated and sorted out based on the corresponding ChemGauss 4 score. The best-docked pose was selected based on the lowest ChemGauss 4 score for deciphering the binding interactions between the compound and the amino acid residues of the targeted enzyme. Discovery Studio Visualizer was used for the visualization of the binding interactions of the best compounds with the amino acid residues of tyrosinase and amylase.

4. Conclusions

In this study, the chemical profile of the EOs obtained from three endemic Prangos species from Turkey was revealed and analyzed. Moreover, the in vitro antioxidant and enzyme inhibitory properties of the EOs were investigated in an attempt to assess the biological potentials of the plant-derived products of these Prangos species. While all the EOs showed moderate to good antioxidant potentials, as evidenced by various assays, their activity as enzyme inhibitors was not as prominent. While all EOs demonstrated anti-tyrosinase and anti-amylase activity, only P. heyniae EO showed anti-BChE activity. Overall, the findings from this study unveiled the varying biological potentials of the studied EOs and some interesting prospects of the investigated EO components as enzyme inhibitors. Particularly, their antioxidant and anti-tyrosinase potentials offer suggestions for their use in the development of topical anti-hyperpigmentation formulations. However, further research is warranted to assess the potency and safety of these EOs and their active components as cosmeceutical agents in vivo.

Supplementary Materials

The following supporting information can be downloaded online.

Author Contributions

Conceptualization, G.Z., M.F.M. and E.Y.; methodology, G.Z., E.Y. and S.U.K.; software, S.U.K.; validation, G.Z., S.D., A.M. and D.M.; formal analysis, G.Z. and D.M.; investigation, G.Z., A.B. and D.M.; resources, G.Z. and E.Y.; data curation, G.Z., A.B. and S.J.; writing—original draft preparation, G.Z., M.F.M. and S.J.; writing—review and editing, G.Z., M.F.M., S.J. and S.U.K.; visualization, S.U.K.; supervision, S.D., A.M., A.B. and D.M.; project administration, G.Z.; funding acquisition, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Djilani, A.; Dicko, A. The therapeutic benefits of essential oils. Nutr. Well Being Health 2012, 7, 155–179. [Google Scholar]
  2. 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]
  3. Fiocco, D.; Arciuli, M.; Arena, M.P.; Benvenuti, S.; Gallone, A. Chemical composition and the anti-melanogenic potential of different essential oils. Flavour Fragr. J. 2016, 31, 255–261. [Google Scholar] [CrossRef]
  4. Yang, S.-K.; Tan, N.-P.; Chong, C.-W.; Abushelaibi, A.; Lim, S.-H.-E.; Lai, K.-S. The missing piece: Recent approaches investigating the antimicrobial mode of action of assential oils. Evol. Bioinf. 2021, 17, 1176934320938391. [Google Scholar] [CrossRef] [PubMed]
  5. De Lavor, É.M.; Fernandes, A.W.C.; de Andrade Teles, R.B.; Leal, A.E.B.P.; de Oliveira Júnior, R.G.; Gama e Silva, M.; De Oliveira, A.P.; Silva, J.C.; de Moura Fontes Araujo, M.T.; Coutinho, H.D.M. Essential oils and their major compounds in the treatment of chronic inflammation: A review of antioxidant potential in preclinical studies and molecular mechanisms. Oxid. Med. Cell. Longev. 2018, 2018, 6468593. [Google Scholar] [CrossRef] [Green Version]
  6. Saad, N.Y.; Muller, C.D.; Lobstein, A. Major bioactivities and mechanism of action of essential oils and their components. Flavour Fragr. J. 2013, 28, 269–279. [Google Scholar] [CrossRef]
  7. Wani, A.R.; Yadav, K.; Khursheed, A.; Rather, M.A. An updated and comprehensive review of the antiviral potential of essential oils and their chemical constituents with special focus on their mechanism of action against various influenza and coronaviruses. Microb. Path. 2021, 152, 104620. [Google Scholar] [CrossRef]
  8. Swamy, M.K.; Akhtar, M.S.; Sinniah, U.R. Antimicrobial properties of plant essential oils against human pathogens and their mode of action: An updated review. Evid.-Based Complement. Altern. 2016, 2016, 3012462. [Google Scholar] [CrossRef]
  9. Mottaghipisheh, J.; Kiss, T.; Tóth, B.; Csupor, D. The Prangos genus: A comprehensive review on traditional use, phytochemistry, and pharmacological activities. Phytochem. Rev. 2020, 19, 1449–1470. [Google Scholar] [CrossRef]
  10. Baser, K.H.C. Essential oils of Anatolian Apiaceae—A profile. Nat. Volatiles Essent. Oils 2014, 1, 1–50. [Google Scholar]
  11. Ulubelen, A.; Topcu, G.; Tan, N.; Ölçal, S.; Johansson, C.; Üçer, M.; Birman, H.; Tamer, Ş. Biological activities of a Turkish medicinal plant Prangos Platychlaena. J. Ethnopharmacol. 1995, 45, 193–197. [Google Scholar] [CrossRef]
  12. Albayrak, G.; Demir, S.; Kose, F.A.; Baykan, S. New coumarin glycosides from endemic Prangos heyniae H. Duman & MF Watson. Nat. Prod. Res. 2021, 13, 1–13. [Google Scholar]
  13. Numonov, S.; Bobakulov, K.; Numonova, M.; Sharopov, F.; Setzer, W.N.; Khalilov, Q.; Begmatov, N.; Habasi, M.; Aisa, H.A. New coumarin from the roots of Prangos pabularia. Nat. Prod. Res. 2018, 32, 2325–2332. [Google Scholar] [CrossRef] [PubMed]
  14. Sevin, G.; Alan, E.; Demir, S.; Albayrak, G.; Demiroz, T.; Yetik-Anacak, G.; Baykan, S. Comparative evaluation of relaxant effects of three Prangos species on mouse corpus cavernosum: Chemical characterization and the relaxant mechanisms of action of P. pabularia and (+)-oxypeucedanin. J. Ethnopharmacol. 2022, 284, 114823. [Google Scholar] [CrossRef] [PubMed]
  15. Karahisar, E.; Kose, Y.B.; Iscan, G.; Kurkcuoglu, M.; Tugay, O. Chemical composition and anticandidal activity of essential oils obtained from different parts of Prangos heyniae H. Duman & M. F. Watson. Rec. Nat. Prod. 2022, 16, 74–83. [Google Scholar]
  16. Mohebi, Z.; Sefidkon, F.; Heshmati, G.A. Investigation of increasing essential oil and maintaining forage quality and digestibility of the Prangos ferulacea using an optimizer apparatus design. J. Med. Plants By-Prod. 2021, 10, 227–235. [Google Scholar]
  17. Bazdar, M.; Sadeghi, H.; Hosseini, S. Evaluation of oil profiles, total phenols and phenolic compounds in Prangos ferulacea leaves and flowers and their effects on antioxidant activities. Biocatal. Agric. Biotechnol. 2018, 14, 418–423. [Google Scholar] [CrossRef]
  18. Bruno, M.; Ilardi, V.; Lupidi, G.; Quassinti, L.; Bramucci, M.; Fiorini, D.; Venditti, A.; Maggi, F. The nonvolatile and volatile metabolites of Prangos ferulacea and their biological properties. Plant. Med. 2019, 85, 815–824. [Google Scholar] [CrossRef] [Green Version]
  19. Delnavazi, M.R.; Soleimani, M.; Hadjiakhoondi, A.; Yassa, N. Isolation of phenolic derivatives and essential oil analysis of Prangos ferulacea (L.) Lindl. aerial parts. Iran. J. Pharm. Res. 2017, 16, 207–215. [Google Scholar]
  20. Davis, P.H. Flora of Turkey and the East Aegean Islands. In Flora of Turkey and the East Aegean Islands; Edinburgh University Press: Edinburgh, UK, 1970; Volume 4. [Google Scholar]
  21. Ahmed, J.; Güvenç, A.; Küçükboyaci, N.; Baldemir, A.; Coşkun, M. Total phenolic contents and antioxidant activities of Prangos Lindl.(Umbelliferae) species growing in Konya province (Turkey). Turk. J. Biol. 2011, 35, 353–360. [Google Scholar]
  22. Uzel, A.; Dirmenci, T.; Çelik, A.; Arabaci, T. Composition and antimicrobial activity of Prangos platychlaena and P. uechtritzii. Chem. Nat. Comp. 2006, 42, 169–171. [Google Scholar] [CrossRef]
  23. Oke Altuntas, F.; Aslim, B.; Duman, H. The anti-lipid peroxidative, metal chelating, and radical scavenging properties of the fruit extracts from endemic Prangos meliocarpoides Boiss var. meliocarpoides. Gazi Univ. J. Sci. 2016, 29, 537–542. [Google Scholar]
  24. Duman, H.; Watson, M. Ekimia, a new genus of Umhelliferae, and two new taxa of Prangos Lindl.(Umbelliferae) from southern Turkey. Edin. J. Bot. 1999, 56, 199–209. [Google Scholar] [CrossRef]
  25. Kafash-Farkhad, N.; Asadi-Samani, M.; Rafieian-Kopaei, M. A review on phytochemistry and pharmacological effects of Prangos ferulacea (L.) Lindl. Life Sci. J. 2013, 10, 360–367. [Google Scholar]
  26. Tan, N.; Bilgin, M.; Tan, E.; Miski, M. Antibacterial activities of pyrenylated coumarins from the roots of Prangos hulusii. Molecules 2017, 22, 1098. [Google Scholar] [CrossRef] [PubMed]
  27. AzarKish, P.; Moghadam, M.; Khakdan, F.; Ghasemi Pirbaloti, A. Variability in morphological traits, total phenolic contents and antioxidant activity in different populations from three species of Prangos from Fars, Kohklouye and Boyerahmad provinces. Eco-Phytochem. J. Med. Plants 2018, 6, 1–20. [Google Scholar]
  28. Baer, K.H.; Demirci, B.; Demirci, F.; Bedir, E.; Weyerstahl, P.; Marschall, H.; Duman, H.; Aytaç, Z.; Hamann, M.T. A new bisabolene derivative from the essential oil of Prangos uechtritzii fruits. Plant. Med. 2000, 66, 674–677. [Google Scholar]
  29. Özek, G.; Bedir, E.; Tabanca, N.; Ali, A.; Khan, I.A.; Duran, A.; Başer, K.H.; Özek, T. Isolation of eudesmane type sesquiterpene ketone from Prangos heyniae H. Duman & MF Watson essential oil and mosquitocidal activity of the essential oils. Open Chem. 2018, 16, 453–467. [Google Scholar]
  30. Dhifi, W.; Bellili, S.; Jazi, S.; Bahloul, N.; Mnif, W. Essential oils’ chemical characterization and investigation of some biological activities: A critical review. Medicines 2016, 3, 25. [Google Scholar] [CrossRef] [Green Version]
  31. Kiliç, Ö.; Özdemir, F.A. Essential oil composition of two Prangos Lindl.(Apiaceae) species from Turkey. Prog. Nutr. 2017, 19, 69–74. [Google Scholar]
  32. Numonov, S.; Sharopov, F.S.; Atolikhshoeva, S.; Safomuddin, A.; Bakri, M.; Setzer, W.N.; Musoev, A.; Sharofova, M.; Habasi, M.; Aisa, H.A. Volatile secondary metabolites with potent antidiabetic activity from the roots of Prangos pabularia Lindl.—Computational and experimental investigations. Appl. Sci. 2019, 9, 2362. [Google Scholar] [CrossRef] [Green Version]
  33. Kasote, D.M.; Katyare, S.S.; Hegde, M.V.; Bae, H. Significance of antioxidant potential of plants and its relevance to therapeutic applications. Int. J. Biol. Sci. 2015, 11, 982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Barreca, D. Mechanisms of Plant Antioxidants Action; Multidisciplinary Digital Publishing Institute: Basel, Switzerland, 2021; Volume 10, p. 35. [Google Scholar]
  35. Amorati, R.; Foti, M.C.; Valgimigli, L. Antioxidant activity of essential oils. J. Agric. Food Chem. 2013, 61, 10835–10847. [Google Scholar] [CrossRef] [PubMed]
  36. Kusumawati, I.; Indrayanto, G. Natural antioxidants in cosmetics. In Studies in Natural Products Chemistry; Elsevier: Amsterdam, The Netherlands, 2013; Volume 40, pp. 485–505. [Google Scholar]
  37. Nirmala, C.; Bisht, M.S.; Bajwa, H.K.; Santosh, O. Bamboo: A rich source of natural antioxidants and its applications in the food and pharmaceutical industry. Trends Food Sci. Technol. 2018, 77, 91–99. [Google Scholar] [CrossRef]
  38. Diniz do Nascimento, L.; Moraes, A.A.B.D.; Costa, K.S.D.; Pereira Galúcio, J.M.; Taube, P.S.; Costa, C.M.L.; Cruz, J.N.; Andrade, E.H.d.A.; Faria, L.J.G.d. Bioactive natural compounds and antioxidant activity of essential oils from spice plants: New findings and potential applications. Biomolecules 2020, 10, 988. [Google Scholar] [CrossRef]
  39. Cardoso-Ugarte, G.A.; Sosa-Morales, M.E. Essential oils from herbs and spices as natural antioxidants: Diversity of promising food applications in the past decade. Food Rev. Int. 2021, 1–31. [Google Scholar] [CrossRef]
  40. Zengin, G.; Sinan, K.I.; Ak, G.; Mahomoodally, M.F.; Paksoy, M.Y.; Picot-Allain, C.; Glamocilja, J.; Sokovic, M.; Jekő, J.; Cziáky, Z. Chemical profile, antioxidant, antimicrobial, enzyme inhibitory, and cytotoxicity of seven Apiaceae species from Turkey: A comparative study. Ind. Crop. Prod. 2020, 153, 112572. [Google Scholar] [CrossRef]
  41. Bahadori, M.B.; Zengin, G.; Bahadori, S.; Maggi, F.; Dinparast, L. Chemical composition of essential oil, antioxidant, antidiabetic, anti-obesity, and neuroprotective properties of Prangos gaubae. Nat. Prod. Commun. 2017, 12, 1934578X1701201233. [Google Scholar] [CrossRef] [Green Version]
  42. Ata, A. Novel plant-derived biomedical agents and their biosynthetic origin. In Studies in Natural Products Chemistry; Elsevier: Amsterdam, The Netherlands, 2012; Volume 38, pp. 225–245. [Google Scholar]
  43. Li, S.; Li, A.J.; Travers, J.; Xu, T.; Sakamuru, S.; Klumpp-Thomas, C.; Huang, R.; Xia, M. Identification of compounds for butyrylcholinesterase inhibition. Adv. Sci. Drug Discov. 2021, 26, 1355–1364. [Google Scholar] [CrossRef]
  44. Pohanka, M. Cholinesterases, a target of pharmacology and toxicology. Biomed. Pap. Med. Fac. Palacky Univ. Olomouc 2011, 155, 219–229. [Google Scholar] [CrossRef] [Green Version]
  45. Tundis, R.; Loizzo, M.; Menichini, F. Natural products as α-amylase and α-glucosidase inhibitors and their hypoglycaemic potential in the treatment of diabetes: An update. Mini Rev. Med. Chem. 2010, 10, 315–331. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, Y.-Q.; Xu, C.-Y.; Liang, F.-Y.; Jin, P.-C.; Qian, Z.-Y.; Luo, Z.-S.; Qin, R.-G. Selecting and characterizing tyrosinase inhibitors from Atractylodis macrocephalae rhizoma based on spectrum-activity relationship and molecular docking. J. Anal. Method. Chem. 2021, 2021, 5596463. [Google Scholar] [CrossRef] [PubMed]
  47. De Oliveira, T.M.; de Carvalho, R.B.; da Costa, I.H.; de Oliveira, G.A.; de Souza, A.A.; de Lima, S.G.; de Freitas, R.M. Evaluation of p-cymene, a natural antioxidant. Pharm. Biol. 2015, 53, 423–428. [Google Scholar] [CrossRef] [PubMed]
  48. Quintans-Júnior, L.; Moreira, J.C.F.; Pasquali, M.A.B.; Rabie, S.M.S.; Pires, A.S.; Schröder, R.; Rabelo, T.K.; Santos, J.P.A.; Lima, P.S.S.; Cavalcanti, S.C.H.; et al. Antinociceptive activity and redox profile of the monoterpenes (+)-camphene, p -cymene, and geranyl acetate in experimental models. ISRN Toxicol. 2013, 2013, 459530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ahmad, K.A.; Ze, H.; Chen, J.; Khan, F.U.; Xuezhuo, C.; Xu, J.; Qilong, D. The protective effects of a novel synthetic β-elemene derivative on human umbilical vein endothelial cells against oxidative stress-induced injury: Involvement of antioxidation and PI3k/Akt/eNOS/NO signaling pathways. Biomed. Pharmacother. 2018, 106, 1734–1741. [Google Scholar] [CrossRef]
  50. Khan, S.U.; Ahemad, N.; Chuah, L.-H.; Naidu, R.; Htar, T.T. Illustrated step by step protocol to perform molecular docking: Human estrogen receptor complex with 4-hydroxytamoxifen as a case study. Prog. Drug Discov. Biomed. Sci. 2020, 3, 1–21. [Google Scholar] [CrossRef]
  51. Khan, S.U.; Ahemad, N.; Chuah, L.-H.; Naidu, R.; Htar, T.T. Sequential ligand-and structure-based virtual screening approach for the identification of potential G protein-coupled estrogen receptor-1 (GPER-1) modulators. RSC Adv. 2019, 9, 2525–2538. [Google Scholar] [CrossRef] [Green Version]
  52. Ak, G.; Zengin, G.; Sinan, K.I.; Mahomoodally, M.F.; Picot-Allain, M.C.N.; Cakır, O.; Bensari, S.; Yılmaz, M.A.; Gallo, M.; Montesano, D. A comparative bio-evaluation and chemical profiles of Calendula officinalis L. extracts prepared via different extraction techniques. Appl. Sci. 2020, 10, 5920. [Google Scholar] [CrossRef]
  53. Grochowski, D.M.; Uysal, S.; Aktumsek, A.; Granica, S.; Zengin, G.; Ceylan, R.; Locatelli, M.; Tomczyk, M. In vitro enzyme inhibitory properties, antioxidant activities, and phytochemical profile of Potentilla thuringiaca. Phytochem. Lett. 2017, 20, 365–372. [Google Scholar] [CrossRef]
  54. Uysal, S.; Zengin, G.; Locatelli, M.; Bahadori, M.B.; Mocan, A.; Bellagamba, G.; De Luca, E.; Mollica, A.; Aktumsek, A. Cytotoxic and enzyme inhibitory potential of two Potentilla species (P. speciosa L. and P. reptans Willd.) and their chemical composition. Front. Pharmacol. 2017, 8, 290. [Google Scholar] [CrossRef]
  55. Mamadalieva, N.Z.; Youssef, F.S.; Hussain, H.; Zengin, G.; Mollica, A.; Al Musayeib, N.M.; Ashour, M.L.; Westermann, B.; Wessjohann, L.A. Validation of the antioxidant and enzyme inhibitory potential of selected triterpenes using in vitro and in silico studies, and the evaluation of their ADMET properties. Molecules 2021, 26, 6331. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The biplot obtained from partial least squared (PLS) regression describing relationship between chemical compounds and bioactivities. For compounds numbers refer to Table 1.
Figure 1. The biplot obtained from partial least squared (PLS) regression describing relationship between chemical compounds and bioactivities. For compounds numbers refer to Table 1.
Molecules 27 01676 g001
Figure 2. The principal component analysis on the biological activities of the EOs of Prangros species. (A). Eigenvalue and percentage of explained variances. (B). Score plot of dim1 and dim2 scores. (C). Corresponding bar plot representing influential bioactivities.
Figure 2. The principal component analysis on the biological activities of the EOs of Prangros species. (A). Eigenvalue and percentage of explained variances. (B). Score plot of dim1 and dim2 scores. (C). Corresponding bar plot representing influential bioactivities.
Molecules 27 01676 g002
Figure 3. Three-dimensional binding interaction of tyrosinase (Blue stick) and α-bisabolol (golden stick). Amylase (Cyan stick model) and elemol (pink stick model). Hydrogen-bonding interactions are shown in green dash lines, while hydrophobic interactions are indicated in light pink dashed lines.
Figure 3. Three-dimensional binding interaction of tyrosinase (Blue stick) and α-bisabolol (golden stick). Amylase (Cyan stick model) and elemol (pink stick model). Hydrogen-bonding interactions are shown in green dash lines, while hydrophobic interactions are indicated in light pink dashed lines.
Molecules 27 01676 g003
Table 1. Chemical composition of the tested Prangos essential oils.
Table 1. Chemical composition of the tested Prangos essential oils.
No.CompoundsRRI aPH (%)PM (%)PU (%)
1α-Pinene10231.66.20.4
2α-Thujene1026-0.3-
3Camphene10680.31.4-
4Hexanal1086-0.1-
5β-Pinene11110.11.00.6
6Sabinene11240.116.70.8
7δ-3-Carene1157-0.4-
8Myrcene11650.10.70.1
9Heptanal11890.1--
10Dehydro 1,8-cineole11901.5--
11Limonene12010.73.73.2
12β-Phellandrene1210--0.8
131,8-Cineole12110.10.1-
142-Pentylfuran12340.20.1
156-Methyl, 2-heptanone12390.1--
16p-Cymene12760.213.224.6
17α, p-dimethylstyrene1447-0.2-
18α-Cubebene14650.2--
19trans-Sabinene hydrate1469-1.0-
20α-Copaene15011.40.42.0
21β-Bourbonene15311.1--
22Camphor15350.21.5-
23β-Cubebene15490.7--
24cis-Sabinene hydrate1554-0.7-
25trans-Chrysanthenyl acetate1581-0.4-
26Pinocarvone1588-0.4-
27Bornyl acetate15930.511.8-
28β-Elemene16012.95.5-
29Terpinen-4-ol1612-3.1-
30β-Caryophyllene16143.8-0.8
31γ-Elemene16504.1--
32Myrtenal1651-0.4-
33Sabina ketone1655-0.9-
34trans-Pinocarveol1670-0.9-
35α-Humulene16896.7--
36trans-Verbenol1690-4.2-
37Cryptone1695--2.6
38γ-Muurolene17040.80.91.4
39Germacrene D17297.8--
407-epi-1,2-Dehydrosesquicineole1730--12.6
41Verbonene1732 1.4-
42β-Bisabolone17385.70.1-
43Valencene1740--0.5
44β-Selinene1743-0.90.4
45Phellandral1745--0.3
46α-selinene1747-0.2-
47Bicyclogermacrene17540.3--
48Carvone1757 1.0-
49δ-Cadinene17732.0-0.7
50γ-Cadinene17790.7-0.1
51Kessane1785-2.5-
52ar-Curcumene1787--0.5
53p-Methylacetophenone1800--0.3
54Cumin aldehyde1807-1.0-
55trans-Carveol1846-0.9-
56Germacrene B18563.3--
57p-Cymen-8-ol1861-6.10.7
58α-Calacorene1943--0.3
591,5-Epoxysalvial-4(14)-ene19471.8--
604-Hydroxy-2-methylacetophenone1950-2.815.1
61Isocaryophyllene oxide2002--1.7
62Caryophyllene oxide20177.93.519.6
63Salvial-4(14)-en 1-one20431.3--
64Humulene epoxide II20744.0--
65Elemol20957.4--
66p-Cresol2101--0.5
67Cumin alchol2121-1.30.2
68Spathulenol21473.61.61.7
69γ-Eudesmol21872.7--
70T-Cadinol2193--0.3
71T-Muurolol22082.3--
72α-Bisabolol2232--3.2
73α-Eudesmol22461.0--
74α-Cadinol22544.0-1.4
75β-Eudesmol22560.4--
76β-Bisabolenal237712.2--
Total identified (%) 95.999.597.4
a Relative retention indices calculated against n-alkanes. PH: Prangos heyniae; PM: Prangos meliocarpoides var. meliocarpoides. PU: Prangos uechtritzii.
Table 2. List if top most abundant selected compounds obtained from chemical profile from the three essential oils.
Table 2. List if top most abundant selected compounds obtained from chemical profile from the three essential oils.
No.CompoundsRRI aPH (%)PM (%)PU (%)
1α-Pinene10231.66.20.4
6Sabinene11240.116.70.8
11Limonene12010.73.73.2
16p-Cymene12760.213.224.6
27Bornyl acetate15930.511.8-
35α-Humulene16896.7--
39Germacrene D17297.8--
407-epi-1,2-Dehydrosesquicineole1730--12.6
57p-Cymen-8-ol1861-6.10.7
62Caryophyllene oxide20177.93.519.6
65Elemol20957.4--
72α-Bisabolol2232--3.2
76β-Bisabolenal237712.2--
a Relative retention indices calculated against n-alkanes.
Table 3. Antioxidant Properties of the Tested Essential Oils.
Table 3. Antioxidant Properties of the Tested Essential Oils.
Essentail OilsDPPH (mg TE/g)ABTS (mg TE/g)CUPRAC (mg TE/g)FRAP (mg TE/g)MCA (mg EDTAE/g)PBD (mmol TE/g)
P. heyniae0.43 ± 0.01 c92.99 ± 1.29 a103.15 ± 3.69 b61.20 ± 0.73 a30.00 ± 5.82 a20.33 ± 0.48 b
P. meliocarpoides var. meliocarpoides1.01 ± 0.06 b24.18 ± 1.10 c113.43 ± 3.37 a47.98 ± 0.89 c28.66 ± 0.46 a24.37 ± 1.23 a
P. uechtritzii1.74 ± 0.10 a58.17 ± 1.46 b109.14 ± 1.00 a,b56.49 ± 0.64 b30.94 ± 0.20 a15.64 ± 0.28 c
Values are reported as mean ± SD. TE: Trolox equivalent; EDTAE: EDTA equivalent; MCA: Metal chelating ability; PBD: Phosphomolybdenum assay. Different superscripts indicate significant differences in the tested essential oils (p < 0.05).
Table 4. Enzyme Inhibitory Effects of the Tested Essential Oils.
Table 4. Enzyme Inhibitory Effects of the Tested Essential Oils.
Essential OilAChE
(mg GALAE/g)
BChE
(mg GALAE/g)
Tyrosinase (mg KAE/g)Amylase
(mmol ACAE/g)
Glucosidase (mmol ACAE/g)
P. heyniaena9.85 ± 0.2053.91 ± 2.11 b0.09 ± 0.01 cna
P. meliocarpoides var. meliocarpoidesnana69.56 ± 4.80 a0.41 ± 0.01 bna
P. uechtritziinana46.34 ± 6.51 b0.61 ± 0.01 ana
Values are reported as mean ± SD. GALAE: Galantamine equivalent; KAE: Kojic acid equivalent; ACAE: Acarbose equivalent; na: Not active. Different superscripts indicate significant differences in the tested essential oils (p < 0.05).
Table 5. Detailed binding score of the topmost compounds in three essential oils based on ChemGauss scores.
Table 5. Detailed binding score of the topmost compounds in three essential oils based on ChemGauss scores.
Compound NameBinding Affinity Based on
ChemGauss 4 Scores
TyrosinaseAmylase
7-epi-1,2-Dehydrosesquicineole−7.29−7.38
Bornyl acetate−6.95−6.67
Caryophyllene oxide−8.36−7.45
Elemol−7.33−8.10
Germacrene D−8.41−7.23
Limonene−8.68−5.73
Sabinene−7.64−5.71
Pinene−6.65−5.39
p-Cymene−7.79−5.56
p-Cymen-8-ol−7.18−5.78
α-Bisabolol−8.94−8.03
α-Humulene−7.78−7.19
Reference (kojic acid)−7.58-
Reference (ascorbic acid)-−8.67
ChemGauss4 for topmost compound against each selected enzyme is shown in bold.
Table 6. Detailed binding Interaction of best docked pose of α-bisabolol and elemol against tyrosinase and amylase, respectively.
Table 6. Detailed binding Interaction of best docked pose of α-bisabolol and elemol against tyrosinase and amylase, respectively.
Interacting Amino Acid Residue of Tyrosinase and α-BisabololDistance between Interacting ResidueType of Bond
A:VAL2834.7077Alkyl
A:ALA2865.3675Alkyl
A:ALA2864.4797Alkyl
A:VAL2834.2341Alkyl
A:HIS614.7969Pi-Alkyl
A:HIS854.9317Pi-Alkyl
A:HIS855.0019Pi-Alkyl
A:HIS2444.3551Pi-Alkyl
A:HIS2444.6949Pi-Alkyl
A:HIS2595.2737Pi-Alkyl
A:HIS2633.8345Pi-Alkyl
A:HIS2633.7349Pi-Alkyl
A:0TR4104.3467Pi-Alkyl
Interacting Amino Acid residue of Amylase and ElemolDistance between Interacting ResidueType of Bond
A:GLN63:NE22.9536Conventional Hydrogen Bond
A:TRP59:O2.123Pi-Alkyl
A:TRP585.4757Pi-Alkyl
A:TRP584.4585Pi-Alkyl
A:TRP585.2415Pi-Alkyl
A:TRP594.7261Pi-Alkyl
A:TYR623.4596Pi-Alkyl
A:HIS2995.2942Pi-Alkyl
A:HIS3054.5785Pi-Alkyl
Table 7. Locations and voucher numbers of the tested Prangos species.
Table 7. Locations and voucher numbers of the tested Prangos species.
Prangos SpeciesLocationsVoucher Numbers
P. meliocarpoides Boiss. var. meliocarpoidesYavşan Location (Tuzgölü), Konya/Turkey, 905 mEY-2998
P. uechtritzii Boiss & HaussknBetween Hadim and Taşkent (2 km), Konya/Turkey, 1490 mEY-3023
P. heyniae H. Duman & M.F. WatsonBetween Hadim—Bozkır, Korualan location, Konya/Turkey, 1545 mEY-3039
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zengin, G.; Mahomoodally, M.F.; Yıldıztugay, E.; Jugreet, S.; Khan, S.U.; Dall’Acqua, S.; Mollica, A.; Bouyahya, A.; Montesano, D. Chemical Composition, Biological Activities and In Silico Analysis of Essential Oils of Three Endemic Prangos Species from Turkey. Molecules 2022, 27, 1676. https://doi.org/10.3390/molecules27051676

AMA Style

Zengin G, Mahomoodally MF, Yıldıztugay E, Jugreet S, Khan SU, Dall’Acqua S, Mollica A, Bouyahya A, Montesano D. Chemical Composition, Biological Activities and In Silico Analysis of Essential Oils of Three Endemic Prangos Species from Turkey. Molecules. 2022; 27(5):1676. https://doi.org/10.3390/molecules27051676

Chicago/Turabian Style

Zengin, Gokhan, Mohamad Fawzi Mahomoodally, Evren Yıldıztugay, Sharmeen Jugreet, Shafi Ullah Khan, Stefano Dall’Acqua, Adriano Mollica, Abdelhakim Bouyahya, and Domenico Montesano. 2022. "Chemical Composition, Biological Activities and In Silico Analysis of Essential Oils of Three Endemic Prangos Species from Turkey" Molecules 27, no. 5: 1676. https://doi.org/10.3390/molecules27051676

Article Metrics

Back to TopTop