Next Article in Journal
Effects of Exponential N Application on Soil Exchangeable Base Cations and the Growth and Nutrient Contents of Clonal Chinese Fir Seedlings
Next Article in Special Issue
Essential Oil Composition of Seven Bulgarian Hypericum Species and Its Potential as a Biopesticide
Previous Article in Journal
Spicy and Aromatic Plants
Previous Article in Special Issue
Current Trends for Lavender (Lavandula angustifolia Mill.) Crops and Products with Emphasis on Essential Oil Quality
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Essential Oil from the Leaves of Gynoxys rugulosa Muschl. (Asteraceae) Growing in Southern Ecuador: Chemical and Enantioselective Analyses

1
Departamento de Química, Universidad Técnica Particular de Loja (UTPL), Calle Marcelino Champagnat s/n, Loja 110107, Ecuador
2
Departamento de Ciencias Biológicas y Agropecuarias, Universidad Técnica Particular de Loja (UTPL), Calle Marcelino Champagnat s/n, Loja 110107, Ecuador
*
Author to whom correspondence should be addressed.
Plants 2023, 12(4), 849; https://doi.org/10.3390/plants12040849
Submission received: 19 January 2023 / Revised: 4 February 2023 / Accepted: 6 February 2023 / Published: 14 February 2023

Abstract

:
An essential oil, distilled from the leaves of the Andean species Gynoxys rugulosa Muschl., is described in the present study for the first time. The chemical composition was qualitatively and quantitatively determined by GC–MS and GC–FID, respectively. On the one hand, the qualitative composition was obtained by comparing the mass spectrum and the linear retention index of each component with data from literature. On the other hand, the quantitative composition was determined by calculating the relative response factor of each constituent, according to its combustion enthalpy. Both analyses were carried out with two orthogonal columns of nonpolar and polar stationary phases. A total of 112 compounds were detected and quantified with at least one column, corresponding to 87.3–93.0% of the whole oil mass. Among the 112 detected components, 103 were identified. The main constituents were α-pinene (5.3–6.0%), (E)-β-caryophyllene (2.4–2.8%), α-humulene (3.0–3.2%), germacrene D (4.9–6.5%), δ-cadinene (2.2–2.3%), caryophyllene oxide (1.6–2.2%), α-cadinol (3.8–4.4%), 1-nonadecanol (1.7–1.9%), 1-eicosanol (0.9–1.2%), n-tricosane (3.3–3.4%), 1-heneicosanol (4.5–5.8%), n-pentacosane (5.8–7.1%), 1-tricosanol (4.0–4.5%), and n-heptacosane (3.0–3.5%). Furthermore, an enantioselective analysis was carried out on the essential oil, by means of two cyclodextrin-based capillary columns. The enantiomers of α-pinene, β-pinene, sabinene, α-phellandrene, β-phellandrene, linalool, α-copaene, terpinen-4-ol, α-terpineol, and germacrene D were detected, and the respective enantiomeric excess was calculated.

1. Introduction

During the last 40 years, the phytochemical investigation has shifted from temperate climates to tropical countries, where most of the botanical species are still unstudied [1,2]. In this sense, a great importance is given to the so-called “megadiverse” countries, a group of 17 countries, including Ecuador, characterized by possessing three-fourths of all higher plant species of the world [3]. For this reason, our group has been investigating the phytochemistry of the Ecuadorian flora for more than 20 years, by describing the major metabolites of unprecedented botanical species [4,5,6,7,8,9]. Together with nonvolatile compounds, we are very interested in essential oils (EOs), defined by the European Pharmacopoeia as “odorous products, usually of complex composition, obtained from a botanically defined plant raw material by steam distillation, dry distillation, or a suitable mechanical process without heating” [10,11,12,13,14,15,16,17]. Our interest in the EOs derives from the commercial importance of these mixtures and, overall, from the fact that they can be sources of new or rare sesquiterpenoids, often biologically active, together with enantiomeric compounds. As discussed in a previous paper, we selected the poorly studied genus Gynoxys as a promising taxon for a systematic investigation. Despite the leaves not usually being very fragrant, a preliminary unpublished analysis indicated that the EOs from this genus are dominated by the sesquiterpene fraction [18]. For what concerns Gynoxys rugulosa Muschl., this species is poorly described also from the botanical point of view. In fact, on the one hand, it is not present in the Catalog of the Vascular Plants of Ecuador. On the other hand, the online database Tropicos only reports three specimens for this plant, from northern Peru and southern Ecuador, where the species grows at an altitude of 2500–3000 m above the sea level [19].
Botanically (see Figure 1), G. rugulosa is a shrub growing up to 2 m tall, with compressed-quadrangular branches and tomentosa, blackish brown in color. The leaves are opposite and petiolate, with an acute apex, rounded, or sometimes subcordate-rounded base, and yellowish tomentose underside, with pinnate venation. The plant presents sub corymbose, compound inflorescence at the apex of the terminal branches. This species only grows in shrubby Paramos, sharing the same ecosystem of typical families such as Melastomataceae, Asteraceae, Orchidaceae, and Ericaceae [20].
Since this plant is little known and quite rare, no traditional use exists to the best of the authors’ knowledge. From the legal point of view, probably due to the lack of botanical information, G. rugulosa is not a protected species, and it does not even appear in the reference publication for threatened taxa (The Red Book of the Endemic Plants in Ecuador). Therefore, the present study presents the first description of an EO distilled from Gynoxys rugulosa Muschl., together with the enantiomeric composition of some chiral terpenes.

2. Results

2.1. Chemical Analysis of the EO

The detailed amount of each component and fraction is represented in Table 1. Overall, with respect to the polar and nonpolar column, the monoterpene fraction ranged between 12.9% and 10.3% of the whole EO respectively, the sesquiterpene fraction ranged between 39.1% and 43.3%, and the other non-terpene compounds ranged between 35.3% and 39.4%. A total of 87.3–93.0% of the oil mass was quantified. The distillation yield of this EO, analytically calculated over four repetitions, was 0.02% ± 0.004% by weight of dry plant material.
According to its chromatographic profiles (Figure 2 and Figure 3), the EO from leaves of G. rugulosa was composed of three main groups of components: a poor monoterpene fraction, an important sesquiterpene fraction, and an abundant heavy fraction, characterized by long-chained alcohols and alkanes.
On the one hand, in the monoterpene fraction, α-pinene (peak 1) was the major compound, corresponding to about 5.3–6.0% by weight of the whole EO. On the other hand, the sesquiterpene fraction was dominated (according to the elution order) by (E)-β-caryophyllene (peak 44, 2.4–2.8%), α-humulene (peak 45, 3.0–3.2%), germacrene D (peak 47, 4.9–6.5%), δ-cadinene (peak 54, 2.2–2.3%), an unidentified compound of MW 220 (peak 56, 3.0–3.5%), caryophyllene oxide (peak 59, 1.6–2.2%), and α-cadinol (peak 68, 3.8–4.4%), altogether contributing for 20.9–24.9% of the EO mass. Lastly, the heavy fraction mainly constituted 1-nonadecanol (peak 97, 1.7–1.9%), 1-eicosanol (peak 99, 0.9–1.2%), n-tricosane (peak 100, 3.3–3.4%), 1-heneicosanol (peak 101, 4.5–5.8%), n-pentacosane (peak 104, 5.8–7.1%), 1-tricosanol (peak 105, 4.0–4.5%), and n-heptacosane (peak 108, 3.0–3.5%). All these heavy aliphatic metabolites, most likely biosynthetically proceeding from the acetate pathway, accounted for about 23.2–27.4% of the whole EO mass.

2.2. Enantioselective Analysis of the EO

For almost all the identified enantiomers, the enantioselective analysis was carried out through a 2,3-diacetyl-6-tert-butyldimethylsilyl-β-cyclodextrin capillary column, with the exception of α-copaene and germacrene D. For these compounds, a 2,3-diethyl-6-tert-butyldimethylsilyl-β-cyclodextrin column was used since their enantiomers are inseparable with the other chiral selector. As a result, eight enantiomeric pairs and two enantiomerically pure terpenes were detected. On the one hand, most of the chiral metabolites were present as scalemic mixtures, whereas α-terpineol was practically a racemate. On the other hand, (1R,5R)-(+)-β-pinene and (S)-(+)-β-phellandrene were enantiomerically pure. All the enantiomers were identified through the MS spectrum and by comparing their linear retention indices (LRI) with those of a mixture of enantiomerically pure standards. The enantiomeric distribution and the enantiomeric excess (e.e.) of the detected enantiomers are reported in Table 2.

3. Discussion

3.1. Chemical Composition and Main Components

The chemical analyses were carried out through two orthogonal columns, affording reciprocally consistent results. Most of the components identified through the nonpolar column were confirmed on the polar one, with few exceptions for some minority compounds. According to our experience, this is not an unusual phenomenon, due to the higher baseline that is sometimes observed with polyethylene glycol stationary phases. As a result, the total quantitative analysis on the nonpolar column resulted a little higher than the one with the polar stationary phase (93.0% vs. 87.3%). This discrepancy is actually acceptable if we consider that it was a 6% difference, spread over 112 compounds. On the other hand, the polar column permitted to separate some constituents that were physically inseparable with the nonpolar phase. Among them, a major compound, corresponding to peak 99, was included.

3.2. Chemical Composition and Main Components

As previously mentioned in Section 2.1, the EO distilled from the leaves of G. rugulosa can be described as composed of three main fractions: a monoterpene fraction, a sesquiterpene fraction, and a heavy fraction, the latter constituting long-chained alcohols and alkanes. This last fraction, despite being very abundant, is not common in most EOs, and its constituents are not known for presenting interesting biological activities or constituting important toxicological issues. For these reasons, the discussion of the present volatile fraction focuses on its terpene components. With this respect, the chemical composition of this EO is coherent with the one discussed, in a previous paper, for the entire genus Gynoxys and especially for the species G. miniphylla [18]. In fact, we can find many common major components, which can be better visualized normalizing each amount to the only terpene fraction, in order to neglect the contribution of the heavy components. The results of this approach are shown in Table 3.
It can be observed that these two EOs share, with the same order of magnitude, α-pinene, (E)-β-caryophyllene, α-humulene, germacrene D, δ-cadinene, and α-cadinol, whereas α-phellandrene and β-phellandrene are only typical of G. miniphylla. Furthermore, on the one hand, trans-myrtanol acetate is only present in G. miniphylla, whereas, on the other hand, caryophyllene oxide was only detected in G. rugulosa.

3.3. Biological Activities of Major Components

According to the chemical composition, we could hypothetically expect for G. rugulosa EO some of the properties expectable for the volatile fraction of G. miniphylla. For example, due to the high amount of α-pinene, the anti-inflammatory, bronchodilator, antibacterial, antifungal, and antileishmanial activities must be considered [92,93,94,95,96,97]. Likewise, a potential cholinergic capacity could be expected [98,99].
For what concerns germacrene D, to the best of the authors’ knowledge, no important biological activities have been described in the literature. This sesquiterpene is mainly known for its ecological role as an attractive for the moths of genus Heliothis and Helicoverpa [100,101,102].
Another important component is (E)-β-caryophyllene, which is probably the most common sesquiterpene hydrocarbon in EOs. This metabolite is known to possess a very wide range of biological activities, such as neuroprotective, anti-inflammatory, sedative, anxiolytic, antidepressant, anticonvulsant, and antitumor. Despite the most important activity probably being the anti-inflammatory one, exerted by (E)-β-caryophyllene via countless different mechanisms, this metabolite became quite known for being a non-psychogenic selective agonist of type 2 cannabinoid receptors (CB2-R) [103].
Another major component is α-humulene, relatively more abundant than (E)-β-caryophyllene in this EO. This metabolite is biogenetically related to (E)-β-caryophyllene, and that is the reason why we often found both sesquiterpenes together in many EOs. Like (E)-β-caryophyllene, the very common α-humulene has also been the object of pharmacological studies [104]. The main biological activity reported for α-humulene is its anticancer property, which it shares with its isomer (E)-β-caryophyllene. Furthermore, α-humulene also synergically enhances the antitumor activity of typical cytotoxic drugs (e.g., paclitaxel), by increasing their bioavailability. Anti-inflammatory, antimicrobial, antileishmanial, antiparasitic, cicatrizing, and gastroprotective activities, among others, have also been demonstrated. Of all these latter activities, the anti-inflammatory one is probably the most promising [104].
Another very common but quite less studied sesquiterpene is δ-cadinene. This metabolite is very abundant in some EOs, such as the one obtained from Kadsura longipedunculata (21.8%) and Cedar atlantica (36.3%) [105,106]. According to the literature, both EOs presented a strong antioxidant and antibacterial activity against Gram-positive bacteria. In addition, on the one hand, the EO from K. longipedunculata demonstrated a potential in vitro anti-inflammatory activity, a pro-apoptosis capacity, and a poor cytotoxic activity [105]. On the other hand, the EO from C. atlantica was mainly interesting for its anti-insect and antibiofilm activities [106].
Lastly, an interesting biological property must be mentioned for α-cadinol. In 2007, Wen et al. investigated the antiviral activity of more than 200 natural products against the severe acute respiratory syndrome coronavirus (SARS-CoV). Of all the assayed products, only 22 showed a strong activity; α-cadinol was among them [107].

3.4. Significance of the Enantiomeric Composition

The description of the enantiomeric profile for a new EO is currently a key aspect of its chemical analysis. The importance of the enantioselective analysis is evident if we consider that two enantiomers, chemically indistinguishable in a nonchiral medium, usually show dramatically different in vivo biological properties. In particular, the optical isomers can present different olfactory properties. For this reason, two EOs, showing a very similar chemical composition, can be characterized by two completely different aromas [108]. This phenomenon cannot be explained by a classical chemical analysis but can be understand comparing the enantioselective profiles.
Comparing the EO from G. rugulosa with the volatile fraction of G. miniphylla, the two enantiomeric profiles appear dramatically different [18]. This variability, which can also depend on ecological and climatic factors, attests to the existence in plants of different biosynthetic pathways, where diverse enzymes catalyze the synthesis of different enantiomers for possibly different functions.

4. Materials and Methods

4.1. Plant Material

The leaves of G. rugulosa were collected on 29 July 2020, from many shrubs in the range of 200 m around a central point of coordinates 03°59′22″ S and 79°08′41″ W, at an altitude of 2820–2900 m above the sea level. After collection, the leaves were dried at 35 °C for 48 h and stored in a dark fresh place until use. The plant was identified by one of the authors (N.C.), and a botanical specimen was deposited at the herbarium of the UTPL, with voucher code 14664. The identification was carried out on the basis of the voucher with code MO-1891627/A:4813456, deposited at the herbarium of the Missouri Botanical Garden, Saint Louis, MO, USA. This investigation was carried out under permission of the Ministry of Environment, Water, and Ecological Transition of Ecuador, with MAATE registry number MAE-DNB-CM-2016-0048.

4.2. EO Distillation and Sample Preparation

The dry, whole leaves were analytically steam-distilled in a glass Marcusson-type apparatus, where the plant material was placed in a separated reservoir, installed between the water heater and the condenser. The bottom of the collection tube was connected to the vapor conduct, such that the aqueous phase was recycled during the process (see Figure 4). Moreover, the collection tube was refrigerated, to avoid overheating of the EO. A volume of 2 mL of cyclohexane, containing n-nonane as an internal standard (0.70 mg/mL), was placed over the aqueous phase in the collection tube. With this configuration, the condensed vapors passed through the cyclohexane layer before collection, and the EO was concentrated in the organic phase. The distillation was repeated four times, for 4 h each, obtaining four samples of EO in cyclohexane, which were directly injected into GC (injection volume: 1 μL). The four distillations were carried out with 50.3 g, 33.4 g, 33.2 g, and 34.5 g of dry leaves respectively.

4.3. Qualitative (GC–MS) and Quantitative (GC–FID) Chemical Analyses

The qualitative analysis of G. rugulosa EO was carried out with gas chromatography–mass spectrometry (GC–MS) equipment, consisting of a Trace 1310 gas chromatograph, coupled to a simple quadrupole mass spectrometry detector, model ISQ 7000 (Thermo Fisher Scientific, Walthan, MA, USA). The mass spectrometer was operated in SCAN mode (scan range 40–400 m/z), with the electron ionization (EI) source set at 70 eV, the ion source at 230 °C, and the transfer line at 200 °C. A nonpolar column, based on 5% phenyl-methylpolysiloxane, and a polar one, based on a polyethylene glycol stationary phase, were applied to both the qualitative and the quantitative analyses. The nonpolar column was DB-5ms (30 m long, 0.25 mm internal diameter, and 0.25 μm film thickness), whereas the polar one was HP-INNOWax (30 m × 0.25 mm × 0.25 μm), both purchased from Agilent Technology (Santa Clara, CA, USA). For the nonpolar column, the GC oven was operated according to the following program: 50 °C for 10 min., followed by a first thermal gradient of 2 °C/min until 170 °C, and then a second gradient of 10 °C/min until 250 °C, which was maintained for 20 min (total time 98 min). With the polar column, the same thermal program was applied, except that the final temperature was set at 230 °C, due to the lower stability of the polyethylene glycol stationary phase. The injector was operated in split mode (40:1), and its temperature was set at 230 °C. The carrier gas (GC grade helium, from Indura, Guayaquil, Ecuador) was maintained at a constant flow of 1 mL/min. The components of the EO were identified by calculating the linear retention indices (LRIs) according to Van den Dool and Kratz, and by comparing these values and the respective mass spectra with data from literature (see Table 1) [109].
The quantitative analysis was conducted with the same instrument, equipped with a flame ionization detector (FID), and the same two columns used for the qualitative one. The injector parameters, carrier gas flow, and thermal programs were the same as the GC–MS analyses, except for the final temperature time, which was set at 30 min. The constituents of the EO were quantified by external calibration, using iso-propyl caproate as the calibration standard and n-nonane as the internal standard. A six-point calibration curve was traced for each column, as previously described in the literature, with a correlation coefficient of 0.998 [16]. The use of iso-propyl caproate as a quantification standard is based on the principle that, with FID detection, the relative response factors (RRFs) of different analytes versus a unique standard only depend on the combustion enthalpy and, consequently, on the molecular formula of each compound. Therefore, the RRF of each EO component was calculated as described in the literature [110,111]. The total amount of EO, against which the percentage of each component was calculated, was analytically determined through the total area of the chromatogram, to which a mean RRF value was applied. All the analytical-grade solvents, the n-alkanes (C9–C30) for retention indices, and the internal standard (n-nonane) were purchased from Sigma-Aldrich (St. Louis, MO, USA). The calibration standard was isopropyl caproate, obtained via synthesis in the authors’ laboratory and purified to 98.8% (GC–FID).

4.4. Enantioselective Analyses

The enantioselective analyses were carried out by GC–MS, through two enantioselective capillary columns. They were based on 2,3-diethyl-6-tert-butyldimethylsilyl-β-cyclodextrin and 2,3-diacetyl-6-tert-butyldimethylsilyl-β-cyclodextrin as chiral selectors (25 m × 250 μm internal diameter × 0.25 μm phase thickness, from Mega, MI, Italy). The GC–MS was operated with the same injector and MS parameters of the qualitative ones. With both enantioselective columns, the following thermal program was applied: 50 °C for 1 min, followed by a thermal gradient of 2 °C/min until 220 °C, which was maintained for 10 min (total time 96 min). Unlike the qualitative and quantitative analyses, a carrier gas constant pressure of 70 kPa was used instead of the constant flow of 1 mL/min. The enantiomers present in the EO, which were separable on the chiral selectors, were identified through the injection of enantiomerically pure standards (1 mg/mL, split 40:1, 1 μL injected). In this case, a mixture of n-alkanes (C9–C21) was also injected to calculate the retention indices.

5. Conclusions

The leaves of the Andean species Gynoxys rugulosa Muschl. produce an essential oil, whose chemical and enantiomeric composition was described in the present study for the first time. Despite the low distillation yield, this volatile fraction could possess some interesting biological properties, due to its chemical composition. In fact, thanks to the presence of (E)-β-caryophyllene, α-humulene, and δ-cadinene, the EO of G. rugulosa could be promising as an antibacterial agent against Gram-positive bacteria and as an anti-inflammatory product. Furthermore, the presence of different biosynthetic pathways, selective for the biosynthesis of specific enantiomers, was proposed. The biological activities, suggested in the present work, should be experimentally verified in future.

Author Contributions

Conceptualization, G.G.; investigation, Y.E.M. and N.C.; data curation, Y.E.M.; writing—original draft preparation, G.G.; writing—review and editing, O.M.; supervision, G.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Raw data are available from the authors (Y.E.M.).

Acknowledgments

The authors are very grateful to Carlo Bicchi (University of Turin, Italy) for his support with enantiomerically pure standards. The authors are also grateful to the Universidad Técnica Particular de Loja (UTPL) for supporting this investigation and open-access publication.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Malagón, O.; Ramírez, J.; Andrade, J.; Morocho, V.; Armijos, C.; Gilardoni, G. Phytochemistry and Ethnopharmacology of the Ecuadorian Flora. A Review. Nat. Prod. Commun. 2016, 11, 297. [Google Scholar] [CrossRef] [PubMed]
  2. Armijos, C.; Ramírez, J.; Salinas, M.; Vidari, G.; Suárez, A.I. Pharmacology and Phytochemistry of Ecuadorian Medicinal Plants: An Update and Perspectives. Pharmaceuticals 2021, 14, 1145. [Google Scholar] [CrossRef] [PubMed]
  3. Megadiverse Countries, UNEP-WCMC. Available online: https://www.biodiversitya-z.org/content/megadiverse-countries (accessed on 17 November 2022).
  4. Chiriboga, X.; Gilardoni, G.; Magnaghi, I.; Vita Finzi, P.; Zanoni, G.; Vidari, G. New Anthracene Derivatives from Coussarea macrophylla. J. Nat. Prod. 2003, 66, 905–909. [Google Scholar] [CrossRef] [PubMed]
  5. Quílez, A.; Berenguer, B.; Gilardoni, G.; Souccar, C.; De Mendonça, S.; Oliveira, L.F.S.; Martin-Calero, M.J.; Vidari, G. Anti-secretory, Anti-inflammatory, and Anti-Helicobacter pylori Activities of Several Fractions Isolated from Piper carpunya Ruiz & Pav. J. Ethnopharmacol. 2010, 128, 583–589. [Google Scholar]
  6. Gilardoni, G.; Tosi, S.; Mellerio, G.; Maldonado, M.E.; Chiriboga, X.; Vidari, G. Lipophilic Components from the Ecuadorian Plant Schistocarpha eupatorioides. Nat. Prod. Commun. 2011, 6, 767–772. [Google Scholar] [CrossRef]
  7. Gilardoni, G.; Chiriboga, X.; Finzi, P.V.; Vidari, G. New 3,4-Secocycloartane and 3,4-Secodammarane Triterpenes from the Ecuadorian Plant Coussarea macrophylla. Chem. Biodivers. 2015, 12, 946–954. [Google Scholar] [CrossRef]
  8. Herrera, C.; Pérez, Y.; Morocho, V.; Armijos, C.; Malagón, O.; Brito, B.; Tacán, M.; Cartuche, L.; Gilardoni, G. Preliminary Phytochemical Study of the Ecuadorian Plant Croton elegans Kunth. (Euphorbiaceae). J. Chil. Chem. Soc. 2018, 63, 3875–3877. [Google Scholar] [CrossRef]
  9. Morocho, V.; Valarezo, L.P.; Tapia, D.A.; Cartuche, L.; Cumbicus, N.; Gilardoni, G. A Rare Dirhamnosyl Flavonoid and Other Radical-scavenging Metabolites from Cynophalla mollis (Kunth) J. Presl and Colicodendron scabridum (Kunt) Seem. (Capparaceae) of Ecuador. Chem. Biodivers. 2021, 16, e2100260. [Google Scholar] [CrossRef]
  10. Council of Europe. European Pharmacopoeia; Council of Europe: Strasbourg, France, 2013; p. 743. [Google Scholar]
  11. Gilardoni, G.; Montalván, M.; Vélez, M.; Malagón, O. Chemical and Enantioselective Analysis of the Essential Oils from Different Morphological Structures of Ocotea quixos (Lam.) Kosterm. Plants 2021, 10, 2171. [Google Scholar] [CrossRef]
  12. Calvopiña, K.; Malagón, O.; Capetti, F.; Sgorbini, B.; Verdugo, V.; Gilardoni, G. A New Sesquiterpene Essential Oil from the Native Andean Species Jungia rugosa Less (Asteraceae): Chemical Analysis, Enantiomeric Evaluation, and Cholinergic Activity. Plants 2021, 10, 2102. [Google Scholar] [CrossRef]
  13. Ramírez, J.; Andrade, M.D.; Vidari, G.; Gilardoni, G. Essential Oil and Major Non-Volatile Secondary Metabolites from the Leaves of Amazonian Piper subscutatum. Plants 2021, 10, 1168. [Google Scholar] [CrossRef]
  14. Espinosa, S.; Bec, N.; Larroque, C.; Ramírez, J.; Sgorbini, B.; Bicchi, C.; Cumbicus, N.; Gilardoni, G. A Novel Chemical Profile of a Selective In Vitro Cholinergic Essential Oil from Clinopodium taxifolium (Kunth) Govaerts (Lamiaceae), a Native Andean Species of Ecuador. Molecules 2021, 26, 45. [Google Scholar] [CrossRef]
  15. Gilardoni, G.; Montalván, M.; Ortiz, M.; Vinueza, D.; Montesinos, J.V. The Flower Essential Oil of Dalea mutisii Kunth (Fabaceae) from Ecuador: Chemical, Enantioselective, and Olfactometric Analyses. Plants 2020, 9, 1403. [Google Scholar] [CrossRef]
  16. Gilardoni, G.; Matute, Y.; Ramírez, J. Chemical and Enantioselective Analysis of the Leaf Essential Oil from Piper coruscans Kunth (Piperaceae), a Costal and Amazonian Native Species of Ecuador. Plants 2020, 9, 791. [Google Scholar] [CrossRef]
  17. Malagón, O.; Bravo, C.; Vidari, G.; Cumbicus, N.; Gilardoni, G. Essential Oil and Non-Volatile Metabolites from Kaunia longipetiolata (Sch.Bip. ex Rusby) R. M. King and H. Rob., an Andean Plant Native to Southern Ecuador. Plants 2022, 11, 2972. [Google Scholar] [CrossRef]
  18. Malagón, O.; Cartuche, P.; Montaño, A.; Cumbicus, N.; Gilardoni, G. A New Essential Oil from the Leaves of the Endemic Andean Species Gynoxys miniphylla Cuatrec. (Asteraceae): Chemical and Enantioselective Analyses. Plants 2022, 11, 398. [Google Scholar] [CrossRef]
  19. Tropicos.org. Missouri Botanical Garden. Available online: https://www.tropicos.org (accessed on 17 November 2022).
  20. Muschler, R. Compositae Peruvianae et Bolivianae. Bot. Jahrb. Syst. 1914, 50, 87. [Google Scholar]
  21. Adams, R.P. Identification of Essential Oil Components by Gas Chromatography/Mass Spectrometry, 4th ed.; Allured Publishing Corporation: Carol Stream, IL, USA, 2007; ISBN 10-193263321. [Google Scholar]
  22. Cozzani, S.; Muselli, A.; Desjobert, J.-M.; Bernardini, A.-F.; Tomi, F.; Casanova, J. Chemical Composition of Essential Oil of Teucrium polium Subsp. capitatum (L.) from Corsica. Flavour Fragr. J. 2005, 20, 436–441. [Google Scholar] [CrossRef]
  23. Viña, A.; Murillo, E. Essential Oil Composition from Twelve Varieties of Basil (Ocimum spp.) Grown in Columbia. J. Braz. Chem. Soc. 2003, 14, 744–749. [Google Scholar] [CrossRef]
  24. Christensen, L.P.; Jakobsen, H.B.; Paulsen, E.; Hodal, L.; Andersen, K.E. Airborne Compositae Dermatitis: Monoterpenes and No Parthenolide are Released from Flowering Tanacetum parthenium (Feverfew) Plants. Arch. Dermatol. Res. 1999, 291, 425–431. [Google Scholar] [CrossRef]
  25. Gonny, M.; Cavaleiro, C.; Salgueiro, L.; Casanova, J. Analysis of Juniperus communis Subsp. alpina Needle, Berry, Wood and Root Oils by Combination of GC, GC/MS and 13C-NMR. Flavour Fragr. J. 2006, 21, 99–106. [Google Scholar] [CrossRef]
  26. Pozo-Bayon, M.A.; Ruiz-Rodriguez, A.; Pernin, K.; Cayot, N. Influence of Eggs on the Aroma Composition of a Sponge Cake and on the Aroma Release in Model Studies on Flavored Sponge Cakes. J. Agric. Food Chem. 2007, 55, 1418–1426. [Google Scholar] [CrossRef] [PubMed]
  27. Zeng, Y.-X.; Zhao, C.-X.; Liang, Y.-Z.; Yang, H.; Fang, H.-Z.; Yi, L.-Z.; Zeng, Z.-D. Comparative Analysis of Volatile Components from Clematis Species Growing in China. Anal. Chim. Acta 2007, 595, 328–339. [Google Scholar] [CrossRef] [PubMed]
  28. Wu, C.-M.; Liou, S.-E. Volatile Components of Water-Boiled Duck Meat and Cantonese Style Roasted Duck. J. Agric. Food Chem. 1992, 40, 838–841. [Google Scholar] [CrossRef]
  29. Gancel, A.-L.; Ollitrault, P.; Froelicher, Y.; Tomi, F.; Jacquemond, C.; Luro, F.; Brillouet, J.-M. Leaf Volatile Compounds of Six Citrus Somatic Allotetraploid Hybrids Originating from Various Combinations of Lime, Lemon, Citron, Sweet Orange, and Grapefruit. J. Agric. Food Chem. 2005, 53, 2224–2230. [Google Scholar] [CrossRef]
  30. Pennarun, A.-L.; Prost, C.; Haure, J.; Demaimay, M. Comparison of Two Microalgal Diets. 2. Influence on Odorant Composition and Organoleptic Qualities of Raw Oysters (Crassostrea gigas). J. Agric. Food Chem. 2003, 51, 2011–2018. [Google Scholar] [CrossRef]
  31. Fernandez-Segovia, I.; Escriche, I.; Gomez-Sintes, M.; Fuentes, A.; Serra, J.A. Influence of Different Preservation Treatments on the Volatile Fraction of Desalted Cod. Food Chem. 2006, 98, 473–482. [Google Scholar] [CrossRef]
  32. Brophy, J.J.; Goldsack, R.J.; Bean, A.R.; Forster, P.I.; Lepschi, B.J. Leaf Essential Oils of the Genus Leptospermum (Myrtaceae) in Eastern Australia. Part 5. Leptospermum continentale and Allies. Flavour Fragr. J. 1999, 14, 98–104. [Google Scholar] [CrossRef]
  33. Píno, J.A.; Marbot, R.; Vázquez, C. Volatile Components of the Fruits of Vangueria madagascariensis J. F. Gmel. from Cuba. J. Essent. Oil Res. 2004, 16, 302–304. [Google Scholar] [CrossRef]
  34. Nielsen, G.S.; Larsen, L.M.; Poll, L. Formation of Aroma Compounds during Long-Term Frozen Storage of Unblanched Leek (Allium ampeloprasum Var. Bulga) as Affected by Packaging Atmosphere and Slice Thickness. J. Agric. Food Chem. 2004, 52, 1234–1240. [Google Scholar] [CrossRef]
  35. Rega, B.; Fournier, N.; Nicklaus, S.; Guichard, E. Role of Pulp in Flavor Release and Sensory Perception in Orange Juice. J. Agric. Food Chem. 2004, 52, 4204–4212. [Google Scholar] [CrossRef]
  36. Fanciullino, A.-L.; Gancel, A.-L.; Froelicher, Y.; Luro, F.; Ollitrault, P.; Brillouet, J.-M. Effects of Nucleo-Cytoplasmic Interactions on Leaf Volatile Compounds from Citrus Somatic Diploid Hybrids. J. Agric. Food Chem. 2005, 53, 4517–4523. [Google Scholar] [CrossRef]
  37. Fröhlich, O.; Duque, C.; Schreier, P. Volatile Constituents of Curuba (Passiflora mollissima) Fruit. J. Agric. Food Chem. 1989, 37, 421–425. [Google Scholar] [CrossRef]
  38. Gauvin-Bialecki, A.; Marodon, C. Essential Oil of Ayapana triplinervis from Reunion Island: A Good Natural Source of Thymohydroquinone Dimethyl Ether. Biochem. Syst. Ecol. 2009, 36, 853–858. [Google Scholar] [CrossRef]
  39. Osorio, C.; Alarcon, M.; Moreno, C.; Bonilla, A.; Barrios, J.; Garzon, C.; Duque, C. Characterization of Odor-Active Volatiles in Champa (Campomanesia lineatifolia R.P.). J. Agric. Food Chem. 2006, 54, 509–516. [Google Scholar] [CrossRef]
  40. Dob, T.; Dahmane, D.; Agli, M.; Chelghoum, C. Essential Oil Composition of Lavandula stoechas from Algeria. Pharm. Biol. 2006, 44, 60–64. [Google Scholar] [CrossRef]
  41. Selli, S.; Rannou, C.; Prost, C.; Robin, J.; Serot, T. Characterization of Aroma-Active Compounds in Rainbow Trout (Oncorhynchus mykiss) Eliciting an Off-Odor. J. Agric. Food Chem. 2006, 54, 9496–9502. [Google Scholar] [CrossRef]
  42. Werkhoff, P.; Güntert, M.; Krammer, G.; Sommer, H.; Kaulen, J. Vacuum Headspace Method in Aroma Research: Flavor Chemistry of Yellow Passion Fruits. J. Agric. Food Chem. 1998, 46, 1076–1093. [Google Scholar] [CrossRef]
  43. Saroglou, V.; Marin, P.D.; Rancic, A.; Veljic, M.; Skaltsa, H. Composition and Antimicrobial Activity of the Essential Oil of Six Hypericum Species from Serbia. Biochem. Syst. Ecol. 2007, 35, 146–152. [Google Scholar] [CrossRef]
  44. Condurso, C.; Verzera, A.; Romeo, V.; Ziino, M.; Trozzi, A.; Ragusa, S. The Leaf Volatile Constituents of Isatis tinctoria by Solid-Phase Microextraction and Gas Chromatography/Mass Spectrometry. Planta Med. 2006, 72, 924–928. [Google Scholar] [CrossRef]
  45. Flamini, G.; Cioni, P.L.; Morelli, I.; Maccioni, S.; Baldini, R. Phytochemical Typologies in Some Populations of Myrtus communis L. on Caprione Promontory (East Liguria, Italy). Food Chem. 2004, 85, 599–604. [Google Scholar] [CrossRef]
  46. Stevanovic, T.; Garneau, F.-X.; Jean, F.-I.; Gagnon, H.; Vilotic, D.; Petrovic, S.; Ruzic, N.; Pichette, A. The Essential Oil Composition of Pinus mugo Turra from Serbia. Flavour Fragr. J. 2005, 20, 96–97. [Google Scholar] [CrossRef]
  47. Chassagne, D.; Boulanger, R.; Crouzet, J. Enzymatic Hydrolysis of Edible Passiflora Fruit Glycosides. Food Chem. 1999, 66, 281–288. [Google Scholar] [CrossRef]
  48. Lota, M.-L.; de Rocca Serra, D.; Tomi, F.; Casanova, J. Chemical Variability of Peel and Leaf Essential Oils of Mandarins from Citrus reticulata Blanco. Biochem. Syst. Ecol. 2000, 28, 61–78. [Google Scholar] [CrossRef]
  49. Salgueiro, L.R.; Pinto, E.; Goncalves, M.J.; Costa, I.; Palmeira, A.; Cavaleiro, C.; Pina-Vaz, C.; Rodrigues, A.G.; Martinez-De-Oliveira, J. Antifungal Activity of the Essential Oil of Thymus capitellatus against Candida, Aspergillus and Dermatophyte Strains. Flavour Fragr. J. 2006, 21, 749–753. [Google Scholar] [CrossRef]
  50. Píry, J.; Príbela, A.; Durcanská, J.; Farkas, P. Fractionation of Volatiles from Blackcurrant (Ribes nigrum L.) by Different Extractive Methods. Food Chem. 1995, 54, 73–77. [Google Scholar] [CrossRef]
  51. Kim, T.H.; Thuy, N.T.; Shin, J.H.; Baek, H.H.; Lee, H.J. Aroma-Active Compounds of Miniature Beefsteak plant (Mosla dianthera Maxim.). J. Agric. Food Chem. 2000, 48, 2877–2881. [Google Scholar] [CrossRef]
  52. Nielsen, G.S.; Larsen, L.M.; Poll, L. Formation of Aroma Compounds and Lipoxygenase (EC 1.13.11.12) Activity in Unblanced Leek (Allium ampeloprasum Var. Bulga) Slices during Long-Term Frozen Storage. J. Agric. Food Chem. 2003, 51, 1970–1976. [Google Scholar] [CrossRef]
  53. Parada, F.; Duque, C.; Fujimoto, Y. Free and Bound Volatile Composition and Characterization of Some Glucoconjugates as Aroma Precursors in Melón de Olor Fruit Pulp (Sicana odorifera). J. Agric. Food Chem. 2000, 48, 6200–6204. [Google Scholar] [CrossRef]
  54. Mayorga, H.; Knapp, H.; Winterhalter, P.; Duque, C. Glycosidically Bound Flavor Compounds of Cape Gooseberry (Physalis peruviana L.). J. Agric. Food Chem. 2001, 49, 1904–1908. [Google Scholar] [CrossRef]
  55. Chen, C.-C.; Kuo, M.-C.; Liu, S.-E.; Wu, C.-M. Volatile Components of Salted and Pickled Prunes (Prunus mume Sieb. et Zucc.). J. Agric. Food Chem. 1986, 34, 140–144. [Google Scholar] [CrossRef]
  56. Bertoli, A.; Menichini, F.; Noccioli, C.; Morelli, I.; Pistelli, L. Volatile Constituents of Different Organs of Psoralea bituminosa L. Flavour Fragr. J. 2004, 19, 166–171. [Google Scholar] [CrossRef]
  57. Shellie, R.; Marriott, P.; Zappia, G.; Mondello, L.; Dugo, G. Interactive Use of Linear Retention Indices on Polar and Apolar Columns with an MS-Library for Reliable Characterization of Australian Tea Tree and Other Melaleuca sp. Oils. J. Essent. Oil Res. 2003, 15, 305–312. [Google Scholar] [CrossRef]
  58. Kurashov, E.A.; Mitrukova, G.G.; Krylova, Y.V. Variations in the Component Composition of Essential Oil of Ceratophyllum demersum (Ceratophyllaceae) during Vegetation. Plant Resour. (Rastit. Resur.) 2014, 1. in press. [Google Scholar]
  59. Pala-Paul, J.; Brophy, J.J.; Perez-Alonso, M.J.; Usano, J.; Soria, S.C. Essential Oil Composition of the Different Parts of Eryngium corniculatum Lam. (Apiaceae) from Spain. J. Chromatogr. A 2007, 1175, 289–293. [Google Scholar] [CrossRef]
  60. Riu-Aumatell, M.; Lopez-Tamames, E.; Buxaderas, S. Assessment of the Volatile Composition of Juices of Apricot, Peach, and Pear According to Two Pectolytic Treatments. J. Agric. Food Chem. 2005, 53, 7837–7843. [Google Scholar] [CrossRef]
  61. Grujic-Jovanovic, S.; Skaltsa, H.D.; Marin, P.; Sokovic, M. Composition and Antibacterial Activity of the Essential Oil of Six Stachys Species from Serbia. Flavour Fragr. J. 2004, 19, 139–144. [Google Scholar] [CrossRef]
  62. Ka, M.-H.; Choi, E.H.; Chun, H.-S.; Lee, K.-G. Antioxidative Activity of Volatile Extracts Isolated from Angelica tenuissimae Roots, Peppermint Leaves, Pine Needles, and Sweet Flag Leaves. J. Agric. Food Chem. 2005, 53, 4124–4129. [Google Scholar] [CrossRef]
  63. Yu, E.J.; Kim, T.H.; Kim, K.H.; Lee, H.J. Characterization of Aroma-Active Compounds of Abies nephrolepis (Khingan fir) Needles Using Aroma Extract Dilution Analysis. Flavour Fragr. J. 2004, 19, 74–79. [Google Scholar] [CrossRef]
  64. Choi, H.-S.; Kim, M.-S.L.; Sawamura, M. Constituents of the Essential Oil of Cnidium officinale Makino, a Korean Medicinal Plant. Flavour Fragr. J. 2002, 17, 49–53. [Google Scholar] [CrossRef]
  65. Orav, A.; Kann, J. Determination of Peppermint and Orange Aroma Compounds in Food and Beverages. Proc. Est. Acad. Sci. Chem. 2001, 50, 217–225. [Google Scholar]
  66. Vichi, S.; Riu-Aumatell, M.; Mora-Pons, M.; Buxaderas, S.; Lopez-Tamames, E. Characterization of Volatiles in Different Dry Gins. J. Agric. Food Chem. 2005, 53, 10154–10160. [Google Scholar] [CrossRef] [PubMed]
  67. Boti, J.B.; Koukoua, G.; N’Guessan, T.Y.; Casanova, J. Chemical Variability of Conyza sumatrensis and Microglossa pyrifolia from Côte d’Ivoire. Flavour Fragr. J. 2007, 22, 27–31. [Google Scholar] [CrossRef]
  68. Hachicha, S.F.; Skanji, T.; Barrek, S.; Ghrabi, Z.G.; Zarrouk, H. Composition of the Essential Oil of Teucrium ramosissimum Desf. (Lamiaceae) from Tunisia. Flavour Fragr. J. 2007, 22, 101–104. [Google Scholar] [CrossRef]
  69. Cavaleiro, C.; Pinto, E.; Gonsales, M.J.; Salguero, L. Antifungal Activity of Juniperus Essential Oils against Dermatophyte, Asrgillus and Candida Strains. J. Appl. Microbiol. 2006, 100, 1333–1338. [Google Scholar] [CrossRef]
  70. Brophy, J.; Goldsack, R.J.; Bean, A.R.; Forster, P.I.; Lepschi, B.J. Leaf Essential Oils of the Genus Leptospermum (Myrtaceae) in Eastern Australia. Part 6. Leptospermum polygalifolium and Allies. Flavour Fragr. J. 2000, 15, 271–277. [Google Scholar] [CrossRef]
  71. Andriamaharavo, N.R. Retention Data; NIST Mass Spectrometry Data Center: Schlieren, Switzerland, 2014. [Google Scholar]
  72. Bendimerad, N.; Bendiab, S.A.T. Composition and Antibacterial Activity of Pseudocytisus integrifolius (Salisb.) Essential Oil from Algeria. J. Agric. Food Chem. 2005, 53, 2947–2952. [Google Scholar] [CrossRef]
  73. Tunaher, Z.; Kirimer, N.; Baser, K.H.C. Wood Essential Oils of Juniperus foetidissima Willd. Holzforschung 2003, 57, 140–144. [Google Scholar]
  74. Kukic, J.; Petrovic, S.; Pavlovic, M.; Couladis, M.; Tzakou, O.; Niketic, M. Composition of Essential Oil of Stachys alpina L. ssp. dinarica Murb. Flavour Fragr. J. 2006, 21, 539–542. [Google Scholar] [CrossRef]
  75. Ledauphin, J.; Basset, B.; Cohen, S.; Payot, T.; Barillier, D. Identification of Trace Volatile Compounds in Freshly Distilled Calvados and Cognac: Carbonyl and Sulphur Compounds. J. Food Comp. Anal. 2006, 19, 28–40. [Google Scholar] [CrossRef]
  76. Lazarevic, J.; Radulovic, N.; Palic, R.; Zlatkovic, B. Chemical Analysis of Volatile Constituents of Berula erecta (Hudson) Coville Subsp. erecta (Apiaceae) from Serbia. J. Essent. Oil Res. 2010, 22, 153–156. [Google Scholar] [CrossRef]
  77. Bendiabdellah, A.; El Amine Dib, M.; Djabou, N.; Allali, H.; Tabti, B.; Costa, J.; Myseli, A. Biological Activities and Volatile Constituents of Daucus muricatus L. from Algeria. Chem. Cent. J. 2012, 6, 48. [Google Scholar] [CrossRef]
  78. Kiss, M.; Csoka, M.; Gyorfi, J.; Korany, K. Comparison of the Fragrance Constituents of Tuber aestivium and Tuber brumale Gathered in Hungary. J. Appl. Bot. Food Qual. 2011, 84, 102–110. [Google Scholar]
  79. Loghmani-Khouzani, H.; Fini, O.; Safari, J. Essential Oil Composition of Rosa damascena Mill Cultivated in Central Iran. Sci. Iran. 2007, 14, 316–319. [Google Scholar]
  80. Castioni, P.; Kapetanidis, I. Volatile Constituents from Brunfelsia grandiflora ssp. grandiflora: Qualitative Analysis by GC-MS. Sci. Pharm. 1996, 64, 83–91. [Google Scholar]
  81. Madruga, M.S.; Arruda, S.G.B.; Narain, N.; Souza, J.G. Castration and Slaughter Age Effects on Panel Assessment and Aroma Compounds of the Mestico Goat Meat. Meat Sci. 2000, 56, 117–125. [Google Scholar] [CrossRef]
  82. Choi, H.-S. Headspace Analyses of Fresh Leaves and Stems of Angelica gigas Nakai, a Korean medicinal herb. Flavour Fragr. J. 2006, 21, 604–608. [Google Scholar] [CrossRef]
  83. Benkaci-Ali, F.; Baaliouamer, A.; Meklati, B.Y.; Chemat, F. Chemical Composition of Seed Essential Oils from Algerian Nigella sativa Extracted by Microwave and Hydrodistillation. Flavour Fragr. J. 2007, 22, 148–153. [Google Scholar] [CrossRef]
  84. Mebazaa, R.; Mahmoudi, A.; Fouchet, M.; Dos Santos, M.; Kamissoko, F.; Nafti, A.; Ben Cheikh, R.; Rega, B.; Camel, V. Characterization of Volatile Compounds in Tunisian Fenugreek Seeds. Food Chem. 2009, 115, 1326–1336. [Google Scholar] [CrossRef]
  85. Schwab, W.; Mahr, C.; Schreier, P. Studies on the Enzymic Hydrolysis of Bound Aroma Components from Carica papaya Fruit. J. Agric. Food Chem. 1989, 37, 1009–1012. [Google Scholar] [CrossRef]
  86. Zheng, Y.; White, E. Retention Data; NIST Mass Spectrometry Data Center: Schlieren, Switzerland, 2008. [Google Scholar]
  87. Rostad, C.E.; Pereira, W.E. Kovats and Lee Retention Indices Determined by Gas Chromatography/Mass Spectrometry for Organic Compounds of Environmental Interest. J. High Resolut. Chromatogr. 1986, 9, 328–334. [Google Scholar] [CrossRef]
  88. Miyazawa, M.; Maehara, T.; Kurose, K. Composition of the Essential Oil from the Leaves of Eruca sativa. Flavour Fragr. J. 2002, 17, 187–190. [Google Scholar] [CrossRef]
  89. Mondello, L.; Dugo, P.; Basile, A.; Dugo, G. Interactive Use of Linear Retention Indices, on Polar and Apolar Columns, with a MS-Library for Reliable Identification of Complex Mixtures. J. Microcolumn Sep. 1995, 7, 581–591. [Google Scholar] [CrossRef]
  90. Vedernikov, D.N.; Roschin, V.I. Extractive Compounds of Birch Buds (Betula pendula Roth.): I. Composition of Fatty Acids, Hydrocarbons, and Esters. Russ. J. Bioorg. Chem. 2010, 36, 894–898. [Google Scholar] [CrossRef]
  91. Sandoval-Montemayor, N.E.; Garcia, A.; Elizondo-Trevino, E.; Garza-Gonzales, E.; Alvarez, L.; Camacho-Corona, M. Chemical Composition of Hexane Extract of Citrus aurantifolia and Anti-Mycobacterium tuberculosis Activity of Some of its Constituents. Molecules 2012, 17, 11173. [Google Scholar] [CrossRef]
  92. Gil, M.L.; Jimenez, J.; Ocete, M.A.; Zarzuelo, A.; Cabo, M.M. Comparative Study of Different Essential Oils of Bupleurum. gibraltaricum Lamarck. Die Pharm. 1989, 44, 284–287. [Google Scholar]
  93. Falk, A.A.; Hagberg, M.T.; Lof, A.E.; Wigaeus-Hjelm, E.M.; Wang, Z.P. Uptake, Distribution, and Elimination of alpha-Pinene in Man after Exposure by Inhalation. Scand. J. Work Environ. Health 1990, 16, 372–378. [Google Scholar] [CrossRef]
  94. Kose, E.O.; Deniz, I.G.; Sarikurkcu, C.; Aktas, O.; Yavuz, M. Chemical Composition, Antimicrobial and Antioxidant Activities of the Essential Oils of Sideritis erythrantha Boiss. and Heldr. (var. erythrantha and var. cedretorum P.H. Davis) Endemic in Turkey. Food Chem. Toxicol. 2010, 48, 2960–2965. [Google Scholar] [CrossRef]
  95. Rivas da Silva, A.C.; Lopes, P.M.; Barros de Azevedo, M.M.; Costa, D.C.; Alviano, C.S.; Alviano, D.S. Biological Activities of alpha-Pinene and beta-Pinene Enantiomers. Molecules 2012, 17, 6305. [Google Scholar] [CrossRef] [Green Version]
  96. Rodrigues, K.A.; Amorim, L.V.; Dias, C.N.; Moraes, D.F.; Carneiro, S.M.; Carvalho, F.A. Syzygium cumini (L.) Skeels Essential Oil and Its Major Constituent alpha-Pinene Exhibit anti-Leishmania Activity through Immunomodulation in vitro. J. Ethnopharmacol. 2015, 160, 32–40. [Google Scholar] [CrossRef]
  97. Govindarajan, M.; Rajeswary, M.; Hoti, S.L.; Bhattacharyya, A.; Benelli, G. Eugenol, alpha-Pinene and beta-Caryophyllene from Plectranthus barbatus Essential Oil as Eco-friendly Larvicides against Malaria, Dengue and Japanese Encephalitis Mosquito Vectors. Parasitol. Res. 2016, 115, 807–815. [Google Scholar] [CrossRef]
  98. Perry, N.S.; Houghton, P.J.; Theobald, A.; Jenner, P.; Perry, E.K. In-vitro Inhibition of Human Erythrocyte Acetylcholinesterase by Salvia lavandulaefolia Essential Oil and Constituent Terpenes. J. Pharm. Pharmacol. 2000, 52, 895–902. [Google Scholar] [CrossRef]
  99. Miyazawa, M.; Yamafuji, C. Inhibition of Acetylcholinesterase Activity by Bicyclic Monoterpenoids. J. Agric. Food Chem. 2005, 53, 1765–1768. [Google Scholar] [CrossRef]
  100. Røstelien, T.; Borg-Karlson, A.K.; Fäldt, J.; Jacobsson, U.; Mustaparta, H. The Plant Sesquiterpene Germacrene D Specifically Activates a Major Type of Antennal Receptor Neuron of the Tobacco Budworm Moth Heliothis virescens. Chem. Senses 2000, 25, 141–148. [Google Scholar] [CrossRef]
  101. Mozuraitis, R.; Stranden, M.; Ramirez, M.I.; Borg-Karlson, A.K.; Mustaparta, H. (-)-Germacrene D Increases Attraction and Oviposition by the Tobacco Budworm Moth Heliothis virescens. Chem. Senses 2002, 27, 505–509. [Google Scholar] [CrossRef]
  102. Stranden, M.; Liblikas, I.; Koenig, W.A.; Almaas, T.J.; Borg-Karlson, A.K.; Mustaparta, H. (–)-Germacrene D Receptor Neurones in Three Species of Heliothine Moths: Structure-activity Relationships. J. Comp. Physiol. A 2003, 189, 563–577. [Google Scholar] [CrossRef]
  103. Francomano, F.; Caruso, A.; Barbarossa, A.; Fazio, A.; La Torre, C.; Ceramella, J.; Mallamaci, R.; Saturnino, C.; Iacopetta, D.; Sinicropi, M.S. β-Caryophyllene: A Sesquiterpene with countless biological properties. Appl. Sci. 2019, 9, 5420. [Google Scholar] [CrossRef]
  104. de Lacerda Leite, G.M.; de Oliveira Barbosa, M.; Pereira Lopes, M.J.; de Araújo Delmondes, G.; Souza Bezerra, D.; Moura Araújo, I.; Carvalho de Alencar, C.D.; Melo Coutinho, H.D.; Rangel Peixoto, L.; Barbosa-Filho, J.M.; et al. Pharmacological and Toxicological Activities of α-Humulene and Its Isomers: A systematic Review. Trends Food Sci. Technol. 2021, 115, 255–274. [Google Scholar] [CrossRef]
  105. Mulyaningsih, S.; Youns, M.; El-Readi, M.Z.; Ashour, M.L.; Nibret, E.; Sporer, F.; Herrmann, F.; Reichling, J.; Wink, M. Biological Activity of the Essential Oil of Kadsura longipedunculata (Schisandraceae) and Its Major Components. J. Pharm. Pharmacol. 2010, 62, 1037–1044. [Google Scholar] [CrossRef]
  106. Kačániová, M.; Galovičová, L.; Valková, V.; Ďuranová, H.; Štefániková, J.; Čmiková, N.; Vukic, M.; Vukovic, N.L.; Kowalczewski, P.Ł. Chemical Composition, Antioxidant, In Vitro and In Situ Antimicrobial, Antibiofilm, and Anti-Insect Activity of Cedar atlantica Essential Oil. Plants 2022, 11, 358. [Google Scholar] [CrossRef]
  107. Wen, C.C.; Kuo, Y.H.; Jan, J.T.; Liang, P.H.; Wang, S.Y.; Liu, H.G.; Lee, C.K.; Chang, S.T.; Kuo, C.J.; Lee, S.S.; et al. Specific Plant Terpenoids and Lignoids Possess Potent Antiviral Activities against Severe Acute Respiratory Syndrome Coronavirus. J. Med. Chem. 2007, 50, 4087–4095. [Google Scholar] [CrossRef] [PubMed]
  108. Brenna, E.; Fuganti, C.; Serra, S. Enantioselective Perception of Chiral Odorants. Tetrahedron Asymmetry 2003, 14, 1–42. [Google Scholar] [CrossRef]
  109. Van Den Dool, H.; Kratz, P.D. A Generalization of the Retention Index System Including Linear Temperature Programmed Gas—Liquid Partition Chromatography. J. Chromatogr. 1963, 11, 463–471. [Google Scholar] [CrossRef] [PubMed]
  110. De Saint Laumer, J.Y.; Cicchetti, E.; Merle, P.; Egger, J.; Chaintreau, A. Quantification in Gas Chromatography: Prediction of Flame Ionization Detector Response Factors from Combustion Enthalpies and Molecular Structures. Anal. Chem. 2010, 82, 6457–6462. [Google Scholar] [CrossRef]
  111. Tissot, E.; Rochat, S.; Debonneville, C.; Chaintreau, A. Rapid GC-FID quantification technique without authentic samples using predicted response factors. Flavour Fragr. J. 2012, 27, 290–296. [Google Scholar] [CrossRef]
Figure 1. G. rugulosa shrub at the collection site (photo by G. Gilardoni).
Figure 1. G. rugulosa shrub at the collection site (photo by G. Gilardoni).
Plants 12 00849 g001
Figure 2. GC–MS chromatogram of the EO from the leaves of G. rugulosa in a 5% phenyl-methylpolysiloxane-based column. The main components are represented with the respective peak number, according to Table 1.
Figure 2. GC–MS chromatogram of the EO from the leaves of G. rugulosa in a 5% phenyl-methylpolysiloxane-based column. The main components are represented with the respective peak number, according to Table 1.
Plants 12 00849 g002
Figure 3. GC–MS chromatogram of the EO from the leaves of G. rugulosa in a polyethylene glycol-based column. The main components are represented with the respective peak number, according to Table 1.
Figure 3. GC–MS chromatogram of the EO from the leaves of G. rugulosa in a polyethylene glycol-based column. The main components are represented with the respective peak number, according to Table 1.
Plants 12 00849 g003
Figure 4. Distillation apparatus used in this study.
Figure 4. Distillation apparatus used in this study.
Plants 12 00849 g004
Table 1. Chemical composition of G. rugulosa EO through 5% phenyl-methylpolysiloxane and polyethylene glycol GC columns.
Table 1. Chemical composition of G. rugulosa EO through 5% phenyl-methylpolysiloxane and polyethylene glycol GC columns.
N.Compound5% Phenyl-MethylpolysiloxanePolyethylene Glycol
LRI aLRI b%σReferenceLRI aLRI b%σReference
1α-pinene9339326.01.22[21]101910195.30.75[22]
2sabinene9749690.30.03[21]111511150.30.03[23]
3β-pinene9789741.60.37[21]110411031.40.20[24]
4myrcene992988trace-[21]116211620.20.04[25]
52-pentyl furan9949840.60.06[21]123112300.40.04[26]
6n-decane100010000.20.01[21]100010000.10.02-
7trans-2-(2-pentenyl)-furan10041004[27]130212820.10.02[28]
8α-phellandrene100610020.10.02[21]115711580.20.04[29]
9(2E,4E)-heptadienal100810050.40.02[21]148514880.10.01[30]
10n-octanal1010998[21]128612860.10.02[31]
11α-terpinene10171014trace-[21]117211740.10.02[32]
12(2E,4Z)-heptadienal102310130.20.02[33]146014640.10.04[34]
13p-cymene102610200.40.01[21]126312650.30.06[35]
14limonene102910240.70.10[21]119111900.10.03[36]
15β-phellandrene10311025[21]120012000.50.09[37]
16(E)-β-ocimene10481044trace-[21]125012500.20.03[38]
17benzene acetaldehyde105510360.30.04[21]163616360.30.08[26]
18terpinolene108510860.10.02[21]127412710.10.02[39]
19linalool110710950.30.10[21]155215520.30.07[40]
20n-nonanal111311000.90.09[21]138913890.90.20[41]
21p-mentha-1,5-dien-8-ol118211660.10.16[21]172217230.10.06[42]
22terpinen-4-ol 11871174[21]159315940.10.03[43]
23octanoic acid119011900.10.12[44]-----
24p-cymen-8-ol119811790.30.06[21]184518450.30.09[45]
25cryptone11991183[21]164716442.42.09[46]
26α-terpineol120411860.20.08[21]169016890.40.10[47]
27n-decanal121412010.30.03[21]149414920.20.12[48]
28verbenone122012040.20.07[21]-----
29pulegone12281233trace-[21]-----
30trans-carveol12301215trace-[21]182918300.10.02[49]
31nerol12331227trace-[21]175517550.30.20[50]
32trans-chrysanthenyl acetate12341235trace-[21]-----
33geraniol12601249trace-[21]184718470.20.06[51]
34(2E)-decenal127212600.30.06[21]163416340.30.06[52]
35nonanoic acid128712670.30.20[21]212621240.10.06[53]
36p-vinylguaiacol132313090.50.37[21]219021900.90.13[54]
37(2E,4E)-decadienal13311315trace-[21]178017800.10.04[55]
38α-copaene137513741.20.16[21]147514750.90.15[51]
39(E)-β-damascenone138613830.80.21[21]180618030.70.09[56]
40β-cubebene138813870.10.05[21]152415220.20.06[36]
41n-tetradecane140014000.20.03[21]140014000.50.36-
42α-gurjunene 140614090.50.03[21]150615070.50.10[57]
434-(2,4,4-trimethylcyclohexa-1,5-dienyl)-but-3-en-2-one141614230.10.18[58]-----
44(E)-β-caryophyllene142014172.80.45[21]157615752.40.54[51]
45α-humulene145714523.20.54[21]164816493.00.66[36]
46γ-muurolene147714780.20.07[21]167216750.30.07[59]
47germacrene D148414806.51.71[21]168916904.91.49[29]
48(E)-β-ionone14871487[21]192019230.40.14[39]
49(Z,E)-α-farnesene149314910.60.18[21]172517250.90.31[60]
50α-zingiberene149714930.90.92[21]171117130.80.50[61]
51α-muurolene150115000.60.06[21]17091706trace-[62]
52γ-cadinene151615130.20.15[21]17401738trace-[63]
53n-tridecanal151815090.60.06[21]18101809trace-[64]
54δ-cadinene152115232.20.49[21]174317442.30.76[45]
55unidentified (MW = 220)1530-0.70.12[21]1805-0.70.10-
56unidentified (MW = 220)1548-3.50.71[21]1894-3.00.62-
57germacrene D-4-ol158315742.00.55[21]203320380.20.04[65]
58spathulenol15851577[21]210521061.60.57[45]
59caryophyllene oxide158915822.20.77[21]195219531.60.61[66]
60n-hexadecane160016000.10.06[21]160016000.20.05-
61viridiflorol160115920.90.06[21]208420840.50.10[67]
62ledol161116020.40.01[21]200420070.20.19[68]
63unidentified (MW = 220)1618-1.50.68[21]2007-1.20.35-
64cubenol163516450.20.10[21]204320430.10.12[68]
65epi-α-cadinol165116380.70.54[21]215921601.40.54[49]
66epi-α-muurolol165316401.30.25[21]217421721.40.31[69]
67α-muurolol (= torreyol)165616440.80.18[21]218721870.90.27[70]
68α-cadinol166616523.80.63[21]221722184.40.85[69]
69unidentified (MW = 220)1668-1.30.55[21]2145-1.00.36-
70unidentified (MW = 220)1670-[21]-----
71α-amyl cinnamyl alcohol16701682[21]-----
72ar-turmerone167516680.10.06[21]-----
73khusinol168116791.20.12[21]2423-1.10.06§
74(1R,7S,E)-7-isopropyl-4,10-dimethylenecyclodec-5-enol169816951.20.33[71]-----
75unidentified (MW = 220)1700-[21]1433-0.40.08-
76amorpha-4,9-dien-2-ol170217000.30.38[21]2345-0.50.26§
77n-pentadecanal172417171.40.21[21]202120201.00.38[72]
78unidentified (MW = 236)1783-0.40.13[21]-----
79n-octadecane18001800trace-[21]180018000.30.15-
8014-hydroxy-δ-cadinene181118030.10.03[21]258826070.20.01[73]
81n-hexadecanal182818220.10.03[74]212921320.10.04[75]
82(2E,6E)-farnesyl acetate184418450.20.02[21]226322650.50.11[25]
836,10,14-trimethyl-2-pentadecanone185118480.30.02[76]212021250.40.11[77]
84n-hexadecanol18911874trace-[21]235623550.70.13[78]
859-nonadecene18931893[79]-----
86n-nonadecane190019000.10.01[21]190019000.30.03-
87unidentified (MW = 216)1908-trace-[21]-----
88(5E,9E)-farnesyl acetone19191913[21]237023750.30.06[80]
89n-heptadecanal192919300.40.05[81]223822470.30.05[82]
90phytol19501942trace-[21]261226110.20.06[68]
91n-hexadecanoic acid197519750.60.11[83]285028710.40.09[84]
92unidentified (MW = 256)1979-0.30.03[21]-----
931-heptadecanol199319930.20.03[81]245424510.10.06[44]
94n-eicosane200020000.10.01[21]200020000.70.46-
951-octadecanol 209320771.00.20[21]255325581.20.32[85]
96n-heneicosane210021000.50.04[21]210021000.20.20-
971-nonadecanol219621951.90.36[86]265426461.71.18[44]
98n-docosane220022000.30.17[21]220022000.50.33-
991-eicosanol229622921.20.22[87]272427170.90.10[88]
100n-tricosane230023003.40.70[21]230023003.30.71-
1011-heneicosanol239723805.81.49[89]2887-4.51.30§
102n-tetracosane24002400trace-[21]240024001.10.27-
1031-docosanol249524930.20.04[90]-----
104n-pentacosane250025007.11.99[21]250025005.81.65-
1051-tricosanol2598-4.51.11§3566-4.01.00§
106n-hexacosane26002600[21]260026000.30.09-
107n-tetracosanal264426500.50.15[91]-----
108n-heptacosane270027003.50.87[21]270027003.00.88-
1091-pentacosanol2797-0.50.11§-----
110n-octacosane28002800[21]280028000.50.17-
1111-hexacosanol286028620.80.23[21]-----
112n-triacontane300030000.30.07[21]-----
Monoterpene hydrocarbons 9.2 8.7
Oxygenated monoterpenes 1.1 4.2
Sesquiterpene hydrocarbons 23.2 19.9
Oxygenated sesquiterpenes 20.1 19.2
Others 39.4 35.3
Total 93.0 87.3
a Calculated linear retention index (LRI); b reference linear retention index (LRI); trace: <0.1%; §: identified by MS spectrum only; %: percentage amount by GC–FID; σ: standard deviation; MW: molecular weight. The compounds in bold represent the main components (≥2.5% on at least one column).
Table 2. Linear retention indices (LRI), enantiomeric distribution (%), and enantiomeric excess (e.e.) of some chiral terpenes in G. rugulosa leaves EO.
Table 2. Linear retention indices (LRI), enantiomeric distribution (%), and enantiomeric excess (e.e.) of some chiral terpenes in G. rugulosa leaves EO.
LRIEnantiomersEnantiomeric Distribution (%)e.e. (%)
918 *(1S,5S)-(−)-α-pinene62.925.8
920 *(1R,5R)-(+)-α-pinene37.1
972 *(1R,5R)-(+)-β-pinene100.0100.0
1008 *(1R,5R)-(+)-sabinene72.244.4
1016 *(1S,5S)-(−)-sabinene27.8
1024 *(S)-(+)-α-phellandrene59.118.2
1026 *(R)-(−)-α-phellandrene40.9
1075 *(S)-(+)-β-phellandrene100.0100.0
1302 *(R)-(−)-linalool52.44.8
1305 *(S)-(+)-linalool47.6
1317 **(1R,2S,6S,7S,8S)-(−)-α-copaene4.391.4
1319 **(1S,2R,6R,7R,8R)-(+)-α-copaene95.7
1335 *(R)-(−)-terpinen-4-ol 42.614.8
1380 *(S)-(+)-terpinen-4-ol 57.4
1396 *(S)-(−)-α-terpineol50.10.2
1401 *(R)-(+)-α-terpineol49.9
1454 **(R)-(+)-germacrene D95.591.0
1462 **(S)-(−)-germacrene D4.5
* 2,3-Diacetyl-6-tert-butyldimethylsilyl-β-cyclodextrin column; ** 2,3-diethyl-6-tert-butyldimethylsilyl-β-cyclodextrin column.
Table 3. Normalized abundance of major components in the EOs of G. rugulosa and G. miniphylla.
Table 3. Normalized abundance of major components in the EOs of G. rugulosa and G. miniphylla.
CompoundNormalized %
G. rugulosaG. miniphylla
α-pinene10.715.3
α-phellandrene0.317.6
β-phellandrene0.93.2
trans-myrtanol acetate-9.3
(E)-β-caryophyllene4.92.7
α-humulene5.92.0
germacrene D11.714.8
δ-cadinene4.24.6
caryophyllene oxide3.9-
α-cadinol7.72.6
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

Maldonado, Y.E.; Malagón, O.; Cumbicus, N.; Gilardoni, G. A New Essential Oil from the Leaves of Gynoxys rugulosa Muschl. (Asteraceae) Growing in Southern Ecuador: Chemical and Enantioselective Analyses. Plants 2023, 12, 849. https://doi.org/10.3390/plants12040849

AMA Style

Maldonado YE, Malagón O, Cumbicus N, Gilardoni G. A New Essential Oil from the Leaves of Gynoxys rugulosa Muschl. (Asteraceae) Growing in Southern Ecuador: Chemical and Enantioselective Analyses. Plants. 2023; 12(4):849. https://doi.org/10.3390/plants12040849

Chicago/Turabian Style

Maldonado, Yessenia E., Omar Malagón, Nixon Cumbicus, and Gianluca Gilardoni. 2023. "A New Essential Oil from the Leaves of Gynoxys rugulosa Muschl. (Asteraceae) Growing in Southern Ecuador: Chemical and Enantioselective Analyses" Plants 12, no. 4: 849. https://doi.org/10.3390/plants12040849

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