Open access peer-reviewed chapter

Chemical Structure, Quality Indices and Bioactivity of Essential Oil Constituents

Written By

Nashwa Fathy Sayed Morsy

Submitted: 04 June 2016 Reviewed: 07 October 2016 Published: 08 March 2017

DOI: 10.5772/66231

From the Edited Volume

Active Ingredients from Aromatic and Medicinal Plants

Edited by Hany A. El-Shemy

Chapter metrics overview

5,367 Chapter Downloads

View Full Metrics

Abstract

Essential oil (EO) is a mixture of low molecular weight constituents that are responsible for its characteristic aroma. These constituents include terpenoid and non‐terpenoid hydrocarbons and their oxygenated derivatives. This chapter focuses on the heterogeneous composition of the essential oils. It discusses the usage of essential oil constituents as a key marker of the oil quality, freshness and unique characteristics. It describes the biological activity and synergistic effect of the essential oil constituents as antioxidant, antibacterial, antifungal and anticancer agents.

Keywords

  • chemical structure
  • quality characteristics
  • bioactivity
  • antioxidant
  • antimicrobial
  • antitumor

1. Introduction

Essential oil (EO) is defined as odorous product, of complex composition, obtained from part of plant or whole plants through various methods. It has the characteristic taste and odor of the plant from which it was derived [1, 2]. This variation in composition could be due to a variety of plants [3], geographical locations [3, 4], harvesting seasons [57], drying methods [8] and extraction methods [9, 10].

The quality, freshness and uniqueness of an essential oil are major considerations pertaining to its value [11]. Essential oils are highly sensitive to heat, moisture and oxygen [12]. Formation of oxygenated terpenes, chemical transformations, or polymerization are features of aging processes that led to quality loss. Therefore, quality control of essential oils needs to be monitored by producers, traders, or essential oil manufacturers [13, 14].

Essential oils have wide variety of bioactivities and play an important role as ideal natural sources of antimicrobial, antioxidant and chemopreventive agents [1517]. They have potential therapeutic applications in the prevention of cancer [18].

Advertisement

2. Chemical composition of essential oil

Essential oils (EOs) are volatile constituents obtained from aromatic plant material, including leaves, rhizomes, flowers, roots, bark, seeds, peel, fruits, wood and whole plants [19]. A few of the essential oils are found in animal sources, for example, musk and sperm whale, or are produced by microorganisms [20]. In plants, essential oils occur in oil cells, secretory ducts or cavities, or in glandular hairs. Essential oils can be isolated by steam distillation from an aromatic plant because of their volatility [21]. Different constituents in EOs exhibit varying smell or flavor [9]. The perception of individual volatile compounds depends on their threshold [22].

Essential oils are a complex mixture of polar and non‐polar compounds [23]. The essential oil composition depends on the species of the extracted plant, the geographic location of this plant, harvest time and extraction technique [24].

EOs can be classified into the four main groups: [25, 26]

  1. Terpenes, related to isoprene

  2. Straight‐chain compounds not containing any side chain

  3. Phenylpropanoids (benzene derivatives)

  4. Miscellaneous group having varied structures not included in first three groups (sulfur‐ or nitrogen‐containing compounds)

2.1. Terpenes, related to isoprene or isopentene

Essential oils constituents can be divided into two major groups: terpene hydrocarbons and oxygenated compounds [27].

2.1.1. Terpene hydrocarbons

Terpenes are the most common class of chemical compounds found in essential oils [28]. Terpenes are synthesized in the cytoplasm of plant cells, through the mevalonic acid pathway [19]. Terpenes have been regarded as polymers of isoprene (C5H8) joined in a repetitive head‐to‐tail manner [29] as shown in Figure 1.

Figure 1.

Link of isoprene molecules.

Figure 2.

Chemical structures of acyclic and monocyclic monoterpenes.

They could be classified according to the fusion of the isoprene units or to the number of the rings [28]. Terpenes can be classified into hemiterpenes (1 unit), monoterpenes (2 units), sesquiterpenes (3 units), diterpenes (4 units), sesterterpenes (5 units), triterpenes (6 units) and polyterpenes (many units) [30]. The combinations of two isoprene units are called a “terpene unit.” Monoterpenes (C10H16) are formed by the attachment of two isoprene units (at least one double bond). These terpenes have a hydrocarbon skeleton which can be rearranged into acyclic, cyclic, or aromatic (Figure 2). Cyclic monoterpenes can be classified according to their ring size such as monocyclic monoterpenes, bicyclic monoterpenes and tricyclic monoterpenes [31] Figure 3. These compounds oxidize easily because of their rapid reaction to air and heat sources [32].

Figure 3.

Chemical structures of bicyclic and tricyclic monoterpenes.

Sesquiterpenes (C15H24) are the second to the dominant monoterpenes. They are formed from the combination of three isoprene units [29]. Sesquiterpenes are unsaturated compounds. There are linear, branched, or cyclic sesquiterpenes (Figure 4). Sesquiterpenes are unsaturated compounds. Cyclic sesquiterpenes can be classified into monocyclic, bicyclic, or tricyclic [31] (Figure 5). Diterpenes are formed by the head‐to‐tail combinations of four isoprene units followed by rearrangement and/or substitutions (Figure 6).

Figure 4.

Chemical structures of acyclic and monocyclic sesquiterpenes.

Figure 5.

Chemical structures of bicyclic and tricyclic sesquiterpenes.

Figure 6.

Chemical structures of acyclic, monocyclic and bicyclic diterpenes.

They are important components of plant resins [33]. Diterpenes, triterpenes and tetraterpenes are present at a very low concentration in essential oils [27, 34]. Their recovery increases with increasing steam distillation times [35] and influenced by the extraction method.

2.1.2. Oxygenated compounds (terpenoids)

The oxygenated compounds are highly odoriferous [36]. Terpenoids are volatile secondary metabolites which give plants their fragrance [37]. Terpenoids can be subdivided into aldehydes, ethers, alcohols, esters, ketones, phenols and epoxides [19]. Examples: (+)‐Borneol occurs in camphor, rosemary and lavender oils. It has a camphor‐like odor, with a slightly peppery note [38]. Carvacrol is found in oregano and thyme [39] oils. It has a herbaceous, phenolic odor [32]. (-)‐Carvone has a herbal scent and is found in caraway seed oil; (+)‐carvone with a spicy‐minty odor can be found in spearmint oil [40]. 1,8‐Cineole is obtained primarily by isolation from eucalyptus oil. It has a camphor odor [41].

2.2. Straight‐chain compounds, not containing any side branches

This group contains only straight chain non‐terpenoid hydrocarbons and their oxygen derivatives: alcohols, aldehydes, ketones, acids, ethers and esters. These hydrocarbons range from n‐heptane, to compounds with 35 carbon atoms. Heptane content represented 3.8 and 36.8% of the wood volatile oils of Pinus jeffreyi and Pinus sabiniana, respectively [42]. The leaf alcohol (3(Z)‐hexen‐1‐ol, Figure 7) and its esters are responsible for the intense grassy‐green odor of freshly cut green grass and leaves via the octadecenoic pathway [4345].

Figure 7.

Chemical structure of leaf alcohol.

Figure 8.

Chemical structures of phenylpropenes.

2.3. Phenylpropenes (benzene derivatives)

These aromatic compounds are an important group to flavor and fragrance industry though it constitutes a relatively small part of the essential oils. This non‐terpenoid group comprises of constituents derived from n‐propyl benzene. The aromatic ring may carry hydroxy, methoxy and methylene dioxy groups; the propyl side chain may contain hydroxyl or carboxyl group [25].

Phenylpropenes constitute a subfamily of phenylpropanoids that are synthesized from the amino acid phenylalanine and l‐tyrosine via the shikimic acid pathway [46]. Examples of this group include trans‐anethole, methyl chavicol, eugenol, isoeugenol, vanillin, safrole, myristicin and cinnamaldehyde [19, 47] (Figure 8).

2.4. Miscellaneous group (sulfur‐ and nitrogen‐containing compounds)

The representatives of this group are the compounds, which are not included in the above mentioned three groups [25]. They are different degradation products originating from unsaturated fatty acids, lactones, terpenes, glycosides and sulfur‐ and nitrogen‐containing compounds [48].

Figure 9.

Chemical structures of sulfur‐ and nitrogen‐containing compounds.

Sulfur‐ and nitrogen‐containing compounds occur mainly as aglycones or glucosinolates, or their breakdown products, which include isothiocyanates [49]. The Brassicaceae is an important source of glucosinolates and isothiocyanates [50]. Garlic, onion, leek and shallots contain volatile sulfur compounds, namely allyl sulfide, dimethyl sulfide, diallyl disulfide and dimethylthiophene [51]. The sulfur compounds are responsible for the characteristic aroma and taste [52]. Cysteine sulfoxides including alliin predominate in mature, intact Allium spp., along with γ‐glutamylcysteines [53]. Upon rupture, such as when chopped or pressed, the action of a class of enzymes known as alliinases catalyzes the conversion of cysteine sulfoxides into the volatile thiosulfinates [54] including allicin. Allicin makes up 70–80% of the thiosulfinates. Allicin and other thiosulfinates decompose diallyl sulfide, diallyl disulfide, diallyl trisulfide, dithiins and ajoene, while the γ‐glutamylcysteines are converted to S‐allyl cysteine through a non‐alliin/allicin pathway [53]. Nitrogen‐containing compounds are found in only a few essential oils. Examples include methyl anthranilate, indole, pyridine and pyrazine. Methyl anthranilate is present in orange, lemon and bergamot oils [35] and jasmine oil [55]. Indole occurs in neroli and some citrus fruit oils [32]. Pyridines and pyrazines occur in black pepper, sweet orange and vetiver oils [35] (Figure 9).

Advertisement

3. Quality indices of essential oils

The studies have shown that a differentiation in oil quality and volatile components is associated with climatic conditions, geographical location of collection sites and other ecological and genetic factors. The influence of those factors on the accumulation of distinct volatile compounds defines its chemotypes [5658].

3.1. Black pepper oil

Black pepper oil is obtained from the dried unripe fruits of Piper nigrum [59]. The monoterpenes hydrocarbons (i.e., limonene, sabinene, β‐pinene, 3‐carene and phellandrene) to sesquiterpenes hydrocarbons (β‐caryophyllene) ratio represents the quality of the oil and indicates the aroma value; the monoterpenes provide the body of the flavor, but the sesquiterpenes provide the spicy note. A lower value represents a higher quality of the essential oil [60, 61]. A high total monoterpenes content results in a strong “peppery” top‐note and a predominantly pinene content of the terpene fraction gives a turpentine‐like note, while a high caryophyllene content results in a sweet, flowery note, which is more desirable in the flavor industry [62]. In addition to the ratio of monoterpenes to sesquiterpenes, the oxygenated compounds are supposed to provide the heart of the aroma of pepper oil [63]. The pepper oil extracted with supercritical fluid carbon dioxide was found to have a higher content of caryophyllene and oxygenated sesquiterpene compounds. The volatile oil obtained under these conditions was considered superior to the steam‐distilled oil, using aroma, taste and monoterpene to sesquiterpene hydrocarbons ratio as the criteria [64, 65]. Tipsrisukond et al. [66] reported that this ratio reached 3.02 and 0.95 in the steam distilled oil and supercritical fluid extract, respectively. On the other hand, this ratio was found to be 2.21 in the hydro‐distilled oil while it was 1.66 in the supercritical fluid extract [67].

Storage of black pepper hammer milled powder for 6 months at 4°C, markedly increased monoterpene to sesquiterpene hydrocarbons ratio from 1.31 to 4.47. They attributed this increase to the degradation of some sesquiterpenes, as a result of oxidative decomposition during storage [68].

3.2. Cardamom oil

Cardamom oil is obtained from the dried ripe fruits of Elettaria cardamomum [59]. Salzer [60] reported that 1,8‐cineole and α‐terpinyl acetate together with the terpene alcohols are important for the evaluation of the aroma quality of cardamom. 1,8‐Cineole is reported to contribute to pungency while terpinyl acetate is known for its desirable, pleasant flavor of cardamom [69, 70]. The higher level of α‐terpinyl acetate compared to that of 1,8‐cineole could be used as an indicator of the superior quality of cardamom essential oil [71, 72].

Monoterpenes hydrocarbons are less odoriferous than oxygenated monoterpenes [73]. Amma et al. [74] recorded superior flavor quality for the Malabar variety with the highest α‐terpinyl acetate/1,8‐cineole ratio (1.55) when compared with the Mysore variety (1.34). Morsy [75] found that this ratio exceeded 1.6 when ultrasound‐assisted extraction (30 W, for 30 min) was used as a pretreatment for hydrodistillation (30 min) compared to 1.13 when hydrodistillation was conducted for 6 h. The presence of a low level of hydrocarbons in cardamom oil could be used as an indicator of its high quality [74].

3.3. Caraway essential oil

Caraway oil is obtained from the dried ripe fruits of Carum carvi [59]. The main components of caraway essential oil are carvone and limonene, whose mixture constituted from 97.69 to 98.62% of total oil composition [76]. The overall quality of fruits is considered to correlate with the content of essential oil and its carvone/limonene (C/L) ratio: the higher the ratio, the better the quality. C/L ratio varies from 0.90 to 2.74 [77, 78]. This ratio is variable during ripening. Limonene is of high level in green immature seeds [79]. Aćimović et al. [76] considered the quality of essential oil poor when C/L ratio recorded 0.47. The essential oil of biennial caraway varieties usually has a higher C/L ratio compared to that of the annual varieties [80, 81]. The C/L ratio decreased with the time of storage (70 days, at room temperature) of caraway seed samples [82].

3.4. Peppermint oil

Peppermint (Mentha piperita) has bright green leaves, with a fresh, slightly sweet, tangy, peppery and strong menthol notes [83]. The principal constituents of peppermint oil are menthol (30–55%), menthone, menthyl acetate and other esters [84]. The ratio of menthol/menthone is the major determinant of the flavoring quality of distilled essential oil [85]. For commercial purposes, a high oil yield with a ratio of menthol to menthone of 2:1 is desired [86]. In general, high‐quality peppermint oil (i.e., high menthol/menthone ratio) develops during full bloom [87]. The English oil contains 60–70% of menthol, the Japanese oil contains 85% and the American only about 50%. The odor and taste afford a good indication of the quality of the oil.

3.5. Lavender oil

Lavender flowers (Lavendula angustifolia) have a strong perfumery odor with crusty woody undertones a camphor‐like note. The leaves have herbaceous and more pronounced bitter notes [83]. Linalool and its ester form, linalyl acetate, are the most abundant monoterpenes in lavender varieties and are the most desired components of the lavender oil, while trace amounts of camphor generally contribute an undesirable odor, diminishing the quality of the oil [8890]. The linalyl acetate to linalool ratio may change in different distillation times and may affect the final odor of the oil [91]. The oil samples that were obtained after 1 and 2 h of steam distillation had linalyl acetate to linalool ratio of 0.57 and 0.44, respectively. The ratio of linalyl acetate to linalool should be higher than one in high‐quality lavender oil [92].

3.6. Rose oil

Rosa damascena is an ornamental plant. This plant is used for food flavoring as dried flowers, dried bud and dried petals [93]. The flavor compounds that contribute to the distinctive scent of rose oil are β‐damascenone, β‐damascone, β‐ionone and rose oxide. Even though these compounds exist in <1% of rose oil, they make up for more than 90% of the odor content due to their low odor thresholds. The concentration of β‐damascenone is considered as the marker for the quality of the rose oil [94, 95].

Citronellol/geraniol (C/G) ratio could be used for evaluating the quality of rose oil [96]. The finest quality rose oil has C/G ratio between 1.25 and 1.30 [97]. In the rose oil trade, the citronellol content should be higher than 35%. The oils from non‐fermented petals generally contain citronellol lower than this level. Therefore, a short‐term fermentation is conducted to increase the citronellol content. The C/G ratios were higher in the oils distilled from long‐term fermented petals (e.g., 10.3 in 36 h fermentation) than non‐fermented petals (0.56). Based on these results, rose oils distilled from long‐term fermented petals are of poor quality [98].

3.7. Ginger oil

Ginger oil is obtained from the rhizomes of Zingiber officinale [59]. α‐Zingiberene is the major sesquiterpene hydrocarbon of ginger oil [99]. Salzer [60] reported that citral, zingiberene, β‐sesquiphellandrene and ar‐curcumene could be used for evaluating the quality of the ginger oil. Govindarajan [100] reported that citral and citronellyl acetate are co‐determinants of the odor of ginger oil, while zingiberene and β‐sesquiphellandrene are the main components of the freshly prepared oil. ar‐Curcumene increased with storage. A good quality oil has a ratio of zingiberene + β‐sesquiphellandrene to ar‐curcumene = 2:3.

The lemony note is attributed to citral together with α‐terpineol, while nerolidol is responsible for the woody note. β‐Sesquiphellandrene and ar‐curcumene contribute to the characteristic ginger flavor [101].

3.8. Juniper oil

Juniper (Juniperus communis) has green and sharp leaves (needles) [102]. Juniper's odor is woody and astringent with sweet, lemon and pinelike overtones. It has a ginlike aroma. Its flavor is released when it is lightly crushed [83].

Butkienë et al. [103] identified 143 components in the juniper leaves essential oil. They found that monoterpenes (M)/sesquiterpenes (S) ratio ranged from 2:1 to 5:1 according to localities in Vilnius district, Lithuania. Sela et al. [104] found that M/S ratio ranged from 1:1 to 3:1 for leaves essential oil of Macedonian juniper, from different localities. Orav et al. [105] reported that M/S ratio of juniper berries oils was higher (4:1) than that obtained from leaves samples (2.5:1). Pourmortazavi et al. [106] found that supercritical fluid extraction products from J. communis L. leaves using carbon dioxide were markedly different from the hydrodistilled oil. The hydrodistilled oil was characterized by a high concentration of β‐phellanderene. The ratio of α‐pinene to 3‐carene in hydrodistilled oil was high (0.62) compared with that of the supercritical carbon dioxide extracts (0.04–0.065). Looman and Svendsen [107] reported that the average ratio of α‐pinene:sabinene:limonene was generally 21:45:5 in the leaf essential oil of Norwegian mountain juniper (Juniperus communis L. var. saxatilis Pall).

3.9. Oregano, thyme and savory oils

Oregano, wild marjoram, Origanum vulgare, has dark green fresh leaves. Fresh oregano is available whole, chopped, or minced. The dried light green leaves are available whole, flaked, or ground [83]. The thyme plant, Thymus vulgaris L., is an evergreen herb that is used for its flavor [108]. Summer savory, Satureja hortensis, is often used as a culinary herb either as a fresh or dried herb [109].

In oregano essential oil, the total content of the thymol and carvacrol was the highest and amounted to 67.51%; in thyme essential oil, 47.47%; and in savory essential oil, 49.71%. The thymol and carvacrol types have sharp, warm and penetrating herbal (thyme type) odors, with woody, spicy and tobacco‐like notes; thymol itself has a powerful, medicated and herbaceous odor while carvacrol itself has a tar‐like odor [110]. The ratio of carvacrol to thymol in oregano, thyme and savory essential oils was 15:1, 1:19 and 1.8:1, respectively [111]. Oils containing predominantly thymol are generally considered of superior quality [112, 113].

3.10. Lemongrass oil

Lemongrass (Cymbopogon spp.) is a tall perennial C4 grass belonging to the family Poaceae (Gramineae), commonly known as the “sweet grass family” [114]. Lemongrass gives a refreshing lemon‐lime‐like taste with a tinge of mint and ginger. It has a citral odor with floral‐like (rose) and a fresh, grassy aroma. Its flavor is found in the lower tender part of the stem. The root‐end stalk gives the most flavor. The whole fresh stalk becomes aromatic when it is crushed or cut. The dried form has a very little aroma [83]. The quality of the lemongrass essential oil is measured by its citral content [115, 116]. Citral is a natural mixture of two monoterpene aldehydes, geranial (trans isomer) and neral (cis isomer). Supercritical fluid extraction was found to be a superior process than steam distillation, producing better quality lemongrass oil containing 90% citral [117]. The leaves yield aromatic oil, containing 70–90% citral. C. flexuosus has higher citral content than C. citratus [118, 119].

Advertisement

4. Bioactivities of essential oil constituents

Essential oils from different plant parts exhibit different biological activities [120]. Biological activities of essential oils include antioxidant, antimicrobial, antiviral, anti‐mutagenic and anticancer [34]. Essential oils are complex mixtures of terpenoids and phenylpropanoids compounds extracted by distillation or solvent extraction [121]. Overall activity cannot be attributed to only one of the major constituents [122]. The inactive compounds might influence resorption, the rate of reactions and biological activity of the active compounds. The combination of the major and minor constituents modifies the activity to exert significant synergistic or antagonistic effect [123].

4.1. Antioxidant activity

Antioxidants are substances that protect cells from being oxidized by free radicals. Reactive oxygen species (ROS) are highly reactive toxic molecules. ROS induced oxidative diseases such as ageing, arteriosclerosis, cancer, Alzheimer's disease and Parkinson's disease [124]. Living cells possess scavenging activity to diminish excess ROS that induced cellular injury. These mechanisms become inefficient, with ageing and under external stress. Therefore, dietary supplementation of natural and synthetic antioxidants is required [125]. Some synthetic antioxidants cause liver damage and have carcinogenic effects. Natural antioxidants are preferred to synthetic ones [126].

The essential oil of lemon balm, containing neral/citral, citronellal, menthone and isomenthone, showed strong antioxidant activity [127].

The essential oils with good radical‐scavenging activity could be arranged in the following order, clove > cinnamon > nutmeg > basil > oregano > thyme [128].

Misharina and Samusenko [129] stored a mixture of lemon, coriander and clove buds essential oils (1:1:1) at room temperature in the light for 145 days. They found that the mixture of the three essential oils increased the stability of limonene and γ‐terpinene significantly higher than in the individual essential oils and inhibited oxidation of hexanal efficiently. It was shown that terpene hydrocarbons break free‐radical chain reactions, which is accompanied by their irreversible oxidation into inert compounds, such as p‐cymene. Therefore, terpene hydrocarbons do not exhibit the properties of prooxidants.

Wei and Shibamoto [130] proposed that terpenes and terpenoids that contribute to the antioxidant activity of essential oils include α‐terpinene, β‐terpinene and β‐terpinolene in tea tree; 1,8‐cineole in water mint, menthone and isomenthone in peppermint; thymol, eugenol and linalool in black cumin, cinnamon bark and ginger; and thymol and eugenol in thyme and clove leaf.

Eugenol, the major constituent of clove essential oil, was found to have an inhibitory activity against lipid peroxidation by interfering with chain reactions of free radicals. The inhibitory activity of eugenol was about fivefold higher than that observed for alpha‐tocopherol [131]. Thymol and carvacrol the main constituents of thyme oil are shown to act as strong antioxidants [132]. The antioxidant activity of caraway oil may be due to the presence of linalool, carvacrol, anethole and estragole [133]. Marjoram and clove essential oils exerted a powerful antioxidant activity in beef burger prepared with sunflower oil during storage at -18°C for 3 months. The addition of marjoram oil at 250 mg/kg kept thiobarbituric acid value of the beef burger samples after 3 months of storage at a level not significantly different from that of control samples stored frozen for 1 month [134]. Flavoring refined corn oil with thyme increased its oxidative stability (induction time) from 4.36 to 6.48 h [135]. Blending cold‐pressed oregano (Origanum vulgare) oil with sunflower oil at 10 and 20% levels increased its induction time from 3 to 6 h and 8 h, respectively. Assiri et al. [136] attributed the superior antioxidant activity of sunflower oil blends to the high levels of phenolic compounds in oregano oil.

4.2. Antibacterial activity

Essential oils display broad‐spectrum inhibitory activities against various bacterial pathogens [137]. Essential oil is easily permeable through the cell wall and cell membrane due to its lipophilic characteristic. Interaction of essential oil components with polysaccharides, fatty acids and phospholipids causes loss of membrane integrity, leakage of the cellular contents, interference in proton pump activity and leads to cell death [124, 138140]. Other important mechanisms of action include denaturation of cellular proteins [9, 34, 141]. Carvone partitioned in the lipid membrane [142], while terpinen‐4‐ol inhibits cellular respiration and both damage cell membrane function as a permeable barrier [143]. Carvacrol and p‐cymene accumulate in the lipid phase of the membrane by causing expansion of the phospholipids bilayer and increasing spaces through which ion leakage might occur [144].

Sage essential oil is rich in antimicrobial agents [145]. The single layer wall of S. aureus is highly sensitive to essential oils with a high content of p‐cymene [34]. Cold‐pressed oregano (Origanum vulgare) oil exhibited antimicrobial activity against Staphylococcus aureus, Escherichia coli, Salmonella enteritidis, and Listeria monocytogenes with minimum lethal concentrations ranging between 160 and 320 μg/mL. The antimicrobial activity of the oregano oil could be due to the presence of phenolic constituents and the differences in the permeability of cell wall of those bacteria [136]. Ghabraie et al. [146] found that essential oils of Red bergamot (Flower top) and Chinese cinnamon (bark) inhibited Escherichia coli, Listeria monocytogenes, Staphylococcus aureus and Salmonella Typhimurium. The inhibition zone ranged from 20 to more than 70 mm depending on the target bacteria. Chinese cinnamon inhibited S. aureus and E. coli at minimum inhibition concentration of 470 ppm in microbroth dilution assay. Antibacterial activity of Red bergamot oil could be due to carvacrol and p‐cymene, while trans‐cinnamaldehyde is responsible for this effect in Chinese cinnamon oil. Radaelli et al. [147] studied the antimicrobial activities of basil, rosemary, marjoram, peppermint, thyme and anise essential oils against Clostridium perfringens strain A. They found that all oils showed bactericidal activity at their minimum inhibitory concentration, except anise oil, which displayed bacteriostatic effect.

4.3. Antifungal activity

Fungi are important causes of human infections [148]. Several crops are susceptible to fungal attack either in the field or during storage [149]. Fungicide residues are problems for the food industry [150]. Prevention of fungal growth is an effective way of impeding mycotoxin accumulation [151]. Essential oils have the ability to attack the life cycle of molds [152].

The high cost of essential oils production and the low concentration of active constituents limit their direct use in the control of fungal diseases of plants and animals. Therefore, investigation of antifungal compounds of the essential oils is considered important because of the possibility of synthesizing these compounds for the use in the control of fungal diseases [153].

Bouchra et al. [154] reported that the major essential oils constituents of Moroccan Labiatae Origanum compactum and Thymus glandulosus consisted of carvacrol, linalyl acetate and thymol. Both oils inhibited completely the growth of the mycelium of Botrytis cinerea at 100 ppm.

Serrano et al. [155] reduced the growth of yeasts and molds for stored cherries by developing active packaging materials containing eugenol, menthol, thymol and eucalyptol.

The essential oil of cinnamon had a high antifungal effect (very low minimum inhibitory concentration) against Aspergillus flavus [156, 157]. The antimycotic activity of Cinnamomum zeylanicum bark essential oil is due to the presence of cinnamaldehyde [158].

Carvacrol, the major active ingredient, of oregano oil was found to cause complete inhibition of Saccharomyces cerevisiae growth at 0.01%. Its potency was 1500 times that of the oregano oil. In contrast, the γ‐terpinene, which is the biosynthetic precursor of carvacrol, was ineffective as a fungicide. Carvacrol interfered in the target of rapamycin signaling pathway, resulting in loss of viability. Eugenol, thymol and carvacrol affect Ca2+ and H+ homeostasis leading to loss of ions and inhibition of Saccharomyces cerevisiae [159]. Citral, citronellol, geraniol and geranyl acetate that are the major components of eucalyptus oil, tea tree oil and geranium oil possess cell cycle inhibitory activities against Candida albicans [160].

Sourmaghi et al. [161] found that hydrodistilled coriander essential oil had a potent antifungal activity against Candida albicans. Linalool was the major component in coriander oil. The essential oil of coriander had synergistic antimicrobial effect with amphotericin B [162].

Rahman et al. [153] reported that the hydrodistilled essential oil of the leaves of Piper chaba Hunter displayed potent antifungal activity against Fusarium oxysporum, Phytophthora capsici, Colletotrichum capsici, Fusarium solani, and Rhizoctonia solani. They attributed this activity to α‐humulene, caryophyllene oxide, viridiflorol, globulol, β‐selinene, spathulenol, (E)‐nerolidol, linalool, 3‐pentanol and p‐cymene that present in the oil. At a concentration of 125 μg/mL, the fungicidal action of oil showed complete spore germination inhibition of Phytophthora capsici.

de Oliveira et al. [163] found that the essential oil of Piper ilheusense was active in combatting activity of Candida albicans, Candida krusei, and Candida parapsilosis. (E)‐Caryophyllene and patchouli alcohol were the major components of the oil. γ‐Cadinene, germacrene B and gleenol were found in significant quantities. Marino et al. [164] reported that the minor components might be involved in some type of synergism with the other active compounds.

4.4. Anticancer activity

Cancer is a multifactorial disease contributing towards the uncontrolled growth of the abnormal cells, leading to the formation of a tumor [165]. Carcinogenesis is a multistep process and oxidative damage that is linked to the formation of tumors. Secondary metabolites from different plants are capable of halting development of cancer [166]. Essential oil constituents have cytotoxic and antitumor activities. They play an important role in cancer prevention and treatment [18, 167]. Essential oils can be used in combination with cancer therapy to decrease the side effects of the drugs [168]. Cytotoxicity of essential oils is due to its action upon cellular integrity, leading to necrosis and apoptosis, cell cycle arrest and loss of key organelles function [169, 170]. Therefore, the evaluation of the anticancer activity of essential oils and their safety on normal cell lines are of great importance [171].

The essential oil of Ricinus communis leaves exhibited a moderate antiproliferative activity against cervical cancer line. The composition of the oil was predominantly by α‐thujone and 1,8‐cineole [172]. The Eugenia caryophyllata essential oil showed significant cytotoxic effects against HT29, A549 and Hep2 cancer cell lines. The cytotoxicity is likely due to the high concentrations of phenolic compounds, particularly eugenol [173].

Eugenol displayed cytotoxic action in a dose‐dependent manner against human hepatoma cells HepG2 and colon cells Caco‐2 [174], HL‐60 leukemia cells [175] and osteoblastic cell line U2OS [176].

Terpenoids of essential oils prevent tumor cell proliferation [34]. Linalool and linalyl acetate represented the major constituents in lavender essential oil. They were more cytotoxic than the whole essential oil against 153BR, HNDF and HMEC‐1 cancer cell lines [177]. Geraniol suppressed the growth of MCF‐7 breast cancer cells [178] and PC‐3 prostate cancer cells [179]. It was reported to interfere with membrane functions, ion homeostasis; inhibit DNA synthesis; and reduce the size of colon tumors [180]. β‐Caryophyllene did not inhibit cell growth of MCF‐7 breast cancer cell lines, but α‐humulene was cytotoxic. However, β‐caryophyllene potentiated the cytotoxicity of α‐humulene [181]. Limonene and linalyl acetate had no effect on neuroblastoma cells. However, their combination induced apoptosis [182]. This synergic effect was consistent in several studies for antitumor properties [183].

Carvacrol is a major component of oregano and thyme essential oils. It inhibits tumor cell proliferation and induces apoptosis in human colon cancer cell lines, HCT116 and LoVo [184] and in human oral squamous cell carcinoma [185]. Perillyl alcohol decreases the growth of HCT116 cells [186].

Advertisement

5. Conclusion

The essential oils are multi‐component systems, while their key components represent single component systems. Quality assurance of essential oils is imperative to ensure authenticity and product quality. The ratio of one special component or group of components to another is one of the quality indices of an essential oil as it affects its aroma.

Some of the essential oil constituents contribute to the essential oil antioxidant, antimicrobial and anticancer activities, regardless their concentration. They contribute by their synergistic effect to the property of the essential oil.

References

  1. 1. British Pharmacopoeia Commission. British pharmacopoeia (vol. II). London (GB): The Stationary Office; 2013.
  2. 2. Burdock GA. Handbook of flavor ingredients. Fenaroli's handbook of flavor ingredients. 5th ed. New York: CRC Press; 2005.
  3. 3. Sarac N, Ugur A. Antimicrobial activities of the essential oils of Origanum onites L., Origanum vulgare L. subspecies hirtum (Link) Ietswaart, Satureja thymbra L. and Thymus cilicicus Boiss. & Bal. growing wild in Turkey. Journal of Medicinal Food. 2008;11:568–573. doi:10.1089/jmf.2007.0520
  4. 4. Mechergui K, Coelho JA, Serra MC, Lamine SB, Boukhchina S, Khouja ML. Essential oils of Origanum vulgare L. subsp. glandulosum (Desf.) Ietswaart from Tunisia: chemical composition and antioxidant activity. Journal of the Science of Food and Agriculture. 2010;90:1745–1749. doi:10.1002/jsfa.4011
  5. 5. Hussaina AI, Anwar F, Sherazi STH, Przybylski R. Chemical composition, antioxidant and antimicrobial activities of basil (Ocimum basilicum) essential oils depends on seasonal variations. Food Chemistry. 2008;108:986–995. doi:10.1016/j.foodchem.2007. 12.010
  6. 6. Figueiredo AC, Barroso JG, Pedro LG, Scheffer JJC. Factors affecting secondary metabolite production in plants: volatile components and essential oils. Flavour and Fragrance Journal. 2008; 23:213–226. doi:10.1002/ffj.1875
  7. 7. Taiz L, Zeiger E. Plant physiology, 5th ed. MA, USA: Sinauer Associates Inc., Publishers Sunderland; 2010. p. 782. doi:10.1086/658450
  8. 8. Di Cesare LF, Forni E, Viscardi D, Nani RC. Changes in the chemical composition of basil caused by different drying procedures. Journal of Agricultural and Food Chemistry. 2003;51:3575–3581. doi:10.1021/jf021080o
  9. 9. Burt S. Essential oils: their antibacterial properties and potential applications in foods—a review. International Journal of Food Microbiology. 2004;94:223–253. doi: 10.1016/j.ijfoodmicro.2004.03.022
  10. 10. Karakaya S, El SN, Karagözlü N, Sahin S. Antioxidant and antimicrobial activities of essential oils obtained from oregano (Origanum vulgare spp. hirtum) by using different extraction methods. Journal of Medicinal Food. 2011;14:645–652. doi:10.1089/jmf.2010.0098
  11. 11. Shahidi F, Zhong Y. Citrus oils and essences. In: Shahidi F. editor. Bailey's industrial oil and fat products. 6th ed. Wiley Interscience. Hoboken, NJ: John Wiley & Sons, Inc.; 2005. pp. 49–66. doi:10.1002/047167849X
  12. 12. Başer KHC, Buchbauer G, editors. Handbook of essential oils science, technology and applications. 2nd ed. New York: CRC Press Taylor & Francis Group, LLC,; 2016. p. 1109. doi:10.1201/b19393
  13. 13. Turek C, Stintzing FC. Stability of essential oils: a review. Comprehensive Reviews in Food Science and Food Safety. 2013;12:40–53. doi:10.1111/1541‐4337.12006
  14. 14. Reineccius G, editor. Source book of flavors. 2nd ed. Dordrecht: Springer Science+Business Media; 1994. 928 p. doi:10.1007/978‐1‐4615‐7889‐5
  15. 15. Bassolé IH, Juliani HR. Essential oils in combination and their antimicrobial properties. Molecules. 2012;17:3989–4006. doi:10.3390/molecules17043989
  16. 16. Iqbal T, Hussain AI, Chatha SAS, Naqvi SAR, Bokhari TH. Antioxidant activity and volatile and phenolic profiles of essential oil and different extracts of wild mint (Mentha longifolia) from the Pakistani Flora. Journal of Analytical Methods in Chemistry. 2013;536490. doi:10.1155/2013/536490
  17. 17. Victoria FN, Lenardão EJ, Savegnago L, Perin G, Jacob RG, Alves D, da Silva WP, da Motta Ade S, Nascente Pda S. Essential oil of the leaves of Eugenia uniflora L.: antioxidant and antimicrobial properties. Food and Chemical Toxicology. 2012;50:2668–2674. doi:10.1016/j.fct.2012.05.002
  18. 18. Bhalla Y, Gupta VK, Jaitak V. Anticancer activity of essential oils: a review. Journal of the Science of Food and Agriculture. 2013;93:3643–3653. doi:10.1002/jsfa.6267
  19. 19. Hyldgaard M, Mygind T, Meyer RL. Essential oils in food preservation: mode of action, synergies and interactions with food matrix components. Frontiers in Microbiology. 2012;3:1–24. doi:10.3389/fmicb.2012.00012
  20. 20. Surburg H, Panten J. Common fragrance and flavor materials. Preparation, properties and uses. 6th ed. Germany: Wiley‐VCH Verlag GmbH & Co.; 2016. pp. 84–85. doi:10.1002/3527608214
  21. 21. Ríos JL. Essential oils: what they are and how the terms are used and defined. In: Preedy VR, editors. Essential oils in food preservation, flavor and safety. New York: Academic Press; 2016. pp. 3–5.
  22. 22. Tongnuanchan P, Benjakul S. Essential oils: extraction, bioactivities and their uses for food preservation. Journal of Food Science. 2014;79:1231–1249. doi:10.1111/1750‐3841.12492
  23. 23. Masango P. Cleaner production of essential oils by steam distillation. Journal of Cleaner Production. 2005;13:833–839. doi:10.1016/j.jclepro.2004.02.039
  24. 24. Dima C, Dima S. Essential oils in foods: extraction, stabilization and toxicity. Current Opinion in Food Science. 2015;5:29–35. doi:10.1016/j.cofs.2015.07.003
  25. 25. Shukla YM, Dhruve JJ, Patel NJ, Bhatnagar R, Talati JG, Kathiria KB. Plant secondary metabolites. New India Publishing Agency, New Delhi; 2009. p. 306.
  26. 26. Schmidt E. Production of essential oils. In: Baser KHC, Buchbauer G, editors. Handbook of essential oils. Science, technology and applications. Taylor & Francis Group, New York; 2010. pp. 83–118.
  27. 27. Mohamed AA, El‐Emary GA, Ali HF. Influence of some citrus essential oils on cell viability, glutathione‐s‐transferase and lipid peroxidation in Ehrlich ascites carcinoma cells. Journal of American Science. 2010;6:820–826.
  28. 28. Chizzola R. Regular monoterpenes and sesquiterpenes (essential oils). In: Ramawat KG, Mérillon JM, editors. Natural products. Berlin Heidelberg: Springer‐Verlag; 2013. pp. 2973–3008. doi:10.1007/978‐3‐642‐22144‐6_130
  29. 29. Croteau R, Kutchan TM, Lewis NG. Natural products (secondary metabolites). In: Buchanan B, Gruissem W, Jones R, editors. Biochemistry and molecular biology of plants. Rockville, MD, USA: American Society of Plant Biologists; 2000. pp. 1250–1268.
  30. 30. Zuzarte M, Salgueiro L. Essential oils chemistry. In: de Sousa DP, Paraíba JP, editors. Bioactive essential oils and cancer. Switzerland: Springer International Publishing; 2015. pp. 19–28.. doi:10.1007/978‐3‐319‐19144‐7_2
  31. 31. George, KW, Alonso‐Gutierrez, J, Keasling JD, Lee TS. Isoprenoid drugs, biofuels and chemicals—artemisinin, farnesene and beyond. Advances in Biochemical Engineering/Biotechnology. 2015;148:355–389. doi:10.1007/10_2014_288
  32. 32. Hunter M. Essential oils: art, agriculture, science, industry and entrepreneurship (a focus on the Asia‐Pacific region). Agriculture issues and policies series. New York: Nova Science Publishers, Inc; 2009.
  33. 33. Langenheim JH. Plant resins: chemistry, evolution, ecology and ethnobotany. Portland, OR, USA: Timber Press; 2003. p. 586.
  34. 34. Bakkali F, Averbeck S, Averbeck D, Idaomar M. Biological effects of essential oils—a review. Food and Chemical Toxicology. 2008;46:446–475. doi: 10.1016/j.fct.2007.09.106
  35. 35. Baser KHC, Demirci F. Chemistry of essential oils. In: Berger RG, editor. Flavours and fragrances—chemistry, bioprocessing and sustainability. Berlin: Springer; 2007. pp. 43–86. doi:10.1007/b136889
  36. 36. Darjazi BB. A comparison of volatile components of flower of page mandarin obtained by ultrasound‐assisted extraction and hydrodistillation. Journal of Medicinal Plants Research. 2011;5:2839–2847.
  37. 37. Srividya N, Davis EM, Croteau RB, Lange BM. Functional analysis of (4S)‐limonene synthase mutants reveals determinants of catalytic outcome in a model monoterpene synthase. Proceedings of the National Academy of Sciences. 2015;112:3332–3337. doi:10.1073/pnas.1501203112
  38. 38. Lafhal S, Vanloot P, Bombarda I, Kister J, Dupuy N. Identification of metabolomic markers of lavender and lavandin essential oils using mid‐infrared spectroscopy. Vibrational Spectroscopy. 2016;85:79–90. doi:org/10.1016/j.vibspec.2016.04.004
  39. 39. Siroli L, Patrignani F, Gardini F, Lanciotti R. Effects of sub‐lethal concentrations of thyme and oregano essential oils, carvacrol, thymol, citral and trans‐2‐hexenal on membrane fatty acid composition and volatile molecule profile of Listeria monocytogenes, Escherichia coli and Salmonella enteritidis. Food Chemistry. 2015;182:185–192. doi:org/10.1016/j.foodchem.2015.02.136
  40. 40. Nogoceke FP, Barcaro IMR, de Sousa DP andreatini R. Antimanic‐like effects of (R)‐(-)‐carvone and (S)‐(+)‐carvone in mice. Neuroscience Letters. 2016;619:43–48. doi:org/10.1016/j.neulet.2016.03.013
  41. 41. Navarro AM, García B, Hoyuelos FJ, Peňacoba IA, Leal J M. Preferential solvation and mixing behaviour of the essential oil 1,8‐cineole with short‐chain hydrocarbons. Fluid Phase Equilibria. 2016;429:127–136. doi: 10.1016/j.fluid.2016.08.035
  42. 42. Adams RP, Wright JW. Alkanes and terpenes in wood and leaves of Pinus jeffreyi and P. sabiniana. The Journal of Essential Oil Research. 2012;24:435–440. doi:10.1080/ 10412905.2012.703512
  43. 43. Hatanaka A, Kajiwara T, Matsui K. The biogeneration of green odour by green leaves and its physiological functions—past, present and future. Zeitschrift für Naturforschung. 1995;50:476–472. doi:10.1515/znc‐1995‐7‐801
  44. 44. Vasiliev AA, Cherkaev GV, Nikitina MA. Novel approach to fragrances with a green odour. Chemistry and Computational Simulation. Butlerov Communications. 2003; 4:33–43
  45. 45. Gigot C, Ongena M, Fauconnier M‐L, Wathelet J‐P, Jardin PD, Thonart P. The lipoxygenase metabolic pathway in plants: potential for industrial production of natural green leaf volatiles. Biotechnology, Agronomy, Society and Environment. 2010;14:451–460.
  46. 46. Vogt T. Phenylpropanoid biosynthesis. Molecular Plant. 2010;3:2–20. doi:10.1093/mp/ssp106
  47. 47. Kfoury M, Sahraoui AL‐H, Bourdon N, Laruelle F, Fontaine J, Auezova L, Greige‐Gerges H, Fourmentin S. Solubility, photostability and antifungal activity of phenylpropanoids encapsulated in cyclodextrins. Food Chemistry.2016;196:518–525. doi:10.1016/j.foodchem.2015.09.078
  48. 48. Caballero B, Trugo LC, Finglas PM. Encyclopedia of food sciences and nutrition. Amsterdam: Academic Press; 2003. pp. 2169–2176. doi:10.1016/B0‐12‐227055‐X/004 24‐7
  49. 49. Mastelič J, Blažević I, Jerković I. Free and bound sulphur containing and other volatile compounds from evergreen candytuft (Iberis sempervirens L.). Croatica Chemica Acta. 2006;79:591–597.
  50. 50. Fahey JW, Zalcmann AT, Talalay P. The chemical diversity and distribution of glucosinolates and isothiocyanates among plants. Phytochemistry. 2001;56:5–51. doi:10.1016/S0031–9422(00)00316–2
  51. 51. Iranshahi M. A review of volatile sulfur‐containing compounds from terrestrial plants: biosynthesis, distribution and analytical methods. The Journal of Essential Oil Research. 2012;24:393–434. doi:10.1080/10412905.2012.692918
  52. 52. Stajner D, Milić N, Canadanović‐Brunet J, Kapor A, Stajner M, Popović BM. Exploring Allium species as a source of potential medicinal agents. Phytotherapy Research. 2006;20:581–584. doi:10.1002/ptr.1917
  53. 53. Amagase H, Petesch BL, Matsuura H, Kasuga S, Itakura Y. Intake of garlic and its bioactive components. The Journal of Nutrition. 2001;131:955S–962S.
  54. 54. Tellez MR, Khan IA, Kobaisy M, Schrader KK, Dayan FE, Osbrink W. Composition of the essential oil of Lepidium meyenii (Walp.). Phytochemistry. 2002;61:149–155. doi:10.1016/S0031‐9422(02)00208‐X
  55. 55. Rath CC, Devi S, Dash SK, Mishra RK. Antibacterial potential assessment of jasmine essential oil against E. coli. Indian Journal of Pharmaceutical Sciences. 2008;70:238–241. doi:10.4103/0250‐474X.41465
  56. 56. Alma MH, Mavi A, Yildirim A, Digrak M, Hirata T. Screening chemical composition and in vitro antioxidant and antimicrobial activities of the essential oils from Origanum syriacum L. growing in Turkey. Biological and Pharmaceutical Bulletin. 2003;26:1725–1729. doi:10.1248/bpb.26.1725.
  57. 57. Başer KHC, Kürkcüoğlu M, Demirci B, Özek T. The essential oil of Origanum syriacum L. var. sinaicum (Boiss.) letswaart. Flavour and Fragrance Journal. 2003;18:98–99. doi:10.1002/ffj.1169.
  58. 58. Loizzo MR, Menichini F, Conforti F, Tundis R, Bonesi M, Saab AM, Statti GA, de Cindio B, Houghton PJ, Menichini F, Frega NG. Chemical analysis antioxidant antiinflammatory and anticholinesterase activities of Origanum ehrenbergii Boiss and Origanum syriacum L. essential oils. Food Chemistry. 2009;117:174–180. doi:10.1016/j.foodchem.2009.03.095
  59. 59. Bhattacharya, S. Culitvation of essential oils. In: Preedy, VR, editor. Essential oils in food preservation, flavor and safety. London, UK: Academic Press, Elsevier; 2016. pp. 25–28.
  60. 60. Salzer U‐J. Analytical evaluation of seasoning extracts (oleoresins) and essential oils from seasonings. I. International Flavours Food Additives. 1975;6:151–157 (II. Ibid. 6, 206–210. III. Ibid. 6, 253–258).
  61. 61. Salzer UJ, Furia TE. The analysis of essential oils and extracts (oleoresins) from seasonings—a critical review. Critical Reviews in Food Science and Nutrition. 1977;9:345–373.
  62. 62. Salzer U. Analytical evaluation of seasoning extracts: oleoresins and essential oils from seasonings. Critical Reviews in Food Science and Nutrition. 1977;9:242–269.
  63. 63. Govindarajan VS. Pepper‐chemistry, technology and quality evaluation. CRC Critical Reviews in Food Science and Nutrition. 1977;9:115–125. doi:10.1080/104083977095 27233
  64. 64. Sankar KU. Studies on the physicochemical characteristics of volatile oil from pepper (Piper nigrum) extracted by supercritical carbon dioxide. Journal of the Science of Food and Agriculture 1989;48:483–493. doi:10.1002/jsfa.2740480411
  65. 65. Kumoro AC, Hassan M, Singh H. Extraction of Sarawak black pepper essential oil using supercritical carbon dioxide. Arabian Journal for Science and Engineering. 2010;35:7–16
  66. 66. Tipsrisukond N, Fernando LN, Clarke AD. Antioxidant effects of essential oil and oleoresin of black pepper from supercritical carbon dioxide extractions in ground pork. Journal of Agricultural and Food Chemistry. 1998;46:4329–4333. doi:10.1021/jf9802880
  67. 67. Perakis C, Louli V, Magoulas K. Supercritical fluid extraction of black pepper oil. Journal of Food Engineering 2005;71:386–393. doi:10.1016/j.jfoodeng.2004.10.049
  68. 68. Liu H, Zeng F, Wang Q, Ou S, Tan L, Gu F. The effect of cryogenic grinding and hammer milling on the flavour quality of ground pepper (Piper nigrum L.). Food Chemistry. 2013;141:3402–3408. doi:10.1016/j.foodchem.2013.06.052
  69. 69. Karibasappa GS. Postharvest studies in large cardamom (Amomum subulatum Roxb.). Journal of Plantation Crops. 1987;15:63–64.
  70. 70. Zacharia TJ, Gopalam A. Nature, production and quality of essential oils of pepper, ginger, turmeric, cardamom and tree spices. Indian Perfumer 1987;31:188–205.
  71. 71. Pillai OGN, Mathulla T, George KM, Balakrishnan KV, Verghese J. Studies in cardamom 2. An appraisal of the excellence of Indian cardamom (Elettaria cardamomum Maton). Indian Spices. 1984;21:17–25.
  72. 72. Korikanthimath VS, Ravindra M, John ZT. Variation in yield and quality characters of cardamom clones. Journal of Medicinal and Aromatic Plant Sciences. 1997;19:1024–1027.
  73. 73. Ferhat MA, Meklati BY, Smadja J, Chemat F. An improved microwave clevenger apparatus for distillation of essential oils from orange peel. Journal of Chromatography A. 2006;1112:121–126. doi:10.1016/j.chroma.2005.12.030
  74. 74. Amma, KPP, Sasidharan I, Sreekumar, MM, Sumathykutty MA, Arumughan C. Total antioxidant capacity and change in phytochemicals of four major varieties of cardamom oils during decortication. International Journal of Food Properties. 2015;18:1317–1325. doi:10.1080/10942912.2011.587621
  75. 75. Morsy NFS. A short extraction time of high quality hydrodistilled cardamom (Elettaria cardamomum L. Maton) essential oil using ultrasound as a pretreatment. Industrial Crops and Products 2015;65:287–292. doi:10.1016/j.indcrop.2014.12.012
  76. 76. Aćimović MG, Oljača SI, Tešević VV, Todosijević MM, Djisalov JN. Evaluation of caraway essential oil from different production areas of Serbia. Horticultural Science (Prague). 2014;41:122–130.
  77. 77. Kallio H, Kerrola K, Alhonmaki P. Carvone and limonene in caraway fruits (Carum carvi L.) analyzed by supercritical carbon dioxide extraction‐gas chromatography. Journal of Agricultural and Food Chemistry. 1994;42:2478–2485. doi:10.1021/jf000 47a021
  78. 78. Sedláková J, Kocourková B, Lojková L, Kubáň V. Determination of essential oil content in caraway (Carum carvi L.) species by means of supercritical fluid extraction. Plant, Soil and Environment. 2003;49:277–282.
  79. 79. Chizolla R. Composition of the essential oil of the wild grown caraway in meadows of the Vienna region (Austria). Natural Product Comunications. 2014;9:581–582.
  80. 80. Bouwmeester HJ, Kuijpers A‐M. Relationship between assimilate supply and essential oil accumulation in annual and biennial caraway (Carum carvi L.). Journal of Essential Oil Research. 1993;5:143–152. doi:10.1080/10412905.1993.9698193
  81. 81. Bouwmeester HJ, Davies JAR, Smid HG, Welten RSA. Physiological limitations to carvone yield in caraway (Carum carvi L.). Industrial Crops and Products. 1995;4:39–51. doi:10.1016/0926‐6690(95)00009‐2
  82. 82. Sedláková J, Kocourková B, Kubán V. Determination of essential oils content and composition in caraway (Carum carvi L.). Czech Journal of Food Sciences. 2001;19:31–36
  83. 83. Raghavan, S. Handbook of spices, seasonings and flavorings. 2nd ed. New York: CRC Press, Taylor & Francis Group; 2007. p. 330.
  84. 84. Barceloux DG. Medical toxicology of natural substances. Foods, fungi, medicinal herbs, plants and venomous animals. Hoboken, New Jersey: John Wiley & Sons, Inc.; 2008. p. 1200. doi:10.1002/9780470330319.
  85. 85. Singhal RS, Kulkarni PR, Rege DV, editors. Spices, flavourants and condiments (Chapter 8). In: Handbook of indices of food quality and authenticity. Abington Hall, Abington, Cambridge, England: Woodhead Publishing Limited; 1997. pp. 386–437.
  86. 86. Lawrence BM. Peppermint oil. Perfumer and flavorist. 1984;9:92–94.
  87. 87. Rohloff J, Dragland S, Mordal R, Iversen T‐H. Effect of harvest time and drying method on biomass production, essential oil yield and quality of peppermint (Mentha x piperita L.). Journal of Agricultural and Food Chemistry. 2005;53:4143–4148. doi:10.1021/jf047998s
  88. 88. Mosandl, A. Lavender oil. In: Berger RG, editor. Flavours and fragrances. Chemistry, bioprocessing and sustainability. New York: Springer; 2007. pp. 400–403. doi:10.1007/b136889
  89. 89. d’Acampora Zellner B, Casilli A, Dugo P, Dugo G, Mondello L. Odour fingerprint acquisition by means of comprehensive two‐dimensional gas chromatography‐olfactometry and comprehensive two‐dimensional gas chromatography/mass spectrometry. Journal of Chromatography A. 2007;1141:279–286. doi:10.1016/j.chroma.2006.12.035
  90. 90. Lane A, Boecklemann A, Woronuk GN, Sarker L, Mahmoud SS. A genomic resource for investigating regulation of essential oil production in Lavandula angustifolia. Planta 2010;231:835–845. doi:10.1007/s00425‐009‐1090‐4
  91. 91. Bowles EJ. The chemistry of aromatherapeutic oils. 3rd ed. Allen & Unwin, Crows Nest, Australia; 2003. pp.164–165.
  92. 92. Porter NG, Shaw ML, Hurndell LC. Preliminary studies of lavender as an essential oil crop for New Zeland. New Zealand Journal of Agricultural Research. 1982;25:389–394. doi:10.1080/00288233.1982.10417902
  93. 93. Nasery M, Hassanzadeh MK, Najaran ZT, Emami SA. Rose (Rosa × damascena Mill.) essential oils. In: Preedy VR, editor. Essential oils in food preservation, flavor and safety. London, UK: Academic Press, Elsevier; 2016. pp. 659–665.
  94. 94. Gupta LM, Raina R. Side effects of some medicinal plants. Current Science. 1998; 75(9):897–900.
  95. 95. David F, De Clercq C, Sandra P. GC/MS/MS analysis of β‐damascenone in rose oil. GC/MS Varian App. Note 2006;52:1–3.
  96. 96. Kováts E. Composition of essential oils. Part 7. Bulgarian oil of rose (Rosa damascena Mill). Journal of Chromatography A. 1987;406:185–222. doi:10.1016/S0021‐9673(00) 94030‐5
  97. 97. Baser KHC. Turkish rose oil. Perfumer Flavorist. 1992;17:45–52.
  98. 98. Baydar H., Baydar NG. The effects of harvest date, fermentation duration and Tween 20 treatment on essential oil content and composition of industrial oil rose (Rosa damascena Mill.). Industrial Crops and Products. 2005;21:251–255. doi:10.1016/j.indcrop.2004.04.004
  99. 99. Parthasarathy VA, Chempakam B, Zachariah TJ, editors. Chemistry of spices. Oxfordshire, UK: CABI Head Office.. 2008. p. 445.
  100. 100. Govindarajan VS. Ginger—chemistry, technology and quality evaluation: part 1. CRC Critical Reviews in Food Science and Nutrition. 1982; 17:1–96. doi:10.1080/10408398 209527343
  101. 101. Zachariah TJ. Ginger. In: Parthasarathy VA, Chempakam B, Zachariah TJ, editors. Chemistry of spices. Oxfordshire, UK: CABI Head Office; 2008. pp. 70–96. doi:10.1079/9781845934057.0000
  102. 102. Ložiene K, Venskutonis PR. Juniper (Juniperus communis L.) oils. In: Preedy, VR, editor. Essential oils in food preservation, flavor and safety. London, UK: Academic Press, Elsevier; 2016. pp. 495–499.
  103. 103. Butkienë R, Nivinskienë O, Mockutë D. α‐Pinene chemotype of leaf (needle) essential oils of Juniperus communis L. growing wild in Vilnius district. Chemija. 2005;16:53–60.
  104. 104. Sela F, Karapandzova M, Stefkov G, Cvetkovikj I, Trajkovska‐Dokik E, Kaftandzieva A, Kulevanova S. Chemical composition and antimicrobial activity of leaves essential oil of Juniperus communis (Cupressaceae) grown in Republic of Macedonia. Macedonian Pharmaceutical Bulletin. 2013;59:23–32.
  105. 105. Orav A, Kailas T, Müürisepp M. Chemical investigation of the essential oil from berries and needles of common juniper (Juniperus communis L.) growing wild in Estonia. Natural Product Research. 2010;24:1789–1799. doi:10.1080/14786411003752037
  106. 106. Pourmortazavi SM, Baghaee P, Mirhosseini MA. Extraction of volatile compounds from Juniperus communis L. leaves with supercritical fluid carbon dioxide: comparison with hydrodistillation. Flavour and Fragrance Journal. 2004;19:417–420. doi:10.1002/ffj.1327
  107. 107. Looman A, Svendsen AB. The needle essential oil of Norwegian mountain juniper, Juniperus communis L. var. saxatilis Pall. Flavour and Fragrance Journal. 1992;7:23–26. doi:10.1002/ffj.2730070106
  108. 108. Mandal S, DebMandal M. Thyme (Thymus vulgaris L.) oils. In: Preedy, VR, editor. Essential oils in food preservation, flavor and safety. London, UK, Academic Press, Elsevier; 2016. pp. 825–832.
  109. 109. Hassanzadeh MK, Najaran ZT, Nasery M, Emami SA. Summer savory (Satureja hortensis L.) oils. In: Preedy, VR, editor. Essential oils in food preservation, flavor and safety. London, UK: Academic Press, Elsevier; 2016. pp. 757–763.
  110. 110. Rhind JP. Aromatherapeutic blending: essential oils in synergy. London: Singing Dragon. 2016. p. 338.
  111. 111. Alinkina ES, Misharina TA, Fatkullina LD. Antiradical properties of oregano, thyme and savory essential oils. Applied Biochemistry and Microbiology. 2013;49: 73–78. doi:10.1134/S000368381301002X
  112. 112. Panda H. Essential oils handbook. Chap. 110, 1st ed. New Delhi: National Institute of Industrial Research; 2003. p. 625.
  113. 113. Pirbalouti AG, Hashemi M, Ghahfarokhi FT. Essential oil and chemical compositions of wild and cultivated Thymus daenensis Celak and Thymus vulgaris L. Industrial Crops and Products 2013;48:43–48. doi:10.1016/j.indcrop.2013.04.004
  114. 114. Abdulazeez MA, Abdullahi AS, James BD. Lemongrass (Cymbopogon spp.) oils. In: Preedy, VR, editor. Essential oils in food preservation, flavor and safety. London, UK: Academic Press, Elsevier; 2016. pp. 509–515.
  115. 115. Chisowa EH, Hall DR, Farman DI. Volatile constituents of the essential oil of Cymbopogon citratus Stapf grown in Zambia. Flavour and Fragrance Journal. 1998;13:29–30. doi:10.1002/(SICI)1099‐1026(199801/02)13:1<29
  116. 116. Desai MA, Parikh J, De AK. Modelling and optimization studies on extraction of lemongrass oil from Cymbopogon flexuosus (Steud.) wats. Chemical Engineering Research and Design 2014;92:793–803. doi:10.1016/j.cherd.2013.08.011
  117. 117. Ha HKP, Maridable J, Gaspillo P, Hasika M, Malaluan R, Kawasaki J. Essential oil from lemongrass extracted by supercritical carbon dioxide and steam distillation. The Philippine Agricultural Scientist. 2008;91:36–41
  118. 118. Weiss EA, editor. Essential oil crops. Wallingford, UK: CAB International; 1997. p. 600.
  119. 119. Taskinen J, Mathela DK, Mathela CS. Composition of the essential oil of Cymbopogon flexuosus. Journal of Chromatography A. 1983;262:364–366. doi:10.1016/S0021‐9673(01) 88121‐8
  120. 120. Cowan MM. Plant products as antimicrobial agents. Clinical Microbiology Reviews. 1999;12:564–582.
  121. 121. Nakatsu T, Lupo AT, Chinn JW, Kang RKL. Biological activity of essential oils and their constituents. Studies in Natural Products Chemistry. 2000;21:571–631. doi:10.1016/S1572–5995(00)80014‐9
  122. 122. Isman MB, Wilson JA, Bradbury R. Insecticidal activities of commercial rosemary oils (Rosmarinus officinalis) against larvae of Pseudaletia unipuncta and Trichoplusia ni in relation to their chemical compositions. Pharmaceutical Biology. 2008;46:82–87. doi:10.1080/13880200701734661
  123. 123. Pandey AK, Singh P, Tripathi NN. Chemistry and bioactivities of essential oils of some Ocimum species: an overview. Asian Pacific Journal of Tropical Biomedicine. 2014;4:682‐694. doi:10.12980/APJTB.4.2014C77
  124. 124. Edris AE. Pharmaceutical and therapeutic potentials of essential oils and their individual volatile constituents: a review. Phytotherapy Research. 2007;21:308–323. doi:10.1002/ptr.2072
  125. 125. Halliwell B. Antioxidants and human disease: a general introduction. Nutrition Reviews. 1997; 55:49–52. doi:10.1111/j.1753‐4887.1997.tb06100.x
  126. 126. Wanasundara UN, Shahidi F. Antioxidant and pro‐oxidant activity of green tea extracts in marine oils. Food Chemistry. 1998;63:335–342. doi:10.1016/S0308‐8146(98) 00025‐9
  127. 127. Mimica‐Dukic N, Bozin B, Sokovic M, Simin N. Antimicrobial and antioxidant activities of Melissa officinalis L. (Lamiaceae) essential oil. Journal of Agricultural and Food Chemistry. 2004;52:2485–2489. doi:10.1021/jf030698a
  128. 128. Tomaino A, Cimino F, Zimbalatti V, Venuti V, Sulfaro V, De Pasquale A, Saija A. Influence of heating on antioxidant activity and the chemical composition of some spice essential oils. Food Chemistry. 2005;89:549–554. doi:10.1016/j.foodchem.2004.03.011
  129. 129. Misharina TA, Samusenko AL. Antioxidant properties of essential oils from lemon, grapefruit, coriander, clove and their mixtures. Applied Biochemistry and Microbiology. 2008;45:438–442. doi:10.1134/S0003683808040182
  130. 130. Wei A, Shibamoto T. Antioxidant/lipoxygenase inhibitory activities and chemical compositions of selected essential oils. Journal of Agricultural and Food Chemistry. 2010;58:7218–7225. doi:10.1021/jf101077s
  131. 131. Nagababu E, Rifkind JM, Boindala S, Nakka L. Assessment of antioxidant activity of eugenol in vitro and in vivo. Methods in Molecular Biology. 2010;610:165–180. doi:10.1007/978‐1‐60327‐029‐8_10
  132. 132. Miguel MG. Antioxidant and anti‐inflammatory activities of essential oils: a short review. Molecules. 2010;15:9252–9287. doi:10.3390/molecules15129252
  133. 133. Johri RK. Cuminum cyminum and Carum carvi: an update. Pharmacognosy Review. 2011;5:63–72. doi:10.4103/0973‐7847.79101
  134. 134. Abdel‐Aziz ME, Morsy NFS. Keeping quality of frozen beef patties by marjoram and clove essential oils. Journal of Food Processing and Preservation. 2015;39:956–965. doi:10.1111/jfpp.12309
  135. 135. Karoui IJ, Dhifi W, Jemia MB, Marzouk B. Thermal stability of corn oil flavoured with Thymus capitatus under heating and deep‐frying conditions. Journal of the Science of Food and Agriculture. 2011;91:927–933. doi:10.1002/jsfa.4267
  136. 136. Assiri AMA, Elbanna K, Al-Thubiani A, Ramadan MF. Cold‐pressed oregano (Origanum vulgare) oil: a rich source of bioactive lipids with novel antioxidant and antimicrobial properties. European Food Research and Technology. 2016;242:1013–1023. doi:10.1007/s00217‐015‐2607‐7
  137. 137. Teixeira B, Marques A, Ramos C, Neng NR, Nogueira JMF, Saraiva JA, Nunes ML. Chemical composition and antibacterial and antioxidant properties of commercial essential oils. Industrial Crops and Products. 2013;43:587–595. doi:10.1016/j.indcrop.2012.07.069
  138. 138. Oussalah M, Caillet S, Lacroix M. Mechanism of action of Spanish oregano, Chinese cinnamon and savory essential oils against cell membranes and walls of Escherichia coli O157:H7 and Listeria monocytogenes. Journal of Food Protection. 2006;69:1046–1055.
  139. 139. Di Pasqua R, Betts G, Hoskins N, Edwards M, Ercolini D, Mauriello G. Membrane toxicity of antimicrobial compounds from essential oils. Journal of Agricultural and Food Chemistry. 2007;55:4863–4870. doi:10.1021/jf0636465
  140. 140. Saad NY, Muller CD, Lobstein A. Major bioactivities and mechanism of action of essential oils and their components. Flavour and Fragrance Journal. 2013;28:269–279. doi:10.1002/ffj.3165
  141. 141. Gustafson JE, Liew YC, Chew S, Markham J, Bell HC, Wyllie SG, Warmington JR. Effects of tea tree oil on Escherichia coli. Letters in Applied Microbiology. 1998;26:194–198. doi:10.1046/j.1472‐765X.1998.00317.x
  142. 142. Oosterhaven K, Poolman B, Smid EJ. S‐Carvone as a natural potato sprout inhibiting, fungistatic and bacteriostatic compound. Industrial Crops and Products. 1995;4:23–31. doi:10.1016/0926‐6690(95)00007‐Y
  143. 143. Cox SD, Mann CM, Markham JL, Bell HC, Gustafson JE, Warmington JR, Wyllie SG. The mode of antimicrobial action of the essential oil of Melaleuca alternifolia (tea tree oil). Journal of Applied Microbiology. 2000;88:170–175.
  144. 144. Ultee A, Bennik MHJ, Moezelaar R. The phenolic hydroxyl group of carvacrol is essential for action against the food‐borne pathogen Bacillus cereus. Applied and Environmental Microbiology. 2002;68:1561–1568. doi:10.1128/AEM.68.4.1561‐1568. 2002
  145. 145. Jassbi AR, Asadollahi M, Masroor M, Schuman MC, Mehdizadeh Z, Soleimani M, Miri R. Chemical classification of the essential oils of the Iranian Salvia species in comparison with their botanical taxonomy. Chemistry and Biodiversity. 2012;9:1254–1271. doi:10.1002/cbdv.201100209.
  146. 146. Ghabraie M, Vu, KD, Tata L, Salmieri S, Lacroix M. Antimicrobial effect of essential oils in combinations against five bacteria and their effect on sensorial quality of ground meat. LWT‐Food Science and Technology. 2016;66:332–339. doi:10.1016/j.lwt.2015. 10.055
  147. 147. Radaelli, M, da Silva B P, Weidlich L, Hoehne L, Flach A, da Costa, LA, Ethur EM. Antimicrobial activities of six essential oils commonly used as condiments in Brazil against Clostridium perfringens. Brazilian Journal of Microbiology. 2016;47:424–430. doi:10.1016/j.bjm.2015.10.001
  148. 148. Cavaleiro C, Pinto E, Gonçalves M, Salgueiro L. Antifungal activity of Juniperus essential oils against dermatophyte, Aspergillus and Candida strains. Journal of Applied Microbiology. 2006;100:1333–1338. doi:10.1111/j.1365‐2672.2006.02862.x
  149. 149. Reyes‐Jurado F, Franco‐Vega A, Ramírez‐Corona N, Palou E, López‐Malo A. Essential oils: antimicrobial activities, extraction methods and their modeling. Food Engineering Reviews. 2015;7:275–297. doi:10.1007/s12393‐014‐9099‐2
  150. 150. Will F, Krüger E. Fungicide residues in strawberry processing. Journal of Agricultural Food Chemistry. 1999;47:858–61. doi:10.1021/jf980830k
  151. 151. Varga J, Kocsubé S, Péteri Z, Vágvölgyi C, Tóth B. Chemical, physical and biological approaches to prevent ochratoxin induced toxicoses in humans and animals. Toxins (Basel). 2010;2:1718–1750. doi:10.3390/toxins2071718
  152. 152. Sridhar SR, Rajagopal RV, Rajavel R., Masilamani S, Narasimhan S. Antifungal activity of some essential oils. Journal of Agricultural Food Chemistry. 2003;51: 7596–7599. doi:10.1021/jf0344082
  153. 153. Rahman A, Al‐Reza SM, Kang SC. Antifungal activity of essential oil and extracts of Piper chaba hunter against phytopathogenic fungi. Journal of the American Oil Chemists’ Society. 2011;88:573–579. doi:10.1007/s11746‐010‐1698‐3
  154. 154. Bouchra C, Achouri M, Idrissi Hassani LM, Hmamouchi M. Chemical composition and anti‐fungal activity of essential oils of seven Moroccan Labiatae against Botrytis cinerea Pers:Fr. Journal of Ethnopharmacology. 2003;89:165–169. doi:10.1016/S0378‐8741(03) 00275‐7
  155. 155. Serrano M, Martínez‐Romero D, Castillo S, Guillén F, Valero D. The use of natural antifungal compounds improves the beneficial effect of MAP in sweet cherry storage. Innovative Food Science and Emerging Technologies. 2005;6:115–123. doi:10.1016/j.ifset.2004.09.001
  156. 156. López P, Sanchez C, Batlle R, Nerín C. Vapor‐phase activities of cinnamon, thyme and oregano essential oils and key constituents against food borne microorganisms. Journal of Agricultural Food Chemistry. 2007;55:4348–4356. doi:10.1021/jf063295u
  157. 157. Nasir M, Tafess K, Abate D. Antimicrobial potential of the Ethiopian Thymus schimperi essential oil in comparison with others against certain fungal and bacterial species. BMC Complementary and Alternative Medicine. 2015;15:260–264. doi:10.1186/s12906‐015‐0784‐3
  158. 158. Rota MC, Herrera A, Martínez RM, Sotomayor JA, Jordán MJ. Antimicrobial activity and chemical composition of Thymus vulgaris, Thymus zygis and Thymus hyemalis essential oils. Food Control. 2008;19(7):7681–687. doi:10.1016/j.foodcont.2007.07.007
  159. 159. Rao A, Zhang Y, Muend S, Rao R. Mechanism of antifungal activity of terpenoid phenols resembles calcium stress and inhibition of the TOR pathway. Antimicrobial Agents and Chemotherapy. 2010;54:5062–5069. doi:10.1128/AAC.01050‐10
  160. 160. Zore GB, Thakre AD, Jadhav S, Karuppayil SM. Terpenoids inhibit Candida albicans growth by affecting membrane integrity and arrest of cell cycle. Phytomedicine 2011;18:1181–1190. doi:10.1016/j.phymed.2011.03.008
  161. 161. Sourmaghi MHS, Kiaee G, Golfakhrabadi F, Jamalifar H, Khanavi M. Comparison of essential oil composition and antimicrobial activity of Coriandrum sativum L. extracted by hydrodistillation and microwave‐assisted hydrodistillation. Journal of Food Science and Technology. 2015;52:2452–2457. doi:10.1007/s13197‐014‐1286‐x
  162. 162. Silva F, Ferreira S, Duarte A, Mendonça DI, Domingues FC. Antifungal activity of Coriandrum sativum essential oil, its mode of action against Candida species and potential synergism with amphotericin B. Phytomedicine. 2011;19:42–47. doi:10.1016/j.phymed.2011.06.033
  163. 163. de Oliveira RA, de Assis AMAD, da Silva LAM andrioli JL, de Oliveira FF. Chemical profile and antimicrobial activity of essential oil of Piper ilheusense. Chemistry of Natural Compounds. 2016; 52:331–333. doi:10.1007/s10600‐016‐1634‐3
  164. 164. Marino M, Bersani C, Comi G. Impedance measurements to study the antimicrobial activity of essential oils from Lamiaceae and Compositae. International Journal of Food Microbiology. 2001;67:187–195. doi:10.1016/S0168–1605(01)00447‐0
  165. 165. Pavet V, Portal MM, Moulin JC, Herbrecht R, Gronemeyer H. Towards novel paradigms for cancer therapy. Oncogene. 2011;30:1–20. doi:10.1038/onc
  166. 166. Rajput S, Mandal M. Antitumor promoting potential of selected phytochemicals derived from spices: a review. European Journal of Cancer Prevention. 2012;21:205–215. doi:10.1097/CEJ.0b013e32834a7f0c
  167. 167. Magalhães HIF, de Sousa ÉBV. Antitumor essential oils. In: de Sousa DP, editor, Bioactive essential oils and cancer. Switzerland: Springer International Publishing; 2015. pp. 135–174. doi:10.1007/978‐3‐319‐19144‐7_7
  168. 168. Hadfield N. The role of aromatherapy massage in reducing anxiety in patients with malignant brain tumours. International Journal of Palliative Nursing. 2001;7:279–285. doi:10.12968/ijpn.2001.7.6.9025
  169. 169. Bhardwaj P, Alok U, Khanna A. In vitro cytotoxicity of essential oils: a review. International Journal of Research in Pharmacy and Chemistry. 2013;3:675–681.
  170. 170. Russo R, Corasaniti MT, Bagetta G, Morrone LA. Exploitation of cytotoxicity of some essential oils for translation in cancer therapy. Evidence‐Based Complementary and Alternative Medicine. 2015. doi:10.1155/2015/397821
  171. 171. Sieniawska E, Świątek Ł, Rajtar B, Kozioł E, Polz‐Dacewicz M, Skalicka‐Wózniak K. Carrot seed essential oil—source of carotol and cytotoxicity study. Industrial Crops and Products. 2016;92:109–115. doi:10.1016/j.indcrop.2016.08.001
  172. 172. Zarai Z, Ben Chobba I, Ben Mansour R, Békir A, Gharsallah N, Kadri A. Essential oil of the leaves of Ricinus communis L.: in vitro cytotoxicity and antimicrobial properties. Lipids in health and disease. 2012;11:102–108. doi:10.1186/1476‐511X‐11‐102
  173. 173. Kouidhi B, Zmantar T, Bakhrouf A. Anticariogenic and cytotoxic activity of clove essential oil (Eugenia caryophyllata) against a large number of oral pathogens. Annals of Microbiology. 2010;60:599–604. doi:10.1007/s13213‐010‐0092‐6
  174. 174. Yoo CB, Han KT, Cho KS, Ha J, Park HJ, Nam JH, Kil UH, Lee KT. Eugenol isolated from the essential oil of Eugenia caryophyllata induces a reactive oxygen species‐mediated apoptosis in HL‐60 human promyelocytic leukemia cells. Cancer Letters. 2005;225:41–52. doi:10.1016/j.canlet.2004.11.018
  175. 175. Okada N, Hirata A, Murakami Y, Shoji M, Sakagami H, Fujisawa S. Induction of cytotoxicity and apoptosis and inhibition of cyclooxygenase‐2 gene expression by eugenol‐related compounds. Anticancer Research. 2005;25:3263–3269.
  176. 176. Ho, YC. Huang FM, Chang YC. Mechanisms of cytotoxicity of eugenol in human osteoblastic cells in vitro. International Endodontic Journal. 2006;39:389–393. doi:10.1111/j.1365‐2591.2006.01091.x
  177. 177. Prashar A, Locke IC, Evans CS. Cytotoxicity of lavender oil and its major components to human skin cells. Cell Proliferation. 2004;37:221–229. doi:10.1111/j.1365‐2184. 2004.00307.x
  178. 178. Duncan RE, Lau D, El‐Sohemy A, Archer MC. Geraniol and beta‐ionone inhibit proliferation, cell cycle progression and cyclin‐dependent kinase 2 activity in MCF‐7 breast cancer cells independent of effects on HMG‐CoA reductase activity. Biochemical Pharmacology. 2004;68:1739–1747. doi:10.1016/j.bcp.2004.06.022
  179. 179. Kim SH, Park EJ, Lee CR, Chun JN, Cho NH, Kim IG, Lee S, Kim TW, Park HH, So I, Jeon JH. Geraniol induces cooperative interaction of apoptosis and autophagy to elicit cell death in PC‐3 prostate cancer cells. International Journal of Oncology. 2012;40:1683–1690. doi:10.3892/ijo.2011.1318
  180. 180. Carnesecchi S, Bras‐Gonçalves R, Bradaia A, Zeisel M, Gossé F, Poupon MF, Raul F. Geraniol, a component of plant essential oils, modulates DNA synthesis and potentiates 5‐fluorouracil efficacy on human colon tumor xenografts. Cancer Letters. 2004;215:53–59. doi:10.1016/j.canlet.2004.06.019
  181. 181. Legault J, Pichette A. Potentiating effect of beta‐caryophyllene on anticancer activity of alpha‐humulene, isocaryophyllene and paclitaxel. Journal of Pharmacy and Pharmacology. 2007;59:1643–7. doi:10.1211/jpp.59.12.0005
  182. 182. Russo R, Ciociaro A, Berliocchi L, Cassiano MG, Rombolà L, Ragusa S, Bagetta G, Blandini F, Corasaniti MT. Implication of limonene and linalyl acetate in cytotoxicity induced by bergamot essential oil in human neuroblastoma cells. Fitoterapia. 2013;89:48–57. doi:10.1016/j.fitote.2013.05.014
  183. 183. El‐Shemy HA, Aboul‐Enein KM, Lightfoot DA. Predicting in silico which mixtures of the natural products of plants might most effectively kill human leukemia cells? Evidence‐Based Complementary and Alternative Medicine. 2013;801501. doi:10.1155/ 2013/801501
  184. 184. Fan K, Li X, Cao Y, Qi H, Li L, Zhang Q, Sun H. Carvacrol inhibits proliferation and induces apoptosis in human colon cancer cells. Anticancer Drugs. 2015;26:813–823. doi:10.1097/CAD.0000000000000263
  185. 185. Dai W, Sun C, Huang S, Zhou Q. Carvacrol suppresses proliferation and invasion in human oral squamous cell carcinoma. Journal of OncoTargets and Therapy. 2016; 9:2297–2304. doi:10.2147/OTT.S98875
  186. 186. Ma J, Li J, Wang KS, Mi C, Piao LX, Xu GH, Li X, Lee JJ, Jin X. Perillyl alcohol efficiently scavenges activity of cellular ROS and inhibits the translational expression of hypoxia‐inducible factor‐1α via mTOR/4E‐BP1 signaling pathways. International Immunopharmacology. 2016;39:1–9. doi:10.1016/j.intimp.2016.06.034.

Written By

Nashwa Fathy Sayed Morsy

Submitted: 04 June 2016 Reviewed: 07 October 2016 Published: 08 March 2017