You are on page 1of 22

,koe00056−aff −0001koe00056−aff −0002koe00056−aff −0002

VS Pragadheeshkoe00056−aff −0001koe00056−aff −0002koe00056−aff −0001 Deepa Bishtkoe00056−aff −0003koe00056−aff −0003 CS Chanotiyakoe00056−aff −0001koe00056−aff −0001

TERPENOIDS FROM ESSENTIAL OILS


1. Introduction

Terpenoids are the major group of natural product compounds present in plants.
They are formed by the combination of isoprene units (Fig. 1), as building blocks,
by joining the isoprene units in a head-to-tail fashion, which is called the isoprene
rule. Based on the number of isoprene units joined together, terpenoids are cat-
egorized into monoterpenoids, sesquiterpenoids, diterpenoids, and triterpenoids.
This article discusses the source, analysis identification, enantiomeric composi-
tions, etc. of monoterpenoids and sesquiterpenoids.
Many structurally new mono- and sesquiterpenoids (MSTs) including men-
thane derivatives, sesquiterpene lactones, and acidic compounds are also reported
from the aromatic plants from time to time (1–3).
Monoterpenoids are C10 compounds present in low boiling fractions of essen-
tial oils. Monoterpenoids are present in acyclic, monocyclic, and bicyclic forms.
Bicyclic monoterpenoids are classified into three groups based on the arrange-
ment of the carbon in the ring as (6 + 3 member), namely, thujane and carane
type; (6 + 4 member) as pinane type; and (6 + 5 member) as camphene, fenchane,
and isocamphane, isobornylane type (4).
Sesquiterpenoids are C15 compounds present in high boiling fractions
of essential oils and composed of three isoprene units linked together in a
head-to-tail fashion. Different structural types are reported for sesquiterpenoids
having open-chain (acyclic), monocyclic, bicyclic, and tricyclic skeleton elucidated
through chemical degradation of the molecules. In spite of these conventionally
reported structures, a tremendous number of possible terpene structures were
identified. Cyclization/isomerization and selective functionalization mainly by
oxygenated functionalities lead to the genesis of novel structural moieties in the
sesquiterpenoid chemistry (5). Furthermore, recent advances in spectroscopic
methods and different hyphenated techniques aid in the precise structural
characterization of natural molecules.
Farnesene, farnesol, and nerolidol are the representative compounds
of open-chain (acyclic) plant sesquiterpenoids. Cyclic sesquiterpenoids were
observed with a different structural pattern in which monocyclic sesquiter-
penoids include bisabolane, elemane, humulane, and germacrene skeletons.
Bicyclic sesquiterpenoids have cadinane, eudesmane, caryophyllene, copaene,
oplopane, cubebane, α-vetivone, β-vetivone, eremophilane, and perhydroazulene
carbon skeletons. Natural sesquiterpenoids with varied structures were isolated
and characterized from a variety of plants (3,6,7).
New unusual cyclic terpenoids with terminal pendant prenyl moieties in
sesquiterpenoids, diterpenoids, sesterterpenoids, and triterpenoids were isolated
and characterized (5). New structural patterns with sesquiterpenyl indole
molecules were identified from different plants.
Another structural pattern observed under sesquiterpenoid chemistry is
“disesquiterpenoids” or “sesquiterpenoid dimers”. These C30 molecules occur
in nature, originating biosynthetically from either two identical or different
sesquiterpenoid molecules. A cadinane sesquiterpenoid dimer, gossypol (Fig. 2)
exists as enantiomers due to restricted rotation around the central binaphthyl
1

Kirk-Othmer Encyclopedia of Chemical Technology. Copyright © 2020 John Wiley & Sons, Inc. All rights reserved.
DOI: 10.1002/0471238961.koe00056
2 TERPENOIDS FROM ESSENTIAL OILS

10
1
6 2

5 3
4

7 8
9

(a) (b)
Fig. 1. (a) Schematic representation of the isoprene rule; (b) p-Menthane skeleton and the
numbering pattern in monoterpenoids.

CHO OH OH CHO
HO OH

HO OH

Fig. 2. Structure of gossypol.

bond. (−)-Gossypol is toxic to nonruminant animals, whereas its enantiomer


exhibits relatively weak toxicity, but it shows equal toxicity to plant pathogens
(8,9).
Many sesquiterpenoid compounds are isolated from Indian aromatic plants
from different geographic conditions. Kanokonyl acetate, maaliol, patchouli alco-
hol, and 8-acetoxyl-patchouli alcohol were isolated from rhizome oil of Valeriana
wallichii (10,11).
Plants produce a variety of secondary metabolites in normal conditions
and are found to be regulated by various biotic and abiotic stress factors. The
microbe-infected roots of Chrysopogon zizanioides secrete different volatilomes,
which are absent otherwise. Direct contact of roots with the soil also attracts
various microbes, which elicit the production of spectrum of sesquiterpenes.
Many bacteria in the rhizosphere are associated with roots of vetiver and
metabolize the vetiver sesquiterpenoids (carbon source for bacteria) and release
the modified sesquiterpenoids. These sesquiterpenoids are valuable compounds
and commonly found in the vetiver essential oils of commercial importance.
These bacteria are also reported to influence gene expression of the vetiver
sesquiterpenoid biosynthesis. This shows that bacteria may have a major role
in the biosynthesis of sesquiterpenoids in vetiver essential oils (12). However,
bacteria can also produce sesquiterpenes of wide structural diversity (13). Thus,
the scientific community persists in the search of novel molecules from aromatic
plants.
1.1. Biosynthesis of Terpenoids. Terpenoids are biosynthesized from
isopentenyl pyrophosphate (IPP) and dimethylallyl pyrophosphate (DMAPP).
IPP and DMAPP are synthesized by two parallel pathways, namely, the meval-
onate (MVA) pathway in cytosol and the methylerythritol 4-phosphate (MEP)
pathway in plastids, respectively. As a general rule, the mevalonate pathway
leads to the biosynthesis of sesquiterpenes, whereas monoterpenes are formed
in the methylerythritol 4-phosphate pathway. Linear geranyl diphosphate is
the precursor of monoterpenoids (C10 ), farnesyl diphosphate is the precursor
TERPENOIDS FROM ESSENTIAL OILS 3

for sesquiterpenoids (C15 ), and geranylgeranyl diphosphate is for diterpenoids


(C20 ). All these compounds were produced by the activities of three prenyl
transferases. Terpene synthases are the enzymes that catalyze the formation
of monoterpenoids, sesquiterpenoids, and diterpenoids from their respective
precursors (14).
Terpenoids are the most structurally diverse compounds with varied
functional groups. More than hundreds of monoterpenoids, sesquiterpenoids,
diterpenoids, and triterpenoids have been identified to date, altogether compris-
ing almost 30,000 compounds. Due to the availability of a large number of terpene
synthases and their function as multiple product enzymes, a diverse class of
terpenoids are biosynthesized with a multitude of terpene carbon skeletons as
acyclic, monocyclic, bicyclic, and tricyclic compounds (Fig. 3).

2. Source of Mono- and Sesquiterpenoids

2.1. Essential Oils. Terpenoids, mainly monoterpenoids and sesquiter-


penoids, are largely present in plant essential oils, which are fragrant exudates
from aromatic plants and impart smell to the plants. They are present in essential
oils as a heterogeneous group of volatile chemicals with straight-chain, mono-
cyclic, bicyclic, tricyclic backbones and exist with various chemical functionalities
such as hydrocarbons, hydroxyl, carbonyl, ester, lactone. Blend of various MSTs
in different proportions attributes to characteristic physical and chemical prop-
erties to essential oils. Essential oils are normally liquid at room temperature,
nonpolar to moderately polar, less soluble in water and more soluble in organic
solvents.
Among various plant families, Apiaceae/Umbelliferae, Asteraceae/
Compositae, Cupressaceae, Geraniaceae, Hypericaceae/Guttiferae, Lamiaceae/
Labiatae, Lauraceae, Leguminosae/Fabaceae, Liliaceae, Malvaceae, Myrtaceae,
Oleaceae, Pinaceae, Poaceae, Rosaceae, Rutaceae, and Zingiberaceae are found to
be rich in mono- and sesquiterpenoids (15). However, the accumulation of mono-
and sesquiterpenoids has been specific to a particular tissue or organ (16,17).
Moreover, the proportion of constituents may be found in all plant parts (most of
the members of family Lamiaceae/Labiatae), or leaves (Cupressaceae, Lauraceae,
Myrtaceae, Geraniaceae, Pinaceae, Poaceae, etc.), or flowers (Rosaceae, Oleaceae,
etc.), or seeds (Apiaceae/Umbelliferae), or roots (Poaceae—C. zizanioides), or
in rhizomes (Zingiberaceae). Since mono- and sesquiterpenoids are secondary
metabolites, their amount will depend on several factors such as developmental
stages of the plant or plant parts, environmental conditions, season, temperature,
daylight (15,18).

3. Extraction of Mono- and Sesquiterpenoids (MSTs)

Mono- and sesquiterpenoids are isolated by various methods such as steam distil-
lation or hydrodistillation, cold pressing, pressurized extraction with organic sol-
vents, or supercritical carbon dioxide extraction of different plant parts. The most
4 TERPENOIDS FROM ESSENTIAL OILS

OH
Ionization Isomerization

PP
1 2 3 9

Cyclization

Cyclization
10
11
6 4
Cyclization

12
7
Isomerization
O
13 14 15
OH

Cyclization

19 21
8 5
OH

20
OH
16 17 18

Fig. 3. Pathway showing the biosynthesis of important monoterpenoids; 1. geranyl diphos-


phate, 2. geranyl cation, 3. linalyl cation, 4. α-terpinyl cation, 5. terpinyl cation, 6. pinyl
cation, 7. bornyl cation, 8. sabinyl cation, 9. linalool, 10. (E)-β-ocimene, 11. β-pinene 12.
α-pinene, 13. limonene, 14. α-terpineol, 15. 1,8-cineole, 16. α-terpinene, 17. γ-terpinene, 18.
terpinen-4-ol, 19. sabinene, 20. α-thujene, 21. borneol (14).

widely used method for extraction is hydrodistillation with Clevenger-type appa-


ratus. It is an environmentally friendly extraction method, in which water is used
to extract the essential oil. Upon heating with water, the secondary metabolites
in the plant materials are vaporized and the vaporized compounds are condensed
to separate the layer of immiscible essential oil from water (19).
TERPENOIDS FROM ESSENTIAL OILS 5

Solid-phase microextraction is a fast, solvent-free, and field-compatible tech-


nique for the extraction of volatile and semivolatile compounds in an analytical
scale. Due to high vapor pressure of monoterpenoid compounds, the headspace
extraction technique is preferred in the analysis. When a conditioned fiber is
exposed to headspace, the fiber gets saturated with the compounds present in
headspace through various processes such as adsorption, partition. The saturated
fiber is transferred to hot gas chromatographic injector for desorption, and the
extracted volatiles can be separated and quantified in gas chromatography (20).
The advantage of SPME is that it is compact and portable and can be success-
fully applied to collect samples from a distant place. Due to its unique sensitivity,
SPME is ahead of all other techniques in monitoring temporal variation in plant
volatiles and in vivo studies. The selection of fiber for the extraction process is the
most crucial task in SPME. Polarity and extraction mechanism of the fibers make
SPME selective to a class of compounds (21,22). For precise measurements and
accurate quantification, flavor chemists have made certain guidelines in SPME
measurements. The absolute amount of the compound present in a sample can
also be measured in SPME when the fiber is calibrated with standards (23,24).

4. Analysis of Monoterpenoids and Sesquiterpenoids

4.1. Gas Chromatography. Because terpenoid compounds are volatile,


gas chromatography is the best analytical technique to analyze the compounds
qualitatively and quantitatively. Essential oils, the major source of MSTs, are
a complex composition of a variable class of chemical compounds. In gas chro-
matography, the vaporized sample is carried by the mobile phase (carrier gas)
through the stationary phase for the separation of components and the separated
components are quantified by a suitable detector. The separation is based on the
difference in the partition of the solutes between the stationary phase and the
mobile phase of the chemical compounds. Retention time of the compound is the
time period in which the peak of the compound reaches its maxima in the chro-
matogram. Retention time can be used for the identification of compounds. For
the analysis of MSTs, a capillary column coated with a polar or nonpolar station-
ary phase is used with flame ionization detection. Mass spectrometer is another
detector mainly used for the identification of compounds. The most important fea-
ture of a detector for the analysis of MSTs is that the detector should be highly
sensitive, robust, and linear in a large range, and response factors for all types of
analytes should be close to unity (25).
4.2. Gas Chromatography–Mass Spectrometry (GC–MS). Gas chro-
matography can be effectively coupled with mass spectrometer for the qualitative
analysis of MST components. The separated compounds eluting out of the column
in GC is ionized by a suitable ionization mode followed by separation and quantifi-
cation of the fragment ions by an analyzer and detector, respectively. The resulting
mass spectrum is distinctive for every compound under standard conditions. This
mass spectral pattern can be used to identify the compounds by comparing it with
the library of the mass spectrum. Electron ionization with quadruple analyzer is
the most common instrumental setup in mass spectrometer of a typical GC–MS
in the analysis of essential oils. Advances in GC–MS have led to improvement
6 TERPENOIDS FROM ESSENTIAL OILS

in the separation technique of GC along with superior detection limits and data
handling tools of mass spectrometer (26).

5. Identification of Terpenoids

5.1. Identification of Mono- and Sesquiterpenoids. Until a few


decades earlier, structure elucidation of natural products was an exceedingly
complex and tedious task by carrying out various synthetic and degradation
reactions. Clues obtained by these experiments were taken together to determine
the structures. With the advent of spectroscopic instruments, natural product
structure elucidation becomes facile. Recent developments in commercial spec-
troscopic instrumentation pave the way for the structural elucidation of even a
little sample of 1 mg in 24 h or less time (27).
5.2. Linear Retention Index. Analysis of MSTs includes separation,
quantification, and identification of the compounds. In any chromatographic
system, a compound is identified based on its retention data. The retention time
of a compound in gas chromatography often varies due to change in instrumental
conditions. As gas chromatographic analysis is performed on different polarity
stationary phases with different flow rates and temperature programs, retention
time cannot be the same all the time and the value will not represent a compound.
To overcome this incongruity, Kovats in 1958 developed an equation to determine
the retention index calculations using a series of n-alkane hydrocarbons in
isothermal conditions, and it was further extended for the temperature gradient
conditions by van den Dool and Kratz in 1963 (28).
Retention index aids in the identification of a constituent through the
known retention time of n-alkane hydrocarbons in the respective column
phase. Representation of gas chromatographic retention index is an important
parameter in any standardized method. Thus, the retention index calculated by
this method remained the same for the particular column in all temperature
programs. The retention index (RI) can be calculated using the following formula
(28,29):
[ ]
tr − tn
RI = 100n + 100
tn+1 − tn

where
n = the number of hydrocarbons eluting before the peak of interest
tr = retention time of the unknown peak
tn = retention time of the hydrocarbon eluting before the peak of interest
tn+1 = retention time of the hydrocarbon eluting after the peak of interest

5.3. Mass Spectrometry Analysis. Retention time or mass spectral


data is not sufficient for the identification of the compounds present in essential
oils. Mass spectrometry is well known for its identification power, compared
to nonspectroscopic methods, but in many terpenoid compounds, especially
sesquiterpene hydrocarbons, having a similar mass spectrum leads to incorrect
identification of the compounds. A conventional detector such as flame ionization
TERPENOIDS FROM ESSENTIAL OILS 7

detector (FID) can give a well-accepted response relationship but lack in reliable
identification of the compounds. For accurate identification of the compounds,
reproducible retention time (retention index) of FID can be coupled with the
specificity of the mass spectrum. This confirms the identity of the compound
based on two independent parameters (26). Various groups have been working
on the development of retention index for essential oil compounds, and the data
compiled by Adams (30,31), which is a combination of retention data with mass
spectrometry, is well accepted and widely followed by most of the laboratories
working on essential oils worldwide.

6. Isolation and Characterization of MSTs

For the isolation of MSTs, essential oils can be fractionated by column chro-
matography (CC) on silica gel or alumina stationary phase and common
solvents used for the mobile phase as eluent are pentane/hexane/diethyl ether/
dichloromethane/ethyl acetate, etc. The isolation and purification of the CC
fractions can be monitored based on their TLC with anisaldehyde as a visualizing
agent, and the purity of the fractions can be assured by gas chromatography.
The presence of the functional group can be ascertained by using infrared
spectroscopy, whereas mass spectrometry will provide the basic structural infor-
mation of the compound. To date, NMR spectroscopy is the eminent technique to
predict and confirm the structure of compounds. The 1 H NMR, 13 C NMR, DEPT,
COSY, HMBC, HSQC spectra of diosphenolene are shown in Figure 4.

7. Biological Activity of Mono- and Sesquiterpenoids

Terpenoids are well known for their antibacterial, antifungal, and antiviral prop-
erties. They are used as a therapeutic agent to prevent the attack of many human
pathogens. Activities of terpenoids as biopreservatives, air disinfectants, water
disinfectants, against skin infections, dental bacteria, and so on have also been
reported (32–34).

8. Plant Protection

Terpenoids are receiving great attention in crop protection as synthetic pesticides


cause long-term health effects and environmental pollution. Natural pesticides
obtained from both plants and microbes are getting great demand in Europe and
North America. The Food Quality Protection Act of 1996 restricts the use of syn-
thetic insecticides, which considerably increases the research and market oppor-
tunity for natural pesticides. Terpenoids may be good substitutes for synthetic
pesticides, as they are well known to protect the plants from microbial infec-
tions and repel the pests. Terpenoids such as α-terpineol, (+)-terpinen-4-ol, and
carvacrol along with eugenol are reported to possess activity against insects and
two-spotted spider mite (35).
(a) (b) (c)

(d) (e)
1 13
Fig. 4. Representative NMR spectra of monoterpenoid compound. (a) H NMR spectrum; (b) C NMR spectrum; (c) DEPT 135∘ -NMR
spectrum; (d) HSQC–NMR spectrum; (e) HMBC–NMR spectrum.
TERPENOIDS FROM ESSENTIAL OILS 9

9. Enantiomers of Mono- and Sesquiterpenoids

Specific three-dimensional structures of enzymes create chiral centers and lead to


various stereoisomers in terpene biosynthesis. Many terpenoid compounds have
more than one chiral centers resulting in a number of diastereomers and their
corresponding enantiomers.
The enantioselective nature of terpenoids is well known because of their
diverse odor and flavor perception. Odor sensation and odor threshold may vary
to a large extent between the enantiomers, and most enantiomers differ in odor
qualities (36). For example, (R)-(−)-carvone has an odor of spearmint, whereas
(S)-(+)-carvone smells like caraway. Similarly, (R)-(−)-linalool have a character-
istic smell of lavender and (S)-(+)-linalool is of coriander (Fig. 5). Terpenoids are
found in almost all types of plants in different chemical proportions. Although the
composition of essential oils differs, enantiomeric excess of a predominant enan-
tiomer always falls in a very narrow range (37–39), which is a distinctive feature
for a particular aromatic plant.
For example, menthol always exists as laevorotatory, (−)-menthol in
both peppermint (Mentha piperita) and cornmint (Mentha arvensis). Similarly,
the most popular European bergamot oil (Citrus bergamia) contains only
(R)-(−)-linalool in its essential oil (40). (R)-(+)-Limonene has an odor of orange,
whereas (S)-(−)-limonene smells like either turpentine or lemon (Fig. 6) (41).

10. Importance of Monoterpenoids and Sesquiterpenoids

MSTs are having a high commercial value in perfumery, cosmetics, food technol-
ogy, pharmaceuticals, phytomedicine, and aromatherapy, etc. These compounds
are also used as a flavoring agent in food, beverages, and confectionary. Further-
more, terpenoids help in preserving food materials against contagious/infectious
pathogens. Besides, perfumes, fine fragrances, personal care, and household

HO OH

(R) (S)

(R)-(–)-Linalool (S)-(+)-Linalool
Fig. 5. Enantiomers of linalool.

(4S)-(–)-Limonene (4R)-(+)-Limonene
Fig. 6. Structure of limonene enantiomers.
10 TERPENOIDS FROM ESSENTIAL OILS

(R) (S) (R) (S) (R) (S)

O (Z) O (Z)

(4R,4αS,6R)-(+)-Nootkatone (4S,4αR,6S)-(–)-Nootkatone
Fig. 7. Enantiomers of nootkatone.

products have essential oils as an important ingredient of their composition.


The world trade of flavor and fragrance compounds increased about fourfold
from 1986 to 2002, showing the current demand for these flavor and fragrance
compounds (42). Natural flavor has a great demand in the food and beverages
because natural flavors cannot be replaced by artificial alternatives. The aroma
chemicals market is currently valued at ∼US$ 4.6 billion, and is estimated to
expand at a CAGR of ∼6 % to reach ∼US$ 7.8 billion by the end of 2027 (43).
Essential oils are the major source of terpenoid compounds that are used for
various ailments (44).
The odor threshold difference was well explained using the enantiomers
of nootkatone (Fig. 7), a sesquiterpenoid ketone of the eremophilane series.
(4R,4αS,6R)-(+)-Nootkatone has a grapefruit odor, while its enantiomer
(4S,4αR,6S)-(−)-nootkatone is woody and spicy. The odor threshold concentration
for (4R,4αS,6R)-(+)-nootkatone in the vapor phase is 30 ppm, but its enantiomer
is observed with 2200-fold and found to be approximately 66,000 ppm (45).
In a forest ecosystem, a group of aroma compounds emitted mainly by vari-
ous tree species constituting a galaxy of organic compounds such as terpenoids are
called phytoncides. Phytoncides are reported to have relaxing effects such as for-
est bathing, have antibacterial activity, and reduce sympathetic responsiveness
under restraint stress (46).
In some cases, a terpenoid mixture composed of α-pinene, β-pinene, cam-
phene, 1,8-cineole, menthone, borneol, and menthol (Fig. 8) is being used for
the treatment of cholesterol stones in the gall bladder and bile ducts. Similarly,
another composition of α-pinene, camphene, and menthol has therapeutic impor-
tance in the treatment of urethral stones. Formulations comprising cajeput oil,
eucalyptus oil, and mentha oil are used in aerosol therapy of the upper respiratory
tract. The mixture of marjoram oil, pine oil, eucalyptus oil, turpentine oil, and
camphor is applied in bronchitis treatment. Camphor-rich Oleum camphoratum
is used for the treatment of muscular aches and pains (44).
(−)-trans-β-Elemene isolated from Curcuma wenyujin and Rhizoma
zedoariae was reported to have antitumor activity, anticancer activity against

HO HO

(+)-Menthol (–)-Menthol
Fig. 8. Enantiomers of menthol.
TERPENOIDS FROM ESSENTIAL OILS 11

(Z) (Z)
(R)
(R)

(S)
(R) HO
(R)
(–)-α-Pinene
(S)-cis-Verbenol
Ips typographus

Microbial transformation
(Z)
(S) (Z)
(S)
(S)
(S) OH
(S)
(+)-α-Pinene (S)-trans-Verbenol

Fig. 9. Transformation of α-pinene to verbenol by Ips typographus.

brain, liver, breast, lung tissues, and also for leukemia (47). (R)-(+)-Pulegone, a
major compound present in the essential oil of Mentha pulegium, was reported to
delay the action potential repolarization with respect to its concentration (48).
10.1. MSTs as Semiochemicals. Semiochemicals are either a single
chemical compound or a mixture of compounds, which involve in the commu-
nication between individuals of the same species or between different species
of plants and animals. Semiochemicals derive specificity from asymmetry, and
this has made the study of chirality an essential objective in semiochemicals.
A mixture of (1S,5S)-α-pinene and (R)-limonene along with the aggregation
pheromone increases attractiveness to the adults of spruce bark beetle, Ips
typographus, whereas (1R,5R)-α-pinene acts as a repellent (49). Bark beetles
such as Tomicus piniperda, Hylurgops palliatus, and Trypodendron domesticum
have been known to contain (1S)-δ-3-carene and terpinolene. In H. palliatus,
such attraction was improved by the addition of ethanol; however, addition of
(1S,5S)-verbenone showed reduction in this attraction in T. domesticum (50).
The female spruce bark beetle, I. typographus, is primarily attracted by
its host tree Picea abies (L.) Karst. and the beetle transforms (−)-α-pinene to
cis-verbenol, which acts as an active constituent in the aggregation of pheromone
thereby, attracting more males and females, also aiding in suitable breeding host.
When the beetles were exposed to chemicals such as (+)-a-pinene, trans-verbenol
(Fig. 9) was formed, which was not a component of aggregation pheromone
and acted complementary (inhibited the aggregation) in combination with
verbenone (51).

11. Distribution of Terpenoid Enantiomers in Indian Aromatic Plants

This article discusses the enantioselective GC method that was successfully


applied for the mirror-image separation and resolution of chiral terpenoids in
various essential oils in a single run. The enantiomers of chiral terpenoids in
various essential oils studied in the past 10 years in the author’s laboratory CS
12 TERPENOIDS FROM ESSENTIAL OILS

OH OH
(1S)-(–)-Borneol (1R)-(+)-Borneol
Fig. 10. Enantiomers of borneol.

6000
5500
5000

(1S)-(–)-Camphor
4500
4000

(1S)-(–)-Borneol
(1R)-(+)-Camphor
3500
3000

(1R)-(+)-Borneol
2500
2000
1500
16.40

1000

20.38
500
0 RT [mm]

13 14 15 16 17 18 19 20 21 22 23
Fig. 11. Chromatogram showing the separation of camphor and borneol racemates in
Artemisia pleiocephala essential oil.

Chanotiya at CSIR-CIMAP, India, are discussed individually as per the details


given under:
11.1. Borneol Enantiomers. Borneol is an important monoterpenoid
alcohol involved as a precursor in the biosynthesis of camphor (Figs. 10 and
11). Ravid and co-workers (52) studied the resolution of borneol racemates in
permethylated-β-cyclodextrin and found that, similar to camphor, borneol was
also observed with a variable enantiomeric ratio in Salvia officinalis L. plant
samples collected from different places. Lavandin essence from France, lavandin
oil from Israel, and lavender “Fine” oil from Switzerland all were observed
with a high enantiomeric ratio of (1R)-(+)-borneol. Essential oils extracted from
Coridothymus capitatus (L.), Origanum vulgare L. oil from Greece, Origanum
syriacum (L.), Mentha longifolia (L.) from Negev Israel, Ocimum canum Sims.
from Thailand, Chrysanthemum parthenium (L.) (feverfew) and Valerian oil
were all having (1S)-(−)-borneol as the major enantiomer with more than 90%
enantiomeric excess (Table 1). Further, the commercial samples of camphor
tree oil, white, rectified, Achillea fragrantissirma (Forsk.). Negev, Israel, and
citronella oils were observed with racemic forms of borneol enantiomers (52).
11.2. Camphor Enantiomers. Due to its importance as a chiral starting
material, enantiomers of camphor (Fig. 12) are always in great demand in
pharmaceutical, biological, and fine chemical applications (54). (−)-Camphor
has been reported as the starting material for many enantioselective reac-
tions such as (−)-laurolenic acid, steroid intermediates, and vitamin B12 ,
TERPENOIDS FROM ESSENTIAL OILS 13

Table 1. Enantiomeric Distribution of Borneol in Different Plantsa from India


(53)
Plants Family Borneol
(+)- (−)-
Coriandrum sativum Apiaceae 12.8 87.2
Artemisia pleiocephala Asteraceae 4.9 95.1
Artemisia annua Asteraceae 34.1 65.9
Ocimum canum Lamiaceae 6 94
Ocimum kilimandscharicum Lamiaceae 8.5 91.5
Cinnamomum camphora Lauraceae 63.6 37.4
Zingiber officinale Zingiberaceae 94.3 5.7
a Source:
VS Pragadheesh, C.S. Chanotiya, V.S. Pragadheesh Thesis, AcSIR, CSIR-CIMAP,
Lucknow, India.

O O
(1S)-(–)-Camphor (1R)-(+)-Camphor
Fig. 12. Enantiomers of camphor.

whereas (+)-camphor has been used in (−)-oestrene synthesis, carotenoid,


and terpenoid intermediates. In a previous report, camphor enantiomers
were characterized on octakis-(3-O-butyryl-2,6-di-O-pentyl)-γ-cyclodextrin chi-
ral phase in the family of Lamiaceae alone where S. officinalis L. essential
oil possessed both enantiomeric forms with a variable ratio in accessions
of different places. C. parthenium (L.) Bernh. (feverfew) oil contained 100%
(S)-(−)-camphor, and Artemisia judaica L. oil had (S)-(−)-camphor around
61% and 69% in collections from Sinai of Egypt and Negev of Israel (55). In
another study, camphor enantiomers were characterized using 30% heptakis
(2,3-di-O-methyl-6-O-t-butyl dimethylsilyl)-β-cyclodextrin in DB-1701 chiral
stationary phase in “over-the-counter” pharmaceutical products and compared
with the enantiomeric ratio of the ingredient essential oils such as Cinnamomum
camphora to verify the natural or synthetic origin of the active ingredients (56).
Lavandin essential oils of Spanish origin contained (S)-(−)-camphor with more
than 99% enantiomeric excess (57). Camphor racemates characterized in another
study showed (1R)-(+)-camphor with more than 90% ee in the samples of Salvia
sclarea and Ocimum basilicum from Italy and more than 50% ee in S. officinalis
oil from France, while Coriandrum sativum samples from Italy had >70% ee for
(1S)-(−)-camphor (58). Camphor enantiomers in plants of Indian origin are given
in Table 2.
11.3. Citronellol Enantiomers. Citronellol enantiomers also exist as a
mixture of enantiomers in most studied essential oils (60). Citronellol in geranium
has been observed with almost racemic form, whereas (+)-citronellol is a major
enantiomer in citronella and (−)-citronellol is a major enantiomer in rose (Table 3,
Figs. 13 and 14).
Table 2. Enantiomeric Distribution of Camphor in Plant Essential Oilsa (53,59)
Plants Family Camphor
(+)- (−)-
Coriandrum sativum Apiaceae 87.4 12.6
Artemisia pleiocephala Asteraceae 84.8 15.2
Artemisia annua Asteraceae 94.2 5.8
Ocimum canum Lamiaceae >99 t
Ocimum kilimandscharicum Lamiaceae 99.5 0.5
Cinnamomum camphora Lauraceae 99.4 0.6
a Source:
V.S. Pragadheesh, C.S. Chanotiya, V.S. Pragadheesh Thesis, AcSIR, CSIR-CIMAP,
Lucknow, India; Pragadheesh and co-workers (59).

Table 3. Enantiomeric Distribution Citronellol in Plant Essential Oilsa (53)


Plants Family Citronellol
(+)- (−)-
Pelargonium graveolens 1 Geraniaceae 42.2 57.8
Pelargonium graveolens 2 Geraniaceae 41.2 58.8
Cympbopogon winterianus 1 Poaceae 83.5 16.5
Cympbopogon winterianus 2 Poaceae 83.9 16.1
Rosa damascena cv. Ranisahiba Rosaceae 7.9 92.1
a Source:
V.S. Pragadheesh, C.S. Chanotiya, V.S. Pragadheesh Thesis, AcSIR, CSIR-CIMAP,
Lucknow, India.

CH2OH CH2OH

(R)-(+)-Citronellol (S)-(–)-Citronellol
Fig. 13. Enantiomers of citronellol.

150
(R)-(+)-Citronellol
(S)-(–)-Citronellol

Standard mixture
100
uVolts

50

750 (S)-(–)-Citronellol in
R. damascena
RaniSahiba
500
uVolts

250

20.0 22.5 25.0 27.5 30.0 32.5 35.0 Minutes


Fig. 14. Separation of citronellol enantiomers in rose oil Rosa damascena cv. Ranisahiba
developed by CSIR-CIMAP.
TERPENOIDS FROM ESSENTIAL OILS 15

11.4. Linalool Enantiomers. Enantiomeric distribution of linalool in


several plants from different origin was studied, where Lavandula angusti-
folia (France), Lavandin grosso (France), Citrus aurantium Bergamia (Italy
and Ivory Coast), Cananga odorata (Madagascar), Michelia alba, Verbena
officinalis (France), Eucalyptus citriodora (Madagascar), Thymus vulgaris
(France), Thymus zygis (Spain and Portugal), O. basilicum (France), Satureja
hortensis (France), Thymus serpyllum (France, Indonesia, and China), S. sclarea
(France), Artemisia dracunculus (France), and Lonicera caprifolium (France)
were observed with more than 90% enantiomeric ratio of (R)-linalool. Plants
such as Elettaria cardamomum (Sri Lanka, China, and India), Citrus reticulata
(France), and Robina flowers were found with more than 90% enantiomeric ratio
of (S)-linalool. Essential oils of Rosa damascena (Bulgaria and Turkey), Citrus
limon (Spain), Citrus paradisi (Israel), Aniba rosaeodora (Brazil), Cinnamomum
zeylanicum (Sri Lanka), and Pelargonium graveolens (Egypt, Morocco, Tunisia,
and France) constitute linalool as a racemic mixture (61,62). Cinnamomum
tamala leaf essential oil from India was found to contain (3S)-(+)-linalool with
100% enantiomeric purity (39).
In another study, (R)-(−)-linalool was identified from lavandin varieties from
Spain with an enantiomeric excess of 93.6–95.4% (57). (R)-(−)-Linalyl acetate of
lavandin essential oils from Spain was also determined by Casabianca et al. (61) to
be 96.7–100% enantiomeric excess. Distribution of linalool enantiomers of Indian
origin is given in Table 4.

Table 4. Enantiomeric Distribution of Linalool in Plant Essential Oilsa (53)


Plants Family Linalool
(+)- (−)-
Acorus calamus Acoraceae 1.1 98.9
Coriandrum sativum Apiaceae 87.4 12.6
Pelargonium graveolens 1 Geraniaceae 49.9 50.1
Pelargonium graveolens 2 Geraniaceae 49.8 50.2
Ocimum basilicum Lamiaceae t > 99
Ocimum kilimandscharicum Lamiaceae 70.3 29.7
Origanum majorana Lamiaceae 59.4 40.6
Salvia sclarea Lamiaceae 28.6 71.4
Pimenta racemosa Myrtaceae 6.7 93.3
C. winterianus 1 Poaceae 62.7 37.3
C. winterianus 2 Poaceae 64.1 35.9
Rosa damascena Rosaceae 49.7 50.3
Citrus medica Rutaceae 92.5 7.5
Citrus reticulata Rutaceae 83.2 16.8
Citrus sinensis Rutaceae 90.0 10.0
Citrus jambhiri Rutaceae 7.7 92.3
Lippia alba Verbenaceae >99 nd
Hedychium coronarium Zingiberaceae >99 <1.0
Zingiber officinale Zingiberaceae 48.9 51.1
Zingiber roseum Zingiberaceae 86.0 14.0
a Source:
V.S. Pragadheesh, C.S. Chanotiya, V.S. Pragadheesh Thesis, AcSIR, CSIR-CIMAP,
Lucknow, India.
16 TERPENOIDS FROM ESSENTIAL OILS

OH OH

(S)-(+)-Terpinen-4-ol (R)-(–)-Terpinen-4-ol
Fig. 15. Enantiomers of terpinen-4-ol.

11.5. Monoterpene Hydrocarbons. Monoterpene hydrocarbons,


namely, α-pinene, β-pinene, camphene, sabinene, limonene, phellandrene, etc. are
the common compounds identified in almost all types of essential oils (Table 5).
Among these compounds, α-pinene, β-pinene, and sabinene usually exist as a
mixture of enantiomers, whereas limonene occurs as a single enantiomer in
many essential oils (63–66).
11.6. Menthol Enantiomers. Menthol is one of the main flavor com-
ponents in confectionery. It occurs as two enantiomers, and in plants, only
(−)-menthol is biosynthesized with high enantiomeric purity. Menthol race-
mates were resolved using various chiral stationary phases such as octakis-
(6-O-methyl-2,3-di-O-pentyl)-γ-cyclodextrin (67), permethylated-β-cyclodextrin,
2,3-di-acetoxy-6-O-tert-butyldimethylsilyl-γ-cyclodextrin (68), and 2,3-di-O-
methyl-6-O-t-butyldimethylsilyl)-β-cyclodextrin (57). Even though the menthol
content was reported to vary up to a large extent due to different agroclimatic
zones of a country (69,70), but in the literature, menthol was reported as a single
enantiomer, (−)-menthol all over the world (40,67,68,70).
11.7. Terpinen-4-ol Enantiomers. Terpinen-4-ol is one of the common
monoterpenoid compounds present in several plants. In most cases, terpeinen-4-ol
is found as an enantiomeric mixture (Fig. 15) rather than as a single enantiomer.
Reports show that lavender essential oil contains (S)-(+)-terpinen-4-ol as the
major enantiomer with optical purity more than 97%, whereas the rest of the
essential oils contained ∼30% enantiomeric excess of either (S)-(+)-terpinen-4-ol
or (R)-(−)-terpinen-4-ol. Juniper berry and Oregano essential oil (thymol type)
were found with almost racemic compositions (72). Another report used lan-
thanide shift reagents in NMR to identify the enantiomeric nature of terpinen-4-ol
and found that both enantiomers are present in Origanum majorana in various
ratios (73). The enantiomeric ratio of terpinen-4-ol enantiomers is given in
Table 6.
11.8. 𝛂-Terpineol Enantiomers

OH OH

α-Terpineol is an intermediate compound formed from linalyl cation and


act as a precursor for most of the acyclic, cyclic, and bicyclic monoterpenoids.
Table 5. Enantiomeric Distribution of Monoterpene Hydrocarbons in the Essential Oilsa (53,71)
Plants α-Pinene β-Pinene Camphene Sabinene Limonene
(+)- (−)- (+)- (−)- (+)- (−)- (+)- (−)- (+)- (−)-
Apiaceae
Daucus carota nd >99 82.9 17.1 nd >99 8.9 91.1 81.5 18.5
Coriandrum sativum 95.2 4.8 >99 nd 32.7 67.3 67.6 32.4 49.1 50.9
Asteraceae
Artemisia pleiocephala 76.1 23.9 26.0 74.0 >99 nd – – – –
Artemisia annua >99 nd 15.8 84.2 98 2.0 – – 47.6 52.4
Tagetues erecta – – – – – – – – >99 nd
Myrtaceae
Pimenta racemosa – – – – – – – – 29.9 70.1
Poaceae
Cymbopogon winterianus 1 – – – – – – – – nd >99

17
Cymbopogon winterianus 2 – – – – – – – – nd >99
Rosaceae
Rosa damascena nd >99 28.7 71.3 – – – – – –
Rutaceae
Citrus medica – – – – – – – – 98.5 1.5
Citrus reticulata >99 nd 59.5 40.5 – – 98.0 2.0 67.0 33.0
Citrus sinensis 99.5 0.5 62.0 38.0 – – 97.9 2.1 62.4 37.6
Citrus jambhiri 5.2 94.8 0.4 99.6 – – 10.1 89.9 95.0 5.0
Murraya koenigii >99 t: trace (<0.1%) 73.6 26.4 – – 87.4 12.6 – –
Zanthoxylum armatum – – – – – – 11.6 88.4 – –
Aegle marmelos nd >99 – – – – – – 28.0 72.0
a Source: V.S. Pragadheesh, C.S. Chanotiya, V.S. Pragadheesh, Thesis, AcSIR, CSIR-CIMAP, Lucknow, India; Pragadheesh and co-workers (71).
18 TERPENOIDS FROM ESSENTIAL OILS

Table 6. Enantiomeric Distribution of Terpinen-4-ol in Plant Essential Oils


Plants Family Terpinen-4-ol
(+)- (−)-
Artemisia pleiocephala Asteraceae 65.7 34.3
Artemisia annua Asteraceae 50.4 49.6
Ocimum canum Lamiaceae 59.4 40.6
Ocimum kilimandscharicum Lamiaceae 84.2 15.2
Origanum majorana Lamiaceae 75.1 24.9
Cinnamomum camphora Lauraceae 45.8 54.2
Citrus reticulata Rutaceae 74.4 25.6
Citrus sinensis Rutaceae 74.6 25.4
Citrus jambhiri Rutaceae 27.2 72.8
Murraya koenigii Rutaceae 68.9 31.1
Zingiber officinale Zingiberaceae 84.8 15.2

Table 7. Enantiomeric Distribution of 𝛂-Terpineol in Plant Essential Oils


Plants Family α-Terpineol
(4R)-(+)- (4S)-(−)-
Coriandrum sativum Apiaceae 56.6 43.4
Artemisia pleiocephala Asteraceae 44.2 55.8
Origanum majorana Lamiaceae 59.1 40.9
Cinnamomum camphora Lauraceae 25.8 74.2
Citrus sinensis Rutaceae 88.2 11.8
Citrus jambhiri Rutaceae 8.9 91.1
Lippia alba Verbenaceae 16.8 83.2
Zingiber officinale Zingiberaceae >99 nd

α-Terpineol also exists as a mixture of enantiomers in most of the Indian essential


oils as shown in Table 7.

12. Future Prospects

Terpenoids have vast applications in our daily life. The present-day market keeps
exploring the prospects of the products in which terpenoids are known to play a
crucial role directly or indirectly. For example, the Indian mint market is one of the
major areas. India is the largest producer of M. arvensis oil including bi-products
(menthol, de-mentholated oils) (74). Menthol is an oxygenated monoterpene pos-
sessing diverse applications. The popularity of mint in consumer goods, partic-
ularly in oral care, has been expanding into the genesis of new products with
distinct flavor characteristics. In fragrance, mint offers a green and herbal quality
to formulations, while imparting a cool, crisp, and refreshing scent profile. With
the consumers’ preference toward unique fragrances and the rise of aromather-
apy, essential oils have seen a boom in the industry for their perceived calming
effects (75).
In India, economically motivated adulteration has been a serious concern
for the past two decades. Synthetic ingredients are intentionally added to con-
sumer products to reduce the cost of production or to increase the economic gain.
TERPENOIDS FROM ESSENTIAL OILS 19

The absence of target-based quality control methods and guidelines help in flour-
ishing this menace with more prominence in larger areas. ASTM D6866-18 is an
international standard recommended for determining the biobased content using
radiocarbon analysis. ISO 16620-2, equivalent to ASTMD6866-18, is standard-
ized as an ideal tool for producers/industry who face the challenge of verifying the
origin of essential oils. Since living carbon-based materials, such as plants, have
a known C14 level and will yield a biobased result of 100%. Compounds synthe-
sized from fossil feedstock do not contain C14 radioisotope because C14 obeys the
laws of radioactive decay. Carbon14 disintegrates completely within 50,000 years.
Petroleum-based aroma chemicals are, therefore, 0% bio-based.
In highly regulated marketplaces where mislabeling could have legal and
commercial repercussions, a reliable test like chiral pair differentiation may pro-
vide a quick solution for high-value essential oils, where the marker compound
exists in high optical purity and small ranges of enantiomeric excess values.

13. Conclusions

Regarding the distribution of camphor enantiomers, genus Ocimum and Cin-


namomum camphora contained >99% of single enantiomer (+)-camphor in Indian
oils. The distribution of (+)-citronellol was high in citronella oil, whereas rose
cv. Ranisahiba contained (−)-citronellol with high (>92%) enantiomeric ratio.
(+)-Linalool was highest (>90% or above) in Hedychium coronarium, Lippia
alba, Citrus medica, Citrus sinensis, etc. On the contrary, O. basilicum and
Acorus calamus were the unique oils since (−)-linalool was observed as the most
exclusive enantiomer. Similarly, terpinen-4-ol and α-terpineol were recorded as
racemic mixtures except for Zingiber officinale, which possessed (+)-α-terpineol
as a major enantiomer. The Indian mint species have not been a subject of
chiral discrimination studies to date. We have identified limonene, menthone,
menthol, and menthyl acetate as a single image represented by (−)-enantiomers,
whereas isomenthone, neomenthol, pulegone, and piperitone were recorded as
pure (+)-enantiomers in menthol mint and peppermint oils, respectively (70).
In conclusion, the presence of both enantiomers of single terpene in Indian
essential oils has been observed as a common feature for most of the monoterpene
hydrocarbons (70). The chiral pairs of high optical purity and enantiomeric excess
values can be used to authenticate the origin of essential oils or to detect the
misbranding of natural and synthetic chiral components. This criterion is met
for components of Indian menthol mint and Indian peppermint essential oils. If
essential oils show vast variations in their optical purity profile, then every chiral
compound must be studied enantioselectively in terms of enantiomeric excess
for delineation of geographical origin and detection of undesired components.
However, several limitations on the chiral differentiation of essential oil exist.
If the compounds (chiral and achiral both) synthesized using petroleum are the
potential adulterants, then C14 -based radiocarbon dating by using accelerator
mass spectrometry method will probably be the best technique of modern time.
Moreover, chiral differentiation using gas chromatography can be used for
the regulatory purposes in quality control studies of commercially important
essential oils rich with compounds of high optical purity.
20 TERPENOIDS FROM ESSENTIAL OILS

REFERENCES
1. B. Weber and co-workers, J. Essent. Oil Res. 18(6), 607–610 (2006). DOI: 10.1080/
10412905.2006.9699180
2. K. Husain and co-workers, Phytochem. Lett. 5(4), 788–792 (2012). DOI: 10.1016/
j.phytol.2012.09.003
3. B. M. Fraga, Nat. Prod. Rep. 30, 1226–1264 (2013). DOI: 10.1039/C3NP70047J
4. I. L. Finar, Organic Chemistry, 3rd ed., Vol. 242, Longmans, Green and Co. Ltd., 1964.
5. V. Kulcitki, P. Harghel, and N. Ungur, Nat. Prod. Rep. 31, 1686–1720 (2014). DOI:
10.1039/C4NP00081A
6. B. M. Fraga, Nat. Prod Rep. 2, 147–161 (1985). DOI: 10.1039/NP9850200147
7. B. M. Fraga, Nat. Prod. Rep. 26, 1125–1155 (2009). DOI: 10.1039/B908720F
8. J. Gershenzon and N. Dudareva, Nat. Chem. Biol. 3(7), 408–414 (2007). DOI: 10.1038/
nchembio.2007.5
9. Z.-J. Zhan and co-workers, Nat. Prod. Rep. 28, 594–629 (2011). DOI: 10.1039/
C0NP00050G
10. C. S. Mathela, R. C. Padalia, and C. S. Chanotiya, Nat. Prod. Commun. 4(9), 1253–1256
(2009).
11. C. S. Mathela and co-workers, Oil Res. 17, 672–675 (2005). DOI: 10.1080/10412905.
2005.9699029
12. L. Del Giudice and co-workers, Environ. Microbiol. 10(10), 2824–2841 (2008). DOI:
10.1111/j.1462-2920.2008.01703.x
13. S. Schulz and J. S. Dickschat, Nat. Prod. Rep. 24, 814–842 (2007). DOI: 10.1039/
B507392H
14. J. Degenhardt, T. G. Köllner, and J. Gershenzon, Phytochemistry 70(15–16), 1621–1637
(2009). DOI: 10.1016/j.phytochem.2009.07.030
15. A. C. Figueiredo and co-workers, Flavour Fragr. J. 23(4), 213–226 (2008). DOI:
10.1002/ffj.1875
16. V. S. Pragadheesh, A. Yadav, and C. S. Chanotiya, J. Essent. Oil. Res. 27(2), 101–106
(2015). DOI: 10.1080/10412905.2014.987929
17. V. S. Pragadheesh and co-workers, Nat. Prod. Commun. 8(2), 221–224 (2013).
18. V. S. Pragadheesh and co-workers, Ind. Crop. Prod. 49, 628–633 (2013). DOI:
10.1016/j.indcrop.2013.06.023
19. H. Sovová and S. A. Aleksovski, Flavour Fragr. J. 21(6), 881–889 (2006). DOI:
10.1002/ffj.1729
20. A. C. Soria, J. Sanz, and I. Martínez-Castro, Eur. Food Res. Technol. 228(4), 579–590
(2009). DOI: 10.1007/s00217-008-0966-z
21. C. Bicchi, C. Cordero, and P. Rubiolo, J. Chromatogr. Sci. 42(8), 402–409 (2004). DOI:
10.1093/chromsci/42.8.402
22. V. S. Pragadheesh and co-workers, Nat. Prod. Commun. 6(9), 1333–1338 (2011).
23. J. T. Romeo, J. Chem. Ecol. 35(12), 1383 (2009). DOI: 10.1007/s10886-009-9733-2
24. IOFI Working Group on Methods of Analysis, Flavour Fragr. J. 25(404–406) (2010).
DOI: 10.1002/ffj.1991
25. C. Bicchi and co-workers, Flavour Fragr. J. 23(6), 382–391 (2008). DOI: 10.1002/ffj.1905
26. P. J. Marriott, R. Shellie, and C. Cornwell, J. Chromatogr. A 936(1–2), 1–22 (2001).
DOI: 10.1016/S0021-9673(01)01314-0
27. R. C. Breton and W. F. Reynolds, Nat. Prod. Rep. 30, 501–524 (2013). DOI: 10.1039/
C2NP20104F
28. B. d.’A. Zellner and co-workers, Flavour Fragr. J. 23(5), 297–314 (2008). DOI:
10.1002/ffj.1887
29. M.-F. Herent, V. D. Bie, and B. Tilquin, J. Pharm. Biomed. Anal. 43(3), 886–892 (2007).
DOI: 10.1016/j.jpba.2006.09.005
TERPENOIDS FROM ESSENTIAL OILS 21

30. R. P. Adams, Identification of Essential Oil Components by Gas Chromatography/Mass


Spectroscopy, Allured Publishing, Carol Stream, IL, 1995.
31. R. P. Adams, Identification of Essential Oil Components by Gas Chromatography/Mass
Spectroscopy, Allured Publishing, Carol Stream, IL, 2007.
32. P. Suppakul and co-workers, J. Agric. Food Chem. 51(11), 3197–3207 (2003). DOI:
10.1021/jf021038t
33. K. Fujita, T. Fujita, and I. Kubo, Phytother. Res. 21(1), 47–51 (2007). DOI: 10.1002/
ptr.2016
34. G. Lang and G. Buchbauer, Flavour Fragr. J. 27(1), 13–39 (2012). DOI: 10.1002/ffj.2082
35. M. B. Isman, Crop Protection 19(8–10), 603–608 (2000). DOI: 10.1016/S0261-2194(00)
00079-X
36. W. A. König, in W. J. Lough and I. W. Wainer, eds. Natural and Applied Sciences,
Oxford, UK, Blackwell Publishers, 2002, pp. 261–284.
37. S. Tamogami, K. Awano, and T. Kitahara, Flavour Fragr. J. 16(3), 161–163 (2001). DOI:
10.1002/ffj.969
38. C. S. Chanotiya and A. Yadav, Nat. Prod. Commun. 4(4), 563–566 (2009).
39. C. S. Chanotiya and A. Yadav, J. Essent. Oil Res. 22(6), 593–596 (2010). DOI:
10.1080/10412905.2010.9700407
40. W. A. Konig and co-workers, J. High Res. Chromatogr. 20(2), 55–61 (1997). DOI:
10.1002/jhrc.1240200202
41. R. Bentley, Chem. Rev. 106(9), 4099–4112 (2006). DOI: 10.1021/cr050049t
42. W. Schwab, R. Davidovich-Rikanati, and E. Lewinsohn, Plant J 54(4), 712–732 (2008).
DOI: 10.1111/j.1365-313X.2008.03446.x
43. https://www.transparencymarketresearch.com/aroma-chemicals-market.html
(accessed 12 August 2020).
44. D. Sybilska and M. Asztemborska, J. Biochem. Biophys. Meth. 54(1-3), 187–195 (2002).
DOI: 10.1016/S0165-022X(02)00141-0
45. H. G. Haring and co-workers, J. Agric. Food Chem. 20(5), 1018–1021 (1972). DOI:
10.1021/jf60183a011
46. K. Kawakami and co-workers, Clin. Exp. Pharmacol. Physiol. 31(s2), S27–S28 (2004).
DOI: 10.1021/jf60183a011
47. A. M. Adio, Tetrahedron 65(27), 5145–5159 (2009). DOI: 10.1016/j.tet.2009.04.062
48. A. Santos-Miranda and co-workers, Phytomedicine 21(10), 1146–1153 (2014). DOI:
10.1016/j.phymed.2014.05.007
49. J. Reddemann and R. Schopf, Entomol. Gener. 21(1–2), 69–80 (1996). DOI: 10.1127/
entom.gen/21/1996/69
50. J. A. Byers and J. Chem, Ecol. 18(12), 2385–2402 (1992). DOI: 10.1007/BF00984957
51. T. Norin, Pure Appl. Chem. 68(11), 2043–2049 (1996). DOI: 10.1351/pac199668112043
52. U. Ravid, E. Putievsky, and I. Katzir, Flavour Fragr. J. 11(3), 191–195 (1996). DOI:
10.1002/(SICI)1099-1026(199605)11:3<191::AID-FFJ568>3.0.CO;2-M
53. V. S. Pragadheesh, C. S. Chanotiya, and V. S. Pragadheesh Thesis, AcSIR,
CSIR-CIMAP, Lucknow, India.
54. M. Bhaya and R. Beniwal, Cardiovasc. Toxicol. 7, 212–214 (2007). DOI: 10.1007/s12012-
007-0029-x
55. U. Ravid, E. Putievsky, and I. Katzir, Flavour Fragr. J. 8(4), 225–228 (1993). DOI:
10.1002/ffj.2730080411
56. J. Fiori, M. Naldi, and R. Gotti, Chromatographia 72(9–10), 941–947 (2010). DOI:
10.1365/s10337-010-1735-2
57. G. Flores and co-workers, J. Sep. Sci. 28(17), 2333–2338 (2005). DOI: 10.1002/jssc.
200500124
58. F. Tateo and co-workers, Anal. Commun. 36, 149–151 (1999). DOI: 10.1039/A901388A
59. V. S. Pragadheesh and co-workers, Ind. Crop. Prod. 50, 333–337 (2013). DOI:
10.1016/j.indcrop.2013.08.009
22 TERPENOIDS FROM ESSENTIAL OILS

60. U. Ravid and co-workers, Flavour Fragr. J. 7(4), 235–238 (1992). DOI: 10.1002/ffj.
2730070413
61. H. Casabianca and co-workers, J. High Res. Chromatogr. 21(2), 107–112 (1998). DOI:
10.1002/(SICI)1521-4168(19980201)21:2<107::AID-JHRC107>3.0.CO;2-A
62. T. Ozek and co-workers, Rec. Nat. Prod. 4(4), 180–192 (2010).
63. D. W. Armstrong and co-workers, Chirality 8(1), 39–48 (1996). DOI: 10.1002/(SICI)
1520-636X(1996)8:1<39::AID-CHIR9>3.0.CO;2-B
64. C. Bicchi, C. Cagliero, and P. Rubiolo, Flavour Fragr. J. 26(5), 321–325 (2011). DOI:
10.1002/ffj.2059
65. W. A. Konig, Chirality 10(5), 499–504 (1998). DOI: 10.1002/(SICI)1520-636X(1998)10:
5<499::AID-CHIR13>3.0.CO;2-V
66. J. M. Finefield and co-workers, Angew. Chem. 51(20), 4802–4836 (2012). DOI:
10.1002/anie.201107204
67. V. S. Pragadheesh, A. Yadav, and C. S. Chanotiya, J. Chromatogr. B 1002, 30–41 (2015).
DOI: 10.1016/j.jchromb.2015.07.034
68. M. L. Ruiz del Castillo, G. P. Blanch, and M. Herraiz, J. Chromatogr. A 1054(1–2),
87–93 (2004). DOI: 10.1016/j.chroma.2004.08.055
69. S. Dwivedi and co-workers, Flavour Fragr. J. 19(5), 437–440 (2004). DOI: 10.1002/
ffj.1333
70. C. S. Chanotiya and co-workers, J. Essent. Oil Res., under review (2020).
71. V. S. Pragadheesh and co-workers, Med. Aromat. Plants 2(1), 1–3 (2013). DOI:
10.4172/2167-0412.1000118
72. U. Ravid and co-workers, Flavour Fragr. J. 7(1), 49–52 (1992). DOI: 10.1002/ffj.
2730070112
73. U. Ravid and co-workers, Flavour Fragr. J. 2(1), 17–19 (1987). DOI: 10.1002/ffj.
2730020104
74. S. Pringle, Perfumer Flavorist 43, 36 (2018).
75. D. Ataman, Perfumer Flavorist 43, 31 (2018).

VS PRAGADHEESH
Laboratory of Aromatic Plants and Chiral
Separation, CSIR-Central Institute of
Medicinal and Aromatic Plants, Lucknow,
India
CSIR-Central Institute of Medicinal and
Aromatic Plants (Research Centre),
Bengaluru, India
Deepa BISHT
IORA-RCSTT Coordination Centre on
Medicinal Plants (ICCMP), CSIR-Central
Institute of Medicinal and Aromatic
Plants, Lucknow, India
CS CHANOTIYA
Laboratory of Aromatic Plants and Chiral
Separation, CSIR-Central Institute of
Medicinal and Aromatic Plants, Lucknow,
India

You might also like