ArticlePDF Available

Ultrasound-assisted nanoemulsion of Trachyspermum ammi essential oil and its constituent thymol on toxicity and biochemical aspect of Aedes aegypti

Authors:

Abstract and Figures

Aedes aegypti is the main vector of yellow fever, chikungunya, Zika, and dengue worldwide and is managed by using chemical insecticides. Though effective, their indiscriminate use brings in associated problems on safety to non-target and the environment. This supports the use of plant-based essential oil (EO) formulations as they are safe to use with limited effect on non-target organisms. Quick volatility and degradation of EO are a hurdle in its use; the present study attempts to develop nanoemulsions (NE) of Trachyspermum ammi EO and its constituent thymol using Tween 80 as surfactant by ultrasonication method. The NE of EO had droplet size ranging from 65 ± 0.7 to 83 ± 0.09 nm and a poly dispersity index (PDI) value of 0.18 ± 0.003 to 0.20 ± 0.07 from 1 to 60 days of storage. The NE of thymol showed a droplet size ranging from 167 ± 1 to 230 ± 1 nm and PDI value of 0.30 ± 0.03 to 0.40 ± 0.008 from 1 to 60 days of storage. The droplet shape of both NEs appeared spherical under a transmission electron microscope (TEM). The larvicidal effect of NEs of EO and thymol was better than BEs (Bulk emulsion) of EO and thymol against Ae. aegypti. Among the NEs, thymol (LC 50 34.89 ppm) had better larvicidal action than EO (LC 50 46.73 ppm). Exposure to NEs of EO and thymol causes the shrinkage of the larval cuticle and inhibited the acetylcholinesterase (AChE) activity in Ae. aegypti. Our findings show the enhanced effect of NEs over BEs which facilitate its use as an alternative control measure for Ae. aegypti.
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
1 3
Environmental Science and Pollution Research
https://doi.org/10.1007/s11356-022-20870-2
RESEARCH ARTICLE
Ultrasound‑assisted nanoemulsion ofTrachyspermum ammi essential
oil andits constituent thymol ontoxicity andbiochemical aspect
ofAedes aegypti
KesavanSubaharan1 · PeriyasamySenthamaraiSelvan1· ThagareManjunathaSubramanya2·
RajendranSenthoorraja1· SowmyaManjunath1· TaniaDas1· VppalayamShanmugamPragadheesh3·
NandagopalBakthavatsalam1· MuthuGounderMohan1· SengottayanSenthil‑Nathan4· SreehariUragayala5·
PaulrajPhilipSamuel6· RenuGovindarajan6· MuthuswamyEswaramoorthy2
Received: 18 August 2021 / Accepted: 12 May 2022
© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2022
Abstract
Aedes aegypti is the main vector of yellow fever, chikungunya, Zika, and dengue worldwide and is managed by using
chemical insecticides. Though effective, their indiscriminate use brings in associated problems on safety to non-target and
the environment. This supports the use of plant-based essential oil (EO) formulations as they are safe to use with limited
effect on non-target organisms. Quick volatility and degradation of EO are a hurdle in its use; the present study attempts
to develop nanoemulsions (NE) of Trachyspermum ammi EO and its constituent thymol using Tween 80 as surfactant by
ultrasonication method. The NE of EO had droplet size ranging from 65 ± 0.7 to 83 ± 0.09nm and a poly dispersity index
(PDI) value of 0.18 ± 0.003 to 0.20 ± 0.07 from 1 to 60days of storage. The NE of thymol showed a droplet size ranging from
167 ± 1 to 230 ± 1nm and PDI value of 0.30 ± 0.03 to 0.40 ± 0.008 from 1 to 60days of storage. The droplet shape of both
NEs appeared spherical under a transmission electron microscope (TEM). The larvicidal effect of NEs of EO and thymol
was better than BEs (Bulk emulsion) of EO and thymol against Ae. aegypti. Among the NEs, thymol (LC50 34.89ppm) had
better larvicidal action than EO (LC50 46.73ppm). Exposure to NEs of EO and thymol causes the shrinkage of the larval
cuticle and inhibited the acetylcholinesterase (AChE) activity in Ae. aegypti. Our findings show the enhanced effect of NEs
over BEs which facilitate its use as an alternative control measure for Ae. aegypti.
Keywords Aedes aegypti· Nanoemulsion· Droplet size· Thymol· Trachyspermum ammi· Biochemical effect
Introduction
Mosquitoes (Diptera: Culicidae) have a significant role
in transmitting vector-borne diseases that cause mortality
and morbidity in humans (Benelli etal. 2018; Pavela etal.
2019). Among them, the yellow fever mosquito, Aedes
aegypti (L.) (Diptera Culicidae) is of concern as it transmits
Zika, dengue chikungunya, and yellow fever (Weaver etal.
2016; Ferreira-de-Brito etal. 2016; Benelli etal. 2019). The
annual economic loss caused due to dengue is estimated to
be US$ 9 billion (Shepard etal. 2016).
Responsible Editor: Philippe Garrigues
* Kesavan Subaharan
Kesavan.Subaharan@icar.gov.in; subaharan_70@yahoo.com
1 Division ofGermplasm Conservation andUtilization,
ICAR-National Bureau ofAgricultural Insect Resources,
Bengaluru, India560024
2 CPMU, Jawaharlal Nehru Centre forAdvanced Scientific
Research, Bangalore, India560065
3 CSIR-Central Institute ofMedicinal andAromatic Plants,
Regional Centre, Bengaluru, India560065
4 Division ofBiopesticides andEnvironmental Toxicology,
Sri Paramakalyani Centre forExcellence inEnvironmental
Sciences, Manonmaniam Sundaranar University,
Alwarkurichi, 627412Tirunelveli, TamilNadu, India
5 ICMR, National Institute forMalaria Research FU,
Bangalore, India562110
6 ICMR – Vector Control Research Centre, Field Station,
Madurai, India625002
Environmental Science and Pollution Research
1 3
The Ae. aegypti adult females are anthropophilic and
use man-made receptacles in households for breeding
(Benelli and Mehlhorn 2016). Hence, the primary level
of management depends on source reduction by involving
community participation (Alvarado-Castro etal. 2017).
When the population is in excess, the synthetic insecti-
cides belonging to pyrethroids, organophosphates, carba-
mates and insect growth regulators are widely used for
Ae. aegypti management (Chen etal. 2009; Marcombe
etal. 2012). However, the continued use of synthetic
insecticides led to the development of insecticide resist-
ance among mosquito species (Thanigaivel etal. 2017)
and negative effects on humans and the environment (Roiz
etal. 2018). In recent years, the resurgence of dengue, chi-
kungunya, the rising incidence of Zika virus, and yellow
fever exposes the limitations in Ae. aegypti management
(Wilder-Smith etal. 2017). This will increase the necessity
to search and develop low-risk and environmentally safe
methods for Ae. aegypti management.
Natural products like essential oil (EO) obtained
from plant parts are an alternative source for Ae. aegypti
management (Silva etal. 2017). Essential oils (EO) are
a complex mixture of compounds that are hydrophobic,
volatile, and of low molecular weight (Rai etal. 2017).
EOs have a wide spectrum of biological action on pests
of agriculture and public health (Kavallieratos etal. 2020)
coupled with safety to the environment and non-target
organisms (Benelli 2018; Chellappandian etal. 2018). This
attribute qualifies them as “low-risk pesticides” (Kalita
etal. 2013) to replace or minimize the use of chemical
insecticides (Isman and Grieneisen 2014).
EOs from Tagetes spp. (Dharmagadda etal. 2005),
Trachyspermum ammi (Pandey et al. 2009) Ocimum
basilicum (Govindarajan et al. 2013), Sphaeranthus
amaranthoides (Thanigaivel et al. 2019) Kaempferia
galanga (AlSalhi etal. 2020) Thymus vulgaris (Pavela
etal. 2009), Eucalyptus sp. (Lucia etal. 2007), clove bud
oil (Kalaiselvi etal. 2019), and Zanthoxylum monophylum
(Pavela and Govindarajan 2016) have a significant effect on
mosquitoes as a larvicide, adulticide, and repellents.
The EO derived from Apiaceae plants has drawn inter-
est among research groups, as they have antimicrobial and
insecticidal properties (Evergetis etal. 2013; Singh etal.
2014). Notably, the EO derived from schizocarps of T. ammi
possesses larvicidal activity against Ae. aegypti (Seo etal.
2012), oviposition deterrence, fumigant toxicity, and repel-
lence activity against Anopheles stephensi (Pandey etal.
2009), and Musca domestica (Chantawee and Soonwera
2018). Thymol, a major constituent of T. ammi EO (Pandi-
yan etal. 2019) has larvicidal action on Ae. aegypti larvae
(Govindarajan etal. 2013; Silva etal. 2017; Junkum etal.
2021) and is 1.65 times more toxic than T. ammi EO when
exposed to An. stephensi larvae (Pandey etal. 2009). This
confirms the potential bioaction of T. ammi EO and thymol
formulations against the mosquitoes.
Though the EO formulations are effective in insect man-
agement, they have limitations in use due to their volatility,
water-solubility, physical destabilization caused by gravi-
tational separation, flocculation, and coalescence (Pavela
and Benelli 2016). The existing gap can be addressed by
developing oil-in-water (o/w) emulsion which will facilitate
improving the dispersion, availability, and scale down the
volatile loss of compounds (Pavoni etal. 2020). Utilizing the
nanotechnological approaches, the EO formulations could be
developed with better physical stability, smaller size, bio-
availability, and lower risk to non-targets (Nenaah 2014).
The nanoformulations like microemulsions (MEs) and
nanoemulsions (NEs) are a recent development in pest
management (Pavela etal. 2019; Jesser etal. 2020). The
MEs are thermodynamically stable, but sensitive to changes
in temperature and required a high amount of surfactants
(Forgiarini etal. 2000). Whereas the NEs are colloidal
delivery matrix prepared by low-energy and high-energy
methods (Fryd and Mason, 2012). NEs have robust stability,
high surface area, and tunable rheology (Gupta etal. 2016).
The high-energy method of NE preparation employs high-
pressure homogenization and ultrasonication (Homs etal.
2018). The ultrasonication provides energy to produce
turbulence that splits the oil and water phase to produce
small oil droplets (Delmas etal. 2011). It is a preferred
method to prepare NE as it is economically viable, energy-
efficient, and amenable to control the formulation variables
(Periasamy etal. 2016).
The NE of Rosmarinus officinalis L. and O. basilicum L.,
Vitex negundo EO and thymol caused higher larval mortality
to Ae. aegypti over the free form of EO (Ghosh etal. 2013;
Duarte etal. 2015; Balasubramani etal. 2017; Lucia etal.
2020). There is scanty work on assessing the efficacy of T.
ammi EO NE to manage Ae. aegypti, except for an attempt
to use β-cyclodextrin formulation of T. ammi to improve its
larvicidal effect on Ae. aegypti (Pandiyan etal. 2019). Previ-
ous studies on emulsion are more focusing on the efficacy
of EO with limited information on the effects of compounds
present in the EO. To develop a novel EO based formulation
for mosquito management, herein we present the methods
to develop NE of T. ammi EO and its constituent thymol by
ultrasonication method and assess their stability, larvicidal,
and biochemical effect on Ae. aegypti.
Materials andmethods
Chemicals
Polysorbate 80 (Tween 80) was purchased from Alpha Aesar.
Ethylenediaminetetraacetic acid (EDTA), phenyl thiourea,
Environmental Science and Pollution Research
1 3
Coomassie brilliant blue (CBB), and bovine serum albumin
(BSA) were procured from Himedia Laboratories, Banga-
lore. Acetylcholineesterase kit and acetone were purchased
from Sigma Aldrich (St. Louis, MO, US).
Extraction andcharacterization ofessential oil
The seeds of T. ammi, sourced from Unjha, 23°4807.6N
72°2303.9E in Gujarat, India were shade dried for
(25 ± 2°C with RH 65 ± 5%) for 3days and cleaned for
physical and biological debris. The coarse ground seeds of
T. ammi (300g) was loaded into a 1000-ml round bottom
flask. To this, 500ml of distilled water was added, and the
temperature of the flask with contents was raised to 100°C
by placing over a heating mantle. The EO was extracted in
the Clevenger apparatus by hydro-distillation for 5h. The oil
collected in the receiver tube was separated from the aque-
ous layer using a separating funnel. The oil yield was 1.8%
(w/w). The EO was dehydrated by passing over anhydrous
sodium sulfate and stored at 4°C until use.
The extracted T. ammi EO was characterized by using
GC–MS (Agilent GC- 7890A and MS-5975) as previously
suggested by Subaharan etal. (2021). The analytical
conditions were maintained the same as those reported
by Senthoorraja et al. (2021). The components in EO
were identified by retention time and comparing the mass
fragmentation pattern using the National Institute of
Standards and Technology (NIST) library.
Preparation ofT. ammi EO andthymol inoil/water
nanoemulsion
The T. ammi EO and its constituent thymol oil-in-water
(O/W) nanoemulsion was prepared by mixing EO/Thymol
(5%) and the surfactant Tween 80 (5%) (v/v) at 400rpm for
20min under constant stirring at room temperature. Then,
water (90% v/v) was added dropwise into the mixtures of
EO/Tween 80 with continuous stirring for another 20min.
The obtained turbid mixtures of macroemulsions were sub-
jected to ultrasonic homogenization using a probe sonica-
tor (Model-VCX 750, Power—750W, Frequency—0kHz,
USA) for 10min at 50°C. The probe with a stepped microtip
(thickness about 3mm) was placed 25mm above the bot-
tom of the reservoir/container containing the liquid disper-
sion. The NE was stored at room temperature for further
characterization.
Nanoemulsion characterization
Droplet size analysis
The average droplet size distribution of NE was determined
using the dynamic light scattering technique (DLS) by
Zetasizer Nano ZS (Malvern Instruments Ltd., India) instru-
ment equipped with a 633-nm laser. The scattering intensity
was recorded with a detector kept at a backscattering angle
(173°). Measurement of NE was done on 1, 3, 5, 7, 15, 30,
and 60days after preparation. The oil droplet size (nm) was
characterized by distribution curves in intensity (%), average
droplet size, and polydispersity index (PDI). Droplet size
was expressed as mean diameter + SE (n = 3).
Transmission electron microscopy
The droplet shape in NE was observed by transmission elec-
tron microscope (TEM) JEOL—010 transmission electron
microscope operating at a voltage of 200kV. TEM samples
were prepared by mixing 100 μL of the sample with 200 μL
of 2 wt% phosphotungstic acids. A drop of the mixture was
placed on a copper grid and dried in a desiccator for removal
of water and then observed under the TEM.
Acute toxicity onAe. aegypti larvae
Aedes aegypti eggs obtained from the Indian Council of
Medical Research (ICMR)—Vector Control Research
Centre—Field station, Madurai was reared at the Veterinary
Entomology laboratory, ICAR—National Bureau of
Agricultural Insect Resources, Bengaluru, India in the
method suggested by Balasubramani etal. (2017). Briefly,
the eggs were seeded in an enamel trap having tap water
free of chlorine. On hatching, the larvae were fed with a
diet consisting of dog biscuit + yeast (3:1 ratio). Third instar
larvae were used to evaluate the efficacy of T. ammi EO,
thymol NE, and BE following the WHO method (WHO
2005). Based on the range-finding test, the NE and BE
were diluted in distilled water to obtain the concentration
ranging from 5 to 100ppm (relative to T. ammi EO and
thymol). The experiments were carried out with a group of
25 larvae of uniform size per replicate. Four replicates were
maintained. A control with surfactant alone was maintained
as a negative control. Imidacloprid tested between 0.2 and
2ppm was maintained as a positive control. The bioassays
were performed at 25 + 2°C, 60 + 5% RH and 16:8h (L:D).
Larval mortality was observed 24h after exposure. When
the mortality rate in control was above 20%, the mortality
was subjected to correction using Abbott’s formula (Abbott
1925). The lethal concentrations (LC50 and LC90) were
calculated by probit analysis (Finney 1971).
Scanning electron microscopy (SEM) ofAe. aegypti
larvae
The third instar Ae. aegypti larvae exposed to EO, thy-
mol, and control were used for SEM imaging in a method
reported by Subaharan etal. (2021). Briefly, the dead larvae
Environmental Science and Pollution Research
1 3
exposed to LC50 dose of EO, thymol NE, and control were
separated and washed with double distilled water 5 × times to
detach the debris attached to the body surface. The primary
fixation of the samples was done with glutaraldehyde 2.5%
in distilled water for 1–2h. The samples were then rinsed
in distilled water for 10–0min, and the second fixation was
done for 1–2h in 1–% osmium tetroxide in distilled water.
The samples rinsed with distilled water (10–20min) were
serially dehydrated with 25, 50, 75, 90, and 100% ethanol
for 10min each. Ethanol was used for critical point drying
(CPD) of the samples. Samples were mounted on an alu-
minum stub and spluttered with gold particles by cathodic
spraying. External structures of Ae. aegypti larva was
observed on a scanning electronic microscope (Carl Zeiss
MERLIN VP Compact) for structural changes. Sample prep-
aration for SEM imaging was done at room temperature.
AChE inhibition
The surviving third instar Ae. aegypti larvae treated with
NE and BE of EO at LC50 (46.7 and 57.5ppm) and thymol
at LC50 (34.8 and 43.9ppm) from bioassay were used to
estimate the acetylcholineesterase (AChE). Larvae exposed
to tap water alone were used as a negative control and those
exposed to imidacloprid (LC50 dose 0.96ppm) was used as a
positive control. Ae. aegypti larvae exposed to the treatments
were homogenized in ice-cold 0.1-M phosphate buffer (pH
8.0), and the contents were centrifuged at 4°C for 20min
at 10,000rpm. The supernatant was used to estimate acetyl-
cholineesterase inhibition activity. The protein content was
estimated as suggested by Bradford (1976). Acetylcholinest-
erase (MAK119, Sigma Aldrich, US) inhibition activity was
measured following the modified method of Ellman etal.
(1961). The reaction mixture consisting of 10µl of enzyme
homogenate and 190µl of reagent (consisting of 5,5-dithio-
bis-2-nitrobenzoic acid, (DTNB) and acetyl thiocholine was
dissolved in Tris–HCl buffer (100Mm, pH 8.0 assay buffer).
Control had 10µl of distilled water and 190µl of reagent,
reaction mixture. The reaction mixture was incubated for
10min at room temperature. The rate of change in absorb-
ance was measured at 415 for 10min in a microplate reader
(iMark, BioRad). The percent inhibition was calculated as
suggested previously (Kim etal. 2013).
Statistical analysis
In toxicity assay when the control mortality was above
20%, it was subjected to correction using Abbott’s formu-
lae (Abbott 1925). The data on Ae. aegypti larval mortality
were subjected to probit analysis (Finney 1971) to determine
the median lethal dose LC50, and their 95% CI values and
chi-square test were calculated using the SPSS software ver-
sion 14.0. The variations in particle droplet size, PDI, and
variation in AChE enzyme activity were compared using
one-way analysis of variance (ANOVA) followed by Tukey’s
post hoc test (P < 0.05) using SPSS.
Results anddiscussion
Characterization ofEO
The constituents of T. ammi EO characterized by GC–MS
are shown in Table1. Twenty compounds were identified in
T. ammi EO. The major constituents were thymol (54.22%),
p-cymene (15.04%), and γ-terpinene (10.46%). The minor
compounds present between 0.5 and 1% include carvac-
rol (1.86%), p-cymenene (1.1%), carane, 4,5-epoxy-, trans
(0.85%), terpinen-4-ol (0.79%) 8,9-dehydrothymol (0.79%),
α-terpineol (0.76), p-cymene-2,5-diol (0.74), p-cymen-8-ol
(0.62), α-pinene (0.58), and camphene (0.51%). The con-
stituents in the EO in our studies agree with the previous
report (Benelli etal. 2017).
Characterization ofnanoemulsion
EOs have a potential effect on insects, but the hydrophobic-
ity and low solubility in water are a limitation for further use
as a larvicide in vector management. An effective mosquito
Table 1 Chemical composition of the T. ammi essential oil
Compounds Reported RI Percentage
composition
α-Pinene 933 0.58
Camphene 951 0.51
β-Pinene 980 0.24
β-Myrcene 993 0.3
α-Phellandrene 1005 0.08
δ-3-Carene 1011 0.03
α-Terpinene 1017 0.16
p-Cymene 1034 15.04
γ-Terpinene 1060 10.89
Terpinolene 1088 0.24
p-Cymenene 1090 1.1
Trans-2-caren-4-ol 1178 0.36
Terpinen-4-ol 1177 0.79
Carane, 4,5-epoxy-, trans 1179 0.85
p-Cymen-8-ol 1183 0.62
α-Terpineol 1190 0.76
8,9-Dehydrothymol 1221 0.79
Thymol 1291 55.84
Carvacrol 1299 1.86
p-Cymene-2,5-diol 1561 0.74
Total 91.78
Environmental Science and Pollution Research
1 3
larvicide formulation needs to have a good dispersion in
water and stability. The O/W nanoemulsion of T. ammi EO
and its constituent thymol is an improved formulation to
overcome the limitations of dispersibility with enhanced
efficacy over free EO formulation.
Droplet size andpolydispersity index
The emulsion stability of EO and thymol NE was estimated
by recording the droplet size and the polydispersity index
(PDI) value on 1, 3, 5, 7, 15, 30, and 60days after stor-
age at 26°C. The droplet size and the PDI of NEs over a
period of time are shown in Table2. There was a signifi-
cant difference in the droplet size and PDI values between
NEs of EO and thymol, with the latter showing larger drop-
let size (F13,28 = 8411.39 P < 0.001) and higher PDI value
(F13,28 = 58.10 P < 0.001). Between the formulations, EO NE
showed smaller droplet size (65 ± 0.7 to 83 ± 0.09nm from 1
and 60days of storage) and lower PDI values (0.18 ± 0.003
to 0.20 ± 0.07 from 1 to 60days of storage). The slow desta-
bilization in EO NE until the 60th day of storage may be due
to the low PDI values (< 0.22) that reflects on the narrow
distribution of the particle size (Table1) which in turn attrib-
utes to the robust physical stability of NE due to reduced
Ostwald’s ripening of the droplets (Jesser etal. 2020). The
droplet size and stability of the EO NE maintained below
90nm on 60days of storage may be due to the steric effect
(Burt 2004).
Correlograms obtained during the droplet size meas-
urement of NEs of EO and thymol present a normal single
exponential decay of the intensities of scattered light over
time owing to the Brownian motion of the droplets (Figs.1
and 2), i.e., the correlogram curves are much steeper and
exhibit flat baseline at the end of the decay, suggesting that
the NEs remained monodispersed without any aggregation
throughout 60days of storage. These agree with the PDI
values of NEs of EO and thymol reported in Table2.
The mean droplet size of thymol NE during stor-
age ranged from 167 ± 1 on day 1 to 185 ± 0.6, 201 ± 1,
228 ± 0.03, and 230 ± 1nm on 5, 15, 30, and 60days after
storage, respectively. The PDI values for thymol NE ranged
from 0.30 ± 0.03 to 0.40 ± 0.008. The increase in droplet size
of thymol NE (230 ± 1nm after 60days of storage) may
be attributed to the droplet aggregation that resulted from
emulsion instability. Thymol has weak dispersibility as it has
a higher PDI value (Kumari etal. 2018), but when mixed
with eugenol, its dispersion in an aqueous medium increased
resulting in monodisperse droplets (Lucia etal. 2020).
Though the droplet size and PDI value of thymol NE
were greater than EO NE, the rate of increase in droplet
size during 60days of storage was 1.37 and 1.27-fold,
respectively. NE of Rosemarinus officinalis oil that had
a larvicidal effect on Ae. aegypti had a 3.6-fold rise in
droplet size 30days after storage (Duarte etal. 2015).
Smaller PDI values indicate that nanodroplets are more
homogeneous and the PDI of less than 0.6 is an aggrega-
ble homogeneity (Suyal etal. 2018). In our studies, the
PDI values of thymol NE was 0.40 ± 0.008 and droplet
size was 230 ± 1.0nm on 60days of storage. The narrow
range of increase in droplet size did not alter the concept
of it being termed a nanoemulsion, as previous studies
have reported that average droplet size with a diameter
lower than 300nm is to be considered as a nanoemul-
sion (Hashem etal. 2018). The EO and thymol NE did
not show any phase separation after 60days of storage.
TEM
Droplet shape of NEs of EO and thymol, when viewed
by transmission electron microscopy appeared well dis-
persed, spherical, and retained the overall size range
within < 250 nm (Fig.3). The transmission electron
microscopy (TEM) images of Tarragon, clove, and black
seed oil nanoemulsion revealed the spherical structure
of the droplets (Mossa etal. 2021). The spherical struc-
ture is due to the reduction of the interfacial area arising
due to a small radius and increased interfacial tension
(Pavoni etal. 2020).
Acute toxicity of nanoemulsion and bulk emulsion of T.
ammi EO and thymol to Ae. aegypti larvae.
The toxicity of NE and bulk emulsion BE of T. ammi
EO and thymol (referred to as EO BE and thymol BE)
on Ae. aegypti larvae is shown in Table3. Irrespective
of emulsion type, both EO and thymol showed larvicidal
Table 2 Nanoemulsion droplet size and polydispersity index
Means within a column followed by the same letter are not signifi-
cantly different by Tukey’s test (P < 0.05)
Nanoemulsion type Days
after
storage
Size (nm) + SE PDI + SE
T. ammi EO + Tween 80 165 ± 0.7a0.18 ± 0.003ab
366 ± 0.73ab 0.17 ± 0.005a
567 ± 0.89ab 0.17 ± 0.012a
769 ± 0.29bc 0.17 ± 0.002a
15 72 ± 0.46cd 0.17 ± 0.04a
30 74 ± 0.39d0.22 ± 0.001ab
60 83 ± 0.09e0.20 ± 0.07ab
Thymol + Tween 80 1167 ± 1.0f0.30 ± 0.03c
3177 ± 1.0g 0.30 ± 0.05c
5185 ± 0.60h 0.36 ± 0.11d
7196 ± 0.70i0.36 ± 0.12d
15 201 ± 1.0 j0.24 ± 0.02bc
30 228 ± 0.03k 0.40 ± 0.02d
60 230 ± 1.0k 0.40 ± 0.008d
Environmental Science and Pollution Research
1 3
activity against Ae. aegypti. The toxicity of T. ammi EO
to Ae. aegypti (Silva etal. 2017; Seo etal. 2012) and
An. stephensi larvae (Pandey etal. 2009) was reported
earlier. Among the formulations tested, NE caused higher
toxicity than BE (Table3). Enhanced larvicidal activ-
ity of eucalyptus oil nanoemulsion than its bulk emul-
sion to Culex quinquefasciatus was reported by Sugu-
mar etal. (2014). Between the NEs tested, thymol NE
(LC50 34.89ppm) was more toxic than EO NE (LC50
46.73ppm). A similar trend of thymol (LC50 52.82ppm)
being more toxic than EO (59.67ppm) was observed in
BEs too. The order of toxicity was thymol NE > T. ammi
EO NE > thymol BE > T. ammi BE. Our observations
are in line with findings by Pandey etal. (2009) who
reported thymol to be 1.65 times toxic to An. stephensi
larvae as compared to T. ammi whole extract.
The previous studies revealed that the nanoemulsion
of thymol showed higher toxicity to Ae. aegypti lar-
vae (LC50 11.1ppm), but the addition of eugenol to it
declined the toxicity (Lucia etal. 2020). EOs constitute
20–60 compounds belonging to phenols, alkaloids, and
terpenes, among them the bioactivity is mainly attributed
to few compounds that are in higher concentration (Pavela
2015); nevertheless, the compounds present in lower con-
centrations also add to its efficacy (Benelli etal. 2017).
Artificial mixtures of T. ammi EO prepared and tested
without the major constituent, viz., thymol, p-cymene and
γ-terpinene caused a decrease in toxicity to Ae aegypti as
compared to their presence in the mixture that enhanced
the toxicity (Seo etal. 2012). Thymol present in EO is a
major contributor in causing toxicity to Ae. aegypti (Park
etal. 2011). In our study, thymol is a major constitu-
ent in T. ammi EO along with p-cymene (15.04%) and
γ-terpinene (10.46%); this explains the enhanced larvi-
cidal activity of NEs of thymol and EO.
The droplet size and PDI are important physical
characteristics to determine the interaction of nanofor-
mulations and insects (Benelli 2018). In our studies,
the droplet size of EO and thymol ranged from 90 to
230nm. Previous studies confirm that droplet size in
nanoemulsion with a size ranging from 100 to 400nm
impute benefits likes dispersion, thermal stability, per-
meability (Bordes etal. 2009). Though there was a dif-
ference in NE droplet size of EO (< 90nm) and thymol
Fig. 1 Size distribution by
intensity and correlogram of T.
ammi EO NE
Environmental Science and Pollution Research
1 3
(230nm) the LC50 values were lower than 100ppm, and
this qualifies both as potent mosquito larvicide, as sug-
gested by Pavela (2015). In such a condition, it permits
the use of NEs of thymol and EO as larvicide against Ae.
aegypti. Though thymol as a single component is effec-
tive larvicide than EO as a whole, the advantage of other
constituents in the EO will add to their efficacy, and the
possibility of developing insecticide resistance can be
scaled down. Added to this the fractionation of thymol
from EO and use would add to cost than use of EO as
Fig. 2 Size distribution by
intensity and correlogram of
thymol NE
Fig. 3 TEM images of a EO NE
and b thymol NE
Table 3 Acute toxicity of emulsions of T. ammi EO and its constitu-
ent thymol
95% CL confidence interval at 95% confidence level
Test sample LC50 (ppm) 95% CL df Chi-square P value
EO NE 46.73 35.78–63.20 4 9.25 0.05
EO BE 59.67 45.98–82.68 4 9.20 0.06
Thymol NE 34.89 26.79–45.87 4 9.0 0.05
Thymol BE 52.82 40.94–71.27 4 8.99 0.05
Imidacloprid 0.96 0.76–1.27 3 7.2 0.06
Environmental Science and Pollution Research
1 3
such. The NEs are economical over Bes as in nanoemul-
sions; the requirement of thymol and EO can be reduced
by 1.51 and 1.28 times over the BE using EO and thymol
as a whole. Downsizing the droplet in nanoemulsion of
eucalyptus oil enhanced the contact of the toxicity to
Tribolium castaneum (Adak etal. 2020).
The synthetic insecticide imidacloprid was more toxic
(LC50 0.96ppm) than NE and BE formulations of EO and
thymol (Table3). Though chemical insecticides are effec-
tive, their indiscriminate use adds to problems like insecti-
cide resistance and harm to humans and the environment.
Scanning electron microscopy
The impact of EO and thymol on III instar larvae of Ae.
aegypti by contact exposure is shown in Fig.4a. EO and
thymol at LC50 dose distorted the cuticle by shrinkage, as
shown in Fig.4b, c. The control larvae had smooth and nor-
mal skin texture. One of the mechanisms for an effective
insecticide is its ability to cross over the cuticle (Kasai etal.
2014). The droplet size of the particles in nm facilitates bet-
ter penetration coupled with lipophilicity. The wider surface
area of the nanoemulsion would have been a possible reason
for better action of the EO and thymol by causing shrinkage
in larval cuticle in Ae. aegypti larvae. A decrease in droplet
size in nanoemulsion enhanced the uptake and penetration
of bioactive compounds in EO into the insect body; thereby,
improving the efficacy of NE over the BE was reported ear-
lier (Pascual-Villalobos etal. 2019).
AChE inhibition assay
The effects of NEs of EO and thymol on larval acetylcho-
lineesterase activity were studied. The inhibition of AChE
activity of NEs of EO and thymol is shown in Fig.5. The
NE of thymol was a potent AChE inhibitor as it caused
the significant inhibition activity (83.48 ± 1.21%), fol-
lowed by imidacloprid that caused 78.78 ± 5.48% inhibi-
tion that was at par. EO NE caused an inhibition activity
(53.62 ± 3.77%) (F2,6 = 18.96 P < 0.003).
Acetylcholineesterase is an enzyme in an insect that
hydrolyzes acetylcholine (Ach). Inhibition of AChE
causes paralysis in insects that leads to death (Huang etal.
2020). Natural products having phenols, monoterpenoids,
and mixtures of compounds inhibit the activity of AChE
(Wu etal. 2020) and block insect octopamine receptors
(Almadiy 2020), thereby causing an insecticidal effect.
Our studies showed that thymol NE caused the high-
est inhibition of AChE (83.48%) followed by EO NE
(53.62 ± 3.775). The insecticidal effect of nanoemulsion
of EO is facilitated through enzymatic inhibition which
has an impact on neural signal transduction (Seo etal.
2015). Thymol inhibited AChE in the third instar larvae
of Ae. albopictus and interacted with GABA-A and octo-
pamine receptors in Culex pipiens (Youssefi etal. 2019).
Pterodon emarginatus nanoemulsion showed anti-acetyl-
cholinesterase activity in Ae. aegypti (Oliveira etal. 2016).
Monoterpenes in essential form hydrophobic interaction
with the catalytic subunit of AChE thereby reducing its
Fig. 4 SEM micrograph of Ae.
aegypti larvae intersegmental
region. a Control (without any
oil), b T. ammi EO, c thymol
Environmental Science and Pollution Research
1 3
activity (Oyedeji etal. 2020). This may be the mechanism
of insecticidal activity of T. ammi EO and thymol.
Conclusion
The NEs of T. ammi EO and thymol optimized in this study
facilitated the development of eco-friendly and effective
nanoformulation for mosquito management. The NEs
stability and enhanced bioactivity of EO and thymol against
Ae. aegypti larvae as compared to BE confirm the effect of
NE in causing acute larval toxicity, impact on the larval
cuticle and its ability to inhibit the acetylcholineesterase
activity. These properties ascribe NEs of EO and thymol
as potential candidates for Ae. aegypti management.
Considering the improved formulation characters of NE,
it is effective and economical as compared to BE which
uses EO as a whole. Though the synthetic insecticide had
better larvicidal activity than the NE, they also bring in
ill effects due to their indiscriminate use like insecticide
resistance and harm environment. The EO with multiple
modes of action prevents or delays the development
of insecticide resistance in mosquitoes. Hence, the
developed NE formulation has the benefit to minimize the
dependence of harmful chemical pesticides used for Ae.
aegypti management. Further research on the impact of
nanoemulsion on human health and the environment will
strengthen its use in area-wide management.
Acknowledgements The work was supported by ICAR–NBAIR as an
institutionally funded project. Imaging services were provided by Dr
Geetha; NICE Lab NCBS is acknowledged
Author contribution Kesavan Subaharan: conceptualization, writ-
ing — original draft, and data curation. Periyasamy Senthamarai
Selvan, Tania Das, and Rajendran Senthoorraja: investigation. T. M.
Subramanya: investigation, Sowmya Manjunath: investigation on bio-
chemical aspects. Muthu Gounder Mohan: resources. Vppalayam Shan-
mugam Pragadheesh: investigation and validation. Nandagopal Bak-
thavatsalam: writing — review and editing. M. Eswaramoorthy: writing
— review and editing, and supervision. Paulraj Philip Samueland Renu
Govindaraju: resources. Sreehari U: writing — review and editing.
Sengottayan Senthil-Nathan: writing — original draft and editing.
All the authors read and approved the final manuscript.
Funding The work was supported by ICAR–NBAIR as an institutional
project.
Data availability Data is available by request to the corresponding
author.
Declarations
Ethics approval and consent to participate Not applicable.
Consent for publication Not applicable
Competing interests The authors declare no competing interests.
References
Abbott WS (1925) A method of computing the effectiveness of an
insecticide. J Econ Entomol 18:265–267. https:// doi. org/ 10. 1093/
jee/ 18.2. 265a
Adak T, Barik N, Patil NB, Govindharaj GPP, Gadratagi BG, Anna-
malai M, Mukherjee AK, Rath PC (2020) Nanoemulsion of euca-
lyptus oil: an alternative to synthetic pesticides against two major
storage insects Sitophilus oryzae (L.) and Tribolium castaneum
Fig. 5 Inhibition of acetylcho-
lineesterase (AChE) enzyme
activity of (LC50 values) NE-EO
and thymol against Ae. aegypti.
Bars having the same letter
do not differ significantly at
P < 0.05 (ANOVA followed by
Tukey test)
Environmental Science and Pollution Research
1 3
(Herbst) of rice. Ind Crops Prod 143:111849. https:// doi. org/ 10.
1016/j. indcr op. 2019. 111849
AlSalhi MS, Elumalai K, Devanesan S, Govindarajan M, Krishnappa
K, Maggi F (2020) The aromatic ginger Kaempferia galanga L.
(Zingiberaceae) essential oil and its main compounds are effective
larvicidal agents against Aedes vittatus and Anopheles maculatus
without toxicity on the non-target aquatic fauna. Ind Crops Prod
158:113012. https:// doi. org/ 10. 1016/j. indcr op. 2020. 113012
Almadiy AA (2020) Chemical composition, insecticidal and biochemi-
cal effects of two plant oils and their major fractions against Aedes
aegypti, the common vector of dengue fever. Heliyon 6:e04915.
https:// doi. org/ 10. 1016/j. heliy on. 2020. e04915
Alvarado-Castro V, Paredes-Solís S, Nava-Aguilera E, Morales-Pérez
A, Alarcón-Morales L, Balderas-Vargas NA, Andersson N (2017)
Assessing the effects of interventions for Aedes aegypti control:
systematic review and meta-analysis of cluster randomised con-
trolled trials. BMC Public Health 17:384. https:// doi. org/ 10. 1186/
s12889- 017- 4290-z
Balasubramani S, Rajendhiran T, Moola AK, Diana RKB (2017)
Development of nanoemulsion from Vitex negundo L. essential
oil and their efficacy of antioxidant, antimicrobial and larvicidal
activities (Aedes aegypti L.). Environ Sci Pollut Res Int 24:15125–
15133. https:// doi. org/ 10. 1007/ s11356- 017- 9118-y
Benelli G (2018) Plant-borne compounds and nanoparticles: chal-
lenges for medicine, parasitology and entomology. Environ
Sci Pollut Res Int 25:10149–10150. https:// doi. org/ 10. 1007/
s11356- 017- 9960-y
Benelli G, Mehlhorn H (2016) Declining malaria, rising of den-
gue and zika virus: insights for mosquito vector control.
Parasitol Res 115:1747–1754. https:// doi. org/ 10. 1007/
s00436- 016- 4971-z
Benelli G, Pavela R, Canale A, Cianfaglione K, Ciaschetti G, Conti
F, Nicoletti M, Senthil Nathan S, Mehlhorn H, Maggi F (2017)
Acute larvicidal toxicity of five essential oils (Pinus nigra, Hysso-
pus officinalis, Satureja montana, Aloysia citrodora and Pelargo-
nium graveolens) against the filariasis vector Culex quinquefascia-
tus: synergistic and antagonistic effects. Parasitol Int 66:166–171.
https:// doi. org/ 10. 1016/j. parint. 2017. 01. 012
Benelli G, Pavela R, Petrelli R, Cappellacci L, Canale A, Senthil-
Nathan S, Maggi F (2018) Not just popular spices! essential oils
from Cuminum cyminum and Pimpinella anisum are toxic to insect
pests and vectors without affecting non-target invertebrates. Ind
Crop Prod 124:236–243. https:// doi. org/ 10. 1016/j. indcr op. 2018.
07. 048
Benelli G, Pavela R, Petrelli R, Cappellacci L, Bartolucci F, Canale
A, Maggi F (2019) Origanum syriacum subsp. syriacum: from an
ingredient of Lebanese ‘manoushe’ to a source of effective and
eco-friendly botanical insecticides. Ind Crops Prod 134:26–32.
https:// doi. org/ 10. 1016/j. indcr op. 2019. 03. 055
Bordes P, Pollet E, Averous L (2009) Nano-biocomposites: biodegrad-
able polyester/nanoclay systems. Prog Polym Sci 34:125–155.
https:// doi. org/ 10. 1016/j. progp olyms ci. 2008. 10. 002
Bradford MM (1976) A rapid and sensitive method for the quantita-
tion of microgram quantities of protein utilizing the principle of
protein-dye binding. Anal Biochem 72:248–254. https:// doi. org/
10. 1006/ abio. 1976. 9999
Burt S (2004) Essential oils: their antibacterial properties and potential
applications in foods-a review. Int J Food Microbiol 94:223–253.
https:// doi. org/ 10. 1016/j. ijfoo dmicro. 2004. 03. 022
Chantawee A, Soonwera M (2018) Larvicidal, pupicidal and ovipo-
sition deterrent activities of essential oils from Umbelliferae
plants against house fly Musca domestica. Asian Pac J Trop Med
11:621–629. https:// doi. org/ 10. 4103/ 1995- 7645. 246338
Chellappandian M, Vasantha-Srinivasan P, Senthil-Nathan S, Karthi
S, Thanigaivel A, Ponsankar A, Kalaivani K, Hunter WB (2018)
Botanical essential oils and uses as mosquitocides and repellents
against dengue. Environ Int 113:214–230. https:// doi. org/ 10.
1016/j. envint. 2017. 12. 038
Chen CD, Lee HL, Chan CK, Ang CL, Azahari AH, Lau KW, Sofian-
Azirun M (2009) Laboratory bioefficacy of nine commercial for-
mulations of temephos against larvae of Aedes aegypti (L.), Aedes
albopictus Skuse and Culex quinquefasciatus Say. Trop Biomed
26:360–365
Delmas T, Piraux H, Couffin AC, Texier I, Vinet F, Poulin P, Cates
ME, Bibette J (2011) How to prepare and stabilize very small
nanoemulsions. Langmuir 27:1683–1692. https:// doi. org/ 10. 1021/
la104 221q
Dharmagadda VS, Naik SN, Mittal PK, Vasudevan P (2005) Larvicidal
activity of Tagetes patula essential oil against three mosquito spe-
cies. Bioresour Technol 96:1235–1240. https:// doi. org/ 10. 1016/j.
biort ech. 2004. 10. 020
Duarte JL, Amado JRR, Oliveira AEMFM, Cruz RAS, Ferreira AM,
Souto RNP, Falcão DQ, Carvalho JCT, Fernandes CP (2015)
Evaluation of larvicidal activity of a nanoemulsion of Rosmarinus
officinalis essential oil. Rev Bras Farmacogn 25:189–192. https://
doi. org/ 10. 1016/j. bjp. 2015. 02. 010
Ellman GL, Courtney KD Jr, Andres V, Feather-Stone RM (1961) A
new and rapid colorimetric determination of acetylcholinesterase
activity. Biochem Pharmacol 7:88–95. https:// doi. org/ 10. 1016/
0006- 2952(61) 90145-9
Evergetis E, Michaelakis A, Haroutounian SA (2013) Exploitation of
Apiaceae family essential oils as potent biopesticides and rich
source of phellandrenes. Ind Crops Prod 41:365–370. https:// doi.
org/ 10. 1016/j. indcr op. 2012. 04. 058
Ferreira-de-Brito A, Ribeiro IP, Miranda RM, Fernandes RS, Campos
SS, Silva KA, Castro MG, Bonaldo MC, Brasil P, Lourenco-de-
Oliveira R (2016) First detection of natural infection of Aedes
aegypti with zika virus in Brazil and throughout South America.
Mem Inst Oswaldo Cruz 111:655–658. https:// doi. org/ 10. 1590/
0074- 02760 160332
Finney DJ (1971) Probit analysis. Cambridge University Press, Cam-
bridge, London, UK
Forgiarini A, Esquena J, Gonzalez C, Solans C (2000) Studies of the
relation between phase behavior and emulsification methods with
nanoemulsion formation. Prog Colloid Polym Sci 115:36–39.
https:// doi. org/ 10. 1007/3- 540- 46545-6_8
Fryd MM, Mason TG (2012) Advanced nanoemulsions. Annu Rev
Phys Chem 63:493–518. https:// doi. org/ 10. 1146/ annur ev- physc
hem- 032210- 103436
Ghosh V, Mukherjee A, Chandrasekaran N (2013) Formulation and
characterization of plant essential oil-based nanoemulsion: evalu-
ation of its larvicidal activity against Aedes aegypti. Asian J Chem
25:S321–S323
Govindarajan M, Sivakumar R, Rajeswary M, Veerakumar K (2013)
Mosquito larvicidal activity of thymol from essential oil of
Coleus aromaticus Benth. against Culex tritaeniorhynchus,
Aedes albopictus, and Anopheles subpictus (Diptera: Culici-
dae). Parasitol Res 112:3713–3721. https:// doi. org/ 10. 1007/
s00436- 013- 3557-2
Gupta A, Eral HB, Hatton TA, Doyle PS (2016) Nanoemulsions: for-
mation, properties and applications. Soft Matter 12:2826–2841.
https:// doi. org/ 10. 1039/ c5sm0 2958a
Hashem AS, Awadalla SS, Zayed GM, Maggi F, Benelli G (2018)
Pimpinella anisum essential oil nanoemulsions against Tribo-
lium castaneum-insecticidal activity and mode of action. Envi-
ron Sci Pollut Res Int 15:18802–18812. https:// doi. org/ 10. 1007/
s11356- 018- 2068-1
Homs M, Calderó G, Monge M, Morales D, Solans C (2018) Influence
of polymer concentration on the properties of nano-emulsions and
nanoparticles obtained by a low-energy method. Colloids Surf
A Physicochem Eng Asp 536:204–212. https:// doi. org/ 10. 1016/j.
colsu rfa. 2017. 06. 009
Environmental Science and Pollution Research
1 3
Huang Y, Lin M, Jia M, Hu J, Zhu L (2020) Chemical composition
and larvicidal activity against Aedes mosquitoes of essential oils
from Arisaema fargesii. Pest Manag Sci 76:534–542. https:// doi.
org/ 10. 1002/ ps. 5542
Isman MB, Grieneisen ML (2014) Botanical insecticide research: many
publications, limited useful data. Trends Plant Sci 19:140–145.
https:// doi. org/ 10. 1016/j. tplan ts. 2013. 11. 005
Jesser E, Lorenzetti AS, Yeguerman C, Murray AP, Domini C, Werdin-
Gonzalez JO (2020) Ultrasound-assisted formation of essential oil
nanoemulsions: emerging alternative for Culex pipiens pipiens
Say (Diptera: Culicidae) and Plodia interpunctella Hübner (Lepi-
doptera: Pyralidae) management. Ultrason Sonochem 61:104832.
https:// doi. org/ 10. 1016/j. ultso nch. 2019. 104832
Junkum A, Intirach J, Chansang A, Champakaew D, Chaithong U, Jit-
pakdi A, Riyong D, Somboon P, Pitasawat B (2021) Enhancement
of temephos and deltamethrin toxicity by Petroselinum crispum
oil and its main constituents against Aedes aegypti (Diptera: Culi-
cidae). J Med Entomol 58:1298–1315. https:// doi. org/ 10. 1093/
jme/ tjab0 08
Kalaiselvi A, Aswin Jeno JG, Roopan SM, Madhumitha G, Nakkeeran
E (2019) Chemical composition of clove bud oil and development
of clove bud oil loaded niosomes against three larvae species. Int
Biodeterior Biodegrad 137:102–108. https:// doi. org/ 10. 1016/j.
ibiod. 2018. 12. 004
Kalita B, Bora S, Sharma AK (2013) Plant essential oils as mosquito
repellent-a review. Int J Res Dev Pharm L Sci 3:741–747
Kasai S, Komagata O, Itokawa K, Shono T, Ng LC, Kobayashi M,
Tomita T (2014) Mechanisms of pyrethroid resistance in the den-
gue mosquito vector, Aedes aegypti: target site insensitivity, pen-
etration, and metabolism. PLoS Negl Trop Dis 8:e2948. https://
doi. org/ 10. 1371/ journ al. pntd. 00029 48
Kavallieratos NG, Boukouvala MC, Ntalli N, Skourti A, Karagianni
ES, Nika EP, Kontodimas DC, Cappellacci L, Petrelli R, Cian-
faglione K, Morshedloo MR, Tapondjou LA, Rakotosaona R,
Maggi F, Benelli G (2020) Effectiveness of eight essential oils
against two key stored-product beetles, Prostephanus truncatus
(Horn) and Trogoderma granarium Everts. Food Chem Toxicol
139:111255. https:// doi. org/ 10. 1016/j. fct. 2020. 111255
Kim KM, Nasir A, Caetano-Anolles G (2013) The importance of using
realistic evolutionary model for retrodicting proteomes. Biochimie
99:129–137. https:// doi. org/ 10. 1016/j. biochi. 2013. 11. 019
Kumari S, Kumaraswamy RV, Choudhary RC, Sharma SS, Pal A, Ral-
iya R, Biswas P, Saharan V (2018) Thymol nanoemulsion exhibits
potential antibacterial activity against bacterial pustule disease
and growth promotory effect on soybean. Sci Rep 8:6650. https://
doi. org/ 10. 1038/ s41598- 018- 24871-5
Lucia A, Gonzalez Audino P, Seccacini E, Licastro S, Zerba E, Masuh
H (2007) Larvicidal effect of Eucalyptus grandis essential oil and
turpentine and their major components on Aedes aegypti larvae.
J Am Mosq Control Assoc 23:299–303. https:// doi. org/ 10. 2987/
8756- 971X(2007) 23[299: LEOEGE] 2.0. CO;2
Lucia A, Girard C, Fanucce M, Coviella C, Rubio RG, Ortega F,
Guzmán E (2020) Development of an environmentally friendly
larvicidal formulation based on essential oil compounds blend
to control Aedes aegypti larvae: correlations between physico-
chemical properties and insecticidal activity. ACS Sustainable
Chem Eng 8:10995–11006. https:// doi. org/ 10. 1021/ acssu schem
eng. 0c037 78
Marcombe S, Mathieu RB, Pocquet N, Riaz MA, Poupardin R, Selior
S, Darriet F, Reynaud S, Yébakima A, Corbel V, David JP, Chan-
dre F (2012) Insecticide resistance in the dengue vector Aedes
aegypti from Martinique: distribution, mechanisms and relations
with environmental factors. PLoS ONE 7:e30989. https:// doi. org/
10. 1371/ journ al. pone. 00309 89
Mossa ATH, Mohafrash SMM, El-SHE Z, Abdelsalam IS, Sahab AF
(2021) Development of eco-friendly nanoemulsions of some
natural oils and evaluating of its efficiency against postharvest
fruit rot fungi of cucumber. Ind Crops Prod 159:113049. https://
doi. org/ 10. 1016/j. indcr op. 2020. 113049
Nenaah G (2014) Chemical composition, toxicity and growth inhibi-
tory activities of essential oils of three Achillea species and their
nano-emulsions against Tribolium castaneum (Herbst). Ind Crops
Prod 53:252–260. https:// doi. org/ 10. 1016/j. indcr op. 2013. 12. 042
Oliveira AE, Duarte JL, Amado JR, Cruz RA, Rocha CF, Souto RN,
Ferreira RM, Santos K, da Conceição EC, de Oliveira LA, Kele-
com A, Fernandes CP, Carvalho JC (2016) Development of a lar-
vicidal nanoemulsion with Pterodon emarginatus Vogel oil. PLoS
ONE 11:e0145835. https:// doi. org/ 10. 1371/ journ al. pone. 01458 35
Oyedeji AO, Okunowo WO, Osuntoki AA, Olabode TB, Ayo-Folo-
runso F (2020) Insecticidal and biochemical activity of essential
oil from Citrus sinensis peel and constituents on Callosobru-
chus maculatus and Sitophilus zeamais. Pestic Biochem Physiol
168:104643. https:// doi. org/ 10. 1016/j. pestbp. 2020. 104643
Pandey SK, Upadhyay S, Tripathi AK (2009) Insecticidal and repellent
activities of thymol from the essential oil of Trachyspermum ammi
(Linn.) Sprague seeds against Anopheles stephensi. Parasitol Res
105:507–512. https:// doi. org/ 10. 1007/ s00436- 009- 1429-6
Pandiyan GN, Mathew N, Munusamy S (2019) Larvicidal activity of
selected essential oil in synergized combinations against Aedes
aegypti. Ecotoxicol Environ Saf 174:549–556. https:// doi. or g/ 10.
1016/j. ecoenv. 2019. 03. 019
Park HM, Kim J, Chang KS, Kim BS, Yang YJ, Kim GH, Shin SC,
Park IK (2011) Larvicidal activity of Myrtaceae essential oils and
their components against Aedes aegypti, acute toxicity on Daph-
nia magna, and aqueous residue. J Med Entomol 48:405–410.
https:// doi. org/ 10. 1603/ me101 08
Pascual-Villalobos MJ, Guirao P, Diaz-Banos FG, Canto-Tejero M, Vil-
lora G (2019) Oil in water nanoemulsion formulations of botanical
active substances. In: Koul O (ed) Nano-biopesticides today and
future perspectives. Academic Press, pp 223–247
Pavela R (2015) Essential oils for the development of eco-friendly
mosquito larvicides: a review. Ind Crops Prod 76:174–187. https://
doi. org/ 10. 1016/j. indcr op. 2015. 06. 050
Pavela R, Benelli G (2016) Essential oils as ecofriendly biopesticides?
Challenges and constraints. Trends Plant Sci 21:1000–1007.
https:// doi. org/ 10. 1016/j. tplan ts. 2016. 10. 005
Pavela R, Vrchotova N, Triska J (2009) Mosquitocidal activities of
thyme oils (Thymus vulgaris L.) against Culex quinquefasciatus
(Diptera: Culicidae). Parasitol Res 105:1365–1370. https:// doi.
org/ 10. 1007/ s00436- 009- 1571-1
Pavela R, Govindarajan M (2016) The essential oil from Zanthoxylum
monophyllum a potential mosquito larvicide with low toxicity
to the non-target fish Gambusia affinis. J Pest Sci 90:369–378.
https:// doi. org/ 10. 1007/ s10340- 016- 0763-6
Pavela R, Benelli G, Pavoni L, Bonacucina G, Cespi M, Cianfaglione K,
Bajalan I, Morshedloo MR, Lupidi G, Romano D, Canale A, Maggi F
(2019) Microemulsions for delivery of Apiaceae essential oils-towards
highly effective and eco-friendly mosquito larvicides? Ind Crops Prod
129:631–640. https:// doi. org/ 10. 1016/j. indcr op. 2018. 11. 073
Pavoni L, Perinelli DR, Bonacucina G, Cespi M, Palmieri GF (2020)
An overview of micro- and nanoemulsions as vehicles for essen-
tial oils: formulation, preparation and stability. Nanomaterials
(basel) 10:135. https:// doi. org/ 10. 3390/ nano1 00101 35
Periasamy VS, Athinarayanan J, Alshatwi AA (2016) Anticancer activ-
ity of an ultrasonic nanoemulsion formulation of Nigella sativa
L. essential oil on human breast cancer cells. Ultrason Sonochem
31:449–455. https:// doi. org/ 10. 1016/j. ultso nch. 2016. 01. 035
Rai M, Paralikar P, Jogee P, Agarkar G, Ingle AP, Derita M, Zacchino
S (2017) Synergistic antimicrobial potential of essential oils in
combination with nanoparticles: emerging trends and future per-
spectives. Int J Pharm 519:67–78. https:// doi. org/ 10. 1016/j. ijpha
rm. 2017. 01. 013
Environmental Science and Pollution Research
1 3
Roiz D, Wilson AL, Scott TW, Fonseca DM, Jourdain F,
Müller P, Velayudhan R, Corbel V (2018) Integrated Aedes
management for the control of Aedes-borne disease. PLoS
Negl Trop Dis 12:e0006845. https:// doi. org/ 10. 1371/ journ
al. pntd. 00068 45
Senthoorraja R, Subaharan K, Manjunath S, Pragadheesh
VS, Bakthavatsalam N, Mohan MG, Senthil-Nathan S,
Basavarajappa S (2021) Electrophysiological, behavioural
and biochemical effect of Ocimum basilicum oil and its
constituents methyl chavicol and linalool on Musca domestica
L. Environ Sci Pollut Res Int 8:1–14. https:// doi. org/ 10. 1007/
s11356- 021- 14282-x
Seo SM, Park HM, Park IK (2012) Larvicidal activity of ajowan (Tra-
chyspermum ammi) and Peru balsam (Myroxylon pereira) oils and
blends of their constituents against mosquito, Aedes aegypti, acute
toxicity on water flea, Daphnia magna, and aqueous residue. J
Agric Food Chem 60:5909–5914. https:// doi. org/ 10. 1021/ jf301
296d
Seo JW, Kim KJ, Kim SH, Hwang KM, Seok SH, Park ES (2015)
Effect of process parameters on formation and aggregation of
nanoparticles prepared with a Shirasu porous glass membrane.
Chem Pharm Bull (Tokyo) 63:792–798. https:// doi. org/ 10. 1248/
cpb. c15- 00297
Shepard DS, Undurraga EA, Halasa YA, Stanaway JD (2016) The
global economic burden of dengue: a systematic analysis. Lancet
Infect Dis 16:935–941. https:// doi. org/ 10. 1016/ S1473- 3099(16)
00146-8
Silva VB, Travassos DL, Nepel A, Barison A, Costa EV, Scotti L,
Scotti MT, Mendonça-Junior FJB, La Corte Dos Santos R, de
Holanda Cavalcanti SC (2017) Synthesis and chemometrics of
thymol and carvacrol derivatives as larvicides against Aedes
aegypti. J Arthropod Borne Dis 11:315–330
Singh S, Das SS, Singh G, Schuff C, de Lampasona MP, Catalan CA
(2014) Composition, invitro antioxidant and antimicrobial activi-
ties of essential oil and oleoresins obtained from black cumin
seeds (Nigella sativa L.). Biomed Res Int 2014:1–10. https:// doi.
org/ 10. 1155/ 2014/ 918209
Subaharan K, Senthoorraja R, Manjunath S, Thimmegowda GG, Pra-
gadheesh VS, Bakthavatsalam N, Mohan MG, Senthil-Nathan
S, David KJ, Basavarajappa S, Ballal C (2021) Toxicity, behav-
ioural and biochemical effect of Piper betle L. essential oil and
its constituents against housefly Musca domestica L. Pestic Bio-
chem Physiol 174:104804. https:// doi. org/ 10. 1016/j. pestbp. 2021.
104804
Sugumar S, Clarke SK, Nirmala MJ, Tyagi BK, Mukherjee A,
Chandrasekaran N (2014) Nanoemulsion of eucalyptus oil
and its larvicidal activity against Culex quinquefasciatus.
Bull Entomol Res 104:393–402. https:// doi. org/ 10. 1017/ S0007
48531 30007 10
Suyal J, Bhatt G, Singh N (2018) Formulation and evaluation of nanoemul-
sion for enhanced bioavailability of Itraconazole. Int J Pharm Sci Res
9:2927–31. https:// doi. org/ 10. 13040/ IJPSR. 0975- 8232. 9(7). 2927- 31
Thanigaivel A, Senthil-Nathan S, Vasantha-Srinivasan P, Edwin ES,
Ponsankar A, Selin-Rani S, Pradeepa V, Chellappandian M,
Kalaivani K, Abdel-Megeed A, Narayanan R, Murugan K (2017)
Chemicals isolated from Justicia adhatoda Linn reduce fitness
of the mosquito, Aedes aegypti L. Arch Insect Biochem Physiol
94(4): e21384. https:// doi. org/ 10. 1002/ arch. 21384
Thanigaivel A, Chanthini KM, Karthi S, Vasantha-Srinivasan P, Pon-
sankar A, Sivanesh H, Stanley-Raja V, Shyam-Sundar N, Naray-
anan KR, Senthil-Nathan S (2019) Toxic effect of essential oil
and its compounds isolated from Sphaeranthus amaranthoides
Burm. f. against dengue mosquito vector Aedes aegypti Linn.
Pestic Biochem Physiol 160:163–170. https:// doi. org/ 10. 1016/j.
pestbp. 2019. 08. 006
Weaver C, Costa F, Garcia-Blanco MA, Ko AI, Ribeiro GS, Saade G,
Shi PY, Vasilakis N (2016) Zika virus: history, emergence, biol-
ogy, and prospects for control. Antiviral Res 130:69–80. https://
doi. org/ 10. 1016/j. antiv iral. 2016. 03. 010
WHO (2005) Guidelines for laboratory and field testing of mosquito
larvicides. WHO/CDS/WHOPES/GCDPP/13. 1–39
Wilder-Smith A, Gubler DJ, Weaver SC, Monath TP, Heymann
DL, Scott TW (2017) Epidemic arboviral diseases: priorities
for research and public health. Lancet Infect Dis 17:e101–
e106. https:// doi. org/ 10. 1016/ S1473- 3099(16) 30518-7
Wu Z, Wei W, Cheng K, Zheng L, Ma C, Wang Y (2020) Insecticidal
activity of triterpenoids and volatile oil from the stems of Tetraena
mongolica. Pestic Biochem Physiol 166:104551. https:// doi. org/
10. 1016/j. pestbp. 2020. 02. 017
Youssefi MR, Tabari MA, Esfandiari A, Kazemi S, Moghadamnia
AA, Sut S, Dall-Acqua S, Benelli G, Maggi F (2019) Effi-
cacy of two monoterpenoids, carvacrol and thymol, and their
combinations against eggs and larvae of the West Nile vector
Culex pipiens. Molecules 24:1867. https:// doi. org/ 10. 3390/
molec ules2 41018 67
Publisher's note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
... 9 In the past two decades, essential oils (EOs) have gained attention as novel biopesticides, aiming to reduce or replace the use of chemical insecticides. 10 EOs offer environmentally-friendly advantages in pest control, such as biodegradability, safety, effectiveness, and delaying pest resistance. 11 EOs have been shown to be effective against Cx. ...
... 20 Ultrasonication is an economically viable, energy-efficient, and controllable technique for reducing droplet size from macroemulsions in NEs. 10 In the scientific literature, few studies analyzed the bioefficacy of the combination of EOs or their major compounds with CP on arthropod pest. Bullangpoti et al. ...
Article
Full-text available
Background The development of novel and ecofriendly tools plays an important role in insect pest management. Nanoemulsions (NEs) based on essential oils (EOs) offer a safer alternative for human health and the environment. This study aimed to elaborate and evaluate the toxicological effects of NEs containing peppermint or palmarosa EOs combined with β‐cypermethrin (β‐CP) using ultrasound technique. Results The optimized ratio of active ingredients to surfactant was 1:2. The NEs containing peppermint EO combined with β‐CP (NEs peppermint/β‐CP) were polydisperse with two peaks at 12.77 nm (33.4% intensity) and 299.1 nm (66.6% intensity). However, the NEs containing palmarosa EO combined with β‐CP (NEs palmarosa/β‐CP) were monodisperse with a size of 104.5 nm. Both NEs were transparent and stable for 2 months. The insecticidal effect of NEs was evaluated against Tribolium castaneum and Sitophilus oryzae adults, as well as Culex pipiens pipiens larvae. On all these insects, NEs peppermint/β‐CP enhanced pyrethroid bioactivity from 4.22‐ to 16‐folds while NEs palmarosa/β‐CP, from 3.90‐ to 10.6‐folds. Moreover, both NEs maintained high insecticidal activities against all insects for 2 months, although a slight increase of the particle size was detected. Conclusion The NEs elaborated in this work can be considered as highly promising formulations for the development of new insecticides. © 2023 Society of Chemical Industry.
... Pyrethroids, specially cypermethrin (CP), are the most widely used synthetic insecticides, since they had lower environmental impact than organochlorine, organophosphate and carbamates (Yadav et al., 2017). On the other hand, in the last 20 years, essential oils (EO) have received particular attention as novel biopesticides to replace or minimize the use of chemical insecticides (Subaharan et al., 2022). EO are effective against Cx. ...
... Generally, high-energy methods (high-pressure homogenizer, high shear homogenizer or ultrasonicator) use lower concentration of surfactant compared with low-energy methods (Jesser et al., 2020). The ultrasonication provide the energy force to reduce the droplets from the macroemulsions.Furthermore, the ultrasonic technique is economically viable, energy e cient, and offer better control over formulation variables (Subaharan et al., 2022 In the attempt to develop effective and safer insect control tools, the aim of this work was to incorporate CP and EO into a single nanoformulation. Consequently, NEs loaded with peppermint or palmarosa oils and β-CP were developed using an ultrasound probe. ...
Preprint
Full-text available
This work developed and evaluated the toxicological effects of single nanoemulsions (NEs) of peppermint or palmarosa essential oils (EO) combined with β-cypermethrin (CP) using ultrasound technique. For NEs formulation ultrasound parameters were: ultrasound power = 65 W, sonication time = 2 min and cycles = 30 on/20 off. The optimized active ingredients:surfactant ratio was 1:2. The NEs of peppermint EO combined with CP showed two peaks at 12.77 (33.4% intensity) and 299.1 nm (66.6% intensity) and polydisperse index (PDI) values of 1, while the NEs of palmarosa EO combined with CP had a size of 104.5 nm and their PDI value, 0.220. Both NEs were transparent and they were stable for 2 months. The insecticidal effect of these NEs (active ingredients:surfactant ratio was 1:2) were evaluated against T. castaneum and S. oryzae adults and Cx p. pipiens larvae. The NEs of peppermint EO combined with CP showed an enhancement of bioactivity of β-CP between 4.22 and 16 folds, while NEs of palmarosa EO + CP between 3.90 and 10.6 folds in all these insects. Moreover, both NEs demonstrated high insecticidal activity in all insect for 2 months, although slightly increase of particle size was detected. In this regard, the NEs elaborated in this work can be considered as highly promising formulation for the development of new insecticides.
... Following the method described by Subaharan et al., the EOs, added to a 0.1% Tween-80 solution (Shanghai Macklin Biochemical Technology, Shanghai, China), were subjected to 15 min of ultrasonication to emulsify the EOs and produce an EO microemulsion with a concentration of 4 µL/mL [44]. ...
Article
Full-text available
Tetranychus urticae, a prominent pest mite in strawberry and vegetable cultivation in China, has developed escalating resistance due to extensive chemical pesticide application. Consequently, there is an urgent need to identify safe and efficacious methods to reduce resistance development. In this study, 38 commercially available plant essential oils (EOs) were screened for their acaricidal potential and ability to inhibit oviposition. The findings revealed that 13 EOs exhibited notable acaricidal activity, with lemon EO demonstrating the highest toxicity, followed by sage, patchouli, frankincense, lemongrass, palmarosa, and oregano EOs. In addition, 18 EOs displayed significant inhibitory effects on oviposition, with lemon EO exhibiting the highest inhibition rate (99.15%) and inhibition index (0.98). Subsequently, sage, frankincense, clove, lemongrass, oregano, patchouli, myrrh, black pepper, palmarosa, and geranium EOs also showed inhibition rates exceeding 50%. Despite black pepper, clove, myrrh, and oregano EOs demonstrating relatively low toxicity against T. urticae, they exhibited heightened efficacy in inhibiting oviposition and suppressing population expansion. This study conducted a comparative assessment of the acaricidal and oviposition inhibition activities of EOs and their principal constituents, thus providing a theoretical basis for the development of botanical acaricides against T. urticae.
... The Global Vector Control Response (GVCR) 2017-2030 recognized by the WHO established strategic approaches for the management of mosquito vectors and their diseases (WHO 2017). Moreover, mosquito borne diseases are posing substantial risk to the public and alternate control measures without using chemical insecticides for the management of vectors (mosquito) at the breeding sites turned out to be very important (Subaharan et al., 2022;Opiyo et al., 2021). ...
... The essential oils were dissolved in a 0.1% Tween 80 water solution to prepare different concentrations of the solution [32], and the solution was emulsified by ultrasonic. A leafdip bioassay was conducted to evaluate the effects of 40 essential oils on field-collected populations of P. citri [33]. ...
Article
Full-text available
The citrus red mite, Panonychus citri (McGregor), is an important pest of citrus in China, where pesticides are commonly used in citrus orchards. In order to reduce the use of chemical pesticides against P. citri and the development of resistance, the screening of biological control agents has attracted the attention of students. In this study, seven plant essential oils with high toxicity were selected from 40 plant essential oils by the leaf-dip bioassay, including plant essential oils of lemongrass, patchouli, juniper berry, sage, clove, frankincense, and citrus. The LC50 after 72 h treatment were 3.198 μL/mL, 8.312 μL/mL, 3.244 μL/mL, 6.701 μL/mL, 8.350 μL/mL, 21.953 μL/mL, and 8.788 μL/mL, respectively. According to the LT50, the essential oils’ acute toxicity to P. citri from high to low were lemongrass, juniper berry, patchouli, citrus, sage, vetiver, and frankincense essential oils. In general, lemongrass and juniper berry essential oils have the best acaricidal effect and have high application value for the biological control of P. citri, which provides a basis for the development of botanical acaricides.
... Acaricidal and repellent properties of nanoemulsions based on different EOs against various pests were investigated (Echeverría and Albuquerque 2019). Subaharan et al. (2022) prepared a nanoemulsion containing 5% Trachyspermum ammi (L.) Sprague [Apiaceae] EO and 5% Tween 80 and tested toxicity against Aedes aegypti. The LC 50 values recorded after 24 h for bulk emulsion and nanoemulsion were 59.67 ppm and 46.73 ppm, respectively. ...
Article
This study was conducted to assess the repellent activity of Pogostemon cablin (Blanco) Benth. [Lamiaceae] essential oil (EO) and its nanoemulsion (NE) against Haemaphysalis longicornis nymphs. Pogostemon cablin EO nanoemulsion (PCNE) was prepared by ultrasonic emulsification using P. cablin EO and a surfactant, Tween 80. The results of gas chromatography–mass spectrometry analysis showed that the major component of P. cablin EO was patchouli alcohol (32.04%), followed by δ-guaiene (18.13%), α-guaiene (16.76%), seychellene (7.49%), and β-patchoulene (5.09%). The droplet sizes of PCNEs formulated using P. cablin EO and Tween 80 in ratios of 1:1, 1:2, and 1:3 were 23.8, 20.1 and 17.7 nm, respectively, indicating that the increase in the ratio of Tween 80 led to a decrease in droplet size. All essential oil preparations had an acceptable polydispersity index of 0.5 or less. The P. cablin EO and Tween 80 ratio of 1:3 was selected as the optimal formulation with the smallest droplet size; transmission electron microscopy indicated the formation of uniform spherical droplets. PCNE (84.0%) showed a higher repellent activity against H. longicornis nymphs than P. cablin EO (72.0%). The nanoemulsion is suitable for improving the repellent activity of P. cablin EO and is a potential repellent for H. longicornis.
... The effect of plant essential oils on insects is largely due to the presence of certain compounds, including mono and sesquiterpenoids in their composition. These compounds may act in certain ways to kill the insects via neurotoxicity or interact with one or more receptors in the nervous system of insects Subaharan et al., 2022). The terpenoids can also influence the activity of acetylcholinesterase enzyme as well as blocking octopamine receptors, GABA-gated chloride channels and nicotinic acetylcholine receptor (Isman, 2020). ...
Article
Extensive usage of synthetic pesticides has proved to be destructive to all living being and the resurgence of pest resistance. Compounds derived from certain plants are usually safer compared to chemical control of pest. The present study thus intended to use Thymus vulgaris essential oil (EO) and two of its derivatives including thymol and carvacrol in order to see their deleterious effects on Glyphodes pyloalis (Walker). We also studied the oil components. This pest has recently become a serious concern for the silk industry. Our results showed that the thyme EO contain several components including thymol (26.9%), ρ-Cymene (14.54%), linalool (13.39%) and carvacrol (5.7%). Our toxicity tests revealed an estimated LD50 values for thyme EO, thymol and carvacrol 2.82, 32.18 and 56.54 μg/larva, respectively. However, the thyme EO was more toxic than its two tested compounds. The activity of certain detoxifying enzymes such as α- and β-esterase, glutathione S-transferase and cytochrome P450 were significantly inhibited by thymol-treated larvae compared to the control group. Similarly, the activity of alanine aminotransferase, aspartate aminotransferase, lactate dehydrogenase and alkaline phosphatases enzymes in thymol-treated larvae decreased while the activity of acid phosphatases increased. Our results suggest that thyme EO and its components have potential for the control of G. pyloalis larvae in mulberry orchards, where no synthetic chemicals are allowed.
Article
The world's population is continuously increasing; therefore, food availability will be one of the major concerns of our future. In addition to that, many practices and products used, such as pesticides and fertilizers have been shown harmful to the environment and human health and are assumed as being one of the main factors responsible for the loss of biodiversity. Also, climate change could agravate the problem since it causes unpredictable variation of local and regional climate conditions,which frequently favor the growth of diseases, pathogens and pest growth. The use of natural products, like essential oils, plant extracts, or substances of microbial-origin in combination with nanotechnology is one suitable way to outgrow this problem. The most often employed natural products in research studies to date include pyrethrum extract, neem oil, and various essential oils, which when enclosed shown increased resistance to environmental factors. They also demonstrated insecticidal, antibacterial, and fungicidal properties. However, in order to truly determine if these products, despite being natural, would be hazardous or not, testing in non-target organisms, which are rare, must start to become a common practice. Therefore, this review aims to present the existing literature concerning nanoformulations of biopesticides and a standard definition for nanobiopesticides, their synthesis methods and their possible ecotoxicological impacts, while discussing the regulatory aspects regarding their authorization and commercialization. As a result of this, you will find a critical analysis in this reading. The most obvious findings are that i) there are insufficient reliable ecotoxicological data for risk assessment purposes and to establish safety doses; and ii) the requirements for registration and authorization of these new products are not as straightforward as those for synthetic chemicals and take a lot of time, which is a major challenge/limitation in terms of the goals set by the Farm to Fork initiative.
Article
Full-text available
Ocimum basilicum essential oil (EO) was evaluated for its biological effects on M. domestica. Characterization of O. basilicum EO revealed the presence of methyl chavicol (70.93%), linalool (9.34%), epi-α-cadinol (3.69 %), methyl eugenol (2.48%), γ-cadinene (1.67%), 1,8-cineole (1.30%) and (E)-β-ocimene (1.11%). The basil EO and its constituents methyl chavicol and linalool elicited a neuronal response in female adults of M. domestica. Adult female flies showed reduced preference to food source laced with basil EO and methyl chavicol. Substrates treated with EO and methyl chavicol at 0.25% resulted in an oviposition deterrence of over 80%. A large ovicidal effect was found for O. basilicum EO (EC50 9.74 mg/dm3) followed by methyl chavicol (EC50 10.67 mg/dm3) and linalool (EC50 13.57 mg/dm3). Adults exposed to EO (LD50 10.01 μg/adult) were more susceptible to contact toxicity than to methyl chavicol and linalool (LD50 13.62 μg/adult and LD50 43.12 μg/adult respectively). EO and its constituents methyl chavicol and linalool also induced the detoxifying enzymes Carboxyl esterase (Car E) and Glutathione S – transferases (GST).
Article
Full-text available
In an attempt to reduce the massive application of the toxic chemical pesticides, essential oils (EOs) of Achillea biebersteinii and Juniperus procera were obtained through hydrodistillation and analyzed using (GC-FID) and (GC-MS). α-terpinene and p-cymene were detected as the major components in the EO of A. biebersteinii, while eugenol and ß-caryophyllene were the major constituents in the EO of J. procera. The plant EOs and major fractions act as considerable mosquitocides against Aedes aegypti L, the common transmitter of Dengue fever. The EOs and major fractions were tested at 6.25, 12.5, 25, 50 and 100 μl/l. Insect mortality was time and dose-dependent, and the adult stage was more sensitive than larvae. At a concentration of 50 μl/l, 24 post treatment larval and adult mortality ranged between (40.3 and 89.3%) and (51.4 and 95.6%), respectively. The LC50 values ranged between 12.2 and 70.1 μl/l against larvae and between 10.1 and 63.12 μl/l against adults. All of the crude EOs were more potent than their major fractions. Eugenol and ß-caryophyllene showed strong mosquitocidal activity than p-cymene and α-terpinene. The corrected percentage mortality was increased over time with all of the test materials. In terms of lethal time required to kill 50% of the population (LT50), a concentration of 100 μl/l of J. procera EO showed LT50 values of 2.3 and 1.7 h against larvae and adult, respectively. The EOs induced considerable inhibition of acetylcholinesterase activity, where J. procera crude oil (IC50 = 13.12mM) and eugenol (IC50 = 19.65mM) were the most potent. Results proved that the test plant oils and their major fractions could be developed as natural pest control agents to control A. aegypti.
Article
Full-text available
Oil in water (o/w) emulsions stabilized by an amphiphilic copolymer have been studied in relation to their potential insecticidal activity against Aedes aegypti mosquito larvae. These emulsions contain as oil phase different blends of two isomeric essential oil compounds, thymol and carvacrol. The results show that the addition of carvacrol facilitates the dispersion of the oil within the aqueous phase, with the stabilization and polydispersity of the emulsions being controlled by the change of the ratio between the copolymer concentration and that of the oil phase (Rcop/EOC). Emulsions containing pure essential oil compounds as oil phase do not present any significant difference on their larvicidal activity against mosquito larvae, with emulsions containing only thymol being slightly more effective than those containing only carvacrol as oil phase. Furthermore, the use of blends containing different weight fractions of thymol and carvacrol as oil phase results in formulations with an additive larvicidal activity in relation to those with the pure compounds. Despite the larvicidal activity of the emulsions, they do not provoke inhibition to the emergence of adult individuals in Aedes aegypti populations. The spreading and evaporation of the emulsions onto solid surface, which may be an important parameter for the performance of larvicidal formulations, was found to be dependent on the same parameters that govern the stability of the emulsions. This study helps on seeking new alternatives for the preparation of new eco-sustainable formulations against insect pest.
Article
Housefly, Musca domestica L. is a pest of public health importance and is responsible for spreading diseases like typhoid, diarrhoea, plague etc. Indiscriminate reliance on synthetic insecticides has led to development of insecticide resistance and ill effect to humans and nontarget animals. This demands an alternative and safer pest control option. This study evaluates the biological effect of Piper betle L essential oil and its constituent eugenol, eugenol acetate, and β - caryophyllene on the housefly. The major components present in P. betel EO were safrole (44.25%), eugenol (5.16%), β -caryophyllene (5.98%), β -selinene (5.93%), α-selinene (5.27%) and eugenol acetate (9.77%). Eugenol caused 4.5fold higher ovicidal activity (EC50 86.99 μg/ml) than P. betle EO (EC50 390.37 μg/ml). Eugenol caused fumigant toxicity to adults (LC50 88.38 mg/dm³). On contact toxicity by topical application, eugenol acetate, eugenol and β-caryophyllene caused higher mortality to larval and adult stages than EO. FESEM (Field Emission Scanning Electron Microscope) images reveal that exposure to P. betle EO causes the shrinkage of the larval cuticle. Both EO and eugenol induced the detoxifying enzymes Carboxyl esterase (Car E) and Glutathione S – transferases (GST) in larvae and adults. EO and eugenol at 0.2% caused effective repellence and oviposition deterrence to M. domestica adults and this merits their use as alternative strategy to manage M. domestica.
Article
Previous work presented the profound antimosquito potential of Petroselinum crispum essential oil (PEO) against either the pyrethroid-susceptible or resistant strains of Aedes aegypti. This plant oil also inhibited the activity of acetylcholinesterase and mixed-function oxidases significantly, thus suggesting its potential as a synergist for improving mosquitocidal efficacy of insecticidal formulations. This study investigated the chemical composition, larvicidal activity, and potential synergism with synthetic insecticides of PEO and its main compounds for the purpose of interacting with insecticide resistance in mosquito vectors. The chemical profile of PEO, obtained by GC-MS analysis, showed a total of 17 bioactive compounds, accounting for 99.09% of the whole oil, with the most dominant constituents being thymol (74.57%), p-cymene (10.73%), and γ-terpinene (8.34%). All PEO constituents exhibited promising larvicidal effects, with LC50 values ranging from 19.47 to 59.75 ppm against Ae. aegypti, in both the pyrethroid-susceptible and resistant strains. Furthermore, combination-based bioassays revealed that PEO, thymol, p-cymene, and γ-terpinene enhanced the efficacy of temephos and deltamethrin significantly. The most effective synergist with temephos was PEO, which reduced LC50 values to 2.73, 4.94, and 3.28 ppb against MCM-S, PMD-R, and UPK-R, respectively, with synergism ratio (SR) values of 1.33, 1.38, and 2.12, respectively. The best synergist with deltamethrin also was PEO, which reduced LC50 values against MCM-S, PMD-R, and UPK-R to 0.008, 0.18, and 2.49 ppb, respectively, with SR values of 21.25, 9.00, and 4.06, respectively. This research promoted the potential for using essential oil and its principal constituents as not only alternative larvicides, but also attractive synergists for enhancing efficacy of existing conventional insecticides.
Article
There are growing concerns about the extensive use of synthetic fungicides worldwide due to their negative health effect on humans, ecosystem, and the development of fungal resistance. Therefore, developing new and safe eco-friendly antifungal agents such as plant-based oils is of major interest. This study was conducted for the first time to develop and evaluate the efficiency of eco-friendly nanoemulsions of some natural oils and evaluating of its efficiency against postharvest fruit rot of cucumber caused by the following fungal isolates: Galactomyces candidum (MF373433), Alternaria tenuissima (MF 373440) and Fusarium solani (MF373443). Nanoemulsions of clove, black seed, lemon, and orange oils were formulated by using ultrasound and subjected to physicochemical and stability studies. The efficiency of these nanoemulsions was evaluated against the postharvest fruit rot fungi of cucumber. The oils of promising nanoemulsions have been analyzed by gas chromatography-mass spectrometry analysis (GC/MS). The acute toxicity study was conducted on rats and on the luminescence emitted of Vibrio fischeri bacteriausing Microtox®. The smaller droplet size of nanoemulsions was found after 10 min of sonication with an oil to Tween 80 ratio of 1:1 and 1:2 for clove and black seed oils, respectively. Nanoemulsion of black seed and clove oils from Egypt at 5000 ppm showed a significantly high reduction in the linear mycelial growth of fungal isolates. The mixture formula (2:1 v/v) of nanoemulsions of clove and black seed oils showed a significantly high reduction in linear mycelial growth, completely inhibited (100%) conidia sporulation of fungal isolates, and highly reduced the percentage of the rot of germinated seeds as well mortality of cucumber developed seedlings. Nanoemulsion formulation of clove and black seed (2:1) at concentration 2.0 % as soaking treatment of cucumber fruits before artificial infestation by inocula of fungal pathogens caused completely (100%) suppress postharvest fruit rot of cucumber. Both clove and black seed oils nanoemulsions and their combination showed no signs of toxicity or mortality in rats and no toxic effects on the luminescence emitted of Vibrio fischeri bacteria using Microtox® assay. Therefore, it can be concluded that the nanoformulation of clove and black seed oils at the ratio of 2:1 is a promising green and safe alternative to synthetic fungicide for application against postharvest rot of cucumber fruits.
Article
Kaempferia galangal L. (Zingiberaceae), also known as aromatic ginger, is used in the food and cosmetics, pharmaceuticals and maquillage materials, as well as in the traditional Indian medicine. The essential oil (EO) obtained from its rhizome by hydrodistillation is of economic interest and contains ethyl p-methoxycinnamate as the main component, which is used as an anticancer, antimicrobial, skin protector and whiteneing and nematicidal agent. To our knowledge, the mosquitocidal properties of the aromatic ginger EO have hitherto been poorly investigated. Therefore, in the present work, we tested this EO and its major chemical constituents (MCCs), namely ethyl p-methoxycinnamate, trans-ethyl cinnamate and trans-cinnamaldehyde for their larval toxicity on Aedes vittatus and Anopheles maculatus. The EO chemical composition, studied by GC-MS analysis, was made up of 39 compounds, with ethyl p-methoxycinnamate (30.6%), trans-ethyl cinnamate (26.8%) and trans-cinnamaldehyde (11.5%) as the major constituents. The EO and its MCCs showed LC50 values of 39.22, 10.36, 28.26 and 36.35 μg/ml on Ae. vittatus, and of 41.36, 12.56, 30.23 and 38.47 μg/ml on An. maculatus third instar larvae, respectively. The ecofriendliness of EO and MCCs on the aquatic fauna was demonstrated as they were safe to Anisops bouvieri, Acilius sulcatus and Gambusia affinis. Our findings showed that the K. galanga rhizome EO and its MCCs are promising larvicidal agents without affecting non-target organisms (NTOs). They also give news insights into the potential economic exploitation of this EO and its MCCs in the fabrication of effective and eco-friendly mosquitocidal formulations.
Article
Essential oils are promising substitute for chemical pesticides with the inherent resistance by pests, environmental and health effects on humans. In this study, the chemical composition of essential oil extracted from Citrus sinensis peel was characterized, the insecticidal activities of the oil and its constituents against Callosobrunchus maculatus (Cowpea weevil) and Sitophilus zeamais (maize weevil) were investigated and the underlying insecticidal mechanism were elucidated. The essential oil was extracted by hydro-distillation and characterized using gas chromatography–mass spectrometry (GC–MS). Insecticidal activity was determined by contact and fumigant toxicity assay. The inhibitory effect of the oil and its constituents on acetylcholinesterase (AChE), Na⁺/K⁺-ATPase and glutathione-S- transferase (GST) activity were assayed using standard protocols. The total number of volatile compounds detected in C. sinensis essential oil was eighteen (18). d-limonene (59.3%), terpineol (8.31%) and linalool (6.88%) were the major compounds present in the essential oil. Among the tested essential oil compounds, terpineol showed highest contact toxicity against C. maculatus (LD50 = 17.05 μg/adult) while 3-carene showed highest contact toxicity against S. zeamais (LD50 = 26.01 μg/adult) at 24 h exposure time. Citral exhibited the highest fumigant toxicity against C. maculatus and S. zeamais with LC50 value 0.19 and 2.02 μL/L air at 24 h respectively. Acetylcholinesterase and Na⁺/K⁺-ATPase activities were significantly inhibited by C. sinensis oil and its constituents in both C. maculatus and S. zeamais as compared to control. This study indicates that C. sinensis essential oil and its constituents have potential to be developed into botanical pesticides.
Article
The use of chemical pesticides to preserve food commodities is a global issue of concern due to their negative effect on the environment and public health. In recent years, the European Union is trying to reduce their use, favoring alternative or complementary approaches based on natural products. In this respect, plant-borne essential oils (EOs) represent valid options for Integrated Pest Management (IPM) programs. In the present study, the insecticidal effect of eight EOs obtained from plants from different parts of the world, namely Mentha longifolia, Dysphania ambrosioides, Carlina acaulis, Trachyspermum ammi, Pimpinella anisum, Origanum syriacum, Cannabis sativa and Hazomalania voyronii, were evaluated against two stored-product insect species of economic importance, Prostephanus truncatus and Trogoderma granarium. Simulating a small-scale stored product conservation environment, an AG-4 airbrush was used to spray maize and wheat with 500 and 1000 ppm of EOs, then T. granarium and P. truncatus were exposed to the stored products and mortality was evaluated over selected time intervals (4, 8, and 16 h, and 1, 2, 3, 4, 5, 6, and 7 days). The EO of C. acaulis exhibited high efficacy against P. truncatus adults at both tested concentrations by killing >97% of the individuals exposed to treated maize within 3 days at 500 ppm. The EO of D. ambrosioides eliminated all T. granarium adults exposed to 1000 ppm-treated wheat 2 days post-exposure. At this exposure interval, 91.1% of the exposed T. granarium adults died on wheat treated with 1000 ppm of C. acaulis EO. The EO of M. longifolia at both tested concentrations was the most effective against T. granarium larvae, leading to 97.8% mortality at 500 ppm after 3 days of exposure, and 100% mortality at 1000 pm 2 days post-exposure. At 1000 ppm, the EOs of D. ambrosioides and P. anisum led to 95.6 and 90% mortality, respectively, to larvae exposed to treated wheat for 7 days. Overall, our research shed light on the potential of selected EOs, with special reference to M. longifolia, D. ambrosioides, C. acaulis and P. anisum, which could be considered further to develop effective and alternative grain protectants to manage P. truncatus and T. granarium infestations.
Article
Tetraena mongolica Maxim is a species of Zygophyllaceae endemic to China. Because few insect pests affect its growth and flowering, we speculated that this plant produces defensive chemicals that are insect repellents or antifeedants. The effects of different fractions from crude stem and leaf extracts on Pieris rapae were examined. The results confirmed that the ethyl acetate (EtOAc) fraction from the stems had insecticidal potential. Five compounds were isolated from the EtOAc fraction: a volatile oil [bis(2-ethylhexyl) benzene-1,2-dicarboxylate (1)], three triterpenoids 2E-3β-(3,4-dihydroxycinnamoyl)-erythrodiol (2), 2Z-3β-(3,4-dihydroxycinnamoyl)-erythrodiol (3), and 2E-3β-(3,4-dihydroxyphenyl)-2-propenoate (4)], and one steroid [β-sitosterol (5)]. Compounds 1–5 exhibited different degrees of insecticidal activity, including antifeedant and growth-inhibition effects. Compounds 1–5 inhibited the activity of carboxylesterase (CarE) and acetylcholinesterase (AChE) to different degrees. Compound 1 had the strongest antifeedant and growth-inhibition effects, and significantly inhibited the activity of CarE and AChE. Our results indicate that compounds 1–4 are the major bioactive insecticidal constituents of Tetraena mongolica. This work should facilitate the development and application of plant-derived botanical pesticides.