A Study On The Antipathogenic Effects of Nanoemuls

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

BioNanoScience

https://doi.org/10.1007/s12668-024-01375-3

RESEARCH

A Study on the Antipathogenic Effects of Nanoemulsion Against


V. parahaemolyticus in Shrimp Aquaculture: Antibacterial
and Antibiofilm Activities
Jahangir Ahmed1 · Karthikeyan Ramalingam1

Accepted: 24 March 2024


© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2024

Abstract
Antimicrobial resistance is a significant concern in aquaculture, prompting the exploration of nanoemulsions (NEs) with
a non-resistant mode of action as a safer and more effective alternative to antibiotics. In Asian countries, the culture of
Penaeus vannamei has faced growing concerns as V. parahaemolyticus (Vp) has been increasingly linked to shrimp diseases,
leading to huge economic losses. In response, stable ozonated oil-in-water NEs (NE-25, 26, 28, 29) were produced using a
microfluidization method and characterized by their polydispersity index and average droplet size (NE-25, 26, 28, 29), which
were found to 0.108, 0.251, 0.223, 0.173, and 229.7nm, 190.2nm, 201.8nm, and 145.8nm, respectively. The therapeutic
potential of NEs is tested against pathogenic strains having tdh and toxR virulence gene, which were isolated from diseased
shrimp (VP-S7 and VP-S14), these were found to be more resistant compared to commercial V. parahaemolyticus (MTCC-
451) strain. Our NEs confirmatory antibacterial tests include bacteriostatic, bactericidal, antibiofilm, adherence, live/dead
assays, and oxygen consumption rate (OCR) studies. All tested nanoemulsions showed effectiveness against planktonic and
biofilm stages of V. parahaemolyticus; NE-25 showed strong adherence inhibition, with results of VP-S7 (49.55%), VP-S14
(53.47%), and MTCC-451 (68.76%). The biofilm inhibition of NE-25 (55.97%) is comparable to inhibition with antibiotics
gentamicin (58.73%). The NE OCR results (1 × 107 cell/ml) compared to antibiotics treated and untreated, the fold reduc-
tion for NE-28 (0.2) is highly significant when compared with commercial antibiotics (0.6) and the growth control (1.0).
Therefore, using NEs, which possess high antibacterial properties, is a promising alternative for treating V. parahaemolyticus
compared to current antibiotics.

Keywords Biocidal alternative · Disease prevention · Nanotechnology · Oxygen consumption rate · Penaeus vannamei ·
Vibrio pathogens

1 Introduction shrimp. In 2018, 9.4 million tons of crustaceans were pro-


duced through aquaculture worldwide, producing USD 69.3
In Asian countries, aquaculture is the most important and billion [2, 3]. Vibriosis is one of the most significant issues
rapidly developing industry, dramatically increasing the caused by a group of bacteria faced by the crustacean indus-
demand for expensive shrimp goods on the global market try. Early mortality syndrome (EMS) occurs within 35 days
[1]. Commercial shrimp production, which was dominated after stocking P. monodon in grow-out ponds, resulting in
by two species of penaeid shrimp in 2015, observed that severe mortality [4]. Vibriosis in India linked history back to
95% of the shrimp come from either the Penaeus monodon Vibrio spp., causing shell disease, white gut, and loose shell
(giant tiger shrimp) or the exotic P. vannamei (white-leg) syndrome in the coastal area [5]. In addition, common vibrio
illnesses caused by Vibrio spp. (Vibrio harveyi, V. campbel-
lii, and V. parahaemolyticus), luminous vibriosis (Vibrio fis-
* Karthikeyan Ramalingam cheri, V. harveyi, V. cholerae) produced in shrimp farms as
karthikeyan.sls@crescent.education mortality led to 50–100% loss to the shrimp industry, bil-
1 lions of dollars worldwide yearly [6–8]. These Vibrionaceae
School of Life Sciences, B.S. Abdur Rahman Crescent
Institute of Science and Technology, Chennai, Tamil Nadu, family facultative anaerobic bacteria frequently cause total
India mortality in prawn hatcheries and grow-out ponds [7].

Vol.:(0123456789)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

Following a disease outbreak and mass mortality on 2 Material and Methods


India’s east coast from October to November 2013, cultures
of V. parahaemolyticus were obtained from the haemolymph 2.1 Bacterial strains and Growth Conditions
of moribund shrimp. Independently, V. parahaemolyticus
strains cause mass mortalities of cultured shrimp by activat- The strain commercial V. parahaemolyticus (MTCC-451)
ing virulent genes like a transmembrane transcriptional regu- was obtained from the Microbial Type Culture Collection
lator gene (toxR), thermostable direct haemolysin (tdh), and (MTCC), Chandigarh, India. A shrimp disease-causing strain
Thermolabile haemolysin gene (tlh) expression of virulence (VP-S7 and VP-S14) was acquired from the Central Institute
factors [9–11]. Among bacterial lethal infections since 2009, of Brackishwater Aquaculture in Chennai (Tamil Naidu),
acute hepatopancreatic necrosis disease (AHPND), caused by India [28]. The sunflower (SF), castor oil (CA), and olive
Photorhabdus insect-related toxins (PirA/B), is a lethal illness oil (OL) are purchased from the local market in Tambaram,
that has risen in Southeast Asian nations [12, 13]. In such a Chennai, Tamil Naidu, India. The surfactant Polysorbate 60
case, various chemicals and antibiotics were used to quickly (Tween-60), polyethylene glycol dodecyl ether (Brij-30), and
control the bacterial infection. Antibiotic overuse in the aqua- co-surfactant cetylpyridinium chloride, agar, and tryptose
culture industry has resulted in momentous economic loss soya broth (TSB) were purchased from HI-media and Sisco
and the establishment of antibiotic-resistant bacterial strains Research Laboratories (SRL) Mumbai, India. The bacterial
[14, 15]. Therefore, finding novel treatment routes is a pre- cultures were preserved in TSB with 50% (v/v) glycerol
requisite for standard antibiotics treatment and guarantees at − 80°C. Inoculated stock mother cultures were cultivated
more sustainable growth for the aquaculture industry [16]. at 37°C in sterile, fresh broths containing 1.5% NaCl. Fresh
The NE’s physicochemical properties, including the inoculums were utilized in every experiment; the bacterium
oil phase, aqueous phase, and emulsifiers, determine how cultured optical density was measured at 600nm, which
NEs form and stabilize [17]. Because of their effective interprets a bacterial concentration of about ­107 CFU/mL.
encapsulation, solubilization, targeted distribution, and
bioavailability, food-grade NEs are increasingly focused in the
feed industry. The nanosized NEs are regarded as one of the
2.2 Production and Preparation of Nanoemulsions
crucial carriers for the maintenance release of phytocompounds
According to earlier studies, four nanoemulsions (NEs) were
[18, 19]. Typically, emulsion-based delivery safeguard from
synthesized and processed as oils in water (O/W) combina-
harmful environmental factors, improve product stability
tions [29]. Nanoemulsion composition was used for one set
during storage, and preserve their functional activity [20].
of combinations such as castor oil (25%) and olive oil (25%)
When lipophilic compounds are encapsulated within NEs,
with Polysorbate 60 (Tween-60–2.5%) as surfactant and
such as Docosahexaenoic acid [21], fish oil [22], beta-carotene
cetylpyridinium chloride (CPC-1%) as co-surfactant. The
[23], curcumin [24], etc., they have a more significant ability
sunflower oil (20%) used in another set of combinations,
to boost bioavailability. Due to their increased bioactivity and
such as NE-28 (Epigallocatechin-gallate-EGCG, 0.4%) and
widespread use in antibacterial, disinfectants, and antiseptics,
NE-29 (Taxifolin-TF, 0.5%) with Brij-30 (3%) as well as
NEs can also be utilized as an antimicrobial agent [25, 26].
CPC (0.5%) was used.
The anti-biofilm and anti-adherence feature of the NEs
The Microfluidizer-LM10 was used to emulsify all NEs
recommends that they may also help develop anticarcinogenic
(Microfluidics Int. Corporation, USA). The solutions were
agents [27].
mixed thoroughly using a magnetic stirrer, until samples
In vitro screening of NEs was performed parallel to
were mixed well with the oil, and the remaining sterile
understand the effect of the administrated dose on three
water ­(DH2O) was added gently to create a homogeneous,
vibrio pathogens’ growth. Our study focuses on developing
transparent solution. Finally, the solution was emulsified at
antibacterial (ozone-enriched oil used) NEs from phytocom-
20,000psi lb/in2, three times at room temperature (one set),
pounds such as epigallocatechin-gallate, and taxifolin, with
whereas another set was produced in the presence of ice
sunflower oil in this set and the next set of olive oil and
(phytocompounds). The Schematic representation is shown
castor oil. The antibacterial NEs synthesized using high-
in Fig. 1. Using a magnetic stirrer, the samples were mixed
pressure micro-fluidization (20,000 psi) techniques. The
well with the oil, and the remaining sterile water solution
oil combinations are said to be tested as a unique class of
was added gently to create a homogeneous, transparent
NEs, because of better antibacterial potential confirmed by
solution. The prepared nanoemulsion was further character-
MIC followed by MBC for further evaluation studies. The
ized by dynamic light scattering for size and zeta-potential
antibiofilm, anti-adherence, and live/dead studies were done
analysis.
by slanted glass coverslip (SGC) methods, and the oxygen
consumption rate was analyzed independently.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

Fig. 1  Schematic representation


of the synthesis of ozonated
oil nanoemulsion, charac-
terization (DLS), and antibac-
terial efficacy tests against V.
parahaemolyticus strains from
diseased shrimp (P. vannamei)
culture

2.3 Fourier Transform Infrared Spectroscopy visible turbidity acting as a growth inhibitor NEs well of the
of Synthesized Nanoemulsions microbes [31]. Following the incubation time, the concentra-
tions of bactericidal nanoemulsions were also determined in
The oils (castor, olive, and sunflower) nanoemulsions were the poor turbid and clear wells. To provide more detail, each
prepared, and FTIR analysis was used to examine whether well was chosen, a 5µl solution was extracted, and doublet
the active ingredients involved in the optimized nanoemul- spots of the suspension were placed on T-SB agar plates.
sion composition were compatible. The prepared nanoemul- These plates were then incubated for 24 h at 37°C [29]. The
sion’s FTIR spectra were performed on a JASCO (JASCO maximum suspension of NEs that resulted in a (99.9%) bac-
INFRARED SPECTRUM) FTIR-spectrophotometer. The terial cell count, or the endpoint of the effective reductions
nanoemulsion’s FTIR spectra were scanned at A-399–4000 of the microbial population, was noted as the bactericidal
­cm−1 [30]. concentration in terms of the solution.

2.4 MIC and MBC Against V. Parahaemolyticus 2.5 Slanted Coverslip Method for Biofilm
Pathogens Quantification

The bacteriostatic and bactericidal approach used the micro- For the slanted glass coverslip (SGC) method, a coverslip
dilution method in a microtiter (96-well) plate to assess the (18 × 18 mm) was used to develop a biofilm. Because the
antibacterial properties of produced nanoemulsions against coverslip (18 mm × 18 mm) bulges roughly one-third of its
V. parahaemolyticus pathogenic strains. Briefly, 100µl of length above the level of the medium, the formation of a
(2 ×) TS-broth and 100µl of NEs from each group were biofilm could begin at the level of the medium on the upper
added to the initial well and orderly diluted up to the 12th surface. After that, a consistent biofilm spread across the
well. In the next step, 5µl of overnight active bacteria with entire coverslip developed both below (into the medium)
­107 CFU/ml of the mixed solution was added to each well, and above this level of the coverslip. To evaluate the biofilm
except the negative (media) control. As a negative control, yield, overnight bacterial cultures containing ­107 CFU/ml
sterile ultrapure water ­(UPH2O) containing medium without of different strains of V. parahaemolyticus (MTCC-451),
bacteria and positive control (gentamicin) media with bacte- shrimp isolated (VP-S7 and VP-S14), were added to 2.5 ml
ria in an identical volume were maintained, respectively. A of media (T-SB) in 12-well microtiter plates with inserted
microtiter plate containing four distinct NE-treated V. para- slanted glass coverslips. Every 12 h, the media containing
haemolyticus strains and an untreated control was incubated the suspended bacterial cells were removed, and an equiva-
at 37° for 24 h. The greatest dilutions that exhibit no bac- lent volume of fresh sterile medium was added [32, 33].
terial growth were used to calculate the MIC, with the no After 72 h, the grown biofilm was treated with NEs in an

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

equal volume covering the biofilm completely in an aseptic 2.7 Live/Dead Assay


condition for a fixed period. After incubation, the glass cov-
erslips and the well plate with the biofilm-developed bacteria Slanted glass coverslip (SGC) was used to build the bio-
were placed for 15 min in 2.5 ml of methanol in each well. films for 72 h, stained with the LIVE/DEAD-BacLight,
After, the well plate was emptied and left to dry naturally, and then treated with ozone-enriched oil-based nanoemul-
with and without glass coverslips. Each well was stained for sions (OZNEs). A multilabel-detector VICTOR X3 (Perkin
5 min with a 0.1% (w/v) crystal violet solution. The plates Elmer) was used to measure the fluorescence intensity [36].
were thoroughly cleaned of extra stains using wash bottles The staining solution containing the components SYTO-9
in a controlled setting. Using glacial acetic acid, 33% (v/v), was mixed in (0.1µl, SYTO-9 can enter all cells and is used
the dye bound to adherent cells was removed from the well for assessing total cell counts), quantities from stock solu-
plate with and without glass coverslips. Finally, the optical tions in each well and maintained in an aseptic condition
densities of the resultant solutions (590nm) were measured for a further 1 h [37]. The untreated bacterial cells in the
[34]. Under each experimental condition, three different tri- untreated wells were designated as controls and counted to
als were conducted. The mean optical density of the negative remove the absorbance readings of lives and dead cells. For
control was subtracted from the readings of all other samples each experimental condition, three separate experiments
in each case. were conducted. In each case, the mean optical density of
the negative control was subtracted from all other sample
results recorded.
2.6 Adherence Assay by Slanted Coverslip Method
2.8 Oxygen Consumption Assay
The adherence inhibition properties of V. parahaemolyti-
cus pathogens grown on the slanted glass coverslip (SGC- The measurement of oxygen consumption rates (OCR) was
18 × 18mm) were performed according to earlier studies [33], measured using a Clarke electrode-based oxygraph instru-
with little modifications. On selected bacteria, the bacte- ment (Hansatech, Norfolk, UK) following the manufactur-
rial adherent behavior was conducted: V. parahaemolyticus er’s instructions. The cells in the exponential growth phase
(MTCC-451) pathogens and two shrimp isolated (VP-S7 and were subjected to centrifugation at 8000 rpm for 2–3 min
VP-S14). Based on predetermined results of MBC, NE-25, and then resuspended in fresh TS-broth with or without the
NE-26, NE-28, and NE-29 were used for the anti-adherence test compounds, such as NE-25, 26, 28, and 29, and antibi-
assay. In 12-well plates with slanted glass coverslips put at otics like gentamicin at a concentration of approximately
their appropriate Sub-MBC dilutions for the MTCC and 1 × ­107 cells. The obtained results were analyzed using ­O2
shrimp isolate strains, overnight-grown bacterial cultures view software version 2.0, and the rates were expressed as
containing ­107 cells/ml, were introduced. Slanted glass cov- nanomoles of O ­ 2 consumed per 1 × ­107 cells/ml/min, based
erslips were used to maintain a total volume of 2.5 ml in each on the findings of previous studies [38].
well, which included positive and negative (sterilized water)
controls [34]. For 1h, the plates were incubated at 37°C. Fol- 3 Statistical Analysis
lowing the fixation procedure is the removing of the residual
medium once the incubation period is complete and using 5 To determine the significance level within each row, statisti-
ml of 100% methanol per well for 15 min. Following medium cal analyses were reviewed using Dunnett’s multiple com-
emptying and air drying, adhering layers remained on the fixa- parisons test., simple effects within the rows, and compara-
tion at the tilted glass coverslip surface. The well was stained ble columns. The two-way and one-way ANOVA method
for 5 min with 3ml/well of 0.1% (w/v) crystal violet. The wells was used to analyze variance for experimental data, and
were uniformly washed using a water bottle to eliminate the statistical significance was considered as P values < 0.05.
extra stains. After air drying, the coverslip was taken out and
transferred to new wells (microtiter 12-well plates), the num-
ber of labelled adhering cells was quantified, and 3mL of 33% 4 Results and Discussions
(v/v) glacial acetic acid was added to each well [35]. Each
experiment was performed in three independent sets for each 4.1 Nanoemulsion Production and Preparation
repeat to conclude the results. The bound bacteria’s solution
was measured using a multilabel-detector (Perkin Elmer) at After passing the solution through Microf luidizer-
590nm. To subtract the reading from all experiment sets, the LM10, the visual appearance of the emulsions changed
­UPH2O (negative control) was used as a blank.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

noticeably, making them physically transparent and leav- 4.3 Fourier Transform Infrared Spectroscopy
ing no turbidity behind. Under high vacuum pressure, the Analysis
combined phase (oil and water) solutions were treated in
the intake of a microfluidizer to produce emulsions with The nanoemulsion of edible oils, including castor, olive,
fine nanodroplets as mentioned in Fig. 2. and sunflower oils exhibits peaks in the FTIR spectrum. The
corresponding groups analyzed O–H stretch, C-H stretch,
C = O stretch, C = C stretch, C-O stretch, and C–Br stretch at
4.2 Nanoemulsion Size and Zeta Potential 3336–3359, 2855–2926, 1460–1461, 1742–1744, 1638–1639,
Determination and 586–626 cm − 1, represented all NEs (NE-25, 26, 28, 29)
(Fig. 2a–d). The O–H stretch of all NEs has a broad and strong
The nanoemulsions were emulsified with the micro- peak in the 3200–3555 cm − 1, indicating NE involvement
fluidizer and tested for size and zeta potential (DLS- with the oil’s functional groups. The shift in peak towards
Malvern Zetasizer Nano ZSP instrument) measurement. the lower wavenumber is due to the increased mass of the
Each NE employed for a dynamic light scattering (DLS) molecule, indicating that the vibration frequency decreases as
instrument for zeta potential analysis (mV) as NE-25, the mass of a molecule increases [41]. The C-H stretch is also
NE-26, NE-28, and NE-29 are 41.8mV, 49.1mV, 40.8mV, observed in the range of 3333–3267 cm − 1, with strong and
and 49.1mV and PDI values of 0.108, 0.251, 0.223, sharp peaks, and in the case of all NEs, this stretch is observed
and 0.173. Dynamic light scattering size distribution at an increased wavenumber of 2921 cm − 1. The FTIR spectra
histogram predicted that the mean diameter size of of the loaded nanoemulsion reveal all the significant active
nanoemulsions would be calculated as an average of ingredient peaks, indicating no incompatibility between the
the sizes [39]. The NE-25, NE-26, NE-28, and NE-29 active and inactive ingredients. Therefore, it can be concluded
showed distinct single peaks and particles of an average that all the characteristic peaks of studied NEs are related to
size of 229.7nm, 190.2nm, 201.8nm, and 145.8nm, functional groups of edible oils involved in nanoformulations
shown in Fig. 2a–d. The produced nanoemulsions [42, 43] mentioned in Fig. 3a–d.
quality was determined by the unique peak discovered
by DLS analysis [40]. The formulations suggested that 4.4 MIC and MBC Determination
the successful preparation of the nanoemulsion at the
nanometres scale is shown by high-intensity dispersion The antibacterial activities of the developed nanoemul-
in smaller droplet size ranges. sions were assessed using MIC and MBC. The lowest

Fig. 2  An analysis of synthe-


sized nanoemulsions’ mean
droplet size distribution was
carried out using a 25, b 26, c
28, and d 29 nanoemulsions. To
create the four nanoemulsions,
three emulsifications at 20,000
lb/in2 were performed using a
microfluidizer LM-10. DLS was
used to analyze the average size
distributions of all the nanoe-
mulsions

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

Fig. 3  Fourier Transform


Infrared Spectra were obtained
for formulated nanoemulsions
containing castor oil (NE-25),
olive oil (NE-26), and sunflower
oil, with incorporated phyto-
compounds NE-28 (Epigallo-
catechin-gallate) and NE-29
(Taxifolin-TF)

concentration, at which no apparent growth of the test Fig. 4 a–c, strains of V. parahaemolyticus pathogens in the
pathogens was seen, was recorded as the MIC. The anti- bacteriostatic and bactericidal assay. The fraction dilution
bacterial activity of all (four) developed nanoemulsions (D) (1:10, 1:100, and 1:1000 D) of NEs also showed suc-
was evaluated using the micro-dilution method [44] against cessful inhibition against shrimp pathogenic strains and
V. parahaemolyticus commercial (MTCC-451) and a path- commercial strains tested in Fig. 5a–c. Individually, castor
ogenic diseased shrimp strain (VP-S7 and VP-S14). All oil, olive, and sunflower (20%) were tested, and no killing
synthesized (NE-25, 26 28, 29) NE efficacy was to exhibit effects were noted against all the pathogens. The presence
noteworthy antibacterial activity as determined by bacte- of castor, olive, and sunflower oils can capsulate active
riostatic and bactericidal investigations; however, NE-25 ingredients, improve antibacterial effects, and function as
showed effectual and significant catalytic activity at high nanoscale catalysis in nanodroplets. The potential use of
dilution ranges against V. parahaemolyticus (MTCC-451, NEs incorporate bioactive components, as a practical and
VP-S14) (MIC-2:4 to 2:4096 and MBC- 1:2 to 1:2048) and effective method either to increase their physical stability
VP-S7 (MIC-2:4 to 2:1024; MBC- 1:2 to 1:512), shown in or to decrease their possible harmful sensory effects [45].

Fig. 4  The bacteriostatic and bactericidal solutions were determined strains: a VP-MTCC-451; b VP-S7; c VP-S14); NE-25-castor oil
at various dilutions against pathogenic strains of V. parahaemolyti- nanoemulsion; NE-26 olive oil nanoemulsion; NE-28, 29-sunflower
cus. The different oil-based nanoemulsions efficacy tested against the oil nanoemulsion
commercial strain of V. parahaemolyticus as well as diseased shrimp

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

Fig. 5  Three fractional dilutions were utilized to check (D-1:10, VP-S14), our NEs showed successful inhibitions at various dilutions,
1:100, 1:1000) minimum inhibitory and bactericidal concentrations comparable to antibiotics (gentamicin): NE-25-(castor oil nanoe-
against three V. parahaemolyticus. Based on the a V. parahaemo- mulsion); NE-26 (olive oil nanoemulsion); NE-28, 29 (sunflower oil
lyticus (MTCC-451) and b, c diseased shrimp strains (VP-S7 and nanoemulsion)

4.5 Adherence Inhibition of Nanoemulsions at (49.55 and 53.48%), and with antibiotics at (53.50;
54.39%) in Fig. 6a. The various dilutions (10, 100, 1000D)
Four nanoemulsions were produced and tested by ant- were also utilized to know adherence inhibitions on all the
adherence assays, based on the results obtained from MBIC strains of V. parahaemolyticus grown here, which are shown
dose inhibitory potential on strains of V. parahaemolyticus in Fig. 6b–d. Amongst all strains, the NE (NE-25)-treated
from shrimp aquaculture. The adherence efficacy of nanoe- maximum elimination was found in MTCC-451, compared
mulsions on three strains of V. parahaemolyticus (MTC- to shrimp strains.
451) and two from shrimp (VP-S7; VP-S14), developing The clinically isolated strains may possess a higher
cells to the glass surfaces of slanted glass coverslips, was degree of resistance due to virulence factors (tlh, toxR) pres-
investigated. The outcomes of our results stated here the ently reported and the influential activity of these genes.
adherence potential of NE-25 (68.76%) and gentamicin as The VP-14 genome was discovered to carry all 39 virulence
(69.70%) inhibition reported against the V. parahaemolyti- factors listed in the Virulence Factors Database as T3SS1
cus (MTCC-451) when treated with 62.5ug/ml. The shrimp (responsible for the cytotoxicity of host cells) [9, 28]. The
obtained strains’ (VP-S7; VP-S14) adherence was reported nanoemulsion is positively charge-bearing particles due to

Fig. 6  Adherence potential


of four nanoemulsions tested
against the V. parahaemo-
lyticus strains. The oil-based
nanoemulsions in effectiveness
tested against the commercial
strain (MTCC-451) as well as
diseased shrimp strains: a V.
parahaemolyticus (VP-S7 &
VP-S14). Additionally, nanoe-
mulsions dilutions were used
1:10, 1:100, and 1:1000 D to
test adherence inhibition poten-
tial against b V. parahaemolyti-
cus (MTCC-451), and shrimp
strains c, d VP-S7 and VP-S14,
respectively. NE (nanoemul-
sion); PC (positive control); GC
(growth control); D (Dilution);
VP (V. parahaemolyticus); S
(shrimp strains 7 and 14)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

cetylpyridinium chloride (CPC). The addition of CPC (cati- following treatments. Quantitative analysis of the biofilm
onic halogen) into the nanoemulsion mixture can enhance assay showed biofilm inhibition in the presence of castor
antibacterial potential [46]. Additionally, nanoemulsions and olive oil nanoemulsion (NE-25, 26), phytocompound
having nano-droplet size can penetrate the biofilm matrix sunflower oil nanoemulsions as (NE-28-Epigallocatchin gal-
and interact with bacteria embedded within it. This can dis- late 0.5%, and NE-29-Taxifolin 0.4%), and nanoemulsions
rupt the matrix and weaken the biofilm’s structural integrity, inhibitions observed dose-dependent manner. Minimum
decreasing adherence to the glass surface [47]. Extracellular biofilm inhibition concentration (MBIC) values for NE-25
curli fibrils, heteropolymeric filaments of major, and minor and 26 are 62.5ug/ml, and NE-28 and 29 were found to be
subunits produced by many bacteria, including V. para- 125 μg/ml, respectively. Therefore, results showed higher
haemolyticus, have been linked to adhesion, biofilm forma- inhibition in NE-25 (55.96%) and less inhibition in NE-29
tion, and virulence. However, the additional components of (44.36%), including gentamicin (58.28%) against V. para-
planktonic form a complex extracellular matrix that supports haemolyticus (MTCC-451). The shrimp obtained pathogenic
and protects the bacterial community [48]. It is important strains with higher inhibition in NE-25 (51.74%), less inhi-
to note that the efficacy of nanoemulsions in reducing bac- bition in NE-29 (44.04%), and gentamicin (54.91%). The
terial biofilm adherence on glass surfaces can be affected VP-S14 maximum biofilm inhibition (52.73%), lowest in
by various factors, including the specific biofilm-forming NE-28 (41.64%), and with gentamicin (54.28%), respec-
bacteria, the composition and concentration of the nanoe- tively shown in Fig. 7a. Furthermore, the biofilm was also
mulsion, contact time, and application method. Resistance quantified at various dilutions (D) against V. parahaemo-
to currently available antibiotics and their spread into the lyticus (MTCC-451) as well as shrimp obtained strains
food chain and environment necessitates the development of (CIBA-VP-S7; VP-S14) (1:10; 1:50; 1:1000D) showed in
safe and effective alternative methods against this bacterium. shown in Fig. 7b–d; and in each dilution, inhibition was
also quantified. The study found that increasing the dilution
4.6 Nanoemulsions Effect on Biofilm of nanoemulsion resulted in a decrease in its effectiveness
Quantifications in inhibiting the biofilm formation against V. parahaemo-
lyticus, thus suggesting that NE-25 has a stronger ability
Antibiotic-resistant gene associations in bacteria promote to inhibit biofilm formation at higher concentrations. The
the formation of a matrix through the production of vari- biofilm of V. parahaemolyticus was treated with NEs and
ous surface components and virulence factors [49, 50]. The stained with crystal violet as quantitative data was esti-
effect of edible oils NE, including antibiotics, on the biofilm mated. These results demonstrated that a novel castor, olive
of V. parahaemolyticus, was quantitatively evaluated by the oil, and phytocompounds loaded poorly with water-soluble

Fig. 7  Percentage of biofilm


formation determined against
three strains of V. parahaemo-
lyticus. The different oil-based
nanoemulsions efficacy tested
against commercial as well as
the shrimp isolated pathogenic
strains: a V. parahaemolyticus
(MTCC-451) and (VP-S7 and
VP-S14). Further, dilutions (D)
as 1:50, 1:100, 1:1000, b rep-
resented V. parahaemolyticus
(MTCC-451), and c, d diseased
shrimp strains. NE-25-(castor
oil nanoemulsion); NE-26 (olive
oil nanoemulsion); NE-28 &
29 (sunflower oil NEs with
phytocompounds). S1-(VP-
MTCC-451), S-shrimp strains
(S7 & 14); D1 (1:10 dilution),
D2 (1:100 dilution), D3 (1:1000
dilution)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

nanoemulsion could improve this agent’s aqueous solubility antibiotics gentamicin-54.06%. The diseased shrimp strains
and antimicrobial activity against V. parahaemolyticus in VP-S7 and VP-S14 at (43.19 and 44.92%), whereas anti-
vitro. So nanoemulsion inhibits effect on biofilm formation biotics at (48.28 and 52.09%) are shown in Fig. 8a respec-
and enhances damage to the structure of biofilm. Similarly, tively. The dilutions (1:10, 1:100, and 1:1000 D) results also
earlier studies reported nanoemulsion biofilm effectuality assessed the reduction efficiency in dead cells against V. par-
against the Streptococcus mutans [29]. The MIC test is used ahaemolyticus strains tested here, and successful inhibitions
to determine planktonic bacteria’s antimicrobial sensitivity. were noted, as shown in Fig. 8b–d. The separate mechanisms
It is the lowest antibiotic concentration required to halt bac- of action of nanoemulsions (EGCG; TF) and CPC may be
terial growth under certain conditions. Unlike planktonic operating to reduce biofilm formation on the glass coverslip.
bacteria cultivated in liquid media, bacterial biofilms can The positively charged emulsion may remain attached to the
withstand antibiotic dosages up to 1000 times the MIC. The biofilm for a longer time than individual phytocompounds
presence of extracellular polysaccharides, the spatial organi- do and, therefore, can prevent the formation of furthermore
zation of the community, and the physiological changes biofilm. Additionally, after treatment, the fractional cell was
caused by biofilm formation all help bacteria survive. As a stained with SYTO 9 and quantified by the multimode reader
result, the MIC is not proportional to the minimal biofilm [27]. We used SYTO 9 to enter into cells continuously while
elimination concentration (MBEC), as the reason biofilm entering quantified cells by fluorescence lost within 0.1ns
changes is specific to environmental conditions [51]. [52]. The SYTO9 is a hydrophobic cell-permeant nucleic
acid dye that exhibits a significant fluorescence enhancement
4.7 Quantifications of Dead Cell Populations after penetrating the cell’s intact membrane and binding to
the nucleic acids. The SYTO9 can be permeable as (i) it can
The biofilm that had grown was treated with NE-25 to NE-29 enter both the living and dead cells. (ii) It penetrates dead
followed treatment; the cells were stained with LIVE/DEAD cells more than live cells due to their impaired membrane.
BacLight TM Bacterial Viability Kit (Molecular Probes Inc.) (iii) It has greater penetration frequently to dead cells than
After the treatment, the well was rinsed and washed once live cells due to their larger membrane pores [37].
with PBS (1 ×) solutions. We utilized the minimum biofilm
inhibition concentration (MBIC) values for NE-25 and 26, 4.8 Oxygen Consumption Rate Determination
62.5ug/ml, and NE-28 and 29 were found to be 125 μg/ml,
to quantify the dead cell numbers, respectively. The results The permeability and efflux of drugs are significantly influ-
determined that NE-25 (48.68%) showed the dead popula- enced by the bacterial membrane, which is highly depend-
tions against V. parahaemolyticus (MTCC-451) and standard ent on the active membrane potential. Previous research

Fig. 8  Assessment of percent-


age of dead cells of pathogens
treated with ozonated oil
nanoemulsions and gentamicin
(PC) by employing already
determined concentrations as
on pre-formed biofilms. The
nanoemulsion potential tested
against commercially purchased
V. parahaemolyticus (MTCC-
451) as well as diseased shrimp
strains (VP-S7 and VP-S14)
a) combine graph of all three
strains and dilutions employed
as 1:10, 1:100, 1:1000D. The b
V. parahaemolyticus (MTCC-
451); c, d shrimp isolated
VP-S7 and VP-S14 strains—D,
dilution; PC, positive control-
(gentamicin); NE, nanoemul-
sion

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

Fig. 9  Nanoemulsion OCR


effect against V. parahaemolyti-
cus a, b (MTCC-451); c, d (VP-
S7); and e, f (VP-S14) culture
were grown for 6-8h and treated
with the desired concentrations
of NEs and PC (gentamicin),
by maintaining 10.7cells/ml
with OD at 600nm. The graph
shows OCR affected; signifying
cell death compared to growth
control. The average values
from each experiment were
plotted after being repeated
three times. The values were
presented as means ± SD, with
n = 3, compared to untreated
cells, *P < 0.05; ***P < 0.001.
MTCC, microbial type of
culture collection; Ab, antibiot-
ics; NE, nanoemulsion; OCR,
oxygen consumption rate; VP,
V. parahaemolyticus; S, shrimp
strains-7 And 14; growth
control (Vp-MTCC, VP-S7, and
VP-S14)

by Kosuru et al., [53] tested the oxygen consumption rate of VP-S14, NE-25, and NE-26 demonstrated reductions of
against P. aeruginosa using gallic acid and tannic acid. In 0.12 and 0.22, and antibiotics resulted in a reduction of 0.62.
this study, nanoemulsions (NE-25, 26), including phyto- These values indicate that NE-25 and NE-26 were 0.5- and
compounds (NE28-EGCG and NE29-TF) used by Ahmed 0.40-fold lesser than the antibiotics, as depicted in Fig. 9e.
et al., [54] at optimized concentrations of 125μg/ml, were Moreover, the impact of OCR on cell death was noticeable,
employed as a therapeutic treatment against V. parahaemo- as there was an increase in the measurement of nanomoles
lyticus strains. The current antibiotic at a concentration of of oxygen/milliliter/minute, as recorded in Fig. 9f. Hence,
175μg/ml impacted the OCR in a commercial strain of VP all tested NEs against V. parahaemolyticus from diseased
(MTCC-451). It resulted in a reduction of 0.58-fold (Fig. 9a), shrimp exhibited a notable reduction in cell count at a con-
indicating a lower rate than the growth control (1.0). The centration of 125μg/ml. Additionally, there was an observed
increase in the nanomoles of oxygen/ml/min was recorded increase in the measurement of nanomoles of oxygen/mil-
(Fig. 9b). Meanwhile, NE-25, NE-28, and NE-29 exhibited liliter/minute compared to the concentration of antibiotics
reductions of 0.38-, 0.16-, and 0.31-fold. Interestingly, these at 175μg/ml and the growth control (0-μg/ml). The NEs
values are 0.2-, 0.42-, and 0.27-fold lower than antibiotics with high-pressure nanodroplet particles, when hitting with
(0.58), as shown in Fig. 9a. Consequently, the reduction in pathogen surfaces, generate a bombing effect and release
the OCR observed in the cells was significantly increased a substantial amount of energy that disrupts and kills the
when measured in terms of nanomoles of oxygen/ml/min pathogens [55].
(Fig. 9b). The results were observed with NE-25, and NE-29
were tested against VP-S7. These NEs exhibited reductions
of 0.23 and 0.20, which were 0.26- and 0.29-fold lower than 5 Conclusion
the reduction caused by antibiotics, which measured at 0.49
(Fig. 9c). Furthermore, the impact of OCR on cell death A recent study produced four nanoemulsions using a high-
was evident as there was an increase in the measurement of pressure homogenization technique. The nanoemulsion’s
nanomoles of oxygen/milliliter/minute (Fig. 9d). In the case zeta potential, droplet sizes, and PDI values indicate the

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

uniformity and stability of the formulations. To assess Data Availability The datasets generated during and/or analyzed dur-
the impact of ozone-enriched nanoemulsions (OZNEs) on ing the current study are available from the corresponding author upon
reasonable request.
V. parahaemolyticus strains from diseased shrimp were
employed, each exhibiting varying degrees of sensitivity to Code Availability Not applicable.
gentamicin. Remarkably, the NE-25 showed strong adher-
ence inhibition, with results of VP-S7 (49.55%), VP-S14 Declarations
(53.47%), and MTCC-451 (68.76%). In contrast, the bio-
Competing interests The authors declare no competing interests.
film inhibition assays revealed that NE-25 and NE-26 at a
concentration of 62.5 μg/ml, as well as NE-28 and NE-29 Research Involving Humans and Animals Statement None.
at a concentration of 125 μg/ml, exhibited independent
efficacy against all strains of V. parahaemolyticus. We Informed Consent Not applicable.
utilized the minimum biofilm inhibition concentration Ethics Approval Not applicable.
(MBIC) values for NE-25 and 26 (62.5μg/ml) and NE-28
and 29 (125 μg/ml) to quantify the dead cell. The results Consent to Participate Not applicable.
determined that NE-25 (48.68%) showed the dead cells of
Consent for Publication All authors read and agreed to the manuscript
V. parahaemolyticus (MTCC-451) and current antibiotics for publication.
gentamicin-54.06%. The diseased shrimp strains VP-S7 and
VP-S14 were at 43.19 and 44.92%), whereas current antibi- Conflict of interest None.
otics were at 48.28 and 52.09%. For the OCR assays, opti-
mized concentrations of NEs, including phytocompounds
NEs (125 μg/ml) and antibiotics gentamicin-175 μg/ml,
were utilized for treatments. The better fold reduction for References
Vp-MTCC strain, NE-25 and NE-28 (0.42; 0.2), indicat-
1. Kooloth, V. R., Stentiford, G. D., & Bass, D. (2021). The rise of
ing lower than the growth control (1.0), is highly signifi- the syndrome–sub-optimal growth disorders in farmed shrimp.
cant when compared with the antibiotics (0.58). Similarly, Reviews in Aquaculture, 13, 41888–41906. https://​doi.​org/​10.​
against VP-S7, the reduction was noted as NE-29 was 0.29- 1111/​raq.​12550
2. Little, D. C., Newton, R. W., & Beveridge, M. C. M. (2016).
fold lower than the reduction caused by antibiotics, which
Aquaculture: a rapidly growing and significant source of sustain-
measured at 0.49, whereas VP-S14, NE-25, and NE-26 dem- able food? Status, transitions and potential. Proceedings of the
onstrated reductions as 0.12 and 0.22, compared to antibi- Nutrition Society, 75(3), 274–286. https://​doi.​org/​10.​1017/​S0029​
otics 0.62 and growth control (1.0). Hence, all tested NEs 66511​60006​65
3. Albalat, A., Zacarias, S., Coates, C. J., Neil, D. M., & Planellas, S.
against V. parahaemolyticus from diseased shrimp exhib-
R. (2022). Welfare in farmed decapod crustaceans, with particular
ited a notable reduction in cell count at a concentration of reference to Penaeus vannamei. Frontiers in Marine Science, 09,
125μg/ml. Additionally, an increase in the measurement of 886024. https://​doi.​org/​10.​3389/​fmars.​2022.​886024
nanomoles of oxygen/milliliter/minute was observed, com- 4. Ghosh, A. K., Panda, S. K., & Luyten, W. (2021). Anti-vibrio and
immune-enhancing activity of medicinal plants in shrimp: A com-
pared to the concentration of antibiotics at 175μg/ml and the
prehensive review. Fish & Shellfish Immunology, 117, 192–210.
growth control (0-μg/ml). Hence, the results from all tests https://​doi.​org/​10.​1016/j.​fsi.​2021.​08.​006
affirm that NEs demonstrate comparable and, in some cases, 5. Gunalan, B., Soundarapandian, P., Anand, T., Kotiya, A. S., &
even superior antibacterial effectiveness compared to cur- Simon, N. T. (2014). Disease occurrence in Litopenaeus vannamei
shrimp culture systems in different geographical regions of India.
rent antibiotics. Therefore, NEs are an extremely promising
International Journal of Aquatic Biology, 10, 4. https://​doi.​org/​
substitute in combatting V. parahaemolyticus strains, which 10.​5376/​ija.​2014.​04.​0004
cause diseases in shrimp aquaculture. 6. Tan, L. T. H., Chan, K. G., Lee, L. H., & Goh, B. H. (2016). Strep-
tomyces bacteria as potential probiotics in aquaculture. Frontiers
Acknowledgements We acknowledge B. S. Abdur Rahman Crescent in Microbiology, 7, 79. https://d​ oi.o​ rg/1​ 0.3​ 389/f​ micb.2​ 016.0​ 0079
Institute of Science and Technology for providing the facility to per- 7. Letchumanan, V., Chan, K. G., & Lee, L. H. (2014). Vibrio par-
form this research; and Dr. Sneha Unnikrishnan for data curation, for- ahaemolyticus: A review on the pathogenesis, prevalence, and
mal analysis, and investigation. We thank Dr. Soumen Bera and Mr advance molecular identification techniques. Frontiers in Micro-
Md. Aashique for their help in doing OCR studies using the Clarke biology, 5, 705. https://​doi.​org/​10.​3389/​fmicb.​2014.​00705
electrode-polygraph instrument. 8. Wang, L., Chen, Y., Huang, H., Huang, Z., Chen, H., & Shao, Z.
(2015). Isolation and identification of Vibrio campbellii as a bac-
Author Contribution Jahangir Ahmed: All experimental work, data terial pathogen for luminous vibriosis of Litopenaeus vannamei.
curation, formal analysis, investigation, methodology, and writing of Aquaculture Research, 46(2), 395–404. https://​doi.​org/​10.​1111/​
the original draft. Dr. Karthikeyan Ramalingam: conceptualization, are.​12191
supervision; validation; writing, review, and editing. 9. Raja, R. A., Sridhar, R., Balachandran, C., Palanisammi, A.,
Ramesh, S., & Nagarajan, K. (2017). Prevalence of Vibrio spp.
Funding This work was supported by the Indian Council of Medical with special reference to Vibrio parahaemolyticus in farmed
Research [Project ID: 2020–4964]. penaeid shrimp Penaeus vannamei (Boone, 1931) from selected

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

districts of Tamil Nadu, India. Indian Journal of Fisheries, 23. Yuan, Y., Gao, Y., Zhao, J., & Mao, L. (2008). Characterization
64(3):122–128. https://​doi.​org/​10.​21077/​ijf.​2017.​64.3.​69011-​18 and stability evaluation of β-carotene nanoemulsions prepared by
10. Narayanan, S. V., Joseph, T. C., Peeralil, S., Koombankallil, high pressure homogenization under various emulsifying condi-
R., Vaiyapuri, M., Mothadaka, M. P., & Lalitha, K. V. (2020). tions. Food Research International, 41(1), 61–68. https://​doi.​org/​
Tropical shrimp aquaculture farms harbour pathogenic Vibrio 10.​1016/j.​foodr​es.​2007.​09.​006
parahaemolyticus with high genetic diversity and Carbapenam 24. Wang, X., Jiang, Y., Wang, Y. W., Huang, M. T., Ho, C. T., &
resistance. Marine Pollution Bulletin, 160, 111551. https://​doi.​ Huang, Q. (2008). Enhancing anti-inflammation activity of cur-
org/​10.​1016/j.​marpo​lbul.​2020.​111551 cumin through O/W nanoemulsions. Food Chemistry, 108(2),
11. Kumar, B. K., Deekshit, V. K., Raj, J. R. M., Rai, P., Shi- 419–424. https://​doi.​org/​10.​1016/j.​foodc​hem.​2007.​10.​086
vanagowda, B. M., Karunasagar, I., & Karunasagar, I. (2014). 25. Lawrence, M. J., & Rees, G. D. (2012). Microemulsion-based
Diversity of Vibrio parahaemolyticus associated with disease media as novel drug delivery systems. Advanced Drug Delivery
outbreak among cultured Litopenaeus vannamei (Pacific white Reviews, 64, 175–193. https://d​ oi.o​ rg/1​ 0.1​ 016/j.a​ ddr.2​ 012.0​ 9.0​ 18
shrimp) in India. Aquaculture, 433, 247–251. https://​doi.​org/​10.​ 26. Teixeira, P. C., Leite, G. M., Domingues, R. J., Silva, J., Gibbs,
1016/j.​aquac​ulture.​2014.​06.​016 P. A., & Ferreira, J. P. (2007). Antimicrobial effects of a micro-
12. Kumar, R., Tung, T. C., Ng, T. H., Chang, C. C., Chen, Y. L., emulsion and a nanoemulsion on enteric and other pathogens and
Chen, Y. M., Lin, S. S., & Wang, H. C. (2021). Metabolic altera- biofilms. International Journal of Microbiology, 118(1), 15–19.
tions in shrimp stomach during acute hepatopancreatic necrosis https://​doi.​org/​10.​1016/j.​ijfoo​dmicro.​2007.​05.​008
disease and effects of taurocholate on Vibrio parahaemolyticus. 27. Ramalingam, K., Amaechi, B. T., Ralph, R. H., & Lee, V. A.
Frontiers in Microbiology, 12, 631468. https://​doi.​org/​10.​3389/​ (2012). Antimicrobial activity of nanoemulsion on cariogenic
fmicb.​2021.​631468 planktonic and biofilm organisms. Archives of Oral Biology,
13. Ahmed, J., Khan, M. H., Unnikrishnan, S., & Ramalingam, K. 57(1), 15–22. https://​doi.​org/​10.​1016/j.​archo​ralbio.​2011.​07.​001
(2022). Acute hepatopancreases necrosis diseases (AHPND) as 28. Jangam, A. K., Bhuvaneswari, T., Krishnan, A. N., Katneni, V.
challenging threat in shrimp. Biointerface Research in Applied K., Avunje, S., Grover, M., Kumar, S., Alavandi, S. V., & Vijayan,
Chemistry, 12(1), 978–991. https://​doi.​org/​10.​33263/​briac​121.​ K. K. (2018). Draft genome sequence of Vibrio parahaemolyti-
978991 cus strain VP14, isolated from a Penaeus vannamei culture farm.
14. Dash, P., Avunje, S., Tandel, R. S., KP, S., & Panigrahi, A. Genome Announcements, 6(11), 00149–00218. https://d​ oi.o​ rg/1​ 0.​
(2017). Biocontrol of luminous vibriosis in shrimp aquaculture: 1128/​genom​eA.​00149-​18
A review of current approaches and future perspectives. Reviews 29. Li, W., Chen, H., He, Z., Han, C., Liu, S., & Li, Y. (2015). Influ-
in Fisheries Science & Aquaculture, 25(3), 245–255. https://​doi.​ ence of surfactant and oil composition on the stability and anti-
org/​10.​1080/​23308​249.​2016.​12779​73 bacterial activity of eugenol nanoemulsions. LWT-Food Science
15. Okocha, R. C., Olatoye, I. O., & Adedeji, O. B. (2018). Food and Technology, 62(1), 39–47. https://d​ oi.o​ rg/1​ 0.1​ 016/j.l​ wt.2​ 015.​
safety impacts of antimicrobial use and their residues in aqua- 01.​012
culture. Public Health Reviews, 39(1), 1–22. https://​doi.​org/​10.​ 30. Kaur, H., Pancham, P., Kaur, R., Agarwal, S., & Singh, M. (2020).
1186/​s40985-​018-​0099-2 Synthesis and characterization of Citrus limonum essential oil
16. Alloul, A., Wille, M., Lucenti, P., Bossier, P., Van Stappen, based nanoemulsion and its enhanced antioxidant activity with
G., & Vlaeminck, S. E. (2021). Purple bacteria as added-value stability for transdermal application. Journal of Biomaterials and
protein ingredient in shrimp feed: Penaeus vannamei growth Nanobiotechnology, 11(4), 215–236. https://d​ oi.o​ rg/1​ 0.4​ 236/j​ bnb.​
performance, and tolerance against Vibrio and ammonia stress. 2020.​114014
Aquaculture, 530, 735788. https://​d oi.​o rg/​1 0.​1 016/j.​a quac​ 31. Gholipourkanani, H., Buller, N., & Lymbery, A. (2019). In vitro
ulture.​2020.​735788 antibacterial activity of four nano-encapsulated herbal essential
17. Kumar, M., Bishnoi, R. S., Shukla, A. K., & Jain, C. P. (2019). oils against three bacterial fish pathogens. Aqua Research, 50(3),
Techniques for formulation of nanoemulsion drug delivery sys- 871–875. https://​doi.​org/​10.​1111/​are.​13959
tem: A review. Preventive Nutrition and Food Science, 24(3), 32. Ramalingam, K., & Lee, V. (2019). Biotic and abiotic substrates
225. https://​doi.​org/​10.​3746/​pnf.​2019.​24.3.​225 for enhancing Acinetobacter baumannii biofilm formation: New
18. Hosseini, S. F., Ramezanzade, L., & McClements, D. J. (2021). approach using extracellular matrix and slanted coverslip tech-
Recent advances in nanoencapsulation of hydrophobic marine nique. The Journal of General and Applied Microbiology, 65(2),
bioactives: Bioavailability, safety, and sensory attributes of 64–71. https://​doi.​org/​10.​2323/​jgam.​2018.​05.​004
nano-fortified functional foods. Trends in Food Science & Tech- 33. Wang, X. H., & Leung, K. Y. (2000). Biochemical characteri-
nology, 109, 322–339. https://​doi.​org/​10.​1016/j.​tifs.​2021.​01.​045 zation of different types of adherence of Vibrio species to fish
19. Karthik, P., Ezhilarasi, P. N., & Anandharamakrishna, C. (2017). epithelial cells. Microbiology, 146(4), 989–998. https://​doi.​org/​
Challenges associated in stability of food grade nanoemulsions. 10.​1099/​00221​287-​146-4-​989
Critical Reviews in Food Science and Nutrition, 57(7), 1435– 34. Khan, M. H., & Ramalingam, K. (2019). Synthesis of antimicro-
1450. https://​doi.​org/​10.​1080/​10408​398.​2015.​10067​67 bial nanoemulsions and its effectuality for the treatment of multi-
20. Madene, A., Jacquot, M., Scher, J., & Desobry, S. (2006). drug resistant ESKAPE pathogens. Biocatalysis and Agricultural
Flavour encapsulation and controlled release–A review. Inter- Biotechnology, 18, 101025. https://​doi.​org/​10.​1016/j.​bcab.​2019.​
national Journal of Food Science & Technology, 41(1), 1–21. 101025
https://​doi.​org/​10.​1111/j.​1365-​2621.​2005.​00980.x 35. Xie, T., Liao, Z., Lei, H., Fang, X., Wang, J., & Zhong, Q. (2017).
21. Zimet, P., Rosenberg, D., & Livney, Y. D. (2011). Re-assembled Antibacterial activity of food-grade chitosan against Vibrio para-
casein micelles and casein nanoparticles as nano-vehicles for haemolyticus biofilms. Microbiology in Clinical Pathology, 110,
ω-3 polyunsaturated fatty acids. Food Hydrocolloids, 25(5), 291–297. https://​doi.​org/​10.​1016/j.​micpa​th.​2017.​07.​011
1270–1276. https://​doi.​org/​10.​1016/j.​foodh​yd.​2010.​11.​025 36. Stiefel, P., Schmidt-Emrich, S., Maniura-Weber, K., & Ren, Q.
22. Belhaj, N., Arab-Tehrany, E., & Linder, M. (2010). Oxidative (2015). Critical aspects of using bacterial cell viability assays with
kinetics of salmon oil in bulk and in nanoemulsion stabilized by the fluorophores SYTO9 and propidium iodide. BMC Microbiol-
marine lecithin. Process Biochemistry, 45(2), 187–195. https://​ ogy, 15, 1–9. https://​doi.​org/​10.​1186/​s12866-​015-​0376-x
doi.​org/​10.​1016/j.​procb​io.​2009.​09.​005

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


BioNanoScience

37. Cal-Sabater, P., Caro, I., Castro, M. J., Cao, M. J., Mateo, J., & 47. Karan, S., Choudhury, D., & Dixit, A. (2021). Immunogenic char-
Quinto, E. J. (2019). Flow cytometry to assess the counts and acterization and protective efficacy of recombinant CsgA, major
physiological state of Cronobacter sakazakii cells after heat expo- subunit of curli fibers, against Vibrio parahaemolyticus. Applied
sure. Foods, 8(12), 688. https://​doi.​org/​10.​3390/​foods​81206​88 Microbiology and Biotechnology, 105, 599–616. https://​doi.​org/​
38. Aashique, M., Roy, A., Kosuru, R. Y., & Bera, S. (2021). Mem- 10.​1007/​s00253-​020-​11038-4
brane depolarization sensitizes Pseudomonas aeruginosa against 48. Yeh, Y. C., Huang, T. H., Yang, S. C., Chen, C. C., & Fang, J. Y.
tannic acid. Current Microbiology, 78, 713–717. https://​doi.​org/​ (2020). Nano-based drug delivery or targeting to eradicate bacte-
10.​1007/​s00284-​020-​02330-7 ria for infection mitigation: A review of recent advances. Frontiers
39. Yuan, Y., Gao, Y., Zhao, J., & Mao, L. (2008). Characterization in chemistry, 8, 286. https://​doi.​org/​10.​3389/​fchem.​2020.​00286
and stability evaluation of β-carotene nanoemulsions prepared by 49. Roy, P. K., Park, S. H., Song, M. G., & Park, S. Y. (2022). Anti-
high pressure homogenization under various emulsifying condi- microbial Efficacy of Quercetin against Vibrio parahaemolyticus
tions. Food Research International, 41(1), 61–68. https://​doi.​org/​ biofilm on food surfaces and downregulation of virulence genes.
10.​1016/j.​foodr​es.​2007.​09.​006 Polymers, 14(18), 3847. https://​doi.​org/​10.​3390/​polym​14183​847
40. Rahnama, M., Anvar, S. A., Ahari, H., & Kazempoor, R. (2021). 50. Liu, H., Zhu, W., Cao, Y., Gao, J., Jin, T., Qin, N., & Xia, X.
Antibacterial effects of extracted corn zein with garlic extract- (2022). Punicalagin inhibits biofilm formation and virulence
based nanoemulsion on the shelf life of Vannamei prawn (Lito- gene expression of Vibrio parahaemolyticus. Food Control, 139,
penaeus vannamei) at refrigerated temperature. Journal of Food 109045. https://​doi.​org/​10.​1016/j.​foodc​ont.​2022.​109045
Science, 86(11), 4969–4990. https://​doi.​org/​10.​1111/​1750-​3841.​ 51. Sahli, C., Moya, S. E., Lomas, J. S., Gravier-Pelletier, C., Brian-
15923 det, R., & Hémadi, M. (2022). Recent advances in nanotechnol-
41. Horison, R., Sulaiman, F. O., Alfredo, D., & Wardana, A. A. ogy for eradicating bacterial biofilm. Theranostics, 12(5), 2383.
(2019). Physical characteristics of nanoemulsion from chitosan/ https://​doi.​org/​10.​7150/​thno.​67296
nutmeg seed oil and evaluation of its coating against microbial 52. Deng, Y., Wang, L., Chen, Y., & Long, Y. (2020). Optimization
growth on strawberry. Food Research, 3(6), 821–827. https://​doi.​ of staining with SYTO 9/propidium iodide: Interplay, kinetics
org/​10.​26656/​fr.​2017.​3(6).​159 and impact on Brevibacillus brevis. Biotechniques, 69(2), 88–98.
42. Rohman, A., & Man, Y. C. (2010). Fourier transform infrared https://​doi.​org/​10.​2144/​btn-​2020-​0036
(FTIR) spectroscopy for analysis of extra virgin olive oil adulter- 53. Kosuru, R. Y., Aashique, M., Fathima, A., Roy, A., & Bera,
ated with palm oil. Food Research International, 43(3), 886–892. S. (2018). Revealing the dual role of gallic acid in modulat-
https://​doi.​org/​10.​1016/j.​foodr​es.​2009.​12.​006 ing ampicillin sensitivity of Pseudomonas aeruginosa biofilms.
43. Osanloo, M., Firooziyan, S., Abdollahi, A., Hatami, S., Nematol- Future Microbiology, 13(3), 297–312. https://​doi.​org/​10.​2217/​
lahi, A., Elahi, N., & Zarenezhad, E. (2022). Nanoemulsion and fmb-​2017-​0132
nanogel containing Artemisia dracunculus essential oil; larvicidal 54. Ahmed, J., Navabshan, I., Unnikrishnan, S., Radhakrishnan, L.,
effect and antibacterial activity. BMC Research Notes, 15(1), 276. Vasagam, K. K., & Ramalingam, K. (2023). In silico and In vitro
https://​doi.​org/​10.​1186/​s13104-​022-​06135-8 investigation of phytochemicals against shrimp AHPND syn-
44. de Meneses, A. C., Sayer, C., Puton, B. M., Cansian, R. L., drome causing PirA/B toxins of vibrio parahaemolyticus. Applied
Araújo, P. H., & de Oliveira, D. (2019). Production of clove oil Biochemistry and Biotechnology, 195(12), 7176–7196. https://d​ oi.​
nanoemulsion with rapid and enhanced antimicrobial activity org/​10.​1007/​s12010-​023-​04458-1
against gram-positive and gram-negative bacteria. Journal of 55. Tavares, T. M. B., Almeida, H. M. D. E. S., Lage, M. V. M., de
Water Process Engineering, 42(6), 13209. https://​doi.​org/​10.​ Carvalho Feitosa, R., & da Silva Júnior, A. A. (2023). Nanoemul-
1111/​jfpe.​13209 sions: A promising strategy in the fight against bacterial infec-
45. Pérez-Córdoba, L. J., Norton, I. T., Batchelor, H. K., Gkatzionis, tions. In Medical sciences forum (Vol. 24, No. 1, p. 18). MDPI.
K., Spyropoulos, F., & Sobral, P. J. (2018). Physico-chemical, https://​doi.​org/​10.​3390/​ECA20​23-​16402
antimicrobial and antioxidant properties of gelatin-chitosan
based films loaded with nanoemulsions encapsulating active Publisher's Note Springer Nature remains neutral with regard to
compounds. Food Hydrocolloids, 79, 544–559. https://​doi.​org/​ jurisdictional claims in published maps and institutional affiliations.
10.​1016/j.​foodh​yd.​2017.​12.​012
46. Hwang, Y. Y., Ramalingam, K., Bienek, D. R., Lee, V., You, T., Springer Nature or its licensor (e.g. a society or other partner) holds
& Alvarez, R. (2013). Antimicrobial activity of nanoemulsion in exclusive rights to this article under a publishing agreement with the
combination with cetylpyridinium chloride in multidrug-resistant author(s) or other rightsholder(s); author self-archiving of the accepted
Acinetobacter baumannii. Antimicrobial agents and chemother- manuscript version of this article is solely governed by the terms of
apy, 57(8), 3568–3575. https://​doi.​org/​10.​1128/​aac.​02109-​12 such publishing agreement and applicable law.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like