You are on page 1of 15

International Journal of Biological Macromolecules 243 (2023) 125228

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


journal homepage: www.elsevier.com/locate/ijbiomac

Eco-friendly and effective antimicrobial Melaleuca alternifolia essential oil


Pickering emulsions stabilized with cellulose nanofibrils against bacteria
and SARS-CoV-2
Greiciele da S. Ferreira a, Daniel J. da Silva a, Alana G. Souza a, Eliana D.C. Yudice b,
Ivana B. de Campos b, Rute Dal Col b, Andre Mourão c, Herculano S. Martinho c, Derval S. Rosa a, *
a
Center for Engineering, Modeling, and Applied Social Sciences (CECS), Federal University of ABC (UFABC), Av. dos Estados, 5001, CEP 09210-210 Santo André, SP,
Brazil
b
Adolfo Lutz Institute, Santo André Regional Center, Av. Ramiro Colleoni, 240, CEP 09040-160 Santo André, SP, Brazil
c
Center for Natural and Human Sciences (CCNH), Federal University of ABC (UFABC), Av. dos Estados, 5001, CEP 09210-210 Santo André, SP, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Melaleuca alternifolia essential oil (MaEO) is a green antimicrobial agent suitable for confection eco-friendly
Melaleuca alternifolia disinfectants to substitute conventional chemical disinfectants commonly formulated with toxic substances
Essential oil that cause dangerous environmental impacts. In this contribution, MaEO-in-water Pickering emulsions were
Nanocellulose
successfully stabilized with cellulose nanofibrils (CNFs) by a simple mixing procedure. MaEO and the emulsions
Oil-in-water Pickering emulsion
SARS-CoV-2
presented antimicrobial activities against Staphylococcus aureus (S. aureus) and Escherichia coli (E. coli). Moreover,
MaEO deactivated the SARS-CoV-2 virions immediately. FT-Raman and FTIR spectroscopies indicate that the
CNF stabilizes the MaEO droplets in water by the dipole-induced-dipole interactions and hydrogen bonds. The
factorial design of experiments (DoE) indicates that CNF content and mixing time have significant effects on
preventing the MaEO droplets' coalescence during 30-day shelf life. The bacteria inhibition zone assays show that
the most stable emulsions showed antimicrobial activity comparable to commercial disinfectant agents such as
hypochlorite. The MaEO/water stabilized-CNF emulsion is a promissory natural disinfectant with antibacterial
activity against these bacteria strains, including the capability to damage the spike proteins at the SARS-CoV-2
particle surface after 15 min of direct contact when the MaEO concentration is 30 % v/v.

1. Introduction to degradation and oxidation that limit their direct use in the manu­
facture of antiseptic, pharmaceutical, cosmetic, and food products. In
Currently, essential oils (EOs) have been considered an innovative this way, they are usually inserted into products in formulations con­
and technologically viable alternative to the manufacture of conven­ taining different components capable of stabilizing them, ensuring the
tional chemical disinfectants (sodium chloride, ethanol, quaternary prolongation of the activity of their active principles and the effective­
ammonium salts, among others) due to their biological activity against a ness of their antimicrobial properties.
wide range of pathogenic bacteria, viruses, and fungi. The antimicrobial Emulsification has been a technology widely used by the food,
capacity of EOs is directly associated with their chemical compositions cosmetic and pharmaceutical industries to stabilize or control the
containing terpenes, terpenoids, and other substances capable of release kinetics of an active ingredient or additive to protect the product.
directly destroying the biological structures of microorganisms (such as In resume, emulsions are liquid mixtures based on an aqueous phase and
proteins, lipids, and nucleic acids). Besides that, these substances can an oil phase, commonly stabilized by amphiphilic molecules or macro­
lead to the generation of reactive oxygen species (ROS) that can cause molecules that are well-established in the industry as surfactants and
irreversible damage to microorganisms, compromising the cellular ma­ emulsifiers. These auxiliary components guarantee outstanding emul­
chinery and virion activities [1–4]. Despite the ability to act as dis­ sion stability over time to increase the shelf life of emulsion-based
infecting agents, EOs have high volatility, low solubility, high tendency products, which is a crucial factor for industries.

* Corresponding author.
E-mail addresses: dervalrosa@yahoo.com.br, derval.rosa@ufabc.edu.br (D.S. Rosa).

https://doi.org/10.1016/j.ijbiomac.2023.125228
Received 23 January 2023; Received in revised form 23 April 2023; Accepted 3 June 2023
Available online 7 June 2023
0141-8130/© 2023 Published by Elsevier B.V.
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 1. Schematic illustration of the adopted methodology and characterizations of the MaEO/water Pickering emulsions stabilized with CNF.

2
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Cellulose is the most abundant natural polysaccharide in the Earth's protecting food against SARS-CoV-2 and bacteria contamination.
crust, presenting several advantages of being applied instead of syn­
thetic emulsion stabilizers (surfactants and emulsifiers), such as biode­ 2. Materials ad methods
gradability, renewable biomass, and non-toxicity that are relevant
characteristics to avoid environmental problems and toxicological issues 2.1. Materials
to biological systems. Among the several cellulose structures currently
and commercially available (cellulose nanocrystals, nanocrystalline Cellulose nanofibrils (CNF) (4.0 ± 1.0 μm of length and 66.3 ± 20.5
cellulose, and cellulose microfibrils, just to name a few), cellulose nm of diameter, length/diameter ratio of 60, ζ potential of − 33.2 ± 3.2
nanofibrils (CNFs) are fibrous nanostructures with a nanoscale diameter mV) were donated by Suzano Papel e Celulose (São Paulo, Brazil) [12].
(lesser than 100 nm) that enable high surface area without high energy Melaleuca alternifolia essential oil (MaEO) was purchased from Ferquima
and financial production costs [5–7]. In addition to being expensive, the Indústria e Comércio Ltda. (São Paulo, SP, Brazil). Rapid Antigen Test
processes for nanocellulose isolation from the vegetable cell wall ma­ kits were supplied by Panbio™ COVID-19 Ag rapid test (Abbott, USA).
trixes generally involve the usage of toxic reagents and a high generation Hexane and resazurin were purchased from Synth (São Paulo, Brazil).
of residues that cause several environmental impacts [5,8,9]. The Lysoform® and hypochlorite were purchased in a local Market (Santo
combination of high surface area and small size of these eco-friendly André, SP, Brazil).
solid cellulose nanostructures is important to their stabilization capac­
ity in EO-water Pickering emulsions, which involve a formation of an 2.2. Bacterial strains
interfacial CNF layer via intermolecular and steric hindrance respon­
sible for hampering the collapsing resistance of EO droplets due to Bacterial isolates were obtained from the Center for Interdisciplinary
gravity and mechanical shear stress [10–12]. Moreover, the oil-in-water Procedures for the Collection of Microorganisms of the Instituto Adolfo
(O/W) emulsion stabilization protects the antimicrobial compounds in Lutz (Brazil), affiliated with World Federation Culture Collections
the EO droplets from volatilization and oxidation. For these peculiar (WFCC) # 282, Depository Collection - # 017/09-SECEX/CGEN/MMA:
characteristics, CNFs have been applied as interfacial stabilizing agents Staphylococcus aureus (ATCC #6538) and Escherichia coli (ATCC
in several Pickering emulsions [6,12–16]. In addition to high stability, #11229).
Pickering emulsions have the advantage of using biodegradable, edible
or renewable solid particles as stabilizing agents instead of being less 2.3. Methods
harmful to the environment and safer for biota and fauna than synthetic
stabilizers [17,18]. 2.3.1. Pickering emulsion preparation
Melaleuca alternifolia essential oil (MaEO) is extracted from the Tea CNF aqueous suspensions were homogenized in a high-intensity ul­
Tree (a native plant from Australia) by steam distillation of the leaves trasound (Sonics Vibra Cell, 400 W power and 24 kHz), for 10 min. The
and terminal branches. MaEO has both bactericidal and fungicidal ac­ MaEO was mixed with a previously dispersed CNF suspension, and this
tivity, including insecticidal capacity [19], standing out from other mixture was homogenized in an Ultra-Turrax Blender, IKA T25 model
essential oils for its strong bactericidal activity associated with shallow (IKA Werke, Staufen, Germany). The following parameters were varied:
minimum bactericidal concentration (MBC) values against several bac­ CNF content in aqueous suspension (0.5 and 1 % wt/v), oil/water (O/W)
terial strains [20]. MaEO is mainly composed of volatile cyclic mono­ ratio (20 and 30 % v/v), mixing rotation (12,000 and 15,000 rpm), and
terpenes, which are summarily constituted by terpinen-4-ol, eucalyptol, mixing time (5 and 7 min). Fig. 1 shows a schematic representation of
and α-terpineol (oxygenated monoterpenes) combined with α-terpinene, the adopted methodology. The emulsions were named XEO/YT/ZCNF/
γ-terpinene, and α-pinene (hydrocarbon monoterpenes) [21,22]. MaEO WRPM, where X corresponds to the MaEO content (% v/v), Y is the
has an antimicrobial activity due to these volatile compounds that lead mixing time (min), Z is the CNF concentration (% wt/v), and W is the
to the inhibition of the cellular respiration processes, microbial mem­ mixing rotation (rpm).
brane rupture, and leakage of potassium ions from cell membranes in
Gram-positive and Gram-negative bacteria [20]. 2.4. Characterization
Until now, few works have been published involving the stability of
Pickering emulsions of MaEO [23]. To the best of our knowledge, the 2.4.1. Gas chromatography coupled to mass spectrometry (GC–MS)
evaluation of the antiviral activity of Pickering emulsions of MaEO GC–MS analyses were carried out on a gas chromatograph coupled to
against Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) a mass spectrometer (3800 CG - 4000 EM system, Varian, USA) using a
is still a gap in the scientific literature. In a previous work [24], we DB-5MS chromatographic column (30 m × 0.25 mm × 0.25 μm) with a
evaluated the usage of Pickering emulsions of MaEO/water stabilized stationary phase composed of 5 % diphenyl and 95 % poly(dimethyl
with CNFs to confect wet disinfectant fabrics against different bacteria siloxane). The ion fragment analyzer was of the “Ion Trap” type, and the
and fungi. Therefore, it is necessary to evaluate the antiviral perfor­ detection range was 50–1000 m/z. The carrier gas used was Helium 6.0
mance of the MaEO/water Pickering emulsion against SARS-CoV-2 (gas flow = 1 mL min− 1). The temperature program was as follows:
responsible for the global epidemy of coronavirus disease 2019 injector temperature = 250 ◦ C; initial column temperature = 60 ◦ C; final
(COVID-19) since this emulsion type can be applied to design antimi­ column temperature = 270 ◦ C; heating rate = 10 ◦ C min− 1; column
crobial cosmetics and pharmaceutical products with sustainable and pressure = 8.2 psi. For the analysis, 5 μL of essential oil was dissolved in
renewable materials. 2 mL of hexane, and only 5 μL of the solution was injected into the
In this contribution, we evaluate several process parameters on the GC–MS oven. The essential oil compounds were identified with the NIST
stability of MaEO/water Pickering emulsions stabilized with CNFs using MS Search 2.0 and MS Data Review Software (Varian), using the NIST
design of experiments (DoE). For this purpose, cremation, sedimenta­ MSMS and NIST 02 MS databases.
tion, stability index, and drop size of the emulsions were monitored over
time. Also, the antimicrobial performance of the MaEO and its emulsion 2.4.2. Optical microscopy
was evaluated against E. coli, S. aureus and SARS-CoV-2. Several exciting The oil droplets in the emulsions were analyzed using a Leica DM KM
points about the preparation and stability of the emulsions were dis­ optical microscope (Leica Microsystems GmbH, Wetzlar, Germany). An
closed here from a DoE perspective. The MaEO emulsion was able to emulsion droplet was added to a glass slide, and representative images
inactivate the coronavirus responsible for the biggest epidemic faced by were obtained on days 0 and 30 (statical storage). The droplet diameters
humanity in the last five years, showing great potential for use in rapid were measured using the ImageJ software as an average number of 120
sanitization products and food packaging formulations capable of droplets.

3
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

2.4.3. Emulsion storage stability 2.4.8. Bacteria inhibition zone


The samples were added to glass tubes and stored at room temper­ Stock bacterial cultures were grown in Mueller-Hinton broth with 15
ature. The stability of the emulsions was evaluated through the crema­ % glycerol and stored at − 70 ◦ C. Stock cultures were grown on Mueller-
tion phenomenon, evaluated in duplicates on 0–30 days. The cremation Hinton agar (Oxoid) and inoculated at 37 ◦ C for 24 h. The inoculating
index (CI) was calculated using Eq. (1), where Htotal and Hcream are the culture (for 24 h with turbidity equivalent to 0.5 Mc Farland, a con­
total heights of the emulsion and the cream layer, respectively. centration equivalent to 1.5 × 108 CFU mL− 1) prepared the isolated and
standard bacterial strain suspensions. Then, agar (30 mL) was melted
Hcream
CI = × 100 (1) and cooled to 48 ± 1 ◦ C. Subsequently, 500 μL of the bacterial inoculum
Htotal
at 1.5 × 108 CFU mL− 1 was mixed with the molten agar. Then, this
The sedimentation ratio (Sr) was measured according to Eq. (2), mixture was transferred to a Petri dish. After solidification, a diameter
where h1 is the sedimentation height (observed at the bottom of the test hole (0.9 cm) was made in the middle of the dishes, and the emulsion
tubes), and H is the height of the total suspension. (0.2–0.5 mL) was allocated into this hole. Finally, the inoculated dishes
were incubated for 24 h at 36 ± 1 ◦ C. The appearance of inhibition halos
h1
Sr = × 100 (2) around the orifice with the product was evaluated. These inhibition
H
halos correspond to the diameter of an inhibition zone merged around
The emulsion stability index (ESI) was measured in duplicates on the antimicrobial samples. We used common commercial sanitizers
0–30 days and calculated according to Eq. (3) [25,26]. (hypochlorite and Lysoform®) as antimicrobial controls.
Hs
ESI = × 100 (3) 2.4.9. Bacteria colony-forming counting assays
Ht
For the bacteria counting, 100 μL of a previously adjusted stock
where, Hs is the height of the aqueous layer and Ht is the total height of bacteria inoculum (105 CFU.mL− 1 in Nutrient Broth, HIMEDIA, India)
the emulsion layer. was mixed with 9.9 mL of nutrient broth (HIMEDIA, India) and 100 μL of
the essential oil or emulsion. The samples were incubated at 36 ± 1 ◦ C
2.4.4. Fourier-transform Raman spectroscopy (FT-Raman) for 24 h with 85 % humidity. Then, 100 μL of the incubated samples
The emulsions were analyzed in the FT-Raman model MultiRaman were mixed with 9,9 mL of agar nutrient medium (HIMEDIA, India),
(Bruker Optics, USA) spectrometer using the laser at the wavelength of pour-plated into Petri dishes (with 90 mm diameter) and incubated at 36
1064 nm as an excitation source (laser power of 150 mW). The data ± 1 ◦ C for 24 h with 85 % humidity. The inoculum viability was
acquisition was carried out in the spectral window of 600–4000 cm− 1 confirmed via a pour-plate method on agar nutrient as the positive
using 64 scans and 2 cm− 1 spectral resolution. control. The number of counting forming units was measured with
ImageJ software. The tests were carried out in triplicate. The total
2.4.5. Fourier-transform infrared spectroscopy (FTIR) bacterial count (in CFU.mL− 1) was determined by multiplying the
Fourier-transform infrared spectroscopy (FTIR) with attenuated total number of bacterial colonies counted in the dishes by the dilution factor
reflectance (ATR) diamond accessory was performed on Frontier 94,942 (10− 4) of the plated bacterial inoculum.
equipment (PerkinElmer Inc., Massachusetts, USA). The spectra were
collected with 4 cm− 1 spectral resolution, 64 scans, from 4000 to 400 2.4.10. SARS-CoV-2 virucidal assays
cm− 1. The emulsions were dripped onto the ATR crystal until covered, Nasopharyngeal samples collected from patients and resuspended in
and then the spectra were collected. saline were tested for SARS-CoV-2 in the laboratory routine diagnostic
assay by the RT-qPCR method (real-time reverse transcription-
2.4.6. Essential oil releasing polymerase chain reaction), and the spare sample was applied in this
The MaEO releasing from the emulsions was monitored in the study. Then, 100 μL of a positive sample were diluted in 100 μL of saline
equipment STA 6000 (PerkinElmer®, Inc., Massachusetts, USA) with using a swab from the kit to mix. This swab was dipped into 300 μL of the
alumina pans, using 20–25 mg of the emulsion. The release was buffer from the kit, and five drops of the diluted positive SARS-CoV-2
measured under isothermal conditions at 37 ◦ C under an N2 atmosphere inoculum were added to the Rapid Antigen Test device to confirm the
(gas flow = 20 mL.min− 1). Since the MaEO was released with the water, integrity of the viral particles, according to the manufacturer's in­
the weight fraction (wt%) of oil released was obtained from the TGA structions. The run-time of chromatography was 15 min. After obtaining
measurements correcting the weight data by considering the specific a positive control, a mixture of 100 μL of the same SARS-CoV-2 positive
mass (ρ) and volume fraction of water (ρ = 1 g.cm− 3) and MaEO (ρ = 0.9 sample and 100 μL of the essential oil or the MaEO/water emulsion was
g.cm− 3) in the emulsions. prepared. The samples were tested at different contact times (0, 5, 15,
and 20 min). The tests were performed in duplicate.
2.4.7. Minimal inhibition concentration (MIC) and minimal bactericidal
concentration (MBC) 2.4.11. Statistical analysis and design of experiments (DoE)
The MIC and MBC assays were determined by the resazurin protocol One-way analysis of variance (ANOVA) was performed in GraphPad
involving successive microdilution in a culture broth [27]. Twelve serial Prism 7 (Dotmatics, Massachusetts, USA) to evaluate the significant
dilutions (40, 16, 6.4, 2.6 % v/v, until 0.002 % v/v) were performed in a difference between data, using Tukey's test with a confidence level of 95
96-well polystyrene plate (Kasvi). The first plate was made up of 40 % v/ %. The effects of the parameters used to prepare the emulsions were
v of the antimicrobial sample (MaEO or emulsion), diluted sequentially evaluated by a factorial DoE, using ANOVA and factor analysis with t-
in 150 μL of Mueller-Hinton nutrient broth until the plates reached a test and Bonferroni test. The effects of these factors and their in­
total volume of 250 μL. Then, all diluted mixtures were inoculated with teractions on the emulsions' stability properties were presented in Pareto
100 μL of a previous bacteria inoculum (1.5 × 108 CFU mL− 1) and charts. DoE analyses were carried out on Design Expert 11 (Stat-Ease,
incubated for 24 h at 36 ± 1 ◦ C. Then, 50 μL of the serial inoculates were Inc., Minnesota, USA). In the Pareto charts, positive effect means that
plated in Petri dishes filled with Mueller-Hinton nutrient agar (Oxoid) the factor increases the property, and negative effect means that the
and incubated at 36 ± 1 ◦ C for 24 h. The MBC corresponds to the factor reduces the property, according to the results obtained by factor
minimum concentration at which there is no formation of bacteria analysis.
colony-forming units (CFUs), using resazurin as an indicator of cell
growth [27,28].

4
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Table 1
Chemical composition of Melaleuca alternifolia essential oil (MaEO) from GC–MS data.
Chemical component Retention time (min) Area Peak area percentage (%)

Terpinen-4-ol 7.671 14,800,000 40.40


γ-Terpinene 5.871 8,134,000 22.20
α-Terpinene 5.309 4,539,000 12.39
Cimene 5.407 2,047,000 5.59
α-Terpineol 7.847 1,877,000 5.12
Terpinolene 6.27 1,659,000 4.53
1R-α-pinene 4.243 1,218,000 3.32
Eucalyptol (cineole) 5.541 973,252 2.66
Limonene 5.477 580,725 1.59
α-Tujene 4.124 198,543 0.54
α-Phellandrene 5.169 130,438 0.36
1S-β-pinene 4.87 122,567 0.33
1R-β-pinene 4.82 103,764 0.28
3-Carene 7.095 68,112 0.19
1S-α-pinene 4.717 57,517 0.16
(Z,Z,Z)-8,11,14-eicosatrienoic acid, methyl ester 4.022 60,373 0.16
4-Carene 6.846 50,472 0.14
Thujone 7.192 8,920 0.02
Camphene 6.493 7,130 0.02

3. Results and discussion ketones, and phenols [30,31].

3.1. GC–MS
3.2. Design of experiments and optical microscopy and emulsion stability

The MaEO chemical composition was evaluated by GC–MS, and the


The CNFs (in 20 and 30 wt/v %) effectively stabilized the MaEO/
nineteen constituent substances identified are detailed in Table 1. The
water Pickering emulsions against creaming and sedimentation within
GC–MS chromatogram and spectra are presented in Material Supple­
the 30-day shelf life under static conditions, as shown in optical mi­
mentary. The main chemical components in the MaEO are ternenen-4-ol
croscopy images and droplet size distributions in Figs. S2 and S3 (Sup­
(40.4 %), γ-terpinene (22.2 %), α-terpinene (12.39 %), cymene (5.59 %),
porting Information). The Pareto charts are shown in Figs. S4 and S5,
and α-terpineol (5.12 %) that are a chemical composition very similar to
indicating the effects of EO content, CNF content, mixing rotation, and
other results reported in the literature [20,22,29], with differences in
mixing time adopted in preparing the emulsions. As can be identified in
the percentage composition that is expected since it depends on the part
Pareto charts, the MaEO average droplet diameter in the emulsions is
of the plant used in extraction, extraction method used, characterization
enhanced by the increase of oil content, mainly at the 30-day storage
method, crop region and climate [22].
time. The increase in CNF content and mixing time prevents the oil
γ-Terpinene, α-terpinene, and cymene are terpenes (also known as
droplets' coalescence due to the Ostwald ripening mechanism [12],
isoprenoids) that are hydrocarbons based on isoprene (2-methylbuta-
principally at the 30-day storage according to the t-value and Bonferroni
1,3-diene) units rearranged into cyclic or aliphatic chemical structures.
limit tests. Also, the mixing rotation and CNF content have significant
Ternenen-4-ol and α-terpineol are terpenoids that are terpene modified
synergetic effects on the droplet stability right after preparing the
with different functional groups with oxygen atoms (i.e., oxygenated
emulsion.
terpene derivatives). Terpenoids are usually found in essential oils of
The emulsion stability index (ESI) tends to be increased by the CNF
different classes, such as alcohols, aldehydes, esters, ether, epoxides,
content that can be connected with the high and preferential

Fig. 2. Optical microscopy images of the main MaEO/water Pickering emulsions: (a) 30EO/5T/0.5CNF/15RPM and (b) 20EO/5T/0.5CNF/12RPM samples.

5
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 3. (a) FTIR spectra of MaEO and MaEO/water emulsions stabilized with CNF. Zoom in the regions: (b) 500–1600 cm− 1; (c) 1500–2500 cm− 1; and (d)
2500–4000 cm− 1. Spectral data from MaEO (black line), 20EO/5T/0.5CNF/12RPM (blue line) and 30EO/5T/0.5CNF/15RPM (red line) samples. Each color of the
sets is associated with signals from different substances: cellulose (light green), water (cyan), terpinen-4-ol (orange), γ-terpinene (pink), α-terpinene (navy blue),
terpineol (yellow) and terpinolene (purple).

interactions between the CNFs, leading to CNF output of the oil/water


Table 2 interface and the cellulose aggregation over storage time. The ESI values
Signals observed in the FTIR spectra from the MaEO and MaEO/W emulsions
are decreased by the MaEO content within the concentration range used
stabilized with CNF.
to prepare the emulsions within the 30-days of emulsion storage. Mixing
Chemical FTIR signals (cm− 1) rotation and mixing time improves the emulsion stability along the
compound
storage time. However, no emulsion preparation factors affect the ESI
Cellulose 1155 (C–O–C asymmetrical stretching) data predominantly, but they act simultaneously and synergetically with
1160 (C–O–C asymmetrical stretching)
very similar intensity over the ESI results in a complex way, according to
1315 (OH vibration in the cellulose rings)
1430 (symmetric CH2 bending) the Pareto charts from DoE data.
2019 and 2129 (C––C stretch vibration) The 30EO/5T/0.5CNF/15RPM and 20EO/5T/0.5CNF/12RPM
2850 (CH2 group vibration) emulsions are the main stable emulsions concerning the smallest droplet
2910 (CH group vibration) sizes and ESI values, including the lowest variations of these parameters
Water 1640 (H2O scissor vibration mode)
2100 (H–O–H bending vibration)
along the 30 days of storage as shown in Fig. 2. Both emulsions were
3330 (OH stretching vibrations) stabilized with 0.5 % wt/v CNF and using 1 % wt/v visually tends to
Terpinen-4-ol 799, 865, 889, 985, 1160 and 1310 (C–O and C–C aromatic increase ESI, as indicated by Pareto diagrams. Besides the lowest
ring vibrations) manufacturing cost, CNFs have a high aspect ratio and flexibility to
3450 (OH vibrations)
stabilize over time by adsorption and fiber entanglement to be more
γ-Terpinene 780, 815, 830,1160 and 1375 (C–C aromatic ring vibrations)
2820 and 2872 (CH3 and CH2 vibrations) effective via steric hindrance much lower than cellulose nanostructures
α-Terpinene 950 and 1470 (C–C aromatic ring vibrations) as CNCs with a lower aspect ratio that cannot involve all the surface area
1640 (C––C vibrations) of the oil droplets by adsorption [26,32].
2967 (CH3 and CH2 vibrations)
3330 (O–H vibrations)
Terpineol 1365 and 1440 (C–O and C–C aromatic ring vibrations)
3450 (O–H vibrations) 3.3. FTIR
Terpinolene 1020, 1125 and 1514 (C–O and C–C aromatic ring vibrations)
1640 and 1750 (C––C vibrations) The FTIR spectra of MaEO and more stable emulsions (30EO/5T/
2910 (CH3 and CH2 vibrations) 0.5CNF/15RPM and 20EO/5T/0.5CNF/12RPM samples) are shown in
3390 (O–H vibrations)
Fig. 3. The FTIR spectra from MaEO reflect the chemical structures of the
main chemical groups identified by GC–MS. The infrared bands between

6
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 4. (a) FT-Raman spectra of MaEO and MaEO/water emulsions stabilized with CNF. Zoom in the spectral regions: (b) 500–1600 cm− 1; (c) 1500–2500 cm− 1; and
(d) 2500–4000 cm− 1. Each color of the sets is associated with signals from different substances: cellulose (light green), water (cyan), terpinen-4-ol (orange),
γ-terpinene (pink), α-terpinene (navy blue), terpineol (yellow) and terpinolene (purple).

1600 and 1800 cm− 1 correspond to alkene chemical bonds (C– – C), while overlap among the water, MaEO, and cellulose absorption bands since
the bands in the 1370–1470 cm− 1 region are associated with –CH2– there is an increase in the OH moiety from the cellulose and water
and –CH3 scissoring vibrations [33]. The FTIR signals around 2900 (Fig. 3). The FTIR band from 2000 to 2300 cm− 1 is associated with the
cm− 1 are from the CH3 and CH2 vibrations on the hydrocarbon struc­ H–O–H bending vibration forming a hydrogen bonding network of
tures of the terpenes and terpenoids [34]. Other notable bands from liquid water in the Pickering emulsions that can be influenced by tem­
C–O aromatic ring vibrations and C–C vibrational modes in hydro­ perature, including the presence and concentration of anions and cations
carbon cyclic skeletons found below 1500 cm− 1 are influenced by the in the water media [37]. Moreover, the high water content in Pickering
position and chemical nature of the aromatic ring's substituent. Table 2 MaEO/water emulsion causes an overlap of the FTIR characteristic ab­
summarizes the FTIR signals detected by infrared spectroscopy. sorption signals from the MaEO compounds located at 500–1500 cm− 1
The band intensity at 3600–3000 cm− 1 is due to the OH vibrations and 3200–3600 cm− 1. Also, due to the outstanding water concentration,
from terpenoids in MaEO, such as terpinen-4-ol, terpineol, α-terpinene, the infrared signals from cellulose nanostructures are not observed in
and terpinolene [35]. However, α-terpinene, terpinolene and terpenes the FTIR spectra of the emulsions. According to the literature [24,38],
strongly absorb infrared radiation with wavenumber from 3200 to 3600 FTIR characteristic signals for the cellulose are C–O–C asymmetrical
cm− 1 due to C–H asymmetric and symmetric stretching modes. It is stretching at 1155 cm− 1, C–O–C asymmetrical stretching at 1160
highlighted in Fig. 3b, c and d the FTIR characteristics signals from cm− 1, OH vibration in the cellulose rings at 1315 cm− 1, symmetric CH2
terpinen-4-ol (799, 865, 889, 985, 1160, 1310, and 3450 cm− 1), γ-ter­ bending at 1430 cm− 1, C– – C stretch vibration at 2019 and 2129 cm− 1,
pinene (780, 815, 830,1160, 1375, 2820, and 2872 cm− 1), α-terpinene CH2 group vibration at 2850 cm− 1, and CH groups at 2910 cm− 1.
(950, 1470, 1640, 2967 and 3330 cm− 1), terpineol (1365, 1440 and The slight shift of the FTIR signals at 2500–3000 cm− 1 spectral re­
3450 cm− 1), and terpinolene (1020, 1125, 1514, 1640, 1750, 2910 and gion (CH2 and CH3 vibrational modes) to higher wavenumbers in Fig. 3d
3390 cm− 1) [35,36]. can be connected with modification in molecular vibrations of the MaEO
The presence of CNF and water in O/W emulsions improves the in­ components (γ-terpinene, terpinolene, and α-terpinene) due to their
tensity of the absorption band from 3200 to 3600 cm− 1 due to the physical and intermolecular interactions with the surface of the CNFs in

7
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Table 3 the aqueous medium, as highlighted in the inset shown in Fig. 4d. This
Raman shift signals observed in the FT-Raman spectra from the MaEO and phenomenon can be attributed to the formation of hydrogen bonds be­
MaEO/W emulsions stabilized with CNF. tween water and cellulose nanofibrils in the emulsion with less energy
Chemical Raman shifts (cm− 1) than the strong hydrogen bonds between water molecules [44].
compound The characteristic bands in the 700–800 cm− 1 region from cellulose
Cellulose 1090 (glycosidic C–O–C, ring vibration) due to OH bond vibrations can be identified by FTIR spectroscopy.
1155 (asymmetric C–C and C–O ring elongation) However, such vibrational modes commonly are not detected by Raman
1375 (CH2, HCC, HCO and COH vibrations) spectroscopies because there is a great dipole moment and low polar­
1460 (amorphous phase)
izability associated with the OH bond among cellulose macromolecules.
2895 (CH and CH2 elongation)
3250 (O–H elongation) However, the FT-Raman spectra of the emulsions display Raman pattern
Water 1600 (H2O scissor vibration mode) shifts at 725 (CCC, COC, OCO, CCO, and OH out-of-plane bending) and
3200–3500 (water network of hydrogen bonds) 800 cm− 1 (OH bending), suggesting strong interactions between the
Terpinen-4-ol 728 (ring's stretching and deformation) CNFs with water and the oxygenated terpenes in the MaEO by hydrogen
1375, 1430 and 1445 (CH3 and CH2 flexion and elongation
modes)
bonds [45]. Fig. 5 illustrates the possible intermolecular interactions
1680 (C––C stretching) between water and MaEO compounds with the hydrophobic and hy­
2880, 2920 and 2960 (C–H symmetrical and asymmetrical drophilic cellulose sites [46], as the FTIR and FT-Raman data suggested.
stretching)
755 (ring's stretching and deformation)
3.5. Essential oil releasing
γ-Terpinene
1420 and 1450 (CH3 and CH2 flexion and elongation modes)
1700 (C––C stretching)
2880, 2920 and 2960 (C–H symmetrical and asymmetrical We propose the usage of thermal analysis to evaluate the controlled
stretching) release of MaEO from emulsions at a temperature similar to that used in
α-Terpinene 1380 and 1465 (CH3 and CH2 flexion and elongation modes) bactericidal assays (37 ◦ C). The method is an accelerated test for the oil
1660 (C––C stretching)
2880, 2920 and 2960 (C–H symmetrical and asymmetrical
release since the carrier gas helps release the essential oil by convection,
stretching) making it the fastest release test that allows a comparative investigation
Terpineol 1310 and 1435 (CH3 and CH2 flexion and elongation modes) of the MaEO release rate and stability between emulsions.
2880, 2920 and 2960 (C–H symmetrical and asymmetrical According to the isothermal curves (Fig. S6), MaEO in the emulsions
stretching)
starts to be liberated at the beginning of the experiment. It takes 64 and
Terpinolene 723 (ring's stretching and deformation)
1445 (CH3 and CH2 flexion and elongation modes) 58 min for the MaEO to be released entirely from the 20EO/5T/0.5CNF/
1660 (C––C stretching) 12RPM and 30EO/5T/0.5CNF/15RPM emulsions, respectively.
2880, 2920 and 2960 (C–H symmetrical and asymmetrical The oil-releasing kinetics follows a zero-order model, i.e., MaEO is
stretching) released from the emulsion by a linear relationship with the time ac­
cording to Eq. (4) [47–49].
the emulsions. These interactions with cellulose cause changes in the Mt/M = k0 t + C0 (4)
vibration modes and vibrational energy levels available to the MaEO ∞

compounds in the multiphasic system stabilized by the CNF. where Mt/ corresponds to the fraction of MaEO released at time t, and
Souza et al. [12] observed that cellulose in the form of CNF prefer­ M∞
entially interacts with OH groups of the components of different k0 is a constant connected with the oil release rate.
essential oils (cinnamon, cardamon, and Ho wood). In contrast, our FTIR A zero-order model suggests that the diffusion rate is practically
results here indicate changes in the vibration modes of CH2 and CH3 due independent of the emulsion stability [24]. The coefficient parameters
to interactions between terpene groups with non-polar groups in cellu­ for the kinetic mechanism involved in the MaEO liberation are listed in
lose, suggesting that the OH groups in the terpinen-4-ol molecules (the Table 4. The k0 values are − 0.6 and − 0.3 (% v/v).min− 1 for the 30EO/
major component of MaEO) probably cannot interact effectively with 5T/0.5CNF/15RPM and 20EO/5T/0.5CNF/12RPM samples, indicating
the polar groups of cellulose due to steric hindrance caused by the iso­ that the MaEO releasing is faster for the emulsion with the highest ratio
propyl substituent group on terpinen-4-ol. between the oil and CNF content what is expected since there is a higher
amount of oil to be stabilized by the same CNF content.

3.4. FT-Raman 3.6. MIC and MBC

The FT-Raman spectrum of the MaEO is shown in Fig. 4. All terpe­ MIC and MBC assays were performed to evaluate the bactericidal
noids also showed bands around 650–800 cm− 1, typically associated performance of the MaEO against E. coli (Gram-negative bacteria strain,
with ring stretching and deformations. The band up to 1160 cm− 1 refers G− ) and S. aureus (Gram-positive bacteria strain, G+). The MIC and
to C–O elongation in terpenoids. Ring vibrations from the cyclohexane MBC values are detailed in Table 5. The MIC value of MaEO is 0.2 ± 0.1
exhibit bands at 1020 and 950 cm− 1. Symmetrical and asymmetrical % v/v against E. coli, but 2.3 ± 0.9 % v/v against S. aureus. The MIC/
C–H stretching modes of terpenes and terpenoids are associated with MBC ratio for the MaEO against both G+ and G− strains evaluated are
bands from 2750 to 3050 cm− 1 [39]. The band observed at 1445 cm− 1 is ≤4, indicating that MaEO can be considered a bactericidal agent in
linked to the flexion and elongation modes of the CH3 and CH2 groups. which it is possible to kill 99.9 % of the bacteria cells under adequate
The bands around 1600 cm− 1 in the MaEO spectra are associated with concentrations [50]. A MIC/MBC ratio higher than 4 is related to a
C–– C stretching [40,41], as listed in Table 3. bacteriostatic drug.
Cellulose Raman spectra bands are seen in Fig. 4a–d at 1090 The O/W emulsions present higher MIC and MBC values than the
(glycosidic COC, ring vibration), 1155 (asymmetric CC and CO ring MaEO due to the lowest amount of antimicrobial compounds in the
elongation), 1375 (CH2, HCC, HCO and COH vibration modes), 1460 emulsified system. The MIC and MBC doses of the emulsions are 6.4 ±
(amorphous cellulose), 2895 (CH and CH2 elongation) and 3250 cm− 1 0.7 % v/v against E. coli and do not vary with the amount of MaEO
(OH elongation) [42,43]. The pure water band is reported at 1600 and content. The MIC and MBC data are identical for 20EO/5T/0.5CNF/
3400 cm− 1 [44], but O/W emulsions show Raman signals at 1600, 3200 12RPM and 30EO/5T/0.5CNF/15RPM emulsion against S. aureus,
and 3350 cm− 1. This reduction in the water band in the region although they are different conceptual concepts. Garzoli et al. [51]
3200–3350 cm− 1 indicates changes in the network of H–O–H bonds in observed a similar phenomenon for MIC and MBC results for essential

8
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 5. Schematic representation of the


chemical interactions at the cellulose/
water and cellulose/MaEO interfaces in
the Pickering emulsions. Hydrogen bonds
between cellulose and the terpenoids and
CH-CH dipole-induced-dipole interactions
(van der Waals forces) between cellulose
terpenes in the MaEO at the cellulose/oil
interface. At the cellulose/water interface,
there are hydrogen bonds between cellu­
lose and water molecules. The hydrogen
bonds between the CH-OH cellulose
groups are the main interactions that
contribute to the cohesion of CNFs sur­
rounding the oil droplets [46].

oils from the Lavandula genus against G− bacteria strains.


Table 4 For these outstanding bactericidal features and broad-spectrum and
Zero-order kinetic model parameters for the MaEO releasing from the emulsion antimicrobial activity, MaEO is commonly applied as an eco-friendly
at 37 ◦ C. topical antiseptic and antibacterial component in cosmetics with po­
Sample Parameters tential use in the food industry to prolong the shelf life of food products
k0 [(% v/v).min− 1] C0 (% v/v)
[52]. S. aureus is an opportunistic pathogen that, when installed in the
R2
host, leads to superficial (suppurative infection in soft tissues) and deep
20EO/5T/0.5CNF/12RPM − 0.3 ± 0.1 20.2 ± 0.1 0.99
(systemic spread by bacteremia) infections, even toxigenic infections
30EO/5T/0.5CNF/15RPM − 0.6 ± 0.1 30.1 ± 0.1 0.99
characterized by the production of toxins by systemic action [53]. E. coli
are facultative anaerobic bacteria and a member of the human intestinal
microbiota, which can cause intestinal and urinary tract infections by
Table 5 different mechanisms [53].
Minimal Inhibition Concentration (MIC) and Minimal Bactericidal Concentra­ The antimicrobial performance of the MaEO has been attributed to
tion (MBC) of Melaleuca alternifolia essential oil (MaEO) against S. aureus (ATCC its components, mainly terpinen-4-ol and α-terpineol, which cause
#6538) and E. coli (ATCC #11229). several injuries to the bacteria cells via membrane-damaging mecha­
Sample S. aureus E. coli nisms [52,54]. It is reported that MaEO inhibits cellular respiration and
MIC (% v/ MBC (% v/ MIC (% v/ MBC (% v/ disbalances the potassium ions at the cytoplasmic membrane that pro­
v) v) v) v) voke cellular morphological damages, accompanied by disruption of the
MaEO 2.3 ± 0.9 4.0 ± 1.5 0.2 ± 0.1 0.3 ± 0.1
cell permeability barrier for the S. aureus and E. coli strains [54]. Other
20EO/5T/0.5CNF/ 16.0 ± 1.6 16.0 ± 1.6 6.4 ± 0.7 6.4 ± 0.7 authors indicate the coagulation of cytoplasmic constituents in E. coli
12RPM cells after direct contact with MaEO due to monoterpene penetration
30EO/5T/0.5CNF/ 2.6 ± 2.5 2.6 ± 2.5 6.4 ± 0.7 6.4 ± 0.7 into the cell associated with a previous cell permeability barrier
15RPM
compromising that was occasioned by irreversible membrane injuries by
MaEO constituents [21]. Moreover, monoterpenes in MaEO can damage
the lipid constituents of the microbial membranes [54]. Then,

9
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 6. Inhibition halos for hypochlorite, Lysoform, MaEO, and MaEO/water emulsions against E. coli (a, c) and S. aureus (b, d). The values are presented as mean ±
standard deviation from triplicate measurements. Only the sample data without statistically significant differences are marked on the graphs (ns, where p > 0.05),
according to Tukey's test with a 95 % confidence level. *The scale bars of graphs (a, b) are in different scales.

alternative effectiveness and synergistic interactions between antimi­ membrane surface (covered by a peptidoglycan cell wall) predominantly
crobial MaEO compounds may result in simultaneous and different charged with negative charges from phosphate groups in teichoic acids,
bactericidal mechanisms that lead to whole-cell lysis or compromising of which are essential components of the lipopolysaccharide cell wall to
the cell machinery, eventually leading to the bacteria's death. stabilize the G+ cell wall layer through electrostatic interactions with
MBC corresponds to the concentration at which the antimicrobial cations, such as potassium [58]. Besides G− strains having an additional
sample can inactivate >99.99 % of the bacteria in the medium [55]. The outer lipidic membrane responsible for protecting the peptidoglycan cell
MBC results in Table 5, the G+ strain seems more resistant to the wall and the inner cytoplasmic membrane, they are also predominantly
bactericidal components in the MaEO, evidencing different susceptibil­ negatively charged since their outer lipidic membranes have a strong
ity to tea tree essential oil as reported by other authors [21,56]. This negative charge conferred by the lipopolysaccharide that, with lipo­
differential susceptibility effect is still not well understood. Even more, it proteins and phospholipids, cover the peptidoglycan cell layer [58].
is an effect contrary to what was expected since G+ bacteria have only Moreover, the plasma membrane of G+ bacteria is surrounded by a
an outer layer, which is ineffective in hampering the interactions be­ thinner cell wall with several peptidoglycan layers, forming a thickness
tween external agents with the cytoplasmic membrane, generally mak­ of 20–100 nm. In contrast, G− strains display a thin cell wall, showing a
ing the G+ bacteria more susceptible and fragile to antimicrobial agents few nanometric thick (1–8 nm) due to the presence of a single pepti­
[56]. Antagonastly, G− bacteria cells have an additional outer mem­ doglycan layer or a reduced number of peptidoglycan layers [59–61].
brane responsible for protecting the inner cytoplasmic membrane, These differences between the cell wall thickness of the G+ and G− cell
which generally makes them more resistant than G+ strains to external walls must play a crucial role in the bactericidal resistance of these
antimicrobial agents [56]. Carson et al. reported that the primary anti­ bacteria strains against MaEO since monoterpenes are potent destroyers
microbial mechanism against S. aureus is a complex delayed rupture of of lipid structures [54].
the cytoplasmic membrane marked by inducing the leakage of cyto­
plasmic material [57]. Terpinen-4-ol cannot lyse the S. aureus cells 3.7. Bacteria inhibition zones
completely, while α-pinene, β-pinene, and limonene inhibit the respi­
ratory activity of the mitochondria. The inhibition zone technique using the agar diffusion test was
Surface charges and chemical composition of the cell and outer-cell performed to evaluate the bacteriostatic capacity of the MaEO and
membranes of the G− and G+ bacterial strains must be connected with MaEO/emulsions. The formation of inhibition halos indicates that the
their susceptibility to MaEO. G+ strains present a cytoplasmatic product under analysis has bacteriostatic properties, evidencing the

10
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

Fig. 7. Bacterial counting and relative bacterial reduction for the MaEO and MaEO/water emulsions against E. coli (a, b) and S. aureus (c, d). The values are presented
as mean ± standard deviation from triplicate measurements. *Sample data are significantly different (*p > 0.05), according to Tukey's test with a 95 % confi­
dence level.

inhibiting of bacterial growth and reproduction without causing the release the volatile MaEO bactericidal agents gradually within the MIC
bacteria death. The inhibition zone diameter depends on the diffusion of values, which is essential to ensure the prevention of bacteria prolifer­
the agar preparation. As a comparative control test, we also evaluated ation of both G− and G+ strains in a long-standing, avoiding a quick loss
the formation of inhibition halos on agar diffusion plate assays with of antimicrobial activities of the MaEO [12]. Then, the protection of the
hypochlorite and Lysoform®, which are effective disinfectants CNFs of these agents against the decomposition occasioned by external
commercially available and commonly used by the Brazilian population. agents (e.g., sunlight) does not compromise the bactericidal perfor­
As shown in the graphs in Fig. 6a, the inhibition zone diameter mance of the MaEO.
observed for the MaEO is equal to 2.95 ± 0.35 cm against E. coli, which
is a value 50 % higher than that detected for Lysoform® (1.65 ± 0.05
cm) but statically equal to that obtained with hypochlorite (3.15 ± 0.52 3.8. Bacteria colony-forming counting assays
cm). Fig. 6c and d presents the images of the agar diffusion dishes with
the inhibition zones. The MaEO/water emulsions (30EO/5T/0.5CNF/ MaEO presents bacterial colony-forming counting of 4.9 ± 0.1 log
15RPM and 20EO/5T/0.5CNF/12RPM samples) also display significant (CFU.mL− 1) and 5.1 ± 0.3 log(CFU.mL− 1) against S. aureus and E. coli,
inhibition zone diameters than that observed for the hypochlorite corresponding to a bacterial reduction of around 40–42 % (in regarding
against E. coli and S. aureus. the bacterial counting from positive controls) with significantly different
According to Fig. 6b, the 30EO/5T/0.5CNF/15RPM and 20EO/5T/ (*p > 0.05), according to Tukey's test with a 95 % confidence level, as
0.5CNF/12RPM samples lead to the formation of inhibition zone di­ shown in Fig. 7. It is important to mention that the bacteria counting
ameters similar to the obtained for the Lysoform® against S. aureus (2.95 tests were performed with MaEO content below the MIC value deter­
± 0.10 cm), but they are less than that identified by the MaEO (4.15 ± mined against the S. aureus. 2-log(CFU/mL− 1), i.e., 102 CFU.mL− 1 that
0.37 cm) that can be justified by the high content of volatile antimi­ corresponds to a bacteriostatic action and indicates a 99 % reduction in
crobial agents in the tea tree essential oil. These highlightable antimi­ bacteria that involves inhibition of the bacteria growth and reproduc­
crobial results indicate that the Pickering emulsions are suitable to tion without killing the bacteria cells. Bactericidal activity is associated
with an antibacterial agent that completely prevents bacterial growth

11
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

counting tests against E. coli, the 20EO/5T/0.5CNF/12RPM and 30EO/


5T/0.5CNF/15RPM emulsion also showed statistically identical relative
bacterial reduction results (5–10 %), which is the antimicrobial per­
formance less than that for the MaEO (42.1 ± 3.1 %).
Similar to the inhibition zone results, bacterial count and ANOVA
data indicate that MaEO and MaEO/water CNF-stabilized emulsions
have significantly greater antimicrobial activity against S. aureus than
against E. coli. In contrast, the MIC and MBC results suggested the lowest
bacteriostatic and bactericidal doses of MaEO against E. coli than those
against S. aureus. D'Arrigo et al. [54] observed similar contradictory
behavior for the antimicrobial performance of MaEO, and they reported
that the E. coli strain is more susceptible than S. aureus to the antimi­
crobial compounds in MaEO over time from using time-killing assays,
while the MIC and MBC indicate that S. aureus is more susceptible to this
essential oil. However, these authors did not explain these conflicting
results for the MaEO antimicrobial sensibility from the MIC, MBC, and
time-kill assay data. In another work, Cox et al. [21] also observed a
lower MIC and MBC of MaEO against E. coli than that against S. aureus,
but identified that E. coli undergoes greater respiratory inhibition and
are more susceptible to K+ efflux due to membrane-damaging effects
than S. aureus. These findings are connected with differences in the
susceptibility of the different bacterial strains to MaEO, which seem not
to be only influenced by the bacterial cell structure and physical/
chemical factors (warmth, moisture, pH and oxygen levels). The anti­
microbial test method also influences the findings on the bactericidal
properties of an antimicrobial agent due to its experimental variables
that affect the results and lead to experimental errors. Just to exemplify,
the final colony-forming units obtained by bacterial colony counting
assays are influenced by the growth and proliferation kinetics of the
bacteria strains, which depends on the bacterial generation time.

3.9. SARS-CoV-2 virucidal assays

The SARS-CoV-2 antigen-detecting rapid diagnostic test kits were


applied to evaluate the integrity of the spike proteins (S proteins) at the
surface of the SARS-CoV-2 virions, identifying their viability to infect a
new host [63]. The rapid test used here is a chromatographic immu­
noassay for qualitatively detecting the S-specific proteins of the SARS-
CoV-2 particles in the nasopharynx inoculum. The denaturation/
disruption of these viral proteins on the surface of the virus promoted by
our samples is indicated by the absence of the red color in the “test line”
on the rapid diagnostic test kits, presented as negative result, because
their antibodies cannot bind to these antigens once damaged, as shown
in Fig. 8.
MaEO deactivated all SARS-CoV-2 virions immediately upon con­
tacting it, as indicated in Fig. 8a. The 30EO/5T/0.5CNF/15RPM emul­
sion could inactivate all virions in 15 min of direct contact since all the S
viral proteins were damaged in the inoculum (Fig. 8b). In contrast, the
20EO/5T/0.5CNF/12RPM emulsion did not inactivate all viral particles
within 20 min (Fig. 8c), which can be explained by its lowest MaOE
Fig. 8. SARS-CoV-2 antiviral tests for the MaEO (a) and MaEO/water emul­ content. However, the reduction in the intensity of the red coloration in
sions (b, c) after 0, 5, 15, and 20 min of direct contact. Negative results are the “test line” indicates partial inactivation of the virus; thus, this 20EO/
highlighted in red squares. 5T/0.5CNF/12RPM emulsion can probably lead to complete inactiva­
tion of all viable virus particles at longer direct contact times.
and reproduction, characterized by a 3-log10 reduction of the bacterial Several studies demonstrated the antiviral activity of MaEO against a
counting (in CFU.mL− 1) in the initial inoculum (i.e., a reduction like ⩾ wide range of viruses, even at doses below the cytotoxic doses, such as
99.9 % in bacterial counting) [62]. Herpes labialis, Tobacco mosaic, Herpes simplex (HSV), and Influenza A
The bacterial counting for 30EO/5T/0.5CNF/15RPM and 20EO/5T/ viruses [64–68]. The viral inactivation mechanism depends on virus
0.5CNF/12RPM emulsions are 7.2 ± 0.3 and 6.9 ± 0.1 log(CFU.mL− 1) morphology, including viral replication and viral infection molecular
against S. aureus, respectively. In contrast, the S. aureus positive control mechanisms. So far, no studies show the antiviral mechanism of MaEO
presents a bacterial counting equal to 8.5 ± 0.2 log(CFU.mL− 1). How­ against SARS-CoV-2. However, our antiviral assay results indicate that
ever, the relative bacterial reduction from 30EO/5T/0.5CNF/15RPM the inactivation mechanism of SARS-CoV-2 virions involves the
and 20EO/5T/0.5CNF/12RPM emulsions are statistically similar destruction/denaturation of Spike protein trimers that play a critical
(15–20 %) against this G+ bacteria strain. Concerning the bacterial role in the host cellular infection mechanism. Consequently, the infec­
tion ability of the SARS-CoV-2 particles is abrogated [63,69].

12
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

4. Conclusions Data availability

In conclusion, we have shown that MaEO and MaEO/water CNF Data will be made available on request.
stabilized emulsion present antibacterial performance against E. coli and
S. aureus, including virucidal performance against SARS-CoV-2. Mela­ Acknowledgment
leuca alternifolia essential oil (MaEO) deactivates the SARS-CoV-2 par­
ticles immediately by damaging their S proteins, but its emulsions take The authors thank the CAPES (Code 001), CAPES-Pandemias
much longer to deactivate. FT-Raman and FTIR data indicate different (8881.504639/2020-01), MULTIUSER CENTRAL FACILITIES (UFABC),
interactions between water and CNF with different essential oil com­ UFABC, and REVALORES Strategic Unit.
ponents in emulsions, depending on their structure and chemical
composition. The MaEO/water emulsion stabilized with a 0.5 % wt/v Appendix A. Supplementary data
CNF suspension, 30 % v/v MaEO content, mixed by 5 min using a mixing
rotation of 15 rpm, presented the highest statical stability and was Supplementary data to this article can be found online at https://doi.
capable of deactivating the SARS-CoV-2 virions in 15 min of direct org/10.1016/j.ijbiomac.2023.125228.
contact. Factorial DoE indicates that the improvement of the mixing
time is essential to stabilizing the emulsions within a 30-days storage References
time.
Here, we evidence that the bactericidal performance of the MaEO [1] N. Zhang, L. Yao, Anxiolytic effect of essential oils and their constituents: a review,
and the Pickering emulsions against E. coli and S. aureus is influenced by J. Agric. Food Chem. 67 (2019) 13790–13808, https://doi.org/10.1021/acs.
jafc.9b00433.
the antimicrobial assay type, which can lead to conflicting conclusions [2] Y. Li, C. Wang, Z. Tao, Z. Zhao, L. You, R. Zheng, X. Guo, Z. Zhang, Enhanced
about the greatest antimicrobial susceptibility among these bacteria antioxidant and antiproliferative activities of Cymbopogon citratus (DC.) Stapf
strains to the “green” bactericidal agents. Then, the antimicrobial sus­ essential oils in microemulsion, ACS Sustain. Chem. Eng. 7 (2019) 15173–15181,
https://doi.org/10.1021/acssuschemeng.9b01606.
ceptibility of one bacteria strain to the essential oil must be evaluated by
[3] F. Froiio, A. Mosaddik, M.T. Morshed, D. Paolino, H. Fessi, A. Elaissari, Edible
different microbiological assays to identify such inconsistencies, Polymers for Essential Oils Encapsulation: Application in Food Preservation, 2019,
enabling a better decision on which antimicrobial agent is most effective https://doi.org/10.1021/acs.iecr.9b02418.
[4] J.B. Reis, L.A. Figueiredo, G.M. Castorani, S.M.O.M. Veiga, Avaliação da atividade
against a given bacterial strain.
antimicrobiana dos óleos essenciais contra patógenos alimentares, Braz. J. Heal.
Rev. 3 (2020) 342–363, https://doi.org/10.34119/bjhrv3n1-025.
Ethical statement [5] K. Heise, E. Kontturi, Y. Allahverdiyeva, T. Tammelin, M.B. Linder, O.
Ikkala Nonappa, Nanocellulose: recent fundamental advances and emerging
biological and biomimicking applications, Adv. Mater. 33 (2021) 2004349,
The antiviral tests against SARS-CoV-2 were performed following the https://doi.org/10.1002/adma.202004349.
standard regulations approved by the Research Ethics Committee of the [6] X. Liu, X. Qi, Y. Guan, Y. He, S. Li, H. Liu, L. Zhou, C. Wei, C. Yu, Y. Chen,
Transparent and strong polymer nanocomposites generated from Pickering
Federal University of ABC and Instituto Adolfo Lutz CAAE:
emulsion gels stabilized by cellulose nanofibrils, Carbohydr. Polym. 224 (2019),
49573421.2.3001.0059, and by the Technical-Scientific Council of 115202, https://doi.org/10.1016/j.carbpol.2019.115202.
Instituto Adolfo Lutz, CTC 18-N/2021. All procedures followed the [7] O. Nechyporchuk, M.N. Belgacem, J. Bras, Production of cellulose nanofibrils: a
review of recent advances, Ind. Crop. Prod. 93 (2016) 2–25, https://doi.org/
biosafety standards of the Instituto Adolfo Lutz of the Regional Center of
10.1016/j.indcrop.2016.02.016.
Laboratories of Santo André. [8] D. Trache, A.F. Tarchoun, M. Derradji, T.S. Hamidon, N. Masruchin, N. Brosse, M.
H. Hussin, Nanocellulose: from fundamentals to advanced applications, Front.
Chem. 8 (2020), https://doi.org/10.3389/fchem.2020.00392.
Funding [9] T. Yi, H. Zhao, Q. Mo, D. Pan, Y. Liu, L. Huang, H. Xu, B. Hu, H. Song, From
cellulose to cellulose nanofibrils—a comprehensive review of the preparation and
This research was funded by Fundação de Amparo à Pesquisa do modification of cellulose nanofibrils, Materials (Basel) 13 (2020) 5062, https://doi.
org/10.3390/ma13225062.
Estado de São Paulo (2020/12208-9), Conselho Nacional de Desenvol­
[10] H. Dai, J. Wu, H. Zhang, Y. Chen, L. Ma, H. Huang, Y. Huang, Y. Zhang, Recent
vimento Científico e Tecnológico (305819/2017-8), CAPES-Pandemias advances on cellulose nanocrystals for Pickering emulsions: development and
(88881.504639/2020-01). challenge, Trends Food Sci. Technol. 102 (2020) 16–29, https://doi.org/10.1016/
j.tifs.2020.05.016.
[11] Y. Zhou, S. Sun, W. Bei, M.R. Zahi, Q. Yuan, H. Liang, Preparation and
CRediT authorship contribution statement antimicrobial activity of oregano essential oil Pickering emulsion stabilized by
cellulose nanocrystals, Int. J. Biol. Macromol. 112 (2018) 7–13, https://doi.org/
10.1016/j.ijbiomac.2018.01.102.
Greiciele da S. Ferreira: Conceptualization, Methodology, Valida­ [12] A.G. Souza, R.R. Ferreira, L.C. Paula, L.F.G. Setz, D.S. Rosa, The effect of essential
tion, Formal analysis, Investigation, Writing – original draft, Writing – oil chemical structures on Pickering emulsion stabilized with cellulose nanofibrils,
review & editing, Visualization. Daniel J. da Silva: Conceptualization, J. Mol. Liq. 320 (2020), 114458, https://doi.org/10.1016/j.molliq.2020.114458.
[13] N. Bu, L. Huang, G. Cao, H. Lin, J. Pang, R. Mu, L. Wang, Konjac glucomannan/
Methodology, Validation, Formal analysis, Investigation, Writing – Pullulan films incorporated with cellulose nanofibrils-stabilized tea tree essential
original draft, Writing – review & editing, Visualization. Alana G. oil Pickering emulsions, Colloids Surf. A Physicochem. Eng. Asp. 650 (2022),
Souza: Conceptualization, Methodology, Validation. Eliana D.C. 129553, https://doi.org/10.1016/j.colsurfa.2022.129553.
[14] Y. Zhang, Y. Jiang, L. Han, B. Wang, H. Xu, Y. Zhong, L. Zhang, Z. Mao, X. Sui,
Yudice: Methodology, Validation, Formal analysis. Ivana B. de Cam­
Biodegradable regenerated cellulose-dispersed composites with improved
pos: Methodology, Validation, Formal analysis. Rute Dal Col: Meth­ properties via a Pickering emulsion process, Carbohydr. Polym. 179 (2018) 86–92,
odology, Validation, Formal analysis. Andre Mourão: Methodology, https://doi.org/10.1016/j.carbpol.2017.09.065.
Validation, Formal analysis, Writing – review & editing. Herculano S. [15] Y. Lu, J. Li, L. Ge, W. Xie, D. Wu, Pickering emulsion stabilized with fibrous
nanocelluloses: insight into fiber flexibility-emulsifying capacity relations,
Martinho: Formal analysis, Supervision, Writing – review & editing. Carbohydr. Polym. 255 (2021), 117483, https://doi.org/10.1016/j.
Derval S. Rosa: Resources, Writing – review & editing, Supervision, carbpol.2020.117483.
Project administration, Funding acquisition. [16] A.G. Souza, R.R. Ferreira, L.C. Paula, S.K. Mitra, D.S. Rosa, Starch-based films
enriched with nanocellulose-stabilized Pickering emulsions containing different
essential oils for possible applications in food packaging, Food Packag. Shelf Life
27 (2021), 100615, https://doi.org/10.1016/j.fpsl.2020.100615.
Declaration of competing interest [17] M. Liao, J.-J. Xiao, L.-J. Zhou, X. Yao, F. Tang, R.-M. Hua, X.-W. Wu, H.-Q. Cao,
Chemical composition, insecticidal and biochemical effects of Melaleuca alternifolia
The authors declare that they have no known competing financial essential oil on the Helicoverpa armigera, J. Appl. Entomol. 141 (2017) 721–728,
https://doi.org/10.1111/jen.12397.
interests or personal relationships that could have appeared to influence [18] A. Oliva, S. Costantini, M. De Angelis, S. Garzoli, M. Božović, M. Mascellino,
the work reported in this paper. V. Vullo, R. Ragno, High potency of Melaleuca alternifolia essential oil against

13
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

multi-drug resistant gram-negative bacteria and methicillin-resistant [40] R. Mogana, A. Adhikari, M.N. Tzar, R. Ramliza, C. Wiart, Antibacterial activities of
Staphylococcus aureus, Molecules 23 (2018) 2584, https://doi.org/10.3390/ the extracts, fractions and isolated compounds from Canarium patentinervium Miq.
molecules23102584. against bacterial clinical isolates, BMC Complement. Med. Ther. 20 (2020) 55,
[19] S.D. Cox, C.M. Mann, J.L. Markham, H.C. Bell, J.E. Gustafson, J.R. Warmington, S. https://doi.org/10.1186/s12906-020-2837-5.
G. Wyllie, The mode of antimicrobial action of the essential oil of Melaleuca [41] N. Puvača, I. Čabarkapa, A. Petrović, V. Bursić, R. Prodanović, D. Soleša, J. Lević,
alternifolia (tea tree oil), J. Appl. Microbiol. 88 (2001) 170–175, https://doi.org/ Tea tree (Melaleuca alternifolia) and its essential oil: antimicrobial, antioxidant and
10.1046/j.1365-2672.2000.00943.x. acaricidal effects in poultry production, Worlds Poult. Sci. J. 75 (2019) 235–246,
[20] X. Zhang, Y. Guo, L. Guo, H. Jiang, Q. Ji, In vitro evaluation of antioxidant and https://doi.org/10.1017/S0043933919000229.
antimicrobial activities of Melaleuca alternifolia essential oil, Biomed. Res. Int. 2018 [42] L.T. Correa, M.A. Nicoletti, C.S. De Amorim, A.R. Da Costa, L.A.B. Leoni, J.W.
(2018) 1–8, https://doi.org/10.1155/2018/2396109. P. Munõz, A.R. Fukushima, Atividade antimicrobiana do óleo essencial de
[21] B. Vörös-Horváth, S. Das, A. Salem, S. Nagy, A. Böszörményi, T. Kőszegi, S. Pál, Melaleuca e sua incorporação em um creme mucocutâneo, Rev. Fitos 14 (2020)
A. Széchenyi, Formulation of tioconazole and Melaleuca alternifolia essential oil 26–37, https://doi.org/10.32712/2446-4775.2020.818.
Pickering emulsions for onychomycosis topical treatment, Molecules 25 (2020) [43] M. D'Arrigo, G. Ginestra, G. Mandalari, P.M. Furneri, G. Bisignano, Synergism and
5544, https://doi.org/10.3390/molecules25235544. postantibiotic effect of tobramycin and Melaleuca alternifolia (tea tree) oil against
[22] G. da S. Ferreira, D.J. da Silva, L. Zanata, A.G. Souza, R.R. Ferreira, D.S. Rosa, Staphylococcus aureus and Escherichia coli, Phytomedicine 17 (2010) 317–322,
Antimicrobial cotton wipes functionalized with Melaleuca alternifolia Pickering https://doi.org/10.1016/j.phymed.2009.07.008.
emulsions stabilized with cellulose nanofibrils, Carbohydr. Polym. Technol. Appl. 3 [44] C. Rodríguez-Melcón, C. Alonso-Calleja, C. García-Fernández, J. Carballo,
(2022), 100208, https://doi.org/10.1016/j.carpta.2022.100208. R. Capita, Minimum inhibitory concentration (MIC) and minimum bactericidal
[23] M.H. Monteiro, H. de Macedo, A. da Silva Junior, F. Paumgartten, Óleos essenciais concentration (MBC) for twelve antimicrobials (biocides and antibiotics) in eight
terapêuticos obtidos de espécies de Melaleuca L. (Myrtaceae Juss.), Rev. Fitos 8 strains of Listeria monocytogenes, Biology (Basel) 11 (2021) 46, https://doi.org/
(2013) 19–32, https://doi.org/10.32712/2446-4775.2013.191. 10.3390/biology11010046.
[24] A. Masyita, R. Mustika Sari, A. Dwi Astuti, B. Yasir, N. Rahma Rumata, T. Bin [45] M.C. Hall, E. Martelli, C. Meneguzzi, M. Zanetti, L.L. Silva, M.A. Fiori, F.R. da S.
Emran, F. Nainu, J. Simal-Gandara, Terpenes and terpenoids as main bioactive Machado Junior, J.M.M. Mello, F. Dalcanton, Avaliação da atividade
compounds of essential oils, their roles in human health and potential application antimicrobiana dos óleos essenciais Nerol e Melaleuca puros e microencapsulados,
as natural food preservatives, Food Chem. X. 13 (2022), 100217, https://doi.org/ Braz. J. Heal. Rev. 3 (2020) 5331–5345, https://doi.org/10.34119/bjhrv3n3-097.
10.1016/j.fochx.2022.100217. [46] C.F. Carson, B.J. Mee, T.V. Riley, Mechanism of action of Melaleuca alternifolia (tea
[25] F. Fotsing Yannick Stephane, B. Kezetas Jean Jules, Terpenoids as important tree) oil on Staphylococcus aureus determined by time-kill, lysis, leakage, and salt
bioactive constituents of essential oils, in: Essent. Oils - Bioact. Compd. New tolerance assays and electron microscopy, Antimicrob. Agents Chemother. 46
Perspect. Appl, IntechOpen, 2020, p. 13, https://doi.org/10.5772/ (2002) 1914–1920, https://doi.org/10.1128/AAC.46.6.1914-1920.2002.
intechopen.91426. [47] G.J. Tortora, B.R. Funke, C.L. Case, Microbiologia, 12th ed., Artmed, Porto Alegre,
[26] D. Gallart-Mateu, C.D. Largo-Arango, T. Larkman, S. Garrigues, M. de la Guardia, 2017.
Fast authentication of tea tree oil through spectroscopy, Talanta 189 (2018) [48] T.J. Silhavy, D. Kahne, S. Walker, The bacterial cell envelope, Cold Spring Harb.
404–410, https://doi.org/10.1016/j.talanta.2018.07.023. Perspect. Biol. 2 (2010) a000414, https://doi.org/10.1101/cshperspect.a000414.
[27] Y. Montero, A.G. Souza, É.R. Oliveira, D. dos S. Rosa, Nanocellulose functionalized [49] A. Mai-Prochnow, M. Clauson, J. Hong, A.B. Murphy, Gram positive and Gram
with cinnamon essential oil: a potential application in active biodegradable negative bacteria differ in their sensitivity to cold plasma, Sci. Rep. 6 (2016)
packaging for strawberry, Sustain. Mater. Technol. 29 (2021), e00289, https://doi. 38610, https://doi.org/10.1038/srep38610.
org/10.1016/j.susmat.2021.e00289. [50] W. Levinson, Review of Medical Microbiology and Immunology, 13th ed.,
[28] S. Tankeu, I. Vermaak, G. Kamatou, A. Viljoen, Vibrational spectroscopy as a rapid McGraw-Hill Education, New York, USA, 2014.
quality control method for Melaleuca alternifolia cheel (tea tree oil), Phytochem. [51] K.Y. Rhee, D.F. Gardiner, Clinical relevance of bacteriostatic versus bactericidal
Anal. 25 (2014) 81–88, https://doi.org/10.1002/pca.2470. activity in the treatment of gram-positive bacterial infections, Clin. Infect. Dis. 39
[29] G. Rytwo, R. Zakai, B. Wicklein, The use of ATR-FTIR spectroscopy for (2004) 755–756, https://doi.org/10.1086/422881.
quantification of adsorbed compounds, J. Spectrosc. 2015 (2015) 1–8, https://doi. [52] D.J. da Silva, A.G. Souza, G. da S. Ferreira, A. Duran, A.D. Cabral, F.L.A. Fonseca,
org/10.1155/2015/727595. R.F. Bueno, D.S. Rosa, Cotton fabrics decorated with antimicrobial Ag-coated TiO2
[30] P.K. Verma, A. Kundu, M.S. Puretz, C. Dhoonmoon, O.S. Chegwidden, C. nanoparticles are unable to fully and rapidly eradicate SARS-CoV-2, ACS Appl.
H. Londergan, M. Cho, The bend+libration combination band is an intrinsic, Nano Mater. 4 (2021) 12949–12956, https://doi.org/10.1021/acsanm.1c03492.
collective, and strongly solute-dependent reporter on the hydrogen bonding [53] B. Nadjib, Effective antiviral activity of essential oils and their characteristic
network of liquid water, J. Phys. Chem. B 122 (2018) 2587–2599, https://doi.org/ terpenes against coronaviruses: an update, J. Pharmacol. Clin. Toxicol. 8 (2020)
10.1021/acs.jpcb.7b09641. 1138.
[31] R. Md Salim, J. Asik, M.S. Sarjadi, Chemical functional groups of extractives, [54] A.R. Wani, K. Yadav, A. Khursheed, M.A. Rather, An updated and comprehensive
cellulose and lignin extracted from native Leucaena leucocephala bark, Wood Sci. review of the antiviral potential of essential oils and their chemical constituents
Technol. 55 (2021) 295–313, https://doi.org/10.1007/s00226-020-01258-2. with special focus on their mechanism of action against various influenza and
[32] K. Bazaka, M.V. Jacob, V.K. Truong, F. Wang, W.A.A. Pushpamali, J.Y. Wang, A. coronaviruses, Microb. Pathog. 152 (2021), 104620, https://doi.org/10.1016/j.
V. Ellis, C.C. Berndt, R.J. Crawford, E.P. Ivanova, Plasma-enhanced synthesis of micpath.2020.104620.
bioactive polymeric coatings from monoterpene alcohols: a combined [55] A. Garozzo, R. Timpanaro, A. Stivala, G. Bisignano, A. Castro, Activity of Melaleuca
experimental and theoretical study, Biomacromolecules 11 (2010) 2016–2026, alternifolia (tea tree) oil on influenza virus A/PR/8: study on the mechanism of
https://doi.org/10.1021/bm100369n. action, Antivir. Res. 89 (2011) 83–88, https://doi.org/10.1016/j.
[33] F. Gloerfelt-Tarp, J.C. Mieog, M. Bigland, S. Wheeler, W.M. Palmer, antiviral.2010.11.010.
T. Kretzschmar, Predicting tea tree oil distillate composition using portable [56] L. Ma, L. Yao, Antiviral effects of plant-derived essential oils and their components:
spectrometric technology, J. Raman Spectrosc. 53 (2022) 771–784, https://doi. an updated review, Molecules 25 (2020) 2627, https://doi.org/10.3390/
org/10.1002/jrs.6308. molecules25112627.
[34] M. Baranska, H. Schulz, H. Kruger, R. Quilitzsch, Chemotaxonomy of aromatic [57] M. Asif, M. Saleem, M. Saadullah, H.S. Yaseen, R. Al Zarzour, COVID-19 and
plants of the genus Origanum via vibrational spectroscopy, Anal. Bioanal. Chem. therapy with essential oils having antiviral, anti-inflammatory, and
381 (2005) 1241–1247, https://doi.org/10.1007/s00216-004-3018-y. immunomodulatory properties, Inflammopharmacology 28 (2020) 1153–1161,
[35] L. Cabrales, N. Abidi, F. Manciu, Characterization of developing cotton fibers by https://doi.org/10.1007/s10787-020-00744-0.
confocal Raman microscopy, Fibers 2 (2014) 285–294, https://doi.org/10.3390/ [58] G.J. Tortora, B.R. Funke, C.L. Case. Microbiologia, 12th ed., Artmed, Porto Alegre,
fib2040285. 2017.
[36] H. Almlöf Ambjörnsson, K. Schenzel, U. Germgård, Carboxymethyl cellulose [59] T.J. Silhavy, D. Kahne, S. Walker, The bacterial cell envelope, Cold Spring Harb.
produced at different mercerization conditions and characterized by NIR FT Raman Perspect. Biol. 2 (2010), a000414, https://doi.org/10.1101/cshperspect.a000414.
spectroscopy in combination with multivariate analytical methods, BioResources 8 [60] A. Mai-Prochnow, M. Clauson, J. Hong, A.B. Murphy, Gram positive and Gram
(2013) 1918–1932, https://doi.org/10.15376/biores.8.2.1918-1932. negative bacteria differ in their sensitivity to cold plasma, Sci. Rep. 6 (2016),
[37] S. Burikov, T. Dolenko, S. Patsaeva, Y. Starokurov, V. Yuzhakov, Raman and IR 38610, https://doi.org/10.1038/srep38610.
spectroscopy research on hydrogen bonding in water-ethanol systems, Mol. Phys. [61] W. Levinson. Review of Medical Microbiology and Immunology, 13th ed.,
108 (2010) 2427–2436, https://doi.org/10.1080/00268976.2010.516277. McGraw-Hill Education, New York, USA, 2014.
[38] J.H. Wiley, R.H. Atalla, Band assignments in the raman spectra of celluloses, [62] K.Y. Rhee, D.F. Gardiner, Clinical relevance of bacteriostatic versus bactericidal
Carbohydr. Res. 160 (1987) 113–129, https://doi.org/10.1016/0008-6215(87) activity in the treatment of gram-positive bacterial infections, Clin. Infect. Dis. 39
80306-3. (2004) 755–756, https://doi.org/10.1086/422881.
[39] J. Xi, W. Du, L. Zhong, Probing the interaction between cellulose and cellulase with [63] D.J. da Silva, A.G. Souza, G. da, S. Ferreira, A. Duran, A.D. Cabral, F.L.A. Fonseca,
a nanomechanical sensor, in: Cellul. - Medical, Pharm. Electron. Appl., InTech, R.F. Bueno, D.S. Rosa, Cotton fabrics decorated with antimicrobial Ag-Coated TiO2
2013, pp. 125–140, https://doi.org/10.5772/50285. nanoparticles are unable to fully and rapidly eradicate SARS-CoV-2, ACS Appl.
Nano Mater. 4 (2021) 12949–12956, https://doi.org/10.1021/acsanm.1c03492.

14
G. da S. Ferreira et al. International Journal of Biological Macromolecules 243 (2023) 125228

[64] B. Nadjib, Effective antiviral activity of essential oils and their characteristic [67] L. Ma, L. Yao, Antiviral Effects of Plant-Derived Essential Oils and Their
terpenes against coronaviruses: An update, J. Pharmacol. Clin. Toxicol. 8 (2020) Components: An Updated Review, Molecules 25 (2020) 2627, https://doi.org/
1138. 10.3390/molecules25112627.
[65] A.R. Wani, K. Yadav, A. Khursheed, M.A. Rather, An updated and comprehensive [68] M. Asif, M. Saleem, M. Saadullah, H.S. Yaseen, R. Al Zarzour, COVID-19 and
review of the antiviral potential of essential oils and their chemical constituents therapy with essential oils having antiviral, anti-inflammatory, and
with special focus on their mechanism of action against various influenza and immunomodulatory properties, Inflammopharmacology 28 (2020) 1153–1161,
coronaviruses, Microb. Pathog. 152 (2021), 104620, https://doi.org/10.1016/j. https://doi.org/10.1007/s10787-020-00744-0.
micpath.2020.104620. [69] D.J. da Silva, A. Duran, A.D. Cabral, F.L.A. Fonseca, R.F. Bueno, S.H. Wang, D.
[66] A. Garozzo, R. Timpanaro, A. Stivala, G. Bisignano, A. Castro, Activity of Melaleuca S. Rosa, Delta SARS-CoV-2 inactivation and bactericidal performance of cotton
alternifolia (tea tree) oil on Influenza virus A/PR/8: Study on the mechanism of wipes decorated with TiO2/Ag nanoparticles like Brazilian heavy-fruited Myrciaria
action, Antiviral Res. 89 (2011) 83–88, https://doi.org/10.1016/j. cauliflora, Mater. Today Commun. 33 (2022), 104288, https://doi.org/10.1016/j.
antiviral.2010.11.010. mtcomm.2022.104288.

15

You might also like