Available via license: CC BY-NC-ND 4.0
Content may be subject to copyright.
Exploring the Capabilities of Nanosized Graphene Oxide as a
Pesticide Nanosorbent: Simulation Studies
Prin Tadawattana, Kyohei Kawashima, Sirin Sittiwanichai, Jiraroj T-Thienprasert, Toshifumi Mori,*
and Prapasiri Pongprayoon*
Cite This: ACS Omega 2025, 10, 8951−8959
Read Online
ACCESS Metrics & More Article Recommendations *
sı Supporting Information
ABSTRACT: The pesticide contamination in the environment
has become a global concern. So far, pesticide adsorption from
waste solution is one of the most economic strategies for pesticide
removal. Carbon-based nanomaterials were reported to be
potential pesticide sorbents. To date, nanosized graphene oxide
(GO) has been discovered. Its nanosize, which is comparable to
pesticide sizes, is attractive enough to explore its performance to be
the pesticide sorbent. Thus, herein, the adsorption mechanisms of
a single pesticide on GO were studied by comparing 6-pesticide
systems. Three types of common pesticides (cyfluthrin (CFT)
(pyrethroid), ivermectin (IVM) (avermectin), and diazinon (DZ)
(organophosphate)) were used as pesticide models. All pesticides
rapidly adhere to GO at the graphene-like region. The π−πand π−
alkyl interactions contribute most to pesticide adhesion. The adsorption of CFT and DZ is led by the π−πstacking, whereas bulky
IVM uses the π−alkyl forces. Having more pesticides results in self-clustering. Pesticides pile up and avoid lying on the oxygenated
area. IVM is the most favorable for GO and shows tight self-packing via dispersion force and hydrogen bonding. Overall, this work
displays the encouraging ability of nanosized GO to eectively absorb all pesticides which will benefit future applications in pest
control.
1. INTRODUCTION
Pesticides are substances that are produced by natural or
synthetic processes. These are used as one of key ingredients in
agricultural developments for shielding agricultural products
from pests, weeds, and bacterial and fungal diseases.
1,2
Pesticides are divided into inorganic and organic classes.
Previously, diverse inorganic compounds were used to
formulate inorganic pesticides including metals (antimony,
arsenic, lead, copper, cadmium, and mercury) and metal
oxides,
3
but the presence of toxic heavy metals causes high
health risks. Organic pesticides were reported to be less toxic
to living organisms and more degradable than some inorganic
substances.
3
Thus, organic pesticides have largely replaced
these inorganic chemicals. Organic pesticides that are currently
in use are organophosphates, organochlorines, pyrethroids, and
carbamates.
4
Although only small amounts of organic
pesticides are required for pest control, such amounts still
contaminate the environment. Farmers normally use pesticides
to achieve high agricultural productivity, but they are unaware
that using excess pesticides leads to more poisonous
contamination unfit for consumption. This pesticide contam-
ination can lead to severe conditions including cancers,
reproductive harm, genetic changes, neurological toxicity, and
endocrine disruption.
4−7
The excess pesticides do not wear o
and cause hazard to living entities, especially humans.
4
Owing
to the high health risks to humans, many attempts have been
made to monitor and detect pesticide content in agricultural
products.
Recently, many attempts have been made to not only
formulate less toxic and eco-friendly pesticides
8−10
but also
collect and quantify pesticides. Biodegradation strategies are
also one of eco-friendly techniques.
11
Carbon-based nanoma-
terials such as graphene oxide (GO) have been reported to be
recyclable substances to remove pesticides from wastewater
due to its large surface area, high eectiveness in adsorption,
and high stability in high pressure and temperature.
9,10,12−15
In
addition, GO can be functionalized to target specific
contaminants,
16
and its abundance shows the potential to
provide cost-eective sorbents.
17
GO is also known to show
the excellent photocatalytic properties which can promote the
pesticide degradation in water and soil.
18
Therefore, GO was
Received: July 1, 2024
Revised: February 15, 2025
Accepted: February 20, 2025
Published: February 25, 2025
Article
http://pubs.acs.org/journal/acsodf
© 2025 The Authors. Published by
American Chemical Society 8951
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
This article is licensed under CC-BY-NC-ND 4.0
reported to be one of tools for environmental remediation
technologies.
19
Conventional spectroscopic techniques such as
gas chromatography−mass spectrometry (GC−MS) and liquid
chromatography−mass spectrometry (LC−MS) are also used
to quantify pesticide levels, but such methods require
sophisticated instruments and professional skills that are not
practical under field conditions. Therefore, many studies have
focused on quick and sensitive portable sensors.
20−22
GO was
also involved in the pesticide sensor design and develop-
ment.
23−25
Besides, GO was reported to be used as sorbents
because they are recyclable substances and show high
eectiveness for pesticide adsorption.
1,26
Earlier, the mecha-
nism of pesticide accumulation on GO was revealed.
26,27
It was
reported that the assembly of pesticides onto GO requires π−π
stacking as a major driving force.
9,26,27
In addition, almost all
studies focused on microsized graphene and GO as a pesticide
sorbent. Recently, nanosized GO has been discovered to
display high biocompatibility and low cytotoxicity.
28−30
It was
found that the microsized GO was toxic and even caused death
to mice, whereas nanosized GO displayed no clear influence on
mice.
31
A previous study also reported that a nanosized GO
hardly influences human hematopoietic stem cells after 36 h
incubation.
32
Importantly, its nanosize which is comparable to
pesticide size is attractive enough to explore its performance to
adsorb small pesticides for future design of pesticide removal
or carrier.
Thus, herein, the capabilities of nanosized GO to adsorb
pesticides in solution are investigated. To obtain the
adsorption mechanism on the microscopic level, molecular
dynamics (MD) simulations were employed. MD simulations
have been successfully used to reveal the binding of small
molecules on GO including pesticides.
26,27,33,34
Here, we
specifically examine the ability of nanosized GO to adhere
dierent types of organic pesticides. Commonly used organic
pesticides from dierent families (cyfluthrin (CFT), ivermectin
(IVM), and diazinon (DZ) belonging to pyrethroid,
avermectin, and organophosphate families, respectively) are
used as pesticide models (Figure 1a). Such pesticides are the
main contact pesticides that have a direct eect on human and
animal.
35−39
A nanosized GO with 25% oxygen contents was
employed as a nanosized GO model (Figure 1a). This GO
structure was built based on the Lerf−Klinowski model where
the most stable alignment with 2:1 ratio of hydroxyl (−OH)
and epoxy (−O−) groups was used.
40
The adsorption
mechanism of a single pesticide on GO was studied in
comparison to a 6-pesticide system in order to study the eect
of pesticide concentration on GO adhesion (Figure 1b). Six
pesticides in each system are sucient to surround a GO sheet
in all dimensions to mimic the random pesticide orientations
in solution. Since GO has been used as a pesticide carrier,
9,15
the adsorption insights obtained here will be beneficial for
designing novel strategies for pesticide removal and carrier.
2. MATERIALS AND METHODS
2.1. Preparation of Nanosized Graphene Oxide and
Pesticide Topologies. The structure of nanosized graphene
oxide (GO) was constructed by HierGO
41
generated based on
the Lerf−Klinowski model. GO with 25% oxygen contents
(GO25) was constructed based on a 2:1 ratio of hydroxyl
(−OH) and epoxy (−O−) groups, where this ratio was
reported to be the most stable alignment.
40
GO has a
dimension of 1.5 ×1.3 nm2(72 carbon atoms) as seen in
Figure 1a. This nanodimension size is comparable to the
pesticides studied here. Geometry optimization and normal-
mode analysis for GO25 were performed by density functional
theory (DFT) at the B3LYP/6-31G(d) level using the
Gaussian16 package.
42
Then, the electrostatic potential
(ESP) was calculated at the HF/6-31G(d) level. All atomic
RESP charges converted from ESP were used to construct
topologies using Antechamber from AmberTool20.
43
The
three-dimensional structures of pesticides which are cyfluthrin
(CFT), diazinon (DZ), and ivermectin (IVM), were built
using Discovery Studio 2021
44
and ACPYPE.
45
AMBER14SB
was used to construct all of the pesticide topologies.
In this work, six systems were set. Three systems contain one
of each pesticide and GO (GO-1CFT, GO-1DZ, and GO-
1IVM) and another three systems contain six pesticides with
one GO sheet localized at the center of the simulation box
(GO-6CFT, GO-1DZ, and GO-6IVM) to mimic high
pesticide concentration. Six molecules of pesticides represent
Figure 1. (a) Chemical structures of nanosized graphene oxide with 25% oxygen contents (GO), cyfluthrin (CFT), diazinon (DZ), and ivermectin
(IVM). (b) Orientations of pesticides in 1-pesticide and 6-pesticide systems. Pesticides were randomly placed at least 1 nm away from GO. Each
bead (P1−P6) represents each pesticide.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8952
the presence of pesticides in all directions of GO. In all
systems, a pesticide was placed at least 1 nm away from GO in
all directions in a simulation box of 6 ×6×6 nm3. The
settings can be seen in Figure 1b. Each system was then soaked
in TIP3P water molecules. For each system, 1000 steps of
energy minimization were run to remove bad contacts with the
steepest descent algorithm (the details of the condition used
are displayed in the “Simulation Protocols” section). The
equilibration run was conducted for 10 ns, followed by the 1 μs
production run.
2.2. Simulation Protocols. All molecular dynamics (MD)
simulations were performed using the GROMACS2022.7
software package.
46
The electrostatics were treated with
Particle mesh Ewald (PME)
47
at 1 nm radius cuto, a Fourier
spacing of 0.12 nm, and fourth-order spline interpolation. The
LINCS method
48
was applied to constrain bond lengths within
each system. Periodic boundary conditions were applied in all
directions. All simulations were performed under constant
particle number, pressure, and temperature (NPT) conditions.
The Berendsen thermostat
49
was used to maintain the
temperature at 300 K with a coupling constant τt= 0.1 ps.
The pressure was coupled using the Parrinello−Rahman
algorithm at 1 bar with a coupling constant of τp= 1 ps.
The time step for integration was set to 2 fs, and coordinates
were saved every 2 ps.
All graphical images were made by Discovery Studio 2021
44
and visual molecular dynamics (VMD).
50
The hydrogen bond
counts were operated based on the hydrogen-donor−acceptor
cuto angle and radius of 30°and 0.35 nm, respectively. The
binding energies of GO pesticides and self-interaction energies
of pesticides were calculated using the Poisson−Boltzmann
solver (MM/PBSA) using “gmx mmpbsa” option in
GROMACS.
51
3. RESULTS AND DISCUSSION
Figure 2 shows the final structures after 1 μs simulations for all
systems. It is clear that the pesticide adsorption onto nanosized
GO is spontaneous in all cases. All pesticides were fully
absorbed onto a GO surface (Figure 2). Among the three, IVM
and DZ are the bulkiest and smallest molecules, respectively.
Thus, a bulky IVM in GO-1IVM is partly exposed to the
aqueous solution, whereas GO-1CFT and GO-1DZ seem to
align the whole structure on a GO surface (Figure 2a). More
detailed analyses will be discussed later in the text. When more
pesticide molecules are added, they can fully physisorb on GO
where GO-6IVM produces the largest cluster due to its large
size (Figure 2b). Most previous theoretical studies explored
the adsorption of one pesticide on microsized graphene oxide
where such large surface area results in the “laying-flat”
conformation of the pesticide.
27
However, this lying-flat
conformation becomes the key bottleneck for eective
pesticide desorption. Due to the smaller GO structure in the
current study, no “laying-flat” conformation is captured in all
simulations. This will allow some pesticides to desorb and be
detected by probes. Moreover, in 6-pesticide systems, all
pesticides can aggregate onto GO and self-assemble at the
same time (Figure 2). GO and all pesticides appear to form a
stable nanocomposite throughout the course of the simu-
lations. Seemingly, this finding suggests that nanosized GO can
act as a good pesticide nanocarrier.
To understand the adsorption mechanism, the number of
contacts and hydrogen bonds between GO and pesticides is
computed in Table 1. In 1-pesticide systems, the larger IVM
molecule shows more GO−pesticide contacts (∼103 contacts)
than DZ and CFT (∼44 contacts for CFT and ∼57 contacts
for DZ). This implies the tight binding of IVM. Furthermore,
bulky IVM seems to be more water-accessible (∼11 hydrogen
bonds with water) than DZ and CFT (∼2 hydrogen bonds
with water). Nonetheless, each type of pesticides forms only
∼1 hydrogen bond with GO. This indicates the minute role of
electrostatic interactions on pesticide adsorption. When
increasing the pesticide concentration (6-pesticide systems),
each pesticide appears to severely lose its contacts to GO due
to self-aggregation (Table 1). Compared to the 1-pesticide
system, almost half of the GO−pesticide contacts in all systems
are lost during the aggregation, but such loss is compensated
by the self-assembly (Table 1). Nonetheless, all pesticides still
maintain a hydrogen bond with the GO (Table 1). CFT and
IVM lose ∼20−30% of water contacts during the aggregation
Figure 2. Final snapshots of 1-pesticide systems (GO-1CFT, GO-1DZ, and GO-1IVM) (a) and 6-pesticide systems (GO-6CFT, GO-6DZ, and
GO-6IVM) (b).
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8953
in GO_6CFT and GO_6IVM, whereas enhancing DZ
concentration does not disrupt DZ−water contacts. DZ in
both GO_1DZ and GO_6DZ remains surrounded by water
via ∼160 contacts (Table 1). Also, it is expected that GO
becomes less water-exposed in 6-pesticide cases (∼100 GO−
water contacts are lost) (Table 1).
To further investigate how each pesticide interacts with GO,
the number of contacts between GO and main functional
groups of each pesticide (cyclopropane-carboxylate ester for
CFT, organophosphorus for DZ, lactone, and a group of
disaccharides, benzofuran, and spiroketal for IVM) is
computed in Figure 3. Compared to the total contacts in
Table 1, the number of DZ−GO and IVM−GO contacts
(parent group−GO contacts, where parent groups are
phosphate ester of DZ and lactone of IVM (red and cyan
labels in Figure 3b)) in Figure 3a is smaller, indicating that
both do not require their parent groups to interact with GO
(Figure 3). Rather, both DZ and IVM employ the rest of the
structure to adhere to GO (Figure 3). For CFT, its carboxylate
ester appears to play a role in the adsorption where 30% of the
total CFT−GO contacts are obtained from its functional group
(cyclopropane-carboxylate ester; Table 1 and Figure 3a).
Moreover, the hydrogen bond found in Figure 3a for DZ and
CFT demonstrates that the total number of GO−pesticide
hydrogen bonds found in Table 1 is rooted from a parent
group of DZ and CFT. Unlike others, IVM seems to employ
the side chain moieties to hydrogen bond to GO (Figure 3).
The results obtained demonstrate that each type of pesticide
employs nonidentical mechanisms to interact with GO. The
hydrophobic interactions were reported to contribute most to
GO−pesticide adsorption.
27,33,52−54
The pesticides studied
here are consistent with this trend since only one hydrogen
bond is captured (Table 1 and Figure 3a) while a large number
of contacts are found. This highlights the role of hydrophobic
interaction for adhesion. In addition, we further decompose
the adsorption mechanism of each type of existing pesticides in
detail later in the text.
To better understand the eect of chemical properties of
each pesticide on the adsorption ability, the binding energies
between each component (GO−pesticide and pesticide−
pesticide) are computed in Table 2. The high Vdw energies
observed in all cases confirm that the hydrophobic interactions
contribute most to pesticide adhesion. These binding energies
Table 1. Number of Contacts and Hydrogen Bonds between
Individual Pesticide (Cyfluthrin (CFT), Diazinon (DZ), and
Ivermectin (IVM)) with Each Component (GO, Water, and
Pesticide) in the Systems
a
system pair no. of contacts no. of hydrogen bond
GO-1CFT GO−CFT 44.13 ±16.00 1.13 ±0.23
GO−water 600.82 ±43.65 26.66 ±2.57
CFT−water 150.88 ±33.59 2.19 ±1.07
GO-1DZ GO−DZ 57.82 ±18.10 1.04 ±0.48
GO−water 610.78 ±41.99 26.35 ±2.58
DZ−water 161.67 ±30.58 1.89 ±0.97
GO-1IVM GO−IVM 103.98 ±18.30 1.00 ±0.45
GO−water 571.51 ±37.82 25.79 ±2.55
IVM-water 512.69 ±47.90 11.11 ±1.85
GO-6CFT GO−CFT 26.48 ±3.17 1.03 ±0.03
CFT−CFT 16.81 ±0.81
GO−water 481.84 ±50.95 24.13 ±2.70
CFT−water 103.54 ±33.85 1.83 ±0.97
GO-6DZ GO−DZ 36.39 ±3.63 1.07 ±0.06
DZ−DZ 21.06 ±1.73
GO−water 477.86 ±47.71 23.18 ±2.65
DZ−water 160.96 ±43.12 2.17 ±1.17
GO-6IVM GO−IVM 52.53 ±29.09 1.19 ±0.67
IVM−IVM 38.11 ±32.66 1.00 ±0.45
GO−water 409.90 ±47.48 21.07 ±2.62
IVM−water 416.48 ±60.25 6.32 ±1.32
a
The contacts are counted when all pairs between two chosen
components are in the cut-o distance of 0.35 nm.
Figure 3. (a) Number of pesticide−GO contacts and hydrogen bonds in all systems. Only the contacts and hydrogen bonds between key chemical
families of each pesticide and GO are computed. The chemical families of each pesticide are shown in red (cyclopropane-carboxylate ester for CFT,
organophosphorus for DZ, a group of disaccharides, benzofuran, and spiroketal for IVM) and cyan (lactone) in (b).
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8954
agree well with previous studies.
27,53,55,56
Among all, IVM
shows better binding anity (binding energies of −161.34 kJ/
mol in GO_1IVM and −77.00 kJ/mol in GO_6IVM) to GO
than to DZ and CFT in all cases. DZ and CFT display a
comparable degree of binding anities in 1-pesticide systems,
but DZ in GO-6DZ becomes more preferable for GO than
CFT in 6-pesticide systems (GO-6CFT) (Table 2). Seemingly,
the binding anities of all pesticides on GO are severely
reduced under a high pesticide concentration (Table 2). A
high number of pesticides induce the self-clustering with
dierent degrees of compactness. Bulky IVM molecules form a
cluster (IVM−IVM binding energy of −31.15 kJ/mol), while
DZs are more dispersed (binding energy of −8.13 kJ/mol)
(Table 2). Although the interactions with GO are the main
attractive forces to trap all pesticides on a GO surface, the self-
clustering interactions also play a role in pesticide adsorption,
especially in GO_6CFT and GO_6IVM (Table 2). DZs in
GO-6DZ can also self-interact but are less favorable than other
pesticides (DZ−DZ binding energy of −8.13 kJ/mol) (Table
2). The standard deviations of the energies are large in the 6-
pesticide systems because of the diusive dynamics of
pesticides. In CFT and DZ, all molecules move diusively
toward GO, where all appear to gain a comparable number of
GO (∼25−32 contacts for CFT and 31−40 contacts for DZ)
and self-contacts (∼15−19 contacts for CFT and ∼19−23
contacts for DZ) (Table 3). Each molecule of CFT and DZ
seems to equally contribute to interact with GO. On the
contrary, dierent bound IVM conformations on GO are
observed (Table 3). IVM5 is fully adhered on one side of the
GO surface at the beginning of a simulation resulting in high
GO contacts (∼105 contacts) (Table 3 and Figure 4b). Unlike
IVM5, IVM3 is literarily packed at the outermost of the IVM
cluster, leading to the least GO contacts (∼22 contacts)
(Table 3 and Figure 4c). Bulky IVMs appear to act dierently
from CFT and DZ. Some IVMs show high self-contacts, while
others display high GO contacts (Table 3). This implies
dierent degrees of GO accessibility. The high IVM−IVM
contacts reflect not only the favorable self-assembly but also
the steric hindrance of its structure that blocks the GO
accessibility. The self-clustering of pesticides can also be seen
in GO-free systems with similar levels of self-binding energies
to GO−pesticide systems (Figure S1 and Table S1 in the
Table 2. Average Binding Energies (kJ/mol) between Each Component (GO and Pesticide) with Standard Deviation in All
Systems Using MMPBSA
a
system VdW elec total
GO-1CFT GO−CFT −94.15 ±30.48 −6.97 ±10.79 −100.52 ±34.28
GO-1DZ GO−DZ −104.44 ±23.88 −12.10 ±5.39 −116.54 ±31.83
GO-1IVM GO−IVM −144.87 ±16.28 −16.48 ±8.37 −161.34 ±18.30
GO-6CFT GO−CFT −34.53 ±37.27 −5.31 ±10.24 −39.84 ±43.80
CFT−CFT −12.54 ±16.69 −1.53 ±3.67 −14.07 ±18.97
GO-6DZ GO−DZ −48.85 ±36.28 −8.67 ±13.62 −57.51 ±43.69
DZ−DZ −6.81 ±11.93 −1.32 ±4.11 −8.13 ±15.02
GO-6IVM GO−IVM −62.48 ±23.76 −14.52 ±10.77 −77.00 ±30.09
IVM−IVM −29.41 ±15.30 −1.73 ±3.51 −31.15 ±17.19
a
VdW and Elec denote the van der Waals and electrostatic energies, respectively, and Total is the sum of VdW and Elec terms.
Table 3. Number of Contacts in 6-Pesticide Systems (GO-6CFT, GO-6DZ, and GO-6IVM)
system CFT1 CFT2 CFT3 CFT4 CFT5 CFT6 water
GO-6CFT GO 24.23 ±17.36 27.13 ±18.86 32.56 ±20.89 24.46 ±17.82 25.79 ±17.80 24.68 ±18.99 481.84 ±50.95
CFT1 15.95 ±12.41 17.51 ±13.76 17.44 ±13.87 16.61 ±12.77 16.24 ±13.02 107.89 ±34.61
CFT2 17.24 ±13.95 15.34 ±12.37 16.44 ±13.10 16.72 ±13.47 104.41 ±34.06
CFT3 16.82 ±13.30 16.89 ±12.86 18.93 ±14.87 95.36 ±33.57
CFT4 16.47 ±13.32 16.56 ±12.95 107.35 ±30.88
CFT5 16.96 ±13.39 105.75 ±34.01
CFT6 100.48 ±36.03
DZ1 DZ2 DZ3 DZ4 DZ5 DZ6 water
GO-6DZ GO 37.80 ±22.04 33.12 ±21.18 40.53 ±24.76 39.30 ±23.84 36.46 ±22.49 31.14 ±19.74 477.86 ±47.71
DZ1 20.81 ±12.18 18.78 ±10.84 21.93 ±12.92 23.30 ±14.58 22.41 ±14.56 154.73 ±40.07
DZ2 19.41 ±11.52 21.65 ±12.70 22.71 ±13.09 19.24 ±11.74 167.46 ±43.50
DZ3 19.16 ±10.80 21.48 ±11.85 23.14 ±13.85 162.88 ±44.74
DZ4 20.06 ±10.80 23.24 ±14.92 155.66 ±43.10
DZ5 18.57 ±10.91 166.00 ±46.30
DZ6 159.06 ±42.05
IVM1 IVM2 IVM3 IVM4 IVM5 IVM6 water
GO-6IVM GO 45.83 ±21.30 32.94 ±17.54 22.88 ±13.66 45.67 ±17.90 105.20 ±15.54 62.67 ±33.09 409.90 ±47.48
IVM1 82.26 ±45.54 63.04 ±33.99 25.93 ±15.85 6.94 ±2.14 18.40 ±13.30 438.29 ±53.17
IVM2 47.83 ±27.24 49.28 ±31.14 8.22 ±3.97 38.72 ±34.83 426.38 ±76.11
IVM3 113.10 ±23.12 0.00 ±0.00 23.64 ±19.98 359.60 ±56.11
IVM4 4.69 ±2.57 69.83 ±18.59 326.16 ±69.27
IVM5 19.81 ±11.32 498.86 ±48.45
IVM6 449.60 ±58.36
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8955
Supporting Information). However, only loose packing of
small CFT and DZ is captured because of the presence of free
CFT and DZ in the bulk (Figure S1 in the Supporting
Information). Having GO appears to eectively collect all
pesticides. This indicates that GO can serve as a promoter for
pesticide aggregation.
In previous studies, not only π−πstacking but also other
hydrophobic forces such as alkyl-πinteractions were reported
to play a role in the GO adsorption.
27,33,52−54
Even though π
interactions cannot be directly captured by molecular
mechanics, such interactions can be evident from the planar
orientations of aromatic moieties toward GO for π−πstacking
Figure 4. Cartoon representations of 1-pesticide (a) and 6-pesticide (b,c) systems where key interactions are shown as insets.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8956
(planar angle ≤20°
54,57
) and constant distance between alkyl
chain and GO for π-alkyl interactions. Recent theoretical
studies have reported that the π−alkyl dispersion forces can
exist within a C−C distance of 0.45 nm.
52,54
In this work, the
π−πand π−alkyl dispersion forces are found to play a vital role
in the pesticide adsorption. High number of contacts between
the graphene-like region and pesticides demonstrate that all
pesticides prefer the graphene-like zone to the polar region
(Figure S2 and Table S2 in the Supporting Information). For
the 1-pesticide system, CFT and DZ are mainly trapped on the
GO surface by planar π−πinteractions in assistance with one
hydrogen bond (Figures 4a and S3 in the Supporting
Information). On the other hand, IVM employs only π−alkyl
forces and a hydrogen bond for GO binding due to the
absence of aromaticity (Figures 4a and S3 in the Supporting
Information). In the case of a hydrogen bond, the single
adsorbate−GO hydrogen bond obtained in CFT and DZ is
formed by their parent groups (carboxylate ester for CFT and
organophosphate for DZ), while IVM utilizes nonlactone
moieties for hydrogen bonding (Figures 3 and 4). The larger
size also allows IVM to form more interactions, resulting in the
best binding energy (Figures 3 and 4).
In the case of 6-pesticide systems, the adsorption on GO
appears to rely on the “first come first served” basis. A first few
molecules start to land on the deoxygenated region and mainly
interact with GO via πinteractions (π−πstacking for CFT and
DZ and π−alkyl interactions for IVM) (Figures 4 and S4 in the
Supporting Information). When the graphene-like zone on GO
is completely occupied, the other pesticides tend to pile up on
top of GO-bound pesticides rather than line up on the
oxygenated area (Figures 1,4c, and S2b in the Supporting
Information). This is why the pesticide adhesion on to GO
lacks hydrogen bonding (only ∼1 GO−adsorbate hydrogen
bond is found for all GO−pesticide systems). The binding
energies display IVM as the most favorable species on GO
(Table 2) because bulky IVMs are packed upright, allowing the
polar moieties to hydrogen bond with adjacent molecules
(Figure S5 in the Supporting Information). This self-packing in
the assistance of a hydrogen bond allows more favorable
IVM−GO clustering. On the contrary, CFT and DZ show no
significant self-hydrogen bond, but their self-aggregation on
GO is facilitated by π−πand π−alkyl interactions (Figures 4c
and S4 in the Supporting Information). This finding can
explain why IVM displays better binding energies than the
other two (Table 2). The force fields used in this work can
capture the key interactions between GO and pesticides as
seen in other quantum calculations and MD simulations with
other generalized force fields.
26,27,58,59
This indicates the
reliability of the force fields used in this work.
4. CONCLUSIONS
Herein, the aggregation mechanisms of commonly used
pesticides (CFT, DZ, and IVM) on nanosized GO are
investigated for the first time. All pesticides are rapidly and
spontaneously adsorbed on GO without desorption through-
out the course of all simulations. Overall, nanosized GO
exhibits promising potential as a pesticide sorbent. All types of
pesticides here prefer a deoxygenated area for adsorption,
while they can also hydrogen bond with an oxygenated area.
The adsorption anities of all of the adsorbates here depend
on the degree of hydrophobicity. GO with high oxygen
contents seems to be inappropriate for eective pesticide
adsorption. Among all of these, IVM appears to act as the best
adsorbate. Although bulky IVM cannot form any π−π
interaction owing to the lack of aromaticity, it can favorably
be stabilized on the GO surface by multiple π−alkyl
interactions and hydrogen bonds. In addition, IVM are also
clustered on GO using both hydrophobic and electrostatic
interactions, whereas the self-packing of CFT and DZ employs
only πinteractions. Thus, more interactions found in the IVM
allow tighter binding. Taken together, both favorable GO-
binding anities and self-clustering allow IVM to become the
best adsorbate.
The results here suggest that nanosized GO can act as a
potential pesticide sorbent. Furthermore, each pesticide
displays dierent degrees of accumulation anities on GO,
suggesting the feasibility of sequential pesticide extraction for
further detection. Recently, microsized GO has been studied as
a pesticide carrier. Such composite was reported to act as a
high-eciency carrier.
9,10
GO also showed the synergistical
eect with pesticides to improve control ecacy and utilization
eciency.
8,60
While this pesticide−GO composite displays the
high fungicide and insecticide activities,
9,10,60
microsized GO
can be toxic to hosts such as crop or livestock. This is because
large-sized GO was reported to be toxic to various species,
while cytotoxicity is reduced in nanosized GO.
61,62
Thus, the
encouraging ability of nanosized GO to eectively adsorb all
pesticides similar to large-sized GO found here suggests the
safer application potential in pest control of nanosized GO.
■ASSOCIATED CONTENT
*
sı Supporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acsomega.4c06036.
Final snapshots of GO-free pesticide systems (6CFT,
6DZ, and 6IVM) at 150 ns; chemical structure of GO
where the graphene-like region; cartoon representatives
of π−πand π−alkyl orientations observed in GO-1CFT,
GO-1DZ, and GO-1IVM; cartoon representatives of
π−πand π−−alkyl orientations observed in GO-6CFT,
GO-6DZ, and GO-6IVM; final snapshots of all 6-
pesticide systems; average binding energies (kJ/mol)
between pesticides with standard deviation in all systems
using MMPBSA; and number of pesticide contacts with
graphene-like and oxygenated regions (PDF)
■AUTHOR INFORMATION
Corresponding Authors
Toshifumi Mori −Institute for Material Chemistry and
Engineering, Kyushu University, Kasuga, Fukuoka 8168580,
Japan; Interdisciplinary Graduate School of Engineering
Science, Kyushu University, Kasuga, Fukuoka 8168580,
Japan; orcid.org/0000-0003-0188-0794; Phone: +66-
2562-5555; Email: toshi_mori@cm.kyushu-u.ac.jp
Prapasiri Pongprayoon −Department of Chemistry, Faculty
of Science, Kasetsart University, Bangkok 10900, Thailand;
Center for Advanced Studies in Nanotechnology for Chemical,
Food and Agricultural Industries, KU Institute for Advanced
Studies, Kasetsart University, Bangkok 10900, Thailand;
orcid.org/0000-0002-1472-8241; Phone: +81-92-583-
7800; Email: fsciprpo@ku.ac.th
Authors
Prin Tadawattana −Department of Chemistry, Faculty of
Science, Kasetsart University, Bangkok 10900, Thailand
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8957
Kyohei Kawashima −Institute for Material Chemistry and
Engineering, Kyushu University, Kasuga, Fukuoka 8168580,
Japan; orcid.org/0000-0002-9335-8145
Sirin Sittiwanichai −Department of Chemistry, Faculty of
Science, Kasetsart University, Bangkok 10900, Thailand
Jiraroj T-Thienprasert −Department of Physics, Faculty of
Science, Kasetsart University, Bangkok 10900, Thailand;
orcid.org/0000-0001-5611-9607
Complete contact information is available at:
https://pubs.acs.org/10.1021/acsomega.4c06036
Notes
The authors declare no competing financial interest.
■ACKNOWLEDGMENTS
We would like to thank Kasetsart University Research and
Development Institute (grant no. FF(KU)51.67), Grant-in-Aid
for Scientific Research (Grant Nos. 22H02035, 23K23303, and
23KK0254) from JSPS, the Oce of the National Economic
and Social Development Council, and the Oce of the Prime
Minister through Kasetsart University under the project
entitled “Driving Research and Development of Cutting-edge
Innovations for ASEAN’s Agricultural Leadership” for financial
support. The calculations were carried out at Kasetsart
University HPC service center (Nontri AI) and the Research
Center for Computational Sciences in Okazaki (Project Nos.
23-IMS-C111 and 24-IMS-C105).
■REFERENCES
(1) Dasriya, V.; Joshi, R.; Ranveer, S.; Dhundale, V.; Kumar, N.;
Raghu, H. V. Rapid detection of pesticide in milk, cereal and cereal
based food and fruit juices using paper strip-based sensor. Sci. Rep.
2021,11 (1), 18855.
(2) Meshram, S.; Bisht, S.; Gogoi, R. Current development,
application and constraints of biopesticides in plant disease manage-
ment. In Biopesticides; Elsevier, 2022; pp 207−224.
(3) Hassan, A. S. Inorganic-based pesticides: a review article. Egypt
Sci. J. Pestic 2019,5, 39−52.
(4) Thorat, T.; Patle, B.; Wakchaure, M.; Parihar, L. Advancements
in Techniques Used for Identification of Pesticide Residue on Crops.
Journal of Natural Pesticide Research 2023,4, 100031.
(5) Damalas, C. A.; Koutroubas, S. D. Farmers’ Exposure to
Pesticides: Toxicity Types and Ways of Prevention. Toxics 2016,4
(1), 1.
(6) Park, B. K.; Kwon, S. H.; Yeom, M. S.; Joo, K. S.; Heo, M. J.
Detection of pesticide residues and risk assessment from the local
fruits and vegetables in Incheon, Korea. Sci. Rep. 2022,12 (1), 9613.
(7) Mostafalou, S.; Abdollahi, M. Pesticides and human chronic
diseases: evidences, mechanisms, and perspectives. Toxicol. Appl.
Pharmacol. 2013,268 (2), 157−177.
(8) Li, X.; Wang, Q.; Wang, X.; Wang, Z. Synergistic Effects of
Graphene Oxide and Pesticides on Fall Armyworm, Spodoptera
frugiperda. Nanomaterials 2022,12 (22), 3985.
(9) Gao, X.; Shi, F.; Peng, F.; Shi, X.; Cheng, C.; Hou, W.; Xie, H.;
Lin, X.; Wang, X. Formulation of nanopesticide with graphene oxide
as the nanocarrier of pyrethroid pesticide and its application in spider
mite control. Rsc Adv. 2021,11 (57), 36089−36097.
(10) Hu, P.; Zhu, L.; Zheng, F.; Lai, J.; Xu, H.; Jia, J. Graphene oxide
as a pesticide carrier for enhancing fungicide activity against
Magnaporthe oryzae. New J. Chem. 2021,45 (5), 2649−2658.
(11) Guerrero Ramírez, J. R.; Ibarra Munoz, L. A.; Balagurusamy,
N.; Frías Ramírez, J. E.; Alfaro Hernández, L.; Carrillo Campos, J.
Microbiology and biochemistry of pesticides biodegradation. Int. J.
Mol. Sci. 2023,24 (21), 15969.
(12) Boruah, P. K.; Sharma, B.; Hussain, N.; Das, M. R. Magnetically
recoverable Fe3O4/graphene nanocomposite towards efficient
removal of triazine pesticides from aqueous solution: Investigation
of the adsorption phenomenon and specific ion effect. Chemosphere
2017,168, 1058−1067.
(13) Song, S.; Wan, M.; Feng, W.; Zhang, J.; Mo, H.; Jiang, X.; Shen,
H.; Shen, J. Graphene oxide as the potential vector of hydrophobic
pesticides: ultrahigh pesticide loading capacity and improved antipest
activity. ACS Agric. Sci. Technol. 2021,1(3), 182−191.
(14) Chen, Z.; Zhao, J.; Liu, Z.; Bai, X.; Li, W.; Guan, Z.; Zhou, M.;
Zhu, H. Graphene-Delivered Insecticides against Cotton Bollworm.
Nanomaterials 2022,12 (16), 2731.
(15) Liu, J.; Luo, Y.; Jiang, X.; Sun, G.; Song, S.; Yang, M.; Shen, J.
Enhanced and sustained pesticidal activity of a graphene-based
pesticide delivery system against the diamondback moth Plutella
xylostella. Pest Manag Sci. 2022,78 (12), 5358−5365.
(16) Jayakaran, P.; Nirmala, G.; Govindarajan, L. Qualitative and
Quantitative Analysis of Graphene-Based Adsorbents in Wastewater
Treatment. Int. J. Chem. Eng. 2019,2019 (1), 1−17.
(17) Anegbe, B.; Ifijen, I. H.; Maliki, M.; Uwidia, I. E.; Aigbodion, A.
I. Graphene oxide synthesis and applications in emerging contaminant
removal: a comprehensive review. Environ. Sci. Eur. 2024,36 (1), 15.
(18) Mohan, V. B.; Lau, K.-t.; Hui, D.; Bhattacharyya, D. Graphene-
based materials and their composites: A review on production,
applications and product limitations. Composites, Part B 2018,142,
200−220.
(19) Singh, R.; Samuel, M. S.; Ravikumar, M.; Ethiraj, S.; Kumar, M.
Graphene materials in pollution trace detection and environmental
improvement. Environ. Res. 2024,243, 117830.
(20) Liu, B.; Zhou, P.; Liu, X.; Sun, X.; Li, H.; Lin, M. Detection of
pesticides in fruits by surface-enhanced Raman spectroscopy coupled
with gold nanostructures. Food Bioprocess Technol. 2013,6, 710−718.
(21) Poudyal, D. C.; Dhamu, V. N.; Samson, M.; Muthukumar, S.;
Prasad, S. Portable Pesticide Electrochem-sensor: A Label-Free
Detection of Glyphosate in Human Urine. Langmuir 2022,38 (5),
1781−1790.
(22) Kilele, J. C.; Chokkareddy, R.; Redhi, G. G. Ultra-sensitive
electrochemical sensor for fenitrothion pesticide residues in fruit
samples using IL@ CoFe2O4NPs@ MWCNTs nanocomposite.
Microchem. J. 2021,164, 106012.
(23) Lang, T.; Xiao, M.; Cen, W. Graphene-Based Metamaterial
Sensor for Pesticide Trace Detection. Biosensors 2023,13 (5), 560.
(24) Loudiki, A.; Azriouil, M.; Matrouf, M.; Laghrib, F.; Farahi, A.;
Saqrane, S.; Bakasse, M.; Lahrich, S.; El Mhammedi, M. Graphene-
based electrode materials used for some pesticide’s detection in food
samples: A review. Inorg. Chem. Commun. 2022,144, 109891.
(25) Xue, R.; Kang, T.-F.; Lu, L.-P.; Cheng, S.-Y. Electrochemical
sensor based on the graphene-nafion matrix for sensitive determi-
nation of organophosphorus pesticides. Anal. Lett. 2013,46 (1), 131−
141.
(26) Kumari, K.; Singh, M. B.; Tomar, N.; Kumar, A.; Kumar, V.;
Dabodhia, K. L.; Singh, P. Adsorption of pesticides using graphene
oxide through computational and experimental approach. J. Mol.
Struct. 2023,1291, 136043.
(27) Wang, H.; Hu, B.; Gao, Z.; Zhang, F.; Wang, J. Emerging role
of graphene oxide as sorbent for pesticides adsorption: Experimental
observations analyzed by molecular modeling. J. Mater. Sci. 2021,63,
192−202.
(28) Henna, T. K.; Pramod, K. Graphene quantum dots redefine
nanobiomedicine. Mater. Sci. Eng. C 2020,110, 110651.
(29) Singh, R. D.; Shandilya, R.; Bhargava, A.; Kumar, R.; Tiwari, R.;
Chaudhury, K.; Srivastava, R. K.; Goryacheva, I. Y.; Mishra, P. K.
Quantum Dot Based Nano-Biosensors for Detection of Circulating
Cell Free miRNAs in Lung Carcinogenesis: From Biology to Clinical
Translation. Front. Genet. 2018,9, 616.
(30) Yan, Y.; Gong, J.; Chen, J.; Zeng, Z.; Huang, W.; Pu, K.; Liu, J.;
Chen, P. Recent Advances on Graphene Quantum Dots: From
Chemistry and Physics to Applications. Adv. Mater. 2019,31 (21),
No. e1808283.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8958
(31) Chong, Y.; Ma, Y.; Shen, H.; Tu, X.; Zhou, X.; Xu, J.; Dai, J.;
Fan, S.; Zhang, Z. The in vitro and in vivo toxicity of graphene
quantum dots. Biomaterials 2014,35 (19), 5041−5048.
(32) Fasbender, S.; Zimmermann, L.; Cadeddu, R.-P.; Luysberg, M.;
Moll, B.; Janiak, C.; Heinzel, T.; Haas, R. The low toxicity of
graphene quantum dots is reflected by marginal gene expression
changes of primary human hematopoietic stem cells. Sci. Rep. 2019,9
(1), 12028.
(33) Jia, Q.; Yang, C.; Venton, B. J.; DuBay, K. H. Atomistic
Simulations of Dopamine Diffusion Dynamics on a Pristine Graphene
Surface. ChemPhysChem 2022,23 (4), No. e202100783.
(34) You, X.; He, M.; Cao, X.; Wang, P.; Wang, J.; Li, L. Molecular
dynamics simulations of removal of nonylphenol pollutants by
graphene oxide: Experimental study and modelling. Appl. Surf. Sci.
2019,475, 621−626.
(35) Xie, Y.; Zhao, J.; Li, X.; Sun, J.; Yang, H. Effects of Cyfluthrin
Exposure on Neurobehaviour, Hippocampal Tissue and Synaptic
Plasticity in Wistar Rats. Toxics 2023,11 (12), 999.
(36) Wijewickrema, A.; Banneheke, H.; Pathmeswaran, A.; Refai, F.
W.; Kauranaratne, M.; Malavige, N.; Jeewandara, C.; Ekanayake, M.;
Samaraweera, D.; Thambavita, D.; et al. Efficacy and safety of oral
ivermectin in the treatment of mild to moderate Covid-19 patients: a
multi-centre double-blind randomized controlled clinical trial. BMC
Infectious Diseases 2024,24 (1), 719.
(37) Wu, X.; Li, J.; Zhou, Z.; Lin, Z.; Pang, S.; Bhatt, P.; Mishra, S.;
Chen, S. Environmental occurrence, toxicity concerns, and degrada-
tion of diazinon using a microbial system. Front. Microbiol. 2021,12,
717286.
(38) Galadima, M.; Singh, S.; Pawar, A.; Khasnabis, S.; Dhanjal, D.
S.; Anil, A. G.; Rai, P.; Ramamurthy, P. C.; Singh, J. Toxicity,
microbial degradation and analytical detection of pyrethroids: A
review. Environ. Adv. 2021,5, 100105.
(39) Yang, Y.; Zhang, X.; Jiang, J.; Han, J.; Li, W.; Li, X.; Yee Leung,
K. M.; Snyder, S. A.; Alvarez, P. J. Which micropollutants in water
environments deserve more attention globally? Environ. Sci. Technol.
2022,56 (1), 13−29.
(40) Boukhvalov, D. W.; Katsnelson, M. I. Modeling of Graphite
Oxide. J. Am. Chem. Soc. 2008,130 (32), 10697−10701.
(41) Garcia, N. A.; Awuah, J. B.; Zhao, C.; Vukovic, F.; Walsh, T. R.
Simulation-ready graphene oxide structures with hierarchical com-
plexity: a modular tiling strategy. 2D Materials 2023,10 (2), 025007.
(42) Gaussian 16; Gaussian, Inc.,: Wallingford CT, 2016 GaussView
5.0. Wallingford, E.U.A., 2016. (accessed).
(43) Amber 2020; University of California: San Francisco, 2020.
(accessed).
(44) Dassault Systemes; Dassault Systemes, San Diego: 2021.
(accessed).
(45) Sousa da Silva, A. W.; Vranken, W. F. ACPYPE - AnteChamber
PYthon Parser interfacE. BMC Research Notes 2012,5(1), 367.
(46) GROMACS 2022 Source code; Zenodo: 2022. (accessed).
(47) Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N·
log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993,
98 (12), 10089−10092.
(48) Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. G. E. M.
LINCS: A linear constraint solver for molecular simulations. J.
Comput. Chem. 1997,18, 1463−1472.
(49) Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling
through velocity rescaling. J. Chem. Phys. 2007,126 (1), 014101.
(50) Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular
dynamics. J. Mol. Graphics 1996,14 (1), 33−38.
(51) Valdés-Tresanco, M. S.; Valdés-Tresanco, M. E.; Valiente, P.
A.; Moreno, E. gmx_MMPBSA: a new tool to perform end-state free
energy calculations with GROMACS. J. Chem. Theory Comput. 2021,
17 (10), 6281−6291.
(52) Moriggi, F.; Barbera, V.; Galimberti, M.; Raffaini, G.
Adsorption Affinities of Small Volatile Organic Molecules on
Graphene Surfaces for Novel Nanofiller Design: A DFT Study.
Molecules 2023,28 (22), 7633.
(53) Subasinghege Don, V.; Kim, L.; David, R.; Nauman, J. A.;
Kumar, R. Adsorption Studies at the Graphene Oxide−Liquid
Interface: A Molecular Dynamics Study. J. Phys. Chem. C 2023,127
(12), 5920−5930.
(54) Alonso, M.; Woller, T.; Martin-Martinez, F. J.; Contreras-
Garcia, J.; Geerlings, P.; De Proft, F. Understanding the fundamental
role of pi/pi, sigma/sigma, and sigma/pi dispersion interactions in
shaping carbon-based materials. Chemistry 2014,20 (17), 4931−
4941.
(55) Bolibok, P.; Koter, S.; Kaczmarek-Kędziera, A.; Kowalczyk, P.;
Łukomska, B.; Łukomska, O.; Boncel, S.; Wisniewski, M.; Kaneko, K.;
Terzyk, A. P. Liquid phase adsorption induced nanosizing of graphene
oxide. Carbon 2021,183, 948−957.
(56) Molla, A.; Li, Y.; Mandal, B.; Kang, S. G.; Hur, S. H.; Chung, J.
S. Selective adsorption of organic dyes on graphene oxide: Theoretical
and experimental analysis. Appl. Surf. Sci. 2019,464, 170−177.
(57) Janiak, C. A critical account on π−πstacking in metal
complexes with aromatic nitrogen-containing ligands. J. Chem. Soc.,
Dalton Trans. 2000, No. 21, 3885−3896.
(58) S. Araujo, W.; Caldeira Rego, C. R.; Guedes-Sobrinho, D.;
Cavalheiro Dias, A.; Rodrigues do Couto, I.; Bordin, J. R.; Ferreira de
Matos, C.; Piotrowski, M. J. Quantum Simulations and Experimental
Insights into Glyphosate Adsorption Using Graphene-Based Nano-
materials. ACS Appl. Mater. Interfaces 2024, 31500.
(59) Zeng, J.; Zhang, Y.; Chen, Y.; Han, Z.; Chen, X.; Peng, Y.;
Chen, L.; Chen, S. Molecular dynamics simulation of the adsorption
properties of graphene oxide/graphene composite for alkali metal
ions. Journal of Molecular Graphics and Modelling 2022,114, 108184.
(60) Wang, X.; Xie, H.; Wang, Z.; He, K. Graphene oxide as a
pesticide delivery vector for enhancing acaricidal activity against
spider mites. Colloids Surf., B 2019,173, 632−638.
(61) Yadav, S.; Singh Raman, A. P.; Meena, H.; Goswami, A. G.;
Bhawna; Kumar, V.; Jain, P.; Kumar, G.; Sagar, M.; Rana, D. K.; et al.
An update on graphene oxide: applications and toxicity. ACS omega
2022,7(40), 35387−35445.
(62) Jiang, T.; Amadei, C. A.; Lin, Y.; Gou, N.; Rahman, S. M.; Lan,
J.; Vecitis, C. D.; Gu, A. Z. Dependence of Graphene Oxide (GO)
Toxicity on Oxidation Level, Elemental Composition, and Size. Int. J.
Mol. Sci. 2021,22 (19), 10578.
ACS Omega http://pubs.acs.org/journal/acsodf Article
https://doi.org/10.1021/acsomega.4c06036
ACS Omega 2025, 10, 8951−8959
8959