Access to this full-text is provided by Wiley.
Content available from Small Science
This content is subject to copyright. Terms and conditions apply.
Instant Upcycling of Microplastics into Graphene and Its
Environmental Application
Muhammad Adeel Zafar and Mohan V. Jacob*
1. Introduction
Plastic pollution has surged to the forefront of concerns for
both the scientific community and environmentalists. What
compounds this issue is the degradation of plastic into smaller
fragments, often reaching micron sizes, driven by a complex
interplay of chemical, physical, and biological processes in
the environment. These minuscule particles, known as
“microplastics”typically consist of solid
polymer particles measuring 1–5 mm.
[1]
Microplastics are notorious for their non-
degradable and insoluble nature in water,
which is an evolving threat to aquatic and
terrestrial ecosystems.
[2]
For instance, the
accumulation of microplastics in aquatic
ecosystems ultimately integrates them into
both marine and human food chains, while
their high surface area-to-volume ratio
enables them to absorb organic pollutants,
further complicating the threat to living
species.
[3]
Microplastics originate from
diverse sources, including raw materials
used in pellets, textiles, and personal care
products, and they pervade various envi-
ronments, from soil and groundwater to
plants, and oceans.
[4–7]
Polyethylene (PE),
polypropylene (PP), polystyrene (PS), and
polyethylene terephthalate (PET) are
among the plastics most commonly identi-
fied as contributors to this escalating envi-
ronmental issue.
[8]
The global recognition
of the eradication of microplastic pollution
as a matter of utmost importance is now
widespread.
Various technologies have been consid-
ered for the remediation of this issue,
including recycling, degradation, and upcycling. The conven-
tional approach of recycling faces significant challenges due to
labor-intensive separation processes and high costs, resulting
in less than 10% of plastic waste being recycled in the United
States in 2018.
[9]
Upcycling, which involves transforming plastic
waste into higher-value materials rather than simply breaking it
down into less valuable forms, is increasingly becoming a
preferred approach. Various methods, such as pyrolysis,
[10–12]
chemical vapor deposition (CVD),
[13]
catalytic carbonization,
[14]
and flash Joule heating (FJH)
[15–17]
have been employed to trans-
form waste plastics into graphene, an incredibly valuable and
futuristic material. However, these methods come with draw-
backs, including the requirement of specific substrates, catalysts,
high vacuum conditions, etc.
[18]
Moreover, these techniques are
susceptible to potential contaminations from various sources,
such as parent materials or during the transfer of graphene
which can compromise the quality of graphene.
[19]
Table 1
provides a comprehensive comparison of these techniques.
A recently emerging technique known as atmospheric
pressure microwave plasma (APMP) offers a significant advan-
tage in graphene production. Using APMP, diverse precursors
like methane, ethanol, and oil vapors have been successfully
M. A. Zafar, M. V. Jacob
Electronics Materials Lab
College of Science and Engineering
James Cook University
Townsville, QLD 4811, Australia
E-mail: mohan.jacob@jcu.edu.au
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/smsc.202400176.
© 2024 The Author(s). Small Science published by Wiley-VCH GmbH.
This is an open access article under the terms of the Creative
Commons Attribution License, which permits use, distribution and
reproduction in any medium, provided the original work is properly cited.
DOI: 10.1002/smsc.202400176
Microplastic pollution poses a growing threat to ecosystems globally,
necessitating sustainable solutions. This study explores upcycling microplastics
into graphene as a promising approach Traditional methods like pyrolysis and
catalytic carbonization are slow and compromise graphene quality. Flash Joule
heating is fast but energy-intensive and hard to control. In contrast, atmospheric
pressure microwave plasma (APMP) synthesis, the proposed technique, offers a
one-step, environmentally friendly alternative. APMP operates at relatively lower
temperatures, reducing energy consumption and providing precise control over
process parameters. This study demonstrates that polyethylene microplastics
from waste dropper bottles can be efficiently transformed into graphene using
APMP synthesis. Raman spectroscopy of synthesized material reveals a spectrum
characteristic of graphene-based materials, with indications of defects and the
presence of oxygen content. X-ray diffraction illustrates the characteristic gra-
phitic lattice, with a slightly larger interlayer spacing attributed to intercalated
functional groups. X-ray photoelectron spectroscopy confirms sp
2
hybridized
carbon as the major component. High-resolution transmission electron
microscopy provides insights into the multilayered structure and variations in
interlayer spacing. The as-synthesized pristine graphene exhibits nearly ten times
greater efficiency in adsorbing perfluorooctanoic acid compared to the oxidized
form of graphene, although it is slightly less effective than graphene-based
nanocomposites.
RESEARCH ARTICLE
Editor’s Choice www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (1 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
transformed into contaminant-free, high-quality graphene.
[20,21]
Notably, this process is conducted under ambient conditions,
eliminating the necessity for high vacuum, substrates, or catalysts.
However, its applicability is restricted to gaseous or vapor-phase
precursors, neglecting the potential of solid precursors.
[22–26]
In this study, the goal is to broaden the application of APMP to
solid precursors, specifically converting microplastics into
graphene. In contrast to the traditional method of initiating gra-
phene production from gaseous-phase products, this approach
involves the transformation of PE microplastics into gases such
as methane, ethylene, and ethane, and then converting them into
graphene within the plasma, all in one step. Furthermore, the
advantages of microwave-based technologies in terms of energy
consumption and cost compared to conventional techniques
for recycling or upcycling polymers can be found in recently
reported studies.
[27–29]
Following the successful synthesis of gra-
phene, we also showcase its effectiveness in adsorbing perfluor-
ooctanoic acid (PFOA), facilitated by ultrasonication.
2. Experimental Section
To begin, clean PE microplastics were obtained by crushing PE
dropper bottles using a household blender. Fourier transform
Table 1. Comparative analysis of microplastics and macroplastics upcycling techniques.
Plastic material Method Advantages Disadvantages References
Polystyrene obtained from Petri
dishes (macrosize feedstock)
CVD –Process feasibility for macrosize plastics
–High I
2D
/I
G
ratio and small I
D
/I
G
ratio
–Substrate dependent
–Requires high vacuum and preheating
–Long process
–Graphene transfer complications
–Unscalable
[47]
Plastic bottles and packaging
material (macrosize feedstock)
CVD –Flexible and foldable paper-like graphene foil
–Various types of plastic materials were tried which
showed good Raman characteristics
–High I
2D
/I
G
ratio and small I
D
/I
G
ratio
–Substrate dependent
–Requires high vacuum and preheating
–Long process
–Graphene transfer complications
–Unscalable
[48]
PE and PS obtained from packaging
material (macrosize feedstock)
Thermal
decomposition þ
CVD
–Atmospheric pressure synthesis
–Absence of D-peak in Raman spectrum
–Two-stage process
–Substrate dependent
–Requires preheating
–Long process
–Graphene transfer complications
–Unscalable
–Requires hydrogen gas
[13]
PET obtained from water bottles
(macrosize feedstock)
CVD –Waste gases from plastic pyrolysis are
converted into graphene
–Synthesis of monolayer graphene
–Requires copper film coating on a
substrate
–Long process
–Graphene transfer complications
–Unscalable
[49]
Polypropylene obtained from waste
bumper and panel (macrosize
feedstock)
Pyrolysis þ
carbonization
Scalable method –Multistage process including pyrolysis,
carbonization, purification
–Use of toxic chemicals
–Poor quality of graphene
–20 layers graphene
[14]
Polypropylene ash obtained after its
pyrolysis (macrosize feedstock)
Pyrolysis þFJH Scalable technique –Moderate I
2D
/I
G
ratio and small I
D
/I
G
ratio
–High risk of doping and contamination
due to no exhaust in the system
–Postsynthesis grinding may contaminate
and damage the graphene structure
[16]
Mixture of six polymers (microsize
feedstock)
FJH –High yield
–Fast process
–Do not require sorting of plastics
–Requires conductive material for synthesis
such as carbon black
–Prone to contamination
–Moderate I
2D
/I
G
ratio and small I
D
/I
G
ratio
[15]
Mixture of polymers obtained from
vehicle (microsize feedstock)
FJH –High yield
–Fast process
–Do not require sorting of plastics
–Requires conductive material for synthesis
such as coke
–Prone to contamination
–Moderate I
2D
/I
G
ratio and small I
D
/I
G
ratio
[17]
PE obtained by crushing dropper
bottles (microsize feedstock)
APMP –Substrate-free synthesis at ambient conditions
–Fast, single-step process
–Contamination-free production
–Low yield of graphene
–Moderate I
2D
/I
G
ratio and small I
D
/I
G
ratio
This work
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (2 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
infrared (FTIR) spectra of the crushed microplastics (Figure S1,
Supporting Information) confirm their PE bonding structure.
These microplastics were then sieved to achieve uniform particle
sizes ranging from 1 to 3 mm. Subsequently, 30 mg of the sieved
microplastics were placed in a ceramic boat positioned at the cen-
ter of a quartz tube within the plasma system. The plasma system
consists of a 2.45 GHz microwave power supply, a tuner, and a
quartz reaction tube (30 mm OD), with argon gas used as the
background gas to provide an oxygen-free atmosphere. The oper-
ating conditions include microwave power levels of 400, 500, and
600 W, with an argon gas flow rate of 2 slm and reaction time of
1 min. The optimal conditions mentioned were established
through extensive parametric investigations, although detailed
parameters are not provided here. The graphene nanosheets
synthesized through this approach are free-standing and can
be collected at the open end of the tube and on silicon substrate,
or they may accumulate on the walls of the quartz tube. Figure 1
schematically depicts the synthesis process.
The synthesis mechanism inside plasma can be divided into
two stages. Initially, the plasma efficiently breaks down the
microplastics, converting them into constituent gases such as
methane (CH
4
), ethylene (C
2
H
4
), ethane (C
2
H
6
), carbon dioxide
Figure 1. a) The figure depicts dropper bottles that were crushed (without caps) into microplastics (1–3 mm sizes). b) Schematic representation of APMP
system for the synthesis of graphene from PE microplastics.
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (3 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
(CO
2
), hydrogen (H
2
), and carbon monoxide (CO). Subsequently,
hydrocarbons, particularly methane undergo further processing
within the system, experiencing plasma dissociation and conver-
sion into graphene. This graphene is deposited onto the walls of
the quartz tube, from which it can be collected for subsequent
analysis and characterization. The conversion of PE into constit-
uent gases can be supported by previous studies,
[30–32]
where PE
is converted into various gases upon severe thermal degradation,
either by plasma or conventional heating. These gases serve as
precursors for further transformations.
2.1. Characterization Techniques
To characterize the vibrational properties of graphene, Raman
spectroscopy was employed (Witec, excitation-beam wavelength:
532 nm). The morphology of the graphene samples was analyzed
using a field-emission scanning electron microscope (SEM,
Hitachi SU 5000). Prior to SEM observation, the samples
underwent sputter coating with platinum. X-ray diffraction
(XRD) analysis was conducted using a Bruker D8-Advance
X-ray diffractometer equipped with Cu Kαradiation
(λ=0.154 nm). FTIR was performed using a Spectrum-100
spectrometer from Perkin Elmer, USA. The surface element
composition of the graphene was assessed using X-ray photoelec-
tron spectroscopy (XPS), which was performed on a Kratos Axis
Ultra XPS instrument featuring an Al KαX-ray source. To delve
into the microstructure of the graphene nanosheets, high-
resolution transmission electron microscopy (HRTEM, Hitachi
HF 5000) operating at 200 kV was employed.
3. Results and Discussion
The as-synthesized material was subjected to Raman spectros-
copy for structural analysis. The Raman spectra of all samples
(shown in Figure 2a), regardless of the plasma power used for
growth, exhibited three characteristic vibrational modes associ-
ated with graphene-based materials: a D peak at ≈1333 cm
1
(indicative of defects), a G peak at ≈1576 cm
1
(related to vertical
vibration), and a 2D peak at ≈2674 cm
1
(associated with two-
phonon vibration).
[33]
The increase in temperature resulting
from the rise in plasma power led to a reduction in the height
of the D peak. This decline in intensity can be attributed to the
removal of oxygen attached to graphene or the potential reduc-
tion in defects within the graphene structure. The intensity ratio
of D and G peaks, i.e., I
D
/I
G
, which serves as a measure of dis-
order, is notably higher in the 400 W sample. This observation is
consistent with prior research, such as the work of Marsden
et al.
[34]
where higher oxygen dosing in graphene exhibited a
higher I
D
/I
G
value. Conversely, the 500 and 600 W samples
exhibited lower I
D
/I
G
values, suggesting a reduced presence
of defects or oxygen content compared to the 400 W sample.
However, it can be noted that there were minimal differences
in the I
D
/I
G
or I
2D
/I
G
ratios between the 500 and 600 W samples,
as shown in Figure 2b. This indicates that microwave power
beyond 500 W has a diminishing impact on the synthesis
process.
The 2D peak in the Raman spectrum is a mark of graphene-
based materials, with characteristics like full width at
half-maximum (FWHM) and intensity ratio (I
2D
/I
G
) linked to
the number of graphene layers. FWHM values of ≈30 cm
1
and I
2D
/I
G
values of 2 or higher typically indicate monolayer
graphene, while FWHM values of ≈50 cm
1
and I
2D
/I
G
ratios
of 1 to 1.5 suggest bilayer structures.
[21]
The 500 and 600 W sam-
ples displayed similar FWHM values, measuring 84.8 and
78.3 cm
1
, respectively, while the FWHM of the 400 W sample
was 65.3 cm
1
. Evaluating the I
2D
/I
G
and FWHM values of sam-
ples, it can be inferred that all samples consist of multiple-layer
graphene. Nonetheless, the microwave power of 500 W is
deemed suitable due to its lower I
D
/I
G
value.
Crystalline materials exhibit an exclusive XRD pattern, often
associated with a fingerprint for material identification. In
Figure 2c, the characteristic peak of 500 W sample, positioned
at a diffraction angle (2θ) of 26.3°, explicitly confirmed the effec-
tive creation of the graphitic lattice. This result aligns well with
previous reports.
[35,36]
The peak position can be attributed to the
(002) basal plane, denoting an interlayer spacing of 0.37 nm, a
measurement further validated by TEM. It is noteworthy that this
layer spacing slightly exceeds the 0.33 nm, usually reported for
graphene. We ascribe this discrepancy to the presence of inter-
calated C─OH and C─O functional groups attached within the
layers.
The 500 W graphene sample underwent XPS analysis to
determine its elemental content and functional group types.
The survey scan XPS (Figure 2d) revealed the presence of
intense carbon (C) and small oxygen (O) peaks at 284.3 and
532.5 eV, respectively. The composition of C and O elements
was ≈98.10% and 1.9%, respectively. High-resolution spectra
of C1s and O1s were further analyzed to study the bonding
construction.
The C1s high-resolution spectrum of 500 W sample was
divided into four constituent peaks (Figure 2e). The primary
C═C peak (sp
2
-C), located at 284.3 eV, is vital for graphene
material, and it indicates the occurrence of a honeycomb lattice
structure. Additionally, the spectrum contained C─C (sp
3
-C) and
C─OH/C─O peaks at 285.2 and 288.1 eV, respectively. The sp
3
-C
peak was ascribed to either the edges of the graphene or doping
effects within the structure.
[26]
The O1s spectrum (Figure 2f ) dis-
played impacts from O═C and C─OH/C─O peaks at 532 and
533 eV, respectively.
[37,38]
Comparing the XPS of 500 W sample
with the 600 W sample (Figure S2, Supporting Information), it
can be observed that the carbon content increased, whereas
the oxygen content decreased slightly in the 600 W sample.
Furthermore, in the C1s high-resolution spectrum, sp
2
bonding
decreased and sp
3
bonding increased compared to the 500 W
sample. This could be related to the decreased 2D peak strength
of the 600 W sample in the Raman spectrum. The differences in
FWHM and peak positions observed in the XPS spectra of the
500 and 600 W samples can be attributed to surface charging
effects, which can lead to shifts in binding energy positions, even
though the samples were prepared on the same silicon substrate.
Additionally, inhomogeneities within the samples, despite simi-
lar morphology in SEM images, may result in discrepancies in
peak positions and FWHM. It should be noted that we used the
same peak fitting conditions for all samples.
Figure 3a,b displays low- and high-resolution SEM images of
the 500 W graphene nanosheets, directly deposited onto silicon
substrate. As expected, the SEM images did not reveal smooth,
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (4 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
two-dimensional films. Instead, all the samples showcased
typical three-dimensional island-like structures, resembling
crumpled and folded sheets of paper scattered across the
surface.
[39]
This unique, curly, and wrinkled morphology pre-
dominantly stems from the inherent thermodynamic instability
of two-dimensional materials.
[40]
Further validation of the
primarily multilayered graphene composition comes from the
HRTEM results, as depicted in Figure 3c–f. The high-resolution
SEM images of the 400 and 500 W samples in Figure S3,
Supporting Information show a significant resemblance to the
500 W sample, with no apparent differences. The differences
in Raman spectra despite the similar SEM images could be
due to variations in the local chemical composition or structural
defects that are not discernible in the SEM images. Raman spec-
troscopy is highly sensitive to molecular vibrations and can detect
subtle changes in the material’s structure and composition that
are not visible in SEM. These variations could result from
localized differences in the distribution of functional groups
(oxygen functional groups in our case), strain, or defects within
the graphene sheets. Thus, while the overall morphology appears
consistent, the Raman spectra reveal underlying heterogeneities
in the material’s structural properties.
Figure 2. a) Raman spectra b) and the intensity ratios of the D and 2D peaks concerning the G peak in the graphene samples which demonstrate the
influence of plasma power on the synthesis process. c) XRD pattern of 500 W graphene sample. 500 W graphene sample d) survey scan XPS, e) C1s high
resolution, and f ) O1s high resolution.
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (5 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
Low-resolution TEM and HRTEM analyses were conducted on
the 500 W graphene sample. The specimens used for TEM stud-
ies were mechanically detached from the silicon substrates. The
TEM images in Figure 3c–f reveal the presence of both large and
small graphene nanosheets. These graphene nanosheets exhibit
an interesting resemblance to rippled silk waves, with some
regions appearing transparent. The most transparent and seem-
ingly featureless areas, as indicated by arrows in Figure 3c, are
likely monolayer graphene nanosheets. In Figure 3d–f, high-
magnification TEM images provide further insights. The
HRTEM images depicted that the graphene material consisted
of multiple layers, as evidenced by the nine-layer structure shown
in the inset of Figure 3d. The images also unveiled an interlayer
spacing ranging from 0.34 to 0.36 nm, nearly matching the
(0 0 2) plane spacing as calculated from XRD analysis. The
observed variances may be attributed to the presence of a small
quantity of oxygen atoms occupying interstitial sites, inducing
structural discontinuities, and promoting dislocation formation
within the graphene lattice. HRTEM also revealed that while some
graphene particles exist independently, most graphene sheets are
interconnected, forming a large, crumpled structure due to
agglomeration. This phenomenon results in partial overlapping
and coalescence of the graphene sheets. The quantity of graphene
produced in this approach mainly relies on the quantity of micro-
plastic feedstock. ≈30 mg of microplastics produced nearly 5 mg of
graphene in 1 min. This production rate is remarkably higher than
that achieved in previously used gas-phase synthesis in APMP,
where the feedstock was ethanol,
[24]
methane,
[22]
or oil vapors.
[41]
Figure 3. a,b) Low- and high-magnification SEM images of 500 W graphene sample. c) Low-magnification TEM and d–f ) HRTEM images of 500 W
graphene sample.
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (6 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
3.1. Experimental Procedure for PFOA Adsorption Using
Synthesized Graphene
The graphene synthesized at 500 W was employed to assess its
adsorption capacity for PFOA. In a related study by Lath et al.
[42]
graphene oxide (GO) was utilized for adsorption using an orbital
shaker, resulting in an adsorption efficiency of only 3.3% in a
3–4 h procedure. In the current investigation, we focused on pris-
tine graphene with the assistance of ultrasonication at varying
durations. Three identical samples were prepared by combining
10 mg of PFOA with 1 mg of graphene in 1 liter of distilled water
within separate containers. Ultrasonication was conducted for
durations of 15 and 30 min at a pH of 7 and at room temperature,
using a specialized 40 kHz ultrasonicator with a power output of
30 W. Additionally, a comparative sample with a 30 min mag-
netic stirring method was prepared. After both ultrasonication
and magnetic stirring, the samples underwent filtration using
a 0.45 μmfilter to effectively remove graphene nanosheets.
Subsequently, the filtered samples were sent to the laboratory
for analysis.
3.2. Adsorption Capacity of Graphene for PFOA
The removal efficiency was calculated using Equation (1),
where C
0
and C
e
are the initial and final concentrations,
respectively.
Rð%Þ¼C0Ce
C0
100 (1)
The graphical representation in Figure 4 represents the results
of the adsorption study. The 15 min ultrasonicated sample exhib-
ited a ≈30% adsorption capacity, which increased to ≈32% when
the sonication time was extended to 30 min. Notably, when com-
pared to the previously reported GO sample,
[42]
which utilized an
orbital shaker for enhanced adsorption and showed an efficiency
of 3.3%, pristine graphene demonstrated significantly higher
results. The increased adsorption capacity in this study is attrib-
uted to graphene’s hydrophobic nature, fostering interactions
with PFOA chains. This supports previous reports highlighting
high adsorption capacities linked to the hydrophobic nature of
carbon materials.
[43,44]
Ultrasonication likely contributed by
increasing surface area and creating additional adsorption sites
through improved dispersion and mass transfer. Ultrasonication
surpassed magnetic stirring in combating agglomeration in gra-
phene nanosheets. While both methods aid deagglomeration,
ultrasonication’s intensity ensures thorough exposure of nano-
sheets, maximizing accessible surface area for enhanced adsorp-
tion. However, it is essential to note that real-world factors,
including pH, temperature, and contaminants, can influence
adsorption capacity.
[45]
Furthermore, while the advantages of
ultrasonication for adsorption are well known, a direct compari-
son of PFOA adsorption capacities between graphene synthe-
sized in this study and GO using both ultrasonication and
orbital shaking would provide a more comprehensive
understanding of the influence of these agitation methods on
adsorption performance.
Previously, graphene derivatives such as GO and fluorogra-
phene (FG), in composite with various other materials, have
demonstrated excellent adsorption capacities exceeding 90%.
However, pristine graphene, which possesses a higher adsorp-
tion capacity than GO, has not been investigated in this regard.
For instance, GO composite with iron oxide
[42]
showed more
than 90% adsorption. Wang et al.
[46]
synthesized a novel adsor-
bent by attaching hydrophobic FG to hydrophilic magnetic nano-
particles (MNPs). The resulting MNPs@FG exhibited impressive
removal efficiencies of 91–97% for PFOA and perfluoro octane
sulfonic acid. In an effort, to enhance this efficiency beyond 97%,
it is proposed that incorporating pristine graphene with other
materials may prove beneficial.
4. Conclusion
In summary, this research has successfully demonstrated the
APMP synthesis as a rapid and facile method for fabricating
graphene from microplastics, marking a significant milestone
in this field. In comparison to established techniques such as
CVD, pyrolysis, and FJH, the plasma-based synthesis displayed
superior characteristics, featuring contamination-free produc-
tion at ambient conditions. The method demonstrated the swift
conversion of PE microplastics into graphene, a transformation
confirmed by Raman spectroscopy, revealing graphene charac-
teristics along with additional defects and oxygen content.
XRD provided insight into the graphitic lattice, while XPS con-
firmed sp
2
hybridized carbon. HRTEM offered a glimpse into the
multilayered structure with interstitial spacing ranging between
0.34 and 0.36 nm. The pristine graphene synthesized through
this process exhibited significant efficiency in adsorbing
PFOA, positioning it as a promising candidate for addressing
environmental challenges linked to microplastics. This research
not only pioneers a novel approach to graphene synthesis but also
contributes to the broader goal of mitigating the adverse effects
of microplastic pollution on our ecosystems.
Figure 4. PFOA adsorption capacity on pristine graphene at different
durations (15 and 30 min) with ultrasonication and magnetic stirring.
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (7 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH
Supporting Information
Supporting Information is available from the Wiley Online Library or from
the author.
Acknowledgements
Open access publishing facilitated by James Cook University, as part of the
Wiley - James Cook University agreement via the Council of Australian
University Librarians.
Conflict of Interest
The authors declare no conflict of interest.
Data Availability Statement
The data that support the findings of this study are available from the
corresponding author upon reasonable request.
Keywords
graphene, microplastics, plasma, sustainable synthesis
Received: April 18, 2024
Revised: July 23, 2024
Published online: August 7, 2024
[1] Z. Lin, T. Jin, T. Zou, L. Xu, B. Xi, D. Xu, J. He, L. Xiong, C. Tang,
J. Peng, Y. Zhou, J. Fei, Environ. Pollut. 2022,304, 119159.
[2] K. Ziani, C.-B. Ionit
¸˘
a-Mîndrican, M. Mititelu, S. M. Neacşu, C. Negrei,
E. Moroşan, D. Dr˘
ag˘
anescu, O.-T. Preda, Nutrients 2023,15, 617.
[3] A. Baruah, A. Sharma, S. Sharma, R. Nagraik, Int. J. Environ. Sci.
Technol. 2022,19, 5721.
[4] X. Z. Lim, Nature 2021,593, 22.
[5] R. W. Chia, J.-Y. Lee, H. Kim, J. Jang, Environ. Chem. Lett. 2021,19,
4211.
[6] I. Nabi, A.-U.-R. Bacha, L. Zhang, J. Cleaner Prod. 2022,337, 130458.
[7] R. Ahmed, A. K. Hamid, S. A. Krebsbach, J. He, D. Wang,
Chemosphere 2022,293, 133557.
[8] M. Padervand, E. Lichtfouse, D. Robert, C. Wang, Environ. Chem. Lett.
2020,18, 807.
[9] J. Choi, I. Yang, S.-S. Kim, S. Y. Cho, S. Lee, Macromol. Rapid
Commun. 2022,43, 2100467.
[10] S. Pandey, M. Karakoti, S. Dhali, N. Karki, B. Santhibhushan,
C. Tewari, S. Rana, A. Srivastava, A. B. Melkani, N. G. Sahoo,
Waste Manage. 2019,88, 48.
[11] N. A. El Essawy, S. M. Ali, H. A. Farag, A. H. Konsowa, M. Elnouby,
H. A. Hamad, Ecotoxicol. Environ. Saf. 2017,145, 57.
[12] A. Rendón-Pati˜no, J. Niu, A. Doménech-Carbó, H. García, A. Primo,
Nanomaterials 2019,9, 101.
[13] S. Sharma, G. Kalita, R. Hirano, S. M. Shinde, R. Papon, H. Ohtani,
M. Tanemura, Carbon 2014,72, 66.
[14] J. Gong, J. Liu, X. Wen, Z. Jiang, X. Chen, E. Mijowska, T. Tang, Ind.
Eng. Chem. Res. 2014,53, 4173.
[15] W. A. Algozeeb, P. E. Savas, D. X. Luong, W. Chen, C. Kittrell, M. Bhat,
R. Shahsavari, J. M. Tour, ACS Nano 2020,14, 15595.
[16] K. M. Wyss, J. L. Beckham, W. Chen, D. X. Luong, P. Hundi,
S. Raghuraman, R. Shahsavari, J. M. Tour, Carbon 2021,174, 430.
[17] K. M. Wyss, R. D. De Kleine, R. L. Couvreur, A. Kiziltas,
D. F. Mielewski, J. M. Tour, Commun. Eng. 2022,1,3.
[18] G. He, M. Yan, H. Gong, H. Fei, S. Wang, Int. J. Extreme Manuf. 2022,
4, 032003.
[19] M. S. A. Kamel, C. T. Stoppiello, M. V. Jacob, Carbon 2023,202, 150.
[20] J. Toman, M. Šnírer, R. Rincón, O. Jašek, D. Všianský, A. M. Raya,
F. J. Morales-Calero, J. Mu˜noz, M. D. Calzada, Fuel Process.
Technol. 2023,239, 107534.
[21] M. A. Zafar, O. K. Varghese, F. C. Robles Hernandez, Y. Liu,
M. V. Jacob, ACS Appl. Mater. Interfaces 2022,14, 5797.
[22] N. Bundaleska, D. Tsyganov, A. Dias, E. Felizardo, J. Henriques,
F. M. Dias, M. Abrashev, J. Kissovski, E. Tatarova, Phys. Chem.
Chem. Phys. 2018,20, 13810.
[23] O. Jašek, J. Toman, J. Jurmanová, M. Šnírer, V. Kudrle, V. Buršíková,
Diamond Relat. Mater. 2020,105, 107798.
[24] C. Melero, R. Rincón, J. Mu˜noz, G. Zhang, S. Sun, A. Perez,
O. Royuela, C. González-Gago, M. D. Calzada, Plasma Phys.
Controlled Fusion 2018,60, 014009.
[25] D. Tsyganov, N. Bundaleska, A. Dias, J. Henriques, E. Felizardo,
M. Abrashev, J. Kissovski, A. M. B. Do Rego, A. M. Ferraria,
E. Tatarova, Phys. Chem. Chem. Phys. 2020,22, 4772.
[26] M. A. Zafar, Y. Liu, S. Allende, M. V. Jacob, Appl. Surf. Sci. 2023,638,
158006.
[27] Y. Luo, E. Selvam, D. G. Vlachos, M. Ierapetritou, ACS Sustainable
Chem. Eng. 2023,11, 4209.
[28] S. S. Alam, A. H. Khan, Int. J. Environ. Sci. Technol. 2024,21, 5311.
[29] Z. Jiang, Y. Liang, F. Guo, Y. Wang, R. Li, A. Tang, Y. Tu, X. Zhang,
J. Wang, S. Li, L. Kong, ChemSusChem 2024, e202400129.
[30] L. Tang, H. Huang, Fuel Process. Technol. 2007,88, 549.
[31] M. He, B. Xiao, Z. Hu, S. Liu, X. Guo, S. Luo, Int. J. Hydrogen Energy
2009,34, 1342.
[32] S. Colantonio, L. Cafiero, D. De Angelis, N. M. Ippolito, R. Tuffi,
S. V. Ciprioti, Front. Chem. Sci. Eng. 2020,14, 288.
[33] J. Romo-Rico, R. Bright, S. M. Krishna, K. Vasilev, J. Golledge,
M. V. Jacob, Carbon Trends 2023,12, 100282.
[34] A. J. Marsden, M. Skilbeck, M. Healey, H. R. Thomas, M. Walker,
R. S. Edwards, N. A. Garcia, F. Vukovi´c, H. Jabraoui, T. R. Walsh,
J. P. Rourke, N. R. Wilson, Phys. Chem. Chem. Phys. 2022,24, 2318.
[35] H. Gao, Y. Tan, S. Wang, L. Sun, J. Jin, Y. Zhao, Y. Zhao, J. Alloys
Compd. 2023,968, 171858.
[36] X. Wang, L. Zhang, RSC Adv. 2019,9, 3693.
[37] J. V. Rojas, M. Toro-Gonzalez, M. C. Molina-Higgins, C. E. Castano,
Mater. Sci. Eng.: B 2016,205, 28.
[38] P. Kumar, N. Divya, J. K. Ratan, J. Iran. Chem. Soc. 2021,18, 201.
[39] F. Wang, F. Wang, R. Hong, X. Lv, Y. Zheng, H. Chen, J. Mater. Res.
Technol. 2020,9, 10004.
[40] X. Gao, C. Xu, H. Yin, X. Wang, Q. Song, P. Chen, Nanoscale 2017,9,
10639.
[41] M. A. Zafar, Y. Liu, F. C. R. Hernandez, O. K. Varghese, M. V. Jacob,
Adv. Mater. Interfaces 2023,10, 2202399.
[42] S. Lath, D. A. Navarro, D. Losic, A. Kumar, M. J. Mclaughlin, Environ.
Chem. 2018,15, 472.
[43] M. Inyang, E. R. V. Dickenson, Chemosphere 2017,184, 168.
[44] M. Park, S. Wu, I. J. Lopez, J. Y. Chang, T. Karanfil, S. A. Snyder, Water
Res. 2020,170, 115364.
[45] D. Dey, T. Shafi, S. Chowdhury, B. K. Dubey, R. Sen, Chemosphere
2024,351, 141164.
[46] W. Wang, Z. Xu, X. Zhang, A. Wimmer, E. Shi, Y. Qin, X. Zhao,
B. Zhou, L. Li, Chem. Eng. J. 2018,343, 61.
[47] G. Ruan, Z. Sun, Z. Peng, J. M. Tour, ACS Nano 2011,5, 7601.
[48] L. Cui, X. Wang, N. Chen, B. Ji, L. Qu, Nanoscale 2017,9, 9089.
[49] Y. You, M. Mayyas, S. Xu, I. Mansuri, V. Gaikwad, P. Munroe,
V. Sahajwalla, R. K. Joshi, Green Chem. 2017,19, 5599.
www.advancedsciencenews.com www.small-science-journal.com
Small Sci. 2024,4, 2400176 2400176 (8 of 8) © 2024 The Author(s). Small Science published by Wiley-VCH GmbH