Uv Radiation GO in Water
Uv Radiation GO in Water
Uv Radiation GO in Water
Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere
h i g h l i g h t s g r a p h i c a l a b s t r a c t
a r t i c l e i n f o a b s t r a c t
Article history: Graphene oxide (GO) is widely used in different applications, however once released into the environ-
Received 4 November 2019 ment it can change its structure and affect the transport of important contaminants such as arsenic. In
Received in revised form this work we show that UV radiation, even in the range of 28-74 mW/cm2 of irradiance up to 120 h of
4 February 2020
exposure, can induce important changes in the structure of graphene oxide, by eliminating -OH and C]O
Accepted 8 February 2020
Available online 10 February 2020
functional groups. This reduction affected the stability of graphene oxide in water by decreasing its zeta
potential from -41 to -37 mV at pH¼7 with the increase of the exposure time. Our results showed that
Handling Editor: Prof. X. Cao after 24 and 120 h of UV exposure, As(III) adsorption capacity decreased from 5 mg/g to 4.7 and 3.8 mg/g,
respectively, suggesting a lower capacity to transport contaminants with time. Computer modelling
Keywords: showed that even a degraded GO structure can have an interaction energy of 223.84 kJ/mol with H3AsO3.
Graphene oxide Furthermore, we observed that the cytotoxicity of graphene oxide changed after being irradiated at
Ultraviolet radiation 74 mW/cm2 for 120 h, showing 20% more cell viability compared to as-produced GO. Our results stress the
Arsenic adsorption importance of considering the microstructural and compositional changes that GO undergoes even
Reduction
under low irradiance and short periods, when studying its fate and behavior in the environment and
Cytotoxicity
possible applications in water treatment.
© 2020 Elsevier Ltd. All rights reserved.
* Corresponding author.
pez-Honorato).
E-mail address: [email protected] (E. Lo
https://doi.org/10.1016/j.chemosphere.2020.126160
0045-6535/© 2020 Elsevier Ltd. All rights reserved.
2 W.R. Gallegos-Perez et al. / Chemosphere 249 (2020) 126160
1300 W with a plasma gas flow rate of 15 L/min, nebulization gas CO2). Production of formazan was quantified by using a microplate
flow rate of 0.55 L/min and wavelengths of 188.9, 193.7, 197.2 and reader (Epoch, Bioteck), monitoring the difference in absorbance at
228.8 nm). Transmission electron microscopy analysis was per- 492 and 690 nm. All experiments were performed in triplicate. The
formed in a FEI-Talos F200S with four in-column Super-X EDS de- effect of each treatment was expressed as a percentage of cell
tectors with a beam current of 300 pA and a collection time of proliferation relative to untreated control cells, differences be-
10 min. tween control and treated groups are shown as a result of a two-
way ANOVA statistics.
2.5. Cell culture
Primary mononuclear cells were isolated using density-gradient 2.6. Computer modelling
separation with the reagent Histopaque-1077 (Sigma-Aldrich,
USA), according to the manufacturer’s instructions. All experiments H3AsO3, H2AsO 3 and HAsO3
2
were independently adsorbed
were set up with mononuclear cells extracted from normal in- over degraded graphene oxide and their geometries were opti-
dividuals (n¼5). Cells were cultured with RPMI 1640 medium plus mized at the LC-uPBE/6-31G (d,p) (Fig. S5) level of theory with the
10% of Fetal Bovine Serum, 10 mM Penicillin Streptomycin, and use of the Gaussian 09 suite of programs (Frisch, 2009) The inter-
10 mM of L-glutamine (Thermofisher Scientific, USA) in a 37 C, 5% action energies were also calculated with the NBODel procedure
CO2 humidified incubator. Cell viability was evaluated by the XTT with the NBO3.1Weinhold and Glendening, 1996 program as pro-
(sodium 2,3,-bis(2-methoxy-4-nitro-5-sulfophenyl)-5-[(phenyl- vided within the aforementioned suite, and the obtained values are
amino)-carbonyl]-2H tetrazolium, Roche™) assay. Briefly, cells collected in Table 3. The NBODel procedure deletes all orbital in-
were seeded in a 96-well plate at a density of 5 103 cells/well, and teractions between both species and the concomitant raise in en-
the compounds were incorporated into the medium at a concen- ergy is ascribed to the interaction between them.
tration of GO of 10, 30 and 50 mg/ml (GO after UV irradiation was
filtered and dried to prepare these solutions in order to ensure the Table 3
same amount of mass regardless of the effect of UV irradiation on Interaction Energies [kJ/mol] calculated at the LC-uPBE/6-31G (d,p) level of theory.
the structure of the material). The compounds employed for the
Functional group As (III)
treatments were freshly prepared in Dimethyl sulfoxide (DMSO);
H3AsO3 H2AsO HAsO2-
control cells were incubated only with DMSO. After 48 h, the me- 3 3
dium was aspirated and cells were washed with PBS. Cells were Graphene 16.32 1064.99 1603.93
cultured in a mix of 60 ml of DMEM/F12 medium without FBS plus Epoxide 1508.62 1380.63 3.01
Hydroxyl 1583.64 1603.93 1585.65
40 ml of XTT-solution for 2 h in a humidified incubator (at 37 C, 5%
Table 1
Percentage of functional groups in GO and irradiated GO (UV3 and UV5) at 37 mW/cm2 and 74 mW/cm (Pendolino and Armata, 2017).
Type of bond GO 72 h (UV3), 37 mW/cm2 72 h (UV3) 74 mW/cm2 120 h (UV5) 37 mW/cm2 120 h (UV5) 74 mW/cm2
% of bond
Table 2
Oxygenated functional groups lost after ultraviolet radiation.
Light source l (nm) Power (W) Irradiance Time (h) Lost functional group
3. Results and discussions OH, C-O-C, O-C]O, COOH groups decreased 17.2, 0.6, 4.2, 1%,
respectively. It should be noted that compared to 72 h of irradiation,
3.1. Changes in composition the CeC bonds showed an increment of 22.5% (see Table 1), and the
concentration of C]C bonds decrease from 32.6 to 12.8% from 72 to
Fig. 1 shows two examples of the FTIR spectra of GO and GO after 120 h. These differences could be related to the bond energy
UV irradiation with different irradiance during 72 (UV3) and 120 h needed to detach a functional group from the C structure (Mulyana
(UV5). In these spectra is possible to identify the signal of the et al., 2014). For example, 360 kJ/mol are required to break a C-O
stretching mode -OH at 3219 cm1, deformation mode of C]O at bond corresponding to the CeOH functional group, whereas 370
1724 cm1, stretching mode of C]C at 1619 cm1, deformation and 680 kJ/mol are needed for a C-C and C]C bond, respectively
mode of C-H at 1372 cm1 and flexion mode of C-O at 1036 cm1, (Glockler, 1958). Therefore, CeOH are transformed faster than the
particularly for the reference material, GO (Konkena and C-C and C]C bonds (Glockler, 1958). After 120 h of irradiation, due
Vasudevan, 2012). As the UV irradiance increased from 28 to to the small concentration of oxygenated functional groups, carbon
74 mW/cm (Pendolino and Armata, 2017), the intensities of these bonds appear to be altered.
bands changed, most of them decreased, giving the first sing of a As can be seen in Fig. 3 and Table 1, as the UV radiation increase
structural transformation. For example, the most notorious to 74 mW/cm2 for 72 h, the C]C decreased 5.7% and the C-C, C-OH,
changes, in both UV3 and UV5, are the disappearance of the -OH C-O-C bonds increased 3.8, 6.6, and 2.7% respectively, compared to
band and considerable reduction of the C]O band at 74 mW/cm the as-produced GO. However, the C]C bonds decrease 26.3% and
(Pendolino and Armata, 2017). Conversely, the signals assigned to the CeC bonds increased 4.3% as the irradiance increased from 37
C]C and CeH bonds increased in relation to the other bands. to 74 mW/cm (Pendolino and Armata, 2017). Conversely, the C]O
Similar behavior is observed for GO irradiated for 24 and 96 h (see and COOH bonds decreased by 5.7 and 1.7%, respectively compared
Figs. S1 and S2 in supporting information). Fig. 1, although quali- to as-produced GO. It should be noted that C-O-C bonds had an
tatively, shows that irradiance plays an important role in GO increase of 3.1% as the UV irradiance increased during the first 72 h
reduction. of irradiation. Similarly, as the irradiation time increased to 120 h
In order to obtain a more quantitative analysis of the effect with 74 mW/cm (Pendolino and Armata, 2017), the C]C bonds
induced by UV irradiation, XPS analysis was performed. Fig. 2 decreased 3.6% and the C-C bonds increased 16.4% compared to the
shows the XPS spectra for carbon (1s) in the as-produced GO and as-produced GO, but with 4.4% and 5.7% less than the value
GO irradiated with 37 mW/cm2 during 72 and 120 h. The as- measured for 120 h with lower irradiance (37 mW/cm2). Similar to
produced GO showed the presence of C]C, C-C bonds (284.4 and the previous cases, the C-OH and C]O functional groups decreased
285.2 eV) attributed to the carbon structure with sp (Pendolino and 11.9, and 6.6%, respectively, compared to pristine GO. Overall, it
Armata, 2017) and sp (Chen et al., 2015) hybridization, respectively. appears that longer periods under UV irradiation reduced the
Additionally, hydroxyl (CeOH 286.4 eV), epoxide (Ce-O-eC oxygen-bearing functional groups, as the C/O ratio changed from
287.1 eV), carbonyl (C]O 288 eV), and carboxylic acid (COOH 1.4 of pristine GO up to 2.2 and 2.7 for 120 h of irradiation. However,
289.2 eV) functional groups (Li and Bubeck, 2013; Liu et al., 2011; as time increased up to 120 h, the C-O-C concentration seems to
Chen and Chen, 2016; Zhang et al., 2016), with a C/O ratio of 1.35, increase up to 7.2%, compared to the results obtained for 72 h of
were observed similar to those previously identified by FTIR in irradiation. The elimination of C-OH, C]O and COOH functional
Fig. 1. The percentage of each type of bond is shown in Table 1. As groups with an increase of C-C and C]C bonds, has also been
the sample got irradiated for 72 h with 37 mW/m (Pendolino and previously observed by Hou et al., even though the irradiance they
Armata, 2017), the C]C and COOH bonds increased 20.6 and used was 65 000 mW/cm (Pendolino and Armata, 2017), 106 times
1.2%, respectively. Conversely, the C-OH, C-O-C, C]O bonds higher than the one used in this study (Hou et al., 2015; Li and
decreased 16.1, 0.4 and 4.7%, respectively (Table 1), as previously Bubeck, 2013), see Table 2.
suggested by FTIR in Fig. 1. The mechanism for GO reduction by UV radiation has not been
Similarly, after 120 h of irradiation with 37 mW/cm (Pendolino elucidated, however, some authors agree that the first step is water
and Armata, 2017), the C-C bonds increased 22%, whereas the C- ionization and the formations of the radicals HO and H, besides
Fig. 1. ATR-FTIR spectra of a) as-produced GO and irradiated GO UV3 (72 h) and b) as-produced GO and irradiated GO UV5 (120 h) at 28, 37, 54, 74 mW/cm2 and at pH¼7.
W.R. Gallegos-Perez et al. / Chemosphere 249 (2020) 126160 5
Fig. 2. HR-XPS spectra for a) GO and irradiated GO, b) UV3 (72 h) and c) UV5 (120 h), at 37 mW/cm2 and at pH¼7.
Fig. 3. HR-XPS spectra for irradiated GO a) UV3 (72 h) and b) UV5 (120 h), at 74 mW/cm2 and at pH¼7.
the formation of solvated electrons (e) (Mohandoss et al., 2017; et al., 2013). The detachment of C-OH groups frequently results in
Hou et al., 2015; Ji et al., 2013; Gengler et al., 2013;Ganguly et al., the formation of new C]C bonds (Silva et al., 2017).
2011; Nyangiwe et al., 2015). The radical H and solvated e act Furthermore, it has been observed that graphene can also be
as reducing agents, especially in groups such as CeOH and C-O-C, oxidized under prolonged exposure to UV radiation and form new
on the basal plane, which are very reactive and are easily detached C-O-C and C-OH functional groups (Hu et al., 2015) by the action of
in the earlier exposure to UV radiation (Mohandoss et al., 2017; Ji the radical HO (Ji et al., 2013). This reduction route could explain
6 W.R. Gallegos-Perez et al. / Chemosphere 249 (2020) 126160
some the changes on C]O and -OH functional groups detected by (Pendolino and Armata, 2017), lower stability is achieved, partic-
XPS, where C-C bonds concentration always increases with irradi- ularly at pH¼5. These changes in zeta potential agree with the re-
ance and time, but the concentration of -OH and -CO, change sults obtained by FTIR and XPS, suggesting the loss of functional
considerably depending on time and irradiance. For example, at groups, particularly C-OH and C]O, which confer its hydrophilic
72 h of irradiation with 74 mW/cm2 the concentration of -OH property to GO and allows it to remain in a stable suspension (Dai
functional groups increased from 24 to 31%, however as time et al., 2015). Furthermore, apparently, regardless of the clear
increased to 120 h, its concentration dropped to 12%. It appears that changes in composition detected by FTIR and XPS, irradiated GO
at short times and low irradiance in the range of 37 mW/cm maintains a certain degree of stability even after 120 h of
(Pendolino and Armata, 2017), GO is reduced However, a prolonged irradiation.
exposure or higher irradiance, allows an increase on oxygen con- Our results suggest that as the composition of the material
tent in GO as an intermediate step on the decomposition and change, its stability in suspension is also affected, therefore, even
reduction of GO, probably due to the formation of a higher con- though low irradiance in water might not mean a full trans-
centration of hydroxyl radicals (Ji et al., 2013) This can be seen in formation into reduced graphene oxide, it confers important
the increase of C-OH bonds detected by XPS at 72 h but the final structural changes that with time affect its composition and
reduction on C-OH concentration at 120 h. It is important to therefore its surface charge. Consequently, as time progresses and
mention that in our experiments it is apparent that C-O-C con- functional groups are loss, GO will become less and less hydro-
centration remained almost the same with even a slight increase in philic, thus favoring the formation of agglomerates and promoting
concentration as irradiance and time increased to 120 h and 72 mW/ the material to resurface. As this happens a higher level of irradi-
cm2 (from 21% for GO to 28%). This could be related to epoxy ring- ance will be received thus accelerating this reduction and
opening/closing reactions induced by OH groups generated by UV decomposition.
irradiation (Hatakeyama et al., 2017).
These changes in composition are expected to have a strong
impact on the stability of this material in suspension. The mea- 3.2. Microstructural changes
surement of the zeta potential (z) is a factor that describes this
stability, which according to the literature, values of z higher than The structural changes previously discussed were visible also by
30 or -30 mV correspond to stable graphene oxide suspensions TEM. Fig. 5a shows the TEM micrographs of as-produced GO, con-
(Hunter, 2013). Fig. 4 shows the effect of pH, irradiance and time on firming the formation of a single layer structure of approximately
z. It is observed that at any given pH the as-produced GO is stable in 25 mm in length with soft edges and distinctive wrinkles, particu-
water since all the pH tested had a zeta potential below -30 mV. larly at the edge (Fig. S3). This tortuosity is distinctive to GO due to
However, at pH¼3 it is observed that the z of most of the irradiated the double carbon vacancy (C2) defects which generates non-
GO crosses the stability line with values higher than -30 mV, while regular rings commonly 5-8-5 member rings, but also Stone-
the as-produced GO remains stable. Similarly, it is observed that at Wales defects formed by a pair of adjacent 5-7 member rings
pH values higher than 3, the stability of the irradiated materials increasing tension in carbon structure (Ji et al., 2013; Gengler et al.,
increases with pH. This effect is related with the concentration of 2013), besides the formation of carbons with sp (Chen et al., 2015)
ionizable species such as -OH and COOH, which are protonated hybridization results in the tortuosity of the carbon structure.
under acid conditions and deprotonated under alkaline pH Overall, most of the structure of as-produced GO is amorphous as
(Konkena and Vasudevan, 2012), since the carboxylic groups have can be seen in Fig. S3.
two pK values of 4.3 and 6.6, whereas the hydroxyl groups have a However, as the sample got irradiated with 74 mW/cm2 for 120 h,
pK value of 9.8 (Konkena and Vasudevan, 2012). Overall, the GO changed considerably as it can be seen in Figs. 5b and S4, where
structure of GO does not change, however its agglomeration is the distinctive single-layer structure of GO was no longer observed.
promoted under acid conditions (Konkena and Vasudevan, 2012; Instead, a large number of irregular particles with rough edges
Jiang et al., 2016; Chowdhury et al., 2015). between 20 nm and 8 mm in length were detected. At higher
By increasing the irradiance of GO from 37 to 74 mW/cm magnification, it was possible to identify that some of the particles
were in fact agglomerates of smaller particles of around 100 nm
Fig. 4. Zeta Potential of GO and GO irradiated for 72 h (UV3) and 120 h (UV5) at 37, 74 mW/cm2 with pH from 2 to 11 at room temperature.
W.R. Gallegos-Perez et al. / Chemosphere 249 (2020) 126160 7
Fig. 5. TEM micrographs of a) a layer of as-produced GO and irradiated GO with 74 mW/cm2 120 h, b) and c) bright and dark field image showing the presence of crystalline
structures within GO agglomerates. d) And e) high magnification micrograph showing the presence of ordered structures within GO particles. And.
(Fig. 5b) and that the irradiated graphene oxide layers were formation of smaller particles with sharp edges (Hatakeyama et al.,
severely defective at the edges and on the plane (Fig. 5d) as a results 2017,Mattevi et al., 2009).
of the decomposition process induced by UV irradiation (Ren et al.,
2018). These new structures showed clear evidence of higher 3.3. As(III) adsorption tests
crystallinity as can be observed in the dark field images (Fig. 5c)
where the bright sections correspond to crystalline structures with Previous studies have shown that As(III) adsorbs onto GO
similar orientation (Lo pez-Honorato et al., 2008). This increase in
through hydrogen bonds, the lone pairs of electrons from the
crystallinity was also evident in higher resolution TEM where epoxide group and p electron density of the sp (Pendolino and
graphene domains were now visible, albeit still in a random s-Arriagada and Toro-Labbe
Armata, 2017) structure (Corte , 2015;
arrangement (Fig. 5e). These results suggest that graphene oxide Reynosa-Martínez et al., 2020), resulting in a stronger As-GO
suffered partial reduction but a considerable degradation and interaction with GO with a higher degree of oxidation. Therefore,
8 W.R. Gallegos-Perez et al. / Chemosphere 249 (2020) 126160
4. Conclusions
Chen, X., Chen, B., 2016. Direct observation, molecular structure, and location of applications. Macromol. Res. 21 (3), 290e297.
oxidation debris on graphene oxide nanosheets. Environ. Sci. Technol. 50 (16), Li, R., Guiney, L.M., Chang, C.H., Mansukhani, N.D., Ji, Z., Wang, X., et al., 2018.
8568e8577. Surface oxidation of graphene oxide determines membrane damage, lipid
Chen, M.L., Sun, Y., Huo, C.B., Liu, C., Wang, J.H., 2015. Akaganeite decorated gra- peroxidation, and cytotoxicity in macrophages in a pulmonary toxicity model.
phene oxide composite for arsenic adsorption/removal and its proconcentra- ACS Nano 12 (2), 1390e1402.
tion at ultra-trace level. Chemosphere 130, 52e58. Li, M., Zhu, J., Wang, M., Fang, H., Zhu, G., Wang, Q., 2019. Exposure to graphene
Choong, T.S.Y., Chuah, T.G., Robiah, Y., Gregory Koay, F.L., Azni, I., 2007. Arsenic oxide at environmental concentrations induces thyroid endocrine disruption
toxicity, health hazards and removal techniques from water: an overview. and lipid metabolic disturbance in Xenopus laevis. Chemosphere 236, 124834
Desalination 217 (1e3), 139e166. 1e9.
Chowdhury, I., Mansukhani, N.D., Guiney, L.M., Hersam, M.C., Bouchard, D., 2015. Liu, S., Zeng, T.H., Hofmann, M., Burcombe, E., Wei, J., Jiang, R., et al., 2011. Anti-
Aggregation and stability of reduced graphene oxide: complex roles of divalent bacterial activity of graphite , graphite oxide , graphene oxide , and reduced
cations, pH, and natural organic matter. Environ. Sci. Technol. 49 (18), graphene Oxide : membrane and oxidative stress. ASC Nano 5 (9), 6971e6980.
10886e10893. pez-Honorato, E., Meadows, P.J., Xiao, P., Marsh, G., Abram, T.J., 2008. Structure
Lo
Corte s-Arriagada, D., Toro-Labbe , A., 2015. Improving As(III) adsorption on gra- and mechanical properties of pyrolytic carbon produced by fluidized bed
phene based surfaces: impact of chemical doping. Phys. Chem. Chem. Phys. 17 chemical vapor deposition. Nucl. Eng. Des. 238 (11), 3121e3128.
(18), 12056e12064. Marcano, D.C., Kosynkin, D.V., Berlin, J.M., Sinitskii, A., Sun, Z., Slesarev, A., et al.,
Dai, J., Wang, G., Ma, L., Wu, C., 2015. Study on the surface energies and dis- 2010. Improved synthesis of graphene oxide. ACS Nano 4 (8), 4806e4814.
persibility of graphene oxide and its derivatives. J. Mater. Sci. 50 (11), Mattevi, C., Eda, G., Agnoli, S., Miller, S., Mkhoyan, K.A., Celik, O., et al., 2009. Evo-
3895e3907. lution of electrical, chemical, and structural properties of transparent and
Díez-Betriu, X., Alvarez-García,
S., Botas, C., Alvarez, P., S
anchez-Marcos, J., Prieto, C., conducting chemically derived graphene thin films. Adv. Funct. Mater. 19 (16),
et al., 2013. Raman spectroscopy for the study of reduction mechanisms and 2577e2583.
optimization of conductivity in graphene oxide thin films. J. Mater. Chem. C 1 Mohandoss, M., Gupta, S Sen, Nelleri, A., Pradeep, T., Maliyekkal, S.M., 2017. Solar
(41), 6905e6912. mediated reduction of graphene oxide. RSC Adv. 7 (2), 957e963.
Ding, Y.H., Zhang, P., Zhuo, Q., Ren, H.M., Yang, Z.M., Jiang, Y., 2011. A green approach Mulyana, Y., Uenuma, M., Ishikawa, Y., Uraoka, Y., 2014. Reversible oxidation of
to the synthesis of reduced graphene oxide nanosheets under UV irradiation. graphene through ultraviolet/ozone treatment and its non- thermal reduction
Nanotechnology 22 (21). through ultraviolet irradiation. J. Phys. Chem. C 118, 27372e27381.
Frisch, M.L., 2009. Gaussian 09, Revision A.02. Gaussian, Inc, Wallingford CT. Nemati, F., Zare-Dorabei, R., Hosseini, M., Ganjali, M.R., 2018. Fluorescence turn-on
Ganguly, A., Sharma, S., Papakonstantinou, P., Hamilton, J., 2011. Probing the ther- sensing of thiamine based on Arginine e functionalized graphene quantum
mal deoxygenation of graphene oxide using high resolution in situ X ray based dots (Arg-GQDs): central composite design for process optimization. Sensor.
spectroscopies. J. Phys. Chem. C 115 (34), 17009e17019. Actuator. B Chem. 255, 2078e2085.
Gengler, R.Y.N., Badali, D.S., Zhang, D., Dimos, K., Spyrou, K., Gournis, D., et al., 2013. Nyangiwe, N.N., Khenfouch, M., Thema, F.T., Nukwa, K., Kotsedi, L., Maaza, M., 2015.
Revealing the ultrafast process behind the photoreduction of graphene oxide. Free-Green synthesis and dynamics of reduced graphene sheets via sun light
Nat. Commun. 4 (May), 1e5. irradiation. Graphene 54e61, 04(03).
Glockler, G., 1958. Carboneoxygen bond energies and bond distances. J. Phys. Chem. Pendolino, F., Armata, N., 2017. In: Graphene Oxide in Environmental Remediation
62 (9), 1049e1054. Process. 1 St. Kacprzyk J, Polish Academy of Science, Systems Research Instutute.
Hatakeyama, K., Awaya, K., Koinuma, M., Shimizu, Y., Hakuta, Y., Matsumoto, Y., Springer International Publishing AG, Warsaw, Poland, pp. 1e56.
2017. Production of water-dispersible reduced graphene oxide without stabi- Pinedo, J.L.V., Mireles, F.G., Ríos, C.M., Quirino, L.L.T., Da vila, J.I.R., 2006. Spectral
lizers using liquid-phase photoreduction. Soft Matter 13 (45), 8353e8356. signature of ultraviolet solar irradiance in Zacatecas. Geofisc. Int. 45 (4),
Herath, I., Vithanage, M., Bundschuh, J., Maity, J.P., Bhattacharya, P., 2016. Natural 263e269.
arsenic in global groundwaters: distribution and geochemical triggers for Ren, X., Li, J., Chen, C., Gao, Y., Chen, D., Su, M., et al., 2018. Graphene analogues in
mobilization. Curr Pollut Reports 2 (1), 68e89. aquatic environments and porous media: dispersion, aggregation, deposition
Hou, W.C., Chowdhury, I., Goodwin, D.G., Henderson, W.M., Fairbrother, D.H., and transformation. Environ Sci Nano 5 (6), 1298e1340.
Bouchard, D., et al., 2015. Photochemical transformation of graphene oxide in Reynosa-Martínez, A.C., Navarro-Tovar, G., Gallegos, W.R., Rodríguez-Melendez, H.,
sunlight. Environ. Sci. Technol. 49 (6), 3435e3443. Torres-Cadena, R., Mondrago n-Solo
rzano, G., et al., 2020. Effect of the degree of
Hu, X., Zhou, M., Zhou, Q., 2015. Ambient water and visible-light irradiation drive oxidation ofgraphene oxide on As(III) adsorption. J. Hazard Mater. 384 (15),
changes in graphene morphology, structure, surface chemistry, aggregation, 1e11. https://doi.org/10.1016/j.jhazmat.2019.121440, 121440, In press.
and toxicity. Environ. Sci. Technol. 49 (6), 3410e3418. Silva, KKH De, Huang, H.-H., Joshi, R.K., Yoshimura, M., 2017. Chemical reduction of
Hunter, R.J., 2013. Zeta Potential in Colloid Science: Principles and Applications, vol. graphene oxide using green reductants. Carbon N Y 119, 190e199.
2. Academic press. Soni, R., Shukla, D.P., 2019. Synthesis of fly ash based zeolite-reduced graphene
Ji, T., Hua, Y., Sun, M., Ma, N., 2013. The mechanism of the reaction of graphite oxide oxide composite and its evaluation as an adsorbent for arsenic removal. Che-
to reduced graphene oxide under ultraviolet irradiation. Carbon N Y 54, mosphere 219, 504e509.
412e418. Weinhold, F., Glendening, E.D., 1996. NBO3.1. Wisconsin.
Jiang, Y., Raliya, R., Fortner, J.D., Biswas, P., 2016. Graphene oxides in water: corre- Xie, X., Ma, X., Guo, L., Fan, Y., Zeng, G., Zhang, M., et al., 2019. Novel magnetic multi-
lating morphology and surface chemistry with aggregation behavior. Environ. templates molecularly imprinted polymer for selective and rapid removal and
Sci. Technol. 50 (13), 6964e6973. detection of alkylphenols in water. Chem. Eng. J. 357, 56e65.
Konkena, B., Vasudevan, S., 2012. Understanding aqueous dispersibility of graphene Zhang, C.Z., Li, T., Yuan, Y., Xu, J., 2016. An efficient and environment-friendly
oxide and reduced graphene oxide through pKa measurements. J. Phys. Chem. method of removing graphene oxide in wastewater and its degradation
Lett. 3 (7), 867e872. mechanisms. Chemosphere 153, 531e540.
Konkena, B., Vasudevan, S., 2015. Engineering a water-dispersible, conducting, Zhang, M., Ma, X., Li, J., Huang, R., Guo, L., Zhang, X., et al., 2019. Enhanced removal
photoreduced graphene oxide. J. Phys. Chem. C 119 (11), 6356e6362. of As(III) and As(V) from aqueous solution using ionic liquid-modified magnetic
Larios, R., Ferna ndez-Martínez, R., Silva, V., Loredo, J., Rucandio, I., 2012. Arsenic graphene oxide. Chemosphere 234, 196e203.
contamination and speciation in surrounding waters of three old cinnabar Zhu, Y., Murali, S., Cai, W., Li, X., Suk, J.W., Potts, J.R., et al., 2010. Graphene and
mines. J. Environ. Monit. 14 (2), 531e542. graphene oxide: synthesis, properties, and applications. Adv. Mater. 22 (35),
Li, H., Bubeck, C., 2013. Photoreduction processes of graphene oxide and related 3906e3924.