You are on page 1of 10

Ceramics International 47 (2021) 10113–10122

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Effects of preparation temperature on production of graphene oxide by


novel chemical processing
Rashad Al-Gaashani a, b, *, Yahya Zakaria a, One-Sun Lee a, Janarthanan Ponraj a,
Viktor Kochkodan a, Muataz A. Atieh c, d
a
Qatar Environment and Energy Research Institute (QEERI), Hamad Bin Khalifa University (HBKU), Qatar Foundation, P.O. Box: 34110, Doha, Qatar
b
Physics Department, Faculty of Education, Thamar University, Dhamar, Republic of Yemen
c
College of Engineering, University of Sharjah, PO Box 27272, Sharjah, United Arab Emirates
d
Desalination Research Group, Research Institute of Sciences and Engineering, University of Sharjah, Sharjah, P.O. Box 27272, United Arab Emirates

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

Keywords: The effect of preparation temperature on the synthesis of high quality graphene oxide (GO) by a chemical
Graphene oxide oxidation method has been studied. GO samples have been produced from graphite using a blend of sulfuric and
Graphite phosphoric acids with a volume ratio of 4:1, where potassium permanganate and hydrogen peroxide were
Chemical method
selected as oxidizing agents. The temperature for GO preparation was changed over a range of 20–85 ◦ C. The
XPS
Optical properties
synthesized GO samples have been characterized by X-ray diffraction, scanning electron microscopy, electron
Atomistic model dispersive spectroscopy, transmission electron microscopy, Fourier transform infrared spectroscopy, X-ray
photoelectron spectroscopy (XPS), and ultraviolet–visible spectroscopy methods. It was shown that high quality
GO can be synthesized at 20 ◦ C but the purity of GO increased with the increase of the preparation temperature.
The atomistic models of GO sheets prepared at different temperatures and according to the stoichiometric ratio of
the functional groups found by XPS have been developed. Based on the developed models it was found that GO
prepared at higher temperatures had more holes in the sheets supported with experimental images.

1. Introduction graphite flakes using three modified Hummers methods. We used a


mixture of three acids, namely: sulfuric acid (H2SO4), phosphoric acid
Graphene oxide (GO) is a single layer of graphite oxide having many (H3PO4) and nitric acid (HNO3) as intercalating agents and potassium
functional groups (complex of carbon, oxygen, and hydrogen) enable permanganate (KMnO4) and hydrogen peroxide (H2O2) as oxidizing
GO has various chemical, physical, and optical properties [1]. GO is agents. The synthesized GO samples have revealed high concentration of
easily synthesized from graphite (top-down method) using strong oxi­ the oxygen functional groups [10].
dizers such as sulfuric acid, nitric acid and hydrogen peroxide. There are GO nanosheets have hydrophilic edges and large surface area which
two major methods for synthesizing GO: “top-down” (Brodie and is expected to be used as an excellent adsorbent for removing pollutants
Hummer technique) in which strong oxidizers are used and “bottom-up” and heavy metals [11,12]. GO and GO-composite materials have
(Tang–Lau process) in which only glucose is used as a material source attracted significant interest due to their prospective and numerous
[2]. In 1859, GO was prepared for the first time by Brodie using the applications in water treatment and purification [13,14], GO-assisted
top-down approach [3]. About a century later, Hummers and Offeman membranes [15–18], electronic and optical devices [19–21], hydrogen
prepared GO from flake graphite using a mixture of sodium nitrate, storage [22–26], coatings [27,28], batteries [29–31], lenses [32,33],
sulfuric acid, and potassium permanganate [4]. Since 1958, Hummers cellular imaging and drug delivery [34], medicine and biotechnology
and Offeman method has become the most popular technique to prepare [35,36], and graphene production [37,38]. Graphene can easily be
GO. Thereafter, several modified Hummers’ methods have been re­ produced by reducing of GO which is one promising material for the
ported [5–9]. In a previous study, we presented the findings of GO and mass manufacture of graphene [39].
reduced-GO (rGO) synthesized by chemical oxidation process from It should be noted that numerous structural models of GO have been

* Corresponding author. Qatar Environment and Energy Research Institute (QEERI), Hamad Bin Khalifa University (HBKU), Qatar Foundation, P.O. Box: 5825,
Doha, Qatar.
E-mail address: ralgaashani@hbku.edu.qa (R. Al-Gaashani).

https://doi.org/10.1016/j.ceramint.2020.12.159
Received 14 September 2020; Received in revised form 16 December 2020; Accepted 16 December 2020
Available online 18 December 2020
0272-8842/© 2020 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

Table 1
XPS survey and high resolution (HR) scan fitting data for C and O of GO samples prepared at 20, 40, 60 and 85 ◦ C.
Sample Name Peak binding Atomic % (Survey) Chemical states Chemical states Peak binding Atomic %
energy (eV) defined from the survey (HR scans fitting) energy (eV) (HR scans fitting)

GO-20 P 2p 134.73 0.61 PO4 Phosphate 0.61


S 2p 169.42 1.82 SO3 Sulfate 1.82
C 1s 285.08 74.92 C–C/C– –C/C–H 284.48 57.50
C–O 286.32 11.89
C=O 288.00 4.19
O–C––O 289.20 1.35
O 1s 532.72 22.65 O–C––O* 531.00 1.36
C=O 531.56 4.21
C–OH 532.72 6.62
C–O–C 533.50 2.59
PO4 530.90 2.41
SO3 532.15 5.46
Total 100 Total 100
GO-40 P 2p 134.29 0.17 PO4 Phosphate 0.17
S 2p 168.8 0.85 SO3 Sulfate 0.85
C 1s 286.01 65.67 C–C/C– –C/C–H 284.50 29.10
C–O 286.60 31.90
C=O 288.04 3.35
O–C––O 289.20 1.32
O 1s 532.54 33.31 O–C––O* 531.08 1.31
C=O 531.60 3.34
C–OH 532.33 19.29
C–O–C 533.10 6.16
PO4 530.92 0.66
SO3 532.16 2.55
Total 100 Total 100
GO-60 S 2p 168.54 1.69 SO3 Sulfate 1.69
C 1s 285.79 61.24 C–C/C– –C/C–H 284.47 23.96
C–O 286.52 30.01
C=O 287.95 5.92
O–C––O 288.97 1.35
O 1s 531.98 37.06 O–C––O* 530.66 1.34
C=O 531.40 5.93
C–OH 532.34 19.60
C–O–C 533.18 5.13
SO3 532.06 5.07
Total 100 Total 100
GO-85 S 2p 168.61 0.47 SO3 Sulfate 0.47
C 1s 285.96 63.33 C–C/C– –C/C–H 284.55 15.82
C–O 286.44 32.48
C=O 287.71 12.28
O–C––O 288.80 2.74
O 1s 532.00 36.20 O–C––O* 531.04 2.72
C=O 532.01 12.27
C–OH 532.53 7.47
C–O–C 533.29 12.31
SO3 532.20 1.43
Total 100 Total 100

suggested, but those are still controversial. The most logical and wide 2. Experiments
spread model describes the GO structure as a random distribution of
oxidized areas due to the oxygenated functional groups, jointed with 2.1. Materials
non-oxidized regions wherein most of the carbon atoms preserve sp2
hybridization [40]. A proposed defect such as holes stabilized by the Natural graphite flakes (10 mesh, 99.9% purity) was purchased from
adjacent carbonyl group and the hemi-acetal due to the loss of one Alfa Aesar, Germany. Hydrochloric acid (HCl) 37 wt % was purchased
carbon atom, ethers and topological defects due to the from Sigma-Aldrich, USA. Sulphuric acid (H2SO4) (95%) and hydrogen
oxidation-reduction process have been reported [39,41,42]. peroxide (H2O2) (30 wt %) was purchased from VWR Chemicals, France.
Here, we report the effect of reaction temperature on the synthesis of Phosphoric acid (H3PO4) 85 wt % and potassium permanganate
high quality GO prepared by a modified-Hummers’ method which we (KMnO4) (≥99%) was purchased from Honeywell Fluka, India. The
recently proposed [10]. This method is environmentally friendly and aqueous solutions were prepared in deionized water (DIW) (18.2 MΩ
safer than traditional approaches since the formation of explosive or cm− 1).
toxic gaseous products is avoided. The GO samples were prepared by
varying the reaction temperature between 20 ◦ C and 85 ◦ C to explore the
effect of reaction temperature on GO structure and its characteristics. 2.2. Preparation of GO at different temperatures
The atomistic models of GO sheets were suggested according to the
functional group compositions revealed by XPS high resolution scan 100 ml mixture of two acids, namely:80 ml H2SO4 and 20 ml H3PO4
data fitting, and it was found that the surface structure of GO samples were placed in a 500 ml glass conical flask and stirred in an ice bath for
prepared at different temperatures notably varies. 10 min. Then, 1 g of graphite flakes and 9 g of KMnO4 were slowly
dispersed into the acid mixture under regular magnetic stirring for 10
min. The flask was then transferred to an oil bath and the mixture was

10114
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

energy is 20 eV for the high-resolution spectra and 100 eV for the survey
spectra. Binding energy values of XPS lines are calibrated using the C 1s
main peak at 284.5 eV as a reference. Light absorption spectra were
collected at 20 ◦ C in the wavelength range of 190–1000 nm using a
JASCO UV–Vis–NIR spectrophotometer (V-650, single monochromator
with low stray light).

2.4. Development of atomistic GO models

Atomistic GO models that obey the chemical composition suggested


by XPS and high resolution scan fitting data were generated. Since the
chemical composition of GO samples varies with the temperature, we
generated four atomistic GO models for difference preparation temper­
atures (20, 40, 60, and 85 ◦ C) where the relative amount of the car­
boxylic, hydroxyl, carbonyl and epoxy groups for each GO sample
follows the stoichiometric ratio shown in Table 1. The functional groups
were added on the randomly chosen carbon of a bare graphene and the
connectivity of carbon atoms was adjusted to obey the coordination
number of 4. In our GO model generation algorithm, the addition of
more than two functional groups on the same carbon atom is avoided.
After building the model, each GO sheet is geometry optimized for 1000
steps using the conjugate gradient method in vacuo with empirical force
field parameters [43]. The conformer obtained from

Fig. 1. XRD patterns of graphite (a) and GO samples synthesized using a


mixture of 80 ml H2SO4 and 20 ml H3PO4 at 20 (RT) (b), 40 (c), 60 (d), and
85 ◦ C (e).

stirred at 20, 40, 60, and 85 ◦ C for 1 h, separately. 100 ml of cool DIW
was then added gradually and the mixture was stirred under the same
condition for 1 h. Then, the flask was put in an ice bath and 120 ml of
DIW and 15 ml of H2O2 were slowly added and the mixture was stirred
for 50 min. After the reaction was completed, the produced slurry was
washed using DIW with 20% HCl solution and centrifuged using
Eppendorf centrifuge 5810R at 5000 rpm for 20 min. The prepared GO
was then washed again with DIW for several times until the pH of the
washing solution was increased to about 7. The washed GO samples
were finally dried in a furnace (BINDER drying and heating chambers
ED115) at 95 ◦ C for 24 h under air.

2.3. Characterization of GO samples

X-ray diffraction (XRD) patterns of the prepared GO samples were


measured with a Bruker D8 Advance X-Ray diffract meter using Cu-Kα
radiation source (λ = 0.15418 nm). High resolution transmission elec­
tron microscopy (HRTEM) (FEI Talos) and field emission scanning
electron microscopy (SEM) (JSM-7800F) were used to study the
morphology of the prepared GO samples. Fourier transform infrared
(FT-IR) spectra were measured using a FT-IR spectrometer (Thermo
Fisher scientific NICOLET iS50 FT-IR). Chemical surface analysis was
conducted by X-ray Photoelectron Spectroscope (XPS) method (Thermo Fig. 2. SEM images of GO samples synthesized at solvent temperature of 20 ◦ C
Fisher Scientific, Escalab 250Xi) where the X-ray source is mono­ (a and b), 40 ◦ C (c and d), 60 ◦ C (e and f), and 85 ◦ C (g and h). The inset in
chromatic Al K alpha and the X-ray energy is 1486.68eV. XPS pass Fig. 2 (a) shows a low magnification SEM image (10μm) of GO multilayers.

10115
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

3. Results and discussion

3.1. Structural studies of the prepared samples

Fig. 1 (a) displays XRD pattern of graphite used as a raw material to


prepare GO samples. The diffraction peaks observed at Bragg angle (2θ)
of 26.50 and 54.64◦ represent, respectively, (002) and (004) planes of
the hexagonal structure of graphite (ICDD card No. 00-056-0159 with a
lattice constant equals to a: 2.5 Å c: 6.7 Å volume: 35.2 Å3 and space
group P63/mmc) (Fig. 1 a). Fig. 1 shows XRD patterns of GO synthesized
using acid mixture of 80 ml H2SO4 and 20 ml H3PO4 (volume ratio 4:1)
at a solvent temperature of 20 (b), 40 (c), 60 (d), and 85 ◦ C (e). The
diffraction peak detected at 2θ of about 11 ◦ can be appropriated to the
(001) plane of the hexagonal crystalline structure of GO (ICDD card No.
00-065-1528 with a: 2.50 Å c: 9.11 Å Volume: 47.40 Å3 c/a: 3.718 and
space group: P31 m). The intensity of the peak (001) of graphene oxide
phase increased with increasing the reaction temperature. However, the
intensity of the peak (002) of graphite phase reduced by rising solvent
temperature. The samples prepared at 20, 40, and 60 ◦ C showed two
phases: graphite and GO. On the other hand, the pure phase of GO was
obtained at 85 ◦ C (Fig. 1 (e)). A shift of the XRD patterns towards higher
2θ was also observed with the increase of the solvent temperature from
20 to 85 ◦ C, suggesting a decrease of the lattice parameters. The results
of XRD showed that GO can be synthesized at 20 ◦ C and the purity of GO
increased with the rising of the solvent temperature.

3.2. Morphological properties of the prepared GO samples

Morphological properties of the synthesized samples have been


investigated by SEM and TEM as shown in Fig. 2 and Fig. 3; respectively.
SEM images show the flat morphology of the prepared GO samples
bearing some wrinkles (Fig. 2). The inset image in Fig. 2 (a) displays that
the GO-graphite structure is formed by several layers/flakes arranged in
stacks. This is due to the oxidation of the material, and usually the
greater the degree of oxidation; the higher are the spaces between the
layers. As seen in Fig. 2 the GO samples prepared at different reaction
temperatures show flat sheet morphology with a wrinkled appearance
with several folds at higher temperatures. The wrinkles and folds in­
crease with rising temperature due to an increase of a content of func­
tional groups and defects. On the other hand, the size of the wrinkles and
folds decreases and becomes smoother with an increase in temperature.
The difference in the sheets’ morphology on four prepared samples is
probable caused by the progressing of oxidation processes with tem­
perature. In the same batch of as-prepared GO at 20 ◦ C, among the
majority of normally oxidized GO sheets, there were a few number of
sheets that have been slightly damaged by oxidation. Those damages/
defects increase with rising temperature as shown in Fig. 2(a–h). The
difference in sheet morphologies is probably because the change in
Fig. 3. TEM images of GO samples prepared using the acid mixture of 80 ml diffusion rates at different temperatures. This causes the differences in
H2SO4 and 20 ml H3PO4 at 20 ◦ C (a and b), 40 ◦ C (c and d), 60 ◦ C (e and f), and the degrees of defects and oxidation [48].
85 ◦ C (g and h). The inset in Fig. 3 (h) shows a high resolution TEM image (20
The TEM images (Fig. 3) show a sheet/flake like morphology of the
nm) of GO wrinkles.
prepared GO samples with different transparencies, which increases by
raising the preparation temperature. Fig. 3(a–h) shows that the number
geometry-optimization was submitted for a short molecular dynamics of layers of the GO patches reduces with increasing the preparation
(MD) simulation (2 ps in NVT ensemble) to relax the structure further. temperature as testified by XRD (Fig. 1). However, the patch dimensions
The dimension of the system is 40 × 40 × 40 Å3, and the periodic decrease with rising the preparation temperature.
boundary conditions are applied. During the optimization, the position
of sp2 carbon atoms of GO sheet is fixed to the initial planar position
3.3. Compositional analysis of the prepared samples
with harmonic constraint of 10 kcal/mol/Å2 to maintain the planarity of
the sheet. With the particle-mesh Ewald method, full electrostatics was
The functional groups on the surfaces of GO samples prepared at
employed with a 1 Å grid width [44]. All of the models were generated
different temperatures were determined by FT-IR spectroscopy. The FT-
using QGO Builder [45,46] and the simulations were performed by
IR spectra of GO prepared at 20, 40, 60 and 85 ◦ C are shown in Fig. 4 (a),
nanoscale molecular dynamics (NAMD) [47].
(b), (c) and (d), respectively. The broad peaks spread at around 3435
cm− 1 result from O–H stretching vibrations are observed for all GO
samples. The carbonyl peaks (C– – O) are observed at about 1720 cm− 1
− 1
(Figs. 4 b) and 1725 cm (Fig. 4 a, c, and d). However, the aromatic

10116
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

Fig. 4. FT-IR spectra of GO synthesized using the acid mixture of 80 ml H2SO4 and 20 ml H3PO4 at 20 (RT) (a), 40 (b), 60 (c), and 85 ◦ C (d).

composition and the present functional groups using the chemical state
analysis. The elemental composition of all samples has been revealed by
the survey scan from 0 to 1350 eV as shown in Fig. 5. It showed the
presence of mainly C and O as well as trace amounts of S and P. The
chemical states of C have been identified by the fitting of high-resolution
spectra C 1s and it shows the presence of C–C/C– – C/C–H, C–O, C– – O and
O–C– – O chemical states. On the other hand, the fitting of high resolution
O1s shows the presence of O–C– – O, C–– O, C–O–H, C–O–C and SO2−
3
3−
species. PO4 has been found in two samples GO-20 and GO-40.
The high-resolution spectra deconvolution of C 1s for all samples are
shown in Fig. 6, where the C 1s chemical states of GO prepared at 20 ◦ C
are shown in Fig. 6 (a). It showed the presence of mainly C and O as well
as trace amounts of S and P. The chemical states of C have been iden­
tified by the fitting of high-resolution spectra C 1s and it shows the
presence of C–C/C– – C/C–H (284.47–284.55 eV), C–O (286.32–286.71
eV), C– – O (287.71–288.04 eV) and O–C– – O (288.80–289.20 eV)
chemical states. On the other hand, the fitting of high resolution O 1s
shows the presence of O–C– – O* (530.66–531.08 eV), C– –O
(531.40–532.01 eV), C–O–H (532.33–532.72 eV), C–O–C
(533.10–533.50 eV) and SO2− 3 (532.06–532.20 eV). PO3− 4
(530.90–530.92 eV) has been found in two samples GO-20 and GO-40
(see Table 1).
Fig. 5. XPS survey spectra of graphite (a) and GO samples synthesized using The high-resolution spectra deconvolution of C 1s for all samples are
the acid mixture of 80 ml H2SO4 and 20 ml H3PO4 at 20 (RT) (b), 40 (c), 60 (d), shown in Fig. 6, where the C 1s chemical states of GO prepared at 20 ◦ C
and 85 ◦ C (e).
are shown in Fig. 6 (a). The chemical state ratios are respectively
76.75% (C–C/C– – C/C–H), 15.87% (C–O), 5.59% (C– – O) and 1.80%
– C) stretching groups are located at 1725 cm− 1for all samples. The
(C– (O–C– – O). Fig. 6 (b) shows the C 1s chemical states of GO prepared at
peaks at about 1385 cm− 1are due to the O–H deformations of the C–OH 40 ◦ C. The chemical state ratios are respectively 44.31% (C–C/C– – C/
groups. Finally, the (C–O–C) stretching vibrating mode at 1040–1060 C–H), 48.58% (C–O), 5.10% (C– – O) and 2.01% (O–C– – O). Fig. 6 (c)
cm− 1are only detected for the GO samples prepared at 40, 60, and 85 ◦ C. shows the C1s chemical states of GO prepared at 60 ◦ C. The chemical
The functional groups present in GO samples confirm their hydrophilic state ratios are respectively 39.13% (C–C/C– – C/C–H), 49.00% (C–O),
character and these findings are consistent with the previous studies 9.67% (C– – O) and 2.20% (O–C– – O). Fig. 6 (d) shows the C1s chemical
[10,49–51]. states of GO prepared at 85 ◦ C. The chemical state ratios are respectively
The surface chemical studies of the GO samples have been conducted 24.98% (C–C/C– – C/C–H), 51.29% (C–O), 19.39% (C– – O) and 4.33%
using X-ray photoelectron spectroscopy (XPS) to identify the elemental (O–C– – O). All temperature treated samples show chemical states

10117
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

Fig. 6. High resolution XPS spectra in C 1s region of GO prepared samples using the acid mixture of 80 ml H2SO4 and 20 ml H3PO4 at 20 (a), 40 (b), 60 (c), and
85 ◦ C (d).

Fig. 7. High resolution spectra in O 1s region of GO samples prepared using the acid mixture of 80 ml H2SO4 and 20 ml H3PO4 at 20 (a), 40 (b), 60 (c), and 85 ◦ C (d).

distributions similar to GO samples which were reported in a previous is shown in Fig. 7, where the chemical states of GO prepared at 20 ◦ C are
article [10]. shown in (a). The chemical state ratios are respectively 10.65% (PO4),
The high-resolution spectra deconvolution of O 1s for all GO samples 5.99% (O–C– – O*), 18.60% (C– – O), 24.10% (SO3), 29.21% (C–OH) and

10118
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

Table 2 of GO-85 shows a stronger oxidation compared to other samples. The


Illustrates the efficiency of the chemical process as a function of the temperature C–C/C–– C/C–H chemical state is decreased dramatically compared to
with respect to the material ratio (produced materials: graphite) (yield), reaction GO-20 and the C–O has increased even more than GO-40 and GO-60.
time, and O/C atomic ratio. The C– – O chemical state increased percentage significantly and is
GO vs T ◦ C Reaction time (h) Yield % O% C% O/C ratio almost 4 times compared to GO-20 and double the percentage of GO-60.
GO at 20 ◦
C 3 91.45 22.65 74.92 0.30 The O–C– – O chemical state increased also significantly compared to GO-
GO at 40 ◦
C 3 69.21 33.31 65.67 0.51 20 and it doubled the percentage of GO-60. Table 1 summarizes the
GO at 60 ◦
C 3 54.15 37.06 61.24 0.60 percentages of elemental composition and all chemical species in each
GO at 85 ◦
C 3 47.12 36.2 66.33 0.55 sample.
Table 2 shows the efficiency of the chemical process as a function of
the temperature, amount of produced materials (GO-rGO-graphite) with
respect to the initial graphite quantity (1 g), and time required for ma­
terial (3 h). The yield obtained at 20 ◦ C is high (91.45%) due to most of it
has graphite characteristics as shown in XRD, XPS, SEM and TEM results,
but the yield of high purity GO obtained at 85 ◦ C is low (47.12%). The
obtained yield of GO decreases but GO purity increases with an increase
of the preparation temperature. Preparation GO samples at 60 and 85 ◦ C
causes a relatively large increase in O/C ratio, suggesting that more
oxygen-containing groups are formed on the graphene sheets compared
to the samples prepared at 20 and 40 ◦ C (Table 2).
Fig. 8 illustrates TEM images of GO samples, which were converted
to rainbow pseudo-color images by using TEM imaging analysis soft­
ware. Fig. 8 (a) and (c) shows the clear atomic distribution of the gra­
phene sheets and visible GO crystalline lattice with very low number of
missing atoms. The orientation change around the image might indicate
the presence of few “interconnected sheets”, this has been reported
previously [52]. An increase in the reaction temperature from 20 to
85 ◦ C facilitate the hole formation in the GO lattice as indicated by the
presence of blue spots in Fig. 8 (b) and (d). A higher fraction of holes and
topological defects shown in Fig. 8 (d) can be clearly attributed to more
efficient graphene disordering by oxidation at higher temperature [42].
It was previously reported that the disordered regions show oxygen
functionalities [52]. The areas are shown in green color in Fig. 8 (b) and
(d) might be due to the presence of amorphous adsorbates, so-called
adventitious carbon, but could also be oxidized carbons formed as a
result of the GO preparation or a mixture of both [53,54]. These findings
are consistent with the modeled GO sheets shown in Fig. 9.
Fig. 8. HRTEM images of GO samples synthesized using the acid mixture of 80
To visualize the GO structures obtained at different temperatures, we
ml H2SO4 and 20 ml H3PO4 at 20 ◦ C (a) and 85 ◦ C (c) and the relevant rainbow
pseudo-color images (b) and (d), respectively. The rainbow pseudo-colors,
generated models of GO that obey the chemical composition suggested
images illustrate the different features of GO structure. The defect free crys­ by XPS and high resolution scan fitting data. The relative amount of the
talline graphite/graphene regions represented in the red color. The holes due to carboxylic, hydroxyl, carbonyl, and epoxy groups was generated on a
oxidation process and incorporation of functional groups are presented in blue randomly chosen carbon atoms of a bare graphene according to the
colored areas. The black areas might be related to the topological defects. (For stoichiometric ratio shown in Table 1. After the addition of the func­
interpretation of the references to color in this figure legend, the reader is tional groups, the connectivity of carbon atoms of the sheet was adjusted
referred to the Web version of this article.) to obey the coordination number of 4. We found that we could not avoid
the generation of holes in the sheet to obey the coordination number and
11.45% (C–O–C). Fig. 7 (b) illustrates the chemical states of GO pre­ the stoichiometric ratio of the functional groups at the same time. Since
pared at 40 ◦ C. The chemical state ratios are respectively 1.97% (PO4), the GO sheets synthesized at higher temperatures (60 and 85 ◦ C) have a
3.93% (O–C– – O*), 10.04% (C– – O), 7.66% (SO3), 57.92% (C–OH) and higher composition of the carbonyl group than those at lower temper­
18.48% (C–O–C). Fig. 7(c) shows O 1s region of GO prepared at 60 ◦ C. atures (20 and 40 ◦ C), GO models for 60 and 85 ◦ C have more holes as
The chemical state ratios are respectively 3.62% (O–C– – O*), 15.99% shown in Fig. 9 (c) and (d).
(C–– O), 13.65% (SO3), 52.90% (C–OH) and 13.84% (C–O–C). Fig. 7(d)
illustrates O 1s region of GO prepared at 85 ◦ C. The chemical state ratios 3.4. Optical analysis of the prepared samples
are respectively 7.50% (O–C– – O*), 33.90% (C– – O), 3.96% (SO3),
20.63% (C–OH) and 34.00% (C–O–C). Fig. 10 illustrates the UV–vis absorption spectra of as-prepared
C 1s of GO prepared at 20 ◦ C (GO-20) shows an intense peak of C–C/ samples in ethanol in the wavelength range from 200 to 800 nm.
C–– C/C–H and smaller peaks (area decreasing respectively) of C–O, Many literatures reported that the UV–vis spectrum of pure GO exhibits
C–– O and O–C– – O. This revealed that the C–C/C– – C/C–H is still pre­ two main absorption bands: the main absorption pack at λmax around
dominant in the GO sample compared to the chemical states and that the 230 nm, attributed to π→π*/C– – C bonds, and a shoulder at about 300
oxidation is very small. C1s of GO prepared at 40 ◦ C (GO-40) shows a nm, corresponding n→π*/C– – O transition of C–
– O bond because of the
substantial increase of the C–O chemical state and decrease of C–C/ presence of the carbonyl groups [55–58]. The UV–vis absorption in­
C–– C/C–H. However, C– – O has slightly decreased and O–C– – O slightly tensity of GO is affected by a conjugative effect associated of chromo­
increased to GO-20. C1s of GO-60 shows similarly to GO-40 a substantial phore aggregation, which influences the π→π* plasmon peak. The
increase of C–O and even more decrease of C–C/C– – C/C–H chemical degree of oxidation of GO materials can be correlated with λmax of
state compared to GO-20. However, the C– – O has increase to almost adsorption in UV–vis spectrum. The more conjugation of C– – C bonds
double and the O–C– – O has slightly increased compared to GO-20. C1s and C–– O groups in oxidized GO sheets, the less energy is required for

10119
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

Fig. 9. Snapshots of GO sheet models where each model obeys the stoichiometric elemental ratio as per XPS data at 20 (a), 40 (b), 60 (c), and 85 ◦ C (d).

Fig. 10. Ultraviolet–visible (UV–vis) absorption spectra of GO-rGO samples Fig. 11. Absorption coefficient of GO-rGO samples prepared at different
prepared using the acid mixture of H2SO4 and H3PO4 at 20 (a), 40 (b), 60 (c), temperatures.
and 85 ◦ C (d).
absorption peak with a shoulder [56]. As seen in Fig. 10, the GO samples
the electronic transition and this will lead to a higher λmax value. The prepared at different temperatures (20, 40, 60 and 85 ◦ C) display strong
absorption peaks for GO-20 and GO-40 samples are at the lower wave­ absorption bands at λmax of 252, 254, 256, 258 nm, receptively [62] and
lengths (252 and 254 nm) compared to GO-60 and GO-85 samples (256 shoulders at about 300 nm. Red shift and hyperchromic shift in UV–vis
and 258 nm). These findings confirm that higher energy was required for absorption spectra of GO samples were observed with increased prepa­
the electronic transition in GO-20 and GO-40 samples as compared with ration temperatures. Such shifts is obviously due to changing the
GO-60 and GO-85. It could be concluded that he GO-60 and GO-85 chemical functional groups on the GO sheets [63]. Johari and Shenoy
samples were more oxidized with more functional groups on the basal reported a red shift in the GO absorption spectra due to oxygen in GO
planes as shown in XPS results [59]. As per the Beer-Lambert’s law [60], sheet is present primarily in the form of carbonyl groups. Moreover, the
higher absorption density in Fig. 10 is evidence of better dispersion of rise in carbonyl groups in the center of the GO sheet forms holes, that
the GO sample in the ethanol as for GO-85 and GO-60 samples [61]. causes the red shift of the absorption [64]. Wavelength shift from
It was previously reported that GO with few-layers (1–3 layers) ex­ around 230 nm to about 270 nm was reported [55,57,65–68] for
hibits a single absorption peak, whereas GO with multi-layers has a main reduced GO because of the increased electron concentration, structural

10120
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

ordering and consistent with the restoration of sp2 carbon atoms, which [12] G. Zhao, J. Li, X. Ren, C. Chen, X. Wang, Few-layered graphene oxide nanosheets as
superior sorbents for heavy metal ion pollution management, Environ. Sci.
is consistent with our results. Furthermore, our samples are a mixture of
Technol. 45 (2011) 10454–10462.
GO, reduced GO, and graphite as confirmed in XRD patterns (Fig. 1). As [13] Y. Chen, L. Chen, H. Bai, L. Li, Graphene oxide–chitosan composite hydrogels as
seen in Fig. 11, the absorption coefficient (α) for GO-rGO samples syn­ broad-spectrum adsorbents for water purification, J. Mater. Chem., A 1 (2013)
thesized at different temperatures increases with preparation 1992–2001.
[14] L. Liu, J. Liu, D.D. Sun, Graphene oxide enwrapped Ag3PO4 composite: towards a
temperature. highly efficient and stable visible-light-induced photocatalyst for water
purification, Catal. Sci. Technol. 2 (2012) 2525–2532.
[15] H.M. Hegab, L. Zou, Graphene oxide-assisted membranes: fabrication and potential
4. Conclusions
applications in desalination and water purification, J. Membr. Sci. 484 (2015)
95–106.
The effect of preparation temperature on the properties of GO syn­ [16] C. Xu, A. Cui, Y. Xu, X. Fu, Graphene oxide–TiO2 composite filtration membranes
thesized by the modified Hummers’ method was studied. The modified and their potential application for water purification, Carbon 62 (2013) 465–471.
[17] J. Yin, G. Zhu, B. Deng, Graphene oxide (GO) enhanced polyamide (PA) thin-film
method of GO synthesis is environmentally friendly and safer than the nanocomposite (TFN) membrane for water purification, Desalination 379 (2016)
conventional methods because the oxidizing agents are excluded in GO 93–101.
preparation to avoid the formation of explosive or toxic gaseous prod­ [18] A. Najjar, S. Sabri, R. Al-Gaashani, V. Kochkodan, M.A. Atieh, Enhanced fouling
resistance and antibacterial properties of novel graphene oxide-Arabic gum
ucts. The synthesis procedure includes only oxidation of graphite and polyethersulfone membranes, Appl. Sci. 9 (2019) 513.
washing/purification of GO. The characterizations of GO sheets pre­ [19] Y. Gao, H.L. Yip, K.S. Chen, K.M. O’Malley, O. Acton, Y. Sun, G. Ting, H. Chen, A.K.
pared at different temperatures (20, 40, 60, and 85 ◦ C) by XRD, SEM, Y. Jen, Surface doping of conjugated polymers by graphene oxide and its
application for organic electronic devices, Adv. Mater. 23 (2011) 1903–1908.
EDS, TEM, FT-IR, and XPS indicate the enhancement of the material [20] T. Musso, P.V. Kumar, A.S. Foster, J.C. Grossman, Graphene oxide as a promising
quality as the preparation temperature increases. It was also found that hole injection layer for MoS2-based electronic devices, ACS Nano 8 (2014)
the percentage of the oxygen containing functional groups increases 11432–11439.
[21] T.-Z. Shen, S.-H. Hong, J.-K. Song, Electro-optical switching of graphene oxide
significantly as the preparation temperature increases. With the atom­ liquid crystals with an extremely large Kerr coefficient, Nat. Mater. 13 (2014)
istic models of GO sheets, it was shown that GO sheets possess more 394–399.
defected structure with a larger number of holes as the preparation [22] L. Wang, K. Lee, Y.-Y. Sun, M. Lucking, Z. Chen, J.J. Zhao, S.B. Zhang, Graphene
oxide as an ideal substrate for hydrogen storage, ACS Nano 3 (2009) 2995–3000.
temperature increases. These findings agree well with HR-TEM images
[23] E.S. Cho, A.M. Ruminski, S. Aloni, Y.-S. Liu, J. Guo, J.J. Urban, Graphene oxide/
of GO samples obtained at different reaction temperatures. metal nanocrystal multilaminates as the atomic limit for safe and selective
hydrogen storage, Nat. Commun. 7 (2016) 1–8.
[24] S.H. Aboutalebi, S. Aminorroaya-Yamini, I. Nevirkovets, K. Konstantinov, H.K. Liu,
Declaration of competing interest Enhanced hydrogen storage in graphene oxide-MWCNTs composite at room
temperature, Adv. Energy Mater. 2 (2012) 1439–1446.
[25] Y. Chan, J.M. Hill, Hydrogen storage inside graphene-oxide frameworks,
The authors declare that they have no known competing financial Nanotechnology 22 (2011) 305403.
interests or personal relationships that could have appeared to influence [26] H. Zhou, J. Zhang, J. Zhang, X. Yan, X. Shen, A. Yuan, High-capacity room-
the work reported in this paper. temperature hydrogen storage of zeolitic imidazolate framework/graphene oxide
promoted by platinum metal catalyst, Int. J. Hydrogen Energy 40 (2015)
12275–12285.
Acknowledgments [27] B.P. Singh, B.K. Jena, S. Bhattacharjee, L. Besra, Development of oxidation and
corrosion resistance hydrophobic graphene oxide-polymer composite coating on
copper, Surf. Coating. Technol. 232 (2013) 475–481.
The authors would like to acknowledge the support of Qatar Envi­ [28] B. Ramezanzadeh, S. Niroumandrad, A. Ahmadi, M. Mahdavian, M.M. Moghadam,
ronmental and Energy Research Institute and Hamad Bin Khalifa Uni­ Enhancement of barrier and corrosion protection performance of an epoxy coating
versity. The authors would also like to thank the director of QEERI’s through wet transfer of amino functionalized graphene oxide, Corrosion Sci. 103
(2016) 283–304.
Core Labs Dr. Said Mansour as well as the following Core Labsmembers: [29] K. Fu, Y. Wang, C. Yan, Y. Yao, Y. Chen, J. Dai, S. Lacey, Y. Wang, J. Wan, T. Li,
Dr. Kamal Mroue (FTIR), Dr. Akshath Raghu Shetty (XRD), Mr. Graphene oxide-based electrode inks for 3D-printed lithium-ion batteries, Adv.
MohammedI. Helal and Mr. Mujaheed Pasha (SEM) for their charac­ Mater. 28 (2016) 2587–2594.
[30] Z.L. Wang, D. Xu, J.J. Xu, L.L. Zhang, X.B. Zhang, Graphene oxide gel-derived, free-
terization support.
standing, hierarchically porous carbon for high-capacity and high-rate
rechargeable Li-O2 batteries, Adv. Funct. Mater. 22 (2012) 3699–3705.
References [31] V. Chabot, D. Higgins, A. Yu, X. Xiao, Z. Chen, J. Zhang, A review of graphene and
graphene oxide sponge: material synthesis and applications to energy and the
environment, Energy Environ. Sci. 7 (2014) 1564–1596.
[1] K. Toda, R. Furue, S. Hayami, Recent progress in applications of graphene oxide for
[32] X. Zheng, B. Jia, H. Lin, L. Qiu, D. Li, M. Gu, Highly efficient and ultra-broadband
gas sensing: a review, Anal. Chim. Acta 878 (2015) 43–53.
graphene oxide ultrathin lenses with three-dimensional subwavelength focusing,
[2] L. Tang, X. Li, R. Ji, K.S. Teng, G. Tai, J. Ye, C. Wei, S.P. Lau, Bottom-up synthesis
Nat. Commun. 6 (2015) 8433.
of large-scale graphene oxide nanosheets, J. Mater. Chem. 22 (2012) 5676–5683.
[33] G. Cao, X. Gan, H. Lin, B. Jia, An accurate design of graphene oxide ultrathin flat
[3] B.C. Brodie, XIII. On the atomic weight of graphite, Phil. Trans. Roy. Soc. Lond.
lens based on Rayleigh-Sommerfeld theory, Opto-Electr. Adv. 1 (2018) 180012.
(1859) 249–259.
[34] X. Sun, Z. Liu, K. Welsher, J.T. Robinson, A. Goodwin, S. Zaric, H. Dai, Nano-
[4] W.S. Hummers Jr., R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc.
graphene oxide for cellular imaging and drug delivery, Nano Res. 1 (2008)
80 (1958), 1339–1339.
203–212.
[5] N. Zaaba, K. Foo, U. Hashim, S. Tan, W.-W. Liu, C. Voon, Synthesis of graphene
[35] Y. Wang, Z. Li, J. Wang, J. Li, Y. Lin, Graphene and graphene oxide:
oxide using modified hummers method: solvent influence, Procedia Eng. 184
biofunctionalization and applications in biotechnology, Trends Biotechnol. 29
(2017) 469–477.
(2011) 205–212.
[6] S.N. Alam, N. Sharma, L. Kumar, Synthesis of graphene oxide (GO) by modified
[36] X. Zhou, F. Liang, Application of graphene/graphene oxide in biomedicine and
hummers method and its thermal reduction to obtain reduced graphene oxide
biotechnology, Curr. Med. Chem. 21 (2014) 855–869.
(rGO), Graphene 6 (2017) 1–18.
[37] S.Y. Toh, K.S. Loh, S.K. Kamarudin, W.R.W. Daud, Graphene production via
[7] T. Chen, B. Zeng, J. Liu, J. Dong, X. Liu, Z. Wu, X. Yang, Z. Li, High throughput
electrochemical reduction of graphene oxide: synthesis and characterisation,
exfoliation of graphene oxide from expanded graphite with assistance of strong
Chem. Eng. J. 251 (2014) 422–434.
oxidant in modified Hummers method, J. Phys. Conf. Ser. 188 (2009), 012051,
[38] H. Zhang, Y. Miyamoto, Graphene production by laser shot on graphene oxide: an
https://doi.org/10.1088/1742-6596/188/1/012051.
ab initio prediction, Phys. Rev. B 85 (2012), 033402.
[8] R. Muzyka, M. Kwoka, Ł. Smędowski, N. Díez, G. Gryglewicz, Oxidation of graphite
[39] S. Eigler, C. Dotzer, A. Hirsch, Visualization of defect densities in reduced graphene
by different modified Hummers methods, N. Carbon Mater. 32 (2017) 15–20.
oxide, Carbon 50 (2012) 3666–3673.
[9] J. Chen, B. Yao, C. Li, G. Shi, An improved Hummers method for eco-friendly
[40] A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited,
synthesis of graphene oxide, Carbon 64 (2013) 225–229.
J. Phys. Chem. B 102 (1998) 4477–4482.
[10] R. Al-Gaashani, A. Najjar, Y. Zakaria, S. Mansour, M. Atieh, XPS and structural
[41] S. Eigler, C. Dotzer, F. Hof, W. Bauer, A. Hirsch, Sulfur species in graphene oxide,
studies of high quality graphene oxide and reduced graphene oxide prepared by
Chem. Eur. J. 19 (2013) 9490–9496.
different chemical oxidation methods, Ceram. Int. 45 (2019) 14439–14448.
[42] C. Gómez-Navarro, J.C. Meyer, R.S. Sundaram, A. Chuvilin, S. Kurasch,
[11] Y. Sun, D. Shao, C. Chen, S. Yang, X. Wang, Highly efficient enrichment of
M. Burghard, K. Kern, U. Kaiser, Atomic structure of reduced graphene oxide, Nano
radionuclides on graphene oxide-supported polyaniline, Environ. Sci. Technol. 47
Lett. 10 (2010) 1144–1148.
(2013) 9904–9910.

10121
R. Al-Gaashani et al. Ceramics International 47 (2021) 10113–10122

[43] J.M. Haile, Molecular Dynamics Simulations: Elementary Methods, John Wiley & [56] Q. Lai, S. Zhu, X. Luo, M. Zou, S. Huang, Ultraviolet-visible spectroscopy of
Sons, Inc, New York, 1992. graphene oxides, AIP Adv. 2 (2012), 032146.
[44] T. Darden, D. York, L. Pedersen, Particle mesh Ewald - an n.Log(n) method for [57] E. Rommozzi, M. Zannotti, R. Giovannetti, C.A. D’Amato, S. Ferraro, M. Minicucci,
Ewald sums in large systems, J. Chem. Phys. 98 (1993) 10089–10092. R. Gunnella, A. Di Cicco, Reduced graphene oxide/TiO2 nanocomposite: from
[45] O.S. Lee, QGO Builder (An In-House Code of QEERI), 2019. Doha, Qatar. synthesis to characterization for efficient visible light photocatalytic applications,
[46] O.S. Lee, Dynamic properties of water confined in graphene-based membrane: a Catalysts 8 (2018) 598.
classical molecular dynamics simulation study, Membranes 9 (2019) 165–177. [58] J. Paredes, S. Villar-Rodil, A. Martínez-Alonso, J. Tascon, Graphene oxide
[47] J.C. Phillips, R. Braun, W. Wang, J. Gumbart, E. Tajkhorshid, E. Villa, C. Chipot, R. dispersions in organic solvents, Langmuir 24 (2008) 10560–10564.
D. Skeel, L. Kale, K. Schulten, Scalable molecular dynamics with NAMD, J. Comput. [59] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, L.
Chem. 26 (2005) 1781–1802. B. Alemany, W. Lu, J.M. Tour, Improved synthesis of graphene oxide, ACS Nano 4
[48] A.M. Dimiev, J.M. Tour, Mechanism of graphene oxide formation, ACS Nano 8 (2010) 4806–4814.
(2014) 3060–3068. [60] D. Calloway, Beer-lambert law, J. Chem. Educ. 74 (1997) 744.
[49] H.L. Guo, X.F. Wang, Q.Y. Qian, F.B. Wang, X.H. Xia, A green approach to the [61] G. Jing, Z. Ye, J. Wu, S. Wang, X. Cheng, V. Strokova, V. Nelyubova, Introducing
synthesis of graphene nanosheets, ACS Nano 3 (2009) 2653–2659. reduced graphene oxide to enhance the thermal properties of cement composites,
[50] S. Hanafi, D. Trache, W. He, W.X. Xie, A. Mezroua, Q.L. Yan, Thermostable Cement Concr. Compos. 109 (2020) 103559.
energetic coordination polymers based on functionalized GO and their catalytic [62] Z.B. Liu, Y. Wang, X.L. Zhang, Y.F. Xu, Y.S. Chen, J.G. Tian, Nonlinear optical
effects on the decomposition of AP and RDX, J. Phys. Chem. C 124 (2020) properties of graphene oxide in nanosecond and picosecond regimes, Appl. Phys.
5182–5195. Lett. 94 (2009).
[51] W. Shao, X.F. Liu, H.H. Min, G.H. Dong, Q.Y. Feng, S.L. Zuo, Preparation, [63] X.M. Sun, Z. Liu, K. Welsher, J.T. Robinson, A. Goodwin, S. Zaric, H.J. Dai, Nano-
characterization, and antibacterial activity of silver nanoparticle-decorated graphene oxide for cellular imaging and drug delivery, Nano Res. 1 (2008)
graphene oxide nanocomposite, ACS Appl. Mater. Interfaces 7 (2015) 6966–6973. 203–212.
[52] K. Erickson, R. Erni, Z. Lee, N. Alem, W. Gannett, A. Zettl, Determination of the [64] P. Johari, V.B. Shenoy, Modulating optical properties of graphene oxide: role of
local chemical structure of graphene oxide and reduced graphene oxide, Adv. prominent functional groups, ACS Nano 5 (2011) 7640–7647.
Mater. 22 (2010) 4467–4472. [65] C. Botas, A.M. Pérez-Mas, P. Álvarez, R. Santamaría, M. Granda, C. Blanco,
[53] N.R. Wilson, P.A. Pandey, R. Beanland, R.J. Young, I.A. Kinloch, L. Gong, Z. Liu, R. Menéndez, Optimization of the size and yield of graphene oxide sheets in the
K. Suenaga, J.P. Rourke, S.J. York, Graphene oxide: structural analysis and exfoliation step, Carbon 63 (2013) 576–578.
application as a highly transparent support for electron microscopy, ACS Nano 3 [66] K. Krishnamoorthy, M. Veerapandian, R. Mohan, S.-J. Kim, Investigation of Raman
(2009) 2547–2556. and photoluminescence studies of reduced graphene oxide sheets, Appl. Phys. A
[54] P. Swift, Adventitious carbon—the panacea for energy referencing? Surf. Interface 106 (2012) 501–506.
Anal. 4 (1982) 47–51. [67] D. Li, M.B. Müller, S. Gilje, R.B. Kaner, G.G. Wallace, Processable aqueous
[55] C. Minitha, M. Lalitha, Y. Jeyachandran, L. Senthilkumar, R.K. RT, Adsorption dispersions of graphene nanosheets, Nat. Nanotechnol. 3 (2008) 101–105.
behaviour of reduced graphene oxide towards cationic and anionic dyes: Co-action [68] F.W. Low, C.W. Lai, S.B. Abd Hamid, Easy preparation of ultrathin reduced
of electrostatic and π–π interactions, Mater. Chem. Phys. 194 (2017) 243–252. graphene oxide sheets at a high stirring speed, Ceram. Int. 41 (2015) 5798–5806.

10122

You might also like