You are on page 1of 11

Carbon Trends 10 (2023) 100251

Contents lists available at ScienceDirect

Carbon Trends
journal homepage: www.elsevier.com/locate/cartre

Physicochemical transformation of graphene oxide during heat treatment


at 110–200 °C
Sohan Bir Singh, Seyed A. Dastgheib∗
Illinois State Geological Survey, Prairie Research Institute, University of Illinois Urbana-Champaign, 615 East Peabody Drive, Champaign, IL 61820, United States

a r t i c l e i n f o a b s t r a c t

Keywords: Graphene oxide (GO) was subjected to different heating pathways to investigate water removal at 110 °C (GO-
Graphene oxide 110), water removal followed by removal of surface groups at 140 °C or 200 °C (GO-110–140 or GO-110–200), and
Reduced graphene oxide simultaneous removal of water and surface groups at 140 °C or 200 °C (GO-140 or GO-200). Physicochemical
Physicochemical properties
characteristics of GO samples at each stage were determined to quantify and understand the impact of each
Functional groups
treatment pathway on porosity development, physical structure, or surface chemistry of GO. Surface elemental
Thermal reduction
composition showed the following order for C/O ratio: GO-140–200 > GO-110–200 > GO-200, suggesting that
a two-step heat treatment results in more oxygen removal from the surface, and FTIR profiles confirmed the
lowest intensities of all functionalities for GO-140–200. Proton NMR spectra of as-received and heat-treated GO
samples confirmed near complete removal of water and hydroxyl groups for samples treated at 200 °C. XRD
results showed that the interlayer spacing of GO reduced from 0.792 nm to the lowest values of 0.371 nm for
GO-200 and 0.367 nm for GO-110–200. Porosity characterization revealed that only two pathways, GO-200 and
GO-110–200, created a porous structure. Further analysis of data suggested that exfoliation is primarily triggered
by decomposition of surface functionalities and release of a substantial volume of CO2 /CO, but water vapor
release helps to enhance the exfoliation. These results demonstrate that direct and rapid heat treatment of GO at
200 °C (or higher temperatures) is the most effective option for developing a porous structure.

1. Introduction The conventional approach for GO preparation is through graphite


oxidation using a combination of strong oxidizers following the method
Graphene is a two dimensional carbon material with zero band gap, of Hummers, Brodie, Staudenmaier, or other similar methods that re-
having sp2 bonded carbon sheets that are arranged in a hexagonal hon- sults in introduction of various hydrophilic oxygen-containing func-
eycomb lattice [1–3]. The major remarkable characteristics of graphene tional groups (e.g.; hydroxyl, epoxy, carbonyl, carboxyl) to the basal
include its high thermal and electrical conductivities and mechanical planes and at the edges of graphene sheets [8–10]. The presence of oxy-
stability [4]. Graphene also has a large theoretical specific surface area gen functionalities makes the GO surface highly hydrophilic [11,12],
of about 2600 m2 /g, potential to be functionalized with non-metals (N, resulting in attraction of water molecules and formation of water clus-
O, S, P, etc.), or impregnated with metals (Pt, Pd, Ni, Fe, etc.) and other ters on the surface.
materials [1,3]. These characteristics make graphene a promising mate- Drying is the final step of the GO preparation process that can be
rial for different advanced industrial applications [1,3]. conducted at different temperatures at atmospheric pressure or under
Graphene oxide (GO) has attracted substantial attention due to its vacuum, by freeze drying or other methods [13–15]. The drying temper-
unique characteristics resulting from its two-dimensional carbon struc- ature is an important parameter that may impact GO physicochemical
ture and existence of various oxygen-containing functionalities at the properties due to decomposition of heat-sensitive functional groups and
defect sites of the basal planes or at the edge of the graphene sheets removal of intercalated water [13,15,16]. Several researchers have re-
[5,6]. Graphene oxide can be functionalized with different functional ported the impact of heat treatment temperature on selected properties
groups to prepare advanced functionalized materials with desired prop- of GO. Tene et al. [16] reported that GO samples subjected to drying
erties for different applications such as solar cells, energy storage, mem- at 80 °C for 0.5–24 h exhibited similar D, G, and G’ bands, but there
brane separation, drug delivery, catalysts, healthcare, desalination, etc. was some variations in D∗ , D∗ ∗ , and 2D2A bands in their Raman spec-
[3,7]. tra. Additionally, a slight shift in the UV–vis profiles was reported as
water was removed. Transmission electron microscopy results of their


Corresponding author.
E-mail address: seyed@illinois.edu (S.A. Dastgheib).

https://doi.org/10.1016/j.cartre.2023.100251
Received 7 December 2022; Received in revised form 13 January 2023; Accepted 5 February 2023
2667-0569/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

work showed a change in the stacking of GO sheets resulting from wa-


ter removal. However, they provided no additional evidence about pos-
sible degradation or transformation of surface functionalities when the
samples were dried below 80 °C. Wang et al. [13] compared drying im-
pacts of two freeze drying methods with a thermal drying method (at
100 °C) for an N–doped graphene hydrogel. They reported that ther-
mal drying at 100 °C resulted in development of a less porous structure,
however, based on Fourier transform infrared (FTIR) and X-ray photo-
electron spectroscopy (XPS) results, drying at 100 °C did not have any
impact on the chemical structure of GO. Jo et al. [17] studied the im-
pact of drying temperatures from 120 to 300 °C on the characteristics
and electrochemical activities of the heat-treated GO in which they re-
ported a maximum porosity development of 149 m2 /g and a minimum
oxygen content of 22.81% for the samples treated at 250 °C. The oxygen
content of the sample treated at 120 °C was 49.46% that was sharply
dropped to 25.77% when the sample was treated at 200 °C. The oxy-
gen content of samples treated at 200 °C and 300 °C were 25.77% and
24.46%, respectively, suggesting similar reduction levels at 200–300 °C
range.
Temperature programmed desorption (TPD) investigation of GO in
a temperature range of 50–250 °C showed no significant dissociation
or transformation of surface functionalities below ∼125 °C but major
transformation and dissociation of oxygen surface functionalities above
125 °C and below 200 °C were reported [8,18,19]. The experimental
Fig. 1. Heat treatment pathways of ∼30 mg graphene oxide (GO) samples for
data (e.g., oxygen content of samples, TPD results, etc.) reported by 30 min. Photographs show significant exfoliation and expansion for GO-200 and
other researchers also confirm that decomposition of oxygen surface GO-110–200 samples.
groups on GO occurs below 200 °C [8,14,18–20]. For conventional car-
bon materials, Figueiredo et al. [21] reported a compilation of decom-
position temperatures (based on TPD data) for oxygen functional groups perature and quickly transferred to a sealed vial with minimal exposure
on carbon that is mostly well above 200 °C. Reported <200 °C temper- to ambient air. Additionally, GO samples that had already been dried at
ature for decomposition of major surface oxygen functionalities on GO 110 °C and 140 °C, were heat treated at 200 °C using the same proce-
is not consistent with the reported dissociation range of similar oxygen dure as described above. Heat-treated samples are identified by adding
functionalities on the conventional carbon materials (e.g., activated car- an extension to GO in which the extension indicates the heat treatment
bons, chars, or graphite). This important observation, surprisingly not temperature. For example, GO-110 is the sample that was heat-treated
discussed by carbon researchers, deserves further investigation. Addi- at 110 °C or GO-110–200 is the sample that was first heat-treated at
tionally, the impact of heat treatment temperature and pathways (that 110 °C, then cooled to the ambient temperature and subjected to a sec-
determine the extent of water removal and decomposition of surface ond heat treatment at 200 °C.
groups) on porosity development and transformation of surface chem-
istry of GO needs to be thoroughly investigated. 2.2. Characterization
Through a systematic work, we have subjected GO to different heat-
ing pathways to investigate the impact of removing only water (at Thermogravimetric analysis (TGA) of the as-received GO sample was
<110 °C), water removal followed by decomposition of surface groups at performed using a Thermo Cahn VersaTherm instrument under nitrogen
140 or 200 °C, and simultaneous removal of water and decomposition of from 30 to 900 °C with a heating rate of 4 °C/min. The CHN bulk ele-
surface groups at 140 or 200 °C. The physicochemical characteristics of mental composition of the samples was determined by a CE–440 CHN
GO samples at each stage of treatment were determined to quantify and analyzer (Exeter Analytical, Inc.). The bulk oxygen content of each sam-
understand the extent of water removal and decomposition of surface ple was estimated by difference.
groups under each drying or heat treatment condition and their impact The atomic surface elemental compositions and distribution of sur-
on porosity development, physical structure, or surface chemistry of GO face functionalities of the samples were measured through XPS using
which are critical parameters for GO applications. a Kratos Axis ULTRA instrument (Kratos Analytical Ltd., Manchester,
UK) with a surface penetration depth of ∼30 Å. Casa XPS software was
2. Experimental used to estimate surface composition. XPS Cls spectra was deconvoluted
to find the distribution of different surface functionalities based on the
2.1. Sample preparation binding energy values assigned to each functional group.
Oxygen–containing functional groups of GO samples were also iden-
Graphene oxide was purchased from Graphene Supermarket (Calver- tified using FTIR analysis. A Bruker ALPHA FTIR spectrometer was used
ton, New York, USA). Fig. 1 shows different pathways for drying or heat to record the spectra of the samples in attenuated total reflectance mode.
treatment of GO at 70, 110, 140, or 200 °C that was carried out using The absorbance measurements were taken over the spectral range of
the sample preparation apparatus of a Gemini instrument (Micromerit- 400–4000 cm−1 with resolution of 4 cm−1 and average of 64 scans at
ics, Norcross, GA). Approximately 30 mg of GO sample was placed inside ambient temperature.
a capped PYREX tube and subjected to vacuum at 0.1 mm Hg at ambi- Solid-state nuclear magnetic resonance (NMR) spectra of the sam-
ent temperature for 30 min, then the sample tube was backfilled with ples were recorded using a Unity Inova 300 MHz NMR spectrometer
nitrogen. The capped sample tube was quickly inserted into the heating (Varian Inc., USA) operating at 10 kHz. Proton spectra was performed
jacket of the sample preparation apparatus that was preheated to the de- at resonance frequency of 300 MHz at room temperature. Powder sam-
sired temperature (i.e.; 70, 110, 140, or 200 °C) in which the sample was ples were packed manually into the NMR rotor and placed inside the
heated to the set temperature with a heating rate of ∼ 33 °C/min, heated NMR spectrometer for analysis. MestreNova software was used to ana-
for 30 min at the desired temperature, then cooled to the ambient tem- lyze samples spectra.

2
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

A powder X–ray diffraction (XRD) instrument (Bruker D8 Advance) drying experiments (Fig. 2a and 2b) show similar water contents of 11%
operating with a step size of 0.05° and a scan rate of 1°/min was used to or 12% for the GO sample. The TGA results also show 3% weight loss
record the XRD profiles of the samples. The XRD profiles were recorded at 110–140 °C, due to dissociation of more heat sensitive oxygen func-
from 5° to 70° using Cu–K𝛼 radiation (𝜆 = 0.154 nm). Interlayer spacing tionalities, and a major weight loss of 21% between 140 °C and 200 °C
(d) of the samples were estimated based on the observed diffraction as decomposition of surface functionalities continues. The weight loss
angle in the XRD profile and by using Bragg’s equation (Eq. (1)): of the GO sample continues, but with a constant lower rate, when the
n𝜆 sample is heated from 200 to 900 °C. The observed TGA profile of the
𝑑= (Eq. (1)) GO is consistent with the TGA results reported for GO in the literature
2 𝑠𝑖𝑛𝜃
Where, 𝜆 = wavelength of the X–ray beam (0.1540 nm, Cu–K𝛼), [9,22,23].
𝜃 = diffraction angle in radians, n = 1 (order of diffraction) and d = in- In the batch heat treatment experiments, when the samples were di-
terlayer spacing in nm. rectly heated at 140 or 200 °C (i.e., samples GO-140 and GO-200 in
The crystallite size (D) of each sample was estimated according to Fig. 2b) they showed weight losses of 15% and 40%, respectively. Con-
Scherrer’s equation (Eq. (2)): sidering the water content of 12% for the GO sample (based on the batch
experiments), the difference between the weight loss at 140 °C or 200 °C
0.89 𝜆
𝐷= (Eq. (2)) with the water content of the sample is about 3% and 28%, respectively,
𝐹 𝑊 𝐻𝑀 𝑐𝑜𝑠(𝜃)
which is indicative of the extent of weight loss due to decomposition of
Where, 𝜆 = wavelength of the X–ray beam (0.1540 nm, Cu–K𝛼), surface functionalities. For the GO sample that is first subjected to drying
FWHM is full width at half maxima and 𝜃 = diffraction angle in radians. at 110 °C, then heat treated at 200 °C (i.e., GO-110–200), the observed
The number of layers (N) was calculated using Eq. (3) [14]: weight loss is 27% that represents the loss of surface oxygen groups and
𝐷 is slightly less than the estimated similar loss of 28% for GO-200 sam-
𝑁= +1 (Eq. (3))
𝑑 ple (i.e., direct heat treatment at 200 °C) discussed above. Similarly,
Where, N and d indicate number of layers and interlayer spacing the observed weight loss for GO-140–200 is 23% that shows the extent
between the layers, respectively, and D is the average crystallite size of decomposition of surface functionalities of GO-140 sample treated
calculated using the observed peaks in the 2𝜃 range of 11–26° at 200 °C. Addition of 3% weight loss (accounting for the decomposed
Raman spectra of samples were recorded by Nanophoton Raman 11 surface functionalities of GO-140) to 23% weight loss observed for GO-
instrument with 532 nm line of Ar+ ions laser as an excitation source. 140–200 results in 26% weight loss value for dissociation of all surface
All the spectra were recorded in the range of 700–3000 cm−1 with an functionalities of GO through this two-stage heat treatment pathway. It
exposure time of 5 s. Before the analysis, each sample was manually is notable that the TGA results showed a weight loss of 21% between
placed at the probe tip on a glass slide at room temperature. The exper- 140 °C to 200 °C (Fig. 2a) which is slightly lower than the discussed 23%
imental Raman spectra of the GO samples were fitted to the Lorentzian weight loss of GO-140–200 (Fig. 2b). The higher observed weight loss
function within the OriginPro 2021 software. for the batch heating experiments is likely due to longer soak time of
Nitrogen adsorption–desorption isotherms at −196 °C were obtained GO-140–200 (i.e., 30 min at 200 °C) and rapid heating (i.e., 33 °C/min)
by a Micromeritics Gemini VII instrument and used to estimate the sur- compared to the TGA experiment that had no additional soak time at
face area and porosity of the samples. No additional drying or degassing 200 °C and was conducted using a slow heating rate of 4 °C/min.
were performed, except for the as-received GO sample that was degassed Similar weight loss results for the overall degradation of surface oxy-
at 70 °C for 2 h under vacuum before analysis. The surface area was eval- gen functionalities under one-step or two-step heat treatment of GO to
uated using Brunauer–Emmett–Teller (BET) equation at relative pres- 200 °C may suggest that the three employed pathways (i.e., GO-200,
sure range of 0.03−0.3, and the total pore volume was estimated at the GO-110–200, and GO-140–200) resulted in the degradation of similar
relative pressure of 0.99. Micropore volume of samples were determined masses (i.e., 26–28 wt.%) of surface groups. However, it cannot neces-
from the Dubinin−Radushkevich equation in the relative pressure range sarily indicate the exact same surface chemistry for these samples. This
of 0 − 0.1. The cumulative meso– and macropore volume was calculated matter is further discussed in light of the surface characterization data.
by subtracting the micropore volume from the total pore volume.
3.2. Impact of heat treatment pathways on GO surface chemistry
3. Result and discussions
Based on the TGA and TPD studies, the temperature range of ∼125–
Graphene oxide samples were subjected to drying ≤110 °C, or to one- 200 °C is the most significant range for decomposition of GO’s surface
stage or two-stage heat treatments ≤200 °C (Fig. 1). Depending on the functionalities. The TGA profile of GO shown in Fig. 2a and the TGA
employed condition, heat treatment may result in the removal of wa- profiles of GO reported in literature [9] exhibit a sharp weight drop
ter or dissociation of surface oxygen-containing groups, both of which between ∼125 °C and 200 °C. Additionally, reported TPD data show a
can impact the physical properties of the GO, namely porosity, surface substantial release of CO2 and CO in ∼125–200 °C range which is indica-
area, interlayer spacing, and other properties. Additionally, dissociation tive of thermal degradation of oxygen-containing functionalities on the
of carbon-oxygen functional groups alters the surface chemistry of the carbon surface [14,18,19]. These TPD results also show a significantly
heat-treated GO. The impact of each heat treatment pathway on physic- higher release of CO2 than CO in the temperature range of ∼125–200 °C
ochemical characteristics of GO is discussed below. due to degradation of mainly carboxylic, lactone, anhydride, and similar
groups that generate CO2 after decomposition. The TGA profile of GO
3.1. Impact of heat treatment pathways on GO mass loss above 200 °C to 900 °C shows a continuous weight loss, but at a signifi-
cantly lower rate which is likely due to further decomposition of surface
The TGA profile of as-received GO sample is shown in Fig. 2a and the groups leading to release of mainly CO and H2 which is confirmed with
observed weight losses of the GO samples subjected to rapid heat treat- the TPD data [14].
ment at different temperatures in batch experiments is shown in Fig. 2b. The observed major degradation of surface oxygen functionalities on
The TGA profile of GO exhibits weight losses of ∼7, ∼11, ∼14 and ∼35% GO in low temperature range of ∼125–200 °C and reported degradation
at 70, 110, 140, and 200 °C, respectively. The observed weight loss at temperature of different oxygen functionalities on GO are not consis-
70 or 110 °C is mainly due to evaporation of free or intercalated water tent with the degradation temperature of the same oxygen functional-
without any considerable dissociation of surface oxygen functionalities, ities on conventional carbon materials (i.e. activated carbon, chars, or
since degradation of these functionalities is reported to start >125 °C graphite). This matter is surprisingly not discussed in the literature. A
[19]. Based on the observed weight loss at 110 °C, both TGA and batch side-by-side comparison for degradation temperatures of surface oxy-

3
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Fig. 2. (a) TGA profiles of fresh graphene oxide (GO), and (b) comparison of mass loss of dried GO at 70 °C, 110 °C, 140 °C and 200 °C from batch heat treatment
experiments shown in Fig. 1.

Table 1
Reported decomposition temperatures ( °C) for various oxygen-containing functional groups on conventional carbons and graphene oxide (GO). Reference for various
values are provided in the brackets.

Functional group Reported decomposition Reported decomposition Structures of oxygen-containing functional groups
temperature of each group on temperature of each group
conventional carbons (°C) on GO (°C)

Carboxylic acid 100–400 [21], 250 [24], 100–250 [33], 200 [34]
200–250 [25], 430 [26], 300
[27], 180–200 [28], 300 [28],
302 [29], 250–300 [30], 277
[31], 247–287 [32]
Phenolic/hydroxyl (C–OH) 617–637 [21],600–700 [28], 550 100–150 [33], 200 [34], 180
[30], 702 & 837 [31], 627–657 [35], 250 [36]
[32]
Carbonyl/ ketone/quinone (C = O) 787–827 [21], 800–900 [25], 120–175 [33]
700–980 [37], 702 & 837 [31]
Carboxylic anhydride (CO–O-CO) 627 [24], 350–400 [25]„ 627
[38,39], 280 [26], 427–502 [29],
437–657 [31], 407–447 [32]
Ether/epoxide (C–O) 827 [21], 550 [30], 700 [38] 100–200 [33], 200 [34], 180
[35], 90 [40], 200 [41]
Lactone 667 [21], 350–400 [25], 400
[27], 450 [30], 473–657 [31],
190–650 [37,42], 627 [38,39]

gen functionalities on GO and conventional carbon materials in shown content and C/O ratio. If carbon is conserved during GO drying at 110 °C
in Table 1. For the conventional carbon materials, for more heat sensi- (i.e., no carbon-oxygen functionalities decomposed), carbon content of
tive groups (i.e., CO2 generating groups) such as carboxylic, anhydride, GO-110 is expected to increase by a factor of 1/(1-x), where x is the frac-
and lactones, decomposition temperatures of well above 200 °C, even tion of weight loss due to water removal. The expected weight percent-
up to ∼650 °C, are reported by the majority of works cited in Table 1. age and atom percentage of carbon for GO-110, based on the GO carbon
The decomposition temperature of other surface functionalities includ- weight percentage of 46.65%, atom percentage of 68.15%, and weight
ing phenolics (hydroxyls), carbonyls (or similar forms of ketones and loss fraction of 0.11 during drying, can be calculated as 52.41 wt.% and
quinones), and etheric (or epoxides) on conventional carbon materials 76.57 atm.% which are close to the measured values of 52.62 wt.% and
as listed in Table 1 are in the range of 550–900 °C. However, decompo- 75.27 atm.% (Table 2). The bulk hydrogen contents of GO before and
sition temperature of all surface functionalities on GO is reported to be after drying at 110 °C also are consistent with the expected decrease in
below 200 °C (Table 1). This shows that oxygen surface functionalities hydrogen content due to ∼11% water loss. These results confirm the ab-
are less stable and more prone to thermal decomposition if they are on sence of any degradation of carbon-oxygen functionalities and carbon
GO that has a limited number of carbon layers (e.g., <10) compared to loss during heat treatment at 110 °C. Compared to GO-110, for sample
those that are on conventional amorphous (e.g., activated carbon) or GO-140 a considerably higher reduction in oxygen content and a higher
graphitic carbon materials that have infinite number of carbon layers. increase in carbon content and C/O ratio was observed due to removal
This important observation deserves further research through additional of water and some degradation of oxygen-containing surface function-
experimental and simulation studies. alities (Table 2).
Bulk weight-based and surface atom-based elemental analysis data Oxygen and carbon contents of samples GO-200 and GO-110–200
of as-received and heat-treated GO samples are shown in Table 2. Drying were similar suggesting that the two pathways of direct heat treatment
at 110 °C is expected to result in removal of water only, leading to a re- at 200 °C or initial drying at 110 °C followed by heat treatment at 200 °C
duction in oxygen and hydrogen contents and an increase in the carbon resulted in degradation of similar amounts of surface functionalities.

4
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Table 2
Bulk and surface elemental composition of as-received and heat-treated graphene oxide (GO) samples.

Bulk weight% Surface atom%


Sample
O C H N O C C/O

GO 39.51 46.65 ± 0.56 2.57 ± 0.23 0.40 ± 0.03 31.85 ± 0.07 68.15 ± 0.26 2.13 ± 0.00
GO-110 34.75 52.62 ± 0.14 1.49 ± 0.01 0.27 ± 0.01 24.73 75.27 3.04
GO-140 21.59 66.50 ± 0.08 0.71 ± 0.01 0.33 ± 0.01 19.98 ± 0.04 80.02 ± 0.17 4.00 ± 0.00
GO-200 14.95 73.28 ± 0.49 0.82 ± 0.06 0.12 ± 0.01 18.21 81.79 4.49
GO–110–200 14.22 74.17 ± 0.30 0.64 ± 0.04 0.10 ± 0.01 17.67 82.33 4.65
GO–140–200 19.75 68.39 ± 0.08 0.65 ± 0.00 0.34 ± 0.01 16.28 83.72 5.14

Fig. 4. FTIR spectra of as-received and heat-treated graphene oxide (GO) sam-
ples.

for C–O (representing epoxy & alkoxy groups), and ∼287.8 eV for C = O
(representing carbonyl and other similar groups) [43–45]. However,
only one small peak at ∼289.5 eV representing COOH or carboxylic
groups was observed for GO before and after drying at 110 °C. The
absence of carboxylic peak for GO-140 and other samples treated at
higher temperatures suggests that all carboxylic groups were removed
or transformed in the temperature range of 110–140 °C.
Fig. 3. Deconvoluted C1s XPS spectra of as-received and heat-treated graphene The XPS results in Table 4 show different distribution of carbon-
oxide (GO) samples.
oxygen functionalities for GO and GO-110 that, as previously discussed,
have similar carbon contents after accounting for the removed water.
GO-110 results show that after removing intercalated water from GO
However, another pathway of initial heat treatment at 140 °C followed (by drying at 110 °C) the most significant change occurs with ∼12% re-
by a second heat treatment at 200 °C (i.e., GO-140–200) resulted in a duction of C–O functionalities (i.e., epoxy, etheric, hydroxyl) and ∼14%
different bulk composition than the direct heat treatment at 200 °C, as increase in C–C/C = C (or the carbon skeletal backbone) percentage.
GO-140–200 has a considerably higher bulk oxygen content and lower These changes might be due to electrostatic interaction of intercalated
carbon content than GO-200. Additionally, the bulk composition of GO- water with partially charged carbon from etheric, epoxide, or other sim-
140–200 is more similar to the composition of GO-140 than GO-200. ilar groups as illustrated in Fig. 5a. After removal of intercalated water,
The surface elemental composition of samples shows the following or- the electrostatic dipole interaction of oxygen from water with carbon
der for C/O ratio: GO-140–200 > GO-110–200 > GO-200 (which is not and its associated contribution to C1s spectra will be eliminated that
consistent with the bulk elemental composition) suggesting that a two- would increase the percentage of C–C/C = C contribution. It has been
step heat treatment results in more oxygen removal from surface than a reported in the literature that electrostatic dipole interactions can im-
direct single-step heat treatment to 200 °C (Table 2). pact the intensity of XPS binding energies [48].
Surface functionalities of GO were characterized by deconvolution Distribution of the surface functional groups for GO-140 sample
of C1s peaks of XPS spectra (Fig. 3), and analysis of FTIR profiles shows a lower concentration of C–O groups (i.e., epoxy and etheric) than
(Fig. 4) which were conducted using an extensive survey of peak GO-110 and a higher concentration of C = O groups (e.g., carbonyl) than
assignments for various surface functionalities of GO reported in GO-110 and even GO (Table 4). This suggests that in addition to the im-
literature (Table 3). XPS analysis results for distribution of surface pact of removal of intercalated water (discussed above for GO-110 and
functionalities of as-received and heat-treated GO samples are shown shown in Fig. 5a), a transformation of epoxy to carbonyl groups (Fig. 5b)
in Table 4. C1s spectra of all samples show three main peaks at: might occur. Additionally, decomposition of some oxygen-containing
284.45 eV, which is assigned to C–C(sp2 )/C = C(sp3 ) bonds, ∼286.4 eV surface functionalities that result in releasing CO2 or CO is expected,

5
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Table 3
A summary of reported XPS C1s binding energies and FTIR wavelength for characterization of carbon-oxygen functionalities of graphene oxide (GO) materials.
Reference for various values are provided in the brackets.

Functional group XPS binding energy (eV) FTIR wavelength (cm−1 )

C = C (sp2 -hybridized 284.5 [8], 284.8 [9], 284.4 [12], ∼284.4 [14], 283.2 [19], 1557 [13], 1621 [19], 1623 [35], 1500–1600 [33,47], 1580
carbon) 284.5 [43], 284.4 [44], 284.5 [45], ∼284.6 [46] [43], 1620 [45]
C–C (sp3 -hybridized carbon) 284.8 [6,9], 284.4 [12], 284.6 [13], 285.5 [14], ∼285.1 [14],
284.5 [15], 284.9 [19], ∼284.6 [46], 284.5 [44], 285.4 [45]
C–O–C (epoxy, ether, alkoxy) 286.2 [6], 286 [9], 287.7 [14], 286.5 [12], 286.5 [13], ∼287.2 1218 (epoxy) [9], 1060 (alkoxy) [9], 1046 [19], 1054 [35],
[14], 286 [17], 286.9 [19], 286.2 [43], 286.5 [44], 286.5 [45], 1280–1320 [33,47], 800–900 [33,47], 1056 [43], 1055 [45]
286.6 [46]
C = O (carbonyl) 287.8 [6], 287.6 [9], 288.2 [12], 287.7 [13], 287.9 [17], 287.8 1728 [9], 1730 [19], 1725 [35], 1750–1850 [33,47], 1703
[19], 287.5 [44], 287.6 [45], ∼288.2 [46] [43], 1725 [45]
COOH (carboxylic) 289.1 [6], 288.7 [8], 288.2 [9], 289.8 [12] ∼288.5 [14], 289
[17], 289.1 [19], 288.6 [43], 289.3 [44], 288.9 [45]
C–OH (hydroxyl) 1222 [19], 1383 [35], 1300–1400 [45]
O–H (COOH & H2 O) 3380 [9], 3300 [19], 3416 [35], 3000–3700 [33,47],
3010–3680 [43], 3000–3500 [45]

Table 4
Distribution of surface oxygen-containing functional groups based on deconvolution of C1s peaks
of XPS spectra of the as-received and heat-treated graphene oxide (GO) samples. Distribution
values are in percentages.

Sample C–C/C = C (284.45 eV) C–O (286.4 eV) C = O (287.8 eV) COOH (289.2 eV)

GO 39.55 43.71 13.18 3.57


GO-110 53.36 31.79 10.73 2.12
GO-140 56.99 27.39 15.62 –
GO-200 45.50 26.58 27.92 –
GO–110–200 47.44 27.00 25.56 –
GO–140–200 47.87 24.12 28.01 –

since the thermal degradation of oxygen-containing functionalities on decreased by nearly two orders of magnitude. Theses samples exhibited
the carbon surface reported to occur above ∼125 °C [14,18,19]. a sharp peak at 4.8 ppm with shoulder in the range of 1–2 ppm but with
Heat treatment at 200 °C through direct or indirect pathways re- a small intensity compared to that of the original GO. The small sharp
sulted in a relatively similar distribution of surface functionalities, a peak at 4.8 ppm is associated with the remaining adsorbed water and
large reduction in the percentage of C–O groups (which is similarly ob- the shoulder at 1–2 ppm is assigned to surface structured water and OH
served for GO-140), and less increase in the carbon skeletal (C–C/C = C) groups [49,50]. These results confirm near complete removal of water
percentage compared to GO-140 (Table 4). However, the percentage of and low concentration of hydroxyl groups for samples treated at 200 °C.
C = O groups for samples treated at 200 °C was almost twice as those for Raman spectra of GO samples are shown in Fig. 7. Raman spectra of
the GO sample before the heat treatment. This may mainly occur due as-received and heat-treated GO samples show similar characteristic D
to more preferential dissociation of C–O groups than C = O groups at and G bands at frequencies of 1343–1348 cm−1 and 1590–1598 cm−1 ,
200 °C or possible transformation of some C–O to C = O groups shown respectively, and a broad G’ or 2D band at ∼2700 cm–1 . The G and G’
in Fig. 5b. Results also show that sample GO-140–200 that was initially bands represent a graphitic structure (i.e., carbon sp2 hybridization) and
subjected to heat treatment at 140 °C prior to the second heat treat- the D band is indicative of the extent of defects in the graphitic struc-
ment at 200 °C had slightly higher concentration of C = O groups and ture, mainly created by the formation of carbon–oxygen functionalities
lower concentration of C–O groups compared to GO-110–200 and GO- at the edges of the graphene sheets or other defects [51]. The intensity
200 (Table 4). ratios of D to G band (ID /IG ) were evaluated based on the multiple pro-
Fig. 4 compares the FTIR spectra of GO samples subjected to dif- files obtained for the samples (Fig. 7). The intensity ratio indicates the
ferent heat treatment pathways. All samples had a peak at 1614 cm−1 degree of graphitic ordering in which lower values show more graphitic
caused by sp2 carbon C = C (i.e., skeletal structure) stretching vibration. ordering and less defects. Considering the uncertainty of the measure-
Additionally, all the samples showed a small peak at 1716 cm−1 (for ments, results show that ID /IG values of the as-received and heat-treated
C = O). As-received GO shows broad intense peaks at ∼1348 cm−1 and GO samples are similar, indicating that Raman spectra do not show any
∼3000–3684 cm−1 that are indicative of hydroxyl groups in the stretch- significant difference in GO’s degree of disorder before and after the
ing vibration mode, mainly for water or some carboxylic groups. After employed heat treatments.
drying at 110 °C, or heat treatment at higher temperatures, these peaks The deconvolution peak fitting of the D, G, and 2D Raman bands
are almost disappeared. The peaks at ∼930–1100 cm−1 and 1224 cm−1 for GO sample is shown in Fig. 8. Seven peaks can be fitted within D, G
are assigned to C–O-C (etheric or epoxy groups). FTIR profiles show the and 2D bands at wavelengths of ∼1161, ∼1350, ∼1514, ∼1600 ∼1622,
lowest intensities of all functionalities for sample GO-140–200, suggest- ∼2700 and ∼2930 cm−1 , corresponding to D∗ , D, D∗ ∗ , G, D’, 2D and
ing that this heat treatment pathway was more effective in removal of D + D’ bands, respectively. Shoulder peaks of D∗ , D∗ ∗ , D’ and D + D’
surface functionalities which is also in agreement with the XPS results have been reported in the range of 1150–1200, 1500–1550, 1610–1620
that showed the highest C/O ratio for this sample. and 2925–2940 cm−1 , respectively, for GO and graphene materials
Proton NMR spectra of as-received and heat-treated GO samples are [16,52–54]. The D∗ and D∗ ∗ bands were attributed to the disordered
shown in Fig. 6. As-received GO sample displays a peak at 5.05 ppm that graphitic lattices induced by sp2 -sp3 bonds and amorphous phases,
corresponds to intercalated or free water. Upon drying at 110 °C or heat respectively, whereas D’ and D + D” bands were associated with defects.
treatment at 140 °C, this peak was shifted to lower values of 3.75 ppm or Similar deconvolution peak fitting was also performed for other samples
3.15 ppm. For heat-treated samples prepared at 200 °C through the three prepared in this work but only a small reduction in intensity or a slight
direct or staged pathways, the intensity of this peak was dramatically shift in positions of these bands without any clear trend was observed.

6
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Fig. 5. a) Proposed mechanism for electrostatic dipole interaction (highlighted with dotted lines) of oxygen from intercalated water with carbon from C–O groups
that contributes to the XPS C–O intensity. Arrows show the direction of electron clouds for creation of partial charges. b) An illustration for possible transformation
of some epoxy groups (highlighted in dotted red circles) to carbonyl groups on the graphene oxide (GO) surface.

Fig. 6. 1 HNMR spectra of as-received and heat-treated graphene oxide (GO) samples. Intensities of samples treated at 200 °C (right figure) are two orders of
magnitude lower than those of the other samples (left figure).

3.3. Impact of heat treatment pathways on GO’s structure and porosity layers (i.e., GO), increase the interlayer space and shift 002 line to
∼11° [55]. A comparison of XRD profiles of GO before and after dry-
As reported in the literature, graphite has a major XRD sharp peak ing at 70 °C or 110 °C reveals a shift in the main peak position from 2𝜃
at 2𝜃 angle of ∼26°, representing 002 line, with interlayer spacing of value of 11.15° to 13.15° (for drying at 70 °C) and 14.05° (for drying at
0.335 nm [46]. Introduction of oxygen functionalities to the graphene 110 °C) (Fig. 9 and Table 5), illustrating the impact of removal of inter-
surface and intercalation of water molecules between oxidized graphene calated water molecules on the interlayer spacing. All XRD profiles of

7
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Table 5
Estimated structural parameters of as-received and heat-treated graphene oxide (GO) samples based on
XRD profiles.

Sample 2𝜃 (degree) Interlayer spacing, d (nm) Crystallite size, D (nm) Number of layers, N

GO 11.15 0.792 6.06 (9) 8.65


GO-70 13.15 0.672 5.09 (9) 8.59
GO-110 14.05 0.629 4.65 (8) 8.39
GO-140 19–24 0.409 2.46 (7) 7.01
GO-200 23.9 0.371 2.03 (6) 6.47
GO–110–200 24.2 0.367 1.52 (5) 5.14
GO–140–200 20–24 0.394 0.95 (3) 3.41

Fig. 9. XRD profiles of as-received and heat-treated graphene oxide (GO) sam-
ples.
Fig. 7. Raman spectra of as-received and heat-treated graphene oxide (GO)
samples.

Interlayer spacing, crystallite size, and number of layers that are es-
heat-treated samples exhibit very broad peaks that are indicative of low timated based on the XRD profiles are tabulated in Table 5. These values
extent of crystallinity and a more amorphous structure which is likely should be considered as only rough estimates that require further confir-
created during removal of intercalated water and dissociation of surface mation with other characterization methods such as transmission elec-
functionalities. tron microscopy. The interlayer spacing of GO is estimated as 0.792 nm
Heat treatment at 140 °C shifted the main GO peak from ∼11° to a which is reduced to 0.672 nm and 0.629 nm after drying at 70 °C and
broad peak from ∼19° to ∼24° which is due to removal of intercalated 110 °C, respectively. Interlayer spacing of GO was further reduced to
water and some surface oxygen functionalities. For sample GO-140–200, 0.409 nm after heat treatment at 140 °C. The lowest interlayer spacing
a similar broad peak from ∼20° to ∼24° was also observed. GO sample was observed for GO samples that were heat treated at 200 °C through
directly heat treated at 200 °C (GO-200) or first dried at 110 °C, then different pathways resulting in removal of both intercalated water and
heated at 200 °C (GO-110–200) exhibited broad and small peaks at ∼24° surface functionalities. Samples GO-200 and GO-110–200 show similar

Fig. 8. Deconvolution peak fitting of the D, G, and 2D Raman bands for graphene oxide (GO) sample.

8
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

Table 6
Surface area and pore size distribution of as-received and heat-treated graphene oxide (GO) samples.

Sample BET surface area (m2 /g) Total pore volume (cm3 /g) Micropore volume (cm3 /g) Meso + macropore volume (cm3 /g)

GO 2.0 0.001 0.001 0.000


GO-110 2.0 0.002 0.001 0.001
GO-140 1.8 0.001 0.000 0.001
GO-200 207 0.976 0.085 0.891
GO–110–200 115 0.362 0.046 0.316
GO–140–200 1.9 0.005 0.001 0.004

Table 7
Estimated volume of water vapor and CO2 /CO released during each heat treatment pathway.

GO heat treatment pathway Estimated volume of water vapor released (cm3 /g) Estimated volume of CO2 + CO released (cm3 /g) Total volume of gas released (cm3 /g)

GO-110 216 0 216


GO-140 233 28 261
GO-200 267 278 545
GO–110–200 0 268 268
GO–140–200 0 223 223

volume expansion that was observed only for these two samples (Fig. 1),
and can be attributed to sudden release of intercalated water and dis-
sociation of surface functionalities. The highest obsered exfoliation and
porosity development for GO-200 is due to release of both water va-
por (from quick evaporation of intercalated water) and CO2 /CO from
dissociation of oxygen-containing surface functionalities. Based on 12
wt% moisture content of GO sample when dried at 110 °C (Fig. 2b) and
the ideal gas law assumption for water vapor, the estimated amout of
released water vapor at 200 °C and atmospheric pressure is about 267
cm3 for each gram of GO precursor (Table7). Similarly, based on the
weight loss results (Fig. 2b) of 27% for GO sample in the temperature
range of 110–200 °C, and assuming a CO2 /CO composition of 71/29
(calculated based on the TPD profiles [19]) for the gas released from
degradation of carbon-oxygen functionalities, the volume of suddenly
released gasses can be estimated as 278 cm3 per each gram of precur-
sor (Table 7). Therefore, GO was subjected to a high interlayer pressure
during GO-200 treatment created by sudden release of ∼545 cm3 gas
in confined lnterlayer space of each gram of GO, whereas for GO-110–
200 the volume of released gas and pressure build up was about half
of GO-200 that resulted in less exfoliation and porosity development.
Fig. 10. N2 adsorption–desorption isotherms of as-received and heat-treated For sample GO-110, a significant volume (216 cm3 /g GO) of water va-
graphene oxide (GO) samples.
por was released at 110 °C without creating any exfolication. A com-
parison of GO-110 and GO-110–200 analysis suggests that exfoliation
is primarily trigerred by dissociation of surface groups and release of
values of 0.371 nm and 0.367 nm, respectively, but sample GO-140–200 a substantial volume of CO2 /CO from decomposition of surface func-
exhibits a slightly higher value of 0.394 nm. tionalities but water vapor release helps to significantly enhance the
Both crystallite size and number of GO layers were reduced dra- exfoliation. A similar calculation for GO-140 and GO-140–200 that did
matically after heat treatment. The crystallite size of GO reduced from not exhibit a measureable exfoliation or porosity development shows
6.06 nm to 0.95 nm and the corresponding number of layers reduced estimated CO2 /CO gas release values of 28 cm3 /g and 223 cm3 /g, re-
from 9 to 3 in which the highest impact was observed for samples treated spectively (Table 7). For GO-140 the amount of water vapor release at
at 200 °C particularly for sample GO-140–200 (Table 5). 140 °C can be estimated at 233 cm3 /g, while GO-140–200 did not have
Nitrogen adsorption-desorption isotherms of GO samples before and any intercalated or free water to be released.
after heat treatments are shown in Fig. 10. As-received GO sample, GO These results suggest that quick decomposition of suface function-
samples dried at 110 °C or heat treated at 140 °C, and GO sample heat alities and release of a substantial volume of gas products are the key
treated at 200 °C after an initial heat treatment at 140 °C exhibit a non- factors for GO exfolication, but release of the intercalated water can en-
porous structure and low surface areas of ∼ 2 m2 /g, whereas GO sam- hance the exfolication phenomenon. Overall, direct rapid heat treatment
ple directly heat treated at 200 °C or heat treated at 200 °C after ini- of GO at 200 °C is the most effective option for developing a mesoporous
tial drying at 110 °C shows a porous structure (Fig. 10 and Table 6). structure among other pathways explored in this work. More experi-
Both GO-200 and GO-110–200 samples exhibit type IV isotherms with mental and simulation studies are needed to investigate the mechanism
H3 hystereresis loops which are indicative of a mesoporous structure of thermal reduction of carbon-oxygen functionalities and the treshold
with grove pores fomed by exfoliation of GO layers. The surface area of pressure needed for GO exfoliation.
GO-200 is roughly twice and its total pore volume is about three times
of sample GO-110–200 in which for both samples the micropore vol- 4. Conclusions
ume accounts for only ∼10% of the total volume (Table 6). The high
porosity of GO-200 and GO-110–200 samples compared to non-porous Through a systematic work, the impact of removing only water, wa-
structure of other samples is consistent with the exfoliation and large ter removal followed by removal of surface groups at 140 °C or 200 °C,

9
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

and simultaneous removal of water and decomposition of surface groups Materials Research Laboratory Central Facilities, University of Illinois,
at 140 °C or 200 °C from GO were investigated. which are partially supported by the U.S. Department of Energy under
Thermogravimetric analysis exhibited a sharp weight drop in low grants DE-FG02–07ER46453 and DE-FG02–07ER46471, and in the lab-
temperature range of ∼125–200 °C that is resulted from decomposition oratories of the School of Chemical Sciences at the University of Illinois
of surface oxygen groups. The observed major degradation of surface Urbana Champaign (UIUC). The authors would like to thank Dr. Guiron-
oxygen functionalities on GO in low temperature range of ∼125–200 °C net at the Chemical and Biomolecular Engineering Department of UIUC
and reported degradation temperatures of different oxygen functional- for providing access to his laboratory for FTIR analysis.
ities on GO are not consistent with the reported degradation tempera-
tures of the same oxygen functionalities on conventional carbon mate-
References
rials (i.e., activated carbon, chars, or graphite) that are at significantly
higher levels. This shows that oxygen-containing surface functionalities [1] F. Rehman, F.H. Memon, Z. Bhatti, M. Iqbal, F. Soomro, A. Ali, K.H. Thebo,
are less stable and more prone to thermal decomposition if they are on Graphene-based composite membranes for isotope separation: challenges and op-
GO that has a limited number of carbon layers (e.g., <10) compared to portunities, Rev. Inorg. Chem. 42 (2022) 327–336, doi:10.1515/revic-2021-0035.
[2] F. Rehman, F. Hussain Memon, S. Ullah, M.A. Jafar Mazumder, A. Al-Ahmed,
those that are on conventional amorphous (e.g., activated carbon) or F. Khan, K. Hussain Thebo, Recent development in laminar transition metal
graphitic carbon materials that have infinite number of carbon layers. dichalcogenides-based membranes towards water desalination: a review, Chem. Rec.
This important observation deserves further research through additional 22 (2022) e202200107, doi:10.1002/tcr.202200107.
[3] S. Gadipelli, Z.X. Guo, Graphene-based materials: synthesis and gas
experimental and simulation studies. sorption, storage and separation, Prog. Mater. Sci. 69 (2015) 1–60,
The surface elemental composition of samples suggested that a two- doi:10.1016/j.pmatsci.2014.10.004.
step heat treatment to 200 °C results in more oxygen removal from sur- [4] 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
face than a direct single-step heat treatment to 200 °C, which is also environment, Energy Environ. Sci. 7 (2014) 1564–1596, doi:10.1039/C3EE43385D.
confirmed with FTIR results that show the lowest intensities of all func- [5] C. Chung, Y.-.K. Kim, D. Shin, S.-.R. Ryoo, B.H. Hong, D.-.H. Min, Biomedical ap-
tionalities for sample GO-140–200. plications of graphene and graphene oxide, Acc. Chem. Res. 46 (2013) 2211–2224,
doi:10.1021/ar300159f.
XRD profiles of heat-treated samples exhibit very broad peaks that
[6] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia,
are indicative of low extent of crystallinity and a more amorphous struc- Y. Wu, S.T. Nguyen, R.S. Ruoff, Synthesis of graphene-based nanosheets via
ture which is likely created from delamination of GO layers during re- chemical reduction of exfoliated graphite oxide, Carbon 45 (2007) 1558–1565,
doi:10.1016/j.carbon.2007.02.034.
moval of intercalated water and dissociation of surface functionalities.
[7] W. Yu, L. Sisi, Y. Haiyan, L. Jie, Progress in the functional modification
The lowest interlayer spacing was observed for GO samples that were of graphene/graphene oxide: a review, RSC Adv. 10 (2020) 15328–15345,
heat treated at 200 °C through different pathways resulting in removal of doi:10.1039/D0RA01068E.
both intercalated water and surface functionalities. Both crystallite size [8] C. Botas, P. Álvarez, P. Blanco, M. Granda, C. Blanco, R. Santamaría, L.J. Romasanta,
R. Verdejo, M.A. López-Manchado, R. Menéndez, Graphene materials with different
and number of GO layers were reduced dramatically after heat treat- structures prepared from the same graphite by the Hummers and Brodie methods,
ment due to delamination or exfoliation of GO. Carbon 65 (2013) 156–164, doi:10.1016/j.carbon.2013.08.009.
As-received GO sample, GO samples dried at 110 °C or heat treated [9] T. Kuila, A. Kumar Mishra, P. Khanra, N. Hoon Kim, J. Hee Lee, Recent advances
in the efficient reduction of graphene oxide and its application as energy storage
at 140 °C, and GO sample heat treated at 200 °C after an initial heat electrode materials, Nanoscale 5 (2013) 52–71, doi:10.1039/C2NR32703A.
treatment at 140 °C exhibit a non-porous structure and low surface ar- [10] D.R. D.reyer, S. Park, C.W. B.ielawski, R.S. R.uoff, The chemistry of graphene oxide,
eas of ∼ 2 m2 /g, whereas GO sample directly heat treated at 200 °C or Chem. Soc. Rev. 39 (2010) 228–240, doi:10.1039/B917103G.
[11] D. Chen, H. Feng, J. Li, Graphene Oxide: preparation, Functionalization, and Electro-
heat treated at 200 °C after initial drying at 110 °C shows a porous struc- chemical Applications, Chem. Rev. 112 (2012) 6027–6053, doi:10.1021/cr300115g.
ture. Both GO-200 and GO-110–200 samples exhibited type IV isotherms [12] R. Liu, T. Gong, K. Zhang, C. Lee, Graphene oxide papers with high wa-
with H3 hysteresis which are indicative of a mesoporous structure with ter adsorption capacity for air dehumidification, Sci. Rep. 7 (2017) 9761,
doi:10.1038/s41598-017-09777-y.
grove pores formed by exfoliation of GO layers. Other results suggest
[13] Z. Wang, Z. Tang, Z. Han, S. Shen, B. Zhao, J. Yang, Effect of drying conditions on
that exfoliation is primarily triggered by release of a substantial vol- the structure of three-dimensional N-doped graphene and its electrochemical per-
ume of CO2 /CO from decomposition of surface functionalities, but wa- formance, RSC Adv. 5 (2015) 19838–19843, doi:10.1039/C4RA15494K.
[14] C. Botas, P. Álvarez, C. Blanco, R. Santamaría, M. Granda, M.D. Gutiérrez,
ter vapor release helps to significantly enhance the exfoliation. These
F. Rodríguez-Reinoso, R. Menéndez, Critical temperatures in the synthesis of
results demonstrate that direct rapid heat treatment of GO at 200 °C (or graphene-like materials by thermal exfoliation–reduction of graphite oxide, Carbon
at higher temperatures) is the most effective pathway for developing 52 (2013) 476–485, doi:10.1016/j.carbon.2012.09.059.
a mesoporous structure. More experimental and simulation studies are [15] Y.N. Lv, J.F. Wang, Y. Long, C.A. Tao, L. Xia, H. Zhu, How graphene layers depend
on drying methods of graphene oxide, Adv. Mater. Res. 554–556 (2012) 597–600,
needed to investigate the mechanism of thermal reduction of carbon- doi:10.4028/www.scientific.net/AMR.554-556.597.
oxygen functionalities and the threshold pressure needed for GO exfoli- [16] T. Tene, M. Guevara, A. Valarezo, O. Salguero, F. Arias Arias, M. Arias, A. Scarcello,
ation. L.S. Caputi, C. Vacacela Gomez, Drying-Time Study in Graphene Oxide, Nanomate-
rials 11 (2021) 1035, doi:10.3390/nano11041035.
[17] J. Jo, S. Lee, J. Gim, J. Song, S. Kim, V. Mathew, M.H. Alfaruqi, S. Kim, J. Lim,
Declaration of Competing Interest J. Kim, Facile synthesis of reduced graphene oxide by modified Hummer’s method
as anode material for Li-, Na- and K-ion secondary batteries, R. Soc. Open Sci. 6
(2019) 181978, doi:10.1098/rsos.181978.
The authors declare no competing financial interest. [18] L. Klemeyer, H. Park, J. Huang, Geometry-dependent thermal reduction of
graphene oxide solid, ACS Mater. Lett. 3 (2021) 511–515, doi:10.1021/acsmate-
Data availability rialslett.0c00423.
[19] V.D. Ebajo, C.R.L. Santos, G.V. Alea, Y.A. Lin, C.-.H. Chen, Regenerable acidity of
graphene oxide in promoting multicomponent organic synthesis, Sci. Rep. 9 (2019)
Data will be made available on request. 15579, doi:10.1038/s41598-019-51833-2.
[20] M. Pedrosa, E.S. Da Silva, L.M. Pastrana-Martínez, G. Drazic, P. Falaras, J.L. Faria,
J.L. Figueiredo, A.M.T. Silva, Hummers’ and Brodie’s graphene oxides as photo-
Acknowledgements
catalysts for phenol degradation, J. Colloid Interface Sci. 567 (2020) 243–255,
doi:10.1016/j.jcis.2020.01.093.
The authors are grateful for the support provided by U.S. Depart- [21] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. Órfão, Modification of
the surface chemistry of activated carbons, Carbon 37 (1999) 1379–1389,
ment of Energy, National Energy Technology Laboratory (Cooperative
doi:10.1016/S0008-6223(98)00333-9.
Agreement DE-FE0031798). This document has not been subjected to [22] D. Mhamane, S.M. U.nni, A. Suryawanshi, O. Game, C. Rode, B. Hannoyer,
the review of the funding agencies and therefore does not necessarily S. Kurungot, S. Ogale, Trigol based reduction of graphite oxide to graphene
reflect their views. Any mention of trade names or commercial prod- with enhanced charge storage activity, J. Mater. Chem. 22 (2012) 11140–11145,
doi:10.1039/C2JM30582H.
ucts does not constitute endorsement or recommendation for use. Some [23] S.B. Singh, M. De, Thermally exfoliated graphene oxide for hydrogen storage, Mater.
material characterization tests were carried out in the Frederick Seitz Chem. Phys. 239 (2020) 122102, doi:10.1016/j.matchemphys.2019.122102.

10
S.B. Singh and S.A. Dastgheib Carbon Trends 10 (2023) 100251

[24] Y. Otake, R.G. Jenkins, Characterization of oxygen-containing surface complexes [41] H. Liang, M. Xu, Y. Bu, B. Chen, Y. Zhang, Y. Fu, X. Xu, J. Zhang, Confined in-
created on a microporous carbon by air and nitric acid treatment, Carbon N Y 31 terlayer water enhances solid lubrication performances of graphene oxide films
(1993) 109–121, doi:10.1016/0008-6223(93)90163-5. with optimized oxygen functional groups, Appl. Surf. Sci. 485 (2019) 64–69,
[25] U. Zielke, K.J. Hüttinger, W.P. Hoffman, Surface-oxidized carbon doi:10.1016/j.apsusc.2019.04.190.
fibers: I. Surface structure and chemistry, Carbon 34 (1996) 983–998, [42] B. Marchon, W.T. Tysoe, J. Carrazza, H. Heinemann, G.A. Somorjai, Reactive and
doi:10.1016/0008-6223(96)00032-2. kinetic properties of carbon monoxide and carbon dioxide on a graphite surface, J.
[26] S. Kundu, Y. Wang, W. Xia, M. Muhler, Thermal stability and reducibility of oxygen- Phys. Chem. 92 (1988) 5744–5749, doi:10.1021/j100331a039.
containing functional groups on multiwalled carbon nanotube surfaces: a quantita- [43] Y.-.H. Yu, Y.-.Y. Lin, C.-.H. Lin, C.-.C. Chan, Y.-.C. Huang, High-performance
tive high-resolution XPS and TPD/TPR study, J. Phys. Chem. C 112 (2008) 16869– polystyrene/graphene-based nanocomposites with excellent anti-corrosion proper-
16878, doi:10.1021/jp804413a. ties, Polym. Chem. 5 (2014) 535–550, doi:10.1039/C3PY00825H.
[27] Y.F. Jia, K.M. Thomas, Adsorption of cadmium ions on oxygen surface sites in acti- [44] V.R. Moreira, Y.A.R. Lebron, M.M. da Silva, L.V. de Souza Santos, R.S. Jacob,
vated carbon, Langmuir 16 (2000) 1114–1122, doi:10.1021/la990436w. C.K.B. de Vasconcelos, M.M. Viana, Graphene oxide in the remediation of nor-
[28] B.R. Puri, Surface complexes on carbons, Chem. Phys. Carbon 6 (1970) 191–282. floxacin from aqueous matrix: simultaneous adsorption and degradation process, En-
[29] C. Moreno-Castilla, M.A. Ferro-Garcia, J.P. Joly, I. Bautista-Toledo, F. Carrasco- viron. Sci. Pollut. Res. 27 (2020) 34513–34528, doi:10.1007/s11356-020-09656-6.
Marin, J. Rivera-Utrilla, Activated carbon surface modifications by nitric acid, hy- [45] N. Díez, A. Śliwak, S. Gryglewicz, B. Grzyb, G. Gryglewicz, Enhanced reduction of
drogen peroxide, and ammonium peroxydisulfate treatments, Langmuir 11 (1995) graphene oxide by high-pressure hydrothermal treatment, RSC Adv 5 (2015) 81831–
4386–4392, doi:10.1021/la00011a035. 81837, doi:10.1039/C5RA14461B.
[30] G. de la Puente, J.J. Pis, J.A. Menéndez, P. Grange, Thermal stability of oxy- [46] Y. Chen, Y. Niu, T. Tian, J. Zhang, Y. Wang, Y. Li, L.-.C. Qin, Microbial reduction
genated functions in activated carbons, J. Anal. Appl. Pyrolysis 43 (1997) 125–138, of graphene oxide by Azotobacter chroococcum, Chem. Phys. Lett. 677 (2017) 143–
doi:10.1016/S0165-2370(97)00060-0. 147, doi:10.1016/j.cplett.2017.04.002.
[31] C. Moreno-castilla, F. Carrasco-marín, F.J. Maldonado-hódar, J. Rivera-utrilla, Ef- [47] M. Acik, C. Mattevi, C. Gong, G. Lee, K. Cho, M. Chhowalla, Y.J. Chabal, The role of
fects of non-oxidant and oxidant acid treatments on the surface properties of intercalated water in multilayered graphene oxide, ACS Nano 4 (2010) 5861–5868,
an activated carbon with very low ash content, Carbon 36 (1998) 145–151, doi:10.1021/nn101844t.
doi:10.1016/S0008-6223(97)00171-1. [48] T.C. Taucher, I. Hehn, O.T. Hofmann, M. Zharnikov, E. Zojer, Understand-
[32] D.M. Nevskaia, A. Santianes, V. Muñoz, A. Guerrero-Ruı́z, Interaction of aqueous ing chemical versus electrostatic shifts in X-ray photoelectron spectra of or-
solutions of phenol with commercial activated carbons: an adsorption and kinetic ganic self-assembled monolayers, J. Phys. Chem. C 120 (2016) 3428–3437,
study, Carbon 37 (1999) 1065–1074, doi:10.1016/S0008-6223(98)00301-7. doi:10.1021/acs.jpcc.5b12387.
[33] M. Acik, G. Lee, C. Mattevi, A. Pirkle, R.M. Wallace, M. Chhowalla, K. Cho, Y. Cha- [49] Y. Liu, Y. Zhang, T. Zhang, Y. Jiang, X. Liu, Synthesis, characterization and cyto-
bal, The role of oxygen during thermal reduction of graphene oxide studied by toxicity of phosphorylcholine oligomer grafted graphene oxide, Carbon 71 (2014)
infrared absorption spectroscopy, J. Phys. Chem. C. 115 (2011) 19761–19781, 166–175, doi:10.1016/j.carbon.2014.01.025.
doi:10.1021/jp2052618. [50] K. Kristinaitytė, L. Dagys, J. Kausteklis, V. Klimavicius, I. Doroshenko, V. Pogorelov,
[34] H.-.K. Jeong, Y.P. Lee, M.H. Jin, E.S. Kim, J.J. Bae, Y.H. Lee, Ther- N.R. Valevičienė, V. Balevicius, NMR and FTIR studies of clustering of water
mal stability of graphite oxide, Chem. Phys. Lett. 470 (2009) 255–258, molecules: from low-temperature matrices to nano-structured materials used in in-
doi:10.1016/j.cplett.2009.01.050. novative medicine, J. Mol. Liq. 235 (2017) 1–6, doi:10.1016/j.molliq.2016.11.076.
[35] T. Kavinkumar, D. Sastikumar, S. Manivannan, Effect of functional groups on dielec- [51] G. Srinivas, Y. Zhu, R. Piner, N. Skipper, M. Ellerby, R. Ruoff, Synthesis of graphene-
tric, optical gas sensing properties of graphene oxide and reduced graphene oxide like nanosheets and their hydrogen adsorption capacity, Carbon 48 (2010) 630–635,
at room temperature, RSC Adv 5 (2015) 10816–10825, doi:10.1039/C4RA12766H. doi:10.1016/j.carbon.2009.10.003.
[36] M.P. Lavin-Lopez, A. Paton-Carrero, L. Sanchez-Silva, J.L. Valverde, A. Romero, In- [52] A. Kaniyoor, S. Ramaprabhu, A Raman spectroscopic investigation of graphite oxide
fluence of the reduction strategy in the synthesis of reduced graphene oxide, Adv. derived graphene, AIP Adv 2 (2012) 032183, doi:10.1063/1.4756995.
Powder Technol. 28 (2017) 3195–3203, doi:10.1016/j.apt.2017.09.032. [53] B. Ma, R.D. R.odriguez, A. Ruban, S. Pavlov, E. Sheremet, The correla-
[37] B. Marchon, J. Carrazza, H. Heinemann, G.A. Somorjai, TPD and XPS studies of O2 , tion between electrical conductivity and second-order Raman modes of laser-
CO2 , and H2 O adsorption on clean polycrystalline graphite, Carbon 26 (1988) 507– reduced graphene oxide, Phys. Chem. Chem. Phys. 21 (2019) 10125–10134,
514, doi:10.1016/0008-6223(88)90149-2. doi:10.1039/C9CP00093C.
[38] M.S. Shafeeyan, W.M.A.W. Daud, A. Houshmand, A. Shamiri, A review on surface [54] S. Claramunt, A. Varea, D. López-Díaz, M.M. Velázquez, A. Cornet, A. Cirera, The
modification of activated carbon for carbon dioxide adsorption, J. Anal. Appl. Py- importance of interbands on the interpretation of the raman spectrum of graphene
rolysis 89 (2010) 143–151, doi:10.1016/j.jaap.2010.07.006. oxide, J. Phys. Chem. C 119 (2015) 10123–10129, doi:10.1021/acs.jpcc.5b01590.
[39] Q.-.L. Zhuang, T. Kyotani, A. Tomita, DRIFT and TK/TPD analyses of surface oxy- [55] S. Liu, J. Wu, Z. Zhou, L. Gao, S. Luo, X. Xu, Z.M. Wang, Influence of graphite oxide
gen complexes formed during carbon gasification, Energy Fuels 8 (1994) 714–718, drying temperature on ultra-fast microwave synthesis of graphene, J. Mater. Sci.
doi:10.1021/ef00045a028. Mater. Electron. 24 (2013) 1298–1302, doi:10.1007/s10854-012-0923-2.
[40] M.T.H. Aunkor, I.M. M.ahbubul, R. Saidur, H.S.C. Metselaar, Deoxygenation of
graphene oxide using household baking soda as a reducing agent: a green approach,
RSC Adv 5 (2015) 70461–70472, doi:10.1039/C5RA10520J.

11

You might also like