ArticlePDF Available

Intercalating negatively charged pillars into graphene oxide sheets to enhance sulfonamide pharmaceutical removal from water

Authors:

Abstract and Figures

Herein, novel composite materials were prepared by intercalating functional pillars, i.e., pentafluorobenzene (PFB) and sodium 2,3,4,5,6-pentafluorobenzoate (PFBS), into graphene oxide (GO) sheets. It led to forming size hives and increased availability of intrinsic area of GO. The synthesized materials (GO-PFB and GO-PFBS) were investigated as adsorbents to eliminate sulfadiazine (SD) from aqueous solutions. The adsorption capacities of GO-PFBS (1002.21 μmol/g) and GO-PFB (564.17 μmol/g) were 6.37 and 3.59 times higher than that of GO (157.21 μmol/g), respectively. The adsorption of SD onto GO-PFBS decreased with increasing solution pH. Density functional theory (DFT) results revealed that the SD adsorption onto the adsorbents was exothermic, and the introduction of the carboxylate groups showed lower binding energy. It was found that hydrophobic interaction fully participates in the adsorption process, and the electrostatic complementation of hydrogen bonding further enhances the SD adsorption. Obtained results showed that intercalating functional rigid molecules as pillars to support GO sheets could improve its adsorption behavior.
Content may be subject to copyright.
Vol.:(0123456789)
1 3
https://doi.org/10.1007/s11356-022-20949-w
RESEARCH ARTICLE
Intercalating negatively charged pillars intographene oxide sheets
toenhance sulfonamide pharmaceutical removal fromwater
WeiWang1· ShiyiWang1· MohammadtaghiVakili2· YanWang1· ChangSun1· HaoruYang3· GuotaoXiao1·
MinjuanGong1· ShuangxiZhou1
Received: 31 March 2022 / Accepted: 16 May 2022
© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2022
Abstract
Herein, novel composite materials were prepared by intercalating functional pillars, i.e., pentafluorobenzene (PFB) and
sodium 2,3,4,5,6-pentafluorobenzoate (PFBS), into graphene oxide (GO) sheets. It led to forming size hives and increased
availability of intrinsic area of GO. The synthesized materials (GO-PFB and GO-PFBS) were investigated as adsorbents to
eliminate sulfadiazine (SD) from aqueous solutions. The adsorption capacities of GO-PFBS (1002.21μmol/g) and GO-PFB
(564.17μmol/g) were 6.37 and 3.59 times higher than that of GO (157.21μmol/g), respectively. The adsorption of SD onto
GO-PFBS decreased with increasing solution pH. Density functional theory (DFT) results revealed that the SD adsorption
onto the adsorbents was exothermic, and the introduction of the carboxylate groups showed lower binding energy. It was
found that hydrophobic interaction fully participates in the adsorption process, and the electrostatic complementation of
hydrogen bonding further enhances the SD adsorption. Obtained results showed that intercalating functional rigid molecules
as pillars to support GO sheets could improve its adsorption behavior.
Keywords Graphene oxide· Rigid molecules· Sulfadiazine· Density functional theory· PPCPs
Introduction
Antibiotics are one of the widely applied PPCPs in humans
and animals to treat/prevent infections (Li etal. 2018; Sar-
mah etal. 2006). Sulfonamides (SNs) are one of the com-
monly used antibiotics (Serna-Carrizales etal. 2021; Wei
etal. 2019). In recent years, massive utilization has led to
the presence of SNs in the environment and gradual accu-
mulation in living organisms, which can threaten human
health and environmental safety (Szymańska etal. 2019).
Traditional wastewater treatment plants cannot remove the
SNs completely, and thus these compounds could appear
in natural waters. Different methods have been developed
to eliminate SNs from aqueous solutions such as biodegra-
dation, electrochemical oxidation, membrane filtration, and
photocatalysis (Zhuang etal. 2020). Low manufacturing
cost, simple operating requirements, high selectivity, envi-
ronmentally friendly, and easy control make the adsorption
more advantageous. However, it is necessary to develop an
effective adsorbent that is inexpensive, easy to prepare, and
regenerable.
Graphene (an atomic layer of sp2 bonded carbon atoms,
tightly bound in a honeycomb or hexagonal lattice) was
first identified in 2004 through mechanical exfoliation
by Novoselov and Geim (Novoselov etal., 2004). Due to
unique physicochemical properties (e.g., high hydropho-
bicity, optical transmittance, chemical inertness, high cur-
rent density, and high thermal conductivity) (Priyadarsini
etal. 2018), it is widely applied in tissue scaffolds (Nayak
etal. 2011), energy storage (Wu etal. 2021), biosensing (Lu
etal. 2009), drug delivery (Sun etal. 2008), antibacterial
compositions (Perreault etal. 2015), and so on. Also, such
material has attracted growing attention as an effective and
potential adsorbent in environmental remediation because
of its high surface area (2630 m2/g), fast adsorption rate,
Responsible Editor: George Z. Kyzas
* Shuangxi Zhou
15105176897@126.com
1 State Key Laboratory ofPlateau Ecology andAgriculture,
Qinghai University, Xi’ning810016, QinghaiProvince,
China
2 Green Intelligence Environmental School, Yangtze Normal
University, Chongqing408100, China
3 Colorado College, ColoradoSprings, CO80903, USA
/ Published online: 24 May 2022
Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
and high adsorption capacity (Verma etal. 2021). However,
the aggregation and hard dispersion of graphene in water
affect its adsorption performance (Gunes etal. 2021). The
methods, i.e., the introduction of abundant oxygen elements
to prepare graphene oxides (GO), and hybridization with
other nano-materials, could promote its application and
help overcome aforementioned problems. For example, the
GO-MnO2 nanocomposite performed better in methylene
blue and methyl orange removal compared to GO (Verma
etal. 2021). Travlou etal. prepared a chitosan-GO composite
adsorbent and applied it for reactive dye adsorption (Travlou
etal. 2013).
In recent years, connecting GO sheets to excavate its
effective surface area has also been considered a feasible
strategy. For instance, the decafluorobiphenyl (a rigid mol-
ecule) was successfully inserted into hydroxylated carbon
nanotubes by a nucleophilic substitution reaction. The
adsorption capacity increased by 2.1 and 2.7 times com-
pared with the adsorption capacity of CNTs powder for car-
bamazepine and tetracycline, respectively (Shan etal. 2018).
In addition, 1,4-phenylenebisboronic acid was successfully
intercalated into rGO-Br and greatly improved the conduc-
tion of UO2+ (Wang etal. 2021c).
The intercalation of rigid molecules into GO sheets could
enhance the availability of intrinsic area (Shan etal.2017).
There is almost no report that using GO (prepared by inter-
calating pre-designed functional pillars) to remove con-
taminants, e.g., pharmaceutical and personal care products
(PPCPs), for the PPCPs in an aqueous solution could be
seriously affected by the solution pH (Berhane etal. 2015).
The use of pillars (containing carboxylate groups) for the
connection of GO sheets would significantly increase the
adsorption of positively charged PPCPs. Therefore, it is a
practical and feasible strategy to enhance the adsorption
capacity of nano-material by a nucleophilic substitution
reaction between rigid molecules with carboxylates and GO
layers to form functionally layered porous structure. Due to
the typical electron-donating of amino groups on SNs, the
introduction of strong electron-acceptor groups into pillars
would seriously improve the adsorption of SNs.
Therefore, the carboxyl functional group was pre-design-
ing on the pillars, which was intercalated into GO sheets
in this work. The rigid molecules, i.e., pentafluorobenzene
(PFB) and sodium 2,3,4,5,6-pentafluorobenzoate (PFBS,
carbonylated pillar), were reacted with hydroxyl groups on
GO by a nucleophilic substitution reaction to obtain the GO-
based materials with layered porous structutre and named as
GO-PFB and GO-PFBS, respectively. Meanwhile, one of
the SNs, i.e., sulfadiazine (SD), was chosen as an objective
contaminate to estimate the adsorption performance of the
prepared GO-based materials. Comparative adsorption stud-
ies between GO, GO-PFB, and GO-PFBS were conducted
to obtain the optimal nano-materials in SD removal. The
adsorption performance of the adsorbents was investigated
by studying the adsorption kinetics, isotherms, effects of
solution pHs, and ionic strength. Experimental data and DFT
calculation proposed the possible adsorption mechanism.
Additionally, the popularized application of adsorbent for
sulfonamide removal was studied.
Materials andmethods
Materials andreagents
Graphene oxide (GO, TNGO-3) was purchased from
Chengdu Organic Chemicals Co. Ltd. Pentafluorobenzene
(PFB, 98%) and sodium 2,3,4,5,6-pentafluorobenzoate
(PFBS, 95%) were obtained from Shanghai Aladdin Bio-
chemical Technology Co., Ltd. and J&K Scientific Co., Ltd.,
respectively. Sulfonamides, i.e., sulfadiazine (SD, 98%), sul-
famethoxazole (SMZ, 96%), sulfamethazine (STZ, 99%),
sulfapyridine (SPD, 98%), and sulfachloropyridazine (SPZ,
98%), were supplied by Aladdin Biochemical Technology.
Potassium carbonate (K2CO3), N,N-dimethylformamide
(DMF), methanol (99.9%), and acetonitrile (99.9%) were
purchased from Sinopharm Co., Ltd.
Preparation ofGO‑PFB andGO‑PFBS
The nucleophilic substitution reaction was referred to in the
previous research (Shan etal. 2018). Initially, 0.1g GO was
fully dispersed in 50mL DMF assisted by ultrasonic for
15min in an ice bath. Then, 20mmol/L of PFB or PFBS and
0.60g of K2CO3 were added into the dispersion liquid and
ultrasonicated for 10min in an ice bath. Next, the gas circuit
filled with N2 was penetrated into the mixture for 10min
to crow out air. And then, the glass vial was kept at 65°C
and stirred at 500rpm for 48h. In the next step, the cooled
mixture was washed with HCl (1M), H2O, methanol, and
CH2Cl2 in turns. Finally, the solid was vacuumed in a liquid
nitrogen bath for 10min and then placed into a vacuum
drying oven under 50°C for 48h. The obtained materials
were denoted as GO-PFB and GO-PFBS for the different
intercalations of steel molecules. The synthetic scheme of
GO-PFB and GO-PFBS is shown in Fig.1.
Characterization
The surface morphology of the materials was observed
by a field emission scanning electron microscope (FE-
SEM, JSM-7900F, Japan). The pore size distribution and
Brunauer–Emmett–Teller (BET) surface area were recorded
by an adsorption instrument (Micromeritics, USA) using
nitrogen gas at 77K. Functional groups on the prepared
adsorbents were characterized by an FTIR spectrometer
72546 Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
(PerkinElmer, USA). Raman spectra were obtained with
an excitation line at 532nm (Horiba, France). The zeta
potential was measured at different pH values (1–9) using a
Zetasizer instrument (Brookhaven, USA) and the point zero
charge was calculated by the curve between pH vs surface
electrical property.
Adsorption experiments andanalysis
The kinetic experiments were carried out in 150-mL glass
vials, where 0.1g/L of the prepared materials were mixed
with 90mL of pharmaceutical solution (0.2mmol/L) at pH
5 with a shaking rate of 180rpm under 25°C for 30h. The
sample was taken out by a syringe with a filter membrane
(PES, 0.45μm) at interval time. The isotherm experiments
were performed by mixing 0.1g/L of the adsorbents and sul-
fonamide solutions (0.02–0.44mmol/L) at pH 5 180rpm for
24h. The favorable adsorption conditions were obtained by
investigating the effects of solution pH (1–11) adjusted with
1M HCl or NaOH and the effect of ion strength by varying
NaCl/CaCl2 concentration ranging from 20 to 80mg/L.
The concentration of sulfonamides was measured by
HPLC (Shimadzu LC-20, Japan). SD was separated by a
C18 column (4.6 × 250mm, Shimadzu, Japan). The injected
volume and the mobile phase flow rate, i.e., methanol/0.1%
formic acid (50:50, v/v) under 40°C, were 10 µL and
1.0mL/min, respectively.
Computational method
The optimal structures of SD and adsorbents were mod-
eled by Gaussian 09 program (Frisch etal. 2010). B3LYP
hybrid functional at the 6-311G level was used to optimize
the frequency calculation of the optimal structures (Wang
etal. 2021a). The polarizable continuum model based on
the integral equation formalism variant (IEFPCM) (Tomasi
etal. 1991) was adopted, and water was used as the solvent
in the system. Ebinding values between GO-PFBS and SD
were calculated using Eq.(1):
To further explore the interaction between SD and the
prepared materials, electrostatic potential (ESP) was calcu-
lated by the Multiwfn program (Lu and Chen 2012) and the
VMD 1.9 software package (Humphrey etal. 1996).
Regeneration experiment
Regeneration experiment was carried out by mixing
0.1g/L GO-PFBS and 0.2mmol/L SD solution at pH 5 for
24h. Then the adsorbent was separated by a centrifuge at
8000rpm for 10min. The collected saturated GO-PFBS was
immersed in the mixture of CH3OH and NaOH (with a mass
ratio of 10:1) and stirred at 180rpm for 5h at 25°C (Meng
etal. 2020). The eluent was abandoned, and the humid GO-
PFBS was dried at 60°C overnight for the next cycle.
Results anddiscussion
Characterization
The SEM images of GO, GO-PFB, and GO-PFBS are pre-
sented in Fig.2a–c. Compared with GO, the combined, shrill
wrinkled sheets (similar to a broken tree trunk) can be seen
in the complete view of GO-PFB and GO-PFBS. This might
be caused by the intercalation of pillars into GO sheets. The
(1)
Ebinding =Ebindingcomplex ESD Eadsorbent
Fig. 1 Synthetic scheme of GO-
PFB and GO-PFBS
72547Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
GO sheets, connected by the rigid molecules through the
nucleophilic substitution reaction, are attached and formed
numerous pores, which would enhance the adsorption capac-
ity of GO.
The prepared materials’ specific surface area and pore size
distribution are shown in Table1 and Fig.2d, respectively.
It was apparent that both of the modified adsorbents dis-
played a relatively high BET specific surface area of 40.23
m2/g (GO-PFB) and 26.10 m2/g (GO-PFBS) compared with
that of GO (7.94 m2/g). Also, the total pore volume showed
similar results, and it indicated that reactions between rigid
molecules and GO effectively increased the granulation and
decreased the aggregation of GO sheets (Shan etal. 2018).
In addition, although the mesoporous volume of GO-PFB
and GO-PFBS was dramatically increased, GO-PFBS
showed relatively concentrated pore sizes within 3–35nm,
and GO-PFB displayed a wide range from 1.2 to 185nm.
This difference might be attributed to the inadequate reac-
tion between rigid molecules and GO or the formed tilt pillar
by the vacancy in the benzene ring of PFB compared to fully
filled with fluorine atoms in PFBS.
The FTIR spectra of chemicals and related prepared
materials are shown in Fig.2e. For GO, the peak at
3434 cm−1 corresponded to the stretching vibration of
Fig. 2 The SEM images of GO
(a), GO-PFB (b), and GO-PFBS
(c). Pore size distribution (d),
the FTIR spectra (e), and the
Raman spectra (f) of GO, GO-
PFB, and GO-PFBS
Table 1 Specific surface area and pore volume of GO, GO-PFB, and
GO-PFBS
* Pore volume was calculated by the DFT method
Materials BET surface area (m.2/g) Total pore
volume
(cm.3/g)
GO 7.94 0.010
GO-PFB 40.23 0.101
GO-PFBS 26.10 0.047
72548 Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
hydroxyl groups, and the peak at 1626 cm−1 might accord
to the skeletal stretching vibration of the aromatic ring
and 1726 cm−1 assigned to the stretching vibration of
O-C = O (Han et al. 2021; Li etal. 2020; Singh etal.
2009). In case of PFB and PFBS molecules, the peaks
at 945 cm−1, 1099 cm−1, 987 cm−1, and 1118 cm−1 are
due to the C-F vibration (Singh etal. 2009). The peaks at
1641 cm−1 and 1656 cm−1 are assigned to the stretching
vibrate of C = C in the benzene ring. Compared to PFB,
some new peaks at 3480 cm−1, 1475 cm−1, 1295 cm−1,
1391 cm−1, and 1603 cm−1 were observed in PFBS (due to
the presence of carboxyl groups in PFBS structure) (Han
etal. 2021; Li etal. 2020; Singh etal. 2009). Moreover,
the disappearance of the hydroxyl group’s peak at
3434 cm−1 on GO and appearance of weakly characteristic
peaks of C-F, i.e., 1045 cm−1 and 987 cm−1 for GO-PFB
and 1053 cm−1 and 983 cm−1 for GO-PFBS, show the
successfully executed intercalation of rigid molecules to
the GO structure.
The ratio between D band and G band (R = ID/IG) in
Raman spectra was used to estimate the degree of defects for
carbon materials, where the D band represents the structural
defects and the G band shows the first-order scattering of
E2g vibrational model (Verma etal. 2021; Stankovich etal.
2007). As shown in Fig.2f, the prepared materials (GO-
PFB and GO-PFBS) exhibited a higher R value than that of
GO, indicating that the disorder of prepared materials was
enhanced (Shan etal. 2018). This demonstrated that rigid
molecules were successfully intercalated into GO sheets and
formed porous structure, which was advantageous to capture
contaminants.
Comparison oftheprepared materials
Adsorption kinetic
To investigate the adsorption mechanism and process of SD
onto GO, GO-PFB, and GO-PFBS, kinetic experiments were
conducted. The values of kinetic constants were obtained
using Webber-Morris intra-particle, pseudo-first-order,
pseudo-second-order, and Elovich models (Fig.3a, Fig.S1,
and TableS1). The removal of SD by the GO-PFB and GO-
PFBS increased sharply in the initial 12h of the adsorption
process, and then all reached the adsorption equilibrium
after 24h, which was faster than the intercalation of rigid
molecule DFB to CNT-OH in CBZ and TC removal (Shan
etal. 2018). It was apparent that the SD adsorption capacity
of GO was significantly enhanced after modification. These
findings suggest that the introduction of rigid molecules
has successfully connected GO sheets and formed cubicles.
Moreover, the adsorption amount of SD onto GO-PFBS
(536.40μmol/g) was 1.41 times higher than that of GO-
PFB (379.93μmol/g) (Fig.3a and TableS1). It could be
due to introducing the rigid molecule containing carboxyl
groups to the GO sheets, significantly improving its adsorp-
tion capacity.
As shown in Fig.S1, the intra-particle curves could be
roughly divided into three stages. None of the fitting lines
pass through the origin, indicating that intra-particle dif-
fusion was not the dominantly rate-limiting factor and the
adsorption process was comprehensively affected by bound-
ary layer and internal pore structures (Ma etal. 2019). On
the other hand, the experimental data fitted well with the
pseudo-second-order model based on the R2 value (> 0.98).
This indicated that the adsorption process could be involved
in chemisorption, which can be proved by the high R2 value
0510 15 20 25 30
0
100
200
300
400
500
600
GO
GO-PFB
GO-PFBS
First-order model
Second-order model
yticapacnoitprosdA (
μ
g/lom )
Time (h)
(a)
0100 200300 400
0
200
400
600
800
1000
GO
GO-PFB
GO-PFBS
Langmuir model
Freundlich model
yticapacnoitprosd
A(
μ
g/lo
m)
C
e
(
μ
mol/L)
(b)
Fig. 3 Pseudo-first-order and pseudo-second-order kinetic models (a), Langmuir and Freundlich isotherm models (b) for SD adsorption onto the
GO, GO-PFB, and GO-PFBS
72549Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
of the Elovich model (0.99). The SD adsorption capacity
of GO, GO-PFB, and GO-PFBS were found to be 52.53,
379.92, and 536.40μmol/g, respectively. Therefore, the rigid
molecule PFB has successfully built cubicles in GO sheets
promoting SD diffusion. The carboxylic acid-functionalized
rigid molecule (PFBS) promoted SD adsorption by the spe-
cific electronic complementation hydrogen bonding affinity.
Adsorption isotherm
To determine the maximum adsorption capacity of the
adsorbents and provide information on the interaction of
SD with GO, GO-PFB, and GO-PFBS, relevant experimen-
tal data were fitted to the Langmuir and Freundlich adsorp-
tion isotherm models, and the results are shown in Fig.3b
and Table2. Compared to the Langmuir model (R2 ≥ 0.92),
the equilibrium adsorption data was described better by the
Freundlich model (≥ 0.97), indicating that the SD adsorp-
tion occurred on heterogeneous surfaces (He etal. 2017).
In addition, the n values for SD adsorption onto the GO-
PFB and GO-PFBS were within 2.19–3.38, confirming
the favorability of the adsorption process (Ai etal. 2019).
According to the Langmuir model, the maximum adsorp-
tion capacity of GO-PFB (1002.21μmol/g) and GO-PFBS
(564.17μmol/g) was 6.37 and 3.59 times higher than that of
GO (157.12μmol/g), respectively.
Effects ofsolution pH andionic strength
Effects ofsolution pH
Solution pH is an important controlling factor in the adsorp-
tion process as it can influence the adsorbate’s dissociation
and the adsorbent’s surface charge, thus affecting the adsorp-
tion mechanism (Wu etal. 2020). For understanding the net
surface charge on the adsorbent, the optimum pH value for
effective adsorption, and the adsorption mechanism, evalu-
ating the adsorbent’s pHpzc could be beneficial (Egbedina
etal. 2021). It has been reported that the pH < pHpzc is
favorable for the adsorption of anionic adsorbates where the
adsorbent surface has a positive charge (Yagub etal. 2014).
The pHpzc of GO-PFBS was found to be 1.9 (Fig.S2). At
pH < 1.9, protonation of SD and functional groups on the
adsorbent could prevent the migration of cation SD to GO-
PFBS. Due to the deprotonation of adsorption sites at alka-
line conditions, the SD molecule was in anionic form, and
the surface charge of the GO-PFBS was found to be nega-
tive. Therefore, the strong electrostatic repulsion between
the negative GO-PFBS and SD could seriously inhibit the
adsorption of SD from aqueous solution onto GO-PFBS,
which significantly decreased the SD adsorption capacity
of the adsorbent (Hu etal. 2019).
The effect of the initial pH value of SD solution on the
adsorption capacity of GO-PFBS is shown in Fig.4a. The
adsorption capacity of SD onto GO-PFBS presented an
upward trend with the increasing solution pH values, and
Table 2 The parameters of Langmuir and Freundlich model for SD
Langmuirmodel qe=qmCe∕(1b+Ce)
Freundlichmodel
q
e
=K
f
C
e
1n
Models Parameters GO GO-PFB GO-PFBS
Langmuir model qm (μmol/g) 157.21 564.17 1002.21
R.20.940 0.920 0.960
b (L/μmol) 0.0053 0.037 0.014
Freundlich model Kf (μmol1−1/n L1/n
g.−1)
5.06 98.82 60.88
n1.93 3.38 2.19
R.20.980 0.970 0.970
0
100
200
300
400
500
600
11
9
753
yticapacnoitprosdA (
μ
g/lom )
Initial pH
1
(a)
0
100
200
300
400
500
600
805020
yticapacnoitprosd
A(
μ
g/lo
m)
Na
+
/Ca
2+
concentration (mg/L)
Blank
Na+
Ca2+
0
(b)
Fig. 4 The influence of solution pH (a) and ion strength (b) on the SD removal by GO-PFBS
72550 Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
the maximum adsorption capacity of SD (564.91μmol/g)
occurred at pH 5. It could be attributed that the main species
of SD is in neutral form under this pH value and thus could
be easily attracted via the electron-accepting of carboxylates
on PFBS pillars, hydrogen bond interaction, and electrostatic
interaction.
Effects ofionic strength
Figure4b shows the effects of ionic strength on the SD
removal by GO-PFBS. With an increase in both salt concen-
trations, the SD adsorption capacity of GO-PFBS showed a
decreasing trend. Both the two Na+ and Ca2+ ions weakened
the adsorption capacity of GO-PFBS. The SD adsorption
capacity reached 412.20μmol/g and 393.73μmol/g when
the salt concentration increased to 80mg/L. This indicated
that cations could occupy the adsorption sites on GO-PFBS,
and due to the higher valence, the Ca2+ performed a slightly
stronger restraint/inhibition than Na+. To verify the conclu-
sion, the same concentrations of Na+ and Ca2+ were added
to the mixture of SD and GO-PFB under the same condition,
respectively. As shown in Fig.S3, the adsorption capacity
of SD by GO-PFB was almost no change, indicating that
Na+ and Ca2+ might occupy the adsorption sites supported
by carboxylate groups of GO-PFBS. In addition, the point
zero charges (pHpzc, Fig.S2) of GO-PFBS were calculated
to about pH 1.9, and the strongest negative charge was under
pH 5, demonstrating that the surface of GO-PFBS might be
occupied by cations (Na+ or Ca2+) and thus decrease the
adsorption of SD.
Regeneration andreuse
FigureS4 illustrates the reusability of GO-PFBS after 5
adsorption–desorption cycles. The results revealed that the
SD adsorption capacity of the GO-PFBS at the first cycle
(535.46μmol/g) decreased to 328 μmol/g at the second
cycle. This reduction could be attributed to the decrease
in porosity of GO-PFBS, saturation, and unavailability of
free functional groups on the GO-PFBS. With increasing
regeneration cycles, the adsorption capacity of the adsorbent
remained almost stable and did not change much.
Adsorption mechanism
The density functional theory (DFT) model was used to cal-
culate the adsorption configurations and the binding energy
between SD and pillars, i.e., PFB and PFBS (Fig.5). The
values of ΔEbinding for SD adsorbed onto PFB and PFBS
were − 4.68 and − 7.09kJ/mol (Fig.5a and b), respectively,
indicating that the adsorption was an exothermic process and
the geometry of GO-PFB-SD and GO-PFBS-SD was stable.
The ΔEbinding of GO-PFBS-SD was lower than GO-PFB-
SD, suggesting that GO-PFBS performed a stronger affinity
Fig. 5 DFT optimized structures of SD adsorption on GO-PFB (a) and GO-PFBS (b). ESP mapping of GO-PFB (c) and GO-PFBS (d) after SD
adsorption
72551Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
for SD than GO-PFB, consistent with their experimental
adsorption capacity (Fig.3b).
As shown in Fig.5c and d, the electrostatic potential
(ESP) of GO-PFB and GO-PFBS after SD adsorption was
marked red and blue, representing nucleophilic and electro-
philic groups, respectively. Compared with PFB, the surface
electrostatic potential of PFBS changed to completely nega-
tive for the introduction of carboxylate and showed a strong
affinity for the electrophilic pollutant. For the adsorption
configurations of GO-PFBS-SD (Fig.5d), the carboxylate of
PFBS becomes the dominant active site for attracting elec-
trophilic amino on SD via the electrostatic complementation
of hydrogen bonding (Wang etal. 2021b), while the surface
electrostatic potential of PFB (Fig.5c) has been affected
by the vacancy on the benzene ring and thus weakened the
electrophilic ability. Hence, assembling carboxylate groups
onto pillars enhanced the nucleophilic ability of GO-PFBS
and made it more feasible to capture objective SD.
It was reported that SD possessed two dissociation con-
stants with pKa1 = 2.0 and pKa1 = 6.8 (He etal. 2021), which
were mainly affected by the groups of amino and sulfonamide
in SD molecules. As shown in Fig.S2, the pHpzc of GO-PFBS
was about 1.9, demonstrating that the surface electrical prop-
erty was weakly positive at pH 1 and negative at alkalinity.
It was evident that GO-PFBS showed low electricity at pH 1
(about 0.70mV), probably attributed to the protonation of
carboxylate. On the other hand, the adsorption amounts of SD
by GO-PFBS were frequently affected by solution pH, espe-
cially in alkaline conditions for the previous results, indicating
that the electrostatic interaction might dominate the adsorption
process (Wu etal. 2020). DFT results showed that carboxyl
groups on GO-PFBS exhibited a stronger attracting ability for
SD than GO-PFB. Meanwhile, the hydrophobic interaction
would participate in the whole process of SD removal. The
proposed adsorption mechanism is shown in Fig.6.
At low pHs (pH < 2.0), the special groups of amino
(-NH2) and sulfonamide (-SO2NH-) on the aromatic rings
of SD could function as strong π-electron receptors, and
thus the π+-π electron donor–acceptor interaction would
dominate the adsorption of positive SD+ (He etal. 2021;
Peiris etal. 2017; Zhang etal. 2016). The alkaline solution
would promote the deprotonation of -SO2NH- and form the
anionic SD (SD), and thus, the migration of SD would be
seriously inhibited by the strongly negative surface of GO-
PFBS, and the hydrophobic interaction force might rarely
participate in the adsorption of SD. When the solution pH
was adjusted to 5, the exposed affine carboxylates of GO-
PFBS could appeal to the migration of SD molecules for
its hydrogen element of amino (-NH2) via the electrostatic
interaction and hydrogen bond. It was indicated that the elec-
trostatic complementation of hydrogen bonding participated
in the adsorption process, which was proved in our previ-
ous report (Wang etal. 2021b). In addition, the electrostatic
repulsion between SD and negative GO-PFBS disappeared
and thus promoted the diffusion of SD in GO-PFBS. Mean-
while, the hydrophobic interaction and π-π stacking also
participate in the adsorption process.
Application ofGO‑PFBS
From the above results, the prepared materials via the suc-
cessful intercalation of rigid molecules into GO sheets
exhibited excellent performance in SD removal. It is attrac-
tive to investigate the wider application of GO-PFBS in
other sulfonamide removal. Thus, four typical sulfonamides,
including sulfamethoxazole (SMZ), sulfamethazine (STZ),
sulfapyridine (SPD), and sulfachloropyridazine (SPZ), were
used as the objectives, and the results are shown in Fig.7.
Surprisingly, the adsorption capacity of SPZ, SMZ, SPD,
and STZ onto GO-PFBS has been improved by 18, 10, 3.7,
and 3.4 times compared with that of GO, respectively. The
prepared material’s outstanding properties revealed that
introducing rigid molecules to broaden the inner surface
and create cubicles is a rational design in future applications.
Conclusions
Intercalation of rigid molecules into GO, GO-PFB, and
GO-PFBS through the nucleophilic aromatic substitution
reaction with pillars of PFB and PFBS was performed
successfully. A series of characterization results revealed
Fig. 6 The proposed adsorption
mechanism of SD on GO-PFBS
72552 Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
that the stacked GO sheets were prospectively isolated and
supported. Both of the prepared materials showed an out-
standing performance in SD removal, and the introduction
of carboxylate functional groups improved the adsorption
capacity of GO. According to the Langmuir model, the
adsorption capacity of GO-PFBS, GO-PFB, and GO was
1002.21, 564.17, and 157.21μmol/g, respectively. In addi-
tion, the solution pH played a critical role in SD removal by
altering the surface charges of GO-PFBS and SD. The DFT
calculation revealed that the adsorption process was spon-
taneous and exothermic, and GO-PFBS-SD showed lower
binding energy than GO-PFB-SD. The possible adsorption
mechanism was proposed that electronic complementation
of hydrogen bonding, π+-π electron donor–acceptor interac-
tion, and hydrophobic action participated in SD removal.
Moreover, the prepared GO-PFBS also showed an excellent
removal performance for other four sulfonamide pharma-
ceuticals. In brief, it is a very effective strategy to solve the
low surface area utilization of GO and enhance the func-
tionalization of nano-materials via inlaying functionally
rigid molecules and can be wide application in the future.
Supplementary Information The online version contains supplemen-
tary material available at https:// doi. org/ 10. 1007/ s11356- 022- 20949-w .
Author contribution Wei Wang: conceptualization, writing (original
draft), funding acquisition. Shiyi Wang: investigation. Mohammadtaghi
Vakili: writing (review and editing). Yan Wang: formal analysis. Chang
Sun: methodology. Haoru Yang: writing (review and editing). Guotao
Xiao: data curation. Minjuan Gong: validation. Shuangxi Zhou: con-
ceptualization, supervision, writing (original draft).
Funding This work was financially supported by the Key Youth Foun-
dation of Qinghai University (grant number: 2020-QGY-2).
Data availability All data generated or analyzed during this study are
available from the corresponding author on reasonable request.
Declarations
Ethics approval and consent to participate Not applicable.
Consent for publication Not applicable.
Competing interests The authors declare no competing interests.
0510 15 20 25 30
0
100
200
300
400
500
yticapacnoitprosdA (
µ
g/lom )
Time (h)
GO
GO-PFBS
(a)
0510 15 20 25 30
0
140
280
420
560
700
yticapacnoitprosd
A(
µ
g/lo
m)
Time (h)
GO
GO-PFBS
(b)
0510 15 20 25 30
0
150
300
450
600
yticapacnoitprosdA (
µ
g/lom )
Time (h)
GO
GO-PFBS
(c)
0510 15 20 25 30
0
80
160
240
320
yticapacnoitprosdA(
µ
g/lom)
Time (h)
GO
GO-PFBS
(d)
Fig. 7 The application of GO-PFBS in sulfonamide removal of SPZ (a), SMZ (b), SPD (c), and STZ (d) compared with GO
72553Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
References
Ai T, Jiang X, Liu Q, Lv L, Wu H (2019) Daptomycin adsorption on
magnetic ultra-fine wood-based biochars from water: Kinetics,
isotherms, and mechanism studies. Bioresource Technol 273:8–
15. https:// doi. org/ 10. 1016/j. biort ech. 2018. 10. 039
Berhane TM, Levy J, Krekeler MPS, Danielson ND, Stalcup A (2015)
Sorption-desorption of carbamazepine by palygorskite–montmo-
rillonite (PM) filter medium. J Hazard Mater 282:183–193. https://
doi. org/ 10. 1016/j. jhazm at. 2014. 09. 025
Egbedina AO, Adebowale KO, Olu-Owolabi BI, Unuabonah EI, Ades-
ina MO (2021) Green synthesis of ZnO coated hybrid biochar
for the synchronous removal of ciprofloxacin and tetracycline in
wastewater. RSC Adv 11:18483–18492. https:// doi. org/ 10. 1039/
D1RA0 1130H
Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb,
M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci,
B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratch-
ian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J.
L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J.,
Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven,
T., Montgomery, J. A., Peralta Jr., J. E., Ogliaro, F., Bearpark,
M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N.,
Keith, T., Kobayashi, R., Normand, J., Raghavachari, K., Rendell,
A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N.,
Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V.,
Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev,
O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin,
R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P.,
Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Fores-
man, J. B., Ortiz, J. V., Cioslowski, J.Fox, D. J (2010) Gaussian
09. Revison B. 01, Gaussian, Inc. Wallingford CT
Gunes B, Jaquet Y, Sánchez L, Pumarino R, Mcglade D, Quilty B,
Morrissey A, Gholamvand Z, Nolan K, Lawler J (2021) Activated
Graphene Oxide-Calcium Alginate Beads for Adsorption of Meth-
ylene Blue and Pharmaceuticals. Materials 14:6343. https:// doi.
org/ 10. 3390/ ma142 16343
Han L, Kong Y, Huang M (2021) Magnetic properties and crystal
structures of two copper coordination compounds with pentafluor-
obenzoate ligand. Inorg Chim Acta 514:120019. https:// doi. org/
10. 1016/j. ica. 2020. 120019
He J, Li Y, Wang C, Zhang K, Lin D, Kong L, Liu J (2017) Rapid
adsorption of Pb, Cu and Cd from aqueous solutions by
β-cyclodextrin polymers. Appl Surf Sci 426:29–39. https:// doi.
org/ 10. 1016/j. apsusc. 2017. 07. 103
He X, Li J, Meng Q, Guo Z, Zhang H, Liu Y (2021) Enhanced adsorp-
tion capacity of sulfadiazine on tea waste biochar from aqueous
solutions by the two-step sintering method without corrosive acti-
vator. J Environ Chem Eng 9:104898. https:// doi. org/ 10. 1016/j.
jece. 2020. 104898
Hu S, Zhang Y, Shen G, Zhang H, Yuan Z, Zhang W (2019) Adsorp-
tion/desorption behavior and mechanisms of sulfadiazine and
sulfamethoxazole in agricultural soil systems. Soil Tillage Res
186:233–241. https:// doi. org/ 10. 1016/j. still. 2018. 10. 026
Humphrey W, Dalke A, Schulten K (1996) VMD: visual molecular
dynamics. J Mol Graph 14(33–8):27–28. https:// doi. org/ 10. 1016/
0263- 7855(96) 00018-5
Li S, Shi W, Liu W, Li H, Zhang W, Hu J, Ke Y, Sun W, Ni J (2018)
A duodecennial national synthesis of antibiotics in China’s major
rivers and seas (2005–2016). Sci Total Environ 615:906–917.
https:// doi. org/ 10. 1016/j. scito tenv. 2017. 09. 328
Li Z, Mao Z, Hu C, Li Q, Chen Z (2020) Fluoro-functionalized sta-
tionary phases for electrochromatographic separation of organic
fluorides. J Chromatogr A 1625:461269. https:// doi. org/ 10. 1016/j.
chroma. 2020. 461269
Lu T, Chen F (2012) Quantitative analysis of molecular surface based
on improved Marching Tetrahedra algorithm. J Mol Graphics
Modell 38:314–323. https:// doi. org/ 10. 1016/j. jmgm. 2012. 07. 004
Lu CH, Hua, Yang HH, Hao, Zhu CH, Ling, Chen X, Chen GH (2009)
A Graphene Platform for Sensing Biomolecules. Angewandte
Chemie 121:4879–4881. https:// doi. org/ 10. 1002/ ange. 20090 1479
Ma H, Yang J, Gao X, Liu Z, Liu X, Xu Z (2019) Removal of chro-
mium (VI) from water by porous carbon derived from corn straw:
Influencing factors, regeneration and mechanism. J Hazard Mater
369:550–560. https:// doi. org/ 10. 1016/j. jhazm at. 2019. 02. 063
Meng Q, Zhang Y, Meng D, Liu X, Zhang Z, Gao P, Lin A, Hou
L (2020) Removal of sulfadiazine from aqueous solution by in-
situ activated biochar derived from cotton shell. Environ Res
191:110104. https:// doi. org/ 10. 1016/j. envres. 2020. 110104
Nayak TR, Andersen H, Makam VS, Khaw C, Bae S, Xu X, Ee PR,
Ahn J, Hong BH, Pastorin, G.özyilmaz, B. (2011) Graphene for
Controlled and Accelerated Osteogenic Differentiation of Human
Mesenchymal Stem Cells. ACS Nano 5:4670–4678. https:// doi.
org/ 10. 1021/ nn200 500h
Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos
SV, Grigorieva IV, Firsov AA (2004) Electric field effect in atomi-
cally thin carbon films. Science 306:666–669. https:// doi. org/ 10.
1126/ scien ce. 11028 96
Peiris C, Gunatilake SR, Mlsna TE, Mohan D, Vithanage M (2017)
Biochar based removal of antibiotic sulfonamides and tetracy-
clines in aquatic environments: A critical review. Bioresource
Technol 246:150–159. https:// doi. org/ 10. 1016/j. biort ech. 2017.
07. 150
Perreault F, de Faria AF, Nejati S, Elimelech M (2015) Antimicrobial
Properties of Graphene Oxide Nanosheets: Why Size Matters.
ACS Nano 9:7226–7236. https:// doi. org/ 10. 1021/ acsna no. 5b020
67
Priyadarsini S, Mohanty S, Mukherjee S, Basu S, Mishra M (2018)
Graphene and graphene oxide as nanomaterials for medicine and
biology application. J Nano Chem 8:123–137. https:// doi. org/ 10.
1007/ s40097- 018- 0265-6
Sarmah AK, Meyer MT, Boxall ABA (2006) A global perspective on
the use, sales, exposure pathways, occurrence, fate and effects of
veterinary antibiotics (VAs) in the environment. Chemosphere
65:725–759. https:// doi. org/ 10. 1016/j. chemo sphere. 2006. 03. 026
Serna-Carrizales JC, Collins-Martínez VH, Flórez E, Gomez-Duran
CFA, Palestino G, Ocampo-Pérez R (2021) Adsorption of sul-
famethoxazole, sulfadiazine and sulfametazine in single and ter-
nary systems on activated carbon. Experimental and DFT compu-
tations. J Mol Liq 324:114740. https:// doi. org/ 10. 1016/j. molliq.
2020. 114740
Shan D, Deng S, Li J, Wang H, He C, Cagnetta G, Wang B, Wang Y,
Huang J, Yu G (2017) Preparation of porous graphene oxide by
chemically intercalating a rigid molecule for enhanced removal
of typical pharmaceuticals. Carbon 119:101–109. https:// doi. org/
10. 1016/j. carbon. 2017. 04. 021
Shan D, Deng S, He C, Li J, Wang H, Jiang C, Yu G, Wiesner MR
(2018) Intercalation of rigid molecules between carbon nanotubes
for adsorption enhancement of typical pharmaceuticals. Chem
Eng J 332:102–108. https:// doi. org/ 10. 1016/j. cej. 2017. 09. 054
Singh A, Sharma RP, Aree T, Venugopalan P (2009) Second sphere
coordination in fluoroanion binding: Synthesis, spectroscopic and
X-ray structural study of [Co(phen)2CO3](Pfbz)·6H2O. J Fluorine
Chem 130:650–655. https:// doi. org/ 10. 1016/j. jfluc hem. 2009. 04.
011
Stankovich S, Dikin DA, Piner RD, Kohlhaas KA, Kleinhammes A,
Jia Y, Wu Y, Nguyen ST, Ruoff RS (2007) Synthesis of graphene-
based nanosheets via chemical reduction of exfoliated graphite
oxide. Carbon 45:1558–1565. https:// doi. org/ 10. 1016/j. carbon.
2007. 02. 034
72554 Environmental Science and Pollution Research (2022) 29:72545–72555
1 3
Sun X, Liu Z, Welsher K, Robinson JT, Goodwin A, Zaric S, Dai
H (2008) Nano-graphene oxide for cellular imaging and
drug delivery. Nano Res 1:203–212. https:// doi. org/ 10. 1007/
s12274- 008- 8021-8
Szymańska U, Wiergowski M, Sołtyszewski I, Kuzemko J, Wiergowska
G, Woźniak MK (2019) Presence of antibiotics in the aquatic
environment in Europe and their analytical monitoring: Recent
trends and perspectives. Microchem J 147:729–740. https:// doi.
org/ 10. 1016/j. microc. 2019. 04. 003
Tomasi J, Bonaccorsi R, Cammi R, de Valle FJO (1991) Theoretical
chemistry in solution. Some results and perspectives of the con-
tinuum methods and in particular of the polarizable continuum
model. J Mol Struct (theochem) 234:401–424
Travlou NA, Kyzas GZ, Lazaridis NK, Deliyanni EA (2013) Graphite
oxide/chitosan composite for reactive dye removal. Chem Eng J
217:256–265. https:// doi. org/ 10. 1016/j. cej. 2012. 12. 008
Verma M, Tyagi I, Kumar V, Goel S, Vaya D, Kim H (2021) Fabrica-
tion of GO-MnO2 nanocomposite using hydrothermal process for
cationic and anionic dyes adsorption: Kinetics, isotherm, and reus-
ability. J Environ Chem Eng 9:106045. https:// doi. org/ 10. 1016/j.
jece. 2021. 106045
Wang W, Zhou S, Li R, Peng Y, Sun C, Vakili M, Yu G, Deng S
(2021a) Preparation of magnetic powdered carbon/nano-Fe3O4
composite for efficient adsorption and degradation of trichloro-
propyl phosphate from water. J Hazard Mater 416:125765. https://
doi. org/ 10. 1016/j. jhazm at. 2021. 125765
Wang W, Zhou Z, Shao H, Zhou S, Yu G, Deng S (2021b) Cationic
covalent organic framework for efficient removal of PFOA sub-
stitutes from aqueous solution. Chem Eng J 412:127509. https://
doi. org/ 10. 1016/j. cej. 2020. 127509
Wang Z, Ma R, Meng Q, Yang Y, Ma X, Ruan X, Yuan Y, Zhu G (2021c)
Constructing Uranyl-Specific Nanofluidic Channels for Unipolar
Ionic Transport to Realize Ultrafast Uranium Extraction. J Am
Chem Soc 143:14523–14529. https:// doi. org/ 10. 1021/ jacs. 1c025 92
Wei X, Zhang Z, Qin L, Dai J (2019) Template-free preparation of yeast-
derived three-dimensional hierarchical porous carbon for highly
efficient sulfamethazine adsorption from water. J Taiwan Inst Chem
E 95:532–540. https:// doi. org/ 10. 1016/j. jtice. 2018. 09. 009
Wu L, Du C, He J, Yang Z, Li H (2020) Effective adsorption of
diclofenac sodium from neutral aqueous solution by low-cost lig-
nite activated cokes. J Hazard Mater 384:121284. https:// doi. org/
10. 1016/j. jhazm at. 2019. 121284
Wu X, Mu F, Lin Z (2021) Three-dimensional printing of graphene-
based materials and the application in energy storage. Mater Today
Adv 11:100157. https:// doi. org/ 10. 1016/j. mtadv. 2021. 100157
Yagub MT, Sen TK, Afroze S, Ang HM (2014) Dye and its removal
from aqueous solution by adsorption: A review. Adv Colloid Inter-
fac 209:172–184. https:// doi. org/ 10. 1016/j. cis. 2014. 04. 002
Zhang C, Lai C, Zeng G, Huang D, Yang C, Wang Y, Zhou Y, Cheng
M (2016) Efficacy of carbonaceous nanocomposites for sorbing
ionizable antibiotic sulfamethazine from aqueous solution. Water
Res 95:103–112. https:// doi. org/ 10. 1016/j. watres. 2016. 03. 014
Zhuang X, Li X, Yang Y, Wang N, Shang Y, Zhou Z, Li J, Wang H
(2020) Enhanced Sulfamerazine Removal via Adsorption-Photo-
catalysis Using Bi2O3-T iO2/PAC Ternary Nanoparticles. Water
12:2273. https:// doi. org/ 10. 3390/ w1208 2273
Publisher's note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
72555Environmental Science and Pollution Research (2022) 29:72545–72555
Article
Due to increasing antibiotic pollution in the water environment, green and efficient adsorbents are urgently needed to solve this problem. Here we prepare magnetic bamboo-based activated carbon (MDBAC) through delignification and carbonization using ZnCl2 as activator, resulting in production of an activated carbon with large specific surface area (1388.83 m2 g-1). The influencing factors, such as solution pH, initial sulfadiazine (SD) concentration, temperature, and contact time, were assessed in batch adsorption experiments. The Langmuir isotherm model demonstrated that MDBAC adsorption capacity on SD was 645.08 mg g-1 at its maximum, being higher than majority of previously reported adsorbents. In SD adsorption, the kinetic adsorption process closely followed the pseudo-second kinetic model, and the thermodynamic adsorption process was discovered to be exothermic and spontaneous in nature. The MDBAC exhibited excellent physicochemical stability, facile magnetic recovery and acceptable recyclability properties. Moreover, the synergistic interactions between MDBAC and SD mainly involved electrostatic forces, hydrogen bonding, π-π stacking, and chelation. Within the benefits of low cost, ease of production and excellent adsorption performance, the MDBAC biosorbent shows promising utilization in removing antibiotic contaminants from wastewater.
Article
A novel magnetic porous graphene oxide (GO-DFe) was prepared by growing Fe3O4 via an impregnation method in the nanoscale channel of porous GO (GO-DFP), which was fabricated by chemically intercalating 4,4'-Diaminooctafluorodiphenyl (DFP) into the GO sheets. In this study, the concentrations of Fe²⁺/Fe³⁺ was optimized to obtain the GO-DFe5 adsorbent with good adsorption and separation performance. The adsorption performance of magnetic GO-DFe5 towards sulfadiazine (SD) and carbamazepine (CBZ) was evaluated, and adsorption kinetic models (the first-order, the second-order, Elovich and intra-particle models) and isotherm models (Langmuir and Freundlich models) were applied to study the adsorption rate and capacity. The effects of solution pH, ions (Cl⁻ and NO3⁻) and co-existed matters in effluent of municipal wastewater on the adsorption behavior of SD and CBZ on GO-DFe5 were further clarified. Additionally, the cyclic reuse performance of GO-DFe5 for SD and CBZ were also evaluated, and the potential adsorption mechanism was proposed by theoretical calculation. The as-prepared GO-DFe5 adsorbent exhibited high adsorption capacities for SD (380.67 μmol/g) and CBZ (350.70 μmol/g). The adsorption amounts of GO-DFe5 for two pharmaceuticals about were 3.0 and 2.1 times higher than that of GO according to the second-order model. π-π interaction, hydrogen bonding and electrostatic interaction were involved in the adsorption of SD and CBZ on GO-DFe5. GO-DFe5 in water can be easily recovered by a magnet and exhibited a well reuse performance for SD and CBZ within 6 cycles.
Article
Full-text available
The remarkable adsorption capacity of graphene-derived materials has prompted their examination in composite materials suitable for deployment in treatment of contaminated waters. In this study, crosslinked calcium alginate–graphene oxide beads were prepared and activated by exposure to pH 4 by using 0.1M HCl. The activated beads were investigated as novel adsorbents for the removal of organic pollutants (methylene blue dye and the pharmaceuticals famotidine and diclofenac) with a range of physicochemical properties. The effects of initial pollutant concentration, temperature, pH, and adsorbent dose were investigated, and kinetic models were examined for fit to the data. The maximum adsorption capacities qmax obtained were 1334, 35.50 and 36.35 mg g−1 for the uptake of methylene blue, famotidine and diclofenac, respectively. The equilibrium adsorption had an alignment with Langmuir isotherms, while the kinetics were most accurately modelled using pseudo- first-order and second order models according to the regression analysis. Thermodynamic parameters such as ΔG°, ΔH° and ΔS° were calculated and the adsorption process was determined to be exothermic and spontaneous.
Article
Full-text available
Preparation of biochar from kaolinite and coconut husk (KCB) and further activated with HCl (KCB-A) and KOH (KCB-B) via a microwave technique for the remediation of ciprofloxacin (CIP) and tetracycline (TET) from water was carried out. Characterization using scanning electron microscopy, energy dispersive X-ray, Fourier transform infrared spectroscopy and X-ray diffraction showed the successful synthesis of functionalized biochars. Batch adsorption experiments demonstrated the potential of the adsorbents for fast and efficient removal of CIP and TET from solution. The adsorption capacities were found to be 71, 140 and 229 mg g-1 for CIP and 118, 117 and 232 mg g-1 for TET removal on KCB, KCB-A and KCB-B, respectively. For KCB, KCB-B and KCB-B, CIP adsorption best followed the pseudo second order kinetic model (PSOM), pseudo first order kinetic model (PFOM) and intraparticle diffusion (IDP) respectively. TET adsorption followed PSOM for KCB, IPD for KCB-B and PFOM for KCB-A. CIP adsorption on KCB, KCB-A and KCB-B best fit the Temkin, Langmuir and Brouers-Sotolongo isotherms, respectively, and TET adsorption on KCB best fit Brouers-Sotolongo while KCB-A and KCB-B best fit Langmuir-Freundlich. Adsorption of both contaminants was thermodynamically feasible showing that these materials are excellent adsorbents for the treatment of pharmaceuticals in water. This journal is
Article
Graphene-based materials have been extensively investigated in the energy-related applications owing to their unique properties, such as high conductivity and mechanical flexibility. Three-dimensional (3D) graphene architectures could further strengthen their performance and facilitate the applications in energy storage. To fabricate 3D graphene architectures, the rapidly developed 3D printing technology presents a lot of advantages and has received much research attention. In this paper, we reviewed the recent research advances of 3D printing of graphene-based materials and discussed the applications in energy storage areas. The main 3D printing techniques applied in constructing graphene-based structures were summarized, and the characteristics of each method were briefly introduced. The current progresses of energy storage applications, focusing on supercapacitors and energy storage batteries, were reviewed in detail. Moreover, the future research challenges and prospects were provided in the last part, aiming at stimulating more significant research and industrial applications in this subject.
Article
The presence of anionic dye and cationic dyes in wastewater has highlighted a great necessity to develop new and effective approaches for their simultaneous removal. Herein, graphene oxide-manganese dioxide (GO-MnO2) nanocomposite was synthesized using the hydrothermal method to reduce pollution load of wastewater. Synthesized material was utilized as adsorbent for the removal of cationic methylene blue (MB) and anionic methyl orange (MO) dyes from aqueous solution that act as model organic pollutants. The morphology, chemical structure, thermal stability, and other properties of the synthesized adsorbent were characterized using FE-SEM, PXRD, Raman spectroscopy, FT-IR, EDS, TGA, and BET surface area techniques. The kinetics results showed the removal efficiency 50.48% and 85.35% in starting 5 min of MO and MB dyes, respectively, and fitted well with pseudo-second-order (PSO) kinetics model. The isotherms adsorption results fitted well with Langmuir isotherm model, confirming the monolayer adsorption and give maximum adsorption capacities 149.253 and 178.253 mg/g for MO and MB dyes, respectively. GO-MnO2 adsorbent shows good reusability and give ˃90% removal efficiency after seven continuous cycles. In last, the adsorption performance for simultaneous adsorption of both dyes gives 100% removal efficiency. All these results give a direct visual impression of the fast kinetics efficiency and high adsorption capacity for real wastewater treatment application.
Article
Trichloropropyl phosphate (TCPP) as a widely used typical chlorinated organophosphate flame retardant has received significant attention because of its widespread presence in water and negative effects on human health. In this study, a ball-milling method was used to prepare a magnetic powdered carbon adsorbent (PC/nano-Fe3O4 composite) for TCPP removal via adsorption and catalytic degradation. The effect of Fe3O4 content on TCPP adsorption and degradation performance by PC/nano-Fe3O4 composite was investigated. The PC/nano-Fe3O4 composite prepared by high Fe3O4 content (25%) was not favorale for TCPP adsorption and degradation. However, the PC/Fe3O4 containing low Fe3O4 content (10%) had insufficient magnetic separation ability from water. The synthesized PC/nano-Fe3O4 composite with a Fe3O4/PC mass ratio of 1/5 exhibited a maximum adsorption capacity of 2682.1 μg/g as well as a complete TCPP degradation within 3 h in a Fenton-like system. Moreover, the possible break sites of TCPP and its degradation pathway were proposed based on theoretical calculation and experimental analysis. Regeneration studies showed that PC/nano-Fe3O4 composite had high reusability and adsorption capacity in six cycles, while its catalytic performance declined in the multiple reuse cycles. This strategy could be extended to prepare other magnetic powdered adsorbents for organic pollutant adsorption and degradation.
Article
The search for a cost-effective and mild activation method of preparing biochar with excellent adsorption capacity is advantageous to broaden its potential industrial application. Accordingly, tea waste was selected as a carbon source, and KHCO3 and CH3COOK were selected as activators in the two-step sintering process because of their mildness and non-corrosive features. Three types of tea waste biochar was prepared, characterised and utilised for the adsorption of sulfadiazine (SDZ) from aqueous solutions. Benefiting from the activation, the specific surface areas of the tea waste biochar with the highest adsorption capacity described as KHCO3-TB-1:2 and CH3COOK-TB-1:2, were 717.636 and 648.415 m²/g, respectively, whereas the specific surface area of tea waste biochar that was not activated in the two-step sintering process was named TB was only 4.833 m²/g. Moreover, the tea waste biochar that was activated by KHCO3 and CH3COOK developed a typical micro-/meso-/macro-hierarchical pore structure. Compared to TB without an adsorption capacity, the maximum adsorption capacities of KHCO3-TB-1:2 and CH3COOK-TB-1:2 for SDZ were 77.52 and 58.14 mg/g, respectively, at an initial concentration of 50 mg/L and pH = 10.97. The data of the batch-adsorption experiments fitted well with the pseudo-second-order kinetics and Langmuir model, thus suggesting that the adsorption process was dominated by chemical and monolayer adsorptions. The π–π electron donor–acceptor interaction, hydrogen bonding, and electrostatic interaction were the possible adsorption mechanisms. The activation of KHCO3 and CH3COOK is a very promising method to limit equipment corrosion and prepare biochar for the removal of SDZ from aqueous solutions.
Article
In this work the single and ternary removal of sulfonamides (sulfamethoxazole, sulfadiazine and sulfametazine) from water was investigated using granular activated carbon. The single adsorption mechanism was elucidated by obtaining the adsorption isotherms supported by computational calculations. The ternary adsorption was analyzed by using an experimental Box-Behnken type response surface design. The results showed that the total adsorption capacity of activated carbon duplicates in ternary systems compared to single systems. Besides, in both cases the activated carbon showed a greater affinity for removing sulfamethoxazole followed by sulfadiazine and sulfametazine, correspondingly. It was shown that hydrogen bonding interactions presented the highest adsorption energies followed by π-π interactions. From the design of experiments three statistically reliable mathematical models were proposed to estimate the adsorption capacity for each sulfonamide as a function of the solution pH, temperature and initial concentration. Finally, it was shown that sulfonamide ternary adsorption is an endothermic process and that the adsorption rate decreases as a result of partial blockage of the pores due to simultaneous adsorption.
Article
Perfluorooctanoic acid (PFOA) substitutes such as hexafluoropropylene oxide dimer acid (GenX) and hexafluoropropylene oxide trimer acid (HFPO-TA) are toxic as environmental contaminants, and efficient adsorbents are required to remove them from waters. In this study, the novel cationic covalent organic framework (COF) was developed via an imine condensation reaction for efficient removal of GenX and HFPO-TA from aqueous solution. The prepared COFs exhibited a 2D ordering hexagonal structure with AA stacking and two classes of pores (mesopore and micropore). The introduction of quaternary ammonium on COF exhibited little effect on its morphology and crystal structure. The cationic COF with quaternary ammonium showed high adsorption capacity for GenX (2.06 mmol/g) and HFPO-TA (2.16 mmol/g), more efficient than conventional activated carbons and resins. The effects of pH and natural organic matter on the removal efficiency of GenX and HFPO-TA on COF were also investigated. Interestingly, the previously adsorbed GenX on the cationic COF was replaced by the longer-chain HFPO-TA, and the underlying competitive mechanism was further clarified by performing a density functional theory calculation. This study highlights the exceptional potential of COFs through topology structural design for the efficient removal of per- and polyfluoroalkyl substances from aqueous solution.
Article
Two copper coordination compounds, [Cu2(OOCC6F5)6](Et3NH)2 (1) and {(Et3NH)[Cu2(OOCC6F5)5]}n (2, [Cu2(OOCC6F5)4] as paddle-wheel secondary building unit, OOCC6F5 = pentafluorobenzoate ligand), have been obtained and characterized. The coordination modes of axial ligands were different in compound 1 and 2, resulting to different crystal structures and magnetic properties. The zero-dimensional molecule compound 1 indicated two magnetic orderings. At low temperature (T < 70 K), a weak ferromagnetic ordering was observed, and an antiferromagnetic behavior took place beyond this temperature (T > 70 K). However, one-dimensional chained compound 2 exhibited intramolecular antiferromagnetic exchange coupling in the range 2–300 K. Moreover, the correlation of magnetism and structure was discussed.