Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

nanomaterials

Article
Enhanced Adsorption of Zn(II) onto Graphene Oxides
Investigated Using Batch and Modeling Techniques
Min Pan 1, * , Guangxue Wu 2 , Chang Liu 3 , Xinxin Lin 1 and Xiaoming Huang 1, *
1 Fujian Engineering and Research Center of Rural Sewage Treatment and Water Safety, School of
Environmental Science and Engineering, Xiamen University of Technology, Xiamen 361024, China;
lin.xinxin@outlook.com
2 Graduate School at Shenzhen, Tsinghua University, Shenzhen 518055, China;
wu.guangxue@sz.tsinghua.edu.cn
3 College of Environmental Science and Engineering, Anhui Normal University, Wuhu 241002, China;

D
lc2014@ahnu.edu.cn
* Correspondence: panmin@xmut.edu.cn (M.P.); huangxm@xmut.edu.cn (X.H.); Tel.: +86-5926291138 (M.P.)

Received: 17 September 2018; Accepted: 3 October 2018; Published: 9 October 2018 

TE
Abstract: Graphene oxide (GO) was synthesized and employed as an adsorbent for Zn(II) removal
from an aqueous solution. The adsorption isotherms showed that Zn(II) adsorption can be better
described using the Freundlich model than the Langmuir model. The maximum adsorption capacity
of Zn(II) on GO determined using the Langmuir model at pH 7.0 and 293 K was 208.33 mg/g.
C
The calculation of thermodynamic parameters revealed that the process of Zn(II) adsorption
on GO was chemisorptions, endothermic, and spontaneous. Kinetic studies indicated that the
pseudo-second-order kinetic model showed a better simulation of Zn(II) adsorption than the
pseudo-first-order kinetic model. On the basis of surface complexation modeling, the double layer
A
model provided a satisfactory prediction of Zn(II) by inner-sphere surface complexes (for example,
SOZn+ and SOZnOH species), indicating that the interaction mechanism between Zn(II) and GO
was mainly inner-sphere complexation. In terms of reusability, GO could maintain 92.23% of its
TR

initial capability after six cycles. These findings indicated that GO was a promising candidate for the
immobilization and preconcentration of Zn(II) from aqueous solutions.

Keywords: graphene oxides; surface complexation modeling; adsorption; Zn(II)


RE

1. Introduction
With the rapid development of agricultural and industrial activities, massive wastewater caused
by heavy metal ions is generated. Increasing attention has been paid to the pollution of heavy
metals due to their poisonous effect on organisms even at trace levels, such as cancer, anemia,
and intellectual disability [1,2]. Zn can be toxic through the food chain via bio-accumulation [3].
Hence, it is mandatory to decontaminate heavy metal ions to permissible limits (5 mg/L in drinking
water according to The World Health Organization). As one of the typical heavy metals, Zn is often
found in effluent from mining, smelting, machinery manufacturing, instrumentation, etc. Currently,
Zn(II) removal from liquid has been investigated through chemical precipitation, ion exchange,
membrane separation, adsorption, and coagulation [3–6]. Among these methods, the adsorption
process is extensively considered as an effective and efficient approach due to its low cost, easy
handing, the availability of various adsorbents, and its environment-friendly properties [1]. To date,
there has been plenty of literature regarding Zn removal on various adsorbents such as active
carbon [7], carbon nano-materials [8], magnetic nanoparticle [9], nano-scale zero-valent iron [10],
mineral [4,11], etc. Ggasemi et al. (2015) demonstrated that the maximum adsorption capacity of Zn(II)

Nanomaterials 2018, 8, 806; doi:10.3390/nano8100806 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 806 2 of 13

was only 24.21 mg/g using dioctylphetalate triethylenetetraamine magnetic nanoparticles [9]. Therefore,
the development of highly efficient adsorbents for Zn(II) removal was still required, especially for actual
application in environmental clean-up operations.
Graphene oxide (GO), as a new two-dimensional carbon-based material, has been widely explored
due to its large specific surface area (about 2600 m2 /g), high mechanical strength (>1060 GPa),
and plentiful oxygenous functionalities (e.g., carboxyl, epoxy, and hydroxyl groups) [12–15]. Thus, GO
is deemed to be a promising sorbent for the removal of heavy metal ions [3,16–19], humic acid [20],
radionuclides [14,21–24], organic dyes [25,26], ammonia [27], etc. Wang (2013) demonstrated that
the maximum adsorption capacity of GO was 246 mg/g for Zn(II) [3]. Zhang (2016) also found that
the removal efficiency of Zn was 98.07% through GO-based framework composite membranes [16].
However, most research has focused on the adsorption mechanisms by investigating adsorption
isotherms and kinetics coupled with the effects of experiment conditions (such as pH, temperature,
time, etc.) on adsorption behaviors. Modeling evidence of Zn(II) adsorption onto GO is limited.

D
Surface complexation modeling, as a powerful tool, has been widely applied to fit the adsorption
behaviors. In our previous study, the highly efficient adsorption of Pb(II) onto GO was well-simulated
by employing surface complexation modeling with SOPb+ species and (SO)2 Pb2 (OH)3 − species [28].

TE
Ding et al. (2014), using surface complexation modeling, determined that the sorption performance
of GO was attributed to the type and amount of the oxygen-bearing functionalities [29]. However,
few studies regarding the prediction of Zn(II) onto GO using surface complexation modeling are
currently available.
This study aims to synthesize GO and explore the interaction mechanisms of Zn(II) adsorption
on GO subject to analytical techniques (such as X-ray diffraction (XRD), scanning electron
C
microscopy (SEM), transmission electron microscopy (TEM), Fourier transformed infrared (FTIR),
X-ray photoelectron spectra (XPS), Raman, and Zeta potential), batch experiments (pH effect, ionic
strength effect, kinetic study, and temperature effect), and surface complexation modeling. The key
A
objective of this work is to reveal the fate and transformation of Zn(II) on GO nano-material,
which can be considered as a promising absorbent used for the removal of heavy metals in
environmental remediation.
TR

2. Materials and Methods

2.1. Materials
Graphite (<20 µm, 99.99% purity) was purchased from Qingdao Tianhe Graphite Co. Ltd.
(Qingdao, China). Concentrated H2 SO4 (98%), KMnO4 , H2 O2 (37 wt%) and NaNO3 were purchased
RE

from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China). The Zn(II) stock solution (50 mg/L)
was prepared by dissolving Zn(NO3 )2 in deionized water under ambient conditions and then diluted
with deionized water to obtain all the working solutions of Zn(II).

2.2. Synthesis and Characterization of GO


The GO was synthesized using the modified Hummers method. Generally, flake graphite was
preoxidized by slowly adding 2.0 g flake graphite and 1.5 g NaNO3 into concentrated H2 SO4 in
vigorous stirring and ice-water conditions. Sequentially, the preoxidized graphite was further oxidized
by adding 9.0 g of KMnO4 , and then the suspensions were reacted under room temperature for 5 days.
Then, 120 mL 5% of H2 SO4 and H2 O2 solution were added to remove the abundant MnO4 − ions.
Details on the synthesis of GO are described in our previous research [28].
The morphology of GO was evaluated using a scanning electron microscope (SEM JEOL-6060SEM,
JEOL Ltd., Tokyo, Japan) and a transmission electron microscope (TEM JEM-2010 JEOL Ltd., Tokyo,
Japan). The FTIR spectra of GO were identified using Bruker Vector 22 (Bruker, Madison, WI, USA).
The XPS spectrum was conducted using a thermo ESCALAB 250 electron spectrometer (Thermo Fisher
Scientific, Waltham, MA, USA). GO was examined using the XRD analysis with a Rapid-II powder
Nanomaterials 2018, 8, 806 3 of 13

diffractometer (D/RAPID II, Rigaku, Tokyo, Japan) with CuKa radiation. The XRD pattern was taken
in the range of 5–70◦ at a scan speed of 4◦ /min. Raman spectroscopy was carried out on the LabRam
HR Raman spectrometer (Horiba, Paris, France) with an Ar+ laser. The zeta potential of GO was
confirmed using a zeta meter (Nano ZS Zetasizer, Malvern, UK).

2.3. Batch Adsorption Experiments


To investigate the adsorption characteristic of Zn(II) on GO, a batch technique in polyethylene
centrifugal tubes under room temperature conditions was conducted. Specifically, the GO solution
was mixed with a freshly prepared Zn(II) solution in the presence of 0.01 M NaClO4 solution. The pH
values in the aqueous solution were regulated by adding 0.01–1.0 mol/L HClO4 or NaOH solution
drop by drop. After being centrifuged at 6000 rpm for 30 min, the supernatant was used to determine
the Zn(II) concentration using atomic absorption spectroscopy (AA-6300C, Shimadzu, Japan). All the

D
experimental data obtained in this study were the mean values of duplicate or triplicate measurements
and the relative error was <5%. The equilibrium adsorptive capacity was calculated using the
following equation:
(C0 − Ct )v
qt = (1)

TE
m
where qt is the adsorptive capacity at time t, mg/g; C0 and Ct (mg/L) are the Zn(II) concentration in
the aqueous solution at the beginning and time t, respectively; v is the volume of the solution, L; and m
is the mass of the adsorbent, g.

2.4. Surface Complexation Modeling


C
The adsorption of Zn(II) on GO is simulated using the diffuse layer model (DLM) of surface
complexation modeling with the help of the FITEQL v 4.0 code [28,30]. The values of the protonation
and deprotonation constants (logK+ and logK− ) are described using Equations (2) and (3), respectively:
A
i
SOH + H+ = SOH2 + logK1 + = log([SOH 2 + ]/[SOH][H + ) (2)
TR

SOH = SO− + H+ log K2 − = log([SO− ][H+ ]/[SOH]) (3)

where SOH is represented as the amphoteric surface groups on GO. The values of logK+ and logK−
were cited from the previous studies. In the existence of 0.01 mol/L NaClO4 solution, the surface
complexation reactions are expressed as Equations (4) and (5):

SOH + Zn2+ = SOZn+ + H+ log K3 = log([SOZn+ ][H+ ]/[SOH][Zn2+ ])


RE

(4)

2
SOH + Zn2 + + H2 O = SOZnOH + 2H+ log K4 = log([SOZnOH][H+ ] /[SOH][Zn2+ ]) (5)

The equilibrium constants (logK values) can be received by optimizing the simulated results with
the adsorption data.

3. Results and Discussion

3.1. Characterization
Figure 1A,B show the observation from the SEM and TEM techniques of GO, respectively.
The wrinkled, aggregated, and thin nanosheets were clearly observed, which corresponded with
the previous study [29]. According to the TEM images of GO, the mean thickness of GO was near
2.5 nm (Figure 1B). As shown in the FTIR spectrum (Figure 1C), the peaks at ~1045, 1450, 1630, 1720,
and 3440 cm−1 contributed to the stretching vibration of the C–O–C, C=O, C=C, COOH and COH
groups, respectively [31]. XPS, as a powerful tool, can characterize the chemical constituents and
electronic structures of GO. Figure 1D shows the deconvolution of high-resolution O 1s XPS spectra of
Nanomaterials 2018, 8, 806 4 of 13

GO. The binding energy of O 1s XPS at 531.25, 532.48, 533.02, and 533.45 eV can be indexed to C=O,
C–O–C, C–OH and adsorbed H2 O (chemisorbed/intercalated adsorbed water molecules), respectively.
Nanomaterials
The findings from2018, the
8, x FOR
XPSPEER REVIEW illustrated that the GO sheets were decorated with 4plentiful
analysis of 13

oxygenic functionalities (such as hydroxyl, epoxy, carboxyl, and carbonyl groups) [32]. Figure 1E
findings from the XPS analysis illustrated that the GO sheets were decorated with plentiful oxygenic
illustrates the XRD pattern of GO. Ragubanshi et al. (2017) found that the diffraction peak of pure
functionalities (such as hydroxyl, epoxy, carboxyl, and carbonyl groups) [32]. Figure 1E illustrates
graphite was pattern
the XRD appropriately 2θ = 26.58et◦ al.
of GO. Ragubanshi [33].
(2017)One canthat
found seethe
that only one
diffraction weak
peak diffraction
of pure peak at
graphite was
2θ = 10.66 ◦
appropriately 2θ = 26.58° [33]. One can see that only one weak diffraction peak at 2θ = 10.66° was(002)
was shown, which corresponded to the GO nanosheets and was attributed to the
plane.shown,
The absence of a peak of pure
which corresponded graphite
to the suggested
GO nanosheets andthat
was an excellent
attributed separation
to the andThe
(002) plane. high quality of
absence
GO was of produced.
a peak of pure graphite suggested
The interplanar spacingthatof GOan was
excellent separation
estimated to beand
0.945high
nm,quality
which of GOconsistent
was was
produced.work
with previous The interplanar
[34]. Ramanspacing of GO was estimated
spectroscopy to be 0.945
of GO indicated nm,
the which was
D-band cm−1 with
consistent
at 1350 and the
previous work − [34].
1 Raman spectroscopy of GO indicated the D-band at 1350 cm
G-band at 1590 cm (Figure 1F). It is well-known that the D-band reflects disorders and local defects
−1 and the G-band at

while 1590 cm−1 (Figure 1F). It is well-known that the D-band


the G-band is related to the vibrations of SP2 ofreflects carbondisorders
atoms in and local defects while the
a graphitic 2D hexagonal
G-band is related to the vibrations of SP2 of carbon atoms in a graphitic 2D hexagonal lattice [33,35].
lattice [33,35]. The ID /IG ratio was calculated as 0.91, implying less disorder and defect concentration

D
The ID/IG ratio was calculated as 0.91, implying less disorder and defect concentration and more
and more perfect aromatic structure in GO.
perfect aromatic structure in GO.

TE
C
A
TR
RE

FigureFigure 1. Characterization
1. Characterization ofofGO
GO((A)
((A) SEM
SEM image;
image;(B) TEM
(B) image;
TEM (C) FTIR
image; spectrum;
(C) FTIR (D) XPS;
spectrum; (D)(E)XPS;
XRD;
(E) XRD; (F)(F)Raman).
Raman).
Nanomaterials 2018, 8, 806 5 of 13

3.2. Effect of pH and Ionic Strength


Figure 2A shows
Nanomaterials 2018, 8, xthe
FOReffect of pH on Zn(II) adsorption onto GO by altering the pH value
PEER REVIEW 5 offrom
13 2.0
to 10.0. It clearly demonstrates that Zn(II) adsorption rapidly increased with an increase in the pH from
3.2. Effect
2.0 to 7.0, of pH and
preserved at aIonic Strength sorption at pH 7.0–8.0, and slightly decreased from pH 8.0 to 10.0.
high-degree
The adsorptionFigurewas determined
2A shows to be
the effect a pH-dependent
of pH process.
on Zn(II) adsorption ontoInGO Figure 2B, thethe
by altering main species
pH value fromof Zn in
liquid2.0
were Zn2+
to 10.0. species
It clearly at pH < 8.5,
demonstrates Zn(OH)
that (s) species
Zn(II) adsorption at
rapidly pH of 8.5–9.5,
increased with Zn(OH)
an increase − in
species
the pH at pH
2 3
from 2.0 to 7.0, 2 −
preserved at a high-degree sorption at pH 7.0–8.0, and
11.0–11.5, and Zn(OH)4 at pH > 12.5. Based on the zeta potential in Figure 2C, GO was negatively slightly decreased from pH 8.0
chargedto 10.0.
at pH The
> adsorption
2.0. Thus, was determined adsorption
the increasing to be a pH-dependent
of Zn onprocess.
GO at pH In Figure
2.0–7.0 2B,could
the main species
be attributed to
of Zn in liquid were Zn2+ species at pH < 8.5, Zn(OH) 2 (s) species at pH of 8.5–9.5, Zn(OH)
2+ and negatively charged GO. The significant Zn(II) 3 − species
the electrostatic attraction between positive Zn
at pH 11.0–11.5, and Zn(OH)42− at pH > 12.5. Based on the zeta potential in Figure 2C, GO was
removal percentage at pH 7.0–8.0 could be attributed to simultaneous precipitation Zn(OH)2 (s),
negatively charged at pH > 2.0. Thus, the increasing adsorption of Zn on GO at pH 2.0–7.0 could be
which
was inattributed
accordance with the previous study [3]. At pH > 8.0, the electrostatic
to the electrostatic attraction between positive Zn2+ and negatively charged GO. The repulsion between
negatively charged GO and Zn(OH) − 2− species resulted in a slight reduction in
significant Zn(II) removal 3 species
percentage and Zn(OH)
at pH 7.0–8.0 could be4 attributed to simultaneous precipitation

D
Zn(II) Zn(OH)
adsorption onto GO.
2(s), which was in accordance with the previous study [3]. At pH > 8.0, the electrostatic
repulsion
Figure 2A between
indicates negatively charged
the influence GO and
ionic Zn(OH)
strength species
on3−the and Zn(OH)
adsorption 42− species
of Zn(II) resulted
onto GO with in a 0.001,
slight reduction in Zn(II) adsorption onto GO.
0.01 and 0.1 mol/L NaClO4 presented. When the pH value was lower than 4, the removal efficiencies
Figure 2A indicates the influence ionic strength on the adsorption of Zn(II) onto GO with 0.001,
of Zn(II) were slightly increased with the NaClO4 concentrations increased. When the pH value was

TE
0.01 and 0.1 mol/L NaClO4 presented. When the pH value was lower than 4, the removal efficiencies
higher than 4, the changes in the removal efficiencies of Zn(II) could be negligible with the NaClO4
of Zn(II) were slightly increased with the NaClO4 concentrations increased. When the pH value was
concentrations
higher thanincreasing.
4, the changes Wangin theetremoval
al. (2013) stated that
efficiencies a foreign
of Zn(II) could ion may affect
be negligible withthetheaffinity
NaClO4 to the
surface of GO [3]. Weng et al. (2004) found that a foreign ion may influence
concentrations increasing. Wang et al. (2013) stated that a foreign ion may affect the affinity to the the electrical double
layer structure
surface of GOand[3]. thenWengimpact
et al. the
(2004)binding of the
found that adsorption
a foreign ion may species
influence[36].
theHowever, Tan etlayer
electrical double al. (2017)
structure
declared that the and then impact
sorption 2+ on
of Cuthe binding of the adsorption
GO changed slightly byspecies Cl− (fromTan
[36]. However,
increasing 0.01etMal.to(2017)
0.1 M) [17].
C
declared
Liu et al. (2013)that the sorption that
demonstrated of Cumonovalent
2+ on GO changed slightly by increasing Cl− (from 0.01 M to 0.1 M)
anions showed a negligible impact on Co(II) adsorption
[17]. Liu et al. (2013) demonstrated that monovalent anions showed a negligible impact on Co(II)
onto carbon nanotube-hydroxyapatite composites [37], which is consistent with the result obtained in
adsorption onto carbon nanotube-hydroxyapatite composites [37], which is consistent with the result
this study. Previous studies proved that ionic strength mainly took charge of the outer-sphere surface
obtained in this study. Previous studies proved that ionic strength mainly took charge of the outer-
A
complexation and ion
sphere surface exchange,and
complexation whereas ionic strength
ion exchange, whereas ionicshowed no impact
strength showed on no inner-sphere
impact on inner- surface
complexation [31]. Therefore, inner-sphere surface complexation
sphere surface complexation [31]. Therefore, inner-sphere surface complexation decided the decided the adsorption process of
Zn(II) adsorption
on GO over a wide
process pH range.
of Zn(II) on GO over a wide pH range.
TR
RE

Figure 2. Cont.
Nanomaterials 2018, 8, 806 6 of 13
Nanomaterials 2018, 8, x FOR PEER REVIEW 6 of 13

2. (A)2.Effect
FigureFigure of pH
(A) Effect on on
of pH Zn(II) adsorption
Zn(II) adsorptiononto GO in
onto GO inthe
thepresence
presence of of 0.001,
0.001, 0.01,0.01,
and and 0.1 mol/L
0.1 mol/L
NaClO4 ; (B) Distribution of Zn(II) species in aqueous solution (C0 = 10 mg/L, m/v = 0.12 g/L, h,
NaClO ; (B) Distribution of Zn(II) species in aqueous solution (C = 10 mg/L, m/v = 0.12 g/L, t = 24

D
4 0 t = 24 h,
T K);
T = 293 = 293(C)
K);Zeta
(C) Zeta potential
potential of of
GOGO dispersion as
dispersion as aa function
functionofofpHpH value.
value.

3.3. Adsorption
3.3. Adsorption Isotherms
Isotherms

TE
The adsorption
The adsorption isotherms
isotherms of of Zn(II)adsorption
Zn(II) adsorption on
onGO
GOwere
were simulated using
simulated the Langmuir
using and and
the Langmuir
Freundlich models, which are described as Equations (6) and (7):
Freundlich models, which are described as Equations (6) and (7):
Ce 1 1
C 1Ce  1 (6)
qe e =qm Ce + qm k (6)
qe qm qm k
1
ln qe  ln K f + ln Ce (7)
C
n 1
ln qe = ln Kf + ln Ce (7)
where Ce is the equilibrium Zn(II) concentration in the aqueous n solution (mg/L); qe and qm refer to the
Ce is the equilibrium
whereequilibrium adsorptionZn(II) concentration
capacity in the aqueous
on the adsorbent and thesolution
maximum (mg/L); qe and qcapacity,
adsorption m refer to the
A
respectively
equilibrium (mg/g).capacity
adsorption The values
on theof kadsorbent
(L/mg) and andKfthe
(mg/g) are the adsorption
maximum adsorption capacity,
constants respectively
of the
Langmuir and Freundlich models, respectively.
(mg/g). The values of k (L/mg) and Kf (mg/g) are the adsorption constants of the Langmuir and
FreundlichBy calculating the slope and intercept of the linear plots, the relative parameters of the Langmuir
models, respectively.
and Freundlich models obtained are exhibited in Table 1. Due to the high correlation coefficients (R2
By calculating the slope and intercept of the linear plots, the relative parameters of the Langmuir
TR

> 0.9272), these two models were sustainably employed to fit the experimental results in this study.
and Freundlich models
In Figure 3, it obtained
is clearly observed are
that exhibited
the behaviors in of
Table
Zn(II)1.sorption
Due toonto theGO high
can correlation coefficients
be better simulated
(R2 > using
0.9272),
the Freundlich model with correlation coefficients higher than 0.998 than Langmuir model within this
these two models were sustainably employed to fit the experimental results
study.correlation
In Figurecoefficients
3, it is clearly
lowerobserved
than 0.943 that the behaviors
in three of Zn(II) sorption
different temperatures, onto
indicating theGO can be better
adsorption
conducted
simulated usingon thea Freundlich
structurally heterogeneous sorbent [38].coefficients
model with correlation The qm of Zn(II)
higheron GO
thancounted usingLangmuir
0.998 than the
modelLangmuir model wascoefficients
with correlation 208.33 mg/g at 293 Kthan
lower (Table 1), which
0.943 was higher
in three thantemperatures,
different other nano-scaleindicating
sorbents the
such as: dioctylphetalate trithylenetetraamine magnetic nanoparticle (24.21 mg/g) [9], nZVI (20.00
adsorption conducted on a structurally heterogeneous sorbent [38]. The qm of Zn(II) on GO counted
RE

mg/g) [7], nanostructured birnessite-type manganese oxide and birnessite/carbon nanotubes


using the Langmuir model was 208.33 mg/g at 293 K (Table 1), which was higher than other nano-scale
(HB/CNTs) (89.50 mg/g) [8]. It was also higher than in other research on GO for Zn(II) removal, for
sorbents such as:
example dioctylphetalate
73 mg/g of GO prepared trithylenetetraamine
using amorphous graphite magnetic [39]nanoparticle
and 88.12 mg/g (24.21
of GO mg/g) [9], nZVI
prepared
(20.00 using
mg/g) [7], nanostructured
glycine, EDC and NHS [40]. birnessite-type
Moreover, the qmanganese oxidefrom
m values increased and208.33
birnessite/carbon nanotubes
mg/g to 211.42 mg/g
(HB/CNTs)
with the(89.50 mg/g)increasing
temperature [8]. It wasfromalso
293higher
K to 313 than in other
K (Table researchthat
1), indicating on higher
GO fortemperature
Zn(II) removal,
promoted
for example a higher
73 mg/g ofadsorption
GO prepared capacity
usingof Zn(II) on GO. graphite [39] and 88.12 mg/g of GO prepared
amorphous
using glycine, EDC and NHS [40]. Moreover, the qm values increased from 208.33 mg/g to 211.42 mg/g
with the temperature increasing from 293 K to 313 K (Table 1), indicating that higher temperature
promoted a higher adsorption capacity of Zn(II) on GO.
Nanomaterials 2018, 8, 806 7 of 13
Nanomaterials 2018, 8, x FOR PEER REVIEW 7 of 13

D
of Zn(II) adsorption on GO.
of Zn(II) adsorption on GO.

TE
Figure 3. (A) Langmuir isotherm plots of Zn(II) adsorption on GO; (B) The Freundlich isotherm plots
Figure 3. (A) Langmuir isotherm plots of Zn(II) adsorption on GO; (B) The Freundlich isotherm plots

Table 1. Relative parameters of the Langmuir and Frenudlich isotherm models.


Table 1. Relative parameters of the Langmuir and Frenudlich isotherm models.
C
Isotherm Models Parameters T = 293 K T = 303 K T = 313 K
Isotherm Models Parameters T = 293 K T = 303 K T = 313 K
qm (mg/g) 208.33 210.97 211.42
Langmuir qm (mg/g)
K (L/mg) 208.33
0.8727 210.97
1.0085 211.42
1.1537
A
Langmuir K (L/mg)R 2 0.8727
0.9096 1.0085
0.9219 1.1537
0.9196
R 2 1−n n
lnKf (mg ·L /g) 0.9096
4.3837 0.9219
4.4883 0.9196
4.5721
Frenudlich lnKf (mg1−n1/n
·Ln/g) 4.3837
0.6226 4.4883
0.6313 4.5721
0.6321
2 0.9991 0.9990 0.9991
Frenudlich 1/n R 0.6226 0.6313 0.6321
TR

R2 0.9991 0.9990 0.9991


3.4. Thermodynamic Parameters
3.4. Thermodynamic Parameters
Thermodynamic parameters involving Gibbs free energy change (∆G0 ), enthalpy change (∆H0 ),
and Thermodynamic (∆S0 ) are determined
entropy change parameters by the
involving Gibbs freetemperature dependent
energy change adsorption
(ΔG0), enthalpy isotherms
change (ΔH0),
according
and to the
entropy following
change (ΔS0) Equations (8)–(10):by the temperature dependent adsorption isotherms
are determined
RE

according to the following Equations (8)–(10):


qe
Kd = (8)
qe eC
Kd  (8)
∆G0 = C −eRT ln Kd (9)
0
G   RT∆H ln 0K d ∆S0 (9)
ln Kd = − 0 +0 (10)
H RT S R
ln K    (10)
where Kd is the distribution coefficient,d mL/g. RT ∆H0 R and ∆S0 can be obtained by calculating the
liner constants for the plot of lnKd versus 1/T (Figure 4). The minus values of ∆G0 implied that the
where Kd is the distribution coefficient, mL/g. ΔH0 and ΔS0 can be obtained by calculating the liner
behavior of Zn(II) adsorption on GO was a thermodynamically favorable and spontaneous process.
constants for the plot of0lnKd versus 1/T (Figure 4). The minus values of ΔG0 implied that the behavior
The magnitudes of ∆G reduced from −26.23 kJ/mol to −28.84 kJ/mol coupled with an increase in
of Zn(II) adsorption on GO was a thermodynamically favorable and spontaneous process. The
temperature from 293 to 313 K (Table 2), implying that higher temperatures favored more energetic
magnitudes of ΔG0 reduced from −26.23 kJ/mol to −28.84 kJ/mol coupled with an increase in
adsorption. As the value of ∆H0 was positive (12.00 kJ/mol), the adsorption of Zn(II) on GO was
temperature from 293 to 313 K (Table 2), implying that higher temperatures favored more energetic
an endothermic process. With the value of ∆S0 (130 J/mol/K) being higher than zero, the extent of
adsorption. As the value of ΔH was positive (12.00 kJ/mol), the adsorption of Zn(II) on GO was an
0
randomness was revealed to rise at the solid-solution interface during the sorption process of Zn(II)
endothermic process. With the value of ΔS0 (130 J/mol/K) being higher than zero, the extent of
on GO. It also demonstrated that increasing temperature benefited the spontaneous proceeding of
randomness was revealed to rise at the solid-solution interface during the sorption process of Zn(II)
adsorption behaviors.
on GO. It also demonstrated that increasing temperature benefited the spontaneous proceeding of
adsorption behaviors.
Nanomaterials 2018, 8, 806 8 of 13
Nanomaterials 2018, 8, x FOR PEER REVIEW 8 of 13

D
TE
Figure4.4.Plots
Figure Plotsof
oflnK
lnKdd vs.
vs. 1/T.
1/T.

Table 2. Thermodynamic parameters for the adsorption of Zn(II) on GO.


Table 2. Thermodynamic parameters for the adsorption of Zn(II) on GO.
T (K) ∆G0 (kJ/mol) ∆S0 (J/mol/K) ∆H0 (kJ/mol)
T (K) ΔG0 (kJ/mol) ΔS0 (J/mol/K) ΔH0 (kJ/mol)
293
293 −26.23
−26.23
C
303 −27.52 130 12.00
303
313
−27.52
−28.84 130 12.00
313 −28.84
A
3.5. Adsorption Kinetics
3.5. Adsorption Kinetics
The transient behavior of Zn(II) adsorption is fitted using two typical kinetic models, which are
The transient behavior of Zn(II) adsorption is fitted using two typical kinetic models, which are
expressed as:
expressed as:
TR

Pseudo first-order equation: ln(qe − qt ) = ln qe − k1 t (11)


Pseudo first-order equation: ln( qe  qtt )  ln1qe  k1tt (11)
Pseudo second-order equation: = + (12)
t qt 1k2 qe t qe
2
Pseudo second-order equation:   (12)
where k1 and k2 are the respective rate constants (g/(mg qt h)).k 2 qe 2 qe
Figure 5 shows the adsorption kinetics of Zn(II) adsorption on GO at 293 K. One can see that
where k1 and k2 are the respective rate constants (g/(mg h)).
the adsorption of Zn(II) on GO rapidly rose up to 75.9% with the reaction time increasing within
RE

Figure 5 shows the adsorption kinetics of Zn(II) adsorption on GO at 293 K. One can see that the
3 h. It should be mentioned here that the low removal percentage obtained was mainly due to
adsorption of Zn(II) on GO rapidly rose up to 75.9% with the reaction time increasing within 3 h. It
the low m/v ratio of 0.12 g/L applied in this study. After 6 h, the adsorption removal efficiency
should be mentioned here that the low removal percentage obtained was mainly due to the low m/v
of Zn(II) was slightly increased to 78.86%. Comparing the kinetic results of these two kinetic
ratio of 0.12 g/L applied in this study. After 6 h, the adsorption removal efficiency of Zn(II) was
models, the pseudo-second-order model can describe the adsorption kinetic of Zn on GO with higher
slightly increased to78.86%. Comparing the kinetic results of these two kinetic models, the pseudo-
correlation coefficients (R2 > 0.999) very well (Table 3). The adsorption behavior was determined to
second-order model can describe the adsorption kinetic of Zn on GO with higher correlation
be chemisorption [28]. Moreover, the adsorption capacity in the theoretical evaluation at equilibrium
coefficients (R2 > 0.999) very well (Table 3). The adsorption behavior was determined to be
(167.50 mg/g at 293 K) counted using the pseudo-second-order model was quite close to the adsorption
chemisorption [28]. Moreover, the adsorption capacity in the theoretical evaluation at equilibrium
capacity at equilibrium gained from the actual experiment (166.84 mg/g at 293 K). Moreover, the values
(167.50 mg/g at 293 K) counted using the pseudo-second-order model was quite close to the
of qe rose from 167.5 mg/g to 174.83 mg/g with the temperature increasing from 293 K to 313 K, which
adsorption capacity at equilibrium gained from the actual experiment (166.84 mg/g at 293 K).
corresponded with the findings obtained above. The kinetic data obtained in this study are also
Moreover, the values of qe rose from 167.5 mg/g to 174.83 mg/g with the temperature increasing from
consistent with other research demonstrating that the pseudo-second-order model can simulate the
293 K to 313 K, which corresponded with the findings obtained above. The kinetic data obtained in
kinetics of heavy metals sorption on GO well [28,32,41].
this study are also consistent with other research demonstrating that the pseudo-second-order model
can simulate the kinetics of heavy metals sorption on GO well [28,32,41].
Nanomaterials 2018, 8, 806 9 of 13
Nanomaterials 2018, 8, x FOR PEER REVIEW 9 of 13

D
TE
Figure 5. Adsorption
Figure 5. Adsorption kinetics
kinetics of
of Zn(II)
Zn(II)on
onGO
GO(C
(C00== 25
25 mg/L, m/v==0.12
mg/L, m/v 0.12g/L,
g/L,I I= =0.01
0.01mol/L
mol/L NaClO
NaClO T,
4, 4
T= 293
= 293
K).K).
Table 3. Kinetic parameters of Zn(II) adsorption on GO.
Table 3. Kinetic parameters of Zn(II) adsorption on GO.
Kinetic Models Parameters T = 293 K T = 303 K T = 313 K
C
Kinetic Models Parameters T = 293 K T = 303 K T = 313 K
qeq(mg/g)
e (mg/g) 78.16
78.16 96.75
96.75 92.12
92.12
Pseudo first-order k1 (h−1−1 ) 0.2947 0.3328 0.292
Pseudo first-order k1 (hR2 ) 0.2947
0.7486 0.3328
0.6888 0.292
0.7688
A
R2 0.7486 0.6888 0.7688
qe (mg/g) 167.50 169.78 174.83
qke (mg/g) 167.50 169.78 174.83
Pseudo second-order 2 (g/mg h) 0.015 0.014 0.018
Pseudo second-order k2 (g/mgR2 h) 0.015
0.9994 0.014
0.9991 0.018
0.9992
R2 0.9994 0.9991 0.9992
TR

3.6. Surface Complexation Modeling


3.6. Surface Complexation Modeling
By optimizing the simulated data under various pH cultures, the equilibrium constants obtained
from By
theoptimizing the simulated
surface complexation data under
modeling various pH
are exhibited cultures,
in Table 4. Asthe equilibrium
shown constants
in Figure 6, Zn(II)obtained
sorption
from the surface complexation modeling are exhibited in Table 4. As shown
on GO in pH-dependent conditions are satisfactorily fitted by applying DLM with mononuclear in Figure 6, Zn(II)
and monodentate complexes, SOZn species and SOZnOH species. It is clearly observed thatwith
sorption on GO in pH-dependent + conditions are satisfactorily fitted by applying DLM the
RE

mononuclear and monodentate complexes,


+ SOZn + species and SOZnOH species. It is clearly observed
main adsorption species was SOZn species at pH < 4, and the main SOZnOH species charged the
that the main
adsorption of adsorption
Zn(II) on GOspecies
at pH was
> 4. SOZn
+ species at pH < 4, and the main SOZnOH species charged
The findings from the surface complexation modeling suggested
the adsorption
that the sorptionofofZn(II)
Zn(II) on GO was
on GO at pH > 4. The
mainly findings surface
inner-sphere from the surface complexation
complexation, modeling
which corresponded
suggested
with that the
the findings ofsorption
the ionicof Zn(II) on
strength GO was experiments.
adsorption mainly inner-sphere surface complexation, which
corresponded with the findings of the ionic strength adsorption experiments.
Table 4. Theoptimized parameters of surface complexation modeling for Zn(II) adsorption on GO.
Table 4. Theoptimized parameters of surface complexation modeling for Zn(II) adsorption on GO.
Equations LogK
Equations LogK
Protonation and deprotonation
Protonation
SOH + Hand deprotonation
+ = SOH + 4.52
2
SOHSOH
+ H+==SO
SOH
− +2+H+ −7.88 4.52
SOHSurface
= SO− +complexation
H+ modeling −7.88
Surface
SOH complexation
+ Zn2+ = SOZn modeling
+ + H+ 3.48
2+ + H+ O = +SOZnOH + 2H+ −8.64 3.48
SOHSOH
+ Zn+2+Zn
= SOZn 2+ H
SOH + Zn + H2O = SOZnOH + 2H+
2+ −8.64
Nanomaterials 2018, 8, 806 10 of 13
Nanomaterials2018,
Nanomaterials 2018,8,
8,xxFOR
FORPEER
PEERREVIEW
REVIEW 10of
10 of13
13

D
TE
Figure 6.
Figure The fitted
6. The fitted results
results of
of surface
surface complexation
complexation modeling
modeling of Zn(II) on
of Zn(II)
Zn(II) on GO
on GO at
GO at different
at different pH
different pH levels.
pH levels.
levels.

3.7. Regeneration
3.7. Regeneration
Regeneration and and Reusability
and Reusability
Reusability
3.7.
The repeated
The repeated availability
repeated availability
availability ofof GO
of GO after
GO after desorption
after desorption
desorption by by using
by using
using 0.10.1
0.1 MM HCl
M HCl
HCl asas an
as an elution
an elution agent
elution agent
agent waswas
was
The
investigated. As shown in Figure 7, the adsorption capacity of Zn(II) on GO was slightly reduced.
investigated. As shown in Figure 7, the adsorption capacity of Zn(II) on GO was slightly reduced. ItIt
investigated. As shown in Figure 7, the adsorption capacity of Zn(II) on GO was slightly reduced.
It clearly
clearly demonstratesthat
demonstrates thatafter
afterthree
threecycles
cyclesof of adsorption-desorption
adsorption-desorptionthe the value
value ofof the adsorption
the adsorption
adsorption
C
clearly demonstrates that after three cycles of adsorption-desorption the value of the
capacity
capacity of
of Zn(II)
Zn(II) changed
changed from
from 166.83
166.83 mg/g
mg/g to
to 164.76
164.76 mg/g
mg/g (appropriate
(appropriate 98.76%
98.76% of
of GO
GO regenerated).
regenerated).
capacity of Zn(II) changed from 166.83 mg/g to 164.76 mg/g (appropriate 98.76% of GO regenerated).
After six
After six cycles, the
six cycles,
cycles, the adsorption
adsorptioncapacity
capacityofofZn(II)
Zn(II)was was153.84
153.84mg/g,
mg/g, which
which counted
counted for
for 92.23%92.23%
92.23% of theof
the
After the adsorption capacity of Zn(II) was 153.84 mg/g, which counted for of
the initial
initial capacity
capacity of the
of the
the GO.GO.GuoGuo et al.
et al.
al. (2017)
(2017) obtained
obtained 86%86% of the
of the
the initial
initial adsorption
adsorption capacity
capacity of Lof
L--
initial capacity of GO. Guo et (2017) obtained 86% of initial adsorption capacity of
A
L -arginine modified magnetic adsorbent after five cycles of adsorption-regeneration on Zn(II) using
arginine modified magnetic adsorbent after five cycles of adsorption-regeneration on Zn(II) using aa
arginine modified magnetic adsorbent after five cycles of adsorption-regeneration on Zn(II) using
a 6mol/L
mol/L HCl solution(the(the maximum adsorption capacity
was was
150.4150.4 mg/g) [42].etLiu et al. (2017)
66 mol/L HCl
HCl solution
solution (the maximum
maximum adsorption
adsorption capacity
capacity was 150.4 mg/g)
mg/g) [42]. Liu
[42]. Liu al. (2017)
et al. (2017) found
found
found
90% of 90%
of the of the
the initial initial
initial Zn(II) Zn(II)
Zn(II) removal removal
removal capacity capacity
capacity using using nanostructured
using nanostructured birnessite-type
nanostructured birnessite-type
birnessite-type manganese manganese
manganese oxide oxide oxide
and
90% and
and birnessite/carbon nanotubes (HB/CNTs) (the maximum adsorption capacity was 89.5 mg/g) [43].
TR

birnessite/carbon nanotubes (HB/CNTs) (the maximum adsorption capacity


birnessite/carbon nanotubes (HB/CNTs) (the maximum adsorption capacity was 89.5 mg/g) [43]. As was 89.5 mg/g) [43]. As
As a whole, the results here demonstrated that GO has satisfactory reuse potential and excellent
aa whole,
whole, the the results
results here
here demonstrated
demonstrated that that GO GO hashas satisfactory
satisfactory reuse
reuse potential
potential and and excellent
excellent
regeneration
regeneration performance
performance with
witha highly
a highlyefficient adsorption
efficient capacity
adsorption for Zn(II)
capacity for removal
Zn(II) from aqueous
removal from
regeneration performance with a highly efficient adsorption capacity for Zn(II) removal from
solutions,
aqueous which corresponded
solutions, which with the previous
corresponded with the study
previous[3].study [3].
aqueous solutions, which corresponded with the previous study [3].
RE

Figure 7.
Figure
Figure 7. Adsorption-desorption
Adsorption-desorption cycles
Adsorption-desorption cycles of
cyclesof Zn(II)
ofZn(II) on
Zn(II)on GO
onGO nanosheets
GOnanosheets (C000===10
(C
nanosheets(C 10 mg/L,
10mg/L,
mg/L,m/v
m/v
m/v == 0.12
0.12 g/L,
= 0.12 II ==
g/L,
g/L,
I0.01
0.01 mol/L
= 0.01 NaClO
mol/L
mol/L 4, T =, 293
NaClO
NaClO 4, T =4 293
K).
T =K).
293 K).
Nanomaterials 2018, 8, 806 11 of 13

4. Conclusions
Based on various characterization techniques, plentiful oxygenic functionalities were presented
on the surface of GO, which contributed to the highly efficient sorption of Zn(II). The sorption of
Zn(II) on GO was irrelevant with ionic strength, revealing that the inner-sphere surface complexation
dominated Zn(II) adsorption. According to the Langmuir model, the maximum adsorption capacity
of GO for Zn(II) was 208.33 mg/g at pH 7.0 and 293 K. Thermodynamic parameters determined that
the adsorption process of Zn(II) on GO was endothermic and spontaneous. The fitting findings of
the surface complexation modeling suggested that pH-dependent sorption of Zn(II) on GO can be
satisfactorily fitted using DLM with mononuclear and monodentate complexes, SOZn+ species and
SOZnOH species, at a lower and higher pH, respectively. The findings of this study showed that GO
has a potential use in heavy metal removal from aqueous solutions.

D
Author Contributions: M.P. and G.W. conceived and designed the experiments; X.L. and X.H. performed the
experiments and contributed to the reagents, materials and analysis; C.L. organized the characterization analysis;
and M.P. and X.H. contributed to the drafting, writing, and editing of the manuscript.
Acknowledgments: The authors would like to extend their thanks for the financial support provided by the

TE
Educational Research Project from the Education Department of Fujian Province (JAT170411), the Open Research
Fund Program from Fujian Engineering and Research Center of Rural Sewage Treatment and Water Safety
(RST201804), the Science and Technology Project of Longyan City (2017LY63), the Natural Science Foundation
of Anhui Province (1808085QD106), and the Natural Science Foundation of Education Department of Anhui
Province (KJ2016A270).
Conflicts of Interest: The authors declare no conflict of interest.
C
References
1. Lingamdinne, L.P.; Koduru, J.R.; Choi, Y.-L.; Chang, Y.-Y.; Yang, J.-K. Studies on removal of Pb(II) and Cr(III)
using graphene oxide based inverse spinel nickel ferrite nano-composite as sorbent. Hydrometallurgy 2016,
A
165, 64–72. [CrossRef]
2. Wu, S.; Kong, L.; Liu, J. Removal of mercury and fluoride from aqueous solutions by three-dimensional
reduced-graphene oxide aerogel. Res. Chem. Intermed. 2015, 42, 4513–4530. [CrossRef]
TR

3. Wang, H.; Yuan, X.; Wu, Y.; Huang, H.; Zeng, G.; Liu, Y.; Wang, X.; Lin, N.; Qi, Y. Adsorption characteristics
and behaviors of graphene oxide for Zn(II) removal from aqueous solution. Appl. Surf. Sci. 2013, 279,
432–440. [CrossRef]
4. Ostroski, I.C.; Barros, M.A.; Silva, E.A.; Dantas, J.H.; Arroyo, P.A.; Lima, O.C. A comparative study for the
ion exchange of Fe(III) and Zn(II) on zeolite NaY. J. Hazard. Mater. 2009, 161, 1404–1412. [CrossRef] [PubMed]
5. Ferreira, L.S.; Rodrigues, M.S.; de Carvalho, J.C.M.; Lodi, A.; Finocchio, E.; Perego, P.; Converti, A. Adsorption
of Ni2+ , Zn2+ and Pb2+ onto dry biomass of Arthrospira (Spirulina) platensis and Chlorella vulgaris. I. Single
RE

metal systems. Chem. Eng. J. 2011, 173, 326–333. [CrossRef]


6. Lu, C.; Chiu, H. Chemical modification of multiwalled carbon nanotubes for sorption of Zn2+ from aqueous
solution. Chem. Eng. J. 2008, 139, 462–468. [CrossRef]
7. Khademi, Z.; Ramavandi, B.; Ghaneian, M.T. The behaviors and characteristics of a mesoporous activated
carbon prepared from Tamarix hispida for Zn(II) adsorption from wastewater. J. Environl. Chem. Eng. 2015,
3, 2057–2067. [CrossRef]
8. Liu, H.; Zhu, Y.; Xu, B.; Li, P.; Sun, Y.; Chen, T. Mechanical investigation of U(VI) on pyrrhotite by batch,
EXAFS and modeling techniques. J. Hazard. Mater. 2017, 322, 488–498. [CrossRef] [PubMed]
9. Ghasemi, N.; Ghasemi, M.; Moazeni, S.; Ghasemi, P.; Alharbi, N.S.; Gupta, V.K.; Agarwal, S.; Burakova, I.V.;
Tkachev, A.G. Zn(II) removal by amino-functionalized magnetic nanoparticles: Kinetics, isotherm,
and thermodynamic aspects of adsorption. J. Ind. Eng. Chem. 2018, 62, 302–310. [CrossRef]
10. Krzisnik, N.; Mladenovic, A.; Skapin, A.S.; Skrlep, L.; Scancar, J.; Milacic, R. Nanoscale zero-valent iron
for the removal of Zn2+ , Zn(II)-EDTA and Zn(II)-citrate from aqueous solutions. Sci. Total Environ. 2014,
476–477, 20–28. [CrossRef] [PubMed]
Nanomaterials 2018, 8, 806 12 of 13

11. Gogoi, H.; Leiviska, T.; Heiderscheidt, E.; Postila, H.; Tanskanen, J. Removal of metals from industrial
wastewater and urban runoff by mineral and bio-based sorbents. J. Environ. Manag. 2018, 209, 316–327.
[CrossRef] [PubMed]
12. Peng, W.; Li, H.; Liu, Y.; Song, S. A review on heavy metal ions adsorption from water by graphene oxide
and its composites. J. Mol. Liq. 2017, 230, 496–504. [CrossRef]
13. Sun, Y.; Li, J.; Wang, X. The retention of uranium and europium onto sepiolite investigated by macroscopic,
spectroscopic and modeling techniques. Geochim. Cosmochim. Acta 2014, 140, 621–643. [CrossRef]
14. Sun, Y.; Wang, X.; Ai, Y.; Yu, Z.; Huang, W.; Chen, C.; Hayat, T.; Alsaedi, A.; Wang, X. Interaction of sulfonated
graphene oxide with U(VI) studied by spectroscopic analysis and theoretical calculations. Chem. Eng. J. 2017,
310, 292–299. [CrossRef]
15. Tang, J.; Huang, Y.; Gong, Y.; Lyu, H.; Wang, Q.; Ma, J. Preparation of a novel graphene oxide/Fe-Mn
composite and its application for aqueous Hg(II) removal. J. Hazard. Mater. 2016, 316, 151–158. [CrossRef]
[PubMed]

D
16. Zhang, Y.; Zhang, S.; Gao, J.; Chung, T.-S. Layer-by-layer construction of graphene oxide (GO) framework
composite membranes for highly efficient heavy metal removal. J. Membrane Sci. 2016, 515, 230–237.
[CrossRef]
17. Tan, P.; Bi, Q.; Hu, Y.; Fang, Z.; Chen, Y.; Cheng, J. Effect of the degree of oxidation and defects of graphene

TE
oxide on adsorption of Cu2+ from aqueous solution. Appl. Surf. Sci. 2017, 423, 1141–1151. [CrossRef]
18. Wan, S.; He, F.; Wu, J.; Wan, W.; Gu, Y.; Gao, B. Rapid and highly selective removal of lead from water using
graphene oxide-hydrated manganese oxide nanocomposites. J. Hazard. Mater. 2016, 314, 32–40. [CrossRef]
[PubMed]
19. Yari, M.; Rajabi, M.; Moradi, O.; Yari, A.; Asif, M.; Agarwal, S.; Gupta, V.K. Kinetics of the adsorption of
C
Pb(II) ions from aqueous solutions by graphene oxide and thiol functionalized graphene oxide. J. Mol. Liq.
2015, 209, 50–57. [CrossRef]
20. Zhang, J.; Gong, J.-L.; Zenga, G.-M.; Ou, X.-M.; Jiang, Y.; Chang, Y.-N.; Guo, M.; Zhang, C.; Liu, H.-Y.
Simultaneous removal of humic acid/fulvic acid and lead from landfill leachate using magnetic graphene
A
oxide. Appl. Surf. Sci. 2016, 370, 335–350. [CrossRef]
21. Sun, Y.; Shao, D.; Chen, C.; Yang, S.; Wang, X. Highly efficient enrichment of radionuclides on graphene
oxide-supported polyaniline. Environ. Sci. Technol. 2013, 47, 9904–9910. [CrossRef] [PubMed]
TR

22. Sun, Y.; Chen, C.; Tan, X.; Shao, D.; Li, J.; Zhao, G.; Yang, S.; Wang, Q.; Wang, X. Enhanced adsorption of
Eu(III) on mesoporous Al2 O3 /expanded graphite composites investigated by macroscopic and microscopic
techniques. Dalton Trans. 2012, 41, 13388–13394. [CrossRef] [PubMed]
23. Sun, Y.; Yang, S.; Chen, Y.; Ding, C.; Cheng, W.; Wang, X. Adsorption and desorption of U(VI) on
functionalized graphene oxides: A combined experimental and theoretical study. Environ. Sci. Technol. 2015,
49, 4255–4262. [CrossRef] [PubMed]
24. Yang, P.; Liu, Q.; Liu, J.; Zhang, H.; Li, Z.; Li, R.; Liu, L.; Wang, J. Bovine serum albumin-coated graphene
RE

oxide for effective adsorption of Uranium(VI) from aqueous solutions. Ind. Eng. Chem. Res. 2017, 56,
3588–3598. [CrossRef]
25. Russo, P.; D’Urso, L.; Hu, A.; Zhou, N.; Compagnini, G. In liquid laser treated graphene oxide for dye
removal. Appl. Surf. Sci. 2015, 348, 85–91. [CrossRef]
26. Parmar, K.R.; Murthy, Z.V.P. Synthesis of acetone reduced graphene oxide/Fe3 O4 composite through simple
and efficient chemical eduction of exfoliated graphene oxide for removal of dye from aqueous solution.
J. Mater. Sci. 2014, 49, 6772–6783. [CrossRef]
27. Peng, Y.; Li, J. Ammonia adsorption on graphene and graphene oxide: A first-principles study. Front. Environ.
Sci. Eng. 2013, 7, 403–411. [CrossRef]
28. Huang, X.; Pan, M. The highly efficient adsorption of Pb(II) on graphene oxides: A process combined by
batch experiments and modeling techniques. J. Mol. Liq. 2016, 215, 410–416. [CrossRef]
29. Ding, C.; Cheng, W.; Sun, Y.; Wang, X. Determination of chemical affinity of graphene oxide nanosheets with
radionuclides investigated by macroscopic, spectroscopic and modeling techniques. Dalton Trans. 2014, 43,
3888–3896. [CrossRef] [PubMed]
30. Huang, X.; Chen, T.; Zou, X.; Zhu, M.; Chen, D.; Pan, M. The Adsorption of Cd(II) on Manganese Oxide
Investigated by Batch and Modeling Techniques. Int. J. Environ. Res. Public Health 2017, 14, 1145. [CrossRef]
[PubMed]
Nanomaterials 2018, 8, 806 13 of 13

31. Li, J.; Zhang, S.; Chen, C.; Zhao, G.; Yang, X.; Li, J.; Wang, X. Removal of Cu(II) and fulvic acid by graphene
oxide nanosheets decorated with Fe3 O4 nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 4991–5000.
[CrossRef] [PubMed]
32. Li, P.; Gao, H.; Wang, Y. Uptake of Ni(II) from aqueous solution onto graphene oxide: Investigated by batch
and modeling techniques. J. Mol. Liq. 2017, 227, 303–308. [CrossRef]
33. Raghubanshi, H.; Ngobeni, S.M.; Osikoya, S.O.; Shooto, N.E.; Dikio, C.W.; Naidoo, E.B.; Dikio, E.D.;
Pandey, R.K.; Prakash, R. Synthesis of graphene oxide and its application for the adsorption pf Pb2+ from
aqueous solution. J. Ind. Eng. Chem. 2017, 47, 169–178. [CrossRef]
34. Jin, Z.; Sheng, J.; Sun, Y. Characterization of radioactive cobalt on graphene oxide by macroscopic and
spectroscopic techniques. J. Radioanal. Nucl. Chem. 2014, 299, 1979–1986. [CrossRef]
35. Qi, Y.; Yang, M.; Xu, W.; He, S.; Men, Y. Natural polyaccharides-modified graphene oxide for adsorption of
organic dyes from aqueous solutions. J. Colloid Interf. Sci. 2017, 486, 84–96. [CrossRef] [PubMed]
36. Weng, C.-H.; Huang, C.P. Adsorption characteristics of Zn(II) from dilute aqueous solution by fly ash.

D
Colloid Surf. A 2004, 247, 137–143. [CrossRef]
37. Liu, Z.; Chen, L.; Zhang, Z.; Li, Y.; Dong, Y.; Sun, Y. Synthesis of multi-walled carbon
nanotube–hydroxyapatite composites and its application in the sorption of Co(II) from aqueous solutions.
J. Mol. Liq. 2013, 179, 46–53. [CrossRef]

TE
38. Pan, M.; Lin, X.; Xie, J.; Huang, X. Kinetic, equilibrium and thermodynamic studies for phosphate adsorption
on aluminum hydroxide modified palygorskite nano-composites. RSC Adv. 2017, 7, 4492–4500. [CrossRef]
39. Yang, Q.; Yang, G.; Peng, W.; Song, S. Adsorption of Zn(II) on graphene oxide prepared from low-purity of
amorphous graphite. Surf. Interface Anal. 2017, 49, 398–404. [CrossRef]
40. Najafi, F. Removal of zinc(II) ion by graphene oxide(GO) and functionalized graphene oxide-glycine(GO-G)
C
as adsorbents from aqueous solution: Kinetics studies. Int. Nano Lett. 2015, 5, 171–178. [CrossRef]
41. Huang, J.; Wu, Z.; Chen, L.; Sun, Y. Surface complexation modeling of adsorption of Cd(II) on graphene
oxides. J. Mol. Liq. 2015, 209, 753–758. [CrossRef]
42. Guo, S.; Jiao, P.; Dan, Z.; Duan, N.; Chen, G.; Zhang, J. Preparation of L-arginine modified magnetics
A
adsorbent by one-step method for removal of Zn(II) and Cd(II) from aqueous solution. Chem. Eng. J. 2017,
317, 999–1011. [CrossRef]
43. Liu, L.; Qiu, G.; Suib, S.L.; Liu, F.; Zhen, L.; Tan, W.; Qin, L. Enhancement of Zn2+ and Ni2+ removal
TR

performance using a deionization pseudocapacitor with nanostructured birnessite and its carbon nanotube
composite electrodes. Chem. Eng. J. 2017, 328, 464–473. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
RE

You might also like