Вы находитесь на странице: 1из 11

Chemical Engineering Journal 358 (2019) 1399–1409

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Novel insight into adsorption and co-adsorption of heavy metal ions and an T
organic pollutant by magnetic graphene nanomaterials in water
Dan Huanga,b, Jizi Wua,b, Lu Wanga,b, Xingmei Liua,b, Jun Menga,b, Xianjin Tanga,b,
Caixian Tanga,c, Jianming Xua,b,

a
Institute of Soil and Water Resources and Environmental Science, College of Environmental and Resource Sciences, Zhejiang University, 866 Yuhangtang Road, Hangzhou
310058, China
b
Zhejiang Provincial Key Laboratory of Agricultural Resources and Environment, Zhejiang University, 866 Yuhangtang Road, Hangzhou 310058, China
c
Department of Animal, Plant and Soil Sciences, La Trobe University, Melbourne Campus, Bundoora, VIC 3086, Australia

HIGHLIGHTS GRAPHICAL ABSTRACT

• MGO showed the best adsorption ca-


pacities for TC, Cd(II) and As(V)
among three magnetic graphene na-
nomaterials.
• Adsorption behaviors of MGO in three
different types of binary systems were
systematically studied.
• Mutual effects of TC and Cd(II) in the
simultaneously added system were
negligible.
• Adsorption of As(V) was significantly
suppressed in the presence of TC.
• Cd(II) and As(V) were co-adsorbed by
MGO via electrostatic attraction and
formation of type A ternary surface
complexation.

ARTICLE INFO ABSTRACT

Keywords: The adsorption of tetracycline (TC), cadmium [Cd(II)] and arsenate [As(V)] onto magnetic graphene oxide
Magnetic graphene nanomaterials (MGO), magnetic chemically-reduced graphene (MCRG) and magnetic annealing-reduced graphene (MARG) was
Co-adsorption investigated to understand the adsorption properties and molecular mechanisms. The adsorption of three con-
Heavy metal taminants was pH-dependent and the adsorption capability followed the order of MGO > MCRG > MARG, and
Organic pollutant
hence MGO was selected to systematically study the adsorption behaviors in three different types of binary
Water purification
systems. The maximum adsorption capacities of MGO were 252 mg/g for TC, 234 mg/g for Cd(II) and 14 mg/g
for As(V). The superiority of MGO was mainly attributed to its high dispersibility, thin nanosheets and various O-
containing functional groups. In addition to H-bonding and π–π interactions, the strong adsorption of TC onto
MGO was mainly due to the n–π electron-donor–acceptor (EDA) effect, with the maximum adsorption around
pKa of TC. The mutual effects of TC and Cd(II) in the simultaneously added system were negligible. Adsorption
of As(V) was significantly suppressed in the presence of TC, whilst As(V) hardly affected TC adsorption. The
adsorptions of Cd(II) and As(V) in the co-adsorption system were increased by 65% and 30%, respectively. This


Corresponding author at: Institute of Soil and Water Resources and Environmental Science, College of Environmental and Resource Sciences, Zhejiang University,
866 Yuhangtang Road, Hangzhou 310058, China.
E-mail address: jmxu@zju.edu.cn (J. Xu).

https://doi.org/10.1016/j.cej.2018.10.138
Received 24 May 2018; Received in revised form 15 October 2018; Accepted 16 October 2018
Available online 17 October 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

synergistic effect resulted from the electrostatic attraction and the formation of type A ternary surface com-
plexation. These new insights were valuable for elucidating the interaction mechanisms and designing novel
adsorbents for traditional and emerging pollutions in practical application.

1. Introduction magnetic annealing-reduced graphene (MARG), which contained dif-


ferent microstructures, iron and oxygen contents, and displayed various
Water quality has been deteriorated continuously due to the rapid sorption capacities. Cadmium [Cd(II)] and arsenate [As(V)] were se-
growth of population, urbanization, industrialization, and other environ- lected as model cation and oxyanion contaminant, respectively [6,42],
mental issues [1–3]. Multiple pollutants often coexist in aquatic environ- and tetracycline (TC) was selected as a model organic contaminant
ments and change chemical, physical or biological properties of water [43]. The main objective was to examine the properties and unique
[4–6]. If inadequately treated, aquatic ecosystems could be adversely af- molecular adsorptive and co-adsorptive mechanisms of these magnetic
fected by the combined pollutions [7,8]. The treatment of soil and water graphene nanomaterials. The adsorption isotherms were established
co-contaminated with heavy metals and organic pollutants poses a sig- and effects of pH were examined to evaluate the adsorption behaviors
nificant challenge because these two types of pollutants have different fates of each of these pollutants on MGO, MCRG, and MARG in single sys-
and transport mechanisms [9]. Sorption is widely used to eliminate both tems. Besides, MGO was selected to further study the adsorption be-
inorganic and organic pollutants from aqueous solutions owing to the haviors, interactions between the pollutants and associated mechanisms
simplicity of design, low cost, high efficiency, and wide adaptability [1,10]. in three diverse types of binary solutions.
As two-dimensional carbon-based materials with honeycomb structures
that are sp2-hybridized with an atom thickness, graphene and its derivatives 2. Materials and methods
were reported to show excellent performance in adsorption of various en-
vironmental contaminants, such as aromatic compounds [11–13], dyes 2.1. Synthesis of magnetic graphene nanocomposites
[14–16], antibiotics [17,18], estrogens [19–21], heavy metal ions [22,23],
anions [24,25] and combined pollutions [6,26]. Especially, graphene oxide Graphene oxide (GO) was synthesized from natural graphite flakes via
(GO), the main precursor of graphene, has been widely applied in waste- a modified Hummers’ method [44]. The adhesive GO was exfoliated by
water treatment or water purification due to its large specific surface area sonication and dialyzed to remove acids and other impurities. The Fe3O4
and a variety of O-containing functional groups such as hydroxide, epoxide, nanoparticles (NPs) were prepared by chemical co-precipitation of
carbonyl and carboxyl groups [4,26,27]. However, the separation proce- FeCl3·6H2O and FeSO4·7H2O. The sol of Fe3O4 NPs was obtained after so-
dures for this type of absorbents are often complicated due to their high nication for 2 h [45]. The GO solution was then added dropwise into the sol
dispersibility in water, which could lead to new environmental risks [4,28]. of Fe3O4 NPs under appropriate concentrations and pH conditions under
Magnetization by introducing Fe3O4 nanoparticles (NPs) onto GO and re- mechanical stirring, and the self-assembly of MGO was conducted [46].
duced graphene oxide (RGO) can bring about separation convenience, even MCRG was obtained by the reduction of as-synthesized MGO with hy-
increased adsorption capacity, which favor the practical applications drazine hydrate. MARG was achieved by a 500 °C expansion under N2
[29–31]. using MGO as the precursor. MGO, MCRG, and MARG of different oxygen
A number of studies have been conducted on the adsorption of contents were used as magnetic graphene adsorbents. The detailed proce-
magnetic graphene nanomaterials (MGNs) for heavy metals and organic dures are presented in the Supporting Information (SI).
contaminants in aqueous systems [32,33]. However, the existing lit-
erature is inadequate in concerning the interactions between multi- 2.2. Characterization of magnetic graphene nanocomposites
pollutants in systems, whereas coexistence of various pollutants is much
more common in real environments [6,12]. Antibiotics (e.g., TC) and The structures and surface morphologies of MGO, MCRG, and
heavy metals (e.g., Cd(II)) are extensively employed as growth pro- MARG were characterized by elemental analyses, the BET-N2 specific
moters and usually coexist in wastewater from livestock farm, causing surface areas, scanning electron microscopy (SEM), transmission elec-
considerable toxicological concerns [34]. Furthermore, multiple con- tron microscopy (TEM), Raman spectra and X-ray diffraction (XRD).
taminations in natural environments influence the removal of in- The surface functional groups were elucidated by X-ray photoelectron
dividual contaminants via electrostatic interaction, cation–π interac- spectroscopy (XPS) and Fourier transform infrared spectroscopy (FTIR).
tion, precipitation or/and complexation [5,35,36]. The magnetic properties of the MGNs were studied using a vibrating
Chemically-synthesized graphene nanosheets are generally defective sample magnetometer (VSM). The amount of Fe was quantified using
and contain surface heterogeneity with various structures such as flat atomic absorption spectrometry (AAS) after microwave digestion.
surface, wrinkles, defects, and O-containing functional groups, which are Surface charges were characterized by the zeta potential (ZP) method.
valuable for the high capacity of pollutant sorption [37–39]. The surface The detailed methods are presented in the SI.
properties could be greatly altered by these structures, and affect the in-
trinsic properties and environmental applications of graphene-based ma- 2.3. Batch experiments
terials. For instance, metal ions are preferentially adsorbed onto GO rather
than RGO, due to the abundant surface functional groups on GO [4,5]. The Tetracycline, Cd(II) and As(V) were selected as model pollutants to
reduction of exfoliated GO removed the O-containing functional groups investigate the adsorption processes of magnetic graphene materials.
and repaired a sp2-hybridized structure [40], leading to weakened surface All the adsorptions were carried out in Teflon screw-cap vials at
complexations of heavy metal ions with oxidized sites [41], enhanced π–π 25 ± 1 °C. Isotherm experiments were conducted with different solid-
interactions between graphene nanosheets and organic pollutants, and re- to-water ratios for various adsorbates in a background solution, which
duced competition of water molecules with organic pollutants at the oxi- contained 0.01 M NaNO3 in Milli-Q water with 200 mg/L NaN3 as a bio-
dized sites [5,11]. Several preparative strategies have been explored for inhibitor at pH 5.0 ± 0.1. The adsorption conditions were all kept the
graphene morphology regulation [37,38], while designing MGNs with same in single-solute and binary systems. The detailed experimental
special physicochemical properties and structural features is needed in processes are described in the SI.
practical applications. The MGO was selected to further study the adsorption behaviors in
In this study, we prepared and characterized magnetic graphene binary solutions due to its superior adsorption capacities for TC, Cd(II)
oxide (MGO), magnetic chemically-reduced graphene (MCRG) and and As(V). Different initial concentrations of TC, Cd(II) and As(V) were

1400
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

pairwise combined in adsorption evaluations, including organic pollu- Table 1


tant-cation [marked as (TC + Cd)], organic pollutant-oxyanion Elemental compositions and atomic ratios, BET-N2 surface area (SA) and ID/IG
[marked as (TC + As)], and cation-oxyanion [marked as (Cd + As)] of Raman spectra of magnetic graphene oxide (MGO), magnetic chemically-
binary systems. The co-adsorption experiments were designed as fol- reduced graphene (MCRG), and magnetic annealing-reduced graphene
(MARG).
lows: (1) TC (5, 50 or 200 mg/L) and Cd(II) (5, 25 or 50 mg/L) were
adsorbed simultaneously (TC + Cd); (2) TC (5, 50 or 200 mg/L) and As Composition MGO MCRG MARG
(V) (5, 25 or 50 mg/L) were adsorbed simultaneously (TC + As); (3) Cd
C (wt%) 15.26 16.43 14.34
(II) (5, 25 or 50 mg/L) and As(V) (5, 25 or 50 mg/L) were adsorbed
H (wt%) 0.901 0.638 0.402
simultaneously (Cd + As). O (wt%) 38.62 32.41 29.24
N (wt%) 0.028 0.617 0.063
Fe (wt%) 45.19 49.91 55.96
3. Results and discussion H/Ca 0.709 0.466 0.336
C]C/C–C (%)b 41.00 64.65 78.21
C–O (%)b 15.01 28.55 21.79
3.1. Characterization of magnetic graphene nanocomposites C]O (%)b 40.23 6.80 –
O–C]O(%)b 3.76 – –
The TEM image (Fig. 1a) of MGO showed that the Fe3O4 NPs SA (m2/g) 148.2 156.6 78.2
(10–20 nm size) were well dispersed in 2–5 layers of GO matrix, and the ID/IG 1.02 1.24 1.03

folding nature of graphene sheets was clearly visible. The basal plane of a
H/C: atomic ratio of hydrogen to carbon calculated from by elemental
the GO was almost unoccupied by the magnetic NPs, consistent with the analysis.
previous finding [47]. Dense aggregates of Fe3O4 NPs were observed in b
C1s bonding state and its relative atomic percentage on activated graphene
MCRG and MARG (Fig. 1b-c), indicating that the reduction of MGO were determined by XPS. – represents none or not detected.
induced serious wrinkles and overlaps on graphene nanosheets.
Thermal reduction of MGO to MARG decreased the BET specific surface MGO, and that the reduction decreased newly formed sp2 domains in
areas (SA) from 148 to 78 m2/g (Table 1), while chemical reduction of size [37]. As shown in Fig. 1f, the values of magnetic saturation (Ms) of
MGO to MCRG increased the SA to 157 m2/g. MGO, MCRG and MARG were 34.0, 55.6 and 67.3 emu/g, respectively,
In the XRD patterns (Fig. 1d), the diffraction peaks at 2ϴ values of suggesting that the magnetism was strong enough for many practical
30.2°, 35.6°, 43.3°, 53.7°, 57.1°, and 62.8° were assigned to the (2 2 0), applications [47].
(3 1 1), (4 0 0), (4 2 2), (5 1 1), (4 4 0), and (5 3 3) reflections, respec- The peaks of Fe 2p, C 1s and O 1s were observed in XPS spectra,
tively. These diffraction peaks were consistent with the characteristic indicating that the obtained MGO, MCRG and MARG contained Fe, O
crystalline spinel structure (JCPDS 88–0315) of Fe3O4, revealing that and C elements, which were in accordance with the result from the
magnetic Fe3O4 NPs were successfully synthesized [48,49]. The in- elemental analysis (Table 1 and Fig. S1). Carbon concentrations of the
tensity ratio of the D and G bands (ID/IG) is often used to estimate the three materials were similar, while oxygen concentrations decreased
amounts of defects on carbon materials [14]. The ID/IG value (Fig. 1e) from 38.6% for MGO to 32.4% for MCRG and 29.2% for MARG, in-
of MCRG (1.24) was higher than those of MGO (1.02) and MARG dicating that O-containing functional groups were diminished by the
(1.03), indicating that MCRG had more graphitic structure defects than

Fig. 1. TEM images of magnetic graphene oxide (MGO) (a), magnetic chemically-reduced graphene (MCRG) (b) and magnetic annealing reduced graphene (MARG)
(c); XRD patterns (d), Raman spectra (e) and magnetization curves (f) of MGO, MCRG and MARG. The inset in f shows photographs of the aqueous solution of the
MGO before (left) and after attracted by a magnetic bar (right).

1401
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

reduction process. In contrast, Fe concentrations increased from 45.2% 252.0 mg/g, which was much higher than those of MCRG (104.5 mg/g)
for MGO to 49.9% for MCRG and 56.0% for MARG, which were con- and MARG (24.2 mg/g) in this study (Table S1). The capacity of MGO to
sistent with the magnetization results. The MARG had lower H/C adsorb TC was significantly greater than those of MCRG and MARG.
atomic ratios and was more aromatic than MCRG and MGO [5]. Such difference in adsorption was primarily due to the following rea-
The compositions of surface functional groups of MGNs were ex- sons. First, the abundant O-containing functional groups on MGO led to
amined by FTIR spectra and XPS (Fig. 2). The broad bands at ap- its stable dispersion in the water, and the actual available surface area
proximately 3402 and 1570 cm−1 were assigned to the –OH bonds and of MGO in the aqueous solution was far above the value measured in
the H–O–H stretching and bending vibrations of the adsorbed water the solid state [52]. Second, in addition to π–π interaction and hydro-
molecules. For MGO, the peaks at 1406 and 1052 cm−1 were attributed phobic interaction, MGO could interact with TC via H-bonding [53].
to the stretching vibration of carboxyl O–C]O and alkoxy C–O bonds, Third, an appreciable amount of Fe3O4 NPs in the composite material
respectively. For MCRG and MARG, the intensity of –OH bonds at might increase the adsorption sites [54]. The Qmax of TC in this study
3402 cm−1 was lower, and those of O–C]O peaks at 1395 cm−1 and was much higher than those observed in previous studies, in which the
C–O at 1162 cm−1 were weaker or absent compared with that of MGO. Qmax of TC was 39.1 mg/g on functionalized magnetic particles [55],
Regarding the C 1s spectrum (Fig. 2b-f), four peaks at 284.7, 285.3, approximately 70 mg/g on GO and RGO [53], and 95 mg/g on 6-nm-
287.0, and 288.9 eV corresponded to the C]C/C–C (aromatic rings), Fe3O4-RGO [56].
C–O (phenolic hydroxyl/epoxy), C]O (carbonyl), and O–C]O (car- The Qmax of Cd(II) onto MGO (234.3 mg/g) was distinctly higher
boxyl groups), respectively [50,51]. The fitting results indicated that than those of Cd(II) onto MCRG (15.2 mg/g) and MARG (5.8 mg/g)
MGO had apparently more C]O and O–C]O groups than MCRG and (Table S1). Compared with other graphene-based materials, such as
MARG, and the quantity of C]C/C–C followed the order of MARG > magnetic graphene oxide nanocomposite (91.3 mg/g) [57], GO
MCRG > MGO. The results also suggested that the aromatic C in- (35.7 mg/g) [5], and amino-functionalized magnetic graphene compo-
creased via reduction and that the thermal reduction could reduce MGO site (27.8 mg/g) [58], MGO performed extremely well on Cd(II) ad-
thoroughly. sorption, probably attributed to the strong electrostatic attraction and
surface complexations between them. Notably, the Qmax of As(V) onto
MGO (13.9 mg/g) was significantly lower than that of Cd(II), mainly
3.2. Adsorption behaviors in single systems
due to the electrostatic repulsion between negative As(V) and MGO (at
pH 5.0) (Fig. S2). Nevertheless, the Qmax of As(V) onto MGO was still
3.2.1. Adsorption isotherms of TC, Cd(II) and As(V)
far greater than those of MCRG and MARG, and even higher than those
The adsorption isotherms of TC, Cd(II) and As(V) onto MGO, MCRG
of 3D Fe3O4-graphene macroscopic composite (0.471 mg/g) [59] and
and MARG were shown in Fig. 3. The parameters of the isotherms fitted
magnetite-reduced graphene oxide composite (5.83 mg/g) [33], sug-
by Langmuir and Freundlich models are listed in Table S1. Generally,
gesting that O-containing functional groups played an important role in
the isotherms of three magnetic graphene materials were fitted better
As adsorption and surface complexation [29]. Additionally, Fe3O4 NPs
by Freundlich (R2, 0.834–0.991) than by Langmuir model (R2,
had been confirmed to be excellent adsorbents for As [30,50]. Inter-
0.674–0.980), especially for the adsorption of TC onto MCRG and of As
estingly, though having the lowest component ratio of Fe (45.2% of
(V) onto MGO (Table S1). It was suggested that the adsorption of TC
weight, Table 1), MGO still exhibited a superior adsorption capacity
and As(V) onto MCRG and MARG tended toward multilayers rather
compared with MCRG and MARG, which might be attributed to more
than monolayers [14,29]. Furthermore, MGO displayed the strongest
available adsorption sites on MGO due to the homogeneous dispersion
capacities to adsorb TC, Cd(II) and As(V), followed by MCRG, and
of MGO.
MARG the smallest (Fig. 3).
The maximum adsorption capacity (Qmax) of TC onto MGO was

Fig. 2. FTIR spectra (a), XPS spectra of wide scan (b) and Fe 2p spectra (c) of MGO, MCRG and MARG, and XPS spectra of MGO (d), MCRG (e) and MARG (f) for C 1s.

1402
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

Fig. S3, TC had four species with their pKa values of 3.3, 7.5, and 9.5 for
the three hydrolyzable moieties, respectively, while the point of zero
charge (PZC) of TC was 5.5 [53]. Therefore, TC changed from posi-
tively-charged molecule to neutrality, and then electronegativity in-
creased with increasing pH, with the interaction between the TC and

Fig. 3. Adsorption isotherms of tetracycline (TC) (a), Cd(II) (b), As(V) (c), onto
MGO, MCRG and MARG. Adsorption conditions: pH = 5.0,
temperature = 25 ± 1 °C, equilibrium time = 24 h. Solid lines represent iso-
therms fitted by Freundlich model, and dashed lines represent isotherms fitted
by Langmuir model.

3.2.2. Effect of pH on adsorption of magnetic graphene materials


The adsorption of TC onto MGNs was pH-dependent (Fig. 4a). In-
creasing pH increased the adsorption of TC onto MGO, reached the Fig. 4. pH-dependent adsorptions of tetracycline (TC) (a), Cd(II) (b), As(V) (c),
maximum (216.7 mg/g) at pH 4.0 and then decreased. The zeta po- onto MGO, MCRG and MARG at a contact time of 24 h. The initial concentra-
tions of TC, Cd(II), As(V) were 50, 5, and 5 mg/L, respectively. The solid lines
tentials of MGO were always negative over the pH range of 2.0–12.0,
represent general trends of adsorption capacities of adsorbates by graphene
and the electronegativity increased with increasing pH. As shown in
materials as a function of pH.

1403
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

MGO being strongest at pH 4.0. For MCRG and MARG, the adsorption The FTIR spectra are shown in Fig. 5a (MGO) and Fig. S4a (MCRG
of TC changed inappreciably at pH 2.0–5.0, and then decreased slightly and MARG). In the case of MGO–TC (Fig. 5a), the intense band at
over the pH range of 5.0–8.0. The phenomenon was mainly because 3421 cm−1 could be assigned to the stretching vibration of the N–H
MCRG and MARG were positively-charged in the acidic range, and and/or O–H [34], as well as intermolecular hydrogen bonding with the
there was electrostatic repulsion between TC and the two nanomater-
ials. Under alkaline conditions, the dominant species of TC are TCH−
and TC2−, the adsorption capacity of TC significantly decreased due to
the strong electrostatic repulsion between TC and MGNs. Notably, there
were still considerable amounts of TC adsorbed on three nanomaterials
at pH > 8.0, while the electrostatic repulsion was very strong. Inter-
estingly, MCRG showed the most effective adsorption of TC in the pH
range of 8.0–12.0, which could be attributed to more defects on MCRG.
The ID/IG value of MCRG (1.24) was higher than those of MGO (1.02)
and MARG (1.03) (Fig. 1e), and MCRG consequently possessed more
active sites for TC.
As shown in Fig. 4b, the adsorptions of Cd(II) on all three MGNs
elevated with the increase in pH, which was in line with previous ob-
servations [6]. The phenomenon was not attributed to the formation of
Cd(OH)2 precipitates, since the initial pH values of Cd(II) were adjusted
to 2.0–8.0 and Cd(OH)2 precipitation were unlikely [5]. The MCRG and
MARG exhibited similar sorption trends for Cd(II) in the pH range of
2.0–5.0, but their adsorption capacities were lower than that of MGO.
The electrostatic repulsion between the positively-charged graphene
nanosheets (Fig. S2) and the cations suppressed the adsorption of Cd
(II). Notably, the adsorption of Cd(II) on MGO was markedly enhanced
with increasing pH because the electrostatic attraction between the
negatively-charged MGO and Cd(II) cations promoted the adsorption
process. The adsorption capacities were in accordance with the surface
oxygen contents. Therefore, the Cd(II) adsorption onto magnetic gra-
phene materials correlated with the surface oxygen content as well as
electrostatic interaction.
The adsorptions of As(V) onto three magnetic graphene nanoma-
terials decreased with increasing pH (Fig. 4c), consistent with previous
studies [29,33]. Notably, the trend was more prominent for the ad-
sorptions onto MGO than MCRG and MARG. It is known that As(V)
speciation can be significantly affected by pH, and is presented in the
negative ionic form under most pH conditions except when the pH is
less than 2.2 (pH < 2.2: H3AsO4, 2.2 < pH < 6.9: H2AsO4−,
6.9 < pH < 11.5: HAsO42−, pH > 11.5: AsO43−). Increasing pH in-
creased the negatively-charged sites and the competition of hydroxide
ions (OH–) with As(V), and hence decreased As(V) adsorption by the
three adsorbents. Similar adsorption capacities of As(V) were also ob-
served for MCRG and MARG over the pH range of 2.0–12.0. Interest-
ingly, the adsorption of As(V) on MGO was apparently higher than
those of MCRG and MARG in the pH range of 2.0–6.0, indicating that
electrostatic interactions were not the major driving force in the ad-
sorption process [29].

3.2.3. Adsorption mechanism of TC, Cd(II) and As(V)


The morphology and surface chemical compositions of MGO after
adsorption were monitored by SEM and XPS, respectively. Besides, the
FTIR spectra of MGO, MCRG and MARG before and after TC, Cd(II) and
As(V) adsorption were analyzed (Fig. 5 for MGO and Fig. S4 for MCRG
and MARG). By comparison, MGO showed excellent adsorption capa-
cities for TC, Cd(II) and As(V), which should be attributed to a larger
number of O-containing functional groups.
Microscopic morphology of MGO after 200 mg/L TC adsorption at
pH 5.0 was similar to that of the fresh sample (Figs. 6 and S1a), while
many dark particles were homogeneously distributed on the surface of
MGO adsorbed with 50 mg/L Cd(II) and 50 mg/L As(V), which should
be ascribed to the sorption of Cd(II) and As(V). The surface chemical
composition of the new particles was quantified using XPS (Fig. 7). The
appearance of characteristic Cd 3d peaks (405.5, 407.2 and 412.1 eV)
and As 3d peak (45.1 eV) in the spectra of (MGO + Cd) and
(MGO + As) implies the adsorption of Cd(II) and As(V) at homo- Fig. 5. FTIR spectra of MGO before and after adsorption of tetracycline (TC)
geneously-distributed active sites on MGO. (a), Cd(II) (b) and As(V) (c).

1404
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

Fig. 6. SEM images of the MGO after the treatment with 200 mg/L TC (a), 50 mg/L Cd(II) (b), 50 mg/L As(V) (c), 200 mg/L TC + 50 mg/L Cd(II) (d), 200 mg/L
TC + 50 mg/L As(V) (e), and 50 mg/L Cd(II) + 50 mg/L As(V) (f). Adsorption conditions: pH = 5.0, temperature = 25 ± 1 °C, equilibrium time = 24 h.

TC molecule. The peak observed at 1604 cm−1 could be assigned to the sharing complexes between As(V) and MNPs surface was the main ad-
stretching vibration of the N–H and/or C]C, which revealed the suc- sorption mechanism. The FTIR spectra revealed that the As(V) ad-
cessful adsorption of TC onto MGO. Besides, the N 1s peaks (407.8 and sorption via complexation with the peak at 45.1 eV being attributed to
399.7 eV) were observed in the XPS spectra (Fig. 7a) of MGO loaded As–O (Fig. 7c) [64]. Compared with original adsorbents, adsorption
with TC, which also agreed with the deduction. Thus, the FTIR and XPS with As(V) significantly increased the −OH peak (at approximately
spectra, highly consistent with the batch adsorption results, clarified 3405 and 1570 cm−1) of MGO as well as MCRG–As(V) and MARG–As
that H-bonds existed between N–H/–OH of TC molecules and O-con- (V). The stretching vibration of –OH was shifted from 1058 to
taining groups in MGO [53,60,61]. Moreover, the peak of carboxyl 1033 cm−1 for MCRG–As(V). In addition, the peak at around
(O–C]O) bending vibration of MGO was blue-shifted from 1396 to 1156 cm−1 could be assigned to the stretching vibration of the C–O
1352 cm−1 and its intensity significantly increased after reaction with and/or Fe–OH, and its intensity was also increased after As(V) ad-
TC, indicating the great contributions of edge carboxyl to TC adsorp- sorption, suggesting that –OH groups participated in the adsorption as
tion. As shown in Fig. S4a, the variations of characteristic peaks in H-bonding sites [30].
MCRG and MARG after TC adsorption were not obvious compared to
those of MGO-TC, which was primarily due to the elimination of most 3.3. Adsorption behaviors and binding mechanisms in binary systems
O-containing functional groups after reduction (Fig. 2) and π–π inter-
action becoming the main interaction between TC and MCRG and Since multiple types of pollutants often coexist in the actual en-
MARG [62]. Typically, the lengths of most H-bonds were shorter than vironments, the adsorption behaviors and interaction mechanisms of
3 Å, while the interval of two attached molecules via π–π interactions MGNs need to be elucidated when they are used in practical applica-
was larger [53]. The main adsorption forces for TC on MGO were H- tions. Thus, MGO, with the highest adsorption capacity, was selected to
bonding, π–π interaction, and n–π electron-donor–acceptor (EDA) ef- further study the adsorption behaviors of TC, Cd(II) and As(V) in binary
fect, which could explain the apparently higher adsorption of TC by systems (Fig. 8).
MGO compared to MCRG and MARG [36,53].
After Cd(II) adsorption onto the MGNs, the stretching vibration of 3.3.1. Mutual effects of TC and Cd(II) in a simultaneously added system
O–H was blue-shifted from around 3402 cm−1 (pure sorbate) to The presence of Cd(II) did not affect the TC adsorption onto MGO,
3418 cm−1 for MGO–Cd (Fig. 5b), and red-shifted from 3406 to nor did TC affect the Cd(II) adsorption (Fig. 8a-b). Zhao et al. [65]
3389 cm−1 for MCRG–Cd, and from 3417 to 3400 cm−1 for MARG–Cd showed a similar effect of Cd(II) on TC adsorption in the study of co-
(Fig. S4b). The behaviors were in accordance with the electrostatic adsorption of TC and Cd(II) onto soils. After co-adsorption of 200 mg/L
attraction between Cd(II) and MGO and electrostatic repulsion between TC and 50 mg/L Cd(II), numerous dark particles of Cd(II) complexes
Cd(II) and MCRG and MARG at pH 5.0. The intensity of O–C]O bands were anchored onto the surface of MGO (Fig. 6d); the amount was
appeared at 1395 cm−1 was much increased after Cd(II) adsorption, comparable to that of single adsorption of 50 mg/L Cd(II) (Fig. 6b). The
especially for MGO. A greater number of –OH/–COOH on MGO relative active sites of MGO were limited, when all the active sites were occu-
to MCRG and MARG conveyed the improved adsorption and dispersion pied, the extra reactants could not be adsorbed on the MGO surface, and
abilities of MGO [38]. In addition to the electrostatic interaction, the O- the co-adsorption would not be improved further [66]. The adsorption
containing functional groups and hydrogen of H2O could also form coefficients (Kd) for MGO at various concentrations of pollutants are
strong surface complexes with Cd(II), which played an important role in shown in Fig. S5. The adsorption affinity of TC onto MGO was much
the adsorption [5,14]. higher than that of Cd(II) at given equilibrium concentrations, in-
Similarly, adsorption with As(V) largely enhanced the O–C]O peak dicating that TC would be more prone to interactions with MGO both
intensity (∼1395 cm−1) for three adsorbents (Figs. 5c and S4c). Liu on the nanosheets and functional groups. The amino and hydroxyl
et al. [63] reported that the formation of bidentate binuclear corner- groups of TC were preferentially attracted to functional groups of MGO

1405
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

affinities for cations [43,61]. The electrostatic interaction between TC


and MGO was weak (at pH 5.0), and a number of edges –COO− on MGO
were available for the adsorption of Cd(II) via electrostatic interaction
and complexation. In the case of MGO covered by TC, the number of
available active sites was decreased, whereas TC molecules themselves
could make up some similar sites because of N–H/–OH and structure of
aromatic rings [53]. Due to the presence of the various compensation
mechanisms, MGO showed a great potential for remediation of the
single contamination of TC or Cd(II) as well as co-contamination of TC
and Cd(II).

3.3.2. Mutual effects of TC and As(V) in a simultaneously added system


In the simultaneous sorption (TC + As) system, increasing As(V)
concentration hardly affected TC adsorption when the initial con-
centrations of TC were in the range of 5–50 mg/L. However, the ad-
sorption of As(V) onto MGO was significantly suppressed at the TC
concentration range of 5–50 mg/L. The MGO preferred to adsorb TC to
As(V), which was in accordance with the Kd–Ce curve (Fig. S5). This
phenomenon was probably due to the formation of a more stable
complex of multiple H-bonds, since TC bears more O-containing groups
than that of As(V) [54]. In addition, TC exists as neutral molecules or
cations at pH 5.0, and the adsorbed negatively-charged As(V) could
slightly enhance the TC adsorption, which might be due to the anion–π
interactions resulting from the interaction of As(V) and benzene rings of
TC [67]. Generally, the anion–π interactions are weak, more targets of
TC molecules were available to interact with As(V) molecules which
were adsorbed onto MGO at 200 mg/L, leading to the enhancement of
TC adsorption. At 50 mg/L As(V), the presence of 200 mg/L TC de-
creased As(V) adsorption by 70% (Fig. 8d). The number of dark parti-
cles of As(V) complexes on the surface of MGO adsorbed with 200 mg/L
TC and 50 mg/L As(V) was obviously less than that of MGO adsorbed
with 50 mg/L As(V) (Fig. 6c, e). The literature showed that Fe3O4 NPs
could effectively adsorb TC [54] and As [30] via the formation of
complexes. The attached TC and As(V) shared similar adsorption sites
of H-bonds which were derived from O-containing functional groups
and Fe3O4 nanoparticles on MGO. However, the adsorption sites of
MGO were limited. The presence of TC on the surface of MGO could not
provide extra active sites for As(V) adsorption, leading to the compe-
tition for the available sorption sites. Consequently, the adsorption of
As(V) onto MGO was distinctly decreased in the presence of TC.

3.3.3. Mutual effects of Cd(II) and As(V) in a simultaneously added system


Fig. 8e-f presents the mutual effects of Cd(II) and As(V) in the
(Cd + As) system. Notably, both Cd(II) and As(V) adsorptions were
enhanced in the entire range of 5–50 mg/L, revealing a synergistic ef-
fect. At 50 mg/L Cd(II), the presence of 50 mg/L As(V) enhanced the
adsorption of Cd(II) onto MGO by 65%. In turn, the adsorption of As(V)
onto MGO was enhanced by 30%. Clearly, the surface of MGO was
almost covered by complexes of Cd(II) and As(V) (Fig. 6f) due to ex-
tremely great amounts of adsorbed Cd(II) and As(V). Similar synergic
effects were also observed for the co-adsorption of Cd(II) and orange G
dye (anion) using a magnetic graphene oxide nanocomposite adsorbent
[57]. Moreover, a similar phenomenon has been reported in Cd(II) and
PO43− co-adsorption onto GO at pH 6.0–10.0 [6]. Ren et al. [68] ob-
served that the adsorption of metal cations could be enhanced, in the
presence of anions through electrostatic attraction, formation of ternary
Fig. 7. XPS spectra of MGO after the treatment with 200 mg/L TC (a), 50 mg/L surface complexation, and/or surface precipitation. The Cd(II) adsorp-
Cd(II) (b), 50 mg/L As(V) (c) with the insets showing high resolution spectra of tion resulted in compression of the double electrostatic layer and neu-
N 1s, Cd 3d, and As 3d, respectively. tralization of negative charges on the MGO surface, which might
weaken the electronic repulsion between the negatively-charged GO
surfaces via H-bonding which was stronger than π–π interaction [53]. and H2AsO4− at pH 5.0 [6]. Typically, the ternary surface complexes
As previously reported, heavy metal ions could bridge organic com- are divided into type A (the metal is near the surface) and type B (the
pounds and functional groups on carbon nanomaterials [5,34]. The ligand is near the surface). In order to assess the contribution of surface
unoccupied benzene ring of TC could interact with Cd(II) via cation–π complexation and precipitation, the elemental compositions of MGO
interaction, since TC was a π-electron-rich donor and showed high after Cd(II) and As(V) sorption were determined using XPS and EDS.
Table S2 showed that the atomic ratio of Cd to As (0.20 for XPS, and

1406
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

Fig. 8. Adsorption capacity of tetracycline (TC) (a) and Cd(II) (b) in the presence of different initial concentrations of Cd(II) and TC; adsorption capacity of TC (c) and
As(V) (d) in the presence of different initial concentrations of As(V) and TC; adsorption capacity of Cd(II) (e) and As(V) (f) in the presence of different initial
concentrations of As(V) and Cd(II). Adsorption conditions: pH = 5.0, temperature = 25 ± 1 °C, equilibrium time = 24 h.

0.63 for EDS) was lower than that of Cd to P in Cd3(PO4)2 precipitate 4. Conclusions
(1.5) at pH 5.0 [6]. Moreover, no new minerals were detected in the
MGO loaded with Cd(II) and As(V) (Fig. S6), indicating that the co- This study presents a new strategy to prepare diverse types of
precipitation was not the dominant mechanism for the synergistic ef- magnetic graphene nanomaterials, MGO, MCRG, and MARG, with dif-
fect. Although both Cd(II) and As(V) adsorptions were enhanced in the ferent microstructures and properties. The three magnetic graphene
(Cd + As) system, the reduced adsorption of Cd(II) and increased ad- nanomaterials had a good magnetic separability and showed excellent
sorption of As(V) were observed using the surface-sensitive XPS tech- adsorption capacities for TC, Cd(II) and As(V), while MGO was shown
nique (only 2–5 nm probing depth) [69]. The results might indicate the to be the best adsorbent. High dispersibility and thin nanosheets con-
formation of a ternary surface complex with Cd(II) as the bridge mo- tributed to the superior adsorption capabilities of MGO. Moreover, 59%
lecule (i.e., MGO–Cd–As, marked as type A). On the basis of the Kd–Ce O-containing functional groups also promoted the strongest adsorption
curve (Fig. S5), Cd(II) had a stronger affinity to MGO than As(V) at 25 capabilities of MGO. In the binary systems of organic–inorganic com-
and 50 mg/L. Therefore, the formation of type A ternary surface com- bined pollutions, TC adsorption was hardly affected by the presence of
plexes was the primary mechanism of Cd(II) and As(V) adsorption in Cd(II) and As(V), nor was Cd(II) adsorption affected by TC, while the
the (Cd + As) system. For Cd(II) adsorption, the exposed As(V) in type adsorption of As(V) was significantly suppressed by TC. In the cation-
A ternary surface complexes could act as the new adsorption sites to oxyanion (Cd + As) binary systems, adsorption by MGO of both Cd(II)
adsorb the dissolved Cd(II) in solution continuously via electrostatic and As(V) were enhanced, which were attributed to the formation of
attraction and complexation at pH 5.0. type A ternary surface complexation (MGO–Cd–As) and electrostatic

1407
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

attraction. By regulating the O-containing functional groups on gra- [17] Y. Gao, Y. Li, L. Zhang, H. Huang, J. Hu, S.M. Shah, X. Su, Adsorption and removal
phene nanosheets, magnetic graphene nanomaterials had a great po- of tetracycline antibiotics from aqueous solution by graphene oxide, J. Colloid
Interf. Sci. 368 (2012) 540–546.
tential in the prevention of various complex pollution water systems. [18] E.E. Ghadim, F. Manouchehri, G. Soleimani, H. Hosseini, S. Kimiagar, S. Nafisi,
The findings provide new insights into designing novel adsorbents for Adsorption properties of tetracycline onto graphene oxide: equilibrium, kinetic and
efficient removal of pollutants in practical applications. Further studies thermodynamic studies, PLoS One 8 (2013).
[19] L. Jiang, Y. Liu, G. Zeng, S. Liu, X. Hu, L. Zhou, X. Tan, N. Liu, M. Li, J. Wen,
are needed on environmental effects and recycling of used magnetic Adsorption of estrogen contaminants (17 beta-estradiol and 17 alpha-ethynyles-
graphene nanomaterials. tradiol) by graphene nanosheets from water: effects of graphene characteristics and
solution chemistry, Chem. Eng. J. 339 (2018) 296–302.
[20] L. Jiang, Y. Liu, G. Zeng, F. Xiao, X. Hu, X. Hu, H. Wang, T. Li, L. Zhou, X. Tan,
Acknowledgement Removal of 17 beta-estradiol by few-layered graphene oxide nanosheets from
aqueous solutions: external influence and adsorption mechanism, Chem. Eng. J. 284
This work was financially supported by the National Natural Science (2016) 93–102.
[21] G. Abdul, X. Zhu, B. Chen, Structural characteristics of biochar-graphene nanosheet
Foundation of China (41721001), the Science and Technology Program
composites and their adsorption performance for phthalic acid esters, Chem. Eng. J.
of Zhejiang Province (2018C03028), and Agriculture Research System 319 (2017) 9–20.
of China (CARS-01-30). [22] D. Vilela, J. Parmar, Y.F. Zeng, Y.L. Zhao, S. Sanchez, Graphene-based microbots for
toxic heavy metal removal and recovery from water, Nano Lett. 16 (2016)
2860–2866.
Appendix A. Supplementary data [23] K. Chang, X. Li, Q. Liao, B. Hu, J. Hu, G. Sheng, W. Linghu, Y. Huang, A.M. Asiri,
K.A. Alamry, Molecular insights into the role of fulvic acid in cobalt sorption onto
Supplementary data: materials and methods, experimental details graphene oxide and reduced graphene oxide, Chem. Eng. J. 327 (2017) 320–327.
[24] Q. Liang, H. Luo, J. Geng, J. Chen, Facile one-pot preparation of nitrogen-doped
and adsorption models; regression parameters (Table S1), elemental ultra-light graphene oxide aerogel and its prominent adsorption performance of Cr
compositions after co-adsorption of Cd(II) and As(V) (Table S2); SEM (VI), Chem. Eng. J. 338 (2018) 62–71.
images (Fig. S1), zeta potentials (Fig. S2), TC properties (Fig. S3), FTIR [25] S.B. Wang, H.Q. Sun, H.M. Ang, M.O. Tade, Adsorptive remediation of environ-
mental pollutants using novel graphene-based nanomaterials, Chem. Eng. J. 226
spectra after adsorption (Fig. S4), variations of Kd (Fig. S5), XRD pat- (2013) 336–347.
terns after adsorption (Fig. S6), dissolved iron concentrations (Fig. S7). [26] V. Kumar, K. Kim, J.W. Park, J. Hong, S. Kumar, Graphene and its nanocomposites
Supplementary data to this article can be found online at https://doi. as a platform for environmental applications, Chem. Eng. J. 315 (2017) 210–232.
[27] B. Hu, C. Huang, X. Li, G. Sheng, H. Li, X. Ren, J. Ma, J. Wang, Y. Huang,
org/10.1016/j.cej.2018.10.138. Macroscopic and spectroscopic insights into the mutual interaction of graphene
oxide, Cu(II), and Mg/Al layered double hydroxides, Chem. Eng. J. 313 (2017)
References 527–534.
[28] J. Zhao, X. Cao, Z. Wang, Y. Dai, B. Xing, Mechanistic understanding toward the
toxicity of graphene-family materials to freshwater algae, Water Res. 111 (2017)
[1] S. Chowdhury, R. Balasubramanian, Recent advances in the use of graphene-family 18–27.
nanoadsorbents for removal of toxic pollutants from wastewater, Adv. Colloid [29] Y. Yoon, W.K. Park, T. Hwang, D.H. Yoon, W.S. Yang, J. Kang, Comparative eva-
Interface 204 (2014) 35–56. luation of magnetite-graphene oxide and magnetite-reduced graphene oxide com-
[2] J. Kang, X. Duan, L. Zhou, H. Sun, M.O. Tade, S. Wang, Carbocatalytic activation of posite for As(III) and As(V) removal, J. Hazard. Mater. 304 (2016) 196–204.
persulfate for removal of antibiotics in water solutions, Chem. Eng. J. 288 (2016) [30] C. Tian, J. Zhao, J. Zhang, S. Chu, Z. Dang, Z. Lin, B. Xing, Enhanced removal of
399–405. roxarsone by Fe3O4@3D graphene nanocomposites: synergistic adsorption and
[3] J. Kang, X. Duan, C. Wang, H. Sun, X. Tan, M.O. Tade, S. Wang, Nitrogen-doped mechanism, Environ. Sci.: Nano 4 (2017) 2134–2143.
bamboo-like carbon nanotubes with Ni encapsulation for persulfate activation to [31] Y. Yao, S. Miao, S. Liu, L.P. Ma, H. Sun, S. Wang, Synthesis, characterization, and
remove emerging contaminants with excellent catalytic stability, Chem. Eng. J. 332 adsorption properties of magnetic Fe3O4@graphene nanocomposite, Chem. Eng. J.
(2018) 398–408. 184 (2012) 326–332.
[4] J. Zhao, Z. Wang, J.C. White, B. Xing, Graphene in the aquatic environment: ad- [32] K. Chang, Y. Sun, F. Ye, X. Li, G. Sheng, D. Zhao, W. Linghu, H. Li, J. Liu,
sorption, dispersion, toxicity and transformation, Environ. Sci. Technol. 48 (2014) Macroscopic and molecular study of the sorption and co-sorption of graphene oxide
9995–10009. and Eu(III) onto layered double hydroxides, Chem. Eng. J. 325 (2017) 665–671.
[5] J. Wang, B. Chen, Adsorption and coadsorption of organic pollutants and a heavy [33] V. Chandra, J. Park, Y. Chun, J.W. Lee, I.C. Hwang, K.S. Kim, Water-dispersible
metal by graphene oxide and reduced graphene materials, Chem. Eng. J. 281 (2015) magnetite-reduced graphene oxide composites for arsenic removal, ACS Nano 4
379–388. (2010) 3979–3986.
[6] X. Ren, Q. Wu, H. Xu, D. Shao, X. Tan, W. Shi, C. Chen, J. Li, Z. Chai, T. Hayat, [34] A. Chen, C. Shang, J. Shao, Y. Lin, S. Luo, J. Zhang, H. Huang, M. Lei, Q. Zeng,
X. Wang, New insight into GO, cadmium(II), phosphate interaction and its role in Carbon disulfide-modified magnetic ion-imprinted chitosan-Fe(III): a novel ad-
GO colloidal behavior, Environ. Sci. Technol. 50 (2016) 9361–9369. sorbent for simultaneous removal of tetracycline and cadmium, Carbohyd. Polym.
[7] Z. Chen, H. Li, Y. Gao, A. Peng, Removal of phenanthrene and acenaphthene from 155 (2017) 19–27.
aqueous solution by enzyme-catalyzed phenol coupling reaction, Chem. Eng. J. 265 [35] J. Wu, D. Huang, X. Liu, J. Meng, C. Tang, J. Xu, Remediation of As(III) and Cd(II)
(2015) 27–33. co-contamination and its mechanism in aqueous systems by a novel calcium-based
[8] Lalhmunsiama, D. Tiwari, S. Lee, Surface-functionalized activated sericite for the magnetic biochar, J. Hazard. Mater. 348 (2018) 10–19.
simultaneous removal of cadmium and phenol from aqueous solutions: mechanistic [36] Q. Zhao, S. Zhang, X. Zhang, L. Lei, W. Ma, C. Ma, L. Song, J. Chen, B. Pan, B. Xing,
insights, Chem. Eng. J. 283 (2016) 1414–1423. Cation-Pi interaction: a key force for sorption of fluoroquinolone antibiotics on
[9] L. Ma, Q. Chen, J. Zhu, Y. Xi, H. He, R. Zhu, Q. Tao, G.A. Ayoko, Adsorption of pyrogenic carbonaceous materials, Environ. Sci. Technol. 51 (2017) 13659–13667.
phenol and Cu(II) onto cationic and zwitterionic surfactant modified montmor- [37] J. Wang, B. Chen, B. Xing, Wrinkles and folds of activated graphene nanosheets as
illonite in single and binary systems, Chem. Eng. J. 283 (2016) 880–888. fast and efficient adsorptive sites for hydrophobic organic contaminants, Environ.
[10] Z. Wu, H. Zhong, X. Yuan, H. Wang, L. Wang, X. Chen, G. Zeng, Y. Wu, Adsorptive Sci. Technol. 50 (2016) 3798–3808.
removal of methylene blue by rhamnolipid-functionalized graphene oxide from [38] X. Chen, B. Chen, Macroscopic and spectroscopic investigations of the adsorption of
wastewater, Water Res. 67 (2014) 330–344. nitroaromatic compounds on graphene oxide, reduced graphene oxide, and gra-
[11] J. Wang, Z. Chen, B. Chen, Adsorption of polycyclic aromatic hydrocarbons by phene nanosheets, Environ. Sci. Technol. 49 (2015) 6181–6189.
graphene and graphene oxide nanosheets, Environ. Sci. Technol. 48 (2014) [39] D. Li, X. Duan, H. Sun, J. Kang, H. Zhang, M.O. Tade, S. Wang, Facile synthesis of
4817–4825. nitrogen-doped graphene via low-temperature pyrolysis: the effects of precursors
[12] S.J. Yu, X.X. Wang, W. Yao, J. Wang, Y.F. Ji, Y.J. Ai, A. Alsaedi, T. Hayat, and annealing ambience on metal-free catalytic oxidation, Carbon 115 (2017)
X.K. Wang, Macroscopic, spectroscopic, and theoretical investigation for the in- 649–658.
teraction of phenol and naphthol on reduced graphene oxide, Environ. Sci. Technol. [40] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu,
51 (2017) 3278–3286. S.B.T. Nguyen, R.S. Ruoff, Synthesis of graphene-based nanosheets via chemical
[13] Y. Shen, X. Zhu, L. Zhu, B. Chen, Synergistic effects of 2D graphene oxide na- reduction of exfoliated graphite oxide, Carbon 45 (2007) 1558–1565.
nosheets and 1D carbon nanotubes in the constructed 3D carbon aerogel for high [41] Y. Sun, D. Shao, C. Chen, S. Yang, X. Wang, Highly efficient enrichment of radio-
performance pollutant removal, Chem. Eng. J. 314 (2017) 336–346. nuclides on graphene oxide-supported polyaniline, Environ. Sci. Technol. 47 (2013)
[14] Y. Shen, B. Chen, Sulfonated graphene nanosheets as a superb adsorbent for various 9904–9910.
environmental pollutants in water, Environ. Sci. Technol. 49 (2015) 7364–7372. [42] W. Jiang, J.T. Lv, L. Luo, K. Yang, Y.F. Lin, F.B. Hu, J. Zhang, S.Z. Zhang, Arsenate
[15] D. Robati, B. Mirza, M. Rajabi, O. Moradi, I. Tyagi, S. Agarwal, V.K. Gupta, Removal and cadmium co-adsorption and co-precipitation on goethite, J. Hazard. Mater. 262
of hazardous dyes-BR 12 and methyl orange using graphene oxide as an adsorbent (2013) 55–63.
from aqueous phase, Chem. Eng. J. 284 (2016) 687–697. [43] Y. Zhang, S.A. Boyd, B.J. Teppen, J.M. Tiedje, H. Li, Role of tetracycline speciation
[16] J. Xiao, W. Lv, Y. Song, Q. Zheng, Graphene/nanofiber aerogels: performance in the bioavailability to Escherichia coli for uptake and expression of antibiotic
regulation towards multiple applications in dye adsorption and oil/water separa- resistance, Environ. Sci. Technol. 48 (2014) 4893–4900.
tion, Chem. Eng. J. 338 (2018) 202–210. [44] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80

1408
D. Huang et al. Chemical Engineering Journal 358 (2019) 1399–1409

(1958) 1339. functionalized magnetic graphenes composite material and its application to re-
[45] J. Liu, Z. Zhao, G. Jiang, Coating Fe3O4 magnetic nanoparticles with humic acid for move Cr(VI), Pb(II), Hg(II), Cd(II) and Ni(II) from contaminated water, J. Hazard.
high efficient removal of heavy metals in water, Environ. Sci. Technol. 42 (2008) Mater. 278 (2014) 211–220.
6949–6954. [59] L. Guo, P. Ye, J. Wang, F. Fu, Z. Wu, Three-dimensional Fe3O4-graphene macro-
[46] N. Ye, Y. Xie, P. Shi, T. Gao, J. Ma, Synthesis of magnetite/graphene oxide/chitosan scopic composites for arsenic and arsenate removal, J. Hazard. Mater. 298 (2015)
composite and its application for protein adsorption, Mat. Sci. Eng. C-Mater. 45 28–35.
(2014) 8–14. [60] Z. Zhang, H. Liu, L. Wu, H. Lan, J. Qu, Preparation of amino-Fe(III) functionalized
[47] S. Zhan, D. Zhu, S. Ma, W. Yu, Y. Jia, Highly efficient removal of pathogenic bac- mesoporous silica for synergistic adsorption of tetracycline and copper,
teria with magnetic graphene composite, ACS Appl. Mater. Interfaces 7 (2015) Chemosphere 138 (2015) 625–632.
4290–4298. [61] M. Li, Y. Liu, G. Zeng, S. Liu, X. Hu, D. Shu, L. Jiang, X. Tan, X. Cai, Z. Yan,
[48] H.M. Sun, L.Y. Cao, L.H. Lu, Magnetite/reduced graphene oxide nanocomposites: Tetracycline absorbed onto nitrilotriacetic acid-functionalized magnetic graphene
one step solvothermal synthesis and use as a novel platform for removal of dye oxide: influencing factors and uptake mechanism, J. Colloid Interface Sci. 485
pollutants, Nano Res. 4 (2011) 550–562. (2017) 269–279.
[49] E. Roy, S. Patra, D. Kumar, R. Madhuri, P.K. Sharma, Multifunctional magnetic [62] B. Li, J. Ma, L. Zhou, Y. Qiu, Magnetic microsphere to remove tetracycline from
reduced graphene oxide dendrites: synthesis, characterization and their applica- water: adsorption, H2O2 oxidation and regeneration, Chem. Eng. J. 330 (2017)
tions, Biosensors Bioelectron. 68 (2015) 726–735. 191–201.
[50] S. Venkateswarlu, D. Lee, M. Yoon, Bioinspired 2D-carbon flakes and Fe3O4 nano- [63] C.H. Liu, Y.H. Chuang, T.Y. Chen, Y. Tian, H. Li, M.K. Wang, W. Zhang, Mechanism
particles composite for arsenite removal, ACS Appl. Mater. Interfaces 8 (2016) of arsenic adsorption on magnetite nanoparticles from water: thermodynamic and
23876–23885. spectroscopic studies, Environ. Sci. Technol. 49 (2015) 7726–7734.
[51] H. Peng, P. Gao, G. Chu, B. Pan, J. Peng, B. Xing, Enhanced adsorption of Cu(II) and [64] T. Wen, X. Wu, X. Tan, X. Wang, A. Xu, One-pot synthesis of water-swellable Mg-Al
Cd(II) by phosphoric acid-modified biochars, Environ. Pollut. 229 (2017) 846–853. layered double hydroxides and graphene oxide nanocomposites for efficient re-
[52] X.R. Xia, N.A. Monteiro-Riviere, S. Mathur, X. Song, L. Xiao, S.J. Oldenberg, moval of As(V) from aqueous solutions, ACS Appl. Mater. Interfaces 5 (2013)
B. Fadeel, J.E. Riviere, Mapping the surface adsorption forces of nanomaterials in 3304–3311.
biological systems, ACS Nano 5 (2011) 9074–9081. [65] Y. Zhao, Y. Tan, Y. Guo, X. Gu, X. Wang, Y. Zhang, Interactions of tetracycline with
[53] X. Zhang, J. Shen, N. Zhuo, Z. Tian, P. Xu, Z. Yang, W. Yang, Interactions between Cd (II), Cu (II) and Pb (II) and their cosorption behavior in soils, Environ. Pollut.
antibiotics and graphene-based materials in water: a comparative experimental and 180 (2013) 206–213.
theoretical investigation, ACS Appl. Mater. Interfaces 8 (2016) 24273–24280. [66] Y.C. Deng, L. Tang, G.M. Zeng, Z.J. Zhu, M. Yan, Y.Y. Zhou, J.J. Wang, Y.N. Liu,
[54] Q. Liu, L. Zhong, Q. Zhao, C. Frear, Y. Zheng, Synthesis of Fe3O4/polyacrylonitrile J.J. Wang, Insight into highly efficient simultaneous photocatalytic removal of Cr
composite electrospun nanofiber mat for effective adsorption of tetracycline, ACS (VI) and 2,4-diclorophenol under visible light irradiation by phosphorus doped
Appl. Mater. Interfaces 7 (2015) 14573–14583. porous ultrathin g-C3N4 nanosheets from aqueous media: performance and reaction
[55] Y. Lin, S. Xu, L. Jia, Fast and highly efficient tetracyclines removal from environ- mechanism, Appl. Catal. B-Environ. 203 (2017) 343–354.
mental waters by graphene oxide functionalized magnetic particles, Chem. Eng. J. [67] G. Shi, Y. Ding, H. Fang, Unexpectedly strong anion-π interactions on the graphene
225 (2013) 679–685. flakes, J. Comput. Chem. 33 (2012) 1328.
[56] Y. Zhang, B. Chen, L. Zhang, J. Huang, F. Chen, Z. Yang, J. Yao, Z. Zhang, [68] X. Ren, S. Yang, X. Tan, C. Chen, G. Sheng, X. Wang, Mutual effects of copper and
Controlled assembly of Fe3O4 magnetic nanoparticles on graphene oxide, Nanoscale phosphate on their interaction with gamma-Al2O3: combined batch macroscopic
3 (2011) 1446. experiments with DFT calculations, J. Hazard. Mater. 237 (2012) 199–208.
[57] J. Deng, X. Zhang, G. Zeng, J. Gong, Q. Niu, J. Liang, Simultaneous removal of Cd [69] J. Li, Q. Fan, Y. Wu, X. Wang, C. Chen, Z. Tang, X. Wang, Magnetic polydopamine
(II) and ionic dyes from aqueous solution using magnetic graphene oxide nano- decorated with Mg-Al LDH nanoflakes as a novel bio-based adsorbent for simulta-
composite as an adsorbent, Chem. Eng. J. 226 (2013) 189–200. neous removal of potentially toxic metals and anionic dyes, J. Mater. Chem. A 4
[58] X. Guo, B. Du, Q. Wei, J. Yang, L. Hu, L. Yan, W. Xu, Synthesis of amino (2016) 1737–1746.

1409

Вам также может понравиться