You are on page 1of 10

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2018, 3, 13045−13054 http://pubs.acs.org/journal/acsodf

Graphene Oxide−Chitosan Composite Material for Treatment of a


Model Dye Effluent
Mina Sabzevari,† Duncan E. Cree,† and Lee D. Wilson*,‡

Department of Mechanical Engineering, University of Saskatchewan, 57 Campus Drive, Saskatoon, Saskatchewan S7N 5A9, Canada

Department of Chemistry, University of Saskatchewan, 110 Science Place, Saskatoon, Saskatchewan S7N 5C9, Canada
*
S Supporting Information

ABSTRACT: Graphene oxide (GO) was cross-linked with chitosan


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

to yield a composite (GO-LCTS) with variable morphology,


enhanced surface area, and notably high methylene blue (MB)
adsorption capacity. The materials were structurally characterized
using thermogravimetric analysis and spectroscopic methods (X-ray
Downloaded via 5.188.219.231 on October 11, 2018 at 19:15:46 (UTC).

diffraction, Fourier transform infrared spectroscopy, Raman spectros-


copy, and 13C solid-state NMR) to support that cross-linking occurs
between the amine groups of chitosan and the −COOH groups of
GO. Equilibrium swelling studies provide support for the enhanced
structural stability of GO-cross-linked materials over the synthetic
precursors. Scanning electron microscopy studies reveal the enhanced
surface area and variable morphology of the cross-linked GO
materials, along with equilibrium and kinetic uptake results with
MB dye in aqueous media, revealing greater uptake of GO-LCTS
composites over pristine GO. The monolayer uptake capacity (Qm; mg g−1) with MB reveals twofold variation for Qm, where
GO-LCTS (402.6 mg g−1) > GO (286.9 mg g−1). The kinetic uptake profiles of MB follow a pseudo-second-order trend, where
the GO composite shows more rapid uptake over GO. This study reveals that the sorption properties of GO are markedly
improved upon formation of a GO−chitosan composite. The facile cross-linking strategy of GO reveals that its physicochemical
properties are tunable and versatile for a wider field of application for contaminant removal, especially over multiple
adsorption−desorption cycles when compared against pristine GO in its highly dispersed nanoparticle form.

1. INTRODUCTION has a low aqueous solubility, GO is highly dispersible in water


The development of advanced materials for the controlled and organic solvents because of the presence of polar
removal of dyes and organic contaminants from effluent functional groups and high surface area.9 The highly negative
originating from printing, food, textile, paper, and pharma- charge density of GO in aqueous solutions provides effective
ceutical industries1,2 is an active area of research. Many dye adsorption sites for cationic dyes,6 such as methylene blue
species are known to be carcinogenic and can affect water (MB) and serves as a material for gas capture (e.g., N2 and
quality, as shown by undesirable human and ecosystem health CO2) applications.10,11 Although GO is a promising material as
effects.3 Adsorption-based removal is an efficient process for an adsorbent, efficient recovery is problematic because of its
the capture of dyes and contaminants from wastewater tendency of forming stable colloids that hinder phase
systems. Solid phase adsorbents represent an effective separation. However, the use of polymers to immobilize GO
remediation strategy because of their technical simplicity, offers an opportunity to produce composite materials with
high efficiency, low cost, and potential reuse of the unique properties for solid phase separation and adsorption.
adsorbent.4,5 Therefore, continued effort is required for the Fabrication of framework structures from individual GO
development of an efficient, novel, and cost-effective solid sheets was studied using physical and chemical interactions to
adsorbent material. yield GO self-assembled films, hydrogels, and aerogels.12,13
Recently, graphene- and graphene oxide (GO)-based However, produced GO frameworks have limited applications
materials have been studied for applications in adsorption in wastewater treatment because of scalability as a result of
and water treatment.6,7 In particular, GO has emerged over the repulsive hydration forces between GO layers that result in
past decade as a next-generation material for wastewater electrostatic separation.14 Therefore, the development of stable
treatment because of its low-cost production, large surface interconnected GO frameworks is one strategy for the
area, and strong interaction with a wide range of anionic,
cationic, or neutral dyes in aqueous media.8,9 Graphene is Received: August 2, 2018
comprised of sp2-hybridized carbon atoms arranged in a two- Accepted: September 21, 2018
dimensional honeycomb lattice. In contrast to graphene, which Published: October 11, 2018

© 2018 American Chemical Society 13045 DOI: 10.1021/acsomega.8b01871


ACS Omega 2018, 3, 13045−13054
ACS Omega Article

Figure 1. Synthetic procedure for cross-linked GO-LCTS composite materials.

successful application of these materials in wastewater drying in a vacuum oven at 60 °C for 12 h to remove
treatment. Previous works reveal that cross-linking and other impurities.
forms of synthetic modification of GO sheets can improve the 2.2. Synthesis of GO. GO was synthesized from graphite
sorption properties, as shown by the variation in methylene flakes using the modified Hummer’s method.24 Briefly, 100 mL
blue uptake (cf. Table S1, Supporting Information).15,16 The concentrated H2SO4 was added into a 500 mL flask filled with
resulting material would offer higher sorption capacity 4 g of graphite in an ice bath, followed by the addition of 2 g of
compared to pure GO because of the increased available NaNO3, and stirred for 4 h. Then, 12 g of KMnO4 was
adsorption sites as well as improved stability and mechanical gradually added while the mixture was stirred for 2 h. The ice
structure for producing thin films.17,18 bath was removed, and the system was heated at 35 °C for
GO framework structures can be achieved by cross-linking another 30 min. Subsequently, 240 mL of distilled water was
GO sheets through noncovalent or covalent bonds by using slowly added to the system, which caused a temperature rise to
biopolymers as cross-linkers.15,19 Chitosan (CTS) is among a 90 °C, and continued to stir for another 30 min. Then, 160 mL
group of abundant renewable biopolymers20 that has been of water and 30 v/v % H2O2 were added to terminate the
used as an adsorbent for removal of organic dyes from aqueous reaction. The solution was stirred overnight and purified using
solution because of the presence of active adsorption sites multiple washings with Millipore water, HCl (30%), and
(hydroxyl and amino groups).21 CTS can undergo an ethanol until it reached neutrality (pH 7). After multiple
amidation reaction with the carboxyl groups of GO to form washings, the solid GO was vacuum-dried at 40 °C to obtain
a homogenous and well-dispersed GO composite.22,23 On the dried GO powder.
basis of the foregoing considerations, a key objective of this 2.3. Preparation of GO−CTS Cross-Linked Composite.
study was to develop a facile method for the preparation of a The composite materials (GO-LCTS) were prepared by
GO composite by cross-linking GO with CTS to obtain a making the GO solution with a concentration of 3 mg/mL.
material with improved physicochemical properties for LCTS solution was produced by dissolving 5 g of LCTS in 500
sorption-based applications. To this end, spectroscopic mL of 1 v/v % glacial acetic acid with stirring. The resulting
methods [Fourier transform infrared spectroscopy (FTIR), LCTS solution (1 v/v %) was added drop-wise to the GO
Raman spectroscopy, and 13C nuclear magnetic resonance solution with continuous stirring for about 4 h. The mixture
(NMR)] were used to characterize the structure and was neutralized to pH ≈ 7 using 1 M NaOH followed by
composition of cross-linked GO composites. Also, thermogra- stirring for 12 h. Herein, several cross-linked GO samples were
vimetric analysis (TGA) and scanning electron microscopy prepared with variable precursor weight ratios. The results
(SEM) were used to analyze the thermal stability and surface reported herein are focused on GO-LCTS with 1:0.3 (w/w)
morphology of the GO composites. The adsorption properties ratio because this condition affords a minimum amount of
were evaluated under equilibrium and kinetic conditions using cross-linker level to stabilize the GO-LCTS composite. The
MB in aqueous solution. It will be shown that this study solution was washed with Millipore water followed by pouring
contributes significantly to the field of sustainable GO onto a glass surface and dried at ambient conditions for 48 h to
composites through the development of a suitable adsorbent obtain the GO-LCTS films with an average thickness ranging
for contaminant removal with tunable properties that exceed from 20 to 60 μm. An outline of the experimental sequence is
those of GO alone, and also can be reused for multiple cycles
shown in Figure 1.
of adsorption−desorption. Also, this study advances the field of
2.4. Characterization. 2.4.1. Surface Charge (ζ Potential)
GO composites by use of a facile cross-linking methodology
Measurement. The ζ potential of the GO solutions (0.01 w/v
that is low cost and ecofriendly, where the structure of the
%) were measured at various pH values (4−12) using a
composites are supported by several complementary methods.
Zetasizer Nano ZS (Nano ZS90, Malvern, UK), before and
after cross-linking. All samples were diluted to 0.01 w/v % and
2. EXPERIMENTAL SECTION measured in Milli-Q water at different pH.
2.1. Materials. Low molecular weight CTS (LCTS) (Mw = 2.4.2. Scanning Electron Microscopy. The surface
950 000 Da, 75−85% deacetylation), sodium nitrite (NaNO3), morphology and surface topography of non-cross-linked and
potassium permanganate (KMnO4), sulfuric acid (98%, ACS cross-linked composite materials were studied using SEM
grade), MB (high purity, biological stain), and hydrogen (Hitachi model SU8010). SEM images of samples were
peroxide (30 v/v %) used in this study were obtained from collected under an accelerating voltage (5 kV).
Sigma-Aldrich Canada Ltd. HPLC-grade methanol was 2.4.3. FTIR Spectroscopy. A Bio-Rad FTS-40 IR spec-
obtained from Fisher Scientific, NJ, USA. Filter paper trophotometer was used to obtain the IR spectra of the
(Ahlstrom grade 613, 7.5 cm) and graphite flakes, natural, composite materials. The sample powder was mixed with pure
325 mesh, 99.8% (metals basis) were obtained from Alfa Aesar spectroscopic-grade KBr (weight ratio: 1:10). The FTIR
Thermo Fisher Scientific and further purified by Soxhlet spectra were obtained in reflectance mode with a resolution
extraction for 24 h using HPLC-grade methanol, followed by of 4 cm−1 over a spectral range of 400−4000 cm−1.
13046 DOI: 10.1021/acsomega.8b01871
ACS Omega 2018, 3, 13045−13054
ACS Omega Article

Figure 2. ζ Potential for GO at variable pH and GO-LCTS composite. The right-hand panel corresponds to solution conditions at ambient pH and
295 K.

2.4.4. Raman Spectroscopy. The Raman spectra of pure 2.5. Sorption Studies of GO-LCTS Composite.
GO and cross-linked GO-LCTS samples were obtained to 2.5.1. MB Sorption of GO-LCTS Composite. The equilibrium
monitor changes in the degree of order−disorder of the GO MB uptake capacity of GO and GO-LCTS composites was
structure after cross-linking using a Renishaw Raman evaluated by batch mode using fixed amounts of adsorbent in
spectrophotometer equipped with an inVia reflex optical sealed glass vials. The adsorption isotherms obtained by such
microscope and a laser excitation wavelength of 514.5 nm batch mode conditions used a dosage of 5 mg adsorbent with 7
over the spectral range 3700−5000 cm−1. mL MB solution (pH = 7) at variable initial MB (100−1000
2.4.5. 13C Solid-State NMR. The 13C solid NMR spectra of μM). Samples were mixed on a horizontal shaker (SCILOGEX
GO and GO-LCTS composite material were obtained using a SK-O330-Pro) for 24 h to ensure equilibrium. A 1 mM stock
Bruker AVANCE III HD spectrometer equipped with a 4 mm solution of MB in water was prepared, and other solutions
DOTY cross-polarization with magic angle spinning (CP/ were obtained by appropriate dilution. Absorption measure-
MAS). Spectra were acquired using CP/MAS using a solid ments were carried out at λ = 664 nm, and a linear calibration
probe operating at 125.77 MHz (1H spectral frequency at curve for MB absorbance at variable concentration was
500.23 MHz). The 13C CP/MAS NMR spectra were acquired obtained with a slope of 0.0615 = Abs664nm/[MB]. After 24
using a contact time of 0.75 ms, MAS at 10 kHz, and 1H 90° h shaking to achieve equilibrium, the sample powders were
pulse of 3.5 μs, with a ramp pulse on the 1H channel. All separated from the solution by centrifugation and the residual
spectral data were recorded using 71 kHz SPINAL-64 (MB) was measured using UV−vis spectroscopy (Varian Cary
decoupling during acquisition with external reference to 100 Scan UV−vis spectrophotometer). The equilibrium uptake
adamantane at δ = 38.48 ppm. of MB was calculated using eq 2.
2.4.6. X-ray Diffraction. The X-ray diffraction (XRD) C0 − Ce
patterns of GO and cross-linked composite materials were Qe = ×m
V (2)
obtained using a PANalytical Empyrean powder X-ray
−1
diffractometer equipped with a Co source and X’Celerator where Qe (mg g ) is the MB adsorbed per unit mass of
detector. The powdered samples were placed in a horizontal adsorbent, C0 (mM) is the initial dye concentration, Ce (mM)
mode and mounted for the test. The PXRD patterns were is the residual amount of MB, V (L) is the volume of the MB
measured in continuous mode over a range of 5−60° 2θ, with solution, and m (g) is the weight of the adsorbent.
a scan speed of 335 ms/step (3°/min). The GO composite was tested for its regeneration properties
2.4.7. Thermogravimetric Analysis. Thermal stability and over several adsorption−desorption cycles with MB, as
decomposition temperature of the materials were measured described in further detail in the Supporting Information.
using a TA Instruments Q50000IR TGA system, which was 2.5.2. Kinetic Uptake Studies of MB. Kinetic adsorption
operated from 23 to 500 °C with a heating rate of 5 °C min−1 studies of GO samples cross-linked with LCTS toward MB as
under a nitrogen atmosphere. the dye probe in aqueous solution to estimate the sorption
2.4.8. Equilibrium Swelling Properties. The swelling capacity of GO before and after cross-linking with LCTS.
properties and water uptake of the GO-LCTS composites Kinetic isotherm profiles were obtained by plotting Qt versus t
and the precursors (graphite, LCTS, and GO) were evaluated using a one-pot method described elsewhere to account for
by immersing 50 mg of the material in 12 mL Millipore water adsorption processes of nanomaterial sorbents.26 In brief, ca.
100 mg of a powdered sample was placed into a folded filter
as the solvent and equilibrated in a horizontal shaker for ∼24
paper with both ends sealed before adding to the MB solution,
h.25 The weight of swollen samples (Ws) was determined by
where it should be noted that the adsorption of MB by the
weighing hydrated samples after removing excess surface water
filter paper was deemed to be negligible overall.27 The sealed
with filter paper. By drying the hydrated samples in an oven at
filter paper was immersed in a fixed volume (250 mL) of an
60 °C, the dry weight (Wd) was obtained and the water aqueous 5 μM MB solution. Aliquots of MB solution were
swelling (Sw) was calculated using eq 1 as below pipetted (3 mL) at variable time intervals and further
Ws − Wd quantified via UV−vis spectrophotometry (Varian Cary 100
Sw (%) = × 100 Scan UV−vis spectrophotometer). It is worth mentioning that
Wd (1) the filter paper used in our experiment provides reliable results
13047 DOI: 10.1021/acsomega.8b01871
ACS Omega 2018, 3, 13045−13054
ACS Omega Article

Figure 3. SEM micrographs of carbonaceous materials (GO, graphite) and the GO-composite material: (a) graphite, (b) GO, and (c) GO-LCTS.

Figure 4. FTIR (a) and Raman spectra (b) of the precursors and GO-LCTS.

because of the fast diffusion processes and negligible groups (such as −OH and COOH) on the GO surface, which
adsorption throughout the experiment.27 The MB adsorption form during the oxidation of graphite. Such groups become
capacity at variable times for the GO and GO-LCTS materials ionized at higher pH and result in enhanced negative charge.7
was calculated using eq 3 Therefore, it is expected that GO and its modified forms will
C0 − Ct display favorable affinity with cationic species, especially over
Qt = ×V the pH range 6−12. GO forms a stable suspension in aqueous
m (3)
solution in the pH range 6−12 because its ζ potential is below
where C0 and Ct are the concentration values of MB initially −30 mV. Accordingly, particles with ζ potential lower than
and at variable time (t), respectively; V is the solution volume; −30 mV are strong enough to maintain a stable colloidal
and m is the weight of the adsorbent. The adsorption kinetics solution because of repulsive forces between them.29 There-
can be described by the pseudo-first-order and the pseudo- fore, it would be appropriate to run the cross-linking
second-order (PSO) model to evaluate parameters of the experiment at any pH above 6. The ζ-potential measurement
isotherm. In this study, the best-fit results were obtained by of GO-LCTS composite further supports the formation of the
PSO model as given in eq 4 GO-LCTS composite. The ζ potential of GO was −37.5 mV
Q e 2 k 2t prior to cross-linking with LCTS, whereas the surface charge of
Qt = the composite was −2.5 mV.
1 + k2 t Q e (4) 3.2. SEM Results. Figure 3 shows SEM micrographs for
−1 GO and its cross-linked form (GO-LCTS). The SEM images
where Qt is the amount of solute adsorbed at time t (mg g ),
depict the graphite starting material and the synthesized GO,
Qe is the amount of solute adsorbed at a pseudo-equilibrium
(mg g−1) condition, and k2 is the rate constant according to the further revealing the layered structure of these materials. By
PSO adsorption model. The kinetic sorption parameters that comparison, the cross-linked GO composites have wrinkled
were deduced from the PSO model in this study provide a edges with irregular shapes of dense interconnected layers. The
comparison of sorption characteristics of GO and its cross- micrographs reveal that cross-linking of GO alters its regular
linked form with LCTS. layered morphology and surface roughness according to the
cross-linker type. The SEM results indicate that the composites
3. RESULTS AND DISCUSSION possess higher surface roughness and porosity when compared
3.1. Surface Charge Material Characterization. The with pristine graphite or GO. These variations in morphology
point of zero charge of GO at different pH conditions is an and textural properties provide support for unique product
important indicator of the nature of adsorptive interactions morphology that occurs upon formation of a cross-linked
because the surface charge of the adsorbent phase (GO and composite between GO and LCTS.
GO-LCTS composite) is likely to affect the adsorption 3.3. FTIR and Raman Spectral Characterization. IR
capacity of MB. As shown in Figure 2, the ζ potential of the spectroscopy was used to monitor the changes in surface
GO materials was highly negative, and this charge decreases functional groups upon the formation of GO from graphite and
continuously with increasing pH from 4 to 10, and is the structural characterization of the GO-LCTS composite.
consistent with previous literature.28 This negative charge is The FTIR spectra of the cross-linked composite material and
primarily because of the introduction of various functional its precursors (GO and LCTS) are shown in Figure 4a.
13048 DOI: 10.1021/acsomega.8b01871
ACS Omega 2018, 3, 13045−13054
ACS Omega Article

First, the GO preparation was characterized by identifying shift variation for the C−OH groups (∼70 ppm) of GO, along
the characteristic IR bands for −OH stretching at ∼3200− with new spectral signatures ca. 30 and 100 ppm assigned to
3700 cm−1, carbonyl groups (CO) at ca. 1680−1730 cm−1, the glucosamine and acetylated forms of the co-monomers of
CC groups at ca. 1550−1650 cm−1, and epoxy groups at ca. LCTS. The upfield signatures (ca. 20−80 ppm) for GO relate
1230−1350 cm−1. For the IR spectrum of LCTS, characteristic to the alkyl and alkenic groups. In addition, the appearance of
bands are centered at 1152 and 895 cm−1, corresponding to downfield signatures at ca. 170−180 ppm relate to the
signatures assigned to a glucopyranose ring unit, whereas the presence of carbonyl (acetyl/amide groups of LCTS and
CO stretching vibration of amide I (NHCO) and amide II −COOH groups of GO). The two 13C carbonyl signatures for
(N−H) bending of NH2 is observed at 1650 and 1590 cm−1, the GO-LCTS composite is related to the presence of at least
respectively. The presence of the band of both precursors (GO two carbonyl groups, one for the acetyl or amide groups. The
and LCTS) is supported by the similar spectral features of NMR results support that cross-linking has occurred, in
cross-linked composite materials. A noteworthy observation for agreement with FTIR bands at 1655 and 1595 cm−1. The
the cross-linked GO composites includes the absence of some above results are in general agreement with 13C NMR results
bands or changes of intensity when compared to similar IR reported by Mahaninia and Wilson35 for unmodified and cross-
bands for the unmodified GO. For instance, the LCTS linked CTS with different bifunctional linker systems. The
glucopyranose band at 895 cm−1 was not observed in GO- results presented herein provide an example of the first
LCTS. The formation of an amide linkage between GO and reported solid NMR spectral results for this type of GO
LCTS can be demonstrated by the absence of GO peaks at composite material (Figure 5).
1730 cm−1, attributed to CO in the −COOH moiety of GO, 3.5. XRD Structural Characterization. XRD character-
where a greater IR intensity of the amide II band at 1595 cm−1 ization (Figure 6a) of GO (after oxidation of graphite) and
is observed for cross-linked GO composites. This provides GO-LCTS (after cross-linking of GO with LCTS) was used to
support for the formation of a linkage between GO and LCTS evaluate the structure of each of the samples, as well as monitor
as the linker, in agreement with previous reports.30 changes in the interlayer spacing of GO after cross-linking. The
The Raman spectrum in Figure 4b for GO consists mainly of XRD pattern of graphite shows a sharp characteristic peak at
D- and G-band signatures at 1350 and 1580 cm−1. The G band 2θ = 31.5°. After the introduction of oxygen functional groups
is characteristic of sp2-hybridized carbon networks of graphene to the structure of graphite, the graphitic peak shifts to 2θ =
sheets, whereas the D-band results because of structural 12.5°, related to the interlayer distance of 0.72 nm between
imperfections created by the attachment of oxygen-based GO sheets.36
functional groups on the carbon basal plane and its partially Also, the XRD pattern of pure LCTS powder displays two
disordered structure.31 In contrast to GO, graphite has an broad peaks at 2θ = 9.5° and 22.5°, related to the amorphous
almost insignificant D band because of its highly crystalline hydrated and anhydrous structure of LCTS, respectively.37 In
structure. However, the intensity of this Raman band comparison to the GO and LCTS patterns, cross-linked GO-
increased, and it became broader upon cross-linking. Various LCTS shows shifting of the sharp XRD line of GO at 12.5° to a
studies report that the removal of functional groups from the lower 2θ (10.2°) value with an absence of broad LCTS peaks.
GO structure results in a greater D/G intensity ratio because of These changes indicate the exfoliation of LCTS into the GO
the onset of defects after reduction and also in self-assembly of sheets38 and an increase in the interplanar distance (0.72−0.87
the GO composites.32,33 In this study, cross-linking of GO led nm) between GO sheets by the introduction of LCTS.
to changes in the D/G intensity ratio, from 0.98 in GO to 1.26 Although the broadening peak in the XRD pattern of GO-
in the cross-linked GO composite. This change in the D/G LCTS suggests slightly lower crystallinity of the GO-LCTS,
signal intensity ratio can be attributed to disruption of the GO shifting of the peak to lower wavenumbers indicates an
layered structure upon cross-linking with CTS. increase in interlayer distance of GO sheets by 0.15 nm upon
3.4. 13C Solid NMR Spectral Characterization. The cross-linking with LCTS. This indicates that LCTS chains are
spectra for pristine GO (Figure 5) present 13C NMR lines for well introduced and strongly bonded to the oxygen functional
epoxide groups (∼60 ppm), −C−OH groups (∼70 ppm), and groups of GO, while maintaining the stacked structure of GO
graphenic group (∼130 ppm), in agreement with a previous sheets in the material.23,39 The higher crystallinity of the
report.34 The spectral differences between the cross-linked GO graphite and GO sheets agrees with the layered cross-sectional
materials with GO relate to the broadening and 13C spectral morphology observed in SEM images (Figure 3a,b), as
compared to the slightly separated layered cross-section
morphology of the GO-LCTS composite (Figure 3c). The
results for GO and GO-LCTS composites suggest that cross-
linking of the GO sheets is a critical technique to tune GO-
based material properties, including morphology, degree of
crystallinity, and interlayer distance between corresponding
GO sheets.
3.6. Thermal Stability Properties. The TGA results for
the cross-linked GO composite are shown in Figure 6b, where
a sharp thermal decomposition occurs starting near 120 °C for
GO, which results in a total weight loss event for the material.
The cross-linked GO-LCTS material does not show a
comparable thermal event, however; two key thermal events
appear at ca. 155 and 250 °C with a weight loss of 15.6 and
Figure 5. Normalized 13C solid NMR spectra of GO and GO-based 35.4%, respectively. These thermal events for the GO
cross-linked composites. composite relate to loss of adsorbed or free water and
13049 DOI: 10.1021/acsomega.8b01871
ACS Omega 2018, 3, 13045−13054
ACS Omega Article

Figure 6. XRD patterns (a) and TGA curves (b) of GO, precursors (graphite, LCTS), and GO-LCTS composite.

decomposition of GO oxygen functionalities and the LCTS composite that may alter water infiltration and hydration
backbone that yields degradation of the GO-LCTS framework properties of the system. The denser arrangement of the GO-
structure. The TGA results confirm that the GO-LCTS LCTS 3D framework (according to SEM results) is expected
structure has greater stability over GO as shown by its stability to contribute steric hydration effects that differ from the 2D
upon heating to the upper-temperature limit (500 °C). nature of GO because of the greater abundance and
3.7. Solvent Swelling and Dye Sorption Properties.
accessibility of polar functional groups in pristine GO.
Equilibrium swelling properties of GO-LCTS composite and
the precursors are listed in Table 1. The results indicate that Differences in the hydrophilic character and network structure
of GO-LCTS and GO provide an account for the water
Table 1. Water Swelling Properties of the GO-LCTS swelling properties of these materials.
Composite and Precursors (Graphite and GO) 3.7.1. Equilibrium Dye Uptake of MB. The MB sorption
behavior of the GO and GO-LCTS adsorbents in aqueous
material swelling (%)
solution was analyzed by fitting the adsorption isotherm results
graphite 510 for uptake of MB at variable concentration by various models
LCTS 345
(Langmuir, Freundlich, and Sips isotherms). The Langmuir
GO 9745
GO-LCTS 6500
model (eq 5) describes monolayer sorption for the adsorbate−
adsorbent system. By comparison, Freundlich and Sips models
describe the relationship for adsorbent−adsorbate systems
graphite and LCTS have the least swelling, whereas GO and with variable adsorption sites (eqs 6 and 7). The Sips model
the cross-linked GO composite (GO-LCTS) display greater
accounts for Langmuir and Freundlich behavior, where the
swelling in water. The GO-LCTS composite has reduced
swelling relative to pristine GO that may relate to the fewer maximum adsorption capacity of the adsorbate (Qm) onto the
ionic sites (−COOH) as a result of amide bond formation adsorbent surface can be estimated.
because of cross-linking, in agreement with the IR and NMR
K1Ce
spectral results above. Qe = Qm
The greater swelling of GO is attributed to its highly 1 + K1Ce (5)
hydrophilic nature imparted by its polar functional groups
(e.g., −OH, −COOH, etc.). By contrast, cross-linking of GO
with LCTS decreases the overall hydrophilic character of the Q e = K f Ce1/ nf (6)

Figure 7. (A) Isotherm sorption results for MB with GO and GO-LCTS sorbents, and (B) decolorization of MB before and after the sorption
process with GO-LCTS.

13050 DOI: 10.1021/acsomega.8b01871


ACS Omega 2018, 3, 13045−13054
ACS Omega Article

Q m K sCe ns microporous structure that is critical for adsorption applica-


Qe = ns tions.23 The interconnected network structure and higher
1 + K sCe (7)
interlayer distance between GO sheets in the GO-LCTS
In eq 5, Qe is the equilibrium uptake of dye per unit mass of composite (as shown in XRD result Figure 6a) allow adsorbate
adsorbent, Ce is the residual equilibrium concentration of dye, molecules (MB) to diffuse into the active sites of GO-LCTS.23
and K1 is the Langmuir adsorption affinity constant. In eq 6, Kf Additionally, both GO and LCTS are known to remove dyes
is the Freundlich adsorption capacity constant and nf relates to via electrostatic interactions.40,41 Therefore, the combined
the intensity of adsorption, where 1/nf < 1.0 represents highly effect of these precursors contribute excellent sorption
favorable adsorption, whereas 1/nf > 2 denotes unfavorable properties that enhance the overall sorption capacity of the
adsorption process, where 1/nf < 1 obtained for both GO and GO-LCTS composite.
GO-LCTS materials showed favorable adsorption for these The high removal efficiency of cross-linked GO samples is
adsorbents. In eq 7, Ks is the Sips adsorption constant that evidenced by Figure 7b. At an initial MB dye concentration
relates to the adsorption energy, and ns indicates surface from 100 to 1000 μM, variable decolorization occurs after the
heterogeneity of the sorbent, where Ce is defined as in eq 5. adsorption process relative to the initially turbid colored
The adsorption isotherms for MB onto GO and GO-LCTS solutions prior to adsorption.
are shown in Figure 7, where the best-fit sorption parameters 3.7.2. Kinetic Dye Uptake of MB. The kinetic uptake
obtained by the Sips model are listed in Table 2, where performance of GO and GO-LCTS materials are shown in
Figure 8, where the MB adsorption capacity increased quickly
Table 2. Isotherm Parameters for MB Adsorption with GO
and GO-LCTS Sorbent Materials at 295 K
sorbent material
adsorbate parameter GO GO-LCTS
MB Qm (mg g−1) 286.9 402.6
Ks (L mg−1) 0.028 0.111
ns 0.90 0.96
R2 0.960 0.972

favorable correlation coefficients (R2) are obtained (R2 ≈


0.960−0.972). The MB adsorption results for GO and GO-
LCTS material adopt behavior described by monolayer
adsorption onto homogeneous adsorption sites. The latter is
supported by values of ns near unity (ns = 0.96) for the surface
cross-linked GO (GO-LCTS). The maximum adsorption Figure 8. Kinetic uptake profile of MB with GO and GO cross-linked
capacity (Qm) of MB, according to the Sips model for GO- composite material, where the solid lines represent the best fit by the
LCTS (Qm = 402.6 mg g−1) exceeds that for pristine GO (Qm PSO model (see eq 4).
= 286.9 mg g−1) and LCTS (Qm = 12.0 mg g−1; data not
shown here). The adsorption capacity of MB by GO-LCTS for both GO and GO-LCTS sorbent materials in the first 50
reported here is notably higher than a number of reported min and decreased slowly thereafter. The rapid MB uptake is
values for other graphene-based adsorbents (cf. Table S1; contributed by the negatively charged adsorption sites
Supporting Information). accessible on the sorbent surface. By contrast, the slower
The Qm value for cross-linked GO increased by 115 mg g−1 uptake is because of the greater adsorption site occupancy as
compared to unmodified GO, reflecting the enhanced the surface sites become more saturated with MB at later stages
adsorption properties of the composite. The superior Qm of the isotherm profile. Consequently, preliminary MB
values may relate to several factors: (i) alteration of the adsorption may avoid the diffusion of further species onto
surface functional groups of GO-LCTS provides additional the dense GO-LCTS structure, resulting in longer equilibrium
adsorption sites that favor adsorption, and (ii) cross-linking of times.
GO with LCTS contributes cooperative sorbent−MB dye Table 3 represents the kinetic adsorption parameters for
interactions because of changes in the macromolecular sorbent materials with MB over a 250 min interval, in which
structure, along with alteration of the hydrophile−lipophile
balance of the composite over GO. It should be noted that Table 3. PSO Kinetic Model Values for GO-Based Materials
pristine LCTS has negligible uptake of MB, whereas the GO- at 295 K
LCTS material has notably higher uptake of MB. There is a
sorbent material Qm (μmol/g) k2 (g/μmol min) R2
strong binding interaction between MB (cationic dye) and the
GO-based materials because of the presence of polar/charged GO-LCTS 5.10 0.019 0.987
functional groups with Lewis base character on the surface of GO 4.17 0.004 0.996
these materials. MB shows a favorable binding with GO,
whereas the GO cross-linked material shows much greater the PSO kinetic model provided reasonable fitting results
uptake of MB after cross-linking with LCTS. The higher based on favorable correlation coefficients (R2), where R2 ≈
adsorption capacity of GO-LCTS is attributed to its composite 0.987−0.996. According to the obtained kinetic parameters (k2
structure, which serves to disperse GO and prevent its and Qe), the GO-LCTS material has greater uptake over
aggregation in aqueous media. The favorable dispersion of pristine GO based on the kinetic profiles. The value of Qe
GO through the composite lends to its large surface area and (μmol g−1) and the PSO rate constant values (k2; g/μmol min)
13051 DOI: 10.1021/acsomega.8b01871
ACS Omega 2018, 3, 13045−13054
ACS Omega Article

for the sorbent materials increased after cross-linking: GO extraction for potential applications in wastewater decontami-
(4.17 μmol g−1) < GO-LTS (5.10 μmol g−1), and GO (0.004 nation, advanced nanomedicine, and drug delivery.45−48
g/μmol min) < GO-LCTS (0.019 g/μmol min), respectively.
These trends in uptake are consistent with the results for the
TGA and spectral characterization (FTIR, 13C NMR, and

*
ASSOCIATED CONTENT
S Supporting Information
Raman) for the GO-LCTS material. Also, a previous report17 The Supporting Information is available free of charge on the
indicates that the intraparticle diffusion and external diffusion ACS Publications website at DOI: 10.1021/acsome-
play an important role in the adsorption kinetics.42 It has been ga.8b01871.
previously reported that the structure of monolith aerogels
containing CTS is affected by the addition of GO. This leads Additional description of composite regeneration
properties for the uptake of MB and tabular comparison
to considerable changes in the morphological characteristics of
of literature values of adsorption properties of composite
both CTS and GO43 that parallel observations noted in the
materials with MB (PDF)
SEM results for GO-LCTS in Figure 3. Changes in the
macromolecular structure of GO and its composite form are
evidenced according to the change in surface area upon
swelling, sorption, and storage capacity,43 as similarly noted in
■ AUTHOR INFORMATION
Corresponding Author
the case of water swelling or strong interactions with gaseous *E-mail: lee.wilson@usask.ca Phone: +1-306-966-2961. Fax:
species.44 Additionally, cross-linking GO with LCTS facilitated +1-306-966-4730.
the nano-dispersion of components that yield more surface ORCID
active sites for adsorption that enhance the adsorption capacity Lee D. Wilson: 0000-0002-0688-3102
of GO-LCTS, as outlined below(cf. Figure 8).
The greater kinetic uptake of the composite material Notes
The authors declare no competing financial interest.


provides support that cooperative effects occur because of
amide bond formation between GO and LCTS that may yield
secondary adsorption sites for MB adsorption. Also, there are ACKNOWLEDGMENTS
more active adsorption sites according to the increasing Qe The authors are grateful for the support of the Natural
values as the GO becomes cross-linked with LCTS. The Sciences and Engineering Research Council of Canada
decreasing kinetic trend of all sorbent materials also suggest (NSERC) Discovery Grant (418729-2012 RGPIN and
that diffusion of MB through the pore network of the sorbent RGPIN 2016-06197) for the financial support of this study.
materials decreased with increasing contact time between Support and expert technical assistance from Leila Dehabadi
sorbent where MB, along with a decreasing number of available for acquiring IR/13C solids NMR spectra and TGA profiles are
adsorption sites over the kinetic profile as time increases. greatly appreciated.

4. CONCLUSIONS ■ ABBREVIATIONS AND NOMENCLATURE


GO, graphene oxide; LCTS, low molecular weight chitosan;
GO and its cross-linked composites were synthesized by cross-
MW, molecular weight; MB, methylene blue


linking LCTS and GO to yield a framework material. The
materials were systematically characterized by several comple- REFERENCES
mentary methods to affirm the structure of the GO-LCTS
composite, where MB removal efficacy was used to study the (1) Dotto, G. L.; Moura, J. M.; Cadaval, T. R. S.; Pinto, L. A. A.
Application of Chitosan Films for the Removal of Food Dyes from
adsorption properties of the materials at equilibrium and Aqueous Solutions by Adsorption. Chem. Eng. J. 2013, 214, 8−16.
kinetic conditions. The FTIR and 13C NMR spectra are (2) Ayad, M. M.; El-Nasr, A. A. Adsorption of Cationic Dye
reported for the first time for GO-LTCS composites that (Methylene Blue) from Water Using Polyaniline Nanotubes Base. J.
provide support for the formation of amide linkages and a Phys. Chem. C 2010, 114, 14377−14383.
unique framework structure relative to the 2D layered structure (3) Namasivayam, C.; Arasi, D. J. S. E. Removal of Congo Red from
of GO. The greater surface roughness of the GO cross-linked Wastewater by Adsorption onto Waste Red Mud. Chemosphere 1997,
composite parallels the variable morphology revealed by SEM 34, 401−417.
that is also supported by the concomitant changes in thermal (4) Ali, I.; Asim, M.; Khan, T. A. Low Cost Adsorbents for the
stability revealed by TGA results. The GO composite displays Removal of Organic Pollutants from Wastewater. J. Environ. Manage.
2012, 113, 170−183.
improved adsorbent properties over GO that are suitable for (5) Dinu, M. V.; Lazar, M. M.; Dragan, E. S. Dual Ionic Cross-
wastewater decontamination, as shown by its effective MB Linked Alginate/Clinoptilolite Composite Microbeads with Improved
removal at equilibrium and kinetic conditions. The monolayer Stability and Enhanced Sorption Properties for Methylene Blue. React.
adsorption capacity (Qm) of the GO-LCTS material with MB Funct. Polym. 2017, 116, 31−40.
was 402.6 mg g−1, which far exceeds that of pristine GO and (6) Ramesha, G. K.; Vijaya Kumara, A.; Muralidhara, H. B.;
CTS. The development GO-based composites derived via Sampath, S. Graphene and Graphene Oxide as Effective Adsorbents
biopolymer cross-linking display enhanced adsorption proper- toward Anionic and Cationic Dyes. J. Colloid Interface Sci. 2011, 361,
ties with improved stability for reuse in multiple adsorption− 270−277.
desorption processes. This strategy extends the field of (7) Sears, K.; Dumée, L.; Schütz, J.; She, M.; Huynh, C.; Hawkins,
S.; Duke, M.; Gray, S. Recent Developments in Carbon Nanotube
application of GO due to the facile nature and versatility of Membranes for Water Purification and Gas Separation. Materials
the cross-linking approach for tuning the structure of GO, 2010, 3, 127−149.
along with the unique adsorption properties of cross-linked (8) Wang, Y.; Li, Z.; Wang, J.; Li, J.; Lin, Y. Graphene and Graphene
GO−CTS. This work demonstrates the potential utility of Oxide: Biofunctionalization and Applications in Biotechnology.
GO−CTS composites as versatile candidates in solid phase Trends Biotechnol. 2011, 29, 205−212.

13052 DOI: 10.1021/acsomega.8b01871


ACS Omega 2018, 3, 13045−13054
ACS Omega Article

(9) Gao, W. The Chemistry of Graphene Oxide. Graphene Oxide: Graphene Oxide/Polyelectrolyte LbL Assemblies by Controlling PH
Reduction Recipes, Spectroscopy, and Applications; Springer, 2015; 61− of GO Suspension to Fabricate Transparent and Super Gas Barrier
95. Films. Nanoscale 2013, 5, 9081−9088.
(10) Pei, S.; Cheng, H.-M. The Reduction of Graphene Oxide. (30) Zuo, P.-P.; Feng, H.-F.; Xu, Z.-Z.; Zhang, L.-F.; Zhang, Y.-L.;
Carbon 2012, 50, 3210−3228. Xia, W.; Zhang, W.-Q. Fabrication of Biocompatible and Mechanically
(11) Li, W.; Zheng, X.; Dong, Z.; Li, C.; Wang, W.; Yan, Y.; Zhang, Reinforced Graphene Oxide-Chitosan Nanocomposite Films. Chem.
J. Molecular Dynamics Simulations of CO2/N2 Separation through Cent. J. 2013, 7, 39−40.
Two-Dimensional Graphene Oxide Membranes. J. Phys. Chem. C (31) Kudin, K. N.; Ozbas, B.; Schniepp, H. C.; Prud’homme, R. K.;
2016, 120, 26061−26066. Aksay, I. A.; Car, R. Raman Spectra of Graphite Oxide and
(12) Kim, S.; Zhou, S.; Hu, Y.; Acik, M.; Chabal, Y. J.; Berger, C.; de Functionalized Graphene Sheets. Nano Lett. 2008, 8, 36−41.
Heer, W.; Bongiorno, A.; Riedo, E. Room-Temperature Metastability (32) Zhao, H.; Jiao, T.; Zhang, L.; Zhou, J.; Zhang, Q.; Peng, Q.;
of Multilayer Graphene Oxide Films. Nat. Mater. 2012, 11, 544−549. Yan, X. Preparation and Adsorption Capacity Evaluation of Graphene
(13) Chen, W.; Li, S.; Chen, C.; Yan, L. Self-Assembly and Oxide-Chitosan Composite Hydrogels. Sci. China Mater. 2015, 58,
Embedding of Nanoparticles by in Situ Reduced Graphene for 811−818.
Preparation of a 3D Graphene/Nanoparticle Aerogel. Adv. Mater. (33) Fan, Z.; Wang, K.; Wei, T.; Yan, J.; Song, L.; Shao, B. An
2011, 23, 5679−5683.
Environmentally Friendly and Efficient Route for the Reduction of
(14) Liu, J.; Li, P.; Xiao, H.; Zhang, Y.; Shi, X.; Lü, X.; Chen, X.
Graphene Oxide by Aluminum Powder. Carbon 2010, 48, 1686−
Understanding Flocculation Mechanism of Graphene Oxide for
1689.
Organic Dyes from Water: Experimental and Molecular Dynamics
(34) He, H.; Riedl, T.; Lerf, A.; Klinowski, J. Solid-State NMR
Simulation. AIP Adv. 2015, 5, 117151.
(15) Park, S.; Lee, K.-S.; Bozoklu, G.; Cai, W.; Nguyen, S. T.; Ruoff, Studies of the Structure of Graphite Oxide. J. Phys. Chem. 1996, 100,
R. S. Graphene Oxide Papers Modified by Divalent Ions-Enhancing 19954−19958.
Mechanical Properties via Chemical Cross-Linking. ACS Nano 2008, (35) Mahaninia, M. H.; Wilson, L. D. Modular Cross-Linked
2, 572−578. Chitosan Beads with Calcium Doping for Enhanced Adsorptive
(16) Mathesh, M.; Liu, J.; Nam, N. D.; Lam, S. K. H.; Zheng, R.; Uptake of Organophosphate Anions. Ind. Eng. Chem. Res. 2016, 55,
Barrow, C. J.; Yang, W. Facile Synthesis of Graphene Oxide Hybrids 11706−11715.
Bridged by Copper Ions for Increased Conductivity. J. Mater. Chem. C (36) Sheng, Y.; Tang, X.; Peng, E.; Xue, J. Graphene Oxide Based
2013, 1, 3084−3090. Fluorescent Nanocomposites for Cellular Imaging. J. Mater. Chem. B
(17) Yang, S.-T.; Luo, J.; Liu, J.-H.; Zhou, Q.; Wan, J.; Ma, C.; Liao, 2013, 1, 512−521.
R.; Wang, H.; Liu, Y. Graphene Oxide/Chitosan Composite for (37) Cairns, P.; Miles, M. J.; Morris, V. J.; Ridout, M. J.; Brownsey,
Methylene Blue Adsorption. Nanosci. Nanotechnol. Lett. 2013, 5, 372− G. J.; Winter, W. T. X-Ray Fibre Diffraction Studies of Chitosan and
376. Chitosan Gels. Carbohydr. Res. 1992, 235, 23−28.
(18) Grande, C. D.; Mangadlao, J.; Fan, J.; De Leon, A.; Delgado- (38) Abolhassani, M.; Griggs, C. S.; Gurtowski, L. A.; Mattei-Sosa, J.
Ospina, J.; Rojas, J. G.; Rodrigues, D. F.; Advincula, R. Chitosan A.; Nevins, M.; Medina, V. F.; Morgan, T. A.; Greenlee, L. F. Scalable
Cross-Linked Graphene Oxide Nanocomposite Films with Anti- Chitosan-Graphene Oxide Membranes: The Effect of GO Size on
microbial Activity for Application in Food Industry. Macromol. Symp. Properties and Cross-Flow Filtration Performance. ACS Omega 2017,
2017, 374, 1600114−1600122. 2, 8751−8759.
(19) Turgut, H.; Tian, Z. R.; Yu, F.; Zhou, W. Multivalent Cation (39) Han, D.; Yan, L.; Chen, W.; Li, W. Preparation of chitosan/
Cross-Linking Suppresses Highly Energetic Graphene Oxide’s graphene oxide composite film with enhanced mechanical strength in
Flammability. J. Phys. Chem. C 2017, 121, 5829−5835. the wet state. Carbohydr. Polym. 2011, 83, 653−658.
(20) Duri, S.; Tran, C. D. Supramolecular Composite Materials from (40) Li, F.; Chung, S.; Oh, G.; Seo, T. S. Three-Dimensional
Cellulose, Chitosan, and Cyclodextrin: Facile Preparation and Their Graphene Oxide Nanostructure for Fast and Efficient Water-Soluble
Selective Inclusion Complex Formation with Endocrine Disruptors. Dye Removal. ACS Appl. Mater. Interfaces 2012, 4, 922−927.
Langmuir 2013, 29, 5037−5049. (41) Crini, G.; Badot, P.-M. Application of Chitosan, a Natural
(21) Wu, F.-C.; Tseng, R.-L.; Juang, R.-S. Kinetic Modeling of Aminopolysaccharide, for Dye Removal from Aqueous Solutions by
Liquid-Phase Adsorption of Reactive Dyes and Metal Ions on Adsorption Processes Using Batch Studies: A Review of Recent
Chitosan. Water Res. 2001, 35, 613−618. Literature. Prog. Polym. Sci. 2008, 33, 399−447.
(22) Yang, X.; Tu, Y.; Li, L.; Shang, S.; Tao, X.-m. Well-Dispersed (42) Chen, L.; Li, Y.; Hu, S.; Sun, J.; Du, Q.; Yang, X.; Ji, Q.; Wang,
Chitosan/Graphene Oxide Nanocomposites. ACS Appl. Mater. Z.; Wang, D.; Xia, Y. Removal of Methylene Blue from Water by
Interfaces 2010, 2, 1707−1713. Cellulose/Graphene Oxide Fibres. J. Exp. Nanosci. 2016, 11, 1156−
(23) Chen, Y.; Chen, L.; Bai, H.; Li, L. Graphene oxide-chitosan
1170.
composite hydrogels as broad-spectrum adsorbents for water
(43) Alhwaige, A. A.; Agag, T.; Ishida, H.; Qutubuddin, S. Biobased
purification. J. Mater. Chem. A 2013, 1, 1992−2001.
chitosan hybrid aerogels with superior adsorption: Role of graphene
(24) Hummers, W. S.; Offeman, R. E. Preparation of Graphitic
Oxide. J. Am. Chem. Soc. 1958, 80, 1339. oxide in CO2 capture. RSC Adv. 2013, 3, 16011−16020.
(25) Udoetok, I. A.; Wilson, L. D.; Headley, J. V. Self-Assembled and (44) Liu, S.; Sun, L.; Xu, F.; Zhang, J.; Jiao, C.; Li, F.; Li, Z.; Wang,
Cross-Linked Animal and Plant-Based Polysaccharides: Chitosan- S.; Wang, Z.; Jiang, X.; Zhou, H.; Yang, L.; Schick, C. Nanosized Cu-
Cellulose Composites and Their Anion Uptake Properties. ACS Appl. MOFs Induced by Graphene Oxide and Enhanced Gas Storage
Mater. Interfaces 2016, 8, 33197−33209. Capacity. Energy Environ. Sci. 2013, 6, 818−823.
(26) Mohamed, M. H.; Dolatkhah, A.; Aboumourad, T.; Dehabadi, (45) An, S. S.; Wu, S.-Y.; Hulme, J. Current Applications of
L.; Wilson, L. D. Investigation of Templated and Supported Graphene Oxide in Nanomedicine. Int. J. Nanomed. 2015, 2015, 9−
Polyaniline Adsorbent Materials. RSC Adv. 2015, 5, 6976−6984. 24.
(27) Mohamed, M.; Wilson, L. Kinetic Uptake Studies of Powdered (46) Puvirajesinghe, T. M.; Zhi, Z. L.; Craster, R. V.; Guenneau, S.
Materials in Solution. Nanomaterials 2015, 5, 969−980. Tailoring Drug Release Rates in Hydrogel-Based Therapeutic
(28) Konkena, B.; Vasudevan, S. Understanding Aqueous Disper- Delivery Applications Using Graphene Oxide. J. R. Soc., Interface
sibility of Graphene Oxide and Reduced Graphene Oxide through 2018, 15, 20170949.
pKa Measurements. J. Phys. Chem. Lett. 2012, 3, 867−872. (47) Guo, R.; Wilson, L. D. Synthetically Engineered Chitosan-
(29) Chen, J.-T.; Fu, Y.-J.; An, Q.-F.; Lo, S.-C.; Huang, S.-H.; Hung, Based Materials and their Sorption Properties with Methylene Blue in
W.-S.; Hu, C.-C.; Lee, K.-R.; Lai, J.-Y. Tuning Nanostructure of Aqueous Solution. J. Colloid Interface Sci. 2012, 388, 225−234.

13053 DOI: 10.1021/acsomega.8b01871


ACS Omega 2018, 3, 13045−13054
ACS Omega Article

(48) Wilson, L. D.; Tewari, B. B. Chitosan-Based Adsorbents:


Environmental Applications for the Removal of Arsenicals. Mater. Res.
Found. 2018, 34, 133−160.

13054 DOI: 10.1021/acsomega.8b01871


ACS Omega 2018, 3, 13045−13054

You might also like