Вы находитесь на странице: 1из 43

 

 
Remarkable performance of magnetized chitosan-decorated lignocellulose
fiber toward biosorptive removal of acidic azo colorant from aqueous environ-
ment

Cheng–Gang Zhou, Qiang Gao, Shi Wang, Yan–Sheng Gong, Kai–Sheng Xia,
Bo Han, Meng Li, Yuan Lin

PII: S1381-5148(15)30072-9
DOI: doi: 10.1016/j.reactfunctpolym.2015.11.010
Reference: REACT 3591

To appear in:

Received date: 22 September 2015


Revised date: 8 November 2015
Accepted date: 30 November 2015

Please cite this article as: Cheng–Gang Zhou, Qiang Gao, Shi Wang, Yan–Sheng Gong,
Kai–Sheng Xia, Bo Han, Meng Li, Yuan Lin, Remarkable performance of magnetized
chitosan-decorated lignocellulose fiber toward biosorptive removal of acidic azo colorant
from aqueous environment, (2015), doi: 10.1016/j.reactfunctpolym.2015.11.010

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Remarkable performance of magnetized


chitosan-decorated lignocellulose fiber toward
biosorptive removal of acidic azo colorant from

T
IP
aqueous environment

R
Cheng–Gang Zhoua, Qiang Gaoa,b,*, Shi Wanga, Yan–Sheng Gonga, Kai–Sheng Xiaa,

SC
Bo Hana, Meng Lia, Yuan Lina

a
Department of Chemistry, Faculty of Material Science and Chemistry, China

NU
University of Geosciences, Wuhan 430074, PR China
MA
b
Engineering Research Center of Nano-Geo Materials of Ministry of Education,

China University of Geosciences, Wuhan 430074, PR China


D
TE

*
Corresponding author

Tel.: +86 027 6788 3731; fax: +86 027 6788 3731.
P
CE

E-mail address: gaoqiang@cug.edu.cn (Q. Gao).


AC

1
ACCEPTED MANUSCRIPT

Abstract

Numerous low-cost adsorbents have been used for removal of harmful azo colorants

from aqueous environment, but they frequently suffer from either low/limited

T
IP
adsorption capacity or slow adsorption kinetics; moreover, almost all of them become

R
problematic on the regard of recycling and reuse. This research explores the

SC
possibility of using a novel low-cost biocomposite, i.e., magnetized chitosan-coated

lignocellulose fiber, for biosorptive removal of acidic azo colorant from aqueous

NU
solution. This chitosan-decorated lignocellulose fiber (CS/LCF) was prepared for the
MA
first time via a facile “surface deposition–crosslinking” method, exhibiting a

well-defined coating structure. Interestingly, the CS/LCF could be easily magnetized


D

by simply blending it with magnetic nanoparticle (MNP) in an aqueous solution


TE

because of the spontaneous adherence of MNP to CS/LCF surface. More importantly,


P

the magnetized CS/LCF was found to be exceptionally effective and efficient in


CE

biosorptive removal of acidic azo compound (i.e., acid red 18), e.g., ultrahigh
AC

adsorption capacity (up to 1184 mg∙g–1), rapid adsorption rate, excellent reusability

(with removal efficiency up to 99.48% in tenth cycle), and convenience in

manipulation (magnetic separation). The inherent advantages of the magnetically

retrievable CS/LCF may pave a new, inexpensive, efficient, and sustainable way

towards removal of acidic azo colorants from Earth’s water environments.

Keywords: Lignocellulose fiber; Chitosan; Magnetized; Biosorption; Azo colorant

2
ACCEPTED MANUSCRIPT

1. Introduction

The ever-growing demand for textile products has created a dramatic increase of

colored effluents discharged from textile industries, making it one of the main sources

T
IP
of water pollution worldwide. Among various organic colorants, the azo colorants

R
represent the majority, making up close to 70% [1]. The principal characteristic for

SC
this kind of compounds is that they contain an azo bond (–N=N–) as a chromophore

group in association with aromatic rings containing some functional groups such as

NU
–OH and –SO3H groups [2]. The recalcitrant azo colorants, when present in receiving
MA
waters, can cause serious aesthetic problems due to their high color. Moreover, these

colorants can strongly absorb sunlight, thus impeding the photosynthetic activity of
D

aquatic plants and seriously threatening the whole ecosystem [3]. Therefore, the
TE

rational and efficient treatment of these discharged azo colorants has been regarded as
P

one of the most urgent and important global problems. For this reason, during the past
CE

decades, various methods based mainly on three different strategies, i.e., biological
AC

decomposition, catalytic oxidation, and direct adsorption, have been applied for the

removal of these colorants [4]. Biological methods are cheap and simple to apply, but

they are generally not sufficiently effective due to the high resistance to aerobic

digestion of azo colorants [5]. Catalytic oxidation-based processes may be able to

eliminate some azo colorants, but their applications are severely restricted by either

the extensive use of chemical reagents (for homogeneous catalysis) or the limited

availability and performance of catalysts (for heterogeneous catalysis) [6]. In sharp

contrast, the removal processes based on adsorption are becoming increasingly

3
ACCEPTED MANUSCRIPT

popular due to the flexibility and simplicity of design, ease of operation, and

insensitivity to toxic pollutants [7,8]. Moreover, adsorption also does not result in the

formation of harmful substances.

T
IP
As far, the majority of commercial systems use activated carbon as an adsorbent

R
to remove azo colorants in wastewaters because of its good adsorption ability [9].

SC
Activated carbon adsorption has been also cited by the US Environmental Protection

Agency as one of the best available control technologies [10]. However, although

NU
activated carbon is a preferred adsorbent, its widespread use is also restricted due to
MA
its high initial cost and the need for a costly regeneration [10]. Consequently,

numerous low-cost alternatives have been proposed, including natural materials such
D

as chitosan, zeolites, clay, or certain waste products from agro-industrial operations


TE

such as fly ash, rice husk, wheat straws, and lignocellulose fiber [11].
P

Chitosan (CS) has gained a significant interest among scientists, mainly due to
CE

its valuable properties such as great hydrophilicity (water loving), non-toxicity, and
AC

the presence of abundant active sites (e.g., –NH2) in its polymeric chain [12]. In some

Asian countries such as China, Japan, and Thailand, the fishery byproducts (e.g.,

shrimp, lobster, and crab shells) obtained for free from local fishery industries have

been developed into one of the promising options to produce CS at low cost [13].

Since CS has these advantages (containing abundant –NH2 groups, great availability,

and low cost), it is now regarded as a very promising biosorbent for wastewater

treatment [14]. Recently, McKay et al. studied the biosorption of CS towards acidic

azo colorants, and found that the adsorption amount of acid red 18 (AR18) on CS was

4
ACCEPTED MANUSCRIPT

up to 693.2 mgg1 [15]. But it should be noted that, almost all the reported pure

CS-based biosorbents are cross-linked into m- even mm-sized beads, in order to

stabilize them in acid solution as well as facilitate their separation processes after

T
IP
adsorption [16]. In such cases, the adsorbate molecules have poor site accessibility,

R
eventually causing a slow adsorption kinetic [16].

SC
Agro-industrial wastes as adsorbents are also drawing considerable attentions

due to their local availability in large quantities, low or no cost, eco-friendliness, and

NU
effectiveness [17]. Lignocellulose fiber (LCF), being composed of two carbohydrate
MA
polymers (i.e., cellulose and hemicellulose) and one aromatic polymer (i.e., lignin),

represents a major or sometimes the sole component of different waste streams from
D

various industries [18]. Recent studies have demonstrated that LCF is practicable for
TE

biosorptive removal of azo colorants, due to its great availability, excellent chemical
P

stability, unique morphology, and good accessibility of adsorption sites [19].


CE

However, pure LCF generally exhibited a low adsorption capacity because of the
AC

evident lack of adsorption sites on its surface [19]. Esters, aspartic acid, anhydrides,

tannin, triethylenetetramine, poly(methacrylic acid) and other functional entities have

been used in the chemical modification on LCF [20]. But these strategies have a

limited success in improving the adsorption efficiency of LCF. For example, Roy et al.

prepared a polyphenolic tannin-modified LCF and applied it to remove the azo

colorant (Congo red) from aqueous solution [20]. Result showed that the maximum

adsorption capacity was 27.12 mgg1 [20], which is quite lower compared with the

5
ACCEPTED MANUSCRIPT

activated carbons (e.g., 300 mgg1 [21]). Obviously, efforts are still needed regarding

the different possibilities of LCF functionalization.

Advantages and disadvantages of CS and LCF introduced above, along with the

T
IP
fact that neither one perfectly act as a biosorbent, strongly inspire us to combine them

R
to into a single material (i.e., LCF coating with CS), maximizing their individual

SC
strengths while avoiding or mitigating their weaknesses. On the other hand,

adsorption processes combined with magnetic separation technology have recently

NU
been developed and show additional advantages of accelerating adsorbent separation
MA
and facilitating the processes of adsorption and desorption [22,23]. Herein, the

objectives of this work are to develop a novel magnetically retrievable CS-decorated


D

LCF biocomposite, and explain its biosorption behaviors to see if desired adsorption
TE

properties such as high adsorption capacity, rapid adsorption rate, good reusability,
P

and manipulative convenience are achieved. As shown in Fig. 1, the CS-decorated


CE

LCF (CS/LCF) is prepared via a simple “surface depositioncrosslinking” procedure.


AC

Furthermore, the CS/LCF is magnetized by simply blending it with magnetic

nanoparticle (MNP) in an aqueous solution taking advantages of the spontaneous

adherence of MNP to CS/LCF surface. To evaluate the biosorption properties of the

magnetically retrievable CS/LCF, the acid red 18 (AR18), a typical acid azo colorant,

was chosen as the model adsorbate. Effects of the operating parameters (e.g., pH,

ionic strength, temperature, and contact time) on adsorption efficiency, isothermal

adsorptions, adsorption kinetics, and reusability of magnetically retrievable CS/LCF

were investigated and discussed in detail.

6
ACCEPTED MANUSCRIPT

2. Experimental

2.1 Synthesis of MNP and CS/LCF

The information of chemicals used in this work can be seen in Text S1. The acid

T
IP
red 18 (AR18) is an important acid azo colorant due to it wide use in textile, cosmetic,

R
and food industries [24]. The improper disposal of this colorant in industrial

SC
wastewater into the environment poses a risk to the aquatic life and contaminates

water for human consumption [25]. This is also the reason why we selected AR18 as

NU
the target azo colorant. The AR18 has a dimension of 1.42 nm (length)  0.765 nm
MA
(width), and the area of AR18 molecule is therefore about 1.09 nm2 (Fig. 1) [7].

Magnetic nanoparticle (MNP), exclusively made up of monodisperse sphere-like


D

particles with a mean size of about 65 nm, was synthesized via a solvothermal process
TE

according to our previously used method [26].


P

Chitosan-coated lignocellulose fiber (CS/LCF) was prepared via a “surface


CE

deposition–crosslinking” procedure. Typically, 1.0 g of LCF and 1.0 g of CS were


AC

successively added into an acetic acid solution (0.2% (v/v); 1 L). Then, the mixture

was stirred vigorously for 1 h to facilitate the dissolution of CS and uniform

dispersion of LCF. Afterwards, 1% (v/v) ammonia solution was added, drop by drop,

into the mixture to adjust the pH up to 11 in order to deposit CS onto the LCF surface.

Thereafter, the mixture was heated to 60 oC and 0.5 mL of glutaraldehyde was

introduced for the crosslinking of deposited CS. Finally, the resulting CS/LCF were

collected by centrifugation and washed with diluted acetic acid to remove the

adsorbed and excessive or uncrosslinked CS. The obtained material, i.e., CS/LCF,

7
ACCEPTED MANUSCRIPT

was dried at 60 oC overnight for subsequent use.

2.2 Magnetization of CS/LCF and batch adsorption procedures

The magnetization of CS/LCF and its adsorption process co-occurred by simply

T
IP
blending CS/LCF with MNP into an AR18 solution (Fig. 1). Typically, 20 mL of

R
AR18 solution (300 mgL–1; pH 3.0) were added into a vial that contained 10 mg of

SC
CS/LCF and 15 mg of MNP, and then the mixture was shaken with speed of 120 rpm

at 303.15 K. After the adsorption reached equilibrium (240 min), the AR18-adsorbed

NU
adsorbent was magnetically retrieved from the solution. The adsorption amount of
MA
AR18 was determined by using a UV–Vis spectrophotometer to monitor the change of

AR18 concentrations before and after adsorption.


D

Through a similar procedure described above, the effect of solution pH on the


TE

adsorption efficiency was investigated at initial concentration of 300 mgL–1 in the pH


P

range from 3.0 to 8.0 by adding a certain amount of 0.1 molL–1 HCl or 0.1 molL–1
CE

NaOH solution. At optimal pH, the effect of ionic strength was also studied by
AC

gradually increasing KCl concentration in adsorption solution from 0 to 400

mmolL–1.

Kinetic experiments were conducted at different time intervals with three

different initial AR18 concentrations (300, 400, and 500 mgL–1), respectively.

Isotherm adsorption experiments were carried out at three different temperatures

(303.15, 323.15, and 343.15 K) over a range of initial AR18 concentration from 300

to 700 mg·L–1.

The experimental details about reusability of magnetically retrieved CS/LCF can

8
ACCEPTED MANUSCRIPT

been seen in Text S2.

2.3 Characterization of materials

The powder X-ray diffraction (XRD) measurements were recorded on an X-ray

T
IP
diffractometer (D8-FOCUS, Bruker, Germany) using Cu-Kα radiation with scattering

R
angles. The FT-IR spectra in the range 4000 to 400 cm–1 were measured on a Fourier

SC
transform infrared spectrometer (VERTEX 70, Bruker, Germany). Thermal

gravimetric analyses were performed with a thermal analyzer system under N2 flow at

NU
a heating rate of 10 oCmin–1 (TGA/DTG, STA 449 F3, Netzsch, Germany). The
MA
micro-morphologies of materials were characterized by a scanning electron

microscope (SEM, SU8010, Hitachi, Japan).


D

3. Results and Discussion


TE

3.1. Characterization of materials


P

The physical characterization of adsorbent is crucial to prove that it is well


CE

developed. Thus, different techniques were applied to examine the morphologies,


AC

compositions, and structures of these materials involved in this work.

3.1.1 SEM images

For a morphological characterization of the samples, a field-emission scanning

electron microscope (SEM) was used. Fig. S1 and Fig. 2 show the SEM images of

MNP, LCF, CS/LCF, and magnetized CS/LCF (denoted as mCS/LCF). Clearly, the

MNP sample is exclusively made up of monodisperse sphere-like particles with a

mean size of about 65 nm (Fig. S1). As shown in Fig. 2a, the LCF sample consists of

one-dimensional fibers with diameters in the micron range (10–40 μm) and lengths in

9
ACCEPTED MANUSCRIPT

millimeter scale (Fig. 2a). Moreover, the SEM image of LCF at larger magnification

shows that the surface of LCF is coarse and heterogeneous (Fig. 2d). After coating

with CS, the resulting CS/LCF is found to still retain the fiber-like morphology, while

T
IP
its surface becomes smooth and homogeneous (Fig. 2b and 2e). Moreover, there are

R
almost no redundant CS particles present in the CS/LCF sample. The result indicates

SC
that CS polymers have been entirely deposited on the surfaces of LCF monomers to

form well-defined coating layers. Previous studies also reported the synthesis of

NU
CS-coated carbon nanotube (CS/CNT) using a similar strategy [27]. However, a
MA
significant amount of CS could not be effectively deposited onto the surface of CNT

perhaps because of the obvious difference between hydrophobic CNT and hydrophilic
D

CS. The better effectiveness of our approach in CS-coating might be due to the fact
TE

that LCF and CS are quite similar in chemical composition [28,29], which should be
P

advantageous for their combination. The magnetized CS/LCF was obtained by simply
CE

blending CS/LCF with MNP in an aqueous solution, and its images are shown in Fig.
AC

2c and 2f. Interestingly, most of the MNP particles are found to adhere to the surface

of CS/LCF. In our previous work, it has been demonstrated that CNT can be

magnetized by simply blending it with MNP in aqueous solution [23]. A possible

explanation for this magnetization behavior was as follows: Because CNT is a typical

solvophobic nano carbon material, CNT monomers could conglobe spontaneously to

form aggregates in the aqueous environment [23]. In the aggregation process, CNT

monomers could wrap MNP into their aggregates, leading to the formation magnetic

composite [23]. Very recently, we reported that the magnetization of CS/CNT could

10
ACCEPTED MANUSCRIPT

also be achieved via the “blending” strategy [7]. This was an interesting result, for it

suggested that although the CS/CNT had a highly hydrophilic surface, it could still be

magnetized by adding MNP. In our latest work, the carbon-decorated LCF (C@LCF)

T
IP
was magnetized by simply adding it and MNP into aqueous solution [29]. Different

R
from CNT and CS/CNT with high surface energies, the C@LCF was micro-size

SC
material and no aggregation between C@LCF monomers took place. Nevertheless,

the MNP still adhered to the surface of C@LCF and enabled the adsorbent to be

NU
magnetically retrievable. Summarizing the above findings, we now concluded that the
MA
magnetization of CNT, CS/CNT, or C@LCF might be similar to a fishing process,

where the “fish” were the magnetic nanoparticles and the “fishing net” was
D

composited of self-assembled aggregates (CNT or CS/CNT aggregates) or micro-size


TE

C@LCF. When an external magnetic field was applied, the magnetic nanoparticles
P

would adhere to the “fishing net”, leading to the magnetization of CNT, CS/CNT, or
CE

C@LCF. Encouraged by these findings, we make a new attempt in this work to


AC

magnetize the CS/LCF via the “blending” strategy. As expected, the MNP could

indeed adhere to the CS/LCF surface, making CS/LCF magnetically retrieved.

3.1.2 XRD patterns

The X-ray diffraction (XRD) patterns in the 2 range of 10o100o of MNP, LCF,

CS, CS/LCF, and mCS/LCF are shown in Fig. 3a. The XRD pattern of the MNP

shows six characteristic peaks that locate at 30.13o, 35.50o, 43.15o, 53.52o, 57.08o, and

62.67o, which can be assigned to the (220), (311), (400), (422), (511), and (440)

planes of spinel Fe3O4 (JCPDS file (PDF No. 65-3107)), respectively [23]. In the

11
ACCEPTED MANUSCRIPT

XRD pattern of LCF, the characteristic peaks at 15.59o, 22.33o, and 34.78o (Fig. 3a)

can be observed, which indicates the existence of crystalline cellulose in LCF

structure [30,31]. The XRD pattern of CS exhibits two strong peaks at 11.09° and

T
IP
20.08°, which are attributed to the hydrated crystal (Forms I) and anhydrous crystal

R
(Forms II), respectively [32]. The CS/LCF exhibits the same characteristic diffraction

SC
peaks as LCF (i.e., 15.59o, 22.33o, and 34.78o), except for the appearance of a peak at

20.08°, conforming the existences of both CS and LCF in this composite. The

NU
decrease of CS crystallinity in CS/LCF (the peak at 11.09° disappears) indicates that
MA
CS was well distributed on the LCF [32]. In the XRD pattern of mCS/LCF, six

characteristic diffraction peaks of MNP as well as the characteristic diffraction peaks


D

of CS/LCF can be observed, which is in accord with our expectations.


TE

3.1.3 TGA/DTG analyses


P

The TGA/DTG profiles of LCF, CS, and CS/LCF are shown in Fig. 3b. In the
CE

TGA curve of LCF, the weight loss below 200 oC should be due to the desorption of
AC

physically adsorbed water, and the significant weight loss between 200 and 450 oC

with a peak at 355 oC should stem from the thermal decomposition of cellulose,

hemicellulose, and lignin [30]. Moreover, an additional weight loss occurred at

temperature higher than 450 oC and should be a consequence of evaporation of the

volatile components [30]. For the pure CS sample, the weight loss occurred between

200 and 360 oC with its maximum value at 304 oC could be due to the degradation of

CS [33]. Besides, some weight loss occurred at higher temperatures is due to the

further condensation of carbonaceous surface species [33]. From the DTG curve of

12
ACCEPTED MANUSCRIPT

CS/LCF, it can be found that the weight loss between 200 and 450 oC appears to

progress two stages: at lower temperature, the weight loss should mainly originate

from the decomposition of CS; while at higher temperature, the decomposition of

T
IP
LCF becomes dominant. By comparing the three TGA curves shown in Fig. 3b, it can

R
be observed that the residue yields of LCF, CS, and CS/LCF are about 21.8 wt%, 34.1

SC
wt%, and 27.7 wt%, respectively, which correspond to the fixed carbon content of

samples [34]. According to the data, the mass of CS in CS/LCF is estimated to be

about 48 wt%.
NU
MA
3.1.4 FT-IR spectroscopy

Fig. 3c shows the FT-IR spectra of LCF and CS/LCF in the region of 4000400
D

cm–1. The LCF displays its typical peaks such as: (i) a broad band is found at 3350
TE

cm–1, which should arise from the stretching vibrations of various hydroxyl (–OH)
P

[35]. The –OH compounds may include absorbed water, aliphatic primary and
CE

secondary alcohols found in cellulose, hemicellulose, and lignin; (ii) the peak at 2909
AC

cm–1 should be associated to C–H stretching vibration; (iii) the peak at 1642 cm–1

corresponds to the stretching vibration of C=C in aromatic skeletal mode of lignin and

C=O of the alpha-keto carbonyl, and the peak at 1438 cm–1 is assigned to the

methoxyl group in lignin; and (iv) the peak at 1169 cm–1 is assigned to the stretching

vibration of C–O–C bridge in hemicellulose and cellulose and to the aromatic C–H

bending vibration of the syringyl and guaiacyl units in lignin, respectively [36,37].

Spectrum of CS/LCF contains all of the above-mentioned adsorption peaks, and two

additional peaks at 1634 and 1603 cm–1 (assigned to the stretching vibration of C=O

13
ACCEPTED MANUSCRIPT

in N-acetyl groups and the bending vibration of N–H in –NH2, respectively) also

appear [7], which confirms that CS is successfully coated on the surface of LCF.

3.2. Biosorption studies

T
IP
Azo colorants are one of the most important groups of pollutants present in water

R
[2,3]. The development of powerful and practical treatments of azo colorants in

SC
wastewater has attracted world-wide attention. Recent studies have demonstrated that

the adsorption method has great potential for removal of azo colorants from

NU
wastewater due to its efficiency, simplicity of design, ease of operation, insensitivity
MA
to toxic pollutants, and without secondary pollution [14]. Numerous adsorbents,

especially low-cost biosorbents, have been widely used for removal of azo colorants
D

from aqueous environment [10]. Unfortunately, these adsorbents frequently suffer


TE

from either low/limited adsorption capacity or slow adsorption kinetics; moreover,


P

almost all of them become problematic on the regard of recycling and reuse [6]. In the
CE

present work, a novel low-cost, amino group-rich, and sites-accessible biocomposite


AC

(CS/LCF) was successfully prepared via a facile “surface depositioncrosslinking”

procedure (Fig. 1). Interestingly, this biocomposite could be magnetized by simply

blending it with MNP in an aqueous solution (Fig. 1). We speculated that the

magnetically retrievable CS/LCF (mCS/LCF) biocomposite should have great

potential for being used as a high-performance biosorbent in biosorptive removal of

acidic azo colorant (e.g., acid red 18 (AR18)). To confirm this, a preliminary

experiment was conducted as follows: 10 mg of CS/LCF and 15 mg of MNP were

added into AR18 solution (300 mgL−1; pH 3.0; 20 mL), and the resulting mixture was

14
ACCEPTED MANUSCRIPT

shaken at room temperature for 240 min. After addition of an external magnetic field,

it was observed that AR18 was significantly removed (the supernatant became nearly

colorless), and all the adsorbent particles were quickly isolated from the solution (Fig.

T
IP
S2). These findings strongly suggested the feasibility and reliability of such magnetic

R
adsorption, and demonstrated its advantages in adsorption ability, adsorption rate, and

SC
manipulative convenience. To fully reveal the adsorption properties of mCS/LCF

toward AR18, the effects of operating parameters on adsorption efficiency, isothermal

NU
adsorptions, adsorption kinetics, and reusability of adsorbent were systematically
MA
investigated.

3.2.1 Effect of pH
D

The effect of solution pH on the adsorption efficiency was firstly investigated


TE

since pH is one of the most important factors that might influence adsorption capacity
P

(Fig. 4a). In view of the fact that MNP will be partially dissolved in bulk solution at
CE

pH less than 2.0 [23], the initial pH was performed over the pH range from 3.0 to 8.0.
AC

During the process of AR18 adsorption, the equilibrium pH values were also

monitored because they were essentially influencing the adsorption efficiency. As

shown in Fig. 4a, the adsorption efficiency of mCS/LCF towards AR18 is

significantly affected by pH. The adsorption capacity gets maximum at initial pH 3.0

(corresponds to equilibrium pH 4.7), and then decreases dramatically as the pH

increases. This can be explained by the fact that at low pH (acidic solution) more

protons will be available to protonate amino groups of chitosan to form –NH3+ groups,

thereby increasing the electrostatic attractions between negatively charged AR18

15
ACCEPTED MANUSCRIPT

anions and positively charged adsorption sites and causing an increase in adsorption

capacity [7]. The similar pH effects were also observed by the adsorption of acidic

azo colorant (acid orange II) on cross-linked chitosan fibers and the adsorption of

T
IP
acidic azo colorant (reactive red 189) on cross-linked chitosan beads [38,39].

R
3.2.2 Effect of ionic strength

SC
The ionic strength of the solution is of significance for its effect on the adsorbent,

as well as the adsorbate [40]. In general, as different kinds of salts are mixed in real

NU
wastewater, its ionic strength is high [40]. At high ionic strength, adsorption sites are
MA
surrounded by counter ions which partially lose their charge, and this weakens the

binding force by an electrostatic interaction [41]. To investigate the effect of ionic


D

strength on AR18 biosorption, equilibrium adsorption studies were performed by


TE

gradually increasing KCl concentration in adsorption solution from 0 to 400 mmolL–1


P

at initial pH 3.0. As expected, it is found that the adsorption efficiency of mCS/LCF


CE

towards AR18 is indeed a function of the added mount of KCl, as shown in Fig. 4b.
AC

The adsorption capacity decreases slightly as KCl concentration increases from 0 to

60 mmol∙L–1, and then goes down obviously with the further increase of KCl

concentration. But, it is worth noting that there is still a considerable adsorption

capacity even when the KCl concentration is up to 400 mmolL–1, which is much

higher than the initial concentration of AR18 (300 mgL–1, also equivalent to 0.496

mmolL–1). This result indicates a strong electrostatic interaction existing between

AR18 and mCS/LCF, and also implies that the mCS/LCF can be used for biosorptive

removal of AR18 from salt-containing wastewater.

16
ACCEPTED MANUSCRIPT

3.2.3 Effect of temperatures and biosorption isotherms

To ascertain the adsorption capacity of mCS/LCF, the isothermal experiments of

AR18 adsorption onto mCS/LCF under different temperatures were performed, and

T
IP
the results are shown in Fig 4c. It is found that varying the adsorption temperature has

R
less impact on the adsorption efficiency. The maximum adsorption capacities of

SC
mCS/LCF towards AR18 at 303.15, 323.15, and 343.15 K are 1184, 1182, and 1179

mgg–1, respectively. Increasing the temperature is known to lower the viscosity of the

NU
solution, which seems to create a favorable condition for transfer and diffusion of
MA
adsorbate from bulk solution to adsorbent surface (“positive effect”) [42]. However,

other researches have also revealed that the increase of temperature can cause greater
D

mobility of the adsorbate molecules previously adsorbed (“negative effect”) [43]. In


TE

our adsorption system, the two effects (i.e., “positive and negative effects”) probably
P

tended to offset each other, thereby causing a temperature-independent adsorption.


CE

Such an adsorption behavior is also in line with other work [44].


AC

To further interpret the adsorption behaviors, three common isotherm models,

Langmuir, Freundlich, and Sips models were applied to fit the experimental data,

respectively. The nonlinear forms of Langmuir and Freundlich models are expressed

as follows, respectively [7]:

qm K LCe
qe = (1)
1  K LCe

qe =K FCe1/ n (2)

where Ce is the equilibrium concentration of AR18 (mg·L–1), qe is the amount of

AR18 adsorbed at equilibrium (mg·g–1), qm is the maximum adsorption capacity


17
ACCEPTED MANUSCRIPT

(mg·g–1), and KL (L·mg–1) is the Langmuir binding constant; KF (mg1–1/n·L1/n·g–1) and

1/n are Freundlich constants, respectively.

Sips equation is represented as follows [45]:

T
IP
qm ( K s Ce )1/ ns
qe = (3)
1  ( K s Ce )1/ ns

R
where qm is the maximum adsorption capacity (mg·g–1), Ks (L·mg–1) is the median

SC
association constant, and 1/ns is the Sips heterogeneity constant.

NU
The adsorption isotherms of AR18 on the mCS/LCF predicted from all the three

models are plotted in Fig. 5. All the correlation coefficients, R2 values, and the
MA
constants obtained for the models are summarized in Table 1. The Sips model yields

the best fit with the highest R2 value of 0.95750.9946, thus it is the most suitable
D
TE

equation to describe the adsorption equilibrium of AR18 on the mCS/LCF. As well

known, Sips model is the combination of Langmuir and Freundlich isotherm models
P
CE

[45]. The heterogeneity factor of 1/ns close to or even 1 indicates adsorbents with

comparatively homogenous binding sites, while 1/ns1 indicates heterogeneous


AC

adsorbents [46]. In this study, the Sips model shows 1/ns values relatively close to 1.0

(Table 1), implying that this model looks more like Langmuir model than Freundlich

model. This also indicates the relatively homogeneous nature of mCS/LCF surface.

Similar phenomena were reported in the adsorptions of food dyes FD&C red no. 40

and acid blue 9 onto Spirulina platensis nanoparticles as well as adsorption of reactive

red 141 on an organoclay [47,48].

3.2.4 Effect of contact time and biosorption kinetics

Removal of AR18 for different initial concentrations (300, 400, and 500 mgL–1)
18
ACCEPTED MANUSCRIPT

as a function of contact time (3600 min) was studied and the results are showed in

Fig. 4d. Obviously, the adsorption capacity of mCS/LCF towards AR18 is closely

dependent on its initial concentration. In each case, the adsorption capacity

T
IP
significantly increases during the initial stage and then slowly reaches equilibrium

R
within 240 min. The phenomenon may be explained as follows: The CS coating of

SC
CS/LCF could swell in acidic solution (initial pH 3.0), which would result in the

appearance of some pores available for AR18 entrance [49]. During the adsorption

NU
process, the adsorption on the exterior surface should take place quickly, but it might
MA
be time-consuming for the AR18 molecules to diffuse into the interior space of

adsorbent.
D

In order to elucidate the kinetic process of AR18 adsorption on the mCS/LCF,


TE

two most widely used kinetic models (i.e., pseudo-first-order and


P

pseudo-second-order models) were used. Nonlinear forms of the two models are
CE

expressed as follows [7]:


AC

qt =qe (1  e  k1t ) (4)

k2 qe2t (5)
qt =
1  k2 qet

where qe and qt are the adsorption capacities (mg·g–1) at equilibrium and at time t,

respectively; k1 (min−1) and k2 (g·mg–1·min–1) are the equilibrium rate constants of

pseudo-first-order and pesudo-second-order models, respectively.

The adsorption kinetic curves of AR18 on the mCS/LCF predicted from the two

kinetic models are plotted in Fig. 6a. All the correlation coefficients (R2) and the

constants obtained for the models are summarized in Table 2. The values of the
19
ACCEPTED MANUSCRIPT

correlation coefficient indicate the applicability of the pseudo-second-order model for

describing the experimental results to a higher degree of accuracy (0.9813<R2<0.9921)

for all studied AR18 concentrations. In addition, Fig. 6a and Table 2, show that the q

T
IP
values (qe,calc,2) determined from pseudo-second-order equation were closer to the

R
experimental q values (qe,exp) than those determined from the pseudo-first-order

SC
equation. This also means that the adsorption rate is proportional to the square of the

number of free sites, which corresponds to the term (qe – qt)2 in the

pseudo-second-order model [50].


NU
MA
Nevertheless, the qt versus t plots do not provide a clear indication regarding

how the AR18 molecules reached the surface of mCS/LCF [51]. Therefore, the
D

diffusion model based on the theory proposed by Weber and Morris was tested.
TE

According to this theory [52]:


P

qt =kit 0.5  I (6)


CE

where qt is the adsorption capacity (mg·g–1) at time t, ki (mg·g–1·min–0.5) and I (mg·g–1)


AC

are the Weber–Morris diffusion constants. The plots of qt versus t0.5 for the adsorption

of AR18 at different initial concentrations are given in Fig. 6b. It is found that the

adsorption data could be fitted by two straight lines, indicating two steps have a

combined influence on the adsorption process: the first straight portion corresponds to

external diffusion and the second portion to intraparticle diffusion [53]. The slope (ki,1

and ki,2) and intercept (I1 and I2) values of two straight lines, along with R2 values are

listed in Table 2. The high values of R2 (0.9436~0.9936) confirm the suitability of this

model. From Table 2, it is also observed that the ki,2 values are much smaller than the

20
ACCEPTED MANUSCRIPT

ki,1 values, indicating that the adsorption process of AR18 onto mCS/LCF was mainly

governed by intraparticle diffusion [53]. This result can be explained as follows:

During the early stage of adsorption, AR18 could be favorably adsorbed onto the

T
IP
exterior surface of adsorbent due to the low mass-transfer resistance. When the

R
adsorption on the exterior surface reached saturation, AR18 molecules would diffuse

SC
into the pores of the adsorbent to interact with the adsorption sites present on the

interior surface. As a result, AR18 molecules had to overcome a considerable

NU
diffusion resistance, and took a longer time to reach the final equilibrium.
MA
3.3. Comparison of adsorption abilities

Firstly, a test for comparing adsorption capacities of mCS/LCF and CS/LCF was
D

carried out under the identical conditions. As shown in Fig. S3, the mCS/LCF shows a
TE

slightly higher removal efficiency as compared with CS/LCF, i.e., 99.69% vs 98.23%.
P

The better adsorption performance of mCS/LCF should be attributed to the fact that
CE

the MNP, in spite of being a poor adsorbent, could also get a removal rate of 3.33%
AC

when used alone (Fig. S3). In addition, it can be also inferred that the MNP should not

cover the adsorption sites of CS/LCF during the process of AR18 adsorption.

Secondly, a comparison between mCS/LCF and other previously reported

adsorbents was also conducted and the results are summarized in Table 3

[15,44,5458]. It can be seen that mCS/LCF has exceptionally high adsorption

capacity (up to 1184 mg∙g–1), which is even much higher than that obtained by pure

nanochitosan (828.1 mg∙g–1) [55]. This result indicates that immobilizing CS on LCF

can make the active sites of CS more fully exposed and thereby facilitates the access

21
ACCEPTED MANUSCRIPT

of AR18.

In our previous work, we reported the synthesis of a magnetized CS/CNT

(mCS/CNT) and its use in adsorptive removal of AR18 [7]. The maximum adsorption

T
IP
capacity of mCS/CNT reached 809.9 mg∙g–1, which was significantly lower than that

R
(1184 mg∙g–1) of mCS/LCF. The difference in adsorption capacity should contribute

SC
to the difference in the CS coating. The better adsorption performance of mCS/LCF

should be due to the fact that LCF had a superiority to the others (e.g., CNT) for the

NU
uniform coating with CS. Besides, the LCF and CS used in this work are freely or
MA
abundantly available, so the resultant adsorbent mCS/LCF is expected to be

economically feasible for removal of AR18 from aqueous solutions.


D

3.4. Reusability of adsorbent


TE

The reusability of adsorbent is an important factor related to its application


P

potential. To evaluate the reusability of the mCS/LCF, ten consecutive


CE

adsorptiondesorption cycles were performed. The ammonia/ethanol (v/v, 2/3)


AC

solution was used as elution solvent. After elution, the mCS/LCF was magnetically

retrieved and reused directly for the next cycle. The relationship between cycle times

and adsorption capacity was shown in Fig. 7. It can be found that the removal of

AR18 keeps relatively constant (99.48% to 99.88%) in ten consecutive cycles. This

result indicates that the reusability of mCS/LCF is quite satisfied [44]. These results

also suggest that the mCS/LCF holds great potential for being used as a cost-effective

biosorbent.

4. Conclusions

22
ACCEPTED MANUSCRIPT

A novel CS-decorated LCF (CS/LCF) biocomposite, with a well-defined coating

structure, was successfully prepared via a facile “surface deposition–crosslinking”

method. Benefiting from the cheap raw materials and the facile preparation processes,

T
IP
this biocomposite is easily available, low-cost, and massively productive.

R
Interestingly, when CS/LCF and MNP were dispersed in an AR18 solution, the MNP

SC
could spontaneously adhere to the surface of CS/LCF, leading to the magnetization of

CS/LCF; meanwhile, the AR18 molecules were significantly adsorbed by the

NU
magnetized CS/LCF. Under optimized conditions, the maximum adsorption capacity
MA
of the magnetized CS/LCF reached 1184 mgg–1, which was even much higher than

that obtained by pure nanochitosan (828.1 mgg–1). Results of adsorption kinetic


D

experiments showed that the adsorption proceeded quite quickly (reaching saturation
TE

within 240 min). Furthermore, this novel biosorbent could be easily separated from
P

adsorption medium by addition of an external magnetic field, without the need of


CE

filtration and centrifugation, thus greatly facilitating the adsorption and desorption
AC

processes. Test of reusability demonstrated that the magnetically retrievable CS/LCF

could be repeatedly used without any significant loss in adsorption capacity. Thus,

this technically feasible, highly efficient, and cost-effective biosorbent is very

attractive and implies a potential of practical application for treatment of acidic azo

colorant-polluted wastewaters.

Acknowledgements

The authors acknowledge the research grant provided by the Fundamental

Research Funds for the Central Universities, China University of Geosciences

23
ACCEPTED MANUSCRIPT

(Wuhan) (No. CUG120115), Special Fund for Basic Scientific Research of Central

Colleges, China University of Geosciences (Wuhan) (No. CUGL090305), Land

Resources Geology Survey Projects of China (Grant No. 12120113015300), and

T
IP
National Natural Science Foundation of China (Grant No. 21303170).

R
References

SC
[1] D. Cui, Y.Q. Guo, H.Y. Cheng, B. Liang, F.Y. Kong, H.S. Lee, A.J. Wang, Azo

dye removal in a membrane-free up-flow biocatalyzed electrolysis reactor

NU
coupled with an aerobic bio-contact oxidation reactor, J. Hazard. Mater. 239240
MA
(2012) 257264.

[2] C. Ramírez, A. Saldaña, B. Hernández, R. Acero, R. Guerra, S. GarciaSegura, E.


D

Brillas, J.M. PeraltaHernández, Electrochemical oxidation of methyl orange azo


TE

dye at pilot flow plant using BDD technology, J. Ind. Eng. Chem. 19 (2013)
P

571579.
CE

[3] R. Jiang, H. Zhu, X. Li, L. Xiao, Visible light photocatalytic decolourization of Cl


AC

acid red 66 by chitosan capped CdS composite nanoparticles, Chem. Eng. J. 152

(2009) 537542.

[4] E. Ghasemi, H. Ziyadi, A.M. Afshar, M. Sillanpää, Iron oxide nanofibers: A new

magnetic catalyst for azo dyes degradation in aqueous solution, Chem. Eng. J.

264 (2015) 146151.

[5] M. Fomina, G.M. Gadd, Biosorption: Current perspectives on concept, definition

and application, Bioresour. Technol. 160 (2014) 314.

[6] Z.G. Geng, Y. Lin, X.X. Yu, Q.H. Shen, L. Ma, Z.Y. Li, N. Pan, X.P. Wang,

24
ACCEPTED MANUSCRIPT

Highly efficient dye adsorption and removal: A functional hybrid of reduced

graphene oxideFe3O4 nanoparticles as an easily regenerative adsorbent, J. Mater.

Chem. 22 (2012) 35273535.

T
IP
[7] S. Wang, Y.Y. Zhai, Q. Gao, W.J. Luo, H. Xia, C.G. Zhou, Highly efficient

R
removal of acid red 18 from aqueous solution by magnetically retrievable

SC
chitosan/carbon nanotube: Batch study, Isotherms, kinetics, and thermodynamics,

J. Chem. Eng. Data 59 (2014) 3951.

NU
[8] Q. Gao, H. Zhu, W.J. Luo, S. Wang, C.G. Zhou, Preparation, characterization, and
MA
adsorption evaluation of chitosan-functionalized mesoporous composites, Micro.

Meso. Mater. 193 (2014) 1526.


D

[9] M. Rafatullah, O. Sulaiman, R. Hashim, A. Ahmad, Adsorption of methylene blue


TE

on low-cost adsorbents: A review, J. Hazar. Mater. 177 (2010) 7080.


P

[10] G. Crini, Non-conventional low-cost adsorbents for dye removal: A review,


CE

Bioresour. Technol. 97 (2006) 10611085.


AC

[11] S. Babel, T.A. Kurniawan, Low-cost adsorbents for heavy metal uptake from

contaminated water: A review, J. Hazard. Mater. 97 (2003) 219243.

[12] M. Kumar, B.P. Tripathi, V.K. Shahi, Crosslinked chitosan/polyvinyl alcohol

blend beads for removal and recovery of Cd(II) from wastewater, J. Hazard.

Mater. 172 (2009) 10411048.

[13] M. Ahmaruzzaman, Adsorption of phenolic compounds on low-cost adsorbents: A

review, Adv. Colloid Interface Sci. 143 (2008) 4867.

[14] W.S.W. Ngah, L.C. Teong, M.A.K.M. Hanafiah, Adsorption of dyes and heavy

25
ACCEPTED MANUSCRIPT

metal ions by chitosan composites: A review, Carbohydr. Polym. 83 (2011)

14461456.

[15] W.H. Cheung, Y.S. Szeto, G. McKay, Intraparticle diffusion processes during acid

T
IP
dye adsorption onto chitosan, Bioresour. Technol. 98 (2007) 28972904.

R
[16] K. Li, Q. Gao, G. Yadavalli, X. Shen, M. Li, H.W. Lei, Selective adsorption of

SC
Gd3+ on a magnetically retrievable imprinted chitosan/carbon nanotube composite

with high capacity, ACS Appl. Mater. Interfaces, 7 (2015) 2104721055.

NU
[17] A. Bhatnagar, M. Sillanpää, Utilization of agro-industrial and municipal waste
MA
materials as potential adsorbents for water treatment  A review, Chem. Eng. J.

157 (2010) 277296.


D

[18] M. Maki, K.T. Leung, W. Qin, The prospects of cellulase-producing bacteria for
TE

the bioconversion of lignocellulosic biomass, Int. J. Biol. Sci. 5 (2009) 500516.


P

[19] A. Roy. S. Chakraborty, S. P. Kundu, B. Adhikari, S.B. Majumder, Adsorption of


CE

anioic-azo dye from aqueous solution by lignocellucose-biomass jute fuber:


AC

Equilibrium, kinetics, and thermodynamics study, Ind. Eng. Chem. Res. 51 (2012)

1209512106.

[20] A. Roy. B. Adhikari, S.B. Majumder, Equilibrium, kinetic, and thermodynamic

studies of azo dye adsorption from aqueous solution by chemically modified

lignocellulosic jute fiber, Ind. Eng. Chem. Res. 52 (2013) 65026512.

[21] M.K. Purkait, A. Maiti, S. DasGupta, S. De, Removal of congo red using

activated carbon and its regeneration, J. Hazard. Mater. 145 (2007) 287295.

[22] J.L. Gong, B. Wang, G.M. Zeng, C.P. Yang, C.G. Niu, Q.Y. Niu, W.J. Zhou, Y.

26
ACCEPTED MANUSCRIPT

Liang, Removal of cationic dyes from aqueous solution using magnetic

multi-wall carbon nanotube nanocomposite as adsorbent, J. Hazard. Mater. 164

(2009) 15171522.

T
IP
[23] S. Wang, Q. Gao, W.J. Luo, J. Xu, C.G. Zhou, H. Xia, Removal of methyl blue

R
from aqueous solution by magnetic carbon nanotube, Water Sci. Technol. 68

SC
(2013) 665673.

[24] J. GrzechulskaDamszel, M. Tomaszewska, A.W. Morawski, Integration of

NU
photocatalysis with membrane processes for purification of water contaminated
MA
with organic dyes, Desalination 241 (2009) 118126.

[25] A.M.M. Vargas, A.L. Cazetta, A.C. Martins, J.C.G. Moraes, E.E. Garcia, G.F.
D

Gauze, W.F. Costa, V.C. Almeida, Kinetic and equilibrium studies: Adsorption of
TE

food dyes acid yellow 6, acid yellow 23, and acid red 18 on activated carbon from
P

flamboyant pods, Chem. Eng. J. 181182 (2012) 243250.


CE

[26] Q. Gao, D. Luo, M. Bai, Z.W. Chen, Y.Q. Feng, Rapid determination of estrogens
AC

in milk samples based on magnetite nanoparticles/polypyrrole magnetic

solid-phase extraction coupled with liquid chromatography–tandem mass

spectrometry, J. Agr. Food Chem. 59 (2011) 85438549.

[27] X.W. Chen, W.J. Wang, Z.N. Song, J.H. Wang, Chitosan/carbon nanotube

composites for the isolation of hemoglobin in the presence of abundant proteins,

Anal. Methods 3 (2011)17691773.

[28] P. Zheng, J. Wang, C. Lu, Y. Xu, Z.H. Sun, Immobilized -glucosidase on

magnetic chitosan microspheres for hydrolysis of straw cellulose, Process

27
ACCEPTED MANUSCRIPT

Biochem. 48 (2013) 683687.

[29] M. Li, S. Wang, W.J. Luo, H. Xia, Q. Gao, C.G. Zhou, Facile synthesis and in situ

magnetization of carbon-decorated lignocellulose fiber for highly efficient

T
IP
removal of methylene blue, J. Chem. Technol. Biotechnol. 90 (2015) 11241134.

R
[30] V. Tserki, P. Matzinos, S. Kokkou, C. Panayiotou, Novel biodegradable

SC
composites based on treated lignocellulosic waste flour as filler. Part I. Surface

chemical modification and characterization of waste flour, Compos. Part A 36

(2005) 11101118.
NU
MA
[31] K.L. Pickering, G.W. Beckermann, S.N. Alam, N.J. Foreman, Optimising

industrial hemp fiber for composites, Compos. Part A 38 (2007) 461468.


D

[32] Y.Q. Chen, L.B. Chen, H. Bai, L. Li, Graphene oxide–chitosan composite
TE

hydrogels as broad-spectrum adsorbents for water purification, J. Mater. Chem. A


P

1 (2013) 19922001.
CE

[33] S.F. Wang, L. Shen, Y.J. Tong, L. Chen, I.Y. Phang, P.Q. Lim, T.X. Liu,
AC

Biopolymer chitosan/montmorillonite nanocomposites: Preparation and

characterization, Polym. Degrad. Stab. 90 (2005) 123131.

[34] J.G. Liu, X.M. Jiang, L.S. Zhou, X.X. Han, Z.G. Cui, Pyrolysis treatment of oil

sludge and model-free kinetics analysis, J. Hazard. Mater. 161 (2009) 12081215.

[35] J. Brugnerotto, J. Lizardi, F.M. Goycoolea, W. ArgüellesMonal, J. Desbrières, M.

Rinaudo, An infrared investigation in relation with chitin and chitosan

characterization, Polymer 42 (2001) 35693580.

[36] M.M. Ibrahim, A. Dufresne, W.K. El-Zawawy, F.A. Agblevor, Banana fibers and

28
ACCEPTED MANUSCRIPT

microfibrils as lignocellulosic reinforcements in polymer composites, Carbohydr.

Polym. 81 (2010) 811819.

[37] F.H. Li, H.J. Hu, R.S. Yao, H. Wang, M.M. Li, Structure and saccharification of

T
IP
rice straw pretreated with microwave-assisted dilute lye, Ind. Eng. Chem. Res. 51

R
(2012) 62706274.

SC
[38] H. Yoshida, A. Okamoto, T. Kataoka, Adsorption of acid dye on cross-linked

chitosan fibers: Equilibria, Chem. Eng. Sci. 48 (1993) 22672272.

NU
[39] M.S. Chiou, H.Y. Li, Equilibrium and kinetic modeling of adsorption of reactive
MA
dye on cross-linked chitosan beads, J. Hazard. Mater. 93 (2002) 233248.

[40] C. Jeon, Y.J. Yoo, W.H. Hoell, Environmental effects and desorption
D

characteristics on heavy metal removal using carboxylated alginic acid, Bioresour.


TE

Tehnol. 96 (2005) 1519.


P

[41] C. Jeon, K.H. Park, Adsorption and desorption characteristics of mercury(II) ions
CE

using aminated chitosan bead, Water Res. 39 (2005) 39383944.


AC

[42] A. Kurniawan, H. Sutiono, Y.H. Ju, F.E. Soetaredjo, A. Ayucitra, A. Yudha, S.

Ismadji, Utilization of rarasaponin natural surfactant for organo-bentonite

preparation: Application for methylene blue removal from aqueous effluent,

Micro. Meso. Mater. 142 (2011) 184193.

[43] S. Ghorai, A.K. Sarkar, A. Panda, S. Pal, Effective removal of Congo red dye

from aqueous solution using modified xanthan gum/silica hybrid nanocomposite

as adsorbent, Bioresour. Technol. 144 (2013) 485491.

[44] N. Li, X.M. Lei, Adsorption of ponceau 4R from aqueous solutions by

29
ACCEPTED MANUSCRIPT

polyamidoamine-cyclodextrin crosslinked copolymer, J. Incl. Phenom. Macrocycl.

Chem. 74 (2012) 167176.

[45] R. Sips, Combined form of Langmuir and Freundlich equations, J. Chem. Phys.

T
IP
16 (1948) 490495.

R
[46] W.X. Zhang, H.J. Li, X.W. Kan, L. Dong, H. Yan, Z.W. Jiang, H. Yang, A.M. Li,

SC
R.S. Cheng, Adsorption of anionic dyes from aqueous solutions using chemically

modified straw, Bioresour. Technol. 117 (2012) 4047.

NU
[47] G.L. Dotto, E.C. Lima, L.A.A. Pinto, Biosorption of food dyes onto spirulina
MA
platensis nanoparticles: Equilibrium isotherm and thermodynamic analysis,

Bioresour. Technol. 103 (2012) 123130.


D

[48] S. Elemen, E.P.A. Kumbasar, S. Yapar, Modeling the adsorption of textile dye on
TE

organoclay using an artificial neural network, Dyes Pigments 95 (2012) 102111.


P

[49] Q. Lei, H.F. Zhang, Green synthesis of chitosan-based nanofibers and their
CE

applications, Green Chem. 12 (2010) 12071214.


AC

[50] S.A. Ntim, S. Mitra, Removal of trace arsenic to meet drinking water standards

using iron oxide coated multiwall carbon nanotubes, J. Chem. Eng. Data 56 (2011)

20772083.

[51] A.B. Albadarin, C. Mangwandi, A.H. Al-Muhtaseb, G.M. Walker, S.J. Allen,

M.N.M. Ahmad, Kinetic and thermodynamics of chromium ions adsorption onto

low-cost dolomite adsorbent, Chem. Eng. J. 179 (2012) 193202.

[52] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, J. Sanit.

Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 3159.

30
ACCEPTED MANUSCRIPT

[53] G.L. Dotto, M.L.G. Vieira, L.A.A. Pinto, Kinetics and mechanism of tartrazine

adsorption onto chitin and chitosan, Ind. Eng. Chem. Res. 51 (2012) 68626868.

[54] S.P. Zhao, F. Zhou, L.Y. Li, M.J. Cao, D.Y. Zuo, H.T. Liu, Removal of anionic

T
IP
dyes from aqueous solutions by adsorption of chitosan-based semi-IPN hydrogel

R
composites, Compos. Part B: Eng. 43 (2012) 15701578.

SC
[55] W.H. Cheung, Y.S. Szeto, G. McKay, Enhancing the adsorption capacities of acid

dyes by chitosan nanoparticles, Bioresour. Technol. 100 (2009) 11431148.

[56] R.M. Cheng, B. Xiang, Y.J.


NU
Li, M.Z. Zhang, Application of
MA
dithiocarbamate-modified starch for dyes removal from aqueous solutions J.

Hazard. Mater. 188 (2011) 254260.


D

[57] X.F. Sun, S.G. Wang, W. Cheng, M.H. Fan, B.H. Tian, B.Y. Gao, X.M. Li,
TE

Enhancement of acidic dye biosorption capacity on poly(ethylenimine) grafted


P

anaerobic granular sludge, J. Hazard. Mater.189 (2011) 2733.


CE

[58] R.M. Cheng, S.J. Ou, B. Xiang, Y.J. Li, Q.Q. Liao, Equilibrium and molecular
AC

mechanism of anionic dyes adsorption onto copper (II) complex of

dithiocarbamate-modified starch, Langmuir 26 (2009) 752758.

Figure Captions

Fig. 1 Principle and process of mCS/LCF formation and its application for AR18

removal.

Fig. 2 SEM images of LCF ((a), (d)), CS/LCF ((b), (e)), and mCS/LCF ((c), (f)).

31
ACCEPTED MANUSCRIPT

Fig. 3 XRD patterns for MNP, LCF, CS, CS/LCF, and mCS/LCF (a); TGA and DTG

of CS, LCF, and CS/LCF, respectively (b); FT-IR spectra of LCF and CS/LCF, and an

enlarged view of the 1500–1760 cm–1 region (c).

T
IP
Fig. 4 Effects of pH (a), KCl concentration (b), temperature (c), and contact time (d)

R
on the adsorption efficiency of AR18 onto mCS/LCF.

SC
Fig. 5 Adsorption isotherm data of AR18 onto mCS/LCF fitting to various isotherm

equations at temperatures 303.15 K (a), 323.15 K (b), and 343.15 K (c).

NU
Fig. 6 Nonlinear fitting of pseudo-first-order and pseudo-second-order models (a),
MA
and plots of Weber–Morris diffusion model (b) for the adsorption of AR18 onto

mCS/LCF.
D

Fig. 7 Effect of recycle number on the adsorption efficiency of AR18 onto mCS/LCF.
P TE
CE
AC

32
ACCEPTED MANUSCRIPT

Table 1 Adsorption isotherm constants for AR18 adsorption onto mCS/LCF at


different temperatures.
T (K)
Isotherm model Model parameter
303.15 323.15 343.15
−1
Langmuir qm (mg∙g ) 1160.5 1138.8 1118.5

T
−1
KL (L∙mg ) 1.48 2.56 1.36
2
R 0.9511 0.8981 0.9788

IP
1–1/n 1/n –1
Freundlich KF (mg ·L ·g ) 762.8 789.6 723.9
1/n 0.106 0.0932 0.106

R
2
R 0.8450 0.8633 0.8853

SC
−1
Sips qm (mg∙g ) 1219.6 1216.8 1165.8
–1
Ks(L∙mg ) 1.56 2.80 1.42
1/ns 0.691 0.576 0.707

NU
R2 MA 0.9724 0.9575 0.9946
D
P TE
CE
AC

33
ACCEPTED MANUSCRIPT

Table 2 Kinetic parameters for adsorption of AR18 onto mCS/LCF at different initial
concentrations.
Initial concentration (mg∙L−1)
Kinetic model Model parameter
300 400 500

T
−1
qe,exp (mg∙g ) 605.6 794.5 996.5
−1
Pseudo-first-order k1 (min ) 0.03516 0.06790 0.07909

IP
−1
qe,calc,1 (mg∙g ) 576.2 767.5 934.8
2
R 0.9522 0.9836 0.9062

R
−1 −1
Pseudo-second-order k2 (g∙mg ∙min ) 0.0000818 0.000125 0.000115

SC
−1
qe,calc,2 (mg∙g ) 617.4 809.8 994.5
2
R 0.9813 0.9921 0.9831
Weber–Morris model ki,1 (mg∙g−1∙min−0.5) 56.11 92.35 85.30

NU
−1
I1 (mg∙g ) 33.85 75.74 228.56
2
R 0.9962 0.9366 0.9027
−1 −0.5
ki,2 (mg∙g ∙min ) 5.21 1.79 11.61
MA
−1
I2 (mg∙g ) 487.93 743.27 774.23
2
R 0.9936 0.9787 0.9436
D
P TE
CE
AC

34
ACCEPTED MANUSCRIPT

Table 3 Comparison of adsorption capacities between various adsorbents.


Adsorbent qm (mg∙g–1) T (K) Refs
Chitosan-based semi-IPN hydrogel composites 358.4 298.15 [54]
Polyamidoamine–cyclodextrin crosslinked copolymer 266.8 303.15 [44]
Chitosan 693.2 298.15 [15]

T
Nanochitosan 828.1 298.15 [55]

IP
Dithiocarbamate-modified starch 271.41 298.15 [56]
Modified anaerobic granule 418.38 293.15 [57]

R
Copper complex of dithiocarbamate-modified starch 107.0 298.15 [58]

SC
mCS/LCF 1184 303.15 This work

NU
MA
D
P TE
CE
AC

35
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA

Fig. 1 Principle and process of mCS/LCF formation and its application for AR18
removal.
D
P TE
CE
AC

36
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
Fig. 2 SEM images of LCF ((a), (d)), CS/LCF ((b), (e)), and mCS/LCF ((c), (f)).
MA
D
P TE
CE
AC

37
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA
D

Fig. 3 XRD patterns for MNP, LCF, CS, CS/LCF, and mCS/LCF (a); TGA and DTG
TE

of CS, LCF, and CS/LCF, respectively (b); FT-IR spectra of LCF and CS/LCF, and an
enlarged view of the 1500–1760 cm–1 region (c).
P
CE
AC

38
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA

Fig. 4 Effects of pH (a), KCl concentration (b), temperature (c), and contact time
(d) on the adsorption efficiency of AR18 onto mCS/LCF.
D
P TE
CE
AC

39
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA
D
P TE
CE
AC

Fig. 5 Adsorption isotherm data of AR18 onto mCS/LCF fitting to various


isotherm equations at temperatures 303.15 K (a), 323.15 K (b), and 343.15 K (c).

40
ACCEPTED MANUSCRIPT

T
R IP
SC
Fig. 6 Nonlinear fitting of pseudo-first-order and pseudo-second-order models
(a), and plots of Weber–Morris diffusion model (b) for the adsorption of AR18 onto

NU
MA mCS/LCF.
D
P TE
CE
AC

41
ACCEPTED MANUSCRIPT

T
R IP
SC
Fig. 7 Effect of recycle number on the adsorption efficiency of AR18 onto mCS/LCF.

NU
MA
D
P TE
CE
AC

42

Вам также может понравиться