Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Journal 339 (2018) 322–333

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Enhanced copper adsorption by DTPA-chitosan/alginate composite beads: T


Mechanism and application in simulated electroplating wastewater

Yanjun Huang, Hailing Wu, Taikang Shao, Xia Zhao, Hong Peng , Yuefa Gong, Huihai Wan
Hubei Key Laboratory of Bioinorganic Chemistry & Materia Medica, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology,
Wuhan 430074, PR China

H I G H L I G H T S

• ACa-DTCS/ALG
novel DTPA-chitosan/alginate composite adsorbent (Ca-DTCS/ALG beads) was prepared.
• Na , Ca andshowed high copper removal efficiency in strongly acidic solutions.
• Ca-DTCS/ALG could
+ 2+
Al 3+
had little effect on copper removal efficiency by Ca-DTCS/ALG.
• DTPA modified chitosan
recover copper from simulated electroplating wastewater high efficiently.
• (DTCS) reached copper adsorption equilibrium within 1 min.

A R T I C L E I N F O A B S T R A C T

Keywords: A novel composite adsorbent of Ca-DTCS/ALG bead was prepared by immobilization of DTPA-modified chitosan
DTPA-modified chitosan micro-gels (DTCS) into the alginate matrix. The copper sorption properties of Ca-DTCS/ALG beads were tested
Alginate and compared with the other two adsorbents of DTCS powder and plain alginate (ALG) beads. In addition, the
Composite beads mechanisms of copper sorption on the adsorbents were investigated by FTIR and XPS analysis. Similar to DTCS
Encapsulation
powder, Ca-DTCS/ALG beads exhibited high copper removal efficiency in strongly acidic solutions, whereas
Removal of copper ions
calcium, sodium and aluminum salts had little effect on its copper adsorption capacities. The maximum ad-
Simulated electroplating wastewater
sorption capacities of Ca-DTCS/ALG and DTCS were 107.5 and 106.0 mg g−1 at pH 3.0, respectively, which were
close to the values predicted by the Sips model. However, the maximum adsorption capacity of ALG beads
predicted by the Sips model was much higher than its experimental one (101.1 mg g−1). The copper adsorption
equilibrium by DTCS powder was achieved within 1 min, nevertheless, ALG and Ca-DTCS/ALG gel beads
reached the adsorption equilibrium within 500 min and 1440 min, respectively. Moreover, Ca-DTCS/ALG gel
beads showed good reusability, but DTCS powder was not easily reused due to its notable swelling property.
Furthermore, the efficiency of copper recovery by Ca-DTCS/ALG beads in the simulated electroplating waste-
water at the copper concentration of 1–5 mg L−1 was much higher than that of DTCS and ALG beads. These
results reveal that immobilization of DTCS into the alginate matrix greatly improves the adsorption and se-
paration performance of the adsorbent.

1. Introduction functionalized chitosan exhibits excellent uptake properties for heavy


metals in strongly acidic solutions (pH 1.5–3.0) [3], and could capture
Chitosan is one of the most widely studied bio-sorbents due to its of Co(II) from its aqueous EDTA-chelate [7]. In addition, DTPA-chit-
advantages of abundant sources, low cost, plenty of reactive groups and osan/silica could be used as a stationary phase for the separation of
biodegradability [1,2]. In order to improve its adsorption performance rare-earth ions by medium pressure liquid chromatography (MPLC) [8].
for heavy metals, chitosan was modified with various aminopoly- The preparation of DTPA- modified chitosan was usually conducted
carboxylic acids chelating agents such as ethylenediaminetetraacetic by the reaction of chitosan with DTPA bis-anhydride in a mixed solvent
acid (EDTA) [3,4], diethylenetriaminepentaacetic acid (DTPA) [3], of methanol and acetic acid solution [3,8]. However, the product of this
ethylene glycol-bis(2-aminoethylether)-N, N, N', N'-tetraacetic acid reaction was not easily separated from aqueous solution by filtration or
(EGTA) [5], and nitrolotriacetic acid (NTA) [6]. Among them, DTPA- centrifugation due to its notable swelling property [8,9]. In order to


Corresponding author.
E-mail address: penghong811@hust.edu.cn (H. Peng).

https://doi.org/10.1016/j.cej.2018.01.071
Received 10 November 2017; Received in revised form 11 January 2018; Accepted 11 January 2018
Available online 02 February 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

solve this problem, magnetic DTPA-cross-linked-chitosan was devel- (4) recovery of copper ions from simulated electroplating wastewater
oped to remove heavy metals from aqueous solutions [9]. In addition, by DTPA-chitosan/alginate composite beads.
this problem may also be overcome by an encapsulation method for the
immobilization of micrometer-sized DTPA-chitosan particles in a sui-
table matrix. Alginate is a natural polysaccharide extracted from brown 2. Materials and methods
seaweeds, which has a large number of free hydroxyl and carboxyl
groups distributed along the backbone as coordination and reaction 2.1. Materials
sites. Thus, alginate is not only highly efficient bio-sorbents for heavy
metal ions [10], but also an excellent encapsulated biopolymer for Chitosan (BR, 80–95% deacetylated, and the viscosity average mo-
nano- or micro-sized adsorbents. Various heavy metal adsorbents such lecular weight measured by Ubbelohde viscometer is 3.69 × 105), so-
as nanohydroxyapatite [11], hydrous MnO2 [12], magnetite graphene dium alginate (CP, viscosity (10 g L−1) > 0.02 pa.s) and glutaraldehyde
oxide [13], PEI derivatives [14], and biomass [15] were immobilized in solution (BR, 25 wt%) were purchased from Sinopharm Chemical
the alginate matrix. These alginate composites exhibited enhanced Reagent Co., Ltd. (China). Diethylenetriaminepentaacetic acid (DTPA,
adsorption and separation performance during heavy metal wastewater AR), 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride
treatment. (EDAC, 98%), and N-hydroxysuccinimide (NHS, 98%) were bought
In this study, DTPA-modified chitosan micro-gels were firstly pre- from Aladdin Chemical Co., Ltd. (China). Copper standard solutions
pared by amidation reaction between the carboxylic acid groups of (1000 mg L−1) were supplied by National Nonferrous Metals and
DTPA and the amino groups of chitosan. In the reaction process, chit- Electronic Materials Analysis and Testing Center. All other chemicals
osan was dissolved in the aqueous solution of DTPA by using NaOH to were commercially available from Sinopharm Chemical Reagent Co.,
adjust the solution pH value to 5–6, which resulted in the formation of Ltd. (China), all of which were analytical grade and used without fur-
the ionic bonds between the carboxylic groups of DTPA and the pro- ther purification. All water mentioned in this paper was deionized
tonated amino groups of chitosan directly. The close contact of the water (18.2 MΩ cm).
carboxylic groups with the amino groups contributed to the subsequent
amidation reactions. Moreover, the water-soluble 1-ethyl-3-(3-di-
methylaminopropyl) carbodiimide hydrochloride (EDAC) and N-hy- 2.2. Synthesis of DTPA-modified chitosan
droxysuccinimide (NHS) were employed together as the cross-linkers in
the amidation reaction, in which the effective NHS-ester was formed as 2 g of chitosan and 7.32 g of DTPA were dispersed in 100 mL of
the reactive intermediate to react with amino groups of chitosan. The deionized water, and then the solution pH value was adjusted to and
synergistic effect of the two crosslinking agents of EDAC and NHS sig- kept at 5–6 with a small amount of sodium hydroxide. At the same time,
nificantly promoted the formation of DTPA-chitosan micro-gels [16]. the solution was continuously stirred until all the chitosan and DTPA
However, the DTPA-chitosan micro-gels had an excellent dispersibility were completely dissolved. Afterwards, 60 mL of cool aqueous solution
in aqueous solution and were difficult to be separated from aqueous containing 5.23 g of EDAC and 0.79 g of NHS was added dropwise into
solutions. Therefore, in order to enhance the separability of DTPA- 60 g of chitosan/DTPA solution (pH 5 – 6, chitosan 1.1 g, DTPA 4.02 g)
chitosan micro-gels in the removal of heavy metals from aqueous so- under the ice-water bath (0–5 °C), and then the mixture was mechani-
lutions, DTPA-chitosan/alginate composite beads were prepared by cally stirred for 12 h at room temperature. The resulting micro-gels
immobilization of DTPA-chitosan micro-gels in the alginate matrix. This were washed with disodium hydrogen phosphate solution (pH 8.0,
work has rarely been reported in the literature. 0.2 M) for 30 min by magnetic stirring, and then washed with deionized
Copper is one of the common heavy metal contaminants in aqueous water several times until the pH value of the rinse water was close to
solutions, which is usually coexisted with a variety of complexing neutral. Finally, the white DTPA-modified chitosan micro-gel (12.56 g)
agents in acidic wastewater discharged from electroplating and printed was obtained after filtration, which was abbreviated as DTCS micro-gel.
circuit board industries. Thus, this study aims to investigate: (1) pre- In addition, 12.56 g of DTCS micro-gel was dried at 45 °C overnight to
paration and characterization of DTPA-chitosan/alginate composite give 1.81 g of the solid, and then the dried DTCS was ground into a
beads; (2) the effects of various factors on removal of copper ions from powder. The process for the synthesis of DTCS micro-gels and its che-
aqueous solutions by DTPA-chitosan/alginate composite beads; (3) the lating reaction with copper ions are shown in Scheme 1.
adsorption mechanisms for DTPA-chitosan/alginate composite beads;

Scheme 1. Synthesis of DTCS and its proposed chelating


reaction with copper ions.

323
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

Scheme 2. Procedure for the preparation of Ca-DTCS/ALG


beads.

2.3. Preparation of DTPA-chitosan/alginate composite beads The kinetic studies were carried out in a series of centrifuge tubes
with lids. 10 mg of powder or gel beads (equivalent to 10 mg dry
The DTPA-chitosan/alginate composite beads were prepared by the weight) were dispersed in 10.0 mL of copper solution in each tube.
following procedure: the as-synthesized DTCS micro-gels were thor- Then, the tubes were kept on a rotary shaker with thermostatic control
oughly dispersed into a sodium alginate solution (1.5 wt%) and me- at 28 °C for a specified contact time. The initial concentration of copper
chanically stirred for 12 h. Three different ratios of DTCS gels to sodium solution was 100 mg L−1 and the solution initial pH value was 3.0. At
alginate (0.5:1, 1:1, and 2:1, dry weight/dry weight) were tested in the designed time intervals, one tube was taken out and the solution sample
preparation of DTPA-chitosan/alginate composite beads. A homo- was filtrated with 0.45 μm filter. After dilution with 2% HNO3, the
geneous DTCS/alginate suspension was then dropwise added to 0.1 M residual metal concentrations in the solutions were analyzed by the
CaCl2 solution using a syringe to instantly form the DTCS/alginate atomic absorption spectrometer (AAS, Thermo Fisher Scientific, icE
composite hydrogel beads. The beads were cured in CaCl2 solution for 3000 Series, USA). To explore the isotherms, the batch adsorption was
12 h to ensure the complete gelation reaction. Afterwards, the hydrogel operated at the copper concentrations varying from 20 to 500 mg L−1 at
beads were washed with deionized water five times, and then immersed pH 3.0. All adsorption tests were carried out in duplicate. The ad-
into a mixture solution containing 2.5 wt% glutaraldehyde and 0.1 M sorption capacities (qe, mg g−1) by the adsorbents were calculated as
HCl for 17 h at room temperature with continuously shaking. After the follows:
crosslinking reaction, the beads were thoroughly washed with water to C0−Ce
remove unreacted glutaraldehyde and excessive hydrogen ions. The qe = V
m (1)
obtained composite beads were abbreviated as Ca-DTCS/ALG beads.
−1
The procedure for the preparation of Ca-DTCS/ALG beads is shown in where C0 and Ce (mg L ) are the initial and the equilibrium con-
Scheme 2. The plain alginate beads (ALG beads) were prepared in the centrations of the copper ions, respectively. V (L) is the volume of the
same procedure as the Ca-DTCS/ALG beads except for DTCS micro-gel solution whereas m (g) represents the weight of the adsorbent.
addition. In order to investigate the effect of the other inorganic ions (Na+,
Ca2+, Fe2+ and Al3+) on the removal efficiency of copper (II) ion, the
2.4. Characterization of the adsorbents experiments followed the same procedure as the adsorption tests
mentioned above. The initial concentration of copper ion solution was
The scanning electron microscope (SEM, Hitachi-S4800, Japan) was 100 mg L−1 (pH = 3), into which Na+ (ranging from 50 mg L−1 to
applied to record the morphology of the adsorbents. The functional 1000 mg L−1), Ca2+, Fe2+, and Al3+ (all ranging from 50 mg L−1 to
groups on the adsorbents were characterized by the Fourier transform 400 mg L−1) was added, respectively.
infrared spectrometer (FTIR, Equinox 55, Bruker Optics, Germany). The
elemental analyses for chitosan after DTPA modification were carried 2.6. Regeneration studies
out with an Elemental analyzer (Vario Micro cube Elementar,
Germany). The surface chemical composition was determined by the X- To evaluate the regeneration and reuse of the spent Ca-DTCS/ALG
ray photoelectron spectroscopy (XPS, AXIS-ULTRA DLD-600W, Kratos, and ALG, the gel beads (equivalent to 10 mg dry weight) were loaded
Japan) with copper ions in 10.0 mL of Cu(II) ion solution (100 mg L−1,
pH = 3.0) agitated at 150 rpm on a shaker at 28 °C for 36 h. The copper
2.5. Batch adsorption studies ion-loaded adsorbent was filtered and washed with water to remove the
un-adsorbed copper ions, then agitated with 10.0 mL of elution solution
The adsorption behavior of Cu(II) ions on adsorbents was in- (2 M HNO3) for another 24 h. Afterwards, the regenerated gel beads
vestigated through batch adsorption studies. In the general batch ex- were filtered and washed with deionized water several times for reuse.
periments, 10 mg of DTCS powder and 10 mL of the solution containing The regeneration of the adsorbent was sequentially conducted 5 times
copper ions were added to the centrifuge tube with lid and agitated at in this way.
150 rpm on a shaker. For the investigation of the adsorbents of Ca-
DTCS/ALG or ALG, the gel beads (equivalent to 10 mg dry weight) were 2.7. Recovery of copper from simulated electroplating wastewater
added to 10 mL of the copper solution. The effect of pH on the ad-
sorption was conducted at the Cu(II) concentration of 100 mg L−1 in the In this work, the simulated electroplating wastewater was prepared
pH range of 1.5–5 at 28 °C for 36 h. The initial pH of copper solution according to the test results of an actual copper electroplating waste-
was adjusted by using 0.1 M HNO3 and/or 0.1 M NaOH aqueous solu- water given by an environmental monitoring center in Guangdong
tions. After adsorption, the final solution pH was also measured to in- Province (China). The detailed composition of the chelating agents or
vestigate the interaction between the adsorbent and the solutes. ligands in the simulated electroplating wastewater is listed in Table S1,

324
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

beads (equivalent to 20 mg dry weight) were added into 10 mL of si-


mulated wastewater (pH = 3.0), and agitated at 150 rpm on a shaker at
28 °C for 36 h. Afterwards, the solution samples were filtrated with
0.45 μm filters, and the residual copper concentrations in the solutions
were analyzed by inductively coupled plasma/optical emission spec-
trometry (ICP-OES, Perkin-Elmer, Optima 8300, USA).

3. Results and discussion

3.1. Characterization of DTPA-chitosan/alginate composite beads

The DTPA-chitosan/alginate composite beads (Ca-DTCS/ALG) were


prepared by encapsulating DTPA-chitosan micro-gels (DTCS) into the
matrix of alginate gel beads (ALG). The FTIR analysis of DTCS powder,
ALG and Ca-DTCS/ALG beads is depicted in Fig. 1. For DTCS powder,
the absorption peaks at 1587 and 1402 cm−1 are attributed to the
asymmetrical and symmetrical stretching vibrations of the carboxylic
groups of DTPA bound to chitosan [17], respectively. The enhanced
absorption peak appearing at 1644 cm−1 represents the carbonyl
Fig. 1. FTIR spectra of DTCS powder, ALG and Ca-DTCS/ALG beads (DTCS/ stretching vibration of the amide bonds, and the peak at 1587 cm−1 is
Alginate = 0.5:1, 1:1, 2:1). also related to the NeH bending vibration of amide [18]. These results
indicate that DTPA was successfully modified on chitosan through
in which the concentration of copper (II) ion varied from 1 to 30 mg amide bonds. In the case of ALG beads, the intensity of the absorption
L−1. In order to investigate of the efficiency of copper recovery by the peak at 1742 cm−1 (assigned to the carbonyl stretch of –COOH group of
adsorbent in the simulated wastewater, 20 mg of the adsorbent or gel alginic acid) is much stronger than that of the peak at 1629 cm−1

Fig. 2. SEM images of (A) DTCS powder; (B) ALG beads; (C)
Ca-DTCS/ALG beads (DTCS/Alginate = 0.5:1); (D) Ca-
DTCS/ALG beads (DTCS/Alginate = 1:1); (E) and (F) Ca-
DTCS/ALG beads (DTCS/Alginate = 2:1).

325
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

(assigned to the asymmetrical stretching vibration of the carboxylate


groups of alginate), suggesting that there are a large number of –COOH
groups and a small amount of calcium carboxylate in ALG beads. In
addition, the peak at 1033 cm−1 might be attributed to the ether
linkage formed by the acetal reaction between glutaraldehyde and al-
coholic hydroxyl group of alginate [19], which might overlap with the
peak assigned to the carbon-oxygen stretching vibration of alcoholic
hydroxyl group of alginate [20].
For Ca-DTCS/ALG beads, the relative intensity of the absorption
peak appearing at about 1733–1735 cm−1 (assigned to the carbonyl
stretch of –COOH group) decreases with the increasing of the ratio of
DTCS to ALG in the composite beads from 0.5:1 to 2:1, whereas the
relative intensity of the absorption peak appearing at about
1631–1635 cm−1 increases, indicating that the carboxylic groups of
DTCS immobilized in the composite beads were more easily chelated
with calcium ions than that of ALG due to its higher Ca(II)-chelating
ability in strongly acidic condition (0.1 M HCl). The XPS analysis of Ca-
DTCS/ALG beads further confirms the existing of calcium element in
the adsorbent, which is illustrated in Section 3.3.
Fig. 2 shows the SEM images of DTCS powder, ALG and Ca-DTCS/
ALG beads. It is observed from Fig. 2A that DTCS powder is an ag-
gregate of submicron particles with a size of about 0.5 μm and has a
porous structure. However, the surface morphology of ALG beads
(Fig. 2B) is nonporous and with wrinkles and cracks. As presented in
Fig. 2C–E, the SEM images of Ca-DTCS/ALG beads show that the
granular polymer (DTCS) is immobilized in the matrix of ALG beads,
and this phenomenon is more obviously observed with the increasing of
the ratio of DTCS to ALG in the composite adsorbents from 0.5:1 to 2:1.
Compared with the DTCS powder, the size of the granular polymer
encapsulating in the matrix (DTCS/ALG = 2:1) does not vary sig-
nificantly, which is also about 0.5 μm (Fig. 2F).

3.2. Copper adsorption properties of DTPA-chitosan/alginate composite


adsorbent

3.2.1. Effect of initial pH on the adsorption


The pH of the aqueous solution is an important operational para-
meter in the adsorption process, since pH values affect not only pro-
tonation of the functional groups of the adsorbent, but also the spe-
Fig. 3. (A) Effect of pH on adsorption of Cu2+ ions by DTCS powder, ALG and Ca-DTCS/
ciation of heavy metal ions in the solution. The pH values in the ALG beads (DTCS/Alginate = 0.5:1, 1:1, 2:1); (B) Comparison of pH values of the copper
following experiments were set below 5.5 due to the probability of the solutions before and after copper adsorption (Initial concentration of copper ions:
formation of metal hydroxide precipitates for Cu(II) ions at higher pH 100 mg L−1; Contact time: 36 h; Dose: 1 g L−1; Temperature: 28 °C).
values. The effect of pH on the removal efficiency of Cu(II) ions by
DTCS powder, ALG and Ca-DTCS/ALG beads is presented in Fig. 3A. In DTCS/ALG beads (DTCS/ALG = 2:1) was mainly investigated in the
the case of DTCS powder, it can be seen that the removal efficiency following experiments. Moreover, it is well known that most of the
increases from 65% to 80% as the pH value increasing from 1.5 to 2.5, practical effluents containing copper ions are strongly acidic, therefore,
and then gently reach a plateau at pH beyond 3.0, which presents a pH 3.0 was selected for the subsequent experiments [5].
similar trend as that obtained from EDTA, DTPA or EGTA-modified The pH value of copper solutions before and after copper adsorption
chitosan materials [3–5]. For ALG gel beads, it is found that the re- on the adsorbents is shown in Fig. 3B. It is found that the final pH
moval efficiency is very low (< 15%) at pH ⩽ 2.0, but rises sharply as values of solutions all decreased to 2.5–3.1 after copper adsorption by
the pH increasing from 2 to 3, and then increase slightly at pH beyond the adsorbents in the initial pH range of 4–6, suggesting that the ion-
3.0, similar results were observed from the other alginate based ad- exchange process between copper ions in aqueous solution and H+ ions
sorbents [21,22]. on the adsorbents was happened.
Compared with ALG beads, DTCS powder exhibits much higher
copper removal efficiency in the wide pH range. This can be explained
by the fact that the DTPA groups attached to DTCS exhibit strong 3.2.2. Uptake kinetics study
chelating ability for Cu(II) ions in the wide pH range [18]. Therefore, The effect of contact time on the adsorption of copper (II) ions by
the greater the content of DTCS immobilized in the alginate matrix, the DTCS powder, ALG and Ca-DTCS/ALG gel beads is shown in Fig. 4. It
higher the copper removal efficiency by the Ca-DTCS/ALG beads in the can be seen that the adsorption equilibrium for DTCS powder was
wide pH range. As presented in Fig. 3A, the removal efficiency by Ca- achieved within 1 min (Fig. S1), and its copper adsorption rate was as
DTCS/ALG beads (DTCS/ALG = 2:1) attains 52% at very low pH values fast as that of chitosan modified magnetic nanoparticles [23], which
(pH 1.5–2.0), and gradually increases to 68% in the pH range of was much faster than that of EDTA, DTPA and EGTA-chitosan reported
2.0–3.5, then reaches a plateau at pH beyond 3.5. Obviously, Ca-DTCS/ by the Refs. [3–5]. Such a fast adsorption rate might be attributed to its
ALG beads (DTCS/ALG = 2:1) show excellent tolerance to low pH va- excellent dispersibility in aqueous solutions and the absence of internal
lues in comparison with ALG and the other two Ca-DTCS/ALG beads diffusion resistance [23]. However, ALG and Ca-DTCS/ALG gel beads
(DTCS/ALG = 1:1 or 0.5:1). Thus, the adsorption performance of Ca- reached the adsorption equilibrium within 500 min and 1440 min,

326
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

The pseudo–first order model:


qt = qe [1−exp−k1 t ] (2)

The pseudo–second order model:

qe2 k2 t
qt =
1 + k2 qe t (3)

The intra–particle diffusion kinetic model:

qt = ki t 1/2 + c (4)
−1
where qt and qe (mg g ) are the adsorption capacity of copper ions
onto the adsorbent at any time t and at equilibrium, respectively. The
parameters of k1 (min−1) and k2 (g mg−1 min−1) are the rate constants
of the pseudo-first order and pseudo-second order models, respectively.
ki (mg g−1 min−1/2) is the intra–particle diffusion rate constant. C is
another constant of the intra–particle diffusion model.
In this study, no-linear fitting method was used to obtained the
kinetic parameters, and the root mean squared error (RMSE) was em-
ployed to ascertain the suitability of kinetic model for describing the
adsorption process. In principle, the lower the RMSE value, the better
the model fits [17]. RMSE could be calculated using the following
equation:
n
1
RMSE =
n
∑ (qcali−qexpi)2
i=1 (5)

where qcali and qexpi are the predicted and measured values of the ad-
sorption capacity at time t, respectively, and n is the number of ex-
perimental data.
The non-linear fitting plots of the pseudo-first and pseudo-second
order models are given in Fig. 4A, and the kinetic parameters of the two
models and their RMSE values are summarized in Table 1. For ALG
beads, the RMSE value of the pseudo-second order model is observed to
be lower than that of the pseudo-first order model. Moreover, the
theoretical adsorption capacity predicted by the pseudo-second order
model is much closer to the experimental one. These results suggest that
the pseudo-second order model is more applicable to describe the ad-
Fig. 4. Kinetic profiles of Cu(II) adsorption by DTCS powder, ALG and Ca-DTCS/ALG
beads (DTCS/Alginate = 2:1): (A) The pseudo-first and pseudo-second order kinetic plots;
sorption process than the pseudo-first-order model, which reveals that
(B) The intra-particle diffusion kinetic model fits (Initial concentration of copper ions: the rate–determine step for ALG beads might be a chemisorption pro-
100 mg L−1; Dose: 1 g L−1; Temperature: 28 °C; pH = 3.0). cess [13].
In the case of Ca-DTCS/ALG, the RMSE values obtained from the
nonlinear regression of the two models are very close to each other,
respectively, which were much longer than that of DTCS powder due to
however, as presented in Fig. 4A, the experiment data follows the fitting
their much large gel bead sizes (2–3 mm). Similar results were observed
plot of the pseudo-first order model much better than that of the
in the kinetic studies for alginate, PEI/alginate, and PEI-CS2/alginate
pseudo- second order model, indicating a good correlation to the
beads [14]. In order to reach the adsorption equilibrium sufficiently,
pseudo-first order model by Ca-DTCS/ALG for copper ions. These re-
the contact time of 36 h was selected in the following isotherm ex-
sults suggested that the rate-controlling step for Ca-DTCS/ALG beads
periments.
was not a simple chemisorption process, and the sophisticated diffusion
For the purpose of investigation of the rate-determine step in the
process was also involved, because the Ca-DTCS micro-gels en-
adsorption mechanisms, the kinetic experimental data were fitted with
capsulated in the ALG matrix were less accessible and required much
Lagergren’s pseudo–first order [24], Ho’s pseudo–second order [25],
longer contact time for equilibrium and saturation [14]. Thus, as listed
and intra–particle diffusion kinetic models [26], respectively, and the
in Table 1, the rate constants of the pseudo-first order and pseudo-
mathematical representations (nonlinear) of these models are described
second order models (k1 and k2) for Ca-DTCS/ALG beads are much
as follows:
lower than that of ALG beads. In addition, the adsorption rate of DTCS

Table 1
Kinetic parameters and RMSE values for adsorption of Cu(II) ions on DTCS powder, ALG and Ca-DTCS/ALG beads (DTCS/Alginate = 2:1).

Adsorbent C0 qe, exp Pseudo-first order model Pseudo-second order model Intra-particle diffusion model
(mg L−1) (mg g−1)
qe, cal k1 RMSE qe, cal k2 RMSE ki, 1 ki, 2 ki, 3 ki, 4
(mg g−1) (min−1) (mg g−1) (g mg−1 min−1)
−1 −1/2
(mg g min )

DTCS 100 86.4 – – – – – – – – – –


ALG 100 42.6 40.1 0.018 2.37 42.9 0.00071 0.96 3.47 0.96 0.023 –
Ca-DTCS/ALG 100 61.1 58.9 0.0088 2.69 63.9 0.00021 2.87 3.77 1.97 0.14 0.043

327
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

qm KL Ce
qe =
1 + KL Ce (6)

where qe is the adsorbed value of copper ions at equilibrium con-


centration (mg g−1), and qm is the maximum adsorption capacity of
adsorbent (mg g−1); Ce represent the equilibrium concentration of
copper ions; KL is the Langmuir constant which is related to the affinity
of binding sites (L mg−1).
The Freundlich isotherm model [28] is based on the assumption of
the adsorption on the heterogeneous surfaces, and the amount of solute
adsorbed on adsorbent increase infinitely with the concentration rise.
The model can be described by the following equation:

qe = KF Ce1/ nF (7)

where KF is the Freundlich constant (L mg−1) related to adsorption


capacity, and nF is the heterogeneity factor related to adsorption in-
tensity. The nF values larger than 1 indicates heterogeneous adsorp-
Fig. 5. Adsorption isotherms of Cu2+ ions on DTCS powder, ALG and Ca-DTCS/ALG tions.
beads (DTCS/Alginate = 2:1) (Contact time: 36 h; Dose: 1 g L−1; Temperature: 28 °C; The Sips model is a hybrid of the Langmuir and the Freundlich
pH = 3.0). isotherms [29]:

qm (K S Ce )nS
powder was very fast, so that it is difficult to determine its rate–limiting qe =
step under this experimental conditions. 1 + (K S Ce )nS (8)
The intra-particle diffusion model is used to determine whether
where KS (L mg−1) is the Langmuir equilibrium constant and nS is
intra-particle diffusion is the rate-controlling step in the adsorption
comparable to the Freundlich heterogeneity factor (nS = 1/nF).
process. As shown in Fig. 4B, it can be observed that both plots are
The adsorption isotherm parameters simulated from the three iso-
divided into three or four linear portions, and do not pass through the
therm models mentioned above are listed in Table 2. For DTCS powder,
origin, indicating that the adsorption was mainly controlled by in-
Ca-DTCS/ALG and ALG beads, their RMSE values obtained from the
tra–particle diffusion, but other mechanisms such as surface chemi-
Sips model are observed to be lowest in comparison with that of the
sorption process might also take place simultaneously. Moreover, the
other two isotherm models, indicating that the isotherms of the three
diffusion rate constants in every portions of the plots are given in
adsorbents follow the Sips model best. As presented in Table 2, the
Table 1, which follow the order of ki,1 > ki,2 > ki,3 > ki,4. These re-
surface affinity values (KS or KL) of the three adsorbents for copper ions
sults can be illustrated as follows: the first portion with steep slope is
follow the order of DTCS powder ≫ Ca-DTCS/ALG beads ≫ ALG beads,
attributed to the fast external surface sorption stage, and the second or
suggesting that the ability of the three adsorbents binding to copper
third portion with gentle slope is related to slow internal diffusion, and
ions follows the same order [3–5]. Moreover, the heterogeneity factor
the slope of the final portion is nearly zero, suggesting that adsorption
(nS or 1/nF) values shown in Table 2 are all less than unit, indicating
equilibrium is achieved. Similar results were also observed in the pre-
that the sorption of Cu(II) ions onto the prepared adsorbents was a
vious studies [3–5].
heterogeneous process [3–5].
Table S2 lists the maximum adsorption capacities (qm) of copper
ions on as-prepared and other commonly used chitosan and alginate
3.2.3. Adsorption isotherms
derivatives adsorbents [30–34]. It can be seen that the maximum ad-
Fig. 5 represents the adsorption isotherms of copper ions on DTCS
sorption capacities of DTCS and Ca-DTCS/ALG was comparable and
powder, ALG and Ca-DTCS/ALG gel beads at 28 °C. It is found that the
both higher than most of other chitosan or alginate based adsorbents
copper adsorption capacities on the adsorbents follow the order of
listed in Table S2 except for alginate-immobilized bentonite clay and
DTCS powder > Ca-DTCS/ALG beads > ALG beads at the same
phenanthroline-modified chitosan, which might be attributed to the
equilibrium concentration when the equilibrium concentration is lower
increased number of adsorption sites in these two adsorbents [32,34].
than 150 mg L−1. The adsorption isotherms data were fitted by the
Nevertheless, the maximum adsorption capacities of DTCS and Ca-
three common adsorption isotherm models: Langmuir, Freundlich and
DTCS/ALG were measured at pH 3.0, while that of the other chitosan
Sips models, and the parameters of adsorption isotherms were obtained
and alginate based adsorbents was determined at pH 4 or pH 5. The
by the means of a nonlinear regression method, which had been em-
maximum adsorption capacity of ALG bead was comparable with that
ployed in the treatment of the kinetic data mentioned above.
of the other two plain alginate beads shown in Table S2, however, it
The Langmuir isotherm model [27] predicts that a monolayer ad-
was much lower than that predicted by the Sips model. This predicted
sorption happens on a homogeneous surface, and its nonlinear form
value might only be obtained at very high concentration of copper ions.
equation can be expressed as follows:

Table 2
Isothermal model constants, RMSE values and adsorption capacities of DTCS powder, ALG and Ca-DTCS/ALG beads (DTCS/Alginate = 2:1).

Materials qe, exp Langmuir model Freundlich model Sips model


(mg g−1)
qe, cal KL RMSE nF KF (L mg−1) RMSE qe, cal ns KS RMSE
(mg g−1) (L mg−1) (mg g−1) (L mg−1)

DTCS 106.0 102.7 5.8 2.1 36.2 90.0 2.4 105.8 0.45 30.4 1.8
Ca-DTCS/ALG 107.5 103.3 0.19 3.8 5.4 38.8 10.5 110.5 0.76 0.15 2.5
ALG 101.0 102.0 0.023 4.9 3.1 14.1 3.7 152.0 0.58 0.0066 1.5

328
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

3.2.4. Effect of inorganic cations on copper removal


Generally speaking, the presence of the inorganic cations such as Na
(I) and Ca(II) ions in heavy metal wastewaters may affect the perfor-
mance of the adsorption process [21]. Fig. 6A and B reports the removal
performance of Cu(II) ions by DTCS powder, Ca-DTCS/ALG and ALG
beads in the presence of different concentrations of Na(I) or Ca(II) ions.
In the case of DTCS powder, no significant difference in Cu (II) removal
efficiency is observed as the concentrations of Na(I) and Ca(II) ions rise
up to 1000 mg L−1 (43.5 mmol L−1) and 400 mg L−1 (10 mmol L−1) in
the copper solutions, respectively. These results suggested that Na(I) or
Ca(II) ions did not compete strongly with Cu(II) in the chelating reac-
tion with DTPA groups bound to the adsorbents, which is related to the
much higher DTPA complex formation constant for Cu(II)
(lgK = 21.10) than that of Ca(II) (lgK = 10.74) and Na(I) ions (its
lgK ≪ 10.74, because Na(I) ion is a monovalent metal ion) [35].
For Ca-DTCS/ALG beads, the increase of the concentration of Na(I)
ions has no significant effect on its Cu (II) removal efficiency, but there
is a slight decrease in copper removal efficiency with the increasing of
Ca(II) concentration in the solution. However, the copper removal ef-
ficiency by ALG beads decreases greatly as Na+ and Ca2+ concentra-
tions are higher than 400 mg L−1 and 50 mg L−1, respectively, which
was attributed to the strong competitive adsorption of Na+ and Ca2+
with Cu2+ for the adsorption sites (–COOH groups) on ALG beads, and
thus negatively affect the adsorption of Cu(II) ions by the ALG beads
[21]. Obviously, the copper adsorption performance of Ca-DTCS/ALG
beads in the presence of Na+ and Ca2+ ions was improved significantly
in comparison with ALG beads.
As the other inorganic metal ions such as Al3+ and Fe2+ ions may
usually coexist with Cu(II) in the electroplating wastewater, the effect
of Al3+ and Fe2+ on the copper removal efficiency of Ca-DTCS/ALG
beads was further investigated. The results are displayed in Fig. 6C. It
can be seen that the effect of Al(III) ions concentration on the copper
removal efficiency is little, but the copper removal efficiency decreases
greatly in the presence of Fe(II) ions. Based on the software calculation
results (Visual MINTEQ ver. 3.1), this can be explained by the much
higher binding ability of Cu2+ ions to DTPA groups in Ca-DTCS/ALG
beads than that of Al3+ ions. Although the chelating ability of DTPA to
Al3+ is higher than that of Fe2+(Visual MINTEQ ver. 3.1), the effect of
Fe(II) ions on Cu(II) adsorption is greater than that of Al3+ ions. This is
because that some of Fe(II) ions were oxidized to Fe(III) ions in the
adsorption process, and the DTPA complex formation constant for Fe
(III) (lgK = 28.60) is much higher than that of Cu(II) (lgK = 21.10).

3.2.5. Reusability study


As DTCS powder was not easily recovered from aqueous solution
due to its notable swelling properties, the reuse experiments of the
adsorbents were only conducted by Ca-DTCS/ALG and ALG beads.
Based on Refs. [3–5], the regeneration of heavy metal-binding EDTA-
chitosan and DTPA-chitosan was usually performed by 2 M HNO3, thus,
it was selected as the eluent for regeneration of copper-binding Ca-
DTCS/ALG and ALG beads in the six adsorption-desorption continuous
cycles. The results of the reusability of the two adsorbents are shown in
Fig. 7. It can be seen that the adsorption capacity of Ca-DTCS/ALG and
ALG beads maintain 88% and 51% of their initial values after six times
reuse, respectively. The decreased adsorption capacity was likely due to
the loss or degradation of the adsorbent [35]. In addition, the in-
completely removal of Cu(II) ions from the adsorbent by the eluent
Fig. 6. Effect of Na+ ions (A) and Ca2+ ions (B) concentrations on sorption of Cu2+ ions
might be another reason for the decreased adsorption capacity. For
by DTCS powder, ALG and Ca-DTCS/ALG beads (DTCS/Alginate = 2:1); (C) Effect of
Al3+ and Fe2+ ions concentrations on sorption of Cu2+ ions by Ca-DTCS/ALG beads example, the copper desorption efficiency from copper-loaded Ca-
(Initial concentration of copper ions: 100 mg L−1; Contact time: 36 h; Dose: 1 g L−1; DTCS/ALG by 2 M HNO3 was about 99.1%, which was less than 100%.
Temperature: 28 °C; pH = 3.0). Overall, due to its high regeneration efficiency, Ca-DTCS/ALG beads
can be used repeatedly as an efficient adsorbent for practical waste-
water treatment.

329
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

carboxylic group in the state of (−1/2OfCfO1/2− Na+) might be


stronger than that in the complex of (O]CeOeCueOeC]O). Thus,
after adsorption of copper ions on DTCS, the peaks for C]O and CeO
bands in the carboxylic group shift to the higher frequency (from 1587
to 1624 cm−1) and the lower frequency (from 1402 to 1401 cm−1),
respectively. Similar results were reported by the Ref. [36]. In fact, after
adsorption of Cu(II) at the initial copper concentration of 100 ppm
(pH = 3.0), 95% of Na(I) ions in DTCS powder was exchanged by Cu(II)
ions based on ICP-OES measurement results.
On the other hand, for Ca-DTCS/ALG beads, after adsorption of
copper ions, the peaks for C]O and CeO bands in the carboxylic group
both shift to lower frequency (from 1633 to 1624 cm−1 and from 1403
to 1401 cm−1, respectively) due to the transformation of
(O]CeOeCaeOeC]O) complex to (O]CeOeCueOeC]O) com-
plex, indicating that the CueO band is stronger than Ca-O band and
thus weakens the adjacent CeO and C]O bands. Haug [37] reported
that the affinity of alginic acid to metal ions follows the order of
Pb2+ > Cu2+ > Cd2+ > Ba2+ > Sr2+ > Ca2+ > Co2+. More-
Fig. 7. Reuse of ALG and Ca-DTCS/ALG beads (DTCS/Alginate = 2:1) in removal of over, the intensity of the peak at 1735 cm−1 assigned to the carbonyl
copper ions from aqueous solution (Initial concentration of copper ions: 100 mg L−1; stretch of –COOH group is weakened after copper adsorption, sug-
Contact time: 36 h; Dose: 1 g L−1; Temperature: 28 °C; pH = 3.0). gesting that the –COOH group was also involved in the formation of the
complex of (O]CeOeCueOeC]O). Thus, the analysis of FTIR spectra
3.3. Adsorption mechanism provides the obvious evidences for the complex reaction between the
carboxyl groups on the adsorbents and copper ions in the solution.
In order to gain a good understanding of the adsorption mechanism, Based on the adsorption experiments, after adsorption of Cu(II) at the
FT-IR and XPS were employed to identify the functional groups on the initial copper concentration of 100 ppm (pH = 3.0), most of Cu(II) in
adsorbents and their interactions with copper ions. aqueous solution was exchanged with hydrogen ions in Ca-DTCS/ALG
beads. Nevertheless, only 37% of Ca(II) in Ca-DTCS/ALG beads (Ca
content: 10 mg g−1(dry bead)) were exchanged by Cu(II) ions.
3.3.1. FTIR analysis
The FTIR spectra of DTCS powder and Ca-DTCS/ALG beads before
3.3.2. XPS analysis
and after adsorption of Cu(II) ions are presented in Fig. 8. The notice-
The wide scan XPS spectra of DTCS powder and Ca-DTCS/ALG
able feature is that the absorption peaks assigned to the stretching vi-
beads show that, the peaks at binding energy of 1071.3 eV for Na1s in
bration of C]O and CeO bonds in the carboxylic group (O]CeO−)
DTCS (Fig.S2A) and 347.98 eV for Ca2p in Ca-DTCS/ALG (Fig. S2B)
bound to the adsorbent show a minor shift due to the copper adsorp-
both disappear after the copper adsorption (Fig.S2). These results in-
tion. In the case of DTCS, before adsorption, the carboxylate group on
dicate that the sodium ions bound to DTCS powder and the calcium ions
virgin DTCS was considered to be in the form of its conjugate state of
bound to Ca-DTCS/ALG beads may be both exchanged by the copper
(−1/2OfCfO1/2− Na+) due to the small electronegativity (0.9) of so-
ions after adsorption of copper ions at pH 3.0 [35]. High resolution
dium atom. After copper adsorption, the carboxylate group on DTCS
spectra of N1s regions of DTCS powder and Ca-DTCS/ALG beads are
combined with copper ions would form a complex of (O]CeOe-
presented in Fig. 9. The N1s spectra can be deconvoluted into four or
CueOeC]O) due to the relatively large electronegativity (1.9) of
five individual component peaks, which come from the different ni-
copper atom. Based on the principle of infrared spectrum, the stretching
trogen-containing groups and overlap on each other [36]. The binding
vibration of C]O band of the carboxylic group in the state of (−1/
2 energy and relative content of N1s in different forms in the two sorbents
OfCfO1/2− Na+) might be weaker than that in the (O]CeOe-
before and after copper adsorption are given in Table S3.
CueOeC]O) complex, but the stretching vibration of CeO band of the
In the case of virgin DTCS powder, as shown in Fig. 9A, the peaks at
binding energy of 402.1, 401.0, 399.9 and 399.2 eV can be assigned to
the nitrogen atom in the form of protonated amine [34], amide bond
(O]CeNHe) [38], tertiary amine (CeN) [38] and primary amine
(–NH2)[38,39], respectively. For copper-loaded DTCS powder, as pre-
sented in Fig. 9B, the peak with binding energy of 402.1 eV disappears
and the content of primary amine and tertiary amine increase, in-
dicating that ion exchange between hydrogen ions from protonated
amine on DTCS and copper ions in solution was happened during the
process of copper adsorption. Moreover, Table S3 shows that the con-
tent of the amide bond does not changed much before and after ad-
sorption of copper ions on DTCS. Furthermore, it can be seen that the
binding energy of the peaks of the tertiary amine (400.3 eV) and pri-
mary amine (399.6 eV) has a certain degree of shift to higher energy
region (compared with virgin DTCS), which is due to the decrease of the
electron density toward the nitrogen atoms after copper adsorption and
thus indicates the complex formation between copper ions and the ni-
trogen atoms of the tertiary amine and primary amine of DTCS [39].
In addition, as presented in Table S3, the content of the nitrogen
atom in the form of tertiary amine of the copper-loaded DTCS powder
Fig. 8. FTIR spectra of DTCS powder and Ca-DTCS/ALG beads before and after adsorption was about 49.6%, indicating that the content of the nitrogen atom
of copper ions from aqueous solution.
coming from the DTPA groups bound to DTCS powder can be estimated

330
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

Fig. 9. N1s XPS spectra of (A) DTCS powder and


(C) Ca-DTCS/ALG beads before adsorption of Cu
(II) ions from aqueous solutions; N1s XPS spectra
of (B) DTCS powder and (D) Ca-DTCS/ALG beads
after adsorption of Cu(II) ions from aqueous so-
lutions.

to 49.6%. Moreover, the total nitrogen content of DTCS obtained by greatly (from 21.2% to 44.9%) in comparison with that of Ca-DTCS/
element analysis was 8.61%. Through the equation (S1) shown in the ALG beads, indicating that protonated tertiary amine in Ca-DTCS/ALG
supplementary material, the surface coverage of DTPA groups on DTCS beads could react with copper ions to form DTPA chelates during the
can be calculated to be 1.0 mmol g−1, which is comparable with that adsorption process. Moreover, a small amount of protonated primary
(0.9–0.96 mmol g−1) of the DTPA-modified chitosan prepared by DTPA amine could also chelate with copper ions due to the increasing of
bis-anhydride [7]. In addition, the degree of substitution (D.S.) of the primary amine from 8.4% to 12.6% after copper adsorption. In addi-
DTPA residue on the modified chitosan (DTCS) can be calculated to be tion, the peaks for Ca2p atom disappear in the wide scan XPS spectra of
32.8% from the content of the tertiary amine nitrogen atom (49.6%) the Cu-DTCS/ALG beads compared with that of Ca-DTCS/ALG (Fig.S2).
and the total contents (50.4%) of the nitrogen atoms in the form of These results suggest that the hydrogen ions or calcium ions bound to
amide (O]CeN, 16.5%) and primary amine (–NH2, 33.9%) bonds in Ca-DTCS/ALG beads could be exchanged by the copper ions during the
the copper loaded DTCS (using the equation (S2) shown in the sup- adsorption process. Furthermore, the primary amine in copper loaded
plementary material). Ca-DTCS/ALG beads might be chelated with copper ions due to a shift
Compared with the N1s XPS spectra of DTCS powder, the new peak of the binding energy of the primary amine from 399.2 (before ad-
with binding energy of 400.5 eV assigned to the C]N groups [40] is sorption) to 399.6 eV (after adsorption).
obviously observed in Ca-DTCS/ALG beads (Fig. 9C), which is formed Based on the analysis of XPS and FTIR spectra of DTCS before and
during the process for the preparation of Ca-DTCS/ALG beads by using after copper adsorption, the assumption that DTPA modified-chitosan
glutaraldehyde as the cross-linker of amino groups on DTCS im- forms the octahedral structure with copper ions is illustrated in Scheme
mobilized in alginate matrix. The relative content of the C]N groups is 1. Moreover, the amino groups on DTCS that has not been replaced by
about 9.3% in Ca-DTCS/ALG beads, while the quantity of –NH2 groups DTPA can also form complexes with copper ions.
decreases from 33.9% in DTCS to 8.4% in Ca-DTCS/ALG beads. This is From the XPS and FTIR analysis of Ca-DTCS/ALG beads before and
because that a part of the –NH2 groups on DTCS was cross-linked by after copper adsorption, it was confirmed that copper ions diffused into
glutaraldehyde in Ca-DTCS/ALG beads, and the other part of –NH2 the alginate matrix and then bound to DTCS immobilized in the alginate
groups on DTCS (16.2%) was protonated (Table S3). In addition, the matrix. Obviously, the adsorption mechanism of copper ions into Ca-
binding energy of the peak of the nitrogen atom in the form of tertiary DTCS/ALG beads is a complicated process, in which ion exchange and
amine in Ca-DTCS/ALG beads also slightly shifted to higher energy complex formation are involved [36].
region (compared with DTCS), suggesting the complex formation be-
tween calcium ions and the nitrogen atoms of DTPA groups. However, 3.4. Recovery of copper from simulated electroplating wastewater
the binding energy at 399.2 eV attributed to the primary amine residues
in Ca-DTCS/ALG beads is the same as that of DTCS, indicating that the To investigate their application in practice, DTCS powder, Ca-
primary amine do not form chelates with calcium ions in Ca-DTCS/ALG DTCS/ALG and ALG beads were tested to recover copper ions from si-
beads. This result is consistent with the fact that the amino groups of mulated electroplating wastewater, which was strongly acidic (pH 3.0)
chitosan are not chelated with calcium ions [39]. and contained a variety of chelating agents or ligands such as sodium
As presented in Fig. 9D and Table S3, after Ca-DTCS/ALG beads di-hydrogen phosphate, EDTA, oxalic acid, tartaric acid, glycine and
adsorption of copper ions, the contents of the nitrogen atoms in the ethylenediamine, etc. The detailed composition of the chelating agents
form of amide (O]CeNHe) and imide (C]N) bonds in the Cu-DTCS/ or ligands in the simulated electroplating wastewater is listed in Table
ALG beads are very similar with that of Ca-DTCS/ALG. However, the S1, and the concentration of copper (II) ion in the simulated electro-
content of protonated nitrogen atom in Cu-DTCS/ALG beads decreases plating wastewater is varied from 1 to 30 mg L−1. The impact of ligands
from 44.3% to 16.2%, while the content of the tertiary amine increase on the recovery efficiency of copper by the adsorbents at different

331
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

higher than that of Ca-DTCS/ALG beads, but surprisingly, the ability of


Ca-DTCS/ALG to overcome the interference from chelating agents is
much stronger than that of DTCS especially at the copper concentration
of 1–5 mg L−1 (Fig. 10A).
Repo [7] investigated the reaction mechanism of the adsorption of
Co(II) ions on DTPA-functionalized silica gel and chitosan in the pre-
sence of EDTA. Their capillary electrophoresis measurement results
confirmed that the surface DTPA groups of the adsorbents were able to
capture Co(II) from its dissolved EDTA chelates due to the high-affinity
of DTPA-groups to Co(II).
In this work, compared with the other ligands in the simulated
electroplating wastewater shown in Table S1, EDTA is considered to be
the strongest chelating agent for copper ions at pH 3.0. Thus, the main
copper species in the simulated electroplating wastewater at the low
copper concentration of 1 mg L−1 might be in the form of its dissolved
EDTA chelates (Cu(II):EDTA = 1:2.86). However, it seems difficult for
DTCS to capture Cu(II) from its dissolved EDTA chelates (Cu
(II):EDTA = 1:2.86) due to the much lower efficiency of copper re-
covery by DTCS (Fig. 10A). The results can be explained as follows: for
DTCS powder in the simulated electroplating wastewater, the excessive
chelating ligands such as EDTA, oxalic acid, glycine, ethylenediamine
and the like were distributed around the DTPA groups bound to DTCS,
and might strongly compete with the DTPA groups for the copper ions
dissociated from its EDTA chelates. Thus, these ligands interfered ser-
iously with the removal of copper ions from the simulated electro-
plating wastewater by DTCS at low copper concentration of 1 mg L−1.
However, in the case of Ca-DTCS/ALG beads, the copper ions dis-
sociated from its EDTA chelates might diffuse into the inside of the gel
matrix more easily than EDTA ligands or Cu-EDTA chelates, which
might be related to the size of copper ion smaller than that of EDTA or
Cu-EDTA chelate. Moreover, it seems that the alginate matrix preferred
the positively charged copper ions to the negatively charged EDTA or
Cu-EDTA chelate at pH 3.0 (Visual MINTEQ ver. 3.1) due to the ne-
gatively charged carboxylic groups of alginate [13]. In addition, the
concentration of EDTA ligand was low outside of Ca-DTCS/ALG gel
beads (shown in Table S1). Thus, few EDTA ligands might diffuse into
the alginate matrix. This speculation was confirmed indirectly by the
results that ALG beads hardly recovered copper ions from the simulated
Fig. 10. The efficiency of copper recovery by DTCS powder, ALG and Ca-DTCS/ALG electroplating wastewater at the copper concentration of 1 mg L−1 (Cu
beads (DTCS/Alginate = 2:1) in the simulated electroplating wastewater (A) and the (II):EDTA = 1:2.86), because there seem to be no Cu-EDTA chelate
aqueous solutions containing copper ions in the absence of chelating ligands (B) (Contact
diffusing into ALG beads. Therefore, the interference of EDTA ligands
time: 36 h; Dose: 2 g L−1; Temperature: 28 °C; pH = 3.0).
on the chelating reaction of copper ions with the DTPA groups of DTCS
in the matrix might be reduced greatly. Under this condition, the
copper concentrations is illustrated in Fig. 10A. As a comparison, the copper ions could be continuously dissociated from the Cu-EDTA
performance of the three adsorbents for the recovery of copper in the complex outside of the Ca-DTCS/ALG gel beads, then diffuse into the
absence of chelating agents or ligands is shown in Fig. 10B. It is ob- alginate matrix and strongly bind to the DTPA groups in the matrix. In
viously observed that the chelating agents or ligands seriously inter- this way, the copper removal efficiency by Ca-DTCS/ALG beads could
fered with the recovery of copper ions from the aqueous solutions by reach 92% in the simulated electroplating wastewater with copper
the adsorbent. In addition, the higher the molar ratio of total chelating concentration of 1 mg L−1 (Cu(II):EDTA = 1:2.86).
agents to copper ions, the lower the efficiency of copper recovery by As the molar ratio of EDTA to copper ions decreased and the con-
ALG beads and DTCS powder. For example, as the molar ratio of total centration of copper ions increased from 10 to 30 mg L−1 in the simu-
chelating agents to copper is 24.95: 1 at the copper concentration of lated electroplating wastewater, the amount of free copper ions in-
1 mg L−1, the ALG beads are almost impossible to recover copper from creased greatly. When these free copper ions diffused into the alginate
the simulated electroplating wastewater, and the efficiency of copper matrix, they preferentially occupy the external active sites of Ca-DTCS/
recovery by DTCS powder is only 38%. However, the efficiency of ALG beads, which might inhibit copper ions dissociated from the Cu-
copper recovery by Ca-DTCS/ALG beads is kept at around 92% at the EDTA chelate outside of Ca-DTCS/ALG beads. Thus, the ability of Ca-
copper concentration ranging from 1 to 30 mg L−1 in the simulated DTCS/ALG beads for capture of copper ions from Cu-EDTA chelate
electroplating wastewater, indicating that Ca-DTCS/ALG beads show would decrease under this condition. Obviously, in order to further
more excellent performance for copper recovery in the presence of improve the efficiency of copper recovery by Ca-DTCS/ALG beads at
chelating agents or ligands in comparison with that of ALG beads and high copper concentration in the simulated electroplating wastewater,
DTCS powder. it is better to increase the dosage of Ca-DTCS/ALG beads.
As summarized in Table 2, the Langmuir equilibrium constant KL or
KS of ALG beads is the lowest in the three adsorbents, suggesting that 4. Conclusions
the affinity of ALG beads to copper (II) ions is the lowest. Consequently,
its ability for copper recovery in the presence of chelating agents is the DTCS powder was prepared by modification of chitosan with DTPA
weakest. Nevertheless, the KL or KS value of DTCS powder is much and using EDAC and NHS as cross-linkers. DTCS was found to be a

332
Y. Huang et al. Chemical Engineering Journal 339 (2018) 322–333

highly effective adsorbent for removal of copper ions from strongly Enhanced sr adsorption performance of MnO2-alginate beads in seawater and
acidic solutions. The maximum adsorption capacity for copper ions by evaluation of its mechanism, Chem. Eng. J. 319 (2017) 163–169.
[13] H.C. Vu, A.D. Dwivedi, T.T. Le, S.-H. Seo, E.-J. Kim, Y.-S. Chang, Magnetite gra-
DTCS at pH 3.0 was 106 mg g−1, and its adsorption rate was so fast that phene oxide encapsulated in alginate beads for enhanced adsorption of Cr(VI) and
the adsorption equilibrium was achieved within 1 min. However, DTCS As(V) from aqueous solutions: role of crosslinking metal cations in pH control,
is not easily separated from aqueous solutions due to its notable swel- Chem. Eng. J. 307 (2017) 220–229.
[14] C. Bertagnolli, A. Grishin, T. Vincent, E. Guibal, Recovering heavy metal ions from
ling property. Therefore, the composite adsorbent of Ca-DTCS/ALG complex solutions using polyethylenimine derivatives encapsulated in alginate
beads was prepared by encapsulating DTCS micro-gels into the alginate matrix, Ind. Eng. Chem. Res. 55 (2016) 2461–2470.
gel beads. Compared with plain alginate (ALG beads), the Ca-DTCS/ [15] M. Yakup Arıca, G. Bayramoǧlu, M. Yılmaz, S. Bektaş, Ö. Genç, Biosorption of
Hg2+, Cd2+, and Zn2+ by Ca-alginate and immobilized wood-rotting fungus funalia
ALG beads (DTCS/Alginate = 2:1) showed high copper removal effi- trogii, J. Hazard. Mater. 109 (2004) 191–199.
ciency in strongly acidic solutions (pH 1.5–3), and its copper capacities [16] X. Zhu, M. Su, S. Tang, L. Wang, X. Liang, F. Meng, Y. Hong, Z. Xu, Synthesis of
decreased very slightly in the presence of sodium (up to 1000 mg L−1), thiolated chitosan and preparation nanoparticles with sodium alginate for ocular
calcium and aluminum (up to 400 mg L−1) salts. Moreover, the Ca- drug delivery, Mol. Vis. 18 (2012) 1973–1982.
[17] Y. Chen, F. He, Y. Ren, H. Peng, K. Huang, Fabrication of chitosan/PAA multilayer
DTCS/ALG beads (DTCS/Alginate = 2:1) showed much higher effi- onto magnetic microspheres by LbL method for removal of dyes, Chem. Eng. J. 249
ciency of copper recovery in the simulated electroplating wastewater (2014) 79–92.
than that of DTCS and plain alginate at low copper concentration of [18] Y. Ren, H.A. Abbood, F. He, H. Peng, K. Huang, Magnetic EDTA-modified chitosan/
SiO2/Fe3O4 adsorbent: preparation, characterization, and application in heavy
1–5 mg L−1. In addition, the Ca-DTCS/ALG beads showed good se- metal adsorption, Chem. Eng. J. 226 (2013) 300–311.
paration performance and reusability in the removal of copper ions [19] T. Lu, T. Xiang, X.L. Huang, C. Li, W.F. Zhao, Q. Zhang, C.S. Zhao, Post-crosslinking
from solutions. Thus, Ca-DTCS/ALG beads not only have the ad- towards stimuli-responsive sodium alginate beads for the removal of dye and heavy
metals, Carbohydr. Polym. 133 (2015) 587–595.
vantages of both DTCS and ALG adsorbents, but also overcome their [20] K. Sakugawa, A. Ikeda, A. Takemura, H. Ono, Simplified method for estimation of
shortcomings. Ca-DTCS/ALG is a good candidate as an adsorbent for composition of alginates by FTIR, J. Appl. Polym. Sci. 93 (2004) 1372–1377.
the removal of copper ions from wastewater. [21] J.H. Chen, Q.L. Liu, S.R. Hu, J.C. Ni, Y.S. He, Adsorption mechanism of Cu(II) ions
from aqueous solution by glutaraldehyde crosslinked humic acid-immobilized so-
dium alginate porous membrane adsorbent, Chem. Eng. J. 173 (2011) 511–519.
Acknowledgements [22] R.D.C. Soltani, G.S. Khorramabadi, A.R. Khataee, S. Jorfi, Silica nanopowders/al-
ginate composite for adsorption of lead (II) ions in aqueous solutions, J. Taiwan
Inst. Chem. Eng. 45 (2014) 973–980.
The authors would like to thank the financial support provided by [23] Y.C. Chang, D.H. Chen, Preparation and adsorption properties of monodisperse
Ministry of Water Resource, Public Interest Scientific Research Fund chitosan-bound Fe3O4 magnetic nanoparticles for removal of Cu(II) ions, J. Colloid
(No.201501019), and also acknowledge the Analytical and Testing Interface Sci. 283 (2005) 446–451.
[24] S. Lagergren, About the theory of so-called adsorption of solution substances, Kung,
Center, Huazhong University of Science and Technology (China) for the
Sven. Veten. Hand. 24 (1898) 1–39.
kind help on sample characterization. [25] Y.S. Ho, G. McKay, Sorption of dye from aqueous solution by peat, Chem. Eng. J. 70
(1998) 115–124.
Appendix A. Supplementary data [26] W. Weber, J. Morris, Kinetics of adsorption on carbon from solution, J. Sanit. Eng.
Div. Am. Soc. Civ. Eng. 89 (1963) 31–60.
[27] I. Langmuir, The constitution and fundamental properties of solids and liquids. Part
Supplementary data associated with this article can be found, in the I. solids, J. Am. Chem. Soc. 38 (1916) 2221–2295.
online version, at http://dx.doi.org/10.1016/j.cej.2018.01.071. [28] H. Freundlich, Uber die adsorption in lasungen, Z. Phys. Chem. 57 (1906) 385–470.
[29] Y.S. Ho, J.F. Porter, G. Mckay, Equilibrium isotherm studies for the sorptionof di-
valent metal ions onto peat: copper, nickel and lead. Single componentsystems,
References Water Air Soil Pollut. 141 (2002) 1–33.
[30] W.S.W. Ngah, S. Fatinathan, Adsorption characterization of Pb(II) and Cu(II) ions
onto chitosan-tripolyphosphate beads: kinetic, equilibrium and thermodynamic
[1] S. Olivera, H.B. Muralidhara, K. Venkatesh, V.K. Guna, K. Gopalakrishna,
studies, J. Environ. Manage. 91 (2010) 958–969.
K.Y. Kumar, Potential applications of cellulose and chitosan nanoparticles/com-
[31] F. Zhao, E. Repo, D. Yin, Y. Meng, S. Jafari, M. Sillanpaa, EDTA-cross-linked beta-
posites in wastewater treatment: a review, Carbohydr. Polym. 153 (2016) 600–618.
cyclodextrin: an environmentally friendly bifunctional adsorbent for simultaneous
[2] J. Wang, C. Chen, Chitosan-based biosorbents: modification and application for
adsorption of metals and cationic dyes, Environ. Sci. Technol. 49 (2015)
biosorption of heavy metals and radionuclides, Bioresour. Technol. 160 (2014)
10570–10580.
129–141.
[32] W.S. Tan, A.S. Ting, Alginate-immobilized bentonite clay: adsorption efficacy and
[3] E. Repo, J.K. Warchol, T.A. Kurniawan, M.E.T. Sillanpää, Adsorption of Co(II) and
reusability for Cu(II) removal from aqueous solution, Bioresour. Technol. 160
Ni(II) by EDTA- and/or DTPA-modified chitosan: kinetic and equilibrium modeling,
(2014) 115–118.
Chem. Eng. J. 161 (2010) 73–82.
[33] J. Ma, G. Zhou, L. Chu, Y. Liu, C. Liu, S. Luo, Y. Wei, Efficient removal of heavy
[4] E. Repo, J.K. Warchol, A. Bhatnagar, M. Sillanpaa, Heavy metals adsorption by
metal ions with an EDTA functionalized chitosan/polyacrylamide double network
novel EDTA-modified chitosan-silica hybrid materials, J. Colloid Interface Sci. 358
hydrogel, ACS Sustain. Chem. Eng. (2016).
(2011) 261–267.
[34] H. Jiang, W. Zhang, P. Chen, W. Zhang, G. Wang, X. Luo, S. Luo, Equipping an
[5] F. Zhao, E. Repo, D. Yin, M.E. Sillanpaa, Adsorption of Cd(II) and Pb(II) by a novel
adsorbent with an indicator: a novel composite to simultaneously detect and re-
EGTA-modified chitosan material: kinetics and isotherms, J. Colloid Interface Sci.
move heavy metals from water, J. Mater. Chem. A 4 (2016) 11897–11907.
409 (2013) 174–182.
[35] Y. Huang, A.A. Keller, EDTA functionalized magnetic nanoparticle sorbents for
[6] V.E. Tikhonov, L.A. Radigina, Y.A. Yamskov, Metal-chelating chitin derivatives via
cadmium and lead contaminated water treatment, Water Res. 80 (2015) 159–168.
reaction of chitosan with nitrilotriacetic acid, Carbohydr. Res. 290 (1996) 33–41.
[36] S.-F. Lim, Y.-M. Zheng, S.-W. Zou, J.P. Chen, Characterization of copper adsorption
[7] E. Repo, L. Malinen, R. Koivula, R. Harjula, M. Sillanpaa, Capture of Co(II) from its
onto an alginate encapsulated magnetic sorbent by a combined FT-IR, XPS, and
aqueous EDTA-chelate by DTPA-modified silica gel and chitosan, J. Hazard. Mater.
mathematical modeling study, Environ. Sci. Technol. 42 (2008) 2551–2556.
187 (2011) 122–132.
[37] A. Haug, J. Bjerrum, O. Buchardt, G.E. Olsen, C. Pedersen, J. Toft, The affinity of
[8] J. Roosen, K. Binnemans, Adsorption and chromatographic separation of rare earths
some divalent metals for different types of alginates, Acta Chem. Scand. 15 (1961)
with EDTA- and DTPA-functionalized chitosan biopolymers, J. Mater. Chem. A 2
1794–1795.
(2014) 1530–1540.
[38] Y. Yuan, G. Zhang, Y. Li, G. Zhang, F. Zhang, X. Fan, Poly(amidoamine) modified
[9] F. Zhao, E. Repo, M. Sillanpää, Y. Meng, D. Yin, W.Z. Tang, Green synthesis of
graphene oxide as an efficient adsorbent for heavy metal ions, Polym. Chem. 4
magnetic EDTA- and/or DTPA-cross-linked chitosan adsorbents for highly efficient
(2013) 2164.
removal of metals, Ind. Eng. Chem. Res. 54 (2015) 1271–1281.
[39] B. Liao, W.-Y. Sun, N. Guo, S.-L. Ding, S.-J. Su, Comparison of Co2+ adsorption by
[10] F. Wang, X. Lu, X.Y. Li, Selective removals of heavy metals (Pb2+, Cu2+, and Cd2+)
chitosan and its triethylene-tetramine derivative: performance and mechanism,
from wastewater by gelation with alginate for effective metal recovery, J. Hazard.
Carbohydr. Polym. 151 (2016) 20–28.
Mater. 308 (2016) 75–83.
[40] Z. Geng, H. Zhang, Q. Xiong, Y. Zhang, H. Zhao, G. Wang, A fluorescent chitosan
[11] F. Googerdchian, A. Moheb, R. Emadi, Lead sorption properties of nanohydrox-
hydrogel detection platform for the sensitive and selective determination of trace
yapatite-alginate composite adsorbents, Chem. Eng. J. 200–202 (2012) 471–479.
mercury(II) in water, J. Mater. Chem. A 3 (2015) 19455–19460.
[12] H.-J. Hong, B.-G. Kim, J. Hong, J. Ryu, T. Ryu, K.-S. Chung, H. Kim, I.-S. Park,

333

You might also like