You are on page 1of 14

International Journal of Biological Macromolecules 80 (2015) 431–444

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


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

Application of chitosan-citric acid nanoparticles for removal of


chromium (VI)
Mahdiye Bagheri a , Habibollah Younesi b,∗ , Shaaker Hajati c,d , Seyed Mehdi Borghei e
a
Department of Environmental Science, Science and Research Branch, Islamic Azad University, Tehran, Iran
b
Department of Environmental Science, Faculty of Natural Resources, Tarbiat Modares University, Imam Reza Street, Noor P.O. Box 46414-356, Iran
c
Department of Physics, Yasouj University, Yasouj 75918-74831, Iran
d
Physics Department, Sharif University of Technology, Tehran P.O. Box 11155-9161, Iran
e
Sharif University of Technology – BBRC, Tehran, Iran

a r t i c l e i n f o a b s t r a c t

Article history: In the present study, CS–CA nanoparticle was prepared for forming a new amide linkage, by grafting
Received 8 April 2015 the amino groups of CS in the presence of carboxylic groups of CA that acts as cross-linking agent. The
Received in revised form 9 July 2015 as-prepared CS–CA nanoparticle samples were characterized by use of dynamic light scattering (DLS),
Accepted 12 July 2015
scanning electron microscopy (SEM), Fourier-transformed infrared spectroscopy (FTIR), X-ray diffraction
Available online 16 July 2015
(XRD) and X-ray photoelectron spectroscopy (XPS) techniques, which showed that the cross-linking agent
preserved during the chemical modifications. The adsorption capacity of the CS–CA nanoparticles for the
Keywords:
removal of Cr (VI) in aqueous solution was studied. The adsorption equilibrium data taken at the optimized
Chitosan–CA nanoparticle
Cr (VI)
condition, i.e., 25 ◦ C and pH of 3, were analyzed with the Langmuir, Freundlich and Redlich–Peterson
Isotherm isotherm models. The kinetics of Cr (VI) adsorption on CS–CA nanoparticles obtained at different initial
Kinetics concentrations were also analyzed using the pseudo-second-order model.
Desorption © 2015 Elsevier B.V. All rights reserved.

1. Introduction and hexavalent chromium [Cr (VI)], Cr (VI) being five times more
toxic than Cr (III).
Water is the most widely abundant occurring substance on Beyond a certain concentration, the toxicity of Cr (VI) has serious
earth, used for drinking, hygiene, industry and agriculture. Unfor- implications for human health such as damage to the physiol-
tunately, developing a modern civilization and societies shifting ogy of the body by its accumulation in the food chain, problems
toward industrialization has led to introducing different types of ranging from simple skin irritation to lung cancer, head and neck,
toxic substances, harmful to the health of living organisms, into kidney, liver and stomach injuries [2]. This element has been indi-
the environment such as heavy metals in drinking water supplies cated to be mutagenic and carcinogenic and is soaked up through
and wastewater discharge. Nowadays, water pollution by heavy the skin and ingestion. So if untreated wastewater containing Cr
metals is one of the most serious environmental problems, as they (VI) is discharged to surface waters, contamination that endangers
are nonbiodegradable, accumulate in the environment with a long ecosystems also jeopardize people’s health seriously. The maxi-
biological half-life, subsequently contaminating the food chain, mum allowable Cr (VI) in drinking water and that discharged to
increasing cancer risk [1]. Chromium is one of the most dangerous the surface waters recommended by World Health Organization
toxic heavy metals that get discharged into the environment as a (WHO) is 0.05 mg/l, while the drinking water standard for total
result of industrial applications, including metal smelting, mineral chromium is 100 ␮g/l according to the United State Environmental
ore, electroplating, paper pulp production, leather tanning, metal- Protection Agency (USEPA) [3]. In surface waters in Iran, levels up
lurgical plants, nuclear power plants, pigment fabrication, timber to 84 ␮g/l have been reported.
treatment and petroleum refining. Chromium exists in the environ- In recent years, widely used treatment technologies such
ment mainly in two stable states, i.e. trivalent chromium [Cr (III)] as chemical precipitation, flotation, electrochemical deposition,
membrane filtration, ion exchange and adsorption have been
developed for removing heavy metals from wastewater. How-
ever, adsorption with unique features such as high metal-binding
∗ Corresponding author. capacity, adsorbents resistance in a wide pH range, easy operating
E-mail addresses: hunesi@modares.ac.ir, hunesi@yahoo.com (H. Younesi). conditions and maintenance, lack of toxicity, availability and low

http://dx.doi.org/10.1016/j.ijbiomac.2015.07.022
0141-8130/© 2015 Elsevier B.V. All rights reserved.
432 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

cost, with the capability of sorbent recovery and reuse, is an effec- In the present study, CS–CA nanoparticles were synthesized in dif-
tive economic and relatively new technique that has been widely ferent weight ratios (w/w) of CS–CA (1:1, 1:2, 1:3, 1:4 and 2:1%).
used as the most versatile method [4]. Chitosan is an anti-bacterial, The CA solution was added drop-wise to that of the CS. Then as an
renewable biopolymer with a proved, attractive feature in sur- initiator, some 0.2 mmol of KPS was added to this CS–CA in a poly-
face adsorption and characteristics such as high biodegradability, merization vessel; after formation of free radicals CA crosslinked
high biocompatibility, non-toxicity, biodhesion, immunogenicity with CS, as the CA could produce esterification with the hydroxyl
and the ability to form metal ion chelation with both cationic groups and transamidation with the amino groups of CS to form the
and anionic compounds [5]. Chitosan can be produced artificially crosslink. After refluxing at 70 ◦ C for 2 h under N2 protection, the
using biotechnological techniques in growth medium in massive transparent mixture was held in an ice bath for 1 h. Then the white
amounts as a result of producing chitin from shellfish and crus- product of CS–CA nanoparticles was collected by centrifugation
tacean exoskeleton wastes and external walls of some fungi. The (236HK, Hermle, Germany) for 40 min at 4000 rpm. The particles
recent efforts have focused on chemical modification of chitosan thus obtained were freeze-dried for 36 h using a freeze–dry system
using the reactive activities of primary and secondary hydroxyl- (Opergn, Korea). The CS–CA nanoparticles were stored in a sealed
and amine groups to control the reactivity of the polymer and drying tube in the refrigerator at 4 ◦ C prior to use.
improve its capacity, selectivity and range of pH for optimal adsorp-
tion. Various physical or chemical modification techniques have 2.3. Characterization
been used to increase the adsorption kinetics [6], by obtaining con-
ditioned polymer forms such as powder, nanoparticles and gels The Fourier transform infrared (FT-IR) spectroscopy (Shimad-
(membranes, sponge, beads, fibers or hollow fibers). Large surface zou, FT-IR1650 spectrophotometer, Japan) was used to identify
area and small particle size favor the adsorption of heavy metal various functional groups of the samples by using KBr pellets made
ions. from a mixture of the powder of each sample and assessed within
The objective of the present study was to modify chitosan with the range of 400–4000 cm−1 . The average diameter of the CS–CA
carboxyl functional groups using citric acid as a crosslinking agent nanoparticles and particle size distribution was got by a Malvern
at different conditions (molar ratio of chitosan to citric acid). We Zetasizer analyzer (Malvern Instruments Ltd) and to analyze the
investigated the adsorptive removal of Cr (VI) from aqueous solu- size of the nanoparticles by using dynamic light scattering (DLS).
tion onto as-prepared carboxylated nanochitosan with different A scanning electron microscope (SEM, KYKY-EM3200, China) was
process parameters such as pH, temperature, contact time, adsor- used to analyze the surface morphology of nanochitosan. The sam-
bent dose and initial concentration of Cr (VI). The experimental data ples were dispersed on a double-faced carbon conductive tape and
were analyzed using the pseudo-second-order kinetic equations then covered with gold film (KYKY-SBC12, China). The X-ray pow-
and the Langmuir, Freundlich and Redlish–Peterson isotherms. der diffraction patterns of CS and synthesized CS–CA nanoparticles
Finally, the reusability of carboxylated nanochitosan was evaluated were obtained by a X’Pert MPD (Philips, Netherlands) diffractome-
after three adsorption–desorption cycle. ter. The X-ray source was Cu-K␣ radiation (40 kV, 30 mA) in the
2 range of 5–70◦ . The spectra were recorded at a scanning rate
of 2.4◦ min−1 . The crystallinity index (CI) was computed using the
2. Materials and methods
following equation:
2.1. Materials I110 − Iam
CI = × 100 (1)
I110
The starting materials for preparing the chitosan–citric acid
where I110 and Iam are the maximum intensity of the diffraction at
(CS–CA) nanoparticle were chitosan polymer (CA) with high
plane 110 and the intensity of the amorphous diffraction, respec-
molecular weight (Mr = 600,000), a deacetilation content at 80%
tively. Results showed that the crystallinity index increases from
purchased from Sigma-Aldrich Chemical Co. (UK) and potassium
18 to 62% when CA was crosslinked with CS.
persulfate (K2 S2 O8 , KPS), acetic acid (AA) and citric acid (CA)
X-ray photoelectron spectroscopy (XPS) spectra were obtained
(anhydrous 99%) purchased from the Merck company (Darmstadt,
with a 8025-BesTec twin anode XR3E2 X-ray source system
Germany). Hydrochloric acid (HCl) and sodium hydroxide (NaOH)
(Germany). XPS data were taken using achromatic Mg-K␣
used for pH adjustment were purchased from the Merck Company
(1253.6 eV) and Al-K␣ (1486.6 eV) X-ray source, which operated at
(Darmstadt, Germany). Potassium dichromate salt (K2 Cr2 O7 ) for
15 kV. The XPS spectra taken from CS and CS–CA nanoparticles were
adsorption experiments was prepared by diluting 1000 mg/l of a
deconvoluted to obtain the contribution of the different chemical
standard solution of Cr (VI), purchased from the Merck Company
bondings involved in the materials studied. It is worth noting that
(Darmstadt, Germany), using de-ionized water. The pH of each
the charging effect was compensated by using N1s corresponding
heavy metal solution was adjusted by adding diluted NaOH (1 M) or
to C N bonding that appeared at the binding energy of 400 eV. The
HCl (1 M) using a pH meter (CyberScan, Singapore). The chemicals
spectra were acquired at low pass energy to enhance their resolu-
used were of analytical grade and not purified further.
tion. Savitzky–Golay filtering was performed to smooth the spectra
[8,9].
2.2. Preparation of chitosan nanoparticles
2.4. Adsorption experiments
To prepare CS–CA nanoparticles was done according to the
methods of Heidari et al. [7] but using anhydrous CA as a crosslink- A stock solution of Cr (VI) was prepared in a 1 L volumetric flask
ing agent. CS was dissolved in an aqueous solution of 1% AA in by dissolving 2.830 g of potassium dichromate in de-ionized water
a 250 ml, doubleneck, round-bottom flask and a condenser, then and diluted up to mark. The sample solutions were transferred into
stirred with a magnetic stirrer at 150 rpm for 3 h under nitrogen gas polyethylene bottles, which were stored in the refrigerator at 4 ◦ C
atmosphere until a complete dissolution of CS to a clear and trans- for future use. The effect of weight ratios, pH, contact time, adsor-
parent solution. The pH value of the above solution was adjusted bent dose, Cr (VI) ion concentration, adsorption kinetics, adsorption
to 4, by adding 1 M HCl at ambient temperature, and stirred mag- isotherm, thermodynamics of adsorption in a batch system and
netically for 7 h at 70 ◦ C. The CA was also dissolved in de-ionized regeneration studies on CS–CA nanoparticles were investigated in
water and the pH of the solution adjusted to 4 by adding 1 M NaOH. this study. The experiments were replicated three times in order to
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 433

carry out the statistical analysis. The statistical analysis of the each namely Langmuir, Freundlich and Sips were applied. The Langmuir
stage was carried out on a Microsoft-Excel 2010 spreadsheet. model assumes that (i) the adsorption is monolayer, (ii) there are a
fixed number of adsorbates places with equal free energy of adsorp-
2.4.1. Effect weight ratio of CS–CA nanoparticles on Cr (VI) tion onto the adsorbent surface, (iii) it has a homogeneous surface
removal structure with equally accessible sites, (iv) the adsorption process is
The effect of CS–CA nanoparticles with different weight ratios reversible and (v) no interactions between adsorbed species takes
(1:1, 1:2, 1:3, 1:4 and 2:1) on the removal of Cr (VI) was studied place. The nonlinear form of Langmuir equation is expressed in the
in different flasks containing 25 ml Cr (VI) solution (50 mg/l) at pH following equation [11]:
5.0.
qm bce
qe = (4)
1 + bCe
2.4.2. Effect of contact time on Cr (VI) removal
The effect of different contact times (10, 20, 40, 60, 110 and where qm is the maximum adsorption capacity in mg/g and b the
120 min) was studied on the removal of Cr (VI) by optimized CS–CA Langmuir constant in l/mg related to surface adsorption energy
nanoparticles in different flasks containing 25 ml Cr (VI) solution and strength of adsorption joint sites. The Freundlich isotherm
(50 mg/l) at a dose of 2 g/l, pH 4 and ambient temperature. model describes adsorption on heterogeneous systems. This model
is expressed as follows:
2.4.3. Effect of pH on Cr (VI) removal 1/n
The effect of different pH (2, 3, 4, 5 and 6) was studied on the qe = Kf Ce (5)
removal of Cr (VI) by optimized CS–CA nanoparticles in different where Kf is the Freundlich constant related to adsorption capacity
flasks containing 25 ml Cr (VI) solution (50 mg/l) at a dose of 2 g/l (mg/g), the larger value of Kf the greater adsorption capacity and n
and at ambient temperature (25 ◦ C). the adsorption intensity, called heterogeneity factor and its value
is in the range of 0–1. The more heterogeneous the surface, the
2.4.4. Effect of dose of CS–CA nanoparticles on Cr (VI) removal closer the 1/n value is to zero [12]. The Redlich–Peterson adsorp-
The effect of different adsorbent doses (0.5, 1, 1.5, 2, 2.5, 3 and tion isotherm is the combination of Langmuir and Freundlich and
3.5 g/l) was studied on the removal of Cr (VI) by optimized CS–CA contains three empirical coefficients A, B and g which derive from
nanoparticles in different flasks containing 25 ml Cr (VI) solution the limiting behavior of the equation. It follows the Freundlich
(50 mg/l) at pH 4 and ambient temperature. isotherm at low sorbate concentrations and thus does not obey
Henry’s law, while it predicts a monolayer adsorption capacity
2.5. Equilibrium and kinetic of adsorption characteristic of the Langmuir isotherm at high adsorbate concen-
trations. The Redlich–Peterson equation may be expressed by the
The equilibrium and kinetics adsorption experiments were car- following equation:
ried out in 50-ml Erlenmeyer flask containing 25 ml of the metal ion
solution with different concentration of Cr (VI) solution (10, 30, 50, ACe
qe = (6)
70, 90 and 110 mg/l) prepared by appropriate dilutions of the stock 1 + BCeG
solution. After adjustment of pH at 5, a 5 g/l of the adsorbent was
added. The effect of different temperature (25, 35 and 45 ◦ C) was where A (L/g) and B (L/mg)ˇ are the Redlich–Peterson empirical
studied on the removal of Cr (VI) by optimized CS–CA nanoparti- coefficients and g is the exponent, that lies between 1 and 0 [13].
cles in different flasks containing 25 ml Cr (VI) solution (50 mg/l) at The controlling mechanism of adsorption process in adsorption
a dose 2 g/l and pH 4. All flasks were agitated on a shaker at 150 rpm kinetic models, e.g., a mass transfer chemical reaction, provides
at room temperature and contact time of 120 min, until equilibrium a great deal for understanding the adsorption process of Cr (VI)
was reached. A 1 ml sample from each flask was taken after inter- onto CS–CA nanoparticles in a batch process. The pseudo-first-
vals of 10, 20, 40, 60, 110 and 120 min, centrifuged at 4000 rpm for order and pseudo-second-order equations were employed for the
40 min at ambient temperature and the residual Cr (VI) measured uptake of Cr (VI) by CS–CA nanoparticles to test all experimental
spectrophotometrically in the supernatant. The atomic absorption data. The advantage of these equations is easy to use and analyze
spectrometry with flame (AAS, Aanalyst 400, Perkin Elmer, USA) the experimental data, involves the calculation of the rate constant
was used to measure the concentration of Cr (VI) with a linear cal- and equilibrium adsorption capacity, estimate the initial adsorp-
ibration curve obtained at concentration ranges varying from 1 to tion rate; all of which can be used to explain rate mechanism
10 mg/l for Cr (VI). The equilibrium adsorption capacity (qe ) onto of adsorption process. The linear form of pseudo-first-order and
adsorbents was computed by the following equation: pseudo-second-order equations can be expressed as follows:

(Co − Ce ) × V k1
qe = (2) log (qe1 − qt ) = log qe1 − t (7)
m 2.303

where qe is the amount of metal ion adsorbed at equilibrium per t 1 t


= + (8)
mass unit of adsorbent at mg/g, Co initial concentration of metal qt k2 q2e2 qe2
ions in mg/L solution, Ce the equilibrium concentration of metal
ions in solution in mg/L, V the volume of the solution in ml, and m where qe and qt are the amounts in mg/g of adsorbed metal ion onto
the adsorbent weight in g. Furthermore, metal removal efficiency adsorbent, at equilibrium and time t, respectively. Kinetic studies
was calculated by the following equation: were carried out using different initial Cr (VI) concentrations in
110 mg/l (10, 30, 50, 70, 90) but constant amount of 0.15 g CS–CA
Co − Ct nanoparticles (ratio of 3–1 CS–CA) with 50 ml working volume at
R= × 100 (3)
Co pH 3.
where R represents the adsorption efficiency percentage of metal
ions, and Ct the concentration of the metal ions at time t in mg/L 2.6. Thermodynamic of adsorption
[10].
In order to evaluate equilibrium adsorption data for adsorbing The effect of temperature (25, 35 and 45 ◦ C) on the adsorp-
Cr (VI) from aqueous solution, three nonlinear isotherm models, tion process was also evaluated in order to understand the
434 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

thermodynamic parameters such as the Gibbs free energy (G◦ ), Particle size and surface morphology of CS–CA nanoparticles
enthalpy (H◦ ) and entropy (S◦ ), using the following equation: with different weight ratios of CS–CA was compared by SEM
images. The pure CS particle showed a continuous and smooth
G◦ = −RT ln Kd (9)
surface (figure not shown), indicating a smooth integrity of sur-
where R is the gas constant (8.314 J mol−1
K), T the absolute tem- face structure. The effect of weight ratio of CS–CA suggested that
perature in Kelvin and Kd the distribution coefficient at different the cross-linking reaction between CS and CA leading to increas-
temperature levels, which can be obtained from the following equa- ing stacked lentil-like nanoparticles with higher content of the
tion: nanoparticles. The lentil-like shape on the images also suggested
that the CS–CA nanoparticle scattered in the same shapes and
Co − Ce V
Kd = × (10) forms, which allow controlling the size of nanoparticles. The SEM
Ce m
images (Fig. 2a–e) reveals lentil-like shape of CS–CA nanopar-
The standard Gibbs free energy was computed at different tem- ticles, which indicated the formation of more uniform particles
peratures according to the following equation [14]: in these cases. The CS–CA nanoparticle with 2: 1 weight ratio
(Fig. 3c) showed lentil-shaped structure, while several fiber-like
G◦ = H ◦ − TS ◦ (11)
shape can be observed in this image, which might be due to
where the values of change in enthalpy (H◦ in J mol−1 ) and change molecular irregularity increasing with an increase of CS content in
in entropy (S◦ in J mol−1 K), respectively, were calculated from CS–CA nanoparticles. This result was inconsistent with Heidari et al.
linear coefficients obtained by regression when a plot of lnKd versus [7], who reported that a spherical and homogeneous shape was
1/T should be a straight line using the van’t Hoff equation: observed in the chitosan-methacrylic acid (CS–MAA) nanoparticles.
Infrared spectroscopy is introduced as a valuable tool and a
S ◦ H ◦
ln Kd = − (12) well stabilized experimental technique for the analysis of struc-
R RT
ture and conformational changes in biopolymers. The FT-IR spectra
The activation energy of adsorption was obtained by the follow- of pure CS and CS–CA nanopartilces are presented in Fig. 3. Com-
ing Arrhenius equation: paring the results of FT-IR spectra of CS with CS–CA suggested
Ea that the presence of carboxyl ( COO− ), amide I ( NH COO− ) and
ln K2 = ln A − (13) amine II ( NH2 ) functional groups in the spectra of all weight
RT
ratios of CS–CA confirms the formation of nanoparticles. It implies
where k2 is the rate constant from pseudo-second-order kinetic
that chitosan was covalently bonded with carboxyl group of cit-
model in g mg−1 min−1 , Ea is the Arrhenius activation energy of
ric acid through cross linkage reaction. Overlap of the symmetrical
adsorption and A the Arrhenius A factor.
peaks at 3400 and 3300 cm−1 are assigned to the stretching of
free amine ( NH2 ) and hydroxyl ( OH) functional groups in the
2.7. Desorption experiments ring of CS. The peaks at 2920 and 2890 cm−1 are due to C H
stretching in the CS ring. However, the peaks at wave numbers
In this study nanochitosan was made for proving it as an effec- 3470 and 618 cm−1 are related to the OH stretching and bend-
tive and cheap adsorbent, but for use in wastewater treatment, it is ing vibration in the CS ring, respectively. In addition, a broadening
necessary to generate this inexpensive adsorbent through a tech- of the 3000–3500 cm−1 peaks, attributed to OH of molecular
nically practical and economically feasible process. The generation water along with 1640 cm−1 attributed to H O H bending mode
and reuse capacity of the adsorbent was examined in three succes- of water molecule, in the nanoparticles structure is in evidence.
sive adsorption/desorption cycles. An adsorption of metal ion was The peak at about 1380 cm−1 is attributed to C O stretching of
carried out in a 10 ml of Cr (VI) solution at 100 mg/l. Then, metal the primary alcohol group in the CS ring. The last peak at about
ions in the supernatant, obtained from centrifuging the sample for 1080 cm−1 is attributed to the C O C stretching of the glycosidic
10 min at a speed of 4000 rpm, was analyzed with AAS to deter- linkage between chitosan monomers. On the contrary, the peak
mine the amount of metal ion left unadsorbed. Desorption was at about 1580 cm−1 , related to the NH2 bending vibration in the
performed using eluent of 1 and 2 M sodium chloride (NaCl) with CS ring, disappeared in the FT-IR spectra of CS–CA nanoparticles.
each fraction volume of 10 ml on a stirrer with a speed of 100 rpm The peak at 1750 cm−1 is due to the stretching of C O from the
for 1 h at room temperature. After that, metal ion in each fraction carbonyl group remaining from CA, which was formerly effective
of elution was analyzed with AAS to calculate the Cr (VI) concen- in the adsorption of Cr (VI). However, the main peaks in the ring
tration. The recovery percentage of metal ion was computed with 1530–1640 cm−1 are attributed to amide I in the CS–CA nanoparti-
the following equation [15]: cles.
Amount of metal ions desorbed The X-ray powder diffraction patterns of CS and CS–CA nanopar-
Metal recovery = × 100 (14) ticles are shown in Fig. 4. Three major peaks in in the XRD pattern of
Amount of metal ions adsorbed
chitosan were observed at 2 = 8.58◦ , 11.61◦ and 19.21◦ , indicating
3. Results and discussion the low degree of crystallinity of chitosan. CA was crosslinked with
CS that may be the cause of the increase in the degree of crystallinity
3.1. Characterization study of the CA–CS nanoparticles, shown on XRD results (Fig. 4). Further-
more, the XRD pattern of CS particles in the same figure shows that
In the present study, a milky white color, powder form of CS–CA there is a broad peak low intensity at 2 = 22.62◦ , which indicates an
nanoparticles was obtained, which was insoluble in water. The amorphous nature of the CS particles. However, the XRD pattern of
particle size and the average particle size distribution for CS–CA the CS–CA particles in Fig. 4 shows that, as CA was crosslinked with
nanoparticles with different molar ratio of CS–CA (1:1, 1:2, 1:3, CS, peak appear both at 2 — о ◦
— 20·12 and 2 = 29.48 . The absence of
1:4 and 2:1) were examined by zetasizer, shown in Fig. 1. As can be a peak at 2 = 22.62◦ for CS–CA nanoparticles may be due to their
seen from Fig. 1a, the smallest particle size (403.6 mm) corresponds crystallinity structure.
to nanochitosan with 1:2 ratio of CS–CA. The largest particle size As seen in Fig. 5a, C1s spectra corresponding to N C O, O C O,
of CS–CA nanochitosan of about 1084 mm was observed with 1:3 C O and C N present also in CS were identified and quantified with
ratio of CS–CA (Fig. 1e). It was revealed that the size of nanoparticles percentages 6.66, 20, 20 and 53.33%, respectively. From the analy-
increased with an increasing weight ratio of CS–CA. sis of C1s, taken from CS–CA–NP, the contribution of O C O was
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 435

Size distribution(s) Size distribution(s)

% in class

% in class
15 15

10 10

5 5

5 10 50 100 500 1000 5 10 50 100 500 1000


Diameter (nm) Diameter (nm)

(a) (b)
Size distribution(s) Size distribution(s)
% in class

% in class
60

60

40

40

20
20

5 10 50 100 500 1000 5 10 50 100 500 1000


Diameter (nm) Diameter (nm)

(c) (d)
Size distribution(s)

80
% in class

60

40

20

5 10 50 100 500 1000


Diameter (nm)

(e)
Fig. 1. Effect of chitosan (CS) to citric acid (CA) ratio on particle size of CS–CA nanoparticles; (a) 1:1 (dave. = 554.2 nm), (b) 1:2 (dave. = 403.6 nm), (c) 1:3 (dave. = 1084 nm), (d)
1:4 (dave. = 749 nm), and (e) 2:1 (dave. = 1037 nm).
436 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

Fig. 2. Scanning electronic microscopy (SEM) images of as-prepared CS–CA nanoparticles with (a) 1:1, (b) 1:2, (c) 1:3, (d) 1:4, (e) 2:1 ratios of chitosan (CS) to citric acid (CA).
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 437

(a)

(b)

(c)

(d)
Transmittance, %

(e)

(f)

4000 3600 3200 2800 2400 2000 1600 1200 800 400

Wavenumber, cm-1
Fig. 3. FT-IR spectra of (a) chitosan, as-prepared CS–CA nanoparticles with (b) 1:1, (c) 1:2, (d) 1:3, (e) 1:4, (f) 2:1 ratios of chitosan (CS) to citric acid (CA).

verified in addition to all above mentioned components, confirming CS and CS–CA–NP represent the components involved correspond-
the formation of CS–CA–NP, as seen in Fig. 5b. The contribution ing to C OH (57.14% for CS and 54.54% for CS–CA–NP), C O C
percentages of O C O, N C O, O C O, C O and C N were found (35.71% for CS and 27.27% for CS–CA–NP) and O C (7.14% for CS
to be 8.58, 14.24, 17.15, 42.88 and 17.15%, respectively, which are and 18.18% for CS–CA–NP). The increase in the percentage of O C
consistent to the chemical structure of CS–CA–NP. A similar trend bonds present in the nanoparticles is in good agreement with what
was applied for the analysis of the peaks in N1, which showed is expected to be observed.
the presence of O C NH and C NH2 in both materials studied
(Fig. 5); in percentages of 33.33, 66.66 and 83.33, 16.67 for CS, and 3.2. Effect of weight ratio CS–CA on adsorption
CS–CA–NP, respectively. This variation is another quantitative con-
firmation on the formation of CS–CA–NP. Furthermore, as seen in The effect of the weight ratio was examined at constant tem-
Fig. 5, the deconvolutions of O1 peaks in the spectra taken from perature of 25 ◦ C and pH 4, with Cr (VI) concentration at 50 mg/l,
438 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

500 in the pH from 2 to 3, the removal percentage increased from 52.89


to 86.83% and then it decreased from 82.24 to 73.95% by an fur-
ther increase in pH from 4 to 6, respectively. The changes in the
400 (b) adsorption rate were proportional to pH changes, so the pH impact
on adsorption is due to its impact on the speciation and surface
Intensity, a.u.

charge of both the adsorbate and the adsorbent. From a thermo-


300 dynamic standpoint, pH determines the stability of chromium in
aqueous solution and is considered an outstanding and effective
parameter in the adsorption process.
200
In aqueous solution, Cr (VI) occurs mainly as chromate
(a)
(CrO4 2− ), dichromate (Cr2 O7 2− ), hydrogen chromate (HCrO4 − ),
chromic acid (H2 CrO7 ), hydrogen dichromate (HCr2 O7 − ), trichro-
100
mate (Cr3 O10 2− ) or tetrachromate (Cr4 O13 2 ). However, these
species depend on both the pH and the analytical concentration
of chromium; for example, when the concentration of Cr (VI) in
0
surface waters is less than 5 ␮g/l, only HCrO4 − and CrO4 2− can be
5 10 15 20 25 30 35 40 45 50 55 60
found, while HCr2 O7 − is predominant at pH < 6, while CrO4 2− ion is
2Theta, degree predominant at pH > 7. Both HCrO4 − and H2 CrO7 are the predomi-
nant species in acidic conditions, while the formation of Cr3 O10 2− ,
Fig. 4. X-ray diffraction patterns of (a) chitosan and (b) as-prepared CS–CA nanopar-
and Cr4 O13 2− species of chromium can occur in a strongly acidic
ticles with 1:1 ratio of chitosan–citric acid.
solution. Because of the importance of stability and formation
of chromium species in various pH conditions, our investigation
CS–CA nanoparticle dose of 2 g/l, contact time 60 min including demonstrated that the adsorption of Cr (VI) at lower pH value was
various weight ratios of CS–CA nanoparticles (1:1, 1:2, 1:3, 1:4 more favorable, being more efficient. As shown in Fig. 6, the maxi-
and 2:1) and CS in order to choose the best candidate for as- mum adsorption capacity and removal percentage of Cr (VI) occurs
prepared nanoparticle on the basis of the maximum Cr (VI) uptake at pH 3. This could be due to the fact that Cr (VI) in acidic pH is found
capacity and removal efficiency (figure not shown). The results in two forms, in HCrO4 −1 and Cr2 O2 −2 ions that are in equilibrium.
revealed that the removal percentage of Cr (VI) reached was 71.28, Acidic pH also causes a protonation of the NH− and COO functional
76.35, 83.54, 82.30, 76.35 and 61.75 for weight ratios of 1:1, 1:2, groups on the surface of nanochitosan, and as a result make the
1:3, 1:4, 2:1 and CS, respectively, after 60 min contact time. The adsorbent surface positively charged. Therefore, these two ions can
weight ratio study showed that the best ratio for Cr (VI) adsorp- be adsorbed on oppositely charged adsorbent surfaces and adsor-
tion on CS–CA nanparticle is 1:3. The Cr (VI) appears to adsorb bates at low pH. This means that due to the negatively charged
onto CS even when CA was not crosslinked with CS. However, the HCrO4 −1 and Cr2 O2 −2 ions and the protonated surface functional
removal efficiency of Cr (VI) by CS–CA nanoparticle was observed groups of the adsorbent and the electrostatic tension between pos-
to be about 10–20%. This indicates that the more positively charged itive and negative ions, the maximum adsorption occurred at pH
nanoparticles on the surface may interact with Cr (VI) ions by elec- 3; so this observed fact was used for the rest of the study. At pH
trostatic force and also providing evidence for the possibility of lower than 3, nanochitosan dissolution occurs slowly and the bulky
forming a Cr (VI)/CS–CA surface complex on the CS–CA nanoparti- natured predominant species of Cr3 O10 2− , and Cr4 O13 2− thus have
cles. a negative effect on the efficiency of adsorption. At pH higher than
3, because ion HCrO4 −1 turns into ion CrO4 −2 and the adsorbent
3.3. Effect of contact time on adsorption surface is deprotonated, adsorption of CrO4 −2 is restricted due to
the presence of negatively charged ions and competitive − OH, the
For designing an efficient filtration system that is economically low positive surface charge density of the adsorbent and the weak-
affordable, determining the required time to reach equilibrium on ening of the electrostatic forces of attraction between the parallelly
chromium adsorption process is critical. Therefore, the effect of charged adsorbate and adsorbent. The results of the study are con-
contact time was carried out at constant temperature of 25 ◦ C and sistent with results obtained in previous researches [17,18].
pH 4, with Cr (VI) concentration of 25 mg/l, CS and CS–CA nanopar-
ticle dose of 2 g/l (in 1:3 weight ratio) and at various time intervals 3.5. Effects of adsorbent dose on adsorption
(0, 10, 20, 40, 60, 110 and 120 min), in order to determine the opti-
mal time for the maximum removal efficiency (figure not shown). The adsorbent dose is an important parameter that strongly
The removal efficiency rate was very rapid in the first 10 min of affects the adsorption capacity. Fig. 7 shows the effect of Cr (VI)
contact time, but with increasing contact time the removal effi- adsorption onto CS–CA nanoparticles (with 1:3 weight ratio) per-
ciency of Cr (VI) increased slowly until the equilibrium time for formed with various doses of adsorbent (0.5, 1, 1.5, 2, 2.5, 3,
both the CS and CS–CA nanoparticle was 60 min, while there was 3.5 g/l) while holding other operating parameters constant dur-
no significant change in adsorption removal efficiency from 60 to ing 60 min contact time, pH 3 and 50 mg/l concentration of Cr
120 min of contact time. At the equilibrium situation, the max- (VI), 25 ◦ C temperature and 150 rpm agitation speed. Experiments
imum removal efficiency of 83.54 and 61.75% for CS and CS–CA showed that by increasing the adsorbent dose form 0.5 to 3 g/l,
nanoparticles, respectively, was attained. This finding is in accor- the removal efficiency increased from 38.51 to 83.33%. But by con-
dance with previous research [16]. In general, the time required to tinuing to increase the adsorbent dose to 3.5 g/l, the percentage
reach equilibrium was less in this research as compared to previous of removal efficiency remains constant and no change is observed
studies. (Fig. 7).
With increasing the dose of adsorbent, the active sites on the
3.4. Effect of pH on adsorption adsorbent surface increases over the number of constant metal
ions, resulting in increasing the metal removal percentage [19]. But
Fig. 6 shows the effect of initial pH of the solution on the adsorp- after a 3 g/l dose of adsorbent, the removal percentage decreases
tion of Cr (VI) was studied in the pH range of 2–6. With an increase due to the fact that the increase in adsorbent doses enhances
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 439

Fig. 5. X-ray photoelectron spectroscopy characterization: C1s of CS (a) and CS–CA nanoparticles (b), N1s of CS (c) and CS–CA nanoparticles (d), and O1s of CS (e) and CS–CA
nanoparticles (f) as-prepared CS–CA nanoparticles with 1:1 ratio of chitosan to citric acid.

the chances of collision between the adsorbent nanoparticles, of adsorbent decreases as the adsorbent dose increases (Fig. 7). This
leading to an aggregation of the adsorbent nanoparticles, besides means that the Cr (VI) concentration in solution drops to a lower
decreasing the surface area of the adsorbent [20]. Unlike the per- value at higher adsorbent dose and the adsorption system reaches
centage of removal, the adsorption capacity decreases from 38.51 to an equilibrium state at a lower value of qe , indicating that the active
3.88 mg/g. In this case the amount of Cr (VI) adsorbed per unit mass sites remain unsaturated.
440 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

40 50
100

Removal efficiency, %
80 40

60
30
Adsorption capacity, mg/g

30

q e, mg/g
40

20 20 Experimental data
Langmuir
20 0 Freundlich
2 3 4 5 6 10 Redlich-Peterson
pH
0
10 0 10 20 30 40 50 60

C e, mg/l

Fig. 8. Langmuir, Freundlich and Redlich–Peterson isotherm models for the Cr (VI)
adsorption onto CS–CA nanoparticles and at dose 0.15 g/50 ml, pH 4, contact time of
contact time of 60 min, agitation speed of 150 rpm and temperature of 25 ◦ C.
0
2 3 4 5 6
removal increases dramatically in the range from 53.42 to 94.46%
pH
at the initial Cr (VI) concentration of 10–70 mg/l. However, the Cr
Fig. 6. Effect of pH on adsorption capacity and removal percentage (inside) of Cr (VI) (VI) removal percentage reaches 90.13 and 88.93 with the initial Cr
adsorption onto CS–CA nanoparticles with 1:1 ratio of chitosan to citric acid and at (VI) concentration of 90 and 110 mg/l, respectively, going through
dose of 0.15 g/50 ml, initial Cr (VI) concentration of 50 mg/l, contact time of 60 min, a slight decrease compared with the 70 mg/l. The explanation lies
agitation speed of 150 rpm, and temperature of 25 ◦ C. in the response of the removal percentage, which increased with
an increase in the Cr (VI) concentration and decreased after satura-
3.6. Effect of initial concentration of Cr (VI) on adsorption rate tion above a certain concentration. This means that by increasing
the amount of available Cr (VI) in the solution, the collision rate of
Effect of initial concentration of Cr (VI) was investigated with metal ions with available functional groups on the adsorbent sur-
fixed pH 3, contact time of 60 min, adsorbent dose of 3 g/l and tem- face, responsible for adsorption, increase which causes an increase
perature of 25 ◦ C over a concentration range of 100–110 mg/l. The of effective diffusion and in the number of bonds in order to form
results indicate that the removal percentage decreases as the initial the complex with metal ions. At higher Cr (VI) concentrations,
concentration of Cr (VI) is decreased (figure not shown). The Cr (VI) adsorption sites on the adsorbent surface are occupied by Cr (VI)
and also the adsorbent has a limited number of active sites, thereby,
consequently leading to a decrease in the removal percentage [21].
50
100 3.7. Adsorption isotherms
Removal efficiency, %

80 In order to investigate different isotherms and their capability to


40 compare with equilibrium data, a theoretical graph is plotted in the
60 form of the amount of Cr (VI) adsorbed onto CS–CA per unit mass of
CS–CA (qe ) against the equilibrium concentration of Cr (VI) in the
40
solution (Ce ). Fig. 8 shows the experimental equilibrium data using
30 Langmuir, Freundlich and Redlich–Peterson and also the calculated
20
qe, mg/g

isotherm parameters by nonlinear models are shown in Table 1,


0
according to which the correlation coefficients for Langmuir and
Redlich–Peterson isotherm are higher (R2 > 0.99) than those of the
0.025
0.05
0.075
0.1

0.15
0.125

0.175

20 Freundlich, indicating them to be suitable isotherms for the data.


The maximum adsorption capacity was obtained with 106.15 mg/g
CS-CA nanop article, g/50 ml using the Langmuir model. Based on the data in Table 1, KF
and n values are 16.29 (mg/g)(L/mg)1/n and 2.30, respectively. KF
10
Table 1
Langmuir and Frendlich and Redlich–Peterson parameters of Cr (VI) adsorption onto
CS–CA nanoparticles.

0 Models Parameters
0.025 0.05 0.075 0.1 0.125 0.15 0.175 Langmuir qm (mg/g) b (L/mg) R2
106.15 0.081 0.9932
CS-CA nanoparticle dose, g/50 ml
Freundlich KF (mg/g)(L/mg)1/n n R2
Fig. 7. Effect of adsorbent dose on uptake capacity and removal efficiency onto 16.29 2.30 0.9740
CS–CA nanoparticles with 1:1 ratio of chitosan to citric acid at initial pH of 4, initial
Redlich–Peterson A (L/g) B (L/mg)ˇ g R2
Cr (VI) concentration of 50 mg/l, contact time of 60 min, agitation speed of 150 rpm,
8.65 0.084 0.992 0.9933
and temperature of 25 ◦ C.
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 441

represents the adsorption capacity of adsorbed Cr (VI) and 1/n Table 2


Pseudo-second-order parameters of Cr (VI) adsorption onto CS–CA nanoparticles.
shows the effect of concentration on the adsorption capacity
exp.
representing adsorption intensity. In order for adsorption to be rec- Concentration, qe (mg/g) qe2 (mg/g) k2 (g mg−1 min−1 ) R2
ognized as the best, the values of n in range 2–10 indicate good, mg/L
1–2 moderately difficult, and <1 poor separation. Table 1 shows 30 8.69 8.61 0.076 0.9908
that the n value is found between 2 and 10, showing good sep- 50 13.60 13.26 0.057 0.9919
aration of Cr (VI) from the CS–CA nanoparticles [22]. In addition, 70 22.04 21.80 0.046 0.9938
90 27.40 27.15 0.068 0.9973
Table 1 shows that the value of ˇ is close to unity, indicating homo-
110 28.68 28.94 0.072 0.9998
geneity of the adsorbent surface and that the Redlich–Peterson
isotherm approaches the Langmuir adsorption and not the Freund-
lich isotherm [23]. obtained from the plot of the pseudo-first-order-equation, Eq. (7),
are presented in Table 2. The pseudo-second-order equation, Eq.
(8), shows good agreement between the calculated and experi-
3.8. Adsorption kinetics
mental equilibrium adsorption capacities, revealed by the very high
correlation coefficients obtained (R2 > 0.99), as the values show in
The kinetics of adsorption is the most important factor for
Table 2. The increase in the value of qe is linearly proportional to the
any adsorption system design in predicting the adsorption rates
increase in the initial concentration. The equilibrium rate constant
at which adsorption takes place in a given system. The adsorp-
of the pseudo-second-order model, k2 , decreases with an increase
tion kinetics depends on type and surface concentration of the
in the initial Cr (VI) concentrations of 30, 50, and 70 mg/L, but then
active sites and the surface area of an absorbent that controls the
the k2 increases at higher initial concentrations of 90 and 110 mg/L,
adsorption mechanism [24]. In order to investigate the controlling
indicating a tendency for the Cr (VI) ion from bulk liquid to the
mechanism and adsorption consistency data, experimental data
adsorbent surface with an increase in the initial concentration.
were applied with pseudo-first-order and pseudo-second-order
adsorption kinetic models [25]. It was found that the regression
coefficient of the pseudo-first-order equation was small (figure 3.9. Adsorption thermodynamics
not shown), so the kinetics of Cr (VI) adsorption onto CS–CA did
not fit the pseudo-first-order model and hence was not charac- In order to describe the thermodynamic behavior of the adsorp-
terized by the diffusion-controlled phenomenon. In this case, the tion of Cr (VI) onto CS–CA nanoparticles, the thermodynamic
pseudo-second-order model was used for analyzing the experi- parameters, including the changes in G◦ , H◦ and S◦ were cal-
mental data in this study, wherein it is assumed that chemical culated at different temperatures (from 298 to 318 K) The results
adsorption may be the rate-controlling step in the adsorption pro- showed that the adsorption capacity decreased with increasing of
cess (with no involvement of mass transfer in solution) and the temperature, indicating that a lower temperature promoted the Cr
adsorption site occupancy rate is proportional to the square of the (VI) adsorption onto CS–CA nanoparticles. Fig. 10a shows the van’t
number of unoccupied sites and the analyte can be bound to differ- Hoff plot, considering a satisfactory correlation coefficient value
ent binding sites. The kinetic studies indicated that the adsorption of 0.9830. The values of the thermodynamic parameters for the
process was rapid during the initial 30 min, reaching equilibrium adsorption of Cr (VI) onto CS–CA nanoparticles were calculated
within 60 min, following the pseudo-second-order rate model. The from the Eqs. (10) and (11), listed in Table 3. The chemisorption
values of qe and k2 were calculated, when the curve of t/qt as func- usually has the heat value in the range of −80 to −800 kJ/mol, while
tion of t gave a straight line. A good relationship between t/qt and heat values between −5 and −40 kJ/mol refer to a physisorption of
t for the pseudo-second-order equation makes possible to confirm adsorption process [26]. The negative H◦ value of −110.29 kJ/mol
that the adsorption process is of chemical nature. The plots of t/qt indicates that Cr (VI) adsorption onto CS–CA nanoparticles is
against time for adsorption of Cr (VI) onto CS–CA for the pseudo- exothermic in nature. Hence, it seems that there is chemisorption
second-order equation is shown in Fig. 9 and the kinetic parameters involved. Furthermore, an exothermic nature of adsorption is prob-
ably related to the large number of the active sites on the surface of
CS–CA nanoparticles, which may improve the adsorption capacity.
8 The negative S◦ value of 365.33 J mol−1 K−1 suggests a decrease
in the disorder at the solid/solution interface during the adsorption
30 mg/l
50 mg/l of Cr (VI) onto CS–CA nanoparticles. The observed negative G◦
70 mg/l value of −1.42 kJ/mol at lower temperature confirms the feasibility
6 90 mg/l of the process and the spontaneous nature of adsorption processes.
110 mg/l
Likewise, the increasing G◦ values with increasing temperature
Pseudo-secon-order kinetic
explains a decrease in feasibility of nonspontaneous adsorption at
higher temperature. The results validate the theory that the thick-
t/qt

4 ness of the boundary layer decreases with increasing temperature,


due to the increased tendency of the Cr (VI) ions to escape from
the adsorbent surface to the solution phase, which results in a
decrease in the adsorption capacity as the temperature increases
2
[27]. Similar results were previously reported by Shahbazi et al.
[15]. Fig. 10b shows the Arrhenius equation, considering a satisfac-
tory correlation coefficient value of 0.9880. Table 3 shows that the
adsorption processes have an adsorption activation energy value
0
0 20 40 60
of 66.51 kJ/mol for Cr (VI) onto CS–CA nanoparticles. If the value of
activation energy of adsorption is between 20 and 80 kJ/mol, it sug-
Time, min gests chemical adsorption, when the adsorption energy is between
8 and 16 kJ/mol, the adsorption type can be explained by ion-
Fig. 9. Pseudo-second-order model plot of Cr (VI) adsorption onto CS–CA nanopar-
ticle at pH 3.0, contact time 60 min, dosage of 1.5 g/L, temperature of 25 ◦ C and exchange; and when the adsorption energy is lower than 8 kJ/mol,
agitation speed of 150 rpm. the type of adsorption can be considered as diffusion-controlled
442 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

4 to 69.4% in the second and to 56.0% in the third one. However,


the adsorption capacity of Cr (VI) removal after three cycles was
2 8.5%, whereas the results of adsorption and desorption of Cr (VI)
Experimental data
Van't Hoff equation
in 2 M NaCl decreased from 94.7% in the first cycle to 83.5% in the
third one, which showed reasonable regeneration capacity for the
0
adsorbent. The desorption mechanisms proposed were (i) electro-
lnKd, L/g

static interactions or van der Waals forces (physical adsorption)


-2 or (ii) covalent bonds (chemical adsorption containing chelating
and ion exchange) [28]. The using of NaCl to desorb Cr (VI) can
-4 be explained by the presence of Cl− ions with high electronegativ-
ity causing a competition to be created between Cl− and HCrO4 −
-6 ions and Cr2 O2 −2 in establishing links with the adsorbent that has
been protonated at pH 3 and is positively charged. This eventu-
ally weakens the Cr (VI) binding and pushes back the metal species
-8
(HCrO4 − and Cr2 O2 −2 ) mentioned above, causing Cl− to be replaced
3.10 3.15 3.20 3.25 3.30 3.35 3.40
in connection with CS–CA nanoparticles.
-1
1000/T, K
(a) 3.11. Comparison of CS–CA nanoparticles with other sorbents

In the present study, the effect of five parameters namely ini-


-1.2
tial solution pH, initial Cr (VI) ion concentration, contact time,
-1.4 adsorbent dose and temperature on adsorption were considered.
Experimental data
-1.6 Arrhenius equation Studies of the removal and recovery of Cr (VI) from aqueous
solution by different adsorbent have recently gained significant
-1

-1.8
lnK2, g.mg .min

attention. Although, difficult is inherent in direct comparisons


-2.0 with other literature data because of the different experimen-
-1

-2.2 tal conditions used in each study, it is important for readers


to consider the applicability of the research results to practice
-2.4
and that the design and methodology are consistent with the
-2.6 study purpose. Table 4 shows a comparison experimental study
-2.8 of maximum removal efficiency and uptake capacity and desorp-
tion efficiency of Cr (VI) ions by various adsorbents with respect
-3.0
to different initial solution pH, initial ion concentration, and
-3.2 absorbent concentration dose reported in the literatures. The
3.10 3.15 3.20 3.25 3.30 3.35 3.40 results exhibit that the maximum removal efficiency and uptake
-1 capacity and desorption efficiency of CS–CA nanoparticles for Cr
1000/T, K
(VI) ions were comparable to other corresponding adsorbents
(b) reported in the literatures. As can be seen from Table 4, varia-
Fig. 10. (a) van’t Hoff and (b) Arrhenius plots of Cr (VI) adsorption onto CS–CA
tions in removal efficiency and uptake capacity and desorption
nanoparticles at adsorbent dose of 1.5 g/L, pH 3.0, initial concentration of 70 mg/L, efficiency on different kinds adsorbents are associated with adsor-
contact time of 60 min and agitation speed of 150 rpm. bent properties, such as physical and chemical structure of the
adsorbents, metal-bonding functional groups, initial pH of the solu-
tion and initial Cr (VI) concentration, Cr (VI) species (i.e., CrO4 2− ,
transport, thus a physical adsorption process. The value of the Cr2 O7 2− , HCrO4 − , H2 CrO7 , HCr2 O7 − , Cr3 O10 2− , Cr4 O13 2− ) and
activation energy was positive and greater than 40 kJ/mol, indicat- adsorbent dose. Several nano-adsorbent such as polyacrylonitrile
ing the feasibility of the adsorption process being predominantly (PAN) nanofibers functionalized with amine groups (PAN-NH2)
chemical in nature. In fact, the values of H◦ , G◦ and Ea indi- [29], manganese oxide nanofibers (MONFs) [30], guar gum–nano
cated that the adsorption of Cr (VI) by CS–CA nanoparticles was a zinc oxide (GG/nZnO) [31], polypyrrole-titanium(IV) phosphate
chemisorption process. (PPy-TiP) nanocomposite [32], polyaniline/zeolite (PANI/zeolite)
nanocomposite [33], nano-hydrotalcite supported on supported
3.10. Desorption on silica (nano-HT/SiO2 ) [34], and maghemite/polydopamine
core/shell nanoparticles (MNP@PDA) [35] have been investigated
The results of desorption in successive adsorption–desorption for their Cr (VI) adsorption capacity. As can be seen from Table 4,
cycles were obtained by repeating three times using the same hydrotalcite [34], magnetotactic bacteria [36] and wheat bran (WB)
adsorbent. The adsorption capacity of Cr (VI) removal after three [37] and polyacrylonitrile nanofibers functionalized with amine
cycles was 22.7 mg/g, whereas the recovery of Cr (VI) in 1 M NaCl groups (PAN-NH2 ) [29] were used for Cr (VI) removal from aque-
solution it was observed as 82.3% in the first cycle, decreasing ous solution and the results showed that the removal efficiency

Table 3
Thermodynamic parameters of Cr (VI) adsorption onto CS–CA nanoparticles.

T (K) kd (L/g) van’t Hoff equation Arrhenius equation

G◦ (kJ/mol) H◦ (kJ/mol) S◦ (J mol−1 K−1 ) R2 Ea (kJ/mol) R2

298 5.68 −3.95 −303.19 −1004 0.9956 66.51 0.9880


308 0.069 6.09
318 0.0026 16.13
M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444 443

Table 4
Comparison of maximum removal efficiencies of Cr (VI) ion by various adsorbent with respect to various initial solution pH, initial Cr (VI) concentration, and adsorbent dose
reported in the literature.

Adsorbent pH Initial Cr (VI) Adsorbent Temperature Uptake Removal Desorption Ref.


concentration dose (g/l) (◦ C) capacity efficiency (%) efficiency (%)
(mg/l) (mg/g)

Polyacrylonitrile nanofibers 2 100 0.02 23 156 (>90) 10 [29]


functionalized with amine
groups
Magnenese oxide nanofibers 2 40 0.02 25 14.6 99.8 n.d. [30]
Wheat bran 2 200 50 23 4.53 51 n.d. [37]
Modified wheat bran 2 200 50 23 5.28 90 n.d. [37]
Guar gum-nano zinc oxide 7 25 1 25 55.56 98.63 n.d. [31]
Polypyrrole-titanium(IV) 2 200 0.2 n.d. 31.64 99 88 [32]
phosphate nanocomposite
Magnetotactic bacteria 6 50 30 (g wet cell l−1 ) 29 70.42 77 n.d. [36]
PANI/zeolite nanocomposite 2 50 0.2 30 24 100 n.d. [33]
Nano-hydrotalcite/SiO2 n.d. 4 1.0 n.d. n.a. 94.6 n.d. [34]
Hydrotalcite n.d. 4 1.0 n.d. n.a. 84.5 n.d. [34]
Shell of Swietenia mahagoni 2 250 0.08 59.87 47.61 99.9 17.44 [40]
Gelatin-impregnated-yeast 1–2 500 0.005 n.d. 500 100 n.d. [39]
Graphene 2 50 0.5 45 666.67 100 n.d. [38]
oxide-␣CD-polypyrrole
Maghemite/polydopamine 3 100 0.1 25 n.a. 97 90 [35]
core/shell nanoparticles
CS–CA nanoparticles 3 70 3 25 22.4 94.46 82.23 This study

n.d.: not determine, n.a.: not available.

were 84.5, 77, 51 and >90%, respectively, that they are in lower References
than CS–CA nanoparticles efficiency removal by 94.46%. Mean-
[1] M. Amini, H. Younesi, N. Bahramifar, Statistical modeling and optimization of
while, other adsorbents showed higher removal efficiency. For
the cadmium biosorption process in an aqueous solution using Aspergillus
example, Graphene oxide–␣CD-polypyrrole (GO–␣CD NC) [38], niger, Colloids Surf. Physicochem. Eng. Aspects 337 (2009) 67–73.
gelatin-impregnated-yeast [39] and PANI/zeolite nanocomposite [2] S. Sadeghi, E. Sheikhzadeh, Solid phase extraction using silica gel modified
[33] was used for the removal of Cr (VI) from aqueous solution, with murexide for preconcentration of uranium (VI) ions from water samples,
J. Hazard. Mater. 163 (2009) 861–868.
and the removal efficiency were found to be 100%. Based on data [3] T.A. Kurniawan, M.E.T. Sillanpää, M. Sillanpää, Nanoadsorbents for
presented in Table 4, CS–CA nanoparticles can remove Cr (VI) ions remediation of aquatic environment: local and practical solutions for global
from water also it is obvious that CS–CA nanoparticles showed water pollution problems, Crit. Rev. Environ. Sci. Technol. 42 (2011)
1233–1295.
a good affinity for binding Cr (VI) ions and desorption. This out- [4] F. Fu, Q. Wang, Removal of heavy metal ions from wastewaters: a review, J.
come showed that CS–CA nanoparticles will be competitively and Environ. Manag. 92 (2011) 407–418.
efficiently adsorbent for the removal of Cr (VI) ions from aqueous [5] D.H.K. Reddy, S.-M. Lee, Application of magnetic chitosan composites for the
removal of toxic metal and dyes from aqueous solutions, Adv. Colloid
solutions. Interface Sci. 201–202 (2013) 68–93.
[6] P. Chassary, T. Vincent, E. Guibal, Metal anion sorption on chitosan and
derivative materials: a strategy for polymer modification and optimum use,
4. Conclusions
React. Funct. Polym. 60 (2004) 137–149.
[7] A. Heidari, H. Younesi, Z. Mehraban, H. Heikkinen, Selective adsorption of
The adsorption potential of CS–CA nanoparticles to remove Cr Pb(II), Cd(II), and Ni(II) ions from aqueous solution using chitosan–MAA
(VI) from aqueous solution was investigated in a batch system. nanoparticles, Int. J. Biol. Macromol. 61 (2013) 251–263.
[8] S. Hajati, S. Tougaard, J. Walton, N. Fairley, Noise reduction procedures
CS–CA nanoparticles were prepared by polymerizing citric acid applied to XPS imaging of depth distribution of atoms on the nanoscale, Surf.
(CA) in chitosan (CS) solution. The effect of pH, adsorbent dose, Sci. 602 (2008) 3064–3070.
type of CS–CA nanparticles, Cr (VI) concentration, contact time and [9] A. Savitzky, M.J.E. Golay, Smoothing and differentiation of data by simplified
least squares procedures, Anal. Chem. 36 (1964) 1627–1639.
temperature was found to play a significant role in controlling the [10] A. Shahbazi, H. Younesi, A. Badiei, Functionalized nanostructured silica by
adsorption process. The results showed that the adsorption process tetradentate-amine chelating ligand as efficient heavy metals adsorbent:
follows the pseudo-second-order rate reaction. The equilibrium applications to industrial effluent treatment, Korean J. Chem. Eng. 31 (2014)
1598–1607.
data was modeled and estimated using three isotherm models, [11] O. Redlich, D.L. Peterson, A useful adsorption isotherm, J. Phys. Chem. 63
the Langmuir and Redlich–Peterson isotherm models yielding the (1959), 1024-1024.
best fit to the experimental data. Thermodynamic parameters [12] H.M.F. Freundlich, Über die Adsorption in Lösungen, Z. Phys. Chem. (Leipzig)
57 (1906) 385–470.
derived from the van’t Hoff and Arrhenius equations showed the
[13] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and
exothermic, spontaneous and chemical nature of Cr (VI) adsorption. platinum, J. Am. Chem. Soc. 40 (1918) 1361–1403.
Consequently, the potential is promising for using CS–CA nanopar- [14] H. Khakpour, H. Younesi, M. Mohammadhosseini, Two-stage biosorption of
selenium from aqueous solution using dried biomass of the baker’s yeast
ticles as clean-up material for Cr (VI) contaminated wastewaters.
Saccharomyces cerevisiae, J. Environ. Chem. Eng. 2 (2014) 532–542.
[15] A. Shahbazi, H. Younesi, A. Badiei, Functionalized SBA-15 mesoporous silica
Acknowledgements by melamine-based dendrimer amines for adsorptive characteristics of Pb(II),
Cu(II) and Cd(II) heavy metal ions in batch and fixed bed column, Chem. Eng.
J. 168 (2011) 505–518.
The authors wish to thank the Islamic Azad University, Science [16] S. Hussain, S. Gul, S. Khan, H.-u. Rehman, M. Ishaq, A. khan, F.A. Jan, Z.U. Din,
and Research Branch, for their financial support, which funding Removal of Cr (VI) from aqueous solution using brick kiln chimney waste as
adsorbent, Desalin. Water Treat. (2013) 1–9.
and a research grant made this study possible, and Ellen Vuosalo [17] D. Chauhan, M. Jaiswal, N. Sankararamakrishnan, Removal of cadmium and
Tavakoli (University of Mazandaran) for final editing of the English hexavalent chromium from electroplating waste water using thiocarbamoyl
text. chitosan, Carbohydr. Polym. 88 (2012) 670–675.
444 M. Bagheri et al. / International Journal of Biological Macromolecules 80 (2015) 431–444

[18] L. Panda, B. Das, D.S. Rao, B.K. Mishra, Application of dolochar in the removal [30] A.H. Qusti, Removal of chromium (VI) from aqueous solution using
of cadmium and hexavalent chromium ions from aqueous solutions, J. manganese oxide nanofibers, J. Ind. Eng. Chem. 20 (2014) 3394–3399.
Hazard. Mater. 192 (2011) 822–831. [31] T.A. Khan, M. Nazir, I. Ali, A. Kumar, Removal of chromium (VI) from aqueous
[19] K.G. Bhattacharyya, A. Sharma, Adsorption of Pb(II) from aqueous solution by solution using guar gum–nano zinc oxide biocomposite adsorbent, Arabian J.
Azadirachta indica (Neem) leaf powder, J. Hazard. Mater. 113 (2004) 97–109. Chem. (2013), http://dx.doi.org/10.1016/j.arabjc.2013.08.019, in press.
[20] M. Özacar, İ.A. Şengil, Adsorption of metal complex dyes from aqueous [32] U. Baig, R.A.K. Rao, A.A. Khan, M.M. Sanagi, M.A. Gondal, Removal of
solutions by pine sawdust, Bioresour. Technol. 96 (2005) 791–795. carcinogenic hexavalent chromium from aqueous solutions using newly
[21] A. Sarı, M. Tuzen, Ö.D. Uluözlü, M. Soylak, Biosorption of Pb(II) and Ni(II) from synthesized and characterized polypyrrole-titanium(IV)phosphate
aqueous solution by lichen (Cladonia furcata) biomass, Biochem. Eng. J. 37 nanocomposite, Chem. Eng. J. 280 (2015) 494–504.
(2007) 151–158. [33] A.A. Shyaa, O.A. Hasan, A.M. Abbas, Synthesis and characterization of
[22] J. Febrianto, A.N. Kosasih, J. Sunarso, Y.-H. Ju, N. Indraswati, S. Ismadji, polyaniline/zeolite nanocomposite for the removal of chromium (VI) from
Equilibrium and kinetic studies in adsorption of heavy metals using aqueous solution, J. Saudi Chem. Soc. 19 (2015) 101–107.
biosorbent: a summary of recent studies, J. Hazard. Mater. 162 (2009) [34] E. Pérez, L. Ayele, G. Getachew, G. Fetter, P. Bosch, A. Mayoral, I. Díaz, Removal
616–645. of chromium (VI) using nano-hydrotalcite/SiO2 composite, J. Environ. Chem.
[23] N.K. Amin, Removal of reactive dye from aqueous solutions by adsorption Eng. 3 (2015) 1555–1561.
onto activated carbons prepared from sugarcane bagasse pith, Desalination [35] A. Nematollahzadeh, S. Seraj, B. Mirzayi, Catecholamine coated maghemite
223 (2008) 152–161. nanoparticles for the environmental remediation: hexavalent chromium ions
[24] M. Arshadi, M.J. Amiri, S. Mousavi, Kinetic, equilibrium and thermodynamic removal, Chem. Eng. J. 277 (2015) 21–29.
investigations of Ni(II), Cd(II), Cu(II) and Co(II) adsorption on barley straw ash, [36] Y. Qu, X. Zhang, J. Xu, W. Zhang, Y. Guo, Removal of hexavalent chromium
Water Resour. Ind. 6 (2014) 1–17. from wastewater using magnetotactic bacteria, Sep. Purif. Technol. 136
[25] A.E. Ofomaja, Intraparticle diffusion process for lead(II) biosorption onto (2014) 10–17.
mansonia wood sawdust, Bioresour. Technol. 101 (2010) 5868–5876. [37] K. Kaya, E. Pehlivan, C. Schmidt, M. Bahadir, Use of modified wheat bran for
[26] I.N. Levine, Physical Chemistry, McGraw-Hill, Boston, MA, 2008. the removal of chromium (VI) from aqueous solutions, Food Chem. 158
[27] G. Akkaya, F. Güzel, Application of some domestic wastes as new low-cost (2014) 112–117.
biosorbents for removal of methylene blue: kinetic and equilibrium studies, [38] V.P. Chauke, A. Maity, A. Chetty, High-performance towards removal of toxic
Chem. Eng. Commun. 201 (2013) 557–578. hexavalent chromium from aqueous solution using graphene oxide-alpha
[28] H.Y. Aghdas Heidari, Z. Mehraban, H. Heikkinen, Selective adsorption of Pb(II), cyclodextrin-polypyrrole nanocomposites, J. Mol. Liq. 211 (2015) 71–77.
Cd(II), and Ni(II) ions from aqueous solution chitosan–MAA nanoparticles, Int. [39] M.E. Mahmoud, Water treatment of hexavalent chromium by
J. Biol. Macromol. (2013) 251–263. gelatin–impregnated-yeast (Gel–Yst) biosorbent, J. Environ. Manag. 147
[29] M. Avila, T. Burks, F. Akhtar, M. Göthelid, P.C. Lansåker, M.S. Toprak, M. (2015) 264–270.
Muhammed, A. Uheida, Surface functionalized nanofibers for the removal of [40] S. Rangabhashiyam, N. Selvaraju, Efficacy of unmodified and chemically
chromium (VI) from aqueous solutions, Chem. Eng. J. 245 (2014) modified Swietenia mahagoni shells for the removal of hexavalent chromium
201–209. from simulated wastewater, J. Mol. Liq. 209 (2015) 487–497.

You might also like