Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Colloid and Interface Science 398 (2013) 152–160

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Equilibrium and kinetic studies on acid dye Acid Red 88 adsorption by magnetic
ZnFe2O4 spinel ferrite nanoparticles
Wojciech Konicki a,⇑, Daniel Sibera b, Ewa Mijowska b, Zofia Lendzion-Bieluń b, Urszula Narkiewicz b
a
Department of Integrated Transport Technology and Environmental Protection, Maritime University of Szczecin, H. Poboz_ nego St. 11, 70-507 Szczecin, Poland
b
Institute of Chemical and Environment Engineering, West Pomeranian University of Technology, Pułaskiego St. 10, 70 322 Szczecin, Poland

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

Article history: A magnetic ZnFe2O4 (MNZnFe) was synthesized by microwave assisted hydrothermal method and was
Received 24 October 2012 used as an adsorbent for the removal of acid dye Acid Red 88 (AR88) from aqueous solution. The effects
Accepted 8 February 2013 of various parameters such as initial AR88 concentration (10–56 mg L1), pH solution (3.2–10.7), and
Available online 22 February 2013
temperature (20–60 °C) were investigated. Prepared magnetic ZnFe2O4 was characterized by XRD, SEM,
HRTEM, ICP-AES, BET, FTIR, and measurements of the magnetic susceptibility. The experimental data
Keywords: were analyzed by the Langmuir and Freundlich models of adsorption. Equilibrium data fitted well with
Adsorption
the Langmuir model. Pseudo-first-order and pseudo-second-order kinetic models and intraparticle diffu-
Acid dye
Magnetic nanoparticles
sion model were used to examine the adsorption kinetic data. The adsorption kinetics was found to fol-
Kinetics low the pseudo-second-order kinetic model. Thermodynamics parameters, DG°, DH° and DS°, indicate
Thermodynamics that the adsorption of AR88 onto MNZnFe was spontaneous and exothermic in nature.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction In recent years, adsorption process combined with magnetic


separation technology has been widely used for dyes removal from
Synthetic dyes are extensively used in various branches of the wastewaters. Nanosized magnetic materials can be used to pro-
textile industry, paper printing, color photography, pharmaceuti- duce magnetically-active carbon composites and as magnetic
cal, leather, cosmetics, plastic, and other industries. The discharges nano-adsorbents. Ai and Jiang have synthesized an activated car-
of industrial wastewater containing dyes cause serious environ- bon/ferrospinel composite as the magnetic adsorbent for fast re-
mental problems because of its high toxicity and possible accumu- moval of organic dyes from aqueous solutions [7]. Yang et al.
lation in the environment. Synthetic dyes are a group of most have prepared magnetic Fe3O4-activated carbon nanocomposite
dangerous pollutants in water. The presence of even very low con- particles for removal of methylene blue from aqueous solution
centrations of dyes in water reduces light penetration through the [8]. Ai et al. have synthesized an activated carbon/CoFe2O4 com-
water surface, precluding photosynthesis of the aqueous flora. posite as adsorbent for removal of malachite green dye from water
Many of these dyes are carcinogenic, mutagenic, and teratogenic [9]. Zhang et al. have developed magnetic CuFe2O4/activated car-
and also toxic to human beings, microorganisms, and fish species. bon composite for removal of acid orange II [10]. Several works
Hence, their removal from aquatic wastewater becomes environ- have been devoted to the adsorption of dyes onto magnetic-mod-
mentally important. ified MWCNTs and graphene [11–14]. Qadri et al. have studied a
Many treatment methods, such as membrane filtration, chemi- separation of a cationic dye acridine orange, by use of magnetic
cal oxidation, photocatalytic degradation and adsorption, have nanoparticles c-Fe2O3 [15]. Zargar et al. have studied the adsorp-
been examined to remove dyes from aqueous solutions. Of these tion and removal of amaranth from an aqueous solution by iron
methods, adsorption has been found to be an efficient and eco- oxide nanoparticles coated with cetyltrimethylammonium bro-
nomic process to remove dyes, pigments, and other colorants and mide [16]. Wang et al. have prepared different MFe2O4 (M = Mn,
also to control the bio-chemical oxygen demand [1]. Various adsor- Fe, Co, Ni) ferrite nanocrystals for the removal of Congo Red [17].
bents, such as chitosan [2], cotton [3], orange peel [4], palm fruit Hou et al. have synthesized by a simple sol–gel method porous spi-
bunch [5], and shale oil ash [6], have been investigated for removal nel ferrites Mn1xZnxFe2O4 (0 6 x 6 0.8) as adsorbent for the re-
of dyes from aqueous solution. moval of methylene blue [18]. Ai et al. have prepared
montmorillonite/CoFe2O4 composite for removal of methylene
blue [19] and prepared Ag–Fe3O4 composite as a magnetically
⇑ Corresponding author. Fax: +48 91 48 09 643. recyclable catalyst for reduction of Rhodamine B [20]. Iram et al.
E-mail address: w.konicki@am.szczecin.pl (W. Konicki). have prepared Fe3O4 hollow nanospheres for application as an

0021-9797/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2013.02.021
W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160 153

adsorbent for the removal of dye contaminants from water [21]. of Acid Red 88 is shown in Fig. 1. All solutions were prepared using
Mak and Chen have studied the adsorption of methylene blue from deionized water and reagent-grade chemicals.
an aqueous solution by polyacrylic acid-bound iron oxide magnetic
nanoparticles [22]. 2.2. Adsorption experiments
Acid Red 88 was selected as an adsorbate for the present study
because it is widely used in dyeing textile fabrics, silk, nylon, wool Adsorption experiments were carried out in Erlenmeyer flask
and leather, and its degradation byproducts have carcinogenic ef- (250 mL), where solution of AR88 (200 mL) with initial dye concen-
fects. Despite this, very few studies have been conducted on the tration of 10–56 mg L1 was placed. The flask with solution was
adsorption of AR88 from water [23–29]. Therefore, in this study, sealed and placed in a constant temperature bath and agitated
AR88 was selected as an adsorbate and used to evaluate the by magnetic stirrer. To observe the effect of temperature, the
adsorption efficiency of magnetic ZnFe2O4 nanoparticles experiments were carried out at five different temperatures, that
(MNZnFe). The effects of pH, temperature, and initial dye concen- is, 20, 30, 40, 50, and 60 °C. Before mixing with the adsorbent, var-
tration on dye adsorption by the MNZnFe were investigated. The ious pH levels (3.2–10.7) of the dye solution were adjusted by add-
experimental data were analyzed using the Lagergren pseudo- ing a few drops of diluted hydrochloric acid (0.1 N HCl) or sodium
first-order kinetic model, the pseudo-second-order model, and hydroxide (0.1 N NaOH). When the desired temperature was
the intraparticle diffusion model. Thermodynamics parameters, reached, about 0.03 g of MNZnFe was added into flask.
DG°, DH° and DS°, were also calculated. At the end of the equilibrium period, aqueous sample (5 mL)
was taken from the solution, the liquid was separated from the
adsorbent magnetically, and concentration of AR88 in solution
2. Experimental was determined spectrophotometrically at a wavelength of
505 nm. The amount of AR88 adsorbed at equilibrium qe (mg g1)
2.1. Materials and methods was calculated by following equation:

Zinc ferrite powder with compositions of MNZnFe was synthe- ðC 0  C e ÞV


qe ¼ ð1Þ
sized by microwave assisted hydrothermal method. The mixture of m
iron and zinc hydroxides was obtained by addition of 2 M solution where C0 (mg L1) is the initial dye concentration, Ce (mg L1) the
of KOH to 1 M solution of proper amount of zinc nitrate (V) hexa- dye concentration at equilibrium, V (L) the volume of the solution,
hydrate and iron (III) nitrate (V) tetrahydrate in water. The ob- and m (g) is the mass of the adsorbent.
tained hydroxides were put in the reactor with microwave The procedures of kinetic experiments were identical with
emission. The hydrothermal syntheses were carried out in a proto- those of equilibrium tests. At predetermined moments, aqueous
type reactor constructed in the Institute of High Pressure Physics samples (5 mL) were taken from the solution, the liquid was sepa-
PAS in Warsaw, in cooperation with the company ERTEC, Poland, rated from the adsorbent magnetically and concentration of AR88
and the Institute for Sustainable Technologies – National Research in solution was determined spectrophotometrically at a wave-
Institute in Radom. This is the reactor with an emission of micro- length of 505 nm. The amount of AR88 adsorbed at time t qt
waves enabling to conduct inorganic and organic processes under (mg g1) was calculated by following equation:
the pressure up to 3.9 MPa and temperatures up to 350 °C. The
flow of the reaction mixture is 2 l/h. In the present work, the ðC 0  C t ÞV
qt ð2Þ
microwave assisted synthesis was conducted under a pressure of m
3.9 MPa for 15 min in air. The obtained product was filtered and where Ct (mg L1) the dye concentration at any time t.
dried. The phase composition of the samples was determined using
XRD (Co Ka radiation, X’Pert Philips). The morphology of samples
3. Results and discussion
was investigated using scanning electron microscopy (LEO 1530)
and a high-resolution transmission electron microscope (HRTEM)
3.1. Characterization of the adsorbent
FEI Tecnai F20 operating at 200 kV. The real chemical composition
of samples was determined using (ICP-AES) inductively coupled
XRD analysis has shown the presence of a spinel phase arising
plasma atomic emission spectroscopy (Yvon-Jobin, France). The
from zinc ferrite. All peaks visible in the XRD pattern correspond
specific surface area of the nanopowders was determined by the
Brunauer–Emmett–Teller (BET) method using the equipment
Gemini 2360 of Micromeritics. The helium pycnometer AccuPyc
1330 of Micromeritics was applied to determine the density of
powders. The functional groups on the MNZnFe surface were
determined using Fourier transform infrared FTIR method (Perkin
Elmer Spectrum One FT-IR spectrometer). FTIR spectra of MNZnFe
were recorded in the range of 4000–350 cm1 using KBr pellets
method. The point of zero charge (pHPZC) of MNZnFe was deter-
mined by the pH drift method [30]. The pH of a solution of
0.01 M NaCl was adjusted between pH 3.0 and pH 12.0 by adding
either HCl or NaOH. Nitrogen was bubbled through the solution
thermostatted at 25 °C to remove dissolved CO2 until the initial
pH (pHi) value of the solution stabilized. Then, 0.15 g of MNZnFe
was added to 50 mL of the solution and the final pH (pHf) mea-
sured after 48 h. The pHPZC of MMWCNTs was obtained from the
plot of DpH = pHi–pHf versus initial pHi. The point of intersection
of the resulting curve at which DpH = 0 gave the pHPZC. Azo dye
Acid Red 88 (C20H13N2NaO4S, molar mass 400.38) supplied by Za-
chem Barwniki was used as an adsorbate. The chemical structure Fig. 1. Chemical structure of Acid Red 88.
154 W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160

to this phase. The XRD results are described in details elsewhere


[31]. The mean size of the MNZnFe crystallites, calculated using
the Scherrer’s formula, was equal to 11 nm. Table 1 summarizes
the results of the density and specific surface area measurements
to gather with sample chemical composition determined by ICP-
AES measurements, in perfect agreement with the nominal one
calculated from the initial concentration of the precursor salts.
SEM images reveal the agglomerated structure of the synthe-
sized material (Fig. 2a). We assume that the small spheroidal crys-
tals correspond to the magnetic MNZnFe spinel. Agglomerates
have size of about 60 nm and are bound to each other creating
large structures. The powder exhibits homogenous structure, with
a narrow distribution of agglomerates sizes. HRTEM images show a
fine crystalline structure with mean crystallite size of about 10 nm,
which is in accordance with XRD measurements. The agglomerates
seen in Fig. 2a are composed of the crystallites visible in Fig. 2b.
Fig. 3. Photograph of magnetic separation of MNZnFe from the aqueous solution
with a permanent magnet.
Table 1
Chemical composition and physical properties of MNZnFe.

Nominal Density (g cm3) Surface area (m2 g1) Measured The magnetic properties of the material are described in our re-
concentration concentration cent paper [32], where the measurements of the samples ZnO–
(wt.%) (wt.%)
Fe2O3 containing 5–70 wt.% Fe2O3 are presented. The dynamic
Fe2O3 ZnO Fe2O3 ZnO magnetic properties were studied by means of AC susceptibility.
70 30 4.6 139 70.1 26.2 The real and imaginary parts of magnetic susceptibility were mea-
sured using a mutual inductance method in an AC magnetic field of
frequency range 7–10.000 Hz and amplitude (HAC) not exceeding
5 Oe. The measurements were performed in the temperature from
4.5 to 160 K. The most pronounced maximum of the intensity of
magnetic susceptibility was observed for the spinel sample, con-
taining 70 wt.% of Fe2O3, taken for the experiments in the recent
studies. For sample synthesized by hydrothermal process, MNZnFe
nanocrystals showed superparamagnetic behavior. Fig. 3 shows
the separation of the MNZnFe from the aqueous solution with a
permanent magnet.
Fig. 4 shows the FTIR spectrum of MNZnFe. Two characteristic
peaks observed at around 410 and 565 cm1 corresponds to octa-
hedral-metal stretching, (Fe3+–O2), and to intrinsic stretching
vibrations of the metal at the tetrahedral site, (Zn2+–O2), respec-
tively [33,34]. The peaks at 1387, 1060, and 845 cm1 present in
the IR spectrum of MNZnFe can be assigned to the vibration of
NO3 group trapped in the prepared sample [35–37]. The peaks ob-
served at around 3415 and 1544 cm1 are ascribed due to the
stretching modes and H–O–H bending vibration of the free or ab-
sorbed water molecules on the surface of the MNZnFe [38]. In
Fig. 5, the plot of DpH versus pHi is showed. The pH at which
DpH becomes 0 is called pHPZC. The zero point charge of MNZnFe
was found to be 8.5.

Fig. 2. (a) SEM and (b) HRTEM images of MNZnFe. Fig. 4. FTIR spectra of MNZnFe.
W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160 155

Fig. 5. The pHPZC of MNZnFe. Fig. 7. The effect of initial pH of dye solution on adsorption of the Acid Red 88 onto
MNZnFe. Experimental conditions: C0,AR88 = 20 mg L1, T = 30 °C.

3.2. Effect of initial dye concentration


The adsorption capacity decreased when the pH of the dye solu-
The effect of initial dye concentration on adsorption of AR88 tion was increased over the whole studied range. The maximum
onto MNZnFe was investigated in the concentration range 10– adsorption capacity of AR88 was 130.6 mg g1, observed at pH
56 mg L1 at pH 7.0 and 30 °C. The results are shown in Fig. 6. 3.2. A large decrease in adsorption capacity from 130.6 to
The adsorption of the AR88 dye onto MNZnFe increased with 21.2 mg g1 was observed when the pH of the dye solutions was
time and then attained equilibrium after 340 min. The adsorption increased from 3.2 to 9.2. When the pH was increased from 9.2
is initially rapid, and then slow, perhaps because a large number to 10.7, the adsorption capacity of the AR88 decreased from 21.2
of vacant surface sites were available for adsorption during the ini- to 17.5 mg g1. Acid Red 88 is an anionic mono azo dye that have
tial stage, and then, the remaining vacant surface sites were diffi- sulfonic acid group (R–SO3Na). In aqueous solution, Acid Red 88
cult to occupy because of the repulsive forces between the dye dissociates to the sodium ions Na+ and the sulfonate anions
molecules on the MNZnFe and the bulk phase [39]. The adsorption R—SO 3 . It should be highlighted that sulfonic groups present neg-

capacity at equilibrium increases evidently from 37.8 to ative charge even at higher acidic solutions, because their pKa val-
96.5 mg g1, with an increase in the initial dye concentration from ues are lower than zero [41]. The pHPZC value for MNZnFe was 8.5
10 to 56 mg L1. The result that adsorption capacity of AR88 in- and it is the pH at which the net surface charge on adsorbent is
creased with the increase in initial AR88 concentration might be zero. The adsorbent surface has a net positive charge at pH < pHPZC,
attributed to an increase in the driving force of concentration gra- while has net negative charge at pH > pHPZC. This means that at
dient with the increase in the initial concentration [40]. pH < 8.5, MNZnFe surface becomes positively charged, favoring
the adsorption of AR88 in anionic form. Hence, when pH is de-
creased, the adsorbent surface is more positively charged resulting
3.3. Effect of pH in greater attraction between the MNZnFe and AR88. On the other
hand, as the solution pH increases, the number of sites positively
The influence of pH on adsorption capacity of AR88 onto charged decreases and thereby the uptake of AR88 decreases. In
MNZnFe was investigated in the range of pH values from 3.2 to addition, when solution pH is above the pHPZC, the MNZnFe surface
10.7 at a fixed dye concentration of 20 mg L1 and 30 °C. The effect is more negatively charged, which does not favor the adsorption of
of the initial pH on the adsorption capacity of AR88 is shown in dye anions due to the electrostatic repulsion.
Fig. 7.

3.4. Effect of ionic strength

The effect of ionic strength on adsorption of AR88 onto MNZnFe


was investigated in NaCl solutions with the concentration range
from 0 to 0.5 mol/L at a fixed dye concentration of 20 mg L1 and
30 °C. The results are shown in Fig. 8.
It was observed that the adsorption equilibrium of AR88 in-
creased from 61.8 to 94.8 mg g1 with an increase in the NaCl con-
centration from 0 to 0.5 mol/L. Theoretically, when the
electrostatic forces between the adsorbent surface and adsorbate
ions are attractive, an increase in ionic strength will decrease the
adsorption capacity. Despite this, our study did not confirm this
assumption. The presence of NaCl in the solution may have two
opposite effects. On the one hand, the salt can neutralize the neg-
atively charged surface of adsorbent, enabling them to adsorb
more dye molecules. Thus, under high-level NaCl concentration
positively charged sodium ions Na+ can react with the negatively
Fig. 6. The effect of initial dye concentration on adsorption of the acid dye Acid Red charged surface of MNZnFe, which results in the reduction of elec-
88 onto MNZnFe. Experimental conditions: T = 30 °C, pH = 7.0. trostatic repulsion between the sulfonate anions R—SO 3 and sur-
156 W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160

Fig. 8. The effect of ionic strength on adsorption of AR88 onto MNZnFe. Experi-
mental conditions: C0,AR88 = 20 mg L1, T = 30 °C.

face of adsorbent. On the other hand, the salt can neutralize the
negative charged dye molecules and the positively charged surface
of adsorbent, which weakens the attractive force between AR88
and MNZnFe. In this study, adsorption capacity of AR88 dye in-
creased significantly onto MNZnFe due to the presence of NaCl. It
means that other interactions besides the interaction between
the dye molecules and the adsorbent surface might control the
adsorption process. Similar phenomena have been observed for
Reactive Black 5 dye adsorption onto two bamboo based active car-
bons and three conventional adsorbents, carbon F400, bone char,
and peat [42], for Reactive Blue 2, Reactive Red 4, and Reactive Yel-
low 2 adsorption onto activated carbon [43], for methylene blue
Fig. 9. (a) Effect of temperature on adsorption of the acid dye Acid Red 88 onto
adsorption onto rice husk [44], for Acid Red 14 adsorption onto
MNZnFe. Experimental conditions: C0,AR88 = 20 mg L1, pH = 7.0. (b) Van’t Hoff plot
surface soils [45], and for Acid Yellow 49 adsorption onto sepiolite for the adsorption of the acid dye Acid Red 88 onto MNZnFe.
[46].

3.5. Effect of temperature where T is the solution temperature (K), Ka is the adsorption equi-
librium constant, and R is the gas constant. Enthalpy (DH°) and en-
In order to observe the effect of temperature, the adsorption tropy (DS°) were calculated from the slope and intercept from the
studies were carried out for the initial dye concentration of plot of lnqe/Ce versus 1/T (Fig. 9b).
20 mg L1 at five different temperatures, that is, 20, 30, 40, 50 The value of Gibbs free energy (DG°) was calculated using Eq.
and 60 °C, at pH 7.0. Fig. 9a shows the adsorption capacity versus (5). The results are shown in Table 2. The negative value of DH°
the adsorption time at different temperatures. It was observed that (30.3 kJ mol1) indicates that adsorption of AR88 onto MNZnFe
the adsorption equilibrium of AR88 decreased from 69.1 to is an exothermic process. The values of Gibbs free energy (DG°)
22.5 mg g1 with the increase in temperature from 20 to 60 °C, were negative in the temperature range of 20–60 °C, confirming
indicating the exothermic nature of the adsorption reaction. In or- that the adsorption of AR88 onto MNZnFe was spontaneous and
der to study the thermodynamics of adsorption of AR88 dye on thermodynamically favorable. When the temperature increased
MNZnFe, three basic thermodynamic parameters, enthalpy (DH°), from 20 to 60 °C, DG° increased from 4.37 to 0.92 kJ mol1, sug-
entropy (DS°) and Gibbs free energy (DG°), were calculated using gesting that adsorption was spontaneous at low temperatures. The
following equations [47]: negative value DS° (86.9 J mol1 K1) indicates a decrease in the
DS DH  degree of freedom of adsorption AR88 onto MNZnFe. The behavior
ln K a ¼  ð3Þ similar to this was also observed by Purkait et al. [48], Iqbal and
R RT
Ashiq [49] and Karaoglu et al. [50]. Molecules before adsorption
qe can move in three dimensions but as they get adsorbed on the sur-
Ka ¼ ð4Þ
Ce face, the motion of molecules is restricted toward the surface and
their disorderness decreased resulting in the decrease in entropy
DG ¼ RT ln K a ð5Þ [49,51].

Table 2
Thermodynamic parameters for the adsorption of AR88 on NZnFe.

Dye concentration (mg L1) DH° (kJ mol1) DS° (J mol1 K1) DG° at temperature (°C) (kJ mol1)
20 30 40 50 60
20 30.3 86.9 4.37 4.34 3.42 2.38 0.92
W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160 157

3.6. Adsorption isotherms Table 3


Langmuir and Freundlich parameters for the adsorption of AR88 on MNZnFe at 30 °C.

In this study, the Langmuir and Freundlich isotherms were used Langmuir isotherm
to describe the equilibrium adsorption. Langmuir isotherm theory Q0 (mg g1) 111.1
is based on the assumption of adsorption on a homogeneous sur- b (L mg1) 0.120
RL 0.130
face. The linear form of Langmuir’s isotherm model is represented R2 0.997
by the following equation:
Freundlich isotherm
Ce 1 Ce KF [(mg g1)(L mg1)1/n] 2.6
¼ þ ð6Þ n 2.42
qe Q 0 b Q 0 R2 0.968

where Q0 (mg g1) is the maximum monolayer adsorption capacity


and b (L/mg) is a constant related to energy of adsorption. The
values of Q0 and b were calculated from the slope and intercept of
Table 4
the linear plot Ce/qe versus Ce (Fig. 10a). The essential characteristics
Comparison of the maximum monolayer adsorption of AR88 onto various adsorbents.
of the Langmuir isotherm can be expressed in terms of a dimension-
less equilibrium parameter (RL), which is defined by the following Adsorbent Q0 (mg g1) Refs.
equation [52]: Charfines 33.3 [23]
Lignite coal 30.09 [23]
1 Bituminous coal 26.1 [23]
RL ¼ ð7Þ Activated carbon 109.0 [23]
1 þ bC 0
Azolla filiculoides 123.5 [24]
where C0 (mg L1) is the highest initial concentration of the adsor- Azolla rongpong 78.74 [25]
bate. The value of RL indicates the type of the isotherm to be either Azolla microphylla 54.89 [26]
Anion exchange membrane 42.01 [27]
unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL < 1), or irre- Natural clay 1133.1 [28]
versible (RL = 0). The values of RL were found to be 0.130 and con- Raw alunite 20.98 [29]
firmed that the MNZnFe is favorable for adsorption of AR88 under Calcined alunite 832.81 [29]
the conditions used in this study. The maximum monolayer adsorp- MNZnFe 111.1 This study
tion capacity Q0 was 111.1 mg AR88 per gram of the MNZnFe.
The Freundlich isotherm describes equilibrium on heteroge-
 
neous surfaces and hence does not assume monolayer capacity 1
ln qe ¼ ln K F þ ln C e ð8Þ
[53]. A linear form of the Freundlich model is represented by the n
following equation:
where KF (mg g1(L/mg)1/n) and n are Freundlich constants, which
represent adsorption capacity and adsorption strength, respec-
tively. The values of KF and n were calculated from the slope and
intercept of the linear plot lnqe versus lnCe (Fig. 10b). The values
of n ranging from 1.0 to 10.0 indicated a good adsorption process
[54].
Table 3 showed the values of Langmuir and Freundlich con-
stants, and the correlation coefficients R2 obtained from the linear
regression. The fit of the experimental data for the adsorption of
AR88 onto MNZnFe suggests that Langmuir model (R2 = 0.997)
gave closer fittings than the Freundlich model (R2 = 0.968).
Table 4 compares the adsorption capacities of the MNZnFe with
different types of adsorbents previously used for removal of AR88.
Adsorption capacity of MNZnFe equal to 111.1 mg g1 was higher
than that of other adsorbents, except for Azolla filiculoides, natural
clay and calcined alunite. This data suggests that the MNZnFe as
adsorbent was suitable for the removal of AR88 from aqueous
solution.

3.7. Adsorption kinetics

Kinetic adsorption data were analyzed using the Lagergren


pseudo-first-order kinetic model, the pseudo-second-order model,
and the intraparticle diffusion model. The Lagergren pseudo-first-
order model is represented by the following equation [55]:
lnðqe  qt Þ ¼ ln qe  k1 t ð9Þ
1
where time t (min) and k1 (min ) is the first-order rate constant
adsorption. Values of k1 and equilibrium adsorption density qe at
30 °C were calculated from the plots of ln(qe–qt) versus t for
different initial concentrations of AR88 (Fig. 11a). The pseudo-sec-
ond-order kinetic model can be expressed as follows:
t 1 1
Fig. 10. (a) Langmuir and (b) Freundlich adsorption isotherm of the acid dye Acid ¼ þ t ð10Þ
Red 88 onto MNZnFe at 30 °C.
qt k2 q2e qe
158 W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160

intercept of the linear plot of t/qt versus t. The plot of t/qt versus t
at 30 °C is shown in Fig. 11b. Intraparticle diffusion model based
on the theory proposed by Weber and Morris is represented by
the following equation [56]:

qt ¼ kp t 0:5 þ C ð11Þ
1 1 0.5
where C (mg g ) is the intercept and kp (mg g min ) is the
intraparticle diffusion rate constant. Value of kp was obtained from
the slope of the linear plot of qt versus t0.5 (Fig. 11c).
The kinetic data from pseudo-first and pseudo-second-order
adsorption kinetic models and the intraparticle diffusion model
are given in Table 5. The linear plots of t/qt versus t (pseudo-sec-
ond-order kinetic model) indicated a good agreement between
the experimental and calculated qe values for different initial dye
concentrations. Furthermore, the correlation coefficients of the
pseudo-second-order kinetic model (R2 > 0.99) were greater than
that of the pseudo-first-order model. As a result, it can be said that
the adsorption fits to the pseudo-second-order better than the
pseudo-first-order kinetic model. Adsorption process of AR88 on
different adsorbents such as A. filiculoides [24], Azolla rongpong
[25], natural clay [28], anion exchange membrane [27], calcined
alunite [29], and sawdust [57] was also described by the pseudo-
first-order and pseudo-second-order kinetic models by other
authors. In any case, they found that experimental data better fit
to pseudo-second-order kinetic model.
The observed data were also analyzed by the intraparticle diffu-
sion model. According to the model based on the theory proposed
by Weber and Morris, if the regression of qt versus t0.5 is linear and
passes through the origin, then intraparticle diffusion is the sole
rate-limiting step. The regression was linear, but the plot did not
pass through the origin, suggesting that adsorption involved
intraparticle diffusion, but that was not the only rate-controlling
step. Values of intercept C give an idea about the thickness of the
boundary layer, the larger intercept the greater is the boundary
layer effect [56]. The C values increased from 25.7 to 57.1 mg g1
with increasing the initial dye concentrations form 10 to
56 mg L1. The results of this study demonstrated that increasing
the initial dye concentrations promoted the boundary layer diffu-
sion effect.
Adsorption kinetics of AR88 onto MNZnFe was best described
by the pseudo-second-order model. Accordingly, the rate constants
k2 of the pseudo-second-order model were adopted to calculate the
activation energy of the adsorption process using the Arrhenius
equation [58]:
Ea
ln k2 ¼ ln A  ð12Þ
RT
Fig. 11. (a) Pseudo-first-order kinetics, (b) pseudo-second-order kinetics, and (c)
intraparticle diffusion model for adsorption of the acid dye Acid Red 88 onto where k2 (g mg1 min1) is the rate constant for the pseudo-sec-
MNZnFe at 30 °C. ond-order adsorption kinetics, A (g mg1 min1) is the tempera-
ture-independent Arrhenius factor, Ea (J mol1) the activation
energy, R (8.314 J mol1 K1) is the gas constant, and T (K) is the
where k2 (g mg1 min1) is the rate constant for the pseudo-sec- temperature. Value of Ea was obtained from the slope of the linear
ond-order adsorption kinetics. Values of k2 and qe for different ini- plot of lnk2 versus 1/T (Fig. 12). The activation energy was
tial concentrations of AR88 were calculated from the slope and 37.5 kJ mol1 (R2 = 0.932). The magnitude of the activation energy

Table 5
Comparison of the pseudo-first-order, pseudo-second-order, and the intraparticle diffusion models for different initial concentrations of AR88 at 30 °C.

Initial concentration qe,exp Pseudo-first-order kinetic model Pseudo-second-order kinetic model Intraparticle diffusion model
(mg L1) (mg g1)
k1 (min1) qe,cal (mg g1) R2 k2 (min1) qe,cal (mg g1) R2 kp (mg g1 min1/2) C (mg g1) R2
10 37,8 0.0137 23.1 0.976 0.0018 38.9 0.999 0.6600 25.7 0.899
20 61,8 0.0142 43.4 0.969 0.0009 64.1 0.998 1.3023 37.7 0.951
30 81,5 0.0111 51.0 0.978 0.0007 84.0 0.998 1.7388 48.9 0.948
44 91,5 0.0090 51.3 0.939 0.0007 92.6 0.998 1.7733 57.0 0.952
56 96,5 0.0109 61.1 0.962 0.0006 99.0 0.997 2.0758 57.1 0.977
W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160 159

4. Conclusion

A magnetic ZnFe2O4 adsorbent for effective acid dye removal


has been prepared by microwave assisted hydrothermal method.
The prepared adsorbent can be easily separated from the aqueous
solution by an external magnetic field. The effects of various oper-
ating conditions, like initial dye concentration, pH and tempera-
ture, were investigated. It was observed that the adsorption of
AR88 is favored at low pH (130.6 mg g1 at pH = 3.2). The maxi-
mum monolayer adsorption capacity of magnetic ZnFe2O4 was
found to be 111.1 mg g1. This value was higher than those re-
ported by other researchers with the exception of A. filiculoides,
natural clay, and calcined alunite. Adsorption data were described
using the pseudo-first-order kinetic model, the pseudo-second-or-
der model, and the intraparticle diffusion model. The adsorption
kinetics was well described by pseudo-second-order model. The
Fig. 12. Arrhenius plot for adsorption of the acid dye Acid Red 88 onto MNZnFe.
data were analyzed by the Langmuir and Freundlich models of
Experimental conditions: C0,AR88 = 20 mg L1, pH = 7.0.
adsorption. The equilibrium data fitted well with the Langmuir
model. The thermodynamic study indicated that the adsorption
of AR88 onto ZnFe2O4 was spontaneous and an exothermic process.
may give an idea about the type of sorption. Two main types of
adsorptions may occur: physical and chemical. The physisorption
process normally has an activation energy of 5–40 kJ mol1, while References
chemisorption has a higher activation energy of 40–800 kJ mol1 [1] P. Cooper, Colour in Dye House Effluent, in: Society of Dyers and Colorists,
[59]. The value of Ea for the adsorption of AR88 onto MNZnFe is Alden Press, Oxford, 1995.
within the range of physisorption. Similar results were observed [2] S. Chatterjee, D.S. Lee, M.W. Lee, S.H. Woo, Bioresour. Technol. 100 (2009)
3862.
for the adsorption of Procion Red MX-5B onto CNTs (33.35 kJ mol1)
[3] M. Ertas, B. Acemioglub, M. Hakki Alma, M. Usta, J. Hazard. Mater. 183 (2010)
[58] and for the adsorption of maxilon blue GRL on sepiolite 421.
(33.96 kJ mol1) [60]. [4] C. Namasivayam, N. Muniasamy, K. Gayatri, M. Rani, K. Ranganathan,
Bioresour. Technol. 57 (1996) 37.
The Gibbs free energy change, (DG°), is the fundamental crite-
[5] M.M. Nassar, Y.H. Magdy, Chem. Eng. J. 66 (1997) 223.
rion of spontaneity of a chemical reaction. Generally, DG° for [6] Z. Al-Qodah, Wat. Res. 34 (2000) 4295.
physisorption is less than that for chemisorption. The change in [7] L. Ai, J. Jiang, Desalination 262 (2010) 134.
free energy for physisorption is between 20 and 0 kJ mol1, [8] N. Yang, S. Zhu, D. Zhang, S. Xu, Mater. Lett. 62 (2008) 645.
[9] L. Ai, H. Huang, Z. Chen, X. Wei, J. Jiang, Chem. Eng. J. 156 (2010) 243.
the physisorption together with chemisorptions is at the range [10] G. Zhang, J. Qu, H. Liu, A.T. Cooper, R. Wu, Chemosphere 68 (2007) 1058.
of 20 to 80 kJ mol1 and chemisorption is at the range of [11] S. Qu, F. Huang, S. Yu, G. Chen, J. Kong, J. Hazard. Mater. 160 (2008) 643.
80 to 400 kJ mol1 [61]. The values of DG° (4.37 to [12] J.-L. Gong, B. Wang, G.-M. Zeng, C.-P. Yang, C.-G. Niu, Q.-Y. Niu, W.-J. Zhou, Y.
Liang, J. Hazard. Mater. 164 (2009) 1517.
0.92 kJ mol1) for the adsorption of AR88 onto MNZnFe were [13] T. Madrakian, A. Afkhami, M. Ahmadin, H. Bagheri, J. Hazard. Mater. 196 (2011)
in the range of physisorption. The value of enthalpy change 109.
(DH°) shows the exothermic or endothermic nature of the [14] Y. Yao, S. Miao, S. Liu, L.P. Ma, H. Sun, S. Wang, Chem. Eng. J. 184 (2012) 326.
[15] S. Qadri, A. Ganoe, Y. Haik, J. Hazard. Mater. 169 (2009) 318.
adsorption process which may signify either physisorption or [16] B. Zargar, H. Parham, A. Hatamie, Chemosphere 76 (2009) 554.
chemisorptions. Kara et al. suggested that the DH° of physisorp- [17] L. Wang, J. Li, Y. Wang, L. Zhao, Q. Jiang, Chem. Eng. J. 72 (2012) 181.
tion is smaller than 40 kJ mol1 [62]. In this study, DH° [18] X. Hou, J. Feng, X. Liu, Y. Ren, Z. Fan, M. Zhang, J. Colloid Interf. Sci. 353 (2011)
524.
(30.3 kJ mol1) for the adsorption of AR88 onto MNZnFe was
[19] L. Ai, Y. Zhou, J. Jiang, Desalination 266 (2011) 72.
less than 40 kJ mol1. Therefore, DH°, DG°, and Ea all suggested [20] L. Ai, C. Zeng, Q. Wang, Catal. Commun. 14 (2011) 68.
the same fact: the adsorption of AR88 onto MNZnFe was a phys- [21] M. Iram, C. Guo, Y. Guan, A. Ishfaq, H. Liu, J. Hazard. Mater. 181 (2010) 1039.
[22] S.-Y. Mak, D.-H. Chen, Dyes Pigments 61 (2004) 93.
isorption process.
[23] S.V. Mohan, P. Sailaja, M. Srimurali, J. Karthikeyan, Environ. Eng. Policy 1
(1999) 149.
[24] T.V.N. Padmesh, K. Vijayaraghavan, G. Sekaran, M. Velan, J. Hazard. Mater. 125
(2005) 121.
3.8. Desorption study [25] T.V.N. Padmesh, K. Vijayaraghavan, G. Sekaran, M. Velan, Chem. Eng. J. 122
(2006) 55.
In order to evaluate the possibility of regeneration of the [26] T.V.N. Padmesh, K. Vijayaraghavan, G. Sekaran, M. Velan, Biorem. J. 10 (2006)
37.
MNZnFe adsorbent, desorption experiments have been performed. [27] T. Xing, H. Kai, G. Chen, Color. Technol. 128 (2012) 295.
Two solvents including acetone and methanol were used for the [28] S.T. Akar, R. Uysal, Chem. Eng. J. 162 (2010) 591.
regeneration of the adsorbent after AR88 adsorption [16,17]. For [29] S.T. Akar, T. Alp, D. Yilmazer, J. Chem. Technol. Biot. (2012).
[30] M.V. Lopez-Ramon, F. Stoeckli, C. Moreno-Castilla, F. Carrasco-Marin, Carbon
this purpose, 0.02 g of dye loaded MNZnFe was added to 20 mL 37 (1999) 1215.
of acetone solutions or methanol and shaken vigorously for [31] D. Sibera, R. Jedrzejewski, J. Mizeracki, A. Presz, U. Narkiewicz, W. Łojkowski,
15 min. The nanoparticles of the adsorbent were separated from Acta Phys. Pol. A 116 (2009) 133.
[32] I. Kuryliszyn-Kudelska, B. Hadzić, D. Sibera, M. Romcević, N. Romcević, U.
the liquid magnetically and concentration of AR88 in the desorbed Narkiewicz, W. Dobrowolski, J. Alloys Compd. 509 (2011) 3756.
solution was measured spectrophotometrically. Desorption effi- [33] Y. Koseoglu, A. Baykal, M.S. Toprak, F. Gozuak, A.C. Basaran, B. Aktas, J. Alloys
ciency of dye from the adsorbent was calculated as the ratio of Compd. 462 (2008) 209.
[34] P. Pascuta, G. Borodi, M. Bosca, L. Pop, S. Rada, E. Culea, J. Phys. Conf. Ser. 182
the amount of the dye desorbed to amount of the dye adsorbed.
(2009) 1.
It was found that the desorption efficiency of AR88 from MNZnFe [35] D. Weinzierl, D. Touraud, A. Lecker, A. Pfitzner, W. Kunz, Mater. Res. Bull. 43
in acetone and methanol was 63% and 81%, respectively. These re- (2008) 62.
sults indicate that the MNZnFe can be potentially used as a [36] D. Varshney, K. Verma, A. Kumar, Mater. Chem. Phys. 131 (2011) 413.
[37] M. Mouallem-Bahout, S. Bertrand, O. Pena, J. Solid State Chem. 178 (2005)
reusable magnetic absorbent for the removal of AR88 from aque- 1080.
ous solution. [38] I. Sharifi, H. Shokrollahi, J. Magn. Magn. Mater. 324 (2012) 2397.
160 W. Konicki et al. / Journal of Colloid and Interface Science 398 (2013) 152–160

[39] A.H. Norzilah, A. Fakhru’l-Razi, T.S.Y. Choong, A. Luqman Chuah, J. Nanomater [51] L. Vlaev, S. Turmanova, A. Dimitrova, J. Polym. Res. 16 (2009) 151.
(2011) 1. 2011 Article ID 495676. [52] B.H. Hameed, A.T.M. Din, A.L. Ahmad, J. Hazard. Mater. 141 (2007) 819.
[40] Y. Yao, F. Xu, M. Chen, Z. Xu, Z. Zhu, Bioresour. Technol. 101 (2010) 3040. [53] E. Demirbas, M.Z. Nas, Desalination 243 (2009) 8.
[41] E.C. Lima, B. Royer, J.C.P. Vaghetti, N.M. Simon, B.M. da Cunha, F.A. Pavan, E.V. [54] V.K. Gupta, I. Ali, V.K. Saini, J. Colloid Interface Sci. 315 (2007) 87.
Benvenutti, R.C. Veses, C. Airoldi, J. Hazard. Mater. 155 (2008) 536. [55] M. Chairat, S. Rattanaphani, J.B. Bremner, V. Rattanaphani, Dyes Pigments 76
[42] A.W.M. Ip, J.P. Barford, G. McKay, J. Colloid Interf. Sci. 337 (2009) 32. (2008) 435.
[43] Y.S. Al-Degs, M.I. El-Barghouthi, A.H. El-Sheikh, G.M. Walker, Dyes Pigments 77 [56] B.H. Hameed, F.B.M. Daud, Chem. Eng. J. 139 (2008) 48.
(2008) 16. [57] N. Modirshahla, M.A. Behnajady, A. Shamel, H. Eskandari, Iran 7 (2010) 77.
[44] R. Han, Y. Wang, W. Yu, W. Zou, J. Shi, J. Hazard. Mater. 141 (2007) 713. [58] C.H. Wu, J. Hazard. Mater. 144 (2007) 93.
[45] B. Qu, J. Zhou, X. Xiang, C. Zheng, H. Zhao, X. Zhou, J. Environ. Sci. 20 (2008) [59] H. Nollet, M. Roels, P. Lutgen, P. Meeren, W. Verstraete, Chemosphere 53
704. (2003) 655.
[46] M. Alkan, Ö. Demirbasß, M. Doğan, Fresen. Environ. Bull. 13 (2004) 1112. [60] M. Dogan, M. Alkan, O. Demirbas, Y. Ozdemir, C. Ozmetin, Chem. Eng. J. 124
[47] S. Karagoz, T. Tay, S. Ucar, M. Erdem, Bioresource Technol. 99 (2008) 6214. (2006) 89.
[48] M.K. Purkait, S. DasGupta, S. De, J. Environ. Manage. 76 (2005) 135. [61] M.J. Jaycock, G.D. Parfitt, Chemistry of Interfaces, Ellis Horwood Ltd.,
[49] M.J. Iqbal, M.N. Ashiq, J. Hazard. Mater. B139 (2007) 57. Onichester, 1981.
[50] M.H. Karaoglu, M. Dogan, M. Alkan, Desalination 256 (2010) 154. [62] M. Kara, H. Yuzer, E. Sabah, M.S. Celik, Water Res. 37 (2003) 224.

You might also like