You are on page 1of 11

Applied Clay Science 109–110 (2015) 22–32

Contents lists available at ScienceDirect

Applied Clay Science

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

Research paper

Determination of adsorptive properties of expanded vermiculite for the


removal of C. I. Basic Red 9 from aqueous solution: Kinetic, isotherm and
thermodynamic studies
Osman Duman ⁎, Sibel Tunç, Tülin Gürkan Polat
Akdeniz University, Faculty of Science, Department of Chemistry, 07058 Antalya, Turkey

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

Article history: The adsorption characteristics of toxic Basic Red 9 (BR9) onto expanded vermiculite were determined. The effect
Received 16 December 2014 of contact time, initial dye concentration, pH and temperature on the adsorption process was investigated. The
Received in revised form 4 March 2015 adsorption amount of BR9 onto expanded vermiculite increased with increasing initial dye concentration and
Accepted 5 March 2015
pH, but it displayed a decrease with increasing temperature. The adsorption kinetics of BR9 followed the pseudo
Available online 25 March 2015
second-order model. Freundlich model provided the best fitting with the experimental adsorption isotherm data
Keywords:
obtained at 25, 35 and 45 °C. Thermodynamic parameters indicated that the adsorption process was exothermic,
Dye removal occurred spontaneously and led to an increase in the randomness at the solid/solution interface. The desorption
Basic Red 9 of dye from spent adsorbent was investigated in water, ethyl alcohol and benzyl alcohol. The sign of the zeta po-
Vermiculite tential value of clay changed from negative to positive due to the adsorption of BR9 onto expanded vermiculite.
Kinetics The adsorption of BR9 led to some changes in the FTIR spectrum of adsorbent material. The maximum adsorption
Isotherm capacity of expanded vermiculite for BR9 adsorption is comparable with other adsorbents reported in the litera-
Thermodynamic ture. Therefore, expanded vermiculite as an efficient and economical adsorbent may be used for the removal of
dyes from wastewaters.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction activated carbon, new adsorbents with low cost and high adsorption ef-
ficiency should be provided (Yu et al., 2010).
Several traditional methods including coagulation, flocculation, Vermiculite, a clay mineral with a porous structure, is a potential
ozonation, membrane-filtration, ion-exchange, oxidation and chemical candidate as an adsorbent material (Yu et al., 2010; Wen et al., 2013).
precipitation are known for the treatment of dye-containing effluents. It is very abundant and much cheaper in comparison with other clays
On the other hand, these techniques have some disadvantages such as such as montmorillonite, hectorite and saponite (Xu et al., 2003). It is
inefficiency at lower concentration, high energy and chemical reagents an expandable 2:1 mineral and often forms from alteration of mica
requirement, generation of toxic sludge or other wastes as byproduct (Howe-Grant, 1992; Aristov et al., 2000). The structure of vermiculite
that need careful treatment and disposal, high capital and operational is shown in Fig. S1 (Supporting Information). Its crystal lattice consists
costs, etc. (Roy et al., 2012). In addition, it is difficult to remove the of one octahedral sheet sandwiched between two opposing tetrahedral
dyes from wastewaters by biological treatment methods, because sheets. Tetrahedral sheets are composed of corner-linked tetrahedral,
many dyes used in the textile industry are stable to light and oxidizing whose central ions are Si4+ or Al3+. The octahedral sheet is composed
agents and resistant to aerobic digestion (Choi and Cho, 1996). of edge-shared octahedra with Mg2+, Al3+ or Fe2+. Due to isomorphic
Adsorption technique is effective and practical in the treatment of substitutions which are Al3+ for Si4+ substitution in tetrahedral sheets
wastewaters containing dyes owing to its high efficiency, simplicity and Mg2+ or Fe2+ for Al3+ substitution in octahedral sheets, vermicu-
and the availability of many adsorbents. The most used adsorbent to re- lite layers have permanent negative charges. These charges are com-
move the dyes from aqueous solutions is activated carbon. On the other pensated by the cations in the interlayer space (Tvardovski et al.,
hand, the application of activated carbon for a large-scale wastewater 1994; Poyato et al., 2002; Osman, 2006; Abollino et al., 2008).
treatment is limited because of its high cost. Therefore, instead of The use of vermiculite as an adsorbent has been studied for the re-
moval of Hg(II) (Do Nascimento and Masini, 2014), Cr(III) (Badawy
et al., 2010; Sis and Uysal, 2014), Cu(II) (Badawy et al., 2010), Pb(II)
⁎ Corresponding author. Tel.: +90 242 3102303; fax: +90 242 2278911.
(Hongo et al., 2012; Sis and Uysal, 2014), Zn(II) (Sis and Uysal, 2014),
E-mail addresses: osmanduman@akdeniz.edu.tr (O. Duman), stunc@akdeniz.edu.tr Cs (Suzuki et al., 2013), Ag(I) (Sari and Tüzen, 2013), Cr(VI) (Dultz
(S. Tunç). et al., 2012), Cd(II) (Panuccio et al., 2009), mineral and canola oils

http://dx.doi.org/10.1016/j.clay.2015.03.003
0169-1317/© 2015 Elsevier B.V. All rights reserved.
O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32 23

(Mysore et al., 2005), phthalate (Wen et al., 2013), basic cationic dye
(Choi and Cho, 1996), herbicides 2,4-D, diuron, alachlor, metolachlor
and triazine (de Rezende et al., 2011), dibenzothiophene (Froehner
et al., 2010), benzodiazepine (Carvalho et al., 2014) and cationic surfac-
tant cetylpyridinium chloride (Tarasevich et al., 2013) from aqueous so-
lution. Although there are many works published about the adsorption
of metal ions and organic compounds onto vermiculite, only one study
has been carried out to determine the dye adsorption properties of
vermiculite.
When pristine vermiculite flakes are strongly heated at high tem-
perature for a short time, expanded vermiculite is obtained (Zhang
et al., 2012a). If a heating process is applied to a vermiculite particle
rapidly above about 200 °C, the interlayer water in vermiculite turns
into steam. The pressure of the steam forces the silicate layers to
separate forming packets. This exfoliation creates a 10 to 20 times
volume expansion, but the basal dimensions of the particles are un-
changed (Kehal et al., 2010). Expanded vermiculite has a hydrophilic
nature and contains slit shaped pores (De Araujo Medeiros et al., Fig. 1. Chemical structure of C.I. Basic Red 9 (BR9).
2010). Expanded vermiculite is commonly used in industrial and ag-
ricultural applications due to its very low density, high absorption hydroxide were purchased from Merck (Germany). All other chemicals
capacity and excellent acoustical and thermal insulation properties were reagent grade. Deionized water was used in the experiments.
(Kehal et al., 2010).
On the other hand, only limited research has been performed on the 2.2. Preparation and purification of expanded vermiculite
adsorptive properties of expanded vermiculite. Nishikawa et al. (2012)
investigated the removal of zinc ions by adsorption onto expanded To prepare expanded vermiculite, vermiculite sample was placed in
vermiculite. They reported that Zn(II) adsorption isotherm data were a furnace heated to 500 °C. Then, the sample was kept at this tempera-
fitted well with the Freundlich model and the adsorption process was ture for 1–2 s and taken out of the furnace. The obtained sample was
spontaneous, favorable and endothermic. The adsorption of cadmium labeled as expanded vermiculite.
ions onto expanded vermiculite was studied by Mathialagan and After the expanded vermiculite was prepared, it was purified ac-
Viraraghavan (2003). In this study, optimum pH value for the adsorp- cording to the procedure described previously (Duman and Tunç,
tion of Cd(II) was found to be 6 and equilibrium time was determined 2008). Briefly, in this procedure, a suspension was obtained by adding
as 4 h. Turan and Ozgonenel (2013) investigated the adsorbent behavior of 100 g clay sample into 1 L deionized water. This suspension was
of expanded vermiculite to remove the copper ions from industrial stirred at 960 rpm for 24 h by Variomag Poly model magnetic stirrer
leachate. They reported that the removal of Cu(II) was 96.87% at pH 8, and filtered through filter paper. Purified vermiculite was dried at
adsorbent dosage of 10 g/L and contact time of 10 min. On the other 110 °C for 24 h. Dried samples were blended using a Waring Commer-
hand, in the literature, there are no reports about dye adsorption onto cial Blendor for 5 min at high speed. Afterwards, clay sample was sieved
expanded vermiculite. by 100 μm sieve. The particles under 100 μm were used in the further
Triarylmethane dyes form an important commercial dye class. experiments.
Triamino derivatives of triphenylmethane have the largest dye group
among the triarylmethane dye classes and the first member of this 2.3. Characterization of expanded vermiculite
group is C.I. Basic Red 9 (BR9) (Duman et al., 2011). BR9 is used com-
monly to color the products in textile, leather, paper and ink industries. X-ray diffraction (XRD) analysis of expanded vermiculite was per-
The discharge of BR9 from these industries into natural streams can formed using a Shimadzu XRD-6000 model diffractometer with Cu Kα
cause serious problems. BR9 is toxic and carcinogenic and has poor bio- tube (λ = 1.5406 Å, 40 kV, 30 mA) in the 2 theta range of 2 to 70° at
degradation (Huang et al., 2012). Therefore, wastewaters containing scanning rate 2°/min. For XRD measurements, divergence and scatter
dangerous dyes require treatment before being released into the envi- slits are 1° and the width of receiving slit is 0.3 mm. Chemical composi-
ronment (Prola et al., 2013). In this study, the use of expanded vermic- tions of vermiculite and expanded vermiculite were determined by
ulite was investigated for the removal of BR9 as a model dye from using X-ray fluorescence (XRF) spectrometry. Powdered samples were
aqueous solutions. pressed and analyzed by a Philips PW2404 X-ray fluorescence (XRF)
The aims of this study were to evaluate the adsorption efficiency of spectrometer equipped with a rhodium X-ray tube, a maximum
expanded vermiculite as an adsorbent for the removal of toxic Basic power of 3600 W and flow and scintillation detectors. Samples were an-
Red 9 (BR9) dye from aqueous solution, to investigate the effect of alyzed at 60 kV and 125 mA. The chemical compositions of investigated
contact time, initial dye concentration, pH and temperature on the ad- samples were determined using IQ + software. The cation exchange
sorption process, to determine the fitting of adsorption kinetics and iso- capacity (CEC) and the density values of vermiculite and expanded
therm data with various models, to calculate the thermodynamic vermiculite were determined by ammonium acetate method and
parameters and to understand the interactions between adsorbent picnometric method, respectively. Scanning electron microscope
material and dye molecule by zeta potential and FTIR analyses. (SEM) study was conducted at 15 kV using a Zeiss Leo-1430 model
scanning electron microscope to analyze the morphology of vermiculite
and expanded vermiculite. Before SEM analysis, the samples were coat-
2. Materials and methods ed with gold. The specific surface area value of expanded vermiculite
was obtained from N2 adsorption isotherm by a Quantachrome
2.1. Materials Autosorb-1-C/MS surface area analyzer. Prior to N2 adsorption isotherm
experiments, vermiculite and expanded vermiculite were degassed at
Vermiculite was supplied from Shanghai (China). C.I. Basic Red 9 130 °C (up to 10−6 Torr) for 12 h. The specific surface areas of samples
(BR9) was obtained from Sigma-Aldrich (Germany). The chemical were calculated from the N2 adsorption isotherm data by Brunauer–
structure of BR9 is shown in Fig. 1. Hydrochloric acid and sodium Emmett–Teller (BET) method, using relative pressure (P/P0) ≤ 0.26
24 O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32

(Brunauer et al., 1938). FTIR analysis was carried out to see the changes particles were measured by the instrument and then automatically con-
in the FTIR spectrum of vermiculite before and after BR9 adsorption. The verted to zeta potential by employing the Smoluchowski equation
FTIR spectra of vermiculite samples were recorded by KBr pellet tech- (Eq. (3));
nique. The concentration of vermiculite in KBr is 1% (w/w). A Bruker-
Tensor 27 model FTIR spectrometer was used to obtain the IR spectra 4πη
Zeta potential ¼ UE ð3Þ
of samples in the range of 400–4000 cm−1 at a resolution of 2 cm−1. ε
In FTIR analysis, 50 scans were accumulated.
where η is the viscosity of suspending liquid and ε is the dielectric con-
2.4. Adsorption experiments stant (Giese and Van Oss, 2002).
To investigate the effect of dye adsorption on the zeta potential value
A stock solution of BR9 (1 × 10−4 mol·L−1) was prepared in deion- of expanded vermiculite particles, a series of clay-BR9 suspension was
ized water. The desired concentrations of dye solution were obtained by prepared by adding of 50 mL dye solutions at different concentrations
diluting the stock BR9 solution in accurate proportions to obtain differ- (40 × 10−6 mol·L−1–140 × 10−6 mol·L−1) into the Erlenmeyer flasks
ent initial dye concentrations. containing 0.1 g clay sample. The suspensions were shaken in Nüve ST
The batch adsorption technique was used to obtain the kinetic 402 shaking water bath at 150 rpm and 25 °C for 180 min. Then, the
and isotherm data. After the temperature of dye solutions (50 mL) zeta potential measurements of expanded vermiculite-dye suspensions
containing different concentrations of BR9 in Erlenmeyer flasks were carried out. The average of 15 measurements was used to repre-
reached the desired temperature value in Nüve ST 402 shaking sent the zeta potential data. The percentage difference in the measured
water-bath, accurately weighed quantities of expanded vermiculite zeta potential values of suspensions was, on the average, less than 2% of
(0.1 g) were added to these flasks. After the Erlenmeyer flasks were the mean of the 15 values.
sealed, they were shaken at a constant shaking speed of 150 rpm.
At certain time intervals, the dye solutions were separated from ad- 2.6. Desorption studies
sorbent by centrifugation at 10,000 rpm for 5 min. Preliminary ex-
periments showed that the effect of separation time on the amount Desorption studies were carried out to determine the desorption ef-
of adsorbed dye by expanded vermiculite was negligible. The absor- ficiency of dye adsorbed by expanded vermiculite in the presence of
bance values of dye solutions in the supernatant solutions were mea- water, ethyl alcohol or benzyl alcohol. Firstly, expanded vermiculite
sured using a Cary 100 UV–visible spectrophotometer by monitoring was loaded with dye. To load the BR9 onto clay, 0.1 g expanded vermic-
the absorbance changes at a wavelength of maximum absorbance ulite samples were shaken in Erlenmeyer flasks at a constant shaking
(540 nm). Then, the absorbance data were converted into concentra- speed of 150 rpm and 25 °C for 3 h with 50 mL 100 × 10−6 mol·L−1
tion data using calibration relation pre-determined at 540 nm for BR9. Then, dye loaded vermiculites were centrifuged at 10,000 rpm for
BR9. The amount of adsorbed dye per gram adsorbent at time t, qt 5 min by a Hettich Universal II model centrifuge and dried at room tem-
(mol·g− 1), was calculated by perature for 72 h. In order to test the dye desorption from adsorbent,
BR9 loaded expanded vermiculite samples were added to Erlenmeyer
ðC0 −Ct ÞV flasks containing 50 mL deionized water, ethyl alcohol or benzyl alcohol
qt ¼ ð1Þ
W and shaked at 25 °C for 3 h. The desorbed BR9 concentrations were
measured spectrophotometrically. The amount of dye desorption from
where C0 and Ct are the concentration of dye at time 0 and time t, re- expanded vermiculite was calculated by Eq. (4)
spectively (mol·L− 1), V is the volume of solution (L) and W is the
mass of adsorbent (g). The amount of adsorbed BR9 per gram ex- qeðdesorptionÞ
panded vermiculite at equilibrium, qe (mol·g− 1), was obtained by Desorptionð%Þ ¼  100 ð4Þ
qeðadsorptionÞ
ðC0 −Ce ÞV
qe ¼ ð2Þ where qe(desorption) and qe(adsorption) are the amount of dye desorbed
W
from adsorbent and the amount of dye adsorbed onto adsorbent,
where Ce is the equilibrium concentration of BR9 (mol·L− 1). For the respectively.
adsorption isotherm studies, the equilibrium time was determined
from kinetic studies. 3. Results and discussion
To investigate the effect of temperature on the adsorption process,
adsorption experiments were conducted at 25, 35 and 45 °C. The effect 3.1. Characterization of expanded vermiculite
of the initial pH value was determined by adjusting the pH value of BR9
solutions in the range of 3.3 to 9.0 by using 0.1 mol·L− 1 HCl or The XRD pattern of expanded vermiculite is shown in Fig. 2A. The re-
0.1 mol·L−1 NaOH before the experiments. The pH values of solutions sults of XRD analysis indicated that sample consisted of mainly vermicu-
were measured with a Jenway 3040 ion analyzer using combined lite and small amounts of mica, illite and hydrobiotite. The chemical
glass electrode, which was calibrated with standard buffer solutions be- composition of expanded vermiculite obtained from XRF analysis is
fore use. In addition, adsorption isotherm experiments for the adsorp- given in Table S1 (Supporting Information). The chemical compositions
tion of BR9 onto untreated expanded vermiculite and untreated of vermiculite and expanded vermiculite are similar. These samples
vermiculite were carried out at 25 °C. All adsorption experiments have essentially Si4+, Mg2+, Al3+, Fe3+ and K+. The CEC values were de-
were repeated two times and reproducible within 4% error at most. termined to be 0.560 meq·(g vermiculite)−1 and 0.568 meq·(g expanded
Mean values are reported in this study. vermiculite)−1. The CEC value is almost the same for vermiculite and ex-
panded vermiculite. The density values of vermiculite and expanded ver-
2.5. Zeta potential measurements miculite are 2.894 g·cm−3 and 0.400 g·cm−3, respectively. After the
pristine vermiculite was kept at high temperature of 500 °C for 1–2 s,
To determine the change in the surface charge of clay particles with the density of expanded vermiculite decreased approximately 7 fold.
the adsorption of BR9 onto the expanded vermiculite, zeta potential SEM pictures of vermiculite and expanded vermiculite are illustrated in
measurements were carried out by a Nano-ZS model (Malvern) Zetasizer Fig. 2B. The expanded vermiculite was more voluminous than vermiculite
instrument which works with the technique of laser doppler electro- due to the vaporization of water from interlayer of vermiculite during the
phoresis (4 mW He–Ne, 633 nm). Electrophoretic mobilities (UE) of preparation of expanded vermiculite. The nitrogen adsorption isotherm
O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32 25

600

Intensity (CPS)
450
V
I V
300 V
V
HB V
HB
150 V

(A)
0
0 10 20 30 40 50 60 70

2 Theta (°)

(a) 2 __
µm 2 __
µm
(b)

(B) (B)

30
Adsorbed volume (cm .g )

Pore size distribution (cm .Å .g )

0.0010
-1
-1

-1

24 0.0008
3
3

0.0006

0.0004
18
0.0002

0.0000

12 0 20 40 60 80 100

Pore diameter (Å)

6
(C)
0
0.00 0.20 0.40 0.60 0.80 1.00
Relative pressure (P/P0)

Fig. 2. (A) XRD pattern of expanded vermiculite (V: Vermiculite, M: Mica, I: Illite, HB: Hydrobiotite), (B) SEM pictures of vermiculite (a) and expanded vermiculite (b) and (C) Nitrogen
adsorption isotherm of expanded vermiculite (Inset shows the pore size distribution of clay according to BJH theory).

data given in Fig. 2C were used to calculate the specific surface area of ex- in Fig. 2C (inset). This graph indicates that expanded vermiculite consists
panded vermiculite. The BET specific surface area was calculated to be of pores mainly with mesopore character. The total pore volume of adsor-
7.8 m2·g−1 for vermiculite and 9.8 m2·g−1 for expanded vermiculite by bent is 0.0043 cm3·g−1.
multipoint BET method. This result can be explained with the increase
in pore volume or the decrease in density of expanded vermiculite due 3.2. The effect of contact time and initial dye concentration
to vaporization of water from interlayer and increased separation of the
silicate layers. BET specific surface area values of vermiculite and expand- The adsorption of BR9 from aqueous solution by expanded vermicu-
ed vermiculite are in agreement with literature values. Perez-Maqueda lite as a function of contact time is shown in Fig. 3A. The amount
et al. (2003) have reported a BET specific surface area of 7 m2·g−1 for nat- adsorbed dye by expanded vermiculite was very high in the first
ural Mg2+-vermiculite. The BET specific surface area of expanded vermic- 10 min owing to the presence of large amount of adsorption sites on
ulite was found to be 9 m2·g−1 by Aristov et al. (2000) and 9.8 m2·g−1 by the adsorbent for the removal of dye. After 10 min, adsorption slightly
Huang et al. (2014). The pore size distribution graph obtained according increased with the increase in contact time until 25 min. After 25 min,
to Barrett–Joyner–Halenda (BJH) theory (Barrett et al., 1951) is shown there was no significant change in the adsorption amount of dye.
26 O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32

equilibrium time required for BR9 adsorption by vermiculite is relative-


qt (10 mol BR9/g vermiculite)

5.0 (A)
ly shorter than other adsorbent materials.
The adsorption amount of BR9 is significantly affected by the initial dye
4.0
concentration (Fig. 3A). As the initial BR9 concentration is increased from
40.0 × 10−6 mol·L−1 to 100 × 10−6 mol·L−1, the adsorption amount on
3.0 the expanded vermiculite increased from 1.84 × 10−5 mol·g−1 to
4.19 × 10−5 mol·g−1 at contact time of 240 min.
2.0
3.3. Adsorption kinetics
40.0x10 -6 mol.L-1 BR9
-5

1.0 60.0x10 -6 mol.L-1 BR9


80.0x10 -6 mol.L-1 BR9 The investigation of adsorption kinetics provides very useful infor-
100x10 -6 mol.L-1 BR9
0.0
mation in the design of adsorption systems. In this study, several kinetic
0 50 100 150 200 250
models including pseudo first-order, pseudo second-order, Elovich and
intra-particle diffusion were used to understand the kinetic behavior of
t (min)
adsorbent in the adsorption process. Pseudo first-order rate equation is
3.6 expressed as (Lagergren, 1898)
qe (10 mol BR9/g vermiculite)

(B)
ln ðqe −qt Þ ¼ ln qe −k1 t ð5Þ

3.5 where qe and qt are the amounts of BR9 adsorbed by adsorbent


(mol·g−1) at equilibrium and time t (min), respectively, and k1 is the
pseudo first-order rate constant (min−1). Rate constant, k1, can be cal-
culated from the slope of the plot of ln(qe − qt) versus t.
3.4 Pseudo second-order model can be expressed by the following linear
equation (McKay et al., 1999)
-5

t 1 1
¼ þ t ð6Þ
3.3 qt k 2 qe 2 qe
2 4 6 8 10
pH where k2 is the pseudo second-order rate constant (g·mol−1·min−1).
Initial sorption rate (h) is equal to k2·q2e (mol·g−1·min−1). qe and k2
3.6
parameters can be determined from the slope and intercept values of
(C)
qt (10 mol BR9/g vermiculite)

the plot of t/qt versus t.


3.4 The linearized form of Elovich equation is given in Eq. (7) (Shek
et al., 2009):
3.2
1 1
qt ¼ ln ðαβÞ þ ln t ð7Þ
β β
3.0
where α (mol·g− 1·min− 1) and β are Elovich constants. These con-
stants can be calculated from the plot of qt versus ln t.
-5

2.8 25 °C
35 °C Intra-particle diffusion equation can be described as (Weber and
45 °C
Morris, 1963):
2.6
0 50 100 150 200 250 0:5
qt ¼ k i t ð8Þ
t (min)
where ki is the intra-particle diffusion rate constant (mol·g−1·min−0.5).
Fig. 3. (A) The effect of contact time on the adsorption of BR9 (initial dye concentrations: The slope of the plot of qt versus t0.5 gives the rate constant, ki.
40.0 × 10−6 mol·L−1–100 × 10−6 mol·L−1, T: 25 °C, pH: 6.8 and adsorbent dosage: 2 g/L),
(B) the effect of pH on the adsorption of BR9 (initial dye concentration:
The values of pseudo first-order rate constant obtained from the
80.0 × 10−6 mol·L−1, T: 25 °C, adsorbent dosage: 2 g/L and contact time: 180 min) and plots of ln(qe − qt) versus t are given in Table 1 together with the
(C) the effect of temperature on the adsorption of BR9 (initial dye concentration: regression coefficients. The regression coefficients for the pseudo first-
80.0 × 10−6 mol·L−1, pH: 6.8, adsorbent dosage: 2 g/L and contact time: 0–240 min). order kinetic model are lower than 0.7822. Therefore, the adsorption
kinetics of BR9 onto expanded vermiculite does not follow the pseudo
first-order kinetic model.
Therefore, the necessary time to reach the equilibrium was 25 min. The The plots of t/qt versus t for the adsorption of BR9 onto clay at differ-
contact time was fixed as 180 min for all adsorption experiments, ex- ent initial dye concentrations give straight lines (Fig. 4). The obtained ki-
cept kinetic studies, to obtain reliable and consistent experimental netic parameters for the pseudo second-order model are listed in Table 1.
results. Contact time is one of the most important parameters for eco- The regression coefficients are higher than 0.9999. This result indicates
nomical wastewater treatment application and a short adsorption equi- that the adsorption kinetics of BR9 follows the pseudo second-order ki-
librium time favors the application of the adsorption process (Wang and netic model. The values of pseudo second-order rate constant and the ini-
Chu, 2011). For the adsorption of BR9 onto various adsorbents, the tial sorption rate displayed a change with the initial dye concentration
adsorption equilibrium time was determined to be 210 min for a and temperature. As the initial dye concentration was increased from
cellulose-based multicarboxyl adsorbent (Zhou et al., 2013), 550 min 40.0 × 10−6 mol·L−1 to 100 × 10−6 mol·L−1, the pseudo second-order
for activated carbon (Kong et al., 2014), 105 min for bottom ash and rate constant value decreased from 271.2 × 103 g·mol−1·min−1 to
90 min for deoiled soya (Gupta et al., 2008). Compared with previous 77.61 × 103 g·mol−1·min−1. In addition, k2 value increased with the in-
studies about BR9 adsorption onto various adsorbents, the adsorption crease in the solution temperature. In brief, the initial sorption rate value
O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32 27

Table 1
Kinetic parameters and regression coefficients for the adsorption of BR9 onto expanded vermiculite at different dye concentrations and temperatures.

Initial dye concentration Pseudo first-order Pseudo second-order Elovich equation


(mol·L−1)
T k1 r k2 h r α β r
(°C) (10−2 min−1) (103 g·mol−1·min−1) (10−5 mol·g−1·min−1) (mol·g−1·min−1) (×106)

40.0 × 10−6 25 1.854 0.7050 271.2 9.171 0.9999 1.423 × 1090 12.45 0.6621
60.0 × 10−6 25 0.8775 0.4910 144.8 10.12 0.9999 1.513 × 1024 2.847 0.8446
80.0 × 10−6 25 0.4112 0.6949 94.76 10.97 0.9999 4.769 × 1061 4.782 0.8455
100 × 10−6 25 0.8897 0.3202 77.61 13.71 1.000 6.844 × 1010 1.033 0.7941
80.0 × 10−6 25 0.4112 0.6949 94.76 10.97 0.9999 4.769 × 1061 4.782 0.8455
80.0 × 10−6 35 1.993 0.7822 102.4 11.28 0.9999 2.969 × 1031 2.778 0.8600
80.0 × 10−6 45 1.347 0.5883 151.6 15.80 1.000 4.545 × 1049 4.167 0.8736

exhibited an increase with the increase of initial dye concentration and 3.4. The effect of pH
temperature.
The parameters of the Elovich model were calculated by applying The pH value of solution can affect the surface charge of clay particles
of linear regression analysis to qt versus ln t data. The obtained pa- and sometimes the structure of dye molecules. Therefore, the pH value
rameters are given in Table 1. For the Elovich equation, the regres- of a solution is an important factor in the adsorption studies. The influ-
sion coefficients are in the range of 0.6621 to 0.8736. This means ence of initial pH value of dye solution on the BR9 adsorption amount by
that the fitting of experimental kinetics data to the Elovich model is vermiculite at a reaction time of 180 min is shown in Fig. 3B. The adsorp-
not good. tion amount of expanded vermiculite increased moderately with the in-
Finally, the adsorption kinetic data were analyzed with the intra- crease in the pH value. This result can be explained by the surface
particle diffusion model. According to this model, the plots of qt ver- charges of the adsorbate and adsorbent. In a previous study by Duman
sus t0.5 must pass through the origin and yield straight lines. On the and Tunç (2008), the isoelectric point (pHIEP) of expanded vermiculite
other hand, the obtained plots of qt versus t0.5 were not linear over was determined to be 2.57 by the zeta potential measurements. The
the whole time range. Therefore, intra-particle diffusion model is in- electrical charge at the clay surface/aqueous phase on protonation or
appropriate to predict the adsorption kinetics of BR9 onto expanded deprotonation of the surface hydroxyl can be ascribed as (Alkan et al.,
vermiculite over the whole range of process. This means that intra- 2005)
particle diffusion is not the rate-limiting step in the adsorption
process. ð10Þ
The magnitude of the activation energy (Ea) is useful in the determi-
nation of adsorption type. To calculate the activation energy of adsorp- ð11Þ
tion, pseudo second-order rate constant (k2) can be used in the
Arrhenius equation: and at the isoelectric point

ð12Þ
E 1
lnk2 ¼ lnA− a ð9Þ
RT where ░-S represents the surface of expanded vermiculite. It can be said
that the reaction in Eq. (10) and the reaction in Eq. (11) are responsible
for the surface charge of expanded vermiculite particles below the iso-
where A is the Arrhenius constant, R is the universal gas constant electric point and above the isoelectric point, respectively.
(8.314 J·mol−1·K−1) and T is the temperature (K). The value of Ea for Expanded vermiculite has a negative surface charge above pHIEP. In
the adsorption of BR9 by expanded vermiculite was calculated to be addition, BR9 had a positive charge in the studied pH range. Therefore,
18.40 kJ·mol− 1 from the plot of ln k2 versus 1/T (Fig. S2, Supporting when the pH value of solution was increased, a moderate increase oc-
Information). Low activation energy (0–40 kJ·mol−1) indicates that curred in the adsorption amount of dye molecules due to the electro-
physical adsorption is effective in the removal of BR9 from aqueous static interaction between expanded vermiculite and BR9.
solution by expanded vermiculite (Hamoudi and Belkacemi, 2013).
3.5. The effect of temperature

40.0x10 -6 mol.L-1 BR9 The adsorption of BR9 onto vermiculite from aqueous solution at 25,
12 60.0x10 -6 mol.L-1 BR9 35 and 45 °C is illustrated in Fig. 3C. As the temperature of solution in-
80.0x10 -6 mol.L-1 BR9
t/qt (10 min.mol .g )
-1

100x10 -6 mol.L-1 BR9 creased from 25 °C to 45 °C, the adsorption amount of expanded vermic-
ulite showed a decrease from 3.40 × 10− 5 mol·g−1 to 3.23 ×
-1

9
10−5 mol·g−1. This result indicates that the adsorption process has an
exothermic nature and the BR9 adsorption onto the expanded vermicu-
6 lite is favored at lower temperatures.
6

3
3.6. Adsorption isotherms

Adsorption isotherms are extremely important in determining the


0 maximum adsorption capacity of an adsorbent, in understanding the
0 50 100 150 200 250 nature of adsorption process and in the adjustment of optimum adsor-
t (min) bent dosage in the adsorption process (Njoku et al., 2014). The adsorp-
tion isotherm curves of BR9 obtained at 25, 35 and 45 °C is shown in
Fig. 4. The fit of pseudo second-order kinetic model for the adsorption of BR9 onto Fig. 5A. In this study, Langmuir, Freundlich and Temkin isotherm models
expanded vermiculite at 25 °C. were used for the interpretation of experimental data. The linear forms
28 O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32
qe (10 mol BR9/g vermiculite)
where qe is the amount of BR9 adsorbed by adsorbent at equilibrium
6 (A)
(mol·g−1), Ce is the equilibrium concentration of adsorbate
5 (mol·L−1), qmax is the maximum adsorption capacity at monolayer cov-
erage (mol·g−1), KL is the adsorption equilibrium constant related to
4 the energy of adsorption (L·mol−1), KF is a Freundlich constant
representing the adsorption capacity ((mol·g−1)(L·mol−1)1/n), n is a
2 constant depicting the adsorption intensity, K1 is the Temkin isotherm
energy constant (L·mol−1) and K2 is the Temkin isotherm constant.
-5

25 °C
1 The main difference among these isotherm models is in the variation
35 °C
45 °C
of heat of adsorption with the surface coverage. The Langmuir model as-
0 sumes uniformity, the Freundlich model assumes a logarithmic de-
0.0 0.8 1.6 2.4 3.2 4.0 crease and the Temkin model assumes a linear decrease in heat of
Ce (10-5 mol.L-1) adsorption with surface coverage (Ayranci and Duman, 2005).
The adsorption isotherm parameters obtained from the linear re-
gression analysis of Ce/qe versus Ce data for Langmuir model, ln qe ver-
qe (10 mol BR9/g adsorbent)

6.0 (B) sus ln Ce data for Freundlich model and qe versus ln Ce data for
Temkin model are given in Table 2 together with their corresponding
4.8 regression coefficients. The regression coefficients calculated from
Langmuir, Freundlich and Temkin isotherm models are very high
3.6 (r N 0.99) for all temperatures. Therefore, regression coefficients are
not decisive about which model represents the experimental data bet-
2.4 ter. These results are in agreement with the recent studies related to
the adsorption of phenolic compounds onto Starbon prepared from
-5

Expanded vermiculite
1.2 Raw expanded vermiculite starch and alginic acid polysaccharides (Parker et al., 2013) and the ad-
Raw vermiculite
sorption of Malachite Green onto rice husk activated carbon (Sharma,
0.0
2011). In addition, similar results were obtained for the adsorption of
0.0 1.0 2.0 3.0 4.0 5.0 6.0
some cationic surfactants, dyes and aromatic organic pollutants onto ac-
Ce (10-5 M)
tivated carbon cloth (Duman and Ayranci, 2006, 2010; Ayranci and
Duman, 2009, 2010). When the regression coefficients are insufficient,
Fig. 5. (A) Adsorption isotherms for the adsorption of BR9 onto expanded vermiculite in
the residual root mean square error (RMSE) and the composite fraction-
water at 25, 35 and 45 °C. (B) Adsorption isotherm parameters of Langmuir, Freundlich
and Temkin equations and error values for the adsorption of BR9 onto expanded vermic- al error (HYBRD) analyses can provide good results for the determina-
ulite, untreated expanded vermiculite and untreated vermiculite at 25 °C. Solid lines show tion of the best isotherm model representing the experimental data.
the Freundlich isotherm. (Dye concentrations: 30.0 × 10−6 mol·L−1– Mathematical expression of RMSE and HYBRD is given in Eqs. (16)
140 × 10−6 mol·L−1, pH: 6.8 and adsorbent dosage: 2 g/L).
and (17), respectively (Hadi et al., 2010; Tumsek and Avci, 2013);
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u n  2
of Langmuir, Freundlich and Temkin equations are given in Eqs. (13), u 1 X
(14) and (15), respectively (Smith, 1981; Allen et al., 2003): RMSE ¼ t q −qeðpredÞ ð16Þ
n−p i¼1 eð expÞ

Ce C 1 2
¼ e þ ð13Þ 2 3
qe qmax KL Xn q −q
6 eð expÞ eð pred Þ 7
HYBRD ¼ 4 5 ð17Þ
qeð expÞ
1 i¼1
lnqe ¼ ln K F þ ln Ce ð14Þ
n
where n is the number of experimental data, p is the number of param-
qe ¼ K1 ln K2 þ K1 ln Ce ð15Þ eters within the model equation, qe(exp) is the experimental qe at any Ce

Table 2
Adsorption isotherm parameters of Langmuir, Freundlich and Temkin equations and error values for the adsorption of BR9 onto expanded vermiculite at 25, 35 and 45 °C.

Langmuir parameters

T (°C) qmax (10−5 mol·g−1) KL (104 L·mol−1) r RMSE (10−6) HYBRD (10−7)

25 7.657 8.635 0.9910 1.968 8.032


35 7.232 8.011 0.9908 1.988 8.791
45 6.840 7.866 0.9959 1.340 4.907

Freundlich parameters

T (°C) KF (mol·g−1)·(L·mol−1)1/n 1/n r RMSE (10−6) HYBRD (10−7)

25 0.01761 0.5449 0.9933 1.536 6.741


35 0.01321 0.5284 0.9943 1.339 5.341
45 0.01041 0.5134 0.9961 1.515 4.834

Temkin parameters

T (°C) K1 (10−5 L·mol−1) K2 (105) r RMSE (10−6) HYBRD (10−7)

25 1.700 8.326 0.9912 2.153 10.09


35 1.603 7.761 0.9907 2.108 10.20
45 1.521 7.521 0.9940 1.621 8.309
O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32 29

and qe(pred) is the calculated qe value from isotherm model using the calculated from the Langmuir isotherm model are listed in Table 4. All
best fitted parameters. The smaller RMSE and HYBRD values indicate of the adsorbent materials have lower qmax value than vermiculite, ex-
the better fitting of isotherm model to the experimental data. cept activated carbon.
The calculated RMSE and HYBRD values are given in Table 2. The
magnitude of RMSE and HYBRD values for the Freundlich model is 3.7. Thermodynamic parameters
lower than those of Langmuir and Temkin models. Therefore,
Freundlich model fits best with the experimental isotherm data. The results obtained from adsorption isotherm study were used to
Freundlich constant, KF , representing the adsorption capacity de- evaluate the thermodynamic parameters including Gibbs free energy
creased with the increase in the solution temperature. In other change (ΔG°), enthalpy change (ΔH°) and entropy change (ΔS°). For
words, an increase in temperature leads to a decrease in the adsorp- adsorbates with weak charge, the Langmuir equilibrium constant (KL)
tion capacity of expanded vermiculite. In addition, adsorption can be used for the determination of ΔG° (Liu, 2009). The values of
process is favorable; because the values of 1/n are lower than 1 ΔG° for the adsorption of BR9 onto expanded vermiculite can be calcu-
(Table 2). lated using the following equation,
Untreated vermiculite may also be a good adsorbent for the adsorp-

tion of BR9 from aqueous solution. Moreover, the preparation of ex- ΔG ¼ −RT lnKL ð18Þ
panded vermiculite needs some energy. For practical applications, the
adsorption isotherm curves of BR9 adsorption onto untreated vermicu- where R is the gas constant (8.314 J·mol−1·K−1) and T is the absolute
lite, untreated expanded vermiculite and expanded vermiculite should temperature (K). ΔH° and ΔS° can be determined from van't Hoff equa-
be obtained to see the effect of exfoliation, washing, drying and blend- tion:
ing processes on the maximum adsorption capacity of adsorbent. The
adsorption isotherm curves of BR9 adsorption onto expanded vermicu- ΔS0 ΔH 0 1
lnKL ¼ − : ð19Þ
lite, untreated expanded vermiculite and untreated vermiculite obtain- R R T
ed at 25 °C are presented in Fig. 5B. Adsorption isotherm parameters of
Langmuir, Freundlich and Temkin models were calculated and listed in ΔG° values were determined from Eq. (18). By using Eq. (19), ΔH°
Table 3. For the adsorption of BR9 onto three adsorbents, the lowest and ΔS° values were calculated from the slope and intercept values of
RMSE and HYBRD values were obtained for Freundlich model the linear plot of ln KL versus 1/T (Fig. S3, Supporting Information).
(Table 3). This result indicates that Freundlich model has the best fitting The negative ΔG° value (−28.17 kJ·mol−1 at 25 °C) indicates that the
with the experimental isotherm data. The Freundlich constant, KF, adsorption process is spontaneous thermodynamically. According to
representing the adsorption capacity increases in the order untreated the negative ΔH° value (− 3.70 kJ·mol−1), the adsorption process is
vermiculite b untreated expanded vermiculite b expanded vermiculite exothermic. The positive ΔS° value (82.01 J·mol− 1·K− 1) illustrates
for the adsorption of BR9 dye. An increase in the specific surface area the increase in randomness at the solid/liquid interface during the ad-
of untreated vermiculite occurred after the exfoliation process. More- sorption process. Similar results were also reported for the adsorption
over, the pores were more accessible after the blending process for the of C.I. Basic Yellow 87 onto aluminiferous MCM41 (commercial meso-
dye molecules. The enhanced adsorption of BR9 is observed after the ex- porous adsorbent) by Wu et al. (2012).
foliation, washing, drying and blending processes of untreated
vermiculite. 3.8. The effect of dye adsorption on the zeta potential of expanded
Although a direct comparison of BR9 adsorption capacity of vermic- vermiculite
ulite with other adsorbents is difficult owing to different experimental
conditions and isotherm models, the maximum adsorption capacity The variation of zeta potential value of expanded vermiculite with
(qmax) is an important parameter. For the adsorption of BR9 onto ver- the adsorption of BR9 having different initial dye concentration in the
miculite and various adsorbents used in literature, the qmax values range of 0 mol·L− 1 to 140 × 10− 6 mol·L− 1 is shown in Fig. 6. The

Table 3
Adsorption isotherm parameters of Langmuir, Freundlich and Temkin equations and error values for the adsorption of BR9 onto expanded vermiculite, raw expanded vermiculite and raw
vermiculite at 25 °C.

Langmuir parameters

Adsorbent qmax KL r RMSE HYBRD


(10−5 mol·g−1) (104 L·mol−1) (10−6) (10−7)

Expanded vermiculite 7.657 8.635 0.9910 1.968 8.032


Raw expanded vermiculite 6.890 5.345 0.9929 1.429 7.141
Raw vermiculite 6.016 4.211 0.9843 1.730 8.333

Freundlich parameters

T KF 1/n r RMSE HYBRD


(°C) (mol·g−1)·(L·mol−1)1/n (10−6) (10−7)

Expanded vermiculite 0.01761 0.5449 0.9933 1.536 6.741


Raw expanded vermiculite 0.01332 0.5524 0.9944 1.090 6.424
Raw vermiculite 0.007394 0.5219 0.9928 1.530 6.328

Temkin parameters

T K1 K2 r RMSE HYBRD
(°C) (10−5 L·mol−1) (105) (10−6) (10−7)

Expanded vermiculite 1.700 8.326 0.9912 2.153 10.09


Raw expanded vermiculite 1.543 5.036 0.9889 2.054 14.94
Raw vermiculite 1.312 4.350 0.9779 2.509 27.23
30 O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32

Table 4
Comparison of the maximum adsorption capacity (qmax) values with literature data.

Adsorbent T (°C) pH qmax (mol·g−1) Reference

Fish bone 25 Natural pH 5.02 × 10−5 Kizilkaya (2012)


Bottom ash 30 9.0 2.12 × 10−5 Gupta et al. (2008)
Deoiled soya 30 9.0 3.99 × 10−5 Gupta et al. (2008)
Jalshakti® (starch based super absorbent polymer) 27 6.5 3.61 × 10−5 Dhodapkar et al. (2007)
Polyacrylamide/laponite nanocomposite hydrogel 30 7.5 7.34 × 10−5 Zhang et al. (2012b)
Activated carbon 32 8.0 44.8 × 10−5 Kumar and Sivanesan (2006)
Expanded vermiculite 25 6.8 7.66 × 10−5 This study

zeta potential value of expanded vermiculite-water suspension is study shows that ethyl alcohol is a potential medium in the regenera-
−37 mV in the absence of dye molecules. This means that the net sur- tion of expanded vermiculite and the recovery of BR9.
face charge of the adsorbent is negative in water. The zeta potential
value of expanded vermiculite increased with the addition of BR9 into 3.10. FTIR study
the clay suspension. For the equilibrium dye concentrations of
1.64 × 10− 5 mol·L−1 and 1.89 × 10− 5 mol·L−1, the zeta potential FTIR spectra of expanded vermiculite before and after BR9 adsorp-
values of expanded vermiculite were measured to be − 3.77 mV and tion are illustrated in Fig. 8. In the FTIR spectrum of expanded vermicu-
+0.897 mV, respectively. These results indicate that the zeta potential lite, the O–H stretching vibration was observed at 3717 cm− 1 for
value of the clay particles changed from negative to positive with the interlayer water molecules and at 3418 cm−1 for silanol groups. The
adsorption of dye molecules onto vermiculite. In the presence of spectrum of expanded vermiculite showed a weak absorption at
140 × 10−6 mol·L−1 BR9 (initial dye concentration), the zeta potential 1640 cm−1 representing the O–H bending vibration of water molecules.
value of vermiculite reached +5.37 mV. According to the results of zeta A strong band observed at 996 cm−1 is related to the Si–O stretching vi-
potential measurements, hydrogen bonding and electrostatic interac- bration. Two weak absorption peaks recorded at 820 cm−1 and
tions may play an important role in the adsorption process. The dye 687 cm−1 are assigned to the bending vibrations of Al–OH and Al–O, re-
molecules can form hydrogen bonding with vermiculite. In addition, spectively. The Si–O–Mg bending vibration in expanded vermiculite
dye dissolved in water has a positive charge on its “N” center. Therefore, was observed at 456 cm−1 (Arab et al., 2002; Da Fonseca et al., 2006a,
electrostatic interactions between negatively charged vermiculite parti- 2006b; Duman and Tunç, 2008).
cles and positively charged dye molecules may contribute to the adsorp- The bands in the FTIR spectrum of pure BR9 (Fig. 8) represent the
tion process. Due to the formation of hydrogen bonding and the stretching vibration of N–H at 3439 cm− 1 and 3307 cm− 1, the C–H
electrostatic attraction forces, the negative charge of the original ver- stretching of aromatic ring at 3186 cm−1, the bending vibration of N–
miculite is neutralized by the adsorption of positively charged BR9 mol- H at 1630 cm− 1, the stretching vibration of C–C in aromatic ring at
ecules. After a certain amount of dye adsorption, the zeta potential of 1585 cm− 1 and 1459 cm− 1, and the stretching vibration of C–N at
vermiculite becomes positive owing to the adsorption of positively 1369 cm−1. The in-plane C–H bending vibration appears at
charged BR9 molecules onto the particle surface of clay. 1169 cm−1. The bands obtained at 907 cm−1 and 843 cm−1 are related
to out-of-plane C–H bending vibration. N–H wagging vibrations are ob-
3.9. Desorption of BR9 from spent expanded vermiculite served at 563 cm−1 and 508 cm−1.
Some changes can be observed between the FTIR spectra of expand-
Desorption studies can contribute to the understanding of the regen- ed vermiculite before and after BR9 adsorption. Compared with the FTIR
eration of adsorbent. In this study, water, ethyl alcohol and benzyl alco- spectrum of pure expanded vermiculite, a shift from 3418 cm− 1 to
hol were used as the solvent to remove the dye from adsorbent. The 3405 cm−1 in the O–H stretching vibration of silanol groups of clay is
desorption amount of BR9 from expanded vermiculite in water, ethyl al- observed after the adsorption of BR9 onto the expanded vermiculite
cohol and benzyl alcohol is displayed in Fig. 7. In comparison with alco- (Özcan et al., 2006; Chen et al., 2007). In addition, the absorption band
hols, the desorption amount in water is very low. A similar result was corresponding to the N–H stretching vibration of BR9 was observed at
reported for the desorption of quinoline from spent adsorbents, granu- 3405 cm−1. As a result of the attachment of BR9 onto the adsorbent, a
lar activated carbon and bagasse fly ash (Rameshraja et al., 2012). The shift from 1369 cm− 1 to 1372 cm− 1 in the absorption peak of BR9
highest desorption amount was obtained with ethyl alcohol. This was observed after adsorption process. Additionally, the adsorption of
BR9 caused the shifts in the Al–O and Si–O–Mg bending vibrations of ex-
10 panded vermiculite to lower wavenumbers (685 cm−1 and 451 cm−1).
The results of FTIR study demonstrate that the silanol groups of
0
Zeta potential (mV)

30

-10
24
% Desorption

-20
18

-30
12

-40 6
0.0 0.6 1.2 1.8 2.4 3.0

Ce (10-5 mol.L-1) 0
Water Ethyl alcohol Benzyl alcohol
Fig. 6. The change of zeta potential value of expanded vermiculite with the adsorption of
BR9 in water (the initial BR9 concentrations: 0 mol·L−1–140 × 10−6 mol·L−1, adsorbent Fig. 7. The desorption of BR9 from expanded vermiculite in water, ethyl alcohol and benzyl
dosage: 2 g/L, contact time: 180 min and T: 25 °C). alcohol at 25 °C.
O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32 31

Expanded vermiculite before BR9 adsorption References

3717 Abollino, O., Giacomino, A., Malandrino, M., Mentasti, E., 2008. Interaction of metal ions

1640

820
3418
with montmorillonite and vermiculite. Appl. Clay Sci. 38, 227–236.

687
Alkan, M., Demirbaş, Ö., Doğan, M., 2005. Electrokinetic properties of sepiolite suspen-
sions in different electrolyte media. J. Colloid Interface Sci. 281, 240–248.
Allen, S.J., Gan, Q., Matthews, R., Johnson, P.A., 2003. Comparison of optimized models for

996
Transmittance

456
Only BR9 basic dye adsorption by kudzu. Bioresour. Technol. 88, 143–152.
Arab, M., Bougeard, D., Smirnov, K.S., 2002. Experimental and computer simulation study
of the vibrational spectra of vermiculite. Phys. Chem. Chem. Phys. 4, 1957–1963.

563
907

508
1459

843
3439

1630
3307

Aristov, Y.I., Restuccia, G., Tokarev, M.M., Buerger, H.D., Freni, A., 2000. Selective water
3186

1169
1369
sorbents for multiple applications. 11. CaCl2 confined to expanded vermiculite.

1585
Expanded vermiculite after BR9 adsorption React. Kinet. Catal. Lett. 71, 377–384.
Ayranci, E., Duman, O., 2005. Adsorption behaviors of some phenolic compounds onto
high specific area activated carbon cloth. J. Hazard. Mater. 124, 125–132.
3717

1585

1372
1640

685
3186

Ayranci, E., Duman, O., 2009. In-situ UV–visible spectroscopic study on the adsorption of

1169
3405

some dyes onto activated carbon cloth. Sep. Sci. Technol. 44, 3735–3752.
Ayranci, E., Duman, O., 2010. Structural effects on the interactions of benzene and naph-

451
thalene sulfonates with activated carbon cloth during adsorption from aqueous solu-

996
tions. Chem. Eng. J. 156, 70–76.
Badawy, N.A., El-Bayaa, A.A., AlKhalik, E.A., 2010. Vermiculite as an exchanger for
4000 3400 2800 2200 1600 1000 400 copper(II) and Cr(III) ions, kinetic studies. Ionics 16, 733–739.
Barrett, E.P., Joyner, L.G., Halenda, P.P., 1951. The determination of pore volume and area
Wavenumber / cm-1 distribution in porous substances. I. Computations from nitrogen isotherms. J. Am.
Chem. Soc. 73, 373–380.
Fig. 8. FTIR spectra of expanded vermiculite before and after BR9 adsorption and pure BR9. Brunauer, S., Emmett, P.H., Teller, E., 1938. Adsorption of gases in multimolecular layers.
J. Am. Chem. Soc. 60, 309–319.
Carvalho, A.J.P., Dordio, A.V., Ramalho, J.P.P., 2014. A DFT study on the adsorption of ben-
expanded vermiculite and the amine groups of BR9 are important in the zodiazepines to vermiculite surfaces. J. Mol. Model. 20, 2336.
Chen, H., Zheng, M., Sun, H., Jia, Q., 2007. Characterization and properties of sepiolite/
adsorption process. The formation of hydrogen bonding between the –
polyurethane nanocomposites. Mater. Sci. Eng. A Struct. 445–446, 725–730.
OH groups in the expanded vermiculite and the –NH2 groups of dye Choi, Y.S., Cho, J.H., 1996. Color removal from dye wastewater using vermiculite. Environ.
molecule plays a major role in the dye adsorption. Technol. 17, 1169–1180.
Da Fonseca, M.G., de Oliveira, M.M., Arakaki, L.N.H., 2006a. Removal of cadmium, zinc,
manganese and chromium cations from aqueous solution by a clay mineral.
4. Conclusions J. Hazard. Mater. 137, 288–292.
Da Fonseca, M.G., Wanderley, A.F., Sousa, K., Arakaki, L.N.H., Espinola, J.G.P., 2006b. Inter-
action of aliphatic diamines with vermiculite in aqueous solution. Appl. Clay Sci. 32,
Kinetic, isotherm, thermodynamic and desorption studies were car-
94–98.
ried out comprehensively for the adsorption of BR9 from aqueous solu- De Araujo Medeiros, M., de Oliveira, D.L., Sansiviero, M.T.C., Araujo, M.H., Lago, R.M., 2010.
tion onto expanded vermiculite. Experimental results showed that a Use of the glycerol by-product of biodiesel to modify the surface of expanded vermic-
short contact time of 25 min was sufficient to reach the equilibrium. ulite to produce an efficient oil absorbent. J. Chem. Technol. Biotechnol. 85, 447–452.
De Rezende, E.I.P., Peralta-Zamora, P.G., Abate, G., 2011. Sorption study of herbicides with
The increase in the temperature of the solution exhibited a negative ef- the clay minerals vermiculite and montmorillonite. Quim. Nova 34, 21–27.
fect on the adsorption process. On the other hand, the increase in the Dhodapkar, R., Rao, N.N., Pande, S.P., Nandy, T., Devotta, S., 2007. Adsorption of cationic
initial dye concentration and pH values caused an increase in the ad- dyes on Jalshakti®, super absorbent polymer and photocatalytic regeneration of the
adsorbent. React. Funct. Polym. 67, 540–548.
sorption amount. Among the studied kinetic models, the best fitting of Do Nascimento, F.H., Masini, J.C., 2014. Influence of humic acid on adsorption of Hg(II) by
experimental data was obtained with the pseudo second-order kinetic vermiculite. J. Environ. Manag. 143, 1–7.
model. The increase in the initial dye concentration and temperature Dultz, S., An, J.H., Riebe, B., 2012. Organic cation exchanged montmorillonite and vermic-
ulite as adsorbents for Cr(VI): effect of layer charge on adsorption properties. Appl.
values led to an increase in the initial sorption rates. Adsorption iso- Clay Sci. 67–68, 125–133.
therm studies displayed that Freundlich isotherm model fitted well Duman, O., Ayranci, E., 2006. Adsorption characteristics of benzaldehyde, sulphanilic acid
with the experimental equilibrium data. As the temperature of solution and p-phenolsulfonate from water, acid or base solutions onto activated carbon cloth.
Sep. Sci. Technol. 41, 3673–3692.
was increased, a decrease was observed in the adsorption capacity value Duman, O., Ayranci, E., 2010. Adsorptive removal of cationic surfactants from aqueous so-
of expanded vermiculite. Negative ΔG° values reflected the spontaneous lutions onto high-area activated carbon cloth monitored by in situ UV spectroscopy.
nature of adsorption process. Exothermic nature of adsorption was re- J. Hazard. Mater. 174, 359–367.
Duman, O., Tunç, S., 2008. Electrokinetic properties of vermiculite and expanded vermic-
vealed from negative ΔH° value. Positive ΔS° value illustrated the in-
ulite: effects of pH, clay concentration and mono- and multivalent electrolytes. Sep.
crease in the randomness at the adsorbent/solution interface. The zeta Sci. Technol. 43, 3755–3776.
potential value of expanded vermiculite was converted from negative Duman, O., Tunç, S., Kanci, B., 2011. Spectrophotometric studies on the interactions of C.I.
to positive due to the adsorption of BR9 onto expanded vermiculite. Basic Red 9 and C.I. Acid Blue 25 with hexadecyltrimethylammonium bromide in cat-
ionic surfactant micelles. Fluid Phase Equilib. 301, 56–61.
For the spent expanded vermiculite, the highest desorption amount Froehner, S., Machado, K.S., Falcao, F., 2010. Adsorption of dibenzothiophene by vermicu-
was obtained by ethyl alcohol solution. FTIR analysis confirmed the ad- lite in hydrophobic form, impregnated with copper ions and in natural form. Water
sorption of BR9. This study showed that expanded vermiculite has a Air Soil Pollut. 209, 357–363.
Giese, R.F., Van Oss, C.J., 2002. Colloid and Surface Properties of Clays And Related
high adsorption capacity and adsorption rate for the BR9 adsorption. Minerals. Marcel Dekker Inc., New York.
Therefore, it is a potential adsorbent candidate for the removal of dyes Gupta, V.K., Mittal, A., Gajbe, V., Mittal, J., 2008. Adsorption of basic fuchsin using waste
from wastewaters. materials—bottom ash and deoiled soya—as adsorbents. J. Colloid Interface Sci. 319,
30–39.
Hadi, M., Samarghandi, M.R., McKay, G., 2010. Equilibrium two-parameter isotherms of
Acknowledgments acid dyes sorption by activated carbons: study of residual errors. Chem. Eng. J. 160,
408–416.
Hamoudi, S., Belkacemi, K., 2013. Adsorption of nitrate and phosphate ions from aqueous
Central Laboratory of Middle East Technical University and TUBITAK solutions using organically-functionalized silica materials: kinetic modeling. Fuel 110,
are acknowledged for determining the some characterization properties 107–113.
of clay sample. Hongo, T., Yoshino, S., Yamazaki, A., Yamasaki, A., Satokawa, S., 2012. Mechanochemical
treatment of vermiculite in vibration milling and its effect on lead (II) adsorption
ability. Appl. Clay Sci. 70, 74–78.
Appendix A. Supplementary data Howe-Grant, M., 1992. Kirk-Othmer Encyclopedia of Chemical Technology. 4th edition.
vol. 6. John Wiley & Sons, New York.
Huang, L.H., Kong, J.J., Wang, W.L., Zhang, C.L., Niu, S.F., Gao, B.Y., 2012. Study on Fe(III)
Supplementary data to this article can be found online at http://dx. and Mn(II) modified activated carbons derived from Zizania latifolia to removal
doi.org/10.1016/j.clay.2015.03.003. basic fuchsin. Desalination 286, 268–276.
32 O. Duman et al. / Applied Clay Science 109–110 (2015) 22–32

Huang, W.Y., Li, D., Liu, Z.Q., Tao, Q., Zhu, Y., Yang, J., Zhang, Y.M., 2014. Kinetics, isotherm, Sari, A., Tüzen, M., 2013. Adsorption of silver from aqueous solution onto raw vermiculite
thermodynamic, and adsorption mechanism studies of La(OH)3-modified exfoliated and manganese oxide-modified vermiculite. Microporous Mesoporous Mater. 170,
vermiculites as highly efficient phosphate adsorbents. Chem. Eng. J. 236, 191–201. 155–163.
Kehal, M., Reinert, L., Duclaux, L., 2010. Characterization and boron adsorption capacity of Sharma, Y.C., 2011. Adsorption characteristics of a low-cost activated carbon for the rec-
vermiculite modified by thermal shock or H2O2 reaction and/or sonication. Appl. Clay lamation of colored effluents containing Malachite Green. J. Chem. Eng. Data 56,
Sci. 48, 561–568. 478–484.
Kizilkaya, B., 2012. Usage of biogenic apatite (fish bones) on removal of basic fuchsin dye Shek, T.H., Ma, A., Lee, V.K.C., McKay, G., 2009. Kinetics of zinc ions removal from effluents
from aqueous solution. J. Dispers. Sci. Technol. 33, 1596–1602. using ion exchange resin. Chem. Eng. J. 146, 63–70.
Kong, J., Huang, L., Yue, Q., Gao, B., 2014. Preparation of activated carbon derived from Sis, H., Uysal, T., 2014. Removal of heavy metal ions from aqueous medium using
leather waste by H3PO4 activation and its application for basic fuchsin adsorption. Kuluncak (Malatya) vermiculites and effect of precipitation on removal. Appl. Clay
Desalin. Water Treat. 52, 2440–2449. Sci. 95, 1–8.
Kumar, K.V., Sivanesan, S., 2006. Isotherm parameters for basic dyes onto activated car- Smith, J.M., 1981. Chemical Engineering Kinetics. McGraw-Hill Company.
bon: comparison of linear and non-linear method. J. Hazard. Mater. 129, 147–150. Suzuki, N., Ozawa, S., Ochi, K., Chikuma, T., Watanabe, Y., 2013. Approaches for cesium
Lagergren, S., 1898. About the theory of so-called adsorption of soluble substances. uptake by vermiculite. J. Chem. Technol. Biotechnol. 88, 1603–1605.
Kungliga Svenska Vetenskapsakademiens Handlingar 24, 1–39. Tarasevich, Y.I., Krysenko, D.A., Polyakov, V.E., Demchenko, V.Y., 2013. A study of
Liu, Y., 2009. Is the free energy change of adsorption correctly calculated? J. Chem. Eng. cetylpyridinium cation sorption on Na-vermiculite. Colloid J. 75, 606–610.
Data 54, 1981–1985. Tumsek, F., Avci, O., 2013. Investigation of kinetics and isotherm models for the Acid Or-
Mathialagan, T., Viraraghavan, T., 2003. Adsorption of cadmium from aqueous solutions ange 95 adsorption from aqueous solution onto natural minerals. J. Chem. Eng. Data
by vermiculite. Sep. Sci. Technol. 38, 57–76. 58, 551–559.
McKay, G., Ho, Y.S., Ng, J.C.Y., 1999. Biosorption of copper from waste waters: a review. Turan, N.G., Ozgonenel, O., 2013. Optimizing copper ions removal from industrial leachate
Sep. Purif. Methods 28, 87–125. by explored vermiculite—a comparative analysis. J. Taiwan Inst. Chem. Eng. 44,
Mysore, D., Viraraghavan, T., Jin, Y.C., 2005. Treatment of oily waters using vermiculite. 895–903.
Water Res. 39, 2643–2653. Tvardovski, A.V., Fomkin, A.A., Tarasevich, Y.I., Polyakova, I.G., Serpinski, V.V., Guseva, I.M.,
Nishikawa, E., Neto, A.F.A., Vieira, M.G.A., 2012. Equilibrium and thermodynamic studies 1994. Investigation of cation-substituted vermiculite deformation upon water vapor
of zinc adsorption on expanded vermiculite. Adsorpt. Sci. Technol. 30, 759–772. sorption. J. Colloid Interface Sci. 164, 114–118.
Njoku, V.O., Foo, K.Y., Asif, M., Hameed, B.H., 2014. Preparation of activated carbons from Wang, Y., Chu, W., 2011. Adsorption and removal of a xanthene dye from aqueous solu-
rambutan (Nephelium lappaceum) peel by microwave-induced KOH activation for tion using two solid wastes as adsorbents. Ind. Eng. Chem. Res. 50, 8734–8741.
acid yellow 17 dye adsorption. Chem. Eng. J. 250, 198–204. Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon solution. J. Sanit. Eng. Div.
Osman, M.A., 2006. Organo-vermiculites: synthesis, structure and properties. Platelike Am. Soc. Civ. Eng. 89, 31–59.
nanoparticles with high aspect ratio. J. Mater. Chem. 16, 3007–3013. Wen, Z.D., Gao, D.W., Li, Z., Ren, N.Q., 2013. Effects of humic acid on phthalate adsorption
Özcan, A., Öncü, E.M., Özcan, A.S., 2006. Adsorption of Acid Blue 193 from aqueous solu- to vermiculite. Chem. Eng. J. 223, 298–303.
tions onto DEDMA-sepiolite. J. Hazard. Mater. 129, 244–252. Wu, X., Hui, K.N., Hui, K.S., Lee, S.K., Zhou, W., Chen, R., Hwang, D.H., Cho, Y.R., Son, Y.G.,
Panuccio, M.R., Sorgona, A., Rizzo, M., Cacco, G., 2009. Cadmium adsorption on vermicu- 2012. Adsorption of basic yellow 87 from aqueous solution onto two different meso-
lite, zeolite and pumice: batch experimental studies. J. Environ. Manag. 90, 364–374. porous adsorbents. Chem. Eng. J. 180, 91–98.
Parker, H.L., Budarin, V.L., Clark, J.H., Hunt, A.J., 2013. Use of Starbon for the adsorption and Xu, J., Meng, Y.Z., Li, R.K.Y., Xu, Y., Rajulu, A.V., 2003. Preparation and properties of
desorption of phenols. ACS Sustain. Chem. Eng. 1, 1311–1318. poly(vinyl alcohol)-vermiculite nanocomposites. J. Polym. Sci. B Polym. Phys.
Perez-Maqueda, L.A., Balek, V., Poyato, J., Perez-Rodriguez, J.L., Subrt, J., Bountsewa, I.M., 41, 749–755.
Beckman, I.N., Malek, Z., 2003. Study of natural and ion exchanged vermiculite by em- Yu, X., Wei, C., Ke, L., Hu, Y., Xie, X., Wu, H., 2010. Development of organovermiculite-
anation thermal analysis, TG, DTA and XRD. J. Therm. Anal. Calorim. 71, 715–726. based adsorbent for removing anionic dye from aqueous solution. J. Hazard. Mater.
Poyato, J., Perez-Maqueda, L.A., Justo, A., Balek, V., 2002. Emanation thermal analysis of 180, 499–507.
natural and chemically-modified vermiculite. Clay Clay Miner. 50, 791–798. Zhang, H., Yu, J., Kuang, D., 2012a. Effect of expanded vermiculite on aging properties of
Prola, L.D.T., Machado, F.M., Bergmann, C.P., de Souza, F.E., Gally, C.R., Lima, E.C., Adebayo, bitumen. Constr. Build. Mater. 26, 244–248.
M.A., Dias, S.L.P., Calvete, T., 2013. Adsorption of Direct Blue 53 dye from aqueous Zhang, X., Zheng, S., Lin, Z., Zhang, J., 2012b. Removal of basic fuchsin dye by adsorption
solutions by multi-walled carbon nanotubes and activated carbon. J. Environ. onto polyacrylamide/laponite nanocomposite hydrogels. Synth. React. Inorg. Met.
Manag. 130, 166–175. 42, 1273–1277.
Rameshraja, D., Srivastava, V.C., Kushwaha, J.P., Mall, I.D., 2012. Quinoline adsorption onto Zhou, Y., Zhang, M., Hu, X., Wang, X., Niu, J., Ma, T., 2013. Adsorption of cationic dyes on a
granular activated carbon and bagasse fly ash. Chem. Eng. J. 181–182, 343–351. cellulose-based multicarboxyl adsorbent. J. Chem. Eng. Data 58, 413–421.
Roy, A., Chakraborty, S., Kundu, S.P., Adhikari, B., Majumder, S.B., 2012. Adsorption of
anionic-azo dye from aqueous solution by lignocellulose-biomass jute fiber: equilib-
rium, kinetics, and thermodynamics study. Ind. Eng. Chem. Res. 51, 12095–12106.

You might also like