Você está na página 1de 11

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

Contents lists available at ScienceDirect

Process Safety and Environmental Protection


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

Adsorptive removal of basic dyes from aqueous


solutions by surfactant modied bentonite clay
(organoclay): Kinetic and competitive adsorption
isotherm
T.S. Anirudhan , M. Ramachandran
Department of Chemistry, University of Kerala, Kariavattom, Trivandrum 695581, India

a r t i c l e

i n f o

a b s t r a c t

Article history:

Cationic surfactant (Hexadecyltrimenthylammonium chloride) modied bentonite clay was

Received 19 December 2014

prepared and systematically studied for its adsorption behavior as an efcient adsorbent

Received in revised form 19

for the removal of basic dyes such as methylene blue (MB), crystal violet (CV) and Rho-

February 2015

damine B (RB) from aqueous phase. Organo modied clay shows better capacity for the

Accepted 1 March 2015

removal of three dyes. The adsorption process was found to be dependent on pH and initial

Available online 14 March 2015

dye concentration. The maximum dye sorption was found to be at a pH of 9.0 (99.99% for
MB, 95.0% for CV and 83.0% for RB). The adsorption capacity for the dyes was found to be

Keywords:

399.74, 365.11 and 324.36 mol/g for MB, CV and RB, respectively at 30 C. The equilibrium

Bentonite

uptake was attained within 240 min. The kinetic studies were revealed that sorption follows

Organoclay

a pseudo-second-order kinetic model which indicates chemisorption between adsorbent

Adsorption isotherm

and adsorbate molecules. Adsorption isotherm indicates non-energetically adsorption sites

Adsorption kinetics

which t with Freundlich isotherm model. The tness of kinetics and isotherm models was

Competitive adsorption isotherm

evaluated by using HYBRID error analysis function. Competitive adsorptions of dyes were

Basic dyes

studied by using binary component systems.


2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1.

Introduction

Industries such as textiles, leather, paper-making, plastics,


food, rubber, and cosmetics use different types of dyestuffs,
which also appear in the efuents discharged from some of
these industries. Generally dyes are stable to light, heat and
oxidizing agents, and are usually non-biodegradable (Wang
et al., 2006). Several dyes make their presence strikingly visible as they impart color to the water bodies. Color in water
bodies affects aquatic diversity by blocking the passage of sunlight. Further, color in water bodies has an adverse aesthetic
effect.
Methylene blue on contact with skin may cause discoloration, feeling of cold, redness or dryness (due to chronic

exposure). Ingestion of this dye may cause gastrointestinal


irritation, discoloration of oral mucosa, irritation of lips,
mouth and throat, paleness of complexion, lack of coordination or drowsiness. Crystal violet dye belongs to triphenyl
carbocation group widely used for the dying of wool, silk,
cotton, nylon, paper, leather, etc. It is also used in biological
staining and as a dermatological agent, and in veterinary
medicine. It is the brightest class of basic dye having high
tinctorial value. Crystal violet can produce human bladder
cancer and cancer in the digestive system of other animals
(Rammel et al., 2011). Higher concentration ingestion causes
nausea, vomiting and central nervous system depression.
Discharge of Rhodamine B into the hydrosphere can cause
environmental degradation. In California, Rhodamine B is

Corresponding author. Tel.: +91 4712418782.


E-mail address: tsani@rediffmail.com (T.S. Anirudhan).
http://dx.doi.org/10.1016/j.psep.2015.03.003
0957-5820/ 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

216

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

suspected to be carcinogenic and thus products containing it


must contain a warning on its label (Ashly et al., 2014).
Removal methods of dyes in efuents may be divided
into three main categories: physical, chemical and biological (Dhodapkar et al., 2006). Physical methods are found to
be easier and economical. Adsorption process belongs to this
category and this is found to be very effective and is applied in
liquid phase. Various adsorbents can be used for the removal
of dyes from industrial efuents. Activated carbon is very
expensive and the regeneration of the adsorbent for several steps may cause difculties (Beyene, 2014). Consequently
this cost problem has led to a search for cheap and efcient
alternate materials including clays such as sepiolite, zeolite,
montmorillonite, smectite, bentonite, alunite, and perlite. The
wide usefulness of these different kinds of clay is a result
of their high specic surface area, their high chemical and
mechanical stability, and a variety of surface and structural
properties (Malik, 2003; zacar and Sengil, 2003; Dogan and
Alkan, 2003; Robinson et al., 2002).
Bentonite, which is primarily smectite clay, is composed
of two silica tetrahedral sheets with a central Al octahedral
sheet. It has permanent negative charges that arise due to the
isomorphous substitution of Al3+ for Si4+ in the tetrahedral
layer and Mg2+ for Al3+ in the octahedral layer. This negative
charge is balanced by the presence of exchangeable cations
(Na+ , Ca2+ , etc.) in the lattice structure. The inorganic cations
present in the interlayer are exchanged by organic cations
by means of ion exchange mechanism. In aqueous solution,
water molecules enter into the lattice structure causing the
clays to swell. The introduction of organic cation changes the
clay from hydrophilic to hydrophobic (Baskaralingam et al.,
2006). The surface modication of natural clays by organic
cations leads to an increase in adsorption capacity. The surface modied clays were used as adsorbents for the removal
of phenolic compounds and pesticides (Ozcan et al., 2004;
Bartelt-Hunt et al., 2003; Akcay, 2004; Lee et al., 2002).
Various adsorbents are used for the removal of basic dyes
from aqueous phase. Natural untreated clay (NUC) for the
removal of Basic Yellow 2 (BY2) from aqueous solution in
batch system was studied by ztrk and Malkoc (2014). Nitrogen sorption measurements were employed to investigate the
variation in surface and pore properties after dye adsorption.
The calculated activation energy of adsorption was found to be
5.24 kJ/mol, and indicates that the adsorption was enhanced
by physisorption. In another study adsorptions of Rhodamine
B (Rh-B) and methylene blue by sugarcane bagasse of different surface areas were compared by Zhang et al. (2013). In their
study there was a small gain in the amount of dye removed by
increasing bagasse surface area from 0.57 m2 /g to 1.81 m2 /g.
Methylene blue adsorption was less sensitive to surface area
change than Rh-B adsorption. They found that the differences in adsorption performances between these dyes have
been related to the molecular structure of the dyes and the
surface chemistry of bagasse. The adsorption of RhodamineB (Rh-B) onto treated rice husk-based activated carbon was
investigated by Wang et al. (2014). They found that pH had a
little effect on Rh-B adsorption. Adsorption kinetics, equilibrium and thermodynamics study was carried out. The results
demonstrated that pseudo-second order kinetic model represented the adsorption kinetics of Rh-B well. Adsorption
equilibrium data were well described by Langmuir isotherm
model.
TiO2 alone and in combination with rare earth ions [La3+ ,
Ce4+ , Pr3+ and Gd3+ ] in the presence and absence of ultrasound

for the adsorption of Reactive Blue 21 dye was studied by


Srivastava et al. (2013). They found that RE-TiO2 has greater
afnity for dye removal than TiO2 alone. The adsorption was
best explained by WeberMorris intraparticle diffusion and
Boyd kinetic models. Natarajan et al. (2014) reported that
surface hydroxyl group enriched titanium dioxide nanotube
(TNT) by hydrothermal method was successfully applied for
preferential adsorption of methylene blue (MB) dye. From the
results it was reported that the preferential interaction of MB
on TNT is due to the electrostatic interaction between the
cationic MB and negatively charged TNT surface. Adsorption
was well studied by Langmuir, Freundlich and Sips isotherm
models and pseudo-rst and second-order kinetic models.
In the present study, the surface of the commercial
bentonite is converted into organo modied one with hexadecyltrimethylammonium chloride surfactant, designated as
organoclay. The objective of the present work is to examine the
effectiveness of organoclay for the removal of basic dyes such
as methylene blue (MB), crystal violet (CV) and Rhodamine B
(RB) from aqueous solutions.

2.

Materials and methods

2.1.

Materials

The basic dyes used in this study were methylene blue (MB),
Rhodamine B (Rh-B) and crystal violet (CV). The dyes were
used as such without any further purication and were purchased from Fluka, Switzerland. The structures of these dyes
are shown in Fig. 1. The aqueous solutions of dyes were
prepared by dissolving denite amount of dyes in distilled
water within the concentration range 2001000 mol/L. The
surfactant hexadecyltrimenthylammonium chloride was purchased from Aldrich Chemicals (USA). Commercially available
bentonite clay was purchased from SigmaAldrich Chemie
(Germany). The cation exchange capacity (CEC) of the bentonite clay was found to be 0.52 mmol/g.

2.2.

Preparation of organoclay

Commercially available bentonite clay was rst converted into


Na+ saturated form by stirring with 0.1 M NaCl for 4 h. The Nabentonite obtained is ltered and dried at a temperature of
70 C in an air oven for overnight. The product is powdered
and again converted into organo form by treating with surfactant. For the synthesis, the amount of HDTMA corresponding
to 100% of the cation exchange capacity of the Na-bentonite
was dissolved in 1 L of distilled water and added to 20 g Nabentonite. The suspension was shaken for 6 h at 60 C, ltered,
washed several times with distilled water and then dried at
70 C for 24 h. The product obtained is powdered and sieved to
get a particle size of 0.096 mm (Scheme 1).

2.3.

Instrumental studies

Surface areas of material used were determined using N2


adsorption isotherm. FTIR spectra were obtained using Nicolet
400 D spectrophotometer. Morphological features of samples
were obtained with a Philips XC 30 CP Scanning Electron
Microscope. X-ray diffractograms were obtained using Rigaku
difractometer with Cu K radiation. Point of zero charge
(pHzpc ) was determined by potentiometric titration.

217

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

H 3C

CH3
N

Cl-

N
C

H3C

CH3

Cl-

CH3

CH3
N
H 3C

H 3C
CH3

(A)

CH3

(B)

COOH
CH 3

CH 3
O

CH 3

ClN

CH 3

(C)

Fig. 1 Structures of basic dyes: (A) Methylene Blue, (B) Crystal violet and (C) Rhodamine B.

2.4.

Point of zero charge (pHzpc )

Point of zero charge of an adsorbent is dened as the pH at


which the surface charge becomes zero. Potentiometric titration was used to calculate the point of zero charge (pHzpc ).
A denite amount of adsorbent (0.1 g) was titrated with 0.01
and 0.001 M NaCl. The titrations were performed by adding
different amounts of acid or base to about 50 mL of each
concentration of NaCl solutions. The solutions were taken in

different bottles and were shaken to an equilibrium time for


8 h. After reaching equilibrium, the pH was recorded after a
stable value was obtained using a -362 pH meter. Blank experiments were also conducted. The net surface charge density
 0 was calculated using the equation

0 =

F(CA CB ) + ([OH ] [H+ ])


A

Scheme 1 Preparation of organoclay.

(1)

218

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

Parameters
Surface area (m2 /g)
Zeropoint charge
Density (g/cm3 )
Porosity (mL/g)
Particle size (mm)
CEC (meq/g)

Na-bentonite
36.5
4.5
1.25
0.28
0.096
0.52

Organoclay
27.9
7.8
1.48
0.40
0.096
0.71

where F is Faradays constant (C/eq) and A is the surface area


of the suspension surface (cm2 /L). CA and CB are the concentrations of acid and base after each addition (eq/L) and [H+ ]
and [OH ] are the equivalents of H+ and OH ions bound to
suspension surface (eq/cm2 ).

2.5.

Volume of N2 adsorbed
[mL(STP)/g]

40

Table 1 Physical properties of Na-bentonite and


organoclay.

30
20
organo clay
10

Na-bentonite

0
0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

P/Po
Fig. 2 Adsorption isotherms of N2 /77 K of Na-bentonite
and organoclay.

Adsorption experiments

The concentration of dyes in samples was determined during adsorption experiments (single-component system) by
using UV-Visible spectrophotometer (model Jasco model V530) at maximum wavelength absorption for each dye. The
max for each dye was 662, 591 and 554 nm for MB, CV and
RB respectively. Adsorption equilibrium studies were done by
treating 0.1 g of organoclay with 50 mL of various dye solutions
(2001000 mol/L) in a stoppered bottle and they were placed
in a temperature controlled water bath shaker (Labline, India).
The pH of the solution was calibrated by using NaOH and HCl
with a -362 pH meter (systronics). After the completion of
the reaction period (4 h), the asks were taken out and the dye
concentrations were determined.
Adsorption studies were also performed to determine the
competition among various binary solution mixtures (MB-RB,
MB-CV and RB-CV). Batch experiments were performed using
50 mL binary mixture solutions containing varied concentration of dyes with 0.1 g of the adsorbent in 100 mL stoppered
asks. The contents were shaken over a water bath shaker
with a constant speed of 200 rpm at pH 9.0 and 30 C. Then, the
equilibrium samples were withdrawn and centrifuged, and
supernatants were analyzed for dye concentrations by using
the following equations.
CA =

KB2 d1 KB1 d2
KA1 KB2 KA2 KB1

(2)

CB =

KA1 d2 KA2 d1
KA1 KB2 KA2 KB1

(3)

where subscripts A and B represent components in a binary


solute system. KA1 , KB1 , KA2 and KB2 are the calibration
constants for components A and B at 1 and 2 , respectively.
d1 and d2 are the optical densities at 1 and 2 , respectively.
CA and CB are the concentrations of components A and B in
binary solution after equilibrium.

3.

Result and discussion

3.1.

Characterization of organoclay

The surface properties and chemical composition of adsorbent


were examined by using different techniques. The important
physical property of the Na-bentonite and organoclay is
listed in Table 1. Na-bentonite contains 53.81% SiO2 , 18.43%
Al2 O3 , 3.76% Fe2 O3 , 2.11% CaO, and 1.76% MgO. The chemical

composition of organoclay indicates that there is a decrease in


content of the Fe2 O3 and Al2 O3 , which implies the adsorption
of surfactant through the exchange of inorganic cations.
Fig. 2 shows the adsorption isotherms of N2 /77 K of Nabentonite and organoclay. Both isotherms correspond to
intermediate shape between groups I and II of the B.D.D.T
classication (Gregg and Singh, 1982), typical of micro porous
materials associated with some mesopores. The BET surface
area was calculated from the adsorption data and the values were found to be 36.5 and 27.9 m2 /g for Na-bentonite and
organoclay respectively. Substances of high adsorption capacity exhibit isotherm with a round knee and those with low
adsorption capacity exhibit isotherm with a sharper knee.
Isotherms of both Na-bentonite and organoclay showed high
adsorption capacity, so that the linear branch at high P/P0 was
reached more gradually with a round knee at low P/P0 values. The higher surface area of Na-bentonite is probably due
to the enhanced adsorption of N2 in the wider micropores
and mesopores. The total pore volume determined from the
N2 adsorption data at P/P0 of 0.95 was found to be 0.37 and
0.24 mL/g for Na-bentonite and organoclay respectively.
FTIR analysis permits spectrophotometric observation and
gives means for the identication of surface functional groups
present on the adsorbent. The spectra analysis was done
within the range of 4004000 cm1 . FTIR spectra are shown
in Fig. 3. An examination of the adsorbent surface before and
after modication provides information regarding the surface
groups that might have participated in the adsorption reaction and also indicates the surface site(s) on which adsorption
has taken place. The broad absorption bands at 3716 and
3624 cm1 are due to the O H stretching vibrations of the
Si OH (silanol) and Al OH groups of Na-B respectively. The
peak at 3294 cm1 corresponds to the OH vibrations band of
the silicate skeleton. The strong band at 1024 cm1 represents the Si O Si groups of the tetrahedral sheets, while the
bands at 715, 616 and 480 cm1 are due to the deformation
and bending modes of the Si O bond. After modication with
surfactant, the asymmetric stretching mode of Si O Si has
shifted from 1024 to 1036 cm1 and the deformation and bending mode of Si O Si band at 715, 616 and 480 cm1 have also
shifted to 726, 628 and 485 cm1 . The additional peaks at 1425
and 1467 cm1 in organoclay, which are absent in Na-B indicate the presence of C N vibration in tertiary amines. FTIR
results clearly indicate that surface modication of bentonite
clay by surfactants was achieved.

219

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

Counts

Organoclay

Transmitance (%)

RB - Organoclay

Na-B

CV - Organoclay

10

15

20

25

30

35

40

45
0

50

2Theta

Fig. 4 XRD patterns of Na-B and organoclay.

MB- Organoclay

O rga noclay

Na-Bentonite

4000 360 0 320 0 2800 2400 200 0 160 0 120 0 800

400

-1

Wave number (cm )

the aluminosilicate layer is maintained while layer spacing


The increase in the basal spacincreased from 15.2 to 24.1 A.
ing of Na-bentonite with HDTMA cations can be attributed to
replacement of the inorganic layer cations and their hydration water with HDTMA cations. On the basis of information
given by earlier works (Yilmez and Yaper, 2004; Lee et al., 2005)
regarding the orientation and the number of molecular layers
in the interlamellar spacing of clay minerals, and considering
molecular size of HDTMA, the increase in the basal spacing
points to a bilayer arrangement for HDTMA (Scheme 1).
The surface morphology of the Na-B and organoclay is
shown in Fig. 5. It can be obviously seen from both gures
that the bentonite has porous structure. After modication,
its form changed to an ultra ne thin corn ake like crystal
with uffy appearance. This uffy appearance probably occurs

Fig. 3 FTIR spectra of Na-B, organoclay, MB-organoclay,


CV-organoclay and RB-organoclay.
By comparing FTIR spectra of organoclay and MB adsorbed
organoclay, additional peaks at 3719 cm1 show NH2 group,
aliphatic and aromatic group, 2932 cm1 shows C H, aliphatic
group, and 1236 cm1 shows aromatic group of MB, 1371 cm1
indicates the presence of CH3 group. The peaks at 1350, 1326,
1246, and 1217 cm1 of MB, indicated the presence of C N
stretching vibration peak and presence of tertiary amine in
MB (Dhodapkar et al., 2006). Shift in peaks from 3716 and
3624 cm1 to 3754 and 3638 cm1 indicates the adsorption of
dyes on silanol group.
New peaks were observed after the adsorption of crystal
violet on the organoclay at 2067, 1589 and 1369 cm1 which
indicates C C in the aromatic ring and C N stretching in aromatic tertiary amine. In the case of Rhodamine B adsorbed
organoclay, weak peak at about 1453 cm1 and another peak
at about 2900 cm1 indicate C H stretching modes of aromatic rings of Rhodamine B (Pavia et al., 2009). 1341 cm1 peak
attributes to C-aryl bond vibration; the peak at 1720 cm1 is
due to C O groups. The peak at 1646 cm1 is caused by vibrations of the C N bond and the heterocycle vibrations cause
the peak ranging at 15301558 cm1 (Jing-yil et al., 2007).
The XRD of Na-B and organoclay is shown in Fig. 4. The
patterns of modied bentonite indicate the modication of
clay by surfactants. The characteristic d spacing of 4.48, 3.71,
conrms that the sample used in this work
3.62 and 1.51 A
further implies the 2:1 minis bentonite. The peak at 4.48 A
eral type. On treatment with HDTMA, the basic structure of

Fig. 5 SEM images of Na-B and organoclay.

220

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

120

120

100

100

Na-bentonite

80

MB
CV
RB

80

organoclay

60

60

MB

Initial concentration : 400 mol/L


pH
: 9.0
Equilibrium time
: 6h

40
20

40

Initial concentration :200 mol/L


T emperature : 30 0 C

Adsorption(%)

Adsorption (%)

20

100
80

Na-benton ite

60

organoclay
Initial concentration : 40 0 mol/L
pH
: 9.0
: 6h
Equilibrium time

40
20

MB
CV
RB

100
80

CV

60
Initial concentration :400 mol/L
T emperature : 30 0 C

40
20

100
0

80

Na-benton ite

60

organoclay
Initial con centration : 40 0 mol/L
pH
: 9.0
Equilibrium time
: 6h

40
20

10 0

20 0

30 0

RB

40 0

10

12

Fig. 7 Effect of pH on the adsorption of MB, CV and RB


onto organoclay.

50 0

Adsorbent dose (mg)


Fig. 6 Effect of surface modication for the adsorption of
MB, CV and RB onto organoclay.
due to the reduction in certain amorphous phase originally
associated with the bentonite.

3.2.
Effect of surface modication on the adsorption of
dyes onto organoclay
Batch experiments were carried out in order to evaluate the
effect of surface modication on the efciency of dye removal.
Fig. 6 shows the percentage removal of dye from aqueous
solutions as a function of adsorbent dose at a dye concentration of 400 mol/L. As expected the adsorption efciency
increased with the increase in the amount of adsorbent, however this did not occur in a steady progression. From the results
obtained it was observed that organoclay is 1.6, 1.7 and 1.75
times effective than Na-bentonite for the removal of MB, CV
and RB respectively. This is due to the fact that after modication, the surface area of the clay increases and its porosity
also increases. Increasing dose increases the partitioning of
per gram of adsorbent, which leads to an increase in adsorption. As a result more adsorbate molecules can be bound to
adsorbent surface through chemical bonding.

3.3.

pH

0
0

Effect of pH

To study the effect of pH on adsorption, experiments were


performed by using an initial dye concentration of 200 and
400 mol/L solutions of MB, RB and CV. For this batch study,

50 mL of each of dye solutions was treated with 0.1 g of adsorbent at different pH (2.011.0) in 100 mL stoppered bottles.
The results are shown in Fig. 7. From the results it was
observed that increase in pH increases the adsorption. As the
pH of the solution increases, the positive charge on the surface decreases and the number of negatively charged sites
increases (Baskaralingam, 2006). A negative charged surface
site on the clay favors the adsorption of cationic dye due to
electrostatic attraction. Earlier works reported that cationic
dye sorption increases with increase in pH (Gupta et al., 2004;
Singh et al., 2003). From Fig. 7, it was found that the adsorption
capacity of the organoclay for dyes at initial concentrations
of 200 and 400 mol/L was 99.0 and 97.0 mol/g for MB, 95.0
and 184.0 mol/g for CV and 90.0 and 174.0 mol/g for RB,
respectively at pH 9.0. The order of afnity based on the
amount of dye uptake is as follows: MB > CV > RB. The order
of afnity is due to the fact that different dyes will experience
different physical and electrostatic forces according to their
structure, molecular size and functional groups (Allen et al.,
1989). Basically, methylene blue and other cationic dyes produce molecular cations (C+ ) and reduced ions (CH+ ) (Kavitha
and Namasivayam, 2007). At lower pH range the decrease in
adsorption is due to the electrostatic repulsion between the
cationic dye species and the protonated adsorbent surface.
However, after the surface modication the adsorbent surface
becomes more positive and the pHzpc is found to be 7.8. But
the optimum removal was observed at pH 9.0. Above the pHzpc
surface is negative and adsorption occurring by electrostatic
attraction between negative surface of adsorbent and cationic
dye species (Janos et al., 2003).
The mechanism of this adsorption process is also
explained on the basis of adsorption/partition model (Zhu and
Chen, 2000). On the basis of this model, they have pointed
out that the organic fraction of the surfactant modied clays

221

with long alkyl chains behaves mechanistically as a partition medium (Zhu and Su, 2002; Rawajh and Nsour, 2006),
where clay mineral acts as the conventional adsorbent in the
uptake of organic compounds (Jang and Kamens, 1999; Grate
et al., 1995; Burg et al., 2000, 2002). Adsorption of organics onto
clay minerals may occur through ion exchange, protonation,
hydrogen bonding and iondipole reactions. Partition occurs
through the interaction of the organic solutes with organic
fraction of the organoclay. For organoclay, the HDTMA surfactant evidently modies interlayer spaces by close packing of
the alkyl chains to create hydrophobic regions. Partition of the
dyes into hydrophobic regions plays an important role in dye
uptake. Moreover, in clay-aqueous systems the potential of the
surface is determined by the activity of ions (e.g., H+ or pH),
which react with the mineral surface. For clay minerals the
potential determining ions are H+ and OH and complex ions
formed by bonding with H+ and OH .
At high pH the reaction is:
SiOH + OH SiO + H2 O

(4)

In the present study the maximum uptake is found to be at


pH 9.0. After this pH the dye uptake remains almost same. At
this high pH, solution in contact with the basal oxygen surface
of the tetrahedral sheet which contain excess hydroxyls. The
surface will then exhibit a cation exchange capacity.
Various researchers found various mechanisms for
removal of basic dyes from aqueous phase. Turabik (2008)
found that adsorption mechanism occurs partly by ion
exchange releasing exchangeable cations in the interlayer
and basal plane surfaces and partly via non-columbic interactions between an adsorbed cation and a neutralized site
for the removal of basic dye onto bentonite. Yener et al.
(2006) observed that plugging of pores took place as a result
of chemisorption and/or dimerization of the dye and thus
adsorption was decreased at temperatures over 30 C. Ozacar
and Sengil (2006) report that when bentonite is added to a MB
solution, a cation in the MB solution drives away a cation on
the surface of any negatively charged external surface of the
bentonite. The process continues until all the other cations
have been expelled. Up to that point all the MB are attached
to the bentonite surfaces. Then the MB ions replace the ions
of the interlayer. In the present study electrostatic attraction
between negatively charged silanol group and positively
charged dye molecule predominates the adsorption process.

200
150

MB

100

CV

Initial concentration : 400 mol/L


pH
: 9.0
: 30 0 C
T emperature

50

RB

0
50

100

150

200

250

300

Time (min)
Fig. 8 Effect of contact time on the adsorption of MB, CV
and RB onto organoclay.

the dye removal. When the concentration increases from


200 to 1000 mol/L, the amount of removal increases from
99.9 to 375.0 mol/g (99.975.0%) for MB, 94.0 to 325 mol/g
(91.065.0%) for CV and 88.0 to 275 mol/g (84.055.0%) for RB.
From this it was observed that dye uptake is highly dependent on concentrations of the adsorbate. The increase in dye
removal with an increase in dye concentration is a common
phenomenon obtained with a variety of adsorbents, such as
clinoptilolite and amberlite (Yener et al., 2006). Unburned carbon (Wang et al., 2005), granular kohlrabi peel (Gong et al.,
2007) used for the adsorption of dyes.

3.5.

Adsorption kinetics

In order to nd out the potential rate-controlling steps


involved in the process of adsorption of basic dyes onto organoclay, pseudo rst-order and pseudo second-order kinetic
models were used to t the experimental data at room temperature by using an initial concentration of 400 mol/L. The
results are shown in Fig. 10. The rate constant of the dye
removal was determined by using pseudo-rst and pseudosecond-order kinetic equations,
qt = qe (1 ek1 t )
qt =

Effect of initial concentration and contact time

From Fig. 8 it was observed that the uptake of dyes onto


organoclay increases rapidly at initial time intervals and
then decreases slowly until the equilibrium is reached. The
time required to reach equilibrium was found to be 4 h. The
amounts of dyes adsorbed at equilibrium are found to be
199.0 mol/g (99.5%), 181.0 mol/g (90.5%) and 166.0 mol/g
(83.0%) respectively for MB, CV and RB at an initial concentration of 400 mol/L at a pH of 9.0 at 30 C. The high efciency in
uptake at an initial step is due to the availability of adsorption
sites on the surface of adsorbent. After a rapid uptake, a transitional phase takes place in which the rate of removal was slow
and reaches a constant value. The effect of dye concentration
on the adsorption by organoclay was investigated by varying
the initial dye concentration between 200 and 1000 mol/L
at an initial pH of 9.0 at 30 C (Fig. 9). From the results it
was observed that the increase in concentration increases

250

(5)

k2 q2e t
1 + k2 qe t

(6)

400

Amount adsorbed ( mol/g)

3.4.

Amount adsorbed (mol/g)

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

350

Equilibrium Time : 4 h

300

Temperature
pH

250

: 30 C
: 9.0

200
150

MB

100

CV

50

RB

0
0

20 0

40 0

60 0

800

100 0

120 0

C0 (mol/L)
Fig. 9 Effect of initial concentration on the adsorption of
dyes onto organoclay.

222

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

250

500

A
200
150

T heoretical
Initial concentration : 400 mol/L
pH
: 9.0
: 30 0 C
T emperature

50

Langmuir

MB
CV
RB

200

Amount adsorbed ( mol/g)

B
150
Exp erimental

MB
CV
RB

T heoretical
Initial concentration : 400 mol/L
pH
: 9.0
T emperature
: 30 0 C

50

Freundlich
Redlich peterson

Equilibrium T ime : 4 h
T emperature
: 30 0 C
pH
: 9.0

100

200

100

Exp erimental

300

Exp erimental

100

Amount adsorbed (mol/g)

MB

400

CV

350
300
250

Experimental
Langmuir
Freundlich
Redlich peterson

200
150
100

Equilibrium T ime : 4 h
T emperature
: 30 0 C
pH
: 9.0

50

0
0

50

100

150

200

250

RB

300

300

250

Time (min)

200
Fig. 10 Plots for (A) pseudo-rst-order, (B)
pseudo-second-order model for the adsorption of MB, CV
and RB onto organoclay.

Experimental

150

Langmuir

100
50

where qt and qe represent the amount of dye adsorbed at


time t and equilibrium time, respectively, k1 and k2 are the
rate constants for pseudo-rst and second-order kinetic rates.
The rate constants are evaluated by using non-linear regression analysis and the results are represented in Table 2. From
the table it was observed that the rate constants values are
independent of initial concentration which indicates that the
adsorption process is through chemical sorption between
adsorbate and adsorbent surface. The positive charged dye
ions at higher pH were electrostatically attracted to the negative sites of adsorbent. The Chi2 (2 ) values indicate that the
process follows pseudo-second-order kinetics.

0
0

Table 2 Kinetics parameters for the adsorption of three


dyes onto organoclay.
Parameters
Pseudo-rst-order
k1
2
HYBRID

MB

CV

RB

1.46 101
19.11
6.19

2.69 102
44.23
7.06

2.49 102
48.17
4.57

Pseudo-second-order
1.43 102
k2
2

0.37
0.37
HYBRID

2.99 104
3.32
1.49

4.96 104
0.75
1.12

200

300

400

500

Fig. 11 Isotherm plots for the adsorption of dyes onto


organoclay.

Peterson isotherms (Fig. 11). The non-linear forms of equations are given as,

Adsorption isotherm

The equilibrium adsorption isotherm is of importance to


establish the most appropriate correlation for the equilibrium
data for adsorption batch system. Adsorption studies were
performed by using three dye concentrations ranging from 200
to 1000 mol/L at 30 C and at a pH of 9.0. Several isotherm
equations are available and three important isotherms are
selected in this study, the Langmuir, Freundlich and Redlich

100

Ce (mol/L)

qe =

3.6.

Freundlich
Redlich peterson

Equilibrium T ime : 4 h
: 30 0 C
T emperature
: 9.0
pH

Q 0 bCe
1 + bCe

(7)

1/n

qe = KF Ce
qe =

(8)

KR Ce

1 + aR Ce

(9)

where Ce and qe are the equilibrium concentration and amount


adsorbed at equilibrium respectively. Q0 and b are Langmuir
constants related to adsorption capacity and sorption binding constant respectively. KF and 1/n are Freundlich constants
related to adsorption capacity and intensity of adsorption
respectively. To determine the parameters in Redlich Peterson equation, a range of KR values was chosen until a value of
KR was obtained which maximized the correlation coefcient
values. The tness of the adsorption isotherm is evaluated by
HYBRID error analysis.

3.6.1.

Error analysis

In order to check the best t isotherm model to the experimental data, the hybrid fractional error analysis (HYBRID) was
chosen.

223

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

400

Isotherm constants

MB

CV

CV adsorbed (mol/g)

Table 3 Isotherm parameters and HYBRID error


analysis coefcient for the adsorption of MB, CV and RB
onto organoclay.
RB

Langmuir
Q0 (mol/g)
b (L/mol)
2
HYBRID

399.74
0.04
8.75
0.86

365.11
0.02
7.68
5.72

324.36
0.01
1.08
1.31

Freundlich
KF
1/n
2
HYBRID

81.76
0.28
0.44
0.29

35.59
0.39
0.73
0.63

18.64
0.46
0.48
1.20

350
Initial CV concentration

300

200
800

250
200
150
100

0.98
0.04
0.84
0.47

0.99
0.023
0.94
1.34

MB adsorbed (mol/g)

50

RedlichPeterson
b
aR
2
HYBRID

0.86
0.02
2.4
1.43

HYBRID is given as,

100  (qe,exp qe,cal )


HYBRID =
np
qe,exp
n

i=1

600
1000

600
500
400
300

Initial CV concentration

0
600
1000

200

100

200
800

(10)

100

200

300

400

500

600

Equ ilibrium conc: MB (mol/L)

Table 3 gives the isotherm parameters along with the error


function. From the table it was observed that the Freundlich
isotherm model gives a best t for the experimental data for
all the three dyes. This results indicate that the adsorption of
dye molecules onto non-energetically equivalent sites (Vinod
and Anirudhan, 2003) of the organoclay.

Fig. 12 (A) Adsorption isotherms of MB in the presence of


CV added and (B) the amount of CV adsorbed in the
presence of MB onto organoclay.

3.7.
Comparison of adsorption capacity with other
available adsorbents

(qe )i = KFi Cei

Adsorption capacity of organoclay correlates with other available adsorbents along with equilibrium time and is listed in
Table 4. From the comparison it was clear that the adsorbent
developed is much more effective for the removal of basic dyes
from aqueous phase.

where (qe )i is the amount of solute i adsorbed per unit weight


of adsorbent in the presence of solute j. KFi is the single component Freundlich constant for solute i. 1/n is the Freundlich
exponential term for solute i. Cei and Cej are the equilibrium
concentrations of solute i and j respectively and aij is the competitive coefcient.
For binary component system the SRS equation can be written as,

3.8.

Binary component adsorption isotherm

Fig. 12 shows the adsorption of MB in the presence of different


concentrations of CV at room temperature. It was found that
the adsorption of MB is decreased in the presence of increasing
concentration of CV. The similar trend was seen in the case
of adsorption of CV. Isotherm characteristics of RB adsorbed
in the presence of different concentrations of CV and that of
RB adsorbed in the presence of different concentrations of MB
were also studied (isotherm plots are not shown). In these two
cases also the adsorption isotherm plots are similar to that of
Fig. 12. Adsorption of RB is reduced in the presence of CV and
that of MB reduced in the presence of RB. From the studies
conducted it was observed that the competitive effect of MB
on RB is less than that of RB on MB.

basis that there is an exponential distribution of adsorption


energies available for each solute. The SRS equation can be
written as
j



aij Cej

[(1/ni )1]

(qe )i = KFi Cei (Cei + aij Cej )

[(1/ni )1]

SheindrofRebhunSheintuch (SRS) (Sheindrof et al., 1981)


equation was used to study the binary component isotherm.
This equation is a multicomponent Freundlich type on the

(12)

In order to compute the competitive adsorption isotherm to


measure any one of the solute adsorbed in the presence of
other, the linearized form of SRS equation was used,
Cei
=
Cej
Cej
Cei




Bi
Cej
Bj
Cei


aij

(13)

aji

(14)

n /(n 1)

3.9.
The extended Freundlich isotherm for binary
components and competitive coefcients

(11)

where Bi = [KFi (Cei /qei )] i i


and aij is the competitive coefcient obtained from the Y intercept of the plot Cei /Cej vs Bi /Cej .
The six-binary solute systems are tested with SRS equation.
The values of competitive coefcients are listed in Table 5. The
competitive coefcient for MB on RB was found to be 1.101,
which indicates the linearity of SRS equation. Fig. 13 shows
the SRS plot for RB-MB system. The correlation coefcients

224

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

Table 4 Comparison of adsorption capacity of organoclay with other available adsorbents.


Adsorbent

Adsorption capacity

Banana peel
Fe(III)/Cr(III) hydroxide
Clay
Fly ash (treated with H2 SO4 )
Sugarcane dust
Activated sludge
Coir pith carbonized
Blast furnace sludge
Carbonaceous adsorbent
Surfactant modied bentonite

20.8 mg/g
22.8 mg/g
6.3 mg/g
0.0021 mmol/g
3.798 mg/g
113.2 mg/g
2.56 mg/g
25 mg/g
161 mg/g
399.4 mol/g

Equilibrium time (min)


24 h

1h
72 h
30 min
6h
10 min
120 min
120 min
4h

Reference
Annadurai et al. (2002)
Namasivayam and Sumithra (2005)
Gurses et al. (2004)
Lin et al. (2008)
Khattri and Singh (1999)
Chu and Chen (2002)
Namasivayam et al. (2001)
Jain et al. (2003)
Jain et al. (2003a)
Present study

Table 5 Competitive coefcients for the binary-solute


systems, derived from SRS equation.
Dye

Competitive coefcients
aij

aji

MB + CV
n
R2

0.103
64
0.985

0.036
64
0.988

RB + CV
n
R2

0.253
64
0.992

0.534
64
0.985

MB + RB
n
R2

0.047
64
0.992

1.101
64
0.995

Fig. 14 Adsorption/desorption cycles.


were found to be in the range of 0.9850.994, which indicates
the applicability of SRS equation to describe the competitive
adsorption behavior of binary mixture.

3.10.

Desorption and regeneration study

Desorption and regeneration studies were performed by using


spent adsorbent. Different reagents were tried for this pur-

C (RB/MB)

12
10

4.

8
6
4
2

0
-2 0

10

B (RB/MB)
5

C (MB/RB)

4
3
2

1
0
-1

pose and complete desorption was achieved by 0.1 M HNO3 .


Repeated adsorption/desorption cycles were performed to
examine the reusability and metal recovery efciency of the
adsorbent. Each cycle consisted of loading with an aqueous
metal solution (C0 -75 mol/L) and elution of the bound metals
with 0.1 M HNO3 . The results are depicted in Fig. 14. During
the adsorption/desorption process, the percentage of removal
was little decreased, which indicates that the mechanism of
adsorption is formation of complexation between dyes and
silanol molecule on the clay surface.

0.5

1.5

Conclusions

Organoclay was used to remove basic dyes from aqueous


phase. Hydrophobic nature of the adsorbent facilitates the
adsorption process. The dye adsorption was also inuenced by
solution pH and concentration. Higher pH will generally result
in higher adsorption for MB and CV and RB. The adsorption
kinetics can be well described by the pseudo-second-order
model equation. Adsorption isotherm can be tted by Langmuir, Freundlich, and RedlichPeterson models, in which the
Freundlich model shows the better ones. The kinetic and
isotherm models were also tted by calculating HYBRID error
analysis. Extended Freundlich adsorption isotherm for binary
mixture was performed in order to calculate the competing
effect of each dye on organoclay. It was found that the organoclay developed can effectively remove basic dyes, and in their
competition also behaves as a good adsorbent.

B (M B/RB)

Fig. 13 SRS competitive isotherm plots for (A) MB


adsorption in the presence of RB and (B) RB adsorption in
the presence of MB.

Acknowledgement
The authors are grateful to the Professor and Head, Department of Chemistry, University of Kerala, Trivandrum, for
providing laboratory facilities for this work.

Process Safety and Environmental Protection 9 5 ( 2 0 1 5 ) 215225

References
Allen, S.J., McKay, G., Khader, K.Y.H., 1989. Equilibrium adsorption
isotherms for basic dyes onto lignite. J. Chem. Technol.
Biotechnol. 45, 291302.
Akcay, M., 2004. Characterization and determination of the
thermodynamic and kinetic properties of p-CP adsorption
onto organophilic bentonite from aqueous solution. J. Colloid
Interface Sci. 280, 299304.
Ashly, S., Prasad, L., Manonmani, S., 2014. A comparative study of
microwave and chemically treated Acacia nilotica leaf as an eco
friendly adsorbent for the removal of Rhodamine B dye from
aqueous solution. Arab. J. Chem. 7, 494503.
Bartelt-Hunt, S.L., Burns, S.E., Smith, J.A., 2003. Nonionic organic
solute sorption onto two organobentonites as a function of
organic-carbon content. J. Colloid Interface Sci. 266, 251258.
Baskaralingam, P., Pulikesi, M., Elango, D., Ramamurthi, V.,
Sivanesan, S., 2006. Adsorption of acid dye onto
organobentonite. J. Hazard. Mater. B 128, 138144.
Beyene, H.D., 2014. The potential of dyes removal from textile
wastewater by using different treatment technology. A review.
Int. J. Environ. Monit. Anal. 2 (6), 347353.
Burg, P., Fydrych, P., Bimer, J., Abraham, M.H., Matt, M., Gruber, R.,
2000. The characterization of an active carbon in terms of
selectivity towards volatile organic compounds using an LSER
approach. Fuel 79, 10411045.
Burg, P., Fydrych, P., Bimer, J., Salbut, P.D., Jankowska, A., 2002.
Comparison of three active carbons using LSER modeling:
prediction of their selectivity towards pairs of volatile organic
compounds (VOCs). Carbon 40, 7380.
Dhodapkar, J.R., Rao, N.N., Pande, S.P., Kaul, S.N., 2006. Removal of
basic dyes from aqueous medium using a novel polymer.
Bioresour. Technol. 97, 877885.
Dogan, M., Alkan, M., 2003. Adsorption kinetics of methyl violet
onto perlite. Chemosphere 50, 517528.
Gong, R., Zhang, X., Liu, H., Sun, Y., Liu, B., 2007. Uptake of
cationic dyes from aqueous solution by biosorption onto
granular kohlrabi peels. Bioresour. Technol. 98, 13191323.
Grate, J.W., Abraham, M.H., Du, C.M., McGill, R.A., Shuely, W.J.,
1995. Examination of vapor sorption by fullerene,
fullerene-coated surface acoustic wave sensors, graphite, and
low-polarity polymers using linear solvation energy
relationships. Langmuir 11, 21252130.
Gregg, S., Singh, K.S.W., 1982. Adsorption. In: Surface Area and
Porosity. Academic Press, New York.
Gupta, V.K., Suhas, A.I., Saini, V.K., 2004. Removal of Rhodamine
B, fast green and methylene blue from wastewater using red
mud an aluminium industry waste. Ind. Eng. Chem. Res. 43,
17401747.
Jang, M., Kamens, R.M., 1999. Adsorptive partitioning of
semi-volatile organic compounds on ne atmospheric
inorganic dust particles. Environ. Sci. Technol. 33, 18251831.
Janos, P., Buchtova, H., Ryznarova, M., 2003. Sorption of dyes from
aqueous solutions onto y ash. Water Res. 37, 49384944.
Jing-yil, L.I., Wan-hong, M.A., Peng xiang, L.E.I., Jin cai, Z., 2007.
Detection of intermediates in the TiO2 -assisted
photodegradation of Rhodamine B under visible light
irradiation. J. Environ. Sci. 19, 892896.
Kavitha, D., Namasivayam, C., 2007. Experimental and kinetic
studies on methylene blue adsorption by coir pith carbon.
Bioresour. Technol. 98, 1421.
Lee, J.J., Choi, J., Park, J.W., 2002. Simultaneous sorption of lead
and chlorobenzene by organobentonite. Chemosphere 49,
13091315.
Lee, S.Y., Cho, W.J., Hahn, P.S., Lee, M., Lee, Y.B., Kim, K.J., 2005.
Microstructural changes of reference montmorillonites by
cationic surfactants. Appl. Clay Sci. 30, 174180.
Malik, P.K., 2003. Use of activated carbons prepared from sawdust
and ricehusk for adsorption of acid dyes: a case study of Acid
Yellow 36. Dyes Pigm. 56, 239249.
Natarajan, T.S., Bajaj, H.C., Tayade, R.J., 2014. Preferential
adsorption behavior of methylene blue dye onto surface

225

hydroxyl group enriched TiO2 nanotube and its photocatalytic


regeneration. J. Colloid Interface Sci. 433, 104114.
Ozcan, A.S., Erdem, B., Ozcan, A., 2004. Adsorption of acid blue
193 from aqueous solutions onto Na-bentonite and
BTMA-bentonite. J. Colloid Interface Sci. 280, 4454.
zacar, M., Sengil, I.A., 2003. Adsorption of reactive dyes on
calcined alunite from aqueous solutions. J. Hazard. Mater. 98,
211224.
Ozacar, M., Sengil, I.A., 2006. A two stage batch adsorber design
for methylene blue removal to minimize contact time. J.
Environ. Manage. 80, 372379.
ztrk, A., Malkoc, E., 2014. Adsorptive potential of cationic Basic
Yellow 2 (BY2) dye onto natural untreated clay (NUC) from
aqueous phase: mass transfer analysis, kinetic and
equilibrium prole. Appl. Surf. Sci. 299, 105115.
Pavia, D.L., Lampman, G.M., Kriz, G.S., Vyvyan, J.R., 2009.
Introduction to Spectroscopy, 4th ed. Brooks/Cole, Belmont,
CA.
Rammel, R.S., Zatiti, S.A., El Jamal, M.M., 2011. Biosorption of
crystal violet by Chaetophora elegans alga. J. Chem. Technol.
Metall. 46, 283292.
Rawajh, Z., Nsour, N., 2006. Characteristics of phenol and
chlorinated phenols sorption onto surfactant-modied
bentonite. J. Colloid Interface Sci. 298, 3949.
Robinson, T., Chandran, B., Nigam, P., 2002. Effect of
pretreatments of three waste residues, wheat straw, corncobs
and barley husks on dye adsorption. Bioresour. Technol. 85,
119124.
Sheindrof, C., Rebhun, M., Sheintuch, M., 1981. A Freundlich type
multicomponent isotherm. J. Colloid Interface Sci. 79,
136142.
Singh, K.P., Mohan, D., Sinha, S., Tondon, G.S., Gosh, D., 2003.
Color removal from wastewater using low-cost activated
carbon derived from agricultural waste material. Ind. Eng.
Chem. Res. 42, 19651976.
Srivastava, P., Goyal, S., Tayade, R., 2013. Ultrasound-assisted
adsorption of Reactive Blue 21 dye on TiO2 in the presence of
some rare earths (La, Ce Pr & Gd). Can. J. Chem. Eng. 92, 4151.
Turabik, M., 2008. Adsorption of basic dyes from single and binary
component systems onto bentonite: simultaneous analysis of
Basic Red 46 and Basic Yellow 28 by rst order derivative
spectrophotometric analysis method. J. Hazard. Mater. 158,
5264.
Vinod, V.P., Anirudhan, T.S., 2003. Adsorption behaviour of basic
dyes on the humic acid immobilized pillared clay. Water Air
Soil Pollut. 150, 193217.
Wang, S., Li, L., Wu, H., Zhua, Z.H., 2005. Unburned carbon as a
low-cost adsorbent for treatment of methylene
blue-containing wastewater. J. Colloid Interface Sci. 292,
336343.
Wang, Y., Mu., Y., Zhao, Q.B., Yu, H.Q., 2006. Isotherms, kinetics
and thermodynamics of dye biosorption by anaerobic sludge.
Sep. Purif. Technol. 50, 17.
Wang, Z., Ding, L., Zou, B., Gao, W., Liu, Q., Guo, Y., Wang, X., Liu,
Y., 2014. Adsorption of Rhodamine-B from aqueous solution
using treated rice husk-based activated carbon. Colloids Surf.
A: Phys. Eng. Aspects 446, 17.
Yener, J., Kopac, T., Dogu, G., Dogu, T., 2006. Adsorption of Basic
Yellow 28 from aqueous solutions with clinoptilolite and
amberlite. J. Colloid Interface Sci. 294, 255264.
Yilmez, N., Yaper, S., 2004. Adsorption properties of tetradecyland hexadecyltrimethylammonium bentonites. Appl. Clay
Sci. 27, 223228.
Zhang, Z., OHara, I.M., Kent, G.A., Doherty, W.O.S., 2013.
Comparative study on adsorption of two cationic dyes by
milled sugarcane bagasse. Ind. Crops Prod. 42, 4149.
Zhu, L., Chen, B., 2000. Sorption behavior of p-nitrophenol on the
interface between anioncation organobentonite and water.
Environ. Sci. Technol. 34, 29973002.
Zhu, L., Su, Y., 2002. Benzene vapor sorption by organobentonites
from ambient air. Clays Clay Miner. 50, 421427.

Você também pode gostar