Sie sind auf Seite 1von 9

Desalination 261 (2010) 52–60

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / d e s a l

Adsorption of dye from aqueous solution by cashew nut shell: Studies on equilibrium
isotherm, kinetics and thermodynamics of interactions
P. Senthil Kumar a,⁎, S. Ramalingam b, C. Senthamarai a, M. Niranjanaa a, P. Vijayalakshmi c, S. Sivanesan c
a
Department of Chemical Engineering, SSN College of Engineering, Chennai, 603 110, India
b
Department of Chemical Engineering, University of Louisiana at Lafayette, LA 70504, United States
c
Environmental Management Laboratory, AC Tech, Anna University-Chennai, 600 025, India

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

Article history: Cashew nut shell (CNS) — a novel, low cost adsorbent prepared from agricultural waste has been utilized as
Received 27 March 2010 the adsorbent for the removal of Congo red (CR) dye from an aqueous solution. The effect of pH, adsorbent
Received in revised form 13 May 2010 dose, initial dye concentration, time and temperature on adsorption was studied. The results indicate that
Accepted 17 May 2010
CNS can be employed as a low cost alternative compared to other commercial adsorbents in the removal of
Available online 9 June 2010
dyes from wastewater. The experimental data were analyzed by Langmuir, Freundlich, Redlich–Peterson,
Koble–Corrigan, Sips, Toth, Temkin and Dubinin–Radushkevich adsorption isotherms. The characteristic
Keywords:
Adsorption
parameters for each isotherm and related correlation coefficients have been determined using MATLAB 7.1.
Cashew nut shell Thermodynamic parameters such as ΔGo, ΔHo and ΔSo have also been evaluated and it has been found that
Congo red dye the sorption process was feasible, spontaneous and exothermic in nature. Pseudo-first-order, pseudo-
Isotherm second-order and intraparticle diffusion models were used to fit the experimental data. Kinetic parameters,
Kinetics rate constants, equilibrium sorption capacities and related correlation coefficients, for each kinetic model
Thermodynamic were calculated and discussed. It was shown that the adsorption of CR could be described by the pseudo-
second-order equation, suggesting that the adsorption process is a presumably chemisorption.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction quite popular due to its simplicity and high efficiency, as well as the
availability of a wide range of adsorbents. Activated carbon is the
Wastewater effluents from different industries such as textiles, most popular adsorbent for removal of dyestuffs from wastewater
leather, rubber, paper and plastics, contain several kinds of synthetic [10–18]. However, adsorbent grade carbon is cost prohibitive and
dye stuffs [1]. A very small amount of dye in water is highly visible. both regeneration and disposal of the used carbon often very
Further, discharging even a small amount of dye into water can difficult. Therefore, there is a growing need to find locally available,
affect aquatic life and food webs due to the carcinogenic and low cost and effective materials for the removal of dyes. A number of
mutagenic effects of synthetic dyes [2]. Synthetic dyes are difficult non-conventional, low cost adsorbents such as wood [19], bagasse
to biodegrade due to their complex aromatic structures, which pith [20], maize cob [21], rice hull ash [22], Azadirachta indica leaf
provide them physico-chemical, thermal and optical stability [3], powder [23], palm kernel fibre [24], fungi [25], bentonite [26],
thus bringing some difficulties for the treatment of these dyes. Organo-attapulgite [27] have been used for the removal of dye from
Hence, it is imperative that a suitable treatment method should be aqueous solutions. However, some of these adsorbents do not have
devised. good adsorption capacities for anionic dyes because most have
In recent years, many methods including coagulation and hydrophobic or anionic surfaces. Hence, there is a need to search for
flocculation [4], reverse osmosis [5], chemical oxidation [6], more effective adsorbents. In the present study, CNS has been used
biological treatments [7], photodegradation [8], and adsorption as an adsorbent for the removal of CR from its aqueous solutions.
[9], have been developed for treating dye containing wastewater. The objective of the present work is to introduce a new low cost
Among various treatment technologies, adsorption technique is adsorbent such as CNS and also to examine its effectiveness in
removing CR from aqueous solution. The influence of experimental
parameters such as pH, contact time, temperature, CNS dosage and
initial CR concentrations were studied. Also the sorption of CR at
solid–liquid interfaces has been studied extensively under equilib-
⁎ Corresponding author. Tel.: +91 44 32909855, +91 9884823425; fax: +91 44
27475063.
rium and various thermodynamic conditions. Further the kinetic
E-mail addresses: senthilkumarp@ssn.edu.in, senthilchem8582@gmail.com and the mechanistic steps involved in the sorption process were
(P. Senthil Kumar). evaluated at different initial CR dye concentrations.

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.05.032
P. Senthil Kumar et al. / Desalination 261 (2010) 52–60 53

2. Experimental

2.1. Adsorbent

Cashew nut shells were collected from Pudukkottai District,


Tamilnadu, India and were used as an adsorbent. This natural waste
was thoroughly rinsed with water to remove dust and soluble
material and was then allowed to dry at room temperature. The
Fig. 1. Chemical structure of Congo red.
above dried natural waste was crushed into powder and sieved
through the sieves. 200–30 mesh size particles of CNS were taken for
the adsorption studies. The proximate and ultimate analysis of cashew added to 100 mL of CR solution in 250 mL flasks. The flasks were
nut shells are shown in Table 1. shaken at 120 rpm for 120 min. The initial CR concentration used in
this study was 20 to 100 mg/L. The equilibrium study was performed
2.2. Adsorbate using various concentrations of CR solutions i.e. from 20 to 100 mg/L.
A 20 g/L CNS with 100 mL CR solutions of various initial concentra-
Congo red (CI = 22,120) was supplied by Merck (India). A stock tions was shaken at 120 rpm for 120 min at 30 °C. All the experiments
solution of CR dye was prepared (100 mg/L) by dissolving a required were performed in duplicates. The amount of dye adsorbed onto CNS,
amount of dye powder in deionized water. The stock solution was qe (mg/g), was calculated by the following mass balance relationship:
diluted with deionized water to obtain desired concentration ranging
from 20 to 100 mg/L. The chemical structure of CR is shown in the ðCi −Ce ÞV
qe = ð1Þ
Fig. 1. W

2.3. Analysis where Ci and Ce are the initial and equilibrium concentrations (mg/L)
of CR dye solution respectively; V is the volume (L); and W is the mass
The concentration of CR in the experimental solution was (g) of the adsorbent.
determined from the calibration curve prepared by measuring
absorbance of different determined concentrations of CR solutions 3. Results and discussion
at λmax 497 nm using a UV–vis Spectrophotometer (Shimadzu, Japan).
The pH of solution was measured with a Hanna pH meter using a 3.1. Characterization of CNS
combined glass electrode (Model HI 9025C, Singapore).
The adsorption capacity of CNS depends upon porosity as well as
2.4. Adsorption experiment chemical reactivity of functional groups at the surface. This reactivity
creates an imbalance between forces at the surface when compared to
Adsorption experiments were conducted by varying pH, contact those within the body, thus leading to molecular adsorption by the
time, CNS dose, temperature, initial CR concentration under the van der Waals force. Knowledge on surface functional groups would
aspects of thermodynamic study, adsorption isotherms and adsorp- give insight to the adsorption capability of the CNS. The FT-IR
tion kinetics. The experiments were carried out in 250 mL Erlenmeyer spectrums of the CNS and CR loaded on CNS are shown in Fig. 2. OH
flasks and the total volume of the reaction mixture was kept at and NH stretching is between 3100 and 3500 cm−1, C–H aromatic is
100 mL. The pH of solution was maintained at a desired value by between 3000 and 3100 cm−1, and C–H aliphatic is between 2800 and
adding 0.1 M NaOH or HCl. The flasks were shaken for the required 3000 cm−1. The spectrum shows a broad band near 3399 cm−1 which
time period in a water bath shaker. The kinetic studies was carried out indicates the presence of hydroxyl groups on the CNS surface. The
by agitating 250 mL flasks containing 20 g/L of CNS and 100 mL CR dye stretching was attributed to the absorbed water on the surface of CNS.
solutions of different concentrations i.e. from 20 to 100 mg/L, in water The peak at 2925 cm−1 is due to C–H stretching of CH2 groups. The
bath shaker. The mixture was agitated at 120 rpm at 30 °C. The stretching frequencies of the aromatic C = C and aromatic C–H groups
contact time was varied from 0 to 120 min. At predetermined time, give rise to peaks at 3011 and 2854 cm−1 respectively. The bands near
the flasks were withdrawn from the shaker and the residual dye 1630 cm−1 indicates fingerprint region of C = O, C–O and O–H
concentration in the reaction mixture was analyzed by centrifuging groups that exist as functional groups of CNS. The peaks at 1542 and
the reaction mixture and then measuring the absorbance of the 1515 cm−1 is assigned to a conjugated hydrogen bonded carboxyl
supernatant at the wavelength that correspond to the maximum group. The peak at 1454 cm−1 (vC–O) indicates the presence of
absorbance of the sample. Dye concentration in the reaction mixture carboxylic groups. The peak at 1374 cm−1 indicates the presence of C–
was calculated from the calibration curve. The λmax values of the
wastewater samples varied ±10 nm from the λmax values of pure
dyes at a fixed pH. For thermodynamic study, the experiments were
performed by varying temperature from 30 to 60 °C using 20 g/L CNS

Table 1
Properties of the cashew nut shells.

Proximate analysis Ultimate analysis Ash chemical composition


(wt.%) (wt.%) (wt.%)

Volatile matter 65.21 Carbon 45.21 Silica 64.53


Moisture 9.83 Hydrogen 4.25 Iron oxide 3.27
Ash 2.75 Oxygen 37.75 Aluminium oxide 2.19
Fixed carbon 22.21 Nitrogen 0.21 Calcium oxide 26.89
Sulphur Nil Magnesium oxide 2.49
Moisture 9.83 Sodium oxide 0.63
Ash 2.75
Fig. 2. FT-IR spectrum of cashew nut shell and CR loaded on cashew nut shell.
54 P. Senthil Kumar et al. / Desalination 261 (2010) 52–60

H aliphatic bending. The peak at 1232 cm−1 indicates the presence of


C–N from amine. The two peaks at 1156 cm−1 (P = O) and 1035 cm−1
(P–OH) was characteristic of PO2 stretching. These results agree with
the surface chemistries of other agricultural by-products, such as
coconut coir pith [28] and rubber wood saw dust [29]. The region
between 700 and 900 cm−1 contains various bands related to
aromatic, out of plane C–H bending with different degrees of
substitution [30]. After adsorption of CR, peaks of –NH and –NH2
disappeared, –NH rocking peak shifted to a higher wave number,
indicating the significant role of amine groups in the adsorption
process. The stretching vibration of the second –OH group shifted to a
lower wave number. But vibration for the first –OH group did not
change much. It can be concluded from the FT-IR spectroscopy that
amine and the second –OH group were included in the adsorption of
CR. Fig. 4. Effect of CNS dose on CR removal (the initial dye concentration = 50 mg/L and
time = 2 h).
The specific surface area and pore structure of the CNS was
determine by using surface area and pore size analyzer (Quanta-
chrome, Autosorb-I) on nitrogen adsorption at 77 K. The specific dose. The % CR removal increases from 56.33 to 99.34% for an increase
surface area was calculated by BET equation [31]. It was found that the in adsorbent dose from 5 to 30 g/L. At higher CNS to CR concentration
BET surface area, pore volume, average pore diameter and bulk ratios, there is very fast superficial sorption onto the CNS surface that
density of the CNS were 395 m2/g, 0.4732 cm3/g, 5.89 nm and 0.415 g/ produces a lower CR concentration in the solution than when the
cm3, respectively. biomass to CR concentration ratio is lower. This is because a fixed dose
of CNS can only adsorb a certain amount of dye. Therefore, the more
3.2. Effect of pH the adsorbent dosage, the larger the volume of effluent can be
purifying with a fixed dosage of CNS. The decrease in amount of CR
The pH of the system exerts profound influence on the adsorptive adsorbed on to an adsorbent qe (mg/g) with increasing CNS dose is
uptake of CR molecule presumably due to its influence on the surface due to a split in the flux or the concentration gradient between CR
properties of the CNS and ionization/dissociation of the CR molecule. concentration in the solution and the CR concentration in the surface
The variations in removal of CR dye from aqueous solution at various of the CNS. Thus with increasing CNS dose, the amount of dye
system pH are shown in Fig. 3. From the figure, it is evident that the adsorbed onto unit weight of CNS gets reduced, thus causing a
maximum removal of dye is observed at pH 3 and below. Low pH decreasing in qe value with increasing CNS dose (not shown in the
leads to an increase in H+ ion concentration in the system and the figure).
surface of the CNS acquires positive charge by absorbing H+ ions. As
the CNS surface is positively charged at low pH, a significantly strong
electrostatic attraction appears between the positively charged CNS 3.4. Effect of initial dye concentration
surface and anionic dye molecule leading to maximum adsorption of
dye. As the pH of the system increases, the number of negatively The effect of initial dye concentration in the range of 20 to 100 mg/
charged sites increases and the number of positively charged sites L on adsorption was investigated and is shown in Fig. 5. It is evident
decreases. A negatively charged surface site on the CNS does not favor from the figure that the percentage CR removal decreased with the
the adsorption of anionic dye molecules due to the electrostatic increase in initial concentration of CR. The initial dye concentration
repulsion. Furthermore, lower adsorption of the CR dye in alkaline provides the necessary driving force to overcome the resistance to the
medium is also due the competition from excess OH− ions with the mass transfer of CR between aqueous phase and the solid phase. The
anionic dye molecule for the adsorption sites. increase in initial dye concentration also enhances the interaction
between CR and CNS. Therefore, an increase in initial concentration of
3.3. Effect of adsorbent dose CR enhances the adsorption uptake of CR. This is due to increase in the
driving force of the concentration gradient, as an increase in the initial
Fig. 4 shows the plot between the % CR removals against adsorbent dye concentration. While the percentage CR removal was found to be
dose. From figure it was noted that the % CR removal is varied with 98.52% for 20 mg/L of initial concentration, this value was 85.44% for
varying sorbent mass and it increased with increase in adsorbent that of 100 mg/L.

Fig. 3. Effect of pH for the adsorption of CR onto CNS (the initial dye concentra- Fig. 5. Effect of initial dye concentration for the adsorption of CR onto CNS (the initial
tion = 50 mg/L, CNS dose = 20 g/L and time = 2 h). dye concentration = 20 to 100 mg/L, CNS dose = 20 g/L and time = 2 h).
P. Senthil Kumar et al. / Desalination 261 (2010) 52–60 55

3.5. Effect of contact time

Effect of contact time for the removal of CR by the CNS at Co = 20


to 100 mg/L for the adsorbent dose of 20 g/L showed rapid adsorption
of dye in the first 60 min and, thereafter, the adsorption rate
decreased gradually and the adsorption reached equilibrium in
about 90 min as shown in Fig. 6. Increase in contact time up to
120 min showed that the CR removal by CNS was only by about 0.2%
over those obtained for 90 min contact time. Aggregation of dye
molecules with the increase in contact time makes it almost
impossible to diffuse deeper into the adsorbent structure at highest
energy sites. This aggregation negates the influence of contact time as
the micropores get filled up and start offering resistance to diffusion of
aggregated dye molecules in the adsorbents. This is the reason why an Fig. 7. Effect of temperature on CR removal by CNS (the initial dye concentration = 20
insignificant enhancement in adsorption is effected in 120 min as to 100 mg/L, CNS dose = 20 g/L and time = 2 h).
compared to that in 90 min. Since the difference in the adsorption
values at 90 min and at 120 min is very small, after 90 min contact, a
steady-state approximation was assumed and a quasi-equilibrium
situation was accepted. Further experiments were conducted for ΔSo ΔH o
90 min contact time only. The adsorption curves were single, smooth log Kc = − ð5Þ
2:303R 2:303RT
and continuous leading to saturation and indicated the possible
monolayer coverage on the surface of adsorbents by the dye
molecules [32,33]. where Kc is the equilibrium constant, Ce is the equilibrium concentration
in solution (mg/L) and CAe is the amount of CR adsorbed on the
3.6. Effect of temperature adsorbent per liter of solution at equilibrium (mg/L). ΔGo, ΔHo and ΔSo
are changes in Gibbs free energy (kJ/mol), enthalpy (kJ/mol) and
The adsorption of CR on CNS was investigated as a function of entropy (J/mol/K), respectively. R is the gas constant (8.314 J/mol/K), T
temperature and maximum removal of CR was obtained at 30 °C. is the temperature (K). The values of ΔHo and ΔSo are determined from
Experiments were performed at different temperatures of 30, 40, 50 the slope and the intercept of the plots of logKc versus 1/T (Fig. 8). The
and 60 °C for the initial CR concentrations of 20–100 mg/L at constant ΔGo values were calculated using Eq. (3). Adsorption of CR on CNS
adsorbent dose of 20 g/L. The adsorption decreased from 98.52 to decreased when the temperature was increased from 303 to 333 K is
88.11%, 94.71 to 84.96, 91.61 to 82.79, 88.64 to 79.28 and 85.44 to shown in Fig. 8. The process was thus exothermic in nature. The plots
76.82% for the initial CR concentrations of 20, 40, 60, 80 and 100 mg/L were used to compute the values of thermodynamic parameters
respectively with the rise in temperature from 30 to 60 °C (Fig. 7). This (Table 2). The value of enthalpy change (ΔHo) and the entropy change
is mainly due to the decreased surface activity suggesting that (ΔSo) recorded from this work were presented in the Table 2. The
adsorption between CR and CNS was an exothermic process. negative ΔGo value indicates the process be feasible and spontaneous
nature of adsorption; negative ΔHo value suggests the exothermic
3.7. Thermodynamic study nature of adsorption and the ΔSo can be used to describe the
randomness at the CNS-solution interface during the sorption.
Thermodynamic parameters such as free energy (ΔGo), enthalpy
(ΔHo) and entropy (ΔSo) change of adsorption can be evaluated from 3.8. Adsorption equilibrium study
the following equations ((2), (3), (4)):
The capacity of the adsorption isotherm is fundamental, and plays
C an important role in the determination of the maximum capacity of
Kc = Ae ð2Þ
Ce adsorption. It also provides a panorama of the course taken by the
o
system under study in a concise form, indicating how efficiently an
ΔG = −RT ln Kc ð3Þ adsorbent will adsorb and allows an estimate of the economic viability
of the adsorbent commercial applications for the specified solute. In
o o o
ΔG = ΔH −TΔS ð4Þ order to adapt for the considered system, an adequate model that can
reproduce the experimental results obtained, equations of Langmuir,

Fig. 6. Effect of contact time for the adsorption of CR onto CNS (the initial dye
concentration = 20 to 100 mg/L, CNS dose = 20 g/L and time = 2 h). Fig. 8. Thermodynamic study.
56 P. Senthil Kumar et al. / Desalination 261 (2010) 52–60

Table 2 β = 1, the Langmuir will be the preferable isotherm, while if β = 0, the


Thermodynamic parameters for the adsorption of CR onto CNS. Freundlich isotherm will be the preferable isotherm.
Initial CR concn. ΔHo ΔSo ΔGo (kJ/mol)
(mg/L) (kJ/mol) (J/mol/K) 3.8.4. Koble–Corrigan model
30 °C 40 °C 50 °C 60 °C
Koble–Corrigan model [37] is another three parameter empirical
20 −59.567 −163.976 −10.576 −7.288 −6.306 −5.545
model depends on the combination of the Langmuir and Freundlich
40 −31.650 −81.069 −7.268 −6.005 −5.404 −4.794
60 −22.651 −55.431 −6.022 −5.102 −4.656 −4.349 isotherm equations in one non-linear equation for representing the
80 −20.066 −49.476 −5.176 −4.503 −3.987 −3.715 equilibrium adsorption data. It is commonly expressed by Eq. (9):
100 −15.844 −37.873 −4.458 −3.867 −3.547 −3.317

n
aCe
qe = ð9Þ
Freundlich, Redlich–Peterson, Koble–Corrigan, Sips, Toth, Temkin and 1 + bCen
Dubinin–Radushkevich have been considered.
where a, b and n are the Koble–Corrigan parameters.
3.8.1. The Langmuir isotherm
The theoretical Langmuir sorption isotherm [34] is based on the 3.8.5. The Sips isotherm
assumption that the maximum adsorption occurs when a saturated The sips isotherm [38] has been used in the following form:
monolayer of solute molecules is present on the adsorbent surface, the
energy of adsorption is constant and there is no migration of
adsorbate molecules in the surface plane. The non-linear equation of Qmax KS Ce1 = n
qe = 1=n
ð10Þ
Langmuir isotherm model is expressed as follow: 1 + KS Ce

qm K L C e
qe = ð6Þ where KS is the Sips constant related with affinity constant (mg/L)−1 / n
1 + KL Ce
and Qmax is the Sips maximum adsorption capacity (mg/g).

where Ce is the supernatant concentration at the equilibrium state of 3.8.6. The Toth isotherm
the system (mg/L), qm and KL are the Langmuir constants, represent- The equation of Toth [39] combines the characteristics of Langmuir
ing the maximum adsorption capacity for the solid phase loading and and Freundlich model, which can be presented as:
the energy constant related to the heat of adsorption respectively.

3.8.2. The Freundlich isotherm fCe


qe =  1 ð11Þ
The Freundlich isotherm model [35] is the earliest known g + ðCe Þd =d
relationship describing the sorption process. The model applies to
adsorption on heterogeneous surfaces with interaction between
where f, g and d are Toth constants.
adsorbed molecules and the application of the Freundlich equation
also suggests that sorption energy exponentially decreases on
3.8.7. The Temkin isotherm
completion of the sorptional centers of an adsorbent. This isotherm
The Temkin isotherm model [40] contains a factor that explicitly
is an empirical equation can be employed to describe heterogeneous
takes into account adsorbing species-adsorbate interactions. This
systems and is expressed as follows:
model assumes the following: (i) the heat of adsorption of all the
1 molecules in the layer decreases linearly with coverage due to
=n
qe = Kf Ce ð7Þ adsorbent–adsorbate interactions, and that (ii) the adsorption is
characterized by a uniform distribution of binding energies, up to
where Kf is the Freundlich constant (L/g) related to the bonding some maximum binding energy. The derivation of the Temkin
energy. Kf can be defined as the adsorption or distribution coefficient isotherm assumes that the fall in the heat of sorption is linear rather
and represents the quantity of dye adsorbed onto adsorbent for unit than logarithmic, as implied in the Freundlich equation. The Temkin
equilibrium concentration. 1/n is the heterogeneity factor and n is a isotherm has commonly been applied in the following form:
measure of the deviation from linearity of adsorption. Its value
indicates the degree of non-linearity between solution concentration qe = B ln ðACe Þ ð12Þ
and adsorption as follows: if the value of n is equal to unity, the
adsorption is linear; if the value is below to unity, this implies that where A and B are Temkin isotherm constants.
adsorption process is chemical; if the value is above unity adsorption
is a favorable physical process. 3.8.8. Dubinin–Radushkevich isotherm
The Dubinin–Radushkevich [41] has the following form
3.8.3. The Redlich–Peterson isotherm −βε2
The Redlich–Peterson isotherm [36] is a combination of Langmuir– qe = qm e ð13Þ
Freundlich model. It approaches the Freundlich model at high
concentration and is in accord with the low concentration limit of where qm is the Dubinin–Radushkevich monolayer capacity (mg/g), β
the Langmuir equation. The equation is given as: a constant related to sorption energy, and ε is the Polanyi potential
which is related to the equilibrium concentration as follows
KR Ce  
qe = ð8Þ 1
1+ αR Ceβ ε = RT ln 1 + ð14Þ
Ce

where KR is Redlich–Peterson isotherm constant (L/g), αR is Redlich– where R is the gas constant (8.314 J/mol K) and T is the absolute
Peterson isotherm constant (L/mg) and β is the exponent which lies temperature. The constant β gives the mean free energy, E, of sorption
between 0 and 1. The constant β can characterize the isotherm as: if per molecule of the sorbate when it is transferred to the surface of the
P. Senthil Kumar et al. / Desalination 261 (2010) 52–60 57

solid from infinity in the solution and can be computed using the Table 3
relationship: The value of parameters for each isotherm models used in the studies.

Isotherm model Parameter R2 Equation


1:415Ce
Langmuir qm = 5.184 0.981 qe =
1 1 + 0:273Ce
E = pffiffiffiffiffiffi ð15Þ KL = 0.273
2β Freundlich KF = 1.357 0.998 qe = 1.357C0.439
e
n = 2.279
5:548Ce
Redlich–Peterson KR = 5.548 0.999 qe = 1 + 3:186Ce0:638
αR = 3.186
The experimental data on the effect of an initial concentration of
β = 0.638
CR on the CNS of the test medium were fitted to the isotherm models Koble–Corrigan model a = 1.454 0.999 qe =
1:454Ce0:543
1 + 0:102Ce0:543
using MATLAB 7.1 and the graphical representations of these models b = 0.102
are presented in Fig. 9a and b. All of the constants are presented in n = 0.543
1:453Ce0:543
Sips Qmax = 14.29 0.999 qe =
Table 3. Since the value of R2 nearer to 1 indicates that the respective 1 + 0:102Ce0:543
KS = 0.102
equation better fits the experimental data. The representations of the n = 1.841
experimental data by all models equation result in non-linear curve Toth f = 67.39 0.999 qe = 67:39Ce
5:618
½1:013 + ðCe Þ0:178 
with R2 values of a least 0.801 as tabulated in Table 3. The g = 1.013
experimental data yielded excellent fits with in the following d = 0.178
Temkin A = 3.767 0.976 qe = 0.444 ln (3.767Ce)
isotherms order: Redlich–Peterson N Toth N Koble–Corri gan N Sips N
B = 0.444
Freundlich N Langmuir N Temkin N Dubinin–Radushkevich, based on Dubinin–Radushkevich qm = 3.852 0.801 qe = 3.852e − 1.656 × 10
−7 2
ε

its R2 values. The plotted equations obtained from the graph are β = 1.656 × 10−7
presented in Table 3. The comparison of maximum monolayer E = 1736.62
adsorption capacity of some dyes onto various adsorbents was
presented in Table 4. It shows that the CNS studied in this work has
large adsorption capacity. This is due to its high surface area (395 m2/g). 3.9. Adsorption kinetics

In order to examine the controlling mechanism of adsorption


processes such as mass transfer and chemical reaction, pseudo-first-
order, pseudo-second-order and intraparticle diffusion kinetic equa-
tions were used to test the experimental data. The pseudo-first-order
kinetic model was suggested by Lagergren [46] for the adsorption of
solid/liquid systems and its formula is given as:

dqt
= kad ðqe −qt Þ ð16Þ
dt

After integration by applying the initial conditions qt = 0 at t = 0


and qt = qt at t = t, Eq. (16) becomes:
 
qe kad
log = t ð17Þ
qe −qt 2:303

Eq. (17) can be rearranged to obtain a linear form:

kad
log ðqe −qt Þ = log qe − t ð18Þ
2:303

where qt is the adsorption capacity at time t (mg/g) and kad (min−1) is


the rate constant of the pseudo-first order adsorption, was applied to
the present study of CR dye adsorption. The rate constant, kad and
correlation coefficients of the dye under different concentrations were
calculated from the linear plots of log(qe − qt) versus t (Fig. 10) and
listed in Table 5. The correlation coefficients for the pseudo-first-order
kinetic model are low. Moreover, a large difference of equilibrium
adsorption capacity (qe) between the experiment and calculation was

Table 4
Comparison of maximum monolayer adsorption of some dyes onto various adsorbents.

Dyes Adsorbents qm (mg/g) Reference

Congo red Acid activated red mud 7.08 [42]


Congo red Activated carbon prepared 6.70 [43]
from coir pith
Congo red Cashew nut shell 5.184 This work
Methylene blue Apricot stones-AC 4.11 [44]
Methylene blue Walnut shell-AC 3.53 [45]
Congo red Activated carbon (LR) 1.88 [44]
Methylene blue Almond shell-AC 1.33 [44]
Fig. 9. (a and b) The non-linear adsorption isotherm for CR with CNS at 30 °C.
58 P. Senthil Kumar et al. / Desalination 261 (2010) 52–60

Fig. 11. Pseudo-second-order kinetic fit for adsorption of CR onto CNS at 30 °C.


Fig. 10. Pseudo-first-order kinetic fit for adsorption of CR onto CNS at 30 °C.

observed, indicating a poor pseudo-first-order fit to the experimental phenomena have been observed in the adsorption of acid dyes on
data. surfactant-modified bentonites [48,49].
The kinetic data were further analyzed using Ho's pseudo-second- In order to gain insight into the mechanisms and rate controlling
order kinetics model [47]. This model is based on the assumption that steps affecting the kinetics of adsorption, the kinetic experimental
the sorption follows second-order chemisorption. It can be expressed results were fitted to the Weber's intraparticle diffusion [50]. The
as: kinetic results were analyzed by the intraparticle model to elucidate
the diffusion mechanism, which model is expressed as:
dqt 2
= kðqe −qt Þ ð19Þ
dt 1
=2 + C
qt = kp t ð22Þ
Integrating Eq. (19) and applying the boundary conditions, gives:
  where C is the intercept and kp is the intraparticle diffusion rate
1 1 constant, (mg/gmin1/2), which can be evaluated from the slope of the
= + kt ð20Þ
qe −qt qe linear plot of qt versus t(1/2) as shown in Fig. 12. The intercept of the
plot reflects the boundary layer effect. The larger the intercept, the
Eq. (20) can be rearranged to obtain a linear form: greater the contribution of the surface sorption in the rate controlling
step. The calculated intraparticle diffusion coefficient kp values are
t 1 1 listed in Table 5. If the regression of qt versus t(1/2) is linear and passes
= + t ð21Þ
qt h qe through the origin, then intraparticle diffusion is the sole rate-limiting
step. However, the linear plots at each concentration did not pass
where h = kq2e (mg g−1 min−1) can be regarded as the initial through the origin. This indicates that the intraparticle diffusion was
adsorption rate as t → 0 and k is the rate constant of pseudo-second- not only a rate controlling step.
order adsorption (g mg−1 min−1). The plot t / qt versus t (Fig. 11)
should give a straight line if pseudo-second-order kinetics is 4. Conclusion
applicable and qe, k and h can be determined from the slope and
intercept of the plot, respectively. At all studied initial dye CNS used in this investigation is freely, abundantly and locally
concentrations, the straight lines with extremely high correlation available; the resulting adsorbent is expected to be economically viable
coefficients (N0.998) were obtained. In addition, the calculated qe for removal of CR from aqueous solution. The results obtained from FT-
values also agree with the experimental data in the case of pseudo- IR spectrum shows that the CNS can be a viable option for its usage as an
second-order kinetics. These suggest that the adsorption data are well alternate adsorbent for CR removal. The specific surface area, pore
represented by pseudo-second-order kinetics and this supports the volume and average pore diameter calculated by BET equation states the
assumption [47] that the rate-limiting step of CR onto CNS may be effectiveness of CNS adsorbent for the removal of CR from aqueous
chemical sorption or chemisorption. From Table 5, the values of the solution. The effect of pH of aqueous solution decreases the removal
rate constant k decrease with increasing initial dye concentration for efficiency with the increase in pH due to the ionic effect. The % CR
the CNS. The reason for this behavior can be attributed to the lower removal increases from 56.33 to 99.34% for an increase in adsorbent
competition for the sorption surface sites at lower concentration. At dose from 5 to 30 g/L because of the concentration gradient between CR
higher concentrations, the competition for the surface active sites will concentration in the solution and the CR concentration in the surface of
be high and consequently lower sorption rates are obtained. Similar the CNS. The increase in initial dye concentration also enhances the

Table 5
Comparison between the adsorption rate constants, qe, estimated and correlation coefficients associated with pseudo-first-order and to the pseudo-second-order rate equations and
intraparticle diffusion.

Initial CR concn. Pseudo-first-order rate equation Pseudo-second-order rate equations Intraparticle diffusion
(mg/L)
kad (min−1) qe (mg/g) R2 k (g mg−1 min−1) qe , cal (mg/g) h (mg g−1 min−1) qe, exp (mg/g) R2 kp (mg/g. min1/2) C R2

20 0.074 1.367 0.912 0.096 1.096 0.115 0.982 0.999 0.063 0.436 0.903
40 0.067 2.483 0.897 0.045 2.123 0.203 1.889 0.999 0.124 0.805 0.937
60 0.071 4.102 0.875 0.031 3.086 0.294 2.741 0.999 0.184 1.145 0.918
80 0.069 5.105 0.901 0.022 4.016 0.361 3.538 0.999 0.245 1.417 0.927
100 0.074 6.637 0.898 0.019 4.831 0.451 4.261 0.999 0.294 1.737 0.910
P. Senthil Kumar et al. / Desalination 261 (2010) 52–60 59

[11] K.C.L.N. Rao, K.K. Ashutosh, Color removal from a dyestuff industry effluent using
activated carbon, Ind. J. Chem. Tech. 1 (1994) 13–19.
[12] S.V. Mohan, J. Karthikeyan, Removal of lignin and tannin colour from aqueous
solution by adsorption onto activated charcoal, J. Environ. Poll. 97 (1997) 183–187.
[13] S.H. Gharaibeh, S.V. Moore, A. Buck, Effluent treatment of industrial wastewater
using processed solid residue of olive mill products and commercial activated
carbon, J. Chem. Tech. Biotech. 71 (1998) 291–298.
[14] M. Sankar, G. Sekaran, S. Sadulla, T. Ramasami, Removal of diazo and
triphenylmethane dyes from aqueous solutions through an adsorption process,
J. Chem. Tech. Biotech. 74 (1999) 337–344.
[15] C. Pelekani, V.L. Snoeyink, A kinetic and equilibrium study of competitive
adsorption between atrazine and Congo red dye on activated carbon: the
importance of pore size distribution, Carbon 39 (2001) 25–37.
[16] A. Pala, E. Tokat, Color removal from cotton textile industry wastewater in an
activated sludge system with various additives, Water Res. 36 (2002) 2920–2925.
[17] K. Kadirvelu, M. Kavipriya, C. Karthika, M. Radhika, N. Vennilamani, S. Pattabhi,
Utilization of various agricultural wastes for activated carbon preparation and
application for the removal of dyes and metal ions from aqueous solution,
Fig. 12. Intraparticle diffusion model for adsorption of CR onto CNS at 30 °C. Bioresour. Technol. 87 (2003) 120–132.
[18] E. Lorenc-Grabowska, G. Gryglewicz, Adsorption characteristics of Congo red on
coal-based mesoporous activated carbon, Dyes Pigments 74 (2007) 34–40.
[19] H.M. Asfour, O.A. Fadali, M.M. Nassar, M.S. El-Geundi, Colour removal from textile
effluents using hardwood sawdust as an absorbent, J. Chem. Tech. Biotech. 35
interaction between CR and CNS which results in the decreased (1985) 28–35.
percentage of CR removal with the increase in initial concentration of [20] M.M. Nassar, M.S. El-Geundi, Comparative cost of colour removal from textile
CR. In the effect of contact time, the percentage removal increases with effluents using natural adsorbents, J. Chem. Tech. Biotech. 50 (1991) 257–264.
[21] M.S. El-Geundi, Colour removal from textile effluents by adsorption techniques,
time up to 90 min and remains almost constant. There is not much Water Res. 25 (1991) 271–273.
significant increase in the % of dye removal in view of the fact that the [22] K.S. Chou, J.C. Tsai, C.T. Lo, The adsorption of Congo red and vacuum pump oil by
aggregation of solid particles on the pore volume counteracts the rice hull ash, Bioresour. Technol. 78 (2001) 217–219.
[23] K.G. Bhattacharyya, A. Sharma, Azadirachta indica leaf powder as an effective
influence of contact time as the micro pores get filled up and start
biosorbent for dyes: a case study with aqueous congo red solutions, J. Environ.
offering resistance to diffusion of aggregated dye molecules in the Manage. 71 (2004) 217–229.
adsorbents. The temperature has adverse effect on the percentage of dye [24] A.E. Ofomaja, Y.S. Ho, Equilibrium sorption of anionic dye from aqueous solution
removal due to the decreased surface activity suggesting that adsorption by palm kernel fibre as sorbent, Dyes Pigments 74 (2007) 60–66.
[25] A.R. Binupriya, M. Sathishkumar, K. Swaminathan, C.S. Ku, S.E. Yun, Comparative
between CR and CNS was an exothermic process. The equilibrium data studies on removal of congo red by native and modified mycelia pellets of
have been analyzed using Langmuir, Freundlich, Redlich–Peterson, Trametes versicolor in various reactor modes, Bioresour. Technol. 99 (2008)
Koble–Corrigan, Sips, Toth Temkin and Dubinin–Radushkevich 1080–1088.
[26] L. Lian, L. Guo, C. Guo, Adsorption of Congo red from aqueous solutions onto Ca-
isotherms. The characteristic parameters for each isotherm and related bentonite, J. Hazard. Mater. 161 (2009) 126–131.
correlation coefficients have been determined. The experimental data [27] H. Chen, J. Zhao, Adsorption study for removal of Congo red anionic dye using
yielded excellent fits within the following isotherms order: Redlich– organo-attapulgite, Adsorption. 15 (2009) 381–389.
[28] T.S. Anirudhan, S.S. Sreekumari, C.D. Bringle, Removal of phenol from water and
Peterson N Toth N Koble–Corrigan N Sips N Freundlich N Langmuir N petroleum industry refinery effluents by activated carbon obtained from coconut
Temkin N Dubinin–Radushkevich, based on its correlation coefficient coir pith, Adsorption. 15 (2009) 439–451.
values. The thermodynamic analysis indicates that the system is [29] Z.A. Zakaria, M. Suratman, N. Mohammed, W.A. Ahmad, Chromium (VI) removal
from aqueous solution by untreated rubber wood sawdust, Desalination. 244
spontaneous and exothermic. The suitability of the pseudo-first- and (2009) 109–121.
second-order equations and intra particle diffusion kinetic model for the [30] M. Mastalerz, R.M. Bustin, Application of reflectance micro-Fourier transform
sorption of CR onto CNS is also discussed. The pseudo-second-order infrared spectrometry in studying coal macerals: comparison with other Fourier
transform infrared techniques, Fuel 74 (1995) 536–542.
kinetic model agrees very well with the dynamical behavior for the
[31] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers, J.
adsorption of CR onto CNS for different initial CR concentrations over the Am. Chem. Soc. 60 (1938) 309–319.
whole range studied. The CNS appeared to be suitable for the removal of [32] Y. Wong, J. Yu, Laccase-catalyzed decolorization of synthetic dyes, Water Res. 33
CR from aqueous solutions. (1999) 3512–3520.
[33] P.K. Malik, Use of activated carbons prepared from sawdust and rice-husk for
adsorption of acid dyes: a case study of Acid Yellow 36, Dyes Pigments 56 (2003)
239–249.
References [34] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum,
J. Ame. Chem. Soc. 40 (1918) 1361–1403.
[1] M.S. Chiou, P. Ho, Y. Ho, H.Y. Li, Adsorption of anionic dyes in acid solutions using [35] H.M.F. Freundlich, Over the adsorption in solution, J. Phys. Chem. 57 (1906)
chemically cross-linked chitosan beads, Dyes Pigments 60 (2004) 69–84. 385–470.
[2] R. Gong, Y. Ding, M. Li, C. Yang, H. Liu, Y. Sun, Utilization of powdered peanut hull [36] O. Redlich, D.L. Peterson, A useful adsorption isotherms, J. Phys. Chem. 63 (1959)
as biosorbent for removal of anionic dyes from aqueous solution, Dyes Pigments 1024–1026.
64 (2005) 187–192. [37] R.A. Koble, T.E. Corrigan, Adsorption isotherm for pure hydrocarbons, Ind. Eng.
[3] R. Han, D. Ding, Y. Xu, W. Zou, Y. Wang, Y. Li, L. Zou, Use of rice husk for adsorption Chem. Res. 44 (1952) 383–387.
of congo red from aqueous solution in column made, Bioresour. Technol. 99 [38] B. Volesky, Biosorption process simulation tools, Hydrometallurgy. 71 (2003)
(2008) 2938–2946. 179–190.
[4] P.J. Halliday, S. Beszedits, Color removal from textile mill wastewaters, Can. Tex. J. [39] J. Toth, State equations of the solid–gas interface layers, Acta Chim. Acad. Sci.
103 (1986) 78–84. Hung. 69 (1961) 311–328.
[5] G.S. Gupta, G. Prasad, V.N. Singh, Removal of chrome dye from aqueous solutions [40] M.J. Temkin, V. Pyzhev, Recent modifications to Langmuir isotherms, Acta
by mixed adsorbents: fly ash and coal, Water Res. 24 (1990) 45–50. Physicochim. URSS. 12 (1940) 217–225.
[6] M. Neamtu, A. Yediler, I. Siminiceanu, M. Macoveanu, A. Kellrup, Decolorization of [41] M.M. Dubinin, L.V. Radushkevich, Equation of the characteristic curve of activated
disperse red 354 azo dye in water by several oxidation processes—a comparative charcoal, Chem Zent. 1 (1947) 875.
study, Dyes Pigments 60 (2004) 61–68. [42] A. Tor, Y. Cengeloglu, Removal of congo red from aqueous solution by adsorption
[7] I.K. Kapdan, R. Ozturk, Effect of operating parameters on color and COD removal onto acid activated red med, J. Hazard. Mater. 138 (2006) 409–415.
performance of SBR: sludge age and initial dyestuff concentration, J. Hazard. [43] C. Namasivayam, D. Kavitha, Removal of congo red from water by adsorption onto
Mater. B123 (2005) 217–222. activated carbon prepared from coir pith, an agricultural solid waste, Dyes
[8] R.K. Wahi, W.W. Yu, Y.P. Liu, M.L. Meija, J.C. Falkner, W. Nolte, V.L. Colvin, Pigments 54 (2002) 47–58.
Photodegradation of Congo Red catalyzed by nanosized TiO2, J. Molecular Catal. A: [44] A. Aygun, S. Yenisoy-Karakas, I. Duman, Production of granular activated carbon
Chem. 242 (2005) 48–56. from fruit stones and nutshells and evaluation of their physical, chemical and
[9] V.V.B. Rao, S.R.M. Rao, Adsorption studies on treatment of textile dyeing industrial adsorption properties, Micropor. Mesopor. Mater. 66 (2003) 189–195.
effluent by flyash, Chem. Eng. J. 116 (2006) 77–84. [45] I.D. Mall, V.C. Srivastava, N.K. Agarwal, I.M. Mishra, Removal of congo red from
[10] J. Karthikeyan, in: R.K. Trivedy (Ed.), Pollution Management in Industries, aqueous solution by bagasse fly ash and activated carbon: kinetic study and
Environmental Publication, Karad India, 1988, p. 189. equilibrium isotherm analyses, Chemosphere 61 (2005) 492–501.
60 P. Senthil Kumar et al. / Desalination 261 (2010) 52–60

[46] S. Lagergren, About the theory of so-called adsorption of soluble substances, [49] P. Baskaralingam, M. Pulikesi, D. Elango, V. Ramamurthi, S. Sivanesan, Adsorption
Kungliga Svenska Vetensk Handl. 24 (1898) 1–39. of acid dye onto organobentonite, J. Hazard. Mater. 128 (2006) 138–144.
[47] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process [50] W.J. Jr, J.C. Morriss Weber, Kinetics of adsorption on carbon from solution, J. Sanit.
Biochem. 34 (1999) 451–465. Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 31–60.
[48] A.S. Ozcan, B. Erdem, A. Ozcan, Adsorption of Acid Blue 193 from aqueous
solutions onto BTMA-bentonite, Colloids Surf. A 266 (2005) 73–81.

Das könnte Ihnen auch gefallen