You are on page 1of 9

Desalination 266 (2011) 63–71

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

Removal of copper(II) ions from aqueous solution by adsorption using cashew


nut shell
P. SenthilKumar a, S. Ramalingam b, V. Sathyaselvabala c, S. Dinesh Kirupha 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: In the present study, cashew nut shell(CNS) was investigated as a biosorbent for the removal of copper ions
Received 15 June 2010 from aqueous solutions. Batch experiments were carried out to investigate the effect of solution pH, CNS
Received in revised form 3 August 2010 concentration, contact time, initial copper(II) ion concentration and temperature on sorption efficiency. The
Accepted 3 August 2010
copper adsorption was favored with maximum adsorption at pH 5.0. The percentage of copper ion removal
onto the CNS was decreased with increasing temperature. Biosorption equilibrium time was observed in
Keywords:
30 min. The equilibrium adsorption data were fitted to Langmuir and Freundlich adsorption isotherm models
Copper(II) ions
CNS
and the model parameters were evaluated. The kinetics of copper(II) ion was discussed using four kinetic
Isotherms models, the pseudo-first-order, the pseudo-second-order, the Elovich kinetic model and the intra-particle
Kinetics diffusion models. It was shown that the adsorption of copper ions could be described by the pseudo-second
Thermodynamics order kinetic model. Thermodynamic quantities such as Gibbs free energy (ΔGo), the enthalpy (ΔHo) and the
entropy change of sorption (ΔSo) have also been evaluated and it has been found that the sorption process
was feasible, spontaneous and exothermic in nature. The results showed that CNS could be employed as a
low-cost alternative adsorbent for the removal of copper ions from aqueous solutions.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction alternative treatment techniques for the effluents containing copper


before discharging into the environment.
Nowadays heavy metal pollution has become one of the most The conventional methods for removing the heavy metals include
important environmental problems. Heavy metal pollution has coagulation, chemical precipitation, ion-exchange, adsorption, mem-
accumulating characteristics in nature and cannot be biodegraded, brane separation, reverse osmosis, oxidation, evaporation, electro-
and this environmental issue is threatening the health of human flotation, solvent extraction, etc. However, most of these methods
beings seriously. Copper(II) is known to be one of the heavy metals have some disadvantages such as complicated treatment process, high
and is widely used in many industries including metal cleaning and cost and huge amount of energy consumption. Among these methods,
plating baths, paints and pigments, fertilizer, paper board, wood pulp, adsorption is the most widely used method because it is an
printed board circuit production, etc. [1,2]. The effluents from these economically feasible, simple, effective, versatile and environmentally
industries usually contain considerable quantity of copper, which friendly method in practice [6–8]. Even though commercially
spreads into the environment through soils and water streams and activated carbon has high surface area, microporous character and
finally accumulates along the food chain which causes human health high adsorption capacity, it has proven its potential as an adsorbent
hazards. The higher concentration of copper will cause severe mucosal for the removal of heavy metals from industrial wastewater, it is
irritation and corrosion, widespread capillary damage, hepatic and expensive with relatively high operating costs and there is need for
renal damage, and central nervous system irritation followed by regeneration after each adsorption cycle [9]. Hence, there is a growing
depression [3]. The maximum recommended concentration for demand to find low-cost and efficient, locally available adsorbents for
drinking water which is regulated by Environmental Quality Act the sorption of copper such as the sugar beet pulp [1], peanut hull [2],
1974 is 0.2 mg/L [3–5]. These limits suggest more stringent require- Cinnamomum camphora leaves powder [6], Litter of poplar forests
ment for the removal of copper from aqueous environment, which [10], Lentil shell [11], Wheat shell [11], Rice shell [11], Herbaceous
necessitated the development of innovative and cost-effective peat [12], T. grandis L.f. leaves powder [13], Base treated rubber leaves
[14], Pine cone powder [15], Spent grain [16], Tree fern [17],
Groundnut shells [18], Pretreated Aspergillus niger [19], Cedar sawdust
⁎ Corresponding author. Tel.: + 91 9444960106. [20], crushed brick [20], saw dust [21], etc. Because of the low cost and
E-mail address: sivanesh@yahoo.com (S. Sivanesan). high availability of these materials, it is not essential to have

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.08.003
64 P. SenthilKumar et al. / Desalination 266 (2011) 63–71

complicated regeneration process. This low cost adsorption method used to calculate the equilibrium metal adsorptive quantity by using
has attracted many scientists and engineers. the following mass balance relationship:
The objective of this research is to develop inexpensive and
effective adsorbents from the sources of natural wastes, such as ðCo −Ce ÞV
qe = ð1Þ
cashew nut shell, to replace the existing commercial materials. In the m
present study, cashew nut shell (CNS) from which the cashew nut
shell liquid (CNSL) was extracted [22] is used as an adsorbent and was where qe is the amount of heavy metal ion adsorbed onto per unit
examined for its sorption properties towards copper(II) ions. The weight of the adsorbent in mg/g, V is the volume of solution treated in
effect of various experimental parameters such as solution pH, CNS liter, Co is the initial concentration of metal ion in mg/L, Ce is the
concentration, contact time, initial copper(II) concentration and equilibrium metal ion concentration in mg/L and m is the mass of the
temperature has been investigated. Equilibrium modelling was adsorbent in grams.
carried out using Langmuir and Freundlich adsorption isotherms.
The nature of the sorption process has been evaluated with respect to 2.4.1. Effect of solution pH on copper(II) adsorption
its kinetic and thermodynamic aspects. The effect of solution pH on the adsorption capacity of CNS was
investigated using a 100 mL solution of 20 mg/L of copper(II) ion for a
2. Experimental pH range of 2.0 to 8.0 at 30 °C. Experiments could not be performed at
higher pH value due to hydrolysis and precipitation of copper(II) ions.
2.1. Adsorbent Flasks were agitated on a rotary shaker for 30 min to ensure that the
equilibrium was reached. The mixtures were then filtered through
The raw CNS was collected from Karaikudi, Sivagangai District, Whatman 42 filter paper and the concentration of copper(II) in the
Tamilnadu, India and the treated CNS was collected from the CNSL filtrates was analyzed using AA6300 Atomic absorption spectrometer.
recovery unit and it was used as an adsorbent. This natural waste was The above procedure was repeated three times and the average value
thoroughly rinsed with water to remove dust and soluble material and was taken.
was allowed to dry at room temperature. The above dried natural waste
was ground to a fine powder in a still mill. The resulting material was 2.4.2. Effect of CNS concentration
sieved in the size range of 200-30 mesh particle size. The proximate and Batch adsorption experiments were done at a different CNS
ultimate analysis of cashew nut shells are shown in Table 1. concentration from 0.1 g to 0.6 g in a 100 mL solution of 20 mg/L of
copper(II) ion at pH 5.0, for a contact time of 30 min at 30 °C. The
2.2. Adsorbate samples were then agitated and filtered and the filtrates were
analyzed as mentioned before.
All the chemicals used were of analytical reagent (AR) grade. Stock
solution of 100 mg/L of copper(II) ion was prepared from CuSO4.5H2O 2.4.3. Effect of contact time
(Merck, India) in double distilled water. Desired test solutions of Batch adsorption experiments were carried out at different contact
copper(II) ions were prepared using appropriate subsequent dilutions times (5, 10, 15, 20, 25, 30, 35, 40, 45, 50, 55, and 60 min) for an initial
of stock solution. The range in concentrations of copper(II) ions concentration in the range of 10–50 mg/L of copper(II) ion solution at
prepared from standard solution varied between 10 to 50 mg/L. pH 5.0, the CNS dose concentration was 0.3 g in 100 mL solution in
Before mixing the adsorbent, the pH of each test solution was adjusted 250 mL conical flask at 30 °C. The samples were then agitated and
to the required value with 0.1 M NaOH or 0.1 M HCl. filtered. The filtrates were analyzed as mentioned before.

2.3. Analysis 2.4.4. Effect of initial copper(II) ion concentration


Batch experiments were carried out by contacting 0.3 g of CNS
The concentration of copper(II) ions in the solutions before and with 100 mL of copper(II) ion solution of different initial concentra-
after equilibrium were determined by AA6300 Atomic absorption tions (10, 20, 30, 40, and 50 mg/L) at pH value of 5.0 at 30 °C. A series
spectrometer (Shimadzu, Japan). The pH of solution was measured of such conical flasks were shaken for 30 min at a speed of 120 rpm.
with a Hanna pH meter using a combined glass electrode. The samples were then agitated and filtered. The filtrates were
analyzed as mentioned before.
2.4. Adsorption experiments
2.4.5. Effect of temperature
Batch adsorption experiments were carried out by shaking the Batch experiments were performed at different temperatures of
flasks at 120 rpm for a fixed period of time using a rotary shaker. 30, 40, 50 and 60 °C for the initial copper(II) ion concentrations in the
Following a systematic process, the removal of copper(II) ions from range of 10–50 mg/L at constant CNS dose of 3 g/L and pH of 5. A series
aqueous solutions by the use of CNS in a batch system was studied in of such conical flasks were shaken for 30 min at a speed of 120 rpm.
the present research work. The data obtained in batch studies were The samples were then agitated and filtered. The filtrates were
analyzed as mentioned before.

2.5. Adsorption isotherms


Table 1
Properties of the Cashew nut shells.
A series of solutions containing different initial concentrations of
Proximate analysis Ultimate analysis Ash chemical composition
copper(II) ions was prepared and batch adsorption studies were
(wt %) (wt %) (wt %)
undertaken at 30 °C employing these solutions to check the
Volatile matter 65.21 Carbon 45.21 Silica 64.53
applicability of the Langmuir and Freundlich adsorption isotherms
Moisture 9.83 Hydrogen 4.25 Iron Oxide 3.27
Ash 2.75 Oxygen 37.75 Aluminium Oxide 2.19
under the specified conditions, i.e. initial solution pH of 5, contact time
Fixed carbon 22.21 Nitrogen 0.21 Calcium Oxide 26.89 of 30 min, adsorbent dose of 0.3 g and an initial copper(II) ion
Sulphur Nil Magnesium Oxide 2.49 concentration range of 10–50 mg/L. Analysis of copper(II) ion content
Moisture 9.83 Sodium Oxide 0.63 of the various solutions was performed by Atomic absorption
Ash 2.75
spectrometer methods.
P. SenthilKumar et al. / Desalination 266 (2011) 63–71 65

2.6. Adsorption Kinetics characteristic of PO2 stretching. FT-IR studies reveal that several
functional groups, which are able to bind with heavy metal ions, in
Kinetic measurements were conducted under static conditions particular copper(II) ions were present in the CNS. These results agree
employing a glass vessel equipped with a rotary shaker. Thus, the with the surface chemistries of other agricultural by-products, such as
adsorbent dose of 0.3 g of CNS was contacted with 100 mL of a copper rubber wood saw dust [23] and heartwood of Areca catechu powder
(II) ion solution of known different concentrations (10–50 mg/L). The [24]. The specific surface area and pore structure of the CNS was
concentration of copper(II) ion in the wastewater was determined at determined by using surface area and pore size analyzer (Quanta-
known time intervals. Analysis of copper(II) ion content in the various chrome, Autosorb-I) on nitrogen adsorption at 77 K. The specific surface
solutions was performed by atomic absorption spectrometer methods. area was calculated by BET equation [25]. It was found that the BET
surface area, pore volume, average pore diameter and bulk density of the
3. Results and discussion CNS were 395 m 2 /g, 0.4732 cm 3 /g, 5.89 nm and 0.415 g/cm3 ,
respectively.
3.1. Characterization of CNS
3.2. Effect of solution pH
The chemical functional groups such as carbonyl, hydroxyl, amine,
amide etc., have been identified as potential adsorption sites to be The pH of the solution is one of the most critical parameters in the
responsible for binding metallic ions to the adsorbent. The adsorption adsorption of metal ions from aqueous solutions. This parameter is
capacity of CNS depends upon porosity as well as chemical reactivity directly related with competition ability of hydrogen ions with metal
of functional groups at the surface. This reactivity creates an ions to actives sites on the CNS surface. Generally, metal biosorption
imbalance between the forces at the surface when compared to involves complex mechanisms of ion-exchange, chelation, adsorption
those within the body, thus leading to molecular adsorption by the by physical forces and ion entrapment in inter and intrafibrillar
van der Waals force. Knowledge on surface functional groups would capillaries and spaces of the structural network of a biosorbent. The
give the insight of the adsorption capability of the CNS. The FT-IR FT-IR spectroscopic analysis showed that the CNS has a variety of
spectrums of the CNS are shown in Fig. 1. OH and NH stretching functional groups such as carboxyl, hydroxyl and amine groups are
between 3100 and 3500 cm−1, C–H aromatic in between 3000 and involved in almost all potential binding mechanisms. Moreover,
3100 cm−1, C–H aliphatic in between 2800 and 3000 cm−1. The depending on the pH value of the aqueous solution these functional
spectrum shows a broad band near 3399 cm−1, which indicates the groups participate in metal ion bindings. The effect of solution pH on
presence of hydroxyl groups on the CNS surface. The stretching was the adsorption of copper(II) ions onto CNS was evaluated within the
attributed to the absorbed water on the surface of CNS. The peak at pH range of 2.0–8.0 and the result is shown in Fig. 2. Copper(II)
2925 cm−1 is due to C–H stretching of CH2 groups. The stretching removal recorded its minimum values at pH 2.0. There was an
frequencies of the aromatic C = C and aromatic C–H groups give rise to increase in the percentage of copper(II) removal with an increase in
peaks at 3011 and 2854 cm−1 respectively. The bands near 1630 cm−1 pH from 2.0 to 5.0 and showed marginal downward trend from pH
indicates fingerprint region of C = O, C–O and O–H groups that exist 6.0. The maximum adsorption was found to be 82.11% for copper(II)
as functional groups of CNS. The peaks at 1542 and 1515 cm−1 is ion at pH 5. At highly acidic pH (b2.0), the overall surface charge on
assigned to a conjugated hydrogen bonded carboxyl group. The peak at the active sites became positive and metal cations and protons
1454 cm−-1 (vC–O) indicates the presence of carboxylic groups. The compete for binding sites on cell wall, which results in lower uptake of
peak at 1374 cm−1 indicates the presence of C–H aliphatic bending. metal. As the solution pH was increased, the ability of copper(II) ions
The peaks at 1232 cm−1 indicate the presence of the C–N from amine. for competition with protons was also increased. Although the
The two peaks at 1156 cm− 1 (P= O) and 1035 cm− 1 (P–OH) was sorption of metal ions raised by increasing solution pH, further

Fig. 1. FT-IR spectrum of Cashew nut shell.


66 P. SenthilKumar et al. / Desalination 266 (2011) 63–71

Fig. 2. Effect of pH for the adsorption of copper(II) ions onto CNS (the initial copper(II) Fig. 4. Effect of contact time for the adsorption of copper(II) ions onto CNS (the initial
ions concentration = 20 mg/L, CNS dose = 3 g/L and time = 30 min). copper(II) ions concentration = 10 to 50 mg/L, pH = 5.0, CNS dose = 3 g/L and
time = 30 min).

increment in pH caused decline in adsorption due to the precipitation


of metal hydroxides. percent of copper(II) ions removal is higher in the beginning due to the
larger surface area of the adsorbent being available for the adsorption of
3.3. Effect of CNS concentration the metals. As the surface adsorption sites become exhausted, the rate
of uptake is controlled by the rate of transport from the exterior to the
The CNS concentration is an important parameter because this interior sites of the adsorbent particles [26]. It is also relevant that, since
determines the capacity of the CNS for a given initial copper(II) ion active sorption sites in a system have a fixed number and each active
concentration. Fig. 3. shows the influence of CNS concentration on sites can adsorb only one ion in a monolayer, the metal uptake by the
copper(II) adsorption and it was examined by varying CNS concen- sorbent surface will be rapid initially, slowing down as the competition
tration from 1 to 6 g/L while keeping the volume of the copper(II) ion for the decreasing availability of active sites intensifies by the metal ions
solutions constant and at pH of 5.0. The percentage of copper(II) ion remaining in the solution [26,27].
removal steeply increases with the CNS loading up to 3 g/L. This result
can be explained by the fact that the biosorption sites remain
unsaturated during the biosorption whereas the number of sites 3.5. Effect of initial copper(II) ion concentration
available for biosorption site increases by increasing the biosorbent
dose. The maximum biosorption efficiency of copper(II) ion onto CNS The copper(II) uptake is particularly dependent on initial copper
was found to be 82% at CNS concentration of 3 g/L. There was a non- (II) ion concentration. At low concentration values, copper(II) ions are
significant increase in the percentage removal of copper(II) ion when adsorbed at specific sites, while with increasing copper(II) ion
the CNS concentration increases beyond the 3 g/L. This suggests that concentrations the specific sites are saturated with and exchange
after a certain dose of biosorbent, the maximum adsorption is attained sites are filled. The different concentrations of copper(II) ions
and hence the amount of ions remains constant even with further solutions of 10, 20, 30, 40 and 50 mg/L were used and the experiments
addition of dose of adsorbent [26]. were performed at room temperature (30 °C) with contact time of
30 min. Fig. 5. shows the effect of metal concentration on the percent
3.4. Effect of contact time removal of copper(II) ions. From figure, it was shown that the
percentage copper(II) ion adsorption decreased from 86.03% to
The effect of contact time on copper(II) biosorption on CNS was 76.17% with the increment of the initial copper(II) ions concentration.
studied and the results were shown in Fig. 4. The data obtained from the At the lower copper(II) ion concentrations, the copper(II) removal
biosorption of copper(II) ions onto the CNS showed that the adsorption percentage was higher due to a larger surface area of the CNS being
increases with increase in contact time. The biosorption of copper(II) available for the adsorption of copper(II) ions. When the concentra-
ions onto CNS was rapid for the first 5 min and equilibrium was nearly tion of the copper(II) ion solution became higher, the copper(II)
reached after 30 min for five different initial copper(II) ion concentra- removal percentage was lower because the available sites of the
tions. Hence, in the present work, 30 min was chosen as the equilibrium adsorption became less. At a higher initial copper(II) ion concentra-
time. Generally the removal rate of sorbate is rapid initially, but it tion, the ratio of initial number of moles of copper(II) ion to the
gradually decreases with time until it reaches equilibrium. The rate in

Fig. 5. Effect of initial copper(II) ions concentration for the adsorption of copper(II) ions
Fig. 3. Effect of CNS concentration on copper(II) removal (the initial copper(II) ions onto CNS (the initial copper(II) ions concentration = 10 to 50 mg/L, pH = 5.0, CNS
concentration = 20 mg/L, pH = 5.0, and time = 30 min). dose = 3 g/L and time = 30 min).
P. SenthilKumar et al. / Desalination 266 (2011) 63–71 67

Table 2
Thermodynamic parameters for the adsorption of copper(II) ions onto CNS.

Initial Copper(II) ΔHo ΔSo ΔGo(kJ/mol)


Concn. (mg/L) (kJ/mol) (J/mol/K)
30 °C 40 °C 50 °C 60 °C

10 −11.417 −22.555 −4.579 −4.368 −4.088 −3.922


20 − 10.052 −20.334 −3.915 −3.659 −3.452 −3.311
30 −9.631 −20.411 −3.467 −3.202 −3.019 −2.851
40 − 10.041 −22.306 −3.301 −3.029 −2.834 −2.626
50 −9.359 −20.947 −3.003 −2.801 −2.602 −2.369

ΔGo,ΔHo and ΔSoare changes in Gibbs free energy (kJ/mol), enthalpy


(kJ/mol) and entropy (J/mol/K), respectively. R is the gas constant
Fig. 6. Effect of temperature on copper(II) ions removal by CNS (the initial copper(II) (8.314 J/mol/K), T is the temperature (K). The values of ΔHo and ΔSo
ions concentration = 10 to 50 mg/L, pH = 5.0, CNS dose = 3 g/L and time = 30 min). are determined from the slope and the intercept of the plots of log Kc
versus 1/T (Fig. 7). The ΔGo values were calculated using Eq. (3).
available adsorption surface area was higher and as a result adsorption Adsorption of copper(II) ion on CNS decreased, when the temperature
percentage was less. was increased from 303 to 333 K which is shown in Fig. 7. The process
was thus exothermic in nature. The plots were used to compute the
3.6. Effect of temperature and thermodynamic parameters values of thermodynamic parameters (Table 2). The value of enthalpy
change (ΔHo) and the entropy change (ΔSo) recorded from this work
The adsorption of copper(II) ion on CNS was investigated as a were presented in the Table 2. The Gibbs free energy ΔGois small and
function of temperature and the maximum removal of copper(II) ion negative but increases with increasing temperature, indicating that
was obtained at 30 °C. The batch adsorption experiments were the adsorption process led to a decrease in Gibbs energy. The negative
performed at different temperatures of 30, 40, 50 and 60 °C for the ΔGo value indicates that the process is feasible and spontaneous in the
initial copper(II) ion concentrations in the range of 10-50 mg/L at nature of adsorption; negative ΔHo value suggests the exothermic
constant CNS concentration of 3 g/L and optimum pH value of 5. The nature of adsorption and the ΔSo can be used to describe the
adsorption decreased from 86.03% to 80.48%, 82.55% to 76.78%, 79.84% randomness at the CNS–solution interface during the sorption. The
to 73.69% and 76.71% to 70.18% for the initial copper(II) ion thermodynamic parameters thus indicate that this adsorption process
concentration in the range of 10–50 mg/L respectively with the rise can be used for the removal of copper(II) ions by CNS.
in temperature from 30 to 60 °C (Fig. 6). This is mainly due to the
decrease in surface activity suggesting that adsorption between 3.7. Adsorption isotherm models
copper(II) and CNS is an exothermic process.
Thermodynamic parameters such as free energy (ΔGo), enthalpy The most commonly used isotherm equations are Langmuir and
(ΔHo) and entropy (ΔSo) change of adsorption can be evaluated from Freundlich models which are used to find out the relation between the
the following Eqs. (2)–(4) equilibrium concentrations of the adsorbate in liquid phase and in the
solid phase.
CAe
Kc = ð2Þ
Ce 3.7.1. The Langmuir isotherm model
The Langmuir isotherm model [27] is valid for monolayer
o
ΔG = −RT ln Kc ð3Þ adsorption onto surface containing finite number of identical sorption
sites which is presented by the following equation:
o o o
ΔG = ΔH −TΔS ð4Þ
qm KL Ce
o o
qe = ð6Þ
ΔS ΔH 1 + KL Ce
log Kc = − ð5Þ
2:303 R 2:303 RT
where qe is the amount of metal adsorbed per specific amount
where Kc is the equilibrium constant, Ce is the equilibrium con- of adsorbent (mg/g), Ce is equilibrium concentration of the solution
centration in solution (mg/L) and CAeis the amount of copper(II) (mg/L), and qm is the maximum amount of metal ions required to form a
adsorbed on the adsorbent per liter of solution at equilibrium (mg/L). monolayer (mg/g). The Langmuir equation can be rearranged to linear
form for the convenience of plotting and determining the Langmuir
constants (KL) and maximum monolayer adsorption capacity of
CNS (qm). The values of qm and KL can be determined from the linear
plot of 1/qe versus 1/Ce:

1 1 1 1
= + ð7Þ
qe qm qm KL Ce

3.7.2. The Freundlich isotherm model


The Freundlich equation [28] is purely empirical based on sorption
on heterogeneous surface, which is commonly described by the
following equation:

1=

Fig. 7. Thermodynamic study. qe = Kf Ce n


ð8Þ
68 P. SenthilKumar et al. / Desalination 266 (2011) 63–71

Table 3
The value of parameters for each isotherm models used in the studies.

Isotherm Model Parameter R2 Equation

Langmuir qm = 20.00 0.996 2:36 C


qe = 1 + 0:118e C
e
KL = 0.118
Freundlich KF = 2.275 0.998 qe = 2.275 C0.701
e
n = 1.427

was found to be 0.459 to 0.145 for concentration of 10–50 mg/L of


copper(II).
The linear plot of log qe versus log Ce (Fig. 9) showed that the
Freundlich isotherm was also well representative for the copper(II)
Fig. 8. The linear Langmuir adsorption isotherm for copper(II) ions with CNS at 30 °C.
ions removal by the CNS. The correlation coefficient was 0.998. Kf and
n were calculated from the slopes of the Freundlich plots (Fig. 9) and
were found to be 2.275 (mg/g)(l/mg)(1/n) and 1.427 respectively. The
where, Kf and 1/n are the Freundlich constants related to adsorption magnitude of Kf and n shows easy separation of heavy metal ion from
capacity and adsorption intensity, respectively. The Freundlich wastewater and high adsorption capacity. The value of n, which is
equilibrium constants evaluated from the intercept and the slope, related to the distribution of bonded ions on the adsorbent surface,
respectively of the linear plot of log qe versus log Ce based on represents beneficial adsorption if it is between 1 and 10 [30]. The n
experimental data. The Freundlich equation can be linearized in value for the CNS used was found to be greater than 1, indicating that
logarithmic form for the determination of the Freundlich constants as adsorption of copper(II) is favourable.
shown below: Table 3 gives isotherm parameters for both Langmuir and
Freundlich isotherms. From linear correlation coefficients of the
1 adsorption isotherm, it is noted that the Langmuir and Freundlich
log qe = log Kf + log Ce ð9Þ isotherm models exhibits a better fit to the sorption data of copper(II)
n
ions. The comparison of maximum monolayer adsorption capacity of
copper(II) ions onto various adsorbents was presented in Table 4. It
The Langmuir and Freundlich equations were used to describe the
shows that the CNS studied in this work has large adsorption capacity.
data derived from the adsorption of copper(II) by CNS over the entire
This is due to its high surface area (395 m2/g).
concentration range studied (10 mg/L to 50 mg/L). The linear plot of
Langmuir and Freundlich isotherm models for sorption of copper(II)
ion on CNS were presented in Figs. 8 and 9. The plot of 1/qe versus 1/Ce
3.8. Adsorption kinetics
(Fig. 8) showed that the experimental data fitted reasonably well to
the linearized equation of the Langmuir isotherm over the whole
Various kinetic models, namely pseudo-first-order, pseudo-second-
copper(II) ion concentration range studied. The correlation coefficient
order, Elovich models and intraparticle diffusion, have been used for
was 0.996. qm and KL were evaluated from the slope and intercept of
their validity with the experimental adsorption data for copper(II) onto
the plot and were found to be 20 mg/g and 0.118 L/mg respectively.
CNS. The study of adsorption kinetics describes the solute uptake rate
The essential characteristics of the Langmuir isotherm parameters
and evidently this rate controls the residence time of adsorbate uptake
can be used to predict the affinity between the sorbate and sorbent
at the solid-solution interface including the diffusion process. The
using separation factor or dimensionless equilibrium parameter, “RL”,
mechanism of adsorption depends on the physical and chemical
expressed as in the following equation [29]:
characteristics of the adsorbent as well as on the mass transfer process.
The results obtained from the experiments were used to study the
1
RL = ð10Þ kinetics of metal ion adsorption. The rate kinetics of copper(II) ion
1 + KL Co
adsorption CNS was analysed using pseudo-first-order [31], pseudo-
second-order [32], Elovich models [33] and intraparticle diffusion [34].
where KL is the Langmuir constant and Co is the initial concentration of
copper(II) ion. The value of separation parameter RL provides
important information about the nature of adsorption. The value of Table 4
RL indicated the type of Langmuir isotherm to be irreversible (RL = 0), Comparison of maximum monolayer adsorption of copper(II) ions onto various
adsorbents.
favourable (0 b RL b 1), linear (RL = 1) or unfavourable (RL N 1). The RL
Adsorbents qm (mg/g) Reference

Sugar beet pulp 31.40 [1]


Peanut hull 21.25 [2]
Cashew nut shell 20.00 This work
Cinnamomum camphora leaves powder 16.756 [6]
Litter of poplar forests 19.53 [10]
T. grandis L.f. leaves powder 15.43 [13]
Base treated rubber leaves 14.97 [14]
Tree fern 10.60 [17]
Spent grain 10.47 [16]
Lentil shell 8.98 [11]
Groundnut shells 7.60 [18]
Wheat shell 7.39 [11]
Pine cone powder 6.80 [15]
Pretreated Aspergillus niger 2.61 [19]
Herbaceous peat 4.84 [12]
Rice shell 1.85 [11]
Fig. 9. The linear Freundlich adsorption isotherm for copper(II) ions with CNS at 30 °C.
P. SenthilKumar et al. / Desalination 266 (2011) 63–71 69

Fig. 11. Pseudo-second-order kinetic fit for adsorption of copper(II) ions onto CNS at
30 °C.
Fig. 10. Pseudo-first-order kinetic fit for adsorption of copper(II) ions onto CNS at 30 °C.

and listed in Table 5. The correlation coefficients for the pseudo-first-


3.8.1. The pseudo-first-order kinetic model order kinetic model are low. Moreover, a large difference of equilibrium
The adsorption of copper(II) ion from an aqueous solution to CNS adsorption capacity (qe) between the experiment and calculation was
can be considered as a reversible process with equilibrium being observed, indicating a poor pseudo first-order fit to the experimental
established between the solution and CNS. Adsorption phenomenon data.
can be described as the diffusion control process assuming a non-
dissociation molecular adsorption of copper(II) on CNS as follows: 3.8.2. The pseudo-second-order kinetic model
The kinetic data were further analyzed using Ho's pseudo-second-
A + S⇔AS ð11Þ order kinetic model. This model is based on the assumption that the
sorption follows second order chemisorption. It can be expressed as:
If initially no copper(II) ions were present on the CNS surface (i.e.,
dqt
CASo = 0 at t = 0), then assuming the first-order rate kinetics the = k ðqe −qt Þ
2
ð15Þ
fractional uptake on the copper(II) by the CNS can be expressed as: dt
 
qt k Integrating Eq.(15) and applying the boundary conditions, gives:
= 1− exp kA CS + A t ð12Þ
qe kS  
1 1
= + kt ð16Þ
qe −qt qe
Equation can be transformed as
kad Eq. (16) can be rearranged to obtain a linear form:
logðqe −qt Þ = log qe − t ð13Þ
2:303
where t 1 1
  = + t ð17Þ
k qt h qe
kad = kA CS + A ð14Þ
kS
where h = kq2e (mg g− 1 min− 1) can be regarded as the initial adsorption
qt = XAt and qe = XAe rate as t → 0 and k is the rate constant of pseudo-second-order
adsorption (g mg− 1 min− 1). The plot of t/qt versus t (Fig. 11) should
where qt is the adsorption capacity at time t (mg/g) and kad (min− 1) is give a straight line if pseudo-second-order kinetics is applicable and qe, k
the rate constant of the pseudo-first order adsorption, was applied to the and h can be determined from the slope and intercept of the plot,
present study of copper(II) adsorption. The rate constant, kad and respectively. At all studied initial copper(II) concentrations, the straight
correlation coefficients of the copper(II) under different concentrations lines with extremely high correlation coefficients (N0.99) were
were calculated from the linear plots of log(qe − qt) versus t (Fig. 10) obtained. In addition, the calculated qe values also agree with the

Table 5
Kinetic models and other statistical parameters at 30 °C and at pH 5.

Kinetic model Parameters Concentration of copper(II) solution

10 mg/L 20 mg/L 30 mg/L 40 mg/L 50 mg/L

Pseudo-first-order equation kad (min−1) 0.124 0.140 0.143 0.145 0.154


qe, cal (mg/g) 1.415 2.264 2.826 3.425 4.100
R2 0.941 0.974 0.968 0.962 0.951
Pseudo-second-order equation k (g mg−1 min−1) 0.088 0.036 0.021 0.014 0.010
qe ,cal (mg/g) 2.962 5.898 8.408 11.004 13.195
h (mg g−1 min−1) 0.770 1.239 1.488 1.751 1.808
qe,exp (mg/g) 2.854 5.519 8.010 10.408 12.664
R2 0.997 0.997 0.995 0.994 0.992
Intraparticle diffusion kp (mg/g.min1/2) 0.205 0.448 0.723 0.988 1.313
C 1.499 2.559 3.212 3.821 3.881
R2 0.808 0.818 0.835 0.849 0.858
Elovich equation α (mg/g.min) 3.657 4.178 4.084 4.438 4.079
β (g/mg) 2.012 0.924 0.578 0.424 0.321
R2 0.917 0.924 0.929 0.936 0.936
70 P. SenthilKumar et al. / Desalination 266 (2011) 63–71

experimental data in the case of pseudo-second-order kinetics. These


suggest that the adsorption data are well represented by pseudo-
second-order kinetics and the supports the assumption [32] that the
rate-limiting step of copper(II) ion onto CNS may be chemisorption.
From Table 5, the values of the rate constant k decrease with increasing
initial copper(II) concentration for the CNS. The reason for this behavior
can be attributed to the lower competition for the sorption surface sites
at lower concentration. At higher concentrations, the competition for
the surface active sites will be high and consequently lower sorption
rates are obtained.

3.8.3. The Elovich kinetic model


Elovich kinetic model is also considered where α is the initial
adsorption rate in mg/(g.min), and β (g/mg) is the desorption Fig. 13. Intraparticle diffusion model for adsorption of copper(II) ions onto CNS at 30 °C.
constant related to the extent of the surface coverage and activation
energy for chemisorptions. Both the kinetic constants (α and β) will sorption or all may be operating simultaneously. Thus based on the
be estimated from the slope and intercept of the plot of qt versus ln t. high co-relation coefficient values (Table 5), it can be inferred that the
adsorption of copper(II) onto CNS followed pseudo-second-order
1 1 model than that of intraparticle diffusion model, pseudo-first-order
qt = lnðα βÞ + ln t ð18Þ
β β model and Elovich kinetic model.

Copper(II) ion adsorption kinetics onto CNS was also tested with
Elovich kinetic model by plotting qt versus ln t as shown in Fig. 12. 4. Conclusions
Recorded R2 values are low which indicates that the experimental
data does not fit the Elovich kinetic model and copper(II) ions This study was focused on the adsorption of copper(II) ions onto
removal using CNS under study cannot be described using this model. cashew nut shell from aqueous solution. The operating parameters
such as solution pH, CNS concentration, contact time, initial copper(II)
3.8.4. The intraparticle diffusion model concentration and temperature, were effective on the adsorption
Taking into account that the kinetic results are fitted very well to a efficiency of copper(II). The monolayer adsorption capacity of cashew
chemisorption model, the intraparticle diffusion model was plotted in nut shell was found to be 20.00 mg/g of copper(II) ions at pH 5. The
order to verify the influence of mass transfer resistance on the binding adsorption of the copper(II) ion on CNS reached equilibrium in
of copper(II) ions to the CNS. The kinetic results were analyzed by the 30 minutes. The Langmuir and Freundlich isotherm models were used
Weber and Morris intraparticle diffusion model to elucidate the for the mathematical description of the adsorption of copper(II) ions
diffusion mechanism, which model is expressed as: onto CNS and the constants were evaluated from these isotherms.
Results indicated that the adsorption equilibrium data fitted very well
1=
qt = kp t 2
+C ð19Þ to both the isotherm models in the studied concentration range at
30 °C. The kinetic data showed that the pseudo-second-order kinetic
model was obeyed better than the pseudo-first-order, Elovich kinetic
where C is the intercept and kp is the intraparticle diffusion rate
and intraparticle diffusion models. The thermodynamic calculations
constant, (mg/g min1/2), which can be evaluated from the slope of the
indicated the feasibility, exothermic and spontaneous nature of
linear plot of qt versus t(1/2) as shown in Fig. 13. The intercept of the
adsorption of copper(II) ions onto CNS at 30-60 °C. It is concluded
plot reflects the boundary layer effect. Larger the intercept, greater is
that the CNS is an effective and alternative adsorbent for the removal
the contribution of the surface sorption in the rate controlling step.
of copper(II) ions from aqueous solutions in terms of high adsorption
The calculated intraparticle diffusion coefficient kp values are listed in
capacity, natural and abundant availability, and low cost.
Table 5. If the regression of qt versus t(1/2) is linear and passes through
the origin, then intraparticle diffusion is the sole rate-limiting step.
However, the linear plots at each concentration did not pass through References
the origin. This deviation from the origin is perhaps due to the
difference in the rate of mass transfer in the initial and final stages of [1] Z. Aksu, I.A. Isoglu, Removal of copper(II) ions from aqueous solution by
biosorption onto agricultural waste sugar beet pulp, Process Biochem. 40
the adsorption. This is indicative of some degree of boundary layer (2005) 3031–3044.
control and this further showed that the intraparticle diffusion was [2] C.S. Zhu, L.P. Wang, W.B. Chen, Removal of Cu(II) from aqueous solution by
not the only rate-limiting step, but also be controlling the rate of agricultural by-product: Peanut hull, J. Hazard. Mater. 168 (2009) 739–746.
[3] A. Ahmad, M. Rafatullah, O. Sulaiman, M.H. Ibrahim, Y.Y. Chii, B.M. Siddique,
Removal of Cu(II) and Pb(II) ions from aqueous solutions by adsorption on
sawdust of Meranti wood, Desalination 247 (2009) 636–646.
[4] P. King, P. Srinivas, Y.P. Kumar, V.S.R.K. Prasad, Sorption of copper(II) ion from
aqueous solution by Tectona grandis I.f. (teak leaves powder), J. Hazard. Mater. 136
(2006) 560–566.
[5] B. Yu, Y. Zhang, A. Shukla, S.S. Shukla, K.L. Dorris, The removal of heavy metal from
aqueous solutions by sawdust adsorption-removal of copper, J. Hazard. Mater.
B80 (2000) 33–42.
[6] H. Chen, G. Dai, J. Zhao, A. Zhong, J. Wu, H. Yan, Removal of copper(II) ions by a
biosorbent-Cinnamomum camphora leaves powder, J. Hazard. Mater. 177 (2010)
228–236.
[7] M. Sciban, M. Klasnja, B. Skrbic, Adsorption of copper ions from water by modified
agricultural by-products, Desalination 229 (2008) 170–180.
[8] M. Sarioglu, U.A. Guler, N. Beyazit, Removal of copper from aqueous solutions
using biosolids, Desalination 239 (2009) 167–174.
[9] I. Ali, V.K. Gupta, Advances in water treatment by adsorption technology, Nat.
Protoc. 1 (2007) 2661–2667.
[10] M. Dundar, C. Nuhoglu, Y. Nuhoglu, Biosorption of Cu(II) ions onto the litter of
Fig. 12. Elovich kinetic model for adsorption of copper(II) ions onto CNS at 30 °C. natural trembling poplar forest, J. Hazard. Mater. 151 (2008) 86–95.
P. SenthilKumar et al. / Desalination 266 (2011) 63–71 71

[11] H. Aydin, Y. Bulut, C. Yerlikaya, Removal of copper(II) from aqueous solution by [22] P.S. Kumar, N.A. Kumar, R. Sivakumar, C. Kaushik, Experimentation on solvent
adsorption onto low-cost adsorbents, J. Environ. Manage. 87 (2008) 37–45. extraction of polyphenols from natural waste, J. Mater. Sci. 44 (2009) 5894–5899.
[12] R. Gundogan, B. Acemioglu, M.H. Amla, Copper(II) adsorption from aqueous [23] Z.A. Zakaria, M. Suratman, N. Mohammed, W.A. Ahmad, Chromium (VI) removal
solution by herbaceous peat, J. Colloid Interface Sci. 269 (2004) 303–309. from aqueous solution by untreated rubber wood sawdust, Desalination 244
[13] Y.P. Kumar, P. King, V.S.R.K. Prasad, Equilibrium and kinetic studies for the (2009) 109–121.
biosorption system of copper(II) ion from aqueous solution using Tectona grandis [24] P. Chakravarty, N.S. Sarma, H.P. Sharma, Removal of Pb(II) from aqueous solution
L.f. leaves powder, J. Hazard. Mater. 137 (2006) 1211–1217. using heartwood of Areca catechu powder, Desalination 256 (2010) 16–21.
[14] W.S.W. Ngah, M.A.K.M. Hanafiah, Biosorption of copper ions from dilute [25] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers, J.
aqueous solutions on base treated rubber (Hevea brasiliensis) leaves powder: Am. Chem. Soc. 60 (1938) 309–319.
kinetics, isotherm, and biosorption mechanisms, J. Environ. Sci. 20 (2008) [26] P. Chakravarty, N.S. Sarma, H.P. Sharma, Removal of Pb (II) from aqueous solution
1168–1176. using heartwood of Areca catechu powder, Desalination 256 (2010) 16–21.
[15] A.E. Ofomaja, E.B. Naidoo, S.J. Modise, Removal of copper(II) from aqueous [27] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum,
solution by pine and base modified pine cone powder as biosorbent, J. Hazard. J. Am. Chem. Soc. 40 (1918) 1361–1368.
Mater. 168 (2009) 909–917. [28] H.M.F. Freundlich, Over the adsorption in solution, J. Phys. Chem. 57 (1906)
[16] S. Lu, S.W. Gibb, Copper removal from wastewater using spent-grain as 385–471.
biosorbent, Bioresour. Technol. 99 (2008) 1509–1517. [29] T.W. Weber, R.K. Chakraborty, Pore and solid diffusion models for fixed bed
[17] Y.S. Ho, C.T. Huang, H.W. Huang, Equilibrium sorption isotherm for metal ions on adsorbents, J. Am. Inst. Chem. Eng. 20 (1974) 228–238.
tree fern, Process Biochem. 37 (2002) 1421–1430. [30] K. Kadirvelu, C. Namasivayan, Agricultural by-products as metal adsorbents:
[18] S.R. Shukla, R.S. Pai, Adsorption of Cu(II), Ni(II) and Zn(II) on dye loaded sorption of lead (II) from aqueous solutions onto coir-pith carbon, Environ.
groundnut shells and sawdust, Sep. Purif. Technol. 43 (2005) 1–8. Technol. 21 (2000) 1091–1097.
[19] M. Mukhopadhyay, S.B. Noronha, G.K. Suraiskhumar, Kinetic modelling for the [31] S. Lagergren, About the theory of so-called adsorption of soluble substances,
biosorption of copper by pretreated Aspergillus niger biomass, Bioresour. Technol. Kungliga Svenska Vetensk Handl. 24 (1898) 1–39.
98 (2007) 1787-1787. [32] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
[20] R. Djeribi, O. Hamdaoui, Sorption of copper(II) from aqueous solutions by cedar Biochem. 34 (1999) 451–465.
sawdust and crushed brick, Desalination 225 (2008) 95–112. [33] Y.S. Ho, G. McKay, Application of kinetic models to the sorption of copper(II) onto
[21] S. Larous, A.H. Meniai, M.B. Lehocine, Experimental study of the removal of copper peat, Adsorpt. Sci. Technol. 20 (2002) 797–815.
from aqueous solutions by adsorption using sawdust, Desalination 185 (2005) [34] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, J. Sanit.
483–490. Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 31–60.

You might also like