Sei sulla pagina 1di 11

Batch and Dynamics Modeling of the Biosorption of

Cr(VI) from Aqueous Solutions by Solid Biomass


Waste from the Biodiesel Production
Muthusamy Shanmugaprakash,a Venkatachalam Sivakumar,b Manickavelu Manimaran,a and
Jeyaseelan Aravinda
a
Department of Biotechnology, Kumaraguru College of Technology, Coimbatore 641049, India
b
Department of Food Technology, Kongu Engineering College, Perundurai, Erode 638052, India; drversusivakumar@yahoo.com
(for correspondence)
Published online 30 April 2013 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/ep.11781
ABSTRACT
Pongamia oil cake (POC), a bio-residual
waste is obtained during the production of biodiesel as a
byproduct. This is used as biosorbent to evaluate the removal
of chromium (VI) ions from an aqueous synthetic solution.
The effects of various process parameters such as pH, contact
time, initial chromium ions concentration and adsorbent
dosage have been investigated. The FT-IR and SEM analysis
of the adsorbents was done in the native- and Cr(VI)-loaded
state, to explore the position of the functional groups available for binding of chromium ions and the structural morphology of the studied adsorbents respectively. Langmuir,
Freundlich, Temkin and DubininRadushkevich isotherm
were used to study the adsorption mechanism, and it was
found that the equilibrium data was better represented by
the Freundlich isotherm. The maximal removal of hexavalent
chromium ion was found to be at a pH of 2.0 within 2 h.
The sorption kinetic follows the pseudo second order kinetic
model. The Cr(VI) ions bound to the biosorbent could be
effectively removed, using dilute H2SO4 (0.05 mM). The ability of POC to adsorb Cr(VI) ions in packed column was also
investigated through the column studies. Bed Depth Service
Time model and the Thomas model were used to analyze the
experimental data and evaluate the model parameters. POC
was shown to be a promising adsorbent for removal of Cr(VI)
C
2013 American Institute of
ions from aqueous solutions. V
Chemical Engineers Environ Prog, 33: 342352, 2014

Keywords: biosorption; pongamia oil cake; hexavalent


chromium; isotherms; kinetics; column studies
INTRODUCTION

The recent rapid industrialization in India has led to the


release of huge amounts of industrial effluents into the environment and the heavy metals such as Cr, Cu, Pd, Ni, Zn,
and so forth, present in these effluents are more stable and
persistent environmental contaminants; since they cannot be
degraded or destroyed. Thus they tend to accumulate in the
soil, ground water, seawater and sediments, which intensify
environmental pollution. Chromium have been widely used
in a variety of industrial applications such as mining, leather

C 2013 American Institute of Chemical Engineers


V

342

July 2014

tanning, cement industries, steel production units, other


metal alloys industries and so on [1]. The highest level of
Cr(VI) ions permitted in a discharge into inland surface and
potable waters, are 0.1 and 0.05 mg/L, respectively [2,3].
Chromium, especially that is present in electroplating effluents, remains in two oxidation states Cr(III) and Cr(VI), of
which the hexavalent form is more toxic, when compared to
the trivalent form. An increased risk of lung cancer is
reported among workers exposed to chromate production
environment [4]. Therefore, there is a major concern for the
elimination of such heavy metal from wastewater before disposal to the environment.
Out of the wide ranges of conventional methods which
are available for the removal of this heavy metal from aqueous solutions, include precipitation [5], oxidation-reduction
[6], ionic exchange [7], electrochemical treatment [8], and
membrane techniques [9]. However, the major drawbacks of
all these methods are that they involve high operating costs,
and may result in a large volume of solid wastes [10,11]. Biosorption is found to be an effective treatment technique
owing to its cost-effectiveness and environmental friendly
characteristics [12]. In biosorption, either metabolically active
or inactive biological material is used, to concentrate and
recover or eliminate these pollutants from effluents. Commercially available activated carbons are found be costly and
hence they have to be regenerated many times for reuse.
Reduction in activity during each stage of regeneration and
the short shelf life of activated carbon affect the economic viability of the process for real time operation. From the analysis of recent publications it was found that different
inexpensive, locally and abundantly available biosorbents
have been employed for removal of various heavy metals,
for example, sugarcane bagasse, maize corn cob, jatropha oil
cake [13], mustard oil cake [14], olivestone [15], sawdust, rice
husk [16], modified coconut coir pith [17], and so forth.
However, it is still found to be inadequate to address
these problems, more work and research are needed to identify other locally available, cheap adsorbents to eliminate
Cr(VI) ions from industrial wastewater samples with different
compositions and characteristics. Therefore, there is a need
to identify suitable sources of materials, which are locally
available in plenty with low-cost or generated as a waste

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

Table 1. Isotherm constants for the removal of Cr(VI) by POC.

Langmuir isotherm
Adsorbent
POC

Freundlich isotherm

Temkin
isotherm

D-R isotherm

Qo
(mg/g)

b
(L/mg)

R2

KF
[(mg/g)(L/mg)1/n]

R2

qd
(mg/g)

ED
(kJ/mol)

R2

A
(L/g)

R2

166.6

0.00233

0.91

0.984

1.33

0.99

78.0

0.286

0.91

33.33

0.95

Table 2. Comparison of adsorption capacity of other


adsorbents for Cr(VI) onto POC.

Adsorbents
Soya cake
Coconut shell carbon
Beech saw dust
Jatropha seed press cake
Rice husk carbon
Saw dust carbon
Modified coconut coir pith
S. platensis
C. varlgaris
POC

Adsorption
capacity
(mg/g)

pH

References

0.28
10.88
16.13
22.727
48.31
53.48
76.3
188.68
163.68
166.6

1.0
4.0
1.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0

[32]
[33]
[34]
[35]
[16]
[16]
[17]
[36]
[36]
Present
work

from any process to remove Cr(VI) ions from industrial


wastewater. In India, particularly in southern part, many
industrial plants have been installed for extraction of oil
from Pongamia seeds to prepare the biodiesel. After the
extraction of oil, the cake generated from this is usually discarded as waste material (POC). Because of the huge volume
of cake generated during oil extraction, this waste could be
used as biosorbent for removal of heavy metals due to its
potential adsorption characteristics.
In this study, an attempt has been made to explore the use
of pongamia oil cake as suitable adsorbent for the removal of
Cr(VI) ions from an aqueous system under different experimental conditions in the batch and continuous modes. The main
process parameters considered in the batch studies, were pH,
contact time, initial Cr(VI) ions concentration, and adsorbent
dose. The Bed Depth Service Time (BDST) and the Thomas
models were analyzed, using the column breakthrough data.
MATERIALS AND METHODS

Preparation of the Adsorbent and Synthetic Wastewater


Pongamia oil cake (POC) was acquired from the local pongamia processing industry. The cake was crushed into fine
particles using a mortar and pestle. It was dried in a hot air
oven at 363 K for 24 h. The dried cake was filtered through a
250 mm size sieve. The sieved material was stored in a desiccator for further adsorption studies. Synthetic wastewater was
prepared by dissolving K2Cr2O7.H2O (AR grade) crystals in
distilled water. A stock solution of 1000 mg/L was prepared,
and from this the working standards were diluted accordingly.
The pH of the solution was adjusted with 0.1 M HCl and 0.1
M NaOH, using the pH meter (ELICO LI120, India), calibrated
with buffers of pH 4.0, 7.0, and 9.2.
Batch Adsorption Studies
Batch experiments were carried out with various pH (210), metal concentration (100500 mg/L), adsorbent doses
(15 g/L) maintaining a temperature of 303 6 2 K, and a stirring speed of 150 rpm. Samples were withdrawn at specified

time intervals and centrifuged in a refrigerated centrifuge


(Kubota 3700, Japan) at 5000 rpm for 15 min. The supernatant was collected, and the amount of Cr(VI) present in the
supernatant was determined, using DPC (1,5-diphenyl carbazide). In this method, concentration of Cr(VI) was determined by developing red-violet color with DPC and
measuring the absorbance at 540 nm using a UVvis Spectrophotometer(Shimadzu UV-1800, Japan). The removal efficiency (E %) was calculated using the following equation,
E%5



Co 2Ce
3100
Co

(1)

where Co and Ce are initial and equilibrium chromium concentrations in the solution (mg/L), respectively. The adsorption capacity, qe (mg/g), of an adsorbent is the amount of
chromium adsorbed per unit mass of the adsorbent, and it
was calculated using the following equation,


Co 2Ce
3V
qe 5
M

(2)

where M is the mass of the adsorbent (g) used and V is the


volume of the solution (L).

Desorption Studies
Chemicals such as sulfuric acid (H2SO4) (0.050.5 mM),
hydrochloric acid (HCl) (0.010.1 mM) and ethylenediaminetetraacetic acid (EDTA) (0.010.1 mM), were used as a
desorbing agents for the recovery of Cr(VI) ions from the
loaded adsorbent. The Cr(VI) loaded biomass was wetted
with distilled water and added to 100 mL of desorbing solution. The flasks were incubated at 303 6 2 K in an orbital
shaker at 150 rpm.

Column Studies
It is necessary to generate the data by conducting the
fixed bed column experiments in order to understand the
adsorption characteristics, and also used to design the industrial fixed bed column. Continuous flow sorption experiments were conducted in a glass column (2 cm internal
diameter and 45-cm height) packed with required POC. A
known concentration of the metal solution was pumped into
the column at a desired flow rate using peristaltic pump
(Miclins PP 20 Ex, India). The residual concentration of
Cr(VI) ions at the outlet of the column was determined at
regular time intervals. The volume of the effluent, Vef (mL)
was calculated using the following equation
Vef 5Qttotal
where ttotal is the total
circulates through the
capacity qe (mg/g) can
batch studies, but with
given as follows

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

(3)
time(min), Q is the flow rate which
column (mL/min). The adsorption
be determined by the equation as in
slight modifications. The equation is

July 2014 343

Table 3. Kinetic parameters for the removal of Cr (VI) ions onto the by POC.

Initial conc.
of metal ion
(mg/L)
100
200
300
400
500

Pseudo I order
qe exp
(mg/g)

kI
(1/min)

qe cal
(mg/g)

R2

KII
(g/mg/min)

qe cal
(mg/g)

R2

ki
(mg/g time1/2)

R2

18.63
32.22
47.89
60.35
70.11

0.044
0.05
0.043
0.048
0.053

21.54
25.53
24.04
33.75
23.31

0.892
0.818
0.882
0..873
0.924

0.0012
0.0022
0.0027
0.0022
0.0046

22.22
34.48
50.00
62.50
71.42

0.943
0.994
0.998
0.998
0.999

1.596
2.351
3.263
4.042
4.316

0.078
6.491
14.30
18.49
28.43

0.891
0.862
0.748
0.751
0.609

Table 4. Desorption of Cr (VI) with time using different eluent.

Eluant

Desorption of
Cr(VI) (%) at

Concentration
of eluant (mM)

20 min

40 min

60 min

0.00
0.05
0.1
0.2
0.3
0.4
0.5
0.00
0.01
0.05
0.075
0.1
0.00
0.01
0.05
0.075
0.1

0.04
22.22
27.88
33.21
39.71
46.33
51.36
0.05
5.21
10.25
30.25
44.21
0.05
7.45
19.54
30.55
45.27

0.25
24.19
30.2
34.99
41.77
48.17
60.28
0.27
6.44
11.88
32.58
56.63
0.27
8.21
23.65
31.45
55.67

0.25
26.22
30.9
36.33
42.81
59.26
84.65
0.27
7.01
11.88
35.58
69.46
0.27
8.21
23.65
31.45
61.22

H2SO4

HCl

EDTA



Co 2Cb
qe 5
3Vef
M

(4)

where M is the mass of adsorbent (g), Cb is the breakthrough


concentration (mg/L) and Vef is the volume of effluent that is
required to attain the exhaustion of the column (L).
Modeling of Isotherm, Kinetic and Breakthrough
Curves
The various theoretical models for both batch and column
adsorption studies (Tables 15) were applied to find a model
adequacy to predict the isotherm, kinetic and breakthrough
data. The validity of models was evaluated based on the
regression coefficient (R2).
RESULTS AND DISCUSSION

Fourier Transforms Infrared Analysis


To determine the surface functional groups on the adsorbent during biosorption and desorption processes of
Cr(VI) onto POC, Fourier transform infrared (FT-IR) analysis
was done (Spectrum one, FT-IR Spectrometer). The spectra
of the dry adsorbents were measured within a range of 500
4000 cm21. An intense broad spectrum observed at 3411
cm21 corresponds to OAH stretching vibrations of cellulose,
pectin, absorbed water and lignin in both native and desorbed biomass was shown in Figure 1 [23]. The sharp peak
at 2989 cm21 is attributed to the CAH stretching. Bands at
344

July 2014

Intraparticle
diffusion constants

Pseudo II order

2292 and 2143 cm21 are due to the CAN stretching vibration.
A peak at 1629 cm21 is due to stretching of carbonyl (C@O)
vibrations and broad peak at 1239 cm21 is due to CAN
stretching [24].
After treating the adsorbent with Cr(VI) solutions at pH
2.0, there is disappearance, shifting and increase in peak
intensities at 3411, 2989, 2292, 2143, 1629, 1563, 1493, and
1239, 1170 cm21. The changes seen in the peak at 2989
cm21 after biosorption indicate that there is possible
involvement between symmetric or asymmetric CAH and
the symmetric stretching vibrations of CH2. This is confirmed by the reappearance of peak at 2989 cm21 after
Cr(VI) desorption from POC. The changes observed on
peaks 1629 and 1563 cm21which reflect stretching vibrations of symmetrical or asymmetric ions carboxylic groups
(ACOOH) present on the surface of POC. A slight increase
in intensity is observed at 1473 cm21 after biosorption, suggesting the coordination of CH3 groups present on the surface of metal ions. The changes in the band at 856 cm21
are indicative of CAH bending during biosorption process.
Scanning Electron Microscopy of the Adsorbents
Scanning electron microscopy (SEM) is a primary tool
for characterizing the surface morphology, and fundamental
physical properties of the given adsorbents (FEI Quanta
FEG 200 High resolution SEM). SEM micrographs of native
POC and the adsorbed POC with Cr(VI) are shown in
Figures 2(a) and 2(b). It is clear that, the POC has a considerable number of pores with high homogeneity and there
is considerable decrease in the porous nature of Cr(VI)
loaded biomass. This may be due to the significant possibility of the chromium ions being adsorbed onto the pores
and remaining trapped [25]. Also, there are shiny bulky
particles accumulated as a layer over the surface of adsorbents due to the presence of chromium ions. The specific
surface area of the POC was determined by BET method
and it is found to be 0.9 m2/g and its bulk density is given
by 0.60 kg/m3.
Effect of pH
pH is an important factor which influences the surface
properties of the adsorbent, as well as the ionic form of
the chromium ions during adsorption studies. It is seen that
the maximum chromium removal is observed at a pH of
2.0 and as pH increased there is a significant decrease in
the removal of chromium (Figure 3). This is due to the fact
that chromium exists as Cr2 O-7 , HCr4- , Cr3 O210 , Cr4 O313 , and
Cr4 O313- in the acidic pH range and HCrO4- is the most predominant. At pH 2.0, the surface of the adsorbent seems
highly protonated, which results in strong electrostatic
attractions between the protons and the chromate ions. At
a higher pH, that is, in the alkaline range, the surface of
the adsorbent will have hydroxyl groups which will not
attract the chromate ions that compete with the hydroxyl
ions [26].

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

Table 5. Effect of Bed height for Cr(VI) adsorption by POC using BDST and Thomas Model.
The BDST model parameters
Bed height (cm)
5
10
15

Sorbent weight (g)

Breakthrough time (min)

Qo (mg/g)

Exhaustion
time (min)

6.5
13
19.5

105
195
300

43.53
58.64
71.61

540
960
1440

Qo (mg/g)

R2

43.53
40.53
37.73

0.857
0.875
0.977

The Thomas model parameters


Flow rate (ml/min)
10
15
20

Bed height (cm)


5
5
5

Figure 1. (A) FT-IR of native POC, (B) FT-IR of Cr(VI)


loaded POC (C) FT-IR of Cr(VI) desorbed POC.

Effect of the Adsorbent Dose


The adsorbent dosage determines the capacity of an adsorbent for an adsorbate at a given initial concentration [27].
From Figure 4, it can be concluded that the removal efficiency increased from 48 to 81% as the adsorbent dose
increased but the amount of Cr(VI) ions adsorbed per unit
mass decreases by increasing the biosorbent concentration.
This is mainly due to an increase in the number of available
adsorption sites with an increase in biosorbent concentration
and resulting in the increase of adsorbed metal concentration. On the contrary, the adsorption qe (mg/g) capacity
decreased from 129.3 to 69.23 as the adsorbent dose
increased from 1.0 to 5.0 g/L, mainly due to the overlapping
of adsorption sites caused by the overcrowding of the
adsorbent particles or unsaturation of adsorption sites
through the adsorption reaction [28]. Another reason may be
particle interactions, such as aggregation, resulting from high
biosorbent concentration and such aggregation would lead
to decrease in total surface area of the sorbent and an
increase in diffusional path length [29]. Further increment in
biosorbent concentration from 3.0 g/L did not cause any significant improvement in adsorption. This is due to the binding of almost all Cr(VI) ions to the adsorbent and the
establishment of equilibrium between the Cr(VI) ions bound
to the sorbent and those remaining unadsorbed in the
solution.

KTh (L/mg h)
24

1 3 10
2 3 1024
2.4 3 1024

Effects of Initial Metal Ion Concentration and


Contact Time
The adsorption of chromium onto the POC for the different initial concentrations (100500 mg/L) as a function of the
contact time was studied, by keeping the other related
parameters constant. The removal of chromium increased
with an increase in the time and reached the maximal level
at 60 min with no further increase, suggesting that equilibrium has been attained. It was seen that as the metal ion
concentration increased, the removal efficiency decreased.
This is due to the fact that when the initial metal ion concentration is low, the ratio of the available surface area for
adsorption to the concentration of chromium ions is higher;
whereas in higher initial metal ion concentrations this ratio is
less and so the efficiency is also lesser. However, the adsorption capacity increased as the metal ion concentration
increases as shown in Figure 5. An increase in the initial
metal ion concentration acts as a prime force in overcoming
all the resistance to adsorption [30]. Moreover, it increases
the number of collisions between the Cr(VI) ions.

Biosorption Isotherm Studies


The adsorption isotherms are important to analyze the
equilibrium data by developing a mathematical equation, to
design and optimize an operating procedure on a large scale.
It also provides the relationship between the amount of the
metal ions adsorbed onto the surface, and the concentration
of the metal ions in the aqueous solution at equilibrium condition. The Langmuir isotherm assumes that the adsorbate
forms a monolayer around the homogenous surface of the
adsorbent, and there is no interaction between the adsorbed
molecules [31]. When the adsorbent surface is completely
saturated with the adsorbate, equilibrium is said to be
achieved. The linear form of the Langmuir model is given by
Ce
1
Ce
5
1
qe Qm b Qo

(5)

where qe is the amount of the solute adsorbed on the adsorbent surface at equilibrium (mg/g), Ce is the concentration
of the un adsorbed solute in solution (mg/L), Qm is the maximum amount of the solute adsorbed per unit mass of the
adsorbent to form a complete monolayer (mg/g) and b is a
constant related to the affinity of the binding sites (L/mg). A
plot between Cqee versus Ce as shown in Figure 6(a) is used to
calculate the constants Qm and b.
The Freundlich model is an empirical one, which assumes
that adsorption takes place on a heterogeneous surface and
also proposes multilayer sorption with interaction among the

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

July 2014 345

Figure 2. (a) SEM images of native POC (a) and (b) SEM images of Cr(VI) loaded PO.

Figure 3. Effect of pH on adsorption of Cr (VI) by POC


[Chromium concentration 5 100 mg/L; stirring speed 5 150
rpm; contact time 5 120 min; adsorbent dose 5 3 g/L, temperature 5 30 6 2 C].

Figure 5. Effect of initial metal ion concentration and contact time on Cr (VI) removal by POC [Adsorbent dose 5 3 g/
L, stirring speed 5 150 rpm, contact time 5 120 min, pH 5
2.0; temperature 5 30 6 2 C].

adsorbed molecules [32]. The logarithmic form of above


equation can be written as
1
ln qe 5ln kf 1 ln Ce
n

(6)

where KF is the Freundlich constant representing the bonding energy((mg/g)(L/mg)(1/n)) and n is a constant which is
an indicator of the adsorption favorability. Values of n > 1
favors adsorption. The values of KF and n are determined
from the plot Figure 6. (b) between lnqe versus lnCe.
The effects of some indirect adsorbate/adsorbent interaction on adsorption isotherms were studied by using isotherm
proposed by Temkin and Pyzhez [33] and it was found that
the heat of adsorption of all the molecules in the layer
decreases linearly with coverage [34]. The linearized Temkin
isotherm is represented by the equation
Figure 4. Effect of adsorbent dose on Cr (VI) removal efficiency (%) and biosorption capacity (qe) by POC [Chromium
concentration 5 500 mg/L; stirring speed 5 150 rpm; contact
time 5 120 min; pH 5 2.0; temperature 5 30 6 2 C].

346

July 2014

qe 5

RT
RT
ln A1
ln Ce
b
b

(7)

where R is the gas constant (kJ/mol), T is the absolute temperature (K), b is a constant related to the heat of adsorption, and A is the equilibrium binding constant (L/g)

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

Figure 6. Linear isotherms for Cr (VI) removal by POC at 30 6 2 C. (a) Langmuir isotherm, (b) Freundlich Isotherm, (c) Temkim Isotherm, and (d) DR isotherm.

corresponding to the maximum binding energy. The values


of the constants can be determined from the plot between qe
and lnCe in Figure 6c.
Another important equation for the analysis of the isotherm of a high degree of rectangularity is the Dubinin
Radush Kevich isotherm [35]. The DR isotherm model does
not assume a homogenous surface, and is expressed as


ln qe 5ln qD -2BD R2 T 2 ln 111 Ce 2

(8)

where qD is the theoretical saturation capacity (mg/g) and


BD is a constant related to the adsorption energy (mol2/kJ2),
R is the gas constant (J/mol/K), and T is the temperature (K).
The values of Bd and qD can be calculated from the slope
and intercept correspondingly from the plot between lnqe
and ln(1 1 1/Ce)2 as shown in Figure 6d. The constant BD is
a representative of the mean free energy ED (kJ/mol) of the
adsorption per molecule of the adsorbate. It is calculated
from the following equation
1
ED 5 p
2BD

(9)

According to the data from Table 1 it seems, that


the Freundlich model fits better with the experimental data

than the Langmuir model. The regression coefficient of


Freundlich model is found to be near unity (R2 value of
0.995). It can be seen from Table 1 that the value of ED is
0.286 kJ/mol and it is known that when ED value is lesser
than 8 kJ/mol then the adsorption process follows physical
adsorption [36]. In the present study, POC has been compared with other adsorbents based on their maximum
adsorption capacity (qm) for Cr(VI) ions (Table 2). Accordingly, this data indicates that POC could be considered as
promising adsorbent in comparison to other available
adsorbents.
Biosorption Kinetics Studies
For constructing a waste water treatment plant, it is important to predict the solute uptake rate, which determines the
residence time of the sorbate at the solidliquid interface. In
order to investigate the mechanism of adsorption, the characteristic constants of the adsorption rate are determined using a
pseudo first-order equation of Lagergren, based on solid
capacity, and pseudo second-order equation based on solid
phase adsorption, and an intraparticle diffusion model [37,38].
The mathematical values of these models are given in Table 3.
The pseudo first order and pseudo second order kinetic
models are applied to study the sorption kinetics of chromium on the POC and the extent of the uptake in the biosorption process.

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

July 2014 347

Figure 7. (a) Pseudo first order and (b) second order kinetic plot (b) for Cr (VI) ions removal onto the POC at 30 6 2 C.

The pseudo first order kinetic model was proposed by


Lagergren [39] and is expressed as
dqt
5kI qe -qt
dt

(10)

The integrating Eq. (10) between the boundary conditions t


5 0 to t 5 t, gives
lnqe -qt 5ln qe -kI t

(11)

where kI is the rate constant (min21), qe is the amount of


chromium adsorbed on the surface at equilibrium(mg/g),
and qt is the amount of chromium adsorbed on the surface
at any time t (mg/g). A plot is drawn between ln (qe 2 qt)
versus time, and the values of k1 and qe are determined from
the slope and intercept, respectively.
The pseudo second order kinetic model [40] is expressed
by the equation
t
1
1
5
1 t
qt qe2 kII qe

(12)

where kII is the pseudo second order rate constant (g/mg/


min), which can be determined from the plot between t/qt
versus t. It was seen that the qe values calculated from the
pseudo first order kinetic model do not correlate well with
the experimental qe values (Table 3). It was also noted that
the coefficients in this model are low, while comparing the
values derived from the pseudo second order kinetic model.
The calculated qe values from the pseudo second order kinetic model agree well with the experimental qe values. It is
obvious that the graph is more linear in the pseudo second
order kinetic model, and based on the above observation it
can be concluded that the pseudo second order kinetic
model fits well in this sorption study [Figures 7(a) and 7(b)].
Thus the adsorption of chromium ions onto the POC is presumably physisorption and the surface functional groups of
the adsorbent. Similar results have been reported in the biosorption of Cr (VI) onto the activated carbon derived from
agricultural waste material [41].
Weber and Morris [42] proposed the intra particle diffusion model, which describes the diffusion mechanism. The
model is expressed as
348

July 2014

Figure 8. Intraparticle diffusion plot for Cr(VI) ions removal


onto the POC at 30 6 2 C.

qt 5ki t 1=2 1C

(13)

where ki is the intraparticle diffusion rate constant ((mg/


g)(time)1/2), and C is a constant which gives an idea about
the thickness of the boundary layer, which can be calculated
from the plot qt versus t1/2. If this plot yields a linear graph
which passes through the origin, then the sorption process is
controlled only by intraparticle diffusion. It can be seen from
Figure 8 that the graph is not completely linear, and also
does not pass through the origin, suggesting that there are
other kinetic factors involvedapart from Intraparticle diffusion, that control the adsorption rate.

Desorption Studies
Desorption studies were carried out using three desorbing
agents (Table 4). Among the three agents tried, for the desorption of Cr(VI) from a laden biosorbent, 0.5 mM H2SO4
showed the maximum elution (84.65%) of Cr(VI) in 60 min,

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

Figure 9. Effect of bed height for Cr(VI) adsorption onto the


POC (flowrate 5 10 mL/min, Co 5 100 mg/L, Bed height 5
5 cm, pH 5 2.0 and temperature 5 30 6 2 C).

Figure 10. BDST model for Cr(VI) adsorption by POC.

which clearly indicates that desorption is a physic-chemical


sequester on the surface [43].
COLUMN STUDIES

Effect of The Bed Height on The Adsorption of Cr(VI)


by the POC
The effect of the bed height was studied by varying column height and by maintaining all other parameters, such as
the flow rate, pH and initial metal ion concentration and rate
constant (Figure 9). The effect of bed height is explained,
using the Bed Depth Service Time (BDST) model [44]. The
BDST model is expressed as
t5



No z
1
Co
ln
-1
Co m Ka Co
Cb

(14)

where t is the service time that can be taken as the time


(min), where the outlet concentration of the column is 5 mg/

Figure 11. The experimental and modeled breakthrough


curves from Thomas Model for the adsorption of Cr(VI) onto
the POC at different flow rates.(Co 5 100 mg/L, Bed height
5 5 cm, pH 5 2.0 and temperature 5 30 6 2 C).

L. Cb is the breakthrough time taken for the outlet concentration in the column to reach 5 mg/L. The time required to
achieve the breakthrough concentration is called as the
breakthrough time. The time required to achieve the initial
metal concentration, which is fed into the column at the outlet of the column, is called the exhaustion time [45,46]. It
means that the adsorbent in the column cannot absorb any
heavy metal beyond the exhaustion time. No is the adsorption capacity of the column (mg/L), Co is the initial metal ion
concentration which (here) is 100 mg/L, v is the velocity in
cm/min, Z is the bed height in cm and ka is a rate constant
(L/mg/min). It is observed from Figure 10, as the bed height
increases both the breakthrough and exhaustion time
increase and also it is seen that the R2 value for the BDST
plot is high, which suggests that BDST model fits well and
thus the column studies are feasible. Only when the bed
height is the maximum, the adsorption capacity is more, so it
is desired to perform column studies with the maximum bed
height. This value is given in Table 5 and it should also be
taken into account, that when the bed height increases the
process becomes time consuming.
The parameters of the BDST model No and the rate constant Ka, were calculated from the BDST plot shown in Figure 10 and were found to be 3923.4 mg/L and 5.88 3 1023
L/mg/min, respectively. The rate constant Ka, which is a
measure of the rate transfer of the solute from the fluid
phase to the solid phase, largely influences the breakthrough
phenomenon in the column study. For a smaller value of Ka,
a relatively longer bed is required to avoid a breakthrough,
whereas the breakthrough can be eliminated even in smaller
bed heights when the value of Ka is high [43,46].
Effect of the Flow Rate on the Adsorption of
Cr (VI) by POC
The flow rate is one of the important characteristics in
evaluating the sorbents for continuous treatment of the
metal-laden effluents on an industrial scale. Adsorption was
studied for various flow rates in the range of 1020 mL/min
for the initial metal ion concentration of 100 mg/L and a bed
height of 5 cm. From Figure 11 it can be observed that as
the flow rate increases both the exhaustion time and breakthrough time decrease. The uptake values were found to be

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

July 2014 349

Table 6. Adsorption and desorption efficiency of Cr(VI) ions


in column studies.
Cycle
Nos.
1
2
3
4
5

Adsorption
of Cr(VI) (%)

Desorption
of Cr(VI) (%)

98.33
87.22
81.22
66.12
40.25

82.49
80.21
75.32
70.12
65.32

CONCLUSION

Cr(VI) 5 100mg/L, pH 5 2.0, bed height 5 10 cm, flow rate


5 10 mL/min.

43.53, 40.43, and 37.73 mg/g for the flow rates 10, 15, and
20 mL/min respectively, which shows the influence of the
flow rate on the sorption capacity. The reason for the
decreased sorption capacity at a higher flow rate may be due
to the unavailability of sufficient retention time, and the limited diffusivity of the solute into the sorptive sites or pores of
the biomass.
The Thomas model is a simple and generally used model
for the prediction of the concentration versus time profile, or
the breakthrough curve for the effluent with basic assumption. The adsorption is described by a pseudo second order
reaction rate principle which reduces a Langmuir isotherm at
equilibrium; constant column void fraction; constant physical
properties of the biomass (solid-phase) and the fluid phase;
isothermal and isobaric process conditions; the intraparticle
diffusion and external resistance during the mass transfer
processes are considered to be negligible.
On the basis of the above assumption, the Thomas model
[47] can be expressed in its linear form as

ln


Co
KTH Qo M KTH Co
-1 5
F
C
F

(15)

where kTh is the Thomas model constant (L/mg h), Qo is the


maximum solid phase concentration of the solute (mg/g)


and V is the throughput volume (L). A plot of ln CCo -1
against t (where t 5 V/F) for a given flow rate is used to
determine the model constants. The Thomas model constants
are given in Table 5 and it seen that a high correlation exists
between the experimental data and the predicted value; both
the experimental and the Thomas model predicted breakthrough curves are shown in Figure 11. From Table 5, it can
be inferred that the bed capacity Qo decreases with an
increasing flow rate, while the Thomas constants KTH
increased with an increasing flow rate of 1020 ml/min.
Similar trends have been observed by other researchers for
various adsorbents [48,49].
Desorption and Column Regeneration Studies
From the economic point of view, the reusability of the adsorbent material is very important in the field of adsorption
technology. In our batch studies, the results showed that
H2SO4 (0.5 mM) has good elution capacity, compared with
other eluant (Table 4). Therefore, the 0.5 mM H2SO4 solution
was used as the eluting agent at a constant flow rate of 10
mL/min and preloaded with Cr(VI)biomass, of a bed height of
10 cm. After the recovery of the metal ion, the regenerated
POC was thoroughly washed with distilled water and again
loaded with the Cr(VI) solution of a concentration of 100 mg/
L. Five adsorptiondesorption studies were carried out repeatedly by maintaining the same operating conditions. From
350

July 2014

Table 6, it can be interpreted that the efficiency of the Cr(VI)


adsorption on the regenerated POC in the fourth and fifth
cycle, was beginning to drop. This is due to the gradual deterioration of the POC, because of continuous usage. As the
adsorption and desorption cycle proceeds, the accessibility of
the metal ions toward the binding sites get reduced [43].
The amount of metal adsorbed was found to vary with
the pH, contact time, initial chromium ions concentration
and dosage concentration. From the FT-IR analysis it was
found, that the hydroxyl and CAH groups present in the surface of the adsorbent contribute to chromium adsorption.
The adsorption equilibrium data were found to fit the
Freundlich isotherm, indicating that the surface is heterogeneous. The pseudo-second order model can be used to predict the adsorption kinetics. The breakthrough curves were
analyzed at different flow rates and bed heights. The BDST
and the Thomas Model were used to evaluate the experimental data and both the models were valid for accurate
evaluation. The Cr(VI) loaded POC was eluted with a 0.5
mM H2SO4 solution through column studies, and reused for
five adsorptiondesorption studies. The present study suggests that pongamia oil cake, an abundant agricultural waste
from biodiesel production, can be used as an adsorbent for
the removal of chromium (VI) from aqueous solutions.
ACKNOWLEDGMENTS

The authors (M.Shanmugaprakash, M.Manimaran, J. Aravind)


are thankful to the Management, Kumaraguru College of Technology, Coimbatore, for providing the facilities to carry out this
research.

LITERATURE CITED

1. ODwyer, T.F., & Hodnett, B.K. (1995). Recovery of chromium from tannery effluents using a redox-adsorption
approach. Journal of Chemical Technology and Biotechnology, 62, 3037.
2. EPA, Environmental Protection Agency, Environmental
Pollution Control Alternatives, EPA/625/5-90/025, EPA/
625/4-89/023, Cincinnati, US, 1990.
3. Indian Standard, Drinking waterspecification (first revision), IS 10500, 1991.
4. Gibb, H.J., Lees, P.S.J., Pinsky, P.F., & Rooney, B.C.
(2000). Lung cancer among workers in chromium chemical production. American Journal of Industrial Medicine,
38, 115126.
5. Zhou, X., Korenaga, T., Takahashi, T., Moriwake, T., &
Shinoda, S. (1993). A process monitoring/controlling system for the treatment of wastewater containing chromium
(VI). Water Research, 27, 10491054.
6. Seaman, J.C., Bertsch, P.M., & Schwallie, L. (1999). In situ
Cr(VI) reduction within coarse-textured, oxide-coated soil
and aquifer systems using Fe (II) solutions. Environmental Science and Technology, 33, 938944.
7. Tiravanti, G., Petruzzelli, D., & Passino, R. (1997). Pretreatment of tannery wastewaters by an ion exchange
process for Cr(III) removal and recovery. Water Science
and Technology, 36, 197207.
8. Kongsricharoern, N., & Polprasert, C. (1996). Chromium
removal by a bipolar electro-chemical precipitation process. Water Science and Technology, 34, 109116.
9. Chakravarti, A.K., Chowdhury, S.B., Chakrabarty, S.,
Chakrabarty, T., & Mukherjee, D.C. (1995). Liquid membrane multiple emulsion process of chromium(VI) separation from waste waters. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 103, 5971.

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

10. Volesky, B. (1990). Removal and recovery of heavy metals by biosorption. Biosorption of Heavy Metals (pp. 7
44), CRC Press Boca Raton, FL, USA.
11. Basso, M.C., Cerrella, E.G., & Cukierman, A.L. (2002).
Lignocellulosic materials as potential biosorbents of trace
toxic metals from wastewater. Industrial and Engineering
Chemistry Research, 41, 35803585.
12. Kim, E.J., Park, S., Hong, H.J., Choi, Y.E., & Yang, J.W.
(2011). Biosorption of chromium (Cr (III)/Cr (VI)) on the
residual microalga Nannochlorisoculata after lipid extraction for biodiesel production. Bioresource Technology,
102, 1115511160.
13. Garg, U.K., Kaur, M.P., Garg, V.K., & Sud, D. (2007). Removal of hexavalent chromium from aqueous solution by
agricultural waste biomass. Journal of Hazardous Materials, 140, 6068.
14. Khan, M.A., Ngabura, M., Choong, T.S.Y., & Masood, H.
(2012). Biosorption and desorption of Nickel on oil cake:
Batch and column studies. Bioresource Technology, 103,
3542.
15. Blazquez, G., Calero, M., Hernainz, F., Tenorio, G., &
Martn-Lara, M.A. (2011). Batch and continuous packed
column studies of chromium(III) biosorption by olive
stone. Environmental Progress and Sustainable Energy,
30, 576585.
16. Bansal, M., Singh, D., & Garg, V.K. (2009). A comparative
study for the removal of hexavalent chromium from
aqueous solution by agriculture wastes^aeTM carbons.
Journal of hazardous materials, 171, 8392.
17. Namasivayam, C., & Sureshkumar, M.V. (2008). Removal
of chromium(VI) from water and wastewater using surfactant modified coconut coir pith as a biosorbent. Bioresource Technology, 99, 22182225.
18. Daneshvar, N., Salari, D., & Aber, S. (2002). Chromium
adsorption and Cr(VI) reduction to trivalent chromium in
aqueous solutions by soya cake. Journal of Hazardous
Materials, 94, 4961.
19. Babel, S., & Kurniawan, T. A. (2004). Cr(VI) removal
from synthetic wastewater using coconut shell charcoal
and commercial activated carbon modified with oxidizing
agents and/or chitosan. Chemosphere, 54, 951967.
20. Acar, F. N., & Malkoc, E. (2004). The removal of chromium(VI) from aqueous solutions by Fagusorientalis L.
Bioresource Technology, 94, 1315.
21. Bose, A., Kavita, B., & Keharia, H. (2011). The Suitability
of Jatropha Seed Press Cake as a Biosorbent for Removal
of Hexavalent Chromium from Aqueous Solutions. Bioremediation Journal, 15, 218229.
22. Gokhale, S. V., Jyoti, K. K., & Lele, S. S. (2008). Kinetic
and equilibrium modeling of chromium (VI) biosorption
on fresh and spent Spirulinaplatensis Chlorella vulgaris
biomass. Bioresource Technology, 99, 36003608.
23. Feng, N., Guo, X., Liang, S., Zhu, Y., & Liu, J. (2011). Biosorption of heavy metals from aqueous solutions by
chemically modified orange peel. Journal of Hazardous
Materials, 185, 4954.
24. Nagpal, U.M.K., Bankar, A.V., Pawar, N.J., Kapadnis, B.P.,
& Zinjarde, S.S. (2011). Equilibrium and Kinetic Studies
on Biosorption of Heavy Metals by Leaf Powder of Paper
Mulberry (Broussonetiapapyrifera). Water, Air, and Soil
Pollution, 215, 177188.
25. Bansal, M., Garg, U., Singh, D., & Garg, V.K. (2009). Removal of Cr(VI) from aqueous solutions using pre-consumer processing agricultural waste: A case study of rice
husk. Journal of Hazardous Materials, 162,312320.
26. Vankar, P. S., Sarswat, R., & Sahu, R. (2012). Biosorption of
zinc ions from aqueous solutions onto natural dye waste of
Hibiscus rosa sinensis: Thermodynamic and kinetic studies.
Environmental Progress and Sustainable Energy, 31, 8999.

27. Bansal, M., Garg, U., Singh, D., & Garg, V.K. (2009). Removal of Cr(VI) from aqueous solutions using pre-consumer processing agricultural waste: A case study of rice
husk. Journal of Hazardous Materials, 162, 312320.
28. Bulut, Y., & Aydin, H. (2006). A kinetics and thermodynamics study of methylene blue adsorption on wheat
shells. Desalination, 194, 259267.
29. Shukla, A., Zhang, Y., Dubey, P., Margrave, J.L., & Shukla,
S.S. (2002).The role of sawdust in the removal of
unwanted materials from water. Journal of Hazardous
Materials, 95, 13752.
30. Kumar, R., Bishnoi, N.R., & Bishnoi, K. (2008). Biosorption of chromium(VI) from aqueous solution and electroplating wastewater using fungal biomass. Chemical
Engineering Journal, 135, 202208.
31. Langmuir, I. (1918). The adsorption of gases on plane
surfaces of glass, mica and platinum. Journal of the
American Chemical Society, 40, 13611403.
32. Freundlich, H.M.F. (1906). Over the adsorption in solution. Journal of Physical Chemistry, 57, 385470.
33. Temkin, M.I., & Pyzhev, V. (1940). Kinetics of ammonia
synthesis on promoted iron catalysts. Acta Physiochimica
URSS, 12, 327356.
34. Kavitha, D., & Namasivayam, C. (2007). Experimental
and kinetic studies on methylene blue adsorption by coir
pith carbon. Bioresource Technology, 98, 1421.
35. Dubinin, M.M., & Radushkevich, L.V. (1947). Equation of
the characteristic curve of activated charcoal. Chem.
Zentr, 1, 875.
36. Hasany, S.M., & Chaudhary, M.H. (1996). Sorption potential of
Haro river sand for the removal of antimony from acidic aqueous solution. Applied Radiation and Isotopes, 47, 467471.
37. Arami, M., Limaee, N.Y., Mahmoodi, N.M., & Tabrizi, N.S.
(2006). Equilibrium and kinetics studies for the adsorption
of direct and acid dyes from aqueous solution by soy
meal hull. Journal of Hazardous Materials, 135, 171179.
38. Namasivayam, C., & Sumithra, S. (2005). Removal of direct
red 12B and methylene blue from water by adsorption
onto Fe (III)/Cr (III) hydroxide, an industrial solid waste.
Journal of Environmental Management, 74, 207215.
39. Lagergren, S. (1898). Zurtheorie der sogenannten adsorption gel
osterstoffe, KungligaSvenskaVetenskapsakademiens. Handlingar, 24, 139.
40. Ho, Y.S., & McKay, G. (1999). Pseudo-second order model
for sorption processes. Process Biochemistry, 34, 451465.
41. Mohan, D., Singh, K.P., & Singh, V.K. (2005). Removal of
hexavalent chromium from aqueous solution using lowcost activated carbons derived from agricultural waste
materials and activated carbon fabric cloth. Industrial and
Engineering Chemistry Research, 44, 10271042.
42. Weber, W.J., Jr., & Moris, J.C. (1963). Kinetics of adsorption on carbon from solution, Journal of the Sanitary Engineering Division, 89, 3160.
43. Suksabye, P., Thiravetyan, P., & Nakbanpote, W. (2008).
Column study of chromium (VI) adsorption from electroplating industry by coconut coir pith. Journal of Hazardous Materials, 160, 5662.
44. Bohart, G.S., & Adams, E.Q. (1920). Some aspects of the
behavior of charcoal with respect to chlorine. 1. Journal
of the American Chemical Society, 42, 523544.
45. Faust, S.D., & Aly, O.M. (1987). Adsorption processes for
water treatment, Butterworth: Boston, MA.
46. Taty-Costodes, V.C., Fauduet, H., Porte, C., & Ho, Y.S.
(2005). Removal of lead (II) ions from synthetic and real
effluents using immobilized Pinussylvestris sawdust:
Adsorption on a fixed-bed column. Journal of Hazardous
Materials, 123, 135144.

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

July 2014 351

47. Thomas, H.C. (1944). Heterogeneous ion exchange in a


flowing system. Journal of the American Chemical Society
66, 16641666.
48. Senthilkumar, R., Vijayaraghavan, K., Jegan, J., & Velan,
M. (2010). Batch and column removal of total chromium
from aqueous solution using Sargassumpolycystum.

352

July 2014

Environmental Progress and Sustainable Energy, 29, 334


341.
49. Mishra, V., Balomajumder, C., & Agarwal, V.K. (2012). Simultaneous adsorption and bioaccumulation: A study on
continuous mass transfer in column reactor. Environmental Progress & Sustainable Energy. doi:10.1002/ep.11671.

Environmental Progress & Sustainable Energy (Vol.33, No.2) DOI 10.1002/ep

Potrebbero piacerti anche