You are on page 1of 10

Separation and Purification Technology 45 (2005) 131140

Kinetic and equilibrium modelling of lead(II) sorption from water and


wastewater by polymerized banana stem in a batch reactor
B.F. Noeline, D.M. Manohar, T.S. Anirudhan
Department of Chemistry, University of Kerala, Kariavattom, Trivandrum 695581, India
Received 31 August 2004; received in revised form 8 March 2005; accepted 10 March 2005

Abstract
The aim of this research work was a kinetic and equilibrium study of the sorption of lead(II) ions from water and wastewater by formaldehyde
polymerized banana stem containing sulphonic acid groups. The adsorbent was characterized using surface area analyzer, infrared spectroscopy
and scanning electron microscopy measurements. The surface charge and the acid groups of the adsorbent were determined using potentiometric
and acidbase titrations, respectively. Batch experiments were performed under kinetic and equilibrium conditions. The optimum pH range
for the maximum removal of lead(II) was 59. The maximum adsorption of 98.5 and 89.9% took place for an initial concentration of 10
and 25 mg/l, respectively, at pH 6.0. The sorption process occurred in two stages: external mass transport occurred in the early stage and
intraparticular diffusion occurred in the long-term stage. The diffusion coefficients, energies of activation and entropies of activation for both
processes were calculated to determine the theoretical behaviour of the sorption process. In the external mass transfer process, the diffusion
coefficient increases with increasing initial concentration while in the intraparticle diffusion process, the diffusion coefficient decreases
with increasing initial concentration. The temperature dependence indicates the endothermic nature of adsorption process. The Langmuir,
Freundlich and RedlichPeterson isotherm models were tried to represent the equilibrium data of lead(II) adsorption. The data fitted very
well to the Freundlich isotherm model in the studied concentration range of lead(II) adsorption. Quantitative removal of 10.0 mg/l lead(II) in
50 ml of battery manufacturing wastewater by 125 mg of the adsorbent was observed at pH 6.0. The adsorbent was suitable for repeated use
(for more than four cycles) without noticeable loss of capacity.
2005 Published by Elsevier B.V.
Keywords: Adsorption; Lead(II) removal; Polymerized banana stem; Kinetics; Mass transport; Isotherm constants

1. Introduction
Lead(II) is a well known highly toxic metal considered as
a priority pollutant. It directly enters the water bodies through
the effluent discharges including battery, paper and pulp, mining, electroplating, lead smelting and metallurgical finishing
industries and caused a marked increase in concentrations. In
most countries, lead(II) level of water is limited with the value
of 0.05 mg/l. It causes various types of acute and chronic
disorders. To remove lead(II) from wastewater, conventional
methods such as chemical precipitation, ion exchange, solvent extraction, membrane process and adsorption by activated carbon can be used. The removal of toxic heavy metals

Corresponding author. Tel.: +91 471 2418782.


E-mail address: tsani@rediffmail.com (T.S. Anirudhan).

1383-5866/$ see front matter 2005 Published by Elsevier B.V.


doi:10.1016/j.seppur.2005.03.004

at very low concentrations from water can be readily accomplished by adsorption method. Adsorption process has many
advantages over other methods including recovery of metal
value, selectivity, sludge free operation, cost effectiveness
and meeting of strict discharge specifications.
A number of adsorbents such as activated carbon [1],
sargassum [2], chitosan [3], metal oxide gel [4], saw dust
[5], humusboehmite complex [6], animal bone powder and
ceramics [7] have been used for lead(II) removal. Among
these materials, agricultural by-products and biomass showed
very high adsorption capacities. However, the applicability of these materials has been found to be limited due to
leaching of organic substances such as lignin, tannin, pectin
and cellulose into the solution. To overcome such problems,
chemical treatment on solid adsorbents has been used as a
technique for improving physical and chemical properties of

132

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140

them and to increase their adsorption capacity. Banana stem


is another commonly available and abundant natural material. Besides its utility in preparing banana fibre with good
strength and lustre, any attempt to find a better alternative
for the use of banana stem would be another milestone in our
march towards economic development. In the present study, a
new adsorbent material prepared from banana stem has been
employed for the removal of lead(II) from wastewater. The
main objective of this preliminary study is to investigate the
feasibility of using formaldehyde polymerized banana stem
with sulphonic acid functionality in removing lead(II) from
water and wastewater.

2. Experimental procedure

using the pressed disc technique in a Perkin-Elmer IR-180


spectrophotometer. The thermal characteristics of the adsorbents were examined by TG peaks using a Metler Toledo
Thermoflex instrument. Surface morphology of the adsorbents was probed by scanning electron microscope (SEM)
using a Philips XL 30 CP scanning electron microscope.
The cation exchange capacity was determined by Na saturation method [10]. The concentration of Na in the solution was determined using a GBC Avanta A5450 (Australia) atomic absorption spectrophotometer. A potentiometric method [11] was used to determine the zero point charge
(pHzpc ). The acidic groups present in the adsorbents were
estimated using the method proposed by Boehm and Voll
[12]. A systronic microprocessor pH meter (model, 362,
India) was used to measure the potential and pH of the
suspension.

2.1. Reagents
2.4. Batch adsorption experiments
All the chemicals used to prepare reagent solutions were
of analytical reagent grade. The stock solution of lead(II)
(1000 mg/l) was prepared by dissolving a weighed quantity
of the chloride salt in distilled water. Lead(II) solutions having concentrations 10400 mg/l were prepared and used for
adsorption experiments. The pH of the adsorbentlead(II)
system was adjusted by using 0.1 M NaOH or 0.1 M HCl.
Constant ionic strength (0.01 M NaCl) was used in all experiments. The formaldehyde procured from Ranbaxy, India,
was used to prepare the adsorbent.
2.2. Adsorbent
The pseudo stem of Musa paradisiaca L. was used for the
preparation of the adsorbent. The material was washed several times with distilled water to remove surface impurities
and dried at 80 C. Banana stem (BS) basically contains cellulose, hemicellulose and lignin which were determined
using the method described by Ott [8] and were found
to be 43.3, 20.6 and 27.8%, respectively. The dried sample was ground in order to increase the surface area and
sieved to 80 + 230 mesh size. Two parts of the BS powder was treated with 20 parts of 0.2N H2 SO4 and five parts
of 39% HCHO. The reaction mixture was then kept in an air
oven at 50 C for 6 h and occasionally stirred. The product,
formaldehyde polymerized BS (FPBS), was washed several
times with distilled water and dried at 60 C. The polymer content of the FPBS was determined using standard
method [9] and was found to be 33.8%. The dried sample
of FPBS was sieved and the fraction with average particle
diameter of 0.096 mm was collected and used for all the
experiments.
2.3. Instrument used for adsorbent characterization
The specific surface area of BS and FPBS was measured by
BET N2 adsorption using Quantasorb Surface area analyzer
(QS/7). The FTIR spectra of the adsorbents were obtained

Batch experiments were conducted to determine the pH


range at which the maximum adsorption of Pb(II) would
take place on the FPBS. To a series of 100 ml flasks, each
containing 0.1 g of the adsorbent, added 50 ml aqueous Pb(II)
solution of desired concentration. The initial pH was adjusted
to values ranging from 2.0 to 9.0 by using 0.1 M NaOH and
HCl. The suspension was shaken at 200 rpm for 1 h using a
temperature controlled flask shaker. Preliminary experiments
showed that adsorption process studied was completed after
1 h. The contents of the flask were filtered through filter paper
and the filtrate was analyzed for final pH and final Pb(II)
concentrations. A GBC Avanta A5450 (Australia) atomic
absorption spectrophotometer equipped with a flame atomizer was used to determine the concentration of Pb(II) in the
filtrate. The amount of adsorption (qe ) was calculated by the
following equation:
C0 Ce
V
(1)
m
where C0 and Ce are the initial and equilibrium Pb(II) concentrations, respectively. V is the volume of the solution and
m is the amount of adsorbent used.
The kinetic experiments were conducted in the concentration range between 10 and 100 mg/l and the contact time
was varied from 1 to 60 min. The pH of the solution was
adjusted to an optimum pH (6.0) from the earlier study.
Aliquots of supernatant were withdrawn at different time
intervals and the amount of Pb(II) ions in the solution was
estimated. Adsorption isotherm experiments were also performed by agitating 0.1 g of the adsorbent with 50 ml of the
varying concentrations of Pb(II) (10400 mg/l) at different
temperatures (3060 C). The initial pH of the suspension
was adjusted to 6.0. After the established contact time (1 h)
was reached, aliquots of supernatants were withdrawn and the
amount of Pb(II) in the solution was estimated. All the experiments were carried out in duplicate and the mean values are
presented.
qe =

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140

133

2.5. Desorption experiments


For any adsorption process, the most important factors are
the recovery of the adsorbent and the regeneration capacity of
the adsorbent. Desorption studies were carried out as follows:
after adsorption experiments with 10 mg/l Pb(II) at 0.1 g of
adsorbent in 50 ml, the Pb(II) loaded samples were separated
by filtration and washed with distilled water to remove any
unadsorbed Pb(II). The adsorbed Pb(II) was eluted with 50 ml
of 0.1 M HCl. After mixing for 4 h, the suspension was filtered
and the filtrate was analyzed for Pb(II). A comparison of the
value with those observed in the initial sorption step was
used to compute the percentage recovery values. The sorbent
sample thus regenerated was reused for adsorption purposes.
The loading and regeneration cycles were repeated four times.

3. Results and discussion


3.1. Adsorbent characterization
The FTIR spectra of BS and FPBS (4004000 cm1 ) are
plotted in Fig. 1. The strong asymmetric absorption band at
3477 cm1 for BS is attributable to the sum of the contribution
from adsorbed water and hydroxyl groups from polyphenols
originally present in BS. The IR exhibits weak absorption
peak characteristic of the C H stretching vibrations of the
cellulose and hemicellulose. The strong bands at 1519 and
1209 cm1 are due to the aromatic C O stretching vibrations
of the lignin component and C O stretching, respectively.
The band characteristic to -glycosidic linkage appears at
885 cm1 . The FTIR spectrum of FPBS shows a broad peak

Fig. 1. FTIR spectra of BS and FPBS.

Fig. 2. Scanning electron micrographs of BS (A) and FPBS (B).

centered at 3400 cm1 is attributed to the overlapped bands


arising from sulphonic acid hydroxyl and free phenolic and
alcoholic hydroxyl groups in the BS matrix. Additional peaks
at 1169 cm1 (as SO2 ), 1027 cm1 (s SO2 ) and 602 cm1
(s S O) show the presence of SO3 H groups in the FPBS.
The peak at 1742 cm1 (C O ) along with another band at
1458 cm1 (C O ) indicates the presence of carboxylic acid
groups in the FPBS.
The SEM micrographs of BS and FPBS, taken at magnitude 500, are presented in Fig. 2. The surface morphology
of BS is different from that of FPBS. FPBS has less void
space than BS, which is consistent with the FPBS sample having smaller pores, and therefore more specific surface area.
The FPBS had a rather granular structure, whereas BS had a
rigid and folded appearance. The SEM image of FPBS clearly
shows that formaldehyde polymerized more on the surface of
the unit cell than in the intercellular gaps.
The TG analysis of both BS and FPBS was done in order
to study the thermal stability of the FPBS. Fig. 3 shows the
TG curves of BS and FPBS. The TG curves indicate the better thermal stability of the FPBS with respect to BS. The
decomposition temperature, TD , for BS and FPBS was found
to be 268 and 312 C, respectively. The weight loss of FPBS
was relatively lower than BS. TG curves exhibit weight loss of
22.6% for BS starting at 40 C and ending at 240 C, whereas
for FPBS a weight loss of 13.0% starting at 60 C and ending at 280 C. The temperature for 10.0% weight loss (T10 )
is the basic criterion used to indicate the thermal stability of

134

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140


Table 1
Effect of biomass pre-treatment on Pb(II) adsorption after 1 h of contact time
(C0 = 10 mg/l; adsorbent dose: 2 g/l)
Adsorbent

COD (mg O2 /l)

Removal (%) of Pb(II)

BS
FPBS

37.4
9.3

62.1
98.6

3.2. Effect of biomass pre-treatment on Pb(II)


adsorption

Fig. 3. TG curves of BS and FPBS.

polymers. The higher the value of T10 , the higher the thermal
stability of the system. The values of T10 were found to be
150 and 200 C for BS and FPBS, respectively.
The surface charge density o was determined by potentiometric method using the following equation:
o = F

(CA CB ) + {OH } {H+ }


A

(2)

where F is the Faraday constant, A the surface area of the


suspension (cm2 /l), CA and CB the concentrations of the acid
and base (eq/l) after each addition during the titration and
[OH] and [H]+ are the ions bound to the suspension surface
(eq/cm2 ). The point of intersection of o versus pH curve
(Fig. 4) showed that the pHzpc of BS and FPBS occurred at
6.6 and 5.0, respectively. The decrease in pHzpc after surface
modification indicates that the surface became more negative after polymerization and it facilitates the electrostatic
interaction with the metal ions. The surface area obtained
from the N2 adsorption isotherm was found to be 108.54
and 214.13 m2 /g for BS and FPBS, respectively. The total
acidity and cation exchange capacity were found to be 2.06
and 0.67 meq/g for BS and 3.12 and 0.93 meq/g for FPBS,
respectively. The apparent density was determined by the
pyknometer method using nitrobenzene as displacing liquid
and was found to be 1.10 and 1.45 g/ml for BS and FPBS,
respectively.

Fig. 4. Potentiometric curves depicting the surface charge density ( o ) as a


function of solution pH.

The adsorption efficiency of BS and FPBS on to Pb(II)


removal was examined by conducting batch experiments
using 50 ml of 10.0 mg/l Pb(II) ions and the results are presented in Table 1. The results clearly show that FPBS is 1.6
times more effective than BS for the removal of Pb(II). This
may be due to the moderate ion exchange capacity of FPBS
compared to BS. Formaldehyde treatment favours the stabilization of the organic substances of BS. Chemical treatment
induces a stabilization of the hydrolysable compounds of BS
by creating covalent bonds on the constitutive units, and make
the BS able to adsorb metal cations without an increase in
chemical oxygen demand (COD) that is due to the release of
the polyphenolic compounds.
3.3. Effect of pH on Pb(II) removal
The effect of pH on the removal of Pb(II) was examined
because the wastewaters from the battery manufacturing and
electroplating industries were of various pH values. Therefore, the effect of pH on the removal of Pb(II) was investigated
over the pH range 2.09.0 and the results are shown in Fig. 5.
For comparison, Pb(II) removal by hydroxide precipitation
as a function of pH with no adsorbent is also given in Fig. 5.
Precipitation curve shows a sharp decrease in concentration
of Pb(II) ions in solution which suggests that Pb(II) is precipitating from solution at this concentration, well before
adsorption is complete. However, at any pH Pb(II) removal
by adsorption is greater than by hydroxide precipitation. The
percentage removal of lead(II) was maximum at pH range
5.09.0. Lower adsorption of lead(II), a cationic species, at

Fig. 5. The effect of pH on the adsorption of lead(II) onto FPBS.

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140

135

highly acidic pH is probably due to the presence of excess


H+ ions competing with the lead(II) cations for the adsorption sites. The perusal of the literature on lead(II) speciation
diagram [13] shows that Pb2+ , Pb(OH)+ and Pb(OH)2 are the
major components for pH range of 2.07.0. The pHzpc value
of FPBS was 5.0. This result indicates that at pH < 5.0, the
FPBS is positively charged and major species of Pb(II) is
presented as Pb2+ . Under such condition, Pb(II) adsorption
must occur by ion exchange mechanism. FPBS is a cation
exchanger with sulphonic acid and carboxylic acid groups. It
has been shown that the final pH is always less than the initial
pH. The final pH of the reaction mixture remained between
1.7 and 3.9 and 1.6 and 3.4 for 10 and 25 mg/l Pb(II) solutions, respectively, during the experiments when the initial
pH varies between 2 and 5. This indicates that as the Pb2+
ions are bound on the FPBS, H+ ions from the peripheral
SO3 H and COOH groups are released into the solution
and it leads to the conclusion that FPBS probably acts as an
acid-form ion exchanger. The removal of lead(II) below pH
5.0 may be represented as:

in approximately 60 min. A further increase in contact time


had a negligible effect on the amount of adsorption. The
equilibrium time was found to be independent of the initial
concentration. According to these results, the agitation time
was fixed at 1 h for the rest of the batch experiments to make
sure that the equilibrium was reached.
The higher the initial concentration of Pb(II), the larger the
amount of Pb(II) adsorption. The increase in uptake capacity of the FPBS with increasing Pb(II) concentration may
be due to higher probability of collision between Pb(II) ions
and adsorbent particles. The variation in the extent of adsorption may also be due to the fact that initially all sites on the
surface of adsorbent were vacant and the solute concentration gradient was relatively high. Consequently, the extent
of Pb(II) uptake decreases significantly with the increase of
contact time, depending on the decrease in the number of
vacant sites on the surface of adsorbent.

2H FPBS + Pb2+  (FPBS)2 Pb + 2H+

(3)

H FPBS + Pb(OH)+  FPBS Pb(OH)+ + H+

(4)

The study of adsorption kinetics is quite significant in


wastewater treatment as it describes the solute uptake rate,
which in turn controls the residence time of solute uptake
at the solidsolution interface. The kinetic reaction for the
adsorption of Pb(II) was studied using a mass transport diffusion model developed by Crank [14]. This model, as an application of the Fickiens laws, expresses the relation between
solute uptake and sorption time as


Mt
D 1/2 1/2
=6
t
(as t  t for smaller times) (5)
M
r 2

As pH increases to the range 5.57.0, where Pb(OH)+ and


Pb(OH)2 species are present in the solution, FPBS surface
starts to acquire a net negative charge, creating a situation
electrostatically favourable for higher adsorption of Pb(II)
ions. The slight increase in adsorption at higher pH 8.0 may
be due to the retention of Pb(OH)2 species in the micropores
of the sorbent particles. The maximum adsorption of 98.5
and 89.9% took place by FPBS, respectively, from an initial
concentration of 10 and 25 mg/l at pH 6.0.
3.4. Effect of initial concentration and contact time
Fig. 6 shows the effect of contact time on batch adsorption
of 10100 mg/l at 30 C. The amount of adsorption sharply
increases with time in the initial stage (010 min range),
and then gradually increases to reach an equilibrium value

Fig. 6. The variation of the amount of adsorption of Pb(II) with time for
different concentrations at 30 C.

3.5. Adsorption kinetic modelling

Mt
ln 1
M

6
= ln 2 +

D2
r2


t

(as t t for larger times)

(6)

where Mt and M are the weight uptake at time t and equilibrium, respectively, D the diffusion coefficient and r is the
particle size radius assuming spherical geometry. The values of diffusion coefficients D1 and D2 for smaller and larger
times were determined from the plots of Mt /M versus t1/2 and
ln [1 (Mt /M )] versus t, respectively, at different concentrations and temperatures (Figs. 7 and 8) and the results are
presented in Table 2. It is obvious from Figs. 7 and 8 that the
Pb(II) adsorption occurs in two stages, i.e. the external mass
transport process in the early stage and intraparticular diffusion process in the long-term stage. The values of D1 were
calculated from the slope of the first straight line (010 min
range) in the plot of Mt /M versus t1/2 and D2 from the slope
of the second straight line (1560 min) range in the plot of ln
[1 (Mt /M )] versus t using least squares method.
When the initial concentration increases from 10 to
100 mg/l, the value of D1 increases from 1.42 108 to
2.12 108 cm2 /s, whereas the value of D2 decreases from
3.22 109 to 2.78 109 cm2 /s. The results indicate that
diffusion rate in the external mass transport process is more

136

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140


Table 2
The values of diffusion coefficients (D1 and D2 ) and correlation coefficients
(R21 and R22 ) at various initial concentrations and temperatures
R21

D2 (cm2 /s)

R22

Initial concentration (mg/l)


10
1.42 108
25
1.72 108
50
1.98 108
100
2.12 108

0.981
0.988
0.995
0.992

3.22 109
3.10 109
2.99 109
2.78 109

0.985
0.981
0.984
0.994

Temperature ( C)
30
2.12 108
40
2.30 108
50
2.47 108
60
2.78 108

0.992
0.991
0.991
0.997

2.78 109
3.01 109
3.40 109
3.87 109

0.994
0.984
0.986
0.991

D1 (cm2 /s)

Fig. 7. The plots of Mt /M vs. time1/2 at different (A) initial concentrations


and (B) temperatures.

concentration dependent than in the intraparticle diffusion


process. In the external mass transport process the diffusion
rate of ions is controlled by the diffusion of ions in aqueous
solutions, whereas in the intraparticle diffusion process, the
diffusion rate is controlled by the diffusion of ions with in
the adsorbent.
The value of D1 increases from 2.12 108 to
2.78 108 cm2 /s and the value of D2 increases from
2.78 109 to 3.87 109 cm2 /s when the temperature

increases from 30 to 60 C. The ion diffusion in both


processes is temperature dependent and also indicates the
endothermic nature of adsorption process. The temperature
dependence of Pb(II) diffusion in both the processes is about
the same. Michelson et al. [15] reported that the diffusion coefficient values should be in the range of 106 to
108 cm2 /s for film diffusion (external mass transport process) and 1011 to 1013 cm2 /s for intraparticle diffusion.
Since the values obtained for the D1 are in the order of
108 cm2 /s, it may be concluded that the process of Pb(II)
adsorption onto FPBS is controlled by external mass transport process occur in the earlier stage.
The linear relationship between ln D and 1/T (regression
analysis) enables the calculation of the activation energy, Ea ,
and pre-exponential factor, D0 , for diffusion using Arrhenius
equation:


Ea
D = D0 exp
(7)
RT
The values of Ea1 and Ea2 for early stage and long-term stage
were calculated from the slope of ln D versus 1/T plot (not
shown) using regression analysis and were found to be 7.32
and 9.23 kJ/mol, respectively. The results indicate that the
external mass transport and intraparticle diffusion processes
are endothermic in nature. The values of Ea seems to be small
and the rate of diffusion processes is not very sensitive to
temperature in the range we studied. The values of D0 for
both stages were calculated from the intercept of the plots
and were found to be 3.84 107 cm2 /s for early stage and
10.70 108 cm2 /s for long-term stage.
The substitution of D0 value in the following equation give
the entropy of activation (S# ):




S #
2 kT
D0 = 2.73d
exp
(8)
h
R

Fig. 8. The plots of ln [1 (Mt /M )] vs. t at different (A) initial concentrations and (B) temperatures.

where d is the distance between the adjacent exchanging


sites in the exchanger [16] which is assumed to be equal to
5 108 cm, k the Boltzmann constant, h the Plancks constant and T is taken as 273 K. The negative values of S#
(95.85 kJ/mol for early stage and 144.72 kJ/mol for longterm stage) indicate a greater order of reaction during the

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140

137

Table 3
Isotherm constants and normalized standard deviation (q, %) for the
adsorption of lead(II) onto FPBS at 30 C
Isotherm constants

Magnitude

Langmuir
Qo (mg/g)
b (l/mg)
R2
q (%)

91.74
0.04
0.974
29.4

Freundlich
KF
1/n
R2
q (%)
Fig. 9. Comparison of the experimental (legends) and the model plots of
Langmuir, Freundlich and RedlichPeterson (lines) isotherm for the adsorption of lead(II) onto FPBS at 30 C.

adsorption of Pb(II) onto FPBS and also reflect that no significant change occurs in the internal structure of the adsorbent
material during adsorption. The negative values of S# also
indicate the presence of more active sites in the adsorbent.
3.6. Adsorption isotherm
Experimental isotherm determined at 30 C is presented in
Fig. 9. It can be seen that the adsorption isotherm exhibits Hshape, which corresponds to the classification of Giles et al.
[17]. The H-type isotherm is characteristic of the cases of high
affinity of solute for an adsorbent. Such isotherm indicates
that there is no competition from the solvent for adsorption sites. Several isotherm equations have been used for
the equilibrium modelling of sorption. Out of these isotherm
equations, three have been applied for this study, the Langmuir, Freundlich and RedlichPeterson isotherms, which are,
respectively, reported in the following:
Langmuir :

Ce
1
Ce
= o + o
qe
Q b Q

Freundlich : log qe = log KF +

(9)
1
log Ce
n

(10)


Ce
1
Redlich Peterson : log KR
qe


= bR log Ce + log aR

(11)

where qe is the equilibrium amount adsorbed (mg/g), Ce


the equilibrium concentration of the adsorbate (mg/l), Qo
and b are the Langmuir constants related to the adsorption capacity and energy of adsorption, respectively. KF and
1/n are Freundlich constants related to adsorption capacity
and intensity of adsorption, respectively. KR , bR and aR are
the RedlichPeterson constants. The estimated parameters of
these models have been evaluated by regression analysis and
the results are shown in Table 3.

RedlichPeterson
KR
bR
aR
R2
q (%)

8.00
0.46
0.991
9.4
3.58
0.99
0.04
0.990
27.3

In order to compare the validity of isotherm equations


more definitely a normalized standard deviation, q (%) is
calculated as follows:

 exp
exp 2
[(qe qecal )/qe ]
q (%) = 100
(12)
N 1
where the superscripts exp and cal show the experimental
and calculated values and N is the number of measurements.
The validity of the isotherm models is tested by comparing
the experimental and calculated data at 30 C (Fig. 9). q
(%) values given in Table 3 attest the applicability of the
Freundlich model. The Langmuir values fit well to experimental results only for limited range of concentration. Values
of 0.1 < 1/n < 1.0 show the favourable adsorption of lead(II)
on FPBS [18]. The Freundlich isotherm was developed to
describe heterogeneous surface isotherms. The most actively
energetic sites are occupied first, and thereafter the surface
is occupied until the lowest energy sites are filled at the final
stage of the adsorption process.
In order to justify the validity of a treatment process, the
adsorptive capacity of the FPBS must be compared with other
adsorbents examined for the treatment of Pb(II) under similar conditions. The ultimate adsorption capacity of FPBS can
be calculated from the isothermal data by substituting the
required equilibrium concentration in the Freundlich equation. Thus, for an equilibrium concentration of 1 mg/l, each
gram of FPBS can remove 8.0 mg (KF ) of lead(II) at 30 C.
The reported KF values for the adsorption of Pb(II) were
6.32, 0.50 and 1.40 for humusboehmite [19], chitosan [20]
and modified biomass P. chrysosporium [21], respectively.
The value of maximum monolayer adsorption capacity, Qo ,
of the FPBS used in the present study at 30 C was found
to be 91.74 mg/g, which is very much higher than the values
reported in the literature. The Qo values for the adsorption of
Pb(II) on surface modified carbon [22], marine algae [23] and
surface modified saw dust [24] were found to be 54.1, 85.7

138

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140


Table 5
Composition of Pb(II) containing real battery manufacturing wastewater and
the corresponding simulated wastewater

Fig. 10. Comparison of the experimental (legends) and the model plots of
Langmuir, Freundlich and RedlichPeterson (lines) isotherm for the adsorption of lead(II) onto Duolite C-26 ion exchange resin at 30 C.

and 22.2 mg/g, respectively. Comparison of Pb(II) adsorption onto FPBS with the literature data indicates that the
adsorption capacity of FPBS is very much greater than other
adsorbents.
3.7. Comparison with commercial ion exchanger
Adsorption experiments were also performed using a commercial ion exchanger (Duolite C-26 ion exchanger) having
sulphonic acid functionality. The experimental isotherm data
for the adsorption of Pb(II) by Duolite C-26 ion exchanger
at 30 C and pH 6.0 are shown in Fig. 10. The Langmuir,
Freundlich and RedlichPeterson isotherm constants for the
adsorption of Pb(II) were calculated by regression analysis
and are summarized in Table 4. In order to compare the different isotherms and their ability to fit with the experimental
data, the computed data of these isotherms are shown with
experimental data in Fig. 10. The adsorption data of Pb(II)
on Duolite were found to fit well with Freundlich equation
Table 4
Isotherm constants and normalized standard deviation (q, %) for the
adsorption of lead(II) onto Duolite C-26 ion exchanger at 30 C
Isotherm constants

Magnitude

Langmuir
Qo (mg/g)
b (l/mg)
R2
q (%)

138.89
0.05
0.983
18.7

Freundlich
KF
1/n
R2
q (%)

12.11
0.53
0.991
8.2

RedlichPeterson
KR
bR
aR
R2
q (%)

7.07
0.90
0.08
0.990
17.7

Species

Real wastewater
(mg/l)

Simulated wastewater (mg/l)


Sample 1

Sample 2

Pb(II)
Ca(II)
Mg(II)
Na(I)
K(I)
SO4 2
Cl
NO3
SiO2
TOC
pH

4.5
4060
815
1020
15
100200
3050
25
810
310
2.83.5

5
60
14
20
3
170
30
3
8

10
70
24
31
6
205
60
6
12

(Table 4), while Langmuir and RedlichPeterson equations


did not fit the adsorption data well. Also the correlation coefficient (R2 ) for the Freundlich equation is high. The values
of Freundlich and Langmuir constants, KF and Qo , for Duolite were found to be 12.11 and 138.89 mg/g, respectively,
which are 1.5 times greater than the values reported for FPBS.
The difference in adsorption capacity may be due to differences in ion exchange behaviour of sulphonate functionality
in styrenedivinylbenzene copolymer based Duolite and
the same functionality in phenolformaldehyde copolymer
based FPBS. Even though the adsorption potential of Duolite
is greater, the cost of commercial material is very high (US$
95 per kg of resin). The raw material used in the present
study, BS, is available free of cost, as a waste, and including
expenses for transport, chemicals for surface modification,
electrical energy for drying adsorbent materials, etc. the final
product, FPBS, would cost approximately US$ 20 per kg.
3.8. Application to wastewater from battery
manufacturing factory
This proposed method was applied to the removal of Pb(II)
in wastewater from battery manufacturing factory. The composition of the water sample [25] is presented in Table 5.
Fig. 11 shows the effect of the mass of FPBS onto adsorp-

Fig. 11. Effect of adsorbent dose for the removal of Pb(II) from battery
manufacturing industry wastewater.

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140


Table 6
Four cycles of Pb(II) adsorptiondesorption with 0.1 M HCl as the desorbing
agent
Cycles

Adsorption, mg/g (%)

Desorption, mg/g (%)

1
2
3
4

4.92 (98.4)
4.87 (97.3)
4.74 (94.8)
4.56 (91.2)

4.88 (99.2)
4.65 (95.5)
4.42 (93.2)
4.20 (92.0)

tion of Pb(II) from wastewater. It is evident that quantitative


removal of Pb(II) from 50 ml simulated water samples containing 5 and 10 mg/l Pb(II) ions and several other ions, a minimum adsorbent dose of 25 mg is sufficient for the removal of
85 and 75% of Pb(II), respectively. Almost complete (100%)
removal of Pb(II) from 50 ml samples was achieved with 75
and 125 mg of the adsorbent dose for samples 1 and 2, respectively, which is in good agreement with that obtained from
the batch experiments mentioned above. The results clearly
indicate that the proposed method can be used for the removal
of Pb(II) from wastewater.
3.9. Desorption capacity and recyclability
Recyclability of an adsorbent is of crucial importance in
industrial practice for metal removal from wastewater. To
test the suitability and stability of the adsorbent, it was subjected to successive adsorption and desorption cycles. The
procedure was carried out four times and 50 ml of 0.1 M HCl
was used as elution solution. The adsorbent was washed with
water before each measurement. The results in Table 6 clearly
show that FPBS can be used repeatedly without significantly
loosing the adsorption capacity for Pb(II). The percentage
adsorption/desorption values were calculated in relation with
the original amount of the adsorbent. At the end of the four
adsorptiondesorption cycles, the recovery and removal were
little decreased. The initial amount of FPBS employed in
the first cycle was 0.1 g and this was reduced to 0.076 g by
the time the FPBS reached the fourth cycle. Further study
is needed to solve the problems associated with the loss of
adsorbent over the course of the repeated cycles. The adsorbent seems to be relatively stable for at least four runs. The
process of adsorption and desorption does not alter the original physical and chemical characteristics of the adsorbent.

4. Conclusions
The preliminary studies presented here show that formaldehyde polymerized banana stem is an effective adsorbent for
the removal of Pb(II) from aqueous solutions. An initial pH in
the range 59 is favourable for the removal of Pb(II) by FPBS.
Removal of greater than 99.0% has been achieved under
optimum conditions. The equilibrium adsorption capacity
increased with increase of the initial concentration. The sorption process occurred in two stages: external mass transport
occurs in the early stage and intraparticle diffusion occurs

139

in the long-term stage. For both stages, the diffusion process is endothermic. The sorption equilibrium data could be
described well by the Freundlich equation. The simulated battery manufacturing industry wastewater samples are treated
by FPBS to demonstrate its efficiency in removing Pb(II)
from wastewater. The adsorbed Pb(II) is desorbed quantitatively by 0.1 M HCl solution and the adsorbent can be reused
for four cycles consecutively. Additional research is needed to
evaluate the sorption efficiency of this adsorbent using other
industrial wastewaters and also with other heavy metals.
Acknowledgement
We thank the Head, Department of Chemistry, University
of Kerala, for providing instrumental and laboratory facilities.
References
[1] B.E. Reed, S. Arunachalam, B. Tomes, Removal of lead and
cadmium from aqueous solutions using granular activated carbon
columns, Environ. Progr. 13 (1994) 6064.
[2] K.H. Park, M.A. Park, H. Jang, E.K. Kim, Y.H. Kim, Removal of
heavy metals cadmium(II) and lead(II) ions in water by sargassum,
Adsorption Sci. Technol. 12 (1999) 196202.
[3] K. Inoue, K. Ohto, K. Yoshizuka, T. Yamaguchi, T. Tanaka, Sorption
of lead(II) ion on complexane type of chemically modified chitosan,
Bull. Chem. Soc. Jpn. 70 (1997) 24432447.
[4] K.P. Shubha, C. Raji, T.S. Anirudhan, Immobilization of heavy metals from aqueous solutions using polyacrylamide grafted hydrous
tin(IV) oxide gel having carboxylate functional groups, Water Res.
35 (2001) 300310.
[5] C. Raji, K.P. Shubha, T.S. Anirudhan, Use of chemically modified
saw dust in the removal of lead(II) ions from aqueous media, Indian
J. Environ. Health 39 (1997) 230238.
[6] B.T. Abraham, T.S. Anirudhan, Sorption recovery of metal ions from
aqueous solution using humusboehmite complex, Indian J. Chem.
Technol. 8 (2001) 286292.
[7] S.H. Abdel-Halim, A.M.A. Shehta, M.F. El-Shahat, Removal of lead
ions from industrial wastewater by different types of natural materials, Water Res. 37 (2003) 16781683.
[8] E. Ott, Cellulose and Cellulose Derivatives, Interscience Publishers,
New York, 1946, 1176 pp.
[9] T.M. Suzuki, O. Itabashi, T. Goto, T. Yokoyama, T. Kimura, Preparation and metal-adsorption properties of the polymer-coated silica
gel having iminodiacetate functional group, Bull. Chem. Soc. Jpn.
60 (1987) 28392842.
[10] F. Helferrich, Ion Exchange, Mc Graw-Hill, New York, 1962, 122
pp.
[11] J.A. Schwarz, C.T. Driscoll, A.K. Bhanot, The zero point charge of
silicaalumina oxide suspension, J. Colloid Interface Sci. 97 (1984)
5561.
[12] H.P. Boehm, M. Voll, Basic surface oxides on carbon: adsorption of
acids, Carbon 8 (1920) 227240.
[13] D. Singh, N.S. Rawat, Sorption of Pb(II) by bituminous coal, Indian
J. Chem. Technol. 2 (1995) 4950.
[14] J. Crank, The Mathematics of Diffusion, second ed., Clarendon,
Oxford, 1975, 91 pp.
[15] L.D. Michelson, P.G. Gideon, E.G. Pace, L.H. Kutal, Removal of
Soluble Mercury from Wastewater by Complexing Technique, Bull.
No. 74, US, Department of Industry, Office of Water Research and
Technology, 1975.
[16] P. Mears, Polymers; Structure and Bulk Properties, Van Nostrand
London, 1965, 21 pp.

140

B.F. Noeline et al. / Separation and Purication Technology 45 (2005) 131140

[17] C.H. Giles, T.H. Mc Ewan, S.W. Nakhwa, D. Smith, Studies on


adsorption. Part II. A system of classification of solution, adsorption
isotherm, J. Chem. Soc. 4 (1960) 39733993.
[18] G. McKay, M.S. Otterburn, A.G. Sweeney, The removal of colour
from effluent using various adsorbents. III. Silica rate process, Water
Res. 14 (1980) 1520.
[19] B.T. Abraham, T.S. Anirudhan, Sorption recovery of metal ions from
aqueous solutions using humusboehmite complex, Indian J. Chem.
Technol. 8 (2001) 280296.
[20] J.Y. Ng, W.H. Cheung, G. McKay, Equilibrium studies for the
adsorption of lead from effluents using chitosan, Chemosphere 52
(2003) 10211030.
[21] L. Qingbiao, W. Singato, L. Gang, L. Xinkai, X. Deng, S. Daohua,
H. Yuelin, Y. Huang, Simultaneous biosorption of cadmium(II) and
lead(II) ions by pretreated biomass of Phanerochaeta chrysosporium,
Sep. Purif. Technol. 34 (2004) 135142.

[22] J. Rivera-Utrilla, I. Bautista-Toledo, M.A. Ferro-Gracia, C. MorenoCastila, Activated carbon surface modification by adsorption of bacteria and their effect on aqueous lead adsorption, J. Chem. Technol.
Biotechnol. 76 (2001) 12091215.
[23] R. Jalali, H. Ghafourian, Y. Asef, S.J. Sepehr, Removal and recovery
of lead using non living biomass of marine algae, J. Hazard. Mater.
92 (2002) 253262.
[24] V. Taty-Costodes, H. Fauduet, C. Porte, A. Delacroix, Removal of
Cd(II) and Pb(II) ions, from aqueous solutions, by adsorption onto
saw dust of Pinus sylvestris, J. Hazard. Mater. B105 (2003) 121
142.
[25] D. Petruzzelli, M. Pagno, G. Tiravante, R. Passino, Lead removal and
recovery from battery wastewaters by natural zeolite clinoptilolite,
Solvent Extr. Ion Exch. 17 (1999) 677694.

You might also like