(Y) Removal of Lead (II) From Aqueous Solution Using Heartwood of Areca Catechu Powder

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Desalination 256 (2010) 16–21

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 e v i e r. c o m / l o c a t e / d e s a l

Removal of lead(II) from aqueous solution using heartwood of Areca catechu powder
Paresh Chakravarty a,⁎, N. Sen Sarma b, H.P. Sarma c
a
Department of Chemistry, Birjhora Mahavidyalaya, Bongaigaon, Assam, India
b
Material Sciences Division, Polymer Section, IASST, Guwahati, Assam, India
c
Department of Environmental Science, Gauhati University, Guwahati, Assam, India

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

Article history: Removal of lead(II) from aqueous solution was studied using the powder of heartwood of Areca catechu as a
Received 29 October 2009 new biosorbent under batch method at room temperature. Various sorption parameters such as contact time,
Received in revised form 16 February 2010 initial concentration of lead(II) ion, effect of pH and amount of the biomass on the adsorption capacity of the
Accepted 17 February 2010
biosorbent were studied. The adsorbent was effective for the quantitative removal of lead(II) ions in acidic
Available online 12 March 2010
conditions and equilibrium has been achieved in 25 min. The equilibrium adsorption data were fitted to
Keywords:
Langmuir and Freundlich adsorption isotherm models and the model parameters were evaluated. The kinetic
Adsorbent study showed that the pseudo-second order rate equation better described the biosorption process. The FT-IR
Areca catechu powder spectra of the adsorbent before and after treatment with lead(II) solution indicated that hydroxyl, carboxyl,
Lead(II) ion amide and amine groups were major binding sites with the metal. This method is quite feasible, economic, time
Removal saving, and low cost.
Wastewater © 2010 Elsevier B.V. All rights reserved.

1. Introduction into the aqueous system from paper and pulp industries, lead smelter,
boat and ship fuels, battery manufacturers and ammunition industries.
Contamination of the environment by heavy metals is a growing Lead can contaminate the environment by anthropogenic sources as
concern because of health risks on humans and animals [1,2]. Most of the well as natural geochemical processes. It can accumulate along the
toxic metal pollutants are waste products of industrial and metallurgical food chain and is not amenable to biological degradation [14]. Lead
processes and in particular, the effluents from tanning industries. exposure causes weakness in fingers, wrists and ankles. The effects of
According to the WHO, the metals of most immediate concern are Al, As, lead toxicity are very wide ranging and include impaired blood
Cd, Cr, Co, Cu, Fe, Pb, Mg, Hg, Ni, and Zn. synthesis, hypertension, severe stomach-ache, and brain and kidney
There are several techniques, which have been utilized to reduce damage and even can cause miscarriage in pregnant women. The
heavy metal ion content in effluents, namely lime stone precipitation, consistent pattern of lower IQ values and other neuropsychological
ion exchange, adsorption on activated carbon, membrane processing deficits among the children exposed to higher levels of lead pollution
and electrolytic methods [1]. Most of these methods have been found to have been reported [15].
be limited since they often involve high capital and operational cost and In the present study, the ability of the heartwood powder of Areca
may also be associated with the generation of secondary wastes. Of catechu (betel-nut tree) (HPAC) to eliminate Pb(II) ion from synthetic
these techniques, activated carbon adsorption appears to be particularly wastewater and the effect of various parameters such as contact time
competitive and an effective process for the removal of heavy metals at of biosorbent and sorbent, pH of the metal solution, initial metal ion
trace level [3]. A good number of studies have been reported for metal concentration and different biomass amount have been investigated.
ion removal by using different bioadsorbents. In this regard, low cost Equilibrium modeling was carried out using Langmuir [16] and
biosorbent such as neem-leaf powder [4], phaseolus vulgaris [5], Freundlich [17] adsorption isotherm. The nature of the sorption
Canadian albicans [6], macro fungus [7], pine bark [8], ficus religiosa process has been evaluated with respect to its kinetic aspects.
leaves [9], green algae [10], citrus peel [11], bael leaves [12], moringa
oleifera bark [13] etc. have been utilized for the removal of lead and 2. Materials and methods
other heavy metals from wastewater.
Lead was chosen for biosorption studies with regard to their wide 2.1. Preparation of the heartwood powder of A. catechu (HPAC)
use in the industry and the potential pollution impact. Lead is released
A healthy and matured A. catechu (betel-nut tree) is collected from
Bakhrapara village of Bongaigaon district of Assam, India and the heart-
⁎ Corresponding author. Tel.: +919435122322 (mobile). wood of the same is carefully separated. The heartwood of A. catechu,
E-mail address: pareshchakravarty@gmail.com (P. Chakravarty). which is soft and spongy, is cut into small pieces and washed with tap

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.02.029
P. Chakravarty et al. / Desalination 256 (2010) 16–21 17

water and then with double distilled water to remove dust and other pH range of 2.0 to 7.0 at 29 °C. Experiments could not be performed at
impurities. The small cut pieces of heartwood of A. catechu are sun-dried higher pH value due to hydrolysis and precipitation of lead ions. Flasks
for seven days and dried again in an electric oven at 70 °C for 48 h. The were agitated on a shaker for 25 min to ensure that the equilibrium
dried materials are grounded in a laboratory blender and made into fine was reached. The mixtures were then filtered and the concentration of
powder. The finely divided powder of the heartwood of A. catechu is metal in the filtrates was measured.
then shorted in desiccator and used for biosorption studies.
2.4. FT-IR measurements
2.2. Metal solution
FT-IR spectra for both free HACP and Pb(II) loaded HPAC were
All chemicals and reagents used for experiments and analyses obtained by KBr pellets methods operated on FT-IR spectrophotometer
were of analytical grade. Stock solution of 1000 mg/L of Pb(II) was (Brucker, Vector 22) to investigate the functional groups present in
prepared from Pb (NO3)2 (E. Merck, Mumbai, India) in double distilled the biomass and to look into possible Pb(II) binding sites.
water that contained few drops of 0.1 N HNO3 to prevent the
precipitation of Pb(II). The solution was diluted as required to obtain 3. Result and discussion
the working solution. The initial pH of the working solution was
adjusted to 5.0 by an addition of the 0.1 N HNO3 or 0.1 N NaOH 3.1. Effect of contact time
solutions except for the experiment examining the effect of pH. Fresh
dilutions were used for each study. The effect of contact time on Pb(II) biosorption on HPAC was
studied and the results were shown in Fig. 1. From Fig. 1, it was found
2.3. Methods of adsorption studies that the adsorption quantity of Pb(II) ion on HPAC increases as the
contact time increased. The biosorption of lead onto HPAC was rapid
Batch adsorption experiments were carried out by shaking the flasks for the first 5 min (88%) and equilibrium was nearly reached after
at 120 rmp for a fixed period of time using a mechanical shaker. 25 min (97%). Hence, in the present study, 25 min was chosen as the
Following a systematic process, the adsorption uptake capacity of Pb(II) equilibrium time. Basically the removal rate of sorbate is rapid, but it
in batch system was studied in the present work. The data obtained in gradually decreases with time until it reaches equilibrium. The rate in
batch mode studies was used to calculate the equilibrium metal percent of metal removal is higher in the beginning due to the larger
adsorptive quantity by using the following expression: surface area of the adsorbent being available for the adsorption of the
metals [18]. It is also relevant that, since active sorption sites in a
ðCo −Ce ÞV system have a fixed number and each active site can adsorb only one
qe =
m ion in a monolayer, the metal uptake by the sorbent surface will be
rapid initially, slowing down as the competition for the decreasing
where qe is the amount of heavy metal ion adsorbed onto per unit availability of active sites intensifies by the metal ions remaining in
weight of the biomass in mg/g, V is the volume of solution treated in the solution [19].
liter, Co is the initial concentration of metal ion in mg/L, Ce is the
equilibrium metal ion concentration in mg/L and m is the biomass in g. 3.2. Effect of initial metal ion concentration

2.3.1. Effect of contact time Biosorption experiments with HPAC were conducted for solutions
Batch biosorption experiments were carried out at different contact containing 10 mg/L to 70 mg/L Pb(II) ion. As seen in Fig. 2, at lower
times (5, 10, 15, 20, 25, 30 and 35 min) at an initial concentration of concentrations of Pb(II) ion (10 mg/L–40 mg/L), biosorption was
20 mg/L of Pb(II) ion solution at pH 5.0, the HPAC dose concentration is completed in about 5 min, but at higher concentrations it took about
0.5 g in 100 ml of solution in 250 ml conical flask at room temperature 25 min. At lower concentrations, all metal ions present in the solution
(29 °C ± 2 °C). The samples were then agitated in a mechanical shaker would interact with binding sites and then facilitated about 99%
at 120 rmp at a regular time interval and filtered through Whatman 42 adsorption. At higher concentration, more Pb(II) ions are left
filter paper and the filtrates were analyzed using flame atomic unabsorbed in the solution due to the saturation of binding sites.
absorption spectrometry (Model: Perkin Elmer 3110). Each determi- This appears due to the increase in the number of ions competing for
nation is repeated three times and the results given were the average available binding sites in the biomass [20].
values.

2.3.2. Effect of initial metal ion concentration


Equilibrium experiments were carried out by contacting 0.5 g of
HPAC with 100 ml of Pb(II) ion solution of different initial concentra-
tions (10 mg/L–70 mg/L) at pH values of 5.0 at room temperature. A
series of such conical flasks were shaken for 25 min at a speed of
120 rpm at room temperature (29 °C ± 2 °C). The samples were
filtered and the filtrates were analyzed as mentioned before.

2.3.3. Effect of biosorbent concentration


Batch adsorption tests were done at a different concentration of
HPAC from 0.1 g to 0.6 g at a 100 ml solution of 20 mg/L of Pb(II) ion at
pH 5.0, for a contact time of 25 min at room temperature (29 °C± 2 °C).
The samples were then agitated and filtered and the filtrates were
analyzed as mentioned before.

2.3.4. Effect of solution pH on biosorption Fig. 1. Effect of contact time on Pb(II) adsorption on the heartwood powder of Areca
The effect of pH on the adsorption capacity of HPAC was catechu (HPAC), 100 ml single metal solution, pH = 5.0, initial metal ion concentration
investigated using a 100 ml solution of 20 mg/L of Pb(II) ion for a (Co) = 20 mg/L, biosorbent amount = 0.5 g, and temperature = 29 °C.
18 P. Chakravarty et al. / Desalination 256 (2010) 16–21

Fig. 4. Effect of pH on Pb(II) adsorption on HPAC; 100 ml of single metal solution; contact
Fig. 2. Effect of Pb(II) ion concentration on HPAC; initial metal ion concentration = 10 mg/L, time = 25 min, pH= 2.0, 3.0, 4.0, 5.0, 6.0 and 7.0; initial metal ion concentration= 20 mg/L,
20 mg/L, 30 mg/L, 40 mg/L, 50 mg/L, 60 mg/L and 70 mg/L; 100 ml of single metal solution; amount of HACP= 0.5 g and temperature = 31 °C.
contact time=25 min; amount of HPAC= 0.5 g; temperature = 30 °C, and pH = 5.0.

downward trend from pH 6.0. In lower pH values, occupation of the


3.3. Effect of biosorbent concentration negative sites of the adsorbent by H+ and H3O+ ions leads to
reduction of vacancies for metal ion and consequently causes a
The influence of biosorbent concentration in equilibrium uptake decrease in metal ion biosorption [23]. As the pH was raised, the
was shown in Fig. 3. It can be observed that the percentage removal of ability of the metal ions for competition with H+ and H3O+ ions was
Pb(II) ion increases with the increase in HPAC doses from 0.1 g to also increased. Although the sorption of metal ions raised by growing
0.5 g. There was a non-significant increase in the removal of pH, further increment of pH caused declining in adsorption due to the
percentages of Pb(II) when the adsorbent dose increases beyond precipitation of metal hydroxides.
0.5 g. A similarity in metal uptake with variation in biosorbent
concentration has been reported for lead biosorption from its 3.5. FT-IR analysis
synthetic aqueous solutions by Spirulina maxima [21] and Cr {VI}
removal from industrial effluent using Saccharomyces Cerevisiae Numerous chemical functional groups such as carbonyl, hydroxyl,
immobilised in chitosan [22]. This suggests that after a certain dose amine, amide, etc. have been identified as potential adsorption sites to
of biosorbent, the maximum adsorption is attained and hence the be responsible for binding metallic ions to the biomass. The FT-IR
amount of ions remains constant even with further addition of dose of spectroscopy is an important analytical technique, which detects the
adsorbent. The increase in Pb(II) removal percentage with increase in vibration characteristics of chemical functional groups present on
adsorbent dose is due to the greater availability of the exchangeable adsorbent surfaces. The FT-IR spectra of both the adsorbents (fresh as
sites or surface area at a higher concentration of the adsorbent. well as lead loaded) are shown in Fig. 5.
FT-IR spectra of fresh HPAC show peaks between the regions 3500
3.4. Effect of solution pH and 3200 cm−1 which represent the overlapping peaks of stretching
vibrations of O–H and N–H groups. The strong absorption peak
The pH of the solution is perhaps the most important parameter at 2919 cm−1 could be assigned to –CH stretching vibrations of –CH3
for adsorption. To understand the adsorption mechanism, the and –CH2 functional groups. The distinct peak observed between
adsorption of Pb(II) as a function of pH was measured and the result 1749 cm−1 and 1627 cm−1 characterizes carbonyl groups stretching from
is shown in Fig. 4. Lead removal recorded its minimum values at pH aldehyde and ketones. The adsorption peaks at 1627–1421 cm−1
2.0. There was an increase in the biosorption capacity of the biomass correspond to the primary and secondary amide bands, respectively.
with an increase in pH from 2.0 to 5.0 and showed marginal 1421–1392 cm−1 was the deformation stretching of C–H, –CH3 and –CH2

Fig. 3. Effect of biosorbent concentration on Pb(II) on HPAC; 100 ml of single metal


solution; contact time = 25 min, pH = 5.0; initial metal ion concentration = 20 mg/L,
and temperature = 30 °C. Fig. 5. FT-IR spectra of fresh-dried (A) HPAC and (B) Pb(II) loaded HPAC.
P. Chakravarty et al. / Desalination 256 (2010) 16–21 19

functional groups. The strong band within 1100–1000 cm−1 is due to 3.7. Adsorption kinetic studies
the C–O group, which are characteristic peaks of polysaccharides.
Changes in intensity and shift in position of the peaks could be Adsorption kinetics, which describes the solute adsorption rate, is
observed in FT-IR spectrum after Pb(II) adsorption in HPAC. The an important characteristic in evaluating the efficiency of the
first change was the enhancement of the intensity at the regions adsorption. In order to clarify the biosorption kinetics of Pb(II) ions
3500–3200 cm−1, indicating an increase of the free hydroxyl group on onto HPAC, the kinetic models, that are, Lagergren's pseudo-first order
the biomass. The shifting of peak at 1749 to 1737 cm−1 and 1627 to [27] and pseudo-second order [28] models were applied to the
1629 cm−1 indicates the involvement of C O of carboxyl group, N–H of experimental data. The linearised form of the pseudo-first order rate
amine and C–O of amides in the adsorption process. The minor shift of equation by Lagergren is given as;
the peak from 1033 to 1031 cm−1 also suggests the involvement of
the C–O group in binding Pb(II). It is clear from the FT-IR analysis that K1 ⋅t
Logðqe −qt Þ = log qe −
the possible mechanism of adsorption of Pb(II) on the biomass HPAC 2:303
may be due to physical adsorption, ion exchange, surface precipita-
tion, complexation with functional groups and chemical reactions where qe and qt are the amounts of metal ion sorbed (mg/g) at
with surface sites [10,24]. equilibrium and at time t, respectively. K1 is the Lagergren rate
constant of the biosorbent (1/min). A plot of log (qe − qt) versus t
3.6. SEM and EDX analysis gives the result shown in Fig. 7. The values of K1 and qe, calculated
from the slope and intercept are presented in Table 1.
Scanning electron microscopy (LEO, Model 1430VP) along with The pseudo-second order kinetic model of McKay and Ho [28] can
energy dispersive X-ray (EDX) analysis has been used by many be expressed as;
researchers for the characterization of the adsorbent as well as the
elucidation of the probable mechanism of biosorption. SEM micro- t 1 1
= + t
graphs and EDX spectra obtained from fresh HPAC and lead loaded qt K2 q2e qe
HPAC are shown in Fig. 6. SEM micrographs of fresh HPAC (Fig. 6a)
reveal the nature of the biomass which is dark, rough, heterogeneous, where the equilibrium biosorption capacity, qe and the pseudo-
very few pores and lots of ups and down, and thus makes possible for second order rate constant K2 (g/mg min) were determined exper-
the adsorption of Pb(II) ions in different parts of the adsorbent. Fig. 6b imentally from the slope and intercept of plot t / qt versus t (Fig. 7).
represents the micrograph of Pb(II) loaded HPAC. The micrograph Calculated correlations are closer to unity for pseudo-second order
clearly shows the presence of shiny particles over the surface of Pb(II) kinetic model; therefore, the biosorption kinetics could well be
loaded biosorbent which are absent in the fresh biosorbent [25,26]. approximated more favourably by the second-order kinetics model
EDX analysis provides the elemental information for the fresh as well rather than the pseudo-first order kinetics for Pb(II). The K2 and qe
as the Pb(II) loaded biosorbent. Fig. 6c indicates the presence of C, O, values for the pseudo-second order kinetic, calculated from Fig. 8 are
N, Cr, Ca, and S in the fresh biosorbent. EDX spectra presented in listed in Table 1. By comparing the constants of the two kinetic
Fig. 6d.revealed the additional Pb signal, which confirms the binding models, the pseudo-second order kinetic model seems to be best
of the metal ion to the surface of the biomass. fitted for the experiment.

Fig. 6. SEM micrographs of (a) fresh HPAC, (b) Pb(II) loaded HPAC; EDX spectra of (c) fresh HPAC and (d) Pb(II) loaded HPAC.
20 P. Chakravarty et al. / Desalination 256 (2010) 16–21

Fig. 8. Pseudo-second order kinetic plot for the biosorption of Pb(II) by HPAC at 31 °C
Fig. 7. Pseudo-first order kinetic plot for the biosorption of Pb(II) by HPAC at 31 °C and and pH 5.0.
pH 5.0.
plot of the Langmuir and Freundlich isotherm models for sorption of Pb
3.8. Biosorption equilibria studies (II) on HPAC are presented in Figs. 9 and 10. The plot of Ce / qe versus Ce
(Fig. 9) showed that the experimental data fitted reasonably well to the
An adsorption isotherm is a graphical representation showing the linearised equation of the Langmuir isotherm over the whole Pb(II)
relationship between the amounts adsorbed by a unit weight of concentration range studied. The correlation coefficient was 0.9993.
adsorbent and the amount of adsorbate remaining in a test medium at qmax and b were evaluated from the slope and intercept of the plot and
equilibrium. The lead(II) uptake capacity of HPAC was evaluated using were found to be 11.7233 mg/g and 1.3023 L/mg respectively.
the Langmuir [16] and Freundlich [17] adsorption isotherms. The The essential characteristics of Langmuir can be explained in terms
Langmuir equation, which is valid for monolayer sorption onto a surface of dimensionless constant separation factor (RL), defined by:
with a finite number of identical sites, which are homogeneously
distributed over the adsorbent surface, is given by the equation: 1
RL =
1 + bCo
qmax bCeq
qe =
1 + bCeq
where b is the Langmuir constant and Co is the initial concentration of
where qe is the amount of metal ion bound to per gram of the biomass metal ion. The value of RL indicated the type of Langmuir isotherm to
at equilibrium and Ce is the residual (equilibrium) metal ion be irreversible (RL = 0), favourable (0 b RL b 1), linear (RL = 1) or
concentration left in the solution after binding, respectively. qmax is unfavourable (RL N 1). The RL was found to be 0.0713 to 0.0108 for
the maximum amount of metal ion per unit weight of sorbent to form a concentrations of 10–70 mg/L Pb(II).
complete monolayer on the surface and b is the constant related to the Linear plots of log qe versus log Ce showed that the Freundlich
affinity of the binding sites. qmax and b can be determined from Ce / qe isotherm was also representative for the Pb(II) by the biosorbent
versus the Ce plot which gives a straight line of slope 1/qmax and tested. The correlation coefficient was 0.9728. Kf and n were calculated
intercept 1/b∙qmax [29]. from the slopes of the Freundlich plots (Fig. 10) and were found to be
On the other hand, the Freundlich equation is an empirical 0.1816 and 1.5562 respectively. The magnitude of Kf and n shows easy
equation based on adsorption on a heterogeneous surface is given separation of heavy metal ion from wastewater and high adsorption
by equation: capacity. The value of n, which is related to the distribution of bonded
ions on the sorbent surface, represents beneficial adsorption if it is
1
qe ¼Kf Ce n between 1 and 10 [30]. The n value for the biosorbent used found to be
greater than 1, indicating that adsorption of Pb(II) is favourable.
where Ce is the equilibrium concentration (mg/L), qe is the amount of
metal ion bound to per gram of the biomass at equilibrium (mg/L) and
Kf and n are the Freundlich constants related to the sorption capacity
and sorption intensity of the sorbent, respectively. The equation can
be linearised in logarithmic form and Freundlich constants can be
evaluated.
The Langmuir and Freundlich equations were used to describe
the data derived from the adsorption of Pb(II) by the biosorbent over
the entire concentration range studied (10 mg/L to 70 mg/L). The linear

Table 1
Kinetic model parameters for the biosorption of Cd (II) by the heartwood powder of
Areca catechu (HPAC).

Pseudo-first order Pseudo-second order Experimental


parameters parameters value

qe (mg/g) K1 (1/min) R2 qe (mg/g) K2 (g/mg min) R2 qe (mg/g)

3.85 0.143 0.9892 20.2 0.0572 0.9998 19.5


Fig. 9. Langmuir isotherm plot for Pb(II) adsorption onto HPAC.
P. Chakravarty et al. / Desalination 256 (2010) 16–21 21

References
[1] J.M. Tobin, J.C. Roux, Mucor biosorbent for chromium removal, Water Res. 32
(1998) 1407–1416.
[2] E. Fourest, et al., Contributions of carboxyl groups to the heavy metal binding sites
in fungal wall, Toxic Environ. Chem. 54 (1996) 1–10.
[3] H.H. Tran, F.A. Roddick, J.A. O'Donnell, Comparism of chromatography and
desiccant silica gels for the adsorption of metal ions-I, adsorption and kinetics,
Water Res. 33 (1999) 2992–3000.
[4] A. Sharma, K.G. Bhattacharya, Interaction of Pb (II), Cd (II) and Cr (II) with neem
(Azadirachta indica) leaf powder; kinetics and thermodynamics, Int. J. Environ.
Pollut. 34 (1–4) (2008) 374–399.
[5] A. Safa Ozcan, S. Tunali, T. Akbar, A. Ozcan, Biosorption of lead (II) ions onto waste
biomass of Phaseolus vulgaris L.: estimation of equilibrium kinetic and thermo-
dynamics parameters, Desalination 244 (1–3) (2009) 188–198.
[6] Z. Baysal, E. Cinar, Y. Bulut, H. Alkan, M. Dogru, Equlibrium and thermodynamics
studies on biosorption of Pb (II) onto Candida albicans biomass, J. Hazard. Mater.
161 (1) (2009) 62–69.
[7] A. Sari, M. Tuzon, Kinetic and equilibrium studies of biosorption of Pb (II) and Cd (II)
from aqueous solution by macro fungus (Amanita rubescens) biomass, J. Hazard.
Mater. 164 (2–3) (2009) 1004–1011.
Fig. 10. Freundlich plot for Pb(II) adsorption onto HPAC.
[8] A. Gundogdu, D. Ozdes, C. Duran, V.N. Bulut, M. Soylak, H.B. Senturk, Biosorption
of Pb(II) ions from aqueous solution by pine bark (Pinus brutia Ten.), Chem. Eng. J.
153 (1–3) (2009) 62–69.
Table 2 [9] S. Qaiser, A.R. Saleemi, M. Umer, Biosorption of lead from aqueous solution by
Isotherm parameters for Pb(II) adsorption on the heartwood of Areca catechu powder Ficus religiosa leaves: batch and column study, J. Hazard. Mater. 166 (2–3) (2009)
(HACP). 998–1005.
[10] V.K. Gupta, A. Rastogi, Biosorption of lead from aqueous solution by green algae
Langmuir isotherm Freundlich isotherm Spirogyra species: kinetic and equilibrium studies, J. Hazard. Mater. 152 (1)
b (L/mg) qmax (mg/g) R2 Kf n R2 (2008) 407–414.
[11] A. Balaria, S. Schiewer, Assessment of biosorption mechanism for Pb binding by
1.3023 11.7233 0.9993 0.1816 1.5562 0.9728 citrus pectin, Sep. Purif. Technol. 639 (3) (2008) 577–581.
[12] S. Chakravarty, A. Mohanty, T. Nag Sudha, A.K. Upadhyay, J. Konar, J.K. Sircar, A.
Madhukar, K.K. Gupta, Removal of Pb(II) ions from aqueous solution by
adsorption by using bael leaves (Aegle marmelos), J. Hazard. Mater. 173 (1–3)
Table 2 gives the isotherm parameters for both Langmuir and (2010) 502–509.
Freundlich isotherms. From linear correlation coefficients of the adsorp- [13] D.H.K. Reddy, K. Seshaiah, A.V.R. Reddy, M.M. Rao, M.C. Wang, Biosorption of Pb2+
tion isotherm, it is noted that the Langmuir isotherm model exhibits a ions from aqueous solution by Moringa Oleifera bark: equilibrium and kinetic
studies, J. Hazard. Mater. 174 (1–3) (2010) 831–838.
better fit to the sorption data of Pb(II) than the Freundlich isotherm [14] A.Y. Dursun, G. Uslu, Y. Cuci, Z. Aksu, Bioaccumulation of copper (II), lead (II) and
model. This phenomenon suggests that monolayer sorption takes place on chromium (VI) by growing Aspergillus niger, Process Biochem. 38 (2003)
the surface of the HPAC. 1647–1651.
[15] Toxicological profile for lead, US Department of Health & Human Services, Public
Health Services (Agency for Toxic Substances and Diseases Registry), Atlanta,
4. Conclusion Georgia, (1999 a).
[16] I. Langmuir, The constitution and fundamental properties of solids and liquids.
Part I: solids, J. Am. Chem. Soc. 38 (11) (1916) 2221–2295.
In this study the ability of the biomass of the heartwood powder of [17] H.M.F. Freundlich, Over the adsorption in solution, J. Phys. Chem. 57A (1906)
A. catechu (HPAC) to remove the synthetic Pb(II) solution has been 385–470.
demonstrated. In this batch mode of studies, the adsorption was [18] E.S.Z. El-Ashtoukhy, N.K. Amin, O. Abdelwahab, Removal of lead (II) and copper
(II) from aqueous solution using pomegranate peel as a new adsorbent,
dependent on contact time, pH, initial metal ion concentration, and
Desalination 223 (1–3) (2008) 162–173.
biosorbent dosage. The biosorbent was successful in removing Pb(II) [19] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum,
with 97% sorption efficiency from the aqueous solution containing J. Am. Chem. Soc. 40 (1918) 1361–1368.
20 mg/l Pb(II). The biosorption was rapid and equilibrium was [20] P.R. Puranik, K.M. Paknikar, Biosorption of lead, cadmium and zinc by Citrobacter
stain MCMB- 181: characterization studies, Biotechnol. Progr. 15 (2) (1999)
achieved within 25 min. It was also observed that the adsorption 228–237.
was pH dependent and the maximum adsorption of 97% occurred at [21] R. Gong, Y. Ding, H. Liu, Q. Chen, Z. Liu, Lead biosorption and desorption by intact
pH 5.0. The results of the study revealed that the removal of Pb(II) from and pretreated Spirulina maxima biomass, Chemosphere 58 (1) (2005) 125–130.
[22] N. Saifuddin, A.Z. Raziah, Removal of heavy metals from industrial effluent using
aqueous solutions by using HPAC as an adsorbent seems to follow the Saccharomyces Cerevisiae (Baker's yeast) immobilized in chitosan/lignosulphonate
pseudo-second order kinetic model and the Langmuir isotherm better matrix, J. Appl. Sci. Res 3 (12) (2007) 2091–2099.
than the Lagergren's pseudo-first order kinetic model and Freundlich [23] J.M. Tobin, D.G. Cooper, R.J. Neufeld, Appl. Environ. Microbiol. 47 (1984) 821–824.
[24] K. Srividya, K. Mohanty, Biosorption of hexavalent chromium from aqueous
isotherm, respectively. FT-IR and SEM characterization of the biosor- solutions by Catla catla scale: equilibrium and kinetics studies, Chem. Eng. J. 155
bent have shown a clear difference in the fresh and Pb(II) loaded (2009) 666–673.
biosorbent. Based on all results, it can be concluded that since the [25] S. Tulani, K. Ismail, T. Akbar, Chromium (VI) biosorption characteristics of
Neurospora crassa fungal biomass, Miner. Eng. 18 (2005) 681–689.
biosorbent, HPAC, is an easily available and low cost adsorbent and has
[26] M. Bansal, D. Singh, V.K. Garg, P. Rose, Use of agricultural waste for the removal of
a considerable high biosorption capacity, it may be treated as an nickel ions from aqueous solutions: equilibrium and kinetics studies, Int. J. Environ.
alternative biosorbent for treatment of wastewater containing Pb(II) Sci. Eng. (2009) 108–114.
[27] S. Lagergren, B.K. Svenska, Zur theorie dersogenannten adsorption geloester
ions.
stoffe, Vetenskapsakad Handl 24 (1898) 1–39.
[28] G. MaKay, T.S. Ho, Pseudo-second order model for sorption processes, Process
Acknowledgement Biochem. 34 (1999) 451–465.
[29] Z. Aksu, Equilibrium and kinetic modeling of cadmium (II) biosorption by C. valgaris
in a batch system: effect of temperature, Sep. Purif. Technol. 21 (2001) 285–294.
The authors are thankful to the Regional Sophisticated Instrumen- [30] K. Kadirvelu, C. Namasivayan, Agricultural by-products as metal adsorbents:
tation Centre (RSIC), Shillong, India, for providing instrumental help sorption of lead (II) from aqueous solutions onto coir-pith carbon, Environ.
during the research work. Technol 21 (10) (2000) 1091–1097.

You might also like