You are on page 1of 9

Chemical Engineering Journal 185–186 (2012) 82–90

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Biosorption potential of the waste biomaterial obtained from Cucumis melo for
the removal of Pb2+ ions from aqueous media: Equilibrium, kinetic,
thermodynamic and mechanism analysis
Sibel Tunali Akar a,∗ , Sercan Arslan b , Tugba Alp b , Derya Arslan b , Tamer Akar a
a
Department of Chemistry, Faculty of Arts and Science, Eskisehir Osmangazi University, Turkey
b
Department of Chemistry, Graduate School of Natural and Applied Sciences, Eskisehir Osmangazi University, 26480 Eskisehir, Turkey

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

Article history: The potential use of a waste biosorbent material obtained from Cucumis melo (C. melo) for the removal
Received 14 November 2011 of Pb2+ ions from aqueous solutions was investigated by considering equilibrium and kinetic aspects.
Received in revised form 2 January 2012 The biosorption showed a pH dependent profile. An increase in biosorbent dosage up to 1.8 g L−1 caused
Accepted 3 January 2012
an increase in the biosorption yield of the biosorbent. The relatively fast biosorption at all studied tem-
peratures follows the pseudo-second-order kinetic model. Biosorption isotherm modeling shows the
Keywords:
equilibrium data fitted to the Langmuir model with a maximum monolayer biosorption capacity of
Biosorption
3.64 × 10−4 mol g−1 . The thermodynamic parameters indicated the biosorption of Pb2+ on the biomass was
Isotherm
Kinetic
a spontaneous and endothermic process. Experiments conducted with multi-metal system demonstrated
Pb2+ that the presence of co-ions slightly reduced the Pb2+ biosorption capacity of the biomass. The possible
Mechanism Pb2+ ion-biomaterial interactions were evaluated by the zeta potential, FTIR, SEM and EDX analysis.
Thermodynamic Results of this work showed the suggested biosorbent could be an effective and eco-friendly alternative
for the removal of Pb2+ ions from contaminated solutions.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction effectively with toxic metals [5]. Among the commonly available
sorbent materials, plant origin biomasses being economic and
Discharge of metal contaminated effluents into aquatic environ- ecofriendly due to their unique chemical composition, availabil-
ment without adequate treatment is one of the significant problems ity in large quantity, renewable, low cost and more efficient are
and must be seriously considered in the water treatment technol- seem to be viable option for heavy metal remediation. Hence the
ogy. Among different organic and inorganic pollutants heavy metal use of plant derived biomaterials in this field has received much
ions are serious poisons capable of being assimilated and stored in more attention in recent years [6]. Symphoricarpus albus berries
the tissues of living organisms, causing noticeable health problems. [7], Tamarindus indica seeds [8], olive mill solid and olive stone
For example, Pb2+ is a potent metal toxin and common contaminant [9], castor tree leaves [10], Capsicum annuum seeds [11], golden
found in the industrial effluents [1]. Due to acute toxicity of lead, it shower pods bark [12], peanut shell [13], coffee husks [14], chaff
has joined mercury and cadmium in forming “the big three” of met- [15], Ficus religiosa leaves [16] and black gram husk [17] are just a
als with the serious potential hazard [2]. The amount of lead used few examples of these materials.
in the 20th century far exceeds the total consumed in all previous Cucumis melo (C. melo) is one of the most important fruit crops
eras [3]. Effluents from production of batteries, gasoline additives, grown in many countries. It is easily cultivated in all the temper-
pigments, alloys and sheets etc. contain often high concentrations ate regions of the world because of its good adaptation to growing
of Pb2+ ions [4]. Therefore, adequate treatment of lead containing conditions [18]. It was reported that about 30% of the world melon
effluents prior to discharge into receiving bodies of water is of great production is in Turkey and it is followed by USA, Spain, and Roma-
importance for human health and environmental quality. nia [19]. The fruits are widely consumed as desert and vegetable in
The biosorption has considerable amount of attention as an summer period and contain seeds in large quantities [20]. The seeds
alternative process to traditional methods in heavy metal removal of C. melo are a rich protein (53.90%) and oil (37.67%) [21] and are
from contaminated water. Different kinds of biomaterials interact sometimes used in Chinese folk medicine as antitussive, digestive,
febrifuge and vermifuge [22].
The objective of the present work was to investigate the pos-
∗ Corresponding author. Tel.: +90 222 2393750/2862; fax: +90 222 2393578. sible use of C. melo seeds as a biosorbent for the removal of Pb2+
E-mail address: stunali@ogu.edu.tr (S. Tunali Akar). ions from contaminated solutions since its biosorption potential

1385-8947/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2012.01.032
S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90 83

has not been previously reported. The batch biosorption param- 20


eters such as pH, biosorbent dosage, contact time, temperature
and ionic strength were investigated. Some instrumental analysis
methods were employed to evaluate the possible biosorbent-metal 10

Zeta potential (mV)


ion interactions. Kinetic, isotherm and thermodynamic parame-
ters were deduced. The biosorption performance of the biosorbent
0
material in multi-metal mixtures was also determined in order to 1 2 3 4 5 6
study the co-ion effect. Overall the first comprehensive report on
pH
the biosorption potential of C. melo seeds was presented. -10

2. Experimental
-20
2.1. Materials used biosorbent in water
biosorbent in Pb 2+ solutions

C. melo fruits were obtained from local markets in Turkey and -30
seeds were removed from their flesh. Separated seeds were washed
Fig. 1. Zeta potential trend for C. melo biomass at different pH values.
several times with deionized water. They were crushed and ground
using a laboratory mill and dried at 60 ◦ C for 24 h to obtain a con-
stant weight. Dried seeds were screened through a sieve of 300 ␮m
Pb2+ biosorption was investigated. The biosorbent dosage was opti-
particle size and then stored for further use. Pb2+ ion solutions were
mized by changing the amount of biosorbent from 0.6 to 4.0 g L−1 .
prepared by dissolving Pb(NO3 )2 in deionized water to make a stock
Pb2+ biosorption kinetics was determined with initial Pb2+ con-
Pb2+ solution of 1 g L−1 . Other concentrations of Pb2+ ion solutions
centration of 200 mg L−1 at three different temperatures. For the
used in the experiments were obtained by dilution of this stock
biosorption isotherms the concentration of Pb2+ solutions was
solution. The pH adjustments were done by adding small amounts
changed between 50 and 250 mg L−1 .
of 0.1 M NaOH and/or 0.1 M HNO3 solutions.
In order to study the competitive biosorption of Pb2+ ions from
their binary mixtures, the initial concentrations of the competing
2.2. Characteristics of material
metal ions (Cd2+ , Cu2+ and Ni2+ ) were varied between 5, 50 and
100 mg L−1 while the Pb2+ ion concentration in each biosorption
Zeta potential measurements were done using a zeta potential
solution was held constant at 100 mg L−1 . Similarly biosorption tri-
analyzer (Malvern zeta sizer) in order to explain the relation-
als with Pb2+ , (Cd2+ , Cu2+ and Ni2+ ) were carried out using solutions
ship between the pH and surface charge of biosorbent. A 0.1 g of
containing four metals simultaneously.
biosorbent was placed into 50 mL of deionized water or 100 mg L−1
The equilibrium concentrations (Ce ) of the biosorbate were
Pb2+ solutions and stirred for 80 min. The pH of the medium was
determined by an atomic absorption spectrometer FAAS (ATI Uni-
adjusted to desired value with the pH range of 2.0–5.5. The small
cam929). Equilibrium biosorption capacities (qe ) were calculated
volumes of the mixtures were used to measure of zeta potentials.
with Eq. (1);
All measurements were done at least three times and mean values
were adopted. V (Ci − Ce )
Original and Pb2+ -loaded biosorbent (filtered and dried after qe = (1)
m
contact with 50 mL of 100 mg L−1 Pb2+ at pH 5.5) samples were
used for FTIR, SEM and EDX analysis. where Ci is initial Pb2+ concentration (mg L−1 ), V is the solution
The surface functional groups on the biosorbent material were volume (L) and m is the amount of biosorbent (g).
identified using FTIR spectroscopic method. About 1 mg biosorbent
sample was ground with about 100 mg of finely divided spectro- 2.4. Statistical analysis
scopic grade KBr powder. The mixture was placed in a die and
pressed under high pressure. Obtained disks were placed into In order to ensure the reproducibility of results, experiments
Bruker Tensor27 FTIR spectrometer and FTIR spectra were recorded were replicated three times and data presented were the mean
in the region of 400–4000 cm−1 . values from these independent experiments. Experimental errors
The surface microstructure of the biosorbent material were were estimated and depicted with error bars and standard devia-
characterized using scanning electron microscope coupled with tions are indicated wherever necessary. All statistical analysis was
energy dispersive X-ray analysis (Zeiss SUPRA 50 VP), at 20 kV and performed using SPSS 10.0 for Windows.
1000 times of magnification. Prior to analyze, biosorbent samples
were mounted on a stainless steel stab with a double stick tape. 3. Results and discussion
Then they were coated with a thin layer of gold under vacuum to
improve electron conductivity and image quality. 3.1. Characterization of biosorbent
The BET surface area, total pore volume, average pore size and
micro pore volume of the biosorbent were determined from the The BET surface area, total pore volume, average pore diam-
N2 adsorption isotherm with a surface area and pore size analyzer eter and micro pore volume of the biosorbent were found to be
(Quantachrome Instruments, Autosorb 1). 8.407 m2 g−1 , 0.01152 cm3 g−1 , 54.83 Å and 3.94 × 10−6 cm3 g−1 ,
respectively.
2.3. Biosorption studies The surface charge trend of the biosorbent in water and Pb2+
solutions as a function of pH was shown in Fig. 1. Firstly, the
The batch biosorption experiments were carried out by using zeta potential measurements were carried out in deionized water.
a sample containing 0.1 g of biosorbent and 50 mL of 100 mg L−1 According to these results the biosorbent surface was positively
Pb2+ solution in 100 mL beakers and stirring on a magnetic stirrer charged at pH 2.0. When the pH was increased to 3.0 the sur-
at 200 rpm. The original pH of Pb2+ solution was used through- face charge of the biomass converted to negative. Above pH 3.0
out all biosorption experiments except when the pH effect on the the zeta potential values of the biosorbent stayed almost constant
84 S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90

Fig. 2. FTIR spectra for waste biomass obtained from C. melo before (a) and after (b) Pb2+ biosorption.

(p > 0.05). Zero point of charge (ZPC) values provide useful informa-
tion related to net surface charge of the biosorbent as a function of
pH. In this study the ZPC value of the biomass in water was observed
at about pH 2.5. The high negative charge on the biosorbent sur-
face above ZPC makes it useful for metal removal because of the
attraction forces between metal cation and biosorbent surface. Sec-
ondly, biosorbent was mixed with Pb2+ ion solutions at different
pHs and zeta potentials of the mixtures were recorded. At these
conditions, ZPC value shifted to about 3.75. This finding supports
the mechanism of the electrostatic interactions between Pb2+ ions
and negatively charged groups on the biosorbent surface.
The FTIR spectrum pattern of the biosorbent material before and
after biosorption process is shown in Fig. 2. FTIR spectrum of C.
melo biomass displays a number of absorption bands, reflecting the
complex biomass structure (Fig. 2(a)). The broad stretching absorp-
tion peak at about 3400 cm−1 represents NH and bounded OH
groups. The bands at 2925 and 2854 cm−1 are assigned to symmet-
ric and asymmetric stretching vibrations of CH3 and CH2 groups
and their bending vibrations are observed at 1380 and 1462 cm−1 .
The band at 1745 cm−1 corresponds to C O stretching vibration of
carboxyl groups. Other significant band positions of C. melo seeds
are noted at 1654 cm−1 (indicative of C O stretching vibration of
amide II), 1238 cm−1 (indicative of SO3 group), 1163 cm−1 (indica-
tive of P O stretching) and 1099 cm−1 (indicative of C O stretching
of carbonyl groups and the bending vibrations of hydroxyl group).
The FTIR spectrum (Fig. 2(b)) of C. melo seeds exposed to Pb2+ ions
indicated no considerable change or shifts in any of the character-
istic absorbance bands present in unloaded biomass. This current
result implied the possibility that Pb2+ biosorption onto C. melo
biomass could be taken place through other mechanism rather than
complexation.
SEM and EDX analysis have been extensively used to identify
the possible metal–biosorbent interactions. In order to examine
the textural structure of biomass, SEM micrographs were taken Fig. 3. SEM micrographs for C. melo biomass before (a) and after (b) Pb2+ biosorption.
before (Fig. 3(a)) and after (Fig. 3(b)) Pb2+ biosorption onto C. melo
biomass. SEM micrograph of the unloaded biosorbent indicates
S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90 85

Fig. 4. EDX spectra for C. melo biomass before (a) and after (b) Pb2+ biosorption.

rough, irregular and porous structure of surface. SEM micrograph constant (p > 0.05). Hence further biosorption studies were carried
of Pb2+ -loaded biosorbent surface is significantly different from out at the original pH value of Pb2+ solutions (∼5.5) as the opti-
that of unloaded biosorbent. Many new shiny small particles was mum pH value. Biosorption studies were not conducted above pH
observed over the surface of biosorbent after the biosorption pro- 5.5 because the neutral hydrolysis product of Pb2+ (Pb(OH)2 ) was
cess. This observation evidenced the surface coverage of biosorbent formed as a solid phase in the medium.
by Pb2+ ions. EDX spectra of the unloaded and Pb2+ loaded biosor-
bent were included in Fig. 4(a) and (b), respectively. New peaks
3.3. Effect of biosorbent concentration
revealed between 2 and 3 keV (Fig. 4(b)) corresponding to lead ions
confirm the coverage of the biosorbent surface by Pb2+ ions. Dis-
Fig. 6 shows the variation of the biosorption yield of Pb2+ versus
appeared signals of Mg (at about 1.3 keV) and K (between 3 and
various biosorbent dosages ranging from 0.6 to 4.0 g L−1 . From this
4 keV) showed that an ion-exchange mechanism was included in
Pb2+ biosorption process rather than complexation with functional
groups on the biosorbent surface. 50

3.2. pH effect 40

Among the parameters affecting the biosorption of metals, ini-


tial pH is the most important due to the pH-dependent of the
q (mg g−1)

30
surface characteristics of the biosorbent and ionization degree of
the metals in aqueous medium [2]. To understand the pH effect on
the biosorption of Pb2+ ions onto the biosorbent, biosorption per- 20
formance of the biomass was investigated at pH values in the range
of 2.0–5.5. Fig. 5 shows the relationship between pH and biosorp-
tion capacity of biosorbent. The positive charge of the biosorbent 10
surface at pH 2.0 makes it impossible to bind with Pb2+ ions. There-
fore, it was seen that the biosorption of Pb2+ at pH 2.0 was almost
0
negligible while the biosorption capacity of the biosorbent started 1 2 3 4 5 6
to increase above this pH value (p < 0.05). This may be explained by
pH
the competition between H+ and Pb2+ ions caused to decreased
levels of biosorption at low pH values. At pH 4.0 the biosorp- Fig. 5. Biosorption capacities of C. melo biomass at different pH values (m: 0.1 g; V:
tion capacity attained maximum level and above it stayed almost 50 mL; t: 60 min, T: 20 ◦ C). The bars represent the standard error of the mean.
86 S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90

4
o 2
20 C y = -0.0487x + 3.6157 r = 0.856
100 o 2
30 C y = -0.0500x + 3.2978 r = 0.648
3 o 2
40 C y = -0.0458x + 3.3205 r = 0.836

80
Biosorption yield (%)

ln (qe-qt)
60

0
40
-1
20
-2

0 -3
0 1 2 3 4 5 0 20 40 60 80 100
m (g L−1) t (min)

Fig. 6. Effect of biosorbent concentration on the biosorption of Pb2+ ions onto C. melo Fig. 8. Pseudo-first-order kinetic plots for the biosorption of Pb2+ ions onto C. melo
biomass (pH: 5.5; V: 50 mL; t: 60 min, T: 20 ◦ C). The bars represent the standard error biomass at different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; T: 20 ◦ C).
of the mean.

liquid phases [25,26]. Therefore, the optimum contact time was


figure it was observed that the biosorption yield firstly increased selected as 80 min for further experiments. This relatively short
from 32.94 ± 0.70% to 94.65 ± 1.20% with an increase in the biosor- contact time may also provide an advantage for the large-scale
bent concentration from 0.6 to 1.8 g L−1 (p < 0.05). This is due to application of the biosorption method to real wastewater treat-
an increase in the surface area of the biosorbent, which conse- ment systems.
quently increases the number of binding sites [23]. Further increase Fig. 7 also describes the temperature effect on the biosorption
in biosorbent concentration did not cause much increase in the capacity of the biomass at various contact times. The tempera-
biosorption yield for Pb2+ (p > 0.05). This can be explained by the ture positively affected the biosorption efficiency of biosorbent
decreased active binding sites on the biosorbent as a result of partial for Pb2+ ions. The experimental equilibrium biosorption capacity
aggregation at higher amount of biosorbent [24]. Therefore 1.8 g L−1 increased from 52.09 ± 1.47 to 75.88 ± 0.75 mg g−1 (p < 0.05) when
was selected as optimum biosorbent concentration that would be the temperature was raised from 20 to 40 ◦ C. Therefore, the Pb2+
required for cost-effective treatment of Pb2+ ions. biosorption occurred endothermically. The similar results for the
metal biosorption onto the various biosorbents were reported in
3.4. Equilibrium time and temperature effect the literature [27–30].

Fig. 7 shows the equilibrium time profile of the biosorption 3.5. Biosorption kinetics
capacity of biomass versus contact time at different temperatures.
The biosorption was rapid in the initial stages of the process and With respect to the kinetic modeling of Pb2+ biosorption, the
increased with an increase in contact time up to 80 min (p < 0.05). pseudo-first-order and the pseudo-second-order rate equations
After this period the biosorption capacity of biomass did not sig- have been used.
nificantly change up to 90 min (p > 0.05). This could be attributed The pseudo first-order model of Lagergren [31] is presented by
to a large number of vacant binding sites, which are available for the following equation:
biosorption during the initial stage, and after an increase in contact
time, the occupation of the remaining vacant sites will be difficult ln(qe − qt ) = ln qe − k1 t (2)
due to the repulsive forces between the Pb2+ ions on the solid and
The plot of ln(qe − qt ) against t provides a linear relationship from
which k1 , rate constant of pseudo-first-order biosorption (min−1 )
100 (Fig. 8) and qe , biosorption capacity at equilibrium (mg g−1 ) are
determined from the slope and intercept of the plot, respectively,
given that qt is the biosorption capacity at time t, (mg g−1 ).
80
The biosorption kinetics can also be described by the pseudo-
second-order rate equation [32];
t
qt (mg g )

60 1 1
−1

= + t (3)
qt k2 q2e qe
40 The initial biosorption rate, h (mg g min−1 ), is presented by:

h = k2 q2e (4)
20
where k2 is the equilibrium rate constant of pseudo-second-order
biosorption (g mg−1 min−1 ). Values of k2 and qe were calculated
0 from the plot of t/qt against t (Fig. 9).
0 20 40 60 80 100 The pseudo-first-order and the pseudo-second-order kinetic
t (min) models were tested in the temperature range of 20–40 ◦ C and
the results were represented in Table 1. This table indicated the
Fig. 7. Effect of contact time on the biosorption of Pb2+ ions onto C. melo biomass at
different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; T: 20 ◦ C). The bars represent
pseudo-first-order kinetic model did not identify the Pb2+ biosorp-
the standard error of the mean. tion in this study. The r2 values of this model were lower than those
S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90 87

Table 1
Kinetic parameters for the biosorption of Pb2+ ions onto C. melo biomass at different temperatures.

T (◦ C) Pseudo-first-order Pseudo-second-order

qe (exp) k1 (min−1 ) qe (mg g−1 ) r12 k2 (g mg−1 min−1 ) qe (mg g−1 ) h (mg g min−1 ) r22

20 52.09 4.87 × 10−2 37.18 0.856 5.07 × 10−4 70.20 2.50 0.997
30 69.55 5.01 × 10−2 27.05 0.648 9.77 × 10−4 81.04 6.42 0.998
40 75.88 4.58 × 10−2 27.68 0.836 1.67 × 10−3 83.61 11.67 0.998

for the pseudo-second-order model. The biosorption of Pb2+ on ‘RL ’, referred to as separation factor or equilibrium parameter and
the biomass follows the pseudo-second-order kinetic model with RL is defined by the following equation:
the higher (>0.996) r2 values for all temperatures. The biosorp-
1
tion of Pb2+ was endothermic since the biosorption capacity values RL = (7)
1 + KL C0
increased with increasing temperature. In addition, temperature
also affected the kinetics of Pb2+ biosorption in such a way that where C0 is the initial solute concentration (mol L−1 ). If the RL values
the rate constant (overall and initial) increased with tempera- lie between 0 and 1, the biosorption process is considered to be
ture. The values of k2 were found to increase from 5.07 × 10−4 to favorable.
1.67 × 10−3 g mg−1 min−1 when the temperature increased from 20 The linear Dubinin–Radushkevich isotherm model can be rep-
to 40 ◦ C. resented as [36]:

3.6. Effect of initial metal concentration ln qe = ln qm − ˇε2 (8)

where ε = RT ln(1 + 1/Ce ) (Polanyi potential), qm is the biosorp-


The effect of initial Pb2+ ion concentration was investigated in tion capacity (mol g−1 ), ˇ is the activity coefficient related to the
the range of 50–250 mg L−1 at predetermined optimum conditions. biosorption energy, R is the gas constant and T is the absolute tem-
Equilibrium data for the biosorption Pb2+ ions were evaluated with perature (K). D–R isotherm constants, ˇ and qm can be calculated
Freundlich, Langmuir and Dubinin–Radushkevich (D–R) isotherm from the slope and intercept of the plot of ln qe against ε2 , respec-
models. tively and the mean free energy (E) can be calculated from the
The linearized form of Freundlich equation [33] is: ˇ-value obtained from the following equation [37–39]:
1 1
ln qe = ln KF + ln Ce (5) E= (9)
n 1/2
(2ˇ)
where KF (L g−1 )
and n (dimensionless) are Freundlich constants
and indicators of the biosorption capacity and biosorption intensity, From the magnitude of E, the type of biosorption such as chemisorp-
respectively. tion or physical sorption can be determined. If E = 8–16 kJ mol−1 ,
The well known expression of the Langmuir model is [34]: then the reaction is due to the chemical ion-exchange; if
  1 E < 8 kJ mol−1 , then biosorption takes place physically.
1 1 1 The Freundlich, Langmuir and D–R isotherms for the
= + (6)
qe qmax qmax KL Ce removal of Pb2+ onto the waste biosorbent were presented in
where qe (mol g−1 ) and Ce (mol L−1 ) are the amount of biosorbed Figs. 10, 11 and 12, respectively. The calculated model constants
Pb2+ per unit weight of biosorbent and residual Pb2+ concentra- and r2 values are included in Table 2. These data indicated that
tion in solution at equilibrium, respectively. qmax (mol g−1 ) is the Langmuir isotherm model well described the experimental data
maximum amount of Pb2+ per unit weight of biosorbent to form a at all studied temperatures, meaning that the surface of the
complete monolayer on the surface and KL (L mol−1 ) is a constant biosorbent is made up of homogeneous biosorption patches. The
related to the energy of biosorption. maximum monolayer biosorption capacity of the biosorbent was
Hall et al. [35] reported that the essential feature of the Langmuir obtained as 3.64 × 10−4 mol g−1 (75.42 mg g−1 ) at 20 ◦ C and found
isotherm can be expressed by means of a dimensionless constant
-7,6
o
20 C y = 0,3211x -5,6729 r ² = 0,872
1,8 o
-7,8 30 C y = 0,2317x - 6,1437 r ² = 0,775
o
1,6 40 C y = 0,2066x - 6,2277 r ² = 0,847
-8,0
1,4
-8,2
t/qt (min g mg -1)

1,2
-8,4
lnqe

1,0
-8,6
0,8
-8,8
0,6
-9,0
0,4
-9,2
0,2
-9,4
0,0 -13 -12 -11 -10 -9 -8 -7
0 20 40 60 80 100 lnCe
t (min)
Fig. 10. Freundlich isotherm plots for the biosorption of Pb2+ ions onto C. melo
Fig. 9. Pseudo-second-order kinetic plots for the biosorption of Pb2+ ions onto C. biomass at different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; t: 80 min; T:
melo biomass at different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; T: 20 ◦ C). 20 ◦ C).
88 S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90

Table 2
Isotherm constants for the biosorption of Pb2+ ions onto C. melo biomass at different temperatures.

T (◦ C) Langmuir Freundlich Dubinin–Radushkevich (D–R)

qmax (mol g−1 ) KL (L mol−1 ) rL2 n KF (L g−1 ) rF2 qmax (mol g−1 ) ˇ (mol2 kJ−2 ) rD2 –R E (kJ mol−1 )

20 3.38 × 10−4 1.32 × 104 0.992 3.11 3.44 × 10−3 0.872 8.79 × 10−4 1.77 × 10−3 0.903 16.80
30 3.63 × 10−4 4.17 × 104 0.992 4.32 2.15 × 10−3 0.775 7.71 × 10−4 1.42 × 10−3 0.824 18.77
40 3.64 × 10−4 9.52 × 104 0.999 4.84 1.97 × 10−3 0.847 7.58 × 10−4 1.27 × 10−3 0.891 19.88

10000 Table 3
Biosorption results of Pb2+ ions by different biosorbents from the literature.
9000
Biosorbent material Biosorption References
8000 capacity (mg g−1 )

Symphoricarpus albus 62.16 [7]


1/qe (g mol )
−1

7000 Chaff 12.50 [15]


Ficus religiosa leaves 37.45 [16]
6000 Cladonia furcata 12.30 [24]
Phaseolus vulgaris L. 42.77 [27]
5000 Seed husk of Calophyllum inophyllum 34.51 [28]
Penicillium simplicissimum 87.72 [29]
4000 Nostoc sp. 93.46 [30]
o
20 C y = 0,2250x + 2959,1387 r ² = 0,992 Aspergillus flavus 13.46 [40]
3000 o
30 C y = 0,0659x + 2749,0200 r ² = 0,992 Bacillus sp. 92.27 [41]
o
40 C y = 0,0288x + 2744,7379 r ² = 0,999
C. melo 75.42 This study
2000
0,0 2,0e+4 4,0e+4 6,0e+4 8,0e+4 1,0e+5 1,2e+5 1,4e+5 1,6e+5 1,8e+5

1/Ce (L mol−1) 0,30


20 C
30 C
Fig. 11. Langmuir isotherm plots for the biosorption of Pb2+ ions onto C. melo
0,25 40 C
biomass at different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; t: 80 min; T:
20 ◦ C).
0,20
RL

to be higher than that of and comparable to many biosorbents 0,15


reported in the literature (Table 3). The RL values in this study
were obtained at the range of 0.250–0.059, 0.095–0.020 and 0,10
0.042–0.008 at the temperatures of 20, 30 and 40 ◦ C, respectively
(Fig. 13), indicating that the biosorption process is favorable. In 0,05
addition, RL values decreased as the initial Pb2+ concentration
increased. Therefore, the biosorption of Pb2+ was more favorable
0,00
at higher concentration. 0 50 100 150 200 250 300
In order to determine the type of biosorption, the equilib- -1
Co (mg L )
rium biosorption data were also tested with the D–R model.
The mean free energy of biosorption (E) is found between 16.80
Fig. 13. RL values for the biosorption of Pb2+ ions onto C. melo biomass at different
and 19.88 kJ mol−1 at studied temperature range, which implies temperatures and initial Pb2+ concentrations.
that the removal process of Pb2+ from aqueous solutions is
chemisorption.
3.7. Thermodynamics of biosorption

Pb2+ biosorption may be represented by the following reversible


-7,6 process:
o
20 C y = -3,141e-9x - 7,0366 r ² = 0,903
-7,8 o
30 C y = -2,015e-9x - 7,1689 r ² = 0,824 Pb2+ in solution  Pb2+ -biosorbent
o
40 C y = -1,601e-9x - 7,1846 r ² = 0,891
-8,0
For such equilibrium reactions, KL (Langmuir constant) can be
-8,2 used as equilibrium constant in order to estimate thermodynamic
parameters, such as change in the free energy (G◦ ), enthalpy
-8,4
(H◦ ) and entropy (S◦ ) associated to the biosorption process.
lnqe

-8,6 They were determined by using following equations and repre-


sented in Table 4.
-8,8

-9,0 G◦ = −RT ln KL (10)

-9,2
Table 4
-9,4 Thermodynamic parameters for the biosorption of Pb2+ ions onto C. melo biomass.
2,0e+8 4,0e+8 6,0e+8 8,0e+8 1,0e+9 1,2e+9
T (◦ C) G◦ (kJ mol−1 ) H◦ (kJ mol−1 ) S◦ (kJ K−1 mol−1 )
ε (J mol )
2 -1 2
20 −23.22
30 −26.59 75.67 0.337
Fig. 12. D–R isotherm plots for the biosorption of Pb2+ ions onto C. melo biomass at
40 −29.97
different temperatures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; t: 80 min; T: 20 ◦ C).
S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90 89

12,0
y = -9101,908x + 40,5755 r ² = 0,994

11,5

11,0
lnKL

10,5

10,0

9,5

9,0
0,00315 0,00320 0,00325 0,00330 0,00335 0,00340 0,00345
1/T (K-1)

Fig. 14. Plot of ln KL versus 1/T for estimation of thermodynamic parameters for the Fig. 16. Competitive biosorption results for the biosorption of Pb2+ ions onto C. melo
biosorption of Pb2+ ions onto C. melo biomass. biosorbent in binary and multi metal mixtures (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; t:
80 min; T: 20 ◦ C). The bars represent the standard error of the mean.
G◦ H ◦ S ◦
ln KL = − =− + (11)
RT RT R
sharply decreased from 80.48± to 57.85± mg g−1 (p < 0.05) when
where R is the universal gas constant, 8.314 J mol−1 K−1 , and T is the ionic strength of the biosorption medium was increased from
the absolute temperature in K. The plot of ln KL as a function of 0.01 to 0.10 mol L−1 . The biosorption potential of the biosorbent
1/T (Fig. 14) yields a straight line from which H◦ and S◦ were was almost constant up to 0.50 mol L−1 of KNO3 (p > 0.05). This
calculated from the slope and intercept, respectively. The negative can be explained by two ways. Firstly, the competition of the K+
values of G◦ at all temperatures studied indicate the Pb2+ biosorp- ions with Pb2+ ions on the biosorbent surface caused a decrease in
tion is a spontaneous process and the biosorbent has higher affinity the biosorption performance of the biosorbent. Secondly, the elec-
at higher temperatures. A decrease in the magnitude of G◦ values trostatic interaction between Pb2+ ions and the active biosorption
with an increase in the temperature may be attributed to a reduc- sites on the biosorbent can be other mechanism for the biosorption
tion in the spontaneity degree at studied temperatures. The positive process. This findings also agree with the literature [43,44].
value of H◦ suggests the endothermic nature of biosorption. The
positive S◦ value indicates that randomness at the solid-solute
interface increases with Pb2+ biosorption onto the biosorbent [42], 3.9. Competitive biosorption
suggests good affinity of Pb2+ toward the biosorbent and reflects
some structural changes in biosorbate and biosorbent. The biosorption yield of Pb2+ onto the biosorbent in the presence
of (Cd2+ , Co2+ and Ni2+ ) at different concentrations was evaluated.
3.8. Effect of ionic strength The experiments were carried out in binary- and multi-metal mix-
ture conditions. The results were presented in Fig. 16 and indicated
The effect of ionic strength on the Pb2+ biosorption was inves- that the presence of other metal ions reduced the Pb2+ biosorption
tigated by changing KNO3 concentration from 0.01 mol L−1 from yield of the biomass. The reduction ratio of biosorption yield of the
0.50 mol L−1 at pH 5.5. Fig. 15 shows that the influence of the ionic biosorbent slightly increased when the concentration of co-ions
strength on the biosorption of Pb2+ is almost negligible at 0.01 was increased from 5 to 100 mg L−1 . The presence of (Cd2+ , Co2+ and
and 0.02 mol L−1 of KNO3 concentrations. The biosorption capacity Ni2+ ) in binary systems reduced the biosorption performance of the
biosorbent for Pb2+ by 28.05%, 12.13% and 25.25%, respectively. In
the simultaneous presence of all metal ions it decreased by 30.50%.
100
This observation could be attributed to the competition between
Pb2+ and the other co-ions for the active biosorption sites on the
80
biosorbent surface. Similar effects of co-cations in the biosorption
Biosorption yield (%)

medium were reported in the literature [40,41,45–47].

60
4. Conclusions

40
This study reveals:

• pH 5.5 was suitable for the Pb2+ removal onto the waste biosor-
20 bent of C. melo, 1.8 g L−1 of biosorbent was enough for the
maximum biosorption of Pb2+ ,
• The pseudo-second-order rate equation best described the Pb2+
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6
biosorption kinetics. The equilibrium data for Pb2+ biosorption
fitted well to the Langmuir model,
Csalt (mol L−1) • An endothermic process was responsible for the biosorption of
Pb2+ on the biomass,
Fig. 15. Effect of ionic strength on the biosorption of Pb2+ ions onto C. melo biosor- • Competitive conditions indicated that the biosorption yield was
bent (pH: 5.5; V: 50 mL; m: 1.8 g L−1 ; t: 80 min; T: 20 ◦ C). The bars represent the
standard error of the mean.
decreased by 30.50% in multi-metal mixtures,
90 S. Tunali Akar et al. / Chemical Engineering Journal 185–186 (2012) 82–90

• Finally, characterization studies show the Pb2+ biosorption takes [22] J.A. Duke, E.S. Ayensu, Medicinal Plants of China, Reference Publications, Inc.,
place with mechanisms including especially electrostatic attrac- 1985.
[23] R. Razmovski, M. Šćiban, Biosorption of Cr(VI) and Cu(II) by waste tea fungal
tion forces and ion-exchange. biomass, Ecol. Eng. 34 (2008) 179–186.
[24] A. Sarı, M. Tuzen, Ö.D. Uluözlü, M. Soylak, Biosorption of Pb(II) and Ni(II) from
References aqueous solution by lichen (Cladonia furcata) biomass, Biochem. Eng. J. 37
(2007) 151–158.
[25] T. Akar, Z. Kaynak, S. Ulusoy, D. Yuvaci, G. Ozsari, S. Tunali Akar, Enhanced
[1] V.K. Gupta, I. Ali, Removal of lead and chromium from wastewater using biosorption of nickel(II) ions by silica-gel-immobilized waste biomass: Biosorp-
bagasse fly ash a sugar industry waste, J. Colloid Interface Sci. 271 (2004) tion characteristics in batch and dynamic flow mode, J. Hazard. Mater. 163
321–328. (2009) 1134–1141.
[2] B. Volesky, Removal and recovery of heavy metals by biosorption, in: B. Volesky [26] K. Vijayaraghavan, J. Mao, Y.S. Yun, Biosorption of methylene blue from aqueous
(Ed.), Biosorption of Heavy Metals, CRC Press, Boca Raton, FL, 1990, pp. 3–43. solution using free and polysulfone-immobilized Corynebacterium glutamicum:
[3] P.-C. Hsu, Y.-L. Guo, Antioxidant nutrients and lead toxicity, Toxicology 180 batch and column studies, Bioresour. Technol. 99 (2008) 2864–2871.
(2002) 33–44. [27] A.S. Özcan, S. Tunali, T. Akar, A. Özcan, Biosorption of lead(II) ions onto waste
[4] A. Keen, Lead and the London Metal Exchange a happy marriage? The outlook biomass of Phaseolus vulgaris L. estimation of the equilibrium, kinetic and ther-
for prices and pricing issues confronting the lead industry, J. Power Sources 88 modynamic parameters, Desalination 244 (2009) 188–198.
(2000) 27–35. [28] O.S. Lawal, A.R. Sanni, I.A. Ajayi, O.O. Rabiu, Equilibrium, thermodynamic and
[5] S.S. Ahluwalia, D. Goyal, Microbial and plant derived biomass for removal kinetic studies for the biosorption of aqueous lead(II) ions onto the seed husk
of heavy metals from wastewater, Bioresour. Technol. 98 (2007) 2243– of Calophyllum inophyllum, J. Hazard. Mater. 177 (2010) 829–835.
2257. [29] T. Fan, Y. Liu, B. Feng, G. Zeng, C. Yang, M. Zhou, H. Zhou, Z. Tan, X. Wang,
[6] D. Sud, G. Mahajan, M.P. Kaur, Agricultural waste material as potential adsor- Biosorption of cadmium(II), zinc(II) and lead(II) by Penicillium simplicissimum:
bent for sequestering heavy metal ions from aqueous solutions—a review, isotherms, kinetics and thermodynamics, J. Hazard. Mater. 160 (2008) 655–661.
Bioresour. Technol. 99 (2008) 6017–6027. [30] V.K. Gupta, A. Rastogi, Biosorption of lead(II) from aqueous solutions by non-
[7] S. Tunali Akar, A. Gorgulu, B. Anilan, Z. Kaynak, T. Akar, Investigation of the living algal biomass Oedogonium sp. and Nostoc sp.—a comparative study,
biosorption characteristics of lead(II) ions onto Symphoricarpus albus: batch Colloids Surf. B 64 (2008) 170–178.
and dynamic flow studies, J. Hazard. Mater. 165 (2009) 126–133. [31] S. Lagergren, Zur theorie der sogenannten adsorption gelöster stoffe Kungliga
[8] S. Chowdhury, P.D. Saha, Biosorption kinetics, thermodynamics and isosteric Svenska Vetenskapsakademiens, Handlingar 24 (1898) 1–39.
heat of sorption of Cu(II) onto Tamarindus indica seed powder, Colloids Surf. B [32] Y.S. Ho, G. McKay, Kinetic models for the sorption of dye from aqueous solution
88 (2011) 697–705. by wood, Process Saf. Environ. Prot. 76 (B2) (1998) 183–191.
[9] G. Blazquez, M. Calero, F. Hernainz, G. Tenorio, M.A. Martin-Lara, Equilibrium [33] H.M.F. Freundlich, Über die adsorption in lösungen, Z. Phys. Chem. 57 (1906)
biosorption of lead(II) from aqueous solutions by solid waste from olive-oil 385–470.
production, Chem. Eng. J. 160 (2010) 615–622. [34] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and plat-
[10] S.W. Al Rmalli, A.A. Dahmani, M.M. Abuein, A.A. Gleza, Biosorption of mercury inum, J. Am. Chem. Soc. 40 (1918) 1361–1403.
from aqueous solutions by powdered leaves of castor tree (Ricinus communis [35] K.R. Hall, L.C. Eagleton, A. Acrivos, T. Vermeulen, Pore- and solid-diffusion kinet-
L.), J. Hazard. Mater. 152 (2008) 955–959. ics in fixed-bed adsorption under constant pattern conditions, Ind. Eng. Chem.
[11] A.S. Ozcan, A. Ozcan, S. Tunali, T. Akar, I. Kiran, T. Gedikbey, Adsorption potential Fundam. 5 (1966) 212–223.
of lead(II) ions from qaueous solutions onto Capsicum annuum seeds, Sep. Sci. [36] M.M. Dubinin, L.V. Radushkevich, Proc. Acad. Sci. USSR Phys. Chem. Sect. 55
Technol. 42 (2007) 137–151. (1947) 331–333.
[12] M.A. Hanif, R. Nadeem, M.N. Zafar, K. Akhtar, H.N. Bhatti, Kinetic studies for [37] J.P. Hobson, Physical adsorption isotherms extending from ultrahigh vacuum
Ni(II) biosorption from industrial wastewater by Cassia fistula (Golden Shower) pressure, J. Phys. Chem. 73 (1969) 2720–2727.
biomass, J. Hazard. Mater. 145 (2007) 501–505. [38] S.M. Hasany, M.H. Chaudhary, Sorption potential of Hare River sand for the
[13] A. Witek-Krowiak, R.G. Szafran, S. Modelski, Biosorption of heavy metals from removal of antimony from acidic aqueous solution, Appl. Radiat. Isot. 47 (1996)
aqueous solutions onto peanut shell as a low-cost biosorbent, Desalination 265 467–471.
(2011) 126–134. [39] S.S. Dubey, R.K. Gupta, Removal behavior of Babool bark (Acacia nilotica) for
[14] W.E. Oliveira, A.S. Franca, L.S. Oliveira, S.D. Rocha, Untreated coffee husks as submicro concentrations of Hg2+ from aqueous solutions: a radiotracer study,
biosorbents for the removal of heavy metals from aqueous solutions, J. Hazard. Sep. Purif. Technol. 41 (2005) 21–28.
Mater. 152 (2008) 1073–1081. [40] T. Akar, S. Tunali, Biosorption characteristics of Aspergillus flavus biomass for
[15] R. Han, J. Zhang, W. Zou, J. Shi, H. Liu, Equilibrium biosorption isotherm for lead removal of Pb(II) and Cu(II) ions from an aqueous solution, Bioresour. Technol.
ion on chaff, J. Hazard. Mater. 125 (2005) 266–271. 97 (2006) 1780–1787.
[16] S. Qasier, A.R. Saleemi, M. Umar, Biosorption of lead from aqueous solution by [41] S. Tunali, A. Çabuk, T. Akar, Removal of lead and copper ions from aqueous solu-
Ficus religiosa leaves: batch and column study, J. Hazard. Mater. 166 (2009) tions by bacterial strain isolated from soil, Chem. Eng. J. 115 (2006) 203–211.
998–1005. [42] N. Tewaria, P. Vasudevana, B.K. Guhab, Study on biosorption of Cr(VI) by Mucor
[17] A. Saeed, M. Iqbal, M.W. Akhtar, Removal and recovery of lead(II) from single hiemalis, Biochem. Eng. J. 23 (2005) 185–192.
and multi metal (Cd, Cu, Ni, Zn) solutions by crop milling waste (black gram [43] T.A. Davis, B. Volesky, A. Mucci, A review of the biochemistry of heavy metal
husk), J. Hazard. Mater. 117 (2005) 65–73. biosorption by brown algae, Water Res. 37 (2003) 4311–4330.
[18] F. Stagnari, M. Pisante, Managing faba bean residues to enhance the fruit quality [44] J.E.B. Cayllahua, M.L. Torem, Biosorption of aluminum ions onto Rhodococcus
of the melon (Cucumis melo L.) crop, Sci. Hortic. 126 (2010) 317–323. opacus from wastewaters, Chem. Eng. J. 161 (2010) 1–8.
[19] C. Fabeiro, F.M. de Santa Olalla, J.A. de Juan, Production of muskmelon (Cucumis [45] T. Akar, A. Cabuk, S. Tunali, M. Yamac, Biosorption potential of the macrofungus
melo L.) under controlled deficit irrigation in a semi-arid climate, Agric. Water Ganoderma carnosum for removal of lead(II) ions from aqueous solutions, J.
Manage. 54 (2002) 93–105. Environ. Sci. Health A 41 (2006) 2587–2606.
[20] K. Tanaka, A. Nishitani, Y. Akashi, Y. Sakata, H. Nishida, H. Yoshino, K. Kato, [46] T. Akar, S. Tunali, Biosorption performance of Botrytis cinerea fungal by-
Molecular characterization of South and East Asian melon, Cucumis melo L., products for removal of Cd(II) and Cu(II) ions from aqueous solutions, Miner.
and the origin of Group Conomon var. makuwa and var. conomon revealed by Eng. 18 (2005) 1099–1109.
RAPD analysis, Euphytica 153 (2007) 233–247. [47] A.H. Hawari, C.N. Mulligan, Effect of the presence of lead on the biosorption of
[21] M.L.S. de Mello, P.S. Bora, N. Narain, Fatty and amino acids composition of melon copper, cadmium and nickel by anaerobic biomass, Process Biochem. 42 (2007)
(Cucumis melo var. saccharinus) seeds, J. Food Comp. Anal. 14 (2001) 69–74. 1546–1552.

You might also like