Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Biosorbent from tomato waste and apple juice residue for lead removal T
a,b,⁎ a,b a a
Eddy Heraldy , Witri Wahyu Lestari , Diah Permatasari , Devita Dwi Arimurti
a
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Sebelas Maret University, Jl. Ir. A. Sutami 36A, Kentingan, Surakarta 57126, Indonesia
b
Porous Materials Reseach Group, Chemistry Department, Sebelas Maret University, Jl. Ir. A. Sutami 36A, Kentingan, Surakarta 57126, Indonesia

A R T I C L E I N F O A B S T R A C T

Keywords: The biosorbents from different fruit waste such as tomato waste and apple juice residue were prepared and had
Biosorbents been investigated in removing of Pb(II) ion from the aqueous solution under optimized conditions. All fruit waste
Apple materials were activated by NaOH in room temperature. The synthesized biosorbents were characterized by FT-
Tomato IR, SEM, and surface area analyzer. Effects of contact time, the pH of the solution, and sorbent dosage on the Pb
Lead
(II) removal process had been optimized. Freundlich isotherm matched the experimental data than Langmuir
isotherm model. A kinetic study presented that Pb(II) sorption follows the pseudo-second-order kinetics. The Pb
(II) removal capacity of tomato waste and apple juice residue biosorbents are 152 and 108 mg/g, respectively.
The results showed that biosorbents have the potential capacity to remove the lead, which further strongly
increases after its activation.

1. Introduction [10,11]. The equilibrium sorption was modeled using the Freundlich
and Langmuir isotherm. The behavior of sorption with the passage of
The heavy metal ions removal from wastewater such as Pb(II) is time was investigated using kinetic studies.
imperative due to their bioaccumulation tendency, toxicity, persistence,
and non–biodegradability in nature [1,2]. Lead accumulation in the 2. Materials and methods
human body also causes reproductive problems, central nervous system
damage [3], retard in the physical and mental activities, mainly in the 2.1. Materials
childhood age. Besides, its potential carcinogenic and neurological ef-
fects [4]; it could even cause death. Therefore, an increased interest had All chemicals and reagents are used in analytical grade without
been focused on effective studies of Pb(II) ion removal. In wastewater, further purification. The pH of solutions was adjusted by addition of
Pb(II) mainly come from metal plating and finishing manufacturing, either HCl or NaOH 0.1 M solution. The fruit waste material required as
battery manufacturing, soldering material, printing and pigment, glass starting material for this work was collected locally. Mainly, the apple
industries and ceramic, ammunition, steel and iron manufacturing units juice residue (AR) was obtained from the destruction of locally apple
[5]. In literature, numerous treatment methods of Pb(II) wastewater (“manalagi”) [12]; while the biomass tomato wastes (TW) was collected
have been reported. These methods include membrane separation, ion from “Pasar Gede Surakarta” market [13].
exchange, chemical precipitation and adsorption [6]. For example, Initially, for the preparation of biosorbents, the fruit wastes were
adsorption was regarded as the superior technique method related with crushed, washed several times with distilled water, and dried in the
low lead concentrations in wastewater. Various adsorbents, for in- oven at 70 °C, respectively. Then, they were washed properly using
stance, activated carbon, montmorillonite, zeolite, wheat straw, waste distilled water and dried at 70 °C for 48 h until the biosorbent constant
biomass, hydrotalcite [7–9] were used. The present study aimed to weight is obtained. Dried-fruit wastes were crushed and sieved to
investigate the possibility of apple juice residue and tomato waste as an 100 mesh sizes.
effective and low-cost sorbent for Pb(II) ion removal. The variables
operation effect such as initial lead ion concentration, adsorbent dose, 2.2. Procedure
solution pH, contact time was investigated. Biosorbents could be in-
creasing their adsorption capacity through chemical modification. 2.2.1. Preparation and modification of biosorbents
Treatment with acid or alkali was the common modification methods NaOH reagent was used to increase the proportion of active


Corresponding author at: Department of Chemistry, Faculty of Mathematics and Natural Sciences, Sebelas Maret University, Jl. Ir. A. Sutami 36A, Kentingan, Surakarta 57126,
Indonesia.
E-mail address: eheraldy@mipa.uns.ac.id (E. Heraldy).

https://doi.org/10.1016/j.jece.2017.12.026
Received 18 August 2017; Received in revised form 31 October 2017; Accepted 13 December 2017
Available online 09 January 2018
2213-3437/ © 2018 Elsevier Ltd. All rights reserved.
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Fig. 1. FTIR spectra of AR biosorbent (A) and TW biosorbent (B) before and after activation and Pb(II) loaded.

surfaces. For samples activation, 10 g biosorbents was poured in dif- certain time intervals, and the concentration of Pb(II) was similarly
ferent beakers containing 100 mL NaOH, 1:10 (m/v) and stirred thor- measured. The quantity of sorption at time t, was computed using the
oughly for 24 h at room temperature. Then, the solution was filtered, equation:
and the activated samples were taken and washed several times using
(Ci − Ct ) V
distilled water until the washing solution pH becomes neutral. Then the qt =
W (3)
biosorbents was oven dried at 70 °C for 24 h. The final product was
sieved through 100 mesh sieve again to ensure uniform particle size. where Ct (mg/L) is the concentration of Pb(II) at any time.

2.2.2. Batch sorption experiments design 2.2.4. Biosorbents characterization


The batch sorption experiment was carried out in a set 250 mL To detect the functional groups present in the prepared biosorbent
conical flask containing 25 mL of Pb(NO3)2 solution to investigate the and interaction between lead and the active functional group on the
effect of biosorbent dosage (0.0125, 0,0250, 0.0500, 0.0750, 0,1000 biosorbent surface were analyzed by Fourier transform IR (FTIR)
and 0.1250 g); pH (2, 4, 6, 8 and 10), contact time (15, 30, 45, 60 and spectrometry. FTIR spectra were recorded using Shimadzu FTIR model
75 min) and initial Pb(II) concentration (20, 40, 60, 80 and 100 ppm) at 8201. The sample of dried biosorbents was grinded sufficiently with
constant agitation speed at 120 rpm and at the finish of pre-determined KBr by pallet method. The spectra were recorded in the range of
time period the reaction mixtures were filtered out. The concentration 4000–400 cm−1. The surface morphologies were observed with the aid
of Pb(II) on supernatant was analyzed by atomic absorption spectro- of field-emission scanning electron microscopy (FE-SEM). SEM images
photometer (AAS). Finally, Pb(II) removal efficiency was calculated were taken using the FEI type INSPECT-s50 at an accelerating voltage of
using the follows equation [14]: 6 kV. The specific surface area (SBET) measurements of the materials
were calculated by Brunauer-Emmett-Teller (BET) method and pore
C − Ce ⎫ size distribution were performed by nitrogen (N2) adsorption-deso-
Removal efficiency(%) = ⎧ i × 100
⎩ Ci ⎬
⎨ ⎭ (1) rption isotherms method through a Monosorb direct reading specific
surface analyzer (Nova Win 1200e).
where Ci and Ce are the initial at t = 0 and equilibrium concentration of
Pb(II) (mg/L) in the aqueous solution, respectively. The sorption ca-
3. Results and discussion
pacity was determined using the weight balance equation for the bio-
sorbent by the follows equation:
3.1. Characterization of biosorbents
(Ci − Ce ) V
q=
W (2) 3.1.1. FTIR analysis
FTIR spectra determination was carried out to analyze the potential
where q is the sorption capacity (mg/g), V is the volume of Pb(II) so-
changes in vibration frequency resulting in the functional groups of the
lution (L), and W is the weight of the sorbent (g).
prepared biosorbents and after biosorption. The FTIR spectra of bio-
sorbents before and after activating by NaOH and after Pb(II) sorption
2.2.2.1. Effect of initial pH. Initial pH effect on Pb(II) removal was
are displayed in Fig. 1. In two biosorbent (before activating), the
investigated at different pH values (2, 4, 6, 8 and 10). The initial
spectra are similar which is the band observed around 3440 cm−1 due
concentration of Pb(II) was 20 ppm. The samples were stirred for 1 h at
to OH stretching. The peaks around 2926 and 1746 cm−1 are relating to
room temperature.
CeH symmetrical stretching and C]O stretching that main character of
hemicellulose, respectively. Absorption bands appeared around 1514
2.2.2.2. Effect contact time. At the optimisation of contact time, and 1016–1163 cm−1 are assigned to C]C aromatic stretching from
adsorption process was performed in different shaking times (15, 30, lignin and CeO. While in biosorbents after activated, the peaks appear
45, 60 and 75 min) at an optimum pH value obtained from the previous around 1504 and 1727 cm−1 are corresponding to C]C and C]O bond
section. The initial Pb(II) concentration was 20 ppm. that indicated delignification process [15]. The bands shifting to lower
frequencies showed bond weakening, while conversely, a shift to higher
2.2.2.3. Determine the optimum biosorbent dosage. Amount of biosorbent frequencies suggests an increase in bond strength [16]. It is shown that
was varied with a range of concentration (0.0250, 0.0500, 0.0750, the decreasing in the bond strength of lignin and hemicellulose in cel-
0.1000 and 0.1250 mg/L) in 25 mL of Pb(II). lulose due to the reaction of delignification by NaOH in the biosorbent.
Hence, delignification by NaOH is able to break the lignin in cellulose
2.2.3. Batch kinetic studies and may solve other substances. In Fig. 1, the intensity of absorption
The processes of the batch kinetic studies were fundamental to those bands OeH, CeH and C]O group in FTIR spectra of biosorbent before
of equilibrium experiments. The aqueous solutions were chosen at and after biosorption of Pb(II) decreased due to the interaction of Pb(II)

1202
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Fig. 2. SEM images of biosorbent (for instance AR) before (A)


and after activation (B) at 2000×. Similar behaviour was
observed for TW.

with the functional groups [17]. It was seen that the band at 3435 cm−1 The SEM images of biosorbents before and after Pb(II) ions sorption
shifted to 3492 cm−1 (for AR) and at 3399 cm−1–3463 cm−1 (for TW) are displayed in Fig. 3. SEM analyses were carried out with the purpose
related to OH functional was shifted after sorption, indicating the in- to confirm the sorption of Pb(II) onto NaOH-activated biosorbents.
volvement of OH group with sorption of Pb(II) [18,19]. There is a definite difference in the surface morphology of biosorbents
before and after Pb(II) loaded. It is clear from SEM images that bio-
3.1.2. SEM analysis sorbents (Fig. 3A and C) have a rough and varied morphology of surface
SEM analysis is a helpful tool for portraying the surface morphology with small granules and many holes, wherein there is a good chance for
of the biosorbents. The SEM was performed to analyze the surface Pb(II) ions to sorb. The Pb(II) loaded images (Fig. 3B and D) demon-
modifications of biosorbents before and after NaOH activation. As could strate the secondary solids appearance on the surface that knowingly
be seen in Fig. 2 that a biosorbent before NaOH activation has a dense different from unloaded biosorbent, representing that precipitation of
and rigid morphology, revealing homogeneous without any porous surface took place during sorption. The surface of Pb(II) loaded bio-
structure and a little-damaged surface, caused by liberation of volatile sorbents had various particle sizes; the morphology of surface had al-
substances in carbonization process. The SEM images of biosorbents tered with no other existing space. It is nothing surface spaces, and the
after NaOH activation obviously indicate surface modifications during particles were surfaces closely packed. The result was in agreement
the activation process. The NaOH-activated biosorbents appearance an with the FT-IR analysis (Fig. 1).
irregular surface with cavities of several sizes, demonstrating more
porous with a well-built porous structure, affecting from the vaporiza- 3.1.3. Surface area analysis
tion of impregnated NaOH and derivative compounds. It is clear that The specific surface area and pore volume of biosorbents were de-
biosorbents possesses different and a rough surface morphology, where termined by the Brunauer-Emmett-Teller (BET) methods were pre-
there is a good possibility for sorbates be adsorbed. The cavities on sented in Table 1. The BET surface areas of biosorbents before and after
biosorbents surface proper canals for diffusion of sorbates into the activation were found to be 7.04 and 11.13 m2/g (for AR) as well as
biosorbents, make available access to its micro and mesopores, wherein 8.83 and 45.00 m2/g (for TW), respectively. Activation of biosorbent
they can interact with the surface of functional groups [20]. using NaOH cause delignification process, increasing in total pore

Fig. 3. SEM images of biosorbent before (A = AR; C = TW)


and after Pb(II) loaded (B = AR; D = TW) at 2000×.

1203
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Table 1
BET Surface Area of Biosorbents.

Biosorbent Surface area (m2/g) Pore volume (cc/g) Pore size (Å)

AR 7.04 6.44 × 10−3 1.827


ARac 11.13 8.34 × 10−3 1.499
TW 8.83 8.13 × 10−3 1.841
TWac 45.00 4.47 × 10−2 1.985

volume and surface area. The activation method by NaOH providing the
growth of various pores with different sizes and high surface area. The
pore size founds by BJH method was typical of mesoporous materials.
The TW biosorbent has a good surface area compared with other sor-
Fig. 5. Effect of sorbent dosage on sorption capacity (mg/g) (thin line) and Pb(II) removal
bent, for instance, activated phosphate (22.22 m2/g) [21], raw clin-
(%) (thick line).
optilolite (15.36 m2/g) [22], non-activated charcoal (21.0 m2/g) [23]
used for Pb(II) removal.
sorbents was low at lower pH and increases at higher pH values [25]
had stated same results. It might be justified by the fact that at acidic
3.2. Biosorption parameter studies pH (mainly in pH 2), the H+ concentration is higher and most of the
active site of the biosorbent might be occupied by H+ [26], making the
3.2.1. Effect of pH biosorbents surface becomes more positively charged. In such a situa-
The aqueous solution pH containing the sorbate is an essential tion, thereby being electrostatic repulsion forces for binding of Pb(II)
parameter in the sorption process [24]. The significant role of pH is ions occurs. However, when the pH value increased (from pH 2 to 4),
very well documented in the literature because it may affect not merely the concentration of H+ ions decreases and negatively charged on
the appearance of the compounds dispersed in water but also the biosorbents surface is exposed due to deprotonated of sites which could
group’s dissociation on the surface of the biosorbents [3]. Besides that, interact with Pb(II) ions [19,27–29]. At the pH > 6, the Pb(II) ions in
the efficiency of sorption activity greatly depends on solution pH, due solution might undergo hydrolysis and precipitation as lead hydroxides
to the variation of pH could further changes in the charges and states of [19,30,31].
the functional groups of biosorbent materials as well as sorbate mole-
cules. Therefore, optimizing of sorbent capacity in Pb(II) removal needs 3.2.2. Effect of biosorbents dosage
between other determination of pH value favorable for contact and Biosorbents dosage had a substantial impact on the sorption of the
interaction among sorbent and sorbate. Influence of initial solution pH Pb(II) at a certain initial concentration. Sorption process is related to
on Pb(II) adsorption onto NaOH-activated biosorbents was evaluated in mass transfer phenomenon. Sorption capacity rises with increasing
the pH value of pH 2–10. Fig. 4 demonstrations the pH effect on the contact among sorbent and sorbate. Nevertheless, excessive usage of
percentage of Pb(II) removal. As shown in Fig. 4 at pH 2.0 (low pH), Pb sorbent adds the expense of sorption process, making contact less at-
(II) sorbed a low amount of Pb(II). However, when the initial pH of Pb tractive. The biosorbents dosage dependency used up on the sorption of
(II) solution was raised, the sorption capacity was raised as well. As can Pb(II) was examined, and the results are presented in Fig. 5. As can be
be seen, NaOH-activated biosorbents demonstrated a percentage of Pb seen in Fig. 5 that the sorption of Pb(II) rose with the increase of sor-
(II) removal rises gradually from pH 2.0, reached a maximum nearly bent dosage from 0.025 mg to 0.100 mg/25 mL and reached a max-
97% at pH 4.0 (for TW biosorbents), and around 99% (for AR biosor- imum at 95.6% removal (for TW biosorbents) and 90.8% (for AR bio-
bents) and then decreases that match of previous research. Thus, the sorbents). The increase in sorption was due to the availability of more
sorption experiments for Pb(II) removal using biosorbent fruit waste amount of sorption active sites and functional groups for the sorption of
were not conducted at pH > 4. Hence, the optimum pH for sorption of Pb(II) ions [27]. Subsequently, with the increase of sorbent dosage from
Pb(II) was observed at pH 4.0. At the pH 4.0, there are other functional 0.100 to 0.125 mg/25 mL, the amount of Pb(II) removal was found to
groups could be revealed and effectively release their protons into the diminished caused by the overlying or aggregation of active sites at
solution to transport negative charges to increase the binding capacity higher dosage [32]. Consequently, an optimum sorbent dosage of
for Pb(II). Other investigators who showed that sorption efficiency of 0.100 mg/25 mL was chosen for further experiments.
However, the amount sorbed per unit mass of sorbent reduces
considerably. This is caused by the increase in some sorption active sites
on the sorbents surface. The results are predictable due to by increasing
the sorbents dosage, the number of sorption sites accessible for sor-
bent–sorbate interaction is increased as well and consequently be-
coming easier penetration of the Pb(II) ions to the sorption sites. The
increase of biosorbents dosage is parallel with the accessibility of other
active sites for the joining of Pb(II) in solution, which justifies the re-
lation of removal rate with biosorbents dosage increasing.

3.2.3. Effect of contact time


One of the critical parameters that affecting sorption is contact time.
The effect of contact time for Pb(II) ions removal onto AR and TW
biosorbents is presented in Fig. 6. In the beginning, the plot reveals that
the sorption of Pb(II) removal by all biosorbents is rapid very quickly.
Thereafter, the sorption reaction becomes slower. It gradually slows
down a slight change in Pb(II) removal efficiency until it gets the re-
mains constant (equilibrium). In all case of biosorbents, about 94% of
Fig. 4. Effect of solution pH on Pb(II) removal (%).
the sorption capacity for Pb(II) removal was reached within the first

1204
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

isotherm shows surface sameness of the sorbent and suggestion to the


conclusions that the surface of sorbent is built up of minor sorption
spots, which are energetically same to other, is respect to sorption oc-
currence. It is assumed that the sorption of Pb(II) ions onto biosorbents
is good and the Langmuir isotherms are related to the sorption process.
The Freundlich isotherm model can be considered as one of the
initial empirical equations and is displayed to be reliable with an ex-
ponential distribution of the active centres [35]. The Freundlich iso-
therm model supposes that the multilayers sorption of Pb(II) ions took
place on heterogeneous surface sites. The well-known linear equation of
Freundlich sorption isotherm can be expressed as follows [34]:

ln qe = ln KF + 1/n ln Ce (5)

where KF and 1/n are the Freundlich isotherm constant for Pb(II) ion
sorption. Hence, a plot of ln qe against ln Ce should be a linear line with
a slope 1/n and the intercept of ln KF as shown in Fig. 8. The corre-
sponding parameters of two isotherm models found in the current study
as shown in Table 2.
Fig. 6. Effect of contact time on Pb(II) removal (%).
From the two isotherm models evaluated, the data matched with
Freundlich isotherm with a high-value correlation coefficient (R2) for
15 min of contact for a Pb(II) solution of 20 ppm. It could be seen that in AR (R2 = 0.992) and TW biosorbent (R2 = 0.985), compared to
all circumstance no desorption occur. When the equilibrium is attained, Langmuir isotherm (R2 = 0.980 and R2 = 0.973), respectively. The
it remains constant. sorption capacity found by Langmuir isotherm is 108 mg/g (for AR) and
It might be explained by initial sorption step, which involved the Pb 152 mg/g (for TW). The best match of equilibrium data with Freundlich
(II) sorption onto the free external surface sites of sorbents are available isotherm indicates that heterogeneous biosorption performs a key part
for sorption. The biosorption is relatively rapid in the initial 15 min due in Pb(II) removal ions. In addition, the KF value shows the sorption
to the active site of biosorbents were unoccupied, and Pb(II) ions could capacity of sorbent and the value of 1/n < 1 or n (dimensionless
easily interact with free active surface sites. After an interval of time, constant) higher than one shows easy separation and favorable multi-
the next step is slower due to much of the surface sites are occupied layer sorption of Pb(II) ion from aqueous medium [36,37]. The smaller
[33]. In addition, it is difficult to occupy the remaining free surface sites value of 1/n obtained (0.474–0.543) shows better sorption and for-
are due to repulsive forces between the molecules of solute on the solid mation of a relatively stronger bond between biosorbents and Pb(II)
and bulk phases. Therefore, the equilibrium contact time is regarded as ions.
60 min (AR biosorbents) and 75 min (TW biosorbents) for all furthest Table 3 presents many others material used for Pb(II) removal from
experiments to confirm that the equilibrium was attained. aqueous solution. When the comparison of sorption capacity in re-
moving Pb(II) ions of AR and TW sorbent with other sorbents in the
3.3. Isotherm analysis studies literature, it is be seen that biosorbents in this studied had a good ad-
sorptive character. These results indicate that biosorbent might be well
To measure the sorption capacity of biosorbent for Pb(II) removal used as a sorbent of Pb(II) removal from aqueous solution even without
ions, the Freundlich and Langmuir isotherm models were used. any modifications.
Freundlich and Langmuir’s isotherm is the most common isotherms
describing solid-liquid sorption system. As two-parameter models,
3.4. Sorption kinetic studies
Freundlich and Langmuir isotherms, which give facts on the sorption,
capacity and constants associated with the activation energy. The
The sorption kinetics might give valuable understanding into the
maximum sorption capacity (qm) values, correlation coefficient (R2),
reaction ways and sorption courses mechanisms. To explain the sorp-
and other constants can be determined from experimental data.
tion kinetic of Pb(II) ion onto biosorbents was studied by two of the
The Langmuir model is possibly the largely widely employed to
most commonly used models, namely, pseudo-first-order model and
sorption isotherms. Langmuir sorption isotherm assumes that the
pseudo-second-order model. The compliance between experimental
sorption takes place at the surface of sorbent, which is specific homo-
data and the model expected values was stated by the correlation
geneous sites, has only one-type sites binding, where the molecules
coefficient (R2).
form a monolayer of sorbate, saturating the pores and avoiding the
migration. No further sorption could take place at this site. It considers
the sorption energy of each molecule is same, independent of the sur- 3.4.1. Pseudo-first-order model
face of the material. The non-linearized and linearized form of The pseudo-first order rate model of Lagergren [30,34] is based on
Langmuir isotherm could be expressed as follow [34]. solid capacity and expresses as follows
Ce C 1 k1
= e + log(qe − qt ) = log qe − t
qe qm KL. qm (4) 2,303 (6)

where qe (mg/g) is the amount of Pb(II) ion sorbed per specified where qe is the quantity of solute sorbed at equilibrium per weight unit
amount of sorbent; Ce (mg/L) is the equilibrium concentration of Pb(II) of biosorbent (mg/g), qt is the quantity of solute sorbed at any time
ions; qm is the amount of sorbate required to form a monolayer (mg/g) (mg/g), and k1 is the sorption constant. It is the very well-known form
and KL (L/mg) is the Langmuir equilibrium constant, respectively. From of the pseudo-first-order kinetic model. The worth of k1 at various times
the linear plots, Ce/qe vs. Ce of Langmuir model (Fig. 7) can obtain the was determined from the plots of log (qe-qt) versus t (Fig. 9) for Pb(II)
value of qm and KL from the slope (1/qm) and the intercept [1/(qm KL)]. removal. Constant k1 and correlation coefficient (R2) had been com-
The results for sorption of Pb(II) ions onto both AR and TW biosorbents puted and presented in Table 4. The correlation coefficient (R2) values
were 4.00 and 1.51 × 10−1 mg/g, respectively. The linearity (R2) of the found were quite small. It indicated that the pseudo-first order rate
plot was 0.986 and 0.973, respectively. The Langmuir type sorption extremely bad correlation coefficient (R2) for best fits data.

1205
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Fig. 7. Langmuir isotherm plots of Pb(II) ion onto AR biosorbent (A) and TW biosorbent (B).

Fig. 8. Freundlich isotherm plots of Pb(II) ion onto AR biosorbent (A) and TW biosorbent (B).

Table 2 3.4.2. Pseudo-second-order model


Isotherm constants and regression data for sorption of Pb(II) onto biosorbents. For the pseudo-second-order model, the kinetic data were examined
using a formula, which could express as below [34,38]:
Biosorbents Sorption parameters
t 1 1
Langmuir Freundlich
= + t
qt k2 qe2 qe (7)
qm (mg/g) KL (L/mg) R2 KF (mg/g) n R2
where k2 (g/(mg.min)) is the pseudo-second-order rate constant; t (min)
AR 108 0.130 0.980 13.00 1.84 0.992 is the sorption time; qe (mg/g) is the equilibrium Pb(II) sorption ca-
TW 152 0.496 0.973 13.81 2.11 0.985 pacity of biosorbent, and qt (mg/g) is the sorption capacity at a time t.
The k2 and qe could be obtained from the slope (1/qe) and intercept (1/
k2.qe2)) of the linear plot of t/qt versus t (Fig. 10). The k2 and qe, which
Table 3 is computed from the model, are displayed in Table 4. It could be shown
Comparison of sorption capacity (mg/g) of Pb(II) ions removal on some sorbent.
that the sorption of Pb(II) ions follow the pseudo-second-order rate
Sorbent Sorption capacity of Pb(II) References model, with a correlation coefficient (R2) of 0.999 for all biosorbent.
(mg/g) These results indicate that Pb(II) removal on biosorbent follows me-
chanism of a pseudo-second-order, and the sorption process is mostly
Anadara inaequivalvis shell 621.1 Bozbas and Boz [42]
governed by chemisorption.
Orange peel xanthate 204.5 Jiang et al. [40]
Tomato waste 152 This study
Citrus maxima peel 126 Chao et al. [41] 3.5. Desorption studies
Apple residue 15.8 Lee et al. [44]
Apple residue 108 This study
One of the fundamental considering for sorbents usage is the
Ponkan peel 112.1 Pavan et al. [43]
Palm shell activated carbon 95.2 Issabayeva et al. [31] probability of their regeneration, where allows the materials to be re-
Magnetic Litchi peel 78.74 Liang et al. [39] used. It is the main factor in supporting the use of a sorption process as
Passion fruit shell 51.3 Chao et al. [41] a cost-effective and highly effective method of Pb(II) removal ions from
Sugarcane bagasse 27.1 Chao et al. [41] aqueous systems. Due to this fact, in this work the desorbing agents
Cedar leaf ash 8.0 Hafshejani et al. [28]
used were water (for capturing mechanism), ammonium acetate 0.1 M
(for ion exchange mechanism), hydrochloric acid 0.1 M (for acid-base

1206
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Fig. 9. Pseudo-first-order kinetics models plot of Pb(II) ion onto AR biosorbent (A) and TW biosorbent (B).

Table 4 Table 5
Pseudo-first and pseudo-second-order models for sorption of Pb(II) ions onto biosorbents. Desorption efficiency (%) of Pb(II) ions from biosorbent depending on the sorbing agent
used.
Biosorbent Pseudo-first-order model Pseudo-double-order model
Biosorbent Desorption efficiency (%)
k1 (1/min) qe (mg/g) R2 k2 (g/mg.min) qe (mg/g) R2
H2O NH4-acetate HCl Na2-EDTA
AR 0.041 0.435 0.754 0.184 5.000 0.999
TW 0.040 0.365 0.873 0.259 4.958 0.999 AR 0.000 0.073 59.647 99.809
TW 0.000 8.118 53.473 94.247

reaction mechanism) and sodium EDTA (Na2-EDTA) 0.1 M (for complex


forming reaction mechanism). The results obtained are presented in • Pre-treatment with NaOH causes the increasing of biosorbents ac-
Table 5. tivation.
As it can be shown when a desorbing agent is a water, the Pb(II) ions • The optimum conditions for maximum removal percentage of Pb(II)
were not eluted. It might demonstrate a strong bond between the sor- by biosorbents were found to be at pH 4.0 with 0.1 g of sorbent and
bent and sorbate. However, when Na2-EDTA and hydrochloric acid 60 min contact time for AR and 90 min for TW.
were used, the desorption results gave satisfactorily. It showed that the • The kinetics of Pb(II) ions removal on biosorbents followed a
biosorbent was obtained highly stable and then it can be reused. Hence, pseudo-second-order model, which shows that the process was
from the desorption results it might be assumed that it is possible to use controlled by chemisorption.
the biosorbent in further sorption process, which is significant im- • The equilibrium data can best describe by the Freundlich isotherm
portance for economic development and ‘green’ technologies in Pb(II) than Langmuir isotherm.
ions removal from aqueous solution. • The sorption capacity of AR and TW were 108 and 152 mg/g, re-
spectively. These values were higher than those of biosorbent, which
had been considered as excellent biosorbent by several researchers.
4. Conclusions
It can be concluded that the AR and TW were a strong candidate, an
The following conclusions are arrived at based on the results of the effective and efficient biosorbents for the Pb(II) removal from the
present study: aqueous system.

Fig. 10. Pseudo-second-order kinetics models plot of Pb(II) ion onto AR biosorbent (A) and TW biosorbent (B).

1207
E. Heraldy et al. Journal of Environmental Chemical Engineering 6 (2018) 1201–1208

Acknowledgments 778–788.
[21] M. Mouflih, A. Aklil, S. Sebti, Removal of lead from aqueous solutions by activated
phosphate, J. Hazard. Mater. B 119 (2005) 183–188.
The authors wish to thank The Sebelas Maret University (Universitas [22] N. Bekta, S. Kara, Removal of lead from aqueous solutions by natural clinoptilolite:
Sebelas Maret) for the financial support to this work under the Grant of equilibrium and kinetic studies, Sep. Purif. Technol. 39 (2004) 189–200.
University Research Committee (‘Hibah Mandatory’ PNBP UNS, 2017). [23] M. Machida, R. Yamazaki, M. Aikawa, H. Tatsumoto, Role of minerals in carbo-
naceous adsorbents for removal of Pb(II) ions from aqueous solution, Sep. Purif.
Technol. 46 (2005) 88–94.
References [24] A. Rahmani, H.Z. Mousavi, M. Fazli, Effect of nanostructure alumina on adsorption
of heavy metals, Desalination 253 (2010) 94–100.
[25] Y. Pang, G. Zeng, L. Tang, Y. Zhang, Y. Liu, X. Lei, Z. Li, J. Zhang, G. Xie, PEI-grafted
[1] M. Suguna, N.S. Kumar, V. Sreenivasulu, A. Krishnaiah, Removal of Pb(II) from
magnetic porous powder for highly effective adsorption of heavy metal ions,
aqueous solutions by using chitosan coated zero valent iron nanoparticles, Sep. Sci.
Desalination 281 (2011) 278–284.
Technol. 49 (2014) 1613–1622.
[26] G.C. Saha, M.I. Ul Hoque, M.M. Miah, R. Holzec, D.A. Chowdhury, S. Khandaker,
[2] R. Tabaraki, A. Nateghi, S. Ahmady-Asbchin, Biosorption of lead(II) ions on
S. Chowdhury, Biosorptive removal of lead from aqueous solutions onto Taro
Sargassum ilicifolium: application of response surface methodology, Int.
(Colocasiaesculenta (L.) Schott) as a low cost bioadsorbent: characterization, equi-
Biodeterior. Biodegrad. 93 (2014) 145–152.
libria, kinetics and biosorption-mechanism studies, J. Environ. Chem. Eng. 5 (2017)
[3] D.H.K. Reddy, Y. Harinath, K. Seshaiah, A.V.R. Reddy, Biosorption of Pb(II) from
2151–2162.
aqueous solutions using chemically modified Moringa oleifera tree leaves, Chem.
[27] T.K. Naiya, A.K. Bhattacharya, S.K. Das, Clarified sludge (basic oxygen furnace
Eng. J. 162 (2010) 626–634.
sludge)—an adsorbent for removal of Pb(II) from aqueous solutions—kinetics,
[4] B.L. Martins, C.C.V. Cruz, A.S. Luna, C.A. Henriques, Sorption and desorption of
thermodynamics and desorption studies, J. Hazard. Mater. 170 (2009) 252–262.
Pb+2 ions by dead Sargassum sp. biomass, Biochem. Eng. J. 27 (2006) 310–314.
[28] L.D. Hafshejani, S.B. Nasab, R.M. Gholami, M. Moradzadeh, Z. Izadpanaha,
[5] T.M. Ansari, M.A. Hanif, A. Mahmood, U. Ijaz, M.A. Khan, R. Nadeem, M. Ali,
S.B. Hafshejani, A. Bhatnagar, Removal of zinc and lead from aqueous solution by
Immobilization of rose waste biomass for uptake of Pb(II) from aqueous solutions,
nanostructured cedar leaf ash as biosorbent, J. Mol. Liq. 211 (2015) 448–456.
Biotechnol. Res. Int. 11 (2011) 1–9.
[29] S. Tasar, F. Kaya, A. Ozer, Biosorption of lead(II) ions from aqueous solution by
[6] A.R. Kul, H. Koyuncu, Adsorption of Pb(II) ions from aqueous solution by native and
peanut shells: equilibrium, thermodynamic and kinetic studies, J. Environ. Chem.
activated bentonite: kinetic, equilibrium and thermodynamic study, J. Hazard.
Eng. 2 (2014) 1018–1026.
Mater. 179 (2010) 332–339.
[30] M.N.M. Ibrahim, W.S. Wan Ngah, M.S. Norliyana, W.R. Wan Daud, M. Rafatullah,
[7] A.M. Azzam, S.T. El-wakeel, B.B. Mostafa, M.F. El-Shahat, Removal of Pb, Cd, Cu
O. Sulaiman, R. Hashim, A novel agricultural waste adsorbent for the removal of
and Ni from aqueous solution using nano scale zero valent iron particles, J. Environ.
lead(II) ions from aqueous solutions, J. Hazard. Mater. 182 (2010) 377–385.
Chem. Eng. 4 (2016) 2196–2206.
[31] G. Issabayeva, M.K. Aroua, N. Meriam, N. Sulaiman, Removal of lead from aqueous
[8] L. Monser, N. Adhoum, Tartrazine modified activated carbon for the removal of Pb
solutions on palm shell activated carbon, Bioresour. Technol. 97 (2006)
(II), Cd(II) and Cr(III), J. Hazard. Mater. 161 (2009) 263–269.
2350–2355.
[9] M.M. Raoa, D.K. Ramanaa, K. Seshaiaha, M.C. Wang, S.W.C. Chien, Removal of
[32] M. Kostić, M. Radović, J. Mitrović, M. Antonijević, D. Bojić, M. Petrović, A. Bojić,
some metal ions by activated carbon prepared from Phaseolus aureus hulls, J.
Using xanthated Lagenaria vulgaris shell biosorbent for removal of Pb(II) ions from
Hazard. Mater. 166 (2009) 1006–1013.
wastewater, J. Iran. Chem. Soc. 11 (2014) 565–578.
[10] M.R. Lasheen, N.S. Ammar, H.S. Ibrahim, Adsorption/desorption of Cd(II), Cu(II)
[33] Q. Li, M. Diao, H. Xiao, K. Gao, Synthesis ofsuperabsorbent and sea shell powder
and Pb(II) using chemically modified orange peel: equilibrium and kinetic studies,
composite and itsadsorption kinetics and isotherms to Pb2+ in water, J. Mater.
Solid State Sci. 14 (2012) 202–210.
Appl. 2 (2) (2013) 45–56.
[11] M. Calero, A. Perez, G. Blazquez, A. Ronda, M.A. Martın-Lara, Characterization of
[34] E. Heraldy, S.J. Sri Juari Santosa, Triyono, K. Wijaya, Anionic and cationic dyes
chemically modified biosorbents from olive tree pruning for the biosorption of lead,
removal from aqueous solutions by adsorption onto synthetic Mg/Al hydrotalcite-
Ecol. Eng. 58 (2013) 344–354.
like compound, Indones. J. Chem. 15 (3) (2015) 234–241.
[12] D.D. Arimurti, E. Heraldy, W.W. Lestari, Biosorption of lead(II) ions by NaOH-ac-
[35] Y.S. Ho, Effect of pH on lead removal from water using tree fern as the sorbent,
tivated apple (Malus domestica) juice residue, AIP Conf. Proc. 1710 (2016) 030036,
Bioresour. Technol. 96 (2005) 1292–1296.
http://dx.doi.org/10.1063/1.4941502.
[36] P. Xu, G.M. Zeng, D.L. Huang, C. Lai, M.H. Zhao, Z. Wei, N.J. Li, C. Huang, G.X. Xie,
[13] D. Permatasari, E. Heraldy, W.W. Lestari, Biosorption of toxic lead(II) ions using
Adsorption of Pb(II) by iron oxide nanoparticles immobilized Phanerochaete
tomato waste (Solanum lycopersicum) activated by NaOH, AIP Conf. Proc. 1710
chrysosporium: equilibrium, kinetic, thermodynamic and mechanisms analysis,
(2016) 030022, http://dx.doi.org/10.1063/1.4943488.
Chem. Eng. J. 203 (2012) 423–431.
[14] S.K. Yadav, D.K. Singh, S. Sinha, Chemical carbonization of papaya seed originated
[37] S.S. Subhashini, M. Velan, S. Kaliappan, Biosorption of lead by Kluyveromyces
charcoals for sorption of Pb(II) from aqueous solution, J. Environ. Chem. Eng. 2
marxianus immobilized in alginate beads, J. Environ. Biol. 34 (2013) 831–835.
(2014) 9–19.
[38] Y.S. Ho, G. McKay, Sorption of dye from aqueous solution by peat, Chem. Eng. J. 70
[15] A.B. Albadarin, A.H. Al-Muhtaseb, N.A. Al-Iaqtah, G.M. Walker, S.J. Allen,
(2) (1998) 115–124.
M.N.M. Ahmad, Biosorption of toxic chromium from aqueous phase by lignin:
[39] S. Liang, X. Guo, N. Feng, Q. Tian, Application of orange peel xanthate for the
mechanism, effect of other metal ions and salts, Chem. Eng. J. 169 (2011) 20–30.
adsorption of Pb2+ from aqueous solutions, J. Hazard. Mater. 170 (2009) 425–429.
[16] G. Yuvaraja, N. Krishnaiah, M.V. Subbaiah, A. Krishnaiah, Biosorption of Pb(II)
[40] R. Jiang, J. Tian, H. Zheng, J. Qi, S. Sun, X. Li, A novel magnetic adsorbent based on
from aqueous solution by Solanum melongena leaf powder as a low-cost biosorbent
waste litchi peels for removing Pb(II) from aqueous solution, J. Environ. Manage.
prepared from agricultural waste, Colloid Surf. B 114 (2014) 75–81.
155 (2015) 24–30.
[17] P. Chand, Y.B. Pakade, Removal of Pb from water by adsorption on apple pomace:
[41] H.P. Chao, C.C. Chang, A. Nieva, Biosorption of heavy metals on Citrus maxima
equilibrium, kinetics, and thermodynamics studies, J. Chem. 2013 (2013) 1–8,
peel, passion fruit shell, and sugarcane bagasse in a fixed-bed column, J. Ind. Eng.
http://dx.doi.org/10.1155/2013/164575 164575.
Chem. 20 (2014) 3408–3414.
[18] M. Suguna, V.N.R. Mudumala, V. Sreenivasulu, A. Krishnaiah, Modified leaf bio-
[42] S.K. Bozbas, Y. Boz, Low-cost biosorbent: Anadara inaequivalvis shells for removal
mass for Pb(II) removal from aqueous solution: application of response surface
of Pb(II) and Cu(II) from aqueous solution, Process Saf. Environ. 103 (2016)
methodology, Ecol. Eng. 83 (2015) 218–226.
144–152.
[19] A. Verma, S. Kumar, S. Kumar, Biosorption of lead ions from the aqueous solution
[43] F.A. Pavan, A.C. Mazzocato, R.A. Jacques, S.L.P. Dias, Ponkan peel: a potential
by Sargassum filipendula: equilibrium and kinetic studies, J. Environ. Chem. Eng. 4
biosorbent for removal of Pb(II) ions from aqueous solution, Biochem. Eng. J. 40
(2016) 4587–4599.
(2008) 357–362.
[20] O. Pezoti, A.L. Cazetta, K.C. Bedin, L.S. Souza, A.C. Martins, T.L. Silva,
[44] S.H. Lee, C.H. Jung, H. Chung, M.Y. Lee, J.W. Yang, Removal of heavy metals from
O.O.S. Júnior, J.V. Visentainer, V.C. Almeida, NaOH-activated carbon of high sur-
aqueous solution by apple residues, Process Biochem. 33 (2) (1998) 2015–2211.
face area produced from guava seeds as a high-efficiency adsorbent for amoxicillin
removal: kinetic, isotherm and thermodynamic studies, Chem. Eng. J. 288 (2016)

1208

You might also like