You are on page 1of 9

Journal of Hazardous Materials 184 (2010) 538–546

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Rapid removal and recovery of Pb(II) from wastewater by magnetic


nanoadsorbents
Nashaat N. Nassar ∗
Department of Chemical and Petroleum Engineering, University of Calgary, Calgary, Alberta, Canada

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

Article history: Iron oxide nanoadsorbents are cost-effective adsorbents that provide high adsorption capacity, rapid
Received 3 March 2010 adsorption rate and simple separation and regeneration. In this study, Fe3 O4 nanoadsorbents have been
Received in revised form 23 June 2010 employed for the removal of Pb(II) ions from aqueous solutions by a batch-adsorption technique. The
Accepted 18 August 2010
effects of contact time, initial concentration of Pb(II) ions, temperature, solution pH and coexisting ions
Available online 26 August 2010
on the amount of Pb(II) adsorbed have been investigated. Pb(II) adsorption was fast, and equilibrium was
achieved within 30 min. The amount of Pb(II) adsorbed increased as temperature increased, suggesting
Keywords:
an endothermic adsorption. The optimal pH value for Pb(II) adsorption was around 5.5. Furthermore,
Lead
Nanoadsorbent
the addition of coexisting cations such as Ca2+ , Ni2+ , Co2+ , and Cd2+ has no remarkable influence on
Iron oxide Pb(II) removal efficiency. The adsorption equilibrium data fitted very well to Langmuir and Freundlich
Wastewater adsorption isotherm models. The thermodynamics of Pb(II) adsorption onto the Fe3 O4 nanoadsorbents
Adsorption indicated that the adsorption was spontaneous, endothermic and physical in nature. The desorption
Equilibrium and regeneration studies have proven that Fe3 O4 nanoadsorbents can be employed repeatedly without
Nanoparticle impacting its adsorption capacity.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction addition, it should generate a minimum amount of sludge. Nanoad-


sorbents could be employed most effectively not only in a very
The presence of toxic metal ions in wastewater remains a serious low concentration range (∼1 ppm) of pollutant, but also in a very
environmental concern. Therefore, it is necessary to develop vari- high concentration range (∼1000 ppm), where other techniques
ous efficient technologies for their removal. A number of techniques are ineffective, time-consuming, or costly. Nanoadsorbents have
have been used to remove the metal ions from wastewater efflu- particularly high adsorption capacities because of their specific
ents; including chemical precipitation [1], ion exchange process functionality and large specific surface areas, i.e., surface area
[2,3], electrolytic methods [4], adsorption onto activated carbon per unit mass. In addition, nanoadsorbents are highly mobile in
[5], organic-based ligand precipitation [6], membrane and reverse porous media because they are much smaller than the relevant
osmosis processes [7]. These methods have been found to be lim- pore spaces, so they can be transported effectively by the flow in
ited, because of the high capital and operating costs and/or the porous media [17]. Therefore, the nanoadsorbents can be employed
ineffectiveness in meeting stringent effluent standards. Therefore, in situ, within the contaminated zone where treatment is needed. In
several approaches have been studied for the development of situ, involves treatment of contaminants in place, in comparison to
inexpensive and abundant metal sorbents, such as sawdust [8], ex situ where treatment occur after transferring the contaminated
live biomass [9], clay [10,11] and agricultural byproducts [12–14]. material to a more convenient location [17], adding more cost and
These approaches, however, suffer from some challenges; includ- environmental impact to the process (e.g. pumping, transportation,
ing the huge amount of sludge waste generated from the spent instrumentations and treatment of contaminants). Certain types
adsorbents, the high regeneration cost and the low adsorption of nanoadsorbents have been found to be effective in metal ion
rate and capacity, especially at very low pollutant concentration removal [15,18–21]. Although previous research has highlighted
[15,16]. An excellent adsorbent should generally possess high sur- the ability of metal oxide nanoadsorbents for metal ion adsorption,
face area and short adsorption equilibrium time, so that it can few metal ions were actually tested [15,20–23]. Furthermore, the
be used to remove high amount of pollutant in shorter time. In reported literature looked into percentage removal without deep
investigation onto the adsorption mechanism and thermodynamic
parameters. In the present study, the adsorption mechanistic,
∗ Corresponding author. Tel.: +1 403 210 9772; fax: +1 403 210 3973. kinetics, equilibria and thermodynamic of Pb(II) ions, as an example
E-mail address: nassar@ucalgary.ca. of a typical heavy metal ion present in wastewater, onto the Fe3 O4

0304-3894/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhazmat.2010.08.069
N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546 539

III Multi Purpose Diffraction System (Rigaku Corp., The Woodlands,


Nomenclature TX) with Cu KR radiation operating at 40 kV and 44 mA with a –2
goniometer. Scanning was performed from 8◦ to 90◦ with a count-
C concentration of Pb(II) in the solution at any time ing time of 0.2 s/step. The structure was identified by comparison
(mg/L) to spectra in the JADE program, Materials Data XRD Pattern Pro-
Cads amount of Pb(II) adsorbed onto the nanoadsorbents cessing Identification & Quantification. The diffraction pattern is
(mg/L) shown in Fig. 1. Clearly, it conforms the peaks of Fe3 O4 as reported
Cdes amount of Pb(II) released into the aqueous solution by the manufacturer. Measurements on particle size indicated that
(mg/L) the Fe3 O4 nanoadsorbents have a size of 22 ± 1.5 nm.
Ce equilibrium concentration of Pb(II) in the solution Separation of nanoadsorbents from aqueous solution was eval-
(mg/L) uated by separating them via a small horseshoe Alnico magnet.
Co initial concentration of Pb(II) in the solution (mg/L) The concentrations of the nanoadsorbents in the supernatant
K adsorption equilibrium constant (dimensionless) were evaluated at different pH values between 2.5 and 6.5 using
KF Freundlich isotherm constant (mmol/g) (L/mmol)1/n the inductively coupled plasma-atomic emission spectroscopy
KL Langmuir isotherm constant (L/mmol) (ICP-AES) (IRIS Intrepid II XDL, Thermo-Instruments Canada Inc.,
m mass of Fe3 O4 nanoparticles (g) Mississauga, ON). No significant Fe concentrations were detected
q amount of Pb(II) adsorbed at any time (mg/g) as confirmed by the ICP, indicating that all the nanoadsorbents were
qe amount of Pb(II) adsorbed at equilibrium time separated by the magnet and the dissolution of the nanoadsorbents
(mg/g) under the considered experimental conditions was insignificant.
qm maximum equilibrium adsorption capacity of Pb(II)
(mg/g) 2.2. Adsorbates
r (%) recovery efficiency of Pb(II)
R ideal gas constant (R = 8.314 J/mol K) Pb(NO3 )2 obtained from ACROS Organics (99%, ACROS Organ-
V total volume of aqueous solution (L) ics, NJ) was selected for this study as a source of Pb(II) ions.
1/n Freundlich heterogeneity factor (unitless) Pb(II) stock solution was prepared by dissolving a specified
o
Gads standard Gibbs free energy change (kJ/mol)
o
weight of Pb(NO3 )2 in 500 mL of deionized water, and subse-
Hads standard enthalpy change (kJ/mol) quently diluted to the required concentrations. The following
o
Sads standard entropy change (J/mol K) chemicals were used as the precursor salts for the coexisting
cations present in wastewater, namely Ni(NO3 )2 ·6H2 O (99.9985%,
Subscripts
Alfa Aesar, Toronto, ON), Cd(NO3 )2 ·4H2 O (99%, Fisher Scientific,
ads adsorption
Toronto, ON), Ca(NO3 )2 ·4H2 O (99.9%, Fisher Scientific, Toronto, ON)
des desorption
and Co(NO3 )2 ·6H2 O (99%, Fisher Scientific, Toronto, ON). All chem-
e equilibrium
icals were used without further purifications.
o initial
m maximum
2.3. pH measurements
F Freundlich
L Langmuir
pH measurements were performed with a AB15 plus pH meter
zpc zeropoint of charge
from Fisher Scientific, Toronto, ON. Aliquots of 0.1 M HNO3 or 0.1 M
NaOH were used to adjust the pH over the range 2.5–6.5. In adjust-
ing the pH, care was taken not to change significantly the initial
nanoadsorbents were investigated. In addition, the effect of solu- concentration of Pb(II) ions in solution. Therefore, different NaOH
tion pH, coexisting cations, and initial Pb(II) concentration on the and HNO3 solution of optimum concentration were used to achieve
adsorption was studied. Furthermore, the Langmuir and Freundlich the desired pH values mentioned above and the added volume of
models were used to describe the equilibrium isotherms. acid or base solution was kept ≤50 ␮L.

2. Experimental 2.4. Adsorption of Pb(II) ions onto Fe3 O4 nanoadsorbents

2.1. Adsorbent Batch-adsorption studies were performed by mixing nanoad-


sorbents with Pb(II) aqueous solution at a ratio 10:1 (g/L, mass
Fe3 O4 nanoparticles obtained from Nanostructured & Amor- of nanoadsorbents/volume of solution). Kinetics and equilibrium
phous Materials, Inc., Houston, TX, were used as nanoadsorbents. adsorption of Pb(II) ions onto the nanoadsorbents were carried
Before any adsorption experiments were conducted, the charac- out at different temperatures, namely 298, 313 and 328 K. 100 mg
teristics of the selected nanoadsorbents were evaluated; namely: dry powder of nanoadsorbents was introduced into a 20-mL vial
particle size, structure, specific surface area (BET), pore volume directly. Then 10 mL of Pb(II) aqueous solution with known initial
and external surface area. Results are presented in Table 1. The concentration, Co (mg/L), ranging from 10 to 800 mg/L was added
surface areas and pore volume of the nanoadsorbents were mea- to each vial. The vials were sealed and were shaken at 200 rpm
sured by performing N2 -adsorption and desorption at 77 K, using in a temperature incubator at a pre-selected temperature for a
a surface area and porosity analyzer (TriStar II 3020, Micromerit- required specific time. The Pb(II)-nanoadsorbents were separated
ics Corporate, Norcross, GA). The samples were degassed at 423 K from the mixture by a small horseshoe Alnico magnet, and the
under N2 flow overnight before analysis. Surface area was cal- supernatant was decanted. The concentration of Pb(II) in the super-
culated using Brunauer-Emmet-Teller (BET) equation. External natant was measured by using inductively coupled plasma-atomic
surface areas were obtained by t-plot method and there was no emission spectroscopy (ICP-AES). Pb(II) stock standard was pre-
significant difference between the surface areas obtained by BET pared by dissolving a specified mass of the corresponding Pb(NO3 )2
and t-plot methods. This indicates that the nanoadsorbents have in deionized water. Standards containing 0.1, 1 and 10 ppm were
no significant porosity and maintain a high external surface area. prepared by diluting 50 ppm of the stock standard in deionized
Structure and particle size were determined by using X-ray Ultima water. Accordingly, a four-point calibration curve was constructed.
540 N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546

Table 1
Characterizations of Fe3 O4 nanoadsorbents considered in this study.

Manufacturer reported surface area (m2 /g) Manufacturer reported X-ray Estimated BET surface External Average Pore
particle size (nm) measured area (m2 /g) surface area volume (cm3 /g)
particle size (m2 /g)
(nm)

40 20–30 22 ± 1.5 43 39 0.38

A typical standard error in the metal concentration measurements ature incubator for 2 h. After desorption, the nanoadsorbents were
was ±0.001 ppm. The equilibrium adsorption capacity, Qe , of Pb(II) separated by a small horseshoe Alnico magnet, and the supernatant
aqueous was calculated using the mass balance, according to the was decanted for Pb(II) concentration measurement. The recovery
following equation [3]: efficiency, r (%), of Pb(II) from the solid phase was calculated as
follows:
Co − Ce
Qe = V (1)
m
Cdes
where V is the sample volume (L), m is the mass of Fe3 O4 nanoad- r (%) = × 100 (2)
sorbents (g), Co is the initial concentration of Pb(II) in the solution Cads
(mg/L), and Ce is the equilibrium concentration of Pb(II) in the solu-
tion (mg/L). For time-dependent data, C replaces Ce and Q replaces where Cdes and Cads are the amount of Pb(II) released into the aque-
Qe in Eq. (1). ous solution and the amount of Pb(II) adsorbed onto the Fe3 O4
nanoadsorbents (mg/L), respectively.
2.5. Effect of coexisting cations

Multi-metal cation adsorption of nanoadsorbents was exam-


ined in the presence of the following coexisting cations Pb2+ , Ca2+ , 2.7. Reusability of nanoadsorbents
Ni2+ , Co2+ , and Cd2+ . A sample containing a mixture of Pb2+ with
the aforementioned cations was prepared at pH of 5.5 and 298 K. To test the reusability of the nanoadsorbents, the Fe3 O4 nanoad-
Concentration of each cation was maintained at approximately sorbents containing Pb(II) were washed with HNO3 solution of pH
100 mg/L. values between 2 and 3 until no detectable Pb(II) was observed in
the wash water, as confirmed by the ICP. After that, Fe3 O4 nanoad-
2.6. Desorption of Pb(II) sorbents were thoroughly washed with deionized water till the
pH of the wash water reached the range of 5.0–6.5. Then, the
In this experiment, desorption of Pb(II) from Pb-loaded washed nanoadsorbents were vacuum dried at 60 ◦ C for 24 h and
nanoadsorbents was performed using HNO3 solution at different re-used for the subsequent adsorption cycle. Five cycles of con-
concentrations. Nanoadsorbents containing Pb(II) was exposed to secutive adsorption–desorption–regeneration were carried out to
10 mL solution of different HNO3 concentrations, namely 0, 1, 5 and validate the reusability of Fe3 O4 nanoadsorbents for the removal
10 mM. The mixture was agitated at 200 rpm and 298 K in a temper- and recovery of Pb(II).

Fig. 1. X-ray diffractograms of selected nanoadsorbents. Reference data for Fe3 O4 are from Materials Data XRD Pattern Processing Identification & Quantification.
N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546 541

Fig. 3. Effect of pH on Pb(II) adsorption onto Fe3 O4 nanoadsorbents. Initial Pb(II)


Fig. 2. Effect of contact time on the removal of Pb(II) at different initial concentra-
concentration: 220 mg/L, adsorbent dose: 10 g/L, shaking rate: 200 rpm, contact
tions. Adsorbent dose: 10 g/L, shaking rate: 200 rpm, pH: 5.5, T: 298 K.
time: 24 h, T = 298 K.

3. Results and discussions


down the adsorption rate. Despite the short equilibrium time a
3.1. Adsorption mechanism 24-h contact time was adopted for the subsequent experiment
to ensure attaining adsorption equilibrium. The short equilibrium
In aqueous solution, the Pb(II) ions undergo solvation and time is in agreement with that reported by other researchers for
hydrolysis as per the following reactions [24,25]: the adsorption of other metal ions onto iron oxide nanoparticles
[23,28,29]. In different to other conventional porous adsorbents
Pb2+ + nH2 O ↔ Pb(H2 O)2+
n (R1) in which adsorption occur through pore diffusion steps, which in
turn slows down the adsorption rate [3]. Fe3 O4 nanoadsorbent is
Pb(H2 O)2+
n ↔ Pb(H2 O)
n−1
+ H+ (R2)
a nonporous adsorbent, as confirmed by surface area and porosity
2+ (2n−m) + measurement, where only external adsorption occurs. This type of
Pb + mH2 O ↔ Pb(OH)m + mH (R3)
adsorption mass transfer requires less time to reach equilibrium
Literature on Pb(II) speciation shows that the dominant species is [30]. This result is promising as equilibrium time plays a major role
Pb(OH)2 at pH > 6.5 and Pb2+ and Pb(OH)+ at pH < 6.5 [24]. In the in wastewater treatment plant economic viability. Also, as seen in
meantime, the surface of metal oxide nanoadsorbents may undergo Fig. 2, the amount adsorbed of Pb(II) increased with the increase in
protonation/deprotonation as per reaction (R4) [26,27]. the initial concentration of Pb(II) in solution. This can be attributed
H+ H+ to the increase in the ion occupancy number which favored the
H2 O + M–O− ←→M–OH←→M–OH+2 (R4) adsorption process.
OH− OH−

where H+ and OH− refer to the potential determining ions.


The adsorption process is likely to be electrostatic attraction. In 3.3. Effect of pH
the basic solution, a significantly high electrostatic attraction exists
between the negatively charged surface of the adsorbent and the It is well known that pH is one of the most important factors
Pb(II). As the pH of the solution decreases, the number of positively which affect the adsorption process. Experiments were performed
charged sites increases and the number of negatively charged sites to find the optimum pH on the adsorption of Pb(II) ions onto Fe3 O4
decreases. A positively charged surface site on the nanoadsorbent nanoadsorbents using different initial pH values changing from 2.5
does not favor the adsorption of Pb(II) ions, due to the electro- to 6.5. A pH value of 6.5 was not exceeded, since increasing pH to
static repulsion. In addition, in an acidic environment, at low pH 7.0, at fixed values of other variables, resulted in the precipitation
values, excess hydrogen ions compete with the Pb(II) ions for the of lead hydroxide [24]. Therefore, it was not possible to carry out
adsorption site resulting in lower adsorption of Pb(II). adsorption experiments for Pb(II) at pH > 6.5. This would introduce
uncertainty into the results.
3.2. Effect of contact time and initial Pb(II) concentration Fig. 3 presents the effects of pH on the adsorption of Pb(II). As
seen, the removal of Pb(II) ions is clearly pH dependent with the
The adsorption of Pb(II) onto nanoadsorbents was monitored highest adsorption occurring at pH ≥ 5.5. Such pH effect has been
for 250 min. The initial Pb(II) concentrations were 95 and 162 mg/L, observed for the adsorption of Pb(II) onto bulk iron oxides [31].
respectively, the initial pH value of solution was 5.5, and the solu- Higher pH is favorable for the deprotonation of sorbent surface
tion temperature was 298 K. As seen in Fig. 2, Pb(II) was adsorbed [26,27]. Increased deprotonation results in the increase of the neg-
onto Fe3 O4 nanoadsorbents quickly, and equilibrium was achieved atively charged sites, which enhances attractive forces between the
within 30 min, independent of the initial concentration of Pb(II). sorbent surface and the Pb(II) ions, and thus results in an increase in
This could be due to the small size of Fe3 O4 nanoadsorbents, which the adsorption capacity. In the lower pH region, on the other hand,
was favorable for the diffusion of Pb(II) ions from bulk solution the positively charged sites dominate, this enhances the repulsion
onto the active sites of the solid surface. Apparently, external forces existing between the sorbent surface and the Pb(II) ions, and
adsorption dominated and no pore diffusion was observed to slow therefore decreases the adsorption of Pb(II) ions.
542 N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546

Fig. 5. Effect of temperature on the adsorption of Pb(II). Adsorbent dose: 10 g/L,


Fig. 4. Effect of coexisting ions on Pb(II) removal. Initial cation concentration: shaking rate: 200 rpm, contact time: 24 h, pH 5.5.
100 mg/L, adsorbent dose: 10 g/L, shaking rate: 200 rpm, contact time: 24 h, pH 5.5,
T = 298 K.
at the adsorbent surfaces. Similar observations have been reported
by other researchers for the adsorption of Pb(II) onto metal oxide
3.4. Effect of coexisting cations
surfaces [36,37].
Generally, wastewater contains more than one cation. Thus, the
presence of other cations may interfere in the removal efficiency of 3.6. Adsorption isotherms
Pb2+ . As a result, the effects of the following coexisting cations on
the Pb2+ adsorption were studied, namely Ca2+ , Ni2+ , Co2+ , and Cd2+ . One of the most important characteristics of an adsorbent
The results are shown in Fig. 4. Results are the means of triplicate is the amount of adsorbate it can accumulate, which is cal-
runs performed on different days with freshly prepared samples. culated from the adsorption isotherms. Adsorption isotherms
Surprisingly, the presence of these coexisting cations has no are constant–temperature equilibrium relationship between the
remarkable influence on Pb(II) removal efficiency. In addition, Ca2+ amount of adsorbate per unit of adsorbent (Qe ) and its equilibrium
ions and some of the Ni2+ , Co2+ , and Cd2+ ions were able to adsorb solution concentration (Ce ). A number of mathematical models
onto the nanoadsorbents. In general, the competitive adsorption have been employed for describing this equilibrium relationship.
ability varies from one metal ion to another and is related to a num- Out of several models, Freundlich and Langmuir models have been
ber of factors, such as molecular mass, ion charges, hydrated ionic reported most frequently [38,39]. Eqs. (3) and (4) represent the Fre-
radius and hydration energy of the metals [32]. Apparently, the undlich and the Langmuir models, respectively. The linear forms of
coexisting cations did not compete with Pb(II) ions for the active the Freundlich and the Langmuir models are given by Eqs. (5) and
sites on the nanoadsorbents, suggesting a multi surface adsorption (6), respectively.
[33]. Qe = KF Ce
1/n
(3)

3.5. Effect of temperature KF Ce


Qe = Qm (4)
1 + KL Ce
Temperature plays key roles on the adsorption process [35]. 1
First, increasing the temperature decreases the viscosity of the solu- log Qe = log KF + log Ce (5)
n
tion which, in turn, enhances the rate of diffusion of the adsorbate
Ce 1 Ce
molecules across the external boundary layer of the adsorbent and = + (6)
Qe Qm KL Qm
resulted in higher adsorption. Second, changing the temperature
may affect the equilibrium adsorption capacity of the adsorbent. For where Qe is the amount of Pb(II) adsorbed onto the nanoadsorbents
instance, the adsorption capacity will decrease upon increasing the (mmol/g), Ce is the equilibrium concentration of Pb(II) in super-
temperature for an exothermic reaction; while it will increase for natant (mM), KF and n are Freundlich constants. KF is roughly an
an endothermic one. Hence, a study of the temperature-dependent indicator of the adsorption capacity ((mmol/g) (L/mmol)1/n ), and
adsorption processes provides valuable information about the 1/n is the adsorption intensity factor (unitless). Specifically, a larger
standard Gibbs free energy, enthalpy and entropy changes accom- KF value suggests a greater adsorption capacity, and a lower 1/n
panying adsorption. In this study, a series of experiments were value indicates stronger adsorption strength. KL is the equilibrium
conducted at 298, 313 and 328 K to investigate the effect of temper- Langmuir adsorption constant related to the affinity of binding
ature on Pb(II) adsorption and determine the adsorption isotherms sites (L/mmol). Qm (mmol/g) is the saturation capacity, represent-
and thermodynamic parameters. Fig. 5 shows the amount of Pb(II) ing the maximum amount of the Pb(II) adsorbed per unit weight of
adsorbed onto Fe3 O4 nanoadsorbents at different temperatures. As nanoadsorbents for complete monolayer coverage. The adsorption
seen in the figure, the amount of Pb(II) adsorbed increases as the isotherms were determined using batch tests at different temper-
temperature increased. This increase suggests that the adsorption atures (Fig. 5) and pH = 5.5. Initial Pb(II) concentration was varied
process is an endothermic one. The increase in the Pb(II) adsorption from 0 to 800 mg/L.
with temperature may be due the increase in ions mobility, which As seen in Fig. 5, Pb(II) adsorption increased sharply at low equi-
in turn increases the number of ions that interacted with active sites librium concentration and starts to level off, suggesting that Fe3 O4
N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546 543

Table 3
Maximum adsorption capacity of Pb(II) ions onto various adsorbents.

Adsorbent Maximum Reference


adsorption
capacity (mg/g)

Iron oxide 36.0 (0.17 mmol/g) This study


nanoparticles
Humic acid 22.7 [11]
Goethite 11.04 [11]
Al2 O3 -supported 28.98 [37]
iron oxide
Sand 0 [41]
Iron-coated sand 1.21 [42]
Manganese 1.34 [43]
oxide-coated sand
Montmorillonite 33 [44]
Al2 O3 17.5 [45]
Diatomite 24 [46]
Manganese 26.24 [47]
oxide–carbon
nanotube
Sawdust 3.19 [48]
Activated carbon 21.5 [49]
(Merck)
Carbon nanotubes 17.44 [50]

the model. The adsorption constants obtained from the isotherms


together with 2 values are listed in Table 2. Clearly, Freundlich
model seems to be the best-fitting model for the experimental
data. It is noteworthy that Langmuir model assumes that adsorp-
tion occurs on a homogeneous surface; while Freundlich model
describes adsorption where the adsorbent has a heterogeneous sur-
face with adsorption sites that have different adsorption energies.
The values of (1/n) in Table 2 indicate that n values are greater
than unity suggesting that Pb(II) ions are favorably adsorbed by the
nanoadsorbents. Furthermore, KF values increased as the temper-
ature increased suggesting an increase in the adsorption capacity
with temperature as was confirmed experimentally. A comparison
of the maximum uptake of Pb(II) ions onto different adsorbents
is given in Table 3. As seen, the maximum uptake of Pb(II) ions
onto Fe3 O4 nanoadsorbents was approximately 1–35 times higher
Fig. 6. Freundlich and Langmuir isotherms of Pb(II) onto Fe3 O4 nanoadsorbents at than that of reported low cost adsorbents [11,37,41–50]. This
different temperatures. Adsorbent dose: 10 g/L, shaking rate: 200 rpm, contact time: may be attributed to effect of particle size and distribution, sur-
24 h, pH 5.5.
face area, morphology, surface structure and properties. However,
higher adsorption capacities of Pb(II) have been reported by other
has a high sorption affinity for Pb(II) even at low concentration. In researchers [24,51–53]. Nonetheless, this comparison is not pre-
addition, an increase in the solution temperature favors the Pb(II) cise, since the experimental conditions are different.
adsorption as explained earlier. The Freundlich and Langmuir lin-
ear plots are presented in Fig. 6a and b. The adsorption data in 3.7. Thermodynamic studies
Fig. 6a and b were fitted very well for both Freundlich and Lang-
muir models with correlation coefficients R2 ∼ 0.99. Nonetheless, Effect of temperature on the adsorption process provides valu-
for comparing the best fitting of Freundlich and Langmuir models able information not only on the type of process but also it provides
non-linear Chi-square analysis [40] were conducted as per Eq. (7) knowledge about standard Gibbs free energy (Gads o ), standard
o ) and standard entropy (So ) changes during the
 (Qe − Q 2 enthalpy (Hads ads
e Model ) o can be determined using the Langmuir constant,
2 = (7) adsorption. Gads
Qe Model
KL , by the following equation [54]:
where Qe and Qe Model are the equilibrium adsorption capacity o
Gads = −RT ln K (8)
obtained experimentally and modeling, respectively. If the model
is in well agreement with the experimental data the 2 value where R is the ideal gas constant (R = 8.314 J/mol K), T is the tem-
would be small, the smaller 2 value the better agreement is with perature in Kelvin and K is the adsorption equilibrium constant

Table 2
Estimated parameters for the Langmuir and Freundlich models at different temperatures and pH = 5.5.

Temperature (K) Freundlich constants Langmuir constants


1/n 2
KF (mmol/g) (L/mmol) 1/n (unitless)  KL (L/mmol) Qm (mmol/g) 2

298 0.128 0.076 0.001 24.6 0.14 0.001


313 0.145 0.074 0.001 28.6 0.15 0.01
328 0.167 0.070 0.001 33.1 0.17 0.1
544 N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546

Fig. 8. Desorption of Pb(II) from Pb(II)-loaded Fe3 O4 nanoadsorbents using differ-


ent concentration of HNO3 solution at fixed initial concentration of Pb(II): 95 and
162 mg/L, adsorbent dose: 10 g/L, shaking rate: 200 rpm, contact time: 2 h, T = 298 K.
Fig. 7. Determination of thermodynamic parameters for the adsorption of Pb(II)
onto Fe3 O4 nanoadsorbents.
3.8. Desorption of Pb(II)

Desorption of Pb(II) has two advantages, namely valuable Pb(II)


(dimensionless). K can be expressed as KL × Cs , where KL is the equi- ions can be recovered and nanoadsorbents can be re-used for
librium Langmuir constant (L/mmol) and Cs is the solvent molar another cycle. From the pH study, it has been found that adsorption
concentration (mM), which can be calculated from the density and of Pb(II) ions is highly dependent on pH. Therefore, the desorption
molecular weight of water at the given temperature. Hads o and of Pb(II) can be achieved by decreasing the pH of solution. Fig. 8
o can be determined from van’t Hoff equation [54] that showed
Sads shows that the desorption efficiency increased with the increase in
the dependence of equilibrium constant of the adsorption process HNO3 concentration. Insignificant desorption was achieved with-
on temperature as follows: out additional HNO3 ; while almost 90% recovery was obtained at
o o
10 mM HNO3 , independent of the adsorbed Pb(II) concentration.
Hads Sads This suggests that the adsorption of Pb(II) onto Fe3 O4 nanoad-
ln K = − + (9)
RT R sorbents is reversible, and the bonding between the active sites
and the adsorbed Pb(II) is not strong. The results also suggest that
The van’t Hoff plot is presented in Fig. 7. The plot of ln(K) versus 1/T the Pb(II)-loaded nanoadsorbents can be easily desorbed using a
was found to be linear with a correlation coefficient R2 = 1.0. Hads o
o
very low concentration of acid, and Fe3 O4 nanoadsorbents has the
and Sads were calculated from the slope and intercept, respec- potential to be used as an adsorbent for the removal and recov-
tively. The thermodynamic parameters at the studied temperatures ery of metal ions from wastewater. It is worth noting that, no
are listed in Table 4. As seen, Gads o at all temperatures was a obvious Fe ions were observed in the solution with the change in
negative value, confirming the spontaneous nature and the ther- pH as detected by the ICP. This indicates that the dissolution of
modynamically favorable adsorption. The further increase in the Fe3 O4 nanoadsorbents during the desorption process was not an
negative value of Gadso the stronger the driving force of adsorption
o
issue.
process. The positive value of Hads implies that the interaction
between Pb(II) ions and nanoadsorbents is endothermic in nature,
o may be attributed to the increase 3.9. Reusability of nanoadsorbents
while the positive value of Sads
in randomness at the solid–liquid interface. This increase results
Fig. 9 shows the removal efficiency, R(%), of Pb(II), dur-
from the extra translational entropy gained by the water molecules
ing a five cycles of adsorption–desorption–regeneration, from
previously adsorbed onto nanoadsorbents but displaced by Pb(II)
o a 10-mL solution of initial Pb(II) concentration 162 mg/L
ions. It is noteworthy that adsorption process with Gads values
at pH 5.5. As seen, no significant decrease in the adsorp-
between −20 and 0 kJ/mol corresponds to spontaneous physical
tion capacity of Fe3 O4 nanoadsorbents during the five
process, while that with values between −80 and −400 kJ/mol
o cycles was observed. The results demonstrated that Fe3 O4
corresponds to chemisorptions [55,56]. From the Gads values
nanoadsorbents can be used for the removal and recovery
obtained in this study, it can be deduced that the adsorption mech-
of metal ions from wastewater over a number of cycles,
anism is dominated by physisorption. This also is supported by the
o < 40 kJ mol, indicating physical adsorption process indicating its suitability for the design of a continuous pro-
fact that Hads
cess.
[55].
It is noteworthy that commercially available adsorbents such
as, activated carbons are costly, time-consuming and its regener-
ation is essential. Actually, the regeneration and recovery of spent
Table 4
o o o
activated carbons is the most expensive part of the adsorption tech-
Estimated values of Gads , Hads , Sads for the adsorption of Pb(II) ions onto Fe3 O4
nanoadsorbents at different temperatures.
nology and it accounts for about 75% of the total operating and
maintenance costs [34]. However, it seems this is not the case for
Temperature (K) −Gads
o
(kJ/mol) o
Hads (kJ/mol) o
Sads (J/mol K) R2 the regeneration of spent Fe3 O4 nanoadsorbents. Therefore, Fe3 O4
298 17.9 8.1 87.1 1.0 nanoadsorbent is a cost-effective alternative with fast adsorption
313 19.2 rate and its regeneration is simple and may be not necessary, espe-
328 20.0
cially for in situ application.
N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546 545

[8] S.R. Shukla, S.P. Roshan, Removal of Pb(II) from solution using cellulose-
containing materials, J. Chem. Technol. Biotechnol. 80 (2005) 176–183.
[9] G. Yan, T. Viraraghavan, Heavy-metal removal from aqueous solution by fungus
Mucor rouxii, Water Res. 37 (2003) 4486–4496.
[10] R. Shawabkeh, A. Al-Harahsheh, M. Hami, A. Khlaifat, Conversion of oil shale
ash into zeolite for cadmium and lead removal from wastewater, Fuel 83 (2004)
981–985.
[11] Z. Wu, Z. Gu, X. Wang, L. Evans, H. Guo, Effects of organic acids on adsorption
of lead onto montmorillonite, goethite and humic acid, Environ. Pollut. 121
(2003) 469–475.
[12] M. Nasiruddin Khan, M. Farooq Wahab, Characterization of chemically modified
corncobs and its application in the removal of metal ions from aqueous solution,
J. Hazard. Mater. 141 (2007) 237–244.
[13] M. Kobya, E. Demirbas, E. Senturk, M. Ince, Adsorption of heavy metal ions from
aqueous solutions by activated carbon prepared from apricot stone, Bioresour.
Technol. 96 (2005) 1518–1521.
[14] B.S. Inbaraj, N. Sulochana, Carbonised jackfruit peel as an adsorbent for the
removal of Cd(II) from aqueous solution, Bioresour. Technol. 94 (2004) 49–52.
[15] J. Hu, G. Chen, I.M.C. Lo, Removal and recovery of Cr(VI) from wastewater by
maghemite nanoparticles, Water Res. 39 (2005) 4528–4536.
[16] R.Q. Long, R.T. Yang, Carbon nanotubes as superior sorbent for dioxin removal,
J. Am. Chem. Soc. 123 (2001) 2058–2059.
[17] P.G. Tratnyek, R.L. Johnson, Nanotechnologies for environmental cleanup, Nano
Today 1 (2006) 44–48.
Fig. 9. Reusability of Fe3 O4 nanoadsorbents for adsorption/desorption of Pb(II) dur- [18] Y.C. Sharma, V. Srivastava, S.N. Upadhyay, C.H. Weng, Alumina nanoparticles
ing five cycles. Adsorption: initial concentration of Pb(II): 162 mg/L, adsorbent dose: for the removal of Ni(II) from aqueous solutions, Ind. Eng. Chem. Res. 47 (2008)
10 g/L, shaking rate: 200 rpm, contact time: 2 h, T = 298 K. Desorption: pH values 8095–8100.
between 2 and 3. [19] Z. Ai, Y. Cheng, L. Zhang, J. Qiu, Efficient removal of Cr(VI) from aqueous solu-
tion with Fe@Fe2 O3 Coreâ shell nanowires, Environ. Sci. Technol. 42 (2008)
6955–6960.
4. Conclusions [20] K. Hristovski, A. Baumgardner, P. Westerhoff, Selecting metal oxide nano-
materials for arsenic removal in fixed bed columns: from nanopowders to
This study showed that Fe3 O4 nanoadsorbents could be used as aggregated nanoparticle media, J. Hazard. Mater. 147 (2007) 265–274.
[21] J. Hu, G. Lo, Chen, Fast removal and recovery of Cr(VI) using surface-modified
an alternate to conventional adsorbents for the removal of metal jacobsite (MnFe2 O4 ) nanoparticles, Langmuir 21 (2005) 11173–11179.
ions from wastewater in a very short time with high removal effi- [22] J. Hu, I.M.C. Lo, G. Chen, Removal of Cr(VI) by magnetite nanoparticles, Water
ciency. The removal of Pb(II), as a typical metal ion commonly Sci. Technol. 50 (2004) 139–146.
[23] A. Uheida, M. Iglesias, C. Fontàs, M. Hidalgo, V. Salvadó, Y. Zhang, M.
present in wastewater, by adsorption onto Fe3 O4 nanoadsorbents Muhammed, Sorption of palladium(II), rhodium(III), and platinum(IV) on Fe3 O4
was successfully accomplished. Adsorption was very rapid and nanoparticles, J. Colloid Interface Sci. 301 (2006) 402–408.
equilibrium was achieved in less than 30 min. Also adsorption was [24] T.K. Naiya, A.K. Bhattacharya, S.K. Das, Adsorption of Pb(II) by sawdust and
neem bark from aqueous solutions, Environ. Prog. 27 (2008) 313–328.
highly dependent on the initial concentration of Pb(II), pH and tem- [25] C.F. Baes, R.E. Mesmer, The Hydrolysis of Cations, Wiley, New York, 1976.
perature. Maximum removal was observed at pH 5.5. Adsorption [26] P. Roonasi, A. Holmgren, An ATR-FTIR study of sulphate sorption on magnetite;
increased as the initial concentration of Pb(II) and temperature rate of adsorption, surface speciation, and effect of calcium ions, J. Colloid
Interface Sci. 333 (2009) 27–32.
increased. The adsorption isotherms were also determined and
[27] S.B. Johnson, G.V. Franks, P.J. Scales, D.V. Boger, T.W. Healy, Surface
were appropriately described by both Langmuir and Freundlich chemistry–rheology relationships in concentrated mineral suspensions, Int. J.
models, with a better fitting to the Freundlich model than the Miner. Process. 58 (2000) 267–304.
[28] A. Uheida, G. Salazar-Alvarez, E. Björkman, Z. Yu, M. Muhammed, Fe3 O4 and g-
Langmuir model. The thermodynamics of Pb(II) adsorption onto
Fe2 O3 nanoparticles for the adsorption of Co2+ from aqueous solution, J. Colloid
the Fe3 O4 nanoadsorbents confirmed the endothermic nature of Interface Sci. 298 (2006) 501–507.
the adsorption process, the spontaneity and the physisorption of [29] H. Sun, X. Zhang, Q. Niu, Y. Chen, J. Crittenden, Enhanced accumulation of arse-
the process. The desorption and regeneration studies indicated nate in carp in the presence of titanium dioxide nanoparticles, Water Air Soil
Pollut. 178 (2007) 245–254.
that nanoadsorbents can be used repeatedly, without impacting [30] P.K. Dutta, A.K. Ray, V.K. Sharma, F.J. Millero, Adsorption of arsenate and
the adsorption capacity. Therefore, Fe3 O4 nanoadsorbents are rec- arsenite on titanium dioxide suspensions, J. Colloid Interface Sci. 278 (2004)
ommended as fast, effective and inexpensive adsorbents for rapid 270–275.
[31] S.-Z. Lee, L. Chang, H.-H. Yang, C.-M. Chen, M.-C. Liu, Adsorption characteristics
removal and recovery of metal ions from wastewater effluents. of lead onto soils, J. Hazard. Mater. 63 (1998) 37–49.
[32] L. Lv, M.P. Hor, F. Su, X.S. Zhao, Competitive adsorption of Pb2+ , Cu2+ , and Cd2+
Acknowledgment ions on microporous titanosilicate ETS-10, J. Colloid Interface Sci. 287 (2005)
178–184.
[33] M.M. Benjamin, J.O. Leckie, Multiple-site adsorption of Cd, Cu, Zn, and Pb
The author thanks Mr. Aashish Ragani and Ms. Keran Masoud on amorphous iron oxyhydroxide, J. Colloid Interface Sci. 79 (1981) 209–
for their technical help in the experimental work. 221.
[34] V.J. Inglezakis, S.G. Poulopoulos, Adsorption, Ion Exchange and Catalysis:
Design of Operations and Environmental Applications, Elsevier, UK, 2006.
References [35] D.P. Rodda, B.B. Johnson, J.D. Wells, Modeling the effect of temperature on
adsorption of lead(II) and zinc(II) onto goethite at constant pH, J. Colloid Inter-
[1] J.R. Anderson, C.O. Weiss, Method for precipitation of heavy metal sulphides, face Sci. 184 (1996) 365–377.
USA, Patent, 3740331 (June, 19 1973). [36] E. Eren, Removal of lead ions by Unye (Turkey) bentonite in iron and magnesium
[2] A. Smara, R. Delimi, E. Chainet, J. Sandeaux, Removal of heavy metals from oxide-coated forms, J. Hazard. Mater. 165 (2009) 63–70.
diluted mixtures by a hybrid ion-exchange/electrodialysis process, Sep. Purif. [37] Y.-H. Huang, C.-L. Hsueh, C.-P. Huang, L.-C. Su, C.-Y. Chen, Adsorption ther-
Technol. 57 (2007) 103–110. modynamic and kinetic studies of Pb(II) removal from water onto a versatile
[3] I. Metcalf, Eddy, Wastewater Engineering Treatment and Reuse, 4th ed., Al2 O3 -supported iron oxide, Sep. Purif. Technol. 55 (2007) 23–29.
McGraw Hill, New York, 2003. [38] I. Langmuir, The constitution and fundamental properties of solids and liquids.
[4] W.R. Peters, E.T. White, Y.K. Carole, S. Shedroll, Wastewater treatment–physical Part I. Solids, J. Am. Chem. Soc. 38 (1916) 2221–2295.
and chemical methods, J. Water Pollut. Control Fed. 58 (1986) 481–489. [39] H.M.F. Freundlich, Uber die adsorption in losungen, Z. Phys. Chem. 57 (A) (1906)
[5] J. Patterson, Industrial Wastewater Treatment Technology, 2nd ed., Butter- 385–470.
worth Publisher, Boston, 1985. [40] D.C. Montgomery, G.C. Runger, Applied Statistics and Probability for Engineers,
[6] J. Esalah, M.M. Husein, Removal of heavy metals from aqueous solutions by 4 ed., John Wiley & Sons, New York, 2006.
precipitation–filtration using novel organo-phosphorus ligands, Sep. Sci. Tech- [41] A. Kaya, S. Durukan, Utilization of bentonite-embedded zeolite as clay liner,
nol. 43 (2008) 3461–3475. Appl. Clay Sci. 25 (2004) 83–91.
[7] B.A. Winfield, The treatment of sewage effluents by reverse osmosis—pH based [42] C.H. Lai, C.Y. Chen, Removal of metal ions and humic acid from water by iron-
studies of the fouling layer and its removal, Water Res. 13 (1979) 561–564. coated filter media, Chemosphere 44 (2001) 1177–1184.
546 N.N. Nassar / Journal of Hazardous Materials 184 (2010) 538–546

[43] R. Han, Z. Lu, W. Zou, W. Daotong, J. Shi, Y. Jiujun, Removal of copper(II) and [49] J. Rivera-Utrilla, I. Bautista-Toledo, M.A. Ferro-García, C. Moreno-Castilla, Acti-
lead(II) from aqueous solution by manganese oxide coated sand: II. Equilibrium vated carbon surface modifications by adsorption of bacteria and their effect on
study and competitive adsorption, J. Hazard. Mater. 137 (2006) 480–488. aqueous lead adsorption, J. Chem. Technol. Biotechnol. 76 (2001) 1209–1215.
[44] K.G. Bhattacharyya, S.S. Gupta, Adsorptive accumulation of Cd(II), Co(II), Cu(II), [50] Y.-H. Li, S. Wang, J. Wei, X. Zhang, C. Xu, Z. Luan, D. Wu, B. Wei, Lead adsorption
Pb(II), and Ni(II) from water on montmorillonite: influence of acid activation, on carbon nanotubes, Chem. Phys. Lett. 357 (2002) 263–266.
J. Colloid Interface Sci. 310 (2007) 411–424. [51] T.K. Naiya, A.K. Bhattacharya, S.K. Das, Adsorption of Cd(II) and Pb(II) from
[45] J. Yin, Z. Jiang, G. Chang, B. Hu, Simultaneous on-line preconcentration aqueous solutions on activated alumina, J. Colloid Interface Sci. 333 (2009)
and determination of trace metals in environmental samples by flow injec- 14–26.
tion combined with inductively coupled plasma mass spectrometry using a [52] T.K. Naiya, A.K. Bhattacharya, S.K. Das, Clarified sludge (basic oxygen furnace
nanometer-sized alumina packed micro-column, Anal. Chim. Acta 540 (2005) sludge) – an adsorbent for removal of Pb(II) from aqueous solutions – kinetics,
333–339. thermodynamics and desorption studies, J. Hazard. Mater. 170 (2009) 252–262.
[46] Y. Al-Degs, M.A.M. Khraisheh, M.F. Tutunji, Sorption of lead ions on diatomite [53] T.K. Naiya, A.K. Bhattacharya, S. Mandal, S.K. Das, The sorption of lead(II) ions
and manganese oxides modified diatomite, Water Res. 35 (2001) 3724–3728. on rice husk ash, J. Hazard. Mater. 163 (2009) 1254–1264.
[47] S.-G. Wang, W.-X. Gong, X.-W. Liu, Y.-W. Yao, B.-Y. Gao, Q.-Y. Yue, Removal [54] J.M. Smith, H.C. VanNess, M.M. Abbott, Introduction to Chemical Engineering
of lead(II) from aqueous solution by adsorption onto manganese oxide-coated Thermodynamics, McGraw Hill, New York, 2005.
carbon nanotubes, Sep. Purif. Technol. 58 (2007) 17–23. [55] Y. Seki, K. Yurdakoç, Adsorption of promethazine hydrochloride with KSF mont-
[48] B. Yu, Y. Zhang, A. Shukla, S.S. Shukla, K.L. Dorris, The removal of heavy metals morillonite, Adsorption 12 (2006) 89–100.
from aqueous solutions by sawdust adsorption—removal of lead and compar- [56] Y. Yu, Y.-Y. Zhuang, Z.-H. Wang, Adsorption of water-soluble dye onto func-
ison of its adsorption with copper, J. Hazard. Mater. 84 (2001) 83–94. tionalized resin, J. Colloid Interface Sci. 242 (2001) 288–293.

You might also like