Está en la página 1de 10

Chemical Engineering Journal 180 (2012) 81–90

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Cadmium removal and recovery from aqueous solutions by novel adsorbents


prepared from orange peel and Fe2 O3 nanoparticles
V.K. Gupta ∗ , Arunima Nayak
Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee 247667, India

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

Article history: An agricultural waste-orange peel powder (OPP) was successfully modified into a novel magnetic nano-
Received 2 September 2011 adsorbent (MNP–OPP) by co-precipitating it with Fe3 O4 nanoparticles (MNP) for cadmium ion removal
Received in revised form 1 November 2011 from aqueous solutions. Characterization of MNP–OPP by FTIR, SEM, XRD, TEM and VSM revealed the
Accepted 2 November 2011
covalent binding of hydroxyl groups of MNP with the carboxyl groups of OPP, and further confirmed its
physico-chemical properties favorable for metal binding. The cadmium adsorption onto MNP–OPP, MNP
Keywords:
and OPP was tested under different pH, ionic strength, natural organic matter, adsorbate concentration,
Magnetic nanoparticles
contact time and temperature conditions. Results revealed a faster kinetics and efficiency of MNP–OPP in
Orange peel
Cadmium ions
comparison to those of MNP and OPP and further confirmed a complexation and ion exchange mechanism
Adsorption kinetics to be operative in metal binding. The adsorption equilibrium data obeyed the Langmuir model and the
Isotherm kinetic data were well described by the pseudo-second-order model. Thermodynamic studies revealed
Thermodynamics the feasibility and endothermic nature of the system. Breakthrough capacity from column experiments,
adequate desorption as well as reusability without significant loss of efficiency established the practi-
cality of the developed system. Cadmium removal was achieved at 82% from a simulated electroplating
industry wastewater. The experimental results reveal the technical feasibility of MNP–OPP, its easy syn-
thesis, recovery, economic, eco-friendly and a promising advanced adsorbent in environmental pollution
cleanup.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction Adsorption is considered an effective, efficient, and economic


method for water purification [4,5]. Since the performance of
Heavy metals, because of their non-degradable, persistent and an adsorptive separation is directly dependent on the quality
accumulative nature are toxic when present in trace amounts and and cost effectiveness of the adsorbent, the last decade has
are a source of environmental concern [1]. Toxicity at trace levels seen a continuous improvement in the development of effective
leading to adverse health effect is usually associated with exposure noble adsorbents in the form of activated carbon [6], zeolites
to lead, cadmium, mercury and arsenic. A wide range of industries [7], clay minerals [8], chitosan [9], lignocelluloses [10], natural
(mining, metal processing, electroplating, electronics, etc.) release inorganic minerals [11], functionalized polymers [12], etc. How-
such metals into the environment in amounts that can pose a risk ever, most of these adsorbents are either not effective (due to
to human health [2]. Therefore, metal remediation of wastewater diffusion limitation or the lack of enough active surface sites)
prior to discharge, is of great importance. Strict limitations imposed or have shown problems like high cost, difficulties of separation
on metal discharge as well as the complexity of effluents make the from wastewater, or generation of secondary wastes. Considering
wastewater treatment process more difficult. Moreover, dissolved such drawbacks, recently nano-adsorbents viz. nano-alumina [13],
and sorbed organic matter like humic acid exists ubiquitously in functionalized carbon nanotubes [14] and hydroxyapatite nanopar-
natural aquatic environment. Such substances, having a variety of ticles [15] have demonstrated high adsorption efficiency for metal
components including quinone, phenol, catechol and sugar moi- ion removal. One such advanced class of adsorbent – magnetic
eties play a vital role in controlling physicochemical behavior of nano-adsorbent with the help of an external magnetic field has
metal ions [3] and thereby necessitate metal remediation in the been further successful in circumventing the solid–liquid separa-
presence of such organic materials. tion problem usually encountered with nanoparticles. Such novel
adsorbent combining nanotechnology and magnetic separation
technique has not only demonstrated high adsorption efficiency
due to its large surface to volume ratio, but have also shown
∗ Corresponding author. Tel.: +91 1332 285801; fax: +91 1332 285801. additional benefits like ease of synthesis, easy recovery and manip-
E-mail addresses: vinodfcy@gmail.com, vinodfcy@iitr.ernet.in (V.K. Gupta). ulation via subsequent coating and functionalization, absence

1385-8947/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.11.006
82 V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90

of secondary pollutants, cost-effectiveness and environmental- The bare MNP was prepared in a similar way except that no
friendliness. OPP was added. Characterizations of the adsorbents are described
Till date, several magnetic nanomaterials, including in Supporting information.
maghaemite nanoparticles [16], Fe3 O4 magnetic nanoparti-
cles [17], Fe3 O4 nanoparticles functionalized and stabilized with 2.3. Characterization of the adsorbent
compounds like humic acid [18], amino-functionalized polyacrylic
acid (PAA) [19], and various biopolymers like gum arabic [20], LEO 435 VP (Leo Elektronenmikroskopie GmbH, Germany) scan-
chitosan [21] and polysaccharides [22] have been explored for the ning electron microscopy was used for scanning the adsorbent
removal of metal ions. surface. The infrared spectra of adsorbents were recorded in KBr
In this work, a novel magnetic nano-adsorbent (MNP–OPP) discs on an infrared spectrophotometer (Model Perkin Elmer-1600
was developed by the surface modification of Fe3 O4 nanoparticles Series). The BET surface area of the adsorbent was measured
(MNP) with orange peel powder (OPP) with the aim of exploring on micromeritics ASAP 2010 (UK). X-ray measurements were
its feasibility as adsorbent for the removal of cadmium taken as a performed by using a Philips X-ray diffractometer employing Ni-
model toxic metal ion. Orange peel – a low cost, non toxic biosor- filtered Cu KR radiation and Ni filters. TEM images of the composite
bent containing active functional groups of hydroxyl and carboxyl were recorded by Transmission Electron Microscope (FEI TECNAI
present in cellulose, hemi-cellulose and pectin components [23,24] G2 microscope operating at 200 kV). The magnetic properties were
is selected for its better application and management for wastew- evaluated using a Vibrating Sample Magnetometer (Model 155,
ater remediation. Princeton Applied Research).
The objectives of this study are: (1) synthesis of MNP and For pHpzc determination, 0.01 M NaCl was prepared and its pH
MNP–OPP by co-precipitation method and their characterization was adjusted in the range of 2–12 by adding NaOH or HCl. 50 mL
with respect to FE-SEM, TEM, XRD, VSM, pHpzc and FTIR, (2) com- of 0.01 M NaCl each was put in three different conical flasks and
parative batch adsorption study of the three adsorbents (MNP, OPP then 0.25 g of the three adsorbents was added to these solutions.
and MNP–OPP) for Cd2+ with respect to various environmental These flasks were kept for 48 h and the final pH of the solution was
parameters, (3) comparative isotherm, kinetic and thermody- measured by using pH meter (Model Cyberscan 510, Singapore).
namic studies, (4) column, desorption and reusability studies Graphs were then plotted for pHfinal vs. pHinitial .
to investigate the practical utility of the developed adsorbent
and lastly (5) to presume the underlying mechanism of metal 2.4. Batch adsorption and kinetic studies
binding.
Batch adsorption of cadmium ions onto the three adsorbents
(MNP–OPP, MNP and OPP) was investigated in aqueous solutions
2. Experimental under various operating conditions viz. pH 2–11, temperatures 298,
308, 318 K for initial Cd2+ ion concentration of 16 mg/L.
2.1. Reagents The adsorbent suspension (10 mg) and NaNO3 (0.001 M, 0.01 M,
and 0.1 M) were pre-equilibrated for 1 day. Cadmium nitrate solu-
Analytical-grade salt of cadmium [Cd(NO3 )2 ·4H2 O] and humic tion (16 mg/L) and HA solution (10 mg/L) were then added in the
acid was obtained from M/s Merck, India. A 200 mg/L stock solution 100 mL beaker and finally placed on an orbital shaker at 200 rpm.
of the salt was prepared in deionized water. All working solutions The pH of the solution was kept constant by adding 0.1 M NaOH or
were prepared by diluting the stock solution with deionized water. 0.1 M HNO3 . After equilibrium, the samples were centrifuged. OPP
Deionized water was prepared using a Millipore Milli-Q (Bedford, was removed by filtration while MNP–OPP and MNP were removed
MA) water purification system. magnetically from the solution. The residual concentration of
All reagents (ferric chloride 6-hydrate, ferrous chloride tetrahy- cadmium ions was determined by an atomic absorption spec-
drate, ammonium hydroxide (29.6%), NaOH, H2 SO4 , NaNO3 , HNO3 ), trophotometer model Z-7000 (Hitachi, Japan) at a wavelength of
nickel chloride, zinc nitrate used in the study were of ana- 228.8 nm. The concentration of HA was determined spectrophoto-
lytical grade and purchased from Wiswani Chemicals. Oranges metrically on a Specord 200 UV–visible spectrophotometer (Model
were purchased from local market. All glassware were purchased UV–vis 200) at wavelength of 254 nm.
from Borosil. Before each experiment, all glassware were cleaned For kinetic studies, experiments were conducted with 50 mL
with dilute nitric acid and repeatedly washed with deionized cadmium nitrate solutions of concentrations (4 mg/L, 16 mg/L),
water. maintained at pH 7, temperature of 318 K at different contact time
of 5–240 min. The adsorption capacity for cadmium uptake, qe
(mg/g), was determined as follows:
2.2. Preparation of the adsorbents
V
qe = (C0 − C) (1)
The orange peel was washed with water and dried in a convec- W
tion oven at 40 ◦ C for 72 h. These were then crushed into smaller where C0 and C are the initial and final concentrations (mg/L),
particles of approximate size between 0.1 and 0.2 mm (hereafter, respectively, V is the volume of solution (L) and W is the weight
abbreviated as OPP). MNP and MNP–OPP were synthesized by co- of adsorbent (g).
precipitation method modified from Refs. [25,26]. Briefly, 6.1 g All the experiments were repeated three times and average val-
of FeCl3 ·6H2 Oand 4.2 g of FeSO4 ·7H2 O were dissolved in 100 mL ues were reported. The standard deviation was found to be ±2.12%;
water and heated to 90 ◦ C. Solution of 10 mL of ammonium hydrox- values of correlation coefficient were in the range 0.98–0.99.
ide (26%) and the solution of 1 g of OPP dissolved in 200 mL of
water were added rapidly and sequentially. The pH of the reac- 2.5. Desorption and reusability studies
tion medium was adjusted to 10. The mixture was stirred at 80 ◦ C
for 30 min and then cooled to room temperature. The black pre- Desorption studies were carried out in 5 mL of 0.1 M HNO3
cipitate Fe3 O4 –OPP (MNP–OPP) was collected by filtering, washed maintained at a constant temperature of 318 K. The Cd2+ adsorbed
to neutral with water, dried at 50 ◦ C for 24 h and finally stored for MNP–OPP were placed in the desorbing medium on a rotary shaker
further use. at 200 rpm for 30 min. The residual Cd2+ in the solution were
V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90 83

measured after MNP–OPP removal in order to estimate the amount Spectroscopic analysis shows the successful binding of OPP on to
of cadmium ions desorbed. the MNP surface. Infrared spectrum of MNP and MNP–OPP (Fig. 3)
Reusability study of MNP–OPP was carried out by following the showed that the characteristic peak of MNP at 3434.76 cm−1 (OH
adsorption–desorption study for 5 cycles. The adsorption efficiency stretch) and 575.15 cm−1 (Fe–O) underwent a significant shift in
in each cycle was analyzed. Both the adsorption and desorption MNP–OPP spectra to 3413.03 cm−1 (OH) and 578.34 cm−1 (Fe–O)
experiments were followed as described above. which indicated the interaction of the hydroxyl groups and metal-
oxide on MNP during the MNP–OPP formation. Comparison of OPP
2.6. Column experiments and MNP–OPP spectra on the other hand revealed the shifting,
disappearance and appearance of certain peaks. Significant band
Column experiments were carried out with a glass column shifting from 3423.19 cm−1 , 2925 cm−1 , 2847 cm−1 , 1626.32 cm−1
(length 30 cm, internal diameter 1 cm) filled with 10 mg of and 1033.02 cm−1 on OPP to 3413.03 cm−1 , 2929.95 cm−1 ,
MNP–OPP and the setup was established as per our previous 2284.47 cm−1 , 1624.23 cm−1 , 1019.54 cm−1 on MNP–OPP corre-
study [27]. Experiments were carried out with adsorbate solu- sponding to the bonded OH stretching, –CH, –CH2 stretching, –C O
tion of known concentration of 16 mg/L at 30 ◦ C at a flow rate of stretching in carboxyl and –C–O–C– stretching has revealed the
1.5 mL/min. Experiments were also carried out with wastewater successful binding of OPP onto MNP to form MNP–OPP [30]. The
simulating cadmium plating industry effluent. Samples of the efflu- disappearance of the peak at 1731.34 cm−1 (originally present in
ent were collected and the effluent concentrations were analyzed OPP) on MNP–OPP revealed the binding of Fe with the OH of car-
for the cadmium content. The column was shut down when the boxyl group [31]. Peaks observed on OPP at 1424.94 cm−1 (C–OH of
runoff concentration matched the initial cadmium concentration. carboxyl), 1312.44 cm−1 (C–O stretch in ester), 1262.67 cm−1 (C–O
in acid) and 1154.54 cm−1 (C–O–C stretch in ether) were found to
3. Results and discussion have disappeared on MNP–OPP. The newer peak at 578.34 cm−1
assigned to Fe–O group on MNP–OPP indicated the presence of
3.1. Characterization of MNP–OPP, MNP and OPP MNP on the composite [32]. This interaction also accounted for the
mechanism of surface modification of MNP with OPP.
The BET surface area of OPP, MNP and MNP–OPP was found to Fig. S2 showing the IR spectrum of the MNP–OPP before and after
be 47.03, 76.32 and 65.19 m2 g−1 respectively. Although MNP–OPP adsorption indicates distinct and significant changes in the absorp-
had a lower surface area than MNP, its adsorption capacity is higher tion peak frequencies of OH, C O stretch in acids, C–O stretch in
which indicates that the multiple functional groups on MNP–OPP phenols and C–O–C stretch in ethers suggesting that such ionizable
played an important role in the enhancement of the adsorption functional groups on the adsorbent surface are able to bind with
capacity. the metal ion. Similar observations were made for the IR spectrum
SEM micrographs of the three adsorbents present the morpho- of OPP and MNP.
logical characteristics favorable for metal adsorption. Fig. 1a shows The pHpzc analysis was further conducted to confirm the attach-
heterogenous, porous surface morphology of OPP having a mean ment of OPP surface onto the surface of MNP. From Fig. S3 it was
particle diameter of 60 ␮m. Fig. 1b shows a spherical morphology obvious that the pHpzc shifted from 6.78 (MNP) to 5.21 (MNP–OPP)
of MNP with particle size of 500 nm and Fig. 1c shows morphology after binding with OPP. This further revealed that MNP–OPP was
and particle size of MNP–OPP similar to those of MNP but with a positively charged at pH < 5.21.
much smoother surface. The hysteresis loop of MNP and MNP–OPP at room temperature
The particle morphology of MNP and MNP–OPP as shown in is shown in Fig. S4. The absence of remainance and coercivity as
the TEM micrographs of Fig. S1 revealed a smaller size of particles observed in the magnetic loop proved that the synthesized MNP
in the nanoscale range. The image of MNP, as shown in Fig. S1a, was superparamagnetic [25]. Also, the respective saturation mag-
showed that the MNP particles were fine, compact, monodisperse netizations of MNP–OPP and MNP were 68.1 and 79.6 emu/g, which
and had a mean diameter of 25–29 nm. The TEM image of MNP–OPP suggested that the OPP content in MNP–OPP was about 72.64%
in Fig. S1b showed that the structure of the nanoparticles was (w/w). MNP–OPP can easily be separated from its aqueous dis-
looser, leading to an increase in size; the average diameter of such persions in a few minutes with permanent hand-held magnets.
a structure was 32–35 nm. Modification of MNP with OPP resulted MNP aqueous suspension was easily oxidized to brown suspensions
in a slight but not significant agglomeration of its particles. Similar having no magnetization, whereas no significant change of the sat-
observations were reported [28]. uration magnetization and color was observed after the MNP–OPP
The XRD pattern of the OPP sample did not show any peak, was stored in water for 30 days that indicates the chemical and
which indicated the amorphous nature of the product (Fig. 2). The magnetic stability of MNP–OPP. Similar phenomenon was reported
phase structure of Fe3 O4 is revealed from XRD patterns for the MNP by [18].
and MNP–OPP samples. Six characteristic peaks for Fe3 O4 at 30.1◦ ,
35.5◦ , 43.1◦ , 53.4◦ , 57.0◦ and 62.6◦ , corresponding to their indices 3.2. Effect of environmental parameters
(2 2 0), (3 1 1), (4 0 0), (4 2 2), (5 1 1) and (4 4 0) were observed for
both samples revealing that the composite MNP–OPP nanoparticles The surface characteristics of the adsorbent surface and its metal
were pure Fe3 O4 with inverse-spinel structure [25]. These peaks are binding capacity are mainly controlled by the contact time, metal
consistent with the database in JCPDS file (PCPDFWIN v.2.02, PDF ion concentration, pH, ionic strength, natural organic matter and
No. 85-1436) indicating that the OPP binding did not result in the other parameters [33,34]. Therefore, a comparative batch adsorp-
phase change of Fe3 O4 in the composite. The average particle size tion study was carried out on OPP, MNP and MNP–OPP to study the
‘d’ of the MNP and MNP–OPP was estimated to be 25 and 32 nm effect of these environmental parameters.
respectively by using the standard Debye–Scherrer equation [29].
3.2.1. Effect of contact time and concentration
k
d= (2) Fig. 4 shows the time profile of Cd2+ removal with 0.2 g/L of OPP,
ˇ cos 
MNP and MNP–OPP at different initial Cd2+ concentrations at 45 ◦ C,
where d is the particles size, k is the Debye–Scherrer constant (0.89), from which, two important conclusions could be drawn. Firstly,
 is the X-ray wavelength (0.15406 nm) and ˇ is the full width at irrespective of initial concentration, the adsorption process for
half maximum,  is the Bragg angle. OPP almost finished within 210 min, whereas, a faster process was
84 V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90

Fig. 1. SEM images of (a) OPP (b) MNP and (c) MNP–OPP.

observed at 90 min for MNP and at 40 min for MNP–OPP. A faster major role in designing a wastewater treatment plant. Secondly, for
adsorption rate for MNP and MNP–OPP could be attributed to the all three adsorbents, the total amount of Cd2+ adsorbed increased
external surface adsorption exhibited by non-porous nano adsor- with the increasing initial Cd2+ concentrations. This is because
bent [35,36], which is different from the microporous adsorption more Cd2+ is available at higher initial concentrations which in
process exhibited by the OPP [37]. Since nearly all the adsorp- turn may have provided higher driving force for the ions from
tion sites of MNP and MNP–OPP existed on their exterior, it was the solution to the adsorbents. The result may be more colli-
easy for the adsorbate to access these active sites, thus result- sions between Cd2+ ions and active sites on the OPP, MNP and
ing in a rapid approach to equilibrium. This result is promising MNP–OPP [20]. Similar phenomena were observed in the literature
by virtue of the economic viability as equilibrium time plays a [20,38].
V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90 85

Fig. 4. Effect of contact time on the adsorption of Cd2+ onto OPP, MNP and MNP–OPP
at two different Cd2+ concentrations of 4 mg/L and 16 mg/L [average value of 3 tests,
error < 2.12%] (NaNO3 concentration 0.001 M; temperature 318 K; adsorbent dose
0.2 g/L and pH 7).

Fig. 2. XRD patterns of OPP, MNP and MNP–OPP.

3.2.2. Effect of pH and underlying mechanism


The adsorption of Cd2+ on OPP, MNP and MNP–OPP is found to
be pH dependent as revealed from Fig. 5. For OPP, the adsorption
capacity was as low as 2.63% at lower pH values but increased with
increasing pH from 36.38% to 48.0% in the pH range of 4–6, and then
appeared to level off at higher pH values. Cd2+ adsorption on MNP
increased slowly from 35.5% to 70.94% in pH range of 2.0–8.0, and
then maintained a high level with increasing pH. The same Cd2+
adsorption onto MNP–OPP increased at a much faster rate with
increasing pH from about 42.38% to 96.0% in the pH range of 2–7
and then leveled off.
The observed trend can be explained by effect of the surface Fig. 5. Effect of pH, ionic strength on the adsorption of Cd2+ onto OPP, MNP and
charge of adsorbent and pHpzc . MNP–OPP in the presence/absence of HA [average value of 3 tests, error < 2.12%]
The sole hydroxyl group on MNP, multi hydroxyl groups on (Cd2+ concentration 16 mg/L; HA concentration 10 mg/L; time 40 min for MNP–OPP,
60 min for MNP and 210 min for OPP).
OPP and MNP–OPP as revealed by FTIR spectra play a dominant
role in the Cd2+ adsorption. Depending upon the solution pH, the

Fig. 3. FT-IR spectra of OPP, MNP and MNP–OPP before adsorption.


86 V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90

adsorbent surface undergoes protonation or deprotonation [20].


The protonation/deprotonation reaction undergoes as follows:

–Fe–OH + H+  –Fe–O–H2 + for MNP (3)

–HO–(C O)–M–(OH)n + nH+  –HO–(C O)–M–(OH2 )n + for OPP

(4)

–Fe–O–(C O)–(OH)n + H+  –Fe–O–(C O)–(OH2 )n +

for MNP–OPP (5)

At pH < pHpzc (pHpzc of OPP, MNP and MNP–OPP is 3.6, 6.78


and 5.21 respectively), –Fe–O–H2 + and –Fe–O–(C O)–(OH2 )n +
are the dominant species for MNP and MNP–OPP respectively.
These species having high positive charge density make the Cd2+
Fig. 6. Adsorption isotherm of OPP–Cd2+ , MNP–Cd2+ and MNP–OPP–Cd2+ at 3 dif-
adsorption unfavorable due to electrostatic repulsion. Also, stiff ferent temperatures [average value of 3 tests, error < 2.12%] (Cd2+ concentration
competition between H+ and Cd2+ for the active sites will decrease 16 mg/L; NaNO3 concentration 0.001 M; time 40 min for MNP–OPP, 60 min for MNP
Cd2+ adsorption [39]. and 210 min for OPP; adsorbent dose 0.2 g/L and pH 7).
But, at pH > pHpzc , Fe–OH and –Fe–O–(C O)–(OH)n are the
dominant species in MNP and MNP–OPP respectively. Such depro- humic acid and MNP–OPP causes repulsion so that the adsorption is
tonated species undergo electrostatic attraction for Cd2+ that result hindered. At still higher pH, more free humic acid molecules form
in the formation of metal–ligand magnetic composite complexes. complexes with Cd2+ which diminish adsorption onto MNP–OPP
This causes enhanced adsorption [22]. [34,42].
Complexation and ion exchange appear to be the principle
mechanism of the adsorption as revealed by Eqs. (6)–(8). 3.3. Adsorption capacity and effect of temperature
2+ + +
–Fe–OH + Cd → –Fe–O–Cd + H for MNP (6)
Fig. 6 demonstrates the adsorption isotherm of the OPP, MNP,
2+ + + MNP–OPP for Cd2+ ions that were determined at the three studied
–HO–(C O)–M–(OH)n + Cd → –HO–(C O)–M–(O–Cd)n + nH
temperature of 218, 308, 328 K. The data of the Cd2+ ions adsorbed
for OPP (7) at equilibrium (qe , mg/g) and the equilibrium metal concentration
(Ce , mg/L) were fitted to the Langmuir adsorption model [43] (Fig. 7)
as per the equation:
–Fe–O–(C O)–(OH)n + Cd2+ → –Fe–O–(C O)–(O–Cd)n + + nH+ 1 1 1
= + (9)
for MNP–OPP (8) qe Q0 bQ0 Ce
where Q0 and b are Langmuir constants, corresponding to maxi-
mum adsorption capacity at complete monolayer coverage (mg/g)
3.2.3. Effect of ionic strength
and equilibrium constant (L/mg) respectively.
The ionic strengths of 0.1 M, 0.01 M and 0.001 M NaNO3
The data fit well to the model as shown in Table 1 with cor-
were chosen to investigate their effect on Cd2+ adsorptions onto
relation coefficients (R2 ) in the range 0.994–0.998, and maximum
MNP–OPP. Fig. 5 reveals that Cd2+ adsorption increased with
adsorption capacity of OPP, MNP and MNP–OPP being 40.00, 58.82
decreasing ionic strength. This phenomenon could be attributed
and 76.92 mg/g respectively. The b values (L/mg) indicated an
to two reasons: Firstly, a decreasing electrolyte concentration
ascending series of affinity to the OPP, MNP and MNP–OPP as fol-
favored the complexation tendency of Cd2+ ions with the hydroxyl
lows: OPP–Cd2+ (8.33) > MNP–Cd2+ (8.50) > MNP–OPP–Cd2+ (8.67).
functional groups on MNP–OPP, which enhances the adsorption
process. This may imply that the interaction between metal and
the magnetic composites is mainly ionic in nature. Secondly, an
increasing ionic strength of solution influenced the activity coef-
ficient of metal ions that may have limited their transfer to the
composite surfaces [40].

3.2.4. Effect of presence/absence of HA


Adsorption of Cd2+ on MNP–OPP in the presence of humic acid
is shown in Fig. 5. Cd2+ adsorption increased steeply at pH < 7
in the presence of humic acid but at pH > 7, the same shows a
slight decrease. In order to understand the mechanism underly-
ing such a phenomenon, individual adsorption of humic acid onto
MNP–OPP was carried out at different pH. Fig. S5 revealed the grad-
ual decrease of humic acid adsorption with increasing pH. At lower
pH (pH < pHpzc ), the positively charged MNP–OPP easily adsorbs
the negatively charged humic acid molecules. The surface adsorbed
humic acid induces more oxygen functional groups on MNP–OPP, Fig. 7. Langmuir plot of OPP–Cd2+ , MNP–Cd2+ , MNP–OPP–Cd2+ and
MNP–OPP–HA–Cd2+ [average value of 3 tests, error < 2.12%] (Cd2+ concentra-
allowing more Cd2+ ions to be adsorbed due to the strong com-
tion 16 mg/L; NaNO3 concentration 0.001 M; HA concentration 10 mg/L; time
plexation of Cd2+ with surface adsorbed humic acid [41]. As pH 40 min for MNP–OPP, 60 min for MNP and 210 min for OPP; temperature 318 K;
increases (pH > pHpzc ), more negative charge on hydroxyl groups of adsorbent dose 0.2 g/L and pH 7).
V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90 87

Table 1
Langmuir isotherm model and thermodynamic parameters for the adsorption of Cd2+ onto OPP, MNP and MNP–OPP.

Adsorbent C0 Temp Langmuir isotherm parameters Thermodynamic parameters


(mg/L) (K)
Q0 (mg/g) b (L/g) R2 RL G (kJ/mol) H (kJ/mol) S (kJ/mol/K)

298 35.71 5.60 0.997 0.011 −4.27 19.67 0.080


OPP 16 308 37.04 7.32 0.995 0.009 −5.10
318 40.00 9.23 0.996 0.007 −5.88

298 50.00 5.00 0.997 0.013 −3.99 22.73 0.089


MNP 16 308 52.63 6.91 0.995 0.009 −4.95
318 58.82 8.92 0.996 0.007 −5.78

298 71.43 4.67 0.997 0.013 −3.82 24.35 0.094


MNP–OPP 16 308 74.63 6.70 0.998 0.009 −4.87
318 76.92 8.67 0.994 0.007 −5.71

The high adsorption capacity exhibited by MNP–OPP may be desolvated, which enhances the sorption process. The positive S◦
explained by its nano-scale particle size giving access to a larger values indicate the affinity of the adsorbents towards Cd2+ ions in
surface area as well as the incorporation of a large number aqueous solutions and may suggest increasing degree of freedom
of hydroxyl functional groups of OPP, which provided effective at the solid–liquid interface during the sorption of metal ions on to
adsorption sites for the binding of Cd2+ ions [20,21,32]. the adsorbents [49–53].
Table 1 as well as Fig. 7 amply demonstrates the sensitivity of
the adsorption process towards temperature and it is observed that
3.5. Kinetic study
in all cases, adsorption increases with increase in temperature.
Favorable adsorption for Cd2+ is revealed further from the values
The adsorption data were simulated by pseudo-second-order
of the dimensionless constant RL (0.009–0.013) as determined [44],
model, which is expressed by the following equation [54–56]:
which is shown in Table 1.
A comparative assessment of Cd2+ with various other adsor- t 1 t
= + (12)
bents reported in the literature reveals the efficacy of the developed q k2,ads qe 2 qe
adsorbent as revealed in Table 2 [45–48].
where k2,ads is the rate constant of second-order adsorption
3.4. Thermodynamic study (g/(mg h)). The values of qe and k2,ads determined from the slope
and intercept of the plot of t/q vs. t (Fig. 9) are shown in Table 3.
The free energy change (G◦ ) for adsorption process was calcu- The calculated qe values are in agreement with the theoretical ones,
lated using the equation: and the plots show good linearity with R2 above 0.99. Therefore, the
adsorption kinetics follows the pseudo-second-order model.
G◦ = −RT ln b (10)

where b is the Langmuir constant. 3.6. Desorption and reusability studies


The values of enthalpy change (H◦ ) and entropy change (S◦ )
were calculated from the slope and intercept of the plot of G◦ vs. Regeneration of the adsorbent for repeated use is of crucial
T (Fig. 8) using the equation: importance in industrial practice for metal removal from wastewa-
G◦ = H ◦ − T S ◦ (11) ter. Desorption experiments carried out with Cd2+ laden MNP–OPP
in 0.1 M HNO3 showed that approximately 98% of the adsorbed
Table 1 which summarizes the values of these parameters cadmium was desorbed.
reveals the endothermic nature of the adsorption process (H◦ val-
ues are positive). A possible explanation is that Cd2+ ions are well
hydrated and will require breaking of the hydration sheath so as to
proceed for adsorption; this in turn requires high energy. High tem-
perature hence favors the dehydration process and ultimately the
adsorption phenomenon too. Similar observation and explanation
have been given [49].
The negative G◦ value suggests that the adsorption process
is a spontaneous process and thermodynamically favorable under
the experimental conditions. The decrease of G◦ with increasing
temperature indicates more efficient adsorption at higher temper-
ature. In addition, at higher temperature, the Cd2+ ions are readily

Table 2
Adsorption capacities for Cd2+ using different adsorbents (at room temperature).

Adsorbent pH Adsorption Reference


capacity (mg/g)

Peat 5.0 22.50 [45]


Hazel nut shells 6.0 5.42 [46]
Bamboo charcoal 8.0 12.08 [47]
Sugarcane bagasse 6.0 2.00 [48]
Fig. 8. Plot of free energy change vs. temperature for Cd2+ adsorption onto MNP–OPP
Orange peel (OPP) 7.0 35.71 This study
(Cd2+ concentration 16 mg/L; NaNO3 concentration 0.001 M; HA concentration
Orange peel–Fe2 O3 (MNP–OPP) 7.0 71.43 This study
10 mg/L; adsorbent dose 0.2 g/L and pH 7).
88 V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90

Table 3
Kinetic parameters for the adsorption of Cd2+ onto OPP, MNP and MNP–OPP.

Adsorbent qe (mg/g) (experimental) C0 (mg/L) Pseudo-second-order

qe (mg/g) (theoretical) k2,ads (g/mg/min) R2

MNP–OPP 79.60 16 80.00 0.0010 0.996


MNP 61.00 16 66.67 0.0013 0.994
OPP 38.50 16 40.00 0.0030 0.998

Table 4
Adsorption–desorption data in consecutive cycles.

Cycle # Adsorption Desorption

C0 (mg/L) Ce (mg/L) qe (mg/g) Cf (mg/L) % desorption

1 16.00 0.80 76.00 15.71 98.19


2 15.71 0.71 75.05 15.50 98.66
3 15.50 0.68 74.10 15.28 98.58
4 15.28 0.65 73.15 15.10 98.82
5 15.10 0.62 72.40 14.84 98.28

Table 5
Quality of urban wastewater simulating a typical Cd-plating industry effluent.

pH Ni2+ (mg/L) Zn2+ (mg/L) Pb2+ (mg/L) Cd2+ (mg/L)

Quality of wastewater before treatment 9.12 9.52 16.28 0.043 35.43


Quality of wastewater after treatment 7.23 6.93 9.72 0.040 6.39

Results of the adsorption capacity of MNP–OPP for five consec- cycles of adsorption–desorption that demonstrates good reusabil-
utive adsorption–desorption cycles are graphically illustrated in ity. This fulfills an important criterion for advanced adsorbents.
Fig. S6. In all cycles, desorption was 98.19–98.66%. A 4.74% decrease
in adsorption efficiency occurred (Table 4) after five consecutive 3.7. Fixed bed adsorption studies

Breakthrough capacity of the prepared column was determined


by plotting C/C0 vs. V (effluent volume) as given in Fig. 10. It is seen
that first 2 L of influent was adsorbed to the tune of 99%. 74% of
Cd2+ adsorption took place in the next 2–6 L influent followed by
a rapid decrease of adsorption (74–1.7%) for 8–20 L. The column
was exhausted and no further adsorption of cadmium occurred. A
breakthrough capacity of 55.38 mg/g was obtained signifying 55.8%
column utilization which is relatively lower than the batch adsorp-
tion capacity (76.92 mg/g). This decrease is usually observed and
generally attributed to relatively less time of interaction between
sorbent and adsorbate surface in column method.

3.8. Testing under electroplating industry effluent simulation


condition

Urban wastewater from IIT Roorkee campus was collected and


Fig. 9. Lagergren’s pseudo-second order plot of OPP–Cd2+ , MNP–Cd2+ and was spiked with nickel chloride, zinc nitrate and cadmium nitrate
MNP–OPP–Cd2+ [average value of 3 tests, error < 2.12%] (Cd2+ concentration 16 mg/L;
NaNO3 concentration 0.001 M; temperature 318 K; adsorbent dose 0.2 g/L and pH
to obtain wastewater which simulated cadmium plating industry
7). effluent. Despite the presence of competitive effect of nickel and
zinc metal ions, about 82% reduction in Cd2+ concentrations was
achieved as a result of treatment with the developed adsorbent.
This is revealed in Table 5.

4. Conclusion

This study focused on assessing the adsorption efficiency of


a novel MNP–OPP adsorbent developed from a local agricultural
waste-orange peel powder for a toxic cadmium metal ion.
Characterization studies revealed the various physico-chemical
properties of MNP–OPP which are favorable for metal binding.
Comparative preliminary batch studies demonstrated not only the
optimal process parameters for maximizing metal sorption but
also the principle underlying mechanism. Langmuir adsorption
plot showed maximum Cd2+ removal by MNP–OPP at 76.92 mg/g
at optimal conditions and thermodynamic studies revealed
Fig. 10. Breakthrough curve of Cd2+ adsorption onto MNP–OPP. the feasibility of the process. Pseudo-second-order model was
V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90 89

determined as the best fit model. Column studies with a break- [21] Y.C. Chang, D.H. Chen, Preparation and adsorption properties of monodisperse
through capacity of 55.38 mg/g as well as 82% Cd2+ removals from chitosan-bound Fe3 O4 magnetic nanoparticles for removal of Cu(II) ions, J. Col-
loid Interface Sci. 283 (2005) 446–451.
an electroplating effluent simulated wastewater revealed the prac- [22] A.Z.M. Badruddoza, A.S.H. Tay, P.Y. Tan, K. Hidajat, M.S. Uddin, Carboxymethyl-
tical utility of the developed adsorbent. Finally, desorption and ␤-cyclodextrin conjugated magnetic nanoparticles as nano-adsorbents for
reusability studies indicate a fulfilling of important criteria for removal of copper ions: synthesis and adsorption studies, J. Hazard. Mater.
185 (2011) 1177–1186.
advanced adsorbents. [23] N.C. Feng, X.Y. Guo, S. Liang, Adsorption study of copper (II) by chemically
The developed MNP–OPP has demonstrated not only high modified orange peel, J. Hazard. Mater. 164 (2009) 1286–1292.
adsorption efficiency, faster kinetics but also have shown additional [24] S. Liang, X.Y. Guo, N.C. Feng, Q.H. Tian, Adsorption of Cu2+ and Cd2+ from aque-
ous solution by mercapto-acetic acid modified orange peel, Colloids Surf. B:
benefits like ease of synthesis, easy recovery, absence of secondary
Biointerfaces 73 (2009) 10–14.
pollutants, cost-effectiveness and environmental-friendliness. It [25] M.-H. Liao, D.-H. Chen, Preparation and characterization of a novel magnetic
can be concluded to be a promising advanced adsorbent in envi- nano-adsorbent, J. Mater. Chem. 12 (2002) 3654–3659.
[26] V.K. Gupta, S. Agarwal, T.A. Saleh, Chromium removal by combining the mag-
ronmental pollution cleanup.
netic properties of iron oxide with adsorption properties of carbon nanotubes,
Water Res. 45 (2011) 2207–2212.
Acknowledgement [27] V.K. Gupta, B. Gupta, A. Rastogi, S. Agarwal, A. Nayak, A comparative investi-
gation on adsorption performances of mesoporous activated carbon prepared
from waste rubber tire and activated carbon for an azo dye, J. Hazard. Mater.
The author (A. Nayak) is thankful to Ministry of Human Resource 186 (2011) 891–901.
Development (MHRD), New Delhi, India, for financial support. [28] D.T.K. Dung, T.H. Hai, L.H. Phuc, B.D. Long, L.K. Vinh, P.N. Truc, Preparation and
characterization of magnetic nanoparticles with chitosan coating, J. Phys.: Conf.
Ser. 187 (2009) 012036.
Appendix A. Supplementary data [29] C.H. Yu, A. Al-Saadi, S. Shih, L. Qiu, K.Y. Tam, S.C. Tsang, Immobilization of BSA
on silica-coated magnetic iron oxide nanoparticle, J. Phys. Chem. C 113 (2009)
537–543.
Supplementary data associated with this article can be found, in [30] X. Li, Y. Tang, X. Cao, D. Lu, F. Luoa, W. Shao, Preparation and evaluation
the online version, at doi:10.1016/j.cej.2011.11.006. of orange peel cellulose adsorbents for effective removal of cadmium, zinc,
cobalt and nickel, Colloids Surf. A: Physicochem. Eng. Aspects 317 (2008)
512–521.
References [31] Y.C. Chang, S.W. Chang, D.H. Chen, Magnetic chitosan nanoparticles: studies on
chitosan binding and adsorption of Co(II) ions, React. Funct. Polym. 66 (2006)
[1] V.K. Gupta, I. Ali, Removal of lead and chromium from wastewater using 335–341.
bagasse fly ash – a sugar industry waste, J. Colloid Interface Sci. 271 (2004) [32] P. Panneerselvam, N. Morad, K.A. Tan, Magnetic nanoparticle (Fe3 O4 ) impreg-
321–328. nated onto tea waste for the removal of nickel(II) from aqueous solution, J.
[2] D.H.F. Liu, B.G. Liptack, P.A. Bouis, Environmental Engineer’s Handbook, 2nd Hazard. Mater. 186 (2011) 160–168.
edn, Lewis, Boca Raton, FL, USA, 1997. [33] V.K. Gupta, A. Rastogi, A. Nayak, Adsorption studies on the removal of hexava-
[3] S.B. Yang, J. Hu, C.L. Chen, D.D. Shao, X.K. Wang, Mutual effect of Pb(II) and lent chromium from aqueous solution using a low cost fertilizer industry waste
humic acid adsorption onto multiwalled carbon nanotubes/poly(acrylamide) material, J. Colloid Interface Sci. 342 (2010) 135–141.
composites from aqueous solution, Environ. Sci. Technol. 45 (2011) 3621–3627. [34] S. Wang, J. Hu, J. Li, Y. Dong, Influence of pH, soil humic/fulvic acid, ionic
[4] V.K. Gupta, P.J.M. Carrott, M.M.L. RibeiroCarrott, Suhas, Low cost adsorbents: strength, foreign ions and addition sequences on adsorption of Pb(II) onto GMZ
growing approach to wastewater treatment – a review, Crit. Rev. Environ. Sci. bentonite, J. Hazard. Mater. 167 (2009) 44–51.
Technol. 39 (2009) 783–842. [35] A. Uheida, M. Iglesias, C. Fontàs, M. Hidalgo, V. Salvadó, Y. Zhang,
[5] I. Ali, V.K. Gupta, Advances in water treatment by adsorption technology, Nat. M. Muhammed, Sorption of palladium (II), rhodium (III), and platinum
Protoc. 1 (2007) 2661–2667. (IV) on Fe3 O4 nanoparticles, J. Colloid Interface Sci. 301 (2006) 402–
[6] X. Huang, N.Y. Gao, Q.L. Zhang, Thermodynamics and kinetics of cadmium 408.
adsorption onto oxidized granular activated carbon, J. Environ. Sci. 19 (2007) [36] N.N. Nassar, Rapid removal and recovery of Pb (II) from wastewater by magnetic
1287–1292. nanoadsorbents, J. Hazard. Mater. 184 (2010) 538–546.
[7] M.R. Panuccio, A. Sorgonà, M. Rizzo, G. Cacco, Cadmium adsorption on vermi- [37] A.L. Nemr, O. Abdelwahab, A. El-Sikaily, A. Khaled, Removal of direct blue-86
culite zeolite and pumice: batch experimental studies, J. Environ. Manage. 90 from aqueous solution by new activated carbon developed from orange peel, J.
(2009) 364–374. Hazard. Mater. 161 (2009) 102–110.
[8] J. Hizal, R. Apak, Modeling of cadmium (II) adsorption on kaolinite-based clays [38] J. Hu, G. Chen, I.M.C. Lo, Selective removal of heavy metals from industrial
in the absence and presence of humic acid, Appl. Clay Sci. 32 (2006) 232–244. wastewater using maghaemite nanoparticle: performance and mechanisms,
[9] J.T. Bamgbose, S. Adewuyi, O. Bamgbose, A.A. Adetoye, Adsorption kinetics of J. Environ. Eng. 132 (2006) 709–715.
cadmium and lead by chitosan, Afr. J. Biotechnol. 9 (2010) 2560–2565. [39] H. Yong-Meia, C. Mana, H. Zhong-Bob, Effective removal of Cu (II) ions from
[10] E.W. Shin, K.G. Karthikeyan, M.A. Tshabalala, Adsorption mechanism of cad- aqueous solution by amino-functionalized magnetic nanoparticles, J. Hazard.
mium on juniper bark and wood, Bioresour. Technol. 98 (2007) 588–594. Mater. 184 (2010) 392–399.
[11] K. Sevgi, Adsorption of Cd (II), Cr (III) and Mn (II) on natural sepiolite, Desali- [40] C. Chen, J. Hu, D. Shao, J. Li, X. Wang, Adsorption behavior of multiwall carbon
nation 244 (2009) 24–30. nanotube/iron oxide magnetic composites for Ni(II) and Sr(II), J. Hazard. Mater.
[12] G.C. Panda, S.K. Das, A.K. Guha, Biosorption of cadmium and nickel by func- 164 (2009) 923–928.
tionalized husk of Lathyrus sativus, Colloids Surf. B: Biointerfaces 62 (2008) [41] A.P. Davis, V. Bhatnagar, Adsorption of cadmium and humic acid onto hematite,
173–179. Chemosphere 30 (1995) 243–256.
[13] V. Srivastava, C.H. Weng, V.K. Singh, Y.C. Sharma, Adsorption of nickel ions from [42] W. Yantasee, C.L. Warner, T. Sangvanich, R.S. Addleman, T.G. Carter, R.J. Wiacek,
aqueous solutions by nano alumina: kinetic, mass transfer, and equilibrium G.E. Fryxell, C. Timchalk, M.G. Warner, Removal of heavy metals from aqueous
studies, J. Chem. Eng. Data 56 (2011) 1414–1422. systems with thiol functionalized superparamagnetic nanoparticles, Environ.
[14] V.K. Gupta, S. Agarwal, T.A. Saleh, Synthesis and characterization of alumina- Sci. Technol. 41 (2007) 5114–5119.
coated carbon nanotubes and their application for lead removal, J. Hazard. [43] I. Langmuir, The constitution and fundamental properties of solids and liquids,
Mater. 185 (2011) 17–23. J. Am. Chem. Soc. 38 (1916) 2221–2295.
[15] Y. Feng, J.-L. Gong, G.-M. Zeng, Q.-Y. Niu, H.-Y. Zhang, C.-G. Niu, J.-H. Deng, [44] T.M. Weber, R.K. Chakrabarti, Pore and solid diffusion models for fixed-bed
M. Yan, Adsorption of Cd (II) and Zn (II) from aqueous solutions using mag- adsorbers, Am. Inst. Chem. Eng. J. 20 (1974) 228.
netic hydroxyapatite nanoparticles as adsorbents, Chem. Eng. J. 162 (2010) [45] T. Gosset, J.L. Transcart, D.R. Thevenot, Batch metal removal by peat: kinetics
487–494. and thermodynamics, Water Res. 20 (1986) 21–26.
[16] J. Hu, G. Chen, I.M.C. Lo, Removal and recovery of Cr(VI) from wastewater by [46] G. Cimino, A. Passerini, G. Toscano, Removal of toxic cations and Cr(VI) from
maghaemite nanoparticles, Water Res. 39 (2005) 4528–4536. aqueous solution by hazelnut shell, Water Res. 34 (2000) 2955–2962.
[17] Y.F. Shen, J. Tang, Z.H. Nie, Y.D. Wang, Y. Ren, L. Zuo, Preparation and application [47] F. Yuan Wang, H. Wang, J.W. Ma, Adsorption of cadmium (II) ions from aqueous
of magnetic Fe3 O4 nanoparticles for wastewater purification, Sep. Sci. Technol. solution by a new low-cost adsorbent – bamboo charcoal, J. Hazard. Mater. 177
68 (2009) 312–319. (2010) 300–306.
[18] J.F. Liu, Z.S. Zhao, G.B. Jiang, Coating Fe3 O4 magnetic nanoparticles with humic [48] V.K. Gupta, C.K. Jain, I. Ali, M. Sharma, V.K. Saini, Removal of cadmium and nickel
acid for high efficient removal of heavy metals in water, Environ. Sci. Technol. from wastewater using bagasse fly ash – a sugar industry waste, Water Res. 37
42 (2008) 6949–6954. (2003) 4038–4044.
[19] S.H. Huang, D.H. Chen, Rapid removal of heavy metal cations and anions from [49] C.L. Chen, X.K. Wang, Adsorption of Ni (II) from aqueous solution using oxidized
aqueous solutions by an amino-functionalized magnetic nano-adsorbent, J. multiwall carbon nanotubes, Ind. Eng. Chem. Res. 45 (2006) 9144–9149.
Hazard. Mater. 163 (2009) 174–179. [50] V.K. Gupta, A. Rastogi, Biosorption of lead from aqueous solutions by non-living
[20] S.S. Banerjee, D.H. Chen, Fast removal of copper ions by gum arabic modified algal biomass Oedogonium sp. and Nostoc sp. – a comparative study, Colloids
magnetic nano-adsorbent, J. Hazard. Mater. 147 (2007) 792–799. Surf. B 64 (2) (2008) 170–178.
90 V.K. Gupta, A. Nayak / Chemical Engineering Journal 180 (2012) 81–90

[51] V.K. Gupta, A. Rastogi, Sorption and desorption studies of chromium (VI) from [54] Y. Ho, G.G. McKay, Pseudo-second-order model for sorption processes, Process
nonviable cyanobacterium Nostoc muscorum biomass, J. Hazard. Mater. 154 Biochem. 34 (1999) 451–465.
(2008) 347–354. [55] V.K. Gupta, I. Ali, Removal of endosulfan and methoxychlor from water on
[52] V.K. Gupta, A. Rastogi, M.K. Dwivedi, D. Mohan, Process development for the carbon slurry, Environ. Sci. Technol. 42 (2008) 766–770.
removal of zinc and cadmium from wastewater using slag – a blast-furnace [56] A.K. Jain, V.K. Gupta, L.P. Singh, Neutral carrier and organic resin based
waste material, Sep. Sci. Technol. 32 (1997) 2883–2912. membranes as sensors for uranyl ions, Anal. Proc. Anal. Commun. 32 (1995)
[53] V.K. Gupta, D. Mohan, S. Sharma, Removal of lead from wastewater using 263–265.
bagasse fly ash – a sugar industry waste material, Sep. Sci. Technol. 33 (1998)
1331–1343.

También podría gustarte