You are on page 1of 9

Chemical Engineering Journal 225 (2013) 607–615

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Ionically modified magnetic nanomaterials for arsenic and chromium


removal from water
Abu Zayed Md. Badruddoza a,1, Zayed Bin Zakir Shawon a,1, Md. Taifur Rahman b, Kow Wei Hao a,
Kus Hidajat a, Mohammad Shahab Uddin c,⇑
a
Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, Singapore 117576, Singapore
b
School of Chemistry and Chemical Engineering, The Queen’s University of Belfast, Belfast BT9 5AG, Northern Ireland, United Kingdom
c
Department of Biotechnology Engineering, Faculty of Engineering, International Islamic University Malaysia, P.O. Box 10, 50728 Kuala Lumpur, Malaysia

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Ionically modified magnetic 60

adsorbent was synthesized for As(V) 50

and Cr(VI) removal. 40


qe (mg/g )

 Maximum uptakes for As(V) and 30

Cr(VI) were 50.5 mg/g and 35.2 mg/g, 20


As(V)
respectively. 10 Cr(VI)
Fitting by Langmuir model
 Adsorption occurred via electrostatic 0
0 50 100 150
interaction and ion exchange Ce (mg/L)

mechanism. M
-

 Adsorption is dependent on pH and O - O -


I- P+ -I M
Fe3O4 O Si Fe3O4 O Si P+
O Ion exchange O

independent of coexisting ions except


phosphate. - - -
M = H2AsO4 or HCrO4 /Cr2O7
2-

 It can be magnetically separable,


reusable and promising for anionic
metal removal.

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

Article history: We describe a method for the removal of arsenic and chromium contaminants from water by using ion-
Available online 6 April 2013 ically modified magnetic nanoparticles (PPhSi-MNPs). We synthesized and covalently attached a cationic
ligand: phosphonium-silane (PPhSi), onto iron oxide magnetic nanoparticles. We exploit the ion-
Keywords: exchange capability of the phosphonium ligand for the removal of the metal ions from water and super-
Magnetic nanoparticles paramagnetic nature of iron oxide for physical separation of the sorbent materials for recovery and
Ionic liquids regeneration. Chemical and physical properties of the phosphonium-coated MNPs are investigated by
Arsenate
FTIR, TEM, XPS, EDX, XRD and zeta potential. The highest As(V) and Cr(VI) adsorption efficiency is
Chromate
Adsorption
observed at pH 3. The adsorption of As(V) and Cr(VI) on PPhSi-MNPs is not affected to any considerable
Desorption extent by the presence of coexisting anions such as chloride, nitrate, and sulfate except phosphate. The
adsorption data are best fitted with the Langmuir isotherm model and a pseudo-second-order kinetic
model is proposed. Based on the zeta-potential and chemical analyses of the adsorbent surface, we pro-
pose that a synergy of electrostatic interaction and ion-exchange between anionic metals and the cation-
modified magnetic nanoparticles allow highly efficient removal of the pollutants. A simple pH-triggered
desorption and regeneration of the adsorbent is also presented. With adsorption capacities for arsenic
(50.5 mg/g) and chromium (35.2 mg/g), coupled with facile magnetic separation, this nanoadsorbent
can be an attractive material for the removal of several anionic metal species from water.
Ó 2013 Elsevier B.V. All rights reserved.

⇑ Corresponding author. Tel.: +60 3 61966519; fax: +60 3 61964442


E-mail address: uddinms@iium.edu.my (M.S. Uddin).
1
These authors contributed equally to this work.

1385-8947/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2013.03.114
608 Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615

1. Introduction Ionic liquids (ILs) are of particular interest in recent years, since
these unique materials have shown potentials as extracting agents
The presence of toxic heavy metals in water poses serious for elimination of heavy metals from wastewater [36–38]. ILs are
health and environmental hazard. Hence, significant research generally defined as organic salts that are liquid below 100 °C
activities are aimed for analysis, minimization, containment and and consist entirely of ions. ILs can absorb contaminants by elec-
mitigation of such pollution. Arsenic stands as one of the most trostatic interaction, and preferential solubilization of polar mole-
toxic and carcinogenic chemical element, which is a common nat- cules and ions. ILs show extraordinary properties such as an
urally occurring material found in the surface and ground water as extremely low vapor pressure, high thermal stability and their
a result of weathering of rocks, industrial waste discharges, agri- physico-chemical properties can be tuned by modifying their
cultural use of arsenical herbicides, pesticides, crop desiccants, chemical structure. ILs have been also widely used as solvents in
etc [1–3]. It is generally mobilized through natural processes and organic and inorganic syntheses, catalysis, and chromatography
results in severe environmental hazards and chronic poisoning [39–41]. Several types of ionic liquids based on imidazolium-,
[4,5]. Like arsenic, chromium is another highly toxic heavy metal pyridinium-, pyrrolidinium- or phosphonium cations have been
and a common element of earth’s crust [6]. Chromium acts as car- investigated for the extraction and separation of inorganic sub-
cinogens, mutagens, and teratogens in biological systems [7]. It stances from aqueous media [42–44]. Although bulk ionic liquid
enters the aquatic systems through discharge of concentrated has been recognized as an efficient fluidic sorbent for removal of
effluents mainly from leather tanning, smelting, electroplating, heavy metals from water and wastewater solutions, grafting ionic
paint, textile and paper industries. Both of these toxic metals grad- liquid molecules on the nanoparticles as a coating layer for metal
ually accumulate in the living organisms and thereby cause disor- ion removal has been rarely exploited. Compared to the use of bulk
ders and life threatening diseases [7,8]. World Health Organization ionic liquid, covalent grafting of IL ligands onto nanoadsorbents
(WHO) has set a maximum limit of 10 lg/L arsenic in the drinking would be operationally simpler, because (i) most ionic liquids are
water, whereas US Environmental Protection Agency (EPA) has set highly viscous materials: biphasic mixing and separation is
the maximum contaminate level for chromium at 0.1 mg/L [2,7,9]. energy-intensive due to mass transfer, (ii) ionic liquid recovery
Due to their toxic, carcinogenicity and harmful nature, and also can be an issue as many IL have finite solubility in water, and most
legislative bindings, the removal of arsenic and chromium pollu- importantly (iii) typically, ILs are expensive materials. In our case
tants from various water sources is very crucial and critical. IL molecules are chemically grafted on the inexpensive iron oxide
Several treatment technologies have been developed for the nanoparticles, which would increase the surface area and the ac-
removal of arsenic and chromium from water: chemical precipita- tive surface sites for metal ion adsorption, and facile magnetic sep-
tion, ion exchange, membrane separation, filtration/ultrafiltration, aration would allow fast and complete separation of the sorbent
electrocoagulation, extraction, sedimentation, reduction, reverse phase for regeneration and recycling.
osmosis, dialysis/electrodialysis, adsorption, etc [3,10]. Among We ionically modified the iron oxide nanoparticles with an io-
these techniques adsorption is most popular because of its simplic- nic liquid ligand which is designed to possess a trimethoxysilane
ity, easy handling and sludge-free operation, regeneration capacity pendant group for attaching onto the iron oxide surface and a
and cost effectiveness [11,12]. For removal of arsenic and chro- phosphonium iodide head group for chemically interacting with
mium from water several adsorbents such as, oxides and hydrox- the metal ions. These easily separable and recyclable nanoadsor-
ides of metals, activated carbon, polymeric resins, clay minerals bents were characterized by FTIR, TEM, XPS and zeta potential.
and zeolites have been reported [3,8,13–21]. Recently, the exploi- Influence of external parameter such as, pH, competitive anions,
tation of nanomaterials for removal of metallic pollutants from contact time on the extent of adsorption of As(V) and Cr(VI) were
water or wastewater has emerged as an intriguing research direc- investigated. Adsorption equilibrium and kinetic properties of this
tion. Because, compared to the bulk counterparts, nanomaterials- adsorption process were also studied. The As(V) and Cr(VI)-sorbed
based adsorbents have a much larger surface area to accommodate PPhSi-MNPs were analyzed with XPS and EDX for elucidating the
far greater number of active sites for interaction/adsorption. Differ- adsorption mechanism.
ent types of nanomaterials such as, carbon nanotube, iron oxide,
aluminum oxide, and titanium oxide were employed in many stud- 2. Materials and methods
ies as nanoadsorbents and showed excellent adsorption capacities
for various heavy metals including arsenic and chromium [22–29]. 2.1. Chemicals
Recently, iron oxide nanoparticles attracted great interest in the
field of environmental remediation because these superparamag- Iron (II) chloride tetrahydrate, iron (III) chloride hexahydrate
netic materials can be easily dispersed in, and quickly retrieved and sodium hydrogen arsenate were purchased from Alfa Aesar
from water simply by switching off and on an external magnetic (Singapore). Ammonium hydroxide (25%) and toluene were
field. Hence, retrieval in most cases nearly 100% and reuse is facile. purchased from Merck (Germany). Triphenylphosphine and
To avoid the oxidation of magnetic nanoparticles and also to in- 3-iodopropyl (trimethoxy)silane were purchased from Sigma–Al-
crease the stability and adsorption capacities of nanoparticles, dif- drich (Singapore). For As(V) and Cr(VI) adsorption/desorption stud-
ferent coating materials or ligands have been employed to modify ies, potassium dichromate (K2Cr2O7) and disodium hydrogen
the surface. Over the past few years several magnetic nanoadsor- arsenate (Na2HAsO4) salts were used. The water used in this entire
bents are developed, e.g., magnetite (Fe3O4), diatomite sup- work was Milli-Q ultrapure water.
ported/unsupported magnetite nanoparticles, flower like Fe2O3
nanoparticles, magnetite-reduced graphene oxides composites, 2.2. Synthesis of phosphonium silane
CTAB modified magnetic nanoparticles, polypyrrole/Fe3O4 mag-
netic nanocomposite, surface modified jecobosite nanoparticles Phosphonium silane was synthesized by the condensation reac-
and calix[4]arene-grafted magnetite nanoparticles [22,29–35]. In tion between triphenylphosphine and 3-iodopropyl(trimeth-
this work, we utilize an ionic liquid analogue (phosphonium based oxy)silane. 1 mL of 3-iodopropyl (trimethoxy)silane and 1.34 gm
silane, PPhSi) as an ion-exchange coating material to modify the triphenylphosphine (1:1 molar ratio) was taken in a schlenk tube.
Fe3O4 nanoparticles to accomplish high adsorption capacity for an- 5 mL of toluene was added into the tube and the components were
ionic pollutants and easy magnetic separation. dissolved completely by magnetic stirrer. Meanwhile, the air was
Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615 609

replaced by nitrogen gas flow to create inert atmosphere. The Ci  Ce


ð%Þ Removal ¼  100% ð2Þ
openings of the tube were sealed to prevent air insertion. Then Ci
the mixture was placed in a silica oil bath at 100 °C. Magnetic stir-
Desorption study was conducted using different desorption elu-
rer was set at 1000 rpm. Within few minutes the clear mixture
ents such as 0.1 M NaOH, Na2HPO4, NaHCO3 and Na2CO3. Adsorp-
became turbid and white in color. The degree of turbidity increased
tion was first conducted using 120 mg of wet PPhSi-MNPs in
with the progress of the reaction and the reaction was continued
10 mL of 200 mg/L As(V) and Cr(VI) solution at pH 3 and shaken
overnight. The product phosphonium silane which is insoluble in
for 2 h. Desorption was then examined by adding 10 mL of either
toluene was separated by centrifugation and washed 4–5 times
desorption eluent to the metal-sorbed PPhSi-MNPs. After shaking
with toluene to remove any unreacted materials. Finally this mate-
at 230 rpm for 4 h, PPhSi-MNPs were collected by magnetic decan-
rial was dried in a vacuum dryer for 6 h at room temperature to
tation and the concentration of each pollutant in the supernatant
obtain the pure product. In the IR spectrum of PPhSi (Fig. S1), the
was measured by ICP-MS.
typical absorption bands are at: 2841, 2885, 2941 (ACH2), 3011
(CAH, arom. ring), 1585 (C@C, arom. ring), 1435, 1483 (ACH2),
1000–1150 (SiAOA), 800 (CAP), 723 (CH, arom. ring). 2.5. Characterization of the materials

2.3. Synthesis of phosphonium silane coated magnetic nanoparticles Field Emission Transmission Electron Microscopy (JEOL 2011F)
(PPhSi-MNPs) was used to determine the size and morphology of MNPs. The sam-
ple was prepared by coating a thin layer of diluted magnetic parti-
Phosphonium silane coated magnetic nanoparticles PPhSi- cle suspension on a copper grid (200 mesh and cover with formvar/
MNPs were fabricated by one step co-precipitation method. Typi- carbon). The copper film was then dried at room temperature for
cally, 0.86 g of iron (II) chloride tetrahydrate (FeCl24H2O), 2.36 g 24 h before the measurement. Chemical attachment of phospho-
of iron (III) chloride hexahydrate (FeCl36H2O) and 0.15 g of phos- nium silane onto the surface of Fe3O4 nanoparticles was examined
phonium silane (PPhSi) were dissolved in 40 mL of de-aerated by Fourier transform infrared spectroscopy (FTIR) and X-ray photo-
Milli-Q water with vigorous stirring at a speed of 1,200 rpm. electron spectroscopy (XPS). FTIR measurements were performed
5 mL of 25% ammonium hydroxide (NH4OH) was added after the using a Shimadzu infrared spectrometer (Model 400) with KBr as
solution temperature was reached to 90 °C. The reaction was con- background over the range of 4000–400 cm1. X-ray photoelectron
tinued for 1 h at 90 °C under constant stirring and inert atmo- spectroscopy (XPS) analysis was carried out on an Axis Ultra DLD
sphere created by flowing N2 gas. The resulting nanoparticles (Kratos) spectrometer with Al mono Ka X-ray source (1486.71 eV
were then washed with Milli-Q water 4–5 times to remove the photons) to determine the elements. All binding energies (BEs)
unreacted chemicals species. Bare Fe3O4 MNPs were synthesized were referenced to the C 1s neutral carbon peak at 284.6 eV. The
following the same procedure described above, except no PPhSi zeta potentials of as-synthesized nanoparticles were measured at
was added in the reaction mixture. different pH using a Malvern ZEN 3600 Zetasizer Nano ZS.

2.4. Adsorption and desorption of As(V) and Cr(VI) ions 3. Results and discussion

As(V) and Cr(VI) adsorption experiments were carried out using 3.1. Synthesis and characterization of magnetic nanoparticles
batch equilibrium technique in aqueous solutions at optimum pH 3
and at 25 °C. Briefly, 120 mg of wet magnetic nanoadsorbents (20% The synthesis of PPhSi-MNPs is carried out by using a simple
dry particle content) was added to 10 mL of As(V) and Cr(VI) solu- one-step method where precipitation of Fe3O4 MNPs and the
tion of different concentrations (i.e., 10–200 mg/L) and shaken in a attachment of PPhSi on Fe3O4 crystal surface occurs at the same
thermostatic water-bath shaker operated at 230 rpm. After equi- process. Fe3O4 nanoparticles are formed in the presence of PPhSi
librium was reached, magnetic nanoadsorbents were removed by in the reaction mixture: iron oxide nanoparticles has free hydroxyl
using a permanent Nd–Fe–B magnet and the supernatant was col- (AOH) groups on its surface, hence, condensation between OH
lected. The concentrations of As(V) and Cr(VI) ions were deter- with hydrolyzed trimethoxy silane group (SiAOH) in the alkaline
mined using Inductive Couple Plasma Mass Spectrometry environment results in the attachment of phosphonium group onto
(Agilent ICP-MS 7700 series). The influence of pH (from 3 to 9), the Fe3O4 surface. The reaction steps are in shown in Scheme 1.
and competitive anions (nitrate, chloride, sulfate, and phosphate) Covalent grafting of the phosphonium silane on the iron oxide
on the As(V) and Cr(VI) adsorption at pH 3 were also investigated surface is evident from the corresponding FTIR traces. FTIR spec-
with the initial concentration of each metal of 100 mg L1. The trum of the PPhSi-MNPs in the range of 400–4000 cm1 is pre-
solution pH was adjusted by NaOH or HCl. For the kinetic experi- sented in Fig. 1a. As shown in Fig. 1a, the peak at 586 cm1 is
ments, the same initial concentrations of As(V) and Cr(VI) were attributed to the FeAO bond vibration. The prominent absorption
used at pH 3. At various time intervals, samples were collected bands were ascribed to OH (3400 cm1), SiAOASi or SiAOAFe (ms
after quick magnetic decantation and the concentrations of studied 1100 cm1, mas 807 cm1), SiAOH (ms 950 cm1) (where ms and mas
pollutants were determined as mentioned previously. The amount refer to the symmetric and asymmetric stretching, respectively)
of metal ions adsorbed onto PPhSi-MNPs was calculated by a mass [45]. The bands between 650–750 cm1 and 1508–1631 cm1
balance relationship: could be assigned to CAH bending and aromatic C@C stretching
respectively, while absorption band at 3014 cm1 is assigned for
V vibration of aromatic CAH stretching.
Q e ¼ ðC t  C e Þ ð1Þ
w TEM image of PPhSi-MNPs is shown in Fig. 1b. It shows that
where V (L) is the volume of the aqueous solution, Ci and Ce (mg/L) Fe3O4 nanocomposites modified with phosphonium based silane
are the initial and final solution concentration of metal ions, respec- are spherical particles with size range from 8 to 15 nm. The crystal-
tively, and w (g) is the dry mass of the solid. The percentage line structure of the adsorbent was characterized by XRD as shown
removal efficiency was calculated as described by the following in Fig. S2 (Supplementary data). The diffraction peaks at 30.2°
equation: (2 2 0), 35.5° (3 1 1), 43.15° (4 0 0), 53.4° (4 2 2), 57° (5 1 1) and
62.9° (4 4 0) are ascribed to Fe3O4 [34]. However, the modification
of magnetic nanoparticle surfaces with PPhSi does not result in the
610 Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615

OCH3 OCH3

Reflux, N2 I-
CH3O Si I + P CH3O Si P+

OCH3 OCH3

PPhSi

OCH3

I- O I-
Fe2+ + Fe3+ + CH3O Si P+ Fe3O4 O Si P+
NH4OH O
OCH3

PPhSi-MNPs

Scheme 1. Synthesis steps of PPhSi and PPhSi-MNPs.

phase change of Fe3O4. The average crystal size of PPhSi-MNPs to be about 91% and 7.1%, respectively from the weight loss of the
calculated to be 10.2 nm from the XRD pattern by using Debye– samples measured by thermogravimetric analysis (Fig. S3, Supple-
Scherrer’s equation, was consistent with that observed from the mentary data). The saturation magnetization curves of bare Fe3O4
TEM images. The BET surface area of PPhSi-MNPs was found to MNPs and PPhSi-MNPs were obtained and presented in Fig. S4
be 105.7 m2/g; the relatively high surface area is beneficial for (Supplementary data). The magnetic hysteresis loop of the samples
the adsorption of toxic chemicals from wastewater. The average indicates no remanence and no coercivity and reveals their super-
mass contents of Fe3O4 and PPhSi on PPhSi-MNPs were calculated paramagnetic nature, which is beneficial for their dispersion/coag-
ulation upon removal/exposure of the colloidal mixture to an
external magnetic field. The saturation magnetization of PPhSi-
(a) MNPs was found to be as high (68.2 emu/g) as that of bare Fe3O4
(74 emu/g) ensuring quick response of these ionically modified
nanoadsorbents to external magnetic fields.
The zeta potentials of bare MNPs and PPhSi-MNPs were mea-
% Transmission

sured in a wide range of pH (from 2 to 11) and the corresponding


860 1111 results are presented in Fig. 2. The zeta potentials of both bare and
941
coated nanoparticles progressively changed from positive to nega-
tive when the pH was increased from 3 to 11. Fig. 2 shows that the
741
pHpzc of bare Fe3O4 MNPs was around 6.1, which is very close to lit-
2840
717
3014
erature value (pHpzc 6.0) [46]. After the coating with PPhSi, the zeta
690 3400 potentials move to significantly higher positive values compared to
the bare Fe3O4 MNPs over a range of pH (2-7), and the pHpzc has
586
shifted to 6.7. Moreover, the surface charge of MNPs is significantly
increased after coating with PPhSi. The cationic charge on phos-
400 900 1400 1900 2400 2900 3400 3900
phonium head groups contribute to enhance the positive surface
Wavenumber, cm-1 charge of the original unmodified Fe3O4 particles. Similar observa-
tion was reported in previous studies. For example, iron oxide
(b) nanoparticles modified with similar ionic liquid-based ligands
showed higher positive charge even at pH > 8 (compared to
unmodified parent particles), due to the presence of cationic head
groups (which were also coordinated with an anion of equal
charge) [45,47].

Fig. 1. (a) FTIR spectrum of phosphonium silane coated magnetic nanoparticles


(PPhSi-MNPs). (b) TEM image of PPhSi-MNPs. Fig. 2. Zeta potentials of bare Fe3O4 MNPs and PPhSi-MNPs at different pH (25 °C).
Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615 611

at the initial stage and reaches equilibrium within 50 min (for


40 AAs(V) As(V)) and 70 min (Cr(VI), respectively. The pseudo-second-order
Cr(VI)
kinetic model was adopted to test the experimental data which
C
helps shed light on the possible adsorption mechanism.
30
qe (mg/g)

The pseudo-second-order adsorption kinetic rate equation in


the integrated form could be expressed as follows [49]:
20

t 1 1
¼ þ t ð3Þ
10 qt k2 q2e qe

0 where qe and qt refer to the adsorption capacity of metal ions (mg/g)


2 4 6 8 10 at equilibrium and at any time, t (min), respectively, and k2 is the
pH rate constant of pseudo-second-order adsorption (g mg1 min1).
The slope and intercept of the plot of t/qt versus t are used to calcu-
Fig. 3. Effect of solution pH on the adsorption of As(V) and Cr(VI) onto PPhSi-MNPs
at 25 °C.
late k2 and qe,cal, respectively (Fig. 4b). The kinetic parameters
obtained in fitting the experimental data are summarized in Table 1.
The applicability of this model was quantified by the correlation
3.2. Adsorption of As(V) and Cr(VI) ions coefficient, R2 values, and the closeness of R2 to 1 indicates that
the model fitted the experimental data accurately. Moreover, the
3.2.1. Effects of pH calculated qe values for both As(V) and Cr(VI) adsorption using
One of the factors that significantly influence the adsorption pseudo-second-order rate equation also were in good agreement
process is the pH of the solution. In the present work, the effect with the experimental values. Therefore the adsorption reaction
of pH has been examined by varying the pH of the solution in
the range 3.0–9.0 using an initial concentration of 100 mg/L of both
As(V) and Cr(VI) (Fig. 3). As seen from the figure, the adsorption of (a) 40
both ions is optimal at lower pH range, i.e., pH 3–4, while the
adsorption of As(V) and Cr(VI) decreases on further increase in
the pH of the solution. The maximum removal percentage of 30
As(V) and Cr(VI) ions for this adsorbent is 97% and 67.8% respec-
qe (mg/g)

tively, at pH  3 (Fig. S5, Supplementary data). However, at higher


20
pH (pH  9) this percentage was reduced drastically for As (7.9%)
and significantly for Cr (41.3%). The effect of pH on adsorption of
both metals by this adsorbent can be rationalized as below. It 10 As(V)
has been reported that in the pH range 3–6, As(V) ions exist mainly
2
in the form of H2 AsO 4 while a divalent anion HAsO4 dominates at
Cr(VI)
higher pH values (8–10.5) and in the intermediate pH region 6–8, 0
both species co-exist [22]. On the other hand, Cr(VI) ions exist 0 50 100 150 200 250
2
mainly in the form of HCrO 4 /Cr2 O7 in the pH range 3–6, while t (min)
2
CrO4 is the dominant species at higher pH range [48]. As the pHzpc
of PPhSi-MNPs is 6.7 (Fig. 2); these nanoadsorbents are positively (b) 14
charged below pH 6.2, they can effectively adsorb both H2 AsO 4
and HCrO 4 ions in the pH range 3–6 through electrostatic attrac-
12
tions and ion-exchange mechanism as shown in Fig. 6. However, 10
as the solution pH increases beyond 6.2 positive surface charges
of PPhSi-MNPs diminish and become progressively negative, so 8
t/qe

the repulsive force between the divalent metal anions and the
6
absorbent surface become dominant resulting in reduced adsorp-
tion of metal anions. In addition, excess AOH ions present in the 4
As(V)
solution at high pH, may compete with the metal anions for the io-
2
dide of the anion-exchange sites of the adsorbent. Further discus- Cr(VI)
sion on adsorption mechanism can be found in Section 3.2.4. It 0
can also be seen that the effect of solution pH on the adsorption 0 50 100 150 200 250
for As(V) is much larger than for Cr(VI). The factor affecting this t (min)
variation of adsorptive capacities towards Cr(VI) and As(V) in dif-
ferent pH may be the adsorption free energy of various chromium Fig. 4. (a) Effect of contact time on the adsorption of As(V) and Cr(VI) by PPhSi-
2 MNPs (pH: 3, temperature: 25 °C, concentration: 100 mg/mL); (b) Linear plot of
species (HCrO 4 , Cr2 O7 and CrO2 
4 ) and arsenic species (H2 AsO4 ,
2 pseudo-second-order kinetic model for the adsorption of As(V) and Cr(VI).
HAsO4 ) existing at different pH [7]. At pH > 6, very less adsorp-
tion of As(V) occurs compared to Cr(VI) ions which may be due
to the fact that As(V) ions has less favorable interactions with Table 1
the adsorbent surface compared to Cr(VI) ions at higher pH. Kinetic parameters for the adsorption of As(V) and Cr(VI) ions onto PPhSi-MNPs.

Metal ions qe,exp (mg/g) Pseudo-second order


3.2.2. Effects of contact time and adsorption kinetics
k2 (g/(mg min)) qe,cal, (mg/g) R2
Fig. 4a shows the effects of contact time on adsorption of As(V)
As(V) 34.5 0.0043 35.59 0.999
and Cr(VI) onto PPhSi-MNPs at an initial concentration of 100 mg/
Cr(VI) 19.0 0.0082 18.90 0.999
mL and at pH 3, 25 °C. It is found that the adsorption rate was fast
612 Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615

belongs to the pseudo second-order kinetic model regime and the


overall adsorption process involves ion exchange step [50,51].
(a) PPhSi-MNPs
O1s
PPhSi-MNPs adsorbed As(V) Fe2p
PPhSi-MNPs adsorbed Cr(VI)
3.2.3. Equilibrium studies of As(V) and Cr(VI)

Intensity (a.u.)
Cr2p
The equilibrium isotherms for the adsorption of As(V) and C1s
Fe3p
Cr(VI) ions by bare MNPs and PPhSi MNPs at 25 °C are presented Si2p
in Fig. 5. It can be seen that the adsorption capacity increases with
increase in equilibrium metal ion concentration for both metal ions As3p As(A) As(A)
As3d
in the solution. The equilibrium data are fitted by Langmuir and
Freundlich adsorption isotherm models. Langmuir and Freundlich
I3d
adsorption isotherms which can be expressed in Eqs. (4) and (5),
respectively are widely used to describe the relationship between
the amount of analyte adsorbed on the adsorbent and its equilib- 0 100 200 300 400 500 600 700 800
rium concentration in aqueous solution [52,53]. Binding energy (eV)
Ce Ce 1
¼ þ ð4Þ (b)
qe qm qm K L
As3d
ln qe ¼ ð1=nÞ ln C e þ ln K F ð5Þ

Intensity (a.u.)
where qe is the amount of adsorbate adsorbed per mass of adsor-
bent at equilibrium (mg/g), Ce is the equilibrium concentration of
adsorbate in aqueous solution (mg/L), qm is the monolayer adsorp-
tion capacity at equilibrium (mg/g), KL is the Langmuir equilibrium
constant, KF is a Freundlich constant (index of adsorption capacity),
n is Freundlich constant (index of adsorption intensity or surface
heterogeneity).
The values of qm and KL are determined from the slope and 40 41 42 43 44 45 46 47 48 49 50 51
intercept of the linear plots of Ce/qe versus Ce and the values of KF
Binding energy (eV)
and 1/n are determined from the slope and intercept of the linear
plot of ln qe versus ln Ce (figures are not shown). The isotherm
parameters and related correlation coefficients (R2) are shown in (c)
Table 2. The adsorption isotherm data of both metal anions on Cr 2p
PPhSi-MNPs and bare MNPs are better fitted to Langmuir isotherm
Intensity (a.u.)

model (R2 > 0.99) compared to Freundlich model. The Freundlich Cr(VI)
Cr(III)

(a) 60 As(V)
50

40
qe (mg/g )

571 574 577 580 583 586 589 592


30
Binding energy (eV)
20
Coated MNPs Fig. 6. (a) Full-range XPS spectra of PPhSi-MNPs and after As(V) and Cr(VI)
10 Bare MNPs adsorption; (b) XPS As 3d spectrum of PPhSi-MNPs after As(V) adsorption; and (c)
Fitting by Langmuir model XPS Cr 2p3/2 spectrum of PPhSi-MNPs after Cr(VI) adsorption.
0
0 50 100 150 200
Ce (mg/L) constant n is found to be greater than 1 which indicates the favor-
able condition for adsorption. Based on Langmuir isotherms, the
(b) 40 maximum uptake (qm) for As(V) and Cr(VI) using bare MNPs were
Cr(VI) 25.1 and 14.9 mg/g. Whereas, the maximum adsorption capacities
of PPhSi-MNPs toward As(V) and Cr(VI) were 50.5 and 35.2 mg/g,
30 respectively at 25 °C which are significantly greater than those
qe (mg/g)

using bare MNPs. These results indicate that PPhSi anchored on


20 the surface of magnetic particles has greater As(V) and Cr(VI)
bonding capacity relative to that of pure Fe3O4. The adsorption
capacities of the PPhSi-MNPs for As(V) and Cr(VI) are comparable
10 to those of several other types of magnetic adsorbents reported
Coated MNPs
Bare MNPs in literature as shown in Table 3.
Fitting by Langmuir model
0
0 50 100 150 200
Ce (mg/L) 3.2.4. Adsorption mechanism
To investigate the adsorption mechanism for heavy metal ions,
Fig. 5. Adsorption isotherms for As(V) and Cr(VI) adsorbed onto bare MNPs and As(V)- and Cr(VI)-laden PPhSi-MNPs were prepared. As discussed
PPhSi-MNPs at pH 3, 25 °C. before, As(V) and Cr(VI) ions exist predominately as H2 AsO
4 and
Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615 613

Table 2
Adsorption isotherm parameters for As(V) and Cr(VI) onto bare MNPs and PPhSi-MNPs at pH 3, 25 °C.

Metal anions Adsorbents Langmuir isotherm Freundlich isotherm


2
KL (L/mg) qm (mg/g) R KF n R2
As(V) PPhSi-MNPs 0.239 50.50 0.995 9.49 2.48 0.970
Bare MNPs 0.822 25.13 0.999 11.52 5.29 0.920
Cr(VI) PPhSi-MNPs 0.100 35.21 0.988 5.66 2.49 0.992
Bare MNPs 0.046 14.93 0.995 5.19 2.61 0.959

Table 3
Comparison of maximum adsorption capacities of PPhSi-MNPs with those of some other adsorbents reported in literature for As(V) and Cr(VI) adsorption.

Adsorbents Initial pH qm (mg/g) Refs.


As(V)
CTAB modified magnetic nanoparticles (Fe3O4@CTAB) 6.0 23.07 [32]
Magnetite-reduced graphene oxide (MARGO) composites 7.0 5.83 [33]
Activated carbon 2.2 39.84 [54]
Flower like a-Fe2O3 3.0 51.0 [29]
Nanoscale zero-valent iron 3.0 3.50 [4]
PPhSi-MNPs 3.0 50.50 This study
Cr(VI)
Flower like a-Fe2O3 3.0 30.0 [29]
Maghemite nanoparticles 2.5 17.43 [30]
Iron oxide/zeolite Y nanocomposites coated with APTES 2.0 37.44 [20]
d-FeOOH-coated c-Fe2O3 nanoparticles 2.5 25.83 [23]
PPhSi-MNPs 3.0 35.20 This study

2
HCrO 4 /Cr2 O7 ions in aqueous solution with pH  3, and the sur- pH accelerates the redox reaction in aqueous and solid phases,
face of PPhSi-MNPs is positively charged. Therefore, the electro- since the protons participate in this reaction [57]. Some portion
static attraction between positively charged PPhSi-MNP surface of Cr(III) ions may release into the aqueous phase from the surface
and negatively charged As(V) or Cr(VI) species presumably played due to electronic repulsion between the positively-charged groups
the initial driving force to bind the anions onto the surface. The and Cr(III) ions. Some other possible mechanisms involved in
wide scan XPS spectra of PPhSi-MNPs after As(V) and Cr(VI) Cr(VI) reduction on a bioadsorbent surface was proposed by Park
adsorption as shown in Fig. 6a confirms the presence of As and et al. [55]. However, the mechanism of such Cr(VI) reduction on
Cr elements on the adsorbent. The high resolution XPS As 3d spec- our as-developed nanoparticle surface was not investigated to
trum after As(V) adsorption showed a peak located at 45.2 eV any details in this study.
(Fig. 6b), which could be attributed to As(V)AP bonding [29]. The Interestingly, no iodine was detected in the XPS spectra after
XPS spectrum of the Cr 2p after the adsorption of Cr(VI) on the sur- adsorption of both metal ions onto PPhSi-MNPs suggesting that
face of PPhSi-MNPs is also shown in Fig. 6c; Cr 2p1/2 and Cr 2p3/2 the adsorption of As(V) and Cr(VI) ions was due to the anion
line peaks are located at 586.9 and 577.2 eV, respectively [50]. exchange property of triphenylphosphine by replacing the doped
The broad peak of Cr 2p3/2 could be fitted to two peaks at binding iodide (I) ions as shown in Scheme 2. Further support for ion
energies of 578.1 and 576.3 eV, which are the characteristic signa- exchange came from energy dispersive X-ray (EDX) analysis
tures of Cr(VI) and Cr(III), respectively. This suggests that both (Fig. S6, Supplementary data), where the anions (I) were also
Cr(VI) and Cr(III) coexist on the surface of Cr(VI)-adsorbed PPhSi- not detected in case with metal-loaded PPhSi-MNPs. Thus, both
MNPs and Cr(VI) anions were reduced to Cr(III) ions to a certain XPS and EDX results indicate that As(V) and Cr(VI) ions were ad-
extent during the sorption process. There is a possibility that after sorbed onto the surface of PPhSi-MNPs through ion-exchange
binding to the positively charged groups present on the solid sur- mechanism.
face, Cr(VI) anions can be reduced to Cr(III) due to adjacent elec-
tron donor groups and also because of high redox potential value 3.2.5. Effect of coexisting ions
(>+1.3 V at standard condition) of Cr(VI) [55]. The process of solid There are various kinds of other anions in the groundwater,
redox reaction might occur due to the presence of electron rich such as chloride, sulfate and nitrate which may lead to competitive
divalent iron in Fe3O4 nanocomposites [56]. Moreover, the acidic adsorption. The influence of such anions (Cl, NO 2
3 , SO4 and

M-

O - I- O M-
I- P+
Fe3O4 O Si Fe3O4 O Si P+
O Ion exchange O

M- = H2AsO4- or HCrO4-/Cr2O72-

Scheme 2. Proposed reaction mechanism of adsorption of As(V) and Cr(VI) ions onto PPhSi-MNPs.
614 Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615

As(V) capacities in each cycle is presented in Fig. S7 (Supplementary


100
data). It can be seen that the regenerated adsorbents can be suc-
Cr(VI)
cessfully reused for up to 2 and 3 adsorption–desorption cycles
80
for Cr(VI) and As(V) removal, respectively with negligible loss of
% Removal

its original adsorption capacities.


60

40 4. Conclusions

20 In summary, phosphonium silane coated magnetic nanoparti-


cles (PPhSi-MNPs) were synthesized by a simple one-step co-pre-
0 cipitation process, and they demonstrated a fast and efficient
removal of both As(V) and Cr(VI) species. Various factors affecting
the metal uptake behavior such as pH, contact time, initial concen-
tration of the metal ions and co-existing ions were investigated.
Fig. 7. Effects of coexisting ions on the adsorption of As(V) and Cr(VI) onto The solution pH plays a very important role upon the adsorption
PPhSi-MNPs. of both As(V) and Cr(VI) from an aqueous solution on PPhSi-MNPs
and the optimal adsorption occurred at pH 3.0. The adsorption
equilibrium data for both anions were well fitted by the Langmuir
Table 4
isotherm model. The maximum adsorbed quantities for As(V) and
Desorption efficiencies of As(V) and Cr(VI) using different buffer eluents.
Cr(VI) were found to be 50.5 mg/g and 35.2 mg/g respectively.
Desorption eluent (0.1 M) % Desorbed The modification of Fe3O4 MNPs by phosphonium silane greatly
As(V) Cr(VI) enhanced the adsorption capacities of both metal anions. The
NaOH 89.4 51.6 kinetic data were closely fitted to the pseudo-second-order model.
Na2HPO4 38.5 82.7 From the mechanistic point of view, a synergy of electrostatic
NaHCO3 39.9 87.7 interaction and ion-exchange between positive ligand and negative
Na2CO3 64.0 75.7 pollutant anions is playing a significant role in enabling high
adsorption. Gratifyingly, PPhSi-MNPs nanocomposites could be
easily magnetically retrieved, regenerated and then reused for
two consecutive adsorption–desorption cycles without any appre-
H2 PO4 , 20 mg/L for each anion) on the overall adsorption of As(V) ciable loss of its original adsorption capacity. Such cationically
and Cr(VI) (100 mg/L of each metal) was studied. These anions may modified nanoadsorbents can be a potential candidate for removal
compete with As(V) or Cr(VI) anions for the positive phosphonium of other anionic pollutants from water.
centres. From Fig. 7 it is evident that removal efficiency of this
2
nanoadsorbent was unaffected by the presence of Cl, NO 3 , SO4
ions, which may be due to the fact that the interaction of Cl, Acknowledgments
2
NO3 and SO4 ions with the adsorption sites of the adsorbent is
weaker than Cr(VI) or As(V) ion species. However, H2 PO AZMB and ZBZS acknowledge the Research Scholarship of
4 ions dra-
matically affected the adsorption of both metals which may be National University of Singapore. This work was financially sup-
attributed to their similar ionic structures and the inhibition effect ported by the National University of Singapore Research Fund
of H2 PO (C279-000-003-001).
4 ions on adsorption can be rationalized by preferential
adsorption of phosphates by phosphonium sites (electrostatic
and/or ion-exchange). Significant reduction of arsenic and chro- Appendix A. Supplementary material
mium removal from aqueous solutions in the presence of phos-
phate has also been previously reported in literature for other Supplementary data associated with this article can be found, in
types of adsorbents [22,32,51]. the online version, at http://dx.doi.org/10.1016/j.cej.2013.03.114.

3.2.6. Desorption References


The regeneration ability of the spent adsorbent material is an
important criterion when it comes to cost effectiveness. As it has [1] V.K. Sharma, M. Sohn, Aquatic arsenic: toxicity, speciation, transformations,
been seen that the adsorption of both As(V) and Cr(VI) ions on and remediation, Environ. Int. 35 (4) (2009) 743–759.
[2] P.L. Smedley, D.G. Kinnuburg, A review of the source, behaviour and
PPhSi-MNPs is highly pH dependent; therefore we assumed that distribution of arsenic in natural waters, Appl. Geochem. 17 (2002) 517–568.
desorption of these ions may be possible by varying the pH of [3] D. Mohan, C.U. Pittman, Arsenic removal from water/wastewater using
the system. Desorption of adsorbed As(V) and Cr(VI) from the sur- adsorbents – a critical review, J. Hazard. Mater. 142 (1–2) (2007) 1–53.
[4] S.R. Kanel, B. Manning, L. Charlet, H. Choi, Removal of arsenic(III) from
face of PPhSi-MNPs was carried out using four different desorption groundwater by nanoscale zero-valent iron, Environ. Sci. Technol. 39 (5)
eluents: 0.1 M NaOH, Na2HPO4, NaHCO3 and Na2CO3 solutions, and (2005) 1291–1298.
the results for desorption efficiencies are presented in Table 4. As [5] S. Kapaj, H. Peterson, K. Liber, P. Bhattacharya, Human health effects from
chronic arsenic poisoning – a review, J. Environ. Science Health, Part A 41 (10)
seen from the table, the maximum efficiency for As(V) desorption, (2006) 2399–2428.
89.4% was achieved using NaOH solution. However, the desorption [6] R. Rakhunde, L. Deshpande, H.D. Juneja, Chemical speciation of chromium in
efficiency of Cr(VI) using NaOH was low (51.6%), this is due to the water: a review, Critical Rev. Environ. Sci. Technol. 42 (7) (2012) 776–810.
[7] J. Hu, G. Chen, I.M. Lo, Removal and recovery of Cr(VI) from wastewater by
reduction of adsorbed Cr(VI) to the Cr(III) species (as observed
maghemite nanoparticles, Water Res. 39 (18) (2005) 4528–4536.
from XPS analyses) which could not be desorbed upon treatment [8] P. Klaus, J. Amita, H. Richard, Arsenite and arsenate adsorption on ferrihydrite:
with NaOH solution. All other desorption eluents show reasonably kinetics, equilibrium, and adsorption envelopes, Environ. Sci. Technol. 32
higher efficiencies for Cr(VI) desorption. The reusability of PPhSi- (1998) 344–349.
[9] E. Malkoc, Y. Nuhoglu, M. Dundar, Adsorption of chromium(VI) on pomace-an
MNPs was examined by conducting the adsorption–desorption olive oil industry waste: batch and column studies, J. Hazard. Mater. 138 (1)
process for five cycles for As(V) and Cr(VI) ions and the adsorption (2006) 142–151.
Abu Zayed Md. Badruddoza et al. / Chemical Engineering Journal 225 (2013) 607–615 615

[10] D. Mohan, C.U. Pittman Jr., Activated carbons and low cost adsorbents for [34] M. Bhaumik, A. Maity, V. Srinivasu, M.S. Onyango, Enhanced removal of Cr(VI)
remediation of tri- and hexavalent chromium from water, J. Hazard. Mater. from aqueous solution using polypyrrole/Fe3O4 magnetic nanocomposite, J.
137 (2) (2006) 762–811. Hazard. Mater. 190 (1–3) (2011) 381–390.
[11] A.Z.M. Badruddoza, A. Tay, P. Tan, K. Hidajat, M.S. Uddin, Carboxymethyl-b- [35] S. Sayin, F. Ozcan, M. Yilmaz, A. Tor, S. Memon, Y. Cengeloglu, et al., Synthesis
cyclodextrin conjugated magnetic nanoparticles as nano-adsorbents for of calix[4]arene-grafted magnetite nanoparticles and evaluation of their
removal of copper ions: synthesis and adsorption studies, J. Hazard. Mater. arsenate as well as dichromate removal efficiency, Clean Soil, Air, Water 38
185 (2–3) (2011) 1177–1186. (7) (2010) 639–648.
[12] A.Z.M. Badruddoza, Z.B.Z. Shawon, W.J.D. Tay, K. Hidajat, M.S. Uddin, Fe3O4/ [36] A.P. de los Rios, F.J. Hernandez-Fernandez, L.J. Lozano, S. Sanchez, J.I. Moreno,
cyclodextrin polymer nanocomposites for selective heavy metals removal C. Godinez, et al., Removal of metal ions from aqueous solutions by extraction
from industrial wastewater, Carbohydr. Polym. 91 (1) (2013) 322–332. with ionic liquids, J. Chem. Eng. Data 55 (2) (2010) 605–608.
[13] Y. Li, B. Gao, T. Wu, D. Sun, X. Li, B. Wang, et al., Hexavalent chromium removal [37] J.M. Reyna-Gonzalez, A.A.J. Torriero, A.I. Siriwardana, I.M. Burgar, A.M. Bond,
from aqueous solution by adsorption on aluminum magnesium mixed Extraction of copper(II) ions from aqueous solutions with a methimazole-
hydroxide, Water Res. 43 (12) (2009) 3067–3075. based ionic liquid, Anal. Chem. 82 (18) (2010) 7691–7698.
[14] B. Manna, U.C. Ghosh, Adsorption of arsenic from aqueous solution on [38] G.T. Wei, Z. Yang, C.J. Chen, Room temperature ionic liquid as a novel medium
synthetic hydrous stannic oxide, J. Hazard. Mater. 144 (1–2) (2007) 522–531. for liquid/liquid extraction of metal ions, Anal. Chim. Acta 488 (2) (2003) 183–
[15] D. Duranoğlu, A.W. Trochimczuk, U. Beker, Kinetics and thermodynamics of 192.
hexavalent chromium adsorption onto activated carbon derived from [39] M.E. Mahmoud, H.M. Al-Bishri, Supported hydrophobic ionic liquid on nano-
acrylonitrile-divinylbenzene copolymer, Chem. Eng. J. 187 (2012) 193–202. silica for adsorption of lead, Chem. Eng. J. 166 (1) (2011) 157–167.
_
[16] D. Duranoğlu, I.G. Buyruklardan Kaya, U. Beker, B.F. Sßenkal, Synthesis and [40] A. Pourjavadi, S.H. Hosseini, M. Doulabi, S.M. Fakoorpoor, F. Seidi, Multi-layer
adsorption properties of polymeric and polymer-based hybrid adsorbent for functionalized poly(ionic liquid) coated magnetic nanoparticles: highly
hexavalent chromium removal, Chem. Eng. J. 181–182 (2012) 103–112. recoverable and magnetically separable brønsted acid catalyst, ACS Catal. 2
[17] B. Doušová, T. Grygar, A. Martaus, L. Fuitová, D. Koloušek, V. Machovič, et al., (6) (2012) 1259–1266.
Sorption of As(V) on aluminosilicates treated with Fe(II) nanoparticles, J. [41] H. Qiu, S. Jiang, X. Liu, N-methylimidazolium anion-exchange stationary phase
Colloid Interface Sci. 302 (2) (2006) 424–431. for high-performance liquid chromatography, J. Chromatogr., A 1103 (2)
[18] V. Lenoble, O. Bouras, V. Deluchat, B. Serpaud, J.C. Bollinger, Arsenic adsorption (2006) 265–270.
onto pillared clays and iron oxides, J. Colloid Interface Sci. 255 (1) (2002) 52– [42] L. Fischer, T. Falta, G. Koellensperger, A. Stojanovic, D. Kogelnig, M. Galanski,
58. et al., Ionic liquids for extraction of metals and metal containing compounds
[19] Y. Zhang, Y. Li, J. Li, G. Sheng, Y. Zhang, X. Zheng, et al., Enhanced Cr(VI) from communal and industrial waste water, Water Res. 45 (15) (2011) 4601–
removal by using the mixture of pillared bentonite and zero-valent iron, Chem. 4614.
Eng. J. 185–186 (2012) 243–249. [43] S.T.M. Vidal, M.J. Neiva Correia, M.M. Marques, M.R. Ismael, M.T. Angelino Reis,
[20] K. Barquist, S.C. Larsen, Chromate adsorption on bifunctional, magnetic zeolite Studies on the use of ionic liquids as potential extractants of phenolic
composites, Microporous Mesoporous Mater. 130 (1–3) (2010) 197–202. compounds and metal ions, Sep. Sci. Technol. 39 (9) (2005) 2155–2169.
[21] P. Chutia, S. Kato, T. Kojima, S. Satokawa, Arsenic adsorption from aqueous [44] U. Domańska, A. Re˛kawek, Extraction of metal ions from aqueous solutions
solution on synthetic zeolites, J. Hazard. Mater. 162 (1) (2009) 440–447. using imidazolium based ionic liquids, J. Sol. Chem. 38 (2009) 739–751.
[22] S.R. Chowdhury, E.K. Yanful, Arsenic and chromium removal by mixed [45] K. Naka, A. Narita, H. Tanaka, Y. Chujo, M. Morita, T. Inubushi, I. Nishimura, J.
magnetite-maghemite nanoparticles and the effect of phosphate on removal, Hiruta, H. Shibayama, M. Koga, S. Ishibashi, J. Seki, S. Kizaka-Kondoh, M.
J. Environ. Manage. 91 (11) (2010) 2238–2247. Hiraoka, Biomedical applications of imidazolium cation-modified iron oxide
[23] J. Hu, I. Lo, G. Chen, Performance and mechanism of chromate (VI) adsorption nanoparticles, Polym. Adv. Technol. 19 (2008) 1421–1429.
by d-FeOOH-coated maghemite (c-Fe2O3) nanoparticles, Sep. Purif. Technol. 58 [46] J.F. Liu, Z.S. Zhao, G.B. Jiang, Coating Fe3O4 magnetic nanoparticles with humic
(1) (2007) 76–82. acid for high efficient removal of heavy metals in water, Environ. Sci. Technol.
[24] J. Hu, I.M.C. Lo, G. Chen, Fast removal and recovery of Cr(VI) using surface- 42 (18) (2008) 6949–6954.
modified jacobsite (MnFe2O4) nanoparticles, Langmuir 21 (24) (2005) 11173– [47] Y. Jiang, C. Guo, H. Xia, I. Mahmood, C. Liu, H. Liu, Magnetic nanoparticles
11179. supported ionic liquids for lipase immobilization: enzyme activity in
[25] Z.C. Di, J. Ding, X.J. Peng, Y.H. Li, Z.K. Luan, J. Liang, et al., Chromium adsorption catalyzing esterification, J. Mol. Catal. B: Enzym. 58 (2009) 103–109.
by aligned carbon nanotubes supported ceria nanoparticles, Chemosphere 62 [48] M. Dakiky, M. Khamis, A. Manassra, M. Mer’eb, Selective adsorption of
(5) (2006) 861–865. chromium(VI) in industrial wastewater using low-cost abundantly available
[26] M.E. Pena, G.P. Korfiatis, M. Patel, L. Lippincott, X. Meng, Adsorption of As(V) adsorbents, Adv. Environ. Res. 6 (2002) 533.
and As(III) by nanocrystalline titanium dioxide, Water Res. 39 (11) (2005) [49] Y. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
2327–2337. Biochem. 34 (1999) 451–465.
[27] S.R. Kanel, J.M. Grenèche, H. Choi, Arsenic(V) removal from groundwater using [50] Z. Ai, Y. Cheng, L. Zhang, J. Qiu, Efficient removal of Cr(VI) from aqueous
nano scale zero-valent iron as a colloidal reactive barrier material, Environ. Sci. solution with Fe@Fe2O3 core–shell nanowires, Environ. Sci. Technol. 42 (18)
Technol. 40 (6) (2006) 2045–2050. (2008) 6955–6960.
[28] Y. Sharma, V. Srivastava, V. Singh, S. Kaul, C. Weng, Nano-adsorbents for the [51] R. Li, Q. Li, S. Gao, J.K. Shang, Exceptional arsenic adsorption performance of
removal of metallic pollutants from water and wastewater, Environ. Technol. hydrous cerium oxide nanoparticles: part A. Adsorption capacity and
30 (6) (2009) 583–609. mechanism, Chem. Eng. J. 185–186 (2012) 127–135.
[29] C.Y. Cao, J. Qu, W.S. Yan, J.F. Zhu, Z.Y. Wu, W.G. Song, et al., Low-cost synthesis [52] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and
of flowerlike a-Fe2O3 nanostructures for heavy metal ion removal: adsorption platinum, J. Am. Chem. Soc. 40 (1918) 1361–1403.
property and mechanism, Langmuir 28 (9) (2012) 4573–4579. [53] H. Freundlich, Über die adsorption in lösungen, Z. Phys. Chem. 57 (1906) 385–
[30] J. Hu, G. Chen, I.M.C. Lo, Selective removal of heavy metals from industrial 471.
wastewater using maghemite nanoparticle, J. Environ. Eng. 132 (2006) 709– [54] J. Pattanayak, K. Mondal, S. Mathew, S.B. Lalvani, A parametric evaluation of
715. the removal of As(V) and As(III) by carbon-based adsorbents, Carbon 38 (2000)
[31] P. Yuan, D. Liu, M. Fan, D. Yang, R. Zhu, F. Ge, et al., Removal of hexavalent 589–596.
chromium [Cr(VI)] from aqueous solutions by the diatomite-supported/ [55] D. Park, Y.S. Yun, J.M. Park, Studies on hexavalent chromium biosorption by
unsupported magnetite nanoparticles, J. Hazard. Mater. 173 (1–3) (2010) chemically-treated biomass of Ecklonia sp., Chemosphere 60 (10) (2005) 1356–
614–621. 1364.
[32] Y. Jin, F. Liu, M. Tong, Y. Hou, Removal of arsenate by [56] N. Mao, L. Yang, G. Zhao, X. Li, Y. Li, Adsorption performance and mechanism of
cetyltrimethylammonium bromide modified magnetic nanoparticles, J. Cr(VI) using magnetic PS-EDTA resin from micro-polluted waters, Chem. Eng. J.
Hazard. Mater. 227–228 (2012) 461–468. 200–202 (2012) 480–490.
[33] V. Chandra, J. Park, Y. Chun, J.W. Lee, I.C. Hwang, K.S. Kim, et al., Water- [57] S.K. Prabhakaran, K. Vijayaraghavan, R. Balasubramanian, Removal of Cr(VI)
dispersible magnetite-reduced graphene oxide composites for arsenic ions by spent tea and coffee dusts: reduction to Cr(III) and biosorption, Ind.
removal, ACS Nano 4 (7) (2010) 3979–3986. Eng. Chem. Res. 48 (2009) 2113–2117.

You might also like