You are on page 1of 11

Chemical Engineering Journal 327 (2017) 286–296

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Selective separation of rare earth ions from aqueous solution using


functionalized magnetite nanoparticles: kinetic and thermodynamic
studies
Radwa M. Ashour a,b, Ramy El-sayed c, Ahmed F. Abdel-Magied b, Ahmed A. Abdel-khalek d, M.M. Ali b,
Kerstin Forsberg e, A. Uheida a, Mamoun Muhammed f, Joydeep Dutta a,⇑
a
Functional Materials, Applied Physics Department, SCI School, Isafjordsgatan 22, SE-164 40 Kista Stockholm, Sweden
b
Nuclear Materials Authority, P.O. Box 530, El Maadi, Cairo, Egypt
c
Experimental Cancer Medicine, KFC, Novum, Department of Laboratory Medicine, Karolinska Institute, 141 86 Stockholm, Sweden
d
Department of Chemistry, Faculty of Science, Beni-Suef University, Beni-Suef, Egypt
e
Department of Chemical Engineering, KTH Royal Institute of Technology, 100 44 Stockholm, Sweden
f
Institute of Graduate Studies and Research, Alexandria University, Alexandria, Egypt

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

 Functionalized Fe3O4 NPs with citric


acid and L-cysteine were prepared.
3+ 3+
 Selective separation of La , Nd ,
Gd3+ and Y3+ by the functionalized
Fe3O4 NPs.
3+
 Cys@Fe3O4 NPs showed high La ,
Nd3+, Gd3+ and Y3+ adsorption
capacity.
 The adsorption process followed a
pseudo-second order rate law.

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

Article history: Separation of rare earth ions (RE3+) from aqueous solution is a tricky problem due to their physico-
Received 11 May 2017 chemical similarities of properties. In this study, we investigate the influence of the functionalized
Received in revised form 18 June 2017 ligands on the adsorption efficiency and selective adsorption of La3+, Nd3+, Gd3+ and Y3+ from aqueous
Accepted 19 June 2017
solution using Magnetite (Fe3O4) nanoparticles (NPs) functionalized with citric acid (CA@Fe3O4 NPs) or
Available online 20 June 2017
L-cysteine (Cys@Fe3O4 NPs). The microstructure, thermal behavior and surface functionalization of the
synthesized nanoparticles were studied. The general adsorption capacity of Cys@Fe3O4 NPs was found
Keywords:
to be high (98 mg g1) in comparison to CA@Fe3O4 NPs (52 mg g1) at neutral pH 7.0. The adsorption
Magnetic nanoparticles
Rare earths
kinetic studies revealed that the adsorption of RE3+ ions follows a pseudo second-order model and the
Functionalization adsorption equilibrium data fits well to the Langmuir isotherm. Thermodynamic studies imply that the
Citric acid adsorption process was endothermic and spontaneous in nature. Controlled desorption within 30 min
L-cysteine of the adsorbed RE3+ ions from both Cys@Fe3O4 NPs and CA@Fe3O4 NPs was achieved with 0.5 M
Adsorption HNO3. Furthermore, Cys@Fe3O4 NPs exhibited a higher separation factor (SF) in the separation of Gd3+/
La3+, Gd3+/Nd3+, Gd3+/Y3+ ions compared to CA@Fe3O4 NPs.
Ó 2017 Elsevier B.V. All rights reserved.

⇑ Corresponding author.
E-mail address: joydeep@kth.se (J. Dutta).

http://dx.doi.org/10.1016/j.cej.2017.06.101
1385-8947/Ó 2017 Elsevier B.V. All rights reserved.
R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296 287

1. Introduction Herein, we report the synthesis and characterization of mag-


netic nanoparticles functionalized with citric acid (CA@Fe3O4
Rare earth elements (REEs) have received attention due to their NPs) and L-cysteine (Cys@Fe3O4 NPs) ligands, respectively. We
use in different applications such as in superconductors, telecom- investigate the influence of CA@Fe3O4 NPs and Cys@Fe3O4 NPs on
munications, nuclear and solar energy conversion, photocatalysis, the selective adsorption of RE3+ (= La3+, Nd3+, Gd3+ and Y3+) and
and biosciences [1,2]. In recent years, RREs separation chemistry the adsorption mechanism from aqueous solution. The dependence
has gained a great attention not only for their industrial applica- of adsorption parameters such as contact time, pH, temperature
tions, but also their environmental mitigation. Gadolinium (Gd) and initial concentration of metal ions on the adsorption efficiency
is useful in ceramic industries, metallurgy, nuclear techniques in were studied. In addition, we report the desorption characteristics
fuel element fabrications and heavy water reactors as neutron poi- of REEs from the loaded CA@Fe3O4 NPs and Cys@Fe3O4 NPs using
son. Aqueous Gd3+ ion, even at trace level, considered as hazardous different eluents such as NaOH and HNO3. Adsorption kinetics
material [3,4]. In view of all the above-mentioned uses and its high and adsorption isotherms were analyzed using a non-linear
toxicity, selective separation of Gd3+ from aqueous solutions is of method and the thermodynamic parameters (DG°, DH° and DS°)
extreme importance. Due to the similar chemical and physical were derived from the obtained experimental results.
properties of REEs, separating them from a mixture is still a chal-
lenging task [5,6]. The most commonly used techniques for the 2. Materials and methods
separation and recovery of REEs are precipitation [7], liquid-
liquid extraction [8], ion exchange [9] and adsorption [10]. 2.1. Chemicals
Amongst these, solid-phase extraction (SPE) is proven to be an
effective and convenient method for REEs recovery. SPE has been Ferric chloride hexahydrate (FeCl36H2O,  99%), ferrous chlo-
proven to be an easy operation offering high adsorption capacities, ride tetrahydrate (FeCl24H2O,  98%), ammonium hydroxide
rapid phase separation and economical cost effectiveness [11]. Var- (25%), citric acid monohydrate (98%), L-cysteine, nitric acid (65%),
ious types of adsorbents (organic, inorganic, carbon based material sodium hydroxide (98%), lanthanum(III) nitrate hexahydrate
and bio-sorbents) have been developed for SPE recovery of REEs (99.9%) [La(NO3)36H2O], neodymium(III) nitrate hexahydrate
[12–17]. Magnetic nanoparticles have been found increased atten- (99.9%) [Nd(NO3)36H2O], gadolinium(III) nitrate hexahydrate
tion as nanoadsorbents for environmental decontamination due to (99.9%) [Gd(NO3)36H2O], yttrium(III) nitrate hexahydrate (99.9%)
the possibilities of using external magnetic field to guide the pro- [Y(NO3)36H2O] and standard solutions (1000 mg/L) of Mg(NO3)2,
cess of separation [18]. Furthermore, upon desorption of the Ca(NO3)2and Ni(NO3)2 were purchased from Sigma Aldrich. All
adsorbed ions, the magnetic nanoparticles can be reused, making chemicals were of analytical grade reagents and used as received
them promising sustainable adsorbents. Functionalized magnetic without further purification. Deionized water with a resistivity of
nanoparticles using inorganic materials have been demonstrated 18 MX cm was used in all the experiments.
to be excellent candidates for selective adsorption of RE3+ ions
and metal traces from aqueous solution [19–23]. Dupont et al.
reported the functionalization of magnetite (Fe3O4) and nonmag- 2.2. Characterization
netic (silica and titanium dioxide) nanoparticles with N-[(3-trime
thoxysilyl)propyl]ethylenediamine triacetic acid (TMS-EDTA) and High resolution transmission electron microscopy (HRTEM)
their behavior towards the adsorption and separation of RE3+ ions (HR-FEG-TEM, JEOL JEM-2100, Tokyo, Japan) was used for the char-
in aqueous media [24]. Selective extraction of heavy samarium- acterization of the morphology and size distribution of the synthe-
holmium (Sm–Ho) and Light lanthanum-niobium (La–Nd) lan- sized nanoparticles. For structural evaluation X-Ray Powder
thanides from aqueous solutions using diethylenetriamine- Diffraction analysis (XRPD, PANalytical Empyrean) was carried
pentaacetic acid (DTPA) functionalized Fe3O4 NPs, achieving high out. Fourier transform infra–red (FT–IR) spectrometer (Nicolet
SF between heavy-lanthanides and light-lanthanides reaching the Instruments model Avatar-100 equipped with diamond ATR, Madi-
maximal value of 11.5 at low pH (2.0) in 30 min was reported by son, WI, USA) was used to study characteristic functional groups on
Qiang et al. recently [25]. the nanoparticle surfaces in the range of 500–4000 cm1. Specific
The choice of appropriate functional chelating groups is a criti- surface area calculations were determined by the Brunauer-
cal factor to obtain maximum adsorption efficiency and selectivity. Emmett-Teller (BET) method with N2 gas using Micromeritics ASAP
Citric acid and L-cysteine are two major classes of chelating agents 2000 surface area and porosity system (Quantachrome, USA) by
that have been studied as chelators for adsorption of metal ions first degassing the samples at 150 °C for 1 hour. Zeta potential of
[26]. Citric acid contains three carboxyl groups and a hydroxyl synthesized nanoparticles and surface charge of nanoparticles
group, while L-cysteine is constituted of three different functional were determined by (Delsa Nano C, Beckman Coulter, CA, USA).
groups, amine, thiol and carboxylic acid (Fig. 1). Both citric acid The percentage weight loss of CA@Fe3O4 NPs and Cys@Fe3O4 NPs
and L-cysteine ligands lead to the availability of functional groups were studied in the temperature range of 30–600 °C at a heating
on Fe3O4 NPs for controlled adsorption but also affects the size, rate of 10 °C min in nitrogen atmosphere, using Thermogravimetric
morphology and colloidal stability of the nanoparticles [21]. analysis (TGA) (Q5000, TA instruments, DE, USA). The concentra-
tions of metal ions were determined by Inductively Coupled
Plasma-Optical Emission Spectroscopy (ICP-OES) iCAP 6500 ICP
from Thermo Scientific (MA, USA). Adjustments of solutions pH
were measured by a pH-meter (ORION 410).

2.3. Synthesis of precursor Fe3O4 NPs

The detailed synthetic process of Fe3O4 NPs (10 nm) have been
described elsewhere [27]. Briefly, 1.988 g (0.125 mol) of FeCl24H2-
O and 5.406 g (0.25 mol) of FeCl36H2O (1:2 molar ratio) were dis-
Fig. 1. Chemical structure of the functionalization ligands used in this study: (a) solved in 80 mL of water in a round bottomed flask at room
Citric acid and (b) L-cysteine. temperature. Then the temperature was slowly increased to
288 R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296

70 °C in nitrogen atmosphere with continuous mechanical stirring The adsorption kinetic experimental data were fitted to a non-
at 370 rpm for 30 min, then 20 mL of ammonium hydroxide solu- linear form of pseudo-first order and pseudo-second order models
tion (25%) was added to the reaction mixture and the mixture using Eqs. (4) and (5), respectively.
was then kept at 70 °C for another 30 min. The obtained black pre-
cipitate was then washed three times with deionized water and qt ¼ qe ð1  ek 1t Þ ð4Þ
separated from the supernatant using a permanent magnet to
obtain the Fe3O4 NPs (about 1.30 g was obtained typically). k2 q2e t
qt ¼ ð5Þ
1 þ k 2 qe t
2.4. Synthesis of CA@Fe3O4 NPs and Cys@Fe3O4 NPs where, k1 (min1) is the pseudo-first order rate constant of
adsorption, k2 (g mg1 min) is the pseudo-second order rate con-
In order to functionalize the synthesized Fe3O4 NPs with stant of adsorption, qe (mg g1) is the adsorption capacity at equi-
citrates, 4 mL of an aqueous solution of citric acid (0.3 g mL1) librium and qt (mg g1) is the adsorption capacity at a given time t.
was added to the Fe3O4 NPs in deionized water and the tempera-
ture then raised slowly to 90 °C under refluxing conditions with 2.6. Selective adsorption studies
continuous mechanical stirring for 1 h. The obtained precipitate
was rinsed with water and separated by an external permanent Experiments were performed to study the selectivity of
magnet to obtain CA@Fe3O4 NPs. CA@Fe3O4 NPs and Cys@Fe3O4 NPs towards RE3+ adsorption. The
Surface functionalization of Fe3O4 NPs with L-cysteine was car- selective adsorption experiments were carried out by mixing a
ried out by mixing a known quantity of Fe3O4 NPs with 50 mL known amount of CA@Fe3O4 NPs or Cys@Fe3O4 NPs with La3+,
(0.1 M) L-cysteine and the reaction mixture was kept under vigor- Nd3+, Gd3+ and Y3+ solutions containing Ni2+, Ca2+ and Mg2+ at
ous shaking overnight. The product was separated from the solu- 1:1 molar ratio using a shaker at room temperature. The aqueous
tion using magnetic decantation and was washed three times phase was separated from the solid phase by using an external
using deionized water to obtain Cys@Fe3O4 NPs. magnet followed by ultra-centrifuging of the supernatant at
10,000 rpm and the concentration of metal ions in the supernatant
2.5. Adsorption studies was determined by ICP-OES.

Stock solutions of lanthanum(III) nitrate, neodymium(III) 2.7. Desorption studies


nitrate, gadolinium(III) nitrate and yttrium(III) nitrate were pre-
pared by dissolving an appropriate amount of the required salt in To examine the feasibility of recycling the coated nanoparticles,
deionized water. Adsorption experiments were conducted by mix- desorption studies were performed using 0.1–1 mol L1 of eluting
ing 10 mL of RE3+ mixture solution containing lanthanum La3+, solution of NaOH or HNO3 for 30 min. The desorption efficiency
neodymium Nd3+, gadolinium Gd3+ and yttrium Y3+ with 2.5 mg of La3+, Nd3+, Gd3+ and Y3+ from the loaded CA@Fe3O4 NPs or
of magnetic nanoadsorbents (CA@Fe3O4 NPs or Cys@Fe3O4 NPs) Cys@Fe3O4 NPs was calculated using Eq. (6).
and the mixture was subjected to continuous shaking. Batch
adsorption experiments were carried out to investigate RE3+ ions Cdes
Desorption% ¼ x100 ð6Þ
adsorption as a function of pH. The pH was adjusted in the range Cads
3–8 by using nitric acid and sodium hydroxide, respectively. Initial where, Cdes is the amount of La3+, Nd3+, Gd3+ and Y3+ released
RE3+ concentration was varied from 5 to 50 mg L1 to study the into aqueous solution (mg L1) and Cads is the amount of La3+,
adsorption isotherms of RE3+ on surface functionalized Fe3O4 Nd3+, Gd3+ and Y3+ adsorbed on the coated Fe3O4 NPs (mg L1).
NPs. Contact time was changed from 1 to 120 min to study the
adsorption kinetic efficiency. Temperature was varied from 278
3. Results and discussion
to 318 K to study thermodynamic parameters. The magnetic
nanoadsorbents CA@Fe3O4 NPs or Cys@Fe3O4 NPs were separated
3.1. Characterization of CA@Fe3O4 NPs and Cys@Fe3O4 NPs
from the aqueous phase with an external magnet. The adsorption
experiments were performed in duplicate and the average values
The TEM images of CA@Fe3O4 NPs and Cys@Fe3O4 NPs
of total adsorption are reported. The adsorption capacity qe
(Fig. 2a and b, respectively), revealed that Fe3O4 NPs are well sep-
(mg g1) of the adsorbents were determined by Eq. (1).
arated spherical particles. The particle size distribution determined
ðCi  Ce ÞV from TEM micrographs, showed that the particles size of CA@Fe3O4
Adsorption capacityðqe Þ ¼ ð1Þ
m NPs and Cys@Fe3O4 NPs are 12 ± 3 nm and 10 ± 5 nm, respectively
(Fig. 2a (inset) and b (inset)).
where, Ci is the initial RE3+ concentration (mg L1), Ce is the equilib-
The XRD patterns of the synthesized CA@Fe3O4 NPs and
rium RE3+ concentration (mg L1) in aqueous solution, V is the total
Cys@Fe3O4 NPs are presented in Fig. 2c and d, respectively. Five
volume of water (mL) and m is the mass of CA@Fe3O4 NPs or
characteristic peaks for Fe3O4 NPs corresponding to (2 Theta)
Cys@Fe3O4 NPs (mg).
= 30.136, 35.554, 43.2, 53.529, 57.012 and 62.961 for (2 2 0),
The adsorption isotherms were analyzed with the nonlinear
(3 1 1), (4 0 0), (4 2 2), (5 1 1) and (4 4 0) planes were observed. These
equations of Langmuir and Freundlich models using Eqs. (2) and
peaks are well matched with the magnetite characteristic peaks
(3).
confirming the functionalization process did not result in phase
qmax kCe alteration of Fe3O4 NPs [28].
qe ¼ ð2Þ The TGA thermograms of pristine Fe3O4 NPs, CA@Fe3O4 NPs and
1 þ kCe
Cys@Fe3O4 NPs are shown in Fig. 2e and f, respectively. About 1%
weight loss for all NPs are observed at 50–200 °C due to the dehy-
qe ¼ K f C e1=n ð3Þ
dration of NP surface. Above 200 °C, 8.8% weight loss of CA@Fe3O4
1
where, qmax (mg g ) is the maximum adsorption capacity, k NPs and 2.8% for Cys@Fe3O4 NPs were observed due to the removal
(L mg1) is the Langmuir equilibrium constant, Kf (mg1n Ln g1) of the ligands. Weight losses above 450 °C is related to transforma-
is Freundlich constant and 1/n is the heterogeneity factor. tion of Fe3O4–Fe2O3 [28]. The obtained TGA results confirm that the
R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296 289

Fig. 2. Characterization of the functionalized CA@Fe3O4 NPs and Cys@Fe3O4 NPs: TEM image of (a) CA@Fe3O4 NPs and (b) Cys@Fe3O4 NPs with particle size distribution
(inset); XRD pattern of (c) CA@Fe3O4 NPs and (d) Cys@Fe3O4 NPs; TGA curves of (e) CA@Fe3O4 NPs, (f) Cys@Fe3O4 NPs and pristine Fe3O4 NPs; FT-IR analysis of (g) CA, Fe3O4
NPs and CA@Fe3O4 NPs and (h) Cys, Fe3O4 NPs and Cys@Fe3O4 NPs.

surface of Fe3O4 NPs was successfully functionalized with CA and appearance at 1582 cm1. The symmetric stretching of the COO
Cys. group of CA at 1401 cm1 is also slightly shifted to lower frequen-
FT-IR spectrum of CA@Fe3O4 NPs (Fig. 2g) showed that car- cies with a strong but broad band appearing at 1389 cm1, con-
boxylic functional group at 1680 cm1 assigned to C@O vibration firming the successful functionalization of Fe3O4 nanoparticles
of CA was shifted to lower wave number due to the covalent bond- with CA [29]. Similarly, the infrared spectrum of Cys@Fe3O4 NPs
ing on the surface of Fe3O4 nanoparticles as confirmed by the (Fig. 2h), reveal a strong broad peak at 3055 cm1, attributed to
290 R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296

the ANH2 stretches, at 1586 cm1 due to the symmetric stretching determined kinetic data were analyzed using nonlinear pseudo
of the C@O, at 1375 cm-1 due to the symmetric stretching of the first-order and pseudo second-order models. The obtained fit of
COO The band at 2555 cm1 attributed to the –SH group of Cys, the kinetic parameters of CA@Fe3O4 NPs and Cys@Fe3O4 NPs are
disappear in Cys@Fe3O4 NPs implying that the Cys molecules shown in Table 1. The pseudo-second-order model fitted well with
attaches to the surface of the Fe3O4 NPs via the formation of FeAS all experimental data (La3+: R2 = 0.999, Nd3+: R2 = 0.996, Gd3+:
covalent bond [26,28]. R2 = 0.995 and Y3+: R2 = 0.991 for CA@Fe3O4 NPs and La3+:
The BET analysis for Fe3O4 NPs, CA@Fe3O4 NPs and Cys@Fe3O4 R2 = 0.993, Nd3+: R2 = 0.995, Gd3+: R2 = 0.987 and Y3+: R2 = 0.998
NPs are shown in Fig. S1a, b and c (ESIy), respectively. The specific for Cys@Fe3O4 NPs), and also the obtained qe,cal values from the
surface area of Fe3O4 NPs, CA@Fe3O4 NPs and Cys@Fe3O4 NPs were pseudo-second-order model were found to agree well to the exper-
87.95, 94.65 and 53.95 m2 g1, respectively. The magnetic proper- imentally obtained qe,exp. The fitting results thus indicate that the
ties of the magnetic nanoparticles have been reported elsewhere adsorption rate was mainly determined by the chemical adsorp-
[30]. tion process through sharing or exchange of electrons between
The zeta potential of CA@Fe3O4 NPs and Cys@Fe3O4 NPs in the La3+, Nd3+, Gd3+ and Y3+ ions on the CA@Fe3O4 NPs and Cys@Fe3O4
pH range of 2–10 are shown in Fig. 3. In case of CA@Fe3O4 NPs, high NPs [37].
negative charges on the surface is observed and the isoelectric
point (IEP) could not be found in the entire pH range studied, as 3.3. Effect of pH
also observed by others [29]. The amount of the negative charges
increased with increasing pH due to the deprotonation of car- pH factor plays an important role in the adsorption of RE3+ ions
boxylic groups in the case of functionalisation with citrates. The on the surface of the NPs. To investigate the effect of pH on the
obtained results thus confirm that grafting of carboxylic groups adsorption of La3+, Nd3+, Gd3+ and Y3+ ions on CA@Fe3O4 NPs and
to the surface of Fe3O4 NPs has occurred [31], as the IEP of Fe3O4 Cys@Fe3O4 NPs, we prepared a series of mixed sample solutions
NPs is expected to be at pH 6.5 [32]. The zeta potential measure- containing 5 mg L1 of La3+, Nd3+, Gd3+ and Y3+ ions
ments of Cys@Fe3O4 NPs showed that the zeta potential values of (Fig. 4c and d). The pH values of the solutions were adjusted by
Cys varies with pH, similar to other amino acids.[33] The IEP of HNO3 or NaOH solution. This study is restricted up to a maximum
Cys has been reported to be at pH 5.1 [34]. Below the IEP, a positive pH 8.0 since beyond this pH, precipitation of RE3+ ions occur as
charge of Cys due to the protonation of amino and carboxylic observed from the speciation of REEs [38]. At pH > 8, REEs could
groups on the surface of the Fe3O4 NPs, render the overall charge be precipitated as insoluble hydroxides i.e. RE(OH)3(s) [38,39].
positive. Above IEP, the surface is negatively charged due to depro- The obtained results revealed that the adsorption capacity of
tonation of carboxylic and amino groups at higher pH values which CA@Fe3O4 NPs and Cys@Fe3O4 NPs for La3+, Nd3+, Gd3+ and Y3+ ions
causes the surface potential to cross from zero to the negatively increased with increasing the pH, so, pH = 7.0 was selected for sub-
charged region [35]. sequent studies.

3.2. Adsorption kinetics 3.4. Temperature dependence

The kinetic data for the adsorption of La3+, Nd3+, Gd3+ and Y3+ The temperature dependence of La3+, Nd3+, Gd3+ and Y3+ adsorp-
ions on CA@Fe3O4 NPs and Cys@Fe3O4 NPs are shown in tion on the surface of CA@Fe3O4 NPs and Cys@Fe3O4 NPs were
Fig. 4a and b, respectively. It is clear that the adsorption of La3+, studied at 278–318 K. From the data represented in Fig. S2 (ESIy),
Nd3+, Gd3+ and Y3+ ions increase rapidly in the first 5 and 15 min the adsorption capacity of CA@Fe3O4 NPs and Cys@Fe3O4 NPs for
for both the Cys@Fe3O4 NPs and CA@Fe3O4 NPs. and adsorption the adsorption of La3+, Nd3+, Gd3+ and Y3+ ions increased upon
equilibrium was achieved within 15 and 30 min. The rapid adsorp- increasing the temperature due to the higher energy of the system
tion of La3+, Nd3+, Gd3+ and Y3+ ions is attributed to the abundance facilitating the adsorption process, as well as due to the higher
of surface sites on the Fe3O4 surface-functionalized with Cys and affinity of active binding sites on the surface of the coated Fe3O4
CA [36]. NPs [40].
The adsorption kinetics is an important aspect that can provide
useful information of the adsorption mechanisms. Experimentally 3.5. Thermodynamics parameters

The standard enthalpy change (DH°), entropy change (DS°), and


Gibb‘s energy change (DG°) were evaluated by applying Eqs. (7)
and (8) [41].

DH  DS
lnKd ¼  þ ð7Þ
RT R

DG ¼ DH  T DS ð8Þ


1 1
where, R is the gas constant (8.314 J mol K ), T is the absolute
temperature (Kelvin) and the distribution coefficients (Kd, mL g1)
were defined by using Eq. (9) [42].

ðCi  Cf Þ
Kd ðRE3þ Þ ¼  V=m ð9Þ
Cf

where, Kd (RE3+) is the distribution coefficient, Ci = initial concentra-


tion, Cf = final concentration, V = volume of solution, and m = mass
of CA@Fe3O4 NPs or Cys@Fe3O4 NPs, which is defined as the ratio
of the concentration of rare earth ions in the solid phase to that
Fig. 3. Zeta potential of CA@Fe3O4 NPs and Cys@Fe3O4 NPs at different pH. in liquid phase.
R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296 291

Fig. 4. Typical pseudo-second order fitting of the adsorption of La3+, Nd3+, Gd3+ and Y3+ ions using (a) CA@Fe3O4 NPs and (b) Cys@Fe3O4 NPs. Experimental conditions: Ci
3+
(RE) = 5 mg L1, pH = 7.0, m of adsorbents = 2.5 mg and v = 10 mL at room temperature. Effect of pH on adsorption of La3+, Nd3+, Gd3+ and Y3+ ions using (c) CA@Fe3O4 NPs,
1
time = 30 min and (d) Cys@Fe3O4 NPs, time = 15 min. Experimental conditions: C3+)i(RE = 5 mg L , m of adsorbents = 2.5 mg and v = 10 mL at room temperature.

Table 1
Kinetic parameters of La3+, Nd3+, Gd3+ and Y3+ adsorption on CA@Fe3O4 NPs and Cys@Fe3O4 NPs.

Model La3+ Nd3+ Gd3+ Y3+


CA@Fe3O4 NPs
Pseudo-first order qe,cal. (mg g1) 9.1 ± 0.01 12 ± 0.02 16.9 ± 0.02 11.1 ± 0.01
qe,exp. (mg g1) 10 13.3 18 12
k1 (min1) 0.15 ± 0.005 0.09 ± 0.005 0.064 ± 0.003 0.1 ± 0.001
R2 0.961 0.982 0.982 0.984
Pseudo-second order qe,cal (mg g1) 10.1 ± 0.30 14 ± 0.24 19.6 ± 0.41 12.4 ± 0.13
qe,exp. (mg g1) 10 13.3 18 12
k2 (g mg1 min1) 0.02 ± 0.001 0.007 ± 0.005 0.003 ± 0.003 0.01 ± 0.001
R2 0.999 0.996 0.995 0.991
Cys@Fe3O4 NPs
Pseudo-first order qe,cal (mg g1) 18.4 ± 0.08 20.1 ± 0.08 21.3 ± 0.02 20 ± 0.04
qe,exp.(mg g1) 18.5 20.3 21.5 20.1
k1 (min1) 2.5 ± 0.15 3.2 ± 0.13 3.8 ± 0.16 2.3 ± 0.14
R2 0.891 0.863 0.833 0.892
Pseudo-second order qe,cal. (mg g1) 18.5 ± 0.04 20.32 ± 0.01 21.45 ± 0.015 20.1 ± 0.006
qe,exp.(mg g1) 18.5 20.3 21.5 20.1
k2 (g mg1 min1) 0.52 ± 0.03 1.03 ± 0.002 1.8 ± 0.15 0.44 ± 0.008
R2 0.993 0.995 0.987 0.998

The standard enthalpy change (DH°) and the entropy change for complexation and stability of the adsorption of RE3+ ions in both
CA@Fe3O4 NPs and Cys@Fe3O4 NPs can be calculated from the slope the systems studied in this work. In addition, the negative charge
and intercept of the plot of ln Kd vs. 1/T using the van’t Hoff plot as of Gibb’s free energy (DG°) values in the temperature range stud-
shown in Fig. 5a and b, respectively. The obtained values of DG° ied, suggest that the adsorption process is spontaneous and more
(kJ mol1), DH° (kJ mol1) and DS° (J mol1) are summarized in favorable at higher temperatures indicating that the adsorption
Table 2. The values of enthalpy change (DH°) are positive, reactions are primarily driven towards the products.
indicating that the adsorption of La3+, Nd3+, Gd3+ and Y3+ ions are
endo-thermally driven and thus increasing the temperature is 3.6. Adsorption isotherms
advantageous for higher adsorption. The positive values of entropy
change (DS°) obtained in this study indicates an increase in the The effect of the concentration of RE3+ (La3+, Nd3+, Gd3+ and Y3+
degree of freedom at the solid-liquid interface [27]. It favors ions) on the adsorption efficiency were investigated under the
292 R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296

Fig. 5. Linear plot of ln Kd versus 1/T for adsorption capacity of La3+, Nd3+, Gd3+ and Y3+ on (a) CA@Fe3O4 NPs, time = 30 min and (b) Cys@Fe3O4 NPs, time = 15 min.
1
Experimental condition: C3+)i(RE = 5.0 mg L , pH = 7.0, m of adsorbents = 2.5 mg and v = 10 mL.

Table 2
Thermodynamic parameters for the adsorption of La3+, Nd3+, Gd3+ and Y3+ by CA@Fe3O4 NPs and Cys@Fe3O4 NPs at 278–318 K.

Parameters La3+ Nd3+ Gd3+ Y3+


CA@Fe3O4 NPs
DH° (kJ mol1) 41.85 53 35 44
DS° (J mol1) 165 201 135 168
DG°(kJ mol1) (–46 to –52) (–56 to –64) (–38 to–43) (–47 to –53)
R2 0. 995 0. 993 0. 995 0. 999
Cys@Fe3O4 NPs
DH° (kJ mol1) 16.7 14 15.5 11
DS° (J mol1) 87 89 92 65
DG°(kJ mol1) (–24 to –28) (–25 to –28) (–25 to –29) (–18 to –20)
R2 0.996 0. 991 0.997 0. 994

optimized conditions. The results demonstrate that the adsorbed predominated by positively charged species such as RE3+ ions that
amount of RE3+ ions (qe) on CA@Fe3O4 NPs and Cys@Fe3O4 NPs are present in the form of RE3+, RE(OH)+, RE(OH)2 and RE(OH)3 at
increased with increasing initial concentration of RE3+ and reached various pH values. At pH < 8, the main species of RE3+ ions are
a plateau at higher concentrations due to saturation of the number RE3+ and RE(OH)+ and the removal of RE3+ is mainly accomplished
of binding sites in a fixed amount of CA@Fe3O4 NPs and Cys@Fe3O4 by surface complexation. For CA@Fe3O4 NPs, the negatively
NPs. The experimental data was interpreted by non-linear Lang- charged COO groups have strong coordinative affinity towards
muir and Freundlich sorption isotherms (Fig. 6a and b) and the cor- La3+, Nd3+, Gd3+ and Y3+ ions. These carboxylate ions capture the
responding parameters related to Langmuir and Freundlich models RE3+ ions by forming complexes and the ability of surface complex-
are summarized in Table 3. It was found that the equilibrium ation increases with increasing pH of the solution. At higher pH,
adsorption capacity was better described by the Langmuir iso- the chelating complexes of carboxylate and La3+, Nd3+, Gd3+ and
therm model, indicating that the monolayer adsorption process Y3+ ions are expected more than at lower pH because the chelation
can explain the metal adsorption. The obtained qmax values sites occupied by H+ are released at higher pH leading to the
(Table 3) suggest that the adsorption affinity of both systems desired chelation [47].
decreased in the order Gd3+ > Nd3+ > Y3+ > La3+ due to the compet- In case of Cys@Fe3O4 NPs, there are two different surface func-
itive adsorption between La3+, Nd3+, Gd3+ and Y3+ ions on the same tional groups, carboxylate (COO) and amine (ANH2) At low pH,
adsorption site on the surface of the NPs. The relative order is con- the Cys@Fe3O4 NPs surface has a net positive charge due to the pro-
sistent with the chemical properties of RE3+ such as the sequence tonation of amine (ANH+3). RE3+ ions adsorption occurred at pH < 5
of the electronegativity and the size of the metal ions. The maxi- in the range 5–10 mg g1. This could be attributed to the electro-
mum adsorption capacity values of RE3+ ions for Cys@Fe3O4 NPs static repulsion between RE3+ and the surface of Fe3O4 NPs, which
is almost double the values of CA@Fe3O4 NPs, which is attributed increase the conversion of ANH2 groups to ANH+3 and there were
to the two different functional groups (amine and carboxylic few of ANH2OH sites available on the Cys@Fe3O4 NPs surface
groups in Cys) that leads to increased active sites on the Fe3O4 for the adsorption of ions through surface complexation [48]. At
NPs surfaces. higher pH the concentration of H+ increases as observed from the
A comparison of the adsorption capacity with different func- zeta potential measurements. The increase in the negative charge
tionalized nanoparticles are represented in Table 4. on the surface of Fe3O4 NPs are possibly due to the carboxylate
sites on Fe3O4 NPs and the formation of NH2OH sites. Also, the
3.7. Adsorption mechanism increased adsorption capacity of metal ions through the electro-
static attraction mechanism between (NH2OH and RE3+) and sur-
Based on the results obtained from zeta potential measure- face complexation mechanism between (COO and RE3+) seems to
ments (Fig. 3), FT-IR spectrum (Fig. S3, ESIy) and the trend of be the major routes for the adsorption of RE3+ ions on the surface of
RE3+ ions adsorption at different pH values (Fig. 4c and d), we pro- Cys@Fe3O4 NPs [27].
pose a schematic representation of the adsorption mechanisms as FT-IR spectra of the RE3+ loaded (CA@Fe3O4 NPs-RE3+) is shown
shown in scheme 1. Generally, The RE3+ ions can be estimated from in Fig. S1a (ESIy). The stretching vibrational of C@O peak was
the stability constants; it is clear that RE3+ ions in solution is longer, broad and shifted from 1582 to 1605 cm1. Additionally,
R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296 293

Fig. 6. The non-linear Langmuir and Freundlich fitting of adsorption capacity of La3+, Nd3+, Gd3+ and Y3+ on (a) CA@Fe3O4 NPs, time = 30 min and (b) Cys@Fe3O4 NPs,
time = 15 min. Experimental condition: pH = 7.0, m of adsorbents = 2.5 mg and v = 10 mL at room temperature.

Table 3
Parameters of Langmuir and Freundlich adsorption isotherm models.

Model La3+ Nd3+ Gd3+ Y3+


CA@Fe3O4 NPs
Langmuir qe, (mg g1) 32.5 ± 0.71 41 ± 0.11 52 ± 0.20 35.8 ± 0.72
kL (L mg1) 0.7 ± 0.08 0.3 ± 0.09 0.3 ± 0.03 0.67 ± 0.02
R2 0.995 0.999 0.993 0.991
Freundlich n 16 13.8 18 17.3
kF (mg g1)(L g1)n 4.526 3.2 3.35 4.53
R2 0.973 0.978 0.978 0.955
Cys@Fe3O4 NPs
Langmuir qe,(mg g1) 57.2 ± 0.31 85.5 ± 0.39 98 ± 0.50 73 ± 0.34
kL(L mg1) 1.73 ± 0.46 1.812 ± 0.4 0.81 ± 0.11 0.47 ± 0.08
R2 0.988 0. 987 0. 998 0.997
Freundlich n 32.75 46.73 18 13.8
kF (mg g1)(L g1)n 5.32 4.0 3.3 2.98
R2 0.921 0.979 0.978 0.946

Table 4
Reported literature review of La3+, Nd3+, Gd3+ and Y3+ ions adsorption capacity by different adsorbents.

Adsorbents Ions Adsorption


capacity (mg g1)
Magnetic nano-hydroxyapatite [40] Nd and Sm 323 and 370
Magnetic silica nanocomposite (P507) [43] La 55.9
Magnetic alginate (P507) microcapsules [44] Nd 149.3
Fe3O4 (humic acid) [45] Eu 10.6
Dibenzo- 18-crown-6 ether onto Cs 50.32
mesoporous silica monoliths [46]
CA@Fe3O4 NPs (this work) La, Nd, Gd, Y 32.5, 41, 52 and 35.8
Cys@Fe3O4 NPs (this work) La, Nd, Gd, Y 57.2, 85.5, 98 and 73

the symmetric stretching of COO group shifted to 1361 cm1, that the v(C@O) band at 1586 cm1 shifted to 1628 cm1. The
indicating that the oxygen in COO and C@O are involved in chelat- broad and strong bands of tertiary amine at 3055 cm1 were
ing adsorption of RE3+. The FTIR spectrum for Cys@Fe3O4 NPs after shifted to 3378 cm1 due to the interaction between carboxylate
adsorption of RE3+ ions are presented in Fig. S1b (ESIy). It is clear and amine groups in Cys and RE3+ ions [49].
294 R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296

Scheme 1. Schematic representations of proposed mechanisms for adsorption of RE3+ ions by (a) CA@Fe3O4 NPs and (b) Cys@Fe3O4 NPs.

3.8. Desorption studies Mg2+, Ca2+ and Ni2+ ions with initial concentration of
0.01 mmol L1 and 0.05 mmol L1 at pH 7.0 was investigated. From
The recovery of adsorbed RE3+ ions from the loaded CA@Fe3O4 the obtained results (Fig. 8), CA@Fe3O4 NPs and Cys@Fe3O4 NPs
NPs and Cys@Fe3O4 NPs was tested using different eluting solu- adsorbed much more RE3+ ions than other base metal ions. This
tions. The desorption experiments were carried out in an aqueous can be explained in terms of strong binding of La3+, Nd3+, Gd3+
solution of sodium hydroxide (0–0.72%), whereby it was found that and Y3+ with O and N donor atoms than with the common interfer-
the recovered amounts of La3+, Nd3+, Gd3+ and Y3+ were negligible. ing ions. The RE3+ ions are hard Lewis acid cations and have high
The low recovery of all metals may be attributed to the strong tendency to complex with hard Lewis base atoms [50]. Here, the
binding between CA@Fe3O4 NPs/Cys@Fe3O4 NPs with the RE3+ ions CA@Fe3O4 NPs and Cys@Fe3O4 NPs containing O and N donor
in alkaline conditions which is in conformation to the observations atoms strongly bind to La3+, Nd3+, Gd3+ and Y3+ than the other com-
made from pH studies and zeta potential measurements as well as peting cations according to Pearson’s hard and soft acid and base
the observations from the FT-IR measurements. Different concen- theory [51]. The results indicate that the CA@Fe3O4 NPs and
trations of nitric acid were tested for the recovery of RE3+ species Cys@Fe3O4 NPs have higher adsorption affinity of heavy rare earth
from the loaded coated Fe3O4 NPs. as shown in Fig. 7a and b. It is ions such as (Gd3+, Y3+) ions than light rare earth ions such as
clear that recovery of all metals can be accomplished with 0.5 M (Nd3+, La3+). Ionic strength does not significantly seem to affect
HNO3 in which 85% and 97% recovery can be achieved for adsorption process.
CA@Fe3O4 NPs and Cys@Fe3O4 NPs, respectively. The selectivity was reported using separation factor (SF),
which compares the molar ratio of two element ions. The
3.9. Selective separation separation of Gd3+ ions was chosen because of the difference
in the ionic radius. SF was defined as the ratio of distribution
The selective separation performance of La3+, Nd3+, Gd3+ and Y3+ coefficients; SF(Gd/La) = Kd(Gd)/Kd(La), SF(Gd/Nd) = Kd(Gd)/Kd(Nd)
ions from a mixture of a base metal solution containing RE3+ and and SF(Gd/Y) = Kd(Gd)/Kd(Y) (see Eq. (9) for more details).

Fig. 7. Recovery efficiency% of La3+, Nd3+, Gd3+ and Y3+ from the loaded (a) CA@Fe3O4 NPs and (b) Cys@Fe3O4 NPs using different concentration of HNO3 as an eluting solution.
R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296 295

Fig. 8. Separation of La3+, Nd3+, Gd3+ and Y3+ from a mixture of (RE3+ ions and Mg2+, Ca2+ and Ni2+) of 0.01 and 0.05 mmol L1 concentrations by (a) CA@Fe3O4 NPs,
time = 30 min and (b) Cys@Fe3O4 NPs time = 15 min and pH = 7.0 at room temperature.

Table 5
Selective SF of Gd3+ from the mixed Gd3+/La3+, Gd3+/Nd3+ and Gd3+/Y3+ solution using CA@Fe3O4 NPs and Cys@Fe3O4 NPs.

CA@Fe3O4 NPs 0.01 M 0.05 M Cys@Fe3O4 NPs 0.01 M 0.05 M


SF(Gd3+/La3+) 3.7 1.98 SF(Gd3+/La3+) 4.8 2.3
SF(Gd3+/Nd3+) 1.4 1.04 SF(Gd3+/Nd3+) 2.3 1.5
SF(Gd3+/Y3+) 1.8 1.66 SF(Gd3+/Y3+) 2.23 1.8

One can notice that SF increases rapidly at the beginning of the Acknowledgments
Gd3+/La3+, Gd3+/Nd3+and Gd3+/Y3+ separation series and stagnates
for the heaviest lanthanides. The obtained results in Table 5, Radwa M. Ashour is grateful to Erasmus Mundus program (EM-
demonstrate that the separation of Gd3+ from mixed Gd3+/La3+, WELCOME) for financial support and KTH for hosting and giving
Gd3+/Nd3+ and Gd3+/Y3+ solution was achieved by using CA@Fe3O4 the opportunity to work in Functional Materials Laboratory.
NPs and Cys@Fe3O4 NPs. The Cys@Fe3O4 NPs have a higher selectiv-
ity separation toward heavier rare earth ions than CA@Fe3O4 NPs,
Appendix A. Supplementary data
which can be explained by the fact that it has amine and carboxylic
functional groups on the surface.
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.cej.2017.06.101.
4. Conclusions
References
In this work, different chelating functional groups with mag-
netic nanoparticles (CA@Fe3O4 NPs and Cys@Fe3O4 NPs) were suc- [1] Y. Li, B. Hu, Cloud point extraction with/without chelating agent on-line
cessfully synthesized and characterized. The role of the functional coupled with inductively coupled plasma optical emission spectrometry for
the determination of trace rare earth elements in biological samples, J. Hazard.
groups on the adsorption of rare earth ion and selective separation Mater. 174 (2010) 534–540.
of Gd3+ from aqueous solutions was investigated. The adsorption [2] S.V. Eliseeva, J.-C.G. Bunzli, Rare earths: jewels for functional materials of the
behaviour was found to be highly dependent on the surface func- future, New J. Chem. 35 (2011) 1165–1176.
[3] P. Sasikumar, S.V. Narasimhan, S. Velmurugan, Development of a Modified Ion
tionality of Fe3O4 NPs. The time needed to reach maximum adsorp-
Exchange Resin Column for Removal of Gadolinium From the Moderator
tion was attained in less than 30 min and high loading capacity System of PHWRs, Sep. Sci. Technol. 48 (2013) 1220–1225.
was obtained (98 mg g1 and 52 mg g1 for Cys@Fe3O4 NPs and [4] W. Yongxing, W. Xiaorong, H. Zichun, Genotoxicity of lanthanum (III) and
CA@Fe3O4 NPs, respectively). Maximum adsorption capacity qmax gadolinium (III) in human peripheral blood lymphocytes, Bull. Environ.
Contam. Toxicol. 64 (2000) 611–616.
of the RE3+ was better descried by Langmuir isotherm model, and [5] K. Binnemans, P.T. Jones, B. Blanpain, T. Van Gerven, Y.X. Yang, A. Walton, M.
the relative order of qmax values Gd3+ > Nd3+ > Y3+ > La3+ was con- Buchert, Recycling of rare earths: a critical review, J Clean. Prod. 51 (2013) 1–
sistent with the sequence of electronegativity and ionic radial sizes 22.
[6] N. Das, D. Das, Recovery of rare earth metals through biosorption: an overview,
of REEs. The adsorption process is of endothermic nature in both J. Rare Earths 31 (2013) 933–943.
the systems are more favourable at higher temperatures. Adsorp- [7] E. Fujimori, T. Hayashi, K. Inagaki, H. Haraguchi, Determination of lanthanum
tion kinetic studies follow a pseudo-second order model suggest- and rare earth elements in bovine whole blood reference material by ICP-MS
after coprecipitation preconcentration with heme-iron as coprecipitant,
ing that the rate-limiting step is chemisorption. The adsorption Fresenius J. Anal. Chem. 363 (1999) 277–282.
of La3+, Nd3+, Gd3+ and Y3+ was not affected by commonly coexist- [8] S. Tong, N. Song, Q. Jia, W. Zhou, W. Liao, Solvent extraction of rare earths from
ing ions such as Ca2+, Mg2+ and Ni2+, which illustrates the selective chloride medium with mixtures of 1-phenyl-3-methyl-4-benzoyl-pyrazalone-
5 and sec-octylphenoxyacetic acid, Sep. Purif. Technol. 69 (2009) 97–101.
adsorption of La3+, Nd3+, Gd3+ and Y3+ from diluted aqueous solu- [9] Q. Jia, X. Kong, W. Zhou, L. Bi, Flow injection on-line preconcentration with an
tions. High desorption efficiency was obtained in which more than ion-exchange resin coupled with microwave plasma torch-atomic emission
90% RE3+ ions recovery was achieved in both systems under acidic spectrometry for the determination of trace rare earth elements, Microchem. J.
89 (2008) 82–87.
conditions. The Cys@Fe3O4 NPs exhibits highly selective separation
[10] X. Sun, B. Peng, Y. Ji, J. Chen, D. Li, The solid–liquid extraction of yttrium from
factor for pairs of Gd3+ /La3+, Gd3+/Nd3+, Gd3+/Y3+ ions more than rare earths by solvent (ionic liquid) impreganated resin coupled with
CA@Fe3O4 NPs. These results therefore demonstrate that Cys@Fe3- complexing method, Sep. Purif. Technol. 63 (2008) 61–68.
O4 NPs possesses high specificity for the separation of RE3+ ions [11] P. Liang, Y. Qin, B. Hu, T. Peng, Z. Jiang, Nanometer-size titanium dioxide
microcolumn on-line preconcentration of trace metals and their
from aqueous solution and could be applied in the field of recovery determination by inductively coupled plasma atomic emission spectrometry
and separation of REEs from mineral sources. in water, Anal. Chim. Acta 440 (2001) 207–213.
296 R.M. Ashour et al. / Chemical Engineering Journal 327 (2017) 286–296

[12] S. Radhika, V. Nagaraju, B. Nagaphani Kumar, M.L. Kantam, B.R. Reddy, Solid- superparamagnetic nanoclusters at high salinity, Ind. Eng. Chem. Res. 49
liquid extraction of Gd(III) and separation possibilities of rare earths from (2010) 12435–12443.
phosphoric acid solutions using Tulsion CH-93 and Tulsion CH-90 resins, J. [32] M. Kosmulski, Chemical Properties of Material Surfaces, CRC press, 2001.
Rare Earths 30 (2012) 1270–1275. [33] R.A. Bini, R.F.C. Marques, F.J. Santos, J.A. Chaker, M. Jafelicci Jr., Synthesis and
[13] P. Liang, J. Cao, R. Liu, Y. Liu, Determination of trace rare earth elements by functionalization of magnetite nanoparticles with different amino-functional
inductively coupled plasma optical emission spectrometry after alkoxysilanes, J. Magn. Magn. Mater. 324 (2012) 534–539.
preconcentration with immobilized nanometer titanium dioxide, Microchim. [34] Z. Durmus, H. Kavas, M.S. Toprak, A. Baykal, T.G. Altınçekiç, A. Aslan, A.
Acta 159 (2007) 35–40. Bozkurt, S. Cosßgun, l-lysine coated iron oxide nanoparticles: synthesis,
[14] S. Tong, S. Zhao, W. Zhou, R. Li, Q. Jia, Modification of multi-walled carbon structural and conductivity characterization, J. Alloys Compd. 484 (2009)
nanotubes with tannic acid for the adsorption of La, Tb and Lu ions, Microchim. 371–376.
Acta 174 (2011) 257–264. [35] A. Mocanu, I. Cernica, G. Tomoaia, L.-D. Bobos, O. Horovitz, M. Tomoaia-Cotisel,
[15] Y. Yao, S. Miao, S. Liu, L.P. Ma, H. Sun, S. Wang, Synthesis, characterization, and Self-assembly characteristics of gold nanoparticles in the presence of cysteine,
adsorption properties of magnetic Fe3O4@graphene nanocomposite, Chem. Colloids Surfaces A 338 (2009) 93–101.
Eng. J. 184 (2012) 326–332. [36] M.R. Awual, T. Kobayashi, Y. Miyazaki, R. Motokawa, H. Shiwaku, S. Suzuki, Y.
[16] V. Diniz, B. Volesky, Biosorption of La Eu and Yb using Sargassum biomass, Okamoto, T. Yaita, Selective lanthanide sorption and mechanism using novel
Water Res. 39 (2005) 239–247. hybrid Lewis base (N-methyl-N-phenyl-1,10-phenanthroline-2-carboxamide)
[17] R.M. Ashour, H.N. Abdelhamid, A.F. Abdel-Magied, A.A. Abdel-Khalek, M.M. Ali, ligand modified adsorbent, J. Hazard. Mater. 252–253 (2013) 313–320.
A. Uheida, M. Muhammed, X. Zou, J. Dutta, Rare Earth Ions Adsorption onto [37] S. Yusan, C. Gok, S. Erenturk, S. Aytas, Adsorptive removal of thorium (IV) using
Graphene Oxide Nanosheets, Solvent Extr. Ion Exch. (2017) 1–13. calcined and flux calcined diatomite from Turkey: Evaluation of equilibrium,
[18] R. Hao, R. Xing, Z. Xu, Y. Hou, S. Gao, S. Sun, Synthesis, functionalization, and kinetic and thermodynamic data, Appl. Clay Sci. 67–68 (2012) 106–116.
biomedical applications of multifunctional magnetic nanoparticles, Adv. [38] E. Kim, K. Osseo-Asare, Aqueous stability of thorium and rare earth metals in
Mater. 22 (2010) 2729–2742. monazite hydrometallurgy: Eh–pH diagrams for the systems Th–, Ce–, La–,
[19] B. Gao, Y. Gao, Y. Li, Preparation and chelation adsorption property of Nd– (PO4)–(SO4)–H2O at 25 °C, Hydrometallurgy 113–114 (2012) 67–78.
composite chelating material poly(amidoxime)/SiO2 towards heavy metal [39] M. Vaca, R. Mier López Callejas, R. Gehr, B.E. Jiménez Cisneros, P.J.J. Alvarez,
ions, Chem. Eng. J. 158 (2010) 542–549. Heavy metal removal with mexican clinoptilolite: multi-component ionic
[20] L. Li, K.Y. Mak, C.W. Leung, K.Y. Chan, W.K. Chan, W. Zhong, P.W.T. Pong, Effect exchange, Water Res. 35 (2001) 373–378.
of synthesis conditions on the properties of citric-acid coated iron oxide [40] C. Gok, Neodymium and samarium recovery by magnetic nano-
nanoparticles, Microelectron. Eng. 110 (2013) 329–334. hydroxyapatite, J. Radioanal. Nucl. Chem. 301 (2014) 641–651.
[21] Y. Su, A.S. Adeleye, A.A. Keller, Y. Huang, C. Dai, X. Zhou, Y. Zhang, Magnetic [41] C.-H. Weng, C.P. Huang, Adsorption characteristics of Zn(II) from dilute
sulfide-modified nanoscale zerovalent iron (S-nZVI) for dissolved metal ion aqueous solution by fly ash, Colloids Surfaces A 247 (2004) 137–143.
removal, Water Res. 74 (2015) 47–57. [42] R. Rahal, F. Annani, S. Pellet-Rostaing, G. Arrachart, S. Daniele, Surface
[22] E.-J. Kim, C.-S. Lee, Y.-Y. Chang, Y.-S. Chang, Hierarchically structured modification of titanium oxide nanoparticles with chelating molecules: New
manganese oxide-coated magnetic nanocomposites for the efficient removal recognition devices for controlling the selectivity towards lanthanides ionic
of heavy metal ions from aqueous systems, ACS Appl. Mater. Interfaces 5 separation, Sep. Purif. Technol. 147 (2015) 220–226.
(2013) 9628–9634. [43] D. Wu, Y. Sun, Q. Wang, Adsorption of lanthanum (III) from aqueous solution
[23] S. Shin, J. Jang, Thiol containing polymer encapsulated magnetic nanoparticles using 2-ethylhexyl phosphonic acid mono-2-ethylhexyl ester-grafted
as reusable and efficiently separable adsorbent for heavy metal ions, Chem. magnetic silica nanocomposites, J. Hazard. Mater. 260 (2013) 409–419.
Commun. 4230–4232 (2007). [44] L. Zhang, D. Wu, B. Zhu, Y. Yang, L. Wang, Adsorption and selective separation
[24] D. Dupont, W. Brullot, M. Bloemen, T. Verbiest, K. Binnemans, Selective uptake of neodymium with magnetic alginate microcapsules containing the
of rare earths from aqueous solutions by edta-functionalized magnetic and extractant 2-ethylhexyl phosphonic acid mono-2-ethylhexyl ester, J. Chem.
nonmagnetic nanoparticles, ACS Appl. Mater. Interfaces 6 (2014) 4980–4988. Eng. Data 56 (2011) 2280–2289.
[25] H. Zhang, R.G. McDowell, L.R. Martin, Y. Qiang, Selective extraction of heavy [45] S. Yang, P. Zong, X. Ren, Q. Wang, X. Wang, Rapid and highly efficient
and light lanthanides from aqueous solution by advanced magnetic preconcentration of Eu(III) by core-shell structured Fe3O4@Humic acid
nanosorbents, ACS Appl. Mater. Interfaces 8 (2016) 9523–9531. magnetic nanoparticles, ACS Appl. Mater. Interfaces 4 (2012) 6891–6900.
[26] X. Shen, Q. Wang, W. Chen, Y. Pang, One-step synthesis of water-dispersible [46] M.R. Awual, S. Suzuki, T. Taguchi, H. Shiwaku, Y. Okamoto, T. Yaita, Radioactive
cysteine functionalized magnetic Fe3O4 nanoparticles for mercury(II) removal cesium removal from nuclear wastewater by novel inorganic and conjugate
from aqueous solutions, Appl. Surf. Sci. 317 (2014) 1028–1034. adsorbents, Chem. Eng. J. 242 (2014) 127–135.
[27] S. Singh, K.C. Barick, D. Bahadur, Surface engineered magnetic nanoparticles [47] W. Brullot, N.K. Reddy, J. Wouters, V.K. Valev, B. Goderis, J. Vermant, T.
for removal of toxic metal ions and bacterial pathogens, J. Hazard. Mater. 192 Verbiest, Versatile ferrofluids based on polyethylene glycol coated iron oxide
(2011) 1539–1547. nanoparticles, J. Magn. Magn. Mater. 324 (2012) 1919–1925.
[28] J. Sangeetha, J. Philip, Synthesis, characterization and antimicrobial property of [48] S.S. Banerjee, D.-H. Chen, Fast removal of copper ions by gum arabic modified
Fe3O4-Cys-HNQ nanocomplex, with l-cysteine molecule as a linker, RSC Adv. 3 magnetic nano-adsorbent, J. Hazard. Mater. 147 (2007) 792–799.
(2013) 8047–8057. [49] F. Zhao, E. Repo, M. Sillanpää, Y. Meng, D. Yin, W.Z. Tang, Green synthesis of
[29] S. Nigam, K.C. Barick, D. Bahadur, Development of citrate-stabilized Fe3O4 magnetic EDTA- and/or DTPA-cross-linked chitosan adsorbents for highly
nanoparticles: conjugation and release of doxorubicin for therapeutic efficient removal of metals, Ind. Eng. Chem. Res. 54 (2015) 1271–1281.
applications, J. Magn. Magn. Mater. 323 (2011) 237–243. [50] Y. Hasegawa, S. Tamaki, H. Yajima, B. Hashimoto, T. Yaita, Selective separation
[30] G. Salazar-Alvarez, J. Sort, A. Uheida, M. Muhammed, S. Surinach, M.D. Baro, J. of samarium (III) by synergistic extraction with b-diketone and
Nogues, Reversible post-synthesis tuning of the superparamagnetic blocking methylphenylphenanthroline carboxamide, Talanta 85 (2011) 1543–1548.
temperature of [gamma]-Fe2O3 nanoparticles by adsorption and desorption of [51] R.G. Pearson, Hard and soft acids and bases, HSAB, part 1: fundamental
Co(ii) ions, J. Mater. Chem. 17 (2007) 322–328. principles, J Chem Educ 45 (1968) 581.
[31] C. Kotsmar, K.Y. Yoon, H. Yu, S.Y. Ryoo, J. Barth, S. Shao, M. Prodanović, T.E.
Milner, S.L. Bryant, C. Huh, K.P. Johnston, Stable citrate-coated iron oxide

You might also like