Skip to main content
Top
Published in: Journal of Materials Science 6/2014

Open Access 01-03-2014

Evaluation of iron-based hybrid materials for heavy metal ions removal

Authors: Dorota Kołodyńska, Maciej Kowalczyk, Zbigniew Hubicki

Published in: Journal of Materials Science | Issue 6/2014

Activate our intelligent search to find suitable subject content or patents.

search-config
download
DOWNLOAD
print
PRINT
insite
SEARCH
loading …

Abstract

The article presents in detail the results of the sorption of heavy metal ions such as Cu(II), Zn(II), Cd(II), and Zn(II) in the presence of ethylenediamine-N,N′-disuccinic acid as well as chromium(VI) on Purolite Arsen Xnp and Lewatit FO36. Factors affecting the sorption equilibrium (sorbent dose, contact time, temperature, pH, and the presence of interfering ions) were studied. To compare the surface morphologies of the studied ion exchangers, scans using atomic force microscope were also recorded, and attenuated total reflectance infrared spectroscopy was applied to establish the adsorption mechanism. The main parameters affecting sorption are initial concentration of the solution, pH, and phase contact time. Temperature has only a slight influence. The kinetic data were fitted using the pseudo-first-order and pseudo-second-order models. In the case of Cr(VI) adsorption, the equilibriums on Purolite Arsen Xnp and Lewatit FO36 were established within 60 min, in the case of Cu(II) and Zn(II) within 30–40 min and for Cd(II) and Pb(II) even less than 30 min. Moreover, it was found that the effectiveness of adsorption in the case of Cr(VI), Cu(II), Zn(II), Cd(II), and Pb(II) on Purolite ArsenXnp was higher than that on Lewatit FO36.

Introduction

Removal of heavy metal ions from various types of waters and waste waters is an essential problem in environmental protection and has a result on human health. On the other hand, stricter requirements concerning the quality of waste waters lead to more refined and complex chemical processes used in their purification. Therefore, the search for new methods and novel materials for heavy metal ions removal, particularly of As(V) and Cr(VI) ions, is very important [18]. Some of them are hybrid ion exchangers (HIX) based on nanosized metal oxides (NMOs) such as ferric, manganese, aluminum, titanium, magnesium, and cerium oxides. For example, the addition of hydrated iron oxides to the polymers (PS and DVB, less frequently polyacrylic, with a macroporous or gel matrix, containing cation or anion exchange functional groups) leads to the obtaining of adsorbents characterized by very good adsorption, mechanical, and hydraulic properties [9, 10]. The quality of the sorbents obtained depends on: (i) the quantity, structure, crystallinity, and degree of dispersion of the iron oxides introduced; (ii) physical form of the base polymer; and (iii) the type of functional groups involved in the preparation of hybrid polymers as well as influence on the process of adsorption. For example, Qiu et al. [11] found that the hybrid sorbent based on the hydrous iron oxide and neutral polystyrene matrix (PS) without an attached functional group (HFO–PS) and based on the negatively charged HFO–PS–SO3 matrix had totally different adsorption properties toward Cu(II) ions in the presence of sulfate(VI) ions. The adsorption of Cu(II) on HFO–PS was promoted in sulfate solutions; however, when influenced by covalently attached sulfonic acid groups, Cu(II) adsorption on HFO–PS–SO3 was inhibited by the added sulfates.
As described by Sarkar et al. [12], the first step in the preparation of HIX based on iron oxides and polystyrene–divinylbenzene copolymers (PS–DVB) includes loading of Fe(III) onto the functional groups of ion exchanger, then desorption and simultaneous precipitation of Fe(III) hydroxides using a solution containing NaCl and NaOH, and finally rinsing and washing the resin with an ethanol–water mixture followed by a mild thermal treatment. The high concentration of the functional groups allowed high and fairly uniform loading of hydrous iron oxide particles (~9–12 % of Fe by mass) within the polymeric matrix. Such ions as Cl, SO4 2−, and HCO3 exhibit only weak affinities for hydrous iron oxides and are characterized by a smaller selectivity. Cumbal and SenGupta [5] reported the preparation of two classes of polymer-supported nanoparticles containing iron for As(III,V) removal: (i) hydrated Fe(III) oxides dispersed in an ion exchange resin, and (ii) polymeric particles with magnetic properties.
Increasing the amount of iron oxide introduced into the polymer increases the adsorption capacity of the sorbent, but after crossing saturation of the polymer matrix followed by the blocking of pores, a significant decrease in adsorption capacity is observed. Typically, the iron oxide content in the copolymer is equal to 100–200 mg Fe g−1 of sorbent. The best adsorption properties of hybrid polymers are found for those in which the inner surfaces of the pores are coated with a thin layer of iron oxides. This effect can be achieved by proper selection of the polymer matrix, the amount of iron placed in a single cycle of synthesis, and by performing appropriate treatment. These factors also affect the degree of dispersion of the introduced iron oxides. The introduction of smaller agglomerates provides a greater surface area and thus a more effective adsorbent.
In this article, the HIX Purolite Arsen Xnp (Purolite) and Lewatit FO36 (Lanxess) were used to remove Cu(II), Zn(II), Cd(II), and Pb(II) ions in the presence of ethylenediamine-N,N′-disuccinic acid (EDDS) which belongs to a group of readily biodegradable complexing agents. The metals for the evaluation were chosen due to their occurrence in waste water streams. Optimal conditions of adsorption were based on the speciation distribution of the studied heavy metal complexes with EDDS. The adsorption properties of the above-mentioned sorbents were also evaluated in the case of arsenic and chromium.
Investigations were carried out using the static method, based on which adsorption parameters were calculated in single versus multicomponent systems at different pH values and temperatures as well as in the presence of interfering ions. To compare the surface morphologies of the studied ion exchangers before and after the adsorption process, scans using atomic force microscope (AFM) were also recorded, and infrared spectroscopy (ATR/FT-IR) was applied to establish the adsorption mechanism.

Materials and methods

Purolite Arsen X np also known as FerriIX™ A33E is a strongly acidic macroporous cation exchanger based on a polystyrene-divinylbenzene skeleton with attached sulfonic acid groups and containing hydrated iron oxide nanoparticles. The resin has the bead size of 0.3–1.2 mm; adsorption capacity 1.56 meq L−1; and iron oxide content not exceeding 42 %. It is also characterized by an enormous capacity for As(V) [13].
Lewatit FO36 is a weakly basic, macroporous anion exchanger also based on PS–DVB and containing the weak amine functional groups and iron oxide nanoparticles. It has the bead size 0.34–0.38 mm, adsorption capacity 1.5 eq L−1 (30 BV h−1, 0.1 mg L−1 As(V), 6 mg L−1 SiO2, and 60 mg L−1 PO4 3−), and iron oxide content not exceeding 23 % [14]. The following salts were used: CuCl2·2H2O, ZnCl2, Cd(NO3)2·4H2O, and Pb(NO3)2 as a source of metal ions, and the S,S-isomer of EDDS (Innospec Speciality Chemicals, UK) used as supplied to react with these metal ions in solution without pH adjustment. In addition, a Cr(VI) working stock solution was prepared by dissolving K2Cr2O7 in deionized water. Experiments were carried out in multicomponent systems containing Cr(VI), Cu(II), and Zn(II) ions, with or without the complexing agent (EDDS) as well as in the presence of accompanying ions Cl, NO3 , and SO4 2−. In the investigations mixtures of Cd(II) and Pb(II), with and without EDDS were also used. In these systems, metal ions are complexed in the form of [M(edds)]2− ions, whereas, Cr(VI) is not complexed.
Batch kinetic experiments were conducted in 250-mL conical flasks containing ion exchanger (2 g) to which a solution of metal ions (200 mL) was added, with and without EDDS and with the initial pH values of 4.5 and 10.0. The flasks were shaked on a rotary shaker at 180 rpm. Batch studies were carried out to investigate the effects of phase contact time (1–180 min), initial metal concentration (1–5 mmol L−1), solution pH (4–12), temperature (20–60 °C), and the presence of other ions (Cl, NO3 , SO4 2−). Solution pH was adjusted with 0.1-mol L−1 HNO3 or 0.1-mol L−1 NaOH.
A laboratory shaker Elpin type 357, (Elpin-Plus, Poland) was used for agitation. The pH values were measured with a PHM 84 pH meter (Radiometer, Denmark) with glass REF 451 and calomel pHG 201-8 electrodes. The accuracy of the pH meter is ±0.01 pH units. Concentrations of the heavy metals were measured with an AAS spectrometer SpectrAA 240FS (Varian, Australia). Measurements were performed using the following parameters: wavelength Cu(II): 324.70 nm, Zn(II): 213.90 nm, Cd(II): 228.80 nm, and Pb(II): 217.00 nm with slit width = 0.50 nm. Cr(VI) ions in the effluent were determined colorimetrically using 1,5-diphenyl-carbazide method at 546 nm with a Specord M42 (Carl Zeiss, Jena) UV–Vis spectrophotometer. The morphology of the hybrid sorbent samples was analyzed with AFM Digital Instruments (USA) NanoScope III. The infrared spectra were recorded by a Nicolet-380 Fourier transform infrared spectrometer using the ATR method. All experiments were made in triplicate, and the results were taken as the average value.
The amounts of metal ions adsorbed at time “t” of hybrid sorbents (q t , mg g−1) were calculated from the mass balance equation:
$$ q_{t} = {{\left( {c_{0} - c_{t} } \right)V} \mathord{\left/ {\vphantom {{\left( {c_{0} - c_{t} } \right)V} m}} \right. \kern-0pt} m}, $$
where c 0 is the initial concentration of a metal ion (mg L−1); c t is the concentration of metal ion at time t (mg L−1), V is the volume of the solution (L); and m is the mass of HIX (g).
The equilibrium data were analyzed using the Langmuir and Freundlich isotherms and the characteristic parameters for each isotherm determined. The linear form of the Langmuir model can be expressed as [15]
$$ {{c_{\text{e}} } \mathord{\left/ {\vphantom {{c_{\text{e}} } {q_{\text{e}} = {{c_{\text{e}} } \mathord{\left/ {\vphantom {{c_{\text{e}} } {q_{0} + {1 \mathord{\left/ {\vphantom {1 {K_{\text{L}} }}} \right. \kern-0pt} {K_{\text{L}} }}q_{0} }}} \right. \kern-0pt} {q_{0} + {1 \mathord{\left/ {\vphantom {1 {K_{\text{L}} }}} \right. \kern-0pt} {K_{\text{L}} }}q_{0} }}}}} \right. \kern-0pt} {q_{\text{e}} = {{c_{\text{e}} } \mathord{\left/ {\vphantom {{c_{\text{e}} } {q_{0} + {1 \mathord{\left/ {\vphantom {1 {K_{\text{L}} }}} \right. \kern-0pt} {K_{\text{L}} }}q_{0} }}} \right. \kern-0pt} {q_{0} + {1 \mathord{\left/ {\vphantom {1 {K_{\text{L}} }}} \right. \kern-0pt} {K_{\text{L}} }}q_{0} }}}}, $$
where c e is the equilibrium concentration of metal ions (mg L−1) and q e is the amount of the metal ions adsorbed on the hybrid sorbent (mg g−1), and q 0 and K L are the Langmuir constants related to the adsorption capacity (mg g−1) and the equilibrium constant (L g−1), respectively.
The adsorption equilibrium data were also applied to the Freundlich model [16]:
$$ { \log }\;q_{\text{e}} = { \log }\;K_{\text{F}} + \left( {{1 \mathord{\left/ {\vphantom {1 n}} \right. \kern-0pt} n}} \right){ \log }\;c_{\text{e}} , $$
where K F and n are the Freundlich constants related to the adsorption capacity and the adsorption intensity, respectively.
The pseudo-first-order model (PF-order) was presented by Lagergren and expressed as follows [17]:
$${\text{log(}}q_{e} - q_{t} ) = {\text{log}}q_{e} - (k_{1} /2.303)t $$
where q t is the amount of metal ions adsorbed on the hybrid sorbents at time t, and k 1 is the rate constant (min−1). The rate constant, k 1, was obtained from the slope of linear plots of log(q e − q t ) against t.
The adsorption data were also analyzed in terms of the pseudo-second order mechanism (PS-order) [18, 19]:
$$ {t \mathord{\left/ {\vphantom {t {q_{t} }}} \right. \kern-0pt} {q_{t} }} = \left( {{1 \mathord{\left/ {\vphantom {1 h}} \right. \kern-0pt} h}} \right) + \left( {{1 \mathord{\left/ {\vphantom {1 {q_{\text{e}} }}} \right. \kern-0pt} {q_{\text{e}} }}} \right)t, $$
and the initial rate of adsorption h is
$$ h = k_{2} q_{\text{e}}^{2} , $$
where k 2 is the rate constant of PS-order adsorption (g(mg min)−1), and h is the initial rate of adsorption (mg(g min)−1).
The constants q e, h, and k 2 can be determined from the plots of t/q t against t.

Results and discussion

The study showed that the hybrid sorbents, including Purolite Arsen Xnp and Lewatit FO36, can be used to remove Cu(II), Zn(II), Cd(II), Pb(II), and Cr(VI) under appropriate conditions. The effectiveness of the adsorption process depends on the initial concentration of solution, pH, phase contact time, addition of a complexing agent EDDS and presence of accompanying ions.
It was found that the optimal pH values for the adsorption of Cu(II), Zn(II), Cd(II), Pb(II), and Cr(VI) ions on Purolite Arsen Xnp and Lewatit FO36 are above 4.5 (Fig. 1).
Under these conditions, Cu(II), Zn(II), Cd(II), and Pb(II) ions are coordinated to the goethite phase. It should be mentioned that, for example, Cu(II) at pH between 2 and 6.5 occurs predominantly as the aqueous Cu2+ cation. With increasing pH, the following hydroxoions occur Cu(OH)+, Cu(OH)2, Cu(OH) 3 , and Cu(OH) 4 2− . At pH values >7, the major hydrolysis product is Cu2(OH) 2 2+ . Therefore, at pH values >4.5, Cu(II) is likely to sorb as the aqueous cation. Under analogous conditions, other ions are also in the form of cations or hydroxoions. Numerous studies show that the Cu(II) ions are sorbed on goethite as the surface complexes ≡FOCu+ and ≡FOCuOH or a combination of both [20, 21].
However, as complexing agents are present, these are adsorbed specifically on metal oxides and hydroxides to form inner sphere complexes. It is also known that they are responsible for increase, decrease, or no change in the adsorption of metal ions, for example on soils. Some authors confirm that the adsorption of metals in systems containing organic ligands increases at pH values below the adsorbent point of zero charge (PZC) and increases at higher pH [22]. However, other authors showed that the addition of organic substances decreases the adsorption even under acid conditions below the PZC [23].
In the present article, ethylenediamine-N,N′-disuccinic acid (H4edds) consists of four isomers: S,S- (25 %), R,R- (25 %), R,S- (50 %), and S,R- (50 %); however, only the S,S-isomer of EDDS is readily biodegradable. It was proved that 83 % of it is converted to CO2 within 20 days [24]. The other isomers are either partly or completely nonbiodegradable [25, 26]. It follows from the EDDS speciation profile at pH 10 that the predominant form is the anion (edds)4− (Fig. 2).
EDDS has four ionisable protons with the following protonation constants:
$$ \begin{aligned} {\text{edds}}^{ 4- } + {\text{H}}^{ + } \rightleftarrows & {\text{Hedds}}^{ 3- } ,\quad & { \log }\;K_{\text{p1}} = 9. 70 \\ {\text{Hedds}}^{ 3- } + {\text{H}}^{ + } \rightleftarrows & {\text{H}}_{ 2} {\text{edds}}^{ 2- } ,\quad & { \log }\;K_{\text{p2}} = 7.0 1\\ {\text{H}}_{ 2} {\text{edds}}^{ 2- } + {\text{H}}^{ + } \rightleftarrows & {\text{H}}_{ 3} {\text{edds}}^{ - } ,\quad & { \log }\;K_{\text{p3}} = 3. 9 2\\ {\text{H}}_{ 3} {\text{edds}}^{ - } + {\text{H}}^{ + } \rightleftarrows & {\text{H}}_{ 4} {\text{edds}},\quad & { \log }\;K_{\text{p4}} = 3.0 3\\ \end{aligned} $$
and (Σ log K pi  = 23.66, i = 1–4) [27, 28].
According to the literature data, EDDS showed good complexing properties toward heavy metal ions, and the reaction is characterized by the formation of stable 1:1 metal to ligand complexes as the major species:
$$ {\text{M}}^{ 2+ } + {\text{edds}}^{ 4- } \rightleftarrows \left[ {{\text{M}}\left( {\text{edds}} \right)} \right]^{ 2- } . $$
For the Cu(II), Zn(II), Cd(II), and Pb(II) complexes with EDDS, the stability constants are as follows: 13.1, 13.4, 16.4, and 12.7, respectively.
According to the literature data, metal capacity on the goethite surface increases with metal electronegativity: Cu(II) > Pb(II) > Cd(II) [29]. In addition, the behavior of the metals in the three-component system consisting of metal solution; organic ligand and hybrid sorbent can be controlled by several processes: (i) the adsorption of the free metal cations on the sorbent surface, (ii) the adsorption of the metal cations by the sorbent surface modified by interaction with the anions of dissolved EDDS, (iii) the formation of metal complexes with EDDS in solution, (iv) the adsorption of free and charged metal complexes with EDDS on the sorbent surface, and finally (v) the interaction of anionic complexes with the functional groups of the hybrid sorbents.
To establish which of the above-mentioned steps has decisive influence on the sorption process in the next stage, the effect of the phase contact time on the effectiveness of Cu(II), Zn(II), Cd(II), Pb(II), and Cr(VI) ions’ adsorption at different initial concentrations (1–5 mmol L−1) was studied, and the examples of results for Purolite Arsen Xnp are presented in Fig. 3.
It was found that in the case of Cr(VI) adsorption, the equilibrium on Purolite Arsen Xnp was established within 60 min, in the case of Cu(II) and Zn(II) within 30–40 min. (at low concentrations), and for Cd(II) and Pb(II) even <30 min. For Lewatit FO36, the equilibrium times were almost analogous. Moreover, it was found that the effectiveness of adsorption for these ions on Purolite ArsenXnp was higher than on Lewatit FO36.
Based on the kinetic analysis, it was found that the rate constants k 1 decreased with increase in the initial metal ion concentrations (data not presented for the concentrations 3 and 5 mmol L−1). The correlation coefficients were high, ranging from 0.8253 to 0.9284 for Purolite Arsen Xnp and from 0.9562 to 0.9982 for Lewatit FO 36. However, the calculated equilibrium capacities (q 1) according to the Lagergren pseudo-first order rate expression were in disagreement with the experimental capacities (q e,exp) for solutions with initial concentrations 1–5 mmol L−1. Better results were obtained using the pseudo-second order equation (Table 1).
Table 1
Kinetic parameters for Cr(II), Cu(II), Zn(II), Cd(II), and Pb(II) ion sorption on Purolite Arsen Xnp and Lewatit FO36 at initial concentration 1 mmol L−1
System
Cr(VI)
Cu(II)
Zn(II)
Cd(II)
Pb(II)
Purolite Arsen Xnp
 PF-order
  q exp
5.19
5.05
8.35
6.42
16.47
  q 1
2.62
2.24
2.98
3.16
3.63
  k 1
0.094
0.051
0.020
0.024
0.065
  R 2
0.9284
0.9109
0.9510
0.9115
0.9049
 PS-order
  q 2
4.80
5.03
8.31
6.44
16.61
  k 2
0.078
0.129
0.211
0.163
0.081
  h
1.806
3.275
14.556
6.775
22.422
  R 2
0.9922
0.9907
0.9910
0.9968
0.9976
Lewatit FO36
 PF-order
  q exp
5.80
2.87
3.05
2.72
18.65
  q 1
4.12
4.37
4.46
1.50
12.11
  k 1
0.141
0.009
0.008
0.069
0.127
  R 2
0.9985
0.9791
0.9562
0.7931
0.9571
 PS-order
  q 2
5.88
3.10
3.06
2.92
18.55
  k 2
0.039
0.122
0.144
0.445
0.011
  h
1.359
1.170
1.356
3.804
3.900
  R 2
0.9982
0.9913
0.9932
0.9957
0.9963
In addition, studies on the effect of ethylenediamine-N,N′-disuccinic acid (EDDS) addition clearly showed that the influence of complexing agent is different depending on the system (Figs. 4, 5). For Purolite ArsenXnp, addition of EDDS had a negative impact on the adsorption of Cu(II) and Pb(II), however, for Cd(II) an increase of the removal efficiency was observed. In the case of Lewatit FO36, much better adsorption in the presence of EDDS was found for Cu(II), Zn(II), and Cd(II), while it had no effect on Cr(VI) adsorption. As for Pb(II) ions, their removal from waste waters decreases. The increase in the metal adsorption is probably related to the formation of stable negatively charged M(II)-EDDS complexes of the type M[(Hedds)]3− or [M(edds)]2−. These complexes increase the adsorption capacity of the surface due to the involvement of positively charged sorption sites into the adsorption process. Their adsorption on the hybrid sorbents surface can occur via the metal–surface interaction and through the organic ligand [20, 30].
$$ \begin{aligned} \equiv & {\text{Fe}}{-}{\text{OH}} + {\text{M}}^{ 2+ } + {\text{edds}}^{{{\text{n}} - }} + {\text{H}}^{ + } \rightleftarrows \equiv {\text{Fe}}{-}{\text{edds}}{-}{\text{M}}^{{({\text{n}} - 3) - }} + {\text{H}}_{ 2} {\text{O}} \\ \equiv & {\text{Fe}}{-}{\text{OH}} + {\text{M}}^{ 2+ } + {\text{edds}}^{{{\text{n}} - }} \rightleftarrows \equiv {\text{Fe}}{-}{\text{O}}{-}{\text{M}}{-}{\text{edds}}^{{({\text{n}} - 1) - }} + {\text{H}}^{ + } \\ \end{aligned} $$
where M2+ is the metal cation and edds4− is the ligand.
Moreover, especially for Lewatit FO36 with weakly basic functional groups, the anion exchange process should be considered, and this mechanism is probably responsible for the enhanced adsorption of the negatively charged complexes:
$$ 2 {\text{RCl}} + \left[ {{\text{M}}\left( {\text{edds}} \right)} \right]^{ 2- } \rightleftarrows {\text{R}}_{ 2} \left[ {{\text{M}}\left( {\text{edds}} \right)} \right] + 2 {\text{Cl}}^{ - } . $$
As was shown by Ali and Dzombak [31], sulfate ligands can dramatically enhance the adsorption of heavy metals on iron oxides by modifying the electrostatic environment at the interface or by forming metal-sulfate ternary surface complexes like ≡Fe–OHCuSO4. As was mentioned earlier, Qiu et al. [11] also found that the Cu(II) adsorption on a hybrid sorbent based on the PS–SO3 matrix is strongly inhibited by the presence of the SO4 2− ligand. Therefore, the effects of accompanying ions such as Cl, NO3 , and SO4 2− were studied. From the obtained results, it should be noted that the effects of accompanying ions are negligible; however, the addition of the complexing agent EDDS to the system in the presence of Cl, NO3 , and SO4 2− ions improves the sorption process parameters for both hybrid sorbents. This is mainly noticeable in the case of Zn(II), where the addition of EDDS significantly improves the sorption capacity from 14.43 to 21.50 mg g−1 for Purolite ArsenXnp as well as from 14.43 to 21.50 mg g−1 for Lewatit FO36 (initial concentration 3 mmol L−1). In the case of Cr(VI), this increase is negligible, and in the case Cu(II) ions the sorption capacity decreases from 18.46 to 9.22 mg g−1 in the same conditions (Figs. 6, 7).
In addition, it is found that the sorption capacity of the sorbents slightly increased when the temperature of heavy metal solutions increased from 20 to 60 °C. The sorption of the above complexes could be described by the Langmuir isotherm with high correlation coefficients ranging from 0.9986 to 1.000, whereas for the Freundlich isotherm they range from 0.8345 to 0.9871. The values of the separation factor, R L, indicate the nature of the adsorption process with: R L > 1 unfavorable; R L = 1 linear; 0 < R L < 1 favorable; and R L = 0 irreversible. In the present study, these values indicate that the adsorption process is favorable for both the Purolite ArsenXnp and Lewatit FO36 samples. The maximum capacities of adsorption of Cr(VI), Cu(II), Zn(II), Cd(II), and Pb(II) on Purolite Arsen Xnp are equal to 25.55, 21.47, 10.11, 28.99, and 81.85 mg g−1, respectively, and they increase with the increasing temperature. Meanwhile, the Langmuir equilibrium constants (K L) (L g−1) for Purolite Arsen Xnp decrease as follows: Cr(VI): 0.240–0.017; Cu(II): 0.185–0.014; Zn(II): 0.122–0.0087; Cd(II): 0.221–0.013; and Pb(II): 0.324–0.025. The Langmuir sorption isotherms of Cd(II) and Pb(II) on Lewatit FO36 and Purolite Arsen Xnp are presented in Fig. 8. Moreover, the Langmuir and Freundlich parameters of sorption isotherms of Cd(II) and Pb(II) on hybrid sorbents are presented in Table 2. The comparison between the determined sorption capacities of Purolite ArsenXnp and Lewatit FO36 toward heavy metal ions is presented in Table 3.
Table 2
The Langmuir and Freundlich parameters for adsorption of Cd(II) and Pb(II) on Lewatit FO36 and Purolite Arsen Xnp (1.0 × 10−3–2.5 × 10−2 M, shaking time 24 h, shaking speed 180 rpm, temperature 295 K)
System
Langmuir model
Freundlich model
q 0
K L
R L
R 2
K F
n
R 2
Lewatit FO36
 Cd(II)
50.00
0.001
0.385
0.8346
6.29
1.48
0.9655
 Pb(II)
138.89
0.004
0.043
0.9882
9.89
3.02
0.9266
Purolite Arsen Xnp
 Cd(II)
11.15
0.001
0.373
0.7918
9.82
0.53
0.9028
 Pb(II)
147.06
0.005
0.035
0.9827
15.77
3.57
0.8953
Table 3
Comparison of adsorption capacities for heavy metal ions onto different sorbents containing hydrous iron oxides
Sorbents
Conditions
Adsorption capacity for experimental conditions
Ref.
HIO
hydrous iron oxide in chloride–sulfate 1:1 system in the presence of phosphate, acetate, and oxalate ions
m = 1 mg
Cu(II) = 10 mg L−1
Zn(II) = 10 mg L−1
V = 50 mL
pH = 5.5
Cu(II) > 90 %
Zn(II) > 90 %
[21]
Nano-Fe3O4–SiO2–TETA
magnetic nano-iron oxide–nano-silicon oxide-triethylenetetramine
m = 5 mg
V = 50 mL
CCu(II) = 127 mg L−1
CPb(II) = 414 mg L−1
pH = 1–7
Pb(II) 269 mg g−1
in pH 7.0
Cu(II) 76 mg g−1
in pH 7.0.
[33]
SCMN
shellac-coated magnetic nanoparticle adsorbent
m = 0.5 mL
V = 50 mL
CCd(II) = 50 mg L−1
pH = 8
Cd(II) 20.6612 mg g−1
[34]
INS
iron-oxide nano-particles-immobilized sand
m = 0.5 g
V = 100 mL
CCu(II) = 20 mg L−1
CCd(II) = 20 mg L−1
CPb(II) = 20 mg L−1
pH = 1–10
Cu(II) 1.26 mg g−1
Cd(II) 0.52 mg g−1
Pb(II) 2.08 mg g−1
[32]
INR
iron nanorods
m = 50 mg
V = 50 mL
CFe(II) = 10 mg L−1
CPb(II) = 10 mg L−1
CZn(II) = 10 mg L−1
CNi(II) = 10 mg L−1
CCd(II) = 10 mg L−1
CCu(II) = 10 mg L−1
pH = 1–8
Fe(II) 43.11 mg g−1
Pb(II) 38.04 mg g−1
Zn(II) 36.12 mg g−1
Ni(II) 32.13 mg g−1
Cd(II) 29.76 mg g−1
Cu(II) 25.62 mg g−1
[35]
HFO–PS and HFOPS0, respectively) hydrous ferric oxide (HFO) nanoparticles—polystyrene beads (negatively charged and covalently bonded with neutral chloromethyl) in the presence of SO4 2− and NO3 ions
m = 5 mg
V = 50 mL
CCu(II) = 10 mg L−1
pH = 5.5
Cu(II) 215.7 mg g−1
(for 100 mg L−1 of Ca(II)),
184.2 mg g−1
(for 200 mg L−1 of Ca(II))
165.2 mg/g
(for 300 mg L−1 of Ca(II))
[9]
MNPs-NH2
amino-functionalized magnetic nano-adsorbent
m = 5 mg
V = 50 mL
CPb(II) = 5 mg L−1
CPb(II) = 15 mg L−1
CPb(II) = 25 mg L−1
pH = 5.0
Pb(II) 8.82 mg g−1
(for Pb(II) C0 5 mg L−1)
Pb(II) 22.78 mg g−1
(for Pb(II) C015 mg L−1) Pb(II) 41.10 mg g−1
(for Pb(II) C025 mg L−1)
[36]
Magnetic maghemite
(γ-Fe2O3) nanotubes
m = 50 mg
V = 100 mL
CCu(II) = 100 mg L−1
CZn(II) = 100 mg L−1
CPb(II) = 100 mg L−1
pH = 6.0
Cu(II) 37.31 mg g−1
Zn(II) 31.25 mg g−1
Pb(II) 24.87 mg g−1
[37]
Fe–Al–NF
Iron oxide–alumina mixed nanocomposite fiber
m = 50 mg
V = 20 mL
CCu(II) = 5 mg L−1
CNi(II) = 5 mg L−1
CPb(II) = 5 mg L−1
CHg(II) = 5 mg L−1
pH = 6.0
Cu(II) 10.2 mg g−1
Ni(II) 13.5 mg g−1
Pb(II) 21.1 mg g−1
Hg(II) 19.4 mg g−1
[38]
Purolite Arsen Xnp
Lewatit FO36
without EDDS and accompanying ions
m = 2 g
V = 200 mL
CCu(II) = 63 mg L−1
CZn(II) = 65 mg L−1
CCr(VI) = 52 mg L−1
CPb(II) = 207 mg L−1
CCd(II) = 112 mg L−1
pH > 4.5
For Purolite Arsen Xnp
CCu(II) = 5.57 mg g−1
CZn(II) = 4.84 mg g−1
CCr(VI) = 4.77 mg g−1
CPb(II) = 17.14 mg g−1
CCd(II) = 8.20 mg g−1
for Lewatit FO36
CCu(II) = 2.94 mg g−1
CZn(II) = 2.37 mg g−1
CCr(VI) = 5.19 mg g−1
CPb(II) = 15.93 mg g−1
CCd(II) = 3.02 mg g−1
Present study
It can be observed that the adsorption capacity of both ion exchangers is equal and also higher than that of some other sorption materials reported in the literature. For example, Lee et al. [32] found that sorption capacity of the iron oxide nanoparticles immobilized sand for Cd(II) 1.26 mg g−1 and for Pb(II) 2.08 mg g−1, respectively. For the used ion exchangers Lewatit FO36 and Purolite Arsen Xnp, these values are equal to 15.93, 3.02, 17.14, and 8.20 mg g−1, respectively. However, it should be mentioned that in the case of applicability of EDDS, the mechanism of sorption is quite different and connected with the presence of anionic complexes (ion exchange). In the literature data, there is no information concerning the applicability of ethylenediaminedisuccinic acid as a complexing agent for heavy metal ion removal on HIX based on the hydrous iron oxides. EDDS was only used for heavy metal ion removal in the anionic complexes on anion exchangers, as shown in [3941]. The preliminary results for heavy metal ion removal on Lewatit FO36 were also presented in our previous article [42]. Such biodegradable chelating agents as EDDS are commonly used to enhance the heavy metal extraction from the contaminated soils. EDDS causes a much lower potential metal leaching than EDTA, and it is less toxic to soil microorganisms [43].
As was previously mentioned, the sorption of anionic complexes of Cu(II), Zn(II), Cd(II), and Pb(II) with EDDS is connected with the ion exchange process. This mechanism was confirmed by the ATR/FT-IR analysis. The exemplary results for Purolite Arsen Xnp and Lewatit FO36 are presented in Fig. 9. In the recorded spectra, the stretching vibrations at 3290 and 3170 cm−1 are connected with the presence of hydroxyl and amine groups (Lewatit FO36). Other characteristic bands appear at the wavelengths 1636 cm−1 and 1478 cm−1. These are related to the presence of the NH deformation vibration and that of symmetric and asymmetric ν asCOO and ν sCOO. In addition, the bands at 1031 and 1086 cm−1 are connected with the stretching vibrations of CO group. Changing the pH of the solution causes a slight shift of various bands, Ø most noticeably in the case with Purolite Arsenic Xnp.
The AFM images illustrate the changes in the surface of the HIX Purolite Arsen Xnp used in the studies before and after sorption (Fig. 10). No significant changes in the surface morphology were observed before and after the sorption. Based on the observations, it was found that Cr(VI) ions present in the solution penetrate inside the ion exchanger pores according to the reaction:
$$ 2 {\text{R}}-{\text{N}}^{ + } \left( {{\text{CH}}_{ 3} } \right)_{ 3} {\text{Cl}}^{ - } + {\text{ Cr}}_{ 2} {\text{O}}_{ 7}^{ 2- } \rightleftarrows [{\text{R}} - {\text{N}}^{ + } \left( {{\text{CH}}_{ 3} } \right)_{ 3} \left] {_{ 2} } \right[{\text{Cr}}_{ 2} {\text{O}}_{ 7} ]^{ 2- } + {\text{ 2Cl}}^{ - } . $$
Desorption of hybrid sorbents proceeds easily on alkalization of the adsorbent surface. In an alkaline environment, the dominant form on the surface of iron oxides are ≡Fe–O groups, and they adsorb anions in solution. Some ions may, however, remain inside the polymer by reducing the effective sorption capacity by 10–20 %. It has been shown in subsequent regeneration cycles that this effect gets smaller. Regeneration is carried out with a mixture of NaOH and NaCl or NaOH (4–10 %). Depending on conditions and sorption properties of the hybrid material, in the regeneration process, 90–98 % ions are removed. In order to increase the efficiency of regeneration, this process is sometimes carried out at elevated temperatures.

Conclusions

Based on the results obtained, both Purolite Arsen Xnp and Lewatit FO36 may find use in removing heavy metal ions such as Cu(II), Zn(II), Cd(II), and Pb(II) in the presence of EDDS. The interactions of metal ions with the hybrid sorbents are complex, probably simultaneously dominated by adsorption and ion exchange. The pH dependence of ion exchange may suggest that the metal ions are adsorbed according to an ion exchange mechanism. The effect of sulfate ligands on metal adsorption to hybrid sorbents is largely dependent on the surface properties of these materials.
On the basis of the studies, it can be stated that the effectiveness of sorption for Cr(VI), Cu(II), Zn(II), Cd(II), as well as Pb(II) ions depends on the conditions under which the process is carried out. The main parameters affecting sorption are initial concentration of the solution, pH, and phase contact time. Temperature has only a slight influence. With respect to effectiveness of the removal of Cr(VI), Cu(II), Zn(II), Cd(II), and Pb(II), the hybrid sorbents can be arranged as Purolite Arsen Xnp > Lewatit FO36. It should also be emphasized that the best results are obtained for Cr(VI) ions. Therefore, they can be applied in systems of drinking and industrial water purification.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution License which permits any use, distribution, and reproduction in any medium, provided the original author(s) and the source are credited.
Literature
2.
go back to reference Clifford DA, Ghurye GL (2002) Metal-oxide adsorption, ion exchange and coagulation-microfiltration for arsenic removal from water. In: Frankenberger WT (ed) Environmental chemistry of arsenic. Marcel Dekker, New York, pp 217–245 Clifford DA, Ghurye GL (2002) Metal-oxide adsorption, ion exchange and coagulation-microfiltration for arsenic removal from water. In: Frankenberger WT (ed) Environmental chemistry of arsenic. Marcel Dekker, New York, pp 217–245
3.
go back to reference Clifford DA, Ghurye GL, Tripp AR (1999) Development of an anion exchange process for arsenic removal from water. In: Chappell WR, Abernathy CO, Calderon RL (eds) Arsenic exposure and health effects. Elsevier Science, Oxford, pp 379–387CrossRef Clifford DA, Ghurye GL, Tripp AR (1999) Development of an anion exchange process for arsenic removal from water. In: Chappell WR, Abernathy CO, Calderon RL (eds) Arsenic exposure and health effects. Elsevier Science, Oxford, pp 379–387CrossRef
4.
go back to reference Cumbal L, Greenleaf J, Leun D, SenGupta AK (2003) Polymer supported inorganic nanoparticles: characterization and environmental applications. React Funct Polym 54:167–180CrossRef Cumbal L, Greenleaf J, Leun D, SenGupta AK (2003) Polymer supported inorganic nanoparticles: characterization and environmental applications. React Funct Polym 54:167–180CrossRef
5.
go back to reference Cumbal L, SenGupta AK (2005) Arsenic removal using polymer supported hydrated iron(III) oxide nanoparticles: role of Donnan membrane effect. Environ Sci Technol 39:6508–6515CrossRef Cumbal L, SenGupta AK (2005) Arsenic removal using polymer supported hydrated iron(III) oxide nanoparticles: role of Donnan membrane effect. Environ Sci Technol 39:6508–6515CrossRef
6.
go back to reference Elangovan R, Philip L, Chandraraj K (2008) Biosorption of hexavalent and trivalent chromium by palm flower (Borassus aethiopum). Chem Eng J 141:99–111CrossRef Elangovan R, Philip L, Chandraraj K (2008) Biosorption of hexavalent and trivalent chromium by palm flower (Borassus aethiopum). Chem Eng J 141:99–111CrossRef
7.
go back to reference US EPA (2000) Arsenic removal from drinking water by ion exchange and activated alumina plants. Prepared by Battelle under contract 68-C7-0008 for EPA ORD US EPA (2000) Arsenic removal from drinking water by ion exchange and activated alumina plants. Prepared by Battelle under contract 68-C7-0008 for EPA ORD
8.
go back to reference Mohan D, Pittman ChU Jr (2007) Arsenic removal from water/wastewater using adsorbents: a critical review. J Hazard Mater 142:1–53CrossRef Mohan D, Pittman ChU Jr (2007) Arsenic removal from water/wastewater using adsorbents: a critical review. J Hazard Mater 142:1–53CrossRef
9.
go back to reference Hua M, Zhang S, Pan B, Zhang W, Lv L, Zhang Q (2012) Heavy metal removal from water/wastewater by nanosized metal oxides: a review. J Hazard Mater 211–212:317–331CrossRef Hua M, Zhang S, Pan B, Zhang W, Lv L, Zhang Q (2012) Heavy metal removal from water/wastewater by nanosized metal oxides: a review. J Hazard Mater 211–212:317–331CrossRef
10.
go back to reference Kołodyńska D (2011) Organic–inorganic composite materials as sorbents and ion-exchangers used in environmental protection. In: Inamuddin (ed) Advanced organic–inorganic composites: materials, devices and allied applications. Nova Science Publishers, New York, pp 69–130 Kołodyńska D (2011) Organic–inorganic composite materials as sorbents and ion-exchangers used in environmental protection. In: Inamuddin (ed) Advanced organic–inorganic composites: materials, devices and allied applications. Nova Science Publishers, New York, pp 69–130
11.
go back to reference Qiu H, Zhang S, Pan B, Zhang W, Lv L (2012) Effect of sulfate on Cu(II) sorption to polymer-supported nano-iron oxides: behavior and XPS study. J Colloid Interface Sci 366:37–43CrossRef Qiu H, Zhang S, Pan B, Zhang W, Lv L (2012) Effect of sulfate on Cu(II) sorption to polymer-supported nano-iron oxides: behavior and XPS study. J Colloid Interface Sci 366:37–43CrossRef
12.
go back to reference Sarkar S, Blaney LM, Gupta A, Ghosh D, SenGupta AK (2008) Arsenic removal from groundwater and its safe containment in a rural environment: validation of a sustainable approach. Environ Sci Technol 49:4268–4273CrossRef Sarkar S, Blaney LM, Gupta A, Ghosh D, SenGupta AK (2008) Arsenic removal from groundwater and its safe containment in a rural environment: validation of a sustainable approach. Environ Sci Technol 49:4268–4273CrossRef
13.
go back to reference Brochure of Purolite International Ltd. 2011 Brochure of Purolite International Ltd. 2011
14.
go back to reference Lanxess (2007) Engineering information, arsenic separation from ground water using Lewatit FO 36 ion exchange/iron oxide hybrids system, Leverkusen Lanxess (2007) Engineering information, arsenic separation from ground water using Lewatit FO 36 ion exchange/iron oxide hybrids system, Leverkusen
15.
go back to reference Langmuir I (1916) The constitution and fundamental properties of solids and liquids. J Am Chem Soc 38:2221–2295CrossRef Langmuir I (1916) The constitution and fundamental properties of solids and liquids. J Am Chem Soc 38:2221–2295CrossRef
16.
go back to reference Freundlich HFM (1906) Über die adsorption in Lösungen. Z Phys Chem 57:385–470 Freundlich HFM (1906) Über die adsorption in Lösungen. Z Phys Chem 57:385–470
17.
go back to reference Lagergren S (1898) Zur Theorie der sogenannten Adsorption gelöster Stoffe. Kungliga Svenska Vetenskapsakademiens Handlingar 24:1–39 Lagergren S (1898) Zur Theorie der sogenannten Adsorption gelöster Stoffe. Kungliga Svenska Vetenskapsakademiens Handlingar 24:1–39
18.
go back to reference Blanchard G, Maunaye M, Martin G (1984) Removal of heavy metals from waters by means of natural zeolites. Water Res 18:1501–1507CrossRef Blanchard G, Maunaye M, Martin G (1984) Removal of heavy metals from waters by means of natural zeolites. Water Res 18:1501–1507CrossRef
19.
go back to reference Ho YS, McKay G (1998) Sorption of dye from aqueous solution by peat. Chem Eng J 70:115–124 Ho YS, McKay G (1998) Sorption of dye from aqueous solution by peat. Chem Eng J 70:115–124
20.
go back to reference Juang R, Wu W (2002) Adsorption of sulphate and copper(II) on goethite in relation to the changes of zeta potentials. J Colloid Interface Sci 249:22–29CrossRef Juang R, Wu W (2002) Adsorption of sulphate and copper(II) on goethite in relation to the changes of zeta potentials. J Colloid Interface Sci 249:22–29CrossRef
21.
go back to reference Boukhalfa Ch (2010) Sulfate removal from aqueous solutions by hydrous iron oxide in the presence of heavy metals and competitive anions. Macroscopic and spectroscopic analyses. Desalination 250:428–432CrossRef Boukhalfa Ch (2010) Sulfate removal from aqueous solutions by hydrous iron oxide in the presence of heavy metals and competitive anions. Macroscopic and spectroscopic analyses. Desalination 250:428–432CrossRef
22.
go back to reference Chardichai P, Ritchie GSP (1992) The effect of pH on zinc adsorption by a lateritic soil in the presence of citrate and oxalate. J Soil Sci 43:713–728 Chardichai P, Ritchie GSP (1992) The effect of pH on zinc adsorption by a lateritic soil in the presence of citrate and oxalate. J Soil Sci 43:713–728
23.
go back to reference Christophi ChA, Axe L (2000) Competition of Cd, Cu, and Pb adsorption on goethite. J Environ Eng 126:66–74CrossRef Christophi ChA, Axe L (2000) Competition of Cd, Cu, and Pb adsorption on goethite. J Environ Eng 126:66–74CrossRef
24.
go back to reference Jaworska JS, Schowaneck D, Feitzel TCJ (1999) Environmental risk assessment for trisodium [S,S]-ethylene diaminedisuccinate, a biodegradable chelator used in detergent applications. Chemosphere 38:3597–3625CrossRef Jaworska JS, Schowaneck D, Feitzel TCJ (1999) Environmental risk assessment for trisodium [S,S]-ethylene diaminedisuccinate, a biodegradable chelator used in detergent applications. Chemosphere 38:3597–3625CrossRef
25.
go back to reference Loonen H, Lindgren F, Hansen B, Karcher W, Niemelä J, Hiromatsu K, Takatsuki M, Peijnenburg W, Rorije E, Struijs J (1999) Prediction of biodegradability from chemical structure: modeling of ready biodegradation test data. Environ Toxicol Chem 18:1763–1768CrossRef Loonen H, Lindgren F, Hansen B, Karcher W, Niemelä J, Hiromatsu K, Takatsuki M, Peijnenburg W, Rorije E, Struijs J (1999) Prediction of biodegradability from chemical structure: modeling of ready biodegradation test data. Environ Toxicol Chem 18:1763–1768CrossRef
26.
go back to reference Tandy S, Ammann A, Schulin R, Nowack B (2006) Biodegradation and speciation of residual S,S-ethylenediaminedisuccinic acid (EDDS) in soil solution left after soil washing. Environ Pollut 142:191–199CrossRef Tandy S, Ammann A, Schulin R, Nowack B (2006) Biodegradation and speciation of residual S,S-ethylenediaminedisuccinic acid (EDDS) in soil solution left after soil washing. Environ Pollut 142:191–199CrossRef
27.
go back to reference Elliot HA, Huang CP (1981) Adsorption characteristics of some Cu(II) complexes on aluminosilicates. Water Res 15:849–855CrossRef Elliot HA, Huang CP (1981) Adsorption characteristics of some Cu(II) complexes on aluminosilicates. Water Res 15:849–855CrossRef
28.
go back to reference Kołodyńska D (2011) Application of strongly basic anion exchangers for removal of heavy metals in the presence of green chelating agent. Chem Eng J 168:994–1007CrossRef Kołodyńska D (2011) Application of strongly basic anion exchangers for removal of heavy metals in the presence of green chelating agent. Chem Eng J 168:994–1007CrossRef
29.
go back to reference McBride MB (1989) Reactions controlling heavy metal solubility in soils. Adv Soil Sci 10:1–56CrossRef McBride MB (1989) Reactions controlling heavy metal solubility in soils. Adv Soil Sci 10:1–56CrossRef
30.
go back to reference Perelomov LV, Pinskiy DL, Violante A (2011) Effect of organic acids on the adsorption of copper, lead, and zinc by goethite. Eur Soil Sci 44:22–28CrossRef Perelomov LV, Pinskiy DL, Violante A (2011) Effect of organic acids on the adsorption of copper, lead, and zinc by goethite. Eur Soil Sci 44:22–28CrossRef
31.
go back to reference Ali MA, Dzombak DA (1996) Interactions of copper, organic acids, and sulfate in goethite suspensions. Geochim Cosmochim Acta 60:5045–5053CrossRef Ali MA, Dzombak DA (1996) Interactions of copper, organic acids, and sulfate in goethite suspensions. Geochim Cosmochim Acta 60:5045–5053CrossRef
32.
go back to reference Lee SM, Laldawngliana C, Tiwari D (2012) Iron oxide nano-particles-immobilized-sand material in the treatment of Cu(II), Cd(II) and Pb(II) contaminated waste waters. Chem Eng J 195–196:103–111CrossRef Lee SM, Laldawngliana C, Tiwari D (2012) Iron oxide nano-particles-immobilized-sand material in the treatment of Cu(II), Cd(II) and Pb(II) contaminated waste waters. Chem Eng J 195–196:103–111CrossRef
33.
go back to reference Mahmoud ME, Abdelwahab MS, Fathallah EM (2013) Design of novel nano-sorbents based on nano-magnetic iron oxide-bound-nano-silicon oxide-immobilized-triethylenetetramine for implementation in water treatment of heavy metals. Chem Eng J 223:318–327CrossRef Mahmoud ME, Abdelwahab MS, Fathallah EM (2013) Design of novel nano-sorbents based on nano-magnetic iron oxide-bound-nano-silicon oxide-immobilized-triethylenetetramine for implementation in water treatment of heavy metals. Chem Eng J 223:318–327CrossRef
34.
go back to reference Gong J, Chen L, Zeng G, Long F, Deng J, Niu Q, He X (2012) Shellac-coated iron oxide nanoparticles for removal of cadmium(II) ions from aqueous solution. J Environ Sci 24:1165–1173CrossRef Gong J, Chen L, Zeng G, Long F, Deng J, Niu Q, He X (2012) Shellac-coated iron oxide nanoparticles for removal of cadmium(II) ions from aqueous solution. J Environ Sci 24:1165–1173CrossRef
35.
go back to reference Karami H (2013) Heavy metal removal from water by magnetite nanorods. Chem Eng J 219:209–216CrossRef Karami H (2013) Heavy metal removal from water by magnetite nanorods. Chem Eng J 219:209–216CrossRef
36.
go back to reference Tan Y, Chen M, Hao Y (2012) High efficient removal of Pb(II) by amino-functionalized Fe3O4 magnetic nano-particles. Chem Eng J 191:104–111CrossRef Tan Y, Chen M, Hao Y (2012) High efficient removal of Pb(II) by amino-functionalized Fe3O4 magnetic nano-particles. Chem Eng J 191:104–111CrossRef
37.
go back to reference Roy A, Bhattacharya J (2012) Removal of Cu(II), Zn(II) and Pb(II) from water using microwave-assisted synthesized maghemite nanotubes. Chem Eng J 211–212:493–500CrossRef Roy A, Bhattacharya J (2012) Removal of Cu(II), Zn(II) and Pb(II) from water using microwave-assisted synthesized maghemite nanotubes. Chem Eng J 211–212:493–500CrossRef
38.
go back to reference Mahapatra A, Mishra BG, Hota G (2013) Electrospun Fe2O3–Al2O3 nanocomposite fibers as efficient adsorbent for removal of heavy metal ions from aqueous solution. J Hazard Mater 258–259:116–123CrossRef Mahapatra A, Mishra BG, Hota G (2013) Electrospun Fe2O3–Al2O3 nanocomposite fibers as efficient adsorbent for removal of heavy metal ions from aqueous solution. J Hazard Mater 258–259:116–123CrossRef
39.
go back to reference Kołodyńska D (2010) Biodegradable complexing agents as an alternative to chelators in sorption of heavy metal ions. Desalination Water Treat Sci Eng 16:146–155CrossRef Kołodyńska D (2010) Biodegradable complexing agents as an alternative to chelators in sorption of heavy metal ions. Desalination Water Treat Sci Eng 16:146–155CrossRef
40.
go back to reference Kołodyńska D (2011) Green complexing agent EDDS in removal of heavy metal ions on strongly basic anion exchangers. Desalination 280:44–57CrossRef Kołodyńska D (2011) Green complexing agent EDDS in removal of heavy metal ions on strongly basic anion exchangers. Desalination 280:44–57CrossRef
41.
go back to reference Kołodyńska D (2011) The chelating agents of a new generation as an alternative to conventional chelators for heavy metal ions removal from different waste waters. In: RY Ning (ed) Expanding issues in desalination. InTech Publishers, ISBN 978-953-307-326-2, pp 339–371 Kołodyńska D (2011) The chelating agents of a new generation as an alternative to conventional chelators for heavy metal ions removal from different waste waters. In: RY Ning (ed) Expanding issues in desalination. InTech Publishers, ISBN 978-953-307-326-2, pp 339–371
42.
go back to reference Kowalczyk M, Hubicki Z, Kołodyńska D (2013) Modern hybrid sorbents: new ways of heavy metal removal from waters. Chem Eng Process 70:55–65CrossRef Kowalczyk M, Hubicki Z, Kołodyńska D (2013) Modern hybrid sorbents: new ways of heavy metal removal from waters. Chem Eng Process 70:55–65CrossRef
43.
go back to reference Grčman H, Vodnik D, Velikonja-Bolta Š, Leštan D (2003) Ethylenediaminedissuccinate as a new chelate for environmentally safe enhanced lead phytoextraction. J Environ Qual 32:500–506 Grčman H, Vodnik D, Velikonja-Bolta Š, Leštan D (2003) Ethylenediaminedissuccinate as a new chelate for environmentally safe enhanced lead phytoextraction. J Environ Qual 32:500–506
Metadata
Title
Evaluation of iron-based hybrid materials for heavy metal ions removal
Authors
Dorota Kołodyńska
Maciej Kowalczyk
Zbigniew Hubicki
Publication date
01-03-2014
Publisher
Springer US
Published in
Journal of Materials Science / Issue 6/2014
Print ISSN: 0022-2461
Electronic ISSN: 1573-4803
DOI
https://doi.org/10.1007/s10853-013-7944-y

Other articles of this Issue 6/2014

Journal of Materials Science 6/2014 Go to the issue

Premium Partners