Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter April 4, 2015

Applications of chelating resin for heavy metal removal from wastewater

  • Jun Jiang , Xiao-Shuang Ma , Ling-Yun Xu , Li-Hua Wang , Gao-Yan Liu , Qing-Feng Xu EMAIL logo , Jian-Mei Lu EMAIL logo and Yong Zhang
From the journal e-Polymers

Abstract

The chelating resin was synthesized by free-radical copolymerization of iminodiacetic acid modified glycidyl methacrylate with a cross-linker N,N′-methylene biscarylamide at 70°C for removal of heavy metal ions from aqueous solutions. The equilibrium adsorption capacities of the chelating resin from their single-metal ion solutions were 3.28 mmol/g for Cd(II), 2.36 mmol/g for Cu(II), 1.71 mmol/g for Mn(II), 1.69 mmol/g for Ni(II), 1.41 mmol/g for Zn(II), 1.24 mmol/g for Co(II), 0.78 mmol/g for Cr(III) and 0.66 mmol/g for Pb(II). Their related absorption behaviors are discussed in this paper such as thermodynamic equilibrium, pH effect and the Langmuir and Freundlich model to evaluate the experimental data. According to the results, this resin could be used as a promising adsorbent for industrial wastewater disposal.

1 Introduction

The disposal of industrial effluents containing heavy metal ions into natural water resources causes serious damage to the aquatic environment. Even a small amount of heavy metals absorbed by human beings would be hazardous to health (1–3). Therefore, the permitted level of heavy metal ions present in wastewater discharged into public watercourse is very strict (4). Conventional methods for the removal of heavy metal ions from waste streams include chemical precipitation, membrane filtration, ion exchange, carbon adsorption and coprecipitation/adsorption, among which sorption operations such as adsorption and ion exchange are the most popular and effective processes (5, 6). The sorption process offers flexibility in design and operation, and in many cases, the treated water via sorption process is colorless, odorless and suitable for reuse (7). In addition, due to the regeneration of adsorbent in the reversible sorption, the operation is relatively economical (8). But in ion-exchange process, modern ion-exchange materials were prepared from synthetic polymer resins such as styrene-divinylbenzene copolymers, polyacrylate, melamine formaldehyde diethylenetriaminepentaacetic acid (MF-DTPA) and poly(glycidyl methacrylate-iminodiacetic acid) [poly(GMA-IDA)] etc. It should be mentioned that IDA, iminodiacetic acid, as an ideal chelating group applied in ion-exchange resins, has gained considerable attention (9, 10). For example, Wang et al. reported a type of resin with mixed various ratios of IDA (chelating groups) and AA (methacrylic acid, ion-exchange groups) in the copolymer resin. The resin shows good absorption capacity for four common environmentally toxic metal ions: Pb2+, Cu2+, Cr3+ and Cd2+ (11).

However, the emulsion polymerization makes the resin hard to control well when applied in industrial production. In this paper, we induced the monomer IDA into a common radical polymerization process to form a chelating resin. The obtained resin was applied in the absorption of several heavy metal ions including Cu2+, Pb2+, Cd2+, Zn2+, Co2+, Ni2+, Cr3+ and Mn2+. From the reported analogous results, we concluded that this resin showed higher adsorption capacity for Cd2+ and higher selectivity for Zn2+ and Cd2+. The absorption mechanism was also studied in detail.

2 Experimental

2.1 Materials

All materials mentioned in the synthetic process such as iminodiacetic acid, NaOH, glycidyl methacrylate, N,N′-Methylene bisacrylamide (MBA) and 4,4′-azobis (4-cyarovaleric acid) (V501) were all reagent grade and used as-received from the supplier (Chemically pure, Shanghai Chemical Reagent Co. Ltd., Shanghai, China) without further treatments. The heavy metal ions solutions are prepared from deionized water and the following metal salts: CuSO4·5H2O, Pb(NO3)2, CdCl2·2.5H2O, ZnSO4·7H2O, CoCl2·6H2O, NiCl2·6H2O, CrCl3·6H2O and Mn(Ac)2·4H2O.

2.2 Measurement

Fourier transform infrared (FT-IR) measurement was performed with KBr pellets on a Perkin-Elmer 577 FT-IR spectrometer in the range 400–4000 cm-1. A JENWAY 3010 model pH meter was used to measure an expected pH value. Here, the initial and final concentrations of metal ions were determined by using a UNICAM 930 model flame atomic adsorption spectrometer equipped with deuterium lamp background correction, Hallow Cathode Lamp and air acetylene burner. Thermal gravimetric analysis (TGA) was conducted on a Universal V3.7A TA instrument under flowing pure N2 gas (20 ml/min) at a heating rate of 10°C min-1 from room temperature to 800°C. Elemental analyses for C, H and N were done on an EA1110-CHNS elemental analyzer. Scanning electron microscope (SEM) images were taken on a Hitachi S-4700 SEM.

2.3 Synthesis of the resin PIDA

The resin of poly(iminodiacetic acid) (PIDA) was prepared by a cross-linking polymerization as shown in Scheme 1. The monomer was obtained from the reaction of glycidyl methacrylete and sodium 2,2′-azanediyldiacetate under 65°C as reported (11). The polymerization procedure was as follows: the mixture of monomer, MBA and V501, with a molar ratio of 100:10:1 (31.9 g:4.63 g:0.84 g), was stirred at 70°C for 3 h. The obtained polymer was washed with distilled water to remove excess monomers for three times and dried before any other chemical treatment. This experiment can also amplify in kilogram reaction. The ratio of the reactants is similar with the above (3.2 kg:5 kg:1 kg), but the reaction temperature will be 80°C, which is higher to ensure complete polymerization.

Scheme 1: The polymerization process of chelating resin PIDA.
Scheme 1:

The polymerization process of chelating resin PIDA.

2.4 Adsorption experiments

In most cases, metal ion solutions were prepared from the reagents with 3000 mg/l and diluted to appropriate concentration with pH adjustment. All adsorption experiments were carried out batchwise, and the adsorption equilibrium was obtained by stirring the mixture. When the adsorption process was completed, the solution was filtered and the metal ions remaining in solution were measured after dilution. Every experiment was repeated at least three times to ensure the accuracy of data. The amount of metal ion adsorption was calculated from the ratio of value difference between the initial and final concentrations in aqueous solution. The amount of metal ion absorbed at time t, qt was calculated from the mass balance equation

[1]Q=(C0-Ct)VM [1]

where Q is the amount of metal ions adsorbed onto the amount of the resin (mmol g-1), C0 and Ct the concentration of metal ions in the initial and equilibrium concentrations of the metal ions in aqueous phase (mmol l-1), V is the volume of the aqueous phase (L), and M is the dry weight of the resin (g).

3 Results and discussion

3.1 Characterization of the resin

The obtained resin was opaque, and the morphology of the resin was observed by SEM scanning, as shown in Figure 1A. The clearly observed porous surface was one of the main reasons for the large chelating capacity and fast adsorption speed. Moreover, we measured the contact angle of the resin. According to Figure 1B, the small contact angle showed that the resin was ultra-hydrophilic and the water could wet the solid easily.

Figure 1: (A) The SEM photograph of bifunctional chelating resin and (B) contact angle of bifunctional chelating resin PIDA with an 8 μl water droplet.
Figure 1:

(A) The SEM photograph of bifunctional chelating resin and (B) contact angle of bifunctional chelating resin PIDA with an 8 μl water droplet.

The thermal analysis TGA and differential scanning calorimetry of the cross-linking polymer PIDA was measured in the range of 20–600°C. As shown in Figure S1, The PIDA showed a good thermal stability with decomposition temperature close to 200°C. Its glass transition temperature is about 72°C, similar to those analogues such as methyl methacrylate polymer (12).

3.2 Effect of pH Value

It is well known that the pH value of solutions plays an important role in coordination reactions and electrostatic interactions in physical adsorption processes. The hydrophilic groups such as hydroxyl, carboxyl and carbonyl groups in PIDA affect the pH value of its solution. The effect of pH value on metal ions adsorption was studied by changing the initial pH value of the solutions with the addition of dilute hydrochloric acid and potassium hydroxide solution (13). Results are shown in Figure 2; it should be noticed that the adsorption of Pb2+ and Co2+ was improved slightly with increased pH value. According to the results of other metal ions, all adsorption capacities of different metal ions were particularly low at low pH values. It might be attributed to the competitive adsorption between H3O+ ions and metal ions for the limited active adsorption sites in the PIDA resin. However, with the increased pH value, by decreasing the proton concentration, the adsorption capacities increased significantly. When the pH values increased in the appropriate range, the interactions between the resin and metal ions increased, which enhanced the adsorption capacity. The optimum pH values at which the resin showed the maximum adsorption capacity were 5.0 for Cu2+and Cd2+, and 6.0 for Mn2+, Co2+, Ni2+, Cr3+, Pb2+and Zn2+. Compared to the previous reports (14), the optimum pH value of the resin adsorption of Cu2+, Cd2+ and Pb2+ was about 5.0, which was consistent with our result except for the minor difference with the Pb2+. It might be due to the different structures of the two kinds of resin.

Figure 2: The effect of pH on the adsorption of Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III) metal ions for 1 h.
Figure 2:

The effect of pH on the adsorption of Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III) metal ions for 1 h.

3.3 Effect of reaction time

The adsorption capacity for different metal ions was measured as a function of time to determine an optimum time for the adsorption (15). Figure 3 showed the effect of reaction time on the removal of Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III) from solutions by PIDA. The removed metal ions increased and achieved equilibrium after about 30 min. As shown in Figure 3, the adsorption rate was fast in the first 6 min, which was because of the exchange ability of chelating groups (-COO-Na+) in PIDA resin. However, the amount of adsorbed metal ions reached a maximum value and was kept constant in the following period, which meant adsorption equilibrium. It was obvious that the largest capacity was shown on the adsorption of Cd (II). The maximum adsorption capacity of resin for each metal ion is shown in Table 1. From the values listed in Table 1, it was easy to conclude that the resin had much higher affinity to Cd (II) than any other ions. And the adsorption capacity of Cd (II) was higher than the previous report, such as 0.92 mmol/g for S930 (16), 0.004 mmol/g for sporopollenin (17) and 0.035 mmol/g for PAA-PVC (18).

Figure 3: Effect of contact time on the different metal ions (Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III)) using PIDA (initial concentration of metal ions: 3000 ppm (100 ml), amount of resin: 1 g, temperature: 25±1°C, stirring time: 1 h, initial pH: the optimum pH values).
Figure 3:

Effect of contact time on the different metal ions (Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III)) using PIDA (initial concentration of metal ions: 3000 ppm (100 ml), amount of resin: 1 g, temperature: 25±1°C, stirring time: 1 h, initial pH: the optimum pH values).

Table 1

The adsorption capacity of PIDA for metal ions.

Cd(II)Cu(II)Mn(II)Ni(II)Zn(II)Co(II)Cr(III)Pb(II)
Adsorption capacity (mmol/g and mg/g)3.282.361.711.691.411.240.780.66
368.7151.094.199.292.273.140.6136.8

FT-IR spectra for the PIDA resin with and without metal ions adsorbed in were measured to study the adsorption mechanism. Carboxylic acid groups of the resin showed two bands: an asymmetrical stretching band near 1650 cm-1 and a weaker symmetrical stretching band near 1400 cm-1. According to previous reports (19), the shift of adsorption bands for the free carbonyl bond (C=O, 1650 cm-1) could illustrate whether the bonding between the ligand and each metal ion in solid phase was covalent or ionic. The larger the red shift is, the more covalent bonds it has. The covalent bonds are formed by the N and O atoms and metal ions. And the infrared data are shown in Table 2, which are in line with the actual situation. The adsorption band of carbonyl bond (C=O, 1650 cm-1) red-shifted due to the interaction with different ions. According to the results, Cd(II) ion showed the strongest affinity to carboxylic acid via covalent bond, and the affinity of Pd (II) to carboxylic acid was the weakest. The infrared data coincided with the descending orders of adsorption of 1 g resin, Cd2+>Cu2+>Mn2+>Ni2+>Zn2+>Co2+>Cr3+>Pb2+, from former experiments. As a result, the properties of covalent bond between carboxylic acid group in resin and metal ion would be used to explain the difference of adsorption capacity between each metal ion. However, the adsorption mechanism of Cd(II) is different from the S930 resin. In that system, the adsorption band of carbonyl bond (C=O, 1650 cm-1) show small changes after adsorbing Cd(II), which reveal that the N and O atoms in the resin and Cd (II) do not show significantly coordinated behavior. It is because the S930 is a medium with more rigid LEWIS alkali which is unlike PIDA, so the Cd (II) (LEWIS acid) will not be coordinated with N and O atoms in LEWIS alkali. And in our system, from the results of FT-IR spectra, Cd(II) ion showed the strongest affinity to carboxylic acid via covalent bond which was embodied by the larger red shift of adsorption band of carbonyl bond (C=O, 1650 cm-1) after adsorption. Moreover, we have also studied on the reuse of the resin. The PIDA resin could be reused for adsorbing metal ions after adding a dilute aqueous hydrochloric acid solution (1%). The Cd ions were used as an example to do the experiment, and we found that the resin could be used repeatedly for more than eight times in which it had the same amount of adsorption.

Table 2

The infrared stretching bands of C=O in resin before and after adsorption.

PIDACd(II)Cu(II)Mn(II)Ni(II)Zn(II)Co(II)Cr(II)Pb(II)
Asymmetrical stretching band (cm-1)165015871590159515951609162316341640
Symmetrical stretching band (cm-1)140014001400140014001400140014001400

3.4 Effect of initial metal concentration and adsorption isotherms

In fact, the adsorption of metal ions highly depended on the initial metal concentration and the amount of resin. Therefore, we changed the initial concentration of metal ions and kept the amount of resin consistent, to examine the conditions of adsorption equilibrium. The initial metal ion concentrations were 500, 1000, 1500, 2000, 2500 and 3000 ppm, respectively (10 ml), and the weight of the resin was 100 mg in all solutions. The adsorption experiments were carried out in the conditions of optimal pH value, and the reaction time was 1 h. As shown in Figure 4, the metal adsorption capacities significantly increased with the increasing initial metal concentration until it reached the saturated adsorption of adsorbent. The attendant might be due to the high driving force for mass transfer. In other words, for the same amount of adsorbent, the higher the initial concentration is, the faster the adsorption equilibrium reaches.

Figure 4: Effect of initial concentration on adsorption of heavy metal by resin PIDA (amount of resin: 100 mg, temperature: 25±1°C, stirring time: 1 h, initial pH: the optimum pH values).
Figure 4:

Effect of initial concentration on adsorption of heavy metal by resin PIDA (amount of resin: 100 mg, temperature: 25±1°C, stirring time: 1 h, initial pH: the optimum pH values).

The adsorption isotherm is defined as the relation between the concentration of adsorbent in the bulk and the amount of adsorbed metal ions on the interface. Therefore, we need to analyze sorption isotherms to form an equation that can accurately represent the results and could be used for the purpose of design. Herein, two popular adsorption models, Langmuir and Freundlich models, were applied to evaluate the experimental data.

3.5 Chelating isotherm

The non-competitive Langmuir has been a useful tool for the description and comparison of heavy metal sorption by different adsorbents (11, 20). Those isotherms elaborate the relation between the amount of metal taken up per unit weight of adsorbent qe and the equilibrium adsorbate concentration in the bulk fluid phase Ce.

The Langmuir isotherm is given as

1qe=1Q0+1bQ0Ce

where qe is the amount of solute adsorbed on the surface of the adsorbent (mmol g-1), Ce is the equilibrium ion concentration in the solution (mmol l-1), Q0 is the maximum surface density at monolayer coverage and b is the Langmuir adsorption constant (l mmol-1). The plots of 1/qe vs. 1/Ce gave a straight line, and the values of Q0 and b can be calculated from the intercept and slope of the plots, respectively. Experimental data obtained from the effect of initial concentrations on the adsorption of heavy metals on PIDA were also evaluated by applying this equation, and the corresponding constants are given in Table 3.

Table 3

Langmuir isotherm constants of the adsorption of different metal ions by resin.

MetalLangmuir isotherm parameters
Q0(mmol g-1) b(l mmol-1) R2
Cd0.003281.250.9856
Cu0.002366.880.9723
Mn0.001715.450.9534
Ni0.001660.850.9722
Zn0.000143.850.9812
Co0.001183.220.9621
Cr0.000784.020.9735
Pb0.000661.270.9827

3.6 Effect of resin amount on ion exchange

As mentioned above, the adsorption of metal ions highly depended on the amount of resin. Figure 5 shows the removal of heavy metal ions from solution as a function of PIDA resin amount at their optimum pH values. The amount of resin varied from 0.0 to 1.2 g, and the reaction time was 1 h constantly. From the results, it was clear that the removal of metal increased along with the increased amount of resin in the same initial concentration, 3000 ppm. Among these ions, Pb(II)showed the maximum adsorption capacity (75%) by using the smallest amount of resin. This phenomenon may be due to the property of Pb as a heavy metal, the nucleus of which has weaker binding capacity to atomic electrons, so the ion exchange between Pb and resin is faster (21, 22). Similarly, the ion adsorption capacity of Cd reached about 48%, which might be because of the formation of stable bonds between Cd and the resin, leading to adsorption equilibrium (23, 24).

Figure 5: Effect of the quantity of resin on the adsorption equilibrium in PIDA (initial concentration: 3000 ppm, temperature: 25+1°C, stirring time: 1 h, initial pH: the optimum pH values).
Figure 5:

Effect of the quantity of resin on the adsorption equilibrium in PIDA (initial concentration: 3000 ppm, temperature: 25+1°C, stirring time: 1 h, initial pH: the optimum pH values).

Other metal ions such as Zn2+, Mn2+ and Co2+ exhibited similar adsorption rate. However, the metal removal rate slowed down when the amount of resin continued to increase. Since the portion of metal ions removed from the aqueous phase increased as the sorbent dosage increased in the batch vessel with a fixed initial concentration, the curves in Figure 5 approach asymptotic values from 0 to 1.2 g resin. Less time was required to reach the equilibrium adsorption when the amount of sorbent increased. Therefore, we concluded that the adsorbent dosage would increase the amount and rate of removed metal ions but would reduce the efficiency of exchange. This was because the ion exchange capacity of PIDA was not saturated. A small amount of adsorbent which attributes to a lower chelating density would increase the efficiency of the adsorbent (8). However, the speed of adsorption might be reduced. As expected, for an identical initial concentration of metal ions, the equilibrated concentration decreased when the dosage of adsorbent increased. The reason lies on the fact that increased adsorbent provides a greater surface area or chelating sites for a fixed concentration (21, 25, 26).

3.7 Thermodynamics of the adsorption

The effect of temperature on the adsorption of heavy metal ions into the resin PIDA is shown in Figure 6. In a typical experiment, 10 ml of aqueous solution containing metal ions with the initial concentration of 3000 ppm and 100 mg of resin was kept at the optimum pH value. The experimental temperatures are 288, 303, 318 and 333 K, respectively. The effect of temperature on the equilibrium constant of ion exchange between metal and resin was investigated in this stage. From Figure 6, it was observed that the metal ions’ adsorbent capacity increased slightly along with the increasing temperature and showed the endothermic nature of the adsorption. This is due to the exothermic chelating reactions of the metal ions with the resin PIDA. The equilibrium constants in the experiments increased with the increase of temperature. The carboxylic groups in the structure of PIDA are partially protonated at all temperature levels, resulting in only a slight increase in metal adsorption capacities at high temperature. Here, only thermodynamics of ion exchange is involved; the temperature effect on dynamics of adsorption is relatively small. The chelating process is the major mechanism for the removal of metal ions from solution, and the basic process is essentially competitive ionic attraction for the ionic site.

Figure 6: Effect of temperature on heavy metal adsorption using resin PIDA (initial concentration: 3000 ppm (10 ml), amount of resin: 100 mg, stirring time: 1 h, initial pH: the optimum pH values).
Figure 6:

Effect of temperature on heavy metal adsorption using resin PIDA (initial concentration: 3000 ppm (10 ml), amount of resin: 100 mg, stirring time: 1 h, initial pH: the optimum pH values).

4 Conclusions

The chelating resin PIDA was obtained by cross-linking of iminodiacetic acid modified glycidyl methacrylate during radical polymerization. The major functional groups, carboxylate acid group, of the resin were verified by FT-IR spectroscopy. Thermal analysis shows that it has good thermal stability at about 200°C. The resin has good adsorption capacities for eight kinds of metal ions such as Cd(II), Co(II), Cu(II), Ni(II), Pb(II), Mn(II), Zn(II) and Cr(III). Among them, the adsorption capacities of Cd(II) and Cu(II) were higher, which were about 368 and 150 mg metal ions per 1 g resin, respectively. The interaction of metal ions and carboxylate would change from covalent to ionic bond with increased amount of chelating groups contained in resin as indicated by the IR spectroscopy results. In addition, we investigated the optimum pH value and the best reaction time of adsorption. The optimum pH values for maximum adsorption capacity were 5.0 for Cu2+and Cd2+, and 6.0 for Mn2+, Co2+, Ni2+, Cr3+, Pb2+ and Zn2+. The adsorption rates were fast in the first 6 min and leveled off afterwards. From the experimental data, we believe that the chelating resin PIDA can be used as a low-cost adsorbent for removal of many kinds of heavy metal ions in wastewater.


Corresponding authors: Qing-Feng Xu and Jian-Mei Lu, Key Laboratory of Organic Synthesis of Jiangsu Province, College of Chemistry, Chemical Engineering and Materials Science and Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow University, Suzhou 215123, China; and Key Laboratory of Absorbent Materials and Techniques for Environment, Soochow University, 199 Renai Road, Suzhou 215123, China, e-mail: ;

Acknowledgments

This work is supported by Collaborative Innovation Center of Suzhou Nano Science and Technology. The authors gratefully thank the Chinese Natural Science Foundation (No. 20876101 and 21071105), major project of college and universities Jiangsu Province (08KJA430004) and project of environmental protection (SZS0808, 2013020).

References

1. Connell DW, Birkinshaw C, Dwyer TF. Heavy metal adsorbents prepared from the modification of cellulose: A review. Bioresour Technol. 2008;99(15):6709–24.10.1016/j.biortech.2008.01.036Search in Google Scholar

2. Gode F, Pehlivan E. Removal of chromium(III) from aqueous solutions using Lewatit S 100: The effect of pH, time, metal concentration and temperature. J Hazard Mater. 2006;B136(2):330–7.10.1016/j.jhazmat.2005.12.021Search in Google Scholar

3. Baraka A, Hall PJ, Heslop MJ. Preparation and characterization of melamine–formaldehyde–DTPA chelating resin and its use as an adsorbent for heavy metals removal from wastewater. React Funct Polym. 2007;67(7):585.10.1016/j.reactfunctpolym.2007.01.015Search in Google Scholar

4. Low KS, Lee CK, Ng AY. Column study on the sorption of Cr(VI) using quaternized rice hulls. Bioresour Technol. 1999;68(2):205–8.10.1016/S0960-8524(98)00128-XSearch in Google Scholar

5. Bailey ES, Olin TJ, Bricka RM, Adrian DD. A review of potentially low-cost sorbents for heavy metals. Water Res. 1999;33(11):2469–79.10.1016/S0043-1354(98)00475-8Search in Google Scholar

6. Chan WC, Fu TP. Adsorption/ion-exchange behavior between a water-insoluble cationic starch and 2-chlorophenol in aqueous solutions. J Appl Polym Sci. 1998;67(6):1085–92.10.1002/(SICI)1097-4628(19980207)67:6<1085::AID-APP16>3.0.CO;2-#Search in Google Scholar

7. Kantipuly C, Katragadda S, Chow A, Gesser HD. Chelating polymers and related supports for separation and preconcentration of trace metals. Talanta 1990;37(5):491–517.10.1016/0039-9140(90)80075-QSearch in Google Scholar

8. Dakiky M, Khamis M, Manassra A. Selective adsorption of chromium(VI) in industrial wastewater using low-cost abundantly available adsorbents. Adv Environ Res. 2002;6(4):533–40.10.1016/S1093-0191(01)00079-XSearch in Google Scholar

9. Wang CC, Chang CY, Chen CY. Study on metal ion adsorption of bifunctional chelating/ion-exchange resins. Macromol Chem Phys. 2001;202(6):882–90.10.1002/1521-3935(20010301)202:6<882::AID-MACP882>3.0.CO;2-KSearch in Google Scholar

10. Tabushi I, Kobuke Y, Nishiya T. Extraction of uranium from seawater by polymer-bound macrocyclic hexaketone. Nature 1979;208:665–6.10.1038/280665a0Search in Google Scholar

11. Chong KH, Volesky B. Description of two-metal biosorption equilibria by Langmuir-type models. Biotechnol Bioeng. 1995;47(4):451–60.10.1002/bit.260470406Search in Google Scholar

12. Li HM, Liu ZJ, Wei DQ. Polymerization of methyl methacrylate and the measurement of glass transition temperature and the molecular weight. Lab Sci. 2008;49(5):89–91.Search in Google Scholar

13. Namasivayam C, Senthilkumar S. Adsorption of Copper(II) by “Waste” Fe(III)/Cr(III) hydroxide from aqueous solution and radiator manufacturing industry wastewater. Sep Sci Technol. 1999;34(2):201–17.10.1081/SS-100100645Search in Google Scholar

14. Zhang DQ, Kang HY. Detection of Heavy Metal in Soil with Different Digestion Methods. Environ Sci Technol (China). 2003;26(4):4–9.Search in Google Scholar

15. Mohan D, Chander S. Single component and multi-component adsorption of metal ions by activated carbons. Colloids Surf A: Phys Eng Aspects. 2001;177(2–3):183–96.10.1016/S0927-7757(00)00670-1Search in Google Scholar

16. Jin XS, Liu FQ. Ion Exchange Adsorpt. 2010;26(6):481.Search in Google Scholar

17. Unlu N, Ersoz M. Adsorption characteristics of heavy metal ions onto a low cost biopolymeric sorbent from aqueous solutions. J Hazard Mater. 2006;B136(2):272–80.10.1016/j.jhazmat.2005.12.013Search in Google Scholar

18. Liu P, Liu YS. Modification of poly(hydroethyl acrylate)-grafted cross-linked poly(vinyl chloride) particles via surface-initiated atom-transfer radical polymerization (SI-ATRP). Competitive adsorption of some heavy metal ions on modified polymers. Ind Eng Chem Res. 2006;45(7):2255–60.10.1021/ie051247dSearch in Google Scholar

19. Silverstein RM, Bassler GC, Morrill TC. Spectrometric identification of organic compounds. 5th ed. New York: John Wiley & Sons Inc. 1991;23–60.Search in Google Scholar

20. Yu Q, Kaewsarn P. Binary Adsorption of Copper(II) and Cadmium(II) from Aqueous Solutions by Biomass of Marine Alga Durvillaea potatorum. Sep Sci Technol. 1999;34(8):1595–1605.10.1080/01496399909353759Search in Google Scholar

21. Cortina JL, Miralles N, Aguilar M, Saster AM. Distribution studies of Zn(II), Cu(II) and Cd(II) with Levextrel resins containing di(2,4,4-trimethylpentyl)phosphonic acid (Lewatit TP807’84).Hydrometallurgy 1996;40(1–2):195–206.10.1016/0304-386X(94)00078-HSearch in Google Scholar

22. Seco A, Gabaldon C, Marzal P, Aucejo A. Effect of pH, cation concentration and sorbent concentration on cadmium and copper removal by a granular activated carbon. J Chem Technol Biotechnol. 1999;74(9):911–18.10.1002/(SICI)1097-4660(199909)74:9<911::AID-JCTB126>3.0.CO;2-RSearch in Google Scholar

23. Lacour S, Bollinger JC, Serpaud B, Chantron P, Arcos R. Removal of heavy metals in industrial wastewaters by ion-exchanger grafted textiles. Anal Chim Acta. 2001;428(1):121–132.10.1016/S0003-2670(00)01215-0Search in Google Scholar

24. Gode F, Pehlivan E. A comparative study of two chelating ion-exchange resins for the removal of chromium(III) from aqueous solution. J Hazard Mater. 2003;100(1–3):231–43.10.1016/S0304-3894(03)00110-9Search in Google Scholar

25. Rengaraj S, Joo CK, Kim Y, Yi J. Kinetics of removal of chromium from water and electronic process wastewater by ion exchange resins: 1200H, 1500H and IRN97H. J Hazard Mater. 2003; 102(2–3):257–75.10.1016/S0304-3894(03)00209-7Search in Google Scholar

26. Yu LJ, Shukla SS, Dorris KL, Shukla A, Margrave JL. Adsorption of chromium from aqueous solutions by maple sawdust. J Hazard Mater. 2003;B100(1–3):53–63.10.1016/S0304-3894(03)00008-6Search in Google Scholar


Supplemental Material

The online version of this article (DOI: 10.1515/epoly-2014-0192) offers supplementary material, available to authorized users.


Received: 2014-11-3
Accepted: 2015-3-5
Published Online: 2015-4-4
Published in Print: 2015-5-1

©2015 by De Gruyter

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 18.4.2024 from https://www.degruyter.com/document/doi/10.1515/epoly-2014-0192/html
Scroll to top button