Adsorption of uranium(VI) from sulphate solutions ...

6 downloads 14868 Views 853KB Size Report
used in order to present a mathematical description of the equilibrium data at three different ... Various methods have been employed for recovery and removal.
Annals of Nuclear Energy 75 (2015) 132–138

Contents lists available at ScienceDirect

Annals of Nuclear Energy journal homepage: www.elsevier.com/locate/anucene

Adsorption of uranium(VI) from sulphate solutions using Amberlite IRA-402 resin: Equilibrium, kinetics and thermodynamics study Mostafa Solgy, Majid Taghizadeh ⇑, Davood Ghoddocynejad Chemical Engineering Department, Babol University of Technology, P.O. Box 484, 4714871167 Babol, Iran

a r t i c l e

i n f o

Article history: Received 5 April 2014 Received in revised form 29 July 2014 Accepted 5 August 2014

Keywords: Uranium adsorption Isotherms Kinetics Amberlite IRA-402 Sulphate solution

a b s t r a c t In the present study, adsorption of uranium from sulphate solutions was evaluated using Amberlite IRA402 resin. The variation of adsorption process was investigated in batch sorption mode. The parameters studied were pH, contact time and adsorbent dosage. Langmuir and Freundlich isotherm models were used in order to present a mathematical description of the equilibrium data at three different temperatures (25 °C, 35 °C and 45 °C). The final results confirmed that the equilibrium data tend to follow Freundlich isotherm model. The maximum adsorption capacity of Amberlite IRA-402 for uranium(VI) was evaluated to be 213 mg/g for the Langmuir model at 25 °C. The adsorption of uranium on the mentioned anion exchange resin was found to follow the pseudo-second order kinetic model, indicating that chemical adsorption was the rate limiting-step. The values of thermodynamic parameters proved that adsorption process of uranium onto Amberlite IRA-402 resin could be considered endothermic (DH > 0) and spontaneous (DG < 0). Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Among heavy metals, considering their radioactivity and toxicity, uranium is undoubtedly one of the most dangerous to the environment. At different stages of uranium production and consumption in the nuclear industries, large volumes of effluents are produced. Due to environmental risks, the wastewater must be treated before releasing back to the environment (Sun et al., 2010; Nakajima and Sakaguchi, 1999). Various methods have been employed for recovery and removal of uranium from effluents and wastewater. Ultrafiltration (Villalobos-Rodríguez et al., 2012), nanofiltration (Favre-Reguillon et al., 2008), resin-in-pulp (Mirjalili and Roshani, 2007), liquid membrane (Hayworth et al., 1983), biosorption (Raff et al., 2003), liquid–liquid extraction (Krea and Khalaf, 2000), ion exchange (Huikuri and Salonen, 2000), solid phase extraction (Aydin and Soylak, 2007), co-precipitation (Reeder et al., 2001) and electrodeposition (Giridhar et al., 2008) are the most commonly applied methods. Among these methods, ion exchange method employing synthetic resin has received considerable interest compared to others, which is due to its selectivity, less sludge volume, metal recovery at low concentrations from aqueous solution and ability to regenerate the resin (Gode and Pehlivan, 2003). ⇑ Corresponding author. Tel.: +98 111 3234204. E-mail address: [email protected] (M. Taghizadeh). http://dx.doi.org/10.1016/j.anucene.2014.08.009 0306-4549/Ó 2014 Elsevier Ltd. All rights reserved.

Uranium found in the environment is usually in the hexavalent form. In industrial effluents, uranium is mainly capable of forming anionic complexes such as uranyl fluoride, uranyl carbonate, uranyl phosphate, uranyl chloride or uranyl sulphate. Due to the presence of this anionic complex, the strong base anion exchange resins are considered to be the most suitable choice for uranium recovery. Anion exchange resins have been successfully employed to recover uranium in mining industry, especially from leach liquor (Zhang and Clifford, 1994; Song et al., 1999; Chellam and Clifford, 2002). The results of a recent research have shown that adsorption of uranium from phosphate solution onto a strongly basic anion exchange resin (Amberlite CG-400) is favourable and the maximum adsorption capacity of 112.36 mg/g was obtained (Semnani et al., 2012). Another study of uranium adsorption by Amberlite IRA-910 anion exchange resin showed that at 25 °C, the maximum adsorption capacity of uranium was 64.26 mg/g (Rahmati et al., 2012). The present study focuses on adsorption of uranium from sulphate solution using Amberlite IRA-402 resin. The experiments were carried out in a batch system. The influence of different variables such as pH, contact time and adsorbent dosage on the adsorption capacity of Amberlite IRA-402 resin was investigated. Kinetic studies have been performed in which the parameters were determined at 25 °C. Furthermore, equilibrium and thermodynamic parameters of uranium(VI) adsorption were calculated at three different temperatures.

133

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138 Table 1 Characteristics properties of the ion exchange resin. Resin Physical form Matrix Functional group Ionic form as shipped Total exchange capacity (eq/L) Moisture holding capacity (%) Shipping weight (g/L) Particle size (mm) Uniformity coefficient Reversible swelling (%) Maximum operating temperature (°C)

Amberlite IRA-402 Pale yellow translucent spherical beads Styrene divinylbenzene copolymer Trimethyl ammonium Cl P1.20 (Cl form) 49–60 (Cl form) 670 0.600–0.750 61.6 Cl ? OH 6 30 60

Each experiment was carried out three times and the average results were reported. The relative standard deviation values of the results were relatively low and of the order of ±2.5%, which shows the good reproducibility and accuracy of the experiments. The adsorption rate percentage, the adsorption capacity (q, mg/g) and the distribution coefficient (Kd, mL/g) were calculated using the following equations:

Adsorption rate ð%Þ ¼

ðC i  C e Þ  100 Ci

ð1Þ

q ðmg=gÞ ¼ ðC i  C e Þ 

V m

ð2Þ

K d ðmL=gÞ ¼ Table 2 ICP–OES operational parameters used for the analysis of uranium. Parameter

Setting

Wavelength (k) RF power Nebulizer Plasma flow Auxiliary flow Nebulizer flow Sample flow Injector Spray chamber

385.958 nm 1400 W Gemcone 15 L/min 0.6 L/min 0.8 L/min 1.0 mL/min 2.0 mm Alumina Scott double-pass

2. Materials and methods 2.1. Materials All chemicals used in the experiments were of analytical reagent (AR) grade and obtained from Merck (Darmstadt, Germany). The physical properties and characteristics of Amberlite IRA-402 anion exchange resin which was used in this study are presented in Table 1. A stock solution of 1000 mg/L was prepared by dissolving 2.151 g of (UO2(NO3)26H2O) in double distilled water. Uranium working solutions were made by diluting this stock solution to the desired concentration. Sodium sulphate (Na2SO4) was used as the source of sulphate. The initial pH of the solutions was adjusted by adding 0.1 mol/L hydrochloric acid (HCl) and 0.1 mol/L sodium hydroxide (NaOH).

ðC i  C e Þ V  Ce m

ð3Þ

where Ci is initial concentration of uranium (mg/L), Ce is equilibrium concentration of uranium after adsorption (mg/L), V is volume of uranium aqueous solution contacted with adsorbent (L), and m is the amount of adsorbent (g). 3. Result and discussion 3.1. Effect of pH The uranium(VI) adsorption efficiency using Amberlite IRA-402 resin found to be strongly dependent on variation of pH of the solution. Therefore, the uranium adsorption experiments were carried out at different pH levels (1.5, 2, 2.5, 2.75. 3, 3.5, 4). The results of uranium adsorption onto Amberlite IRA-402 resin at different pH levels are illustrated in Fig. 1. As can be seen in Fig. 1, the maximum uptake value is observed at pH = 3. In acidic solution and at pH levels less than 2, the predominant species seem to be UO2 2þ ions. Although, the presence of high contents of sulphate in the solution enhances formation of complexes such as UO2 ðSO4 Þ2 2 and UO2 ðSO4 Þ3 4 (Nascimento et al., 2004). Eqs. (4)–(6) indicate the equilibrium state between uranium and sulphate ions in the solution. The constants K1, K2 and K3 are called stability constants. In a pH range of 3–4 no significant change in adsorption of uranium complexation was noticed. Increasing pH level above 4.0 would lead to uranium precipitation which is the reason why experiments were not performed at values higher than 4 (Ladeira and Gonçalves, 2007).

UO2 2þ þ SO4 2 $ UO2 ðSO4 Þ K 1 ¼ 50

ð4Þ

2.2. Batch adsorption experiments 100

80

Adsorption rate (%)

In order to determine the optimal values for pH, adsorbent dosage and contact time, the adsorption experiments were performed at ambient temperature (25 °C) in a batch system. Only one of the parameters was changed at a time while others were kept constant during the experiments. In all experiments 25 mL of uranium solution was poured into stoppered Pyrex glass flasks in the presence of 3000 mg/L sulphate anion and then the flasks were placed inside a mechanical shaker where the agitation speed was set at 300 rpm. The adsorbent and solution were separated by filtration through a filter paper. After each experiment, the concentration of uranium ions in the solution was determined by inductively coupled plasma–optical emission spectrometry (ICP–OES). The measurements were performed using a Perkin Elmer Optima 2000 DV ICP–OES according to ASTM C1109. The calibration range was between 1 mg/L and 1000 mg/L. Samples over the range of the standards were diluted to be within range. The operating conditions for the ICP–OES were shown in Table 2. Furthermore, all standards, control samples, and samples were measured in triplicate for statistical purposes.

60

40

20

0 1

2

3

4

5

pH Fig. 1. The effect of pH on the adsorption of uranium(VI) onto Amberlite IRA-402. T = 25 °C, Time = 120 min, Ci = 100 mg/L, and mass of resin = 50 mg.

134

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138

100

2 200 mg/L 1

60

60

40

40

20

20

log (qe-qt)

80

80

q (mg/g)

Adsoption rate (%)

100

100 mg/L

0 -1 -2

0

-3 -4

0 0

1

2

3

4

0

5

100

200

300

400

Time (min)

Absorbent dosage (g/L) Fig. 2. Effect of the adsorbent dose on uranium(VI) sorption by Amberlite IRA-402 resin. pH = 3, T = 25 °C, Time = 120 min, and Ci = 100 mg/L.

Fig. 4. Pseudo-first order kinetics of uranium(VI) adsorption onto Amberlite IRA402 resin. pH = 3, T = 25 °C, and mass of resin = 50 mg.

100

ent time intervals up to 360 min. In the present step, the experiments were conducted under optimum conditions and at two different initial concentrations of uranium (100 and 200 mg/L). From Fig. 3, it could be observed that uranium adsorption percentage grew rapidly by increasing contact time and then reached the saturation point in 100 min. After that, no significant change in adsorption of uranium(VI) was noticed. It should be mentioned that the data obtained in this step was used further for kinetic studies.

60

40

20

3.4. Adsorption kinetics study

100 mg/L 200 mg/L

0 0

50

100

150

200

250

300

350

400

Time (min) Fig. 3. Effect of contact time on the adsorption of uranium onto Amberlite IRA-402 resin. pH = 3, T = 25 °C, and mass of resin = 50 mg.

UO2 ðSO4 Þ þ SO4

2

$ UO2 ðSO4 Þ2

2

K 2 ¼ 350

UO2 ðSO4 Þ2 2 þ SO4 2 $ UO2 ðSO4 Þ3 4

K 3 ¼ 3500

ð5Þ ð6Þ

3.2. Effect of adsorbent dosage The relation between adsorbent dosage and adsorption efficiency has also been investigated. The volume of solution (25 mL), the concentration of uranium (100 mg/L), the pH of solution (3) and the time of adsorption (120 min) were kept constant while the amount of resin varied from 0.5 to 4 g/L. As shown in Fig. 2, by increasing the amount of adsorbent the adsorption percentage of uranium increases. However, amount of uranium adsorption per unit weight of adsorbent declines. The increase observed in adsorption due to increasing the amount of adsorbent could be attributed to the increase in surface area and the availability of adsorption sites (Fan et al., 2012). While at a constant concentration, increasing adsorbent leads to more unsaturated sites which could be the reason why the amount of uranium adsorption per unit weight of adsorbent decreases.

To determine the kinetics of uranium adsorption on Amberlite IRA-402 resin, the kinetics experimental data were simulated using pseudo-first order and pseudo-second order models. The linear form of pseudo-first order model can be represented by following equation (Ho and McKay, 1999):

logðqe  qt Þ ¼ logðqe Þ 

k1 t 2:303

ð7Þ

where k1 represents the pseudo-first order adsorption rate constant (1/min), qe (mg/g) is the adsorption capacity of uranium at equilibrium, t is the time (min) and qt is the amount of adsorbed uranium at time t (mg/g). k1 could be obtained by plotting log(qeqt) versus t (Fig. 4).

9 8 7

t/qt (min g/mg)

Adsorption rate (%)

80

6 5 4 3 2

100 mg/L

1

200 mg/L

0

3.3. Effect of contact time Fig. 3 provides data on the effect of contact time on batch adsorption of uranium(VI) onto Amberlite IRA-402 resin at differ-

0

100

200

300

400

Time (min) Fig. 5. Pseudo-second order kinetics of uranium(VI) adsorption onto Amberlite IRA402 resin. pH = 3, T = 25 °C, and mass of resin = 50 mg.

135

200

0.5

160

0.4

Ce/qe (g/L)

qe (mg/g)

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138

120

80

40

0.3

0.2

45 C 35 C 25 C

0.1

45 C 35 C 25 C

0

0

0

20

40

60

80

100

0

20

C e (mg/L)

40

60

80

100

Ce (mg/L)

Fig. 6. Adsorption isotherm for uranium(VI) adsorption onto Amberlite IRA-402 resin at different temperatures.

Fig. 7. Langmuir adsorption isotherms for uranium(VI) adsorption onto Amberlite IRA-402 resin at different temperatures.

The linear pseudo-second-order kinetic model could be shown

5.5

as:

t 1 1 ¼ þ t qt k2 q2e qe

5

where k2 is the pseudo-second order adsorption rate constant (g/mg min) (Ho, 2006). k2 can be determined from the slope of plot t/qt versus t (Fig. 5). The parameter values of the pseudo-first order and pseudo second order models at two initial concentrations of uranium (100 and 200 mg/L) are presented in Table 3. As illustrated in Table 2, the regression correlation coefficients (R2) at two initial concentrations of uranium for pseudo-firs order seem to be lower and experimental qe does not agree with calculated qe, on the other hand, for pseudo-second order model at the two initial concentrations of uranium, the values for regression correlation coefficients coefficient are equal to 0.999 and qe,cal (theoretical equilibrium adsorption capacity) shows good agreement with qe,exp (experimental equilibrium adsorption capacity). According to the results, uranium adsorption follows pseudo-second order kinetics model and the rate-limiting step in the process might be the chemical adsorption (Wu et al., 2001).

4.5

3.5. Adsorption isotherm study Adsorption isotherms are most commonly used to select adsorbent and also to study methods of equilibrium adsorption data. They are of great importance for practical design of adsorption systems (Yusan and Akyil Erenturk, 2010). For the present purpose, the equilibrium experiments were conducted at three temperatures of 25 °C, 35 °C, and 45 °C. The results of the tests are shown in Fig. 6. In order to analyze the experimental equilibrium adsorption data in the present study, the Langmuir and Freundlich isoterm models were applied. Langmuir adsorption isotherm assumes that adsorption takes place on a surface that is energetically

ln (qe)

ð8Þ

4

3.5

45 C 35 C 25 C

3 0

1

2

3

4

5

ln (C e) Fig. 8. Freundlich adsorption isotherms for uranium(VI) adsorption onto Amberlite IRA-402 resin at different temperatures.

homogeneous, and there is no interaction between neighbouring adsorbed molecules on the surface of adsorbent (Langmuir, 1918), while the Freundlich model assumes that adsorption occurs on a heterogeneous surface (Freundlich, 1906). The linear form of Langmuir and Freundlich isotherms are presented by Eqs. (9) and (10), respectively:

Ce 1 Ce ¼ þ qe bqmax qmax ln qe ¼ ln K F þ

ð9Þ

1 ln C e n

ð10Þ

where Ce is the equilibrium concentration of uranium in the solution (mg/L), qe is the amount of uranium adsorbed per weight unit of resin at equilibrium time (mg/g), qmax is the saturated monolayer

Table 3 Values of pseudo-first order and pseudo second order kinetic constants of uranium(VI) adsorbed onto Amberlite IRA-resin. pH = 3, T = 25 °C, and mass of resin = 50 mg. Ci (mg/L)

100 200 a

Pseudo-first-order rate model

Pseudo-second-order rate model

k1 (1/min)

qe,exp (mg/g)

qe,cal (mg/g)

R2

SDa

k2 (g/mg min)

qe,cal (mg/g)

R2

SD

0.03 0.022

43 74.5

25.2 24.19

0.929 0.851

0.06 0.07

0.0039 0.0017

43.85 76.33

0.999 0.999

0.04 0.03

Standard deviation.

136

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138

Table 4 Langmuir and Freundlich isotherm constants for uranium(VI) adsorption onto Amberlite IRA-402 resin. Temperature (°C)

25 35 45

Langmuir constants

Freundlich constants 2

qmax (mg/g)

b (L/mg)

R

SD

n

KF (mg/g) (L/mg)1/n

R2

SD

213 200 172

0.05 0.08 0.06

0.989 0.980 0.959

0.01 0.02 0.03

2.15 2.71 3.00

23.80 36.97 44.70

0.993 0.999 0.999

0.02 0.004 0.006

Table 5 Comparison of adsorption capacities of various adsorbents used for uranium adsorption. Adsorbent

qmax (mg/g)

References

APM Chitosan/amine resin RIa-mag AXAD-16-3,4-dihydroxybenzoyl methyl phosphonic acid PANSIL at pH = 7.5 AXAD-16 functionalized with (bis-2,3,4-trihydroxy benzyl) ethylenediamine Chitosan/azole resin PANSIL at pH = 6 Merrifield chloromethylated Amberlite XAD-16 Ria Doulite A101 XAD-16 functionalized with (bis-3,4-dihydroxy benzyl)p-phenylene diamine Merrifield polymer-N,N,N,N-tetrahexyl malonamide Polystyrene–divinylbenzene resin containing 1-(2-thiazolylazo)-2-naphthol Amberlite XAD-4-Bicene Amberlite XAD-4-octa carboxy methyl-c-methyl Calix resorcinarene Merrifield polymer-TTA PANSIL at pH = 4.5 Amberlite XAD-4 functionalized with succinic acid Amberlite XAD-2-tiron Amberlite XAD-4-OVSC Amberlite IRA-910 Amberlite CG-400 AK-1 and AK-2 Cross-linked chitosan (CCTS) Talc Amberlite IRA-402

24.276 428.4 399.84 395.08 357 340.34 309.4 238 228.48 214.2 195.16 161.84 159.46 154.7 97.58 90.44 64.26 33.32 95.2 11.9 7.14 2.38 64.26 112.36 10.5 72.46 41.49 213

Hazer and Kartal (2009) Atia (2005) Donia et al. (2009) Maheswari and Subramanian (2005) Barton et al. (2004) Prabhakaran and Subramanian (2003a) Atia (2005) Barton et al. (2004) Raju and Subramanian (2007) Merdivan et al. (2001) Donia et al. (2009) Khalifa (1998) Prabhakaran and Subramanian (2003b) Raju and Subramanian (2005) Lee et al. (2001) Metilda et al. (2005) Jain et al. (2001) Prabhakaran and Subramanian (2004) Barton et al. (2004) Dev et al. (1999) Kumar et al. (2000) Merdivan et al. (2001) Rahmati et al. (2012) Semnani et al. (2012) Yusan and Akyil Erenturk (2010) Wang et al. (2009) Sprynskyy et al. (2011) This work

Table 7 Thermodynamics parameters for uranium(VI) adsorption on Amberlite IRA-402 resin at 250 mg/L initial uranium concentration.

Table 6 Values of RL at three different temperatures. Temperature (°C)

Ci (mg/L)

RL

25 35 45

50–250 50–250 50–250

0.07–0.28 0.04–0.2 0.25–0.06

Thermodynamics parameters

7.3

lnKd = −1122.9/T + 10.795 R² = 0.9961

7.25 7.2

lnK d

7.15 7.1 7.05 7 6.95 6.9 0.0031

0.00315

0.0032

0.00325

1/T

0.0033

0.00335

0.0034

(K −1 )

Fig. 9. Plot of LnKd versus 1/T for uranium sorption onto Amberlite IRA-402 resin.

T (°C)

DGo (kJ/mol)

DHo (kJ/mol)

DSo (J/mol K)

25 35 45

17.41 18.34 19.2

9.33 – –

89.75 – –

adsorption capacity (mg/g) and b is the Langmuir constant (L/mg). KF ((mg/g) (L/mg)1/n) and n are the Freundlich constants related to the adsorption capacity and adsorption intensity, respectively. The Langmuir and Freundlich plots at three temperatures (25 °C, 35 °C, and 45 °C) are presented in Figs. 7 and 8, respectively and their parameters and regression correlation coefficients are given in Table 4. As presented in Table 4, at all tested temperatures the Freundlich model gave slightly better fittings than the Langmuir model. The maximum uranium adsorption capacity of the Amberlite IRA-402 resin in the presence of sulphate anion was 213 mg/g. Comparisons between maximum uptake capacities (qm) of Amberlite IRA-402 and other adsorbents for uranium(VI) reported in the literature are presented in Table 5. The result shows that Amberlite IRA-402 exhibits a reasonable capacity for uranium adsorption from aqueous solutions. Also, the values of n being larger than 1, shows the favourable nature of uranium adsorption onto Amberlite IRA-402 resin.

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138

The essential characteristic parameter of the Langmuir isotherm that indicates the type of isotherm, is the dimensionless factor RL. The favourable adsorption will occur when RL value is in the range of 0–1 (Webi and Chakravort, 1974). This parameter is defined by the following equation:

RL ¼

1 1 þ bC i

ð11Þ

where Ci is the initial uranium concentration (mg/L) and b is the Langmuir constant (L/mg). As shown in Table 6, at all temperatures RL is in the range of 0–1, which indicates that the adsorption of uranium on the Amberlite IRA-402 is favourable. 3.6. Thermodynamics study The adsorption experiments were carried out at constant initial uranium concentration of 250 mg/L and three different temperatures (25 °C, 35 °C, and 45 °C). The adsorption standard free energy (DGo) was determined using the following equation:

DGo ¼ DHo  T DSo o

ð12Þ o

where DH is the enthalpy of adsorption, DS is the entropy change and T is the temperature (K). From the slope and intercept of the Van’t Hoff’s equation which is shown in Eq. (13), DHo and DSo can be calculated (Fig. 9).

ln K d ¼

DSo DH  RT R

ð13Þ

where R is universal gas constant (8.314 J/mol K) and T is the absolute temperature (K). The values of thermodynamics parameters are given in Table 7. The positive value of DHo indicates that the adsorption of uranium on the Amberlite IRA-402 is endothermic. In addition, the negative standard free energy and the positive standard entropy indicate that the adsorption reaction is a spontaneous process and is more favourable at higher temperatures. 4. Conclusions The present work confirmed that uranium(VI) adsorption onto amberlite IRA-402 resin in the presence of sulphate anions is affected by different parameters (pH, adsorbent dosage, contact time). The adsorptions were rapid during the first 30 min, and reached equilibrium in 100 min. The adsorption kinetics can be well described by the pseudo-second-order model, indicating that chemical adsorption was the rate limiting-step. Equilibrium studies showed that uranium adsorption process follows the Freundlich model and the maximum adsorption capacity obtained by the Langmuir equation is 213 mg/g. Uranium adsorption onto Amberlite IRA-402 resin in the presence of sulphate is a favourable process at higher levels of temperature during the experiments (35 °C and 45 °C). Furthermore it is an endothermic and spontaneous process. Finally, it can be concluded that Amberlite IRA-402 anion exchange resin is a suitable candidate for recovery of anionic complexes of uranyl sulphate from aqueous solution. References Atia, A.A., 2005. Studies on the interaction of mercury(II) and uranyl(II) with modified chitosan resins. Hydrometallurgy 80, 13–22. Aydin, F.A., Soylak, M., 2007. Solid phase extraction and preconcentration of uranium(VI) and thorium (IV) on Duolite XAD761 prior to their inductively coupled plasma mass spectrometric determination. Talanta 72, 187–192. Barton, C.S., Stewart, D.I., Morris, K., Bryant, D.E., 2004. Performance of three resin based materials for treating uranium contaminated groundwater within a PRB. J. Hazard. Mater. 116, 191–204. Chellam, S., Clifford, D.A., 2002. Physical–chemical treatment of groundwater

137

contaminated by leachate from surface disposal of uranium tailings. J. Environ. Eng. ASCE 128, 942–952. Dev, K., Pathak, R., Rao, G.N., 1999. Sorption behaviour of lanthanum(III), neodymium (III), terbium(III), thorium(IV) and uranium(VI) on Amberlite XAD-4 resin functionalized with bicine ligands. Talanta 48, 579–584. Donia, A.M., Atia, A.A., Moussa, E.M.M., El-Sherif, A.M., Abd El-Magied, M.O., 2009. Removal of uranium(VI) from aqueous solutions using glycidyl methacrylate chelating resins. Hydrometallurgy 95, 183–189. Fan, F.L., Qin, Z., Bai, J., Rong, W.D., Fan, F.Y., Tian, W., Zhao, L., 2012. Rapid removal of uranium from aqueous solutions using magnetic Fe3O4@SiO2 composite particles. J. Environ. Radioact. 106, 40–46. Favre-Reguillon, A., Lebuzit, G., Murat, D., Foos, J., Mansour, C., Draye, M., 2008. Selective removal of dissolved uranium in drinking water by nanofiltration. Water Res. 42, 1160–1166. Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys. Chem. 57, 292. Giridhar, P., Venkatesan, K.A., Subramaniam, S., Srinivasan, T.G., Vasudeva Rao, P.R., 2008. Extraction of uranium(VI) by 1.1 M tri-n-butyl phosphate/ionic liquid and the feasibility of recovery by direct electrodeposition from organic phase. J. Alloys Compd. 448, 104–108. Gode, F., Pehlivan, E., 2003. A comparative study of two chelating ion-exchange resins for the removal of chromium (III) from aqueous solution. J. Hazard. Mater. 100, 231–243. Hayworth, H.C., Ho, W.S., Burns Jr, W.A., Li, N.N., 1983. Extraction of uranium from wet process phosphoric acid by liquid membranes. Sep. Sci. Technol. 18, 493– 521. Hazer, O., Kartal, S., 2009. Synthesis of a novel chelating resin for the separation and preconcentration of uranium(VI) and its spectrophotometric determination. Anal. Sci. 25, 547–551. Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem. 34, 451–465. Ho, Y.S., 2006. Review of second-order models for adsorption systems. J. Hazard. Mater. 136, 681–689. Huikuri, P., Salonen, L., 2000. Removal of uranium from Finnish groundwaters in domestic use with a strong base anion resin. J. Radioanal. Nucl. Chem. 245, 385– 393. Jain, V.K., Handa, A., Sait, S.S., Shrivastav, P., Agarval, Y.K., 2001. Preconcentration, separation and trace determination of lanthanum(III), cerium(III), thorium(IV) and uranium(VI) on polymer supported ovanillinsemicarbazone. Anal. Chim. Acta 429, 237–246. Khalifa, M.E., 1998. Selective separation of uranium using alizarin red S (ARS) – modified anion-exchange resin or by flotation of U-ARS chelate. Sep. Sci. Technol. 33, 2123–2141. Krea, M., Khalaf, H., 2000. Liquid–liquid extraction of uranium and lanthanides from phosphoric acid using a synergistic DOPPA–TOPO mixture. Hydrometallurgy 58, 215–225. Kumar, M., Rathore, D.P.S., Singh, A.K., 2000. Metal ion enrichment with Amberlite XAD-2 functionalized with tiron: analytical applications. Analyst 125, 1221– 1226. Ladeira, A.C.Q., Gonçalves, C.R., 2007. Influence of anionic species on uranium separation from acid mine water using strong base resins. J. Hazard. Mater. 148, 499–504. Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 40, 1361–1403. Lee, W., Lee, S., Lee, C., Kim, Y., Lee, Y., 2001. A chelating resin containing 1(2thiazolyazo)-2-naphthol as the functional group; synthesis and sorption behavior for trace metal ions. Microchem. J. 70, 195–203. Maheswari, M.A., Subramanian, M.S., 2005. AXAD-16-3, 4-dihydroxy benzoyl methyl phosphonic acid: a selective preconcentrator for U and Th from acidic waste streams and environmental samples. React. Funct. Polym. 62, 105–114. Merdivan, M., Duz, M.Z., Hamamci, C., 2001. Sorption behaviour of uranium(VI) with N, N-dibutyl-N-benzoylthiourea Impregnated in Amberlite XAD-16. Talanta 55, 639–645. Metilda, P., Sanghamitra, K., Gladis, J.M., Naidu, G.R.K., Prasada Rao, T., 2005. Amberlite XAD-4 functionalized with succinic acid for the solid phase extractive preconcentration and separation of uranium(VI). Talanta 65, 192–200. Mirjalili, K., Roshani, M., 2007. Resin-in-pulp method for uranium recovery from leached pulp of low grade uranium ore. Hydrometallurgy 85, 103–109. Nakajima, A., Sakaguchi, T., 1999. Recovery of uranium from uranium refining waste water by using immobilized persimmon tannin. J. Radioanal. Nucl. Chem. 242, 623–626. Nascimento, M.R.L., Fatibello-Filho, O., Teixeira, L.A., 2004. Recovery of uranium from acid mine drainage waters by ion exchange. Miner. Process. Extr. Metall. 25, 129–142. Prabhakaran, D., Subramanian, M.S., 2003a. A column system for the selective extraction of U(VI) and Th(IV) using a new chelating sorbent. Talanta 61, 423– 430. Prabhakaran, D., Subramanian, M.S., 2003b. Selective extraction and sequential separation of e and transition ions using AXAD-16-BTBED polymeric sorbent. React. Funct. Polym. 57, 147–155. Prabhakaran, D., Subramanian, M.S., 2004. Extraction of U(VI), Th(IV), and La from acidic streams and geological samples using AXAD-16-POPDE polymer. Anal. Bioanal. Chem. 380, 578–585. Rahmati, A., Ghaemi, A., Samadfam, M., 2012. Kinetic and thermodynamic studies of uranium (VI) adsorption using Amberlite IRA-910 resin. Ann. Nucl. Energy 39, 42–48.

138

M. Solgy et al. / Annals of Nuclear Energy 75 (2015) 132–138

Raju, K.Ch., Subramanian, M.S., 2005. Selective preconcentration of U(VI) and Th(IV) in trace and macroscopic levels using malonamide grafted polymer from acidic matrices. Microchim. Acta 150, 297–304. Raju, K.Ch., Subramanian, M.S., 2007. Sequential separation of lanthanides, thorium and uranium using novel solid phase extraction method from high acidic nuclear wastes. J. Hazard. Mater. 145, 315–322. Raff, J., Soltmann, U., Matys, S., Selenska-Pobell, S., Böttcher, H., Pompe, W., 2003. Biosorption of uranium and copper by biocers. Chem. Mater. 15, 240–244. Reeder, R.J., Nugent, M., Tait, C.D., Morris, D.E., Heald, S.M., Beck, K.M., Lanzirotti, A., 2001. Coprecipitation of uranium (VI) with calcite: XAFS, micro-XAS, and luminescence characterization. Geochim. Cosmochim. Acta 65, 3491–3503. Semnani, F., Asadi, Z., Samadfam, M., Sepehrian, H., 2012. Uranium (VI) sorption behavior onto Amberlite CG-400 anion exchange resin: effects of pH, contact time, temperature and presence of phosphate. Ann. Nucl. Energy 48, 21–24. Song, Y., Wang, Y., Wang, L., Song, C., Yang, Z.Z., Zhao, A., 1999. Recovery of uranium from carbonate solutions using strongly basic anion exchanger 4. Column operation and quantitative analysis. React. Funct. Polym. 39, 245–252. Sprynskyy, M., Kowalkowski, T., Tutu, H., Cukrowska, E.M., Buszewski, B., 2011. Adsorption performance of talc for uranium removal from aqueous solution. Chem. Eng. J. 171, 1185–1193.

Sun, X., Huang, X., Liao, X.P., Shi, B., 2010. Adsorptive recovery of UO2+ 2 from aqueous solutions using collagen–tannin resin. J. Hazard. Mater. 179, 295–302. Villalobos-Rodríguez, R., Montero-Cabrera, M.E., Esparza-Ponce, H.E., HerreraPeraza, E.F., Ballinas-Casarrubias, M.L., 2012. Uranium removal from water using cellulose triacetate membranes added with activated carbon. Appl. Radiat. Isotopes 70, 872–881. Wang, G., Liu, J., Wang, X., Xie, Z., Deng, N., 2009. Adsorption of uranium (VI) from aqueous solution onto cross-linked chitosan. J. Hazard. Mater. 168 (2), 1053– 1058. Webi, T.W., Chakravort, R.K., 1974. Pore and solid diffusion models for fixed-bed adsorbers. AIChE J. 20, 228–238. Wu, F.C., Tseng, R.L., Juang, R.S., 2001. Enhanced abilities of highly swollen chitosan beads for color removal and tyrosinase immobilization. J. Hazard. Mater. 81, 167–177. Yusan, S., Akyil Erenturk, S., 2010. Adsorption equilibrium and kinetics of U (VI) on beta type of akaganeite. Desalination 263, 233–239. Zhang, Z., Clifford, D.A., 1994. Exhausting and regenerating resin for uranium removal. J. Am. Water Works Assoc. 86, 228–241.