Journal of Molecular Liquids 234 (2017) 172–178
Contents lists available at ScienceDirect
Journal of Molecular Liquids journal homepage: www.elsevier.com/locate/molliq
Facile synthesis of gelatin modified attapulgite for the uptake of uranium from aqueous solution Huaxuan Han a,b, Cheng Cheng a, Shuheng Hu b, Xiaolong Li c, Wenjuan Wang c, Chengjian Xiao c,⁎, Zimu Xu b,⁎, Dadong Shao a,c,⁎⁎ a b c
Institute of Nuclear Physics and Chemistry, China Academy of Engineering Physics, Mianyang 621900, China School of Resources and Environmental Engineering, Hefei University of Technology, Hefei 230009, China Institute of Plasma Physics, Chinese Academy of Sciences, Hefei 230031, China
a r t i c l e
i n f o
Article history: Received 23 January 2017 Received in revised form 7 March 2017 Accepted 9 March 2017 Available online 22 March 2017 Keywords: Uranium Adsorption Attapulgite Gelatin Plasma technique
a b s t r a c t Gelatin modified attapulgite (ATT@Gel) with high adsorption capability for U(VI) was synthesized by plasma technique. The ATT and ATT@Gel composites were characterized by scanning electron microscope (SEM), Xray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and thermogravimetric analysis (TGA). The characterization results indicated that ATT@Gel was synthesized successfully. The adsorption capability of ATT@Gel for U(VI) was studied by batch adsorption technique. Experiment results indicate that the modified gelatin on ATT surface can markedly improves the adsorption efficiency of ATT@Gel for U(VI). © 2017 Published by Elsevier B.V.
1. Introduction The behaviors of long-lived actinides in environmental has aroused great interest in radioactive waste management [1,2]. The migration and diffusion of actinides on the surface of minerals are of primary importance for its environmental behavior [3]. The fate of actinides in environment solution is mainly dominated by its adsorption and complexation on mineral surfaces [4–7]. Uranium, a typical actinide elements, exists as hexavalent uranyl (U(VI)) in environment solution. U(VI) can be retained and enriched in environment for a long time due to its long half-life (such as 238Ut1/2 = 4.51 × 109 years) [8–10]. Therefore, it is very important to uptake uranium ions from environment solution. Attapulgite, a number of sepiolite families in mineralogy, its basic structural unit is 2: 1 layer type which is two layers of silicon oxide tetrahedraclip a layer of magnesium oxide octahedron. Si4+ is replaced by Al3+ on the surface of attapulgite and thereby its appears negative charge, which ensures it presents adsorption capability for trace heavy metal ions [11,12] and radionuclides (such as U(VI)) in environment solution [13]. However, the low adsorption capability of attapulgite greatly limits its application in the management of
radioactive pollution. Amino groups are widely considered as effective chelating functional groups to separate various contaminants from solution because of their high reactivity [14–16]. Gelatin, a heterogeneous mixture of animal origin protein, contains abundant amounts of amino groups which can form strong complex with metal ions in solution. To further enhance the adsorption capability of attapulgite, attapulgite was modified with gelatin (denoted as ATT@Gel) in this work. Plasma technique was used in synthesis process because it is an environmental friendly and effective technique which can modify material surface without altering its bulk properties [17]. The energy of activated particles formed in plasma is much greater than chemical bond energy and can break chemical bonds to form active species on ATT surface. The active species would then react with the functional groups in gelatin, which result in the modification of gelatin on ATT surface. The prepared ATT@Gel composites were applied to uptake U(VI) from solution to evaluate its potential application in radioactive pollution management. 2. Experimental 2.1. Chemicals
⁎ Corresponding authors. ⁎⁎ Correspondence author at: Institute of Nuclear Physics and Chemistry, China Academy of Engineering Physics, Mianyang, 621900, China. E-mail addresses:
[email protected] (C. Xiao),
[email protected] (Z. Xu),
[email protected] (D. Shao).
http://dx.doi.org/10.1016/j.molliq.2017.03.076 0167-7322/© 2017 Published by Elsevier B.V.
All chemical reagents used in this paper were in analytic purity. The attapulgite (ATT) was received from Kai-Xi Co. China. ATT@Gel was synthesized by plasma technique. Briefly, 1.0 g attapulgite was treated by nitrogen plasma (100 W) for 30 min the
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178
173
Fig. 1. The XRD patterns (A) and TGA curves (B) of ATT and ATT@Gel.
treated ATT was added into 100 mL 10.0 g/L gelatin solution and reacted at 50 °C for 3 h under magnetic stirring, and then washed with Milli-Q water and dried at 50 °C. The resulted materials were denoted as ATT@Gel. 2.2. Adsorption experiment The adsorption experiment was performed by batch adsorption technique. After the adsorbent (ATT and ATT@Gel) was pre-equilibrated with NaCl for 12 h in polyethylene tubes, U(VI) stock solution and MilliQ water were added to obtain the desired composition, and the pH values were adjusted by diluted nitric acid or sodium hydroxide. After shaken for requested time, the suspensions were centrifuged at 18,000 rpm for 10 min (Beckman Coulter 64R) at the temperature in adsorption experiments. The residual U(VI) concentration in supernatant was measured by arsenazo III spectrophotometric method. All adsorption were repeated for three times, and the uncertainties of adsorption data were b5%.
2.3. Characterization The ATT and ATT@Gel samples were characterized by SEM, XRD, TGA, and XPS. The SEM images were obtained on a FEI Sirion 200 FEG scanning electron microscope. The XRD patterns were measured on a D/MAX-2500 V equipped with Cu Kα radiation. The TGA measurements were performed on a Shimadzu TGA-50 thermogravimetric analyzer. The heating rate and air flow rate were 10 °C /min and 40 mL/min, respectively. XPS measurements were performed on an ESCALab220i-XL surface microanalysis system equipped with an Al Kα source.
3. Results and discussion Fig. 1A shows the XRD patterns of ATT and ATT@Gel samples. The XRD peaks at 2θ = 8.4°, 13.8°, 19.7°, 21°, 27.5°, 35.0° and 42.6° indicate the crystal structure of ATT [18]. Quartz and montmorillonite are also found in the attapulgite samples. The XRD peak positions of ATT and
Fig. 2. SEM images of ATT (A, B) and ATT@Gel (C, D).
174
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178
Fig. 3. XPS C 1s spectra (A) and O 1s spectra (B) of ATT and ATT@Gel.
ATT@Gel are fairly similar, which reveals that the crystal structure of ATT was not destroyed during plasma indiced modification process. Fig. 1B shows the TGA patterns of ATT and ATT@Gel. The first weight losses below 150 °C can be assigned to the releasing of surface adsorbed water, which were estimated to be 6.63% and 6.37% in ATT and ATT@ Gel, respectively. The second weight losses at 150–350 °C can be attributed to the releasing of structure water, which were detected to be 3.02% and 2.44% in ATT and ATT@Gel, respectively. The third weight loss at N 350 °C can be due to other impurity minerals and the decomposition of gelatin, which were estimated to be 5.30% and 17.8% in ATT and ATT@Gel, respectively. According to the difference weight losses, the weight percent of gelatin in dry ATT@Gel is calculated to be ~13.6%. Fig. 2 depicted SEM images of ATT and ATT@Gel. One can see that ATT (Fig. 2A, B) shows a fibrous morphology and ATT fibers also form straight parallel aggregations. It reveals the weak cohesive force between raw ATT can be neglected, and the very low content of organic materials in raw ATT. However, ATT@Gel particles (Fig. 2C, D) are agglomerated and covered by organic films, which can be due to the cohesive force generated from functional groups in modified gelatin. It is well known, gelatin is composed of many kinds of amino acids, and
Table 1 Curve fitting results of XPS C 1 s spectra.
ATT
ATT@Gel
a b
Peak
BEa (eV)
FWHMb (eV)
%
C_C C\ \C C\ \OH NC_O \ \COOH C_C C\ \C C\ \OH NC_O \ \COOH
284.42 285.24 286.50 287.30 289.20 284.26 285.47 286.10 287.44 288.80
1.73 1.77 1.55 0.93 2.29 1.75 1.27 1.25 1.75 0.82
43.1 38.8 7.23 2.57 8.26 51.1 16.4 8.70 22.9 0.85
various interaction forces, such as Van der Waals forces, π-π interactions, hydrophobic interaction, and Lewis acid–base interaction, are exist among the amino acids in gelatin. Thereby, the SEM images indicate that gelatin was successfully modified on to the surface of ATT. Shao et al. [19] reported a similar microstructure transformation of carboxymethyl cellulose modified multiwalled carbon nanotubes. Fig. 3A show the XPS C 1 s spectra of ATT and ATT@Gel. The peak fractions of hydroxyl groups (C-OH) and sp2- hybridized (C = C) increase from 7.23% to 8.70% and from 43.1% to 51.1%, respectively (Table 1), which can be assigned to effect of modified gelatin on ATT surfaces because there gelatin contains abundance aromatic organics. Meanwhile, the XPS O 1 s spectra of ATT and ATT@Gel (Fig. 3B and Table 2) show that hydroxyl groups (C\\OH) and carbonyl groups (C_O) are the dominant oxygen species on the surface of ATT and ATT@Gel, respectively. It further prove the successfully modification of gelatin on ATT surface. 3.1. Adsorption experiment The adsorption of U(VI) on ATT and on ATT@Gel surface as a function of contact time is depicted in Fig. 4. The adsorption of U(VI) was rapid increased at initial contact time, and reached equilibrium in 12 h. Thereby, 24 h was selected as contact time to study the adsorption of U(VI) on ATT and ATT@Gel in following adsorption experiment. The adsorption of U(VI) on ATT@Gel is much higher than that on ATT, which can be explained by the forming binding between modified gelatin and U(VI).
Binding energy. Full width at half-maximum.
Table 2 Curve fitting results of XPS O 1 s spectra.
ATT
ATT@Gel
peak
BE (eV)
FWHM (eV)
%
C\ \OH C_O COOH C\ \OH C_O COOH
531.30 532.40 533.80 531.25 532.40 533.72
1.71 1.78 2.17 2.18 1.00 1.67
29.3 58.0 12.7 95.9 0.00 4.07
Fig. 4. Effect of contact time on U(VI) adsorption on ATT and on ATT@Gel. C[U(VI)]initial = 6.25 mg/L, I = 0.01 M NaCl, m/V = 0.5 g/L, pH = 5.0 ± 0.1, T = 293 ± 1 K.
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178 Table 3 Kinetic parameters of pseudo-first-order and pseudo-second-order kinetic models. Pseudo–first–order model
ATT ATT@Gel
t 1 t ¼ þ t qt 2k2 q2e qe
Eq: ð2Þ
Pseudo–second–order model
qe (mg/g)
k1 (1/h)
R2
qe (mg/g)
k2 (mg/g h)
R2
5.76 7.53
1.11 1.06
0.796 0.716
5.92 7.76
0.267 0.188
0.999 0.999
where k1 and k2 are the pseudo-first and pseudo-second order kinetic constants, respectively. qt and qe are the amount of U(VI) adsorbed at time t and at equilibrium. According to the correlation coefficients (R2) [23] in Table 3, the experimental data can be fitted very well by the pseudo-second-order rate model than by the pseudo-second-order rate model, which implies that the adsorption of U(VI) on ATT and on ATT@Gel surfaces is limited by a chemisorption mechanism. Fig. 5 shows the U(VI) adsorption onto ATT@Gel as a function of adsorbent content. The adsorption of U(VI) on ATT@Gel increases with increasing adsorbent content. At same experimental conditions, more adsorption sites can be obtained at higher adsorbent content, which in favor of U(VI) adsorption. Distribution coefficient (Kd) is obtained from Eq. 3 [24] and shown in Fig. 5.
Kd ¼
Fig. 5. Effect of solid content on U(VI) adsorption to ATT@Gel. C[U(VI)]initial = 6.25 mg/L, I = 0.01 M NaCl, pH = 5.0 ± 0.1, T = 293 ± 1 K.
Pseudo-first-order kinetic equation (Eq. 1) and pseudo-second-order kinetic equation (Eq. 2) is used to simulate experimental data [20–22]. ln ðqe −qt Þ ¼ lnqe −k1 t
175
Eq: ð1Þ
Co −Ce V þ Ce m
Eq: ð3Þ
where V and m are the volume of suspension and the mass of adsorbent, respectively. It is obviously that the Kd values are decrease slightly with increasing adsorbent content, which could be due to the competition among the adsorption sites on the surface of ATT@Gel. Meanwhile, adsorbent particle would aggregate at high solid content that can decrease the efficiency of adsorbent, and thereby suppress the increasing of adsorption at high solid content [25,26]. The effect of solution pH on U(VI) adsorption on ATT and ATT@Gel was studied in NaCl solution (0.001, 0.01 and 0.1 M) and depicted in Fig. 6A. The adsorption of U(VI) is strongly dependent on pH, which
Fig. 6. Adsorption of U (VI) on ATT@Gel in different NaCl solutions as a function of pH values (A). CU(initial) = 6.25 mg/L, m/V = 0.5 g/L, T = 293 ± 1 K. The zeta potentials of ATT and ATT@ Gel (B). The relative species of 6.25 mg/L U(VI) as a function of pH in aqueous solutions (C).
176
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178
Table 4 Aqueous complexation reactions of U(VI).
Table 5 The relative parameters for Langmuir and Freundlich isotherms at different temperatures.
Reactions
Log K (I = 0)
+ H2O = UO2(OH)+ + H+ UO2+ 2 + 2H2O = UO2(OH)02 + 2H+ UO2+ 2 + UO2+ 2 + 3H2O = UO2(OH)3− + 3H + 2− UO2+ 2 + 4H2O = UO2(OH)4 + 4H 2UO2+ + H2O = (UO2)2(OH)3+ + H+ 2 + 2H2O = (UO2)2(OH)2+ + 2H+ 2UO2+ 2 2 + 5H2O = (UO2)3(OH)5+ + 5H+ 3UO2+ 2 3UO22+ + 7H2O = (UO2)3(OH)7− + 7H+ + 7H2O = (UO2)4(OH)7+ + 7H+ 4UO2+ 2
−5.25 −12.15 −20.25 −32.4 −2.70 −5.62 −15.55 −32.20 −21.90
increases quickly and slowly at pH b 6.0 and at pH = 6.0–7.5, respectively. The adsorption of U(VI) was then decreased with further increasing pH, which can be due to electrostatic repulsion. As can see from Fig. 6B, the zeta potentials of ATT and ATT@Gel becomes more negative with the increase of pH value in pH 3–10. Meanwhile, there have no significant differences of the zeta potentials between ATT and ATT@Gel. Zhu et al. [23] also reported similar U(VI) adsorption behavior on attapulgite and they contributed it to the different existing forms of U(VI) at different pH values. To further study U(VI) adsorption behavior, the distribution of U(VI) species as a function of pH were calculated according the thermodynamic data (Table 4) and showed in Fig. 6C. The distribution of U(VI) species are fairly affected by pH values. Free uranyl ion (UO22 +) and U(VI) hydrolysis complexes are dominant species at pH b 5 and pH N 5, respectively. Moreover, U(VI) species are positively charged at pH b 8, and then change gradually from positively charge to negatively charge. According to the adsorption isotherms of U(VI) on ATT and ATT@Gel are depicted in Fig. 7A, the adsorption on ATT@Gel are significantly higher than on ATT. Meanwhile, the adsorption isotherms are increased with increasing temperature, which illustrate that high temperature is
ATT
ATT@Gel
T (K)
Langmuir model Cs max (mg/g)
b (L mg−1)
R2
Freundlich model KF (mg1−nLng−1)
n
R2
293 313 333 293 313 333
9.35 10.4 11.2 23.9 24.5 25.3
0.422 0.520 0.093 0.402 0.519 0.626
0.996 0.995 0.999 0.990 0.992 0.995
3.79 4.91 5.97 7.59 8.79 9.75
0.269 0.222 0.207 0.381 0.351 0.338
0.977 0.992 0.916 0.968 0.969 0.967
in favor of U(VI) adsorption on ATT@Gel and on ATT. These can be interpreted by the fact that U(VI) is hydrated in solution and the losing of U(VI) hydration shell need energy during adsorption [27]. The Langmuir and Freundlich isotherm models are applied to analysis adsorption data to get a better understanding of the adsorption mechanism of U(VI) on ATT@Gel. The Langmuir isotherm model is often applied to simulate the adsorption in a monolayer. The model can be described by Eq. 4: Ce Ce 1 þ ¼ Cs Cs max bCs max
Eq: ð4Þ
where Ce is the concentration in supernatant after centrifugation, Cs is the equilibrium concentration adsorbed on adsorbent, Cs max and b are constants related to maximum adsorption capacity and adsorption energy, respectively. Freundlich isotherm model [28] is often used to illustrate the adsorption on heterogeneous adsorption surface, which usually expressed as Eq. 5: LnCs ¼ nLnCe þ LnK F
Eq: ð5Þ
Fig. 7. Adsorption isotherms of U(VI) adsorption (A). Liner plots of lnKd versus Ce of U(VI) adsorption (B). Langmuir isotherm model (C) Freundlich isotherm model (D) onto ATT@Gel at three different temperatures.
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178
177
Table 6 Thermodynamic parameters for U(VI) adsorption on ATT and on ATT@Gel.
ATT
ATT@Gel
T (K)
△G° (kJ/moL)
△H° (kJ/moL)
△S° (J/moL/K)
293 303 333 293 303 333
−21.2 −23.7 −26.1 −23.4 −25.1 −28.3
10.8
111
11.5
120
where KF and n represent the adsorption capacity and the degree of adsorption relates to equilibrium concentration, respectively. Fig. 7B and C and Table 5 show that the results of Langmuir and Freundlich models. According to the correlation coefficients (R2) in Table 5, the Langmuir isotherm model fits experiment data better than the Freundlich isotherm model. The Cs,max values of U(VI) on ATT@Gel surface obtained from the Langmuir isotherm model are 23.9, 24.5, and 25.3 mg/g at 293, 313, and 333 K, which are comparable to many adsorbent under similar experimental conditions, such as NaATT (14.42 mg/g) [23], illite (5.27 mg/g) [29], hematite (5.59 mg/g) [30], and hierarchical Fe4(P2O7)3 (14.92 mg/g) [31]. The thermodynamic parameters for U(VI) adsorption on ATT@Gel are obtained from the temperature-dependent adsorption isotherms. The free energy change (ΔG°) is firstly obtained from following relationship (Eq. 6) [32]: ΔG ° ¼ −RTLnK °
Eq: ð6Þ
where R and T are the ideal gas constant (8.314 J mol−1 K−1) and the reaction temperature in Kelvin, respectively. LnK° values are got by plotting LnKd versus Ce at T = 293, 313 and 333 K (Fig. 7D) and extrapolating the Ce values to zero. The standard entropy change (ΔS°) is obtained from following equation (Eq. 7): ∂ΔG ° ΔS ¼ − ∂T
!
°
Eq: ð7Þ p
The average standard enthalpy change (ΔH°) is obtained from the values of ΔG° and ΔS° (Eq. 8): ΔH ° ¼ ΔG ° þ TΔS °
Eq: ð8Þ
The relative thermodynamic parameters in Table 6 give an explanation about the adsorption mechanism of U(VI) on ATT@Gel. The positive
Fig. 9. Recycling of ATT@Gel for the removal of U(VI). C[U(VI)]initial = 6.25 mg/L, I = 0.01 M NaCl, m/V = 0.5 g/L, pH = 5.0 ± 0.1, T = 293 ± 1 K.
ΔH° value reveals that the adsorption is endothermic process. This positive ΔS° can serve as an explanation for how well U(VI) adsorption onto ATT@Gel surface from aqueous solution. High temperature is in favor of the dehydration process that needs energy [33]. The negative ΔG° values are believed to be a spontaneous process. Furthermore, the ΔG° values are also decreased with increasing reaction temperature, which indicates that the adsorption is increases at higher temperature. It is well known that many metal ions, such as Sr(II), Ni(II), Pb(II), Cu(II), and Fe(III), co–exists with U(VI) in radioactive wastewater, and these metal ions can compete with U(VI) for the adsorption sites on adsorption surface. Therefore, the selective adsorption of those metal ions and U(VI) on ATT@Gel surface was studied. As can be seen from Fig. 8, the adsorption of U(VI) on ATT@Gel surface is much higher than other metal ions under same experimental conditions, which reveals the high selectivity of ATT@Gel toward U(VI) in radioactive wastewater. The result highlights the potential application of ATT@Gel in selective uptake of U(VI) from radioactive wastewater. The recycling of ATT@Gel is critical for its usage in potential application. The regeneration of ATT@Gel was achieved by dispersed 0.1 g Uladen ATT@Gel in 100 mL 0.1 M HNO3 solution and rinsed with MilliQ water thoroughly, and dried at 50 °C for 24 h. the regenerated ATT@ Gel was applied in adsorption experiment. As shown in Fig. 9, the adsorption of U(VI) on regenerated ATT@Gel just slightly decreases with increasing rounds, and the regenerated ATT@Gel still presents higher adsorption capability for U(VI) even after (at least) eight cycles of applications. From the industrial point of view, ATT@Gel has excellent stability and the regenerated ATT@Gel still meets the requirement of real application. 4. Conclusions In summary, the modified gelatin on ATT surface can sound improve its adsorption capability for U(VI). The results of XRD pattern show that the crystal structure of ATT was not destroyed during plasma induced modification process. The SEM images indicate that gelatin was successfully modified on ATT surface. The adsorption of U(VI) on ATT@Gel increases with increasing adsorbent content. The adsorption of U(VI) increases at pH b 7.5, and then decrease at pH N 7.5. The adsorption of U(VI) onto ATT@Gel is an endothermic and spontaneous process. It is followed the pseudo-second-order kinetic model and Langmuir isotherm model, and is dependent on solution pH and ionic strength. Acknowledgements
Fig. 8. Selective adsorption comparison of metal ions and U(VI) on ATT@Gel surface. m/V = 0.50 g/L, C0 = 6.25 mg/L, T = 293 ± 1 K, pH = 5.0 ± 0.1, I = 0.1 mol/L NaCl, contact time 24 h.
Financial supports from the NSAF (U1530131), the National Natural Science Foundation of China (11675210), the Radiochemistry 909
178
H. Han et al. / Journal of Molecular Liquids 234 (2017) 172–178
Project in the China Academy of Engineering Physics, the Science and Technology Development Foundation of China Academy of Engineering Physics (2014B0301034), and the China Postdoctoral Science Foundation (2015M582769XB, 2016T90872) are acknowledged. References [1] D. Shao, D. Xu, S. Wang, Q. Fan, W. Wu, Y. Dong, X. Wang, Modeling of radionickel sorption on MX-80 bentonite as a function of pH and ionic strength, Sci. China B 52 (2009) 362–371. [2] K. Wendt, T. Gottwald, C. Mattolat, S. Raeder, Ionization potentials of the lanthanides and actinides towards atomic spectroscopy of super-heavy elements, Hyperfine Interact. 227 (2014) 55–67. [3] A. Singh, K.U. Ulrich, D.E. Giammar, Impact of phosphate on U(VI) immobilization in the presence of goethite, Geochim. Cosmochim. Acta 74 (2010) 6324–6343. [4] G. Sheng, J. Hu, X. Wang, Sorption properties of Th(IV) on the raw diatomite-Effects of contact time, pH, ionic strength and temperature, Appl. Radiat. Isot. 66 (2008) 1313–1320. [5] K. Mori, K. Tada, Y. Tawara, K. Ohno, M. Asami, K. Kosaka, H. Tosaka, Integrated watershed modeling for simulation of spatiotemporal redistribution of post-fallout radionuclides: application in radiocesium fate and transport processes derived from the Fukushima accidents, Environ. Model. Softw. 72 (2015) 126–146. [6] D. Shao, Q. Fan, J. Li, Z. Niu, W. Wu, Y. Chen, X. Wang, Removal of Eu(III) from aqueous solution using ZSM-5 zeolite, Microporous Mesoporous Mater. 123 (2009) 1–9. [7] B.D. Stewart, M.A. Mayes, S. Fendorf, Impact of uranyl−calcium−carbonato complexes on uranium(VI) adsorption to synthetic and natural sediments, Environ. Sci. Technol. 44 (2010) 928–934. [8] S.D. Yusan, S. Akyil, Sorption of uranium(VI) from aqueous solutions by akaganeite, J. Hazard. Mater. 160 (2008) 388–395. [9] A. Krestou, A. Xenidis, D. Panias, Mechanism of aqueous uranium(VI) uptake by natural zeolitic tuff, Miner. Eng. 16 (2003) 1363–1370. [10] R. Han, W. Zou, Y. Wang, Removal of uranium(VI) from aqueous solutions by manganese oxide coated zeolite: discussion of adsorption isotherms and pH effect, J. Environ. Radioact. 93 (2007) 127–143. [11] L. Zhu, J. Guo, P. Liu, Effects of length and organic modification of attapulgite nanorods on attapulgite/polystyrene nanocomposite via in-situ radical bulk polymerization, Appl. Clay Sci. 119 (2016) 87–95. [12] E. Pehlivan, S. Cetin, K. Yan, Equilibrium studies for the sorption of zinc and copper from aqueous solutions using sugar beet pulp and fly ash, J. Hazard. Mater. 135 (2006) 193–199. [13] C. Pang, Y. Liu, X. Cao, R. Hua, C. Wang, Adsorptive removal of uranium from aqueous solution using chitosan-coated attapulgite, J. Radioanal. Nucl. Chem. 286 (2010) 185–193. [14] D. Shao, G. Hou, J. Li, T. Wen, X. Ren, X. Wang, PANI/GO as a super adsorbent for the selective adsorption of uranium(VI), Chem. Eng. J. 255 (2014) 604–612. [15] D. Shao, C. Chen, X. Wang, Application of polyaniline and multiwalled carbon nanotube magnetic composites for removal of Pb(II), Chem. Eng. J. 185–186 (2012) 144–150.
[16] D. Shao, J. Hu, C. Chen, G. Sheng, X. Ren, X. Wang, Polyaniline multiwalled carbon nanotube magnetic composite prepared by plasma-induced graft technique and its application for removal of aniline and phenol, J. Phys. Chem. C 114 (2010) 21524–21530. [17] L. Li, F. Chen, J. Shao, Z. Tang, Attapulgite clay supported Ni nanoparticles encapsulated by porous silica: thermally stable catalysts for ammonia decomposition to COx free hydrogen, Int. J. Hydrog. Energy 14 (2016) 21157–21165. [18] S. Yang, J. Li, Y. Lu, Y. Chen, X. Wang, Sorption of Ni(II) on GMZ bentonite: Effects of pH, ionic strength, foreign ions, humic acid and temperature, Appl. Radiat. Isot. 67 (2009) 1600–1608. [19] Q. Fan, D. Shao, J. Hu, W. Wu, X. Wang, Comparison of Ni2+ sorption to bare and ACT-graft attapulgites: effect of pH, temperature and foreign ions, Surf. Sci. 602 (2008) 778–785. [20] J.P. Simonin, On the comparison of pseudo-first order and pseudo-second order rate laws in the modeling of adsorption kinetics, Chem. Eng. J. 300 (2016) 254–263. [21] A. Duconseille, M. Traikia, M. Lagree, C. Jousseb, G. Pagès, P. Gatellier, T. Astruc, V. Santé-Lhoutellier, The impact of processing and aging on the oxidative potential, molecular structure and dissolution of gelatin, Food Hydrocoll. 11 (2016) 1–13. [22] Y.S. Ho, G. Mckay, Kinetic models for the sorption of dye from aqueous solution by wood, Process. Saf. Environ. Prot. 76 (1998) 183–191. [23] W. Zhu, Z. Liu, L. Chen, Y. Dong, Sorption of uranium(VI) on Na-attapulgite as a function of contact time, solid content, pH, ionic strength, temperature and humic acid, J. Radioanal. Nucl. Chem. 289 (2011) 781–788. [24] X. Yang, S. Yang, S. Yang, J. Hu, X. Tan, X. Wang, Effect of pH, ionic strength and temperature on sorption of Pb(II) on NKF-6 zeolite studied by batch technique, Chem. Eng. J. 168 (2011) 86–93. [25] B. Hu, W. Cheng, H. Zhang, G. Sheng, Sorption of radionickel to goethite: effect of water quality parameters and temperature, J. Radioanal. Nucl. Chem. 285 (2010) 389–398. [26] M. Wang, H. Xie, L. Tan, J. Qiu, X. Tao, C. Wu, Uptake properties of Eu(III) on Naattapulgite as a function of pH, ionic strength and temperature, J. Radioanal. Nucl. Chem. 292 (2012) 763–770. [27] L. Tan, Y. Jin, J. Chen, X. Cheng, J. Wu, L. Feng, Sorption of radiocobalt(II) from aqueous solutions to Na-attapulgite, J. Radioanal. Nucl. Chem. 289 (2011) 601–610. [28] S. Yang, D. Zhao, H. Zhang, S. Lu, L. Chen, X. Yu, Impact of environmental conditions on the sorption behavior of Pb(II) in Na-bentonite suspensions, J. Hazard. Mater. 183 (2010) 632–640. [29] Y. Gao, Z. Shao, Z. Xiao, U(VI) sorption on illite: effect of pH, ionic strength, humic acid and temperature, J. Radioanal. Nucl. Chem. 303 (2015) 867–876. [30] D. Zhao, X. Wang, S. Yang, Z. Guo, G. Sheng, Impact of water quality parameters on the sorption of U(VI) onto hematite, J. Environ. Radioact. 103 (2012) 20–29. [31] X. Liu, Y. Xu, R. Jin, P. Yin, L. Sun, T. Liang, S. Gao, Facile synthesis of hierarchical Fe4(P2O7)3 for removal of U(VI), J. Mol. Liq. 200 (2014) 311–318. [32] D. Shao, Z. Jiang, X. Wang, J. Li, Y. Meng, Plasma induced grafting carboxymethyl celfrom aqueous solulose on multiwalled carbon nanotubes for the removal of UO2+ 2 lution, J. Phys. Chem. B 113 (2009) 860–864. [33] G. Sheng, S. Wang, J. Hu, Y. Lu, J. Li, Y. Dong, X. Wang, Adsorption of Pb(II) on diatomite as affected via aqueous solution chemistry and temperature, Colloids Surf. A Physicochem. Eng. Asp. 339 (2009) 159–166.