Synthesis and characterization of a starch–AlOOH–FeS2 nanocomposite for the adsorption of congo red dye from aqueous solution

Rajeev Kumar*a, Jamshaid Rashida and M. A. Barakatabc
aDepartment of Environmental Sciences, Faculty of Meteorology, Environment and Arid Land Agriculture, King Abdulaziz University, Jeddah, 21589, Saudi Arabia. E-mail: olifiaraju@gmail.com; Fax: +966(02) 6952364; Tel: +966 (02)6400000 ext. 64821
bCentral Metallurgical R & D Institute, Helwan, Cairo, 11421, Egypt
cCenter of Excellent in Environmental Studies (CEES), King Abdulaziz University, Saudi Arabia

Received 31st May 2014 , Accepted 12th August 2014

First published on 15th August 2014


Abstract

This work described the synthesis and characterization of a starch–AlOOH–FeS2 nanocomposite for the adsorption of congo red (CR) dye from aqueous solution. The morphology of the starch–AlOOH–FeS2 was characterized by using scanning and transmission electron microscopy, N2 adsorption–desorption isotherms, X-ray photoelectron spectroscopy, and fourier transform-infrared spectroscopy. The adsorption of CR onto starch–AlOOH–FeS2 was evaluated as a function of contact time, solution pH, concentration and temperature. The adsorption results demonstrate that the maximum removal of CR was found to be at pH 5. The adsorption kinetics data fitted well to the pseudo first-order equation whereas the Freundlich equation exhibits better correlation to the experimental data. Thermodynamic parameters, such as the standard free energy change (ΔG°), the standard entropy change (ΔS°) and the standard enthalpy change (ΔH°), were also evaluated. The results suggested that starch–AlOOH–FeS2 is a potential adsorbent for CR dye removal from aqueous solution.


Introduction

Organic dyes and pigments discharged from sources such as textile, cosmetic, food, pharmaceutical, and paper industries are serious environmental problems because their unwanted colour and resistance to decomposition. Synthetic dyes are toxic and carcinogenic in nature towards the organism and mammals.1,2 Congo red (CR) (sodium salt of benzidinediazobis-1-naphthylamine-4-sulfonic acid) is widely used in the textile, leather, paper, and printing industries and around 15% CR is discharged in wastewater during operation. CR is well known carcinogenic and its exposure may cause some allergic responses.2,3 Moreover, due to its complex aromatic structure, eliminate or degrade (bio/photo) of CR from aqueous medium by conventional methods is difficult. Several methods such as adsorption, oxidation, photocatalysis, membrane filtration, biological degradation etc. have been used for the treatment of dyeing effluents.1,4–7 Among all of them, adsorption is a very promising method for the removal of dyes from wastewater. Several adsorbents such as carbons, metal oxides, polymers, clays, composites etc. have been used for the removal of CR from aqueous solutions.1,2,4,8–10

However, low or moderate adsorption capacities of these materials towards the CR removal promote the researchers to develop novel materials with high efficiency and economical removal of CR from dyeing effluents.

The recently discovered hybrid materials showed many advantages in wastewater treatment.3 The adsorption ability of hybrid materials relies on the smart manipulation of the structure of the embedded compounds such as charge, functionality, hydrophobic–hydrophilic nature etc.5 Boehmite (AlOOH) is non-hazardous material, exhibit good adsorption capacity (85–99 mg g−1) toward CR from aqueous solution.11 The adsorption of the CR onto AlOOH takes place due to interaction between metal sites and amine group of dye. The adsorption performance of AlOOH can be enhanced by tailoring its surface. The high functionality of starch and pyrite combined with AlOOH can much effectively remove the CR from wastewater. Pyrite and starch are most abundant sulfide mineral and biopolymer, respectively. Starch and starch based adsorbent have been investigated for the removal of various classes of dyes.12–14 To the best of our knowledge, pyrite or pyrite based adsorbent has not been investigated for the removal of dyes. However, few studies were reported for the removal of metal ions such as molybdate, tetrathiomolybdate,15 Cu(II), Cd(II) and Pb(II),16 As(III) and As(V),17 etc. Pyrite showed good adsorption capacity for negatively charged metal ions. Therefore, it is supposed to be good adsorbent for the removal of negatively charged CR dye. In this work, a simple route for the synthesis of starch–AlOOH–FeS2 composite and its adsorption efficiency for the removal of CR is reported. The adsorption of CR onto starch–AlOOH–FeS2 is discussed in terms of kinetics, thermodynamics and isotherm studies.

Materials and methods

Materials

Starch (potato) and aluminum nitrate were purchased from Panreac Quimica S.A. Congo red and Sodium dodecyl sulfate (SDS) were obtained from Techno Pharmachem Haryana, India and Scharlab S.L. Spain. FeSO4·7H2O and NH4HCO3 were supplied by BDH chemical Ltd and Fisher Scientific Company, New Jersey.

Synthesis of starch–AlOOH–FeS2nanocomposite

In a typical synthesis, 13.9 g Al2(NO3)3 was dissolved in 25 mL water under sonicated (Elma S30H Elmasonic) for 20 min, and then 1 g starch was added into the solution. After sonication at 40 °C for 30 min, 1.2 g SDS was added into the solution to prevent the agglomeration of particles. Thereafter, 100 mL 2.0 M NH4HCO3 was added to the mixture and sonicated for 2 h at 60 °C. Resulting white precipitate was filtered by Buchner funnel with a sintered glass disk and thoroughly washed with hot de-ionized water and ethanol for the removal of excess starch and SDS. The obtained white precipitate was transferred into ethanol solution (200 mL) containing 8.0 g Na2S2O3 and 4.4 g FeSO4·7H2O under sonication at 40 °C for 30 min. Thereafter, the mixture was refluxed at 180 °C for 1 h and finally cooled to room temperature. The precipitate was filtered and thoroughly washed with hot de-ionized water and ethanol, and dried at 200 °C for 4 h. Dry sample was collected and stored for further adsorption studies.

Characterization

The morphology of starch–AlOOH–FeS2 was investigated by field emission scanning electron microscopy (FESEM) (JSM-7500 F; JEOL-Japan) without spurting coated gold or carbon and transmission electron microscopy (TEM) (model Tecnai G2 F20 Super Twin) at 200 kV with LaB6 emitter. X-ray photoelectron spectroscopy (XPS) measurements were recorded on SPECS GmbH, (Germany) spectrometer operated at a base pressure of 10−10 mbar and step size 1 eV. A non-monochromatic Mg-Kα (1253.6 eV) X-ray source was used to irradiate the sample surface with 13.5 kV, 100 W X-ray power. The nitrogen adsorption–desorption isotherm analysis [Brunauer, Emmett and Teller (BET) and Density Functional Theory (DFT)] of starch–AlOOH–FeS2 was performed at 77 K Quantachrome Nova Win 2 (Quantachrome instruments) equipment by degassing the sample at 200 °C. FTIR spectra of starch–AlOOH–FeS2 before and after CR adsorption were recorded using a Perkin Elmer Spectrum 100 FTIR Spectrometer over a range of 500–4000 cm−1.

Adsorption studies

Adsorption of CR was investigated under a wide variety of experimental conditions. A series of experiments were performed by reacting 0.02 g starch–AlOOH–FeS2 with 20 mL of CR solution (50–500 mg L−1) as a function of solution pH ranging from 5–9. The solution pH was adjusted using 0.1 M HCl and NaOH. Adsorption isotherms were performed at pH 5, 7, 9 and at temperature 30, 40, 50 °C. The effect of contact time was studied in a series of flasks contain 20 mL dye solution of 500 mg L−1 in water bath shaker at 200 rpm at temperature 25, 40, 50 °C. After a fixed time interval, the samples were filtered and the amount of dye remaining in the supernatant solution was analyzed using a HACH LANGES DR-6000 UV-visible spectrophotometer at 495 nm. The amount of CR adsorbed on starch–AlOOH–FeS2 can be estimated by:
 
qe = (C0Ce)V/m (1)
where, qe is the amount of CR adsorbed per unit mass of starch–AlOOH–FeS2 (mg g−1), C0 and Ce is the initial and equilibrium dye concentration (mg L−1), respectively. V is the volume of dye solution (L) and m is the mass of starch–AlOOH–FeS2 (g).

Results and discussion

Characterization

The morphology of the prepared composite material was investigated by SEM and TEM as shown in Fig. 1. The SEM images (Fig. 1a) shows the needle like clusters of starch–AlOOH–FeS2. The TEM image (Fig. 1b) also reflects the needle like structure. The high resolution TEM image (Fig. 1c) shows the polycrystalline nature of starch–AlOOH–FeS2 composite.
image file: c4ra05183a-f1.tif
Fig. 1 SEM (a), TEM (b) and HRTEM (c) of starch–AlOOH–FeS2.

The specific BET surface area and pore structure of the starch–AlOOH–FeS2 was investigated using N2 adsorption–desorption measurements. Fig. 2 shows the type IV isotherm. The BET, DFT surface area and total pore volume are 40.013, 25.33 and 0.055 m2 g−1, respectively. The pore diameter is found to be in the range of 1.348 to 3.39 nm and the total micropore and mesopore are 0.0549 and 0.0484 cm3 g−1, respectively. The surface analysis of starch–AlOOH–FeS2 by XPS is shown in Fig. 3. The sulfur peak located at 163.82 eV is corresponding to the 2p1/2 which is identical with the S2− in the pyrite.18 Broad peaks located at 168.69 and 530 eV are due to the presence of SO42− impurities of iron precursor and FeOOH (O 1s). Moreover, the Fe 2p3/2 and 2p1/2 peaks are located at 708.2 eV and 719.6 eV are consistent with bulk pyrite.19 Two peaks appeared at 118.5 eV and 74.68 eV are corresponding to the Al 2s and 2p, and confirm the presence of AlOOH in the composite.20 The binding energies for C 1s of the starch functional groups at 284.8, 286.6 and 288.38 eV are corresponding to the C–C/C–H, C–O (alcohol) and O–C–O (ester), respectively.21 In addition, the presence of these functional groups can be confirmed by FTIR analysis of starch–AlOOH–FeS2 nanocomposite. Fig. 4 shows the FTIR spectra of starch–AlOOH–FeS2. The broad and strong peaks centered at 3213 and 1633 cm−1 are attributed to the O–H vibrations of starch–AlOOH or water. The broad peak at 1079 cm−1 can be assigned to the C–O–C linkage of the starch.22,23 Moreover, the broad peak with small shoulders in the range of 1000–1200 may be attributed to the pyrite.24 The strong band at 766 cm−1 may be ascribed to the AlO6 vibration mode which shifted to 757 cm−1 after adsorption. After dye adsorption new peaks appear in the region 1500–1600 and 1417, 1371, 1040 cm−1 may be assign to the NH, NH2 and C–H (bending), S–O (stretching) vibrations of the dye.2,25,26 From the FTIR spectra, it is noted that most of the peaks shifted or diminished after the adsorption of dye. The shift in wave numbers and intensity of the peaks could be due to the interaction of dye functional groups (NH, NH2, SO3) with the adsorbent functional groups.2


image file: c4ra05183a-f2.tif
Fig. 2 Nitrogen adsorption–desorption isotherm plot of starch–AlOOH–FeS2.

image file: c4ra05183a-f3.tif
Fig. 3 XPS analysis of starch–AlOOH–FeS2.

image file: c4ra05183a-f4.tif
Fig. 4 FTIR spectra of starch–AlOOH–FeS2 before and after CR adsorption.

Effect of solution pH

CR is highly sensitive to the solution pH and shows colour change from red (λmax – 496 nm) to blue (λmax – 580 nm) in the presence of inorganic acids in the pH ranges from 2–4 due protonation of azo group (π–π transition).2,8 Fig. 5 shows the pH dependent adsorption of CR onto starch–AlOOH–FeS2 nanocomposite. The adsorption of CR decreases with the increase in the solution pH from 5 to 9 and maximum adsorption is found to be at pH 5 as shown in Fig. 5. The adsorptive removal of CR shows a slight decrease at pH 9 compared to pH 7. Several investigations revealed that adsorption of negatively charged CR usually increases as the solution pH decreased.2,8,27 This behavior can be explained on the basis of change in the functionality of adsorbent with respect to the solution pH. According to Gissinger et al.,28 S–S–H groups are predominated in acidic medium and pyrite surface shows a deficiency of iron and formation of iron vacancies Fe1−xS2, while O–H groups are predominated in basic medium. When the solution pH increases, the pyrite surface shows the excess of FeOOH. Both S–S–H and O–H groups are coexisting over the entire pH range at pyrite surface. Therefore, high adsorption is observed in acidic medium and both groups are thus probably involved in the binding of CR by the surface complexation. The adsorbent shows effective removal of CR at natural and basic condition (pH 9). This may be explain on the basis of positive surface charge of AlOOH (in basic medium at pH < 10) while CR is anionic charged. Thus, CR is electrostatically adsorbed on adsorbent surface (Dye–SO3 + AlOOH2+ ↔ Dye–SO3+H2OOAl).27,29 Furthermore, CR may be adsorbed on the surface of metal oxides due to a coordination effect of metal atoms with –NH2 and –SO3 groups of CR. Moreover, the adsorption of CR may be largely ascribed to the strong hydrogen bond formed between the electronegative atoms of dye and adsorbent such as oxygen, nitrogen or sulfur and hydrogen atoms.9 From the results mentioned above and previous reports9,10,27 hydrogen bonding, electrostatic interaction, complexation and coordination effect of metal atoms with –NH2 and –SO3 groups are mainly responsible for the adsorption of CR in neural CR solution. Therefore, pH 7 is selected for the further adsorption studies.
image file: c4ra05183a-f5.tif
Fig. 5 Effect of solution pH on CR adsorption.

Adsorption kinetics

The Effect of contact time on CR adsorption is shown in Fig. 6. The removal of CR is very fast during the first 100 min due to the adsorption of dye molecules onto the vacant adsorbent sites and thereafter rate of dye adsorption decreases due to the saturation of active sites (internal surface adsorption).2 The equilibrium is attained within 300 min. Adsorption dynamics of CR removal on starch–AlOOH–FeS2 are performed to find the rate controlling step. The adsorption data has been analyzed using pseudo first-order and pseudo second-order kinetic models. The linear equations for pseudo first-order30 and pseudo-second order31 kinetic models, respectively, are given as:
 
log(qeqt) = log[thin space (1/6-em)]qe − (k1t/2.303) (2)
 
t/qt = (1/k2qe2) + t/qe (3)
where and qe (mg g−1) and qt (mg g−1) refer to the adsorption capacity at equilibrium and after time t (min) respectively. k1 (min−1) and k2 (g mg−1 min−1) are the pseudo-first and pseudo-second order rate constant, respectively. The values for pseudo first-order and pseudo second-order kinetic parameters obtained from their respective linear plots log(qeqt) against t and t/qe against t, respectively are tabulated in Table 1. The calculated adsorption capacities (qcale) from pseudo first order kinetic equation are close to the experimental (qexpe) values. Furthermore, R2 values obtained for pseudo first-order kinetic equation are higher than pseudo second-order kinetic equation. These results revealed that the ongoing adsorption reaction proceeds via a pseudo-first order kinetic rather than a pseudo second-order kinetic mechanism. Similar type of adsorption behavior for CR adsorption onto bio-polymeric composite was reported by Chatterjee et al.,3 and Du et al.32

image file: c4ra05183a-f6.tif
Fig. 6 Effect of Contact time on CR adsorption.
Table 1 Kinetic parameters for CR adsorption onto starch–AlOOH–FeS2
Temp. (°C) q(exp)e (mg g−1) Pseudo-first order model Pseudo-second order model
q(cal)e (mg g−1) k1 (min−1) R2 q(cal)e (mg g−1) k2 (g mg−1 min−1) R2
25 261 266.07 11.51 × 10−3 0.994 333.33 3.474 × 10−5 0.985
40 313.2 305.49 9.212 × 10−3 0.996 500 1.550 × 10−5 0.990
50 346.2 320.62 9.212 × 10−3 0.987 500 19.704 × 10−5 0.984


Adsorption isotherms

The adsorption isotherms analysis has been performed at different concentrations and temperatures as shown in Fig. 7. The removal of CR increases with the increases in solution concentration and temperature. This can be due to the greater driving force through a higher concentration gradient at high concentration.2 An increase in the temperature, enhance the mobility of the dye molecule and decrease in the retarding forces acting on the diffusing dye molecules. This results in the increase in the adsorption capacity of starch–AlOOH–FeS2 for CR.33
image file: c4ra05183a-f7.tif
Fig. 7 Effect of concentration of CR adsorption.

In attempting to describe the adsorption isotherm of CR removal by starch–AlOOH–FeS2, Langmuir, Freundlich and Temkin isotherm models are tested. The linear equation for Langmuir,34 Freundlich35 and Temkin36 isotherm models, respectively are given as:

 
1/qe = (1/qmbCe) + 1/qm (4)
 
ln[thin space (1/6-em)]qe = ln KF + (1/n)ln[thin space (1/6-em)]Ce (5)
 
qe = B[thin space (1/6-em)]ln[thin space (1/6-em)]A + B[thin space (1/6-em)]ln[thin space (1/6-em)]Ce (6)
where qe (mg g−1) and Ce (mg L−1) are the CR adsorption capacity and concentration at equilibrium, qm (mg g−1) and b (L mg−1) are Langmuir constant. Freundlich isotherm constants KF (mg1−1/n L1/ng−1) and n are representing the capacity and intensity of adsorption. Temkin isotherm parameters, A (L mg−1) and B are related to the equilibrium binding constant and the heat of adsorption, respectively. The parameters for Langmuir, Freundlich and Temkin isotherms are calculated for their respective linear plots 1/qe vs. 1/Ce, ln[thin space (1/6-em)]qe vs. ln[thin space (1/6-em)]Ce and qe vs. ln[thin space (1/6-em)]Ce. From the Table 2, it can be seen that the values of correlation coefficient (R2) for the Freundlich isotherm model are higher than the Langmuir and Temkin adsorption models, revealing that experimental equilibrium data fits better with Freundlich isotherm. This reveled that adsorption of CR occur through multilayer formation on the heterogeneous surface of starch–AlOOH–FeS2composite.37 Moreover, the magnitude of Freundlich constant (n) should be more than one for favorable adsorption process.38,39 The values of n (Table 2) are greater than one at all the studied temperatures revealing the feasible adsorption of CR onto starch–AlOOH–FeS2 composite.

Table 2 Adsorption isotherm parameters for CR adsorption onto starch–AlOOH–FeS2
Temp. (°C) Langmuir isotherm model Freundlich isotherm model Temkin isotherm model
qm (mg g−1) b (L mg−1) R2 RL KF (mg1−1/n L1/ng−1) n R2 A (L mg−1) B R
30 333.33 0.032 0.951   44.880 2.288 0.970 0.226 78.06 0.955
40 333.33 0.040 0.957   45.422 2.293 0.980 0.274 81.66 0.970
50 333.33 0.166 0.912   70.035 3.246 0.956 1.455 61.65 0.923


Adsorption thermodynamics

The adsorption of CR onto starch–AlOOH–FeS2 increases from 298 mg g−1 to 346 mg g−1 with the increase in solution temperature from 30 °C to 50 °C as shown in Fig. 7. These results indicate that the adsorption process is endothermic in nature. Adsorption thermodynamic parameters for CR removal, standard Gibbs free energy change (ΔG°), standard entropy change (ΔS°) and standard enthalpy change (ΔH°) are calculated from following equations:
 
ΔG° = −RT[thin space (1/6-em)]ln[thin space (1/6-em)]Kc (7)
 
Kc = (Cae/Ce) (8)
 
ln[thin space (1/6-em)]Kc = (ΔS°/R) − (ΔH°/RT) (9)
where, Kc and R (8.314 J mol−1 K−1) are the distribution coefficient and gas constant, respectively. T is the temperature (K). Cae and Ce are the equilibrium concentration of CR (mg g−1) on adsorbent and in solution, respectively. The value of ΔH° and ΔS° calculated from the slope and intercept of ln[thin space (1/6-em)]Kc verses 1/T are in Table 3. As can be seen from Table 3, the values of ΔG° are negative indicating the feasibility and spontaneous nature of adsorption process for all the studied temperatures. The positive value of ΔH° and ΔS° confirming the endothermic process and increase in randomness at solid–solution interface.40 To a certain extent, the forces involve in the adsorption process can be determined by the magnitude of ΔH° and ΔG°. For the physisorption, the ΔG° value should be lower than −20 kJ mol−1 and those for chemisorptions ranges from −80 to −400 kJ mol−1. The ΔG° values for CR adsorption are in the range of −0.891 to −2.172 kJ mol−1, indicating that physical forces are involved in CR adsorption onto starch–AlOOH–FeS2. The energy (ΔH°) associated with physisorption: van der Waals forces (4–10 kJ mol−1), hydrophobic interaction (5 kJ mol−1), hydrogen bonding (2–40 kJ mol−1), coordination exchange (40 kJ mol−1), dipole bond forces (2–29 kJ mol−1), and electrostatic interaction (20–80 kJ mol−1), while for chemisorptions bond strength can be ranges from 80 to 450 kJ mol−1 (ref. 40 and 41). The value of ΔH° is 17 kJ mol−1 confirming the involvement of physical forces in CR adsorption.
Table 3 Thermodynamic parameters for CR adsorption onto starch–AlOOH–FeS2
Temperature (°C) ΔG° kJ mol−1 ΔH° kJ mol−1 ΔS° J mol−1 K−1 R2
30 −0.891      
40 −1.358 17.002 59.137 0.948
50 −2.173    


Conclusion

This article described the facile method for the synthesis of starch–AlOOH–FeS2 nanocomposite for the high efficient removal of CR from aqueous solution. The prepared material shows good surface area with the mesoporous structure. FTIR analysis shows that the NH, NH2 and SO3 groups are involved in the interaction of dye on the surface of starch–AlOOH–FeS2. The adsorption results demonstrate that CR removal is highly dependent on the solution pH, contact time, concentration and temperature. Adsorption isotherm study revealed that surface of starch–AlOOH–FeS2 is heterogeneous and multilayer adsorption of CR supposes to occur on the adsorbent surface. Thermodynamic study suggested that adsorption of CR onto starch–AlOOH–FeS2 is spontaneous and endothermic in nature. The magnitude of standard Gibbs free energy change (ΔG°), and standard enthalpy change (ΔH°) revealed that the physical forces are involved in the adsorption of CR onto starch–AlOOH–FeS2 surface. These results therefore suggested that starch–AlOOH–FeS2 can be considered as a potential material for the scavenging of CR from aqueous solution.

References

  1. M. Dong, Q. Lin, D. Chen, X. Fu, M. Wang, Q. Wu, X. Chen and S. Li, RSC Adv., 2013, 3, 11628–11633 RSC.
  2. R. Ahmad and R. Kumar, Appl. Surf. Sci., 2010, 257, 1628–1633 CrossRef CAS PubMed.
  3. S. Chatterjee, M. W. Lee and S. H. Woo, Bioresour. Technol., 2010, 101, 1800–1806 CrossRef CAS PubMed.
  4. W. Cai, Y. Hu, J. Chen, G. Zhang and T. Xia, CrystEngComm., 2012, 14, 972–977 RSC.
  5. Z. Y. Chen, H. W. Gao and Y. Y. He, RSC Adv., 2013, 3, 5815–5818 RSC.
  6. M. Solís, A. Solís, H. I. Pérez, N. Manjarrez and M. Flores, Process Biochem., 2012, 47, 1723–1748 CrossRef PubMed.
  7. E. Forgacsa, T. Cserhati and G. Oros, Environ. Int., 2004, 30, 953–971 CrossRef PubMed.
  8. M. Toor and B. Jin, Chem. Eng. J., 2012, 187, 79–88 CrossRef CAS PubMed.
  9. J. Wu, J. Wang, H. Li, Y. Du, K. Huang and B. Liu, J. Mater. Chem. A, 2013, 1, 9837–9847 CAS.
  10. A. Mahapatra, B. G. Mishra and G. Hota, Ceram. Int., 2013, 39, 5443–5451 CrossRef CAS PubMed.
  11. C. Wang, S. Huang, L. Wang, Z. Deng, J. Jin, J. Liu, L. Chen, X. Zheng, Y. Li and B. L. Su, RSC Adv., 2013, 3, 1699–1702 RSC.
  12. V. Janaki, K. Vijayaraghavan, B. T. Oh, K. J. Lee, K. Muthuchelian, A. K. Ramasamy and S. K. Kannan, Carbohydr. Polym., 2012, 90, 1437–1444 CrossRef CAS PubMed.
  13. P. R. Chang, P. Zheng, B. Liu, D. P. Anderson, J. Yu and X. Ma, J. Hazard. Mater., 2011, 186, 2144–2150 CrossRef CAS PubMed.
  14. S. Xu, J. Wang, R. Wu, J. Wang and H. Li, Chem. Eng. J., 2006, 117, 161–167 CrossRef CAS PubMed.
  15. N. Xu, C. Christodoulatos and W. Braida, Chemosphere, 2006, 62, 1726–1735 CrossRef CAS PubMed.
  16. A. Ozverdi and M. Erdem, J. Hazard. Mater., 2006, 137, 626–632 CrossRef PubMed.
  17. D. S. Han, J. K. Song, B. Batchelor and A. A. Waha, J. Colloid Interface Sci., 2013, 392, 311–318 CrossRef CAS PubMed.
  18. S. C. Hsiao, C. M. Hsu, S. Y. Chen, Y. H. Perng, Y. L.Chueh, L. J. Chen and L. H. Chou, Mater. Lett., 2012, 75, 152–154 CrossRef CAS PubMed.
  19. A. R. Pratt, H. W. Nesbitt and J. R. Mycroft, J. Geochem. Explor., 1996, 56, 1–11 CrossRef CAS.
  20. C. Gao, X. Y. Yu, R. X. Xu, J. H. Liu and X. J. Huang, ACS Appl. Mater. Interfaces, 2012, 4, 4672–4682 CAS.
  21. A. Uliniuca, M. Popa, E. Drockenmuller, F. Boisson, D. Leonard and T. Hamaidea, Carbohydr. Polym., 2013, 96, 259–269 CrossRef PubMed.
  22. Q. Lin, J. Pan, Q. Lin and Q. Liu, J. Hazard. Mater., 2013, 263, 517–524 CrossRef CAS PubMed.
  23. L. O. Filippov, V. V. Severov and I. V. Filippova, Int. J. Miner. Process., 2013, 123, 120–128 CrossRef CAS PubMed.
  24. R. K. Rath, S. Subramanian and T. Pradeep, J. Colloid Interface Sci., 2000, 229, 82–91 CrossRef CAS PubMed.
  25. Y. Yang, G. Wang, B. Wang, Z. Li, X. Jia, Q. Zhou and Y. Zhao, Bioresour. Technol., 2011, 102, 828–834 CrossRef CAS PubMed.
  26. F. Zhang, Y. Liu, Y. Cai, H. Li, X. Cai, I. Djerdj and Y. Wang, Powder Technol., 2013, 235, 121–125 CrossRef CAS PubMed.
  27. Y. Wang, W. Li and X. J. Chen, J. Mater. Chem. A, 2013, 1, 10720–10726 CAS.
  28. P. B. Gissinger, M. Alnot, J. E. Ehrhardt and P. Behra, Environ. Sci. Technol., 1998, 32, 2839–2845 CrossRef.
  29. K. Rezwan, L. P. Meier and L. J. Gauckler, Biomaterials, 2005, 26, 4351–4357 CrossRef CAS PubMed.
  30. S. Lagergren and K. Sven, Vetenskapsakad. Handl., 1898, 24, 1–39 Search PubMed.
  31. G. McKay and Y. S. Ho, Process Biochem., 1999, 34, 451–465 CrossRef.
  32. Q. Du, J. Sun, Y. Li, X. Yang, X. Wang, Z. Wang and L. Xia, Chem. Eng. J., 2014, 245, 99–106 CrossRef CAS PubMed.
  33. V. S. Mane and P. V. V. Babu, J. Taiwan Inst. Chem. Eng., 2013, 44, 81–88 CrossRef CAS PubMed.
  34. I. Langmuir, J. Am. Chem. Soc., 1918, 40, 1361–1403 CrossRef CAS.
  35. M. I. Temkin and V. Pyzhev, Acta Physiochem. USSR, 1940, 12, 217–222 Search PubMed.
  36. H. Freundlich, Uber die adsorption in loseungen, J. Phys. Chem., 1907, 57, 385–470 CAS.
  37. L. Yu, W. Xue, L. Cui, W. Xing, X. Cao and H. Li, Int. J. Biol. Macromol., 2014, 64, 233–239 CrossRef CAS PubMed.
  38. S. Dawood and T. K. Sen, Water Res., 2012, 46, 1933–1946 CrossRef CAS PubMed.
  39. V. Vimonses, S. Lei, B. Jin, C. W. K. Chow and C. Saint, Chem. Eng. J., 2009, 148, 354–364 CrossRef CAS PubMed.
  40. S. Liu, Y. Ding, P. Li, K. Diao, X. Tan, F. Lei, Y. Zhan, Q. Li, B. Huang and Z. Huang, Chem. Eng. J., 2014, 248, 135–144 CrossRef CAS PubMed.
  41. F. M. Machado, C. P. Bergmann, E. C. Lima, B. Royer, F. E. de Souza, I. M. Jauris, T. Calvete and S. B. Fagan, Phys. Chem. Chem. Phys., 2012, 14, 11139–1 RSC.

This journal is © The Royal Society of Chemistry 2014