Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Reusable bentonite clay: modelling and optimization of hazardous lead and p-nitrophenol adsorption using a response surface methodology approach

Mohamed Zbair*a, Zakaria Anfarb and Hassan Ait Ahsaineb
aLaboratory of Catalysis and Corrosion of Materials, Chouaïb Doukkali University, Faculty of Sciences El Jadida, BP. 20, El Jadida 24000, Morocco. E-mail: zbair.mohamed@gmail.com
bMaterials and Environment Laboratory, Ibn Zohr University, Faculty of Sciences, Agadir, Morocco

Received 4th January 2019 , Accepted 12th February 2019

First published on 15th February 2019


Abstract

In this work, bentonite clay (BC) calcined at 500 °C was used as an adsorbent (BC-500) for the adsorption of Pb2+ and p-nitrophenol. The ability of BC-500 for the removal of Pb2+ and p-nitrophenol has been investigated. The adsorption studies tailored well the pseudo-first-order and the Langmuir model for Pb2+ and p-nitrophenol both. In addition, the optimal removal of Pb2+ and p-nitrophenol was found at pH 5 for Pb2+ and pH 6 for p-nitrophenol. However, the change of temperature (20–60 °C) was found to have a negative effect on the adsorption process on BC-500. Based on the Dubinin–Radushkevich model the adsorption occurs via a physical process. Accordingly, the adsorption mechanism was proposed using N2-physisorption analysis before and after adsorption of Pb2+ and p-nitrophenol. The reusability of BC-500 was examined and the outcomes recommended that BC-500 had good potential as an economic and proficient adsorbent for Pb2+ or p-nitrophenol from contaminated water. Finally, the experimental Pb2+ and p-nitrophenol removal efficiency were found to be 90.93 ± 2.15% and 98.06 ± 1.87% while the predicted value by model equals 91.28 ± 1.68 and 97.24 ± 2.54, respectively, showing that the predicted model values are in good agreement with the experimental value.


1. Introduction

At present, pollution caused by organic and inorganic compounds is a major problem for the world and is likely to cause health hazards, harm to ecology, damage to structures or amenities, and interference with the legitimate use of water.1 Industries such as papermaking, electroplating, metal finishing, textiles, storage batteries, lead smelting, mining, plating, ceramics, leather, food, rubber, cosmetics, and plastics use different chemical compounds, which makes it possible to produce toxic effluents containing organic and inorganic contaminants.2 Lead (Pb) is one of the most toxic elements, even at low concentrations, because of its presence in automobile fuel and the subsequent emission into the atmosphere in the exhaust gases.3 Pb(II) metals affect the central nervous system, kidneys, liver, and gastrointestinal system, and may directly or indirectly cause diseases such as anemia, encephalopathy, hepatitis, and the nephrotic syndrome.4 p-Nitrophenol is another toxic pollutant (found particularly in the effluents from pesticides, pharmaceuticals, petrochemicals, and other industries) with an impact on many organisms that live in aquatic ecosystems. The presence of p-nitrophenol and its analogues in the environment has become a great concern in recent years because of their increased discharge, toxic nature and adverse effects on the receiving water bodies.5 Moreover, p-nitrophenol is responsible to cause degeneration in human proteins, tissue erosion, paralysis of the central nervous system, and damage to the liver, kidney and pancreas.6 As a priority toxin (US Environmental Protection Agency) because of its potential harm on aquatic life, plants and human beings even at low concentrations. p-Nitrophenol is a common constituent of contaminated soil, ground water or leachate.7,8 Hence, p-nitrophenol may cause severe environmental and public health problems, and it is mandatory to be treated before being discharged into the environment.

The wide usage of Pb(II) and p-nitrophenol in various industries have triggered the necessities of developing an efficient method to remove these pollutants from wastewaters.9 The most common methods for removal of these pollutants from industrial effluents include chemical precipitation, oxidation, and ion exchanges and adsorption etc.7,8,10 Adsorption process belongs to this category and it is found to be very effective and is applied in the liquid phase.1,11–20 Many adsorbents have been tested on the adsorption of Pb2+ and p-nitrophenol. For instance, Tran et al.11 have been reported 4.112 mmol g−1 of Pb2+ adsorption capacity using phenylalanine–Mg/AL(LDH). Yuanqing Huang et al.21 have utilized a nitrilotriacetic acid anhydride modified ligno-cellulosic material (NTAA-LCM) toward Pb2+ removal and indicated an adsorption capacity of 303.5 mg g−1 at 298 K. Also, Pb2+ was biosorbed using waste biomass (160 mg g−1 at 25 °C).22 For p-nitrophenol removal, Q. Zhang et al.23 have been found an adsorption capacity of 102 mg g−1 by ZnAl-layered double hydroxides. B. Liu et al.24 reached a capacity of 11 mg g−1 by coal fly ash. Furthermore, Muthanna J. Ahmed et al.25 have been utilized microporous activated carbon with an adsorption capacity of 185 mg g−1.

In recent years, clays are alternative solutions for acting as activated carbon. Clays have been used for thousands of years and continue to be among the leading industrial material because they are natural, earthy and fine-grained material.26 For a long time, clays have been also used for adsorption of various organic, inorganic substances due to its low cost and high removal efficiency (high surface area and exchange capacities).10,27–31 However, on our knowledge there is lack of information on the adsorption of Pb(II) and p-nitrophenol especially on bentonite clay and the effect of some variables on its adsorption capacity, optimization using response surface methodology and regeneration. The Response Surface Methodology (RSM) is a collection of mathematical and statistical techniques useful for the modeling and analysis of problems in which response of interest is affected by several variables (Kumar et al., 2017). Several published paper has been reported the utilization of RSM for the optimization and verification of scientific researches and industrial studies.32–34

In this research, bentonite clay (BC) from the region of Safi (Morocco) has been used as an adsorbent to remove Pb2+ and p-nitrophenol from aqueous solution. Adsorption experiments have been performed to look for equilibrium as well as adsorption kinetics. The experimental data were compared with by others studies and adsorption mechanism was proposed. RSM approach also was applied to determine the optimum removal efficiency and also to explain the relations between Pb2+ and p-nitrophenol removal and the pertinent parameters.

2. Experimental

2.1. Adsorbent's preparation

The adsorbent used in this study is the bentonite clay (BC) from the region of Safi (Morocco). The BC was primarily washed several times and then milled and sieved, acquiring the average diameter of 10 μm. The clay was subsequently calcined over 500 °C in a muffle for a period of 6 h. The calcination process is used in order to increase the adsorbent's stability. The obtained material was labeled: BC-500.

2.2. Characterization of adsorbent

BC-500 was analyzed using the X-ray diffraction (diffractometer Bruker D8 Advanced). Copper Kα radiation (λ = 1.5406 nm) produced at 50 kV and 20 mA. Infrared Spectra of BC-500 was recorded using Fourier Transform Spectrometer (Shimadzu FTIR-8400S) with a resolution of 4 cm−1 and 20 scans (400 to 4000 cm−1). The surface area of BC-500 was determined using Micromeritics ASAP 2010. Scanning Electron Microscopy analysis was done using (Hitachi S2500, Japan). Scanning electron microscopy (SEM) analyses were performed using FEI, Quanta 200-ESEM operated at 20 kV. X-Ray fluorescence elemental analysis of BC-500 was done using Philips spectrophotometer equipment. The PZC (point zero charge) is the pH at which the surface has zero net charge; known as pHpzc it is characteristic for amphoteric surfaces and affected by the type of surface sites and the structure. For the PZC determination 0.20 g of BC-20 were mixed with 50 mL of 0.01 M NaCl solution for 2 days. The starting solutions pH (2.0 to 12.0) was adjusted using HCl and NaOH. After 2 days, the final pH was measured. The point of zero charge of the AC-HP (pHpzc) is the point where the curve pHfinal versus pHinitial crosses the line pHinitial = pHfinal.16,35–37

2.3. Adsorbate

Stock solutions (1000 mg L−1) of tested adsorbate were prepared by dissolving p-nitrophenol (Sigma-Aldrich) in double distilled water and Pb2+ was prepared by dissolving Pb(NO3)2 (Sigma-Aldrich) in 1% HNO3 solution to avoid hydrolysis formation and precipitation. The working solution was prepared by diluting stock solution to appropriate volumes.

2.4. Adsorption experiments

Adsorption experiment was executed in duplicate, and the results were averaged. If the bias of the repeated experiment exceeded 15%, a triplicate run was performed.

The influence of the initial pH on adsorption of Pb2+ and p-nitrophenol onto BC-500 was observed by adjusting the pH value (2.0–12.0) at 20 °C by using 0.5 M HCl or 0.5 M NaOH prior to the experiments. Kinetics essays were carried out at three temperatures (20, 40, and 60 °C) in an Erlenmeyer by mixing 100 mg of BC-500 with 200 mL of Pb2+ and p-nitrophenol solution (concentration of 50 mg L−1). The system was constantly shaken in order to keep the mixture's homogeneity and aliquots were collected at different time intervals (Pb2+: 0–500 min and p-nitrophenol: 0–180 min). For adsorption equilibrium studies: various Pb2+ and p-nitrophenol solutions with different initial concentration at 20 °C (5–200 mg L−1 for Pb2+ and from 10–600 mg L−1 for p-nitrophenol). After the specified time, suspensions were filtered through filter paper Whatman 44. The concentration of the supernatant liquid of Pb2+ was determined by the absorption spectrophotometer AA-670. However, the concentration of p-nitrophenol was determined at λmax = 315 nm using UV-Vis spectrophotometer spectrometry (Shimadzu-2600). Diverse theoretical kinetic and isotherm models are applied to experimental data in order to determine the best-fitting model Table 1.

Table 1 Mathematical equations and models using in this study
Formula Utility Parameters Ref.
image file: c9ra00079h-t1.tif Adsorption capacity C0 (mg L−1) and Ce,t (mg L−1) are the initial and equilibrium concentrations, respectively. m (g) is the weight of adsorbent and V (L) is the volume 38
image file: c9ra00079h-t2.tif Removal efficiency 39
Qt = Qcal(1 − expK1t) Pseudo-first-order Qe (mg g−1) and Qt (mg g−1) are the adsorbed amounts at equilibrium and at times t (min), respectively. K1 (min−1): the rate constant; K2 (g mg−1 min−1): rate constant 40
image file: c9ra00079h-t3.tif Pseudo-second-order 41
image file: c9ra00079h-t4.tif Intraparticle diffusion Kip (mg g−1 min−1/2): rate coefficient; C: thickness of the boundary layer 42
image file: c9ra00079h-t5.tif Langmuir isotherm KL: direct measure of the intensity of the adsorption process; Qm: maximum adsorption capacity 43
image file: c9ra00079h-t6.tif Adsorption feasibility The adsorption process can be de fined as irreversible (RL = 0), favorable (RL between 0 and 1), linear (RL = 1) or unfavorable (RL < 1) 44
image file: c9ra00079h-t7.tif Freundlich isotherm KF: adsorption capacity; n: adsorption behavior 45
image file: c9ra00079h-t8.tif Dubinin–Radushkevich isotherm QDR (mg g−1) the theoretical adsorption capacity; KDR (mol2 kJ−2) the constant related to adsorption energy and ε the Polanyi potential; EDR (kJ mol−1) is the mean free energy of adsorption 46–48
image file: c9ra00079h-t9.tif
image file: c9ra00079h-t10.tif
ΔG° = −RT[thin space (1/6-em)]ln[thin space (1/6-em)]Kc Gibbs free energy ΔG°: Gibbs free energy change; Kc: an equilibrium constant (dimensionless); R: gas constant; T: temperature 49–51
image file: c9ra00079h-t11.tif Van't Hoff ΔS°: entropy change; ΔH°: enthalpy change


2.5. Response surface methodology (RSM) technique

Among the methods used to optimize and modeling the adsorption process is the experimental methodology designs, more precisely the optimization by the RSM.15 This method has attracted the attention of researchers in recent years.32–34 Response surface methodology (RSM) was used to obtain optimum operational parameters.52,53 The experimental levels for lead and p-nitrophenol adsorption were presented in Tables 2 and 3, respectively. The experimental results, as well as the CCD matrix, were shown in Table 4. Therefore, the quadratic polynomial equation that links the response lead and p-nitrophenol removal with three studied parameters were presented in equations below.
image file: c9ra00079h-t12.tif
where Y is the predicted response lead or p-nitrophenol adsorption, b0 is the intercept coefficient, bi the linear terms, bii the squared terms, bij the interaction terms, and X1, X2 and X3 represent the coded independent variables of pH, temperature and concentration, respectively.
Table 2 Experimental levels for p-nitrophenol adsorption
Variable Factor Unit −1.68(α) −1 0 1 +1.68(α)
X1 pH 2.63 4 6 8 9.36
X2 Temperature °C 11.59 15 20 25 28.40
X3 Concentration mg L−1 39.90 44 50 56 60.09


Table 3 Experimental levels for lead adsorption
Variable Factor Unit −1.68(α) −1 0 1 +1.68(α)
X1 pH 1.63 3 5 8 8.36
X2 Temperature °C 11.59 15 20 25 28.40
X3 Concentration mg L−1 39.90 44 50 56 60.09


Table 4 Experimental results of lead and p-nitrophenol adsorption using CCD matrix
No. exp. X1 X2 X3 Lead adsorption removal% p-Nitrophenol adsorption removal%
1 −1 −1 −1 32.52 48.23
2 1 −1 −1 80.54 72.71
3 −1 1 −1 43.30 51.21
4 1 1 −1 81.62 63.62
5 −1 −1 1 36.46 45.45
6 1 −1 1 67.50 66.60
7 −1 1 1 33.41 48.48
8 1 1 1 58.56 56.82
9 α 0 0 3.44 10.20
10 α 0 0 61.47 30.21
11 0 α 0 76.82 77.85
12 0 α 0 75.84 73.20
13 0 0 α 73.48 88.86
14 0 0 α 52.02 80.29
15 0 0 0 91.68 97.64
16 0 0 0 92.92 97.74
17 0 0 0 91.48 96.44


3. Results and discussion

3.1. Characterization of BC-500

XRD pattern of the BC-500 exhibited mostly smectite described by the reflection of d(001) = 1.54 nm and minor impurities of quartz (Fig. 1a). Moreover, XRD of BC-500 presented the specific reflections consistent to d values of 1.54 nm (001), 0.50 nm (003), 0.37 nm (004) and 0.29 nm (005) for the basal values for smectites.54 However, the reflexion of d(003) = 0.44 nm, d(110) = 0.25 nm, and d(210) = 0.16 nm is attributed to the presence of montmorillonite.55,56 The reflection of d(060) = 0.14 nm is linked to a dioctahedral smectite in which the high intensity of d060 reflection designated a large size of the coherent domains along the b direction.54,57
image file: c9ra00079h-f1.tif
Fig. 1 (a) XRD pattern, (b) FTIR spectra, (c) N2 physisorption isotherm and (d) the surface morphology of BC-500.

FTIR spectra can reveal the surface functional groups on BC-500 surface qualitatively based on the typical absorbed energy for each bond in certain groups.58 The surface functional groups in BC-500 are depicted in Fig. 1b. The peak located at 3637 cm−1 is ascribed to the groups stretching vibrations OH coordinated octahedral layer to Al(Mg)–O–H.59,60 At 3434 cm−1 the peak corresponds to OH stretching of structural hydroxyl groups. The peak at 1604 cm−1 is associated with H–O–H deformation of water. However, the peaks observed at around 998 and 749 cm−1 are linked to Si–O–Si and Si–O stretching of silica and quartz.61 The peaks at 503 and 429 cm−1 are assigned to Al–O–Si and Si–O deformation, respectively.56

Nitrogen physisorption isotherm of BC-500 presented in Fig. 1c, represent type IV adsorption isotherm with type H3 hysteresis loops which are characteristics of slit-like pore materials. The noticeable hysteresis loop can be seen on the BC-500 isotherm, signifying the existence of capillary condensation and the presence of mesopores. The BET surface area, total pore volume, and pore size of BC-500 are 101.2 m2 g−1, 0.26 cm3 g−1, and 4.8 nm, respectively. The surface morphology of BC-500 (Fig. 1d) shows porous and heterogeneous structures. This result is also in accord with the specific surface area of BC-500.

Elementary chemical analysis by X-ray fluorescence (Table 5) displays that Al2O3 and SiO2 are the principal elements with a ratio of SiO2/Al2O3 equal to 2.11. This value and the low content of K2O (0.20%) designate that the BC-500 used in this work belongs to the family of smectite.62 Moreover, the existence of exchangeable cations such as Ca2+, Na+, and Mg2+ indicates that the analyzed BC-500 has a high cation exchange capacity. These outcomes are entirely consistent with those stated by Er-ramly and Ider.63

Table 5 Elementary chemical analysis by X-ray fluorescence
SiO2 Al2O3 MgO Na2O Fe2O3 TiO2 CaO K2O P2O5 LOI Total
54.11 25.62 5.12 1.02 1.20 0.29 2.62 0.20 0.03 9.48 99.69


3.2. Adsorption of Pb2+ and p-nitrophenol

3.2.1. Effect of pH. It is well recognized that the pH of the adsorbate solution plays an essential role in the adsorption process.64–66 For additional understanding, we conducted adsorption of Pb2+ and p-nitrophenol onto BC-500 at various pH values.

The removal efficiency increased from 13.4 to 91.74% for Pb2+ and 16.2–97.64% for p-nitrophenol by increasing the solution pH from 2.0 to 5.0 for Pb2+ and 2.0 to 6.0 for p-nitrophenol. However, the adsorption starts decreasing by 16.2 and 26.4% at pH 12.0 for Pb2+ and p-nitrophenol, respectively (Fig. 2a). The surface of the BC-500 has the point of zero charges (pHpzc) at about 5.89 (Fig. 2b). The decrease in H3O+ with increasing pH in the acidic solution might increase the electrostatic attraction between the positively charged surface of the BC-500 and the partially negative charge of the p-nitrophenol molecules, consequently increasing the adsorption. The BC-500 surface starts to develop a partial negative charge at pH higher than pHpzc (5.89) and the p-nitrophenol molecules exists as deprotonated from (p-nitrophenolate anion) at pH > pKa = 7.15,67 this forms an electrostatic repulsions between the negatively charges p-nitrophenol molecules and negative surface of BC-500, which is consistent with previous studies.5,68,69


image file: c9ra00079h-f2.tif
Fig. 2 (a) effect of pH solution and (b) the point of zero charge (pHpzc) of BC-500.

At pH value between 2.0 to 4.0, the low removal of Pb2+ may be due to the competition with H+ ions on active sites of BC-500. Furthermore, the decrease in the removal efficiency at pH values (6.0–12.0) might be linked to the formation of Pb(OH)2 and an insoluble precipitate of Pb2+.70 In most of the stated studies regarding heavy metal ions adsorption onto various adsorbent, the adsorption removal displays a maximum for an initial pH at around 5.70–74

3.2.2. Effect of contact time and temperature. Contact time effect was performed in 200 mL of Pb2+ and p-nitrophenol solutions of 50 mg L−1 with varying contact time at optimal pH and three different temperatures (20, 40, and 60 °C). Fig. 3 displays the effect of different contact times on the adsorption of Pb2+ and p-nitrophenol on BC-500. The adsorption was rapid and approximately accomplished (adsorption equilibrium) in 140 min (Fig. 3a) and 30 min (Fig. 3b) at 20 °C for Pb2+ and p-nitrophenol, respectively. However, increasing temperature influence negatively on adsorption capacity from 92 to 38 mg g−1 in the case of Pb2+ and from 195 to 100 mg g−1 for p-nitrophenol, respectively. This might be related to desorption produced by an increase in the available thermal energy which modify the position of adsorption–desorption equilibrium.66 This result shows that a low temperature favours the adsorption of for Pb2+ and p-nitrophenol on BC-500.
image file: c9ra00079h-f3.tif
Fig. 3 Effect of contact time and temperature: (a): Pb(II); (b): p-nitrophenol.
3.2.3. Adsorption kinetics. The adsorption kinetics of Pb2+ and p-nitrophenol onto BC-500 were also examined. Pseudo-first (PFO), second order (PSO), and intraparticle diffusion (IPD) models were employed to correlate the kinetics data (Table 6). Adsorption kinetics data on BC-500 along with the linked model by means of the PFO, PSO, and IPD models were presented in Fig. 4.
Table 6 Different constants adsorption kinetics
    Pseudo-first-order Pseudo-second-order
Qe,exp (mg g−1) Qe,cal (mg g−1) K1 (min−1) R2 Qe,cal (mg g−1) K2 (g mg−1 min−1) R2
Pb2+ 20 °C 91.74 90.66 ± 1.07 0.043 0.988 97.96 ± 1.84 0.0006 0.980
40 °C 71.74 75.09 ± 1.76 0.031 0.976 67.51 ± 7.12 0.0005 0.930
60 °C 38.04 38.78 ± 1.45 0.008 0.982 47.90 ± 2.24 0.0002 0.973
p-Nitrophenol 20 °C 195.36 197.66 ± 7.29 0.066 0.961 219.13 ± 12.07 0.0004 0.948
40 °C 169.71 168.08 ± 3.04 0.044 0.993 196.66 ± 5.86 0.0003 0.990
60 °C 100.36 104.53 ± 5.18 0.034 0.993 128.18 ± 12.23 0.0003 0.939

    Intraparticle diffusion
Kip C R2
Pb2+ 20 °C 3.47 33.39 ± 8.19 0.705
40 °C 2.99 18.48 ± 4.08 0.877
60 °C 1.94 1.079 ± 1.36 0.861
p-Nitrophenol 20 °C 13.91 54.18 ± 21.22 0.740
40 °C 13.44 24.10 ± 1.78 0.876
60 °C 8.84 5.93 ± 1.42 0.828



image file: c9ra00079h-f4.tif
Fig. 4 Non-linear kinetics models adsorption of Pb2+ and p-nitrophenol onto BC-500.

To approve the best model that describe the adsorption kinetics of Pb2+ and p-nitrophenol onto BC-500, we usually compare the value of correlation coefficients and standard error of estimate (SEE) allowing the correlation between experimental data and the model-predicted values. The nonlinear plots of Pb2+ and p-nitrophenol adsorption kinetics and the calculated kinetic parameters are given in Fig. 4 and Table 6. As can be seen, the correlation coefficients found from PFO model was higher compared with R2 gotten from PSO model. Additionally, the SEE of PFO model was the lowest one compared to PSO model. Consequently, PFO model describe well the adsorption data of Pb2+ and p-nitrophenol onto BC-500 at all studied temperature. Besides, the experimental adsorption capacity of Pb2+ and p-nitrophenol onto BC-500 were very close to the calculated one by PFO model (Table 6).

In other hand, the increase in temperature decrease the acceleration of Pb2+ and p-nitrophenol adsorption on BC-500, which was certificated by the decrease in the second-order rate constants (K2). This might be due to the increase of solution viscosity at high temperature and thus diminish the mass transfer and diffusion of the Pb2+ and p-nitrophenol inside BC-500.

Moreover, Fig. 4 displays the curve-fitting plots of IPD model for Pb2+ and p-nitrophenol onto BC-500 adsorption. The values of the intercept C of Pb2+ and p-nitrophenol give an idea about the boundary layer thickness: the larger intercept, the greater is the boundary layer effect in adsorption by BC-500. In the meantime, the plots do not pass over the origin, this specifies that there lies some degree of boundary layer control and the IPD is not the lonely rate-controlling step, but also other processes may control the rate of adsorption of Pb2+ and p-nitrophenol.75 Hence, the IPD model is not appropriate for describing the Pb2+ and p-nitrophenol removal from water onto BC-500. It can be concluding that the PFO model described well the kinetics data of Pb2+ and p-nitrophenol compared to PFO and IPD models.

3.2.4. Adsorption isotherms. To investigate the most suitable correlation for the equilibrium data of Pb2+ and p-nitrophenol adsorption onto BC-500 (Fig. 5), three non-linear models were used, Langmuir, Freundlich, and Dubinin–Radushkevich (D–R). The calculated values for Langmuir, Freundlich, and Dubinin–Radushkevich parameters were summarized in Table 4. From Table 4, the adsorption of Pb2+ and p-nitrophenol onto BC-500 were well fitted to the Langmuir isotherm model with high R2, which endorse the homogeneous distribution of active sites on the surface of BC-500.4 In addition, the nature of the adsorption process on BC-500 was investigated by calculating the values of RL and Freundlich constant n. As a result, the adsorption Pb2+ and p-nitrophenol onto BC-500 were favorable (RL between 0 and 1, and 1/n parameter was smaller than 1).17 Based on Langmuir isotherm (Table 7), the maximum adsorption capacities Qmax was equal 94.01 mg g−1 (for Pb2+ adsorption) and 284.25 mg g−1 (for p-nitrophenol adsorption).
image file: c9ra00079h-f5.tif
Fig. 5 Non-linear isotherms models used for adsorption of Pb2+ and p-nitrophenol onto BC-500.
Table 7 Isotherms parameters of Pb2+ and p-nitrophenol adsorption onto BC-500
  Langmuir isotherm Freundlich isotherm
Qmax (mg g−1) KL (L mg−1) RL R2 KF ((mg g−1) (mg L−1)n) n R2
Pb2+ 94.01 2.97 0.002 0.942 46.24 5.93 0.760
p-Nitrophenol 284.25 16.05 0.0001 0.971 145.01 8.01 0.886

  Dubinin–Radushkevich isotherm
QDR (mg g−1) KDR (mol2 kJ2) R2 E (kJ mol−1)
Pb2+ 91.63 0.260 0.928 1.386
p-Nitrophenol 289.82 138.74 0.910 0.060


To elucidate the physical and chemical adsorption nature of Pb2+ and p-nitrophenol onto BC-500, Langmuir and Freundlich isotherms are insufficient, for this purpose D–R isotherm is normally used.30 The mean free energy of adsorption EDR (EDR of transference for 1 mol of solute from solution to the surface of CB-500 (ref. 76)), for Pb2+ and p-nitrophenol adsorption onto BC-500 was found equal to 1.386 kJ mol−1 (for Pb2+ adsorption) and 0.060 kJ mol−1 (for p-nitrophenol adsorption) (Table 7). These values are matching with other earlier studies using natural clay for adsorption in aqueous solutions.30 Following the D–R isotherm model, |E| values between 1–8 kJ mol−1 specify that the adsorption is a physical process, whereas, when |E| are between 8.0 kJ mol−1 and 16 kJ mol−1 the adsorption occurs at an ion-exchange-type surface.77 In our study, Pb2+ and p-nitrophenol adsorption onto BC-500, we found the values of EDR is in the range of 1–8 kJ mol−1, and consequently, mean that in our case the adsorption occurs via a physical process on BC-500. The adsorption capacity of BC-500 is compared to several adsorbents presented in Table 8. We found that the sorption capacity of BC-500 is better compared to others adsorbents.

Table 8 Comparison of the maximum adsorption capacity of BC-500 with other adsorbents
Adsorbent Pb2+ p-Nitrophenol Ref.
BC-500 94 284 This study
Phaseolus hulls activated carbon 21 78
Ulva lactuca (AAC) 83 79
Apricot stone 23 80
Pecan shell 64 81
Nano illite/smectite clay 2 82
ZnAl-layered double hydroxides 102 23
Microporous activated carbon 185 25
Hypercrosslinked magnetic polymer 153 83
Coal fly ash 11 24
H3PO3-treated coal 256 84


3.2.5. Thermodynamic parameters. The comportment of adsorption (i.e., physical or chemical) can be understood over the study of adsorption thermodynamics. The thermodynamic parameters (ΔG°, ΔH°, and ΔS°) can be calculated by the Van't Hoff approach.50,85 The Van't Hoff equation and Gibbs energy equations (Table 9) were used to calculate the thermodynamic parameters of Pb2+ and p-nitrophenol adsorption on BC-500 (Fig. 6). From Table 9, the negative ΔG° values at all studied temperatures for Pb2+ and p-nitrophenol stipulate that the adsorption occurrence occurred spontaneously. Furthermore, the negative (−ΔS°) values disclose that the organization of Pb2+ and p-nitrophenol at the solid/solution interface through the adsorption process on BC-500 develops less random when the temperature increases. A negative value of ΔH° indicate that the adsorption process occurred exothermically. An exothermic process is obviously related to physical adsorption (physisorption) with the occurrence of relatively weak interactions (i.e., van der Waals force).86
Table 9 Thermodynamic parameters for Pb2+ and p-nitrophenol adsorption onto BC-500
  ΔH (kJ mol−1) ΔS (kJ mol−1 K−1)   ΔG (kJ mol−1)
293 K 313 K 333 K
Ketoprofen −104.499 −0.201 KC 184[thin space (1/6-em)]568[thin space (1/6-em)]580 3[thin space (1/6-em)]415[thin space (1/6-em)]381.2 1[thin space (1/6-em)]126[thin space (1/6-em)]960.8
  −46.365 −39.148 −38.580
p-Nitrophenol −132.467 −0.303 KC 123[thin space (1/6-em)]915[thin space (1/6-em)]710.3 362[thin space (1/6-em)]868.435 199[thin space (1/6-em)]963.6695
  −463.657 −33.313 −33.792



image file: c9ra00079h-f6.tif
Fig. 6 Van't Hoff plot for Pb2+ and p-nitrophenol adsorption onto BC-500.
3.2.6. Regeneration of BC-500 and adsorption mechanism. Renewability is an essential concern to evaluate the adsorbent performance for wastewater treatment.14,87 In order to distinguish the renewability of BC-500, the used BC-500 was washed out and then reused in the new experiments to evaluate its cycle performance. In this work, 0.5 M HCl and ethanol were used for Pb2+/BC-500 and p-nitrophenol/BC-500 regeneration, respectively. The outcomes are illustrated in Fig. 7a. The Pb2+/BC-500 regenerated by 0.5 M HCl revealed a slight decrease (−5.82%) from 91.74% to 86.40 after 5 cycles. Likewise, the p-nitrophenol/BC-500 regenerated by ethanol-exposed a small decreased (−5.54%). This suggests that BC-500 is an outstanding recycling adsorption material.
image file: c9ra00079h-f7.tif
Fig. 7 (a) Regeneration of BC-500 and (b) adsorption mechanism of Pb2+ and p-nitrophenol onto BC-500.

To advance the investigation of adsorption comportment of Pb2+ and p-nitrophenol onto BC-500, Fig. 7b displays the nitrogen physisorption isotherm for BC-500 before and after adsorption of Pb2+ and p-nitrophenol. After the adsorption of Pb2+ and p-nitrophenol onto BC-500, the surface area decreases noticeably from 101.2 to 94.4 m2 g−1 and from 101.2 to 88.6 m2 g−1, respectively. The total pore volume also diminishes from 0.26 to 0.19 cm3 g−1 for Pb2+/BC-500 and from 0.26 to 0.14 cm3 g−1, which proves that the adsorbed Pb2+ and p-nitrophenol choose to be stored in the pores of BC-500 which is consistent with the outcomes of Dubinin–Radushkevich model.

3.2.7. Optimization of adsorption by response surface methodology (RSM). In this part of work, response surface methodology technique was used to optimize the adsorption process of Pb2+ and p-nitrophenol in order to modelled the reaction pathways taking into account the influence of different parameters as a function of lead and p-nitrophenol removal efficiency onto BC-500.88,89 Therefore, the analysis of variance results (ANOVA) for lead or p-nitrophenol adsorption were presented in Table 10. Based on this statically analysis we conclude that:
Table 10 ANOVA analyses
NOM Coefficient P-value
Lead adsorption
Model <0.0001
b0 92.081 <0.0001
b1 17.583 <0.0001
b2 −0.130 0.709
b3 −5.722 <0.0001
b1-1 −21.246 <0.0001
b2-2 −5.733 <0.0001
b3-3 −10.535 <0.0001
b1-2 −1.949 0.00293
b1-3 −3.769 <0.0001
b2-3 −2.981 0.000248
Lack of fit   0.26
R2 0.999  
RAdj2 0.998  
[thin space (1/6-em)]
p-Nitrophenol adsorption
Model <0.0001
b0 53.532 <0.0001
b1 17.249 <0.0001
b2 10.802 0.0200
b3 8.605 0.0206
b1-1 −2.998 <0.0001
b2-2 −2.689 <0.0001
b3-3 −4.416 <0.0001
b1-2 −0.720 0.0218
b1-3 −2.542 0.2040
b2-3 −4.090 0.9070
Lack of fit   0.104
R2 0.997  
RAdj2 0.994  


• The term ‘P-value’ less than 0.05 implies that designed model is significant for both lead or p-nitrophenol.

• The correlation coefficient (R2) and the adjusted correlation coefficient (RAdj2) prove the existence of a good correlation between data.

• The coefficients significations of models were studied at the 5% confidence level and their P-values less than 0.05, except b2 for lead adsorption and b2-1, b1-3 for p-nitrophenol adsorption.

In addition, Pareto diagram presented in Fig. 8 indicates that the parameters including square effect of pH, pH and square effect of concentration represent more than 90% of response variation in the case of lead adsorption. In the case of p-nitrophenol adsorption, pH, temperature and concentration influence the removal efficiency by 54.03%, 21.30% and 13.52, respectively. As results, Pareto analysis showed that the influence of pH was very high compared with other factors in both systems, which confirms the effect of electrostatic attractions mechanism during the adsorption of lead and p-nitrophenol.


image file: c9ra00079h-f8.tif
Fig. 8 Pareto chart for (a) lead and (b) p-nitrophenol adsorption.

The developed models for lead or p-nitrophenol adsorption were giving below using the quadratic polynomial equation.

R% (lead) = 92.081 + 17.583XpH − 5.722Xconcentration − 21.246XpH × XpH − 5.733XTemperature × XTemperature − 10.535Xconcentration × Xconcentration −1.949XpH × Xconcentration − 3.769XpH × Xconcentration − 2.981Xconcentration × XTemperature

R% (p-nitrophenol) = 53.532 + 17.249XpH + 10.802XTemperature + 8.605Xconcentration − 2.998XpH × XpH − 2.689XTemperature × XTemperature − 4.416Xconcentration × Xconcentration − 0.720XpH × XTemperature − 2.542XXpH × Xconcentration − 4.090Xconcentration × XTemperature

The both equations were used to elaborate the response surfaces which allows to obtain the optimization parameters of lead and p-nitrophenol adsorption and well understood the process (Fig. 9 and 10). Based on RSM presentation, the optimum conditions in term of lead and p-nitrophenol removal were showed in Table 11. Under these conditions the experimental and predicted removal by the obtained model showing a high correlation, which indicate the importance of this study in terms of modeling and optimization.


image file: c9ra00079h-f9.tif
Fig. 9 RSM presentations for lead adsorption on different possible plans.

image file: c9ra00079h-f10.tif
Fig. 10 RSM presentations for p-nitrophenol adsorption on different possible plans.
Table 11 Experimental and predicted removal by the model for lead and p-nitrophenol adsorption
Adsorption of lead (pH = 5.0, temperature = 20.0 °C, concentration = 50.0 mg L−1)
Predicted Experimental
91.28 ± 1.68 90.93 ± 2.15

Adsorption of p-nitrophenol (pH = 6.0, temperature = 20.0 °C, concentration = 50.0 mg L−1)
Predicted Experimental
97.24 ± 2.54 98.06 ± 1.87


4. Conclusion

In this study, BC-500 as the low-cost adsorbent is talented to remove Pb2+ and p-nitrophenol from contaminated water very efficiently. The adsorption kinetic studies demonstrate that the removal of Pb2+ and p-nitrophenol is a rapid process and the adsorption process follows the pseudo-first-order model, demonstrating that Pb2+ and p-nitrophenol has a strong affinity on the BC-500 surface. It was found that the experimental isotherm data can be tailored well to the Langmuir equilibrium equation and BC-500 has very high Pb2+ and p-nitrophenol adsorption capacity of approximately 94 mg g−1 and 284 mg g−1 at 20 °C, respectively. Increasing temperature (20–60 °C) has an effect negative on adsorption capacity Pb2+ and p-nitrophenol. The removal of Pb2+ and p-nitrophenol depend clearly on the pH of the solution. The study on adsorption mechanism displays that the adsorption of Pb2+ and p-nitrophenol on BC-500 is dominated by physisorption process. Finally, the optimized parameters by the RSM method showed a good agreement between the predicted and experimental results.

Conflicts of interest

Authors declare that they have no conflict of interest.

Acknowledgements

This work was supported by the Ibn Zohr University, Faculty of Sciences-Morocco.

References

  1. T. S. Anirudhan and M. Ramachandran, J. Colloid Interface Sci., 2006, 299, 116–124 CrossRef CAS PubMed.
  2. T. S. Anirudhan and M. Ramachandran, Process Saf. Environ. Prot., 2015, 95, 215–225 CrossRef CAS.
  3. A. H. Jawad, R.A. Rachid, M. A. M. Ishak and L. D. Wilson, Desalin. Water Treat., 2016, 1, 25194–25206 CrossRef.
  4. L. Wang, J. Zhang, R. Zhao, Y. Li, C. Li and C. Zhang, Bioresour. Technol., 2010, 101, 5808–5814 CrossRef CAS PubMed.
  5. B. Zhang, F. Li, T. Wu, D. Sun and Y. Li, Colloids Surf., A, 2015, 464, 78–88 CrossRef CAS.
  6. A. dos Santos, M. F. Viante, D. J. Pochapski, A. J. Downs and C. A. P. Almeida, J. Hazard. Mater., 2018, 355, 136–144 CrossRef CAS PubMed.
  7. S. Dutta, J. K. Basu and R. N. Ghar, Sep. Purif. Technol., 2001, 21, 227–235 CrossRef CAS.
  8. S. Haydar, M. A. Ferro-García, J. Rivera-Utrilla and J. P. Joly, Carbon, 2003, 41, 387–395 CrossRef CAS.
  9. W. S. W. Ngah and S. Fatinathan, J. Environ. Manage., 2010, 91, 958–969 CrossRef CAS PubMed.
  10. S. S. Tahir and R. Naseem, Sep. Purif. Technol., 2007, 53, 312–321 CrossRef CAS.
  11. H. N. Tran, C.-C. Lin and H.-P. Chao, Sep. Purif. Technol., 2018, 192, 36–45 CrossRef CAS.
  12. H. J. He, Z. H. Xiang, X. J. Chen, H. Chen, H. Huang, M. Wen and C. P. Yang, Int. J. Environ. Sci. Technol., 2018, 15, 1491–1500 CrossRef CAS.
  13. M. Zbair, Z. Anfar, H. Khallok, H. A. Ahsaine, M. Ezahri and N. Elalem, Fullerenes, Nanotubes, Carbon Nanostruct., 2018, 26, 433–442 CrossRef CAS.
  14. Z. Anfar, M. Zbair, H. A. Ahsaine, M. Ezahri and N. El Alem, Fullerenes, Nanotubes, Carbon Nanostruct., 2018, 26, 389–397 CrossRef CAS.
  15. M. Zbair, Z. Anfar, H. A. Ahsaine, N. El Alem and M. Ezahri, J. Environ. Manage., 2018, 206, 383–397 CrossRef CAS PubMed.
  16. M. Zbair, H. A. Ahsaine and Z. Anfar, J. Cleaner Prod., 2018, 202, 571–581 CrossRef CAS.
  17. Z. Anfar, R. El Haouti, S. Lhanafi, M. Benafqir, Y. Azougarh and N. El Alem, J. Environ. Chem. Eng., 2017, 5, 5857–5867 CrossRef CAS.
  18. H. Ait Ahsaine, M. Zbair, Z. Anfar, Y. Naciri, R. El haouti, N. El Alem and M. Ezahri, Mater. Today Chem., 2018, 8, 1–12 CrossRef.
  19. I. Anastopoulos, A. Bhatnagar, B. Hameed, Y. S. Ok and M. Omirou, J. Mol. Liq., 2017, 240, 179–188 CrossRef CAS.
  20. L. Bulgariu and D. Bulgariu, J. Cleaner Prod., 2018, 197, 875–885 CrossRef CAS.
  21. Y. Huang, C. Yang, Z. Sun, G. Zeng and H. He, RSC Adv., 2015, 5, 11475–11484 RSC.
  22. C. Yan, Y. Chunping, H. Huijun, Z. Guangming, Z. Kun and Y. Zhou, J. Environ. Eng., 2016, 142, C4015001 CrossRef.
  23. Q. Zhang, C. Peter Okoli, L. Wang and T. Liang, Desalin. Water Treat., 2015, 55, 1575–1585 CrossRef CAS.
  24. B. Liu, F. Yang, Y. Zou and Y. Peng, J. Chem. Eng. Data, 2014, 59, 1476–1482 CrossRef CAS.
  25. M. Ahmed and S. Theydan, Adsorptive removal of p-nitrophenol on microporous activated carbon by FeCl3 activation: equilibrium and kinetics studies, 2014, vol. 55 Search PubMed.
  26. A. Sari, M. Tuzen and M. Soylak, J. Hazard. Mater., 2007, 144, 41–46 CrossRef CAS PubMed.
  27. G. Bereket, A. Z. Aroǧuz and M. Z. Özel, J. Colloid Interface Sci., 1997, 187, 338–343 CrossRef CAS PubMed.
  28. M. Tuzen, E. Melek and M. Soylak, J. Hazard. Mater., 2006, 136, 597–603 CrossRef CAS PubMed.
  29. N. L. D. Filho, W. L. Polito and Y. Gushikem, Talanta, 1995, 42, 1031–1036 CrossRef CAS PubMed.
  30. S. Veli and B. Alyüz, J. Hazard. Mater., 2007, 149, 226–233 CrossRef CAS PubMed.
  31. S. S. Tahir and N. Rauf, Chemosphere, 2006, 63, 1842–1848 CrossRef CAS PubMed.
  32. K. Zhao, Y. Cheng, H. Liu, C. Yang, L. Qiu, G. Zeng and H. He, RSC Adv., 2015, 5, 66013–66023 RSC.
  33. L. Qiu, Y. Cheng, C. Yang, G. Zeng, Z. Long, S. Wei, K. Zhao and L. Luo, RSC Adv., 2016, 6, 17036–17045 RSC.
  34. C. Yang, K. Zhao, Y. Cheng, G. Zeng, M. Zhang, J. Shao and L. Lu, Sep. Purif. Technol., 2016, 163, 153–161 CrossRef CAS.
  35. T. K. M. Prashantha Kumar, T. R. Mandlimath, P. Sangeetha, P. Sakthivel, S. K. Revathi, S. K. Ashok Kumar and S. K. Sahoo, RSC Adv., 2015, 5, 108034–108043 RSC.
  36. D. Mohan, A. Sarswat, V. K. Singh, M. Alexandre-Franco and C. U. Pittman Jr, Chem. Eng. J., 2011, 172, 1111–1125 CrossRef CAS.
  37. J. S. Noh and J. A. Schwarz, J. Colloid Interface Sci., 1989, 130, 157–164 CrossRef CAS.
  38. J. Wang, C. P. Huang, H. E. Allen, D. K. Cha and D. W. Kim, J. Colloid Interface Sci., 1998, 208, 518–528 CrossRef CAS PubMed.
  39. V. K. Garg, R. Gupta, A. B. Yadav and R. Kumar, Bioresour. Technol., 2003, 89, 121–124 CrossRef CAS PubMed.
  40. S. Lagergren, K. Sven. Vetenskapsakad. Handl., 1898, 24, 1 Search PubMed.
  41. G. McKay, Process Biochem., 1999, 34, 451 CrossRef.
  42. J. C. Weber and W. J. Morris, J. Sanit. Eng. Div., Am. Soc. Civ. Eng., 1963, 89, 31–60 Search PubMed.
  43. I. Langmuir, J. Am. Chem. Soc., 1916, 38, 2221–2295 CrossRef CAS.
  44. S. Fan, Y. Wang, Z. Wang, J. Tang, J. Tang and X. Li, J. Environ. Chem. Eng., 2017, 5, 601–611 CrossRef CAS.
  45. H. Freundlich, Z. Phys. Chem., 1906, 57, 385 CAS.
  46. L. V. Dubinin and M. M. Radushkevich, Proc. Acad. Sci. USSR, Phys. Chem. Sect., 1947, 55, 331–333 Search PubMed.
  47. G. F. Cerofolini, Surf. Sci., 1975, 51, 333–335 CrossRef CAS.
  48. M. Ghasemi, M. Naushad, N. Ghasemi and Y. Khosravi-fard, J. Ind. Eng. Chem., 2014, 20, 2193–2199 CrossRef CAS.
  49. E. C. Lima, A. Hosseini-Bandegharaei, J. C. Moreno-Piraján and I. Anastopoulos, J. Mol. Liq., 2019, 273, 425–434 CrossRef CAS.
  50. I. Anastopoulos and G. Z. Kyzas, J. Mol. Liq., 2016, 218, 174–185 CrossRef CAS.
  51. P. S. Ghosal and A. K. Gupta, J. Mol. Liq., 2017, 225, 137–146 CrossRef CAS.
  52. Z. Anfar, M. Zbair, H. A. Ahsaine, M. Ezahri and N. E. Alem, Fullerenes, Nanotubes, Carbon Nanostruct., 2018, 26, 389–397 CrossRef CAS.
  53. M. Zbair, K. Ainassaari, Z. El Assal, S. Ojala, N. El Ouahedy, R. L. Keiski, M. Bensitel and R. Brahmi, Environ. Sci. Pollut. Res., 2018, 25, 35657–35671 CrossRef CAS PubMed.
  54. F. G. Alabarse, R. V. Conceição, N. M. Balzaretti, F. Schenato and A. M. Xavier, Appl. Clay Sci., 2011, 51, 202–208 CrossRef CAS.
  55. X. Wang, S. Lu, L. Chen, J. Li, S. Dai and X. Wang, J. Radioanal. Nucl. Chem., 2015, 306, 497–505 CrossRef CAS.
  56. M. El Miz, H. Akichoh, D. Berraaouan, S. Salhi and A. Tahani, Am. J. Chem., 2017, 7, 105–112 Search PubMed.
  57. L. M. Calarge, A. Meunier and M. L. L. Formoso, J. South Am. Earth Sci., 2003, 16, 187–198 CrossRef.
  58. H. A. Ahsaine, M. Zbair and R. El Haouti, Desalin. Water Treat., 2017, 85, 330–338 CrossRef.
  59. J. Madejová, W. P. Gates and S. Petit, in Infrared and Raman Spectroscopies of Clay Minerals, ed. W. P. Gates, J. T. Kloprogge, J. Madejová and F. Bergaya, Elsevier, 2017, vol. 8, pp. 107–149 Search PubMed.
  60. C. T. Johnston, in Infrared and Raman Spectroscopies of Clay Minerals, ed. W. P. Gates, J. T. Kloprogge, J. Madejová and F. Bergaya, Elsevier, 2017, vol. 8, pp. 288–309 Search PubMed.
  61. M. L. Cantuaria, A. F. de Almeida Neto, E. S. Nascimento and M. G. A. Vieira, J. Cleaner Prod., 2016, 112, 1112–1121 CrossRef CAS.
  62. L. Bounab, K. Draoui, M. Ahrouch, M. Hadri, D. Bouchta and A. Barhoun, J. Mater. Environ. Sci., 2017, 8, 244–256 CAS.
  63. A. Er-ramly and A. Ider, Phys. Chem. News, 2011, 61, 112–119 CAS.
  64. H. N. Tran, C.-C. Lin, S. H. Woo and H.-P. Chao, Appl. Clay Sci., 2018, 154, 17–27 CrossRef CAS.
  65. I. Anastopoulos, I. Margiotoudis and I. Massas, Int. J. Phytorem., 2018, 20, 831–838 CrossRef CAS PubMed.
  66. M. Zbair, K. Ainassaari, A. Drif, S. Ojala, M. Bottlinger, M. Pirilä, R. L. Keiski, M. Bensitel and R. Brahmi, Environ. Sci. Pollut. Res., 2018, 25, 1869–1882 CrossRef CAS PubMed.
  67. D. Tang, Z. Zheng, K. Lin, J. Luan and J. Zhang, J. Hazard. Mater., 2007, 143, 49–56 CrossRef CAS PubMed.
  68. T.-H. Pham, B.-K. Lee and J. Kim, Process Saf. Environ. Prot., 2016, 104, 314–322 CrossRef CAS.
  69. G. Xue, M. Gao, Z. Gu, Z. Luo and Z. Hu, Chem. Eng. J., 2013, 218, 223–231 CrossRef CAS.
  70. M. Momčilović, M. Purenović, A. Bojić, A. Zarubica and M. Ranđelović, Desalination, 2011, 276, 53–59 CrossRef.
  71. A. Üçer, A. Uyanik and Ş. F. Aygün, Sep. Purif. Technol., 2006, 47, 113–118 CrossRef.
  72. K. Kadirvelu, C. Faur-Brasquet and P. Le Cloirec, Langmuir, 2000, 16, 8404–8409 CrossRef CAS.
  73. C. Faur-Brasquet, Z. Reddad, K. Kadirvelu and P. Le Cloirec, Appl. Surf. Sci., 2002, 196, 356–365 CrossRef CAS.
  74. S.-T. Lin, H. N. Tran, H.-P. Chao and J.-F. Lee, Appl. Clay Sci., 2018, 162, 443–453 CrossRef CAS.
  75. A. Sharma, Z. Syed, U. Brighu, A. B. Gupta and C. Ram, J. Cleaner Prod., 2019 DOI:10.1016/j.jclepro.2019.01.236.
  76. F. Granados-Correa and J. Jiménez-Becerril, J. Hazard. Mater., 2009, 162, 1178–1184 CrossRef CAS PubMed.
  77. Y. Zeng, Z. Zeng, T. Ju and F. Zhang, Microporous Mesoporous Mater., 2015, 210, 60–68 CrossRef CAS.
  78. M. M. Rao, D. K. Ramana, K. Seshaiah, M. C. Wang and S. W. C. Chien, J. Hazard. Mater., 2009, 166, 1006–1013 CrossRef CAS PubMed.
  79. W. M. Ibrahim, A. F. Hassan and Y. A. Azab, Egypt. J. Basic Appl. Sci., 2016, 3, 241–249 CrossRef.
  80. M. Kobya, E. Demirbas, E. Senturk and M. Ince, Bioresour. Technol., 2005, 96, 1518–1521 CrossRef CAS.
  81. R. R. Bansode, J. N. Losso, W. E. Marshall, R. M. Rao and R. J. Portier, Bioresour. Technol., 2003, 89, 115–119 CrossRef CAS PubMed.
  82. J. Yin, C. Deng, Z. Yu, X. Wang and G. Xu, Water, 2018, 10, 210 CrossRef.
  83. Y. Ma, Q. Zhou, A. Li, C. Shuang, Q. Shi and M. Zhang, J. Hazard. Mater., 2014, 266, 84–93 CrossRef CAS PubMed.
  84. S.-H. Lin and R.-S. Juang, J. Environ. Manage., 2009, 90, 1336–1349 CrossRef CAS PubMed.
  85. H. N. Tran, S. J. You, A. Hosseini-Bandegharaei and H. P. Chao, Water Res., 2017, 120, 88–116 CrossRef CAS PubMed.
  86. H. N. Tran, S.-J. You and H.-P. Chao, J. Environ. Chem. Eng., 2016, 4, 2671–2682 CrossRef CAS.
  87. A. Elouahli, M. Zbair, Z. Anfar, H. A. Ahsaine, H. Khallok, R. Chourak and Z. Hatim, Surf. Interfaces, 2018, 13, 139–147 CrossRef CAS.
  88. H. Ait Ahsaine, Z. Anfar, M. Zbair, M. Ezahri and N. El Alem, J. Chem., 2018, 1–14 Search PubMed.
  89. M. Zbair, Z. Anfar, H. A. Ahsaine and H. Khallok, RSC Adv., 2019, 9, 1084–1094 RSC.

This journal is © The Royal Society of Chemistry 2019