GET THE APP

The Utility of Novel Superabsorbent Core Shell Magnetic Nanocompo
Journal of Chromatography & Separation Techniques

Journal of Chromatography & Separation Techniques
Open Access

ISSN: 2157-7064

+44 1300 500008

Research Article - (2013) Volume 4, Issue 5

The Utility of Novel Superabsorbent Core Shell Magnetic Nanocomposites for Efficient Removal of Basic Dyes from Aqueous Solutions

Magda A Akl1*, Ayman Atta2, Abd ElFattah M Youssef1 and Mohammed A Ibraheim1
1Chemistry Department, Faculty of Science, Mansoura University, Mansoura, Egypt, E-mail: shaikvali2008@gmail.com
2Egyptian Petroleum Research Institute, Petroleum Application Department, Nasr City 11727, Cairo, Egypt, E-mail: shaikvali2008@gmail.com
*Corresponding Author: Magda A Akl, Chemistry Department, Faculty of Science, Mansoura University, Mansoura, Egypt Email:

Abstract

A series of novel superabsorbents Fe3O4/poly(Acrylamide-co-sodium acrylate) core shell magnetic nanogel based on crosslinked poly (sodium acrylate-co-acrylamide) are investigated for the removal of methylene blue from aqueous solutions. The superabsorbents are characterized by FTIR, HR-TEM and SEM. The adsorbents exhibited high sorption capacities toward basic dyes, viz.: 2167.802-2765.893 mg g-1. The adsorption kinetics followed a pseudo-second order equation. Equilibrium isotherm data are analyzed according to Langmuir and Freundlich equations. The thermodynamic parameters including ΔG°, ΔH° and ΔS° for the adsorption processes of MB on the nanocomposite hydrogel were also calculated, and the negative value of ΔG° indicated the spontaneous nature of adsorption.

Keywords: Methylene blue; Nanocomposite Hydrogel; Magnetic nanogel; Temperature

Introduction

AIn recent years, pollution from dye wastewater has become a serious environmental problem due to the vast and increasing uses of a variety of dyes. [1] Methylene blue (MB) is the most commonly used substance for dying cotton, wood and silk. It may cause some hazardous problems to human and animal health [2]. Amongst the numerous techniques of dye removal [3-10], adsorption is a wellknown equilibrium separation process and is an effective method for water decontamination applications [11,12].

Recently, there has been an increasing interest in the design and fabrication of magnetic polymeric particles consisting of one or more magnetic nanoparticles as a core with a coating matrix of polymer as a shell. The main advantage of magnetic nanocomposite particles (MCPs) over conventional polymer nanocomposite particles is that, because of their magnetic properties, they can be rapidly separated from the mixtures by magnetic extraction [13]. They have been tried extensively in various fields including industrial water treatment [14,15].

To our best knowledge, there is no literature focusing on the use of Fe3O4/ poly(sodium acrylate co acrylamide) core-shell magnetic nanogel for the removal of MB from aqueous solutions. The objective of this work is to develop crosslinked p(Am-co-Na Ac) nanocomposite hydrogel that contain different amounts of Fe3O4/p(Am-co-Na Ac) core–shell magnetic nanogels as potential effective adsorbents to remove MB from aqueous solution attributing to their following unique properties: (i) Fe3O4 increase mechanic strength to enhance the usable life. (ii) The p(Am-co-Na Ac) shells with nano-scaled size provide much larger specific surface area. (iii) The raw materials are low-cost and the synthetic approach is simple, which make these nanoparticles potentially commercializable. (iv) These nanoparticles do not cause environmental pollution during the course of treatment.

Experimental

Materials

All chemicals were of analytical grade and used as received, Iron (III) chloride hexahydrate (FeCl3.6H2O>99%), Iron (II) chloride tetrahydrate (FeCl2.4H2O>99%) were purchased from Merck. Acrylic acid (AA), acrylamide (AM) monomers and N,N-methylene-bis- (acrylamide)(MBA) as a crosslinker, Potassium peroxydisulfate (KPS) as initiator and N,N,N`,N`-tetramethylethylenediamine (TEMED) as an accelerator were purchased from Aldrich Chemical Company. MB was purchased from Sigma-Aldrich. Doubly Distilled Water (DDW) was used throughout the present work.

Preparations

Superparamagnetic Fe3O4 nanoparticles were prepared by chemical co-precipitation of iron (III) and iron (II) in an alkaline solution and then treating it under hydrothermal condition. In a typical experiment, 5.2 g of FeCl3.6H2O, 2 g of FeCl2.4H2O and 0.85 ml HCl (12 mol/L) were dissolved in 25 ml of DDW (degassed with nitrogen gas before use) to prepare a stock solution. 250 ml of 1.5 mol/L NaOH solution was heated to 80°C in a beaker. The stock solution was added drop wise under nitrogen gas protection and vigorous stirring. After the reaction, the mixture was cooled to room temperature with stirring, and the resulting magnetic Fe3O4 nanoparticles were separated magnetically and washed with DDW four times to remove excess NaOH and finally dispersed in 100 ml DDW to be used for nanogel preparation.

Core shell poly(acrylamide-co-sodium acrylate) magnetic nanogel was prepared by solution polymerization. In a typical experiment, 1.52 g acrylamide, 1.54 g acrylic acid and 0.198 g MBA (3 mol% with respect to total monomer amount) were dissolved in 72 ml DDW in an ice cooled bath followed by addition of drops of 10 M NaOH till pH=9. The reaction system was bubbling nitrogen gas to exclude the air inside the flask for 20 minute, Then 75 ml of magnetite ferrofluid (10 mg/ml) was added under vigorous stirring. Afterwards, 3 ml aqueous solution of KPS (2 wt% with respect to total monomer amount) and 150 μl of TEMED were added to the solution to start polymerization. The reaction was carried out at room temperature under N2 gas for 15 hour. After preparation of stable nanogel dispersion, the resultant polymer coated magnetic nanoparticles were collected with the aid of an external magnetic field, washed with DDW several times and finally dispersed in 45 ml DDW.

Sodium acrylate(Na-Ac) was prepared as following:12.01 g (0.3003 mol) NaOH and 10 ml water were added to a 50 ml Erlenmeyer flask equipped with a magnetic stirring bar, and the mixture was stirred to dissolution. The NaOH solution was carefully added to a 250 ml beaker containing 28.03 g (0.3893 mol) acrylic acid (exothermic reaction), with continuous stirring. The mixture was allowed to cool, 50 ml of acetone were added, and the precipitate vacuum filtered. The wet sodium acrylate was first air dried and subsequently dried in an oven at 55 C-60 C for 12-15 hours to obtain 27.31 g (96.8% yield) of sodium acrylate.

Superabsorbents nanocomposite hydrogels (Ac blank, Ac4% and Ac8%) containing 0, 4, 8 wt% of nanogel, respectively, were prepared by in situ free radical copolymerization of acrylamide and sodium acrylate in an aqueous dispersion of nanogel particles using MBA as crosslinker and KPS as the initiator. The total mass percentage of both monomers in the reaction mixture was 30%. In brief, to prepare a sample(S), 2.15 g of acrylamide, 2.84 g of sodium acrylate, 0.05 g of crosslinker MBA (1 wt % based on total monomer amount) were dissolved in water in an ice cooled bath under nitrogen atmosphere. Then, aqueous dispersion of magnetic nanogel particles was added to the mixture under vigorous stirring. Afterwards, 3 ml aqueous solution of initiator (0.05 g KPS, 1wt% with respect to total monomer amount) was added and the whole reaction mixture was transferred into a test tube (internal diameter 10 mm) and kept in an electric oven at 50°C for 10 hours. The hydrogel codes and complete compositions are listed in Table 1.

Sample code Monomers feed molar ratio MBA (g) KPS (g) Nanogel (wt %)
Na-AC Acm
Ac blank 90 10 0.05 0.05 0
Ac 4% 90 10 0.05 0.05 4
Ac 8% 90 10 0.05 0.05 8

Table 1: Composition of nanocomposite hydrogels.

The formed nanocomposite hydrogels were recovered carefully by breaking the test tubes and cut into small disks of 3 mm thickness and 10 mm diameter. These hydrogel discs were safely transferred into DDW and allowed to equilibrate over a period of 5 days (water was repeatedly changed every 10 h) to remove unreacted monomers, crosslinker, initiator and soluble or un-crosslinked polymers, etc. Finally, the sample was dried in air followed by overnight drying in an electric oven to get completely dried gels and stored in airtight plastic bags.

Characterization

FTIR spectra were analyzed with a Nicolet FTIR spectrophotometer using KBr in a wavenumber range of 4000-5000 cm-1 with a resolution accuracy of 4 cm-1. The images for magnetite and Fe3O4/p(Amco- Na Ac) nanocomposites were recorded using High resolution transmission electron microscopy (HR-TEM) (JEM-2100F, JEOL, Japan) at an acceleration voltage of 120 kV. The surface morphologies of nanocomposite hydrogels were investigated using scanning electron microscopy (SEM, model JSM-T 220A, JEOL, Japan) equipped with an EDX detector. Zeta potential measurements of nanocomposite hydrogels were measured at different pH values using Laser Zeta meter Malvern Instruments Model Zetasizer 2000.

Adsorption studies

Batch sorption experiments were done by shaking 0.02 g of nanocomposite with 50 ml aqueous solution of MB in 250 ml- Erlenmeyer flasks placed in a temperature controlled shaking water bath at different concentrations (between 100 and 1600 mg/l), pHs (between 2 and 12), ionic strength (between 0.005 and 0.3 mole/l), temperatures (between 25°C and 45°C) and sorbent doses (between 0.01 and 0.05 g) at a constant shaking rate of 125 rpm. The concentration of MB in the residual solution was analyzed spectrophotometrically at 662 nm. The amounts of dye removed by superabsorbents (qe) and the percent of extraction (%E) are calculated using the following equations:

qe = Co−Ce / m ×100       (1)

%E = (Co −Ce ) / Co×100 (2)

Where qe is the amount of dye adsorbed (mg/g). Co and Ce are the initial and equilibrium liquid-phase concentrations of dye (mg/g), respectively. V is the volume of the solution (l), and m is the weight of the sorbent used (g).

Results and Discussion

Characterization of adsorbents

The chemical structure of the prepared magnetite nanoparticles was confirmed by FT-IR analysis. The adsorption of AM, AA-Na and MBA as a ligand onto Fe3O4 colloidal nanoparticles clusters has been studied by means of FTIR. Figure 1 shows the FT-IR spectra of Fe3O4 nanoparticles, Ac blank, Ac 4% and Ac 8% nanocomposites. The FTIR spectrum of Fe3O4 (Figure 1a) shows characteristic absorption band at 586 cm-1 corresponding to the Fe–O bond. Also, the presence of a band in the region of 3250-3600 Cm-1 indicates the presence of water of crystallization in the magnetite sample. The small peaks in the region 1500-1400 cm-1 can also be correlated to the different modes of bonded water molecules existing in the ferrofluid. In FT-IR spectrum of Ac 4% and Ac 8% nanocomposites (Figures 1b and 1c) there is shifting in ferrite peak by 15 Cm-1. The shifting as well as broadening of the peaks can be attributed to interaction of 3d orbit of ferrite with N atom in AM to form coordinate bond [13]. In the FT-IR spectra of Ac blank, Ac 4% and Ac 8% (Figures 1b-d), the peaks found at 3400, 1650, and 1600 cm-1 indicate N–H stretching, C=O stretching and N–H bending of the amide bonds, respectively, which are characteristics of the–CONH group in the AM. In addition, the peak at 1411 cm-1 is due to the–C–N stretching, and the 765-710 cm-1 band represents the out of plane bending of the weak bond N–H. Based on the above observations, the Fe3O4 nanoparticles are located in the polymer shell. The HR-TEM micrograph of the magnetite nanoparticles, Figure 2a, suggests that Fe3O4 nanoparticles are spherical in shape with mean diameter of 8 nm. Figure 2b confirms that the magnetic Fe3O4 nanoparticles coated with poly(Acrylamide-co-sodium acrylate) polymer are nanocapsules of spherical shape with an average diameter of 11 nm. On encapsulation with polymer, the size of the nanocapsules is increased from 8 to 11 nm. This is an indirect confirmation of the encapsulation of Fe3O4 magnetic core with a 3 nm thick polymer shell.

chromatography-separation-techniques-FTIR-spectrum

Figure 1: FTIR spectrum of (a) Fe3O4 nanoparticles, (b) Ac blank, (c) Ac 4%, and (d) Ac 8% superabsorbents.

chromatography-separation-techniques-core-shell

Figure 2: TEM of (a) Fe2O3 nanoparticles, and (b) core-shell magnetic nanogel.

The HR-TEM micrograph of the magnetite nanoparticles, Figure 2a, suggests that Fe3O4 nanoparticles are spherical in shape with mean diameter of 8 nm. Figure 2b confirms that the magnetic Fe3O4 nanoparticles coated with poly(Acrylamide-co-sodium acrylate) polymer are nanocapsules of spherical shape with an average diameter of 11 nm. On encapsulation with polymer, the size of the nanocapsules is increased from 8 to 11 nm. This is an indirect confirmation of the encapsulation of Fe3O4 magnetic core with a 3 nm thick polymer shell.

The HR-TEM micrograph of the magnetite nanoparticles, Figure 2a, suggests that Fe3O4 nanoparticles are spherical in shape with mean diameter of 8 nm. Figure 2b confirms that the magnetic Fe3O4 nanoparticles coated with poly(Acrylamide-co-sodium acrylate) polymer are nanocapsules of spherical shape with an average diameter of 11 nm. On encapsulation with polymer, the size of the nanocapsules is increased from 8 to 11 nm. This is an indirect confirmation of the encapsulation of Fe3O4 magnetic core with a 3 nm thick polymer shell.

The morphologies of Ac blank, Ac 4%, Ac 8% nanocomposite, as observed by SEM, are shown in Figure 3. The pulverized Ac blank, obtained from the radical chain polymerization in aqueous solution, was essentially a smooth sheet free of any significant particulate nature (Figure 3a). The magnetic nanogel loadings at 4 and 8 %wt were immobilized in the Ac 4% and Ac 8% nanocomposite showed the nanogel particles to be well dispersed and so were found to reside on the surface and the inner depths (Figures 3b and 3c). The nanogel particle distribution increased with increasing nanogel content. The figures reveal that small spherical particles of 30 to 50 nm are imbedded within the polymer matrix with absence of phase separation or agglomerations of magnetic particles on the surface.

chromatography-separation-techniques-SEM-images

Figure 3: SEM images of (a) Ac blank, (b) Ac 4%, (c) Ac 8% nanocomposites, respectively.

Zeta potential of Ac blank, Ac 4 % and Ac 8% has been investigated; the results are shown in Figure 4. It could be clearly seen that, with increasing the amount of nanogel on nanocomposite adsorbents the zeta potential shift to more negative values. So, Ac blank has less negative potential whereas Ac 8% has the highest negative potential. The presence of negatively charged carboxylate groups on the surface of nanogel is the main reason for shift in zeta potential to more negative values. This implies that the adsorption capacity of adsorbents toward positively charged MB will increase in the order Ac 8% Ac 4% Ac blank due to presence of more negatively charged active sites in that order. Also, the pH causes a dramatic change in zeta potential. As pH increase from pH 4 to pH 8, zeta potential increase from -20.7 to -29.7 mv, from -21.9 to -31.9 mv and from -22.8 to -36.1 mv for Ac blank, Ac 4% and Ac 8% respectively. With increasing the pH, zeta potential become more negative due to ionization of carboxylic group of acrylic acid. In contrast to that, as pH decrease, zeta potential decrease due to conversion of ?COO? to ?COOH and protonation of NH2 group of acrylamide.

chromatography-separation-techniques-Zeta-potential

Figure 4: Zeta potential of Ac blank, Ac 4% and Ac 8% nanocomposites at different pH.

Adsorption studies

Effect of pH: Figure 5 shows the effect of pH on the removal of MB onto Ac blank, Ac 4% and Ac 8% nanocomposites from aqueous solutions. The adsorption capacity of MB onto nanocomposite hydrogels increased significantly with increasing pH. The high adsorption capacity is due to the strong electrostatic interaction between the nanocomposite hydrogels and cationic dye molecules. Moreover, when pH value is increased, the carboxylic groups of the nanocomposites become ionized and the electrostatic repulsion between the molecular chains is predominated which leads to more expansion of the network and high adsorption of MB. When the pH increases from pH=3 to pH=4, The amount of MB adsorbed at equilibrium (qe) increase from 570.38 to 1389.89 mg/g, from 590.64 to 1396.41 mg/g and from 614.68 to 1443.31 mg/g, for Ac blank, Ac 4% and Ac 8% nanocomposites, respectively. At acidic pH, the lower dye removal is probably due to the presence of excess H+ ions competing with positively charged MB molecules for the sorption sites of sorbents.

chromatography-separation-techniques-pH-values

Figure 5: Effect of the pH values on adsorption capacity of MB by Ac blank, Ac 4% and Ac 8% nanocomposites. (Conditions: Cₒ=600 mg/l, T=25°C; adsorbent dose=0.02 g/50 ml).

Effect of initial dye concentration: The effect of initial concentration on the removal of MB by Ac blank, Ac 4% and Ac 8% nanocomposites is shown in Figure 6. The adsorption capacities of the sorbents increase with an increase in initial dye concentration and the adsorption at different concentrations is rapid in the initial stages and gradually decreases with the progress of adsorption until the equilibrium is reached. This may be attributed to the fact that, the higher the initial MB concentration, the greater the driving force of the concentration gradient at solid-liquid interface which cause an increase of the amount of MB adsorbed on the adsorbent [16]. When the initial concentration of MB increases from 100 to 1000 mg/l at 25°C, the amount of MB adsorbed at equilibrium (qe), increase from 244.54 to 2153.33 mg/g, from 244.87 to 2374.87 mg/g and from 248.23 to 2729.16 mg/g for Ac blank, Ac 4% and Ac 8%, nanocomposites, respectively. At higher MB concentrations adsorption capacity reached a plateau indicating saturation of the available binding sites on the adsorbent.

chromatography-separation-techniques-initial-dye

Figure 6: Effect of initial dye concentration on the removal of MB by Ac blank, Ac 4% and Ac 8% nanocomposites. (Conditions: T=25°C; adsorbent dose=0.02 g/50 ml; time=2 h).

Effect of contact time: The adsorption capacity of Ac blank, Ac 4% and Ac 8% nanocomposites increased rapidly with the increase of contact time from 0 to 20 min (Figure 7) and more than 90% of the equilibrium adsorption capacity for MB occurred within 15 min. After 60 min, the adsorption capacity became constant and the adsorption reached equilibrium. Therefore, 60 min was selected as the contact time for the adsorption of MB onto the composites under our experimental conditions.

chromatography-separation-techniques-adsorption-capacity

Figure 7: Effect of the contact time on adsorption capacity of Ac blank, Ac 4% and Ac 8% nanocomposites. (Conditions: T=25°C; adsorbent dose=0.02 g/50 ml; pH=7).

Effect of adsorbent dosage: When the sorbent dose increases from 0.01 to 0.02 g (Figure 8), the percent dye removals by Ac blank, Ac 4% and Ac 8% nanocomposites increase from 43.33% to 86.06%, from 47.30% to 95.21% and from 55.01% to 99.01%, respectively. This can be simply attributed to the increased sorbent surface area and availability of more sorption sites [17]. The decrease in adsorption capacity with increase in adsorbent dosage is due to the high number of unsaturated adsorption sites.

chromatography-separation-techniques-sorbent-dosage

Figure 8: Effect of sorbent dosage on the removal of MB by Ac blank, Ac 4% and Ac 8% nanocomposites. (Conditions: T=25°C; Cₒ=1000 mg/l; time=2 h).

Effect of ionic strength: Figure 9 presents the effect of ionic strength on the uptake of MB. The adsorption capacity decreased with the increase in ionic strength. As the concentration of NaCl ions increased from 0.01 to 0.1 M, the adsorption capacities decreased from 2164.08 to 2053.215 mg/g, from 2137.772 to 1825.543 mg/g and from 2454.712 to 2136.364 mg/g for Ac blank, Ac 4% and Ac 8% nanocomposites, respectively. Further increase in ionic strength above 0.2 M lowers the adsorption capacities sharply until 0.4 M where any further increase cause steep decrease in adsorption capacities because of salting out of dye molecules from aqueous solution.

chromatography-separation-techniques-ionic-strength

Figure 9: Effect of ionic strength on the removal of MB by Ac blank, Ac 4% and Ac 8% nanocomposites. (Conditions: T=25°C; adsorbent dose=0.02 g/50 ml; Cₒ=1000 mg/l; time=2 h).

Effect of temperature: The adsorption capacity of the nanocomposites increased with increasing of the temperature from 25 to 45°C (Table 2). This is attributed to the increase in the kinetic energy of MB with increasing temperature and the flexibility of the hydrogel increases as well. Consequently, the rate of diffusion of MB molecules across the external boundary layer and in the internal pores in the nanocomposites increases with temperature [18], and the amount of MB uptake increases. When the temperature increased from 25°C to 45°C, the maximum amounts of MB removed by Ac blank, Ac 4% and Ac 8% nanocomposites are found to be increased from 2167.802 to 2176.677 mg/g, from 2377.078 to 2390.351 mg/g and from 2759.505 to 2765.893 mg/g, respectively.

Sample code Qe(mg g-1)
25°C 35°C 45°C
Ac blank 2167.802 2171.987 2176.677
Ac 4% 2377.078 2384.123 2390.351
Ac 8% 2759.505 2763.716 2765.893

Table 2: Effect of temperature on maximum adsorption capacities of MB by Ac blank, Ac 4%, Ac 8%.

Adsorption isotherms

Adsorption is usually described through isotherms, that is, the amount of adsorbate on the adsorbent as a function of its pressure (if gas) or concentration (if liquid) at constant temperature. The two most common isotherm equations namely, Langmuir and Freundlich are tested here [19]. The Langmuir equation can be written as follows:

qe = Q0Kl Ce/1+ Kl Ce (3)

Where qe is the amount of dye adsorbed on nanocomposites at equilibrium (mg/g), Q0 is the maximum adsorption capacity (mg/g), Kl is the Langmuir constant, and Ce is the concentration of dye solution at equilibrium (mg/L). The linear form of Langmuir equation is:

Ce / q e = 1/ Q0 K l+ Ce/ Q0                      (4)

The monolayer adsorption capacity, Q and the Langmuir constant Kl can be obtained from the linear plot of Ce/qe against Ce. An essential characteristic of Langmuir isotherm can be expressed by a dimensionless constant called equilibrium parameter, RL [20], defined by:

Rl = 1/1+ Kl Ce            (5)

Where C0 is the highest initial dye concentration (mg/L), RL values indicate the type of isotherm to be either unfavorable (RL>1), linear (Rl=1), favorable (0L<1) or irreversible (RL=0) [21].

The Freundlich equation is given by [19]:

q e = Kf C1/n e               (6)

Where KF is the Freundlich constant which is a comparative measure of the adsorption capacity of the adsorbent, and n is an empirical constant which gives valuable information about the isotherm shape. 1/n values indicate the type of isotherm to be irreversible (1/n=0), favorable (0<1/n<1) and unfavorable (1/n>1).

The Freundlich parameters be obtained from the following linearized equation:

By linear plotting ln qe as the function of ln Ce, the values of Kf and n can be obtained from the slope and the intercept of the plot.

ln qe = lnKf +1/ n lnCe  (7)

Figures 10 and 11 show the linear plot of Langmuir and Freundlich isotherms for adsorption of MB on nanocomposites. The model parameters obtained by applying Langmuir and Freundlich models to the experimental data are given in Tables 3 and 4. The regression coefficients R2 obtained from Langmuir model are closer to 1 suggesting that the Langmuir isotherm fits better with the adsorption of MB on Ac blank, Ac 4% and Ac 8% nanocomposites. In addition, the qmax values for the adsorption of MB onto the nanocomposites calculated from the Langmuir model are all the same as the experimental data. Also, RL values obtained are in the range of 0.1195-0.2534, thereby confirming that the adsorption is a favorable process. It can be concluded that the monolayer Langmuir adsorption isotherm is more suitable to explain the adsorption of MB onto the proposed nanocomposites.

chromatography-separation-techniques-Langmuir-plot

Figure 10: Langmuir plot for the adsorption of MB at (a) 25°C, (b) 35°C and (c) 45°C.

chromatography-separation-techniques-Freundlich-plot

Figure 11: Freundlich plot for the adsorption of MB at (a) 25°C, (b) 35°C and (c) 45°C.

Sample Code T(°C) Qe,exp (mg g-1) Qmax,Fitted(mg g-1) KL (L mg-1) Ln KL (L mol-1) R2
  Ac blank 25
35
45
2167.802
2171.987
2176.677
2173.435
2174.811
2178.910
0.4745
0.9123
1.7867
11.9301
12.5838
13.2559
0.9999
1
1
  Ac 4% 25
35
45
2377.078
2384.123
2390.351
2403.5053
2401.1795
2399.2322
0.191732
0.28921
0.610744
11.02395
11.435
12.18253
0.9996
0.9998
0.9999
 Ac 8% 25
35
45
2759.505
2763.716
2765.893
2771.7953
2774.9720
2772.8021
0.4673
0.5404
0.9319
11.91473
12.06022
12.60508
0.9999
0.9999
0.9999

Table 3: Langmuir isotherm parameters and correlation coefficient for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different temperatures.

Sample Code T(°C) 1/n Ln Kf R2
  Ac blank 25
35
45
0.03912
0.03425
0.03177
7.44614
7.48023
7.49916
0.6006
0.63252
0.5780
  Ac 4% 25
35
45
0.23825
0.21875
0.1957
6.49468
6.61173
6.75972
0.55408
0.55314
0.58564
  Ac 8% 25
35
45
0.02935
0.02785
0.02652
7.75168
7.76279
7.773
0.88331
0.88524
0.87547

Table 4: Freundlich isotherm parameters and correlation coefficient for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different temperatures.

Adsorption kinetics

The pseudo-first order kinetic model was suggested by Lagergren its linear form can be formulated as

log[(qe − qt )] = log qe −K1t / 2.303        (8)

Where qe is the adsorption uptake of MB at time t (mol g-1) and k1 (min-1) is the rate constant of the pseudo-first-order adsorption. A plot of log (qe-qt) versus t should be linear; the parameters k1 and R2 (correlation coefficient) calculated from the data (Figure 12) are listed in Table 5. The large differences between the experimental qe values (qe,exp) and the calculated qe values (qe,cal) indicate that the pseudo-first order kinetic model was poor fit for the adsorption processes of Ac blank, Ac 4% and Ac 8% nanocomposites for MB.

chromatography-separation-techniques-Pseudo-first-order

Figure 12: Pseudo-first-order kinetic model.

Sample code Cₒ qe, cal K R2
Ac blank 1000
600
300
1816.519
524.3484
133.3828
0.05087
0.05382
0.04581
0.98013
0.96391
0.96653
Ac 4% 1000
600
300
1791.637
1053.416
205.0548
0.03422
0.04998
0.0415
0.98686
0.95768
0.96405
Ac 8% 1000
600
300
1907.129
1323.64
377.3896
0.02211
0.02978
0.03167
0.9828
0.99697
0.95513

Table 5: First order kinetic parameters for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different concentrations.

The Ho’s pseudo-second-order kinetic model [22] can be expressed as:

t / qt = 1/ K2qe 2 + t /qe (9)

Where k2 (g mol-1min-1) is the rate constant of pseudo-secondorder adsorption. Figure 13 shows the plot of t/q versus t for MB onto Ac blank, Ac 4% and Ac 8% nanocomposites. The correlation coefficients(R2), Table 6, for the pseudo-second order kinetic model are all over 0.9999, moreover, the qe,cal values for the pseudo-second order kinetic model are all consistent with the qe,exp values. These findings suggest that the adsorption processes of Ac blank, Ac 4% and Ac 8% nanocomposites for MB can be well described by the pseudo-second order kinetic model

chromatography-separation-techniques-Pseudo-second-order

Figure 13: Pseudo-second-order kinetic model.

Sample code Cₒ k2 (g mg−1 min−1) qe,cal. (mg g−1) R2
Ac blank 1000
600
300
0.4943
2.2354
7.6909
2428.5638
1510.7657
751.8797
0.99926
0.99971
0.99992
Ac 4% 1000
600
300
0.2753
0.7798
4.8753
2653.0055
1595.0693
763.3588
0.99948
0.99836
0.99996
Ac 8% 1000
600
300
0.1334
0.2634
1.2721
2775.8039
1682.6802
793.6508
0.99585
0.99577
0.9978

Table 6: Second order kinetic parameters for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different concentrations.

The intra-particle diffusion parameter, Kp (mmol.g-1 h-0.5) is defined by equation:

qt = Kpt0.5 +C                (10)

Where kp is the intra-particle diffusion rate constant (mmol.g-1 h-0.5) and C is a constant. According to this model qt versus t0.5 should be linear if intra-particle diffusion is involved in the adsorption process [23]. From Eq. (5), if pore diffusion is the rate limiting step, then a plot of qt against t0.5 must give a straight line with a slope that equals kp and the intercept value c represents the resistance to mass transfer in the external liquid film.

Figure 14 shows the plot of qt against t0.5 of Ac blank, Ac 4% and Ac 8% nanocomposites for MB at three different initial concentrations; the plots are multi-linear, containing at least three linear segments which indicate that three steps occur during adsorption process. The first sharper portion is the transport of dye molecules from the bulk solution to the adsorbent external surface by diffusion through the boundary layer (film diffusion). The second portion is the diffusion of the dye molecules from the external surface into the pores of the adsorbent. The third portion is the final equilibrium stage, where the dye molecules were adsorbed on the active sites on the internal surface of the pores and the intra-particle diffusion starts to slow down due to the solute concentration getting lower and lower in solution [24]. The linear portions of curves did not pass through the origin, suggesting that pore diffusion is not the step controlling the overall rate of mass transfer at beginning of adsorption.

chromatography-separation-techniques-Intra-particle

Figure 14: Intra-particle diffusion plots.

Thermodynamic studies

The data obtained from the temperature studies were used for thermodynamic analysis. The Gibbs free energy is presented in Eq. (12). The Gibbs free energy can also be expressed using enthalpy and entropy at a constant temperature. The linearized form of Eqs. (11) and (12) results in Eq. (13), which is the Van’t Hoff equation [25]:

ΔGo = RTlnKc   (13)    

ΔGo = ΔHo − TΔSo     (14)  

lnKc = ΔS / R − ΔRK1 / RT                   (15)

where ΔG° (kJ/mol) is the change in Gibbs Free Energy, ΔH° (kJ/mol) is the enthalpy change of MB adsorption, ΔS° (J/mol K) is the entropy change of MB adsorption, R is the universal gas constant (8.314J/mol K), T is the absolute temperature (K), and Kl is the Langmuir constant.

By plotting a graph of ln Kc versus 1/T (Figure 15), the values of ΔH° and ΔS° can be estimated from the slope and intercept of Van’t Hoff plots, respectively. The thermodynamic parameters are given in Table 7 and 8. The negative values of ΔG° for Ac blank, Ac 4% and Ac 8% nanocomposites at various temperatures indicate the spontaneous nature of the sorption process. The fact that the values of the ΔG° decrease with increasing temperature indicates the increase of spontaneous effect. For all sorbents, the positive value of ΔH° suggested the endothermic nature of the adsorption process (Tables 9 and 10). Moreover, the positive values of ΔS° point out the increased randomness at the solid/liquid interface during the sorption of MB on Ac blank, Ac 4% and Ac 8% nanocomposites.

chromatography-separation-techniques-lnKa-versus

Figure 15: The plots of lnKa versus T−1 for estimations of thermodynamic parameters.

Sample code Cₒ K(mg g-1 min-0.5 ) R2
Ac blank 1000
600
300
375.68769
261.21426
127.78369
0.96163
0.97107
0.92865
Ac 4% 1000
600
300
421.3177
303.93001
106.87931
0.99132
0.99621
0.92895
Ac 8% 1000
600
300
384.32703
185.944
162.28759
0.96557
0.9686
0.99397

Table 7: Intraparticle diffusion parameters for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different concentrations.

Sample Code ΔH° (KJ/mol) ΔS° (KJ/mol K) ΔG° (KJ/mol)
298 K 308 K 318 K
Ac blank 52.2015 0.2742 - 29.5576 -32.2234 - 35.0466
Ac 4% 45.4809 0.243 -27.3126 -29.2817 -30.1830
Ac 8% 27.0154 0.1891 -29.5196 -30.8827 -33.3259

Table 8: Thermodynamic parameters for adsorption of MB by Ac blank, Ac 4%, Ac 8% at different temperatures.

Cₒ qe for first cycle (mg/g) qe for second cycle(mg/g)
Treatment with HNO3 Treatment with NaOH
300 748.20 481.23 748.15
400 997.43 486.95 994.43
500 1246.81 495.61 1234.34

Table 9: The adsorption capacity of MB using Ac 8% nanocomposite before and after treatment with NaOH at three different concentrations.

Cₒ Regeneration efficiencies in each cycle (%)
2nd 3rd 4th
300 100 99.9 99.9
400 99.7 99.3 98.1
500 99 98.4 97.1

Table 10: Regeneration efficiencies of Ac 8% nanocomposite after repeated adsorption desorption cycles.

Conclusion

The results of the present study reveals that Fe3O4/poly (Acrylamideco- sodium acrylate) core shell magnetic nanogel (Fe3O4/p(Am-co-Na Ac)) based on crosslinked poly (sodium acrylate-co-acrylamide) may be an extremely viable adsorbent for application in the treatment of water and industrial wastewater contaminated with dyes. The amount of MB adsorbed was found to increase with increasing the wt% of nanogel from 4% to 8%. Moreover, the adsorption of MB was dependent on initial concentration, reaction temperature and pH. The MB adsorption capacity increased with the increase of pH in the range of 3–8, where ionization of carboxylic group occurs. The adsorption of MB onto the nanocomposites reached equilibrium within about 60 min. The adsorbents exhibited high sorption capacities toward basic dyes, viz.: 2167.802-2765.893 mg g-1. The adsorption equilibrium could be well described by Langmuir adsorption isotherms, namely monolayer adsorption on a homogenous surface. The adsorption kinetics followed a pseudo-second order kinetic model and intra-particle diffusion was involved in the adsorption process. Thermodynamic results indicated that the adsorption process was spontaneous and endothermic in nature.

Acknowledgement

The author would like to thank Dr MAGabr providing drug samples.

References

  1. Sanghi R, Bhatttacharya B, Singh V (2002) Cassia angustifoliaseed gum as an effective natural coagulant for decolourisation of dye solutions Green Chem 4: 252-254.
  2. Ghosh D, Bhattacharyya KG (2002) Adsorption of methylene blue on kaolinite. App Clay Sci 20: 295-300.
  3. Kornaros M, Lyberatos G (2006) Biological treatment of wastewaters from a dye manufacturing company using a trickling filter. J Hazar Mater 136: 95-102.
  4. Guibal E, Roussy J (2007) Coagulation and flocculation of dye-containing solutions using a biopolymer (Chitosan). React and Func Pol 67: 33-42.
  5. Zhao W, Wu Z,Wang D (2006) Ozone direct oxidation kinetics of Cationic Red X-GRL in aqueous solution. J Hazar Mater 137: 1859-1865.
  6. Dutta K, Mukhopadhyay S, Bhattacharjee S, Chaudhuri B (2001) Chemical oxidation of methylene blue using a Fenton-like reaction J Hazar Mater 84: 57-71.
  7. Capar G, Yetis U, Yilmaz L (2006) Membrane based strategies for the pre-treatment of acid dye bath wastewaters, J Hazar Mater135: 423-430.
  8. Liu CH, Wu JS, Chiu HC, Suen SY, Chu KH (2007) Removal of anionic reactive dyes from water using anion exchange membranes as adsorbers. Water res 41: 1491-1500.
  9. Muruganandham M, Swaminathan M (2006) TiO2-UV photocatalytic oxidation of Reactive Yellow 14: Effect of operational parameters. J Hazard Mater 135: 78-86.
  10. De Lisi R, Lazzara G, Milioto S, Muratore N (2007) Adsorption of a dye on clay and sand. Use of cyclodextrins as solubility-enhancement agents.Chemosphere 69: 1703-1712.
  11. Dabrowski A (2001) Adsorption--from theory to practice. AdvColl Inter Sci 93: 135-224.
  12. Ahmad A, Rafatullah M, Danish M (2007) Sorption studies of Zn (II)-and Cd (II) ions from aqueous solution on treated sawdust of wood. Eur J Wood and Wood Prod 65: 429-436.
  13. Xie G, Zhang Q, Luo Z, Wu M, Li T (2003) Preparation and characterization of monodisperse magnetic poly (styrene butyl acrylate methacrylic acid) microspheres in the presence of a polar solvent, J appl pol sci 87: 1733-1738.
  14. Lee Y, Rho J, Jung B (2003) Preparation of magnetic ion-exchange resins by the suspention polymerization of styrenewith magnetite. J appl pol sci89: 2058-2067.
  15. Hameed B (2009) Removal of cationic dye from aqueous solution using jackfruit peel as non-conventional low-cost adsorbent. J Hazar Mater 162: 344-350.
  16. Ho YS, Chiang TH, Hsueh YM (2005) Removal of basic dye from aqueous solution using tree fern as a biosorbent. Prod Biochem 40: 119-124.
  17. Garg V, Amita M, Kumar R, Gupta R (2004) Basic dye (methylene blue) removal from simulated wastewater by adsorption using Indian Rosewood sawdust: a timber industry waste. Dyes and pig 63: 243-250.
  18. Özer A, Akkaya G, Turabik M (2006) Biosorption of Acid Blue 290 (AB 290) and Acid Blue 324 (AB 324) dyes on Spirogyra rhizopus. J Hazard Mater 135: 355-364.
  19. Arami M, Limaee NY, Mahmoodi NM, Tabrizi NS (2006) Equilibrium and kinetics studies for the adsorption of direct and acid dyes from aqueous solution by soy meal hull. J Hazard Mater 135: 171-179.
  20. Webi TW, Chakravort RK (1974) Pore and solid modes for fixed adsorbers. AIChEJ 20: 228-238.
  21. Mckay G (1982) Adsorption of dyestuffs from aqueous solutions with activated carbon I: Equilibrium and batch contact-time studies. J Chem Techn and Biotechn 32: 759-772.
  22. Ho Y, McKay G (1998) A Comparison of Chemisorption Kinetic Models Applied to Pollutant Removal on Various Sorbents. Process Safety and Environmental Protection 76: 332-340.
  23. Kiran I, Akar T, Ozcan A, Tunali S (2006) Biosorption kinetics and isotherm studies of Acid Red 57 by dried Cephalosporiumaphidicola cells from aqueous solutions. Biochem Engin J 31:197-203.
  24. Sun Q, Yang L (2003) The adsorption of basic dyes from aqueous solution on modified peat-resin particle. Water res 37: 1535-1544.
  25. Hameed B, El-Khaiary M (2008) Kinetics and equilibrium studies of malachite green adsorption on rice straw-derived. Char J Hazar Mater 153: 701-708.
Citation: Akl MA, Atta A, Youssef AEM, Ibraheim MA (2013) The Utility of Novel Superabsorbent Core Shell Magnetic Nanocomposites for Efficient Removalof Basic Dyes from Aqueous Solutions. J Chromat Separation Techniq 4:185.

Copyright: © 2013 Akl MA, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top