Synthesis and characterization of Fe3O4@TSC nanocomposite: highly efficient removal of toxic metal ions from aqueous medium

Ayoub Abdullah Alqadami, Mu Naushad*, Mohammad Abulhassan Abdalla, Tansir Ahamad, Zeid Abdullah Alothman and Saad M. Alshehri
Department of Chemistry, College of Science, Bld #5, King Saud University, Riyadh-11451, Kingdom of Saudi Arabia. E-mail: shad81@rediffmail.com; Tel: +966 14674198

Received 23rd December 2015 , Accepted 17th February 2016

First published on 17th February 2016


Abstract

In the present study, trisodium citrate (TSC) modified magnetite (Fe3O4) nanocomposite was synthesized and characterized using various analytical techniques. The transmission electron microscope images show that the Fe3O4@TSC nanoparticles are well dispersed due to the presence of the TSC coating on Fe3O4, and the particles sizes are in the range of 5–10 nm. The SBET and the total pore volume of Fe3O4@TSC are 245.42 m2 g−1 and 0.368 cm3 g−1, respectively. The saturation magnetization values of Fe3O4 and Fe3O4@TSC are 78.4 and 55.4 emu g−1, respectively. The Fe3O4@TSC is a magnetic adsorbent and was used for the removal of Cr3+ and Co2+ metal ions from aqueous medium. The adsorption of both metal ions onto Fe3O4@TSC is rapid and efficient. The adsorption process is performed at diverse temperatures and the outcomes are investigated kinetically. The results show that adsorption was exothermic and followed the pseudo-second-order kinetic model. Isotherm modelling reveals that the Langmuir equation described the adsorption of both metal ions. Moreover, the loaded Fe3O4@TSC could be recovered easily from aqueous solution by magnetic separation and regenerated by simply washing with 0.1 M HCl solution. Consequently, Fe3O4@TSC nanocomposite could be utilized as an efficient and recyclable adsorbent for the removal of toxic metal ions from aqueous solution.


1. Introduction

Magnetic nanoparticles (MNPs) have been widely studied due to their many potential applications in the chemical, biological and medical sciences such as pre-concentration and separation of ions,1–3 bioseparation4–6 protein separation,7 DNA recovery,8 cancer treatment,9 biomedicine and targeted drug delivery.10,11 MNPs are effective adsorbents due to their magnetic properties and large specific surface area that enables effective separation in a short time using an external magnetic field.12–14 Nano-sized iron oxide nanoparticles have gained significant consideration owing to their simple, low-cost synthesis in comparison to commercially accessible activated carbon.13 Nevertheless, nano-sized iron oxide nanoparticles suffer from some shortcomings such as agglomeration, and low chemical stability making industrial applications inconvenient.15 Hence, to attain stability, MNPs are usually functionalized with specific groups. Citric acid and trisodium citrate (TSC) have proven to be suitable and easily accessible materials to stabilize MNPs.16–20 Due to several advantages over bulk magnetite –greater adsorption capacity, less equilibrium time and easy magnetic separation of solids after adsorption – MNPs have also been utilized as adsorbents for the removal of heavy metal ions and other pollutants from aqueous systems.21–23 Heavy metal ions are pollutants with toxic to living organisms if present at a certain concentration.24–26 The existence of toxic metals in the environment can be harmful to human beings and other living species even at low concentrations.27,28 The most toxic heavy metals are chromium, mercury, lead, nickel, copper, cobalt and cadmium which can be distinguished from other pollutants, because they are non-degradable and accumulate in living organisms. The main sources of heavy metal pollution in water are agricultural and industrial wastewaters; agricultural sources arise from fertilizer and fungicidal spray and industrial sources include paint, plating, tanneries, chemical, mining and petroleum refining.29 So far, a variety of techniques, such as ion-exchange, chemical coagulation, membrane separation, extraction, chemical precipitation, electro deposition and adsorption have been used for the removal of heavy metals from water.30–34 Most of these techniques have limitations such as high cost, complicated treatment processes and high energy requirements, or they risk introducing secondary pollution. Adsorption is a very efficient, simple, and cost-effective method for the removal of contaminants from aqueous medium.35–39

Herein, we have synthesized MNPs modified with TSC solution. The modified magnetite nanocomposite (Fe3O4@TSC) was characterized using various analytical techniques, and was utilized for the removal of Cr3+ and Co2+ metal ions from aqueous medium. Kinetic and thermodynamic studies were conducted to obtain optimum metal ion removal capacity and certain constants relating to the adsorption phenomena.

2. Experimental

2.1. Materials and methods

All chemicals and reagents used for this study were of analytical reagent grade. Metal nitrates, chlorides and TSC were purchased from Sigma-Aldrich, USA. All solutions were prepared in Milli-Q water.

Fe3O4 nanoparticles were synthesized in a three-necked round bottom flask equipped with mechanical stirrer, nitrogen gas inlet and dropping funnel by using a co-precipitation process. In the first step, iron(II)chloride and iron(III)chloride were mixed in 1[thin space (1/6-em)]:[thin space (1/6-em)]2 molar ratio. This ratio was achieved by dissolving 10.8 g FeCl3·6H2O in 100 mL 0.5 M HCl and 3.98 g FeCl2·4H2O in 100 mL 0.5 M HCl under N2 gas for 30 min. Nitrogen gas was used to remove O2 and to prevent the unwanted critical oxidation of Fe(II). Then, these two solutions were mixed with vigorous stirring and heated at 75 °C under N2 for 1 h. An aqueous solution of NH4OH (28%, 20 mL) was added drop-wise into solution at the same temperature (75 °C) for 5 h. Black colored precipitates were obtained and were isolated from the supernatant liquid by decantation. The precipitates were washed with demineralized water and separated using centrifugation at 5000 rpm for 10 min. Then, 0.05 M HCl solution was added to neutralize the anionic charge on the particle surface. The formed magnetite nanoparticles (Fe3O4) were re-dispersed in 150 mL of 0.5 M solution of TSC (to prevent Fe3O4 NPs from possible oxidation in air as well as from agglomeration) and heated at 80 °C under ultra-sonication vibration for 60 min. The as-formed reaction product was kept in excess of citric acid and so, the nanoparticle dispersion was centrifuged. The Fe3O4@TSC nanocomposites were washed with deionized water and acetone to remove the excess citrate and then dried under vacuum oven for 12 h at 50 °C.

Throughout the study, the pH measurements were performed using a single electrode pH meter (Orion 2 star, Thermo Scientific, USA). The concentrations of chromium and cobalt were determined using atomic absorption spectrophotometry (AAS, model AAnalyst 400, Perkin Elmer). The FTIR spectra were obtained between 500 to 4000 cm−1 by KBr disc method using an FTIR spectrophotometer (Nicolet 6700, Thermo Scientific, USA). The surface area and other related parameters of the magnetite and Fe3O4@TSC were evaluated using an automated surface area and porosity analyzer, (Quadrasorb evo, Quantachrome, USA). The scanning electron micrographs were obtained using a high-performance scanning electron microscope with a high resolution of 3.0 nm (JSM-6380 LA, Tokyo, Japan). TEM (JEM-2100F, JEOL, Japan) and X-ray diffraction (XRD, Shimadzu model 6000) characterizations were also performed on Fe3O4 and Fe3O4@TSC nanocomposites.

2.2. Batch adsorption experiment

The adsorption of Cr3+ and Co2+ metal ions using Fe3O4@TSC was determined by performing adsorption tests in 100 mL Erlenmeyer flasks. 150 mL of 10 mg L−1 metal ion solutions were shaken with 30 mg Fe3O4@TSC nanocomposite at 100 rpm for 24 h to reach equilibrium. After the adsorption process, the samples were separated using a magnetic field, filtered and analyzed by AAS. All experiments were done in triplicate and the average values were taken. The pH of the solution was adjusted using 0.1 M HNO3/NaOH solutions. The amount of the metal ions adsorbed at equilibrium (qe, mg g−1) was evaluated by the following equation:40
 
image file: c5ra27525c-t1.tif(1)
where, Co and Ce are the initial and equilibrium concentrations of metal ions in the solution phase, respectively, V is the initial volume of the metal ion solution (L), and m is the mass of the adsorbent (g).

Batch studies were also performed to obtain kinetic, thermodynamic and adsorption isotherm parameters for the adsorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC.

2.3. Desorption studies

The desorption efficiency was carried out by batch process using HCl and HNO3 solutions of different concentrations as the eluents. 100 mL of 10 mg L−1 metal ion solutions were treated with 100 mg of Fe3O4@TSC nanocomposites in an Erlenmeyer flask at 120 rpm for 180 min. After 180 min, Fe3O4@TSC nanocomposites were filtered off and washed several times with Milli-Q water to remove the excess metal ions. Then, Fe3O4@TSC nanocomposites were treated with 100 mL of the aforementioned eluents in an Erlenmeyer flask under the above mentioned conditions. After 180 min, the solution was filtered and the residual concentration of metal ions in the solution phase was determined by AAS as described above.

3. Results and discussion

3.1. Characterization

The procedure for the preparation of the Fe3O4@TSC nanocomposite is illustrated in Scheme 1. The FTIR spectra of Fe3O4, Fe3O4@TSC and Fe3O4@TSC after Cr3+ and Co2+ adsorption are illustrated in Fig. 1. In the FTIR spectrum of Fe3O4 nanoparticles, the broad band at 595–610 cm−1 corresponds to Fe–O bonds. In the Fe3O4@TSC spectrum, the bands at 1410 and 1620 cm−1 come from symmetric and asymmetric vibrations of carboxylate groups on ligands on the surface of Fe3O4 due to the presence of TSC. The other two bands at 2925 and 2854 cm−1 were due to the presence of C–H stretching vibrations.41 Other bands at 3325 and 3380 cm−1 support the presence of stretching vibrations of O–H and the broadening in this region shows hydrogen bonding. Comparing the original spectrum of Fe3O4@TSC before and after adsorption of Cr3+ and Co2+, the FTIR spectrum of Fe3O4@TSC@Cr/Co shows characteristic absorption bands at 540/567 cm−1 corresponding to Cr–O and Co–O.
image file: c5ra27525c-s1.tif
Scheme 1 Preparation of Fe3O4@TSC nanocomposite.

image file: c5ra27525c-f1.tif
Fig. 1 FTIR spectra of (A) Fe3O4, (B) Fe3O4@TSC, (C) Cr3+ adsorbed Fe3O4@TSC, (D) Co2+ adsorbed Fe3O4@TSC.

The thermal stability of Fe3O4 and Fe3O4@TSC powder was characterized by thermogravimetric analysis (TGA) as shown in Fig. 2. The weight loss during the thermal treatment was observed for Fe3O4 and Fe3O4@TSC up to 750 °C.


image file: c5ra27525c-f2.tif
Fig. 2 TGA and XRD patterns of Fe3O4 and Fe3O4@TSC.

In the case of Fe3O4 nanoparticles, the thermal degradation was divided into two stages, the first stage is up to 200 °C, where 4.3% weight loss was observed due to the loss of adsorbed water of humidity. The second phase was the key degradation stage in the case of Fe3O4, where 4.8% weight loss was observed up to 400 °C, possibly due to the removal of crystalline water and conversion of some hydroxide to oxide.

On the other hand, the thermal degradation of Fe3O4@TSC nanocomposites was divided into three stages. The first weight loss (up to 200 °C) is associated with the evaporation of physio absorbed water and the solvent. The second stage (between 200 and 300 °C) links to a similar removal of TSC, and further 18% weight loss occurs on increasing the temperature from 300 to 750 °C. During this stage, TSC was slowly removed, resulting in an uninterrupted weight loss owing to the elimination of various volatile compounds including H2O, CO2, CO, C2H6, H2 and CH4.

Fig. 2 shows the XRD studies of Fe3O4 and Fe3O4@TSC. The X-ray diffraction lines at 2θ; 18.28, 30.06, 35.48, 43.70, 53.81, 57.51, 62.80, 70.80 and 74.92 correspond to (111), (220), (311), (400), (422), (511), (440), (620) and (533) planes of the face-centered cubic (fcc) lattice of Fe3O4, respectively (JCPDS card no. 19-0629). TSC stabilized Fe3O4 shows low intense XRD peaks due to the low content of Fe3O4. Moreover, the peaks became broader due to the amorphous nature of TSC and more broadening occurred at a 2θ value of 20–30. The FTIR and XRD results support the formation of magnetic Fe3O4 nanoparticles and Fe3O4 nanoparticles being stabilized by TSC.

The magnetic properties of Fe3O4 nanoparticles and Fe3O4@TSC were subsequently studied by using a vibrating sample magnetometer at room temperature. As shown in Fig. 3, no obvious magnetic hysteresis loops were observed for Fe3O4 indicating that all nanoparticles exhibited superparamagnetic behavior. The magnetic Fe3O4 nanoparticles had a saturation magnetization (M) of 78.4 emu g−1. The saturation magnetization value of the Fe3O4@TSC nanocomposite was 55.4 emu g−1. The saturation magnetization value of Fe3O4@TSC nanocomposite was lower in comparison to the original magnetic Fe3O4 nanoparticles which might be due to the decrease of Fe3O4 content in the nanocomposite, but the magnetic attraction of Fe3O4@TSC was strong enough to be proficiently separated from a solution using an external magnetic field.


image file: c5ra27525c-f3.tif
Fig. 3 Magnetization behavior of Fe3O4 and Fe3O4@TSC.

The morphology of Fe3O4 and the Fe3O4@TSC nanocomposite was characterized using scanning electron microscopy (SEM). The results show that the prepared nanoparticles were spherical, within the diameter range of 5–8 nm. Monodisperse Fe3O4@TSC magnetic microsphere composites showed a similar spherical form with an average diameter of 5–10 nm (Fig. 4A and B).


image file: c5ra27525c-f4.tif
Fig. 4 (A) SEM of Fe3O4, (B) SEM of Fe3O4@TSC, (C and D) TEM of Fe3O4, (E and F) TEM of Fe3O4@TSC.

TEM and high resolution transmission electron microscopy (HRTEM) of the Fe3O4, Fe3O4@TSC are illustrated in Fig. 4C–F and 5, respectively. The TEM images of Fe3O4 nanoparticles were agglomerated due to the involvement of the hydroxyl groups and magnetic interaction. The average particle sizes were 5–8 nm for the Fe3O4 (Fig. 4C and D). On the other hand, the Fe3O4@TSC nanoparticles were well dispersed due to the presence of TSC coating on Fe3O4 and the particles size were 5–10 nm (Fig. 4E and F). The lattice fringes with an interfering distance of 0.25 ± 0.05 nm corresponding to the (311) plane of face centered (fcc) of Fe3O4, similarly, an interfering distance of 0.25 ± 0.05 nm corresponding to the (311) plane of Fe3O4@TSC as shown in Fig. 5A and B, respectively.


image file: c5ra27525c-f5.tif
Fig. 5 HRTEM of (A) Fe3O4, (B) Fe3O4@TSC.

The composition of Fe3O4, Fe3O4@TSC, Fe3O4@TSC@Cr and Fe3O4@TSC@Co were measured by EDX. The elements detected were Fe and O in the case of Fe3O4, and Fe, O and C in the case of Fe3O4@TSC. The EDX and elemental mapping images illustrated in Fig. 6, supported the presence of Cr3+ and Co2+ metal ions after the adsorption onto Fe3O4@TSC.


image file: c5ra27525c-f6.tif
Fig. 6 EDX spectra and elemental mapping of (A) Fe3O4, (B) Fe3O4@TSC, (C) Cr3+ adsorbed Fe3O4@TSC, (D) Co2+ adsorbed Fe3O4@TSC.

Nitrogen adsorption–desorption measurements were carried out to measure the porosity and surface nature of the Fe3O4 and Fe3O4@TSC as shown in Fig. 7. The adsorption isotherms were type IV in the case of both Fe3O4 and Fe3O4@TSC which supports the nanoporous nature. The corresponding Barrett–Joyner–Halenda (BJH) pore size distribution was calculated using the adsorption–desorption isotherm which revealed a uniform pore size of 5–10 nm for nanocomposites prepared after coating with TSC. The Brunauer–Emmett–Teller (BET) surface area and the total pore volume for Fe3O4@TSC was higher than that of the parent Fe3O4, found to be 245.42 m2 g−1 and 0.368 cm3 g−1, respectively.


image file: c5ra27525c-f7.tif
Fig. 7 Adsorption–desorption isotherms.

3.2. Effects of operating parameters

The time dependent adsorption behavior was measured by performing a series of adsorption experiments at different contact times (1–240 min). It was found that the removal of Cr3+ and Co2+ metal ions was very fast for both metal ions. However, the adsorption of Cr3+ was faster than Co2+. The equilibrium was attained within 120 min for Cr3+ metal ions where the adsorption was 100% (qe, 50 mg g−1). While, for the Co2+ metal ion the equilibrium was achieved in 240 min where the adsorption was 95% (qe, 47.4 mg g−1). Thus, 240 min was taken as the optimal contact time for the removal of both metal ions. The high removal efficiency of Fe3O4@TSC nanocomposite for these metal ions may be attributed to the presence of various acetate groups on the Fe3O4@TSC nanocomposite (Fig. 8A). The oxygen atoms of the acetate groups are electron rich (electronegative) and the metal ions (Cr3+ and Co2+) are electropositive. So there is an electrostatic attraction between the oxygen atoms and the metal ions. In addition, there are several other oxygen atoms on the carbonyl group and internal ester groups on the core of Fe3O4@TSC which were also responsible for the bonding of Fe3O4@TSC to the metal ions through electrostatic attraction. The fast adsorption at the initial stage of adsorption was attributed to the large number of active sites available. Thus, the progressive decrease of adsorption sites resulted in a slower adsorption reaction.
image file: c5ra27525c-f8.tif
Fig. 8 Removal of Cr3+ and Co2+ metal ions using Fe3O4@TSC nanocomposite at different (A) time, (B) pH, (C and D) Fe3O4@TSC dose.

The pH of the solution plays an important role in the adsorption of any adsorbate. Fig. 8B shows the effect of pH on the adsorption of Cr3+ and Co2+ metal ions from aqueous solution onto Fe3O4@TSC nanocomposite by keeping all other parameters constant (contact time 240 min, Fe3O4@TSC dose 1 g L−1, temperature 25 ± 2 °C). It was found that the adsorption capacity for Cr3+ metal ions was only 0.20 mg g−1 at pH 2.5 and reached up to 24.8 mg g−1 increasing the pH to 6.0. After that, it started to decrease. The same trend was observed for Co2+ metal ions. The adsorption capacity was 1.5 mg g−1 at pH 2.5 and increased up to 22.4 mg g−1 at pH 6.0, and then decreased with further pH increase. At pH > 6.0, the adsorption of both metal ions decreased due to the formation of metal hydroxides.42 As expected, the adsorptive removal of Cr3+ and Co2+ metal ions was low in the acidic medium; the concentration of H+ is high, so protonation of the active sites of Fe3O4@TSC dominates the adsorption process.

The effect of Fe3O4@TSC dose on the adsorption capacity of Cr3+ and Co2+ metal ions is shown in Fig. 8C and D. It was observed that adsorption of Cr3+ and Co2+ metal ions increases from 89 to 96% and 91 to 97%, respectively, as the dose of Fe3O4@TSC increased from 100 to 500 mg. While the adsorption capacity for Cr3+ and Co2+ sharply decreased from 22.4 mg g−1 to 4.8 mg g−1 and 22.7 mg g−1 to 4.8 mg g−1, respectively. This decrease in the adsorption capacity of both metal ions with the increase in the adsorbent dose might be due to the fact that some adsorption sites remained unsaturated during the adsorption process when the number of available adsorption sites increased.43,44 Moreover, an increase in Fe3O4@TSC dose resulted in aggregation of adsorbent particles, leading to a decrease in total surface area of the Fe3O4@TSC and increased diffusional path length.45 Thus, a dose of 100 mg/25 mL was chosen as the optimum dose of Fe3O4@TSC nanocomposite for the further studies as it brought down the Cr3+ and Co2+ metal ion concentration to below the WHO permissible limit.

The effects of initial metal ion concentrations at different temperatures (298–328 K) were also studied. The adsorption capacity of the Fe3O4@TSC nanocomposite for both metals increased with increasing metal ion concentration (Fig. 9). This could be ascribed to more targets of metal ions affording a higher driving force to enable the ion diffusion from the solution phase to Fe3O4@TSC, and more collisions between metal ions and active sites of Fe3O4@TSC.46 However, in relation to the temperature, it was found that the adsorption capacity for Cr3+ and Co2+ metal ions decreased with increasing temperature, showing the exothermic nature of adsorption. The highest adsorption was achieved at 298 K for both metal ions.


image file: c5ra27525c-f9.tif
Fig. 9 The effect of initial ion concentration on the adsorption of Cr3+ and Co2+ metal ions onto the Fe3O4@TSC nanocomposite at different temperatures.

3.3. Adsorption kinetics

The kinetics of Cr3+ and Co2+ metal ions adsorption onto the Fe3O4@TSC nanocomposite was examined using pseudo first-order and pseudo second-order kinetic models. The conformity between experimental data and the model predicted values was expressed by the correlation coefficients (R2). A relatively high R2 value showed that the model effectively described the kinetics of Cr3+ and Co2+ metal ions adsorption.
3.3.1. The pseudo-first-order equation. It was given by Lagergren:47
 
image file: c5ra27525c-t2.tif(2)
where, qt and qe are the amounts of metal ions adsorbed at time t, and at equilibrium, respectively, and k1 is the rate constant for pseudo-first-order adsorption (min−1). The values of rate constant were determined from the slope of the plot log(qeqt) versus t.
3.3.2. The pseudo-second-order equation. The pseudo-second-order kinetic rate equation is given by Ho and Mckay:48
 
image file: c5ra27525c-t3.tif(3)
where k2 is the pseudo-second-order rate constant. The values of k2 for all studied concentrations were determined from the intercepts of the plot t/qt versus time.

The kinetic parameters are given in Table 1. The fitted linear regression plot shows data that fit the pseudo-second-order model well with a better value of correlation coefficient, compared to the pseudo-first-order (Fig. S1). The low values of correlation coefficient (R2) for pseudo-first-order ascertained the unsuitability of this model for the adsorption of Cr3+ and Co2+ metal ions adsorption onto the Fe3O4@TSC nanocomposite. The experimental equilibrium adsorption capacity values (qe,exp) for Cr3+ and Co2+ metal ions at various concentrations are in good agreement with the calculated equilibrium adsorption capacity values (qe,cal) for the pseudo-second-order kinetics model.

Table 1 Isotherm constants for the adsorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC
Metal ions Temperature (K) qm,exp (mg g−1) Langmuir isotherm Freundlich isotherm
qm,cal (mg g−1) b (L mg−1) RL R2 Kf (mg g−1)(L mg−1)1/n n R2
Cr(III) 298 549.125 1667 0.004 0.961 0.994 32.94 1.117 0.976
308 525.325 1000 0.0091 0.911 0.994 15.16 0.941 0.967
318 507.6 625 0.024 0.805 0.993 9.28 0.883 0.965
328 487.2 495 0.118 0.456 0.997 6.88 0.871 0.973
Co(II) 298 452.5 588 0.007 0.939 0.997 22.42 1.33 0.984
308 399.2 555 0.011 0.904 0.990 15.46 1.334 0.992
318 382.5 380 0.036 0.735 0.993 6.01 1.08 0.998
328 352.5 370 0.059 0.631 0.993 3.52 1.012 0.993


3.4. Adsorption isotherms

Isotherm results were examined using Langmuir and Freundlich isotherms.49,50

The linear form of the Langmuir isotherm model can be written as:

 
image file: c5ra27525c-t4.tif(4)
where, qe is the amount of adsorbed phenol (mg g−1), Ce is the equilibrium concentration of phenol (mg L−1), Qm and b are the Langmuir constants related to maximum monolayer adsorption capacity and energy of adsorption, respectively.

The values of Qm and b were evaluated from the intercept and slope of the linear plots of 1/qe vs. 1/Ce, respectively (Table 2). To see the efficiency of the adsorption process, the value of dimension less equilibrium parameter (RL) was determined using the following equation:51

 
image file: c5ra27525c-t5.tif(5)
where, Co (mg L−1) is the lowest initial metal ion concentration and b is the Langmuir constant (L mg−1). The values of RL in the present study were <1 which shows a favorable adsorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC (Table 2).

Table 2 Kinetic parameters for the adsorption of Cr3+ and Co2+ metals onto Fe3O4@TSC
Metal ions (Co, mg L−1) qe,exp (mg g−1) Pseudo-first-order Pseudo-second-order
qe,cal (mg g−1) K1 (min−1) R2 qe,cal (mg g−1) K2 (g mg−1 min−1) R2
Cr(III) 10 50 4.50 0.012 0.363 38.91 0.032 0.975
Co(II) 10 46.58 13.31 0.024 0.926 46.73 0.0079 0.999


The linear form of the Freundlich isotherm is given by the following equation:

 
image file: c5ra27525c-t6.tif(6)

The Freundlich isotherm constants (Kf and n) were determined from the intercept and slope of the linear plots of log[thin space (1/6-em)]qe vs. log[thin space (1/6-em)]Ce, respectively (Table 2). Kf decreased with increasing temperature which shows the exothermic nature of the adsorption process. The value of n was found to be greater than one which suggested the favorable adsorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC nanocomposites.52 As seen in Table 2, the Langmuir isotherm fitted well with the experimental data (correlation coefficient R2 > 0.99), whereas, the low correlation coefficients showed poor agreement with the Freundlich isotherm experimental data (Fig. S2).

3.5. Adsorption thermodynamics

The values of enthalpy change (ΔH°) and entropy change (ΔS°) were calculated from the slopes and intercepts of the plots of ln[thin space (1/6-em)]Kc versus 1/T using the following equation:
 
image file: c5ra27525c-t7.tif(7)

The values of ΔG° (free energy change) were calculated from the following relation:

 
ΔG° = ΔH° − TΔS° (8)

The results for the thermodynamic parameters are presented in Table 3. The values of Δ were negative and decreased with the increase in temperature showing that the spontaneous adsorption process and spontaneity decreased with increasing temperature. The negative values of Δ suggest a decrease in randomness at the solid/solution interface. The adsorption of both metal ions onto Fe3O4@TSC nanocomposite was physical and exothermic in nature as the values of Δ were negative. The decrease in adsorption capacity with the increase in temperature may be ascribed to the weakening of adsorptive forces between the active sites present on the Fe3O4@TSC nanocomposite surface and the metal ions, and also between the adjacent molecules of the adsorbed phase.

Table 3 Thermodynamic parameters for the adsorption of Cr3+ and Co2+ metals onto Fe3O4@TSC
Metal ions Co (mg L−1) −ΔH° (kJ mol−1) −ΔS° (J mol−1 K−1) −ΔG° (kJ mol−1)
298 K 308 K 318 K 328 K
Cr(III) 10 41.84 119.67 6.385 4.691 3.715 2.753
100 21.76 59.81 3.858 3.451 2.801 2.068
250 19.12 47.93 4.903 4.256 3.870 3.443
Co(II) 10 47.56 141.63 5.256 4.208 2.252 1.185
100 41.19 124.29 4.495 2.391 1.625 0.658
250 18.26 53.73 2.389 1.459 1.204 0.702


3.6. Desorption study

To recover the adsorbed Cr3+ and Co2+ metal ions, desorption studies were performed. It was noticed that better recovery of both metal ions was observed in HCl as compared to HNO3 (Fig. 10). It might be due to the smaller size of Cl ions in comparison to the NO3 ion. It is also interesting to note that as the concentration of HCl increased, the recovery was decreased from 88.3% to 80.2% for Cr3+, and 90% to 70.3% for Co2+, which might be due to the deterioration of Fe3O4@TSC surface functional groups at higher concentrations of HCl solution.
image file: c5ra27525c-f10.tif
Fig. 10 Adsorption/desorption studies.

3.7. Proposed adsorption/desorption mechanism

The suggested mechanism for the adsorption/desorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC is given in Scheme 2. It can be seen that Fe3O4@TSC had three types of active groups: –COONa, –C[double bond, length as m-dash]O and –OH which were capable of bonding with the metal ions. However, –COONa was more active towards metal ions in comparison to –C[double bond, length as m-dash]O and –OH groups. Actually, the oxygen atoms of acetate groups are electron rich (electronegative) and the metal ions (Cr3+ and Co2+) are electropositive. So, there is electrostatic attraction between the electron rich oxygen atoms and the electropositive metal ions. As can be seen in Section 3.6, the desorption of both metal ions was performed efficiently using 0.1 M HCl solution.
image file: c5ra27525c-s2.tif
Scheme 2 Proposed mechanism for the adsorption/desorption of Cr3+ and Co2+ metal ions onto Fe3O4@TSC nanocomposite.

4. Conclusion

In the present study, TSC modified magnetite nanocomposite was synthesized and used as a magnetic adsorbent for the removal of toxic Cr3+ and Co2+ metal ions from aqueous medium. The Fe3O4@TSC was characterized using various analytical techniques and it was found that Fe3O4@TSC particles were in the range of 5–10 nm. The saturation magnetization value of Fe3O4@TSC microspheres was lower in comparison to the original magnetic Fe3O4 nanoparticles which is due to the decrease of Fe3O4 content in the nanocomposites, the magnetic receptiveness of Fe3O4@TSC is still sufficient for separation from a solution using external magnetic fields. The surface area and the total pore volume for Fe3O4@TSC was higher than that of parental Fe3O4 and found to be 245.42 m2 g−1 and 0.368 cm3 g−1. The prepared magnetic adsorbent could be well dispersed in the water and easily separated magnetically from the medium after adsorption. Isotherm modelling revealed that the Langmuir equation could better describe the adsorption of both metal ions onto Fe3O4@TSC as compared to the Freundlich isotherm model. The adsorption process was kinetically studied by fitting the experimental data with two kinetic models, and the results designated that the adsorption followed the pseudo second order kinetic model well.

Acknowledgements

This project was supported by King Saud University, Deanship of Scientific Research, College of Science Research Center.

References

  1. B. R. White, B. T. Stackhouse and J. A. Holcombe, J. Hazard. Mater., 2009, 161, 848 CrossRef CAS PubMed.
  2. L. Zhou, Y. Wang, Z. Liu and Q. Huang, J. Hazard. Mater., 2009, 161, 995 CrossRef CAS PubMed.
  3. T. Tuutijarvi, J. Lu, M. Sillanp and G. Chen, J. Hazard. Mater., 2009, 166, 1415 CrossRef CAS PubMed.
  4. S. Bucak, D. A. Jones, P. E. Laibinis and T. A. Hatton, Biotechnol. Prog., 2003, 19, 477 CrossRef CAS PubMed.
  5. X. Song, X. Luo, Q. Zhang, A. Zhu, L. Ji and C. Yan, J. Magn. Magn. Mater., 2015, 388, 116 CrossRef CAS.
  6. S. V. Salihov, Y. A. Ivanenkov, S. P. Krechetov, M. S. Veselov, N. V. Sviridenkova, A. G. Savchenko, N. L. Klyachko, Y. I. Golovin, N. V. Chufarova, E. K. Beloglazkina and A. G. Majouga, J. Magn. Magn. Mater., 2015, 394, 173 CrossRef CAS.
  7. H. Gu, K. Xu, C. Xu and B. Xu, Chem. Commun., 2006, 9, 941 RSC.
  8. Y. Maeda, T. Toyoda, T. Mogi, T. Taguchi, T. Tanaami, T. Yoshino, T. Matsunaga and T. Tanaka, Colloids Surf., B, 2016, 139, 117 CrossRef CAS PubMed.
  9. I. Akira, T. Kouji, K. Kazuyoshi, S. Masashige, H. Hiroyuki, M. Kazuhiko, S. Toshiaki and K. Takeshi, Cancer Sci., 2003, 94, 308 CrossRef.
  10. D. K. Kim, M. Mikhaylova, F. H. Wang, J. Kehr, B. Bjelke, Y. Zhang, T. Tsakalakos and M. Muhammed, Chem. Mater., 2003, 15, 4343 CrossRef CAS.
  11. A. K. Gupta and S. Wells, IEEE Transactions on NanoBioscience, 2004, 3, 66 CrossRef PubMed.
  12. B. Zargar, H. Parham and A. Hatamie, Chemosphere, 2009, 76, 554 CrossRef CAS PubMed.
  13. A. Afkhami and R. Moosavi, J. Hazard. Mater., 2010, 174, 398 CrossRef CAS PubMed.
  14. C. H. Weng, Y. T. Lin, C. L. Yeh and Y. C. Sharma, Water Sci. Technol., 2010, 62, 844 CrossRef CAS PubMed.
  15. J. P. Jolivet, C. Chanéac and E. Tronc, Chem. Commun., 2004, 481 CAS.
  16. L. C. A. Oliveira, D. I. Petkowicz, A. Smaniotto and S. B. C. Pergher, Water Res., 2004, 38, 3699 CrossRef CAS PubMed.
  17. L. C. A. Oliveira, R. V. R. A. Rios, J. D. Fabris, V. Garg, K. Sapag and R. Lago, Carbon, 2002, 40, 2177 CrossRef CAS.
  18. G. Crini, Dyes Pigm., 2008, 77, 415 CrossRef CAS.
  19. G. Crini and P. M. Badot, Prog. Polym. Sci., 2008, 33, 399 CrossRef CAS.
  20. P. R. Chang, J. Yu, X. Ma and D. P. Anderson, Carbohydr. Polym., 2011, 83, 640 CrossRef CAS.
  21. M. H. Liao and D. H. Chen, J. Mater. Chem., 2002, 12, 3654 RSC.
  22. Y. F. Shen, J. Tang, Z. H. Nie, Y. D. Wang, Y. Ren and L. Zuo, Sep. Purif. Technol., 2009, 68, 312 CrossRef CAS.
  23. Y. F. Shen, J. Tang, Z. H. Nie, Y. D. Wang, Y. Ren and L. Zuo, Bioresour. Technol., 2009, 100, 4139 CrossRef CAS PubMed.
  24. M. Kumari, C. U. Pittman Jr and D. Mohan, J. Colloid Interface Sci., 2015, 442, 120 CrossRef CAS PubMed.
  25. S. A. Nabi and M. Naushad, Colloids Surf., A, 2007, 293, 175 CrossRef CAS.
  26. S. A. Nabi, M. Naushad and A. M. Khan, Colloids Surf., A, 2006, 280, 66 CrossRef CAS.
  27. A. A. ALqadami, M. A. Abdalla, Z. A. ALOthman and K. Omer, Int. J. Environ. Res. Public Health, 2013, 10, 361 CrossRef CAS PubMed.
  28. A. Shahat, M. R. Awual, M. A. Khaleque, M. Z. Alam, M. Naushad and A. M. Sarwaruddin Chowdhury, Chem. Eng. J, 2015, 273, 286 CrossRef CAS.
  29. S. Babel and T. A. Kurniawan, J. Hazard. Mater., 2003, 97, 219 CrossRef CAS PubMed.
  30. F. An, R. Du, X. Wang, M. Wan, X. Dai and J. Gao, J. Hazard. Mater., 2012, 201, 74 CrossRef PubMed.
  31. O. B. piridon, E. Preda, A. Botez and L. Pitulice, Environ. Sci. Pollut. Res., 2013, 20, 1 CrossRef PubMed.
  32. M. Caetano, C. Valderrama, A. Farran and J. L. Cortina, J. Colloid Interface Sci., 2009, 338, 402 CrossRef CAS PubMed.
  33. M. Ehtash, M. C. Fournier-Salaün, K. Dimitrov, P. Salaün and A. Saboni, Chem. Eng. J., 2014, 250, 42 CrossRef CAS.
  34. Y. Ng, N. Jayakumar and M. Hashim, J. Hazard. Mater., 2010, 184, 255 CrossRef CAS PubMed.
  35. Z. A. ALOthman, M. Naushad and R. Ali, Environ. Sci. Pollut. Res., 2013, 20, 3351 CrossRef CAS PubMed.
  36. M. Naushad, T. Ahamad, Z. A. Alothman, M. A. Shar, N. S. AlHokbany, S. M. Alshehri and J. Ind, J. Ind. Eng. Chem., 2015, 29, 78 CrossRef CAS.
  37. M. Rabiul Awual, M. M. Hasan, A. Shahat, M. Naushad, H. Shiwaku and T. Yaita, Chem. Eng. J., 2015, 265, 210 CrossRef.
  38. M. Naushad, Z. A. ALOthman and G. Sharma, Ionics, 2015, 21, 1453 CrossRef CAS.
  39. M. R. Awual, M. M. Hasan, M. Naushad, H. Shiwaku and T. Yaita, Chem. Eng. J., 2015, 209, 790 CAS.
  40. R. Bushra, M. Naushad, R. Adnan, Z. A. ALOthman and M. Rafatullah, J. Ind. Eng. Chem., 2015, 21, 1112 CrossRef CAS.
  41. Y. Meng, D. Chen, Y. Sun, D. Jiao, D. Zeng and Z. Liu, Appl. Surf. Sci., 2015, 324, 745 CrossRef CAS.
  42. M. R. Awual, G. E. Eldesoky, T. Yaita, M. Naushad, H. Shiwaku, Z. A. AlOthman and S. Suzuki, Chem. Eng. J., 2015, 279, 639 CrossRef CAS.
  43. S. E. Bailey, T. J. Olin, R. M. Bricka and D. D. Adrian, Water Res., 1999, 33, 2469 CrossRef CAS.
  44. M. M. Saeed, Nucl. Chem., 2003, 256, 73 CrossRef CAS.
  45. A. Ozer, G. Akkaya and M. Turabik, J. Hazard. Mater., 2005, 126, 119 CrossRef PubMed.
  46. J. Gong, X. Wang, X. Shao, S. Yuan, C. Yang and X. Hub, Talanta, 2012, 101, 45 CrossRef CAS PubMed.
  47. S. Lagergren, Handlingar, Band, 1898, 24, p. 1 Search PubMed.
  48. Y. S. Ho and G. McKay, Chem. Eng. J., 1998, 70, 115 CrossRef CAS.
  49. I. Langmuir, J. Am. Chem. Soc., 1918, 40, 1361 CrossRef CAS.
  50. H. M. F. Freundlich, J. Phys. Chem., 1906, 57, 385 CAS.
  51. M. Naushad, T. Ahamad, Z. A. Alothman and A. Aldalbahi, Chem. Eng. J., 2014, 254, 181 CrossRef.
  52. C. Namasivayam, R. T. Yamura and D. J. S. E. Arasi, Environ. Geol., 2001, 41, 269 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra27525c

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.