Next Article in Journal
Realization of Lasing Emission from One Step Fabricated WSe2 Quantum Dots
Next Article in Special Issue
Investigation on Microstructure of Beetle Elytra and Energy Absorption Properties of Bio-Inspired Honeycomb Thin-Walled Structure under Axial Dynamic Crushing
Previous Article in Journal
Subwavelength Nanostructuring of Gold Films by Apertureless Scanning Probe Lithography Assisted by a Femtosecond Fiber Laser Oscillator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient and Reusable Montmorillonite/Fe3O4/Humic Acid Nanocomposites for Simultaneous Removal of Cr(VI) and Aniline

1
School of Chemical Engineering and Technology, Tianjin University, Tianjin 30072, China
2
State Key Laboratory of Chemical Engineering, Tianjin University, Tianjin 30072, China
3
Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 30072, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2018, 8(7), 537; https://doi.org/10.3390/nano8070537
Submission received: 15 June 2018 / Revised: 13 July 2018 / Accepted: 14 July 2018 / Published: 17 July 2018
(This article belongs to the Special Issue Nanotechnology for Clean Energy and Environmental Applications)

Abstract

:
Recyclable nanomaterials are in great need to develop clean technology for applications in the removal of water contaminants. In this work, easily separable montmorillonite/Fe3O4/humic acid (MFH) nanocomposites were fabricated through a facile hydrothermal route. It was found the adsorption ability and stability of MFH was significantly enhanced due to the synergistic effects between montmorillonite, Fe3O4 nanoparticles and humic acid. The MFH nanocomposites are highly efficient and recyclable as they can remove at least 82.3% of Cr(VI) and 95.1% of aniline in six consecutive runs. The adsorption mechanism was investigated by analyzing the kinetic parameters of pseudo first-order, pseudo second-order, and intraparticle diffusion models and describing the equilibrium isotherms of Langmuir and Freundlich models. Results indicated different adsorption mechanisms of Cr(VI) and aniline by MFH. The readily synthesized MFH nanocomposites can act as effective and practical materials for environmental applications.

Graphical Abstract

1. Introduction

As one of the most toxic heavy metals and common contaminants in wastewater, Cr can be found in wastewater from painting, tanning, and metallurgical industries [1,2]. It generally exists in the form of Cr(III) or Cr(VI) with the latter hundreds of times more toxic than the former. Cr(VI) is highly hazardous to organs, such as skin, liver, and lung, and even mutagenic to organisms [3]. In addition, it causes liver damage, pulmonary congestion, and skin irritation. Converting Cr(VI) to Cr(III) is a common and efficient strategy to solve the problem of Cr(VI). Meanwhile, aniline is a group of emerging contaminants in wastewater from many industrial processes, such as painting, pharmacy, and rubber [4]. Aniline can greatly harm human and aquatic life as it is highly toxic and it can accumulate in the environment [5]. It can enter the body through inhalation, digestion and skin absorption, convert hemoglobin to methemoglobin and result in cyanosis [4]. In recent years, growing attention has been focused on aniline removal strategies. In practical situations, heavy metals like Cr(VI) and aniline can coexist in industrial wastewater [6]. Thus, it is significant to achieve simultaneous removal of Cr(VI) and aniline from industrial wastewater to prevent their damage to the ecosystem and human health.
In recent years, various techniques of removing contaminants in wastewater have been investigated, including adsorption [7,8], biodegradation [9,10], photocatalysis [11,12], and biocatalytic oxidation [13,14]. Among them, adsorption has been the most widely used and effective approach in wastewater treatment. A great variety of adsorbents (e.g., active carbon [15], metal oxides [16], polymeric resins [17], zeolites [18], clays [19] and nano zerovalent iron [20]) have been utilized to remove Cr(VI) and aniline in wastewater. The application of Fe3O4 nanoparticles (NPs) as a highly efficient adsorbent attracts great research interest due to the advantages of relatively low cost, large specific surface area and good adsorption ability.
However, the small size of Fe3O4 NPs may cause problems such as easy oxidation, aggregation and difficult separation [21]. Dispersing Fe3O4 NPs on clays like montmorillonite (Mt) is an effective and economic strategy to overcome these problems [22,23]. Like Fe3O4 NPs, Mt can also adsorb contaminants in wastewater and further enhance the adsorption ability [24,25]. What is more, the stability and dispersity of Fe3O4 NPs can be improved by coating humic acid (HA) on the surface [26,27]. HA has strong complexation ability towards heavy metals and organic pollutants, which can restrict their mobility in water [28]. Besides, both Fe3O4 and HA have been reported to effectively reduce Cr(VI) to less toxic Cr(III) in many studies, and HA coated magnetite has also been utilized for Cr(VI) adsorption and reduction (See References [29,30,31] for detailed mechanisms). Studies demonstrated that substituted phenols, α-hydroxyl carboxylic acids, oxalic acid, and α-carbonyl carboxylic acids in HA participated in the reduction reaction of Cr(VI) and lower pH can make Cr(VI) more easily reduced by improving the redox potential of Cr(VI)/Cr(III) [30,31]. However, to get better adsorption capacity, stability, dispersity, and reducibility, Fe3O4, HA, and their composite need further modification.
In this study, montmorillonite/Fe3O4/humic acid (MFH) nanocomposites were readily prepared through a simple hydrothermal route. They were obtained by coating HA on Fe3O4 NPs surface and then dispersing the particles onto Mt. Apart from the advantage of easy synthesis, MFH also possessed strong adsorption ability, good stability and high recyclability due to the synergetic effects between the components. MFH showed great promise for applications in practical contaminants removal in wastewater.

2. Chemicals and Methods

2.1. Chemicals

FeCl3·6H2O (ACS), FeSO4·7H2O (≥99.0%), ammonia solution (25%) and humic acid sodium salt (AR) were obtained from Aladdin Biochemical Technology Co., Ltd., Shanghai, China. Na-montmorillonite (cation exchange capacity of 0.9 meq∙g−1) was produced in Beishan, Gansu, China. K2Cr2O7 (≥99.95%) and hydrochloric acid (36.5–38.0 wt%) were provided by Tianjin Guangfu Fine Chemical Research Institute, Tianjin, China. 1,5-diphenylcarbazide (AR) was provided by Tianjin Chemart Chemical Technology Co., Ltd., Tianjin, China. Aniline (≥99.5%) was purchased from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China. Ethanol absolute (≥99.8%) and NaOH (≥98.0%) were provided by Tianjin Real&Lead Chemical Technology Co., Ltd., Tianjin, China. Deionized water was utilized throughout the study.

2.2. Fabrication of MFH

As shown in Scheme 1, MFH nanocomposites of different Fe3O4/Mt mass ratios were obtained using a facile hydrothermal method as follows: 2.7962 g FeCl3·6H2O, 1.9580 g FeSO4·7H2O, 0.2280 g humic acid sodium salt and 1.5 g Mt (Fe3O4:Mt = 1:20)/0.75 g Mt (Fe3O4:Mt = 1:10)/0.3 g Mt (Fe3O4:Mt = 1:4)/0.075 g Mt (Fe3O4:Mt = 1:1) was dissolved in 60 mL 50% ethanol in water (v/v) and sonicated for 3 h. 4.6 mL of ammonia solution was added into the solution kept at 90 °C. Then, the suspension was transferred into a Teflon-lined autoclave (100 mL) and kept at 120 °C for 2 h. After the autoclave was cooled down naturally, the product was separated by centrifugation, washed by water and ethanol (three times each), and dried in a vacuum oven (60 °C, 6 h). For comparison, 2.7962 g FeCl3·6H2O and 1.9580 g FeSO4·7H2O was dissolved in 50% ethanol in water (v/v, 60 mL). Following the same procedure as described above, bare Fe3O4 NPs were obtained.

2.3. Characterization

Transmission electron microscopy (TEM) was conducted using a Tecnai G2 F20 (FEI, Eindhoven, The Netherlands) at 200 kV. The Brunauer–Emmett–Teller (BET) method was utilized to calculate the specific surface area (SBET). Nitrogen adsorption isotherms (T = 77 K) were obtained using an ASAP 2020 adsorption analyzer (Micromeritics, Norcross, GA, USA). Total pore volume (Vtp) and average pore diameter (Dp) were obtained using Barrett–Joyner–Halenda (BJH) method. A D/max-2500 X-ray diffraction analyzer (Cu Kα λ = 0.154 nm, 100 mA, 40 kV, 6° min−1, Rigaku, Tokyo, Japan) was used to collect X-ray diffraction (XRD) patterns. Fourier transform infrared spectroscopy (FTIR) was recorded by Tensor 27 (Bruker Optics, Ettlingen, Germany). X-ray photoelectron spectroscopy (XPS) was performed on an ESCALAB 250XI (Thermo Fisher, Waltham, MA, USA) for surface analysis. C element content was measured by TOC-VCPH (Shimadzu, Kyoto, Japan) and those of Si, Al and Fe elements were detected using inductively coupled plasma optical emission spectrometry (ICP-OES) iCAP 7400 (Thermo Fisher, Waltham, MA, USA). The content of Cr(VI) was determined using 1,5-diphenylcarbazide by a U-3010 UV-vis spectrophotometer (Hitachi, Tokyo, Japan) (λ = 540 nm) [32]. The concentration of aniline in samples was measured using a U-3010 UV-vis spectrophotometer (Hitachi, Tokyo, Japan) (λ = 230 nm) [33].

2.4. Batch Adsorption Experiments

One-hundred mL water with Cr(VI) (0–40.0 mg∙L−1) and aniline (0–100.0 mg∙L−1) at pH = 3.0–11.0 was added into in 250 mL flasks. The value of pH was controlled by adding 0.05 M HCl or NaOH. The solution was agitated at 400 rpm and kept at 25 °C. 0.05–0.40 g∙L−1 adsorbents were added in the flasks. Every 5 min during the experiment process (100 min), samples (1 mL) were taken out for UV analysis. After the nanocomposites were collected using a magnet, they were washed with 50% ethanol in water (v/v) several times and then dried. To evaluate the reusability, they were reused in the next five runs. Removal efficiency and equilibrium adsorption capacity qe (mg∙g−1) were obtained using Equations (1) and (2). All experiments were performed in triplicate.
Removal   efficiency = c 0 c t c 0 × 100 %
q e = ( c 0 c e ) V m
c0, ct and ce (mg∙L−1): initial, final and equilibrium concentrations, respectively; m (g): the amount of adsorbent added; V (L): the volume of solution.

3. Results and Discussion

3.1. Characterization

As shown in Figure 1a, bare Fe3O4 NPs tended to aggregate, which would reduce specific surface area, dispersity, mobility, and adsorption ability. Following the synthesis procedure in Section 2.2, HA-coated Fe3O4 NPs were dispersed onto Mt layers to overcome this problem. After sonication, Mt of fewer layers (See Figure 1b) can be obtained, which possessed higher specific surface area than normal Mt powder. As both Mt and HA were negatively charged in the form of Na-Mt and humic acid sodium salt respectively, HA was less likely to compete with Fe3O4 NPs to bind with Mt. The electrostatic attraction between Fe3O4 NPs and Mt, and between Fe3O4 NPs and HA, along with the electrostatic repulsion between Mt and HA, functioned as the main interaction forces to disperse HA-coated Fe3O4 NPs onto Mt and maintain the structure stability. The ratio of HA-coated Fe3O4 NPs loaded on Mt was controlled by varying the ratio between Mt and Fe sources. In Figure 1c–f, different mass ratios of HA-coated Fe3O4 NPs were loaded onto Mt. According to Table 1, MFH (Fe3O4:Mt = 1:4) had the highest specific surface area (SBET) among MFH of different mass ratios. With the increase of Fe3O4:Mt ratio in MFH, the total pore volume (Vtp) of MFH increased, and the average pore diameter (Dp) narrowed (Fe3O4:Mt < 1:4) and then widened (Fe3O4:Mt > 1:4). Compared with MFH in Figure 1c,d, the larger SBET and Vtp of MFH in Figure 1e was due to more Fe3O4 NPs loaded onto Mt. However, when Mt was overloaded with Fe3O4 NPs, the aggregation of Fe3O4 NPs would result in loss of SBET and increase of Dp compared with MFH in Figure 1e. Therefore, the MFH nanocomposites investigated in this paper were prepared according to the optimum mass ratio of 1:4 (Fe3O4:Mt). The SBET and Vtp of as-prepared MFH were increased significantly compared with those of Mt and bare Fe3O4 NPs. MFH enriched the pollutants by adsorption and significantly improved the removal efficiency [34,35].
The XRD patterns of Mt, Fe3O4 NPs and MFH are shown in Figure 2. For the XRD pattern of Mt, there were four diffraction peaks at 2θ = 20.4°, 35.9°, 54.1° and 62.2°. As for bare Fe3O4 NPs, the characteristic peaks at 2θ = 30.2°, 35.8°, 43.3°, 52.6°, 57.2°, and 62.6° can be found. In the XRD pattern of MFH, the typical diffraction peaks of both Mt and Fe3O4 NPs can be seen, demonstrating their high crystallinity and the successful dispersion of Fe3O4 NPs onto Mt. Particularly, the peak of Mt at 2θ = 20.4° showed a slight shift to 2θ = 19.6° in MFH. It could be explained by that as the insertion of larger hydrolyzed iron species replaced Na+ in initial Mt, the planar stress increased and thus resulted in the shift to lower angle in XRD pattern [36,37]. Note that as an organic component HA does not diffract and thus cannot be distinguished by XRD.
To clarify the coating of HA on Fe3O4 NPs, FTIR was carried out. In Figure 3a, the band at ~1030 cm−1 was attributed to the stretching vibration absorption of Si-O of Mt [38]. In the Fe3O4 curve (Figure 3b), the main absorption was observed at around 582 cm−1, corresponding to Fe–O bending vibration [39]. In Figure 3c, the typical C=O stretching of carboxylate in HA was located at ~1680 cm−1, and the band (~1350 cm−1) can be attributed to –CH2– scissoring [40]. In Figure 3d, all the characteristic bands of Mt, Fe3O4 NPs and HA mentioned above can be seen, verifying the formation of MFH. The C=O stretching in MFH (featured at ~1550 cm−1) showed blue shift compared with that of free HA (~1680 cm−1). This change can also provide evidence that the carboxylate anions of HA interacted with the surface of Fe3O4 NPs.
Besides, XPS was performed for surface composition analysis. The existence of Mt in MFH can be verified by the characteristic peaks of Al 2p and Si 2p (Figure 4b,c). Fe3O4 NPs can be confirmed by the peaks at 724.8 eV (Fe 2p1/2), 710.7 eV (Fe 2p3/2) and 530.2 eV (O 1s) (Figure 4d,e). The two split peaks of O 1s at 530.0 eV and 532.0 eV were assigned to Fe–O and hydroxyls, respectively. Besides, since spectrum of C 1s in MFH (Figure 4f) was distinguished, the existence of organic HA can also be verified. In summary, XPS indicated that the oxidation of Fe3O4 NPs was prevented by coating HA on Fe3O4 NPs surface and dispersing HA-coated Fe3O4 NPs onto Mt layers. As shown in Figure 4g, the Cr 2p3/2 peak of MFH after removal experiments can be split into two peaks at 574.3 and 573.2 eV (the characteristic peaks of Cr(VI) and Cr(III)), respectively, suggesting that both Cr(VI) and Cr(III) existed on MFH surface [41,42]. Therefore, it can be concluded that Cr(VI) can be reduced to Cr(III) species (e.g., Cr(OH)3) and precipitated on MFH surface.
The total organic carbon (TOC) content of raw HA was measured to be 43.81% (mass fraction). According to the dosage of HA used in MFH synthesis, the TOC content of as-prepared MFH was calculated to be 5.68%, which was very close to the measured value (5.34%). Besides, according to ICP-OES, the contents of Si, Al and Fe in MFH were 4.24%, 1.60% and 50.13%, respectively. As the contents of Si and Al in raw Mt were 24.18% and 8.86% respectively, theoretical contents of Si, Al and Fe in MFH were calculated to be 4.13%, 1.51%, and 49.4%, respectively. Therefore, the chemical compositions of as-prepared MFH were close to the initial dosage, which suggested the firm combination between Mt, Fe3O4 NPs and HA.
To study the recyclability, magnetic hysteresis loops of MFH were plotted in Figure 5. The saturation magnetization (Ms) of Fe3O4 NPs and MFH were measured to be 74.3 and 51.1 emu·g−1, respectively. The decrease of Ms was resulted from the total contents of Mt and HA in MFH, which turned out to be 31.2% after calculation [28]. The composition value agreed well with that the initial dosage and calculation value of ICP-OES and TOC (30.6%). With a relatively high Ms, MFH can be easily separated from water by a magnet.

3.2. Simultaneous Removal of Cr(VI) and Aniline

In Figure 6, the performance of MFH for removing Cr(VI) and aniline was demonstrated and compared with that of Mt, Fe3O4 NPs and HA. The removal efficiency of both Cr(VI) and aniline was MFH > Fe3O4 > Mt > HA. 84.8% of Cr(VI) and 89.2% of aniline could be removed by MFH in about 40 and 50 min, respectively. By comparing with the longer time and lower removal efficiency by Fe3O4 NPs, Mt and HA, the enhanced adsorption ability of MFH can be verified. As stated above, the specific surface area of MFH (98.86 m2∙g−1) was much larger than that of both Mt (53.72 m2∙g−1) and Fe3O4 NPs (77.24 m2∙g−1), which enriched the pollutants by adsorption and significantly improved the removal efficiency. Mt can adsorb contaminants in wastewater and further enhance the adsorption ability of Fe3O4 NPs. Coating HA onto Fe3O4 NPs greatly improved the stability and dispersity and further adsorbed more heavy metals and organic pollutants. Besides, HA also worked as reductant to reduce Cr(VI) to Cr(III) [29,30,31]. Therefore, the synergistic effects between Fe3O4 NPs, Mt and HA contributed greatly to the improvement of adsorption ability and removal efficiency in MFH.

3.3. Adsorption Kinetics

Adsorption kinetics helps to understand the mechanism, and the obtained data can be used to build mathematical models for interpreting interactions. To figure out the mechanisms in adsorption, pseudo first-order, pseudo second-order and intraparticle diffusion models (Equations (3)–(5)) were used for fitting the data from experiments.
ln ( q e q t ) = ln q e k 1 t
t q t = 1 k 2 q e 2 + t q e
q t = k i t 1 / 2 + C
qe and qt (mg∙g−1): the mass of contaminants adsorbed on adsorbent at equilibrium and t, respectively; k1 (min−1), k2 (g∙mg−1∙min−1) and ki (mg∙g−1·min−1/2): the pseudo first-, second-order and the intraparticle diffusion rate constants, respectively; C: a constant.
These three models were tested for the adsorption of Cr(VI) and aniline on MFH in this study. The results are shown in Table 2.
Seen from Table 2, for Cr(VI), the pseudo first-order model showed the best fit and a close value of qe,cal to qe,exp. As for aniline, R2 of the pseudo second-order model (R2 = 0.982) was the highest, indicating that the adsorption of aniline by MFH followed the pseudo second-order kinetic model. It suggested that chemisorption may limit the rate of aniline adsorption. Meanwhile, the intraparticle diffusion model (R2 = 0.975) also exhibited a good fit. Therefore, the adsorption of aniline on MFH may mainly occur in two steps: first, diffusion from water to MFH surface; second, the intraparticle diffusion between the pores of MFH. Cr-bentonite [43], activated carbon [44] and multi-walled carbon nanotubes [45] with similar mechanisms to MFH can be referenced for better understanding.

3.4. Adsorption Isotherm

For the evaluation of adsorption capacity and the interpretation of mechanism, Langmuir and Freundlich models were utilized in this study. The former (Equation (6)) is based on the assumptions of monolayer adsorption and identical adsorption sites. The latter (Equation (7)) describes heterogeneous surface and adsorption sites [43,46].
c e q e = 1 b q m + c e q m
ln q e = ln K + 1 n ln c e
qm (mg∙g−1): the maximum capacity for monolayer adsorption; b (L∙mg−1): a constant; K (mg1−(1/n) L1/n∙g−1): a constant; n: an empirical parameter [47,48].
According to Table 3, the adsorption of Cr(VI) and aniline on MFH followed Langmuir and Freundlich models, indicating monolayer and multilayer adsorption, respectively. As displayed in Table 4, MFH possessed higher adsorption capacity towards both Cr(VI) and aniline than some reported magnetic materials [49,50,51,52].

3.5. Effect of pH

Figure 7 shows that removing of both Cr(VI) and aniline was affected by pH. The adsorption rate was slowed dramatically if pH increased. Zero point charge pH (pHZPC) of MFH in this study was measured to be 4.3 (±0.4) by acid–base titration method [53]. Solid surfaces of MFH had positive and negative charges when pH was higher and lower than pHZPC, respectively. As the surface charge can be affected by pH, the surface of MFH will become more positively charged when pH decreases and thus will have stronger electrostatic attraction with Cr(VI) anions. Adsorption is often affected by electrostatic repulsion and attraction between adsorbent and adsorbate [30]. Stronger electrostatic repulsion due to higher pH will lead to less adsorption of Cr(VI) anions. As Equation (8) shows, studies have revealed that lower pH could enhance the redox potential of Cr(VI)/Cr(III), making Cr(VI) more easily to be reduced to Cr(III) [30]. Although unmodified Fe3O4 has reduction capacity and can reduce Cr(VI), Fe2+ in HA-Fe3O4 turned out not participate in Cr(VI) reduction [30,54,55]. On the contrary, higher pH leads to stronger electrostatic repulsion between them and thus the decrease of adsorption [2]. Moreover, at higher pH, the adsorption of Cr(III) species (e.g., Cr(OH)3) onto MFH surface reduces the adsorption ability towards contaminants [56]. On the other hand, aniline will be more protonated at lower pH (Equation (9)), which strengthens the electrostatic attraction with the negative carboxylic ions of HA in MFH. Increasing pH will significantly reduce the interaction between aniline and MFH. It can be concluded that lower pH is preferred to remove both Cr(VI) and aniline by MFH in wastewater. This tendency agrees well with previous studies on their individual removal [49,50,51,52,56].
HCrO 4 + 7 H + + 3 e Cr 3 + + 4 H 2 O
C 6 H 5 - NH 2 + H + C 6 H 5 - NH 3 +

3.6. Effect of MFH Dosage

The effect of MFH dosage on the simultaneous removal of Cr(VI) and aniline was investigated, as shown in Figure 8. The results indicate that no great improvement was achieved at a higher dosage, which should provide more adsorption sites and thus contribute to contaminants removal. However, excessive MFH cannot effectively improve the adsorption efficiency due to the concentration limit of contaminants. Therefore, 0.10 g∙L−1 turns out to be the optimized dosage of MFH for both Cr(VI) and aniline.

3.7. Mutual Effect between Cr(VI) and Aniline

Experiments were carried out to find whether Cr(VI) and aniline affected the removal of each other. As shown in Figure 9, when c0(aniline) and c0(Cr(VI)) were increased from 0 to 200 mg∙L−1 and 0 to 100 mg∙L−1 respectively, their corresponding removal efficiency barely improved. Since the upper limits of their initial concentrations were relatively high compared with actual wastewater, it can be assumed that they have no remarkable influence on the adsorption of each other by MFH. It is most likely that they are adsorbed by MFH in different mechanisms (Cr(VI), Langmuir, pseudo first-order kinetic; aniline, Freundlich, pseudo second-order kinetic), and the adsorption force is much stronger than the interaction between them.

3.8. Effect of NaCl Content

Inorganic salts of high contents, especially NaCl, can also be found in actual wastewater. In general, NaCl mainly comes from the salinity in initial water or that produced by adding chemical agents in the process. Since the high salinity may have influence on the removal of contaminants in actual wastewater, the effect of NaCl was also studied. Seen from Figure 10, the efficiency of Cr(VI) removal was slightly reduced with NaCl content increasing from 0 to 30% in wastewater. Therefore, it can be concluded that NaCl in wastewater has low effect on Cr(VI) removal, which follows Langmuir monolayer adsorption. On the other hand, the removal efficiency of aniline dramatically decreased with NaCl content increasing from 0% to 30% in wastewater. Negatively charged Cl may compete with carboxylic ions of HA in MFH to bind with the protonated aniline at low pH, and positively charged Na+ may compete against the protonated aniline for adsorption sites. As a result, the diffusion of aniline to MFH exterior surface and even the intraparticle diffusion in pore structures is impeded. Thus, the aniline removal efficiency drops and the influence will become more significant when the content of NaCl in wastewater increases.

3.9. Reusability

In a single run, the batch experiment was carried out as described in Section 2.4 and then the MFH were recycled from solution using a magnet, washed with 50% ethanol in water (v/v) several times and dried. The reusability was evaluated by reusing the reclaimed MFH in the next five runs. As Figure 11 shows, ≥82.3% of Cr(VI) and ≥95.1% of aniline can be removed in six consecutive cycles, demonstrating excellent reusability and structural stability.

4. Conclusions

In this study, montmorillonite/Fe3O4/humic acid (MFH) nanocomposites were easily fabricated through a hydrothermal route. The coating of HA onto Fe3O4 NPs and dispersion of HA-coated Fe3O4 NPs onto Mt can protect Fe3O4 NPs from oxidation and inhibit their aggregation. The synergistic effects between Mt, Fe3O4 NPs and HA turned out to effectively enhance the reactivity and efficiency of Fe3O4 NPs for the simultaneous removal of Cr(VI) and aniline from wastewater. The adsorption of Cr(VI) and aniline by MFH obey different mechanisms, Cr(VI)—Langmuir isotherm, pseudo first-order kinetic; aniline—Freundlich isotherm, pseudo second-order kinetic. Furthermore, the MFH nanocomposites possess good separation ability, stability, and reusability, making them promising materials for applications in water contaminants removal.

Author Contributions

H.L. designed and carried out the experiments and wrote this paper; X.H. and B.T. helped with characterizations; J.W., F.L. and H.H. provided suggestions on writing and revised the paper.

Funding

This research was funded by National Key Research and Development Program of China. Grant number 2016YFB0600504.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kazemi, M.; Jahanshahi, M.; Peyravi, M. Hexavalent chromium removal by multilayer membrane assisted by photocatalytic couple nanoparticle from both permeate and retentate. J. Hazard. Mater. 2018, 344, 12–22. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, X.; Luo, H.; Yin, Y.; Zhou, W. Facile synthesis of MoS2/reduced graphene oxide composites for efficient removal of Cr(VI) from aqueous solutions. RSC Adv. 2017, 7, 24149–24156. [Google Scholar] [CrossRef]
  3. Lu, H.J.; Wang, J.K.; Hao, H.X.; Wang, T. Magnetically Separable MoS2/Fe3O4/nZVI Nanocomposites for the Treatment of Wastewater Containing Cr(VI) and 4-Chlorophenol. Nanomaterials 2017, 7, 303. [Google Scholar] [CrossRef] [PubMed]
  4. Gu, Z.; Gao, M.L.; Luo, Z.X.; Ye, Y.G.; Liu, Y.N. Bis-pyridinium dibromides modified organo-bentonite for the removal of aniline from wastewater: A positive role of π–π polar interaction. Appl. Surf. Sci. 2014, 290, 107–115. [Google Scholar] [CrossRef]
  5. Jiang, L.; Lu, L.; Xiao, S.; Chen, J.M. Preparation of a novel manganese oxide-modified diatomite and its aniline removal mechanism from solution. Chem. Eng. J. 2016, 284, 609–619. [Google Scholar] [CrossRef]
  6. Huang, R.; Zheng, D.; Yang, B.; Wang, B. Preparation and simultaneous sorption of CTMAB-HTCC bentonite towards aniline and Cr (VI). Energy Source Part A 2016, 38, 519–526. [Google Scholar] [CrossRef]
  7. Parlayıcı, Ş.; Karakuzu, E.; Baybara, A.S.; Tuna, K.; Pehlivan, E. Utilization of eco-friendly gelatin for Cr(VI) adsorption. Desalin. Water Treat. 2017, 73, 308–315. [Google Scholar] [CrossRef]
  8. Chen, C.Y.; Geng, X.H.; Huang, W.L. Adsorption of 4-chlorophenol and aniline by nanosized activated carbons. Chem. Eng. J. 2017, 327, 941–952. [Google Scholar] [CrossRef]
  9. Ontañon, O.M.; González, P.S.; Barros, G.G.; Agostini, E. Improvement of simultaneous Cr (VI) and phenol removal by an immobilised bacterial consortium and characterisation of biodegradation products. New Biotechnol. 2017, 37, 172–179. [Google Scholar] [CrossRef] [PubMed]
  10. Li, X.H.; Jin, X.D.; Zhao, N.N.; Angelidaki, I.; Zhang, Y.F. Efficient treatment of aniline containing wastewater in bipolar membrane microbial electrolysis cell-Fenton system. Water Res. 2017, 119, 67–72. [Google Scholar] [CrossRef] [PubMed]
  11. Xie, Q.; Zhou, H.; Lv, Z.; Liu, H.; Guo, H. Sn4+ self-doped hollow cubic SnS as an efficient visible-light photocatalyst for Cr (VI) reduction and detoxification of cyanide. J. Mater. Chem. A 2017, 5, 6299–6309. [Google Scholar] [CrossRef]
  12. Pirsaheb, M.; Shahmoradi, B.; Beikmohammadi, M.; Azizi, E.; Hossini, H.; Md, A.G. Photocatalytic degradation of Aniline from aqueous solutions under sunlight illumination using immobilized Cr: ZnO nanoparticles. Sci. Rep. 2017, 7, 1473. [Google Scholar] [CrossRef] [PubMed]
  13. Pérez-Prior, M.T.; Gómez-Bombarelli, R.; González-Sánchez, M.I.; Valero, E. Biocatalytic oxidation of phenolic compounds by bovine methemoglobin in the presence of H2O2: Quantitative structure–activity relationships. J. Hazard. Mater. 2012, 241–242, 207–215. [Google Scholar] [CrossRef] [PubMed]
  14. Valero, E.; Pérez-Prior, M.T.; González-Sánchez, M.I. Removal of aromatic compounds from wastewater by hemoglobin soluble and immobilized on Eupergit® CM. Environ. Eng. Manag. J. 2014, 13, 2459–2466. [Google Scholar]
  15. Kołodyńska, D.; Krukowska, J.; Thomas, P. Comparison of sorption and desorption studies of heavy metal ions from biochar and commercial active carbon. Chem. Eng. J. 2017, 307, 353–363. [Google Scholar] [CrossRef]
  16. Wang, D.; Zhang, G.; Zhou, L.; Wang, M.; Cai, D.; Wu, Z. Synthesis of a Multifunctional Graphene Oxide-Based Magnetic Nanocomposite for Efficient Removal of Cr(VI). Langmuir 2017, 33, 7007–7014. [Google Scholar] [CrossRef] [PubMed]
  17. Kumar, R.; Arya, D.K.; Singh, N.; Vats, H.K. Removal of Cr (VI) Using Low Cost Activated Carbon Developed by Agricultural Waste. IOSR J. Appl. Chem. 2017, 10, 76–79. [Google Scholar] [CrossRef]
  18. Lv, G.; Li, Z.; Jiang, W.; Ackley, C.; Fenske, N.; Demarco, N. Removal of Cr(VI) from water using Fe(II)-modified natural zeolite. Chem. Eng. Res. Des. 2014, 92, 384–390. [Google Scholar] [CrossRef]
  19. Ma, L.; Xi, Y.; He, H.; Ayoko, G.A.; Zhu, R.; Zhu, J. Efficiency of Fe-montmorillonite on the removal of Rhodamine B and hexavalent chromium from aqueous solution. Appl. Clay Sci. 2016, 120, 9–15. [Google Scholar] [CrossRef]
  20. Lu, H.J.; Wang, J.K.; Ferguson, S.; Wang, T.; Bao, Y.; Hao, H.X. Mechanism, Synthesis and Modification of Nano Zerovalent Iron in Water Treatment. Nanoscale 2016, 8, 9962–9975. [Google Scholar] [CrossRef] [PubMed]
  21. Kalantari, K.; Ahmad, M.B.; Shameli, K.; Hussein, B.M.Z.; Khandanlou, R.; Khanehzaei, H. Size-Controlled synthesis of Fe3O4 magnetic nanoparticles in the layers of montmorillonite. J. Nanomater. 2014, 2014, 181. [Google Scholar] [CrossRef]
  22. Zhang, P.; Mo, Z.L.; Han, L.J.; Zhu, X.B.; Wang, B.; Zhang, C. Preparation and Photocatalytic Performance of Magnetic TiO2/Montmorillonite/Fe3O4 Nanocomposites. Ind. Eng. Chem. Res. 2014, 53, 8057–8061. [Google Scholar] [CrossRef]
  23. Mahdavinia, G.R.; Hasanpour, S.; Behrouzi, L.; Sheykhloie, H. Study on adsorption of Cu(II) on magnetic starch-g-polyamidoxime/montmorillonite/Fe3O4 nanocomposites as novel chelating ligands. Starch Stärke 2016, 68, 188–199. [Google Scholar] [CrossRef]
  24. Cai, J.; Lei, M.; Zhang, Q.; He, J.R.; Chen, T.; Liu, S.; Fu, S.H.; Li, T.T.; Liu, G.; Fei, P. Electrospun composite nanofiber mats of Cellulose@Organically modified montmorillonite for heavy metal ion removal: Design, characterization, evaluation of absorption performance. Compos. Part. A Appl. Sci. Manuf. 2017, 92, 10–16. [Google Scholar] [CrossRef]
  25. Zhu, R.; Chen, Q.; Zhou, Q.; Xi, Y.; Zhu, J.; He, H. Adsorbents based on montmorillonite for contaminant removal from water: A review. Appl. Clay Sci. 2016, 123, 239–258. [Google Scholar] [CrossRef]
  26. Ills, E.; Tombacz, E. The role of variable surface charge and surface complexation in the absorption of humic acid on magnetite. Colloids Surf. A 2003, 230, 99–109. [Google Scholar] [CrossRef]
  27. Ills, E.; Tombacz, E. The effect of humic acid adsorption on pH-dependent surface charging and aggregation of magnetite nanoparticles. J. Colloid Interface Sci. 2006, 295, 115–123. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, S.T.; Zong, P.F.; Ren, X.M.; Wang, Q.; Wang, X.K. Rapid and Highly Efficient Preconcentration of Eu(III) by Core−Shell Structured Fe3O4@Humic Acid Magnetic Nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 6891–6900. [Google Scholar] [CrossRef] [PubMed]
  29. Hu, J.; Lo, I.; Chen, G. Removal of Cr(VI) by magnetite. Water Sci. Technol. 2004, 50, 139–146. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, M.; Li, G.X.; Jiang, X.L.; Xiao, Q.Q.; Niu, M.X.; Wang, Z.Y.; Wang, Y.Y. Non-biological reduction of Cr(VI) by reacting with humic acids composted from cattle manure. RSC Adv. 2017, 7, 26903–26911. [Google Scholar] [CrossRef]
  31. Jiang, W.J.; Cai, Q.; Xu, W.; Yang, M.W.; Cai, Y.; Dionysiou, D.D.; O’Shea, E.K. Cr(VI) Adsorption and Reduction by Humic Acid Coated on Magnetite. Environ. Sci. Technol. 2014, 48, 8078–8085. [Google Scholar] [CrossRef] [PubMed]
  32. Franson, M.A.H. Standard Method for the Examination of Water and Wastewater, 19th ed.; American Public Health Association: Washington, DC, USA, 1995. [Google Scholar]
  33. Belkova, A.A.; Sergeeva, A.I.; Apel, P.Y.; Beklemishev, M.K. Diffusion of aniline through a polyethylene terephthalate track-etched membrane. J. Membr. Sci. 2009, 330, 145–155. [Google Scholar] [CrossRef]
  34. Lian, G.; Zhang, X.; Zhang, S.J.; Liu, D.; Cui, D.L.; Wang, Q.L. Controlled fabrication of ultrathin-shell BN hollow spheres with excellent performance in hydrogen storage and wastewater treatment. Energy Environ. Sci. 2012, 5, 7072–7080. [Google Scholar] [CrossRef]
  35. Wang, M.; Li, M.H.; Xu, L.Q.; Wang, L.C.; Ju, Z.C.; Li, G.D.; Qian, Y.T. High yield synthesis of novel boron nitride submicro-boxes and their photocatalytic application under visible light irradiation. Catal. Sci. Technol. 2011, 1, 1159–1165. [Google Scholar] [CrossRef]
  36. Masih, D.; Izumi, Y.; Aika, K.; Seida, Y. Optimization of an Iron Intercalated Montmorillonite Preparation for the Removalof Arsenic at Low Concentration. Eng. Life Sci. 2007, 7, 52–60. [Google Scholar] [CrossRef]
  37. Ebitani, K.; Ide, M.; Mitsudome, T.; Mizugaki, T.; Kaneda, K. Creation of a chain-like cationic iron species in montmorillonite as a highly active heterogeneous catalyst for alkane oxygenations using hydrogen peroxide. Chem. Commun. 2002, 7, 690–691. [Google Scholar] [CrossRef]
  38. Khang, V.C.; Korovkin, M.V.; Ananyeva, L.G. Identification of clay minerals in reservoir rocks by FTIR spectroscopy. In IOP Conference Series: Earth and Environmental Science; IOP Publishing: Bristol, UK, 2016; Volume 43, p. 012004. [Google Scholar]
  39. Liu, Y.; Li, Y.; Zhao, X.; Chi, W.; Huang, Q.; Yu, C.; Xiang, Y. Effect of magnetite nanoparticles on dye absorption properties of magnetite@carbon composites. Bull. Mater. Sci. 2017, 40, 367–373. [Google Scholar] [CrossRef]
  40. Liu, J.F.; Zhao, Z.S.; Jiang, G.B. Coating Fe3O4 Magnetic Nanoparticles with Humic Acid for High Efficient Removal of Heavy Metals in Water. Environ. Sci. Technol. 2008, 42, 6949–6954. [Google Scholar] [CrossRef] [PubMed]
  41. Yao, X.; Deng, S.; Wu, R.; Hong, S.; Wang, B.; Huang, J.; Wang, Y.; Yu, G. Highly efficient removal of hexavalent chromium from electroplating wastewater using aminated wheat straw. RSC Adv. 2016, 6, 8797–8805. [Google Scholar] [CrossRef]
  42. Kumar, A.S.K.; Jiang, S.J.; Warchoł, J.K. Synthesis and Characterization of Two-Dimensional Transition Metal Dichalcogenide Magnetic MoS2@Fe3O4 Nanoparticles for Adsorption of Cr(VI)/Cr(III). ACS Omega 2017, 2, 6187–6200. [Google Scholar] [CrossRef]
  43. Zheng, H.; Liu, D.H.; Zheng, Y.; Liang, S.P.; Liu, Z. Sorption isotherm and kinetic modeling of aniline on Cr-bentonite. J. Hazard. Mater. 2009, 167, 141–147. [Google Scholar] [CrossRef] [PubMed]
  44. Valderrama, C.; Barios, J.I.; Caetano, M.; Farran, A.; Cortina, J.L. Kinetic evaluation of phenol/aniline mixtures adsorption from aqueous solutions onto activated carbon and hypercrosslinked polymeric resin (MN200). React. Funct. Polym. 2010, 70, 142–150. [Google Scholar] [CrossRef]
  45. Al-Johani, H.; Abdel, S.M. Kinetics and thermodynamic study of aniline adsorption by multi-walled carbon nanotubes from aqueous solution. J. Colloid Interface Sci. 2011, 360, 760–767. [Google Scholar] [CrossRef] [PubMed]
  46. Stromer, B.S.; Woodbury, B.; Williams, C.F. Tylosin sorption to diatomaceous earth described by Langmuir isotherm and Freundlich isotherm models. Chemosphere 2018, 193, 912–920. [Google Scholar] [CrossRef] [PubMed]
  47. Baseri, H.; Tizro, S. Treatment of nickel ions from contaminated water by magnetite based nanocomposite adsorbents: Effects of thermodynamic and kinetic parameters and modeling with Langmuir and Freundlich isotherms. Process Saf. Environ. 2017, 109, 465–477. [Google Scholar] [CrossRef]
  48. Fakhri, A. Adsorption characteristics of graphene oxide as a solid adsorbent for aniline removal from aqueous solutions: Kinetics, thermodynamics and mechanism studies. J. Saudi Chem. Soc. 2017, 21, S52–S57. [Google Scholar] [CrossRef]
  49. Li, L.; Duan, H.; Wang, X.; Luo, C. Adsorption property of Cr(VI) on magnetic mesoporous titanium dioxide–graphene oxide core–shell microspheres. New J. Chem. 2014, 38, 6008–6016. [Google Scholar] [CrossRef]
  50. Fan, L.L.; Luo, C.N.; Sun, M.; Qiu, H.M. Synthesis of graphene oxide decorated with magnetic cyclodextrin for fast chromium removal. J. Mater. Chem. 2012, 22, 24577–24583. [Google Scholar] [CrossRef]
  51. Bhaumik, M.; Maity, A.; Srinivasu, V.V.; Onyango, M.S. Enhanced removal of Cr (VI) from aqueous solution using polypyrrole/Fe3O4 magnetic nanocomposite. J. Hazard. Mater. 2011, 190, 381–390. [Google Scholar] [CrossRef] [PubMed]
  52. Kakavandi, B.; Jafari, A.J.; Kalantary, R.R.; Nasseri, S.; Ameri, A.; Esrafily, A. Synthesis and properties of Fe3O4-activated carbon magnetic nanoparticles for removal of aniline from aqueous solution: Equilibrium, kinetic and thermodynamic studies. Iran. J. Environ. Health Sci. Eng. 2013, 10, 19. [Google Scholar] [CrossRef] [PubMed]
  53. Stumm, W.; Morgan, J.J. Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters, 3rd ed.; John Wiley & Sons: New York, NY, USA, 1996. [Google Scholar]
  54. He, Y.T.; Traina, S.J. Cr(VI) reduction and immobilization by magnetite under alkaline pH conditions: The role of passivation. Environ. Sci. Technol. 2005, 39, 4499–4504. [Google Scholar] [CrossRef] [PubMed]
  55. Eary, L.; Rai, D. Chromate reduction by subsurface soils under acidic conditions. Soil Sci. Soc. Am. J. 1991, 55, 676–683. [Google Scholar] [CrossRef]
  56. Hashem, F.S. Removal of methylene blue by magnetite-covered bentonite nano-composite. Eur. Chem. Bull. 2013, 2, 524–529. [Google Scholar]
Scheme 1. Fabrication of MFH nanocomposites.
Scheme 1. Fabrication of MFH nanocomposites.
Nanomaterials 08 00537 sch001
Figure 1. TEM of (a) bare Fe3O4 NPs; (b) Mt single layer obtained by sonication, different mass ratios of MFH nanocomposites: (c) Fe3O4:Mt = 1:20; (d) Fe3O4:Mt = 1:10; (e) Fe3O4:Mt = 1:4; (f) Fe3O4:Mt = 1:1.
Figure 1. TEM of (a) bare Fe3O4 NPs; (b) Mt single layer obtained by sonication, different mass ratios of MFH nanocomposites: (c) Fe3O4:Mt = 1:20; (d) Fe3O4:Mt = 1:10; (e) Fe3O4:Mt = 1:4; (f) Fe3O4:Mt = 1:1.
Nanomaterials 08 00537 g001
Figure 2. XRD patterns of (a) Mt; (b) Fe3O4 NPs and (c) MFH nanocomposites.
Figure 2. XRD patterns of (a) Mt; (b) Fe3O4 NPs and (c) MFH nanocomposites.
Nanomaterials 08 00537 g002
Figure 3. FTIR of (a) Mt; (b) Fe3O4 NPs; (c) HA; and (d) MFH nanocomposites.
Figure 3. FTIR of (a) Mt; (b) Fe3O4 NPs; (c) HA; and (d) MFH nanocomposites.
Nanomaterials 08 00537 g003
Figure 4. XPS of the MFH nanocomposites before (af) and after (g) removal experiments.
Figure 4. XPS of the MFH nanocomposites before (af) and after (g) removal experiments.
Nanomaterials 08 00537 g004
Figure 5. Magnetic hysteresis loops of Fe3O4 NPs and MFH nanocomposites.
Figure 5. Magnetic hysteresis loops of Fe3O4 NPs and MFH nanocomposites.
Nanomaterials 08 00537 g005
Figure 6. Simultaneously removing (a) Cr(VI) and (b) aniline using various adsorbents. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(adsorbent) = 0.1 g∙L−1, T = 25 °C, pH = 3.0).
Figure 6. Simultaneously removing (a) Cr(VI) and (b) aniline using various adsorbents. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(adsorbent) = 0.1 g∙L−1, T = 25 °C, pH = 3.0).
Nanomaterials 08 00537 g006
Figure 7. Simultaneously removing (a) Cr(VI) and (b) aniline by MFH at various pH. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, T = 25 °C).
Figure 7. Simultaneously removing (a) Cr(VI) and (b) aniline by MFH at various pH. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, T = 25 °C).
Nanomaterials 08 00537 g007
Figure 8. Simultaneously removing (a) Cr(VI) and (b) aniline at various MFH dosages. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, pH = 3.0, T = 25 °C).
Figure 8. Simultaneously removing (a) Cr(VI) and (b) aniline at various MFH dosages. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, pH = 3.0, T = 25 °C).
Nanomaterials 08 00537 g008
Figure 9. (a) Removing Cr(VI) with various initial concentrations of aniline by MFH. (c0(Cr(VI)) = 40.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C). (b) Removing aniline with various initial concentrations of Cr(VI) by MFH. (c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Figure 9. (a) Removing Cr(VI) with various initial concentrations of aniline by MFH. (c0(Cr(VI)) = 40.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C). (b) Removing aniline with various initial concentrations of Cr(VI) by MFH. (c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Nanomaterials 08 00537 g009
Figure 10. Simultaneously removing (a) Cr(VI) and (b) aniline at various NaCl contents. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Figure 10. Simultaneously removing (a) Cr(VI) and (b) aniline at various NaCl contents. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Nanomaterials 08 00537 g010
Figure 11. (a) Recycle of MFH from solution; (b) Reusability of MFH. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Figure 11. (a) Recycle of MFH from solution; (b) Reusability of MFH. (c0(Cr(VI)) = 40.0 mg∙L−1, c0(aniline) = 100.0 mg∙L−1, c(MFH) = 0.10 g∙L−1, pH = 3.0, T = 25 °C).
Nanomaterials 08 00537 g011
Table 1. The specific surface area and porous features of different materials.
Table 1. The specific surface area and porous features of different materials.
MaterialSBET (m2∙g−1)Vtp (cm3∙g−1)Dp (nm)
Bare Fe3O4 NPs77.240.247.5
Mt powder53.720.128.6
MFH nanocomposites (Fe3O4:Mt, mass ratio)1:2062.130.138.0
1:1076.450.157.6
1:498.860.196.5
1:173.460.207.7
Table 2. Adsorption kinetic models using MFH at T = 298 K.
Table 2. Adsorption kinetic models using MFH at T = 298 K.
ModelParameterCr(VI)Aniline
qe,exp (mg∙g−1)339.25 ± 2.42292.16 ± 2.06
Pseudo first-orderqe,cal (mg∙g−1)340.31 ± 2.97293.83 ± 2.64
K1 (min−1)0.132 ± 0.0020.109 ± 0.001
R20.986 ± 0.0010.911 ± 0.002
Pseudo second-orderqe,cal (mg∙g−1)341.19 ± 2.80294.22 ± 2.31
k2 (g∙mg−1∙min−1)0.0165 ± 0.00030.0274 ± 0.0005
R20.929 ± 0.0010.982 ± 0.001
Intra particle diffusionki (mg∙g−1·min−1/2)8.24 ± 0.0111.37 ± 0.02
C0.103 ± 0.0010.307 ± 0.002
R20.942 ± 0.0010.975 ± 0.001
Table 3. Adsorption isotherm models using MFH at T = 298K.
Table 3. Adsorption isotherm models using MFH at T = 298K.
ModelParameterCr(VI)Aniline
Langmuirqm (mg∙g−1)374.19 ± 3.11393.53 ± 3.32
b (L∙mg−1)0.018 ± 0.0010.096 ± 0.001
R20.995 ± 0.0010.941 ± 0.001
FreundlichK (mg1−(1/n) L1/n g−1)0.121 ± 0.00116.58 ± 0.09
n1.87 ± 0.013.74 ± 0.02
R20.907 ± 0.0020.993 ± 0.001
Table 4. Adsorption capacities of various magnetic adsorbents for Cr(VI) and aniline from wastewater.
Table 4. Adsorption capacities of various magnetic adsorbents for Cr(VI) and aniline from wastewater.
AdsorbateAdsorbentAdsorption Capacity (mg∙g−1)ConditionReference
Cr(VI)Fe3O4@mTiO2@GO117.94pH = 2.0, T = 303 K[49]
Graphene oxide decorated with magnetic cyclodextrin120pH = 3.0, T = 298 K[50]
PPY/Fe3O4243.9pH = 2.0, T = 298 K[51]
MFH nanocomposites374.19pH = 3.0, T = 298 Kthis study
AnilineFe3O4-activated carbon90.91pH = 6.0, T = 293 K[52]
Fe3O4/graphene202.84pH = 3.0, T = 298 K[46]
MFH nanocomposites393.53pH = 3.0, T = 298 Kthis study

Share and Cite

MDPI and ACS Style

Lu, H.; Wang, J.; Li, F.; Huang, X.; Tian, B.; Hao, H. Highly Efficient and Reusable Montmorillonite/Fe3O4/Humic Acid Nanocomposites for Simultaneous Removal of Cr(VI) and Aniline. Nanomaterials 2018, 8, 537. https://doi.org/10.3390/nano8070537

AMA Style

Lu H, Wang J, Li F, Huang X, Tian B, Hao H. Highly Efficient and Reusable Montmorillonite/Fe3O4/Humic Acid Nanocomposites for Simultaneous Removal of Cr(VI) and Aniline. Nanomaterials. 2018; 8(7):537. https://doi.org/10.3390/nano8070537

Chicago/Turabian Style

Lu, Haijiao, Jingkang Wang, Fei Li, Xin Huang, Beiqian Tian, and Hongxun Hao. 2018. "Highly Efficient and Reusable Montmorillonite/Fe3O4/Humic Acid Nanocomposites for Simultaneous Removal of Cr(VI) and Aniline" Nanomaterials 8, no. 7: 537. https://doi.org/10.3390/nano8070537

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop