Next Article in Journal
Role of Microalloying Elements on Recrystallization Kinetics of Cold-Rolled High Strength Low Alloy Steels
Previous Article in Journal
Fatigue Resistance Analysis of the Orthotropic Steel Deck with Arc-Shaped Stiffener
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Pb(II) from Water by FeSiB Amorphous Materials

1
State Key Laboratory of Advanced Processing and Recycling of Nonferrous Metals, Lanzhou University of Technology, Lanzhou 730050, China
2
Wenzhou Engineering Institute of Pump & Valve, Lanzhou University of Technology, Wenzhou 325105, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(10), 1740; https://doi.org/10.3390/met12101740
Submission received: 15 August 2022 / Revised: 11 October 2022 / Accepted: 13 October 2022 / Published: 17 October 2022

Abstract

:
Amorphous materials have shown great potential in removing azo dyes in wastewaters. In this study, the performance of FeSiB amorphous materials, including FeSiB amorphous ribbons (FeSiBAR), and FeSiB amorphous powders prepared by argon gas atomization (FeSiBAP) and ball-milling (FeSiBBP), in removing toxic Pb(II) from aqueous solution was compared with the widely used zero valent iron (ZVI) powders (FeCP). The results showed that the removal efficiency of all the amorphous materials in removing Pb(II) from aqueous solution are much better than FeCP. Pb(II) was removed from aqueous solution by amorphous materials through the combined effect of absorption, (co)precipitation and reduction. Furthermore, FeSiBAP and FeSiBBP have relatively higher removal efficiencies than FeSiBAR due to a high specific surface area. Although the FeSiBBP has the highest removal efficiency up to the first 20 min, the removal process then nearly stopped due to aggregation.

1. Introduction

The rapid development of the modern industry brings more and more threat to our environment. For example, a great deal of lead-containing sewage is produced by smelting, manufacturing, and the oil industry. Excessive lead intake may cause neurological, hematological, and immunological disease. Therefore, these lead ions in the sewage must be removed before it is discharged into the environment. In previous decades, ZVI was widely used to remove various heavy metal ions in water, but it can be easily oxidized, which causes the degradation capability to decay rapidly [1]. In order to improve its degradation capability, various methods have been proposed, including nanotechnology [2,3] and bimetallic technology [4,5], but new limitations such as high production costs, physiological toxicity, and aggregation restrict their application.
Recent reports show that Fe-based amorphous materials in ribbon state or powder state can not only degrade azo dyes from the aqueous solution more efficiently than the wildly used FeCP but also exhibit relatively stable reusability due to their homogeneous microstructure, thermodynamically metastable nature, and the existence of metalloid elements [6,7,8,9,10]. Furthermore, researchers found that ball-milled amorphous powders can further improve the degradation capacity in degrading azo dyes due to the uneven topography and the stored deformation energy induced by ball milling [11,12,13,14,15]. In view of this, Fe-based amorphous materials in various states may also be able to efficiently remove lead ions in the wastewater.
In this paper, the removal capability and mechanism of lead ion (Pb(II)) from aqueous solution by FeSiB amorphous ribbons were investigated and compared with those of the widely used FeCP. Furthermore, considering that a more specific surface area always leads to a higher reaction efficiency, the removal capacity and mechanism of FeSiB amorphous powders prepared by argon gas-atomization and ball-milling were also studied.

2. Materials and Methods

The amorphous ribbons and argon gas-atomized amorphous powders with a composition of Fe78Si9B13 (atomic ratio) were provided by Jiangsu Chuangling Technology Co. and Haoxi Nano technology Co. of China, respectively. The FeCP was provided by Tianjin Chemical Reagent Research Institute. The amorphous ribbons were cut into pieces of 3 mm × 3 mm × 0.03 mm. Of these, part of the ribbons were used directly as removal material for Pb(II), and the rest were annealed at 653 K (below the crystallization temperature) for 1 h in an argon atmosphere and then ground by a high-energy ball mill under an argon atmosphere for 4 h at 200 r/min, with a ball-to-powder mass ratio of 10:1. Afterwards, the pulverized fine powders were filtered by mesh screens of 300 meshes per inch and used as removal material for Pb(II).
Pb(II) solutions were prepared by dissolving Pb(NO3)2 salt in ultra-pure water. The reaction tests were conducted in 1000 mL beakers placed in a temperature-controlled water-bath device and continuously stirred at 200 r/min. All tests were conducted with removal materials of 1 g/L and Pb(II) solutions of 500 mL, with a concentration of 100 mg/L at pH 5. About 5 mL of solution was extracted at fixed time intervals by a syringe and filtered through a 0.2 μm membrane. Then, they were tested by Inductively Coupled Plasma Optical Emission Spectrometer (ICP-OES, ICAP-7000, Thermo Fisher Scientific, Waltham, MA, USA) to determine their concentration. Cycling experiments were carried out one after another with the same ribbons.
Structure and morphology of the materials were investigated by X-ray diffraction (XRD, D/MAX-2400, monochromated Cu-Kα radiation) (Rigaku Ltd., Tokyo, Japan) equipped with an Ni filter and a graphite crystal monochromator, and by scanning electron microscopy (SEM, QUANTA FEG 450, FEI Ltd., Hillsborough, Atlanta, GA, USA) equipped with an Energy Dispersive X-ray Spectrometer (EDS). Specific surface areas of materials were recorded on the Brunauer–Emmett–Teller (BET, ASAP2020, Micromeritics Ltd., Atlanta, GA, USA) using the nitrogen method. Surface elemental information of the materials was determined by X-ray photoelectron spectroscopy (XPS, AXIS SUPRA, Shimadzu Ltd., Manchester, UK).

3. Results and Discussion

3.1. Removal Capacity of the Materials

Figure 1 shows the XRD spectrum of the four kinds of removal materials. Sharp crystalline peaks of a single α-Fe phase are observed on the spectrum of FeCP. The spectra of the other materials present a broad hump in the 2θ range of 40–50°, indicating their amorphous structure. Furthermore, there are tiny crystalline peaks superimposed on the broad humps, indicating that a small part of the samples may crystallize during the preparation process of the amorphous materials.
Figure 2 shows SEM images of the materials. The morphologies of FeCP, FeSiBAR, FeSiBAP, and FeSiBBP are shown in Figure 2a–d. All the particles are well-dispersed with no aggregation. The particle size of FeCP is about 20–50 μm. Grain boundaries can be obviously seen from the enlarged surface morphology in Figure 2a, indicating that these particles are polycrystalline, composed of crystal grains with a size of about 2–5 μm, whereas the thickness of FeSiBAR is about 30 μm (Figure 2b). SEM images of both the top and bottom sides of the ribbon are shown in the inset of Figure 2b. It can be seen that both sides of the ribbon are generally smooth, even if ripples are shown on the roll-contact surface. Compared with the round and ellipsoid morphology of the 10–20 μm sized FeSiBAP particles (Figure 2c), the morphology of the FeSiBBP appears to be rather irregular (Figure 2d). Furthermore, compared with the as-quenched ribbon, thickness of the FeSiBBP reduced to between 5 and 15 μm due to severe mechanical deformation during the ball-milling process, as shown in Figure 2e. Additionally, the surfaces of FeSiBAP are smooth, but the surfaces of FeSiBBP are full of protrusions and microcracks, as shown in the high-magnitude image of Figure 2f.
The removal effects of the four materials on Pb(II) solutions at room temperature were investigated. Figure 3 compares the removal performance, dependent on reaction time, for the four materials. Ct is the concentration of Pb(II) at a reaction time of t, and C0 is the initial concentration of Pb(II). It can be seen obviously that the removal efficiency of all the amorphous materials are much better than that of FeCP, among which FeSiBBP has the highest removal efficiency up to the first 20 min. Wang et al. [15] also found that the dye degradation capability of Fe-based amorphous ribbons could be enhanced by ball-milling. However, different from Wang’s reported results, the removal process seems to stop working and increasing the reaction time cannot effectively reduce the Pb(II) when the Pb(II) concentration is reduced to about 89% at 20 min. Considering the high residual Pb(II) concentration, despite its high removal rate, FeSiBBP is not an optimal removal material for Pb(II) from aqueous solution. Furthermore, except for FeSiBBP, the removal process by other materials can be well-fitted by the first-order reaction equation (Equation (1)) [16]:
Ct/C0 = exp(−kt)
where k denotes the reaction rate constant, which is estimated to be 0.026, 0.079, and 0.116 min−1 for the removal processes by FeCP, FeSiBAR, and FeSiBAP by nonlinear regression analysis, respectively. Therefore, the removal rates of FeSiBAR and FeSiBAP are approximately 21 and 87 times as fast as the widely used FeCP.
Considering that the particle size of FeSiBAP is much smaller than that of FeCP, specific surface areas of FeCP and FeSiBAP are measured to be 0.339 and 0.329 m2/g by a BET analysis, respectively. Thus, the surface area normalized rate constants kSA of FeCP and FeSiBAP are calculated to be 0.07 and 0.353 L/(m2·min) by the equation: kSA = k/ρa, where ρa is the surface area concentration of the sample [17]. Therefore, the reaction rate of per-unit surface area of FeSiBAP is about 99 times that of FeCP.

3.2. Removal Mechanism of the Amorphous Ribbons

Reaction active energy (ΔE) is an effective parameter for exploring the removal mechanism of various reactions. Figure 4 shows the dependence of normalized concentration of Pb(II) on reaction time at different temperatures for FeCP and FeSiBAR. Based on the calculated reaction rate constants at different temperatures, the reaction active energy ΔE can be derived using the Arrhenius-type equation, ln k = −ΔE/RT + lnA, where R is the gas constant, and A is a constant [16]. ΔE is calculated to be 31.2 and 22.4 kJ/mol for FeCP and FeSiBAR, indicating that the reaction between FeSiBAR and the Pb(II) solution is easier. According to the perspective of thermodynamics, diffusion-controlled reactions in solution have relatively lower activation energies (~8–21 kJ/mol), whereas the surface-controlled chemical reactions have larger activation energies (>29 kJ/mol) [18,19]. Consequently, the reaction between FeCP and the Pb(II) solution is surface-controlled, whereas interface resistance in the reaction between FeSiBAR and the Pb(II) solution has been greatly reduced. The results are consistent with the report that ZVI removes Pb(II) from solution, mainly by absorption and coprecipitation [20].
In order to investigate the removal mechanism of the amorphous ribbon, XRD patterns of the reacted FeSiBAR are compared with the reacted FeCP. Figure 5 shows XRD patterns of the FeSiBAR and FeCP after reaction with the aqueous solution for 40 min at room temperature. Diffraction peaks of FeO(OH) appear on both the XRD patterns of the reacted FeSiBAR and FeCP. According to the studies of treatment on Pb(II)-contaminated solution by ZVI, the FeO(OH) complexes with strong flocculation coprecipitate with Pb(II) during the reaction process (Equation (2)) [19]. Therefore, a small part of Pb(II) in aqueous solution was also removed by FeSiBAR by absorption and coprecipitating with Pb(II). Furthermore, compared with FeCP, a large number of diffraction peaks of Pb are found on XRD patterns of the reacted FeSiBAR and the precipitate of FeSiBAR, indicating that reduction (Equation (3)) played an important role in the removal process of Pb(II) by FeSiBAR. In addition, there are some tiny unknown diffraction peaks found on the XRD patterns of the reacted FeSiBAR and the precipitate. Combined with XRD patterns of the reacted FeSiBAP and FeSiBBP, which will be discussed later, these tiny diffraction peaks may represent SiO2 and PbO.
Absorption :   Fe OOH + Pb 2 + Fe OOPb + 2 H +
Reduction :   Fe 0 + Pb 2 + Fe 2 + + Pb 0
In addition, it was reported that hydroxide ions (OH) were generated by the reaction of Fe0 with H2O (Equations (4) and (5)) at a pH above 4.5 [21], leading to the increase in the pH of the solution. Then, oxides and hydroxides of iron and lead cover the ZVI particles or form precipitates (Equations (6)–(11)) [2,22,23]. However, some of the reactions cannot be observed by the XRD patterns.
Fe 0 + O 2 + 2 H 2 O 2 e Fe 2 + + 4 OH + H 2
Fe 2 + + O 2 + H 2 O + 3 e Fe 3 + + O 2 + 2 OH
Fe 2 + + 2 OH Fe ( OH ) 2
Fe 3 + + OH Fe ( OH ) 3
Fe ( OH ) 3 FeOOH + H 2 O
FeOOH Fe 2 O 3 + H 2 O
( Co ) precipitation :   Pb 2 + + OH Pb ( OH ) 2 + PbO xH 2 O
Re - dissolution :   Pb ( OH ) 2 + OH Pb ( OH ) 4 2
In view of the limited resolution of XRD, XPS is used to further investigate the surface chemical composition of FeSiBAR and FeCP. Figure 6 shows Pb 4f7/2 spectra of the reacted FeSiBAR and FeCP. It can be observed that only Pb(II) (138.7 eV) can be found on the surface of the reacted FeCP. However, both Pb(II) (138.5 eV) and Pb(136.7 eV) [24] are found on the surface of the reacted FeSiBAR. This discovery further confirms the XRD results showing that the removal of Pb(II) by FeCP is mainly through absorption and (co)precipitation, while the removal mechanism of Pb(II) by FeSiBAR involves absorption, (co)precipitation, and reduction.
Furthermore, the XPS spectra of FeSiBAR before and after reaction are compared in Figure 7. The surface of the as-received FeSiBAR was surrounded by oxide and hydroxide of iron, silicon oxide, and boron oxide. However, the XPS peak of O 1s disappeared after Ar+ sputtered at 0.4 nm/s for 20 s, suggesting that the thickness of the oxidation layer on the surface of the FeSiBAR produced in the atmosphere is about 8 nm. Furthermore, compared with the nominal composition of Fe78Si9B13 amorphous ribbon, the atomic ratios of Fe:Si:B on the surface of the as-received FeSiBAR is 47:37:16, indicating that Si and B atoms are enriched on the surface due to a lower binding energy and the larger diffusion coefficient of Si and B elements [25]. The enrichment of Si and B element may leave an incompact Fe layer on the surface, which may provide an incompact deposition site for chemical reaction. The possible compositions of the surface layer for the as-received FeSiBAR are ascribed to be Fe2O3(713.45 eV), FeOOH (711.02, 724.35 eV), Fe2+ (710.4 eV), Fe0 (706.65 eV), and oxides of boron and silicon [26,27,28]. After reaction, there are a few changes in the chemical compositions on the surface of the Fe-Si-BAR. It contains Fe2O3(713.6, 726.61 eV); FeOOH (711.21, 724.25 eV); Fe2+ (710.51 eV); and oxides of boron, silicon, and lead [26,27,28], indicating the reactions of Equations (4)–(11) involved in the Pb(II) removal process by FeSiBAR. Therefore, the combined effect of absorption, (co)precipitation, and reduction lead to the high removal efficiency of FeSiBAR, which is similar to the removal mechanism of Pb(II) by nano ZVI [29].
SEM images of the precipitate and reacted FeSiBAR were also observed. Figure 8 shows microscopic morphology of the precipitate. The main features observed are a large number of particles with a size of about 5 μm surrounded by a thin layer of reaction product. It can be seen from the enlarged image in Figure 8b that the particles are surrounded by a villous product layer. EDS elemental analyses in Figure 8d show that the particles are mainly Pb crystallines. This result is in accordance with the XRD analyses that show that the precipitates are mainly Pb. The irregular multi-prism morphology of crystal Pb can be clearly seen from the backscattered electron image of the particles in Figure 8c due to the large difference in atomic number between Pb and the other elements.
Figure 9 shows SEM images of the reacted FeSiBAR. Parts of the product layer have peeled off the ribbon (region B) and have formed bright and dark regions on the surface of the FeSiBAR, as shown in Figure 9a. The micrograph of region A and B are shown in high resolution in Figure 9b,c, respectively. Cracks have emerged on the surface of the residual product layer, as shown in Figure 9b. Following this, pieces of product layer tear along the cracks due to agitation and expose a fresh amorphous matrix. Then, a new product layer generates, as shown in Figure 9c. It can be seen that lamellar PbO, irregular multi-prism Pb, and worm-like iron oxide or silicon oxide have covered the fresh amorphous matrix. Furthermore, the high-magnitude image of region C in Figure 9d shows that the new product layer is a loose porous structure. The easily detached and loose porous product layer not only helps the elements exchange and accelerates the reaction process but also contributes to the reuse and durability performance of the FeSiBAR.
ZVI is the typical material for water purification, but fast corrosion always leads to rapid decay of its efficiency [28]. To verify the reusability of FeSiBAR, repeated removal experiments were carried out for six 90 min cycles. Figure 10 shows the removal efficiency of FeSiBAR for six cycles. Although the removal efficiency gradually decreases after every cycle, removal efficiency up to 54% is still present until the sixth cycle, indicating that FeSiBAR can be reused conveniently for a few times without any treatment.
The summarized removal mechanism of FeSiBAR is illustrated in Figure 11. FeSiBAR can be rapidly covered by oxidations of Fe, Si, and B in the atmosphere. However, in an oxygenated solution environment, Fe2+ and B3+ preferentially dissolved into the solution, leaving a porous and incompact layer of Si oxide on the surface of FeSiBAR. This not only provides an incompact transport channel for element exchange, but also makes the product layer easy to be peeled off from the ribbon and keeps the ribbon having a relatively stable reusability. Furthermore, the Fe0 beneath the oxide layer acts as an electron donor through the anodic reaction, Fe→Fe2+ + 2e, and releases a steady stream of Fe2+ ions into the solution. At the same time, Pb2+ is adsorbed on the oxidation layer of FeSiBAR and reduced by Fe0. In addition, the reaction of Fe0 with H2O generates hydroxide ions (OH) and increases the pH of the solution. Then, oxides and hydroxides of iron and lead may cover the FeSiBAR or form precipitates. In this process, the hydroxides of iron can form surface complexes with Pb(II) and act as an adsorbent of Pb(II), while the precipitate of lead oxide can separate Pb(II) from the solution.
In conclusion, the removal of Pb(II) by FeCP is a surface-controlled process and is mainly controlled by absorption and coprecipitation, but the results suggest that, because of the special amorphous structure, FeSiBAR not only removes Pb(II) in solution by absorption, (co)precipitation, and reduction, but also produces a loose porous product layer. The combination of these effects leads to the high removal efficiency of FeSiBAR.

3.3. Removal Mechanism of the Amorphous Powders

Figure 12 shows XRD patterns of the FeSiBAP and FeSiBBP after a reaction with the aqueous solution for 40 min at room temperature. The particles are too small to be distinguished from the precipitate, so the samples used for the XRD tests, included the reacted particles and the precipitate. XRD patterns of the FeSiBAP and FeSiBBP show nearly the same diffraction peaks that Pb, PbO, FeO(OH), and SiO2 are superimposing on the broad hump of amorphous particles. The reaction product is consistent with the reaction product of FeSiBAR.
Figure 13 shows SEM images of the FeSiBAP after reaction with the aqueous solution for 20 min at room temperature. The reacted FeSiBAP maintains its round or ellipsoid morphology and is covered with heterogeneous phases. It can be seen from Figure 13b that the phases appear to have the same morphologies as the product layers of the reacted FeSiBAR. Minor amounts of irregular multi-prism Pb are also present on surface of the reacted FeSiBAP, as shown in high resolution in Figure 13c. However, the sizes of the Pb particles are much smaller than those produced by the reaction between FeSiBAR and the Pb(II) solution, which may be related to the short reaction time and relatively dispersed Pb crystalline due to the large specific surface area of the FeSiBAP. Additionally, it can be seen from region B that pieces of product layer have peeled off and a new product layer composed of worm-like oxides has generated, as shown in high resolution in Figure 13d. The same phenomenon also appears on the surface of the reacted FeSiBAR. Therefore, it can be concluded that FeSiBAR and FeSiBAP have nearly the same reaction mechanism. Compared with FeSiBAR, it is the high specific surface area that leads to the higher removal efficiency of FeSiBAP.
It is interesting that the specific surface area of FeSiBBP is measured to be 0.0649 m2/g, corresponding to one fifth of the specific surface area of FeSiBAP, but FeSiBBP has the highest removal efficiency after the first 20 min. This phenomenon must be related to the special microstructure induced by the ball-milling process. According to Wang [11], huge residual stress and plastic deformation energy are stored in the FeSiBBP due to intense plastic deformation during the ball-milling process, which may facilitate the reaction activity and contribute to the low reaction active energy. Figure 14 shows SEM images of the FeSiBBP after reaction for 20 min at room temperature. Compared with FeSiBAP, the reacted FeSiBBP and the reaction products gather together, as shown in Figure 14a, which may be responsible for the slowing of the removal process after reacting for 20 min.
It can be seen from the enlarged image of region B in Figure 14c that the aggregation is composed of reaction product particles (marked as P), flakes, and the FeSiBBP residue (marked as R).
The EDS elemental analysis of the reaction product particles in Figure 15 shows that the particles are mainly Pb crystalline, according with the reaction precipitates of FeSiBAR. Additionally, there are more Pb particles found on the surface layer of FeSiBBP than on surface of FeSiBAP, as shown in Figure 14b. It can be deduced from so large a number of reaction products of Pb crystalline that the reduction is enhanced and plays a main role in the removal process of Pb(II) by FeSiBBP, which may be induced by the high deformation energy stored in the FeSiBBP.
Furthermore, microcracks formed in the ball-milling process on the original FeSiBBP also affect the reaction process. These microcracks may grow and propagate due to the crevice corrosion effect and the huge residual stress in the agitated Pb(II) solution, especially when the FeSiBBP is very thin. The thinner the FeSiBBP, the easier it is for the cracks to penetrate the FeSiBBP and break it into small pieces. Figure 14d shows surface details of the flake. It is obvious that the flake would have broken into many small pieces if it had been further agitated. The flakes in Figure 14c have a similar morphology. The increased specific surface areas induced by the breaking of the FeSiBBP may also contribute to the reaction process.
Therefore, even though the original specific surface areas of FeSiBBP are low, the combined effect of high deformation energy stored in the FeSiBBP and increased specific surface areas make them the most reactive material. However, the aggregation of the reaction product and the FeSiBBP residue leads to the rapid decay of its efficiency after reaction for 20 min.

4. Conclusions

In summary, the removal efficiency of all the amorphous materials, including FeSiBAR, FeSiBAP, and FeSiBBP, in removing Pb(II) from aqueous solution are much better than the widely used FeCP. Furthermore, it is interesting to find that FeSiBBP has the highest removal efficiency after the first 20 min, but the removal process stops at 20 min, when the Pb(II) concentration is reduced to about 89%.
The removal mechanism of FeSiBAR was discussed, and the results show that, different from the surface-controlled chemical reaction of FeCP, Pb(II) was removed from aqueous solution by FeSiBAR through the combined effect of absorption, (co)precipitation, and reduction. In addition, the product layer on the surface of FeSiBAR is easily detached and is a loose porous structure, which not only accelerates the reaction process but also makes FeSiBAR maintain a relatively stable reusability.
FeSiBAP and FeSiBBP have nearly the same reaction mechanism as FeSiBAR, but they have a relatively higher removal efficiency than FeSiBAR, due to high specific surface areas. Although FeSiBBP has the highest removal efficiency up to the first 20 min, aggregation of reaction product and the FeSiBBP residue makes the reaction process nearly halt after the first 20 min of reaction.
This study suggests that all the amorphous materials have great application potential in removal of Pb(II) from wastewater. Despite this, the long-term use of FeSiBBP for Pb(II) removal may need further treatment due to the rapid decay of its efficiency.

Author Contributions

Original draft preparation and Funding Acquisition, X.-Y.Z.; supervision, Z.-Z.Y.; investigation, J.-Y.D.; formal analysis, L.-L.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (NSFC) (No. 52061024 and No. 52161027), the Natural Science Foundation of Zhejiang Province (No. LQ20E010002), the Natural Science Foundation of Gansu Province (No. 21JR7RA260) and the Hongliu Research Funds of Lanzhou University of Technology for Distinguished Young Scholars.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors acknowledge all the authors, editors, and reviewers who contributed to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lv, Z.W.; Yan, Y.Q.; Yuan, C.C.; Huang, B.; Yang, C.; Ma, J.; Wang, J.Q.; Huo, L.S.; Cui, Z.Q.; Wang, X.L.; et al. Making Fe-Si-B amorphous powders as an effective catalyst for dye degradation by high-energy ultrasonic vibration. Mater. Des. 2020, 194, 108876. [Google Scholar] [CrossRef]
  2. Wang, J.; Liu, G.; Li, T.; Zhou, C.; Qi, C. Zero-Valent Iron Nanoparticles (NZVI) Supported by Kaolinite for CuII and NiII Ion Removal by Adsorption: Kinetics. Aust. J. Chem. 2015, 68, 1305–1315. [Google Scholar] [CrossRef]
  3. Wang, W.; Hua, Y.L.; Li, S.L.; Yan, W.L.; Zhang, W.X. Removal of Pb(II) and Zn(II) using lime and nanoscale zero-valent iron (nZVI): A comparative study. Chem. Eng. J. 2016, 304, 79–88. [Google Scholar] [CrossRef]
  4. Ma, H.; Huang, Y.; Shen, M.; Hu, D.; Yang, H.; Zhu, M.; Yang, S.; Shi, X. Enhanced decoloration efficacy of electrospun polymer nanofibers immobilized with Fe/Ni bimetallic nanoparticles. RSC Adv. 2013, 3, 6455–6465. [Google Scholar] [CrossRef]
  5. Pooja; Goel, A. A Highly Ultrafine Core–Shell Ir–Cu Bimetallic Nanoparticles and Their Application in Catalytic Oxidation of Textile Dye, Congo Red, by Hexacyanoferrate(III) Ions: A Kinetic Approach. Kinet. Catal. 2021, 62, 592–603. [Google Scholar] [CrossRef]
  6. Wang, F.; Wang, H.; Zhang, H.F.; Dan, Z.H.; Weng, N.; Tang, W.Y.; Qin, F.X. Superior azo-dye degradation of Fe-Si-B-P amorphous powders with graphene oxide addition. J. Non-Cryst. Solids 2018, 491, 34–42. [Google Scholar] [CrossRef]
  7. Zhang, L.C.; Jia, Z.; Lyu, F.; Liang, S.X.; Lu, J. A review of catalytic performance of metallic glasses in wastewater treatment: Recent progress and prospects. Prog. Mater. Sci. 2019, 105, 100576. [Google Scholar] [CrossRef]
  8. Liu, P.; Zhang, J.L.; Zha, M.Q.; Shek, C.H. Synthesis of an Fe rich amorphous structure with a catalytic effect to rapidly decolorize Azo dye at room temperature. ACS Appl. Mater. Inter. 2014, 6, 5500–5505. [Google Scholar] [CrossRef]
  9. Qin, X.D.; Li, Z.K.; Zhu, Z.W.; Fang, D.W.; Zhang, H.F. Fenton-like degradation of coke wastewater using Fe78Si8B14 and Fe73.5Nb3Cu1Si13.5B9 metallic glasses. J. Phys. Chem. Solids 2019, 133, 85–91. [Google Scholar] [CrossRef]
  10. Qin, X.D.; Li, Z.K.; Zhu, Z.W.; Fu, H.M.; Li, H.; Wang, A.M.; Zhang, H.W.; Zhang, H.F. Mechanism and kinetics of treatment of acid orange II by aged Fe-Si-B metallic glass powders. J. Mater. Sci. Technol. 2017, 33, 1147–1152. [Google Scholar] [CrossRef]
  11. Wang, J.Q.; Liu, Y.H.; Chen, M.W.; Xie, G.Q.; Louzguine-Luzgin, D.V.; Inoue, A.; Perepezko, J.H. Rapid Degradation of Azo Dye by Fe-Based Metallic Glass Powder. Adv. Funct. Mater. 2012, 22, 2567–2570. [Google Scholar] [CrossRef]
  12. Wang, J.Q.; Liu, Y.H.; Chen, M.W.; Louzguine-Luzgin, D.V.; Inoue, A.; Perepezko, J.H. Excellent capability in degrading azo dyes by MgZn-based metallic glass powders. Sci. Rep. 2012, 2, 418. [Google Scholar] [CrossRef] [Green Version]
  13. Qin, X.D.; Zhu, Z.W.; Liu, G.; Fu, H.M.; Zhang, H.W.; Wang, A.M.; Li, H.; Zhang, H.F. Ultrafast degradation of azo dyes catalyzed by cobalt-based metallic glass. Sci. Rep. 2015, 5, 18226. [Google Scholar] [CrossRef] [Green Version]
  14. Xie, S.H.; Huang, P.; Kruzic, J.J.; Zeng, X.R.; Qian, H.X. A highly efficient degradation mechanism of methyl orange using Fe-based metallic glass powders. Sci. Rep. 2016, 6, 21947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Miao, F.; Wang, Q.Q.; Di, S.Y.; Yun, L.; Zhou, J.; Shen, B.L. Enhanced dye degradation capability and reusability of Fe-based amorphous ribbons by surface activation. J. Mater. Sci. Technol. 2020, 53, 163–173. [Google Scholar] [CrossRef]
  16. Jia, Z.; Kang, J.; Zhang, W.C.; Wang, W.M.; Yang, C.; Sun, H.; Habibi, D.; Zhang, L.C. Surface aging behaviour of Fe-based amorphous alloys as catalysts during heterogeneous photo Fenton-like process for water treatment. Appl. Catal. B Environ. 2017, 204, 537–547. [Google Scholar] [CrossRef]
  17. Zhang, X.Y.; Liu, J.K.; Li, J.Q.; Li, C.Y.; Yuan, Z.Z. Excellent capability in remediating Cu2+ from aqueous solution by Fe–Si–B amorphous alloys. Appl. Phys. A 2020, 126, 291. [Google Scholar] [CrossRef]
  18. Tang, Y.; Shao, Y.; Chen, N.; Liu, X.; Chen, S.Q.; Yao, K.F. Insight into the high reactivity of commercial Fe-Si-B amorphous zero-valent iron in degrading azo dye solutions. RSC Adv. 2015, 5, 34032–34039. [Google Scholar] [CrossRef]
  19. Zhang, C.Q.; Zhu, Z.W.; Zhang, H.F. Effects of the addition of Co, Ni or Cr on the decolorization properties of Fe-Si-B amorphous alloys. J. Phys. Chem. Solids 2017, 110, 152–160. [Google Scholar] [CrossRef]
  20. Sun, J.F.; Tian, A.R.; Feng, Z.Y.; Zhang, Y.; Jiang, F.Y.; Tang, Q. Evaluation of Zero-Valent Iron for Pb(II) Contaminated Soil Remediation: From the Analysis of Experimental Mechanism Hybird with Carbon Emission Assessment. Sustainability 2021, 13, 452. [Google Scholar] [CrossRef]
  21. Liu, X.J.; Lai, D.G.; Wang, Y. Performance of Pb(II) Removal by an Activated Carbon Supported Nanoscale Zero-Valent Iron Composite at Ultralow Iron Content. J. Hazard. Mater. 2019, 361, 37–48. [Google Scholar] [CrossRef]
  22. Scott, T.B.; Popescu, I.C.; Crane, R.A.; Noubactep, C. Nano-scale metallic iron for the treatment of solutions containing multiple inorganic contaminants. J. Hazard. Mater. 2011, 186, 280–287. [Google Scholar] [CrossRef] [PubMed]
  23. Cheng, Y.; Jiao, C.; Fan, W.J. Synthesis and characterization of coated zero-valent iron nanoparticles and their application for the removal of aqueous Pb2+ ions. Desalin Water Treat. 2015, 54, 502–510. [Google Scholar] [CrossRef]
  24. Liu, M.H.; Wang, Y.H.; Chen, L.T.; Zhang, Y.; Lin, Z. Mg(OH)2 Supported Nanoscale Zero Valent Iron Enhancing the Removal of Pb(II) from Aqueous Solution. ACS Appl. Mater. Interfaces 2015, 7, 7961–7969. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, S.L.; Liu, H.P.; Fu, X.T. Oxidation Behavior of Fe-based Amorphous Ribbons. J. Iron Steel Res. Int. 2016, 23, 1219–1225. [Google Scholar] [CrossRef]
  26. Dai, M.Y.; Liu, Y.G.; Zeng, G.M.; Liu, S.B.; Ning, Q.M. Adsorption studies of 17β-dstradiol from aqueous solution using a novel stabilized Fe-Mn binary oxide nanocomposite. Environ. Sci. Pollut. Res. 2019, 26, 7614–7626. [Google Scholar] [CrossRef]
  27. Miao, S.; An, H.L.; Zhao, X.Q.; Wang, Y.J. Preparation of highly selective and stable Cu-Mg-Fe catalyst and its catalytic performance for one-step synthesis of 2-ethylhexanol from n-butyraldehyde. React. Kinet. Mech. Catal. 2019, 128, 395–412. [Google Scholar] [CrossRef]
  28. Houda, J.; Joaquin, M.-O.; Imne, M.; Mourad, M.; Teresa, B.; Grard, D. On the performance of Fe-Cu-ZSM-5 catalyst for the selective catalytic reduction of NO with NH3: The influence of preparation method. Res. Chem. Intermed. 2019, 45, 1057–1072. [Google Scholar] [CrossRef]
  29. Tarekegn, M.M.; Hiruy, A.M.; Dekebo, A.H. Nano zero valent iron (nZVI) particles for the removal of heavy metals (Cd2+, Cu2+ and Pb2+) from aqueous solutions. RSC Adv. 2021, 11, 18539. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD spectrum of the removal materials.
Figure 1. XRD spectrum of the removal materials.
Metals 12 01740 g001
Figure 2. SEM images of the removal materials: (a) FeCP, (b) FeSiBAR, (c) FeSiBAP, and (d) FeSiBBP. (e) and (f) are enlarged images of the particles in (d). The insets highlight surface details of the materials.
Figure 2. SEM images of the removal materials: (a) FeCP, (b) FeSiBAR, (c) FeSiBAP, and (d) FeSiBBP. (e) and (f) are enlarged images of the particles in (d). The insets highlight surface details of the materials.
Metals 12 01740 g002
Figure 3. The normalized concentration of Pb2+ ions dependence on the reaction time.
Figure 3. The normalized concentration of Pb2+ ions dependence on the reaction time.
Metals 12 01740 g003
Figure 4. Effect of temperature on reaction process of (a) FeCP and (b) FeSiBAR in the Pb(II) solution. Insert: Arrhenius plots of −lnk versus 1000/RT for the two kinds of materials, respectively.
Figure 4. Effect of temperature on reaction process of (a) FeCP and (b) FeSiBAR in the Pb(II) solution. Insert: Arrhenius plots of −lnk versus 1000/RT for the two kinds of materials, respectively.
Metals 12 01740 g004
Figure 5. XRD patterns of the reacted FeSiBAR and FeCP, and XRD pattern of the precipitate from the aqueous solution reacted with FeSiBAR.
Figure 5. XRD patterns of the reacted FeSiBAR and FeCP, and XRD pattern of the precipitate from the aqueous solution reacted with FeSiBAR.
Metals 12 01740 g005
Figure 6. Pb 4f7/2 spectra of the reacted FeSiBAR and FeCP.
Figure 6. Pb 4f7/2 spectra of the reacted FeSiBAR and FeCP.
Metals 12 01740 g006
Figure 7. XPS spectra of FeSiBAR before and after reaction: (a) whole XPS, (b) Fe 2p, (c) O 1s.
Figure 7. XPS spectra of FeSiBAR before and after reaction: (a) whole XPS, (b) Fe 2p, (c) O 1s.
Metals 12 01740 g007
Figure 8. (a) SEM images of the precipitate after FeSiBAR reacted with the Pb(II) solution. (b) Enlarged secondary electron image and (c) backscattered electron image of the particles in (a). (d) EDS elemental mapping images of the precipitate.
Figure 8. (a) SEM images of the precipitate after FeSiBAR reacted with the Pb(II) solution. (b) Enlarged secondary electron image and (c) backscattered electron image of the particles in (a). (d) EDS elemental mapping images of the precipitate.
Metals 12 01740 g008
Figure 9. (a) Surface layer of the reacted FeSiBAR and (b,c) enlarged views of the corresponding region A and B in (a), respectively. (d) Enlarged view of the corresponding region C in (c).
Figure 9. (a) Surface layer of the reacted FeSiBAR and (b,c) enlarged views of the corresponding region A and B in (a), respectively. (d) Enlarged view of the corresponding region C in (c).
Metals 12 01740 g009
Figure 10. Removal efficiency of FeSiBAR reacting with Pb(II) solutions for six 90 min cycles.
Figure 10. Removal efficiency of FeSiBAR reacting with Pb(II) solutions for six 90 min cycles.
Metals 12 01740 g010
Figure 11. Removal mechanism of FeSiBAR in Pb(II) aqueous solution.
Figure 11. Removal mechanism of FeSiBAR in Pb(II) aqueous solution.
Metals 12 01740 g011
Figure 12. XRD patterns of the reacted FeSiBAP and FeSiBBP.
Figure 12. XRD patterns of the reacted FeSiBAP and FeSiBBP.
Metals 12 01740 g012
Figure 13. (a,b) SEM images with different magnifications of the reacted FeSiBAP and (c,d) enlarged views of the corresponding regions A and B, respectively.
Figure 13. (a,b) SEM images with different magnifications of the reacted FeSiBAP and (c,d) enlarged views of the corresponding regions A and B, respectively.
Metals 12 01740 g013
Figure 14. (a) SEM images with different magnifications of the reacted FeSiBBP and (bd) enlarged views of the corresponding regions A, B, and C, respectively.
Figure 14. (a) SEM images with different magnifications of the reacted FeSiBBP and (bd) enlarged views of the corresponding regions A, B, and C, respectively.
Metals 12 01740 g014
Figure 15. SEM image of the reaction products for FeSiBBP and the corresponding EDS elemental mapping images. The insets highlight the details of the particles.
Figure 15. SEM image of the reaction products for FeSiBBP and the corresponding EDS elemental mapping images. The insets highlight the details of the particles.
Metals 12 01740 g015
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, X.-Y.; He, L.-L.; Du, J.-Y.; Yuan, Z.-Z. Removal of Pb(II) from Water by FeSiB Amorphous Materials. Metals 2022, 12, 1740. https://doi.org/10.3390/met12101740

AMA Style

Zhang X-Y, He L-L, Du J-Y, Yuan Z-Z. Removal of Pb(II) from Water by FeSiB Amorphous Materials. Metals. 2022; 12(10):1740. https://doi.org/10.3390/met12101740

Chicago/Turabian Style

Zhang, Xiang-Yun, Liang-Liang He, Jin-Ying Du, and Zi-Zhou Yuan. 2022. "Removal of Pb(II) from Water by FeSiB Amorphous Materials" Metals 12, no. 10: 1740. https://doi.org/10.3390/met12101740

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