Next Article in Journal
Effects of Enrofloxacin and Ciprofloxacin on Growth and Toxin Production of Microcystis aeruginosa
Previous Article in Journal
A Unified General Resistance Formula for Uniform Coarse Porous Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Removal of Cadmium (II) and Arsenic (V) from Water by the Composite of Iron Manganese Oxides Loaded Muscovite

1
Yellow River Institute of Hydraulic Research, Zhengzhou 450003, China
2
Research Center on Levee Safety and Disaster Prevention, Ministry of Water Resources, Zhengzhou 450003, China
3
Henan Yellow River Bureau Engineering Construction Center, Zhengzhou 450003, China
*
Author to whom correspondence should be addressed.
Submission received: 7 September 2023 / Revised: 1 October 2023 / Accepted: 5 October 2023 / Published: 13 October 2023

Abstract

:
It is a great challenge for a material with high adsorption capacity for cadmium (Cd) and arsenic (As) simultaneously. To address the simultaneous removal of Cd and As from water and the utilization of waste muscovite, the muscovite-supported iron manganese oxides (FMM) were prepared. The FMM was easy to adsorb Cd(II) and As(V), and the adsorption capacity of Cd(II) and As(V) was 32.47 mg/g and 28.57 mg/g, respectively. Iron manganese oxides significantly improved the Cd(II) and As(V) adsorption capacity, specific surface area, and pore volume of the material. Moreover, the adsorption process of FMM for Cd(II) and As(V) fitted well with the pseudo-first-order model and pseudo-second-order model. The mechanism including chemical adsorption, electrostatic adsorption and covalent bond adsorption was proposed for As(V) removal, while Cd(II) removal was based on the combination of electrostatic adsorption and surface precipitation. In conclusion, FMM is a promising material for the treatment of As and Cd-contaminated wastewater, which provides insight into the resource utilization of waste muscovite.

1. Introduction

Heavy metals and arsenic are commonly present in natural water bodies [1]. Heavy metals cannot be biodegraded, and heavy metals in water bodies can have negative impacts on aquatic organisms, such as growth and reproduction, which can seriously lead to death. Heavy metals entering the plant body can cause toxicity to plants by altering cell vision, inhibiting nutrient absorption, and reducing transpiration rate. Heavy metals can harm human health through the food chain [2,3]. Among these heavy metals, cadmium and arsenic are internationally recognized carcinogenic, teratogenic, and mutagenic elements [4], and human industrial activities produce Cd and As, resulting in Cd and As pollution. The use of water polluted by Cd and As to irrigate crops will result in crop reduction and pose a threat to food security production, and polluted drinking water will cause harm to human health. Therefore, the prevention and control of Cd-As polluted water bodies has important environmental significance.
In order to address the problem of Cd and As pollution in water, various heavy metal removal technologies have been developed, including membrane filtration [5], chemical precipitation [6], solvent extraction [7] and adsorption [8]. Membrane filtration, chemical precipitation, and solvent extraction methods can quickly remove heavy metals from water. However, these technologies have obvious drawbacks, such as low removal rate and high cost, and only removal effects under specific conditions [9]. Due to its advantages in design, recyclability and high efficiency, adsorption technology has the advantages of easy operation and low cost in the treatment of heavy metal-polluted water [10]. The key to the adsorption method is high efficiency and low-cost adsorption material. Due to the opposite chemical properties of Cd and As, the same adsorption material is difficult to effectively remove Cd and As [11]. At present, a variety of new adsorption materials have been developed, such as biochar [12], zeolite [13], porous silica sorbents [14] and metal oxides [15]. Among them, iron oxides and manganese oxides have been used to study the adsorption of heavy metals in water bodies, but iron and manganese oxides are easy to agglomerate in water bodies, reducing their adsorption capacity [16].
Mica is the general name of a class of typical water-rich potassium-rich phyllosilicate minerals, of which muscovite is the most common and has the highest potassium content [17]. Muscovite is widely used in electronic appliances, rubber, plastics and other industries. In the processing of muscovite, such as beneficiation, purification and peeling, a large amount of waste broken muscovite is produced. It is necessary to develop new utilization ways for muscovite and various unutilized potassium-containing silicate minerals.
Some studies have been carried out on the adsorption of heavy metals on muscovite. For example, Yang conducted studies on the adsorption of Cd(II), Cu(II), Pb(II), As(III) and As(V) by muscovite. It is found that the adsorption of Cd(II), Cu(II), Pb(II), As(III) and As(V) on muscovite is mainly the inner layer complexation [18]. The adsorption capacity of muscovite for heavy metals is limited. In order to improve the adsorption capacity of mica minerals for heavy metal ions, mica is often modified by high-temperature method and chemical methods. The expanded mica base sample can be prepared by high-temperature modification, and the cation exchange reaction efficiency can be significantly improved. Chemical modification is a simple and convenient method of adding inorganic salt or oxide and loading it on the surface of mica to increase its adsorption performance.
In recent years, many scholars have conducted extensive research on the treatment of heavy metal ion pollution by muscovite. The muscovite/zeolite aluminosilicate composite material prepared by Salam [19] can efficiently remove As(Ⅴ), Hg(Ⅱ) and U(Ⅵ) plasma from water, with the maximum adsorption capacity being 117, 122.5 and 138.5 mg/g, respectively. To solve the problem of Cd-As pollution in water, the development of modified muscovite with good adsorption capacity for both Cd and As can provide a new way for the resource utilization of muscovite.
Previous studies have found that iron and manganese oxides have good adsorption capacity for Cd and As, but their stability is poor. In this study, it is assumed that the muscovite modified by iron and manganese can reduce the agglomeration of iron and manganese oxides and has an adsorption capacity for cadmium and arsenic by using muscovite as the matrix. Therefore, the following work is carried out in this study: (1) preparation and characterization of muscovite supported iron–manganese oxide materials; and (2) exploration of the adsorption characteristics of materials and the adsorption mechanism.

2. Materials and Methods

2.1. Materials

Waste muscovite was taken from Lingshou, Hebei, China. KMnO4, FeSO4·7H2O and KOH are all analytically pure and purchased from Kemio. The experimental water is Ultrapure water.

2.2. Preparation of Adsorption Materials

Waste Muscovite was ground into powder and passed through a 200 mesh sieve. Muscovite (M) in this step was used to prepare muscovite-modified materials.
Weighed 10 g muscovite and added it into 100 mL KMnO4 solution with a concentration of 0.08 mol/L, and stirred thoroughly for 0.5 h. Then 75 mL of FeSO4·7H2O solution with a concentration of 0.32 mol/L was added, and MnO2 precipitation was generated in this step (chemical reactions (1) and (2)). After 1 h, 110 mL of KOH solution with a concentration of 0.5 mol/L was added, the mixture was stirred for 1h and placed at 70–80 °C for 12 h, and then the solid–liquid was separated and the solid was dried, that is, the muscovite supported iron manganese oxides material (FMM) was obtained.
Muscovite-supported iron oxides (FM) and muscovite-supported manganese oxides (MM) preparation methods can be found in Supplementary Materials.
MnO4 + 5Fe2+ + 8H+ = Mn2+ + 5Fe3+ + 4H2O
3Mn2+ + 2MnO4 + 2H2O = 5MnO2 ↓ + 4H+

2.3. Characterization Methods

XPS (X-ray photoelectron spectroscopy) (ESCALAB 250 XI, Thermo Fisher Scientific, Waltham, MA, USA) characterized the chemical composition of the surface elements for the materials. The specific surface area analyzer (ASAP 2046, Micromeritics, Norcross, GA, USA) was used to test the material’s specific surface area and pore diameter distribution characteristics. The surface morphology of the material was characterized by SEM (Verios G4, Thermo Fisher Scientific, USA). XRD (X-ray diffraction) (D8 QUEST, Bruker, Mannheim, Germany) characterized the crystal structure of the materials. The samples were dried and sieved through a 100 mesh sieve before characterization.

2.4. Adsorption Experiments

Screening adsorption materials: Cd(II) and As(V) were adsorbed with M, FM, MM and FMM. The concentration of adsorbent was 0.5 g/L and the initial pH value of Cd(II) and As(V) solutions was 6. After a 12 h reaction, the solution was filtered to be measured. The adsorption capacity of Cd(II) and As(V) by these four materials was compared.
Adsorption kinetics experiment: The concentration of FMM was 0.5 g/L. Samples were taken at intervals to test the concentration of Cd(II)/As(V) from the beginning of the adsorption reaction to 19 h (the initial pH value of Cd(II)/As(V) solution was 6).
Isothermal adsorption experiment: The concentration of FMM was 0.5 g/L, the initial concentration range of Cd(II)/ As(V) was 5–100 mg/L, and the initial pH value of the solution was 6. After a 12 h reaction, the solution was filtered and measured.
The effect of the initial pH value of the solution on the adsorption of Cd(II)/As(V) by FMM: The concentration of FMM is 0.5 g/L, the initial pH value of Cd(II) solution is 3–8, and the initial pH value of As(V) solution is 3–9. After a 12 h reaction, the solution was filtered and measured.
The influence of solution ion concentration on the adsorption of Cd(II)/As(V) by FMM: The concentration of FMM is 0.5 g/L, the concentration of Na+ ions in Cd(II)/As(V) solution is 0.03 mol/L–2 mol/L and the initial pH value of the solution is 6. After a 12 h reaction, the solution was filtered and measured.
The adsorption experiment was conducted in three parallel groups, oscillating at 200 r/min and 25 °C. After the shock, the concentration of Cd(II)/As(V) was measured by inductively coupled plasma mass spectrometry (ICP-MS, Agilent 7500cx, Santa Clara, CA, USA). The adsorption capacity of materials for Cd(II) and As(V) was calculated using Formula (S1) in Supplementary Materials.

3. Results and Discussion

3.1. Screening Experiments of Cd(II) and As(V) Adsorption Materials

The results of the adsorption amounts of Cd(II) and As(V) by the four materials are shown in Figure 1. The adsorption capacity of muscovite for Cd(II) was 8.70 mg/g, and the adsorption capacity of Cd(II) did not increase after loading iron oxides or manganese oxides on muscovite. The adsorption capacity of Cd(II) by muscovite-loaded iron and manganese oxides was 30.50 mg/g, which was 3.5 times the adsorption quantity of muscovite.
The adsorption of M for As(V) was 0.41 mg/g. The adsorption amount of FM for As(V) was 4.29 mg/g, which was 10 times the amount of M adsorption, indicating that the iron oxides increase the adsorption capacity of the material for the As(V) (p < 0.05). The adsorption of the MM to As(V) was 0.73 mg/g, with no significant difference from M. The amount of As(V) adsorbed by FMM was 17.79 mg/g, which was 43.39 times that of M adsorption capacity. The amount of adsorption of the FMM on As(V) was significantly improved, which was higher than the muscovite with manganese oxides or iron oxides.
The above results show that the adsorption amount of As(V) and Cd(II) was significantly increased after muscovite was loaded with iron and manganese oxides. In the following experiments, FMM was selected as the research object to study its adsorption characteristics and mechanism.

3.2. Adsorption Kinetics

Figure 2 shows the adsorption capacity of FMM for As(V) and Cd(II) with time. The adsorption rate of Cd(II) by FMM was very fast in the initial stage (0–120 min) and approached equilibrium after 120 min, indicating that adsorption mainly occurs on the outer surface of the material in the initial stage. With the increase in adsorption time, Cd2+ slowly diffuses into the pore size of the FMM, further reacting with internal active sites, and the reaction is very slow in this stage. The formula of pseudo-first-order kinetic model (S2) and pseudo-second-order kinetic model (S3) are provided in Supplementary Materials. Both the pseudo-first-order kinetic model (R2 = 0.912) and the pseudo-second-order kinetic model (R2 = 0.988) (Table 1) can effectively simulate the variation of Cd(II) adsorption by FMM with time.
The adsorption rate of As(V) by FMM was very fast in the initial stage (0–120 min) and reached equilibrium at 360 min. The pseudo-second-order kinetic model (R2 = 0.940) can better describe the variation of As(V) adsorption by FMM over time than the pseudo-first-order kinetic model (R2 = 0.857), indicating that the adsorption mechanism of FMM for As(V) is chemisorbed. This chemisorption involves Lewis acid–base coordination, the mechanism of which is to form ligand bonds between metal oxides and arsenic by sharing electrons, thereby allowing arsenic to bind to metal oxides to form stable and strong complexes [20].
The adsorption rate constant K2 of FMM for Cd(II) was 0.019, which was higher than the adsorption constant of 0.005 for As(V). This indicated that the adsorption sites of Cd(II) on the surface of FMM are more than those of As(V), resulting in a faster adsorption rate of Cd(II) on FMM than on As(V).

3.3. Adsorption Isotherms

The adsorption isotherm of FMM for As(V) and Cd(II) are shown in Figure 3. From low concentration to high concentration, the adsorption amount of FMM for As(V) and Cd(II) gradually increases until equilibrium. Using the Langmuir model and the Freundlich model to simulate the isothermal adsorption of As(V) and Cd(II), the formulas are provided in Supplementary Materials. The isothermal adsorption parameters are shown in Table 2.
As can be seen from Table 2, both the Langmuir model and the Freundlich model can fit the adsorption of FMM for As(V) and Cd(II) well. This indicated that the adsorption of As(V) and Cd(II) by FMM involved both single-layer and multi-layer adsorption, and the adsorption process was a multi-step process that may include external surface diffusion and particle internal diffusion. The maximum adsorption capacity of FMM fitted by the Langmuir model for Cd(II) was 32.47 mg/g, and the maximum adsorption capacity of FMM fitted by the Langmuir model for As(V) was 28.57 mg/g. Table 3 lists the maximum adsorption of some adsorbed materials on the Cd(II) and As(V). The adsorption amount of Cd(II) and As(V) on muscovite-loaded ferromanganese oxide (FMM) were 43.3 times and 36.1 times that on natural muscovite respectively. Compared with other materials in the table, FMM had higher or closer adsorption capacity than these materials and had better adsorption capacity for both Cd(II) and As(V).
The Langmuir model correlation coefficients for the adsorption of As(V) and Cd(II) by FMM were higher than those of Freundlich, indicating that the adsorption process of As(V) and Cd(II) by FMM was closer to that of monolayer chemisorption [21]. The parameters n fitted by the Freundlich model for FMM to As(V) and Cd(II) were 5.07 and 1.89 respectively, both between 1 and 10, indicating that the adsorption of FMM to As(V) and Cd(II) is easy to achieve [22].
Table 3. Comparison of the maximum adsorption capacity of As(V) and Cd(II) by FMM and other adsorption materials.
Table 3. Comparison of the maximum adsorption capacity of As(V) and Cd(II) by FMM and other adsorption materials.
Adsorption MaterialsQmax-Cd(II) (mg/g)Qmax-As(V) (mg/g)Ref.
Hydrated iron oxide gel beads 14.83[20]
Iron manganese oxide gel beads 26.65[19]
Fe3O4-reduced graphite oxide–MnO2 nanocomposites 12.22[23]
Magnetite/non-oxidative graphene 9.44[24]
Calcium-based magnetic biochar (Ca-MBC)10.07 [25]
Zeolite-supported nanoscale zero-valent iron48.63 [26]
Mn-doped iron oxide35 [27]
Mn3O4/Fe3O418.8 [28]
Manganese modified wood28.1 [29]
Natural muscovite0.7500.791[18]
FMM32.4728.57This study

3.4. The Effect of Initial pH Value of Solution on Adsorption

The forms of Cd and As in the solution will change with the change of the pH value of the solution, so the pH value of the solution has a great influence on the ability of the material to adsorb cadmium and arsenic [30]. In order to explore the appropriate pH range of FMM for removing Cd and As, experiments were carried out on the influence of initial pH values of different solutions on the removal capacity of FMM for As(V) and Cd(II). The experimental results are shown in Figure 4.
As the pH value of the solution increased, the adsorption capacity of Cd(II) by FMM gradually increased. The initial pH value of the solution was from 3 to 5, the adsorption capacity rose sharply, and from 5 to 7, the adsorption capacity rose slowly. At low pH values, the proton content in the solution was high, and the proton and Cd(II) in the solution competed for adsorption sites, reducing the adsorption capacity of the material to Cd(II). When the initial pH value of the solution is above 5, the adsorption capacity of FMM for Cd(II) is significantly improved, which may be the formation of hydrolyzed products such as Cd(OH)+ and precipitation such as Cd(OH)2 [31]. Under high pH conditions, FMM can remove Cd(II) through precipitation and electrostatic interaction [32].
The As(V) adsorption of FMM decreased with the increase in pH value, which was the opposite of Cd(II). When the solution pH value increased from 3 to 4, the adsorption capacity of FMM for As(V) decreased by 47.6%. When the solution pH value increased from 4 to 9, the adsorption capacity of FMM for arsenic slowly decreased, and the adsorption capacity decreased from 10.50 mg/g to 4.34 mg/g. As(V) exists in the solution in the form of arsenic ions (H3AsO4, H2AsO4, HAsO42−), and arsenic ions are negatively charged. When the pH value of the solution increases, there are more OH ions in the solution, which reduces the positive charge on the surface of the material and the negative charge adsorption point. Therefore, when the pH value of the solution is lower, the concentration of OH ions in the solution is lower, resulting in protonation on the surface of FMM, and weaker competition between OH- and As(V) adsorption sites [33]. The effect of pH on Cd(II) and As(V) adsorption results show that the adsorption mechanism of FMM for Cd(II) and As(V) includes electrostatic adsorption.

3.5. The Effect of Solution Ion Concentration on Adsorption

Ionic strength can affect the double-layer thickness of adsorbed material [34]. In order to explore the effect of ion strength on the adsorption capacity of FMM for As(V) and Cd(II), experiments were carried out to determine the effect of Na+ ion concentration in the range of 0.03 mol/L–2 mol/L on the adsorption of As(V) and Cd(II). The experimental results are shown in Figure 5. The adsorption capacity of Cd(II) by FMM decreased with the increase in Na+ concentration in the solution. When the Na+ concentration increased from 0.03 mol/L to 2 mol/L, the adsorption capacity of FMM for Cd(II) decreased by 16.04%, indicating that when the concentration of Na+ in the solution was high, it would compete with Cd(II) for the adsorption site and reduce the electrostatic adsorption effect, thus reducing the adsorption capacity of FMM for Cd(II) [35].
The adsorption quantity of FMM for As(V) increased with the increase in ion concentration, which increased from 0.03 mol/L to 2 mol/L, and the adsorption capacity of FMM for As(V) increased by 11.85%. High ionic strength can promote the adsorption of As(V) by materials, which indicates that the adsorption of As(V) by FMM may have covalent bond adsorption [34].

3.6. Surface Morphology

As can be seen from Figure 6, muscovite presents a layered structure with a smooth surface. After modification, the surface of muscovite was loaded with iron and manganese oxide particles, and the particles were uniformly dispersed on the surface of muscovite, and no agglomeration of iron and manganese oxides was observed, which indicated that muscovite prevents agglomeration of iron and manganese oxides, thereby increasing the specific surface area of the material and enhancing the stability of iron and manganese oxides. Figure 7 is the surface element distribution diagram of iron and manganese oxides supported by muscovite, with an iron content of 8.51% and a manganese content of 1.86%.

3.7. Structure Characteristics

The adsorption and desorption isotherms of M and FMM are shown in Figure 8. The N2 adsorption and desorption isotherm of FMM was convex upward in the low P/P0 region and belongs to type IV isotherm. The isotherm of M was convex towards the relative pressure axis and belongs to the type V isotherm. The adsorption curve of FMM sharply increases in the relative pressure P/P0 range of 0.0–0.1, indicating the existence of a microporous structure in the material.
The aperture distribution characteristics of M and FMM were similar (Figure 9). In the micropore region (pore size < 2 nm) and mesoporous region (2 nm–50 nm), the number of micropores and mesoporous pores of FMM was higher than that of M, and the volume of micropores and mesoporous pores of muscovite increased after loading Fe-Mn oxides. Corresponding to Table 4, the specific surface area of FMM was 150.82 m2/g, which is 8.3 times that of M. The average pore size of FMM was 3.60 nm, and the pore size of M was 13.20 nm. Both M and FMM belong to mesoporous materials. The above results show that the specific surface area of the material is increased, the pore size is reduced, and the total pore volume is increased after loading iron manganese oxides. The increase in specific surface area and total pore volume promotes the contact between the adsorbate and the adsorbent, improving the physical adsorption capacity of the adsorbent for Cd(II) and As(V). Therefore, the adsorption capacity of FMM for Cd(II) and As(V) was higher than that of M.

3.8. XPS Analysis

Figure 10 is a narrow spectrum of Fe2p. FMM peaks at 710.4 ev and 711.4 ev, corresponding to Fe2O3 and FeOOH, accounting for 37% and 63%, respectively. After the adsorption of Cd(II) by FMM, the proportion of FeOOH decreased to 51.79%, and after the adsorption of As(V), the proportion of FeOOH decreased to 49.48%. These results indicated that FeOOH in iron oxides is involved in the adsorption reaction of Cd(II) and As(V).
Figure 11 is a narrow spectrum of O1S. The characteristic peak position of the O1S for M was 531.5 ev and 532.5 ev, corresponding to Al-O [36] and Si-O [37]. The O1S peak of FMM was located at 529.7 ev, 531.5 ev and 532.5 ev, and the FMM loaded with iron and manganese oxides has a new peak at 529.7 ev, where the O corresponds to Fe(Ⅲ)-O/Mn(Ⅳ)-O [38]. The peak area ratio of O1S for FMM at the binding energy of 529.7 ev was 43.91%, which decreased to 35.73% and 27.94% after the adsorption of Cd(II) and As(V), respectively. These results indicate that Fe/Mn-O is involved in the adsorption of Cd(II) and As(V).
The morphological characteristics and adsorption mechanism of elements on the surface of materials were investigated by XPS. Figure 12 is the Mn2p narrow spectrum, the Mn2p peak location was 642.2 ev and 653.6 ev, which represents the MnO2 [39]. No peak of MnO2 was found in the XRD characterization (Supplementary Materials S1) of FMM, indicating that MnO2 exists on the surface of FMM in the amorphous state.
After the adsorption of As(V) by FMM, the As3d absorption peak was at 44.52 ev binding energy, which corresponds to HAsO42−. The result indicated that monoprotonated bidentate complexes are the dominant species on the surface of FMM. The same study also found that regardless of whether in acidic or alkaline environments, arsenic mainly exists on the surface of the material in the form of HAsO42− after being adsorbed by the adsorbent for As(V) [40].
After the adsorption of Cd(II) by FMM, Cd3d had a characteristic peak at the binding energy of 405.0 ev, indicating that Cd(II) exists on the surface of FMM mainly in the form of Cd(OH)2 and Cd-O. The above indicated that FMM adsorbs Cd(II) through precipitation and surface complexation mechanisms.

4. Conclusions

In this study, the iron–manganese oxides modification of muscovite significantly improved Cd(II) and As(V) adsorption with maximum capacities of 32.47 mg/g and 28.57 mg/g. Compared with the muscovite, FMM has a larger specific surface area and pore volume. Furthermore, electrostatic adsorption, covalent bond adsorption and chemical adsorption were involved in As(V) removal, while electrostatic adsorption and surface precipitation were responsible for Cd(II) removal. The above results provide a novel strategy for the resource utilization of muscovite as an adsorbent for heavy metals and arsenic.
The above research was conducted in the laboratory. In future research, pilot-scale testing will be conducted on FMM treatment of heavy metal wastewater, and its environmental footprint, scalability, and cost-effectiveness assessment will be conducted to provide more comprehensive parameters for its on-field applications.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/w15203579/s1. Supplementary Materials, Figure S1: The XRD pattern of M, FMM, FMM after Cd(II) adsorption(FMM-Cd), FMM after As(V) adsorption(FMM-As).

Author Contributions

All authors contributed to this study. The experiment was designed by J.G., Y.W., Y.Z., Z.X. and S.F. conducted experiments. The experimental data were analyzed by R.W. and H.Z., J.G. and Y.W. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study received funding from The Science and Technology Development Project of Yellow River Institute of Hydraulic Research, grant number HKF202206”, “Youth Science Fund project of Henan Province, grant number 232300421404” and “Special Project of Basic Scientific Research Business Expenses of Yellow River Institute of Hydraulic Research, grant number HKYJBYW-2021-03”.

Data Availability Statement

Data are available.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Neelgund, G.M.; Jimenez, E.A.; Ray, R.L.; Kurkuri, M.D. Facilitated Adsorption of Mercury(II) and Chromium(VI) Ions over Functionalized Carbon Nanotubes. Toxics 2023, 11, 18. [Google Scholar] [CrossRef] [PubMed]
  2. Sajjan, V.A.; Aralekallu, S.; Nemakal, M.; Palanna, M.; Prabhu, C.P.K.; Sannegowda, L.K. Nanomolar detection of lead using electrochemical methods based on a novel phthalocyanine. Inorg. Chim. Acta 2020, 506, 7. [Google Scholar] [CrossRef]
  3. Bastami, K.D.; Bagheri, H.; Kheirabadi, V.; Zaferani, G.G.; Teymori, M.B.; Hamzehpoor, A.; Soltani, F.; Haghparast, S.; Harami, S.R.M.; Ghorghani, N.F.; et al. Distribution and ecological risk assessment of heavy metals in surface sediments along southeast coast of the Caspian Sea. Mar. Pollut. Bull. 2014, 81, 262–267. [Google Scholar] [CrossRef] [PubMed]
  4. Mlangeni, A.T. Health risk assessment of toxic metal(loids) (As, Cd, Pb, Cr, and Co) via consumption of medicinal herbs marketed in Malawi. Toxicol. Rep. 2023, 11, 145–152. [Google Scholar] [CrossRef] [PubMed]
  5. Jia, F.; Li, J.F.; Wang, J.L. Recovery of boric acid from the simulated radioactive wastewater by vacuum membrane distillation crystallization. Ann. Nucl. Energy 2017, 110, 1148–1155. [Google Scholar] [CrossRef]
  6. Chen, Q.Y.; Yao, Y.; Li, X.Y.; Lu, J.; Zhou, J.; Huang, Z.L. Comparison of heavy metal removals from aqueous solutions by chemical precipitation and characteristics of precipitates. J. Water Process. Eng. 2018, 26, 289–300. [Google Scholar] [CrossRef]
  7. Bahsaine, K.; Mekhzoum, M.E.; Benzeid, H.; Qaiss, A.; Bouhfid, R. Recent progress in heavy metals extraction from phosphoric acid: A short review. J. Ind. Eng. Chem. 2022, 115, 120–134. [Google Scholar] [CrossRef]
  8. Liu, X.J.; Wu, J.L.; Wang, J.L. Electro-adsorption of Sr(II) from aqueous solution by activated carbon cloth/nickel hexacyanoferrate composite electrode through capacitive deionization. J. Clean. Prod. 2022, 380, 10. [Google Scholar] [CrossRef]
  9. Badescu, I.S.; Bulgariu, D.; Ahmad, I.; Bulgariu, L. Valorisation possibilities of exhausted biosorbents loaded with metal ions—A review. J. Environ. Manag. 2018, 224, 288–297. [Google Scholar] [CrossRef]
  10. Wang, K.; Zhang, F.B.; Xu, K.X.; Che, Y.J.; Qi, M.Y.; Song, C. Modified magnetic chitosan materials for heavy metal adsorption: A review. RSC Adv. 2023, 13, 6713–6736. [Google Scholar] [CrossRef]
  11. Yang, D.; Yang, S.Y.; Wang, L.; Xu, J.M.; Liu, X.M. Performance of biochar-supported nanoscale zero-valent iron for cadmium and arsenic co-contaminated soil remediation: Insights on availability, bioaccumulation and health risk. Environ. Pollut. 2021, 290, 9. [Google Scholar] [CrossRef]
  12. Sugawara, K.; Ichio, K.; Ichikawa, Y.; Ogawa, H.; Suzuki, S. Effects of Pyrolysis Temperature and Chemical Modification on the Adsorption of Cd and As(V) by Biochar Derived from Pteris vittata. Int. J. Environ. Res. Public Health 2022, 19, 16. [Google Scholar] [CrossRef]
  13. Soni, R.; Shukla, D.P. Synthesis of fly ash based zeolite-reduced graphene oxide composite and its evaluation as an adsorbent for arsenic removal. Chemosphere 2019, 219, 504–509. [Google Scholar] [CrossRef]
  14. Li, S.; Li, S.; Wen, N.; Wei, D.; Zhang, Y. Highly effective removal of lead and cadmium ions from wastewater by bifunctional magnetic mesoporous silica. Sep. Purif. Technol. 2021, 265, 118341. [Google Scholar] [CrossRef]
  15. Michalkova, Z.; Komarek, M.; Veselska, V.; Cihalova, S. Selected Fe and Mn (nano)oxides as perspective amendments for the stabilization of As in contaminated soils. Environ. Sci. Pollut. Res. 2016, 23, 10841–10854. [Google Scholar] [CrossRef]
  16. Zhou, Q.W.; Liao, B.H.; Lin, L.N.; Qiu, W.W.; Song, Z.G. Adsorption of Cu(II) and Cd(II) from aqueous solutions by ferromanganese binary oxide-biochar composites. Sci. Total Environ. 2018, 615, 115–122. [Google Scholar] [CrossRef]
  17. Chakraborty, S.; Wolthers, M.; Chatterjee, D.; Charlet, L. Adsorption of arsenite and arsenate onto muscovite and biotite mica. J. Colloid Interface Sci. 2007, 309, 392–401. [Google Scholar] [CrossRef]
  18. Yang, J.S.; Lee, J.Y.; Park, Y.T.; Baek, K.; Choi, J. Adsorption of As(III), As(V), Cd(II), Cu(II), and Pb(II) from Aqueous Solutions by Natural Muscovite. Sep. Sci. Technol. 2010, 45, 814–823. [Google Scholar] [CrossRef]
  19. Salam, M.A.; Abukhadra, M.R.; Mostafa, M. Effective decontamination of As(V), Hg(II), and U(VI) toxic ions from water using novel muscovite/zeolite aluminosilicate composite: Adsorption behavior and mechanism. Environ. Sci. Pollut. Res. 2020, 27, 13247–13260. [Google Scholar] [CrossRef]
  20. Liu, Y.Q.; Cai, L.K.; Wang, X.Y.; Chen, Z.H.; Yang, W.Z. Efficient adsorption of arsenic in groundwater by hydrated iron oxide and ferromanganese oxide chitosan gel beads. Sep. Purif. Technol. 2023, 315, 12. [Google Scholar] [CrossRef]
  21. Lilhare, S.; Mathew, S.B.; Singh, A.K.; Carabineiro, S.A.C. Calcium Alginate Beads with Entrapped Iron Oxide Magnetic Nanoparticles Functionalized with Methionine-A Versatile Adsorbent for Arsenic Removal. Nanomaterials 2021, 11, 21. [Google Scholar] [CrossRef]
  22. Gupta, K.; Maity, A.; Ghosh, U.C. Manganese associated nanoparticles agglomerate of iron(III) oxide: Synthesis, characterization and arsenic(III) sorption behavior with mechanism. J. Hazard. Mater. 2010, 184, 832–842. [Google Scholar] [CrossRef]
  23. Luo, X.B.; Wang, C.C.; Luo, S.L.; Dong, R.Z.; Tu, X.M.; Zeng, G.S. Adsorption of As (III) and As(V) from water using magnetite Fe3O4-reduced graphite oxide-MnO2 nanocomposites. Chem. Eng. J. 2012, 187, 45–52. [Google Scholar] [CrossRef]
  24. Yoon, Y.J.; Zheng, M.L.; Ahn, Y.T.; Park, W.K.; Yang, W.S.; Kang, J.W. Synthesis of magnetite/non-oxidative graphene composites and their application for arsenic removal. Sep. Purif. Technol. 2017, 178, 40–48. [Google Scholar] [CrossRef]
  25. Wu, J.Z.; Huang, D.; Liu, X.M.; Meng, J.; Tang, C.X.; Xu, J.M. Remediation of As(III) and Cd(II) co-contamination and its mechanism in aqueous systems by a novel calcium-based magnetic biochar. J. Hazard. Mater. 2018, 348, 10–19. [Google Scholar] [CrossRef]
  26. Li, Z.T.; Wang, L.; Meng, J.; Liu, X.M.; Xu, J.M.; Wang, F.; Brookes, P. Zeolite-supported nanoscale zero-valent iron: New findings on simultaneous adsorption of Cd(II), Pb(II), and As(III) in aqueous solution and soil. J. Hazard. Mater. 2018, 344, 1–11. [Google Scholar] [CrossRef]
  27. Warner, C.L.; Chouyyok, W.; Mackie, K.E.; Neiner, D.; Saraf, L.V.; Droubay, T.C.; Warner, M.G.; Addleman, R.S. Manganese Doping of Magnetic Iron Oxide Nanoparticles: Tailoring Surface Reactivity for a Regenerable Heavy Metal Sorbent. Langmuir 2012, 28, 3931–3937. [Google Scholar] [CrossRef]
  28. Gong, J.L.; Chen, L.; Zeng, G.M.; Long, F.; Deng, J.H.; Niu, Q.Y.; He, X. Shellac-coated iron oxide nanoparticles for removal of cadmium(II) ions from aqueous solution. J. Environ. Sci. 2012, 24, 1165–1173. [Google Scholar] [CrossRef]
  29. Wang, H.Y.; Gao, B.; Wang, S.S.; Fang, J.; Xue, Y.W.; Yang, K. Removal of Pb(II), Cu(II), and Cd(II) from aqueous solutions by biochar derived from KMnO4 treated hickory wood. Bioresour. Technol. 2015, 197, 356–362. [Google Scholar] [CrossRef]
  30. Raju, N.J. Arsenic in the geo-environment: A review of sources, geochemical processes, toxicity and removal technologies. Environ. Res. 2022, 203, 15. [Google Scholar] [CrossRef]
  31. Liang, J.; Li, X.M.; Yu, Z.G.; Zeng, G.M.; Luo, Y.; Jiang, L.B.; Yang, Z.X.; Qian, Y.Y.; Wu, H.P. Amorphous MnO2 Modified Biochar Derived from Aerobically Composted Swine Manure for Adsorption of Pb(II) and Cd(II). ACS Sustain. Chem. Eng. 2017, 5, 5049–5058. [Google Scholar] [CrossRef]
  32. Su, Y.M.; Adeleye, A.S.; Huang, Y.X.; Sun, X.Y.; Dai, C.M.; Zhou, X.F.; Zhang, Y.L.; Keller, A.A. Simultaneous removal of cadmium and nitrate in aqueous media by nanoscale zerovalent iron (nZVI) and Au doped nZVI particles. Water Res. 2014, 63, 102–111. [Google Scholar] [CrossRef]
  33. Ben Issa, N.; Rajakovic-Ognjanovic, V.N.; Jovanovic, B.M.; Rajakovic, L.V. Determination of inorganic arsenic species in natural waters-Benefits of separation and preconcentration on ion exchange and hybrid resins. Anal. Chim. Acta 2010, 673, 185–193. [Google Scholar] [CrossRef]
  34. Chen, Q.Q.; Yin, D.Q.; Zhu, S.J.; Hu, X.L. Adsorption of cadmium(II) on humic acid coated titanium dioxide. J. Colloid Interface Sci. 2012, 367, 241–248. [Google Scholar] [CrossRef]
  35. Park, C.M.; Han, J.H.; Chu, K.H.; Al-Hamadani, Y.A.J.; Her, N.; Heo, J.Y.; Yoon, Y.M. Influence of solution pH, ionic strength, and humic acid on cadmium adsorption onto activated biochar: Experiment and modeling. J. Ind. Eng. Chem. 2017, 48, 186–193. [Google Scholar] [CrossRef]
  36. Taylor, J.A. An XPS study of the oxidation of ALAs thin-films grown by MBE. J. Vac. Sci. Technol. 1982, 20, 751–755. [Google Scholar] [CrossRef]
  37. Dang, T.A.; Chau, C.N. Electron spectroscopy for chemical analysis of cool white phosphors coated with SiO2 thin film. J. Electrochem. Soc. 1996, 143, 302–305. [Google Scholar] [CrossRef]
  38. Chastain, J.; King, R.C., Jr. Handbook of X-ray Photoelectron Spectroscopy; Perkin-Elmer Corporation: Waltham, MA, USA, 1992; Volume 40, p. 221. [Google Scholar]
  39. Tan, B.J.; Klabunde, K.J.; Sherwood, P.M.A. XPS studies of solvated metal atom dispersed catalysts—Evidence for layered cobalt manganese particles on alumina and silica. J. Am. Chem. Soc. 1991, 113, 855–861. [Google Scholar] [CrossRef]
  40. Zhang, S.J.; Li, X.Y.; Chen, J.P. An XPS study for mechanisms of arsenate adsorption onto a magnetite-doped activated carbon fiber. J. Colloid Interface Sci. 2010, 343, 232–238. [Google Scholar] [CrossRef]
Figure 1. Adsorption effects of different adsorption materials on Cd(II) and As(V). (Different letters are significantly different at p < 0.05 according to SNK test.)
Figure 1. Adsorption effects of different adsorption materials on Cd(II) and As(V). (Different letters are significantly different at p < 0.05 according to SNK test.)
Water 15 03579 g001
Figure 2. The adsorption capacity of FMM for Cd(II) and As(V) with time.
Figure 2. The adsorption capacity of FMM for Cd(II) and As(V) with time.
Water 15 03579 g002
Figure 3. The adsorption isotherm of FMW for As(V) and Cd(II).
Figure 3. The adsorption isotherm of FMW for As(V) and Cd(II).
Water 15 03579 g003
Figure 4. The effect of initial pH value of solution on the adsorption of As(V) and Cd(II) by FMM.
Figure 4. The effect of initial pH value of solution on the adsorption of As(V) and Cd(II) by FMM.
Water 15 03579 g004
Figure 5. The effect of ion concentration of solution on the adsorption of As(V) and Cd(II) by FMM.
Figure 5. The effect of ion concentration of solution on the adsorption of As(V) and Cd(II) by FMM.
Water 15 03579 g005
Figure 6. Scanning electron microscope image of muscovite (a) and muscovite-loaded iron and manganese oxides (b).
Figure 6. Scanning electron microscope image of muscovite (a) and muscovite-loaded iron and manganese oxides (b).
Water 15 03579 g006
Figure 7. EDX diagram of muscovite-loaded iron and manganese oxides.
Figure 7. EDX diagram of muscovite-loaded iron and manganese oxides.
Water 15 03579 g007
Figure 8. BET nitrogen adsorption–desorption isotherms of M and FMM.
Figure 8. BET nitrogen adsorption–desorption isotherms of M and FMM.
Water 15 03579 g008
Figure 9. Pore size distribution of M and FMM.
Figure 9. Pore size distribution of M and FMM.
Water 15 03579 g009
Figure 10. XPS spectra of Fe2p for FMM (a), FMM adsorbed Cd(II) (b), FMM adsorbed As(V) (c).
Figure 10. XPS spectra of Fe2p for FMM (a), FMM adsorbed Cd(II) (b), FMM adsorbed As(V) (c).
Water 15 03579 g010
Figure 11. XPS spectra of O1s for muscovite (a), FMM (b), FMM adsorbed Cd(II) (c), FMM adsorbed As(V) (d).
Figure 11. XPS spectra of O1s for muscovite (a), FMM (b), FMM adsorbed Cd(II) (c), FMM adsorbed As(V) (d).
Water 15 03579 g011
Figure 12. XPS spectra of Mn2p for FMM (a), Cd3d for FMM adsorbed Cd(II) (b) and As3d for FMM adsorbed As(V) (c).
Figure 12. XPS spectra of Mn2p for FMM (a), Cd3d for FMM adsorbed Cd(II) (b) and As3d for FMM adsorbed As(V) (c).
Water 15 03579 g012
Table 1. Adsorption kinetics constants for Cd(II) and As(V) adsorption onto FMM.
Table 1. Adsorption kinetics constants for Cd(II) and As(V) adsorption onto FMM.
Pseudo-First-Order
Model
Pseudo-Second-Order
Model
Exp.
Qe
(mg/g)
k1
(1/min)
R2Qe
(mg/g)
k2
(g/(mg min))
R2Qe
(mg/g)
Cd(II)19.090.0900.91219.430.0190.98819.29
As(V)6.980.0260.8577.390.0050.9407.37
Table 2. Isothermal adsorption constant of FMM.
Table 2. Isothermal adsorption constant of FMM.
LangmuirFreundlich
Qe (mg/g)KL (L/mg)R2nKF (mg/g)R2
Cd32.470.1910.9965.0713.110.997
As28.570.0460.9781.892.550.966
Table 4. Pore structure parameters of M and FMM.
Table 4. Pore structure parameters of M and FMM.
Surface Area of BET(m2/g)Average Pore Size (nm)Vtotal (m3/g)
M18.1213.200.060
FMM150.823.600.136
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, Y.; Zhao, Y.; Xu, Z.; Wang, R.; Zhang, H.; Feng, S.; Guo, J. Efficient Removal of Cadmium (II) and Arsenic (V) from Water by the Composite of Iron Manganese Oxides Loaded Muscovite. Water 2023, 15, 3579. https://0-doi-org.brum.beds.ac.uk/10.3390/w15203579

AMA Style

Wu Y, Zhao Y, Xu Z, Wang R, Zhang H, Feng S, Guo J. Efficient Removal of Cadmium (II) and Arsenic (V) from Water by the Composite of Iron Manganese Oxides Loaded Muscovite. Water. 2023; 15(20):3579. https://0-doi-org.brum.beds.ac.uk/10.3390/w15203579

Chicago/Turabian Style

Wu, Yan, Yue Zhao, Zhuben Xu, Rui Wang, Han Zhang, Shuaitao Feng, and Jianhua Guo. 2023. "Efficient Removal of Cadmium (II) and Arsenic (V) from Water by the Composite of Iron Manganese Oxides Loaded Muscovite" Water 15, no. 20: 3579. https://0-doi-org.brum.beds.ac.uk/10.3390/w15203579

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