Next Article in Journal
Motor Development Comparison between Preterm and Full-Term Infants Using Alberta Infant Motor Scale
Next Article in Special Issue
Influence of Zinc and Humic Acids on Dye Adsorption from Water by Two Composts
Previous Article in Journal
Health Care Organization in Poland in Light of the Refugee Crisis Related to the Military Conflict in Ukraine
Previous Article in Special Issue
Efficiency Recycling and Utilization of Phosphate from Wastewater Using LDHs-Modified Biochar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Thallium Removal from Water Environment by Metal Oxide Material

1
College of Environmental Science and Engineering, Hunan University, Changsha 410082, China
2
Key Laboratory of Environmental Biology and Pollution Control, Ministry of Education, Hunan University, Changsha 410082, China
3
Aerospace Kaitian Environmental Technology Co., Ltd., Changsha 410100, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Environ. Res. Public Health 2023, 20(5), 3829; https://doi.org/10.3390/ijerph20053829
Submission received: 28 December 2022 / Revised: 16 February 2023 / Accepted: 20 February 2023 / Published: 21 February 2023

Abstract

:
Thallium is widely used in industrial and agricultural development. However, there is still a lack of systematic understanding of its environmental hazards and related treatment methods or technologies. Here, we critically assess the environmental behavior of thallium in aqueous systems. In addition, we first discuss the benefits and limitations of the synthetic methods of metal oxide materials that may affect the practicality and scalability of TI removal from water. We then assess the feasibility of different metal oxide materials for TI removal from water by estimating the material properties and contaminant removal mechanisms of four metal oxides (Mn, Fe, Al, and Ti). Next, we discuss the environmental factors that may inhibit the practicality and scalability of Tl removal from water. We conclude by highlighting the materials and processes that could serve as more sustainable alternatives to TI removal with further research and development.

1. Introduction

Thallium is a highly toxic heavy metal that can cause chronic poisoning and is widely dispersed in very low levels in the environment [1]. Thallium was originally used in medicine. It can be used to treat diseases such as ringworm of the head. It was later found to be highly toxic and used mainly in agriculture as a rodenticide, insecticide, and anti-mold agent It is currently used in large numbers, mainly in the industry [2,3]. Thallium enters the water environment mainly through the exploitation of mines containing thallium, dust deposition in ore smelting, waste discharge from sulfur mineral areas, and waste discharge from the printing and dyeing industry. It can also pollute the environment to varying degrees through the combustion of coal and sulfurous iron ore ash [4,5,6].
Studies have shown that thallium contamination has posed a serious threat to drinking water safety, and the scope of contamination has spread to countries around the world [7,8,9]. In Canada, New Brunswick, the United Kingdom, Cornwall, Idaho, China, Qianxinan, Guangdong, and Xiangnan, as well as other waters near the mine and downstream water bodies, the monitoring value of drinking water thallium exceeded 10.2–10.4 times. Even in some sections of the river, such as the Canon and the Red River in England, the single-factor pollution index for thallium has reached 14–15 times [10]. Trace amounts of thallium have also been detected in areas such as Poland [11]. The environmental ecology of these areas are also affected by thallium contamination, with levels of thallium in water bodies and plants well above background values. Thallium, along with other elements, such as arsenic, cadmium, nickel, mercury, or lead, are very harmful to mammals. However, in some specific cases, thallium is more toxic than the other elements [12]. The relocation and transport of thallium in the environment will eventually accumulate in the food chain and thus enter the human body, posing a significant threat to human health [13]. Figure 1 illustrates the impact of thallium on the environment and humans.
Thallium-containing wastewater with excessive concentrations can seriously threaten human health and environmental water quality, while traditional treatment technologies have shown significant drawbacks (e.g., poor selectivity and interference from impurity ions) [14]. Using new treatment technology to make up for the defects of traditional wastewater treatment methods has become one of the current popular research topics. In addition, there is still a lack of systematic understanding of the advantages and disadvantages of different Tl removal methods.
The current methods for thallium removal include chemical precipitation, ion exchange, solvent extraction, and adsorption [15,16,17]. The chemical precipitation method mainly removes thallium from the water by adding Cl, S2−, potassium iron vanadium, and Prussian blue into the thallium-containing wastewater [18,19]. However, it is usually difficult for its deep treatment capability to meet the stringent thallium control standards, and there is a risk of secondary contamination. Although the ion exchange was recommended as one of the optimal treatment methods for thallium removal, the ion exchange is highly susceptible to the influence of other co-existing alkaline earth metals in thallium-containing waters. Furthermore, it is not highly selective for thallium and has the disadvantage of frequent regeneration. The solvent extraction method is usually used to recover Tl from water with high Tl concentrations, but it is difficult to apply to purifying conventional thallium-contaminated water bodies [20,21]. The adsorption method has the advantages of large adsorption capacity and high treatment depth, and it is currently widely used. Adsorption methods can be divided into physical and chemical adsorption. Physical adsorption depends on the material’s porous structure, while chemical adsorption is mainly determined by the chemical properties and state of the material surface [22]. The adsorbents currently available for thallium removal are biochar, activated carbon, metal oxides, and other materials. However, biochar [23] and activated carbon [24] showed poor performance for thallium adsorption. Recently, metal oxides have aroused wide attention due to their variable size and shape, abundant availability, sufficient surface active sites, and economic feasibility [3]. They can be used as ideal substitutes for carbon materials to remove heavy metals, and the most widely studied metal oxides include the four metal oxides of manganese, iron, aluminum, and titanium [25]. However, there are rarely studies concentrated on the feasibility and removal mechanism of metal oxide materials for TI removal from water.
Thus, this paper aims to review the progress of research on the removal of thallium from wastewater by metal oxides and to assess the feasibility of different metal oxide materials for TI removal from water by estimating the material properties and contaminant removal mechanisms of four metal oxides (Mn, Fe, Al, and Ti). Then, different environmental factors that may inhibit the practicality and scalability of Tl removal from the water will be discussed. It is hoped that this review could serve as a more sustainable alternative to TI removal with further research and development. It is also hoped that this review will provide new opportunities and possibilities for the removal and transformation of thallium to minimize the damage caused by thallium to humans and the environment.

2. Environmental Behavior of Thallium in Aqueous Systems

In natural water bodies, thallium is generally divided into two states: dissolved and particulate. Thallium exists mostly in its dissolved state in two forms: monovalent and trivalent thallium [26]. Tl(I) is highly soluble and mobile. In contrast, the reduction in Tl(III) is extremely fast and its solubility in water is related to the kinetics of the reaction, which is very rapid [27]. The hydrolysis of thallium is an important acid-base reaction for the generation of hydroxy complexes in the aqueous environment. It is therefore an essential factor in the de-termination of the chemical species, biological effectiveness, and toxicity of Tl in the environment [26]. In the aqueous environment, the hydrolysis of Tl(III) can be described by the following equation.
Tl ( OH ) n 1 4 n + H 2 O Tl ( OH ) n 3 n + H +
Solution pH and the hydrolysis constant are the main factors that determine the type of Tl in aqueous solutions. Tl(I) is soluble in water at the pH range of 3–12 to form Tl+, Tl(OH)(aq) [28]. In the available studies, the solubility of Tl is 350 mg/L, indicating that it cannot be precipitated under normal environmental conditions. Lin and Nriagu found that at pH < 7, Tl(III) it was converted to Tl(OH)2+, to Tl(OH)3 at pH in the range 7.4–8.8, and to Tl(OH)4 at pH 8.8 [28]. Tl can form complexes with halides (F, Cl, Br, I), sulphate ions (SO42−), and humic acids, thus affecting the removal of Tl from the wastewater [29,30]. Tl(III)-complexes bind more strongly than Tl(I)-complexes, but Tl(I)-halides have a higher lattice energy than Tl(III)-halides. The chloride and bromide ions can inhibit the hydrolysis of Tl(III), which can form [TlCl(OH)]+ and [TlBr(OH)]+ at low halide concentrations [3]. In addition, the solubilization–precipitation behavior of Tl(III) is modified by the presence of the ligand, which becomes less oxidized in the fact of Cl (concentration of 250 mg/L).
Due to its ionic radius and electronegativity constant, Tl(I) exhibits properties similar to potassium and, to a lesser extent, to rubidium and silver in aquatic ecosystems [21]. As with alkali metal compounds, TlOH and Tl2O are readily soluble in water, but Tl+ can form strong complexes. On the other hand, monovalent thallium is similar to silver in the formation of insoluble sulfides (solubility product of 10−21.2 for Tl2S) and small amounts of soluble photosensitized halide complexes. Due to the large difference in the electronegativity of K+ and Tl+, monovalent thallium readily replaces potassium in clays and secondary silica. Tl(III) is mainly bound to Fe/Mn hydroxides, and thallium is adsorbed by the accumulation of mineral oxides, which are then oxidized on the mineral surface. Tl-partitioning in different sediment and suspended particle phases may provide clues to this element’s redox form in the water column.

3. The Synthesis and Environmental Applications of Metal Oxides

Metal oxides are binary compounds composed of oxygen and other metallic chemical elements. Metal oxides are widely used for the removal of pollutants from aqueous environmental systems because of their large specific surface area and high activity.

3.1. Synthesis Methods of Metal Oxides

Different methods of synthesis can lead to the different physicochemical properties of these metal oxides in different forms (e.g., manganese dioxide particles, titanium carbonate nanotubes, etc.), and differences in the specific surface area, morphology, lattice structure, size, and the zero charge point of the metal oxides can affect the mechanism of their removal of thallium [31]. The synthesis method can, therefore, impact the mechanism and performance of metal oxide removal. In past research into the removal of thallium by metal oxides, the methods of synthesis of metal oxide materials can be divided into two main types: physical and chemical methods.

3.1.1. Physical Synthesis

Physical synthesis for metal oxides mainly includes high-energy ball-milling, inert gas condensation, ultrasonic blasting, severe plastic deformation, etc. The inert gas condensation method controls the particle size of the material, mainly by varying the gas pressure. Xiang et al. prepared zinc nanoparticles with different structures using an inert condensing gas method [32]. High-energy ball-milling changes the size and shape of the material by continuously grinding the material until plastic deformation occurs. Ni et al. improved the capacity and efficiency of LiMnPO4 by high-energy ball-milling. Ultrasonic blasting is a technique that improves the overall mechanical properties of the surface of metallic materials, resulting in a simultaneous change in the shape and dimensions of the material [33]. These methods all affect the performance of metal oxides in removing heavy metals by changing the physical characteristics of the material. For instance, the adsorption site increases as the specific surface area increases and changes in morphology result in the formation of different cavities, which facilitates the formation of surface complexes and adsorption. However, so far, there have been few studies on thallium adsorption by the physical synthesis of prepared metal oxides. Subsequent studies could focus more on thallium adsorption by the physical synthesis of metal oxides.

3.1.2. Chemical Synthesis

The chemical synthesis of metal oxides for thallium removal has been relatively well studied compared to the physical synthesis. Chemical methods include controlled chemical co-precipitation, ant gelatinization (or microemulsion), chemical vapor condensation, liquid flame spraying, pulsed electrode positions, gas-phase reduction, and liquid-phase reduction [15]. Among these synthetic methods, techniques such as co-precipitation, thermal decomposition (zero-valent manganese) [34], and hydrothermal synthesis are widely used and studied due to their simplicity of operation and high yields. Li et al. used co-precipitation to synthesize a layered composite material with amorphous flower-like MnO2 as the outer coating and magnetite as the inner core. The mechanism of thallium removal by this material mainly occurs through the hydroxylation of the manganese dioxide surface and the formation of Mn-O surface complexes in the MnO6 octahedra, which are removed as Tl2O3 and Tl2O precipitates. In this process, Tl(I) is oxidized to Tl(III) and thus adsorbed by Mn(VI) [35]. Titanium carbonate nanotubes prepared by hydrothermal synthesis have a large specific surface area. The material removes thallium mainly by: (1) Tl+ exchange with Na+ ions; (2) Exchange of excess Tl+ with H+; and (3) Tl+ being oxidized to Tl3+ and co-precipitating with TNTs [36]. The mechanism of thallium removal by chemically synthesized metal oxides mainly involves: (i) Oxidation of Tl+ to Tl3+, thereby co-precipitating with the metal oxide; (ii) Ion exchange with ions on the metal oxide surface; (iii) Surface complexation of Tl with hydroxyl groups on the metal oxide surface; (iiii) Surface electrostatic attraction of negatively charged metal oxides to Tl+.

3.2. Removal of Thallium from Water/Wastewater by Metal Oxides

Metal oxides as adsorbents for thallium removal have the advantages of high adsorption capacity, high selectivity, and easy regeneration. The pH, co-existing ions, and organic matter affect the sorption performance and removal mechanism of thallium by different metal oxides. The studies on thallium removal by four metal oxides (manganese, aluminum, iron, and titanium) are summarized in the following section. The advantages, disadvantages, and feasibility of thallium removal by different metal oxides are analyzed to obtain the feasible method for thallium removal in real wastewater.

3.2.1. Thallium Removal by Manganese Oxides

Manganese oxides are natural oxidants with good cation exchange properties and have good adsorption and oxidation properties for Tl(I). In aqueous solutions, Mn(IV) can adsorb Tl(I) by surface complexation and oxidative precipitation. At the same time, the higher valence metal in the manganese oxide can oxidize Tl(I) to form Tl2O3 precipitates, and then Tl2O3 can be absorbed on the surface of the material or in the pores [37]. The current research on thallium adsorption by manganese oxides has been carried out by four types of synthetic manganese oxides, mineral manganese oxides, modified manganese oxides, and manganese oxide composites. Reasonable modifications can improve the ability of the manganese oxides to adsorb Tl, and the selectivity and reusability of the adsorbent can be well improved. However, the lattice structure of the manganese oxides also affects the removal results. Liu et al. investigated the difference in the adsorption effect of five crystal structures of MnO2 (δ-, γ-, α-, β- and λ-MnO2) on Tl(I) [38,39]. The lattice structures of the four MnO2 species are shown in Figure 2. Freundlich’s isothermal adsorption curves show that δ-MnO2 has the maximum adsorption capacity for Tl(I) at a pH of 12 [40]. This is mainly due to the fact that the lamellar structure favors δ-MnO2 to carry more ions and there are different numbers of octahedral cation vacancies within its layers. These cation vacancies are considered to be strong adsorption sites for heavy metal [41]. Compared to conventional adsorbents, nanoscale adsorbents have a larger specific surface area, higher porosity, and a hollow structure, resulting in high-adsorption performance for pollutants, with the highest adsorption capacity of 672 mg/g for thallium [3]. However, nMnO2 is prone to coalescence in water and its separability is poor. Furthermore, there is a lack of suitable methods to remove Tl from nMnO2. The reproducibility of nMnO2 is an issue that can be further investigated. Wan et al. prepared amorphous hydrated MnO2 (HMO) using oxidation and chemical precipitation. Experimental results indicated that the adsorption capacity of HMO for Tl reached 352 mg/g. The mechanism of thallium adsorption by this material is mainly through the ion exchange of Mn-OH and Mn-O bonds, forming an inner sphere complex [42]. Chen et al. prepared a MnO2–vermiculite composite, which circumvents the problems of conventional MnO2 coalescence. At the same time, the specific surface area of MnO2–vermiculite (298.18 m2/g) was much larger than that of single MnO2, exposing more active sites and facilitating the adsorption of Tl. The Tl-O bond is formed by ion exchange on the material surface, and an oxidative precipitation process also takes place [43]. Li et al. prepared zero-valent manganese (nZVMn) in synergy with different oxidants to remove Tl from wastewater. The main mechanisms for the removal of Tl include oxidation-induced precipitation, electrostatic attraction, and surface adsorption with a removal rate of over 95.7%. It has been demonstrated that the composite material of MnO2 and iron can effectively remove heavy metals from water [34]. Chen et al. developed FeOOH-loaded MnO2 composites as emergency adsorbent materials for Tl(I) removal. Studies have shown that the Fe/Mn molar ratio of 1:2 FeOOH loaded MnO2 has a significantly enhanced ability to remove Tl(I) compared to pure MnO2, and the dissolution of MnO2 in this material forms new adsorption sites. Tl(I) can be oxidized to Tl(III) by this material to produce a Tl(OH)3 precipitate [44].
It has been found from the above studies that the predominant mechanism for thallium removal by manganese oxides is the pre-oxidation of Tl(I) to Tl(III) [40]. Subsequently, Tl(III) can either be adsorbed onto the negatively charged manganese group on the surface or generate a Tl(III) hydroxide precipitate. Tl(I) can also be adsorbed through the interaction between ions. At present, research on the removal of Tl by modified manganese oxides is still in the exploratory stage, and strengthening the selection of manganese-oxide-modified materials and modification methods can be the focus of future research. However, research cannot only be carried out under laboratory conditions but should also consider the application in real wastewater. The +1 oxidation state of thallium can be photo-oxidized (to the stable +3 state) or photo-reduced (to the elemental form). Some studies have already reported the photoelectrochemical oxidation of Tl(I) to (Tl2O3). Therefore, the effect of UV-visible light on the valence state of thallium also needs to be considered in practical water-treatment processes [45]. Finally, the manganese oxides loaded with thallium are harmful to the environment and to humans. Used manganese oxides can only be discharged into the environment once the heavy metals have been fully recovered. It is important that used sorbents are disposed of in an environmentally friendly manner and that this needs to be further investigated in subsequent research. The advantages and disadvantages of manganese oxide for thallium removal are shown in Table 1.

3.2.2. Thallium Removal by Iron Oxides

Iron-oxide-based materials have advantages over other adsorbent materials due to their inherent magnetic properties, which can be easily separated after the adsorption process by applying an external magnetic field (Table 2). Previous studies have found that common iron-based materials, such as FeCl3, do not adsorb thallium very well. In their study [46], Coup et al. found that the adsorption constants of hydrous iron ore for Tl+ could be as high as four orders of magnitude [47]. Back in 2007, Yantasee et al. found that superparamagnetic iron oxide (Fe3O4) nanoparticles functionalized with dimercaptosuccinic acid (DMSA) surfaces could effectively adsorb thallium. The results indicate that the optimized adsorption of Tl is possible at a neutral pH [48]. However, as iron oxides can undergo Fenton reactions. The disadvantages of Fenton technology are the narrow pH range (pH = 2.8–3.5), the low efficiency of iron ion recycling and the formation of undesirable iron sludge. Therefore, the modification of iron oxides is also a priority [49,50]. Solid-phase extraction methods for ferrous metal cyanides have attracted a lot of attention due to their high selectivity, rapid separation, high thermal stability, and radiation stability [51,52]. Magnetic Prussian blue (PB) was synthesized at room temperature by Zhang et al. [53]. The particles can be used for a highly selective removal of Tl. The experimental Tl removal in this study reached 528  mg/g. The high selectivity of magnetic PB for Tl is attributed to the fact that hydrated Tl has a smaller hydrated diameter and a lower hydration-free energy than other coexisting ions, and hence, can easily enter the open pores on magnetic PB. Li et al. investigated the reversible adsorption–desorption process mediated by magnetite (Fe3O4) for the removal of Tl(I) from wastewater. The results showed that adsorption under alkaline conditions achieved a rapid and effective removal of Tl(I), and desorption under acidic conditions achieved a rapid and effective enrichment of Tl(I). This study found no significant loss of magnetite in the cycling experiments, indicating that the removal and recovery of Tl(I) can be effectively and consistently repeated. This makes the reversible magnetite-based adsorption–desorption process a promising technology for further development [35].
Recently, low-cost, environmentally friendly, in situ operationally feasible, and easily magnetically selectable nano-zero-valent iron (nZVI) has started to be widely used in contaminated soil and water [54]. nZVI could activate oxygen to generate H2O2 through a two-electron reaction pathway, then H2O2 could react with Fe (II) released by nZVI, and the generated·OH could easily oxidize Tl (I) to Tl (III) (reactions 3-1 and 3-2). nZVI has a detoxifying effect on thallium-containing wastewater and groundwater, but its low adsorption capacity and rapid passivation rate limit the large-scale application of nZVI [55,56]. Therefore, Deng et al. have synthesized a nanoscale zero-valent iron, Fe@Fe2O3 core–shell nanowire (FCSN) [57]. The results showed that FCSN was able to adsorb Tl(I) with a maximum adsorption capacity of 591.7 mg/g. The material was not affected by the pH, and coexisting ions could reach a maximum adsorption rate of 95.6% when a suitable adsorbent was added. Based on the experimental data, it can be assumed that since the adsorption of Tl(I) by FCSNs is a chemical process mainly based on particle diffusion, it is not influenced by pH or coexisting ions. The material can be regenerated by desorption with H3PO4, with a desorption efficiency of 90.7%. However, the reproducibility of the material was also poor, with the removal of Tl decreasing to 52.3% after five cycles. Further studies could be explored to investigate the renewability of the material [58]. The material can be regenerated by desorption with H3PO4, with a desorption efficiency of 90.7%. However, the reproducibility of the material was also poor, with the removal of Tl decreasing to 52.3% after five cycles. Further studies could be explored to investigate the renewability of the material [59]. The mechanism of thallium adsorption by 3-ZVIMn is shown in Figure 3. The adsorption mechanism of thallium by 3-ZVIMn is shown in Figure 3. The results showed that the maximum adsorption of thallium is 990 mg/g. The Langmuir model is more appropriate for the description of thallium removal by 3-ZVIMn, indicating that the adsorption process is monolayer adsorption on a homogeneous surface. The main adsorption mechanisms are: (i) The oxidation of Tl(I) to produce Tl2O3 precipitation on the surface of the material and then be adsorbed by 3-ZVIMn; (ii) The complexation of Tl(I) with the hydroxyl group; (iii) The electrostatic attraction between the negatively charged 3-ZVIMn (pH = 10) and Tl(I). The material can be used in elution experiments with low concentrations of HCl, and the material can still adsorb 88.9% of thallium after five cycles. Liu et al. prepared amorphous zero-valent iron sulfide materials by one-step vulcanization using Na2S2O3 as the vulcanizing agent [56]. The S-ZVI material with S/Fe = 3 is effective in Tl(I) removal. It can directly treat high concentrations of Tl-containing wastewater with a maximum adsorption capacity of 654.4 mg/g of Tl. They also investigated the main mechanisms of thallium adsorption by S-ZVI: (i) Reaction of Tl(I) ions with sulfides in the material to form thallium sulfide precipitates (Tl2S); (ii) Surface complexation of Tl(I) with sulfur-iron compounds in the adsorbent; (iii) Electrostatic attraction of Tl+ to negatively charged S-ZVI (pH = 7). The mechanistic contribution of S-ZVI to the removal of Tl is in the order of sulfide precipitation > surface complexation > electrostatic attraction. The mechanism of thallium adsorption by S-ZVI is shown in Figure 4.
Fe ( 0 ) + O 2 + 2 H +   Fe ( II ) + H 2 O 2
H 2 O 2 + Fe ( II )   OH + OH + Fe ( III )
There is also an increasing interest in studying materials for the adsorption of thallium by combining iron oxides with other metals through modification. Fe-Mn binary oxides have the strong oxidation power of manganese dioxide and the high adsorption capacity of iron oxides. Li et al. used simultaneous chemical oxidation and the precipitation method to synthesize of Fe-Mn bimetallic materials for the removal of Tl(I). Tl adsorption to Fe-Mn adsorbents is fast, efficient, and selective. Equilibrium adsorption can be achieved by over 95% at a wide range of the operating pH values (3–12) and high ionic strengths (0.1–0.5 mol·L−1). This study demonstrates that the mechanism of Tl adsorption is mainly in the complexation of hydroxyl groups and Tl(I) on the Fe-Mn surface, as well as the oxidation of Tl(I) to Tl(Ⅲ) and the generation of Tl2O3 precipitates on the surface of the material to be adsorbed by the material [60]. Li et al. fabricated a graded composite containing magnetic pyrite slag and MnO2 to remove Tl(I) [35]. The results show that thallium is well adsorbed at a pH from 2 to 12, independent of coexisting cations and humic acids. The mechanism of Tl adsorption by this material is shown in Figure 5. In addition, after five cycles of regeneration experiments, there was no significant variation in the adsorption capacity of the Fe-Mn composites for Tl(I). Fe-Mn-binary-oxide-activated aluminosilicate minerals (FMAAM) were prepared using the hydrothermal synthesis of KMnO4, activated aluminosilicates, and Fe(NO3)3. The mesoporous structure of FMAAM provided a large number of surface-binding sites for Tl(I), and the maximum capacity of FMAAM for Tl(I) adsorption was 78.06 mg/g. The results of the XPS analysis indicate that the main mechanisms for the removal of Tl(I) include oxidative precipitation, ion exchange, and surface complexation [61]. Fe3O4@TiO2, synthesized by loading iron oxide on the surface of reduced graphene oxide (RGO) nanosheets and subsequently coated with titanium oxide, had a maximum capacity of 673.2 mg/g of Tl adsorbed at pH 8.0. The analysis results show that the -COOH group on the material present can partially oxidize Fe(II) to Fe(III). Meanwhile, the formation of more oxygen-containing functional groups on the RGO nanosheets leads to the rapid oxidation of Tl+ to Tl3+. Therefore, oxidation and precipitation are the mechanisms by which RGO nanosheets remove Tl from wastewater. The material maintains a stable Tl removal efficiency (three cycles) and does not require regeneration or activation procedures to restore its performance in the cycling test [62]. However, the cost of the material has not been considered in any of the above studies. Further research will have to consider the economic feasibility and viability of the actual wastewater treatment process.
For iron oxides, current studies have found that the mechanisms of thallium adsorption include surface complexation, electrostatic attraction, oxidative precipitation (sulfide precipitation), and ion exchange. However, few materials are unaffected by pH, co-existing ions, etc., and the reproducibility of the materials requires further discussion. Materials unaffected by pH, co-existing ions, and organic matter that have good regenerative properties are still need to be developed. Furthermore, the use of these materials in actual thallium-containing water/wastewater needs to be further verified.
Table 2. Iron oxide removal thallium performance and its advantages and disadvantages.
Table 2. Iron oxide removal thallium performance and its advantages and disadvantages.
Metal OxidesDosageThallium Concentration in the Tested WaterPerformanceAdvantagesDisadvantagesRef.
Magnetic Prussian blue0.6 g/L1000 μg/L528 mg/LHigh removal efficiencyPoor reproducibility[53]
Fe@Fe2O3 Core–Shell0.75 g/L10 mg/L95.60%Less influenced by environmental conditionsPoor reproducibility[57]
3-ZVIMn0.1 g/L200 mg/L990 mg/gHigh removal efficiencyHighly influenced by environmental conditions[59]
Fe-Mn binary oxides0.5 g/L10 mg/L197.6 mg/gLess influenced by environmental conditionsPoor reproducibility[60]
Titanium iron magnetic0.1 g/L12.5 mg/L111.3 mg/gLess influenced by environmental conditionsPoor reproducibility[58]
Fe-Mn binary oxides activated aluminosilicate mineral1.0 g/L10 mg/L78.06 mg/gLess influenced by environmental conditionsPoor removal ability[61]
Fe3O4@TiO2 decorated RGO nanosheets0.2 g/L_673.2 mg/gHigh removal efficiencyHighly influenced by environmental conditions[62]

3.2.3. Thallium Removal by Aluminum Oxides

Activated alumina (AA), with its large surface area, pore size, and wide distribution of micropores, is a traditional adsorbent used to remove heavy metals [63]. The USEPA has recommended AA as the best available technology for Tl removal from water [64]. However, many studies have demonstrated the low efficiency of activated alumina for thallium removal [65]. The adsorption of thallium by nano-Al2O3 has attracted widespread attention. Zhang et al. chose nano-Al2O3 as the object of study for the adsorption of thallium. The results showed that the adsorption of trivalent thallium at pH = 4.5 was as high as 99.56%. This is mainly due to the fact that most atoms are not saturated and are highly susceptible to chemisorption with other atoms on the surface of the nanoparticles [66].
Aluminum composites can also be used as a method of removing thallium. Senol and Ulusoy investigated the adsorption properties of PAAm-Z and PAAm-B on Tl+ and Tl3+ by the direct polymerization of polyacrylamide (PAAm), zeolite (Z), and bentonite (B) composites (PAAm-Z, PAAm-B) prepared from AAm monomers in clay and zeolite suspensions [67]. SEM images of PAAm, PAAm-Z and PAAm-B are shown in Figure 6. The high affinity of PAAm-Z for Tl is probably due to the wide distribution of the “zeolite” particles. This leads to an increase in the specific surface area and the number of active sites for adsorption. The microporous structure formed in the material also facilitates ion exchange with Tl. The researchers performed four regeneration experiments on PAAm-Z and PAAm-B using 0.25 mol·L−1 HCL (the materials were used five times in total). The results of the IR spectrograms show that the efficiency of PAAm-Z decreases significantly after five uses, which may be due to blockage of the adsorption center during acid regeneration or due to the breakdown of some of the active sites in the zeolite. Other researchers have utilized aluminum beverage can powder (AlCP) as a replacement for zero-valent aluminum and the oxidation of Tl(I) at pH 9.5 followed by an alkaline-induced precipitation of Tl(III) to achieve thallium removal. The results show a maximum removal rate of 92% for thallium. The mechanism of thallium removal from water by aluminum oxides is mainly through surface complexation, ion exchange, and electrostatic attraction processes [68]. There are relatively few studies on aluminum oxides, probably due to their instability. Moreover, there are few studies on the effects of organic matter, co-existing ions, and pH on thallium removal by aluminum oxides. However, it was shown that aluminum oxide (Al2O3) with an average particle size of 63 μm was modified with the anionic surfactant sodium dodecyl sulfate (SDS) and applied to the solid-phase extraction separation of (I) and (III) thallium and the pre-enrichment of Tl(III) in wastewater. Only Tl(III) was finally absorbed left on the adsorbent, and the material was also well tolerated to Pb and Cd ions, making it a fast and effective material for Tl extraction and morphological analysis [69]. The renewable nature of the material and the practical application implications are also rarely mentioned and will be discussed further in future studies [38]. The advantages and disadvantages of aluminum oxide for thallium removal are shown in Table 3.

3.2.4. Thallium Removal by Titanium Oxides

Titanium oxides can also play a role in the removal of thallium [38]. As early as 2003, Kajitvichyanukul et al. studied the titanium dioxide particles adsorbed Tl(I) in aqueous media. The results show that up to 95% of Tl(I) is bound to titanium dioxide at low-phosphate concentration levels [71]. The removal mechanism is mainly through surface complexation and electrostatic attraction. In recent years, researches have shifted to materials at the nanoscale. Zhang et al. investigated the removal of Tl(III) from water by titanium dioxide nanoparticles. They showed that at pH = 4.5, the removal of Tl3+ from the solution by titanium dioxide nanoparticles was close to 100%, but the experiment did not consider the regeneration of the titanium dioxide nanomaterials [72].
Titanium carbonate nanotubes (TNTs) are also some of the materials used to remove thallium from wastewater [73]. Liu et al. demonstrated that hydrothermally synthesized TNTs performed well in the adsorption of highly toxic Tl+/Tl3+ [36]. TNTs can be synthesized hydrothermally using TiO2 and NaOH solutions at moderate temperatures. Hydrothermally synthesized titanate nanotubes are effective in removing Tl(I) and Tl(III) from aqueous solutions because of their layered structure consisting of small diameter tubes with large surface areas. The mechanism of Tl removal by TNTs is shown in Figure 7. The structural changes of TNTs after the adsorption of Tl(I) and Tl(III) are shown in Figure 8. The maximum capacity of Tl(I) adsorbed by TNTs was 709.2 mg/g. The mechanism of the Tl(I) removal from wastewater by TNTs is mainly due to the ion exchange between Na+ and Tl+ in the interlayer of TNTs. XPS analysis showed that the removal of Tl(III) by TNTs consisted of two stages: a co-precipitation stage at high concentrations and an ion exchange stage at low concentrations. The experiment also considered the regeneration of TNTs. The results showed that TNTs also exhibit a significant adsorption capacity for thallium after the desorption by HNO3 and the regeneration by NaOH.
Titanium peroxide, synthesized by simple oxidation combined with precipitation, has also been used as a candidate material for the removal of thallium. Zhang et al. investigated the performance and mechanism of Tl(I) adsorption by titanium peroxide. The main adsorption process is the ionic exchange of the hydroxyl groups on the surface of titanium peroxide with Tl(I) to form Ti-O-Tl complexes. The results showed that Tl(I) adsorption increased with increasing pH. The maximum capacity of Tl(I) adsorbed by titanium peroxide at pH = 7 was 412 mg/g, indicating that titanium peroxide was very effective in removing Tl(I) from water. The mechanism is shown in Figure 9 [74]. However, the material is subjected to a maximum of three adsorption–regeneration cycles, HNO3 is not suitable for regenerating titanium peroxide.
The adsorption of thallium by titanium oxides is mainly accomplished by oxidative precipitation, ion exchange reactions, and electrostatic attraction. The most used in current research is titanium carbonate nanotubes. However, this is due to the high cost of titanium carbonate nanotubes, the influence of organic matter, co-existing ions, and pH on it. Therefore, the above issues need to be considered in subsequent studies.
The removal of thallium from water using metal oxides has been extensively studied, but most of them were studied based on laboratory conditions. As thallium is present in much lower concentrations in actual wastewater than under laboratory conditions, the feasibility of the studied methods for thallium removal of thallium in actual wastewater needs to be further investigated. Only methods with reasonable costs and high removal efficiency are widely used in the actual treatment of water/wastewater. For sustainability reasons, the reuse of materials should also be of concern. In the published literature, only a few studies have been carried out on material recyclability, and materials with low recyclability are not suitable for practical wastewater removal from thallium. The exhausted material is enriched with large amounts of thallium, which can easily cause secondary contamination, but there is very little mention of how to deal with the depleted material in current research. It is necessary to consider this in subsequent studies. Table 4 shows the advantages and disadvantages of titanium oxide for thallium removal.

3.3. Factors Affecting the Effectiveness of Thallium Removal by Metal Oxides

The main factors affecting the effectiveness of metal oxides in removing thallium from aqueous systems are pH, co-existing ions, and organic matter. Therefore, we analyzed these three factors in the removal of thallium by metal oxides, summarized some conclusions, identified some issues that can be addressed in future research, and enabled the better application of metal oxides for thallium removal.

3.3.1. Effect of pH

The surface charge of metal oxides is significantly influenced by the pH of the solution. When the pH < pHPZC, H+ on the adsorbent surface would prevent the Tl(I) adsorption by electrostatic repulsion. Possible ion exchange between Tl(I) and the hydroxyl group on the adsorbent surface contributes to the removal of Tl. In addition, Tl(I) can also be oxidized to Tl(III), which then precipitates Tl(OH)3 on the surface of the metal oxides. As the pH of the solution increases, the positive charge on the adsorbent surface will be consumed with OH reaction. When pH > pHPZC, the metal oxides’ surface is negatively charged and has great affinity for positive ions, and Tl(I) could be adsorbed with high affinity through electrostatic attraction. The adsorption mechanisms for Tl(I) at low and high pH values are shown in Figure 10 [21].
The effect of pH is different for different metal oxides. For manganese oxides, on the one hand, pH changes the oxidation of the metal oxide and its surface charge; on the other hand, the solution contains a large amount of H+, which leads to the oxidation of Tl+, and then the generated Tl3+ forms sediment [21]. Redox reactions between heavy metals and manganese oxide at different pH conditions play an important role in the adsorption of Tl [75]. Previous studies have indicated that different manganese oxide adsorbent materials have different pH ranges for which they are required, but that the majority of manganese oxides are best removed under alkaline conditions. For example, the removal efficiency of Tl(I) by nano-MnO2 was lower under acidic conditions than under neutral and alkaline conditions [76,77]. For nanoparticles with zero-valent manganese, at pH > 12, it was observed that more than 95.7% of thallium was removed [34]. At a high pH, the negative-charge sites enhance the attraction between Tl(I) and the hydroxyl groups on the surface of FeOOH-loaded MnO2 composites [43]. Materials such as δ-MnO2 [41] and MnO2@slag [35] also demonstrated that their adsorption capacity reached its maximum at pH = 10.
For iron and titanium oxides, the reason for the increase in negative-charge density on the adsorbent surface with the increasing solution pH is the constant deprotonation of functional groups [78]. The enhanced electrostatic attraction between Tl(I) and the sorbent favors the adsorption of Tl(I). When the initial pH < 6, the hydroxyl groups on the surface of the titanium peroxide react with the protonate and H+. The ions adsorbed by forming the outer spherical complex are sensitive to ionic strength because the background electrolyte ions also form outer spherical ions via electrostatic forces. In contrast, the inner sphere complexes are less sensitive to changes in ionic strength and adsorb higher ionic strengths [74]. The adsorption of thallium by iron and titanium oxides under pH < 7 also behaves consistently with manganese oxides. Li et al. demonstrated that the removal of thallium by Fe-Mn dioxides was almost independent of pH. The removal of Fe-Mn dioxides reached more than 99% in the pH range of 5–12 [35]. For magnetite, the adsorption rate is close to 100% at pH > 12 [60]. The best adsorption results were obtained for titanium–iron magnetic adsorbents at a pH of 10 [60]. Both titanium dioxide [71] and titanium peroxide [79] reached their maximum adsorption under alkaline conditions.
It has also been documented that some materials are more conducive to Tl(I) removal under weakly acidic and weakly basic conditions. This was demonstrated, for example, by studies on the adsorption of thallium by HMO [80] and fancy MnO2-coated magnetic pyrite ash. In a study of titanium carbonate nanotubes it was found that thallium removal efficiency reached 100% at pH > 5 [36].
Unlike Mn, Fe, and Ti oxides, Tl(I) adsorption for Al oxides is more favorable under acidic conditions [70]. The pH affects alumina nanoparticles’ surface active-site distribution, with adsorption reactions occurring more readily at a pH of 3–4.5 [63]. The polyacrylamide–aluminosilicate reaction reaches equilibrium was much lower when the pH = 3 [67].
In summary, pH is a very important factor for thallium removal. Different metal oxides require different pH ranges, and some show good removal performance under acidic and alkaline conditions and poor performance under neutral conditions, such as perovskite nanoparticles [46]. Overall, the effect of pH on thallium removal by metal oxides is still relative. In subsequent studies, we have to consider the effect of pH and select the optimum pH as the condition for thallium removal (Table 5).

3.3.2. Effect of Co-Existing Ions

The adsorption of metal ions under real water conditions is a complex process. The coexisting ions, dissolved organic matter, and many other components can occupy the active site of the adsorbent. All of these factors can reduce the effectiveness of the material for the adsorption of the target contaminant. Alkali or alkali metal ions in the water column have similar physicochemical properties to heavy metal ions, and there is competition for adsorbent sorption sites for Tl+. The metal oxide adsorbents were easily disturbed by coexisting ions. Lower concentrations of coexisting ions (0.001 mol·L−1) had less effect on the flowery MnO2-coated magnetic pyrite ash slag [81], γ-MnO2 [41], and MnO2@slag [35], but a coexisting ion concentration of 0.1 mol·L−1 resulted in a significant decrease in adsorption capacity. For manganese oxides, Ca2+ and Na+ are the most widely studied co-occurring cations affecting thallium removal efficiency. Na+ may weaken the electrostatic adsorption of δ-MnO2, with only 1 mmol·L−1 Ca2+ greatly reducing the adsorption rate of δ-MnO2 on thallium under medium and alkaline conditions [41]. Ca2+ may compete with nano-MnO2 for adsorption sites and affect the aggregation of nano-MnO2, which significantly reduces the removal of Tl [57,77]. Mg2+ can also affect thallium removal. In a study of zero-valent Mn nanoparticles it was found that adsorption decreased from 72% to 8.9% when Mg2+ existed [43]. For aluminum oxides, the effect of Ca2+ was the most significant. γ-Al2O3 adsorption of Tl(I) (6–26 mg/g) decreased with increasing Ca2+ concentration (0–5 mol·L−1). For nano-Al2O3, the interference of Cd2+, Cu2+, Mn2+, and Pb2+ should not be neglected [66]. In studies of iron oxides it was found that in addition to Ca2+ and Na+, which affect removal, Cu and Zn ions also compete with thallium for adsorption on copper ferricyanide at pH near neutral. For titanium oxides, the main consideration is the effect of the co-existence of K+, Na+, and Ca2+ ions: (1) K+ has similar chemical properties and ionic radii to the Tl+, so K+ interferes with the reaction between Tl(I) and the binding site; (2) The removal of Tl(I) was enhanced at low Na+ concentrations and slightly inhibited at high concentrations; (3) The effect of Ca2+ on titanium oxides is due to the different valence states and the affinity for Tl3+. The complexation between Ca2+ ions and trivalent ions is inhibited strongly due to the ionic radius and different valence states. Titanium oxide adsorbs ions by electrostatic-driven outer sphere complexation and/or cation exchange. The decrease in its adsorption capacity with increasing ionic strength is due to its sensitivity to changes in ionic strength. In addition to the above three ions, the remaining coexisting ions also affect the removal efficiency of thallium. For example, at a Cu2+ concentration of 10 mg/L, the adsorption capacity of the titanium–iron magnetic material for Tl decreased by approximately 43.8% [71].
However, some adsorbents are less disturbed by co-existing ions. For example, when magnetite was used to remove thallium, the adsorption of thallium by magnetite remained as high as 92% as the ionic strength increased. It is indicated that the effect of ionic strength is slight [35]. The coexisting ions Mg2+, PO43−, and turbidity negatively affected the removal of Tl(I) and K2FeO4, while Ca2+, Na+, Cl, HCO3, and NO3 had little effect on the removal of Tl(I) [70]. Co-existing ions such as Pb, Cd, and Zn also have less effect on thallium removal using Al2O3-SDS solid-phase extraction as the sorbent. The material is highly tolerant to these interfering ions [69]. Since the adsorption of Tl(I) by titanium peroxide is mainly achieved by the formation of intra-sphere complexes on its surface, the effect of ionic strength is not significant [75]. The ionic strength plays an important role in removing thallium by metal oxides. In subsequent studies, attention should be paid to the effect of competing cations such as Na+, Ca2+, and Mg2+ on the adsorption of Tl by the material. Although the effect of co-existing ions on thallium removal has been studied in previous studies for the above ions, in practice, however, copper, zinc, cadmium, and lead ions should be considered in studies as strong interferers/competitors for oxide retention, especially when thallium is used as tailings. We cannot only consider common co-existing ions alone in future studies but also the influence of common metal ions in tailings. Researchers can develop materials with little influence of co-existing ions to remove thallium from water/wastewater.

3.3.3. Effect of Organic Matters

It is well known that natural organic matter (NOM) plays an important part in the transfer of metal ions [82]. NOM is a complex mixture of compounds. However, the most studied and reactive fraction is the humus consisting of fulvic acid (FA) and humic acid (HA). The complexation of thallium with HA increases with increasing HA concentration, indicating that thallium binds well to HA. If HA is present, thallium will bind to HA and lead to a reduction in adsorption efficiency.
The presence of organic matter affects the oxidation and removal of Tl(I). The organic matter is adsorbed on the adsorbent surface, thus affecting the contact of Tl with the adsorption sites. The functional groups in the organic matter can complex with Tl, forming stable compounds and reducing the removal rate. For manganese oxides, FA and HA contain complexing functional groups such as -COOH and -OH, which complex manganese and make it more soluble in the solution. NOM can reduce dissolved manganese oxides. EDTA and DTPA have a greater inhibitory effect on thallium adsorption than HA [35,41,63,76]. However, strong resistance to HA was demonstrated in studies of flowery MnO2-coated magnetic pyrite ash slag [83]. Tl removal decreases from 90% to 30% as HA concentration increases from 0 to 10 mg/L. EDTA and DTPA complex with Tl(III) and Tl(I), both of which have an inhibitory effect on Tl removal [35]. However, very few studies have investigated the effect of organic substances on the removal of aluminum and titanium oxides, and further efforts in this area are needed in future studies.
In summary, in actual wastewater, organic substances compete with thallium for adsorption. They can affect thallium removal, but only some of the studies have considered this aspect, and more consideration will have to be given to this issue in subsequent studies to achieve high thallium removal rates.
The pH, co-existing ions, and organic matter are important influences on thallium removal. Most current studies have considered the effect of pH and were able to identify the optimal pH for removal needs to be studied. However, since most of the current studies have been carried out under laboratory conditions and do not consider actual water background conditions, the effects of organic matter and co-existing ions have been ignored. The influence of the above factors should be considered to develop suitable techniques that effectively remove thallium from real wastewater.

4. Conclusions

Thallium pollution caused by industrial development is becoming increasingly serious and endangers people’s daily lives. Numerous studies have proven that the use of synthetic composites of metal oxides loaded on the surface of large organic materials to treat thallium contamination is one of the feasible technologies. Moreover, because of the different removal mechanisms of thallium by different metal oxides, it can even realize the repair and chemical extraction of thallium. Of the four metal oxides, modified iron oxides are the most promising metal oxides for future thallium removal from water/wastewater due to their high efficiency, low cost, and ease of recovery. However, the application of these metal oxide materials to actual thallium-contaminated water sources will be a major challenge for the future. In future research, the cost and regeneration of metal oxide materials will need to be considered, and the effects of pH, coexisting ion concentrations, and organic concentrations, as well as the actual background conditions of the wastewater, should also be taken into account. Furthermore, from a green and sustainable perspective, exploring the use of lower-cost nanometallic oxides for the treatment of Tl in wastewater, combined with the effects of organic matter, co-existing ions and pH, may be a more promising study for the future due to the current high price of nanometallic oxides.

Author Contributions

Conceptualization, X.R., H.F. and J.W.; Formal analysis, X.R., H.F. and M.Z.; Data curation X.R. and H.F.; Methodology, X.R. and H.F.; Writing-original draft, X.R.; Supervision, M.Z., X.Z. (Xin Zhou), X.Z. (Xu Zhu) and J.W.; Validation, M.Z., X.Z. (Xin Zhou), X.Z. (Xu Zhu), X.O., J.T. and C.L.; Writing-review and editing, X.Z. (Xin Zhou), X.Z. (Xu Zhu), X.O., J.T., C.L., J.W., W.T. and L.T.; Project administration, J.W.; Funding acquisition, J.W., W.T. and L.T.; Resources, W.T. and L.T. All authors have read and agreed to the published version of the manuscript.

Funding

The study was financially supported by the National Natural Science Foundation of China (52100008, 52100142, 52100184, 22276048), the Project of the National Key Research and Development Program of China (2021YFC1910404), the Funds of Hunan Science and Technology Innovation Project (2021GK4055), Natural Science Foundation of Hunan Province, China (2021JJ40091).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data will be made available when requested.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Liu, J.; Luo, X.; Sun, Y.; Tsang, D.C.; Qi, J.; Zhang, W.; Li, N.; Yin, M.; Wang, J.; Lippold, H.; et al. Thallium pollution in China and removal technologies for waters: A review. Environ. Int. 2019, 126, 771–790. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, A.; Ye, J.; Ren, G.; Qi, Y.; Chen, Y.; Zhou, S. Metal-Free Semiconductor-Based Bio-Nano Hybrids for Sustainable CO2-to-CH4 Conversion with High Quantum Yield. Angew. Chem. Int. Ed. 2022, 61, e202206508. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, H.; Luo, Y.; Wang, P.; Zhu, J.; Yang, Z.; Liu, Z. Removal of thallium in water/wastewater: A review. Water Res. 2019, 165, 114981. [Google Scholar] [CrossRef] [PubMed]
  4. Campanella, B.; Casiot, C.; Onor, M.; Perotti, M.; Petrini, R.; Bramanti, E. Thallium release from acid mine drainages: Speciation in river and tap water from Valdicastello mining district (northwest Tuscany). Talanta 2017, 171, 255–261. [Google Scholar] [CrossRef]
  5. Perotti, M.; Petrini, R.; D’Orazio, M.; Ghezzi, L.; Giannecchini, R.; Vezzoni, S. Thallium and Other Potentially Toxic Elements in the Baccatoio Stream Catchment (Northern Tuscany, Italy) Receiving Drainages from Abandoned Mines. Mine Water Environ. 2018, 37, 431–441. [Google Scholar] [CrossRef]
  6. Rinklebe, J.; Shaheen, S.M.; El-Naggar, A.; Wang, H.; Du Laing, G.; Alessi, D.S.; Ok, Y.S. Redox-induced mobilization of Ag, Sb, Sn, and Tl in the dissolved, colloidal and solid phase of a biochar-treated and un-treated mining soil. Environ. Int. 2020, 140, 105754. [Google Scholar] [CrossRef]
  7. Tatsi, K.; Turner, A. Distributions and concentrations of thallium in surface waters of a region impacted by historical metal mining (Cornwall, UK). Sci. Total. Environ. 2014, 473–474, 139–146. [Google Scholar] [CrossRef]
  8. Xiao, J.; Jin, Z.; Wang, J. Geochemistry of trace elements and water quality assessment of natural water within the Tarim River Basin in the extreme arid region, NW China. J. Geochem. Explor. 2014, 136, 118–126. [Google Scholar] [CrossRef]
  9. Xiao, T. Environmental Impact of Thallium Related to the Mercury-Thallium-Gold Mineralization in Southwest Guizhou Province, China; Universite du Quebec a Chicoutimi: Saguenay, QU, Canada, 2001. [Google Scholar]
  10. Peacock, C.L.; Moon, E.M. Oxidative scavenging of thallium by birnessite: Explanation for thallium enrichment and stable isotope fractionation in marine ferromanganese precipitates. Geochim. Cosmochim. Acta 2012, 84, 297–313. [Google Scholar] [CrossRef]
  11. Khavidaki, H.D.; Aghaie, H. Adsorption of Thallium(I) Ions Using Eucalyptus Leaves Powder. Clean-Soil Air Water 2013, 41, 673–679. [Google Scholar] [CrossRef]
  12. Krasnodebska-Ostrega, B.; Dmowski, K.; Stryjewska, E.; Golimowski, J. Determination of Thallium and Other Elements (As, Cd, Cu, Mn, Pb, Se, Sb, and Zn) in Water and Sediment Samples from the Vicinity of the Zinc-Lead Smelter in Poland (3 pp). J. Soils Sediments 2005, 5, 71–73. [Google Scholar] [CrossRef]
  13. Yao, J.; Wang, H.; Ma, C.; Cao, Y.; Chen, W.; Gu, L.; He, Q.; Liu, C.; Xiong, J.; Ma, J.; et al. Cotransport of thallium(I) with polystyrene plastic particles in water-saturated porous media. J. Hazard. Mater. 2022, 422, 126910. [Google Scholar] [CrossRef] [PubMed]
  14. Yu, M.; Wang, J.; Tang, L.; Feng, C.; Liu, H.; Zhang, H.; Peng, B.; Chen, Z.; Xie, Q. Intimate coupling of photocatalysis and biodegradation for wastewater treatment: Mechanisms, recent advances and environmental applications. Water Res. 2020, 175, 115673. [Google Scholar] [CrossRef] [PubMed]
  15. Khan, F.S.A.; Mubarak, N.M.; Tan, Y.H.; Khalid, M.; Karri, R.R.; Walvekar, R.; Abdullah, E.C.; Nizamuddin, S.N.; Mazari, S.A. A comprehensive review on magnetic carbon nanotubes and carbon nanotube-based buckypaper for removal of heavy metals and dyes. J. Hazard. Mater. 2021, 413, 125375. [Google Scholar] [CrossRef]
  16. Tereshatov, E.E.; Boltoeva, M.Y.; Folden, C.M. Resin Ion Exchange and Liquid-Liquid Extraction of Indium and Thallium from Chloride Media. Solvent Extr. Ion Exch. 2015, 33, 607–624. [Google Scholar] [CrossRef]
  17. Yang, L.; Xiao, J.; Shen, Y.; Liu, X.; Li, W.; Wang, W.; Yang, Y. The efficient removal of thallium from sintering flue gas desulfurization wastewater in ferrous metallurgy using emulsion liquid membrane. Environ. Sci. Pollut. Res. 2017, 24, 24214–24222. [Google Scholar] [CrossRef]
  18. Ma, C.; Huang, R.; Huangfu, X.; Ma, J.; He, Q. Light- and H2O2-Mediated Redox Transformation of Thallium in Acidic Solutions Containing Iron: Kinetics and Mechanistic Insights. Environ. Sci. Technol. 2022, 56, 5530–5541. [Google Scholar] [CrossRef]
  19. Meng, C.; Hu, X.; Yu, S.; Co, D.S. Removal of thallium in raw water by powdered activated carbon. Water Technol. 2015, 9, 1–3. [Google Scholar]
  20. Ye, J.; Wang, C.; Gao, C.; Fu, T.; Yang, C.; Ren, G.; Lü, J.; Zhou, S.; Xiong, Y. Solar-driven methanogenesis with ultrahigh selectivity by turning down H2 production at biotic-abiotic interface. Nat. Commun. 2022, 13, 6612. [Google Scholar] [CrossRef]
  21. Zhao, Z.; Xiong, Y.; Cheng, X.; Hou, X.; Yang, Y.; Tian, Y.; You, J.; Xu, L. Adsorptive removal of trace thallium(I) from wastewater: A review and new perspectives. J. Hazard. Mater. 2020, 393, 122378. [Google Scholar] [CrossRef]
  22. Zhang, H.; Tang, L.; Wang, J.; Yu, J.; Feng, H.; Lu, Y.; Chen, Y.; Liu, Y.; Wang, J.; Xie, Q. Enhanced surface activation process of persulfate by modified bagasse biochar for degradation of phenol in water and soil: Active sites and electron transfer mechanism. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 599, 124904. [Google Scholar] [CrossRef]
  23. Soltani, R.; Marjani, A.; Shirazian, S. Facile one-pot synthesis of thiol-functionalized mesoporous silica submicrospheres for Tl(I) adsorption: Isotherm, kinetic and thermodynamic studies. J. Hazard. Mater. 2019, 371, 146–155. [Google Scholar] [CrossRef] [PubMed]
  24. Sabermahani, F.; Mahani, N.M.; Noraldiny, M. Removal of thallium (I) by activated carbon prepared from apricot nucleus shell and modified with rhodamine B. Toxin Rev. 2016, 36, 154–160. [Google Scholar] [CrossRef]
  25. Gupta, K.; Joshi, P.; Gusain, R.; Khatri, O.P. Recent advances in adsorptive removal of heavy metal and metalloid ions by metal oxide-based nanomaterials. Co-ord. Chem. Rev. 2021, 445, 214100. [Google Scholar] [CrossRef]
  26. Liu, J.Y.; Chang, X.Y.; Tu, X.L. Thallium pollution and its countermeasures. Soils 2007, 39, 528–535. [Google Scholar]
  27. Nowicka, A.M.; Krasnodebska-Ostrega, B.; Wrzosek, B.; Jastrzebska, M.; Sadowska, M.; Maćkiewicz, M.; Stojek, Z. Detection of Oxidative Damage of Synthetic Oligonucleotides Caused by Thallium(III) Complexes. Electroanalysis 2013, 26, 340–350. [Google Scholar] [CrossRef]
  28. Lin, T.-S.; Nriagu, J. Revised Hydrolysis Constants for Thallium(I) and Thallium(III) and the Environmental Implications. J. Air Waste Manag. Assoc. 1998, 48, 151–156. [Google Scholar] [CrossRef]
  29. Liu, J.; Lippold, H.; Wang, J.; Lippmann-Pipke, J.; Chen, Y. Sorption of thallium(I) onto geological materials: Influence of pH and humic matter. Chemosphere 2011, 82, 866–871. [Google Scholar] [CrossRef] [PubMed]
  30. Xiong, Y. The aqueous geochemistry of thallium: Speciation and solubility of thallium in low temperature systems. Environ. Chem. 2009, 6, 441–451. [Google Scholar] [CrossRef]
  31. Feng, H.-P.; Tang, L.; Zeng, G.-M.; Tang, J.; Deng, Y.-C.; Yan, M.; Liu, Y.-N.; Zhou, Y.-Y.; Ren, X.-Y.; Chen, S. Carbon-based core–shell nanostructured materials for electrochemical energy storage. J. Mater. Chem. A 2018, 6, 7310–7337. [Google Scholar] [CrossRef]
  32. Xiang, X.; Xia, F.; Zhan, L.; Xie, B. Preparation of zinc nano structured particles from spent zinc manganese batteries by vacuum separation and inert gas condensation. Sep. Purif. Technol. 2015, 142, 227–233. [Google Scholar] [CrossRef]
  33. Ni, J.; Kawabe, Y.; Morishita, M.; Watada, M.; Sakai, T. Improved electrochemical activity of LiMnPO4 by high-energy ball-milling. J. Power Sources 2011, 196, 8104–8109. [Google Scholar] [CrossRef]
  34. Li, K.; Li, H.; Xiao, T.; Zhang, G.; Liang, A.; Zhang, P.; Lin, L.; Chen, Z.; Cao, X.; Long, J. Zero-valent manganese nanoparticles coupled with different strong oxidants for thallium removal from wastewater. Front. Environ. Sci. Eng. 2020, 14, 34. [Google Scholar] [CrossRef]
  35. Li, H.; Li, X.; Xiao, T.; Chen, Y.; Long, J.; Zhang, G.; Zhang, P.; Li, C.; Zhuang, L.; Li, K. Efficient removal of thallium(I) from wastewater using flower-like manganese dioxide coated magnetic pyrite cinder. Chem. Eng. J. 2018, 353, 867–877. [Google Scholar] [CrossRef]
  36. Liu, W.; Zhang, P.; Borthwick, A.G.; Chen, H.; Ni, J. Adsorption mechanisms of thallium(I) and thallium(III) by titanate nanotubes: Ion-exchange and co-precipitation. J. Colloid Interface Sci. 2014, 423, 67–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zhao, Y.; Cheng, F.; Men, B.; Yi, H.E.; Wang, D. The water-sediment interfacial process of thallium. Environ. Chem. 2019, 38, 2047–2054. [Google Scholar]
  38. Chen, W.; Xiong, J.; Liu, J.; Wang, H.; Yao, J.; Liu, H.; Huangfu, X.; He, Q.; Ma, J.; Liu, C.; et al. Thermodynamic and kinetic coupling modeling for thallium(I) sorption at a heterogeneous titanium dioxide interface. J. Hazard. Mater. 2022, 428, 128230. [Google Scholar] [CrossRef]
  39. Liu, F.; Li, H.; Liao, D.; Xu, Y.; Yu, M.; Deng, S.; Zhang, G.; Xiao, T.; Long, J.; Zhang, H.; et al. Carbon quantum dots derived from the extracellular polymeric substance of anaerobic ammonium oxidation granular sludge for detection of trace Mn(vii) and Cr(vi). RSC Adv. 2020, 10, 32249–32258. [Google Scholar] [CrossRef]
  40. Liu, J.; Chen, W.; Hu, X.; Wang, H.; Zou, Y.; He, Q.; Ma, J.; Liu, C.; Chen, Y.; Huangfu, X. Effects of MnO2 crystal structure on the sorption and oxidative reactivity toward thallium(I). Chem. Eng. J. 2021, 416, 127919. [Google Scholar] [CrossRef]
  41. Song, L.; Duan, Y.; Liu, J.; Pang, H. Transformation between nanosheets and nanowires structure in MnO2 upon providing Co2+ ions and applications for microwave absorption. Nano Res. 2020, 13, 95–104. [Google Scholar] [CrossRef]
  42. Wan, S.; Ma, M.; Lv, L.; Qian, L.; Xu, S.; Xue, Y.; Ma, Z. Selective capture of thallium(I) ion from aqueous solutions by amorphous hydrous manganese dioxide. Chem. Eng. J. 2014, 239, 200–206. [Google Scholar] [CrossRef]
  43. Chen, M.; Wu, P.; Li, S.; Yang, S.; Lin, Z.; Dang, Z. The effects of interaction between vermiculite and manganese dioxide on the environmental geochemical process of thallium. Sci. Total. Environ. 2019, 669, 903–910. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, M.; Wu, P.; Yu, L.; Liu, S.; Ruan, B.; Hu, H.; Zhu, N.; Lin, Z. FeOOH-loaded MnO2 nano-composite: An efficient emergency material for thallium pollution incident. J. Environ. Manag. 2017, 192, 31–38. [Google Scholar] [CrossRef]
  45. Kajitvichyanukul, P.; Chenthamarakshan, C.R.; Rajeshwar, K.; Qasim, S.R. Photocatalytic reactivity of thallium(I) species in aqueous suspensions of titania. J. Electroanal. Chem. 2002, 519, 25–32. [Google Scholar] [CrossRef]
  46. Liu, Y.; Wang, L.; Wang, X.; Huang, Z.; Xu, C.; Yang, T.; Zhao, X.; Qi, J.; Ma, J. Highly efficient removal of trace thallium from contaminated source waters with ferrate: Role of in situ formed ferric nanoparticle. Water Res. 2017, 124, 149–157. [Google Scholar] [CrossRef] [PubMed]
  47. Coup, K.M.; Swedlund, P.J. Demystifying the interfacial aquatic geochemistry of thallium(I): New and old data reveal just a regular cation. Chem. Geol. 2015, 398, 97–103. [Google Scholar] [CrossRef]
  48. Yantasee, W.; Warner, C.L.; Sangvanich, T.; Addleman, R.S.; Carter, T.G.; Wiacek, R.J.; Fryxell, G.E.; Timchalk, C.; Warner, M.G. Removal of Heavy Metals from Aqueous Systems with Thiol Functionalized Superparamagnetic Nanoparticles. Environ. Sci. Technol. 2007, 41, 5114–5119. [Google Scholar] [CrossRef] [PubMed]
  49. Feng, H.; Yu, J.; Tang, J.; Tang, L.; Liu, Y.; Lu, Y.; Wang, J.; Ni, T.; Yang, Y.; Yi, Y. Enhanced electro-oxidation performance of FeCoLDH to organic pollutants using hydrophilic structure. J. Hazard. Mater. 2022, 430, 128464. [Google Scholar] [CrossRef]
  50. Tang, J.; Wang, J.; Tang, L.; Feng, C.; Zhu, X.; Yi, Y.; Feng, H.; Yu, J.; Ren, X. Preparation of floating porous g-C3N4 photocatalyst via a facile one-pot method for efficient photocatalytic elimination of tetracycline under visible light irradiation. Chem. Eng. J. 2022, 430, 132669. [Google Scholar] [CrossRef]
  51. Feng, H.; Yu, J.; Tang, L.; Wang, J.; Dong, H.; Ni, T.; Tang, J.; Tang, W.; Zhu, X.; Liang, C. Improved hydrogen evolution activity of layered double hydroxide by optimizing the electronic structure. Appl. Catal. B Environ. 2021, 297, 120478. [Google Scholar] [CrossRef]
  52. Feng, H.; Yu, J.; Tang, L.; Zeng, G.; Tang, W.; Wang, J.; Luo, T.; Peng, B.; Song, B.; Wang, L.; et al. Tuning Electron Density Endows Fe1–xCoxP with Exceptional Capability of Electrooxidation of Organic Pollutants. Environ. Sci. Technol. 2019, 53, 13878–13887. [Google Scholar] [CrossRef]
  53. Zhang, L.; Yang, Y.; Wu, S.; Xia, F.; Han, X.; Xu, X.; Deng, S.; Jiang, Y. Insights into the synergistic removal mechanisms of thallium(I) by biogenic manganese oxides in a wide pH range. Sci. Total. Environ. 2022, 831, 154865. [Google Scholar] [CrossRef]
  54. Li, X.-Q.; Elliott, D.W.; Zhang, W.-X. Zero-Valent Iron Nanoparticles for Abatement of Environmental Pollutants: Materials and Engineering Aspects. Crit. Rev. Solid State Mater. Sci. 2006, 31, 111–122. [Google Scholar] [CrossRef]
  55. Feng, H.; Tang, L.; Zeng, G.; Yu, J.; Deng, Y.; Zhou, Y.; Wang, J.; Feng, C.; Luo, T.; Shao, B. Electron density modulation of Fe1-xCoxP nanosheet arrays by iron incorporation for highly efficient water splitting. Nano Energy 2020, 67, 104174. [Google Scholar] [CrossRef]
  56. Liu, S.; Feng, H.; Tang, L.; Dong, H.; Wang, J.; Yu, J.; Feng, C.; Liu, Y.; Luo, T.; Ni, T. Removal of Sb(III) by sulfidated nanoscale zerovalent iron: The mechanism and impact of environmental conditions. Sci. Total. Environ. 2020, 736, 139629. [Google Scholar] [CrossRef] [PubMed]
  57. Deng, H.; Zhang, Z.; Xu, N.; Chen, Y.; Wu, H.; Liu, T.; Ye, H. Removal of Tl(I) Ions From Aqueous Solution Using Fe@Fe2O3Core-Shell Nanowires. CLEAN Soil Air Water 2016, 44, 1214–1224. [Google Scholar] [CrossRef]
  58. Tang, L.; Feng, H.; Tang, J.; Zeng, G.; Deng, Y.; Wang, J.; Liu, Y.; Zhou, Y. Treatment of arsenic in acid wastewater and river sediment by Fe@Fe2O3 nanobunches: The effect of environmental conditions and reaction mechanism. Water Res. 2017, 117, 175–186. [Google Scholar] [CrossRef]
  59. Li, Y.; Li, H.; Liu, F.; Zhang, G.; Xu, Y.; Xiao, T.; Long, J.; Chen, Z.; Liao, D.; Zhang, J.; et al. Zero-valent iron-manganese bimetallic nanocomposites catalyze hypochlorite for enhanced thallium(I) oxidation and removal from wastewater: Materials characterization, process optimization and removal mechanisms. J. Hazard. Mater. 2020, 386, 121900. [Google Scholar] [CrossRef]
  60. Li, H.; Chen, Y.; Long, J.; Li, X.; Jiang, D.; Zhang, P.; Qi, J.; Huang, X.; Liu, J.; Xu, R.; et al. Removal of thallium from aqueous solutions using Fe-Mn binary oxides. J. Hazard. Mater. 2017, 338, 296–305. [Google Scholar] [CrossRef]
  61. Zou, Y.; Li, Q.; Tao, T.; Ye, P.; Zhang, P.; Liu, Y. Fe-Mn binary oxides activated aluminosilicate mineral and its Tl(I) removal by oxidation, precipitation and adsorption in aqueous. J. Solid State Chem. 2021, 303, 122383. [Google Scholar] [CrossRef]
  62. Li, L.; Liu, C.; Ma, R.; Yu, Y.; Chang, Z.; Zhang, X.; Yang, C.; Chen, D.; Yu, Y.; Li, W.; et al. Enhanced oxidative and adsorptive removal of thallium(I) using Fe3O4@TiO2 decorated RGO nanosheets as persulfate activator and adsorbent. Sep. Purif. Technol. 2021, 271, 118827. [Google Scholar] [CrossRef]
  63. Mohan, D.; Pittman, C.U., Jr. Arsenic removal from water/wastewater using adsorbents—A critical review. J. Hazard. Mater. 2007, 142, 1–53. [Google Scholar] [CrossRef] [PubMed]
  64. Sinha, R.; Gupta, A.K.; Ghosal, P.S. A review on Trihalomethanes and Haloacetic acids in drinking water: Global status, health impact, insights of control and removal technologies. J. Environ. Chem. Eng. 2021, 9, 106511. [Google Scholar] [CrossRef]
  65. Haldar, D.; Duarah, P.; Purkait, M.K. MOFs for the treatment of arsenic, fluoride and iron contaminated drinking water: A review. Chemosphere 2020, 251, 126388. [Google Scholar] [CrossRef]
  66. Zhang, L.; Huang, T.; Zhang, M.; Guo, X.; Yuan, Z. Studies on the capability and behavior of adsorption of thallium on nano-Al2O3. J. Hazard. Mater. 2008, 157, 352–357. [Google Scholar] [CrossRef]
  67. Şenol, Z.M.; Ulusoy, U. Thallium adsorption onto polyacryamide–aluminosilicate composites: A Tl isotope tracer study. Chem. Eng. J. 2010, 162, 97–105. [Google Scholar] [CrossRef]
  68. Chen, W.; Huangfu, X.; Xiong, J.; Liu, J.; Wang, H.; Yao, J.; Liu, H.; He, Q.; Ma, J.; Liu, C.; et al. Retention of thallium(I) on goethite, hematite, and manganite: Quantitative insights and mechanistic study. Water Res. 2022, 221, 118836. [Google Scholar] [CrossRef]
  69. Biaduń, E.; Sadowska, M.; Ospina-Alvarez, N.; Krasnodębska-Ostręga, B. Direct speciation analysis of thallium based on solid phase extraction and specific retention of a Tl(III) complex on alumina coated with sodium dodecyl sulfate. Microchim. Acta 2015, 183, 177–183. [Google Scholar] [CrossRef] [Green Version]
  70. Liu, Y.; Zhang, J.; Huang, H.; Huang, Z.; Xu, C.; Guo, G.; He, H.; Ma, J. Treatment of trace thallium in contaminated source waters by ferrate pre-oxidation and poly aluminium chloride coagulation. Sep. Purif. Technol. 2019, 227, 115663. [Google Scholar] [CrossRef]
  71. Kajitvichyanukul, P.; Chenthamarakshan, C.; Rajeshwar, K.; Qasim, S. Adsorption of Thallium(I) Ions on Titania Particle Surfaces in Aqueous Media. Adsorpt. Sci. Technol. 2003, 21, 217–228. [Google Scholar] [CrossRef]
  72. Zhang, L.; Huang, T.; Liu, N.; Liu, X.; Li, H. Sorption of thallium(III) ions from aqueous solutions using titanium dioxide nanoparticles. Microchim. Acta 2009, 165, 73–78. [Google Scholar] [CrossRef]
  73. Ferro, S.; Donatoni, M.; De Battisti, A.; Andreev, V.N. Adsorption of thallium cations on RuO2–TiO2 electrodes. J. Appl. Electrochem. 2007, 37, 1389–1394. [Google Scholar] [CrossRef]
  74. Zhang, G.; Fan, F.; Li, X.; Qi, J.; Chen, Y. Superior adsorption of thallium(I) on titanium peroxide: Performance and mechanism. Chem. Eng. J. 2018, 331, 471–479. [Google Scholar] [CrossRef]
  75. Murray, J.W. The surface chemistry of hydrous manganese dioxide. J. Colloid Interface Sci. 1974, 46, 357–371. [Google Scholar] [CrossRef]
  76. Huangfu, X.; Jiang, J.; Lu, X.; Wang, Y.; Liu, Y.; Pang, S.-Y.; Cheng, H.; Zhang, X.; Ma, J. Adsorption and Oxidation of Thallium(I) by a Nanosized Manganese Dioxide. Water Air Soil Pollut. 2015, 226, 1–9. [Google Scholar] [CrossRef]
  77. Huangfu, X.; Ma, C.; Ma, J.; He, Q.; Yang, C.; Jiang, J.; Wang, Y.; Wu, Z. Significantly improving trace thallium removal from surface waters during coagulation enhanced by nanosized manganese dioxide. Chemosphere 2017, 168, 264–271. [Google Scholar] [CrossRef]
  78. Zhang, H.; Qi, J.; Liu, F.; Wang, Z.; Ma, X.; He, D. One-pot synthesis of magnetic Prussian blue for the highly selective removal of thallium(I) from wastewater: Mechanism and implications. J. Hazard. Mater. 2022, 423, 126972. [Google Scholar] [CrossRef]
  79. Wang, N.; Su, Z.; Deng, N.; Qiu, Y.; Ma, L.; Wang, J.; Chen, Y.; Hu, K.; Huang, C.; Xiao, T. Removal of thallium(I) from aqueous solutions using titanate nanomaterials: The performance and the influence of morphology. Sci. Total. Environ. 2020, 717, 137090. [Google Scholar] [CrossRef]
  80. Pan, B.; Wan, S.; Zhang, S.; Guo, Q.; Xu, Z.; Lv, L.; Zhang, W. Recyclable polymer-based nano-hydrous manganese dioxide for highly efficient Tl(I) removal from water. Sci. China Chem. 2014, 57, 763–771. [Google Scholar] [CrossRef]
  81. Zhang, P.; Chen, Y.H.; Liang, M.H. Reflectance Spectroscopic Characterization of Thallium Wastewater Treatment by Pyrite and Pyrite Slag. Guang Pu Xue Yu Guang Pu Fen Xi Guang Pu 2007, 27, 747–749. [Google Scholar]
  82. Marc, F.; Benedetti Chris, J.; Milne David, G.; Kinniburgh, W. Metal Ion Binding to Humic Substances: Application of the Non-Ideal Competitive Adsorption Model. Environ. Sci. Technol. 1995, 29, 446–457. [Google Scholar]
  83. Law, S.; Turner, A. Thallium in the hydrosphere of south west England. Environ. Pollut. 2011, 159, 3484–3489. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The effects of thallium entering the environment and the human nervous system.
Figure 1. The effects of thallium entering the environment and the human nervous system.
Ijerph 20 03829 g001
Figure 2. Five crystal structures of MnO2 (δ-, γ-, α-, β- and λ-MnO2): (a) α-MnO2, (b) γ-MnO2, (c) β-MnO2, (d) δ-MnO2, and (e) λ-MnO2 [40].
Figure 2. Five crystal structures of MnO2 (δ-, γ-, α-, β- and λ-MnO2): (a) α-MnO2, (b) γ-MnO2, (c) β-MnO2, (d) δ-MnO2, and (e) λ-MnO2 [40].
Ijerph 20 03829 g002
Figure 3. Mechanism of adsorption of thallium by 3-ZVIMn adsorbent [59].
Figure 3. Mechanism of adsorption of thallium by 3-ZVIMn adsorbent [59].
Ijerph 20 03829 g003
Figure 4. Schematic diagram of thallium adsorption mechanism of S-ZVI material [56].
Figure 4. Schematic diagram of thallium adsorption mechanism of S-ZVI material [56].
Ijerph 20 03829 g004
Figure 5. Mechanism diagram for the adsorption of Tl(I) by MnO2@pyrite cinder [35].
Figure 5. Mechanism diagram for the adsorption of Tl(I) by MnO2@pyrite cinder [35].
Ijerph 20 03829 g005
Figure 6. SEM views of polyacrylamide(PAAm) (a), polyacrylamide-bentonite(PAAm-B) (b), and polyacrylamide-zeolite(PAAm-Z) (c) [67].
Figure 6. SEM views of polyacrylamide(PAAm) (a), polyacrylamide-bentonite(PAAm-B) (b), and polyacrylamide-zeolite(PAAm-Z) (c) [67].
Ijerph 20 03829 g006
Figure 7. Schematic diagram of the adsorption of Tl(I) an Tl(III) by TNTs [36].
Figure 7. Schematic diagram of the adsorption of Tl(I) an Tl(III) by TNTs [36].
Ijerph 20 03829 g007
Figure 8. Electron micrographs of TNTs: (a) Before adsorption of Tl; (b) After absorption of thallium Tl(I); and (c) After absorption of thallium Tl(III) [36].
Figure 8. Electron micrographs of TNTs: (a) Before adsorption of Tl; (b) After absorption of thallium Tl(I); and (c) After absorption of thallium Tl(III) [36].
Ijerph 20 03829 g008
Figure 9. Tl removal mechanism by titanium peroxide [74].
Figure 9. Tl removal mechanism by titanium peroxide [74].
Ijerph 20 03829 g009
Figure 10. Schematic diagram of the effect of pH on the adsorption of Tl(I) by metal oxides: (a) At low pH; (b) At high pH [21].
Figure 10. Schematic diagram of the effect of pH on the adsorption of Tl(I) by metal oxides: (a) At low pH; (b) At high pH [21].
Ijerph 20 03829 g010
Table 1. Manganese oxide removal thallium performance and advantages and disadvantages.
Table 1. Manganese oxide removal thallium performance and advantages and disadvantages.
Metal OxidesDosageThallium Concentration in the Tested WaterPerformanceAdvantagesDisadvantagesRef.
nMnO20.05 mmol·L−13.5 mg/L672 mg/gHigh removal efficiencyPoor reproducibility[3]
HMO50 mg0.5 mmol·L−1348.84 mg/gWith reusable propertiesHighly influenced by environmental conditions[42]
VER-MnO220 μg/L144.29 mg/gLess influenced by environmental conditionsPoor removal ability[43]
Zero-valent manganese nanoparticles2 g/L10 mg/L95.7%Less influenced by environmental conditionsPoor reproducibility[34]
Flower-like manganese dioxide coated magnetic pyrite cinder0.5 g/L10 mg/L320 mg/gWith reusable propertiesHighly influenced by environmental conditions[35]
FeOOH-loaded MnO2 nano-composite10 mg0~150 mg/L450 mg/gWith reusable propertiesHighly influenced by environmental conditions[44]
δ-MnO2500 μmol·L−125 mg/L581 mg/gHigh removal efficiencyHighly influenced by environmental conditions[41]
MnO2@slag1 g/L10 mg/L99.5%High removal efficiencyPoor reproducibility[35]
Table 3. Aluminum oxide removal thallium performance and advantages and disadvantages.
Table 3. Aluminum oxide removal thallium performance and advantages and disadvantages.
Metal OxidesDosageThallium Concentration in the Tested WaterPerformanceAdvantagesDisadvantagesRef.
Nano-Al2O30.03 g10 mg/L6.28 mg/gLess influenced by environmental conditionsPoor reproducibility[66]
PAAm-B0.1 g0.005 mol·L−1197.88 mg/g-Tl(I), 32.64 mg/g-Tl(III)High removal efficiencyHighly influenced by environmental conditions[67]
PAAm-Z0.1 g0.005 mol·L−1337.4 mg/g-Tl(I), 73.44 mg/g-Tl(III)High removal efficiencyHighly influenced by environmental conditions[67]
Ferrate pre-oxidation and poly aluminum chloride coagulation0.5–4 mg0.76 μg/L87%Less influenced by environmental conditionsPoor reproducibility[70]
Table 4. Titanium oxide removal thallium performance and its advantages and disadvantages.
Table 4. Titanium oxide removal thallium performance and its advantages and disadvantages.
Metal OxidesDosageThallium Concentration in the Tested WaterPerformanceAdvantagesDisadvantagesRef.
Titanium iron magnetic0.1 g/L12.5 mg/L111.3 mg/gLess influenced by environmental conditionsPoor reproducibility[58]
Fe3O4@TiO2 decorated RGO nanosheets0.2 g/L_673.2 mg/gHigh removal efficiencyHighly influenced by environmental conditions[62]
TNTs0.2 g/L100 mg/L709.2 mg/gHigh removal efficiencyHighly influenced by environmental conditions[36]
Titanium peroxide0.2 g/L0.046–20 mg/L412 mg/gLess influenced by environmental conditionsPoor reproducibility[74]
Table 5. Factors affecting the effectiveness of metal oxide removal.
Table 5. Factors affecting the effectiveness of metal oxide removal.
Metal OxidesEffect of pH ValueEffect of Co-Existing IonsEffect of Organic MattersRef.
nMnO2pH < 7, little influenceCa2+, Mg2+HA of 3 mg/L[3]
HMOlittle influenceCa2+, Mg2+NA[42]
VER-MnO2little influenceCa2+NA[43]
Zero-valent manganese nanoparticleslittle influenceNANA[34]
Flower-like manganese dioxide coated magnetic pyrite cinderpH < 12, little influenceCa2+, Mg2+, Na+EDTA[35]
FeOOH-loaded MnO2 nano-compositelittle influenceCa2+HA[44]
δ-MnO2little influenceCa2+, Na+NA[41]
MnO2@slaglittle influenceCa2+HA, FA[35]
Nano-Al2O3little influenceNAHA, FA, EDTA and DTPA[66]
PAAm-Blittle influenceNANA[67]
PAAm-Zlittle influenceCd(II), Cu(II), Pb(II) and Mn(II)NA[67]
Ferrate pre-oxidation and poly aluminum chloride coagulationlittle influenceFe2+, Pb2+, Zn2+NA[70]
Magnetic Prussian bluelittle influenceNANA[53]
Fe@Fe2O3 Core–ShellNAMg2+, PO43, Ca2+, Na+, ClNA[57]
3-ZVIMnlittle influenceFe2+, Fe3+NA[59]
Fe-Mn binary oxideslittle influenceCu2+, Zn2+NA[60]
Titanium iron magneticlittle influenceNa+, NO3NA[58]
Fe-Mn-binary-oxides-activated aluminosilicate minerallittle influenceMg2+, Ca2+HA[61]
Fe3O4@TiO2-decorated RGO nanosheetslittle influenceno influenceEDTA and DTPA[62]
TNTlittle influenceNa+, K+, Ca2+NA[36]
Titanium peroxidelittle influenceNANA[74]
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

Ren, X.; Feng, H.; Zhao, M.; Zhou, X.; Zhu, X.; Ouyang, X.; Tang, J.; Li, C.; Wang, J.; Tang, W.; et al. Recent Advances in Thallium Removal from Water Environment by Metal Oxide Material. Int. J. Environ. Res. Public Health 2023, 20, 3829. https://doi.org/10.3390/ijerph20053829

AMA Style

Ren X, Feng H, Zhao M, Zhou X, Zhu X, Ouyang X, Tang J, Li C, Wang J, Tang W, et al. Recent Advances in Thallium Removal from Water Environment by Metal Oxide Material. International Journal of Environmental Research and Public Health. 2023; 20(5):3829. https://doi.org/10.3390/ijerph20053829

Chicago/Turabian Style

Ren, Xiaoyi, Haopeng Feng, Mengyang Zhao, Xin Zhou, Xu Zhu, Xilian Ouyang, Jing Tang, Changwu Li, Jiajia Wang, Wangwang Tang, and et al. 2023. "Recent Advances in Thallium Removal from Water Environment by Metal Oxide Material" International Journal of Environmental Research and Public Health 20, no. 5: 3829. https://doi.org/10.3390/ijerph20053829

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