Next Article in Journal
Optimizing Liquid Electron Ionization Interface to Boost LC-MS Instrumental Efficiency
Previous Article in Journal
Nonylphenol Removal from Water and Wastewater with Alginate-Activated Carbon Beads
Previous Article in Special Issue
Irradiated Gao Miao Zi Bentonite for Uranium Retention: Performance and Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Two-Dimensional Thiotitanate Ion Exchanger with High Cs+ Removal Performance

1
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
2
Fujian College, University of Chinese Academy of Sciences, Fuzhou 350002, China
3
HTA Co., Ltd., Beijing 102413, China
4
Fujian Province Joint Innovation Key Laboratory of Fuel and Materials in Clean Nuclear Energy System, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Separations 2025, 12(5), 104; https://doi.org/10.3390/separations12050104
Submission received: 16 March 2025 / Revised: 12 April 2025 / Accepted: 18 April 2025 / Published: 22 April 2025
(This article belongs to the Special Issue Separation Technology for Metal Extraction and Removal)

Abstract

:
137Cs is a persistent β/γ-emitter (t1/2 = 30.1 years) generated from 235U and 239Pu fission. It is a critical challenge to efficiently capture 137Cs+ for nuclear waste management due to its high solubility, environmental mobility, and propensity for biological accumulation. Herein, we prepare a two-dimensional (2D) thiotitanate Rb0.32TiS2·0.75H2O (denoted Rb-TiS2) using a special molten salt synthesis method, “Mg + RbCl”. Rb-TiS2 can selectively capture Cs+ from aqueous solutions. Its structure features a flexible anionic thiotitanate layer with Rb+ as counter ions located at the interlayer spaces. As an ion exchanger, it possesses high adsorption capacity (qmCs = 232.70 mg·g−1), rapid kinetics (the removal rate R > 72% within 10 min), and a wide pH tolerance range (pH = 4–12) for Cs⁺ adsorption. Through a single-crystal X-ray structural analysis, we elucidated the mechanism of Cs⁺ capture, revealing the ion exchange pathways between Cs⁺ and Rb+ in Rb-TiS2. This work not only provides an important reference for the synthesis of transition metal sulfides with alkali metal cations but also proves the application prospect of transition metal sulfides in radionuclide remediation.

Graphical Abstract

1. Introduction

Nuclear energy, with its high energy density and low carbon emissions, is a key solution for addressing global energy demands and climate change. However, the radioactive waste generated simultaneously presents significant environmental and public health challenges [1,2,3,4,5]. Among the fission products of spent fuel, 137Cs emerges as one of the particularly problematic radionuclides due to its long half-life (30.1 years), high decay heat, and potent β/γ radiation emissions [6]. It predominantly exists as the form of Cs+ cation. Its high-water solubility facilitates rapid migration through environmental media and entry into ecosystems via the hydrological cycle, posing a serious threat to environmental security and human health [7]. In addition, as alkali metal ions, Cs+ ions have similar chemical properties to K+ ions, which makes it easy to accumulate in living organisms, increasing the risk of radiation exposure, especially the potential damage to bone [8], thyroid [9], and germ cells [10]. Historical nuclear accidents, such as the Chernobyl accident, Goiânia radiation accident, and Fukushima nuclear accident, have amply demonstrated that large-scale releases of radiocesium can result in long-term environmental degradation and serious public health effects [11,12]. Further, 137Cs can be applied in specific fields; for example, industrial irradiation [13], medical radiation therapy [14], radioisotope thermoelectric generators [15], etc. Therefore, the rapid and efficient removal of cesium from complex nuclear wastewater is of vital importance.
At present, the common methods for Cs+ ion removal include chemical precipitation, solvent extraction, membrane separation, and ion exchange [16]. Among these, chemical precipitation suffers from poor selectivity despite its operational simplicity [17]. Solvent extraction offers high efficiency but has prohibitive costs [18]. Membrane separation enables continuous operation yet faces challenges such as fouling and clogging [19]. By contrast, ion exchange techniques have garnered significant attention due to their high selectivity, operational simplicity, and no secondary pollution [20]. Ion exchange materials with exchangeable cations and specific recognition sites can efficiently adsorb Cs⁺ [20,21]. Particularly, metal sulfide ion exchange materials (MSIEs) have attracted research interest due to their numerous soft basic S2− sites and flexible framework for Cs+ capture [22,23], such as KMS series (KMS-1: K2xMnxSn3−xS6 [24], KMS-2: K2xMgxSn3−xS6 (x = 0.5–1), etc. [25]), KTS series (KTSS: K1.93Ti0.22Sn3S6.43 [26], KTS-3: K2xsn4−xS8−x (x = 0.65–1) [27], KTS-2: K2Sn2S5 [28], etc.), and FJSM series (FJSM-SnS: (Me2NH2)4/3(Me3NH)2/3Sn3S7·1.25H2O [29], FJSM-GAS-1: [Me2NH2]2[Ga2Sb2S7]·H2O [30], etc.).
Nevertheless, the current research on MSIEs is predominantly focused on those based on main group metal ions such as In3+, Sn4+, and Sb3+ [20,22,31]. However, transition metal sulfides based on IVB group ions that share similar valence states and coordination modes with Sn have been scarcely explored as MSIEs. Thus far, the only example is a Ti-doped ion-exchange material that contains a trace amount of titanium [26]. Furthermore, it is notable that the alkali metal ions as the exchangeable cations in reported MSIEs are predominantly Na⁺ and K⁺, while the alkali ion Rb+ hardly serves as an exchangeable cation in MSIEs. According to the reported K3Rb3Zn4Sn3Se13 [32] and KCsRb1.2(NH4)2.71Sn5Zn4S17 [33], however, Rb+ with a larger ionic radius also has mobile properties. Additionally, the crystal structures of some ion exchange materials after trapping radioactive ions remain unresolved [34,35], hindering the in-depth exploration of the structure–property relationships and impeding the design and synthesis of advanced materials. Therefore, expanding the diversity of exchangeable cations and designing novel transition metal sulfides are critical for developing advanced crystalline MSIEs with efficient Cs⁺ capture properties.
Herein, we report a two-dimensional (2D) thiotitanate Rb0.32TiS2·0.75H2O (Rb-TiS2) synthesized via a unique “Mg + RbCl” molten salt method [36], which overcomes the operational complexity and safety limitations of traditional solid-state synthesis for thiotitanates while enabling precise single-crystal structure determination. Rb-TiS2 exhibits good adsorption performance, including a high adsorption capacity of 232.70 mg·g−1, rapid kinetics (81.61% removal rate within 5 min), excellent selectivity, and good adsorption–desorption behavior. Notably, we also obtained the crystal structure of Cs0.28TiS2·0.6H2O (denoted as Cs-TiS2) after Cs+ capture by Rb-TiS2, providing atomic-level visualization of the radioactive ion capture process. Then, according to elemental chemical analysis and other characterizations, the ion exchange mechanism of Rb-TiS2 for Cs+ capture is elucidated. The flexible and tunable layered framework of Rb-TiS2, along with its abundant and accessible S2− sites, endows its excellent Cs+ ion adsorption properties. This work not only provides a new crystalline thiotitanate ion exchange material through a unique synthetic approach but also demonstrates the great potential of Group IVB transition metal sulfides for capturing radioactive cations.

2. Materials and Methods

2.1. Materials

Ti (99.00%), S (99.50%), NaCl (Analytical Research grade (AR)), MgCl2·6H2O (AR), and HNO3 (69%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). RbCl (99.50%) was purchased from Adamas Reagent Co., Ltd. (Shanghai, China). Mg (99.5%) was purchased from Tianjin Fuchen Chemical Reagent Factory (Tianjin, China). KCl (AR) was purchased from Titan Co., Ltd. (Shanghai, China). CaCl2·2H2O (74%) was purchased from Shanghai Sili Chemical Plant (Shanghai, China). CsCl (99.99%) was purchased from Shanghai Longjin Metallic Material Co., Ltd. (Shanghai, China). NaOH (AR) was purchased from Guangfu Co., Ltd. (Tianjin, China). All the chemicals were used without further purification.

2.2. Methods

2.2.1. Synthesis of Rb-TiS2

The synthesis was performed by sealing a stoichiometric mixture of Ti (2 mmol), S (6 mmol), RbCl (4 mmol), and Mg (1 mmol) in an evacuated quartz tube (outer diameter 12 mm; inner diameter 10 mm) using an oxygen–hydrogen flame. The sealed ampoule was subjected to a programmed thermal treatment in a muffle furnace with the following temperature program: (i) gradual heating from room temperature to 900 °C over 24 h; (ii) maintaining at 900 °C for 48 h; (iii) controlled cooling to 550 °C at a rate of 7.3 °C·h−1; (iv) isothermal annealing at 550 °C for 10 h; and then (v) natural cooling to ambient temperature. The resulting product was thoroughly washed with deionized water and absolute ethanol, followed by air-drying, yielding lustrous black-gold plate-like crystals.

2.2.2. Characterization Techniques

Morphological and elemental analyses were conducted using a field-emission scanning electron microscope (FE-SEM, JSM-6700F, JEOL, Akishima, Japan) equipped with energy-dispersive X-ray spectroscopy (EDS) for elemental mapping. X-ray powder diffraction (PXRD) was performed using a Rigaku Miniflex II X-ray diffractometer (Rigaku, Tokyo, Japan) with Cu Kα (λ = 1.54178 Å) operating at 30 kV and 15 mA, collecting data over a 2θ range of 5–65°. Single-crystal diffraction data were collected with a Rigaku diffractometer (XtaLAB Synergy Custom, Rigaku, Tokyo, Japan) with graphite monochromatic CuKα (λ = 1.54178 Å). Thermogravimetric analysis (TGA) was carried out on a NETZSCH STA 449 F3 (Netzsch, Selb, Germany) instrument under N2 atmosphere with a heating rate of 10 °C·min−1. X-ray photoelectron spectroscopy (XPS) was conducted on an ESCALAB 250Xi (Thermo Scientific, Waltham, MA, USA) spectrometer (AlKα). The pH value of the solution was measured with a PHS-2F pH meter (PHS-2F, Shanghai, China). Elemental concentrations were quantified using inductively coupled plasma optical emission spectroscopy (ICP-OES, Thermo 7400, Thermo Scientific, Waltham, MA, USA) and inductively coupled plasma mass spectrometry (ICP-MS, Thermo XSeries II, Thermo Scientific, Waltham, MA, USA) with appropriate quality control measures, including calibration standards and replicate analyses.

2.2.3. Batch Adsorption Experiments

In the ion exchange experiments, non-radioactive Cs+ was used as a substitute for radioactive Cs+ ions, and all tests were carried out at room temperature. A typical ion exchange experiment was performed in a 25 mL polyethylene bottle, where the ion exchange material was mixed with a Cs+-containing aqueous solution at a solid-to-liquid ratio (V/m) of 1000 mL·g−1. Then, the bottle was put on the shaking table for a certain time and then taken out and stood for about 3 min. The supernatant was filtered through a 0.22 μm syringe filter and diluted with 2% HNO3 to an appropriate concentration for analysis by ICP-MS or ICP-OES. The solid sample was washed several times with deionized water and ethanol followed by natural drying, and then subjected to PXRD, EDS, and TG analyses.
Isothermal adsorption experiments: Cs+ solutions with varying concentrations were prepared using deionized water. An amount of 10 mg of Rb-TiS2 sample was added to 10 mL of the solution and oscillated for approximately 6 h. The solution was then analyzed using ICP-MS to measure the change in Cs+ concentration. The adsorption data were fitted using the Langmuir, Freundlich, and Langmuir–Freundlich models to calculate the maximum adsorption capacity.
Kinetic experiments: An amount of 50 mg of Rb-TiS2 sample was added to 50 mL of a 7 ppm Cs⁺ solution and stirred using a magnetic stirrer to ensure full solid–liquid contact. Approximately 0.3 mL of the mixture was sampled at regular intervals (2, 5, 10, 15, 20, 30, 40, 60, 120, 240, and 480 min), and the Cs+ concentration was measured. The kinetic data were fitted using the pseudo-first-order and pseudo-second-order models.
Acid–Base Stability Tests: Solutions containing Cs⁺ ions with different acidity and alkalinity were prepared using CsCl, NaOH, and HNO3 and the pH value of the solution was measured using a pH meter. An amount of 10 mg of Rb-TiS2 sample was added to 10 mL of the solution, respectively, and oscillated for approximately 6 h. The ion concentration in the solution was measured using ICP-MS. The solid sample was washed, dried, and analyzed with PXRD.
Competitive Experiments: The Cs–Na (molar ratio = 1:7.53, 1:105, 1:300 and 1:516), Cs–Ca (1:10.64) and Cs–Mg (1:4.80) solutions containing a single competitor ion were prepared by the corresponding halide salts, in which the concentration of Cs+ ions was about 6 mg·L−1. In addition, by adding a small amount of Cs+ ions in tap water, river water (Wulong River), and lake water (Qishan Lake), the polluted environment water body was simulated. An amount of 10 mg of Rb-TiS2 sample was added to each solution and stirred for 6 h. Afterward, the solutions were analyzed using ICP-MS or ICP-OES.
Adsorption–Desorption Experiments: (1) Desorption Experiments: An amount of 50 mg of the Rb-TiS2 sample was treated with 50 mL of an approximately 1000 ppm Cs+ solution to obtain the adsorbed product (denoted as Cs-TiS2). An amount of 20 mg of the Cs-TiS2 sample was immersed in 20 mL of a 0.5 M KCl and 0.5 M RbCl solution for about 6 h, respectively, followed by PXRD analysis and further ion exchange experiments. (2) Adsorption–Desorption Experiments: An amount of 500 mL Cs+ ion solution (about 20 ppm) and 500 mL of a 1 M KCl solution were prepared for adsorption and desorption experiments, respectively. After each adsorption analysis test, the liquid sample was measured using ICP-MS; the solid sample was washed with deionized water and ethanol, dried, and weighed and part of the sample was extracted for the PXRD test.

3. Results

3.1. Synthetic Optimization

Currently, alkali metal-containing titanium sulfides (e.g., K4Ti3S14 [37] and K2TiS3 [38]) are typically synthesized via traditional high-temperature solid-state methods. These approaches require the pre-preparation of polysulfide precursors using liquid ammonia and must be conducted under strict glovebox conditions, resulting in complex procedures and significant safety risks. Additionally, the intercalation method (e.g., KxTiS2 [39]) involves the initial preparation of TiS2 single crystals through iodine vapor transport, followed by secondary reactions to introduce alkali metal ions. This process involves high energy consumption and is technically demanding.
In this study, we use a synthesis strategy based on the “Mg + RbCl” molten salt method [36], which enables the direct preparation of millimeter-scale alkali-metal-ion-containing plate-like crystals. The unique advantage of this method lies in the direct generation of abundant S2− and Mg2+ ions through the redox reaction between sulfur (S) and magnesium (Mg). Excess S2− ions provide the thermodynamic driving force for the formation of alkali metal-containing titanium sulfide. Using this strategy, we successfully synthesized Rb-TiS2 crystals (Figure 1a). Compared to traditional methods, this approach eliminates the need for liquid ammonia, cumbersome glovebox operations, and multi-step synthesis, significantly enhancing safety and efficiency.
To further validate the critical role of Mg, we conducted control experiments under identical reaction conditions without Mg. Interestingly, even in the presence of excess RbCl without Mg, only golden TiS₂ crystals (Figure 1b) were obtained, rather than the target product. The corresponding PXRD (Figure 1c) and EDS data (Figure S1) confirm the indispensable role of Mg in the synthesis of Rb-TiS2.

3.2. Structural Description and Removal Mechanism

Rb0.32TiS2·0.75H2O (Rb-TiS2, Tables S1 and S2) exhibits significant differences in structure compared with the previously reported RbTiS2 [40]. Rb-TiS2 crystallizes in the space group P63/mmc (a = b = 3.4210(1) Å, c = 17.4272(10) Å), while the anhydrous RbTiS2 (ICSD#: 77990) crystallizes in the space group R3m (a = b = 3.43(1) Å, c = 24.2(1) Å). A detailed comparison of their powder XRD patterns is provided in the Supporting Information (Figure S2). In Rb-TiS2, [TiS6] octahedra share edges to form a two-dimensional anionic layer of [TiS2]n0.32n (Figure 2a). The adjacent layers exhibit different orientations, which are stacked in an ABAB sequence, forming a staggered arrangement (Figure 2b). The interlayer space is occupied by disordered Rb+ ions and lattice water molecules. Rb+ ions are located at two distinct sites of Rb(1) and O(1)/Rb(2) (Figure S3). The Cs⁺ adsorbed product, Cs0.28TiS2·0.60H2O (Cs-TiS2, Tables S1 and S2), also crystallizes in the P63/mmc space group (a = b = 3.4177(2) Å, c = 17.8784(17) Å), retaining the anionic layered structure of pristine Rb-TiS2, while Rb+ ions are replaced by Cs+ ions. Notably, the interlayer distance increases from 5.864 Å in Rb-TiS2 to 6.087 Å in Cs-TiS2 after Cs+ capture (Figure 2c), likely due to the larger ionic radius of Cs+ (ionic radius: 1.69 Å) compared to that of Rb+ (ionic radius: 1.48 Å) [41]. This confirms the 2D flexible layered structure of Rb-TiS2.

3.3. Characterization of Rb-TiS2 and Cs-TiS2

TG shows that the actual mass loss of Rb-TiS2 and Cs-TiS2 at 200 °C is 7.85% and 6.97%, respectively (Figure S4), which is close to the theoretical values of loss of their corresponding lattice water molecules in the structure (8.83% and 6.76%). Experimental PXRD patterns of Rb-TiS2 and Cs-TiS2 agree with the simulated ones from the corresponding single-crystal X-ray structures (Figure 1c and Figure S5), respectively. Furthermore, the (002) basal Bragg peak positions of Cs-TiS2 shift to lower angles compared to that of Rb-TiS2 (Figure 3a), consistent with the observed expansion of interlayer distances. EDS and elemental mapping analyses reveal that both Rb-TiS2 and Cs-TiS2 maintain well-defined crystal morphologies and uniform elemental distributions (Figure 3b). In the XPS spectrum of Rb-TiS2, characteristic peaks of Rb 3d5/2 and Rb 3d3/2 are observed at 109.4 eV and 110.9 eV, respectively (Figure 3c,d). By contrast, these Rb peaks disappear in the XPS spectrum of Cs-TiS2, while new peaks corresponding to Cs 3d5/2 and Cs 3d3/2 appear at 724.35 eV and 738.35 eV, respectively (Figure 3e). In summary, single-crystal analysis, PXRD, EDS, and XPS results collectively demonstrate that the Cs+ capture mechanism in Rb-TiS2 is ion exchange between Rb+ and Cs+ ions. The 2D flexible layered structure of Rb-TiS2 contains abundant soft basic S2− sites, which exhibit a stronger affinity for Cs+ ion than Rb+ ion according to the Lewis acid–base theory [42], enabling the effective capture of Cs+.

3.4. Adsorption Isotherm Study

The adsorption capacity (q, Equation (S1)), a critical parameter for evaluating the performance of an adsorbent, was determined through isothermal adsorption experiments to assess the maximum Cs⁺ uptake by Rb-TiS2. The experimental data are summarized in Table S3. The adsorption isotherm data are fitted using the Langmuir (Equation (S2)), Freundlich (Equation (S3)), and Langmuir–Freundlich models (Equation (S4)) [43], and the corresponding adsorption constants and parameters are listed in Table S4. The Langmuir–Freundlich model exhibits the highest correlation coefficient (R2 = 0.993), outperforming the Langmuir (R2 = 0.991) and Freundlich (R2 = 0.903) models, indicating that the adsorption process of Cs⁺ on Rb-TiS2 is best described by the Langmuir–Freundlich model (Figure 4a). Based on this model, the maximum Cs+ adsorption capacity of Rb-TiS2 is calculated to be 232.70 mg·g−1, which surpasses that of many commonly reported Cs+ adsorbents, such as sulfide KMS-1 (226 mg·g−1) [24], oxide NaFeTiO (52.8 mg·g−1) [44], MOF Nd-BTC (86 mg·g−1) [45], zeolite (Turkish samples: 89.18 mg·g−1) [46], and commercial AMP-PAN (ammonium molybdophosphate-polyacrylonitrile) (81 mg·g−1) [47] (the more compounds are compared in Table S5).

3.5. Adsorption Kinetics Studies

Adsorption kinetics is a crucial parameter for evaluating the adsorption rate of an adsorbent, whereas rapid kinetics is essential to minimize the adverse effects of radiation exposure. The kinetic data for Rb-TiS2 at room temperature are summarized in Table S6. Remarkably, Rb-TiS2 achieves a removal efficiency (R, Equation (S5)) of 72.16% within 10 min and reaches adsorption equilibrium in approximately 40 min (Figure 4b). The kinetic data are fitted using the pseudo-first-order (Equation (S6)) and pseudo-second-order models (Equation (S7)); the latter exhibits a higher correlation coefficient (R2 = 0.997, Figure 4c and Table S7), indicating that the adsorption process is predominantly governed by chemisorption. The adsorption kinetics of Rb-TiS2 significantly outperforms that of many reported adsorbents, such as KMS-1/r-GO (34% in 1 h) [48], CdSnSe-1 (71.06% in 4 h) [49], and K-SGU-45 (28% in 15 h) [50].

3.6. Effect of pH on the Cs+ Adsorption

Given the diverse origins of radioactive waste liquids and the significant pH variations among different waste streams, we systematically evaluated the pH-dependent performance of Rb-TiS2. The experimental results are summarized in Table S8 and the PXRD results are shown in Figure S6a. Within the pH range of 3–12, Rb-TiS2 maintains good Cs⁺ adsorption performance, with distribution coefficients (KdCs, Equation (S8)) consistently exceeding 103 mL·g−1 (Figure 4d). However, at pH values below 3, the crystal color of Rb-TiS2 undergoes a noticeable change (Figure S6b), accompanied by an increased leaching of Ti (Figure S6c).

3.7. Competitive Adsorption Studies

Considering that Cs+ ions in actual waste liquids often coexist with various interfering ions, we conducted competitive adsorption experiments with individual competing Na+, Ca2+, and Mg2+ ions. The data are summarized in Table S9. When individual competitive Na+, Ca2+, and Mg2+ ions exist in a small amount, Rb-TiS2 can still efficiently remove Cs+ ions, and its KdCs and RCs values are greater than 103 mL·g−1 and 80%, respectively (Figure 5a). Further, the effect of a high concentration of Na+ ions on Cs+ removal is investigated (Table S10). As illustrated in Figure 5b, even when the Cs+/Na+ molar ratio reaches 1:105, the separation factor (SFCs/Na, Equation (S9)) remains at a high level of 138. And at Cs+/Na+ molar ratios of 1:300, KdCs values of Rb-TiS2 could still reach the level of 103 mL·g−1, whereas in the Cs+/Na+ molar ratio range of 1:7.53–1:516, the RNa are always at very low levels (RNa < 10%, Table S10). This shows that Rb-TiS2 has a high selectivity for Cs+ ion and a good Cs/Na separation effect. This selective ion exchange behavior can be attributed to two factors: (1) size-matching effects and (2) differential affinity toward S2− soft basic sites. According to the Lewis acid–base theory, as the radius of alkali metal ions increases (Na+: 0.95 Å < Rb+: 1.48 Å < Cs+: 1.69 Å [41]), the soft acid–soft base interaction between them and S2− becomes stronger. Consequently, Cs+, which is closer to the radius of Rb+ ion and has a stronger affinity for S2−, is more likely to replace the interlayer Rb+ ion, while Na+ has a weaker exchange capacity due to its smaller size and weaker affinity. In addition, in the simulated polluted tap water, river water (Wulong River), and lake water (Qishan Lake), RCs are all greater than 84% (Figure S9 and Table S11). This indicates that Rb-TiS2 can effectively remove Cs+ ions when the natural water body is polluted with a low concentration of Cs+.

3.8. Adsorption and Desorption

The elution cycle ability of an adsorbent is the key index in practical application. To evaluate this, desorption experiments were conducted on Cs-TiS2 using 0.5 M RbCl and 0.5 M KCl solutions as the eluent, respectively. The sample treated with RbCl solutions is named Rb-Cs-TiS2 and the sample treated with KCl is named K-Cs-TiS2. EDS analysis reveals uniform elemental distribution in the Rb-Cs-TiS2 and K-Cs-TiS2 samples. Among them, there is a small amount of Cs+ ion residue in K-Cs-TiS2, while there is almost no Cs+ ion residue in Rb-Cs-TiS2 (Figure 5c and Figure S7). Moreover, the powder diffraction peak of Rb-Cs-TiS2 is in good agreement with that of Rb-TiS2 (Figure S8), indicating that the framework structure remains unchanged. Considering the cost of the eluent, a 1 M KCl solution was selected as the eluent to carry out the elution cycle experiment (Figure S10, Tables S12 and S13) in the two elution cycles, and the adsorption rate and desorption rate of Cs+ (Equation (S10)) were maintained at a high level (RCs/ECs > 92%). These results further demonstrate the current framework flexibility and practical application potential of this material.

4. Conclusions

In this work, we developed a 2D thiotitanate (Rb-TiS2) for efficient Cs+ removal, which is synthesized using a unique “Mg + RbCl” molten salt method. This approach addresses the limitations of traditional solid-state synthesis, such as operational complexity and high safety risks, providing a valuable reference for the synthesis and design of alkali-metal-ion-containing transition metal sulfides. Notably, Rb-TiS2 represents the first example of alkali-metal-ion-containing ternary Group IVB MSIEs. Rb-TiS2 exhibits high adsorption capacity, broad pH activity range, rapid kinetics, excellent selectivity, and good adsorption–desorption behavior for Cs+ ion removal. Importantly, single-crystal analysis combined with XPS and EDS characterization techniques unambiguously reveals that the Cs+ capture arises from ion exchange between Cs+ and Rb+ ions located in the interlayer spaces of the Rb-TiS2. The efficient capture of Cs+ ions is enabled by the abundance of soft basic S2− sites, size-matched and exchangeable Rb+ ions, appropriate interlayer spacing, and the flexible tunable layered structure of Rb-TiS2. This work offers a valuable reference for synthesizing and designing alkali-metal-ion-containing transition metal sulfides. Moreover, it deepens the understanding of ion exchange mechanisms in these materials and paves the way for designing new materials for nuclear waste management and environmental remediation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/separations12050104/s1, Figure S1: EDS spectra for (a) TiS2 sample and (b) Rb-TiS2 sample; Figure S2: Comparison of PXRD patterns of simulated RbTiS2 (ICSD#: 77990), experimental Rb-TiS2 sample and simulated Rb-TiS2; Figure S3: Structural diagrams of (a) Rb-TiS2 and (b) Cs-TiS2 showing the positions for interlayer cations. Coordination modes of Rb+ or Cs+ ion in (c) Rb-TiS2 and (d) Cs-TiS2 with M-S distances in Å marked (For clarity, only S atoms coordinated to the corresponding Rb(1) or Cs(1) are shown); Figure S4: TG curves of Rb-TiS2 and Cs-TiS2; Figure S5: Experimental PXRD pattern for the Cs-TiS2 sample compared with the simulated one from single-crystal X-ray structure; Figure S6: (a) PXRD patterns of experimental Rb-TiS2 sample after being soaked in different pH solutions. (b) Photographs and (c) Ti leaching percentage curves of Rb-TiS2 crystals treated with different pH solutions; Figure S7: EDS spectra for (a) Cs-TiS2 sample, (b) Rb-Cs-TiS2 sample, and (c) K-Cs-TiS2 sample; Figure S8: PXRD patterns of K-Cs-TiS2, Rb-Cs-TiS2, and Rb-TiS2 samples; Figure S9: Kd and R values of Cs+ ions captured by Rb-TiS2 in simulated polluted environmental water samples; Figure S10: (a) Adsorption rates (RCs) and desorption rates (ECs) of Rb-TiS2 in two adsorption and resolution cycles. (b) PXRD patterns of experimental Rb-TiS2 samples, the sample after each adsorption/desorption and the simulated powder of KTiS2 (ICSD#: 641335); Table S1: Crystallographic data and structural refinements for compounds Rb-TiS2, Cs-TiS2; Table S2: Atomic coordinates (×104), equivalent isotropic displacement parameters (Å2 × 103), SOFs (Site Occupancy Factors) and atomic sites for Rb-TiS2 and Cs-TiS2, U(eq) is defined as 1/3 of the trace of the orthogonalized Uij tensor; Table S3: Experimental results of isothermal adsorption of Cs+ by Rb-TiS2 (V/m = 1000 mL·g−1 at room temperature, 6 h contact time); Table S4: Fitting parameters of the Langmuir, Freundlich and Langmuir–Freundlich models for the isothermal adsorption of Cs+ by Rb-TiS2; Table S5. Structure, synthesis, adsorption capacity, and pH range of ion exchange/adsorption materials for Cs+ ion removal in recent 5 years; Table S6: Experimental results on the kinetics of Cs+ capture by Rb-TiS2 (V/m = 1000 mL·g−1; at room temperature); Table S7: Kinetic fitting parameters of Rb-TiS2 for Cs+ capture; Table S8: pH-dependent adsorption results of Rb-TiS2 for Cs+ removal (V/m = 1000 mL·g−1, at room temperature and 6 h contact time); Table S9: Adsorption results of Rb-TiS2 for Cs+ removal under conditions with no competing ions and with individual competing M (Na+, Ca2+, and Mg2+) ions (V/m = 1000 mL·g−1, at room temperature and 6 h contact time); Table S10: Adsorption results of Rb-TiS2 under different Na+/Cs+ molar ratios (V/m = 1000 mL·g−1, at room temperature and 6 h contact time); Table S11: Adsorption data of Cs+ ions by Rb-TiS2 in simulated polluted environmental water samples (tap water, Wulong River water, Qishan Lake water; V/m = 1000 mL·g−1, at room temperature and 6 h contact time); Table S12: Results of Cs+ adsorption-desorption cycle experiments of Rb-TiS2 (V/m = 1000 mL·g−1, at room temperature and 8 h contact time); Table S13 Leaching of Ti during the desorption in the cycle experiment of Rb-TiS2 (m is the total mass of the sample; V is the volume of solution). The authors have cited additional references within the Supporting Information [31,49,51,52,53,54,55,56,57,58,59,60].

Author Contributions

Writing, original draft preparation, C.W.; data curation, C.W.; visualization, C.W.; review and editing, C.W., S.J., Y.Z., J.L., H.S., M.F. and X.H.; project administration, M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 22325605, U21A20296, 22076185, and 22406185), the Natural Science Foundation of Fujian Province (No. 2024J08105), and the Strategic Priority Research Program of the Chinese Academy of Sciences (No. XDB1170000).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Author Shaoqing Jia was employed by the HTA Co., Ltd. The remaining 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. Bruno, J.; Ewing, R.C. Spent nuclear fuel. Elements 2006, 2, 343–349. [Google Scholar] [CrossRef]
  2. Burns, P.C.; Ewing, R.C.; Navrotsky, A. Nuclear Fuel in a Reactor Accident. Science 2012, 335, 1184–1188. [Google Scholar] [CrossRef] [PubMed]
  3. Koo, Y.H.; Yang, Y.S.; Song, K.W. Radioactivity release from the Fukushima accident and its consequences: A review. Prog. Nucl. Energy 2014, 74, 61–70. [Google Scholar] [CrossRef]
  4. Matsuda, K.; Yamamoto, S.; Miyamoto, K. Comparison of Cs-137 uptake, depuration and continuous uptake, originating from feed, in five salmonid fish species. J. Environ. Radioact. 2020, 222, 106350. [Google Scholar] [CrossRef]
  5. Morimoto, M.; Kobayashi, J.; Kino, Y. Radiation dose and gene expression analysis of wild boar 10 years after the Fukushima Daiichi Nuclear Plant accident. Sci. Rep. 2022, 12, 18653. [Google Scholar] [CrossRef]
  6. Nash, K.L.; Braley, J.C. Chemistry of Radioactive Materials in the Nuclear Fuel Cycle; Nash, K.L.L., Gregg, J., Eds.; Woodhead Publishing: Sawston, UK, 2011; pp. 3–22. [Google Scholar]
  7. Ashraf, M.A.; Akib, S.; Maah, M.J.; Yusoff, I.; Balkhair, K.S. Cesium-137: Radio-Chemistry, Fate, and Transport, Remediation, and Future Concerns. Crit. Rev. Environ. Sci. Technol. 2014, 44, 1740–1793. [Google Scholar] [CrossRef]
  8. Mitrovic, B.M.; Vitorovic, G.; Vicentijevic, M.; Vitorovic, D.; Pantelic, G.; Lazarevic-Macanovic, M. Comparative study of 137Cs distribution in broilers and pheasants and possibilities for protection. Radiat. Environ. Biophys. 2012, 51, 79–84. [Google Scholar] [CrossRef]
  9. Cardis, E.; Howe, G.; Ron, E.; Bebeshko, V.; Bogdanova, T.; Bouville, A.; Carr, Z.; Chumak, V.; Davis, S.; Demidchik, Y.; et al. Cancer consequences of the Chernobyl accident: 20 years on. J. Radiol. Prot. 2006, 26, 127–140. [Google Scholar] [CrossRef]
  10. Murase, K.; Murase, J.; Horie, R.; Endo, K. Effects of the Fukushima Daiichi nuclear accident on goshawk reproduction. Sci. Rep. 2015, 5, 9405. [Google Scholar] [CrossRef]
  11. Haque, M.; Dayem, S.B.; Tasnim, N.T.; Islam, M.R.; Shakil, M.S. Biological impact of Chernobyl radiation: A review of recent progress. Int. J. Radiat. Biol. 2024, 100, 1405–1415. [Google Scholar] [CrossRef]
  12. Clemons, J.; Blumenberg, A. The Goiania incident, the semiotics of danger, and the next 10,000 years. Clin. Toxicol. 2023, 61, 551–558. [Google Scholar] [CrossRef] [PubMed]
  13. Tosri, C.; Chusreeaeom, K.; Limtiyayotin, M.; Sukin, N.; Jompuk, P. Comparative effect of high energy electron beam and 137Cs gamma ray on survival, growth and chlorophyll content in curcuma hybrid ‘laddawan’ and determine proper dose for mutations breeding. Emir. J. Food Agric. 2019, 31, 321–327. [Google Scholar] [CrossRef]
  14. Fentiman, I.S.; Deshmane, V.; Tong, D.; Winter, J.; Mayles, H.; Chaudary, M.A. Caesium137 implant as sole radiation therapy for operable breast cancer:: A phase II trial. Radiother. Oncol. 2004, 71, 281–285. [Google Scholar] [CrossRef] [PubMed]
  15. Artun, O. A study of nuclear structure for 244Am, 238Pu, 210Po, 147Pm, 137Cs, 90Sr and 63Ni nuclei used in nuclear battery. Mod. Phys. Lett. A 2017, 32, 750117. [Google Scholar] [CrossRef]
  16. Wang, J.L.; Zhuang, S.T. Removal of cesium ions from aqueous solutions using various separation technologies. Rev. Environ. Sci. Bio-Technol. 2019, 18, 231–269. [Google Scholar] [CrossRef]
  17. Fang, D.; Wang, Y.; Liu, H.; Zhang, H.; Ye, X.; Li, Q.; Li, J.; Wu, Z. Efficient extraction of Rb+ and Cs+ by a precipitation flotation process with ammonium phosphowolframate as precipitant. Colloid Surf. A-Physicochem. Eng. Asp. 2021, 608, 125581. [Google Scholar] [CrossRef]
  18. Wang, J.L.; Zhuang, S.T. Cesium separation from radioactive waste by extraction and adsorption based on crown ethers and calixarenes. Nucl. Eng. Technol. 2020, 52, 328–336. [Google Scholar] [CrossRef]
  19. Zhang, X.Y.; Gu, P.; Liu, Y. Decontamination of radioactive wastewater: State of the art and challenges forward. Chemosphere 2019, 215, 543–553. [Google Scholar] [CrossRef]
  20. Tang, J.H.; Sun, H.Y.; Ma, W.; Feng, M.L.; Huang, X.Y. Recent Progress in Developing Crystalline Ion Exchange Materials for the Removal of Radioactive Ions. Chin. J. Struct. Chem. 2020, 39, 2157–2171. [Google Scholar] [CrossRef]
  21. Manos, M.J.; Kanatzidis, M.G. Metal sulfide ion exchangers: Superior sorbents for the capture of toxic and nuclear waste-related metal ions. Chem. Sci. 2016, 7, 4804–4824. [Google Scholar] [CrossRef] [PubMed]
  22. Tang, J.; Feng, M.; Huang, X. Metal chalcogenides as ion-exchange materials for the efficient removal of key radionuclides: A review. Fundam. Res. 2024, in press. [CrossRef]
  23. Zhao, Q.; Wang, S.; Wu, Y.C.; Wang, Y.X.; Ma, S.S.; Shih, K.M. Layered metal sulfides with MaSbc− framework (M = Sb, In, Sn) as ion exchangers for the removal of Cs(I) and Sr(II) from radioactive effluents: A review. Front. Chem. 2023, 11, 1292979. [Google Scholar] [CrossRef] [PubMed]
  24. Manos, M.J.; Kanatzidis, M.G. Highly Efficient and Rapid Cs+ Uptake by the Layered Metal Sulfide K2xMnxSn3−xS6 (KMS-1). J. Am. Chem. Soc. 2009, 131, 6599–6607. [Google Scholar] [CrossRef] [PubMed]
  25. Mertz, J.L.; Fard, Z.H.; Malliakas, C.D.; Manos, M.J.; Kanatzidis, M.G. Selective Removal of Cs+, Sr2+, and Ni2+ by K2xMgxSn3−xS6 (x = 0.5–1) (KMS-2) Relevant to Nuclear Waste Remediation. Chem. Mater. 2013, 25, 2116–2127. [Google Scholar] [CrossRef]
  26. Jiang, Z.Z.; Liu, G.L.; Ma, C.; Guo, Y.F.; Duo, J.; Li, M.L.; Deng, T.L. Cesium removal from wastewater: High-efficient and reusable adsorbent K1.93Ti0.22Sn3S6.43. Chemosphere 2022, 305, 135406. [Google Scholar] [CrossRef]
  27. Sarma, D.; Malliakas, C.D.; Subrahmanyam, K.S.; Islama, S.M.; Kanatzidis, M.G. K2xSn4−xS8−x (x = 0.65–1): A new metal sulfide for rapid and selective removal of Cs+, Sr2+ and UO22+ ions. Chem. Sci. 2016, 7, 1121–1132. [Google Scholar] [CrossRef]
  28. Yu, J.M.; Luo, D.; Ma, Z.J.; Zheng, B.; Cheng, F.F.; Xiong, W.W. Effective Enrichment of Low-Concentration Rare-Earth Ions by Three-Dimensional Thiostannate K2Sn2S5. ACS Appl. Mater. Interfaces 2021, 13, 55188–55197. [Google Scholar] [CrossRef]
  29. Qi, X.H.; Du, K.Z.; Feng, M.L.; Li, J.R.; Du, C.F.; Zhang, B.; Huang, X.Y. A two-dimensionally microporous thiostannate with superior Cs+ and Sr2+ ion-exchange property. J. Mater. Chem. A 2015, 3, 5665–5673. [Google Scholar] [CrossRef]
  30. Feng, M.L.; Sarma, D.; Gao, Y.J.; Qi, X.H.; Li, W.A.; Huang, X.Y.; Kanatzidis, M.G. Efficient Removal of [UO2]2+, Cs+, and Sr2+ Ions by Radiation-Resistant Gallium Thioantimonates. J. Am. Chem. Soc. 2018, 140, 11133–11140. [Google Scholar] [CrossRef]
  31. Tang, J.H.; Jin, J.C.; Li, W.A.; Zeng, X.; Ma, W.; Li, J.L.; Lv, T.T.; Peng, Y.C.; Feng, M.L.; Huang, X.Y. Highly selective cesium(I) capture under acidic conditions by a layered sulfide. Nat. Commun. 2022, 13, 658. [Google Scholar] [CrossRef]
  32. Wu, M.; Sul, W.P.; Jasutkar, N.; Huang, X.Y.; Li, J. An open-framework bimetallic chalcogenide structure K3Rb3Zn4Sn3Se13 built on a unique Zn49Sn3Se1612− cluster: Synthesis, crystal structure, ion exchange and optical properties. Mater. Res. Bull. 2005, 40, 21–27. [Google Scholar] [CrossRef]
  33. Manos, M.J.; Iyer, R.G.; Quarez, E.; Liao, J.H.; Kanatzidis, M.G. {Sn[Zn4Sn4S17]}6−: A robust open framework based on metal-linked penta-supertetrahedral [Zn4Sn4S17]10− clusters with ion-exchange properties. Angew. Chem.-Int. Edit. 2005, 44, 3552–3555. [Google Scholar] [CrossRef] [PubMed]
  34. Liang, C.Q.; Jia, M.C.; Wang, X.W.; Du, Z.H.; Men, J.F.; Ding, H. Preparation of potassium niobium sulfide and its selective adsorption properties for Sr2+ and Co2+. J. Radioanal. Nucl. Chem. 2019, 322, 377–387. [Google Scholar] [CrossRef]
  35. Zhang, M.D.; Gu, P.; Yan, S.; Dong, L.H.; Zhang, G.H. Na/Zn/Sn/S (NaZTS): Quaternary metal sulfide nanosheets for efficient adsorption of radioactive strontium ions. Chem. Eng. J. 2020, 379, 122227. [Google Scholar] [CrossRef]
  36. Wei, C.; Liu, J.; Zhao, Y.; Jia, S.; Gao, Y.; Saha, R.A.; Torchio, R.; Zhang, T.; Yang, L.; Sun, H.; et al. Efficient capture of Cs+ and Sr2+ by layered thioniobates and thiotantalate and insight into the structure-property relationship. Sci. China-Chem 2025. [Google Scholar] [CrossRef]
  37. Sunshine, S.A.; Kang, D.; Ibers, J.A. A New Low-Temperature Route to Metal Polychalcogenides: Solid-State Synthesis of K4Ti3S14, A Novel One-Dimensional Compound. J. Am. Chem. Soc. 1987, 109, 6202–6204. [Google Scholar] [CrossRef]
  38. Klepp, K.O. K2TiS3, A New Thiotitanate(lV)with Pentacoordinate Titanium. Z. Naturforsch. (B) 1992, 47, 201–204. [Google Scholar] [CrossRef]
  39. Itoh, Y.; Suematsu, H. Stability of Two-Dimensional Intercalant Structure in KxTiS2. Mater. Sci. Forum 1992, 91–93, 375–380. [Google Scholar] [CrossRef]
  40. Bichon, J.; Danot, M.; Rouxel, J. Systematique Structurale Pour Les Series D’intercalaires MxTiS2 (M = Li, Na, K, Rb, Cs). C. R. Des Seances De L’academie Des Sci. Ser. C Sci. Chim. 1973, 276, 1283–1286. [Google Scholar]
  41. Nightingale, E.R. Phenomenological Theory of Ion Solvation-Effective of Radii of Hydrated Ions. J. Phys. Chem. 1959, 63, 1381–1387. [Google Scholar] [CrossRef]
  42. Pearson, R.G. Hard and soft acids and bases. J. Am. Chem. Soc. 1963, 85, 3533–3539. [Google Scholar] [CrossRef]
  43. Duong, D.D. Adsorption Analysis: Equilibria and Kinetics; World Scientific: Singapore, 1998. [Google Scholar]
  44. Amesh, P.; Suneesh, A.S.; Venkatesan, K.A.; Maheswari, R.U.; Vijayalakshmi, S. Preparation and ion exchange studies of cesium and strontium on sodium iron titanate. Sep. Purif. Technol. 2020, 238, 116393. [Google Scholar] [CrossRef]
  45. Asgari, P.; Mousavi, S.H.; Aghayan, H.; Ghasemi, H.; Yousefi, T. Nd-BTC metal-organic framework (MOF); synthesis, characterization and investigation on its adsorption behavior toward cesium and strontium ions. Microchem. J. 2019, 150, 104188. [Google Scholar] [CrossRef]
  46. Bayülken, S.; Basçetin, E.; Güçlü, K.; Apak, R. Investigation and Modeling of Cesium(I) Adsorption by Turkish Clays: Bentonite, Zeolite, Sepiolite, and Kaolinite. Environ. Prog. Sustain. Energy 2011, 30, 70–80. [Google Scholar] [CrossRef]
  47. Park, Y.; Lee, Y.C.; Shin, W.S.; Choi, S.J. Removal of cobalt, strontium and cesium from radioactive laundry wastewater by ammonium molybdophosphate-polyacrylonitrile (AMP-PAN). Chem. Eng. J. 2010, 162, 685–695. [Google Scholar] [CrossRef]
  48. Gupta, K.; Yuan, B.L.; Chen, C.; Varnakavi, N.; Fu, M.L. K2xMnxSn3−xS6 (x = 0.5–0.95) (KMS-1) immobilized on the reduced graphene oxide as KMS-1/r-GO aerogel to effectively remove Cs+ and Sr2+ from aqueous solution. Chem. Eng. J. 2019, 369, 803–812. [Google Scholar] [CrossRef]
  49. Zhu, J.Y.; Cheng, L.; Zhao, Y.M.; Li, M.Y.; Wang, Z.Z.; Wang, J.; Wang, C.; Wang, K.Y. Structural investigation of the efficient capture of Cs+ and Sr2+ by a microporous Cd-Sn-Se ion exchanger constructed from mono-lacunary supertetrahedral clusterst. Inorg. Chem. Front. 2022, 9, 2880–2894. [Google Scholar] [CrossRef]
  50. Datta, S.J.; Moon, W.K.; Choi, D.Y.; Hwang, I.C.; Yoon, K.B. A Novel Vanadosilicate with Hexadeca-Coordinated Cs+ Ions as a Highly Effective Cs+ Remover. Angew. Chem.-Int. Edit. 2014, 53, 7203–7208. [Google Scholar] [CrossRef]
  51. Sun, H.Y.; Chen, Z.H.; Hu, B.; Tang, J.H.; Yang, L.; Guo, Y.L.; Yao, Y.X.; Feng, M.L.; Huang, X.Y. Boosting selective Cs+ uptake through the modulation of stacking modes in layered niobate-based perovskites. Nat. Commun. 2024, 15, 8681. [Google Scholar] [CrossRef]
  52. Chen, Z.H.; Jia, S.Q.; Sun, H.Y.; Tang, J.H.; Guo, Y.L.; Yao, Y.X.; Pan, T.Y.; Feng, M.L.; Huang, X.Y. All-in-one treatment: Capture and immobilization of 137 Cs by ultra-stable inorganic solid acid materials HMMoO6nH2O (M = Ta, Nb). Water Res. 2024, 255, 121459. [Google Scholar] [CrossRef]
  53. Yan, J.; Wu, J.T.; Li, B.H.; Zhang, Y.D.; Yang, Y.; Li, J.; Zhang, B. Economically friendly and facilely synthesized layered metal sulfide for efficient removal of aqueous cesium. J. Solid State Chem. 2024, 337, 124834. [Google Scholar] [CrossRef]
  54. Wang, K.Y.; Li, M.Y.; Cheng, L.; Hao, X.; Wang, C. Tailoring supertetrahedral cadmium/tin selenide clusters into a robust framework for efficient elimination of Cs+, Co2+, and Ni2+ ions. Inorg. Chem. Front. 2024, 11, 3229–3244. [Google Scholar] [CrossRef]
  55. Yan, J.; Zhang, B.; Li, J.; Yang, Y.; Wang, Y.N.; Zhang, Y.D.; Liu, X.Z. Rapid and Selective Uptake of Radioactive Cesium from Water by a Microporous Zeolitic-like Sulfide. Inorg. Chem. 2023, 62, 12843–12850. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, J.; Zhang, J.F.; Ni, S.; Xing, H.F.; Meng, Q.Y.; Bian, Y.Y.; Xu, Z.H.; Rong, M.; Liu, H.Z.; Yang, L.R. Cation-Intercalated Lamellar MoS2 Adsorbent Enables Highly Selective Capture of Cesium. ACS Appl. Mater. Interfaces 2023, 15, 49095–49106. [Google Scholar] [CrossRef]
  57. Sun, H.Y.; Hu, B.; Lv, T.T.; Guo, Y.L.; Yao, Y.X.; Yang, L.; Feng, M.L.; Huang, X.Y. Efficient Co-Adsorption and Highly Selective Separation of Cs+ and Sr2+ with a K+-Activated Niobium Germanate by the pH Control. Small 2023, 19, 2208212. [Google Scholar] [CrossRef]
  58. Liu, Y.; Shi, F.Q.; Hao, X.; Li, M.Y.; Cheng, L.; Wang, C.; Wang, K.Y. Open-framework hybrid zinc/tin selenide as an ultrafast adsorbent for Cs+, Ba2+, Co2+, and Ni2+. J. Hazard. Mater. 2023, 458, 132038. [Google Scholar] [CrossRef]
  59. Zhao, Y.M.; Cheng, L.; Wang, K.Y.; Hao, X.; Wang, J.; Zhu, J.Y.; Sun, M.; Wang, C. pH-Controlled Switch over Coadsorption and Separation for Mixed Cs+ and Sr2+ by an Acid-Resistant Potassium Thioantimonate. Adv. Funct. Mater. 2022, 32, 2112717. [Google Scholar] [CrossRef]
  60. Li, W.A.; Li, J.R.; Zhang, B.; Sun, H.Y.; Jin, J.C.; Huang, X.Y.; Feng, M.L. Layered Thiostannates with Distinct Arrangements of Mixed Cations for the Selective Capture of Cs+, Sr2+, and Eu3+ Ions. ACS Appl. Mater. Interfaces 2021, 13, 10191–10201. [Google Scholar] [CrossRef]
Figure 1. Photographs of (a) Rb-TiS2 crystal sample and (b) TiS2 crystal sample. (c) Comparison of PXRD patterns of experimental TiS2 and Rb-TiS2 samples with simulated TiS2 (ICSD#: 52195) and simulated Rb-TiS2.
Figure 1. Photographs of (a) Rb-TiS2 crystal sample and (b) TiS2 crystal sample. (c) Comparison of PXRD patterns of experimental TiS2 and Rb-TiS2 samples with simulated TiS2 (ICSD#: 52195) and simulated Rb-TiS2.
Separations 12 00104 g001
Figure 2. (a) An anionic layer of [TiS2]n0.32n. (b) View of the layers stacking in Rb-TiS2 along the c axis. (c) Schematic diagram of the Cs+ ions captured by Rb-TiS2 and the change in interlayer distances of Rb-TiS2 and Cs-TiS2.
Figure 2. (a) An anionic layer of [TiS2]n0.32n. (b) View of the layers stacking in Rb-TiS2 along the c axis. (c) Schematic diagram of the Cs+ ions captured by Rb-TiS2 and the change in interlayer distances of Rb-TiS2 and Cs-TiS2.
Separations 12 00104 g002
Figure 3. (a) PXRD patterns of Rb-TiS2 and Cs-TiS2. (b) Elemental distribution maps of Rb-TiS2 and Cs-TiS2. (c) X-ray photoelectron spectra of Rb-TiS2 and Cs-TiS2. Narrow scan XPS spectra of (d) Rb 3d and (e) Cs 3d of Rb-TiS2 and Cs-TiS2, respectively.
Figure 3. (a) PXRD patterns of Rb-TiS2 and Cs-TiS2. (b) Elemental distribution maps of Rb-TiS2 and Cs-TiS2. (c) X-ray photoelectron spectra of Rb-TiS2 and Cs-TiS2. Narrow scan XPS spectra of (d) Rb 3d and (e) Cs 3d of Rb-TiS2 and Cs-TiS2, respectively.
Separations 12 00104 g003
Figure 4. (a) The Cs+ adsorption isothermal data of Rb-TiS2 fitted by the Langmuir, Freundlich, and Langmuir–Freundlich models. (b) Kinetics of Cs+ removal by Rb-TiS2 plotted as Cs+ concentration and the removal rate of Cs+ vs. the time t (min). (c) Pseudo-first-order kinetics model and pseudo-second-order kinetics model fitting curves for kinetic data of Rb-TiS2 for Cs+ capture. (d) KdCs (column) and RCs (line) values of Rb-TiS2 in solutions with different initial pH values.
Figure 4. (a) The Cs+ adsorption isothermal data of Rb-TiS2 fitted by the Langmuir, Freundlich, and Langmuir–Freundlich models. (b) Kinetics of Cs+ removal by Rb-TiS2 plotted as Cs+ concentration and the removal rate of Cs+ vs. the time t (min). (c) Pseudo-first-order kinetics model and pseudo-second-order kinetics model fitting curves for kinetic data of Rb-TiS2 for Cs+ capture. (d) KdCs (column) and RCs (line) values of Rb-TiS2 in solutions with different initial pH values.
Separations 12 00104 g004
Figure 5. (a) Kd (column) and R (line) values of Rb-TiS2 for Cs+ removal under conditions with no competing ions and with individual competing M (Na+, Ca2+, and Mg2+) ions. C0Cs is about 6 mg·L−1 and the Cs:M molar ratio is 1:7.53, 1:10.64, and 1:4.80, respectively. (b) Kd (column) and R (line) values of Cs+ and Na+ ions removed by Rb-TiS2 with different molar ratios of Cs/Na. (c) Elemental distribution maps of Rb-Cs-TiS2 and K-Cs-TiS2.
Figure 5. (a) Kd (column) and R (line) values of Rb-TiS2 for Cs+ removal under conditions with no competing ions and with individual competing M (Na+, Ca2+, and Mg2+) ions. C0Cs is about 6 mg·L−1 and the Cs:M molar ratio is 1:7.53, 1:10.64, and 1:4.80, respectively. (b) Kd (column) and R (line) values of Cs+ and Na+ ions removed by Rb-TiS2 with different molar ratios of Cs/Na. (c) Elemental distribution maps of Rb-Cs-TiS2 and K-Cs-TiS2.
Separations 12 00104 g005
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

Wei, C.; Jia, S.; Zhao, Y.; Liu, J.; Sun, H.; Feng, M.; Huang, X. A Two-Dimensional Thiotitanate Ion Exchanger with High Cs+ Removal Performance. Separations 2025, 12, 104. https://doi.org/10.3390/separations12050104

AMA Style

Wei C, Jia S, Zhao Y, Liu J, Sun H, Feng M, Huang X. A Two-Dimensional Thiotitanate Ion Exchanger with High Cs+ Removal Performance. Separations. 2025; 12(5):104. https://doi.org/10.3390/separations12050104

Chicago/Turabian Style

Wei, Chang, Shaoqing Jia, Yingying Zhao, Jiating Liu, Haiyan Sun, Meiling Feng, and Xiaoying Huang. 2025. "A Two-Dimensional Thiotitanate Ion Exchanger with High Cs+ Removal Performance" Separations 12, no. 5: 104. https://doi.org/10.3390/separations12050104

APA Style

Wei, C., Jia, S., Zhao, Y., Liu, J., Sun, H., Feng, M., & Huang, X. (2025). A Two-Dimensional Thiotitanate Ion Exchanger with High Cs+ Removal Performance. Separations, 12(5), 104. https://doi.org/10.3390/separations12050104

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