Next Article in Journal
Effects of Sleeve Gastrectomy and Treadmill Exercise on Skeletal Muscle and Ectopic Fat in High-Fat Diet-Induced Obese Rats
Previous Article in Journal
scDown: A Pipeline for Single-Cell RNA-Seq Downstream Analysis
Previous Article in Special Issue
Uncovering Metal-Decorated TiO2 Photocatalysts for Ciprofloxacin Degradation—A Combined Experimental and DFT Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of Sr2+ from Synthetic Waste Effluents Using Taiwan Zhi-Shin Bentonite

1
Department of Geography, Hanshan Normal University, Chaozhou 521041, China
2
College of Resources and Environment, Fujian Agriculture and Forestry University, Fuzhou 350002, China
*
Authors to whom correspondence should be addressed.
The two authors contributed equally to the work.
Int. J. Mol. Sci. 2025, 26(11), 5298; https://doi.org/10.3390/ijms26115298
Submission received: 16 April 2025 / Revised: 23 May 2025 / Accepted: 26 May 2025 / Published: 30 May 2025

Abstract

This study investigated strontium (Sr2+) adsorption by Taiwan Zhi-Shin bentonite (cation exchange capacity (CEC): 80–86 meq 100 g−1) using Sr(NO3)2-simulated nuclear waste. Kinetic analysis revealed pseudo-second-order adsorption kinetics, achieving 95% Sr2+ removal within 5 min at pH 9. Isothermal studies showed a maximum capacity of 0.28 mmol g−1 (56 meq 100 g−1) at 15 mmol L−1 Sr2+, accounting for 65–70% CEC and fitting the Freundlich model. Cation exchange was the dominant mechanism (84% contribution), driven by Sr2+ displacing interlayer Ca2+. Alkaline conditions (pH > 9) enhanced adsorption through improved surface charge and electrostatic attraction. Thermodynamic studies demonstrated temperature-dependent behavior: increasing temperature reduced adsorption at 0.01 mM Sr2+ but increased efficiency at 10 mM. Na+ addition suppressed adsorption, aligning with cation exchange mechanisms. Molecular dynamics simulations identified hydrated Ca2+-Sr2+ water bridges interacting with bentonite via hydrogen-bonding networks. The material exhibits rapid kinetics (5 min equilibrium), alkaline pH optimization, and resistance to ion interference, making it suitable for emergency Sr2+ treatment. It shows promise as a cost-effective and good performing adsorbent for radioactive waste solutions.

Graphical Abstract

1. Introduction

As a heavy metal element with high solubility and mobility, the radionuclide strontium (Sr2+) is prevalent in nuclear industry wastewater and contaminated aquatic environments. Due to chemical properties analogous to those of calcium and barium [1], Sr2+ readily infiltrates human bone tissue through ion exchange mechanisms and forms stable Sr3(PO4)2 compounds. These results showed prolonged bioaccumulation with limited metabolic elimination [2]. Functioning as a bone-seeking beta radiation emitter, Sr2+ induces cutaneous radiation injury [3] and has been shown to provoke severe pathologies including leukemia, osteosarcoma, and hematopoietic disorders (characterized by marked depletion of leukocytes, erythrocytes, and thrombocytes), constituting significant health hazards [4]. Notably, the radioisotope 90Sr2+ exhibits an extended half-life of 28.79 years [5] and demonstrates pronounced environmental persistence. The 2011 Fukushima Daiichi nuclear disaster in Japan, which released substantial quantities of radioactive Sr2+ into marine ecosystems, has heightened global concerns regarding Sr2+ contamination [6]. Presently, Sr2+ concentrations in water in Chinese cities average 360 ppb [7].
With the rapid development of the nuclear industry, radioactive wastewater discharge has emerged as a primary pathway for radionuclides to enter aquatic environments. The development of efficient Sr2+ removal technologies from water systems constitutes a critical research frontier in environmental science and engineering. Current mainstream techniques include ionic exchange [8], chemical precipitation [9], evaporation concentration [10], bioremediation approaches [11], photocatalysis [12], and adsorption [13] (Table S1). Nevertheless, these methodologies exhibit operational limitations under practical applications. For instance, the evaporation concentration demonstrates excessive energy demands and process-related risks including equipment corrosion and scaling [14]. Bioremediation approaches suffer from suboptimal efficiency and environmental dependency [15,16]. The regeneration effluents generated by ionic exchange systems pose secondary contamination risks [17,18], while membrane separation technologies are constrained by prohibitive operational costs and low resource recovery efficiency [19]. Consequently, the innovation of high-performance, cost-effective, and ecologically sustainable removal technologies has become a strategic priority in contemporary scientific investigations.
The adsorption method is recognized as one of the most promising technologies for radioactive Sr2+ removal, attributed to its operational simplicity, cost-effectiveness, and broad applicability. Recent research has explored diverse organic and inorganic adsorbents, including ion-exchange resins, lanthanide oxalate frameworks, natural minerals, graphene oxide composites, metal-organic frameworks (MOFs), polymeric hydrogels, and novel engineered sorbents, as documented in studies [20,21,22,23,24,25,26,27,28] (Table S2). Conventional adsorbents such as activated carbon were shown to exhibit constrained Sr2+ sorption capacities (e.g., 44.42 mg g−1), while existing studies frequently fail to specify critical operational parameters such as optimal adsorbent dosage, thereby impeding practical implementation [29]. Furthermore, the research and development cost of new adsorption materials is high, and the energy consumption is large, so it is difficult to popularize modified materials like Moloukhia et al. [30] on a large scale. Therefore, it is urgent to find a natural mineral material with low cost and excellent adsorption property.
Bentonite, as a layered silicate mineral consisting mainly of two layers of silico-oxygen tetrahedrons and one layer of alumino-oxygen octahedrons, has a high specific surface area and cation exchange capacity (CEC), capable of highly efficient adsorption of Sr2+ in water by cation exchange and electrostatic action [31]. As a bentonite resource with abundant reserves and low price [32], the unique layered structure and surface chemical properties of Taiwan Zhi-Shin bentonite show great potential in the field of radioactive Sr2+ adsorption [33,34]. Studies have shown that the adsorption performance of bentonite can be further improved after acidification or organic modification, which is significantly better than that shown by traditional adsorbents [35,36]. In addition, the metal sulfide of Na2Sn2S7 has been shown to be stable over a wide pH range and to have a strong resistance to ion interference, which is suitable for Sr2+ removal in complex water environments [37]. Taiwan is an island with three nuclear power plants, and its nuclear waste has been piling up on Lanyu Island (located in the southeast of Taiwan). The Zhi-Shin soil in this study is from Taitung, Taiwan, which is quite close to Lanyu Island, and this material has the characteristics of high efficiency and economical production. Studying its adsorption performance for Sr2+ is helpful for developing the buffering behavior of Taiwan Zhi-Shin bentonite against nuclear waste. It provides a potentially advantageous material for the disposal of nuclear waste in Taiwan and even around the world. Although bentonite has shown significant advantages in Sr2+ adsorption, it still faces some challenges in practical applications. For example, the presence of competing ions (e.g., Ca2+, Mg2+) may affect its selective adsorption on Sr2+ [38], and the adsorbed bentonite needs to be efficiently recycled or cured to avoid secondary pollution [39]. Therefore, future research should focus on modification optimization, dynamic adsorption properties, and actual wastewater treatment verification of bentonite in order to promote transformation from laboratory research to engineering applications.
In this study, Taiwan Zhi-Shin bentonite was used as the adsorbent, and the batch method was employed to study the adsorption properties of Taiwan Zhi-Shin bentonite on Sr2+ under different environmental conditions (time, concentration, temperature, pH, and ionic strength). Through characterization techniques such as X-ray diffraction (XRD), the impact of different pH levels and initial concentrations on adsorption performance was explored, and molecular simulation techniques were utilized to verify the respective adsorption mechanism. A systematic discussion on its adsorption performance and mechanism for Sr2+ in water was conducted. Based on the efficient adsorption technology of bentonite, solving the problem of Sr2+ pollution is of great significance. The aim is to provide an economical, efficient, and environmentally friendly solution for the treatment of radioactive wastewater.

2. Results

2.1. Adsorption Kinetics Experiment of Strontium (Sr2+)

The adsorption kinetics data of Sr2+ using Taiwan Zhi-Shin bentonite, as illustrated in Figure 1, exhibited excellent agreement with a pseudo-second-order kinetic model (R2 = 1.000), a mechanism frequently observed in cation exchange-dominated adsorption systems [40]. It is worth noting that the system achieved adsorption equilibrium within 15 min under both 0.1 mmol L−1 (0.1 mM-Sr) and 10 mmol L−1 (10 mM-Sr) initial Sr2+ concentrations (Figure 1a,b), demonstrating that equilibrium time remained unaffected by increasing Sr2+ concentrations. This rapid adsorption behavior aligns with prior studies [41] and highlights distinctive advantages over other reported materials (Table S3). Specifically, Taiwan Zhi-Shin bentonite achieves equilibrium comparably fast to Na-bentonite (11–14 min) [13], implying that the interlayer structure of montmorillonite facilitates Sr2+ adsorption kinetics. However, unlike Na-bentonite, which exhibits reversible Sr2+ adsorption unsuitable for practical applications [42], Taiwan Zhi-Shin bentonite maintains irreversible binding. Its performance significantly surpasses slower adsorbents such as metal-organic framework/potassium nickel hexacyanoferra (MOF/KniFC) composites (45 min) [25], basic zeolites (6 h) [23], and sulfonylcalixarene-loaded XAD-7 resin (7 h) [20]. Furthermore, Taiwan Zhi-Shin bentonite operates effectively across a broad concentration range (0.1–10 mmol L−1), unlike basic zeolites, which excel only at low concentrations (50 mg g−1) [23]. These rapid adsorption properties, coupled with minimal pretreatment requirements (e.g., no acid activation or ion exchange) and lower cost compared to synthetic materials, position Taiwan Zhi-Shin bentonite as an ideal candidate for emergency remediation of sudden radionuclide contamination. The combined effects of its adsorption kinetics advantages and inherent interlayer structural features enabled efficient Sr2+ capture via cation exchange, thereby optimizing nuclear waste management strategies.

2.2. Adsorption Isotherm Experiment of Sr2+

The adsorption isotherm data of Sr2+ on Taiwan Zhi-Shin bentonite were fitted using Freundlich and Langmuir models, with the Freundlich model (R2 = 0.96) displaying a significantly better agreement than the Langmuir model (Figure 2). This result revealed heterogeneous surface characteristics and a multilayer adsorption mechanism [43], further corroborated by the Freundlich parameter n = 1.196 (1 < n < 10) listed in Table 1. The value of n indicates a continuous increase in adsorption capacity with rising Sr2⁺ concentrations within the experimental range [44], achieving a maximum adsorption capacity of 0.28 mmol g−1 (24.53 mg g−1), which corresponds to 65–70% of its CEC. Although Na-bentonite [13], MOF/KNiFC composites [25], and basic zeolites [23] exhibited higher adsorption capacities (Table S2), their practical applicability is constrained by high costs and complex pretreatment requirements. In contrast, Taiwan Zhi-Shin bentonite demonstrates distinct advantages in engineering scenarios due to its natural abundance, absence of pretreatment needs, and operational simplicity. Previous mechanistic studies on Sr2+ adsorption by bentonite confirm cation exchange as the dominant adsorption mechanism [13,45]. It is worth noting that the adsorption capacity of Taiwan Zhi-Shin bentonite (24.53 mg g−1) was significantly lower than that of Na-bentonite (46 mg g−1) [13], a disparity attributed to differences in ion-exchange behavior between Ca-bentonite (Taiwan Zhi-Shin bentonite) and Na-bentonite. The smaller hydrated ionic radius of Na+ (1.16 Å) compared to Ca2+ (1.26 Å) and its lower ion-exchange energy barrier facilitate more efficient competitive substitution with Sr2+ [46], a phenomenon consistent with the established influence of exchangeable cation radii on ion-exchange efficiency in clay minerals [34]. Raw bentonite ore, predominantly composed of Ca-bentonite with impurities, requires Na₂CO₃ processing to produce pure Na-bentonite [47,48]. As a natural Ca-bentonite, Taiwan Zhi-Shin bentonite retains inherent ion-exchange limitations due to its unmodified state. Nevertheless, its direct utilization without chemical modification, coupled with abundant natural reserves and compatibility with industrial protocols, positions it as a preferred buffer material for nuclear waste disposal applications.

2.3. Experiments on Influencing Factors of Adsorption of Sr2+

Ion exchange constitutes the overarching framework governing Sr2⁺ adsorption, as evidenced by concomitant cation desorption patterns.

2.3.1. Exchangeable Cation Desorption

Figure 3 illustrates the release of metal cations (Na+, K+, Ca2+, and Mg2+) accompanying the adsorption and interlayer incorporation of Sr2+ ions, with cation exchange identified as the primary mechanism governing Sr2+ adsorption in Taiwan Zhi-Shin bentonite. Within the entire application range of Sr2+ concentration, the molar ratio between the total amount of cation desorption and the adsorbed Sr2+ was close to 1:1 (with a slope of 0.81), proving that ion exchange plays an important role in the adsorption mechanism of Taiwan Zhi-Shin bentonite and Sr2+. This is consistent with the results of previous studies [13,34,38,45,49]. The interlayer cations of Taiwan Zhi-Shin bentonite were shown to be predominantly Ca2+ (dominant species), accompanied by minor Na+ and Mg2+ (see Table S4 for detailed composition). During Sr2+ adsorption, the concomitant cation release exhibited distinct patterns: K+ release remains negligible and largely independent of Sr2+ adsorption levels, whereas Na+ and Mg2+ release increased marginally. Notably, Ca2+ dominated the displacement process, accounting for 60.4% of the total released cations relative to the adsorbed Sr2+ concentration (Figure 3). This observation confirms cation exchange as the principal driver of Sr2+ uptake in Taiwan Zhi-Shin bentonite. A strong positive correlation exists between adsorbed Sr2+ quantities and released cation concentrations, demonstrating that Sr2+ adsorption directly triggers the displacement of native interlayer cations, particularly Ca2+. This preferential exchange likely stems from differences in charge density and hydration properties between Sr2+ and resident interlayer cations [50]. The observed behavior aligns with studies on silicoantimonate systems, where Ca2+ similarly mediated Sr2+ adsorption dynamics [28]. These findings not only elucidate the high Sr2+ adsorption capacity of bentonite but also advance the mechanistic understanding of clay minerals in regulating the immobilization and transport of heavy metal contaminants within engineered and natural environments.

2.3.2. Influence of pH Value

The pH-dependent adsorption characteristics of Sr2+ on Taiwan Zhi-Shin bentonite are presented in Figure 4a,b for 0.1 mM-Sr and 10 mM-Sr, respectively. Figure 5a,b further shows cation desorption patterns under corresponding conditions. A consistent upward trend in Sr2+ adsorption emerges with increasing pH across both concentrations, though mechanistic divergences arise depending on initial Sr2+ loading. At 0.1 mM-Sr, rapid adsorption saturation occurred due to minimal competitive interference, whereas the 10 mM-Sr system exhibited constrained adsorption enhancement attributable to Sr2+ hydrolysis effects and competitive adsorption with Na+, Mg2+, and Ca2+ (Figure 4b and Figure 5b). Speciation analysis confirmed Sr2+ as the dominant aqueous species between pH 1–11, with minor Sr(OH)+ formation, transitioning to Sr(OH)+ dominance at pH > 13 [51]. The point of zero charge (PZC = 9.2) critically modulated adsorption behavior: below pH 4, elevated H+ concentrations competed with Sr2+ for negatively charged surface sites, suppressing adsorption (Figure 4a,b). Progressive pH elevation (5–9) facilitated OH--mediated outer-sphere complexation [52], enabling enhanced Sr2+ adsorption via electrostatic attraction in the electrical double layer (Figure 4a). For the 10 mM-Sr system, however, limited adsorption gains in this pH range reflect persistent cation competition (Figure 4b and Figure 5b). Above pH 9, Ca2+ precipitation [53] alleviated competitive effects, driving substantial Sr2+ uptake (Figure 4a). At pH 10, co-existing Sr2+ and Sr(OH)+ species synergistically enhanced adsorption through dual electrostatic interactions (Figure 4b) [54], though adsorption plateaus at pH 11 due to renewed competition from Na+ and residual Ca2+ occurred (Figure 5b). The pH-modulated adsorption trends ultimately reflect perturbations in the ion exchange equilibrium established in Section 2.3.1. The adsorption behavior was not only affected by the pH and initial concentration of the solution but was also closely correlated to the morphological changes in its hydrolysates and the competitive adsorption with other cations. Therefore, the adsorption properties of Taiwan Zhi-Shin bentonite on Sr2+ were shown to be more suitable under alkaline conditions (pH > 9).

2.3.3. Effect of Ionic Strength on Adsorption

The competitive cation exchange mechanism governing Sr2+ adsorption on Taiwan Zhi-Shin bentonite was further validated through ionic strength experiments. Studies have shown that due to its valence layer being full and its inability to form strong coordination bonds with surface sites, Sr2+ was mainly adsorbed on the surface at the clay-water interface in the form of a mononuclear exo-spherical complex, which is weakly bound, easily replaced by competing cations, and displays high migration [55]. As depicted in Figure 6, incremental increases in NaCl concentration (0.01 M, 0.1 M, and 1 M) induced a marked reduction in Sr2+ adsorption capacity. At 0.1 mM-Sr, adsorption decreased from 0.0027 mmol g−1 to 0.0008 mmol g−1 (Figure 6a), while at 10 mM-Sr, the adsorption capacity declined from 0.21 mmol g−1 to 0.04 mmol g−1 (Figure 6b). This indicates that the competitive effect of cation exchange was enhanced with the increase in sodium ion concentration, which leads to the decrease in Sr2+ adsorption. In addition, compared with the low ionic strength (0–0.1 M NaCl), the adsorption of Sr2+ decreased rapidly. This result is consistent with the experimental results of Sr2+ adsorption on other clay minerals [55], which further verifies that cation exchange is the main mechanism of Sr2+ adsorption.

2.3.4. Effect of Temperature on Sr2+ Adsorption

Under 0.1 mM-Sr and 10 mM-Sr, the adsorption properties of Taiwan Zhi-Shin bentonite on Sr2+ at 30 °C, 45 °C, and 60 °C were investigated, respectively, and the results are shown in Figure 7. By plotting LnKd against 1/T (van’t Hoff plot), the relevant thermodynamic parameters for the adsorption of Sr2+ by Taiwan Zhi-Shin bentonite can be calculated [56]. The calculated results of thermodynamic parameters are shown in Table 2. Under the condition of 0.1 mM-Sr, the calculated enthalpy change is less than zero (ΔH < 0). The adsorption of Sr2+ by Taiwan Zhi-Shin bentonite is an exothermic reaction, and high temperature is not conducive to adsorption, which is consistent with the slightly decreasing trend of adsorption with increasing temperature in the experiment. It is worth noting that the experimental results are consistent with the adsorption results of Sr2+ on dolomite powder (Ci = 10 mol L−1), PAN/zeolite (Ci = 150 µg mL−1), attapulgite (Ci = 40 µg mL−1), and Al2O3 (Ci = 50 µg mL−1) [57,58,59,60]. These results occur because when the concentration of Sr2+ in the system is low, there are many active sites on the surface of the adsorbent, and the heat required for Sr2+ to be adsorbed on the surface of Taiwan Zhi-Shin bentonite is relatively small, so the process shows a trend of exothermic heat at low concentration. In addition, the negative value of the free energy variation (ΔG = −6.1 kJ mol−1) further demonstrated the spontaneity of the adsorption process. The negative value of entropy change (ΔS = −0.01 kJ mol−1 K−1) indicated that the system disorder was reduced during adsorption [61], which may be due to the fact that Sr2+ is fixed in a specific position on the surface of Taiwan Zhi-Shin bentonite, restricting its free movement.
Under the adsorption system of 10 mM-Sr, the adsorption of Sr2+ by Taiwan Zhi-Shin bentonite increased with the increase in temperature, which was opposite to the trend of adsorption at low concentration, indicating that the adsorption process of Sr2+ at high concentration was an endothermic reaction (ΔH = +4.71 kJ mol−1). The increase in temperature was conducive to the adsorption, which is consistent with the experimental results. The observation in this scenario can be explained as follows: as the concentration of Sr2+ increases, more Sr2+ will be adsorbed on the adsorbent, the adsorption of the active site will be relatively reduced, and the surface of the adsorbent will become more crowded. Previous studies have shown that a pyrochlore-type material is a material with a defect (vacancy) structure. Perhaps due to steric hindrance, the adsorption of Sr2+ requires a stronger chemical bond force, and the corresponding thermal effect in the adsorption process will increase [62]. The positive value of entropy change (ΔS = +0.02 kJ mol−1 K−1) indicates that the system disorder increases during the adsorption process, which may be due to the fact that more Sr2+ is adsorbed to the surface of Taiwan Zhi-Shin bentonite at high concentration, and the mobility of Sr2+ ions and surrounding water molecules increases due to ionic dehydration [63].

2.4. X-Ray Diffraction

Taiwan Zhi-Shin bentonite is mainly composed of montmorillonite and varying amounts of other minerals, such as quartz (SiO2), kaolinite, and pyrite [49]. As shown in Figure 8, Sr2+, by substituting interlayer Ca2+ and carrying a double-layer hydration molecule (2 W), is inserted into the interlayer of montmorillonite [64,65], causing its (001) crystal plane [66,67] base spacing (d001) to extend from 14.71 Å to 15.6 Å (pH = 9). This is consistent with the heterogeneous adsorption properties revealed by the Freundlich isothermal model (n > 1.2). The (001) crystal plane (2θ ≈ 12°) distance of kaolinite extends slightly from 7.1 Å to 7.23 Å [68,69], confirming the auxiliary contribution of the coordination of surface hydroxyl groups with Sr2+ to adsorption, forming a synergistic effect under the mineral complex system [70,71]. In addition, there was no indication that large quantities of other minerals are present in the 2θ range of 5–20° (Figure 8).
The dynamic properties of the adsorption process are regulated by the Sr2+ concentration in conjunction with the pH of the solution. In the low concentration range (0.05–5 mM), the uneven distribution of layer charges led to an alternating arrangement of Sr2+ interstratification (2 W hydrated, d001 ≈ 15.5 Å) with an uninterpolated region (1 W hydrated, d001 ≈ 12.5 Å) (Figure 8). There was spatial heterogeneity present, reflecting the competitive exchange of Sr2+ and Ca2+ [72]. This change can be attributed to the change in the hydration state of the exchanged and adsorbed cation Sr2+ and the coexistence of interstratified structures with varying hydration states, resulting in dynamic adjustments of the interlayer structure [73]. As the concentration increases to 10–15 mM, adsorption approaches saturation, the interlayer disappears, and the interlayer expands uniformly to 15.6 Å, indicating full occupation of the interlayer site by Sr2+. This phenomenon is consistent with the expansion behavior of montmorillonite reported in previous literature [74,75].
The effect of pH on the adsorption system of 0.1 mM-Sr is manifested in two-stage differentiation (Figure 9a): under acidic conditions (pH < 7), H+ competition leads to partial desorption of Ca2+ and Sr2+ insertion into the interlayer with restricted hydration (d001 ≈ 15.0 Å). However, under alkaline conditions (pH > 7), Ca2+ retention stabilizes, and Sr2+ deep intercalation efficiency increases (d001 ≈ 15.6 Å). Under an extremely alkaline environment (pH > 10), the interlayer hydroxylation leads to the reconfiguration of hydration balance, and the interlayer phenomenon reappears, suggesting the reorganization of the interlayer chemical microenvironment after adsorption saturation. This is attributed to the fact that there was little desorption of Ca2+ after pH > 7, while adsorption saturation was achieved at pH > 10, as indicated by studies of pH absorption edge effects and cation resolution (Figure 4a and Figure 5a).
The comparative study with raw bentonite before adsorption showed that Taiwan Zhi-Shin bentonite displayed higher stability (pH > 7 without interlayer) under the 0.1 mM-Sr system, which may be due to the cushioning effect of kaolinite on the interlayer charge and the synergistic effect of minerals. However, the raw bentonite generally exhibited interlayer shrinkage under the same conditions (d001 recovered to 12.0 Å when pH > 10), highlighting the structural advantages of Taiwan Zhi-Shin bentonite (Figure 9b and Figure 10b). Under the 10 mM-Sr system, the adsorption capacity of Taiwan Zhi-Shin bentonite increased abruptly to 0.23 mmol g−1 when at pH > 9 (Figure 4b), resulting in the expansion of d001 spacing from 12.0 Å to 15.5 Å and the emergence of the interlayer at pH 5, and the interlayer homogenization was gradually completed after pH > 8 (Figure 10a). Combined with the evidence of precipitation of no Sr(OH)2, it was further confirmed that the adsorption mechanism is dominated by ion exchange rather than surface precipitation.
Thus, the adsorption of Sr2+ on Taiwan Zhi-Shin bentonite is based on interlayer ion exchange in the form of montmorillonite (Sr2+ replaces Ca2+ and carries hydrated molecules to vacate the layer intervals), supplemented by a synergistic effect of surface adsorption in kaolinite. The adsorption dynamics were regulated by concentration and pH, with intercalations coexisting at low concentrations and high concentrations or alkaline conditions towards uniform intercalation. Its multi-mechanism coordination and interlayer dynamic adjustment characteristics make it have high adsorption potential in the remediation of radioactive Sr2+ contamination.

2.5. Molecular Simulation

Through AIMD simulation, the adsorption mechanism of Sr2+ on the bentonite surface was studied, with special attention to the molecular scale interaction. As can be seen from Figure 11a, on the surface of Ca-bentonite, hydrated Ca2+ and Sr2+ formed hydrogen bonds with oxygen atoms on the surface of bentonite through surface water. As the simulation progresses, Figure 11b shows that the number of hydrogen bonds formed by Ca2+ and Sr2+ with the bentonite surface increases. Surprisingly, we did not find a directly related adsorption mechanism of Sr2+ on bentonite, given that Sr2+ has similar properties to calcium [2], Moreover, it has been reported that the adsorption of Ca2+ on bentonite is achieved through hydrogen bonding [76], so it can be inferred that Sr2+ is mainly bonded with oxygen atoms on the surface of bentonite through hydrogen bonding by binding water.
To quantify the interaction of Sr2+ on bentonite, adsorption energy calculations were performed on a representative simulation framework exhibiting hydrogen bonds. According to the energy analysis, the relative adsorption energy of Sr2+ on Ca-bentonite is −16.08 eV. For comparison, the adsorption energy of Ca2+ on a molybdenite surface is −10.14 eV [77]. The adsorption energies of Cr2+ on asphaltene-pyrrole and asphaltene-thiophene surfaces are −5.89 eV and −3.08 eV, respectively [78]. The adsorption energy of Ca2+ on the bentonite surface is −15.55 eV [79]. These data indicate that Sr2+ has a strong interaction on Ca-bentonite.
Figure 11c shows that hydrated Ca2+ and Sr2+ are connected by surface water through a bond called a bridge. These bridges are able to increase the adsorption of Sr2+ between bentonite layers. This suggests that hydrogen bonds and water bridges coexist during the adsorption of Sr2+, and the two mechanisms synergistically contribute to the adsorption of Sr2+ to the bentonite surface. Figure 11d shows that there also is a water bridge between Ca2+ and bentonite. Due to the short time of molecular dynamic simulation, we were not able to determine whether hydrogen bonds are formed between Sr2+ and bentonite, but given that Sr2+ has similar properties to calcium [2], it is highly likely that water bridge bonds also exist between Sr2+ and bentonite.
In addition, no other adsorption mechanism was observed through this simulation, so a DFT energy analysis was conducted, and it was determined that the energy of Ca2+ on bentonite is −15.55 eV, which is consistent with the results of Xing et al. [79]. Using the same method, it could be shown that the energy of Sr2+ on bentonite is −15.30 eV. The results show that the adsorption energy of Ca2+ was slightly greater than that of Sr2+ [80]. In addition, through the molecular simulation comparison of Na-bentonite and Ca-bentonite on Sr2+, it was determined that Na-bentonite displayed the same mechanism on Sr2+, mainly hydrogen bonding and water bridging (Figure S1). According to DFT energy analysis, the adsorption energy of Sr2+ on bentonite (15.27 eV) was twice that of Na+ (7.1 eV), so the adsorption of Sr2+ on Na-bentonite involves ion exchange.
Therefore, according to the results of molecular dynamics simulation, it can be inferred that the adsorption mechanism of Sr2+ on Ca-bentonite mainly involves hydrogen bonding and a water bridge. During the adsorption process, the water molecules between the bentonite layers play a synergistic role, which enhances the interaction between Sr2+ and the Taiwan Zhi-Shin bentonite surface.

3. Discussion

The adsorption mechanism of Sr2+ by Taiwan Zhi-Shin bentonite exhibited multi-scale synergistic characteristics. The slope of 0.84 between the adsorption capacity of Sr2+ and the desorption of cations confirms that cation exchange was the primary mechanism for Sr2+ adsorption on Taiwan Zhi-Shin bentonite, with Ca2+ (accounting for 60.4% of the desorbed cations) being the main exchanged ion, while Na+ desorption was relatively low. However, the total desorbed cation quantity was less than the adsorbed Sr2+ concentration, indicating that surface complexation may have also played a significant role alongside cation exchange. This dual-mechanism synergy is further corroborated by XRD data: as Sr2+ concentration increases, the expansion of the (001) crystal plane basal spacing (d001) from 14.71 Å to 15.6 Å demonstrates the existence of interlayer adsorption sites in Taiwan Zhi-Shin bentonite, confirming Sr2+ intercalation into the montmorillonite interlayers. The selective replacement of Ca2+ by Sr2+ in the montmorillonite interlayer space, attributed to the Sr2+ larger ionic radius and its bilayer hydration molecules (adsorption energy: −15.3 eV), leads to ion exchange-driven interlayer expansion and potential local structural reorganization caused by surface complexation, resulting in interstratification phenomena [73,81]. This process aligns with the high adsorption capacity (n > 1.2) revealed by the Freundlich isotherm model, indicating multilayer heterogeneity and energetic heterogeneity of surface adsorption sites [43]. Additionally, the presence of minor kaolinite in Taiwan Zhi-Shin bentonite further enhanced Sr2+ adsorption through synergistic effects.
The adsorption behavior of Sr2+ by Taiwan Zhi-Shin bentonite exhibited significant concentration dependence. At low concentrations, the process was exothermic and entropy-decreasing, dominated by spontaneous adsorption driven by electrostatic interactions and hydration effects. In contrast, at high concentrations, the adsorption transitioned to an endothermic and entropy-increasing mechanism characterized by strong chemical bonding (outer-sphere complexation) [52], demonstrating Taiwan Zhi-Shin bentonite’s broad adaptability in complex aqueous environments. The enthalpy change (ΔH) correlated with concentration gradients, active site density, and chemical bond strength, while the entropy variation (ΔS) likely stemmed from Sr2+ dehydration and enhanced mobility of interlayer water molecules [63], further validating mechanistic differences across concentration regimes. AIMD simulations of Sr2+-loaded adsorption at the molecular scale revealed that Sr2+ and Ca2+ form hydrogen-bond networks via interfacial water molecules bridging oxygen atoms on Taiwan Zhi-Shin bentonite surfaces. The relative adsorption energy of Sr2+ (−15.3 eV) exceeded that of Ca2+ (−12.1 eV), confirming the preferential binding of Sr2+ through stronger interactions. Moreover, Sr2+ and Ca2+ are interconnected by water-bridge structures, which enhance the ordered arrangement of interlayer water molecules, stabilize Sr2+ adsorption sites through coordinated hydration shells, and enable high-efficiency adsorption. This dual mechanism—thermodynamic adaptability and molecular-level structural optimization—explains Taiwan Zhi-Shin bentonite’s exceptional performance in Sr2+ sequestration under varying environmental conditions.
The adsorption kinetics of Sr2+ on Taiwan Zhi-Shin bentonite were shown to be significantly regulated by Sr2+ concentration and solution pH. The adsorption process followed a pseudo-second-order model (R2 = 1.000), achieving rapid equilibrium within 15 min due to the high accessibility of montmorillonite interlayer cation exchange sites [82]. The XRD results confirmed that the hydrated ions of Sr were inserted into the interlayers of Taiwan Zhi-Shin bentonite during adsorption, while pH-dependent studies further revealed the dynamic heterogeneity in adsorption behavior: under low Sr2+ concentrations (0.1 mM), acidic conditions (pH < 7) induced partial Ca2+ desorption via H⁺ competition, allowing Sr2+ insertion in a restricted hydration state (d001 ≈ 15.0 Å); alkaline conditions (pH > 7) enhanced Sr2+ deep intercalation efficiency (d001 ≈ 15.6 Å) by stabilizing Ca2+ retention. Under high-concentration systems (10 mM), the adsorption capacity increased markedly with pH (reaching 0.23 mmol g−1 at pH > 9), accompanied by d001 expansion from 12.0 Å to 15.5 Å, indicating complete Sr2+ occupancy of interlayer sites and suppression of interstratification. Additionally, elevated Sr(OH)⁺ concentrations at pH > 9 enhanced adsorption via electrostatic attraction, while weakened Ca2+ competitive displacement due to precipitation formation in alkaline conditions further optimized efficiency. The synergistic interplay of cation exchange, surface complexation, electrostatic interactions, water-bridge assistance, and dynamic interlayer adjustments enables Taiwan Zhi-Shin bentonite to achieve high-efficiency Sr2+ removal (>95% within 5 min) across 0.1–10 mM concentrations, with superior ion interference resistance compared to synthetic materials, providing a natural mineral-based solution for rapid Sr2+ treatment in nuclear wastewater.

4. Materials and Methods

4.1. Materials

4.1.1. Preparation of Simulated Nuclear Waste Liquid

Sr is an alkaline earth metal with a molecular weight of 87.62 g mol−1, a density of 2.60 g cm−3, a boiling point of 1384 °C, and a melting point of 769 °C [83]. It is an active positively charged metal ion that is easily oxidized to the stable and colorless Sr2+, with chemical properties similar to calcium and barium [1]. In this study, Sr(NO3)2 was used to prepare a simulated nuclear waste solution containing Sr2+ with a pH of 7 and a mass concentration of 0.5 mM.

4.1.2. Taiwan Zhi-Shin Bentonite

Taiwan Zhi-Shin bentonite is produced by Zhi Hing Mining Co., Ltd. (Yunlin, Taiwai), which is located in the Duluan Mountain layer of Zhangyuan Village, Changbin Township, Taitung County. Taiwan Zhi-Shin bentonite was adopted for these experiments. The sample was lumpy and was pulverized into powder for use without further purification. Through our literature investigation, the CEC of Taiwan Zhi-Shin bentonite was determined to be 59, 60.1~72.0, 86 meq 100 g−1. This may be due to differences in sampling time and location. The Taiwan Zhi-Shin bentonite in this study is 80~86 meq 100 g−1 after three measurements with the ammonium saturation method, and the main exchange cation was shown to be Ca2+ [84]. The experimental results show that the PZC is 9.2. The chemical composition of the Taiwan Zhi-Shin bentonite is given in Table S4.

4.2. Experimental Methods

4.2.1. Batch Experiment

An amount of 1 g of Taiwan Zhi-Shin bentonite was added to several centrifuge tubes containing 30 mL of simulated nuclear waste liquid (S/L = 33.3 g L−1), and then the centrifuge tubes were placed in an oscillator (150 r min−1) to carry out oscillatory adsorption experiments at room temperature. In the dynamic experiment, the centrifugal tubes were removed successively after 1, 2, 5, and 10 min and 0.25, 0.5, 1, 2, 4, 8, 16, and 24 h of oscillations. In the other batches of experiments, the tubes were removed after 24 h. The tube was centrifuged at 8000 r min−1 for 5 min, and the supernatant was collected in plastic bottles.
In the adsorption isotherm experiment for Sr2+, the initial solution concentration was set at 0.05, 0.08, 0.1, 0.5, 1, 5, 10, and 15 mmol L−1, and the solid-liquid ratio was 33.3 g L−1. The oscillatory adsorption experiment was carried out in an oscillator (150 r min−1) at room temperature. After 24 h, the centrifuge tubes were removed and centrifuged at 8000 r min−1 for 5 min, and the supernatant was collected in the spare plastic bottle.
In the other group of experiments, the initial concentration of Sr2+ was about 10 mmol L−1, and the remaining experimental conditions were the same. An amount of 30 mL of Sr2+ solution with a concentration of 0.1 mmol L−1 was added to the plastic centrifuge tube, and 1 g of Taiwan Zhi-Shin bentonite was added to each centrifuge tube, which was put into the shaking table. The temperature was maintained at 25 °C, and the rotating speed was set at 150 rpm. The pH was adjusted to between 2 and 11 using diluted hydrochloric acid or sodium hydroxide solution. After oscillating for 1, 2, 5, and 10 min and 0.25, 0.5, 1, 2, 4, 8, 16, and 24 h, the centrifuge tubes were removed successively, the supernatant was centrifuged on the centrifuge (8000 r min−1) for 5 min, and the supernatant was collected in the spare plastic bottle.

4.2.2. Analytical Methods

Sr2+ concentrations were measured by inductively coupled plasma optical emission spectroscopy (Vista-MPX, Agilent Technologies Inc., Jinan, China). Samples were diluted and acidified in 1% HNO3 solution. The analysis device adopted an MPX CCD detector and RF generator and applied direct serial coupling technology. The V-groove atomizer and the two-pass Sturman–Masters atomizer were introduced, and the peristaltic pump was also set up. In addition, the spectrometer operated in a transient signal acquisition mode [85].
XRD analysis was performed on a Rigaku Ultima IV (Nanjing, China) diffractometer equipped with a D/teX Ultra (CuKα radiation) detector using X-ray diffractometer (XRD) at a scanning speed of 2° min−1 and a step increment of 0.01 in the range of 2θ 2–65°. The adsorption mechanism of Sr2+ on the Taiwan Zhi-Shin bentonite surface was analyzed.

4.2.3. Molecular Simulation

The structure of bentonite consists of two layers of Si-O tetrahedrons and an Al-O octahedron sandwiched between them [31]. In order to improve computational efficiency, we constructed a cell model with a size of 2a × 1b × 1c (17.96 Å × 10.36 Å × 39.00 Å) supercell. The distance between bentonite layers measured by experiment is 14.7 Å. In the lattice, four Al3+ are replaced by Mg2+ and Ca2+ ions are placed in the interlayer space to maintain charge balance. In addition, 22 water molecules were introduced into the interlayer region, of which six formed a hydrated shell around the Ca2+ ion and another six formed a hydrated shell around the Sr2+ [86,87].
The present study utilized the Quickstep module of the CP2K package [88] to perform ab initio molecular dynamics (AIMD) simulations of Sr2+ ions in Ca-bentonite based on density functional theory (DFT) [89]. The Perdew–Burke–Ernzerhof (PBE) [90] exchange correlation function combined with Grimme’s DFT-D3 correction [91] was used to describe the dispersion effect. The Goedeck–Teter–Hutter (GTH) [92] pseudopotential was used to model the interaction between the core electrons and the nucleus. Due to the large number of electrons involved, we used GTH bi-zeta valence polarization (DZVP) Gaussian basis combined with plane wave basis function and set the energy cutoff value to 500 Ry. The orbital transformation (OT) method [93] was used to optimize the wave function, and the strict convergence threshold was set to 1 × 10−6 atomic units. In the simulation of molecular dynamics, we used a canonical (NVT) [94] ensemble with a temperature set at 330 K and a Nosé–Hoover thermostat [95] to maintain the temperature. The total simulation time was approximately 856 fs with a time step of 0.5 fs to ensure a thorough analysis.

5. Conclusions

The absorption of Taiwan Zhi-Shin bentonite, which is dominated by calcium-based montmorillonite, is mainly based on ion exchange and surface complexation of montmorillonite crystals at low pH and Sr2+ concentration. In highly concentrated Sr2+ solutions, the adsorption is mainly based on ion exchange between the clay-crystal cations and the external medium Sr2+. At pH > 10 electrostatic interactions were also shown to be involved, and the hydration ions of Sr2+ and Ca2+ may form a water bridge co-existing between mineral layers. Minerals like bentonite are abundant in output. Moreover, compared with other adsorbents used in similar processes, Zhi-Shin bentonite has the characteristics of simple operation and low cost.
However, the expansion performance of Taiwan Zhi-Shin bentonite is not as good as that of sodium bentonite abroad, so it is usually necessary to improve the expansion characteristics through physical, chemical, heat treatment, and ion exchange methods to improve its competitiveness as a buffer material.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms26115298/s1. References [96,97,98,99,100,101] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, Y.L. (Yihui Lin) and P.-H.C.; methodology, Y.L. (Yihui Lin); software, Y.L. (Yuhan Li); validation, Y.Y. and Y.L. (Yuhan Li); formal analysis, Y.Y. and Y.L. (Yuhan Li); investigation, P.-H.C.; resources, P.-H.C.; data curation, Y.Y. and Y.L. (Yuhan Li); writing—original draft preparation, Y.Y. and Y.L. (Yuhan Li); writing—review and editing, Y.L. (Yihui Lin), P.-H.C., Y.Y. and Y.L. (Yuhan Li); All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

We sincerely want to thank Christopher Rensing, a distinguished highly-cited scholar and native English speaker, for his meticulous language refinement and grammatical expertise.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mandakhbayar, N.; El-Fiqi, A.; Lee, J.-H.; Kim, H.-W. Evaluation of Strontium-Doped Nanobioactive Glass Cement for Dentin–Pulp Complex Regeneration Therapy. ACS Biomater. Sci. Eng. 2019, 5, 6117–6126. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, M.; Gu, P.; Zhang, Z.; Liu, J.; Dong, L.; Zhang, G. Effective, rapid and selective adsorption of radioactive Sr2+ from aqueous solution by a novel metal sulfide adsorbent. Chem. Eng. J. 2018, 351, 668–677. [Google Scholar] [CrossRef]
  3. Ghandhi, S.A.; Weber, W.; Melo, D.; Doyle-Eisele, M.; Chowdhury, M.; Guilmette, R.; Amundson, S.A. Effect of 90Sr internal emitter on gene expression in mouse blood. BMC Genom. 2015, 16, 586. [Google Scholar] [CrossRef]
  4. Ghoti, H.; Amer, J.; Winder, A.; Rachmilewitz, E.; Fibach, E. Oxidative stress in red blood cells, platelets and polymorphonuclear leukocytes from patients with myelodysplastic syndrome. Eur. J. Haematol. 2007, 79, 463–467. [Google Scholar] [CrossRef]
  5. Aarkrog, A.; Baxter, M.S.; Bettencourt, A.O.; Bojanowski, R.; Bologa, A.; Charmasson, S.; Cunha, I.; Delfanti, R.; Duran, E.; Holm, E.; et al. A comparison of doses from 137Cs and 210Po in marine food: A major international study. J. Environ. Radioactiv. 1997, 34, 69–90. [Google Scholar] [CrossRef]
  6. Povinec, P.P.; Hirose, K.; Aoyama, M. Radiostrontium in the western North Pacific: Characteristics, behavior, and the Fukushima impact. Environ. Sci. Technol. 2012, 46, 10356–10363. [Google Scholar] [CrossRef] [PubMed]
  7. Peng, H.; Yao, F.; Xiong, S.; Wu, Z.; Niu, G.; Lu, T. Strontium in public drinking water and associated public health risks in Chinese cities. Environ. Sci. Pollut. Res. 2021, 28, 23048–23059. [Google Scholar] [CrossRef]
  8. Adabbo, M.; Caputo, D.; de Gennaro, B.; Pansini, M.; Colella, C. Ion exchange selectivity of phillipsite for Cs and Sr as a function of framework composition. Micropor. Mesopor. Mater. 1999, 28, 315–324. [Google Scholar] [CrossRef]
  9. Rout, T.K.; Sengupta, D.K.; Kaur, G.; Kumar, S. Enhanced removal of dissolved metal ions in radioactive effluents by flocculation. Int. J. Miner. Process. 2016, 80, 215–222. [Google Scholar] [CrossRef]
  10. Jia, F.; Li, J.; Wang, J.; Sun, Y. Removal of strontium ions from simulated radioactive wastewater by vacuum membrane distillation. Ann. Nucl. Energy 2017, 103, 363–368. [Google Scholar] [CrossRef]
  11. Su, Z.; Deng, Z.; Wang, Y.; Ji, C.; Li, F.; Yang, G.; Huang, L. Effects of the Sr/Ca ratio on the bioremediation of strontium based on microbially-induced carbonate precipitation. J. Environ. Chem. Eng. 2023, 11, 108990. [Google Scholar] [CrossRef]
  12. Chen, K.; Chen, C.; Ren, X.; Alsaedi, A.; Hayat, T. Interaction mechanism between different facet TiO2 and U (VI): Experimental and density-functional theory investigation. Chem. Eng. J. 2019, 359, 944–954. [Google Scholar] [CrossRef]
  13. Marinović, S.S.; Ajduković, M.J.; Jović-Jovičić, N.J.; Mudrinić, T.M.; Nedić-Vasiljević, B.N.; Banković, P.T.; Milutinović-Nikolić, A.D. Adsorption of strontium on different sodium enriched bentonites. J. Serb. Chem. Soc. 2017, 82, 449–463. [Google Scholar] [CrossRef]
  14. Ma, H.; Shen, M.; Tong, Y.; Wang, X. Radioactive wastewater treatment technologies: A review. Molecules 2023, 28, 1935. [Google Scholar] [CrossRef] [PubMed]
  15. Shukla, A.; Parmar, P.; Saraf, M. Radiation, radionuclides and bacteria: An in-perspective review. J. Environ. Radioact. 2017, 180, 27–35. [Google Scholar] [CrossRef]
  16. Vanhoudt, N.; Vandenhove, H.; Leys, N.; Janssen, P. Potential of higher plants, algae, and cyanobacteria for remediation of radioactively contaminated waters. Chemosphere 2018, 207, 239–254. [Google Scholar] [CrossRef]
  17. Cheng, Y.; Chuah, G.K. The synthesis and applications of α-zirconium phosphate. Chin Chem. Lett. 2020, 31, 307–310. [Google Scholar] [CrossRef]
  18. He, W.; Ai, K.; Ren, X.; Wang, S.; Lu, L. Inorganic layered ion-exchangers for decontamination of toxic metal ions in aquatic systems. J. Mater. Chem. A 2017, 5, 19593–19606. [Google Scholar] [CrossRef]
  19. Zakrzewska-Trznadel, G. Advances in membrane technologies for the treatment of liquid radioactive waste. Desalination 2013, 321, 119–130. [Google Scholar] [CrossRef]
  20. Liu, C.; Su, G.; Chen, T.; Liang, M.; Zhao, L.; Zhang, D. Preparation of sulfonylcalix [4] arene-loaded XAD-7 resin for strontium (II) adsorption. IOP Conf. Ser. Earth Environ. Sci. 2019, 267, 032084. [Google Scholar] [CrossRef]
  21. Hafizi, M.; Abolghasemi, H.; Moradi, M.; Milani, S.A. Strontium adsorption from sulfuric acid solution by Dowex 50W-X resins. Chin. J. Chem. Eng. 2011, 19, 267–272. [Google Scholar] [CrossRef]
  22. Lv, T.T.; Ma, W.; Zhang, D.; Zhang, T.; Tang, J.H.; Zeng, X.; Feng, M.L.; Huang, X.Y. Rapid and highly selective Sr2+ uptake by 3D microporous rare earth oxalates with the facile synthesis, high water stability and radiation resistance. Chem. Eng. J. 2022, 435, 134906. [Google Scholar] [CrossRef]
  23. Ogata, F.; Kobayashi, Y.; Uematsu, Y.; Nakamura, T.; Kawasaki, N. Zeolite produced from fly ash by thermal treatment in alkaline solution and its capability to adsorb Cs (I) and Sr (II) in aqueous solution. Yakugaku Zasshi-J. Pharm. Soc. Jpn. 2020, 140, 729–737. [Google Scholar] [CrossRef]
  24. Abu-Nada, A.; Abdala, A.; McKay, G. Isotherm and Kinetic Modeling of Strontium Adsorption on Graphene Oxide. Nanomaterials 2021, 11, 2780. [Google Scholar] [CrossRef] [PubMed]
  25. Naeimi, S.; Faghihian, H. Modification and magnetization of MOF (HKUST-1) for the removal of Sr2+ from aqueous solutions. Equilibrium, kinetic and thermodynamic modeling studies. Sep. Sci. Technol. 2017, 52, 2899–2908. [Google Scholar] [CrossRef]
  26. Zheng, B.; Yin, J.; Zhu, L.; Zhou, B.; Shen, H.; Harbottle, D.; Hunter, T.N.; Sheng, Y.; Zhu, Q.; Zhang, H. Thiol-rich and ion-imprinted alginate hydrogel as a highly adsorptive and recyclable filtration membrane for rapid and selective Sr (II) removal. Chem. Eng. J. 2023, 465, 142752. [Google Scholar] [CrossRef]
  27. Zhang, J.; Chen, L.; Dai, X.; Zhu, L.; Xiao, C.; Xu, L.; Zhang, Z.; Alekseev, E.V.; Wang, Y.; Zhang, C.; et al. Distinctive two-step intercalation of Sr2+ into a coordination polymer with record high 90Sr uptake capabilities. Chem 2019, 5, 977–994. [Google Scholar] [CrossRef]
  28. Zhang, L.; Wei, J.; Zhao, X.; Li, F.; Jiang, F. Adsorption characteristics of strontium on synthesized antimony silicate. Chem. Eng. J. 2015, 277, 378–387. [Google Scholar] [CrossRef]
  29. Chegrouche, S.; Mellah, A.; Barkat, M. Removal of strontium from aqueous solutions by adsorption onto activated carbon: Kinetic and thermodynamic studies. Desalinatio 2009, 235, 306–318. [Google Scholar] [CrossRef]
  30. Moloukhia, H.; Hegazy, W.S.; Abdel-Galil, E.A.; Mahrous, S.S. Removal of Eu3+, Ce3+, Sr2+, and Cs+ ions from radioactive waste solutions by modified activated carbon prepared from coconut shells. Chem. Ecol. 2016, 32, 324–345. [Google Scholar] [CrossRef]
  31. Van, O.H.; Fripiar, J.J. Data handbook for clay materials and other non-metallic minerals. Soil Sci. 1981, 131, 62. [Google Scholar]
  32. CMS Industries. Bentonite Manufacturers, Suppliers and Exporters in Taiwan Materials. Available online: https://www.cmsindustries.in/ (accessed on 10 January 2024).
  33. Murali, M.S.; Mathur, J.N. Sorption characteristics of Am (III), Sr (II) and Cs (I) on bentonite and granite. J. Radioanal. Nucl. Chem. 2002, 254, 129–136. [Google Scholar] [CrossRef]
  34. Khan, S.A.; Khan, M.A. Sorption of strontium on bentonite. Waste Manag. 1995, 15, 641–650. [Google Scholar] [CrossRef]
  35. Mohajeri, P.; Smith, C.; Selamat, M.R.; Abdul Aziz, H. Enhancing the adsorption of lead (II) by bentonite enriched with pH-adjusted meranti sawdust. Water 2018, 10, 1875. [Google Scholar] [CrossRef]
  36. He, H.; Xu, E.; Qiu, Z.; Wu, T.; Wang, S.; Lu, Y.; Chen, G. Phenol adsorption mechanism of organically modified bentonite and its microstructural changes. Sustainability 2022, 14, 1318. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Gu, P.; Zhang, M.; Yan, S.; Dong, L.; Zhang, G. Synthesis of a robust layered metal sulfide for rapid and effective removal of Sr2+ from aqueous solutions. Chem. Eng. J. 2019, 372, 1205–1215. [Google Scholar] [CrossRef]
  38. Izosimova, Y.; Gurova, I.; Tolpeshta, I.; Karpukhin, M.; Zakusin, S.; Zakusina, O.; Samburshiy, A.; Krupskaya, V. Adsorption of Cs (I) and Sr (II) on bentonites with different compositions at different pH. Minerals 2022, 12, 862. [Google Scholar] [CrossRef]
  39. Yang, X.; Zhou, Y.; Hu, J.; Zheng, Q.; Zhao, Y.; Lv, G.; Liao, L. Clay minerals and clay-based materials for heavy metals pollution control. Sci. Total Environ. 2024, 954, 176193. [Google Scholar] [CrossRef]
  40. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  41. Wu, P.; Dai, Y.; Long, H.; Zhu, N.; Li, P.; Wu, J.; Dang, Z. Characterization of organo-montmorillonites and comparison for Sr (II) removal: Equilibrium and kinetic studies. Chem. Eng. J. 2012, 191, 288–296. [Google Scholar] [CrossRef]
  42. Pabalan, R.T.; Bertetti, F.P. Cation-exchange properties of natural zeolites. Rev. Mineral. Geochem. 2001, 45, 453–518. [Google Scholar] [CrossRef]
  43. Freundlich, H. Über die adsorption in lösungen. Z. Phys. Chem. 1907, 57, 385–470. [Google Scholar] [CrossRef]
  44. Rahman, R.O.A.; Ibrahim, H.A.; Hanafy, M.; Monem, N.M.A. Assessment of synthetic zeolite Na A–X as sorbing barrier for strontium in a radioactive disposal facility. Chem. Eng. J. 2010, 157, 100–112. [Google Scholar] [CrossRef]
  45. Molera, M.; Eriksen, T. Diffusion of 22Na+, 85Sr2+, 134Cs+ and 57Co2+ in bentonite clay compacted to different densities: Experiments and modeling. Radiochim. Acta 2002, 90, 753–760. [Google Scholar] [CrossRef]
  46. Li, Y.; Wang, X.; Wang, J. Cation exchange, interlayer spacing, and thermal analysis of Na/Ca-montmorillonite modified with alkaline and alkaline earth metal ions. J. Therm. Anal. Calorim. 2012, 110, 1199–1206. [Google Scholar] [CrossRef]
  47. Tong, Y.; Zhang, H.; Li, X.; Jia, Q. Experimental study on sodium modification and purification of GMZ bentonite. Constr. Build. Mater. 2023, 367, 130060. [Google Scholar] [CrossRef]
  48. Alvarez-Ayuso, E.; Garcia-Sanchez, A. Removal of heavy metals from waste waters by natural and Na-exchanged bentonites. Clays Clay Miner. 2003, 51, 475–480. [Google Scholar] [CrossRef]
  49. Lee, C.P.; Tsai, S.C.; Wu, M.C.; Tsai, T.L. A study on removal of Cs and Sr from aqueous solution by bentonite–alginate microcapsules. J. Radioanal. Nucl. Chem. 2018, 318, 2381–2387. [Google Scholar] [CrossRef]
  50. Young, D.A.; Smith, D.E. Simulations of clay mineral swelling and hydration: Dependence upon interlayer ion size and charge. J. Phys. Chem. B 2000, 104, 9163–9170. [Google Scholar] [CrossRef]
  51. Nilsson, K.K.; Jensen, B.S.; Carlsen, L. The Migration Chemistry of Strontium; European applied research reports nuclear science and technology section; Harwood Academic Publishers: Reading, UK, 1985; pp. 149–200. Available online: https://books.google.rs/books/about/The_Migration_Chemistry_of_Strontium.html?id=gV970AEACAAJ&redir_esc=y (accessed on 21 May 2025).
  52. Carroll, S.A.; Roberts, S.K.; Criscenti, L.J.; O’Day, P.A. Surface complexation model for strontium sorption to amorphous silica and goethite. Geochem. Trans. 2008, 9, 2. [Google Scholar] [CrossRef]
  53. Zhang, S. A new nano-sized calcium hydroxide photocatalytic material for the photodegradation of organic dyes. Rsc Adv. 2014, 4, 15835–15840. [Google Scholar] [CrossRef]
  54. Janusz, W.; Skwarek, E. Study of sorption processes of strontium on the synthetic hydroxyapatite. Adsorption 2016, 22, 697–706. [Google Scholar] [CrossRef]
  55. Chen, C.C.; Hayes, K.F. X-ray absorption spectroscopy investigation of aqueous Co (II) and Sr (II) sorption at clay–water interfaces. Geochim. Cosmochim. Acta 1999, 63, 3205–3215. [Google Scholar] [CrossRef]
  56. Edet, U.A.; Ifelebuegu, A.O. Kinetics, isotherms, and thermodynamic modeling of the adsorption of phosphates from model wastewater using recycled brick waste. Processes 2020, 8, 665. [Google Scholar] [CrossRef]
  57. Ghaemi, A.; Torab-Mostaedi, M.; Ghannadi-Maragheh, M. Characterizations of strontium (II) and barium (II) adsorption from aqueous solutions using dolomite powder. J. Hazard. Mater. 2011, 190, 916–921. [Google Scholar] [CrossRef]
  58. Yusan, S.; Erenturk, S. Adsorption characterization of strontium on PAN/zeolite composite adsorbent. World J. Nucl. Sci. Technol. 2011, 1, 6–12. [Google Scholar] [CrossRef]
  59. Kilincarslan Kaygun, A.; Eral, M.; Akyil Erenturk, S. Removal of cesium and strontium using natural attapulgite: Evaluation of adsorption isotherm and thermodynamic data. J. Radioanal. Nucl. Chem. 2017, 311, 1459–1464. [Google Scholar] [CrossRef]
  60. Khan, S.; Riaz-Ur-Reman; Khan, M. Adsorption of Cs (I), Sr (II) and Co (II) on Al2O3. J. Radioanal. Nucl. Chem. 1995, 190, 81–96. [Google Scholar] [CrossRef]
  61. Celebi, O.; Kilikli, A.; Erten, H.N. Sorption of radioactive cesium and barium ions onto solid humic acid. J. Hazard. Mater. 2009, 168, 695–703. [Google Scholar] [CrossRef]
  62. Möller, T.; Clearfield, A.; Harjula, R. Preparation of hydrous mixed metal oxides of Sb, Nb, Si, Ti and W with a pyrochlore structure and exchange of radioactive cesium and strontium ions into the materials. Micropor. Mesopor. Mat. 2002, 54, 187–199. [Google Scholar] [CrossRef]
  63. İnan, S.; Altaş, Y. Preparation of zirconium–manganese oxide/polyacrylonitrile (Zr–Mn oxide/PAN) composite spheres and the investigation of Sr (II) sorption by experimental design. Chem. Eng. J. 2011, 168, 1263–1271. [Google Scholar] [CrossRef]
  64. Bérend, I.; Cases, J.M.; François, M.; Uriot, J.P.; Michot, L.; Masion, A.; Thomas, F. Mechanism of adsorption and desorption of water vapor by homoionic montmorillonites: 2. The Li+, Na+, K+, Rb+ and Cs+-exchanged forms. Clays Clay Miner. 1995, 43, 324–336. [Google Scholar] [CrossRef]
  65. Cases, J.M.; Bérend, I.; François, M.; Uriot, J.P.; Michot, L.J.; Thomas, F. Mechanism of adsorption and desorption of water vapor by homoionic montmorillonite: 3. The Mg2+, Ca2+, Sr2+ and Ba2+ exchanged forms. Clays Clay Miner. 1997, 45, 8–22. [Google Scholar] [CrossRef]
  66. Sánchez, R.T. Mechanochemical effects on physicochemical parameters of homoionic smectite. Colloids Surf. Physicochem. Eng. Asp. 1997, 127, 135–140. [Google Scholar] [CrossRef]
  67. Sohrabnezhad, S.; Pourahmad, A.; Razavi, M. Silver bromide in montmorillonite as visible light-driven photocatalyst and the role of montmorillonite. Appl. Phys. A Mater. Sci. Process. 2016, 122, 822. [Google Scholar] [CrossRef]
  68. Berghout, A.; Tunega, D.; Zaoui, A. Density functional theory (DFT) study of the hydration steps of Na+/Mg2+/Ca2+/Sr2+/Ba2+-exchanged montmorillonites. Clays Clay Miner. 2010, 58, 174–187. [Google Scholar] [CrossRef]
  69. Singh, B.; Gilkes, R.J. Properties of soil kaolinites from south-western Australia. J. Soil Sci. 1992, 43, 645–667. [Google Scholar] [CrossRef]
  70. Ning, Z.; Ishiguro, M.; Koopal, L.K.; Sato, T.; Kashiwagi, J.I. Strontium adsorption and penetration in kaolinite at low Sr2+ concentration. Soil Sci. Plant Nutr. 2017, 63, 14–17. [Google Scholar] [CrossRef]
  71. Keçeli, G. Adsorption kinetics and equilibria of strontium onto kaolinite. Sep. Sci. Technol. 2015, 50, 72–80. [Google Scholar] [CrossRef]
  72. Holmboe, M.; Wold, S.; Jonsson, M. Porosity investigation of compacted bentonite using XRD profile modeling. J. Contam. Hydrol. 2012, 128, 19–32. [Google Scholar] [CrossRef]
  73. Ferrage, E.; Lanson, B.; Sakharov, B.A.; Drits, V.A. Investigation of smectite hydration properties by modeling experimental X-ray diffraction patterns: Part I. Montmorillonite hydration properties. Am. Mineral. 2005, 90, 1358–1374. [Google Scholar] [CrossRef]
  74. Laird, D.A. Influence of layer charge on swelling of smectites. Appl. Clay Sci. 2006, 34, 74–87. [Google Scholar] [CrossRef]
  75. Saiyouri, N.; Tessier, D.; Hicher, P.Y. Experimental study of swelling in unsaturated compacted clays. Clay Miner. 2004, 39, 469–479. [Google Scholar] [CrossRef]
  76. Chen, W.C.; Huang, W.H. Effect of groundwater chemistry on the swelling behavior of a Ca-bentonite for deep geological repository. Phys. Chem. Earth. Parts A/B/C 2013, 65, 42–49. [Google Scholar] [CrossRef]
  77. Sun, L.; Cao, Y.; Li, L.; Zeng, Q. Adsorption Characteristics and Mechanism of Calcium Ions on Different Molybdenite Surfaces via Experiments and DFT Simulations. Separations 2021, 8, 107. [Google Scholar] [CrossRef]
  78. Yaphary, Y.L.; He, M.; Lu, G.; Zou, F.; Liu, P.; Tsang, D.C.; Leng, Z. Experiment and multiscale molecular simulations on the Cu absorption by biochar-modified asphalt: An insight into removal capability and mechanism of heavy metals from stormwater runoff. Chem. Eng. J. 2023, 462, 142205. [Google Scholar] [CrossRef]
  79. Xing, X.; Lv, G.; Zhu, W.; He, C.; Liao, L.; Mei, L.; Li, Z.; Li, G. The binding energy between the interlayer cations and montmorillonite layers and its influence on Pb2+ adsorption. Appl. Clay Sci. 2015, 112, 117–122. [Google Scholar] [CrossRef]
  80. Li, Y.W.; Schulthess, C.P. Ion-exchange modeling of divalent cation adsorption on SWy-3 montmorillonite. Clays Clay Miner. 2021, 69, 167–187. [Google Scholar] [CrossRef]
  81. Aristilde, L.; Lanson, B.; Charlet, L. Interstratification patterns from the pH-dependent intercalation of a tetracycline antibiotic within montmorillonite layers. Langmuir 2013, 29, 4492–4501. [Google Scholar] [CrossRef]
  82. Wang, C.J.; Li, Z.; Jiang, W.T.; Jean, J.S.; Liu, C.C. Cation exchange interaction between antibiotic ciprofloxacin and montmorillonite. J. Hazard. Mater. 2010, 183, 309–314. [Google Scholar] [CrossRef]
  83. Lide, D.R. (Ed.) CRC Handbook of Chemistry and Physics, 85th ed.; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  84. Goo, J.Y.; Kim, B.J.; Kwon, J.S.; Jo, H.Y. Bentonite alteration and retention of cesium and iodide ions by Ca-bentonite in alkaline and saline solutions. Appl. Clay Sci. 2023, 245, 107141. [Google Scholar] [CrossRef]
  85. Sun, A.; Xu, Q.C.; Xu, S.J.; Shang, G.X.H.; Sun, J. Ultrasonic-assisted dispersive liquid phase microextration with low-volume toxic solvent for the separation of strontium and its application for strontium isotope determination. Asian J. Chem. 2014, 26, 169–172. [Google Scholar] [CrossRef]
  86. Katz, A.K.; Glusker, J.P.; Beebe, S.A.; Bock, C.W. Calcium ion coordination: A comparison with that of beryllium, magnesium, and zinc. J. Am. Chem. Soc. 1996, 118, 5752–5763. [Google Scholar] [CrossRef]
  87. Borchert, M.; Wilke, M.; Schmidt, C.; Kvashnina, K.; Jahn, S. Strontium complexation in aqueous solutions and silicate glasses: Insights from high energy-resolution fluorescence detection X-ray spectroscopy and ab-initio modeling. Geochim. Cosmochim. Acta 2014, 142, 535–552. [Google Scholar] [CrossRef]
  88. Kühne, T.D.; Iannuzzi, M.D.; Ben, M.; Rybkin, V.V.; Seewald, P.; Stein, F.; Hutter, J. CP2K: An electronic structure and molecular dynamics software package-Quickstep: Efficient and accurate electronic structure calculations. J. Chem. Phys. 2020, 152, 194103. [Google Scholar] [CrossRef] [PubMed]
  89. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef]
  90. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  91. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  92. Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703. [Google Scholar] [CrossRef]
  93. VandeVondele, J.; Hutter, J. An efficient orbital transformation method for electronic structure calculations. J. Chem. Phys. 2003, 118, 4365–4369. [Google Scholar] [CrossRef]
  94. Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511–519. [Google Scholar] [CrossRef]
  95. Hoover, W.G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695. [Google Scholar] [CrossRef] [PubMed]
  96. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  97. Ali, M.M.S.; Abdel-Galil, E.A.; Hamed, M.M. Removal of strontium radionuclides from liquid scintillation waste and environmental water samples. Appl. Radiat. Isot. 2020, 166, 109357. [Google Scholar] [CrossRef]
  98. Aytas, S.; Yurtlu, M.; Donat, R. Adsorption characteristic of U (VI) ion onto thermally activated bentonite. J. Hazard. Mater. 2009, 172, 667–674. [Google Scholar] [CrossRef] [PubMed]
  99. Simon, J.A.; Abrams, S.; Bradburne, T.; Lester, D.; Storch, P.; Remy, J.; Dindal, A. PFAS Experts Symposium: Statements on regulatory policy, chemistry and analytics, toxicology, transport/fate, and remediation for per-and polyfluoroalkyl substances (PFAS) contamination issues. Remediat. J. 2009, 29, 31–48. [Google Scholar] [CrossRef]
  100. Rogers, H.; Bowers, J.; Gates-Anderson, D. An isotope dilution–precipitation process for removing radioactive cesium from wastewater. J. Hazard. Mater. 2012, 243, 124–129. [Google Scholar] [CrossRef]
  101. Oleksiienko, O.; Wolkersdorfer, C.; Sillanpää, M. Titanosilicates in Cation Adsorption and Cation Exchange—A Review. Chem. Eng. J. 2017, 317, 570–585. [Google Scholar] [CrossRef]
Figure 1. The adsorption kinetics of 0.1 mM-Sr (a) and 10 mM-Sr (b) on Taiwan Zhi-Shin bentonite. The solid line is pseudo-second-order fit to the observed data. The dashed green line represents the time when equilibrium was reached. Insert is the linear plot of the equation.
Figure 1. The adsorption kinetics of 0.1 mM-Sr (a) and 10 mM-Sr (b) on Taiwan Zhi-Shin bentonite. The solid line is pseudo-second-order fit to the observed data. The dashed green line represents the time when equilibrium was reached. Insert is the linear plot of the equation.
Ijms 26 05298 g001
Figure 2. Adsorption of Sr2+ on Taiwan Zhi-Shin bentonite. The solid lines are the Langmuir and Freundlich fit to the observed data.
Figure 2. Adsorption of Sr2+ on Taiwan Zhi-Shin bentonite. The solid lines are the Langmuir and Freundlich fit to the observed data.
Ijms 26 05298 g002
Figure 3. Desorption of metal cations from Taiwan Zhi-Shin bentonite as affected by the amount of Sr2+ adsorption.
Figure 3. Desorption of metal cations from Taiwan Zhi-Shin bentonite as affected by the amount of Sr2+ adsorption.
Ijms 26 05298 g003
Figure 4. The 0.1 mM-Sr (a) and 10 mM-Sr (b) adsorption on Taiwan Zhi-Shin bentonite as affected by equilibrium solution pH.
Figure 4. The 0.1 mM-Sr (a) and 10 mM-Sr (b) adsorption on Taiwan Zhi-Shin bentonite as affected by equilibrium solution pH.
Ijms 26 05298 g004
Figure 5. Desorption of metal cations from Taiwan Zhi-Shin bentonite as affected by the pH values under the amount of 0.1 mM-Sr adsorption (a) and 10 mM-Sr (b).
Figure 5. Desorption of metal cations from Taiwan Zhi-Shin bentonite as affected by the pH values under the amount of 0.1 mM-Sr adsorption (a) and 10 mM-Sr (b).
Ijms 26 05298 g005
Figure 6. 0.1 mM-Sr adsorption (a) and 10 mM-Sr (b) adsorption on Taiwan Zhi-Shin bentonite as affected by solution ionic strength.
Figure 6. 0.1 mM-Sr adsorption (a) and 10 mM-Sr (b) adsorption on Taiwan Zhi-Shin bentonite as affected by solution ionic strength.
Ijms 26 05298 g006
Figure 7. The 0.1 mM-Sr adsorption and 10 mM-Sr adsorption on Taiwan Zhi-Shin bentonite as affected by equilibrium temperature.
Figure 7. The 0.1 mM-Sr adsorption and 10 mM-Sr adsorption on Taiwan Zhi-Shin bentonite as affected by equilibrium temperature.
Ijms 26 05298 g007
Figure 8. XRD pattern of raw Taiwan Zhi-Shin bentonite and Sr-adsorbed Taiwan Zhi-Shin bentonite under different adsorbed amounts.
Figure 8. XRD pattern of raw Taiwan Zhi-Shin bentonite and Sr-adsorbed Taiwan Zhi-Shin bentonite under different adsorbed amounts.
Ijms 26 05298 g008
Figure 9. XRD pattern of 0.1 mM-Sr adsorbed on Taiwan Zhi-Shin bentonite (a) and raw bentonite (b) under different pHs.
Figure 9. XRD pattern of 0.1 mM-Sr adsorbed on Taiwan Zhi-Shin bentonite (a) and raw bentonite (b) under different pHs.
Ijms 26 05298 g009
Figure 10. XRD pattern of 10 mM-Sr adsorbed on Taiwan Zhi-Shin bentonite (a) and raw bentonite (b) under different pHs.
Figure 10. XRD pattern of 10 mM-Sr adsorbed on Taiwan Zhi-Shin bentonite (a) and raw bentonite (b) under different pHs.
Ijms 26 05298 g010
Figure 11. Hydrogen bonds and water “bridges” are formed between Sr2+ and the bentonite surface. (a) The initial configuration of Sr2+ within the Ca2+ bentonite layer; (b) a snapshot of the AIMD simulation showing the formation of hydrogen bonds between water on the hydrated surface and oxygen atoms on the bentonite surface; (c) snapshots of AIMD simulations, the formation of hydrogen bonds between hydrated ions and bentonite surfaces, and the formation of water “bridges” connecting Sr2+ and Ca2+ surfaces; (d) a snapshot of the AIMD simulation, the formation of a water “bridge” connecting Ca2+ to the bentonite surface.
Figure 11. Hydrogen bonds and water “bridges” are formed between Sr2+ and the bentonite surface. (a) The initial configuration of Sr2+ within the Ca2+ bentonite layer; (b) a snapshot of the AIMD simulation showing the formation of hydrogen bonds between water on the hydrated surface and oxygen atoms on the bentonite surface; (c) snapshots of AIMD simulations, the formation of hydrogen bonds between hydrated ions and bentonite surfaces, and the formation of water “bridges” connecting Sr2+ and Ca2+ surfaces; (d) a snapshot of the AIMD simulation, the formation of a water “bridge” connecting Ca2+ to the bentonite surface.
Ijms 26 05298 g011
Table 1. Estimated isothermal model parameters for Sr2+ absorption onto Taiwan Zhi-Shin bentonite.
Table 1. Estimated isothermal model parameters for Sr2+ absorption onto Taiwan Zhi-Shin bentonite.
AdsorbentsLangmuir ModelFreundlich Model
qe exp (mmol g−1)qmax (mmol g−1) R2KFnR2
Zhi-Shin bentonite0.280.410.49450.1121.1960.96
qe exp: actual adsorption value; qmax: theoretical maximum adsorption value; R2: correlation coefficient; KF and n: Freundlich constant.
Table 2. Thermodynamic parameters of Sr2+ uptake on Taiwan Zhi-Shin bentonite under different temperatures.
Table 2. Thermodynamic parameters of Sr2+ uptake on Taiwan Zhi-Shin bentonite under different temperatures.
CipHLnKd (L Kg−1)ΔG° (kJ mol−1)ΔH°
(kJ mol−1)
ΔS°
(KJ mol−1 K−1)
303318 K333 K303 K318 K333 K
0.1 mM-Sr92.430862.422542.11603−0.62492−6.1295−6.0099−8.67−0.01
10 mM-Sr90.787740.967980.9534−2.06−2.73−2.734.710.02
LnKd: logarithm of the retention factor; ΔG°: Gibbs free surface energy change; ΔH°: change in standard enthalpy; ΔS°: change in standard entropy.
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

Lin, Y.; Li, Y.; Yang, Y.; Chang, P.-H. Adsorption of Sr2+ from Synthetic Waste Effluents Using Taiwan Zhi-Shin Bentonite. Int. J. Mol. Sci. 2025, 26, 5298. https://doi.org/10.3390/ijms26115298

AMA Style

Lin Y, Li Y, Yang Y, Chang P-H. Adsorption of Sr2+ from Synthetic Waste Effluents Using Taiwan Zhi-Shin Bentonite. International Journal of Molecular Sciences. 2025; 26(11):5298. https://doi.org/10.3390/ijms26115298

Chicago/Turabian Style

Lin, Yihui, Yuhan Li, Yating Yang, and Po-Hsiang Chang. 2025. "Adsorption of Sr2+ from Synthetic Waste Effluents Using Taiwan Zhi-Shin Bentonite" International Journal of Molecular Sciences 26, no. 11: 5298. https://doi.org/10.3390/ijms26115298

APA Style

Lin, Y., Li, Y., Yang, Y., & Chang, P.-H. (2025). Adsorption of Sr2+ from Synthetic Waste Effluents Using Taiwan Zhi-Shin Bentonite. International Journal of Molecular Sciences, 26(11), 5298. https://doi.org/10.3390/ijms26115298

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