Next Article in Journal
Enhanced Optoelectronic Synaptic Performance in Sol–Gel Derived Al-Doped ZnO Thin Film Devices
Previous Article in Journal
Microstructure Engineered Nanoporous Copper for Enhanced Catalytic Degradation of Organic Pollutants in Wastewater
Previous Article in Special Issue
Sea Sand as a Silica Source to Hydrothermally Synthesize Analcime
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impact of Zinc(II) Chloride Contamination on Bentonites: Formation of Simonkolleite and Effects on Porosity and Chemical Composition

1
Faculty of Environmental Engineering, Geodesy and Renewable Energy, Kielce University of Technology, 25-314 Kielce, Poland
2
Faculty of Civil Engineering and Architecture, Kielce University of Technology, 25-314 Kielce, Poland
3
Faculty of Mining, Ecology, Process Control and Geotechnologies, Technical University of Kosice, Letna 9, 042 00 Kosice, Slovakia
*
Author to whom correspondence should be addressed.
Materials 2025, 18(13), 2933; https://doi.org/10.3390/ma18132933
Submission received: 9 May 2025 / Revised: 4 June 2025 / Accepted: 16 June 2025 / Published: 20 June 2025
(This article belongs to the Special Issue Application and Modification of Clay Minerals)

Abstract

This study examines the formation of the clay mineral simonkolleite (Skl) in bentonites contaminated with zinc(II) chloride (ZnCl2), a process that has been little documented in heterogeneous systems such as contaminated bentonites. We explain the contamination mechanisms and provide new insights into the mineralogical, structural, and physicochemical transformations occurring within these materials. The objective, explored for the first time, was to assess how the ZnCl2-induced mineral phase formation influences the properties of bentonites used as sealing materials, particularly regarding changes in specific surface area and porosity. Three bentonites were analyzed: Ca-bentonite from Texas (STx-1b), Na-bentonite from Wyoming (SWy-3), and Ca-bentonite from Jelsovy Potok, Slovakia (BSvk). Treatment with ZnCl2 solution led to ion exchange and the formation of up to ~30% simonkolleite, accompanied by a concurrent decrease in montmorillonite content by 9–30%. A suite of analytical techniques, including X-ray diffraction (XRD), scanning electron microscopy (SEM), thermogravimetric analysis (TGA), X-ray fluorescence (XRF), and energy-dispersive X-ray spectroscopy (EDS), was employed to characterize these transformations. The contamination mechanism of ZnCl2 involves an ion exchange of Zn2+ within the montmorillonite structure, the partial degradation of specific montmorillonite phases, and the formation of a secondary phase, simonkolleite. These transformations caused a ~50% decrease in specific surface area and porosity as measured by the Brunauer–Emmett–Teller (BET) nitrogen adsorption and Barrett–Joyner–Halenda (BJH) methods. The findings raise concerns regarding the long-term performance of bentonite-based barriers. Further research should evaluate hydraulic conductivity, mechanical strength, and the design of modified bentonite materials with improved resistance to Zn-induced alterations.

1. Introduction

Bentonites are clays that belong to the smectite group of minerals, with the 2:1 layered montmorillonite (Mnt) being the main component; the general formula for bentonites is (Na, Ca)0.4(Al4-xMgxSi8-xFexO20)(OH)4·nH2O. The structure of Mnt consists of two layers of silicon oxide (formed by tetrahedral units) with a layer of aluminum oxide (in an octahedral structure) in between. Substitution of atoms (x), such as aluminum (Al) with magnesium (Mg) or iron (Fe), creates a positive charge deficit, causing the Mnt layers to acquire a negative charge. Charge compensation mainly occurs through the adsorption of cations, such as sodium (Na) and calcium (Ca), onto the surface of the layers, which increases the size of the diffuse double layer (DDL) and enhances the mineral’s swelling capacity [1]. The unique sorptive properties and ability to exchange ions in bentonites make them valuable in various industrial fields, including medicine, construction, and materials technology, as well as having environmental applications [2,3,4,5]. One important use of Na-bentonites (>70% Mnt) is as a component of geosynthetic clay liners (GCLs), which are employed as sealing materials in landfills [6]. Egglofstein [7] observed that sodium ions in bentonite are replaced by Ca ions from cover soil through ion exchange over approximately 1–3 years. This results in a slight increase in the hydraulic conductivity of the bentonite barrier. Additionally, high-quality Na-bentonite, such as Wyoming bentonite (SWy-3), is widely used in the United States to produce barrier materials. However, it is not available on a large scale in other countries, where lower quality Ca-bentonites predominate [8]. These observations emphasize the importance of researching both Na- and Ca-bentonites used as GCLs. The widespread use of this material for sealing industrial waste landfills highlights the need to study the impact of potentially toxic metals (PTMs) and leachates with different chemical compositions on bentonite behavior [9,10].
Zinc(II) chloride (ZnCl2) is a common industrial waste generated as a by-product of textile waste recycling [11], as well as from galvanization processes and the dry battery industry [12,13]. As such, it can accumulate in landfills, where its presence poses a potential environmental risk. In industrial areas or closed mining sites, authors of different studies observed concentrations of Zn up to 5000 mg/kg [14], up to 7300 mg/kg [15], or up to 10,638 mg/kg [16] in various types of soils. Bentonites, due to their high specific surface area, exhibit increased reactivity, which can lead to the accumulation of significantly higher concentrations of PTMs, including Zn, particularly over extended exposure periods [8,13,17]. The mechanism of Zn adsorption includes both physical and chemical adsorption [18,19]. Physical adsorption involves the attraction of Zn ions to the surface of bentonite particles through van der Waals forces and ion exchange between Zn and metal cations in bentonite (mainly Na and Ca), which allows for their retention on the bentonite surface. Chemical adsorption, on the other hand, involves the formation of more stable Zn complexes, such as with oxygen groups (Si–O, Al–O) and hydroxyl groups (–OH). It is also possible for Zn complexes to form with Cl present in the environment, which has a limited tendency to bind with the solid phase of the soil [19]. Studies by Sun et al. [13] on bentonites contaminated with different concentrations of ZnCl2 (100–600 mg/L) showed that Zn adsorption is primarily the result of cation exchange regardless of the pH range (3–8). Chai et al. [20] observed that at pH ≥ 8, Zn adsorption is associated with the formation of metal complexes (such as Zn(OH)+, Zn(OH)3, etc.) and the precipitation of an insoluble zinc hydroxide (Zn(OH)2). Similarly, Kaya and Ören [21] observed that between pH 4 and 7, the basic mechanism is an ion exchange process. At higher pH values (i.e., pH 8), it may lead to the formation of zinc hydroxyl species.
Huang et al. [3] indicated that an important direction for future research is assessing the impact of ion substitution in the structure of clay minerals on their behavior. Cui and Chen [8] suggested that high concentrations of Zn when exposed to a reactive bentonite barrier may alter its original properties. Kaya and Ören [21] observed that the negative charge on the clay surface is partially neutralized by positively charged Zn ions, which reduces the repulsion between clay particles and promotes the formation of aggregates. The research by Chai et al. [20] confirmed that 90% of bentonite particles (d90) have a diameter smaller than 2.79 μm. However, following the absorption of Zn ions, this value increases to 5.38 μm. X-ray diffraction analysis (XRD) conducted after the adsorption of 600 mg/dm3 Zn showed an increase in the basal spacing from 1.4471 nm to 1.48661 nm. These findings suggest that Zn ions were adsorbed into the interlayer due to electronic attraction. The study conducted by Nartowska et al. [17] confirmed an increase in the particle size of bentonite and a decrease in the clay fraction (CLY) after saturation with ZnCl2. Depending on the type of bentonite, d10 (the particle diameter below which 10% of the material’s mass is found) increases from 0.96 to 2.83 μm; CLY decreases from 42.6% to 6.6% for SWy-3 bentonite (Na-bentonite from Wyoming, USA), and d10 rises from 1.39 to 1.91 μm; CLY decreases from 18.5% to 10.7% for STx-1b bentonite (Ca-bentonite from Texas, USA), and d10 increases from 1.33 to 2.18 μm; and CLY decreases from 19.8% to 8.9% for BSvk bentonite (Ca-bentonite from Jelsovy Potok, Slovakia). Although the studies presented focus on different types of bentonites and vary in the concentrations of Zn achieved through different saturation methods, the results consistently indicate a concerning increase in the size of clay particles due to the bentonite’s Zn ion saturation. This may increase porosity, reduce the effectiveness of the clay geotechnical barrier, and pose an environmental threat. Further analysis is required to assess how ZnCl2 contamination, a common industrial contaminant, affects porosity and what the causes of this observed variability are.
In recent years, there has been growing interest in Zn-enriched bentonites, which may exhibit additional, previously unknown, properties, particularly in forming new mineral phases, such as simonkolleite (Zn5(OH)8Cl2·H2O) [22]. This mineral belongs to the layered double hydroxide salts, also known as “anionic clays.” It exhibits hydration and adsorption properties similar to natural clays, with a significant interlayer spacing (0.79 nm) [23]. Certain information regarding the presence of simonkolleite (Skl) has been reported in electronics and steel materials. Typically, Skl formation has been documented as a corrosion product of galvanized steel bolts exposed to a marine atmosphere [24]. However, Skl may form in systems where bentonite, zinc, and chloride ions coexist. For instance, Hsiao et al. [25] reported the formation of Skl in the zinc anode of a battery (Na-Mnt with a grain size below 10 μm) following treatment with hydrochloric acid. The formation of this mineral phase was associated with specific conditions. Oueslati et al. [26], in studies on the treatment of Wyoming bentonite (SWy-2) with a 0.5 M ZnCl2 solution, observed that a Skl signal appeared in XRD diffractograms. The authors conducted analyses on a single Na-bentonite sample and assessed its structural and hydration properties. The influence of Skl on the porosity and chemical properties of the bentonite was not considered.
The objective of this paper, explored for the first time, was to assess how ZnCl2-induced mineral phase formation influences the properties of bentonites used as sealing materials, particularly regarding changes in specific surface area and porosity. Understanding these effects is crucial for the design of sealing materials. While many studies have focused on the influence of Zn, pH, and granulometric composition on the quality of bentonite, there remains a critical need to assess porosity, which has not yet been evaluated in the case of bentonites containing Skl. Additionally, there is insufficient knowledge regarding the potential formation of these new mineral phases without additional treatments, such as temperature changes, HCl addition, particle size changes, or contact with steel corrosion products. This highlights the need for further research to document this mineral phase and assess its potential impact on bentonite quality, environmental safety, and the effectiveness of sealing materials. The novelty of this study lies in its unique focus on the in situ formation of Skl in ZnCl2-contaminated bentonites, without the use of additional treatments. This is the first study to explore the direct impact of this mineral phase on the structural and chemical properties of bentonites and its implications for geotechnical applications.
This study is based on the following research hypotheses:
  • Contamination of Na- and Ca-bentonites with ZnCl2 may lead to the formation of a new mineral phase—simonkolleite (Skl);
  • Saturation of bentonites with ZnCl2 leads to changes in their chemical composition, which may reduce their sorption capacity and the soil specific surface area (SSA);
  • The presence of Skl affects the porosity of the bentonites. This potentially alters their effectiveness in geotechnical barriers and poses a potential environmental threat.

2. Materials and Methods

2.1. Chemical Modification

The research material consisted of source clays obtained from the Clay Mineral Society (Purdue University, West Lafayette, Indiana), USA (Ca-bentonite from Texas (STx-1b) and Na-bentonite from Wyoming (SWy-3), as well as Ca-bentonite from Jelsovy Potok, Slovakia (BSvk)). All clays had particle sizes below 63 μm and contained at least 70% Mnt [17].
Fifty-gram samples were collected from each type of clay and subjected to an ion exchange process at room temperature (≈22 °C). Ten liters of a 1 mol L¹ solution of ZnCl2 (Chempur, Poland, CAS No. 7646-85-7) (pH = 4.51 ± 0.01) was added to each sample, thoroughly mixed, and left to sit for 48 h. After this period, the clear liquid was decanted, leaving a sediment layer approximately 2 cm thick. Water was then carefully siphoned off from above the sediment, ensuring the sediment remained undisturbed. This procedure was repeated three times. The average pH value of the leachate from three decantations was 4.90 ± 0.07 for calcium bentonites and 5.05 ± 0.06 for sodium bentonite.
To eliminate chloride ions (Cl) from the soil, the remaining material was placed in permeable membranes and stored in 20 L containers filled with circulating distilled water, allowing for Cl removal via osmosis. Cl levels were monitored using silver nitrate (AgNO3) until no Cl was detected. The water in the containers was changed every 2–3 days. The entire ion exchange process took approximately 35–40 days. Once the water no longer contained Cl ions, the soil pastes were transferred to glass cylinders and air-dried.
Additionally, the Cl concentration was analyzed in water extracts using the Mohr method. This method involves titration with silver nitrate (AgNO3) in the presence of chromate as an indicator, as described in PN-ISO 9297:1994 [27]. Before analysis, soil samples (2 g each) were shaken for 24 h on a rotary shaker in 20 mL of distilled water (conductivity 0.06 μS·cm−1). After shaking, the samples were filtered through Whatman filter paper, grade 1.6 μm, 50 mm diameter circle (Maidstone, Kent, UK), until a clear solution was obtained.

2.2. Mineralogical Analysis

The mineral composition of the tested clays was determined using X-ray diffraction (XRD) in the Bragg–Brentano geometry. X-ray diffraction patterns were recorded with a Bruker D8 Advance X-ray diffractometer (Bruker, Berlin, Germany) equipped with a Johansson-type monochromator (Bruker, Berlin, Germany) to generate CuKα1 radiation (λ = 1.5406 Å) and a position-sensitive LynxEye detector (Bruker, Berlin, Germany). Measurements were carried out in the 2θ range from 4.51° to 70° with 0.02° steps. The applied voltage was 3.540 kV, and the current was 530 mA. The identification of mineral phases in the clays was based on data from the PDF-4+ database of the International Center for Diffraction Data (ICDD, Newtown Square, PA, USA). The X-ray diffractograms of the clays were subjected to semi-quantitative mineralogical analysis using Diffrac Eva V6 software (Bruker, Berlin, Germany). Mineral content was calculated based on the intensity of their characteristic peaks in the diffractograms.

2.3. Thermal Analysis

Analysis of bentonites before and after ZnCl2 treatment was performed using a Thermal Gravimetric Analyzer SDT Q600 V20.9 (TA Instruments, New Castle, PA, USA). Before the measurements, calibrations were conducted to ensure the accuracy and reproducibility of the results: (1) DTA baseline correction, performed by the manufacturer’s protocol to minimize signal drift and ensure thermal stability across the temperature range (50–1200 °C) at a heating rate of 20 °C/min; (2) TGA weight calibration using certified standard weights; (3) temperature calibration using zinc as a reference material (melting point: 419.5 °C); and (4) cell constant adjustment to ensure optimal thermal response and signal consistency.
Samples of approximately 15 mg were placed in alumina pans. The reference sample was Calcium Oxalate Monohydrate (No. 900905.901, TA Instruments). The specimens were heated at a constant rate of 20 °C/min from 20 °C to 800 °C in a nitrogen atmosphere with a flow rate of 100 mL/min. The accuracy of the sample mass measurements was 0.1 μg, and the ∆T sensitivity (DTA) was 0.001 °C. Each bentonite sample was tested three times.

2.4. Scanning Electron Microscopy

The structural and chemical analysis was performed using scanning electron microscopy (SEM) with a field emission microscope JSM-7100F (JEOL, Tokyo, Japan) operating at a voltage of 15 kV. The surfaces of air-dried samples were coated with a thin layer of carbon to prevent charging using a JEOL JEC-530 sputter coater (JEOL, Tokyo, Japan). Images were analyzed at a magnification of ×5000. Such magnification allowed clear visualization of the morphology characteristic of simonkolleite, enabling its distinction from other phases present in the sample. To complement the analysis, SEM images were also acquired at magnifications of ×10,000 and ×30,000, allowing for a detailed assessment of the microcrystal morphology using the approach presented by Sun et al. [28].

2.5. Chemical Analyses of Bentonites

The chemical composition was analyzed using energy-dispersive X-ray spectroscopy (EDS) and X-ray fluorescence (XRF). inductively coupled plasma optical emission spectroscopy (ICP-OES) was also employed to support the interpretation of the results; the methodology and detailed procedure are comprehensively described in the authors’ previous publication [17]. The EDS method provides compositional information at the microscopic level, covering selected micro-areas of the sample. XRF was used to determine the elemental composition from the surface of the solid material. ICP-OES, in turn, enabled precise quantification of trace elements in the digested samples, thus complementing the data obtained by EDS and XRF.
In the case of EDS, the analysis was performed using an X-Max system (Oxford Instruments, Oxford, UK). Each element in the sample emits characteristic X-ray spectral lines when excited, with their energy distribution and intensity reflecting the type and concentration of the respective elements [29]. For the bentonite sample, a magnification of ×2000 was applied, corresponding to a scale of 50 µm visible in the SEM image. Under these conditions, the entire analyzed area, approximately 150 µm × 150 µm, was examined, enabling simultaneous assessment of the microstructure and the average chemical composition of several bentonite grains. An accelerating voltage of 20 kV was used, providing sufficient energy to excite characteristic X-ray emission from most elements typically found in bentonite (e.g., Si, Al, Mg, Fe, Ca, Na), thus allowing for a complete and representative analysis. This approach is particularly important for characterizing mineralogically heterogeneous materials.
The chemical composition of the clays was also analyzed using X-ray fluorescence (XRF) with a SPECTRO iQ II (Ametek, Meerbusch, Germany) equipped with a silicon drift detector (SDD) offering a resolution of 145 eV at 10,000 pulses. A Bragg crystal and Highly Ordered Pyrolytic Graphite (HOPG) target were used to polarize the primary beam. The measurements were conducted for 300 s under a helium atmosphere at voltages of 25 kV and 50 kV and currents of 0.5 mA and 1.0 mA, respectively. The analysis followed the standardized fundamental parameter method for powdered samples. The sample mass used for the analysis was approximately 2 g.

2.6. Granulometric and Plastic Properties

The physicochemical properties of clays are largely determined by the interactions between the solid phase (clay particles) and the liquid phase (water and dissolved substances) [17]. Their intensity depends on granulometric composition and plasticity parameters.
The granulometric composition was determined using a HELOS/BF SUCELL laser diffractometer from Sympatec GmbH (Clausthal-Zellerfeld, Germany), equipped with a “wet” test accessory. Clay samples were prepared as pastes 48 h before analysis to full saturation. For testing, 3 g of clay paste was mixed with 50 mL of distilled water, and a small sample was taken to ensure that the particle concentration in the test chamber did not exceed 25%. Reference measurements were performed before each new analysis. The effective diameter d10 was derived from the granulometric curves, indicating that 10% of the mass consisted of grains equal to or smaller than this diameter within the soil sample. This parameter is used to assess the filtration properties of soil, and smaller values generally indicate lower permeability.
The plasticity characteristics were measured using standard methods: Casagrande’s cup device for determining the liquid limit (LL) and the test for the rolling plastic limit (PL) according to standard no. EN ISO/TS 17892-12 [30].

2.7. Specific Surface Area and Porosity

The specific surface area of the clays (SSA, [m2/g]) studied was determined using the following:
  • The water vapor sorption test (WST) according to Stępkowska [31] (SSAWST) (1),
S S A W S T = 6 · ( w 50 · 5.85 )
where w50 is the sorption moisture at a vapor pressure p p 0 = 0.5 in a desiccator over saturated magnesium nitrate (V) solution determined by drying at 220 °C (sorption takes 10 days).
Additionally, the sorptive moisture at p p 0 = 0.95 was determined in a desiccator over a 10% sulfuric acid (VI) solution (with sorption taking 14 days). This moisture can be identified as strongly bound water.
  • The Brunauer–Emmett–Teller (BET) method (SSABET) (2)
S S A B E T = a m · ω · N A
where am is the amount of adsorbed gas (N2), ω is the surface occupied by a single adsorbate molecule in a monomolecular layer, also known as the “molecular cross-sectional area”, and NA is Avogadro’s number.
Additionally, the Barrett–Joyner–Halenda (BJH) model was used to determine the surface area of mesopores (SABJH) using NOVA Touch Software firmware v1.09, S/N 1050020598 (Anton Paar GmbH, Graz, Austria). This model assumes that pores are filled with gas according to the principles of capillary condensation, which occurs as pressure increases. Based on the Kelvin equation, which relates vapor pressure to the curvature radius of the liquid meniscus, the pore size can be determined (3).
ln p p 0 = 2 γ V m r R T
where γ is the surface tension of the liquid nitrogen, V m is the molar volume, R is the gas constant, T is the temperature, and p p 0 is the relative pressure.
The thickness of the adsorbate layer on the pore walls is then added to the capillary radius, assuming cylindrical pores. The analysis is performed using the nitrogen desorption isotherm, enabling the determination of the pore size distribution and pore volume. The mesoporous surface area is calculated as the sum of the surface areas of pores of specific sizes derived from this distribution.
The BET method was used to calculate the external specific surface area of the bentonites (SSABET), which corresponds to the upper boundary of micropores (>0.7 nm) and includes mesopores (2–50 nm), where inter-packet spaces are not accessible to water. The WST test, on the other hand, provides more comprehensive information, accounting for both external and internal surfaces. The BJH method was applied to determine the surface area of mesopores (SABJH). This parameter is important because it reflects the portion of the material’s surface potentially involved in ion exchange, adsorption, or catalytic processes, features not fully captured by BET when microporosity dominates. The accurate determination of SABJH helps distinguish surface area contributions from different pore size distributions, complementing the BET and WST results. The relevance of combined surface analyses (SABJH and SSABET) was demonstrated in the studies by Asabina et al. [32] and Borinelli et al. [33].
SSABET, SABJH, and the pore space distribution of soil samples were investigated using low-temperature nitrogen (77 K; p/p0 = 0.047–0.95) adsorption–desorption, performed with a NOVA Touch 2LX (Anton Paar, Graz, Austria) instrument. Before measurements, samples weighing approximately 0.35 g were heated to 105.0 °C at a rate of 10.0 °C/min and then held at this temperature for 300 min for degassing.

2.8. Data Quality Control

Bentonites had grain sizes of d < 63 μm, ensuring high homogeneity and uniformity. Additionally, this study was conducted on two source clays, STx-1b and SWy-3, obtained from the Clay Mineral Society collection, guaranteeing their well-verified quality, stability, and representativeness as well-characterized reference materials. High-purity reagents (analytical or chemically pure, Chempur, Piekary Śląskie, Poland)), free from contaminants that could affect analytical results, were used throughout the analyses. Materials with known composition and structural properties provided a solid reference for monitoring structural transformations in bentonites contaminated with ZnCl2 (via XRD and TG analyses). Thermogravimetric analysis (TG) results were compared with X-ray diffraction (XRD) data, allowing a comprehensive interpretation of mineralogical changes and confirming the phases present in the studied samples. Chemical analyses were performed using ICP-OES and XRF techniques. For ICP-OES, a multi-element calibration standard (Instrument Calibration Standard 2, No. N9301721, PerkinElmer, Waltham, MA, USA) was employed, and calibration curves were evaluated using linear regression at a significance level of p = 0.01; Grubbs’s test was applied to exclude outliers. Three replicate measurements were performed for each bentonite specimen, and the calculated standard deviations (SD) were used as practical estimates of measurement uncertainty. Similarly, XRF measurements were conducted in triplicate. All resulting standard deviations are reported in Section 3.2. In SEM-EDS analyses, ZAF correction and normalization were applied to compensate for matrix effects and reduce systematic errors. Elemental composition was determined over the largest possible representative area of the sample, allowing averaging of results and minimizing the influence of local heterogeneities. For specific surface area measurements by the BET method, uncertainty was indirectly estimated based on the quality of the BET model fit by analyzing parameters such as slope, intercept, and the BET constant (C). A high correlation coefficient (r = 0.99) indicates excellent agreement of the model with the data, confirming the reliability and accuracy of the obtained surface area values.

3. Results and Discussion

3.1. Mineralogical and Microstructural Analysis

3.1.1. XRD

Figure 1 shows the XRD diffractograms for bentonites before and after contamination with ZnCl2. XRD-identified mineral phases in bentonites, along with corresponding PDF references, are listed in Table A1. The montmorillonite content (Mnt) in uncontaminated SWy-3 and STx-1b is nearly consistent with the values reported by Chipera and Bish [34] (SWy-2—75% Mnt; STx-1—67% Mnt) and by Castellini et al. [35] (STx-1b—73% Mnt).
In bentonites contaminated with ZnCl2, a new mineral phase, Skl, forms and coexists in an amount of approximately 30% with the main mineral, Mnt (Figure 1a–c). Simonkolleite occurring in calcium (Ca) bentonites (PDF #01-077-2311) and sodium (Na) bentonites (PDF #00-007-0155) exhibits minor differences in crystallographic parameters, which may influence the physical properties of these bentonites. Considering the general formula of Mnt, the probable mechanism for the formation of Skl from Mnt can be predicted:
  • Saturating montmorillonite (pH 7.92 ± 0.12) with a ZnCl2 solution (pH = 4.51 ± 0.01) causes only a slight increase in the suspension’s pH to 4.98 ± 0.11. Under these mildly acidic conditions, intensive ion exchange and partial dissolution of montmorillonite components occur. Chemical reactions, such as the formation of Zn(OH)2 according to Equation (4), take place primarily at the montmorillonite phase boundary or within its structure, where the local pH is higher.
Z n C l 2 + 2 O H Z n ( O H ) 2 + 2 C l
2.
The formation of Skl is most likely caused by Zn ions, which partially replace Al, Mg, Fe, and Si ions (as indicated by the XRF data described in the next section) in the Mnt structure, as well as SiO2, as shown by the XRD data (Figure 1). Only a portion of Mnt and SiO2 contributes to the formation of Skl, so the general reaction equation may proceed according to scheme (5).
Z n ( O H ) 2 + 2 C l + M n t * + S i O 2 S k l * * + t h e   r e m a i n i n g   M n t + t h e   r e m a i n i n g   S i O 2
Notes: The general formula for Mnt montmorillonite is * (Na, Ca)0.4(Al4−xMgxSi8−xFexO20)(OH)4·nH2O; the general formula for Skl simonkolleite is ** (Zn5(OH)8Cl2·H2O).
A decrease in silica content is also observed, likely associated with its dissolution. Bagheri et al. [36] showed that at alkaline pH values, calcium increases the rate of silica dissolution, while aluminum slows it down. Since the greatest decrease in silica content was observed in the Texas bentonite (Figure 1b), it can be assumed that one possible reason is the presence of the highest concentration of calcium ions and the lowest concentration of aluminum ions in this sample (Section 3.2) [17].
To evaluate the behavior of the Mnt phases as a result of their saturation with ZnCl2, Table A1 was analyzed, which summarizes their occurrence.
As shown in Figure 1a, the Mnt 2 montmorillonite phase (NaMgAlSiO2(OH)H2O, PDF no. 00-002-0014) disappears from the XRD pattern in the BSvk sample after the treatment with ZnCl2. This phenomenon is likely due to the destabilization of the Mnt 2 structure caused by the exchange of Na and Mg cations for Zn. The Zn ion has a higher charge and a different ionic radius compared to Na and Mg—its ionic radius (~0.74 Å) is slightly larger than that of Mg (~0.72 Å) but significantly smaller than that of Na+ (~1.02 Å). This results in stronger electrostatic interactions between the mineral layers and introduces structural strain, disrupting the regular arrangement of the montmorillonite layers. Consequently, the interlayer spacing contracts, and the crystalline order is lost, leading to the disappearance of characteristic peaks in the XRD pattern. On the other hand, the Mnt 1 phase with the composition (Na, Ca)0.3(Al, Mg)2Si4O10(OH)2·xH2O (PDF no. 00-003-0015) shows greater stability. The presence of Ca (~1.00 Å) and the mixed cation composition provide increased structural resistance to cation exchange and disturbances. Ca, with its larger ionic radius and charge, stabilizes the layer structure, maintaining the regular arrangement of layers despite the presence of Zn. As a result, this phase retains its crystallinity and its peaks remain visible in the XRD analysis.
Selected forms of montmorillonite—Mnt 3: Na-Al-Si-O-OH-H2O (PDF no. 00-003-0019); Mnt 4: 15 Å Ca0.2(Al, Mg)2Si4O10(OH)2·4H2O (PDF no. 00-013-0135); and Mnt 5: CaMg2AlSi4(OH)2·H2O (PDF no. 00-002-0239)—likely undergo crystal structure breakdown or transform into an amorphous state as a result of the reaction of STx-1b with ZnCl2 (Figure 1b). This process is driven by intense ion exchange, where Zn displaces Na, Ca, and Mg from the interlayer spaces. Removing interlayer cations without effective compensation of the negative charge in the tetrahedral sheets leads to the weakening of interlayer bonds and the gradual disintegration of the crystal structure. These changes result in the loss of layered ordering, observed in XRD analysis as a decrease in intensity or complete disappearance of diffraction peaks characteristic of Mnt 3, 4, and 5 (Figure 1b). In contrast, the Mnt 1 phase ((Na, Ca)0.3(Al, Mg)2Si4O10(OH)2·xH2O, PDF no. 00-003-0015) showed no significant changes in its crystal structure after treatment with ZnCl2—only a reduction in peak intensity was noted, which may indicate a decrease in crystalline material quantity or partial dehydration due to Zn incorporation. This suggests that secondary mineralization of simonkolleite (Zn5(OH)8Cl2·H2O) occurs independently of this particular montmorillonite form and does not significantly affect its structural stability.
Based on Figure 1c, in the SWy-3, the Mnt 6—22 Å montmorillonite phase (Na0.3(Al, Mg)2Si4O10(OH)2·H2O, PDF no. 00-029-1499) undergoes structural modifications as a result of intensive ion exchange triggered by the introduction of ZnCl2 into the bentonite matrix. This process is accompanied by a decrease in pH (from 8.05 to 6.84) and an increase in salinity, which further facilitates the Mnt phase transformation. As a result, a transition occurs to a new, structurally equivalent Mnt 8—14 Å montmorillonite phase (Na0.3(Al, Mg)2Si4O10(OH)2·xH2O, PDF no. 00-013-0259)—characterized by reduced interlayer water content and altered physicochemical properties. Additionally, other montmorillonite-related phases identified in the SWy-3, Mnt 7—(Al(OH)2)0.33Al2(Si3.₆₇O10)·8H2O (PDF no. 00-011-0303)—and Mnt 3—Na-Al-Si-O-OH-H2O (PDF no. 00-003-0019)—are likely transformed or degraded into a deionized or structurally simplified aluminosilicate phase such as Mnt 9—AlSi2O₆(OH)2 (PDF no. 00-002-0037)—due to intensive ion exchange and partial dehydration. The release of OH ions and structural degradation products from the aluminosilicate framework promotes the crystallization of Skl (Zn5(OH)8Cl2·H2O) as a secondary mineral phase, likely forming as microscale aggregates within the pores and interlayer spaces of montmorillonite. These structural transformations are also reflected in changes in the interlayer spacing (d001), as summarized below.
The highest interlayer distances of Mnt (d001) before saturation with ZnCl2 were 14.87, 14.88, and 11.44 Å, while after saturation, they measured 14.29, 14.17, and 13.21 Å for the bentonites BSvk, STx-1b, and SWy-3, respectively [17]. After saturation, a slight decrease in interlayer spacing was observed in Ca-Mnt, which may be due to the smaller ionic radius of Zn compared to Ca. Voora et al. [37] showed that the larger the interlayer cation radius (Li, Na, K, Mg, Ca), the greater the interlayer distance in Mnt. However, our study did not confirm this relationship for Na-Mnt, where, despite the smaller ionic radius of Zn (0.74 Å) compared to Na (0.95 Å), an increase in interlayer spacing was observed. This may be due to microstructural changes and improved hydration of Zn ions in Na-bentonite after saturation with ZnCl2, leading to an increase in the effective ionic radius. Similarly, Oueslati et al. [26] observed larger interlayer distances in ZnCl2 contaminated SWy-2 bentonite with rises in relative humidity (15–75%). The results may also suggest that in SWy-3, Zn ions mainly adsorb in the interlayer spaces of Mnt, most likely due to the easier penetration of Zn ions, which have a smaller ionic radius than Na, potentially leading to an increase in interlayer distance. The main mechanisms responsible for this process may include electronic attraction and interlayer cation exchange [20]. In contrast, in STx-1b and BSvk, Zn primarily adsorbs on the surface of Mnt particles because in Ca-bentonites, the larger radius of Ca (compared to Na) makes it more difficult for Zn ions to access the interlayer spaces. However, this does not exclude the involvement of Zn in cation exchange in these Mnt types, as indicated by the observed, albeit slight, decrease in interlayer spacing, the high Zn content in the saturated samples, and its presence across the entire surface shown in the EDX images, which will be presented in the following section of this work (Section 3.2).

3.1.2. TGA

The thermogravimetry (TG) and derivative thermogravimetry (DTG) curves of bentonites, both before and after treatment with ZnCl2, are shown in Figure 2. For Mnt, three thermal decomposition steps were recorded in bentonites (Figure 2):
  • Desorption (at ~75 °C): This stage corresponds to the release of physically adsorbed water (molecular water on or within clay mineral particles) due to weak interactions, such as van der Waals forces or hydrogen bonds.
  • Dehydration (at ~150 °C): In this step, water that was desorbed from the mineral surface, potentially weakly bound by electrostatic forces or other physical interactions, is further released. This water may include interlayer water between the mineral layers, although it is not chemically bonded to the Mnt. In Zn-bentonites, the dehydration peak occurs at around 125 °C, which may indicate structural changes in the Mnt layers due to the interaction with Zn ions.
  • Thermal decomposition between 550 °C and 800 °C can be attributed to dehydroxylation which refers to the release of chemically bound water associated with the mineral’s structure. This water is part of the hydroxyl groups (-OH) in the Mnt layers, forming part of the interlayer cations and the mineral structure.
The maximum peak temperatures for each step of the thermal decomposition of Mnt depend on the type of bentonite and are as follows for BSvk, STx-1b, and SWy-3: (desorption) T1 = 87.4 °C, 79.5 °C, 70.5 °C; (dehydration) T2 = 154.2 °C, 146 °C, 145.5 °C; (dehydroxylation) T3 = 676.2 °C, 677 °C, and 689.9 °C. Contamination of the bentonites with ZnCl2 lowered the dehydration temperature of Mnt (T2 = 124 °C, 129.5 °C, 126.2 °C) (Figure 2, Figure A1). This phenomenon can be explained by the osmotic forces created after the contamination was introduced into the bentonite and by structural changes. Higher ionic strength weakens the electrostatic repulsion between particles, allowing for easier migration of molecules and ions through the material, which may contribute to the lowering of the dehydration temperature. A similar phenomenon was observed by Zhu et al. [38], who studied bentonite mixtures contaminated with Cr(VI) and found that the compression of the diffusion double layer (DDL) under high ionic strength conditions enlarges the diffusion path, enhancing the migration of Cr(VI) through the pore flow paths.
As shown in Figure 2a–c, ZnCl2 contamination alters the shape of the dehydroxylation peak. In the bentonites BSvk, STx-1b, and SWy-3, dehydroxylation occurred at temperatures above 550 °C prior to contamination. After treatment, the SWy-3 sample exhibits a distinct peak in the range of 400–500 °C, which can be attributed to the thermal instability of the secondary phase—simonkolleite (T > 400 °C [23])—along with a broader montmorillonite dehydroxylation peak beginning at around 480 °C. In contrast, in the BSvk and STx-1b samples, a broad dehydroxylation peak is observed across a wide temperature range above 400 °C. This is most likely the result of the overlap between the simonkolleite decomposition peak and the montmorillonite dehydroxylation peak, the precise onset of which is difficult to determine. However, the shape of the peaks indicates a lowering of the montmorillonite dehydroxylation temperature in all the studied bentonites. This phenomenon can be attributed to two overlapping effects: (1) direct interactions of Zn2+ ions with the aluminosilicate layers, which disrupt interlayer bonding and weaken the mineral’s structural framework, an effect enhanced by the strong hydration shell of Zn [39], and (2) microstructural changes caused by the decomposition of simonkolleite, leading to the formation of defects and local distortions in the vicinity of montmorillonite. Both factors reduce the energy required to initiate dehydroxylation, resulting in its occurrence at lower temperatures.
Thermogravimetric measurements confirm the formation of Skl in ZnCl2-treated bentonites (Figure 2). At temperatures ranging from 105 to 161 °C, Skl loses part of its water (H2O) and transforms into the form Zn5(OH)8Cl2 (6).
Simonkolleite (Skl)
Z n 5 ( O H ) 8 C l 2 · H 2 O 105 161 ° C Z n 5 ( O H ) 8 C l 2 + H 2 O
A peak in the temperature range from 161 to 200 °C (with a maximum of 181 °C) corresponds to the first decomposition of Skl (7).
Z n 5 ( O H ) 8 C l 2 161 200 ° C 3 Z n O + Z n O · Z n C l 2 · 2 H 2 O + 2 H 2 O
Analogous to the mechanism proposed by Moezzi et al. [23] for Skl, it is suggested that for bentonite containing Skl (Zn SWy-3—Na-bentonite treated with ZnCl2), in the temperature range of 205–230 °C, part of the water is released from the mineral ZnO·ZnCl2·2H2O. As a result, the mineral structure transitions to the form Zn(OH)2·ZnCl2·2H2O, which is observed as a distinct peak on the graph (e.g., in Figure 2c) (8).
Z n O · Z n C l 2 · 2 H 2 O 205 230 ° C Z n ( O H ) 2 · Z n C l 2 · 2 H 2 O + H 2 O
The subsequent gradual mass loss observed in the treated SWy-3 above 400 °C can be attributed to the decomposition of a new amorphous phase (9).
Z n ( O H ) 2 · Z n C l 2 a m o r p h o u s T > 400 ° C 2 Z n O + 2 H C l
In Zn STx-1b and Zn BSvk (Ca-bentonites treated with ZnCl2), the peak in the 205–230 °C temperature range is not as distinct as in Zn SWy-3, which may suggest that in these clays, water removal occurs gradually. The larger size of the Ca ion (1.12 Å) compared to Na (0.95 Å) leads to stronger bonds with the bentonite layers, making the exchange of Zn ions more difficult. Zn ions (0.74 Å) that have been exchanged in Ca-containing bentonite become more stable as a result, and the bentonite structure is more difficult to change. This may also contribute to the difficulty in releasing water and Zn ions [17].
Since in ZnCl2-treated bentonites dehydration occurs at temperatures above 105 °C and dehydroxylation occurs above 400 °C, mass loss in these temperature ranges was analyzed to assess differences between untreated and ZnCl2-treated bentonites and to demonstrate the presence of Skl (Figure A1). In the 105–400 °C range, mass losses were observed for BSvk, STx-1b, and SWy-3 before (6.0%, 3.8%, 1.3%) and after (8.9%, 7.9%, 8.9%) ZnCl2 contamination. In the 400–800 °C range, similar increases were observed: before (4.4%, 3.5%, 4.8%) and after (7.1%, 6.7%, 6.6%) treatment. These higher mass losses indicate the formation of simonkolleite. The newly formed mineral phase exhibits thermal properties distinct from montmorillonite, as evidenced by differences in thermal decomposition. Montmorillonite typically loses thermal stability at temperatures above 550 °C. In contrast, secondary phases such as simonkolleite decompose at significantly lower temperatures, starting at 161 °C, with complete decomposition occurring above 400 °C. This process releases Zn, which may affect the ability of ZnCl2-treated bentonite to retain zinc under these conditions. We cannot exclude the fact that these thermal processes also induce changes in the microstructure, chemical composition, and physical properties of bentonite, further impacting the stability of the matrix.
The total mass loss after the decomposition of Skl in the studied bentonites ranges from 21.7% to 23.3%. Previous studies report values between 25.5% and 37.5%, but these refer only to the Skl phase. The authors explain that the mass loss depends on how chlorine behaves during the decomposition process. If chlorine is released as HCl, the mass loss is smaller, whereas if chlorine transforms into ZnCl2 and then evaporates, the mass loss is greater [23].

3.1.3. SEM

In the SEM images (Figure 3b,d,f–h) of bentonites treated with ZnCl2, thin-layered plates are visible that correspond to the needle-like structure of Skl [40]. Aggregates with a noticeable degree of ordering can be observed within the structure. The micrographs of Zn-saturated bentonites exhibit better electron-reflecting ability compared to the unsaturated bentonites. In the latter (Figure 3a,c,e), schistose structures, chaotically arranged aggregates, and low contrast between the layers are observed. The formation of aggregates in Zn-contaminated bentonites was also documented by Chai et al. [20]. The process of clay particle aggregation can be explained by the adsorption of Zn ions onto the clay surface, which neutralizes part of its negative charge, reducing repulsion between particles.
Simonkolleite is a mineral that naturally occurs in thin flakes, scales, or layers, forming crystalline aggregates ranging in size from several tens of nanometers to a few micrometers. These aggregates most commonly adopt a regular hexagonal platelet morphology; however, in practice, more complex structures may also arise due to the crystal growth process. Figure 4 illustrates the morphology of the obtained aggregates at higher magnifications (scale bars: 5 µm and 2 µm), allowing for a detailed analysis of their structure and uniformity. It should be noted, however, that most current synthesis methods for simonkolleite require precise pH adjustment of zinc salt solutions in an alkaline environment. Despite these efforts, such methods often result in not only the desired hexagonal morphologies but also undesired forms with irregular shapes, making it difficult to control the final structure of the material [41].

3.2. Chemical Analysis

Energy-dispersive X-ray spectroscopy (EDS) provides localized, high-resolution elemental data from specific micro-areas (in our study, the analyzed surface area was approximately 150 μm × 150 μm), enabling spatial mapping of metal distribution and the identification of microscale heterogeneities in the elemental composition of materials [28,42]. EDS results for unmodified bentonites and for bentonites treated with ZnCl2 are presented in Figure A2, Figure A3 and Figure A4 and Figure 5, Figure 6 and Figure 7, respectively.
Surface analysis revealed that zinc and chlorine—the main components of simonkolleite—are distributed uniformly across the entire surface of the montmorillonite, including within the spaces between its aggregates (Figure 5, Figure 6 and Figure 7). This distribution, together with the XRD and SEM results, confirms the formation of the simonkolleite phase in this region.
Table 1 shows the chemical composition of the bentonites before and after contamination with ZnCl2, determined by two independent methods: inductively coupled plasma optical emission spectrometry (ICP-OES) and X-ray fluorescence (XRF). ICP-OES measures the concentration of ions in extracts from bentonite solutions, while XRF analyzes the composition of the solid sample in the surface region. This enables a complementary assessment of elemental composition changes in the material and the effectiveness of the ion exchange process.
The results of ICP-OES and XRF analyses, combined with XRD data and interlayer spacing measurements, indicate that ion exchange occurred in sodium and calcium bentonites. Quantitative ICP-OES analysis revealed that the calcium content in Ca-bentonites decreased nearly fourfold (BSvk: from 11,945 ± 140 to 2778 ± 31 mg/kg dm; STx-1b: from 11,802 ± 101 to 2985 ± 23 mg/kg dm), while the sodium content in Na-bentonite SWy-3 decreased approximately tenfold (from 10,086 ± 81 to 995 ± 9 mg/kg dm) after saturation with ZnCl2. In contrast, the zinc content significantly increased in all Zn-saturated bentonites (BSvk: from 64.54 ± 0.69 to 17,857 ± 89 mg/kg dm; STx-1b: from 73.68 ± 0.27 to 16,153 ± 75 mg/kg dm; SWy-3: from 163.66 ± 1.50 to 44,463 ± 124 mg/kg dm). These substantial increases in Zn concentration—considering that simonkolleite constitutes only about 30% of the sample volume (XRD, Figure 1)—strongly suggest that a significant portion of the metal is most likely located within the interlayer spaces of the montmorillonite. This hypothesis is further supported by the XRF results (Table 1), as well as by the interlayer spacing data discussed in Section 3.1.1.
The XRF suggests that the contamination of bentonites with ZnCl2 leads to structural changes in Mnt. As a result of ion exchange with Zn, the concentrations of Si, Al, Mg, and Fe decreased in each of the studied bentonites, which may indicate changes in the tetrahedral and octahedral layers of Mnt. Furthermore, the reduced concentrations of Ca, Mg, Na, and K, along with the simultaneous increase in Zn ion concentration, suggest competition between these ions and their replacement by Zn. The presence of Zn and Cl indicates a new mineral phase, identified by XRD as Skl. The concentrations of heavy metals are low and do not show significant changes after contamination of the bentonites with ZnCl2. A slight decrease in Mn, Ni, and Cu may indicate replacement by Zn.
A 1 molar (1 M) concentration of ZnCl2 allows for an effective ion exchange process, replacing Ca and Na ions with Zn in Na- and Ca-bentonites. Previous studies have shown that concentrations of approximately 0.004 M ZnCl2 did not allow for full ion exchange in bentonites, and the formation of Skl was not observed [13]. In contrast, Oueslati et al. [26] when saturating bentonites with 0.5 M ZnCl2, noticed the appearance of Skl. They did not evaluate the completeness of the ion exchange or the quantity of the newly formed mineral phase. Previous studies also indicate that the effectiveness of ion exchange in bentonites depends not only on the molar concentration but also on the type of exchangeable cation. In the case of CuCl2-contaminated bentonites, the exchange of Ca ions for Cu was fully effective at molar concentrations of 0.10–0.25–0.50 M, while at a concentration of 1 M, not all Ca ions were replaced by Cu [43].

3.3. Physical Analysis

Table 2 shows how the granulometric, sorption, and plastic properties of bentonites change after saturation with ZnCl2.

3.3.1. Granulometric Composition, Plastic Parameters, and Soil Type

In ZnCl2-treated bentonites, a decrease in the clay fraction content (CLY), an increase in silt (SIL) and sand fraction content (SA), and an increase in the effective particle size (d10) (Table 2) were observed. The results may suggest the formation of aggregates. Tang et al. [47] discussed how aggregates may form and break down as a result of complex processes, including clay flocculation, swelling, dispersion, and salt crystallization in the soil. Angon et al. [48] confirm that Zn influences soil texture by altering the proportions of SA, SIL, and CLY. The reorganization of fractions ultimately leads to a change in soil classification (Table 2). For example, according to the United States Department of Agriculture (USDA) classification [44], before treatment, the bentonites are classified as silty loam (BSvk, STx-1b) or clay loam (SWy-3), while after treatment with ZnCl2, all bentonites are classified as silt in this system. Considering the USDA classification, the most significant hydrological changes occurred in the SWy-3 bentonite. This is concerning, as Na-bentonites are often used as geotechnical barriers [49]. Initially, bentonite was classified in category D as a material with a very low infiltration rate. After treatment with ZnCl2, it moved to category B as a material that is moderately to well permeable. Similarly, according to the classification by the EN-ISO 14688-1 standard [45], ZnCl2-treated bentonites are considered materials with poorer engineering properties, which affects their effectiveness as sealing and contaminant-absorbing materials.
Contamination of bentonites with 1M ZnCl2 led to a shift in the chemical equilibrium, and the pH changed from alkaline (~7.9) to neutral (~6.7). Nartowska et al. [19] observed a pH decrease from 7.21 to 6.09 in clay contaminated with 0.08 M ZnCl2. A similar pH decrease, despite a significantly higher salt concentration, may suggest that bentonites exhibit some resistance to drastic pH changes. However, even a slight reduction in pH to neutral levels may increase metal availability [50]. Previous studies by Nartowska et al. [17] showed that bentonites contaminated with ZnCl2 contain as much as 49.9–61.6% of potentially mobile fractions, with pH change being one of the contributing factors.
In bentonites treated with ZnCl2, a decrease in the liquid limit (LL) and an increase in the plastic limit (PL) is observed. The change in PL is most likely related to a reduction in intermolecular forces caused by a change in surface charge due to interactions with Zn ions. This effect results in a more stable structure, leading to smaller pores and reduced water availability, lowering the LL. The LL decreases up to fivefold in Na-bentonite, while in Ca-bentonite, the reduction is around 30%. Stronger changes in Na-bentonites are likely due to the smaller ionic radius of Na (0.95 Å) compared to Ca (1.0 Å), as well as weaker interactions with the bentonite particles, making Na-bentonites more susceptible to exchange with Zn ions. Ca-bentonites are promising, as the increase in the PL can enhance the flexibility and the ability of the bentonite to retain its properties under changing moisture conditions, providing better sealing. On the other hand, the decrease in the LL reduces excessive swelling, which is important to avoid excessive settlement.

3.3.2. Specific Surface Area (SSA) and Porosity

SSA and porosity are key in assessing the interaction of clay particles with water and ZnCl2. SSAWST is the total area of clay particles, and thus pores, available for this interaction (m2/g). SSABET measures the surface area, which includes both micropores (>0.7 nm) and mesopores (ranging from 2 to 50 nm in diameter). On the other hand, the SSABJH method covers the surface area of mesopores. Soil porosity determines how water and contaminants will move.
Figure 8 shows the nitrogen sorption/desorption isotherms for bentonites before and after ZnCl2 treatments. The results show that adsorption–desorption in bentonites contaminated with ZnCl2 is less effective compared to uncontaminated bentonites, likely due to reduced porosity (Figure 9) and changes in the surface properties of the bentonite caused by the interaction with Zn ions (Table 2). Similarly, Angon et al. [48] stated that high concentrations of potentially toxic metals (including Zn) can reduce soil porosity and pore connectivity, leading to decreased water-holding capacity and increased water runoff.
In Figure 8, adsorption–desorption hysteresis is observed in the higher relative pressure range (p/p0 > 0.4) for both contaminated and uncontaminated bentonites. This phenomenon is typical of porous materials and indicates the occurrence of capillary condensation [51]. High pressure can promote the condensation of molecules in micropores, hindering their removal during desorption, which leads to hysteresis in this range.
The SSABET results for SWy-3 (23.30 m2/g) and STx-1b (82.06 m2/g) are consistent with those reported by Byun et al. [52], who obtained slightly higher values: SWy-3 (28 m2/g) and STx-1b (93.2 m2/g). The minor differences are attributed to the different sample degassing temperatures used (105 °C in our study and 130 °C by Byun et al. [52]), as higher degassing temperatures generally remove more adsorbed water molecules and volatile substances, leading to slightly higher measured surface areas. The contamination of bentonite with ZnCl2 leads to a reduction in the specific surface area (SSABET) and the surface area of mesopores (SABJH) (Table 2), suggesting that contamination may block or alter the pore structure of bentonite, decreasing the availability of adsorption sites. Possible mechanisms include particle aggregation and reduced access to micropores (Table 2, an increase in d10), as well as the blocking of micropores in Mnt by Skl. This assumption may be supported by the higher SABJH values compared to SSABET, which indicates a higher proportion of mesopores in adsorption, as also shown in Figure 8. A similar reduction in SSAWST after ZnCl2 contamination is observed in Ca-bentonites (Table 2). However, in Na-bentonite (SWy-3), an increase in SSAWST is visible. This can be explained by the fact that SSAWST also includes the internal surface of the bentonite, which is unavailable to nitrogen, including interlayer spaces, whose distance increased after contamination of SWy-3 (11.44 Å to 13.2 Å) [17]. It is particularly noticeable in Na-bentonites, where a significant portion of Zn or Skl accumulates in the interlayer space. Similarly, studies by Hsiao et al. [25] demonstrated that the sodium silicate electrolyte in a zinc battery exhibited an increased specific surface area after HCl treatment. The increased SSA in Na-bentonites may enhance their ability to adsorb contaminants, which is beneficial for improving soil quality and removing toxic substances from groundwater. Such changes could increase the effectiveness of bentonites as protective barriers against contaminant migration; however, Skl may be less resistant than Mnt to the release of metals. Studies by Yamamoto et al. [53] confirm that Skl can gradually release Zn even when its structure is not modified.
The following summary explains the mechanisms that occur in bentonites contaminated with ZnCl2, detailing the processes at the mineralogical and physicochemical levels:
  • When Ca- or Na-bentonite with a pH of 7.92 ± 0.12 is treated with 1 M ZnCl2 (pH = 4.51 ± 0.01), the solution predominantly contains zinc ions (Zn2+) and chloride ions (Cl). This treatment initiates an almost complete ion exchange process, in which Zn2+ ions replace Ca2+ or Na+ ions within the montmorillonite (Mnt) structure. As a result, the concentrations of Si, Al, Mg, and Fe decrease in all the studied bentonites. XRF analysis confirms that the content of Ca or Na approaches zero, indicating that the tetrahedral and octahedral layers of Mnt have undergone significant structural alterations. Unstable Mnt phases disappear in the XRD analysis.
  • In addition to the structural changes in montmorillonite (Mnt), a new mineral phase, simonkolleite (Skl), forms during the contamination process. Chemical reactions take place primarily at the montmorillonite phase boundary or within its structure, where the local pH is higher. EDS analysis confirms elevated levels of Zn and Cl in the contaminated bentonites, accompanied by a near-complete depletion of Ca or Na. This chemical transformation is a key indicator of Skl formation and has significant implications for the material’s physical properties, particularly the clay’s surface area and porosity.
  • Upon lowering the pH to 6.78 ± 0.05, SEM imaging reveals the characteristic needle-like morphology and hexagonal platelets of simonkolleite (Skl), confirming the formation of this new mineral phase. XRD analysis further supports its presence, indicating that approximately 30% of the contaminated bentonite consists of Skl.
  • Thermal analysis (TGA) confirms the presence of simonkolleite (Skl) and reveals thermal alterations in the montmorillonite (Mnt) structure resulting from physicochemical changes induced by ZnCl2 contamination. In particular, Zn2+ ions replacing Na+ in Na-bentonite accumulate within the interlayer spaces of Mnt, increasing the interlayer spacing and thereby enhancing the specific surface area for water (SSAWST). These structural modifications facilitate more efficient water flow through the material.
  • In Ca-bentonite, Zn replacing Ca predominantly accumulates on the surface of Mnt particles, leading to a decrease in specific surface area (SSABET) and a reduction in surface area of mesopores (SABJH), and porosity. This reduction is also observed in Na-bentonite. Possible mechanisms contributing to these changes include particle aggregation, reduced access to micropores, and the blocking of micropores in Mnt by Skl. This hypothesis is supported by the higher SABJH values than SSABET, suggesting a higher proportion of mesopores in sorption. The aggregation, increasing effective diameter (d10), driven by processes like flocculation, swelling, dispersion, and salt crystallization, further reduces the ability of the material to adsorb water in micropores, which can increase its permeability.
Based on the observed mineralogical and physicochemical transformations induced by ZnCl2 contamination—such as extensive ion exchange, structural degradation of montmorillonite, simonkolleite formation, and reduced surface area and porosity—we suggest that bentonite barriers exposed to metal-rich leachates may experience long-term performance degradation. These changes could result in increased permeability and reduced contaminant retention, which are critical parameters for barrier integrity.
Consequently, our results emphasize the need for careful geochemical compatibility assessments between bentonite and expected leachate compositions during repository design. Furthermore, regulatory guidelines may need to account for potential mineralogical transformations and include long-term durability testing under realistic chemical conditions. This could help ensure the continued effectiveness of bentonite-based barriers in isolating hazardous waste over extended timescales.

4. Conclusions

This study is the first to analyze the impact of zinc(II) chloride (ZnCl2) contamination on the chemical composition, structure, and porosity of Na- and Ca-bentonites, with a particular focus on the formation of simonkolleite (Skl). Experimental evidence shows that ZnCl2 contamination can significantly impair the physical and chemical properties of these materials by inducing the formation of secondary mineral phases. The novelty of this research lies in establishing a direct link between Zn-induced phase transformation and the deterioration of key functional properties, which is crucial for the long-term reliability of bentonite-based barriers in geotechnical applications.
The main findings of this study are as follows:
  • Contamination with 1 M ZnCl2 significantly alters the mineralogical composition and structural stability of bentonites, as evidenced by the formation of simonkolleite (~30%). These changes were confirmed by XRD, TGA, SEM, and complementary analytical techniques.
  • Ion exchange between Zn2+ and Ca2+/Na+ within the montmorillonite structure leads to a marked reduction in Si, Al, Mg, and Fe content, along with the near-complete depletion of Ca and Na, as demonstrated by XRF and EDS data. EDS surface mapping revealed a uniform distribution of zinc and chlorine across the entire montmorillonite surface, including inter-aggregate spaces, supporting the formation of the simonkolleite phase.
  • The extent of montmorillonite degradation depends on the specific type of bentonite and the montmorillonite phases present. BSvk and STx-1b bentonites (calcium forms) appear to undergo more advanced structural damage than SWy-3 (sodium form), as evidenced by the complete disappearance of certain montmorillonite reflections.
  • The formation of Skl and the associated structural modifications cause a ~50% reduction in specific surface area and porosity, as determined by BET and BJH analyses, potentially compromising the material’s performance as a sealing barrier.
Future research should focus on the following:
  • Quantitative evaluation of hydraulic conductivity and mechanical strength in ZnCl2-contaminated bentonites;
  • Long-term stability assessments under variable environmental conditions;
  • Development of modified bentonite materials with improved resistance to potentially toxic metal-induced structural transformations.

Author Contributions

Conceptualization, E.N.; methodology, E.N. and P.S.; software, E.N., P.S. and M.K.; validation, E.N. and P.S.; formal analysis, E.N. and P.S.; investigation, E.N., P.S. and M.K.; resources, E.N., P.S. and M.K.; data curation, E.N. and P.S.; writing—original draft preparation, E.N.; writing—review and editing, E.N., P.S. and M.K.; visualization, E.N. and P.S.; supervision, E.N.; project administration, E.N.; funding acquisition, E.N., P.S. and M.K. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was partially funded by the Scientific Grant Agency of the Ministry of Education, Science, Research, and Sport of the Slovak Republic and the Slovak Academy of Sciences as part of the research project VEGA 1/0247/23. The research was funded by the Faculty of Environmental Engineering, Geodesy, and Renewable Energy of the Kielce University of Technology. No 05.0.12.00/1.02.001/SUBB.IKGT.25.002.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MntMontmorillonite
SklSimonkolleite
SWy-3Na-bentonite from Wyoming, USA
STx-1bCa-bentonite from Texas, USA
BSvkCa-bentonite from Jelsovy Potok, Slovakia
PTMsPotentially toxic metals
SSASpecific surface area of clay
SASurface area of mesopores
CLYClay fraction
SILSilt fraction
d10Effective diameter—the particle diameter below which 10% of the mass is finer
XRDX-ray diffraction
SEMScanning electron microscopy
TGAThermogravimetric analysis
XRFX-ray fluorescence
EDSEnergy-Dispersive X-ray spectroscopy
BETBrunauer–Emmett–Teller adsorption nitrogen method
BJHBarrett–Joyner–Halenda method
WSTWater sorption test method

Appendix A

Figure A1. TG and DTG curves of bentonite before and after treatments with ZnCl2: (a) natural calcium bentonite from Texas, USA (STx1b), before and (b) after; (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3), before and (d) after; (e) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), before and (f) after.
Figure A1. TG and DTG curves of bentonite before and after treatments with ZnCl2: (a) natural calcium bentonite from Texas, USA (STx1b), before and (b) after; (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3), before and (d) after; (e) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), before and (f) after.
Materials 18 02933 g0a1
Table A1. XRD-identified mineral phases in bentonites with PDF references.
Table A1. XRD-identified mineral phases in bentonites with PDF references.
Soil TypeMineral PhaseKeyChemical FormulaPDF Number *
BSvkMontmorilloniteMnt 1(Na,Ca)0.3(Al,Mg)2Si4O10(OH)2·H2O00-003-0015 **
Mnt 2NaMgAlSiO2(OH)·H2O00-002-0014
QuartzQzSiO204-012-0490
BiotiteBtK(Mg,Fe)3(AlSi3O10)(OH)201-086-2115
STx-1bMontmorilloniteMnt 1(Na,Ca)0.3(Al,Mg)2Si4O10(OH)2·xH2O00-003-0015 **
Mnt 3Na-Al-Si-O-OH-H2O00-003-0019
Mnt 415A Ca0.2(Al,Mg)2Si4O10(OH)4H2O00-013-0135
Mnt 5CaMg2AlSi4(OH)H2O00-002-0239
Opal CTOpal CTSiO200-061-0035
QuartzQzSiO201-085-0457
SWy-3MontmorilloniteMnt 622A Na0.3(Al,Mg)2 Si4O10(OH)2·H2O00-029-1499 **
Mnt 7(Al(OH)2)0.33Al2(Si3.67O10)·8H2O00-011-0303
Mnt 3Na-Al-Si-O-OH-H2O00-003-0019
QuartzQz 1SiO2 01-085-0865
Qz 2SiO201-070-2538
BSvk treated with ZnCl2MontmorilloniteMnt 1(Na,Ca)0.3(Al,Mg)2Si4O10(OH)2·xH2O00-003-0015 **
SimonkolleiteSklZn5(OH)8Cl2·H2O01-077-2311
QuartzQzSiO204-005-0490
BiotiteBtK(Mg,Fe)3(AlSi3O10)(OH)201-086-2115
* STx-1b
treated with ZnCl2
MontmorilloniteMnt 1(Na,Ca)0.3(Al,Mg)2Si4O10(OH)2·xH2O00-003-0015 **
SimonkolleiteSklZn5(OH)8Cl2·H2O01-077-2311
Opal CTOpal CTSiO200-061-0035
QuartzQzSiO201-085-0457
SWy-3
treated with ZnCl2
MontmorilloniteMnt 814A Na0.3(Al,Mg)2Si4O10(OH)2·xH2O00-013-0259 **
Mnt 9AlSi2O6(OH)200-002-0037
SimonkolleiteSklZn5(OH)8Cl2·H2O00-007-0155
QuartzQz 3SiO200-005-0490
* the PDF-4+ database of the International Center for Diffraction Data (ICDD, Newtown Square, PA, USA); ** the major phase of montmorillonite (Mnt) for which the interplanar distances d001 were calculated.
Figure A2. EDS Analysis of BSvk bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure A2. EDS Analysis of BSvk bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g0a2
Figure A3. EDS Analysis of STx-1b bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure A3. EDS Analysis of STx-1b bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g0a3
Figure A4. EDS Analysis of SWy-3 bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure A4. EDS Analysis of SWy-3 bentonite before treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g0a4

References

  1. Neretnieks, I.; Moreno, L. Ion Composition at Bentonite-Water Interface and Its Impact on Erosion; SKB Report P-20-26; Svensk Kärnbränslehantering AB (Swedish Nuclear Fuel and Waste Management Co.): Stockholm, Sweden, 2021; pp. 1–66. [Google Scholar]
  2. Geng, K.; Jin, J.; Chai, J.; Qin, Y. Adsorption characteristics and mechanism analysis of heavy metal Zn2+ by cement-soil and alkali activated slag-bentonite-soil. Case Stud. Constr. Mater. 2024, 21, e03583. [Google Scholar] [CrossRef]
  3. Huang, R.; Wu, L.; Wang, X.; Tang, N.; Gao, L.; Wang, A.; Lu, Y. Review on the effect of isomorphic replacement on the structure and application performance of typical clay minerals. Prog. Nat. Sci. Mater. Int. 2024, 34, 251–262. [Google Scholar] [CrossRef]
  4. Klik, B.; Holatko, J.; Jaskulska, I.; Gusiatin, M.Z.; Hammerschmiedt, T.; Brtnicky, M.; Liniauskienė, E.; Baltazar, T.; Jaskulski, D.; Kintl, A.; et al. Bentonite as a Functional Material Enhancing Phytostabilization of Post-Industrial Contaminated Soils with Heavy Metals. Materials 2022, 15, 8331. [Google Scholar] [CrossRef]
  5. Andriulaityte, I.; Valentukeviciene, M.; Zurauskiene, R. Research on the Reusability of Bentonite Waste Materials for Residual Chlorine Removal. Materials 2024, 17, 5647. [Google Scholar] [CrossRef] [PubMed]
  6. Mooshaee, M.R.; Sabour, M.R.; Kamza, E. The swelling performance of raw and modified bentonite of geosynthetic clay liner as the leachate barrier exposed to the synthetic E-waste leachate. Heliyon 2022, 8, e11937. [Google Scholar] [CrossRef] [PubMed]
  7. Egloffstein, T.A. Natural bentonites. Influence of the ion exchange and partial desiccation on permeability and self-healing capacity of bentonites used in GCL. Geotext. Geomembr. 2001, 19, 427–444. [Google Scholar] [CrossRef]
  8. Cui, Q.; Chen, B. Review of polymer-amended bentonite: Categories, mechanism, modification processes and application in barriers for isolating contaminants. Appl. Clay Sci. 2023, 235, 106869. [Google Scholar] [CrossRef]
  9. Li, D.; Zhao, H.; Tian, K. Hydraulic conductivity of bentonite-polymer geosynthetic clay liners to aggressive solid waste leachates. Geotext. Geomembr. 2024, 52, 900–911. [Google Scholar] [CrossRef]
  10. Somani, M.; Hölzle, I.; Datta, M.; Ramana, G.V. An investigation on mobility of heavy metals for assessing the reusability of soil-like material reclaimed from mining of municipal solid waste dumpsites. Waste Manag. 2023, 167, 113–121. [Google Scholar] [CrossRef]
  11. Bågenholm-Ruuth, E.; Sanchis-Sebastiá, M.; Hollinger, N.; Teleman, A.; Larsson, P.T.; Wallberg, O. Transforming post-consumer cotton waste textiles into viscose staple fiber using hydrated zinc chloride. Cellulose 2024, 31, 737–748. [Google Scholar] [CrossRef]
  12. Lobato, N.C.C.; Villegas, E.A.; Mansur, M.B. Management of solid wastes from steelmaking and galvanizing processes: A brief review. Resour. Conserv. Recycl. 2015, 102, 49–57. [Google Scholar] [CrossRef]
  13. Sun, W.J.; Tang, Q.T.; Lu, T.H.; Fan, R.-D.; Sun, G.-G.; Tan, Y.-Z. Adsorption performance of bentonite and clay for Zn(II) in landfill leachate. Geoenviron. Disasters 2024, 11, 4. [Google Scholar] [CrossRef]
  14. Wuana, R.A.; Okieimen, F.E. Heavy metals in contaminated soils: A review of sources, chemistry, risks, and best available strategies for remediation. ISRN Ecol. 2011, 2011, 402647. [Google Scholar] [CrossRef]
  15. Oyourou, J.-N.; McCrindle, R.; Combrinck, S.; Fourie, C.J.S. Investigation of zinc and lead contamination of soil at the abandoned Edendale mine, Mamelodi (Pretoria, South Africa) using a field-portable spectrometer. J. S. Afr. Inst. Min. Metall. 2019, 119, 55–62. [Google Scholar] [CrossRef]
  16. Pietrzykowski, M.; Antonkiewicz, J.; Gruba, P.; Pająk, M. Content of Zn, Cd and Pb in purple moor-grass in soils heavily contaminated with heavy metals around a zinc and lead ore tailing landfill. Open Chem. 2018, 16, 1143–1152. [Google Scholar] [CrossRef]
  17. Nartowska, E.; Podlasek, A.; Vaverková, M.D.; Koda, E.; Jakimiuk, A.; Kowalik, R.; Kozłowski, T. Mobility of Zn and Cu in Bentonites: Implications for Environmental Remediation. Materials 2024, 17, 2957. [Google Scholar] [CrossRef]
  18. Sahmoune, M.N. Evaluation of thermodynamic parameters for adsorption of heavy metals by green adsorbents. Environ. Chem. Lett. 2019, 17, 697–704. [Google Scholar] [CrossRef]
  19. Nartowska, E.; Podlasek, A.; Vaverková, M.D.; Koda, E.; Kanuchova, M.; Jakimiuk, A.; Gawdzik, J. Identification of factors affecting the properties of soils contaminated with Zn(II) and Cu(II) chlorides. J. Soils Sediments 2025, 25, 1175–1200. [Google Scholar] [CrossRef]
  20. Chai, W.; Huang, Y.; Su, S.; Han, G.; Liu, J.; Cao, Y. Adsorption behavior of Zn(II) onto natural minerals in wastewater. A comparative study of bentonite and kaolinite. Physicochem. Probl. Miner. Process. 2017, 53, 264–278. [Google Scholar] [CrossRef]
  21. Kaya, A.; Ören, A.H. Adsorption of zinc from aqueous solutions to bentonite. J. Hazard. Mater. 2005, 125, 183–189. [Google Scholar] [CrossRef]
  22. Cousy, S.; Gorodylova, N.; Svoboda, L.; Zelenka, J. Influence of synthesis conditions over simonkolleite/ZnO precipitation. Chem. Pap. 2017, 71, 2325–2334. [Google Scholar] [CrossRef]
  23. Moezzi, A.; Cortie, M.; McDonagh, A. Transformation of zinc hydroxide chloride monohydrate to crystalline zinc oxide. Dalton Trans. 2016, 45, 7385–7390. [Google Scholar] [CrossRef] [PubMed]
  24. Enobong, F.D.; Wang, C.; Li, C.; Dong, J.; Udoh, I.I.; Zhang, D.; Zhong, W.; Zhong, S. Evolution of corrosion degradation in galvanised steel bolts exposed to a tropical marine environment. J. Mater. Res. Technol. 2023, 27, 5177–5190. [Google Scholar] [CrossRef]
  25. Hsiao, Y.-F.; Hung, F.-Y.; Zhao, J.-R. Chlorination for all solid-state zinc-graphite battery: Investigate into the charge-discharge mechanism of zinc electrodes and sodium silicate electrolyte. Mater. Sci. Eng. B 2024, 301, 117146. [Google Scholar] [CrossRef]
  26. Oueslati, W.; Rhaiem, H.B.; Amara, A.B.H. Effect of relative humidity constraint on the metal exchanged montmorillonite performance: An XRD profile modeling approach. Appl. Surf. Sci. 2012, 261, 396–404. [Google Scholar] [CrossRef]
  27. PN-ISO 9297:1994; Water Quality—Determination of Chloride—Silver Nitrate Titration with Chromate Indicator (Mohr’s Method). Polish Committee for Standardization: Warsaw, Poland, 1994.
  28. Sun, Q.; Peng, Y.; Georgolamprou, X.; Li, D.; Kiebach, R. Synthesis and characterization of a geopolymer/hexagonal-boron nitride composite for free forming 3D extrusion-based printing. Appl. Clay Sci. 2020, 199, 105870. [Google Scholar] [CrossRef]
  29. Torres-Rivero, K.; Bastos-Arrieta, J.; Fiol, N.; Florido, A. Chapter Ten—Metal and metal oxide nanoparticles: An integrated perspective of the green synthesis methods by natural products and waste valorization: Applications and challenges. In Comprehensive Analytical Chemistry, 1st ed.; Verma, S.K., Das, A.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; Volume 94, pp. 433–469. [Google Scholar] [CrossRef]
  30. EN-ISO 17892-12:2018/Amd 2:2022; Geotechnical Investigation and Testing—Laboratory Testing of Soil—Part 12: Determination of Liquid and Plastic Limits. International Organization for Standardization: Geneva, Switzerland, 2022.
  31. Stępkowska, E.T. Simple method of crystal phase water, specific surface and clay mineral content estimation in natural clays. Stud. Geotechn. 1973, IV, 21–36. [Google Scholar]
  32. Asabina, E.A.; Zhukova, A.I.; Sedov, V.A.; Pet’kov, V.I.; Osaulenko, D.A.; Markova, E.B.; Fukina, D.G.; Koshkin, V.A. Catalytic characteristics of NASICON-type phosphates with rare earth elements in ethanol conversion. Solid State Sci. 2025, 162, 107865. [Google Scholar] [CrossRef]
  33. Borinelli, J.B.; Portillo-Estrada, M.; Costa, J.O.; Pajares, A.; Blom, J.; Hernando, D.; Vuye, C. Emission reduction agents: A solution to inhibit the emission of harmful volatile organic compounds from crumb rubber modified bitumen. Constr. Build. Mater. 2024, 411, 134455. [Google Scholar] [CrossRef]
  34. Chipera, S.J.; Bish, D.L. Baseline studies of the Clay Minerals Society source. Clays: Powder x-ray diffraction analyses. Clays Clay Miner. 2001, 49, 398–409. [Google Scholar] [CrossRef]
  35. Castellini, E.; Malferrari, D.; Bernini, F.; Brigatti, M.F.; Castro, G.R.; Medici, L.; Mucci, A.; Borsari, M. Baseline Studies of the Clay Minerals Society Source Clay Montmorillonite STx-1b. Clays Clay Miner. 2017, 65, 220–233. [Google Scholar] [CrossRef]
  36. Bagheri, M.; Lothenbach, B.; Shakoorioskooie, M.; Scrivener, K. Effect of different ions on dissolution rates of silica and feldspars at high pH. Cem. Concr. Res. 2022, 152, 106644. [Google Scholar] [CrossRef]
  37. Voora, V.K.; Al-Saidi, W.A.; Jordan, K.D.J. Density Functional Theory Study of Pyrophyllite and M-Montmorillonites (M = Li, Na, K, Mg, and Ca): Role of Dispersion Interactions. J. Phys. Chem. A 2011, 115, 9695–9703. [Google Scholar] [CrossRef]
  38. Zhu, K.; He, Y.; He, Q.; Lou, W.; Zhang, Z.; Zhang, K. Effects of ionic strength and bentonite ratio on the migration of Cr(VI) in clayey soil-bentonite engineered barrier. Environ. Sci. Pollut. Res. 2024, 31, 45310–45325. [Google Scholar] [CrossRef]
  39. Liu, L.; Zhang, C.; Jiang, W.; Li, X.; Dai, Y.; Jia, H. Understanding the sorption behaviors of heavy metal ions in the interlayer and nanopore of montmorillonite: A molecular dynamics study. J. Hazard. Mater. 2021, 416, 125976. [Google Scholar] [CrossRef] [PubMed]
  40. Gorodylova, N.; Cousy, S.; Šulcová, P. Thermal transformation of layered zinc hydroxide chloride. J. Therm. Anal. Calorim. 2017, 127, 675–683. [Google Scholar] [CrossRef]
  41. Qu, S.; Hadjittofis, E.; Malaret, F.; Hallett, J.; Smith, R.; Sedransk Campbell, K. Controlling simonkolleite crystallisation via metallic Zn oxidation in a betaine hydrochloride solution. Nanoscale Adv. 2023, 5, 2437–2452. [Google Scholar] [CrossRef] [PubMed]
  42. Sun, Q.; Mao, X.; Yao, Y.; Li, J.; Shen, Z.; Tu, K.-N.; Liu, Y. Effect of WC thickness on the microstructure and properties of WC-C/DLC coated 304 steel. Appl. Surf. Sci. 2025, 704, 163418. [Google Scholar] [CrossRef]
  43. Nartowska, E.; Kozłowski, T. The effect of the concentration of copper ions on the unfrozen water content in bentonites measured with the use of DSC method. Minerals 2022, 12, 632. [Google Scholar] [CrossRef]
  44. United States Department of Agriculture (USDA). Soil Taxonomy: A Basic system of Soil Classification for Making and Interpreting Soil Surveys; U.S. Department of Agriculture: Washington, DC, USA, 1999.
  45. EN-ISO 14688-1; Geotechnical Investigation and Testing—Identification and Classification of Soil—Part 1: Identification and Description. International Organization for Standardization: Geneva, Switzerland, 2002.
  46. EN-ISO 10390:2021; Water quality—Determination of pH. International Organization for Standardization: Geneva, Switzerland, 2021.
  47. Tang, S.; She, D.; Wang, H. Effect of salinity on soil structure and soil hydraulic characteristics. Can. J. Soil Sci. 2020, 101, 62–73. [Google Scholar] [CrossRef]
  48. Angon, P.B.; Islam, M.S.; Kc, S.; Das, A.; Anjum, N.; Poudel, A.; Suchi, S.A. Sources, effects and present perspectives of heavy metals contamination: Soil, plants and human food chain. Heliyon 2024, 10, e28357. [Google Scholar] [CrossRef] [PubMed]
  49. Nath, H.; Kabir, M.H.; Kafy, A.-A.; Rahaman, Z.A.; Rahman, M.T. Geotechnical properties and applicability of bentonite-modified local soil as landfill and environmental sustainability liners. Environ. Sustain. Indic. 2023, 18, 100241. [Google Scholar] [CrossRef]
  50. Wei, L.; Li, Y.; Noguera, D.R.; Zhao, N.; Song, Y.; Ding, J.; Zhao, Q.; Cui, F. Adsorption of Cu2+ and Zn2+ by extracellular polymeric substances (EPS) in different sludges: Effect of EPS fractional polarity on binding mechanism. J. Hazard. Mater. 2017, 321, 473–483. [Google Scholar] [CrossRef] [PubMed]
  51. Viisanen, Y.; Lbadaoui-Darvas, M.; Alvarez Piedehierro, A.; Welti, A.; Nenes, A.; Laaksonen, A. Water vapor adsorption-desorption hysteresis due to clustering of water on nonporous surfaces. Langmuir 2024, 40, 20311–20321. [Google Scholar] [CrossRef] [PubMed]
  52. Byun, Y.; Seo, C.; Yun, T.; Joo, Y.; Jo, H.Y. Prediction of Na- and Ca-montmorillonite contents and swelling properties of clay mixtures using Vis-NIR spectroscopy. Geoderma 2023, 430, 116294. [Google Scholar] [CrossRef]
  53. Yamamoto, O.; Nagashima, M.; Nakata, Y.; Udagawa, E. The significant potential of Simonkolleite powder for deep wound healing under a moist environment: In vivo histological evaluation using a rat model. Bioengineering 2023, 10, 375. [Google Scholar] [CrossRef]
Figure 1. XRD diffractograms of bentonites before and after treatments with ZnCl2: (a) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) natural calcium bentonite from Texas, USA (STx-1b), (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3).
Figure 1. XRD diffractograms of bentonites before and after treatments with ZnCl2: (a) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) natural calcium bentonite from Texas, USA (STx-1b), (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3).
Materials 18 02933 g001aMaterials 18 02933 g001b
Figure 2. TG and DTG curves of bentonite before and after treatments with ZnCl2: (a) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) natural calcium bentonite from Texas, USA (STx-1b), (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3).
Figure 2. TG and DTG curves of bentonite before and after treatments with ZnCl2: (a) natural calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) natural calcium bentonite from Texas, USA (STx-1b), (c) natural sodium–calcium bentonite from Wyoming, USA (SWy-3).
Materials 18 02933 g002
Figure 3. SEM microphotographs of bentonites before and after treatment with ZnCl2: (a) calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) BSvk after treatment, (c) calcium bentonite from Texas, USA (STx-1b), (d) STx-1b after treatment, (e) sodium bentonite from Wyoming, USA (SWy-3), (f) SWy-3 after treatment, (g) the needle-like structure of simonkolleite in BSvk bentonite after treatment, (h) the needle-like structure of simonkolleite in SWy-3 bentonite after treatment.
Figure 3. SEM microphotographs of bentonites before and after treatment with ZnCl2: (a) calcium bentonite from Jelsový Potok, Slovakia (BSvk), (b) BSvk after treatment, (c) calcium bentonite from Texas, USA (STx-1b), (d) STx-1b after treatment, (e) sodium bentonite from Wyoming, USA (SWy-3), (f) SWy-3 after treatment, (g) the needle-like structure of simonkolleite in BSvk bentonite after treatment, (h) the needle-like structure of simonkolleite in SWy-3 bentonite after treatment.
Materials 18 02933 g003aMaterials 18 02933 g003b
Figure 4. Microcrystal morphology of simonkolleite (Skl) observed in SEM microphotographs of bentonites after treatment with ZnCl2: (a,c) sodium bentonite from Wyoming, USA (SWy-3); (b,d) calcium bentonite from Jelsový Potok, Slovakia (BSvk). Note: The white arrow indicates hexagonal platelets of simonkolleite.
Figure 4. Microcrystal morphology of simonkolleite (Skl) observed in SEM microphotographs of bentonites after treatment with ZnCl2: (a,c) sodium bentonite from Wyoming, USA (SWy-3); (b,d) calcium bentonite from Jelsový Potok, Slovakia (BSvk). Note: The white arrow indicates hexagonal platelets of simonkolleite.
Materials 18 02933 g004
Figure 5. EDS analysis of BSvk bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure 5. EDS analysis of BSvk bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g005
Figure 6. EDS analysis of STx-1b bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure 6. EDS analysis of STx-1b bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g006
Figure 7. EDS analysis of SWy-3 bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Figure 7. EDS analysis of SWy-3 bentonite after treatment with ZnCl2: (a) area of analysis with microphotograph; (b) elemental composition spectrum; (c) spatial distribution of selected elements on the sample surface.
Materials 18 02933 g007
Figure 8. Nitrogen sorption/desorption isotherms for bentonites before and after ZnCl2 treatment: (a) calcium bentonite from Jelsový Potok, Slovakia (BSvk); (b) calcium bentonite from Texas, USA (STx1b); (c) sodium bentonite from Wyoming, USA (SWy-3).
Figure 8. Nitrogen sorption/desorption isotherms for bentonites before and after ZnCl2 treatment: (a) calcium bentonite from Jelsový Potok, Slovakia (BSvk); (b) calcium bentonite from Texas, USA (STx1b); (c) sodium bentonite from Wyoming, USA (SWy-3).
Materials 18 02933 g008
Figure 9. Total porosity of bentonites before and after treatment with ZnCl2 calculated using BJH method.
Figure 9. Total porosity of bentonites before and after treatment with ZnCl2 calculated using BJH method.
Materials 18 02933 g009
Table 1. Results of ICP-OES and XRF analysis for bentonites.
Table 1. Results of ICP-OES and XRF analysis for bentonites.
Chemical
Analysis
ElementBSvk
(Ca Form)
BSvk Treated with ZnCl2STx-1b
(Ca Form)
STx-1b
Treated with ZnCl2
SWy-3
(Na Form)
SWy-3
Treated with ZnCl2
ICP-OES &
[17]
[mg/kg dry mass]
Ca11,945 ± 1402778 ± 3111,802 ± 1012985 ± 238282 ± 574526 ± 34
Na1151 ± 91204 ± 221970 ± 18885 ± 1610,086 ± 81995 ± 9
Zn64.54 ± 0.6917,857 ± 8973.68 ±0.2716,153 ± 75163.66 ± 1.5044,463 ± 124
XRF *Ca1.140.021.330.031.020.04
Mg2.320.762.150.721.740.53
K0.190.090.140.060.420.18
Na0.170.311.22
Al8.423.687.162.858.313.39
Si24.2911.3837.2015.1727.1411.73
Fe1.670.990.820.412.551.42
Cl2.162.152.88
XRF (potentially toxic metals)Zn21.2221.2421.01
Mn0.05110.02730.01340.00620.01480.0044
Co<0.0001<0.0001<0.0001<0.0001<0.0001<0.0001
Ni0.0010<0.00010.0009<0.00010.0010<0.0001
Cu0.0006<0.00010.0005<0.00010.0006<0.0001
Cd<0.00005<0.00002<0.00005<0.00002<0.00002<0.00002
Pb0.00280.00200.00120.00280.003670.00113
− below detection limits; & total content of major elements in the dry clay matrix determined with ICP-OES [17]; * standard deviation: Zn, Al ± 0.01; Cl ± 0.001; other elements < ±0.001.
Table 2. Physical properties of bentonites before and after treatments with ZnCl2.
Table 2. Physical properties of bentonites before and after treatments with ZnCl2.
PropertiesBSvk BSvk
Treated with ZnCl2
STx-1bSTx-1b Treated with ZnCl2SWy-3 SWy-3 Treated with ZnCl2
Soil classification $ USDAsilt loamsiltsilt loamsiltclay loamsilt
EN-ISO 14688-1clayey siltsiltclayey siltclayey siltclaysilt
pHKCl(-) *7.826.767.886.758.056.84
alkalineneutralalkalineneutralalkalineneutral
Granulometric **CLY (%)19.88.918.510.742.66.6
SIL (%)80.290.281.58757.487
SA (%)0.00.90.02.30.36.4
d10 (μm)1.332.181.391.910.962.83
Plasticity §PL (%)466144733554
LL (%)165119142101519104
Sorption/desorption #SABJH (m2/g)58.8022.1996.2757.2131.2415.33
SSABET (m2/g)41.3914.3882.0640.5523.3017.16
SSAWST (m2/g)671557568538307516
w95 (%)30.14 20.10 29.0822.6721.6717.05
Porosity ϵtotal (cm3/g)0.1370.0690.1960.1300.0760.048
Pore radiusDv (nm)1.9211.9231.9311.9251.9241.920
Notes: $ the United States Department of Agriculture (USDA) [44], EN-ISO 14688-1 [45]; * EN ISO 10390:2021 [46]. The pH of the bentonites treated with ZnCl2 was determined after rinsing them to remove excess chloride ions. ** Laser diffraction method; CLY—clay (d ≤ 0.002 mm), SIL—silt (0.002 mm < d < 0.063 mm), SA—sand (0.063 mm < d < 2 mm), d10 (effective diameter—10% of the particles have a smaller diameter than d10) [17]. § Rolling test; Casagrande method acc. to EN ISO/TS 17892-12 [30]. # SSA—soil specific surface area. The Brunauer–Emmett–Teller (BET) method, the water vapor sorption test method (WST); SA—surface area of mesopore based on the Barrett–Joyner–Halenda (BJH) model (desorption); w95—the sorptive moisture at p/p0 = 0.95 [30].
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

Nartowska, E.; Stępień, P.; Kanuchova, M. Impact of Zinc(II) Chloride Contamination on Bentonites: Formation of Simonkolleite and Effects on Porosity and Chemical Composition. Materials 2025, 18, 2933. https://doi.org/10.3390/ma18132933

AMA Style

Nartowska E, Stępień P, Kanuchova M. Impact of Zinc(II) Chloride Contamination on Bentonites: Formation of Simonkolleite and Effects on Porosity and Chemical Composition. Materials. 2025; 18(13):2933. https://doi.org/10.3390/ma18132933

Chicago/Turabian Style

Nartowska, Edyta, Piotr Stępień, and Maria Kanuchova. 2025. "Impact of Zinc(II) Chloride Contamination on Bentonites: Formation of Simonkolleite and Effects on Porosity and Chemical Composition" Materials 18, no. 13: 2933. https://doi.org/10.3390/ma18132933

APA Style

Nartowska, E., Stępień, P., & Kanuchova, M. (2025). Impact of Zinc(II) Chloride Contamination on Bentonites: Formation of Simonkolleite and Effects on Porosity and Chemical Composition. Materials, 18(13), 2933. https://doi.org/10.3390/ma18132933

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