Next Article in Journal
Optimizing the Y Content of Welding Wire for TIG Welding of Sand-Cast Mg-Y-RE-Zr Alloy
Previous Article in Journal
Study on Multi-Factor Coupling Fatigue Properties of Weathering Steel Welded Specimens
Previous Article in Special Issue
Materials for Heavy Metals Removal from Waters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparison of Cu(II) Adsorption Using Fly Ash and Natural Sorbents During Temperature Change and Thermal–Alkaline Treatment

by
Anna Ďuricová
1,
Veronika Štefanka Prepilková
2,
Michal Sečkár
1,
Marián Schwarz
1,
Dagmar Samešová
1,
Tomáš Murajda
1,
Peter Andráš
3,
Adriana Eštoková
4,
Miriama Čambál Hološová
4,
Juraj Poništ
1,*,
Andrea Zacharová
1,
Jarmila Schmidtová
5,
Darina Veverková
6 and
Adrián Biroň
7
1
Department of Environmental Engineering, Faculty of Ecology and Environmental Sciences, Technical University in Zvolen, T. G. Masaryka 24, 96001 Zvolen, Slovakia
2
Fortischem a.s., M. R. Štefánika 1, 97271 Nováky, Slovakia
3
Department of Geography and Geology, Matej Bel University in Banská Bystrica, Tajovského 40, 97401 Banská Bystrica, Slovakia
4
Department of Material Engineering, Institute for Sustainable and Circular Construction, Faculty of Civil Engineering, Technical University of Košice, Vysokoškolská 4, 04200 Košice, Slovakia
5
Department of Mathematics and Descriptive Geometry, Faculty of Wood Sciences and Technology, Technical University in Zvolen, T. G. Masaryka 24, 96001 Zvolen, Slovakia
6
Institute of Foreign Languages, Technical University in Zvolen, T. G. Masaryka 24, 96001 Zvolen, Slovakia
7
Earth Science Institute v.v.i., Slovak Academy of Sciences, Ďumbierska 1, 97411 Banská Bystrica, Slovakia
*
Author to whom correspondence should be addressed.
Materials 2025, 18(19), 4552; https://doi.org/10.3390/ma18194552
Submission received: 11 August 2025 / Revised: 16 September 2025 / Accepted: 17 September 2025 / Published: 30 September 2025
(This article belongs to the Special Issue Materials for Heavy Metals Removal from Waters (2nd Edition))

Abstract

Mine effluents represent a serious environmental problem on a global scale. Therefore, the effective treatment of this water is a serious issue in the scientific field. The adsorption process seems to be one of the attractive methods, especially due to the simplicity of design, affordability or high efficiency. The latest scientific knowledge has shown that the use of waste and natural adsorbents is economical and effective. This study aimed to evaluate the efficiency of the adsorption process of natural and waste materials—zeolite, bentonite and fly ash—under the influence of temperature and modification of these adsorbents. The novelty of this study resides in an adjustment of the modification method of adsorbents compared to previous research: thermal–alkaline treatment versus hydrothermal one. Another novelty is the use of modified fly ash from biomass combustion as an adsorbent in comparison with the previously used fly ash from coal combustion. The modification of the adsorbents made the adsorption process more effective at all experimental concentrations. The characterisation of adsorbent samples was performed using X-ray diffraction (XRD). The parameters of the adsorption isotherms, Langmuir, Freundlich and Temkin, were estimated by nonlinear regression analysis. The adsorption capacity of Cu(II) of fly ash was comparable to natural adsorbents. Adsorption processes were better described by pseudo-second-order kinetics. At the end of this study, the suitability of using the adsorbents to reduce the concentration of Cu(II) in neutral mine effluents was observed in the following order at 30 °C: unmodified fly ash > modified bentonite > unmodified zeolite. At the temperatures of 20 °C and 10 °C, the same trend of the suitability of adsorbents use was confirmed: modified bentonite > modified zeolite > modified fly ash. The practical applicability of this study lies in the expansion of knowledge in the field of adsorption processes and in the improvement of waste management efficiency of heating plants not only in Slovakia, but also globally.

1. Introduction

Mine wastewater can have very low or, conversely, high pH, depending on complex interactions between hydrological, chemical and biological processes. These waters are typically rich in heavy metals and sulphur (present in minerals) but low in nutrients [1]. Mine drainages also contain chemicals that can pose potential risks, such as salts, nitrates and sulphates [2]. Due to the neutralizing properties of waste minerals or human interventions, such as the addition of limestone to precipitate metals, the pH of these waters can rise to values between 4.5 and 8.5, in which case the water is referred to as neutral mine drainage (NMD) [1].
NMD is less widespread globally than AMD, and since its environmental impacts are not as severe, it has received relatively little attention. Nevertheless, NMD often contains high levels of dissolved metals [3]. NMD is commonly associated with low dissolved oxygen and medium to high concentrations of sulphate and metals that are soluble (e.g., Cu, Zn, Cd) or insoluble (Fe, Al, Mn) under aerobic conditions [4]. Such drainage poses a potentially serious environmental risk because heavy metals, often present at high concentrations, can remain dissolved under suitable redox conditions even in an alkaline environment [1]. Neutral mine drainage (NMD) can pose an environmental risk if the disposal of low-sulphide waste rock at mining sites is not managed properly. Unlike acid mine drainage (AMD), predicting the behaviour of NMD using conventional kinetic tests is challenging, as processes of contaminant immobilization—particularly through sorption and precipitation—occur in these waters [5]. NMD causes surface water contamination, with most cases producing drainage of nearly neutral pH, particularly with elevated zinc concentrations [3].
Excessive levels of the heavy metal Cu(II) in groundwater and surface water can lead to serious environmental contamination [6]. In the site that we monitored, the preliminary measurements in our study revealed increased concentrations of Cu(II) exceeding the limit values. In the human body, copper is essential for the development of connective tissue, nerve sheaths and bones, and it is also a part of several enzymes that catalyse important biological reactions [7].
At present, only limited attention is given to the adsorption of Cu from neutral mine drainage [8,9] compared with its removal from acid mine drainage (AMD) [10,11,12,13,14,15]. Therefore, we decided to focus our study on the removal of ions of this potentially toxic element from NMD. In addition to process efficiency, economic demand is also important in the feasibility of wastewater treatment. Therefore, it is important to consider this factor. Among the costly traditional methods used to remove heavy metal ions from wastewater are chemical precipitation [16], solvent extraction [17] and ion exchange [18]. Adsorption can be considered a favourable alternative for the removal of heavy metal ions from wastewater [19], in which natural materials, agricultural by-products and industrial waste of biological origin can be used as adsorbents of high efficiency [20]. It has been proven that adsorption is an effective cleaning method because it provides significant advantages, including stability, utility, low costs, ease of operation and performance. Adsorption technology has great advantages, since it reduces heavy metal ion concentrations to very low levels and uses various low-cost adsorption materials, including biosorbents, clays, activated carbon, zeolites and metal oxides [21]. Marine red algae [22] and Saccharomyces cerevisiae [23] are among the biomaterials recently used in biosorption. We decided to use bentonite and zeolite as natural adsorbents due to their availability. In the past, both materials were confirmed as suitable adsorbents for the treatment of mine effluents. For the treatment of NMD, we also used biomass waste ash, which is available in our location from the heating plant, and also has a positive impact on reducing this waste when used as a secondary product. The positive effect of the fly ash used as a waste adsorbent was confirmed by studies in the past [24,25]. To enhance the adsorption process, thermal–alkaline treatment was employed. The positive effect of the thermal–alkaline treatment on adsorption was confirmed by studies [26,27].
In addition to thermal–alkaline treatment, the effect of temperature change on the sorption course was monitored. Selected studies confirmed the positive effect of temperature on the adsorption of heavy metals [28,29]. However, beyond a certain temperature, a negative effect was confirmed (when the temperature increased) [30]. On the contrary, in the study [31], the direct effect of temperature on improvement or deterioration of adsorption was not confirmed, although the affinity for the binding site decreased significantly with increasing temperature. The main targets of our study are to expand knowledge in the field of adsorption processes (monitoring the effectiveness of adsorption by temperature change, using thermal–alkaline treatment), and to support waste management of heating plants (reducing the volume of waste fly ash by using it as an adsorbent). The practical applicability of this study is that its findings will support the scientific field in the remediation of AMD and the use of waste fly ash as a secondary product, not only in Slovakia but also globally.
The novelty of this study lies in the modification of the adsorbent preparation method compared to previous research, using a thermal–alkaline treatment instead of a hydrothermal one. It has been confirmed that thermal–alkaline treatment leads to an increase in the BET surface area, total pore volume and adsorption capacity of the material [32]. The alkaline activation of the adsorbents was carried out by applying solid pure NaOH and NaNO3 (sample ratio, NaOH: NaNO3—7:3:10) in an autoclave at a temperature of 350 °C for 12 h. The same methodology was used in the study by Choi et al. [33], but in their case, fly ash was obtained from burning anthracite coal, whereas our fly ash was obtained from burning biomass. The mentioned method of thermal–alkaline treatment in our case differed from hydrothermal–alkaline treatments compared to other studies [34,35]. In the mentioned studies, a highly alkaline environment was created using 5 M NaOH, but the temperature acting on our adsorbents was higher compared to the mentioned studies. In our case, the adsorbents were not introduced into the solution but homogenised with solid alkali. In the study [34], adsorbents were modified at temperatures of 100 and 200 °C during 3–48 h of action, and in the study [35], hydrothermal–alkaline treatment was performed at temperatures of 120–250 °C, during 10–36 h. Unlike the hydrothermal method, our thermal–alkaline pre-treatment method uses a mixture of salts as a reaction medium without any addition of water. The goal of this difference was to monitor the adsorption mechanism under thermal–alkaline pre-treatment conditions without the need to introduce the adsorbent into the solution and to monitor the effect of high temperature on the structure of the adsorbents.
Another innovation is the use of modified fly ash from biomass combustion as an adsorbent, rather than the coal-derived fly ash used earlier. Ash produced from coal and biomass combustion exhibits considerable variability in its physical and chemical properties. Its chemical composition is crucial, as it determines potential applications across different sectors. While coal ash is richer in SiO2, Fe2O3 and Al2O3, biomass ash contains higher levels of Na2O, K2O, MgO and P2O5 [36].

2. Materials and Methods

2.1. Sampling Locations

The neutral mine drainage was collected from the Voznicka heritage tunnel in the Štiavnica–Hodruša mining district (central Slovakia), as shown in Figure 1. Ash from biomass combustion was obtained from the thermal power plant in Zvolen (Figure 1). Natural bentonite was obtained from mining in the Kopernica area (Figure 1). Natural zeolite was obtained from mining in the Nižný Hrabovec area (Figure 1).
Extraction of neutral mine drainage: The neutral mine drainage was taken from the Voznicka heritage tunnel in the Štiavnica–Hodruša mining district (central Slovakia)—48°27′ N latitude and 18°42′ E longitude. This water represents a legacy of mining activities in the area. The total historical production of mines in the mining district is estimated at 4000 t of Ag and 80 t of Au. During the mining of non-ferrous metals (19th century, until 1992), approximately 70,000 t of Zn, 55,000 t of Pb and 8000 t of Cu(II) were mined [37]. The central Slovak volcanic field contains several Ag–Au deposits of the epithermal vein type, which in the past were a source for mining of both precious and base metals [38].
The sampling was carried out at the site where the monitoring of Dionýz Štúr was conducted by the State Geological Institute. The samples were collected in plastic bottles, stored in the cold and transported to the laboratory in accordance with the requirements of STN EN ISO 5667-1:2007 [39]. Guidelines for the design of sampling programmes and sampling techniques. For microbiological analysis, water samples were taken from the heritage drainage tunnels and from drainage water originating in the tailings pond. The samples were collected in sterile glass bottles, kept in the cold and transported to the laboratory in accordance with the requirements of STN EN ISO 19458 [40]. Sampling for microbiological analysis.

2.2. Analysis Methods

The individual characteristics of the samples were obtained by standard methods, which are presented in Table 1 and Table 2.
Atomic Absorption Spectrometry (AAS): An AAS AVANTA Σ flame atomisation spectrometer (GBC Scientific) was used to determine metal concentrations during the adsorption process.
Atomic emission spectrometry with the inductively coupled plasma (AES-ICP): A method of emission spectroscopy that excites atoms and ions with plasma, causing them to emit electromagnetic radiation with wavelengths specific to a particular element.
Elemental analysis with thermal conductivity detection (EA-TCD): A thermal conductivity detector (TCD), a katharometer, is a volumetric property detector and a chemical-specific detector commonly used in gas chromatography. This detector detects changes in the thermal conductivity of the eluent of the column and compares it with the reference flow of the carrier gas. Since most compounds have a thermal conductivity much lower than the common carrier gases such as helium or hydrogen, as the analyte elutes from the column, the thermal conductivity of the effluent is reduced, and a detectable signal is produced.
Brunauer–Emmett–Teller (BET) analysis: Multipoint measurement of the specific surface area of an analyte (m2·g−1) through gas adsorption analysis, where an inert gas such as nitrogen continuously flows through a solid sample, or the solid sample is suspended in a defined volume of gas. Small gas molecules are adsorbed on the solid substrate and its porous structures due to weak Van der Waals forces and form a monolayer of adsorbed gas. This monolayer and adsorption rate can be used to calculate the specific surface area of the solid sample and its porous geometry, providing information for reactivity and bioavailability studies of pharmaceutical products.
Thermal–alkaline activation: In order to modify the adsorbent, thermal–alkaline treatment without the addition of water was used. Typically, mixtures containing 0.7 g of adsorbent, 0.3 g of NaOH, and 1 g of NaNO3 were ground to a fine powder in a Pt crucible and heat-treated at 350 °C (±5 °C) for 12 h.
Adsorption of heavy metals: Adsorption was carried out at temperatures of 30 °C, 20 °C and 10 °C in Erlenmeyer flasks. These temperatures were chosen to simulate the climate conditions of the studied area in the spring, summer and autumn periods. The purpose of these temperatures was to better understand the adsorption mechanism under real temperature conditions. The experimental conditions of sorption included stirring at 500 rpm (revolutions per minute) using a magnetic stirrer, a pH of 7.6 and an initial adsorbate concentration (mg·L−1) to adsorbent dosage (g) ratio of 20, 36, 48, 64 and 72. During the adsorption, the change of pH in the adsorbates was also monitored. The reason for this is the change in the pH of mine effluents due to specific adsorbents. The sample volume in each flask was 100 cm−3. Before adsorption, 500 cm−3 of solution was prepared for each input concentration—a total of 5 input concentrations. A volume of 100 cm−3 was taken to determine the initial concentration. Then, 100 cm−3 of solution was added to each flask and mixed with 0.25 g of adsorbent. After 30, 60, 90 and 120 min, adsorption was stopped by filtering out the adsorbent in individual bottles, and then the concentration of metals—copper—was measured. The temperature was reduced to 10 °C in an incubator. The adsorption of each input concentration was repeated 5 times for the relevance of the evaluation—a total of 6 measurements were performed for each input concentration for each adsorbent at each temperature. The addition of CuSO4·5H2O was used to change the initial concentrations for Cu adsorption. Adsorption took place under continuous stirring to optimise the contact between the aqueous phase and the adsorbent. Mine water solutions with modified Cu concentrations for the needs of 5 input concentrations were used for adsorption processes.
Determination of pH and oxidation–reduction potential follows from STN EN ISO 10390 using a potentiometrically combined electrode. The oxidation–reduction potential is a measure of electron activity and an indicator of the ability of a biogeochemical system to receive or transfer electrons. It is the electrical potential required to transfer electrons from one compound/element (oxidising agent) to another compound/element (reducing agent). Oxidation–reduction potential values close to −300 mV indicate a highly reducing environment or the ability to supply electrons, while values from +100 mV to +300 mV indicate a highly oxidising environment or the ability to accept electrons [46]. The inoLab pH Level 1 device (WTW Weilheim, Weilheim in Oberbayern, Germany) was used to determine pH and oxidation–reduction potential.
X-ray powder diffraction: Mineralogical composition of both initial and treated samples was determined using the X-ray powder diffraction technique. Sample preparation for XRPD analysis was conducted as follows. Exactly 0.5000 g of internal standard Al2O3 (AL-OX-03-P, nominal grain-size 3–4 µm; producer: American Elements Corp., Los Angeles, CA, USA) was added to 2.0000 g of sample mixed with 4 cm−3 of ethanol and ground for 5 min in McCrone Micronising Mill (©Retsch GmbH, Haan, Germany) using a set of zirconia grinding elements. The resulting slurry was dried at 75 °C overnight. To achieve maximum randomisation of the sample, water slurry was prepared and spray-dried using the Spray Drying Kit (The James Hutton Institute, Aberdeen, Scotland [47]).
The XRPD analyses were conducted on a Bruker D8 Advance diffractometer (Karlsruhe, Germany) using β-filtered CuKα radiation generated at 40 kV and 40 mA and a position-sensitive SSD160 detector working in the 1D regime. The beam was collimated with a slit assembly 0.3–6 mm PSD opening 0.4984°, with primary and secondary Soller slits 2.5°. Diffraction patterns were recorded from 3–70°2θ with a step size of 0.01948°2θ and counting 2s per step. Samples were analysed in a top-loaded holder.
The evaluation of XRPD patterns and the identification of individual phases were carried out with DIFFRAC.EVA software Version 4.2.1.10 [48]. The quantitative Rietveld analysis was performed with TOPAS 5 software using a fundamental parameters approach (internal standard method) [49,50]. As starting structural models, the following published data were used: McGinnety (1972) for arcanite [51], Hassan & Grundy (1991) for cancrinite [52], Comodi et al. (2004) [53] for biotite (phlogopite), Hughes et al. (1989) for hydroxylapatite [54], Angel et al. (2013) for K-feldspar (orthoclase) [55], Hazen (1976) for periclase [56], Gualtieri (2000) for plagioclase [57], Desgranges et al. (1993) for portlandite [58], Antao et al. (2008) for quartz [59] and Gournis et al. (2008) for smectite (montmorillonite) [60]. During the Rietveld refinement, the following parameters were considered: emission profile of radiation, scale factor, polynomial background, correction of the zero-point of the goniometer, sample displacement error, lattice parameters and atomic positions (major phases only). However, atomic occupancy and isotropic displacement factors were kept fixed. No attempt to model the turbostratic disorder common for smectite-group phyllosilicates was made.
Samples revealing high amounts of smectite or clinoptilolite were quantitatively evaluated with RockJock Version 1111 software (method based on full-pattern summation and mineral intensity factors [61,62]). The mineral composition of initial samples/materials and its changes upon thermochemical treatment is shown in Table 3.
The powder X-ray diffraction analysis of adsorbent samples (Table 3) before adsorption experiments confirmed the change in the mineralogical structure of the adsorbents after the alkaline modification process.
The XRPD patterns of initial and treated adsorbents are shown in Figure 2, Figure 3 and Figure 4.
The XRPD pattern of original bentonite, shown in Figure 2 (Be1), indicated dioctahedral smectite (d(06,33) = 1.495 Å) with minor admixtures of K-feldspars, quartz, opal-CT, biotite and kaolinite. The quantitative analysis revealed the presence of 73 wt.% smectite, 11 wt.% K-feldspars (orthoclase and sanidine), 6 wt.% quartz and opal-CT and 2 wt.% biotite and kaolinite. The XRPD pattern of the treated sample (Figure 2, Be2) displayed a significant decrease in the intensity of both basal and non-basal smectite reflections. The decrease in basal spacing from 14.7 Å to 12.1 Å indicates changes in the interlayer cation composition and hydration state (i.e., divalent cations were replaced by monovalent ones). Quantitatively, smectite formed only app. 1 wt.% of the treated sample, and it was replaced by the newly crystallized phase, identified as cancrinite (16 wt.%). Traces of calcite (<1 wt.%) were also recorded. Minor phases do not show significant quantitative changes except for the disappearance of opal-CT and kaolinite. However, it should be noted that in treated bentonite, the most abundant reaction product is the amorphous component (72 wt.%, possibly silica). This is documented on the XRPD pattern by an elevated background in the interval 19–35°2θ.
The XRPD analysis of zeolite from the Nižný Hrabovec deposit showed dominancy of clinoptilolite, which formed 76 wt.% of the sample (Figure 3). Supplementary phases are represented by minor opal-CT (15 wt.%) and accessory plagioclase (4 wt.%), smectite (3 wt.%), biotite and K-feldspar (both <1 wt.%). The treatment resulted in a reduction in clinoptilolite content to 29 wt.%, an increase in plagioclase content to 11 wt.% and the formation of 4 wt.% calcite (Figure 3). Again, the most pronounced response to the treatment procedure was the appearance of 50 wt.% of amorphous material.
The XRPD analysis of original ash allowed the identification of several mineral phases, which were subsequently completely removed from the sample after thermochemical treatment (Figure 4, Po1): arcanite (17 wt.%), portlandite (16 wt.%), periclase (6 wt.%) and hydroxylapatite (5 wt.%). Instead, 35 wt.% plagioclase and 7 wt.% K-feldspar were identified as reaction products (Figure 4, Po2). On the other hand, only minor increases in calcite (from 29 to 34 wt.%) and quartz (from 22 to 24 wt.%) contents were detected.
In zeolite, clinoptilolite (76 wt.%) and opal-CT were initially dominant. After modification, the clinoptilolite content decreased to 29 wt.%, the amount of plagioclase increased to 11 wt.%, and calcite formed at 4 wt.%. These structures are favourable for the incorporation of Cu(II) into the newly formed minerals. The most significant response to the modification procedure was the occurrence of 50 wt.% amorphous material, which likely limited the higher adsorption capacity of the modified zeolite. A temperature of 20 °C appears to be the most favourable for adsorption by the modified zeolite.
The modification processes of both bentonite and zeolite altered the mineralogical composition of their phases. In bentonite, opal-CT and kaolinite disappeared, and there was a significant decrease in the intensity of both basal and non-basal reflections of smectite. These minerals were replaced by a newly crystallized cancrinite phase (16 wt.%), which could effectively immobilize Cu(II) in the final matrices enriched with cancrinite. The immobilization mechanism involves the speciation of silicates, hydroxides or oxides and the encapsulation of such substances within cancrinite agglomerates and other aluminosilicate agglomerates. The lower BET surface area is likely due to the presence of the amorphous component.
The behaviour of fly ash is related to the mineralogical changes occurring during its thermal modification. While the raw fly ash contained 44 wt.% mineral phases such as arcanite, portlandite, periclase and hydroxylapatite, in the modified fly ash, the disappearance of these phases was accompanied by an increase in calcite and, especially, plagioclase content. The adsorption of Cu(II) in fly ash at relatively high pH (9.2–9.75) was likely influenced by the presence of periclase and portlandite, and elevated temperature promoted surface reaction mechanisms on the adsorbent.
In the case of modified fly ash, the increased content of plagioclase and calcite contributed to the immobilization of Cu(II) through its incorporation into the structures of aluminosilicate agglomerates. At 30 °C, desorption processes likely occurred due to the increased heterogeneous surface area.
Most of the identified mineral phases are sodium/calcium aluminosilicates (cancrinite, smectite and plagioclase). In the amorphous structure of modified zeolites and bentonites, the negative charge does not appear to be localized but is more or less evenly distributed within the structure. The Si(IV) and Al(III) cations in the aluminosilicate framework are tetrahedrally coordinated and interconnected by oxygen bridges. The negative charge on the AlO4 group is balanced by alkali cations (typically Na+ or K+). Based on XRD analysis, we identified the presence of an amorphous Al-Si-O structure, which is also responsible for the adsorption of the tested copper ions. The presence of the amorphous component enhances the adsorption properties of the materials [63,64].
Microscopic images of crystal size: measured using a VHX-7000 digital microscope (Keyence, Osaka, Japan). Microscopy was performed on different types of fractions: bentonite (1.2–2 mm); modified bentonite (0.3–0.5 mm); fly ash and modified fly ash (0.1–0.5 mm); zeolite and modified zeolite (0.5–1 mm). The fly ash fraction was obtained in its original form (without crushing). Natural bentonite was crushed to bring its grain size closer to that of the other adsorbents used. The zeolite fraction was not adjusted, as the particles already in these sizes approached the size of bentonite.
Microscopic analyses are presented in Figure 5, Figure 6, Figure 7, Figure 8, Figure 9 and Figure 10. Microscopic analysis was used to improve knowledge about the structural properties of sorbents.
Figure 5 and Figure 6 illustrate the changes in the microstructure of fly ash before and after thermal treatment. After modification, the particle size of the fly ash decreased, which was also reflected in an increased BET surface area. The surface became more consistent and compact. Part B of the images highlights the formation of a multilayered heterogeneous surface after fly ash modification.
Figure 7 and Figure 8 show that in bentonite, similar to fly ash, grain fragmentation and disruption of the surface structure occur; however, this does not compromise the specific surface area of the modified bentonite adsorbent.
Figure 9 and Figure 10 indicate that thermal treatment of zeolite did not significantly alter its particle size, nor was the adsorbent’s morphology substantially changed. Part D of the images demonstrates that the zeolite surface was consolidated during modification (losing its porous structure), which is also evidenced by the reduction in specific surface area measured by BET analysis.

2.3. Calculations

2.3.1. Freundlich, Langmuir and Temkin Adsorption Isotherms

To express the dependence of the amount of adsorbed metal ion on its equilibrium concentration in the solution, the Freundlich and Langmuir isotherms were constructed for all adsorbents used. The isotherms were evaluated at five initial concentrations.

2.3.2. Freundlich Adsorption Isotherm

The Freundlich isotherm is usually valid for physical adsorption and for adsorption on heterogeneous surfaces with different active sites. It can be expressed by the formula:
q e = K f · c e   1 n
To verify that the experimental data satisfy this isotherm, the relationship is linearised:
l o g   q e = l o g   K f + 1 n l o g   c e

2.3.3. Langmuir Adsorption Isotherm

The Langmuir isotherm usually applies to chemisorption or electrostatic adsorption, in which only a monomolecular layer is formed on the surface of the adsorbent, and all active centres are equivalent. The Langmuir isotherm is expressed by the formula
q e = q m · b · c e 1 + b · c e  
Or by the linearised formula according to Hanes–Wolf:
c e q e = 1 b · q m + 1 q m · c e  
The Langmuir constant b (L/mg) is used to calculate RL, the dimensionless separation factor, which is given by the equation [65,66]
R L = 1 1 + b · C o  
KR values indicate whether adsorption is unfavourable (RL > 1), linear (RL = 1), favourable (0 < RL < 1) or irreversible (RL = 0) [67].

2.3.4. Temkin Adsorption Isotherm

According to the Temkin isotherm, the heat of sorption should decrease linearly with sorption coverage on the adsorbent due to interactions between the adsorbent and the adsorbate [68,69]. The Temkin isotherm equation is expressed as follows:
q e = R · T b T   ( ln   K T   · c e )  
Equation (8) can be presented in the linear form as follows [70,71]:
q e = R · T b T   ln   K T   +   ( R · T b T   ln   c e )  
where
KT (dm3·g−1) is the equilibrium binding energy, which corresponds to the optimum binding energy; bT (J·mol−1) is a constant, which is related to the heat of adsorption;
R is the ideal gas constant;
T is the absolute temperature.

2.3.5. Adsorption Capacity

From the measured concentrations, the adsorption capacity in the equilibrium state (qe), the amount of metal adsorbed per unit of adsorbent at time t (qt) and the percentage removal efficiency of Cu(II) ions from the solution (Ads. %) were calculated. The adsorption capacity at equilibrium and at time t was calculated according to the equation:
q e = c o c e · V m
where the variable qe represents the adsorption capacity at equilibrium and is expressed in milligrams of adsorbed substance per gram of adsorbent (mg·g−1). It quantifies the amount of heavy metal that binds to one gram of the adsorbent once the system reaches equilibrium, meaning that the concentration in the solution no longer changes significantly. The initial concentration of the substance in the solution, denoted as co and measured in milligrams per cubic decimetre (mg·dm−3), refers to the concentration of the ionic species present before the adsorption process begins, at time zero. The equilibrium concentration ce, also in mg·dm−3, is the concentration of the substance remaining in the solution after adsorption has occurred, and equilibrium has been established, i.e., when the concentration remains stable or changes insignificantly over time. The volume of the solution V, expressed in cubic decimetres (dm3), indicates the total volume of the solution in which adsorption takes place. Lastly, the variable m stands for the mass of the adsorbent added to the solution, measured in grams (g), which is the amount of material used to remove the targeted substance from the solution.
Experiments focused on copper adsorption were carried out with natural, unmodified adsorbents in a closed system with constant stirring of the suspension at room temperature. We monitored the course of sorption depending on the adsorbent used.

2.3.6. Kinetics of Sorption

We used two kinetic models to describe the adsorption kinetics: the pseudo-first-order model [72] and the pseudo-second-order model [73]. The pseudo-first-order model assumes that the rate of metal adsorption is directly proportional to the number of unoccupied binding sites on the surface of the adsorbent. The pseudo-second-order model assumes that the rate of metal adsorption is directly proportional to the square of the remaining binding sites.
Pseudo-first-order equation:
d q t d t = k 1 q e q t
After adjustment, including the separation of variables, integration and substitution of boundary conditions (t = 0 to t; qt = 0 to qe), the equation takes the following form:
ln q e q t =   k 1 ( q e   q t )
and the rate constant k1 is determined from the linear dependence guideline:
ln q e q t =   f ( t )
The rate equation for the pseudo-second-order kinetic model is expressed by the formula
d q t d t = k 2 q e q t 2
After adjustment, including separation of variables, integration and substitution of boundary conditions (t = 0 to t; qt = 0 to qe), the equation takes the following form:
t q t = 1 k 2 . q e 2 + 1 q e t
The dependence t/qt = f (t) is linear, and the values of qe (from the direction) and k2 (from the section on the y-axis) can be calculated from it.

2.3.7. Thermodynamic Quantities

For the adsorption of copper on bentonites, we also calculated the values of thermodynamic variables ΔH0, ΔS0 and ΔG0 according to
l n K C = S 0 R H 0 R . T
G = R . T . l n K c
From the linear dependence ln Kc = f(1/T), we determined the values of ΔH0 (from the direction of the straight line) and ΔS0 (from the section cut by the straight line on the y-axis).

2.4. Statistical Analysis

In addition to descriptive characteristics, selected methods of statistical inference were applied in the statistical processing of data.
Linear and nonlinear regression methods were used to model the investigated dependencies. The statistical significance of the created models was subsequently tested using the F-test. The F-statistic test represents the ratio of two variance estimates—the variance estimate explained by the regression model (MS effect) and the residual variance estimate (MS error). If the MS effect is significantly larger than the MS error, the regression model is suitable for the mathematical definition of the observed dependency. When evaluating other areas of research (adsorption capacities, pH), interval estimates of population means with a 95% confidence level were used. Some of them are illustrated using box plots. All analyses were performed using STATISTICA 14 statistical software. The alpha level of 0.05, traditionally used in similar studies as a decision rule, was applied. The output tables were edited in the Microsoft Excel spreadsheet editor for better clarity [74].

3. Results

3.1. Characteristics of Neutral Mine Drainage

An increased concentration of Cu was measured in the water from the Voznicka heritage tunnel of 17.2 µg·dm−3, which is the limit value for all hardness classes (Table 4). The mining discharges at the monitored site are of a specific neutral nature, which is unusual given their mining origin.
Among all the monitored metals, copper (Cu) stands out, with a concentration of 17.2 µg·dm−3 in the sample, significantly exceeding the limits for water hardness classes 1 and 2 (1.1 µg·dm−3) and for classes 4 and 5 (8.8 µg·dm−3) (Table 4). Other metal elements, including Ag, Ba, Cd, Co, As, Pb, Se, Sb, Sn, Sr, Fe, Zn, Tl and Al, are either within permissible limits or their concentrations do not pose an environmental risk. The sample has a neutral pH (7.6), confirming that the exceedance of the copper limit is not associated with extreme acidity or alkalinity. This result indicates that Cu represents the most significant environmental concern in the analysed NMD water and requires targeted measures for its reduction.

3.2. Characteristics of Natural Sorbents

The composition of the adsorbents used is presented in Table 5.
The chemical composition indicates that SiO2 is the dominant component in both sorbents, with bentonite containing 60.18–73.86% and zeolite 64.18–75.50%. The Al2O3 content ranges from 11.56 to 24.90% in bentonite and from 10.93 to 14.80% in zeolite. Fe2O3 is higher in bentonite (2.15–3.39%) than in zeolite (0.12–2.45%). Other oxides present include CaO, MgO, TiO2, Na2O, K2O and MnO, with concentrations varying between the sorbents, reflecting differences in mineral structure and potential adsorption capacity. Zeolite contains higher amounts of CaO and K2O, whereas bentonite has higher Al2O3 and Fe2O3 contents.
Elemental analysis shows that bentonite contains more aluminium (121,198 mg·kg−1). The contents of toxic elements such as Pb, Cd and As are low in both sorbents, with bentonite having slightly higher levels of lead (12.4 mg·kg−1) and arsenic (6.3 mg·kg−1) compared to zeolite. These data suggest that both sorbents possess suitable chemical properties for adsorption, while the differences in mineral composition may influence their selectivity toward various ions.

3.3. Characteristics of Fly Ash

The ash was produced by burning wood chips (biomass). This is the finest ash captured by the electrofilter before it enters the chimney (Table 6 and Table 7). The properties of the biomass combustion ash used for Cu adsorption are shown in Table 6 and Table 7.
The fly ash used in this study has a very high dry matter content of 99.97%, indicating that it is practically completely dry. It is strongly alkaline, with a pH of 12.11, which may affect its chemical reactivity and ion mobility in the environment.
Table 7 presents the BET surface area of the sorbents used in this study, expressed in m2·g−1. Among the natural sorbents, bentonite has the largest surface area (43.10 m2·g−1), followed by zeolite (33.51 m2·g−1) and fly ash (15.43 m2·g−1). After modification, the surface areas of most sorbents changed: the surface area of modified fly ash slightly increased to 16.45 m2·g−1, while the surface area of modified bentonite decreased to 27.47 m2·g−1. The largest reduction was observed for the modified zeolite, whose surface area dropped to 10.10 m2·g−1. These differences indicate that the modification procedures affect the structure and availability of surface active sites of the sorbents, which can have a significant impact on their adsorption properties.

3.4. Nonlinear Regression Analysis, Nonlinear Adsorption Isotherm Models

To express the dependence of the adsorbed amount of the metal ion on its equilibrium concentration in the solution, we constructed the Freundlich, Langmuir and Temkin isotherms for Cu(II) within the concentration range from 5 to 18 mg·dm−3 for all the adsorbents used (Table 8).
The equilibrium concentrations of Cu2+ at different temperatures were fitted with nonlinear curves of three isothermal models. Temperature had a positive effect on adsorption by fly ash: with increasing temperature, the adsorption capacity rose, reaching a maximum Langmuir capacity (qm) of 36.467 mg·g−1 at 30 °C. For both fly ash and modified fly ash, the Freundlich heterogeneity parameter (1/n) fell within 1 < n < 10, indicating favourable adsorption. The highest n value was observed for modified fly ash at 30 °C, suggesting increased surface heterogeneity after thermal–alkaline treatment.
Among the tested adsorbents, the highest qm was obtained for fly ash (36.467 mg·g−1, 30 °C), followed by modified bentonite (34.109 mg·g−1, 30 °C) and zeolite (22.162 mg·g−1, 30 °C). The tThermal–alkaline treatment of fly ash improved qm at 10 and 20 °C but decreased it at 30 °C, likely due to enhanced desorption. The Temkin model showed a strong increase in sorption heat at 30 °C: the untreated fly ash released 694.579 J·mol−1, while the modified fly ash released 3975.660 J·mol−1, indicating a transition from physical to chemical adsorption.
For bentonite, alkaline treatment substantially increased adsorption at all temperatures. At 30 °C, qm rose from 7.541 to 34.109 mg·g−1, accompanied by higher sorption heat (1523.130 vs. 3289.896 J·mol−1), again pointing to a shift toward chemisorption. The Freundlich constant n ranged from 1.447 to 4.970, with the Langmuir model providing the best fit for untreated bentonite, while the Freundlich model better described modified bentonite.
Zeolite showed lower adsorption than fly ash or bentonite. The Freundlich n values (1.316–2.194) confirmed favourable, though weaker, adsorption. Alkaline treatment enhanced capacity at 20 °C, more than doubling qm, but reducing it at 30 °C. For zeolite, the Temkin model provided the best fit, while modified zeolite at 20 °C was better described by the Freundlich model. Sorption heat varied little with temperature: 998.073–1152.540 J·mol−1 for untreated and 708.488–1046.620 J·mol−1 for modified zeolite, indicating only minor thermal effects.
The separation factor is an important indicator for determining the favourability of adsorption (Table 9).
The RL values indicate whether the adsorption is favourable (0 < RL < 1), unfavourable (RL > 1), linear (RL = 1) or irreversible (RL = 0) [75]. The RL values at concentrations of 5–18 mg dm−3 Cu(II) for the sorbents used indicate that the Langmuir model is favourable, as shown by 0 < RL < 1; however, the Freundlich model also implies favourability, as indicated by 0.1 < 1/n < 1.
The isotherms of adsorbents in relation to temperature before and after treatment are shown in Supporting Information, Figures S1–S27.
The statistical characteristics of the isotherms depending on the adsorption conditions are presented in Table 10.

3.5. Models of Pseudo-First-Order and Pseudo-Second-Order Kinetics

The models of pseudo-first-order and pseudo-second-order kinetics are indicated in Table 11.
The rate constant of PSO k2 is used to describe the rate of adsorption equilibrium. In most observations, the speed increased with rising temperature. The highest achieved values were monitored at the highest initial concentration of the adsorbate (18 mg·dm−3) with modified fly ash at 10 and 20 °C: 54.778 and 63.965 g·mg−1·min−1 (Table 11). Modified fly ash had the highest value of this parameter among all adsorbents tested.
A slight effect of the temperature was observed at higher initial concentrations of adsorbate, i.e., at 12, 16, and 18 mg·dm−3 and unmodified adsorbents. The values of the rate constant for modified zeolite showed a decrease with rising temperature at all initial concentrations except for the highest one of 18 mg·dm−3.
The modification of the adsorbents had a clearly positive effect on the adsorption rate, as described by the PSO model. At all initial concentrations, higher adsorption rates were observed for the modified adsorbents compared with their unmodified forms.

3.6. Thermodynamic Parameters

The evaluation of thermodynamic parameters was performed to better understand the adsorption mechanism. The thermodynamic parameter values at three different adsorption temperatures for the three different adsorbents and their modified forms are summarised in Table 12.
All the results of the free Gibbs energy have a negative value, which indicates a spontaneous adsorption process. The more negative the value of G ° , the more energy-efficient the process. For all adsorbents, an increase in adsorption temperature has a positive effect, with the G ° values becoming more negative. The effect of thermal–alkaline treatment was likewise positive compared with the untreated equivalents of the adsorbents.

3.7. Adsorption Capacities

Monitoring the adsorption capacities is necessary from the point of view of the overall course of adsorption (Table 13). It is important to identify when adsorption begins, or at what time the equilibrium of the adsorption process occurs. Descriptive statistics of the adsorption capacity are presented in Table 13.
At temperatures of 20, 40 and 60 °C, adsorption capacities of approximately 5, 5.6 and 6.4 mg·g−1 were observed at an initial concentration of 3 mg·dm−3. Differences from our study may be caused by the nature of the sorbed medium. In our study, a neutral mine effluent containing other metals was used. In our study, at a similar concentration, higher adsorption capacity was achieved at 10 °C than in the reported study, so the subsequent increase in adsorption capacity as temperatures increased was less significant.
By evaluating the adsorption capacity of fly ash in terms of the influence of temperature, an increase in this indicator due to increasing temperature was found. At the same time, it was observed that the effect of temperature on the increased adsorption capacity of fly ash was more noticeable at higher input concentrations (12, 16 and 18 mg·dm−3) than at lower input concentrations (5 and 9 mg·dm−3). At the initial concentration of 18 mg·dm−3, an increase in adsorption capacity was observed at 10, 20 and 30 °C averaging 6.654, 7.123 and 7.067 mg·g−1.

3.8. Effect of pH on the Adsorption Process

pH is a necessary indicator for adsorption monitoring. For this reason, pH was evaluated as a function of the duration of adsorption. The course of pH for individual adsorbents, depending on time, is presented in Figure 11, Figure 12 and Figure 13. The pH monitoring was carried out at a medium temperature of 20 °C, as there is no assumption of significant differences in pH depending on temperature.
The pH of the drainage was neutral, with a value of 7.35. The pH values of all three monitored adsorbents and their modified forms were measured after moistening with 100 cm3 of original mine water. The values of pH adsorbates with unmodified bentonite and zeolite decreased slightly at time 0 (by 2.7 and 1%). During the following periods of time, they rose at a moderate pace: bentonite by 12.9% and zeolite by 8.7%. Fly ash is alkaline in nature, and its insertion into neutral mine water sharply increased the pH value of the adsorbate to 9.32 (by 26.8%). During the next half-hour measurements, it oscillated around the value of 9.7.
The zeolithization of bentonite increased the pH of the mine water solution compared to unmodified bentonite, with the greatest effect observed after 30 min of stirring: an increase of 2.16 pH units (27.6%). Other half-hour samples showed a decrease in pH differences. In the case of modified zeolite, similar to the modified bentonite, higher pH values were measured, compared to the unmodified form. The largest pH difference was recorded at time 0; in other half-hour samples, we observed a steady-state difference in pH values. The difference was in the range of ⟨0.49–0.89⟩. The modified fly ash behaved differently. Due to the unmodified form of fly ash, the pH values were lower and stabilised at an average difference of 0.74 pH units.

4. Discussion

4.1. Evaluation of the Nonlinear Correlation of Adsorption Isotherms

In any single-component isotherm study, determining the most appropriate model is key for the mathematical description of the respective sorption system. According to the study by Vitek, Masini [76], nonlinear regression analysis should be preferred over linearised equations (Hanes–Wolf, Lineweaver–Burk, Eadie–Hoffsiee, Scatchard equations) for the accurate characterization of adsorption capacities and adsorption affinities. The use of nonlinear procedures for the estimation of adsorption isotherm parameters increases the quality of the obtained results.

4.2. Evaluation of the Parameters of Adsorption Isotherms

The adsorption process forms a layer of adsorbate (metal ions) on the surface of the adsorbents. Adsorption can be reproduced for multiple applications through a desorption method (reverse adsorption, in which adsorbed ions are transported from the adsorbent surface), as adsorption is, under certain conditions, a reversible process. Adsorption onto a solid adsorbent involves three main steps: the transport of the contaminant to the adsorbent surface from the aqueous solution (external diffusion), adsorption onto the solid surface and transport within the adsorbent particle (internal diffusion). In general, electrostatic attraction drives the adsorption of charged contaminants onto differently charged adsorbents, as heavy metals have a strong affinity for hydroxyl (OH) or other functional groups on surfaces.
Adsorption is generally divided into two types: physical adsorption and chemisorption (also described as activated adsorption). Physical adsorption is the adhesion of the adsorbate to the adsorbent surface due to non-specific (i.e., independent of material type) van der Waals forces, whereas chemisorption occurs when chemical bonds create strong attractive forces; that is, chemical adsorption forms ionic or covalent bonds through chemical reactions. Physical adsorption is reversible but less specific, while chemisorption is irreversible but more specific.
Physical adsorption, chemisorption, electrostatic interactions, simple diffusion, intraparticle diffusion, hydrogen bonding, redox interactions, complexation, ion exchange, precipitation and pore adsorption are all possible mechanisms for the adsorption of heavy metal ions [77].
The modification of adsorbents in a NaOH and NaNO3 mixture alters the ligand environment of the adsorbent framework, thereby enhancing the adsorption properties for Cu(II). In bentonites and zeolites, a new amorphous network structure of Si–O–Si and Si–O–Al is formed, which disrupts molecular ordering and increases the adsorption capacity [24].
Adsorption at a given temperature can be quantified using mathematical equations in the form of adsorption isotherms, which relate the amount of adsorbate retained by the adsorbent (qeq_eqe) to the equilibrium concentration in solution (CeC_eCe). Two empirical models most commonly used to describe the adsorption of heavy metals at a specific temperature on various adsorbents are the Freundlich and the Langmuir isotherms. Additionally, the Temkin, Dubinin–Radushkevich, Redlich–Peterson, Koble–Corrigan and Toth isotherms are used to describe the interaction between toxic contaminants and adsorbents.
Adsorption isotherms play a key role in interpreting the mechanism of metal ion adsorption on different adsorbents. These models help elucidate the surface properties of the adsorbents and the intermolecular interactions between the adsorbed molecules and the adsorbent matrix. Isotherm and kinetic models contribute to understanding the adsorption process and depend on various factors, including the structure of the adsorbent and the physical and chemical properties of the dissolved substance.
The Langmuir model is applied to solid–liquid systems, explaining that all sites on the adsorbent surface have an equal probability of being occupied by heavy metals. In contrast, the Freundlich model characterizes a non-ideal process occurring on heterogeneous surfaces, often involving the formation of multiple adsorption layers [77].
Various kinetic models have been developed to describe the adsorption process, such as the pseudo-first-order (PFO) model, the pseudo-second-order (PSO) model, the mixed-order (MO) model, Ritchie’s equation, the Elovich model [78] and phenomenological mass-transfer models. However, there are certain issues when applying these kinetic models. The first is that the most commonly used models, PFO and PSO, are empirical and lack specific physical meaning.
From the empirical adsorption kinetic models, the widely used PFO and PSO models can be derived using Langmuir kinetics. Theoretical analysis and the literature review indicate that the PFO model represents the following conditions:
  • High initial adsorbate concentrations;
  • The initial stage of adsorption;
  • An adsorbent that contains few active sites.
In some cases, the PFO model can describe diffusion-controlled kinetics.
The PSO model, on the other hand, may represent the following conditions:
  • High initial adsorbate concentrations;
  • The final stage of adsorption;
  • An adsorbent surface rich in active sites.
In most cases, the PSO model reveals the adsorption mechanism occurring at the active sites [27].
Melichová et al. [79] utilised adsorption processes for the capture of Cu(II) using different adsorbents, including bentonite and zeolite at 20 °C, with 0.5 g of adsorbent and 100 cm3 of solution. By studying the adsorption of heavy metals (including Cu II), Vengris et al. [80] confirmed the suitability of using the Langmuir isotherm when using modified clays (zeolites). By studying the adsorption of Cu(II), they found the maximum capacity of the monolayer qm in the one-component system to be at the level of 83.3 and in the three-component system at the level of 80.3 mg·g−1. As reported in the study of Melichová et al. [79], neither positive nor negative dependence of parameters qm and n on temperature and thermal–alkaline treatment was confirmed in our case.
The Langmuir isotherms were the most relevant for fly ash. This indicates the formation of a monolayer of adsorbate on the surface of the adsorbate at all three temperatures. As the temperature increased, the nonlinear shape of the models gradually equalised to a linearly occurring dependence. For the alkaline-treated fly ash, the measured sorption data were more in accordance with the Freundlich isotherm model. Adsorption occurs on a heterogeneous surface by a multilayer mechanism. In the modified fly ash, this trend was observed at all three temperatures. By comparing the individual isotherms of fly ash and modified fly ash at particular temperatures, a positive effect of fly ash treatment is clearly observed, especially at 20 °C. At a higher temperature of 30 °C, no significant effect of thermal treatment was observed.
Isotherm models were evaluated using the statistical parameter R2, the coefficient of determination, which assesses the percentage of variability in the concentration of the adsorbed component at equilibrium. A higher value of the coefficient of determination correlates with a better fit of the regression model. The quality of the model is related to the chemical and mineralogical changes in the adsorbents during their thermal–alkaline treatment, which positively influences the adsorption process.

4.3. Evaluation of the Separation Factor

The separation factor RL was calculated and evaluated from the Langmuir constant b, whose value indicates the favourability, unfavorability or linearity of the adsorption process for Cu(II) adsorption according to [67,81]. RL levels, which we found in all cases, confirm the favourability of adsorption when using selected adsorbents. The stated findings are consistent with the findings of the authors presented below. In our study, according to Table 9, it was found out that the untreated original adsorbents—fly ash, bentonite and zeolite—show a decrease in the RL parameter with increasing initial concentration of the sorbed medium. This fact confirms the relevance of the description of the process by the Langmuir isotherm. It quantitatively describes the formation of an adsorbate monolayer on its outer surface, where no further adsorption takes place after it is filled with Cu(II) molecules. The influence of temperature is ambiguous in the case of the original adsorbents. Alkaline-treated adsorbents exhibit more favourable values of the RL parameter (close to zero) compared to their untreated forms. Such results indicate the irreversibility of the adsorption process, which was also confirmed in the evaluation by equilibrium isotherms. Treated adsorbents, especially at a temperature of 30 °C, show the chemical character of an adsorption process. Koyuncu and Kul [82] stated that the RL separation factor showed that Cu(II) ions were favourably adsorbed on natural and acid-treated bentonite. Darmayanti et al. [83] determined RL of Cu(II) when using fly ash at input concentrations of 10 or 20, with levels of 0.16 or 0.09 mg·dm−3. The favourability of Cu(II) adsorption at the indicated concentrations was also confirmed by our study.

4.4. Evaluation of the Models of Pseudo-First-Order and Pseudo-Second-Order Kinetics

Melichová and Ľuptáková [79] demonstrated that the values of the correlation coefficients R2 for a pseudo-first-order model were very low. Therefore, this kinetic model was not suitable for characterising the adsorption process. The pseudo-second-order model was more suitable for experimental data obtained during the experiments, and the correlation coefficients R2 obtained for the pseudo-second-order kinetical model were higher (~1). According to the study, the effect of temperature on the adsorption of copper on bentonites was minimal, due to the calculated qe values, which changed only slightly with temperature change. Harja et al. [84] proved the appropriateness of the description of the adsorption mechanism of synthesised adsorbents (zeolites from ash) using pseudo-second-order kinetics. Vavouraki et al. [85] synthesised zeolites by the fusion of lignite fly ash and NaOH or KOH at 600 °C. They evaluated the effectiveness in terms of decontamination of solutions containing Cu(II) ions. The kinetic data showed that the pseudo-second-order equations described the removal process well. Copper removal was achieved mainly by the combined action of chemisorption and intraparticle diffusion. Based on a study by Joseph et al. [86], a pseudo-second-order equation was the most appropriate for describing the adsorption mechanism of Cu(II) onto fly ash derived from coal combustion. In the study, k2 was found to be 0.012 g·mg−1 min−1. Liu and Zhou [87] achieved a level of k1 of 8.9 dm3·min−1 and k2 of 0.0053 g·mg−1 min−1 in the adsorption of Cu(II) by bentonite. That study demonstrated a favourable description of the adsorption mechanism for pseudo-first- and pseudo-second-order models. Koyuncu and Kul [82] presented kinetic constants using natural bentonite, k1 0.037 dm3·min−1 and k2 0.0876 g·mg−1·min−1. Zou et al. [88] used manganese oxide-coated zeolite for Cu(II) adsorption. The k2 level was reported in the range of 0.625–3.29 g·mg−1·min−1. Adamczuk and Kołodyńska [89] used fly ash for the adsorption of heavy metals. They were able to better describe the adsorption of Cu(II) using the pseudo-second-order model. The constants k2 reached values in the range of 0.03–0.059 g·mg−1·min−1. Apiratikul and Pavasant investigated the use of modified zeolite from coal fly ash for the adsorption of heavy metals, including Cu(II) [90]. A better description of the adsorption process was confirmed for the pseudo-second-order model; the kinetics of k2 reached levels in the range of 0.246–0.538 kg·mol−1·min−1.
The study of adsorption kinetics provides information about the adsorption rate, the efficiency of the adsorbent used, and the mass transfer mechanisms. Knowledge of the kinetics of the process is essential for the design of adsorption systems. From the measured equilibrium data of Cu(II) concentrations, pseudo-first-order (PFO) and pseudo-second-order (PSO) kinetic models were constructed in our study (Table 11). These are the most widely used empirical models, modified by linear regression, which is the most often used method for calculating model parameters, due to its simplicity. The results show that the most suitable model was PSO, except for bentonite and zeolite at the lowest observed adsorbate initial concentration and a temperature of 10 °C, where the PFO model yielded a higher R2 value. According to the study [91], the PFO model is more suitable for describing adsorption at the initial state of adsorption. This mechanism was probably dominant in the case of bentonite and zeolite in the conditions described. PFO also indicates the same rising trend of experimentally obtained values qe with the calculated qe values. At all input concentrations, which were relatively low, the dependence of the k2 parameter on temperature and adsorbent treatment is similar. According to the study [91], the PSO model represents three conditions of the process:
  • Low initial concentration of the adsorbate;
  • Description of the process in the final phase;
  • Adsorbents rich in active sites.
In general, modified materials are rich in active sites. Therefore, adsorption kinetics is dominated by adsorption at the active site. The R2 parameter shows a value of 1 for all modified adsorbents; for unmodified adsorbents, these values are slightly lower but still close to 1. Adsorption at the active sites of the adsorbent involves three steps:
4.
External diffusion of the adsorbate through the liquid film around the solid particle of the adsorbent;
5.
Internal diffusion into the pores of the adsorbent;
6.
Adsorption of the adsorbate at the active sites of the adsorbent.
The difference in concentrations between the solution volume and the adsorbent surface is the driving force of external diffusion, as demonstrated by the increase in k2 with the increasing value of c0 of the adsorbate.
Kinetic models were evaluated using the statistical parameter R2, the coefficient of determination, which assesses the percentage of variability in the concentration of the adsorbed component at equilibrium. A higher value of the coefficient of determination correlates with a better fit of the regression model. The quality of the model is related to the chemical and mineralogical changes of the adsorbents during their thermal–alkaline treatment, which positively influences the adsorption process.

4.5. Evaluation of the Thermodynamic Parameters—Changes in ΔG° ΔH° and ΔS°

Thermodynamic studies provide insights into the minimum kinetic energy required for an adsorbate to bind to an adsorption site. The nature of the adsorption process (spontaneity, randomness, endothermicity or exothermicity) can be evaluated by estimating thermodynamic parameters, such as the change in Gibbs free energy (ΔG°, kJ·mol−1), the change in standard enthalpy (ΔH°, kJ·mol−1) and the change in standard entropy (ΔS°, J·mol−1·K−1).
The thermodynamic properties associated with the removal of metal ions vary due to the composition, structure and surface characteristics of the adsorbent, leading to different affinities for metal ion removal. An increased surface area and porous structure can enhance interactions and subsequent adsorption. Factors such as the presence of competing ions and shifts in pH can modify the charge distribution on both the metal ions and the adsorbent surface, thereby affecting thermodynamic equilibrium. Additionally, different metal ions exhibit distinct thermodynamic affinities due to their unique electronic configurations and charge densities [77].
The most significant difference in G ° values between the thermally treated and original adsorbent was observed in the case of bentonite (3–4 kJ·mol−1), followed by fly ash (1.5–5 kJ·mol−1) and finally, zeolite (2–2.5 kJ·mol−1). Thus, if the adsorption is a spontaneous process, H ° > 0 points to an endothermic process. The released energy, due to thermal–alkaline treatment, was increased by 41.0 kJ·mol−1 in the case of fly ash. In the case of bentonite, it was increased by only 16.7 kJ·mol−1. Alkaline thermal treatment had the least effect on zeolite in terms of enthalpy change. Even untreated zeolite, probably due to the high porosity of the material, had a ∆H° value higher than its alkaline thermally treated form. Entropy is a measure of the disorder of a system. The higher the disorder is, the higher the entropy. It follows from our research that thermal–alkaline treatment of fly ash and bentonite leads to an increased value of S . Zeolite shows the opposite trend, the thermal–alkaline treatment reduced the degree of the system disorder. In their study, Darmayanti et al. [92] adsorbed Cu(II) at 25 °C, 45 °C and 60 °C. The highest value of ΔG0 was −16.493 kJ·mol−1 (60 °C). In our research, the highest value of this parameter was −21.586 kJ·mol−1 in the case of the fly ash (30 °C) and −26.888 kJ·mol−1 in the case of modified fly ash (30 °C). The H ° value in the reference study [92] was 53.834 kJ·mol−1. In our study, fly ash achieved a H ° value of 31.09, while modified fly ash reached 72.1 kJ·mol−1. The highest value of ΔS0 in a study [92] was 211.4 J·mol−1·K−1, while in our study, the values were 173.89 (fly ash) and 324.08 (modified fly ash).
Karapinar and Donat studied the adsorption of Cu(II) and Cd(II) on natural bentonite [93]. The highest ΔG0 value of −29.64 kJ·mol−1 was observed at 60 °C. Our ΔG0 value was lower—for bentonite: −21.400 and for modified bentonite: −25.551 kJ·mol−1 (at 30 °C). In the study [93], a H ° level of 0.0146 kJ·mol−1 was achieved by adsorption. This value is much lower than that obtained in our research on bentonite (32.34 kJ·mol−1) and modified bentonite (49.10 kJ·mol−1). In the reference study [93], the ΔS0 value was 89.06 J·mol−1·K−1, while in our study, the values were 177.61 (bentonite) and 246.63 (modified bentonite). Panayotova studied the kinetics and thermodynamics of the removal of Cu ions from wastewater using zeolite [94]. The highest ΔG0 value, −4.591 kJ·mol−1, was achieved at 50 °C. In our study, we achieved higher ΔG0 values for zeolite (−16.768 kJ·mol−1) and modified zeolite (−18.748 kJ·mol−1) at 30 °C. According to the reference study, the highest H ° level was 27.355 kJ·mol−1 (32 °C). This is lower than the value obtained in our study for zeolite: 37.88 kJ·mol−1 (30 °C); and of modified zeolite: 29.11 kJ·mol−1 (30 °C). The highest level of the ΔS0 parameter was 99.6 J·mol−1·K−1 in the reference study, while we achieved the values of 179.28 J·mol−1·K−1 (zeolite) and 157.23 J·mol−1·K−1 (modified zeolite) at 30 °C.

4.6. Evaluation of Adsorption Capacity

In the study by Al-Harahsheh et al. [95], the positive effect of increasing temperature on Cu(II) adsorption was also confirmed. In this study, adsorption was carried out for 120 min using fly ash from coal combustion with an adsorbent concentration of 6.2 g·dm−3 (more than twice our concentration) with a temperature increase from 25.35 to 45 °C. No significant increase in adsorption capacity due to temperature was observed in the sorption of modified fly ash. This consequence was probably caused by an increase in adsorption capacity due to thermal–alkaline treatment. Thermal–alkaline treatment had a positive effect on the increase in adsorption capacity for Cu(II) sorption at all three temperatures. In the past, this finding was confirmed by a study by Chen et al. [26]. The positive effect of thermal–alkaline treatment in fly ash treatment was observed in this study [26] with the same addition of the adsorbent—2.5 g·dm−3 as in our research, at a temperature of 30 °C. No significant effect of temperature was observed during the adsorption of Cu(II) by bentonite at individual initial concentrations. In some cases, at the end of the sorption, which lasted for 120 min, the adsorption capacity decreased slightly to the average values. Likewise, no significant effect of temperature on adsorption by modified bentonite was observed. Li et al. studied the effect of different conditions (including temperature) on Cu(II) sorption using bentonite [96]. Farsi et al. also found an increase in Cu(II) adsorption by increasing the temperature using activated bentonite as an adsorbent [97]. In our study, a positive effect of thermal–alkaline treatment on the adsorption capacity of bentonite was observed. Thermal–alkaline treatment had a significant effect on the sorption of Cu(II) at 30 °C at an initial concentration of 18 mg·dm−3, when an increase in average adsorption capacity from 6.5 mg·g−1 to 7.2 mg·g−1 was observed. The increase in adsorption by thermal–alkaline treatment of bentonite was also confirmed by Abdallah [98]. In that study, the maximum percentage of removed metals was found to be more than two to three times higher compared with original bentonite, with different removal patterns for individual metals. In contrast to our study, Abdallah studied the adsorption of other PTEs (Co, Mn and Cd). This may have caused the difference in sorption between modified and unmodified bentonite in our study to be less significant. Temperature had a positive effect on the increase in adsorption capacity for the adsorption of Cu(II) by zeolite at the initial concentrations of 5, 9, 16 and 18 mg·dm−3. At the initial concentration of 12 mg·L−1, a reduction in the adsorption capacity of bentonite was observed at the temperature of 20 °C compared to the temperatures of 10 °C and 30 °C. The highest increase in adsorption capacity in the sorption of Cu(II) by zeolite was monitored at the initial concentration of 18 mg·dm−3, when an increase in the adsorption capacity was recorded at the levels of 5.513; 5.967 and 6.729 mg·g−1. Panayotova [94] investigated the removal of Cu(II) by zeolite under different conditions. The study confirmed the positive effect of increasing temperature on the sorption of Cu(II). The findings of that study are consistent with our results. No significant effect of sorption temperature on adsorption capacity was observed in the adsorption of Cu(II) by modified zeolite. A significant positive effect of thermal–alkaline treatment on the adsorption capacity of zeolite was monitored. At a temperature of 10 °C, an increase in the adsorption capacity of zeolite after thermal–alkaline treatment was observed at an initial concentration of 18 mg·dm−3, from 5.513 to 7.200 mg·g−1.

4.7. Evaluation of the Effect of pH on the Adsorption Process

The pH value of the aqueous solution is a significant variable that governs the adsorption of Cu(II) at the adsorbent–solution interface. The results of the study [99] revealed that the maximum adsorption of copper(II) ions onto maple sawdust reached 9.51 mg·g−1 at pH 6.0. The variation in the Langmuir isotherm parameters indicate the fact that the affinity of metal ions onto maple wood sawdust is pH-dependent [99]. The study investigated the adsorption behaviour of Cu(II) on Cs/PVA/PEG beads as a function of solution pH, temperature and contact time. According to the study, the maximum adsorption reached 99.99% for an initial copper ion concentration of 25 mg·dm−3 at pH 5, a temperature of 45 °C, a contact time of 5 h and an adsorbent dose of 1 mg·dm−3 [100]. Eloussaief et al. [101] studied the pH and temperature effects on the adsorption of copper ions on green and red clays from Tejera-Esghira in Medenine, Tunisia. At 20 °C, pH 5.5 and before acid-activation, the amount adsorbed by the green and red clays was 28.3 and 13.2 mg·g−1, respectively, and increased to 40.6 and 26.8 mg·g−1 after acid activation. According to the study [102], under optimal conditions (pH 7, 1 g·dm−3 adsorbent dose, 120 min contact time, 20 mg·dm−3 initial metal concentration and 20 °C), the maximum adsorption capacity of the activated bentonite was 14 ± 0.03 mg·g−1 for Cu2+, exceeding that of the natural bentonite, which had a capacity of 9.2 ± 0.04 mg·g−1.
Our pH measurement results fall within the alkaline range. The mine drainage itself exhibited a pH of 7.35, which, according to study [63], is at the threshold of Cu(II) precipitation into insoluble Cu(II) species. The modification of the adsorbents increased the pH, and copper was present in the form of the soluble salt Cu(OH)2, which had the most positive effect on adsorption onto the modified bentonite at pH 8.1–10.

5. Conclusions

The problem of pollution of surface streams by PTEs from mine effluents or other industrial activities can be solved by an adsorption process, using available adsorbents—zeolite, bentonite and economically and ecologically advantageous fly ash. In this study, the adsorbents zeolite, bentonite and fly ash, as well as the effect of temperature and modifications on the adsorption process, were observed. Alkaline thermal treatment of bentonite and zeolite increased the pH of the mine neutral drainage solutions, as expected due to the alkaline modification process. However, the pH of the fly ash was reduced by the modification. It was found that rising temperature increases the adsorption capacity of unmodified fly ash adsorbents, bentonite and zeolite. With modified adsorbents, the effect of temperature was not significant. However, an essential factor in increasing the adsorption capacity was the alkaline treatment of the adsorbents. The RL separation factor also confirms this finding. The RL values indicate the irreversibility of the adsorption process when using alkaline-treated adsorbents, which is economically advantageous, especially for waste-type materials suitable for adsorption.
The original adsorbents conformed to the Langmuir isotherm model, except for zeolite, which behaved differently and was best suited for the Temkin isotherm model. For all modified adsorbents, the Freundlich model was the most suitable one. The effect of temperature was positive for the original adsorbents, and it was the most evident in the case of fly ash. In the alkaline-treated forms of the adsorbents, the highest observed temperature of −30 °C was unfavourable for the physical course of the process. The chemical essence of the process was manifested there, which was confirmed by an increase in the value of adsorption heat from the Temkin isothermal model.
For the experimental data obtained in the experiments, a pseudo-second-order kinetic model was more suitable, describing the removal process well. Copper removal was achieved mainly by the combined action of chemisorption and intraparticle diffusion. The k2 levels for the modified zeolite showed a decrease with rising temperature in all cases except at the highest initial concentrations. The effect of adsorbent modification was definitely positive on the adsorption rate described by the PSO model.
Thermodynamic parameters confirm the positive effect of increasing process temperature, especially for modified adsorbents. The most energy-efficient process was observed in the case of modified fly ash and modified bentonite. Alkaline treatment has the least effect on zeolite in terms of enthalpy change. Alkaline thermal treatment of bentonite and zeolite increased the pH of the neutral mine drainage solutions, as expected from the alkaline modification process. However, the pH of the fly ash was reduced by the modification. The suitability of using adsorbents to reduce the concentration of Cu(II) in neutral mine effluents was, under the monitored conditions, as follows: at 30 °C, unmodified fly ash > modified bentonite > unmodified zeolite; at 20 and 10 °C, the same trend in the suitability of adsorbents to reduce the concentration of Cu(II) was observed—modified bentonite > modified zeolite > modified fly ash.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma18194552/s1, Isotherms of adsorbents in relation to temperature before and after treatment are presented in Figures S1–S27.

Author Contributions

Conceptualisation, A.Ď.; methodology, V.Š.P.; software, J.S.; validation, M.S. (Michal Sečkár); formal analysis, J.P.; investigation, A.Z.; resources, M.S. (Marián Schwarz); data curation, T.M.; writing—original draft preparation, A.E. and M.S. (Michal Sečkár); writing—review and editing, methodology M.Č.H.; visualisation, J.S.; supervision, D.S. and D.V.; project administration, software, visualisation A.B.; funding acquisition, supervision P.A. and M.S. (Marián Schwarz). All authors have read and agreed to the published version of the manuscript.

Funding

This work was created with the support of the project Comprehensive Research of Determinants for Ensuring Environmental Health (ENVIHEALTH), ITMS 313011T721, supported by the Operational Programme Integrated Infrastructure (OPII), funded by the ERDF.

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/Supplementary Materials; further inquiries can be directed to the corresponding author.

Acknowledgments

The research presented in this paper is an outcome of the project Comprehensive Research of Determinants for Ensuring Environmental Health (ENVIHEALTH), ITMS 313011T721, supported by the Operational Programme Integrated Infrastructure (OPII), funded by the ERDF.

Conflicts of Interest

Author Veronika Štefanka Prepilková was employed by the Fortischem a.s. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Pereira, L.B.; Vicentini, R.; Ottoboni, L.M.M. Changes in the Bacterial Community of Soil from a Neutral Mine Drainage Channel. PLoS ONE 2014, 9, e96605. [Google Scholar] [CrossRef]
  2. Cheng, K.Y.; Acuña, C.R.; Kaksonen, A.H.; Esslemont, G.; Douglas, G.B. Sequential Hydrotalcite Precipitation, Microbial Sulfate Reduction and In Situ Hydrogen Sulfide Removal for Neutral Mine Drainage Treatment. Sci. Total Environ. 2024, 926, 171537. [Google Scholar] [CrossRef] [PubMed]
  3. Jarvis, A.P.; Gandy, C.J.; Webb, J.A. Controls on the Generation and Geochemistry of Neutral Mine Drainage: Evidence from Force Crag Mine, Cumbria, UK. Minerals 2023, 13, 592. [Google Scholar] [CrossRef]
  4. Opitz, J.; Timms, W. Mine Water Discharge Quality—A Review of Classification Frameworks. In Mining Meets Water—Conflicts and Solutions; Drebenstedt, C., Paul, M., Eds.; IMWA (International Mine Water Association): Freiberg, Germany, 2016. [Google Scholar]
  5. Marmier, V.; Plante, B.; Demers, I.; Benzaazoua, M. Neutral Mine Drainage Prediction for Different Waste Rock Lithologies—Case Study of Canadian Malartic. J. Geochem. Explor. 2025, 271, 107685. [Google Scholar] [CrossRef]
  6. Ding, Y.; Shen, S.Z.; Sun, H.; Sun, K.; Liu, F. Synthesis of L-Glutathione-Capped-ZnSe Quantum Dots for the Sensitive and Selective Determination of Copper Ion in Aqueous Solutions. Sens. Actuators B Chem. 2014, 203, 35–43. [Google Scholar] [CrossRef]
  7. Akar, S.T.; Akar, T.; Kaynak, Z.; Anilan, B.; Cabuk, A.; Tabak, Ö.; Demir, T.A.; Gedikbey, T. Removal of Copper(II) Ions from Synthetic Solution and Real Wastewater by the Combined Action of Dried Trametes versicolor Cells and Montmorillonite. Hydrometallurgy 2009, 97, 98–104. [Google Scholar] [CrossRef]
  8. Prepilková, V.; Poništ, J.; Belčáková, I.; Ďuricová, A.; Salva, J.; Schwarz, M.; Samešová, D. Comparison of Copper and Zinc Sorption Depending on Temperature and Sorbents. Pol. J. Environ. Stud. 2024, 33, 2241–2254. [Google Scholar] [CrossRef]
  9. Samešová, D.; Pochyba, A.; Ďuricová, A.; Poništ, J.; Prepilková, V.Š.; Schwarz, M.; Veverková, D.; Salva, J.; Schmidtová, J. Comparative Adsorption of Cu(II), Zn(II), Cd(II), and Mn(II) from Aquatic Solution and Neutral Mine Drainage Using Paper Sludge. Water 2025, 17, 1471. [Google Scholar] [CrossRef]
  10. Ryu, S.; Naidu, G.; Hasan Johir, M.A.; Choi, Y.; Jeong, S.; Vigneswaran, S. Acid Mine Drainage Treatment by Integrated Submerged Membrane Distillation–Sorption System. Chemosphere 2019, 218, 955–965. [Google Scholar] [CrossRef]
  11. Olegario-Sanchez, E.; Pelicano, C.M. Characterization of Philippine Natural Zeolite and Its Application for Heavy Metal Removal from Acid Mine Drainage (AMD). Key Eng. Mater. 2017, 737, 407–411. [Google Scholar] [CrossRef]
  12. Feng, G.; Ma, J.; Zhang, X.; Zhang, Q.; Xiao, Y.; Ma, Q.; Wang, S. Magnetic Natural Composite Fe3O4-Chitosan@bentonite for Removal of Heavy Metals from Acid Mine Drainage. J. Colloid Interface Sci. 2019, 538, 132–141. [Google Scholar] [CrossRef]
  13. Wulandari, E.; Hidayat, A.E.; Moersidik, S.S. Comparison of Copper Adsorption Effectivity in Acid Mine Drainage Using Natural Zeolite and Synthesized Zeolite. IOP Conf. Ser. Earth Environ. Sci. 2020, 473, 012143. [Google Scholar] [CrossRef]
  14. Mohammed, N.H.; Atta, M.; Yaacub, W.Z.W. Remediation of Heavy Metals by Using Industrial Waste by Products in Acid Mine Drainage. Am. J. Eng. Appl. Sci. 2017, 10, 1001–1012. [Google Scholar] [CrossRef]
  15. Zhang, T.; Tu, Z.; Lu, G.; Duan, X.; Yi, X.; Guo, C.; Dang, Z. Removal of Heavy Metals from Acid Mine Drainage Using Chicken Eggshells in Column Mode. J. Environ. Manag. 2017, 188, 1–8. [Google Scholar] [CrossRef] [PubMed]
  16. Fu, F.; Wang, Q. Removal of Heavy Metal Ions from Wastewaters: A Review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef] [PubMed]
  17. Jiang, F.; Yin, S.; Zhang, L.; Peng, J.; Ju, S.; Miller, J.D.; Wang, X. Solvent Extraction of Cu(II) from Sulfate Solutions Containing Zn(II) and Fe(III) Using an Interdigital Micromixer. Hydrometallurgy 2018, 177, 116–122. [Google Scholar] [CrossRef]
  18. Edebali, S.; Pehlivan, E. Evaluation of Chelate and Cation Exchange Resins to Remove Copper Ions. Powder Technol. 2016, 301, 520–525. [Google Scholar] [CrossRef]
  19. Wang, J.; Chen, C. Biosorbents for Heavy Metals Removal and Their Future. Biotechnol. Adv. 2009, 27, 195–226. [Google Scholar] [CrossRef]
  20. Gupta, V.K.; Nayak, A.; Agarwal, S. Bioadsorbents for Remediation of Heavy Metals: Current Status and Their Future Prospects. Environ. Eng. Res. 2015, 20, 1–18. [Google Scholar] [CrossRef]
  21. Ince, M.; İnce, O.K. An Overview of Adsorption Technique for Heavy Metal Removal from Water/Wastewater: A Critical Review. Int. J. Pure Appl. Sci. 2017, 3, 10–19. [Google Scholar] [CrossRef]
  22. Lucaci, A.R.; Bulgariu, D.; Popescu, M.-C.; Bulgariu, L. Adsorption of Cu(II) Ions on Adsorbent Materials Obtained from Marine Red Algae Callithamnion corymbosum Sp. Water 2020, 12, 372. [Google Scholar] [CrossRef]
  23. Savastru, E.; Bulgariu, D.; Zamfir, C.-I.; Bulgariu, L. Application of Saccharomyces cerevisiae in the Biosorption of Co(II), Zn(II) and Cu(II) Ions from Aqueous Media. Water 2022, 14, 976. [Google Scholar] [CrossRef]
  24. Zhao, X.; Zhao, H.; Huang, X.; Wang, L.; Liu, F.; Hu, X.; Li, J.; Zhang, G.; Ji, P. Effect and Mechanisms of Synthesis Conditions on the Cadmium Adsorption Capacity of Modified Fly Ash. Ecotoxicol. Environ. Saf. 2021, 223, 112550. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, X.; Zhao, H.; Zhang, G.; Li, J.; Yang, Y.; Ji, P. Potential of Removing Cd(II) and Pb(II) from Contaminated Water Using a Newly Modified Fly Ash. Chemosphere 2020, 242, 125148. [Google Scholar] [CrossRef]
  26. He, K.; Chen, Y.; Tang, Z.; Hu, Y. Removal of Heavy Metal Ions from Aqueous Solution by Zeolite Synthesized from Fly Ash. Environ. Sci. Pollut. Res. 2016, 23, 2778–2788. [Google Scholar] [CrossRef]
  27. Król, M. Hydrothermal Synthesis of Zeolite Aggregate with Potential Use as a Sorbent of Heavy Metal Cations. J. Mol. Struct. 2019, 1183, 353–359. [Google Scholar] [CrossRef]
  28. Naiya, T.K.; Bhattacharya, A.K.; Mandal, S.; Das, S.K. The Sorption of Lead(II) Ions on Rice Husk Ash. J. Hazard. Mater. 2009, 163, 1254–1264. [Google Scholar] [CrossRef]
  29. Rate, A.W. Sorption of Cadmium(II) and Copper(II) by Soil Humic Acids: Temperature Effects and Sorption Heterogeneity. Chem. Ecol. 2010, 26, 371–383. [Google Scholar] [CrossRef]
  30. Le, A.D.T.; Samri, D.; Rahim, M.; Douzane, O.; Promis, G.; Langlet, T. Effect of Temperature-Dependent Sorption Characteristics on the Hygrothermal Behavior of Hemp Concrete. Energy Procedia 2015, 78, 1449–1454. [Google Scholar] [CrossRef]
  31. Aston, J.E.; Apel, W.A.; Lee, B.D.; Peyton, B.M. Effects of Cell Condition, pH, and Temperature on Lead, Zinc, and Copper Sorption to Acidithiobacillus caldus Strain BC13. J. Hazard. Mater. 2010, 184, 34–41. [Google Scholar] [CrossRef]
  32. Xiao, B.; Dai, Q.; Yu, X.; Yu, P.; Zhai, S.; Liu, R.; Guo, X.; Liu, J.; Chen, H. Effects of Sludge Thermal-Alkaline Pre-treatment on Cationic Red X-GRL Adsorption onto Pyrolysis Biochar of Sewage Sludge. J. Hazard. Mater. 2018, 343, 347–355. [Google Scholar] [CrossRef]
  33. Choi, C.L.; Park, M.; Lee, D.H.; Kim, J.-E.; Park, B.-Y.; Choi, J. Salt-Thermal Zeolitization of Fly Ash. Environ. Sci. Technol. 2001, 35, 2812–2816. [Google Scholar] [CrossRef]
  34. Querol, X.; Moreno, N.; Umaña, J.C.; Alastuey, A.; Hernández, E.; López-Soler, A.; Plana, F. Synthesis of Zeolites from Coal Fly Ash: An Overview. Int. J. Coal Geol. 2002, 50, 413–423. [Google Scholar] [CrossRef]
  35. Števulová, N.; Junák, J. Alkalicky aktivované spojivo na báze popolčeka. Chem. Listy 2014, 108, 620–623. [Google Scholar]
  36. Singh, A.K.; Masto, R.E.; Hazra, B.; Esterle, J.; Singh, P.K. Genesis and Characteristics of Coal and Biomass Ash. In Ash from Coal and Biomass Combustion; Singh, A.K., Masto, R.E., Hazra, B., Esterle, J., Singh, P.K., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 15–36. ISBN 978-3-030-56981-5. [Google Scholar]
  37. Lexa, J.; Štohl, J.; Konečný, V. The Banská Štiavnica Ore District: Relationship between Metallogenetic Processes and the Geological Evolution of a Stratovolcano. Miner. Depos. 1999, 34, 639. [Google Scholar] [CrossRef]
  38. Števko, M.; Sejkora, J.; Malíková, R. New Data on Supergene Minerals from the Banská Štiavnica Deposit (Slovak Republic). Bull. Mineral. Petrol. 2018, 26, 90–101. [Google Scholar]
  39. STN EN ISO 5667-1 (757051); Kvalita Vody. Odber Vzoriek. Časť 1: Pokyny na Návrhy Programov Odberu Vzoriek a Techniky Odberu Vzoriek. NORMSERVIS s.r.o.: Hamry nad Sázavou, Czech Republic, 2023. Available online: https://eshop.normservis.sk/norma/stneniso-5667-1-1.1.2023.html (accessed on 9 July 2025).
  40. STN EN ISO 19458 (757770); Kvalita Vody. Odber Vzoriek na Mikrobiologickú Analýzu. NORMSERVIS s.r.o.: Hamry nad Sázavou, Czech Republic, 2007. Available online: https://eshop.normservis.sk/norma/stneniso-19458-1.2.2007.html (accessed on 16 September 2025).
  41. ISO 11885:2007; Water Quality—Determination of Selected Elements by Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES). ISO: Geneva, Switzerland, 2007. Available online: https://www.iso.org/standard/36250.html (accessed on 16 September 2025).
  42. STN ISO 11465 (465211); Kvalita Pôdy. Stanovenie Obsahu Sušiny a Hmotnostného Obsahu Vody. Gravimetrická Metóda. NORMSERVIS s.r.o.: Hamry nad Sázavou, Czech Republic, 2001. Available online: https://eshop.normservis.sk/norma/stniso-11465-1.3.2001.html (accessed on 30 June 2025).
  43. STN EN ISO 10390 (838445); Zemina, Upravené Bioodpady a Kaly. Stanovenie PH. NORMSERVIS s.r.o.: Hamry nad Sázavou, Czech Republic, 2022. Available online: https://eshop.normservis.sk/norma/stneniso-10390-1.6.2022.html (accessed on 10 December 2024).
  44. ISO 15178:2000; Soil Quality—Determination of Total Sulfur by Dry Combustion. ISO: Geneva, Switzerland, 2000. Available online: https://www.iso.org/standard/26682.html (accessed on 16 September 2025).
  45. eStudio.cz ČSN 75 7440 (757440). Available online: https://www.technicke-normy-csn.cz/csn-75-7440-757440-226378.html (accessed on 16 September 2025).
  46. Skousen, J.; Jacobs, J.A. Stream Characterization for Acid Mine Drainage. In Acid Mine Drainage, Rock Drainage, and Acid Sulfate Soils; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2014; pp. 105–118. ISBN 978-1-118-74919-7. [Google Scholar]
  47. Hillier, S. Use of an Air Brush to Spray Dry Samples for X-Ray Powder Diffraction. Clay Miner. 1999, 34, 127. [Google Scholar] [CrossRef]
  48. Yumpu.com. DIFFRAC.EVA Tutorial. Available online: https://www.yumpu.com/en/document/view/34435692/diffraceva-tutorial (accessed on 16 July 2025).
  49. DIFFRAC.TOPAS. Available online: https://www.bruker.com/en/products-and-solutions/diffractometers-and-x-ray-microscopes/x-ray-diffractometers/diffrac-suite-software/diffrac-topas.html (accessed on 16 July 2025).
  50. Cheary, R.W.; Coelho, A. A Fundamental Parameters Approach to X-Ray Line-Profile Fitting. J. Appl. Crystallogr. 1992, 25, 109–121. [Google Scholar] [CrossRef]
  51. McGinnety, J.A. Redetermination of the Structures of Potassium Sulphate and Potassium Chromate: The Effect of Electrostatic Crystal Forces upon Observed Bond Lengths. Acta Crystallogr. B 1972, 28, 2845–2852. [Google Scholar] [CrossRef]
  52. Hassan, I.; Grundy, H.D. The Crystal Structure of Basic Cancrinite, Ideally Na8Al6Si6O242·3H2O. Can. Mineral. 1991, 29, 377–383. [Google Scholar]
  53. Comodi, P.; Fumagalli, P.; Montagnoli, M.; Zanazzi, P.F. A Single-Crystal Study on the Pressure Behavior of Phlogopite and Petrological Implications. Am. Mineral. 2004, 89, 647–653. [Google Scholar] [CrossRef]
  54. Hughes, J.M.; Cameron, M.; Crowley, K.D. Structural Variations in Natural F, OH, and Cl Apatites. Am. Mineral. 1989, 74, 870–876. [Google Scholar]
  55. Angel, R.J.; Ross, N.L.; Zhao, J.; Sochalski-Kolbus, L.; Krüger, H.; Schmidt, B.C. Structural Controls on the Anisotropy of Tetrahedral Frameworks: The Example of Monoclinic Feldspars. Eur. J. Mineral. 2013, 25, 597–614. [Google Scholar] [CrossRef]
  56. Hazen, R. Effects of Temperature and Pressure on the Cell Dimension and X-Ray Temperature Factors of Periclase. Am. Mineral. 1976, 61, 266–271. [Google Scholar] [CrossRef]
  57. Gualtieri, A.F. Accuracy of XRPD QPA Using the Combined Rietveld–RIR Method. J. Appl. Cryst. 2000, 33, 267–278. [Google Scholar] [CrossRef]
  58. Desgranges, L.; Grebille, D.; Calvarin, G.; Chevrier, G.; Floquet, N.; Niepce, J.-C. Hydrogen Thermal Motion in Calcium Hydroxide: Ca(OH)2. Acta Crystallogr. B 1993, 49, 812–817. [Google Scholar] [CrossRef]
  59. Antao, S.M.; Hassan, I.; Wang, J.; Lee, P.L.; Toby, B.H. State-of-the-Art High-Resolution Powder X-Ray Diffraction (HRPXRD) Illustrated with Rietveld Structure Refinement of Quartz, Sodalite, Tremolite, and Meionite. Can. Mineral. 2008, 46, 1501–1509. [Google Scholar] [CrossRef]
  60. Gournis, D.; Lappas, A.; Karakassides, M.A.; Többens, D.; Moukarika, A. A Neutron Diffraction Study of Alkali Cation Migration in Montmorillonites. Phys. Chem. Miner. 2008, 35, 49–58. [Google Scholar] [CrossRef]
  61. Eberl, D.D. User Guide to RockJock—A Program for Determining Quantitative Mineralogy from X-Ray Diffraction Data; U.S. Geological Survey: Reston, VA, USA, 2003.
  62. Środoń, J.; Drits, V.A.; McCarty, D.K.; Hsieh, J.C.C.; Eberl, D.D. Quantitative X-Ray Diffraction Analysis of Clay-Bearing Rocks from Random Preparations. Clays Clay Miner. 2001, 49, 514–528. [Google Scholar] [CrossRef]
  63. Šuránek, M.; Melichová, Z.; Mirković, M.M.; Ivanović, M.; Pavlović, V.B.; Kljajević, L.; Nenadović, S. The Study of Cu(II) Adsorption onto Synthetically Modified Geopolymers. Sustainability 2023, 15, 2869. [Google Scholar] [CrossRef]
  64. Duxson, P.; Fernández-Jiménez, A.; Provis, J.L.; Lukey, G.C.; Palomo, A.; van Deventer, J.S.J. Geopolymer Technology: The Current State of the Art. J. Mater. Sci. 2007, 42, 2917–2933. [Google Scholar] [CrossRef]
  65. Sari, A.; Tuzen, M.; Citak, D.; Soylak, M. Equilibrium, Kinetic and Thermodynamic Studies of Adsorption of Pb(II) from Aqueous Solution onto Turkish Kaolinite Clay. J. Hazard. Mater. 2007, 149, 283–291. [Google Scholar] [CrossRef] [PubMed]
  66. Han, R.; Zhang, J.; Zou, W.; Shi, J.; Liu, H. Equilibrium Biosorption Isotherm for Lead Ion on Chaff. J. Hazard. Mater. 2005, 125, 266–271. [Google Scholar] [CrossRef] [PubMed]
  67. Melichová, Z.; Ľuptáková, A. Removing Lead from Aqueous Solutions Using Different Low-Cost Abundant Adsorbents. Desal. Water Treat. 2016, 57, 5025–5034. [Google Scholar] [CrossRef]
  68. Obaid, S.A. Langmuir, Freundlich and Temkin Adsorption Isotherms and Kinetics for the Removal of Artichoke Tournefortii Straw from Agricultural Waste. J. Phys. Conf. Ser. 2020, 1664, 012011. [Google Scholar] [CrossRef]
  69. Sampranpiboon, P. Equilibrium Isotherm Models for Adsorption of Zinc(II) Ion from Aqueous Solution on Pulp Waste. Wseas Trans. Environ. Dev. 2014, 10, 35–47. [Google Scholar]
  70. 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]
  71. Javadian, H.; Ghorbani, F.; Tayebi, H.; Asl, S.H. Study of the Adsorption of Cd(II) from Aqueous Solution Using Zeolite-Based Geopolymer, Synthesized from Coal Fly Ash; Kinetic, Isotherm and Thermodynamic Studies. Arab. J. Chem. 2015, 8, 837–849. [Google Scholar] [CrossRef]
  72. Lagergren, S. Zur Theorie der sogenannten Adsorption gelöster Stoffe. Z. Phys. Chem. Kolloid. 1907, 2, 15. [Google Scholar] [CrossRef]
  73. Ho, Y.S.; McKay, G. Pseudo-Second Order Model for Sorption Processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  74. Walpole, R.E. Probability & Statistics for Engineers & Scientists; Pearson: London, UK, 2016. [Google Scholar]
  75. El-Enein, S.A.; Okbah, M.A.; Hussain, S.G.; Soliman, N.F.; Ghounam, H.H. Adsorption of Selected Metals Ions in Solution Using Nano-Bentonite Particles: Isotherms and Kinetics. Environ. Process. 2020, 7, 463–477. [Google Scholar] [CrossRef]
  76. Vitek, R.; Masini, J.C. Nonlinear Regression for Treating Adsorption Isotherm Data to Characterize New Sorbents: Advantages over Linearization Demonstrated with Simulated and Experimental Data. Heliyon 2023, 9, e15128. [Google Scholar] [CrossRef] [PubMed]
  77. Raji, Z.; Karim, A.; Karam, A.; Khalloufi, S. Adsorption of Heavy Metals: Mechanisms, Kinetics, and Applications of Various Adsorbents in Wastewater Remediation—A Review. Waste 2023, 1, 775–805. [Google Scholar] [CrossRef]
  78. Elovich, S.Y.; Larinov, O.G. Theory of Adsorption from Solutions of Non Electrolytes on Solid (I) Equation Adsorption from Solutions and the Analysis of Its Simplest Form, (II) Verification of the Equation of Adsorption Isotherm from Solutions. Izv. Akad. Nauk SSSR Otd. Khimicheskikh Nauk 1962, 2, 209–216. Available online: https://www.scirp.org/reference/referencespapers?referenceid=1418178 (accessed on 16 September 2025).
  79. Melichová, Z.; Ďuricová, A.; Samešová, D.; Nagyová, I. Hodnotenie Rizík Vybraných Kovových Prvkov vo Vodách; Univerzita Mateja Bela, Fakulta prírodných vied: Banská Bystrica, Slovakia, 2017; Volume 1, p. 1. Available online: https://www.library.sk/arl-sldk/sk/detail-sldk_un_epca-0020016 (accessed on 11 August 2025).
  80. Vengris, T.; Binkienė, R.; Sveikauskaitė, A. Nickel, Copper and Zinc Removal from Waste Water by a Modified Clay Sorbent. Appl. Clay Sci. 2001, 18, 183–190. [Google Scholar] [CrossRef]
  81. Melichová, Z.; Handzušová, M. Removal of Cu(II) Ions from Aqueous Solutions by Adsorption onto Natural Bentonites. Solid State Phenom. 2016, 244, 205–212. [Google Scholar] [CrossRef]
  82. Koyuncu, H.; Kul, A.R. An Investigation of Cu(II) Adsorption by Native and Activated Bentonite: Kinetic, Equilibrium and Thermodynamic Study. J. Environ. Chem. Eng. 2014, 2, 1722–1730. [Google Scholar] [CrossRef]
  83. Darmayanti, L.; Notodarmodjo, S.; Damanhuri, E.; Mukti, R.R. Removal of Copper(II) Ions in Aqueous Solutions by Sorption onto Alkali Activated Fly Ash. MATEC Web Conf. 2018, 147, 04007. [Google Scholar] [CrossRef]
  84. Harja, M.; Buema, G.; Sutiman, D.-M.; Munteanu, C.; Bucur, D. Low Cost Adsorbents Obtained from Ash for Copper Removal. Korean J. Chem. Eng. 2012, 29, 1735–1744. [Google Scholar] [CrossRef]
  85. Vavouraki, A.; Bartzas, G.; Komnitsas, K. Synthesis of Zeolites from Greek Fly Ash and Assessment of Their Copper Removal Capacity. Minerals 2020, 10, 844. [Google Scholar] [CrossRef]
  86. Joseph, I.V.; Tosheva, L.; Doyle, A.M. Simultaneous Removal of Cd(II), Co(II), Cu(II), Pb(II), and Zn(II) Ions from Aqueous Solutions via Adsorption on FAU-Type Zeolites Prepared from Coal Fly Ash. J. Environ. Chem. Eng. 2020, 8, 103895. [Google Scholar] [CrossRef]
  87. Liu, Z.; Zhou, S. Adsorption of Copper and Nickel on Na-Bentonite. Process Saf. Environ. Prot. 2010, 88, 62–66. [Google Scholar] [CrossRef]
  88. Zou, W.; Han, R.; Chen, Z.; Jinghua, Z.; Shi, J. Kinetic Study of Adsorption of Cu(II) and Pb(II) from Aqueous Solutions Using Manganese Oxide Coated Zeolite in Batch Mode. Colloids Surf. A Physicochem. Eng. Asp. 2006, 279, 238–246. [Google Scholar] [CrossRef]
  89. Adamczuk, A.; Kołodyńska, D. Equilibrium, Thermodynamic and Kinetic Studies on Removal of Chromium, Copper, Zinc and Arsenic from Aqueous Solutions onto Fly Ash Coated by Chitosan. Chem. Eng. J. 2015, 274, 200–212. [Google Scholar] [CrossRef]
  90. Apiratikul, R.; Pavasant, P. Sorption of Cu2+, Cd2+, and Pb2+ Using Modified Zeolite from Coal Fly Ash. Chem. Eng. J. 2008, 144, 245–258. [Google Scholar] [CrossRef]
  91. Wang, J.; Guo, X. Adsorption Kinetic Models: Physical Meanings, Applications, and Solving Methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  92. Darmayanti, L.; Notodarmodjo, S.; Damanhuri, E. Removal of Copper(II) Ions in Aqueous Solutions by Sorption onto Fly Ash. J. Eng. Technol. Sci. 2017, 49, 546–559. [Google Scholar] [CrossRef]
  93. Karapinar, N.; Donat, R. Adsorption Behaviour of Cu2+ and Cd2+ onto Natural Bentonite. Desalination 2009, 249, 123–129. [Google Scholar] [CrossRef]
  94. Panayotova, M.I. Kinetics and Thermodynamics of Copper Ions Removal from Wastewater by Use of Zeolite. Waste Manag. 2001, 21, 671–676. [Google Scholar] [CrossRef]
  95. Al-Harahsheh, M.S.; Al Zboon, K.; Al-Makhadmeh, L.; Hararah, M.; Mahasneh, M. Fly Ash Based Geopolymer for Heavy Metal Removal: A Case Study on Copper Removal. J. Environ. Chem. Eng. 2015, 3, 1669–1677. [Google Scholar] [CrossRef]
  96. Li, J.; Hu, J.; Sheng, G.; Zhao, G.; Huang, Q. Effect of pH, Ionic Strength, Foreign Ions and Temperature on the Adsorption of Cu(II) from Aqueous Solution to GMZ Bentonite. Colloids Surf. A Physicochem. Eng. Asp. 2009, 349, 195–201. [Google Scholar] [CrossRef]
  97. Farsi, A.; Javid, N.; Malakootian, M. Investigation of Adsorption Efficiency of Cu2+ and Zn2+ by Red Soil and Activated Bentonite from Acid Copper Mine Drainage. Desal. Water Treat. 2019, 144, 172–184. [Google Scholar] [CrossRef]
  98. Abdallah, S.M. Towards a Safer Environment: Zeolitization of Bentonite and Its Potentialities for Removing Heavy Metals from Wastewater. J. Soil Sci. Agric. Eng. 2010, 1, 533–553. [Google Scholar] [CrossRef]
  99. Rahman, M.S.; Islam, M.R. Effects of pH on Isotherms Modeling for Cu(II) Ions Adsorption Using Maple Wood Sawdust. Chem. Eng. J. 2009, 149, 273–280. [Google Scholar] [CrossRef]
  100. Trikkaliotis, D.G.; Christoforidis, A.K.; Mitropoulos, A.C.; Kyzas, G.Z. Adsorption of Copper Ions onto Chitosan/Poly(Vinyl Alcohol) Beads Functionalized with Poly(Ethylene Glycol). Carbohydr. Polym. 2020, 234, 115890. [Google Scholar] [CrossRef]
  101. Eloussaief, M.; Jarraya, I.; Benzina, M. Adsorption of Copper Ions on Two Clays from Tunisia: pH and Temperature Effects. Appl. Clay Sci. 2009, 46, 409–413. [Google Scholar] [CrossRef]
  102. Eleraky, M.I.; Razek, T.M.A.; Hasani, I.W.; Fahim, Y.A. Adsorptive Removal of Lead, Copper, and Nickel Using Natural and Activated Egyptian Calcium Bentonite Clay. Sci. Rep. 2025, 15, 13050. [Google Scholar] [CrossRef]
Figure 1. The locations of the sampling sites for the adsorbents and mine discharges.
Figure 1. The locations of the sampling sites for the adsorbents and mine discharges.
Materials 18 04552 g001
Figure 2. XRPD patterns of initial (Be1) and treated (Be2) bentonite (Kopernica deposit). *—internal standard: corundum. The dashed line schematically depicts the amorphous component.
Figure 2. XRPD patterns of initial (Be1) and treated (Be2) bentonite (Kopernica deposit). *—internal standard: corundum. The dashed line schematically depicts the amorphous component.
Materials 18 04552 g002
Figure 3. XRPD patterns of initial (Ze1) and treated (Ze2) zeolite (Nižný Hrabovec deposit). *—internal standard: corundum. The dashed line schematically depicts the amorphous component.
Figure 3. XRPD patterns of initial (Ze1) and treated (Ze2) zeolite (Nižný Hrabovec deposit). *—internal standard: corundum. The dashed line schematically depicts the amorphous component.
Materials 18 04552 g003
Figure 4. XRPD patterns of initial (FA1) and treated (FA2) fly ash. *—internal standard: corundum.
Figure 4. XRPD patterns of initial (FA1) and treated (FA2) fly ash. *—internal standard: corundum.
Materials 18 04552 g004
Figure 5. Fly ash, before treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 245 μm; (C) colour resolution of surface irregularities; (D) microstructure at 100-fold magnification.
Figure 5. Fly ash, before treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 245 μm; (C) colour resolution of surface irregularities; (D) microstructure at 100-fold magnification.
Materials 18 04552 g005
Figure 6. Fly ash, after treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D with a max depth of 220 μm; (C) colour resolution of surface irregularities; (D) microstructure at 150-fold magnification.
Figure 6. Fly ash, after treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D with a max depth of 220 μm; (C) colour resolution of surface irregularities; (D) microstructure at 150-fold magnification.
Materials 18 04552 g006
Figure 7. Bentonite, before treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D, with a max depth of 570 μm; (C) colour resolution of surface irregularities; (D) microstructure at 200-fold magnification.
Figure 7. Bentonite, before treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D, with a max depth of 570 μm; (C) colour resolution of surface irregularities; (D) microstructure at 200-fold magnification.
Materials 18 04552 g007aMaterials 18 04552 g007b
Figure 8. Bentonite, after treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D, with a max depth of 225 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Figure 8. Bentonite, after treatment: (A) grain size of the sorbent at 20-fold magnification; (B) surface structure in 3D, with a max depth of 225 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Materials 18 04552 g008
Figure 9. Zeolite, before treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 249 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Figure 9. Zeolite, before treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 249 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Materials 18 04552 g009
Figure 10. Zeolite, after treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 183 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Figure 10. Zeolite, after treatment: (A) sorbent grain size at 20-fold magnification; (B) surface structure in 3D, with a max depth of 183 μm; (C) colour resolution of surface irregularities; (D) microstructure at 500-fold magnification.
Materials 18 04552 g010
Figure 11. Course of pH depending on the temperature of fly ash.
Figure 11. Course of pH depending on the temperature of fly ash.
Materials 18 04552 g011
Figure 12. Course of pH depending on the temperature of bentonite.
Figure 12. Course of pH depending on the temperature of bentonite.
Materials 18 04552 g012
Figure 13. Course of pH depending on the temperature of zeolite.
Figure 13. Course of pH depending on the temperature of zeolite.
Materials 18 04552 g013
Table 1. List of test methods for the analysis of neutral mine drainage.
Table 1. List of test methods for the analysis of neutral mine drainage.
Parameter
/Metal
Method
PrincipleMarking
CuAES-ICPSWP 6 (ISO 11885) [41]
MnAES-ICPSWP 6 (ISO 11885)
ZnAES-ICPSWP 6 (ISO 11885)
FeAES-ICPSWP 6 (ISO 11885)
PbAES-ICPSWP 6 (ISO 11885)
CdAES-ICPSWP 6 (ISO 11885)
AlAES-ICPSWP 6 (ISO 11885)
Explanations: AES-ICP—Atomic emission spectrometry with the inductively coupled plasma.
Table 2. List of test methods for the analysis of fly ash.
Table 2. List of test methods for the analysis of fly ash.
Parameter
/Metal
Method
PrincipleMarking
dry mattergravimetricallySTN ISO 11465 < (SWP 2) [42]
pH/H2OpotentiometricallyISO 10390 (SWP 8) [43]
SEA-TCDSWP 4 (ISO 15178) [44]
AlAES-ICPSWP 6 (ISO 11885)
BAES-ICPSWP 6 (ISO 11885)
CrAES-ICPSWP 6 (ISO 11885)
MnAES-ICPSWP 6 (ISO 11885)
CuAES-ICPSWP 6 (ISO 11885)
NaAES-ICPSWP 6 (ISO 11885)
ZnAES-ICPSWP 6 (ISO 11885)
FeAES-ICPSWP 6 (ISO 11885)
CdAES-ICPSWP 6 (ISO 11885)
PbAES-ICPSWP 6 (ISO 11885)
NiAES-ICPSWP 6 (ISO 11885)
HgAAS AMASWP 15 (CSI) 75 7440 l [45]
MgAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
CaAES-ICPSWP 6 (ISO 11885). SWP 19 Mchlich III
KAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
PAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
CuAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
FeAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
MnAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
ZnAES-ICPSWP 6 (ISO 11885). SWP 19 Mehlich III
Explanations: EA-TCD—elemental analysis with thermal conductivity detection; AES-ICP—Atomic emission spectrometry with the inductively coupled plasma; SWP—standard work procedure; AAS-AMA—atomic absorption spectrometry–advanced mercury analyser.
Table 3. Normalised results of quantitative X-ray powder diffraction analysis (weight %).
Table 3. Normalised results of quantitative X-ray powder diffraction analysis (weight %).
Adsorbent SampleFly Ash, Before TreatmentFly Ash, After TreatmentBentonite, Before TreatmentBentonite, After TreatmentZeolite, Before TreatmentZeolite, After Treatment
MethodRieRieRJRieRJRJ
Mineral[wt.%]
quartz222463--
K-feldspar-7115<13
plagioclase535--411
biotite--22<13
clinoptilolite/heulandite----7629
opal-CT--6-15-
arcanite17-----
cancrinite---16--
calcite2934-<1-4
portlandite16-----
hydroxylapatite5-----
periclase6-----
kaolinite--2---
smectite--7313<1
amorphous content---72-50
Sum100100100100100100
RJ—RockJock [61]; Rie—Rietveld—[61].
Table 4. Concentrations of metal elements in the NMD sample and their physicochemical characteristics.
Table 4. Concentrations of metal elements in the NMD sample and their physicochemical characteristics.
Metal Content [µg·dm−3]AgBaCdCoAs
LV5 0.08 for the 1st hardness class up to 0.25
5th hardness class
50 7.5
MV<0.0512.6621.4311.141
Metal Content [µg·dm−3]PbSeSbSnSr
LV7.2 20
MV1.71.30.610.391814.39
Metal Content [µg·dm−3]Cu *FeZnTlAl
LV1.1 for the 1st and 2nd hardness class and 8.8 for 4th and 5th hardness classtotal Fe
2 mg·dm−3
7.8 for the 1st and 2nd hardness class and 52 for 4th and 5th hardness class 200
MV17.2<104821.60.48141
pH [-]7.6
* Cu exceeded the limit values; LV—limit value according to the Regulation of the Government of the Slovak Republic No. 269/2010 Coll. MV—measured value.
Table 5. Chemical composition and elemental analysis of bentonite and zeolite obtained from the supplier of a specific bearing.
Table 5. Chemical composition and elemental analysis of bentonite and zeolite obtained from the supplier of a specific bearing.
Chemical Composition [%]
AdsorbentSiO2Al2O3Fe2O3CaOMgOTiO2Na2OK2OMnO
Bentonite60.18–73.811.56–24.902.15–3.391.23–2.171.09–3.290.11–0.200.29–0.910.42–1.180.02–0.06
Zeolite64.18–75.5010.93–14.800.12–2.451.43–11.680.29–1.43-0.10–2.971.24–4.24
Elemental Analysis [mg·kg−1]
AdsorbentCopperAntimonyZincIronLeadCadmiumArsenicAluminium
Bentonite7.4023.464.423.71112.4-6.3121.198
Zeolite3.120.15-8.0448.390.0450.9166.685
Table 6. Characteristics of fly ash.
Table 6. Characteristics of fly ash.
Parameter
Dry matterpH/H2OSAlBCr
%-mg·kg−1mg·kg−1mg·kg−1mg·kg−1
99.9712.1119,40017,44829514.5
MnCuNaZnFeCd
mg·kg−1mg·kg−1mg·kg−1mg·kg−1mg·kg−1mg·kg−1
47711172099164911,25311.8
PbNiHgMgCaK
mg·kg−1mg·kg−1μg·kg−1mg·kg−1mg·kg−1mg·kg−1
43.015.63.040.56162,41988,632
Table 7. BET characteristics of the sorbents used [m2·g−1].
Table 7. BET characteristics of the sorbents used [m2·g−1].
AdsorbentBET Surface Area (m2·g−1)
fly ash15.433
bentonite43.098
zeolite33.506
modified fly ash16.449
modified bentonite27.472
modified zeolite10.098
Table 8. Langmuir, Freundlich and Temkin constants for the adsorption of Cu(II) on adsorbents at different temperatures—10, 20 and 30 °C.
Table 8. Langmuir, Freundlich and Temkin constants for the adsorption of Cu(II) on adsorbents at different temperatures—10, 20 and 30 °C.
AdsorbentTemperatureFreundlich ConstantsLangmuir ConstantsTemkin Constants
t
[°C]
Kf
[mg·g−1]
nqm
[mg·g−1]
b
[dm3·mg−1]
A
[dm3·g−1]
bt
[J·mol−1]
Fly ash104.6302.0917.8191.61117.5461451.230
206.5823.6886.5499.828103.5711743.130
3019.0591.13436.4670.73521.089694.579
Modified fly ash109.3682.0578.4845.78372.4221401.750
2010.1271.93410.0874.23646.2141140.310
308.1624.5647.68515.94918 7783975.660
Bentonite105.0614.7665.9375.831136.0822350.720
205.5302.8107.2223.18424.2691390.920
305.5123.1997.5413.56731.8751523.130
Modified bentonite1014.2331.44715.9622.14223.974771.812
2018.6811.79823.0832.923130.581150.080
307.7324.97034.1091.3199483.33289.896
Zeolite102.9912.0838.8390.4822.951998.073
202.0961.4828.5400.4782.6511152.540
305.6651.31622.1620.34710.3141020.170
Modified zeolite1016.5301.59512.5884.52029.748708.488
209.4471.29919.4050.81213.139806.982
308.0242.1949.6103.31125.9021046.620
Table 9. RL values calculated from the Langmuir constants.
Table 9. RL values calculated from the Langmuir constants.
Temperature (°C)Adsorbent
c0 [mg·dm−3]
Fly ash BentoniteZeolite
591216185912161859121618
100.0830.0480.0370.0280.0250.0860.0490.0380.0280.0250.4290.2940.2380.1900.173
200.0520.0300.0230.0170.0150.0530.0300.0230.0170.0150.3790.2530.2020.1600.145
300.3030.1950.1530.1200.1080.1060.0620.0470.0360.0320.2050.1250.0970.0750.067
Modified fly ashModified bentoniteModified zeolite
591216185912161859121618
100.0240.0140.0100.0080.0070.1180.0690.0530.0400.0360.0030.0020.0010.0010.001
200.0400.0220.0170.0130.0116.18 × 10−43.44 × 10−42.58 × 10−41.93 × 10−41.72 × 10−41.80 × 10−11.09 × 10−18.37 × 10−26.41 × 10−25.74 × 10−2
300.0050.0030.0020.0010.001*****0.1050.0610.0470.0350.032
* Absolute removal of Cu(II) from the solution was achieved after the first half an hour of adsorption. Therefore, it was not possible to determine the RL.
Table 10. Statistical data of Langmuir, Freundlich and Temkin constants for the adsorption of Cu(II) on adsorbents at different temperatures—10, 20 and 30 °C.
Table 10. Statistical data of Langmuir, Freundlich and Temkin constants for the adsorption of Cu(II) on adsorbents at different temperatures—10, 20 and 30 °C.
AdsorbentIsothermF-Testp-Level *
Bentonite 10 °CFreundlich89.460.000
Langmuir91.090.000
Temkin89.910.000
Bentonite 20 °CFreundlich107.390.000
Langmuir112.480.000
Temkin110.300.000
Bentonite 30 °CFreundlich276.680.000
Langmuir389.020.000
Temkin335.480.000
Zeolite 10 °CFreundlich207.760.000
Langmuir242.030.000
Temkin250.500.000
Zeolite 20 °CFreundlich175.210.000
Langmuir180.630.000
Temkin184.420.000
Zeolite 30 °CFreundlich1439.300.000
Langmuir1217.710.000
Temkin693.330.000
Fly ash 10 °CFreundlich311.940.000
Langmuir318.060.000
Temkin309.290.000
Fly ash 20 °CFreundlich107.230.000
Langmuir113.410.000
Temkin109.950.000
Fly ash 30 °CFreundlich1220.390.000
Langmuir1261.660.000
Temkin1087.930.000
Modified bentonite 10 °CFreundlich469.780.000
Langmuir485.050.000
Temkin494.010.000
Modified bentonite 20 °CFreundlich19.040.000
Langmuir500.270.000
Temkin610.230.000
Modified bentonite 30 °CFreundlich92.310.000
Langmuir408.560.000
Temkin320.790.000
Modified zeolite 10 °CFreundlich126.540.000
Langmuir90.400.000
Temkin70.450.000
Modified zeolite 20 °CFreundlich1777.370.000
Langmuir1597.50.000
Temkin1371.740.000
Modified zeolite 30 °CFreundlich167.970.000
Langmuir167.560.000
Temkin169.880.000
Modified fly ash 10 °CFreundlich387.570.000
Langmuir379.540.000
Temkin365.310.000
Modified fly ash 20 °CFreundlich1328.070.000
Langmuir1550.380.000
Temkin1243.120.000
Modified fly ash 30 °CFreundlich286.930.000
Langmuir185.540.000
Temkin357.030.000
* All p-level values were below 0.000.
Table 11. Pseudo-first-order and pseudo-second-order models.
Table 11. Pseudo-first-order and pseudo-second-order models.
AdsorbentTemperatureqe Experim.
[mg·g−1]
Pseudo-First-OrderPseudo-Second-Order
[°C]k1 [dm3·min−1]qe [mg·g−1]R2Equationk2
[g·mg−1 min−1]
qe [mg·g−1]R2Equation
5 [mg·dm−3]
Fly ash101.9340.01091.7450.579y = 0.0109x + 0.55650.9030.3890.998y = 2.5728x + 7.3327
201.9510.01081.8300.534y = 0.0108x + 0.60452.5790.3931y = 2.546x + 2.513
301.9680.01081.8370.532y = 0.0108x + 0.60842.4080.3961y = 2.5264x + 2.6508
Modified fly ash101.9840.01071.8760.512y = 0.0107x + 0.62896.0690.3971y = 2.5191x + 1.0457
201.9810.01071.8790.509y = 0.0107x + 0.63097.7740.3961y = 2.5243x + 0.8197
302.0000.01071.8990.502y = 0.0107x + 0.641324.3000.4001y = 2.4995x + 0.2571
Bentonite101.8600.01331.2140.886y = 0.0133x + 0.19390.0890.4390.824y = 2.2801x + 58.422
201.8710.01111.6530.634y = 0.0111x + 0.50270.5690.3840.994y = 2.6014x + 11.893
301.8320.01061.7550.559y = 0.0106x + 0.56251.7070.3901y = 2.5664x + 3.8582
Modified bentonite101.9670.01061.8810.505y = 0.0106x + 0.63213.1840.3941y = 2.5412x + 0.4898
202.0000.62951.8770.516y = 0.0108x + 0.62955.3420.1611y = 2.4949x + 1.1651
302.0000.01081.8940.506y = 0.0108x + 0.638515.7770.4011y = 2.4967x + 0.3951
Zeolite101.5780.01081.1820.922y = 0.0108x + 0.16740.1400.3300.839y = 3.032x + 65.646
201.6560.00961.4840.673y = 0.0096x + 0.39480.4160.3160.953y = 3.1679x + 24.097
301.9260.01151.5180.727y = 0.0115x + 0.41770.2840.3900.974y = 2.5667x + 23.22
Modified zeolite101.9790.01071.8800.5068y = 0.0107x + 0.63158.1610.3951y = 2.5314x + 0.7852
201.9420.01061.8660.509y = 0.0106x + 0.62368.3940.3881y = 2.5741x + 0.7894
301.9380.01081.8060.5441y = 0.0108x + 0.59111.6760.3901y = 2.5627x + 3.9187
9 [mg·dm−3]
Fly ash103.3310.01452.1860.552y = 0.0145x + 0.78211.0430.3730.998y = 2.6843x + 6.9117
203.4760.01452.3510.509y = 0.0145x + 0.85497.8480.3871y = 2.5811x + 0.8489
303.5130.01472.2970.531y = 0.0147x + 0.83151.8450.3930.999y = 2.5437x + 3.5069
Modified fly ash103.5300.01452.3770.504y = 0.0145x + 0.86614.6730.3921y = 2.5498x + 0.4431
203.5390.01452.3800.503y = 0.0145x + 0.867314.3250.3931y = 2.544x + 0.4518
303.5620.01462.3900.502y = 0.0146x + 0.871517.8200.3961y = 2.5258x + 0.358
Bentonite103.2440.01461.8790.659y = 0.0146x + 0.63050.1270.2980.961y = 2.8046x + 26.378
203.2280.01472.0730.593y = 0.0147x + 0.72910.6630.3680.995y = 2.7159x + 11.127
303.4780.01522.1300.592y = 0.0152x + 0.75613.0920.3851y = 2.5943x + 2.177
Modified bentonite103.5180.01452.3810.502y = 0.0145x + 0.867537.3020.3911y = 2.5579x + 0.1754
203.6000.01482.3360.524y = 0.0148x + 0.84845.3220.1611y = 2.49x + 2.5051
303.6000.01472.3990.503y = 0.0147x + 0.875121.8110.4001y = 2.4976x + 0.286
Zeolite103.1130.01631.4850.849y = 0.0163x + 0.39570.1010.3890.841y = 2.5689x + 65.018
203.0020.01451.9290.635y = 0.0145x + 0.6570.4630.3480.987y = 2.8755x + 17.859
303.3760.01462.1000.583y = 0.0146x + 0.74180.5960.3730.991y = 2.68x + 12.044
Modified zeolite103.5990.01472.3850.507y = 0.0147x + 0.8697.0850.4001y = 2.4989x + 0.8814
203.5100.01462.3230.519y = 0.0146x + 0.84292.4720.3901y = 2.5611x + 2.6529
303.4780.01452.3250.515y = 0.0145x + 0.84383.1400.3861y = 2.5929x + 2.1411
12 [mg·dm−3]
Fly ash104.1810.01841.7500.772y = 0.0184x + 0.55970.1170.4000.865y = 2.5027x + 53.539
204.6160.01662.5450.531y = 0.0166x + 0.93425.7280.4011y = 2.4952x + 1.0869
304.7070.01692.5200.545y = 0.0169x + 0.92421.1610.3970.999y = 2.5212x + 5.4732
Modified fly ash104.6350.01632.6750.497y = 0.0163x + 0.983914.8580.3861y = 2.5906x − 0.4517
204.7160.01642.6750.501y = 0.0164x + 0.9845.7280.4011y = 2.4952x + 1.0869
304.7260.01652.6640.504y = 0.0165x + 0.97987.8410.3931y = 2.5438x + 0.8253
Bentonite104.5970.01772.1910.643y = 0.0177x + 0.78440.3030.4050.979y = 2.4698x + 20.162
204.6740.01792.1500.661y = 0.0179x + 0.76545.7280.4011y = 2.4952x + 1.0869
304.2000.01622.3940.553y = 0.0162x + 0.87315.5270.3911y = 2.5566x + 1.1827
Modified bentonite104.7430.01652.6710.504y = 0.0165x + 0.98239.6910.3951y = 2.5311x + 0.6611
204.8000.01662.6520.511y = 0.0166x + 0.97525.7280.4011y = 2.4952x + 1.0869
304.8000.01662.6720.508y = 0.0166x + 0.9835.7280.4011y = 2.4952x + 1.0869
Zeolite104.1110.01642.2680.588y = 0.0164x + 0.81890.6820.3520.994y = 2.8409x + 11.84
203.5850.01532.0500.621y = 0.0153x + 0.71765.7280.4011y = 2.4952x + 1.0869
304.4530.01662.4110.561y = 0.0166x + 0.88010.8790.3750.997y = 2.6643x + 8.0793
Modified zeolite104.7210.01662.6590.509y = 0.0166x + 0.97788.5400.3961y = 2.5251x + 0.7466
204.6130.01642.6160.511y = 0.0164x + 0.96155.7280.4011y = 2.4952x + 1.0869
304.6850.01652.6420.509y = 0.0165x + 0.97175.5420.3911y = 2.5559x + 1.1788
16 [mg·dm−3]
Fly ash105.8980.01922.3550.642y = 0.0192x + 0.85640.2530.3830.964y = 2.6128x + 27.027
206.2050.01872.8000.542y = 0.0187x + 1.02950.9440.3920.997y = 2.5529x + 6.9036
306.3020.01862.9010.523y = 0.0186x + 1.06511.7950.3960.999y = 2.5253x + 3.5534
Modified fly ash106.2550.01832.9820.504y = 0.0183x + 1.09259.8810.3901y = 2.5639x + 0.6653
206.2100.01823.0110.497y = 0.0182x + 1.102310.9710.3871y = 2.581x − 0.6072
306.2960.01843.0050.501y = 0.0184x + 1.100229.1170.3941y = 2.5407x + 0.2217
Bentonite105.9910.01892.5690.591y = 0.0189x + 0.94370.4380.3860.988y = 2.5883x + 15.311
206.1050.01942.4120.635y = 0.0194x + 0.88040.2490.4000.959y = 2.4998x + 25.106
306.2350.01872.8610.532y = 0.0187x + 1.05123.2470.3941y = 2.5391x + 1.9858
Modified bentonite106.2990.01852.9800.507y = 0.0185x + 1.0927.3200.3951y = 2.5344x + 0.8775
206.3820.01852.9880.508y = 0.0185x + 1.09454.8310.4001y = 2.5015x + 1.2952
306.4000.01862.9880.509y = 0.0186x + 1.09464.9730.4011y = 2.4921x + 1.2489
Zeolite105.7030.01872.2420.659y = 0.0187x + 0.80730.2200.3610.937y = 2.7723x + 34.893
205.7040.01892.2680.655y = 0.0189x + 0.81870.2180.3690.930y = 2.7077x + 33.705
306.0000.01842.7850.536y = 0.0184x + 1.02441.1270.3770.998y = 2.65x + 6.2315
Modified zeolite106.3740.01853.0040.505y = 0.0185x + 1.10019.8290.3991y = 2.5081x + 0.64
206.1640.01832.9530.507y = 0.0183x + 1.08297.0390.3861y = 2.589x + 0.9523
306.3190.01852.9560.511y = 0.0185x + 1.0843.2700.3941y = 2.5386x + 1.9709
18 [mg·dm−3]
Fly ash106.6540.01932.6240.589y = 0.0193x + 0.96470.3910.3670.978y = 2.7266x + 19.033
207.1230.01963.0000.533y = 0.0196x + 1.09851.3610.3990.999y = 2.5055x + 4.6108
307.0670.01923.1040.509y = 0.0192x + 1.13284.2180.3931y = 2.5451x + 1.5356
Modified fly ash107.0550.01913.1490.500y = 0.0191x + 1.14763.9650.3911y = 2.5543x + 0.102
207.0450.01913.1510.499y = 0.0191x + 1.147854.7780.3911y = 2.5585x + 0.1195
307.1440.01933.1270.507y = 0.0193x + 1.14015.6270.3971y = 2.5182x + 1.127
Bentonite107.0260.02022.6130.614y = 0.0202x + 0.96050.2850.4070.970y = 2.4587x + 21.188
206.8970.02072.4110.664y = 0.0207x + 0.87990.1930.4110.951y = 2.4345x + 30.643
306.4820.01922.8230.552y = 0.0192x + 1.037829.8360.3841y = 2.6019x + 0.2269
Modified bentonite107.0510.01953.0040.530y = 0.0195x + 1.10011.5600.3951y = 2.5321x + 4.1095
207.2000.01943.1060.513y = 0.0194x + 1.13343.1570.1601y = 2.4964x + 1.9742
307.2000.01933.1620.503y = 0.0193x + 1.151313.0320.4001y = 2.4979x + 0.4788
Zeolite105.5130.0192.2730.657y = 0.019x + 0.82120.2970.3250.969y = 3.079x + 31.961
205.9670.01862.5560.586y = 0.0186x + 0.93830.4670.3340.982y = 2.9962x + 19.214
306.7290.01892.9830.518y = 0.0189x + 1.0931.8700.3710.998y = 2.6971x + 3.8898
Modified zeolite107.2000.01933.1600.503y = 0.0193x + 1.150711.7970.4001y = 2.4976x + 0.5288
206.9550.01923.0870.5096y = 0.0192x + 1.12725.0630.3881y = 2.5794x + 1.3141
306.9180.0193.1190.5017y = 0.019x + 1.137527.1240.3851y = 2.6004x + 0.2493
Table 12. Values of the thermodynamic parameters ΔG0, ΔH0 and ΔS0.
Table 12. Values of the thermodynamic parameters ΔG0, ΔH0 and ΔS0.
AdsorbentTemperatureΔG0ΔH0ΔS
[°C][kJ·mol−1][kJ·mol−1][J·mol−1·K−1]
Fly ash10−18.11331.09173.89
20−19.969
30−21.586
Modified fly ash10−20.31272.10324.08
20−21.522
30−26.888
Bentonite10−17.86132.34177.61
20−19.924
30−21.400
Modified bentonite10−20.63449.10246.63
20−23.434
30−25.551
Zeolite10−13.14437.88179.28
20−14.119
30−16.768
Modified zeolite10−15.57829.11157.23
20−16.614
30−18.748
Table 13. Descriptive statistics of adsorption capacity.
Table 13. Descriptive statistics of adsorption capacity.
AdsorbentTime [min]Valid NMean [mg·g−1]ConfidenceConfidenceStd. Dev.
Fly ash3061.671.641.700.03
Fly ash3062.932.922.950.02
Fly ash3061.771.741.790.02
Fly ash3063.923.843.990.07
Fly ash3065.004.935.070.06
Fly ash6061.791.791.800.01
Fly ash6063.093.073.110.02
Fly ash6063.523.433.600.08
Fly ash6064.454.444.470.02
Fly ash6065.305.285.320.02
Fly ash9061.871.861.870.01
Fly ash9063.263.253.260.01
Fly ash9064.074.044.090.02
Fly ash9065.645.615.670.03
Fly ash9065.855.845.870.01
Fly ash12061.931.931.940.01
Fly ash12063.333.323.350.01
Fly ash12064.184.174.190.01
Fly ash12065.905.895.900.01
Fly ash12066.656.646.670.02
Bentonite3060.740.700.780.04
Bentonite3062.252.222.270.02
Bentonite3062.992.963.020.03
Bentonite3064.584.554.600.03
Bentonite3065.014.875.150.13
Bentonite6061.441.421.460.02
Bentonite6062.412.382.430.03
Bentonite6064.164.144.180.02
Bentonite6065.135.105.150.02
Bentonite6065.465.425.510.04
Bentonite9061.861.851.870.01
Bentonite9062.742.712.770.03
Bentonite9064.604.584.610.01
Bentonite9065.995.976.010.02
Bentonite9067.037.027.030.01
Bentonite12061.861.851.870.01
Bentonite12063.243.163.320.08
Bentonite12064.604.584.610.01
Bentonite12065.995.976.010.02
Bentonite12067.037.027.030.01
Zeolite3060.810.800.830.01
Zeolite3061.341.321.360.02
Zeolite3063.153.123.180.03
Zeolite3063.663.633.690.03
Zeolite3063.423.393.440.03
Zeolite6060.890.850.920.03
Zeolite6062.011.982.030.02
Zeolite6063.983.964.010.03
Zeolite6063.863.733.990.12
Zeolite6064.584.544.630.04
Zeolite9061.351.321.380.03
Zeolite9062.912.902.920.01
Zeolite9064.064.034.080.02
Zeolite9064.974.935.010.04
Zeolite9065.435.375.500.06
Zeolite12061.581.561.590.02
Zeolite12063.113.073.150.04
Zeolite12064.114.114.120.00
Zeolite12065.705.665.750.04
Zeolite12065.515.475.560.04
Modified fly ash3061.941.931.940.01
Modified fly ash3063.493.493.500.01
Modified fly ash3064.664.654.660.00
Modified fly ash3066.166.166.160.00
Modified fly ash3067.047.037.060.01
Modified fly ash6061.961.961.970.00
Modified fly ash6063.523.513.520.00
Modified fly ash6064.684.674.690.01
Modified fly ash6066.266.206.320.05
Modified fly ash6067.077.037.110.04
Modified fly ash9061.971.971.970.00
Modified fly ash9063.523.513.520.00
Modified fly ash9064.634.634.640.00
Modified fly ash9066.186.176.180.01
Modified fly ash9067.037.027.030.01
Modified fly ash12061.981.981.990.00
Modified fly ash12063.533.533.530.00
Modified fly ash12064.644.634.640.01
Modified fly ash12066.256.256.260.00
Modified fly ash12067.057.057.060.00
Modified bentonite3061.951.951.950.00
Modified bentonite3063.503.493.510.01
Modified bentonite3064.684.674.690.01
Modified bentonite3066.146.136.160.01
Modified bentonite3066.316.286.340.03
Modified bentonite6061.961.951.960.00
Modified bentonite6063.523.513.520.00
Modified bentonite6064.704.704.710.00
Modified bentonite6066.286.276.280.00
Modified bentonite6066.936.906.960.03
Modified bentonite9061.961.961.960.00
Modified bentonite9063.513.513.510.00
Modified bentonite9064.724.714.720.00
Modified bentonite9066.296.296.300.00
Modified bentonite9066.976.966.970.01
Modified bentonite12061.971.971.970.00
Modified bentonite12063.523.523.520.00
Modified bentonite12064.744.744.750.00
Modified bentonite12066.306.306.300.00
Modified bentonite12067.057.047.060.01
Modified zeolite3061.951.951.950.00
Modified zeolite3063.533.533.530.00
Modified zeolite3064.614.614.620.01
Modified zeolite3066.286.286.280.00
Modified zeolite3067.137.137.130.00
Modified zeolite6061.961.961.960.00
Modified zeolite6063.563.563.570.00
Modified zeolite6064.724.724.720.00
Modified zeolite6066.346.336.340.00
Modified zeolite6067.147.147.150.00
Modified zeolite9061.961.951.960.00
Modified zeolite9063.583.583.590.00
Modified zeolite9064.794.794.790.00
Modified zeolite9066.366.366.360.00
Modified zeolite9067.177.127.220.05
Modified zeolite12061.981.981.980.00
Modified zeolite12063.603.603.600.00
Modified zeolite12064.724.584.860.13
Modified zeolite12066.376.376.380.01
Modified zeolite12067.177.127.220.05
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

Ďuricová, A.; Prepilková, V.Š.; Sečkár, M.; Schwarz, M.; Samešová, D.; Murajda, T.; Andráš, P.; Eštoková, A.; Hološová, M.Č.; Poništ, J.; et al. Comparison of Cu(II) Adsorption Using Fly Ash and Natural Sorbents During Temperature Change and Thermal–Alkaline Treatment. Materials 2025, 18, 4552. https://doi.org/10.3390/ma18194552

AMA Style

Ďuricová A, Prepilková VŠ, Sečkár M, Schwarz M, Samešová D, Murajda T, Andráš P, Eštoková A, Hološová MČ, Poništ J, et al. Comparison of Cu(II) Adsorption Using Fly Ash and Natural Sorbents During Temperature Change and Thermal–Alkaline Treatment. Materials. 2025; 18(19):4552. https://doi.org/10.3390/ma18194552

Chicago/Turabian Style

Ďuricová, Anna, Veronika Štefanka Prepilková, Michal Sečkár, Marián Schwarz, Dagmar Samešová, Tomáš Murajda, Peter Andráš, Adriana Eštoková, Miriama Čambál Hološová, Juraj Poništ, and et al. 2025. "Comparison of Cu(II) Adsorption Using Fly Ash and Natural Sorbents During Temperature Change and Thermal–Alkaline Treatment" Materials 18, no. 19: 4552. https://doi.org/10.3390/ma18194552

APA Style

Ďuricová, A., Prepilková, V. Š., Sečkár, M., Schwarz, M., Samešová, D., Murajda, T., Andráš, P., Eštoková, A., Hološová, M. Č., Poništ, J., Zacharová, A., Schmidtová, J., Veverková, D., & Biroň, A. (2025). Comparison of Cu(II) Adsorption Using Fly Ash and Natural Sorbents During Temperature Change and Thermal–Alkaline Treatment. Materials, 18(19), 4552. https://doi.org/10.3390/ma18194552

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