Next Article in Journal
Shear-Induced Degradation and Rheological Behavior of Polymer-Flooding Waste Liquids: Experimental and Numerical Analysis
Previous Article in Journal
Interfacial Solar Evaporation for Treating High-Salinity Wastewater: Chance and Necessity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on the Performance of Copper(II) Sorption Using Natural and Fe(III)-Modified Natural Zeolite–Sorption Parameters Optimization and Mechanism Elucidation

1
Department of Environmental Engineering, Faculty of Chemistry and Technology, University of Split, Ruđera Boškovića 35, 21 000 Split, Croatia
2
Institute for Technology of Nuclear and Other Mineral Raw Materials, 86 Franchet d’Esperey St., 11 000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Processes 2025, 13(9), 2672; https://doi.org/10.3390/pr13092672
Submission received: 24 July 2025 / Revised: 18 August 2025 / Accepted: 21 August 2025 / Published: 22 August 2025
(This article belongs to the Section Environmental and Green Processes)

Abstract

This study evaluates and compares the sorption performance of natural zeolite (NZ) and Fe(III)-modified zeolite (FeZ) in removing Cu(II) ions from aqueous solutions, with the goal of assessing their potential for environmental remediation. NZ was modified with Fe(NO3)3, NaOH and NaNO3 solutions to improve its sorption properties. The modification led to a slight decrease in crystallinity (XRD), increase in pore volume (BET), functional groups (FTIR) and negative surface charge (zeta potential), thereby improving the affinity of FeZ towards Cu(II). Batch sorption experiments were conducted to optimize key parameters including pH, solid/liquid ratio (S/L), contact time, and initial Cu(II) concentration. The pHo and S/L ratio were identified as key factors significantly influencing Cu(II) sorption on both zeolites, with a particularly pronounced effect observed for FeZ. The optimal conditions determined were pHo = 3–5 for NZ, pHo = 3 for FeZ, S/L = 10 g/L and a contact time of 600 min. Experimental results confirmed that FeZ has almost twice the sorption capacity for Cu(II) compared to NZ (0.271 mmol/g vs. 0.156 mmol/g), as further supported by elemental analysis, SEM-EDS and mapping analysis of saturated samples. The sorption of Cu(II) followed a mechanism of physical nature driven by ion exchange, dominated by intraparticle diffusion as the rate-controlling step. Leaching of copper-saturated zeolites according to the standard leaching method, DIN 38414 S4, demonstrated the ability of both zeolites to fully retain Cu(II) within their structure over a wide pH range, 4.01 ≤ pHo ≤ 10.06. These findings highlight the superior performance of FeZ and its potential as an effective material for the remediation of copper-contaminated environments.

1. Introduction

Copper is an essential biochemical nutrient of vital importance to humans, animals and plants, and is considered fairly harmless [1,2,3]. It is essential for plants to form chlorophyll, photosynthesis, cell wall metabolism, etc. [4]. Since the human body is unable to synthesize copper, it must be ingested through diet as it is substantial for the synthesis of so-called cuproenzymes such as cytochrome C oxidase and dopamine β-hydroxylase, for bone and tissue development, and supports iron absorption and hemoglobin synthesis [3,5,6]. Due to the fact that copper deficiency/excess negatively affects human health, various organizations prescribe a minimum daily copper intake in the form of a U-shaped curve [2,7]. The recommendation for an average dietary intake for adults is 0.9 to 1.3 mg Cu/day, since up to 1 mg of copper is excreted daily from the liver into bile and feces [8,9]. Copper deficiency most often occurs in children, and is manifested by failure to thrive, hypothermia, neutropenia and mental illness, and is also known as Menkes syndrome [6,7,10]. In adults, copper deficiency is quite rare, and symptoms include anemia conditions, immunodeficiency, osteoporosis, and neurological disorders [5,6]. However, in some cases, daily copper intake ranges from 1 to even 5 mg/day, most often through the consumption of contaminated food products (vegetables, seeds, and tap water) and the use of vitamin supplements and food additives [10]. Namely, cases of direct copper intoxication by inhalation, ingestion and skin contact have been reported [11]. Furthermore, acute copper intoxication (>8–10 mg Cu/kg of body weight) [12] causes a number of consequences such as acute kidney damage, gastrointestinal bleeding, cardiac arrhythmia, hypotension, hemolytic anemia, hematuria, methemoglobinemia, hepatocellular toxicity, autism and hypothyroidism [2,3,8,10]. In contrast to acute copper intoxication, chronic copper overload with its abnormal increased accumulation in the liver and brain is an autosomal recessive genetic disorder, and refers to hepatolenticular degeneration, the so-called Wilson’s disease [2,7,8,10,13]. This disease first results in liver dysfunction, and then neurological disorders that lead to Alzheimer’s, Parkinson’s and Huntington’s disease [7,8,14]. Although copper is essential for the functioning of cellular metabolism, its excessive amounts in the human body lead to serious disorders mentioned above [8]. For this reason, the US Environmental Protection Agency has set health guidelines for the maximum copper level in drinking water up to 1.3 mg Cu/L [15], while the World Health Organization prescribes 2 mg Cu/L [16]. Likewise, Croatian legislation prescribes a maximum permitted level of copper in drinking water up to 2 mg/L [17].
Namely, copper enters the environment less frequently from natural sources (volcanoes, forest fires, rocks weathering and erosion, and vegetation decomposition) than from anthropogenic sources (mining and smelting activities of copper ores, wastewater, and agriculture) [4,18]. In proximity to iron and aluminum, copper is the third most used metal worldwide [5]. It is used for the production of water pipes, electrical wirings, alloys (brass and bronze), electronics devices, copper coins, electroplating, azo dyes, explosives, phytopharmaceuticals, food additives (coloring agent), etc. [2,3,5,10,19]. Due to its extremely high usage, copper is considered one of the major anthropogenic environmental contaminants from the group of heavy metals [5].
Karim’s [2] review paper documented the copper concentration distribution in all parts of the environment: air, soil and water. Thus, the average copper concentrations in the air ranging from a few ng/m3 to 200 ng/m3 were found, while in the industrial area they rise even up to 5000 ng/m3. The production and processing of copper causes the emission of fumes and dust particles enriched with copper [2]. Likewise, soil contamination with copper is the primary result of excessive use of Cu-based agrochemicals for agricultural crop protection [4,18,20]. Since copper has antifungal properties, Cu-based fungicides such as Bordeaux mixture (CuSO4 + CaO) and other Cu-based agrochemicals are used to treat various plant diseases, and algicides to suppress algae growth in swimming pools [4,18,19]. As a consequence of the widespread use of Cu for agricultural purposes, the average Cu concentration in soil worldwide is estimated to be approximately 30 mg Cu/kg [18], and values ranging from 2 to 17,000 mg/kg have been recorded. Consequently, increased copper concentrations in the soil lead to its accumulation in vegetables used for human consumption [20]. Furthermore, the concentration of copper in water bodies (rivers, lakes) varies from 0.5 to 1000 µg/L. An interesting fact is the presence of copper in tap water in the range of 20–75 µg/L [2]. Therefore, the main routes of human copper intoxication are inhalation of polluted air, consumption of contaminated vegetables and nuts, and tap water from copper plumbing. Moreover, a review paper by Liu et al. [5] documented copper content in wastewater in the range of 2.5–10,000 mg/L. Globally, it is estimated that approximately 90% of wastewater produced remains untreated and is discharged directly into the environment [4]. This finding is quite concerning and indicates the need to raise awareness of the need for adequate wastewater treatment in the 21st century, which is in accordance with Sustainable Development Goal 6 on water and sanitation (SDG 6) and the idea of a circular economy [21,22]. For instance, Croatian legislation prescribes a maximum permissible copper level in treated industrial wastewater of 1 mg/L for discharge into the public sewage system, and even more rigorously for discharge into water bodies of 0.5 mg/L [23]. Therefore, remediation of copper-contaminated areas is necessary to mitigate its negative effects on the all environmental compartments.
Various physical, chemical and biological methods for treating copper-contaminated waters have been investigated. Comprehensive review papers by Liu et al. [5], Li et al. [24] and Al-Saydeh et al. [25] reported that chemical precipitation, adsorption/ion exchange, cementation, membrane and electrochemical methods are the most commonly used for copper removal from industrial wastewater. Notwithstanding, the generation of sludge and the need for its further treatment and/or disposal after chemical precipitation, as well as the high energy requirements for electrochemical and membrane methods, put these methods into consideration and the need to find more environmentally friendly ones. In contrast, the advantage of adsorption is simple design and operation, low capital costs, and removal of cationic and anionic contaminants from relatively large volumes at the parts per billion (ppb) level [25]. Since the 1940s, adsorption and ion exchange has gained importance for the treatment of industrial wastewaters with the use of activated carbon and synthetic ion exchangers [26]. Despite this, the high production costs of these materials call into question the cost-effectiveness of the adsorption implementation, whereby the adsorbent cost price often plays a decisive role in its selection [27]. Consequently, in order to increase the profitability of adsorption, it is crucial to find natural, highly selective, easily accessible and non-toxic green low-cost sorbents.
Specifically, Krstić et al. [28] highlighted adsorption as a promising green technology for the treatment of wastewater containing copper ions using environmentally friendly and price affordable natural adsorbents such as zeolites, clays and biosorbents. In the review paper by Irannajad and Haghighs [29], a sequential order of adsorption capacity of mineral adsorbents for heavy metals is given, among which the following stand out: zeolite > clay > diatomite. Accordingly, natural zeolites appear to be a promising option since they meet the previously mentioned conditions of sustainability and economic viability along with environmental friendliness.
Natural zeolites consist of (Si and Al)O4 tetrahedra connected by all of their oxygen vertices, with generating channels where exchangeable cations (such as Na+, K+, Ca2+ and Mg2+) and H2O molecules balance the negative charge produced by isomorphous substitution of silicon with aluminum. Zeolites are used in many different applications, except as a sorbent for heavy metals, they are also used as chemical sieves and water softeners, due to their structural characteristics and adsorption capabilities [27,30,31]. According to the findings, copper was primarily removed from water solutions using zeolites by ion-exchange process, wherein copper ions from the solution replace ions like Na+, K+, Ca2+, and others that are present in the pores of zeolite crystalline lattices [28].
Modifying natural zeolites to enhance their adsorption capabilities and selectivity to a certain substance or group of compounds has been the subject of numerous investigations in recent decades. The goal of each modification is to overcome the typical traditional properties of natural zeolites, such as the lack of active sites or the absence of selective active sites. Accordingly, modifications cause changes in surface physicochemical characteristics as a result of the creation of new active sites, changes in the zeolite surface charge, and an increase in porosity, which ultimately affects the increase in sorption efficiency [28,32,33,34]. Common methods for activating natural zeolites include single or combined chemical (acids, bases, or inorganic salts), thermal, surfactant modification and metal oxide modification. The intended application and economic cost-effectiveness should be taken into account while selecting modification methods [28]. The modification of zeolite with inorganic salts, among which the most researched ones are those with Fe-salts, is primarily intended to create new active sites, remove impurities from zeolite channels and increase the negative charge of the zeolite lattice. Namely, Fe-zeolites include incorporation of iron on the outer and inner zeolite surface in the form of mononuclear and binuclear oxygen-bridges iron species. The integration of Fe into the zeolite framework effectively addresses the aforementioned limitations of traditional aluminosilicate zeolites and expands their use in environmental applications [34]. In recent years, many researchers have investigated the possibility of copper removal using natural zeolites [35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52] as opposed to acid-modified [53], thermally activated [54], and especially iron-modified natural zeolites [55,56,57,58]. This fact provides an opportunity for a more systematic investigation of the application of Fe(III)-modified natural zeolites as sorbents for copper binding, which serves as the foundation of this paper.
Hence, the aim of this study is to comparatively investigate the impact of sorption parameters (pHo, solid/liquid ratio, contact time, and initial concentration) and their optimization on the amount of sorbed copper onto natural and Fe(III)-modified natural zeolite, as well as on the removal efficiency. In addition, the sorption mechanism will be clarified by applying kinetic and isotherm models as well as SEM-EDS of copper-saturated zeolites. Finally, the leaching experiments of copper-saturated zeolites will provide a judgment on the possibility of their application for in situ and ex situ remediation of copper-contaminated environments.

2. Materials and Methods

2.1. Zeolite Preparation

The initial sample, natural zeolite (NZ), was sourced from the Zlatokop deposit located in Vranjska Banja, Serbia. The sample was ground, sieved, and a particle size fraction of 0.6–0.8 mm was isolated following the standard procedure (DIN 66165-2) [59]. Subsequently, the sample was rinsed with ultrapure water and dried at 60 °C.
Fe(III)-modified natural zeolite (FeZ) was prepared by treating NZ based on the procedure outlined in a previous study [60]. Briefly, NZ is first mixed with a freshly prepared 1 mol/L solution of Fe(NO3)3·9H2O in acetate buffer at pH = 3.6, followed by the addition of 1 mol/L NaOH solution and finally 4% NaNO3. The prepared sample was dried at 40 °C and stored in a desiccator. Physico-chemical characterization of both NZ and FeZ (chemical composition, SEM/EDS, XRD, FTIR, BET surface area, TG/DTG) was detailed in a previous study [60]. XRD analysis showed that NZ primarily consists of clinoptilolite (80%), with quartz, feldspar, and carbonates as secondary minerals. Modification resulted in slight crystallinity loss (XRD), broader hydroxyl bands (FTIR), a small increase in surface area, a significant increase in pore volume (BET), and higher mass loss (TG/DTG). FTIR analyses are shown in Figures S1 and S2 and are explained in detail in the Supplementary Materials. Zeta potential in a wide pH range (2–12) revealed that both zeolites have a negative charge, of which FeZ is more negative. In summary, the modification aimed to bind the predominantly present Fe3(OH)45+ species at pH = 3.6 to the negative surface of NZ. The addition of NaOH caused the hydroxylation of Fe species, forming negative Fe oxo- and hydroxo-complexes, which increased the negative charge of FeZ.

2.2. Batch Sorption Experiments

The salt, Cu(NO3)2·3H2O purchased from Kemika was used to prepare a stock solution with an initial concentration of 15.744 mmol/L. Solutions of lower concentrations were prepared by diluting the stock solution. Ultrapure water was used to prepare all solutions, and a 0.1 or 1 mol/L HNO3 solution was used to adjust the pH. All experiments were performed in an incubator shaker at 25 °C, 230 rpm, and for 24 h. Cu(II) concentrations before and after equilibrium were determined using a Flame Atomic Absorption Spectrophotometer, AAS, model PinAAcle 900F, (Perkin Elmer Inc., Waltham, MA, USA).

2.2.1. Impact of pH

The impact of pHo on the Cu(II) sorption onto NZ and FeZ was examined in the pHo range 2.25 ≤ pHo ≤ 5.01 and at two initial concentrations of 4.017 mmol Cu/L and 8.137 mmol Cu/L. A mass of 1.0 g of NZ or FeZ was mixed with 100 mL of Cu(II) solution (solid/liquid ratio = 10 g/L). After the equilibrium was established, the equilibrium pHe was measured, as well as the amount of released exchangeable cations (Na, K, Ca and Mg) by ion chromatography (Metrohm 761 Compact IC, Metrohm Ltd., Herisau, Switzerland).

2.2.2. Impact of Solid/Liquid Ratio

The impact of the solid/liquid ratio (S/L) was examined at pHo = 4.83 for NZ and pHo = 2.98 for FeZ, and at an initial concentration of 4.017 mmol Cu/L. The specified pHo values were determined as optimal based on the experiment described in Section 2.2.1. Experiments for both zeolites were carried out at S/L = 2–18 g/L (m = 0.2; 0.6; 1.0; 1.4 and 1.8 g). After equilibrium was established, the equilibrium pHe was measured.

2.2.3. Impact of Contact Time

The impact of contact time on the Cu(II) sorption onto NZ and FeZ was carried out at a previously determined optimal pHo (pHo = 4.83 for NZ and pHo = 2.98 for FeZ) and at S/L = 10 g/L based on the conducted experiment described in Section 2.2.2. A mass of 20 g of NZ or FeZ was mixed with 2 L of Cu(II) solution with an initial concentration of 10.061 mmol/L at 550 rpm. Liquid samples were collected at selected time intervals within 24 h, whereby the total volume of the samples not exceeding 5–6% of the total volume of the suspension. The saturated samples, designated as NZCu and FeZCu, were collected, washed several times in ultrapure water, dried at 40 °C, and used for leaching experiments.

2.2.4. Impact of Initial Concentration

The impact of initial Cu(II) concentration on Cu(II) sorption onto NZ and FeZ was tested in the concentration range 0.508–15.744 mmol Cu/L at the previously determined optimal S/L = 10 g/L and pHo (pHo = 4.83 for NZ and pHo = 2.98 for FeZ). A volume of 100 mL of Cu(II) solution, of different initial concentrations, was mixed with 1.0 g of NZ or FeZ. After the equilibrium was established, in addition to measuring the equilibrium pHe, the concentrations of released zeolite cations (Na, K, Ca, and Mg) were also measured by ion chromatography. Saturated samples with the highest initial concentration of 15.744 mmol Cu/L were collected, washed several times in ultrapure water, dried at 40 °C, and labeled as NZCu and FeZCu. Then, the chemical composition of the starting samples, NZ and FeZ, and the saturated samples, NZCu and FeZCu, was determined by classical chemical analysis of aluminosilicates [61]. A comparison of the chemical composition of the initial and saturated samples is shown in Table 1.
Scanning electron microscopy (SEM) combined with energy-dispersive X-ray analysis (EDS) of copper-saturated samples, NZCu and FeZCu, was performed on a JEOL JSM-6610 microscope (JEOL Ltd., Akishima, Tokyo, Japan). On selected surfaces of the analyzed samples, the semi-quantitative chemical composition was determined, and the distribution of elements on the surface was determined by mapping analysis.

2.3. Leaching Experiments

Leaching of Cu(II) from the collected saturated samples (NZCu and FeZCu) after conducting the experiment described in Section 2.2.3., were subjected to the standard leaching method, DIN 38414 S4 [62]. Ultrapure water with pH pre-adjusted with 0.1 mol/L HNO3 or 0.1 mol/L KOH in the range of 2.06–12.09 was used as leachant. A mass of 1.0 g of NZCu or FeZCu was mixed with 10 mL of ultrapure water of different pHo for 24 h at 25 rpm and at 25 °C. After 24 h, in the filtered suspensions the equilibrium pHe and the concentration of leached Cu(II) were determined using a Flame Atomic Absorption Spectrophotometer.

2.4. Calculation of Sorption Parameters

The quantity of Cu(II) sorbed onto zeolites at time t (up to 24 h), denoted as qt (mmol/g), as well as the removal efficiency at time t, denoted as αt (%), were determined using Equations (1) and (2):
q t = c o c t V m
α t = c o c t c o 100
Here, co and ct represent the concentrations of Cu(II) at t = 0 and at time t (mmol/L), V is the volume of the solution (L), and m is the mass of the zeolite (g). When t = 24 h, qt and αt represents the equilibrium quantity of Cu(II) sorbed (qe) and equilibrium removal efficiency (αe).
The quantity of Cu(II) leached from the saturated zeolites, qleach (mmol/g), and the percentage of Cu(II) leached, αleach (%), were calculated using Equations (3) and (4):
q leach = c leach V m
α leach = q leach q e 100
Here, cleach represents the concentration of Cu(II) leached from the saturated zeolites (mmol/L).

2.5. Theoretical Backround—Kinetic Analysis for Identifying Rate-Controlling Steps

To identify the rate-controlling step, whether it is mass transfer, diffusion, or chemical reaction, various reaction and diffusion kinetic models are used to analyze the experimental data. The most commonly used reaction kinetic models to verify if a chemical reaction is the limiting step are the pseudo-first-order (PFO) model (Lagergren), the pseudo-second-order (PSO) model (Ho), and the Elovich model. The pseudo-first-order non-linear equation is given by the following expression [63,64]:
q t = q m ( 1 e k 1 t )
In this equation, qt represents the quantity of Cu(II) sorbed onto the zeolites at time t (mmol/g), qm is the amount of Cu(II) sorbed on the zeolite obtained from the model, k1 is the rate constant of the PFO (1/min), and t is the time (min).
The non-linear form of the pseudo-second-order model is expressed by Equation (6) [63,64]:
q t = k 2 q m 2 t 1 + k 2 q m t
where k2 is the rate constant of the PSO [g/(mmol·min)].
The Elovich model describes the chemical sorption of Cu(II) onto a heterogeneous zeolite surface, and the non-linear form is represented by Equation (7) [64]:
q t = 1 β E ln 1 + α E β E t
where αE represents the initial chemisorption rate [mmol/(g·min)] and βE is associated with the surface coverage (g/mmol).
Since zeolites are porous materials, sorption is frequently restricted by mass transfer to the active sites within the zeolite structure. In well-stirred systems, the rate-controlling step is typically linked to film or intra-particle diffusion, as the resistance to mass transfer from the bulk to the particle surface is negligible. To determine whether film or intra-particle diffusion is the slowest step of sorption process, different diffusion kinetic models can be applied. The rate-limiting step in sorption, whether due to film or intraparticle diffusion, can be determined using the linear equation of the Weber–Morris model, as shown below [65]:
q t = k WM t 1 / 2 + I
where kWM is the Weber–Morris diffusion constant [mmol/(g∙min1/2)], and I represents the boundary layer thickness (mmol/g).
If I = 0, film diffusion is the sole rate-controlling step; otherwise, both film and intraparticle diffusion contribute to mass transfer. The contribution of film or intraparticle diffusion can be estimated using Equation (9) [65]:
R C = I q e 100
where RC denotes the relative coefficient, expressed as a percentage. Smaller RC values indicate that film diffusion contributes less to the overall mass transfer process.
The Weber–Morris diffusion coefficient, kWM (cm2/min) is calculated using the following equation [65]:
D WM = π d p k WM 12 q e
where dp is the zeolite particle diameter (cm).
The two-step sorption kinetics are represented by a double-exponential non-linear model, as shown below [66,67]:
q t = q m B 1 m z e k B 1 t B 2 m z e k B 2 t
Here, B1 and B2 represent the concentrations of the Cu(II) sorbed in the fast and slow steps (mmol/L), while kB1 and kB2 are the rate constants for the fast and slow steps (1/min), respectively.
The total sorption rate, r [mmol/(g∙min)] is the sum of the fast step, r1, and the slow step, r2 [66,67]:
r = r 1 + r 2 = B 1 m z k B 1 + B 2 m z k B 2
Furthermore, the percentage of Cu(II) sorbed in the fast (RF) and slow (SF) steps can be calculated as follows [66,67]:
R F = B 1 B 1 + B 2 100
S F = B 2 B 1 + B 2 100
Vermeulen’s approximation assumes intraparticle diffusion as the rate-controlling step and is given by Equation (15) [68]:
q t = q m 1 e 4 D V π 2 t r p 2 1 2
In this equation, DV is the intraparticle diffusion coefficient (cm2/min), and rp refers to the radius of the zeolite particle (cm).
The model parameters were calculated using the mathematical tool MathCad 15. The agreement of the model with the experimental data was determined using the linear and non-linear correlation coefficients, R2 and r2. In addition to this, two error functions were used, non-linear chi-square test (χ2) and root mean square error (RMSE) as follows [64]:
χ 2 = i = 1 n q m q e 2 q m
RMSE = i = 1 n q m q e 2 n

3. Results and Discussion

3.1. Determination of Optimal Sorption Parameters

Determining the optimal process parameters (pHo, solid/liquid ratio, contact time and concentration range) is crucial to maximize both, the amount of metal ions sorbed onto the sorbent as well as the metal removal efficiency from the aqueous phase. Although it is important to fully utilize the capacity of the sorbent, from a practical point of view, removal efficiency is much more important because it reflects the success of the sorption process. Hence, sorption efficiency will be considered as a more critical factor in determining optimal sorption parameters.

3.1.1. Determination of Optimal pH

The pH of the suspension is one of the most important factors in the sorption process, as it directly affects the zeolite surface charge as well as the charge and type of Cu(II) species. Accordingly, based on the hydrolysis constants of Cu(II) in the aqueous medium shown by Equations (18)–(21) [69], the distribution of Cu(II) species as a function of pH was constructed as shown in Figure 1.
Cu 2 + + H 2 O CuOH + + H +                        pK 1 = 8.1
Cu 2 + + 2 H 2 O Cu OH 2 o + 2 H +                  pK 2 = 16.1
Cu 2 + + 3 H 2 O Cu OH 3 - + 3 H +                  pK 3 = 26.7
Cu 2 + + 4 H 2 O Cu OH 4 2 - + 4 H +                  pK 4 = 39.6
Figure 1 illustrates the change in the Cu(II) species distribution in aqueous solution depending on pH. Up to pH = 6, copper is present primarily in the form of Cu2+ species with a content of 99%. Above pH = 5, Cu2+ starts to convert into hydroxylated species. Thus, in the pH range from 5.2 to 10.7, a small amount of CuOH+ species appears with a maximum content of 31% at pH = 8. Cu(II) in the form of Cu(OH)2 starts to precipitate at pH = 6.5 with a maximum content of 90% at pH = 9.5. At pH > 8, anionic Cu(II) species are formed. The Cu(OH)3 species is dominant at pH = 11.8 with a content of 88%. At pH = 10.2, the Cu(OH)42− species appears, which has an increasing trend with increasing pH. Therefore, this suggests that the Cu(II) sorption process needs to be conducted below pH = 6.5 to exclude Cu(II) precipitation.
In order to obtain additional confirmation of the pH value at which Cu(II) precipitation occurs (pHppt), taking into account the initial Cu(II) concentration, the following expression is used [70]:
pH ppt = 14 log c o Cu ( II ) K sp Cu OH 2
where co[Cu(II)] is the initial Cu(II) concentration and Ksp is the solubility product constant of Cu(OH)2, Ksp = 2.20∙10−20 [71].
The measured pHe and calculated pHppt for the two initial Cu(II) concentrations (4.017 mmol/L and 8.137 mmol/L), along with the removal efficiency (αe) as a function of pHo, are presented in Figure 2a,b.
In Figure 2a, it is noticeable that for both initial Cu(II) concentrations, pHe < pHppt for NZ, which clearly excludes the precipitation of Cu(II) in suspensions for all pHo. In the case of FeZ, according to the calculated pHppt, for both initial Cu(II) concentrations at pHo ≥ 4, Cu(II) precipitation occurs. The calculated pHppt value for 4.017 mmol Cu/L is pHppt = 5.6, while pHppt = 5.8 for 8.137 mmol Cu/L. Namely, according to the Cu(II) speciation diagram depending on pH (Figure 1), the proportion of Cu(OH)2 at pH = 5.6 is 0.001%, while at pH = 5.8 it is 0.003%. Although the proportions of the precipitate formed are negligible, in order to reliably claim that no precipitate was formed during the sorption process of Cu(II) on FeZ, an optimal range of pHo values up to pHo = 3 was selected. The optimal pHo for Cu(II) sorption on NZ and FeZ will be determined based on the results of Cu(II) removal efficiency depending on pHo (Figure 2b). Therefore, the optimal pHo is the value where maximum removal efficiency is achieved without the possibility of precipitation. For both initial concentrations, the optimal pH range is 3 ≤ pHo ≤ 5 for NZ, while for FeZ it is pHo = 3.
Furthermore, for both samples, NZ and FeZ, the removal efficiency increases with increasing pHo value. Comparing the Cu(II) removal efficiency values for both samples and for both initial concentrations, significantly higher removal efficiency was achieved for FeZ compared to NZ. This indicates that the zeolite modification significantly contributed to the reduction in Cu(II) concentration. Namely, the zeta potential of both samples has negative values in the pH range 2–12, while FeZ having a more pronounced negative charge. This evidence argues that the electrostatic attraction between the positive Cu2+ species and the negatively charged surface of NZ and FeZ will be favorable in the acidic pH range where Cu(II) precipitation does not occur. In order to obtain a more complete insight into the effect of pHo on the sorption of Cu(II) on NZ and FeZ, the interrelationship between in-going Cu(II) ions and out-going exchangeable ions (Na, K, Ca, and Mg) from the zeolite structure was analyzed and shown in Figure 3 and Figure 4 for two initial Cu(II) concentrations.
Interrelationship between in-going Cu(II) ions and out-going exchangeable cations for both initial concentrations and both samples is mostly non-stoichiometric. At extremely acidic pH, where the lowest removal efficiency was recorded, the highest non-stoichiometry is observed for both samples, which is a consequence of the competition between Cu2+ and H+ ions. Non-stoichiometry is more pronounced for FeZ, which is a result of a more pronounced negative charge of the FeZ surface, and thus the more significant influence of the competitive effect. Already at pHo ≥ 3.05 for NZ and at pHo = 3.05 for FeZ, the competitive effect is almost negligible since the non-stoichiometric interrelationship is the least pronounced. Furthermore, for FeZ at pH ≥ 4.04, a non-stoichiometric relationship is again observed, but of a reverse character. The non-stoichiometry is reflected in a higher amount of sorbed Cu(II) ions compared to the released exchangeable cations. This is a consequence of Cu(II) precipitation, which additionally contributes to the increase in removal efficiency, which is in accordance with the results shown in Figure 2. These findings suggest and reaffirms that the optimum pH for NZ is in the range of pHo = 3–5, and for FeZ, it is pHo = 3.
The emphasis on the importance of testing optimal pH is also supported by relatively recent research. Thus, Panayotova [35] reported the optimal pH for Cu(II) removal on Bulgarian natural zeolite in the range 5.5–7.5, which is quite debatable considering the area of copper precipitation according to Figure 1. Forughirad et al. [36] investigated the removal of Cu(II) at pH = 7 and pH = 5 on Azerbaijan natural zeolite clinoptilolite, and found the highest percentage of removed Cu(II) at pH = 7. This is proof that the tests of the effect of pH on sorption processes must be accompanied by theoretical foundations and knowledge of the speciation of any metal cation depending on pH. Ultimately, this is also the only way to confidently claim that the sorption mechanism is solely responsible for the reduction in metal concentrations, without the possible contribution of precipitation.

3.1.2. Determination of Optimal Solid/Liquid Ratio

The effect of the solid/liquid (S/L) ratio was determined at an initial concentration of 4.017 mmol Cu/L and at pHo = 4.83 for NZ and pHo = 2.98 for FeZ. The comparison of pHe with pHppt and the removal efficiency, αe of Cu(II) on NZ and FeZ at different S/L ratios is shown in Figure 5.
For both zeolite samples, an almost linear increase in removal efficiency is observed with increasing S/L up to S/L = 14 g/L, and then a more abrupt increase in removal efficiency is recorded. Namely, since pHo and Cu(II) concentration are constant at all S/L ratios, an increase in the S/L ratio is accompanied by a greater amount of available free active sites on zeolites for Cu(II) sorption, which is reflected in an increase in removal efficiency. On the other hand, although the increase in S/L has a positive effect on the increase in removal efficiency, it also affects the increase in pHe value (Figure 5). The increase in pHe is a consequence of both the decrease in the concentration of Cu(II) in the solution and the competition with H+ ions, which was discussed earlier. However, it can be seen from Figure 5 that pHe > pHppt for both samples at S/L = 18 g/L, indicating the occurrence of Cu(II) precipitation, which is consistent with a more abrupt increase in removal efficiency compared to S/L = 14 g/L. Accordingly, in addition to determining optimal pH, determining optimal S/L is essential in order to achieve maximum sorption removal efficiency without the occurrence of Cu(II) precipitation. Keeping the above in mind, for both NZ and FeZ, the optimal S/L is 10 g/L, since at S/L = 14 despite slight increase in αe, the pHe is very close to the pHppt.

3.1.3. Determination of Optimal Contact Time

Determining the impact of contact time is important for balancing the desired efficiency and process cost-effectiveness. Figure 6 depicts time-dependent changes in the amount of Cu(II) sorbed on NZ and FeZ as well as the removal efficiency.
Time-dependent changes in qt indicate rapid sorption of Cu(II) onto NZ and FeZ up to 120 min, followed by slower sorption up to 600 min, after which equilibrium is established. The rapid initial sorption phase is probably attributed to the easily accessible zeolite surface active sites, while the second phase corresponds to the gradual saturation of the less accessible active sites located within the zeolite structure. The minimum contact time to reach equilibrium is 600 min, achieving almost double the amount of sorbed Cu(II) on FeZ compared to NZ (0.279 mmol/g vs. 0.164 mmol/g). In the initial phase, the majority of Cu(II) is sorbed, 13% on NZ and 23% on FeZ, while the maximum removal efficiency was almost double for FeZ compared to NZ (28% vs. 16%).
Figure 7 shows the fitting of kinetic reaction and diffusion models with the experimental kinetic data of Cu(II) sorption on NZ and FeZ, while the calculated parameters of the kinetic models along with the agreement indicators are listed in Table 2.
The kinetic reaction models were compared based on the fitting parameters (qexp vs. qm, r2, RMSE, and χ2) shown in Table 2. Therefore, the experimental data agree with the kinetic reaction models as follows: PSO > PFO > Elovich model. Since the Elovich model showed the highest disagreement with the experimental results of the three kinetic reaction models considered, chemisorption is certainly not exclusively the slowest limiting step of the sorption process. Furthermore, the deviation of the sorption capacity calculated from the model, qm, and experimentally determined one, qexp, is more significant for PFO than for PSO. This indicates that simple surface sorption of Cu(II) on both zeolites does not take place, but rather more complex interactions such as chemisorption or diffusion. Since the possibility of chemisorption as a limiting step is excluded based on agreement with the Elovich model, it is therefore necessary to fit the experimental data according to diffusion kinetic models.
Accordingly, the kinetic data for NZ and FeZ were fitted to the Weber–Morris model and plotted as a dependence of qt vs. t1/2 as shown in Figure 7b where multicollinearity indicates a more complex kinetic mechanism involving film and intraparticle diffusion. The first linear part corresponds to fast sorption, the second to slower sorption, and the third one to the establishment of equilibrium, which is in line with time-dependent changes in the amount of Cu(II) sorbed on NZ and FeZ (Figure 6a). The values of the parameters kWM1 > kWM2 as well as DWM1 > DWM2 confirm faster sorption in the first phase of the process. In addition, the mentioned parameters are higher in the case of FeZ than NZ, which is reflected in the higher quantity of Cu(II) sorbed on FeZ as well as the higher removal efficiency of Cu(II). Since the value of the calculated relative coefficient (RC) is 13% for NZ and 2% for FeZ, this implies that the proportion of mass transferred by film diffusion is negligible compared to intraparticle diffusion.
The Vermenulen’s approximation model considers intraparticle diffusion as the slowest step in mass transfer. According to the values of correlation parameters shown in Table 2, it could be concluded that intraparticle diffusion is the rate controlling step for NZ and FeZ. Nonetheless, in order to confirm the above, the experimental results were also fitted according to the double-exponential model.
The values of the calculated correlation indicators for the double-exponential model are the most satisfactory of all the tested models. This confirms the two-step sorption of Cu(II) on both zeolites. The first sorption phase, which corresponds to mass transfer through the film, is faster (KB1 > KB2) than the second phase, which is attributed to mass transfer by intraparticle diffusion. This is in line with the conclusions drawn from the Weber–Morris model. Since the values, B1 > B2, as well as RF > SF, the majority of Cu(II) was sorbed on both zeolites during the first, faster sorption phase. The values of the kinetic parameters (KB1, KB2, B1, and B2) for both phases of two-step sorption are higher for FeZ, which is attributed to a higher number of active sites on FeZ due to modification. Taking into account all applied models, the sorption of Cu(II) on both zeolites is complex and takes place in active sites located on the outer and inner surfaces of the zeolite grains. From a kinetic perspective, intraparticle diffusion is the slowest mass transfer step.

3.1.4. Determination of Optimal Concentration Range and Sorption Mechanism

The results of the effect of the initial Cu(II) concentration on the quantity of Cu(II) sorbed per gram of NZ and FeZ, as well as the removal efficiency, are shown in Figure 8.
The effect of the initial Cu(II) concentration is manifested by a gradual increase in the amount of Cu(II) sorbed on NZ up to co = 10 mmol Cu/L and up to co = 12 mmol Cu/L for FeZ. Above the specified concentrations, maximum saturation of zeolite active sites is achieved. The maximum amount of sorbed Cu(II) on FeZ compared to NZ (0.271 mmol/g vs. 0.156 mmol Cu/g) is almost twice as high under optimal Cu(II) sorption conditions. Furthermore, from the aspect of Cu(II) removal efficiency, the highest removal efficiency is achieved at the lowest initial concentration, while an increase in the initial concentration decreases the removal efficiency. This indicates the need to optimize the two mentioned parameters, i.e., maximizing both sorption capacity and removal efficiency.
To place the current research in the context of similar studies, background information from recently conducted research is listed in Table 3 and discussed below.
Namely, numerous studies of copper removal have been carried out on natural zeolites from different regions of the world. Zendelska et al. [37] determined the capacity of Bulgarian natural zeolite clinoptilolite to be 0.150 mmol/g at pH = 3.5 and S/L = 50 g/L. Kabwadza-Corner et al. [38] notified the capacity of natural zeolite clinoptilolite from Shimane Prefecture in Japan towards Cu(II) of 0.074 mmol/g at pH = 5 and S/L = 0.2 g/L. Bakalár et al. [40] obtained the maximum capacity of natural zeolite clinoptilolite from Slovakia towards Cu(II) of 0.037 mmol/g without adjusting the pH. Wilopo et al. [42] determined the Cu(II) capacity of natural zeolite from the Gedangsari deposit, Japan of 0.064 mmol/g at pH = 5. Kyzioł-Komosińska et al. [46] studied the sorption of Cu(II) on Ukrainian natural zeolite clinoptilolite and found a capacity of 0.337 mmol/g at pH = 5 and S/L = 10 g/L. Experimental findings indicate that sorption capacities are primarily related to the geological origin of natural zeolites, wherein each zeolite has unique properties. Furthermore, different experimental conditions yield different outcomes. Notwithstanding, the sorption capacities of different natural zeolites are relatively comparable and consistent with our findings.
Regarding natural zeolites modified with Fe(III) salts, apart to the geological origin, the method of modification also plays a significant role. For example, Doula [55] used Greek natural zeolite clinoptilolite that was modified sequentially with 1 mol/L Fe(NO3)3 and 5 mol/L KOH solutions. At S/L = 10 g/L, the sorption capacity of the modified zeolite towards Cu(II) was 2.8 times higher than that of the parent zeolite (0.59 mmol/g vs. 0.21 mmol/g). It is noteworthy that the equilibrium pH ranged from 6.53 to 8.37 after Cu(II) sorption on the Fe-modified zeolite. Lipovský et al. [56] used the same modification procedure as Doula [55] to obtain Fe-clinoptilolite from natural zeolite originating from Nižný Hrabooc (Slovakia). The authors did not define pH, but they found a 1.05-fold increase in capacity of the modified zeolite compared to the natural zeolite (0.0061 mmol Cu/g vs. 0.0058 mmol Cu/g). Nguyen et al. [57] also used the same procedure as the previous two studies to obtain iron-coated zeolite using natural zeolite heulandite from the Werris Creek deposit, New South Wales, Australia. The research showed a sorption capacity of 0.063 mmol Cu/g of natural zeolite and 0.077 mmol Cu/g of iron-coated zeolite maintaining pH at 6.5 ± 0.2. Thus, the modification contributed to the increase in capacity by 1.22 times. On the other hand, Han et al. [58] used 1 mol/L FeCl3 and 3 mol/L NaOH to produce iron oxide-coated zeolite at 500 °C using natural zeolite clinoptilolite from Xinyang city, China. At S/L = 4 g/L, they found a 1.33-fold higher sorption capacity of the modified sample compared to the starting material (0.081 mmol Cu/g vs. 0.061 mmol Cu/g).
In summary, the modifications contributed to the increase in the capacity of the zeolites, despite differences caused by the origin, type of modification, and experimental conditions. A possible drawback of these studies is that optimization of sorption parameters was not performed, which would be a contribution of the current research. Of the aforementioned studies, Doula [55] achieved the highest increase in capacity, as much as 2.8 times. However, the equilibrium pH values (6.53–8.37) were in the area where copper precipitation is most likely to occur. In this study, the pH was maintained in a range where no precipitation occurred, and the sorption capacity was found to be twice as high. Hence, the almost two-fold increase in the sorption capacity of FeZ is the result of a modification that caused an increase in the negative zeolite surface due to the incorporation of Fe species on the outer and inner surface of the zeolite particle [60]. Since the negative charge is partially compensated by the presence of the exchangeable alkali and alkaline earth cations, the results of interrelationship between in-going Cu(II) ions and out-going exchangeable ions (Na, K, Ca, and Mg) from the NZ and FeZ as a function of the initial Cu(II) concentrations are compared in Figure 9.
For both zeolites, an almost complete stoichiometric interrelationship between the amount of sorbed Cu(II) and released exchangeable cations was observed over the entire concentration range. This implies that ion exchange is the dominant mechanism of Cu(II) sorption on NZ and FeZ, with calcium being the main exchangeable cation for NZ while sodium for FeZ.
It was found that optimal pH is 3 ≤ pHo ≤ 5 for NZ, and for FeZ pHo = 3. Since the Cu(II) sorption experiment at different co was performed at the optimal pH, which was pHo = 4.83 for NZ, and pHo = 2.98 for FeZ, the competition effect between H+ and Cu2+ was not observed for NZ and FeZ. Thus, although the initial conditions in terms of pHo are different, the quantity of Cu(II) sorbed is higher on FeZ making it a more effective sorbent for Cu(II) ions. According to the above, the main mechanism of Cu(II) sorption on NZ and FeZ is ion exchange. As already mentioned, since both zeolites possess a negative charge in the range pHo = 2–12, the electrostatic attraction between the negatively charged surface and the dominantly present copper in the form of Cu2+ at pHo < 5 initiates the sorption of Cu(II). However, the ion exchange mechanism dominates as confirmed by the results shown in Figure 9. For the intention to understand the nature (physical or chemical) of Cu(II) sorption on both zeolites, the equilibrium data shown in Figure 8a were fitted in accordance with the Temkin and Dubinin–Radushkevich isotherm models and are explained in detail in the supplementary file (Table S1). These two isotherms clarified that the sorption is of a physical nature on both zeolites.
In order to fully understand the mechanism of Cu(II) sorption on zeolites, based on their chemical composition shown in Table 1, the amount of elements for the starting and copper-saturated zeolites was calculated and is shown in Table 4.
The results indicate that the modification caused an increase in iron content and that the dominant exchangeable cation is sodium for FeZ, and calcium for NZ. The amount of calcium decreased slightly for FeZ compared to NZ. This is a consequence of the second phase of modification in an alkaline medium, during which sodium is exchanged with calcium and its precipitation on the surface of the FeZ particle occurs [60]. This is supported by the results of almost unchanged calcium amounts in copper-saturated FeZ, since sodium is the main exchangeable cation. The decrease in the amount of sodium for FeZCu and calcium for NZCu is a consequence of ion exchange, and this is in accordance with the results shown in Figure 9. The amount of copper is 2.5 times higher for FeZCu compared to NZCu, which is in agreement with the twice higher sorption capacity of FeZ compared to NZ (Figure 8a). This suggests that Cu(II) is sorbed on active sites located on both the outer and inner surface of the zeolite particle.
In order to gain insight into the distribution of elements on the surface of copper-saturated NZ and FeZ, EDS was performed on four marked surfaces on both zeolites at a magnification of 35× as shown in Figure 10. The results of the elemental composition on the four marked surfaces on NZCu and FeZCu given in wt % are listed in Table 5 and Table 6.
The results indicate an almost uniform distribution of the analyzed elements on the four marked surfaces (spectra, Sp) for both samples. This reveals an even distribution of active centers, and thus an even binding of copper on the outer surface of both zeolites. The mean value of the copper mass percentage is three times higher on the outer surface of FeZ than on NZ, which once again confirms the more pronounced sorption ability of the modified sample. For saturated samples, among exchangeable cations, calcium dominates in both samples, which is in line with the chemical analysis results shown in Table 4. Namely, Ca is the dominant exchangeable cation of NZ, and due to its lower sorption capacity towards Cu(II) compared to FeZ, the amount of Ca decreased slightly on the NZCu surface. In the case of FeZ, Na is the main exchangeable cation. The amount of Na decreased significantly on the FeZCu surface due to the exchange with Cu(II), while the ion exchange of Cu(II) with Ca is significantly lower due to the formed Ca precipitate on the FeZ surface as a result of the modification. In order to confirm the uniform distribution of the detected elements (Table 5 and Table 6), additional SEM imaging was performed at a magnification of 1000 × with EDS and mapping analysis of the detected elements on the surface of NZCu and FeZCu, as shown in Figure 11 and Figure 12.
The results of the mapping analysis confirmed the uniform distribution of the detected elements on the surface of both samples (Figure 11 and Figure 12). This indicates that the exchangeable cations, and thus the active sites, are uniformly distributed on the surface of both samples. In addition, a uniform distribution of copper is observed on both zeolites, with the density of observed dots corresponding to Cu(II) being higher for FeZ compared to NZ. Ultimately, based on the results of monitoring the amount of released exchangeable cations and sorbed Cu2+ ions as well as the results of chemical composition and EDS, it can be concluded that the dominant mechanism is ion exchange initiated by electrostatic attraction between the negatively charged zeolite surface and positive Cu2+ ions.
Since both zeolites have a negative charge, this indicates a partial compensation of the active sites with alkaline and alkaline earth easily mobile cations and the possession of deprotonated, i.e., negatively charged active sites. As it was previously established that sorption is initiated by electrostatic attraction, an illustration of electrostatic attraction reactions is shown by Equation (23) for NZ and Equations (23) and (24) for FeZ as follows:
A l O 4 5 + C u 2 + A l O 4 5
2 F e O + C u 2 + F e O 2
Thus, the enhanced electrostatic attraction in the case of FeZ is due to the possession of (Fe–O) active sites in addition to those shown by Equation (23) as a result of modification. On the other hand, ion exchange takes place at active sites whose negative charge is compensated by the presence of exchangeable cations. The ion exchange reactions that take place on NZ are illustrated by Equations (25) and (26), and on FeZ by Equations (25)–(27).
( S O ) 2 X + C u 2 + ( S O ) 2 C u + X 2 +
2 S O Y + C u 2 + ( S O ) 2 C u + 2 Y +
2 F e O N a + C u 2 + ( F e O ) 2 C u + 2 N a +
where S denotes a Si or Al atom, X is Ca or Mg, and Y is Na or K.
Furthermore, in acidic media, the competition effect of Cu2+ with H+ ions hinder the sorption efficiency. Competitive reactions that take place on NZ are shown by Equations (28) and (29), and (28)–(30) in the case of FeZ as follows:
( S O ) 2 X + 2 H + 2 S O H + X 2 +
S O Y + H + S O H + Y +
F e O N a + H + F e O H + N a +
The formed hydroxylated neutral active sites shown by reactions (28)–(30) are converted into protonated positive active sites by further protonation reactions in acidic media as follows:
S O H + H + S O H 2 +
F e O H + H + F e O H 2 +
Protonated positively charged active sites affect electrostatic repulsions, which are more pronounced in more acidic media and cause a decrease in the zeolites sorption efficiency. Ultimately, this confirms once again that one of the key sorption parameters is determining the optimal pH.

3.2. Leaching Behavior of Copper-Saturated Zeolites

Leaching properties of copper-saturated zeolites, NZCu and FeZCu were tested according to the standard leaching method DIN 38414 S4 [62] in a wide pHo range of 2.06–12.09. The results of the percentage of Cu(II) leached from saturated zeolites are shown in Figure 13.
From Figure 13 it is clearly seen that in a wide pHo range, 4.04 ≤ pHo ≤ 10.06, Cu(II) leaching from both copper-saturated zeolites was not detected. Since in the mentioned pHo range the pHe values are around 7, Cu(II) leaching is probably prevented by the formation of Cu(OH)2 precipitates on the zeolite surface. This is supported by the fact that Cu(II) starts to precipitate at pHo = 6.5 according to the Cu(II) speciation diagram depending on pHo as shown in Figure 1. In an acidic medium, pHo ≤ 3.00, Cu(II) is leached from both zeolites in the amount of 17–36%. Namely, at pHo < 3.00, equilibrium pHe values are below pHe = 6, which indicates that insoluble Cu(OH)2 cannot be formed. As it is well known that metal solubility increases in acidic media, the results indicate that the percentage of leached Cu(II) is higher at lower pHo. Moreover, in acidic medium, enhanced protonation of the zeolite structure occurs with the replacement of H+ ions with sorbed Cu(II) ions, which is reflected in more pronounced leaching at pHo = 2.06. Moreover, in extremely acidic conditions, pHo = 2.06, partial dealumination of the zeolite structure cannot be ignored, which additionally contributes to increased leaching. Unlike extremely acidic conditions, in extremely alkaline conditions, pHo ≥ 11.18, leaching in the range of 1.7–3.8% is observed. Although pHe values indicate that Cu(II) leaching should be prevented by the formation of Cu(OH)2, the highly alkaline conditions caused desilication, i.e., the cracking of Si–O bonds, which resulted in Cu(II) leaching. Comparing the results of the percentage of Cu(II) leached from NZCu and FeZCu, a higher percentage of Cu(II) leached was recorded from NZCu even though twice as much Cu(II) was sorbed on FeZ. Therefore, in addition to superior sorption efficiency of FeZ, its higher ability to retain Cu(II) makes it an even more attractive sorbent for Cu(II) ions. The obtained results imply the possibility of using FeZ for in situ and ex situ remediation of copper-contaminated environments in the pH range 4.04 ≤ pHo ≤ 10.06. Moreover, both zeolites show the property of neutralizing the surrounding medium, which is evident from the pHe range of pHe = 7.5–8.5, for all tested pHo values.

4. Conclusions

This study provided a comparison of the sorption performance of Cu(II) on natural and Fe(III)-modified natural zeolite with optimization of sorption parameters to maximize zeolite utilization while achieving high Cu(II) removal efficiency. The removal of copper from suspension on both zeolites, especially FeZ, is most influenced by pH, followed by the S/L ratio. Optimization of sorption parameters demonstrated an optimal pHo range of 3–5 for NZ, while pHo = 3 for FeZ. For both zeolites, the optimal S/L was 10 g/L, the contact time was 600 min, and the concentration range in which Cu(II) removal was effective was up to 10 mmol/L for NZ and 12 mmol/L for FeZ. The kinetics of Cu(II) sorption on both zeolites involves a complex two-step mechanism, where intraparticle diffusion dominates as the rate-controlling sorption step with an almost negligible contribution of film diffusion in the initial sorption phase. Almost twice the sorption capacity of FeZ compared to NZ towards Cu(II) (0.271 mmol/g vs. 0.156 mmol/g) was confirmed experimentally and supported by determining the elemental composition, SEM/EDS and mapping analysis of saturated samples. Under optimal conditions, sorption was found to be driven by the electrostatic attraction of positive Cu2+ ions and the negatively charged zeolite structure, while the ion exchange mechanism was dominant. The results of the stoichiometric ratio of sorbed Cu(II) and released exchangeable cations under optimal sorption conditions confirmed ion exchange as the primary mechanism of Cu(II) sorption on NZ and FeZ. This was additionally supported by fitting the equilibrium data according to Temkin’s and Dubinin–Radushkevich isotherm, which confirmed the physical process of Cu(II) sorption on both zeolites. Finally, the leaching behavior of saturated NZCu and FeZCu highlight the ability to completely retain Cu(II) in the structure of both zeolites in a wide pHo range, 4.04 ≤ pHo ≤ 10.06. Ultimately, both zeolites contributed to Cu(II) removal, with FeZ standing out as highly promising sorbent for both in situ and ex situ remediation of copper-contaminated environments.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/pr13092672/s1: Figure S1: XRD spectra of NZ and FeZ; Figure S2: FTIR spectra of NZ and FeZ; Table S1: Isotherm models constants and error analysis for Cu(II) sorption onto NZ and FeZ [72,73].

Author Contributions

Conceptualization, investigation, methodology, formal analysis, writing—original draft preparation M.U.; experimental analysis, formal analysis J.M.; writing—review and editing, M.U., I.N., and J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia (grant number 451-03-136/2025-03/200023).

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fitzgerald, D.J. Safety guidelines for copper in water. Am. J. Clin. Nutr. 1998, 67, 1098S–1102S. [Google Scholar] [CrossRef]
  2. Karim, N. Copper and Human Health—A Review. J. Bahria Univ. Med. Dent. Coll. 2018, 8, 117–122. [Google Scholar] [CrossRef]
  3. Kumar Anant, J.; Inchulkar, S.R.; Bhagat, S. An overview of copper toxicity relevance to public health. Eur. J. Pharm. Med. Res. 2018, 5, 232–237. [Google Scholar]
  4. Rehman, M.; Liu, L.; Wang, Q.; Saleem, M.H.; Bashir, S.; Ullah, S.; Peng, D. Copper environmental toxicology, recent advances, and future outlook: A review. Environ. Sci. Pollut. Res. 2019, 26, 18003–18016. [Google Scholar] [CrossRef]
  5. Liu, Y.; Wang, H.; Cui, Y.; Chen, N. Removal of Copper Ions from Wastewater: A Review. Int. J. Environ. Res. Public Health 2023, 20, 3885. [Google Scholar] [CrossRef]
  6. Binesh, A.; Venkatachalam, K. Copper in Human Health and Disease: A Comprehensive Review. J. Biochem. Mol. Toxicol. 2024, 38, e70052. [Google Scholar] [CrossRef] [PubMed]
  7. Kumar, V.; Pandita, S.; Singh Sidhu, G.P.; Sharma, A.; Khanna, K.; Kaur, P.; Bali, A.S.; Setia, R. Copper bioavailability, uptake, toxicity and tolerance in plants: A comprehensive review. Chemosphere 2021, 262, 127810. [Google Scholar] [CrossRef]
  8. Sailer, J.; Nagel, J.; Akdogan, B.; Jauch, A.T.; Engler, J.; Knolle, P.A.; Zischka, H. Deadly excess copper. Redox Biol. 2024, 75, 103256. [Google Scholar] [CrossRef] [PubMed]
  9. Trumbo, H.P.; Yates, A.A.; Schlicker, S.; Poos, M. Dietary reference intakes: Vitamin A, vitamin K, arsenic, boron, chromium, copper, iodine, iron, manganese, molybdenum, nickel, silicon, vanadium, and zinc. J. Am. Diet Assoc. 2001, 101, 294–301. [Google Scholar] [CrossRef]
  10. WHO. Copper in drinking-water. In Background Document for Preparation of WHO Guidelines for Drinking-Water Quality; World Health Organization: Geneva, Switzerland, 2004. [Google Scholar]
  11. Perestrelo, A.P.; Miranda, G.; Goncalves, M.I.; Belino, C.; Ballesteros, R. Chronic copper sulfate poisoning. Eur. J. Case Rep. Intern. Med. 2021, 8, 002309. [Google Scholar] [CrossRef] [PubMed]
  12. Boveris, A.; Musacco-Sebio, R.; Ferrarotti, N.; Saporito-Magrina, C.; Torti, H.; Massot, F.; Repetto, M.G. The acute toxicity of iron and copper: Biomolecule oxidation and oxidative damage in rat liver. J. Inorg. Biochem. 2012, 116, 63–69. [Google Scholar] [CrossRef]
  13. Taylor, A.A.; Tsuji, J.S.; Garry, M.R.; McArdle, M.E.; Goodfellow, W.L., Jr.; Adams, W.J.; Menzie, C.A. Critical Review of Exposure and Effects: Implications for Setting Regulatory Health Criteria for Ingested Copper. Environ. Manag. 2020, 65, 131–159. [Google Scholar] [CrossRef]
  14. Bulcke, F.; Dringen, R.; Scheiber, I.F. Neurotoxicity of Copper. Adv. Neurobiol. 2017, 18, 313–343. [Google Scholar] [CrossRef]
  15. USEPA. National Primary Drinking Water Regulations; U.S. Environmental Protection Agency: Washington, DC, USA, 2017. Available online: https://www.epa.gov/ground-water-and-drinking-water/national-primary-drinking-water-regulations (accessed on 30 June 2025).
  16. WHO. Drinking Water Quality Guidelines Copper Guideline Value; World Health Organization: Geneva, Switzerland, 2022; Available online: https://www.who.int/publications/i/item/9789240045064 (accessed on 30 June 2025).
  17. Croatian Regulation on the Health Safety of Drinking Water. 2023. Available online: https://narodne-novine.nn.hr/clanci/sluzbeni/2023_06_64_1057.html (accessed on 30 June 2025).
  18. Pelosi, C.; Gavinelli, F.; Petit-dit-Grezeriat, L.; Serbource, C.; Schoffer, J.T.; Ginocchio, R.; Yáñez, C.; Concheri, G.; Rault, M.; van Gestel, C.A.M. Copper toxicity to earthworms: A comprehensive review and meta-analysis. Chemosphere 2024, 362, 142765. [Google Scholar] [CrossRef]
  19. Panagos, P.; Ballabio, C.; Lugato, E.; Jones, A.; Borrelli, P.; Scarpa, S.; Orgiazzi, A.; Montanarella, L. Potential Sources of Anthropogenic Copper Inputs to European Agricultural Soils. Sustainability 2018, 10, 2380. [Google Scholar] [CrossRef]
  20. Poggere, G.; Gasparin, A.; Zimmer Barbosa, J.; Wellington Melo, G.; Studart Corrêa, R.; Vargas Motta, A.S. Soil contamination by copper: Sources, ecological risks, and mitigation strategies in Brazil. J. Trace Elem. Miner. 2023, 4, 100059. [Google Scholar] [CrossRef]
  21. United Nations. The UN Sustainable Development Goals; United Nations: New York, NY, USA, 2015; Available online: http://www.un.org/sustainabledevelopment/summit/ (accessed on 8 March 2025).
  22. Ugrina, M.; Milojković, J. Advances in Wastewater Treatment, 2024. Energies 2024, 17, 1400. [Google Scholar] [CrossRef]
  23. Croatian Regulation on Wastewater Emission Limit Values. 2020. Available online: https://narodne-novine.nn.hr/clanci/sluzbeni/2020_03_26_622.html (accessed on 30 June 2025).
  24. Li, Q.; Wang, Y.; Chang, Z.; Kolaly, W.E.; Fan, F.; Li, M. Progress in the treatment of copper(II)-containing wastewater and wastewater treatment systems based on combined technologies: A review. J. Water Process Eng. 2024, 58, 104746. [Google Scholar] [CrossRef]
  25. Al-Saydeh, S.A.; El-Naas, M.H.; Zaidi, S.J. Copper removal from industrial wastewater: A comprehensive review. J. Ind. Eng. Chem. 2017, 56, 35–44. [Google Scholar] [CrossRef]
  26. Shrestha, R.; Ban, S.; Devkota, S.; Sharma, S.; Joshi, R.; Prasad Tiwari, A.; Yong Kim, H.; Kumar Joshi, M. Technological trends in heavy metals removal from industrial wastewater: A review. J. Environ. Chem. Eng. 2021, 9, 105688. [Google Scholar] [CrossRef]
  27. Bahmanzadegan, F.; Ghaemi, A. A comprehensive review on novel zeolite-based adsorbents for environmental pollutant. J. Hazard. Mater. Adv. 2025, 17, 100617. [Google Scholar] [CrossRef]
  28. Krstić, V.; Urošević, T.; Pešovski, B. A review on adsorbents for treatment of water and wastewaters containing copper ions. Chem. Eng. Sci. 2018, 192, 273–287. [Google Scholar] [CrossRef]
  29. Irannajad, M.; Haghighi, H.K. Removal of Heavy Metals from Polluted Solutions by Zeolitic Adsorbents: A Review. Environ. Process 2021, 8, 7–35. [Google Scholar] [CrossRef]
  30. Velarde, L.; Sadegh Nabavi, M.; Escalera, E.; Antti, M.L.; Akhtar, F. Adsorption of heavy metals on natural zeolites: A review. Chemosphere 2023, 328, 138508. [Google Scholar] [CrossRef]
  31. Abdelwahab, O.; Thabet, W.M. Natural zeolites and zeolite composites for heavy metal removal from contaminated water and their applications in aquaculture Systems: A review. Egypt. J. Aquat. Res. 2023, 49, 431–443. [Google Scholar] [CrossRef]
  32. Senthil Rathi, B.; Senthil Kumar, P.; Natanya Ida Susana, J.; Francia Virgin, J.; Dharani, R.; Sanjay, S.; Rangasamy, G. Recent research progress on the removal of heavy metals from wastewater using modified zeolites: A critical review. Des. Water Treat. 2024, 319, 100573. [Google Scholar] [CrossRef]
  33. Solihin, K.I.; Mardiana, S.; Rusli, H.; Kadja, G.T.M. Removal of Heavy Metal Ions Using Pristine and Functionalized Natural Zeolites. Indones. J. Chem. 2023, 23, 863–880. [Google Scholar] [CrossRef]
  34. Zhang, J.; Tang, X.; Yi, H.; Yu, Q.; Zhang, Y.; Wei, J.; Yuan, Y. Synthesis, characterization and application of Fe-zeolite: A review. Appl. Cat. A-Gen. 2022, 630, 118467. [Google Scholar] [CrossRef]
  35. Panayotova, I. Kinetics and thermodynamics of copper ions removal from wastewater by use of zeolite. Waste Manag. 2001, 21, 671–676. [Google Scholar] [CrossRef]
  36. Forughirad, A.; Bahrami, A.; Farhadi, K.; Fathi Azerbaijani, A.; Kazemi, F. Comparative of natural zeolite—Clinoptilolite elimination of metal ions/especially Cu (II) with D-Penicillamine from biological environments. MGPB 2021, 36, 11–19. [Google Scholar] [CrossRef]
  37. Zendelska, A.; Golomeova, M.; Blazev, K.; Krstev, B.; Golomeov, B.; Krstev, A. Adsorption of copper ions from aqueous solutions on natural zeolite. Environ. Prot. Eng. 2015, 41, 17–36. [Google Scholar] [CrossRef]
  38. Kabwadza-Corner, P.; Wazingwa Munthali, M.; Johan, E.; Matsue, N. Comparative study of copper adsorptivity and selectivity toward zeolites. Am. J. Analy. Chem. 2014, 5, 395–405. [Google Scholar] [CrossRef]
  39. Stojaković, Đ.; Milenković, J.; Daneu, N.; Rajić, N. A study of the removal of copper ions from aqueous solution using clinoptilolite from Serbia. Clays Clay Min. 2011, 59, 277–285. [Google Scholar] [CrossRef]
  40. Bakalár, T.; Pavolová, H.; Kyšel’a, K.; Hajduová, Z. Characterization of Cu(II) and Zn(II) sorption onto zeolite. Crystals 2022, 12, 908. [Google Scholar] [CrossRef]
  41. Zabukovec Logar, N.; Arčon, I.; Kovač, J.; Popova, M. Removal of Copper from Aqueous Solutions with Zeolites and Possible Treatment of Exhaust Materials. Chem. Ing. Tech. 2021, 93, 941–948. [Google Scholar] [CrossRef]
  42. Wilopo, W.; Nur Haryono, S.; Prakasa Eka Putra, D.; Warmada, I.W.; Hirajima, T. Copper (Cu2+) removal from water using natural zeolite from Gedangsari, Gunungkidul, Yogyakarta. J. Appl. Geol. 2010, 2, 117–120. [Google Scholar] [CrossRef]
  43. Doula, M.; Ioannou, A.; Dimirkou, A. Copper adsorption and Si, Al, Ca, Mg, and Na release from clinoptilolite. J. Colloid Interface Sci. 2002, 245, 237–250. [Google Scholar] [CrossRef] [PubMed]
  44. Pandová, I.; Panda, A.; Valíček, J.; Harničárová, M.; Kušnerová, M.; Palková, Z. Use of sorption of copper cations by clinoptilolite for wastewater treatment. Int. J. Environ. Res. Public Health 2018, 15, 1364. [Google Scholar] [CrossRef] [PubMed]
  45. Amin, N.A.M.; Kamarudzaman, A.N.; Rahmat, N.R.; Hassan, Z.; Zainon Najib, N.W.A.; Amirah, A.S.N.; Ab Jalil, M.F. Batch adsorption studies on copper removal from an aqueous solution using natural zeolite: Process optimization. IOP Conf. Ser. Earth Environ. Sci. 2024, 1369, 012011. [Google Scholar] [CrossRef]
  46. Kyzioł-Komosińska, J.; Rosik-Dulewska, C.; Franus, M.; Antoszczyszyn-Szpicka, P.; Czupioł, J.; Krzyżewska, I. Sorption capacities of natural and synthetic zeolites for Cu(II) ions. Pol. J. Environ. Stud. 2015, 24, 1111–1123. [Google Scholar] [CrossRef]
  47. Zinicovscaia, I.; Yushin, N.; Grozdov, D.; Safonov, A.; Ostovnaya, T.; Boldyrev, K.; Kryuchkov, D.; Popova, N. Bio-zeolite use for metal removal from copper-containing synthetic effluents. J. Environ. Health Sci. Eng. 2021, 19, 1383–1398. [Google Scholar] [CrossRef]
  48. Elboughdiri, N. The use of natural zeolite to remove heavy metals Cu (II), Pb (II) and Cd (II), from industrial wastewater. Cogent Eng. 2020, 7, 1782623. [Google Scholar] [CrossRef]
  49. Tekin, B.; Açikel, Ü. Intake of divalent copper and nickel onto natural zeolite from aqueous solutions: A study in mono- and dicomponent systems. Turk. J. Chem. 2022, 46, 1042–1054. [Google Scholar] [CrossRef]
  50. Abd-Elaziz, A.S.; Atress, M.S.; Haggag, E.A.; Soliman, K.G. Copper removal from aqueous solution of contaminated soil using natural zeolite. Zagazig J. Agric. Res. 2023, 50, 665–684. [Google Scholar] [CrossRef]
  51. Belova, T.P. Adsorption of heavy metal ions (Cu2+, Ni2+, Co2+ and Fe2+) from aqueous solutions by natural zeolite. Heliyon 2019, 5, e02320. [Google Scholar] [CrossRef] [PubMed]
  52. Milićević, S.; Povrenović, D.; Milošević, V.; Martinović, S. Predicting the copper adsorption capacity on different zeolites. J. Min. Metall. 2017, 53, 57–63. [Google Scholar] [CrossRef]
  53. Yulianti, E.; Yusniyyah, S.I.; Aini, N.; Khalifah, S.N.; Istighfarini, V.N. Removal of Cu and Pb from Wastewater Using Modified Natural Zeolite. In Advances in Social Science, Education and Humanities Research, Proceedings of the International Conference on Engineering, Technology and Social Science (ICONETOS 2020), Malang, Indonesia, 31 October 2020; Atlantis Press: Dordrecht, The Netherlands, 2020; Volume 529, pp. 363–369. [Google Scholar] [CrossRef]
  54. Kuldeyev, E.; Seitzhanova, M.; Tanirbergenova, S.; Tazhu, K.; Doszhanov, E.; Mansurov, Z.; Azat, S.; Nurlybaev, R.; Berndtsson, R. Modifying natural zeolites to improve heavy metal adsorption. Water 2023, 15, 2215. [Google Scholar] [CrossRef]
  55. Doula, M.K. Simultaneous removal of Cu, Mn and Zn from drinking water with the use of clinoptilolite and its Fe-modified form. Water Res. 2009, 43, 3659–3672. [Google Scholar] [CrossRef] [PubMed]
  56. Lipovský, M.; Sirotiak, M.; Soldán, M. Removal of copper from aqueous solutions by using natural and Fe-modified clinoptilolite. Res. Pap. MTF STU 2015, 23, 33–40. [Google Scholar] [CrossRef]
  57. Nguyen, T.C.; Loganathan, P.; Nguyen, T.V.; Vigneswaran, S.; Kandasamy, J.; Naidu, R. Simultaneous adsorption of Cd, Cr, Cu, Pb, and Zn by an iron-coated Australian zeolite in batch and fixed-bed column studies. Chem. Eng. J. 2015, 270, 39357–63404. [Google Scholar] [CrossRef]
  58. Han, R.; Zou, L.; Zhao, X.; Xu, Y.; Xu, F.; Li, Y.; Wang, Y. Characterization and properties of iron oxide-coated zeolite as adsorbent for removal of copper(II) from solution in fixed bed column. Chem. Eng. J. 2009, 149, 123–131. [Google Scholar] [CrossRef]
  59. DIN 66165-2; Particle Size Analysis-Sieving Analysis-Part 2: Procedure. Deutsches Institut für Normung: Berlin, Germany, 2016.
  60. Ugrina, M.; Vukojević Medvidović, N.; Daković, A. Characterization and environmental application of iron-modified zeolite from the Zlatokop deposit. Desalin. Water Treat. 2015, 53, 3557–3569. [Google Scholar] [CrossRef]
  61. Voinovitch, I.; Debrad-Guedon, J.; Louvrier, J. The Analysis of Silicates; Israel Program for Scientific Translations: Jerusalem, Israel, 1966; pp. 127–129. [Google Scholar]
  62. DIN 38414 S4; German Standard Procedure for Water, Wastewater and Sediment Testing–Sludge and Sediment. Determination of Leachability. Institut für Normung: Berlin, Germany, 1984.
  63. Kumar, D.; Gaur, J.P. Chemical reaction- and particle diffusion-based kinetic modeling of metal biosorption by a phormidium sp.-dominated cyanobacterial mat. Bioresour. Technol. 2011, 102, 633–640. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, J.; Guo, X. Adsorption kinetic models: Physical meanings, applications, and solving methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  65. 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]
  66. Chiron, N.; Guilet, R.; Deydier, E. Adsorption of Cu(II) and Pb(II) onto a grafted silica: Isotherms and kinetic models. Water Res. 2003, 37, 3079–3086. [Google Scholar] [CrossRef]
  67. Tosun, I. Ammonium removal from aqueous solutions by clinoptilolite: Determination of isotherm and thermodynamic parameters and comparison of kinetics by the Double exponential model and conventional kinetic models. Int. J. Environ. Res. Public Health 2012, 9, 970–984. [Google Scholar] [CrossRef]
  68. Helferich, F. Ion Exchange; Mc Graw-Hill Inc.: New York, NY, USA, 1962; pp. 250–322. [Google Scholar]
  69. Barnum, D.W. Hydrolysis of Cations. Formation Constants and Standard Free Energies of Formation of Hydroxy Complexes. Inorg. Chem. 1983, 22, 2297–2305. [Google Scholar] [CrossRef]
  70. Minceva, M.; Fajagar, R.; Markovska, L.; Meshko, V. Comparative study of Zn2+, Cd2+, and Pb2+ removal from water solution using natural clinoptilolitic zeolite and commercial granulated activated carbon. Equilibrium and adsorption. Sep. Sci. Technol. 2008, 43, 2117–2143. [Google Scholar] [CrossRef]
  71. Speight, J.G. Lange’s Handbook of Chemistry, 16th ed.; McGraw-Hill, Inc.: New York, NY, USA, 2005; p. 1334. Available online: https://www.labxing.com/files/lab_data/1340-1625805401-Qqoazhfj.pdf (accessed on 20 August 2025).
  72. Temkin, M.I.; Pyzhev, V. Kinetic of Ammonia Synthesis on Promoted Iron Catalyst. Acta Phys. Chem. 1940, 12, 327–356. [Google Scholar]
  73. Dubinin, M.M.; Radushkevich, L.V. The equation of the characteristic curve of activated charcoal. Dokl. Akad. Nauk Sssr. 1947, 55, 331–337. [Google Scholar]
Figure 1. Distribution of Cu(II) species in relation to pH.
Figure 1. Distribution of Cu(II) species in relation to pH.
Processes 13 02672 g001
Figure 2. (a) pHe vs. pHo after sorption of Cu(II) onto NZ and FeZ. (b) The effect of pHo on removal efficiency, αe of Cu(II) onto NZ and FeZ.
Figure 2. (a) pHe vs. pHo after sorption of Cu(II) onto NZ and FeZ. (b) The effect of pHo on removal efficiency, αe of Cu(II) onto NZ and FeZ.
Processes 13 02672 g002
Figure 3. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different pHo for initial concentration of 4.017 mmol Cu/L for NZ (left) and FeZ (right).
Figure 3. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different pHo for initial concentration of 4.017 mmol Cu/L for NZ (left) and FeZ (right).
Processes 13 02672 g003
Figure 4. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different pHo for initial concentration of 8.137 mmol Cu/L for NZ (left) and FeZ (right).
Figure 4. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different pHo for initial concentration of 8.137 mmol Cu/L for NZ (left) and FeZ (right).
Processes 13 02672 g004
Figure 5. Comparison of pHe with pHppt and removal efficiency, αe at different solid/liquid ratios for (a) NZ and (b) FeZ.
Figure 5. Comparison of pHe with pHppt and removal efficiency, αe at different solid/liquid ratios for (a) NZ and (b) FeZ.
Processes 13 02672 g005
Figure 6. Time-dependent changes in (a) the quantity of Cu(II) sorbed on NZ and FeZ and (b) the removal efficiency.
Figure 6. Time-dependent changes in (a) the quantity of Cu(II) sorbed on NZ and FeZ and (b) the removal efficiency.
Processes 13 02672 g006
Figure 7. Comparison of experimental data with (a) kinetic model curves (pseudo-first-order, pseudo-second-order, Elovich, Vermeulen’s approximation and double-exponential model), and (b) the Weber–Morris model.
Figure 7. Comparison of experimental data with (a) kinetic model curves (pseudo-first-order, pseudo-second-order, Elovich, Vermeulen’s approximation and double-exponential model), and (b) the Weber–Morris model.
Processes 13 02672 g007
Figure 8. (a) Effect of initial Cu(II) concentration on: quantity of Cu(II) sorbed per gram of NZ and FeZ and (b) Cu(II) removal efficiency.
Figure 8. (a) Effect of initial Cu(II) concentration on: quantity of Cu(II) sorbed per gram of NZ and FeZ and (b) Cu(II) removal efficiency.
Processes 13 02672 g008
Figure 9. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different initial Cu(II) concentration for NZ (left) and FeZ (right).
Figure 9. Interrelationship between in-going Cu(II) ions and out-going exchangeable cations as a function of different initial Cu(II) concentration for NZ (left) and FeZ (right).
Processes 13 02672 g009
Figure 10. Backscattered electrons mode image (BSE) with four marked surfaces (Spectra, Sp) on NZCu (left) and FeZCu (right) for EDS.
Figure 10. Backscattered electrons mode image (BSE) with four marked surfaces (Spectra, Sp) on NZCu (left) and FeZCu (right) for EDS.
Processes 13 02672 g010
Figure 11. SEM image, EDS of NZCu surface with corresponding mapping analysis.
Figure 11. SEM image, EDS of NZCu surface with corresponding mapping analysis.
Processes 13 02672 g011
Figure 12. SEM image, EDS of FeZCu surface with corresponding mapping analysis.
Figure 12. SEM image, EDS of FeZCu surface with corresponding mapping analysis.
Processes 13 02672 g012
Figure 13. Percentage of leached copper from NZCu and FeZCu, and changes in pHe as a function of pHo.
Figure 13. Percentage of leached copper from NZCu and FeZCu, and changes in pHe as a function of pHo.
Processes 13 02672 g013
Table 1. Chemical composition of starting and copper-saturated zeolites.
Table 1. Chemical composition of starting and copper-saturated zeolites.
SampleContent, wt. %
SiO2Al2O3Fe2O3Na2OK2OCaOMgOTiO2CuLoss of Ignition
NZ65.4014.002.161.501.103.560.850.32-11.09
NZCu65.9814.632.181.091.112.830.370.170.6111.03
FeZ62.8013.903.223.680.942.930.800.17-11.56
FeZCu64.5713.032.681.840.973.840.690.1611.5110.71
Table 2. Calculated kinetic model parameters and agreement indicators for Cu(II) sorption on NZ and FeZ.
Table 2. Calculated kinetic model parameters and agreement indicators for Cu(II) sorption on NZ and FeZ.
Kinetic Model/ParametersNZFeZ
co [mmol/L]10.06110.061
qexp [mmol/g]0.1640.279
Pseudo-first-order model (PFO)
qm [mmol/g]0.1530.264
k1 [1/min]0.0290.320
r20.9870.979
RMSE × 1037.33916.000
χ2 × 1047.7018.838
Pseudo-second-order model (PSO)
qm [mmol/g]0.1650.281
k2 [g/(mmol∙min)]0.2330.161
r20.9930.998
RMSE × 1035.1654.993
χ2 × 1051.9612.300
Elovich model
αE [mmol/(g·min)]0.0210.058
βE [g/mmol]39.21724.814
r20.9450.966
RMSE0.0150.020
χ2 × 1031.5372.638
Weber–Morris intra-particle
diffusion model
kWM1 [mmol/(g∙min1/2)]0.0190.029
DWM1 × 106 [cm2/min]1.5459.782
I0.0210.008
RC [%]13.0602.392
R20.9650.957
kWM2 [mmol/(g∙min1/2)]0.0020.003
DWM2 × 108 [cm2/min]1.0191.309
R20.9930.973
Double-exponential model
qm [mmol/g]0.1650.277
kB1 [1/min]0.0410.070
B1 [mmol/L]1.3251.594
kB2 × 103 [1/min]2.3357.374
B2 [mmol/L]0.3731.167
r1 × 103 [mmol/(g∙min)]5.4430.011
r2 × 105 [mmol/(g∙min1/2)]0.8718.605
r × 103 [mmol/(g∙min1/2)]5.5200.012
RF [%]78.03357.733
SF [%]21.96742.267
r20.9980.998
RMSE × 1032.7834.461
χ2 × 1071.3291.875
Vermeulen’s approximation
qm [mmol/g]0.1640.279
DV × 106 (cm2/min)1.2881.487
r20.9760.993
RMSE0.1240.217
χ2 × 1071.2011.898
Table 3. Comparison of sorption capacities of different natural and modified zeolites towards Cu(II).
Table 3. Comparison of sorption capacities of different natural and modified zeolites towards Cu(II).
Zeolite TypeZeolite OriginSorption
Capacity, mmol/g
References
Natural zeoliteBulgarian clinoptilolite0.150[37]
Japanese clinoptilolite0.074[38]
Slovak clinoptilolite0.037[40]
Japanese clinoptilolite0.064[42]
Ukrainian clinoptilolite0.337[46]
Greek clinoptilolite0.210[55]
Slovak clinoptilolite0.0058[56]
Australian heulandite0.063[57]
Chinese clinoptilolite0.061[58]
Serbian clinoptilolite0.156Current study
Modified zeoliteGreek clinoptilolite0.590[55]
Slovak clinoptilolite0.0061[56]
Australian heulandite0.077[57]
Chinese clinoptilolite0.081[58]
Serbian clinoptilolite0.271Current study
Table 4. Element quantity of starting and copper-saturated zeolites.
Table 4. Element quantity of starting and copper-saturated zeolites.
SampleElement Quantity, mmol/g
NaKMgCaSiAlOFeTiCu
NZ0.4840.2340.2110.63510.8882.74627.5900.2710.040-
NZCu0.3520.2360.0910.50510.9812.87027.6100.2730.0210.097
FeZ1.1870.2000.1980.52210.4522.72627.0600.4030.021-
FeZCu0.5940.2060.1710.68510.7462.55627.1300.3360.0200.238
Table 5. Semi-quantitative chemical composition (given in wt. %) of the four analyzed surfaces on the NZCu sample shown in Figure 10 (Spectra, Sp; analyzed with EDS).
Table 5. Semi-quantitative chemical composition (given in wt. %) of the four analyzed surfaces on the NZCu sample shown in Figure 10 (Spectra, Sp; analyzed with EDS).
ElementNaKMgCaSiAlOFeCu
Sp 1-0.760.581.9730.126.0158.890.710.96
Sp 20.600.940.552.0132.036.0055.551.281.04
Sp 3-0.960.572.0731.645.8157.021.000.93
Sp 40.531.010.592.2632.896.2154.800.441.27
Mean0.280.920.572.0831.676.0156.570.861.05
Table 6. Semi-quantitative chemical composition (given in wt. %) of the four analyzed surfaces on the FeZCu sample shown in Figure 10 (Spectra, Sp; analyzed with EDS).
Table 6. Semi-quantitative chemical composition (given in wt. %) of the four analyzed surfaces on the FeZCu sample shown in Figure 10 (Spectra, Sp; analyzed with EDS).
ElementNaKMgCaSiAlOFeCu
Sp 10.650.920.451.1628.975.9656.472.662.76
Sp 20.780.920.591.1730.287.0753.052.913.23
Sp 30.890.550.411.0730.115.5153.783.713.97
Sp 40.960.620.681.2029.466.7854.922.492.89
Mean0.820.750.531.1529.716.3354.562.943.21
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

Ugrina, M.; Nuić, I.; Milojković, J. Study on the Performance of Copper(II) Sorption Using Natural and Fe(III)-Modified Natural Zeolite–Sorption Parameters Optimization and Mechanism Elucidation. Processes 2025, 13, 2672. https://doi.org/10.3390/pr13092672

AMA Style

Ugrina M, Nuić I, Milojković J. Study on the Performance of Copper(II) Sorption Using Natural and Fe(III)-Modified Natural Zeolite–Sorption Parameters Optimization and Mechanism Elucidation. Processes. 2025; 13(9):2672. https://doi.org/10.3390/pr13092672

Chicago/Turabian Style

Ugrina, Marin, Ivona Nuić, and Jelena Milojković. 2025. "Study on the Performance of Copper(II) Sorption Using Natural and Fe(III)-Modified Natural Zeolite–Sorption Parameters Optimization and Mechanism Elucidation" Processes 13, no. 9: 2672. https://doi.org/10.3390/pr13092672

APA Style

Ugrina, M., Nuić, I., & Milojković, J. (2025). Study on the Performance of Copper(II) Sorption Using Natural and Fe(III)-Modified Natural Zeolite–Sorption Parameters Optimization and Mechanism Elucidation. Processes, 13(9), 2672. https://doi.org/10.3390/pr13092672

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