Next Article in Journal
Oceanography and Culture Shape Morphometric Divergence in Portunus pelagicus: Defining Actionable Management Units for Climate-Resilient Recreational Fisheries in Asia
Previous Article in Journal
A Method for Improving the Efficiency and Effectiveness of Automatic Image Analysis of Water Pipes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Abnormal Adsorption Characteristics of Copper, Zinc, and Manganese Ions on Natural Diatomite in a Liquid/Solid Heterogeneous System

1
School of Ecology and Environment, Central South University of Forestry and Technology, Changsha 410004, China
2
Hunan Provincial Key Laboratory of Wetland and Soil Ecological Remediation, Changsha 410004, China
3
School of Soil and Water Conservation, Central South University of Forestry and Technology, Changsha 410004, China
*
Authors to whom correspondence should be addressed.
These authors contribute equally to this paper.
Water 2025, 17(18), 2782; https://doi.org/10.3390/w17182782
Submission received: 1 August 2025 / Revised: 15 September 2025 / Accepted: 17 September 2025 / Published: 20 September 2025
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

In order to investigate the adsorption characteristics of Cu2+, Zn2+, and Mn2+ on natural diatomite in liquid/solid systems and to provide reliable theoretical support for the application of these materials, we conducted a series of adsorption studies. The results revealed a non-monotonic relationship between the adsorption capacity of natural diatomite and ion concentration. The maximum adsorption capacities for Cu2+, Zn2+, and Mn2+ were found to be 3.56, 6.23, and 3.82 mg·g−1, at concentrations of 200, 500, and 300 mg·L−1. Optimal adsorption conditions were determined by investigating environmental factors such as pH and temperature: pH 6, temperature 30 °C, and contact time 40 min. The adsorption kinetics were found to be in accordance with the pseudo-second-order model (R2 > 0.997). Fitting adsorption isotherms for Cu2+, Zn2+, and Mn2+ using various models revealed that the Langmuir (R2 > 0.993), Temkin (R2 > 0.953), and Freundlich (R2 > 0.997) models most accurately describe their adsorption behaviour. Thermodynamic analysis confirmed that adsorption is a spontaneous, endothermic, physical process (ΔG° < 0, ΔH° > 0, ΔS° > 0) and that the overall adsorption rate is limited by micropore adsorption. Consequently, natural diatomaceous earth can serve as an efficient, low-cost adsorbent for removing heavy metals from contaminated water.

1. Introduction

Heavy metal ions, such as Cu2+, Zn2+, and Mn2+, primarily originate from acidic mine drainage, industrial wastewater from metal electroplating, and agricultural runoff containing pesticides and fertilisers [1,2,3]. These ions can migrate with bodies of water, persist long-term at the sediment/water interface and accumulate through the food chain. This can lead to ecological toxicity and chronic health risks [4,5,6,7,8,9]. Traditional treatment technologies such as chemical precipitation [10,11], ion exchange [12,13], membrane separation [14,15,16], and biological treatment [17,18] face challenges such as high chemical consumption, membrane fouling, and secondary sludge generation [19]. In contrast, adsorption methods have gained attention due to their simplicity, low energy consumption, and ease of modularisation [20,21,22,23,24]. To date, researchers have developed high-performance adsorbents such as activated carbon [25], biochar [26], and metal/organic frameworks (MOFs) [27]. However, their complex synthesis routes and high costs make them difficult to deploy rapidly in remote mining areas or in the event of sudden water pollution.
Natural diatomite, a biogenic siliceous rock formed from diatom remains through geological diagenesis, features three-dimensional, interconnected, multipore channels (coexisting micropores, mesopores, and macropores); a high specific surface area (>100 m2·g−1); and abundant ≡Si–OH surface functional groups. These characteristics can significantly reduce material costs and environmental impact [28,29,30,31,32,33]. Nevertheless, existing studies have primarily focused on modifying or compositing diatomite to enhance its performance. The non-monotonic adsorption behaviour and underlying mechanisms of unmodified natural diatomite in complex liquid/solid systems remain poorly understood [34]. Therefore, this study uses untreated natural diatomite as an adsorbent to systematically evaluate its adsorption behaviour towards Cu2+, Zn2+, and Mn2+ in liquid/solid heterogeneous systems. This study focuses on elucidating the following: (1) the microscopic mechanism behind the decrease in adsorption capacity at high ion concentrations (>200 mg·L−1); (2) the controlling diffusion coupling on the overall rate; (3) the synergistic effects of multiple factors, such as temperature, pH, and the liquid/solid ratio. Using a combination of Langmuir–Temkin–Freundlich isotherms, pseudo-second-order kinetics and thermodynamic calculations, this study quantitatively determines the maximum adsorption site utilisation efficiency and energy barriers of natural diatomite in complex wastewater scenarios. This provides a theoretical benchmark for subsequent “precision functionalisation” or “framework-pore synergistic regulation” modification strategies. The research results will also provide a prototype of zero-modification, low-cost, scalable emergency water treatment technology for remote mining areas, rural regions, or sudden heavy metal leakage scenarios.

2. Materials and Methods

2.1. Experimental Material

The diatomite used in this work was purchased from Shengzhou, Zhejiang, and the basic physicochemical characteristics are shown in Table 1. Before repeatedly scrubbing it by using ultra-pure water, the impurities from the surface of diatomite shell were removed first. Then, the pretreated diatomite was put into the oven to dry at 105 °C. After grinding, the soil was screened using a 100-mesh sieve. Finally, the sieved soil was put into a dryer which was in the standby state.

2.2. Aqueous Solutions of Cu2+, Zn2+, Mn2+

The analytical-grade chemicals Cu(NO3)2·3H2O, Zn(NO3)2·6H2O, and MnSO4·H2O were used to prepare the concentrated standard solutions of Cu2+, Zn2+, and Mn2+, respectively. Stock solutions of 1000 mg·L−1 Cu2+, Zn2+, and Mn2+ were prepared by accurately weighing 3.8020 g of Cu(NO3)2·3H2O, 4.5495 g of Zn(NO3)2·6H2O, and 3.0750 g of MnSO4·H2O, and then dissolving them in ultra-pure water.

2.3. Adsorption Experiment

Different dosages of diatomite (corresponding to dosages from 1 to 20 g/L) were added to 100 mL stoppered flasks containing standard solutions of Cu2+, Zn2+, and Mn2+ with individual metal concentrations ranging from 10 to 500 mg·L−1. We adjusted the initial pH of the solution to 5.5 using HCl and NaOH, and maintained this value. We stirred the reaction solution at 150 rpm under temperatures ranging from 15 to 65 °C for 15 to 120 min. Once equilibrium had been reached, we separated the diatomite from the suspension liquid using medium-speed qualitative filter paper. We determined the residual concentrations of Cu2+, Zn2+, and Mn2+ ions in the aqueous phase in order to calculate the adsorption capacity of the diatomite ultimately. Each treatment group included three parallel samples. Unless otherwise specified, the standard conditions were as follows: diatomite dosage of 1 g∙L−1, temperature of 30 °C, ion concentration of 200 mg∙L−1, adsorption time of 2 h, and pH of 5.5. Throughout the experiment on the dosage of the adsorbent, the pH of the solution was continuously monitored and kept at a constant value of 5.5 in order to eliminate its influence as a variable.

2.4. Determination of the Concentration of Heavy Metal Ions

The concentration of Cu2+, Zn2+, and Mn2+ in water solutions was quantified using atomic absorption spectroscopy. The corresponding absorbance of a sample was determined with a flame atomic absorption spectrophotometer (Model AA-7002, Beijing East Instrument Analysis Co., Ltd., Beijing, China). With Cu2+, Zn2+, and Mn2+ standard solutions formulating a series of standard samples, the concentration of the heavy metal ions in the sample could be estimated with a calibration curve. The amount of metal adsorbed onto per unit mass of diatomite was calculated using the following equation:
q e   =   C 0     C e   ×   V W ,
where qe represents the diatomite equilibrium adsorption capacity, mg·g−1. C0 represents the initial heavy metal ion concentration in the liquid phase at the beginning of the adsorption, mg·L−1; Ce represents the equilibrium point of liquid phase concentration of heavy metal ions, mg·L−1; V indicates the volume of adsorption solution, mL; and W is the amount of diatomite, g.

2.5. Determination of Isoelectric Point

We weighed 0.1 g of diatomite and added it to 100 mL of a 0.05 mol∙L−1 KCl solution. We adjusted the pH to the desired value with hydrochloric acid or NaOH solution, and immediately put it into a micro-electrophoresis instrument (JS94H, Shanghai Zhongchen Digital Technology Equipment Co., Ltd., Shanghai, China) to determine the electrophoresis velocity v (L∙s−1) of the particles in the electric field and record the electric field potential gradient E (V∙cm−1) at the same time. This was calculated with the following equation:
ζ   = 4 π η v ε E × 300 2 ,
where η represents the liquid viscosity coefficient (Pa·s), and ε represents the water dielectric constant. Taking pH as abscissa and ζ potential (mV) as ordinate, the pH value at 0 ζ potential is obtained, i.e., the isoelectric point (IEP).

2.6. Characterization of Diatomite

The micromorphology of diatomite was studied using a scanning electron microscope (JSM-6380LV, Shimadzu, Kyoto, Japan). Phase composition was determined by an X-ray diffraction spectrometer (X'Pert Pro, Malvern Panalytical, Malvern, UK) and a Fourier transform infrared spectroscopy (FTIR-650, Shanghai Precision Instrumentation Co., Ltd., Shanghai, China).

3. Results

3.1. Characterization

3.1.1. Physiochemical Analysis of the Diatomite

The micromorphology of diatomite is closely related to the shape of the diatom shell, which is formed by diatomaceous remains after a long period of geological movement and diagenesis. The shells of the diatomite used in this study were mainly coronal and disc-shaped, with a small number of cylindrical shells, suggesting a main composition of Stephanodiscus- and Melosira-derived showed that the surface of the diatom shell was covered with impurities that may block the micropore (Figure 1).
As shown in the FTIR spectra (Figure 2a), the sharp peaks at 470.6 cm−1, 773.5 cm−1, 1028.1 cm−1, and 1093.6 cm−1 corresponded to the Si-O bond of amorphous SiO2. The Si-O bond of feldspar appeared near at 694.4 cm−1, which indicated that the main component of diatomite was amorphous SiO2, and had a small amount of feldspar. The band of the Al-O bond was at 1508.3 cm−1, and the band of free water, which was formed by a single water molecule and a single Si-O-H bond, was at 1630.1 cm−1. The band at 3419.8 cm−1 was bound water. Furthermore, the bands at 3570–3750 cm−1 could be attributed to the stretching vibration of the hydroxyl (-OH), of which the band at 3579.8 cm−1 was twin silicon hydroxyl (=Si(OH)2), at 3632.0 cm−1 was connected silicon hydroxyl (Si–O–H…OH–Si≡), and at 3749.6 cm−1 was a surface isolated silicon hydroxyl (≡Si–O–H).
The distinct peak intensities that appeared at 19.9°, 25.7°, and 35.6° are assigned to the (101), (100), and (110) crystal planes of quartz (SiO2, ICDD PDF #46−1045), respectively. The broad dispersion peaks observed in the 2θ range of 21° to 25° represent amorphous silica (SiO2), indicating that quartz coexists with amorphous silica in the main crystalline phase of diatomite. Additionally, corundum (Al2O3) was detected, with characteristic peaks at 35.5°, 49.1°, and 58.8° corresponding to the (104), (113), and (116) crystal planes (ICDD PDF #10−0173).

3.1.2. Surface Electrochemical Analysis

The diatomite was typically associated with a negative surface charge due to the dissociation of H+ from the siloxy group. Decreasing the pH caused protonation of the siloxy groups and an increase in the ζ potential. Further decreases in pH would start to protonate the siloxy groups to form oxonium ions, which would then rapidly increase the ζ potential (Figure A1). The isoelectric point of diatomite used in this paper was about 2.0, which is characteristic of siliceous materials and indicates a strong permanent negative surface charge at pH values above 2.0. This property is crucial for the adsorption of cationic heavy metal ions.

3.2. Factors Influencing the Cu2+, Zn2+, and Mn2+ Adsorption Process on Diatomite

3.2.1. Effect of Adsorbent Dosage (W0) on the Adsorption Process of Cu2+, Zn2+, and Mn2+ by Diatomite

As the dosage increased, the adsorption capacity of diatomite for Cu2+, Zn2+, and Mn2+ first increased slowly and then decreased rapidly (Figure 3). The adsorption capacity reached a maximum of 3.69, 23.50, and 5.08 mg∙g−1 for Cu2+, Zn2+, and Mn2+, respectively, at dosages of 8, 2, and 2 g∙L−1.

3.2.2. Effect of Initial Ion Concentration (C0) on the Adsorption Process of Cu2+, Zn2+, and Mn2+ by Diatomite

With the increase in the initial ion concentration, the adsorption of the heavy metal ions on diatomite initially increased, but when the initial ion concentration exceeded a certain threshold, the adsorption capacity decreased rapidly (Figure 4). When the concentration of Cu2+, Zn2+, and Mn2+ was 200, 500, and 300 mg∙L−1, respectively, the adsorption capacity reached a maximum value of 3.56, 6.23, and 3.82 mg∙g−1. What could be deduced from Figure 4 was that when C0 was low, there existed a near-linear correlation between qe and C0. At high C0 values, a plateau was reached, and further increases in C0 led to lower qe in the case of Cu2+, Mn2+, and Zn2+.

3.2.3. Effect of Initial pH Value of Solution on the Adsorption Process of Cu2+, Zn2+, and Mn2+ by Diatomite

It was noteworthy that the acidic condition (pH < 4) was not conducive to adsorption, and showed a poor adsorption property of the adsorbent, possibly through displacement of the adsorbed ion (Figure 5). When the pH was 3, the adsorption capacity of diatomite on each ion was only 1.28, 6.55, and 1.89 mg∙g−1. When the pH falls below the isoelectric point, protonation of the diatomite surface results in low adsorption capacity and a positive charge. This positive charge creates a strong electrostatic repulsion with cationic metal ions, which effectively prevents them from approaching the surface. Weak acidic conditions (4 < pH < 6) were beneficial for diatomite adsorption and the capacity increased significantly (Figure 5). At a pH of 6, the adsorption capacity for Cu2+, Zn2+, and Mn2+ reached 6.56, 7.60, and 2.59 mg∙g−1, respectively. Above pH 6, although there was a great increase in adsorption capacity, the increase appeared to be mainly due to the formation of metal oxides and hydroxides in the form of precipitation.
To gain a deeper mechanistic understanding of pH-dependent adsorption behaviour, the speciation of the solutions containing Cu2+, Zn2+, and Mn2+ was evaluated (Figure A2). At low pH (<4), the predominant species for all three metals are the free divalent cations (M2+), which compete with H+ ions for protonated, less negatively charged surface sites on diatomite. This results in low adsorption.
In the optimal pH range of 4–6, adsorption increases not only with the increasingly negative zeta potential of diatomite (Figure A1) but also with the onset of hydrolysis. For Cu2+ and Zn2+, the formation of monovalent hydroxo-complexes (e.g., CuOH+ and ZnOH+) becomes significant. These hydrolysed species often exhibit a higher adsorption affinity due to their lower hydration energy.
Crucially, Figure A2 confirms that the sharp rise in apparent “adsorption” at pH > 6 coincides with the thermodynamic prediction of precipitation for these metals. In this alkaline region, the formation of insoluble hydroxide precipitates (e.g., Cu(OH)2(s), Zn(OH)2(s), and Mn(OH)2(s)) becomes the dominant removal mechanism, surpassing surface adsorption itself.

3.2.4. Effect of Solution Temperature (T) on the Adsorption Process of Cu2+, Zn2+, and Mn2+ by Diatomite

The adsorption capacity exhibited a unique non-monotonic trend as the temperature increased, which is atypical for conventional adsorption systems (Figure 6) [4,30,33].

3.3. Adsorption Mechanism of Diatomite

3.3.1. Attributes of the Isotherm Adsorption

The Freundlich, Langmuir, Temkin, and D-R models were utilised, as these are widely employed isothermal adsorption models for the description of liquid/solid adsorption behaviour. The linear equations of these models are as follows:
logqe = logkF + (I/n)logCe,
1/qe = 1/(kL·qm)·1/Ce + 1/qm,
qe = A + BlnCe, A = (RT/b)·lnKTe, B = RT/b,
lnqe = lnqmβε2, ε = RT·ln(1 + 1/Ce),
where Ce and qe have the same meaning as mentioned before; qm represents the maximum adsorption capacity of diatomite, mg·g−1; R represents the thermodynamic constant, 8.314 J∙mol−1·K−1; T represents the solution temperature, K; kL represents the Langmuir model constant (L·mg−1); kF (mg·g−1) and n represent Freundlich model constants; KTe represents the equilibrium constants (L·mg−1); b represents the Temkin constant (J·mol−1); β reflects the adsorption activity coefficient (mol2·J−2); ε represents the Polanyi potential.
The experimental data of the isothermal adsorption were analysed and fitted using these models, and the results are shown in Table 2. Table 2 shows that the adsorption isotherms of Cu2+ were best described by the Langmuir model (R2 > 0.99), suggesting a monolayer adsorption mechanism onto a surface with homogeneous sites. In contrast, the adsorption of Zn2+ and Mn2+ showed a better fit with the Temkin and Freundlich models, respectively, indicating interactions with a more heterogeneous surface or that adsorption heat decreases with coverage.

3.3.2. Physicochemical Properties of Diatomite Adsorption Processes

The adsorption characteristics of Cu2+, Zn2+, and Mn2+ on diatomite were best described by the Langmuir, Temkin, and Freundlich isotherm models, respectively (Table 2). Consequently, the discussion of the adsorption mechanisms for each ion is based on the parameters derived from its respective best-fit model.
The Langmuir model assumes monolayer adsorption on a homogeneous surface with identical sites. The maximum monolayer adsorption capacity (qm) of diatomite for Cu2+ decreased slightly from 4.335 to 3.724 mg·g−1 as the temperature increased from 25 to 55 °C. This suggests that higher temperatures might inhibit the uptake of Cu2+ to some extent. The equilibrium constant kL, which relates to the affinity between the adsorbent and the adsorbate, increased with temperature (from 0.254 to 0.302 L·mg−1). This indicates that the adsorption process for Cu2+ becomes more favourable at higher temperatures [34]. The separation factor RL for all concentrations and temperatures ranged between 0.011 and 0.282—all values between 0 and 1—confirming that the adsorption of Cu2+ onto diatomite was favourable.
The Temkin model considers the effects of indirect adsorbate/adsorbate interactions, suggesting that the heat of adsorption decreases linearly with coverage due to these interactions. The Temkin constant B, which is related to the heat of adsorption, increased from 4371.23 to 4763.88 J·mol−1 as the temperature increased. This increase in the B value with temperature indicates either a stronger adsorbate/adsorbate interaction or a change in the adsorption mechanism at higher temperatures. The increase in the equilibrium binding constant (KTe) from 37.892 to 75.272 L·mg−1 also supports the finding that adsorption is enhanced at elevated temperatures.
The Freundlich model is empirical and applies to adsorption on heterogeneous surfaces. The Freundlich constant n is an indicator of adsorption intensity or surface heterogeneity. The value of n for Mn2+ was greater than 1 across all temperatures (ranging from 2.769 to 3.213), confirming that adsorption was favourable. The increase in the n value with temperature (from 2.769 at 25 °C to 3.213 at 40 °C) indicates that the adsorption process became more favourable, possibly due to the activation of more diverse adsorption sites, and that the surface heterogeneity might have been affected by temperature.
Despite the different best-fit models, the nature of the adsorption process can be inferred from the Freundlich exponent (n) and the Dubinin–Radushkevich (D-R) isotherm. The Freundlich adsorption constants (n) for the three ions were all greater than 1, indicating favourable adsorption. Furthermore, the mean free energy (E) derived from the D-R isotherm was below 8 kJ·mol−1 for all ions: Cu2+: 2.236–2.500 kJ·mol−1; Zn2+: 5.774–7.906 kJ/mol−1; Mn2+: 3.492–5.000 kJ·mol−1). As E values in the range of 1–8 kJ·mol−1 typically indicate physical adsorption [35], it can be concluded that adsorption of all three metal ions onto natural diatomite was predominantly physical, likely involving van der Waals forces or electrostatic interactions. Table 3 summarized the adsorption parameters for each ion.

3.3.3. Mass Transfer and Rate-Limiting Step Analysis of Diatomite

The pseudo-first-order kinetics model was used to explore the rate control step of the ion adsorption process. The linear equations of the liquid film diffusion, the particle diffusion, and adsorption reaction are as follows:
l n ( 1 F ) = k t,
1 3 ( 1 F ) 2 / 3 + 2 ( 1 F ) = k t,
1 ( 1 F ) 1 / 3 = k t,
where F = qt/qe is the adsorption fraction at t moment, and k is the rate constant.
The rate constants (k) for the liquid film diffusion, particle diffusion and adsorption reaction steps were calculated (Table A1). The adsorption reaction step values (Equation (9)) were the smallest of the three for each ion: Cu2+ (0.0100), Zn2+ (0.0039), and Mn2+ (0.0124). This suggests that the adsorption reaction within the pores was the rate-limiting step that controlled the overall kinetics of the process. Therefore, improving the internal pore structure of diatomite is crucial to enhancing its adsorption rate and capacity.

3.3.4. Adsorption Kinetics of Diatomite

The adsorption time is of consideration while using diatomite to treat wastewater, because it will not only affect the quality of the effluent but also affect the unit volume of treatment and the occupation of land. Therefore, adsorption kinetics is an important factor to consider. Figure 7 shows that the adsorption for Cu2+ and Zn2+ were initially very rapid, achieving half-maximum within 1–2 min, while it took 10 min for Mn2+ to reach half-maximum. The adsorption reached the plateau after 40, 20, and 20 min for Cu2+, Zn2+, and Mn2+, and the equilibrium adsorption densities were 2.62, 7.80, and 3.96 mg∙g−1, respectively. On the whole, the adsorption equilibrium could be reached within 40 min. To fit the kinetics data, first-order, second-order, Elovich, and double constant formulas were used, and the results are summarized in Table 4. Among these, the second-order fitting achieved the best fit (R2 = 0.995), suggesting an apparent overall second-order kinetics.

3.3.5. Adsorption Thermodynamics of Diatomite

The thermodynamic parameters (Gibbs free energy, enthalpy, and entropy) were studied by carrying out the adsorption at varying temperatures.
G ° = R T l n k d ,
l n k d = S ° R H ° R T ,
where R represents the thermodynamic constant (8.314 J·mol−1·K−1), T is the absolute temperature (K), and kd is the thermodynamic equilibrium constant (L·g−1).
Table 5 shows that for all ions, the ΔG° was less than 0, indicating the adsorption was spontaneous. With a rise in temperature, ΔG° became lower and the spontaneity was enhanced, which suggested that an increase in temperature was beneficial to the whole adsorption process. ΔH° > 0 indicated that the adsorption process was endothermic, which meant that heating was favourable for the adsorption process. ΔS° > 0, which indicated that the increase in chaotic degree occurred after the ions were adsorbed. This phenomenon could be explained by the ion exchange process, which caused the release of previously adsorbed ions, such as Potassium (K), Sodium (Na), Calcium (Ca), and Magnesium (Mg) ions.

4. Discussion

4.1. Adsorbent Dosage and Heavy Metal Adsorption Capacity Exhibit a Non-Linear Relationship

In the liquid/solid system, the dosage of the adsorbent (diatomite) and the concentration of the heavy metal ions are two factors which have mutual effects. Accordingly, based on the initial concentration of heavy metal ions, the optimum dosage of diatomite may need to be optimized for practical application. Previous studies showed that the amount of adsorbed heavy metal ions decreased with an increase in ion dosage [36,37]. The rationale was that with the increase in the dosage, diatomite could provide more space for the adsorption of heavy metal ions, more active groups, exchangeable ions, and negatively charged surfaces, while the amount of heavy metal ions in a solution was relatively constant. Thus, the adsorption capacity per unit area or mass would fall. However, our study discovered a different phenomenon. Under the condition of a constant initial ion concentration, when the adsorbent dosage was low, the adsorption capacity of diatomite for heavy metal ions increased with the dosage rising, and then reached a peak. After that, with the increase in dosage, the adsorption quantity decreased quickly.
The effect of the dosage on the adsorption of heavy metal ions on the diatomite was related to the liquid/solid ratio between the solution containing the heavy metal ions and diatomite. When the liquid/solid ratio was relatively large, that is, the diatomite dosage was relatively small, there was a surplus of heavy metal ions on the surface. Only a small proportion of heavy metal ions could be in contact with diatomite. This was because of the repulsion between the same kind of ions congested in the periphery of the diatomite particles, which expelled other heavy metal ions, so the adsorption capacity was small. The liquid/solid ratio decreased with the increase in the adsorbent dosage until it reached the optimum ratio. At this point, the number of adsorption sites and the amount of heavy metal ions were approximately those under which heavy metal ions could be evenly distributed on the surface and be fully absorbed, so the diatomite adsorption space was fully utilized, and the adsorption capacity reached its maximum value. As the liquid/solid ratio was further reduced, the adsorption sites provided by the diatomite were in surplus, and diatomite particles would agglomerate, resulting in uneven distribution, the loss of valid adsorption sites, and a reduction in the adsorption capacity.

4.2. High Initial Ion Concentration and Low Diatomite Dosage Will Inhibit the Adsorption Capacity of Diatomite

In principle, at constant temperature and pressure, the true thermodynamic equilibrium between a solution and a surface should shift towards higher surface coverage as the bulk concentration increases (Gibbs adsorption isotherm). However, the observed decrease in adsorption capacity at high concentrations suggests that kinetic and physical constraints take precedence over thermodynamic advantages in these conditions. We attribute this to agglomeration of diatomite particles induced by high ionic strength and unfavourable liquid/solid ratios. This drastically reduces the effective surface area and accessible pores, preventing the system from reaching its theoretical thermodynamic equilibrium. Consequently, the measured values represent a pseudo-equilibrium state that is constrained by kinetics. Figure A3 illustrates the influence of the liquid/solid ratio on the adsorption process.

4.3. Multiple Effects of Temperature on the Ability of Diatomite to Adsorb Heavy Metal Ions

Adsorption of heavy metal ions by clay minerals is generally divided into three processes [38,39]: (1) the ions diffuse from the solution to the adsorbent surface by the liquid film; (2) the ions diffuse from the surface to internal pore networks; (3) ions interact with the active groups. Figure A4 shows the adsorption process of metal ions by diatomite. Therefore, the influence of the temperature of the solution on the adsorption process is mainly manifested in the following three aspects: Firstly, the movement rate of the heavy metal ions has something to do with the solution temperature. Elevate temperatures can accelerate ion movement, increasing the surface of contact of the ion and diatomite, which is favourable for adsorption. Secondly, when the ion diffuses to the surface of diatomite by the liquid membrane, it needs to take off its own water membrane in order to enter the pore channel to react. This is a heat adsorption process, for which the increase in temperature is beneficial. Thirdly, when the reaction between the active group and the metal ion entering the pore is a spontaneous process, that is, ΔG < 0, upon adsorption, the metal ions, which were freely moving in a three-dimensional aqueous solution, become confined to a two-dimensional interface. This loss of translational freedom typically results in a decrease in entropy (ΔS < 0). According to the thermodynamic equation ΔG = ΔH − TΔS, for the spontaneous process (ΔG < 0), the enthalpy change (ΔH) must be negative and its magnitude must exceed |TΔS|. This confirms that the overall adsorption process is exothermic (ΔH < 0). In accordance with the thermodynamic equation ΔH = ΔG + TΔS, we could know that ΔH < 0, which showed that the final adsorption reaction was an exothermic reaction. Elevating the temperature was not conducive to the adsorption process. Since the effect of the solution temperature on the adsorption of heavy metal ions by diatomite was multifaceted, the regularity was not obvious. But it was certain that a too high or too low temperature was not conducive to carrying out adsorption.

4.4. Future Perspectives

This study provides a foundational understanding of the adsorption kinetics on natural diatomite. In order to further deepen the mechanistic insights, it is recommended that future investigations focus on deconvoluting the complex mass transfer processes. While the pseudo-second-order model offered an excellent fit for the overall kinetics, more sophisticated models are required to precisely identify the rate-limiting step(s). Subsequent research will incorporate the Boyd model to clearly differentiate between boundary layer diffusion and intra-particle diffusion control. Furthermore, a more nuanced interpretation of the Weber–Morris model, encompassing the analysis of multi-linear segments and their respective intercepts, is imperative for comprehending the diffusion pathways within the macro-, meso-, and micropores of diatomite. The integration of these kinetic analyses with in situ characterisation techniques is expected to yield a more comprehensive model of the adsorption process.

5. Conclusions

In summary, we performed adsorption studies using diatomite for three Group 7 ions that were environmentally important. Ion concentration and the adsorption capacity of natural diatomite dosage had a non-monotonic relationship. In addition, environmental factors such as pH and temperature were investigated. We found that pH within the range of 4–6 can maximize the adsorption efficiency of diatomite, while temperature has no significant effect on the adsorption of heavy metal ions by diatomite. We also studied the adsorption process using Langmuir, Temkin, and Freundlich isothermal models. The adsorption process was mainly a physical process. The overall adsorption rate is controlled by a combination of film diffusion, intra-particle diffusion, and surface complexation. The adsorption reactions were pseudo-second order, and were spontaneous, endothermic, and with increased entropy. This study confirms the potential of natural diatomite as an effective heavy metal adsorbent in environmental remediation, especially treating polluted water bodies containing Cu2+, Zn2+, and Mn2+. This finding guides the design and implementation of diatomite-based strategies for polluted water body remediation. Future work should focus on the modification, regeneration, and scaling-up of diatomite to maximize its utilization in environmental remediation.

Author Contributions

Conceptualisation, Q.H.; methodology, J.H. and J.W. (Jiacheng Wang); validation, W.L.; formal analysis, Y.X.; resources, M.L.; writing—original draft preparation, J.W. (Jieying Wang); writing—review and editing, Q.H. and H.H.; visualization, X.H.; supervision, J.Z. and M.L.; project administration, J.Z.; funding acquisition, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22276220), the Natural Science Foundation of Hunan Province (2023JJ30989), and the Hunan Ecology and Environment Department (HBKYXM-2023004).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Figure A1. ζ-pH curve of natural diatomite used in this study.
Figure A1. ζ-pH curve of natural diatomite used in this study.
Water 17 02782 g0a1
Figure A2. Thermodynamic equilibrium speciation of (a) Cu(II), (b) Zn(II), and (c) Mn(II) as a function of pH.
Figure A2. Thermodynamic equilibrium speciation of (a) Cu(II), (b) Zn(II), and (c) Mn(II) as a function of pH.
Water 17 02782 g0a2
Figure A3. Effect of liquid/solid ratio on adsorption process: (a) liquid/solid ratio is relatively large; (b) best ratio; (c) ratio is relatively small.
Figure A3. Effect of liquid/solid ratio on adsorption process: (a) liquid/solid ratio is relatively large; (b) best ratio; (c) ratio is relatively small.
Water 17 02782 g0a3
Figure A4. Adsorption process of heavy metal ions on diatomite: (a) the ions diffuse from the solution to the diatomite surface by the liquid membrane; (b) the ions remove the water film carried by itself; (c) the ions react with the active groups on adsorption.
Figure A4. Adsorption process of heavy metal ions on diatomite: (a) the ions diffuse from the solution to the diatomite surface by the liquid membrane; (b) the ions remove the water film carried by itself; (c) the ions react with the active groups on adsorption.
Water 17 02782 g0a4
Table A1. Value of velocity constant k of ion adsorption process control step (T = 30 °C).
Table A1. Value of velocity constant k of ion adsorption process control step (T = 30 °C).
Control ProcedureRate Constant k
Cu2+Zn2+Mn2+
Liquid Film Diffusion0.06010.06190.1206
Particle Diffusion0.01070.00730.0153
Adsorption Reaction0.01000.00390.0124

References

  1. Jiang, Y.; Di, J.; Gao, M.; Dong, Y. Study on the new slow-release carbon source biochemistry and its improvement of SRB on the acid mine drainage treatment. J. Environ. Manag. 2024, 370, 122860. [Google Scholar] [CrossRef] [PubMed]
  2. Tang, X.; Xu, T.; Hu, S.; Liu, K.; Zeng, Z.; Wu, Q.; Zhou, Y.; He, M.; Cao, X.; Yu, G. Design and implementation of control system for electroplating wastewater treatment by photovoltaic energy sinusoidal alternating current coagulation technology. Process Saf. Environ. Prot. 2024, 182, 405–415. [Google Scholar] [CrossRef]
  3. Edward, K.; Yuvaraj, K.M.; Kapoor, A. Chitosan-blended membranes for heavy metal removal from aqueous systems: A review of synthesis, separation mechanism, and performance. Int. J. Biol. Macromol. 2024, 279, 134996. [Google Scholar] [CrossRef]
  4. Biswal, B.K.; Balasubramanian, R. Use of biochar as a low-cost adsorbent for removal of heavy metals from water and wastewater: A review. J. Environ. Chem. Eng. 2023, 11, 110986. [Google Scholar] [CrossRef]
  5. Luo, K.H.; Tu, H.P.; Chang, H.C.; Yang, C.C.; Weng, W.C.; Chen, T.H.; Yang, C.H.; Chuang, H.Y. Mediation analysis for TNF-α as a mediator between multiple metal exposure and kidney function. Ecotoxicol. Environ. Saf. 2024, 283, 116837. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, X.; Han, X.; Guo, S.; Ma, Y.; Zhang, Y. Associations between patterns of blood heavy metal exposure and health outcomes: Insights from NHANES 2011–2016. BMC Public Health 2024, 24, 558. [Google Scholar] [CrossRef]
  7. Fontes, A.; Pierson, H.; Bierła, J.B.; Eberhagen, C.; Kinschel, J.; Akdogan, B.; Rieder, T.; Sailer, J.; Reinold, Q.; Cielecka-Kuszyk, J.; et al. Copper impairs the intestinal barrier integrity in Wilson disease. Metabolism 2024, 158, 155973. [Google Scholar] [CrossRef] [PubMed]
  8. Kawahara, M.; Tanaka, K.; Kato-Negishi, M. Zinc, Copper, and Calcium: A Triangle in the Synapse for the Pathogenesis of Vascular-Type Senile Dementia. Biomolecules 2024, 14, 773. [Google Scholar] [CrossRef]
  9. Qi, Y.; Si, H.; Jin, X.; Guo, Y.; Xia, J.; He, J.; Deng, X.; Deng, M.; Yao, W.; Hao, C. Changes in serum TIM-3 and complement C3 expression in workers due to Mn exposure. Front. Public Health 2023, 11, 1289838. [Google Scholar] [CrossRef]
  10. Ko, Y.G. Hybrid method integrating adsorption and chemical precipitation of heavy metal ions on polymeric fiber surfaces for highly efficient water purification. Chemosphere 2024, 363, 142909. [Google Scholar] [CrossRef]
  11. Jin, H.; Yu, Y.; Chen, X. Electrochemical precipitation for water and wastewater treatment. Process Saf. Environ. Prot. 2024, 184, 1011–1016. [Google Scholar] [CrossRef]
  12. Kim, J.G.; Ku, J.; Jung, J.; Park, Y.S.; Choi, G.H.; Hwang, S.S.; Lee, J.H.; Lee, A.S. Ion-exchangeable and sorptive reinforced membranes for efficient electrochemical removal of heavy metal ions in wastewater. J. Clean. Prod. 2024, 438, 140779. [Google Scholar] [CrossRef]
  13. Moore, R.G.; Crawford, J.M. Isolated and Paired Metal Sites in Zeolites Using Solid-State Ion Exchange. Angew. Chem. Int. Ed. 2025, 64, e202505186. [Google Scholar] [CrossRef] [PubMed]
  14. Wanjiya, M.; Zhang, J.C.; Wu, B.; Yin, M.J.; An, Q.F. Nanofiltration membranes for sustainable removal of heavy metal ions from polluted water: A review and future perspective. Desalination 2024, 578, 117441. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Sun, S.; Wang, M.; Shen, Q.; Cong, S.; Zhao, Y.; Peng, J.; Pang, H. Metal-organic framework (MOF) membranes for Mg2+/Li+ separation. Sep. Purif. Technol. 2025, 374, 133737. [Google Scholar] [CrossRef]
  16. Cai, Q.; Shi, C.; Cao, Z.; Li, Z.; Zhao, H.P.; Yuan, S. Electrokinetic bioremediation of trichloroethylene and Cr/As co-contaminated soils with elevated sulfate. J. Hazard. Mater. 2024, 468, 133761. [Google Scholar] [CrossRef]
  17. Taieb, I.; Ben Younes, S.; Messai, B.; Mnif, S.; Mzoughi, R.; Bakhrouf, A.; Jabeur, C.; Serrano, J.A.A.; Ellafi, A. Isolation, characterization and identification of a new lysinibacillus fusiformis strain zc from metlaoui phosphate laundries wastewater: Bio-treatment assays. Sustainability 2021, 13, 10072. [Google Scholar] [CrossRef]
  18. Liyanaarachchi, H.; Thambiliyagodage, C.; Lokuge, H.; Vigneswaran, S. Kinetics and thermodynamics study of methylene blue adsorption to sucrose-and urea-derived nitrogen-enriched, hierarchically porous carbon activated by KOH and H3PO4. ACS Omega 2023, 8, 16158–16173. [Google Scholar] [CrossRef]
  19. Fan, W.; Li, S.; Yuan, Q.; Wu, P.; Zhang, X. Light-driven in-situ synthesis of nano-sulfur and graphene oxide composites for efficient removal of heavy metal ions. J. Hazard. Mater. 2025, 487, 137079. [Google Scholar] [CrossRef]
  20. Abu Elella, M.H.; Abdallah, H.M.; Ali, E.A.; Makhado, E.; Abd El-Ghany, N.A. Recent developments in conductive polysaccharide adsorbent formulations for environmental remediation: A review. Int. J. Biol. Macromol. 2025, 304, 140915. [Google Scholar] [CrossRef]
  21. Chen, J.; Bao, C.; Chen, M.; Wang, Y.; Xu, X.; Yang, T.; Shang, C.; Zhang, Q. β-cyclodextrin-scaffolded crosslinked poly (ionic liquid) s for ultrafast removal of multiple pollutants: Insight into adsorption performance and mechanism. Chem. Eng. J. 2023, 464, 142526. [Google Scholar] [CrossRef]
  22. Umejuru, E.C.; Mashifana, T.; Kandjou, V.; Amani-Beni, M.; Sadeghifar, H.; Fayazi, M.; Karimi-Maleh, H.; Sithole, N.T. Application of zeolite based nanocomposites for wastewater remediation: Evaluating newer and environmentally benign approaches. Environ. Res. 2023, 231, 116073. [Google Scholar] [CrossRef]
  23. Zhao, C.; Liu, G.; Tan, Q.; Gao, M.; Chen, G.; Huang, X.; Xu, X.; Li, L.; Wang, J.; Zhang, Y.; et al. Polysaccharide-based biopolymer hydrogels for heavy metal detection and adsorption. J. Adv. Res. 2023, 44, 53–70. [Google Scholar] [CrossRef]
  24. Kong, Q.; Zhang, X.; Ma, K.; Gong, Y.; Peng, H.; Qi, W. PEI-modified chitosan/activated carbon composites for Cu (II) removal from simulated pyrophosphate plating rinsing wastewater. Int. J. Biol. Macromol. 2023, 251, 126429. [Google Scholar] [CrossRef] [PubMed]
  25. Ravindiran, G.; Rajamanickam, S.; Ramalingam, M.; Hayder, G.; Sathaiah, B.K.; Gaddam, M.K.R.; Muniasamy, S.K.; Arunkumar, P. Conversion of seaweed waste to biochar for the removal of heavy metal ions from aqueous solution: A sustainable method to address eutrophication problem in water bodies. Environ. Res. 2024, 241, 117551. [Google Scholar]
  26. Guo, X.; Feng, S.; Peng, Y.; Li, B.; Zhao, J.; Xu, H.; Meng, X.; Zhai, W.; Pang, H. Emerging insights into the application of metal-organic framework (MOF)-based materials for electrochemical heavy metal ion detection. Food Chem. 2025, 463, 141387. [Google Scholar] [PubMed]
  27. Zahajská, P.; Opfergelt, S.; Fritz, S.C.; Stadmark, J.; Conley, D.J. What is diatomite? Quat. Res. 2020, 96, 48–52. [Google Scholar] [CrossRef]
  28. Mazkad, D.; El Idrissi, A.; Marrane, S.E.; Lazar, N.E.; El Ouardi, M.; Dardari, O.; Channab, B.E.; Layachi, O.A.; Farsad, S.; Baqais, A.; et al. An innovative diatomite-polypyrrole composite for highly efficient Cr (VI) removal through optimized adsorption via surface response methodology. Colloids Surf. A Physicochem. Eng. Asp. 2024, 685, 133172. [Google Scholar]
  29. Tang, X.; Hu, G.; Chen, Z.; Qian, L.; He, C.; Chen, T.; Gao, J.; Zhao, Y.; Han, X. High adsorption of Cu(II) in novel thiacalix[4]arene/acrylic polymer/diatomite fast fabricated by electron beam irradiation: Controllable microstructures, adsorption performance and mechanism. Chem. Eng. J. 2024, 495, 153090. [Google Scholar] [CrossRef]
  30. Youcef, H.; Al-Senani, G.M.; Al-Qahtani, S.D.; Kiari, M.; Sabantina, L.; Benyoucef, A.D.; Hadjel, M. Highly efficient removal of Cadmium (II) ions from aqueous solutions using an eco-friendly waste-derived carbon-based from date seed and Mn-modified diatomite: Equilibrium, kinetics and characterization studies. Colloids Surf. A Physicochem. Eng. Asp. 2025, 724, 137439. [Google Scholar] [CrossRef]
  31. Saidi, M.; Reguig, B.A.; El Amine Monir, M.; Althagafi, T.M.; Fatmi, M.; Remil, A.; Zehhaf, A.; Ghebouli, M.A. Kinetics thermodynamics and adsorption study of raw treated diatomite as a sustainable adsorbent for crystal violet dye. Sci. Rep. 2025, 15, 21991. [Google Scholar] [CrossRef]
  32. Wang, H.; Deng, Q.; Niu, Y.; Wang, X.; Hu, W.; Chen, L. Investigation of mechanical properties and solidification/stabilization mechanisms of Pb2+-contaminated red clay by diatomite-fly-ash-based geopolymer. Constr. Build. Mater. 2025, 492, 142843. [Google Scholar] [CrossRef]
  33. Li, B.; Li, M.; Zhang, P.; Pan, Y.; Huang, Z.; Xiao, H. Remediation of Cd (II) ions in aqueous and soil phases using novel porous cellulose/chitosan composite spheres loaded with zero-valent iron nanoparticles. React. Funct. Polym. 2022, 173, 105210. [Google Scholar] [CrossRef]
  34. e Silva, D.C.T.; da Silva, M.L.M.; de Farias, P.H.M.; Galvao, C.C.; dos Santos Costa, E.M.; Melo, R.A.; Medeiros, E.B.M.; de Lima Filho, N.M. Synthesis and characterization of polyaniline, sucrose octaacetate and chitosan blend for removal of remazol black by adsorption: Equilibrium, kinetics, and regeneration. Int. J. Biol. Macromol. 2025, 289, 138863. [Google Scholar]
  35. Pedroza, R.H.; David, C.; Lodeiro, P.; Rey-Castro, C. Interactions of humic acid with pristine poly (lactic acid) microplastics in aqueous solution. Sci. Total Environ. 2024, 908, 168366. [Google Scholar]
  36. Jing, Q.; Wang, Y.; Chai, L.; Tang, C.; Huang, X.; Guo, H.; Wang, W.; You, W. Adsorption of copper ions on porous ceramsite prepared by diatomite and tungsten residue. Trans. Nonferrous Met. Soc. China 2018, 28, 1053–1060. [Google Scholar] [CrossRef]
  37. Zhu, J.; Lei, M.J.; Wang, P.; Zhang, W.L.; Chen, Y. Preparation of poly-hydroxy-aluminum pillared diatomite and characteristics of Cu2+, Zn2+ adsorption on the pillar in aqueous solutions. Huan Jing Ke Xue Huanjing Kexue 2016, 37, 3177–3185. [Google Scholar]
  38. Yang, X.; Zhou, Y.; Hu, J.; Zheng, Q.; Zhao, Y.; Lv, G.; Liao, L. Clay minerals and clay-based materials for heavy metals pollution control. Sci. Total Environ. 2024, 954, 176193. [Google Scholar] [CrossRef]
  39. Alshameri, A.; He, H.; Zhu, J.; Xi, Y.; Zhu, R.; Ma, L.; Tao, Q. Adsorption of ammonium by different natural clay minerals: Characterization, kinetics and adsorption isotherms. Appl. Clay Sci. 2018, 159, 83–93. [Google Scholar] [CrossRef]
Figure 1. SEM images of natural diatomite: (a) 4500×; (b) 2500×; (c) 2500×; and (d) 9000×.
Figure 1. SEM images of natural diatomite: (a) 4500×; (b) 2500×; (c) 2500×; and (d) 9000×.
Water 17 02782 g001
Figure 2. (a) FTIR spectra of natural diatomite used in this study; (b) XRD spectra of natural diatomite used in this study.
Figure 2. (a) FTIR spectra of natural diatomite used in this study; (b) XRD spectra of natural diatomite used in this study.
Water 17 02782 g002
Figure 3. Effect of diatomite adding quantity: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Figure 3. Effect of diatomite adding quantity: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Water 17 02782 g003
Figure 4. Effect of initial ion concentration: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Figure 4. Effect of initial ion concentration: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Water 17 02782 g004
Figure 5. Effect of solution initial pH: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Figure 5. Effect of solution initial pH: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Water 17 02782 g005
Figure 6. Effect of solution temperature: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Figure 6. Effect of solution temperature: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Water 17 02782 g006
Figure 7. Effect of contact time: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Figure 7. Effect of contact time: (a) Cu2+; (b) Zn2+; (c) Mn2+.
Water 17 02782 g007
Table 1. General physical and chemical properties of natural diatomite.
Table 1. General physical and chemical properties of natural diatomite.
Physical PropertyChemical Composition
Bulk Density (g/cm3)Specific Surface Area (m2/g)Particle Size (M)Pore Size (nm)Content of SiO2 (%)Content of Al2O3 (%)Content of Fe2O3 (%)LOI (%)
0.5758.007.530–40064.8016.402.913.10
Table 2. Parameters for the four evaluated adsorption isotherm models.
Table 2. Parameters for the four evaluated adsorption isotherm models.
IonsTemperatureFreundlich ModelLangmuir ModelTemkin ModelD-R Model
R2kFnR2kLqmR2KTeBR2Βqm
°C mg∙g−1 L∙mg−1mg∙g−1 L∙mg−1J∙mol−1 mol2∙J−2mg∙g−1
Cu2+250.9280.7812.5060.9930.2544.3350.9833.8923514.920.8572.0 × 10−72.949
400.9140.7752.6720.9930.2774.0100.9954.7294022.900.8741.8 × 10−72.799
550.8910.7752.8480.9940.3023.7240.9956.0394607.060.8901.6 × 10−72.654
Zn2+250.9211.3993.2480.8493.9463.0310.95337.8924371.230.8173.0 × 10−83.139
400.9111.4673.3430.8094.1263.1110.96845.8534424.2040.8452.7 × 10−83.290
550.9331.5393.5320.7969.4242.7930.93775.2724763.880.7991.6 × 10−83.130
Mn2+250.9970.5992.7690.8761.2261.8440.9124.3795039.290.6148.2 × 10−82.078
400.9900.6503.2130.8382.8091.6420.9189.1956495.980.6074.0 × 10−81.899
550.9820.5423.1130.6371.3601.5710.9375.9156937.870.6216.5 × 10−81.724
Table 3. Summary of adsorption parameters for Cu2+, Zn2+, and Mn2+.
Table 3. Summary of adsorption parameters for Cu2+, Zn2+, and Mn2+.
IonsTemperaturenKLRL maxRL minΒE
°C mol2∙J−2kJ∙mol−1
Cu2+252.5060.2540.2820.0132.0 × 10−72.236
402.6720.2770.2650.0121.8 × 10−72.357
552.8480.3020.2490.0111.6 × 10−72.500
Zn2+253.2483.9460.0250.0013.0 × 10−85.774
403.3434.1260.0240.0012.7 × 10−86.086
553.5329.4240.0100.0001.6 × 10−87.906
Mn2+252.7691.2260.0750.0038.2 × 10−83.492
403.2132.8090.0340.0014.0 × 10−85.000
553.1131.3600.0680.0026.5 × 10−83.922
Table 4. Estimated parameters of different kinetic equations (T = 30 °C).
Table 4. Estimated parameters of different kinetic equations (T = 30 °C).
Models and Linear ExpressionsParametersCu2+Zn2+Mn2+
First-order kinetics equation
ln(qeqt) = lnqek1t
R20.8150.8090.652
k10.0390.0300.049
qe0.510 (2.635) 10.362 (7.810) 10.283 (3.971) 1
Second-order kinetics equation
t/qt = 1/k2·1/qe2 + t/qe
R20.9980.9970.999
qe2.650 (2.635) 17.819 (7.810) 14.065 (3.971) 1
k20.3090.4180.098
Elovich equation
qt = a + blnt
R20.7990.7540.641
a1.9737.1651.800
b0.1390.1400.512
Double constant equation
logqt = logk3 + mlogt
R20.8990.9530.909
m0.0580.0190.173
k31.9957.1731.913
Note: 1 The values in brackets are qe measured values.
Table 5. Thermodynamic parameters of ion adsorption on natural diatomite at different temperatures.
Table 5. Thermodynamic parameters of ion adsorption on natural diatomite at different temperatures.
IonsΔG°ΔH° (kJ∙mol−1)ΔS°
kJ∙mol−1J∙mol−1∙K−1
298 K313 K328 K298 K313 K328 K298 K313 K328 K
Cu2+ *−1.783−1.989−2.2552.8962.8962.89615.67215.67215.672
Zn2+ *−5.890−6.480−7.0966.0916.0916.09140.19040.19040.190
Mn2+ *−0.846−0.905−1.6350.0080.0080.0083.5833.5833.583
Note: * The goodness-of-fit coefficients (R2) for the ions are as follows: Cu2+ (0.998), Zn2+ (0.997), Mn2+ (0.993).
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

Wang, J.; He, Q.; Lei, M.; Han, J.; Wang, J.; Li, W.; Xiao, Y.; Huang, H.; Huang, X.; Zhu, J. Abnormal Adsorption Characteristics of Copper, Zinc, and Manganese Ions on Natural Diatomite in a Liquid/Solid Heterogeneous System. Water 2025, 17, 2782. https://doi.org/10.3390/w17182782

AMA Style

Wang J, He Q, Lei M, Han J, Wang J, Li W, Xiao Y, Huang H, Huang X, Zhu J. Abnormal Adsorption Characteristics of Copper, Zinc, and Manganese Ions on Natural Diatomite in a Liquid/Solid Heterogeneous System. Water. 2025; 17(18):2782. https://doi.org/10.3390/w17182782

Chicago/Turabian Style

Wang, Jieying, Qihao He, Mingjing Lei, Jing Han, Jiacheng Wang, Wenmin Li, Ying Xiao, Hongchun Huang, Xindeng Huang, and Jian Zhu. 2025. "Abnormal Adsorption Characteristics of Copper, Zinc, and Manganese Ions on Natural Diatomite in a Liquid/Solid Heterogeneous System" Water 17, no. 18: 2782. https://doi.org/10.3390/w17182782

APA Style

Wang, J., He, Q., Lei, M., Han, J., Wang, J., Li, W., Xiao, Y., Huang, H., Huang, X., & Zhu, J. (2025). Abnormal Adsorption Characteristics of Copper, Zinc, and Manganese Ions on Natural Diatomite in a Liquid/Solid Heterogeneous System. Water, 17(18), 2782. https://doi.org/10.3390/w17182782

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