Next Article in Journal
Awareness and Understanding of Climate Change for Environmental Sustainability Using a Mix-Method Approach: A Study in the Kathmandu Valley
Previous Article in Journal
Balancing Efficiency and Sustainability: Multicriteria Decision-Making for Pumping Station Upgrades
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Eco-Friendly Utilization of Phosphogypsum via Mechanical Activation for Sustainable Heavy Metal Removal from Wastewater

by
Abdulrahman M. Alotaibi
1,
Abdulrahman A. Aljabbab
1,
Mamdoh S. Alajmi
1,
Ayman N. Qadrouh
1,*,
Mohsen Farahat
2,
Mohamed Abdeldayem Abdel Khalek
2,
Hassan Baioumy
3,
Mansour S. Alhumimidi
4,
Ramzi S. Almutairi
1 and
Sultan A. Alkhammali
1
1
King Abdulaziz City for Science and Technology (KACST), Riyadh 12354, Saudi Arabia
2
Central Metallurgical Research and Development Institute (CMRDI), Helwan 11421, Egypt
3
Natural Resources, Arabian Geophysical & Surveying Co. (ARGAS), Dhahran 34232, Saudi Arabia
4
Ministry of Environment, Water, and Agriculture, Riyadh 12628, Saudi Arabia
*
Author to whom correspondence should be addressed.
Sustainability 2025, 17(7), 2817; https://doi.org/10.3390/su17072817
Submission received: 18 February 2025 / Revised: 13 March 2025 / Accepted: 17 March 2025 / Published: 22 March 2025
(This article belongs to the Section Health, Well-Being and Sustainability)

Abstract

:
This study examined significant changes in phosphogypsum, a byproduct of the phosphoric acid industry, induced via mechanical activation through intensive grinding using a planetary ball mill. Alterations in crystallinity, surface area, and zeta potential were monitored using X-ray diffraction, Brunauer–Emmett–Teller analysis, zeta potential measurements, X-ray photoelectron spectroscopy, and scanning electron microscopy. The severe grinding of this mining waste led to the conversion of gypsum (CaSO4·2H2O) to anhydrite (CaSO4), an increase in surface area from 5.8 m2/g to 17.8 m2/g, and a decrease in pore radius from 76.6 nm to 9.3 nm. The zeta potential shifted as the isoelectric point changed from pH 8.5 to pH 4.3. These modifications enhanced the material’s potential as a cost-effective and eco-friendly adsorbent for wastewater treatment. The enhanced adsorption capabilities for Cd and Pb were evaluated, revealing a higher adsorption capacity (~40 mg/g for both) and removal efficiency (~90% for Cd and ~80% for Pb) for activated phosphogypsum. The adsorption process followed the Freundlich isotherm and pseudo-second-order kinetic model, indicating its physisorption nature and spontaneous thermodynamic characteristics, and highlighting its potential for wastewater treatment. The mechanically activated adsorbent demonstrated over 90% desorption efficiency over five cycles, ensuring effective regeneration and reusability for Cd and Pb removal. Real tannery wastewater was treated using mechanically activated phosphogypsum at pH 6 and 70 °C for 60 min, achieving a 94% Cd and 92% Pb removal efficiency, with an overall heavy metal removal efficiency of up to 83%. This study demonstrates the sustainable utilization of phosphogypsum, contributing to green wastewater management and environmental protection.

1. Introduction

The contamination of water with Cd and Pb significantly threatens aquatic ecosystems and human health owing to their toxic nature and persistence in the environment. Cd, a highly toxic metal, can accumulate in aquatic organisms through the food chain, leading to kidney, liver, and lung cancers [1,2]. Additionally, Pb, which is often present alongside Cd in industrial discharge, exhibits similar toxicological properties and can contaminate water sources, posing serious health risks even at low concentrations [3]. It is therefore necessary to develop cost-effective approaches for the removal of Cd and Pb from water to safeguard human life and ecosystems.
Several methods have been developed to remove Cd and Pb from wastewater, thereby mitigating their harmful impacts on the environment and human health, including coagulation [4,5], precipitation [6,7], bioremediation [8,9], and adsorption [10,11]. Among these approaches, adsorption techniques have attracted the attention of researchers owing to their low cost, high efficiency, and ease of operation. Several materials, such as clay [12,13], biochar [14], and activated carbon [15], have been used as adsorbents for the removal of heavy metals, including Cd and Pb, from contaminated waters. However, many of these adsorbents face limitations such as high production costs, low adsorption capacities, and environmental concerns during their synthesis and disposal [16,17].
Phosphogypsum, a byproduct of phosphate fertilizer production, presents significant environmental concerns because of its large production volume. It is estimated that producing 1 ton of phosphoric acid results in approximately 4–5 tons of phosphogypsum [18,19], while the chemical properties of phosphogypsum and its metal ion content make its direct use difficult in industrial applications [19]. Despite these challenges, efforts have been made to explore their application in construction and agriculture [18,20,21,22]. Additionally, phosphogypsum has been tested as an adsorbent for removing heavy metals, including Cd [23,24], Cu [25], As [26], Pb, and Zn [24,27], and for organic pollutant removal [28]. However, these attempts have shown that its effectiveness as an adsorbent is generally lower than that of other adsorbents. In response, some researchers have investigated chemical modifications to improve the adsorption capabilities of phosphogypsum for the removal of cadmium [29]; nickel, chromium, and zinc [30]; fluoride [31], arsenic [32]; and malachite green [33]; however, these methods can be expensive and environmentally harmful.
More recently, an ecofriendly, cost-effective drying process involving mechanical activation through severe grinding was demonstrated to lead to significant alterations in materials’ surface properties and crystallinity, positively impacting their adsorption capabilities [34,35,36,37].
While progress has been made, the use of phosphogypsum for heavy metal removal, particularly through mechanical activation, remains insufficiently studied. This study addresses this gap by demonstrating the potential of mechanically activated phosphogypsum as an efficient and sustainable adsorbent for Cd and Pb removal. The key innovation of this research is the use of mechanical activation to enhance phosphogypsum’s adsorption capacity. Unlike chemical modifications, which can be expensive and environmentally harmful, mechanical activation offers a simple, eco-friendly, and scalable alternative. This study optimizes key adsorption parameters and analyzes the underlying mechanisms, contributing to a more sustainable approach for heavy metal remediation.
The aim of this study was to employ a mechanical activation technique to boost the adsorption capacity of phosphogypsum for Cd and Pb in wastewater. Key adsorption parameters such as pH, contact time, adsorbent dosage, and temperature were optimized. Additionally, this study investigates the adsorption kinetics and thermodynamics to unravel the underlying mechanisms.

2. Materials and Methods

2.1. Materials

A representative sample of phosphogypsum was obtained from the fertilizer industry (Maaden Company, Riyadh, Saudi Arabia), KSA, characterized by its typical composition and the impurities associated with the phosphoric acid production process.
Analytical-grade Cd nitrate ([Cd (NO3)2]) and Pb nitrate ([Pb (NO3)2]) (Sigma-Aldrich (St. Louis, MO, USA) and Merck (Darmstadt, Germany), respectively) were employed as the Cd and Pb ion sources for the adsorption experiments.

2.2. Methods

2.2.1. Mechanical Activation

A Planetary Micro Mill PULVERISETTE 7 ball mill (FRITSCH, Idar-Oberstein, Germany), equipped with 250 mL Zr oxide jars and balls, was used for the mechanical activation of phosphogypsum. A predetermined amount of phosphogypsum was placed in the milling jar along with stainless steel balls of 5 mm diameter. A ball-to-powder ratio of 15:1 and a rotating velocity of 600 rpm were maintained throughout this study. Activation experiments were conducted for three milling durations (3 h, 5 h, and 7 h).

2.2.2. Characterization

An X-ray diffraction (XRD) analysis was performed using a Bruker D8-ADVANCE diffractometer (Karlsruhe, Germany) with a Cu Kα radiation source (λ = 1.54056 Å) to investigate the changes in phosphogypsum crystallinity before and after mechanical activation. X-ray photoelectron spectroscopy (XPS) was conducted using a Thermo Scientific K-Alpha spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) with monochromatic Al Kα radiation to analyze the elemental composition and chemical states of the materials. Scanning electron microscopy (SEM) was performed using a Quanta 250 Field Emission Gun (FEG) (Eindhoven, The Netherlands) to examine the morphology of phosphogypsum before and after mechanical activation. The surface area and pore characteristics were determined using a Brunauer–Emmett–Teller (BET) analysis with a BELSORP-MR 6 surface area analyzer. Degassing was performed at 105 °C for 4 h. Zeta potential measurements were conducted using a zeta potential analyzer (Malvern Instrument Co., Ltd., Worcestershire, UK), and 1.0 × 10−3 M KCl was used as an electrolyte solution, whereas HCl and KOH were used as pH regulators to assess the surface charge of the phosphogypsum particles.

2.2.3. Adsorption Experiments

Adsorption experiments were conducted using 0.05 g of phosphogypsum and 50 mL of the desired Cd or Pb ion concentration in a round flask. The flask containing the slurry was shaken (200 rpm) for the selected interaction time. The adsorbent sample was then separated from the solution via filtration using a 0.45 μm Whatman filter paper. The concentration of the remaining ions in the synthetic solution was determined using an atomic absorption spectrophotometer (PerkinElmer, AA Analyst 200) (Waltham, MA, USA). The key adsorption parameters, including the pH, ion concentration, contact time, and temperature, were optimized.

3. Results and Discussion

3.1. Physico-Chemical Characterizations

3.1.1. XRD Analysis

The XRD pattern of the pristine phosphogypsum is shown in Figure 1. The sharp peaks at approximately 11.6°, 20.8°, 23.2°, and 29.2° indicated well-crystallized gypsum (JCPDS 01-070-0982). These peaks correspond to the (020), (021), (040), and (041) planes, respectively. The presence of only gypsum peaks in the pattern suggests a relatively pure phosphogypsum sample. However, XRF analysis (Table 1) shows that silica and phosphorous are the main impurities in phosphogypsum.
Figure 2 compares the XRD patterns of the raw phosphogypsum samples and the mechanically activated samples after 3 h, 5 h, and 7 h of grinding. As shown in the figure, the appearance of anhydrite peaks even after 3 h of grinding suggests that the mechanical stress, and possibly the heat generated during severe grinding, were sufficient to initiate the conversion of gypsum into anhydrite. This transformation may be due to the dehydration of gypsum under mechanical force, which removes crystalline water molecules from its structure. With an increase in the grinding time, the peaks in anhydrite became more pronounced, indicating a further conversion of gypsum to anhydrite. This suggests that it is possible to achieve phase transformations in phosphogypsum via mechanical activation, potentially enhancing its reactivity and suitability for various applications.

3.1.2. XPS Analysis

The high-resolution XPS Ca2p, S2p, and O1s spectra of mechanically activated phosphogypsum and the original phosphogypsum (Raw) are displayed in Figure 3. As shown in the figure, the Ca2p spectrum of raw sample exhibits peaks at 348.0 eV and 351.0 eV, typically attributed to the Ca2p3/2 and Ca2p1/2 spin–orbit coupling components of calcium in CaSO4·2H2O (gypsum) [38,39]. Peaks were observed at the same binding energies, indicating that the oxidation state of Ca remained unchanged as Ca2+ even after mechanical activation. The notable increase in peak intensity might be attributed to the surface becoming more enriched in Ca due to the conversion of gypsum CaSO4·2H2O to anhydrite CaSO4, which has a higher relative content of Ca compared to hydrous gypsum.
The peaks at 168.0 eV are characteristic of S2p in sulfate groups (SO42−) in gypsum and the mechanically activated sample, indicating that the sulfate groups remain chemically unchanged; however, the slight broadening of the peaks might suggest some changes in the sulfate environment due to the dehydration process, consistent with the XRD results.
The O1s spectrum of the raw sample shows a broad peak at approximately 533 eV, which typically corresponds to O in sulfate groups. The peaks located at 531.6 eV and 533.9 eV are attributed to SO42− of gypsum [38,40] and the adsorbed water molecules, respectively [41]. The higher intensity of the 533 eV peak in the mechanically activated sample suggests enhanced surface adsorption of water molecules, likely due to increased surface reactivity after mechanical activation.
These XPS findings corroborate the XRD results, confirming that mechanical activation primarily induces physical changes (e.g., water removal) rather than chemical alterations, leading to the transformation of gypsum into anhydrite.
These XPS findings corroborate the XRD results showing that mechanical activation leads to the dehydration of gypsum into anhydrite, with the primary change being the physical (water removal) rather than chemical alteration of the constituent ions.

3.1.3. Microstructure Analysis

The surface morphologies of the raw and mechanically activated phosphogypsum were examined via field-emission scanning electron microscopy (SEM), as shown in Figure 4. The raw phosphogypsum particles appeared as clumped, relatively large aggregates with a layered structure (Figure 4A). A closer view of the raw phosphogypsum particles shows that the particle shapes were more defined, exhibiting sharp edges and smooth surfaces, which are characteristic of crystalline structures (Figure 4B). Conversely, the particles of the mechanically activated phosphogypsum were notably finer and more dispersed compared to the raw material with high surface roughness (Figure 4C,D). This surface morphology alteration results from the severe bombardment of the particles by the grinding media. Additionally, these dramatic changes are expected to improve the adsorption capability of the phosphogypsum particles.

3.1.4. Zeta Potential

The zeta potentials of the raw and mechanically activated phosphogypsum samples are shown in Figure 5. Pristine gypsum consistently exhibited a low negatively charged surface at various pH levels, except for a positive zeta potential at pH 10. Two isoelectric points (IEPs) were observed: one below pH 2 and the other at pH 9. These results agree with those of previous studies [42]. In contrast, the mechanically activated gypsum exhibited a more dynamic and pH-dependent pattern. At an acidic pH, the zeta potential became highly positive, signifying a shift towards a positively charged surface compared with that of pristine gypsum. However, at an alkaline pH, the zeta potential reverted to a negative value, and the IEP was observed at pH 4.3. The significant shift towards more negative values, particularly at higher pH levels, can be attributed to the higher dissolution rates of Ca and P ions in the mechanically activated sample solution. When leached into the solution, these ions leave behind a more negatively charged surface.
The zeta potential data complement the findings from the XRD, XPS, and SEM analyses, showing that mechanical activation significantly enhances the ionic reactivity and surface charge properties of phosphogypsum. These changes are expected to improve the adsorption performance of the mechanically activated phosphogypsum in wastewater treatment, where charge interactions are crucial for the adsorption and removal of contaminants.

3.1.5. Surface Area

The N2 adsorption–desorption isotherms of raw and mechanically activated phosphogypsum are shown in Figure 6. The inset graphs show a plot of the pore size distribution, and the BET surface area, average pore diameter, and total pore volume are presented in Table S1. The isotherm of raw phosphogypsum showed a low adsorption capacity, which remained flat over a wide range of relative pressures, indicating few accessible pores. The shape of the isotherm suggests a material with low porosity, which is typical of dense non-porous materials.
In contrast, the isotherm of the mechanically activated sample exhibits significantly higher adsorption, particularly at higher relative pressures, indicating the creation of more accessible pore structures. The shape of this isotherm is typical of mesoporous materials (type B of DeBoer’s classification). The average pore radius decreased significantly, from 76.62 nm in the raw sample to 9.34 nm in the mechanically activated sample, indicating the creation of finer pores as larger aggregates were broken down. The total pore volume more than doubled, from 0.0258 cm3/g to 0.0574 cm3/g, indicating smaller and more numerous pores in the mechanically activated sample. Consequently, a significant increase in surface area, from 5.81 m2/g in the raw sample to 17.86 m2/g in the mechanically activated gypsum, was observed. These observations confirm that mechanical activation causes a dramatic destruction of the particle structure, rendering it with a more complicated surface morphology and an altered pore structure, and in turn improving its surface reactivity, which aligns with the improved adsorption performance observed in the study.

3.2. Adsorption Study

3.2.1. Effect of pH

pH is one of the most important working parameters in the adsorption process because it can cause dissociation at the adsorbent sites and alter the solution chemistry of metal ions. Figure 7 shows the effects of pH on the adsorption capacity and removal efficiency of Cd and Pb using raw and mechanically activated phosphogypsum. These experiments were conducted with a metal ion concentration of 50 mg/L, an adsorbent dose of 1 g/L, a contact time of 60 min, and a temperature of 30 °C. As shown in the figure, the adsorption capacity and removal efficiencies increase as the solution pH transitions from acidic to alkaline regions for the two samples. Moreover, the mechanically activated phosphogypsum sample exhibited a higher adsorption efficiency than the raw sample. This behavior can be explained by the zeta potential results (Figure 5), where the shift towards alkaline regions resulted in the particles of mechanically activated phosphogypsum becoming more negatively charged, thus enhancing the electrostatic attraction forces between the metal ions (Cd or Pb) and the adsorbent, thereby improving their adsorption efficiency. Furthermore, the high surface roughness, increased surface area, and smaller pore size of mechanically activated phosphogypsum contributed to its higher removal efficiency compared to the raw phosphogypsum sample, even in the acidic and neutral pH regions, where both samples exhibited positive charges.

3.2.2. Effect of Initial Concentration

The effects of the initial concentrations of Cd and Pb ions on the adsorption capacity and removal efficiency are shown in Figure 8. The maximum adsorption capacities achieved at an initial metal ion concentration of 250 mg/L were 215 mg/g for Cd and 170 mg/g for Pb using mechanically activated phosphogypsum, and 138 mg/g for Cd and 171 mg/g for Pb using raw phosphogypsum. The increased adsorption capacity of the mechanically activated phosphogypsum compared to that of the raw sample was due to its unique surface features, created via mechanical activation.
Figure 8 shows that increasing the initial ion concentration enhanced the uptake capacity but reduced the removal efficiency. The decline in the removal efficiency resulted from the increase in residual metal ions as the initial ion concentration increased. At higher concentrations, more ions are available, thereby increasing the amount absorbed by the sorbent. The driving force required to overcome the barrier for ion migration through the medium to the solid surface of the sorbent increases, thereby progressively loading the sites until saturation is reached [43].

3.2.3. Isotherm Models Study

The sorption process was analyzed using the Langmuir and Freundlich isotherms to describe the behavior [44]. According to [45], the Langmuir isotherm is described by the following equation:
C f q t = C f q m a x + 1 b q m a x
where Cf (mg/L) is the final ion concentration, qt (mg/g) is the number of ions sorbed at time t, qmax (mg/g) is the maximum sorption for the monolayer adsorption capacity, and b (L/mg) is the binding constant related to the adsorption free energy. The Langmuir model suggests that sorption occurs on a uniform adsorbent surface, where ions move through the pores and openings of the structure to replace the ions of the adsorbent. As shown in Figure S1 and Table 2, the R2 values from the linear fits of the Langmuir model are 0.84 and 0.90. In contrast, the Freundlich isotherm model presented by [46] is determined using the following equation:
l n   q t = l n   k + 1 n l n   C f
where qt (mg/g) is the sorbed amount, Cf (mg/L) is the final concentration, K is the extent of the sorption, and n is the sorption intensity. The function 1/n represents sorption capability [47]. If n = 1, the interface between the two phases is unaffected by the concentration changes. An n value of <1 indicates typical adsorption, whereas values between 1 and 10 suggest a favorable sorption process [48]. From Table 2, the n values are 1.27 and 1.35, and the R2 values exceed 0.99 for the two samples, respectively, indicating a favorable and predominantly physical sorption process [49].
The Freundlich model provides a better fit (R2 > 0.99) compared to the Langmuir model, suggesting that adsorption occurs on a heterogeneous surface with non-uniform energy distribution. This is consistent with the observed changes in surface morphology and pore structure after mechanical activation, as discussed in the SEM and BET analyses.

3.2.4. Effect of Contact Time

Figure 9 shows the effects of contact time on the adsorption capacity and removal efficiency of Cd and Pb using raw and mechanically activated samples. The equilibrium time was 60 min. A higher adsorption rate was observed within the first 20 min, which was attributed to readily available adsorbate ions and vacant adsorption sites on the surface of the adsorbent. However, from 20 to 60 min, the adsorption rate gradually decreased as the available active sites became increasingly occupied. The extent of adsorption was measured based on the number of ions transferred from the solution to the active sites. Consequently, the adsorption continued to increase with time until saturation was reached [50]. Moreover, the ions required additional time to penetrate smaller pores.
The higher initial rate suggests that adsorption initially occurred on the external surface, followed by penetration into the internal pores [51]. Additionally, the larger adsorption volumes observed in the initial period indicated more significant adsorption on the external surface than within the pores [52]. The maximum removal efficiencies achieved after a 60 min contact period were 86% for Cd and 68% for Pb with mechanically activated phosphogypsum, and 68.5% for Cd and 55% for Pb with raw phosphogypsum.

3.2.5. The Adsorption Kinetics

The extent of the sorption of Cd and Pb ions by raw and mechanically activated phosphogypsum was examined using the Lagergren pseudo-first-order (PFO) and pseudo-second-order (PSO) models. The Lagergren values for the PFO and PSO models are denoted by the following equations:
The PFO equation [53]:
ln q e q t = ln q e k 1 t
The PSO equation [54]:
t q t = 1 k 2 q e 2 t q e
where qt (mg/g) and qe (mg/g) are the number of ions adsorbed by the adsorbent at time t (min) and at equilibrium, respectively, and k1 (min−1) and k2 (g·mg−1·min−1) are the equilibrium rate constants.
In addition, their equivalent limits are listed in Table 3. Although the PFO and PSO models are the most widely applied models that forecast closer values of the maximum sorption capacity, the best fit was found using PSO linear retrogression, based on the R2 closest value to unity. The kinetic fitting quality changes in the following order: PSO > PFO.

3.2.6. Effect of Temperature

The effects of temperature on the removal efficiency and sorption capacity of Cd and Pb ions by raw and mechanically activated phosphogypsum samples are illustrated in Figure 10. In these experiments, the solution pH was maintained at 6, the initial ion concentration was 250 mg/L, and the adsorbent dose was 1 g/L. As depicted in the figure, the removal efficiency and adsorption capacity are directly proportional to temperature, with the maximum adsorption capacity being reached at 70 °C. This indicates that the adsorption is endothermic. Most adsorption studies have suggested that increasing the temperature enhances the sorption process [55]. At higher temperatures, the uptake is typically greater, owing to an increase in the energetic sites of the sorbent material. Increased temperatures also energize the system, promoting the attachment of ions to mineral surfaces [56]; moreover, the ion movement accelerates as the viscosity of the solution decreases [57], resulting in higher removal efficiencies [58].

3.2.7. Thermodynamic Study

Changes in thermodynamics parameters, such as Gibb’s free energy (ΔG°), entropy (ΔS°), and enthalpy (ΔH°) [55], were determined. ΔH° and ΔS° were computed from the Van’t Hoff equation [59]:
ln K c = S ° R H ° R T
where kc = F/(1 − F) and F = (Co − Ce)/Co, R is the universal gas constant, and T is the temperature in K. The relationship of lnkc versus 1/T (Figure S3) provides a straight line with a slope of −ΔH°/R and an intercept equal to ΔS°/R. The positive values of ΔH° in Table S2 specify the endothermic sorption process. Additionally, the ΔH° values are approximately ≥40 kJ.mol−1, which verifies the physisorption process, as shown by the Freundlich isotherm [60].
Additionally, ΔS° exhibited positive values, suggesting a degree of unpredictability at the interface between the sorbent and adsorbate, inferring that sorption is less advantageous at lower temperatures. The positive ΔS value shows that the stochastic affinity of this process, which suggests adsorption, is a favorable spontaneous reaction.
The standard Gibbs free energy change ΔG° is estimated using the following equation [60]:
G ° = R T   l n K c
Table S2 shows that the sorption is unprompted because the ΔG° has negative values. Notably, the negative values of ΔG° increased with an increasing temperature, suggesting that the sorption is more favorable at higher temperatures [60].
These thermodynamic findings align with the Freundlich isotherm results, confirming that the adsorption process is predominantly physical, spontaneous, and temperature-dependent. The endothermic nature of the process and the increased spontaneity at higher temperatures further explain the enhanced adsorption performance observed using phosphogypsum.

3.2.8. Regeneration and Reusability of Adsorbent

The regeneration and reusability of the mechanically activated sample for cadmium and lead ion removal were examined under the adsorption conditions of an initial ion concentration of 250 mg/L, pH 6, temperature 70 °C, and a contact time of 60 min. The regeneration process was conducted using 0.1 M HNO3 with a solid/liquid ratio of 1:10, a contact time of 15 min, and a temperature of 65 °C. As shown in Figure 11, the adsorbed ions were successfully desorbed with an efficiency exceeding 90% for the first five cycles. These results indicate that the mechanically activated adsorbent can be easily separated and reused multiple times.

3.2.9. Comparison Studies

To evaluate the performance of phosphogypsum adsorbent, its adsorption capacity for lead and cadmium ions was compared with that of other reported adsorbents, as summarized in Table 4. The results demonstrate that phosphogypsum, both in its raw and activated forms, exhibits superior performance in removing lead and cadmium ions compared to other materials. Notably, many of these alternative adsorbents were synthesized using high-cost chemical methods and advanced techniques. This comparison emphasizes the significant improvements achieved through mechanical activation and highlights the competitive advantage of our approach, which relies on a low-cost and sustainable material.

3.2.10. Application in Real Wastewater

Real wastewater from a tannery workshop was used for the application. The treatment was performed by adding a mechanically activated sample to the wastewater after adjusting the pH to 6 and maintaining a temperature of 70 °C for 60 min. As shown in Table S3, cadmium and lead ions were removed with an efficiency of 94% and 92%, respectively, while other heavy metal ions were also successfully removed, with an overall removal efficiency of up to 83%.

4. Conclusions

In this study, the mechanical activation of phosphogypsum through intensive grinding significantly enhanced its physicochemical properties, improving its effectiveness as an eco-friendly adsorbent for Cd and Pb removal from synthetic wastewater. The transformation from gypsum (CaSO4·2H2O) to anhydrite (CaSO4) increased the surface area and reduced the pore radius, enhancing the adsorption capacity.
The optimization of adsorption parameters (pH, contact time, adsorbent dosage, and temperature) confirmed that mechanically activated phosphogypsum outperformed its raw form. The adsorption process followed the Freundlich isotherm (physisorption) and PSO kinetic model, highlighting the importance of surface interactions.
The adsorbent achieved 94% Cd and 92% Pb removal efficiency from real tannery wastewater at pH 6 and 70 °C, with over 90% desorption efficiency over five cycles, thus demonstrating excellent regeneration and reusability. This study highlights the innovative use of mechanical activation to transform phosphogypsum into an effective, eco-friendly adsorbent, offering a sustainable solution for heavy metal remediation and waste valorization.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su17072817/s1, Figure S1. Freundlich and Langmuir isotherms; Figure S2. The pseudo-first-order and pseudo-second-order models’ plotting data; Figure S3. Plot of the thermodynamic parameters for raw and mechanically activated (MA) samples; Table S1. Textural properties of raw and mechanically activated phosphogypsum; Table S2. Thermodynamic parameters for raw and mechanically activated (MA) samples; Table S3. Heavy metal ions in real tannery wastewater samples before and after treatment with MA samples.

Author Contributions

A.M.A.: Initiate the project, collect data, write the manuscript, A.A.A.: collect data, writing—reviewing and editing, M.S.A. (Mamdoh S. Alajmi): writing—reviewing and editing, A.N.Q.: collect data, writing—reviewing and editing, M.F.: analysis—writing—reviewing and editing, M.A.A.K.: writing—reviewing and editing, H.B.: interpret data, write the manuscript, M.S.A. (Mansour S. Alhumimidi): reviewing and editing, R.S.A.: reviewing and editing, S.A.A.: reviewing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The field trips were funded by King Abdulaziz City for Science and Technology.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be available upon request.

Acknowledgments

The authors thank King Abdulaziz City for Science and Technology (KACST) for the funding support.

Conflicts of Interest

Author Hassan Baioumy was employed by the company Arabian Geophysical & Surveying Co. (ARGAS). The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Godt, J.; Scheidig, F.; Grosse-Siestrup, C.; Esche, V.; Brandenburg, P.; Reich, A.; Groneberg, D.A. The toxicity of cadmium and resulting hazards for human health. J. Occup. Med. Toxicol. 2006, 1, 22. [Google Scholar] [CrossRef] [PubMed]
  2. Waalkes, M.P. Cadmium carcinogenesis in review. J. Inorg. Biochem. 2000, 79, 241–244. [Google Scholar] [CrossRef]
  3. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy metals. Interdiscip. Toxicol. 2014, 7, 60–72. [Google Scholar] [CrossRef]
  4. AlJaberi, F.Y.; Hawaas, Z.A. Electrocoagulation removal of Pb, Cd, and Cu ions from wastewater using a new configuration of electrodes. MethodsX 2023, 10, 101951. [Google Scholar] [CrossRef]
  5. Liang, Y.; Jun, M.; Liu, W. Enhanced Removal of Lead(II) and Cadmium(II) from Water in Alum Coagulation by Ferrate(VI) Pretreatment. Water Environ. Res. 2007, 79, 2420–2426. [Google Scholar] [CrossRef] [PubMed]
  6. Kavak, D. Removal of lead from aqueous solutions by precipitation: Statistical analysis and modeling. Desalination Water Treat. 2013, 51, 1720–1726. [Google Scholar] [CrossRef]
  7. Habte, L.; Shiferaw, N.; Thriveni, T.; Mulatu, D.; Lee, M.h.; Jung, S.h.; Ahn, J.W. Removal of Cd(II) and Pb(II) from Wastewater via Carbonation of Aqueous Ca(OH)2 Derived from Eggshell. Process Saf. Environ. Prot. 2020, 141, 278–287. [Google Scholar] [CrossRef]
  8. Chellaiah, E.R. Cadmium (heavy metals) bioremediation by Pseudomonas aeruginosa: A minireview. Appl. Water Sci. 2018, 8, 154. [Google Scholar] [CrossRef]
  9. Kumar, A.; Subrahmanyam, G.; Mondal, R.; Cabral-Pinto MM, S.; Shabnam, A.A.; Jigyasu, D.K.; Malyan, S.K.; Fagodiya, R.K.; Khan, S.A.; Yu, Z.G. Bio-remediation approaches for alleviation of cadmium contamination in natural resources. Chemosphere 2021, 268, 128855. [Google Scholar] [CrossRef]
  10. Chowdhury, I.R.; Chowdhury, S.; Mazumder MA, J.; Al-Ahmed, A. Removal of lead ions (Pb2+) from water and wastewater: A review on the low-cost adsorbents. Appl. Water Sci. 2022, 12, 185. [Google Scholar] [CrossRef]
  11. Li, S.; Li, S.; Wen, N.; Wei, D.; Zhang, Y. Highly effective removal of lead and cadmium ions from wastewater by bifunctional magnetic mesoporous silica. Sep. Purif. Technol. 2021, 265, 118341. [Google Scholar] [CrossRef]
  12. Khalfa, L.; Sdiri, A.; Bagane, M.; Cervera, M.L. A calcined clay fixed bed adsorption studies for the removal of heavy metals from aqueous solutions. J. Clean. Prod. 2021, 278, 123935. [Google Scholar] [CrossRef]
  13. Pfeifer, A.; Škerget, M.; Čolnik, M. Removal of iron, copper, and lead from aqueous solutions with zeolite, bentonite, and steel slag. Sep. Sci. Technol. 2021, 56, 2989–3000. [Google Scholar] [CrossRef]
  14. Cheng, S.; Liu, Y.; Xing, B.; Qin, X.; Zhang, C.; Xia, H. Lead and cadmium clean removal from wastewater by sustainable biochar derived from poplar saw dust. J. Clean. Prod. 2021, 314, 128074. [Google Scholar] [CrossRef]
  15. Karnib, M.; Kabbani, A.; Holail, H.; Olama, Z. Heavy Metals Removal Using Activated Carbon, Silica and Silica Activated Carbon Composite. Energy Procedia 2014, 50, 113–120. [Google Scholar] [CrossRef]
  16. Zhang, H.-Y.; Huang, Z.; Liu, Y.-H.; Hu, L.-X.; He, L.-Y.; Liu, Y.-S.; Zhao, J.-L.; Ying, G.-G. Occurrence and risks of 23 tire additives and their transformation products in an urban water system. Environ. Int. 2023, 171, 107715. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, Y.; Hu, Y.; Chen, A.J.Y.; Cheng, Z.; Bi, Z.; Zhang, R.; Lou, Z. Environmental impacts and nutrient distribution routes for food waste separated disposal on large-scale anaerobic digestion/composting plants. J. Environ. Manag. 2022, 318, 115624. [Google Scholar] [CrossRef]
  18. Al-Masri, M.S.; Amin, Y.; Ibrahim, S.; Al-Bich, F. Distribution of some trace metals in Syrian phosphogypsum. Appl. Geochem. 2004, 19, 747–753. [Google Scholar] [CrossRef]
  19. Rosales, J.; Gázquez, M.; Cabrera, M.; Bolivar, J.P.; Agrela, F. Application of phosphogypsum for the improvement of eco-efficient cements. In Waste and Byproducts in Cement-Based Materials; Woodhead Publishing: Sawston, UK, 2021; pp. 153–189. [Google Scholar] [CrossRef]
  20. Akfas, F.; Elghali, A.; Aboulaich, A.; Munoz, M.; Benzaazoua, M.; Bodinier, J.L. Exploring the potential reuse of phosphogypsum: A waste or a resource? Sci. Total Environ. 2024, 908, 168196. [Google Scholar] [CrossRef]
  21. Outbakat, M.B.; Choukr-Allah, R.; Bouray, M.; El Gharous, M.; El Mejahed, K. Phosphogypsum: Properties and potential use in agriculture. In Biosaline Agriculture as a Climate Change Adaptation for Food Security; Springer International Publishing: Cham, Switzerland, 2023; pp. 229–255. [Google Scholar] [CrossRef]
  22. Pliaka, M.; Gaidajis, G. Potential uses of phosphogypsum: A review. Journal of Environmental Science and Health. Part A Toxic/Hazard. Subst. Environ. Eng. 2022, 57, 746–763. [Google Scholar] [CrossRef]
  23. Rashad, A.M. Phosphogypsum as a construction material. J. Clean. Prod. 2017, 166, 732–743. [Google Scholar] [CrossRef]
  24. Es-said, A.; Nafai, H.; Lamzougui, G.; Bouhaouss, A.; Bchitou, R. Comparative adsorption studies of cadmium ions on phosphogypsum and natural clay. Sci. Afr. 2021, 13, e00960. [Google Scholar] [CrossRef]
  25. Es-said, A.; Nafai, H.; Zerki, N.; Bchitou, R. Chemometrics approach for multi-response optimisation of heavy metals Zn(II), Cu(II) and Cd(II) removal by phosphogypsum: Ternary aqueous solution. Int. J. Environ. Anal. Chem. 2021, 103, 3044–3058. [Google Scholar] [CrossRef]
  26. Fernández-Martínez, A.; Román-Ross, G.; Cuello, G.J.; Turrillas, X.; Charlet, L.; Johnson, M.R.; Bardelli, F. Arsenic uptake by gypsum and calcite: Modelling and probing by neutron and X-ray scattering. Phys. B Condens. Matter 2006, 385–386, 935–937. [Google Scholar] [CrossRef]
  27. Cesur, H.; Balkaya, N. Zinc removal from aqueous solution using an industrial by-product phosphogypsum. Chem. Eng. J. 2007, 131, 203–208. [Google Scholar] [CrossRef]
  28. Balkaya, N.; Cesur, H. Adsorption of cadmium from aqueous solution by phosphogypsum. Chem. Eng. J. 2008, 140, 247–254. [Google Scholar] [CrossRef]
  29. Dai, Q.; Xie, L.; Ma, L.; Yang, J.; Yang, X.; Ren, N.; Tian, G.; Guo, Z.; Ning, P. Effects of flocculant-modified phosphogypsum on sludge treatment: Investigation of the operating parameters, variations of the chemical groups, and heavy metals in the sludge. Environ. Sci. Water Res. Technol. 2021, 7, 184–196. [Google Scholar] [CrossRef]
  30. Jiang, W.; Jiang, Y.; Li, P.; Liu, D.; Ren, Y.; Li, D.; Liu, Z.; Chen, Y.; Ye, Y. Reuse of phosphogypsum and phosphorus ore flotation tailings as adsorbent: The adsorption performance and mechanism of phosphate. J. Phys. Chem. Solids 2023, 178, 111313. [Google Scholar] [CrossRef]
  31. Liu, Z.; Zhang, J.; Mou, R. Phosphogypsum-Modified Vinasse Shell Biochar as a Novel Low-Cost Material for High-Efficiency Fluoride Removal. Molecules 2023, 28, 7617. [Google Scholar] [CrossRef]
  32. Shang, C.; Geng, Z.; Sun, Y.; Che, D.; Zhao, Q.; Chen, T.; Tang, M.; Huo, L. Lanthanum-Modified Phosphogypsum Red Mud Composite for the Co-Adsorption of Cadmium and Arsenic: Mechanism Study and Soil Remediation. Agriculture 2024, 14, 464. [Google Scholar] [CrossRef]
  33. Zhao, L.; Zhang, Q.; Li, X.; Ye, J.; Chen, J. Adsorption of Cu(II) by phosphogypsum modified with sodium dodecyl benzene sulfonate. J. Hazard. Mater. 2020, 387, 121808. [Google Scholar] [CrossRef]
  34. Huang, P.; Li, Z.; Chen, M.; Hu, H.; Lei, Z.; Zhang, Q.; Yuan, W. Mechanochemical activation of serpentine for recovering Cu (II) from wastewater. Appl. Clay Sci. 2017, 149, 1–7. [Google Scholar] [CrossRef]
  35. Pechishcheva, N.V.; Burdina, L.G.; Kel’, P.V.; Estemirova, S.K.; Konysheva, E.Y.; Sushnikova, A.A. Mechanical activation of anatase as a way to improve hexavalent chromium removal from solutions by the photoreduction and adsorption under near visible LED irradiation. Mater. Chem. Phys. 2024, 314, 128872. [Google Scholar] [CrossRef]
  36. Xiyili, H.; Çetintaş, S.; Bingöl, D. Removal of some heavy metals onto mechanically activated fly ash: Modeling approach for optimization, isotherms, kinetics and thermodynamics. Process Saf. Environ. Prot. 2017, 109, 288–300. [Google Scholar] [CrossRef]
  37. Zhang, W.; Li, C.; Xu, Q.; Hu, K.; Chen, H.; Liu, Y.; Wan, Y.; Zhang, J.; Li, X. Effective adsorption and recovery of rare earth elements from wastewater by activated talc. Appl. Clay Sci. 2024, 251, 107312. [Google Scholar] [CrossRef]
  38. Demri, B.; Muster, D. XPS study of some calcium compounds. J. Mater. Process. Technol. 1995, 55, 311–314. [Google Scholar] [CrossRef]
  39. Guan, B.; Ye, Q.; Zhang, J.; Lou, W.; Wu, Z. Interaction between α-calcium sulfate hemihydrate and superplasticizer from the point of adsorption characteristics, hydration and hardening process. Cem. Concr. Res. 2010, 40, 253–259. [Google Scholar] [CrossRef]
  40. Jia, X.J.; Wang, J.; Wu, J.; Teng, W.; Zhao, B.; Li, H.; Du, Y. Facile synthesis of MoO2/CaSO4 composites as highly efficient adsorbents for congo red and rhodamine B. RSC Adv. 2018, 8, 1621. [Google Scholar] [CrossRef]
  41. Jin, Y.; Zheng, Y.; Podkolzin, S.G.; Lee, W. Band gap of reduced graphene oxide tuned by controlling functional groups. J. Mater. Chem. C 2020, 8, 4885–4894. [Google Scholar] [CrossRef]
  42. Salopek, B.; Krasic, D.; Filipovic, S. Measurement and application of zeta-potential. Rud.-Geol.-Naft. Zb. 1992, 4, 147–151. [Google Scholar]
  43. Mahmoud, G.A.; Abdel Khalek, M.A.; Shoukry, E.M.; Amin, M.; Abdulghany, A.H. Removal of Phosphate Ions from Wastewater by Treated Hydrogel Based on Chitosan. Egypt. J. Chem. 2019, 62, 1537–1549. [Google Scholar] [CrossRef]
  44. Hałas, P.; Kołodyńska, D.; Płaza, A.; Gęca, M.; Hubicki, Z. Modified fly ash and zeolites as an effective adsorbent for metal ions from aqueous solution. Adsorpt. Sci. Technol. 2017, 35, 519–533. [Google Scholar] [CrossRef]
  45. Chen, X.; Guo, Y.; Cheng, F.; Song, H.; Zheng, N.; Wang, X. Application of Modified Coal Fly Ash as an Absorbent for Ammonia-Nitrogen Wastewater Treatment. Adv. Mater. Res. 2012, 518–523, 2380–2384. [Google Scholar] [CrossRef]
  46. Visa, M. Synthesis and characterization of new zeolite materials obtained from fly ash for heavy metals removal in advanced wastewater treatment. Powder Technol. 2016, 294, 338–347. [Google Scholar] [CrossRef]
  47. Tsai, W.T.; Chen, H.R. Adsorption kinetics of herbicide paraquat in aqueous solution onto a low-cost adsorbent, swine-manure-derived biochar. Int. J. Environ. Sci. Technol. 2013, 10, 1349–1356. [Google Scholar] [CrossRef]
  48. Goldberg, S. Equations and models describing adsorption processes in soils. Chem. Process. Soils 2018, 8, 489–517. [Google Scholar] [CrossRef]
  49. Ho, Y.S.; Mckay, G. The kinetics of sorption of basic dyes from aqueous solution by sphagnum moss peat. Can. J. Chem. Eng. 1998, 76, 822–827. [Google Scholar] [CrossRef]
  50. Zare, E.N.; Lakouraj, M.M.; Masoumi, M. Efficient removal of Pb(II) and Cd(II) from water by cross-linked poly (N-vinylpyrrolidone-co-maleic anhydride)@eggshell/Fe3O4 environmentally friendly nanocomposite. Desalination Water Treat. 2018, 106, 209–219. [Google Scholar] [CrossRef]
  51. Chung, N.T. Adsorptive removal of five heavy metals from water using blast furnace slag and fly ash. Vietnam J. Sci. Technol. 2018, 54, 314. [Google Scholar] [CrossRef]
  52. Nazarzadeh Zare, E.; Mansour Lakouraj, M.; Ramezani, A. Efficient sorption of Pb(II) from an aqueous solution using a poly(aniline-co-3-aminobenzoic acid)-based magnetic core–shell nanocomposite. New J. Chem. 2016, 40, 2521–2529. [Google Scholar] [CrossRef]
  53. Lopes EC, N.; dos Anjos FS, C.; Vieira EF, S.; Cestari, A.R. An alternative Avrami equation to evaluate kinetic parameters of the interaction of Hg(II) with thin chitosan membranes. J. Colloid Interface Sci. 2003, 263, 542–547. [Google Scholar] [CrossRef] [PubMed]
  54. Issaoui, H.; Sallem, F.; Lafaille, J.; Grassl, B.; Charrier–El Bouhtoury, F. Biosorption of Heavy Metals from Water onto Phenolic Foams Based on Tannins and Lignin Alkaline Liquor. Int. J. Environ. Res. 2021, 15, 369–381. [Google Scholar] [CrossRef]
  55. Plaza, L.; Castellote, M.; Nevshupa, R.; Jimenez-Relinque, E. High-capacity adsorbents from stainless steel slag for the control of dye pollutants in water. Environ. Sci. Pollut. Res. 2021, 28, 23896–23910. [Google Scholar] [CrossRef] [PubMed]
  56. Arief, V.O.; Trilestari, K.; Sunarso, J.; Indraswati, N.; Ismadji, S. Recent Progress on Biosorption of Heavy Metals from Liquids Using Low Cost Biosorbents: Characterization, Biosorption Parameters and Mechanism Studies. CLEAN Soil Air Water 2008, 36, 937–962. [Google Scholar] [CrossRef]
  57. Fakari, S.; Nezamzadeh-Ejhieh, A. Synergistic effects of ion exchange and complexation processes in cysteine-modified clinoptilolite nanoparticles for removal of Cu (II) from aqueous solutions in batch and continuous flow systems. New J. Chem. 2017, 41, 3811–3820. [Google Scholar] [CrossRef]
  58. Rukayat, O.O.; Usman, M.F.; Elizabeth, O.M.; Abosede, O.O.; Faith, I.U. Kinetic Adsorption of Heavy Metal (Copper) On Rubber (Hevea brasiliensis) Leaf Powder. S. Afr. J. Chem. Eng. 2021, 37, 74–80. [Google Scholar] [CrossRef]
  59. Murray, J.W.; Dillard, J.G. The oxidation of cobalt (II) adsorbed on manganese dioxide. Geochim. Cosmochim. Acta 1979, 43, 781–787. [Google Scholar] [CrossRef]
  60. Alkan, M.; Demirbaş, Ö.; Çelikçapa, S.; Doğan, M. Sorption of acid red 57 from aqueous solution onto sepiolite. J. Hazard. Mater. 2004, 116, 135–145. [Google Scholar] [CrossRef]
  61. Mehrasbi, M.R.; Farahmandkia, Z.; Taghibeigloo, B.; Taromi, A. Adsorption of Lead and Cadmium from Aqueous Solution by Using Almond Shells. Water. Air Soil Pollut. 2009, 199, 343–351. [Google Scholar] [CrossRef]
  62. Karunanayake, A.G.; Todd, O.A.; Crowley, M.; Ricchetti, L.; Pittman, C.U.; Anderson, R.; Mohan, D.; Mlsna, T. Lead and Cadmium Remediation Using Magnetized and Nonmagnetized Biochar from Douglas Fir. Chem. Eng. J. 2018, 331, 480–491. [Google Scholar] [CrossRef]
  63. Cui, L.; Chen, T.; Yin, C.; Yan, J.; Ippolito, J.A.; Hussain, Q. Mechanism of Adsorption of Cadmium and Lead Ions by Iron-Activated Biochar. BioResources 2019, 14, 842–857. [Google Scholar] [CrossRef]
  64. Abdulrahman Oyekanmi, A.; Abd Latiff, A.A.; Daud, Z.; Saphira Radin Mohamed, R.M.; Ismail, N.; Ab Aziz, A.; Rafatullah, M.; Hossain, K.; Ahmad, A.; Kamoldeen Abiodun, A. Adsorption of Cadmium and Lead from Palm Oil Mill Effluent Using Bone-Composite: Optimisation and Isotherm Studies. Int. J. Environ. Anal. Chem. 2019, 99, 707–725. [Google Scholar] [CrossRef]
  65. Chen, Z.L.; Zhang, J.Q.; Huang, L.; Yuan, Z.H.; Li, Z.J.; Liu, M.C. Removal of Cd and Pb with Biochar Made from Dairy Manure at Low Temperature. J. Integr. Agric. 2019, 18, 201–210. [Google Scholar] [CrossRef]
  66. Kavand, M.; Eslami, P.; Razeh, L. The Adsorption of Cadmium and Lead Ions from the Synthesis Wastewater with the Activated Carbon: Optimization of the Single and Binary Systems. J. Water Process Eng. 2020, 34, 101151. [Google Scholar] [CrossRef]
  67. Gul, S.; Ahmad, Z.; Asma, M.; Ahmad, M.; Rehan, K.; Munir, M.; Bazmi, A.A.; Ali, H.M.; Mazroua, Y.; Salem, M.A.; et al. Effective Adsorption of Cadmium and Lead Using SO3H-Functionalized Zr-MOFs in Aqueous Medium. Chemosphere 2022, 307, 135633. [Google Scholar] [CrossRef]
  68. Vishwakarma, M.C.; Joshi, H.K.; Tiwari, P.; Bhandari, N.S.; Joshi, S.K. Thermodynamic, Kinetic, and Equilibrium Studies of Cu(II), Cd(II), Ni(II), and Pb(II) Ion Biosorption onto Treated Ageratum Conyzoid Biomass. Int. J. Biol. Macromol. 2024, 274, 133001. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of raw phosphogypsum sample (Gy = gypsum, JCPDS 01-070-0982).
Figure 1. XRD pattern of raw phosphogypsum sample (Gy = gypsum, JCPDS 01-070-0982).
Sustainability 17 02817 g001
Figure 2. XRD patterns for mechanically activated phosphogypsum at different grinding times (Gy = gypsum and A = anhydrite).
Figure 2. XRD patterns for mechanically activated phosphogypsum at different grinding times (Gy = gypsum and A = anhydrite).
Sustainability 17 02817 g002
Figure 3. XPS spectra for Ca2p, S2p, and O1s of raw and mechanically activated (MA) phosphogypsum.
Figure 3. XPS spectra for Ca2p, S2p, and O1s of raw and mechanically activated (MA) phosphogypsum.
Sustainability 17 02817 g003
Figure 4. SEM images of raw phosphogypsum (A,B) and mechanically activated phosphogypsum (C,D).
Figure 4. SEM images of raw phosphogypsum (A,B) and mechanically activated phosphogypsum (C,D).
Sustainability 17 02817 g004
Figure 5. Zeta potential of raw and mechanically activated (MA) phosphogypsum.
Figure 5. Zeta potential of raw and mechanically activated (MA) phosphogypsum.
Sustainability 17 02817 g005
Figure 6. Adsorption–desorption isotherms and pore size distribution for raw and mechanically activated gypsum.
Figure 6. Adsorption–desorption isotherms and pore size distribution for raw and mechanically activated gypsum.
Sustainability 17 02817 g006
Figure 7. Effect of pH on sorption capacity and removal efficiency of Cd and Pb using raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 50 mg/L at 30 °C for 60 min).
Figure 7. Effect of pH on sorption capacity and removal efficiency of Cd and Pb using raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 50 mg/L at 30 °C for 60 min).
Sustainability 17 02817 g007
Figure 8. Effect of initial ions concentration on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L at 30 °C and pH = 6 for 60 min).
Figure 8. Effect of initial ions concentration on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L at 30 °C and pH = 6 for 60 min).
Sustainability 17 02817 g008
Figure 9. Effect of time on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 250 mg/L at 30 °C and pH = 6).
Figure 9. Effect of time on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 250 mg/L at 30 °C and pH = 6).
Sustainability 17 02817 g009
Figure 10. Effect of temperature on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 250 mg/L at pH = 6 for 60 min).
Figure 10. Effect of temperature on sorption capacity and removal efficiency of raw and mechanically activated phosphogypsum (adsorbent dose = 1 g/L and initial ion concentration = 250 mg/L at pH = 6 for 60 min).
Sustainability 17 02817 g010
Figure 11. Performance of mechanically activated adsorbent after five cycles of regeneration.
Figure 11. Performance of mechanically activated adsorbent after five cycles of regeneration.
Sustainability 17 02817 g011
Table 1. XRF analysis of phosphogypsum analysis.
Table 1. XRF analysis of phosphogypsum analysis.
ItemCaONa2OMgOAl2O3SiO2P2O5SO2LOI
%36.50.210.120.633.34.234.8516.5
Table 2. Parameters of the Langmuir and Freundlich isotherm models.
Table 2. Parameters of the Langmuir and Freundlich isotherm models.
IsothermParameterCadmiumLead
RawMARawMA
LangmuirR20.85450.89550.84790.9043
qmax (Cal.)36510044
qmax (Exp.)171214138169
b6.5 × 10−34.5 × 10−3
FreundlichR20.99660.99400.99860.9965
n1.351.341.321.27
KF6.9115.773.915.78
Table 3. The parameters of pseudo-first-order and pseudo-second-order models.
Table 3. The parameters of pseudo-first-order and pseudo-second-order models.
ItemCadmium IonsLead Ions
Raw MARaw MA
The pseudo-first-order model
R20.97700.97600.97460.9798
K1−5.12 × 10−2−5.36 × 10−2−4.02 × 10−2−5.18 × 10−2
Calculated qe218286216239
Experimental qe176219150174
The pseudo-second-order model
R20.99290.99590.99480.9944
K22.98 × 10−42.69 × 10−42.33 × 10−43.18 × 10−4
Calculated qe188250172191
Experimental qe176219150174
Table 4. Comparison of phosphogypsum with other reported adsorbents for Pb2+ and Cd2+ removal.
Table 4. Comparison of phosphogypsum with other reported adsorbents for Pb2+ and Cd2+ removal.
AdsorbentAds. Capacity (mg/g)pHReference
Pb2+Cd2+
Alkali-modified almond shells 975-6[61]
Magnetic biochar40165.0[62]
Modified reed biochar1737.0[63]
Bone composite29424.0[64]
Modified dairy manure biochar175687.0[65]
Activated carbon9.39.26.3[66]
Zirconium organic frameworks1761956.0[67]
Ageratum conyzoid biomass30366.0[68]
Phosphogypsum2302346.0This work
Activated phosphogypsum2352436.0This work
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

Alotaibi, A.M.; Aljabbab, A.A.; Alajmi, M.S.; Qadrouh, A.N.; Farahat, M.; Abdel Khalek, M.A.; Baioumy, H.; Alhumimidi, M.S.; Almutairi, R.S.; Alkhammali, S.A. Eco-Friendly Utilization of Phosphogypsum via Mechanical Activation for Sustainable Heavy Metal Removal from Wastewater. Sustainability 2025, 17, 2817. https://doi.org/10.3390/su17072817

AMA Style

Alotaibi AM, Aljabbab AA, Alajmi MS, Qadrouh AN, Farahat M, Abdel Khalek MA, Baioumy H, Alhumimidi MS, Almutairi RS, Alkhammali SA. Eco-Friendly Utilization of Phosphogypsum via Mechanical Activation for Sustainable Heavy Metal Removal from Wastewater. Sustainability. 2025; 17(7):2817. https://doi.org/10.3390/su17072817

Chicago/Turabian Style

Alotaibi, Abdulrahman M., Abdulrahman A. Aljabbab, Mamdoh S. Alajmi, Ayman N. Qadrouh, Mohsen Farahat, Mohamed Abdeldayem Abdel Khalek, Hassan Baioumy, Mansour S. Alhumimidi, Ramzi S. Almutairi, and Sultan A. Alkhammali. 2025. "Eco-Friendly Utilization of Phosphogypsum via Mechanical Activation for Sustainable Heavy Metal Removal from Wastewater" Sustainability 17, no. 7: 2817. https://doi.org/10.3390/su17072817

APA Style

Alotaibi, A. M., Aljabbab, A. A., Alajmi, M. S., Qadrouh, A. N., Farahat, M., Abdel Khalek, M. A., Baioumy, H., Alhumimidi, M. S., Almutairi, R. S., & Alkhammali, S. A. (2025). Eco-Friendly Utilization of Phosphogypsum via Mechanical Activation for Sustainable Heavy Metal Removal from Wastewater. Sustainability, 17(7), 2817. https://doi.org/10.3390/su17072817

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