Next Article in Journal
The Efficiency of Chlorella vulgaris in Heavy Metal Removal: A Comparative Study of Mono- and Multi-Component Metal Systems
Previous Article in Journal
Member Size Effect in Seebeck Coefficient of Cement Composites Incorporating Silicon Carbide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydroxyapatite-Based Adsorbent Materials from Aquaculture Waste for Remediation of Metal-Contaminated Waters: Investigation of Cadmium Removal

1
Department of Chemical, Pharmaceutical and Agricultural Sciences, University of Ferrara, Via Luigi Borsari 46, 44121 Ferrara, Italy
2
Department of Environmental and Prevention Sciences, University of Ferrara, Via Luigi Borsari 46, 44121 Ferrara, Italy
3
Physics and Earth Sciences Department, University of Ferrara, Via Saragat 1, 44122 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Clean Technol. 2025, 7(2), 34; https://doi.org/10.3390/cleantechnol7020034
Submission received: 22 November 2024 / Revised: 27 February 2025 / Accepted: 17 March 2025 / Published: 14 April 2025

Abstract

:
Adsorption represents an effective strategy for water remediation applications, particularly when utilising eco-friendly materials in a circular economy framework. This approach offers significant advantages, including low cost, material availability, ease of operation, and high efficiency. Herein, the performance of cadmium ion adsorption onto hydroxyapatites, derived through a calcination-free process from shells of two mollusc species, Queen Scallop (Aequipecten opercularis) and Pacific Oyster (Magallana gigas), is examined. The phase and morphology of the synthesised adsorbents were investigated. The results showed that hydroxyapatites obtained from mollusc shells are characterised by high efficiency regarding cadmium removal from water, exhibiting rapid kinetics with equilibrium achieved within 5 min and high adsorption capacities up to 334.9 mg g−1, much higher than many waste-based adsorbents reported in literature. Structural investigation revealed the presence of Cadmium Hydrogen Phosphate Hydrate in the hydroxyapatite derived from oyster shells loaded with Cd, indicating the formation of a solid solution. This finding suggests that the material not only has the capability to decontaminate but also to immobilise and store Cd. Overall, the results indicate that hydroxyapatites prepared via a synthetic route in mild conditions from waste shells are an economical and efficient sorbent for heavy metals encountered in wastewater.

1. Introduction

The introduction of toxic transition metals into the environment can arise from both natural and anthropogenic sources. Metal contamination is particularly relevant in freshwater ecosystems where it can be caused by atmospheric agents, or through the discharge of agricultural, domestic, and industrial waste [1]. Due to their toxicity and tendency to bioaccumulate, heavy metals represent a major concern for human health and ecosystems. They can cause adverse effects on plants, animals [2] and humans; indeed, toxic metals interfere with physiological processes, particularly hematopoietic, hepatic, and renal functions, and the nervous system [3]. Among heavy metals, cadmium (Cd) has an extremely long biological half-life, making it a cumulative toxic metal [4]. Furthermore, Cd has been identified as a human carcinogen by the World Health Organization’s International Agency for Research on Cancer and the United States [4]. Various strategies can be employed to remove metal pollutants from water [5]. Chemical precipitation is cost-effective, has low energy requirements and is safe and simple, but it produces high volumes of sludge, incurring associated disposal costs [5,6]. Coagulation and flocculation effectively reduce the settling time of suspended solids; however, the cost of sludge disposal remains a concern [6]. Membrane filtration is a highly efficient process that requires minimal operational space, but it is expensive and prone to fouling [5,6,7]. Electrochemical methods are highly efficient and selective processes, which do not produce sludge, but they are costly and characterised by high energy consumption [5,8]. Adsorption is another viable solution for the remediation of waters contaminated by metals [5,9,10], particularly when treating large volumes. It is characterised by simple design and operation, high efficiency and selectivity. Nonetheless, efficient adsorbents can be expensive [5], making it crucial to find economic and suitable alternatives. Therefore, developing solutions and evaluating adsorbent materials to remove pollutants within a circular economy context is a global challenge, and the valorisation of waste and biomass through recycling and transformation into new materials is a promising approach [11]. Recent studies on water remediation strategies offer various solutions for the reuse of waste materials to obtain eco-friendly adsorbents for heavy metals removal. Various bio-waste materials have been investigated including, for example, egg and mollusc shells [10,12,13], peanut shells [14,15], peanut shells biochar [16], and coconut husk [17]. Additionally, many calcium carbonate derived materials, such as apatite, have also been evaluated for the removal of Cd from aqueous matrix [18]. Among these materials, hydroxyapatite (HA, Ca10(PO4)6(OH)2) [19] is more stable than carbonates and is a low cost, largely available and biocompatible material [20]. Several studies have investigated the application of HA synthetised using pure reagents, or commercial HA for water remediation [21,22,23,24]. In the contest of a circular economy, biogenic calcium carbonate derived from seafood processing can indeed be valorised as a reagent for the synthesis of HA. However, recent literature, indicates that the synthesis of HA from biomass waste requires high temperatures to prepare the reagents through calcination [25,26,27,28], as well as a hydrothermal method for synthesis, both of which significantly impact the energy footprint of the process [28,29,30,31]. Well-known materials for the synthesis of hydroxyapatite (HA) are mollusc shells waste produced by aquaculture activities, which typically consists of 95–99% calcium carbonate (CaCO3) by weight. These materials are widely available with an estimated global production of 10 million tons/annually and constitute critical wastes in areas where shellfish cultivation is a well-established activity. Indeed, aquaculture activities have to bear disposal costs to discard mollusc shells; the valorisation of this waste material through transformation into efficient adsorbents for water remediation can significantly reduce disposal costs [32]. Since shell morphology and composition play a critical role in influencing the interface properties and the chemical and mechanical behaviour [10,33] of the resulting material, HA prepared from shells of different species may show variations in the contaminant uptake capacity. For this reason, in the present study shells of two different species, namely Queen Scallop (Aequipecten opercularis) and Pacific Oyster (Magallana gigas), were evaluated as substrate for the synthesis under mild conditions of HA based adsorbent materials.
In this work, HA was synthesised from both raw and calcined shells (the latter requires an energetically demanding process). Cd adsorption experiments were conducted on both materials to compare the efficiency of the two adsorbents and to assess the possible use of the lower energy impact HA in water remediation applications. The mechanism of Cd uptake onto this novel promising material was investigated combining kinetic and isotherm adsorption experiments with structural and thermal analyses on pristine and Cd-loaded HAs. The features that influence the sorption capacity of the adsorbents were evaluated.

2. Materials and Methods

2.1. Materials

Pacific oyster (Magallana gigas) shells (OS), shown in Figure 1a, and Queen scallop (Aequipecten opercularis) shells (SS), depicted in Figure 1b, were provided from a shellfish farm and a fish products retailer located on the North Adriatic Sea coast (Naturedulis S.r.l., Goro (FE), Italy). The shells were thoroughly washed with ultrapure water (MilliQ®, Merck KGaA, Darmstadt, Germany) and dried at 105 °C. The dried materials were then ground into a fine powder using a jade ball mortar (Retsch GmbH, Haan, Germany).
Cadmium, Calcium and Phosphorus standard solutions (1000 mg L−1) for ICP were purchased from Carlo Erba Reagents S.r.l., Italy; Multielement standard solution (10–100 mg L−1) for ICP, Sodium and Potassium standard solution (1000 mg L−1) for AAS, were purchased from Merck, Darmstadt, Germany. Nitric Acid (69%, Suprapur®); Cadmium nitrate tetrahydrate and Sodium phosphate dibasic were purchased from Merck, Darmstadt, Germany; Hydrochloric Acid (37%, Superpure) was obtained from Carlo Erba Reagents, Milan, Italy. Ultrapure water was used to prepare all solutions.

2.2. Preparation of the Adsorbents

2.2.1. Shells Calcination

The calcination is an endothermic reaction, as shown in Equation (1) [34], therefore is favoured by higher temperatures; in this work, powdered shells were calcined in a muffle at 900 °C for one hour.
CaCO3 → CaO + CO2 ΔH = 182.1 kJ mol−1
The weight of both materials were monitored, and 69.1% by weight of calcined oyster shells (cOS) and 76.5% by weight of calcined scallop shells (cSS) were obtained.

2.2.2. Hydroxyapatite Synthesis

Hydroxyapatites (HAs) were synthesised from OS, SS, cOS and cSS following the method proposed by Zhang et al. [31]. 5 g of powdered shells were treated with 40 mL of 1.33 M Na2HPO4 solution to achieve a 1.67 Ca/P molar ratio; the mixture was kept under stirring at room temperature (22.0 ± 0.5 °C) for 48 h. Subsequently, the heterogeneous solution was filtered with ashless filter paper with pore size 4–12 μm (Whatman, Maidstone, UK). The obtained materials are hereafter referred as: HAOS (hydroxyapatite from oyster shells), HASS (hydroxyapatite from scallop shell), HAcOS (hydroxyapatite from calcined oyster shell) and HAcSS (hydroxyapatite from calcined scallop shell). All materials were dried in the oven at 105 °C for 24 h.

2.3. Characterisation

2.3.1. Scanning Electron Microscope and Transmission Electron Microscopy Measurements

Scanning Electron Microscopy (SEM) imaging was performed with a Zeiss Evo 40 electron microscope (Carl Zeiss AG, Oberkochen, Germany) equipped with Energy Dispersive x-ray Spectroscopy (EDS) (Oxford Instruments plc, Abingdon, UK). For SEM imaging acquisition the samples were coated with Au for 2 min (5–6 nm min−1) with turbomolecular pump sputter coater system (Q150TS Plus, Quorum Technologies Ltd., Laughton, UK). The images were acquired using Backscattered electrons (BSE) and Secondary Electrons (SE) with an Electron High Tension (EHT) of 15 kVat magnifications from 1.00 K X to 11.00 K X.
Transmission electron microscopy (TEM) imaging was performed with Talos L120C (Thermo Fisher Scientific Inc., Waltham, MA, USA). The images were acquired at magnifications from 120 K X to 190 K X.

2.3.2. Particle Size Distribution Measurements

Particle size distribution of HAs samples were evaluated with a Mastersizer 3000 laser diffraction particle size analyser (Malvern Panalytical Ltd., Malvern, UK); the analyses were performed in triplicate using 0.2 g of sample for each replicate [35].

2.3.3. Chemical Characterisation

The shells’ composition was determined after sample mineralisation. 0.3 g of HASS and HAOS were weighed in TMF (tetrafluoro-methoxyl) vessels, added with 7 mL of HNO3 and 3 mL of HCl and mineralised with a microwave-assisted digestion system (details can be found in Supplementary Materials, Experimental S1). Samples were mineralised in triplicate.
Trace and major elements were quantified by inductively coupled plasma mass spectrometry and atomic spectroscopy; the instrumental conditions are reported in Supplementary Materials (Experimental S1).

2.3.4. X-Ray Powder Diffraction (XRPD)

As synthesised and Cd loaded HAOS and HASS were ground in an agate mortar with a pestle for X-Ray Powder Diffraction (XRPD) analyses. The details of the instrumental conditions are described in Supplementary Materials, in the Experimental S1 Section.
Data elaboration was performed with Bruker AXS EVA software (version 5) and TOPAS (version 5.0). All diffraction peak profiles were modelled employing Lorentzian broadening coefficients in conjunction with shifted Chebyshev polynomials. In each Rietveld structure refinement (RSR), we refined several parameters, including the background polynomial, scale factor, unit cell parameters, zero-point correction, and site occupancy factors, while atom positions and isotropic thermal parameters were held constant. The weight percentage (wt %) of each Bragg diffracting phase was calculated based on the refined scale factors.
The identified phases were further refined using the scale factor, unit-cell parameters, crystallinity percentage, and crystallite size (in nanometres). The crystallite size was determined utilising the Scherrer equation applied to the Full Width at Half Maximum (FWHM) of the hydroxyapatite characteristic peak within the 25–26° 2θ range. The detected phases were refined starting from the following space groups: R-3c (ICSD code 20179) for calcite, Pmcn (ICSD code 15198) for aragonite, P_63/m (ICSD code 26204) for hydroxyapatite, and C 2/c (ICSD code 10045) for cadmium hydrogen phosphate hydrate.

2.3.5. Thermal Analysis

Thermogravimetric (TGA) and Differential Thermal Analysis (DTA) were conducted on both bare and Cd-loaded HAOS and HASS samples under constant and controlled synthetic air flow conditions, utilising a STA 409 PC LUXX® instrument from Netzch (Gerätebau, Germany). The heating rate was maintained at 10 °C min−1, with temperatures ranging from room temperature (RT) to 1100 °C. Approximately 30 mg of each sample was employed for the analyses, with alumina (Al2O3) serving as the control and standard reference.

2.4. Batch Adsorption Experiments

Adsorption experiments were performed using batch method [36] in 20 mL glass vials equipped with PP screw caps with PTFE septa. All the tests were conducted at pH 7 to prevent the precipitation of Cd2+ as cadmium hydroxide [37,38]. Additionally, the pH value selected for the adsorption tests is within the pH range typical of surface waters [39,40,41]. To determine the adsorption kinetics, experiments were carried out using solutions with initial Cd concentration of 10 mg L−1; the solution was mixed with the adsorbent material with a solid/liquid ratio of 5:1 (mg mL−1) and kept under continuous stirring at 25.0 ± 0.5 °C. Cd2+ uptake was measured after contact times equal to 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 15, 20, 30, 40, 50, 60, 90, 120, 150 and 180 min, after each contact time the solution was filtered with a 25 mm PVDF syringe filter with a pore size of 0.45 µm (Agilent Technologies, Santa Clara, CA, USA). The pH value was measured at the beginning and at the end of each experiment, the recorded values were in the range 7.0 ± 0.5. Equilibrium adsorption tests were performed by dispersing HA in Cd2+ aqueous solution at different initial concentrations (5, 20, 40, 60, 100, 125, 150, 175, 200, 250, 300, 350 and 400 mg L−1) using a solid to liquid ratio of 1:1 (mg mL−1). The solutions were maintained at 25.0 ± 0.5 °C under stirring for 24 h, a contact time adequately longer compared to the equilibration time evaluated by kinetic experiments. The solutions were then filtered, and the pH was measured for each batch. All experiments were carried out in triplicate. ICP-OES was employed to quantify Cd and Ca in solution before and after the contact with HAs for both kinetic and equilibrium adsorption experiments, with conditions reported in Section 2.3.5.
The amount of Cd2+ adsorbed per gram at equilibrium, defined as q (mg g−1) was calculated with the mass balance Equation (2) [42]:
q = C 0 C e V m
where C0 (mg L−1) is the initial solution concentration, Ce (mg L−1) is the concentration at the equilibrium, V (L) is the solution volume of the batch and m (g) the mass of adsorbent.

3. Results and Discussion

3.1. Materials Characterisation

Unprocessed grounded shells of each species (SS, OS, cSS, cOS) and synthesised HAs (HAOS, HAcOS and HASS, HAcSS) were firstly characterised by SEM to obtain information on the morphology. Subsequently, the particle size distribution was evaluated for each sample of HAs synthesised. Structural investigation and elemental bulk composition of HAOS and HASS were carried out before the adsorption tests to support the results obtained; the results are reported in Section 3.3 and Supplementary Discussion S1.

3.1.1. SEM/EDS and TEM Imaging

Surface morphology of the synthesised HAs and unprocessed shells were analysed by scanning electron microscopy (SEM). The images collected for each material are reported in Figure 2 and Figures S1–S8. Specifically, HAOS (Figure 2a) and HASS (Figure 2b) exhibited particles characterised by irregular shape and variable size. Additionally, in Figures S1–S8 the elements distributions determined by scanning electron microscopy (SEM) equipped with energy dispersive X-ray spectrometry (EDS) are reported. The calcium is homogeneously distributed in each shell sample (Figures S1–S4), which is reasonable considering that calcium carbonate constitutes 95% of the mollusc shell weight [43]. Furthermore, the elemental maps of HAs (Figures S5–S8), showed a high concentration and homogeneous distribution of phosphorus, oxygen and calcium, which are the main components of HAs.
Transmission electron microscopy analyses were performed on all sythesised materials, the images were acquired on the smaller particles to obtain maximum resolution (Figures S9 and S10). Differences can be observed between the raw materials (i.e., CaCO3 from SS and OS) and HASS and HAOS. Moreover, the comparison between HAs from raw powdered shells and HAs from calcined powdered shows that the materials synthesised from CaCO3 present a more laminar structure, whilst those prepared from CaO have a more compacted structure.

3.1.2. Particle Size Distribution

Figure 3 shows the particle size distribution of each HA. The data obtained suggest that the particle size dimensions of HAs span over a broad range, with HAs from calcined shells more homogeneously distributed than HASS and HAOS, despite a similar average distribution (Dv 50) as shown in Table 1. Furthermore, about 50% of each material is less than 10 µm.

3.2. Batch Adsorption Results

3.2.1. Effect of Calcination in Adsorption Experiments

To compare the adsorption properties of the adsorbent materials, preliminary tests were performed in triplicate at three different initial concentrations (5, 20 and 40 mg L−1), following the method reported in paragraph 2.4. The results reported in Figure 4 show that all the materials exhibit similar adsorption quantities of Cd2+ in the examined concentration range. Since, HAs synthesised from both uncalcined mollusc shells have a lower environmental impact than those obtained from calcined shells, only the most sustainable materials, namely HASS and HAOS were evaluated in the following.

3.2.2. Kinetics Experiments

To gain information on the equilibrium time and the kinetics behaviour of the adsorption process, the Cd2+ uptake onto HASS and HAOS was investigated. Batch experiments were conducted as indicated in Section 2.4.
The experimental data collected were fitted using pseudo second order (PSO) kinetic model [10], given by Equation (3):
q t = k 2   q e q   2 t 1 + k 2   q e q   t
where q t and q e q are the adsorbed quantity per unit mass after a specific contact time t (min) and at the equilibrium respectively, and k 2 (g mg−1 min−1) is the second order adsorption rate constant. The PSO model has already proved to be suitable to describe the adsorption of heavy metals [10]. The adsorption kinetics of HASS and HAOS are reported in Figure 5; it can be seen that for both materials the adsorption equilibrium was reached in about 5 min. The kinetic parameters obtained by nonlinear fitting regression of the experimental data using Equation (3) are reported in Table 2. From the observed value of k 2 , the uptake of Cd2+ onto HAOS is faster than HASS. Furthermore, HAs synthesised in this work showed faster kinetics compared to those of HAs reported in literature. For example, Li et al. [22] reported values of k2 of 1.64 and 4.28 g mg−1·min−1 for HAs synthesised from pure reagents, while in Silva-Holguín et al. [21] HA supported on alumina spheres showed a k2 of 0.012 g mg−1·min−1 [21]. Moreover, the obtained k 2 values are much higher compared to the PSO rate constants of the precursor materials; in particular, HASS and HAOS showed an extremely faster Cd2+ adsorption compared to SS [10] and OS [44,45]. Indeed, at the same Cd2+ initial concentration, Chenet et al. [10] reported a k2 for SS of 0.87 g mg−1·min−1, [10] while for OS values of k2 of 0.0015 and 0.018 g mg−1·min−1 are indicated [44,45]. Therefore, the HAs derived from mollusc shells enabled an improvement in the kinetics of the adsorption process when compared to the starting materials intended for the same application. This very fast Cd2+ uptake suggests that both HASS and HAOS are suitable materials for water remediation applications; the very short time required to reach the equilibrium indicates a suitability also for continuous flow systems [46].

3.2.3. Adsorption Isotherms

The adsorption capacities of the adsorbent materials were obtained by batch experiments. The experimental data collected were fitted using the Langmuir isotherm model, Equation (4):
q e = q s   b   C e 1 + b   C e
where qs (mg g−1) is the saturation capacity, Ce (mg L−1) is the Cd concentration at the equilibrium, and b (L mg−1) is the Langmuir constant which represents the affinity between adsorbent material and adsorbate.
This isotherm model describes a homogeneous monolayer adsorption of the adsorbate onto the material surface, without adsorbate-solvent and lateral interactions between adsorbate molecules. The Langmuir model is suitable to describe adsorption isotherms characterised by a concave curve with a steep first interval followed by a plateau which represents the saturation point, qs [42,47,48].
Isotherms data were fitted also with the Freundlich model, expressed by Equation (5):
q e = k F C e 1 n
where kF ((mg g−1) (L mg−1)1/n) is the Freundlich constant which reflects the affinity between adsorbent and adsorbate and n is related to the sorption intensity, particularly, values of n > 1 indicate favourable adsorption conditions. The Freundlich isotherm model is suitable for describing heterogeneous adsorption with an increase of the uptake onto the adsorbent when the species concentration in the aqueous phase is increased [47,48].
Figure 6 shows the experimental data fitted with Equations (4) and (5), and the estimated parameters are reported in Table 3. The correlation coefficient, R2, allows to compare the goodness of the fitting between models; the R2 values obtained show that each isotherm model selected is suitable for describing the experimental data; however, Langmuir model describes better the adsorption onto HAOS, whereas the uptake onto HASS is better described by the Freundlich model.
The parameters obtained with the Freundlich isotherm model show n values higher than 1, particularly 10.63 ± 3.72 and 3.97 ± 0.26 for HASS and HAOS, respectively. These results suggest that adsorption is achieved under favourable conditions, confirmed by high KF values of 112.90 ± 14.87 and 105.90 ± 12.19 for HASS and HAOS, respectively (Table 3). Indeed, in the literature there are many works that demonstrate the high affinity between metal cations and HA adsorbents. For example, Núñez at al. (2019) [49] obtained an KF value of 15.86 (mg g−1) (L mg−1)1/n for HA from clam shells, and Li et al. (2017) [22] found a KF of 1.04 (mg g−1) (L mg−1)1/n from HA synthesised from pure reagents.
Langmuir isotherm model allows to obtain the values of qs (mg g−1); the results indicate high saturation capacities for the materials investigated, especially for HAOS with a qs of 334.3 ± 46.20 mg g−1, which is higher than that obtained with other adsorbent materials reported in the literature (Table 4). Due to the biogenic origin of the starting materials used for the synthesis of HASS and HAOS, and the consequent complexity of the adsorbent structure which will be discussed in Section 3.3, the Cd adsorption may involve more than one mechanism.
A mechanism for the metal uptake onto hydroxyapatites was proposed by Marchat et al. [50], suggesting that the adsorption of Cd onto HA occurs through two mechanisms: adsorption and co-precipitation.
At the beginning, Cd2+ occupies individual sites on the HA surface through exchange with Ca2+ or by forming complexes. Then, residual Cd2+ may diffuse into the adsorbent structure or form a Cd containing HA phase [50]. Considering the adsorption results described above, the HAs synthesised from scallop and oyster shells are indeed very promising materials for water remediation in the case of heavy metals contamination.
Even though a specific cost analysis is beyond the objective of the present work, the high adsorption capacities and fast adsorption kinetics of HASS and HAOS indeed indicate their suitability for water remediation in a cost-effective way since the materials show very high saturation capacities compared to other materials.
Moreover, the high efficiency demonstrated by the HAs synthetised directly from mollusc shells and skipping the costly calcination step, suggest a lower synthesis cost since it utilises “zero-cost” raw materials; the major contributions to the total cost would be given by transportation and reagents [51].
Table 4. Adsorption capacity (qs (mg g−1)) towards Cd2+ of different adsorbent materials fitted with Langmuir model.
Table 4. Adsorption capacity (qs (mg g−1)) towards Cd2+ of different adsorbent materials fitted with Langmuir model.
Adsorbentsqs (mg g−1)References
Bamboo charcoal12.08[52]
Oyster shells (OS)15.03[45]
Cashew nutshell21.11[53]
Sugar cane48.31[54]
Scallop shells (SS)55 ± 17.4[10]
Orange peel56.5 ± 0.50[55]
Mussel shells HA62.5[25]
Clam shells HA62.5[49]
Small surface area HA71.94[22]
Pine bark biochar (600 °C)85.5[56]
Alumina-HA Spheres89.37[21]
Canna indica biochar (600 °C)140.01 ± 14.42[57]
HASS173.7 ± 16.95Present study
Large surface area HA207.97[22]
nano-HA243.90[18]
HAOS334.3 ± 46.20Present study

3.3. Structural and Thermal Characterisation

The X-ray diffraction patterns of pristine and Cd-loaded HAOS and HASS are given in Figure 7. The pattern profiles are characterised by broad peaks, which can be attributed to the diffracting particles poor crystallinity and high defect density (crystallites). A close examination of the diffraction profiles present in Figure 7 reveals a pronounced anisotropy in the peak broadening.
Hydroxyapatite’s 001 reflections exhibit significantly sharper peaks compared to other reflections within the patterns. This striking anisotropy in the profile function of the hydroxyapatite samples indicates a presence of defects in the a, b planes and/or an anisotropic shape of the crystallites, with coherent domains that are elongated along the c axis. Such features are commonly observed in hydroxyapatite derived from biomass waste [58], suggesting that the coherent diffraction domains preferentially extend along the crystallographic c direction within the hydroxyapatite crystals.
Quantitative phase analysis (QPA) was conducted by the Rietveld method, wherein the weight fraction wi of all ith crystalline component in the multiphase system is calculated from the corresponding refined scale parameter Si, according to the Equation (6):
w i = S i M i V i j S j M j V j   w i t h   t h e   n o r m a l i s a t i o n   c o n d i t i o n   Σ w i   =   1.0  
where Mi and Vi are the unit cell mass and volume, respectively.
Calcite was observed at approximately 23% weight, alongside hydroxyapatite (HA), as indicated in Table 5. This observation suggests an incomplete transition to HA in both the HAOS and HASS samples. Additionally, a low fraction of aragonite, approximately 4% weight, was detected in the HASS sample, also detailed in Table 5. The diffraction peaks corresponding to these secondary phases are distinctly marked in Figure 7.
Commonly, cadmium fixation in hydroxyapatite occurs in accordance with the following mechanisms [50,59]: Cd2+ adsorption on specific P-OH groups of HA surface, diffusion of the excess Cd2+ into the apatite structure, ion exchange, and heterogeneous dissolution and precipitation.
The exchange reaction impacts the structural integrity of hydroxyapatite and its chemical composition, leading to a solid solution incorporating Cd2+. This incorporation can alter the crystallinity and porosity of HA (Table 6), which are critical factors in its adsorption capacity. Additionally, the presence of carbonate ions in the solution can facilitate the formation of mixed carbonate-phosphate phases, further influencing the removal efficiency of cadmium.
Following cadmium adsorption, the cell volume of hydroxyapatite decreased in both HASS and HAOS. This reduction can be attributed to the smaller ionic radii of cadmium (0.95 Å and 1.10 Å for six- and eight-coordination, respectively) compared to calcium (1.00 Å and 1.12 Å for six- and eight-coordination). These findings indicate an exchange reaction between the two elements, resulting in the formation of a Ca1−xCdxCO3 and Ca(5−x)Cdx(PO4)3 (OH) solid solution [59].
This process introduces strain into the HA network structure and increases the HA decomposition [60].
The dissolution of CaCO3 and subsequent Cd adsorption on the surface progressively altered the regular rhombohedral etch pits of calcite, as cadmium ions pinned the acute angles. This process effectively passivated the calcite surface through the epitaxial growth of a less soluble (Ca,Cd)CO3 layer [61,62]. After Cd2+ adsorption, the crystallite size increased in both HAOS and HASS samples, ranging from approximately 25 to 32 nm, while the crystallinity percentage remained nearly unchanged (Table 5).
In the Cd-treated HAOS sample, a new solid phase has been identified as Cadmium Hydrogen Phosphate Hydrate, Cd5H2(PO4)4·4H2O, constituting approximately 13% in weight. This finding suggests a dissolution-precipitation mechanism involving hydroxyapatite. The formation of this new phase is influenced by multiple factors, including the liquid-to-solid ratio, pH, ionic strength, temperature, duration, concentration of the aqueous phase, and the presence of other ions. Additionally, the types and pathways of geochemical reactions play a significant role in this process.
The formation of this phase in Cd-treated HAOS could be due to the different composition of oyster and scallop shells [10,63] (see Discussion S1) and it could explain the higher saturation capacity of HAOS compared to HASS, despite the fact that roughly 4% of aragonite was detected in HASS, but not in HAOS. Aragonite, indeed, facilitates additional Cd sequestration during the Ca/Cd replacement reaction compared to calcite, leading to fracturing and interfacial stress, as well as the incorporation of a greater Cd content [61,64]. It has been already pointed out that biogenic carbonates showed higher adsorbent capacities than non-biogenic one’s due to structural and compositional features. The results of the present study also suggest that hydroxyapatites synthesised from different biogenic carbonates have different adsorption performances.
The thermal characterisation of bare and loaded samples of HAOS and HASS revealed comparable weight loss, approximately 20% by weight, at the maximum investigated temperature of 1100 °C. Throughout the heating process, several distinct steps were observed (see Figure 8 and Figure 9). Initially, the release of physically adsorbed water from hydroxyapatite occurs between 30 and 200 °C, associated with an endothermic reaction evident in the differential thermal analysis (DTA) curve.
At temperatures ranging from 200 to 400 °C, a second thermal event occurs, characterised by an exothermic effect in the DTA, indicating the loss of structural water from HA. As the temperature increases to approximately 600 °C, hydroxyapatite undergoes gradual dehydration, resulting in the release of hydroxide ions (OH) and subsequently transforming into oxyapatite. Above this temperature, significant weight loss is primarily attributed to the thermal decomposition of residual CaCO3 and potentially amorphous carbonate. According to the literature, calcite is recognised as the most thermodynamically stable polymorph of calcium carbonate under ambient conditions [65]. Consequently, the decarbonation of the less stable polymorph, aragonite, begins at around 450 °C, while the decomposition of calcite occurs at elevated temperatures ranging from 720 to 875 °C. It is important to note that the temperature at which the exothermic peak occurs can vary depending on the sample’s composition [66]. Furthermore, cadmium ions bound to HA and CaCO3 can induce a shift in the decarboxylation and dihydroxylation reactions to lower temperatures [67]. The differential thermal gravimetry (DTG) curve of Cd-loaded hydroxyapatite (HASS) exhibited two prominent peaks at 430 °C and 600 °C. In contrast, the incorporation of Cd2+ into the hydroxyapatite structure, along with the mineralisation of Cd5H2(PO4)4·4H2O, resulted in a shift of the curves to 368 °C, 460 °C, and 520 °C. This was accompanied by three exothermic reactions observed in the differential thermal analysis (DTA).

4. Conclusions

The Hydroxyapatites studied in this work were synthesised from not calcined shells and without any thermal process. The results showed higher saturation capacities towards the adsorption of Cd2+ onto HAs and faster kinetics than those reported for mollusc shells as adsorbents. The adsorption performances of the synthesised hydroxyapatites are strongly dependent on the type of the shell employed for the synthesis; indeed, saturation capacities of 173.7 ± 16.95 mg g−1 and 334.3 ± 46.2 mg g−1 were obtained for HASS and HAOS, respectively. These differences may be attributed to variations in the composition and structure of the starting materials, as evidenced from diffractometric and bulk composition characterisation. These findings suggest that hydroxyapatites from mollusc shells are suitable materials for water remediation applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cleantechnol7020034/s1, Experimental S1: Instrumental conditions; Discussion S1: Elements bulk composition; Table S1: Elements with respective emission lines for ICP-OES analyses; Table S2: Elements bulk composition of HASS and HAOS. The molar Ca/P ratios were calculated considering the percentage of HAs indicated by XRD (HASS 73%, HAOS 75.5%); Figure S1: SEM images of SS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), C (e), O (f). Figure S2: SEM images of cSS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), C (e), O (f); Figure S3: SEM images of OS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), C (e), O (f); Figure S4: SEM images of cOS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), C (e), O (f); Figure S5: SEM images of HASS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), O (e), P (f); Figure S6: SEM images of HAcSS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), O (e), P (f); Figure S7: SEM images of HAOS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), O (e), P (f); Figure S8: SEM images of HAcOS (a,b) 5.00 K X and 11.00 K X respectively; SEM image of the area analysed for elemental map collection (c); elemental mapping of Ca (d), O (e), P (f); Figure S9: TEM images of SS (a,b), cSS (c,d), OS (e,f), cOS (g,h) with magnification of 120.00 K X and 190.00 K X. Acquisition parameters are reported on each image; Figure S10: TEM images of HASS (a,b), HAcSS (c,d), HAOS (e,f), HAcOS (g,h) with magnification of 120.00 K X and 190.00 K X. Acquisition parameters are reported on each image.

Author Contributions

Conceptualisation, L.P., T.C. and M.C.; methodology, M.C., M.M. and T.C.; software, M.C., T.C. and M.M.; validation, F.C., T.C., M.C. and A.M.; formal analysis M.C., T.C., M.M. and F.C.; investigation, M.C., T.C., M.M. and F.C.; resources, L.P. and A.M.; data curation, M.C., T.C., M.M. and C.S.; writing—original draft, M.C.; writing—review & editing, L.P., T.C., C.S., F.C., M.M. and A.M.; visualisation, C.S., M.C., T.C. and M.M.; supervision, L.P. and A.M.; project administration, L.P. and A.M.; funding acquisition, L.P., C.S., A.M., M.C. and F.C. All authors have read and agreed to the published version of the manuscript.

Funding

The contribution at this work of M.C. and C.F. was supported by the program PON “Research and Innovation” 2014–2020 (PON R&I), Action IV.6 “Contratti di ricerca su tematiche Green”; L.P., C.S. and A.M. acknowledge financial support from PNRR MUR project ECS_00000033_ECOSISTER.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to acknowledge Leonardo Aguiari of Naturedulis S.r.l. company for the supply of waste material and Antonella Pagnoni for practical help and scientific suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kumar, V.; Parihar, R.D.; Sharma, A.; Bakshi, P.; Singh Sidhu, G.P.; Bali, A.S.; Karaouzas, I.; Bhardwaj, R.; Thukral, A.K.; Gyasi-Agyei, Y.; et al. Global Evaluation of Heavy Metal Content in Surface Water Bodies: A Meta-Analysis Using Heavy Metal Pollution Indices and Multivariate Statistical Analyses. Chemosphere 2019, 236, 124364. [Google Scholar] [CrossRef] [PubMed]
  2. Suhani, I.; Sahab, S.; Srivastava, V.; Singh, R.P. Impact of Cadmium Pollution on Food Safety and Human Health. Curr. Opin. Toxicol. 2021, 27, 1–7. [Google Scholar] [CrossRef]
  3. Sinicropi, M.S.; Amantea, D.; Caruso, A.; Saturnino, C. Chemical and Biological Properties of Toxic Metals and Use of Chelating Agents for the Pharmacological Treatment of Metal Poisoning. Arch. Toxicol. 2010, 84, 501–520. [Google Scholar] [CrossRef] [PubMed]
  4. Waalkes, M.P. Cadmium Carcinogenesis. Mutat. Res.-Fundam. Mol. Mech. Mutagen. 2003, 533, 107–120. [Google Scholar] [CrossRef]
  5. Vareda, J.P.; Valente, A.J.M.; Durães, L. Assessment of Heavy Metal Pollution from Anthropogenic Activities and Remediation Strategies: A Review. J. Environ. Manag. 2019, 246, 101–118. [Google Scholar] [CrossRef]
  6. Fu, F.; Wang, Q. Removal of Heavy Metal Ions from Wastewaters: A Review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef]
  7. Kurniawan, T.A.; Chan, G.Y.S.; Lo, W.H.; Babel, S. Physico-Chemical Treatment Techniques for Wastewater Laden with Heavy Metals. Chem. Eng. J. 2006, 118, 83–98. [Google Scholar] [CrossRef]
  8. Al-Saydeh, S.A.; El-Naas, M.H.; Zaidi, S.J. Copper Removal from Industrial Wastewater: A Comprehensive Review. J. Ind. Eng. Chem. 2017, 56, 35–44. [Google Scholar] [CrossRef]
  9. Burakov, A.E.; Galunin, E.V.; Burakova, I.V.; Kucherova, A.E.; Agarwal, S.; Tkachev, A.G.; Gupta, V.K. Adsorption of Heavy Metals on Conventional and Nanostructured Materials for Wastewater Treatment Purposes: A Review. Ecotoxicol. Environ. Saf. 2018, 148, 702–712. [Google Scholar] [CrossRef]
  10. Chenet, T.; Schwarz, G.; Neff, C.; Hattendorf, B.; Günther, D.; Martucci, A.; Cescon, M.; Baldi, A.; Pasti, L. Scallop Shells as Biosorbents for Water Remediation from Heavy Metals: Contributions and Mechanism of Shell Components in the Adsorption of Cadmium from Aqueous Matrix. Heliyon 2024, 10, e29296. [Google Scholar] [CrossRef]
  11. Morris, J.P.; Backeljau, T.; Chapelle, G. Shells from Aquaculture: A Valuable Biomaterial, Not a Nuisance Waste Product. Rev. Aquac. 2019, 11, 42–57. [Google Scholar] [CrossRef]
  12. Annane, K.; Lemlikchi, W.; Tingry, S. Efficiency of Eggshell as a Low-Cost Adsorbent for Removal of Cadmium: Kinetic and Isotherm Studies. Biomass Convers. Biorefin. 2021, 13, 6163–6174. [Google Scholar] [CrossRef]
  13. Nakajima, S.; Araki, S.; Sasamoto, R.; Kanda, Y.; Yamanaka, S. Key Particle Properties of Shells for Cadmium Chemisorption. Chemosphere 2022, 287, 132257. [Google Scholar] [CrossRef] [PubMed]
  14. Witek-Krowiak, A.; Szafran, R.G.; Modelski, S. Biosorption of Heavy Metals from Aqueous Solutions onto Peanut Shell as a Low-Cost Biosorbent. Desalination 2011, 265, 126–134. [Google Scholar] [CrossRef]
  15. Villar da Gama, B.M.; Elisandra do Nascimento, G.; Silva Sales, D.C.; Rodríguez-Díaz, J.M.; Bezerra de Menezes Barbosa, C.M.; Menezes Bezerra Duarte, M.M. Mono and Binary Component Adsorption of Phenol and Cadmium Using Adsorbent Derived from Peanut Shells. J. Clean. Prod. 2018, 201, 219–228. [Google Scholar] [CrossRef]
  16. Sattar, M.S.; Shakoor, M.B.; Ali, S.; Rizwan, M.; Niazi, N.K.; Jilani, A. Comparative Efficiency of Peanut Shell and Peanut Shell Biochar for Removal of Arsenic from Water. Environ. Sci. Pollut. Res. 2019, 26, 18624–18635. [Google Scholar] [CrossRef]
  17. Sousa, F.W.; Oliveira, A.G.; Ribeiro, J.P.; Rosa, M.F.; Keukeleire, D.; Nascimento, R.F. Green Coconut Shells Applied as Adsorbent for Removal of Toxic Metal Ions Using Fixed-Bed Column Technology. J. Environ. Manag. 2010, 91, 1634–1640. [Google Scholar] [CrossRef]
  18. Salah, T.A.; Mohammad, A.M.; Hassan, M.A.; El-Anadouli, B.E. Development of Nano-Hydroxyapatite/Chitosan Composite for Cadmium Ions Removal in Wastewater Treatment. J. Taiwan Inst. Chem. Eng. 2014, 45, 1571–1577. [Google Scholar] [CrossRef]
  19. Ferri, M.; Campisi, S.; Gervasini, A. Nickel and Cobalt Adsorption on Hydroxyapatite: A Study for the de-Metalation of Electronic Industrial Wastewaters. Adsorption 2019, 25, 649–660. [Google Scholar] [CrossRef]
  20. Kavasi, R.M.; Coelho, C.C.; Platania, V.; Quadros, P.A.; Chatzinikolaidou, M. In Vitro Biocompatibility Assessment of Nano-hydroxyapatite. Nanomaterials 2021, 11, 1152. [Google Scholar] [CrossRef]
  21. Silva-Holguín, P.N.; Medellín-Castillo, N.A.; Zaragoza-Contreras, E.A.; Reyes-López, S.Y. Alumina-Hydroxyapatite Millimetric Spheres for Cadmium(II) Removal in Aqueous Medium. ACS Omega 2023, 8, 44675–44688. [Google Scholar] [CrossRef] [PubMed]
  22. Li, H.; Guo, X.; Ye, X. Screening Hydroxyapatite for Cadmium and Lead Immobilization in Aqueous Solution and Contaminated Soil: The Role of Surface Area. J. Environ. Sci. 2017, 52, 141–150. [Google Scholar] [CrossRef]
  23. Corami, A.; Mignardi, S.; Ferrini, V. Cadmium Removal from Single- and Multi-Metal (Cd + Pb + Zn + Cu) Solutions by Sorption on Hydroxyapatite. J. Colloid. Interface Sci. 2008, 317, 402–408. [Google Scholar] [CrossRef]
  24. Gibert, O.; Valderrama, C.; Martínez, M.M.; Darbra, R.M.; Moncunill, J.O.; Martí, V. Hydroxyapatite Coatings on Calcite Powder for the Removal of Heavy Metals from Contaminated Water. Water 2021, 13, 1493. [Google Scholar] [CrossRef]
  25. Meski, S.; Tazibt, N.; Khireddine, H.; Ziani, S.; Biba, W.; Yala, S.; Sidane, D.; Boudjouan, F.; Moussaoui, N. Synthesis of Hydroxyapatite from Mussel Shells for Effective Adsorption of Aqueous Cd(II). Water Sci. Technol. 2019, 80, 1226–1237. [Google Scholar] [CrossRef]
  26. Wu, S.C.; Hsu, H.C.; Hsu, S.K.; Tseng, C.P.; Ho, W.F. Effects of Calcination on Synthesis of Hydroxyapatite Derived from Oyster Shell Powders. J. Aust. Ceram. Soc. 2019, 55, 1051–1058. [Google Scholar] [CrossRef]
  27. Mohd Pu’ad, N.A.S.; Alipal, J.; Abdullah, H.Z.; Idris, M.I.; Lee, T.C. Synthesis of Eggshell Derived Hydroxyapatite via Chemical Precipitation and Calcination Method. Mater. Today Proc. 2019, 42, 172–177. [Google Scholar] [CrossRef]
  28. Alif, M.F.; Aprillia, W.; Arief, S. A Hydrothermal Synthesis of Natural Hydroxyapatite Obtained from Corbicula Moltkiana Freshwater Clams Shell Biowaste. Mater. Lett. 2018, 230, 40–43. [Google Scholar] [CrossRef]
  29. Chen, J.; Wen, Z.; Zhong, S.; Wang, Z.; Wu, J.; Zhang, Q. Synthesis of Hydroxyapatite Nanorods from Abalone Shells via Hydrothermal Solid-State Conversion. Mater. Des. 2015, 87, 445–449. [Google Scholar] [CrossRef]
  30. Wang, H.; Yan, K.; Chen, J. Preparation of Hydroxyapatite Microspheres by Hydrothermal Self-Assembly of Marine Shell for Effective Adsorption of Congo Red. Mater. Lett. 2021, 304, 130573. [Google Scholar] [CrossRef]
  31. Zhang, H.-B.; Zhou, K.-C.; Li, Z.-Y.; Huang, S.-P. Plate-like Hydroxyapatite Nanoparticles Synthesized by the Hydrothermal Method. J. Phys. Chem. Solids 2009, 70, 243–248. [Google Scholar] [CrossRef]
  32. Summa, D.; Lanzoni, M.; Castaldelli, G.; Fano, E.A.; Tamburini, E. Trends and Opportunities of Bivalve Shells’ Waste Valorization in a Prospect of Circular Blue Bioeconomy. Resources 2022, 11, 48. [Google Scholar] [CrossRef]
  33. Kaplan, D.L. Mollusc Shell Structures: Novel Design Strategies for Synthetic Materials. Curr. Opin. Solid State Mater. Sci. 1998, 3, 232–236. [Google Scholar] [CrossRef]
  34. Stanmore, B.R.; Gilot, P. Review-Calcination and Carbonation of Limestone during Thermal Cycling for CO2 Sequestration. Fuel Process. Technol. 2005, 86, 1707–1743. [Google Scholar] [CrossRef]
  35. Ahmed, J.; Al-Attar, H.; Arfat, Y.A. Effect of Particle Size on Compositional, Functional, Pasting and Rheological Properties of Commercial Water Chestnut Flour. Food Hydrocoll. 2016, 52, 888–895. [Google Scholar] [CrossRef]
  36. Chenet, T.; Sarti, E.; Costa, V.; Cavazzini, A.; Rodeghero, E.; Beltrami, G.; Felletti, S.; Pasti, L.; Martucci, A. Influence of Caffeic Acid on the Adsorption of Toluene onto an Organophilic Zeolite. J. Environ. Chem. Eng. 2020, 8, 104229. [Google Scholar] [CrossRef]
  37. Zeng, C.; Ramos-Ruiz, A.; Field, J.A.; Sierra-Alvarez, R. Cadmium Telluride (CdTe) and Cadmium Selenide (CdSe) Leaching Behavior and Surface Chemistry in Response to PH and O2. J. Environ. Manag. 2015, 154, 78–85. [Google Scholar] [CrossRef]
  38. Smičiklas, I.D.; Milonjicámilonjicá, S.K.; Pfendt, P.; Raičevicá, S.R. The Point of Zero Charge and Sorption of Cadmium (II) and Strontium (II) Ions on Synthetic Hydroxyapatite. Sep. Purif. Technol. 2000, 18, 185–194. [Google Scholar] [CrossRef]
  39. Grassi, S.; Amadori, M.; Pennisi, M.; Cortecci, G. Identifying Sources of B and As Contamination in Surface Water and Groundwater Downstream of the Larderello Geothermal-Industrial Area (Tuscany-Central Italy). J. Hydrol. 2014, 509, 66–82. [Google Scholar] [CrossRef]
  40. La Vigna, F.; Ciadamidaro, S.; Mazza, R.; Mancini, L. Water Quality and Relationship between Superficial and Ground Water in Rome (Aniene River Basin, Central Italy). Environ. Earth Sci. 2010, 60, 1267–1279. [Google Scholar] [CrossRef]
  41. Ioele, G.; De Luca, M.; Grande, F.; Durante, G.; Trozzo, R.; Crupi, C.; Ragno, G. Assessment of Surface Water Quality Using Multivariate Analysis: Case Study of the Crati River, Italy. Water 2020, 12, 2214. [Google Scholar] [CrossRef]
  42. Chenet, T.; Martucci, A.; Cescon, M.; Vergine, G.; Beltrami, G.; Gigli, L.; Ardit, M.; Migliori, M.; Catizzone, E.; Giordano, G.; et al. Supramolecular Assembly of L-Lysine on ZSM-5 Zeolites with Different Si/Al Ratio. Microporous Mesoporous Mater. 2021, 323, 111183. [Google Scholar] [CrossRef]
  43. Ding, J.; Zhao, L.; Chang, Y.; Zhao, W.; Du, Z.; Hao, Z. Transcriptome Sequencing and Characterization of Japanese Scallop Patinopecten Yessoensis from Different Shell Color Lines. PLoS ONE 2015, 10, 0116406. [Google Scholar] [CrossRef] [PubMed]
  44. Xu, X.; Liu, X.; Oh, M.; Park, J. Oyster Shell as a Low-Cost Adsorbent for Removing Heavy Metal Ions from Wastewater. Pol. J. Environ. Stud. 2019, 28, 2949–2959. [Google Scholar] [CrossRef]
  45. Zhao, G.; Wang, X.; Chen, Q.; Cheng, J. Efficacy of Phosphorus Loaded Oyster Shell in Heavy Metal Removal: A Combined Study on Adsorption Behavior, Structure Characterization and Comparative Mechanism. Environ. Technol. Innov. 2024, 33, 103484. [Google Scholar] [CrossRef]
  46. Beni, A.A.; Esmaeili, A. Biosorption, an Efficient Method for Removing Heavy Metals from Industrial Effluents: A Review. Environ. Technol. Innov. 2020, 17, 100503. [Google Scholar] [CrossRef]
  47. Foo, K.Y.; Hameed, B.H. Insights into the Modeling of Adsorption Isotherm Systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  48. Araújo, C.S.T.; Almeida, I.L.S.; Rezende, H.C.; Marcionilio, S.M.L.O.; Léon, J.J.L.; de Matos, T.N. Elucidation of Mechanism Involved in Adsorption of Pb(II) onto Lobeira Fruit (Solanum Lycocarpum) Using Langmuir, Freundlich and Temkin Isotherms. Microchem. J. 2018, 137, 348–354. [Google Scholar] [CrossRef]
  49. Núñez, D.; Serrano, J.A.; Mancisidor, A.; Elgueta, E.; Varaprasad, K.; Oyarzún, P.; Cáceres, R.; Ide, W.; Rivas, B.L. Heavy Metal Removal from Aqueous Systems Using Hydroxyapatite Nanocrystals Derived from Clam Shells. RSC Adv. 2019, 9, 22883–22890. [Google Scholar] [CrossRef]
  50. Marchat, D.; Bernache-Assollant, D.; Champion, E. Cadmium Fixation by Synthetic Hydroxyapatite in Aqueous Solution-Thermal Behaviour. J. Hazard. Mater. 2007, 139, 453–460. [Google Scholar] [CrossRef]
  51. GadelHak, Y.; El-Azazy, M.; Shibl, M.F.; Mahmoud, R.K. Cost estimation of synthesis and utilization of nano-adsorbents on the laboratory and industrial scales: A detailed review. Sci. Total Environ. 2023, 875, 162629. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, F.Y.; Wang, H.; Ma, J.W. Adsorption of Cadmium (II) Ions from Aqueous Solution by a New Low-Cost Adsorbent-Bamboo Charcoal. J. Hazard. Mater. 2010, 177, 300–306. [Google Scholar] [CrossRef] [PubMed]
  53. Kumar, P.S.; Ramalingam, S.; Sathyaselvabala, V.; Kirupha, S.D.; Murugesan, A.; Sivanesan, S. Removal of Cadmium(II) from Aqueous Solution by Agricultural Waste Cashew Nut Shell. Korean J. Chem. Eng. 2012, 29, 756–768. [Google Scholar] [CrossRef]
  54. Batool, F.; Akbar, J.; Iqbal, S.; Noreen, S.; Bukhari, S.N.A. Study of Isothermal, Kinetic, and Thermodynamic Parameters for Adsorption of Cadmium: An Overview of Linear and Nonlinear Approach and Error Analysis. Bioinorg. Chem. Appl. 2018, 2018, 3463724. [Google Scholar] [CrossRef]
  55. Tran, H.N.; You, S.J.; Chao, H.P. Thermodynamic Parameters of Cadmium Adsorption onto Orange Peel Calculated from Various Methods: A Comparison Study. J. Environ. Chem. Eng. 2016, 4, 2671–2682. [Google Scholar] [CrossRef]
  56. Park, J.H.; Wang, J.J.; Kim, S.H.; Kang, S.W.; Jeong, C.Y.; Jeon, J.R.; Park, K.H.; Cho, J.S.; Delaune, R.D.; Seo, D.C. Cadmium Adsorption Characteristics of Biochars Derived Using Various Pine Tree Residues and Pyrolysis Temperatures. J. Colloid Interface Sci. 2019, 553, 298–307. [Google Scholar] [CrossRef]
  57. Cui, X.; Fang, S.; Yao, Y.; Li, T.; Ni, Q.; Yang, X.; He, Z. Potential Mechanisms of Cadmium Removal from Aqueous Solution by Canna Indica Derived Biochar. Sci. Total Environ. 2016, 562, 517–525. [Google Scholar] [CrossRef]
  58. Rujitanapanich, S.; Kumpapan, P.; Wanjanoi, P. Synthesis of Hydroxyapatite from Oyster Shell via Precipitation. Energy Procedia 2014, 56, 112–117. [Google Scholar] [CrossRef]
  59. Wang, M.; Wu, S.; Guo, J.; Zhang, X.; Yang, Y.; Chen, F.; Zhu, R. Immobilization of Cadmium by Hydroxyapatite Converted from Microbial Precipitated Calcite. J. Hazard. Mater. 2019, 366, 684–693. [Google Scholar] [CrossRef]
  60. Evis, Z.; Yilmaz, B.; Usta, M.; Levent Aktug, S. X-Ray Investigation of Sintered Cadmium Doped Hydroxyapatites. Ceram. Int. 2013, 39, 2359–2363. [Google Scholar] [CrossRef]
  61. Julia, M.; Putnis, C.V.; King, H.E.; Renard, F. Coupled Dissolution-Precipitation and Growth Processes on Calcite, Aragonite, and Carrara Marble Exposed to Cadmium-Rich Aqueous Solutions. Chem. Geol. 2023, 621, 121364. [Google Scholar] [CrossRef]
  62. Wu, S.C.; Hsu, H.C.; Hsu, S.K.; Tseng, C.P.; Ho, W.F. Preparation and Characterization of Hydroxyapatite Synthesized from Oyster Shell Powders. Adv. Powder Technol. 2017, 28, 1154–1158. [Google Scholar] [CrossRef]
  63. Da Silva, P.S.; de Moura Farias, W.; Gomez, M.R.; Torrecilha, J.K.; Rocha, F.R.; Scapin, M.A.; Garcia, R.H.; De Simone, L.R.; De Amaral, V.S.; Vincent, M.; et al. Oyster shell element composition as a proxy for environmental studies. J. S. Am. Earth Sci. 2024, 134, 104749. [Google Scholar] [CrossRef]
  64. Filiberto, L.H.; Putnis, C.V.; Julia, M. Factors Controlling Reaction Pathways during Fluid–Rock Interactions. Contrib. Mineral. Petrol. 2023, 178, 53. [Google Scholar] [CrossRef]
  65. Kaddah, F.; Roziere, E.; Ranaivomanana, H.; Amiri, O. Complementary Use of Thermogravimetric Analysis and Oven to Assess the Composition and Bound CO2 Content of Recycled Concrete Aggregates. Dev. Built Environ. 2023, 15, 100184. [Google Scholar] [CrossRef]
  66. Tõnsuaadu, K.; Peld, M.; Bender, V. Thermal analysis of apatite structure. J. Therm. Anal. Calorim. 2003, 72, 363–371. [Google Scholar] [CrossRef]
  67. Lafon, J.P.; Champion, E.; Bernache-Assollant, D.; Gibert, R.; Danna, A.M. Thermal decomposition of carbonated calcium phosphate apatites. J. Therm. Anal. Calorim. 2003, 72, 1127–1134. [Google Scholar] [CrossRef]
Figure 1. (a) Pacific Oyster shells (Magallana gigas); (b) Queen Scallop shells (Aequipecten opercularis).
Figure 1. (a) Pacific Oyster shells (Magallana gigas); (b) Queen Scallop shells (Aequipecten opercularis).
Cleantechnol 07 00034 g001
Figure 2. SEM images of HAOS (a) and HASS (b) at 11.00 K X.
Figure 2. SEM images of HAOS (a) and HASS (b) at 11.00 K X.
Cleantechnol 07 00034 g002
Figure 3. Particle size distribution of HASS and HAOS.
Figure 3. Particle size distribution of HASS and HAOS.
Cleantechnol 07 00034 g003
Figure 4. Uptake of Cd2+ onto: HASS, HAOS, HAcSS and HAcOS.
Figure 4. Uptake of Cd2+ onto: HASS, HAOS, HAcSS and HAcOS.
Cleantechnol 07 00034 g004
Figure 5. Uptake of Cd2+ on HASS (red dots) and HAOS (blue dots) vs. time, Cd2+ initial concentration 10 mg L−1, experimental data fitted with PSO kinetic model.
Figure 5. Uptake of Cd2+ on HASS (red dots) and HAOS (blue dots) vs. time, Cd2+ initial concentration 10 mg L−1, experimental data fitted with PSO kinetic model.
Cleantechnol 07 00034 g005
Figure 6. Adsorption isotherms of Cd2+ onto HASS (red dots) and HAOS (blue dots), experimental data described with Freundlich and Langmuir isotherm models.
Figure 6. Adsorption isotherms of Cd2+ onto HASS (red dots) and HAOS (blue dots), experimental data described with Freundlich and Langmuir isotherm models.
Cleantechnol 07 00034 g006
Figure 7. X-ray powder diffraction patterns of HASS (blue), HAOS (green), Cd-HASS (red), and Cd-HAOS (violet); (a) comparison between HASS and HAOS before and after Cd loading; (b) comparison between HASS and Cd-HASS; (c) comparison between HAOS and Cd-HAOS.
Figure 7. X-ray powder diffraction patterns of HASS (blue), HAOS (green), Cd-HASS (red), and Cd-HAOS (violet); (a) comparison between HASS and HAOS before and after Cd loading; (b) comparison between HASS and Cd-HASS; (c) comparison between HAOS and Cd-HAOS.
Cleantechnol 07 00034 g007aCleantechnol 07 00034 g007b
Figure 8. TG, DTA and DTG plots of bare and cadmium loaded samples of Cd-HASS and HASS.
Figure 8. TG, DTA and DTG plots of bare and cadmium loaded samples of Cd-HASS and HASS.
Cleantechnol 07 00034 g008
Figure 9. TG, DTA and DTG plots of bare and cadmium loaded samples of Cd-HAOS and HAOS.
Figure 9. TG, DTA and DTG plots of bare and cadmium loaded samples of Cd-HAOS and HAOS.
Cleantechnol 07 00034 g009
Table 1. Particle size distribution of HASS, HAOS, HAcSS and HAcOS.
Table 1. Particle size distribution of HASS, HAOS, HAcSS and HAcOS.
ParametersHASSHAOSHAcSSHAcOS
D [4;3] (µm)22.389 ± 1.00917.291 ± 0.76513.519 ± 0.69613.578 ± 0.779
D [3;4] (µm)4.895 ± 0.1456.588 ± 0.3014.695 ± 0.3254.883 ± 0.344
Dv (10) (µm)1.913 ± 0.0963.063 ± 0.1532.017 ± 0.1102.023 ± 0.125
Dv (50) (µm)9.582 ± 0.37910.113 ± 0.40610.330 ± 0.6179.506 ± 0.467
Dv (90) (µm)67.212 ± 3.26143.153 ± 1.87829.910 ± 1.66931.418 ± 1.171
Dv (98) (µm)110.072 ± 5.30678.134 ± 3.70945.799 ± 2.15651.669 ± 2.283
Dv (100) (µm)162.552 ± 7.128125.641 ± 6.58266.712 ± 3.03675.878 ± 3.294
Volume Below 1 μm (%)2.930.753.212.67
Volume Below 10 μm (%)51.2649.5448.6652.02
Volume Above 45 μm (%)16.809.302.223.64
Table 2. Parameters obtained by non-linear fitting of the uptake data using PSO model. The error is given as confidence interval at 95% of probability.
Table 2. Parameters obtained by non-linear fitting of the uptake data using PSO model. The error is given as confidence interval at 95% of probability.
Materialsk2 (g mg−1 min−1)qe (mg L−1)R2
HASS25.8 ± 1.62.0170 ± 0.00051
HAOS189.9 ± 8.42.0870 ± 0.00011
Table 3. Isotherm parameters obtained from Langmuir and Freundlich isotherm models for HASS and HAOS, with confidence intervals at 95% of probability.
Table 3. Isotherm parameters obtained from Langmuir and Freundlich isotherm models for HASS and HAOS, with confidence intervals at 95% of probability.
IsothermHASSHAOS
Langmuir
qs (mg g−1)173.70 ± 16.95334.30 ± 46.20
b (L mg−1)2.04 ± 1.580.31 ± 0.10
R20.91540.9443
Freundlich
kF (mg g−1) (L mg−1)1/n112.90 ± 14.87105.90 ± 12.19
n10.63 ± 3.723.97 ± 0.26
R20.95770.9219
Table 5. Crystallite size (nm) Crystallinity degree (%) and weighted amount (wt%) of phases obtained using RSR for bare and Cd-loaded HAOS and HASS samples.
Table 5. Crystallite size (nm) Crystallinity degree (%) and weighted amount (wt%) of phases obtained using RSR for bare and Cd-loaded HAOS and HASS samples.
HAOSHASSCd-HAOSCd-HASS
Calcite (wt%)24.5 (4)22.8 (2)18.4 (4)17.0 (4)
Aragonite (wt%)-4.2 (2)-4.2 (5)
Hydroxyapatite (wt%)75.5 (4)73.0 (3)68.1 (5)78.8 (6)
Cadmium Hydrogen Phosphate Hydrate (wt%)--13.5 (3)-
Crystallite size (nm)24.825.231.932.0
Crystallinity (%)63.265.463.162.1
Table 6. Unit-cell parameters for bare and Cd-loaded HAOS and HASS samples.
Table 6. Unit-cell parameters for bare and Cd-loaded HAOS and HASS samples.
Calcite CaCO3
R-3 c Ha = b (Å)c (Å)Volume (Å3)
HAOS4.9989 (7)17.111 (6)370.3 (3)
Cd-HAOS4.9927 (8)17.082 (3)368.8 (1)
HAAS4.9917 (1)17.0846 (4)368.6 (1)
Cd-HAAS4.9890 (4)17.0746 (2)368.0 (1)
Hydroxyapatite Ca10(PO4)6(OH)2
P 63/ma = b (Å)c (Å)Volume (Å3)
HAOS9.404 (5)6.891 (4)527.8 (6)
Cd-HAOS9.455 (7)6.717 (6)520.1 (9)
HAAS9.396 (1)6.877 (1)525.8 (3)
Cd-HAAS9.400 (3)6.805 (4)520.7 (4)
Aragonite CaCO3
P mcna (Å)b (Å)c (Å)Volume (Å3)
HAAS4.968 (3)7.962 (7)5.752 (4)227.5 (3)
Cd-HAAS4.964 (2)7.955 (3)5.745 (2)226.9 (1)
Cadmium Hydrogen Phosphate Hydrate Cd5H2(PO4)4(H2O)4
C 1 2/c 1a (Å)b (Å)c (Å)Volume (Å3)β (°)
Cd-HAOS17.9553 (1)9.4258 (4)9.7121 (4)1632.9 (4)96.57 (1)
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

Cescon, M.; Chiefa, F.; Chenet, T.; Mancinelli, M.; Stevanin, C.; Martucci, A.; Pasti, L. Hydroxyapatite-Based Adsorbent Materials from Aquaculture Waste for Remediation of Metal-Contaminated Waters: Investigation of Cadmium Removal. Clean Technol. 2025, 7, 34. https://doi.org/10.3390/cleantechnol7020034

AMA Style

Cescon M, Chiefa F, Chenet T, Mancinelli M, Stevanin C, Martucci A, Pasti L. Hydroxyapatite-Based Adsorbent Materials from Aquaculture Waste for Remediation of Metal-Contaminated Waters: Investigation of Cadmium Removal. Clean Technologies. 2025; 7(2):34. https://doi.org/10.3390/cleantechnol7020034

Chicago/Turabian Style

Cescon, Mirco, Francesco Chiefa, Tatiana Chenet, Maura Mancinelli, Claudia Stevanin, Annalisa Martucci, and Luisa Pasti. 2025. "Hydroxyapatite-Based Adsorbent Materials from Aquaculture Waste for Remediation of Metal-Contaminated Waters: Investigation of Cadmium Removal" Clean Technologies 7, no. 2: 34. https://doi.org/10.3390/cleantechnol7020034

APA Style

Cescon, M., Chiefa, F., Chenet, T., Mancinelli, M., Stevanin, C., Martucci, A., & Pasti, L. (2025). Hydroxyapatite-Based Adsorbent Materials from Aquaculture Waste for Remediation of Metal-Contaminated Waters: Investigation of Cadmium Removal. Clean Technologies, 7(2), 34. https://doi.org/10.3390/cleantechnol7020034

Article Metrics

Back to TopTop