Next Article in Journal
Petrology and Geochemistry of an Unusual Granulite Facies Xenolith of the Late Oligocene Post-Obduction Koum Granodiorite (New Caledonia, Southwest Pacific): Geodynamic Inferences
Previous Article in Journal
The Geology and Mineral Chemistry of Beryl Mineralization, South Eastern Desert, Egypt: A Deeper Insight into Genesis and Distribution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Separation of Rare Earth Ions from Mine Wastewater Using Synthetic Hematite Nanoparticles from Natural Pyrite

1
School of Minerals Processing & Bioengineering, Central South University, Changsha 410083, China
2
Key Lab of Biohydrometallurgy of Ministry of Education, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(5), 464; https://doi.org/10.3390/min14050464
Submission received: 27 March 2024 / Revised: 12 April 2024 / Accepted: 24 April 2024 / Published: 28 April 2024
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
The separation of rare earth ions (RE3+) from aqueous solutions poses a significant challenge due to their similar chemical and physical characteristics. This study presents a method for synthesizing hematite nanoparticles (Fe2O3 NPs) through the high-temperature phase transition of natural pyrite for adsorbing RE3+ from mine wastewater. The characteristics of Fe2O3 NPs were studied using XRD, SEM, BET, XPS, FTIR, and Zeta potential. The optimal condition for RE3+ adsorption by Fe2O3 NPs was determined to be at pH 6.0 with an adsorption time of 60 min. The maximum adsorption capacities of Fe2O3 NPs for La3+, Ce3+, Pr3+, Nd3+, Sm3+, Gd3+, Dy3+, and Y3+ were 12.80, 14.02, 14.67, 15.52, 17.66, 19.16, 19.94, and 11.82 mg·g−1, respectively. The experimental data fitted well with the Langmuir isotherm and pseudo-second-order models, suggesting that the adsorption process was dominated by monolayer chemisorption. Thermodynamic analysis revealed the endothermic nature of the adsorption process. At room temperature, the adsorption of RE3+ in most cases (La3+, Ce3+, Pr3+, Nd3+, Sm3+, and Y3+) onto Fe2O3 NPs was non-spontaneous, except for the adsorption of Gd3+ and Dy3+, which was spontaneous. The higher separation selectivity of Fe2O3 NPs for Gd3+ and Dy3+ was confirmed by the separation factor. Moreover, Fe2O3 NPs exhibited excellent stability, with an RE3+ removal efficiency exceeding 94.70% after five adsorption–desorption cycles, demonstrating its potential for the recovery of RE3+ from mine wastewater.

1. Introduction

Rare earth elements (REEs) play a crucial role in advancing modern technology and have widespread applications in environmental protection, energy production, digital technology, and medical applications [1]. However, the traditional extraction process of REEs leads to a loss of approximately 10% in yield, resulting in over 20 million tons of wastewater annually with REEs concentrations ranging from 1 to100 mg·L−1 [2,3]. With the growing demand for REEs and their compounds, the recycling of REEs from mine wastewater has received considerable attention [4]. Additionally, similar to other heavy metals, REEs can be accumulated in ecosystems, posing potential risks to the environment and human health [5,6]. Studies have shown that REEs can cause liver and lung cytotoxicity, damage the reproductive system, and lead to neural tube defects [7,8]. Therefore, the recovery of REEs from mine wastewater is becoming an increasingly promising option [9].
To date, various methods including ion exchange, chemical precipitation, solvent extraction, and adsorption have been developed for the separation and recovery of REEs [6]. Each method has its advantages and disadvantages. For example, ion exchange is known for its simple operation and effective separation but lacks continuous processing capability and is time-consuming [10]. The chemical precipitation method is cost-effective and easy to operate but is primarily suitable for high-concentration REEs separation [11]. Solvent extraction offers low energy consumption and wide applicability but poses environmental pollution problems due to the use of organic solvents [12]. While new unconventional processes have recently been developed by Faur et al. and Xu et al., they are not yet industrialized for reasons of optimization [13,14,15]. Among these methods, adsorption stands out as a practical, efficient, environmentally friendly, and cost-effective approach for treating wastewater with low REE concentrations [16]. Numerous adsorbents such as O-modified coordination polymer [11], Ca-alginate/carboxymethyl chitosan/Ni0.2Zn0.2Fe2.6O4 magnetic bionanocomposite [16], magnetic Fe3O4/MnO2 decorated reduced graphene oxide [9], functionalized magnetite nanoparticles [17], and magnetic mesoporous Fe3O4@mSiO2–DODGA nanoparticles [12] have been developed for the adsorption of REEs. Magnetic nanoparticles in particular have received interest for their excellent separation and regeneration capabilities [17]. However, researchers still encounter the challenge of devising simple and cost-effective methods for preparing magnetic nanoparticle adsorbents with desirable adsorption properties.
Natural pyrite is abundant, inexpensive [18], and considered as an industrial solid waste [19]. Furthermore, rare earth minerals and pyrite sometimes coexist. For instance, the conglomerate beds of rare earth minerals in Eco Ridge Mine contain 5% to 15% pyrite [20]. Studies have demonstrated that pyrite can be transformed into hematite magnetic nanoparticles through the high-temperature phase transition [21], which occurs without the use of chemical reagents. Hematite has been proven to be an effective and environmentally friendly adsorbent for removing inorganic heavy metal ions such as Cr4+, As3+, As5+, UO22+, Pb2+, and Cu2+ [22,23,24,25]. In particular, rare earth ions (RE3+) have an affinity for adsorption with iron oxides according to previous studies [20,26]. However, the potential of hematite converted from natural pyrite to adsorb RE3+ remains unexplored.
Therefore, in this paper, hematite nanoparticles (Fe2O3 NPs) were synthesized by the high-temperature phase transition of pyrite for the recovery of mixed RE3+ from mine wastewater. The synthesized product was characterized using scanning electron microscopy (SEM), X-ray diffraction (XRD), Brunauer–Emmett–Teller (BET), Fourier-transform infrared spectroscopy (FTIR), and Zeta potential. The effect of time, pH, initial RE3+ concentration, and temperature on the adsorption efficiency of RE3+ was investigated through batch adsorption experiments. Furthermore, the experimental data were fitted with classical kinetic and isotherm models, and the thermodynamic parameters (ΔS°, ΔG°, and ΔH°) were evaluated to provide insights into the adsorption process. The proposed method of utilizing Fe2O3 NPs derived from pyrite as an adsorbent for RE3+ provides a cost-effective and efficient solution for the recovery of RE3+ in aqueous solutions while also increasing the utilization value of pyrite.

2. Material and Methods

2.1. Materials and Reagents

The RE3+ solution used for the adsorption experiments was prepared by diluting a standard RE3+ solution (100 mg·L−1) [27]. The pH of the solution was adjusted using either 0.1 M sodium hydroxide or nitric acid solution and monitored with a pH meter [28]. Anhydrous ethanol was obtained from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. All chemicals and reagents were of analytical grade and used as received. The mine wastewater used in this study was collected from an ion adsorption rare earth mine in Hunan, China. The concentration of the main metal ions in the wastewater was determined by inductively coupled plasma mass spectrometry (ICP-MS) after filtration through a 0.22 μm filter membrane [29], and the results are presented in Table 1. Pyrite samples were taken from Hubei, China, ground using a vibratory mill, and sieved to ensure that the particle size was between 38 μm and 74 μm for subsequent studies. The selected pyrite was found to be of high purity based on XRD analysis (Figure S1).

2.2. Preparation of Fe2O3 NPs

The pyrite was further ground using a planetary ball mill (YXQM-4L, MITR) [30]. In the grinding process, an orthogonal experiment L9(34) was designed by specifying ball-to-material ratios (10:1, 15:1, and 20:1), rotation speeds (400, 500, and 600 rpm), and grinding times (1, 2, and 3 h) as factors A, B, and C, respectively. Experimental details are presented in Tables S1 and S2. Subsequently, the finely ground pyrite was subjected to high-temperature calcination in a muffle furnace at various temperatures and durations to convert it into hematite [21]. The heating and cooling rates were both 10 °C·min−1, and the sample was taken out after dropping to room temperature. Finally, Fe2O3 NPs were obtained by grinding the calcined product in an agate mortar using anhydrous ethanol as a dispersant [23].

2.3. Adsorption and Desorption Experiments

The adsorption capacity of Fe2O3 NPs to mixed RE3+ in solution was investigated by batch adsorption experiments. The experiments were conducted in 250 mL Erlenmeyer flasks placed on a magnetic stirrer at room temperature. The initial concentration of RE3+ was 5 mg·L−1 each, with 50 mg of Fe2O3 NPs, a pH of 5.0, and a solution volume of 50 mL. Various adsorption times (2–240 min) were tested to determine the time required for adsorption equilibrium and for kinetic analyses. The optimal pH for adsorption was determined by testing different initial pH values ranging from 2.0 to 7.0 [31]. Additionally, thermodynamic adsorption experiments were conducted at three different temperatures (298, 308, and 318 K). Isothermal adsorption experiments were performed at initial RE3+ concentrations of 5–40 mg·L−1. All experiments were performed in triplicate. Kinetic and isothermal models were fitted using the Origin software (version 2021), optimizing the model parameters to minimize the difference between the predicted and observed values. After reaching adsorption equilibrium, the concentration of RE3+ in the solution was measured by ICP-MS after filtration with a 0.22 μm filter membrane. The adsorption efficiency and capacity of RE3+ adsorbed on Fe2O3 NPs were calculated according to Equations (1) and (2) [32].
A d s o r p t i o n   e f f i c i e n c y ( % ) = C 0 C e C 0 × 100 %
q t = C 0 C t V m
where q t (mg·g−1) represents the adsorption capacity at time t; C 0 (mg·L−1), C e (mg·L−1), and C t (mg·L−1) denote the initial, equilibrium, and momentary concentrations of RE3+, respectively; V (L) refers to the solution volume; and m (g) is the adsorbent mass.
The recycling and regeneration experiments of Fe2O3 NPs were carried out as follows: a 50 mL solution containing 5 mg·L−1 of RE3+ was mixed with 100 mg of Fe2O3 NPs for 1 h in each cycle. The mixed solution was then filtered using a 0.22 μm membrane, and the residual concentration of RE3+ was analyzed by ICP-MS. Desorption regeneration was achieved by agitating with 0.2 M HNO3 for 1 h, followed by washing with pure water [16]. The cleaned and regenerated Fe2O3 NPs were then reintroduced into a fresh mixed RE3+ solution with the same concentration and volume, allowing for adsorption for 1 h. This process was repeated five times. Finally, the effectiveness of the Fe2O3 NPs for practical applications was evaluated by mixing 200 mg of the adsorbent with 50 mL of actual mine wastewater.

2.4. Characterizations

The particle size analysis of ground pyrite was performed with a laser particle size analyzer (MASTERSIZER 2000) (Malvern, Worcestershire, UK) [33]. Thermal behavior analysis was conducted using a thermal gravimetric analyzer (TGADTA, STA 449C/4/MFC/G/Jupiter®, Bavaria, Germany) with a gas flow rate of 100 mL·min−1 in an air atmosphere [34]. XRD analysis at 4–80° was used to determine the physical phase of pyrite after heat treatment [28]. The specific surface area and pore size distribution of the adsorbent were measured at 77 K using a specific surface area analyzer (Autosorb iQ) (Anton Paar, Graz, Austria) at a relative pressure of 0~1.0 [9]. Morphology and surface element composition were characterized with a scanning electron microscope (SEM; Nova NanoSEM 230, FEI Company, Hillsboro, OR, USA) equipped with an X-ray energy dispersive spectroscopy system (EDS). Nanoparticle size was analyzed by transmission electron microscopy (TEM, Talos F200i) (Thermo Fisher Scientific, Waltham, MA, USA). FTIR spectra were recorded using an FT-IR spectrometer (iS50, Thermo Fisher Scientific, Waltham, MA, USA) in the range of 4000–400 cm−1 [35]. Elemental composition on the sample surface was determined by X-ray photoelectron spectroscopy (ESCALAB250) (Thermo Fisher Scientific, Waltham, MA, USA), with an analysis of the oxidation state of oxygen [28]. Zeta potential was measured using a Zeta point position analyzer (Nano ZS, Malvern Panalytical, Worcestershire, UK).

3. Results and Discussion

3.1. Preparation and Characterization of Fe2O3 NPs

For larger particles, the pyrolysis of pyrite is inhibited or shifted to higher temperatures due to slow gas diffusion [36]. As a result, the pyrite was crushed using a planetary ball mill. To investigate the impact of the ball-to-material ratio, rotation speed, and grinding time on the grinding process of pyrite, the particle size distribution and median size (D50) of the milled pyrite were analyzed [33]. D50 is a crucial parameter for measuring the size distribution, representing the point where half of the particle diameter is less than or equal to this value. A smaller D50 value indicates a finer particle distribution, while a larger D50 value suggests a coarser distribution [37]. The histograms of the particle size distribution and the cumulative distribution curves of the milled pyrite are presented in Figure 1. The analysis revealed a segment with a steep slope in the cumulative distribution curve, while the other segments appeared as nearly horizontal straight lines. This indicated significant fluctuations in the number of particles in the sloped segment, whereas there were almost no particles in the horizontal straight segment. The results of the orthogonal experiment demonstrated that the ball-to-material ratio had the most significant influence, with the rotation speed and grinding time showing a relatively lower degree of influence on the grinding particle size (Table 2). Based on the D50 results, A2B3C1 (ball-to-material ratio of 15:1, rotation speed of 600 rpm, and grinding time of 1 h) was selected for grinding to obtain finely ground pyrite samples suitable for calcination.
Figure 2A illustrates the TG-DTG-DSC curves of fine-grained pyrite from room temperature to 1000 °C in an air atmosphere. The results indicated that a reaction occurred at 386.4 °C, leading to an increase in mass as pyrite transformed into sulfate (FeSO4, Fe2(SO4)3) [38], which further decomposed into iron oxide (Fe2O3) and sulfur dioxide (SO2) [39]. Oxidation occurred at 493.8 °C, converting iron-containing substances into hematite (Fe2O3) at 699.9 °C [36]. Consequently, the thermal decomposition of fine-grained pyrite began at 386.4 °C and ended at 699.9 °C.
Various temperatures and calcination times were chosen to observe the phase transition of pyrite thermal decomposition, as depicted in Figure 2B,C. The results indicated that after being calcined at 450 °C for 60 min and at 650 °C for 10 min, some pyrite could still be detected. Additionally, FeSO4 and hematite were detected. When the calcination conditions were set at 650 °C for 20 min, the pyrite completely disappeared, and the physical phase consisted solely of hematite and FeSO4. When the calcination conditions were 650 °C for 30 min, 650 °C for 60 min, and 900 °C for 20 min, the pyrite phase was completely transformed into hematite. These findings were consistent with the results obtained from the thermogravimetric analysis. Thus, for economic reasons, the optimal calcination conditions were determined to be 650 °C for 30 min. Finally, the calcined products were ground with ethanol as a dispersant to prevent agglomeration [40] in the agate mortar [23]. The overall preparation scheme is illustrated in Figure 3.
Figure 4 depicts the characteristics of fine-grained pyrite and Fe2O3 NPs. It was observed that the particle size of the sample significantly decreased after calcination (Figure 4A,B). The Fe2O3 NPs exhibited fine particles with a rough surface, ranging from 10 to 200 nm in size (Figure 4C,D), confirming the successful preparation of Fe2O3 NPs. Furthermore, the atomic ratio of Fe to O was approximately 2:3 (Figure 4E), indicating the complete conversion of pyrite to hematite, which aligned with the XRD results. The specific surface area, total pore volume, and average pore diameter of the Fe2O3 NPs were analyzed using the N2 adsorption–desorption isotherm to be 14.12 m2·g−1, 0.061 cm3·g−1, and 17.36 nm, respectively (Figure 4F).

3.2. Factors Affecting the Adsorption Efficiency of RE3+

3.2.1. Time

Figure 5A depicts the correlation between the adsorption efficiency of RE3+ by Fe2O3 NPs and the adsorption time. The results showed that the adsorption of RE3+ was rapid within the first 10 min, attributed to the abundance of active sites available on the adsorbent surface [16]. Subsequently, the adsorption efficiency gradually decreased due to the reduction in active sites and the weakening of the driving force, reaching equilibrium in approximately 60 min. Based on these findings, an optimal adsorption time of 60 min was determined for subsequent experiments.
To further understand the adsorption kinetics of RE3+ on the Fe2O3 NPs, the results of the batch experiments were fitted with two classical models: the pseudo-first-order (PFO) model (Equation (3)) and the pseudo-second-order (PSO) model (Equation (4)) [41].
ln ( q e q t ) = ln q e k 1 t
t q t = 1 k 2 q e 2 + t q e
where k 1 (min−1) and k 2 (g·mg−1·min−1) denote the pseudo-first-order and pseudo-second-order kinetic rate constants, respectively.
Figure 5B–I illustrates the nonlinear fit of the PFO and PSO models to the adsorption of RE3+ on the Fe2O3 NPs, and the kinetic parameters are summarized in Table 3. The correlation coefficient (R2) values clearly indicated that the adsorption kinetics of RE3+ on the Fe2O3 NPs could be more accurately explained by the PSO kinetic model. Furthermore, the experimental values of qe (qe,exp) were in close agreement with the theoretical values of qe (qe,cal) calculated by the PSO model. This suggested that the adsorption efficiency limiting step was a chemisorption process through electron sharing or exchange between the Fe2O3 NPs and RE3+ [42].

3.2.2. pH

The pH of the solution can affect the adsorption capacity of the adsorbent by affecting the activity of functional groups and the surface charge [43]. Therefore, the adsorption efficiencies of the Fe2O3 NPs for mixed RE3+ at various pH values were investigated and are shown in Figure 6A. The results revealed that the adsorption efficiency of RE3+ by the Fe2O3 NPs increased as the initial pH of the solution increased, reaching 100% at pH 7.0. The isoelectric point of the Fe2O3 NPs was found to be 5.64 (Figure 6B). At lower pH values, the functional groups on the surface of the Fe2O3 NPs became protonated and positively charged, leading to competition between H+ and RE3+ for active sites on the adsorbent surface, thereby hindering the adsorption of RE3+ [16]. This resulted in lower adsorption efficiencies. As the pH increased, the positive charge on the surface of the Fe2O3 NPs decreased gradually and became negatively charged at pH > 5.64 (Figure 6B). This resulted in electrostatic attraction between the Fe2O3 NPs and RE3+, increasing the adsorption efficiency of RE3+ [29]. The adsorption efficiency reached 100% rapidly at pH 7.0. Notably, pH also influences the distribution of RE3+ species in the solution [2]. Hence, the distribution of RE3+ species at thermodynamic equilibrium was calculated using Visual MINTEQ [29], and the results are shown in Figure 7. It was observed that in different pH ranges, RE3+ existed as different species, including RE3+, RE(OH)2+, and RE(OH)3(s) [2]. The precipitation of RE(OH)3(s) began at pH values of 8.2, 8.2, 8.0, 7.4, 6.8, 6.4, 6.8, and 7.4 for La3+, Ce3+, Pr3+, Nd3+, Sm3+, Gd3+, Dy3+, and Y3+, respectively. In solutions with pH ≤ 6.0, RE3+ was the dominant species. This conclusion is in agreement with previous research [2]. Therefore, a pH of 6.0 was identified as the optimal pH because it achieved a high adsorption efficiency and prevented the precipitation of RE(OH)3(s).

3.2.3. Initial RE3+ Concentration

The equilibrium adsorption capacity increased with RE3+ concentration, as shown in Figure 8A. This could be attributed to the greater driving force at higher initial ion concentrations, which facilitated overcoming the mass transfer resistance between the solid and liquid phases [16]. Consequently, more interactions occurred between RE3+ and the active sites of the adsorbent, leading to a higher adsorption capacity. This finding was consistent with previous research [11].
Two well-known isotherm models, the Langmuir (Equation (5)) and Freundlich (Equation (6)) models [11], were utilized to analyze the experimental data. The dimensionless separation factor R L (Equation (7)) was also employed to assess the affinity between the adsorbent and RE3+ [17]. The Langmuir model demonstrated higher correlation coefficients (0.990 < R2 < 0.994) compared to the Freundlich model (Figure 8B–I and Table 4), indicating better agreement with the experimental data. This suggested that the adsorption of RE3+ on the Fe2O3 NPs was a monolayer adsorption [9]. Once the active sites were covered by ions, no further adsorption occurred. Moreover, the positive K L values for all eight types of RE3+ resulted in R L values between 0 and 1, indicating that the adsorption process of RE3+ on Fe2O3 NPs was favorable [12]. According to the Langmuir isotherm model, the maximum adsorption capacities of the Fe2O3 NPs for La3+, Ce3+, Pr3+, Nd3+, Sm3+, Gd3+, Dy3+, and Y3+ were calculated to be 12.80, 14.02, 14.67, 15.52, 17.66, 19.16, 19.94, and 11.82 mg·g−1, respectively. The adsorption capacities of the Fe2O3 NPs for RE3+ decreased in the order of Dy3+ > Gd3+ > Sm3+ > Nd3+ > Pr3+ > Ce3+ > La3+ > Y3+, corresponding to the increasing radius of RE3+ [44], except for Y3+. This might be due to the different electronic configuration of Y3+ compared to other lanthanide ions [45]. Furthermore, compared with the adsorption capacities of RE3+ by different iron oxides adsorbents reported in the literature (Table 5), the Fe2O3 NPs synthesized in this study exhibit competitive advantages in the adsorption of RE3+ [41,46].
C e q e = C e q m + 1 K L q m
ln q e = ln K F + 1 n ln C e
R L = 1 ( 1 + K L C 0 )
where q m (mg·g−1) is the maximum adsorption capacity, K L (L·mg−1) and K F (mg·g−1) represent the adsorption constants of the Langmuir and Freundlich models, respectively, and n denotes adsorption intensity [16].

3.2.4. Temperature

The adsorption of metal ions at the solid–liquid interface was influenced by thermodynamic interactions between molecules [25]. To understand the energy change in Fe2O3 NPs adsorbing RE3+, the thermodynamic behavior was investigated [9]. As illustrated in Figure 9, the adsorption efficiency of all RE3+ increased with temperature, which could be attributed to the increased mobility of metal ions [16]. The thermodynamic parameters such as Gibbs free energy ( Δ G 0 ), enthalpy ( Δ H 0 ), and entropy ( Δ S 0 ) were calculated using Equations (8)–(10) [16]. The thermodynamic parameters at different temperatures, presented in Table 6, revealed that the Δ G 0 values for La3+, Ce3+, Pr3+, Nd3+, Sm3+, and Y3+ were positive, indicating non-spontaneous adsorption at room temperature. However, at higher temperatures, Δ G 0 for Sm3+ became negative, suggesting spontaneous adsorption. Gd3+ and Dy3+ exhibited negative Δ G 0 values within the tested temperature range, indicating spontaneous adsorption on the Fe2O3 NPs. Particularly, the Δ G 0 value for Dy3+ was more negative, suggesting stronger adsorption feasibility [16]. As the temperature increased from 298 K to 318 K, the negative Δ G 0 values became larger, and the positive Δ G 0 values became smaller for all eight types of RE3+, indicating that higher temperatures could enhance adsorption efficiency [9]. The positive Δ H 0 value suggested that the adsorption process of RE3+ on the Fe2O3 NPs was endothermic [9]. Additionally, the positive Δ S 0 value indicated that the randomness and disorder of the solid–solution interface increased during the adsorption process of RE3+ on the Fe2O3 NPs [16].
K d = q e C e
ln K d = Δ S 0 R Δ H 0 R T
Δ G 0 = R T ln K d = Δ H 0 T Δ S 0
where R stands for the gas constant (8.314 J·mol−1·K−1), T (K) is the absolute temperature, and K d represents the distribution coefficient.

3.3. Selective Separation

To investigate the selective adsorption of RE3+ on the Fe2O3 NPs, Gd3+ and Dy3+ were chosen as representatives. The separation factors (SF) (Equation (11)) for these two ions were calculated and compared (Table 7) [17,48]. The results showed the successful separation of Gd3+ and Dy3+ by the Fe2O3 NPs in a solution with eight types of mixed RE3+. The separation selectivity, reflected in the SF values, increased with the distance between the two REEs, suggesting that the difference in size between the ions played a role in the high separation selectivity [17]. Additionally, Gd3+ and Dy3+ could be easily separated from solutions containing Y3+. Notably, the Fe2O3 NPs exhibited a higher selectivity for Dy3+ separation compared to Gd3+, possibly due to the smaller ionic radius of Dy3+, which facilitated its binding to active sites on the surface of the Fe2O3 NPs.
S F ( A / B ) = K d ( A ) K d ( B )
where Kd is the adsorption partition coefficient; Kd(A) and Kd(B) denote the Kd values of A and B ions, respectively; and SF(A/B) is the separation coefficient of A and B ions.

3.4. Reusability

The reusability of adsorbents is essential for their practical application [9]. Figure 10 displays the results for the adsorption efficiency of RE3+ on the Fe2O3 NPs as the number of adsorption–desorption cycles increases. In the first cycle, the adsorption efficiencies of the Fe2O3 NPs for the eight types of RE3+ were 97.32%, 97.81%, 98.99%, 99.03%, 99.24%, 99.30%, 99.32%, and 96.53%, respectively. After five adsorption–desorption cycles, a slight decrease in adsorption efficiency was observed, with 95.47%, 95.95%, 97.11%, 97.15%, 97.36%, 97.42%, 97.43%, and 94.70%, respectively. This slight decrease could be attributed to the incomplete release of RE3+ from the adsorption sites during the desorption process, leading to the deactivation of some surface sites [16]. These findings indicated that Fe2O3 NPs could be considered as a reusable and recyclable adsorbent for RE3+.

3.5. Application of Fe2O3 NPs to Actual Mine Wastewater

In actual rare earth mine wastewater, RE3+ commonly co-exists with transition metal and alkaline earth ions, which may interfere with the adsorption of RE3+ [11]. Therefore, the adsorption efficiency of Fe2O3 NPs for RE3+ in the actual mine wastewater was investigated (Figure 11). The main components of metal ions in the mine wastewater are listed in Table 1. After treatment, the adsorption efficiency of all RE3+ in the mine wastewater was found to be above 90.5%, indicating that Fe2O3 NPs effectively adsorb RE3+ from the actual mine wastewater. Moreover, adsorption efficiencies for Ca2+ (4.9%), K+ (10.3%), Na+ (3.7%), Mg2+ (11.5%), Al3+ (17.5%), and Mn2+ (23.7%) were low. Evidently, Fe2O3 NPs exhibited a superior adsorption of RE3+ compared to other metal ions. This could be explained by the fact that the binding of RE3+ to the O donor atoms was stronger than that of the common interfering ions [17].

3.6. Adsorption Mechanism

The morphology and elemental composition of the Fe2O3 NPs before and after the adsorption of RE3+ were investigated using SEM. SEM images revealed that there were more small particles on the surface of the Fe2O3 NPs after adsorbing RE3+ (Figure 12A) compared to the original Fe2O3 NPs (Figure 4D), possibly due to magnetic stirring during adsorption. EDS analysis (Figure 12B) and the full XPS spectrum (Figure 12C) confirmed the presence of REEs on the surface of the Fe2O3 NPs, providing evidence for the adsorption of RE3+. Furthermore, the elemental mapping of the Fe2O3 NPs loaded with RE3+ demonstrated that RE3+ was uniformly adsorbed on the surface of the Fe2O3 NPs (Figure S2).
The O 1s peaks before and after RE3+ adsorption on the Fe2O3 NPs are shown in Figure 12D,E. The O 1s spectrum of the Fe2O3 NPs could be deconvoluted into four peaks located at approximately 530.0, 531.5, 532.3, and 533.5 eV, which corresponded to oxygen atoms bonded to iron atoms (Fe-O), surface hydroxyl (Fe-OH), Fe-SO4, and adsorbed water (H2O), respectively [23]. The presence of Fe-SO4 was attributed to residual iron sulfate on the surface of Fe2O3 NPs, formed after the calcination of pyrite. After the adsorption of RE3+, the surface oxygen content of the Fe2O3 NPs decreased slightly. Furthermore, the Fe-O content decreased from 35.86% to 33.84%, while the Fe-OH and Fe-SO4 contents increased from 33.76% and 21.92% to 36.07% and 30.09%, respectively (Figure 12F). These findings suggested that the O-groups on the surface of Fe2O3 NPs were consumed by binding to RE3+ during adsorption.
Figure 12G presents the FTIR spectra of the Fe2O3 NPs before and after the adsorption of RE3+. The absorption band at 3375 cm−1 was attributed to the stretching vibrations of -OH, and the peak at 1643 cm−1 corresponded to the deformation of the H2O molecule. The absorption bands observed at 1189 and 1095 cm−1 were attributed to the double vibration of SO42- [49], which aligned with the XPS results (Figure 12D,E). Additionally, the fluctuation at 1018 cm−1 represented the in-plane bending vibration of -OH, and the absorption peaks at 475 and 530 cm−1 were related to the tensile vibration of Fe-O [22]. After the adsorption of RE3+, the intensity of the -OH and SO42- absorption bands on the surface of the Fe2O3 NPs decreased significantly, demonstrating the substantial contribution of -OH and SO42- to the adsorption process. The adsorption of RE3+ by -OH and SO42- could be explained by the surface complexation mechanism, which was consistent with the hard–soft acid–base (HSAB) theory. According to HSAB theory, hard acids, such as RE3+, tend to form strong ionic complexes with hard bases like O-, OH-, and SO42- [50,51]. Moreover, at a high pH, -OH ionized to O-, which, together with the O- presented in the Fe2O3 NPs itself, absorbed RE3+ by forming a complex. In addition, no Fe3+ was detected in the solution after adsorption, suggesting that the adsorption mechanism did not involve ion exchanges between Fe3+ and RE3+.
Zeta potential measurements were conducted to examine the presence of electrostatic attraction during the adsorption process [29]. Figure 12H shows that as the pH increased, the concentration of OH- also increased, leading to a decrease in the Zeta potential of Fe2O3 NPs [52]. The zero potential value of the Fe2O3 NPs was −2.17 mV at pH 6.0. However, after the adsorption of RE3+, the zeta potential increased to 12.73 mV, indicating the occurrence of charge neutralization in the adsorption process, and electrostatic attraction was one of the reasons for the adsorption [11]. Therefore, the use of pyrite calcination to prepare Fe2O3 NPs as an adsorbent offered multiple advantages, including electrostatic adsorption, the complexation of RE3+ by -OH and O-, and the presence of additional SO42- adsorption sites due to the incomplete thermal decomposition of the pyrite surface. Based on these findings, a schematic diagram of the adsorption mechanism of Fe2O3 NPs for RE3+ was proposed (Figure 13).

4. Conclusions

In this study, Fe2O3 NPs were successfully synthesized from natural pyrite through a high-temperature phase transition to adsorb mixed RE3+ from mine wastewater. The optimal time and pH for the adsorption of RE3+ by Fe2O3 NPs were 60 min and 6.0, respectively. The maximum adsorption capacities for La3+, Ce3+, Pr3+, Nd3+, Sm3+, Gd3+, Dy3+, and Y3+ were 12.80, 14.02, 14.67, 15.52, 17.66, 19.16, 19.94, and 11.82 mg·g−1, respectively. Fe2O3 NPs exhibited a stronger affinity for Gd3+ and Dy3+, enabling the effective separation of these ions from the mixed RE3+ aqueous solution. Compared to iron oxides adsorbents synthesized using chemical reagents reported in the literature, Fe2O3 NPs demonstrated competitive advantages. The adsorption process was characterized by monolayer chemisorption, which was achieved by surface complexation and electrostatic attraction. After five adsorption–desorption cycles, the removal efficiency of RE3+ remained above 94.7%, indicating the reusability of Fe2O3 NPs. Furthermore, Fe2O3 NPs effectively adsorbed RE3+ from actual mine wastewater, even in the presence of high concentrations of transition metal and alkaline earth ions. Therefore, Fe2O3 NPs, as a simple, reusable, and easily separable adsorbent, show significant potential for the adsorption of RE3+ from actual mine wastewater. This study provided a strategy for the recovery and removal of RE3+ from aqueous solutions and the resource utilization of pyrite. Future research could focus on developing eluents with specific selectivity to improve the purity of rare earth elements.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min14050464/s1, Figure S1: XRD pattern of pyrite; Figure S2: SEM-Mapping images of Fe2O3 NPs after adsorption of RE3+; Table S1: Different levels and factors of the orthogonal experiments on the grinding process of pyrite; Table S2: Parameters of the orthogonal experiments on the grinding process of pyrite.

Author Contributions

C.Z.: Investigation, data analysis, interpretation of results, and writing—original draft. J.W.: resources, funding acquisition, and project administration. B.Y.: writing—review and editing. Y.L.: writing—review and editing and project administration. G.Q.: resources and conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (No. 2022YFC2105300), the National Natural Science Foundation of China (No. 52274288), and Central South University Innovation-Driven Research Program (No. 2023CXQD070).

Data Availability Statement

All data are provided within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Y.; Weng, X.; Chen, Z. Recovery of rare earth elements from mine wastewater using biosynthesized reduced graphene oxide. J. Colloid Interface Sci. 2023, 638, 449–460. [Google Scholar] [CrossRef] [PubMed]
  2. Cui, J.; Wang, Q.; Gao, J.; Guo, Y.; Cheng, F. The selective adsorption of rare earth elements by modified coal fly ash based SBA-15. Chin. J. Chem. Eng. 2022, 47, 155–164. [Google Scholar] [CrossRef]
  3. Liu, C.; Liu, W.S.; Huot, H.; Guo, M.N.; Zhu, S.C.; Zheng, H.X.; Morel, J.L.; Tang, Y.T.; Qiu, R.L. Biogeochemical cycles of nutrients, rare earth elements (REEs) and Al in soil-plant system in ion-adsorption REE mine tailings remediated with amendment and ramie (Boehmeria nivea L.). Sci. Total Environ. 2022, 809, 152075. [Google Scholar] [CrossRef] [PubMed]
  4. Xie, X.; Tan, X.; Yu, Y.; Li, Y.; Wang, P.; Liang, Y.; Yan, Y. Effectively auto-regulated adsorption and recovery of rare earth elements via an engineered E. coli. J. Hazard. Mater. 2022, 424 Pt C, 127642. [Google Scholar] [CrossRef]
  5. Guo, Z.; Li, Q.; Li, Z.; Liu, C.; Liu, X.; Liu, Y.; Dong, G.; Lan, T.; Wei, Y. Fabrication of efficient alginate composite beads embedded with N-doped carbon dots and their application for enhanced rare earth elements adsorption from aqueous solutions. J. Colloid Interface Sci. 2020, 562, 224–234. [Google Scholar] [CrossRef] [PubMed]
  6. Iftekhar, S.; Srivastava, V.; Hammouda, S.B.; Sillanpaa, M. Fabrication of novel metal ion imprinted xanthan gum-layered double hydroxide nanocomposite for adsorption of rare earth elements. Carbohydr. Polym. 2018, 194, 274–284. [Google Scholar] [CrossRef]
  7. Wang, Y.; Wang, G.; Sun, M.; Liang, X.; He, H.; Zhu, J.; Takahashi, Y. Environmental risk assessment of the potential “Chemical Time Bomb” of ion-adsorption type rare earth elements in urban areas. Sci. Total Environ. 2022, 822, 153305. [Google Scholar] [CrossRef] [PubMed]
  8. Liu, X.-R.; Liu, W.-S.; Tang, Y.-T.; Wang, S.-Z.; Cao, Y.-J.; Chen, Z.-W.; Xie, C.-D.; Liu, C.; Guo, M.-N.; Qiu, R.-L. Effects of in situ leaching on the origin and migration of rare earth elements in aqueous systems of South China: Insights based on REE patterns, and Ce and Eu anomalies. J. Hazard. Mater. 2022, 435, 128959. [Google Scholar] [CrossRef]
  9. Liu, Z.; Gao, Z.; Xu, L.; Hu, F. Efficient and rapid adsorption of rare earth elements from water by magnetic Fe3O4/MnO2 decorated reduced graphene oxide. J. Mol. Liq. 2020, 313, 113510. [Google Scholar] [CrossRef]
  10. Zhao, F.; Repo, E.; Meng, Y.; Wang, X.; Yin, D.; Sillanpää, M. An EDTA-β-cyclodextrin material for the adsorption of rare earth elements and its application in preconcentration of rare earth elements in seawater. J. Colloid Interface Sci. 2016, 465, 215–224. [Google Scholar] [CrossRef]
  11. Xu, W.; Yu, C.; Liu, L.; Zhai, Y.; Xu, R.; Hou, H. An O- modified coordination polymer for rapid and selective adsorption of rare earth elements from aqueous solution. Colloids Surf. A 2020, 607, 125464. [Google Scholar] [CrossRef]
  12. Li, J.; Gong, A.; Li, F.; Qiu, L.; Zhang, W.; Gao, G.; Liu, Y.; Li, J. Synthesis and characterization of magnetic mesoporous Fe3O4@mSiO2-DODGA nanoparticles for adsorption of 16 rare earth elements. RSC Adv. 2018, 8, 39149–39161. [Google Scholar] [CrossRef] [PubMed]
  13. Gomes Rodrigues, D.; Monge, S.; Pellet-Rostaing, S.; Dacheux, N.; Bouyer, D.; Faur, C. A new carbamoylmethylphosphonic acid-based polymer for the selective sorption of rare earth elements. Chem. Eng. J. 2019, 371, 857–867. [Google Scholar] [CrossRef]
  14. Gomes Rodrigues, D.; Monge, S.; Pellet-Rostaing, S.; Dacheux, N.; Bouyer, D.; Faur, C. Sorption properties of carbamoylmethylphosphonated-based polymer combining both sorption and thermosensitive properties: New valuable hydrosoluble materials for rare earth elements sorption. Chem. Eng. J. 2019, 355, 871–880. [Google Scholar] [CrossRef]
  15. Chen, G.E.; Sun, D.; Xu, Z.L. Rare Earth Ion from Aqueous Solution Removed by Polymer Enhanced Ultrafiltration Process. Adv. Mater. Res. 2011, 233–235, 959–964. [Google Scholar] [CrossRef]
  16. Javadian, H.; Ruiz, M.; Saleh, T.A.; Sastre, A.M. Ca-alginate/carboxymethyl chitosan/Ni0.2Zn0.2Fe2.6O4 magnetic bionanocomposite: Synthesis, characterization and application for single adsorption of Nd+3, Tb+3, and Dy+3 rare earth elements from aqueous media. J. Mol. Liq. 2020, 306, 112760. [Google Scholar] [CrossRef]
  17. Ashour, R.M.; El-sayed, R.; Abdel-Magied, A.F.; Abdel-khalek, A.A.; Ali, M.M.; Forsberg, K.; Uheida, A.; Muhammed, M.; Dutta, J. Selective separation of rare earth ions from aqueous solution using functionalized magnetite nanoparticles: Kinetic and thermodynamic studies. Chem. Eng. J. 2017, 327, 286–296. [Google Scholar] [CrossRef]
  18. Chen, G.; Jiang, C.; Liu, R.; Xie, Z.; Liu, Z.; Cen, S.; Tao, C.; Guo, S. Leaching kinetics of manganese from pyrolusite using pyrite as a reductant under microwave heating. Sep. Purif. Technol. 2021, 277, 119472. [Google Scholar] [CrossRef]
  19. Tang, J.; Zhao, B.; Lyu, H.; Li, D. Development of a novel pyrite/biochar composite (BM-FeS2@BC) by ball milling for aqueous Cr(VI) removal and its mechanisms. J. Hazard. Mater. 2021, 413, 125415. [Google Scholar] [CrossRef]
  20. Karagiorgakis, A.L.; Schindler, M.; Spiers, G.A. Retention of rare earth elements in authigenic phases following biogeochemical dissolution of ore from Elliot Lake, Ontario. Hydrometallurgy 2018, 177, 9–20. [Google Scholar] [CrossRef]
  21. Li, H.; Li, Y.; Guo, J.; Song, Y.; Hou, Y.; Lu, C.; Han, Y.; Shen, X.; Liu, B. Effect of calcinated pyrite on simultaneous ammonia, nitrate and phosphorus removal in the BAF system and the Fe2+ regulatory mechanisms: Electron transfer and biofilm properties. Environ. Res. 2021, 194, 110708. [Google Scholar] [CrossRef] [PubMed]
  22. Trang, V.T.; Tam, L.T.; Van Quy, N.; Phan, V.N.; Van Tuan, H.; Huy, T.Q.; Dinh, N.X.; Le, A.-T. Enhanced adsorption efficiency of inorganic chromium (VI) ions by using carbon-encapsulated hematite nanocubes. J. Sci. Adv. Mater. Devices 2020, 5, 392–399. [Google Scholar] [CrossRef]
  23. Liang, Y.; Wang, M.; Xiong, J.; Hou, J.; Wang, X.; Tan, W. Al-substitution-induced defect sites enhance adsorption of Pb2+ on hematite. Environ. Sci. Nano 2019, 6, 1323–1331. [Google Scholar] [CrossRef]
  24. Yuan, X.; Luo, F.; Chen, X.; Xia, W.; Zou, Y.; Zhou, X.; Liu, H.; Song, Y.; He, J.; Ma, S. Effective Cu(II) ions adsorption from aqueous solutions using low grade oolitic hematite tailing with phosphorus: Response surface methodology. Desalination Water Treat. 2022, 265, 57–70. [Google Scholar] [CrossRef]
  25. Estes, S.L.; Powell, B.A. Enthalpy of Uranium Adsorption onto Hematite. Environ. Sci. Technol. 2020, 54, 15004–15012. [Google Scholar] [CrossRef] [PubMed]
  26. Denys, A.; Janots, E.; Auzende, A.-L.; Lanson, M.; Findling, N.; Trcera, N. Evaluation of selectivity of sequential extraction procedure applied to REE speciation in laterite. Chem. Geol. 2021, 559, 119954. [Google Scholar] [CrossRef]
  27. Viana, T.; Henriques, B.; Ferreira, N.; Pinto, R.J.B.; Monteiro, F.L.S.; Pereira, E. Insight into the mechanisms involved in the removal of toxic, rare earth, and platinum elements from complex mixtures by Ulva sp. Chem. Eng. J. 2023, 453, 139630. [Google Scholar] [CrossRef]
  28. Zhao, C.; Yang, B.; Liao, R.; Hong, M.; Yu, S.; Liu, S.; Wang, J.; Qiu, G. Combined effect and mechanism of visible light and Ag+ on chalcopyrite bioleaching. Miner. Eng. 2022, 175, 107283. [Google Scholar] [CrossRef]
  29. Yan, Q.; Yang, Y.; Chen, W.; Weng, X.; Owens, G.; Chen, Z. Recovery and removal of rare earth elements from mine wastewater using synthesized bio-nanoparticles derived from Bacillus cereus. Chem. Eng. J. 2023, 459, 141585. [Google Scholar] [CrossRef]
  30. Pechishcheva, N.V.; Estemirova, S.K.; Kim, A.V.; Zaitceva, P.V.; Sterkhov, E.V.; Shchapova, Y.V.; Zhidkov, I.S.; Skrylnik, M.Y. Adsorption of hexavalent chromium on mechanically activated graphite. Diam. Relat. Mater. 2022, 127, 109152. [Google Scholar] [CrossRef]
  31. El Ouardi, Y.; Lamsayah, M.; Butylina, S.; Geng, S.; Esmaeili, M.; Giove, A.; Massima Mouele, E.S.; Virolainen, S.; El Barkany, S.; Ouammou, A.; et al. Sustainable composite material based on glutenin biopolymeric-clay for efficient separation of rare earth elements. Chem. Eng. J. 2022, 440, 135959. [Google Scholar] [CrossRef]
  32. Ren, J.; Zheng, L.; Su, Y.; Meng, P.; Zhou, Q.; Zeng, H.; Zhang, T.; Yu, H. Competitive adsorption of Cd(II), Pb(II) and Cu(II) ions from acid mine drainage with zero-valent iron/phosphoric titanium dioxide: XPS qualitative analyses and DFT quantitative calculations. Chem. Eng. J. 2022, 445, 136778. [Google Scholar] [CrossRef]
  33. Smith, Y.R.; Bhattacharyya, D.; Willhard, T.; Misra, M. Adsorption of aqueous rare earth elements using carbon black derived. Chem. Eng. J. 2016, 296, 102–111. [Google Scholar] [CrossRef]
  34. Ambroz, A.; Ban, I.; Luxbacher, T. Assessment of the Capability of Magnetic Nanoparticles to Recover Neodymium Ions from Aqueous Solution. Acta Chim. Slov. 2022, 69, 826–836. [Google Scholar] [CrossRef] [PubMed]
  35. He, P.; Zhu, J.; Chen, Y.; Chen, F.; Zhu, J.; Liu, M.; Zhang, K.; Gan, M. Pyrite-activated persulfate for simultaneous 2,4-DCP oxidation and Cr(VI) reduction. Chem. Eng. J. 2021, 406, 126758. [Google Scholar] [CrossRef]
  36. Labus, M. Pyrite thermal decomposition in source rocks. Fuel 2021, 287, 119529. [Google Scholar] [CrossRef]
  37. Abeshu, G.W.; Li, H.-Y.; Zhu, Z.; Tan, Z.; Leung, L.R. Median bed-material sediment particle size across rivers in the contiguous US. Earth Syst. Sci. Data 2022, 14, 929–942. [Google Scholar] [CrossRef]
  38. Wang, Z.; Xie, X.; Xiao, S.; Liu, J. Comparative study of interaction between pyrite and cysteine by thermogravimetric and electrochemical techniques. Hydrometallurgy 2010, 101, 88–92. [Google Scholar] [CrossRef]
  39. Cheng, H.; Liu, Q.; Huang, M.; Zhang, S.; Frost, R.L. Application of TG-FTIR to study SO2 evolved during the thermal decomposition of coal-derived pyrite. Thermochim. Acta 2013, 555, 1–6. [Google Scholar] [CrossRef]
  40. Yan, X.; Shao, J.; Wen, Q.; Shen, J. Stabilization of soil arsenic by natural limonite after mechanical activation and the associated mechanisms. Sci. Total Environ. 2020, 708, 135118. [Google Scholar] [CrossRef]
  41. Yu, S.H.; Yao, Q.Z.; Zhou, G.T.; Fu, S.Q. Preparation of hollow core/shell microspheres of hematite and its adsorption ability for samarium. ACS Appl. Mater. Interfaces 2014, 6, 10556–10565. [Google Scholar] [CrossRef] [PubMed]
  42. Tan, K.L.; Hameed, B.H. Insight into the adsorption kinetics models for the removal of contaminants from aqueous solutions. J. Taiwan Inst. Chem. Eng. 2017, 74, 25–48. [Google Scholar] [CrossRef]
  43. Poudel, M.B.; Awasthi, G.P.; Kim, H.J. Novel insight into the adsorption of Cr(VI) and Pb(II) ions by MOF derived Co-Al layered double hydroxide @hematite nanorods on 3D porous carbon nanofiber network. Chem. Eng. J. 2021, 417, 129312. [Google Scholar] [CrossRef]
  44. Hughes, I.D.; Dane, M.; Ernst, A.; Hergert, W.; Luders, M.; Poulter, J.; Staunton, J.B.; Svane, A.; Szotek, Z.; Temmerman, W.M. Lanthanide contraction and magnetism in the heavy rare earth elements. Nature 2007, 446, 650–653. [Google Scholar] [CrossRef] [PubMed]
  45. Kujawa, J.; Al Gharabli, S.; Szymczyk, A.; Terzyk, A.P.; Boncel, S.; Knozowska, K.; Li, G.; Kujawski, W. On membrane-based approaches for rare earths separation and extraction—Recent developments. Coord. Chem. Rev. 2023, 493, 215340. [Google Scholar] [CrossRef]
  46. Molina, L.; Gaete, J.; Alfaro, I.; Ide, V.; Valenzuela, F.; Parada, J.; Basualto, C. Synthesis and characterization of magnetite nanoparticles functionalized with organophosphorus compounds and its application as an adsorbent for La (III), Nd (III) and Pr (III) ions from aqueous solutions. J. Mol. Liq. 2019, 275, 178–191. [Google Scholar] [CrossRef]
  47. Javadian, H.; Ruiz, M.; Sastre, A.M. Response surface methodology based on central composite design for simultaneous adsorption of rare earth elements using nanoporous calcium alginate/carboxymethyl chitosan microbiocomposite powder containing Ni0.2Zn0.2Fe2.6O4 magnetic nanoparticles: Batch and column studies. Int. J. Biol. Macromol. 2020, 154, 937–953. [Google Scholar] [CrossRef] [PubMed]
  48. Han, L.; Peng, Y.; Ma, J.; Shi, Z.; Jia, Q. Construction of hypercrosslinked polymers with styrene-based copolymer precursor for adsorption of rare earth elements. Sep. Purif. Technol. 2022, 285, 120378. [Google Scholar] [CrossRef]
  49. Zhu, J.; Chen, F.; Gan, M. Controllable biosynthesis of nanoscale schwertmannite and the application in heavy metal effective removal. Appl. Surf. Sci. 2020, 529, 147012. [Google Scholar] [CrossRef]
  50. Zhou, F.; Xiao, Y.; Guo, M.; Wang, S.; Qiu, R.; Morel, J.L.; Simonnot, M.O.; Zhang, W.X.; Zhang, W.; Tang, Y.T. Insights into the Selective Transformation of Ceria Sulfation and Iron/Manganese Mineralization for Enhancing the Selective Recovery of Rare Earth Elements. Environ. Sci. Technol. 2023, 57, 3357–3368. [Google Scholar] [CrossRef]
  51. Gunawardhana, B.P.N.; Gunathilake, C.A.; Dayananda, K.E.D.Y.T.; Dissanayake, D.M.S.N.; Mantilaka, M.M.M.G.P.G.; Kalpage, C.S.; Rathnayake, R.M.L.D.; Rajapakse, R.M.G.; Manchanda, A.S.; Etampawala, T.N.B.; et al. Synthesis of Hematite Nanodiscs from Natural Laterites and Investigating Their Adsorption Capability of Removing Ni2+ and Cd2+ Ions from Aqueous Solutions. J. Compos. Sci. 2020, 4, 57. [Google Scholar] [CrossRef]
  52. Zhang, C.; Yu, Z.; Zeng, G.; Huang, B.; Dong, H.; Huang, J.; Yang, Z.; Wei, J.; Hu, L.; Zhang, Q. Phase transformation of crystalline iron oxides and their adsorption abilities for Pb and Cd. Chem. Eng. J. 2016, 284, 247–259. [Google Scholar] [CrossRef]
Figure 1. Particle size distribution histograms and cumulative distribution curves for milled pyrite ((AI) represent the tests 1–9 of the orthogonal experiments of the grinding process of pyrite (Table S2)).
Figure 1. Particle size distribution histograms and cumulative distribution curves for milled pyrite ((AI) represent the tests 1–9 of the orthogonal experiments of the grinding process of pyrite (Table S2)).
Minerals 14 00464 g001
Figure 2. (A) TG-DTG-DSC curves for pyrite in the air atmosphere, (B) XRD patterns and (C) semi-quantitation analysis of pyrite after being calcined in the air atmosphere at different temperatures and times.
Figure 2. (A) TG-DTG-DSC curves for pyrite in the air atmosphere, (B) XRD patterns and (C) semi-quantitation analysis of pyrite after being calcined in the air atmosphere at different temperatures and times.
Minerals 14 00464 g002
Figure 3. Schematic illustration of the preparation of Fe2O3 NPs.
Figure 3. Schematic illustration of the preparation of Fe2O3 NPs.
Minerals 14 00464 g003
Figure 4. SEM image of (A) fine-grained pyrite and (B) calcined pyrite, (C) TEM image, (D) SEM image, (E) EDS analysis, and (F) N2 adsorption–desorption isotherms and Barrett–Joyner–Halenda plot of Fe2O3 NPs.
Figure 4. SEM image of (A) fine-grained pyrite and (B) calcined pyrite, (C) TEM image, (D) SEM image, (E) EDS analysis, and (F) N2 adsorption–desorption isotherms and Barrett–Joyner–Halenda plot of Fe2O3 NPs.
Minerals 14 00464 g004
Figure 5. (A) Effect of adsorption time on the adsorption of RE3+ by Fe2O3 NPs and (BI) pseudo-first-order and pseudo-second-order fitting of the adsorption data on Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 50 mg, adsorption time = 0–240 min, 298 K, pH = 5.0, V = 50 mL).
Figure 5. (A) Effect of adsorption time on the adsorption of RE3+ by Fe2O3 NPs and (BI) pseudo-first-order and pseudo-second-order fitting of the adsorption data on Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 50 mg, adsorption time = 0–240 min, 298 K, pH = 5.0, V = 50 mL).
Minerals 14 00464 g005
Figure 6. (A) Effect of pH on the adsorption of RE3+ by Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 50 mg, adsorption time = 60 min, 298 K, pH = 2.0–7.0, V = 50 mL) and (B) Zeta potential of Fe2O3 NPs before the adsorption of RE3+.
Figure 6. (A) Effect of pH on the adsorption of RE3+ by Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 50 mg, adsorption time = 60 min, 298 K, pH = 2.0–7.0, V = 50 mL) and (B) Zeta potential of Fe2O3 NPs before the adsorption of RE3+.
Minerals 14 00464 g006
Figure 7. Distribution of hydrolysis products of RE3+ in aqueous solutions: (A) La3+, (B) Ce3+, (C) Pr3+, (D) Nd3+, (E) Sm3+, (F) Gd3+, (G) Dy3+, (H) Y3+ (experimental conditions: initial RE3+ concentration = 5 mg·L−1, 298 K).
Figure 7. Distribution of hydrolysis products of RE3+ in aqueous solutions: (A) La3+, (B) Ce3+, (C) Pr3+, (D) Nd3+, (E) Sm3+, (F) Gd3+, (G) Dy3+, (H) Y3+ (experimental conditions: initial RE3+ concentration = 5 mg·L−1, 298 K).
Minerals 14 00464 g007
Figure 8. (A) Effect of initial RE3+ concentration on the adsorption of RE3+ by Fe2O3 NPs and (BI) the adsorption data of RE3+ onto Fe2O3 NPs fitting by Langmuir and Freundlich isotherm models (experimental conditions: initial RE3+ concentration = 5–40 mg·L−1, adsorbent dosage = 50 mg, V = 50 mL, adsorption time = 60 min, pH = 6.0, 298 K).
Figure 8. (A) Effect of initial RE3+ concentration on the adsorption of RE3+ by Fe2O3 NPs and (BI) the adsorption data of RE3+ onto Fe2O3 NPs fitting by Langmuir and Freundlich isotherm models (experimental conditions: initial RE3+ concentration = 5–40 mg·L−1, adsorbent dosage = 50 mg, V = 50 mL, adsorption time = 60 min, pH = 6.0, 298 K).
Minerals 14 00464 g008
Figure 9. (A) Effect of temperature on the adsorption of RE3+ by Fe2O3 NPs and (B) linear plot of ln Kd versus 1/T for adsorption capacity of RE3+ on Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 20 mg·L−1, adsorbent dosage = 50 mg, T = 60 min, V = 50 mL, pH = 6.0).
Figure 9. (A) Effect of temperature on the adsorption of RE3+ by Fe2O3 NPs and (B) linear plot of ln Kd versus 1/T for adsorption capacity of RE3+ on Fe2O3 NPs (experimental conditions: initial RE3+ concentration = 20 mg·L−1, adsorbent dosage = 50 mg, T = 60 min, V = 50 mL, pH = 6.0).
Minerals 14 00464 g009
Figure 10. Reusability of Fe2O3 NPs for adsorption of RE3+ (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 100 mg, T = 60 min, V = 50 mL, pH = 6.0, 298 K).
Figure 10. Reusability of Fe2O3 NPs for adsorption of RE3+ (experimental conditions: initial RE3+ concentration = 5 mg·L−1, adsorbent dosage = 100 mg, T = 60 min, V = 50 mL, pH = 6.0, 298 K).
Minerals 14 00464 g010
Figure 11. Adsorption of Fe2O3 NPs for RE3+ in rare earth mine wastewater.
Figure 11. Adsorption of Fe2O3 NPs for RE3+ in rare earth mine wastewater.
Minerals 14 00464 g011
Figure 12. (A) SEM image and (B) EDS spectra of Fe2O3 NPs after the adsorption of RE3+, (C) full-range survey XPS spectrum of Fe2O3 NPs before and after the adsorption of RE3+, XPS spectra of the O 1s peak of Fe2O3 NPs (D) before and (E) after the adsorption of RE3+, (F) contents of different oxygen species on the Fe2O3 NPs surface (atom%), (G) FTIR spectra and (H) zeta potential of Fe2O3 NPs before and after the adsorption of RE3+.
Figure 12. (A) SEM image and (B) EDS spectra of Fe2O3 NPs after the adsorption of RE3+, (C) full-range survey XPS spectrum of Fe2O3 NPs before and after the adsorption of RE3+, XPS spectra of the O 1s peak of Fe2O3 NPs (D) before and (E) after the adsorption of RE3+, (F) contents of different oxygen species on the Fe2O3 NPs surface (atom%), (G) FTIR spectra and (H) zeta potential of Fe2O3 NPs before and after the adsorption of RE3+.
Minerals 14 00464 g012
Figure 13. Schematic diagram of the adsorption mechanism of Fe2O3 NPs for RE3+.
Figure 13. Schematic diagram of the adsorption mechanism of Fe2O3 NPs for RE3+.
Minerals 14 00464 g013
Table 1. Concentrations of various ions in the mine wastewater used in this study.
Table 1. Concentrations of various ions in the mine wastewater used in this study.
IonsConcentration (mg·L−1)
Y3.019
La1.630
Nd1.461
Dy0.432
Gd0.399
Pr0.356
Sm0.341
Ce0.327
Ca34.884
K11.151
Na7.502
Mg5.982
Al3.060
Mn2.025
Table 2. Results of the orthogonal design experiments for pyrite ground by a planetary ball mill.
Table 2. Results of the orthogonal design experiments for pyrite ground by a planetary ball mill.
TestBall-to-Material Ratio (A)Rotation Speed, Rpm (B)Grinding Time, h (C)D50 of Pyrite, μm
110:140017.563
210:150025.132
310:160035.956
415:140024.803
515:150035.912
615:160014.669
720:140035.841
820:150015.137
920:160027.282
I18.65118.20717.369T = 52.295
II15.38416.18117.217
III18.26017.90717.709
K16.2176.0695.790
K25.1285.3945.739
K36.0875.9695.903
R1.0890.6750.164
Notes: I, II, and III are the estimated values of the horizontal effects of 1, 2, and 3 on each corresponding column (factor), which are calculated as Ii (IIi, IIIi) = sum of data corresponding to level 1 (2, 3) in column i; Ki is the combined average of the data of level i = i/number of repetitions of level i; the R line is called the extreme difference (maximum–minimum), indicating the magnitude of the effect of the factor on the results.
Table 3. Kinetic parameters for the adsorption of RE3+ onto Fe2O3 NPs.
Table 3. Kinetic parameters for the adsorption of RE3+ onto Fe2O3 NPs.
Ionsqe,exp/
(mg·g−1)
Pseudo-First OrderPseudo-Second Order
qe,cal/
(mg·g−1)
k1/
min−1
R2qe,cal/
(mg·g−1)
k2/
g·mg−1·min−1
R2
La3+0.990.950.480.9841.000.850.999
Ce3+1.191.150.700.9881.191.230.996
Pr3+1.431.360.880.9831.391.470.992
Nd3+1.531.480.870.9881.511.360.997
Sm3+1.831.780.750.9911.830.910.998
Gd3+1.931.890.720.9971.940.840.998
Dy3+2.072.010.680.9942.080.710.998
Y3+0.580.530.420.9470.561.120.986
Table 4. Adsorption isotherm constants for RE3+ adsorption on Fe2O3 NPs.
Table 4. Adsorption isotherm constants for RE3+ adsorption on Fe2O3 NPs.
REEsLangmuir ConstantsFreundlich Constants
qm/(mg·g−1)kL/(L·mg−1)R2KF/(mg·g−1)1/nR2
La12.8040.0400.9920.8400.6250.991
Ce14.0230.0430.9930.9860.6170.991
Pr14.6650.0450.9911.0720.6120.990
Nd15.5200.0480.9921.1980.6040.991
Sm17.6590.0570.9901.5950.5830.988
Gd19.1600.0670.9911.9800.5640.990
Dy19.9400.0740.9902.2370.5520.986
Y11.8160.0380.9940.7360.6320.991
Table 5. Reported literature review of mixed RE3+ adsorption capacities by different iron oxide adsorbents.
Table 5. Reported literature review of mixed RE3+ adsorption capacities by different iron oxide adsorbents.
AdsorbentsIonsAdsorption Capacity
(mg·g−1)
Ref.
Hollow core/shell hematite microspheresSm14.48[41]
Functionalized Fe3O4 NPsLa, Nd, Gd, Y32.5, 41.0, 52.0, 35.8[17]
Calcium alginate/carboxymethyl chitosan/Ni0.2Zn0.2Fe2.6O4 Nd, Tb, Dy23.15, 24.41, 25.24[47]
Magnetite nanoparticles functionalized with organophosphorus compoundsLa, Pr, Nd8.3, 8.7, 8.9[46]
Fe2O3 NPs (this work)La, Ce, Pr, Nd, Sm, Gd, Dy, Y12.8, 14.0, 14.7, 15.5, 17.7, 19.2, 19.9, 11.8
Table 6. Thermodynamic parameters for the adsorption of RE3+ by Fe2O3 NPs at 298–318 K.
Table 6. Thermodynamic parameters for the adsorption of RE3+ by Fe2O3 NPs at 298–318 K.
RE3+ Δ G 0 (kJ·mol−1) Δ H 0
(kJ·mol−1)
Δ S 0
(J·mol−1·K−1)
R2
298K308K318K
La1.8571.5881.27410.55129.1490.997
Ce1.4351.1300.77411.27833.0020.997
Pr1.2120.8860.50611.72335.2430.997
Nd0.9100.5550.13812.40038.5250.996
Sm0.111−0.336−0.87114.73549.0270.994
Gd−0.525−1.065−1.72817.37860.0120.992
Dy−0.902−1.508−2.26719.41368.0910.990
Y2.2021.9591.67510.05026.3130.997
Table 7. Selective separation factors (SF) of Gd3+ and Dy3+ from the mixed RE3+ solutions using Fe2O3 NPs.
Table 7. Selective separation factors (SF) of Gd3+ and Dy3+ from the mixed RE3+ solutions using Fe2O3 NPs.
SF298K308K318KSF298K308K318K
Gd3+/La3+2.622.823.11Dy3+/La3+3.053.353.82
Gd3+/Ce3+2.212.362.58Dy 3+/Ce3+2.572.803.16
Gd3+/Pr3+2.022.142.33Dy 3+/Pr3+2.352.552.85
Gd3+/Nd3+1.781.882.03Dy 3+/Nd3+2.082.242.48
Gd3+/Sm3+1.291.331.38Dy3+/Sm3+1.511.581.70
Gd3+/Dy3+0.860.840.82Dy 3+/Gd3+1.161.191.23
Gd3+/Y3+3.013.263.62Dy 3+/Y3+3.503.874.44
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

Zhao, C.; Wang, J.; Yang, B.; Liu, Y.; Qiu, G. Selective Separation of Rare Earth Ions from Mine Wastewater Using Synthetic Hematite Nanoparticles from Natural Pyrite. Minerals 2024, 14, 464. https://doi.org/10.3390/min14050464

AMA Style

Zhao C, Wang J, Yang B, Liu Y, Qiu G. Selective Separation of Rare Earth Ions from Mine Wastewater Using Synthetic Hematite Nanoparticles from Natural Pyrite. Minerals. 2024; 14(5):464. https://doi.org/10.3390/min14050464

Chicago/Turabian Style

Zhao, Chunxiao, Jun Wang, Baojun Yang, Yang Liu, and Guanzhou Qiu. 2024. "Selective Separation of Rare Earth Ions from Mine Wastewater Using Synthetic Hematite Nanoparticles from Natural Pyrite" Minerals 14, no. 5: 464. https://doi.org/10.3390/min14050464

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