Next Article in Journal
Impact of Recycled Process Water on Electrochemical Reactivity of Sulphide Ore
Previous Article in Journal
Prospectivity Mapping of Heavy Mineral Ore Deposits Based upon Machine-Learning Algorithms: Columbite-Tantalite Deposits in West- Central Côte d’Ivoire
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of Hydrated Pr3+ and NH4+/Mg2+ Ions onto the (001) Surface of Montmorillonite: A DFT Analysis with Experimental Verification

1
School of Resource and Environmental Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
2
Ganzhou Rare Earth Mining Co., Ltd., Ganzhou 341000, China
3
School of Architecture and Design, Jiangxi University of Science and Technology, Ganzhou 341000, China
4
WA School of Mines: Minerals, Energy and Chemical Engineering, Faculty of Science and Engineering, Curtin University, Perth, WA 6150, Australia
5
School of Civil and Surveying & Mapping Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
6
Jiangxi Ecological Environment Monitoring Center, Nanchang 330000, China
7
School of Emergency Management and Safety Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
*
Authors to whom correspondence should be addressed.
Minerals 2022, 12(11), 1454; https://doi.org/10.3390/min12111454
Submission received: 9 October 2022 / Revised: 5 November 2022 / Accepted: 15 November 2022 / Published: 18 November 2022
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
Montmorillonite is a major mineral present in ion-adsorption rare earth ores, and the microscopic adsorption states of rare earth ions on its surface are of a great significance for the efficient exploitation of ion-adsorption rare earth ores. In this article, density functional theory calculations were used to investigate the adsorption mechanisms and bonding characteristics of hydrated Pr, Mg and NH4 ions on the (001) surface of montmorillonite. Pr3+ exhibited a directed tendency geometry with Pr(H2O)103+, which was adsorbed onto montmorillonite by hydrogen bonding with an adsorption energy of −1182 kJ/mol, and one coordinated H2O ligand was separated from the first hydration layer of Pr. Both hydrated Mg and NH4 ions were adsorbed onto the montmorillonite surface through hydrogen bonds, and the adsorption energies were −206 and −188 kJ/mol, respectively, indicating that the adsorption stability of the hydrated Mg ion was slightly higher than that of the hydrated NH4 ion, but both were lower than that of hydrated Pr (−1182 kJ/mol). Hence, higher concentrations of Mg and NH4 ions than rare earth ions would be necessary in the leaching process of ion-adsorption rare earth ores. Additionally, desorption experiments revealed that the recovery of Pr3+ by Mg2+ with a concentration of 38 mmol/L is 80%, while it is only 65% with the same concentration of NH4+, and the concentrations of Mg2+ and NH4+ were much higher than that of Pr3+ in lixivium, which is consistent with the DFT calculations.

1. Introduction

Rare earths elements (REEs) are known as industrial vitamins and are widely used in the fields of permanent magnet materials, petroleum catalysis, metal products and optical devices due to their unique 4f electronic layer structure [1,2,3]. Ion-adsorption rare earth (RE) ores, which are the main source of REEs in the world, have the characteristics of a complete distribution and low mining cost and have always attracted the attention of scholars. Ion-adsorption RE ores are unique ores that are formed by the following processes: first, RE original rocks (such as igneous rocks and granite) are gradually weathered in a warm and humid environment, and the RE ions in them are released into the solution. With the continuous migration of rainwater, adsorption and desorption occur continuously on the surfaces of kaolinite, montmorillonite (Mt) and other clays during the migration process [4,5]. The occurrence forms of RE in ion-adsorption rare earth ores can be divided into four types, namely, the water-soluble phase, hydrated or hydroxyl-hydrated ionic phase, colloidal sedimentary phase and mineral phase, among which the hydrated or hydroxyl-hydrated ionic phase dominates [6], which can be desorbed by an ion exchange with highly active electrolytes (ammonium sulphate and magnesium sulphate) [7,8]. However, the use of ammonium sulphate would lead to the eutrophication of water around a mine lot.
Mt is a typical 2:1-type layered silicate mineral, which is composed of two silicon–oxygen tetrahedra mixed with an aluminium-oxygen octahedron in the middle [9]. Mt has a large number of negative charges due to the isomorphic substitution of silicon and aluminium by low-valence aluminium and magnesium, respectively, hence, Mt can adsorb positively charged metal cations [10]. Alshameri et al. [11] systematically studied and compared the adsorption/desorption of La3+ and Yb3+ by kaolinite, Mt, muscovite and illite and found that Mt has the highest adsorption capacity; they also found that the maximum adsorption capacity of Yb3+ is approximately 8.5 mg/g, which is slightly higher than that of La3+, and that the desorption of the RE ions from Mt is more difficult than that from kaolinite or illite. In addition, Zhou et al. [12] investigated the adsorption of rare earth elements in sulfuric acid-modified Mt and found that the adsorption capacities of Mt for La3+ and Y3+ were significantly improved. The adsorption conformed to the kinetic second-order reaction model. Additionally, a chemical adsorption was observed, and the Arrhenius activation energies of La3+ and Y3+ were determined to be 14.259 kJ/mol and 22.845 kJ/mol, respectively. These studies mainly revealed the adsorption/desorption characteristics of RE in Mt through macroscopic test methods, while research on the adsorption of RE in Mt at the atomic and molecular scales has been rarely reported, especially regarding the comparison of RE ions and cations of the leaching agents.
With the development of computer technology, quantum chemical calculations have been gradually applied to scientific research [13,14], especially the density functional theory (DFT), which is an important method for supplementing experimental results and gaining new insights into chemistry-related problems [15,16]. Qiu et al. [17] studied the adsorption behaviour of hydroxyl-hydrated lutetium ions on the surface of kaolinite and found that on an ideal kaolinite (001) surface, the lutetium ions are mainly adsorbed through hydrogen bonding, which is called an outer-sphere adsorption, and the adsorption energy on the silicon–oxygen surface is lower than that on the surface of aluminium hydroxyl. In contrast, coordination bonds are formed between lutetium ions and the kaolinite surface if the surface is deprotonated, which is called an inner-sphere adsorption, and the adsorption energy is much lower than that of outer-sphere adsorption. In addition, Peng et al. [18] investigated the adsorption of [Y(OH)2]+ onto the kaolinite (001) surface as a function of the pH value and found that [Y(OH)2]+ ions are adsorbed onto the aluminium hydroxyl surface by coordination and electrostatic bonding, while the adsorption onto the silicon–oxygen surface is mainly through an electrostatic attraction. These works used computer simulation technology to investigate the adsorption behaviour of RE ions onto the surface of kaolinite at the atomic level and clarified the adsorption mechanism. Similar to kaolinite, Mt is a host mineral of RE ions in ion-adsorption RE ores. Therefore, it is of great significance to simulate the adsorption behaviours of RE ions and other cations onto the surface of Mt by the DFT.
In this article, the first-principles method based on the density functional theory was used to investigate the hydrated structure of Pr(III) ions in aqueous solution, and the adsorption energy, bonding mechanisms, Mulliken charge and density of the states of hydrated Pr(III) ions adsorbed onto the Mt (001) surface were studied. Based on the leaching agent used in the mining of ion-adsorption RE ores, the adsorption behaviours of hydrated magnesium ions and hydrated ammonium ions onto the (001) surface of Mt were also simulated using the DFT, which were compared with the adsorption behaviour of hydrated Pr(III) ions. In addition, to verify the accuracy of the calculation results, the adsorption and desorption experiments of Pr(III) ions on Mt were studied. This study provides theoretical guidance for the leaching of RE ions from Mt and the selection of leaching agent cations.

2. Models and Experiments

2.1. DFT Model Methodology

First-principles calculations were performed in the framework of the DFT using the Cambridge Sequential Total Energy Package (CASTEP) (Cambridge, UK) [19,20,21]. The Perdew–Burke–Ernzerhof (PBE) generalised gradient approximation (GGA) was applied to analyse the exchange-correlation potentials [22]. The interactions between the nuclei and valence electrons were approximated with the Vanderbilt ultrasoft pseudopotential [23]. The cut-off energy of the plane-wave basis set was 360 eV. K-point grids (3 × 2 × 1) were used for the Mt crystal and surface models. The atomic positions were optimised using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm until the energy, atomic force and displacement converged to less than 5 × 10−6 eV/atom, 0.07 eV/Å and 0.005 Å, respectively. The system was optimised in a 15 × 15 × 15 Å3 periodic box. When MS stated “Geometry optimization completed successfully”, that means the system is in equilibrium, and the “Final Enthalpy” was the energy of the system. The hydrated Pr3+, NH4+ and Mg2+ complex in the centre of a silicate ring was used as the initial adsorption structure.
The stability of the hydrated species was evaluated in terms of the binding energy (Ebind), which is defined as follows (1):
E b i n d i n g = E t o l     E P r n E H 2 O
where Ebinding is the binding energy of hydrated Pr3+, Etol is the total energy of the hydrated Pr3+ system, EPr is the energy of Pr3+, EH2O represents the energy of H2O and n is the number of H2O molecules.
The feasibility of the adsorption was evaluated based on the adsorption energy (Eads), which is defined as Equation (2):
E a d s = E M / S     E M     E S
where Eads represents the adsorption energy of the hydrated Pr3+/NH4+/Mg2+ complex, EM/S is the total energy of the hydrated Pr3+/NH4+/Mg2+ complex system after the adsorption and EM and ES denote the total energies of the hydrated Pr3+/NH4+/Mg2+ complex and Mt (001) surface, respectively.

2.2. Experimental Details

First, 5 g Mt and 125 mL of chloride praseodymium with a concentration of 2 g/L at pH 5 were mixed in a conical flask in a mechanical shaker at a speed of 180 r/min and a temperature of 25 °C for 240 min. The solution’s pH was adjusted by sulfuric acid, and no other ions were added. Then, the adsorbent was separated by vacuum filtration with a sand core funnel and rinsed with deionised water several times. After that, the adsorbent was dissolved with nitric acid, hydrochloric acid, sulfuric acid and perchloric acid, and the concentration of Pr3+ was determined by inductively coupled plasma–optical emission spectrometry (ICP–OES). The equilibrium adsorption capacities (qe) of the adsorbents were calculated by using the following equation:
q e = m M
where qe is the adsorption capacity (mg/g) and m and M are the masses of the adsorbed Pr3+ and Mt, respectively.
The desorption efficiencies of Pr3+ from Mt by MgSO4 and (NH4)2SO4 were investigated with a solid/liquid ratio of 1:50 at a pH value of 2.5. In the desorption experiments, the molarity of the Mg ions was the same as that of the NH4 ions in the range of 0.015 to 0.11 mol/L. When the desorption reached equilibrium after 240 min, the mixed liquid was vacuum filtrated with a sand core funnel, the supernatant was collected and the Pr3+ content was determined by ICP–OES. The desorption efficiencies were calculated by the following equation:
Recovery   ( % ) = C V m   q e × 100
where C is the concentration of Pr3+ in the solution after desorption, V is the volume of the desorption solution and m and qe are the mass and adsorption capacity, respectively, of Mt.

3. Results and Discussion

3.1. Geometries of Pr(H2O)n3+

Qiu et al. [17,24,25] studied the hydration numbers of hydrated La3+, Y3+ and Lu3+ ions and found that the most stable hydrated structures of those RE ions have 9–10 H2O ligands. Hence, Pr3+ coordination numbers larger than six were studied. Figure 1 shows the equilibrium geometric structure of Pr(H2O)6−103+. It was found that Pr3+ is connected with H2O molecules by covalent bonds. The hydrate maintains a high symmetry when the number of H2O molecules is six, whereas the space around Pr3+ becomes more crowded with the increase in the number of H2O ligands.
The equilibrium geometric parameters and binding energy of [Pr(H2O)6−10]3+ are shown in Table 1. The average bond length (Rmean) of the Pr–O bond increases with the increasing number of H2O ligands. Because there is a steric hindrance between the H2O ligands, which is not conducive to the subsequent coordination of H2O and Pr3+, the average bond length increases. Additionally, the bond length between one H2O molecule and Pr3+ was found to be 3.03 Å when the coordination number was 10, indicating that H2O escapes the first hydration layer of Pr3+ when the coordination number of H2O is higher than 10. More H2O ligands provide electrons to Pr3+ with an increasing coordination number, resulting in decreases in the charge and ionicity of the Pr3+, making it more stable. In addition, the binding energy of hydrated [Pr(H2O)6−10]3+ was found to decrease with an increasing number of H2O ligands, indicating that it becomes more stable. It was found that [Pr(H2O)10]3+ is the most stable structure with the lowest binding energy of −2680 kJ/mol, which was used as the initial structure of hydrated Pr3+ adsorbed on the Mt (001) surface.
The Mulliken charge population and difference charge density of [Pr(H2O)10]3+ are shown in Table 2 and Figure 2. Compared with a Pr(III) ion, the charge of hydrated Pr decreases from +3 to +2.13, that is, Pr loses a total of 2.13 electrons; specifically, the 2s, 2p, 5d and 4f orbitals obtain 0.13 and lose 0.29, 0.17 and 1.81 electrons, respectively. As a result, the coordinated O of H2O gains electrons overall, with a charge between −0.92 and 0.99, that maybe due to the electron density lowering of the adjacent H atoms and Pr atom. Figure 2 shows the difference charge density of [Pr(H2O)10]3+; the red area represents the gain of the electrons, and the blue area represents the loss of the electrons. It is obvious that the H atoms are surrounded by blue areas because they are attracted by the electronegative O atom; hence, the O atom is surrounded by a red area. There is also a blue area in the periphery of Pr, which is consistent with the results in Table 2.

3.2. [Pr(H2O)n]3+ Adsorbed on the (001) Surface of Mt

The (001) surface of Mt is rich in oxygen atoms, which can form hydrogen bonds with the H2O ligands of [Pr(H2O) 10]3+ as this ion approaches the (001) surface. Figure 3 shows the equilibrium geometry of [Pr(H2O) 10]3+ adsorbed on the (001) surface. Mt has a strong attraction to the positively charged [Pr(H2O) 10]3+ due to the large number of negative charges on the surface of Mt, hence, [Pr(H2O) 10]3+ is relatively close to the surface. As a result, four hydrogen bonds were observed between the H atom in the H2O ligand and the surface O atom, with distances of 2.397, 1.534, 1.772 and 2.424 Å, two of which are close to the range of chemical bonds, indicating a strong interaction between [Pr(H2O) 10]3+ and the surface. Due to the steric hindrance effect on the surface of Mt, the coordinating H2O molecules are squeezed to varying degrees when [Pr(H2O)10]3+ approaches the surface; consequently, a coordinating H2O molecule with a weak connection to the Pr atom moves away from the first hydration layer and becomes a free H2O molecule.
The adsorption energy and structural parameters of [Pr(H2O)10]3+ adsorbed on the (001) surface are listed in Table 3. The number of H2O ligands of the Pr atom is reduced from 10 to 9, and the average length of the Pr–O bond is reduced from 2.78 Å to 2.59 Å after the adsorption. The adsorption energy is −1182 kJ/mol, indicating that the adsorption of [Pr(H2O)10]3+ on the (001) surface of Mt is a spontaneous exothermic process. The adsorption energy is significantly lower than the values reported by other scholars on the Si–O surface of kaolinite (−473.37 kJ/mol) [24], suggesting that the adsorption stability of RE ions on the surface of Mt is significantly higher than that on kaolinite. This is consistent with a previous study [12], showing that it is more difficult to desorb RE ions from Mt than from kaolinite.
The partial densities of states (PDOSs) of Pr and the (001) surface before and after adsorption are shown in Figure 4. The 4f orbital of Pr is sharp and thin near the Fermi level before the adsorption, indicating that the 4f orbital is very active because 4f belongs to the valence electron orbital of Pr, which easily gains and loses electrons. However, the 5d and 6p orbitals are in the inner layer, and the activity is lower, hence, their PDOS is far from the Fermi level. Similarly, the 2p orbital of the (001) surface is also near the Fermi level before the adsorption. Notable changes occur after the adsorption, with the 4f orbital of the Pr atom shifting towards a higher energy, which is far from the Fermi level, while the energy of the (001) surface decreases to maintain the stability of the system. Additionally, the energy scope of the Pr 4f orbital expands to 11.2–14 eV, implying that the non-locality of the Pr 4f orbital is increased, which may be beneficial for bonding. According to the Mulliken charge of Pr (Table 4), following the adsorption, the Pr 4f orbital loses 0.15 e, while the 6s and 5d orbitals lose 0.02 e and 0.06 e, respectively. The total charge of Pr increases obviously from 2.28 to 2.52. A large number of electron transfers lead to the peak shape of the density of states of the Pr atoms to change significantly.

3.3. [Mg(H2O)6]2+ Adsorbed on the (001) Surface of Mt

A previous work [26] revealed that the most stable structure of hydrated Mg2+ is [Mg(H2O)6]2+, which was used as the initial adsorption structure in this article. Table 5 shows the equilibrium geometry of [Mg(H2O)6]2+ after the adsorption on the montmorillonite (001) surface, which is similar to the adsorption of hydrated Pr ions. As shown in Figure 5, two hydrogen bonds form between the hydrogen of H2O ligands and oxygen of the surface, with lengths of 2.016 and 2.375 Å, respectively, so that the hydrated Mg complex can be stably adsorbed on the (001) surface of Mt. According to the structural parameters and adsorption energy of [Mg(H2O)6]2+ before and after the adsorption (Table 5), the bond length of Pr-Ow increases from 2.26 Å before the adsorption to 2.40 Å due to the steric hindrance. Because of the lower number of H2O ligands in [Mg(H2O)6]2+ than in [Pr(H2O)n]3+, no H2O ligands escape from the first hydration layer of Mg2+ ions after the adsorption.
The calculation results show that the adsorption energy of [Mg(H2O)6]2+ on the (001) surface of Mt is −206 kJ/mol, which is much lower than that of [Pr(H2O)10]3+ (−1182 kJ/mol), indicating that it is difficult to thermodynamically desorb Pr3+ ions from Mt by Mg2+ ions; thus, dynamic approaches should attract more attention, such as enhancing the concentration of the leaching agent, which may be beneficial for the desorption. Additionally, the calculation results also show that Pr3+ is prone to a reverse adsorption if the concentration of Mg2+ ions is the same as that of Pr3+ ions. Hence, it is necessary to maintain the concentration of the leaching agent in a high range to prevent the reverse adsorption of rare earth ions.
Figure 6 shows the PDOSs of Mg2+ and the (001) surface before and after adsorption. The 2s and 2p orbitals of the Mg2+ ion were far from the Fermi level before adsorption because the ionic radius of Mg2+ is short and it has difficulty gaining and losing electrons. After the adsorption, the 2p orbital of the Mg2+ ion obviously moved in the lower energy direction away from the Fermi level. Additionally, the 2s and 2p orbitals of the (00) surface also moved away from the Fermi level in the lower energy direction, and the intensity was obviously weakened, indicating that the energy of the system decreased, and it became more stable after the adsorption. The Mulliken charges of Mg2+ are listed in Table 6. The 2s and 2p orbitals of the Mg2+ ion lost 0.04 e and 0.06 e, respectively, and the charge increased from 1.68 to 1.78.

3.4. Hydrated NH4+ Adsorbed on the (001) Surface of Mt

Different from Mg2+, NH4+ interacts with H2O ligands through hydrogen bonds; in simulations, the coordination number of NH4+ was found to be approximately five [27,28], whereas the experiments indicated that the coordination number is larger [29]. Hence, the coordination number of NH4+ was chosen to be six, which is beneficial for a comparison with Mg2+. The equilibrium geometry of hydrated NH4+ after the adsorption is shown in Figure 7. Although there were no coordination bonds between the NH4+ ion and H2O, several hydrogen bonds formed. The hydrogen bonds were very short, with lengths ranging from 1.207 Å to 1.943 Å, indicating a strong interaction between the NH4+ ion and H2O ligands. In addition, three hydrogen bonds formed between the H2O ligands and surface O, with an adsorption energy of −188 kJ/mol. This adsorption energy is more positive than that of [Mg(H2O)6]2+, indicating that the adsorption stability of [Mg(H2O)6]2+ is higher than that of the hydrated NH4+ ion, that is, the desorption efficiency of [Mg(H2O)6]2+ for RE ions is stronger than that of the hydrated NH4+ ion. Similar to [Mg(H2O)6]2+, the adsorption energy of the hydrated NH4+ ion is much lower than that of [Pr(H2O)10]3+ (−1182 kJ/mol). Hence, it is difficult to thermodynamically desorb Pr3+ ions from Mt by NH4+ ions.
The PDOSs of NH4+ and the (001) surface before and after the adsorption are shown in Figure 8. Since there were no coordination bonds between the NH4+ ion and H2O, the PDOS of the NH4+ ion showed almost no change after the adsorption. In contrast, the PDOS of the surface shifted towards the lower energy direction, and its 2s and 2p orbitals were obviously far away from the Fermi level. The change in PDOS indicated that the system became more stable after the adsorption. The Mulliken charge of N is listed in Table 7. Similar to the slight change in the PDOS, the electron transfer of the N atom was very low, and its 2s and 2p orbitals only obtained 0.01 e and 0.05 e, respectively. The N atom obtained 0.06 e, and the charge decreased from −0.79 to −0.81 after the adsorption.

3.5. Desorption Efficiencies of Mg2+ and NH4+

A series of desorption experiments were carried out to verify the DFT calculation result that the desorption efficiencies of Mg2+ are higher than those of NH4+. Raw Mt was mixed with a chloride praseodymium solution to obtain Pr–Mt, which was used in the desorption process. Figure 9 shows the recovery of Pr3+ from Mt as a function of the NH4+/Mg2+ concentration. The recoveries of Pr3+ were 55% and 80%, while the concentrations of Mg2+ were 15 mmol/L and 38 mmol/L, respectively, which were much higher than those of NH4+ (23% and 65%), whereas the desorption efficiencies of NH4+ and Mg2+ were equal under high concentrations. The concentrations of NH4+, Mg2+ and Pr3+ in lixivium are listed in Table 8. The concentration of Pr3+ was only 1.64 mmol/L (with a recovery of 55%), while the concentration of Mg2+ was 15 mmol/L, which was nearly ten times larger than that of Pr3+. Additionally, the concentration of NH4+ was forty times larger than that of Pr3+, while the recovery was 95%. These results imply that the desorption efficiency of Mg2+ is higher than that of NH4+ and that the concentration of NH4+/Mg2+ should be much higher than that of Pr3+ in the desorption process, which is consistent with the DFT calculation result that the adsorption stability of hydrated magnesium ions is higher than that of hydrated NH4+, which is lower than that of Pr3+.

4. Conclusions

The first-principles method based on the plane-wave pseudopotential DFT was applied to investigate the adsorption of hydrated Pr3+, Mg2+ and NH4+ ions onto Mt. The DFT calculations showed that Pr3+ ions tend to coordinate with ten H2O molecules with a binding energy of −2680 kJ/mol in aqueous systems. [Pr(H2O)10]3+ was adsorbed on the Mt (001) surface through four hydrogen bonds, with a negative adsorption energy of −1182 kJ/mol. One H2O ligand escaped from the first hydration layer of the Pr ion after the adsorption. Similarly, [Mg(H2O)6]2+ and the hydrated NH4+ ion were also adsorbed onto the Mt (001) surface through hydrogen bonds, and the adsorption energies were −206 and −188 kJ/mol, respectively, implying that the adsorption stability of [Mg(H2O)6]2+ was slightly higher than that of the hydrated NH4+ ion, while both were lower than that of [Pr(H2O)10]3+ (−1182 kJ/mol), indicating that the desorption of Pr3+ from Mt by Mg2+ and NH4+ ions was thermodynamically difficult and that much higher concentrations of Mg2+ and NH4+ ions were necessary. These DFT conclusions have also been confirmed by desorption experiments: the desorption efficiency of Mg2+ is higher than that of NH4+, and the concentration of NH4+/Mg2+ should be much higher than that of Pr3+ in the desorption process.
However, the theoretical investigation of RE ions desorbed from the surface of clays would be improved in the future research work, while a molecule dynamic simulation, which is a supplement for the DFT, would be an important method.

Author Contributions

Z.Z.: writing—original draft and visualization; L.H.: conceptualization ideas; X.L.: investigation; J.X.: investigation; X.Z.: investigation; R.S.: methodology and writing—review and editing; B.X.: investigation; J.O.: investigation; W.H.: funding acquisition and methodology. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The National Natural Science Foundation of China (grant No. 41662004 and 41362003). The Personnel Training Program of Suqu Zhi Guang in Ganzhou City Jiangxi Province.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the editors and reviewers for their excellent support. The authors are also grateful to the Jiangxi University of Science and Technology.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Brown, B.; Ma, B.M.; Chen, Z.M. Developments in the processing and properties of NdFeb-type permanent magnets. J. Magn. Magn. Mater. 2002, 248, 432–440. [Google Scholar] [CrossRef]
  2. Zeng, S.H.; Du, D.P.; Bai, F.H.; Su, H.Q. Bridging complexes of rare earth and cobalt cluster as catalyst precursors for Fischer-Tropsch synthesis. J. Rare Earth 2011, 29, 349–353. [Google Scholar] [CrossRef]
  3. Chen, H.; Wang, Y.H. Photoluminescence and cathodoluminescence properties of novel rare-earth free narrow-band bright green-emitting ZnB2O4:Mn2+ phosphor for LEDs and FEDs. Chem. Eng. J. 2019, 361, 314–321. [Google Scholar] [CrossRef]
  4. Chi, R.A.; Tian, J.; Li, Z.J.; Cui, P.; Wu, Y.X.; Li, S.R.; Wang, C.W.; Zhou, Z.A. Existing State and Partitioning of Rare Earth on Weathered Ores. J. Rare Earth 2005, 6, 756–759. [Google Scholar]
  5. Huang, L.J.H.; Liu, H.L.; Huang, W.F.; Zhao, B.X.; Shen, Z.F.; Bao, Y.Q. Recovery of lanthanum cations by functionalized magnetic multi-walled carbon nanotube bundles. RSC Adv. 2021, 11, 4751–4759. [Google Scholar] [CrossRef]
  6. Chi, R.A.; Tian, J. Review of Weathered Crust Rare Earth Ore. J. Chin. Rare Earth Soc. 2007, 25, 641–650. [Google Scholar]
  7. Xiao, Y.F.; Chen, Y.Y.; Feng, Z.Y.; Huang, X.W.; Huang, L.J.H.; Long, Z.Q.; Cui, D.L. Leaching characteristics of ion-adsorption type rare earths ore with magnesium sulfate. Nonferr. Metal. Soc. 2015, 25, 3784–3790. [Google Scholar] [CrossRef]
  8. Xiao, Y.F.; Feng, Z.Y.; Huang, X.W.; Huang, L.J.H.; Chen, Y.Y.; Wang, L.S.; Long, Z.Q. Recovery of rare earths from weathered crust elution-deposited rare earth ore without ammonia-nitrogen pollution: I. leaching with magnesium sulfate. Hydrometallurgy 2015, 153, 58–65. [Google Scholar]
  9. Zhou, Q.; Shen, W.; Zhu, J.X.; Zhu, R.L.; He, H.P.; Zhou, J.H.; Yuan, P. Structure and dynamic properties of water saturated CTMA-montmorillonite: Molecular dynamics simulations. Appl. Surf. Sci. 2014, 97–98, 62–71. [Google Scholar] [CrossRef]
  10. Qiu, S.; Yan, H.S.; Hong, B.G.; Long, Q.B.; Xiao, J.; Li, F.J.; Tong, L.C.; Zhou, X.W.; Qiu, T.S. Desorption of REEs from halloysite and illite: A link to the exploitation of ion-adsorption RE ore based on clay species. Minerals 2022, 12, 1003. [Google Scholar] [CrossRef]
  11. Alshameri, A.; He, H.; Xin, C.; Zhu, J.; Wei, X.; Zhu, R.; Wang, H. Understanding the role of natural clay minerals as effective adsorbents and alternative source of rare earth elements: Adsorption operative parameters. Hydrometallurgy 2019, 185, 149–161. [Google Scholar] [CrossRef]
  12. Zhou, F.; Huang, S.H.; Liu, X.; Feng, J.; Liu, Q.; Wang, Z.W.; Li, C.C.; Xu, Y.L. Adsorption kinetics and thermodynamics of rare earth on Mt modified by sulfuric acid. Colloid Surface A 2021, 627, 127063. [Google Scholar]
  13. Karaca, H.; Altintig, E.; Turker, D.; Teker, M. An evaluation of coal fly ash as an adsorbent for the removal of methylene blue from aqueous solutions: Kinetic and thermodynamic studies. J. Disper. Sci. Technol. 2018, 39, 1800–1807. [Google Scholar] [CrossRef]
  14. Altintig, E.; Balta, S.; Balta, M.; Aydemir, Z. Methylene blue removal with ZnO coated montmorillonite: Thermodynamic, kinetic, isotherm and artificial intelligence studies. Int. J. Phytoremediat. 2022, 24, 867–880. [Google Scholar] [CrossRef]
  15. Geysermans, P.; Noguera, C. Advances in atomistic simulations of mineral surfaces. J. Mater. Chem. 2009, 19, 7807–7821. [Google Scholar] [CrossRef]
  16. Chen, J.; Min, F.F.; Liu, L.Y.; Jia, F.F. Adsorption of methylamine cations on kaolinite basal surfaces: A DFT study. Physicochem. Probl. Miner. Process. 2020, 56, 338–349. [Google Scholar] [CrossRef]
  17. Qiu, S.; Wu, H.; Yan, H.S.; Li, X.B.; Zhou, X.W.; Qiu, T.S. Theoretical investigation of hydrated [Lu(OH)2]+ adsorption on kaolinite(0 0 1) surface with DFT calculations. Appl. Surf. Sci. 2021, 565, 150473. [Google Scholar] [CrossRef]
  18. Peng, C.L.; Zhong, Y.H.; Wang, G.S.; Min, F.F.; Qin, L. Atomic-level insights into the adsorption of rare earth Y(OH)3-nn+ (n=1-3) ions on kaolinite surface. Appl. Surf. Sci. 2019, 469, 357–367. [Google Scholar] [CrossRef]
  19. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  20. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A113. [Google Scholar] [CrossRef] [Green Version]
  21. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallogr. 2005, 220, 567. [Google Scholar] [CrossRef] [Green Version]
  22. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Vanderbilt, D. Soft self-consistent pseudo-potentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892. [Google Scholar] [CrossRef] [PubMed]
  24. Qiu, T.S.; Qiu, S.; Wu, H.; Yan, H.S.; Li, X.B.; Zhou, X.W. Adsorption of hydrated [Y(OH)2]+ on kaolinite (001) surface: Insight from DFT simulation. Powder Technol. 2021, 387, 80–87. [Google Scholar] [CrossRef]
  25. Qiu, S.; Yan, H.S.; Qiu, X.H.; Wu, H.; Zhou, X.W.; Wu, H.Q.; Li, X.B.; Qiu, T.S. Adsorption of La on kaolinite (001) surface in aqueous system: A combined simulation with an experimental verification. J. Mol. Liq. 2022, 347, 117956. [Google Scholar] [CrossRef]
  26. Zhang, Z.J.; Zhou, Q.; Yuan, Z.T.; Zhao, L.; Dong, J.D. Adsorption of Mg2+ and K+ on the kaolinite (001) surface in aqueous system: A combined DFT and AIMD study with an experimental verification. Appl. Surf. Sci. 2021, 538, 148158. [Google Scholar] [CrossRef]
  27. Bruge, F.; Bernasconi, M.; Parrinello, M. Ab initio simulation of rotational dynamics of solvated ammonium ion in water. J. Am. Chem. Soc. 1999, 121, 10883. [Google Scholar] [CrossRef]
  28. Ekimova, M.; Quevedo, W.; Szyc, Ł.; Iannuzzi, M.; Wernet, P.; Odelius, M.; Nibbering, E.T.J. Aqueous solvation of ammonia and ammonium: Probing hydrogen bond motifs with FT-IR and soft X-ray spectroscopy. J. Am. Chem. Soc. 2017, 139, 12773. [Google Scholar] [CrossRef]
  29. Pálinkás, G.; Radnai, T.; Szász, G.I.; Heinzinger, K. The structure of an aqueous ammonium chloride solution. J. Chem. Phys. 1981, 74, 3522. [Google Scholar] [CrossRef]
Figure 1. The equilibrium geometries of [Pr(H2O)6−10]3+ from periodic DFT calculations.
Figure 1. The equilibrium geometries of [Pr(H2O)6−10]3+ from periodic DFT calculations.
Minerals 12 01454 g001
Figure 2. The difference charge density of [Pr(H2O)10]3+.
Figure 2. The difference charge density of [Pr(H2O)10]3+.
Minerals 12 01454 g002
Figure 3. Equilibrium geometry of [Pr(H2O)10]3+ adsorbed on the (001) surface of Mt.
Figure 3. Equilibrium geometry of [Pr(H2O)10]3+ adsorbed on the (001) surface of Mt.
Minerals 12 01454 g003
Figure 4. PDOS of Pr and (001) surface of Mt before and after adsorption.
Figure 4. PDOS of Pr and (001) surface of Mt before and after adsorption.
Minerals 12 01454 g004
Figure 5. Equilibrium geometry of [Mg(H2O)6]2+ adsorbed on the (001) surface of Mt.
Figure 5. Equilibrium geometry of [Mg(H2O)6]2+ adsorbed on the (001) surface of Mt.
Minerals 12 01454 g005
Figure 6. PDOS of Mg and (001) surface of Mt before and after adsorption.
Figure 6. PDOS of Mg and (001) surface of Mt before and after adsorption.
Minerals 12 01454 g006
Figure 7. Equilibrium geometry of hydrated NH4+ adsorbed on the (001) surface of Mt.
Figure 7. Equilibrium geometry of hydrated NH4+ adsorbed on the (001) surface of Mt.
Minerals 12 01454 g007
Figure 8. PDOS of NH4 and (001) surface of Mt before and after adsorption.
Figure 8. PDOS of NH4 and (001) surface of Mt before and after adsorption.
Minerals 12 01454 g008
Figure 9. Influence of NH4+/Mg2+ concentration on the recovery of Pr3+ from Mt.
Figure 9. Influence of NH4+/Mg2+ concentration on the recovery of Pr3+ from Mt.
Minerals 12 01454 g009
Table 1. The equilibrium geometries parameters and binding energy of [Pr(H2O)6−10]3+.
Table 1. The equilibrium geometries parameters and binding energy of [Pr(H2O)6−10]3+.
nR(Pr-Ow a)minR(Pr-Ow)maxR(Pr-Ow)avg bEbinding/kJ·mol−1Pr Charge/e
62.432.462.44−21872.08
72.442.532.49−23942.05
82.402.662.53−25112.05
92.492.792.58−24862.06
102.453.032.68−26802.04
a Distance of Pr to oxygen centre of H2O ligands. b Average distance of Pr to oxygen centre of H2O ligands.
Table 2. The Mulliken charge population of Pr and O in [Pr(H2O)10]3+.
Table 2. The Mulliken charge population of Pr and O in [Pr(H2O)10]3+.
AtomNSPDFTotalCharge/e
O11.865.09006.95−0.95
O21.865.07006.93−0.93
O31.845.08006.92−0.92
O41.855.08006.93−0.93
O51.855.08006.93−0.93
O61.855.08006.93−0.93
O71.865.09006.95−0.95
O81.865.08006.94−0.94
O91.865.08006.94−0.94
O101.855.14006.99−0.99
Pr12.135.710.832.1910.872.13
Table 3. Adsorption energy and structural parameters of [Pr(H2O)10]3+ adsorbed on the (001) surface of Mt.
Table 3. Adsorption energy and structural parameters of [Pr(H2O)10]3+ adsorbed on the (001) surface of Mt.
StructureNPr-O/ÅPr-OmEads/kJ·mol−1
Before102.67, 3.83, 2.67, 2.67, 2.67
2.67, 2.67, 2.67, 2.67, 2.67
2.78−1182
After92.58, 2.55, 2.59, 2.51
2.69, 2.55, 2.67, 2.62, 2.58
2.59
Table 4. Mulliken charges of Pr adsorbed on the (001) surface of Mt.
Table 4. Mulliken charges of Pr adsorbed on the (001) surface of Mt.
ElementSPDFTotalCharge/e
Pr before2.135.640.782.1610.712.28
Pr after2.115.640.722.0110.482.52
Table 5. Adsorption energy and structural parameters of [Mg(H2O)6]2+ adsorbed on the (001) surface of Mt.
Table 5. Adsorption energy and structural parameters of [Mg(H2O)6]2+ adsorbed on the (001) surface of Mt.
StructureNMg-O/ÅMg-OmEads/kJ·mol−1
Before62.26, 2.26, 2.26
2.26, 2.26, 2.26
2.26−206
After62.31, 2.41, 2.42
2.47, 2.46, 2.33
2.40
Table 6. Mulliken charges of Mg adsorbed on the (001) surface of Mt.
Table 6. Mulliken charges of Mg adsorbed on the (001) surface of Mt.
ElementSPDFTotalCharge/e
Mg before0.335.99006.321.68
Mg after0.295.93006.221.78
Table 7. Mulliken charges of NH4 adsorbed on the (001) surface of Mt.
Table 7. Mulliken charges of NH4 adsorbed on the (001) surface of Mt.
ElementSPDFTotalCharge/e
N before1.634.15005.79−0.79
N after1.644.17005.81−0.81
Table 8. Concentration of NH4+, Mg2+ and Pr3+ in desorption solution.
Table 8. Concentration of NH4+, Mg2+ and Pr3+ in desorption solution.
NH4+/Mg2+ (mmol/L)0153875110
Pr3+ desorbed by NH4+ (mmol/L)0.030.691.952.742.85
Pr3+ desorbed by Mg2+ (mmol/L)0.031.642.372.682.73
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zou, Z.; Huang, L.; Li, X.; Xu, J.; Zeng, X.; Shu, R.; Xiao, B.; Ou, J.; Huang, W. Adsorption of Hydrated Pr3+ and NH4+/Mg2+ Ions onto the (001) Surface of Montmorillonite: A DFT Analysis with Experimental Verification. Minerals 2022, 12, 1454. https://doi.org/10.3390/min12111454

AMA Style

Zou Z, Huang L, Li X, Xu J, Zeng X, Shu R, Xiao B, Ou J, Huang W. Adsorption of Hydrated Pr3+ and NH4+/Mg2+ Ions onto the (001) Surface of Montmorillonite: A DFT Analysis with Experimental Verification. Minerals. 2022; 12(11):1454. https://doi.org/10.3390/min12111454

Chicago/Turabian Style

Zou, Zhiqiang, Lijinhong Huang, Xindong Li, Jie Xu, Xiangrong Zeng, Ronghua Shu, Bin Xiao, Jiacai Ou, and Wanfu Huang. 2022. "Adsorption of Hydrated Pr3+ and NH4+/Mg2+ Ions onto the (001) Surface of Montmorillonite: A DFT Analysis with Experimental Verification" Minerals 12, no. 11: 1454. https://doi.org/10.3390/min12111454

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