Next Article in Journal
Dithiocarbamates: Properties, Methodological Approaches and Challenges to Their Control
Next Article in Special Issue
Study on Dynamic Column Behavior and Complexation Mechanism of DBS-Modified Crown Ether-Based Silica to 90Sr
Previous Article in Journal / Special Issue
Arsenic Removal Using Unconventional Material with Iron Content: Batch Adsorption and Column Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Density Functional Theory Studies of Aluminosilicate-Based Ceramic Solidified Products for Sr Immobilization

School of Nuclear Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Toxics 2023, 11(10), 850; https://doi.org/10.3390/toxics11100850
Submission received: 7 September 2023 / Revised: 28 September 2023 / Accepted: 8 October 2023 / Published: 11 October 2023
(This article belongs to the Special Issue Novel Adsorbents and Adsorption Methods for Pollutants Removal Ⅱ)

Abstract

:
Strontium is a common radionuclide in radioactive waste, and its release into the environment can cause enormous damage to the ecosystem environment. In this study, the natural mineral allophane was selected as the substrate to prepare solidified ceramic products by cold pressing/sintering to solve the problem of the final disposal of radioactive strontium. Ceramic solidified products with various crystal structures were successfully prepared, and the microscopic morphology and energy-dispersive spectroscopy images of the samples showed a uniform distribution of Sr in the solidified products. Sr2Al2SiO7 and SrAl2Si2O8, which can stably solidify strontium, were formed in the solidified products, and the structural characteristics and stability of the above-mentioned substances were analyzed from the perspective of quantum chemical calculations using density functional theory. The calculation results showed that the overall deformation resistance of Sr2Al2SiO7 was higher than that of SrAl2Si2O8. Considering the isomorphic substitution effect of CaO impurities, we inferred that a mixed-crystalline structure of Ca2−xSrxAl2SiO7 may be present in the solidified products.

1. Introduction

Nuclear energy is a crucial high-quality energy source, and its development provides a new impetus for the progress of human society. However, nuclear energy production inevitably generates hazardous radioactive waste, posing substantial environmental and biological risks. Thus, it must be treated safely and effectively [1]. The radionuclide 90Sr, characterized by radiotoxicity, high heat generation, and a long half-life of 28.8 years, is a β-emitter generated by uranium and plutonium fission reactions in nuclear reactors [2,3,4]. As an alkaline-earth cation with chemical properties similar to calcium, 90Sr easily accumulates in human bones [5,6]. 90Y, a daughter nuclide of 90Sr, produces high-energy β particles that can damage the bone marrow. Therefore, considering environmental protection and human health, the safe disposal of 90Sr from radioactive waste is valuable to a great extent [7].
Stabilization and solidification are considered the most promising technologies for radionuclide management [8,9]. In addition to traditional glass and cement solidification, 90Sr solidification primarily focuses on geopolymer and ceramic solidifications. Geopolymers are amorphous inorganic binder materials that are usually composed of SiO4 and AlO4 tetrahedra and can be prepared from solid wastes, such as fly ash, metakaolin, and slags via alkali activation at ambient temperature [10,11,12]. These materials can be considered precursors to zeolites with different affinities toward various ions. Notably, the Si/Al ratio is the controlling factor affecting the adsorption ability of geopolymers [13,14,15,16]. In the field of radioactive waste immobilization, geopolymers that outperform Portland cement can be prepared by changing the precursor materials and activators [15]. The absorbed waste elements can be further converted into stable ceramic-phase components, thereby increasing the difficulty of ion leaching by strengthening chemical bonds. The ceramic solidification approach has received extensive research attention due to its excellent chemical stability, radiation stability, and leaching resistance [17]. At the atomic scale, radioactive elements can enter the lattice structure of ceramic matrix materials, thereby forming a safer immobilization barrier than in glass [18]. At present, various ceramic materials, such as phosphates, aluminosilicates, titanates, zirconates, and vanadates [7,17,18,19,20], have been extensively studied as matrices for immobilizing radioactive wastes through methods, such as cold pressing/sintering, microwave sintering, and spark plasma sintering [21,22,23,24].
As inorganic polymers, aluminosilicates have layered and skeletal crystal structures that endow them with high specific surface areas. These materials can achieve the selective adsorption of multiple ions through a molecular sieve mechanism; that is, certain ion sizes can enter the cavities, pores, and channels formed by the aluminosilicate framework. Allophane is a short-range order aluminosilicate and an affordable material for obtaining solid-state matrices based on aluminosilicate ceramics. It is primarily found in volcanic ash soil, and the basic structure of its skeleton comprises Al(OH)3 and SiO4 [25]. The special frame structure of allophane confers good radiation resistance stability and the ability to capture gases [26]. Owing to these advantages, it is one of the most promising materials for metal-ion removal and immobilization. Our group previously treated a mixture of allophane and cesium secondary waste at high temperatures, resulting in the breakdown of the allophane framework and the formation of stable crystalline phases that encapsulate cesium [27,28]. Excellent leaching resistance and mechanical properties are the main characteristics of the prepared solidified ceramic bodies. However, systematic studies on the application of allophane for the immobilization of highly radioactive 90Sr have rarely been reported.
The use of density functional theory (DFT) to study the binding behavior of metal ions in lattice-defect structures and lattices can enable a deeper understanding of the solidification structures and bonding mechanisms at the molecular level. The lattice parameters, structural stability, mechanical properties, thermo-physical properties, and electronic structure of different materials have been extensively studied through DFT experiments to evaluate their comprehensive properties [29,30,31,32,33,34]. However, few studies have been conducted on the immobilization mechanism of Sr in high-temperature ceramics. Considering the stability and advantages of capturing Sr in ceramic matrices, it is important to explore the immobilization mechanism of solidified ceramic products at the molecular level.
In this study, the natural mineral allophane was used to synthesize solidified Sr ceramic products through cold pressing/sintering. The performance and mechanism of the allophane on Sr were explored, and the structural characteristics and stability of the sintered products were analyzed from the perspective of quantum chemical calculations using DFT.

2. Materials and Methods

2.1. Materials

Nonradioactive 87Sr was used in all samples instead of radioactive 90Sr. Analytical-reagent-grade strontium nitrate (Sr(NO3)2) was purchased from Sinopharm Chemical Reagent Co. Ltd., China. Allophane (1–2SiO2·Al2O3·5–6H2O), with an average particle size of 5.9 µm, was obtained from Hattori Company, Ltd., Japan. The smaller particle size ensured a larger specific surface area, which was conducive to improving the adsorption capacity of metal ions. The surface morphology and main components of the allophane used as matrices are shown in Figure 1.

2.2. Preparation of Solidified Products

The solidified product was prepared as follows: Allophane and Sr(NO3)2 were uniformly mixed in various mass ratios. Then, the uniform mixture was ground into a fine powder and molded into a disk via the cold pressing method using hydraulic pressure tablets under 4 kN with a holding time of 6 min. The resulting 10 mm diameter molded discs were placed in a muffle furnace for sintering at specific temperatures and a heating rate of 10 °C/min for 1 h.

2.3. Characterization

Scanning electron microscopy and energy-dispersive spectrometry (SEM and EDS, Mira 3, TESCAN ORSAY HOLDING a.s., Brno, Czech Republic) were used to measure the surface morphology and elemental distribution of the solidified products. The mass fractions of each element in the samples before and after sintering were determined using an X-ray fluorescence spectrometer (XRF-1800, Shimadzu Corporation, Kyoto, Japan). The crystal structures of the samples were examined using powder X-ray diffraction (XRD) spectroscopy (D8 ADVANCED DAVINCI, Bruker Corporation, Karlsruhe, Germany).

2.4. DFT Calculation

DFT is a classical quantum mechanical modeling method for calculating and predicting the electronic properties of crystal structures with high precision. It has been used extensively in computational research on ceramic materials. This method calculates the energy of the system as a function of electron density without solving the complex many-body Schrodinger equation, thereby significantly simplifying the calculation. In this study, DFT calculations were performed to analyze the crystal structure and mechanical properties of the solidified ceramic products. The gain and loss of electrons and chemical bond types of the Sr atoms were also studied to analyze the chemical stability of the solidified products. All calculations in this study were carried out based on the DFT method using the Vienna ab initio simulation package (VASP). Electron exchange and related energy were computed using the projector augmented wave (PAW) method with the Perdew–Burke–Ernzerhof generalized gradient approximation (PBE-GGA) function [35,36]. The Brillouin zone sampling was performed using a Gamma grid. The cut-off energy was set to 400 eV, as confirmed by the convergence test, and the force convergence of the structural optimization was set to be lower than 0.02 eV/Å.

3. Results and Discussion

3.1. Characterization and Self-Sintering Behavior of Allophane

Self-sintering experiments were performed on the allophanes at different temperatures. The XRD patterns of allophanes with different sintering temperatures are shown in Figure 2. No obvious diffraction peaks were observed in the XRD patterns of fresh allophane, indicating that allophane was an amorphous substance. Two weak diffraction peaks were observed at 26.6° and 28°, which may be attributed to the spherical shell structure of allophane, comprising silica and alumina, characterized by short-range ordering. When the sintering temperature was lower than 1000 °C, wide and diffuse peaks were observed in the XRD pattern, suggesting no mineral-phase transformation or crystal-phase substance formation under these experimental conditions. After sintering at 1200 °C, the main crystal phases in the self-sintering products of allophane were primarily transformed into two phases, namely, mullite (Al6Si2O13) and cristobalite (SiO2). The increase in the sintering temperature may have caused the Si–O–Si and Si–O–Al bonds to break and the allophane framework to collapse. The decomposition products Al2O3 and SiO2 with high chemical reaction activity also recrystallize at high temperatures to yield a mullite crystal phase [26,37]. The content of the cristobalite crystal phase was relatively low, and the cristobalite phase may be attributed to the high-temperature calcination of excessive SiO2 [38]. Moreover, the volume reduction rate of the sintered solidified body increased from 20.6% to 52.2% with increased sintering temperature from 700 °C to 1000 °C. This phenomenon has benefited waste minimization and improved economic efficiency. Hence, the formation of stable crystalline phases and good volume reduction rates indicate that allophane has excellent self-sintering properties, making it a potential matrix material for the solidification of radionuclide ceramics.

3.2. Sr Immobilization Ratio and Solidification

Figure 3 shows the Sr immobilization ratios obtained via XRF for various mixtures of Sr(NO3)2–allophane sintered at 600–1200 °C, with a gradual increase in Sr content from 10wt% to 20 wt%, the immobilization ratio of Sr remained at approximately 100% at all sintering temperatures. This finding indicates that allophane can be effectively used for the immobilization treatment of Sr.
XRD was used to study the crystal phases of the sintered products and to understand the immobilization mechanism of Sr. Figure 4 shows the XRD patterns of various mixtures of Sr(NO3)2–allophane sintered at 1200 °C. The crystal structure of the solidified products changed after sintering. When the Sr content was 1 wt%, no obvious crystal phase of Sr was detected in the solidified product; it had primarily mullite and cristobalite crystal phases. Thus, the self-sintering behavior of allophane dominated the immobilization when the Sr content was low. When the Sr content ranged from 5 wt% to 30 wt%, in addition to the mullite phase, strontium feldspar (SrAl2Si2O8) and strontium aluminum silicate oxide (Sr2Al2SiO7) crystal phases that can stably solidify Sr also formed in the solidified product. With increasing Sr content in the solidified product, the proportion of the mullite phase gradually decreased and that of Sr2Al2SiO7 gradually increased; however, the proportion of SrAl2Si2O8 initially increased and then decreased. This finding may be related to the value of Si/Al in the solidified product. The deflection of the diffraction peaks can also be observed in the solidified products with different compositions. According to the Scherrer formula, the decreased interplanar spacing led to an increased diffraction angle, indicating that the crystal-phase structure of the solidification products changed to a certain extent.
The surface morphologies and main elemental compositions of the solidified products after sintering at 1200 °C are shown in Figure 5. The surface of the solidified product appeared to melt and transform into a ceramic phase, and a non-uniform porous surface was formed. Si, Al, and Sr were detected and uniformly distributed on the surface of the solidified product, indicating that well-crystallized phases containing Sr were produced, which is consistent with the XRD results.

3.3. Calculation of Elastic Properties of Solidified Product

Sr was primarily immobilized in the solidified product in the form of SrAl2Si2O8 and Sr2Al2SiO7. Therefore, the Young’s modulus, shear modulus, and other parameters of the two phases were simulated and calculated using DFT to analyze the change in mechanical properties of the solidified product with increased Sr content. The crystal-structure parameters of SrAl2Si2O8 and Sr2Al2SiO7 after optimization are listed in Table 1, and the structures of SrAl2Si2O8 and Sr2Al2SiO7 are shown in Figure 6. The calculated elastic-tensor parameters of the two crystals after optimizing the cell structure are listed in Table 2.
In general, the hardness of a material increases with an increase in Young’s modulus (E). Conversely, a smaller shear modulus (G) corresponds to more easily formed dislocation slips, causing the material to become ductile [36]. The bulk modulus (B) of polycrystalline crystals was used to measure the relationship between the bulk strain and the average stress of the crystal. The bulk modulus decreased with an increase in the crystal volume. The bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio  ν  were evaluated using the Voigt–Reuss–Hill approximation (Equations (1)–(3)) [36].
B V = 1 9 C 11 + C 22 + C 33 + 2 9 C 12 + C 13 + C 23 B R = 1 S 11 + S 22 + S 33 + 2 S 12 + S 13 + S 23
G V = 1 15 C 11 + C 22 + C 33 C 12 C 13 C 23 + 1 5 C 44 + C 55 + C 66
G R = 15 4 S 11 + S 22 + S 33 4 S 12 + S 13 + S 23 + 3 S 44 + S 55 + S 66
B = B V + B R 2 , G = G V + G R 2 ,
E = 9 B G 3 B + G , ν = 3 B E 6 B
where Cij and Sij represent the elastic constant and elastic compliance, respectively.
The anisotropy indices of the bulk and shear moduli can be expressed using the following formula [39]:
A B = B V B R B V + B R , A G = G V G R G V + G R
The calculated bulk, shear, and Young’s moduli of the SrAl2Si2O8 and Sr2Al2SiO7 polycrystals are shown in Figure 7. The results showed that the overall anti-deformation ability of Sr2Al2SiO7 was better than that of SrAl2Si2O8. Moreover, a larger value of A corresponds to greater anisotropy of the system, whereas when A is equal to zero, it is completely isotropic [36]. The anisotropy percentages of the bulk and shear moduli of Sr2Al2SiO7 are close to 0, further indicating that this crystalline phase is isotropic. By contrast, SrAl2Si2O8 showed obvious anisotropy.
To observe the anisotropy property more clearly, the three-dimensional distributions of Young’s modulus and shear modulus were obtained through further calculation [40], and the results are displayed in Figure 8 and Figure 9, respectively. The closer the three-dimensional figure was to a spherical shape, the more isotropic the system. Obviously, SrAl2Si2O8 was more anisotropic than Sr2Al2SiO7 in terms of Young’s modulus and shear modulus.

3.4. Simulation and Calculation of Sr/Ca Mixed-Crystal Structures

Given that allophane contains some CaO impurities, considering the isomorphism substitution, a Sr-Ca mixed-crystal phase may have existed in the solidified product after sintering. According to the doping levels of different Ca atoms, the XRD patterns of Ca1−xSrxAl2Si2O8 and Ca2−xSrxAl2SiO7 after optimization are shown in Figure 10A,B. As shown in Figure 10A, for the Sr/Ca mixed-crystal structure of SrAl2Si2O8 (simulated), an obvious diffraction peak was observed at 2θ = 6.8° with a change in the Ca/Sr ratio. The intensity of the diffraction peak increased as the Ca/Sr approached one. The XRD pattern of the experimental test revealed no obvious diffraction peak at 2θ = 6.8°. Therefore, we speculated that no mixed-crystal formation of Ca1−xSrxAl2Si2O8 occurred.
As shown in Figure 10B, unlike SrAl2Si2O8, obvious diffraction peaks were observed at 2θ = 11.38°, 16.12°, and 17.38° in the XRD patterns of Sr2Al2SiO7 (experimental) and Ca2−xSrxAl2SiO7 (simulated). This finding suggested that a mixed-crystal structure of Ca2−xSrxAl2SiO7 formed in the sintered solidified product. The mechanical properties of the Ca2−xSrxAl2SiO7 mixed crystals that appeared in the experiment were calculated, and the elastic moduli are shown in Figure 11A. With increased Sr content, the bulk modulus, Young’s modulus, and shear modulus of the polycrystal showed a downward trend, indicating that adding Sr caused a deterioration in mechanical properties. According to the Pugh criterion, the ratio of B/G was generally 1.75 as the boundary ratio between ductility and brittleness. These results confirmed that the brittle crystal transformed into a ductile one with increased Sr content, as shown in Figure 11B.

3.5. Analysis of Electron Localization Function (ELF)

The degree of electron localization at different locations in a three-dimensional real space can be expressed by the ELF, with a value ranging from 0 to 1, which quantitatively characterizes the degree of electron localization. A closer ELF value to 1 corresponds to a stronger localization of electrons in this region and greater difficulty for electrons to escape. Issues, such as charge-transfer bonds, metallic bonds, and hydrogen bonds, have been extensively studied using ELF [41]. Figure 12 shows the three-dimensional ELF diagrams of SrAl2Si2O8 obtained via VESTA as well as the two-dimensional ELF diagrams in sections of 3.24 and 9.72 Å away from the setting plane in the direction of the (0 1 0) crystal plane. The ELF values near the Sr and O atoms were relatively high, indicating that the localization of electrons near the two atoms was relatively strong.
Moreover, the Sr atoms were contracted at the side that was neighboring the O atoms, and the same contraction of ELF was also observed at the O atom in Figure 12B,C [42]. Thus, Sr and O were most likely to combine with each other in the form of ionic bonds. Figure 13A shows the three-dimensional ELF diagrams of Sr2Al2SiO7, and the four Sr atoms in the crystal lattice are marked in Figure 13B–E. Similar to SrAl2Si2O8, the localization of electrons around the Sr and O atoms was relatively high. Therefore, we inferred that Sr and O also combined with each other in the form of ionic bonds in the Sr2Al2SiO7 crystal.

3.6. Comparison of Different Sr-Containing Ceramic Solidified Products

The selection of mineral substrates is currently the primary issue in the preparation and research of Sr-containing solidified products. In this study, three typical Sr-containing ceramic solidified products, Sr0.5Zr2(PO4)3, SrTiO3, and SrZrO3, were selected, and the bond energy and mechanical properties of Sr in different mineral phases were calculated using VASP and compared with those of SrAl2Si2O8 and Sr2Al2SiO7. Then, the chemical stabilities and mechanical properties of the solidified products were evaluated. The Sr bond energies in the different mineral phases are shown in Figure 14, and the elastic moduli of the different mineral phases are shown in Figure 15A. The Sr bond energies of SrAl2Si2O8 and Sr2Al2SiO7 were higher than those of the other three mineral phases, i.e., 11.6 and 12.1 eV, respectively. This order of magnitude is higher than that of Sr0.5Zr2(PO4)3 with apatite structures and SrZrO3 with zirconium-based perovskite structures. This finding indicates good chemical stability in the solidified product containing Sr produced in this study.
The bulk modulus B, shear modulus G, and Young’s modulus E of several Sr-containing mineral phases all exhibited the following trend: SrTiO3 > Sr0.5Zr2(PO4)3 > Sr2Al2SiO7 > SrAl2Si2O8 > SrZrO3. Among these mineral phases, SrTiO3 exhibited the best mechanical properties, whereas SrZrO3 exhibited the worst. The proportions of Sr2Al2SiO7, SrAl2Si2O8, and mullite in the solidified product changed continuously with increased Sr content. The formation of the mullite crystal phase can lead to enhanced hardness and modulus in the solidified product [28]. Therefore, a decrease in the mullite proportion may be the main reason for the decline in the mechanical properties of the solidified product. Therefore, to increase the mechanical properties of the solidified product, the content of allophane during the immobilization of Sr should be appropriately increased. The Pugh’s ratio in Figure 15B also shows that only SrTiO3 had partial brittleness, and the other ceramic solidified products had a certain ductility.

4. Conclusions

The immobilization performance of the natural mineral allophane on the heat-generated nuclide strontium was investigated through ceramic solidification. Stable crystals called Sr2Al2SiO7 and SrAl2Si2O8 were formed in the solidified products after sintering at 1200 °C. These crystalline materials enabled the effective immobilization of strontium, and the immobilization ratio reached 100%. DFT simulations further revealed the structural characteristics and stabilities of the sintered products. The increased strontium content caused a transition from brittle to ductile crystal. From the ELF diagrams, we inferred that Sr and O are highly likely to be bonded to each other through ionic bonds in the two crystals mentioned above. Moreover, the Sr bond energies of SrAl2Si2O8 and Sr2Al2SiO7 were 11.6 and 12.1 eV, respectively, higher than those of the other three common mineral phases. These results indicate that the natural mineral allophane is promising for the final disposal of strontium.

Author Contributions

Conceptualization, Y.W. (Yan Wu); methodology, Y.W. (Yan Wu) and H.W.; investigation, H.S. and Z.G.; writing—original draft preparation, H.S. and Y.W. (Yan Wu); writing—review and editing, Y.W. (Yuezhou Wei); validation, and formal analysis, S.Y. and J.Z.; supervision, Y.W. (Yan Wu) and Y.W. (Yuezhou Wei). 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 number 12175143).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from the corresponding author on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shiner, M.E.; Klein-BenDavid, O.; L’Hôpital, E.; Dauzères, A.; Neji, M.; Teutsch, N.; Peled, A.; Bar-Nes, G. Retention of strontium in high- & low-pH cementitious matrices—OPC vs. model systems. Cem. Concr. Res. 2022, 152, 106659. [Google Scholar] [CrossRef]
  2. Wang, Y.; Wen, Y.; Mao, C.; Sang, H.; Wu, Y.; Li, H.; Wei, Y. Development of chromatographic process for the dynamic separation of 90Sr from high level liquid waste through breakthrough curve simulation and thermal analysis. Sep. Purif. Technol. 2022, 282, 120103. [Google Scholar] [CrossRef]
  3. Wu, Y.; Kim, S.-Y.; Tozawa, D.; Ito, T.; Tada, T.; Hitomi, K.; Kuraoka, E.; Yamazaki, H.; Ishii, K. Equilibrium and kinetic studies of selective adsorption and separation for strontium using DtBuCH18C6 loaded resin. J. Nucl. Sci. Technol. 2012, 49, 320–327. [Google Scholar] [CrossRef]
  4. Schmidt, B.; Kegler, F.; Steinhauser, G.; Chyzhevskyi, I.; Dubchak, S.; Ivesic, C.; Koller-Peroutka, M.; Laarouchi, A.; Adlassnig, W. Uptake of Radionuclides by Bryophytes in the Chornobyl Exclusion Zone. Toxics 2023, 11, 218. [Google Scholar] [CrossRef] [PubMed]
  5. Froidevaux, P.; Pittet, P.-A.; Bühlmann, D.; Bochud, F.; Straub, M. Ion-imprinted resin for use in an automated solid phase extraction system for determining 90Sr in environmental and human samples. J. Radioanal. Nucl. Chem. 2021, 330, 797–804. [Google Scholar] [CrossRef]
  6. Pant, A.D.; Ruhela, R.; Tomar, B.S.; Anilkumar, S. Determination of 90Sr in environmental samples using solid phase extraction chromatography. J. Radioanal. Nucl. Chem. 2019, 322, 49–55. [Google Scholar] [CrossRef]
  7. Prekajski Đorđević, M.; Maletaškić, J.; Stanković, N.; Babić, B.; Yoshida, K.; Yano, T.; Matović, B. In-situ immobilization of Sr radioactive isotope using nanocrystalline hydroxyapatite. Ceram. Int. 2018, 44, 1771–1777. [Google Scholar] [CrossRef]
  8. Tian, Q.; Sasaki, K. Application of fly ash-based materials for stabilization/solidification of cesium and strontium. Environ. Sci. Pollut. Res. 2019, 26, 23542–23554. [Google Scholar] [CrossRef]
  9. Lin, S.-L.; Lai, J.S.; Chian, E.S.K. Modifications of sulfur polymer cement (SPC) stabilization and solidification (S/S) process. Waste Manag. 1995, 15, 441–447. [Google Scholar] [CrossRef]
  10. Kuenzel, C.; Cisneros, J.F.; Neville, T.P.; Vandeperre, L.J.; Simons, S.J.R.; Bensted, J.; Cheeseman, C.R. Encapsulation of Cs/Sr contaminated clinoptilolite in geopolymers produced from metakaolin. J. Nucl. Mater. 2015, 466, 94–99. [Google Scholar] [CrossRef]
  11. Tian, Q.; Sasaki, K. Application of fly ash-based geopolymer for removal of cesium, strontium and arsenate from aqueous solutions: Kinetic, equilibrium and mechanism analysis. Water Sci. Technol. 2019, 79, 2116–2125. [Google Scholar] [CrossRef] [PubMed]
  12. Xu, Z.; Jiang, Z.; Wu, D.; Peng, X.; Xu, Y.; Li, N.; Qi, Y.; Li, P. Immobilization of strontium-loaded zeolite A by metakaolin based-geopolymer. Ceram. Int. 2017, 43, 4434–4439. [Google Scholar] [CrossRef]
  13. Abdelrahman, E.A.; Abou El-Reash, Y.G.; Youssef, H.M.; Kotp, Y.H.; Hegazey, R.M. Utilization of rice husk and waste aluminum cans for the synthesis of some nanosized zeolite, zeolite/zeolite, and geopolymer/zeolite products for the efficient removal of Co(II), Cu(II), and Zn(II) ions from aqueous media. J. Hazard. Mater. 2021, 401, 123813. [Google Scholar] [CrossRef]
  14. Wang, L.; Geddes, D.A.; Walkley, B.; Provis, J.L.; Mechtcherine, V.; Tsang, D.C.W. The role of zinc in metakaolin-based geopolymers. Cem. Concr. Res. 2020, 136, 106194. [Google Scholar] [CrossRef]
  15. Tian, Q.; Pan, Y.; Bai, Y.; Sasaki, K. Immobilization of strontium in geopolymers activated by different concentrations of sodium silicate solutions. Environ. Sci. Pollut. Res. 2022, 29, 24298–24308. [Google Scholar] [CrossRef]
  16. Tian, Q.; Sasaki, K. Structural characterizations of fly ash-based geopolymer after adsorption of various metal ions. Environ. Technol. 2021, 42, 941–951. [Google Scholar] [CrossRef]
  17. Sang, H.; Mao, C.; Ming, F.; Xu, L.; Wei, Y.; Wu, Y. Selective separation and immobilization process of 137Cs from high-level liquid waste based on silicon-based heteropoly salt and natural minerals. Chem. Eng. J. 2022, 449, 137842. [Google Scholar] [CrossRef]
  18. Ma, J.; Fang, Z.; Yang, X.; Wang, B.; Luo, F.; Zhao, X.; Wang, X.; Yang, Y. Investigating hollandite–perovskite composite ceramics as a potential waste form for immobilization of radioactive cesium and strontium. J. Mater. Sci. 2021, 56, 9644–9654. [Google Scholar] [CrossRef]
  19. Keskar, M.; Patkare, G.; Shafeeq, M.; Phatak, R.A.; Kannan, S. Structural and thermal study of Sr(Th1−xUx)(PO4)2 compounds. J. Solid State Chem. 2021, 300, 122228. [Google Scholar] [CrossRef]
  20. Yang, J.; Shu, X.; Luo, F.; Wang, L.; Gu, Y.; Wu, J.; Lu, X. Solubility of Sr2+ in the Gd2Zr2O7 ceramics via appropriate occupation designs. J. Alloys Compd. 2019, 808, 151563. [Google Scholar] [CrossRef]
  21. Mao, X.; Li, Z.; Yi, F.; Wei, L.; Zhou, Y. Chemical durability of strontium-contaminated soil vitrified by microwave sintering. J. Radioanal. Nucl. Chem. 2023, 332, 435–445. [Google Scholar] [CrossRef]
  22. Papynov, E.K.; Shichalin, O.O.; Buravlev, I.Y.; Belov, A.A.; Portnyagin, A.S.; Fedorets, A.N.; Azarova, Y.A.; Tananaev, I.G.; Sergienko, V.I. Spark plasma sintering-reactive synthesis of SrWO4 ceramic matrices for 90Sr immobilization. Vacuum 2020, 180, 109628. [Google Scholar] [CrossRef]
  23. Papynov, E.K.; Belov, A.A.; Shichalin, O.O.; Buravlev, I.Y.; Azon, S.A.; Golub, A.V.; Gerasimenko, A.V.; Parotkina, Y.A.; Zavjalov, A.P.; Tananaev, I.G.; et al. SrAl2Si2O8 ceramic matrices for 90Sr immobilization obtained via spark plasma sintering-reactive synthesis. Nucl. Eng. Technol. 2021, 53, 2289–2294. [Google Scholar] [CrossRef]
  24. Papynov, E.K.; Belov, A.A.; Shichalin, O.O.; Buravlev, I.Y.; Azon, S.A.; Gridasova, E.A.; Parotkina, Y.A.; Yagofarov, V.Y.; Drankov, A.N.; Golub, A.V.; et al. Synthesis of Perovskite-Like SrTiO3 Ceramics for Radioactive Strontium Immobilization by Spark Plasma Sintering-Reactive Synthesis. Russ. J. Inorg. Chem. 2021, 66, 645–653. [Google Scholar] [CrossRef]
  25. Baldermann, A.; Grießbacher, A.C.; Baldermann, C.; Purgstaller, B.; Letofsky-Papst, I.; Kaufhold, S.; Dietzel, M. Removal of Barium, Cobalt, Strontium, and Zinc from Solution by Natural and Synthetic Allophane Adsorbents. Geosciences 2018, 8, 309. [Google Scholar] [CrossRef]
  26. Wu, Y.; Lee, C.-P.; Mimura, H.; Zhang, X.; Wei, Y. Stable solidification of silica-based ammonium molybdophosphate by allophane: Application to treatment of radioactive cesium in secondary solid wastes generated from fukushima. J. Hazard. Mater. 2018, 341, 46–54. [Google Scholar] [CrossRef]
  27. Cheng, Y.; Wang, Y.; Sang, H.; Wu, Y.; Wei, Y. Study on the immobilization of cesium absorbed by copper ferrocyanide using allophane through pressing/sintering method. J. Nucl. Mater. 2020, 532, 152008. [Google Scholar] [CrossRef]
  28. Xu, M.; Wu, Y.; Wei, Y. Stable solidification of silica-based ammonium molybdophosphate absorbing cesium using allophane: Mechenical property and leaching studies. J. Radioanal. Nucl. Chem. 2018, 316, 1313–1321. [Google Scholar] [CrossRef]
  29. Dinsdale, A.; Fang, C.; Que, Z.; Fan, Z. Understanding the Thermodynamics and Crystal Structure of Complex Fe Containing Intermetallic Phases Formed on Solidification of Aluminium Alloys. JOM 2019, 71, 1731–1736. [Google Scholar] [CrossRef]
  30. Li, P.; Zhao, F.; Xiao, H.; Zhang, H.; Gong, H.; Zhang, S.; Liu, Z.; Zu, X. First-Principles Study of Thermo-Physical Properties of Pu-Containing Gd2Zr2O7. Nanomaterials 2019, 9, 196. [Google Scholar] [CrossRef]
  31. Zhao, F.A.; Xiao, H.Y.; Liu, Z.J.; Li, S.; Zu, X.T. A DFT study of mechanical properties, thermal conductivity and electronic structures of Th-doped Gd2Zr2O7. Acta Mater. 2016, 121, 299–309. [Google Scholar] [CrossRef]
  32. Zhao, F.A.; Xiao, H.Y.; Jiang, M.; Liu, Z.J.; Zu, X.T. A DFT+U study of Pu immobilization in Gd2Zr2O7. J. Nucl. Mater. 2015, 467, 937–948. [Google Scholar] [CrossRef]
  33. Sun, Q.; Liu, C.; Fan, T.; Cheng, H.; Cui, P.; Gu, X.; Chen, L.; Ata-Ul-Karim, S.T.; Zhou, D.; Wang, Y. A molecular level understanding of antimony immobilization mechanism on goethite by the combination of X-ray absorption spectroscopy and density functional theory calculations. Sci. Total Environ. 2023, 865, 161294. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, W.; Duan, H.; Wei, D.; Cui, B.; Wang, X. Stability of diethyl dithiocarbamate chelates with Cu(II), Zn(II) and Mn(II). J. Mol. Struct. 2019, 1184, 375–381. [Google Scholar] [CrossRef]
  35. Pfrommer, B.G.; Côté, M.; Louie, S.G.; Cohen, M.L. Relaxation of Crystals with the Quasi-Newton Method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef]
  36. Liao, M.; Liu, Y.; Min, L.; Lai, Z.; Han, T.; Yang, D.; Zhu, J. Alloying effect on phase stability, elastic and thermodynamic properties of Nb-Ti-V-Zr high entropy alloy. Intermetallics 2018, 101, 152–164. [Google Scholar] [CrossRef]
  37. Liu, K.-C.; Thomas, G.; Caballero, A.; Moya, J.S.; de Aza, S. Time-Temperature-Transformation Curves for Kaolinite-α-Alumina. J. Am. Ceram. Soc. 1994, 77, 1545–1552. [Google Scholar] [CrossRef]
  38. Han, L.-F.; Xu, Z.-L.; Cao, Y.; Wei, Y.-M.; Xu, H.-T. Preparation, characterization and permeation property of Al2O3, Al2O3–SiO2 and Al2O3–kaolin hollow fiber membranes. J. Membr. Sci. 2011, 372, 154–164. [Google Scholar] [CrossRef]
  39. Ravindran, P.; Fast, L.; Korzhavyi, P.A.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  40. Dobson, P.J. Physical Properties of Crystals—Their Representation by Tensors and Matrices. Phys. Bull. 1985, 36, 506. [Google Scholar] [CrossRef]
  41. Li, H.; Sang, H.; Mao, C.; Wang, Y.; Xu, L.; Wu, Y. Study on stable solidification of silica-based ammonium molybdophosphate adsorbing cesium: Micromechanics and density functional theory modeling. J. Nucl. Mater. 2022, 560, 153501. [Google Scholar] [CrossRef]
  42. Koumpouras, K.; Larsson, J.A. Distinguishing between chemical bonding and physical binding using electron localization function (ELF). J. Phys. Condens. Matter 2020, 32, 315502. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM images (left) and principal compositions of natural allophanes (right).
Figure 1. SEM images (left) and principal compositions of natural allophanes (right).
Toxics 11 00850 g001
Figure 2. XRD patterns of allophanes sintered at different temperatures.
Figure 2. XRD patterns of allophanes sintered at different temperatures.
Toxics 11 00850 g002
Figure 3. Sr immobilization ratios at different sintering temperatures.
Figure 3. Sr immobilization ratios at different sintering temperatures.
Toxics 11 00850 g003
Figure 4. XRD patterns of solidified products at 1200 °C.
Figure 4. XRD patterns of solidified products at 1200 °C.
Toxics 11 00850 g004
Figure 5. (A) SEM images and (B) EDS mapping of solidified product.
Figure 5. (A) SEM images and (B) EDS mapping of solidified product.
Toxics 11 00850 g005
Figure 6. Structures of (A) SrAl2Si2O8 and (B) Sr2Al2SiO7 after optimization.
Figure 6. Structures of (A) SrAl2Si2O8 and (B) Sr2Al2SiO7 after optimization.
Toxics 11 00850 g006
Figure 7. Elastic modulus and the anisotropy of SrAl2Si2O8 and Sr2Al2SiO7.
Figure 7. Elastic modulus and the anisotropy of SrAl2Si2O8 and Sr2Al2SiO7.
Toxics 11 00850 g007
Figure 8. Anisotropy of the Young’s modulus in (A) SrAl2Si2O8 and (B) Sr2Al2SiO7.
Figure 8. Anisotropy of the Young’s modulus in (A) SrAl2Si2O8 and (B) Sr2Al2SiO7.
Toxics 11 00850 g008
Figure 9. Anisotropy of the shear modulus in (A) SrAl2Si2O8 and (B) Sr2Al2SiO7.
Figure 9. Anisotropy of the shear modulus in (A) SrAl2Si2O8 and (B) Sr2Al2SiO7.
Toxics 11 00850 g009
Figure 10. Comparison of experimental and simulated XRD patterns: (A) SrAl2Si2O8 and Ca1−xSrxAl2Si2O8, and (B) Sr2Al2SiO7 and Ca2−xSrxAl2SiO7. (Green column: 2θ = 6.8°; pink column: 2θ = 11.38°; red column: 2θ = 16.12° and 17.38°).
Figure 10. Comparison of experimental and simulated XRD patterns: (A) SrAl2Si2O8 and Ca1−xSrxAl2Si2O8, and (B) Sr2Al2SiO7 and Ca2−xSrxAl2SiO7. (Green column: 2θ = 6.8°; pink column: 2θ = 11.38°; red column: 2θ = 16.12° and 17.38°).
Toxics 11 00850 g010
Figure 11. (A) Elastic modulus and (B) Pugh’s ratio of Ca2−xSrxAl2SiO7.
Figure 11. (A) Elastic modulus and (B) Pugh’s ratio of Ca2−xSrxAl2SiO7.
Toxics 11 00850 g011
Figure 12. (A) Three-dimensional ELF diagrams of SrAl2Si2O8 (B,C) are two-dimensional electron localization diagrams of SrAl2Si2O8 on the cross-section at 3.24 and 9.72 Å from the set (0 1 0) crystal plane, respectively.
Figure 12. (A) Three-dimensional ELF diagrams of SrAl2Si2O8 (B,C) are two-dimensional electron localization diagrams of SrAl2Si2O8 on the cross-section at 3.24 and 9.72 Å from the set (0 1 0) crystal plane, respectively.
Toxics 11 00850 g012
Figure 13. (A) Three-dimensional ELF diagrams of Sr2Al2SiO7 (BE) are two-dimensional electron localization diagrams of Sr2Al2SiO7 on the cross-section at 1.175, 2.74, 5.09, and 6.66 Å from the set (1 0 0) crystal plane, respectively.
Figure 13. (A) Three-dimensional ELF diagrams of Sr2Al2SiO7 (BE) are two-dimensional electron localization diagrams of Sr2Al2SiO7 on the cross-section at 1.175, 2.74, 5.09, and 6.66 Å from the set (1 0 0) crystal plane, respectively.
Toxics 11 00850 g013
Figure 14. Bond energies of Sr in different mineral phases.
Figure 14. Bond energies of Sr in different mineral phases.
Toxics 11 00850 g014
Figure 15. (A) Elastic modulus and (B) Pugh’s ratios of different mineral phases.
Figure 15. (A) Elastic modulus and (B) Pugh’s ratios of different mineral phases.
Toxics 11 00850 g015
Table 1. The cell parameters of SrAl2Si2O8 and Sr2Al2SiO7 after optimization.
Table 1. The cell parameters of SrAl2Si2O8 and Sr2Al2SiO7 after optimization.
CrystalCrystal SystemSpace GroupLattice Parameter
a (Å)b (Å)c (Å)αβγ
SrAl2Si2O8MonoclinicC2/c9.3859.3859.65081.056°81.056°74.481°
Sr2Al2SiO7TriclinicP17.8375.2647.84890.231°90.000°90.000°
Table 2. The stiffness tensor parameters of SrAl2Si2O8 and Sr2Al2SiO7 after optimization.
Table 2. The stiffness tensor parameters of SrAl2Si2O8 and Sr2Al2SiO7 after optimization.
CrystalStiffness Tensor Cij/GPa
SrAl2Si2O8C11C12C13C15C22C23C25
154.74243.41472.79313.905187.34652.2−16.89
C33C35C44C46C55C66
129.94−0.0114.4299.93145.87213.305
Sr2Al2SiO7C11C12C13C14C15C16C22
193.27163.59978.31.87800160.477
C23C24C25C26C33C34C35
62.6740.00100191.154−1.1410
C36C44C45C46C55C56C66
046.0790068.894−0.54645.721
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

Wu, Y.; Sang, H.; Zheng, J.; Yang, S.; Gu, Z.; Wu, H.; Wei, Y. Preparation and Density Functional Theory Studies of Aluminosilicate-Based Ceramic Solidified Products for Sr Immobilization. Toxics 2023, 11, 850. https://doi.org/10.3390/toxics11100850

AMA Style

Wu Y, Sang H, Zheng J, Yang S, Gu Z, Wu H, Wei Y. Preparation and Density Functional Theory Studies of Aluminosilicate-Based Ceramic Solidified Products for Sr Immobilization. Toxics. 2023; 11(10):850. https://doi.org/10.3390/toxics11100850

Chicago/Turabian Style

Wu, Yan, Hongji Sang, Jiawei Zheng, Shuyi Yang, Zhengcheng Gu, Hao Wu, and Yuezhou Wei. 2023. "Preparation and Density Functional Theory Studies of Aluminosilicate-Based Ceramic Solidified Products for Sr Immobilization" Toxics 11, no. 10: 850. https://doi.org/10.3390/toxics11100850

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