Next Article in Journal
The Casting and Hot Forging of Low-Carbon Copper-Bearing Steel and Its Substructural Characterization
Next Article in Special Issue
Mechanics and Crack Analysis of Irida Graphene Bilayer Composite: A Molecular Dynamics Study
Previous Article in Journal
The Effects of Process Parameters on the Porosity of a VBO Prepreg/Fiber-Interleaved Layup Composite
Previous Article in Special Issue
Numerical Modeling of Single-Lap Shear Bond Tests for Composite-Reinforced Mortar Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atomic Insights into the Structural Properties and Displacement Cascades in Ytterbium Titanate Pyrochlore (Yb2Ti2O7) and High-Entropy Pyrochlores

1
Department of Civil, Architectural, and Environmental Engineering, Missouri University of Science and Technology, Rolla, MO 65409, USA
2
College of Nuclear Science and Technology, Harbin Engineering University, Harbin 150001, China
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(10), 413; https://doi.org/10.3390/jcs7100413
Submission received: 15 July 2023 / Revised: 7 September 2023 / Accepted: 21 September 2023 / Published: 5 October 2023
(This article belongs to the Special Issue Theoretical and Computational Investigation on Composite Materials)

Abstract

:
Pyrochlore oxides ( A 2 B 2 O 7 ) are potential nuclear waste substrate materials due to their superior radiation resistance properties. We performed molecular dynamics simulations to study the structural properties and displacement cascades in ytterbium titanate pyrochlore ( Yb 2 Ti 2 O 7 ) and high-entropy alloys (HEPy), e.g., YbYTiZr O 7 , YbGdTiZr O 7 , and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 . We computed lattice constants (LC) ( a o ) and threshold displacement energy (Ed). Furthermore, the calculation for a o and ionic radius ( r i o n i c ) were performed by substituting a combination of cations at the A and B sites of the original pyrochlore structure. Our simulation results have demonstrated that the lattice constant is proportional to the ionic radius, i.e., a o   α   r i o n i c . Moreover, the effect of displacement cascades of recoils of energies 1 keV, 2 keV, 5 keV, and 10 keV in different crystallographic directions ([100], [110], [111]) was studied. The number of defects is found to be proportional to the energy of incident primary knock-on atoms (PKA). Additionally, the E d of pyrochlore exhibits anisotropy. We also observed that HEPy has a larger E d as compared with Yb 2 Ti 2 O 7 . This establishes that Yb 2 Ti 2 O 7 has characteristics of lower radiation damage resistance than HEPy. Our displacement cascade simulation result proposes that HEPy alloys have more tendency for trapping defects. This work will provide atomic insights into developing substrate materials for nuclear waste applications.

1. Introduction

Despite the environment-friendly and high-efficacy energy source, nuclear energy has a few distinctive concerns [1]. Commercial use of nuclear power is only evident when radioactive waste management is done under a proper strategy. Therefore, the reprocessing of spent fuel is an important part of ensuring nuclear safety and environmental safety. Nuclear wastes have different forms depending on the sources and radioactive concentration [2]. Solidification of liquid waste is a common process for managing nuclear waste disposal. In this regard, the most common processes are vitrification and synroc methods [3]. Although borosilicate glass is being used frequently as a nuclear waste substrate matrix, having low solubility of actinide elements restricts their usage [4]. An alternate substitute to solve this limitation is pyrochlore ceramics due to their superior durability, and better potential at high temperatures and humid environments [5,6]. Moreover, ceramics have high values of waste loading as compared to glass [7]. The ceramic-based substrate with the minor addition of an ionic concentration has higher radiation stability and excellent chemical and physical properties [8]. Long-term radiation damage is critical for a nuclear waste disposal strategy [9]. Pyrochlore has been developed rapidly as high-entropy ceramics waste substrate materials in recent years [10,11]. Future state-of-the-art nuclear reactors involve the recycling of nuclear fuels, burning of minor actinide series, and decontamination of fission fragments [12]. Ytterbium titanate pyrochlores ( Yb 2 Ti 2 O 7 ) have been demonstrated to be an effective waste substrate material for pyrochemical reactions [13,14]. They are used as waste forms for immobilizing radioactive waste, particularly high-level radioactive waste (HLW) [13,14]. It is also notable that the aforementioned pyrochlore serves as a surrogate for actinides like plutonium and americium. Thus, making it a suitable compound for studying waste immobilization behaviors. Moreover, ytterbium titanate pyrochlores have inherent radiation resistance due to their open crystal structure, which allows them to accommodate radiation-induced defects [15].
Minervini et al. reported disorder in different pyrochlore oxides using the atomic-scale simulation method. Additionally, pyrochlore demonstrated stability range dependence on the relative size of cations [16]. Brendan et al. have investigated the relationship between the structural and bonding energy in lanthanide pyrochlore oxides ( S n 2 O 7 ) and found that the position parameter of the oxygen vacancies is inversely proportional to the lattice parameter. Recently, Zhang et al. have reported on the machine learning (ML) methods to determine lattice constants of different multi-substitutional pyrochlores in the range of 9–11 Å [17]. Moreover, the pyrochlore stannate (Ce2Sn2O7) demonstrated a temperature-dependent anisotropic nature [18]. Superconductivity of ternary pyrochlore oxide ( Cd 2 Re 2 O 7 ) was reported by Sakai et al., at 1.1 K, and the lattice constant was found to be 10.22 Å at room temperature [19]. Liyuan et al. performed molecular dynamics to study structural and elastic properties of different pyrochlores under numerous combinations of A and B cations. It was found that displacement energy in pyrochlores strongly depends on the energy of the incident PKA’s as well as their atomic masses [20]. They also reported that lattice parameters and the atomic radius predominantly affect structural as well as thermal properties. Chartier et al. performed molecular dynamics simulations to study displacement cascades in lanthanum zirconate pyrochlore (Ln2Zr2O7) with uranium ions bombarded with 6 keV at 350 K along different orientations. It was observed that the model does not lose its crystallinity [21]. It is further reported that the displacement cascades are determinants of material radiation stability at elevated temperatures [22]. Atomic-scale radiation damage studies in pyrochlores (Gd2Zr2O7) have demonstrated healing mechanism during radiation [9,23]. Moreover, a combination of different anions and cations results in a radiation-resistant response of pyrochlores [24]. Pyrochlores have disordered structures due to irradiation and alteration of thermodynamics [25]. Pyrochlores are known for their radiation-tolerant behavior primarily because of their inherent structural stability. Their structure allows migration and annihilation of defects, minimizing their impact on the material’s overall properties.
High-entropy pyrochlores refer to a class of materials that exhibit high configurational entropy due to the random distribution of multiple cations on the crystal lattice sites. These materials have been studied for various applications, including radiation tolerance. High-entropy pyrochlore (HEPy) oxides synthesized with Yb 2 Ti 2 O 7 are obtained by substituting the cations in the pyrochlore structure and they have displayed higher radiation resistance than individual pyrochlores [26,27,28,29]. Atomic-scale simulation results have also displayed that radiation resistance is affected by the addition of Zr content in HEPy [30]. Defect dynamics in pyrochlore-type structures are quite difficult to determine in experimental studies due to their complex structure and irradiation-associated phenomena [9,20]. The displacement cascades involve the initiation of damage in the materials alternately determine the long-term possible outcomes. Atomic-scale simulation methods can shed light on determining the fundamental material properties and interaction mechanisms in pyrochlores [24,31] It is also important to mention that a limited number of experimental and simulation studies have been performed on the behavior of different pyrochlores under radiation [9,26,32,33,34]. Molecular dynamics simulations have been applied to study the influence of displacement cascades on the microstructural properties of different pyrochlores [9,32]. Computer simulations interpret the radiation damage as a predictive tool for processing the experiments. The simulation models describe the physics behind the phenomena responsible for the radiation damage mechanism in irradiated material by providing valuable tools observed in nuclear power plants (NPP). Numerous radiation damage studies have been reported on metals and alloys [35,36,37,38,39,40,41]. This study will serve as a predictive model for providing insight into expected behavior in radioactive environments.
This article studies three aspects of pyrochlore oxides. Primarily, interatomic potential was established by parametrizing existing literature and later validated through calculation for LC, rionic, and Ed of different combinations of A and B in pyrochlore structures. Furthermore, displacement damage cascade simulations in Yb 2 Ti 2 O 7 were performed with each constituent element. The primary knock-on atom’s energies (EPKA~1, 2, 5, and 10 keV) with different incident directions were used for studying the effect of energy dependence. Moreover, displacement cascades keeping the same energy were studied for HEPy to determine its radiation resistance as compared to Yb 2 Ti 2 O 7 .

2. Computational Details and Methodology

2.1. Crystal Structures

Pyrochlores have cubic fluorite-type structures containing 5 or more elements with oxygen-deficient vacancies having A 2 B 2 O 7 type structure with A (rare-earth) and B (transition metal) being different cations [42,43]. The ‘A’ position coordinated with 3+ cation atoms while the ‘B’ position coordinated with 4+ cation atoms, e.g., 3+ cations (La3+, Nd3+, Gd3+, Sm3+, and Y3+) and 4+ cations (Zr4+, Ti4+, and Mo4+) [44]. Moreover, 1/8 of the oxygen atoms are vacant to balance the charges [5]. The different ion combinations are used to compute different structural properties. The crystal structure is developed by Visualization of Electronic and Structural Analysis (VESTA) [45]. The lattice parameter for pyrochlore oxide was set as 10.194 Å. The initial model comprised eight unit cells with 88 atoms per unit cell with the Fd 3 ¯ m group having composition Yb16Ti16O56 with a unit cell of volume 1059.578190 3 . Periodic boundary conditions were applied along axes. The initial structure was equilibrated using the steep decent method (SD) since this method is considered a quick approach for an optimal efficient solution [46]. For HEPy modeling, the selection of compound Yb0.5Y0.5Eu0.5Gd0.5TiZrO7 represented in equivalent to original Yb2Ti2O7 structure with a stoichiometric ratio replacing Yb2     Yb Yb     (Yb0.5Y0.5) (Eu0.5Gd0.5) and Ti2 Ti Ti Ti Zr. In other words, we can understand it by considering that the atomic number ratio at the A-site is doped with three elements Y, Eu, and Gd with a total of four elements while the B-site is doped with two elements Ti and Zr with a total of three elements. The element in atomic proportion is represented as Yb:Y:Eu:Gd:Ti:Zr     1:1:1:1:2:2 resulting in the formation of Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 Ti   Zr O 7 . Figure 1 displays the computational models of unit cells of all models, where (a) is an MD snapshot of a relaxed computational model under study and (b) is a unit cell of a generic representation of pyrochlore [47], and (c–f) represent unit cells of Yb 16 Ti 16 O 56 , YbYTiZr O 7 , YbGdTiZr O 7 , and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 , respectively. The composition of each pyrochlore is mentioned as labelled in Figure 1.
The characteristic crystal structure of pyrochlores consists of a network of corner-sharing tetrahedra, which provides an open framework that can accommodate defects induced by radiation damage [48,49]. Molecular dynamics simulations were performed by a large-scale atomic/molecular massively parallel simulator (LAMMPS) developed by Plimpton [50] and structure analysis and visualization were done by OVITO [51].

2.2. Interatomic Potential

In MD simulation, realistic interatomic potentials are very crucial for reliable results [52]. The interatomic potential was modified by a coupling potential developed in earlier studies [24,53]. This potential was tested for replicating equilibrium lattice constants and lattice constants reported in earlier studies [54,55]. The potential function is divided into two parts and all pair interactions are described separately. The short-range combination of ZBL (Ziegler-Biersack-Littmark) potential and Buckingham potential are adjoined with a fitting function for smooth truncation. The long-range Coulomb potential is described by particle–particle–particle–mesh solver (PPPM) summation algorithm having accuracy up to 10−4 and cutoff radius (rcutoff = 10 Å) [56]. The Buckingham potential results in negative infinity due to short-range interactions whereas ZBL potential is more realistic for charged modes at short-range interactions [57]. The piecewise interatomic potential of Yb 2 Ti 2 O 7 is expressed in Equation (1), where r 1 , 2 is the active range of potentials. The ZBL potential and Buckingham potential are fitted with a fourth-order exponential spline function represented as V spline   fun . r ij = e x p   ( c 1 + c 2 r i j + c 3 r i j 2 + c 4 r i j 3 ) for smooth interconnection, where c 1 4 are spline constants.
V r i j = V Z B L r i j r i j <   r 1 V s p l i n e   fun . r i j r 1 < r ij <   r 2 V B u c k . r i j     r 2 < r i j <   r i j 0      r cut < r i j
The ZBL potential is defined as V ZBL = 1 4 π ε 0 Z 1 Z 2 e 2 r i j G ( α ) , where G α = 0.18118 e 32 α + 0.5099 e 0.9423 α + 0.2802 e 0.4029 α + 0.02817 e 0.2016 α and α = r i j ( Z 1 0.23 + Z 2 0.23 ) 0.8854 a 0 . Regarding the notation in the above expression, ε 0 represents dielectric constant, a 0 represents Bohr’s radius, e is the electronic charge, and Z 1   ,   Z 2 are the atomic number of elements. The atomic numbers of Yb70, O8, and Ti48. The potential function of Buckingham V B u c k . r ij potential is represented as a function of distance in Equation (2), where A ij , ρ , and C xy are element-dependent potential parameters.
V B u c k . r ij = A ij e r ij ρ     C xy r ij 6
The Buckingham potential function for Yb and O satisfy conditions in Equation (2). The values of the parameters are set for each computed pair in Table 1 as reported by earlier studies [20,53]. For Ti–O, the value of C xy is adjusted for smooth truncation. Before testing the potential further, interatomic potentials were coupled through a splined function to ZBL repulsive and Buckingham functions. The resulting functions are plotted in Figure 2. The connected potentials for cation–anion/anion–anion are computed through Table 1, while cation–cations are defined by pair style automatically computed by LAMMPS as mentioned in the previous studies [58,59].

2.3. Displacement Cascades Simulations

Displacement cascade simulations for all cases were performed through molecular dynamics code LAMMPS [60]. A system having dimensions 10 × 10 × 10 with a total of 88,000 were constructed. Our calculation for lattice parameters resulted in 10.09 Å with periodic boundary set along all axes. The interatomic-potential function was employed as mentioned in the previous section. After completing the preliminary settings, thermal relaxation was performed initially using a canonical (NVT) ensemble at 300 K at a timestep of 0.001 ps. After the thermal relaxation, the model was relaxed again using a microcanonical ensemble (NVE) at a time step of 0.0001 ps to ensure the system had attained sufficient equilibrium before proceeding with the displacement cascades simulation. For damage cascade simulations, the recoil energies were chosen as 1 keV, 2 keV, 5 keV, and 10 keV. The selected PKA’s Yb atom and the PKA’s angles are along the positive Z axis. The direction of PKA was set at 45° along the xy plane at 7° to avoid channeling effects.

3. Results and Discussion

3.1. Calculation for Equilibrium Lattice Constant and Ionic Radii

The sites on A and B in the pyrochlore structure have different ions and their combination affects the properties of the pyrochlores. An important parameter of the unit cell is the lattice constant and its equilibrium properties. Figure 3 displays total energy as a function of the LC of Yb 2 Ti 2 O 7 . The concave curve indicates that the energy of the system will reach the minimum at a certain equilibrium lattice constant. The curve fitting is obtained by a polynomial function. The equilibrium lattice constant was found to be 10.05 Å which agrees with the results reported earlier [47,53].
We tested changing different cations at the A site, i.e., Lu3+, Yb3+, Er3+, Y3+, Gd3+, Eu3+, Sm3+, Nd3+, La3+, and Ce3+ while the cation at the B site was Ti4+, e.g., the first entry in Table 2 is represented as Lu2Ti2O7. The equilibrium lattice constant (LC) and ionic radius (rionic) were calculated for both combinations and compared with the existing literature. Table 2 represents the calculation for the LC and rionic of different ions at site A. The calculations indicate that LC is caused by the replacement of ions proportional to rionic. Our calculations are in good agreement with other studies [18,61,62,63,64]. The preliminary fitting leads to the desired value and replicates the validity of the potential for further testing the displacement cascades of models.
Figure 4 displays the XRD pattern of Yb 2 Ti 2 O 7 , YbYTiZrO7, YbGdTiZrO7, and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 . The peaks of HEPy exhibit the characteristics of original cubic pyrochlores with Fd-3m and an LC of 10.19478 Å, 10.19497 Å, 10.19504 Å, 10.19513 Å were found for simulated structures. X-ray dispersion coefficients for λ = 1.54059 Å with CuKα were chosen during the XRD pattern. The obtained values are in good agreement with other studies on pyrochlores [65,66].

3.2. Threshold Displacement Energy (Ed)

The incident energy of primary knock-on atoms (EPKA) exceeds the constraints of the entire force field and leaves a vacancy. This minimum energy (Emin) required to displace an atom from its lattice site is known as threshold displacement energy (Ed) [32]. The theoretical calculations displayed that off-site threshold energy depends on crystallographic direction, temperature, orientation of incident PKA, energy, and type of incident particles.
We have computed the threshold energy of Yb 2 Ti 2 O 7 initially by selecting Yb and Ti with O atoms as incident PKA along [100], [110], and [111] axes. The Ed is displayed along the orientation in Figure 5. It can be seen that Ed (Ti) >> Ed (Yb) > Ed (O). We have seen that Ed is anisotropic. Our findings are in agreement with other studies [67,68].
The calculation of the threshold energy (Ed) of high-entropy pyrochlore YbYTiZr O 7 was formed by doping the Y element at the A site and the Zr element at the B site. The atomic number ratio for simulated pyrochlore with substitution is set as Yb:Y:Ti:Zr 1:1:1:1. Later, A-site is doped with three elements as Y, Eu, and Gd while B-site is doped with a total of four elements in equal atomic proportion of seven constituent elements, i.e., Yb:Y:Gd:Eu:Ti:Zr 1:1:1:1:2:2 resulting in the formation of Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 . Table 3 compares the average Ed of two simulated models with incident PKA’s directions. It is established that E d ( Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 ) > E d ( Yb 2 Ti 2 O 7 ). This establishes that HEPy is less prone to defects [67]. The values of Yb are not much different from the results in Figure 4, but the value of Ti is 2 to 4 times higher, and the Ed value of O is 10, 50, or more times that of Yb 2 Ti 2 O 7 . This difference is due to the fact that high-entropy alloys have a more complex structure. This complexity can lead to higher Ed due to strong bonding, defects, and strains [43,47,69,70].

3.3. Effect of Displacement Cascades on Energy and Direction of PKA

Irradiated materials are associated with displacement cascades. The displacement cascades are divided into different subcascades during bombardment of incident neutrons and ion implantation. The displacement cascades with different incident PKA’s are represented in Figure 6.
As the energy of the initial PKA increases, the total number of defects of various atoms increases due to an increase in the mean free path between atoms. Higher EPKA produces larger area cascades which have a higher possibility of reaching Ed [32]. The total number of defects will have a characteristic of rapid growth and then annihilate with time. This situation is generally referred to as defect recombination and permanent defect after radiation. In Figure 7, it can be noted that during the initial stage of the simulation cascade, collision initiated a rapid accumulation of defects which occurred in the pyrochlore lattice. The number of defects reached a peak at 0.3 ps and afterwards the size of the cascade continuously reduced. The defects stabilized at 0.9 ps leaving permanent defects.
During the cascade simulations, incident PKA was located at the lower left corner and moved along the diagonal. The effect of incident direction is investigated with Yb incident with 5 keV along the X-axis inclined at 45° at an interval of 5° with a 7° channeling angle along the Z-axis. The relation between incident angle as a function of time is expressed in Figure 8. It can be seen that smaller inclination angles have a lower number of defects. This is due to the fact that higher inclination angles are associated with lower threshold energies of incident PKA’s as well as successive overlapping cascades [32].
To observe the effect of ions on inclination angle, we compared the effect of defect evolution at lower and higher inclination angles by a 5 keV Yb atom. It is also known from Table 3 that Ti has higher values of Ed as compared to other constituents. The relation between the number of defects as a function of time is displayed in Figure 9 for the different incident directions of PKA. It is seen that Ti atoms and Yb atoms are almost the same in 55° (Figure 9a) whereas Yb is double in Figure 9b. In both cases, O has a higher number of defects than other constituents and a higher angle has twice the number of defects. Therefore, it is more difficult to produce more defects in 55° inclination under the same energy of incident PKA.

3.4. Displacement Cascades in HEPy

This effect of displacement cascades on HEPy is studied by doping different elements on A and B sites of the pyrochlore Yb 2 Ti 2 O 7 structure. We have selected Zr at B sites while the A site is changed forming three kinds of HEPy, i.e., YbYTiZr O 7 , YbGdTiZr O 7 , and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 . The displacement cascades formed by Yb with 5keV and defects morphology in HEPy are compared with Yb 2 Ti 2 O 7 . The comparison is based on the analysis and defect morphology.
Figure 10 displays the displacement cascade at its peak and after defect annihilation with Yb incident PKA of 5 keV. The defect forms of Yb 2 Ti 2 O 7 and the other three high-entropy pyrochlores are compared. The subscript for each displays the peak defects Nmax and after annihilation. It is noted that the area of displacement cascades is larger for Yb 2 Ti 2 O 7 whereas the number of defects formed by displacement cascades in HEPy is higher than pyrochlore whereas Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 has a lower number of surviving defects. For all the models, the direction of PKA coincides with the defects cluster direction. We have reason to believe that high-entropy pyrochlore has a better radiation resistance compared with Yb 2 Ti 2 O 7 . It is also seen from the displacement cascade snapshots the characteristic of limiting the circumferential development of defects and reducing the damage volume in HEPy.
Higher energy incident PKA are reported to have a higher number of surviving defects [14,71,72]. The number of defects as a function of time are compared for HEPy and pyrochlore as is plotted in Figure 11. It can be seen from Figure 10 that the number of defects corresponding to YGdEu doped HEPy is lesser in all cases. The surviving defects in undoped pyrochlore are relatively larger than doped ones. The defect growth rate, defect disappearance rate, and defect number annihilation is faster in HEPy, especially Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 . It is preliminarily due to the fact that HEPy provides a large number of available atomic configurations, which can accommodate point defects and compensate for local structural changes induced by radiation. This configurational flexibility enhances the material’s ability to self-heal and recover from radiation damage, further contributing to its lower radiation damage radius. This also establishes that HEPy shows a promising characteristic for radiation tolerance. Thus, MD results in this study can be used to optimize the radiation-resistant nature of HEPy by testing other models with orientation and other constituent elements in the composition.

4. Conclusions

Ytterbium titanate pyrochlore oxides with different compositions and high-entropy pyrochlores significantly expand the studied ceramics for the immobilization of Low-Level Radioactive Waste (LRW) ensuring better waste containment and optimizing storage solutions. Studying radiation resistance and the defects formation mechanism is crucial for designing pyrochlore for Radioactive Waste (RW) immobilization materials. Moreover, analyzing the crystal structure of the high-entropy pyrochlores under various conditions helps to understand any changes due to radiation environments. High-entropy pyrochlores contain multiple elements, potentially providing a greater capacity to accommodate and immobilize various radioactive isotopes through diverse interactions.
Molecular dynamics simulations were performed to study the effect of displacement cascades in ytterbium titanate oxide ( Yb 2 Ti 2 O 7 ) and high-entropy alloys. Calculations were made for equilibrium lattice constant pyrochlore oxides. The displacement cascades in Yb 2 Ti 2 O 7   with Yb, Ti, O of 1 keV, 2 keV, 5 keV, and 10 keV along [100], [110], and [111] were used to determine the threshold energy (Ed) on the radiation damage response of pyrochlore. The effect of displacement cascades on the formation of high-entropy pyrochlore YbYTiZr O 7 , YbGdTiZr O 7 , Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 , and Yb 2 Ti 2 O 7 was compared to analyze the radiation stability. The displacement cascades of different incident PKA displayed that the number of defects is proportional to the incident energy. Higher incident angles have a lower number of defects. The position and inclination angle of the incident angle are important in determining the surviving defects. The defects morphology of HEPy was compared with Yb 2 Ti 2 O 7 .
Our study will provide valuable insight and contribute to the advancement of nuclear waste materials. High-entropy pyrochlores offer a broader range of compositions and structures, which could make them more versatile for different nuclear applications. This versatility allows you to fine-tune the material properties based on the application’s requirements. High-entropy pyrochlores are still a relatively new area of research and it requires an innovative material design approach. We still believe that more extensive research is required to explore their behavior under radiation and in nuclear environments.

Author Contributions

Conceptualization, supervision, and project administration, Funding, Q.W.; Methodology, software, validation, and preparation of the manuscript, M.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by Fundamental Research Funds for the Central Universities (Grant No. 3072022JC1502).

Data Availability Statement

Data is contained within this article.

Acknowledgments

The authors would like to thank Muhammad Asim (HEU) and Muhammad Zeshan Khalid (Department of Materials Science and Engineering, NTNU) for the useful discussion. Q.W. acknowledges the support from Fundamental Research Funds for the Central Universities (Grant No. 3072022JC1502).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abram, T.; Ion, S. Generation-IV nuclear power: A review of the state of the science. Energy Policy 2008, 36, 4323–4330. [Google Scholar] [CrossRef]
  2. Dungan, K.; Gregg, R.W.H.; Morris, K.; Livens, F.R.; Butler, G. Assessment of the disposability of radioactive waste inventories for a range of nuclear fuel cycles: Inventory and evolution over time. Energy 2021, 221, 119826. [Google Scholar] [CrossRef]
  3. Amiard, J. Nuclear Waste Disposal Methods. Manag. Radioact. Waste 2021, 5, 23–66. [Google Scholar] [CrossRef]
  4. Kong, L.; Karatchevtseva, I.; Zhang, Y.; Wei, T. The incorporation of Nd or Ce in CaZrTi2O7 zirconolite: Ceramic versus glass-ceramic. J. Nucl. Mater. 2021, 543, 152583. [Google Scholar] [CrossRef]
  5. Liu, Z.; Liu, K.; Zheng, X.; Wang, Y.; Sun, X.; Xue, P.; Li, C.; Du, M. Formulation of Poly(ionic liquids)@COF Nanotrap for Efficient Perrhenate Sequestration from Alkaline Nuclear Waste. Chem. Mater. 2022, 34, 5452–5456. [Google Scholar] [CrossRef]
  6. Guo, X.; Gin, S.; Frankel, G.S. Review of corrosion interactions between different materials relevant to disposal of high-level nuclear waste. npj Mater. Degrad. 2020, 4, 34. [Google Scholar] [CrossRef]
  7. Frankel, G.S.; Vienna, J.D.; Lian, J.; Guo, X.; Gin, S.; Kim, S.H.; Du, J.; Ryan, J.V.; Wang, J.; Windl, W.; et al. Recent Advances in Corrosion Science Applicable To Disposal of High-Level Nuclear Waste. Chem. Rev. 2021, 121, 12327–12383. [Google Scholar] [CrossRef]
  8. Xu, H.; Wang, Y.; Zhao, P.; Bourcier, W.L.; Van Konynenburg, R.; Shaw, H.F. Investigation of Pyrochlore-Based U-Bearing Ceramic Nuclear Waste: Uranium Leaching Test and TEM Observation. Environ. Sci. Technol. 2004, 38, 1480–1486. [Google Scholar] [CrossRef]
  9. Dholakia, M.; Chandra, S.; Jaya, S.M. Comparative study of radiation damage in Gd2Zr2O7 and Y2Ti2O7 crystal: A molecular dynamics investigation. AIP Conf. Proc. 2020, 2265, 30373. [Google Scholar] [CrossRef]
  10. Zhang, Y. Ceramic Materials for Nuclear Energy Applications. JOM 2019, 71, 4806–4807. [Google Scholar] [CrossRef]
  11. Marra, J. Advanced ceramic materials for next-generation nuclear applications. IOP Conf. Ser. Mater. Sci. Eng. 2011, 18, 16001. [Google Scholar] [CrossRef]
  12. Lee, W.H.; Cheong, J.H. Potential radiological hazard and options to cope with consequences from recycling of activated metal waste disposed of at a near-surface disposal facility. Ann. Nucl. Energy 2021, 152, 107993. [Google Scholar] [CrossRef]
  13. Hand, M.L.; Stennett, M.C.; Hyatt, N.C. Rapid low temperature synthesis of a titanate pyrochlore by molten salt mediated reaction. J. Eur. Ceram. Soc. 2012, 32, 3211–3219. [Google Scholar] [CrossRef]
  14. Dholakia, M.; Chandra, S. Structural stability of titanate pyrochlores undergoing radiation damage. Comput. Mater. Sci. 2022, 201, 110889. [Google Scholar] [CrossRef]
  15. Mohan, S.; Kaur, G.; David, C.; Panigrahi, B.K.; Amarendra, G. Ab initio molecular dynamics simulation of threshold displacement energies and defect formation energies in Y4Zr3O12. J. Appl. Phys. 2020, 127, 235901. [Google Scholar] [CrossRef]
  16. Minervini, L.; Grimes, R.W.; Sickafus, K.E. Disorder in pyrochlore oxides. J. Am. Ceram. Soc. 2000, 83, 1873–1878. [Google Scholar] [CrossRef]
  17. Zhang, Y. Machine learning the lattice constant of cubic pyrochlore compounds. Int. J. Appl. Ceram. Technol. 2021, 18, 661–676. [Google Scholar] [CrossRef]
  18. Kennedy, B.J.; Hunter, B.A.; Howard, C.J. Structural and Bonding Trends in Tin Pyrochlore Oxides. J. Solid State Chem. 1997, 130, 58–65. [Google Scholar] [CrossRef]
  19. Sakai, H.; Yoshimura, K.; Ohno, H.; Kato, H.; Kambe, S.; Walstedt, R.E.; Matsuda, T.D.; Haga, Y.; Onuki, Y. Superconductivity in a pyrochlore oxide, Cd2Re2O7. J. Phys. Condens. Matter 2001, 13, L785. [Google Scholar] [CrossRef]
  20. Dong, L.; Li, Y.; Devanathan, R.; Setyawan, W.; Gao, F. Low energy ion-solid interactions and chemistry effects in a series of pyrochlores. J. Am. Ceram. Soc. 2017, 100, 3132–3144. [Google Scholar] [CrossRef]
  21. Chartier, A.; Meis, C.; Crocombette, J.P.; Corrales, L.R.; Weber, W.J. Atomistic modeling of displacement cascades in (formula presented) pyrochlore. Phys. Rev. B-Condens. Matter Mater. Phys. 2003, 67, 174102. [Google Scholar] [CrossRef]
  22. Whittle, K.R.; Blackford, M.G.; Aughterson, R.D.; Lumpkin, G.R.; Zaluzec, N.J. Ion irradiation of novel yttrium/ytterbium-based pyrochlores: The effect of disorder. Acta Mater. 2011, 59, 7530–7537. [Google Scholar] [CrossRef]
  23. Archer, A.; Foxhall, H.R.; Allan, N.L.; Shearer, J.R.W.; Gunn, D.S.D.; Harding, J.H.; Todorov, I.T.; Travis, K.P.; Purton, J.A. Multiple cascade radiation damage simulations of pyrochlore. Mol. Simul. 2021, 47, 273–283. [Google Scholar] [CrossRef]
  24. Subramani, T.; Voskanyan, A.; Jayanthi, K.; Abramchuk, M.; Navrotsky, A. A Comparison of Order-Disorder in Several Families of Cubic Oxides. Front. Chem. 2021, 9, 719169. [Google Scholar] [CrossRef]
  25. Annamareddy, A.; Eapen, J.; Eapen, J. Decoding ionic conductivity and reordering in cation-disordered pyrochlores Subject Areas: Author for correspondence. Philos. Trans. A 2021, 379, 20190452. [Google Scholar] [CrossRef] [PubMed]
  26. Teng, Z.; Tan, Y.; Zhang, H. High-entropy pyrochlore a2b2o7 with both heavy and light rare-earth elements at the a site. Materials 2022, 15, 129. [Google Scholar] [CrossRef]
  27. Xiang, H.; Xing, Y.; Dai, F.; Wang, H.; Su, L.; Miao, L.; Zhang, G.; Wang, Y.; Qi, X.; Yao, L.; et al. High-Entropy Ceramics: Present Status, Challenges, and a Look Forward. J. Adv. Ceram. 2021, 10, 385–441, ISBN 4014502104. [Google Scholar] [CrossRef]
  28. Widom, M. Modeling the structure and thermodynamics of high-entropy alloys. J. Mater. Res. 2018, 33, 2881–2898. [Google Scholar] [CrossRef]
  29. Zhi, Q.; Tan, X.; Liu, Z.; Liu, Y.; Zhang, Q.; Chen, Y.; Li, M. Effect of Zr content on microstructure and mechanical properties of lightweight Al2NbTi3V2Zrx high entropy alloy. Micron 2021, 144, 103031. [Google Scholar] [CrossRef]
  30. Ji, G.; Zhou, Z.; Meng, F.; Yang, X.; Sheng, R.; Qiao, J.; Liaw, P.K.; Li, M.; Jiang, L.; Chen, S.; et al. Effect of Zr addition on the local structure and mechanical properties of Ti–Ta–Nb–Zr refractory high-entropy alloys. J. Mater. Res. Technol. 2022, 19, 4428–4438. [Google Scholar] [CrossRef]
  31. Brixner, L. Preparation and Properties of the Ln2Ti20-Type Rare Earth Titanates. Inorg. Chem. 1964, 3, 1065–1067. [Google Scholar] [CrossRef]
  32. Devanathan, R.; Weber, W.J. Insights into the radiation response of pyrochlores from calculations of threshold displacement events. J. Appl. Phys. 2005, 98, 086110. [Google Scholar] [CrossRef]
  33. Zheng, S.; Wang, S. First-principles design of refractory high entropy alloy VMoNbTaW. Entropy 2018, 20, 965. [Google Scholar] [CrossRef] [PubMed]
  34. Lipkina, K.; Savchenko, A.; Skupov, M.; Glushenkov, A.; Vatulin, A.; Uferov, O.; Ivanov, Y.; Kulakov, G.; Ershov, S.; Maranchak, S.; et al. Metallic inert matrix fuel concept for minor actinides incineration to achieve ultra-high burn-up. J. Nucl. Mater. 2014, 452, 378–381. [Google Scholar] [CrossRef]
  35. Mustafa Azeem, M.; Li, Z.; Wang, Q.; Zubair, M. Molecular dynamics studies and irradiation effects in ODSS alloys. Int. J. Nucl. Energy Sci. Technol. 2018, 12, 381–399. [Google Scholar] [CrossRef]
  36. Azeem, M.M.; Wang, Q.; Li, Z.; Zhang, Y. Dislocation-oxide interaction in Y2O3 embedded Fe: A molecular dynamics simulation study. Nucl. Eng. Technol. 2020, 52, 337–343. [Google Scholar] [CrossRef]
  37. Azeem, M.M.; Wang, Q.; Zubair, M. Atomistic simulations of nanoindentation response of irradiation defects in iron. Sains Malays. 2019, 48, 2029–2039. [Google Scholar] [CrossRef]
  38. Azeem, M.M.; Wang, Q.; Zhang, Y.; Liu, S.; Zubair, M. Effect of Grain Boundary on Diffusion of P in Alpha-Fe: A Molecular Dynamics Study. Front. Phys. 2019, 7, 97. [Google Scholar] [CrossRef]
  39. Mustafa Azeem, M.; Abd El Gawad, K.; Zubair, M.; Ibrahim, S.A.; Ado, M.; Mehdi, G. Radiation damage effects in oxide dispersion strengthened steel alloys. In Proceedings of the International Conference on Nuclear Engineering, Tsukuba, Japan, 19–24 May 2019. [Google Scholar]
  40. Mustafa, A.M.; Li, Z.; Shao, L. Molecular Dynamics Simulations of Damage Cascades Creation in Oxide-Particle-Embedded Fe. In Proceedings of the 25th International Conference on Nuclear Engineering, Shanghai, China, 2–6 July 2017; Volume 5, pp. 1–5. [Google Scholar]
  41. Azeem, M.M.; Li, Z.; Wang, Q.; Amjad, Q.M.N.; Zubair, M.; Ahmed, O.M.H. Classical molecular dynamics study for defect sink behavior in oxide dispersed strengthened alloys. In Proceedings of the Proceedings of 2018 15th International Bhurban Conference on Applied Sciences and Technology, Islamabad, Pakistan, 9–13 January 2018; pp. 12–15. [Google Scholar]
  42. Talanov, M.V.; Talanov, V.M. Structural Diversity of Ordered Pyrochlores. Chem. Mater. 2021, 33, 2706–2725. [Google Scholar] [CrossRef]
  43. Connor, T.; Cheong, O.; Bornhake, T.; Shad, A.C.; Tesch, R.; Sun, M.; He, Z.; Bukayemsky, A.; Vinograd, V.L. Pyrochlore Compounds From Atomistic Simulations. Front. Chem. 2021, 9, 733321. [Google Scholar] [CrossRef]
  44. Ewing, R.C.; Weber, W.J.; Lian, J. Nuclear waste disposal-pyrochlore (A 2B 2O 7): Nuclear waste form for the immobilization of plutonium and “minor” actinides. J. Appl. Phys. 2004, 95, 5949–5971. [Google Scholar] [CrossRef]
  45. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  46. Wiberg, K.B. A Scheme for Strain Energy Minimization. Application to the Cycloalkanes. J. Am. Chem. Soc. 1965, 87, 1070–1078. [Google Scholar] [CrossRef]
  47. Kocevski, V.; Pilania, G.; Uberuaga, B.P. Modeling Disorder in Pyrochlores and Other Anion-Deficient Fluorite Structural Derivative Oxides. Front. Chem. 2021, 9, 712543. [Google Scholar] [CrossRef]
  48. Chakoumakos, B.C. Systematics of the pyrochlore structure type, ideal A2B2X6Y. J. Solid State Chem. 1984, 53, 120–129. [Google Scholar] [CrossRef]
  49. Sickafus, K.E.; Minervini, L.; Grimes, R.W.; Valdez, J.A.; Ishimaru, M.; Li, F.; McClellan, K.J.; Hartmann, T. Radiation tolerance of complex oxides. Science 2000, 289, 748–751. [Google Scholar] [CrossRef]
  50. Plimpton, S.; Hendrickson, B. A New Parallel Method for Molecular Dynamics Simulation of Macromolecular Systems. J. Comput. Chem. 1996, 17, 326–337. [Google Scholar] [CrossRef]
  51. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO the Open Visualization Tool. Model. Simul. Mater. Sci. Eng. 2010, 18, 015012. [Google Scholar] [CrossRef]
  52. Dholakia, M.; Chandra, S.; Jaya, S.M. A comparative study of topology and local disorder in Y2O3, Y2TiO5, and Y2Ti2O7 crystals Manan. J. Alloys Compd. 2018, 739, 1037–1047. [Google Scholar] [CrossRef]
  53. Minervini, L.; Zacate, M.O.; Grimes, R.W. Defect cluster formation in M2O3-doped CeO2. Solid State Ion. 1999, 116, 339–349. [Google Scholar] [CrossRef]
  54. Minervini, L.; Grimes, R.W.; Tabira, Y.; Withers, R.L.; Sickafus, K.E. The oxygen positional parameter in pyrochlores and its dependence on disorder. Philos. Mag. A Phys. Condens. Matter Struct. Defects Mech. Prop. 2002, 82, 123–135. [Google Scholar] [CrossRef]
  55. Balducci, G.; Kašpar, J.; Fornasiero, P.; Graziani, M.; Islam, M.S. Surface and reduction energetics of the CeO2-ZrO2 catalysts. J. Phys. Chem. B 1998, 102, 557–561. [Google Scholar] [CrossRef]
  56. Luty, B.A.; Davis, M.E.; Tironi, I.G.; Van Gunsteren, W.F. A comparison of particle-particle, particle-mesh and ewald methods for calculating electrostatic interactions in periodic molecular systems. Mol. Simul. 1994, 14, 11–20. [Google Scholar] [CrossRef]
  57. Ziegler, J.F.; Biersack, J.P. The Stopping and Range of Ions in Matter BT—Treatise on Heavy-Ion Science: Volume 6: Astrophysics, Chemistry, and Condensed Matter; Bromley, D.A., Ed.; Springer: Boston, MA, USA, 1985; pp. 93–129. ISBN 978-1-4615-8103-1. [Google Scholar]
  58. Cooper, M.W.D.; Rushton, M.J.D.; Grimes, R.W. A many-body potential approach to modelling the thermomechanical properties of actinide oxides. J. Phys. Condens. Matter 2014, 26, 105401. [Google Scholar] [CrossRef] [PubMed]
  59. Ado, M.; Wang, Q.; Ibrahim, S.A.; Adede, S.O. Effect of radiation and substitution of Ce4+ at Zr site in Y4Zr3O12 using collision cascades: A molecular dynamics simulation study. J. Nucl. Sci. Technol. 2022, 60, 415–424. [Google Scholar] [CrossRef]
  60. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef]
  61. Anupama, M.K.; Rudraswamy, B.; Dhananjaya, N. Influence of trivalent (Gd3+, Cr3+) ion substitution on the structural and magnetic properties of Ni0.4Zn0.6Cr0.5GdxFe1.5−xO4 ferrite nanoparticles. Int. J. Nanotechnol. 2017, 14, 867–874. [Google Scholar] [CrossRef]
  62. Nakamoto, R.; Xu, B.; Xu, C.; Xu, H.; Bellaiche, L. Properties of rare-earth iron garnets from first principles. Phys. Rev. B 2017, 95, 024434. [Google Scholar] [CrossRef]
  63. Patel, K.; Prosandeev, S.; Yang, Y.; Xu, B.; Íñiguez, J.; Bellaiche, L. Atomistic mechanism leading to complex antiferroelectric and incommensurate perovskites. Phys. Rev. B 2016, 94, 054107. [Google Scholar] [CrossRef]
  64. Lee, W.; Chen, S.Y.; Tseng, E.; Gloter, A.; Chen, C.L. Study of defect structure in ferromagnetic nanocrystalline CeO2: Effect of ionic radius. J. Phys. Chem. C 2016, 120, 14874–14882. [Google Scholar] [CrossRef]
  65. Martel, L.; Naji, M.; Popa, K.; Vigier, J.F.; Somers, J. Fingerprint of local disorder in long range ordered isometric pyrochlores. Sci. Rep. 2017, 7, 12269. [Google Scholar] [CrossRef] [PubMed]
  66. Anupama, M.K.; Rudraswamy, B. Effect of Gd3+-Cr3+ ion substitution on the structural, electrical and magnetic properties of Ni-Zn ferrite nanoparticles. IOP Conf. Ser. Mater. Sci. Eng. 2016, 149, 012194. [Google Scholar] [CrossRef]
  67. Dong, L.; Li, Y.; Devanathan, R.; Gao, F. Molecular dynamics simulation of the structural, elastic, and thermal properties of pyrochlores. RSC Adv. 2016, 6, 41410–41419. [Google Scholar] [CrossRef]
  68. Dong, L.; Setyawan, W.; Li, Y.; Devanathan, R.; Gao, F. Molecular dynamics simulation of low-energy recoil events in titanate pyrochlores. RSC Adv. 2017, 7, 35403–35410. [Google Scholar] [CrossRef]
  69. El-Atwani, O.; Li, N.; Li, M.; Devaraj, A.; Baldwin, J.K.S.; Schneider, M.M.; Sobieraj, D.; Wróbel, J.S.; Nguyen-Manh, D.; Maloy, S.A.; et al. Outstanding radiation resistance of tungsten-based high-entropy alloys. Sci. Adv. 2019, 5, eaav2002. [Google Scholar] [CrossRef] [PubMed]
  70. Pickering, E.J.; Carruthers, A.W.; Barron, P.J.; Middleburgh, S.C.; Armstrong, D.E.J.; Gandy, A.S. High-entropy alloys for advanced nuclear applications. Entropy 2021, 23, 98. [Google Scholar] [CrossRef]
  71. Mandal, B.P.; Bhattacharyya, K.; Zain, J.H.; Sudarsan, V.; Nigam, S.; Nayak, C.; Tyagi, A.K. Variation of structural disorder in Zr substituted Y2Sn2O7: Its impact on photocatalysis. J. Solid State Chem. 2021, 303, 122472. [Google Scholar] [CrossRef]
  72. Cai, L.; Arias, A.L.; Nino, J.C. The tolerance factors of the pyrochlore crystal structure. J. Mater. Chem. 2011, 21, 3611–3618. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the computational model and a unit cell of pyrochlore, (a) Snapshot of a computational unit cell model, Yb 16 Ti 16 O 56 , (b) Generic representation of a unit cell of pyrochlore A 2 B 2 O 7 where A and B are transition metals [43], (c) labeled structure of Yb 16 Ti 16 O 56 , (d) Yb 8 Y 8 Ti 8 Z r 8 O 56 , (e) Yb 8 Gd 8 Z r 8 Ti 8 O 56 , and (f) Yb 4 Y 4 Gd 4 Eu 4 Z r 8 Ti 8 O 56 equivalent of Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 ZrTi O 7 . The subscript indicates the number of particular elements in the unit cell.
Figure 1. Schematic representation of the computational model and a unit cell of pyrochlore, (a) Snapshot of a computational unit cell model, Yb 16 Ti 16 O 56 , (b) Generic representation of a unit cell of pyrochlore A 2 B 2 O 7 where A and B are transition metals [43], (c) labeled structure of Yb 16 Ti 16 O 56 , (d) Yb 8 Y 8 Ti 8 Z r 8 O 56 , (e) Yb 8 Gd 8 Z r 8 Ti 8 O 56 , and (f) Yb 4 Y 4 Gd 4 Eu 4 Z r 8 Ti 8 O 56 equivalent of Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 ZrTi O 7 . The subscript indicates the number of particular elements in the unit cell.
Jcs 07 00413 g001
Figure 2. Potential energy (Vij) as a function of distance (rij) between pairs fitted by splined ZBL and Buckingham for (a) Yb-O, (b) O-O, and (c) Ti-O.
Figure 2. Potential energy (Vij) as a function of distance (rij) between pairs fitted by splined ZBL and Buckingham for (a) Yb-O, (b) O-O, and (c) Ti-O.
Jcs 07 00413 g002
Figure 3. Relationship between system energy as a function of LC in Yb2Ti2O7.
Figure 3. Relationship between system energy as a function of LC in Yb2Ti2O7.
Jcs 07 00413 g003
Figure 4. XRD pattern of Yb 2 Ti 2 O 7 , YbYTiZrO7, YbGdTiZrO7, and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 .
Figure 4. XRD pattern of Yb 2 Ti 2 O 7 , YbYTiZrO7, YbGdTiZrO7, and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 .
Jcs 07 00413 g004
Figure 5. Ed as a function of incident PKA for Yb 2 Ti 2 O 7   along with bombardment directions.
Figure 5. Ed as a function of incident PKA for Yb 2 Ti 2 O 7   along with bombardment directions.
Jcs 07 00413 g005
Figure 6. Number of defects ( N d e f e c t s ) as a function of time under different energy PKA damage of Yb 2 Ti 2 O 7 bombarded with Yb ions of different energies.
Figure 6. Number of defects ( N d e f e c t s ) as a function of time under different energy PKA damage of Yb 2 Ti 2 O 7 bombarded with Yb ions of different energies.
Jcs 07 00413 g006
Figure 7. Displacement cascades in Yb 2 Ti 2 O 7 bombarded with 5 keV Yb at different timesteps. Blue colors are vacancies whereas purple is interstitials.
Figure 7. Displacement cascades in Yb 2 Ti 2 O 7 bombarded with 5 keV Yb at different timesteps. Blue colors are vacancies whereas purple is interstitials.
Jcs 07 00413 g007
Figure 8. Number of surviving defects as a function of (a) Time and (b) inclination angle of incident 5 keV Yb in Yb 2 Ti 2 O 7 .
Figure 8. Number of surviving defects as a function of (a) Time and (b) inclination angle of incident 5 keV Yb in Yb 2 Ti 2 O 7 .
Jcs 07 00413 g008
Figure 9. The number of defects of different types of atoms at an inclination of (a) 55° and (b) 85°.
Figure 9. The number of defects of different types of atoms at an inclination of (a) 55° and (b) 85°.
Jcs 07 00413 g009
Figure 10. Evolution of Displacement cascades in Yb 2 Ti 2 O 7   YbYTiZr O 7 , YbGdTiZr O 7 , and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 at timesteps 0.3 ps and 0.9 ps. The number of surviving defects considerably reduced with the increase of alloying constituent.
Figure 10. Evolution of Displacement cascades in Yb 2 Ti 2 O 7   YbYTiZr O 7 , YbGdTiZr O 7 , and Yb 0.5 Y 0.5 Eu 0.5 Gd 0.5 TiZr O 7 at timesteps 0.3 ps and 0.9 ps. The number of surviving defects considerably reduced with the increase of alloying constituent.
Jcs 07 00413 g010
Figure 11. Defect evolution as function time in undoped Yb 2 Ti 2 O 7 and HEPy doped with Y, Gd, and YGdEu at the A sites.
Figure 11. Defect evolution as function time in undoped Yb 2 Ti 2 O 7 and HEPy doped with Y, Gd, and YGdEu at the A sites.
Jcs 07 00413 g011
Table 1. Buckingham potential and splined function parameters [20,53].
Table 1. Buckingham potential and splined function parameters [20,53].
Pair (s) A i j (eV) ρ   (Å) C xy   ( e V Å 6 ) c 1   ( e V Å 1 ) c 2   ( e V Å 2 ) c 3   ( e V Å 3 ) c 4   ( e V Å 4 ) r 1   (Å) r 2   (Å)
Yb-O1649.800.338616.5710.40−8.383.08−0.550.61.3
Ti-O2131.400.30380.40010.62−14.80−14.74−6.280.32.0
O-O9547.960.219232.009.355−10.716.23−1.680.82.1
Table 2. Calculation for LC and rionic by changing A-site cations while keeping B site as Ti4+.
Table 2. Calculation for LC and rionic by changing A-site cations while keeping B site as Ti4+.
Ion (s)rionic (Å)ao (Å)
Lu 3 + 0.977, 0.861 [18], 1.032 [62]9.988, 10.291 [18], 10.011 [61]
Yb 3 + 0.985, 0.87 [18], 1.042 [62]10.007, 10.34 [18], 10.03 [63]
Er 3 + 1.004, 0.89 [18], 1.004 [62]10.055, 10.35 [18], 10.069 [63]
Y 3 + 1.019, 0.90 [18]10.056, 10.37 [18], 10.080 [61]
Gd 3 + 1.053, 1.107 [61]10.136, 8.34 [61], 10.171 [61]
Eu 3 + 1.066, 1.120 [62]10.159, 10.192 [63]
Sm 3 + 1.062, 1.079 [62]10.184, 10.211 [63]
Nd 3 + 1.109, 0.98 [18], 1.163 [62]10.231, 10.56 [18]
Ce 3 + 1.143, 1.196 [62], 1.14 [63]10.343, 10.65 [18], 10.6 [63]
La 3 + 1.164, 1.03 [18], 1.16 [64]10.324, 10.7026 [18]
Table 3. Ed as a function of different incident PKA’s in HEPy models.
Table 3. Ed as a function of different incident PKA’s in HEPy models.
Types of PKAPKA DirectionEd (eV)Types of PKAPKA DirectionEd (eV)
Yb[100]60Ti[100]1182
[110]60[110]1177
[111]66[111]625
Y[100]146Zr[100]430
[110]146[110]431
[111]174[111]426
Gd[100]93O[100]1769
[110]93[110]1521
[111]84[111]343
Eu[100]85
[110]85
[111]128
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

Azeem, M.M.; Wang, Q. Atomic Insights into the Structural Properties and Displacement Cascades in Ytterbium Titanate Pyrochlore (Yb2Ti2O7) and High-Entropy Pyrochlores. J. Compos. Sci. 2023, 7, 413. https://doi.org/10.3390/jcs7100413

AMA Style

Azeem MM, Wang Q. Atomic Insights into the Structural Properties and Displacement Cascades in Ytterbium Titanate Pyrochlore (Yb2Ti2O7) and High-Entropy Pyrochlores. Journal of Composites Science. 2023; 7(10):413. https://doi.org/10.3390/jcs7100413

Chicago/Turabian Style

Azeem, M. Mustafa, and Qingyu Wang. 2023. "Atomic Insights into the Structural Properties and Displacement Cascades in Ytterbium Titanate Pyrochlore (Yb2Ti2O7) and High-Entropy Pyrochlores" Journal of Composites Science 7, no. 10: 413. https://doi.org/10.3390/jcs7100413

Article Metrics

Back to TopTop