Next Article in Journal
Microstructure and Mechanical Properties of TiAl Matrix Composites Reinforced by Carbides
Next Article in Special Issue
The Role of Grain Boundaries in the Corrosion Process of Fe Surface: Insights from ReaxFF Molecular Dynamic Simulations
Previous Article in Journal
Effect of Thermomechanical Treatments on Microstructure, Phase Composition, Vickers Microhardness, and Young’s Modulus of Ti-xNb-5Mo Alloys for Biomedical Applications
Previous Article in Special Issue
Molecular Dynamics Simulations of Xe Behaviors at the Grain Boundary in UO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Point Defects on the Stable Occupation, Diffusion and Nucleation of Xe and Kr in UO2

1
College of Materials Science and Engineering, Hunan University, Changsha 410082, China
2
The First Sub-Institute, Nuclear Power Institute of China, Chengdu 610041, China
3
School of Physics and Electronics, Hunan University, Changsha 410082, China
*
Authors to whom correspondence should be addressed.
Metals 2022, 12(5), 789; https://doi.org/10.3390/met12050789
Submission received: 16 March 2022 / Revised: 25 April 2022 / Accepted: 2 May 2022 / Published: 4 May 2022
(This article belongs to the Special Issue Numerical Modeling of Materials under Extreme Conditions)

Abstract

:
Xe and Kr gases produced during the use of uranium dioxide (UO2)-fuelled reactors can easily form bubbles, resulting in fuel swelling or performance degradation. Therefore, it is important to understand the influence of point defects on the behaviour of Xe and Kr gases in UO2. In this work, the effects of point defects on the behavioural characteristics of Xe/Kr clusters in UO2 have been systematically studied using molecular dynamics. The results show that Xe and Kr clusters occupy vacancies as nucleation points by squeezing U atoms out of the lattice, and the existence of vacancies makes the clusters more stable. The diffusion of interstitial Xe/Kr atoms and clusters in UO2 is also investigated. It is found that the activation energy is ~2 eV and that the diffusion of the interstitial atoms is very difficult. Xe and Kr bubbles form at high temperatures. The more interstitial Xe/Kr atoms or vacancies in the system, the easier the clusters form.

Graphical Abstract

1. Introduction

With the increasing consumption of energy on Earth, the development of nuclear energy has attracted considerable research attention from all walks of life. Nuclear fission is a critical way to generate clean energy, and uranium dioxide (UO2) is the standard nuclear fuel used in pressurised water reactors [1]. Fission gases, such as Xe and Kr, are among the essential fission products in UO2 fuel, which can exacerbate the fuel swelling, thereby leading to the interaction between the fuel and the cladding [2,3,4,5,6]. As the fission products deposit energy in the surrounding material, point defects that control the microstructural evolution of the fuel occur. The point defects that survive the initial damage from irradiation in nuclear fuel form extended defects, such as vacancy clusters, dislocation loops, and voids [7]. Numerous experimental and modeling studies have been conducted to improve the understanding of the behaviour of Xe and Kr gases [7,8,9,10,11,12,13,14].
Among all volatile fission products, Xe and Kr have the highest concentration and are mainly studied herein. Previous literature mainly focused on stable configurations with a constant Xe-vacancy ratio; for example, Moore et al. [15] found that clusters of Xe atoms are formed by single Xe atoms occupying Schottky positions, which is caused by the supersaturation of Schottky vacancies in UO2. Due to the complexity of the behavioural characteristics of UO2 fuel materials and Xe bubbles, it is difficult to determine the behavioural mechanism of Xe gases. Consequently, several separation effect experiments have been proposed to simplify complex material systems by describing the physical processes of one or more fission gases to elucidate the underlying behavioural mechanisms. Thus, Zhang et al. [16] briefly explained the mechanism of UO2 by simulating molybdenum. They simulated the stable configuration of Mo by adding Xe atoms and found that Mo was most stable when the Xe-to-vacancy ratio was unity. This study compares the stable occupation of Xe/Kr clusters in perfect and varying defect-containing systems.
The diffusion of the Xe atoms in bulk UO2 or Xe-vacancy clusters formed by Xe atoms in Schottky vacancies has been studied in previous literature, and even the self-diffusion behaviour of U and O in UO2 has been studied [17,18,19]. Yun et al. [20] have investigated the vacancy-assisted diffusion of Xe in UO2, and calculated the incorporation, binding, and migration energies. They found that the tri-vacancy is a significant diffusion pathway of Xe in UO2. Lawrence [21] discussed the uncertainty in fission gas diffusion coefficients as a function of temperature. Higher activation energies in computing diffusion are usually compensated by higher pre-exponential factors [22]. The diffusion of cations in UO2 and other related compounds is very slow, at <1015 or <1017 cm2/s [23,24], even at high temperatures from 1800 to 2000 K, which is one of the highest temperatures achieved in crystal correlation experiments. One of the most commonly used models in fuel performance codes was published by Massih and Forsberg [25,26,27]. Turnbull et al. [7] analysed this model and other models, and then computed the bulk fission gas diffusion rates, which capture both intrinsic and radiation-enhanced diffusion. This model divides the diffusivity into three regimes. Davies et al. [28] experimentally studied the diffusivity of UO2 at high temperatures (D1, T > 1650 K) and concluded that its activation energy (Ea) and pre-exponential factor (D0) were 3.04 eV and 7.6 × 10−10 m2/s, respectively. This study provides significant guidance for the subsequent diffusion studies of UO2 by many researchers [29,30,31]. The in-pile diffusion coefficient of UO2 is close to the intrinsic diffusion coefficient, so it is considered that the radiation-enhanced diffusion coefficient has high uncertainty [32]. However, due to the complex diffusion of Xe at interstitial sites, the diffusion of interstitial clusters has not presently been well described. Therefore, it is necessary to explore the diffusion behaviour of Xe/Kr clusters at octahedral interstitial sites in UO2 and to explain the relationship between the interstitial and vacancy diffusion mechanisms.
There are two main nucleation mechanisms of fission gases, such as homogeneous and heterogeneous nucleation [33]. Transmission electron microscopy (TEM) images of UO2–irradiated bubbles show that they are characterised by their high density and small, almost uniform bubble size. Nelson [34] predicted that the nucleation density of bubbles was almost independent of irradiation temperature and fission rate. Evans [35] observed bubbles in Kr- and Xe-irradiated UO2 using TEM and found that the threshold temperature for bubble nucleation was in the range of 350 °C–500 °C. Michel et al. [36] found subnanometer Xe bubbles in polycrystalline UO2 under low flux irradiation at 600 °C. Previous studies focused on the vital role of temperature and irradiation dose in the nucleation and growth of bubbles in UO2, but there are few studies on defect concentration. Hence, studying the formation of Xe/Kr clusters in systems with different defect concentrations is important for subsequent nucleation studies.
Thus, the influence of defects in materials on Xe/Kr gas clusters is worth further investigation. Therefore, it is crucial to investigate the effect of point defects on Xe/Kr clusters in UO2 to understand the evolution of fuels.

2. Simulation Method

2.1. Interatomic Potential

The interatomic interaction potentials of UO2 have been widely reported previously. Among them, the potentials reported by Basak et al. [37], Morelon et al. [38], and Cooper et al. [39] are more commonly used. For the further addition of fission gas, Xe, based on UO2, and UO2–Xe interatomic interaction potentials have been mainly developed by Geng et al. [40], Chartier et al. [41], Thompson et al. [42], and Cooper et al. [43]. For the UO2–Kr system, only one interatomic interaction potential developed by Cooper et al. [43] can be presently used. These UO2–Xe and UO2–Kr potentials use the Xe and Kr potentials proposed by Tang and Toennies [44].
Herein, the UO2 potential reported by Cooper et al. [39] is adopted to describe the U–U, U–O, and O–O interactions. This potential reproduces a range of thermophysical properties, such as the lattice parameter, bulk modulus, enthalpy, and specific heat at temperatures between 300 and 3000 K, as well as some defect properties in UO2. In addition, this potential’s bulk modulus and elastic constant are more accurate and in accordance with experimental values [45]. The Xe–Xe interaction is described by the Tang–Toennies potential [44]. Further, the interactions of Xe–U and Xe–O have been described by Thompson et al. [42], and are very flexible and can be applied to a wide variety of potential forms and materials systems, including metals and EAM potentials. For the UO2–Kr system, the interatomic interaction potential developed by Cooper et al. [43] is adopted.

2.2. MD Simulation Setup

An MD simulation programme, LAMMPS (7Aug19 version) [46], is employed herein for all simulations. Images of atomic configurations were produced with the visualisation tool OVITO (3.5.0 version, Darmstadt, Germany) [46]. The Wigner–Seitz (W–S) cell method determined the type and position of the interstitial atoms or vacancies [47,48]. The time step was set as 0.001 ps for all MD simulations. The temperature was controlled via a Nose/Hoover temperature thermostat, and the periodic boundary condition was used. For static relaxation, the energy minimisation was performed, and the minimisation algorithm was set as the conjugate gradient method (cg). The specific simulation processes for different behaviours differ, and the details of the different simulations are described as follows.

2.2.1. Stable Occupation of Xe/Kr Cluster in UO2-Containing Point Defects

In addition to studying the stable occupation of Xe(Kr) clusters in defect-free bulk UO2, the influence of different defects in UO2 on the stable occupation of Xe(Kr) clusters was studied, such as U, O, UO double, Schottky, and double Schottky vacancies. A cubic box of 25 a0 × 25 a0 × 25 a0 (a0 is the lattice constant of the UO2 fluorite structure at 0 K) containing 187,500 atoms was used. MD simulation was performed after generating each configuration to equilibrate the system. After energy minimisation, the first atom was introduced into the system. The site with the lowest energy formation was searched through energy minimisation again to determine the stable site of the first atom. Afterwards, the second atom was introduced around the first atom, and the process was repeated to obtain a stable space for the two atoms. Further, the same process was performed until stable positions for six atoms were found successfully.
The formation energy of a Xe/Kr cluster in defect-free UO2 bulk is defined as follows:
E N   Xe / Kr f = E N     Xe / Kr I n t E P N E Xe / Kr
where E N   Xe / Kr I n t   is the energy of the UO2 system containing the Xe (or Kr) cluster, E P is the energy of perfect UO2, N is the number of Xe (or Kr) atoms, and E Xe / Kr is the energy of a single-isolated Xe (or Kr) atom (this value is zero for the interatomic potential under consideration).
The formation energy of a Xe/Kr cluster in defective UO2 is defined as
E N   Xe / Kr f = E N     Xe / Kr I n t E V D m N E Xe / Kr
where E N   Xe / Kr I n t   is the total energy of the system with the Xe (or Kr) cluster added on VD (represent different vacancy-type defects), E V D m is the total energy of systems containing the VD, m is the number of vacancies for VD, N is the number of Xe (or Kr) atoms, and E Xe / Kr is the energy of a single-isolated Xe (or Kr) atom.
The binding energy of an additional X (X = Xe or Kr) atom to a VD-X cluster in UO2 is defined as follows:
E b ( X + VD - X   cluster ) = E f ( X ) + E f ( VD - X   cluster ) E f ( X + VD - X   cluster )
where Ef(X) is the formation energy of a Xe (or a Kr) atom located on the most stable interstitial site in bulk UO2.

2.2.2. Diffusion of Xe/Kr Cluster in UO2

Generally, diffusion in solids occurs with point defects [18]. The point defect concentration is thermally activated, and it increases as the temperature increases. Migration is also a thermally activated process, accelerated by an increasing temperatures. Hence, the diffusion coefficients and diffusion energy barriers of small Xe and Kr clusters (number of atoms < 6) in UO2 are calculated by the mean square displacement (MSD) method, which can intuitively reflect the strength of the self-diffusion ability of particles. A box of 10 a0 × 10 a0 × 10 a0 (a0 is the lattice constant of the UO2 fluorite structure at different temperatures from 1800 to 2300 K) containing 12,000 atoms was used. The total simulation time is up to 5 ns, with a timestep of 1 fs.
The Arrhenius’s equation can express the temperature dependence of the diffusivity,
D = D 0 exp ( E a k B T )
where D is the diffusion coefficient, E a is the diffusion barrier, T is the temperature, D 0 is a pre-diffusion factor, and k B is the Boltzmann constant. Taking the logarithms of both sides of the above equation give
  ln D = ln D 0 E a k B T
Therefore, if the ln   D at different simulated temperatures is obtained, E a can be obtained by linear fitting, whereas the diffusion coefficient D at different temperatures can be obtained by the MSD method:
D T = ( M S D ) T 2 d t = < Δ r ( t ) T 2 > 2 d t
In the simulation process, a long simulation time and short coordinate position output intervals are used to obtain the atomic coordinate information. The results are obtained by averaging the MSD trajectory segmentation severally.

2.2.3. Nucleation of Xe/Kr Cluster in UO2-Containing Point Defects

MD simulation was performed to simulate the nucleation process of Xe bubbles. Five systems with sizes of 10 a0 × 10 a0 × 10 a0 (a0 is the lattice constant of UO2 fluorite structure at the temperature of 2500 K) were studied, which are mainly perfect bulk without defects, with 1% vacancies concentration, with 2% vacancies concentration, with 1% interstitial concentration, and with 2% interstitial concentration. Then, 2% and 5% concentration of Xe/Kr atoms were added to the different systems, and the Xe/Kr atoms were randomly and uniformly located at octahedral interstitial sites. The specific simulation process is as follows.
After static relaxation of the initial system relaxation of the above configuration with a high temperature, the temperature was raised to 2500 K to accelerate the diffusion of Xe/Kr atoms. Then, the simulation was conducted in the NVT ensemble for 20 ns. Finally, the obtained configuration was subjected to static relaxation to ensure a stable configuration.

3. Results and Discussion

3.1. Stable Occupation of Xe/Kr Cluster in UO2-Containing Point Defects

By studying the position and energy changes of Xe/Kr atoms in UO2, it is found that, after adding Xe/Kr atoms into UO2, the Xe/Kr atoms move to the octahedral interstitial sites after structural optimization. Figure 1a shows that, when a Xe atom is randomly inserted into the bulk UO2, it moves to the nearest octahedral site. The energy is lowest at the octahedral interstitial site, with a formation energy of 9.79 eV. The phenomenon is the same for the Kr atom, but, because the Kr atom is smaller than the Xe atom, its formation energy is smaller at 8.45 eV. When two interstitial Xe or two interstitial Kr atoms are added into UO2, after relaxation, the energy of the two atoms forming a dimer is the lowest. The formation energies of Xe2 and Kr2 are 16.44 and 15.60 eV, respectively. The stable structure of three interstitial atoms is shown as an equilateral triangle. The formation energies of Xe3 and Kr3 are 22.49 and 21.53 eV, respectively. After the addition of the fourth Xe atom and complete relaxation (Figure 1b), the positions of the four atoms appear as triangular cones at different octahedral sites. The first lattice U atom is squeezed out of the cluster, and the Xe atoms cluster around the U vacancy. However, the interstitial U atom moves away from the cluster, and the formation energy of Xe4 is 27.92 eV. The Kr4 cluster slightly differs from Xe4. The four Kr atoms are located on different octahedral gaps, forming a planar quadrilateral. However, no interstitial atoms are squeezed out of clusters, and the formation energy of Kr4 is 27.38 eV. After inserting the fifth atom, the first lattice U atom in the Kr cluster is squeezed out. After adding the sixth atom, the second lattice U atom is squeezed out of the Xe cluster (Figure 1c). The formation energies of Xe5, Kr5, Xe6, and Kr6 are 32.19, 32.41, 39.79, and 37.21 eV, respectively.
Additionally, the stabilities of Xe/Kr atoms in configurations with different defects are compared. As shown in Figure 1d, in the configuration containing a single U vacancy, the addition of one Xe/Kr atom occupies the vacancy. When adding three or six Xe/Kr atoms in this system (Figure 1e,f), the atoms occupy the vacancy and distribute in the octahedral interstitial sites around the vacancy. The configuration of a single O vacancy is consistent with that of a single U vacancy. In the case of the UO double vacancy, one Xe/Kr atom was added to occupy the U vacancy. Two Xe/Kr atoms are evenly distributed into the central region of the two vacancy centres; when multiple atoms are added, they take the central region as the origin and occupy the surrounding octahedral interstitial sites. Figure 1g–i shows that, when there is a double Schottky vacancy, the Xe/Kr atom moves to the position near the central vacancy region. When more than one atom is present in the box, the Xe/Kr atoms are mainly distributed in the central vacancy region or the surrounding octahedral interstitial sites.
Figure 2 shows that, as the number of Xe/Kr atoms increases, the formation energy of the configuration with various defect types increases gradually. The formation energy of the O vacancy was 5 eV larger than that of the U vacancy on average at each stage. The Xe/Kr atom was more accessible to form in the U vacancy than in the O vacancy. Further, Figure 2 shows that the formation energy of Xe/Kr clusters in the six systems can be divided into three layers. The first layer contains the bulk UO2 and the system with O vacancy. They are characterized by no U vacancy, and the formation energy difference is very small. The second layer contains U, UO double, and Schottky vacancies, which contain only one U vacancy. The third layer contains double Schottky vacancies, which contain two U vacancies. Additionally, the volume of the U vacancy is much larger than that of the O vacancy, which provides more space for clusters. Thus, the stability of Xe/Kr clusters depends on the number of U vacancies. Figure 3 shows that, as the number of Xe/Kr atoms increases, the binding energy of Xe/Kr and cluster decreases and stabilises. The overall trends of adding Xe or Kr atoms are consistent, and the changes caused by different defects in the system are similar. The double Schottky configuration has a more vital ability to adsorb Xe atoms and weaken. When additional atoms are adsorbed to a certain extent, the adsorption capacities of all defective configurations tend to be the same, which mainly depends on the number of vacancy defects contained in the configurations.

3.2. Diffusion of Xe/Kr Cluster in UO2

Based on empirical potential calculations, the diffusivity of Xe/Kr clusters in bulk UO2 has been calculated. Since the first U interstitial atom was excited by adding four Xe atoms or five Kr atoms to the bulk UO2, we studied the Xe/Kr clusters with less than four atoms. Xe usually diffuses due to a vacancy-assisted mechanism. The diffusion of Xe at U, O, UO, one U, two O vacancies, and vacancy clusters (comprising two U vacancies and zero, one, or two O vacancies) has been studied in most studies. Earlier studies concluded that Xe atoms occupied trap sites that contained at least one uranium vacancy and, in many cases, one or two additional oxygen vacancies [49,50]. The conclusion showed that triple vacancy was the main diffusion pathway of Xe in UO2. Previous DFT data [51,52,53] have shown the activation energies of Xe from 2.87 to 3.95 eV and pre-diffusion factors from 5 × 104 m2/s to 2.9 × 1012 m2/s. Due to experimental factors, Lawrence et al. [21] found that the diffusion coefficients between different studies have many orders of magnitude. Herein, the interstitial diffusion mechanism of Xe/Kr was investigated. The simulation estimated the diffusion barrier of Xe atoms as 2.11 eV and the pre-diffusion factor index as 1.8 × 105 m2/s at temperatures between 1800 and 2300 K, and the simulation estimated the diffusion barrier of Kr atoms as 2.31 eV and the pre-diffusion factor index as 0.12 × 103 m2/s. Table 1 and Table 2 show the detailed data of the Xe/Kr atom and clusters. Torres et al. [54] calculated the migration energies of Xe/Kr in bulk UO2 by a direct mechanism, and the results were 4.09 and 4.72 eV, respectively.
The activation energy of the Xe cluster was ~2 eV, and the diffusion coefficient can be seen in Figure 4, which shows the diffusion coefficient for Xe and Kr clusters in bulk UO2. Figure 4 illustrates the difficulty of cluster diffusion. It is consistent with the data proposed by Davies et al., indicating that clusters are not easy to diffuse. By analysing the movement of the atoms during migration, we find that, when studying the diffusion of individual atoms, evidently, individual atoms are fast and have a wide range of motion. When studying clusters with two atoms, the atoms move mainly by the rotating bypass method. In the diffusion process, atom A was stapled at random, and then atom B rotated around atom A to find a stable position, and then spread over continuously. When there were three atoms in a cluster, a small cluster was formed with one of the atoms pinned together, and the remaining atoms rotated slightly, causing the whole cluster to move and spread out. These trajectories suggested that the diffusion of interstitial clusters is more complicated and may require more complex conditions. In fact, there are little data on experimental interstitial diffusion.

3.3. Nucleation of Xe/Kr Cluster in UO2-Containing Point Defects

Here, the Xe/Kr atoms cluster together at high temperatures. A similar phenomenon occurred while studying Mo. Zhang et al. [16] studied the clustering process of Xe atoms dispersed in Mo at high temperatures. They observed the formation of Xe bubbles when the concentration of Xe atoms exceeded the threshold concentration value. In this paper, we studied randomly distributed Xe/Kr atoms at octahedral interspaces in UO2. Figure 5 shows that, as the relaxation progresses to 5 ns, small Xe clusters form, and then the tiny clusters gradually grow larger by absorbing extra Xe atoms. The clusters are more evident and are larger, almost stable clusters at relaxation to 10 ns. To ensure the stability of clusters, we observed the clustering phenomenon until 20 ns, which was almost not very different from that at 10 ns. The simulation results were consistent with the growth model proposed by Turnbull [55]. The bubbles were heterogeneously nucleated in the wake of fission fragments. They grew by collecting gas by atomic diffusion for a time controlled by a resolution process.
In the system with the same defect concentration, the number of clusters formed increases as the interstitial Xe/Kr atomic concentration increases. Figure 6 shows that the number and size of clusters in the system with 5% interstitial Xe atoms added significantly exceeded those with 2% interstitial Xe atoms. The former are more likely to form larger clusters, with a considerable number of clusters over 50 atoms or even over 100 atoms in size. In comparison, the latter are mainly distributed in 2 to 50 atoms.
When there are equal interstitial atom concentrations, the system with more vacancies is more likely to form larger clusters. Similarly, the more interstitial atoms prearranged in the system, the more difficult it is for the Xe/Kr atoms to aggregate during relaxation. The system mainly forms many small clusters ranging in size from 2 to 10 atoms.
The W–S cell method was used to analyse the defect results of the five systems. Three systems were selected for a detailed demonstration, namely the defect-free system (bulk), the system with 1% vacancy concentration (1% vac), and the system with 2% vacancy concentration (2% vac). Figure 7 shows the final distribution of Xe atoms at 5% concentration and the distribution of vacancy atoms and interstitial atoms after defect analysis, respectively. The Xe/Kr atom cluster region overlaps with the position of the vacancy cluster. In other words, the nucleation of the Xe/Kr atom mainly occupied the vacancy position by squeezing out the atoms on the original lattice, which is consistent with the analysis of the stable occupation above. The Xe/Kr atoms squeeze out U atoms to form vacancies, and interstitial atoms moved away from clusters. In addition, Xe/vac is ~1.

4. Conclusions

In this paper, molecular dynamics simulations were used to study the stable occupancy of Xe/Kr clusters in defect-free configurations and configurations with five different defects. The results show that the system was energetically favourable when Xe/Kr clusters were trapped by vacancies, especially U vacancies, because they can provide a larger space, and the complex of Xe/Kr and vacancies were more stable. The diffusion of Xe/Kr atoms and small clusters at octahedral interstitial sites in bulk UO2 was also studied. The diffusion of Xe/Kr atoms in bulk UO2 was relatively complex. The activation energy of Xe/Kr small clusters is ~2 eV, and the Xe/Kr clusters are difficult to diffuse. The nucleation of Xe/Kr in systems containing different defect concentrations was also investigated. At 2500 K, the Xe/Kr atoms dispersed in the system gather into clusters after a 20 ns relaxation. A comparison of the different systems revealed that the vacancy-containing system is more likely to form large clusters than a system containing interstitial atoms.

Author Contributions

Conceptualization, W.H. and H.D.; methodology, L.W. (Li Wang), Z.W., Y.X., Y.C., Z.L., Q.W. and W.H.; validation, Y.X., Z.W., L.W. (Li Wang), Y.C., Z.L., Q.W. and L.W. (Lu Wu); formal analysis, L.W. (Li Wang), Z.W., Y.X., Y.C., Z.L. and Q.W.; investigation, L.W. (Li Wang); resources, Z.W., Z.L. and Q.W.; data curation, Y.C. and W.H.; writing—original draft, L.W. (Li Wang); writing—review & editing, L.W. (Li Wang), L.W. (Lu Wu) and H.D.; supervision, H.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matthews, J.R. Technological problems and the future of research on the basic properties of actinide oxides. J. Chem. Soc. Faraday Trans. 2 1987, 83, 1273–1285. [Google Scholar] [CrossRef]
  2. Matzke, H. Diffusion processes in nuclear fuels. In Diffusion Processes in Nuclear Materials; Agarwala, P., Ed.; Elsevier (North Holland Publishing): Amsterdam, The Netherlands, 1992. [Google Scholar]
  3. Rest, J.; Hofman, G. An alternative explanation for evidence that xenon depletion, pore formation, and grain subdivision begin at different local burnups. J. Nucl. Mater. 2000, 277, 231–238. [Google Scholar] [CrossRef]
  4. Matzke, H.J. Gas release mechanisms in UO2—a critical review. Radiat. Eff. 1980, 53, 219–242. [Google Scholar] [CrossRef]
  5. White, R.; Tucker, M. A new fission-gas release model. J. Nucl. Mater. 1983, 118, 1–38. [Google Scholar] [CrossRef]
  6. White, R. The development of grain-face porosity in irradiated oxide fuel. J. Nucl. Mater. 2004, 325, 61–77. [Google Scholar] [CrossRef]
  7. Turnbull, J.; Friskney, C.; Findlay, J.; Johnson, F.; Walter, A. The diffusion coefficients of gaseous and volatile species during the irradiation of uranium dioxide. J. Nucl. Mater. 1982, 107, 168–184. [Google Scholar] [CrossRef]
  8. He, L.; Pakarinen, J.; Kirk, M.; Gan, J.; Nelson, A.; Bai, X.; El-Azab, A.; Allen, T. Microstructure evolution in Xe-irradiated UO2 at room temperature. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2014, 330, 55–60. [Google Scholar] [CrossRef]
  9. Millett, P.C.; Tonks, M. Meso-scale modeling of the influence of intergranular gas bubbles on effective thermal conductivity. J. Nucl. Mater. 2011, 412, 281–286. [Google Scholar] [CrossRef]
  10. Chockalingam, K.; Millett, P.C.; Tonks, M. Effects of intergranular gas bubbles on thermal conductivity. J. Nucl. Mater. 2012, 430, 166–170. [Google Scholar] [CrossRef]
  11. Lucuta, P.; Matzke, H.; Hastings, I. A pragmatic approach to modelling thermal conductivity of irradiated UO2 fuel: Review and recommendations. J. Nucl. Mater. 1996, 232, 166–180. [Google Scholar] [CrossRef]
  12. Hoh, A.; Matzke, H. Fission-enhanced self-diffusion of uranium in UO2 and UC. J. Nucl. Mater. 1973, 48, 157–164. [Google Scholar] [CrossRef]
  13. Jackson, R.; Catlow, C. Trapping and solution of fission Xe in UO2.: Part 1. Single gas atoms and solution from underpressurized bubbles. J. Nucl. Mater. 1985, 127, 161–166. [Google Scholar] [CrossRef]
  14. Djourelov, N.; Marchand, B.; Marinov, H.; Moncoffre, N.; Pipon, Y.; Nédélec, P.; Toulhoat, N.; Sillou, D. Variable energy positron beam study of Xe-implanted uranium oxide. J. Nucl. Mater. 2013, 432, 287–293. [Google Scholar] [CrossRef]
  15. Moore, E.; Corrales, L.R.; Desai, T.; Devanathan, R. Molecular dynamics simulation of Xe bubble nucleation in nanocrystalline UO2 nuclear fuel. J. Nucl. Mater. 2011, 419, 140–144. [Google Scholar] [CrossRef]
  16. Zhang, W.; Yun, D.; Liu, W. Xenon Diffusion Mechanism and Xenon Bubble Nucleation and Growth Behaviors in Molybdenum via Molecular Dynamics Simulations. Materials 2019, 12, 2354. [Google Scholar] [CrossRef] [Green Version]
  17. Govers, K.; Verwerft, M. Classical molecular dynamics investigation of microstructure evolution and grain boundary diffusion in nano-polycrystalline UO2. J. Nucl. Mater. 2013, 438, 134–143. [Google Scholar] [CrossRef]
  18. Vincent-Aublant, E.; Delaye, J.-M.; Van Brutzel, L. Self-diffusion near symmetrical tilt grain boundaries in UO2 matrix: A molecular dynamics simulation study. J. Nucl. Mater. 2009, 392, 114–120. [Google Scholar] [CrossRef]
  19. Murphy, S.T.; Jay, E.E.; Grimes, R.W. Pipe diffusion at dislocations in UO2. J. Nucl. Mater. 2014, 447, 143–149. [Google Scholar] [CrossRef]
  20. Yun, Y.; Kim, H.; Kim, H.; Park, K. Atomic diffusion mechanism of Xe in UO2. J. Nucl. Mater. 2008, 378, 40–44. [Google Scholar] [CrossRef]
  21. Lawrence, G. A review of the diffusion coefficient of fission-product rare gases in uranium dioxide. J. Nucl. Mater. 1978, 71, 195–218. [Google Scholar] [CrossRef]
  22. Andersson, D.; Garcia, P.; Liu, X.-Y.; Pastore, G.; Tonks, M.; Millett, P.; Dorado, B.; Gaston, D.; Andrs, D.; Williamson, R.; et al. Atomistic modeling of intrinsic and radiation-enhanced fission gas (Xe) diffusion in UO2±x: Implications for nuclear fuel performance modeling. J. Nucl. Mater. 2014, 451, 225–242. [Google Scholar] [CrossRef]
  23. Matzke, H. Atomic transport properties in UO2 and mixed oxides (U, Pu)O2. J. Chem. Soc. Faraday Trans. 2 1987, 83, 1121–1142. [Google Scholar] [CrossRef]
  24. Sabioni, A.; Ferraz, W.; Millot, F. First study of uranium self-diffusion in UO2 by SIMS. J. Nucl. Mater. 1998, 257, 180–184. [Google Scholar] [CrossRef]
  25. Forsberg, K.; Massih, A. Diffusion theory of fission gas migration in irradiated nuclear fuel UO2. J. Nucl. Mater. 1985, 135, 140–148. [Google Scholar] [CrossRef]
  26. Forsberg, K.; Massih, A. Fission gas release under time-varying conditions. J. Nucl. Mater. 1985, 127, 141–145. [Google Scholar] [CrossRef]
  27. Lanning, D.D.; Beyer, C.E.; Painter, C.L. FRAPCON-3: Modifications to Fuel Rod Material Properties and Performance Models for High-Burnup Application; Technical Report No. NUREG/CR-6534-Vol. 1; Division of Systems Technology, Office of Nuclear Regulatory Commission, U.S. Nuclear Regulatory Commission: Washington, DC, USA, 1997. [Google Scholar] [CrossRef] [Green Version]
  28. Davies, D.; Long, G. The Emission of Xenon-133 from Lightly Irradiated Uranium Dioxide Spheroids and Powders; Technical Report No. AERE-R-4347; Atomic Energy Research Establishment, United Kingdom Atomic Energy Authority: Harwell, England, 1963. [Google Scholar]
  29. Williams, N.R.; Molinari, M.; Parker, S.C.; Storr, M. Atomistic investigation of the structure and transport properties of tilt grain boundaries of UO2. J. Nucl. Mater. 2014, 458, 45–55. [Google Scholar] [CrossRef]
  30. Bertolus, M.; Freyss, M.; Dorado, B.; Martin, G.; Hoang, K.; Maillard, S.; Skorek, R.; Garcia, P.; Valot, C.; Chartier, A.; et al. Linking atomic and mesoscopic scales for the modelling of the transport properties of uranium dioxide under irradiation. J. Nucl. Mater. 2015, 462, 475–495. [Google Scholar] [CrossRef] [Green Version]
  31. Boyarchenkov, A.; Potashnikov, S.; Nekrasov, K.; Kupryazhkin, A. Investigation of cation self-diffusion mechanisms in UO2±x using molecular dynamics. J. Nucl. Mater. 2013, 442, 148–161. [Google Scholar] [CrossRef] [Green Version]
  32. Childs, B. Fission product effects in uranium dioxide. J. Nucl. Mater. 1963, 9, 217–244. [Google Scholar] [CrossRef]
  33. Olander, D.; Wongsawaeng, D. Re-solution of fission gas—A review: Part I. Intragranular bubbles. J. Nucl. Mater. 2006, 354, 94–109. [Google Scholar] [CrossRef]
  34. Nelson, R. The stability of gas bubbles in an irradiation environment. J. Nucl. Mater. 1969, 31, 153–161. [Google Scholar] [CrossRef]
  35. Evans, J.H. Effect of temperature on bubble precipitation in uranium dioxide implanted with krypton and xenon ions. J. Nucl. Mater. 1992, 188, 222–225. [Google Scholar] [CrossRef]
  36. Michel, A.; Sabathier, C.; Carlot, G.; Kaïtasov, O.; Bouffard, S.; Garcia, P.; Valot, C. An in situ TEM study of the evolution of Xe bubble populations in UO2. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2012, 272, 218–221. [Google Scholar] [CrossRef]
  37. Basak, C.; Sengupta, A.; Kamath, H. Classical molecular dynamics simulation of UO2 to predict thermophysical properties. J. Alloy. Compd. 2003, 360, 210–216. [Google Scholar] [CrossRef]
  38. Morelon, N.-D.; Ghaleb, D.; Delaye, J.-M.; Van Brutzel, L. A new empirical potential for simulating the formation of defects and their mobility in uranium dioxide. Philos. Mag. 2003, 83, 1533–1555. [Google Scholar] [CrossRef]
  39. Cooper, M.W.D.; Rushton, M.; 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]
  40. Geng, H.; Chen, Y.; Kaneta, Y.; Kinoshita, M. Molecular dynamics study on planar clustering of xenon in UO2. J. Alloy. Compd. 2008, 457, 465–471. [Google Scholar] [CrossRef] [Green Version]
  41. Chartier, A.; Van Brutzel, L.; Freyss, M. Atomistic study of stability of xenon nanoclusters in uranium oxide. Phys. Rev. B 2010, 81, 174111. [Google Scholar] [CrossRef]
  42. Thompson, A.E.; Meredig, B.; Stan, M.; Wolverton, C. Interatomic potential for accurate phonons and defects in UO2. J. Nucl. Mater. 2014, 446, 155–162. [Google Scholar] [CrossRef]
  43. Cooper, M.W.D.; Kuganathan, N.; Burr, P.A.; Rushton, M.J.D.; Grimes, R.W.; Stanek, C.R.; Andersson, D.A. Development of Xe and Kr empirical potentials for CeO2, ThO2, UO2 and PuO2, combining DFT with high temperature MD. J. Phys. Condens. Matter 2016, 28, 405401. [Google Scholar] [CrossRef] [Green Version]
  44. Tang, K.T.; Toennies, J.P. The van der Waals potentials between all the rare gas atoms from He to Rn. J. Chem. Phys. 2003, 118, 4976–4983. [Google Scholar] [CrossRef]
  45. Padel, A.; De Novion, C. Constantes elastiques des carbures, nitrures et oxydes d’uranium et de plutonium. J. Nucl. Mater. 1969, 33, 40–51. [Google Scholar] [CrossRef]
  46. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO—The Open Visualization Tool. Model. Simul. Mater. Sci. Eng. 2009, 18, 015012. [Google Scholar] [CrossRef]
  47. Kittel, C. Introduction to Solid State Physics, 8th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2005. [Google Scholar]
  48. Nordlund, K.; Averback, R.S. Point defect movement and annealing in collision cascades. Phys. Rev. B 1997, 56, 2421–2431. [Google Scholar] [CrossRef] [Green Version]
  49. Ball, R.G.J.; Grimes, R.W. Diffusion of Xe in UO2. J. Chem. Soc. Faraday Trans. 1990, 86, 1257–1261. [Google Scholar] [CrossRef]
  50. Catlow, C.R.A. Fission gas diffusion in uranium dioxide. Proc. R. Soc. Lond. A. Math. Phys. Eng. Sci. 1978, 364, 473–497. [Google Scholar] [CrossRef]
  51. Miekeley, W.; Felix, F. Effect of stoichiometry on diffusion of xenon in UO2. J. Nucl. Mater. 1972, 42, 297–306. [Google Scholar] [CrossRef]
  52. Cornell, R.M. The growth of fission gas bubbles in irradiated uranium dioxide. Philos. Mag. 1969, 19, 539–554. [Google Scholar] [CrossRef]
  53. Kaimal, K.; Naik, M.; Paul, A. Temperature dependence of diffusivity of xenon in high dose irradiated UO2. J. Nucl. Mater. 1989, 168, 188–190. [Google Scholar] [CrossRef]
  54. Torres, E.; Kaloni, T. Thermal conductivity and diffusion mechanisms of noble gases in uranium dioxide: A DFT+U study. J. Nucl. Mater. 2019, 521, 137–145. [Google Scholar] [CrossRef]
  55. Turnbull, J. The distribution of intragranular fission gas bubbles in UO2 during irradiation. J. Nucl. Mater. 1971, 38, 203–212. [Google Scholar] [CrossRef]
Figure 1. Relaxation configuration diagrams of different UO2 systems after adding Xe atoms. (ac) are 1, 4, and 6 Xe in the defect-free system, respectively; (df) are 1, 3, and 6 Xe in the system containing U vacancies, respectively. (gi) are 1, 3, and 6 Xe in the system containing double Schottky vacancies, respectively. The dashed frames are the schematics of the squeezed interstitial atoms. The red, yellow, green, and purple balls represent U atoms, Xe atoms, U vacancies, and U interstitial atoms, respectively. The O atoms are ignored.
Figure 1. Relaxation configuration diagrams of different UO2 systems after adding Xe atoms. (ac) are 1, 4, and 6 Xe in the defect-free system, respectively; (df) are 1, 3, and 6 Xe in the system containing U vacancies, respectively. (gi) are 1, 3, and 6 Xe in the system containing double Schottky vacancies, respectively. The dashed frames are the schematics of the squeezed interstitial atoms. The red, yellow, green, and purple balls represent U atoms, Xe atoms, U vacancies, and U interstitial atoms, respectively. The O atoms are ignored.
Metals 12 00789 g001
Figure 2. The formation energy (eV) of (a) Xe clusters and (b) Kr clusters in UO2 with and without defects as a function of the number of Xe or Kr atoms in the formed cluster.
Figure 2. The formation energy (eV) of (a) Xe clusters and (b) Kr clusters in UO2 with and without defects as a function of the number of Xe or Kr atoms in the formed cluster.
Metals 12 00789 g002
Figure 3. The binding energy (eV) of (a) an additional Xe atom and (b) an additional Kr atom to a cluster as a function of the number of Xe or Kr atoms in the formed cluster.
Figure 3. The binding energy (eV) of (a) an additional Xe atom and (b) an additional Kr atom to a cluster as a function of the number of Xe or Kr atoms in the formed cluster.
Metals 12 00789 g003
Figure 4. The diffusion coefficient for (a) Xe and (b) Kr clusters in bulk UO2. The lines are linear Arrhenius fits.
Figure 4. The diffusion coefficient for (a) Xe and (b) Kr clusters in bulk UO2. The lines are linear Arrhenius fits.
Metals 12 00789 g004
Figure 5. System evolution of Xe clusters at different times at 2500 K (the yellow balls represent Xe atoms, the red circles mark the clusters); (a) 0 ns, (b) 5 ns, (c) 10 ns, (d) 20 ns.
Figure 5. System evolution of Xe clusters at different times at 2500 K (the yellow balls represent Xe atoms, the red circles mark the clusters); (a) 0 ns, (b) 5 ns, (c) 10 ns, (d) 20 ns.
Metals 12 00789 g005
Figure 6. The cluster sizes distribution of (a) 2% Xe, (b) 5% Xe, (c) 2% Kr, and (d) 5% Kr in different systems, respectively. The pink, light blue, dark blue, green and red bars are the systems with 2% interstitial concentration, with 1% interstitial concentration, perfect bulk without defects, with 1% vacancies concentration, with 2% vacancies concentration, respectively.
Figure 6. The cluster sizes distribution of (a) 2% Xe, (b) 5% Xe, (c) 2% Kr, and (d) 5% Kr in different systems, respectively. The pink, light blue, dark blue, green and red bars are the systems with 2% interstitial concentration, with 1% interstitial concentration, perfect bulk without defects, with 1% vacancies concentration, with 2% vacancies concentration, respectively.
Metals 12 00789 g006
Figure 7. Xe clusters distribution and defects distribution at 5% concentration of Xe atoms in different systems. The green, red, and yellow balls represent the lattice U vacancies, U interstitials, and Xe atoms.
Figure 7. Xe clusters distribution and defects distribution at 5% concentration of Xe atoms in different systems. The green, red, and yellow balls represent the lattice U vacancies, U interstitials, and Xe atoms.
Metals 12 00789 g007
Table 1. Diffusion energy barrier and diffusion prefactor of small interstitial Xe clusters in UO2.
Table 1. Diffusion energy barrier and diffusion prefactor of small interstitial Xe clusters in UO2.
Number of Xe Diffusion Energy Barrier (eV) Diffusion Prefactor (m2/s)
Xe12.111.8 × 10−5
Xe22.150.35 × 10−5
Xe32.070.25 × 10−5
Table 2. Diffusion energy barrier and diffusion prefactor of small interstitial Kr clusters in UO2.
Table 2. Diffusion energy barrier and diffusion prefactor of small interstitial Kr clusters in UO2.
Number of Kr Diffusion energy Barrier (eV) Diffusion Prefactor (m2/s)
Kr12.310.12 × 10−3
Kr21.890.20 × 10−5
Kr31.950.12 × 10−5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, L.; Wang, Z.; Xia, Y.; Chen, Y.; Liu, Z.; Wang, Q.; Wu, L.; Hu, W.; Deng, H. Effects of Point Defects on the Stable Occupation, Diffusion and Nucleation of Xe and Kr in UO2. Metals 2022, 12, 789. https://doi.org/10.3390/met12050789

AMA Style

Wang L, Wang Z, Xia Y, Chen Y, Liu Z, Wang Q, Wu L, Hu W, Deng H. Effects of Point Defects on the Stable Occupation, Diffusion and Nucleation of Xe and Kr in UO2. Metals. 2022; 12(5):789. https://doi.org/10.3390/met12050789

Chicago/Turabian Style

Wang, Li, Zhen Wang, Yaping Xia, Yangchun Chen, Zhixiao Liu, Qingqing Wang, Lu Wu, Wangyu Hu, and Huiqiu Deng. 2022. "Effects of Point Defects on the Stable Occupation, Diffusion and Nucleation of Xe and Kr in UO2" Metals 12, no. 5: 789. https://doi.org/10.3390/met12050789

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