Abstract
The Bredig transition to the superionic phase indicated with the
Introduction
Plutonium dioxide
Recently, the melting behavior of stoichiometric PuO2 has been studied for the first time by the fast laser heating and multi-wavelength spectro-pyrometry to reduce the side effect mentioned earlier [6]. Based on this measurement the melting point of PuO2 has been determined as
At this point, molecular dynamics simulation turns out to be a useful tool for studying the properties that may be essential for nuclear facilities to run safely. Molecular dynamics simulation of such systems requires developing a reliable potential for the interactions. Within the nuclear fuel materials, uranium dioxide UO2 is one of the most studied systems. There have been a number of pair potentials developed to understand thermophysical and transport properties of UO2 at solid and liquid phases. Govers et al. [9,10] have made a broad comparison to assess the range of applicability of these potentials. For PuO2, previous studies have come up with some pair potentials, generally in the form of Buckingham, Morse, embedded atom model (EAM) and shell model to investigate PuO2 [11–15]. Recently, Cooper et al. [15] have reported a many-body potential model to describe the thermomechanical properties of actinide oxides between 300 and 3,000 K. However, the Bredig transition has not been considered at any stage of this work.
PuO2, like other actinide dioxides, is believed to be a type II superionic conductor, which displays a rapid but continuous increase in the ionic conductivity by heating at about
The latest value of the melting temperature of PuO2 determined by laser-heating technique gave us the opportunity to investigate the thermally induced Bredig transition and the melting. For this purpose, with the present article, a new semi-empirical rigid ion potential is introduced, and thermophysical properties calculated via classical molecular dynamics simulation in constant pressure–temperature ensemble are presented for the wide range of temperature (300—3,600 K).
Potential model
There exist a few semi-empirical potential model in literature developed for PuO2 [11–15]. Yamada, Kurosaki and Arima [11–13] have used Morse- and Buckingham-type pair interaction potential function. Buckingham type with shell model pair interaction potential has been introduced by Chu [14]. Cooper added the term many-body EAM to Buckingham- and Morse-type pair interaction potentials.
In general, the parameters for these potential functions O–O interaction have been generally taken from the potential developed earlier forUO2. On the other side, Pu–Pu and Pu–O parameters have been determined by trial-and-error method, aiming to reproduce the experimental lattice constant at low temperatures. In obtaining the parameters of the pair potential in the present study, we also take the O–O parameters from the potential developed for UO2 in our previous work [16]. The other parameters were adjusted in order to reproduce low-temperature experimental values of lattice parameter, bulk modulus, elastic constants and cohesive energy. Moreover, we have also aimed to model the Bredig transition to the superionic phase, which is highly expected but has not been yet observed experimentally. The potential function has been originally proposed by Vashishta-Rahman [17] and is given by
The first term stands for the Coulomb interactions, second term contains core repulsions, third term is the effective monopol-induced dipole interaction and the last term is Van der Waals interaction. The potential parameters are given in Table 1.
The potential parameters used in the present work. ZPu = 1.96e and ZO = –0.98e.
Aij (eV) | Pij (eV Å4) | Cij (eV Å6) | ηij | σi,j (Å) | |
---|---|---|---|---|---|
Pu–Pu | 1.52658 | 0.0 | 0.0 | 7 | 1.37 |
Pu–O | 0.5238 | 0.0 | 0.0 | 7 | |
O–O | 4.0859 | 40 | 8.3 | 7 | 1.00 |
Molecular dynamics simulation
Molecular dynamics technique was performed for 5 × 5 × 5 cell constructed with 500 cations (Pu4+) and 1,000 anions (O2–) arranged as CaF2-type structure. The MD code called as MOLDY [18] was adopted to carry out the calculations. The Ewald’s sum technique is used to account for the long-range Coulomb interactions. The positions and velocities of the ions are calculated by integrating the Newton’s equation of motion using Beeman’s algorithm, which is predictor–corrector type, with the time step
Results and discussion
The values of lattice parameter, bulk modulus, elastic constants and cohesive energy are given in Table 2 and compared with the available experimental data and ab initio results. The bulk modulus and elastic constants were calculated from the utilization of known Birch–Murnaghan equation of state [19, 20], which is in good agreement with experiment and ab initio results [21–25]. Contrary to the previous MD results [11–15], in the present paper the lattice parameter is underestimated about 1.3 % and 3 % compared to the ab initio calculations and experimental data, respectively [22, 23]. However, bulk modulus obtained with the previous potentials has been overestimated and varied between 200 and 239 GPa [11–15] when compared with the experimental result 178 GPa [21]. In the present study, the amount of deviation of bulk modulus from the experimental study is about 5 %, whereas other studies are between 12 % and 34 %. The main reason for these results is the weight of observables while finding the potential parameters. Groups place great emphasis on the lattice constant while developing potential parameters. In this study, we distribute the weight constant unevenly on lattice constant, bulk modulus, elastic constants and energy as we are performing fitting procedure. Temperature dependence of energy is also taken into account in order to mimic the fast-ion phase transition. Discrepancy between the calculated value and experimental data of C12 are presented in Table 2. This surely can be modified by changing the potential parameters but in that the indication of the phase transition can be lost.
Comparison of calculated results with MD simulation, experimental and ab initio data.
Present | Exp./ab initio | MD simulation | |
---|---|---|---|
B0 (GPa) | 168 | 178 [21] | 200–239 [11–15] |
a0 (Ả) | 5.235 | 5.307–5.398 [22,23] | 5.389–5.39 [11–15] |
C11 (GPa) | 325 | 256–386 [24] | 424.3 [15] |
C12 (GPa) | 88 | 112–177 [24] | 111.7 [15] |
C44 (GPa) | 74 | 53–74 [24] | 69.2 [15] |
EC (eV/PuO2) | 31.8 | 19.24–24 [23,25] | – |
![Figure 1: Lattice parameter evolution with temperature. Experimental data are from TPRC data series [22].](/document/doi/10.1515/htmp-2015-0133/asset/graphic/htmp-2015-0133_figure1.gif)
Lattice parameter evolution with temperature. Experimental data are from TPRC data series [22].

Evolution of relative linear thermal expansion of present results, MD data and experimental data with temperature.
The temperature dependence of the pair correlation functions

Pair correlation functions of
The radial distribution function is defined as follows:
For a species of N particle, the mean square displacement (MSD) is calculated as
where

Mean square displacement versus time at different temperatures.

Evolution of diffusion coefficient with temperature for both plutonium and oxygen ions.
It is clearly evident from the anomalous increase of diffusion coefficient of oxygen ions at ~2,100 K that there is an onset of Bredig transition to the superionic phase. While the oxygen diffusion continuously increases beyond 2,100 K, the plutonium ions show almost no diffusive behavior. As experimental data for oxygen diffusion coefficient for PuO2 is only available up to 1,400 K [26], we are not able to compare self-diffusion coefficient. As the temperature increases, Pu ions also show diffusive behavior after 3,000 K and DPu is less than DO about an order of 10.
The enthalpy change

Enthalpy change of
The existence of the thermally activated transition into superionic phase can be confirmed by a λ-peak in the heat capacity at constant pressure

Temperature dependence of heat capacity for
A very weak increase in Cp up to 1,750 K is interpreted as increase in the anharmonicity of the lattice vibrations. Further increase in temperature results with the additional increase in calculated Cp and a λ shape peak is clearly produced at the critical temperature Tc = 2,055 K, which is close to the expected value of the phase transition temperature in fluorite type of ionic crystals [7]. In order to make sure that it is a
Conclusions
A model of PuO2 as a superionic conductor has been constructed via parameterization of a rigid ion potential with partially ionized atoms. The thermophysical properties of the model have been investigated up to 3,600 K within the classical molecular dynamics in the NPT ensemble. A sign of an anomalous behavior in the lattice parameter and oxygen–oxygen pair correlation function was observed at about 2,100 K. This was reflected as a sharp increase in the self-oxygen diffusion coefficient to the value of a typical liquid system, a discontinuity in enthalpy change and as a λ-type peak in the constant pressure heat capacity at about the same temperature. These features indicate that the system undergoes a thermally activated Bredig transition to the superionic phase and validates the constructed superionic model of PuO2. As it is also pointed out by Fink [7], more accurate high-temperature enthalpy measurements are needed to put this issue beyond the doubts.
Funding statement: Funding: This research was supported by Yildiz Technical University Scientific Research Projects Coordination Department. Project Number: 2013-01-01-GEP01.
References
[1] W.L. Lyon and W.E. Baily, J. Nucl. Mater., 22 (1967) 332–339.10.1016/0022-3115(67)90051-7Search in Google Scholar
[2] E.A. Aitken and S.K. Evans, The thermodynamic data program involving plutonium and urania at high temperatures, USAEC Report, General Electric, GEAP-5672 (1968).10.2172/4503229Search in Google Scholar
[3] B. Riley, Sci. Ceram., 5 (1970) 83–109.10.3406/ecoru.1970.2094Search in Google Scholar
[4] M. Kato, K. Morimoto, H. Sugata, K. Konashi, M. Kashimura and T. Abe, J. Nucl. Mater., 373 (2008) 237–245.10.1016/j.jnucmat.2007.06.002Search in Google Scholar
[5] F. De Bruycker, K. Boboridis, P. Pöml, R. Eloirdi, R.J.M. Konings and D. Manara, J. Nucl. Mater., 416 (2011) 166–172.10.1016/j.jnucmat.2010.11.030Search in Google Scholar
[6] D. Manara, R. Böhler, K. Boboridis, L. Capriotti, A. Quaini, L. Luzzi, F. De Bruycker, C. Guéneau, N. Dupin and R. Konings, Proc. Chem., 7 (2012) 505–512.10.1016/j.proche.2012.10.077Search in Google Scholar
[7] J.K. Fink, Int. J. Thermophys., 3 (2) (1982) 165–200.10.1007/BF00503638Search in Google Scholar
[8] J. Ralph, J. Chem. Soc. Faraday Trans., 2 (83) (1987) 1253–1262.10.1039/F29878301253Search in Google Scholar
[9] K. Govers, S. Lemehov, M. Hou and M. Verwerft, J. Nucl. Mater., 366 (1–2) (2007) 161–177.10.1016/j.jnucmat.2006.12.070Search in Google Scholar
[10] K. Govers, S. Lemehov, M. Hou and M. Verwerft, J. Nucl. Mater., 376 (1) (2008) 66–77.10.1016/j.jnucmat.2008.01.023Search in Google Scholar
[11] K. Yamada, K. Kurosaki, M. Uno and S. Yamanaka, J. Alloys Compd., 307 (2000) 1–9.10.1016/S0925-8388(00)00805-7Search in Google Scholar
[12] K. Kurosaki, K. Yamada, M. Uno, S. Yamanaka, K. Yamamoto and T. Namekawa, J. Nucl. Mater., 294 (2001) 160–167.10.1016/S0022-3115(01)00451-2Search in Google Scholar
[13] T. Arima, S. Yamasaki and Y. Inagaki, J. Alloys Compd., 400 (2005) 43–50.10.1016/j.jallcom.2005.04.003Search in Google Scholar
[14] M. Chu, D. Meng, S. Xiao, W. Wang and Q. Chen, J. Alloys Compd., 539 (2012) 7–11.10.1016/j.jallcom.2012.05.094Search in Google Scholar
[15] M.W.D. Cooper, M.J.D. Rushton, R.W. Grimes, J. Phys. Condens. Matter, 26 (2014) 105401.10.1088/0953-8984/26/10/105401Search in Google Scholar
[16] S.D. Günay, Ü. Akdere, H.B. Kavanoz and Ç. Taşseven, Int. J. Mod. Phys. B, 25 9 (2011) 1201–1210.10.1142/S0217979211100138Search in Google Scholar
[17] P. Vashishta and A. Rahman, Phys. Rev. Lett., 40 (1978) 1337–1340.10.1103/PhysRevLett.40.1337Search in Google Scholar
[18] K. Refson, Comput. Phys. Commun., 126 (3) (2000) 310–329.10.1016/S0010-4655(99)00496-8Search in Google Scholar
[19] J.H. Li, S.H. Liang, H.B. Guo and B.X. Liu, Appl. Phys. Lett., 87 (2005) 194111.10.1063/1.2128071Search in Google Scholar
[20] J.D. Gale and A.L. Rohl, Mol. Simul., 29 (2003) 291–341.10.1080/0892702031000104887Search in Google Scholar
[21] M. Idiri, T. Le Bihan, S. Heathmen and J. Rebizant, Phys. Rev. B, 70 (2004) 014113.10.1103/PhysRevB.70.014113Search in Google Scholar
[22] Y.S. Touloukian and C.Y. Ho, Thermophysical properties of matter, The TPRC data series, vol. 13. IFI/Plenum Data, New York (1970).Search in Google Scholar
[23] P.J. Kelly and M.S.S. Brooks, J. Chem. Soc., Faraday Trans., 2 (83) (1987) 1189–1203.10.1039/f29878301189Search in Google Scholar
[24] P. Zhang, B-T. Wang and X-G. Zhao, Phys. Rev. B, 82 (2010) 144110.10.1103/PhysRevB.82.144110Search in Google Scholar
[25] M. Freyss, N. Vergnet and T. Petit, J. Nucl. Mater., 352 (2006) 144–150.10.1016/j.jnucmat.2006.02.048Search in Google Scholar
[26] G.E. Murch and C.R.A. Catlow, J. Chem. Soc., Faraday Trans., 2 (83) (1987) 1157–1169.10.1039/f29878301157Search in Google Scholar
[27] M.A. Bredig, Colloq. Int. CNRS, 205 (1971) 183–191.Search in Google Scholar
©2016 by De Gruyter
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Frontmatter
- Research Articles
- Effects of Extrusion–Shear Process Conditions on the Microstructures and Mechanical Properties of AZ31 Magnesium Alloy
- Degradation of TiB2/TiC Composites in Liquid Nd and Molten NdF3–LiF–Nd2O3 System
- Investigation of Oxygen Diffusion in Irradiated UO2 with MD Simulation
- Grain Size Effect on Fracture Behavior of the Axis-Tensile Test of Inconel 718 Sheet
- Modeling Superionic Behavior of Plutonium Dioxide
- Influence of Grain Refinement on Oxidation Behavior of Two-Phase Cu–Cr Alloys at 973–1,073 K in Air
- Synthesis and Characterization of Cadmium Sulfide Nanoparticles via a Simple Thermal Decompose Method
- Short Communication
- Simple Thermal Decompose Method for Synthesis of Nickel Disulfide Nanostructures
- Research Articles
- Dynamic Recrystallization Behavior of Ti22Al25Nb Alloy during Hot Isothermal Deformation
- Reduction Smelting Low Ferronickel from Pre-concentrated Nickel-Iron Ore of Nickel Laterite
- Physics-Based Constitutive Model to Predict Dynamic Recovery Behavior of BFe10-1-2 Cupronickel Alloy during Hot Working
- Influence of Oxides on Microstructures and Mechanical Properties of High-Strength Steel Weld Joint
Articles in the same Issue
- Frontmatter
- Research Articles
- Effects of Extrusion–Shear Process Conditions on the Microstructures and Mechanical Properties of AZ31 Magnesium Alloy
- Degradation of TiB2/TiC Composites in Liquid Nd and Molten NdF3–LiF–Nd2O3 System
- Investigation of Oxygen Diffusion in Irradiated UO2 with MD Simulation
- Grain Size Effect on Fracture Behavior of the Axis-Tensile Test of Inconel 718 Sheet
- Modeling Superionic Behavior of Plutonium Dioxide
- Influence of Grain Refinement on Oxidation Behavior of Two-Phase Cu–Cr Alloys at 973–1,073 K in Air
- Synthesis and Characterization of Cadmium Sulfide Nanoparticles via a Simple Thermal Decompose Method
- Short Communication
- Simple Thermal Decompose Method for Synthesis of Nickel Disulfide Nanostructures
- Research Articles
- Dynamic Recrystallization Behavior of Ti22Al25Nb Alloy during Hot Isothermal Deformation
- Reduction Smelting Low Ferronickel from Pre-concentrated Nickel-Iron Ore of Nickel Laterite
- Physics-Based Constitutive Model to Predict Dynamic Recovery Behavior of BFe10-1-2 Cupronickel Alloy during Hot Working
- Influence of Oxides on Microstructures and Mechanical Properties of High-Strength Steel Weld Joint