Sunteți pe pagina 1din 8

Nuclear Engineering and Technology xxx (xxxx) xxx

Contents lists available at ScienceDirect

Nuclear Engineering and Technology


journal homepage: www.elsevier.com/locate/net

Original Article

Thermal transport study in actinide oxides with point defects


Alex Resnick 1, Katherine Mitchell 1, Jungkyu Park*, Eduardo B. Farfa
n, Tien Yee
Kennesaw State University, Department of Mechanical Engineering, Kennesaw, GA, 30144, USA

a r t i c l e i n f o a b s t r a c t

Article history: We use a molecular dynamics simulation to explore thermal transport in oxide nuclear fuels with point
Received 25 October 2018 defects. The effect of vacancy and substitutional defects on the thermal conductivity of plutonium di-
Received in revised form oxide and uranium dioxide is investigated. It is found that the thermal conductivities of these fuels are
11 March 2019
reduced significantly by the presence of small amount of vacancy defects; 0.1% oxygen vacancy reduces
Accepted 14 March 2019
Available online xxx
the thermal conductivity of plutonium dioxide by more than 10%. The missing of larger atoms has a more
detrimental impact on the thermal conductivity of actinide oxides. In uranium dioxide, for example, 0.1%
uranium vacancies decrease the thermal conductivity by 24.6% while the same concentration of oxygen
Keywords:
Nuclear fuel
vacancies decreases the thermal conductivity by 19.4%. However, uranium substitution has a minimal
Thermal conductivity effect on the thermal conductivity; 1.0% uranium substitution decreases the thermal conductivity of
Defects plutonium dioxide only by 1.5%.
Molecular dynamics © 2019 Korean Nuclear Society, Published by Elsevier Korea LLC. This is an open access article under the
CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction multi-point vacancies, interstitial vacancies, and substitutional


atoms. While these defects alter the lattice vibrations in the
Effective control of operating temperatures within a nuclear structures, many of the actinide oxides research focuses on for-
reactor is a fundamental requirement of power plant operations mation energy [1], diffusion, and uranium doping [2,3]. Ma and
and is strongly correlated to the thermal conductivity of the nuclear Asok [4] used the all-electron full-potential linearized augmented
fuel being used within the reactor core. Among other thermo- plane wave plus local orbitals basis method while Petit et al. [5]
physical characteristics of nuclear fuels, thermal conductivity is a used local-density approximation to calculate formation energies.
key component when managing the temperature safety margins Duriez et al. [6] experimented on low-Pu content MOX fuel with
during a power plant's operation. During operation, a large tem- oxygen vacancy point defects to develop quantitative input for a
perature gradient is induced with the fuel during reactor cycle and simple phonon conduction model.
results in cracking because of the increased thermal stress. Fuel In the present study, uranium dioxide (UO2), plutonium dioxide
pellet swelling also occurs due to the limited dissipation of thermal (PuO2), and thorium dioxide (ThO2) are selected to study the effect
energy as uneven heat distribution increases the release rate of of point defects on the thermal transport properties in actinide
fission gases. Moreover, the increased fission gas release is expected oxides. The both oxygen point defects and cation (plutonium/ura-
to cause embrittlement in grain boundaries, significantly short- nium) defects are investigated. Although the majority of re-
ening the creep life of nuclear fuel pellets. Therefore, ensuring searchers reported the effect of anion point defects in the fluorite-
efficient thermal transport in nuclear fuel pellets will stand as a type oxides using atomistic simulations, cation defects such as
major hurdle for the next generation reactors currently in uranium vacancy and plutonium vacancy can affect thermal
development. transport in oxide fuels since they alter the lattice structures of
Temperature, lattice structure, point defect, etc. are all ther- oxide fuels largely. Additionally, the phonon density of states
mophysical or chemo-physical material characteristic traits that (PDOS) for UO2 and PuO2 were examined to better understand the
can alter thermal conductivity. Point defects in actinide oxides are effects of the uranium substitution. To obtain the thermal con-
interruptions in the crystalline materials. Point defects include ductivities of bulk structures, simulations are run for the simulation
structures with several sample lengths, ranging from 10 to 100 unit
cells in length.
* Corresponding author.
E-mail address: jpark186@kennesaw.edu (J. Park).
1
These authors contributed equally to the study.

https://doi.org/10.1016/j.net.2019.03.011
1738-5733/© 2019 Korean Nuclear Society, Published by Elsevier Korea LLC. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/
licenses/by-nc-nd/4.0/).

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
2 A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx

2. Simulation method

In this research, reverse non-equilibrium molecular dynamics


(RNEMD) [7] is employed to estimate the thermal conductivity of
actinide oxides with various defect types and concentrations. In
RNEMD, a heat flux is imposed across a simulation box by
exchanging energies between the hot bath and the cold bath
(Fig. 1). Once the heat flow in the structure reaches a steady state,
the thermal conductivity (k), is calculated using the Fourier's heat
conduction law, k ¼  <q> ∕ <dT/dx> where <q> is averaged heat
flux and <dT/dx> is temperature gradient.
The potential field, or set of interatomic energy potentials,
developed by Cooper et al. [8e10] is used together with LAMMPS
simulator [11] for the all simulations in the present study. The
potential field is constructed by combining many body interaction
and pairwise interactions as shown in the following equation.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X  ffi
1X  
Ei ¼ fab rij  Ga sb rij (1)
2 j j

Fig. 2. Unit cell of PuO2 or UO2 with various defects. (a) Stoichiometric unit cell. (b)
Here, Ei represents the potential energy of the ith atom when Oxygen vacancy. (d) Combined vacancies of primary and oxygen atom. (e) Uranium
compared to all other atoms, and rij is the distance between atoms. substitutional defect.
a and b represents the species of ith atom and jth atom respectively.
The first term describes pairwise interaction between ith atom and
jth atom, given by fab(rij), while the second term, sb(rij), in- PuO2 and UO2 respectively (Fig. 2c). Combined vacancy defects
corporates many body perturbations into the pairwise interaction. occur when both primary atoms and oxygen atoms are missing in
Ga is the constant of proportionality. This force field has success- the molecular structure (Fig. 2d). For the case of PuO2, uranium
fully predicted thermos-physical properties of actinide solids [8] substitutional defects are explored as shown in Fig. 2e; PuO2
and the thermal conductivity of UO2 [12]. Rahman et al. [13] structures with uranium substitutional defects are obtained by
showed the effectiveness of this potential model. The authors replacing plutonium atoms with uranium atoms. To obtain a better
also successfully estimated the thermal properties of ThO2 using understanding on the effect of uranium substitution, ThO2 with
the same potential [14,15]. uranium substitution is also simulated in the present study. The
Simulation structures with different types of defects and defect point defects shown in Fig. 2 are not likely to maintain their original
concentrations are constructed using a custom MATLAB program. positions during molecular dynamics simulations; vacancy and/or
Fig. 2 shows the unit cell of stoichiometric sample (Fig. 2a) and the interstitial diffusions of these point defects will rearrange cations
unit cells with three different types of vacancy defects (Fig. 2bed) and anions in the structure. Since the diffusion process is expected
and substitutional defect (Fig. 2e) studied in this research study. to be accelerated at high temperatures, molecular arrangement (or
Vacancy defects occur when atoms are missing from their lattice the positions of cations and anions) in a structure with high tem-
sites. For example, oxygen vacancy defects appear when some of perature will be different from that in a structure with low tem-
the oxygen atoms in the simulation structure are absent from their perature. This difference between the arrangements of cations and
lattice sites while plutonium vacancy defects are found when anions can cause a difference in the structure's thermal transport
plutonium atoms are missing from their lattice sites. Simulation property by altering its phonon scattering mechanism. Therefore,
structures with oxygen vacancy defects are constructed for the both more simulation studies are required to carefully track the rear-
of PuO2 and UO2 with different sample lengths (Fig. 2b). Plutonium rangement of cations and anions during molecular dynamics
vacancy defects and uranium vacancy defects are simulated for simulations.
The simulation utilizes actinide dioxides of the FCC fluorite-type
(F3m3) with the notation of M1-yO2±x where M ¼ U or Pu and x or
y ¼ 0, 0.001, 0.005, 0.01, 0.02, or 0.05 [16] for all vacancy defect
types. An example of a simulation sample with a 1% oxygen vacancy
defect would be U1O2-0.01. Similarly, a simulation sample of PuO2
with uranium substitution would be P1-xUxO2 in which x ¼ 0,
0.001, 0.005, 0.01, 0.02, or 0.05. The defect concentration is defined
by using the formula shown below.

Number of defect sites



Total number of atoms in the stoichiometric structure
 100:
(2)
Defect concentrations are determined by dividing the number of
defect sites by the total number of atoms in the stoichiometric
structure. In the simulation structure with 7200 unit cells
(6  6  200 supercell) of PuO2, for example, 1% oxygen vacancy
Fig. 1. The schematic of RNEMD in actinide oxide structure with vacancy defects. defects indicate that there are 864 oxygen atoms missing in the

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx 3

simulation structure when compared to its stoichiometric structure RNEMD is sufficient for thermal conductivity estimation. Table 1
since there are 86,400 atoms (28,800 oxygen atoms þ 57,600 lists the lattice constants for the various structures simulated in
plutonium atoms) present in the stoichiometric structure. Likewise, this study after the equilibration. As can be seen in Table 1, the
the concentration of substitutional defects is determined by lattice constants of simulation structures with defects do not
dividing the number of substitutional defect sites by the total exhibit a notable difference from the lattice constant of their stoi-
number of atoms in the stoichiometric structure. For the case of chiometric structures because of the small amount of defect con-
combined vacancies, two different concentrations are selected; centrations simulated in this study.
PuO2 with 1% Pu and 1% O vacancies, PuO2 with 5% Pu and 5% O Fig. 3 shows the representative temperature profiles after 2 ns of
vacancies, UO2 with 1% U and 1% O vacancies, and UO2 with 5% U RNEMD simulations. The distinct irregularities in the simulation
and 5% O vacancies are simulated. structures with lower defect concentrations make thermal con-
To estimate the bulk thermal conductivities of actinide oxides ductivity estimation more uncertain when compared to the case of
with defects, the size of simulation structures is increased pro- higher defect concentrations, thus resulting in bigger standard
gressively in the direction of the thermal conductivity measure- deviations in the structures with lower defect concentrations. This
ment (x direction in this study). Here, the number of unit cells (N) is indicates that the thermal transport in the simulation structures
used to represent a structure size instead of its physical length (L) to with high purity is ballistic and phonons in these structures
avoid the complexity that stems from the different sample lengths encounter more disruptive phonon scatterings at defect locations
after the equilibration of the simulation structure (or after the than in the structures with many defects. As the defect concen-
simulation system reaches a state that does not show changes in its tration increases, the thermal transport in the structures becomes
thermophysical properties such as internal energy and temperature diffusive, resulting in a smooth linear temperature profile. Tem-
gradient); a simulation structure with a fixed number of unit cells perature gradient is obtained from the linear region of temperature
often exhibits different simulation box sizes depending on the profile (red line shown in Fig. 3a) using a linear regression function
defect concentrations once the simulation structure is equilibrated. in MATLAB.
Nx, Ny, and Nz indicate the number of unit cells of a simulation It should be noted that Pu-239 has been selected as a constit-
structure in x, y, and z directions, respectively. In all structures uent for PuO2 instead of the most stable isotope of plutonium, i.e.
simulated, the number of unit cells in y (Ny) and z (Nz) directions are Pu-244 in the present study since the composition of Pu-239 is the
fixed to be 6. In their previous study [14], the authors observed that largest among the isotopes of Pu in MOX fuels, therefore most
thermal transport property is not affected much by the boundary commonly used in nuclear reactors. To examine the effect of atomic
scatterings from x-y or x-z planes. This is mainly because the sim- mass in Pu, the authors simulated the 100 unit cells of PuO2 with
ulations conducted here are carefully designed to maintain 1D heat Pu-244 and identified that the thermal conductivity of PuO2 with
conduction. The same phenomenon has been observed repeatedly Pu-239 is 5% higher than the thermal conductivity of PuO2 with Pu-
by other researchers [17e19]. 244. The authors will provide a more systematic study on the effect
The half of the total number of unit cells in x direction is defined of different isotopic compositions in MOX fuels in the future.
as Nx. In RNEMD algorithm, the hot bath is located at the center of
the simulation box while two cold baths are located at two ends of
the simulation box. Therefore, the characteristic length for the 3. Results and discussion
thermal conductivity estimation becomes the half of the total
simulation box size. Seven different lengths (Nx ¼ 10, 20, 30, 40, 50, The thermal conductivities of UO2 and PuO2 with different va-
75, and 100) are used to estimate the thermal conductivities of cancies of various concentrations are plotted together with the
PuO2 and UO2 with various defect types and concentrations. thermal conductivities of their stoichiometric samples in Fig. 4 as a
It should be noted that the charge imbalance caused by missing function of the number of unit cells in x direction, Nx. The highest
atoms and/or substitutions are ignored in this research study conductivities of stoichiometric PuO2 and UO2 are found to be
because of the relatively small amount of defect concentrations. In 17.15 ± 0.29 W/m-K and 14.99 ± 0.42 W/m-K respectively for the
reality, charge compensation is expected to occur in actinide oxides simulation structures with the longest sample length, i.e. Nx ¼ 100.
with randomly distributed point defects. To accurately analyze the For the case of stoichiometric UO2 and PuO2, the only source of
effect of charged defects, it is necessary to consider the corre- phonon scattering is intrinsic phonon-phonon scattering and the
sponding charge compensating defects. For example, actual charge phonon scattering from the simulation boxes. Consequently, the
compensating defects can be included in the same simulation thermal conductivities of stoichiometric UO2 and PuO2 increase
structure to balance the charge of the simulation structure. In this with increasing the sample length because of the ballistic thermal
case, however, another undesired effect such as phonon scattering transport when the sample length is smaller than their intrinsic
from these additional defect sites need to be carefully examined. phonon mean free path. When the sample length is smaller than
All structures simulated in this study (including structures with the phonon mean free path, phonons are scattered mainly by the
defects) are prepared by the following equilibration procedure. To simulation boundaries, and the length of the simulation structure
achieve well-equilibrated simulation structures, energy-
minimization is performed first to get a near 0 K structure.
Table 1
Energy-minimization is a process of finding an arrangement of Representative lattice constants after equilibration.
atoms that gives the almost zero net interatomic force. Then, the
temperature is elevated to 500 K using isenthalpic (NPH) ensemble Lattice constant, a (Å)

with Langevin thermostat to accelerate the equilibration process. % Defect UO2 (Nx ¼ 100) PuO2 (Nx ¼ 100)
Thereafter, the temperatures of the simulation structures are O vacancy U vacancy O vacancy U substitution
decreased to 300 K with isothermal-isobaric (NPT) ensemble. 0.5 fs
0% 5.469 5.469 5.395 5.395
is used as the time-step during the entire equilibration steps. 0.1% 5.468 5.468 5.394 5.394
RNEMD is carried out on the equilibrated structures for 2 ns with a 0.5% 5.466 5.469 5.392 5.395
time-step of 1 fs A constant amount of swapped energy is observed 1% 5.466 5.468 5.390 5.396
for all simulations in the present study and temperature profiles 2% 5.462 5.469 5.385 5.399
5% 5.445 5.468 5.362 5.405
reach a steady state after 0.2 ns; a total time duration of 2 ns in

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
4 A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx

Fig. 3. Representative temperature profiles during RNEMD.

limits the thermal conductivity. The authors estimated the phonon when defect concentration is high (1%, 2%, and 5%) (Fig. 4). The
mean free path of ThO2 to be 7e8.5 nm at 300K in their previous diminishing slopes of thermal conductivity as defect concentra-
study [14], and they believe the phonon mean free paths of UO2 and tions increase demonstrate that thermal transport becomes more
PuO2 are not deviated much from that of ThO2. This length diffusive in structures with higher vacancy defect concentrations;
dependent thermal conductivity of nanostructured solids is the diffusive thermal transport in these structures can be identified
constantly identified by researchers [18,20,21] who used atomistic in Fig. 3b and d where temperature slips are not identifiable at the
simulations. The thermal conductivity ceases to increase linearly locations of hot bath and cold bath. This diffusive thermal behavior
once the structure size becomes larger than the phonon mean free in actinide oxides with large amount of defects indicate that the
path, and eventually reach its bulk thermal conductivity. intrinsic phonon mean free path in defective actinide oxides be-
The simulation method used in this research study has already comes much shorter when compared to stoichiometric samples.
been validated in the authors’ previous articles [14,15]. The thermal Vacancy defects decrease the thermal conductivities of UO2 and
conductivities of stoichiometric UO2 and PuO2 obtained in the PuO2 for all defect concentrations and lengths simulated in this
present research study are supported by other researchers who study. It is interesting to note that oxygen vacancies degrade the
reported the thermal conductivities of UO2 and PuO2 using atom- thermal conductivities of these nuclear fuels significantly in spite of
istic simulations. For example, Arima et al. [3] reported 15.2 W/m-K their small sizes. For example, the thermal conductivity of UO2 with
as the thermal conductivity of PuO2 using equilibrium molecular Nx ¼ 100 is decreased by 19.4% when 0.1% oxygen vacancies are
dynamics. Qin et al. [12] reported 14.6 W/m-K as the thermal introduced into the simulation structure. For the case of PuO2 with
conductivity of UO2 using molecular dynamics simulations. Cooper Nx ¼ 100, the thermal conductivity decreases by 16.5% when 0.1%
et al. [22] used the same RNEMD and estimated the thermal con- oxygen vacancy is added. This detrimental effect of oxygen va-
ductivities of UO2 and PuO2 to be ~17.5 W/m-K and ~22.5 W/m-K cancies have been identified in the authors’ previous research study
respectively. Fink [23] reported ~13.5 W/m-K as the thermal con- [15] that estimated the thermal conductivity of ThO2 with defects.
ductivity of UO2 by adding the contribution of a small polaron to Since the thermal transport in actinide oxides is largely dependent
molecular dynamics estimation. Wang et al. [24] reported ~12.5 W/ on the lattice vibrations, any small perturbations to the lattice
m-K and 7.5 W/m-K for the thermal conductivities of UO2 and PuO2 structure is expected to alter phonon transport notably. Park et al.
respectively from first-principles simulations. [25]. also showed that a small hydrogen functionalization brings a
The similar dependency on sample length is observed in UO2 detrimental impact on the thermal conductivity of carbon nano-
and PuO2 with the lowest defect concentrations, i.e. 0.1%. This in- tubes, comparable to the impact caused by large phenyl groups. The
dicates that thermal transport is still ballistic when a small number thermal conductivities of UO2 and PuO2 decrease rapidly by
of defects is introduced in the simulation structure. This is well increasing the concentration of vacancy defects. For example, the
supported by the temperature profiles (Fig. 3); when the defect thermal conductivity of UO2 decreases by nearly 90% when 5% of
concentration is only 0.1% (Fig. 3a and c), slips at the locations of the oxygen vacancy is present in the sample (Fig. 4a).
hot and cold bath are notable unlike the simulation structures with The complete thermal conductivity estimation results for UO2
the highest defect concentrations, i.e. 5% (Fig. 3b and d). However, and PuO2 with Nx ¼ 100 are listed in Table 2. It is apparent that a
the size effect on the thermal conductivity becomes negligible vacancy of primary element degrades thermal conductivity more

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx 5

between their experimental results and the current simulation


results is not possible since it is unattainable to control the amount
of vacancies independently during experiments. For example,
thermogravimetry method allows the control of oxygen-to-metal
ratio of oxide fuels only and does not provide information about
the actual amount of oxygen vacancies; with the same amount of
deficit in oxygen, different amounts of oxygen vacancies are
possible depending on the oxygen interstitials present in the fuel.
Fig. 5 compares the effects of vacancy defects and uranium
substitutional defects on the thermal conductivities of PuO2 and
ThO2. Both PuO2 and ThO2 show that the impact of uranium sub-
stitution on thermal conductivity is nominal compared to any other
vacancy defect. For the case of PuO2 with Nx ¼ 100, 0.5% oxygen
vacancy causes a decrease of 45.6% in thermal conductivity when
compared to stoichiometric PuO2, while 5% uranium substitution
causes a thermal conductivity decrease of only 12.3%. The similar
phenomenon is observed for the case of ThO2; the effect of uranium
substitution on the thermal conductivity of ThO2 is minimal when
compared to the effect of vacancy defects. However, the impact of
the uranium substitution on the thermal conductivity of ThO2 is
much larger when compared to its effect on the thermal conduc-
tivity of PuO2; 5% uranium substitution decreases the thermal
conductivity of ThO2 by 33% while the uranium substitution with
the same concentration decreases the thermal conductivity of PuO2
by 12.3% only. This difference can be attributed to the difference
between the atomic sizes and ionic radii; the atomic mass of
plutonium (239 amu) is nearly the same as the atomic mass of
uranium (238 amu) while the atomic mass of thorium (232 amu) is
slightly different from the atomic mass of uranium (238 amu). The
larger atomic mass interferes with the propagation of phonons
more severely, when compared to the smaller atomic mass. It is also
important to note the difference in ionic radii of each element,
based on its lattice constant, the difference between ThO2 (5.5974 Å
Fig. 4. The thermal conductivities of UO2 and PuO2 with various vacancies as a [30]) and UO2 (5.471 Å [31]) lattice constant is about 12 p.m., while
function of sample length. the difference in PuO2 (5.3975 Å [32]) and UO2 lattice constant is 8
p.m. The effect of the difference in ionic radii is consistent with
other studies conducted involving thermal conductivity and va-
Table 2
Thermal conductivities of UO2 and PuO2 with vacancy. cancy defect relationships as observed in Tada et al. [33], wherein it
is examined that the greater difference in ionic radius results in a
Thermal conductivity, k (W/m-K)
more pronounced change in thermal conductivity. As observed in
% Defect UO2 (Nx ¼ 100) PuO2 (Nx ¼ 100) Klemen's perturbation theory [34], the perturbation energy in
U vacancy O vacancy Pu vacancy O vacancy

0% 14.99 ± 0.42 14.99 ± 0.42 17.15 ± 0.29 17.15 ± 0.29


0.1% 11.30 ± 0.33 12.08 ± 0.47 12.82 ± 0.14 14.32 ± 0.35
0.5% 6.02 ± 0.18 7.46 ± 0.13 6.67 ± 0.20 8.83 ± 0.13
1% 3.92 ± 0.08 5.12 ± 0.09 4.38 ± 0.26 5.82 ± 0.07
2% 2.30 ± 0.05 3.42 ± 0.05 2.22 ± 0.03 3.93 ± 0.12
5% 1.22 ± 0.01 2.03 ± 0.01 1.32 ± 0.01 2.16 ± 0.03

severely than the oxygen vacancy equivalent. For example, 0.1%


uranium vacancies decrease the thermal conductivity of UO2 with
Nx ¼ 100 by 24.6% while the same concentration of oxygen va-
cancies decreases the thermal conductivity by 19.4%. This can be
easily explained by the difference between atomic sizes of oxygen
atom and primary atom, i.e. uranium or plutonium; the absence of
larger atoms is expected to alter the lattice vibrations in the
simulation structure more critically when compared to the absence
of smaller atoms. The same phenomenon has been observed in the
authors’ previous research [15] that studied thorium vacancies and
oxygen vacancies in ThO2, and also constantly observed in other
nanostructured materials [25e27]. Using experiments, moreover,
researchers repeatedly reported that the thermal transport prop- Fig. 5. Comparison between the effects of oxygen vacancy and uranium substitutional
erties of oxide nuclear fuels are degraded by the deviation from the on the thermal conductivity of PuO2 and ThO2. The thermal conductivities are
normalized by the thermal conductivities of their stoichiometric samples. Kmax is the
stoichiometry [28,29]. However, the direct quantitative comparison
thermal conductivity of the stoichiometric sample.

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
6 A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx

phonon scattering by point defects is proportional to the difference stoichiometric PuO2. The peak locations of the PDOS is similarly
between the mass of the substituted atom and average atomic found in other works, such as Antropov et al. [35], where the PDOS
mass. The difference between the mass sizes of thorium and for the solid uranium examined at 300K had a peak around 10 meV.
plutonium compared to uranium supports the data in that the ef- That is similar to the peak locations of UO2 in Fig. 6a and b. The
fect of the uranium substitution is less substantial in PuO2. PDOS for plutonium as observed by Manley et al. [36] has a peak
Important mechanisms that influence thermal transport in location also around 10 meV. This observation is consistent with
actinide oxides with defects include magnitude of phonon fre- the low-energy peak for PuO2 found in this study.
quencies; phonon group velocities; and phonon relaxation times. It is noted that vacancy defects reduce mode degeneracies
Point defects reduce thermal conductivity by limiting intrinsic identified as peaks, especially at energies around 20 meV. However,
phonon mean free path, resulting in reduced phonon relaxation these reduced peaks are not likely to affect the thermal conduc-
times as observed in previous sections; the thermal conductivities tivities of these nuclear fuels much as long as phonon group ve-
of actinide oxides with high defect concentrations reach their bulk locities remain unchanged. In the both UO2 and PuO2, the high
thermal conductivities when the sample length is increased while frequency (energy) peaks (~75 meV) are shifted in all cases
the thermal conductivities of stoichiometric actinide oxides although they are not expected to contribute to the reduction in
continue to increase with an increase in their sample lengths. thermal conductivity because high energy phonons generally have
However, the effect of point defects on magnitude of phonon fre- low group velocities. More important peak shift we have to focus on
quencies and phono group velocities still need to be examined. are observed in low frequencies (energies) below 20 meV. Peak
Although phonon dispersion is a powerful tool that is used to shift (or frequency reduction) is more noticeable for the vacancies
examine these parameters, the production of phonon dispersion of larger atoms (UO2 with uranium vacancy and PuO2 with pluto-
relations is prohibited for defective structures with randomly nium vacancy). This reduction in phonon mode frequencies may
distributed point defects. result from reduced phonon group velocities that are caused by
In light of this, we compare the phonon density of states (PDOS) increased structural compliance with vacancy defects. The phonons
in actinide oxides with different types of defects that are obtained from low frequency (energy) modes contribute more to thermal
from the Fourier transform of the autocorrelation function of atom transport when compared to higher frequency (energy) phonon
velocities. The velocity vectors are obtained at 300K using molec- modes, thus it is expected that the uranium and plutonium va-
ular dynamics simulations together with the force field used in cancies contribute to the previously observed reduction in thermal
RNEMD in the present study. Fig. 6a and b shows the PDOS of UO2 conductivity for both fuel types. When observing the PDOS for UO2
with oxygen and uranium vacancies. The shaded area represents and PuO2 with oxygen vacancies, the same differences between the
the PDOS of stoichiometric UO2. Fig. 6c, d, and e illustrate the PDOS graphs can be seen, however, the degree to which the graphs vary is
of PuO2 with oxygen vacancies, plutonium vacancies, and uranium significantly smaller compared to PDOS for UO2 and PuO2 with
substitutional defects. The shaded area represents the PDOS of vacancies of larger atoms, i.e. uranium or plutonium vacancies. This
implies that group velocities and phonon frequencies are not
largely affected by oxygen vacancies, and reduced phonon relaxa-
tion times are assumed to be the dominant mechanism that limits
the thermal conductivity of UO2 and PuO2 with vacancy defects.
The PDOS in PuO2 with uranium substitutional defects do not
show the same extent of variations in the low energy region as
vacancy defects did. Although there is a slight peak shift in the high
energy region in between 70 meV and 80 meV, it is unlikely to have
an effect on the phonon transport of PuO2 since high-frequency
(energy) phonons carry vibrational energy inefficiently when
compared to low-frequency (energy) phonons because of their low
group velocities; phonon dispersion relations in PuO2 repeatedly
showed that acoustic phonon modes (the most important heat
carrier) have high group velocities when their frequencies are low
[37,38]. This minimal effect of substitutional defects on phonon
spectrum is well supported by our thermal conductivity calcula-
tions that showed a small change in the thermal conductivity of
PuO2 by uranium substitutional defects observed in Fig. 5.
To provide a better understanding on phonon scattering in
defective actinide oxides, we repeat the RNEMD calculations at
sample temperatures varying from 100 K to 1200 K. The charac-
teristic length during the thermal conductivity calculation is fixed
to be Nx ¼ 50. (Note that Table 2 lists the thermal conductivities for
the structures with Nx ¼ 100). Defect concentrations are controlled
to be 0.1%, 1%, and 5% for uranium vacancies and oxygen vacancies.
In stoichiometric samples, as shown in Fig. 7a, high temperature
degrades the thermal conductivity of UO2. This is explained by
Umklapp phonon-phonon scattering. High temperatures increase
the population of phonons with large momenta which often lose
their energy by being collided with other high momentum pho-
Fig. 6. Phonon density of states (PDOS) as a function of phonon energy for (a) UO2 nons. Therefore, this Umklapp phonon-phonon scattering reduces
with oxygen vacancy, (b) UO2 with uranium vacancy, (c) PuO2 with oxygen vacancy,
(d) PuO2 with plutonium vacancy, and (e) PuO2 with uranium substitutional defects.
the phonon mean free path of the system, resulting in decreased
The area plot indicates the PDOS of stoichiometric samples as a function of the energy thermal conductivities. Kim et al. [39] also demonstrated that the
of the phonon. acoustic contribution to the thermal conductivity of UO2 has a

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx 7

conductivity of structures with a large number of defects cannot be


increased much by decreasing the temperature.
In Fig. 7b, thermal conductivity calculations for UO2 and PuO2 by
other researchers are plotted together with the results in this study.
The results in the current study show a good match with the results
by other researchers. However, our thermal conductivity calcula-
tions for PuO2 exhibit a deviation from the calculations by Cooper
et al. [22] although the same interatomic potential has been used.
This can be explained by different simulation factors we have
chosen in this study such as simulation box size, equilibration
methods, and amount of energy swap during RNEMD. The authors
will provide more systematic study on the effect of different
simulation settings on the thermal conductivity calculation by
RNEMD in the near future.

4. Conclusions

Through the application of RNEMD simulations, the thermal


conductivities for three actinide oxides (UO2, PuO2, and ThO2) with
vacancy and uranium substitutional defects were found. The effects
of these defects were observed to be significant across all three
nuclear fuels, and even the smallest percentage of defect has a
noteworthy impact on thermal conductivity. Among the three
defect types examined, the vacancies of primary atoms (uranium,
plutonium, and thorium atoms) decreased the thermal conductiv-
ities of actinide oxide fuels the most, compared to the oxygen va-
cancy and the uranium substitutional defect. In the case of UO2
with 100 unit cell in length, for example, a 0.5% uranium vacancy
decreased the thermal conductivity by 85.4% while the oxygen
vacancy of the same concentration decreased the thermal con-
ductivity by 67.1%. It can be concluded that although oxygen va-
cancies have a substantial negative effect on the thermal
conductivity of actinide oxides, the vacancies of primary atoms
degrade the thermal conductivity of actinide oxides more severely.
Fig. 7. (a) Temperature effect on the thermal conductivity of UO2 with defects. Data are Further conclusions can be drawn by investigating the uranium
curve-fitted using a correlation, k ¼ k100 K/(A þ BT) where k100 K is the thermal con-
ductivity at 100 K, and A and B are fitting parameters. (b) Comparison with other
substitutional defects on PuO2 and ThO2. A 1.0% uranium substi-
experimental (Fukushima et al. [41], Gibby [42]) and simulation results (Arima et al. tutional defect in PuO2 (Nx ¼ 100) results in a thermal conductivity
[3], Cooper et al. [22]). of 16.89 W/m-K; this is a decrease of 1.5% compared to the stoi-
chiometric PuO2 of the same length (17.15 W/m-K). ThO2 of the
same length and the same defect concentration has a thermal
strong T1 dependence and dominates the lattice thermal con- conductivity of 14.02 W/m-K while the thermal conductivity of its
ductivity of UO2. The fitting curve for the stoichiometric sample in stoichiometric sample is 15.96 w/m-K. The resulting difference of
Fig. 7a also demonstrates this T1 dependence. This is explained 12.9% reveals that uranium substitution has a greater effect on ThO2
more precisely by considering dominant phonon wavelength. than PuO2.
Dominant phonon wavelength, ld can be evaluated from ħud ¼ hn/
ld z kBT [40], where ħ is the modified Planck constant, kB is the
Acknowledgements
Boltzmann constant, v is the phonon group velocity, ud is the
dominant phonon frequency, and T is the temperature. As shown in
This study was part of collaborative research effort among
this relation, the dominant phonon wavelength increases with
members of the Nuclear Energy, Science, and Engineering Labora-
decreasing the temperature. Therefore, at low temperatures, long
tory at Kennesaw State University. This work made use of the High
wavelength phonons are expected to dominate the thermal trans-
Performance Computing Resource at Kennesaw State University.
port in the system and they are not likely to participate in Umklapp
scattering because of their small wavenumbers. At high tempera-
tures, however, the dominant phonons are short wavelength pho- Appendix A. Supplementary data
nons that are usually engaged in Umklapp scattering, and thus the
intrinsic phonon mean free path is decreased with an increase in Supplementary data to this article can be found online at
temperature. https://doi.org/10.1016/j.net.2019.03.011.
Unlike the stoichiometric sample, the thermal conductivities of
samples with high defect concentrations do not exhibit a clear References
dependency on temperature (Fig. 7a). The parameter “B” which is
[1] Y. Lu, Y. Yang, P. Zhang, Thermodynamic properties and structural stability of
the fitting parameter related to the temperature dependence is thorium dioxide, J. Phys. Condens. Matter 24 (22) (2012) 225801.
only 0.00016 for the UO2 with 5% oxygen vacancy while the [2] P. Martin, D.J. Cooke, R. Cywinski, A molecular dynamics study of the thermal
parameter “B” is 0.00476 for the stoichiometric UO2. In samples properties of thorium oxide, J. Appl. Phys. 112 (7) (2012), p. 073507.
[3] T. Arima, S. Yamasaki, Y. Inagaki, K. Idemitsu, Evaluation of thermal properties
with high defect concentrations, the activation of long wavelength of UO2 and PuO2 by equilibrium molecular dynamics simulations from 300 to
phonons is physically prohibited. Therefore, the thermal 2000 K, J. Alloy. Comp. 400 (1e2) (2005) 43e50.

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011
8 A. Resnick et al. / Nuclear Engineering and Technology xxx (xxxx) xxx

[4] L. Ma, A.K. Ray, Formation energies and swelling of uranium dioxide by point [24] B.-T. Wang, J.-J. Zheng, X. Qu, W.-D. Li, P. Zhang, Thermal conductivity of UO2
defects, Phys. Lett. 376 (17) (2012) 1499e1505. and PuO2 from first-principles, J. Alloy. Comp. 628 (2015) 267e271.
[5] T. Petit, C. Lemaignan, F. Jollet, B. Bigot, A. Pasturel, Point defects in uranium [25] J. Park, M.F. Bifano, V. Prakash, Sensitivity of thermal conductivity of carbon
dioxide, Phil. Mag. B 77 (3) (1998) 779e786. nanotubes to defect concentrations and heat-treatment, J. Appl. Phys. 113 (3)
[6] C. Duriez, J.-P. Alessandri, T. Gervais, Y. Philipponneau, Thermal conductivity (2013), p. 034312.
of hypostoichiometric low Pu content (U, Pu) O2 x mixed oxide, J. Nucl. [26] H.-p. Li, R.-q. Zhang, Vacancy-defecteinduced diminution of thermal con-
Mater. 277 (2e3) (2000) 143e158. ductivity in silicene, EPL (Europhysics Letters) 99 (3) (2012) 36001.
[7] F. Müller-Plathe, A simple nonequilibrium molecular dynamics method for [27] N. Wei, Y. Chen, K. Cai, J. Zhao, H.-Q. Wang, J.-C. Zheng, Thermal conductivity
calculating the thermal conductivity, J. Chem. Phys. 106 (14) (1997) of graphene kirigami: ultralow and strain robustness, Carbon 104 (2016)
6082e6085. 203e213.
[8] M. Cooper, M. Rushton, R. Grimes, A many-body potential approach to [28] R. Kavazauri, S. Pokrovskiy, V. Baranov, A. Tenishev, Thermal properties of
modelling the thermomechanical properties of actinide oxides, J. Phys. Con- nonstoichiometry uranium dioxide, in: IOP Conference Series: Materials Sci-
dens. Matter 26 (10) (2014) 105401. ence and Engineering, vol. 130, IOP Publishing, 2016 no. 1, p. 012025.
[9] M.W. Cooper, S.T. Murphy, P.C. Fossati, M.J. Rushton, R.W. Grimes, Thermo- [29] T. Pavlov, et al., Measurement and interpretation of the thermo-physical
physical and anion diffusion properties of (Ux, Th1x) O2, in: Proceedings of properties of UO2 at high temperatures: the viral effect of oxygen defects,
the Royal Society of London A: Mathematical, Physical and Engineering Sci- Acta Mater. 139 (2017) 138e154.
ences, vol. 470, The Royal Society, 2014, p. 20140427, no. 2171. [30] T. Yamashita, N. Nitani, T. Tsuji, H. Inagaki, Thermal expansions of NpO2 and
[10] M. Cooper, S. Murphy, M. Rushton, R. Grimes, Thermophysical properties and some other actinide dioxides, J. Nucl. Mater. 245 (1) (1997) 72e78.
oxygen transport in the (U x, Pu 1x) O 2 lattice, J. Nucl. Mater. 461 (2015) [31] G. Leinders, T. Cardinaels, K. Binnemans, M. Verwerft, Accurate lattice
206e214. parameter measurements of stoichiometric uranium dioxide, J. Nucl. Mater.
[11] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, 459 (2015) 135e142.
J. Comput. Phys. 117 (1) (1995) 1e19. [32] J. Haschke, T.H. Allen, L.A. Morales, Surface and corrosion chemistry of
[12] M. Qin, et al., Thermal conductivity and energetic recoils in UO2 using a plutonium, Los Alamos Sci. 26 (2) (2000) 252e273.
many-body potential model, J. Phys. Condens. Matter 26 (49) (2014) 495401. [33] M. Tada, M. Yoshiya, H. Yasuda, Effect of ionic radius and resultant two-
[13] M. Rahman, B. Szpunar, J. Szpunar, The induced anisotropy in thermal con- dimensionality of phonons on thermal conductivity in M x CoO 2 (M¼ Li,
ductivity of thorium dioxide and cerium dioxide, Mater. Res. Express 4 (7) Na, K) by perturbed molecular dynamics, J. Electron. Mater. 39 (9) (2010)
(2017), p. 075512. 1439e1445.
[14] J. Park, E.B. Farf
an, C. Enriquez, Thermal transport in thorium dioxide, Nuclear [34] P. Klemens, The scattering of low-frequency lattice waves by static imper-
Engineering and Technology 50 (2018) 731e737. fections, Proc. Phys. Soc. 68 (12) (1955) 1113.
[15] J. Park, E.B. Farfan, K. Mitchell, A. Resnick, C. Enriquez, T. Yee, Sensitivity of [35] A. Antropov, K. Fidanyan, V. Stegailov, Phonon density of states for solid
thermal transport in thorium dioxide to defects, J. Nucl. Mater. 504 (2018) uranium: accuracy of the embedded atom model classical interatomic po-
198e205. tential, in: Journal of Physics: Conference Series, vol. 946, IOP Publishing, 2018
[16] B. Willis, Structures of UO2, UO2þ x andU4O9 by neutron diffraction, J. Phys. no. 1, p. 012094.
25 (5) (1964) 431e439. [36] M. Manley, et al., Phonon density of states of a-and d-plutonium by inelastic
[17] P.K. Schelling, S.R. Phillpot, P. Keblinski, Comparison of atomic-level simula- x-ray scattering, Phys. Rev. B 79 (5) (2009), p. 052301.
tion methods for computing thermal conductivity, Phys. Rev. B 65 (14) (2002) [37] M. Manley, et al., Measurement of the phonon density of states of PuO 2 (þ 2%
144306. Ga): a critical test of theory, Phys. Rev. B 85 (13) (2012) 132301.
[18] T. Watanabe, S.B. Sinnott, J.S. Tulenko, R.W. Grimes, P.K. Schelling, [38] P. Zhang, B.-T. Wang, X.-G. Zhao, Ground-state properties and high-pressure
S.R. Phillpot, Thermal transport properties of uranium dioxide by molecular behavior of plutonium dioxide: density functional theory calculations, Phys.
dynamics simulations, J. Nucl. Mater. 375 (3) (2008) 388e396. Rev. B 82 (14) (2010) 144110.
[19] M. Cooper, S. Middleburgh, R. Grimes, Modelling the thermal conductivity of [39] H. Kim, M.H. Kim, M. Kaviany, Lattice thermal conductivity of UO2 using ab-
(U x Th 1x) O 2 and (U x Pu 1 x) O 2, J. Nucl. Mater. 466 (2015) 29e35. initio and classical molecular dynamics, J. Appl. Phys. 115 (12) (2014) 123510.
[20] J. Park, V. Prakash, Phonon scattering and thermal conductivity of pillared [40] R. Prasher, T. Tong, A. Majumdar, An acoustic and dimensional mismatch
graphene structures with carbon nanotube-graphene intramolecular junc- model for thermal boundary conductance between a vertical mesoscopic
tions, J. Appl. Phys. 116 (1) (2014) 014303. nanowire/nanotube and a bulk substrate, J. Appl. Phys. 102 (10) (2007), pp.
[21] J. Park, V. Prakash, Thermal transport in 3D pillared SWCNTegraphene 104312e10.
nanostructures, J. Mater. Res. 28 (7) (2013) 940e951. [41] S. Fukushima, T. Ohmichi, A. Maeda, H. Watanabe, The effect of yttrium
[22] M. Cooper, S. Middleburgh, R. Grimes, Modelling the thermal conductivity of content on the thermal conductivity of near-stoichiometric (u,y) o2 solid
(UxTh1x) O2 and (UxPu1x) O2, J. Nucl. Mater. 466 (2015) 29e35. solutions, J. Nucl. Mater. 102 (1e2) (1981) 30e39.
[23] J. Fink, Thermophysical properties of uranium dioxide, J. Nucl. Mater. 279 (1) [42] R. Gibby, The effect of plutonium content on the thermal conductivity of (U,
(2000) 1e18. Pu) O2 solid solutions, J. Nucl. Mater. 38 (2) (1971) 163e177.

Please cite this article as: A. Resnick et al., Thermal transport study in actinide oxides with point defects, Nuclear Engineering and Technology,
https://doi.org/10.1016/j.net.2019.03.011

S-ar putea să vă placă și