Better ways are needed to capture radioactive volatile fission products (Kr, Xe, Br, I, Te, Rb, and Cs) discharged during the reprocessing of spent nuclear fuel in order to reduce the volumes of produced waste and minimize environmental impact. Using density functional theory, we examine the efficacy of a one-dimensional yttrium silicide electride (Y5Si3:e) as a host matrix to encapsulate these species. Endoergic encapsulation energies calculated for Kr, Xe, Rb, and Cs imply they are not captured by Y5Si3:e. Encapsulation is exoergic for Br, I, and Te with respect to their atoms and dimers as reference states, meaning that they can be captured effectively due to their high electronegativities. This is further supported by the formation of anions due to charge transfer between Y5Si3:e and Br (I and Te). The selectivity of this material for these volatile species makes it promising for use in nuclear filters.

Spent nuclear fuel is reprocessed to extract the unused fissile 235U and reduce waste volumes.1–5 This gives rise to a substantial volume of fission product containing effluents. If fission products can be separated, the effluent can be treated as a lower hazard material. The fission products (Xe, Kr, Br, I, Rb, Cs, and Te) are of particular concern as they can escape from the fuel as a consequence of various issues, including nuclear power plant accidents and natural crisis. Radioactive iodine (131I) is important because it can be absorbed at high concentration by the human thyroid gland leading to several thyroid disorders.3,6,7 Radioactive cesium (137Cs) is also of concern as it causes damage to human organs, such as kidney and lungs.8–10 Thus, special provision must be made for effective and safe disposal of volatile fission products in order to reduce the risk of release.

Impregnated carbon or charcoal filters are most commonly used to trap iodine in a nuclear plant.11 The preference for this type of filter is due to its large trapping surface area and low production cost.11 Alternative materials, such as silica, zeolites, alumina, silver nitrate, metal organic frameworks, and porous organic polymers, have been examined in order to find a more efficient trapping medium.12–15 The efficiency of the filters depends upon the properties of the filter material: their thermal and mechanical stability, surface area, and chemical species specificity. The search for alternative filter materials continues in order to reduce the potential impact on the environment and humans.

Electrides are a class of materials where electrons serve as loosely bound anions.16 Based on electron localization, they fall into three categories: zero-dimensional (electrons residing in cavities),17,18 one-dimensional (electrons in a channel),19,20 and two-dimensional (electrons localizing in a layer).21 A zero-dimensional “mayenite” type electride, [Ca24Al28O64]4+•(e)4, has been widely studied due to its utility as a catalyst for NH3 synthesis and CO2 depletion, an electron emitter, and a superconductor.22–28 Different types of atoms and molecules have been encapsulated experimentally to examine the effectiveness of encapsulation and studied theoretically to understand how encapsulation modifies the electronic properties of this electride.27–32 

Zhang et al.20 first reported the structure of the apatite based one-dimensional electride [La8Sr2(SiO4)6]4+•(e)4, though its practical utility had not yet been established. An experimental report of the two-dimensional dicalcium nitride electride [(Ca2N)+•e] by Lee et al.21 shows that this electride is stable at room temperature and exhibits an open layer structure with high electron concentration and a low work function of 2.6 eV. This electride has been tested with regard to application in sodium ion batteries33 and solid lubricants.34 

The air and water stable one-dimensional electride yttrium silicide (Y5Si3:e) was first reported by Lu et al.35 This material is a candidate for trapping volatile fission products due to its chemical stability, facilitated by the strong hybridization between yttrium 4d electrons and anionic electrons available in the quasi-one-dimensional channels. The number of anionic electrons per formula is predicted to be 0.79 with the effective formula of the electride then described as [Y5Si3]0.79+:0.79e.35 Ru-loaded Y5Si3:e is catalytically highly efficient for ammonia synthesis because the anionic electrons enhance the nitrogen cleavage and reduce the associated activation energy.35 

In this study, spin-polarized mode density functional theory (DFT) is used to calculate the structures associated with volatile fission products (Kr, Xe, Br, I, Rb, Cs, and Te) encapsulated in Y5Si3:e, their associated encapsulation energies, charge transfer, electronic structures, and charge densities.

All calculations were performed using the DFT code VASP (Vienna ab initio Simulation Package).36 This code solves standard Kohn–Sham equations using plane wave basis sets and projected augmented wave (PAW) pseudopotentials.37 In the PAW method, projectors and auxiliary functions are introduced similar to the ultra-soft pseudopotential method and the total energy function includes auxiliary functions. Furthermore, this method keeps the full all-electron wave function as opposed to the other pseudopotential methods in which only valence pseudowave functions are kept. In all calculations, a plane wave basis set with a cutoff of 500 eV and an 4 × 4 × 8 Monkhorst–Pack38 k-point mesh were employed. Further increase in the k-points to 6 × 6 × 12 resulted in a total energy difference of only 0.8 meV per atom. The exchange–correlation energy was modeled using the generalized gradient approximation (GGA) scheme as defined by Perdew, Burke, and Ernzerhof (PBE).39 All defect calculations were performed using a 2 × 2 × 2 supercell containing 128 atoms (Y:80 and Si:48). The conjugate gradient algorithm40 was used to perform full geometry optimization (both atom positions and lattice constants were relaxed simultaneously). In all relaxed configurations, forces on the atoms were less than 0.001 eV/Å. The following equation was used to calculate single atom encapsulation energies within Y5Si3:e:

EEnc=E(X@Y5Si3:e)E(Y5Si3:e)E(X),
(1)

where E(X@Y5Si3:e) is the total energy of a single fission atom encapsulated in a 2 × 2 × 2 supercell of Y5Si3:e, E(Y5Si3:e)is the total energy of a 2 × 2 × 2 supercell of Y5Si3:e, and E(X) is the energy of an isolated gas phase fission atom. A semiempirical dispersion term was used to describe short-range interactions.41 

At room temperature, yttrium silicide exhibits an hexagonal phase with the space group P63/mcm (193), a = b = 8.403 Å, c = 6.303 Å, α = β = 90°, and γ = 120°, as shown in Fig. 1(a).42 There are two non-equivalent Y sites. The first Y site forms three coordination with adjacent Si atoms. The second Y is bonded to six nearest neighbor Si atoms forming a six-coordinate geometry. The crystal structure of Y5Si3:e was relaxed under constant pressure to obtain equilibrium lattice constants to validate the quality of the basis sets and pseudopotentials. Table I reports the calculated lattice parameters together with experimental values: it is evident that there is a good agreement between the calculated and experimental values.

FIG. 1.

(a) The relaxed structure of bulk Y5Si3:e, (b) one-dimensional electrons confined within the crystal, and (c) DOS plot of bulk Y5Si3:e. Black dashed lines correspond to the Fermi energy level.

FIG. 1.

(a) The relaxed structure of bulk Y5Si3:e, (b) one-dimensional electrons confined within the crystal, and (c) DOS plot of bulk Y5Si3:e. Black dashed lines correspond to the Fermi energy level.

Close modal
TABLE I.

Calculated and experimental lattice parameters of the hexagonal crystal structure (space group P63/mcm) of Y5Si3:e together with corresponding experimental values.

ParameterCalc.Expt.42 |Δ|(%)
a = b (Å) 8.385 8.403 0.21 
c (Å) 6.309 6.303 0.09 
α = β (deg) 90.0 90.0 0.00 
γ (deg) 120.0 120.0 0.00 
V (Å3383.96 385.43 0.38 
ParameterCalc.Expt.42 |Δ|(%)
a = b (Å) 8.385 8.403 0.21 
c (Å) 6.309 6.303 0.09 
α = β (deg) 90.0 90.0 0.00 
γ (deg) 120.0 120.0 0.00 
V (Å3383.96 385.43 0.38 

The one-dimensional electron distribution and the total density of states (DOS) plots are shown in Figs. 1(b) and 1(c), respectively. Y5Si3:e exhibits metallic behavior in agreement with a DFT study performed by Lu et al.35 

Relaxed structures of Y5Si3:e with encapsulated Kr and Xe are shown in Fig. 2. Endoergic encapsulation energies are calculated for both Kr and Xe inferring their instability within the lattice (see Table II). The encapsulation energy calculated for Kr is less unfavorable than Xe due to the smaller atomic radius of Kr than Xe.43 This is further evidenced by the shorter Y–Kr bond length and smaller volume change predicted for Kr (see Table II). According to Bader charge analysis,44 both noble gas atoms gain a small negative charge (Kr: −0.40 and Xe: −0.48) due to the polarization between the one-dimensional holes along the c axis and the noble gas atoms. The smaller value for Kr reflects the smaller polarizability of Kr.

FIG. 2.

(a) Relaxed structure of Kr encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Kr, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for Xe encapsulated within Y5Si3:e.

FIG. 2.

(a) Relaxed structure of Kr encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Kr, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for Xe encapsulated within Y5Si3:e.

Close modal
TABLE II.

Calculated electron affinities of Kr and Xe, encapsulation energies (calculated using the isolated gas phase atom as the reference), Bader charges on the encapsulated atoms, the shortest Y–X bond distances (X = Kr and Xe), and relative volume changes upon encapsulation.

Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV)Bader charge (|e|)Y–X (Å) (X = Kr and Xe)Relative volume change (Δ%)
Kr 2.02 1.08 2.84 −0.40 2.90 0.62 
Xe 2.16 1.24 3.89 −0.48 2.95 0.97 
Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV)Bader charge (|e|)Y–X (Å) (X = Kr and Xe)Relative volume change (Δ%)
Kr 2.02 1.08 2.84 −0.40 2.90 0.62 
Xe 2.16 1.24 3.89 −0.48 2.95 0.97 

Total DOS plots predicted for structures with Xe and Kr are almost identical to the DOS plot of encapsulant free Y5Si3:e (compare Figs. 1 and 2). Atomic DOS plots show that p-states belonging to Kr and Xe appear deep (∼−5 eV) in the valence band. Charge density plots show that encapsulation resulted in no charge transfer from the lattice to Kr (or Xe).

The encapsulation energies for Br and I are highly negative (favorable) with respect to their isolated atom reference states. This is due to the strong electron affinities of these species (see Table III), reflected in the negative Bader charges. Both Br and I gain almost one electron from the extra-framework electrons localized in the one-dimensional channel, meaning that both Br and I are tending toward Br and I, respectively, close to completing their p-shells (p6). The encapsulation energy of Br is more favorable than that of I due to its smaller atomic radius.43 This is reflected in the shorter bond length of Y–Br than Y–I.

TABLE III.

Calculated electron affinities of Br and I, encapsulation energies (calculated using the isolated gas phase atoms and dimers as reference states), Bader charges on the encapsulated atoms, the shortest Y–X bond distances (X = Br and I), and relative volume changes upon encapsulation.

Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV/atom)Bader charge (|e|)Y–X (Å) (X = Br and I)Relative volume change (%)
AtomDimer (12 X2)
Br 1.83 4.97 −3.61 −2.35 −1.03 2.80 0.47 
1.98 4.65 −1.89 −0.77 −1.05 2.90 0.92 
Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV/atom)Bader charge (|e|)Y–X (Å) (X = Br and I)Relative volume change (%)
AtomDimer (12 X2)
Br 1.83 4.97 −3.61 −2.35 −1.03 2.80 0.47 
1.98 4.65 −1.89 −0.77 −1.05 2.90 0.92 

The exothermic endoergic encapsulation energy calculated using ½X2 (X = Br and I) as the reference indicates that both Br and I are still stable inside Y5Si3:e, despite there being an energy penalty to dissociate their diatomic molecules. The calculated dissociation energies (per atom) for Br2 and I2 are 1.26 and 1.12 eV, respectively. Despite the higher dimer dissociation energy for Br than I, Br exhibits a strong encapsulation energy.

Although encapsulation of Br and I results in a reduction of states associated with free anionic electrons at the Fermi level in the total DOS plots, at this concentration of encapsulation, these materials retain their metallic character (see Fig. 3). There is, however, an accumulation of charge density around the encapsulated atoms in the one-dimensional channel. Encapsulation introduces a slight increase in the overall volume. The larger atomic radius of I is reflected in a larger increase in the total volume (see Table III).

FIG. 3.

(a) Relaxed structure of Br encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Br, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for the I encapsulated Y5Si3:e.

FIG. 3.

(a) Relaxed structure of Br encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Br, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for the I encapsulated Y5Si3:e.

Close modal

Te demonstrates a strong negative (favorable) encapsulation energy (see Table IV). This is due to the transfer of 1.42 electrons to Te. The large Bader charge of 1.42 |e| means that Te is tending to the stable Te2− ion electronic configuration. A highly exothermic encapsulation energy is also calculated with respect to the dimer, despite the strong Te2 dissociation energy of 1.46 eV per atom.

TABLE IV.

Calculated electron affinity of Te, encapsulation energy (calculated using the isolated gas phase atom and dimer as reference states), Bader charges on the Te atom, the shortest Y–Te bond distances, and relative volume changes upon encapsulation.

Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV/atom)Bader charge (|e|)Y–Te (Å)Relative volume change (%)
AtomDimer (12 Te2)
Te 2.06 3.49 −3.64 −1.88 −1.42 2.87 0.96 
Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV/atom)Bader charge (|e|)Y–Te (Å)Relative volume change (%)
AtomDimer (12 Te2)
Te 2.06 3.49 −3.64 −1.88 −1.42 2.87 0.96 

At this level of Te encapsulation, the material maintains its metallic character [see Fig. 4(b)]. That is, free electrons are left in the one-dimensional channel. Though a single Te atom can gain only part of the electron density from the channel, the volume increase for Te is larger than for I, due to the larger size of Te2− than I. The accumulation of electrons by Te is confirmed in the charge density plot where charge density is more localized on Te than the remaining lattice.

FIG. 4.

(a) Relaxed structure of Te encapsulated within Y5Si3:e, (b) total DOS plot, (c) constant charge density plot showing electron distribution upon encapsulation, and (d) atomic DOS plot of Te.

FIG. 4.

(a) Relaxed structure of Te encapsulated within Y5Si3:e, (b) total DOS plot, (c) constant charge density plot showing electron distribution upon encapsulation, and (d) atomic DOS plot of Te.

Close modal

The relaxed configurations associated with Rb and Cs encapsulation are shown in Fig. 5 with the encapsulation energies and the Bader charges reported in Table V. Both Rb and Cs exhibit positive encapsulation energies (2.28 and 2.51 eV, respectively). This is due to their low electron affinities and large size. Bader charge analysis shows that both Rb and Cs are polarized with small positive charges (0.29 and 0.01, respectively). While Rb and Cs readily donate their outmost s1 electrons to form stable noble gas electronic configurations, this is not the case in this electride material as channel space has already accommodated electrons.

FIG. 5.

(a) Relaxed structure of Rb encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Rb, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for the Cs encapsulated Y5Si3:e.

FIG. 5.

(a) Relaxed structure of Rb encapsulated within Y5Si3:e, (b) total DOS plot, (c) atomic DOS plot of Rb, (d) constant charge density plot showing electron distribution upon encapsulation, and (e)–(h) corresponding structures and plots calculated for the Cs encapsulated Y5Si3:e.

Close modal
TABLE V.

Calculated electron affinities of Rb and Cs, encapsulation energies (calculated using the isolated gas phase atoms), Bader charges on the encapsulated atoms, the shortest Y–X bond distances (X = Rb and Cs), and relative volume changes upon encapsulation.

Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV)Bader charge (|e|)Y–X (Å) (X = Rb and Cs)Relative volume change (%)
Rb 3.03 2.57 2.28 +0.29 3.05 0.93 
Cs 3.43 2.22 2.51 +0.01 3.05 0.96 
Fission productAtomic radius (Å)43 Electron affinity (eV)Encapsulation energy (eV)Bader charge (|e|)Y–X (Å) (X = Rb and Cs)Relative volume change (%)
Rb 3.03 2.57 2.28 +0.29 3.05 0.93 
Cs 3.43 2.22 2.51 +0.01 3.05 0.96 

The total DOSs calculated for Rb and Cs containing materials remain metallic (see Fig. 5). Both total DOS and charge distributions are not significantly altered in either case.

The relaxed structures of Br2, I2, and Te2 dimers occupying channel positions are shown in Fig. 6. The encapsulation energies and Bader charges are reported in Table VI. Encapsulation energies calculated using the molecular reference state predict that the encapsulation is favorable and only slightly less favorable than when these anions are encapsulated as separated species (compare with energies in Tables III and IV). The molecular dimer bond lengths inside the channel are significantly longer than their corresponding gas phase dimers (which are 2.31, 2.69, and 2.58 Å for Br2, I2, and Te2, respectively). Furthermore, negative Bader charges (see Table VI) on the atoms of dimers confirm the formation of adjacent anion pairs rather than molecules. Nevertheless, this suggests that the electride is capable of accommodating high concentrations of Br and I as anions.

FIG. 6.

Relaxed structures of (a) Br2, (b) I2, and (c) Te2 encapsulated in Y5Si3:e.

FIG. 6.

Relaxed structures of (a) Br2, (b) I2, and (c) Te2 encapsulated in Y5Si3:e.

Close modal
TABLE VI.

Encapsulation energies and Bader charges calculated for dimers encapsulated in Y5Si3:e.

ReactionEncapsulation energy (eV/atom) with respect to dimerBader charges |e|
on both Br or I or Te atoms
Br2 + Y5Si3:e → Br2• Y5Si3:e −2.23 −0.98, −0.97 
I2 + Y5Si3:e → I2• Y5Si3:e −0.58 −0.98, −0.97 
Te2 + Y5Si3:e → Te2• Y5Si3:e −1.75 −1.32, −1.31 
ReactionEncapsulation energy (eV/atom) with respect to dimerBader charges |e|
on both Br or I or Te atoms
Br2 + Y5Si3:e → Br2• Y5Si3:e −2.23 −0.98, −0.97 
I2 + Y5Si3:e → I2• Y5Si3:e −0.58 −0.98, −0.97 
Te2 + Y5Si3:e → Te2• Y5Si3:e −1.75 −1.32, −1.31 

The encapsulation of fission products is expected to take place via the surface of the Y5Si3:e with the involvement of kinetic barrier. There is lack of experimental date available on the surface structure of this electride leaving the modeling of surface structures difficult for this electride. Future work should model the surface structures and calculate the activation energies and the diffusion pathways for these fission products.

Atomic scale simulation based on DFT has been used to examine the thermodynamic efficacy of Y5Si3 as a filter material to encapsulate volatile fission products. It is found that encapsulation is unfavorable for Kr, Xe, Rb, and Cs. Conversely, strong encapsulation energies are predicted for Br, I, and Te with significant charge transfer from this electride's confined electrons to the encapsulated species. This results in the formation of stable Br, I, and Te2− anions. Encapsulation of Br2, I2, and Te2 dimers resulted in the dissociation of molecules and the formation of separated anion pairs. In conclusion, Y3Si3 electride is a promising filter material to trap Br, I, and Te from effluent gases released in the processing of spent nuclear fuel.

Computational facilities and support were provided by the High Performance Computing Centre at Imperial College London. The authors declare that there is no competing financial interest.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
V. M.
Erfrmenkov
,
IAEA Bull.
4
,
37
(
1989
).
2.
P. A.
Baisden
,
C. E.
Atkins-Duffin
 et al, in
Handbook of Nuclear Chemistry
, edited by
A.
Vértes
,
S.
Nagy
, and
Z.
Klencsár
(
Springer US
,
Boston
,
MA
,
2011
), p.
2797
.
3.
M. I.
Ojovan
and
W. E.
Lee
,
An Introduction to Nuclear Waste Immobilisation
, 2nd ed. (
Elsevier
,
Oxford
,
2014
).
4.
E. D.
Collins
,
G. D.
Del Cul
, and
B. A.
Moyer
, in
Advanced Separation Techniques for Nuclear Fuel Reprocessing and Radioactive Waste Treatment
, edited by
K. L.
Nash
and
G. J.
Lumetta
(
Woodhead Publishing
,
2011
), p.
201
.
5.
C.
Corkhill
and
N.
Hyatt
,
Nuclear Waste Management
(
IOP Publishing
,
2018
), p.
1
.
6.
B. H.
Hamling
and
G. F.
Jenkins
,
J. Air Pollut. Control Assoc.
7
,
256
(
1958
).
7.
S.
Hertz
,
A.
Roberts
, and
R. D.
Evans
,
Proc. Soc. Exp. Biol. Med.
38
(
4
),
510
(
1938
).
8.
C. D.
Whitney
and
S.
Landsberger
,
J. Radioanal. Nucl. Chem.
280
,
281
(
2009
).
9.
K.
Knebel
,
J.
Jokiniemi
, and
P. D.
Bottomley
,
J. Nucl. Sci. Technol.
56
,
772
(
2019
).
10.
J.
Aaseth
,
V. M.
Nurchi
, and
O.
Andersen
,
Biomolecules
9
,
856
(
2019
).
11.
A.
Karhu
, “
Methods to prevent the source term of methyl lodide during a core melt accident
(IAEA, Denmark, 1999), Vol. 31(27), p. 66; available at
https://inis.iaea.org/search/search.aspx?orig_q=RN:31031774
12.
J.
Huve
,
A.
Ryzhikov
,
H.
Nouali
,
V.
Lalia
,
G.
Augé
, and
T. J.
Daou
,
RSC Adv.
8
,
29248
(
2018
).
13.
D. F.
Sava
,
M. A.
Rodriguez
,
K. W.
Chapman
,
P. J.
Chupas
,
J. A.
Greathouse
,
P. S.
Crozier
, and
T. M.
Nenoff
,
J. Am. Chem. Soc.
133
,
12398
(
2011
).
14.
B. H. M.
Billinge
,
J. B.
Docherty
, and
M. J.
Bevan
,
Carbon
22
,
83
(
1984
).
15.
R. B.
Hahn
,
S.
Levin
, and
R. L.
Friedlander
,
J. Am. Water Works Assn.
50
(
11
),
1499
(
1958
).
17.
A.
Ellaboudy
,
J. L.
Dye
, and
P. B.
Smith
,
J. Am. Chem. Soc.
105
,
6490
(
1983
).
18.
S.
Matsuishi
,
Y.
Toda
,
M.
Miyakawa
,
K.
Hayashi
,
T.
Kamiya
,
M.
Hirano
,
I.
Tanaka
, and
H.
Hosono
,
Science
301
,
626
(
2003
).
19.
J.
Wang
,
K.
Hanzawa
,
H.
Hiramatsu
,
J.
Kim
,
N.
Umezawa
,
K.
Iwanaka
,
T.
Tada
, and
H.
Hosono
,
J. Am. Chem. Soc.
139
,
15668
(
2017
).
20.
Y.
Zhang
,
Z.
Xiao
,
T.
Kamiya
, and
H.
Hosono
,
J. Phys. Chem. Lett.
6
,
4966
(
2015
).
21.
K.
Lee
,
S. W.
Kim
,
Y.
Toda
,
S.
Matsuishi
, and
H.
Hosono
,
Nature
494
,
336
(
2013
).
22.
T.
Tada
,
S.
Takemoto
,
S.
Matsuishi
, and
H.
Hosono
,
Inorg. Chem.
53
,
10347
(
2014
).
23.
W.
Ming
,
M.
Yoon
,
M.-H.
Du
,
K.
Lee
, and
S. W.
Kim
,
J. Am. Chem. Soc.
138
,
15336
(
2016
).
24.
Y.
Toda
,
H.
Hirayama
,
N.
Kuganathan
,
A.
Torrisi
,
P. V.
Sushko
, and
H.
Hosono
,
Nat. Commun.
4
,
2378
(
2013
).
25.
M.
Kitano
,
S.
Kanbara
,
Y.
Inoue
,
N.
Kuganathan
,
P. V.
Sushko
,
T.
Yokoyama
,
M.
Hara
, and
H.
Hosono
,
Nat. Commun.
6
,
6731
(
2015
).
26.
N.
Kuganathan
,
H.
Hosono
,
A. L.
Shluger
, and
P. V.
Sushko
,
J. Am. Chem. Soc.
136
,
2216
(
2014
).
27.
E.
Feizi
and
A. K.
Ray
,
J. Disp. Technol.
12
,
451
(
2016
).
28.
M.
Hara
,
M.
Kitano
, and
H.
Hosono
,
ACS Catal.
7
,
2313
(
2017
).
29.
N.
Kuganathan
,
A.
Chroneos
, and
R. W.
Grimes
,
Sci. Rep.
9
,
13612
(
2019
).
30.
N.
Kuganathan
,
R. W.
Grimes
, and
A.
Chroneos
,
J. Appl. Phys.
125
,
165103
(
2019
).
31.
F.
Hayashi
,
Y.
Tomota
,
M.
Kitano
,
Y.
Toda
,
T.
Yokoyama
, and
H.
Hosono
,
J. Am. Chem. Soc.
136
,
11698
(
2014
).
32.
C.
Song
,
J.
Sun
,
S.
Qiu
,
L.
Yuan
,
J.
Tu
,
Y.
Torimoto
,
M.
Sadakata
, and
Q.
Li
,
Chem. Mater.
20
,
3473
(
2008
).
33.
G.
Chen
 et al,
ACS Appl. Mater. Interfaces
9
,
6666
(
2017
).
34.
J.
Wang
,
L.
Li
,
Z.
Shen
,
P.
Guo
,
M.
Li
,
B.
Zhao
,
L.
Fang
, and
L.
Yang
,
Materials
11
,
2462
(
2018
).
35.
Y.
Lu
,
J.
Li
,
T.
Tada
,
Y.
Toda
,
S.
Ueda
,
T.
Yokoyama
,
M.
Kitano
, and
H.
Hosono
,
J. Am. Chem. Soc.
138
,
3970
(
2016
).
36.
G.
Kresse
and
J.
Furthmüller
,
Phys. Rev. B
54
,
11169
(
1996
).
37.
P. E.
Blöchl
,
Phys. Rev. B
50
,
17953
(
1994
).
38.
H. J.
Monkhorst
and
J. D.
Pack
,
Phys. Rev. B
13
,
5188
(
1976
).
39.
J. P.
Perdew
,
Int. J. Quantum Chem.
28
,
497
(
1985
).
40.
W. H.
Press
,
S. A.
Teukolsky
,
W. T.
Vetterling
, and
B. P.
Flannery
,
Numerical Recipes in C: The Art of Scientific Computing
, 2nd ed. (
Cambridge University Press
,
1992
).
41.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
,
J. Chem. Phys.
132
,
154104
(
2010
).
42.
E.
Parthé
,
Acta Crystallogr.
13
,
868
(
1960
).
43.
M.
Mantina
,
A. C.
Chamberlin
,
R.
Valero
,
C. J.
Cramer
, and
D. G.
Truhlar
,
J. Phys. Chem. A
113
,
5806
(
2009
).
44.
R. F. W.
Bader
,
Theor. Chem. Acc.
105
,
276
(
2001
).