Trapping of volatile ssion products by C
60
Navaratnarajah Kuganathan
a
,
*
gens and noble gases [1]. Among these, species such as iodine and
toxic impact [2]. To minimize the hazard, radioactive species should
species [3]. However, other porous materials including zeolites,
), consists of 60 carbon atoms and has been widely
studied [6]. The interior and exterior of a C
60
molecule are sur-
rounded by an electron charge density of conjugate
p
bonds each
formed between two adjacent carbon atoms. The charge density of
the inner space is therefore higher than the outer surface. The
encapsulation of species within C
60
has been described in various
experimental and theoretical studies: alkali metals (e.g. Li, K) [7,8],
alkali earth metals (e.g. Ca, Sr and Ba) [7,8], radioactive isotopes (e.g.
159
Gd,
161
Tb) [9,10], actinide metals (e.g. Lr) [11], non-metals (e.g. N,
P) [12,13], noble gases (e.g. Ne, Ar) [14,15], transition metals (e.g. Tc,
Fe) [16,17] and rare earth metals (e.g. La) [18]. These have all been
considered either in the form of atoms or clusters. The external
surface of C
60
has been studied theoretically for its interaction with
transition metal atoms and clusters, and shown to provide a good
catalyst for the activation of di-nitrogen molecules [19] and the
storage of hydrogen [20]. Furthermore, the theoretical study by
Ozdamar et al. [21] demonstrated the stability and electronic
properties of Pt and Pd in the form of atoms and dimers interacting
with the outer surface of C
60
. Likewise, hydrogen absorption has
* Corresponding author.
E-mail address: n.kuganathan@imperial.ac.uk (N. Kuganathan).
Contents lists available at ScienceDirect
Carbon
journal homepage: www.elsevier.com/locate/carbon
https://doi.org/10.1016/j.carbon.2018.02.098
0008-6223/© 2018 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Carbon 132 (2018) 477e485
been considered on the surface of Pd-decorated C
60
by El Mahdy
[22]. Thus, C
60
provides different sites to capture species both
internally (endohedral) and externally (exohedral).
This study uses density functional theory (DFT) together with
dispersion corrections, to investigate the thermodynamical stabil-
ity of gaseous volatile ssion products (Kr, Xe, Br, I, Rb, Cs and Te) in
order to understand whether they can be trapped by lters that
incorporate C
60
. For those species that may become bound to C
60
,
calculations have been performed to establish whether there is a
preference for positions inside or on the surface of the fullerene
cages and whether they become bound as atoms or in the form of
dimers. DFT calculations, in addition to giving structural informa-
tion, provide data on electronic properties, which are also
discussed.
2. Computational methods
Calculations were carried out using the spin-polarized mode of
DFT as implemented in the VASP [23,24] package. The exchange-
correlation term was modelled using the generalized gradient
approximation (GGA) parameterized by Perdew, Burke, and Ern-
zerhof (PBE) [25]. In all cases, we have used a plane-wave basis set
with a cut-off value of 500 eV. Structural optimisations were per-
formed using a conjugate gradient algorithm [26] and the forces on
the atoms were obtained via the Hellman-Feynmann theorem
including Pulay corrections. In all optimized structures, forces on
the atoms were smaller than 0.001 eV/Å and all the values in the
atomic stress tensor were less than 0.0 02 GPa. All calculations were
performed without using symmetry restrictions. All geometry op-
timisations were performed using a single k-point. To calculate the
density of states (DOS), a 4 4 4 Monkhorst-Pack [27] k point
mesh, which contains 36 k points was employed. A cubic super cell
with a length of 25 Å was used for all congurations to ensure that
adjacent structures do not interact (the largest linear dimension of
aC
60
molecule was 7.10 Å). We dene encapsulation energy (for
singe atom or dimer trapping inside the cage) and association en-
ergy (for single atom or dimer trapping outside the cage) by the
following equation:
E
Enc/Assoc
¼ E (FP-C
60
) eE(C
60
) en E (FP) (1)
where E (C
60
) is the total energy for the isolated C
60
molecule, E (FP-
C
60
) is the total energy of the gaseous atom or atoms occupying the
centre of the C
60
cage or interacting on the surface of C
60
, E (FP) is
the total energy of an isolated ssion product (the reference state)
and n is the number of ssion product atoms considered in the
process.
The inclusion of van der Waals (vdW) forces are particularly
important for the interactions of the highly polarizable noble gases
and transition metal atoms with C
60
. Here, dispersion has been
included by using the pair-wise force eld as implemented by
Grimme et al. [28] in the VASP package.
3. Results and discussion
3.1. Structural and electronic properties of C
60
fullerene molecule
AC
60
fullerene molecule is spherical with truncated icosahedral
(I
h
) symmetry. It consists of 12 pentagonal rings, 20 hexagonal rings
and 32 faces as shown in Fig. 1a. There are two distinct alternate
CeC and C¼C bonds present in the molecule with the experimental
bond lengths of 1.43 Å and 1.39 Å respectively [29]. In order to
validate the computational parameters used in this study, we car-
ried out an energy minimisation calculation to optimize the
structure of a C
60
molecule. In the relaxed structure, we calculated
the equilibrium bond distances, energy gap between the highest
occupied molecular orbital (HOMO) and the lowest unoccupied
molecular orbital (LUMO) and the net magnetic moment, to
compare with experimental and other theoretical values.
The calculated bond lengths of CeC and C¼C were 1.45 Å and
1.40 Å respectively, which agree well with experimental values [29]
and other theoretical calculations [22]. The calculated HOMO-
LUMO gap of the C
60
molecule is 1.55 eV, in good agreement with
the value of 1.64 eV calculated by the DFT calculations of Goclon
et al. [30]. Our calculation shows that the C
60
molecule is non-
magnetic (see Fig. 1b) as predicted in previous studies [31].
3.2. Initial congurations
We considered six different sites for the absorption of ssion
products (FP). In the rst conguration, the FP occupies the center
of the cage [endohedral C
60
(E)], as shown in Fig. 2 (a). In terms of
surface sites, there are ve possible initial positions as shown in
Fig. 2 (b) e (f). These are: (b) on top of the center hexagonal ring (H),
(c) on top of the center pentagonal ring (P), (d) on top of a C-C bond
between two hexagonal rings (66), (e) on top of a C-C bond be-
tween an hexagonal ring and a pentagonal ring (65) and (f) on top
of a C atom on the C
60
cage (C).
3.3. Formation of single gaseous ssion products interacting with
C
60
First we consider the stability of the ssion products (Xe, Kr, Br, I,
Cs, Rb and Te) as single atoms occupying the centre of the C
60
cage.
In relaxed congurations, the positions of the ssion product atoms
remain very close to the centre of the cage. The encapsulation en-
ergy for the ssion products within the C
60
cage depends on the
relative ionization potentials and electron afnities of the atom and
the C
60
molecule but also the size of the atom or ion relative to the
size of the internal volume of C
60
. Fig. 3 (a) shows the calculated
ionization potentials and the electron afnities of the ssion
product atoms and C
60
together with values reported in the data
book [32]. Though the trend agrees well with the values reported in
the data book, the current DFT calculations underestimate the rst
ionization potentials and overestimate the electron afnities. This is
because of the ionization energies and electron afnities calculated
from DFT orbital energies are usually poorer than those of Koop-
mans theorem, depending on the exchange-correlation approxi-
mation employed [33,34].
Both Rb and Cs have lower ionization potentials and lower
electron afnities than the C
60
molecule. A lower ionization po-
tential enables the (easy) removal of the outermost electron from
Fig. 1. (a) Optimised structure of a C
60
molecule and (b) its density of state spectra. The
vertical dotted line, at ~ 4.10 eV, corresponds to the Fermi energy. (A colour version of
this gure can be viewed online.)
N. Kuganathan et al. / Carbon 132 (2018) 477e485478
Fig. 2. Initial congurations considered for the FP: (a) absorbed in C
60
endohedrally, (b) exohedrally adjacent to a six membered ring, (c) a ve membered ring, (d) a C-C bond
between two hexagons, (e) a C-C bond between an hexagon and a pentagon and (f) a C on the fullerene cage.
Fig. 3. (a) Calculated ionisation potential and electron afnity of volatile ssion product atoms and C
60
together with the values reported in the data book [32], (b) encapsulation
energies, (c) Bader charge on atoms, (d) magnetic moment of the supercell and (e), (f) and (g) the total DOS of Br, Rb and Cs encapsulated in C
60
respectively. (A colour version of this
gure can be viewed online.)
N. Kuganathan et al. / Carbon 132 (2018) 477e485 479
Rb and Cs which is gained by the C
60
molecule as it has a higher
electron afnity. The calculations show that the encapsulation en-
ergies of Rb and Cs are highly exothermic [see Fig. 3b] meaning that
they are more stable in the inner cage of C
60
than as isolated atoms.
Bader analysis also shows that both Rb and Cs gain a charge close
to þ1 by donating their outer s-electron to the C
60
cage. Cs and Rb
are therefore ions, which exhibit smaller radii than their corre-
sponding atoms. The resultant complex can be considered as a
donor-acceptor endohedral complex (M
þ
-C
60
). Wang et al. [7]have
calculated the formation energies for the endohedral encapsulation
of Rb and Cs using Hartree-Fock calculations together with the
Born-Haber cycle. In their calculations, they assumed that both
metals lose one electron to form a þ1 charge and C
60
gains that lost
electron to form 1 charge. The calculated formation energies by
Wang et al. [7] for Rb and Cs are 1.91 eV and 1.64 eV respectively,
in good agreement with our corresponding calculated values
of 1.82 eV and 1.86 eV though the trend is slightly reversed.
Figs. 1 and 3d report that the magnetic moment of C
60
is zero
before and after the incorporation of Rb and Cs, indicating that
encapsulation does not affect the magnetism. Calculations indicate
that C
60
exhibits zero magnetic moment. As Rb
þ
and Cs
þ
exhibit
zero magnetic moment, overall complexes Rb
þ
-C
60
e
and Cs
þ
-C
60
e
are
thus also non-magnetic. The total DOS for the Rb and Cs encapsu-
lated in C
60
are shown in Fig. 3f and (g) respectively. The encap-
sulation of Rb and Cs shift their DOS towards higher energies and
the lowest unoccupied band of C
60
becomes occupied by the s
1
electron transferred to C
60
.
Turning now to the inert gasses, the outer electronic congu-
rations of Xe and Kr are unaffected by encapsulation. This is a
consequence of their higher ionization potentials. This is further
supported by the very low encapsulation energies and the Bader
charge. The encapsulation energies are a balance between attrac-
tive van der Waals forces of the polarisable atoms and Pauli
repulsion of the electron clouds. Since Xe is large the Pauli repul-
sion is greater than exhibited by Kr and so Xe exhibits a slightly
positive (unfavourable) encapsulation energy whereas for Kr the
energy is slightly favourable.
The electron afnities of Br and I are reected in the encapsu-
lation energies and the Bader charges shown in Fig. 3; the encap-
sulation energy of Br is much more favourable than that of I due to
the high electron afnity of Br but also because the smaller Br ts
better inside C
60
. The Bader charges on Br and I are 0.50 and 0.21
respectively. The outer electronic conguration of both Br and I are
s
2
p
5
with one unpaired electron. As C
60
and C
60
þ
are non-magnetic
and both Br and I gain electronic charge from C
60,
the overall
magnetic moment is reduced. The magnitude of the magnetic
moment is reected in the degree of the Bader charge. The total
DOS (see Fig. 3e) shows that incorporation of Br introduces addi-
tional bands in the HOMO of C
60
but the band gap is not much
affected.
The positive encapsulation energy for Te reveals that it is un-
stable inside the cage due to its larger size and its inability to donate
charge to C-C 66 single bonds. A small Bader charge and magnetic
moment show that its outer electronic conguration is not altered.
Next, we consider how ssion products can be trapped (or not)
by the outer surface of C
60
(i.e. the association energy). The
behaviour of ssion products occupying positions on the inner and
outer faces of the C
60
cage are different as the faces are convex and
concave respectively. As explained earlier, we consider ve possible
initial exohedral congurations. These have been relaxed to give
the nal congurations and association energies. The relative as-
sociation energies of each atom, reported in Table 1, are energies
relative to the most stable site (which therefore has zero relative
energy but not zero absolute association energy (see Fig. 4)).
The most stable exohedral position for Xe and Kr is on top of the
hexagonal ring (H). However, differences in energy compared with
the other sites are very small, a consequence of how unreactive are
the noble ssion product gas atoms. The shortest FP-C distance,
formation energies, Bader charge, magnetic moment of the super-
cell and the total DOS for the most stable congurations (identied
in Table 1) are plotted in Fig. 4. While the (absolute) association
energy is very small, it is negative due to the vdW interaction (see
Fig. 4 a). Both a zero Bader charge and magnetic moment validate
this conclusion. This is why the two noble gas atoms are almost
4.00 Å away from the surface of C
60
. Interestingly the exohedral
association energy for Kr is smaller than the encapsulation energy
because Kr only interacts on the surface with six C atoms while
inside C
60
it interacts with the whole molecule. For Xe the situation
is reversed: larger Xe is stable on the surface while its encapsula-
tion energy is positive due to the large Pauli repulsion (compare
values in Figs. 3 and 4).
Br and I show a preference to be on top of the C-C bond, site (66),
and gain a small charge (see Fig. 4c). That their magnetic moments
are closer to 1 and Bader charge small means that their s
2
p
5
conguration with one spin up electron is unaltered. The formation
energy for Br is 0.93 eV, a more favourable value than that for I
(0.56 eV). The association energy calculated for I by Kobayashi
et al. [35]is0.10 eV (without dispersion correction). This deviates
from our calculated value of 0.56 eV where a dispersion correc-
tion is included. The stronger absorption of Br is due to its higher
electron afnity compared to I and the shorter bond distance (C-Br
vs C-I). Compared to Br and I inside C
60
, less charge is transferred.
This can be due to the halide atoms inside the cage being adjacent
to and thus able to gain electron density from all the C atoms of C
60
.
However, inside the cage the bonding process is in opposition to
Pauli repulsion; this is enough that I is more stable outside C
60
.
Conversely exohedral absorption of smaller Br is less favourable
than inside the cage. Furthermore, with Br the Fermi level is shifted
slightly to higher energies and the DOS spectrum of C
60
is now
dispersed with additional peaks. This is due to the chemical inter-
action of Br with the C atoms.
The (66) position is the preferred site for the Te, which loses ~1
electron (see Bader charge in Fig. 4 c) due to its lower electron af-
nity than C
60
. Its formation energy is 1.39 eV showing strong
absorption by the C
60
surface, in contrast to endohedral adsorption.
There is a signicant bonding interaction between Te and carbon in
the (66) position. This geometry of Te associated with (66) carbon
can be described as an icosahedral triangle with the C-Te bond
distance of 2.24 Å and C-C bond distance of 1.51 Å. The observed C-C
bond distance of 1.51 Å is 0.11 Å longer than the C-C distance pre-
dicted for the (66) positions in other parts of the relaxed C
60
conguration. It is further evidence of the signicant interaction
with Te. The Bader charge approximation shows that Te has lost 1.11
|e| and the carbons in the (66) position have gained those electrons
with the ratio of 0.66:0.51. As the carbon in (66) position are now
four coordinated, the
p
-
p
interaction is reduced with the elonga-
tion in the C-C bond. However, Te is unstable inside the cage and
shows no interaction with C
60
as evidenced by the Bader change
and magnetic moment. This is due to the larger size of the Te and its
inability to donate charge to C-C 66 single bonds (in the relaxed
conguration, Te occupies the center of the C
60
and the distance
between the Te and C-C 66 bond is 3.55 Å).
Rb and Cs exhibit strong association energies, essentially to the
same extent as they did for encapsulation. This is because the alkali
atoms have low ionization potentials and electron afnities. Bader
analysis shows that both Rb and Cs form ~þ1 charge donating their
single s electrons to C
60
. The resultant supercells exhibit zero
magnetic moment. Fig. 4g shows there is a substantial Fermi energy
shift to the higher energy levels upon absorption of Cs due to
electron transfer to the C
60
cage. The calculated association energy
N. Kuganathan et al. / Carbon 132 (2018) 477e485480
of Cs derived by Kobayashi et al. [35]is1.85 eV including the
dispersion correction (without the dispersion correction the values
is 1.49 eV) which is in good agreement with our current DFT þ D
value of 1.96 eV, and shows the necessity of including dispersion.
Association energies calculated for of Rb and Cs by Sankar De et al.
[36]are1.59 eV and 1.70 eV respectively slightly deviating from
our corresponding values of 1.72 eV and 1.96 eV due to the
dispersion correction not being included in their calculations.
The encapsulation and association energies reported in Figs. 3
and 4 are for single atoms from single gas phase reference states.
While this is physically realistic for Xe and Kr. It is clearly not for the
other ssion products that will form more complex gas phase
species or react to form solids in their particular waste streams.
Nevertheless, the energies reported thus are important compo-
nents in relevant thermodynamic cycles and we hope prove useful
to other studies. So, for example, it is possible that the reactive
ssion products may form diatomic species that are encapsulated
or associated. The associated energies are also important
components in future thermodynamic cycles. Thus, as a rst step
towards more complex cycles we next report encapsulation and
association energies of homonuclear diatomic molecules. These
will also enable us to determine if encapsulation or association of
atoms or molecules is the more favourable.
3.4. Formation of homonucler dimers
Calculations were performed to investigate the energetics
associated with homonuclear dimers both inside and outside C
60
,
with respect to isolated atoms and with respect to isolated dimer
molecules (i.e. isolated atoms and molecules as reference states but
normalised per atom). Only the reactive ssion products were
considered (i.e. those that form gas phase molecules). We consid-
ered Br, Rb and Cs for the formation of a dimer inside C
60
(see Fig. 5)
whereas Br, I, Te, Rb and Cs dimers were considered as exohedral
species (see Fig. 6).
The relaxed structures of dimers introduced inside the C
60
cage,
Table 1
Initial and nal congurations (see Fig. 2 for denitions) of volatile ssion products absorbed on C
60
exohedrally. The most stable congurations are highlighted in bold.
Initial Conguration Final conguration and relative association energy (eV)
Xe Kr Br I Te Rb Cs
H H, 0.00 H, 0.00 H, þ0.52 H, þ0.30 H, þ1.19 H, 0.00 H, þ0.02
PP,þ0.07 P, þ0.03 P,þ0.58 P, þ0.32 P, þ1.42 P, þ0.07 P, þ0.10
(66) 66, þ0.02 66, þ0.03 66, þ0.23 66, þ0.06 66, 0.00 66, þ0.03 66, 0.00
(65) 65, þ0.05 65, þ0.03 C, 0.00 65, þ0.13 65, þ0.62 65, þ0.18 65, þ0.20
CC,þ0.06 C, þ0.04 C, 0.00 C, 0.00 66, 0.00 C, þ0.20 C, þ0.22
Fig. 4. (a) Shortest distance between a C atom in C
60
and the exohedral FP, (b) association energy, (c) Bader charge on FP, (d) magnetic moment of the supercell and (e), (f) and (g)
the total DOS of the most stable congurations of Br-C
60
, Te-C
60
and Cs-C
60
respectively. (A colour version of this gure can be viewed online.)
N. Kuganathan et al. / Carbon 132 (2018) 477e485 481
their encapsulation energies, and the diameter of the host C
60
molecules upon encapsulation, the dimer bond distances and the
Bader charges on the dimers are shown in Fig. 5. Calculations show
that all three dimers cause signicant distortion of C
60
; the degree
of distortion is reected in the large positive encapsulation energies
and the diameters of the dimers containing C
60
(d) compared to
dimer free C
60
(7.10 Å) (see Fig. 5d). The encapsulation energies for
Br, Rb and Cs are positive meaning that they are not stable with
respect to gas phase atoms or molecules. Endohedral encapsulation
will therefore not be observed.
A dimer conguration for each species was allowed to absorb on
the surface of C
60
. Fig. 6 shows: the relaxed structures of Br
2
,Te
2
and Cs
2
and reports association energies (both from isolated atoms
and isolated dimer reference states), the Bader charges on the di-
mers, shortest FP-C distance and the dimer bond length (D). The
relaxed structures for I
2
and Rb
2
are similar to Br
2
and Cs
2
respectively.
Rb and Cs exhibit slightly smaller association energies per atom
if absorbed as a dimer to C
60
(Fig. 6) compared to single atoms
(Fig. 4). Further, the separations of Rb-Rb (4.81 Å) and Cs-Cs (5.10 Å)
(see Fig. 6d), are longer than in the corresponding gas phase
diatomic molecules (4.30 Å for Rb
2
and 4.83 Å for Cs
2
) consistent
with these alkali metal species being non-bonded. This is because
Rb and Cs interact strongly with C
60
with each atom donating ~1
electron (so that the Bader charge, reported in Fig. 6 for pairs, is
much greater than 1). The association energies per atom for Rb and
Cs pairs are 1.20 eV and 1.48 eV respectively, less favourable
than the values calculated for the corresponding single
atoms, 1.72 eV and 1.95 eV respectively. Thus, while there is a
driving force for a second alkali metal to join an alkali metal
Fig. 5. Relaxed structures of (a) Br
2
(b) Rb
2
(c) Cs
2
encapsulated in C
60
, (d) encapsulation energies from atoms, encapsulation energies of gas phase dimer (per atom), diameters of
relaxed C
60
molecule, D, dimer bond lengths, B, and the Bader charge on dimers. (A colour version of this gure can be viewed online.)
Fig. 6. Relaxed structures of (a) Br
2
(b) Te
2
(c) Cs
2
on the surface of C
60
and (d) as-
sociation energies of FP dimers, shortest FP-C distances, FP-FP distances and the Bader
charges on dimers. (A colour version of this gure can be viewed online.)
N. Kuganathan et al. / Carbon 132 (2018) 477e485482
exohedrally (i.e. 0.68 eV for the second Rb and 1.01 eV for the
second Cs) if there are other unoccupied C
60
molecules available,
the second alkali metal will become associated with an unoccupied
C
60
molecule.
Br, I and Te all form covalently bonded dimer molecules with
bond lengths values when exohedrally absorbed (Fig. 6), that are
very similar to data from experimental (Br
2
2.28 Å, I
2
2.66 Å and Te
2
2.56 Å) [37] and calculated (Br
2
2.31 Å,I
2
2.68 Å and Te
2
2.58 Å) gas
phase species. This indicates these molecules are not bonded to C
60
.
The long FP-C distances are further evidence of this (see Fig. 6).
While there are strong association energies for Br
2
,I
2
and Te
2
per
atom, from the reference states of atoms, for the dimers to asso-
ciate, the association energies per dimer from an existing dimer are
small (i.e. from the dimer reference state). The dimer association is
a result of the induced-induced polarization observed on each of
the atoms of the dimer, though the overall Bader charge is zero,
consistent with a non-bonded interaction. The rationale for this is
that a single atom has considerable electronegativity based on
gaining charge that will occupy a p-orbital, whereas additional
negative charge to a dimer has to occupy a higher energy
s
*
mo-
lecular orbital. Thus, while Br, I and Te demonstrate a preference to
form exohedral dimers rather than forming two single atoms on the
surfaces of two different C
60
molecules, that driving force is a
consequence of the covalent bond between the two ssion product
species. Thus, the association of these dimers is quite different to
that of their atoms, in contrast to the situation for the alkali metals.
Table 2 reports the energetics for taking homonuclear diatomic
gas phase molecules and associating them with C
60
either as atoms
or molecules, assuming an abundance of C
60
molecules. These
values show that Rb
2
and Cs
2
dissociate and associate as single
positive ions. Conversely, Br
2
,I
2
and Te
2
stay as strong covalently
bonded molecules, weakly associating via vdW interactions.
3.5. Heteronuclear dimer association and endohedral/exohedral
pairs
Finally, we have considered interactions from heteronuclear
diatomic molecule reference states. Consider rst CsBr: there are
two different ssion atoms (Cs and Br) interacting with C
60
in four
different ways. The four congurations for Cs and Br are: Fig. 7 (a)
with Br encapsulated by C
60
while Cs is associated exohedrally,
Fig. 7b where Cs is encapsulated by C
60
and Br associated exohe-
drally, Fig. 7c where both Cs and Br are associated as single atoms
exohedrally and Fig. 7 (d) Cs and Br are associated exohedrally as a
pair. This enables an investigation of whether Cs encapsulation
facilitates Br association or vice versa. We have also considered Cs
and I as a pair but only associating with C
60
exohedrally since
neither Cs nor I show any preference for endohedral encapsulation.
The energies to encapsulate and associate CsBr and CsI are re-
ported in Table 3. Total energies for congurations (geometries) as
in Fig. 7d are always more favourable than those as in Fig. 7c. Thus,
exohedral molecule energies discussed will be with respect only to
Fig. 7d. The values with respect to single atoms can be used to rst
assess how much more strongly a Cs, Br or I atom will become
trapped if an opposite heteronuclear species is already present
(note: energies with respect to single gas atom reference states).
For example, Table 3 reports Br is encapsulated to C
60
with an en-
ergy of 0.98 eV but if Cs is already present as an exohedral species
the Br encapsulation energy increases to 2.29 eV. Similarly for the
Cs exohedral association energy raises from 1.96 eV to 3.28 eV if
Br is already present as an endohedral species. Equivalent combi-
nations show similar gains in energy. Table 3 also shows that Bader
charges on Br and Cs increase when both species are present. The
same effect is reported for Cs and I.
Table 3 also reports energies to trap CsBr and CsI molecules. For
CsBr both Br endohedrally and exohedrally are considered with Cs
considered exohedrally. In both cases CsBr is trapped but not
strongly and there is a preference for both species being exohedral.
The energy for Cs encapsulated and Br associated exohedrally is
also reported in Table 3. In this case the energy positive (i.e.
unfavourable) reecting the strong preference for Cs to be an
exohedral species rather than endohedral. The association energy
of CsI (both exohedral species) is similar to that of CsBr and small
consistent with these molecules being trapped by vdW forces. The
net Bader charge on the CsBr and CsI molecules exohedral to C
60
is
almost zero, again consistent with a non-bonded interaction. The
energies are similar to those observed for homonuclear dimers (Br
2
,
I
2
and Te
2
). For CsI, the adsorption energy with respect to the dimer
derived by Kobayashi et al. [35]is0.12 eV without dispersion
and 0.62 eV with dispersion (though we assume only for Cs). The
current calculation shows that the association of CsI is 0.42 eV
though we include dispersion for Cs and I thus importantly be-
tween Cs and I.
Finally, we calculate the energy difference (reaction) between ½
Br
2
(and ½ I
2
) exohedrally trapped and Cs (and Rb) trapped as ions
to form CsBr (and CsI) trapped exohedrally. In both cases the het-
eronuclear molecules are more stable; by 1.24 eV for CsBr
and 0.98 eV CsI.
4. Conclusions
C
60
is a candidate active material for lters used in nuclear
plants where volatile ssion products must be separated from
efuent gases. In this context it has good mechanical and chemical
properties. Using density functional theory together with a
dispersion correction and vdW forces, we have investigated the
capacity of C
60
to trap key volatile ssion products.
Consider rst trapping of atoms from the reference state of
atoms. Xe and Kr are weakly trapped via vdW forces with exohedral
association energies of about 0.05 eV. Kr is endohedrally trapped
with a stronger encapsulation energy of 0.20 eV but Xe, being
much larger, is not trapped endohedrally. Single gaseous Rb and Cs
atoms exhibit endohedral encapsulation energies of 1.82 eV
and 1.86 eV respectively (i.e. strongly exothermic). The outer
surface of the C
60
also exothermically absorbs Rb and Cs as single
gaseous atoms with similar energies of 1.72 eV for Rb
and 1.96 eV for Cs. Thus endohedral and exohedral behaviour is
similar. Calculated trapping energies for Br, I and Te as single atom
Table 2
Exohedral association energies of FP atoms and homonuclear diatomic molecules from molecule reference states.
Most stable position for atoms/dimers associating with C
60
exohedrally
Rb Cs Br I Te
Most stable site H (66) C C (66)
Association energy of an atom (eV/atom) with respect to dimer 1.44 1.70 þ0.34 þ0.56 þ0.38
Rb
2
Cs
2
Br
2
I
2
Te
2
Association energy of dimer (eV/atom) with respect to dimer 1.20 1.48 0.06 0.07 0.13
N. Kuganathan et al. / Carbon 132 (2018) 477e485 483
species from the gas phase showed different behaviours. Br is
exohedrally and endohedrally trapped with energies of 0.93 eV
and 0.98 eV respectively. I exhibits an encapsulation energy
of 0.2 eV but an association energy of 0.6 eV. For Te the differ-
ence is even greater with an endothermic encapsulation energy and
an exothermic association energy of 1.4 eV.
A dimer will not form inside the cage of C
60
for any species
because there is insufcient volume, as evidenced by the distortion
introduced by the dimer (see Fig. 5). However, the outer surface has
the ability to trap Br
2
,I
2
or Te
2
dimers via non-bonding van-der
Waals interactions with association energies per atom of only
around 0.1 eV per atom. Conversely, Rb
2
and Cs
2
are strongly
absorbed but as single atoms, forming positive charged ions as
these ions are highly stable.
Finally, trapping of CsBr and CsI molecules was considered. Both
showed a preference to be endohedral molecules with encapsula-
tion energies of just under 0.2 eV per atom.
Acknowledgements
We thank the EPSRC for funding as part of the PACIFIC pro-
gramme (grant code EP/L018616/1). Computational facilities and
support were provided by High Performance Computing Centre at
Imperial College London. Mr. Conor O. T. Galvin is acknowledged for
helpful discussions.
References
[1] V.M. Erfrmenkov, Radioactive waste management at nuclear plants, IAEA Bull.
4 (1989) 37e42.
[2] M.I. Ojovan, W.E. Lee, An Introduction to Nuclear Waste Immobilisation,
second ed., Elsevier, Oxford, U.K, 2014.
[3] A. Karhu, Methods to Prevent the Source Term of Methyl Lodide during a Core
Melt Accident (NKSe13) Denmark, 1999.
[4] D.F. Sava, M.A. Rodriguez, K.W. Chapman, P.J. Chupas, J.A. Greathouse,
P.S. Crozier, T.M. Nenoff, Capture of volatile iodine, a gaseous ssion product,
by zeolitic imidazolate Framework-8, J. Am. Chem. Soc. 133 (2011)
12398e12401.
[5] M.S. Dresselhaus, G. Dresselhaus, P.C. Eklund, Science of Fullerenes and Car-
bon Nanotubes, Academic Press, San Diego, 1996.
[6] H.W. Kroto, A.W. Allaf, S.P. Balm, C
60
: buckminsterfullerene, Chem. Rev. 91
(1991) 1213e1235.
[7] Y. Wang, D. Tom
anek, R.S. Ruoff, Stability of M@C60 endohedral complexes,
Chem. Phys. Lett. 208 (1993) 79e85.
[8] E. Broclawik, A. Eilmes, Density functional study of endohedral complexes M@
C60 (M¼Li, Na, K, Be, Mg, Ca, La, B, Al): electronic properties, ionization po-
tentials, and electron afnities, J. Chem. Phys. 108 (1998) 3498 e3503.
[9] K. Kikuchi, K. Kobayashi, K. Sueki, S. Suzuki, H. Nakahara, Y. Achiba, K. Tomura,
M. Katada, Encapsulation of radioactive 159gd and 161tb atoms in fullerene
cages, J. Am. Chem. Soc. 116 (1994) 9775e9776.
[10] S.K. Saha, D.P. Chowdhury, S.K. Das, R. Guin, Encapsulation of radioactive
isotopes into C60 fullerene cage by recoil implantation technique, Nucl. Ins-
trum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 243 (2006)
277e281.
[11] A.K. Srivastava, S.K. Pandey, N. Misra, Encapsulation of lawrencium into C60
fullerene: Lr@C60 versus Li@C60, Mater. Chem. Phys. 177 (2016) 437e441.
[12] J. Lu, Y. Zhou, X. Zhang, X. Zhao, Structural and electronic properties of
endohedral phosphorus fullerene P@C60: an off-centre Displacement of P
inside the cage, Mol. Phys. 99 (2001) 1199e1202.
[13] S.C. Cho, T. Kaneko, H. Ishida, R. Hatakeyama, Nitrogen-atom endohedral
fullerene synthesis with high efciency by controlling plasma-ion irradiation
energy and C60 internal energy, J. Appl. Phys. 117 (2015), 123301.
[14] T. Braun, H. Rausch, Endohedral incorporation of argon atoms into C60 by
neutron irradiation, Chem. Phys. Lett. 237 (1995) 1995.
[15] Z.Y. Wang, K.H. Su, H.Q. Fan, Y.L. Li, Z.Y. Wen, Mechanical, Electronic prop-
erties of endofullerene Ne@C60 studied via structure distortions, Mol. Phys.
106 (2008) 703e716.
[16] P.F. Weck, E. Kim, K.R. Czerwinski, D. Tom
anek, Structural and magnetic
Fig. 7. Relaxed congurations of (a) endohedral Br with exohedral Cs, (b) endohedral Cs with exohedral Br and (c) exohedral Cs with exohedral Br as single atoms and (d) exohedral
Cs with exohedral Br as a pair. (A colour version of this gure can be viewed online.)
Table 3
Encapsulation or association energies and Bader charge calculated for the ssion atoms interacting C
60.
Reaction Encapsulation/Association energy (eV/atom) with respect to atoms Bader charge |e|
Br or I Cs
Br þ C
60
/ Br
endo
C
60
0.98 0.50 e
Br þ Cs
exo
C
60
/ Br
endo
Cs
exo
C
60
2.29 0.71 þ0.94
Cs þ C
60
/ Cs
exo
C
60
1.96 e þ0.91
Cs þ Br
endo
C
60
/ Cs
exo
Br
endo
C
60
3.28 0.71 þ0.94
Br þ C
60
/ Br
exo
C
60
0.93 0.17 e
Br þ Cs
exo
C
60
/ Br
exo
Cs
exo
C
60
1.31 0.37 þ0.94
Br þ Cs
endo
C
60
/ Br
exo
Cs
endo
C
60
1.59 0.26 þ0.91
Cs þ C
60
/ Cs
endo
C
60
1.85 e þ0.91
Cs þ Br
exo
C
60
/ Cs
endo
Br
exo
C
60
2.50 0.26 þ0.91
Cs þ Br
exo
C
60
/ Cs
exo
Br
exo
C
60
2.34 0.37 þ0.94
I þ C
60
/ I
exo
C
60
0.56 0.09 e
I þ Cs
exo
C
60
/ I
exo
Cs
exo
C
60
1.03 0.37 þ0.94
Cs þ I
exo
C
60
/ Cs
exo
I
exo
C
60
2.43 0.37 þ0.94
with respect to dimer (eV/atom)
CsBr þ C
60
/ Cs
exo
Br
endo
C
60
0.07 0.71 þ0.94
CsBr þ C
60
/ Cs
endo
Br
exo
C
60
þ0.34 0.26 þ0.91
CsBr þ C
60
/ (Cs
exo
Br
exo
)C
60
0.17 0.77 þ0.86
CsI þ C
60
/ (Cs
exo
I
exo
)C
60
0.21 0.71 þ0.85
N. Kuganathan et al. / Carbon 132 (2018) 477e485484
properties of Tc
n
@C
60
endohedral metalofullerenes: rst-principles pre-
dictions, Phys. Rev. B 81 (2010), 125448.
[17] H. Minezaki, K. Oshima, T. Uchida, T. Mizuki, R. Racz, M. Muramatsu, T. Asaji,
A. Kitagawa, Y. Kato, S. Biri, Y. Yoshida, Synthesis of FeeC60 complex by ion
irradiation, Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater.
Atoms 310 (2013) 18e22.
[18] J.R. Heath, S.C. O Brien, Q. Zhang, Y. Liu, R.F. Curl, F.K. Tittel, R.E. Smalley, J. Am.
Chem. Soc. 107 (1985) 7779e7780.
[19] N. Kuganathan, J.C. Green, H.-J. Himmel, N. J. Chem. 30 (2006) 1253e1262.
[20] Q. Sun, Q. Wang, P. Jena, Y. Kawazoe, Clustering of Ti on a C60 surface and its
effect on hydrogen storage, J. Am. Chem. Soc. 127 (2005) 14582e14583.
[21] B.
Ozdamar, M. Boero, C. Massobrio, D. Felder-Flesch, S.L. Roux, Systems as
tunable- gap building blocks for nanoarchitecture and nanocatalysis, J. Chem.
Phys. 143 (2015) 114308.
[22] A.M.E. Mahdy, Dft study of hydrogen storage in Pd-Decorated C60 fullerene,
Mol. Phys. 113 (2015) 3531e3544.
[23] G. Kresse, J. Furthmüller, Efcient iterative schemes for ab initio total-energy
calculations using a plane-wave basis set, Phys. Rev. B 54 (1996)
11169e11186.
[24] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector
augmented-wave method, Phys. Rev. B 59 (1999) 1758e1775.
[25] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
simple, Phys. Rev. Lett. 77 (1996) 3865e3868.
[26] W.H. Press, B.P. Flannery, S.A. Teukolsky, W.T. Vetterling, Numerical Recipes.
The Art of Scientic Computing, Cambridge Univ. Press, Cambridge, 1986, 818
pp.
[27] H.J. Monkhorst, J.D. Pack, Special points for brillouin-zone integrations, Phys.
Rev. B 13 (1976) 5188e5192.
[28] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab initio
parametrization of density functional dispersion correction (Dft-D) for the 94
elements H-Pu, J. Chem. Phys. 132 (2010), 154104.
[29] J.M. Hawkins, Accounts Chem. Res. 25 (1992) 150e156.
[30] J. Goclon, K. Winkler, J.T. Margraf, Theoretical investigation of interactions
between palladium and fullerene in polymer, RSC Adv. 7 (2017) 2202e2210.
[31] N.M. Umran, R. Kumar, Theoretical investigation of endohedral complexes of
Si and Ge with C
60
molecule, Phys. B Condens. Matter 437 (2014) 47e52.
[32] D.R. Lide, CRC Handbook of Chemistry and Physics, 86th ed., CRC, Boca Raton,
FL, 2005.
[33] P. Politzer, F. Abu-Awwad, A comparative analysis of Hartree-Fock and Kohn-
Sham orbital energies, Theo. Chem. Acc. 99 (2) (1998) 83e87.
[34] S. Hamel, P. Duffy, M.E. Casida, D.R. Salahub, KohneSham orbitals and orbital
energies: ctitious constructs but good approximations all the same,
J. Electron. Spectrosc. Relat. Phenom. 123 (2002) 345.
[35] T. Kobayashi, K. Yokoyama, Theoretical study of the adsorption of Cs, Csþ,I,
I, and CsI on C60 fullerene, J. Nucl. Sci. Technol. 53 (2016) 1489e1493.
[36] D. Sankar De, J.A. Flores-Livas, S. Saha, L. Genovese, S. Goedecker, Stable
structures of exohedrally decorated C60-fullerenes, Carbon 129 (2018)
847e853.
[37] K.P. Huber, G. Herzberg, Molecular Spectra and Molecular Structured IV.
Constants of Diatomic Molecules, Van Nostrand Reinhold, New York, 1979.
N. Kuganathan et al. / Carbon 132 (2018) 477e485 485