You are on page 1of 21

Science Progress (2018), 101(2), 101 – 120

Paper 1800262 https://doi.org/10.3184/003685018X15173976099750

Oganesson: a most unusual ‘inert gas’


RODERICK M. MACRAE and TERENCE J. KEMP

Dr Roderick Macrae is Lechleiter Endowed Professor of Chemistry


at Marian University in Indianapolis. He received his PhD in
Chemistry from the University of Glasgow in 1990. During
his subsequent travels, he spent six years in Japan, at KEK (the
High-Energy Accelerator Research Organization) and RIKEN (the
Institute for Physical and Chemical Research). His research work
has mainly been divided between experimental studies using the
muon spin rotation method on materials ranging from fullerenes to
semiconductors, and the application of computational chemistry to a
variety of systems. Additionally, he is co-founder of the Institute for Green and Sustainable
Science at Marian University, a research, education, and outreach centre engaging in the
study of sustainability issues. He may be contacted at the College of Arts and Sciences,
Marian University, Indianapolis, IN 46222, USA. E-mail: rmacrae@marian.edu

Terence Kemp is an Emeritus Professor of Chemistry at the


University of Warwick. At Jesus College Oxford he gained
the degrees of MA, DPhil and DSc. After postdoctoral work
at Leeds University on radiation chemistry, he joined the then
new University of Warwick for his remaining academic career.
He has published 250 papers covering radiation chemistry,
free radical solution kinetics, laser flash photolysis, ESR of
(i) reactions of solvated electrons in liquid ammonia and (ii) solution and frozen redox
photosystems, particularly of uranyl ion, metal complexes and diazonium salts, the latter
yielding the triplet state aryl cation. His later interests covered semiconductor photocatalysis
and polymer systems including linear polysulfides and nitrated polyethers. He is currently
Editor of Progress in Reaction Kinetics and Mechanism and an Editor of Science Progress.
E-mail: t.j.kemp@sciencereviews.co.uk

1. Introduction
Oganesson, with the symbol Og, is the artificially prepared (i.e. ‘man-made’) element
with atomic number 118 (see refs 1 and 2), previously known as eka-radon or Uuo
(‘ununoctium’), and recently ‘blessed’ with an official name by IUPAC3. Atoms of
the element were first produced using the U400 cyclotron at the Joint Institute for
Nuclear Research (JINR) in Dubna, Russia, via a heavy-ion fusion reaction utilising
collision of a 48
20 Ca beam with a 98Cf target:
249

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 101


249
98Cf + 48
20 Ca →
294
Og + 3𝑛𝑛 .

The choice of projectile and target nuclides was made on the basis of maximisation
of neutron excess in the fusion products. Initial identification of the new element
was via its characteristic decay pattern, successive alpha decays into the even–
even nuclides 116
290
114 Fl, and 112 Cn, the last of which decays 100% through
Lv, 286 282

spontaneous fission (SF).

2. Background on production method


Within the nuclear shell model, nuclei are sometimes referred to as ‘singly magic’
or ‘doubly magic’ because they possess numbers of protons and/or neutrons
corresponding to filled shells. Doubly magic nuclei are spherical, and correspond to
‘islands of stability’ on the table of all possible isotopes (created by plotting number of
protons on one axis and number of neutrons on the other). One productive approach
to the synthesis of long-lived superheavy nuclei is the bombardment of a singly
or doubly magic nucleus such as 208Pb or 209Bi with stable neutron-rich projectiles
such as 70Zn (see refs 4 and 5); however, this method leads to ‘diminishing returns’
beyond around Z = 113 (see ref. 6). A new approach began to be pursued around the
turn of the 21st century at the Flerov Laboratory of Nuclear Reactions (FLNR) of
the JINR at Dubna, involving the bombardment of actinide targets, some of which
were synthesised and purified at the Radiochemical Engineering Development
7 48
Center at Oak Ridge National Laboratory, USA, with 48 20 Ca . 20 Ca is a ‘nearly
stable’ (i.e. long-lived) ‘doubly magic’ isotope of Ca, which decays via double-beta
decay with a half-life of 4.4 × 1019 years and forms 0.187% of naturally-occurring
Ca. This is the approach that led to the discovery of oganesson, as well as multiple
new isotopes of elements 112–117. Previous reports of the production of 293118 in
the Berkeley separator8 failed to be confirmed by experiments at the Gesellschaft für
Schwerionenforschung (GSI) in Darmstadt at the same energies9,10.
Oganesson has a closed-shell electronic structure [Rn]5f146d107s27p66 and should
therefore be regarded formally as the bottom-most member of the periodic group
of inert gases. What makes Og chemistry a demanding subject to investigate is the
very short lifetime (0.89 ms) of the only currently confirmed isotope owing to alpha-
decay1,2:
294 290 4
118 Og → 116 Lv + 2He .

Other superheavy transactinides such as Cn and Fl possess relatively long-lived


isotopes which permit the application of a variety of ‘one atom at a time’ methods that
have been developed for the study of such exotic species – see for example the 1999
article by Darleane Hoffman and Diana Lee in the Journal of Chemical Education
for an early overview11. (It is also evident from this article that ‘unusual’ behaviour
of transactinides is the rule rather than the exception.) These methods include gas
phase thermochromatography, a technique in which a negative temperature gradient
is used to concentrate analytes in narrow spatial zones depending on volatility; the

102 Roderick M. Macrae and Terence J. Kemp


method has a pedigree dating back to the Curies12, but was established as an effective
technique in radiochemistry by Merinis and Bouissieres in 196113.
Because of its instability and the fact that only a few atoms of Og have ever been
produced, defining its chemistry has become the task of theoreticians. This article
is drawn extensively from the recent article in Physical Review Letters by Jerabek
et al.14, recast and simplified for a more general scientific audience. The article is
quite unusual, in that it is a paper in which electronic structure and nuclear structure
of an atom are discussed together, and in similar terminology, that of particle
localisation functions. Additionally, a substantial amount of historical information
and background theory is supplied.

3. Electronic properties
The Schrödinger theory of atoms and molecules is a non-relativistic theory that,
moreover, does not incorporate electron spin. Despite these shortcomings, it
gives energy solutions for the hydrogen atom that accurately reproduce the gross
features of its spectrum (in exact agreement with Bohr theory). Fine structure
observable in the hydrogen atom spectrum is a result of electron spin, relativistic
effects, and smaller quantum electrodynamic (QED) corrections such as the
Lamb shift. The Dirac theory of the electron is both relativistic and naturally
incorporates spin, but leads to a four-component (i.e. 4 × 1 matrix) wavefunction
that is more complicated to implement in computational methods. Any treatment
of many-electron atoms is of necessity approximate. (The intrinsic ‘spin’ angular
momentum of the electron already appears naturally in the solutions to the non-
relativistic Levy–Leblond equation as a consequence of coupling between the
electron charge and the external electromagnetic field15.)
The inclusion of relativistic effects in quantum-mechanical calculations on
atoms and molecules becomes more important with increasing atomic number
Z. A qualitative understanding of relativistic effects on electronic structure
can be obtained through the Bohr model of the H atom16,17. In Bohr theory, the
electron’s (non-relativistic) velocity is directly proportional to nuclear charge
according to

(In atomic units, v/c = Z/137 for an electron in the n = 1 orbit, so even for the H
atom the electron velocity is close to 1.0% of c. The speed of light in atomic units is
137.035 999 138 a.u.) The corresponding relativistic mass increase of the electron is
given by
,

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 103


which leads to a shrinkage of the orbital radius according to

and an increased Coulomb binding energy of the electron

where n, h, and e mean respectively the principal quantum number, Planck’s constant,
and the charge of the electron.
In the Dirac theory of one-electron atoms, the ‘orbitals’ of the Schrödinger
theory (eigenfunctions of n, l, and ml) are replaced by ‘spinors’, four-component
vector-valued wavefunctions which are eigenfunctions of n, j, and mj (where
j = l + s). (These are the large and small components of α and β spin.) These have
rather different properties from the more familiar non-relativistic orbitals: for
example, the density distributions associated with them are nodeless, because the
large and small components have nodes in different places18,19. Additionally, because
the densities depend on j and not on l, the s1/2 and p1/2 spinors are both spherical.
The analogues to the effects seen in the Bohr model now manifest as shrinkage and
increased binding of spinors with significant density near the nucleus (s1/2 and p1/2)
and smaller effects on others (p3/2, d3/2, etc.). A related consequence is that unlike
the non-relativistic p-orbitals, the p1/2 and p3/2 spinors are unequal in energy (with the
p3/2 lying above the p1/2), while the s1/2 and the p1/2 are degenerate. (For the H atom
itself, the separation between the 2P1/2 and 2P3/2 states is 4.5 × 10–5 eV, leading to a fine
structure splitting of about 0.016 nm in the 1S-2P transition.)
In addition to the relativistic mechanical effects on inertial mass and kinetic
energy, there are several quantum relativistic effects that appear in the solutions of
the Dirac equation: these can be categorised in terms of their dependence on the fine-
structure constant α = 1/c (see ref. 15).The lowest-order effects (depending on α2)
are the Darwin effect (coupling between the electron’s spin and the inhomogenous
nuclear electric field) and the spin-orbit coupling (coupling between the electron’s
spin and a magnetic field or moving electric field; the primary contribution to this
effect is from the electron’s own angular momentum around the nucleus, although
two-electron contributions to SOC also exist). Other, smaller effects arise through
QED (self-energy and vacuum polarisation); while in some cases such as the
hydrogen Lamb shift these lead to clearly observable effects, more generally these
interactions lie at the frontier of comparison between theory and experiment20.
In many-electron atoms relativistic effects on valence orbitals can be classified as
direct or indirect15,21. The direct effect is the orbital contraction and energy lowering
described above, which is also seen in one-electron atoms and has effects on orbitals
in the order s > p > d; this effect can sometimes be greater in many-electron atoms
than in hydrogenic atoms. The indirect effect is primarily a destabilisation of outer
orbitals due to more effective screening of the nuclear charge as a result of inner
orbital contraction. While relativistic effects are small in the lighter elements (on

104 Roderick M. Macrae and Terence J. Kemp


the order of 1% of the total energy in Cu), they become much more significant in
high Z elements (exceeding 10% in the actinides). Relativistic contributions are
implicated in phenomena such as the inert pair effect (relative difficulty of removal
of ns2 valence electrons in heavier metals), the yellow colour and chemical inertness
of gold (see e.g. ref. 22) and quite possibly the fact that mercury is a liquid at room
temperature17,23. Indeed, according to recent work, 1.7 V of the 2.1 V of the lead-acid
cell originates in relativistic effects, indicating that ‘without relativity, cars would not
start’20.
Relativity also exerts a significant effect on the chemical bonding properties of
heavy atoms. For example, while hybridisation of s and p orbitals in an adjacent
pair of atoms of a light element such as O leads to the familiar σ and π bonding
combinations, where spin-orbit coupling is strong two p1/2 spinors on adjacent atoms
1 2 2 1
combine to yield 𝜎𝜎 + 𝜋𝜋 * and 𝜋𝜋 + 𝜎𝜎 * combinations, which have weakened
3 3 3 3
bonding and antibonding power, while for p3/2 spinors the relevant bonding
2 1
combinations are π and 3 𝜋𝜋 + 3 𝜎𝜎 * (see ref. 15). In order to obtain full σ-bonding,
1 2
the combination p1/2 + p3/2 must be taken24. The net result of these effects is that
3 3
bonds are typically weakened and ‘softened’ (smaller vibrational force constants) by
the influence of spin-orbit coupling. An apparently converse effect, however, is the
‘transactinide break’20,25, in which covalent radii of the 6d elements are shorter than
those of the 5d elements in groups 9 through 12.
The best computational approaches to the electronic structures of superheavy
elements are those which treat both relativistic effects and electron correlation at
a high level18,25,26. These are typically based on the Dirac–Coulomb–Breit (DCB)
Hamiltonian
1

where the one-electron Dirac operator is


,

with α and β the Dirac matrices and Vn (i) the nuclear attraction operator. The two-
electron term in hDCB consists of the Coulomb repulsion of the electrons, together
with the frequency-independent Breit operator
1
,
2

which is the lowest-order relativistic correction to the two-electron term in the


Coulomb gauge. The nuclear potential term can incorporate finite nuclear size
effects. All operators in these expressions are 4 × 4 matrix spinor operators, and
the corresponding wavefunction is a four-component wavefunction. The DCB
approximation contains all terms up to second order in the fine structure constant.

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 105


Correlation effects can then be taken into account by any of the usual approaches
(configuration interaction [CI] or many-body perturbation theory [MBPT]), with
the most powerful and accurate method currently available being the Fock-Space
Coupled Cluster (FSCC) method26,27. This is an infinite-order size-extensive
covariant many-body method suitable for systems with a variable number of
particles (and thus is a good candidate for extension to include QED terms). First
an initial single-determinant reference state is correlated, then electrons are added
one at a time, with the whole system being recorrelated at each stage. Solution of the
FSCC equation leads to an effective Hamiltonian which is then diagonalised to yield
transition energies.
The Fock-space can be considered as partitioned into ‘sectors’ designated (m,n),
where m and n refer to the number of electrons removed from occupied orbitals
(valence holes) or added to virtual orbitals (valence particles) with respect to the
reference determinant26.
Less rigorous methods, often used for molecular calculations, include the use of
relativistic effective core potentials (ECPs) to replace inner electrons not involved
in bond formation, and the use of two-component (rather than four-component)
wavefunctions in the Dirac–Hartree–Fock scheme.
Another approach is the relativistic extension of density functional theory (DFT)
methods. These are based on the 1964 theorems of Hohenberg and Kohn which
showed (1) that the total energy of any system of electrons subject to the influence
of an external potential can be described as a functional of the density only and (2)
that because the exact ground state density is the one that minimises the total energy,
the ground state energy can be obtained variationally28. The method is in principle
exact, but is limited by the fact that the true form of the exchange-correlation (XC)
functional is unknown; a variety of approximate XC functionals have been created
over the intervening decades for different applications. In general, these functionals
can be written in relativistic or non-relativistic form29,30. In the relativistic case,
electron spin should in principle be treated by inclusion of the appropriate terms from
the four-current density in the functionals themselves, but the best currently-available
methods make use of an alternative approach in which the spin is included in the
exchange and correlation functionals via the magnetisation density29,31. This ‘non-
collinear spin-polarised’ method permits each electron to be assigned a wavefunction
with quantum numbers j and mj.
A positive electron affinity was predicted over 20 years ago for ‘element
118’, as it was known then, by Eliav and Pyykkö32. In chemistry, the electron
affinity (EA) is conventionally defined as –DE for the process in which an
atom in the gas phase attaches an electron to form an anion. This process is
energetically favourable for most elements, but unfavourable for the noble gases
as well as elements such as beryllium where anion formation involves occupancy
of a subshell of significantly higher energy than those occupied in the neutral
atom. In elements where the process is energetically unfavourable, it is typical
not to assign an electron affinity, although sometimes a negative EA estimated

106 Roderick M. Macrae and Terence J. Kemp


from calculations is seen in tables. In the case of oganesson, Eliav and Pyykkö
computed a value of 0.056(10) eV (about 5.4 kJ mol–1) using a relativistic Fock-
space coupled-cluster method based on the Dirac–Coulomb–Breit Hamiltonian,
with a large (34s26p20d14f9g6h4i) basis set of Gaussian type orbitals, and with
the outermost 40 electrons treated in a correlated fashion32. The Fock space in this
case took into account the (0,0) sector (the neutral atom) and the (0,1) sector (the
anion). The calculation included a nuclear potential Vnuc that took into account the
finite size of the nucleus by treating it as a sphere of uniform charge density. The
value obtained is about the same as the EA of Sr, and larger than that of Ca, the
smallest positive EA in the periodic table33.
The calculations showed the effect of nuclear mass to be negligible – about
35 J mol–1 over an isotopic mass number range of 283 to 302 (302Og is the closed-
shell isotope), bracketing the mass of the experimentally-known isotope. This
is, for example, 200 × smaller than the finite nuclear size effect. The computed
positive EA requires the inclusion of both relativity and electron correlation in the
calculation, with non-relativistic CC calculations yielding no electron affinity, and
the uncorrelated Dirac–Hartree–Fock one-electron Hamiltonian leading to an 8s
orbital with a positive energy. The primary reason for the positive EA of Og is the
relativistic stabilisation of this vacant 8s orbital.
In a subsequent paper, Eliav and Pyykkö together with three other authors
considered QED corrections to the electron binding energy34. As QED effects
are typically smaller than and in the opposite direction to relativistic effects,
relativistic DCB calculations without QED correction have been referred to as
‘101% accurate’20,35. In this particular case, the QED effect (including self-energy
and vacuum polarisation corrections) leads to a reduction of the binding energy of
around 9%, yielding an optimal final EA value of 0.064(2) eV (about 6.2 kJ mol–1).
These calculations used an improved basis set compared to the previous paper and
correlated all 119 electrons. Due to the one-electron character of parts of the QED
component of the approach, ‘tricks’ including placement of partial charges on the
external electrons compensated by a perturbation term in the energy expression
had to be used in order to ensure that the 8s orbital was bound in the reference
Dirac–Fock–Breit wavefunction. (Nonetheless, the final answer is to be considered
reliable26,34.)
Other exceptionally large relativistic effects seen in the electronic structure of Og
include the huge ‘spin-orbit’ splitting of the 7p shell, amounting to 10.125 eV at the
FSCC level, as computed from the ionisation potentials of the 7p1/2 and 7p3/2 spinors14.
(Such splitting is conventionally referred to as ‘spin-orbit splitting’ by analogy with
H, but in a many-electron atom the energy difference between valence p1/2 and p3/2
subshells is largely Coulombic, due to the different degrees of screening of the two
spinors from the nuclear charge.) Comparison of the rightmost panel of Figure 114
with the other two panels shows both the stabilisation of the virtual (n + 1)s orbital
and the large increase in valence p1/2–p3/2 splitting in going down group 18 from Xe
to Rn to Og. The relativistic contribution to the total ground state electron binding

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 107


Figure 1 Orbital energies of Xe, Rn, and Og in their ground states calculated at the non-
relativistic (NR) and scalar relativistic (SR) Hartree–Fock approximations and in the fully
relativistic Dirac–Hartree–Fock (R) approximations. Energies of unoccupied orbitals are
taken from calculations on the first excited state. Reprinted figure with permission from
Jerabek et al.14. Copyright 2018 by the American Physical Society.

energy of Og is estimated to be 227 keV (over 8,000 hartree, or 22 GJ mol–1), while


for Pb the relativistic contribution is ‘only’ 40 keV.
The earliest computational studies of the transactinides date back to the work
of Fricke, Waber, Greiner, and others36, who predicted electron configurations
and other properties. This was at a time – the late 1960s and early 1970s – when
the CRC Handbook periodic table of the elements listed elements by name up to
103 (Lawrencium, the last of the actinides), and elements 104–106 (which are 6d
transition metals) had been ‘pencilled in’ but not yet named.
In 1975, Kenneth Pitzer asked the question ‘Are elements 112, 114, and 118
relatively inert gases?’, and pursued this inquiry through a study of the results of
then-current relativistic electronic structure calculations24,37. This was a somewhat
heterodox suggestion, as noted by Clinton S. Nash38, as although element 118
(Og) is formally a noble gas, elements 112 and 114 (copernicium and flerovium as
they are known today) are respectively the last element in the 6d transition series
block (falling directly below Hg) and the second element in the 7p block (falling
directly below Pb). That they should be inert gases in their normal states would only
be made possible by extreme relativistic behaviour in their valence shell electrons.
Analogously with Pt and Au in the row above, in Ds and Rg both the 7s and 6d
shells are open, and only close with Cn. In Fl and Og, the 7p shell is being filled. In
Fl, the 7p1/2 spinor subshell is complete, while in Og both the 7p1/2 and 7p3/2 spinor
components have been filled.
Based on this ‘closed spinor shell’ picture of elements 112, 114, and 118, Pitzer
compared the predicted bonding energy of the element with the promotion energy of
the atom to its valence state; if the former exceeded the latter, he concluded that the

108 Roderick M. Macrae and Terence J. Kemp


element in its natural state should be a gas or volatile liquid, bound only by London
forces or possibly very weak metallic interactions.
Nash’s 2005 calculations used the MOLFDIR package39, with RECPs replacing
the first 92 electrons, and the remainder treated in a correlated fashion with a ‘core
space’ consisting of the (n – 1)s, p, and d orbitals, and a ‘valence space’ consisting of
the ns and np: the methods used were Hartree–Fock, MP2, CCSD, and CCSD(T). He
computed ionisation energies, proton affinities, and dimerisation energies for these
three elements, and concluded that Fl would be an ‘anomalously inert, compact, and
unpolarisable’ element compared to lead, while Og would be similarly anomalously
polarisable and diffuse compared to radon.
Later fully relativistic DFT calculations by Pershina et al.40,41 cast some doubt
on the proposed inertness of Cn and Fl; both elements were found to exhibit some
reactivity toward metal surfaces, with the more reactive Fl favouring bridge positions
in adsorption on Au clusters, while Cn was found to favour adsorption in fcc hollow
sites on the Au(111) surface. This ‘metallic’ (Hg-like) behaviour is a relativistic effect;
non-relativistic calculations show Fl to exhibit noble gas-like behaviour more akin to
Rn42. Recent experimental studies using ‘one atom at a time’ thermochromatography
analysis confirm that, while Cn and Fl are more inert than their homologues Hg and
Pb, both elements can bind to Au surfaces and are best described as volatile noble
metals25,43,44. Yakushev et al. obtain an adsorption enthalpy of at least (–)48 kJ mol–1
for Fl, indicating binding weaker than the chemisorption of Hg (98 kJ mol–1) but
stronger than the physisorption of Rn (19.2 kJ mol–1) (see ref. 44).
The static dipole polarisability of Og has been computed several times in
recent years, using a variety of methods14,38,41. Nash, in a study already referenced
above, used a slightly unorthodox variation of the finite field approach with a
RECP description of the 92 inner electrons and a 6sd6p2pf1g description of the
outermost 16 electrons to obtain a value of 52.4 a.u. for Og (described by the author
as ‘enormous’) at the CCSD(T) level, almost twice the value for Rn (28.6 a.u.) and
slightly larger than the value for Pb (see ref. 38). The field in this study was that of
a unit negative test charge whose distance from the subject atom was systematically
varied between 2.6 and 12.6 Bohr, with α taken as the second derivative of the
resulting curve (energy vs field strength) in the zero field limit. Nash notes that by
extrapolation from the empirical linear relationship between atomic polarisability
and the known boiling points of the noble gases He–Xe (Figure 2), a value between
178 K and 221 K is obtained for Rn (for which the accepted value is 211 K) and
between 320 K and 380 K for Og. Remarkably, this would imply that noncovalent
interactions (London forces) in Og are sufficiently strong to make it a liquid at room
temperature!
The polarisability, α has units of volume, and for a uniformly charged sphere
of radius R is given by α = R3 = 3V/4π. Dmitrieva and Plindov45 have shown that
for atoms , where r0 is the atomic radius and g(Z) is a slowly-varying
function of the nuclear charge. However, atomic radius is not a physical observable,
and tests of this relationship using, for example, van der Waals radii obtained from

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 109


Figure 2 Boiling points of the noble gases He–Xe plotted against static polarisabilities. (Data
from P. Schwerdtfeger, CTCP table of experimental and calculated dipole polarisabilities
for the electronic ground states of the neutral elements, http://ctcp.massey.ac.nz/dipole-
polarizabilities.) A value of 52 a.u. for Og leads to a predicted boiling point of 320 K.
crystal data46 show that better linear fits are obtained with the use of a substantial
offset47. Using only noble gas data for He through Xe, such a linear fit yields
in atomic units. Extrapolation of this graph using the
value of α obtained by Nash results in a van der Waals radius of 4.91 bohr (2.60 Å)
for Og.
Subsequent all-electron Dirac–Coulomb CCSD(T) calculations by Pershina et
al. using a more conventional finite field approach yielded values of 46.33 a.u. for
41

Og and 35.04 a.u. for Rn. These calculations used the DIRAC04 program48, with a
26s24p18d13f5g2h basis set. Most recently, Jerabek et al.14 obtained a value of this
property for Og of 57.98 a.u. using ‘state-of-the-art’ 4c Dirac–Coulomb CCSD(T)
calculations incorporating the Gaunt term in the relativistic two-electron operator.
The basis set was the uncontracted quadruple-zeta basis known as DYALL.ACV4Z49,
and 50 electrons were correlated. Computations were carried out using a more
recent version of the DIRAC program, DIRAC1550. A very significant part of the
polarisability appears to require a full relativistic treatment to be captured accurately,
with non-relativistic and scalar relativistic approaches (in the latter of which the
spin-orbit operator is omitted but the Darwin and mass-velocity terms are included)
giving considerably smaller values of 45.30 a.u. and 43.78 a.u. respectively.
The exploration of noble gas chemistry began with Neil Bartlett’s discovery
in 1962 that PtF6 is a strong enough oxidising agent to oxidise Xe, leading to the
formation of the orange-yellow ionic solid xenon hexafluoroplatinate, XePtF6 (see
ref. 51). (The formula of the substance is more accurately written [XeF]+[PtF5]–.)
A rash of research over the subsequent decade led to the discovery of a plethora of
xenon compounds, mostly with oxygen and fluorine, as well as a small number of

110 Roderick M. Macrae and Terence J. Kemp


krypton compounds52. The most iconic of these compounds is perhaps XeF4, first
synthesised in 1962 by Claassen, Selig, and Malm at Argonne National Laboratory53.
Xenon tetrafluoride is a somewhat volatile colourless molecular solid, with a melting
point near 387 K and a vapour pressure of about 340 Pa at room temperature54,55.
Remarkably, its structure accords with the predictions of VSEPR theory and is
square planar, as confirmed by Claassen et al. from an analysis of the infrared and
Raman spectra56. (The absence of any shared transition between the two spectra
indicates that the molecule belongs to the centrosymmetric point group D4h.) Xenon
difluoride has similar properties to the tetrafluoride; it is colourless and molecular,
with a melting point near 402 K and a room temperature vapour pressure of around
600 Pa (see refs 55 and 57). Its gas-phase vibrational transitions (two IR-active and
one Raman-active) again indicate centrosymmetry and thus a linear geometry, in
agreement with VSEPR predictions.
The possibility that radon might exhibit similar chemistry was first explored
by Lawrence Stein and co-workers58. They discovered experimental evidence
that radon forms a difluoride, RnF2, which is considerably less volatile than XeF2,
remaining stable at 200 °C but decomposing to its elements in vacuo above 250 °C
(see ref. 59). This prompted Pitzer to investigate the energetics of fluoride formation
of Xe, Rn, and Og using relativistically-calculated energies24. Based on the energy
cost of formation of the valence state of the noble gas atom for bonding (equal to
+ , and estimated at 0.44 eV, 1.42 eV, and 3.93 eV for Xe,
Rn, and Og, respectively), he concluded that ionic rather than molecular crystalline
forms were likely for RnF2 and OgF2. (Figure 3 shows the relative energies of these
species as obtained by Pitzer.)

Figure 3 Energies of states related to the fluorides of Xe, Rn, and Og; v.s. refers to the
valence state formed by mixing the P1/2 and P3/2 states (after Pitzer37).

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 111


In 1999, Nash and Bursten considered relativistic effects on the favoured
structures of the noble gas tetrafluorides using RCI (relativistic configuration
interaction) methods, with all but the valence electrons in Og treated using shape-
consistent RECPs60. These calculations were carried out using a modified version
of the quantum chemistry program COLUMBUS61,62. They calculated optimal
structures constrained to D4h (square planar) or Td (tetrahedral) symmetries, and
concluded that while in XeF4 and RnF4 the D4h geometry is strongly favoured (with
no clear Td minimum being found in either case, and with the Td curve lying about
5 eV above the D4h curve in the XeF4 case, and by about half of that amount in the
RnF4 case), in OgF4 the Td and D4h curves almost coincide, with the D4h curve lying
higher in energy by about 0.25 eV (24 kJ mol–1) and both curves showing energy
minima at a bond length near 2.14 Å. While the authors acknowledge that these
calculations may not be definitive, this result at least suggests that relativistic effects
may lead to stereochemical non-rigidity in OgF4 and, more generally, to deviation
from the predictions of VSEPR theory.
The initial impetus to explore noble gas chemistry came from Bartlett’s
observation that the ionisation potentials of O2 and Xe were amost equal – 12.5
eV for O2 and 12.1 eV for Xe – implying that if the former could be oxidised by
platinum hexafluoride, so could the latter. In general, the first ionisation energy
decreases steadily down the noble gas group (as it does down other groups on the
periodic table), leading to extreme unreactivity in the lighter noble gases He, Ne, and
Ar, and to increasingly rich chemistry for Kr and Xe. Cationic Rn (ionisation energy
10.75 eV, not much higher than that of Hg) has been shown to displace Na+ and K+
ions in resins, leading to the suggestion that it can be characterised as a metalloid
element59,63. The FSCC calculations of Jerabek et al. yield an ionisation potential of
8.842 eV for the 7p3/2 spinor of Og (see ref. 14) (about 1 eV larger than the values
obtained by Nash and Bursten60); this is close to the values of the noble metals Os,
Ir, and Pt.

4. Electron and nucleon localisation


The notion of localised groups of electrons has been central to chemistry since
the discovery by Gilbert Lewis of the central role of the electron pair in covalent
bonding64. The electron pair concept is counterintuitive in the light of the Coulomb
repulsion between particles of like charge, and within the quantum picture
emerges from the interplay of the electron kinetic energy and the Pauli principle.
Finding chemically transparent descriptions of such localised electrons within
quantum-mechanical wavefunctions obtained by electronic structure theory is not
straightforward because the canonical molecular wavefunctions are delocalised
(molecular orbitals) and the unitary transformations that can localise them are not
unique, and can lead to apparently different qualitative pictures. The idea of seeking
such a description within the density matrix was first advanced by Lennard-Jones in
195265; by integrating over the spatial and spin co-ordinates of all electrons but two,
a function is obtained which gives the probability of finding a pair of electrons, with
particular assigned spins, in some region of space.

112 Roderick M. Macrae and Terence J. Kemp


A useful measure of electron localisation was given in 1990 by Becke and
Edgecombe66. These authors developed an expression incorporating the electronic
density pσ of spin σ, its gradient , and the kinetic energy density τσ, which serves
as an electron localisation function (ELF). This function has the form

where
1
4
(an expression that gets smaller with increasing localisation) and
is a term corresponding to the uniform electron gas (Thomas–
Fermi gas) with spin density equal to the local value of pσ (r). The ELF has an upper
limit of 1 corresponding to complete localisation, while the value of 1/2 corresponds
to a pair probability equal to that of the uniform electron gas. Becke and Edgecombe
applied this model to the noble gas atoms Ne, Ar, Kr, Xe, and Rn, and found it to
reveal clear evidence of the existence of 2, 3, 4, 5, and 6 distinct electronic shells in
these atoms. For a general discussion of the ELF and its possible limitations, see the
2005 article by Savin67.
Jerabek et al. utilised an appropriately modified version of the ELF [denoted
] as implemented in the the relativistic ab-initio quantum chemistry program
DIRAC1550 to compare electron localisation in Xe, Rn, and Og, and to consider
relativistic effects on electron localisation14. Figures 4 and 5 show the ELFs predicted
from the authors’ calculations. From Figure 4 it can be seen that electron localisation
in Xe barely changes in going from the non-relativistic to the four-component
relativistic framework. Rn represents an intermediate case, with a decrease in the
distinctness of the shell structure being visible in the outer core region near 50 pm.
For Og, by contrast, the ELF values are much smaller than for the non-relativistic
case across the whole range of radii from 0 to 100 pm, making the atomic shell
structure barely recognisable.
Figure 5 gives some further clarification of this effect by showing a comparison
of the ELFs for Xe and Og computed using non-relativistic (NR), scalar-relavistic
(SR) and fully relativistic four-component Dirac–Hartree–Fock (R) models. In Xe,
the inclusion of relativistic effects in the calculation has remarkably little effect on
electron localisation. In Og, on the other hand, even in the SR picture, the relativistic
contraction of the inner shells and smearing out of the shell structure in the outer
shells is quite clear. The ELF in Og is more homogeneously distributed over the
entire atomic range compared to Xe, resulting in values undergoing fairly small
oscillations around the Thomas–Fermi limit; this is particularly true in the three
outermost shells.
Note that the ELF is a completely different quantity from electron density, which
decreases steadily with increasing distance from the nucleus, and from the radial
density distribution 4πr 2p(r), which is a good indicator of shell structure in light
atoms but is much less successful for heavy atoms68.

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 113


Figure 4 ELFs from non-relativistic (NR, left) and Dirac–Hartree–Fock calculations (R,
right) for the heavy rare gas atoms Xe (top), Rn (middle), and Og (bottom). Reprinted figure
with permission from Jerabek et al.14. Copyright 2018 by the American Physical Society.

The discovery of ‘magic numbers’ of protons and neutrons leading to


exceptional isotopic stability led to the development of a ‘shell model’ of the nucleus
in which protons and neutrons (which are fermions) occupy energy levels somewhat
analogous to the electron shells in the atom. (Simpler models, such as the liquid drop
model and the Fermi gas model, predict that the nuclear properties vary smoothly
with the number of nucleons69.) In this picture the nuclear force, dominated by
the strong interaction, acts primarily as an average central potential, most simply
modelled by the three-dimensional harmonic oscillator potential, which has exact
solutions, or more realistically by potentials such as the Woods–Saxon potential. To
this potential is added the spin-orbit coupling interaction, which splits levels in the
opposite way to the electronic case (p3/2 lower than p1/2). The protons and neutrons are
treated essentially independently. With a realistic potential and the incorporation of
spin-orbit coupling, the correct magic numbers 2, 8, 20, 28, 50, 82, 126 – the same
for protons and neutrons – are found. ‘Doubly magic’ nuclei, such as He, exhibit
exceptional stability. The development of this model was the work for which Maria

114 Roderick M. Macrae and Terence J. Kemp


Figure 5 ELFs for Xe (a) and Og (b) from non-relativistic (NR), scalar-relativistic (SR)
and Dirac–Hartree–Fock (R) calculations as a function of the distance from the nucleus.
The relativistic contraction and smearing-out of the shell structure in the valence and
sub-valence shells of Og are clear. Reprinted figure with permission from Jerabek et al.14.
Copyright 2018 by the American Physical Society.

Goeppert Mayer shared the Nobel Prize in Physics in 1963 with Hans Jensen and
Eugene Wigner.
While the shell model can predict a variety of nuclear properties such as spin and
parity, as with electrons in atoms and molecules quantitative predictions of properties
require an improved approach. The methods available for such computations
are essentially the same as those used in electronic structure theory, namely self-
consistent mean-field models such as Hartree–Fock theory (with an additional
‘Skyrme force’ appearing in the two-body term) and density functional theory.
Such calculations have recently been applied to superheavy nuclei in order to study
stability and shell structure70,71.
Jerabek et al.14 extended their calculations to the issue of nuclear localisation
using nuclear density functional theory with the optimised global Skyrme energy
density functionals UNEDF1 and SV-min (see refs 72 and 73), making use of
the density functional theory (DFT) solver of Reinhard74 constrained to spherical
geometry.
Figure 6 shows the nucleon localisation functions [NLFs, denoted n and
p ] for three spherical nuclei: the medium-mass doubly magic nucleus 132
Sn, the

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 115


Figure 6 NLFs of 132Sn, 302Og, and 472164 calculated with the energy density functional
UNEDF1. Reprinted figure with permission from Jerabek et al.14. Copyright 2018 by the
American Physical Society

superheavy nucleus 302Og, and the (highly) theoretical nucleus 472164. In contrast
to the ELFs, the number of closed shells cannot be determined from the number of
radial maxima owing to the different radial behaviour of electron and nucleon orbits.
While the radii of electronic orbits in atoms belonging to different shells are spatially
well separated, the radii of nucleonic orbits scale only as (2nr +l)1/2, in other words
they increase very gradually with n. Thus, there is a large spatial overlap between
single-nucleon wave functions, and reduced localisation as compared to the radial
distribution of the ELFs. A noteworthy feature of the NLFs is the local enhancement
at the nuclear surface, owing to the relatively small number of valence neutrons
contributing to the total density at distances greater than the nuclear radius.
Figure 6 reveals the following: (1) The pattern of concentric rings is more
distinct in the proton system compared with the neutron system. This is because the
number of occupied proton shells is less than for neutrons; this effect becomes more
prominent for superheavy nuclei where the neutron excess is large. (2) The NLF

116 Roderick M. Macrae and Terence J. Kemp


for the medium-mass 132Sn nucleus shows a clear shell structure; however, the shell
structure becomes fuzzier as the atomic number increases, as can be seen for 302Og
and 472164, particularly in the case of the neutrons. (3) While the neutron NLF for
302
Og retains a structure in its interior, this has vanished for 472164.
5. Conclusions
From their calculations based on fermion localisation, Jerabek et al.14 conclude:
(1) Relativistic effects play a major role in inducing large spin-orbit splitting of the
electrons, resulting in the behaviour approaching that of an electron gas in the valence
region. (2) Associated with point 1, a large static polarisability will be conferred upon
the Og atom, enabling it to participate significantly in van der Waals interactions.
This is consistent with earlier results by Pershina et al.31. (3) A gradual transition
towards the uniform-gas regime is predicted for nucleonic localisation functions in
superheavy nuclei. Because N is greater than Z, more neutrons are confined to the
same volume as protons and will therefore be more delocalised. While in general
the Thomas–Fermi limit is only reached at very high nucleon numbers, in the case
of the neutron NLFs the nucleon gas limit is reached locally in the valence region
(r > 3 fm), just before the final surface peak.
Thus, while Og fits into the general trend of the heavier noble gases to exhibit
chemistry and behaviour to some extent resembling that of a metalloid, Og has
quantitatively much larger relativistic effects in its electronic structure, leading to a
huge splitting of the 6p1/2 and 6p3/2 spinor levels, and as a result exhibits much weaker
shell structure as measured by the ELF. Additionally, as measured by the NLF its
nuclear structure shows a ‘bubble-like’ character, with a Thomas–Fermi-like interior
and a ‘skin’ exhibiting a local peak in nucleon localisation. This is particularly true
for the neutrons. As with the ELF, this should not be confused with the nucleon
density distribution, where in turn it should be clarified whether it is the distribution
of charge (protons) or matter (protons + neutrons) that is being discussed75; based on
electron scattering and muon capture experiments, these quantities are both relatively
flat in the nuclear interior (and only weakly dependent on Z or A), tapering off fairly
rapidly at the nuclear surface.
Calculations on element 118, now known to the world as oganesson, suggest
that it might be a liquid at or near room temperature, and that it might exhibit
interesting chemistry, possibly including the formation of a tetrahedral (or fluxional)
tetrafluoride. However, the question remains of whether it can ever be studied
experimentally. The only isotope presently known, 294Og, has a very short half-life
that precludes experimental study using the established methods. It is possible that
the closed-shell isotope 302Og may have a longer half-life, if it can ever be made, at
which point experimental exploration of the properties of this interesting element
might be able to begin.
6. Footnote: nomenclature
Oganesson is named after the Russian scientist Yuri Oganessian (Figure 7) who led
the group working on superheavy elements at the JINR in Dubna. Element 118

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 117


Figure 7 Professor Yuri Oganessian after whom oganesson was named. Credit: VPRO,
https://commons.wikimedia.org/wiki/File:Yuri_Oganessian.jpg.

was first made in 2002, but it received its name in recognition of the achievements
of Professor Oganessian and his group by IUPAC in 2015. It is currently the only
element named after a living person.

Published online: 18 April 2018

7. References
1. Oganessian, Y.T., Utyonkov, V.K., Lobanov, Yu.V., et al. (2006) Phys. Rev. C, 74, 044602.
2. Oganessian, Y.T. (2006) Pure Appl. Chem., 78, 889.
3. Karol, P., Barber, R., Sherrill, B., et al. (2016) Pure Appl. Chem., 88, 155–160.
4. Hofmann, S. and Münzenberg, G. (2000) Rev. Mod. Phys., 72, 733–767.
5. Morita, K., Morimoto, K., Kaji, D., et al. (2007) J. Phys. Soc. Jpn, 76, 045001.
6. Oganessian, Y.T., Abdullin, F.Sh., Bailey, P.D., et al. (2010) Phys. Rev. Lett., 104, 142502.
7. Oganessian, Y.T. (2007) J. Phys. G, 34, R165.
8. Ninov, V., Gregorich, K.E., Loveland, W., et al. (1999) Phys. Rev. Lett., 83, 1104–1107.
9. Hofmann, S. (1999) Experimental nuclear physics in Europe ENPE 99. In: Rubio, B., Lozano, M.
and Gelletly, W. (eds), AIP conference proceedings No. 495, p. 137. AIP, Melville, NY.
10. Oganessian, Y.T., Utyonkov, V.K., Lobanov, Yu.V., et al. (2000) Phys. Rev. C, 62, 041604(R).
11. Hoffman, D.C. and Lee, D.M. (1999) J. Chem. Educ., 76, 332–347.
12. Curie, P. and Curie, M. (1898) C. R. Acad. Sci. Paris, 127, 1215–1217.
13. Merinis, J. and Bouissieres, G. (1961) Anal. Chim. Acta, 25, 498–504.
14. Jerabek, P., Schuetrumpf, B., Schwerdtfeger, P. and Nazarewicz, W. (2018) Phys. Rev. Lett., 120,
053001. https://dopi.org/ 10.1103/PhysRevLett.120.053001.
15. Schwarz, W.H.E. (2010) An introduction to relativistic quantum chemistry. In: Barysz, M. and
Ishikawa, Y. (eds), Relativistic methods for chemists, pp. 1–62. Springer, Dordrecht.
16. Pershina, V. (2016) EPJ Web Conf., 131, 07002.
17. Pyykkö, P. (1988) Chem. Rev., 88, 563–594.
18. Dyall, K.G. and Fægri, K., Jr. (2007) Introduction to relativistic quantum chemistry. Oxford
University Press, Oxford.
19. Powell, R.E. (1968) J. Chem. Educ., 45, 558–563.
20. Pyykkö, P. (2012) Chem. Rev., 112, 371–384.
21. Rose, S.J., Grant, I.P. and Pyper, N.C. (1978) J. Phys. B, 11, 1171–1176.
22. Pyykkö, P. and Desclaux, J.-P. (1979) Acc. Chem. Res., 12, 276–281.

118 Roderick M. Macrae and Terence J. Kemp


23. Pyykkö, P. (1978) Adv. Quantum Chem., 11, 353–409.
24. Pitzer, K.S. (1975) J. Chem. Soc. Chem. Commun., 0, 760b–761.
25. Türler, A. and Pershina, V. (2013) Chem. Rev., 113, 1237–1312.
26. Eliav, E., Fritsche, S. and Kaldor, U. (2015) Nucl. Phys. A, 944, 518–550.
27. Eliav, E. and Kaldor, U. (2010) Four-component electronic structure methods. In: Barysz, M. and
Ishikawa, Y. (eds), Relativistic methods for chemists, pp. 279–349. Springer, Dordrecht.
28. Hohenberg, P. and Kohn, W. (1964) Phys. Rev., 136, B864–B871.
29. Anton, J., Fricke, B. and Engel, E. (2004) Phys. Rev. A, 69, 012505.
30. Engel, E. (2002) Theor. Comput. Chem., 11, 523–561.
31. Pershina, V. (2015) Nucl. Phys. A, 944, 578–613.
32. Eliav, E. and Pyykkö, P. (1996) Phys. Rev. Lett., 77, 5350–5352.
33. Haynes, W.M. (2011–2012) Handbook of chemistry and physics, 92nd edition. CRC Press, Boca
Raton, FL.
34. Goidenko, I., Labzowsky, L., Eliav, E., et al. (2003) Phys. Rev. A, 67, 020102(R).
35. Indelicato, P., Bieroń, J. and Jönsson, P. (2011) Theor. Chem. Acc., 129, 495–505.
36. Hemingway, J.D. (1972) Superheavy elements. In: Newton, G.W.A. and Hemingway, J.D. (eds)
Radiochemistry, vol. 1, pp. 38–68. Royal Society of Chemistry Specialist Periodical Reports,
London.
37. Pitzer, K.S. (1975) J. Chem. Phys., 63, 1032–1033.
38. Nash, C.S. (2005) J. Phys. Chem. A, 109, 3493–3500.
39. Visscher, L., Visser, O., Aerts, P.J.C., et al. (1994) Comput. Phys. Commun., 81, 120–144.
40. Pershina, V., Anton, J. and Jacob, T. (2009) J. Chem. Phys., 131, 084713–84720.
41. Pershina, V., Borschevsky, A., Eliav, E. and Kaldor, U. (2008) J. Chem. Phys., 128, 024707–
24715.
42. Pershina, V. and Bastug, T. (2005) Chem. Phys., 311, 139–150.
43. Gäggeler, H.W. (2013) Status and perspective of chemical studies of heaviest elements. In:
Penionzhkevich, Yu.E. and Sobolev, Yu.G. (eds), Exotic nuclei: EXON-12 – Proceedings of the
international symposium, pp. 137–146. World Scientific, Singapore.
44. Yakushev, A., Gates, J.M., Türler, A., et al. (2014) Inorg. Chem., 53, 1624–1629.
45. Dmitrieva, I.K. and Plindov, G.I. (1983) Phys. Scr., 27, 402–406.
46. Bondi, A. (1964) J. Phys. Chem., 68, 441–451.
47. Politzer, P., Jin, P. and Murray, J.S. (2002) J. Phys. Chem., 117, 8197–8202.
48. Jensen, H.J.A., Saue, T. and Visscher, L. (n.d.) http://dirac.chem.sdu.dk [accessed 1 February
2018].
49. Dyall, K.G. (2006) Theor. Chem. Accts., 115, 441–447.
50. Bast, R., Saue, T., Visscher, L., et al. (2015) DIRAC15: A relativistic ab initio electronic structure
program. http://www.diracprogram.org [accessed 1 February 2018].
51. Bartlett, N. (1962) Proc. Chem. Soc., 218.
52. Moody, G.J. (1964) J. Chem. Educ., 51, 628–630.
53. Claassen, H.H., Selig, H. and Malm, J.G. (1962) J. Amer. Chem. Soc., 84, 3593.
54. Haner, J. and Scholbilgen, G.J. (2015) Chem. Rev., 115, 1255–1295.
55. Schreiner, F., McDonald, G.N. and Chernick, C.L. (1968) J. Phys. Chem., 78, 1162–1166.
56. Claassen, H.H., Chernick, C.L. and Malm, J.G. (1963) J. Amer. Chem. Soc., 85 1927–1928.
57. Tramšek, M. and Žemva, B. (2006) Acta. Chim. Slov., 53, 105–116.
58. Fields, P.R., Stein, L. and Zirin, M.H. (1962) J. Amer. Chem. Soc., 84, 4164–4165.
59. Stein, L. (1987) Chemical properties of radon. In: Hopke, P.K. (ed), Radon and its decay products,
ACS Symposium Series, Vol. 331, pp. 240–251. ACS Publications, Washington DC.
60. Nash, C.S. and Bursten, B.E. (1999) J. Phys. Chem. A, 103, 402–410.
61. Shepard, R., Shavitt, I., Pitzer, R.M., et al. (1988) Int. J. Quantum Chem., 34, 149–165.
62. Ermler, W.C., Ross, R.B. and Christiansen, P.A. (1988) Adv. Quantum Chem., 19C, 139–182.
63. Stein, L. (1985) J. Chem. Soc. Chem. Commun., 1631–1632.

www.scienceprogress.co.uk Oganesson: a most unusual ‘inert gas’ 119


64. Lewis, G.N. (1916) J. Amer. Chem. Soc., 38, 762–785.
65. Lennard-Jones, J.E. (1952) J. Chem. Phys., 20, 1024–1029.
66. Becke, A. and Edgecombe, K.E. (1990) J. Chem. Phys., 92, 5397–5403.
67. Savin, A. (2005) J. Mol. Struct.–THEOCHEM, 727, 127–131.
68. Simas, A.M., Sagar, R.P., Ku, A.C.T. and Smith, V.H., Jr. (1988) Can. J. Chem., 66, 1923–1930.
69. Hodgson, P.E., Gadioli, E. and Gadioli Erba, E. (1997) Introductory nuclear physics. Oxford
Science Publications, Clarendon Press, Oxford.
70. Rutz, K., Bender, M, Bürvenich, T., et al. (1997) Phys. Rev. C, 56, 238–243.
71. Bender, M., Rutz, K., Reinhard, P.-G., et al. (1999) Phys. Rev. C, 60, 034304.
72. Kortelainen, M., McDonnell, J., Nazarewicz, W., et al. (2012) Phys. Rev. C, 85, 024304.
73. Klüpfel, P., Reinhard, P.-G., Bürvenich, T.J. and Maruhn, J.A. (2009) Phys. Rev. C, 79, 034310.
74. Reinhard, P.-G. (1991) The Skyrme–Hartree–Fock model of the ground state. In: Langanke, K.,
Koonin, S. and Maruhn, J. (eds), Computational nuclear physics I – nuclear structure, pp. 28–50.
Springer, Berlin.
75. Wilets, L. (1959) Science, 129, 361–367.

120 Roderick M. Macrae and Terence J. Kemp


Reproduced with permission of copyright owner. Further
reproduction prohibited without permission.

You might also like