Aguado - 2013 - Structures, Relative Stabilities, and Electronic Properties of Potassium Clusters KN (13 N 80)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Computational and Theoretical Chemistry 1021 (2013) 135–143

Contents lists available at SciVerse ScienceDirect

Computational and Theoretical Chemistry


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

Structures, relative stabilities, and electronic properties of potassium


clusters Kn (13 6 n 6 80)
Andrés Aguado
Departamento de Física Teórica, Atómica y Óptica, Universidad de Valladolid, Valladolid 47071, Spain

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

Article history: We report the global minimum (GM) structures, energetic and electronic properties of Kn clusters with up
Received 24 April 2013 to 80 atoms, obtained by combining basin hopping unbiased optimizations (based on a many-body
Received in revised form 26 June 2013 empirical potential) with subsequent reoptimization of several candidate structures employing a density
Accepted 29 June 2013
functional theory method which accounts for van der Waals dispersion interactions (vdW-DFT). A com-
Available online 5 July 2013
parison with other alkali cluster systems shows that the GM structures of Kn coincide with those of Nan
more often than with those of Csn clusters. Nevertheless, the inclusion of dispersion interactions produces
Keywords:
for several clusters changes in the GM structures, which become either identical to those of Csn clusters,
Potassium clusters
Density functional theory
or completely novel (i.e. not reported for other alkali systems). Many GM structures are based on poly-
van der Waals interactions icosahedral packing; some others contain Kasper polyhedral units with disclination lines. Atoms with a
Magic clusters low coordination number (or adatoms) are avoided in the GM structures following a geometric shell clos-
Global optimization ing: such additional atom is instead inserted into the outermost atomic shell of the cluster. The calculated
Basin hopping ionization potentials, electron affinities and HOMO–LUMO gaps are found to reproduce perfectly the elec-
tron shell closings observed in photoionization measurements. The electric dipole moments of most clus-
ters are close to zero, a distinguishing feature of metallicity previously observed for Nan clusters.
However, some clusters like K26, K47 or K49 have sizable electric dipole moments. The cluster stabilities
are found to explain the experimental abundances inferred from mass spectra. The enhanced stabilities
(magic numbers) can be interpreted in terms of electron shell closing effects for the smaller clusters and
of geometrical packing effects for the larger clusters. Although there is not a sharply defined critical size
separating both regimes, it is estimated to be around 55 atoms.
Ó 2013 Elsevier B.V. All rights reserved.

1. Introduction ject of study because they are representative of simple metal clus-
ters with delocalized electrons. In particular, first principles
There has been an ongoing interest during last years in under- methods such as quantum chemistry methods and density func-
standing the geometric packing preferences of small atomic clus- tional theory have been applied to study many structural, thermal,
ters. Together with the electronic structure, cluster geometry electronic and optical properties of sodium clusters [4–31], con-
determines several physical and chemical properties such as reac- tributing much to our present understanding of the physico-chem-
tivity and optical absorption, which are of general interest in Nano- ical properties of alkali metal systems.
science. Typically, the structure and other cluster properties can The number of studies devoted to potassium clusters is compar-
differ substantially from those of isolated molecules and macro- atively much smaller. This is somehow surprising, because on the
scopic systems, due to quantum confinement effects on the elec- experimental front, results for the abundance mass spectra, ioniza-
tron cloud and also to the large proportion of surface atoms. tion potentials (IPs), polarizabilities, optical absorption properties,
Alkali metal clusters were amongst the first nanoscopic systems and relevant fragmentation channels are available since the early
investigated with mass spectroscopic techniques [1]. Those semi- days of cluster physics [32–37]. In particular, the measured IP val-
nal studies revealed that clusters with 8, 20, 34, 40, 58, . . . electrons ues show sharp drops (an indication of an electronic shell closing)
have an enhanced stability as compared to other cluster sizes. The right after the sizes N = 8, 18, 20, 34, 40, 58, 92 and weaker but sig-
magic sizes were understood in terms of a well defined electron nificant features for N = 14, 26, 30 in good agreement with the pre-
shell structure with the help of the structureless jellium model dictions from a spheroidal Nilsson model [38]. The electron shell
[2,3]. Since then, alkali clusters have remained an interesting ob- closings expected at N = 68, 70 on the base of jellium model predic-
tions are not observed in the experiments, however. A liquid drop
model study incorporating shell-correction terms and a consider-
E-mail address: aguado@metodos.fam.cie.uva.es ation of electronic entropy [39] was also able to reproduce all the

2210-271X/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.comptc.2013.06.040
136 A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143

experimental maxima in the IP, and additionally predicting new results on Na and Cs clusters cast some doubts on this assumption.
weaker maxima at sizes N = 46, 50 and 54, which would be ob- It is therefore interesting to explicitly check its validity.
served only at very low temperatures. The abundance mass spectra The paper is organized as follows: The theoretical approach for
show that all clusters with an electron shell closing are magic, locating the GM structures is briefly summarized in Section 2,
which demonstrates that the enhanced stabilities are correlated where we also include some benchmark tests of the accuracy of
with the electronic shell structure. No enhanced abundance is ob- our DFT calculations. Section 3 describes the structures, stabilities,
served at N = 55, where an icosahedral geometric shell closing is and also some electronic properties of Kn clusters. Finally, Section 4
expected, probably because the clusters in the experimental beam summarizes the main conclusions of our work.
are in a hot liquid phase. Secondary magic numbers (associated
with less prominent maxima in the abundance mass spectra) are
observed for N = 14, 26, 30 [36]. These features at N = 14, 26, 30 2. Theoretical methods
have been recently confirmed in an experimental study of the

abundances of Kþ N and KN cluster ions extracted from helium drop- The methods and theoretical level employed to locate the puta-
lets [40]. This study identifies also a substantial odd–even alterna- tive GM structures are the same as in our previous works on so-
tion in the abundances of clusters containing from 34 to 40 dium [27] and cesium [55] clusters. Briefly stated, the first-
electrons, and a local stability maximum for clusters with about principles calculations are performed at the Kohn–Sham [56] DFT
46 electrons. [57] level. We employ the SIESTA code [58] with different imple-
The first theoretical calculations [38,39,41,42] on potassium mentations of the exchange–correlation functional: the spin-polar-
clusters were based on the jellium model and its variants and, as ized local density approximation (LDA functional), as
mentioned in the previous paragraph, were quite successful in parameterized by Perdew–Zunger [59] and based on Ceperley–Al-
reproducing and explaining the results of photoionization experi- der simulations of the electron gas [60]; the spin-polarized gener-
ments. Nevertheless, those models do not take into account the de- alized gradient approximation (GGA), as implemented by Perdew,
tailed structure of the ionic skeleton, which is essential to Burke and Ernzerhof (PBE functional) [61]; and the fully non-local
understand the properties of solid-like clusters and the possible van der Waals functional (vdW-DFT), as proposed by Klimes, Bow-
impact of geometric packing on their stabilities. First-principles ler, and Michaelides (KBM functional) [62]. This functional has
studies based on either density functional theory (DFT) or quantum been implemented by Román-Pérez and Soler [63] in the SIESTA
chemical methods were reported for very small clusters containing code. The main results are obtained with the KBM functional.
up to about 10 atoms only [43–51]. Banerjee et al. [52] reported ab LDA and PBE are used only for benchmark purposes and also to
initio results for KN clusters containing an even number of atoms perform an initial, lower-cost optimization of cluster structures,
ranging from N = 2 to N = 20. In this study, no global optimization which are then reoptimized with the KBM functional.
of cluster structures was attempted, however, and the structures of Each potassium atom contributes one electron to the valence
KN were simply assumed to be the same as those of NaN clusters. electron density. The effect of core electrons is represented by
Very recently, Abdalla and coworkers [53] have reported a global norm-conserving pseudopotentials [65] which contain scalar rela-
optimization study of potassium clusters with up to 20 atoms, tivistic and non-linear core corrections [66]. The cluster wave func-
employing a genetic algorithm coupled to a density functional tion is expanded into a basis set of localized atom-centered
tight binding (DFTB) method. In this work, the sampling of the po- orbitals, of DZP2 quality. A 100 Ryd cutoff is employed for the fast
tential energy surface is extensive, so the quality of the results de- Fourier transform grid. Table 1 shows benchmark results obtained
pends mainly on the accuracy of the approximations employed to for the dimer and bulk limits with these computational settings.
build the DFTB model. Specifically, the model is parametrized by The KBM functional is found to perform much better than either
first assuming that all potential energy matrix elements involving LDA or GGA approximations in describing the several properties,
non-nearest-neighbor atoms are negligible, and then fitting to DFT suggesting that vdW corrections are not negligible for potassium.
results on the K2 molecule obtained with the B3LYP functional. We This observation is in line with the results of previous works
have not found any other first-principles study dealing with potas- [67,68].
sium clusters of larger size. We would like to emphasize that the improvement obtained
The aim of the present work is to provide a more definitive when using the dispersion corrected functional is a direct conse-
assessment of the structures of Kn clusters in a broad size range quence of a more accurate representation of the interactions, and
n = 13–80, employing global optimization methods and ab initio does not happen by chance. In a metal, the dispersion interaction
DFT calculations that incorporate a description of van der Waals can be understood as the response of the density of the electron
dispersion interactions (vdW-DFT) [54]. The main motivation for
this study comes from the scarcity of theoretical results for
N = 13–20 and the total absence of them for bigger clusters. We
Table 1
would like to reproduce and interpret the experimental results Equilibrium bond distance d, cohesive energy Ecoh, and vertical ionization potential
about ionization potentials and cluster abundances in a broad size (VIP) of the K2 dimer, and lattice constant a, bulk modulus B and cohesive energy Ecoh
range, and also assess the possible impact of geometric shell struc- of bulk potassium, as predicted by several exchange–correlation functionals, are
ture onto the stability of cold, solid-like, clusters. Another motiva- compared to experimental results. Experimental values are taken from Ref. [64] for
bulk potassium, and from Ref. [48] for the dimer.
tion is to offer a comparison with our previous results on sodium
[27] and cesium [55] clusters. In those works, we demonstrated K2 d (Å) Ecoh (eV) VIP (eV)
that NaN and CsN clusters adopt very different global minimum LDA 3.799 0.331 4.498
(GM) structures for most sizes, a result which was a priori quite PBE 3.959 0.238 4.180
unexpected. Also, while most NaN clusters adopt the lowest possi- KBM 3.930 0.251 4.092
Exp. 3.924 0.260 4.061
ble spin multiplicity, many CsN clusters have high magnetic mo-
ments and behave indeed as magnetic superatoms. The K (bulk) a (Å) Ecoh (eV) B (GPa)

additional effects of vdW interactions and sd-hybridization, which LDA 5.157 1.03 4.41
are much more important for cesium than for sodium, explain the PBE 5.264 0.87 3.39
KBM 5.219 0.92 3.83
differences. Although potassium clusters have been many times
Exp. 5.212 0.94 3.70
assumed to be very similar to sodium clusters [52], the previous
A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143 137

gas (formed by the valence electrons) to its own fluctuations, up to 100 (depending on cluster size) varied structural motifs for
mediated by the correlations that exist between charge density subsequent optimization at the DFT level. The selection is based
fluctuations localized at distant (non-overlapped) regions. This on several structural indicators as detailed in our previous work
interaction, which is carried out by plasmons, is a non-local [27], which aims at maintaining a high structural diversity within
many-body correlation effect which cannot be described by any the considered sample. DFT optimizations were first performed
semilocal correlation functional. In the work of Dion et al. [54], dis- with the LDA functional because of its lower computational ex-
persion is defined through a many-body treatment of the electron pense. Finally, we re-optimized at the vdW-DFT level all the iso-
gas involving the response function of the metal. The functional of mers which, according to the LDA calculations, have total
Dion et al. [54] is an explicit non-local functional of the electron energies up to 0.4 eV above the putative GM. All the equilibrium
density, of the following form: geometries shown in the next section were obtained through
Z Z unconstrained conjugate gradients optimization, until the force
1 on each atom was smaller than 0.005 eV/Å.
EvdW
nl ¼ dr1 dr2 nðr1 Þ/ðr1 ; r2 Þnðr2 Þ; ð1Þ
2
where n(r) is the valence electron density. The kernel / satisfies two 3. Results and discussion
important properties: (1) it correctly tends to zero for homogeneous
densities, so the interaction occurs between density fluctuations 3.1. Global minimum structures
and (2) at long distances, it reproduces the correct asymptotic
dependence of the dispersion energy, proportional to 1=r 612 . Notice Fig. 1 shows a representative sample of the GM structures lo-
that the functional depends exclusively on electronic variables, cated in the size range N = 13–34. For N = 13–15 we obtain oblate
and in particular r12 does not correspond to an interatomic distance deltahedral structures based on the packing of pentagonal bipyra-
but to the distance between two localized charge density fluctua- mid units. In the size range N = 16–18, the clusters have a single
tions. The accuracy of this method for metallic solids has been re- endohedral atom enclosed inside a high-symmetry polyhedral
cently demonstrated [68]. shell. The shell of K16 is the so-called Z15 Kasper polyhedron
Just to complete our description, we notice that the functional [74], containing three disclination lines, although an isomer based
form of Dion et al. [54] is appropriate for the electron gas of a me- on a defective icosahedron is almost equally stable. Similarly, K17
tal. The group of Perdew has proposed an alternative point of view and the C2v isomer of K18 contain four and five disclination lines,
to describe the physics of dispersion interactions in metals [67], respectively. The Cs isomer of K18 is instead built by adding one
which is more appropriate to the limit of a Wigner crystal, where atom to the structure of K17. Another structure with a single central
electrons localize at lattice sites and are separated by regions of atom is the C3 isomer of K20, which was also obtained in a recent
very low density. Within this framework, the dispersion interac- study of sodium clusters [29]. From N = 19 to N = 34 the structures
tion in solid potassium would be considered to arise between are dominated by poly-icosahedral (pIh) packing. For example, the
K+–K+ ion pairs, and would be screened (as compared to its value structures of K19–K23 are based on the double icosahedron, a triple
in free space) by the valence electron density. This screening effect icosahedron is obtained for K24, while one of the degenerate struc-
results in an effective interaction between ions which is calculated tures of K27 contains four K13 icosahedral units. Clusters with
via second-order time-dependent perturbation theory and involves N = 28–33 atoms are obtained by simply removing atoms from
once more plasmonic excitations, somehow connecting this point the perfect core–shell pancake structure of K34. This structure is
of view to that of Dion et al. The difference in philosophy between formed by interpenetrating seven K13 icosahedral units. The seven
both methods is clear. In the method of Perdew, the dispersion en- core atoms are in a pentagonal bipyramid arrangement, which re-
ergy correction is written in a more ‘‘classical’’ way as sults in a high D5h point group symmetry.
EvdW
nl ¼ C 6 =R6ij , where Rij is an interatomic distance, and the Comparing our results for N = 13–20 with those of Abdalla et al.
many-body screening effect of the valence electron gas is included [53] we find a qualitative agreement for N = 13–16 and N = 20.
only in the calculation of the C6 coefficient. Both methodologies, These authors also locate deltahedral structures for N = 13–16,
however, provide results of similar accuracy for the alkali metals. although not exactly coinciding with our predictions. For K20, the
In this work, we have followed the functional proposed by Dion GM of Abdalla and coworkers coincides with our C3 isomer. How-
et al. [54]. ever, for N = 16–19 their structures are quite different from ours,
For each cluster size, we need to generate first a number of rea- and of much lower symmetry. For all clusters with N = 13–20, we
sonable trial structures, which are then introduced as initial geom- have checked that Abdalla’s structures are at least local minima
etries for the first-principles unconstrained relaxations. To this at the level of DFT-KBM calculations. The differences may originate
end, we have employed unbiased basin hopping optimizations from the different levels of theory employed and from the approx-
[69] based on a many-body Gupta empirical potential [70–72]. imations inherent to the parametrization of their tight-binding
The appropriate parameters for potassium are given by Li et al. model. A better agreement is found between our results and those
[73]. From the pool of structures thus generated, we have selected reported by Banerjee et al. [52], who considered just clusters with

13(C2 ) 14(C2 ) 16(D3h ) 16(C1) 17(D2) 18(Cs) 18(C2v ) 20(Cs)

20(C3) 21(C2v ) 24(C2v ) 26(Cs) 27(C2v ) 27(C2v ) 30(Cs) 34(D5h )

Fig. 1. Putative GM structures for KN clusters with N = 13–34 atoms. The size and approximate point group symmetry are provided below each cluster. Several isomers are
shown (for example, for N = 16) when they are nearly degenerate (total energy difference smaller than 0.01 eV).
138 A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143

35(C3v ) 37(Cs) 38(C3v ) 38(Cs) 39(C3 ) 40(C5v ) 42(C1)

44(C2v ) 46(Cs) 49(C1) 50(C1 ) 52(Cs) 52(Cs) 55(Ih )

Fig. 2. A representative selection of the putative GM structures for clusters with N = 35–55 atoms. The rest of the caption is the same as in Fig. 1.

an even number of atoms but a fully ab initio technique as in the to be icosahedral, so pIh packing still prevails in some parts of the
present work. We locate the same GM structure for N = 14 and cluster. K44 has a 9-atom core and a respectable C2v symmetry. An
18. For K20, our structure coincides with one of the nearly degener- analysis of the local environment of those inner atoms reveals that
ate structures of Banerjee. For K16, however, Banerjee et al. failed to the cluster contains a non-trivial combination of Ih13 and Z14 and
consider the Kasper polyhedron, so a complete comparison of the Z15 Kasper polyhedral units [74]. This type of structures persists
two calculation levels is not possible. up to N = 52, where a transition towards Mackay icosahedral struc-
Fig. 2 shows relevant structures in the size range N = 35–55. tures occurs which is maintained until the Mackay icosahedron is
Similarly to K34, K35 also consists of seven K13 icosahedral units, completed at N = 55. In the size range N = 44–52 the number of
resulting in another perfect core–shell arrangement, but the core core atoms steadily increases from nine to eleven, and the struc-
atoms adopt a different C3v symmetry. Poly-icosahedral packing tures become more and more prolate. The transition at N = 52 is
continues to dominate the structures up to N = 41, and most struc- thus accompanied by a significant change in the cluster shape to-
tures in this size range can indeed be obtained by removing atoms wards nearly spherical structures.
from a 45-atom perfect anti-Mackay icosahedron, a structure ob- The putative GM structures in the size range N = 56–80 are
tained by covering a 13-atom icosahedron with a complete anti- shown in Fig. 3. Immediately after the icosahedral shell closing at
Mackay overlayer. The only exception to this trend occurs for N = 55, potassium clusters avoid to place the first few additional
K39, which has the same 7-atom core as K35 and is obtained from atoms on a new, more external, atomic shell. For K56–K58, those
it by incorporating the four additional atoms into a crowded shell, atoms are admitted into the surface layer of K55 through the gen-
which lowers the global symmetry from C3v to C3. eration of a number of rosette-like defects [75], which consist of
We observe a structural transition at N = 42, a critical size be- 6-fold vertices that substitute the 5-fold vertices of the original ico-
yond which the clusters cannot anymore accomodate the strain sahedron. For K57 only, there is a nearly degenerate structure with
inherent to pIh structures. Nevertheless, the transition does not in- two adatoms occupying neighboring AM sites. K59 has two degen-
volve just a change from anti-Mackay to Mackay overlayers. K42, erate structures: theC6v isomer is a perfect core–shell structure
for example, has eight core atoms in a hexagonal bipyramid with 15 internal atoms. Its core is the Z14 Kasper polyhedron with
arrangement, and although the shell is quite amorphous, an D6d symmetry, but the growth of the shell follows a different pat-
approximate disclination line can be easily distinguished in tern (Mackay-like and anti-Mackay-like, respectively) on both
Fig. 2. The local environment of some of the core atoms continues sides of the disclination line, lowering the global symmetry to

56(C2v ) 57(Cs) 57(C2v ) 58(C2 ) 59(C3v ) 59(C6v ) 60(C1 )

61(C2v ) 61(C1) 64(C1 ) 66(C1 ) 67(C1) 71(C5) 73(Cs)

74(D3d ) 75(C2) 76(C1 ) 77(C2 ) 78(C1 ) 80(D3 )

Fig. 3. A representative selection of the putative GM structures for clusters with N = 56–80 atoms. The rest of the caption is the same as in Fig. 1.
A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143 139

C6v. The C3v structure contains 14 internal atoms and is obtained by detailed investigation is warranted for each individual material if
adding three adatoms on top of a C3v isomer of K56. Similarly, most accurate results are to be obtained.
clusters with N = 60–62 show adatoms on top of K56, which is thus The effect of van der Waals interactions can be assessed by
a more favorable seed for cluster growth than the Mackay comparing the results obtained at the vdW-DFT level of theory
icosahedron. The C2v isomer of K61 is, however, an exception to this (shown in Figs. 1–3) with those obtained under the LDA and GGA
rule. approximations (not shown explicitly). Qualitatively, the main ef-
A structural transition, similar to the one that occurs at N = 42, fect is the same as that reported for cesium clusters [55], namely
is observed at N = 63 and extends up to N = 66. The structures in a higher stabilization of more compact structures. For example,
this size range contain 17 or 18 internal atoms and a Z15 Kasper the GM structure of Cs15 is a Z14 Kasper polyhedron with one cen-
polyhedral unit at their cores. The 6-fold disclination lines can be tral atom [55], while Na15 adopts a less compact deltahedral struc-
clearly appreciated in Fig. 3. A new transition occurs at N = 67 ture without internal atoms. For K15, the deltahedral structure is
and persists up to N = 72. In fact, all the structures in this size range 0.1 eV more stable than the Kasper polyhedron at the LDA or
can be obtained by removing atoms from (or adding one atom to) GGA levels, but only 0.06 eV more stable at the vdW-DFT level. This
the perfect core–shell arrangement of K71. The packing in this clus- is a general observation for all sizes. In some cases, this extra sta-
ter is of an interesting mixed type: the core, with D5h symmetry, bilization is enough to invert the energetic order of isomers ob-
contains 19 atoms in a double icosahedral arrangement, so it tained at the LDA or GGA levels. One example is K16, whose
shows pIh packing; the shell tries to provide a Mackay-like over- structure coincides with Na16 if vdW effects are neglected but with
layer on top of the pIh core, which is not exactly possible without Cs16 when vdW effects are included. Considering that the bench-
generating a high strain in the bonds along the equatorial region. mark results in Table 1 demonstrate that the KBM functional pro-
This strain and the resulting geometrical frustration are relaxed vides a more accurate description of potassium properties, we
by twisting the two sides of the Mackay shell by opposite angles believe that the KBM predictions on cluster structures are also
about the 5-fold axis, which lowers the global symmetry to C5. more reliable. vdW effects are therefore not negligible for potas-
We would like to emphasize that the structure of K71 is different sium clusters, and should be included in calculations aimed at
from the one that would be obtained by interpenetrating two obtaining high quality results.
Ih55 units (a higher order version of pIh packing), as in that case
the 19-atom core would not show pIh packing.
Our calculations predict that this type of mixed packing domi- 3.2. Electronic properties
nates the structures of the remaining sizes up to N = 80. All these
clusters attempt to build a Mackay-like overlayer on top of a more In order to identify the main electronic shell closings, we show
compactly packed core. Perfect core–shell arrangements, which in Fig. 4 the energy gap between the highest occupied and lowest
can be considered geometrical shell closings, are obtained for unoccupied molecular orbitals (HOMO–LUMO gap), the vertical
K74, K77 and K80. The core in K74 contains 20 atoms and is based ionization potential (VIP) and the vertical electron affinity (VEA)
on two fused and distorted Ih13 units. The two K13 icosahedra share of potassium clusters. We observe a substantial odd–even alterna-
only six atoms, as opposed to seven in a double icosahedron, tion, mostly in the HOMO–LUMO gaps, which gradually dissap-
resulting in a slightly lower D3d symmetry. For this size in particu- pears as the cluster size increases. In fact, the HOMO–LUMO gaps
lar, the core symmetry is fully respected by the shell, which means of clusters with more than 70 atoms are already very small and
that the structure of K74 is just a bigger size version of the structure
of the K20 core, or in other words, that it consists of two Ih55 units,
interpenetrated in exactly the same way as the two Ih13 units of 0.5 14 20
18
εHOMO-LUMO (eV)

K20. The core of K77 has 21 atoms but otherwise is quite similar 0.4
to the core of K74. K78 and K79 have 22 atoms in a low-symmetry 30 34 40 58
0.3
core. Finally, the core of K80 contains 23 atoms and shows perfect
46 50
pIh packing of three Ih13 units. The symmetry of the core is there- 0.2 60 64
fore D3h. Similarly to the situation discussed for K71, the growth of 0.1
a Mackay-like overlayer involves some geometrical frustration that
0
lowers the global symmetry down to D3. Also, for the same reasons 3.5 18
stated for K71, this cluster is not obtained by interpenetrating three 3.4 14
Ih55 units (moreover, such a structure would contain 24 core atoms
3.3 20
VIP (eV)

instead of 23). 34 40 58
It is interesting to compare the structures of KN clusters with 3.2 26 30 46 50
those previously reported by us for clusters of other alkali metals 3.1 60 64
[27,29,55]. In those studies, sodium and cesium clusters were 3
found to adopt very different structures for most sizes, despite
both being alkali metals. Traditionally, the structures of potassium
and sodium clusters have been assumed to be akin to each other 1.6
VEA (eV)

(see, for example, [52]). This assumption is based on the similari- 1.4 60 64
46 50 58
ties observed in the abundance mass spectra, ionization potentials
and other properties of these two alkali clusters. A detailed com- 1.2 34 40
30
parison of our ab initio results confirms that KN structures coincide 24
1
with NaN structures much more frequently than with CsN struc- 14 20
0.8
tures. Exceptions occur only for N = 16, 39, 45, 48, 49, 52, 76, 78, 10 20 30 40 50 60 70 80
where the GM structure of KN coincides with that of CsN; also, for N
N = 37, 42–44, 58, 61, 63, 66, 67–69, 79, the GM structures here re-
Fig. 4. HOMO–LUMO gaps (upper graph), vertical ionization energies (middle
ported are specific to potassium, as they do not coincide with those graph) and vertical electron affinities (lower graph) of Kn clusters. The main
of sodium or cesium clusters. All in all, we conclude that, due to the electron shell closings (sizes preceeding a large drop in the HOMO–LUMO and/or
huge complexity of the potential energy landscape of clusters, a VIP values, or marked minima in the VEA values) are explicitly indicated.
140 A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143

the clusters seem to be approaching the gap closure characteristic while for c = p/3 we have an oblate axially symmetric shape (pan-
of a metallic state. The odd–even alternation of electronic proper- cake). For intermediate c values, cluster deformation is triaxial.
ties occurs on top of an average trend which coincides with the The overall trend for b is to decrease, indicating a tendency for
predictions of classical charged sphere models. bigger clusters to adopt more spherical shapes. On top of that aver-
Large drops in the VIP curve are associated with the opening of age trend, a series of oscillations between spherical and non-spher-
an electron shell. Electronically magic clusters will also have a ical shapes is observed in Fig. 5, which is in fair general agreement
large HOMO–LUMO gap and will display a local minimum in the with the predictions of jellium models although geometric shell
VEA curve. According to these criteria, the cluster sizes with a par- closings introduce some additional features. For example, clusters
ticularly stable electron shell structure are N = 14, 18, 20, 26, 30, with about 20, 40 and 58 electrons are more spherical than the
34, 40, 46, 50 and 58. Some of these numbers (N = 18 and 26, for rest, while clusters in between two spherical shell closings (for
example) do not show up as marked minima in the VEA curve, example, K13, K26 or K48) are the least spherical ones. However,
which instead shows a local minimum for K24. After N = 58 the K39 is slightly more spherical than K40 due to a geometric shell
curves are smoother although the stabilities of K60 and K64 are also closing. Similarly, the perfect icosahedron at N = 55 is more spher-
somewhat enhanced as compared to neighboring sizes. No feature ical than K58. We do not observe any enhanced sphericity at N = 68,
is detected at N = 68 or N = 70 even though these are electron shell 70 even if these sizes correspond to electron shell closings in some
closings according to some jellium models. jellium models [2,3]. These observations suggest that geometric
All the maxima observed in the experimental photoabsorption packing arguments dominate the structures of the biggest clusters
spectra [33,36,38], as well as the absence of specific features for here considered.
N = 68, 70, are reproduced by our calculations. We predict addi- The prolate/oblate character of the quadrupolar deformation is
tional maxima in the VIP (N = 24, 46, 50, 64, 74) which are not ob- also in general agreement with the structureless electron shell
served in the experiments but are predicted by the liquid drop models. In general terms, the structures tend to have prolate shape
model with shell corrections devised by Yannouleas et al. [39]. deformations immediately after an electron shell closing, and ob-
As explained by these authors, the weaker nature of these local late shapes when approaching an electron shell closing from below
maxima is expected to be detected only in experiments performed (see also [26]). Nevertheless, for clusters with more than 55 atoms
at very low temperatures. In this connection, we point out that, gi- the presence of geometric shell closings causes several exceptions
ven the estimated melting points of potassium clusters [76], the re- to such general rule. We conclude that, although the jellium model
ported experimental VIP values most probably refer to liquid-like continues to be a valuable guide for the cluster shapes even at zero
clusters. Kelvin, conclusions based on this model should not be overempha-
Fig. 5 shows the size dependence of the cluster shape, as quan- sized. An ab initio description, including the ionic structures explic-
tified by the Hill–Wheeler parameters [2,77]. b measures the de- itly, is required in order to obtain an accurate description of alkali
gree of quadrupolar deviation from sphericity, with a value b = 0 cluster structures.
indicating a perfectly spherical cluster; c determines the ‘‘sense’’ A resurgence of interest in the screening properties of small al-
of that deviation, i.e. whether the shape is prolate or oblate. c = 0 kali particles comes from recent beam deflection experiments per-
corresponds to an axially symmetric prolate aggregate (cigar-like), formed by Bowlan et al. [78]. In that work, the electric dipole
moments of small cold Nan clusters were measured and found to
be essentially zero (60.002 Debye/atom for n 6 20). This is a distin-
guishing feature of metallicity and apparently implies that the per-
0.4 fect screening properties of alkali metals are maintained down to
13 the nanoscale region. We have recently shown that an exhaustive
25-26 search for the GM structures employing ab initio methods can
0.3
47-48 approximately reproduce the experimental observations about
the electric dipole moments of sodium clusters [29], and predict
0.2 similar results for cesium clusters [55]. Nevertheless, a detailed
β

comparison of the results obtained at different levels of theory


16 45 [30], including several exchange–correlation functionals within
0.1 80 DFT and a variety of quantum-chemical methods, suggests that
66 the theoretical dipoles obtained from the most accurate methods
20 39-40 are significant up to an error of about 10 mDebye/atom. The results
0 55-56 of Fig. 6 show that Bowlan’s expectations can be also transferred to
Oblate
1

15
μ (mDebye/atom)

0.5
γ

10

0 5
Prolate

10 20 30 40 50 60 70 80
N
0
10 20 30 40 50 60 70 80
Fig. 5. Hill–Wheeler deformation parameters b (top panel) and c (bottom panel) of N
potassium clusters. The dashed red lines indicate those c values corresponding to
perfectly oblate, triaxial or oblate shapes. (For interpretation of the references to Fig. 6. Electron dipole moments of Kn clusters. In cases of near-degeneracy, the
color in this figure legend, the reader is referred to the web version of this article.) different isomers typically have very similar l values.
A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143 141

potassium clusters, confirming the generality of the result, at least and 55–60 atoms clearly stand on top of the average behavior as
for the family of alkali clusters. Nearly all the dipole moments of KN regards absolute stability. These ranges are in qualitative agree-
clusters are below 10 mDebye/atom, and our recent work [29] sug- ment with the main spherical electron shell closings.
gests that these values will be even smaller when zero-point vibra- The three indicators essentially predict the same set of magic
tional effects are taken into account. There are only three clusters numbers, which occur for N = 14, 18, 20, 26, 30, 34, 40, 46, 50,
(K26, K47 and K49) with a more sizable dipole moment. If the find- 55, 58, 71, 74, 77 and 80, with some secondary magic numbers also
ings of Bowlan et al. were completely general, it might imply that noticed in the figure. For clusters with up to 58 atoms, nearly all
we have not located the correct GM structures for these sizes. the magic numbers have an electronic origin, as they coincide with
Otherwise, they may just represent exceptions to the ‘‘zero dipole’’ the predictions of spheroidal Nilsson model [38] while their struc-
rule. tures (see Section 3.1) do not correspond in general to geometric
Finally, we mention here that most potassium clusters have the shell closings. An exception for small sizes is K34, a geometric shell
lowest spin multiplicity, i.e. a singlet for clusters with an even closing with high D5h symmetry, whose enhanced stability may
number of atoms and a doublet for clusters with an odd number thus have an important contribution coming from geometric
of atoms. In this sense KN clusters are also more similar to sodium packing.
than to cesium clusters [55]. Exceptions to this rule occur for some In the size range N P 55, we notice that K55 is a more prominent
sizes where geometric packing effects dominate the structure: K54 magic number than K58 based on the three different indicators
is a spin quintet, K55 and K71 are spin quartets, and K16 and K56 are shown in Fig. 7. The enhanced stability of K55 is clearly associated
spin triplets. to the icosahedral shell closing, as no specific feature is observed in
the electronic properties for this size (see Section 3.2). The same is
true of other prominent magic numbers observed in this size range,
3.3. Cluster stability i.e. N = 71, 74, 77, 80. In all these cases, the structures present a
perfect core–shell arrangement, with a complete Mackay overlayer
Fig. 7 shows three different stability measures: the cohesive en- covering a compact core, so they can be identified as geometric
ergy (or binding energy per atom) Ecoh ðNÞ ¼ E1  ENN , the evapora- shell closings. The experimental abundance spectra do not identify
tion energy Eevap(N) = (E1 + EN1)  EN, and the second energy significant features at these sizes, suggesting that the cluster beam
difference D2(N) = EN1 + EN+1  2EN = Eevap(N)  Eevap(N + 1). While is made up of liquid-like clusters and that the experiments are
the cohesive energy quantifies the total internal energy content of blind to the geometric magic numbers typical of cold, solid clus-
a cluster and is therefore a measure of its global (or absolute) sta- ters. These features are expected to be recovered in experiments
bility, Eevap and D2 provide ‘‘local’’ stability measures by comparing that determine the dissociation energies of solid clusters by
the energy of a cluster of size N to that of clusters with neighboring accounting for the latent heat of melting, as exemplified by exper-
sizes, and so they are more suitable quantities to interpret the re- iments on aluminum clusters [79–81]. In summary, we conclude
sults of abundance mass spectra. that electron shells become of secondary importance as compared
Cohesive energies are referenced to the results of a four param- to geometric shells in determining the relative stabilities of solid
eter, liquid drop like, smooth fit. This particular representation is potassium clusters with more than N  55–58 atoms.
convenient to highlight the oscillatory pattern of the cohesive It may be interesting to explicitly comment about the accuracy
energies about the average trend. Clusters with 18–20, 34–40 of the structures derived with the Gupta potential. It is well known
that this type of semiempirical potentials cannot be made fully
20 transferable to different atomic environments, and most of the
0.01 18 available parametrizations have been fitted to bulk properties. This
Ecoh-Efit (eV)

55 58 observation alone casts some reasonable doubts about the ex-


0.005 34 60
40 64 pected accuracy of the Gupta potential for clusters. But even more
14
7174
26 30 77 important, at least for small alkali clusters, is the observation that
0 46 50
electronic shell effects are so important that the cluster shapes fol-
-0.005 low very well the jellium model predictions (see Fig. 5). Obviously,
1 the electron shell effects typical of superatoms are not incorpo-
55 rated into the Gupta potential, which is an explicit function of
0.9 atomic coordinates only. Therefore, it is no surprise that the DFT-
18 34 38
Eevap (eV)

20 26 30 58 7780
0.8 14 44 46 50 64 7174 derived GM structures do not coincide with the Gupta-derived
40
GM structures, except for a few sizes such as N = 55, where the ico-
0.7
sahedral geometric shell closing is well described by the Gupta
0.6 model. Nevertheless, as we emphasized in Section 2, the Gupta po-
tential is a very useful external tool to generate a diversity of plau-
0.3
20 55 sible structural motifs, many of which would be overlooked if
14 34
0.2 using just some sort of intuition or educated guess. This is why
18 40
26 30 46
50 58 71 74 we employ structural indicators rather than just Gupta energies
Δ2 (eV)

0.1 64 77
in order to choose the structures that are re-optimized at the
0 DFT level. A more complete discussion about these issues is to be
-0.1 found in our previous work [27].
-0.2
10 20 30 40 50 60 70 80
N 4. Conclusions

Fig. 7. Stability measures for Kn clusters: the upper graph shows the cohesive We have reported the GM structures, electronic properties and
energies, referenced to Efit, which is a four-parameter fit of the form Efit = A0 + A1N1/
3
+ A2N2/3 + A3N; the middle graph shows the evaporation energies and the lower
stabilities of potassium clusters containing from 13 to 80 atoms,
graph shows the second-energy differences. Clusters with enhanced stability obtained employing a first-principles DFT method that incorpo-
(magic sizes) are shown for each of the stability indicators. rates a description of van der Waals interactions. KN clusters adopt
142 A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143

a variety of structural motifs in different size ranges: deltahedral [2] W.A. de Heer, The physics of simple metal clusters: experimental aspects and
simple models, Rev. Mod. Phys. 65 (1993) 611.
and poly-icosahedral packing dominates for N 6 41; Mackay icosa-
[3] M. Brack, The physics of simple metal clusters: self-consistent jellium model
hedra are the GM for N = 52–55; structures containing Kasper poly- and semiclassical approaches, Rev. Mod. Phys. 65 (1993) 677.
hedral units and disclination lines have the lowest energy for [4] V. Bonacić-Koutecký, P. Fantucci, J. Koutecký, Quantum chemistry of small
N = 16–18, 42–51, 63–66; for N = 56–62 the structures contain ro- clusters of elements of groups Ia, Ib, and IIa: fundamental concepts,
predictions, and interpretation of experiments, Chem. Rev. 91 (1991) 1035.
sette-like defects on the outermost atomic shell; finally, for N P 67 [5] U. Röthlisberger, W. Andreoni, Structural and electronic properties of sodium
we observe a mixed packing pattern in which a distorted Mackay- microclusters (n = 2–20) at low and high temperatures: new insights from ab
like overlayer is grown around a poly-icosahedral core. Our results initio molecular dynamics studies, J. Chem. Phys. 94 (1991) 8129.
[6] U. Röthlisberger, W. Andreoni, P. Giannozzi, Thirteen-atom clusters:
improve over previous theoretical determinations of global min- equilibrium geometries, structural transformations, and trends in Na, Mg, Al,
ima for several sizes with N 6 20. A comparison of the structures and Si, J. Chem. Phys. 96 (1992) 1248.
of potassium clusters with those previously obtained for sodium [7] F. Spiegelman, R. Poteau, B. Montag, P.-G. Reinhard, Global structure of small
Na clusters in different approaches, Phys. Lett. A 242 (1998) 163.
and cesium clusters demonstrates that KN clusters share more sim- [8] L. Kronik, I. Vasiliev, M. Jain, J.R. Chelikowsky, Ab initio structures and
ilarities with NaN than with CsN clusters on average, but at the polarizabilities of sodium clusters, J. Chem. Phys. 115 (2001) 4322.
same time there are numerous exceptions to this simple rule [9] M. Schmidt, H. Haberland, Phase transitions in clusters, C.R. Phys. 3 (2002) 327.
[10] I.A. Solov’yov, A.V. Solov’yov, W. Greiner, Structure and properties of small
which justify a separate and dedicated study of potassium clusters. sodium clusters, Phys. Rev. A 65 (2002) 053203.
The superior accuracy of the vdW-DFT functional as compared to [11] M. Moseler, B. Huber, H. Häkkinen, U. Landman, G. Wrigge, M.A. Hoffmann, B.
usual LDA and GGA approximations has been demonstrated von Issendorff, Thermal effects in the photoelectron spectra of Na N clusters
(N = 4–19), Phys. Rev. B 68 (2003) 165413.
through benchmark calculations on the dimer and bulk. Inclusion
[12] K.R.S. Chandrakumar, T.K. Ghanty, S.K. Ghosh, Static dipole polarizability and
of vdW interactions has been found to modify the GM structure binding energy of sodium clusters Nan (n = 1–10): a critical assessment of all
of several KN clusters. Therefore, we have concluded that vdW electron based post Hartree–Fock and density functional methods, J. Chem.
interactions should be included in accurate studies of potassium Phys. 120 (2004) 6487.
[13] A. Aguado, Competing thermal activation mechanisms in the melting-like
clusters. transition of NaN (N = 135–147) clusters, J. Phys. Chem. B 109 (2005) 13043.
The identified electronic shell closings are in very good agree- [14] M.S. Lee, S. Chacko, D.G. Kanhere, First-principles investigation of finite-
ment with the experimentally determined photoionization spectra. temperature behavior in small sodium clusters, J. Chem. Phys. 123 (2005)
164310.
For clusters with more than 58 atoms, the size dependence of the [15] F. Baletto, R. Ferrando, Structural properties of nanoclusters: energetic,
electronic properties becomes substantially smoother and feature- thermodynamic and kinetic effects, Rev. Mod. Phys. 77 (2005) 371.
less. The cluster shapes are also in good agreement with those pre- [16] A. Aguado, J.M. López, Small sodium clusters that melt gradually: melting
mechanisms in Na30, Phys. Rev. B 74 (2006) 115403.
dicted by spheroidal jellium models, although the detailed atomic [17] M.S. Lee, D.G. Kanhere, Effects of geometric and electronic structure on the
structure (which is neglected in jellium models) is found to intro- finite temperature behavior of Na58, Na57, and Na55 clusters, Phys. Rev. B 75
duce non-negligible modifications on top of the average jellium (2007) 125427.
[18] S. Zorriasatein, M.S. Lee, D.G. Kanhere, Electronic structures, equilibrium
trends. With few exceptions, KN clusters are found to adopt the geometries, and finite-temperature properties of Nan (n = 39–55) from first
lowest possible spin multiplicity. They also possess negligible elec- principles, Phys. Rev. B 76 (2007) 165414.
tric dipole moments, as was also found for sodium and cesium [19] E.G. Noya, J.P.K. Doye, D.J. Wales, A. Aguado, Geometric magic numbers of
sodium clusters: interpretation of the melting behaviour, Eur. Phys. J.D 43
clusters. Metallic-like screening seems thus to be a rather general
(2007) 57.
property of small alkali clusters. [20] O. Kostko, B. Huber, M. Moseler, B. von Issendorff, Structure determination of
We have also reported the relative stabilities and magic num- medium-sized sodium clusters, Phys. Rev. Lett. 98 (2007) 043401.
bers of KN clusters. The theoretical results reproduce all the magic [21] S.M. Ghazi, M.S. Lee, D.G. Kanhere, The effects of electronic structure and
charged state on thermodynamic properties: An ab initio molecular dynamics
numbers observed in experimental abundance spectra and which investigation on neutral and charged clusters of Na39, Na40 and Na41, J. Chem.
coincide with electronic shell closings, but additionally predict Phys. 128 (2008) 104701.
new magic numbers which are associated with geometric shell [22] S. Núñez, J.M. López, A. Aguado, Atomic layering and related postmelting
effects in small liquid metal clusters, Phys. Rev. B 79 (2009) 165429.
closings. The electron shell structure is found to dominate the sta- [23] C. Hock, C. Bartels, S. Strassburg, M. Schmidt, H. Haberland, B. von Issendorff, A.
bilities of clusters with less than about 55–58 atoms, while geo- Aguado, Premelting and postmelting in clusters, Phys. Rev. Lett. 102 (2009)
metric packing effects explain the magic numbers for bigger 043401.
[24] M. Itoh, V. Kumar, T. Adschiri, Y. Kawazoe, Comprehensive study of sodium,
clusters. Theoretical calculations thus become an ideal comple- copper, and silver clusters over a wide range of sizes 2 6 N 6 75, J. Chem. Phys.
ment to the experimental mass spectra which typically miss the 131 (2009) 174510.
geometric shell closings if the cluster beam contains hot, liquid- [25] S.M. Ghazi, S. Zorriasatein, D.G. Kanhere, Building clusters atom-by-atom:
from local order to global order, J. Phys. Chem. A 113 (2009) 2659.
like, clusters. [26] B. Huber, M. Moseler, O. Kostko, B. von Issendorff, Structural evolution of the
sodium cluster anions Na 
20  Na57 , Phys. Rev. B 80 (2009) 235425.

Acknowledgements [27] A. Aguado, O. Kostko, First-principles determination of the structure of NaN


and NaN clusters with up to 80 atoms, J. Chem. Phys. 134 (2011) 164304.
[28] A. Aguado, Stepwise melting in Naþ 41 : a first-principles critical analysis of
I gratefully acknowledge the support of the Spanish ‘‘Ministerio available experimental results, J. Phys. Chem. C 115 (2011) 13180.
de Ciencia e Innovación’’, the European Regional Development [29] A. Aguado, A. Vega, L.C. Balbás, Structural and zero-point vibrational effects on
the electric dipole moments and static dipole polarizabilities of sodium
Fund and ‘‘Junta de Castilla y León’’ (Project Nos. FIS2011-22957 clusters, Phys. Rev. B 84 (2011) 165450.
and VA104A11-2). [30] A. Aguado, A. Largo, A. Vega, L.C. Balbás, On the electric dipole moments of
small sodium clusters from different theoretical approaches, Chem. Phys. 399
(2012) 252.
Appendix A. Supplementary material [31] A. Aguado, M.F. Jarrold, Melting and freezing of metal clusters, Annu. Rev.
Phys. Chem. 62 (2011) 151.
[32] W.D. Knight, W.A. de Heer, K. Clemenger, W.A. Saunders, Electronic shell
Supplementary data associated with this article can be found, in structure in potassium clusters, Solid State. Commun. 53 (1985) 445.
the online version, at http://dx.doi.org/10.1016/j.comptc.2013. [33] W.A. Saunders, K. Clemenger, W.A. de Heer, W.D. Knight, Photoionization and
06.040. shell structure of potassium clusters, Phys. Rev. B 32 (1985) 1366.
[34] M.M. Kappes, M. Schär, P. Radi, E. Schumacher, On the manifestation of
electronic structure effects in metal clusters, J. Chem. Phys. 84 (1986) 1863.
References [35] W.D. Knight, W.A. de Heer, W.A. Saunders, K. Clemenger, M.Y. Chou, M.L.
Cohen, Alkali metal clusters and the jellium model, Chem. Phys. Lett. 134
(1987) 1.
[1] W.D. Knight, K. Clemenger, W.A. de Heer, W.A. Saunders, M.Y. Chou, M.L.
[36] M.L. Cohen, M.Y. Chou, W.D. Knight, W.A. de Heer, Physics of metal clusters, J.
Cohen, Electronic shell structure and abundances of sodium clusters, Phys. Rev.
Phys. Chem. 91 (1987) 3141.
Lett. 52 (1984) 2141.
A. Aguado / Computational and Theoretical Chemistry 1021 (2013) 135–143 143

[37] C. Bréchignac, P. Cahuzac, J.P. Roux, Photoionization of potassium clusters: [60] D.M. Ceperley, B.J. Alder, Ground state of the electron gas by a stochastic
neutral and ionic cluster stabilities, J. Chem. Phys. 87 (1987) 229. method, Phys. Rev. Lett. 45 (1980) 566.
[38] W.D. Knight, W.A. de Heer, W.A. Saunders, Shell structure and response [61] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
properties of metal clusters, Z. Phys. D 3 (1986) 109. simple, Phys. Rev. Lett. 77 (1996) 3865.
[39] C. Yannouleas, U. Landman, Electronic entropy, shell structure, and size- [62] J. Klimes, D.R. Bowler, A. Michaelides, Chemical accuracy for the van der Waals
evolutionary patterns of metal clusters, Phys. Rev. Lett. 78 (1997) 1424. density functional, J. Phys.: Condens. Matter 22 (2010) 022201.
[40] L.A. der Lan, P. Bartl, C. Leidlmair, H. Schöblel, S. Denifl, T.D. Märk, A.M. Ellis, P. [63] G. Román-Pérez, J.M. Soler, Efficient Implementation of a van der Waals
Scheier, Submersion of potassium clusters in helium nanodroplets, Phys. Rev. density functional: application to double-wall carbon nanotubes, Phys. Rev.
B 85 (2012) 115414. Lett. 103 (2009) 096102.
[41] M.Y. Chou, A. Cleland, M.L. Cohen, Total energies, abundances, and electronic [64] G.I. Csonka, J.P. Perdew, A. Ruzsinszky, P.H.T. Philipsen, S. Lebègue, J. Paier, O.A.
shell structure of lithium, sodium, and potassium clusters, Solid State. Vydrov, J.G. Ángyán, Assessing the performance of recent density functionals
Commun. 52 (1984) 645. for bulk solids, Phys. Rev. B 79 (2009) 155107.
[42] S. Saito, M.L. Cohen, Jellium-model calculation for dimer decays of potassium [65] R. Hamann, M. Schlüter, C. Chiang, Norm-conserving pseudopotentials, Phys.
clusters, Phys. Rev. B 38 (1988) 1123. Rev. Lett. 43 (1979) 1494.
[43] M. Manninen, Structures of small alkali-metal clusters, Phys. Rev. B 34 (1986) [66] S.G. Louie, S. Froyen, M.L. Cohen, Nonlinear ionic pseudopotentials in spin-
6886. density-functional calculations, Phys. Rev. B 26 (1982) 1738.
[44] C. Bréchignac, P. Cahuzac, J.P. Roux, D. Pavolini, F. Spiegelmann, Adiabatic [67] J. Tao, J.P. Perdew, A. Ruzsinszky, Long-range van der Waals attraction and
decomposition of mass-selected alkali clusters, J. Chem. Phys. 87 (1987) 5694. alkali-metal lattice constants, Phys. Rev. B 81 (2010) 233102.
[45] F. Spiegelmann, D. Pavolini, Ab initio calculations of the electronic structure of [68] J. Klimes, D.R. Bowler, A. Michaelides, Van der Waals density functionals
small Nan, Naþ þ
n , Kn, and Kn clusters (n 6 6) including core-valence interaction, applied to solids, Phys. Rev. B 83 (2011) 195131.
J. Chem. Phys. 89 (1988) 4954. [69] D.J. Wales, J.P.K. Doye, Global optimization by basin-hopping and the lowest
[46] A.K. Ray, S.D. Altekar, Ab initio many-body perturbation-theoretic study of energy structures of Lennard-Jones Clusters containing up to 110 atoms, J.
small potassium clusters, Phys. Rev. B 42 (1990) 1444. Phys. Chem. A 101 (1997) 5111.
[47] A. Kornath, R. Ludwig, A. Zoermer, Small potassium clusters, Angew. Chem. Int. [70] R.P. Gupta, Lattice relaxation at a metal surface, Phys. Rev. B 23 (1981) 6265.
Ed. 37 (1998) 1575. [71] V. Rosato, M. Guillopé, B. Legrand, Thermodynamical and structural properties
[48] I.S. Lim, P. Schwerdtfeger, T. Söhnel, H. Stoll, Ground state properties and static of f.c.c transition metals using a simple tight-binding model, Philos. Mag. A 59
dipole polarizabilities of the alkali dimers from Kn2 to Frn2 ðn ¼ 0; þ1Þ from scalar (1988) 321.
relativistic pseudopotential coupled cluster and density functional studies, J. [72] F. Cleri, V. Rosato, Tight-binding potentials for transition metals and alloys,
Chem. Phys. 122 (2005) 134307. Phys. Rev. B 48 (1993) 22.
[49] E. Flrez, P. Fuentealba, A theoretical study of alkali metal atomic clusters: from [73] Y. Li, E. Blaisten-Barojas, D.A. Papaconstantopoulos, Structure and dynamics of
Lin to Csn (n = 2–8), Int. J. Quantum Chem. 109 (2009) 1080. alkali-metal clusters and fission of highly charged clusters, Phys. Rev. B 57
[50] L. Padilla-Campos, E. Chávez, Electronic properties of small Kn (n 6 8) and (1998) 15519.
bimetallic KnCum (n, m 6 4) clusters, J. Mol. Struct. THEOCHEM 958 (2010) 92. [74] D.J. Wales, Energy Landscapes, Cambridge University Press, Cambridge, 2003.
[51] J. Centeno, P. Fuentealba, Big bang methodology applied to atomic clusters, Int. [75] E. Aprà, F. Baletto, R. Ferrando, A. Fortunelli, Amorphization mechanism of
J. Quantum Chem. 111 (2011) 1419. icosahedral metal nanoclusters, Phys. Rev. Lett. 93 (2004) 065502.
[52] A. Banerjee, T.K. Ghanty, A. Chakrabarti, Ab initio studies and properties of [76] A. Aguado, Orbital-free molecular dynamics study of melting in K20, K55, K92,
small potassium clusters, J. Phys. Chem. A 112 (2008) 12303. K142, Rb55, and Cs55 clusters, Phys. Rev. B 63 (2001) 115404.
[53] S. Abdalla, M. Sprinborg, Y. Dong, Isolated and deposited potassium clusters: [77] D.L. Hill, J.A. Wheeler, Nuclear constitution and the interpretation of fission
energetic and structural properties, Surf. Sci. 608 (2013) 255–264. phenomena, Phys. Rev. 89 (1953) 1102.
[54] M. Dion, H. Rydberg, E. Schröder, D.C. Langreth, B.I. Lundqvist, Van der Waals [78] J. Bowlan, A. Liang, W.A. de Heer, How metallic are small sodium clusters?,
density functional for general geometries, Phys. Rev. Lett. 92 (2004) 246401. Phys Rev. Lett. 106 (2011) 043401.
[55] A. Aguado, Discovery of magnetic superatoms and assessment of van der [79] A.K. Starace, C.M. Neal, B.P. Cao, M.F. Jarrold, A. Aguado, J.M. López, Correlation
Waals dispersion effects in Csn clusters, J. Phys. Chem. C 116 (2012) 6841. between the latent heats and cohesive energies of metal clusters, J. Chem.
[56] W. Kohn, L.J. Sham, Self-consistent equations including exchange and Phys. 129 (2008) 144702.
correlation effects, Phys. Rev. 140 (1965) A1133. [80] A.K. Starace, C.M. Neal, B.P. Cao, M.F. Jarrold, A. Aguado, J.M. López, Electronic
[57] P. Hohenberg, W. Kohn, Inhomogeneous electron gas, Phys. Rev. 136 (1964) effects on melting: comparison of aluminum cluster anions and cations, J.
B864. Chem. Phys. 131 (2009) 044307.
[58] J.M. Soler, E. Artacho, J.D. Gale, A. García, J. Junquera, P. Ordejón, D. Sánchez- [81] B.P. Cao, A.K. Starace, O.H. Judd, I. Bhattacharyya, M.F. Jarrold, J.M. López, A.
Portal, The SIESTA method for ab initio order-N materials simulation, J. Phys.: Aguado, Activation of dinitrogen by solid and liquid aluminum nanoclusters: a
Condens. Matter 14 (2002) 2745. combined experimental and theoretical study, J. Am. Chem. Soc. 132 (2010)
[59] J.P. Perdew, A. Zunger, Self-interaction correction to density-functional 12906.
approximations for many-electron systems, Phys. Rev. B 23 (1981) 5048.

You might also like