Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Surface Science 608 (2013) 255–264

Contents lists available at SciVerse ScienceDirect

Surface Science
journal homepage: www.elsevier.com/locate/susc

Isolated and deposited potassium clusters: Energetic and structural properties


Sahar Abdalla ⁎, Michael Springborg, Yi Dong
Physical and Theoretical Chemistry, University of Saarland, 66123 Saarbrücken, Germany

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

Article history: With the ultimate purpose of studying the properties of potassium clusters deposited on a surface, we have at
Received 26 July 2012 first studied the energetic and structural properties of isolated potassium clusters with up to 20 atoms. The global
Accepted 23 October 2012 total-energy-minima structures have been determined by using the Density Functional Tight Binding method
Available online 13 November 2012
(DFTB) combined with genetic algorithms. For the isolated clusters in the gas phase we analyze the binding
energy as well as the stability function. Moreover, structural similarity is studied using so-called similarity func-
Keywords:
Clusters
tions. Also the overall shape of the clusters and the radial distribution of the atoms are studied. Subsequently, we
Surface study the changes in the structure when these clusters are deposited on one out of two different potassium
Deposition surfaces. Finally, the energy related to the deposition is studied in detail.
Stability © 2012 Elsevier B.V. All rights reserved.

1. Introduction of electronic shell structures [1], similar to what has been reported for
sodium clusters, with peaks or steps for those clusters which contain
Whereas many theoretical studies on the properties of metal clusters N= 2, 8, 20, 40, ⋯ atoms [2,3]. Early, this prediction was confirmed
in gas phase exist, much less studies have been performed for clusters experimentally through mass spectra from potassium cluster beams
that are deposited on some surface, although the properties of the clus- [4]. Also later, potassium clusters in gas phase have been of considerable
ters on surfaces are relevant, both because such systems provide one experimental interest [5–12].
way of studying them experimentally and because of their importance From the theoretical point of view, potassium clusters have been in-
for various chemical processes, including catalysis. Furthermore, the vestigated with different theoretical methods, although most of these
response of the clusters to being deposited on the surface can also studies were limited to small sizes [13–21]. However, although potassi-
provide relevant information on the clusters themselves. um clusters have been the subject of several theoretical studies, it is sur-
The deposition of clusters on a surface influences both the energetic prising that there has been no attempt to identify the structures of the
and the structural properties of the clusters so that the deposited clus- global total-energy-minima using unbiased structure–optimization
ters have properties different from those obtained in gas phase. methods in combination with electronic-structure calculations.
Since in experiment often clusters are studied that are interacting On the other hand, potassium clusters with up to 60 atoms were
with some surfaces, the ultimate goal of the present work is to study investigated using a Gupta potential for describing the interatomic
the effects of depositing potassium clusters on a potassium surface. interactions combined with both genetic and basin hopping algo-
We are particularly interested in the changes in structures and energies rithms [22] to search for the global minima. Such a model does not
due to the absorption. To this end we, thus, need both the properties of include an explicit description of electronic effects, and may, in
the isolated clusters in the gas phase and those of the clusters on the addition, tend to overestimate the role of packing effects. Therefore
surfaces. Therefore, the first part of this work presents results on the the particularly stable clusters are those that are particularly closed
global total-energy-minima structures of isolated potassium clusters. packed and, hence, not in agreement with those obtained with the
In the second part, these optimized structures were deposited on two spherical jellium model and, it turns out, experiment [1,4]. This
different potassium surfaces, i.e., K(100) and K(110) surfaces. suggests that electronic effects are important for the properties of
Over the years, several studies on the properties of potassium potassium clusters.
clusters in the gas phase have been presented. One of the simplest Accordingly, in the present study, we shall include electronic de-
theoretical approaches treats the isolated potassium clusters within grees of freedom explicitly for clusters with up to 20 atoms by utilizing
a spherical jellium model. With this model, discontinuities in the the Density Functional Tight Binding (DFTB) method. To search for the
total energy as a function of cluster size are related to the existence global minima structures we use genetic algorithms. As we shall see,
the sizes of the particularly stable clusters are found to be in agreement
with those obtained experimentally. In order to obtain further informa-
⁎ Corresponding author.
E-mail addresses: s.abdalla@mx.uni-saarland.de (S. Abdalla),
tion on the properties of the clusters, we shall use various descriptors
m.springborg@mx.uni-saarland.de (M. Springborg), y.dong@mx.uni-saarland.de for analyzing both energetic and structural properties as function of
(Y. Dong). the cluster size.

0039-6028/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.susc.2012.10.016
256 S. Abdalla et al. / Surface Science 608 (2013) 255–264

To include the effect of surface on the properties of potassium approach. The optimized lattice constant was found to be a = 5.29 Å
clusters, we have considered the clusters when being deposited which compares well with the experimental value of a = 5.33 Å.
very softly on K(100) and K(110) surfaces. To our knowledge, this is Even if our approach is constructed as providing accurate results for
the first study devoted to potassium clusters on a potassium surface. the K 2 molecule and is found to be accurate for the crystal, too, it is not
Thereby, we shall focus on the changes in the structures due to the obvious that accurate results will be found for intermediate-sized sys-
adsorption as well as in the energetics related to the adsorption. We tems like the clusters of the present study. However, as we shall discuss
emphasize that we will make use of some approximations that further below, we find a good agreement in the overall structures
ultimately could affect our results. Since, however, our goal is to when comparing our results for the potassium clusters with those of
obtain general trends, we are convinced that these will given correct other studies for isolated clusters [13–21]. Moreover, the structure we
by our results, although details could change with the use of more find for the K 6 cluster has a high symmetry with only two bond lengths.
accurate (and, hence, computationally much more demanding) The same structure was also found by Spiegelmann and Pavolini [15] in
approaches. their ab initio study and by Banerjee et al. [21] in their parameter-free
This paper is organized as follows: In Section 2 our theoretical density-functional calculations. Our calculations give 4.48 and 4.41 Å,
method is briefly described. Subsequently, in Section 3, we present respectively, for the two bond lengths, whereas Spiegelmann and
and discuss our results for both isolated and deposited clusters. Finally, Pavolini found them to be identical and equal to 4.50 Å, and Banerjee
our results are summarized in Section 4. et al. reported values of 4.47 and 4.31 Å. Thus, our values are seen to
match very well those of more accurate studies, giving further support
2. Theoretical methods for our approach.
To determine the global minima structures we have used genetic
Our theoretical method combines the Density Functional Tight algorithms [28–32], which provide an efficient tool for global geometry
Binding (DFTB) method for the calculation of the total energy for a optimizations. Our computer code that combines the DFTB and the
given structure with genetic algorithms for the determination of the genetic algorithm is home-made and has previously been found to
structure of the global total-energy minimum. The DFTB method is provide accurate results for monatomic (Au, Al, Na, Ge, Si), diatomic
based on the density functional theory of Hohenberg and Kohn [23] (SiGe, AlO), and triatomic (HAlO) clusters [33–39].
in the formulation of Kohn and Sham [24] and has been developed At this place it is relevant to discuss briefly some of the consequences
by Seifert and coworkers [25–27]. Within the DFTB, the total energy of our approximations. At first, no theoretical method exists that with
of a given system relative to that of the isolated, non-interacting absolute certainty can provide the structure of the global total-energy
atoms is given as minimum for systems with more than just some few atoms, even if
assuming that the approach for calculating the total energy of a given
X
occ  
1   structure is accurate. Thus, also our version of the genetic algorithms
Etot ¼ εi −∑ ∑ εjm þ ∑ U jj′ Rj −Rj ′  ð1Þ
i j m 2 j≠j′ may in certain cases predict structures that are not those of the global
total-energy minima. This issue has been discussed further by Roberts
where εi are the single particle energies of the system of interest, εjm are et al. [31]. Our experience is, however, that in most cases our approach
those of the isolated atoms (i.e., jth eigenvalue of the mth atom), and the is reliable. Second, different approximate descriptions of the total energy
last term is a pair potential, which describes short-range interactions. In as a function of structure may lead to different structures for the global
the DFTB, only the valance electrons are considered, whereas the other total-energy minimum implying that the true global total-energy-
electrons are treated as a frozen core. minimum structure in some cases has not been identified. This issue
The orbital energies are calculated by expanding the orbital has been discussed in detail by Hellmann et al. [40] for Si 13 and Si 19
wavefunctions (Kohn–Sham orbitals ψi) in a set of atom-centered clusters. Therefore, a comparison with any other available results is
basis functions, {χjm}, that in practical calculations are written as linear very important and will in our case, as we shall discuss below, demon-
combinations of larger number of Slater-type orbitals. For {χjm} m strate that our approach gives results that agree with those of other
marks the atom and j distinguishes between different functions studies. Third, when more structures have total energies close to that
centered at the same atom. Accordingly, the Kohn–Sham orbitals are of the global total-energy minimum, experimental studies may provide
written as results for a statistical mixture of those. Our genetic algorithms are, in
principle, able to provide estimates for as many isomers as members of
ψi ðr Þ ¼ Σj C ij χ jm ðr Þ ð2Þ one generation. In many cases, however, two or more of those structures
are identical. At the costs of significantly increased computational efforts
Moreover, we assume that the potential experienced by the electrons we could, in principle, identify different structures and, subsequently,
can be written as a superposition of those of the isolated atoms, provide statistically averaged results for the (then temperature-
  dependent) properties. Here, we have, however, decided to not follow
→ → → a such approach.
V r ¼ ∑ V m r −R m : ð3Þ
m The main idea behind our version of the genetic algorithms is to
generate a larger number of structures (parent structures) and from
Finally, we assume that bχ j1 m1 jV m jχ j2 m2 > vanishes unless m1 = m those to generate a new set through cutting and pasting. From the
and/or m2 = m. Then, all information on the orbital energies can be total set of new and old structures, those with the lowest energy are
extracted from accurate (parameter-free density-functional) calcula- kept (which then form the next so-called generation). In the present
tions on two-atomic systems, i.e., in our case on the K 2 molecule, as a study, this process is repeated 8000 times for each cluster size until
function of interatomic distance. It is worth to mention that we have the lowest energy is unchanged for 4000 generations. Finally, we
used B3LYP-functional in our accurate calculations for the K 2 molecule. used here our approach for K N clusters for N up to 21, but shall
The last (repulsive) term in Eq. (1) is determined so that the total here concentrate on the properties of the clusters with N up to 20.
energy as a function of interatomic distance for the diatomic as obtained For the calculations of the potassium clusters deposited on a
from the parameter-free density-functional calculations is reproduced surface we considered two types of potassium surfaces, i.e., the
accurately. Our approach is, accordingly, based on parametrizing results (110) and (100) surface. We modeled the surface through a film
from the diatomic K 2 molecule and, subsequently, using those for larger with a thickness of three layers (i.e., 3a) and with each layer
K-based systems. In order to check whether this approach is reasonable, containing 72 atoms. During the surface optimization, the structure
we calculated the lattice constant for crystalline K using the present of the two bottom layers was kept fixed at the experimental crystal
S. Abdalla et al. / Surface Science 608 (2013) 255–264 257

where Etot is the total energy of the complete system (cluster+surface),


EIS is that of the isolated surface, and EIC is that of the isolated cluster. ED
can be considered as consisting of three contributions, i.e., an energy
related to changing the structure of the surface from that of the isolated
system to that of the cluster+surface system, a similar energy for the
cluster, and an interaction energy between the two. Whereas the first
two per construction are positive (see, however, later), the latter should
be negative so that ED becomes negative.
Hence, the interaction energy between the surface and the cluster
can be defined as

Eint ¼ Etot −EDS −EDC : ð5Þ

Here, EDS is the energy of the surface with the structure it has
after the deposition, and EDC is, equivalently, the energy of the cluster
with the structure after the deposition. The energy costs related to
restructuring the surface and the cluster, respectively, are then given as

ERC ¼ EDC −EIC ð6Þ

and

ERS ¼ EDS −EIS : ð7Þ

Fig. 1. The upper panel shows the binding energy per atom for potassium clusters with 3. Results and discussion
up to 20 as function of the size of the clusters, whereas the lower panel shows the
stability function for the same clusters.
3.1. Isolated potassium clusters

structure, whereas the first layer was allowed to relax. Subsequently, At first, we shall discuss the properties of isolated K N clusters with
a potassium cluster with the structure that was optimized in the gas N up to 20. We shall discuss their energetic and structural properties
phase was placed in the closest vicinity to the surface, and the and also compare with those of Na N clusters, as the two types of
resulting system was allowed to relax. systems often are considered as being very similar.
We emphasize that the goal of our work is obtain general trends
about what can happen when a cluster is deposited on a surface. Thus, 3.1.1. Energetic properties
we have not attempted to identify the optimal orientation and/or posi- The binding energy per atom, i.e., − Etot(N)/N with Etot(N) being
tion of the cluster on the surface. Instead, the fact that we considered the total energy of Eq. (1) for the cluster with N atoms, is an overall
just a single deposition geometry will for sure influence our results. increasing function of N, as can be seen in the upper part of Fig. 1.
Moreover, by allowing only the top-most layer to relax, we make use However, as always found for such small systems, additional size-
of the near-sightedness of electronic interactions, although further dependent features are seen in the figure, suggesting that certain
relaxation effects may exist. Finally, as discussed, e.g., by Roberts et al. cluster sizes are more stable than the neighboring sizes.
[31], there is no guarantee that even the unbiased structure–optimization The stability of the clusters can at best be quantified through the
method indeed identifies the structure of the global-total-energy mini- stability function,
mum. Nevertheless, we believe that our study provides general informa-
tion on the effects on structural and energetic properties when a cluster is SðNÞ ¼ Etot ðN þ 1Þ þ Etot ðN−1Þ−2Etot ðNÞ: ð8Þ
deposited on a surface.
In order to analyze the changes due to the deposition of the cluster This function has maxima for particularly stable clusters and is
on the energy we at first define a deposition energy as shown in the lower panel in Fig. 1. At first, a clear even–odd oscillatory
pattern is identified, which can be related to electronic shell-filling
effects, but, in addition, more pronounced peaks are found for N = 8,
ED ¼ Etot −EIS −EIC ð4Þ 18, and 20, similar to our earlier results for Na N clusters using the

Table 1
Comparison of the structures for K N for N ≤ 8.

KN Structure Point group References

K3 Isosceles triangle C2v Present, Refs. [12,15,17,19]


Linear D∞h Ref. [20]
K4 Rhombus D2h Present, Refs. [15,17,21,41]
K5 Trapezoidal C2v Present, Refs. [15,20]
Trigonal pyramidal D3h Ref. [16]
Deformed tetrahedral Ref. [19]
K6 Pentagonal pyramidal C5v Present, Refs. [15,21]
Planar C2v Ref. [20]
Octahedral Ref. [20]
K7 Pentagonal bipyramid D5h Present, Ref. [20]
K8 Compact structure C2v Present, Ref. [20]
258 S. Abdalla et al. / Surface Science 608 (2013) 255–264

a) b) c) d) e)

f) g) h) i) j)

k) l) m) n) o)

p) q) r)
Fig. 2. The structures of the global total-energy minima of K N (2 ≤ N ≤ 20) clusters obtained with DFTB. Below each structure, both N and the corresponding point group are given.

same theoretical approach [35,36]. That these cluster sizes should In order to compare our results with those of other, previous studies,
be particularly stable is in agreement with what is obtained from we tabulate the structures of K N for N ≤8 and their corresponding point
the simplest spherical jellium model as well as with mass spectra of groups in Table 1. There we have also listed the results of previous
potassium cluster beams [1,4]. theoretical studies [19,17,16,15,12,20,41,21]. It is seen that our results
According to the spherical jellium model both systems (i.e., K N and are in good agreement with what has been found in most other studies.
Na N) can be considered as roughly spherical clusters. However, as we There are also deviations, although in no case our results differ from
shall see below, this is not the case and, indeed, K N and Na N clusters those of all other theoretical studies. We interpret this as being in
are less similar than what could be suggested based on their energetic support of our theoretical approach.
properties. For the triangular structure for K 3 we find indications of a weak
Jahn–Teller distortion. For K 4 we obtained a rhombus structure with
D2h symmetry. The rhombus sides have all the same length, a, which
3.1.2. Structural properties is longer than that of the diagonal, b (see Fig. 3). We find a = 4.45 Å
The global minima structures of the potassium cluster with up to 20 and b being 4.16 Å. A similar difference has also been reported in
atoms are shown in 2. We see that up to N= 5, the structures are planar. some earlier studies [21,17,41], whereas others studies have found
In the figure, also the point group symmetries of the individual clusters a = b [14,15,19].
are given. Comparing those with the similar results for Na N clusters, The transition from 2D to 3D occur at K 6, which is predicted to have
about half-part of the clusters have the same symmetries for the two pentagonal pyramidal structure with C5v symmetry in agreement with
systems, although this similarity in most cases results from a particularly the result obtained by Banerjee et al. [21] and in contrast with that
low symmetry (C1 or Cs). obtained from Florez et al. [20] and Richtsmeier et al. [16] These authors
found planar and octahedral structures, respectively. For K 6, the length
a and b (see Fig. 3) are found to be a = 4.48 Å and b= 4.41 Å. A similar
difference in the lengths has been also found in some previous studies
[17,41,21], where in other studies [19,14,15] a = b was found. It is
worth noting that the transition from 2D to 3D has been also predicted
for sodium clusters to occur at Na 6 [42].
Low symmetric structures are found for the clusters in the size range
9≤ N ≤20, where the most dominant point groups are C1, C2, and Cs. The
structures of K 9 and K 10 are characterized by a Cs symmetry similar to
a) K4 b) K6 that obtained for sodium clusters [35,36]. The structure of K 11 has C2v
symmetry. For the magic clusters with N=18 and 20, the corresponding
Fig. 3. The structures of K 4 and K 6 showing the definition of the lengths a and b. point groups are C1 and C3, i.e., these structures are not of particularly
S. Abdalla et al. / Surface Science 608 (2013) 255–264 259

a c

b d

Fig. 4. Different similarity functions used in quantifying the similarity between different objects. In (a), K N and K N−1 clusters are compared with the thin curve showing the results using
Eq. (10) and the thick curve showing those for Eq. (12). In (b), K N and K N−2 clusters are compared, and in (c), K N and Na N clusters are compared. Finally, in (d) the K N clusters after
deposition on the surfaces are compared with those in the gas phase. In this case, the thick line represents the similarity function for the clusters on the K(100) surface, the thin line
represents the similarity function for the clusters on the K(110) surface, and the dashed-line compares the structures of the K N clusters on the two surface. In (b), (c), and (d), only
the similarity function according to Eq. (12) is used.

high symmetry. Our finding for K 18 is in contrast to the C2v symmetry compared with the (N-1)-atomic cluster, and out of all the N different
obtained for Na 18 [35], although both studies agree that the structure values of q, the lowest one, qmin, is chosen to define a similarity
for this size is of relatively low symmetry and far from the roughly function
spherical symmetry assumed by the spherical jellium model.
To identify how the clusters grow and to quantify whether the 1
cluster with N atoms can be derived from the cluster with N-1 atoms S¼ ð10Þ
1 þ qmin =ul
by simply adding one atom, we have used two different theoretical
approaches. In the first case, we have used a so-called similarity func-
tion that we have found to be useful, at least for relatively compact where ul is a length unit.
systems [43,44]. For the comparison of two systems with P atoms, we This approach is not easily applied for comparing structures with
calculate and sort the interatomic distances of the two systems, {di} very different sizes or that are open. We have therefore developed a
and {di}, respectively. Subsequently we calculate new approach for the comparison of two systems A N and B M for
which we will assume that M ≥ N [45]. At first, the two structures
" P ðP−1
# 1 are scaled, and subsequently the two resulting structures are placed
2 XÞ=2 2

q¼ ðdi −di Þ : ð9Þ upon each other so that


P ðP−1Þ i¼1

N 
X → → 2

When comparing clusters with N and N − 1 atoms, we consider all Q¼ 1 R − 1 R  ð11Þ
d A;i d B;i 
A B
those N structures that can be obtained from the N-atomic cluster by i¼1

removing a single atom. For each of those, the resulting structure is

Fig. 6. Properties related to the eigenvalues Iαα/N5/3. The solid line lines show the eigen-
Fig. 5. The distribution of radial distances (in Å̊) for K N clusters as function of clusters values themselves, whereas the dashed line shows their average. In the upper part we
size. For each value of N, each small line represents at least one atom that has that mark whether the clusters have an overall spherical shape (S), a lens-like shape (L), or
radial distance. cigar-like shape (C).
260 S. Abdalla et al. / Surface Science 608 (2013) 255–264

b c

d e

Fig. 7. Various energetic properties related to the deposition of the K N clusters on a K surface as function of clusters size, N. The solid curves represent results for the K(100) surface
and the dashed curves represent those for the K(110) surface. Panel (b) shows the deposition energy, ED, that is split into the interaction energy, Eint [shown in (c)], and the
restructuring energy of the cluster, ERC [shown in (d)], and of the surface, ERS [shown in (e)]. Panel (a) shows the deposition energy per cluster atom.

is minimized.

Here, dA and dB are the two scaling factors for the two→
Since simple chemical reasoning may suggest that Na N and K N
systems, R A;i is the position of the ith atom of the A cluster, and R B;i clusters have very similar properties, we also applied the similarity
is the position of that atom of the B cluster that (after scaling) is clos- function of Eq. (12) to compare the structures of the two sets of
est to the ith atom of the A cluster (also after scaling). From the min- systems. For each of the two systems, we used a scaling factor being
imum value of Q we define a similarity function similar to Eq. (10), equal to the average of the nearest and next-nearest neighbor
distance in the crystal structure. The resulting similarity function is
shown in Fig. 4(c), where it can be seen that the clusters become
1
S¼ : ð12Þ increasingly different as N grows. Thus, the two systems are not
1 þ ðQ=NÞ1=2
completely analogous to each other. In order to get further information
on the structure, we study the so-called radial distances. For each K N
At first, we compare the clusters with K N and K N − 1 atoms. The cluster we first define its center,
results are shown in Fig. 4(a), where we compare the two approaches
of Eqs. (10) and (12). In the former, we have used ul = 1 Å, and in the → 1X N →
R0 ¼ R ð13Þ
latter we have scaled all atomic coordinates by a factor of 4.97 Å N i¼1 i
(which is the average of the nearest and next-nearest neighbor
distance in the crystal structure). The shapes of the two curves are
very similar, although the differences in the scaling factors result in and subsequently for each atom its radial distance,
differences in the absolute values. That the two approaches give
→ 
similar results can be explained from the fact that the present clusters   → →
r i ¼  r i  ¼ jR i −R 0 j: ð14Þ
are fairly closed packed. It is interesting to observe the even–odd
oscillatory pattern in particular for the smallest clusters. This suggests
that the cluster with, e.g., 5 atoms has a structure very similar to that The radial distances themselves are shown in Fig. 5. For some of the
of the 4-atomic cluster, although a new electronic orbital is being smallest clusters (N= 6, 7, 8, and 11, for instance), we have only few
occupied. For the 6-atomic cluster, a new structure is found, which, different values of the radial distances, which is a consequence of
to a large extend, is found for the 7-atomic cluster, too. This trend the high symmetry of these clusters. This is, on the other hand, not
continues up to N around 16, although it becomes less pronounced for the case for the larger clusters. Here, instead there is a tendency for
the larger values of N. The fact that the structure changes whenever developing structures with a small core (with corresponding small
two atoms are added can be seen in Fig. 4(b) where we compare the values of radial distances) covered by a shell of atoms (with larger radial
structure of K N with that of K N −2. Finally, a similar oscillatory pattern distances). This tendency is most pronounced for the largest cluster
was also found for the Na N clusters [36]. with N ≥15.
S. Abdalla et al. / Surface Science 608 (2013) 255–264 261

3 4 5 6

7 8 9 10

11 12 13 14

15 16 17 18

19 20
Fig. 8. A top view of the structures of the K N clusters on the K(100) surface.

Further information on the shape can be obtained as follows. For 3.2. Deposited potassium clusters
each cluster size we first define the 3 × 3 matrix containing the
elements Next, we shall discuss the changes in the properties of the potassium
clusters after they have been deposited softly on a (100) or (110)
surface of a potassium crystal.
1 X N
Ist ¼ 2
si t i ð15Þ
ul i¼1 3.2.1. Energetic properties
When the cluster is deposited on the surface, the total energy of the
cluster+ surface system is lowered, and the system is stabilized, which
can be seen from the negative deposition energies in Fig. 7(b). In that
with ul being a length unit, and s and t being x, y, or z. The three figure it is also seen that the deposition energy is largely independent
eigenvalues of this matrix, Iαα, can be used in separating the clusters of the crystal surface. The deposition energy per cluster atom [shown
into being overall spherical, cigar-shaped, or lens-shaped. The fact that in Fig. 7(a)] is an overall increasing function of N since the fraction
no cluster has three identical values of Iαα implies that no cluster has of the cluster atoms that is in contact with the surface is an overall
an overall spherical shape, which in turn suggests that the spherical decreasing function of N. There are, however, certain extra features in
jellium model at most can be of limited success for these systems. Fig. 7(a) of which the maximum for N = 8 and the minimum for N =5
Instead, in particularly the clusters for N b 15 have markedly different are the most pronounced ones. For N≤ 5 the gas-phase clusters are
values for Iαα so that their structure is significantly different from planar and can, thus, be placed on the surface so that all cluster atoms
spherical, and first for N >15 the three eigenvalues approach each experience (attractive) interactions. This result in the particularly low
other. The fact that the smallest clusters are planar manifests itself values for the deposition energies in this size range. The maxima for
here in that their smallest value of Iαα equals 0 (Fig. 6). N= 8 in Fig. 7(a) might be due to the high stability of this cluster size
262 S. Abdalla et al. / Surface Science 608 (2013) 255–264

3 4 5 6

7 8 9 10

11 12 13 14

15 16 17 18

19 20
Fig. 9. A top view of the structures of the K N clusters on the K(110) surface.

in the gas phase, suggesting that this cluster will not benefit much from deposited on the (100) surface, whereas a smaller value is found for
interactions with the surface. the same cluster on the (110) surface. We suggest that the difference
The fact that the deposition energy becomes increasingly negative as between those two values is a realistic estimate for the scatter in the
a function of cluster size is consistent with the results for the energetic restructuring energies when considering different surfaces and/or
properties of the gas-phase clusters, Fig. 1. Thus, the K systems are deposition geometries.
increasingly more stable, the larger the systems are. Compared to the total deposition energy, the two restructuring
As described in Section 2, it can be useful to split the deposition energies are accordingly the smaller parts and, instead, the dominating
energy into three different contributions: a restructuring energy of the part is the interaction energy between cluster and surface. Thus, this
cluster describing the energy costs for the cluster to change the struc- energy (Fig. 7(c)) follows close the deposition energy.
ture from that of the gas phase to the one it has on the surface, a similar
energy for the surface, and the remaining part that then describes the 3.2.2. Structural properties
interaction between the cluster and the substrate. These three contribu- We shall now discuss the structures that result from the deposition
tions are shown in Figs. 7(d), (e), and (c), respectively. of the clusters on the two surfaces. At first, we shall compare the struc-
At first, it is seen that the restructuring energy of the surface, ture of the clusters before and after deposition as well as the structures
Fig. 7(e), is essentially vanishing. In some of the calculations it takes of the clusters on the two different surfaces. To this end we use the
even negative values, which we shall explain below. The small values similarity function of Eq. (12). The results are shown in Fig. 4(d).
of the restructuring energy of the surface suggest that the changes in Compared to the other panels in that figure, the values of Fig. 4(d) are
the surface structures indeed are small, so that our approach based on in general lower, implying that the changes in the cluster structures
letting only the top-most layer relax is justified. are significant. Moreover, the structures on the two different surfaces
On the other hand, the restructuring energy of the cluster, are clearly different, suggesting that deposited clusters of the same
Fig. 7(d), is clearly non-zero and positive, although also here there size may not take only one structure but the resulting structure depends
are cases where it is fairly small. The largest value is found for K 11 also on the orientation and position of the cluster on the surface.
S. Abdalla et al. / Surface Science 608 (2013) 255–264 263

That the structures of the clusters change upon deposition on 4. Conclusions


the surfaces can also be recognized in Figs. 8 and 9, where we show
these. It is clear that the structures differ from those of Fig. 2, and a In summary, we have presented results for the structural and
careful examination of the structures on the surfaces suggests that the energetic properties of potassium clusters with up to 20 atoms that
clusters there have a structure that is partly dictated by a tendency of are either isolated or deposited on a surface of a potassium crystal.
the cluster atoms to sit on positions on the surface where they have a The structure optimization was performed by using a combination
large number of surface-neighbors. This is the case for bridged and of DFTB calculations for the determination of the total energy for a
hole positions on the surface. One may thus suggest that there is a given structure and genetic algorithms for the determination of the
certain tendency for the cluster atoms to place themselves epitaxial global total-energy-minima structures. From the stability function it
on the surface. is clearly seen that the K N clusters show an even–odd oscillatory
In order to quantify to which extend the surface changes structure pattern and that more pronounced peaks are found for the sizes 8,
upon the deposition and to which extend the cluster atoms are placed 18, and 20, which is in agreement with the results obtained from
epitaxial on the surface, we apply certain so called indices that are the spherical jellium model and mass spectra of potassium cluster
defined as follows. At first we define an index of epitaxy, IE. Taking beams. However, the structural analysis showed that these clusters
the positions of the two bottom layers of atoms for the slab model are not even approximately spherical. Moreover, the similarity func-
that is used for the surface, we imagine the positions of these atoms tion demonstrated that there is a tendency for the smaller clusters
being continued virtually into those parts of space where the upper- to pairwise possess similar structures. This is the case for the clusters
most surface layer and the cluster are placed. Subsequently, we define with 4 and 5 atoms, with 6 and 7 atoms, with 8 and 9 atoms, etc.
When the potassium clusters are deposited softly on a potassium
N → 
2 1X  → 2 surface, both the energetic and the structural properties change. The
qE ¼  R −R iX0  ; ð16Þ deposition energy becomes more negative with the cluster size increase
N i¼1 i
and, in general, the deposition of a clusters on the K(110) surface leads
→ to a larger decrease in the total energy compared to the deposition of
where R i is→the position of the ith atom of the cluster that has N atoms. the cluster on the K(100) surface. By separating the deposition energy
Moreover, R iX0 is the position→
of that atom of the virtual continuation of
the crystal that is closest to R i . Subsequently, we define

1
IE ¼ : ð17Þ
1 þ qE =ul

In a similar way we define an index, IC, for the surface describing


whether the positions of the atoms of the uppermost surface layer in
the presence of the cluster are close to those of the virtual continuation
of the crystal. The only difference is that the summation in Eq. (16) then
runs over the 72 atoms of the upper surface layer. Finally, we compare
the structure of the surface layer without and with the cluster through
a further index, IS, defined as
a
1
IS ¼ ð18Þ
1 þ qS =ul b
with

72 → → 2
2 1 X 
qS ¼  R iS0 −R iS  ; ð19Þ
72 i¼1

→ →
where R iS0 and R iS is the position of the ith atom of the upper surface
layer in the absence and in the presence of the cluster, respectively.
These three indices are shown in Fig. 10. At first it is seen that IE takes
fairly low values, implying that the cluster atoms are not placed at
positions close to those of the crystal. However, IC is even smaller, giving
that the outermost surface layer experiences strong structural relaxa-
tions compared to the crystal. As suggested already by the results of
Fig. 7(e), the structural relaxation of the surface upon deposition of
the cluster is small with some exceptions. In fact, these exceptions are
closely related to those for which the restructuring energy of the surface
was particularly low. A careful inspection of the structural changes in
those case reveals that the changes are related to a single atom at the
boundary of the finite surface model. This atom changes in some cases c
place and lowers thereby the total energy of the complete surface. That
this is possible is not surprising: after all, the structure of the surface
model was not determined as that of the lowest total energy for the
system of 216 K atoms, but rather a local total-energy minimum. Thus, Fig. 10. Various quantities describing the structural properties of the cluster+surface
a small perturbation, for instance the presence of a cluster, may change system. The full and the dashed curve shows the results for the (100) and (110) surface,
this structure. respectively. For further details, see the text.
264 S. Abdalla et al. / Surface Science 608 (2013) 255–264

into two structural reorganization energies (for the surface and the [11] M.M. Kappes, P. Radi, M. Schär, E. Schumacher, Chem. Phys. Lett. 113 (1985) 243.
[12] A. Kornath, R. Ludwig, A. Zoermer, Angew. Chem. Int. Ed. 37 (1998) 1575.
cluster, respectively) and an interaction energies, we found that the [13] H. Soll, J. Flad, E. Golka, Th. Krüger, Surf. Sci. 106 (1981) 251.
interaction energy is the dominating one, although the restructuring [14] G. Pacchioni, H.O. Beckmann, J. Koutecký, Chem. Phys. Lett. 87 (1982) 151.
energy of the clusters occasionally is fairly large. The analysis of the [15] F. Spiegelmann, D.J. Pavolini, J. Chem. Phys. 89 (1988) 4954.
[16] S.C. Richtsmeier, D.A. Dixon, J.L. Gole, J. Phys. Chem. 86 (1982) 3942.
structures of the clusters on the surfaces gave that the clusters change [17] J. Flad, G. Igel, M. Dolg, H. Stoll, H. Preuss, Chem. Phys. 75 (1983) 331.
their structures upon the deposition. In addition, we found that the [18] A. Pellegatti, B.N. McMaster, D.R. Salahub, Chem. Phys. 75 (1983) 83.
structures of the clusters of the two surfaces are different. [19] A.K. Ray, S.D. Altekar, Phys. Rev. B 42 (1990) 1444.
[20] E. Florez, P. Fuentealba, Int. J. Quant. Chem. 109 (2009) 1080.
Finally, our results showed that the structure of the surface relaxes [21] A. Banerjee, T.K. Ghanty, A. Chakrabarti, J. Phys. Chem. A 112 (2008) 12303.
significantly compared to the crystal structure, but only little upon depo- [22] S.K. Lai, P.J. Hsu, K.L. Wu, W.K. Liu, M. Iwamatsu, J. Chem. Phys. 117 (2002) 10715.
sition of the clusters. The cluster atoms did not show any tendency [23] P. Hohenberg, W. Kohn, Phys. Rev. 136 (1964) B864.
[24] W. Kohn, L.J. Sham, Phys. Rev. 140 (1965) A1133.
toward an epitaxial growth of the crystal.
[25] G. Seifert, R. Schmidt, New J. Chem. 16 (1992) 1145.
[26] D. Porezag, Th. Frauenheim, Th. Köhler, G. Seifert, R. Kaschner, Phys. Rev. B 51
Acknowledgements (1995) 12947.
[27] G. Seifert, D. Porezag, Th. Frauenheim, Int. J. Quant. Chem. 58 (1996) 185.
[28] D.M. Deaven, K.M. Ho, Phys. Rev. Lett. 75 (1995) 288.
One of the authors (SA) would like to thank Deutscher Akademischer [29] J.R. Morris, D.M. Deaven, K.M. Ho, Phys. Rev. B 53 (1996) R1740.
Austauschdienst (DAAD) for the financial support. [30] J.A. Niesse, H.R. Mayne, Chem. Phys. Lett. 261 (1996) 576.
[31] C. Roberts, R.L. Johnston, N.T. Wilson, Theor. Chem. Acc. 104 (2000) 123.
[32] B. Hartke, Chem. Phys. Lett. 240 (1995) 560.
References [33] Y. Dong, M. Springborg, J. Phys. Chem. C 111 (2007) 12528.
[34] Y. Dong, M. Springborg, Eur. Phys. J. D43 (2007) 15.
[1] M.Y. Chou, A. Cleland, M.L. Cohen, Solid State Commun. 52 (1984) 645. [35] V. Tevekeliyska, Y. Dong, M. Springborg, V.G. Grigoryan, Eur. Phys. J. D 43 (2007) 19.
[2] W.D. Knight, K. Clemenger, W.A. de Heer, W.A. Saunders, M.Y. Chou, M.L. Cohen, [36] V. Tevekeliyska, Y. Dong, M. Springborg, V.G. Grigoryan, in Aromaticity and Metal
Phys. Rev. Lett. 52 (1984) 2141. Clusters, Taylor & Francis, Ed. P. K. Chattaraj, 2011. 161.
[3] W.D. Knight, K. Clemenger, W.A. de Heer, W.A. Saunders, Phys. Rev. B 31 (1985) [37] H. ur Rehman, M. Springborg, Y. Dong, Eur. Phys. J. D 52 (2009) 39.
2539. [38] H. ur Rehman, M. Springborg, Y. Dong, J. Phys. Chem. A 115 (2011) 2005.
[4] W.D. Knight, W.A. de Heer, K. Clemenger, W.A. Saunders, Solid State Commun. 53 [39] Y. Dong, M. Springborg, Am. Inst. Phys. Conf. Proc. 1148 (2009) 342.
(1985) 445. [40] W. Hellmann, R.G. Hennig, S. Goedecker, C.J. Umrigar, Phys. Rev. B 75 (2007) 085411.
[5] G.A. Thompson, D.M. Lindsay, J. Chem. Phys. 74 (1981) 959. [41] A. Jiemchooroj, B.E. Sernelius, P. Norman, J. Comput. Method Sci. Eng. 7 (2007) 475.
[6] W.A. Saunders, K. Clemenger, W.A. de Heer, W.D. Knight, Phys. Rev. B 32 (1985) 1366. [42] R. Poteau, F. Spiegelmann, Phys. Rev. B 45 (1992) 1878.
[7] W.D. Knight, W.A. de Heer, W.A. Saunders, Z. Phys. D 3 (1986) 109. [43] V.G. Grigoryan, M. Springborg, Phys. Rev. B 70 (2004) 205415.
[8] C. Brèchignac, Ph. Cahuzac, Chem. Phys. Lett. 117 (1985) 365. [44] M. Springborg, in Handbook of Computational Chemistry, Springer Verlag, Ed.
[9] C. Brèchignac, Ph. Cahuzac, Z. Phys. D 3 (1986) 121. J. Leszczynski, 2012. 955.
[10] M.M. Kappes, M. Schär, P. Radi, E. Schumacher, J. Chem. Phys. 84 (1986) 1863. [45] Y. Dong, M. Springborg, Y. Pang, F. M. Morillon, in preparation.

You might also like