Anda di halaman 1dari 7

Computational and Theoretical Chemistry 1053 (2015) 238–244

Contents lists available at ScienceDirect

Computational and Theoretical Chemistry


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

Chemical bonding and surface interactions in Bi2Se3 and Bi4Se3


Matthew S. Christian a, Sarah R. Whittleton a, Alberto Otero-de-la-Roza b, Erin R. Johnson a,⇑
a
Chemistry and Chemical Biology, School of Natural Sciences, University of California, Merced, 5200 North Lake Road, Merced, CA 95343, USA
b
National Institute for Nanotechnology, National Research Council of Canada, 11421 Saskatchewan Drive, Edmonton, Alberta T6G 2M9, Canada

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

Article history: Topological insulators are a new class of materials that are insulators in the bulk but have near zero elec-
Received 8 August 2014 tron transport dissipation behavior on the surface. These surface states are topologically robust (resistant
Received in revised form 17 September 2014 to impurities, defects, and geometry deformations), which makes these materials ideal candidates for a
Accepted 18 September 2014
number of technological applications. Well-known three-dimensional topological insulators are bismuth
Available online 30 September 2014
selenide (Bi2Se3), which is composed of five-atom-thick layers that interact non-covalently with each
other, and Bi–Se alloys resulting from combining these quintuple layers with elemental bismuth bilayers.
Keywords:
In this article, we examine the surface sliding and binding energetics of the combinations of quintuple
Chemical bonding
Density-functional theory
layers and bisumth bilayers found in Bi2Se3 and Bi4Se3. In addition, we investigate the nature of the
Topological insulators chemical bonding in these systems and its relation to the surface energetics.
Van der Waals Ó 2014 Elsevier B.V. All rights reserved.
Tribology

1. Introduction electronic structure, the signature of a QSH state is the crossing


of bands corresponding to surface states, forming what is called a
Topological insulators (TIs) [1–7] are a new class of materials Dirac cone [5]. In addition to being theoretically interesting [3],
characterized by metallic surface states that are ‘‘topologically pro- TIs have great potential for use in future technological applications
tected,’’ that is, resistant to impurities and surface defects, and by [4,6] such as, for instance, spintronics [8] and thermoelectrics
being an insulator in the bulk. The mechanism behind the TI prop- [9,10].
erties is similar to the quantum Hall (QH) effect. In the QH state, The strong spin–orbit interaction required for the TI state limits
electrons in a bidimensional material subject to a strong magnetic the availability of such materials. Heavy elements, where relativis-
field perpendicular to the surface are forced to move in quantized tic effects are important, are usually involved to generate strong
orbits, making the material an insulator in the bulk, but conducting spin–orbit coupling. Bi2Se3 is the most-studied three-dimensional
on the surface. In these metallic surface states, electrons are forced topological insulator [3,11,12], presenting a Dirac cone composed
to move exclusively in one direction, thus preventing backscatter- of surface states and a relatively large band gap in the bulk. Its
ing at defects and achieving an almost dissipationless electron structure, shown in Fig. 1, consists of ABCABC stacks of five-atom
flow. Topological insulators show a closely-related behavior, called layers, called quintuple layers (QL). The QLs interact with each
the quantum spin Hall (QSH) effect. In the QSH state, as in the QH other primarily via van der Waals forces [13]. Each Bi atom in a
state, dispersionless metallic surface states are present, but a QL forms close contacts with six Se atoms, in an octahedral
strong external magnetic field is not required, thus making these arrangement. The Bi4Se3 material is formed by combining QLs
materials much more interesting from a practical point of view. and elemental bismuth bilayers (Bi2) in a 1:1 ratio (Fig. 1), and also
The QSH state arises from strong spin–orbit interactions present presents QSH states and topological insulator behavior [14,15]. The
in TIs. In a TI, the electron spin is locked perpendicular to the crys- bismuth bilayer is a known 2D topological insulator [16]. Because
tal momentum, and electrons of different spins are forced to move of the relatively weak inter-layer binding, the QL and Bi2 can be
exclusively in opposite directions on the surface [5]. Backscatter- stacked to form Bi–Se alloys with variable Bi and Se concentration,
ing, which would involve a spin-flip, is forbidden due to the pres- resulting in extremely long cell lengths in the stacking direction
ence of time-reversal symmetry. The QSH state is resistant to and possibly incommensurate phases [13]. Material properties
impurities, surface defects and disorder [5]. In terms of the could be controlled by varying the stoichiometry, though more
information about the surface energetics and chemical behavior
⇑ Corresponding author. is needed; [16] this will be addressed in the present work via
E-mail address: ejohnson29@ucmerced.edu (E.R. Johnson). density-functional theory (DFT).

http://dx.doi.org/10.1016/j.comptc.2014.09.023
2210-271X/Ó 2014 Elsevier B.V. All rights reserved.
M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244 239

Several experimental and theoretical studies on the surface 2. Computational methods


behavior of Bi–Se materials have been conducted in recent years.
It is known that the Bi2Se3 surface is relatively inert [17] and resis- Common density functionals have several well-known short-
tant to oxidation [18,19]. As expected, Bi2Se3 cleaves preferentially comings [35], of which the lack of dispersion interactions and the
at the van der Waals inter-QL contacts. It is known experimentally systematic underestimation of the band gap are particularly rele-
that the geometry and electronic structure of the surface are sensi- vant to the modeling of topological insulators. In recent years,
tive to cleavage and show ageing effects, with the Bi concentration great progress has been made towards the efficient and accurate
increasing over time, although the Se termination can be recovered modeling of non-covalent interactions by using dispersion-energy
after annealing [8,20]. He et al., on the other hand, report a prefer- functionals [36,37]. These methods capture long-range correlation
ence for a surface Bi termination in Bi2Se3 at room temperature effects and are coupled to a common density functional (called the
that may be attributed to a bismuth bilayer [21], but the Se termi- base functional) that, in principle, accounts for the non-covalent
nation is preferred at low temperatures. The position of the Dirac energy terms other than dispersion.
cone relative to the Fermi level, which is important in the charac- The structures of the topological insulators considered in this
terization of the TI behavior, is also affected by charge transfer article (Bi2Se3 and Bi4Se3) consist of stacked QLs and bismuth
from surface adatoms [20,22]. bilayers that interact with each other via non-covalent interac-
The nature of the inter-layer interaction was previously tions (see Fig. 1). In this work, we account for the dispersion
explored by Lind et al. [13] using local and semi-local density energy by using the exchange-hole dipole moment (XDM) model,
functionals. However, a functional that includes dispersion attrac- proposed by Becke and Johnson [25,26,38–47]. In XDM, the total
tion is required to accurately describe the inter-layer van der energy from the base functional is supplemented by the
Waals interactions. Liu et al. [23] and Luo et al. [24] used non- dispersion correction,
local van der Waals density functionals to study the band struc-
ture, equilibrium geometry, and exfoliation of Bi2Se3, although lit- E ¼ Ebase þ EXDM ; ð1Þ
tle attention has been paid to the surface energetics and the
nature of the inter-layer chemical bonding. In this work, we apply that is calculated using an atom-based pairwise expression:
the exchange-hole dipole (XDM) dispersion model [25], which
has been shown to give accurate exfoliation energies for graphite X X C n;ij f ðRij Þ
n
EXDM ¼  : ð2Þ
and sublimation energies for molecular crystals [26,27], to study
n¼6;8;10 i<j
Rnij
the binding energies and sliding behavior of the QL and Bi2 sur-
faces. Furthermore, the nature of the chemical bonding in Bi2Se3 The sum runs over all pairs of atoms in the system, Rij is the dis-
and Bi4Se3 is relatively unexplored. We present an analysis of tance between atoms i and j; f n is a damping function that deacti-
the chemical bonding properties in these materials using a range vates the dispersion correction at short range, and the C n;ij are the
of techniques, including the Quantum Theory of Atoms in Mole- dispersion coefficients for each pairwise interaction. The latter are
cules (QTAIM) [28–30], Non-Covalent Interactions (NCI) [31,32], approximated non-empirically by the electrostatic energy of the
and the Electron Localization Function (ELF) [33,34], to study electron distributions and their exchange holes in the two interact-
the atomic charges and inter-layer interactions in these Bi–Se ing atoms. It has been shown that the XDM model offers excellent
materials. accuracy for gas-phase [46], as well as condensed-matter, systems
[26,27].
In this article, we use the XDM dispersion functional to model
the inter-layer interactions in the Bi–Se materials. Our calculations
were performed under periodic boundary conditions using the
pseudopotentials-plane waves approach as implemented in the
Quantum ESPRESSO program [48]. We used B86b [49] exchange
coupled with the PBE correlation [50] functional. Both Bi2Se3 and

Bi4Se3 crystallize in the rhombohedral (R3m) space group and we
have used the rhombohedral (primitive) and hexagonal (conven-
tional) cells where convenient. The primitive-cell calculations were
carried out using a 4  4  4 k-point grid while the grid for the
hexagonal cell was 4  4  2. We used the Projector-Augmented
Wave method [51] (PAW) and scalar-relativistic pseudopotentials.
The cutoff energies were 60 Ry for the plane-waves and 600 Ry for
the density.
Regarding the surface energetics, we optimized the bulk geom-
etries and found the equilibrium structures of Bi2Se3 and Bi4Se3. In
Bi2Se3, the most stable cleavage occurs between QLs because of
their weak non-covalent binding. The Bi4Se3 crystal presents alter-
nating QL and bismuth bilayers, so cleavage disrupts van der Waals
contacts between a QL and a Bi2 bilayer.
It is also interesting to examine the non-covalent interactions
and energy landscape for sliding, which determines the tribological
behavior of these materials. In Bi2Se3, we considered two QLs slid-
ing on top of each other, and for Bi4Se3 we examine a bismuth
bilayer sliding over a QL. These calculations were performed using
Fig. 1. Structures of bulk Bi2Se3 (left) and Bi4Se3 (right), showing the quintuple a supercell with a length of 75 bohr atomic units perpendicular to
layer (QL) in the former and the QL and intercalated Bi2 layers in the latter. Bi atoms
are represented in violet and Se atoms in yellow. (For interpretation of the
the slabs. The sliding energetics were computed by displacing the
references to color in this figure legend, the reader is referred to the web version of top layer and fixing the x and y atomic coordinates while relaxing
this article.) the vertical coordinate.
240 M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244

techniques. QTAIM atomic charges were calculated using the Yu–


Trinkle algorithm [52] implemented in the CRITIC2 program
[53,54]. In addition, the electron localization function (ELF)
[33,34] was used to determine the relative importance of the cova-
lent, inter-layer charge transfer, and ionic contributions to the
bonding by mapping the position of the localized electron pairs.
Calculations using norm-conserving pseudopotentials and a larger
cutoff energy were used to compute the ELF.
Finally, we used the non-covalent interaction (NCI) index
[31,32] to map the intra- and inter-layer interactions. The NCI
index is a measure of non-covalent contacts, similar to the topol-
ogy analysis of the electron density in QTAIM. It is based on two
quantities: the electron density and the reduced density gradient
(s, RDG), defined as

1 jrqj
s¼ : ð3Þ
2ð3p2 Þ
1=3 q4=3
NCI plots comprise low reduced-gradient isosurfaces, for a certain
small isovalue of s. The coloring is determined by the electron den-
sity times the sign of the second Hessian eigenvalue [31]. Strong
Fig. 2. Inter-layer and intra-layer distances (in angstrom) in Bi2Se3 and Bi4Se3 at the non-covalent interactions appear as localized blue domains
bulk geometry. The atomic Bader charges for the atoms in each layer are shown in whereas weak molecular contacts are represented by extended
red. The distance between QLs in Bi2Se3 and between QL and Bi2 in Bi4Se3 are
green regions. The NCI plots were generated using CRITIC2 [53,54].
indicated by dashed arrows. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)
3. Results and discussion
Table 1
Inter-layer binding energies for different interactions between the QL and the 3.1. Structure and chemical bonding in bulk Bi2Se3 and Bi4Se3
bismuth bilayer (in kcal/mol). The base functional (‘‘Base’’) and dispersion (‘‘Disp’’)
contribution to the binding energy are also given. The inter-layer distances and Bader atomic charges for bulk
Layers Total Base Disp Bi2Se3 and Bi4Se3 are shown in Fig. 2. The inter-QL distance for
Bi2Se3 (2.52 Å) is in close agreement with experimental results
(QL)–(QL) 6.33 0.52 6.86
(QL)(QL)–(QL) 6.36 0.48 6.85 (2.58 Å [55], 2.25 Å [56]) and previous DFT results: C09-vdW [24]
(QL)–(QL)–(QL) 12.70 1.01 13.71 (2.58 Å), PBE-D2 with spin–orbit coupling [24] (2.57 Å), and
(QL)–(Bi2) 8.60 0.07 8.67 optPBE-vdW [23] (2.41 Å). Non-dispersion-corrected functionals
(Bi2)(QL)–(Bi2) 8.68 0.26 8.94 such as PBE with spin–orbit coupling give very long inter-layer
(Bi2)–(QL)–(Bi2) 17.28 0.33 17.61 separations [24] (3.30 Å), as does the non-local van der Waals
(QL)(Bi2)–(QL) 8.99 0.15 9.14 functional coupled with revPBE [24] (3.72 Å).
(QL)–(Bi2)–(QL) 17.60 0.22 17.81
The inter-layer distances between two QLs and between one QL
and the bismuth bilayer are significantly longer than the intra-
In order to examine the nature of the chemical bonds in these layer distances, as expected because of the non-covalent character
bismuth selenide systems, we employed a number of different of the bonding. However, the intra-QL distances differ depending

Fig. 3. Non-covalent interaction (NCI) plots within one QL (top left), the bismuth bilayer (top right), and the between one QL and Bi2 in Bi4Se3 (bottom left), and the inter-
layer interaction in Bi2Se3 (bottom right). The isovalue is s ¼ 0:3. Bi atoms are represented in violet and Se atoms in yellow. Strong non-covalent interactions are represented
in blue and weak interactions are in green. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244 241

Fig. 4. Contour plots of the electron localization function (ELF) in the plane bisecting two QLs in Bi2Se3 (left) and two QLs and the intercalated bismuth bilayer in Bi4Se3
(right). The green dotted lines correspond to the paths used in Fig. 5. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)

on whether the QL is adjacent to a Bi2 bilayer or to another QL. The


1
QL intra-layer distances in Bi2Se3 (1.58 and 1.92 Å) are closer to the
0.9
equivalent distances in an isolated QL (1.59 and 1.93 Å) than to
Electron localization function
0.8
those in Bi4Se3 (1.54 and 1.83 Å), indicating that the donor– 0.7
acceptor interactions between two QLs are much smaller than 0.6
between a bismuth bilayer and a QL. Likewise, the Bi2 intra-layer 0.5
distance adjacent to a QL in Bi4Se3 (1.76 Å) is shorter than its value 0.4
in vacuum (1.79 Å). In Table 1, discussed later, this inter-layer 0.3
donor–acceptor interaction results in a higher binding energy 0.2
between a QL and Bi2 than between QLs themselves. 0.1
0
These geometric differences can be explained readily by inter-
layer charge-transfer effects. The Se atoms at the center or edges Se Bi Se Bi Se Se
of a QL have different charges, with the center seleniums being
Se Bi Se Bi Se Bi Bi
Bi2Se3
more ionic than the edge seleniums because they are coordinated Bi4Se3
to a higher number of electropositive Bi atoms, which essentially 0 2 4 6 8 10 12 14 16 18 20
behave as +1 cations. The bismuth bilayer, however, is formally Interatomic distance (Å)
composed of elemental Bi and our results indicate that there is a
charge transfer of 0.13 electrons from the Bi2 to each of the adja- Fig. 5. Electron localization function (ELF) along the paths connecting bonded
cent QLs. This charge transfer increases the ionicity of the QL and atoms in successive atomic layers (green paths in Fig. 4). The intra-QL ELF profiles
are represented first, while the inter-layer regions and the intra-Bi2 region are
decreases the intra-layer distances, while at the same time weak-
shown at the end of the plot. The profiles for Bi2Se3 (red) and Bi4Se3 (blue) are
ening the intra-layer bonding of Bi2. These observations point to superimposed. The atomic positions are shown below the plot (red for atoms in
the ionic nature of the binding in each QL and the metallicity of Bi2Se3 and blue for Bi4Se3). (For interpretation of the references to color in this
the bismuth bilayer. figure legend, the reader is referred to the web version of this article.)
The atomic charges in Fig. 2 can be compared to the QL charges
in the vacuum: 1.03 for Bi, 0.84 (Se, center), and 0.61 (Se, edge).
The charge of Bi2 in vacuum is naturally zero. These results indicate the surface of the QL may be related to the metallic conductance.
that charge transfer from Bi2 to the QL increases its ionicity, short- The charge transfer from the Bi2 bilayer to the QL in Bi4Se3 induces
ens the intra-layer distances, and improves the intra-layer binding. changes in the bonding restricted to the edge Se-Bi bond. As seen in
The strength and spatial extent of the inter- and intra-layer Fig. 5, the ELF value at the bond midpoint decreases by about 0.1,
interactions can be visualized using the NCI plots shown in indicating an increase in ionicity.
Fig. 3. Again, the intra-layer interactions are significantly stronger, The inter-layer ELF is consistently smaller than 0.5, but is higher
leading to small, localized high-density (blue) NCI domains, for the QL–Bi2, than the QL–QL, non-covalent interaction, consis-
whereas the inter-layer contacts are weaker (green). The NCI plots tent with the inter-layer charge transfer behavior in the former.
also indicate that the QL–Bi2 interaction is stronger than the QL–QL The Bi–Bi ELF along the bond path is higher than the equivalent
interaction. However, in both cases the inter-layer interactions are Bi–Se intra-layer values in the QL, indicating relatively-strong Bi–
resolved into localized atom–atom contacts, instead of extended Bi bonding.
contact regions as are common in molecular crystals [32].
The ELF maps the degree of electron localization in real space. 3.2. Binding energies
Fig. 4 shows a colormap of the ELF values in bulk Bi2Se3 and Bi4Se3,
and Fig. 5 represents the same quantity along a path connecting Table 1 shows the inter-layer binding energies for different
bonded atoms in successive atomic layers. The figures show that, combinations of QLs and Bi2 bilayers. The interaction between
within a QL, bonds between the edge Se and Bi are more covalent two QLs is weaker than that between a QL and Bi2 by about
than between the central Se and Bi, which according to the charges 2.3 kcal/mol, as predicted by the NCI plots and the atomic charges
in Fig. 2 are more ionic. The increased electron delocalization on in the previous section.
242 M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244

The ability to bind to a single QL is mostly unaffected by adja- 3.3. Sliding


cent layers. The change in the QL–QL binding energy caused by a
third QL is only 0.03 kcal/mol. It is known experimentally that Another way of examining the inter-layer interactions in Bi2Se3
the intra-layer phonon frequencies change only slightly (a few and Bi4Se3 is to study the effect of displacing combinations of the
cm1) with increasing QL thickness in thin films [57], which is con- QL and Bi2 layers with respect to each other. In addition to provid-
sistent with our results. Our observations also rule out any kind of ing further insight regarding the bonding, study of their tribologi-
inter-layer cooperative effects in multi-QL thin films, suggesting cal behavior has relevance to the processing and manufacturing of
that the onset of the TI state with thickness [58] comes from the these materials. The potential energy surfaces (PES) for sliding of
physical separation of the surfaces on both sides of the slab. either a QL or Bi2 bilayer over a QL are shown in Fig. 6, together
Contrary to what happens between QLs, the binding energy of with the dispersion and base-functional contributions and the
QL–Bi2 is affected if the bismuth bilayer is in contact with another inter-layer distances.
QL. This is not surprising given the charge transfer between these The sliding energetics of the QL–Bi2 model are different from
layers, seen in Figs. 2 and 5. A QL adjacent to the Bi2 increases the QL–QL PES. The inter-layer contact between two QLs is through
the binding energy of the QL on the opposite side of the bismuth negatively-charged Se atoms and, as a consequence, the configura-
bilayer by 0.39 kcal/mol. These inter-layer cooperative effects are tion where two Se atoms are on top of each other is strongly disfa-
not expected to propagate in the Bi4Se3 bulk. In all cases studied, vored, having a binding energy of only 0.5 kcal/mol. The Se atoms
the inter-layer binding comes almost exclusively from the disper- at the edge of the QL prefer to be positioned over a void in the sur-
sion energy contribution, consistent with the dominant van der face below and voids with an underlying Bi cation are preferred to
Waals interaction between layers. the empty voids by about 1 kcal/mol.

4.0 3.0 4.0 4.5


BE per slab (kcal/mol)

BE per slab (kcal/mol)


Se Se Se Se
4.0
3.0 2.5 3.0
3.5
2.0
y (Å)
y (Å)

2.0 2.0 3.0


1.5 2.5
1.0 Bi 1.0 Bi
2.0
0.0
(a) Se Se
1.0
0.0 Se Se 1.5
0.5 1.0
−3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0 −3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0
x (Å) x (Å)
BE per slab (kcal/mol, XDM)

BE per slab (kcal/mol, XDM)


4.0 3.5 4.0 4.5
Se Se Se Se 4.0
3.0 3.0 3.0 3.5
2.5 3.0
y (Å)
y (Å)

2.0 2.0
2.5
Bi 2.0 Bi 2.0
1.0 1.0
1.5
0.0
(b) Se Se
1.5
0.0 Se Se 1.0
1.0 0.5
−3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0 −3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0
x (Å) x (Å)
BE per slab (kcal/mol, bare)

BE per slab (kcal/mol, bare)

4.0 −0.2 4.0 0.20


Se Se −0.3 Se Se
3.0 −0.4 3.0 0.10
−0.5
y (Å)
y (Å)

2.0 2.0 0.00


−0.6
Bi −0.7 Bi −0.10
1.0 1.0
−0.8
0.0
(c) Se Se −0.9 0.0 Se Se
−0.20

−1 −0.30
−3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0 −3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0
x (Å) x (Å)
4.0 3.6 4.0 3.6
Se Se Se Se
3.4 3.4
3.0 3.0
3.2
z (Å)
y (Å)

3.2
y (Å)

2.0
z (Å)

2.0
3.0
Bi 3.0 Bi
1.0 1.0 2.8

0.0
(d) Se Se
2.8
0.0 Se Se
2.6
2.6 2.4
−3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0 −3.0 −2.0 −1.0 0.0 1.0 2.0 3.0 4.0
x (Å) x (Å)

Fig. 6. Plots in the left column correspond to the sliding of two QLs relative to each other and the right column presents data for a Bi2 layer on top of a QL. In all the plots, the
origin corresponds to a configuration where the atoms in the top layer are exactly on top of the atoms in the bottom layer. The four rows show, in order: (a) the potential
energy surfaces for sliding represented as binding energy relative to the isolated layers, (b) the dispersion contribution to the binding energy, (c) the base functional
contribution to the binding energy, and (d) the inter-layer distance. All energies are in kcal/mol and all distances are in angstroms.
M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244 243

The geometry of the Bi2 surface contact is roughly the same as adjacent to the QL. However, there is a small degree of non-additiv-
that of the QL, except that the Se atoms are replaced by Bi. As in the ity in the Bi2–QL interaction where the presence of another QL
QL–QL case, Bi atoms in the bilayer prefer to be in direct contact adjacent to the Bi2 increases the binding energy by 0.4 kcal/mol.
with the voids possessing an underlying Bi atom, but with a stron- Finally, we calculated the energetics of sliding for two QLs and
ger binding energy consistent with the results in Table 1. However, for QL–Bi2. In both cases, the orientation where the surface atoms
the geometry with the Bi directly over the other QL void is very of one layer face the QL voids containing the underlying Bi atoms is
unstable and configurations with Bi on top of the Se atoms are favored. In the QL–QL system, the second most-stable position on
favorable. The energy barrier for sliding is smaller in (QL)2 the PES is for the surface atoms to face the other void, with an
(1 kcal/mol) than in QL–Bi2 (2 kcal/mol). underlying vacancy, while in QL–Bi2, it is the orientation where
For both (QL)2 and QL–Bi2, the dominant features of the poten- the Bi atoms are on top of the Se atoms on the surface of the QL.
tial energy surface are captured by the dispersion-energy term. The The energy barrier for sliding is significantly larger in QL–Bi2
lowest-energy arrangements occur when the corrugated surfaces (2 kcal/mol) than in QL–QL (1 kcal/mol). The inter-layer binding
interlock, such that the number of close atomic contacts, and hence and sliding energetics are dominated by dispersion interactions.
the dispersion energy, is maximized. Interestingly, the base func- This implies that formation of other Bi–Se alloys, mixing QLs and
tional gives a contribution to the PES that is roughly an inverted Bi2 bilayers with more complex stoichiometry, should be facile.
version of the dispersion-corrected PES. The minimum-energy ori- This could allow systematic tuning of materials properties to meet
entations occur where the number of close atomic contacts, which specific manufacturing requirements for spintronic and thermo-
are repulsive in the absence of dispersion, is minimized. This con- electric applications.
firms the dominant dispersion contribution to the inter-layer inter-
actions. In the most stable orientation, the two surfaces approach Acknowledgements
closer to each other and the dispersion attraction serves to shift
the position of the Pauli repulsive wall of the potential. This The authors thank the Spanish Malta/Consolider initiative (No.
increases the repulsive contribution from the base functional. CSD2007-00045) and the UC Lab Fees Program, Award number
The dispersion contribution is not as favorable in the on-top 237789.
configuration and the repulsive wall is encountered at longer
inter-layer distances, as shown in the last two panes of Fig. 6. References

[1] B.A. Bernevig, T.L. Hughes, S.-C. Zhang, Quantum spin Hall effect and
topological phase transition in HgTe quantum wells, Science 314 (2006)
4. Conclusions
1757–1761.
[2] M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buhmann, L.W. Molenkamp, X.-L.
We have presented in this article a study of chemical bonding Qi, S.-C. Zhang, Quantum spin Hall insulator state in HgTe quantum wells,
and surface energetics in topological insulators based on Bi–Se Science 318 (2007) 766–770.
[3] J. Maciejko, T.L. Hughes, S.-C. Zhang, The quantum spin Hall effect, Ann. Rev.
alloys, with particular emphasis on Bi2Se3 and Bi4Se3. The structure Condens. Matter Phys. 2 (2011) 31–53.
of Bi2Se3 is a stack of five-atom Se–Bi–Se–Bi–Se layers (quintuple [4] D. Kong, Y. Cui, Opportunities in chemistry and materials science for
layers, QL) whereas Bi4Se3 comprises a 1:1 ratio of QLs and bis- topological insulators and their nano structures, Nature Chem. 3 (2011)
845–849.
muth bilayers (Bi2). It is experimentally known that these two [5] X.-L. Qi, S.-C. Zhang, The quantum spin Hall effect and topological insulators,
layer types can liberally amalgamate, so thorough knowledge of Phys. Today 63 (2010) 33–38.
the inter-layer chemistry is essential. [6] M.Z. Hasan, C.L. Kane, Colloquium: topological insulators, Rev. Mod. Phys. 82
(2010) 3045.
The inter-layer binding is dominated by non-covalent interac- [7] B. Yan, S.-C. Zhang, Topological materials, Rep. Prog. Phys. 75 (2012) 096501.
tions, so description of dispersion contributions to the binding is [8] A.S. Hewitt, J. Wang, J. Boltersdorf, P.A. Maggard, D.B. Dougherty, Coexisting Bi
key. Using the exchange-hole dipole moment (XDM) dispersion and Se surface terminations of cleaved Bi2Se3 single crystals, J. Vac. Sci.
Technol. B 32 (2014) 04E103.
model combined with the B86bPBE density functional, we calcu- [9] S. Urazhdin, D. Bilc, S. Mahanti, S. Tessmer, T. Kyratsi, M. Kanatzidis, Surface
lated the structure and binding properties of both the bulk and effects in layered semiconductors Bi2Se3 and Bi2Te3, Phys. Rev. B 69 (2004)
model slabs containing combinations of QLs and Bi2 bilayers. 085313.
[10] Y. Saeed, N. Singh, U. Schwingenschl, Thickness and strain effects on the
Because of the dispersion corrections, the bulk structures are in
thermoelectric transport in nanostructured Bi2Se3, Appl. Phys. Lett. 104 (2014)
agreement with the known experimental results. 033105.
The chemical bonding properties were examined using the non- [11] Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. Hor, R.
covalent interaction (NCI) index, the electron localization function Cava, et al., Observation of a large-gap topological-insulator class with a single
Dirac cone on the surface, Nature Phys. 5 (2009) 398–402.
(ELF), and atomic charges from Bader’s Quantum Theory of Atoms [12] H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, S.-C. Zhang, Topological insulators
in Molecules (QTAIM). The NCI plots show that inter-layer binding in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface, Nature
is weaker than the intra-layer and that binding between QLs is Phys. 5 (2009) 438–442.
[13] H. Lind, S. Lidin, A general structure model for Bi–Se phases using a superspace
weaker than between a QL and Bi2. QTAIM charges and the intra- formalism, Solid State Sci. 5 (2003) 47–57.
layer distances explain this observation: the interaction between [14] T. Hirahara, G. Bihlmayer, Y. Sakamoto, M. Yamada, H. Miyazaki, S.-i. Kimura, S.
two QLs barely affects the intra-layer geometry whereas some Blügel, S. Hasegawa, Interfacing 2D and 3D topological insulators: Bi(1 1 1)
bilayer on Bi2 Te3 , Phys. Rev. Lett. 107 (2011) 166801.
charge is transferred from Bi2 to the QL, increasing the binding [15] T. Valla, H. Ji, L. Schoop, A. Weber, Z.-H. Pan, J. Sadowski, E. Vescovo, A. Fedorov,
and decreasing the Se–Bi bond lengths at the edge of the QL A. Caruso, Q. Gibson, et al., Topological semimetal in a Bi–Bi2Se3 infinitely
directly facing the bismuth bilayer. However, the changes to the adaptive superlattice phase, Phys. Rev. B 86 (2012) 241101.
[16] Q.D. Gibson, L.M. Schoop, A.P. Weber, H. Ji, S. Nadj-Perge, I.K. Drozdov, H.
electronic structure induced by Bi2 do not propagate further than Beidenkopf, J.T. Sadowski, A. Fedorov, A. Yazdani, T. Valla, R.J. Cava,
the first two atomic layers of the QL. The ELF and QTAIM charges Termination-dependent topological surface states of the natural superlattice
indicate that the nature of the intra-layer bonding is ionic, with a phase Bi4Se3, Phys. Rev. B 88 (2013) 081108.
[17] L.V. Yashine, M.R. Sánchez-Barriga, J ad Scholz, A.A. Volykhov, A.P. Sirotina, V.S.
certain covalent component, and that the charge donation from
Neudachina, M.E. Tamm, A. Varykhalov, D. Marchenko, G. Springholz, G. Bauer,
Bi2 increases the stability of the QL and decreases the bismuth A. Knop-Gericke, O. Rader, Negligible surface reactivity of topological
bilayer binding energy. insulators Bi2Se3 and Bi2Te3 towards oxygen and water, ACS Nano 7 (2013)
Consistent with the chemical bonding picture, the binding ener- 5181–5191.
[18] O. Tereshchenko, K. Kokh, V. Atuchin, K. Romanyuk, S. Makarenko, V.
gies are 6.3 kcal/mol between two QLs and 8.6 kcal/mol between a Golyashov, A. Kozhukhov, I. Prosvirin, A. Shklyaev, Stability of the (0 0 0 1)
QL and Bi2 and are independent of the presence of additional layers surface of the Bi2Se3 topological insulator, JETP Lett. 94 (2011) 465–468.
244 M.S. Christian et al. / Computational and Theoretical Chemistry 1053 (2015) 238–244

[19] V. Atuchin, V. Golyashov, K. Kokh, I. Korolkov, A. Kozhukhov, V. Kruchinin, S. [39] A.D. Becke, E.R. Johnson, A density-functional model of the dispersion
Makarenko, L. Pokrovsky, I. Prosvirin, K. Romanyuk, et al., Formation of inert interaction, J. Chem. Phys. 123 (2005) 154101.
Bi2Se3 (0 0 0 1) cleaved surface, Cryst. Growth Des. 11 (2011) 5507–5514. [40] A.D. Becke, E.R. Johnson, Exchange-hole dipole moment and the dispersion
[20] K. Park, C. De Beule, B. Partoens, The ageing effect in topological insulators: interaction: high-order dispersion coefficients, J. Chem. Phys. 124 (2006)
evolution of the surface electronic structure of Bi2Se3 upon K adsorption, New 014104.
J. Phys. 15 (2013) 113031. [41] E.R. Johnson, A.D. Becke, A post-hartree-fock model of intermolecular
[21] X. He, W. Zhou, Z.Y. Wang, Y.N. Zhang, J. Shi, R.Q. Wu, J.A. Yarmoff, Surface interactions: inclusion of higher-order corrections, J. Chem. Phys. 124 (2006)
termination of cleaved Bi2 Se3 investigated by low energy ion scattering, Phys. 174104.
Rev. Lett. 110 (2013) 156101. [42] E.R. Johnson, A.D. Becke, Van der waals interactions from the exchange hole
[22] I. Rusinov, I. Nechaev, E. Chulkov, Theoretical study of influencing factors on dipole moment: application to bio-organic benchmark systems, Chem. Phys.
the dispersion of bulk band-gap edges and the surface states in topological Lett. 432 (2006) 600–603.
insulators Bi2Te3 and Bi2Se3, J. Exp. Theor. Phys. 116 (2013) 1006–1017. [43] A.D. Becke, E.R. Johnson, A unified density-functional treatment of dynamical,
[23] W. Liu, X. Peng, X. Wei, H. Yang, G.M. Stocks, J. Zhong, Surface and substrate nondynamical, and dispersion correlations, J. Chem. Phys. 127 (2007) 124108.
induced effects on thin films of the topological insulators Bi2Se3 and Bi2Te3, [44] F.O. Kannemann, A.D. Becke, Van der Waals interactions in density-functional
Phys. Rev. B 87 (2013) 205315. theory: rare-gas diatomics, J. Chem. Theory Comput. 5 (2009) 719–727.
[24] X. Luo, M.B. Sullivan, S.Y. Quek, First-principles investigations of the atomic, [45] F.O. Kannemann, A.D. Becke, van der Waals interactions in density-functional
electronic, and thermoelectric properties of equilibrium and strained Bi2Se3 theory: intermolecular complexes, J. Chem. Theory Comput. 6 (2010) 1081–
and Bi2Te3 including van der waals interactions, Phys. Rev. B 86 (2012) 1088.
184111. [46] A. Otero-de-la Roza, E.R. Johnson, Non-covalent interactions and
[25] A.D. Becke, E.R. Johnson, Exchange-hole dipole moment and the dispersion thermochemistry using XDM-corrected hybrid and range-separated hybrid
interaction revisited, J. Chem. Phys. 127 (2007) 154108. density functionals, J. Chem. Phys. 138 (2013) 204109.
[26] A. Otero-de-la Roza, E.R. Johnson, Van der Waals interactions in solids using [47] A. Otero-de-la Roza, E.R. Johnson, Many-body dispersion interactions from the
the exchange-hole dipole moment, J. Chem. Phys. 136 (2012) 174109. exchange-hole dipole moment model, J. Chem. Phys. 138 (2013) 054103.
[27] A. Otero-de-la Roza, E.R. Johnson, A benchmark for non-covalent interactions [48] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli,
in solids, J. Chem. Phys. 137 (2012) 054103. G. Chiarotti, M. Cococcioni, I. Dabo, et al., Quantum ESPRESSO: a modular and
[28] R.F.W. Bader, Atoms in Molecules. A Quantum Theory, Oxford University Press, open-source software project for quantum simulations of materials, J. Phys.:
Oxford, 1990. Condens. Matter 21 (2009) 395502.
[29] R.F.W. Bader, A quantum-theory of molecular-structure and its applications, [49] A. Becke, On the large-gradient behavior of the density functional exchange
Chem. Rev. 91 (1991) 893–928. energy, J. Chem. Phys. 85 (1986) 7184.
[30] R. Boyd, C. Matta, The Quantum Theory of Atoms in Molecules: from Solid [50] J. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
State to DNA and Drug Design, Wiley-VCH, Weinheim, Germany, 2007. simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
[31] E. Johnson, S. Keinan, P. Mori-Sánchez, J. Contreras-García, A. Cohen, W. Yang, [51] P. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (1994) 17953.
Revealing noncovalent interactions, J. Am. Chem. Soc. 132 (2010) 6498–6506. [52] M. Yu, D.R. Trinkle, Accurate and efficient algorithm for Bader charge
[32] A. Otero-de-la Roza, E.R. Johnson, J. Contreras-García, Revealing non-covalent integration, J. Chem. Phys. 134 (2011) 064111.
interactions in solids: NCI plots revisited, Phys. Chem. Chem. Phys. 14 (2012) [53] A. Otero-de-la Roza, M.A. Blanco, A. Martín Pendás, V. Luaña, Critic: a new
12165–12172. program for the topological analysis of solid-state electron densities, Comput.
[33] A.D. Becke, K.E. Edgecombe, A simple measure of electron localization in Phys. Commun. 180 (2009) 157–166.
atomic and molecular systems, J. Chem. Phys. 92 (1990) 5397. [54] A. Otero-de-la Roza, E.R. Johnson, V. Luaña, Critic2: a program for real-space
[34] B. Silvi, A. Savin, Classification of chemical bonds based on topological analysis analysis of quantum chemical interactions in solids, Comput. Phys. Commun.
of electron localization functions, Nature 371 (1994) 683–686. 185 (2014) 1007–1018.
[35] A.J. Cohen, P. Mori-Sánchez, W. Yang, Challenges for density functional theory, [55] R. Wyckoff, Crystal Structures, Interscience Publishers, New York, 1960.
Chem. Rev. 112 (2012) 289–320. [56] S. Nakajima, Crystal structure of Bi2Te3-xSex, J. Phys. Chem. Solids 24 (1963)
[36] E.R. Johnson, I.D. Mackie, G.A. DiLabio, Dispersion interactions in density- 479.
functional theory, J. Phys. Org. Chem. 22 (2009) 1127–1135. [57] J. Zhang, Z. Peng, A. Soni, Y. Zhao, Y. Xiong, B. Peng, J. Wang, M.S. Dresselhaus,
[37] G.A. DiLabio, A. Otero-de-la Roza, Non-covalent interactions in density- Q. Xiong, Raman spectroscopy of few-quintuple layer topological insulator
functional theory, in: K.B. Lipkowitz, (Ed.), Rev. Comput. Chem., Wiley-VCH, Bi2Se3 nanoplatelets, Nano Lett. 11 (2011) 2407–2414.
Hoboken, NJ, 2014, (in press, arXiv: 1405.1771). [58] S.S. Hong, W. Kundhikanjana, J.J. Cha, K. Lai, D. Kong, S. Meister, M.A. Kelly,
[38] A.D. Becke, E.R. Johnson, Exchange-hole dipole moment and the dispersion Z.-X. Shen, Y. Cui, Ultrathin topological insulator Bi2Se3 nanoribbons exfoliated
interaction, J. Chem. Phys. 122 (2005) 154104. by atomic force microscopy, Nano Lett. 10 (2010) 3118–3122.

Anda mungkin juga menyukai