Sei sulla pagina 1di 12

First-principles investigation of the size-dependent structural stability and electronic

properties of O-vacancies at the ZnO polar and non-polar surfaces


Kin Mun Wong, S. M. Alay-e-Abbas, A. Shaukat, Yaoguo Fang, and Yong Lei
Citation: Journal of Applied Physics 113, 014304 (2013); doi: 10.1063/1.4772647
View online: http://dx.doi.org/10.1063/1.4772647
View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/113/1?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
First-principles study of vacancy-assisted impurity diffusion in ZnO
APL Mat. 2, 096101 (2014); 10.1063/1.4894195
Spatial distribution of neutral oxygen vacancies on ZnO nanowire surfaces: An investigation combining confocal
microscopy and first principles calculations
J. Appl. Phys. 114, 034901 (2013); 10.1063/1.4813517
First-principles study of electronic structures and photocatalytic activity of low-Miller-index surfaces of ZnO
J. Appl. Phys. 113, 034903 (2013); 10.1063/1.4775766
Size effects on formation energies and electronic structures of oxygen and zinc vacancies in ZnO nanowires: A
first-principles study
J. Appl. Phys. 109, 044306 (2011); 10.1063/1.3549131
Oxygen vacancies in ZnO
Appl. Phys. Lett. 87, 122102 (2005); 10.1063/1.2053360

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

JOURNAL OF APPLIED PHYSICS 113, 014304 (2013)

First-principles investigation of the size-dependent structural stability and


electronic properties of O-vacancies at the ZnO polar and non-polar surfaces
Kin Mun Wong,1,a) S. M. Alay-e-Abbas,2,3 A. Shaukat,3 Yaoguo Fang,1 and Yong Lei1,b)
1

Institut f
ur Physik & IMN MacroNanoV (ZIK), Technische Universit
at Ilmenau, Prof. Schmidt-Str. 26,
Ilmenau 98693, Germany
2
Department of Physics, GC University Faisalabad, Allama Iqbal Road, Faisalabad 38000, Pakistan
3
Department of Physics, University of Sargodha, 40100 Sargodha, Pakistan

(Received 11 November 2012; accepted 29 November 2012; published online 3 January 2013)
In this paper, all electron full-potential linearized augmented plane wave plus local orbitals method
has been used to investigate the structural and electronic properties of polar (0001) and non-polar
(10
10) surfaces of ZnO in terms of the defect formation energy (DFE), charge density, and
electronic band structure with the supercell-slab (SS) models. Our calculations support the sizedependent structural phase transformation of wurzite lattice to graphite-like structure which is a
result of the termination of hexagonal ZnO at the (0001) basal plane, when the stacking of ZnO
primitive cell along the hexagonal principle c-axis is less than 16 atomic layers of Zn and O atoms.
This structural phase transformation has been studied in terms of Coulomb energy, nature of the
bond, energy due to macroscopic electric field in the [0001] direction, and the surface to volume
ratio for the smaller SS. We show that the size-dependent phase transformation is completely
absent for surfaces with a non-basal plane termination, and the resulting structure is less stable.
Similarly, elimination of this size-dependent graphite-like structural phase transformation also
occurs on the creation of O-vacancy which is investigated in terms of Coulomb attraction at the
surface. Furthermore, the DFE at the (1010)/(1010) and (0001)/(0001) surfaces is correlated with
the slab-like structures elongation in the hexagonal a- and c-axis. Electronic structure of the neutral
O-vacancy at the (0001)/(0001) surfaces has been calculated and the effect of charge transfer
between the two sides of the polar surfaces (0001)/(0001) on the mixing of conduction band
through the 4s orbitals of the surface Zn atoms is elaborated. An insulating band structure profile
for the non-polar (1010)/(1010) surfaces and for the smaller polar (0001)/(0001) SS without
O-vacancy is also discussed. The results in this paper will be useful for the tuning of the structural
C 2013
and electronic properties of the (0001) and (1010) ZnO nanosheets by varying their size. V
American Institute of Physics. [http://dx.doi.org/10.1063/1.4772647]

I. INTRODUCTION

Zinc oxide (ZnO) is crystallographically a wurtzite


structure with the Zn and O ionic planes stacked alternatively
along the principal axis (cHex) of hexagonal symmetry. As it
is a wide band gap (3.37 eV) semiconducting oxide having
a large exciton binding energy of 60 meV at room temperature which is chemically stable, biocompatible, with high
thermal stability, and environmentally friendly, hence ZnO
has been frequently utilized in thin films and in lowdimensional nanostructures.1,2 The assemble of functional
nanostructures by various techniques,24 in particular, the
use of ZnO as a fundamental nanoscale building block has
led to the increase in the packing density of the thin films
and nanostructures for potential applications in nanosensors,5
ultra-violet lasers,6 and solar cells.7 In the process of implementing ZnO for device integration in these applications,
ZnO has been grown in many different configurations such
as nanowires, nanodots, and nanosheets (ultra thin films).8

a)

Electronic addresses: kin-mun.wong@tu-ilmenau.de and km2002wong


@yahoo.com.sg.
b)
Electronic mail: yong.lei@tu-ilmenau.de.
0021-8979/2013/113(1)/014304/11/$30.00

Growth of the ZnO nanosheets along the [0001] direction9,10 or where the nanosheets are enclosed entirely by the
(0001) facets11 have been achieved by a number of different
techniques such as the pulsed laser deposition method10 or
by a thermal decomposition and reduction process of the initial ZnS powder followed by a simple oxidation process.11 It
is also well known that the ZnO polar surfaces (0001)/(0001)
formed by repeating the wurtzite primitive unit cell along the
c-axis leads to a macroscopic electric field perpendicular to
the (0001) surface due to the non-zero dipole moment, thus,
resulting in electrostatic instability.12 In spite of this, the low
Miller indexed surfaces (1010) and (0001) of ZnO are especially important as they usually appear as the dominate surfaces in the synthesized two-dimensional (2D) nanosheets and
one-dimensional (1D) nanostructures.13 For example, the
growth of highly crystalline ZnO nanorods with the microwave assisted hydrothermal method result in a predominately hexagonal phase along the (1010) plane.14 In
addition, synthesis of uniform hexagonal prismatic ZnO
whiskers, highly crystalline hexagonal ZnO quantum dots or
particles, as well as some of the sides of the hexagonal columns of the ZnO powder particles are also primarily made
up of these (1010) facets and surfaces.1517 Conversely,

113, 014304-1

C 2013 American Institute of Physics


V

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-2

Wong et al.

according to the models of the idealized growth mechanism


of the ZnO prismatic crystals, the outlook of the crystallite is
related to the relative growth rate of the different crystal facets where the growth rate along the [0001] direction is almost
twice as compared to the other directions.15 In a recent experimental work, we had also shown that the predominate
growth direction of ZnO nanowires is along the [0001]
direction18 which agrees with other studies reported in the
literature and is shown to be energetically favourable. In
addition, some of the synthesized ZnO nanosheets are
observed to be enclosed entirely by the (0001) facets.10
Consequently, there are some stabilization mechanisms of
the oxide polar surfaces such as the creation of surface
states by the intrinsic charge transfer of negative charges
from the O-terminated to the Zn-terminate surface, adsorption of the charged species (e.g., H) as well as the modification of the surface stoichiometry from the creation of
vacancies (e.g., O-vacancies).1921
The feasibility of single-walled ZnO nanotubes had
been demonstrated with the deposition of ultra-thin ZnO
(0001) films on Ag (111) where the Zn and O atoms are
arranged in planar sheets.22 Importantly, the demonstration
that ZnO can be grown in a sheet-like form having either
(10
10) or (0001) termination or are entirely enclosed by the
(0001) surfaces10 makes it worth investigating for spintronics, electronics, and other important technological applications.23,24 This is also due to the well defined geometry of
the 2D ZnO nanosheets and their ease of assembly which
render them as useful materials for the synthesis of nanodevices. These ultra-thin ZnO nanosheets have a large surface
to volume ratio, therefore, the electrical, optical, and magnetic properties as well as their performance in device applications are intimately related and are strongly affected by the
presence of surface defects such as oxygen vacancies.25
Hence it is important to investigate the formation energy of
the oxygen vacancies in the ZnO (1010) and (0001) surfaces
with different sizes.
Due to the nanometer size of the ZnO nanosheets, there
are many difficulties related with present experimental techniques to elucidate physical properties associated with the
causes and effects of the surface defects. Moreover, the
functionalization of ZnO nanosheets requires a thorough
understanding of the electronic structure of the ZnO surfaces. In this paper, theoretical investigations involving the
first-principles full-potential linearized augmented plane
wave plus local orbitals (FP-LAPW lo) method have been
carried out for a deeper understanding of the size-dependent
structural and electronic properties of the (1010) and (0001)
ZnO surfaces with and without the surface O-vacancy.
Comparison of our results with earlier theoretical studies
carried out using different ab-initio methods and various
structural models provide unambiguous evidence of the
suitability of computationally accurate FP-LAPW lo and
the supercell-slab (SS) models. Therefore, we hope to
provide a further insight into the size-dependent structural
and electronic properties of the (1010) and (0001) ZnO
surfaces pertaining to the physical mechanisms behind the
geometrical distortions and structural phase transformation
of, especially, the (0001) SS. In addition, our results pro-

J. Appl. Phys. 113, 014304 (2013)

vide useful and better understanding in tailoring the


structural and electronic properties of ZnO surfaces or
nanosheets by controlling their size for important device
application.
II. CALCULATIONAL DETAILS AND STRUCTURAL
MODELS

The properties of all the ZnO SSs in an artificial symmetry


have been calculated using the density functional theory
(DFT) based WIEN2K computer program.26 The exchangecorrelation potential has been treated by utilizing the PBEsol
generalized gradient approximation (GGA) parametrization
scheme as this functional is known to provide good results for
solids and their surfaces as compared to hybrid DFT functional
which are 2 to 3 orders more expansive computationally.27
The SSs for imitating (0001)/(0001) polar surface have
been constructed by stacking ZnO primitive unit cells and a
along the c-axis of hexagonal latvacuum region of several A
tice. In this case we represent different lengths of SS as
1  1  2, 1  1  3, 1  1  4, and 1  1  5 (1  1  N)
which contain 8, 12, 16, and 20 atoms, respectively. In order
to verify stability of a 1  1 surface unit cell of the (0001) and
(0001) surfaces that represents correct cleavage of a hexagonal
lattice at the basal plane, we consider two possible SS which
are labeled as type-A and type-B, respectively, and both SS
belong to trigonal space group # 156 (P3m1). In case of typeA SSs, the bulk unit cell of wurtzite ZnO was stacked along
that
the c-axis followed by a vacuum region of about 10 A
resulted in the 1  1  3 type-A SS shown in Fig. 1(a). A simple stacking of the bulk unit cell along the c-axis results in the
(0001) surface being terminated with an O atom, while the
(0001) surface is terminated by a Zn atom which is contradictory to the required termination of the wurtzite structure at the
basal plane.19,20 The type-B 1  1  2 SS shown in Fig. 1(b)
can be obtained by constructing a 1  1  3 type-A SS and
then clipping off Zn and O atoms at both ends of the slab to
obtain a 1  1  2 type-B SS with the Zn (O) atom layer terminating at the basal plane along [0001] ([0001]) directions. Similar procedure was adopted for obtaining the type-B 1  1  3,
1  1  4, and 1  1  5 SSs. In Fig. 1(c), the 1  1 surface
unit cell of ZnO (0001) surface is shown (Tasker Type-III12).
The non-polar (1010)/(1010) surface terminations have
been modeled by stacking six, eight, and ten primitive unit
cells of wurtzite ZnO along the a-axis of hexagonal lattice
to decouple the
followed by a vacuum region of several A
interactions between SSs. We represent these SSs as
6  1  1, 8  1  1, and 10  1  1 (in general, N  1  1,
where N is the number of ZnO primitive unit cells stacked
along the a-axis of the hexagonal lattice) which contain 24,
32, and 40 atoms, respectively. The N  1  1 SSs obtained
in this way belong to triclinic space group # 1 (P1) and have
equal number of Zn and O atoms on the surface (Tasker
Type-I12) and consists of rows of ZnO dimmers along
the [1010]/[1010] direction as shown in Fig. 2. We
would like to stress that our N  1  1 SSs are repeated infinitely both along the b- and c-axis whereas the 1  1  N SSs
are repeated infinitely along a- and b-axis of hexagonal
lattice.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-3

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

FIG. 1. (a) Side view of the 1  1  3


type-A supercell-slab and (b) 1  1  2
type-B supercell-slab in the trigonal
P3m1 setting. d1 , d2 , and d3 are the first
three interlayer distances from the
(0001) surface. (c) The top view of
the Zn-terminated (0001) surface of
1  1  2 type-B supercell-slab showing
a 1  1 surface unit cell. In case of
1  1  N SS, the hexagonal c-axis is
parallel to the trigonal c-axis as shown
in (a) and (b). The O, Zn and surface
O-vacancy atoms are represented by the
large red, small grey, and crossed yellow
spheres, respectively.

In the FP-LAPW lo method implemented in this work,


the core states have been treated fully relativistically relaxed
in the spherical approximation, while the valence states have
been treated scalar-relativistically. For all the SS calculations,
the parameter RO x Kmax 7.0 for the plane wave basis set
is muffin-tin radius
expansion was applied, where RO 0.79 A
of O atom and Kmax is the maximal reciprocal lattice vector.
For the first Brillouin zone integration, a (9  5  1) Monkhorst-Pack28 k-point mesh has been used for the N  1  1
SSs in the P1 structure, while a (14  14  1) k-point mesh
has been used for 1  1  N SSs in the P3m1 structure. Subsequently, the k-point mesh was increased to 10  6  1 and
17  17  1 for the N  1  1 and 1  1  N SS, respectively,
as a test of the convergence where same results were obtained.
Both Kmax and the number of k-points were found to be sufficient to achieve a self-consistent minimum total energy below
0.1 mRy, and hence, our calculations are credible.

III. RESULTS AND DISCUSSIONS


 surfaces
A. Perfect polar (0001)/(0001)

In both the type-A and type-B 1  1  N SS, the Zn and


O atoms were initially located at their bulk positions which
were then fully relaxed by minimizing the atom forces to 1
mRy/a.u. The type-A and type-B 1  1  N SS shown in Fig.
3 deform along the c-axis after optimization of internal geometry. The structural data of the four type-A and type-B SS
are tabulated in Table I where it can be clearly seen that the
2

x100, where L is the length of


% volume collapse (LDLD
LD2
the SS along c-axis, DL is the length contraction, and D is
the width which is same for all 1  1  N SSs since the slabs
are infinite along a- and b-axis), in type-A SS do not follow a
trend on going from smaller structure to the larger ones
which provide a first evidence regarding the non-suitability
of type-A SS for the (0001) surfaces. On the other hand, for

FIG. 2. 6   1  1 supercell-slab in triclinic P1 setting showing the (10


10) and
(
1010) non-polar surfaces consisting of a
ZnO pair at the surface. The hexagonal
a-axis is shown along which the
O-vacancy has been studied. b.l1, b.l2, and
b.l3 are the bond lengths between surface
O atom and Zn atoms bonded to it, while
U1 and U2 are the surface OZnO bond
angles at the (10
10) and (
1010) surfaces.
The O, Zn, and surface O-vacancy atoms
are represented by the large red,
small grey, and crossed yellow spheres,
respectively.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-4

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

FIG. 3. (a) 1  1  3 type-A SS before


relaxation, (b) 1  1  3 type-A SS after
relaxation, (c) 1  1  3 type-B SS
before relaxation, and (d) 1  1  3 typeB SS after relaxation in the trigonal
P3m1 setting. The hexagonal c-axis is
also shown which is parallel to the trigonal c-axis. The O and Zn atoms are represented by the large red and small grey
spheres, respectively.

the type-B SS the contraction in length is maximum


(14.48%) in the case of 1  1  2 SS and reduces with the
increasing length of the structure which should be expected
as increase in the length of SS along the c-axis results in a
bulk like environment. The volume collapses in the
1  1  N SS are a direct consequence of the vacuum region
between successive slabs that vanishes the outwards electric
attraction for the surface atoms. We stress here again that for
the perfect type-A and type-B structures we have not used
any non-stoichiometric SSs which may result in a false stabilizing mechanism. For a complete graphical representation
of the deformation along the c-axis corresponding to the type
B 1  1  N SS before and after minimization, please refer
to Figs. S1 and S2 in the supplementary information.44
In order to provide a further insight into the stability of
different lengths of the SS, the heat of formation (Hf ) values
have been computed using29
Hf Ept =n  EZnbulk  EO:

(1)

In Eq. (1), Ept is the minimum total energies of the perfect


(stoichiometric) ZnO SS, EZnbulk is the energy of hexagonal Zn and EO EO2 =2, where the energy of gas-phase
O2 has been computed by placing two O atoms at a distance
in artificial periodic conditions in a cubic box of
of 1.21 A
and n is the number of formula units in a 1  1  N
side 10 A
SS. Our calculated value of Hf for bulk ZnO obtained using
the PBEsol27 functional is 3.14 eV which is in good agreement with theoretical as well as experimental values.29 This
benchmark test of the PBEsol functional for the bulk ZnO

verifies the suitability of this functional for the subsequent


calculation performed for any of the 1  1  N and N  1  1
SS in this work. The heat of formation values tabulated in
Table I further strengthen the suitability of type-B SS for
imitating the polar (0001) surface of ZnO.
The three interlayer spacing at the (0001) surface
depicted in Figs. 1(a) and 1(b) are designated as d1 , d2 , and
d3 for the type-A and type-B SS, respectively. From previous
theoretical accounts, it is well established that the interlayer
separations d1 and d3 contract, while d2 expand along the
[0001] direction and approach their bulk values with increasing length.19,30,31 For the case of type-A SS, d1 and d3
increase with the increasing length of the structure approach , whereas the d2 decrease to
ing the bulk value of 1.99 A
which provides further eviequal the bulk value of 0.62 A
dence that the type-A SS do not imitate a polar surface and
mimic only the local environment of wurtzite ZnO.32 In contrast, the small negative d1 values for type-B 1  1  2 and
1  1  3 SS imply a surprising slight crossover between surface layers at the Zn-terminated (0001) surface, while all the
positive d1 (for the 1  1  4 and 1  1  5 SS), d2 and d3
values indicate that there is a contraction or expansion of
interlayer spacing which is never accompanied by a crossover. Importantly, d2 values for the type-B SS approach the
), when the length of the SS increases
bulk values (1.99 A
from 1  1  2 to 1  1  5. The d1 and d3 values also
) on going from smaller
approach their bulk value (0.62 A
type-B 1  1  N SS to larger ones except for 1  1  3 SS.
This non-linear behavior hints at the competition between
bonding energy and the electrostatic energy that finally

TABLE I. Structural data for the type-A and type-B 1  1  N SS along with the heat of formation (eV) computed using the PBEsol functional. The interlayer
distances depicted in Figs. 1(a) and 1(b) and the surface ZnO bond lengths are presented.

1  1  N SS
1  1  2 Type-A
1  1  3 Type-A
1  1  4 Type-A
1  1  5 Type-A
1  1  2 Type-B
1  1  3 Type-B
1  1  4 Type-B
1  1  5 Type-B

% Volume
collapse

Interlayer
)
distance d1 (A

Interlayer
)
distance d2 (A

Interlayer
)
distance d3 (A

0.534
0.429
0.449
0.646
14.487
12.272
3.853
3.551

1.960
1.972
1.985
1.992
0.066
0.070
0.269
0.3112

0.662
0.644
0.632
0.620
2.423
2.416
2.162
2.125

1.937
1.949
1.956
1.964
0.024
0.010
0.466
0.504

)
Surface ZnO bond lengths (A
Zn-term

O-term

Hf (eV)

1.960
1.972
1.985
1.992
1.882
1.882
1.900
1.907

1.831
1.833
1.830
1.830
1.882
1.882
1.904
1.906

1.321
1.829
2.085
2.239
2.584
2.595
2.623
2.668

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-5

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

triggers a structural phase transformation after a certain


threshold length during the growth of wurtzite ZnO along
the [0001] direction. In spite of the fact that relaxed nonpolar type-B 1  1  2 and 1  1  3 SS have the same surface bond lengths (Table I), a larger stacking of the 1  1  3
SS have significant effects related to the non-linear behavior
of defect formation energy (DFE) in Table II and the band
structure profiles after the creation of O-vacancy at the Znterminated (0001) surface which are discussed in Sec. III B.
Unlike the type-A SS, the most striking feature of the
type-B SS is the difference in the structural deformation
when the atomic layers of Zn and O atoms is less than 16. In
the case of type-B 1  1  2 and 1  1  3 SS, the structures
flatten after geometry optimization (Fig. 3) resulting in a
graphite-like ZnO. As mentioned previously, the volume collapse in type-B 1  1  N SS decreases as the stacking of
bulk ZnO unit cell along the c-axis is increased from 2 to 5
(Table I) which is a direct consequence of the surface to volume ratio (D2 =LD2 1=L) that is larger in the case of the
type-B 1  1  2 and 1  1  3 SSs as compared to all other
type-A and type-B structures. Since the extreme surface Zn
and O atoms of the 1  1  2 and 1  1  3 SS have already
lost one of the four bonds due to the surface termination as
compared to bulk ZnO, a larger surface to volume ratio for
these SS ensures that they are unable to compensate for these
broken surface bonds which together with the stronger Coulombs attraction from the interior atomic layers then trigger a
collapse of surface atomic layers towards the interior of SS
resulting in flattening of the atomic layers. Since this could
have been a consequence of insufficient vacuum region
between the SS, we verified our results by varying the vacuum
for both type-A and type-B SS and
region from 7 to 13 A
found the same volume collapse for all the 1  1  N SSs.
In case of type-B 1  1  4 and 1  1  5 SS, a smaller
surface to volume ratio for these structures does compensate
for the broken surface bonds, and thus, is able to retain the
wurtzite structure for these SS. These results are consistent
with earlier theoretical reports on AlN, GaN, and ZnO,31,33
however, unlike Ref. 31 our results clearly show that the
transformation from wurtzite to graphite-like ZnO structure
occurs for <16 atomic layers (corresponding to the type-B
1  1  4 SS). Importantly, recently reported nonpolar stabilization of the ZnO (0001) surfaces in an experimental study22
revealed a flat surface morphology for unreconstructed wurtzite ZnO films having 4 or less atomic monolayers (ML). In
for a
spite of the fact that the experimental values d2 2.40 A

3.5 ML and d2 2.26 A for a 4.5 ML were obtained for sur-

face with considerable roughening,22 our computed d2 values


) and 1  1  4 (2.162 A
) SS
for type-B 1  1  2 (2.423 A
without any surface roughening provide an excellent agreement between theory and experiment of the wurtzite to
graphite-like transformation in ZnO thin films terminating at
the basal plane. The phase transformation and the resulting
depolarization of smaller 1  1  N type-B SS have important
implication regarding the electronic band structure profile of
these SS with and without O-vacancy.

B. Neutral O-vacancy at the polar (0001)/(0001)
surfaces

After confirming that type-B SSs are the correct representatives for simulating the effects of shallow O-vacancy at
the (0001) surface on the electronic properties of ZnO, we
have computed the DFE, charge density and electronic band
structures. It is very well known that the O-vacancy in ZnO
can exist either in the neutral state (Vo ) or the two charge
states (Vo or Vo 2 ).34 In this paper, the formation energy of
neutral O-vacancy have only been considered as some studies suggest that Vo is the major component in the defect
structure of ZnO and Vo is more stable as compared to
Vo .35,36 The DFE have been estimated using32
EVo Edt  Ept  RDno lo qEF ;

(2)

where Edt and Ept are the minimum total energies of the O deficient and the perfect ZnO type-B 1  1  N SS, respectively. EF and q are the Fermi energy and charge state of the
defect, respectively. Dno is the difference in the number of O
atoms in the perfect and defective SS, while lo is the chemical potential of oxygen which in two extreme cases of the
growth conditions (i.e., O-rich and O-poor condition) is
given by Eqs. (3) and (4), respectively
lopoor EO Hf ;

(3)

lorich EO:

(4)

In Table II, we present the DFE of the shallow Vo at the


(0001) surface for the type-B 1  1  N SS for both the
O-poor and O-rich conditions. Contrary to the systematic
increase of Hf in Table I, the DFE first increases on going
from 1  1  2 to 1  1  3 SS and then decreases up to
1  1  5 SS. This trend in the DFE is similar to the variation
of d3 for the type-B 1  1  N SSs reported in Table I and
provides a hint on the relationship of DFE with the ZnO

TABLE II. Vo defect formation energies for type-B 1  1  N SS.


Type-B SS defect formation energy (eV)
at the Zn-terminated (0001) surface
SS
112
113
114
115

Type-B SS defect formation energy (eV) at the


O-terminated (000
1) surface

Vo (lo-poor)

Vo (lo-rich)

Vo (lo-poor)

Vo (lo-rich)

1.053
1.111
0.864
0.801

3.636
3.705
3.487
3.468

1.052
0.572
0.232
0.143

3.635
3.166
2.854
2.810

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-6

Wong et al.

FIG. 4. The charge density plots of the relaxed type-B 1  1  N SSs along
the (1120) plane (top panel) without an O-vacancy (perfect) and (bottom
panel) with a shallow O-vacancy. (a) 1  1  2, (b) 1  1  3, (c) 1  1  4,
and (d) 1  1  5 SS.

bond lengths and the bonding nature at the (0001) surfaces.


Therefore, we predict here that the competition between the
bond energy, the Coulomb energy, and the energy originating from the macroscopic electric field of the dipole surface

J. Appl. Phys. 113, 014304 (2013)

must be responsible for the non-linear variation in the DFE


for O-vacancy at the Zn-terminated (0001) surface. On the
other hand, the smaller values of DFE for the O-terminated
(0001) surface indicates that an O-vacancy can be easily produced as compared to the (0001) surface.
From Table I, it can be easily observed that the surface
ZnO bond lengths for type-B 1  1  N SS increases on
going from 1  1  3 to 1  1  5 SS which is an indication
of stronger bond energy for the 1  1  5 SS and decrease in
the Coulombs energy. This bonding nature is supported by
charge density plots in the (1120) and (0001) planes as
shown in Figs. 4 and 5, respectively. For the charge density
plots in the (0001) plane, please refer to Fig. S3 in the supplementary information/material.44
From the above figures, the transformation of the
smaller 1  1  2 and 1  1  3 SS to graphite-like structure
is also visible. In the case of 1  1  2 and 1  1  3 SS, after
relaxation, the bond lengths of the surface Zn and O atoms
which signify a stronger Coulombs energy over
are 1.882 A
the 1  1  4 and 1  1  5 SS. Importantly, the creation of
surface O-vacancy as depicted in Fig. 1 results in the removal of the stronger Coulombs attraction at the terminated
surface and thus eliminates the graphite-like structure for the
defective SS. However, the effect of the removal of the O
atom at the (0001) surface for these SS is different. This difference in the charge density distribution at the surface of defective type-B 1  1  2 SS gives it a band structure profile
entirely different from the rest of the type-B SSs which are
presented later in Fig. 6. Finally, the smaller Coulomb
energy results in smaller values of the defect formation
energy for the larger structures. It should also be mentioned
that due to the phase transformation of the 1  1  2 and
1  1  3 SS to graphite-like, the trend for the DFE at the Zn
and O-terminated surface could be different from the
1  1  4 and 1  1  5 SS. The values computed for these
two surfaces in the case of 1  1  2 SS are the same, while
for 1  1  3 SS, the DFE value on both surfaces are different which is indicative of the fact that upon creation of vacancy the 1  1  3 SS starts to retain its wurtzite phase.
In bulk wurtzite ZnO, large charge transfer from Zn
to O atom is expected due to the large electronegativity

FIG. 5. The charge density plots of the


relaxed type-B 1  1  N SSs along the
(0001) plane (top panel) without an
O-vacancy (perfect) and (bottom panel)
with a shallow O-vacancy. (a) 1  1  2,
(b) 1  1  3, (c) 1  1  4 and (d)
1  1  5 SS.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-7

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

FIG. 6. The electronic band structure of the type-B 1  1  N SS (top panel) without, (middle panel) with an O-vacancy at the (0001) surface, respectively.
(Bottom panel) With an O-vacancy at the (000
1) surface. (a) 1  1  2, (b) 1  1  3, (c) 1  1  4, and (d) 1  1  5 SS.

differences of 1.79 on the Pauling scale between the Zn and


O atom,37 which results in the bulk ZnO having ionic and
covalent bond simultaneously with about 55% ionic character. In the middle of the charge density plots for the larger
type-B SSs as shown in Fig. 4, a mixed ionic and covalent
character is seen which is maintained at the surface region
also. At the Zn-terminated and O-terminated surfaces, charge
accumulation and charge depletion, respectively, is also
clearly seen which is consistent with other studies38 for the
(0001) and (000
1) surfaces. Fig. 5 shows the charge density
plots for the type-B SS without and with the removal of an O
atom below the Zn-terminated (0001) surface. Comparison
of the two panels shows that the charge accumulation is
larger for the case of vacancy SS. This should be expected as
the ZnZn bond is created at the surface and the charges
become highly delocalized which is characteristic of shallow
O-vacancy. Moreover, the graphite-like nature of the charge
density at the (0001) surface of the 1  1  3 SS transforms
it back to the wurtzite charge distribution with the creation
of vacancy which has already been discussed. In addition,

the (0001) and (0001) surfaces for the type-B 1  1  2 and


1  1  3 SS show almost similar charge density profile
which is a direct consequence of the flattening of the polar
surfaces that gives rise to an insulating band structures for
these two SS as shown in Fig. 6.
Fig. 6 shows the electronic band structure for the perfect
and defective type-B SS where it is clearly observed that the
1  1  2 and 1  1  3 SS are insulating with a band gap of
1.25 eV and 1.15 eV, respectively. These band gaps are
larger as compared to our computed band gap (0.75 eV) for
bulk ZnO at the gamma point (C) which can be attributed to
the quantum size confinement effect.39 The band structure
diagrams for perfect SS show no sign of the Fermi level
being shifted into the conduction or valence band for the
1  1  2 and 1  1  3 SS. This is due to the structural
transformation of the 1  1  2 and 1  1  3 SS into
graphite-like structure that eliminates the macroscopic electric field along the c-axis inside these structures. Conversely,
the band structures of the 1  1  4 and 1  1  5 SS do have
clear evidence of the electric field along the c-axis causing

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-8

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

TABLE III. The volume collapse and Hf of the relaxed N  1  1 SS after running the minimization program. The bond lengths and the bond angles at the
(1010) and (1010) surfaces.
At the (10
10) surface
N  1  1 SS
611
811
10  1  1

At the (
1010) surface

% Volume collapse

)
b.l1 (A

)
b.l2 (A

)
b.l3 (A

U1 (deg)

)
b.l1 (A

)
b.l2 (A

)
b.l3 (A

U2 (deg)

Hf (eV)

1.744
1.681
1.651

1.846
1.849
1.852

1.913
1.912
1.912

1.913
1.912
1.912

118.751
118.563
118.654

1.845
1.847
1.844

1.913
1.913
1.911

1.912
1.912
1.911

118.782
118.631
118.778

2.688
2.732
2.758

the valence band to move above the Fermi level. Moreover,


a mixing of conduction and valence band states at the C
point leading to surface metallization is observable. This surface metallization results from the charge transfer to the Znterminated surface from the O-terminated surface, which
leads to strongly dispersing 4s orbitals of the Zn atom on the
surface, hence leaving almost no occupancy of the 4s states
in the subsurface layers.19 From the band structures of the
defective 1  1  N SS, it can be seen that the removal of an
O atom below the Zn-terminated surface have different
effects on the different sizes of the slabs. It is clearly observable from the figure that the creation of vacancy results in a
band structure with no band gap for all the type B 1  1  N
SS. This could have been possibly a result of the metallic
ZnZn bond at the surface due to the creation of the
O-vacancy which enhances the surface metallization.


C. Perfect non-polar (1010)/(
1010)
surfaces

In order to understand the effects of O-vacancy at nonpolar (10


10)/(
1010) surfaces, we present the bond lengths of
the O atom with its neighbouring Zn atoms at the surface
(labelled as b:l1 , b:l2 , and b:l3 in Fig. 2) in Table III along
with other structural properties of the N  1  1 SS. From
the decreasing volume collapse and increasing Hf values, it
is evident that the SS become more and more stable with
increasing length along the hexagonal a-axis. The bond
lengths listed in Table III together with the charge density
centered at the surface Zn atom at the (1010) and (1010)

surface as shown in Fig. 7 are helpful in understanding the


bonding nature of the surface O atoms. The increasing value
of the b:l1 between the surface O and Zn indicates a tendency
of becoming more covalent as the length of the N  1  1 SS
increases. For example, the charge density contours for the
1  1  N SS which encompass both the O and Zn atom in a
8-like shape for the 6  1  1 SS (see Fig. 7) appears to split
for the 10  1  1 SS. This is a clear indication of reduced
charge transfer (reduced Coulomb energy) between the surface ZnO dimer in larger SS, hence making the bonding nature more covalent in the 10  1  1 SS. This is further
confirmed by the charge density contours around the Zn
atom in the 10  1  1 SS above the O2 atom, which appear
to be slightly more non-spherical as compared to the
8  1  1 and 6  1  1 SS. Furthermore, in the case of
6  1  1 SS, the charge density around the Zn atom above
the O2 and O11 atoms at the (1010) and (1010) surfaces,
respectively, show similar trends from which we can predict
that the DFE for these atoms will be approximately equal.
On the other hand, for the 8  1  1 (10  1  1) supercellslab comparison of charge densities around the Zn atom
above the O2 and O15 (O19) atoms clearly show different
bonding natures. The O2 atoms at the (1010) surface appears
to have stronger covalent bond as compared to the surface O
atoms at the (1010) surface from which one can infer that
the O15 (O19) atoms will have larger DFE.
From Table III, the bond lengths between the O2 and
O11, O15 or O19 atom with its neighbouring Zn atoms generally increases on going from 6  1  1 to 10  1  1 SS,

FIG. 7. The charge density plot centered


around the surface Zn atom at the (top
panel) (10
10) surface and (bottom panel)
(
1010) surface. (a) 6  1  1, (b) 8  1
 1, and (c) 10  1  1 SS.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-9

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

thus, confirming the above mentioned more ionic nature of


the 6  1  1 surface ZnO dimmer. However, the bond
lengths are not sufficient to present a clear picture of the
trend in the DFE of the prospective neutral O-vacancy at
the (10
10)/(
1010) surfaces. For this reason, we also present
the OZnO bond angle (U1/U2) in Table III which is constituted between the two outermost ZnO dimers at the (1010)/
(
1010) surfaces.
For the unrelaxed N  1  1 SS shown in Fig. 2 (and
also in our optimized ZnO bulk unit cell), these angles are
108.243 . Upon geometry optimization, the surface OZnO
bond angle tilts by an amount x1/x2 at the (1010)/(1010)
surfaces resulting in a larger U1/U2 value as compared to the
bulk. This tilt angle is a consequence of the outward movement of surface O atom and the relative movements of the
internal layers of the atoms below the (1010)/(1010) surfaces. Earlier theoretical studies carried out using local density
approximation,40 B3LYP,30 Perdew-Wang GGA,31 and all
electron Hartree-Fock method41 predict a tilt angle 3.59 ,
2.13 , 9 , and 2.48 , respectively, which are much smaller
than the strong tilt angles (11.5 6 5 ) observed in lowenergy-electron diffraction (ELEED) experiments for the
(10
10) surfaces dimers of wurtzite structures.42 Our computed tilt angles ranging between 10.2 and 10.6 again provide evidence of the excellent performance of the PBEsol
functional for surface studies.
All in all, a close inspection of the bond lengths and the
bond angles for the surface O atoms listed in Table III show
no particular trends on going from smaller SS to larger ones
as was observed in the case of the polar surfaces. Therefore,
the trendy competition between the bond energy and the
Coulomb energy as seen in Fig. 7 that controls the DFE values of the neutral surface O-vacancy presented in Sec. III D
are a complex interplay between the bond lengths and bond
angles listed in Table III.


D. Neutral O-vacancy at non-polar (1010)/(
1010)
surfaces

The DFE due to a neutral surface O-vacancy at the


(10
10) and (
1010) surface was computed by removing the O
atom highlighted by a yellow cross in Fig. 2. Our computed
DFE values for the O-vacancy (Table IV) reduce with
increasing length of the SS along the hexagonal a-axis which
are in good agreement with earlier works.39 However, in the
case of 8  1  1 and 10  1  1 SS, the DFE values at the
(
1010) surface are slightly higher than the DFE values at
TABLE IV. O-vacancy DFE for the N  1  1 SSs at the (10
10) and (
1010)
surfaces in both the O-poor and O-rich conditions.
SS defect formation
energy (eV) at the (10
10)
surface
SS
611
811
10  1  1

SS defect formation
energy (eV) at the (
1010)
surface

Vo (O-poor)

Vo (O-rich)

Vo (O-poor)

Vo (O-rich)

0.546
0.261
0.208

3.233
2.991
2.965

0.546
0.487
0.467

3.233
3.217
3.224

the (1010) surface. This is clearly due to the smaller bond


lengths and larger tilt angles (Table III) of the O15 and O19
atoms corresponding to the 8  1  1 and 10  1  1 SS,
respectively, as discussed in Sec. III C.
From a comparison between Tables II and IV, it can be
observed that the DFE for the 1  1  N SS are larger than
the N  1  1 SS which was due to the extra energy originating from the macroscopic electric field along the [0001]
direction as the centers of the positive and negative charges
do not coincide at the surface for the 1  1  N SS. The presence of this electrostatic field in the perfect type-B 1  1  N
SS results in scaling the DFE to higher values as compared
to the N  1  1 SS. However, from the charge density plots
(Figs. 5 and 7), we can easily deduce that the decreasing
value of DFE on going from smaller supercells to larger ones
is controlled by the decreasing Coulomb energy.
The charge density plots with the creation of the
O-vacancy at the (1010) and (1010) surface of the 6  1  1,
8  1  1, and 10  1  1 SS clearly shows a charge delocalization on the surfaces which is a characteristic of surface
O-vacancy. For further information about the charge density
plots and the corresponding electronic band structures of the
surface O-vacancy for the N  1  1 SS at the (1010) surface, please refer to Figs. S4 and S5 in the supplementary information/material.44 The O-vacancy DFE values obtained
using the SS model in this work is in good agreement with
the more complicated structure of similar cross-sectional
,39 which validates the suitability of the SS
width at 16.4 A
(considered in this work) that require a fraction of the extensive computational time required for the more complicated
structure.
Finally we discuss the effects of surface O-vacancy on
the electronic band structures of the N  1  1 SS which are
shown in Fig. 8. The C symmetry point direct band gap of
6  1  1, 8  1  1, and 10  1  1 SS are 0.78 eV, 0.71 and
0.68 eV, respectively. Thus, all the perfect SS show insulating nature and the band gap increase with decreasing length
of the structure in the hexagonal a-axis which is due to the
quantum confinement effect. The conduction band is predominantly made up of the empty Zn 4s-orbitals,19 while the
valence band is made up of the occupied O 2p-orbitals43 and
the band gap is created due to the repulsion between the conduction band states and the valence band states. For the
entire perfect N  1  1 SS, the Fermi level is located inside
the valence band which is due to the fact that both surfaces
of the relaxed N  1  1 SS are made up of the O atoms that
results in the shifting of the O 2p-states above the Fermi
level and therefore giving these materials (without any O-vacancy) a p-type semiconducting character.
A band gap for all the horizontal N  1  1 SS also
shows that there is no charge transfer between the (1010)
and (1010) surfaces as compared to the polar 1  1  N typeB SS, which is primarily due to the equal number of similar
atoms on both the non-polar (1010)/(1010) surface. After
creation of the O vacancy (O2 and O11, O15 or O19 atom),
one of the bands made up of Zn 4s-orbitals in the perfect
N  1  1 SS split from the conduction band and moves
downwards on the energy scale making it an occupied state.
The difference in the energy positioning of this vacancy state

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-10

Wong et al.

J. Appl. Phys. 113, 014304 (2013)

FIG. 8. The electronic band structure of the N  1  1 SS (top panel) without, (bottom panel) with O-vacancy at the (10
10) surface. (a) 6  1  1, (b) 8  1  1,
and (c) 10  1  1 SS.

with the bottom of the conduction band give rise to a band


gap. This bandgap is symmetric for both the (1010) and
(
1010) surface and decreases from 0.78 eV to 0.42 eV corresponding to the 6  1  1 and 10  1  1 SS, respectively. In
addition, the bandgap for the N  1  1 SS with the surface
O-vacancy is smaller as compared to the bandgap for the perfect N  1  1 SS. With the creation of surface O-vacancy,
the semiconducting nature of the defective N  1  1 SS
becomes n-type. The shift of the Fermi level into the conduction band is a consequence of the O-vacancy that results in
the creation of a broken ZnO bond which leads to a ZnZn
metallic bond on the surface. The free electrons as a result of
the creation of the O-vacancy do not lead to the mixing of
the conduction and valence bands as observed in the vertical
1  1  N SS.
IV. CONCLUSION


In this study, the polar (0001) and non-polar (1010)
surfaces of ZnO have been modelled as supercell-slabs to
investigate the effects of O-vacancy at these surfaces. For
the two perfect polar (0001) SSs, type-A and -B 1  1  N

SS with a non-basal and basal plane termination, respectively, a wurtzite to graphite-like phase transformation is
observed only for the more stable type-B SS. Our results
show a clear evidence that the competition between the
bonding nature, the Coulomb energy, energy originating
from the macroscopic electric field of the dipole and the
larger surface to volume ratio for the type-B 1  1  2 and
1  1  3 SS triggers a structural phase transformation from
the wurtzite structure to graphite-like structure below a
which is longer and shorter
certain threshold length (<16 A
than the length of the type-B 1  1  3 SS and 1  1  4 SS,
respectively). This phase transformation results in an insulating band structure profile at the C point for the smaller
1  1  N type-B SS. However, the creation of surface
O-vacancy results in the removal of the stronger Coulombs
attraction at the surface and thus eliminates the graphite-like
structure for defective type-B SS even when they are smaller
.
than 16 A
In the case of perfect type-B 1  1  4, 1  1  5 SS and
all defective type B 1  1  N SS, a mixing of conduction
and valence band states at the C point leads to surface
metallization which is caused by the charge transfer to the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

014304-11

Wong et al.

Zn-terminated surface from the O-terminated surface. Importantly, a good correlation is observed in the type-B SS
between the longer bond lengths which is associated with a
smaller/larger Coulomb/bond energy that yields smaller values of the defect formation energy for the larger structures.
The higher values of DFE at the polar (0001) surfaces as
compared to the non-polar surface are a consequence of the
extra energy originating from the macroscopic electric field
along the [0001] direction in the polar SS.
On the contrary, for the non-polar (1010) and (1010)
surfaces, the computed DFE values for O-vacancy of the
N  1  1 SS reduce with increasing length of the SS along
the hexagonal a-axis and appears to be dependent on the
ionic and covalent nature of the surface ZnO bonds. However, the bond lengths and the bond angles for the surface O
atoms show no particular trends on going from the smaller
SS to larger ones. Therefore, the competition between the
bond energy and the Coulomb energy that controls the DFE
values of the neutral surface O-vacancy are a complex interplay of the ZnO bond lengths and the OZnO bond angles.
Importantly, the O-vacancy DFE values, the tilt angle, and
the interlayer spacing obtained using the SS model and the
PBEsol functional in this work are in good agreement with
the computationally more demanding calculations reported
in earlier studies. Importantly, our findings open up the possibility of tuning the structural and electronic properties of
ZnO polar and non-polar surfaces by controlling their size
which is essential for technological application.
ACKNOWLEDGMENTS

Financial support from the European Research Council


Grant (ThreeDSurface), Volkswagen Stiftung, and BMBF
(ZIK: 3DNano-Device) is gratefully acknowledged.
1

F. M. Hossain, J. Nishii, S. Takagi, A. Ohtomo, T. Fukumura, H. Fujioka,


H. Ohno, H. Koinuma, and M. Kawasaki, J. Appl. Phys. 94, 7768 (2003).
Y. C. Wang, I. C. Leu, and M. H. Hon, J. Appl. Phys. 95, 1444 (2004).
3
Y. Lei, C. H. Liang, Y. C. Wu, L. D. Zhang, and Y. Q. Mao, J. Vac. Sci.
Technol. B 19, 1109 (2001).
4
Y. Lei, Z. Jiao, M. H. Wu, and G. Wilde, Adv. Eng. Mater. 9, 343 (2007).
5
K. Subannajui, C. Wongchoosuk, N. Ramgir, C. Wang, Y. Yang, A. Hartel, V. Cimalla, and M. Zacharias, J. Appl. Phys. 112, 034311 (2012).
6
M. Haupt, A. Ladenburger, R. Sauer, K. Thonke, R. Glass, W. Roos, J. P.
Spatz, H. Rauscher, S. Riethm
uller, and M. M
oller, J. Appl. Phys. 93,
6252 (2003).
7
W. W. Wenas, A. Yamada, K. Takahashi, M. Yoshino, and M. Konagai,
J. Appl. Phys. 70, 7119 (1991).
8
U. Ozgur, Ya. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Do
gan,
V. Avrutin, S.-J. Cho, and H. Morkoc, J. Appl. Phys. 98, 041301 (2005).
9
J. Zhang, X. Zhang, L. Chen, J. Xu, L. You, H. Ye, and D. Yu, Appl.
Phys. Lett. 90, 233104 (2007).
10
S. Masuda, K. Kitamura, Y. Okumura, S. Miyatake, H. Tabata, and
T. Kawai, J. Appl. Phys. 93, 1624 (2003).
2

J. Appl. Phys. 113, 014304 (2013)


11

J. Q. Hu, Y. Bando, J. H. Zhan, Y. B. Li, and T. Sekiguchi, Appl. Phys.


Lett. 83, 4414 (2003).
12
P. W. Tasker, J. Phys. C: Solid State 12, 4977 (1979).
13
B. Meyer and D. Marx, Phys. Rev. B 67, 035403 (2003).
14
N. Moloto, S. Mpelane, L. M. Sikhwivhilu, and S. S. Ray, Int. J. Photoenergy 2012, 189069 (2012).
15
J. Q. Hu, Q. Li, N. B. Wong, C. S. Lee, and S. T. Lee, Chem. Mater. 14,
1216 (2002).
16
L. Zhang, L. Yin, C. Wang, N. Lun, and Y. Qi, ACS Appl. Mater. Interfaces 2, 1769 (2010).
17
C. W
oll, Prog. Surf. Sci. 82, 55 (2007).
18
K. M. Wong, Y. Fang, A. Devaux, L. Wen, J. Huang, L. D. Cola, and
Y. Lei, Nanoscale 3, 4830 (2011).
19
A. Wander, F. Schedin, P. Steadman, A. Norris, R. McGrath, T. S.
Turner, G. Thornton, and N. M. Harrison, Phys. Rev. Lett. 86, 3811
(2001).
20
G. Kresse, O. Dulub, and U. Diebold, Phys. Rev. B 68, 245409 (2003).
21
B. Meyer, Phys. Rev. B 69, 045416 (2004).
22
C. Tusche, H. L. Meyerheim, and J. Kirschner, Phys. Rev. Lett. 99,
026102 (2007).
23
Z. Qin, Y. Huang, J. Qi, H. Li, J. Su, and Y. Zhang, Solid State Sci. 14,
155 (2012).
24
F.-B. Zheng, C.-W. Zhang, P.-J. Wang, and H.-X. Luan, J. Appl. Phys.
111, 044329 (2012).
25
L. Schmidt-Mende and J. L. MacManus-Driscoll, Mater. Today 10, 40
(2007).
26
P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz,
WIEN2K, an Augmented Plane Wave p Local Orbitals Program for Calculating Crystal Properties (Karlheinz Schwarz, Techn. Universitat Wien,
Austria, 2001).
27
J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov, G. E. Scuseria,
L. A. Constantin, X. Zhou, and K. Burke, Phys. Rev. Lett. 100, 136406
(2008).
28
H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).
29
F. Oba, A. Togo, I. Tanaka, J. Paier, and G. Kresse, Phys. Rev. B 77,
245202 (2008).
30
A. Wander and N. M. Harrison, Surf. Sci. 457, L342 (2000).
31
F. Claeyssens, C. L. Freeman, N. L. Allan, Y. Sun, M. N. R. Ashfold, and
J. H. Harding, J. Mater. Chem. 15, 139 (2005).
32
C. G. Van de Walle and J. Neugebauer, J. Appl. Phys. 95, 3851 (2004).
33
Y. Wu, G. Chen, H. Ye, Y. Zhu, and S.-H. Wei, Appl. Phys. Lett. 94,
253101 (2009).
34
S. J. Clark, J. Robertson, S. Lany, and A. Zunger, Phys. Rev. B 81, 115311
(2010).
35
L. E. Halliburton, N. C. Giles, N. Y. Garces, M. Luo, C. Xu, L. Bai, and
L. A. Boatner, Appl. Phys. Lett. 87, 172108 (2005).
36
Y. Wang, B. Yang, N. Can, and P. D. Townsend, J. Appl. Phys. 109,
053508 (2011).
37
D. R. Lide, CRC Handbook of Chemistry and Physics: A Ready-Reference
Book of Chemical and Physical Data, 89 ed. (CRC, Boca Raton, FL, 2008).
38
J. M. Carlsson, Comput. Mater. Sci. 22, 24 (2001).
39
D. Q. Fang and R. Q. Zhang, J. Appl. Phys. 109, 044306 (2011).
40
P. Schr
oer, P. Kr
uger, and J. Pollmann, Phys. Rev. B 49, 17092 (1994).
41
J. E. Jaffe, N. M. Harrison, and A. C. Hess, Phys. Rev. B 49, 11153
(1994).
42
C. B. Duke, D. E. Lessor, T. N. Horsky, G. Brandes, K. F. Canter, P. H.
Lippel, A. P. Mills, A. Paton, and Y. R. Wang, J. Vac. Sci. Technol. A 7,
2030 (1989).
43
S. E. Chamberlin, C. J. Hirschmugl, S. T. King, H. C. Poon, and D. K. Saldin, Phys. Rev. B 84, 075437 (2011).
44
See supplementary material at http://dx.doi.org/10.1063/1.4772647 for
additional figures of the type-B 1  1  N SS before and after minimization as well as further charge density plots and electronic band structures
of the surface O-vacancy for the N  1  1 SS.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
62.113.206.61 On: Mon, 08 Jun 2015 18:44:08

Potrebbero piacerti anche