You are on page 1of 7

Applied Surface Science 255 (2009) 57025708

Contents lists available at ScienceDirect

Applied Surface Science


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

Molecular dynamics study on surface structure and surface energy of


rutile TiO2 (1 1 0)
Dai-Ping Song *, Ying-Chun Liang, Ming-Jun Chen, Qing-Shun Bai
Precision Engineering Research Institute, Harbin Institute of Technology, Harbin 150001, China

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 28 August 2008
Received in revised form 1 December 2008
Accepted 23 December 2008
Available online 31 December 2008

The formula for surface energy was modied in accordance with the slab model of molecular dynamics
(MDs) simulations, and MD simulations were performed to investigate the relaxed structure and surface
energy of perfect and pit rutile TiO2(1 1 0). Simulation results indicate that the slab with a surface more
than four layers away from the xed layer expresses well the surface characteristics of rutile TiO2 (1 1 0)
surface; and the surface energy of perfect rutile TiO2 (1 1 0) surface converges to 1:801  0:001 J m2.
The study on perfect and pit slab models proves the effectiveness of the modied formula for surface
energy. Moreover, the surface energy of pit surface is higher than that of perfect surface and exhibits an
upper-concave parabolic increase and a step-like increase with increasing the number of units deleted
along [0 0 1] and [1 1 0], respectively. Therefore, in order to obtain a higher surface energy, the direction
along which atoms are cut out should be chosen in accordance with the pit sizes: [1 1 0] direction for a
small pit size and [0 0 1] direction for a big pit size; or alternatively the odd units of atoms along [1 1 0]
direction are removed.
2008 Elsevier B.V. All rights reserved.

PACS:
31.15.xv
68.35.Md
68.35.bd
61.72.Ff
Keywords:
Molecular dynamics
TiO2
Surface energy
Surface structure
Pit

1. Introduction
Ti-materials have extensive biomedical applications [1],
especially in medical implant and prosthesis [2], for their good
biocompatibility, biological responses, osseointegration, excellent
mechanical properties and corrosion resistance. In fact, there is a
layer of TiO2on Ti and Ti-alloy surfaces, which has a strong direct
effect on the success of implant [3,4].
The surface topographic characters of an implant at the micro/
nano-scale are related to the biological response of a host. The
roughness of more than 10 mm of micro-rough surface will has an
inuence on the biocompatibility, mechanical characteristics and
mechanical interlocking effect between implant and tissue. Due to
the same size as a cell or a biomacromolecule, the micro-rough
surface with a roughness ranging from 10 nm to 10 mm has
remarkable inuences on the biocompatibility and weakens the
mechanical characteristics of the interface [5]. The micro-roughness of less than 10 nm of the surface has a signicant inuence on
the interface structures of the implant. The reason is that the
defects in a crystal structure, such as vacancy, grain boundary, step

* Corresponding author. Tel.: +86 451 86413840; fax: +86 451 86415244.
E-mail address: songdp@yahoo.cn (D.-P. Song).
0169-4332/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2008.12.062

and pit, are in this size range, which are active region for
adsorption and inuence further the integration of biomolecular
and implant surface [6].
Moreover, the surface energy of biomaterials signicantly
inuences the adhesion, spreading and growth of a cell. The surface
with a higher surface energy promotes adhesion and spreading of a
cell [7,8]. For example, good spreading only occurs when surface
energy is higher than approximately 0.057 J m2[9]. Baier et al.
[10,11] also showed the relationship between critical surface
energy and cell adhesion: materials with a low surface energy
show a low cell attachment.
The rutile (1 1 0) surface is a very stable crystal face. Many
scholars [1217] have researched the surface structure of rutile
(1 1 0) by ab initio calculations, such as density functional theory
(DFT) local density approximation (LDA) [12,13], generalized
gradient approximation (GGA) [14], full-potential linear augmented plane wave (FP-LAPW) [15,16] and linear combination of
atomic orbitals (LCAO) [17]. On the other hand, Matsui and Akaogi
[18] and Kim et al. [19] have developed the molecular dynamics
(MDs) potential energy function and its parameters for TiO2 since
1991. A few scholars [20,21] studied the relaxed structure of the
TiO2 (1 1 0) surface by MDs. However, there was a discrepancy
between their theoretical [1217,2023] and experimental [24]
results.

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

5703

The surface energy of the rutile TiO2 (1 1 0) surface has been


calculated recently by rst principles [1214,25,26], but there is
few systematic calculations by MD possibly because the calculation accuracy of MD simulation is lower than that of rst principles
for small-scale models. However, the larger amount of computational work and the smaller atom amount of rst principles make it
very difcult to calculate the large-scale surface structures and
evaluate the changes in surface energies resulting from the
difference of the large-scale complex surface structures.
Therefore, we modify the formula for surface energy in
accordance with the slab model of MD simulations. Then, we
study the relaxed structures of the rutile TiO2(1 1 0) surface by MD
and investigate the relationship between surface energy and slab
thickness. Finally, we also cut out some atoms to create pits,
calculate the surface energies and relaxed structures of pit surfaces
and study the inuence of pit size on the surface energy.
Fig. 1. Rutile TiO2(1 1 0) surface (1  1): O, big red ball; Ti, small grey ball. Atom
labels follow as bridging oxygen O1, six-coordinate titanium Ti2, ve-coordinate
titanium Ti3, three-coordinate in-plane oxygen O4, O5, out-of-plane oxygen O6,
and reference atom O7.

2. Methods and surface modeling


2.1. Molecular dynamics simulation method
A survey of the literature indicated that several force elds have
been published for titanium dioxide and titanium oxide [18,19,27
32]. A detailed analysis of the different available force elds has
been published by Smith and his co-worker [31]. They concluded
that the force eld of Matsui and Akaogi [18] is the most suitable of
the available force elds for use in classical molecular dynamics
simulation. Matsui and Akaogi [18] developed the following force
elds for TiO2 in 1991.
Vij





qi q j C i C j
Ai A j  r i j
 6 f Bi B j exp
ri j
Bi B j
ri j

(1)

where the terms represent Coulomb, dispersion, and repulsion


interactions, respectively. Here r i j is the interatomic distance
between atoms i and j, and qi , Ai , Bi , and C i are the effective charge,
repulsive radius, softness parameter, and van der Waals coefcient
of atom i, respectively. Ai and Bi are relative to atom size and
compressibility. The quantity f is a standard force of
4.184 kJ A1 mol1. For many oxides, Coulomb and repulsion
interactions are necessary. Consequently, in this paper we used the
reduced form for Eq. (1), i.e. Buckingham potential together with
Coulombic interactions from partial charges
!
r i j
C i j qi q j
V i j Ai j exp
(2)
 6
ri j
ri j
ri j
where V i j is the interaction energy between atoms i and j, and the
parameters Ai j , ri j , C i j , qTi and qO are shown in Table 1[18].
MD simulations were carried out at 310 K for further study on
the interaction between TiO2 surface and biomolecules. Simulations were carried out in the canonical ensemble (i.e. NVT
ensemble), and the thermostat derived by Hoover [34] was
employed to maintain a constant temperature. The velocity Verlet
algorithm was used to calculate the atomic motions and the
particleparticle particlemesh (PPPM) [35] solver was applied for
the calculation of electrostatic interactions. A 10.0 A cutoff
distance was used for van der Waals interactions.
Table 1
Interaction parameters for TiO2.

2.2. Perfect surface models


The most stable (1 1 0) surface of rutile TiO2 was investigated in
this
p
paper. The (1 1 0) surface unit cell (1  1) has a dimension of
2a  c which corresponds to 6.497 and 2.959 A along [1 1 0] and
[0 0 1] directions, respectively [36,37]. The surface contains veand six-coordinate Ti atoms and two types of O atoms (in-plane
and bridging). The six-coordinate Ti atoms are covered by the
outermost bridging oxygen on the surface, while the vecoordinate Ti atoms are coordinated to in-plane oxygen atoms
on the surface. One TiO layer of rutile TiO2(1 1 0) surface unit cell
(1  1) is built up of symmetrical three-plane OTi2O2O unit
along [1 1 0] direction, containing 6 atoms (2 Ti and 4 O), so that
the sequence of TiO layers is OTi2O2OOTi2O2O    , as
shown in Fig. 1.
We constructed (10  18) surface slabs with the thickness of n
TiO layers based on the rutile (1 1 0) surface unit cell (1  1), as
shown in Table 2. For example, when n 13, this corresponds to a
containing 14040
dimension of 64:969 A  53:262 A  42:229 A,
atoms. Periodic boundary conditions were applied along [1 1 0] and
[0 0 1] directions of the surface, with a periodic vacuum gap of
threefold slab thickness (3  3:2484n A) in the direction normal to
the (1 1 0) surface. The simulations were carried out for 50 ps with
time step of 0.5 fs, and the simulation results of last 30 ps were
analyzed.
2.3. Pit surface models
Based on the perfect TiO2(1 1 0) surface model, we cut out some
atoms from the relaxed surface to create pit, as shown in Table 3.
Fig. 2 shows the unrelaxed pit surface model (PitL16B2S1x
(5  6  1)) in which the deleted atoms range over (5  6  1).
For the pit surface models, due to the longer time to equilibrate the
Table 2
Perfect surface slab models.
Model set ID

Number of
total layers n

Number of xed layers nfixed


1 for n 2; 2 for n 36
1 for n 25; oor n=2
for n 616
1

Interation

Ai j (kcal mol1)

ri j (A)

C i j (kcal mol1 A6)

Perfect I
Perfect II

216
216

TiTi
TiO
OO

717653.9571
391052.7442
271718.8311

0.154
0.194
0.234

120.9967
290.3920
696.9407

Perfect III

217

Notes: Atomic charges, qTi 2:196, qO 1:098.

Notes: Type of xed layers: Perfect I and II are one side xed, and Perfect III are the
central ceil(n=2)th layer xed.

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

5704
Table 3
Pit surface slab models.
Model set ID
a

PitL16B2S1x
PitL16B2S1ya
PitL16B2S1za
PitL17C1S1xb
PitL17C1S1yb
PitL17C1S1zb
PitL17C1S2xc
PitL17C1S2yc
PitL17C1S2zc

Pit size (1 1 0  0 0 1  1 1 0)

N delunit

N delunit  6  1
5  N delunit  1
5  6  N delunit
N delunit  6  1
5  N delunit  1
5  6  N delunit
N delunit  6  1
5  N delunit  1
5  6  N delunit

17
214
16
17
214
14
17
214
14

Notes:a is based on the 16-layer slab with the bottom 2 layers xed and pit locates
on the top surface.b,c is based on the 17-layer slab with the central 9th layer
xed.bPit locates on the top surface.cPits locate on the top and bottom surfaces.

surface atoms, the simulations were carried out for 200 ps, and the
results of last 50 ps were analyzed. The other conditions of MD
simulations for the pit surfaces are the same as those for the perfect
surfaces given above.
2.4. Surface energy calculations
According to the Gibbs denition [38], the surface energy
density (s ) of a solid, which corresponds to the energy variation
(per unit area) due to the creation of a surface, is given by

Enslab  nEbulk
Aslab

(3)

where Aslab is the total area of the surface considered, Enslab is the
total energy of a n-layer slab and Ebulk is the bulk energy per layer of
an innite solid.
However, the surface energy calculated in Eq. (3) is very
sensitive to the accuracy in determination of the bulk energy term
[39,40]. A small error in that term can make the calculated surface
energies diverge linearly with increasing slab thickness. To avoid
the divergence problem, we have employed the method [26,41,42]
that makes use of Eq. (3) rewritten in the following form:
Enslab Aslab s nEbulk

(4)

which implies that the bulk energy can be extracted from the slope
of a linear t of the slab_s total energy plotted versus n. This value
is subsequently used in Eq. (3).
For molecular dynamics simulations, a slab model, especially
one with defects, usually needs to contain xed layers, otherwise
the slab structure will be distorted. To calculate the effect of the
xed layers on Enslab and surface energy, we simulated the following

perfect surface slab models which will not be distorted: (i) a


sufciently large number of layers are unconstrained and other
layers are xed, totally m layers; (ii) all m layers are unconstrained.
The effect (DEfixed ) of xed layers can be estimated from the
difference between the total energies of the former and the latter.
DEfixed is relative to the type and number of xed layers (Tables 2
and 3). The effect of xed layers on Enslab and surface energy can
then be avoided by subtracting DEfixed from Enslab . Therefore, the
surface energy of slab model containing xed layers can be given
by the following modied equation:

Enslab  nEbulk  DEfixed


Aslab

(5)

For a large-scale complex surface structure, Eq. (5) can be rewritten


as

unit
EN
slab  NEbulk  DEfixed
Aslab

(6)

where N is the total number of units in slab.


For our slab models, the numerators in Eqs. (3), (5) and (6) are
the energy variation due to the creation of top and bottom surfaces,
denoted as DE. Because the structure and area of the top surface is
the same as those of the bottom surface in the perfect surface slab,
the contribution of each perfect surface to DE is 50% of the energy
variation (DEPerfM ) of the slab model with two-sided perfect
surfaces. However, for the slab model with a perfect surface on one
side and a pit surface on the other side, the contribution of the pit
surface to DE is not 50% of the energy variation (DEHybridM ) of this
slab model and the surface energy of the pit surface cannot be
directly obtained using Eqs. (5) and (6).
To calculate the surface energy of pit surface and compare DE and
s with those of other slab models, we removed the contribution of the
perfect surface to DE by transforming the slab model with one-sided
pit (PitL16B2S1x, PitL17C1S1x, PitL16B2S1y, PitL17C1S1y,
PitL16B2S1z and PitL17C1S1z) into the slab model with twosided pits (PitL16B2S1TS2x, PitL17C1S1TS2x, PitL16B2S1TS2y,
PitL17C1S1TS2y, PitL16B2S1TS2z and PitL17C1S1TS2z, respectively). 2DEHybridM  DEPerfM =2 and twice the surface area of the
pit surface are used as the energy variation DE and the surface area
Aslab of the transformed slab model, respectively. So, the surface
energy of the pit surface of slab model with one-sided pit can be
calculated using Eq. (6). The surface energy of the pit surface is the
surface energy of the slab model with two-sided pits or the
transformed slab model with two-sided pits.
3. Results and discussion
3.1. Analysis on perfect TiO2(1 1 0) surface structure

Fig. 2. Unrelaxed pit surface model PitL16B2S1x (5  6  1): O, big red ball; Ti, small
grey ball. The box with green lines represents the pit.

The perfect surface unit cell (1  1) of the rutile TiO2 (1 1 0)


contains bridging oxygen O1, six-coordinate titanium Ti2, vecoordinate titanium Ti3, three-coordinate in-plane oxygen O4 and
O5, and out-of-plane oxygen O6, as shown in Fig. 1. Supposing that
O7 are reference atoms, Fig. 3 shows the relationship between the
average displacements of (1 1 0) surface unit cells (10  18) along
[1 1 0] and the slab thickness, which is a function of slab thickness.
Although the average displacements of O1, Ti2 and O6 change
considerably, the displacements of all atoms reach certain values if
n  6 for Perfect I and n  9 for Perfect III.
We compared further the average displacements of surface
atoms in slab models Perfect I containing 616 layers with those of
other scholars works in Table 4. The average displacements of all
surface atoms are qualitatively in agreement with the experimental nding of Charlton [24] and the previous calculations [12
15,17,2023]. Moreover, our calculations coincide quantitatively

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

5705

Moreover the average displacements of the atoms of (1 1 0) surface


are less than 6e  4 A along [0 0 1]. Thus it may seem that all atoms
oscillate slightly in bulk positions along [1 1 0] and [0 0 1] except
the displacements of in-plane oxygen atoms along [1 1 0].
3.2. Surface energy of perfect TiO2(1 1 0) surfaces

Fig. 3. Calculated displacements of perfect surfaces.

with the experimental work [24] for the surface atoms except
O1 and Ti3. On the outermost surface layer, the bridging
oxygen atoms O1 and the ve-coordinate titanium atoms Ti3
relax inward, while the six-coordinate titanium atoms Ti2 and the
neighboring three-coordinate oxygen atoms O4 and O5 move
outward, which causes a rumpled appearance of the surface.
Although the average displacements of O1 and Ti3 are different
considerably from those of experimental work [24], the average
displacements of O6, O4, O5, and Ti2 along [1 1 0] are in good
agreement with those of experimental work [24]. Langer et al. [43]
reported that the bridging oxygen displaces laterally into an
asymmetric position and the TiO distances are 1.75 and 2.1 A.
However, in the experimental work [24], the bridging oxygen
locates in the center of the neighboring six-coordinate titanium
along [0 0 1], and the TiO distance is reduced to 1:71  0:07 A
mainly by relaxed inward largely (0:27 A). In our calculations, the
O1 also locates laterally in a symmetric position and the Ti2O1
average distance is 1.845 A, which is close to 1.82 and 1.83 A for a
ve-layer slab and a three-layer slab, respectively by GGA
calculations [37] and also in good agreement with 1.85 A by MD
simulation [18].
In addition, the average displacements of the three-coordinate
in-plane oxygen atoms O4 and O5 along [1 1 0] are very close to
those of some works [12,14,15,22,23], and the in-plane oxygen
atoms displace toward the neighboring ve-coordinate titanium
by 0.044 A, while the rest atoms displace by less than 1e  4 A.

Using 1828:6568 kcal mol1for Eunit


bulk , as determined from a
linear t (Eq. (4)) to the slab energies (the energy of a thinnest fourlayer slab), yields very well converged DE and s (Figs. 4(a) and 5(a),
respectively). Thus, the above values of Eunit
bulk is adopted in all
further calculations of s .
Our calculations of energy variation DE and surface energies s
versus slab thickness n are shown in Figs. 4(a) and 5(a),
respectively. For the one-sided relaxed perfect surface models
with different xed layers and the two-sided relaxed perfect
surface models (Table 2), their surface energies are converged to a
value of 1:801  0:001 J m2as the slab thickness increases, which
is very close to 1.78 J m2 from the MD simulations of Oliver et al.
[44].
Although the surface energies obtained by MD simulations
[33,44] are more than those by ab initio calculations, such as
0.89 J m2 from the average of the 5- and 6-layer LDA calculations
of Ramamoorthy et al. [12], 0.73 J m2from the 7-layer GGA
calculations of Bates et al. [14], 0.64 J m2from the average of the
9- and 10-layer PWGGA calculations of Bredow et al. [45],
0.57 J m2by GGAPW91 and 0.47 J m2by GGAPBE from the
average of the 10- and 11-layer of Kiejna et al. [26], large-scale
surface slabs with complex morphologic surfaces can be
evaluated more easily by MD simulations than by ab initio
calculations.
In addition, the calculation temperature results partly in the
difference between our calculations and rst-principles calculations
[12,14,26,45]. The theoretical results of our reference literature of
rst-principles calculations are strictly valid only at zero temperature, while our MD simulations were carried out at 310 K.
The surface energies of one-sided relaxed models are much the
same as those of two-sided relaxed models by MD simulations,
which is different from the result obtained by ab initio
calculations from Kiejna et al. [26] that the surface energies of
one-sided relaxed models are much more than those of two-sided
relaxed models. Moreover, the slab model with one-sided
relaxation is usually used as the adsorption substrate [46,47].
Therefore we will continue our study on the surface structure and
surface energy of one-sided relaxed slab model with surface
micro/nano-morphology.

Table 4
Atomic displacements for the perfect rutile TiO2 (1 1 0) surface (A).
Atoma

918b

Expc[24]

4d[20]

6e[21]

7f[15]

5g[12]

5h[14]

3i[22]

4j[23]

O1
Ti2
Ti3
O4, O5
O4, O5 (1 1 0)
O6
O7

0.032
0.134
0.085
0.055
0:044
0.038
0

0:27  0:08
0:12  0:05
0:16  0:05
0:05  0:05
0:16  0:08
0:03  0:08
0  0:08

0.03
0.22
0.19
0.10

0.00
0.04

0.11
0.02
0.26
0.0

0.0
0.0

0.02
0.23
0.17
0.03
0:05
0.02
0

0.06
0.13
0.17
0.12
0:04
0.07
0.02

0.02
0.23
0.11
0.18
0:05
0.03
0.03

0.09
0.09
0.12
0.11
0:05
0.05

0.04
0.19
0.14
0.15
0:05

a
b
c
d
e
f
g
h
i
j

Atom labels refer to Fig. 1.


This paper: MD, average displacements of 616 layers in Perfect I, S:D:  0:001 A.
Charlton: SXRD experimental results [24].
Yin: MD, four layers [20].
San Miguel et al.: MD, six layers [21].
Harrison et al.: LCAO, seven layers [15].
Ramamoorthy et al.: PWPPLDA, ve layers [12].
Bates et al.: PWGGA, ve layers [14].
Lindan et al.: PWPPGGA, three layers [22].
Maria et al.: GGAPBE, four layers [23].

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

5706

Fig. 4. Energy variations DE of rutile TiO2 surfaces: (a) perfect (1 1 0) surface slabs, (b) surface slabs with pit by cutting out atoms along [1 1 0], (c) surface slabs with pit by
cutting out atoms along [0 0 1] and (d) surface slabs with pit by cutting out atoms along [1 1 0].

To determine the contribution of each atomic layer to the


surface energy, we calculated the energetic contribution of any
atomic layer to the surface energy using the formula in Refs. [48
50]. We found that, for the perfect TiO2(1 1 0) surface, s is stored
mainly in the rst four layers, especially the two outer layers, with
86.4, 9.9, 1.9 and 1.0% of the energy, respectively.
As shown in Figs. 3 and 5(a), the changes in the atomic
displacements govern the changes in the surface energy as a
function of slab thickness, which is in agreement with the results of
Bates et al. [14].
Therefore, it can be concluded that the slab with a surface more
than four layers away from the xed layer can express well the
surface characteristics of rutile TiO2(1 1 0) surface.
3.3. Surface energy of pit TiO2(1 1 0) surfaces
In order to describe the atomic displacements in the pit surface,
according to the formula for slip vector [51], we dened the atomic
displacement vector as
sa 

k
1X
xab  Xab
k b 6 a

(7)

where k is the number of the nearest neighbors to atom a, and xab


and Xab are the vector differences of atoms a and b at current and
reference positions, respectively. The reference conguration is the
arrangement of atomic positions with unrelaxation.
The unstable atoms with the dangling bonds around the pit
result in the maximum displacement vectors. Fig. 6 shows the

displacement vector of the relaxed pit surface model PitL16B2S1x


(5  6  1) with atoms colored by jsa j, if jsa j  0:2.
As shown in Figs. 4 and 5, due to the effect of the perfect surface
on energy variation DE and surface energies s , DE and s of the slab
models with one-sided pit are lower than those of the slab models
with two-sided pits.
For the pit surfaces, we also calculated the energetic contribution of any atomic layer to the surface energy. We found that s is
stored mainly in the rst two layers. For example, for the pit
surface PitL16B2S1x (5  6  1), the energetic contribution of the
rst two layers is 97.9 and 1.8% of the energy, respectively.
It can be seen from Figs. 4 and 5 that the energy variations and
surface energies of the pit surfaces are higher than those of the
perfect surfaces. The reason is that the pits created by removing
some atoms from the perfect surfaces weaken the interaction of
the atoms on the interface of the pits and many unstable atoms
remain, and after relaxation these atoms offset their initial
positions as shown in Fig. 6, which lead to a signicant increase
in the surface energy. Furthermore, according to the theory of solid
surface physics [52], the closer the arrangement of atoms is, the
stronger the interaction of atoms is and the lower the surface
energy is, which is in agreement with our calculations.
Figs. 4(bd) and 5(bd) show the relationship between DE and
s of pit surfaces and the number (N delunit ) of units deleted along
[1 1 0], [0 0 1] and [1 1 0]. The energy variations and surface
energies of the surfaces with pit by cutting out atoms along [1 1 0]
change slightly along a middle-convex parabola, and the changes
in their surface energies are less than about 0.1 J m2. When the
number of units deleted along [0 0 1] is 6 and that along [1 1 0] is 1,

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

5707

Fig. 5. Surface energies s of rutile TiO2 surfaces: (a) perfect (1 1 0) surface slabs, (b) surface slabs with pit by cutting out atoms along [1 1 0], (c) surface slabs with pit by cutting
out atoms along [0 0 1] and (d) surface slabs with pit by cutting out atoms along [1 1 0].

the quadratic regressions of N delunit values along [1 1 0] with DE


and s values of the pit surfaces are shown as the solid lines in
Figs. 4(b) and 5(b), respectively.
The energy variations and surface energies of the surfaces with
pit by cutting out atoms along [0 0 1] exhibit an upper-concave
parabolic increase with the increasing number of units deleted.
When the number of units deleted along [1 1 0] is 5 and that along
[1 1 0] is 1, the quadratic regressions of N delunit values along [0 0 1]
with DE and s values of the pit surfaces are shown as the solid lines
in Figs. 4(c) and 5(c), respectively.
The energy variations DE of the surfaces with pit by cutting out
atoms along [1 1 0] is linearly relative to the number of units
deleted. When the number of units deleted along [1 1 0] is 5 and
that along [0 0 1] is 6, the linear regression of Ndelunit values along

[1 1 0] with DE values is shown as the solid line in Fig. 4(d).


However, as their surface areas increases signicantly, their
surface energies s show a step-like increase with the increasing
number of units deleted. When N delunit is odd, s increases
signicantly. The means of s at Ndelunit are shown as the solid
line in Fig. 5(d). If the number of deleted atoms is more than 180
atoms as shown in Fig. 5, the surface energy of the pit surface from
which the atoms are removed along [0 0 1] is higher than that
along [1 1 0]. In order to obtain a higher surface energy, the
direction along which atoms are cut out should be chosen in
accordance with the pit sizes: [1 1 0] direction for a small pit size
and [0 0 1] direction for a big pit size. We can also remove the odd
units of atoms along [1 1 0] direction. Because the pit surface has
higher surface energy and more adsorption sites than the perfect
surface, one can design the surface structure of implant using
Figs. 4 and 5, and the regression equations therein, and investigate
further the effect of surface structure and surface energy on
biomolecular adsorption on Ti-materials.

4. Conclusions
We modied the formula used for calculation of surface energy
in accordance with the slab model of MD simulation and studied
the surface energy and relaxed structure of perfect and pit rutile
TiO2(1 1 0) through MD simulation. The following are our ndings:

Fig. 6. Displacement vector of pit surface PitL16B2S1x (5  6  1) with atoms


colored by jsa j.

1. The outermost surface bridging oxygens and ve-coordinate


titaniums relax inward, whereas the six-coordinate titaniums
and the in-plane oxygens move outward. The slab with a surface

5708

D.-P. Song et al. / Applied Surface Science 255 (2009) 57025708

more than four TiO layers away from the xed layer expresses
well the surface characteristics of rutile TiO2 (1 1 0) surface. To
reduce the amount of the computational work and improve
computational efciency, it is feasible to use the slab model with
one-sided relaxation as the surface model. The unstable atoms
with the dangling bonds around the pit in the pit surface have
the maximum displacement vectors. Larger pit size generally
results in an increase in the region of atoms with large
displacement vectors.
2. The study on perfect and pit slab models proves the effectiveness of the modied formula for surface energy. The Surface
energy of the perfect rutile TiO2 (1 1 0) surface converges to
1:801  0:001 J m2.
3. The surface energy of the pit surface is higher than that of the
perfect surface and exhibits an upper-concave parabolic
increase and a step-like increase with increasing the number
of units deleted along [0 0 1] and [1 1 0], respectively. The
surface energy of the surface with pit by cutting out atoms along
[0 0 1] or [1 1 0] changes more signicantly than that along
[1 1 0]. Moreover, if the number of deleted atoms is over a
certain value (180 atoms in this paper), the surface energy of the
pit surface from which the atoms are cut out along [0 0 1] is
higher than that along [1 1 0]. Thus, in order to obtain a higher
surface energy, the direction along which atoms are cut out
should be chosen in accordance with the pit sizes: [1 1 0]
direction for a small pit size and [0 0 1] direction for a big pit
size; or alternatively the odd units of atoms along [1 1 0]
direction are removed.
It can therefore be concluded that the modied formula for
surface energy and the displacement vector can be used to
calculate the surface energy of the large-scale complex structures
and evaluate these surface structures by MD, respectively, which
will be aid to analyze further the effect of surface micro/nanomorphology of the TiO2 covering titanium implant surfaces on cell
adhesion and protein adsorption.
Acknowledgments
This work is funded by the National Natural Science Foundation
of China (No. 50675050) and the Multidiscipline Scientic
Research Foundation of Harbin Institute of Technology (No. HIT.
MD 2003. 10).
References
[1] C. Leyens, M. Peters, Translated by Z.H. Chen et al. Titanium and Titanium Alloys,
Chemical Industry Press, Beijing (2005).
[2] U. Diebold, Surf. Sci. Rep. 48 (2003) 53.
[3] A. Hunter, C.W. Archer, P.S. Walker, G.W. Blunn, Biomaterials 16 (1995) 287.
[4] B.D. Boyan, T.W. Hummert, D.D. Den, Z. Schwartz, Biomaterials 17 (1996) 137.
[5] B. Kasemo, J. Prosth. Dent. 49 (1983) 832.
[6] B. Kasemo, J. Lausmaa, Material selection surface characteristics and chemical
processes at implant surface, in: P.I. Branemark, G.A. Zarb, T. Albrektsson (Eds.),

[7]
[8]
[9]
[10]

[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]

Tissue Integrated Prostheses: Osseointergtation in Clinical Dentistry, Quintessence Publishing Co. Inc., Chicago, 1985, p. 99.
T.G. van Kooten, J.M. Schakenraad, H.C. van der Mei, H.J. Busscher, Biomaterials 13
(1992) 897.
S.A. Redey, S. Razzouk, C. Rey, D. Bernache-Assollant, G. Leroy, M. Nardin, G.
Cournout, J. Biomed. Mater. Res. 45 (1999) 140.
J.M. Schakenraad, H.J. Busscher, C.R.H. Wildevuur, J. Arends, J. Biomed. Mater. Res.
20 (1986) 773.
R.E. Baier, Surface properties inuencing biological adhesion, in: R.S. Mansly
(Ed.), Adhesion in Biological Systems, Academic Press, New York, 1970 , pp.
1548.
R.E. Baier, A.E. Meyer, J.R. Natiella, R.R. Natiella, J.M. Carter, J. Biomed. Mater. Res.
18 (1984) 327.
M. Ramamoorthy, D. Vanderbilt, R.D. King-Smith, Phys. Rev. B 49 (1994) 16721.
M. Ramamoorthy, R.D. King-Smith, D. Vanderbilt, Phys. Rev. B 49 (1994) 7709.
S.P. Bates, G. Kresse, M.J. Gillan, Surf. Sci. 385 (1997) 386.
N.M. Harrison, X.G. Wang, J. Muscat, M. Schefet, Faraday Discuss. Chem. Soc. 144
(1999) 305.
D. Vogtenhuber, R. Podlouchy, A. Nechel, S.G. Steinemann, A.J. Freeman, Phys. Rev.
B 49 (1994) 2099.
P. Reinhardt, B.A. Heb, Phys. Rev. B 50 (1994) 12015.
M. Matsui, M. Akaogi, Mol. Simul. 6 (1991) 239.
D.W. Kim, N. Enomoto, Z. Nakagawa, K. Kawamura, J. Am. Ceram. Soc. 79 (1996)
1095.
X. Yin, R. Miura, A. Endou, I. Gunji, R. Yamauchi, M. Kubo, A. Stirling, A. Fahmi, A.
Miyamoto, Appl. Surf. Sci. 119 (1997) 199.
M.A. San Miguel, C.J. Calzado, J.F. Sanz, Surf. Sci. 409 (1998) 92.
P.J.D. Lindan, N.M. Harrison, M.J. Gillan, J.A. While, Phys. Rev. B 55 (1997) 15919.
L.S. Maria, Y.G. Andrey, L.S. Alexander, J. Phys. Chem. B 110 (2006) 4853.
G. Charlton, Phys. Rev. Lett. 78 (1997) 495.
J. Goniakowski, J.M. Holender, L.N. Kantorovich, M.J. Gillan, Phys. Rev. Lett. 53
(1996) 957.
A. Kiejna, T. Pabisiak, S.W. Gao, J. Phys.: Condens. Matter 18 (2006) 4207.
M. Mostoller, J.C. Wang, Phys. Rev. B 32 (1985) 6773.
V. Swamy, J.D. Gale, Phys. Rev. B 62 (2000) 5406.
S. Ogata, H. Iyetomi, K. Tsuruta, F. Shimojo, R.K. Kalia, A. Nakano, P. Vashishta, J.
Appl. Phys. 86 (1999) 3036.
H. le Roux, L. Glasser, J. Mater. Chem. 7 (1997) 843.
D.R. Collins, W. Smith, Council for the Central Laboratory of Research Councils
(1996).
F.H. Streitz, J.W. Mintmire, J. Adhes. Sci. Technol. 8 (1994) 853.
P.K. Naicker, P.T. Cummings, H. Zhang, J.F. Baneld, Phys. Phys. Chem. B 109
(2005) 15243.
W.G. Hoover, Phys. Rev. A 31 (1985) 1695.
H. Eastwood, Computer Simulation Using Particles, Adam Hilger, NY, 1989.
S.C. Abrahams, J.L. Bernstein, J. Chem. Phys. 55 (1971) 3206.
A.V. Bandura, J.D. Kubicki, J. Phys. Chem. B 107 (2003) 11072.
J.W. Gibbs, Scientic Papers, Thermodynamics, vol. 1, Dover Publictions, New
York, 1926, pp. 55.
J.C. Boettger, Phys. Rev. B 49 (1994) 16798.
V. Fiorentini, M. Methfessel, J. Phys.: Condens. Matter 8 (1996) 6525.
J.C. Boettger, Phys. Rev. B 53 (1996) 13133.
J.G. Gay, J.R. Smith, R. Richter, F.J. Arlinghaus, R.H. Wagoner, J. Vac. Sci. Technol. A 2
(1984) 931.
W. Langel, Surf. Sci. 496 (2002) 141.
P.M. Oliver, G.W. Watson, E.T. Kelsey, S.C. Parker, J. Mater. Chem. 7 (1997) 563.
T. Bredow, L. Giordano, F. Cinquini, G. Pacchioni, Phys. Rev. B 70 (2004) 035419.
J.R.B. Gomes, J.P. Prates Ramalho, Phys. Rev. B 71 (2005) 235421.
M. Predota, A.V. Bandura, P.T. Cummings, J.D. Kubicki, D.J. Wesolowski, A.A.
Chialvo, M.L. Machesky, J. Phys. Chem. B 108 (2004) 12049.
O. Politano, S. Garruchet, J.M. Salazar, Mat. Sci. Eng. A 387389 (2004) 749.
S. Garruchet, O. Politano, J.M. Salazar, T. Montesin, Surf. Sci. 586 (2005) 15.
S. Garruchet, O. Politano, J.M. Salazar, A. Hasnaoui, T. Montesin, Appl. Surf. Sci. 252
(2006) 5384.
J.A. Zimmerman, C.L. Kelchner, P.A. Klein, J.C. Hamilton, S.M. Foiles, Phys. Rev. Lett.
87 (2001) 165507.
L.-S. Wen, Physics Basis of Solid Materials Interface Research, Science Press,
Beijing, 1991, p. 21.

You might also like