Sei sulla pagina 1di 5

China Particuology 5 (2007) 242246

Atomic resolution in noncontact AFM by


probing cantilever frequency shifts
Hong Yong Xie
Department of Environmental Engineering, Shanghai Second Polytechnic University,
2360 Jin Hai Road, Pudong Shanghai 201209, China
Received 13 July 2005; accepted 30 September 2006

Abstract
Rutile TiO2 (0 0 1) quantum dots (or nano-marks) in different shapes were used to imitate uncleaved material surfaces or materials with rough
surfaces. By numerical integration of the equation of motion of cantilever for silicon tip scanning along the [1 1 0] direction over the rutile TiO2
(0 0 1) quantum dots in ultra high vacuum (UHV), scanning routes were explored to achieve atomic resolution from frequency shift image. The
tipsurface interaction forces were calculated from LennardJones (12-6) potential by the Hamaker summation method. The calculated results
showed that atomic resolution could be achieved by frequency shift image for TiO2 (0 0 1) surfaces of rhombohedral quantum dot scanning in a
vertical route, and spherical cap quantum dot scanning in a superposition route.
2007 Chinese Society of Particuology and Institute of Process Engineering, Chinese Academy of Sciences. Published by Elsevier B.V.
All rights reserved.
Keywords: Nc-AFM; Frequency shift image; Atomic resolution; Quantum dots; Uncleaved material surface

1. Introduction
An atomic-force microscope (AFM) can operate in tapping
mode or noncontact mode. The tapping mode was introduced
in 1993 and has become a standard tool for investigating surfaces of soft materials, specially, for model block copolymer
materials (Wang, Song, Li, & Shen, 2003). Since achieving
true atomic resolution with noncontact atomic-force microscopy
(nc-AFM) using frequency modulation technique (Giessibl,
1995), nc-AFM has become one of the most promising tools
for surface analysis on atomic scale for cleaved surface of
various types of materials of Si (1 1 1)-7 7 (Kitamura &
Iwatsuki, 1995), Si (1 0 0)-2 1 (Sugawara, Ohta, Ueyama,
& Morita, 1995), InP (1 1 0) (Schwarz, Allers, Schwarz, &
Wiesendanger, 1999; Ueyama, Ohta, Sugawara, & Morita,
1995), GaAs (1 1 0) (Morita, Abe, Yokoyama, & Sugawara,
2000), InAs (1 1 0) (Foster, Barth, Schluger, & Reichling, 2001),
and ionic crystals, as well as defects on these surfaces (Morita,
Sugawara, Yokoyama, & Uchihashi, 2000). Experiments show
that frequency shift in nc-AFM increases abruptly, e.g., in
a discontinuous manner, or quickly as the closest distance

E-mail address: hyxie@eed.sspu.cn.

approaching the sample surface, in achieving true atomic resolutions (Morita, Sugawara, et al., 2000; Uchihashi et al., 1999).
For understanding nc-AFM and achieving atomic resolution
in surface analysis and characterization, tipsurface interaction
has been simulated by first-principal electronic state calculations (Sasaki, Watanabe, Aizawa, & Tsukada, 2001), and by
the density-function theory (DFT) in plane-wave pseudopotential formalism with generalized gradient approximation (Stich,
Dieska, & Perez, 2002). The interactions have also been modeled from van der Waals interactions (Giessibl, 1997; Sasaki &
Tsukada, 1999; Xie, Sun, & Ning, 2005). The cantilever dynamics has been approached by the classical perturbation theory
(Giessibl, 1997; Stich, Tobik, Perez, Terakura, & Ke, 2000), and
by forced vibration equations (Nunes, Zanette, Caride, Prioli, &
Rivas, 2003; Sasaki & Tsukada, 1999; Xie, 2005). The firstprincipal electronic state calculations and the DFT theory in
plane-wave pseudopotential formalism have the advantage of
no empirical constants and surface relaxations are embodied
in the interaction simulations, which have been shown to be
important for image corrugations (Ke, Uda, & Terakura, 2002).
The van der Waals interactions based on LennardJones (12-6)
potential by Hamaker summation method, however, have empirical constants, but are simple in mathematical descriptions, and
the theory could produce atomic resolution by frequency shift

1672-2515/$ see front matter 2007 Chinese Society of Particuology and Institute of Process Engineering, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved.

doi:10.1016/j.cpart.2006.09.001

H.Y. Xie / China Particuology 5 (2007) 242246

image for Si-tip over Si (1 1 1)-7 7 surface (Sasaki & Tsukada,


1999) and for Si-tip over cleaved rutile TiO2 (0 0 1) surface (Xie,
2005).
For un-cleaved material surfaces or materials with rough
surfaces, topographic images could well be acquired by AFM
with colloidal probes (Jones, Pollock, Geldart, & Verlinden,
2003). For some applications such as catalysis, surface properties such as morphology, structure, etc., will play an important
role. Achieving atomic resolution for such surfaces would be
highly significant both in science and in practical applications.
In this paper, rutile TiO2 (0 0 1) quantum dots (or nano-marks)
in different shapes were used to imitate un-cleaved material
surfaces or materials with rough surfaces. By numerical integration of the equation of motion of cantilever for silicon tip
scanning along the [1 1 0] direction over the rutile TiO2 (0 0 1)
quantum dots in ultra high vacuum (UHV), scanning routes
have been explored to achieve atomic resolution from frequency
shift image for rutile TiO2 (0 0 1) quantum dots in different
shapes. The tipsurface interaction forces were calculated from
the Hamaker method by the summation of the tip with the Ti4+
and O2 ions on rutile TiO2 (0 0 1) surface (Xie et al., 2005),
and LennardJones (12-6) potential was used for the interactions
between the tip and the Ti4+ and O2 ions.
2. Equation of motion
A model of nc-AFM is shown in Fig. 1, where z0 and u0
are the equilibrium positions of the tip and the cantilever base,
respectively. The equation of motion of the cantilever tip is
m

d2 z m0 dz
+
+ k(z u) = F + mg
dt 2
Q dt

(1)

where m is the cantilevers mass, Q its Q-factor, k its zcomponent spring constant, F
the interaction force between
tip and sample surface, 0 (= k/m) the angular resonance
frequency of the free cantilever, and u is the position of the
cantilever base which is excited with angular frequency and
amplitude a.
The equation of motion of the cantilever can be normalized
to
F (z0 + a) F (z0 )
+ + = cos +
,
ak
2

0
, = 02 , = t,
z = z0 + a, =
Q

(2)

243

with the initial conditions = 0, = 0 and = 0. The frequency


shift of the cantilever is obtained from the classical perturbation
theory as (Giessibl, 1997)
a0
 = 2 F (z0 + a)
(3)
kA0
where 0 (=0 /2) is the resonance frequency of the free cantilever, and A0 is the amplitude of free cantilever under excitation
at the cantilever base, which is obtained from Eq. (1) by setting
the interaction force between the tip and sample surface F = 0.
3. Tipsurface interaction forces
Both theoretical (Stich et al., 2002) and experimental (Morita,
Abe, et al., 2000; Morita, Sugawara, et al., 2000) work have
shown that atomic resolution for both semiconductor and metal
surfaces is primarily mediated by short range chemical-type
interactions between tip and surface. Moreover, the influence
of long-range electrostatic forces has been demonstrated to play
an insignificant role by applying a bias voltage (Ke et al., 2002;
Morita, Sugawara, et al., 2000). Therefore, LennardJones (126) potential is used for the interactions between tip and the Ti4+
ions and the O2 ions on rutile TiO2 (0 0 1) surface, and the
Hamaker summation method is used to calculate the interactions between silicon tip with rutile TiO2 (0 0 1) surface (Xie et
al., 2005).
In general, parameters in LennardJones potential for an ion
would be different if it is in a different electronic environment,
and this becomes more significant when surface relaxations
occur. In this study parameters in LennardJones potential are
taken to be constant and surface relaxations are not considered.
LennardJones radius parameters O2 Si and Ti4+ Si were
estimated from the radius of O2 ion, the covalent radius of Si,
and the bond length of T4+ O2 in rutile TiO2 by the empirical
and 1.74 A,

combining law of AB = 1/2( A + B ) to be 2.58 A


respectively (Xie et al., 2005). LennardJones well depth was
estimated by Hamaker constant, London van der Walls constant,
and Berthelots rule (Xie, 1997) to be 0.001 eV and 0.11 eV for
silicon tip-O2 ion and silicon tip-Ti4+ ion, respectively, with
a modification for the well depth SiO2 that the potentials of
SiTi4+ and SiO2 have similar values in the repulsive regime
(Xie et al., 2005).
The calculation results show that the Hamaker summation
method gives the periodical nature of the tipsurface force with
abrupt peaks at both Ti4+ and O2 ion positions on flat surface as
shown in Fig. 2. But the peaks deviate slightly from Ti4+ and O2
ion positions for an inclined surface, and the more inclined, the
greater the deviation as shown in Fig. 2. The calculation results
also show that the number of unit cells on rutile TiO2 (0 0 1)
surface has little effect on the tipsurface interactions as can be
seen from Fig. 3. In Figs. 2 and 3, angle is the angle formed
by an inclined surface and a horizontal plane.
4. Frequency shift

Fig. 1. A model of nc-AFM by forced vibration at the cantilever base.

Eq. (2) of the motion of cantilever is integrated numerically by the 4th order RungeKutta method at tip radius of

244

H.Y. Xie / China Particuology 5 (2007) 242246

Fig. 2. Interaction force calculated between Si-tip and rutile TiO2 (0 0 1) surface
(the
(flat and inclined) along [1 1 0] direction with the tipsurface distance of 1 A
small bold points stand for Ti4+ ions and the large bold points stand for O2
ions).

0 = 150 kHz, k = 35 N/m, Q-factor = 38000, a = 10 A,

100 A,
and 0 / = 1.5, which has been used in achieving atomic resolution for cleaved surfaces (Morita, Abe, et al., 2000; Morita,
Sugawara, et al., 2000; Uchihashi et al., 1999). The tip is positioned vertically above an O2 ion or a Ti4+ ion. Displacement
versus time is calculated from Eq. (2) and frequency shift is calculated from Eq. (3) in a time period when the average frequency
shift approaches constancy.
The calculation results show that the frequency shift increases
slowly as the equilibrium position z0 decreases for larger z0
values, and increases dramatically as z0 approaches the sample surface, e.g., in a discontinuous manner as shown in Fig. 4,
and then the frequency shift levels off as z0 decreases further,
as shown in Fig. 4. The change of frequency shift with tip
equilibrium position z0 is qualitatively in agreement with those
measured on silicon tip with Zn-doped p-GaAs (1 1 0) cleaved
surface (Morita, Abe, et al., 2000) and with Si (1 1 1)-7 7
leaved surface (Morita, Sugawara, et al., 2000), and on oxidized silicon tip with Si (1 1 1)-7 7 cleaved surface (Morita,
Sugawara, et al., 2000). The contact point is defined as the point
where the frequency shift starts to level off. The insert on the
top right in Fig. 4 is the closest distance of tip to surface versus
tip equilibrium position z0 . It shows that the closest distance of

Fig. 3. Effects of unit cell number on the interaction force between Si-tip and
rutile TiO2 (0 0 1) surface along [1 1 0] direction with the tipsurface distance
(the small bold points stand for Ti4+ ions and the large bold points stand
of 1 A
for O2 ions).

Fig. 4. Frequency shift vs. tip equilibrium position z0 (tip over an O2 ion,
/0 = 1.5).

the tip to surface decreases linearly as z0 decreases down to a


point and then is almost constant after that point. This point is
equivalent to the contact point from the frequency shift curve
(Xie, 2005).
The calculation results in Fig. 4 show that the frequency shift
curve is only slightly affected as the number of unit cells on TiO2
(0 0 1) surface increases in the calculations for the tipsurface
interactions, even up to 441, as the number of unit cells has but
little effect on the tipsurface interactions.
5. Frequency shift image
When the tip scans along the [1 1 0] direction on rutile TiO2
(0 0 1) surface, the tip will pass alternately the Ti4+ and O2 ions
on the surface. If the scanning speed is v, within the time interval
t the tip travels through a distance interval y, then t = y/v
and the normalized time  is then  = t = 2y/v.
Within this time interval, Eq. (2) is applied to calculate the
motion of the cantilever tip and the tip displacement could be
obtained by successive numerical integration of Eq. (2) from the
beginning, and the frequency shift could then be calculated from
Eq. (3). In this calculation, the distance interval y was taken to
and 1.2855 A
between Ti4+ and O2 ions and between
be 0.98 A
2
two O ions, respectively.
In the rhombohedral quantum dot of rutile TiO2 (0 0 1) surface illustrated in Fig. 5, scanning is carried out along a vertical

Fig. 5. Illustration of a rhombohedral quantum dot of rutile TiO2 (0 0 1) surface


(scanning along [1 1 0] direction; the small bold points stand for Ti4+ ions and
the large bold points stand for O2 ions).

H.Y. Xie / China Particuology 5 (2007) 242246

Fig. 6. Frequency shift vs. scanning distance along the [1 1 0] direction on a

rhombohedral quantum dot of rutile TiO2 (0 0 1) surface (scanning speed 30 A/s,


0 / = 1.5, scanning in vertical route; the small bold points stand
z0 = 30.18 A,
for Ti4+ ions and the large bold points stand for O2 ions).

route, i.e., along the [1 1 0] direction, with a constant distance


between tip and surface. Calculation results (Liu, 2005) show
that the frequency shift along the [1 1 0] direction has a periodical nature with peaks at Ti4+ and O2 ion positions on the
flat surface, but it deviates from Ti4+ and O2 ion positions for
inclined surfaces as shown in Fig. 6, which is in accordance with
the periodical nature of the interaction force (Fig. 2). Again, the
calculation results show that the frequency shift image is not
affected by the number of unit cells on the TiO2 (0 0 1) surface used in the calculations for the tipsurface interactions as
illustrated by the insert in Fig. 6.
A spherical cap quantum dot of rutile TiO2 (0 0 1) surface is
illustrated in Fig. 7, showing the [1 1 0] direction on its equator. In this figure, the horizontal distance is the distance in the
projected plane parallel to the [1 1 0] plane on the equator with
origin at the center of the spherical cap, while the vertical distance is the distance in the projected plane normal to the [1 1 0]
plane on the equator with origin at the center of the spherical cap.
The vertical route represents the distance between the tip and the
surface is kept constant vertically in scanning, while the radial
route represents the distance between the tip and the surface is
kept constant radially in scanning. The superposition route is
the superposition of the vertical route and a radial route. It can
be seen from Figs. 8 and 9 that (Liu, 2005), in the superposition route, the tipsurface force is regularly in periodical nature

245

Fig. 8. Interaction force between Si-tip and rutile TiO2 (0 0 1) surface of a spherical cap quantum dot (scanning along [1 1 0] direction at the tipsurface distance
the small bold points stand for Ti4+ ions and the large bold points stand
of 1 A;
for O2 ions).

Fig. 9. Frequency shift vs. scanning distance along the [1 1 0] direction on a

spherical cap quantum dot of rutile TiO2 (0 0 1) surface (scanning speed 30 A/s,
0 / = 1.5, r0 = 10 A,
r = 10.5 A
in the superposition route; the
z0 = 30.08 A,
small bold points stand for Ti4+ ions and the large bold blue points stand for
O2 ions).

with peaks appearing abruptly at both Ti4+ and O2 ion positions (Fig. 8) and the frequency shift along the [1 1 0] direction
has periodical nature with peaks at Ti4+ and O2 ion positions
(Fig. 9).
6. Conclusion remarks
Based on calculation results of frequency shift image in ncAFM for silicon tip scanning over rutile TiO2 (0 0 1) surfaces
of quantum dots under ultra high vacuum (UHV) conditions, it
can be concluded that scanning route is vital to achieve atomic
resolution by frequency shift image. For un-cleaved surfaces or
materials with rough surfaces, after topographic images have
been acquired by AFM with colloidal probes, atomic resolution
could be achieved experimentally by a proper scanning route,
and such experimental attempts would be significant both in
materials sciences and in practical applications.
Acknowledgments

Fig. 7. Illustration of a spherical cap quantum dot of rutile TiO2 (0 0 1) surface


(scanning along [1 1 0] direction; the small bold points stand for Ti4+ ions and
the large bold points stand for O2 ions).

The author is grateful for a research grant from Material


Science and Engineering Fund and a research grant from NanoMaterials and Science Fund of Dalian University of Technology.

246

H.Y. Xie / China Particuology 5 (2007) 242246

References
Foster, A. S., Barth, C., Schluger, A. L., & Reichling, M. (2001). Unambiguous interpretation of atomically resolved force microscopy images of an
insulator. Physical Review Letters, 86, 23732376.
Giessibl, F. J. (1995). Atomic resolution of silicon (1 1 1)-(7 7) surface by
atomic force microscopy. Science, 267, 6871.
Giessibl, F. J. (1997). Forces and frequency shifts in atomic-resolution dynamicforce microscopy. Physical Review, B56, 1601016015.
Jones, R., Pollock, H. M., Geldart, D., & Verlinden, A. (2003). Inter-particle
forces in cohesive powders studied by AFM: Effects of relative humidity,
particle size and wall adhesion. Powder Technology, 132, 196210.
Ke, S. H., Uda, T., & Terakura, K. (2002). STMAFM image formation on
TiO2 (1 1 0)1 1 and 1 2 surfaces. Applied Surface Science, 188, 319324.
Kitamura, S., & Iwatsuki, M. (1995). Observation of 7 7 reconstructed structure on the silicon (1 1 1) surface using ultrahigh vacuum noncontact atomic
force microscopy. Japanese Journal of Applied Physics, 34, L145L148.
Liu, H. (2005). Cantilever dynamics of AFM scanning over quantum dots. Master
Thesis. Dalian University of Technology (in Chinese).
Morita, S., Abe, M., Yokoyama, K., & Sugawara, Y. (2000). Defects and
their charge imaging on semiconductor surfaces by noncontact atomic
force microscopy and spectroscopy. Journal of Crystal Growth, 210, 408
415.
Morita, S., Sugawara, Y., Yokoyama, K., & Uchihashi, T. (2000). Correlation of
frequency shifts discontinuity to atomic positions on a Si(1 1 1)7 7 surface
by noncontact atomic force microscopy. Nanotechnology, 11, 120123.
Nunes, V. B., Zanette, S. I., Caride, A. O., Prioli, R., & Rivas, A. M. F. (2003).
Cantilevers behavior in the AC mode of an AFM. Materials Characterization, 50, 173177.
Sasaki, N., & Tsukada, M. (1999). Theory for the effect of the tipsurface interaction potential on atomic resolution in forced vibration system of noncontact
AFM. Applied Surface Science, 140, 339343.

Sasaki, N., Watanabe, S., Aizawa, H., & Tsukada,


M.
(2001). Simulation of
interaction force between Si tip and Si(111) 3 3 Ag surface of IET
structure in noncontact AFM. Surface Science, 493, 188193.
Schwarz, A., Allers, W., Schwarz, U. D., & Wiesendanger, R. (1999). Simulation
imaging of the In and As sublattice on InAs(1 1 0)-(1 1) with dynamic
scanning force microscopy. Applied Surface Science, 140, 293297.
Stich, I., Dieska, P., & Perez, R. (2002). Tipsurface interactions in atomic force
microscopy: Reactive vs. metallic surface. Applied Surface Science, 188,
325330.
Stich, I., Tobik, J., Perez, R., Terakura, K., & Ke, S. H. (2000). Tipsurface
interactions in noncontact atomic force microscopy on reactive surfaces.
Progress in Surface Science, 64, 179191.
Sugawara, Y., Ohta, M., Ueyama, H., & Morita, S. (1995). Defect motion on
an InP(1 1 0) surface observed with noncontact atomic force microscopy.
Science, 270, 16461648.
Uchihashi, T., Sugawara, Y., Tsukamoto, T., Minobe, T., Orisaka, S., Okada,
T., et al. (1999). Imaging of chemical reactivity and buckled dimmers on
Si(1 0 0)2 1 reconstructed surface with noncontact AFM. Applied Surface
Science, 140, 304308.
Ueyama, H., Ohta, M., Sugawara, Y., & Morita, S. (1995). Atomically resolved
InP(1 1 0) surface observed with noncontact ultrahigh vacuum atomic force
microscopy. Japanese Journal of Applied Physics, 34, L1086L1088.
Wang, Y., Song, R., Li, Y., & Shen, J. (2003). Understanding tapping-mode
atomic force microscopy data on the surface of soft block copolymers.
Surface Science, 530, 136148.
Xie, H. (2005). Frequency shifts of cantilever in AFM. Applied Surface Science,
252, 372378.
Xie, H. (1997). The role of interparticle forces in the fluidization of fine particles.
Powder Technology, 94, 99108.
Xie, H., Sun, X., & Ning, G. (2005). Hamaker theory for atom-tip interactions
of nonreactive surface in noncontact AFM. Applied Surface Science, 239,
129131.

Potrebbero piacerti anche