Scanning Tunneling Microscopy in Surface Science: Peter Sutter
Scanning Tunneling Microscopy in Surface Science: Peter Sutter
1 Introduction
After its invention in 1982 by Gerd Binnig and Heinrich Rohrer (Binnig
et al., 1982, 1983), who were awarded the 1986 Nobel Price for this
discovery, scanning tunneling microscopy (STM) has rapidly become
a standard technique for high-resolution imaging of conducting
surfaces. As such, it has revolutionized the way surface science is
conducted. Having relied almost entirely on diffraction methods for
determining surface structures, and electron spectroscopy for measur-
ing surface chemistry and electronic structure, surface scientists imme-
diately embraced the powerful imaging and spectroscopy capabilities
of STM. Earlier techniques for high-resolution surface microscopy
include field ion microscopy (Muller and Tsong, 1969), low-energy
electron microscopy (Bauer, 1998), ultrahigh-vacuum scanning elec-
tron microscopy (Venables, 2000), and reflection electron microscopy
(Yagi, 1982). Field ion microscopy, the first surface imaging technique
to routinely provide atomic resolution on metals over small sample
areas, has recently seen a revival in the form of modern atom probes
that provide tomographic images of three-dimensional sample volumes
(Cerezo et al., 2001), nicely complementing the surface imaging capa-
bilities of STM. The other techniques, none of which achieves atomic
resolution, have been developed further for specific applications. As an
example, low-energy electron microscopy has become a powerful tech-
nique for real-time microscopy of fast surface processes, such as epi-
taxial growth or surface reactions (Bauer, 1998). STM itself has recently
been at the center of yet another scientific revolution, enabling system-
atic studies on individual structures composed of a small number of
atoms or molecules, and on the order of a few nanometers in size.
This chapter attempts to provide an introduction of the basic con-
cepts of STM, together with illustrations of applications of the tech-
nique. Over two decades after its invention, the applications of STM
for imaging, spectroscopy, and manipulation at the atomic level have
clearly become too numerous to allow for a comprehensive review. In
view of the wide range of applications and overall maturity of the
969
970 P. Sutter
Figure 15–1. Band diagrams (energy vs. distance along the tunneling direc-
tion) for vacuum tunneling between two metal electrodes, A and B.
a voltage, VA. The bias offsets the Fermi energies on either side of the
vacuum barrier by eVA. Under steady-state conditions, this potential
difference establishes a net tunneling current, I, between A and B.
In a simple planar tunneling model using the Wentzel–Kramers–
Brillouin (WKB) approximation, the tunneling current is given by an
integral over the energy range between the Fermi energies on either
side of the vacuum gap,
eVA
I= ∫ ρA ( E ) ρΒ ( E − eVA )T ( E, eVA ) dE (1)
0
2 z 2m φ A + φB eVA
T ( E, eVA ) = exp − + −E (2)
2 2
Here, φA,B are the work functions of the two solids, z is their separation,
and m is the mass of the electron. Evaluation of T for positive (VA > 0)
and negative bias (VA < 0) shows that the transmission probability is
highest for electrons at the Fermi energy of the material that is nega-
tively biased, and falls off exponentially for lower energies down to a
lower cutoff at the Fermi energy of the positively biased electrode. This
general observation has important consequences in tunneling spec-
troscopy and spectroscopic imaging, as discussed in Section 3.
The tunneling current depends strongly on the separation, z, between
the two solids. To illustrate this fact, we simplify Eq. (1) by assuming
constant densities of states, independent of energy. In the limit of low
bias voltage, VA/φA,B << 1, the tunneling current is then given by
(
I = ρA ρBVA exp −2 m ( φ A + φB ) / 2 z ) (3)
Figure 15–5. (a) Ball-and-stick model of the Si(111)–(7 × 7) surface. Top: top view; bottom: side view.
a, adatom; r, rest atom; c, corner hole. (b) Filled-state constant-current STM on Si(111)–(7 × 7); V = −1.0 V;
I = 0.4 nA. (c) Empty-state constant-current STM; V = +1.2 V; I = 0.4 nA.
978 P. Sutter
Rohrer, 1983) and electronic (Hamers et al., 1986) structure was solved
in pioneering STM experiments, and which continues to play an impor-
tant role in STM technique development. Figure 15–5a shows a struc-
ture model of the (7 × 7) reconstruction (Takayanagi et al., 1985). Figure
15–5b, obtained at negative sample bias (V = −1.0 V), i.e., with electrons
tunneling from occupied sample states to unoccupied tip states, shows
the corrugation associated with filled states of the sample. Conversely,
Figure 15–5c, obtained at positive sample bias (V = +1.2 V), maps the
empty states of the sample. Both images show atomic resolution, clearly
resolving the twelve adatoms (“a”) per unit cell. In addition character-
istic deep “corner holes” (“c”) bounding the diagonals of the 4.6 nm ×
2.9 nm rhombohedral unit cell are imaged.
While all adatoms are mapped uniformly in the empty state image,
the filled state scan shows one-half of the unit cell somewhat higher
than the other, an effect on the charge density due to the different
stacking sequence of atomic layers in the two halves of the unit cell. A
comparison of the images obtained at opposite bias polarity suggests
that in addition to surface topography, the electronic structure of the
sample surface adds substantially to the contrast observed in STM
imaging. This is implicit also in Eq. (1) via the dependence of the tun-
neling current on the local densities of states of both tip and sample at
the tunneling contact. Note also that the rest atoms (“r”), a second
near-surface species with dangling bonds protruding into the vacuum,
are imaged neither at positive nor at negative sample bias. A detailed
discussion of bias-dependent imaging and other tunneling spectro-
scopy methods used to assess the local electronic structure of a sample
surface is given in Section 3.
Apart from the structure of clean Si surfaces, Si-based surface chem-
istry has been studied widely by STM. Initial studies provided a knowl-
edge base for technological processes, such as reactive ion etching,
doping, and chemical vapor deposition (CVD). STM was used to probe
interactions of Si with halogens (Boland, 1993) and with small mole-
cules involved in surface passivation (H2, H; Laracuente and Whitman,
2001), doping (PH3; Wang et al., 1994a), and CVD growth (Si2H6; Wang
et al., 1994b). More recent research directions include the integration
of molecular electronic elements with Si. Under this perspective, the
covalent bonding of a wide variety of organic molecules on Si has been
studied (for a review, see Wolkow, 1999), including small molecules
such as ethylene (Mayne et al., 1993) and acetylene (Li et al., 1997),
simple alkenes (e.g., propylene; Lopinski et al., 1998) and polyenes (e.g.,
1,3-cyclohexadiene; Hovis and Hamers, 1997), pentacene (Kasaya et al,
1998), and benzene (Borovski et al., 1998). On clean Si(001), adsorption
is from the gas phase in UHV. H-passivated Si can also be modified by
ex situ wet-chemical techniques, and reintroduced into UHV for STM
imaging.
Traditionally, an important aspect of the surface science of semicon-
ductors has been epitaxial growth. STM observations at initial growth
stages, i.e., at coverages of fractions of a monolayer (ML) up to several
ML, can provide direct insight into fundamental processes such as
adatom diffusion, incorporation into steps, and nucleation of mono-
Chapter 15 Scanning Tunneling Microscopy in Surface Science 979
layer islands (for a review, see Zhang and Lagally, 1998). Observations
of growth and equilibrium structures allow identifying kinetic and
thermodynamic factors affecting the growth process, as well as the role
of defects, of elastic strain, etc.
STM studies of Si epitaxy are commonly performed in either of two
modes. Conceptually preferred is the dynamic observation, via time-
lapse STM “movies,” of the growth surface with the sample held in the
microscope at high temperature and in the presence of a deposition
flux. This difficult technique will be discussed in more detail in Section
4. As an alternative, STM imaging can be performed at room tempera-
ture on samples that have been quenched at key stages in the growth
process. If measures are taken to verify that the quenched-in morphol-
ogy is indeed representative of that at high temperature during growth,
this “quench-and-look” technique offers the advantage of higher spatial
resolution and larger scan sizes over high-temperature STM.
Early STM experiments focused on the initial stages of Si homoepi-
taxy and Ge/Si heteroepitaxy, mostly on the technologically important
(111) or (001) surfaces. Key results include the identification of regimes
of step-flow growth and nucleation, determination of the activation
energy and atomistic pathway of surface diffusion (Mo et al., 1991),
explanation of growth and equilibrium shapes of two-dimensional
(2D) islands (Mo et al., 1989), and exploration of the modification of the
growth process by surfactants (Horn-von Hoegen, 1994). Figure 15–6
illustrates the identification of the initial stage of island formation in
Figure 15–6. STM images of 0.01 monolayers Ge on Si(001). The long diagonal
bands are substrate dimer rows. Filled-state images (A–C; V = −2 V; I = 0.2 nA)
and corresponding empty-state images (D– F; V = +2 V; I = 0.2 nA), showing
rows of symmetric (bean-shaped) and asymmetric (buckled) substrate dimers,
and Ge adatom-induced chain-like structures (marked by arrows). The chain-
like paired adatom structures are seen as metastable precursors to the forma-
tion of monolayer islands in Ge/Si homoepitaxy (Reprinted with permission
from Qin and Lagally, © 1997 AAAS).
980 P. Sutter
Figure 15–7. Large-area atom-resolved STM. (a) STM image with 0.75-µm
field of view and 0.05-nm pixel size, showing the surface morphology of Si(111)
after deposition of 1.5 ML Ge at 550°C. An array of parallel steps, running
diagonally from the upper left to lower right, coexists with islands nucleated
during Ge deposition. (b) Zoomed-in view (50 nm) of the area marked by a
square in (a), showing one of the islands. (c) Further zoom-in (10 nm), showing
the coexistence of (7 × 7) and (5 × 5) reconstructed domains in the area marked
in (b).
Figure 15–8. Transition from 2D to 3D morphology in the growth of Ge on Si(001). Large-area atom-
resolved STM images (250 nm × 125 nm sections are shown here) illustrate the surface morphology at
(a) 1.5 ML, (b) 2.3 ML, (c) 3.5 ML, and (d) 4.0 ML Ge coverage. The transition to 3D growth is preceded
by the formation of highly correlated, anisotropic surface roughness. With the superior statistics of
large-area high-resolution images, the repulsive interaction between surface steps and defects (dimer
vacancy lines) was identified as the origin of the ordering of this surface roughness. (Reprinted with
permission from Sutter et al., © 2003a by the American Physical Society.)
982 P. Sutter
Figure 15–9. Shapes of coherent (i.e., dislocation-free) 3D islands induced by lattice mismatch strain
during Si1−xGex/Si(001) heteroepitaxy. (a) Unit stereographic triangle for Si, showing the major and
minor stable facets (Gai et al., 1998). (b) STM top view, showing a shallow (105) faceted “hut.” (c) Mul-
tifaceted “dome” shape, terminated by (105), (113), and (15 3 23) facets. (d) “Barn”-shaped 3D island,
observed for low-misfit Si1−xGex alloys on Si(001), transformed from a “dome” by introduction of addi-
tional (111) facets (Reprinted with permission from Sutter et al., © 2004 American Institute of
Physics.).
3 Tunneling Spectroscopy
Equation (1) gives an expression for the tunneling current I in the WKB
approximation for planar tunneling. In this approximation the tunnel-
ing current is an integral, over the energy range of width |eV| in which
tunneling can occur, of a product of the densities of states (DOS) of tip
[ρT(E)] and sample [ρS (E)] multiplied by the transmission probability,
T(E, eV). While this simple approximation cannot address the high
spatial resolution of STM—a more elaborate theory capable of address-
ing this aspect will be discussed in Section 6—it demonstrates that the
tunneling current is affected by the electronic structure of both the tip
and the sample. In a large number of applications it would be desirable
to merely map the atomic scale “topography” of a sample. The sensitiv-
ity of the tunneling current to the DOS can then be a disadvantage,
since it often prevents a simple interpretation of a tunneling image as
Chapter 15 Scanning Tunneling Microscopy in Surface Science 983
Figure 15–12. Energy-filtered STM. (a) Bulk band structure of monocrystalline InAs probe tips, pro-
jected along the (111) zone axis parallel to the tunneling direction. Ep denotes a projected gap in the
bulk conduction band. (b) Band diagram and STM image for low-bias filled state STM with InAs tips,
demonstrating preferential imaging of the adatom state on Si(111)-(7 × 7). (c) High-bias filled state
InAs-tip STM on Si(111)-(7 × 7): alignment of the adatom state with the projected tip gap reduces its
contribution to the tunneling current, giving rise to preferential imaging of the energetically lower
rest-atom state (Reprinted with permission from Sutter et al., © 2003b by the American Physical
Society).
not exist. Low-energy electrons can propagate into extended states, and
the reflectivity in the same energy range is low. Similarly, it can be
expected that a lack of final states due to projected band gaps in the
STM tip material can significantly affect the tunneling probability in
the energy range spanned by such gaps. This modulation of the trans-
mission probability could be used for efficient energy filtering of the
tunneling current in constant-current STM, specifically in filled-state
imaging.
Calculations by density functional theory (DFT) show a number of
gaps in the [111]-projected band structure of III–V compound semicon-
ductors such as InAs. III–V compound semiconductors cleave easily at
(110) planes, and cleavage corners bounded by {110} and (100) planes
are sufficiently sharp for atomic-resolution STM (Nunes and Amer,
1993). In addition, InAs(110) has the advantage of an unpinned surface
with minimal band bending and slight electron accumulation. The
feasibility of energy-filtered STM has been demonstrated by using a
cleaved InAs tip to selectively image different occupied surface states
on Si(111)-(7 × 7) (Sutter et al., 2003b). In these experiments, summa-
rized in Figure 15–12, a [111]-projected gap in the InAs conduction
band, centered at Γ and about 2.6 eV wide, is used to modulate the
tunneling current in filled-state constant current STM. The “adatom”
state (“a”) on Si(111)-(7 × 7), which lies only 0.35 eV below the Fermi
energy and is imaged by STM using a metal tip, dominates the tunnel-
ing current at low tip–sample bias. At higher bias, this state aligns with
a projected gap in the InAs conduction bands and its tunneling
probability is reduced sharply. Under these conditions the “rest atom”
986 P. Sutter
Figure 15–13. (a) Bias-dependent STM on GaAs(110): selective imaging of Ga and As sublattices at
positive and negative sample bias, respectively (Reprinted with Permission from Feenstra et al., ©1987
by the Armeucion Physics Socuity). (b) Compound STM image of the InP(110) surface, assembled from
separate positive and negative bias scans (Reprinted from Ebert et al., ©1992 with permission from
Elsevier). (See color plate.)
state (“r”), a surface state 0.8 eV below EF that is not accessible in con-
ventional STM imaging, contributes a majority of the tunneling current
and can be imaged selectively. Importantly, these results suggest that
robust bulk band structure effects, which are independent of the atomic
structure of the tip, can be used to modulate the transmission probabil-
ity in vacuum tunneling. Once developed into a routine imaging tech-
nique, energy-filtered STM has the potential to provide rich spectroscopic
contrast within the relatively simple framework of bias-dependent
constant current STM.
A classic example of a scenario in which conventional bias-
dependent imaging provides directly interpretable spectroscopic infor-
mation is imaging of cation and anion sites on (110) cleavage planes of
compound semiconductors such as GaAs (Feenstra et al., 1987) or InP
(Ebert et al., 1992) (Figure 15–13). For (110) surfaces of compound semi-
conductors, such as GaAs, the occupied state density (imaged at nega-
tive sample voltage) is localized on the anions (As, P), while the
unoccupied state density (positive sample voltage) is localized on
the cations (Ga, In). As a result, bias-dependent STM imaging can be used
to selectively image the cation and anion sublattices on these surfaces.
Figure 15–14. (a) Local normalized conductance spectra obtained on n- and p-doped GaAs(110),
showing three peaks (V, D, C) originating from surface states (Feenstra et al., 1987). (b) Normalized
conductance spectra obtained on InP(110), illustrating the valence and conduction band edges delimit-
ing the bulk bandgap (1.4 eV), and the A5 and C3 surface states defining the surface gap (1.9 eV)
(Reprinted from Ebert et al., © 1992 with permission from Elsevier).
988 P. Sutter
of the surface state density. The spectra show three distinct peaks
associated with surface states. Peak C is related to a state in the GaAs
conduction band, localized on the Ga atoms, while peak V stems from
a valence state localized on the As anions. Peak D is associated with
tunneling of dopant-induced carriers within the bulk band gap. The
separation between the leading edges of the C and V peaks is close to
the bulk band gap of GaAs (Eg = 1.43 eV), i.e., in this system tunneling
spectroscopy can be used to determine the band gap. Figure 15–14b)
shows tunneling conductance spectra on the (110) cleavage surface of
another III–V compound, InP. Also for InP, the band gap (∼1.4 eV) is
evident from sharp onsets of significant state density at the valence
and conduction band edges. However, for this system it was argued
that a wider surface band gap (∼1.9 eV) is indirect, delimited by a
surface state C3 at the edge of the surface Brillouin zone and a broad-
ened state A5 at the zone center.
A natural, albeit experimentally much more complex extension of
local I(V) spectroscopy is the measurement of tunneling spectra, as
discussed above, at each image pixel of a constant-current STM scan.
This measuring scheme is called current-imaging tunneling spectro-
scopy (CITS). Since complete I(V) spectra are obtained at each image
point, the corresponding data sets can provide a full range of spectro-
scopic information. As an example, current maps I(x, y) at fixed tip–
sample bias can be produced. Instead, the voltage dependent tunneling
conductance dI(V)/dV can be calculated numerically and mapped as a
function of sample position. If the chosen voltage corresponds to the
energy of a surface state of the sample, conductance maps will provide
a direct image of the spatial distribution of that state. Due to the com-
plexity of the data acquisition, the experimental requirements for CITS
are quite stringent. Since complete I(V) curves are measured at each
image pixel, requiring the stopping of a scan and deactivation of the
feedback loop for a fraction of a second per spectrum, a very high sta-
bility of the tunneling gap and low lateral drift of the tip relative to
the sample are of key importance. These conditions are more easily
fulfilled at cryogenic temperatures, where a z-stability of the order of
1 pm and lateral drift velocities of the order of few Å/hour are possible.
Low temperatures also reduce the thermal broadening of the tunneling
transmission coefficient T(E), and thus narrow the linewidth of fea-
tures in the tunneling spectra.
A classic example of the application of CITS, the measurement of the
electronic structure of Si(111)-(7 × 7), is shown in Figure 15–15 (Hamers
et al., 1986). Shown are a constant current image of the surface at +2 V
sample bias (a), as well as current images acquired during the same
scan at bias voltages of +1.45 V and −1.45 V, showing spatial maps of
unoccupied (+)/occupied (−) sample states at energies up to 1.45 eV
above/below the Fermi energy. CITS difference images, a precursor to
modern dI/dV maps, discussed below, calculated by numerically sub-
tracting current images at energies bracketing those of specific surface
states, allowed the mapping of the adatom state, dangling bond, and
backbond state on this complex reconstructed surface.
Chapter 15 Scanning Tunneling Microscopy in Surface Science 989
Figure 15–15. Current-imaging tunneling spectroscopy on Si(111)-(7 × 7) (Hames et al., 1986). (a)
Constant-current STM image (V = +2 V), with one (7 × 7) unit cell outlined (F, faulted; U, unfaulted
half). (b and c) Current images obtained at +1.45 V and −1.45 V, respectively. Note the strong similarity
between the rest-atom contrast in the current image (c), and the contrast obtained in rest-atom imaging
by energy-filtered constant-current STM (Figure 15–12).
More recently, this complex data acquisition mode has been used,
for example, to map modifications in the electronic structure of indi-
vidual carbon nanotubes arising with the supramolecular assembly of
C60 molecules inside the hollow tube to form nanotube “peapods”
(Hornbaker et al., 2002). dI/dV spectra obtained along the symmetry
axis of C nanotubes show distinct peaks in the density of unoccupied
states with a spatial periodicity corresponding to the average spacing
of embedded C60 molecules observed by transmission electron micros-
copy, suggesting that the insertion of C60 causes a significant modula-
tion of the electronic structure. A control measurement was constructed
by using the STM tip to push the C60 away, and remeasuring dI/dV
spectra along the same nanotube section without embedded C60. The
empty tube obtained in this way shows small and smooth variations
in the density of states along the nanotube axis. Qualitatively, the elec-
tronic structure of the peapods was explained by the single wall C-
nanotube acting as a conduit that enhances the coupling between C60
molecules nested inside it, and the Bragg scattering of nanotube states
due to the periodic arrangement of the C60. Calculations of the peapod
electronic structure confirm the formation of a hybrid electronic band,
which derives its character from both the nanotube states and the C60
molecular orbitals (Figure 15–16).
Figure 15–16. I(V) tunneling spectroscopy on C60/C-nanotube “peapods” (Reprinted with permission
from Hornbaker et al., © 2002 AAAS). (A) Map of an array of full dI/dV spectra along the axis of a C-
nanotube “peapod.” Sample bias voltage is plotted on the horizontal axis and displacement along the
tube on the vertical axis. (B) Representative dI/dV spectra at selected positions along the tube. Large
conductance peaks are found at positions of embedded C60 molecules. (C) Variation of tunneling con-
ductance along the tube axis. (D) Reference spectroscopic map on an empty C-nanotube section
without embedded C60, in which no strong modulation of the tunneling conductance is observed. (See
color plate.)
chains (Crain and Pierce, 2005) or single molecules (Repp et al., 2005).
A particularly widespread application is the imaging of quasiparticle
interference. In an ideal metal, the Landau-quasiparticle eigenstates are
Bloch wavefunctions with specific wavevectors k and energies ε. Their
dispersion cannot be measured directly by STM. However, in the pres-
ence of disorder, e.g., impurities or crystal defects, elastic scattering
mixes eigenstates with different k that are located on the same quasi-
particle contour of constant energy in k-space. When states k1 and k 2
are mixed by scattering, an interference pattern with wavevector q =
k 2 − k1 appears in the norm of the quasiparticle wavefunction. The
interference leads to modulations in the local density of states with
wavelength λ = 2π/|q|, which can be observed as modulations in the
differential tunneling conductance. Via differential conductance
mapping, interference of electronic eigenstates has been probed in
metals (e.g., Cu(111), Crommie et al., 1993; Au(111), Hasegawa and
Avouris, 1993) and semiconductors (e.g., InAs(110), Wittneven et al.,
1998). When density of states modulations are detected by STM, certain
contours of constant energy in k-space can be reconstructed by analyz-
ing the Fourier transform of the real-space density of states map. Such
Fourier transform spectroscopy simultaneously yields real-space and
momentum-space information on wavefunctions, scattering processes,
and quasiparticle dispersion.
Elastic scattering of Bogoliubov quasiparticles in superconductors
can also give rise to conductance modulations (Hoffman et al., 2002a),
suggesting that STM could be used as a local probe to study electron
correlation in superconductors. Quasiparticle inteference has indeed
been demonstrated for cuprate high-temperature superconductors, such
as Bi2Sr2CaCu2O8+δ (Hoffman et al., 2002a; Vershinin et al., 2004) and
Ca2−xNaxCuO2Cl2 (Hanaguri et al., 2004). On Bi2Sr2CaCu2O8+δ crystals
cleaved at the BiO plane at cryogenic temperature in UHV, a constant-
current image (Figure 15–17A) and differential conductance maps
(Figure 15–17B) are obtained simultaneously in a low-temperature
Figure 15–18. Spin-polarized STM on a Gd(0001) sample with an exchange-split surface state and a
magnetic Fe tip with constant spin polarization close to EF. (a) Due to the spin-valve effect the tun-
neling current of the surface state spin component parallel to the tip magnetization is enhanced. (b)
Illustration of the reversal in the dI/dV signal at the surface state peak position upon switching the
sample magnetically. (c) Experimental observation of this reversal in tunneling into an isolated Gd
island (Reprinted with permission from Bode et al., ©1998 by the American Physical Society). (See
color plate.)
Chapter 15 Scanning Tunneling Microscopy in Surface Science 993
state will be high, while that of the minority state is low. After reversal
of the applied magnetic field, the minority spin will align with the tip
magnetization. The conductance of the minority state thus increases,
while that of the majority state is reduced. Difference spectra obtained
by subtracting the traces, measured at two polarities of the external
field, provide a direct measure of the sample magnetization, and can
be used to build maps of magnetic order. Spin polarized STM has been
shown to be a very powerful technique for mapping magnetic and
antiferrromagnetic order with atomic resolution, as demonstrated by
studies on 2D antiferromagnetic ordering in Mn/W(110) (Heinze et al.,
2000) and on magnetic hysteresis in Fe/W(110) (Pietzsch et al., 2001).
Figure 15–19. Single molecule vibrational spectroscopy and microscopy. (a) d2 I/dV 2 recorded over
individual C2H2 and C2D2 molecules on Cu(100) at 8 K. (1) Signature of inelastic electron tunneling by
exciting the C—H stretch mode with an energy of 358 meV. (2) Isotope shifted (266 meV) inelastic tun-
neling peak of the C—D stretch mode. (b) Constant current (CC, 4.8 nm) STM of adjacent C2H2 and
C2D2 molecules on Cu(100), along with simultaneously acquired d2 I/dV 2 maps obtained at 358 mV and
266 mV, imaging the C—H and C—D bonds selectively. The symmetric round appearance is attributed
to rotations of the molecules during the experiment. A d2 I/dV 2 map obtained at intermediate energy
(311 mV) shows no vibrational contrast. (Reprinted with permission from Stipe et al., © 1998 AAAS.)
bond. With the tip–sample distance too large to allow true chemical
bonding, either atomic relaxation or the local charge density can change
as the tip is moved across different atomic species, which in turn
affects the tunneling current and gives rise to apparent height differ-
ences in constant current STM. Qualitatively, surface atoms with
higher chemical affinity to the tip atom should have higher tunneling
conductance and thus will be imaged at larger apparent height. Several
metal alloy systems, such as PtNi(111) and PtRh(111), are believed
to show element contrast that is based on tip–sample interaction.
Clearly, the nature of the tip apex would be of key importance for this
type of contrast. This may explain why for some systems element
contrast is obtained only after deliberate chemical modification of the
STM tip.
Figure 15–21. STM on Rutile TiO2(110). (a) Structure of the TiO2(110) surface. V, a single oxygen
vacancy. (b) Constant-current STM on TiO2(110). Bright rows are assigned to five-fold coordinated Ti
atoms and dark rows to bridging oxygen atoms. Oxygen vacancies are imaged as protrusions between
the Ti rows. Adsorbed O2 molecules are associated with bright protrusions on the Ti rows. (Reprinted
with permission from Schaub et al., © 2003 AAAS.)
Figure 15–22. O2-mediated diffusion of oxygen vacancies on TiO2(110). (a and b) Consecutive frames
of an STM movie on the motion of oxygen vacancies on TiO2(110) (T = 300 K, 8.5 s/frame). (c) Difference
image of (a) and (b), highlighting changes due to the diffusion of single oxygen vacancies. Vacancies
diffuse perpendicular to the bridging oxygen rows. (d–f) Time-lapse STM of the O2-assisted diffusion
of an individual oxygen vacancy (T = 230 K, 1.1 s/frame). (g) Schematic illustration of the O2-mediated
vacancy diffusion mechanism. (Reprinted with permission from Schaub et al., © 2003 AAAS.)
1000 P. Sutter
Figure 15–23. Simultaneous mass spectrometry and STM imaging in a catalytic flow reactor: CO
oxidation on Pt(110). (Top) Mass spectrometer signals of O2, CO, and CO2 measured at the output of
the flow reactor cell. (Bottom) STM images on the Pt(110) catalyst surface acquired at selected stages
of the process: CO adsorption (A), O2 flow (B–D), and repetition of the sequence (E–H). Note the strong
surface roughening in (D), associated with increased CO2 evolution. (Reprinted with permission from
Hendriksen and Frenken, © 2002 by the American Physical Society.)
the tip over that species and tracks its coordinates as it diffuses over
the substrate. In the atom-tracking mode, the STM spends all of its time
measuring the diffusion trajectory, which results in substantially
improved time-resolution compared to time-lapse STM movies.
Atom tracking STM has been used to measure the diffusion kinetics
of Si (Swartzentruber, 1996) and SiGe (Qin et al., 2000) dimers on
Si(001) above room temperatures, of water molecules on Pd(111) (Mitsui
et al., 2002), and of Pd atoms in a Pd/Cu(001) surface alloy (Grant
et al., 2001). The latter example is illustrated in Figure 15–24. Pd atoms
in the surface alloy are imaged as protrusions in STM. Locking the
STM tip onto individual Pd atoms, their diffusion pathway can be
tracked. Analyzing the residence time (the time between hops) and
jump length leads to the conclusion that there is no time correlation
between individual Pd diffusion events, and that the diffusion is medi-
ated by surface vacancies rather than Cu adatoms, i.e., involves rapidly
diffusing vacancies visiting Pd atoms in the surface layer. From
the temperature dependence of the Pd hop rate—determined from
temperature-dependent tracking experiments—the activation energy
of the overall process, in this case equal to the sum of the vacancy for-
mation energy and the energy barrier for lateral Pd-vacancy exchange,
can be measured.
In contrast to time-lapse STM, in which the tip is scanned rapidly
across a larger field of view, the tip maintains close contact with the
diffusing entity in atom tracking. Hence, the observed diffusion process
could be affected by tip–sample interactions. While this question has
to be studied on a case-by-case basis, at least one system [SiGe dimer
Figure 15–26. Molecule cascades. (Left) (A) Configurations of neighboring CO molecules on Cu(111):
isolated CO molecule, dimer, and trimer in a “chevron” configuration. Large circles mark the positions
of the molecules and small dots indicate the surface layer Cu atoms (1.9 nm scans; T = 5 K). (B) The
same area after one CO molecule in the trimer has hopped to generate a stable close-packed trimer.
(Right) Demonstration of a molecular mechanical AND gate, implemented via cascades of “chevron”-
type CO trimers. (A) Model of the AND gate. (B–D) Sequence of STM images (5.1 nm by 3.4 nm)
showing the operation of the AND gate: (B) Initialization. (C) Result after input X was triggered with
the STM tip. (D) When input Y was triggered, the cascade propagated all the way to the output. (E)
Result of triggering only input Y. (Reprinted with permission from Heinrich et al., © 2002 AAAS.)
1006 P. Sutter
Figure 15–27. Charging of individual Au atoms on ultrathin NaCl/Cu(111). (a) Constant-current STM
on a pair of Au atoms, one of which was charged negatively with the other remaining neutral. (b)
Signature of the charging event. Successful charging, achieved by a voltage pulse while maintaining
the tip above a single Au atom, is indicated by a sudden drop in tunneling current. (c and d) Quasi-
particle scattering: neutral Au adatoms do not scatter NaCl/Cu(111) interface-state electrons (c),
whereas the negatively charged adatom acts as a scatterer (d). (Reprinted with permission from Repp
et al., © 2004 AAAS.)
The STM tip is scanned at constant current It, measured between the
tip and base electrode, over the surface of the base layer. The tip serves
as a point source of hot carriers, electrons or holes, injected into the
metal base. If the metal thickness is small compared to the mean free
path for electron–phonon, electron–electron, or impurity scattering,
most carriers reach the metal–semiconductor interface ballistically, i.e.,
with their original energy and momentum distribution. In metals,
the mean free path for electron–phonon scattering of electrons with
energies of a few eV is typically of the order of 10 nm. In significantly
thicker films, electron–phonon scattering would broaden the momen-
tum distribution of the injected carriers, i.e., mostly affect the spatial
resolution of BEEM, while having little effect on the energy distribu-
tion. Hence, the injected ballistic carriers generally have a well-defined
energy distribution and can be used to perform spectroscopic mea-
surements on buried potential barriers.
At the metal–semiconductor interface, ballistic carriers with energies
below a threshold equal to the Schottky barrier height eφB are reflected
back into the metal base, while carriers with energy above the thresh-
old are transmitted into the semiconductor collector (Figure 15–29). If
other barriers exist in the collector layer, the transport through that
material can involve additional interfacial reflections. Carriers that are
transmitted through the entire collector layer contribute to the collector
current (or BEEM current), which is measured between base and col-
lector contacts.
The energy of the injected carriers is varied by changing the voltage
applied between the STM tip and metal base. For low bias, none of the
injected carriers is transferred across the metal–semiconductor inter-
face. At voltages above the threshold, some carriers have energy above
the Schottky barrier and can cross the interface, causing an increase in
collector current with increasing tip–sample bias above threshold.
From the onset of the BEEM current, the barrier height eφB can be
determined, e.g., by fitting the measured Ic(V) to a theoretical expres-
sion. In a simple one-dimensional theory,
I c (V ) = RIt ∫ dE [ f ( E ) − f ( E − eV )] × θ [ E − ( EF − eV + eφ B )] (6)
where R is a bias independent parameter and f(E) is the Fermi
function.
Figure 15–30 shows experimental BEEM spectra on an Au/n-Si(001)
junction, measured at room temperature in N2 atmosphere, fitted using
Eq. (6) to determine a Schottky barrier eφB = 0.92 eV. Due to its low
reactivity and oxidation resistance, Au was used in many of the early
BEEM experiments as a nonepitaxial metal base on a variety of other
semiconductors. Schottky barrier heights were determined, for example,
for Au/n-GaAs(001) (eφB = 0.70 eV at 77 K) (Bell et al., 1990) and Au/n-
GaP(110) (eφB = 1.41 eV) (Prietsch and Ludeke, 1991). For Au/n-CdTe(001)
(eφB = 0.69–1.07 eV) (Fowell et al., 1990), strong lateral variations in the
Figure 15–32. BEEM on self-assembled quantum dots. (a) STM and BEEM images of an individual
InAs/GaAs quantum dot under a polycrystalline Au base. (b) Comparison of BEEM spectra obtained
by ballistic electron injection into an InAs dot and into the wetting layer between dots. (Reprinted
with permission from Rubin et al., © 1996 by the American Physical Society.)
Figure 15–33. Electron interference in a Pb “quantum wedge” on Si(111). (Left) Geometry of the
Pb wedge on a stepped Si(111) substrate. The Pb thickness varies in increments of roughly one Si(111)
step height, giving rise to electron interference fringes in the direction of the substrate steps. (Right)
Constant-current STM images (730 nm × 1100 nm) at −5 V (a) and +5 V (b) sample bias. The image
obtained at positive bias shows apparent height changes at the surface of the wedge due to electron
interference in the Pb film. (Reprinted with permission from Altfeder et al., © 1997 by the American
Physical Society.)
Chapter 15 Scanning Tunneling Microscopy in Surface Science 1013
Figure 15–34. Cross-sectional STM on self-assembled quantum dots. (Left) Illustration of the possible
apparent geometries observed due to cleavage at random positions through quantum dots with pyra-
midal and truncated pyramidal shape. (Right) Filled-state constant-current STM of an InAs quantum
dot embedded in GaAs. Part of the image in (a) is treated by a local mean equalization filter to accen-
tuate the atomic corrugations in the dot and the surrounding GaAs matrix, as shown in (b). (Reprinted
with permission from Bruls et al., © 2002 American Institute of Physics; reprinted with permission
from Gong et al., © 2004 Amercian Institute of Physics.)
Chapter 15 Scanning Tunneling Microscopy in Surface Science 1015
Figure 15–35. Imaging of individual dopants by X-STM. (a) Filled-state constant-current STM image
(V = −1.6 V) of cleaved GaAs(110). Individual SiGa substitutional donors appear as protrusions of dif-
ferent apparent height. (b) Band diagram illustrating the electron tunneling process between a metal
tip and the GaAs(110) surface in the presence of tip-induced band bending. The Coulomb potential of
a donor ion locally alters the band bending, causing an increase in tunneling current above the donor.
(Reprinted with permission from Zheng et al., © 1994 by the American Physical Society.)
where f(E) is the Fermi function, V denotes the tunneling bias, Mµv the
tunneling matrix element between a tip state ψµ and a sample state ψv,
and Eµ(v) the energy of state ψµ(v). In the second line, the Fermi function
was replaced by a step function (zero-T approximation), and the limit
of low tunneling bias has been assumed. If, after Bardeen (1961), the
tunneling matrix element is written as a surface integral, and the
sample and tip wavefunctions are expanded in plane waves, one
obtains
4 π 2 2
d q a qb q* e − κ q z t e iq ⋅x t
m ∫
M µv = − (8)
References
Altfeder, I.B., Matveev, K.A. and Chen, D.M. (1997). Phys. Rev. Lett. 78, 2815.
Altfeder, I.B., Chen, D.M. and Matveev, K.A. (1998). Phys. Rev. Lett. 80, 4895.
Bardeen, J. (1961). Phys. Rev. Lett. 6, 57.
Bauer, E. (1994). Rep. Prog. Phys. 57, 895.
Bauer, E. (1998). Surf. Rev. Lett. 5, 1275.
Becker, R.S., Swartzentruber, B.S., Vickers, J.S. and Klitsner, T. (1989). Phys. Rev.
B 39, 1633.
Bell, L.D., Hecht, M.H., Kaiser, W.J. and Davis, L.C. (1990). Phys. Rev. Lett. 64,
2679.
Bell, L.D., Kaiser, W.J., Hecht, M.H. and Davis, L.C. (1991). J. Vac. Sci. Technol.
B 9, 594.
Binnig, G. and Rohrer, H. (1983). Surf. Sci. 126, 236.
Binnig, G., Rohrer, H., Gerber, Ch. and Weibel, E. (1982). Appl. Phys. Lett. 40,
178.
Binnig, G., Rohrer, H., Gerber, Ch. and Weibel, E. (1983). Phys. Rev. Lett. 50,
120.
Binnig, G., Garcia, N. and Rohrer, H. (1985). Phys. Rev. B 32, 1336.
Binnig, G. and Smith, D.P.E. (1986). Rev. Sci. Instrum. 57, 1688.
Bode, M., Getzlaff, M. and Wiesendanger, R. (1998). Phys. Rev. Lett. 81, 4256.
Boland, J.J. (1993). Science 262, 1703.
Boland, J.J. and Weaver, J.H. (1998). Phys. Today 51, 34.
Bonnell, D.A., Ed. (2001). Scanning Probe Microscopy and Spectroscopy—Theory,
Techniques, and Applications, 2nd ed., Wiley-VCH, New York.
Borovski, B., Krueger, M. and Ganz, E. (1997). Phys. Rev. Lett. 78, 4229.
Borovski, B., Krueger, M. and Ganz, E. (1998). Phys. Rev. B 57, 4269.
Brookes, I.M., Muryn, C.A. and Thornton, G. (2001). Phys. Rev. Lett. 87,
266103.
Bruls, D.M., Vugs, J.W.A.M., Roenraad, P.M., Salemink, H.W.M., Wolter, J.H.,
Hopkinson, M., Skolnick, M.S., Long, F. and Gill, S.P.A. (2002). Appl. Phys.
81, 1708.
Cerezo, A., Larson, D.J. and Smith, G.D.W. (2001). MRS Bull. 26, 102.
Chambliss, D.D., Wilson, R.J. and Chiang, S. (1995). IBM J. Res. Dev. 39, 639.
Chen, C.J. (1993). Introduction to Scanning Tunneling Microscopy (Oxford Uni-
versity Press, Oxford).
Chen, Y., Ohlberg, D.A.A., Medeiros-Ribeiro, G., Chang, Y.A. and Williams,
R.S. (2000). Appl. Phys. Lett. 76, 4004.
Ciobanu, C.V. and Predescu, C. (2004). Phys. Rev. B 70, 085321.
Crain, J.N. and Pierce, D.T. (2005). Science 307, 703.
Crommie, M.F., Lutz, C.P. and Eigler, D.M. (1993). Nature 363, 524.
Cuberes, M.T., Bauer, A., Wen, H.J., Prietsch, M. and Kaindl, G. (1994). J. Vac.
Sci. Technol. B 12, 2646.
Cygan, M.T., Dunbar, T.D., Arnold, J.J., Bumm, L.A., Shedlock, N.F., Burgin,
T.P., Jones, L., Allara, D.L., Tour, J.M. and Weiss, P.S. (1998). J. Am. Chem.
Soc. 120, 2721.
Davies, J.H., Bruls, D.M., Vugs, J.W.A.M. and Koenraad, P.M. (2002). J. Appl.
Phys. 91, 4171.
Diebold, U. (2003). Surf. Sci. Rep. 48, 53.
Ebert, Ph., Cox, G., Poppe, U. and Urban, K. (1992). Surf. Sci. 271, 587.
Chapter 15 Scanning Tunneling Microscopy in Surface Science 1021
Eder, C., Smoliner, J. and Strasser, G. (1996). Appl. Phys. Lett. 68, 2876.
Eigler, D.M. and Schweizer, E.K. (1990). Nature 344, 524.
Erwin, S.C., Baski, A.A. and Whitman, L.J. (1996). Phys. Rev. Lett. 77, 687.
Feenstra, R.M., Stroscio, J.A., Tersoff, J. and Fein, A.P. (1987). Phys. Rev. Lett.
58, 1192.
Feenstra, R.M., Collins, D.A., Ting, D.Z.-Y., Wang, M.W. and McGill, T.C.
(1994). Phys. Rev. Lett. 72, 2749.
Festy, F. and Palmer, R.E. (2004). Appl. Phys. Lett. 85, 5034.
Fowell, A.E., Williams, R.H., Richardson, B.E. and Shen, T.-H. (1990). Semicond.
Sci. Technol. 5, 348.
Fujikawa, Y., Akiyama, K., Nagao, T., Sakurai, T., Lagally, M.G., Hashimoto,
T., Morikawa, Y. and Terakura, K. (2002). Phys. Rev. Lett. 88, 176101.
Gai, Z., Li, X., Zhao, R.G. and Yang, W.S. (1998). Phys. Rev. B 57, 15060.
Gewirth, A.A. and Niece, B.K. (1997). Chem. Rev. 97, 1129.
Gimzewski, J.K., Joachim, C., Schlittler, R.R., Langlais, V., Tang, H. and
Johannsen, I. (1998). Science 281, 531.
Goldman, R.S., Feenstra, R.M., Briner, B.G., O’Steen, M.L. and Hauenstein, R.
J. (1996). Appl. Phys. Lett. 69, 3698.
Gong, Q., Offermans, P., Nötzel, R., Koenraod, P.M. and Loolter, J.H. (2004).
Appl. Phys. 85, 5697.
Grant, M.L., Swartzentruber, B.S., Bartelt, N.C. and Hannon, J.B. (2001). Phys.
Rev. Lett. 86, 4588.
Gregory, S. (1990). Phys. Rev. Lett. 64, 689.
Güntherodt, H.-J. and Wiesendanger, R., Eds. (1992). Scanning Tunneling
Microscopy Vol. I & II (Springer, Berlin).
Hamers, R.J. (1989). Annu. Rev. Phys. Chem. 531.
Hamers, R.J., Tromp, R.M. and Demuth, J.E. (1986). Phys. Rev. Lett. 56, 1972.
Hanaguri, T., Lupien, C., Kohsaka, Y., Lee, D.-H., Azuma, M., Takano, M.,
Takagi, H. and Davis, J.C. (2004). Nature 430, 1001.
Hasegawa, Y. and Avouris, Ph. (1993). Phys. Rev. Lett. 71, 1071.
Heer, R., Smoliner, J., Strasser, G. and Gornik, E. (1998). Appl. Phys. Lett. 73,
3138.
Heinrich, A.J., Lutz, C.P., Gupta, J.A. and Eigler, D.M. (2002). Science 298, 1381.
Heinze, S., Bode, M., Kubetzka, A., Pietzsch, O., Nie, X., Blügel, S. and Wiesen-
danger, R. (2000). Science 288, 1805.
Hendriksen, B.L.M. and Frenken, J.W.M. (2002). Phys. Rev. Lett. 89, 046101.
Hla, S.-W. and Rieder, K.-H. (2003). Annu. Rev. Phys. Chem. 54, 307.
Hoffman, J.E., McElroy, K., Lee, D.-H., Lang, K.M., Eisaki, H., Uchida, S. and
Davis, J.C. (2002a). Science 297, 1148.
Hoffman, J.E., Hudson, E.W., Lang, K.M., Madhavan, V., Eisaki, H., Uchida, S.
and Davis, J.C. (2002b). Science 295, 466.
Horch, S., Lorensen, H.T., Helveg, S., Lægsgaard, E., Stensgaard, I., Jacobsen,
K.W., Nørskov, J.K. and Besenbacher, F. (1999). Nature 398, 134.
Hornbaker, D.J., Kahng, S.-J., Misra, S., Smith, B.W., Johnson, A.T., Mele, E.J.,
Luzzi, D.E. and Yazdani, A. (2002). Science 295, 828.
Horn-von Hoegen, M. (1994). Appl. Phys. A 59, 503.
Hou, J.G., Jinlong, Y., Haiqian, W., Qunxiang, L., Changgan, Z., Hai, L., Wang,
B., Chen, D.M. and Qinshi, Z. (1999). Phys. Rev. Lett. 83, 3001.
Hovis, J.S. and Hamers, R.J. (1997). Surf. Sci. 402, 1.
Jaklevic, R.C. and Lambe, J. (1966). Phys. Rev. Lett. 17, 1139.
Jaklevic, R.C., Lambe, J., Mikkor, M. and Vassell, W.C. (1971). Phys. Rev. Lett.
26, 88.
Jensen, J.A., Rider, K.B., Chen, Y., Salmeron, M. and Somorjai, G.A. (1999). J.
Vac. Sci. Technol. B 17, 1080.
1022 P. Sutter
Johnson, M.B., Albrektsen, O., Feenstra, R.M. and Salemink, H.W.M. (1993).
Appl. Phys. Lett. 63, 2923.
Jung, T.A., Schlittler, R.R., Gimzewski, J.K., Tang, H. and Joachim, C. (1996).
Science 271, 181.
Kaiser, W.J. and Bell, L.D. (1988). Phys. Rev. Lett. 60, 1406.
Kasaya, M., Tabata, H. and Kawai, T. (1998). Surf. Sci. 400, 367.
Kawamura, M., Paul, N., Cherepanov, V. and Voigtländer, B. (2003). Phys. Rev.
Lett. 91, 096102.
Klijn, J., Sacharov, L., Meyer, C., Blügel, S., Morgenstern, M. and Wiesendan-
ger, R. (2003). Phys. Rev. B 68, 205327.
Kochanski, G.P. (1989). Phys. Rev. Lett. 62, 2285.
Komeda, T., Kim, Y., Kawai, M., Persson, B.N.J. and Ueba, H. (2002). Science
295, 2055.
Kubby, J.A. and Greene, W.J. (1992). Phys. Rev. Lett. 68, 329.
Kühnle, A., Linderoth, T.R., Hammer, B. and Besenbacher, F. (2002). Nature
415, 891.
Kuipers, L., Hoogeman, M.S. and Frenken, J.W.M. (1993). Phys. Rev. Lett. 71,
3517.
Kuk, Y. and Silverman, P.J. (1986). Appl. Phys. Lett. 48, 1597.
Lægsgaard, E., Österlund, L., Rasmussen, P.B., Stensgaard, I. and Besenbacher,
F. (2001). Rev. Sci. Instrum. 72, 3537.
Laracuente, A. and Whitman, L.J. (2001). Surf. Sci. 476, L247.
Lee, E.Y., Sirringhaus, H. and von Känel, H. (1994). Phys. Rev. B 50, 5807.
Lee, H.J. and Ho, W. (1999). Science 286, 1719.
Legrand, B., Grandidier, B., Nys, J.P., Stiévenard, D., Gérard, J.M. and Thierry-
Mieg, V. (1998). Appl. Phys. Lett. 73, 96.
Li, L., Tindall, C., Takaoka, O., Hasegawa, Y. and Sakurai, T. (1997). Phys. Rev.
B 56, 4648.
Lopinski, G.P., Moffatt, D.J., Wayner, D.D.M. and Wolkow, R.A. (1998). Nature
392, 909.
Lu, X., Grobis, M., Khoo, K.H., Louie, S.G. and Crommie, M.F. (2003). Phys.
Rev. Lett. 90, 096802.
Mayne, A.J., Avery, A.R., Knall, J., Jones, T.S., Briggs, G.A.D. and Weinberg,
W.H. (1993). Surf. Sci. 284, 247.
Medeiros-Ribeiro, G., Bratkovski, A.M., Kamins, T.I., Ohlberg, D.A.A. and
Williams, R.S. (1998). Science 279, 353.
Meinel, K., Wolter, H., Ammer, Ch., Beckmann, A. and Neddermeyer, H.
(1997). J. Phys. Condens. Matter 9, 4611.
Men, F.K., Packard, W.E. and Webb, M.B. (1988). Phys. Rev. Lett. 61, 2469.
Meyer, G. and Rieder, K.-H. (1997). Surf. Sci. 377–379, 1087.
Meyer, G. and Rieder, K.-H. (1998). MRS Bull. 23, 28.
Mitsui, T., Rose, M.K., Fomin, E., Ogletree, D.F. and Salmeron, M. (2002).
Science 297, 1850.
Mo, Y.-W., Swartzentruber, B.S., Kariotis, R., Webb, M.B. and Lagally, M.G.
(1989). Phys. Rev. Lett. 63, 2393.
Mo, Y.-W., Savage, D.E., Swartzentruber, B.S. and Lagally, M.G. (1990). Phys.
Rev. Lett. 65, 1020.
Mo, Y.-W., Kleiner, J., Webb, M.B. and Lagally, M.G. (1991). Phys. Rev. Lett. 66,
1998.
Moresco, F., Meyer, G., Rieder, K.-H., Ping, J., Tang, H. and Joachim, C. (2002).
Surf. Sci. 499, 94.
Muller, E.W. and Tsong, T.T. (1969). Field Ion Microscopy (Elsevier, New York).
Nagl, C., Haller, O., Platzgummer, E., Schmid, M. and Varga, P. (1994). Surf.
Sci. 321, 237.
Chapter 15 Scanning Tunneling Microscopy in Surface Science 1023
Nakayama, K. and Weaver, J.H. (1999). Phys. Rev. Lett. 82, 980.
Narayanamurti, V. and Kozhevnikov, M. (2001). Phys. Rep. 349, 447.
Nazin, G.V., Qiu, X.H. and Ho, W. (2003). Science 302, 77.
Nunes, G., Jr. and Amer, N.M. (1993). Appl. Phys. Lett. 63, 1851.
Olsson, F.E., Lorente, N. and Persson, M. (2003). Surf. Sci. 522, L27.
Olsson, F.E., Persson, M., Repp, J. and Meyer, G. (2005). Phys. Rev. B 71,
075419.
Pascual, J.I., Gomez-Herrero, J., Rogero, C., Baro, A.M., Sanchez-Portal, D.,
Artacho, E., Ordejon, P. and Soler, J.M. (2000). Chem. Phys. Lett. 321, 78.
Pfister, M., Johnson, M.B., Alvarado, S.F., Salemink, H.W.M., Marti, U., Martin,
D., Morier-Genoud, F. and Reinhard, F.K. (1995). Appl. Phys. Lett. 67, 1459.
Pietzsch, O., Kubetzka, A., Bode, M. and Wiesendanger, R. (2001). Science 292,
2053.
Prietsch, M. and Ludeke, R. (1991). Phys. Rev. Lett. 66, 2511.
Qin, X.R. and Lagally, M.G. (1997). Science 278, 1444.
Qin, X.R., Swartzentruber, B.S. and Lagally, M.G. (2000). Phys. Rev. Lett. 85,
3660.
Rasmussen, P.B., Hendriksen, B.L.M., Zeijlemaker, H., Ficke, H.G. and Frenken,
J.W.M. (1998). Rev. Sci. Instrum. 69, 3879.
Repp, J., Meyer, G., Olsson, F.E. and Persson, M. (2004). Science 305, 493.
Repp, J., Meyer, G., Stojkovic, S.M., Gourdon, A. and Joachim, C. (2005). Phys.
Rev. Lett. 94, 026803.
Rippard, W.H. and Buhrmann, R.A. (2000). Phys. Rev. Lett. 84, 971.
Robinson, K.M., Robinson, I.K. and O’Grady, W.E. (1992). Surf. Sci. 262, 387.
Rößler, M., Geng, P. and Wintterlin, J. (2005). Rev. Sci. Instrum. 76, 023705.
Rost, M.J., Crama, L., Schakel, P., van Tol, E., van Velzen-Williams, G.B.E.M.,
Overgauw, C.F., ter Horst, H., Dekker, H., Okhuijsen, B., Seynen, M., Vijft-
igschild, A., Han, P., Katan, A.J., Schoots, K., Schumm, R., van Loo, W.,
Oosterkamp, T.H. and Frenken, J.W.M. (2005). Rev. Sci. Instrum. 76,
053710.
Rubin, M.E., Medeiros-Ribeiro, G., O’Shea, J.J., Chin, M.A., Lee, E.Y., Petroff,
P.M. and Narayanamurti, V. (1996). Phys. Rev. Lett. 77, 5268.
Sajoto, T., O’Shea, J.J., Bhargava, S., Leonard, D., Chin, M.A. and Narayana-
murti, V. (1995). Phys. Rev. Lett. 74, 3427.
Salemink, H. and Albrektsen, O. (1991). J. Vac. Sci. Technol. B 9, 779.
Sanders, L.M., Stumpf, R., Mattson, T.R. and Swartzentruber, B.S. (2003). Phys.
Rev. Lett. 91, 206104.
Schaub, R., Thostrup, P., Lopez, N., Stensgaard, I., Nørskov, J.K. and
Besenbacher, F. (2001). Phys. Rev. Lett. 87, 266104.
Schaub, R., Wahlström, E., Rønnau, A., Lægsgaard, E., Stensgaard, I. and
Besenbacher, F. (2003). Science 299, 377.
Schintke, S. and Schneider, W.-D. (2004). J. Phys: Cond. Matter 16, R49.
Schunack, M., Linderoth, T.R., Rosei, F., Lægsgaard, E., Stensgaard, I. and
Besenbacher, F. (2002). Phys. Rev. Lett. 88, 156102.
Sirringhaus, H., Lee, E.Y. and von Känel, H. (1994). Phys. Rev. Lett. 73, 577.
Smith, D.P.E. and Binnig, G. (1986). Rev. Sci. Instrum. 57, 2630.
Stipe, B.C., Rezaei, M.A. and Ho, W. (1998a). Science 280, 1732.
Stipe, B.C., Rezaei, M.A. and Ho, W. (1998b). Science 279, 1907.
Stroscio, J.A. and Celotta, R.J. (2004). Science 306, 242.
Sutter, P., Schick, I., Ernst, W. and Sutter, E. (2003a). Phys. Rev. Lett. 91,
176102.
Sutter, P., Zahl, P., Sutter, E. and Bernard, J.E. (2003b). Phys. Rev. Lett. 90,
166101.
Sutter, E., Sutter, P. and Bernard, J.E. (2004). Appl. Phys. Lett. 84, 2262.
1024 P. Sutter