ArticlePDF Available

Role of Molecular Orbitals Near the Fermi Level in the Excitation of Vibrational Modes of a Single Molecule at a Scanning Tunneling Microscope Junction

Authors:

Abstract and Figures

Inelastically tunneled electrons from the tip of a scanning tunneling microscope were used to induce S-S bond dissociation of a (CH(3)S)(2) and lateral hopping of a CH(3)S on Cu(111) at 4.7 K. Both experimental results and theoretical calculations confirm that the excitation mechanism of the vibrationally induced chemistry reflects the projected density of states of molecular orbitals that appear near the Fermi level as a result of the rehybridization of the orbitals between the adsorbed molecules and the substrate metal atoms.
Content may be subject to copyright.
Role of Molecular Orbitals Near the Fermi Level in the Excitation of Vibrational Modes
of a Single Molecule at a Scanning Tunneling Microscope Junction
Michiaki Ohara,
1,2
Yousoo Kim,
1
Susumu Yanagisawa,
3
Yoshitada Morikawa,
3,4
and Maki Kawai
1,2
1
Surface Chemistry Laboratory, RIKEN, Saitama 351-0198, Japan
2
Department of Advanced Materials Science, The University of Tokyo, Chiba 277-8561, Japan
3
The Institute of Scientific and Industrial Research, Osaka University, Osaka 567-0047, Japan
4
Research Institute for Computational Sciences, National Institute of Advanced Industrial Science and Technology,
Ibaraki 305-8568, Japan
(Received 1 February 2007; published 2 April 2008)
Inelastically tunneled electrons from the tip of a scanning tunneling microscope were used to induce S-
S bond dissociation of a CH3S2and lateral hopping of a CH3Son Cu(111) at 4.7 K. Both experimental
results and theoretical calculations confirm that the excitation mechanism of the vibrationally induced
chemistry reflects the projected density of states of molecular orbitals that appear near the Fermi level as a
result of the rehybridization of the orbitals between the adsorbed molecules and the substrate metal atoms.
DOI: 10.1103/PhysRevLett.100.136104 PACS numbers: 68.37.Ef, 68.35.Ja, 68.43.Fg, 68.43.Pq
Mode-selective chemistry, whereby a specific chemical
bond of a molecule is excited to select a desired reaction
pathway, has been widely investigated in the gas phase
[1,2]. The scanning tunneling microscope (STM) provides
a way to achieve mode-selective chemistry of single mole-
cules adsorbed on surfaces because inelastically tunneled
electrons from the STM tip can excite molecular vibrations
that lead to a variety of dynamic motions and chemical
reactions [316]. Molecular vibrations excited through
inelastic electron tunneling processes can be accompanied
by changes in the differential conductance of a molecule.
Inelastic electron tunneling spectroscopy with STM (STM-
IETS) [17], which records the second derivative of the
tunneling current (I) with respect to bias voltage (V), has
proven useful for detecting vibrational signals of individual
molecules. However, it does not always reveal all the
vibrational modes in the spectra, nor is it applicable to
mobile or reactive molecules [46,11,15,17]. As an alter-
native to the STM-IETS, action spectroscopy, which mea-
sure the response of vibrationally mediated molecular
motion to applied bias voltage, has been used for obtaining
vibrational spectra of mobile molecules. Sainoo et al. have
shown that action spectroscopy can even detect vibrational
modes that are not visible using STM-IETS [11].
Persson and Baratoff [18] developed a theory of STM-
IETS using an adsorbate-induced resonance model, and
before the first experiment by Stipe et al. [3], they pre-
dicted that the STM-IET spectrum would exhibit a peak-
shaped signal at the particular bias voltage where the
vibrational excitation of a molecule takes place.
Similarly, for the action spectrum obtained from the reac-
tion induced by vibrational excitation, Ueba and Persson
[19] have shown that the second derivative of the reaction
rate with respect to the bias voltage provides direct access
to the vibrational density of states. Moreover, the width of
this action spectrum displays the intrinsic vibrational
broadening of a single adsorbed molecule on the surface.
Although action spectra offer rich information of this
kind, the precise mechanism by which a particular vibra-
tional mode induces molecular reactions by inelastic tun-
neling electrons has not yet been experimentally examined.
Many theoretical attempts to explain the tunneling
electron-vibration coupling [20] in a molecule have been
based on a resonance model, in which an incident electron
becomes temporarily captured by molecular orbitals (MO)
localized in a resonant state resulting in excitation of
molecular vibrations. Thus, a detailed knowledge of the
MOs of the adsorbate is necessary for understanding the
underlying mechanism of electron-vibration coupling and
for predicting which vibrational mode is actually excited
by tunneling electrons.
This Letter reports on quantitative investigations of the
vibrationally induced bond dissociation of an isolated
CH3S2molecule and the lateral hopping of an isolated
CH3Smolecule adsorbed on the Cu(111) surface with a
low-temperature STM. Based on the experimental findings
and on density functional theory (DFT) calculations, we
discuss how the spatial distribution of MOs of an adsorbate
at resonant states can affect electron-vibration coupling
and thereby affect the resulting molecular motions and
reactions.
All STM measurements were carried out at 4.7 K [21,22]
with a low-temperature STM (LT-STM, Omicron GmbH)
in an ultrahigh-vacuum chamber ( <31011 Torr).
Typical conditions for obtaining STM images were
Vsample 20 mV and Itunnel 0:2nA. The clean
Cu(111) surface was exposed to CH3S2vapor at a tem-
perature below 50 K. After the STM image was taken, the
STM tip was positioned over the center of a target mole-
cule. The feedback loop was then turned off, and the
tunneling electrons with the defined tunneling current
were injected into the molecule. During electron injection,
this tunneling current stays constant except when the target
molecule reacts, for example, by bond dissociation or
PRL 100, 136104 (2008) PHYSICAL REVIEW LETTERS week ending
4 APRIL 2008
0031-9007=08=100(13)=136104(4) 136104-1 ©2008 The American Physical Society
molecular hopping, in which case the tunneling current
suddenly decreases. In the present study, we calculated the
mean time required for bond dissociation (before molecu-
lar hopping actually occurred) for each applied bias volt-
age over 10 trial events with the STM tip in the same
position. Multiplying the value of the current (I) and the
mean time taken for the reaction (T) gives the number of
electrons necessary for the dissociation and for the hop-
ping, from which the reaction yield (R) is determined. The
reaction probability was plotted against applied bias volt-
age to produce an action spectrum. DFT calculations were
carried out using the program package DMOL3in the
Materials Studio of Accelrys Inc. [23,24] for optimizing
the adsorption system, and the program code STATE
(Simulation Tool for Atom TEchnology) [2527] for the
projected density of states (PDOS) analysis and the draw-
ing of MOs for an isolated CH3S2and CH3Son Cu(111).
Individual CH3S2molecules on the Cu(111) surface
appear as elliptic protrusions in STM images, as shown in
Fig. 1(a). The intersection points of the mesh in Fig. 1(a)
indicate the center positions of the Cu atoms on the surface
[21]. By injecting tunneling electrons into the target mole-
cule, we find that a single CH3S2molecule is broken into
two identical ball-shaped protrusions, implying S-S bond
dissociation to form two CH3Smolecules [Fig. 1(b)]. For
the adsorption site of CH3S, fcc and hcp hollow sites are
equally eligible [28].
Action spectra for dissociation of CH3S2and CD3S2
are shown in Fig. 1(c). Each spectrum shows one clear
threshold voltage at 357.5 mV for CH3S2and at 275 mV
for CD3S2. The threshold voltage at 357.5 mV corre-
sponds to the reported vibrational excitation energy of a
C-H stretching mode (C-H) of the CH3group in the
molecule. In high-resolution electron energy loss spectros-
copy (HREELS), the vibrational frequencies of the C-H
mode are observed at 365 meV for monolayers of CH3S2
on Au(111) and at 366 meV for those on Cu(100) [29,30].
Moreover, the shift of the threshold voltage from 357.5 to
275 mV reasonably corresponds to the isotope shift of
C-Hto C-D. Note that the small increase in the
reaction yield (-- and --) clearly appears at around
410 mV for CH3S2and at 330 mV for CD3S2, respec-
tively [31]. In the negative bias voltage region, however,
the reaction yield was 1:61012 and 0:91012 for
Vsample 800 mV and 600 mV, respectively. This
negligibly small dissociation yield in the negative bias
voltage region suggests that the reaction yield is strongly
influenced by whether the electron was injected to the
molecule or extracted from it.
Excited molecular vibrations can cause dynamic pro-
cesses by overcoming the potential barrier along the reac-
tion coordinate (RC) of a specific adsorbate motion. It has
been shown that the vibrational mode along the RC can be
excited not only directly but also indirectly through cou-
pling to the higher frequency (HF) vibrational mode ex-
cited by tunneling electrons [9]. In the present study, the
thresholds observed in the action spectra as a sharp peak at
357.5 meV due to C-Hand as a shoulder at 410 meV
due to C-HS-Sreveal that S-S bond dissociation
in a CH3S2molecule is induced by excitation of two HF
modes, not by direct excitation of the S-Smode, which
is the RC mode for S-S bond dissociation. This is the first
observation of an indirect pathway for intramolecular bond
dissociation on a surface attained by means of tunneling
electrons. So far, only a direct pathway, resulting from the
direct excitation of a specific RC mode, has been reported,
in experiments involving O-O bond dissociation of O2on
Pt(111) [12] and C-H bond dissociation of trans-2-butene
on Pd(110) [15].
Figure 2(a) shows an STM image of two CH3Smole-
cules obtained by breaking the S-S bond of a CH3S2
molecule on Cu(111). In previous work, we reported that
the injection of tunneling electrons can induce individual
CH3Smolecules to hop randomly between the fcc and hcp
hollow sites of Cu(111) [22], as shown in Figs. 2(b)2(d).
An action spectrum for the hopping of CH3Sis shown in
Fig. 2(e). In the spectrum, threshold voltages are observed
at both 85 mV, the absolute value of which corresponds
to the vibrational excitation energy of the C-S stretching
mode (C-S). It is important to note that the reaction
yield was identical for both bias polarities, which is very
different from what occurs in CH3S2dissociation.
According to DFT calculations, the vibrational excitation
energy of C-Sis 85 meV for CS3Son a Au substrate and
81 meV for CS3Son a Cu substrate [26,30]. Experi-
mentally, the C-Smode of various alkylthiolates has
been observed at similar energies in HREELS: for ex-
ample, 88 meV for methylthiolate on Au(111) and
81 meV for butylthiolate and hexylthioate on Au(111)
[32]. Since the frustrated translation and/or rotation modes
are the RC mode for the motion [7,9], as for the dissocia-
FIG. 1. Topographic STM images of CH3S2molecules on Cu
(111) before (a) and after (b) injection of tunneling electrons.
(c) Action spectra for the S-S bond dissociation of CH3S2and
CD3S2molecules. Initial current was set to 4 nA. The lower
traces with ‘‘--’’ and ‘‘--’’ represent the slopes of the reaction
for CH3S2and CD3S2molecules, respectively. In both spec-
tra, each data point consists of 10 trial events, and thus to obtain
these spectra, a total of 130 molecules for CH3S2and of 110
molecules for CD3S2respectively were dissociated by tunnel-
ing electrons, with the STM tip condition kept constant.
PRL 100, 136104 (2008) PHYSICAL REVIEW LETTERS week ending
4 APRIL 2008
136104-2
tion of the S-S bond in a CH3S2molecule, the hopping of
CH3Sis evidently also the result of an indirect mechanism
of RC mode excitation via anharmonic coupling between
HF C-Sand RC modes.
As mentioned above, the S-S bond dissociation of
CH3S2is induced by vibrational excitation of the
C-Hmode, whereas lateral hopping of CH3Sis not,
although both molecules involve C-H bonds of the methyl
(CH3) group. Sainoo et al. have pointed out that vibrational
modes observed in action spectra can be explained by
means of a resonant model of electron-vibration coupling
[11]. When a molecule is chemisorbed onto the metal
surface, it is distorted by molecule-metal bond formation,
and the rehybridized MO consists of several fundamental
MOs that appear around the Fermi level (Ef). Thus, to
address the applicability of this resonant model to our
system, we calculated the projected density of states
(PDOS) and the spatial distribution of the MOs for each
molecular adsorbate, as shown in Figs. 3(a) 3(d).
Figure 3(a) shows that the lowest unoccupied molecular
orbital (LUMO) of CH3S2is located just above Ef, and it
also contributes to states near Ef. In addition, the small
contributions from the LUMO 1and LUMO 2states
are broadly distributed over the energy region shown. The
reaction yield for dissociation showed distinctly different
dependence on polarity, where the contribution of the
LUMO state apparently dominates, obscuring the rela-
tively small contributions from LUMO 1and LUMO
2as well as from the HOMO state which is distributed
along Cu-S bond. In the case of CH3S, the HOMO is
located just below Efand, in addition to the LUMO and
LUMO 1, it also contributes to states near Ef, as shown
in Fig. 3(c). The reaction yield for hopping is identical for
the applied bias voltages of both polarities, which parallels
the homogeneous distribution of the DOS across the Efof
the LUMO and/or LUMO 1, while that of the HOMO
contributes much more in the occupied state. The spatial
distribution of individual MOs for CH3S2and CH3Sis
depicted in Fig. 3(b) and 3(d), respectively. As shown in
Fig. 3(b), the LUMO of CH3S2is clearly localized at both
the S-S and C-H bonds in the molecule, suggesting that a
resonantly captured electron in the LUMO will influence
the S-S and C-H bond. However, in the case of CH3S
[Fig. 3(d)], there is no contribution of the MOs to the C-
H bond in the molecule. This explains why CH3S2shows
aC-Hsignal in the action spectrum, whereas CH3Sdoes
not. Figure 3(d) also reveals that the LUMO and LUMO
1are rather localized at the C-S bond in CH3S, which
supports the experimental observation of a C-Ssignal in
the action spectrum of the molecule. It is important to note
that the LUMO 1and LUMO 2of CH3S2are like-
wise localized at the C-S bond in the molecule, as shown in
Fig. 3(b). One would therefore expect that the C-Smode
also would be detected in the action spectrum of CH3S2.
However, in the present study, this was not the case.
Especially in the positive bias region, the cross-section
with the LUMO state dominates, and as a result, the con-
FIG. 3 (color online). (a) PDOS of an isolated CH3S2mole-
cule on Cu(111). (b) The spatial distributions of LUMO 2,
LUMO 1, LUMO, and HOMO for an isolated CH3S2mole-
cule on Cu(111). MOs on C-H bond in CH3S2molecule are
indicated by arrows. (c) PDOS of an isolated CH3Smolecule on
Cu(111). (d) The spatial distributions of LUMO 1, LUMO,
and HOMO for an isolated CH3Smolecule on Cu(111). H, C, S,
and Cu atoms are drawn in blue, yellow, green, and gray,
respectively.
FIG. 2. (a) –(d) Sequential STM images of the hopping of a
CH3Smolecule on Cu (111). (e) Action spectrum for the hop-
ping of CH3Smolecules in both positive and negative bias.
Initial current was set to 4 nA. The lower trace with ‘‘--’
represents the slope of the hopping for CH3Smolecules in both
positive and negative bias. In the spectrum, each data point
consists of 10 trial events; thus, to obtain the spectrum in the
positive bias region, a total of 200 CH3Smolecules were
counted, with the STM tip maintained in a constant state. A
similar number of trials was undertaken for the negative bias
region.
PRL 100, 136104 (2008) PHYSICAL REVIEW LETTERS week ending
4 APRIL 2008
136104-3
tribution of LUMO 1and LUMO 2diminishes pro-
portionally. Thus, detailed interpretation of the spatial
distribution of individual MOs can successfully explain
our experimental action spectra, where the C-Hmode
is active for S-S bond dissociation of CH3S2and the
C-Smode is active for lateral hopping of CH3Sby
injecting inelastically tunneled electrons from the tip of
an STM.
In conclusion, a combination of quantitative STM ex-
periments and theoretical calculations has been used to
investigate the S-S bond dissociation of an isolated
CH3S2molecule and the lateral hopping of an isolated
CH3Smolecule on Cu(111) in terms of excitation of mo-
lecular vibrations induced by inelastically tunneled elec-
trons. Our experiments reveal that vibrational modes are
resonantly excited through a temporal occupation of the
MOs by the tunneling electrons and that the molecular
vibrations are selectively excited depending on the spatial
distribution and the population of the MOs near Efby
hybridization with the substrate metal.
Our present finding that molecular vibrations can be
selectively excited depending on the spatial distribution
of the MOs near Efoffers a way to systematically realize
mode-selective chemistry of individual molecules ad-
sorbed on surfaces, and it also may be applied to character-
ize molecules bridging two electrodes [33 35].
The present work was supported in part by the Grant-in-
Aid for Scientific Research on Priority Areas, ‘‘Electron
Transport Through a Linked Molecule in Nano-Scale’’
(Grant No. 17069006), from the Ministry of Education,
Culture, Sports, Science and Technology, and ‘‘Control
and Application of Nano-Structural Materials for
Advanced Data Processing and Communications,’’
CREST program from Japan Science and Technology cor-
poration (JST). We are grateful for the computational
resources of the RIKEN Super Combined Cluster (RSCC).
[1] A. Sinha, R. L. Vander Wal, and F. F. Crim, J. Chem. Phys.
92, 401 (1990).
[2] P. R. Fleming, M. Li, and T. R. Rizzo, J. Chem. Phys. 94,
2425 (1991).
[3] B. C. Stipe, M. A. Rezaei, and W. Ho, Science 279, 1907
(1998).
[4] B. C. Stipe, M. A. Rezaei, and W. Ho, Phys. Rev. Lett. 81,
1263 (1998).
[5] J. Gaudioso, H. J. Lee, and W. Ho, J. Am. Chem. Soc. 121,
8479 (1999).
[6] W. Ho, J. Chem. Phys. 117, 11033 (2002), and references
therein.
[7] J. I. Pascual et al., Nature (London) 423, 525 (2003).
[8] Y. Kim, T. Komeda, and M. Kawai, Jpn. J. Appl. Phys. 41,
4924 (2002).
[9] T. Komeda et al., Science 295, 2055 (2002).
[10] C. Matsumoto et al., Surf. Sci. 587, 19 (2005).
[11] Y. Sainoo et al., Phys. Rev. Lett. 95, 246102 (2005).
[12] B. C. Stipe et al., Phys. Rev. Lett. 78, 4410 (1997).
[13] S.-W. Hla, L. Bartels, G. Meyer, and K.-H. Rieder, Phys.
Rev. Lett. 85, 2777 (2000).
[14] Y. Kim, T. Komeda, and M. Kawai, Surf. Sci. 502,7
(2002).
[15] Y. Kim, T. Komeda, and M. Kawai, Phys. Rev. Lett. 89,
126104 (2002).
[16] T. Komeda et al., J. Chem. Phys. 120, 5347 (2004).
[17] B. C. Stipe, M. A. Rezaei, and W. Ho, Science 280, 1732
(1998).
[18] B. N. J. Persson and A. Baratoff, Phys. Rev. Lett. 59, 339
(1987).
[19] H. Ueba and B. N. J. Persson, Phys. Rev. B 75, 041403(R)
(2007).
[20] H. Ueba, Surf. Rev. Lett. 10, 771 (2003), and references
therein.
[21] M. Ohara, Y. Kim, and M. Kawai, Langmuir 21, 4779
(2005).
[22] M. Ohara, Y. Kim, and M. Kawai, Chem. Phys. Lett. 426,
357 (2006).
[23] B. Delley, J. Chem. Phys. 92, 508 (1990).
[24] B. Delley, J. Chem. Phys. 113, 7756 (2000).
[25] Y. Morikawa, Phys. Rev. B 51, 14802 (1995).
[26] T. Hayashi, Y. Morikawa, and H. Nozoye, J. Chem. Phys.
114, 7615 (2001).
[27] S. Yanagisawa and Y. Morikawa, Chem. Phys. Lett. 420,
523 (2006).
[28] M. Ohara, Y. Kim, and M. Kawai, Jpn. J. Appl. Phys. 45,
2022 (2006).
[29] R. G. Nuzzo, B. R. Zegarski, and L. H. Dubois, J. Am.
Chem. Soc. 109, 733 (1987).
[30] B. A. Sexton and G. L. Nyberg, Surf. Sci. 165, 251 (1986).
[31] Second thresholds at 410 mV for CH3S2and at 330 mV
for CD3S2do not match with any of the vibrational
normal modes of CH3S2and CD3S2, respectively.
These higher voltages may correspond to a combination
mode of C-Hand S-S stretching [S-S] modes of
CH3S2. This is the first observation of a combination
mode related to bond dissociation. According to DFT
calculations, the vibrational excitation energies of the
S-Smode of isolated CH3S2and CD3S2molecules
on Cu(111) are 53 and 51 meV, respectively, which
suggests the higher threshold corresponds to the vibra-
tional excitation energy of the C-HSSmode.
[32] H. S. Kato et al., J. Phys. Chem. B 106, 9655 (2002).
[33] R. H. M. Smit et al., Nature (London) 419, 906 (2002).
[34] W. H. A. Thijssen, D. Djukic, A. F. Otte, R. H. Bremmer,
and J. M. van Ruitenbeek, Phys. Rev. Lett. 97, 226806
(2006).
[35] M. Kiguchi, R. Stadler, I. S. Kristensen, D. Djukic, and
J. M. van Ruitenbeek, Phys. Rev. Lett. 98, 146802 (2007).
PRL 100, 136104 (2008) PHYSICAL REVIEW LETTERS week ending
4 APRIL 2008
136104-4
... In both spectra, each data point consists of 10 trial events, and thus to obtain these spectra, a total of 130 molecules for (CH 3 S) 2 and of 110 molecules for (CD 3 S) 2 respectively were dissociated by tunneling electrons, with the STM tip condition kept constant. Transcribed from Ref.32 ...
Article
Probe microscopes are now one of the essential observation tools for materials science research. Scanning tunneling microscope is not limited to surface morphology observations, keeping the highest spatial resolution as an analytical tool with sub-Å spatial resolution in the plane and the availability as a tool for spectroscopy provides the STM a unique standpoint. Furthermore, the possibility of using it as a tool for manipulating surface atoms and molecules was thoroughly pursued. After 40 years since its invention, I would provide some of the amazing innovations that this analytical instrument has brought to the field of surface science.
... Timedependent transition coefficient from all the occupied states i to LUMO+9 demonstrates that such a charge transfer is mainly from the HOMO−1 state contributed mainly by the Ag 20 nanocluster, see Figure 5c. These findings indicate that the charge is indirectly transferred from the Ag 20 to the CO 2 species via IET, [30,42,57] as demonstrated by the significantly larger energy of the applied pulse than the (HOMO−1)−(LUMO+9) energy gap, by about 3.25 eV. Here, the photogenerated hot electrons within the metal nanoparticle transfer to the unoccupied states of adsorbates via inelastic tunneling, where an energy barrier may exist between the related orbitals. ...
Article
Full-text available
Understanding hot carrier dynamics between plasmonic nanomaterials and its adsorbate is of great importance for plasmon‐enhanced photoelectronic processes such as photocatalysis, optical sensing and spectroscopic analysis. However, it is often challenging to identify specific dominant mechanisms for a given process because of the complex pathways and ultrafast interactive dynamics of the photoelectrons. Here, using CO2 reduction as an example, the underlying mechanisms of plasmon‐driven catalysis at the single‐molecule level using time‐dependent density functional theory calculations is clearly probed. The CO2 molecule adsorbed on two typical nanoclusters, Ag20 and Ag147, is photoreduced by optically excited plasmon, accompanied by the excitation of asymmetric stretching and bending modes of CO2. A nonlinear relationship has been identified between laser intensity and reaction rate, demonstrating a synergic interplay and transition from indirect hot‐electron transfer to direct charge transfer, enacted by strong localized surface plasmons. These findings offer new insights for CO2 photoreduction and for the design of effective pathways toward highly efficient plasmon‐mediated photocatalysis. Upon illumination, hot electrons generated by the plasmon decay around the silver clusters transfer into the molecular orbitals of CO2, resulting in activation of the asymmetrical stretching and bending mode of CO2 and leading to its final splitting. Cooperative interaction and transition from indirect hot‐electron transfer to direct charge transfer cause the nonlinear relationship between laser intensity and reaction rate.
... The examples of switching include cis-trans isomerization induced with an electric field [95], [96] or tunneling electrons [97], [98]; conformational switching induced with an electric field [99], direct electron injection [100], [101] or electron injection via surface states [102]. [109] and for bigger molecules as well [110], [111]. Apart from bond breaking, STM can also be used to form chemical bonds in a local and controlled manner. ...
Article
This paper describes a single-molecule study of N-methylbutylamine molecular rotors supported on a Cu(111) surface. It is first demonstrated that the chirality of the individual rotating molecules can be directly determined by scanning tunneling microscopy (STM) imaging and understood with density functional theory (DFT) simulations. Tunneling electrons from the STM tip are then utilized to excite vibrational modes of the molecule that drives the rotational motion. Experimental action spectra were used to demonstrate that the electrically induced rotational motion of N-methylbutylamine occurs above 360 meV, which coincides with C–H stretching vibrational modes. The measurements also reveal that, above this 360 meV threshold, the excitation occurs via a one-electron process. DFT calculations indicated that the rotation barrier is over an order of magnitude smaller, meaning that the rotor is excited via high-energy vibrational modes that then couple to the low energy rotational mode. Furthermore, by adjusting the electron flux, individual rotational motions between the six different stable orientations of the molecule on the Cu(111) surface were monitored in real time. It was found that, for most STM tips used to electrically excite the rotors, the rotation of one enantiomer is faster than the other. This confirms an earlier report that STM tips can themselves be chiral and illustrates the fact that diastereomerism arising from a chiral STM tip interacting with a chiral molecule can lead to significant physical differences in the rotation rates of R versus S molecular rotors. This result has ramifications for interpreting the data from experiments where nanoscale electrical contacts to chiral molecules are made in devices like break junctions and scanning probe experiments.
Article
Breaking bonds selectively in molecules is vital in many chemistry reactions and custom nanoscale device fabrications. The scanning tunneling microscope (STM) has proved to be an ideal tool to initiate and view bond-selective chemistry at the single-molecule level, offering opportunities for the further study of the dynamics in single molecules on metal surfaces. We demonstrate H─HS and H─S bond breaking on Au(111) induced by tunneling electrons using low-temperature STM. An experimental study combined with theoretical calculations shows that the dissociation pathway is facilitated by vibrational excitations. Furthermore, the dissociation probabilities of the two different dissociation processes are bias dependent due to different inelastic-tunneling probabilities, and they are also closely linked to the lifetime of inelastic-tunneling electrons. Combined with time-dependent ab initio nonadiabatic molecular dynamics simulations, the dynamics of the injected electron and the phonon-excitation-induced molecule dissociation can be understood at the atomic scale, demonstrating the potential application of STM for the investigation of excited-state dynamics of single molecules on surfaces.
Article
Surface and interfacial water is ubiquitous in nature and modern technology. It plays vital roles in an extremely wide range of basic and applied fields including physics, chemistry, environmental science, material science, biology, geology, etc. Therefore, the studies of surface/interfacial water lies at the heart of water science. When water molecules are brought into contact with various materials, a variety of phenomena can show up, such as wetting, corrosion, lubrication, nanofluidics, ice nucleation, to name just a few. Due to the complexity of hydrogen-bonding interactions between water molecules and the competition between water-water interaction and water-solid interaction, surface/interfacial water is very sensitive to local environment, which makes it necessary to study the structure and dynamics of water at the molecular level. In recent years, the development of new scanning probe techniques allows detailed real-space research on surface/interfacial water at single-molecule or even submolecular scale. In Section 2, several representative scanning probe techniques and their applications in surface/interfacial water are reviewed. The first one is ultra-high vacuum scanning tunneling microscopy, which allows molecular imaging of single water molecules, water clusters, wetting layers, and even water multilayers on metal surfaces as well as ultrathin insulating films. Based on scanning tunneling microscopy, the single-molecule vibrational spectroscopy can be further developed to probe the vibration and movement of individual water molecules, which assist us in understanding water diffusion, dissociation and quantum nature of hydrogen bonds. As a versatile tool at liquid/solid interfaces, electrochemical scanning tunneling microscopy opens up the unique possibility of probing the double electric layer and identifying water dynamics during electrochemical reactions. Moreover, non-contact atomic force microscopy yields higher resolution than scanning tunneling microscopy, such that the topology of hydrogen-bonding skeleton of surface/interfacial water and even the degree of freedom of hydrogen atoms can be discerned. To conclude this review, the challenges and future directions of this field are discussed in Section 3, focusing on non-invasive imaging under ambient conditions, ultrafast molecular dynamics, and novel structures under high pressures.
Article
Full-text available
The reversible rotation of a single isolated acetylene molecule between two diagonal sites on the Cu(100) surface at 8 K was induced and monitored with tunneling electrons from a scanning tunneling microscope (STM). Excitation of the C—H (C—D) stretch mode of C2H2 ( C2D2) at 358 meV (266 meV) led to a 10-fold (60-fold) increase in the rotation rate. This increase is attributed to energy transfer from the C—H (C—D) stretch mode to the hindered rotational motion of the molecule. Inelastic electron tunneling spectroscopy with the STM provides the energies of the stretch modes and allows a quantitative determination of the inelastic tunneling and coupling probabilities.
Article
Full-text available
The tunneling current from a scanning tunneling microscope was used to image and dissociate individual O2 molecules on fcc three-fold sites of the Pt(111) surface in the temperature range of 40 K to 150 K. After dissociation, the two oxygen atoms are found one to three lattice constants apart on both fcc and hcp three-fold sites. Conversion from hcp to fcc sites was possible with additional excitation by the tunneling current. The dissociation rate was measured as a function of current and voltage and found to vary as I^0.8±0.2, I^1.8±0.2, and I^2.9±0.3 for sample biases of 0.4, 0.3, and 0.2 V, respectively. These rates are explained using a general model for dissociation induced by intramolecular vibrational excitations via resonant inelastic electron tunneling.(B.C. Stipe, M.A. Rezaei, W. Ho, S. Gao, M. Persson, and B.I. Lundqvist, to be published.) Chemisorbed oxygen molecules were also found on bridge sites and at steps with (111) microfacets. The induced dissociation and desorption of these species with tunneling electrons will also be discused.
Article
Recent progress in the study of motions and reactions of single adsorbed molecules on metal surfaces induced by inelastic tunneling electrons with a scanning tunneling microscope (STM) is given an overview, with the focus on our current theoretical understanding of the elementary processes behind these phenomena. The selected topics include rotation and dissociation of O2 on Pt(111), rotation of a C2H(D)2 on Cu(100), lateral hopping of CO on Pd(110), lateral translation and desorption of NH3 on Cu(100), and controlled manipulation of chemical transformation as well as bimolecular reaction of coadsorbed species on metal surfaces. Brief descriptions are presented of how an adsorbate to overcome the potential barrier for motion and reaction by incoherent stepwise and coherent single multistep climbing of the vibrational ladders in the potential well along the reaction coordinate, and indirect excitation of the reaction coordinate mode via anharmonic coupling to the vibrational mode excited by tunneling current. Elementary processes of the mode-selective control of different motions are also discussed in conjunction with a recent experimental result of lateral hopping and desorption of a single NH3 molecule on Cu(100). Although still at a premature stage, these novel phenomena open a new world of "nano-surface-science," in which the manipulation and reaction of single adsorbates, and synthesis of a new molecular system are realized by a selective excitation of the relevant vibrational mode by tunneling electrons with an STM.
Article
Controlled chemical reaction of single trans-2-butene molecules on the Pd(110) surface was realized by dosing tunneling electrons from the tip of a scanning tunneling microscope at 4.7 K. The reaction product was identified as a 1,3-butadiene molecule by inelastic electron tunneling spectroscopy. Threshold voltage for the reaction is ∼365 mV, which coincides with the vibrational excitation of the C-H stretching mode. The reaction was ascertained to be caused by C-H bond dissociation by multiple vibrational excitations of the C-H stretching mode via inelastic electron tunneling process.
Article
Scanning tunneling microscope images of isolated molecules of trans-2-butene (C4H8) and 1,3-butadiene (C4H6) adsorbed on the Pd(1 1 0) surface are presented at 4.7 K. The former appears as a dumbbell-shaped protrusion perpendicular to the Pd row, while the latter appears as an elliptical protrusion parallel to the Pd row. By dosing tunneling electrons, the trans-2-butene molecule is transformed into a 1,3-butadiene molecule via dehydrogenation reaction. The threshold voltage (∼365 mV) is so small that the electronic excitation to the C–H antibonding state cannot be attributed to the mechanism. Instead, we consider that the multiple excitation of C–H stretching mode by inelastically tunneled electron is responsible for the dehydrogenation.
Article
The adsorption of individual acetylene molecules at a Pd(111) surface has been studied with scanning tunneling microscopy (STM) at 4.7K. The adsorbed acetylene molecules appear in the STM images as a combination of a protrusion and a pair of depressions. To determine the adsorption site and configuration of the molecule, we performed rotational manipulation of the molecule by injecting tunneling electrons into the molecule as well as performing atomically resolved STM imaging. The STM images revealed that the molecules have a total of six equivalent adsorption orientations on Pd(111), indicating three equivalent rotational states on two (both fcc and hcp) 3-fold hollow adsorption sites. From the atomically resolved STM images of the surface with a monatomic step edge, we have distinguished the two types of 3-fold hollow sites and confirmed that the fcc site is more stable than the hcp site at a low temperature.
Article
Vibrational (EELS) and TDS data for methyl mercaptan (CH3SH), dimethyl sulfide (CH3)2S and dimethyl disulfide (CH3S)2 are analyzed to determine the nature of the adsorption states on Cu(100). Dimethyl sulfide is reversibly adsorbed on Cu(100); no dissociation (CS bond breaking) was found. By contrast, methyl mercaptan and dimethyl disulfide dissociate below 300 K to form adsorbed CH3S (methyl mercaptide) species. Depending on the coverage, two orientations of methyl mercaptide are found: linear and bent. The two different orientations can be distinguished via the surface dipole selection rule by different intensities of the methyl rocking and deformation vibrations. By contrast with the methoxy species, which on Cu(100) decomposes to formaldehyde, no H2C=S is liberated during decomposition of CH3S. The mercaptide is stable to ∼ 350 K, but decomposes at higher temperatures to form adsorbed sulfur and recombinant methane, hydrogen and ethane. The methane appears to be formed by methyl-hydrogen recombination when the C-S bond scission occurs. TDS results show that sulfur released from the decomposition poisons the surface toward further adsorption. In addition, the selectivity toward methane versus ethane can be altered by pre-titrating the adsorbed hydrogen with oxygen, thereby changing the relative methyl-hydrogen and methyl-methyl recombination probabilities.
Article
The hopping motion of methylthiolate, CH3S, on Cu(111) was effectively induced by injecting tunneling electrons from the tip of a scanning tunneling microscope (STM). When the tunneling electrons are injected into an isolated CH3S at liquid He temperature in which the motion is originally frozen, the hopping phenomenon of CH3S from a hollow site to a neighboring hollow site is observed. This phenomenon shows a clear threshold in the applied bias voltage corresponding to the vibrational excitation energy of the C-S stretching mode in CH3S. We suggest a model in which the hindered translational mode, directly related to the hopping motion, is excited in accordance with the excitation of the C-S stretching mode through anharmonic coupling.
Article
We report here single-molecule imaging and repositioning of 1,3-butadiene molecules adsorbed on a palladium (110) surface by scanning tunneling microscopy (STM) at cryogenic temperature of 4.7 K. The atomically resolved STM images reveal that the adsorbed 1,3-butadiene molecules have two types of adsorption sites, where the molecular axis is oriented along [1\bar{1}0] (N type) or [001] (R type). Based on the STM images, we determine the adsorption sites of each type. In addition, repositioning of R-type 1,3-butadiene molecules is demonstrated by controlling tip-adsorbate distance and by moving the tip laterally. The target molecules have been laterally moved and simultaneously rotated to change their adsorption sites from R type to N type on the next row.
Article
We propose an action spectroscopy for single-molecule motion induced by vibrational excitation with a scanning tunneling microscope (STM). Calculations of the inelastic tunneling current for excitation of the C-O stretch mode of the CO molecule on metal surfaces are combined with a theory which describes how the energy in the vibrational mode is transferred to a reaction coordinate mode to overcome the activation barrier. The calculated rate for CO hopping on Pd (110) as a function of the bias voltage agrees with the experimental result. It is proposed that the second derivative of the reaction rate with respect to the bias voltage is related to the vibrational density of states, which usually cannot be directly observed in STM inelastic electron tunneling spectroscopy when a molecule motion is induced by vibrational excitation.