Nanoscale Coherent Phonon Spectros

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

SCIENCE ADVANCES | RESEARCH ARTICLE

SURFACE CHEMISTRY Copyright © 2022


The Authors, some
Nanoscale coherent phonon spectroscopy rights reserved;
exclusive licensee
American Association
Shuyi Liu1, Adnan Hammud2, Ikutaro Hamada3, Martin Wolf1, for the Advancement
Melanie Müller1*, Takashi Kumagai1,4* of Science. No claim to
original U.S. Government
Coherent phonon spectroscopy can provide microscopic insight into ultrafast lattice dynamics and its coupling to Works. Distributed
other degrees of freedom under nonequilibrium conditions. Ultrafast optical spectroscopy is a well-established under a Creative
method to study coherent phonons, but the diffraction limit has hampered observing their local dynamics directly. Commons Attribution
Here, we demonstrate nanoscale coherent phonon spectroscopy using ultrafast laser–induced scanning tunnel- License 4.0 (CC BY).
ing microscopy in a plasmonic junction. Coherent phonons are locally excited in ultrathin zinc oxide films by the
tightly confined plasmonic field and are probed via the photoinduced tunneling current through an electronic

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
resonance of the zinc oxide film. Concurrently performed tip-enhanced Raman spectroscopy allows us to identify
the involved phonon modes. In contrast to the Raman spectra, the phonon dynamics observed in coherent pho-
non spectroscopy exhibit strong nanoscale spatial variations that are correlated with the distribution of the elec-
tronic local density of states resolved by scanning tunneling spectroscopy.

INTRODUCTION vibrational motion of single molecules with angstrom spatial and


Tracking the time evolution of coherent lattice vibrations can pro- femtosecond temporal resolution (26, 29, 33). More recently, it was
vide valuable insight into the mutual coupling between electronic demonstrated that coherent acoustic phonons can be launched in a
and structural degrees of freedom in condensed matter. Collective thin metal film by local femtosecond Coulomb forces using a
oscillations of atoms can be induced by ultrafast excitation with a terahertz-gated STM (34). However, ultrafast photoexcitation and
laser pulse whose duration is shorter than the period of nuclear real-time observation of optically excited CPs on the nanoscale re-
vibration. Coherent phonons (CPs) can be observed in various ways, main to be demonstrated. Here, we show nanoscale coherent
for example, via their coupling to the electronic structure of the sam- phonon spectroscopy (CPS) on ultrathin zinc oxide (ZnO) films by
ple as used in optical spectroscopy and time-resolved photoelectron coupling ultrashort NIR laser pulses into a plasmonic STM junction.
spectroscopy (1) or directly via the lattice response using time-­ The strong confinement of the plasmonic field enables us to locally
resolved diffraction techniques (2, 3). CPs have been investigated in excite and probe the ultrafast lattice dynamics of few-monolayer
a wide range of materials including metals (4), semimetals (5, 6), (ML) ZnO films with ~2-nm and ~10-fs spatiotemporal resolution.
semiconductors (7–9), oxides (10–12), strongly correlated materials Furthermore, the combination of nanoscale CPS with scanning
(13, 14), carbon materials (15, 16), molecular crystals (17), and for tunneling spectroscopy (STS) and tip-enhanced Raman spectroscopy
atoms/molecules adsorbed on surfaces (18). On femtosecond time (TERS) allows us to correlate the local equilibrium electronic struc-
scales, the electronic and phononic systems after impulsive excitation ture and phonon modes concurrently with the CPs excited under
are not in equilibrium, and the dynamics of CPs evolve through nonequilibrium conditions.
electron-phonon or phonon-phonon scattering. It has been shown
recently that CPs can also be used to control the rearrangement of
atoms during a photoinduced phase transition (3). The dynamics of RESULTS
CPs will be sensitive to nanoscale perturbations of the local lattice Our approach to probing local ultrafast dynamics with photoinduced
potential, e.g., originating from point defects in a crystal (19). How- STM is based on femtosecond laser excitation of a plasmonic tun-
ever, the real-space observation of CPs approaching atomic length neling junction as illustrated in Fig. 1A. The STM junction consists
scales remains a challenging task. of the clean Ag(111) surface and a nanoscopically sharp Ag tip fab-
To attain ultrafast spectroscopy at the atomic scale, continuous ricated by focused ion beam milling (35). The junction is illuminated
efforts have been devoted to combine scanning tunneling microscopy by two collinear NIR pulses of 10-fs duration at a center wavelength
(STM) with ultrashort laser excitation in the past decades (20–26). of 780 nm and a repetition rate of 80 MHz. The pulse pair is created
More recently, coupling phase-stable, single-cycle terahertz or near-­ by separating a laser beam using a balanced-dispersion Michelson
infrared (NIR) pulses to the STM junction has allowed to trace ultra- interferometer with variable delay in one arm. Both pulses are linearly
fast charge carrier dynamics (27–29), plasmon dynamics (30), and polarized along the tip axis. The resulting highly confined plas-
electronic coherences (31, 32) with subcycle temporal resolution. In monic field enables optical excitation in the STM junction at a scale
addition, it was demonstrated that ultrafast STM can probe coherent of a few nanometers (36) and generates an optically induced tunneling
current at low incident laser power densities (Pinc < 0.2 mW/m2,
1
which corresponds to the laser power of <2 mW incident into the
Department of Physical Chemistry, Fritz Haber Institute of the Max Planck Society,
Faradayweg 4-6, 14195 Berlin, Germany. 2Department of Inorganic Chemistry, Fritz
STM junction). This low incident power is highly beneficial for
Haber Institute of the Max Planck Society, Faradayweg 4-6, 14195 Berlin, Germany. keeping the junction stable and minimizing photoinduced thermal
3
Department of Precision Engineering, Graduate School of Engineering, Osaka Uni- effects (23). The detected STM current (ISTM) under illumination is
versity, 2-1 Yamada-Oka, Suita, Osaka 565-0871, Japan. 4Center for Mesoscopic the sum of the static tunneling current (IDC) induced by the applied
Sciences, Institute for Molecular Science, Okazaki 444-8585, Japan.
*Corresponding author. Email: m.mueller@fhi-berlin.mpg.de (M.M.); kuma@ims. DC bias (Vbias) and the photoinduced current (Iphoto). The Iphoto
ac.jp (T.K.) originates from transfer of photoexcited electrons across the junction

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 1 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

tunneling regime, and the plasmonically enhanced interferometric


Iphoto can be detected at a sufficient signal-to-noise ratio without
lock-in detection methods. The photothermal modulation (noise)
of the junction during the IAC measurement is estimated to be less
than 0.05 nm at Pinc = 0.12 mW/m2 (see fig. S1), which does not
significantly affect the Iphoto because of the shallow ISTM–z slope in
the used distance range. Thus, the contribution from thermal ex-
pansion is almost negligible in the IAC measurement when it is re-
corded in the photoinduced tunneling regime, in which the Iphoto is
dominant. The IAC trace measured for the Ag tip–Ag(111) junc-
tion can be fitted by
+∞
​​I​ IAC​​(t ) ∝ ​∫−∞ ​​​ ​∣(​E​ pl​​(t ) + ​E​ pl​​(t − t ) )∣​​ 2​n​ eff​​​  dt ​

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
(1)

where Epl is the plasmonic field and neff is the effective nonlinearity
of the photoinduced process. The noninstantaneous plasmonic re-
sponse of the Ag tip–Ag(111) junction contributes to the IAC trace,
leading to broadening of the central envelope and the appearance of
satellite tails extending to tens of femtoseconds (inset to Fig. 1C)
(39). We approximate the plasmon dynamics by a damped harmon-
Fig. 1. Ultrashort pulsed laser–induced STM in a plasmonic junction. (A) Sche- ic oscillator driven by the laser pulse field E(t), resulting in the
matic of the experiment. (B) ISTM–z curves recorded with (red) and without (black) plasmon-enhanced field
illumination. z = 0 corresponds to the STM set point of Vbias = 2 V and ISTM = 6 nA at

​​   ​ ​​ E(​t′ ​ ) exp​(​​− ​ ─ ​​  pl​​ )


t
Pinc = 0.16 mW/m2. The red circles show the neff extracted from the peak-to-baseline ​​​E​ pl​​(t ) ∝ ​∫−∞​​​ ​ ─ 1 t − ​t ​​′ ​ ​​sin(​​  ​​(t − ​t′ ​ ) ) d​t′ ​ ​​ (2)
pl
ratio r of the IAC trace using the formula neff = [log2(r) + 1]/2. (C) IAC trace of the pl
photocurrent from the tip recorded for the Ag tip–Ag(111) junction at Vbias = 2 V

where ​​E(​t′ ​ ) = sech​(​​​t​_′ ​)​​ ​​sin(​t′ ​)​​is the electric field of the individual


with fixed tip height at z = 1 nm. The dashed line indicates the fitting result using
Eqs. 1 and 2. The inset is the magnified IAC trace after a delay time of 40 fs.
laser pulses with a hyperbolic secant shape of duration  (full width
at half maximum) and pl and pl are the decoherence time and the
either above the vacuum barrier or via photoassisted tunneling (37). resonance frequency of the plasmon, respectively. The fitting anal-
With the excitation wavelength at 780 nm, one-, two- and three-photon– ysis can reproduce the IAC trace reasonably well (black dashed line
excited electrons gain an excess energy of ~1.6, ~3.2, and ~ 4.8 eV, in Fig. 1C) and yields pl~25 fs, pl~ 421.6 THz (corresponding to a
respectively. Because the work function of the Ag surface is 4.2 to photon wavelength of 711 nm). This dephasing time is consistent
4.7 eV (38), the one- and two-photon processes result in the Iphoto with the reported dephasing time of Ag nanoparticle plasmons
through photoassisted tunneling, whereas the three-photon process (40, 41). To clarify the contribution from the localized surface plas-
can lead to the Iphoto through over-barrier emission. The proportion mon (LSP), we record an IAC trace for nonlinear photoemission
of electrons that are emitted above or tunnel through the barrier de- from a Cu(111) surface that exhibits a much weaker plasmonic re-
pends on the barrier width (i.e., tip–surface distance). Electron tun- sponse. The result reveals a negligible plasmon response, and the
neling becomes dominant at a small tip–surface distance where the IAC trace can be used to evaluate the pulse duration at the junction
tunneling probability is large. Under illumination, the Iphoto consists (see fig. S2). We further evaluate the effective nonlinearity neff of the
of electrons transferred from both tip and surface, but the electron Iphoto from the peak-to-baseline ratio r of the IAC trace, revealing
transfer from the tip (surface) can be suppressed by applying a neg- that the neff decreases from 2.8 to 1.6 as z decreases (Fig. 1B). This
ative (positive) bias to the surface. The exact magnitude of the nega- observation evidences the increasing contribution of lower-order
tive (positive) bias required to suppress the Iphoto contribution from processes due to photoassisted tunneling at short tip–surface distances.
the tip (surface) depends on the incident laser power, the tip–surface The femtosecond time resolution achievable with ultrafast photo­
distance, and the tip conditions (plasmon enhancement). induced STM allows us to trace ultrafast dynamics at surfaces in real
To characterize the plasmonic response of the junction, we mea- time and in real space. We apply the above measurement scheme to
sure Iphoto from the tip to the surface at a positive sample bias. The locally excite and probe ultrafast lattice dynamics in ultrathin ZnO
relative contribution from IDC and Iphoto can be assessed by record- films epitaxially grown on the Ag(111) surface (42). A coherent lattice
ing ISTM as a function of the relative tip–surface distance (z) with oscillation was previously proposed in femtosecond laser-coupled
and without illumination (Fig. 1B). Here, z = 0 corresponds to the STM (24), but it has remained uncorroborated. Figure 2A illustrates
tip–surface distance given by the STM set-point of Vbias = 2 V and the experiment, which is the same setup as depicted in Fig. 1 but
ISTM = 5 nA. At z < 0.1 nm, the total current (ISTM) is dominated with a negative Vbias applied to the sample to suppress the transfer
by the IDC. The Iphoto mainly contributes to the ISTM at z > 0.3 nm of photoexcited electrons from the tip. Figure 2B compares the
and extends more than 5 nm outside the tunneling regime (red ISTM–z curves under femtosecond laser illumination recorded over
curve in Fig. 1B). Figure 1C shows the Iphoto recorded for the Ag the bare Ag(111) surface and over the regions of 2- and 3-ML ZnO
tip–Ag(111) junction as a function of the time delay (t) between (hereafter denoted by 2-ML and 3-ML ZnO). Note that the current
the two NIR pulses, yielding the interferometric autocorrelation is given by the absolute value and the electrons flow from the surface
(IAC) of the Iphoto. The data are recorded at z = 1 nm outside the to the tip. The simulated structure of 2- and 3-ML ZnO is reported

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 2 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
Fig. 2. Photoinduced resonant electron tunneling from the ultrathin ZnO films. (A) Schematic of the experiment. (B) ISTM–z curves recorded over the bare Ag(111)
surface, 2-ML, and 3-ML ZnO, respectively. The measurement locations are indicated by the stars in the inset STM image (scale bar, 5 nm). z = 0 corresponds to the
tip–surface distance given by the STM set point of Vbias = −2 V and ISTM = 3 nA at Pinc = 0.08 mW/µm2. Note that the current is given by the absolute value and the electrons
flow from the surface to the tip. (C) Schematic energy diagram. s/t, work function of surface/tip; Evac, vacuum level. The vertical and horizontal arrows indicate electronic
excitation and electron transfer, respectively.

in (36) and in fig. S3, respectively. We found that the Iphoto is larger modulation of the Iphoto. The coherent lattice displacement is esti-
for 3-ML ZnO by one order of magnitude compared to the other mated to be 10−4 to 10−3 nm (44, 45), which is much smaller than
two cases. The large Iphoto from 3-ML ZnO can be explained by res- the tip–surface distance (~1 nm) used in the IAC measurements. As
onant photoassisted tunneling, which is absent for 2-ML ZnO and discussed above, at the distance of the IAC measurement, the slope
the Ag(111) surface. Figure 2C depicts the schematic energy dia- of the Iphoto-z curve is rather small; hence, a CP-induced displace-
gram of the Ag tip–vacuum–ZnO–Ag(111) junction. The position ment of the atoms cannot result in a measurable change of ISTM. The
of the conduction band edge (CBE) of the ultrathin ZnO films de- negligible contribution is also corroborated by the fact that the am-
pends on the film thickness, and its onset can be observed in STS at plitude of the in-plane modes is larger than the out-of-plane modes
~1.8 and ~1.4 eV above the Fermi level (EF) for 2- and 3-ML ZnO, in some cases. Instead, we deduce that the CPs modulate the elec-
respectively (42). Furthermore, the interface state (IS) between the tronic structure of the ZnO film, eventually modifying the magni-
ZnO film and the Ag(111) surface is located at ~0.2 eV below EF, tude of the Iphoto through a change in the transition probability from
originating from the Shockley surface state of Ag(111) (42). In con- IS to CB. The local work function may also be modulated by the CPs,
trast to 2-ML ZnO, the transition from the IS to the CBE is resonant which could affect the apparent barrier height and, thus, the tunnel-
with the excitation wavelength (780 nm) for 3-ML ZnO, resulting in ing probability. However, because, at a positive bias, the Iphoto caused
a much larger Iphoto. We also confirm that the local work function of by electron transfer from the tip is not modulated by the CPs, the
the surface (s) is not primarily related to the transfer of photoex- local work function change plays a minor role. The involved phonon
cited electrons in the junction (see fig. S4). modes are further analyzed by fitting the IAC trace with a damped
The resonant electronic excitation of 3-ML ZnO yields the en- harmonic oscillator model. Assuming that the phonon-induced
hanced Iphoto and allows us to observe the CPs excited impulsively modulation of the Iphoto is, in first-order approximation, linearly
via the resonant optical excitation. Figure 3A shows representative proportional to the phonon amplitude, we can fit the IAC traces in
IAC of the Iphoto recorded at three different locations on 3-ML ZnO Fig. 3A to a model of multiple damped oscillators
with z = 0.5 nm in the regime where IDC is negligible. In addition
to the coherent electronic center part of the IAC including the re- ​​I​ IAC​​ = ​∑ i​ ​​ ​A​ i​​ sin(2 ​​  i​​ t + ​φ​  i​​ ) ​e​​  −t/​​  i​​​ + ​I​ 0​​​ (3)
sponse of the LSP (<25 fs), all traces exhibit pronounced oscillations
extending up to ~1.5 ps, and their detailed profiles differ at the three where the ith damped oscillator is characterized by the initial ampli-
locations. In the Fourier transform (FT) of the IAC traces (Fig. 3B), tude Ai, the frequency i, the initial phase φi, and the damping
several peaks appear in the range of 7.5 to 21.0 THz. The peak posi- time i. We deduce the phonon frequencies that are involved in the
tions in the IAC-FT spectra are consistent with the phonon modes IAC trace based on the analysis of the IAC-FT and TERS spectra in
observed in TERS measured at the same position (Fig. 3C). TERS Fig. 3 (B and C). The fitting results (the black dashed curves in
provides equilibrium phonon spectra in the frequency domain with Fig. 3A) reproduce fairly well the IAC trace using five phonon
ultrahigh spatial resolution reaching ~1 nm (36). In Fig. 3C, the modes at frequencies of 300, 340, 400, 433, and 530 cm−1. Note that
bands at 250 to 460 cm−1 and 470 to 700 cm−1 can be assigned to the there is one measurably different peak at 340 cm−1 in the FT spectra
out-of-plane and in-plane phonon modes of the ZnO film, respec- instead of 360 cm−1 in TERS [these peaks are highlighted by vertical
tively (36, 43). Therefore, we conclude that the long-lasting oscilla- black lines in Fig. 3 (B and C)]. The fitting procedure and parame-
tions observed in the IAC traces result from CPs of 3-ML ZnO. In ters for all phonon modes are described in the Supplementary
the out-of-plane and in-plane modes, atoms in the ZnO film are Materials, showing that the values of Ai, φi, and i largely depend on
mainly displaced perpendicular and parallel to the surface, respec- the measurement site.
tively. The out-of-plane modes will slightly change the tip–surface The observed spatial variation of the IAC trace should be associ-
distance, which, in principle, can modify the tunneling probability ated with the underlying mechanisms of the CP excitation and re-
(i.e., Iphoto). However, this should not contribute to the observed laxation. To rationalize the CP excitation, two key processes have

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 3 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

spectra, indicates that resonant Raman scattering cannot be a dom-


inant mechanism responsible for CP generation in the ZnO film. As
the excitation wavelength of 780 nm is resonant with the electronic
transition between the IS and the CBE of 3-ML ZnO, we believe that
the DECP mechanism is a major driving force. DECP and ISRS are
often distinguished by the initial phonon phase, which are zero for
the ISRS mechanism under nonresonant conditions. The phases
(φi) in Eq. 3 are all nonzero (see table S1). The original DECP theory
predicts ​​φ​  i​​ = ​_2​​when the lifetime of a coupled electronically excited
state is much longer than the period of the coherent lattice oscilla-
tion (49). However, all φi obtained from the fitting analysis in Fig. 3A
largely deviate from _​​2 ​​(table S1). This could be explained by a short

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
lifetime of the resonantly excited charge carriers compared to that of
the coherent lattice oscillation (49). The absence of an (incoherent)
electronic decay in the IAC traces implies that the lifetime of the
excited electrons is shorter than the plasmon dephasing (i.e. <25 fs).
In this case, the resulting displacive force becomes more impulsive,
thus yielding ​​φ​  i​​ ≠ ​_2​​ (49, 50). The short electron lifetime could be
explained by fast relaxation of the excited state(s) into the Ag sub-
strate and/or by the nature of the highly localized excitation inside
the plasmonic field, for which lateral transport of excited electrons
out of the excitation volume could effectively reduce the lifetime of
the locally excited charge density. We note that the CPs are excited
by the local electromagnetic field enhanced by the LSP in the STM
junction, which differs significantly from far-field excitation. The
strong field confinement (to x) leads to momentum uncertainty,
which is given by k ~ /x in the optical excitation, which could
relax the momentum conservation required in far-field optical exci-
tation (51, 52). The momentum uncertainty could contribute to the
electronic transition and result in excitation of phonons away from
the  point.
In the DECP model, the amplitude of CPs is assumed to be lin-
early proportional to the excited charge density (46, 49). We thus
examine how the localized plasmonic field will affect CP excitation.
According to previous TERS measurements (53), the LSP is expo-
nentially enhanced as the tip–surface distance is reduced. Figure 3D
Fig. 3. CPs measured in ultrathin 3-ML ZnO. (A) IAC traces measured over differ-
shows the z-dependent IAC traces obtained on 3-ML ZnO. The
ent locations over 3-ML ZnO (set point: Vbias = −4 V, ISTM = 2 nA, z = 0.5 nm,
Pinc = 0.08 mW/m2). The dashed lines are fits obtained using Eq. 3 with the parameters
shape of the IAC traces is independent of z; thus, the CP dynamics
provided in table S1. (B) FT spectra obtained from the IAC traces in (A). (C) TERS are not noticeably modified by the change of the LSP enhancement.
spectra recorded with a narrowband continuous wave laser at 780 nm (Pinc = The amplitude of the CPs can be evaluated by analyzing the tran-
0.8 mW/m2) at a set point of Vbias = 50 mV and ISTM = 3 nA. The vertical solid lines sient photocurrent change, i.e., the ratio between the coherent oscil-
in (B) and (C) indicate the frequencies used for the fitting analysis in (A). (D) IAC lation amplitude and the IAC baseline current. Provided IDC is
traces recorded over 3-ML ZnO at different z (z = 0 corresponds to the tip–surface negligible, and the relative oscillation amplitude of the IAC induced
distance given by the STM set point of Vbias = −1 V, ISTM = 3 nA, Pinc = 0.08 mW/m2). by the CPs is approximately proportional to the nuclear displace-
(E) Semi-log plot of the amplitude-to-baseline ratio of the traces in (D) as a function ment (Q); thus
of z. Note that the current is given by the absolute value and the electrons flow
from the surface to the tip. ​I​ dis​​  − ​I​ eq​​ 2(​I​ IAC​​  − ​I​ 0​​)
Q ∝ ​─​ = ​─
​ ​​ (4)
​Ie​  q​​ ​I​ 0​​
been discussed, namely displacive excitation of CPs (DECP) (46)
and impulsive stimulated Raman scattering (ISRS) (47, 48). The where IIAC = Idis + Ieq, Idis is the photocurrent from the probe pulse
​I​  ​​
DECP mechanism involves real electronic transitions that create an in the presence of the CPs, and ​​I​ eq​​ = ​_20​​is the photocurrent from a
excited state population. The corresponding rearrangement of charges single pulse at equilibrium lattice positions. As plotted in Fig. 3E, we
leads to a displaced potential energy surface, which can eventually found that Q increases as z is decreased for z > 1 nm, which is
drive coherent lattice motion. In contrast, ISRS drives Raman active explained by the increased LSP field strength with decreasing the
modes via a virtual excitation, which is typically the operating mecha- tip–surface distance. However, Q saturates for z < 1 nm, implying
nism in transparent media. In the present case, the pronounced dis- that the phonon amplitude is saturated when the local field becomes
crepancy in the spectral distributions between the frequency- and very strong. At extremely small cavity distances, the LSP will be
time-domain measurements, i.e., the frequency at 340 cm−1 observed damped because of quantum mechanical effects (54, 55). As previ-
in the IAC-FT spectra in Fig. 3B being different from the TERS ously reported (53), damping is expected to occur at Ag tip–Ag(111)

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 4 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
Fig. 4. Nanoscale CPS. (A) STS map (Vbias = 1.5 V, ISTM = 0.3 nA), representing the LDOS of the CBE of 3-ML ZnO. (B) Iphoto map recorded over the same 3-ML ZnO island in
constant height mode (z = 0 corresponds to the tip–surface distance given by the STM set point of Vbias = −3 V, ISTM = 4 nA, z = 0.5 nm, Pinc = 0.16 mW/m2). (C) Blurred
STS map of (A) using a mean filter with the diameter of 4 nm. (D) The region of interest is indicated by the white dashed box in (B). Scale bars, 5 nm (A to D). (E) IAC traces
at different positions marked in the enlarged Iphoto image in (D). Note that the current is given by the absolute value, and the electrons flow from the surface to the tip.
The dashed lines are the fitting curves, and the parameters are provided in table S2. (F) Corresponding FT spectra of the IAC traces in (E). The FT magnitude is normalized
with respect to the IAC baseline. (G) Lateral distance dependence of the filtered STS intensity, Iphoto, and the FT peak magnitude along the line scan in (E). The magenta
and cyan markers in the bottom panel are the FT magnitudes integrated over the range of 254 to 371 cm−1 and 391 to 469 cm−1, respectively, as shaded in (F).

surface distances smaller than ~0.5 nm, at which the tip is nearly in and further corroborating that the Iphoto is enhanced through the
atomic contact to the surface of 2-ML ZnO film. However, all the electronic resonance. The Iphoto is larger at positions of higher LDOS
IAC measurements are performed at a distance of ~1 nm; hence, the because of a larger transition matrix for the photoexcitation. The
Q saturation cannot be explained by attenuation of the LSP. We Iphoto and blurred STS maps exhibit “domains” in their contrast on
speculate that the saturating behavior may be related to limited car- a length scale of 2 to 8 nm. Figure 4 (D to F) display the IAC traces
rier generation due to the strong spatiotemporal confinement of the and corresponding FT spectra recorded across the different do-
exciting plasmon field. Above a certain strength of the extremely mains with a 2-nm interval. Whereas the IAC traces do not change
localized and enhanced field, the locally excited charge density significantly inside each domain, an abrupt change of the beating
might be saturated because of Coulomb repulsion. pattern occurs within a 2-nm length scale between neighboring do-
The generation of CPs via optical excitation is fundamentally as- mains. In the FT spectra, this is manifested as a significant change
sociated with the electronic structure of the sample. The versatility in the relative peak intensities of the involved phonon modes
of ultrafast STM enables the concurrent observation of local femto- (Fig. 4F; see also table S2, which lists the fitting parameters of the
second dynamics and equilibrium electronic states. Figure 4A shows IAC traces in Fig. 4D). As can be seen in Fig. 4G, the change in the
an STS intensity map recorded at Vbias = 1.5 V without laser illumi- relative peak intensities (and in the fitting parameters in table S2)
nation, representing the distribution of the local density of states correlates with the spatial variation of the STS intensity and Iphoto.
(LDOS) of the CB of 3-ML ZnO. The contrast modulation inside The dephasing of CPs can involve complex contributions, such
the 3-ML ZnO arises from inhomogeneities originating from atom- as anharmonic coupling between phonon modes, which is domi-
istic defects in the film (56). Figure 4B displays the corresponding nant in bulk ZnO (12). However, the phonon dephasing time in
Iphoto map recorded at z = 0.5 nm, exhibiting pronounced na- bulk ZnO is estimated to be several picoseconds, thus much longer
noscale variations. However, the spatial resolution of the Iphoto map than the i obtained in the present case. Therefore, anharmonic
is lower than that of the STS map and will be determined by the coupling within 3-ML ZnO should not be a major damping chan-
lateral size of the plasmonic field in the junction (37). This assump- nel. However, phonon scattering at atomistic defects is reported as
tion can be verified in the artificially “blurred” STS map using a a damping channel of CPs (19), which potentially contributes to the
mean filter with the diameter of 4 nm that accounts for the field dephasing process of the CPs. For adsorbates on metal surfaces,
localization (the size of the plasmonic field). The blurred STS map multiple channels may contribute simultaneously to the vibrational
(Fig. 4C) is indeed very similar to the Iphoto map (Fig. 4B), indicating damping, namely, interadsorbate interactions (57), electron–hole pair
that the plasmonic field is confined to ~4 nm underneath the tip (EHP) excitation in a metal (58), anharmonic coupling to surface

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 5 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

phonons (59), and hot-electron scattering (60). Damping through small but unavoidable thermal effects. To overcome these issues,
hot-electron scattering can be excluded by the z-dependence of lock-in detection of the laser-induced tunneling current without
the IAC traces in Fig. 3E. At smaller tip–surface distances, hot elec- amplitude modulation (31) or combination with terahertz-gated
trons will be excited more efficiently because of the increased LSP STM might be promising approaches. We envision that real-time
field in the junction, but the observed CP dynamics are not affected. and real-space observation of ultrafast coherent lattice dynamics
The contribution from EHP excitation and anharmonic coupling to using ultrafast photoinduced STM may pave the way for studying
surface phonons depends on the interaction between the adsorbate the coupling of fundamental degrees of freedom in solids—lattice,
and the surface (61). Because the ZnO film is physisorbed onto the charge, orbital, and spin—on the atomic scale.
Ag(111) surface, these channels may not be dominant. The locally
excited CPs in the ZnO film may efficiently couple to the surround-
ing lattice because they have the same phonon modes (the phonon MATERIALS AND METHODS
frequency is not influenced by the inhomogeneity as evidenced by All experiments were performed inside an ultrahigh vacuum (UHV)
(base pressure of <5 × 10−10 mbar) chamber equipped with a low-­

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
the TERS spectra in Fig. 3C). However, to gain further insight into
the complex phonon dynamics in the ZnO film is challenging in the temperature STM system from UNISOKU Ltd. (modified USM-
present work. Polarization-dependent measurements, which are 1400) that is operated with Nanonis SPM Controller from SPECS
commonly used to distinguish CP modes with different symmetry, GmbH. The bias voltage (Vbias) was applied to the sample, and the
are not straightforward to implement in the present case because of tip was grounded. The ISTM was collected from the tip through a
the nontrivial polarization of the tightly confined plasmonic near current amplifier from FEMTO Messtechnik GmbH (DLPCA-200).
field. We expect that more insight into the underlying mechanism The Ag tips are prepared in a two-step process. First, an Ag wire was
can be gained by variation of the resonant excitation condition, e.g., chemically etched to obtain a sharp tip, which was then further
by using a tunable wavelength femtosecond laser source. sharpened by focused ion beam milling to yield a nanoscopically
sharp plasmonic tip as detailed in (35).
The STM junction was illuminated by a titanium-sapphire laser
DISCUSSION oscillator from Laser Quantum (Venteon Pulse One), which pro-
We demonstrated the capability of ultrasfast NIR laser-induced vides 10-fs laser pulses at 780-nm center wavelength and 80-MHz
STM to resolve the spatial inhomogeneity of ultrafast lattice dynamics repetition rate. A spectral phase interferometry for direct electric-­
in ultrathin ZnO films with ~2-nm and ~10-fs spatiotemporal reso- field reconstruction (SPIDER, APE GmbH) was used to measure
lution. We exploited the strong confinement and enhancement of the duration of the laser pulse at a reference beam path with identi-
the LSP in the STM junction to obtain spatially localized ultrafast cal dispersion outside UHV. To couple with the low-temperature
photocurrents at very low excitation power. The LSP field allowed STM, the laser beam enters the UHV chamber through a fused silica
us to operate the photoinduced STM outside the normal tunneling window. A chirped mirror pair is used for dispersion control and to
regime while maintaining a high spatial resolution that is deter- precompensate the dispersion of the setup. The laser polarization
mined by the localization of the photoassisted tunneling current. was aligned along the tip axis (p-polarized). The beam is focused
This operation mode largely mitigated the thermal effects that usually inside the UHV by a custom-designed parabolic mirror with a focal
disturb the measurement and interpretation of two-pulse autocor- length of 8 mm to ~3-m spot size in the junction. The parabolic
relation experiments performed in the femtosecond laser-coupled mirror was precisely aligned using piezo motors (Attocube GmbH)
STM. Furthermore, resonant optical excitation of photoassisted enabling three translational and two rotational motions. Collinear
tunneling channels via the transition from the IS to the CB within pairs of identical pulses are generated inside a dispersion-balanced
the ZnO film yields sufficient photoinduced current to detect the Michelson interferometer. The laser power is equal in both inter-
ultrafast coherent lattice vibrations. The resonant excitation condi- ferometer arms, and before measurements, it was checked that the
tion, together with the observed discrepancies between the TERS photocurrent from both arms is additive at large delays far away
and time-domain spectra of CPs, indicated that the DECP mecha- from temporal overlap of the two pulses. In the IAC measure-
nism is involved in the CP generation in 3-ML ZnO films. The con- ments, the delay is controlled by a closed-loop piezo stage (PI-Hera)
current real-space observation of the local electronic structure and operated at 0.5- to 2-Hz scanning frequency, and the ISTM is time-­
equilibrium phonon modes by STS and TERS is a unique capability of averaged over several scanning cycles. The maximum scanning speed
STM and enables to investigate the complex correlation between ultra- is thereby limited by the bandwidth of the preamplifier (1 kHz here)
fast dynamics of electron–phonon and phonon–phonon coupling. and the fastest frequency components that need to be resolved in
The presented experimental concept for nanoscale CPS can be the IAC trace.
applied to investigate the excited state dynamics and the interplay TERS was measured using a narrow-band solid-state laser at
between electronic and phononic degrees of freedom at photoexcited 780 nm. The beam was focused by the same parabolic mirror as
surfaces on length scales approaching a few unit cells. To extend the mentioned above. The Raman signal was collected in the back-­
approach to other sample systems and to control the resonant con- scattering geometry and was detected outside of the UHV chamber
dition in the excitation process, it will be beneficial to combine with a grating spectrometer (AndorShamrock 303i) equipped with
wavelength-tunable pulsed lasers with a broadband LSP resonance the back-illuminated charge-coupled device camera (Newton 970).
in the STM junction. We expect that higher spatial resolution The scattered light was separated by a beamsplitter (90:10) and fil-
approaching the atomic scale can be achieved by using plasmonic tered by a longpass filter before coupling to the spectrometer via an
picocavities (62), confining the plasmonic near-field down to a few optical fiber.
atoms (63, 64). Operating at closer tip–surface distances to increase The Ag(111) surface was first cleaned by repeated cycles of Ar+
spatial resolution will remain a significant challenge due to the remaining sputtering and annealing up to ~700 K. Ultrathin ZnO layers were

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 6 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

grown on the clean Ag(111) surface by reactive deposition method 24. S. W. Wu, W. Ho, Two-photon-induced hot-electron transfer to a single molecule
as described in (65). in a scanning tunneling microscope. Phys. Rev. B 82, 085444 (2010).
25. Y. Terada, S. Yoshida, O. Takeuchi, H. Shigekawa, Real-space imaging of transient carrier
dynamics by nanoscale pump–Probe microscopy. Nat. Photonics 4, 869–874 (2010).
26. S. Li, S. Chen, J. Li, R. Wu, W. Ho, Joint space-time coherent vibration driven
SUPPLEMENTARY MATERIALS
conformational transitions in a single molecule. Phys. Rev. Lett. 119, 176002 (2017).
Supplementary material for this article is available at https://science.org/doi/10.1126/
27. T. L. Cocker, V. Jelic, M. Gupta, S. J. Molesky, J. A. J. Burgess, G. D. L. Reyes, L. V. Titova,
sciadv.abq5682
Y. Y. Tsui, M. R. Freeman, F. A. Hegmann, An ultrafast terahertz scanning tunnelling
microscope. Nat. Photonics 7, 620–625 (2013).
28. S. Yoshida, Y. Arashida, H. Hirori, T. Tachizaki, A. Taninaka, H. Ueno, O. Takeuchi,
REFERENCES AND NOTES H. Shigekawa, Terahertz scanning tunneling microscopy for visualizing ultrafast electron
1. U. Bovensiepen, P. S. Kirchmann, Elementary relaxation processes investigated by motion in nanoscale potential variations. ASC Photonics 8, 315–323 (2021).
femtosecond photoelectron spectroscopy of two-dimensional materials. Laser Photonics 29. T. L. Cocker, D. Peller, P. Yu, J. Repp, R. Huber, Tracking the ultrafast motion of a single
Rev. 6, 589–606 (2012). molecule by femtosecond orbital imaging. Nature 539, 263–267 (2016).
2. A. M. Lindenberg, I. Kang, S. L. Johnson, T. Missalla, P. A. Heimann, Z. Chang, J. Larsson, 30. M. Garg, K. Kern, Attosecond coherent manipulation of electrons in tunneling

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
P. H. Bucksbaum, H. C. Kapteyn, H. A. Padmore, R. W. Lee, J. S. Wark, R. W. Falcone, microscopy. Science 367, 411–415 (2020).
Time-resolved x-ray diffraction from coherent phonons during a laser-induced phase 31. M. Garg, A. Martin-Jimenez, Y. Luo, K. Kern, Ultrafast photon-induced tunneling
transition. Phys. Rev. Lett. 84, 111–114 (2000). microscopy. ACS Nano 15, 18071–18084 (2021).
3. J. G. Horstmann, H. Böckmann, B. Wit, F. Kurtz, G. Storeck, C. Ropers, Coherent control 32. M. Garg, A. Martin-Jimenez, M. Pisarra, Y. Luo, F. Martín, K. Kern, Real-space
of a surface structural phase transition. Nature 583, 232–236 (2020). subfemtosecond imaging of quantum electronic coherences in molecules. Nat. Photon.
4. M. Hase, K. Ishioka, J. Demsar, K. Ushida, M. Kitajima, Ultrafast dynamics of coherent 16, 196–202 (2022).
optical phonons and nonequilibrium electrons in transition metals. Phys. Rev. B 72, 33. L. Wang, Y. Xia, W. Ho, Atomic-scale quantum sensing based on the ultrafast coherence
189902 (2005). of an H2 molecule in an STM cavity. Science 376, 401–405 (2022).
5. M. Hase, K. Mizoguchi, H. Harima, S. Nakashima, M. Tani, K. Sakai, M. Hangyo, Optical 34. S. Sheng, A.-C. Oeter, M. Abdo, K. Lichtenberg, M. Hentschel, S. Loth, Launching coherent
control of coherent optical phonons in bismuth films. Appl. Phys. Lett. 69, 2474–2476 acoustic phonon wave packets with local femtosecond coulomb forces. Phys. Rev. Lett.
(1996). 129, 043001 (2022).
6. E. Papalazarou, J. Faure, J. Mauchain, M. Marsi, A. Taleb-Ibrahimi, I. Reshetnyak, 35. H. Böckmann, S. Liu, M. Müller, A. Hammud, M. Wolf, T. Kumagai, Near-field manipulation
A. van Roekeghem, I. Timrov, N. Vast, B. Arnaud, L. Perfetti, Coherent phonon coupling in a scanning tunneling microscope junction with plasmonic fabry-Pérot tips. Nano Lett.
to individual bloch states in photoexcited bismuth. Phys. Rev. Lett. 108, 256808 (2012). 19, 3597–3602 (2019).
7. T. Pfeifer, W. Kütt, H. Kurz, R. Scholz, Generation and detection of coherent optical 36. S. Liu, M. Müller, Y. Sun, I. Hamada, A. Hammud, M. Wolf, T. Kumagai, Resolving the correlation
phonons in germanium. Phys. Rev. Lett. 69, 3248–3251 (1992). between tip-enhanced resonance Raman scattering and local electronic states with 1 nm
8. M. Hase, M. Kitajima, A. M. Constantinescu, H. Petek, The birth of a quasiparticle in silicon resolution. Nano Lett. 19, 5725–5731 (2019).
observed in time–frequency space. Nature 426, 51–54 (2003). 37. B. Schröder, O. Bunjes, L. Wimmer, K. Kaiser, G. A. Traeger, T. Kotzott, C. Ropers,
9. K. J. Yee, K. G. Lee, E. Oh, D. S. Kim, Coherent optical phonon oscillations in bulk GaN M. Wenderoth, Controlling photocurrent channels in scanning tunneling microscopy.
excited by far below the band gap photons. Phys. Rev. Lett. 88, 105501 (2002). New J. Phys. 22, 033047 (2020).
10. I. H. Lee, K. J. Yee, K. G. Lee, E. Oh, D. S. Kim, Y. S. Lim, Coherent optical phonon mode 38. A. W. Dweydari, C. H. B. Mee, Work function measurements on (100) and (110) surfaces
oscillations in wurtzite ZnO excited by femtosecond pulses. J. Appl. Phys. 93, 4939–4941 of silver. Phys. Stat. Sol. 27, 223–230 (1975).
(2003). 39. B. Lamprecht, J. R. Krenn, A. Leitner, F. R. Aussenegg, Resonant and off-resonant
11. C. Aku-Leh, J. Zhao, R. Merlin, J. Menendez, M. Cardona, Long-lived optical phonons light-driven plasmons in metal nanoparticles studied by femtosecond-resolution
in ZnO studied with impulsive stimulated Raman scattering. Phys. Rev. B 71, 205211 third-harmonic generation. Phys. Rev. Lett. 83, 4421–4424 (1999).
(2005). 40. R. Mittal, R. Glenn, I. Saytashev, V. V. Lozovoy, M. Dantus, Femtosecond nanoplasmonic
12. K. Ishioka, H. Petek, V. E. Kaydashev, E. M. Kaidashev, O. V. Misochko, Coherent optical dephasing of individual silver nanoparticles and small clusters. J. Phys. Chem. Lett. 6,
phonons of ZnO under near resonant photoexcitation. J. Phys. Condens. Matter 22, 1638–1644 (2015).
465803 (2010). 41. M. Merschdorf, C. Kennerknecht, W. Pfeiffer, Collective and single-particle dynamics
13. L. Rettig, S. O. Mariager, A. Ferrer, S. Grübel, J. A. Johnson, J. Rittmann, T. Wolf, in time-resolved two-photon photoemission. Phys. Rev. B 70, 193401 (2004).
S. L. Johnson, G. Ingold, P. Beaud, U. Staub, Ultrafast structural dynamics of the Fe- 42. T. Kumagai, S. Liu, A. Shiotari, D. Baugh, S. Shaikhutdinov, M. Wolf, Local electronic
pnictide parent compound BaFe2As2. Phys. Rev. Lett. 114, 067402 (2015). structure, work function, and line defect dynamics of ultrathin epitaxial ZnO layers
14. Y. Zhang, X. Shi, W. You, Z. Tao, Y. Zhong, F. C. Kabeer, P. Maldonado, P. M. Oppeneer, on a Ag(111) surface. J. Phys. Condens. Matter 28, 494003 (2016).
M. Bauer, K. Rossnagel, H. Kapteyn, M. Murnane, Coherent modulation of the electron 43. C. Peng, G. Qin, L. Zhang, G. Zhang, C. Wang, Y. Yan, Y. Wang, M. Hu, Dependence
temperature and electron–phonon couplings in a 2D material. Proc. Natl. Acad. Sci. U.S.A. of phonon transport properties with stacking thickness in layered ZnO. J. Phys. D Appl.
117, 8788–8793 (2020). Phys. 51, 315303 (2018).
15. K. Ishioka, M. Hase, M. Kitajima, L. Wirtz, A. Rubio, H. Petek, Ultrafast electron-phonon 44. T. K. Cheng, J. Vidal, H. J. Zeiger, G. Dresselhaus, M. S. Dresselhaus, E. P. Ippen, Mechanism
decoupling in graphite. Phys. Rev. B 77, 121402 (2008). for displacive excitation of coherent phonons in Sb, Bi, Te, and Ti2O3. Appl. Phys. Lett. 59,
16. A. Gambetta, C. Manzoni, E. Menna, M. Meneghetti, G. Cerullo, G. Lanzani, S. Tretiak, 1923–1925 (1991).
A. Piryatinski, A. Saxena, R. L. Martin, A. R. Bishop, Real-time observation of nonlinear
45. A. V. Kuznetsov, C. J. Stanton, Theory of coherent phonon oscillations in semiconductors.
coherent phonon dynamics in single-walled carbon nanotubes. Nat. Phys. 2, 515–520
Phys. Rev. Lett. 73, 3243–3246 (1994).
(2006).
46. H. J. Zeiger, J. Vidal, T. K. Cheng, E. P. Ippen, G. Dresselhaus, M. S. Dresselhaus, Theory
17. H. Seiler, M. Krynski, D. Zahn, S. Hammer, Y. W. Windsor, T. Vasileiadis, J. Pflaum,
for displacive excitation of coherent phonons. Phys. Rev. B 45, 768–778 (1992).
R. Ernstorfer, M. Rossi, H. Schwoerer, Nuclear dynamics of singlet exciton fission
47. G. A. Garrett, T. F. Albrecht, J. F. Whitaker, R. Merlin, Coherent THz phonons driven by
in pentacene single crystals. Sci. Adv. 7, eabg0869 (2021).
18. K. Watanabe, N. Takagi, Y. Matsumoto, Impulsive excitation of a vibrational mode of Cs light pulses and the Sb problem: What is the mechanism? Phys. Rev. Lett. 77, 3661–3664
on Pt(111). Chem. Phys. Lett. 366, 606–610 (2002). (1996).
19. M. Hase, M. Kitajima, Interaction of coherent phonons with defects and elementary 48. T. E. Stevens, J. Kuhl, R. Merlin, Coherent phonon generation and the two stimulated
excitations. J. Phys. Condens. Matter 22, 073201 (2010). Raman tensors. Phys. Rev. B 65, 144304 (2002).
20. R. J. Hamers, Ultrafast time resolution in scanned probe microscopies. Appl. Phys. Lett. 57, 49. D. M. Riffe, A. J. Sabbah, Coherent excitation of the optic phonon in Si: Transiently
2031–2033 (1990). stimulated Raman scattering with a finite-lifetime electronic excitation. Phys. Rev. B 76,
21. G. Nunes Jr., M. R. Freeman, Picosecond resolution in scanning tunneling microscopy. 085207 (2007).
Science 262, 1029–1032 (1993). 50. E. M. Bothschafter, A. Paarmann, E. S. Zijlstra, N. Karpowicz, M. E. Garcia, R. Kienberger,
22. S. Weiss, D. F. Ogletree, D. Botkin, M. Salmeron, D. S. Chemla, Ultrafast scanning probe R. Ernstorfer, Ultrafast evolution of the excited-state potential energy surface of TiO2
microscopy. Appl. Phys. Lett. 63, 2567–2569 (1993). single crystals induced by carrier cooling. Phys. Rev. Lett. 110, 067402 (2013).
23. V. Gerstner, A. Knoll, W. Pfeiffer, A. Thon, G. Gerber, Femtosecond laser assisted scanning 51. J. Mertens, M.-E. Kleemann, R. Chikkaraddy, P. Narang, J. J. Baumberg, How light is
tunneling microscopy. J. Appl. Phys. 88, 4851–4859 (2000). emitted by plasmonic metals. Nano Lett. 17, 2568–2574 (2017).

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 7 of 8


SCIENCE ADVANCES | RESEARCH ARTICLE

52. S. Liu, A. Hammud, M. Wolf, T. Kumagai, Anti-stokes light scattering mediated by electron 65. A. Shiotari, B.-H. Liu, S. Jaekel, L. Grill, S. Shaikhutdinov, H.-J. Freund, M. Wolf,
transfer across a biased plasmonic nanojunction. ACS Photonics 8, 2610–2617 (2021). T. Kumagai, Local characterization of ultrathin ZnO layers on Ag(111) by scanning
53. S. Liu, B. Cirera, Y. Sun, I. Hamada, M. Müller, A. Hammud, M. Wolf, T. Kumagai, Dramatic tunneling microscopy and atomic force microscopy. J. Phys. Chem. C 118, 27428–27435
enhancement of tip-enhanced raman scattering mediated by atomic point contact (2014).
formation. Nano Lett. 20, 5879–5884 (2020). 66. P. E. Blöchl, Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994).
54. M. Urbieta, M. Barbry, Y. Zhang, P. Koval, D. Sánchez-Portal, N. Zabala, J. Aizpurua, 67. G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations
Atomic-scale lightning rod effect in plasmonic picocavities: A classical view to a quantum using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996).
effect. ACS Nano 12 (1), 585–595 (2018). 68. G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals
55. W. Zhu, R. Esteban, A. G. Borisov, J. J. Baumberg, P. Nordlander, H. J. Lezec, J. Aizpurua, and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15–50
K. B. Crozier, Quantum mechanical effects in plasmonic structures with subnanometre (1996).
gaps. Nat. Commun. 7, 11495 (2016). 69. J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple.
56. S. Liu, A. Shiotari, D. Baugh, M. Wolf, T. Kumagai, Enhanced resolution imaging Phys. Rev. Lett. 77, 3865–3868 (1996).
of ultrathin ZnO layers on Ag(111) by multiple hydrogen molecules in a scanning
tunneling microscope junction. Phys. Rev. B 97, 195417 (2018).
Acknowledgments: We thank D. Wegkamp for technical support, and Y. Sun, A. Paarmann,
57. M. Bonn, C. Hess, M. Wolf, The dynamics of vibrational excitations on surfaces: CO

Downloaded from https://www.science.org at University of Science and Technology of China on November 07, 2024
and S. Mährlein for valuable discussions. We thank R. Schlögl for supporting the development
on Ru(001). J. Chem. Phys. 115, 7725–7735 (2001).
and preparation of the FIB tips. Funding: T.K. acknowledges the support by JST FOREST
58. B. N. J. Persson, M. Persson, Vibrational lifetime for CO adsorbed on Cu(100). Solid State
Program (grant number JPMJFR201J, Japan) and the Grants-in-Aid for Scientific Research (JSPS
Commun. 36, 175–179 (1980).
KAKENHI grant number 19 K24684) from the Ministry of Education, Culture, Sports, Science,
59. B. N. J. Persson, R. Ryberg, Brownian motion and vibrational phase relaxation at surfaces:
and Technology of Japan. Author contributions: S.L., T.K., and M.M. conceived the
CO on Ni(111). Phys. Rev. B 32, 3586–3596 (1985).
experiment and performed the measurements. S.L. and T.K. performed the data analysis. A.H.
60. K. Watanabe, N. Takagi, Y. Matsumoto, Direct time-domain observation of ultrafast
prepared the Ag tip. S.L., T.K., and M.M. wrote the manuscript. All authors were involved in the
dephasing in adsorbate-substrate vibration under the influence of a hot electron bath: Cs
discussion and commented on the manuscript. Competing interests: The authors declare
adatoms on Pt(111). Phys. Rev. Lett. 92, 057401 (2004).
that they have no competing interests. Data and materials availability: All data needed to
61. I. Lončarić, M. Alducin, J. I. Juaristi, D. Novko, CO stretch vibration lives long on Au(111).
evaluate the conclusions in the paper are present in the paper and/or the Supplementary
J. Phys. Chem. Lett. 10, 1043–1047 (2019).
Materials.
62. J. J. Baumberg, Picocavities: A primer. Nano Lett. 22, 5859–5865 (2012).
63. J. Lee, K. T. Crampton, N. Tallarida, V. A. Apkarian, Visualizing vibrational normal modes
of a single molecule with atomically confined light. Nature 568, 78–82 (2019). Submitted 15 May 2022
64. Y. Zhang, B. Yang, A. Ghafoor, Y. Zhang, Y.-F. Zhang, R.-P. Wang, J.-L. Yang, Y. Luo, Accepted 2 September 2022
Z.-C. Dong, J. G. Hou, Visually constructing the chemical structure of a single molecule by Published 21 October 2022
scanning raman picoscopy. Natl. Sci. Rev. 6, 1169–1175 (2019). 10.1126/sciadv.abq5682

Liu et al., Sci. Adv. 8, eabq5682 (2022) 21 October 2022 8 of 8

You might also like