Tip-Enhanced Raman Spectromicroscopy of Co (II) - Tetraphenylporphyrin On Au (111) Toward The Chemists' Microscope

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article
Cite This: ACS Nano 2017, 11, 11466-11474 www.acsnano.org

Tip-Enhanced Raman Spectromicroscopy of


Co(II)-Tetraphenylporphyrin on Au(111):
Toward the Chemists’ Microscope
Joonhee Lee,† Nicholas Tallarida,† Xing Chen,‡ Pengchong Liu,‡ Lasse Jensen,*,‡
and Vartkess Ara Apkarian*,†

Department of Chemistry, University of California, Irvine, Irvine, California 92697, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Department of Chemistry, Pennsylvania State University, University Park, Pennsylvania 16802, United States
*
S Supporting Information

ABSTRACT: Atomically terminated and nanoscopically smooth silver tips effectively


Downloaded via 49.147.101.35 on June 28, 2018 at 19:54:24 (UTC).

focus light on the angstrom scale, allowing tip-enhanced Raman spectromicroscopy


(TER-sm) with single molecule sensitivity and submolecular spatial resolution. Through
measurements carried out on cobalt-tetraphenylporphyrin (CoTPP) adsorbed on
Au(111), we highlight peculiarities of vibrational spectromicroscopy with light confined
on the angstrom scale. Field-gradient-driven spectra, orientational fingerprinting, and
sculpting of local fields by atomic morphology of the junction are elucidated through
measurements that range from 2D arrays at room temperature to single molecule
manipulations at 5 K. Notably, vibrational Stark tuning within molecules, reflecting
intramolecular charge distributions, becomes accessible when light is more localized
than the interrogated normal modes. The Stark images of CoTPP reveal that it is
saddled, and the distortion is accompanied by charge transfer to gold through the two
anchoring pyrroles.
KEYWORDS: tip-enhanced Raman spectroscopy, spectromicroscopy, confined light, Stark shift, field-gradient-driven Raman,
quadrupolar scattering, scanning tunneling microscopy

by a gap of ∼1 nm, is prototypical.2 By coupling light to surface

T o the extent that molecules can be regarded as


networks of balls (atoms) and springs (bonds), they
can be characterized by the frequencies of their internal
motions. This is the realm of vibrational spectroscopy, which
remains one of chemists’ sharpest tools for fingerprinting
charge oscillations (plasmons) of the dipolar nanoantenna,
electromagnetic fields are effectively focused at the nano-
junction. The large field enhancement due to confinement is
sufficient to record Raman spectra of individual molecules,3−5
molecules and characterizing their inner workings. As such, the through the surface-enhanced Raman scattering (SERS) effect,
ability to record vibrational spectra of individual molecules, by which was discovered nearly four decades ago.6 Although there
sounding off individual atoms within molecules, can be has been an explosion of developments and applications of
regarded as the chemists’ ultimate microscopethe chemi- SERS,7 the fundamentals of the effect remain the subject of
scope for short. The task is challenging for several reasons. The active research. It is recognized that the primary challenge
universally applicable method for recording vibrational spectra stems from the atomistic morphology of junction shaping the
is based on the Raman effect, which involves the excitation of local fields,8 which determines the SERS signal.9 This is
molecules with light and recording the spectrum of reradiation precisely the sensitivity required to image the internal motions
by the vibrating molecule. The effect is feeble. To obtain a of individual molecules; at the same time, spectroscopy with
useful signal at the level of 103 photons/s, it is necessary to confined light is poorly understood, and characterization or
irradiate a molecule at intensities of terawatt/cm2. Moreover, control of nanojunctions with atomistic precision is a challenge.
photons are much larger than atoms. The Abbe principle Resorting to the more controllable junction of a scanning
recognizes that the wavelength of the photon limits the focus of tunneling microscope (STM), under the rubric of tip-enhanced
the optical microscope to a spot of ∼λ/2, with λ ∼ 500 nm in Raman spectroscopy (TERS),10 has led to accelerated discovery
the visible region. To reach the 0.1 nm size scale of the atom, through TER spectromicroscopy (TER-sm).11,12 Noteworthy
photons must be squeezed down to a thousandth of their size.
These challenges are met by resorting to plasmonic Received: August 30, 2017
nanoantennas.1 The nanoscale analogue of the antenna used Accepted: October 4, 2017
by Hertz, consisting of a pair of metallic nanospheres separated Published: October 4, 2017

© 2017 American Chemical Society 11466 DOI: 10.1021/acsnano.7b06183


ACS Nano 2017, 11, 11466−11474
ACS Nano Article

among these was the realization of TERS with submolecular


spatial resolution13an effect that was not anticipated by then
standard classical models of tip plasmons.14 In the subnan-
ometer junction gaps necessary to reach single molecule
sensitivity, quantum effects cannot be dismissed.15−17 The
dynamics of optically driven electrons, which tunnel across
junctions, must be explicitly treated to describe the relevant
local fields.18,19 Atomistic electrodynamics simulations,20
coupled with quantum chemical treatments of the molecule
and the scattering process,21,22 are necessary for reliable
descriptions. Moreover, given the strong fields, field and matter
states can be entangled.23,24 The quantum electrodynamics of
plasmonic cavities, which confine photons, electrons, and the
molecule under study, provides a frontier in science that must
be mastered along the way to develop the chemiscope.
We use cobalt(II)-tetraphenylporphyrin (CoTPP) to carry
out a joint experimental and theoretical study of TERS at the
STM junction of a silver tip and an atomically flat Au(111)
substrate. CoTPP has been previously investigated through
STM,25 TERS,26 and theory.27 It saddles upon adsorption,
whereby the phenyl groups twist and tilt successive pyrroles up
and down (see Figure S1 in Supporting Information). The
saddled molecule contains the necessary geometric features to
scrutinize principles of vibrational spectromicroscopy with
Figure 1. Vibrational spectromicroscopy at 300 K. (a) Topographic
confined light. As an important milestone, we show room image (57 × 57 Å2) recorded at 10 pA, +0.15 V. (b) Simultaneously
temperature TER-sm at the same resolution as STM, but now recorded Raman map integrated over the full window (300−1700
with vibrational hyperspectral information. The spectra are cm−1). (c) Mode-specific map of the highest frequency vibration
thermally activated. At 80 K, they degrade to field-gradient- (integrated over the 1550−1600 cm−1 window colored by light pink
driven modes, which is Stark tuned. This allows TERS-relayed and blue bands). (d) Comparison between TERS of CoTPP on
vibrational Stark microscopy to image intramolecular charge Au(111) and simulated spectrum. Calculated normal modes nearest
distributions, with prior examples of such information extracted to the color-coded windows are illustrated accordingly. (e)
using atomically resolved Kelvin probe force microscopy Exponential decay of TERS intensity integrated over the 700−
(KPFM).28,29 At 5 K, we manipulate individual molecules to 1650 cm−1 window. The dashed line indicates the level for zero
intensity. Images are filtered due to the thermal noise at 300 K (raw
highlight the tensorial nature of single molecule spectroscopy
images are presented in Figure S4 in the Supporting Information).
with confined light, and show that the spatial resolution of
TERS reaches the 1 Å limit. The range of observed effects can
be rationalized through simulations using the atomic electro- presented spatial maps are the spectrally integrated TERS
dynamics model20,30 to describe confined fields and the dressed intensity (Figure 1b) and the mode-specific map of the
tensor formalism31 to treat spectroscopy with confined light in vibration at 1576 cm−1 (Figure 1c). The mode-specific
terms of local field/field gradient enhancement driven through vibrational images of all eight observed lines are provided in
derivatives of electric dipole−dipole, electric dipole−quadru- Figure S3 in the Supporting Information. They show subtle
pole, and electric quadrupole−quadrupole polarizabilities. variations with respect to the STM topography. Inspection of
Notwithstanding the necessary approximations, the combined the images in Figure 1a−c is sufficient to infer that the spatial
experiment and theory study yields deep insights on the salient resolution in the TER-sm is nearly the same as the STM.
features and prospects of TER-sm. A satisfactory reproduction of the spectrum is possible under
the assumption of electric field polarized along the tip axis and
RESULTS AND DISCUSSION molecules tilted out of plane. The best match is obtained upon
TER-sm. At room temperature, CoTPP forms an ordered tilting the molecule out of the substrate plane by 32° relative to
monolayer on Au(111), as seen in the simultaneously recorded the surface normal, followed by 42° rotation in the tilted
STM and hyperspectral TERS images in Figure 1a−c. The molecular plane. The field-active normal modes dominate the
latter consist of complete spectra recorded on each 1.4 Å × 1.4 spectrum, and at room temperature, the thermal population of
Å pixel with an acquisition time of 1 s/pixel. The TERS molecules that tilt out of the plane is observable.33,34 The
spectrum (Figure 1d) disappears upon 1 nm retraction of the observed vibrations are normal modes that are principally
tip (Figure 1e). The gap dependence of the signal intensity can localized on the phenyl rings, as illustrated in Figure 1.
be approximated as exponential, with a decay constant of 3.5 Å. Nevertheless, there are subtle spatial variations in the mode-
The scatter in the retraction data is due to motion of the specific images (see the Supporting Information). The maps of
molecule, both thermal and lift-off by the van der Waals the 1252 cm−1 (1220 cm−1 calculated) and 1576 cm−1 (1592
attraction to the tip.32 The spectrum (Figure 1d) consists of cm−1 calculated) mode are generally anticorrelated with the
three distinct components: the exponentially decaying back- location of the molecules, whereas the map of the mode at 1092
ground of electronic Raman scattering on the metal at shifts cm−1 (absent in the calculations) is closely correlated. The
larger than 1700 cm−1 (better shown in Figure 6c), a flat 1550 cm−1 is the only mode that does not involve phenyl
background with a cutoff near 1600 cm−1, and a set of thermally vibrations. It is mainly localized on the pyrroles, which pull out
broadened molecular vibrational lines within this window. The of the porphyrin plane in the saddled molecule.
11467 DOI: 10.1021/acsnano.7b06183
ACS Nano 2017, 11, 11466−11474
ACS Nano Article

Confined Field, Photon, and Signal. The enhanced smaller, ∼5 × 10−3 V/Å. Only a small fraction by volume, 10−4,
background without vibrational peaks tracks the STM top- of one photon is ever present at the junction.
ography with high fidelity. This is clarified by the superimposed The observed localization of the signal is consistent with the
contour maps of the STM topography and the background field confinement expected from the atomically terminated tip
(cyan band in Figure 1d) map in Figure 2a and associated line- presented in Figure 3. We use an icosahedral gold nanoparticle
consisting of 2057 atoms against a planar gold slab to calculate
the local fields. The dimensions and gap-independent
resonance of the structure, which matches the experiment,
2.32 eV (534 nm), is shown in Figure S5. The strongest and
most confined near-field is obtained at a slightly lower energy
than the plasmonic resonance.21 At the depicted gap distance of
5.5 Å, which corresponds to the experiment, the hot spot of the
Au-atom-terminated tip has a lateral width at half-maximum of
E21/2 ∼ 6 Å (Figure S5b), with associated field gradients that are
more tightly confined. When the quartic dependence of the
Figure 2. Raman signal of confined cavity mode. (a) Overlay of the
STM topography (black contours) and the cavity mode map (cyan
signal on the local field is taken into account, the predicted
contours, integrated over the cyan band in Figure 1d, 320−500 lateral confinement of the field-induced signal is ∼3 Å, in
cm−1). (b) Line profiles of topography (black) and cavity map agreement with the experiment. The signal is localized down to
(cyan) taken along the two diagonal lines in (a), which are the size of an atom, nearly 4 orders of magnitude sharper than
displaced by 4 Å. allowed by the Abbe principle of optics and significantly smaller
than the spatial extent of vibrational normal modes that extend
cuts shown in Figure 2b. The background is assigned to the over the molecule.
cavity mode, namely, the confined optical field in the junction. Silence and Activation of TERS with CO. Upon being
Its intensity perfectly tracks the STM topography, but with a cooled to 80 K, the molecular TERS signal disappears. Due to
diagonal displacement of ∼4 Å. The apparent displacement of the improved thermal stability and planarity of the cold
the optical signal from the terminal tip Ag atom by the molecules on the substrate, sharp STM images of the molecular
array can be recorded, as in Figure 4a. However, the
dimensions of one atom provides a direct measure of the lateral
simultaneously collected TERS images are completely feature-
confinement of the TERS signal to an area of ∼16 Å2. Together less. Indeed, if the scattering was electronically resonant via the
with the decay length of the intensity with the vertical tip Q-band, then cancellation of the in-plane transition dipole via
retraction, a signal confinement volume of ∼100 Å3 can be its image in the gold mirror could explain the total silence.
inferred. Although it is tempting to assign this volume to the However, the excitation wavelength at 634 nm is removed from
confinement of a photon, the local field that would arise from the Q ← X transition of CoTPP, which occurs at 530 nm
(Figure S2). Flattening the molecule from its room temperature
one such photon, E = ℏω/2ε0V = 1.2 V/Å , is not sensible
tilt into planarity reduces the signal by a factor of 5−10, but this
because it would ionize any molecule. From the observed does not explain the complete silence of the spectra given the
scattered intensity and the Raman cross section of the dynamic range of detection. In the depicted junctions (Figure
molecule, the local field can be estimated to be 100 times 3c,d), the hot spots of field and field gradient miss the aromatic

Figure 3. Sculpted junction fields. Cross-sectional views of near-field and field gradient distributions are illustrated for a flat gold tip (a,b),
atomically terminated tip (c,d), and CO-terminated tip (e,f).

11468 DOI: 10.1021/acsnano.7b06183


ACS Nano 2017, 11, 11466−11474
ACS Nano Article

the center frequency on the order of 1 cm−1. The large dynamic


range of the spectra identifies over 30 vibrational lines, with
direct lineage to the room temperature spectrum (see traces 3
and 4 in Figure 4d). Their single molecule TERS nature is
established by the retraction dependencethe spectrum
disappears upon 5 Å retraction of the tip (Figure 4d, trace
5). That the intercalated CO is only loosely attached to the
junction or to CoTPP is demonstrated by the trajectory in
Figure 4f: the activated spectrum is lost as soon as the scan
starts. Despite the high S/N ratio of the activated spectra, we
do not observe the characteristic CO vibrational line, which we
see when it binds to the apex at 5 K (see below). The possible
role of CO as a local field spoiler is considered by comparing
the field of an empty cavity and one in which a single CO
molecule or a single Au atom is introduced (Figure 3c−h). The
field lines and field gradients are sculpted quite differently in
each case. Their coexistence would generate significant
inhomogeneity on the molecular scale and destroy the extended
SS.
Alternatively, the spectra may be activated by chemical
binding of CO. As many as two CO molecules bind to the
central Co atom of CoTPP, with two possible binding
geometries: axial binding on opposite sides and bridge binding
with both carbonyls on the same face. The optimized structures
are shown in Figure 5, along with predicted spectra, generated
Figure 4. CO as a cavity spoiler. (a) Topography of CoTPP
monolayer at 80 K with resolved phenyl groups. The set point is 16
pA and 0.2 V. (b) Degraded topography of CoTPP obtained at 80
pA, 4.9 mV, due to CO lattice gas on the surface. (c)
Simultaneously recorded Raman map with flashes as CO molecules
pass by. Each pixel is the Raman intensity integrated from 600 to
1650 cm−1. (d) Extracted spectra from the map (c), 1 and 2,
highlight the flashes. Acquisition time is 1 s. Spectra 4 and 5 are
recorded with a stationary tip engaged and retracted by 5 Å,
respectively (10 s integration). The 80 K spectra 1 and 4 agree with
the spectrum acquired at room temperature, 3. (e) TERS trajectory
while the tip is stationary. (f) Disappearance of Raman signal upon
initiating a scan. Scale bars in (a−c) are 1 nm long.

macrocycle, which remains in the electron spill-out region


where the electric field is screened. Not included in the model
is the Schockley surface state (SS) of gold,35 which provides a
nearly free 2D electron density, famously visualized by the
construction of quantum corrals.36 The coupling of the
slouched molecule with the SS is suspected as the origin of
the silence, which would also explain the prior TERS imaging of Figure 5. CO-activated spectrum at 80 K and simulations assuming
planar porphyrins with bulky substituents that raise the (a) bare CoTPP and its dicarbonyl with (b) axial and (c) bridge
macrocycle plane.13 Consistent with this picture, the vibrational binding. The spectra are dominated by field contributions.
spectra are activated upon introducing a 2D lattice gas of
mobile molecules, ostensibly by destroying the SS or by lifting
molecules out of the zone of silence. by dressing molecular polarizabilities with field and field
Upon introducing CO molecules, which are mobile on gold gradient and adjusting their orientation (details in Table S1 of
at 80 K but eventually desorb, complete vibrational spectra of the Supporting Information). The comparison between experi-
CoTPP emerge. This is illustrated in the simultaneously ment and simulation is informative. Although the direct lineage
recorded images of Figure 4b,c: the STM image is streaky, due between the 80 K and room temperature bare molecule is
to the CO gas, and the simultaneously recorded TERS data established, the best agreement in Figure 5 is with the axial
show no correlation with the STM image or any spatial dicarbonyl. The contradiction illustrates that, in the single
features. However, the bright streaks that appear sporadically molecule limit, intensity patterns in spectra can be more
during the scan carry the sharp and rather complete vibrational strongly influenced by orientation than chemical structure.
spectrum of the cold molecule (Figure 4d, trace 1). The Here, the axially bound CO leads to tilting the molecule and
activated spectra are stable when probed with a stationary tip, raising its slouch. As a result, the in-plane normal modes are
as illustrated by the TERS trajectory in Figure 4e. The lines are enhanced (e.g., 1342 cm−1 mode) relative to bare CoTPP and
instrument limited to FWHM = 6.7 cm−1, with fluctuations in bridging dicarbonyl. The calculations clarify that the local field
11469 DOI: 10.1021/acsnano.7b06183
ACS Nano 2017, 11, 11466−11474
ACS Nano Article

Figure 6. Field-gradient-driven TER-sm and intramoleculear Stark imaging at 80 K. (a) STM topography of CoTPP lattice at 0.1 nA, 9.8 mV.
(b) Raman image anticorrelated with topography produced by integrating 1270 cm−1 mode intensity (634 nm laser at 5 μW/μm2, 1 s
accumulation per pixel). (c) Degraded TERS spectrum. (d) Background-subtracted experimental spectrum and simulation. (e) Zoomed-in
topography. (f−h) Super-resolution TER-sm. The 1270 cm−1 mode is fitted to a Gaussian (h), and the extracted intensity and frequency shift
are mapped in (f) and (g), respectively. The high pyrroles of saddled CoTPPs are resolved in the frequency map. (h) Line profiles from two
pixels in (g). Scale bars are 1 nm long. Images are filtered due to invasive imaging conditions. Raw images are presented in Figure S7 in the
Supporting Information.

dominates the spectra (E/∇E = 14 Å). Either geometry could, local contact potential. Due to the difference in work functions,
in principle, lift the macrocycle out of the zone of silence. a contact potential of 0.57 V can develop between Au(111) and
Vibrational Stark Microscopy: Polarization of the Ag tip if we assume [111] termination.39 Stark tuning of a
Pyrroles. TERS microscopy, albeit with a degraded vibrational mode within the molecule implies variation of charge density,
spectrum, becomes possible once the CO gas subsides. The therefore potential, inside the molecule. The largest frequency
simultaneously recorded STM and TERS images are shown in shift, 10 cm−1 relative to gold, occurs on the raised pyrroles of
Figure 6a,b along with the observed spectrum in Figure 6c. An the saddled molecule. The lower pyrroles appear inequivalent
adequate reproduction of the spectrum is obtained assuming due to the extent of their contacts to gold, whereas the cobalt
quadrupole−quadrupole scattering in extremely confined light, center appears at an intermediate potential. Clearly, the
E/∇E = 1.4 Å. Note that the contrast in the TERS images of molecule is charged, evidently by electron transfer to gold
Figure 6 is inverted. The TERS intensity is largest when the tip through the anchoring pyrroles. These findings agree with the
traces the metal intervening the molecules. This is, in part, due recent theoretical analysis of CoTPP on gold, which predicts
to the conductivity of the junction, which controls field nearly 1 electron transferred to gold, almost entirely from the
confinement, and it is consistent with field-gradient-driven anchoring pyrroles.27 In effect, through vibrational Stark tuning,
scattering, which has previously been demonstrated to be more super-resolved TER-sm maps out polarization and submolec-
efficient when the tip is placed outside the molecule.21 Despite ular charge distributions. An example of such a map was
the weak signal, spatially well-resolved images can be obtained produced through KPFM using CO-terminated tips.28 There,
by mapping out the Gaussian fit to the vibrational line at 1270 the force neutrality point as a function of applied potential is
cm−1, with advantage analogous to that of super-resolution in used to map the intramolecular charge distribution. Here, a
fluorescence microscopy.37 A given line now generates two normal mode of the targeted molecule is probed with the bare
maps, one of intensity and one of frequency shift, as shown for tip, with spatial resolution more localized than the mode itself
the magnified area scan in Figure 4f,g. The spatial detail in the (see Figure 2g).
frequency shift image is significantly higher than that in the Isolated Molecules on Terraces at 5 K. The governing
STM. The intensity map shows that the signal is peaked in the principles of TER-sm are tested through measurements on
intermolecular pockets where phenyl groups converge. The isolated single molecules immobilized at 5 K on the Au(111)
vibrational frequency shift arises from the Stark effect, due to substrate. The manipulation sequence on the flat terrace is
the local electrostatic field, F, sampled by the vibration. The summarized in Figure 7a−c. In the scanned area, we see three
effect is first-order in dipolar transitions of odd (u) symmetry, molecules, but TERS is silent. When the tip is placed on one of
Δν = ΔμF/h, second-order in strictly Raman-active modes of the phenyl groups, we only see the electronic Raman
even (g) symmetry, Δν = ΔαF2/2h.38 The large vibrational continuum of the tip, void of molecular lines (orange trace in
shift of 10 cm−1 observed across the image identifies an odd first panel). Upon pressing on the phenyl ring (Vb = −2 mV, I
transition that scatters into the far-field through the local field = 4 nA), a rather complete vibrational spectrum emerges (blue
gradient. We show in Figure S8 in the Supporting Information trace). The spectrum remains stationary as we translate the tip,
that the 1270 cm−1 mode becomes the main spectral line when indicating a molecule snagged by the tip. After measuring the
|∇E/E| ∼ 2. The large field gradients alter the traditional retraction-dependent sequence of spectra (middle panel), we
Raman selection rule so that quadupolar scattering dominates. repeat the same area scan (Figure 7b). The distorted image of
Because the measurement is carried out at a small bias, Vb = the remaining molecules confirms that the missing CoTPP
10 mV, the source of the electrostatic field is dominated by the molecule is firmly attached to the tip apex. Upon applying a
11470 DOI: 10.1021/acsnano.7b06183
ACS Nano 2017, 11, 11466−11474
ACS Nano Article

intensity arises from the relative orientation of molecule to


the local field. In the simulated spectra (Figure 7e), we employ
a vertically oriented free CoTPP, and |E/∇E|< 0.2 Å. The field
gradients play the key role in this case, which again indicates
that the field gradients break the traditional Raman selection
rules. However, when the molecule is lifted away by a few
angstrom, TERS dramatically changes. A larger ratio of field
and field gradient reproduces the spectrum at 2.13 Å, indicating
that the field contribution becomes increasingly dominant
during the tip retraction.
Tensorial TER-sm on the Angstrom Scale. As illustrated
in Figure 7, molecules isolated on terraces cannot be imaged at
gap distances required to see the TERS signal. They are
displaced or picked up by the tip even at 5 K. In contrast,
CoTPP molecules at step edges are strongly bound and remain
stationary under harsh tunneling conditions (4 nA, −10 mV).
The stationary molecules allow the direct demonstration of the
spatial resolution in TER-sm governed by field gradients.
Measurements on molecules decorating a step edge are
presented in Figure 8. We zoom in on three phenyl groups
from two different molecules that straddle the edge (Figure
8a−c) and record hyperspectral images in constant height
mode, enabled by the stability of the junction at 5 K. The
spectra observed on the three phenyls are displayed in Figure
8e. The mode-specific map of the shaded spectral window in
Figure 8d shows all three phenyls (P1-3). The TERS signal on
P2 arises from a 1 Å × 1 Å area, localized on the right side of
Figure 7. TER spectra upon transferring a CoTPP molecule to the the topographic image of P2 (see contour lines). The image of
tip at 5 K. The common set point for topographic images (104 × P3, which is closer to the tip, is larger; accordingly, the TERS
104 Å2) is 0.1 nA and +0.485 V. (a) First image of the transfer signal arises from a somewhat larger area, ∼2 Å × 2 Å, now
sequence. The Ag tip is placed on the phenyl ring denoted by the localized on the left side of P3. The spatial resolution is atomic
asterisk (*). The TER spectrum appears upon approaching the in scale, and the excitation is tensorial; it depends on the
molecule by changing the set point from 1 nA, +0.485 V to 4 nA, orientation of the molecule and confined field. Had we mapped
−2 mV, as demonstrated at the beginning of the spectral sequence
denoted by *, as well. (b) Image scanned after picking up a CoTPP the modes at 1500 cm−1, only P3 would have been imaged.
molecule. The disappearance of a molecule is marked by a dashed This captures the excitement and challenge of vibrational
circle. With molecule held at the tip apex, TER spectra are recorded spectromicroscopy with atomistically confined light.
as the tip is incrementally retracted and brought back to the
tunneling position. (c) Image taken after dropping the CoTPP by CONCLUSIONS
applying a −3 V pulse. In the process, the porphyrin is accidentally The reported measurements were selected to underscore the
exchanged with a CO molecule as identified by the TERS of C−O promise of TER-sm as the chemist’s ultimate microscope. We
stretch and enhanced resolution of the image. The landing position
is displaced from the position where the pulse was applied directly demonstrate that the method can yield vibrational
(denoted by x), which shows the CoTPP was adsorbed slightly off spectroscopy with angstrom-scale spatial resolution, signifi-
the apex and possibly tilted. (d) Gap dependences of two cantly sharper than the extent of molecular normal modes,
vibrational peaks at 807 and 1517 cm−1, respectively. The peak therefore at the ultimate limit of relevance to molecular matter.
intensities are obtained by integrating over shaded windows in the We have shown that cavity spectra alone lead to imaging of
spectral sequence. The exponential fit to the decay of 808 cm−1 molecular structure with resolution comparable to that of STM.
peak intensity yields 2.2 Å in decay length. (e) Simulated spectra Beyond structure, the inner workings of molecules are
compared with the experimental retraction dependence. accessible through vibrational Stark microscopy. Befitting a
chemist’s microscope, TER-sm provides structure, bond
voltage pulse, the molecule is dropped and verified to be the strengths, and charge distributions with submolecular reso-
intact CoTPP (Figure 7c). In this manipulation, the CoTPP lution. In the specific example of CoTPP considered here, we
spectrum is replaced with the single vibrational line of CO at show that the saddling of the molecule on Au(111) is
2065 cm−1. Now the tip is terminated with a CO molecule, accompanied by charge transfer through the pyrroles in contact
which also explains the sharper contrast in topography than in with the metal; the central cobalt atom does not participate in
the case of the bare tip (compare Figure 7a and 7c). Several the transfer, and the molecule and the surrounding gold atoms
observations are noteworthy about the gap dependence of the are polarized.27
spectra of the molecule attached to the tip apex. The entire The exposition also highlights the challenges that must be
spectrum disappears upon 2.4 Å retraction (Figure 7d), even overcome to develop a quantitative tool with analytical utility.
though the molecule remains on the apex. While the vibration Spectroscopy with confined light does not obey standard
localized on the phenyl rings at 807 cm−1 decays, the pyrrolic selection rules. A propensity rule previously noted is the silence
mode at 1517 cm−1 increases in intensity during the first 1 Å of cold molecules that lie flat on metallic surfaces.40 Here, we
retraction, and both decay exponentially over the next ∼1 Å suggest that this arises from coupling to the SS. They can be
retraction (Figure 7d). The mode-selective variation in activated by lifting or tilting with the tip using large tunneling
11471 DOI: 10.1021/acsnano.7b06183
ACS Nano 2017, 11, 11466−11474
ACS Nano Article

recognized that with simple molecular reporters attached to a


substrate, TERS may be used to characterize tip morphology.
Alternatively, with a simple reporter molecule attached to the
tip, TERS of the tip terminus can be used to probe molecular
matter. The CO-terminated tip, which we encountered here, is
particularly valuable in this regard because it can act as a
transducer of electrostatic forces with atomic resolution, as
already demonstrated in our validation of TERS-bright silver
tips.42 Quite clearly, image quality and resolution will be
directly related to the signal strength and acquisition speed, and
significant improvement over the present can be expected
through tip-enhanced stimulated Raman spectroscopy, as
already demonstrated in related work.43

METHODS
The instrumentation used in the current studies has been described
previously.40 We use a cryogenic, ultrahigh vacuum STM equipped
with a parabolic collector mounted on a piezoelectric stack. The focus
of the parabola can be precisely centered on the tip apex by imaging
electroluminescence (EL) and modeling the image obtained through
the train of optical elements.40 The laser excitation is focused through
an aspheric lens (focal length = 79 mm) at 45° incidence angle with an
estimated spot size of ∼10 μm. The Raman spectra are recorded using
continuous wave diode lasers, at 532 or 634 nm, with a typical
intensity of 5 μW/μm2 delivered at the sample. The principle advance
that has allowed us to obtain reproducible TERS results is the
fabrication of sharp silver tips with nanoscopically smooth surfaces, as
recently described.42 After electrochemically etching and polishing tips
with an automated setup,44 we use field-directed sputter sharpening as
the final process. A detailed description of the procedure has been
reported.42 A good indicator of the formation of junction cavities is the
observation of gap-dependent shift in the EL spectra,45 and this
criterion is used to establish the TERS utility of a tip when reshaped
through field emission. The fine-tuning of the cavity is accomplished in
operando, while monitoring the electronic Raman scattering
continuum of the junction. TERS requires alignment of the cavity
window with the Stokes scattering range. Either the broad resonance
of the cavity window is moved toward the Stokes window or the color
of the laser is chosen to match it. In the present case, the excitation
wavelength was shifted from 532 nm, where no signal could be seen
despite the Q-band electronic resonance of CoTPP (see Figure S2 in
the Supporting Information), to 634 nm to match with the cavity
Figure 8. Hyperspectral imaging on the angstrom scale of CoTPP window.
molecules adsorbed at a step edge. (a) Topography of molecules
decorating a step edge (66 × 66 Å2). (b) Close-up of the area ASSOCIATED CONTENT
marked by the yellow box in (a). White dashed lines trace two
CoTPP molecules that straddle the edge, with a pair of phenyls
*
S Supporting Information

each on upper and lower terraces (highlighted by black dashed The Supporting Information is available free of charge on the
lines). (c) Close-up of the area indicated by the yellow oval in (b); ACS Publications website at DOI: 10.1021/acsnano.7b06183.
note the scale bar. The common imaging condition for (a−c) is 0.1 Optimized geometry of CoTPP on Au(111); absorption
nA and +0.48 V. (d) Raman map of the same area as in (c) spectrum of CoTPP; mode-specific maps of all nine
recorded in constant height mode (4 nA, −10 mV). (e) Spectra vibrations seen at room temperature; raw images
recorded on the three phenyl groups (P1, P2, and P3) of two
recorded at room temperature; theoretical model of the
different molecules. The shaded band is the spectral window used
in the TERS map in (d). TERS junction; simulation of TERS spectra in the
dressed tensor formalism, junction resonance, field
currents near zero-bias, and more interestingly, the spectra can profile, and gap dependence of field strengths (Figure
be activated by the introduction of a TERS-invisible mobile 2D S5); raw images acquired at 80 K; field gradient driven
gas. In the present case, rather complete and sharp vibrational spectra (Figure S7); multipolar decomposition of Raman
spectra of individual CoTPP molecules have been obtained, scattering spectra (Figure S8); CO binding energies to
with high S/N ratio. Nevertheless, spectral interpretations are CoTPP and predicted orientations of CoTPP and
nontrivial as multipolar scattering driven by field gradients bicarbonyl CoTPP molecules at 80 K (PDF)
dominates spectroscopy with confined light. In the absence of
orientational averaging, the tensorial nature of scattering makes AUTHOR INFORMATION
it sensitive to 3D orientation of the targeted sample and the Corresponding Authors
cavity-confined fields.41 These are ultimately advantages, given *E-mail: jensen@chem.psu.edu.
atomistically characterized and controlled junctions. It may be *E-mail: aapkaria@uci.edu.
11472 DOI: 10.1021/acsnano.7b06183
ACS Nano 2017, 11, 11466−11474
ACS Nano Article

ORCID Chemical Mapping of a Single Molecule by Plasmon-Enhanced Raman


Joonhee Lee: 0000-0003-4685-3511 Scattering. Nature 2013, 498, 82−86.
(14) Chu, P.; Mills, D. Plasmonic Response of STM Tips. Phys. Rev.
Lasse Jensen: 0000-0003-1237-5021 B: Condens. Matter Mater. Phys. 2011, 84, 045430.
Vartkess Ara Apkarian: 0000-0002-7648-5230 (15) Song, P.; Nordlander, P.; Gao, S. Quantum Mechanical Study of
Author Contributions the Coupling of Plasmon Excitations to Atomic-Scale Electron
J.L. and N.T. equally contributed to the experimental effort. Transport. J. Chem. Phys. 2011, 134, 074701.
X.C. and P.L. carried out the modeling of the junctions and the (16) Savage, K. J.; Hawkeye, M. M.; Esteban, R.; Borisov, A. G.;
TERS calculations. V.A.A. and L.J. conceived and coordinated Aizpurua, J.; Baumberg, J. J. Revealing the Quantum Regime in
the work and authored the paper with input from all co-authors. Tunnelling Plasmonics. Nature 2012, 491, 574−577.
(17) Zhu, W.; Esteban, R.; Borisov, A. G.; Baumberg, J. J.;
Notes Nordlander, P.; Lezec, H. J.; Aizpurua, J.; Crozier, K. B. Quantum
The authors declare no competing financial interest. Mechanical Effects in Plasmonic Structures with Subnanometre Gaps.
Nat. Commun. 2016, 7, 11495.
ACKNOWLEDGMENTS (18) Gieseking, R. L.; Ratner, M. A.; Schatz, G. C. Review of
Plasmon-Induced Hot-Electron Dynamics and Related SERS Chem-
This research was supported by the grant of NSF Center for ical Effects. In Frontiers of Plasmon Enhanced Spectroscopy Volume 1;
Chemical Innovation dedicated to Chemistry at the Space- ACS Symposium Series; American Chemical Society: Washington,
Time Limit (CHE-1414466). We gratefully acknowledge Dr. DC, 2016; Vol. 1245, pp 1−22.
Laura Rios for her measurements of the porphyrin absorption (19) Barbry, M.; Koval, P.; Marchesin, F.; Esteban, R.; Borisov, A. G.;
spectrum, and Prof. Everly Fleischer for many stimulating Aizpurua, J.; Sánchez-Portal, D. Atomistic Near-Field Nanoplas-
discussions and for donating his collection of metalloporphyr- monics: Reaching Atomic-Scale Resolution in Nanooptics. Nano
ins. The cryo-UHV STM was constructed under Prof. W. Ho’s Lett. 2015, 15, 3410−3419.
supervision. Portions of this work were conducted with (20) Chen, X.; Moore, J. E.; Zekarias, M.; Jensen, L. Atomistic
Advanced CyberInfrastructure computational resources pro- Electrodynamics Simulations of Bare and Ligand-Coated Nano-
vided by The Institute for CyberScience at The Pennsylvania particles in the Quantum Size Regime. Nat. Commun. 2015, 6, 8921.
State University (http://ics.psu.edu). (21) Liu, P.; Chulhai, D. V.; Jensen, L. Single-Molecule Imaging
Using Atomistic Near-Field Tip-Enhanced Raman Spectroscopy. ACS
Nano 2017, 11, 5094−5102.
REFERENCES (22) Duan, S.; Tian, G.; Ji, Y.; Shao, J.; Dong, Z.; Luo, Y. Theoretical
(1) Novotny, L.; van Hulst, N. Antennas for Light. Nat. Photonics Modeling of Plasmon-Enhanced Raman Images of a Single Molecule
2011, 5, 83−90. with Subnanometer Resolution. J. Am. Chem. Soc. 2015, 137, 9515−
(2) Alù, A.; Engheta, N. Hertzian Plasmonic Nanodimer as an 9518.
Efficient Optical Nanoantenna. Phys. Rev. B: Condens. Matter Mater. (23) Roelli, P.; Galland, C.; Piro, N.; Kippenberg, T. J. Molecular
Phys. 2008, 78, 195111. Cavity Optomechanics as a Theory of Plasmon-Enhanced Raman
(3) Kneipp, K.; Wang, Y.; Kneipp, H.; Perelman, L. T.; Itzkan, I.; Scattering. Nat. Nanotechnol. 2015, 11, 164−169.
Dasari, R. R.; Feld, M. S. Single Molecule Detection Using Surface- (24) Benz, F.; Schmidt, M. K.; Dreismann, A.; Chikkaraddy, R.;
Enhanced Raman Scattering (SERS). Phys. Rev. Lett. 1997, 78, 1667− Zhang, Y.; Demetriadou, A.; Carnegie, C.; Ohadi, H.; de Nijs, B.;
1670. Esteban, R.; Aizpurua, J.; Baumberg, J. J. Single-Molecule Opto-
(4) Nie, S.; Emory, S. R. Probing Single Molecules and Single mechanics In “picocavities. Science 2016, 354, 726−729.
Nanoparticles by Surface-Enhanced Raman Scattering. Science 1997, (25) Seufert, K.; Bocquet, M.-L.; Auwärter, W.; Weber-Bargioni, A.;
275, 1102−1106. Reichert, J.; Lorente, N.; Barth, J. V. Cis-Dicarbonyl Binding at Cobalt
(5) Le Ru, E. C.; Etchegoin, P. G. Single-Molecule Surface-Enhanced and Iron Porphyrins with Saddle-Shape Conformation. Nat. Chem.
Raman Spectroscopy. Annu. Rev. Phys. Chem. 2012, 63, 65−87. 2011, 3, 114−119.
(6) Jeanmaire, D. L.; Van Duyne, R. P. Surface Raman (26) Domke, K. F.; Pettinger, B. In Situ Discrimination between
Spectroelectrochemistry: Part I. Heterocyclic, Aromatic, and Aliphatic
Axially Complexed and Ligand-Free Co Porphyrin on Au(111) with
Amines Adsorbed on the Anodized Silver Electrode. J. Electroanal.
Tip-Enhanced Raman Spectroscopy. ChemPhysChem 2009, 10, 1794−
Chem. Interfacial Electrochem. 1977, 84, 1−20.
1798.
(7) Surface-Enhanced Raman Scattering: Physics and Applications;
(27) Janet, J. P.; Zhao, Q.; Ioannidis, E. I.; Kulik, H. J. Density
Kneipp, K., Moskovits, M., Kneipp, H., Eds.; Springer: Berlin, 2010.
(8) Zhang, P.; Feist, J.; Rubio, A.; García-González, P.; García-Vidal, Functional Theory for Modelling Large Molecular Adsorbate−surface
F. J. Ab Initio Nanoplasmonics: The Impact of Atomic Structure. Phys. Interactions: A Mini-Review and Worked Example. Mol. Simul. 2017,
Rev. B: Condens. Matter Mater. Phys. 2014, 90, 161407. 43, 327−345.
(9) Trautmann, S.; Aizpurua, J.; Götz, I.; Undisz, A.; Dellith, J.; (28) Mohn, F.; Gross, L.; Moll, N.; Meyer, G. Imaging the Charge
Schneidewind, H.; Rettenmayr, M.; Deckert, V. A Classical Distribution within a Single Molecule. Nat. Nanotechnol. 2012, 7,
Description of Subnanometer Resolution by Atomic Features in 227−231.
Metallic Structures. Nanoscale 2017, 9, 391−401. (29) Albrecht, F.; Repp, J.; Fleischmann, M.; Scheer, M.; Ondrácě k,
(10) Pettinger, B.; Schambach, P.; Villagómez, C. J.; Scott, N. Tip- M.; Jelínek, P. Probing Charges on the Atomic Scale by Means of
Enhanced Raman Spectroscopy: Near-Fields Acting on a Few Atomic Force Microscopy. Phys. Rev. Lett. 2015, 115, 076101.
Molecules. Annu. Rev. Phys. Chem. 2012, 63, 379−399. (30) Payton, J. L.; Morton, S. M.; Moore, J. E.; Jensen, L. A Hybrid
(11) Zrimsek, A. B.; Chiang, N.; Mattei, M.; Zaleski, S.; McAnally, M. Atomistic Electrodynamics−Quantum Mechanical Approach for
O.; Chapman, C. T.; Henry, A. I.; Schatz, G. C.; Van Duyne, R. P. Simulating Surface-Enhanced Raman Scattering. Acc. Chem. Res.
Single-Molecule Chemistry with Surface- and Tip-Enhanced Raman 2014, 47, 88−99.
Spectroscopy. Chem. Rev. 2017, 117, 7583−7613. (31) Chulhai, D. V.; Jensen, L. Determining Molecular Orientation
(12) Jiang, N.; Kurouski, D.; Pozzi, E. A.; Chiang, N.; Hersam, M. C.; With Surface-Enhanced Raman Scattering Using Inhomogenous
Van Duyne, R. P. Tip-Enhanced Raman Spectroscopy: From Concepts Electric Fields. J. Phys. Chem. C 2013, 117, 19622−19631.
to Practical Applications. Chem. Phys. Lett. 2016, 659, 16−24. (32) Rios, L.; Lee, J.; Tallarida, N.; Apkarian, V. A. Hovering and
(13) Zhang, R.; Zhang, Y.; Dong, Z. C.; Jiang, S.; Zhang, C.; Chen, L. Twirling of Tethered Molecules by Confinement between Surfaces. J.
G.; Zhang, L.; Liao, Y.; Aizpurua, J.; Luo, Y.; Yang, J. L.; Hou, J. G. Phys. Chem. Lett. 2016, 7, 2461−2464.

11473 DOI: 10.1021/acsnano.7b06183


ACS Nano 2017, 11, 11466−11474
ACS Nano Article

(33) Shao, F.; Müller, V.; Zhang, Y.; Schlüter, A. D.; Zenobi, R.
Nanoscale Chemical Imaging of Interfacial Monolayers by Tip-
Enhanced Raman Spectroscopy. Angew. Chem. 2017, 129, 9489−9494.
(34) Jiang, N.; Chiang, N.; Madison, L. R.; Pozzi, E. A.; Wasielewski,
M. R.; Seideman, T.; Ratner, M. A.; Hersam, M. C.; Schatz, G. C.; Van
Duyne, R. P. Nanoscale Chemical Imaging of a Dynamic Molecular
Phase Boundary with Ultrahigh Vacuum Tip-Enhanced Raman
Spectroscopy. Nano Lett. 2016, 16, 3898−3904.
(35) Yan, B.; Stadtmüller, B.; Haag, N.; Jakobs, S.; Seidel, J.;
Jungkenn, D.; Mathias, S.; Cinchetti, M.; Aeschlimann, M.; Felser, C.
Topological States on the Gold Surface. Nat. Commun. 2015, 6, 10167.
(36) Crommie, M. F.; Lutz, C. P.; Eigler, D. M. Confinement of
Electrons to Quantum Corrals on a Metal Surface. Science 1993, 262,
218−220.
(37) Betzig, E.; Trautman, J. K. Near-Field Optics: Microscopy,
Spectroscopy, and Surface Modification beyond the Diffraction Limit.
Science 1992, 257, 189−195.
(38) Bishop, D. M. The Vibrational Stark Effect. J. Chem. Phys. 1993,
98, 3179.
(39) Michaelson, H. B. The Work Function of the Elements and Its
Periodicity. J. Appl. Phys. 1977, 48, 4729−4733.
(40) Tallarida, N.; Rios, L.; Apkarian, V. A.; Lee, J. Isomerization of
One Molecule Observed through Tip-Enhanced Raman Spectroscopy.
Nano Lett. 2015, 15, 6386−6394.
(41) Banik, M.; El-Khoury, P. Z.; Nag, A.; Rodriguez-Perez, A.;
Guarrottxena, N.; Bazan, G. C.; Apkarian, V. a. Surface-Enhanced
Raman Trajectories on a Nano-Dumbbell: Transition from Field to
Charge Transfer Plasmons as the Spheres Fuse. ACS Nano 2012, 6,
10343−10354.
(42) Tallarida, N.; Lee, J.; Apkarian, V. A. Tip-Enhanced Raman
Spectromicroscopy on the Å Scale: Bare and CO-Terminated Ag Tips.
Accepted for publication. ACS Nano DOI: 10.1021/acsnano.7b06022.
(43) Wickramasinghe, H. K.; Chaigneau, M.; Yasukuni, R.; Picardi,
G.; Ossikovski, R. Billion-Fold Increase in Tip-Enhanced Raman
Signal. ACS Nano 2014, 8, 3421−3426.
(44) Sasaki, S. S.; Perdue, S. M.; Rodriguez Perez, A.; Tallarida, N.;
Majors, J. H.; Apkarian, V. A.; Lee, J. Note: Automated Electro-
chemical Etching and Polishing of Silver Scanning Tunneling
Microscope Tips. Rev. Sci. Instrum. 2013, 84, 096109.
(45) Rendell, R.; Scalapino, D.; Mühlschlegel, B. Role of Local
Plasmon Modes in Light Emission from Small-Particle Tunnel
Junctions. Phys. Rev. Lett. 1978, 41, 1746−1750.

11474 DOI: 10.1021/acsnano.7b06183


ACS Nano 2017, 11, 11466−11474

You might also like