You are on page 1of 8

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: J. Phys. Chem. C 2018, 122, 3138−3145 pubs.acs.org/JPCC

Visualizing the Effect of Partial Oxide Formation on Single Silver


Nanoparticle Electrodissolution
Vignesh Sundaresan, Joseph W. Monaghan, and Katherine A. Willets*
Department of Chemistry, Temple University, Philadelphia, Pennsylvania 19122, United States
*
S Supporting Information

ABSTRACT: The impact of heterogeneous surface oxide


formation on the electrochemical performance of single silver
nanoparticles is explored using in situ superlocalization optical
microscopy. Silver nanoparticles are well-known to form a
natural oxide layer on their surface, but the effect of this oxide
layer on electrochemical reactions is not well understood. Here
we track the temporal and spatial dependence of electro-
dissolution of single silver nanoparticles in order to study the
role of surface oxide layers on electrochemical reactions.
Heterogeneity in electrodissolution kinetics is observed by
following the time-dependent loss in scattering intensity from
individual silver nanoparticles using dark-field scattering. Both
fast and slow dissolution kinetics are observed, with the
dominant pathway changing as a function of applied potential.
To understand this, superlocalization imaging is employed to follow the spatial variance of the electrodissolution process and
reveals that the silver nanoparticles undergo electrodissolution in either a spatially symmetric or asymmetric manner. We
hypothesize that asymmetric dissolution events are due to heterogeneity in the exposed silver sites on the nanoparticle surface
because of incomplete surface oxide formation. Polarization-resolved measurements support this hypothesis by revealing
anisotropic dissolution of the nanoparticles over time. By tracking the electrodissolution behavior of silver nanoparticles in both
the temporal and spatial domains, we provide an improved understanding of how heterogeneity in electrochemical reactions is
impacted by nanoparticle surface properties.

■ INTRODUCTION
Plasmonic nanoparticles (NPs) are used in a variety of
Single nanoparticle electrochemistry is essential to under-
stand nanoscale electrochemical mechanisms because it exposes
any surface-induced heterogeneity, which would be hidden in
applications such as substrates for surface-enhanced spectros-
an ensemble-averaged measurement.14−17 Electrodissolution/
copies (e.g., surface-enhanced Raman scattering, SERS),1−3
corrosion of single Ag NPs has been studied by several other
plasmonic heating,4−6 and plasmon-mediated electrochemical
groups using electrochemical,18−23 optical,24−31 and correlated
reactions.7,8 Silver NPs are of particular interest due to their electrochemical−optical approaches.32−34 Electrochemical
localized surface plasmon resonance lying in the visible region, measurements of single nanoparticle electrodissolution typically
the tunability of their spectral properties based on the particle involve measuring changes in current as a silver nanoparticle
size and shape, and their utility in plasmon-based electro- collides with an ultramicroelectrode and is subsequently
chemical reactions due to high generation of hot electrons.9,10 oxidized. The charge associated with each collision event
However, the main challenge associated with Ag NPs is the reveals the extent to which the particle is oxidized at the
formation of silver oxide on the NP surface, which damps the surface.18 Although the collision-based experiments show
plasmon response of the nanoparticles and reduces their heterogeneity in the magnitude of the oxidation current, the
efficacy as nanoscale electrodes.11 It is well-known that the collision of inactive particles cannot be observed, and particles
native oxide layer on the Ag NP lowers the enhancement factor that show sluggish or delayed oxidation kinetics cannot be
of the electromagnetic field surrounding the nanoparticle and easily identified. To overcome this limitation, optical methods
thus affects the SERS enhancement.12,13 However, the role of such as dark-field scattering or photoluminescence imaging are
the surface oxide on electrochemical reactions involving silver used,27 in which electrochemical reactions involving the
NPs is less well-known. Here we study the electrodissolution of nanoparticle modulate the intensity of its optical signature.
single Ag NPs to uncover both temporal and spatial
heterogeneity in the silver oxidation process and to understand Received: November 30, 2017
how surface oxide formation impacts local electrochemical Revised: January 4, 2018
reactions. Published: January 5, 2018

© 2018 American Chemical Society 3138 DOI: 10.1021/acs.jpcc.7b11824


J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

Figure 1. (A) Schematic of the experimental setup in which an electrochemical cell is mounted on an inverted optical microscope and dark field
scattering from single nanoparticles is observed as a function of an applied potential. (B) Scattering intensity vs time for three representative Ag NPs,
showing different oxidation behaviors: fast oxidation (<0.5 s, green trace), slow oxidation (>0.5 s, blue trace), and no oxidation (orange trace). The
corresponding optical images for each of the three cases are shown and color-coded accordingly. The applied potentials are indicated at the top of
the graph and are as follows: 0 V for 10 s, 1.2 V for 120 s, and open circuit potential (ocp) for 20 s. (C) Oxidation behavior summary for 250 Ag NPs
taken at four different potentials: 0.8, 1.0, 1.2, and 1.4 V. (D) Histogram of the time required for the particle intensity to drop to 50% of its original
value at 1.2 V (blue) and 1.4 V (pink). (E) Histogram of the time required for the particle intensity to drop to 90% of its original value at 1.2 V
(blue) and 1.4 V (pink). The red dashed lines indicate particles that either do not oxidize or do not reach the intensity threshold within 120 s.

This strategy allows every NP on the surface to be observed coated glass coverslips and attaching a silicone elastomer well
(provided they generate sufficient signal) and readily exposes with adequate dimensions as previously described.7 Two
inactive particles or delayed oxidation events. However, while separate Pt wires were inserted through the sidewalls of the
both the electrochemical and optical methods reveal hetero- silicone elastomer well and serve as the counter and reference
geneity in the electrodissolution of nanoparticles on an electrodes. The exposed area of the planar ITO electrode is
electrode surface, neither provides insight into why this ∼84 mm2, and the Pt wire counter electrode is ∼5 mm long
heterogeneous behavior occurs. In this report, we use dark- and 2 mm in diameter. The counter electrode is inserted
field superlocalization optical imaging to probe electro- through one wall of the silicone elastomer well, with a vertical
dissolution of single silver nanoparticles on an electrode surface distance between the ITO electrode and counter electrode of
and reveal both temporally and spatially heterogeneous ∼3.5 mm. An electrical connection was made to the working
behavior, which we attribute to differences in the formation electrode (ITO with the Ag NPs) by attaching copper wire to
of oxide layers on the nanoparticle surfaces. This method offers the ITO using an electrically conductive silver epoxy. All
a spatial precision better than 10 nm, which makes it a useful electrodes were connected to a potentiostat (CH750E, CH
tool for understanding real-time nanoscale variations in the Instruments). A 20 mM potassium nitrate solution (pH ∼ 7)
electrodissolution properties of single NPs on the electrode was used as the electrolyte in all experiments unless otherwise
surface. stated. During all experiments, the electrochemical cell is open

■ EXPERIMENTAL SECTION
Chemicals and Materials. Citrate-capped Ag NPs (73 ± 8
to the ambient atmosphere, and oxygen was not removed from
the electrolyte solution.
The spectroelectrochemical cell was mounted on top of an
nm diameter, λmax = 449 nm) were purchased from inverted optical microscope (Olympus IX-73) equipped with a
nanoComposix. Tin-doped indium oxide (ITO) coated glass 100× objective (Olympus oil immersion, NA 0.6), a tungsten−
cover slides (22 × 22 mm, 15−30 Ω/sq, and thickness #1) halogen white light source, and a dark-field condenser
were purchased from SPI Supplies. Conductive silver epoxy (Olympus U-DCD). White light passing through the dark-
(Electrically/Thermally Conductive Bond 556) was purchased field condenser was used to illuminate the working electrode,
from Electron Microscopy Sciences. A two-part silicone and the low angle scattered light from the Ag NPs was collected
elastomer (SYLGARD 184) and two-part epoxy adhesive through the objective and imaged onto an EM-CCD camera
(DEVCON 5 min epoxy) were purchased from Fisher (Princeton Instruments PhotonMAX) after passing through a
Scientific. The 2 mm platinum wire purchased from Alfa relay lens. The schematic of the experimental setup is shown in
Aesar was used for both reference and counter electrodes. Figure 1A. The images on the CCD camera were acquired with
Potassium nitrate (ReagentPlus, ≥99.0%), silver nitrate an acquisition time of 0.1 s per frame. The CCD was triggered
(99.9999% trace metal basis), and sodium hydroxide (BioXtra, using the potentiostat output signal to obtain synchronized
≥98%) were purchased from Sigma-Aldrich. All solutions were measurements.
prepared with nanopure water (18 MΩ·cm, arium pro, Superlocalization Imaging. The center position of each
Sartorius). Ag NP was obtained by localizing each diffraction-limited image
Experimental Setup. The spectroelectrochemical cell was of a single scatterer to a two-dimensional Gaussian equation
prepared by drop casting 73 nm diameter Ag NPs on ITO- (eq 1).
3139 DOI: 10.1021/acs.jpcc.7b11824
J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

⎡ ⎡ ⎛ y − y ⎞ 2 ⎤⎤ intensity drop by the time the potential was returned to open


⎢ 1 ⎢⎛ x − x 0 ⎞
2
0 ⎟ ⎥⎥
I(x , y) = z 0 + I0 exp⎢ − ⎜ ⎟ + ⎜⎜ ⎟ ⎥
circuit.) Moreover, when the 90% intensity drop criterion is
2 ⎢⎣⎝ sx ⎠ ⎝ y ⎠ ⎥⎦⎦
s imposed, there is a clear shift to shorter oxidation times for the

nanoparticles oxidized at 1.4 V vs 1.2 V. This study shows that
(1) kinetics of nanoparticle oxidation is heterogeneous over
In this equation, I(x,y) is the intensity of the diffraction-limited different oxidation potentials and that at least 1.2 V is required
spot, z0 is the background intensity, sx and sy are the standard to oxidize nearly all of the particles on the ITO substrate, with
deviation (or width) of the spot, and x0 and y0 are the center the significant overpotential (E > 1 V) most likely due to the
position of the spot. We approximate that the center of the sluggish electron transfer kinetics of the ITO substrate.27,35
Gaussian fit (e.g., x0, y0) represents the center of the scattering Spatial and Temporal Dependence of Single Ag NP
nanoparticle. Home-written MATLAB code was used to fit Electrodissolution. In order to further investigate the
each nanoparticle region-of-interest by a least-squares algorithm heterogeneity of electrochemical dissolution of single nano-
to obtain the center position. To ensure the quality of the fits, particles, a dark-field superlocalization technique with a spatial
only the center positions obtained with a coefficient of precision better than 10 nm was used. Each diffraction limited
determination (R2) above a threshold of 0.85 were considered. dark-field image of a single Ag NP was fit to a 2-D Gaussian

■ RESULTS AND DISCUSSION


Heterogeneity in the Oxidation Kinetics of Individual
equation, and the center position of the NP was determined
(Figure S3). To determine the precision of this approach, 1500
images of several single Ag NPs were recorded without
Ag NPs. To uncover the heterogeneity in the oxidation kinetics applying a potential, and their center positions were determined
of individual Ag NPs and their relationship to the applied (Figure S4). The calculated standard deviation in the center
oxidation potential, potential-dependent dark-field imaging was positions of the particles was less than 10 nm, which both
employed. The oxidation of Ag NPs was observed as a decrease reveals the precision of our technique and shows that
in the dark-field scattering intensity as each particle shrinks in nanoparticles are stationary under conditions when no potential
size during the electrodissolution process. In addition to is applied.
particle shrinkage, scattering intensity changes can be observed Next, an oxidation potential of 1.2 V was applied, and the
due to the shifts in the refractive index of the particles due to center position of the Ag NP undergoing electrochemical
changes in the native surface oxide layer as well as changes in dissolution was extracted by fitting the sequential dark-field
particle shape. Figure 1B shows three different types of optical images. A time- and potential-dependent scatter plot
oxidation behaviors that occur as the potential is stepped from was created for each Ag NP from the extracted center points as
0 to 1.2 V: (1) fast oxidation, in which the scattering intensity shown in Figure 2, left column. In all three cases shown, slow
drops to background in less than 0.5 s (green trace); (2) slow oxidation kinetics were observed (Figure 2, right column). The
oxidation, in which the scattering intensity falls gradually over behavior in Figure 2A (which we call symmetric) shows an
time but does not show complete oxidation by 0.5 s (blue example in which the spatial origin of the scattering does not
trace); and (3) no oxidation, indicated by no change in the change over time, even as the integrated dark field scattering
scattering intensity (orange trace). We note that the criterion intensity of the NP decreases (Figure 2B). In comparison,
for distinguishing fast oxidation from slow oxidation (0.5 s) is Figure 2C shows an example in which the spatial origin of the
arbitrarily chosen but allows us to easily discriminate the two NP scattering shifts as the Ag NP oxidizes, as indicated by the
oxidation behaviors in a large population of nanoparticles. color coding of the points; we call this behavior asymmetric.
Figure 1C summarizes the fraction of nanoparticles under- Figure 2E shows a third type of behavior (“jump” and
going fast, slow, and no oxidation at four different oxidation asymmetric), where the spatial origin of the NP scattering
potentials. At the lowest applied potential (0.8 V), about 85% changes abruptly as soon as the potential is switched from 0 V
of the particles show no oxidation, indicating that the (black dots) to 1.2 V (yellow points) at 10 s. After this initial
supporting ITO substrate requires a significant overpotential jump, the spatial origin shifts gradually over time as the particle
to generate silver oxidation. For potentials above 1.0 V, where oxidizes, similar to asymmetric behavior in Figure 2C.
silver oxidation is observed for most of the nanoparticles Additional examples of all three oxidation behaviors are
studied, slow oxidation behavior dominates fast oxidation (this shown in Figure S5. We note that particles undergoing fast
can also be observed in the graphs in Figure S1), although fast oxidation show only symmetric behavior, while particles that do
oxidation becomes more prominent at a potential of 1.4 V. not oxidize also show symmetric behavior indicating that they
Because slow oxidation events do not always show a continuous are stationary under an applied oxidation potential (Figure S6).
drop in intensity or fall completely to background (Figure S2), To differentiate the three types of behaviors described, we
we cannot fit the data to a single exponential (or other first quantified the shift in the center position (e.g., the
function) to extract kinetic data. Instead, we calculate the time magnitude of the “jump”), d, of each NP upon applying an
required for the scattering intensity to drop to either 50% or oxidation potential (Figure 3A). The d value is obtained by
90% of its original value for 250 individual nanoparticles and calculating the difference between the average center position
plot the results for the two most oxidizing potentials (1.2 and of the NP during the first 10 s at 0 V (red X in Figure 3A) and
1.4 V) in Figures 1D and 1E, respectively. The histogram in the average center position of the NP for the next 10 s at 1.2 V
Figure 1D shows that most particles reach 50% of their original (green X in Figure 3A). The calculated d values for the three
intensity by 40 s for both applied potentials (1.2 V (blue) and examples in Figure 2 were found to be 3, 6, and 33 nm for the
1.4 V (pink)). However, the histogram in Figure 1E shows the particles shown in panels A, C, and E, respectively (Table S1
emergence of multiple time scales required for the nano- has additional values for the examples in Figure S5). Figure 3B
particles to drop to 90% of their original intensity. (The data shows a histogram of d values for Ag NPs at three different
shown after the red dotted line correspond to particles that oxidation potentials. For all particles included in this histogram,
showed no oxidation or did not reach the respective threshold slow oxidation kinetics was observed, with a minimum
3140 DOI: 10.1021/acs.jpcc.7b11824
J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

Figure 3. (A) Illustration of how the d value is calculated to


differentiate “jump” behavior. The red X and green X indicate the
average center positions of the Ag NP at 0 V (0.1−10 s) and 1.2 V
(10.1−20 s), respectively. The distance between the red and green X’s
is calculated and indicated as d. (B) Histograms showing d values with
the bin size of 10 nm for a population of Ag NPs at three different
oxidation potentials (1.0 V, green; 1.2 V, red; 1.4 V, blue). (C)
Schematic showing the calculation of the asymmetry ratio. The red
Figure 2. Left: scatter plots showing the (x, y) center positions of a border in panel i corresponds to the center positions obtained after 1.2
single Ag NP at 0 V for the first 10 s (black dots) and at 1.2 V after 10 V was applied (starting from 11 s) until the particle showed a 90%
s (orange to red points). The color represents the evolution of time as drop from its initial scattering intensity. The area encompassed by the
shown in the color bar. Right: intensity−time traces obtained by red border is rotated as shown in panel ii (see Figure S8 for details),
integrating the intensity of the dark-field image of each respective and the asymmetry ratio was determined by calculating average
particle shown on the left. The applied potentials are shown at the top distance of each center position in x and y directions. (D) Histograms
of each graph. (A) The center position of the Ag NP is stationary of the asymmetry ratio (bin size is 0.4) of Ag NPs oxidizing at 1.0 V
during oxidation. (C) The center position changes continuously, (blue) and 1.2 V (pink).
moving toward the upper right as the particle oxidizes. (E) The center
position show a sudden change after 1.2 V is initially applied and then
gradually shifts upward as the particle oxidizes.
when the potential is held at an oxidizing value (see Figure 3C
oxidation time of 10 s. At 1.0 V (green data), the calculated d and Figure S8 for more details). In the case of ideal symmetric
for all particles except one are less than 10 nm, indicating the behavior, the calculated asymmetry ratio will be one, whereas
particles do not change their position when a positive potential for asymmetric behavior this ratio will be greater than one. The
is applied. However, at the higher potentials (1.2 and 1.4 V, red calculated asymmetry ratio for the particles shown in Figures
and blue, respectively), the d values shift toward larger values, 2A, 2C, and 2E are 0.8, 1.6, and 2.2, respectively, which agrees
indicating that many particles show a “jump” from their initial with the previous qualitative assignments of symmetric (panel
position as the potential is stepped from 0 to >1 V. Control A) and asymmetric (panels C and E) behavior. The asymmetry
studies carried out at no potential, 0.3, 0.6, and 0.9 V are shown ratio was calculated for >20 individual Ag NPs at two different
in Figure S7, and the value of d (for 25 particles in each case) is potentials (1.0 and 1.2 V), and the results are shown as
found to be <10 nm, similar to what is observed at 1.0 V. Since histograms in Figure 3D. These data show that both symmetric
d is dependent on the applied potential, we hypothesize that and asymmetric behaviors are observed, with ∼70% of the
the initial jump in the position of the NP upon applying particles showing asymmetric behavior. From the histograms
potentials ≥1.2 V is due to the electrostatic movement of the we note that the relative distribution of behaviors (symmetric
particle on the electrode surface. Given the heterogeneity of the vs asymmetric) does not depend on the applied oxidation
ITO film, we hypothesize there are local potential differences potential, suggesting that the origin of the asymmetry is not
across the substrate surface. At sufficiently positive potentials potential-dependent but is instead related to the properties of
(≥1.2 V), these regions of local substrate potentials exert an the Ag NPs.
electrostatic force on the negatively charged citrate-capped Ag Next, we examined the trajectories of the dissolution process,
NPs (ζ-potential= −41 mV) that overcomes the adhesion of in order to determine if the asymmetric trajectories originate
the particle to the surface, causing a shift in its position. For from particles shifting in position over time due to motion on
potentials <1.2 V, the electrostatic force is not high enough to the surface induced by electro-osmotic or electrophoretic flow.
overcome adhesive forces, and thus d < 10 nm (e.g., the To do this, the raw x and y coordinates were smoothed using a
particles do not move when the positive potential is applied). 9 point moving average filter, and the average center position of
Our next goal is to differentiate symmetric and asymmetric the Ag NP during the first 10 s at an oxidizing potential was set
behavior. To do this, we calculate an asymmetry ratio which we as the origin with all other positions calculated relative to this
define as the ratio of the spread in the center positions along value. The distance r and orientation θ relative to this origin
the long and short axis of the distribution during the times were calculated over time using eqs 2 and 3, respectively.
3141 DOI: 10.1021/acs.jpcc.7b11824
J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

r (t ) = (x(t ) − xi)2 + (y(t ) − yi )2 time to reach 50% of the original intensity). For example, the
(2) red circle in Figure 4B indicates a radius of 40 nm. Eight
⎡ y(t ) ⎤ trajectories reach this rmax value but show different oxidation
θ(t ) = tan−1⎢ ⎥ times, ranging from 10 to 50 s, as represented in different
⎣ x(t ) ⎦ (3) colors. Figure 4C shows a histogram of rmax values at 1.0 V
(blue) and 1.2 V (pink) and indicates that the distribution of
In these equations, xi and yi correspond to the origin, while x
rmax values is relatively independent of the applied oxidation
and y correspond to the center positions for each frame.
potential, further supporting our claim that electro-osmotic or
Polar plots showing the time-dependent trajectories of r and
electrophoretic motion is not responsible for the shift in the
θ for >20 representative Ag NPs at two different oxidation
nanoparticle position. Moreover, we do not observe trajectories
potentials are shown in Figures 4A and 4B. We note that the
that extend beyond the physical dimensions of the particles
(∼73 nm diameter).
Proposed Mechanism for Spatially Varying Electro-
dissolution. To explain both the changes in the spatial origin
of Ag NP scattering during the electrodissolution process as
well as why we observe both symmetric and asymmetric
oxidation behavior, we propose the mechanism depicted in
Figure 5 in which the thickness and/or heterogeneity of the

Figure 5. (A) Schematic showing how the native oxide layer on the
surface of the Ag NP influences electrodissolution of a single Ag NP to
undergo either symmetric (top) or asymmetric (bottom) dissolution
pathways. (B) Schematic showing the comparison between the
changes in the center position (peak of the Gaussian) of the Ag NP
undergoing symmetric vs asymmetric oxidation.

surface oxide layer impacts the electrodissolution behavior. We


also note that depending on the experimental conditions, Ag
NPs surface may have hydroxide, oxyhydride, or carbonate
layers in addition to a silver oxide layer which would impact the
Figure 4. (A, B) Polar plots showing calculated r and θ for 20 and 28 electrodissolution behavior. If the oxide layer on the NP is
particles undergoing electrodissolution at (A) 1.0 and (B) 1.2 V, either thin and uniform (top schematic of Figure 5A) or
respectively. The average center position of each Ag NP between 10.1 nonexistent, the Ag oxidizes to Ag+ uniformly across the
and 20 s (1.2 V) was set as its origin for the polar plot. The color of nanoparticle surface, yielding symmetric dissolution of the Ag
each trajectory represents the time taken for the particle to show a NP. In this case, the particle surface would dissolve uniformly
90% drop of its original intensity as indicated in the respective color
around the core, reducing the size (and thus scattering
bars. The maximum radius of each polar plot is 50 nm. (C)
Histograms showing rmax values for 15 representative Ag NPs at two intensity) from the Ag NP but leaving the center position in
different oxidation potentials (1.0 V, blue; 1.2 V, pink). a fixed position (Figure 5B, left pathway). On the other hand, if
the oxide layer is nonuniform across the particle surface
(bottom schematic of Figure 5A), the Ag will preferentially
orientations of the Ag NP trajectories show no directional oxidize to Ag+ from the more exposed side, leading to
preference and cover 360°, indicating that the change in the asymmetric dissolution behavior. During this asymmetric
spatial position of each Ag NP is not due to electro-osmotic or oxidation (right pathway of Figure 5B), the particle shape
electrophoretic flow within the electrochemical cell. From this and size change anisotropically, shifting the center position of
data, we also see that the length of each trajectory (rmax) does the scattering as the geometry of the nanoparticle changes and
not depend on the time it takes for each Ag NP to reach 90% of generating the observed asymmetric dissolution behavior. We
its original intensity, as indicated by the color of each trajectory note that in both pathways the presence of the oxide layer is
(see Figure S9 for the same data but colored according to the expected to slow the oxidation kinetics of the silver, consistent
3142 DOI: 10.1021/acs.jpcc.7b11824
J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

with both symmetric and asymmetric oxidation behavior shown in Figure 6B. The intensity of both polarization-resolved
observed for the slow oxidation events. images decreases at the same rate, and the calculated M value
To confirm our hypothesis that the spatial origin of the dark (Figure 6B bottom panel) is ∼0 throughout the oxidation of
field scattering changes due to anisotropic dissolution of the the particle, clearly indicating that the particle is oxidizing in an
particle and not due to particle movement, polarization- isotropic manner, similar to the proposed symmetric pathway
resolved studies were carried out by inserting a birefringent shown in Figure 5. Conversely, the Ag NP shown in Figures 6C
crystal in the detection path of our optical microscope. The and 6D shows variation between the intensities of the two
scattered light from the Ag NP is collected through the polarization-resolved images over time. The calculated M value
objective and sent through a calcite crystal, which splits the (Figure 6D, bottom panel) before oxidation is 0, showing that
scattered light into two orthogonal polarizations. Imaging a the particle is initially isotropic; however, upon applying an
single Ag NP through the calcite crystal produces two oxidizing potential at 10 s, the M value shifts toward negative
diffraction-limited images as shown in Figure 6A-i. For an values, confirming that the NP undergoes a change in its shape
and becomes increasingly anisotropic during the electro-
dissolution process. This Ag NP shows behavior consistent
with the asymmetric dissolution pathway in Figure 5.
Verifying the Role of Silver Oxide on Heterogeneous
Electrodissolution. To validate that the nature of particle
dissolution (symmetric vs asymmetric) is influenced by the
silver oxide layer on the NP surface, electrodissolution
experiments were carried out at basic pH. The formation of
the oxide layer is accelerated under basic conditions,39−41 which
results in the formation of a thick silver oxide layer on the NP
surface and prevents electrodissolution of the particle. For these
experiments, the electrolyte pH was adjusted to pH 9 by adding
NaOH to the 20 mM KNO3 solution. Figure S10 shows the
results of these experiments with two key results: First, the
particles show intensity damping during the first 10 s when no
oxidation potential is applied, consistent with the accelerated
formation of the native oxide layer on the silver in the higher
pH solution. (We note that this initial loss of intensity was
Figure 6. (A, C) Series of dark-field optical images of a single Ag NP observed in other experiments as well and suggests that light
undergoing (A) symmetric oxidation and (C) asymmetric oxidation as helps drive oxide formation at 0 V, although this is beyond the
imaged through a birefringent crystal at (i) 0, (ii) 30, and (iii) 130 s. scope of the present study.) Second, upon applying 1.2 V, no
Scale bars = 1 μm. (B, D) Intensity−time traces of the left image, right electrodissolution of the Ag NPs was observed, which suggests
image, and sum of both images (top panel) and changes in the that the surface silver oxide layer is thick enough to prevent
calculated M value over time (bottom panel) for (B) symmetric and further oxidation of Ag NPs to silver ions.
(D) asymmetric oxidation corresponding to panels A and C,
To further confirm the role of the surface silver oxide on Ag
respectively. The small variation around 110 s was due to a different
particle diffusing in the background during the measurement. NP electrodissolution, we studied the electrodissolution
process of Ag particles without any surface oxide layer. Silver
particles were electrodeposited at −0.2 V from a 1 μM AgNO3
isotropic nanoparticle, the two polarization-resolved images will solution on the ITO electrode to generate surface oxide free Ag
have identical intensities; however, anisotropic nanoparticles particles (see Movie S1 in the Supporting Information).
will show different intensities between the two channels Immediately following electrodeposition, electrodissolution of
depending on the orientation of the particle.36,37 To quantify the pristine Ag particles was performed in 20 mM KNO3,
the anisotropy of a nanoparticle, we define the M value in eq 4 where the electrode was held at 0 V for the first 10 s (Figure
Ileft − Iright 7A) and then switched to an oxidizing potential of 1.2 V at 10 s
M= (Figure 7B). The intensity time traces of 10 different
Ileft + Iright (4) electrodeposited particles during electrodissolution (Figure
where Ileft and Iright correspond to the intensity of the left and 7C) show that the intensity of all particles drop sharply to
right polarization-resolved images, respectively. In the case of a background (within 100 ms) upon applying an oxidation
perfectly symmetric particle (spherical shape) the M value is 0, potential, indicating the spontaneous electrodissolution of Ag
whereas an asymmetric particle will have an |M| value between particles. This further supports our claim that surface oxide
0 and 1, depending upon the extent of the anisotropy.38 formation is responsible for the sluggish electrodissolution
During the silver oxidation experiments, two images are kinetics of the Ag NPs and is thus responsible for the observed
asymmetric electrodissolution trajectories.


obtained for each individual Ag NP, as shown in Figure 6A, i−
iii. The potential of the working electrode is held at 0 V for the
first 10 s and then switched to the oxidation potential of 1.2 V. CONCLUSIONS
As the particle starts to oxidize, the intensity of the two images In summary, we used optical dark-field scattering to study the
decreases, indicating a decrease in particle size as shown in the oxidation of individual Ag NPs on ITO at various potentials
series of dark-field optical images in Figure 6A (i, ii, and iii and revealed that surface oxide formation impacts both the
correspond to 0, 30, and 130 s, respectively). The kinetics and the spatially dependent oxidation behavior.
corresponding intensity−time trace is obtained by integrating Intensity−time traces of individual NPs show that electro-
the intensity of left and right images separately over time as dissolution kinetics are highly heterogeneous, with some
3143 DOI: 10.1021/acs.jpcc.7b11824
J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

■ ACKNOWLEDGMENTS
The support of this work by the AFOSR MURI (FA9550-14-1-
0003) is gratefully acknowledged. The authors thank Dr. Yun
Yu, Dr. Martin A. Edwards, and Dr. Donald A. Robinson for
discussions and their valuable suggestions.

■ REFERENCES
(1) Halas, N. J.; Lal, S.; Chang, W.-S.; Link, S.; Nordlander, P.
Plasmons in Strongly Coupled Metallic Nanostructures. Chem. Rev.
2011, 111, 3913−3961.
(2) Laurence, T. A.; Braun, G.; Talley, C.; Schwartzberg, A.;
Moskovits, M.; Reich, N.; Huser, T. Rapid, Solution-Based Character-
ization of Optimized SERS Nanoparticle Substrates. J. Am. Chem. Soc.
2009, 131, 162−169.
(3) Sakir, M.; Pekdemir, S.; Karatay, A.; Kücu̧ ̈köz, B.; Ipekci, H. H.;
Elmali, A.; Demirel, G.; Onses, M. S. Fabrication of Plasmonically
Figure 7. (A) Dark-field optical image of electrodeposited Ag particles Active Substrates Using Engineered Silver Nanostructures for SERS
at 0.1 s held at 0 V. (B) Optical image showing bare ITO after the Applications. ACS Appl. Mater. Interfaces 2017, 9, 39795−39803.
electrodissolution of all the particles in panel A within 100 ms upon (4) Herzog, J. B.; Knight, M. W.; Natelson, D. Thermoplasmonics:
applying oxidation potential at 10 s. Scale bars = 5 μm. (C) Intensity Quantifying Plasmonic Heating in Single Nanowires. Nano Lett. 2014,
time traces showing rapid electrodissolution of 10 different electro- 14, 499−503.
deposited Ag particles. (5) Aibara, I.; Chikazawa, J.; Uwada, T.; Hashimoto, S. Localized
Phase Separation of Thermoresponsive Polymers Induced by
Plasmonic Heating. J. Phys. Chem. C 2017, 121, 22496−22507.
(6) Chang, C.; Yang, C.; Liu, Y.; Tao, P.; Song, C.; Shang, W.; Wu, J.;
Deng, T. Efficient Solar-Thermal Energy Harvest Driven by Interfacial
particles undergoing fast oxidation while others oxidize over Plasmonic Heating-Assisted Evaporation. ACS Appl. Mater. Interfaces
much longer time scales. Superlocalization and polarization- 2016, 8, 23412−23418.
resolved optical imaging of individual Ag NPs reveal that the (7) Wilson, A. J.; Willets, K. A. Visualizing Site-Specific Redox
particle shape can change anistropically during electro- Potentials on the Surface of Plasmonic Nanoparticle Aggregates with
dissolution based on the nature of surface oxide layer. This Superlocalization SERS Microscopy. Nano Lett. 2014, 14, 939−945.
finding is useful to explain the variation in the magnitude of the (8) Wilson, A. J.; Molina, N. Y.; Willets, K. A. Modification of the
observed current during collision-based single Ag NPs Electrochemical Properties of Nile Blue through Covalent Attachment
electrodissolution experiments and provides insight into the to Gold As Revealed by Electrochemistry and SERS. J. Phys. Chem. C
effect of surface oxides on electrochemical reactions using silver 2016, 120, 21091−21098.
nanoparticle electrodes. Lastly, the approach described herein is (9) Hartland, G. V.; Besteiro, L. V.; Johns, P.; Govorov, A. O. What’s
so Hot about Electrons in Metal Nanoparticles? ACS Energy Lett.
versatile and can be extended to understand various electro- 2017, 2, 1641−1653.
chemical and galvanic reactions occurring at nanoscale (10) Willets, K. A.; Wilson, A. J.; Sundaresan, V.; Joshi, P. B. Super-
electrodes.


Resolution Imaging and Plasmonics. Chem. Rev. 2017, 117, 7538−
7582.
ASSOCIATED CONTENT (11) Yin, Y.; Li, Z.-Y.; Zhong, Z.; Gates, B.; Xia, Y.; Venkateswaran, S.
* Supporting Information
S Synthesis and Characterization of Stable Aqueous Dispersions of Silver
The Supporting Information is available free of charge on the Nanoparticles through the Tollens Process. J. Mater. Chem. 2002, 12,
ACS Publications website at DOI: 10.1021/acs.jpcc.7b11824. 522−527.
(12) Erol, M.; Han, Y.; Stanley, S. K.; Stafford, C. M.; Du, H.;
Statistics for three different oxidation behaviors at various Sukhishvili, S. SERS Not To Be Taken for Granted in the Presence of
potentials, diverse slow oxidation behaviors, schematics Oxygen. J. Am. Chem. Soc. 2009, 131, 7480−7481.
of the superlocalization imaging, superlocalization (13) Han, Y.; Lupitskyy, R.; Chou, T.-M.; Stafford, C. M.; Du, H.;
imaging of Ag NP under no potential and 1.2 V, Sukhishvili, S. Effect of Oxidation on Surface-Enhanced Raman
asymmetry ratio calculation details, histograms of d Scattering Activity of Silver Nanoparticles: A Quantitative Correlation.
under various electrochemical potentials, polar plots of r Anal. Chem. 2011, 83, 5873−5880.
(14) Mirkin, M. V.; Sun, T.; Yu, Y.; Zhou, M. Electrochemistry at
at two different potentials, oxidation of Ag NP at basic One Nanoparticle. Acc. Chem. Res. 2016, 49, 2328−2335.
pH 9 (PDF) (15) Anderson, T. J.; Zhang, B. Single-Nanoparticle Electrochemistry
Video showing the electrodeposition of Ag particle on through Immobilization and Collision. Acc. Chem. Res. 2016, 49,
ITO electrode (AVI) 2625−2631.


(16) Fang, Y.; Wang, H.; Yu, H.; Liu, X.; Wang, W.; Chen, H.-Y.;
AUTHOR INFORMATION Tao, N. J. Plasmonic Imaging of Electrochemical Reactions of Single
Nanoparticles. Acc. Chem. Res. 2016, 49, 2614−2624.
Corresponding Author (17) Yu, Y.; Sundaresan, V.; Bandyopadhyay, S.; Zhang, Y.; Edwards,
*E-mail: kwillets@temple.edu (K.A.W.). M. A.; McKelvey, K.; White, H. S.; Willets, K. A. Three-Dimensional
ORCID Super-Resolution Imaging of Single Nanoparticles Delivered by
Vignesh Sundaresan: 0000-0001-9390-1681 Pipettes. ACS Nano 2017, 11, 10529−10538.
(18) Oja, S. M.; Robinson, D. A.; Vitti, N. J.; Edwards, M. A.; Liu, Y.;
Katherine A. Willets: 0000-0002-1417-4656 White, H. S.; Zhang, B. Observation of Multipeak Collision Behavior
Notes during the Electro-Oxidation of Single Ag Nanoparticles. J. Am. Chem.
The authors declare no competing financial interest. Soc. 2017, 139, 708−718.

3144 DOI: 10.1021/acs.jpcc.7b11824


J. Phys. Chem. C 2018, 122, 3138−3145
The Journal of Physical Chemistry C Article

(19) Robinson, D. A.; Liu, Y.; Edwards, M. A.; Vitti, N. J.; Oja, S. M.; (38) Brasselet, S. Polarization-Resolved Nonlinear Microscopy:
Zhang, B.; White, H. S. Collision Dynamics during the Electro- Application to Structural Molecular and Biological Imaging. Adv.
oxidation of Individual Silver Nanoparticles. J. Am. Chem. Soc. 2017, Opt. Photonics 2011, 3, 205.
139, 16923−16931. (39) Peretyazhko, T. S.; Zhang, Q.; Colvin, V. L. Size-Controlled
(20) Zhang, F.; Edwards, M. A.; Hao, R.; White, H. S.; Zhang, B. Dissolution of Silver Nanoparticles at Neutral and Acidic PH
Collision and Oxidation of Silver Nanoparticles on a Gold Nanoband Conditions: Kinetics and Size Changes. Environ. Sci. Technol. 2014,
Electrode. J. Phys. Chem. C 2017, 121, 23564−23573. 48, 11954−11961.
(21) Ngamchuea, K.; Clark, R. O. D.; Sokolov, S. V.; Young, N. P.; (40) Badawy, A. M. E.; Luxton, T. P.; Silva, R. G.; Scheckel, K. G.;
Batchelor-McAuley, C.; Compton, R. G. Single Oxidative Collision Suidan, M. T.; Tolaymat, T. M. Impact of Environmental Conditions
Events of Silver Nanoparticles: Understanding the Rate-Determining (PH, Ionic Strength, and Electrolyte Type) on the Surface Charge and
Chemistry. Chem. - Eur. J. 2017, 23, 16085−16096. Aggregation of Silver Nanoparticles Suspensions. Environ. Sci. Technol.
(22) Ustarroz, J.; Kang, M.; Bullions, E.; Unwin, P. R. Impact and 2010, 44, 1260−1266.
(41) Bard, A. J.; Parsons, R.; Jordan, J. Standard Potentials in Aqueous
Oxidation of Single Silver Nanoparticles at Electrode Surfaces: One
Solution; Monographs in Electroanalytical Chemistry and Electrochemistr;
Shot versus Multiple Events. Chem. Sci. 2017, 8, 1841−1853.
Taylor & Francis: New York, 1985.
(23) Ma, W.; Ma, H.; Chen, J.-F.; Peng, Y.-Y.; Yang, Z.-Y.; Wang, H.-
F.; Ying, Y.-L.; Tian, H.; Long, Y.-T. Tracking Motion Trajectories of
Individual Nanoparticles Using Time-Resolved Current Traces. Chem.
Sci. 2017, 8, 1854−1861.
(24) Hao, R.; Fan, Y.; Zhang, B. Imaging Dynamic Collision and
Oxidation of Single Silver Nanoparticles at the Electrode/Solution
Interface. J. Am. Chem. Soc. 2017, 139, 12274−12282.
(25) Fang, Y.; Wang, W.; Wo, X.; Luo, Y.; Yin, S.; Wang, Y.; Shan, X.;
Tao, N. Plasmonic Imaging of Electrochemical Oxidation of Single
Nanoparticles. J. Am. Chem. Soc. 2014, 136, 12584−12587.
(26) Patel, A. N.; Martinez-Marrades, A.; Brasiliense, V.; Koshelev,
D.; Besbes, M.; Kuszelewicz, R.; Combellas, C.; Tessier, G.; Kanoufi,
F. Deciphering the Elementary Steps of Transport-Reaction Processes
at Individual Ag Nanoparticles by 3D Superlocalization Microscopy.
Nano Lett. 2015, 15, 6454−6463.
(27) Hill, C. M.; Bennett, R.; Zhou, C.; Street, S.; Zheng, J.; Pan, S.
Single Ag Nanoparticle Spectroelectrochemistry via Dark-Field
Scattering and Fluorescence Microscopies. J. Phys. Chem. C 2015,
119, 6760−6768.
(28) Brasiliense, V.; Berto, P.; Combellas, C.; Tessier, G.; Kanoufi, F.
Electrochemistry of Single Nanodomains Revealed by Three-Dimen-
sional Holographic Microscopy. Acc. Chem. Res. 2016, 49, 2049−2057.
(29) Smith, J. G.; Jain, P. K. The Ligand Shell as an Energy Barrier in
Surface Reactions on Transition Metal Nanoparticles. J. Am. Chem.
Soc. 2016, 138, 6765−6773.
(30) Smith, J. G.; Yang, Q.; Jain, P. K. Identification of a Critical
Intermediate in Galvanic Exchange Reactions by Single-Nanoparticle-
Resolved Kinetics. Angew. Chem., Int. Ed. 2014, 53, 2867−2872.
(31) Smith, J. G.; Jain, P. K. Kinetics of Self-Assembled Monolayer
Formation on Individual Nanoparticles. Phys. Chem. Chem. Phys. 2016,
18, 23990−23997.
(32) Brasiliense, V.; Patel, A. N.; Martinez-Marrades, A.; Shi, J.;
Chen, Y.; Combellas, C.; Tessier, G.; Kanoufi, F. Correlated
Electrochemical and Optical Detection Reveals the Chemical
Reactivity of Individual Silver Nanoparticles. J. Am. Chem. Soc. 2016,
138, 3478−3483.
(33) Batchelor-McAuley, C.; Martinez-Marrades, A.; Tschulik, K.;
Patel, A. N.; Combellas, C.; Kanoufi, F.; Tessier, G.; Compton, R. G.
Simultaneous Electrochemical and 3D Optical Imaging of Silver
Nanoparticle Oxidation. Chem. Phys. Lett. 2014, 597, 20−25.
(34) Sun, L.; Fang, Y.; Li, Z.; Wang, W.; Chen, H. Simultaneous
Optical and Electrochemical Recording of Single Nanoparticle
Electrochemistry. Nano Res. 2017, 10, 1740−1748.
(35) Hill, C. M.; Pan, S. A Dark-Field Scattering Spectroelec-
trochemical Technique for Tracking the Electrodeposition of Single
Silver Nanoparticles. J. Am. Chem. Soc. 2013, 135, 17250−17253.
(36) Sönnichsen, C.; Alivisatos, A. P. Gold Nanorods as Novel
Nonbleaching Plasmon-Based Orientation Sensors for Polarized
Single-Particle Microscopy. Nano Lett. 2005, 5, 301−304.
(37) Sinkó, J.; Gajdos, T.; Czvik, E.; Szabó, G.; Erdélyi, M.
Polarization Sensitive Localization Based Super-Resolution Micros-
copy with a Birefringent Wedge. Methods Appl. Fluoresc. 2017, 5,
017001.

3145 DOI: 10.1021/acs.jpcc.7b11824


J. Phys. Chem. C 2018, 122, 3138−3145

You might also like