Sie sind auf Seite 1von 9

191

Full Paper

Spatial Imaging of Cu2-Ion Release by Combining Alternating


Current and Underpotential Stripping Mode Scanning
Electrochemical Microscopy
Dirk Ruhlig, Wolfgang Schuhmann*
Analytische Chemie Elektroanalytik & Sensorik, Ruhr-Universitt Bochum, Universittsstr. 150, 44780 Bochum, Germany
*e-mail: wolfgang.schuhmann@rub.de
Received: July 17, 2006
Accepted: October 24, 2006
Abstract
Anodic underpotential stripping voltammetry was integrated into SECM in order to characterize local corrosion of
metallic copper deposits on metal surfaces as a model for copper containing alloys. Primarily, the alternating current
mode of SECM was applied in an electrolyte of low ionic strength for localizing possible corrosion sites without any
perturbation of the corroding surface, e.g., by the presence of any redox mediator. Sequentially, the release of Cu2ions was confirmed and locally visualized at the previously detected electrochemically active sites by means of
spatially resolved anodic underpotential stripping voltammetry performed during SECM scanning. Underpotential
stripping voltammetry of Cu2-ions was performed at a specifically developed 15 mm gold-coated Pt microelectrode
used as SECM tip with a detection limit of 0.15 nM Cu2 (N 4, RSD 6%) for an accumulation of 45 s at  0.4 V.
SECM images of model samples such as copper coated microelectrodes and lacquered metallic copper workpieces
demonstrated the feasibility and applicability of combining AC- and underpotential stripping mode of SECM for local
visualization of Cu2-ion release from corroding surfaces.
Keywords: Underpotential stripping voltammetry, Scanning electrochemical microscopy, AC-SECM, SM-SECM,
Corrosion, Copper determination
DOI: 10.1002/elan.200603693

1. Introduction
Metal alloys are used in numerous applications as materials
with specific mechanical properties. However, they may
release metal ions during corrosive degradation. For example, shape memory alloys based on NiTi or NiTiCu are used
to manufacture vascular stents, staples, orthopedic clamps,
or orthodontic wires [1, 2]. Corrosion causes the release of
hypoallergenic or biotoxic Ni2- and Cu2-ions into the body
[3], or may even lead to a breakdown of structural integrity
of the material [4, 5]. Thus, metal alloys are usually coated
with a very thin passivating film [6] to enhance corrosion
resistance. However, in the case of shape memory alloys the
superelasticity of the material combined with a temperature
dependent Martensite/Austenite transition may cause severe stress for the protecting layer. The composition,
stability, and homogeneity of the passivating or protective
layer are vital for the stability and corrosion resistance of
metal alloy surfaces. Hence, variations in the local thickness
of passivating layers and inclusions are said to promote local
corrosion. In order to identify possible corrosion sites,
spatially resolved characterization of the local activity of the
metal surface is needed. Several techniques such as local
electrochemical impedance spectroscopy (LEIS) [7], scanning reference (SRET) and scanning vibrating electrode
techniques (SVET) [8], scanning Kelvin probe (SKP) [9],
Electroanalysis 19, 2007, No. 2-3, 191 199

scanning droplet cell [10, 11], and scanning electrochemical


microscopy (SECM) [12, 13] were used for imaging local
surface properties with high lateral resolution. Previously,
SECM was successfully applied for simultaneous initiation
and visualization of corrosion pits [14, 15]. The visualization
of corroding sites was done using feedback mode SECM in
the presence of redox mediators [16 19] or substrategeneration-tip-collection mode SECM measuring the fluxes
of dissolved Fe2 ions [20]. Alternating current was used
before in SECM as a tool to control the tip-to-sample
distance in electrolyte solutions of high ionic strength [21,
22]. Recently, alternating current mode SECM (AC-SECM)
was applied for the visualization of local variations in
electrochemical surface activity [23 29]. Due to the fact
that AC-SECM is performed in the absence of any freediffusing redox mediator and an electrolyte solution of low
ionic strength the corrosion processes remain unimpaired.
Furthermore the low ionic strength of the used electrolyte
(1 mM NaCl) of the bulk solution is not expected to force
corrosion in the timescale of the experiment. However, ACSECM does not provide information on the nature of the
released ionic species. In order to gain a more detailed
insight in local changes of ion concentrations, methods
providing possibilities for a selective determination of the
concentration or fluxes of metal ions close to the sample
surface have to be integrated into SECM.
H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

192

D. Ruhlig, W. Schuhmann

Anodic stripping voltammetry (ASV) is a well known


technique for quantifying small concentrations of a variety
of different metal ions. Initially, stripping voltammetry was
applied in combination with SECM for the sensitive
determination of the release of a phenoxazine dye from a
carbon paste matrix after its accumulation on the surface of
a positioned SECM tip [30]. First approaches for combining
anodic stripping voltammetry and SECM have been shown
by Daniele and co-workers [31], who used a hemispherical
Hg microelectrode for detecting Pb2-ions released from
soil sediments in dependence from the tip-to-sample
distance. Janotta et al. showed the feasibility of detecting
the local release of Zn2-ions using anodic stripping
voltammetry, scanning over a scratch in a DLC coated
ZnSe surface [32]. Rudolph et al. demonstrated the detection of Mn2-ions in a generator collector mode, also using
ASV [33].
Here, we have chosen underpotential anodic stripping
voltammetry (UP-ASV) for the detection of the release of
Cu2-ions from surfaces. Underpotential anodic stripping
voltammetry exhibits a very low detection limit, short
accumulations times, and a comparatively low deposition
potential. Only very small amounts of Cu are deposited on
the electrode surface forming a monolayer in the absence of
any complexing agent [34, 35]. Therefore, corrosive dissolution at surfaces will not be induced by UP-ASV in direct
vicinity of the surface.
In this communication, the feasibility of combining ACSECM for the detection of active sites of enhanced electrochemical activity followed by local visualization of Cu2-ion
release from these active surface sites using stripping-mode
SECM (SM-SECM) will be demonstrated.

difficult. Thus, following previous ideas for the fabrication


of microelectrodes of different materials [36] Au microelectrodes were obtained by galvanic plating of Pt-microelectrodes fabricated following a previously published
procedure [37]. In a first attempt, Au plating was performed
directly on the polished Pt-disk from a commercial Au
plating solution (Aurocor K24 HF, Atotech, Feucht, Germany) applying a sequence of potential pulse to  1.1 V vs.
Ag/AgCl. However, during continuous potential cycling of
these tips in 1 M H2SO4 the deposited Au was lost. We
attributed this observation to an evolution of H2 at the
electrode tip during the deposition pulses, potentially
leading to an insufficient adhesion of the deposited Au
clusters on the Pt surface. To overcome this problem the Pt
surface was first covered with a Bi film, from a bismuth
plating solutions containing LiBr and Bi(NO3)3 in 1 M HCl
by applying a potential of  0.23 V for 10 s following a
procedure previously published by Krolicka et al. for coating of glassy carbon electrodes [38]. The Bi film on the one
hand increases the overvoltage for H2 formation and on the
other hand serves as adhesion layer for the deposited Au.
After washing off the Bi plating solution from the SECM tip,
an Au film was deposited on the Bi layer from an Aurocor
solution by applying 40 potential pulses to  1.1 V vs. Ag/
AgCl for 0.5 s followed up by a resting potential of 0 V vs.
Ag/AgCl for 1 s. The modified electrodes were rinsed with
water and dried in air.

2.3. Test Samples

All solutions used for Cu2-ion determination were prepared using ultra pure water (Ultrex II, J. T. Baker,
Deventer, Netherlands) and chemicals of analytical grade.
For evaluating the detection limit of the systems, we used a
1 M stock solution of CuSO4 3H2O (J. T. Baker, Deventer,
Netherlands). The coating of the electrode was carried out
in a bismuth plating solutions based on LiBr and Bi(NO3)3 in
1 M HCl (Aldrich, Steinheim, Germany) and a commercial
available Au plating solution (Aurocor K24 HF, Atotech,
Feucht, Germany). AC-SECM was performed in 1 mM
NaCl as electrolyte solution (J. T. Baker, Deventer, Netherlands) while SM-SECM was done in 10 mM HNO3 (Fluka,
Steinheim, Germany) containing 10 mM NaCl.

Test samples (provided by M. Koudelka, Institute of Microtechnology, University of Neuchatel, Switzerland) consisting of 36 individually addressable gold microdisk electrodes
with a diameter of 50 mm each were used. They were
fabricated on glass using lithographic techniques. Each
individual Au microdisk electrode is surrounded by a zone
of 100 mm in diameter which is not covered with the terminal
insulator (Si3N4). Thus, a small part of the connecting lead
and the microelectrode surface are exposed to solution.
Individual electrodes were coated with metallic Cu by
applying 20 potential pulses to  0.7 V vs. Ag/AgCl for 0.5 s
followed by a resting phase at 0 V vs. Ag/AgCl for 1 s in a
solution of 1 M CuSO4.
In addition, a metallic copper workpiece was coated with a
layer of nail varnish for corrosion protection. After drying,
small holes (about 100 200 mm diameter) were pierced into
the coating using a small syringe needle for generating
models of locally deteriorated surface coatings. The samples
were kept at open circuit potential (OCP) throughout the
SECM measurements.

2.2. Fabrication of Au-Covered SECM Tips

2.4. Instrumentation

UP-ASV of Cu is conventionally performed at Au electrodes. Due to the fact that Au wires show a poor adhesion to
glass the fabrication of glass insulated Au SECM tips is

AC-SECM was done in constant-height mode using a homebuilt SECM setup (details see [39]) equipped with three
computer-controlled bidirectional stepper motors with a

2. Experimental
2.1. Chemicals

Electroanalysis 19, 2007, No. 2-3, 191 199

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Spatial Imaging of Cu2-Ion Release

193

nominal resolution of 10 nm per microstep for positioning


the sample in x-, y-, z-direction underneath the fixed SECM
tip. A sinusoidal frequency perturbation was applied to the
SECM tip via the external potential input of a PGSTAT 12
potentiostat/galvanostat (Metrohm, Filderstadt, Germany)
using a Model 7265 DSP Lock-In Amplifier (Signal
Recovery, Wokingham, UK). The AC current response
from the potentiostat was fed into the lock-in amplifier
which provided the current magnitude R and the phase shift
q amplified with respect to the internal reference signal. All
experiments were carried out in a one-compartment cell
equipped with the 15 mm diameter Au-coated Pt microdisk
electrode as working electrode, a chlorinated silver wire
pseudo-reference electrode, and a Pt mesh (10  5 mm2) as
counter electrode.
For stripping-mode SECM (SM-SECM) the software of
the potentiostat (GPES 4.9) had to be synchronized with the
SECM software. This was achieved by connecting the digital
input/output port of the PGSTAT 12 with a digital input/
output port of the computer controlling the SECM for
sending and receiving 5 V trigger signals. A communication
module was integrated into the SECM software which was
used to stop and restart the SECM software and simultaneously starting the GPES 4.9 software of the PGSTAT 12
using trigger signals. The PGSTAT 12 potentiostat was set to
internal control via the GPES 4.9 software and an analytical
script was developed controlling all parameters of the UPASVand the square wave voltammetry (SWV) for stripping
off the accumulated Cu. A trigger signal is sent from the
SECM to the potentiostat after finishing a movement in xdirection for positioning the SECM tip at a predefined grid
point. This trigger initiates the start of the predefined
measuring script. After finishing the SWV and storing the
data the potentiostat sends a trigger signal to the SECM setup which continues with the scanning. Due to limitations in
the script size in the GPES software the overall measuring
cycle can be repeated up to 360 times.

2.5. Sequential AC- and SM-SECM Measurements


Sequential AC- and SM-SECM measurements were carried
out using the same Au-coated microelectrode without
changing the relative position of SECM-tip and sample
between the steps of the overall procedure: The Aumodified SECM tip was slowly approached to the sample
surface in AC-mode until the observed signal (current
magnitude or phase shift) stopped decreasing indicating
contact between tip and surface. The tip then was retracted
from the surface to a predefined distance of generally 3 mm.
The AC measurements were carried out in 1 mM NaCl at a
frequency of 10 to 30 kHz and an amplitude of 200 mV rms.
Scanning was done with a speed of 6.4 mm s1 between the
grid points. After recording the AC-SECM image the tip was
repositioned to a x-, y-position close to a site of increased
local activity as derived from the AC-SECM image. The
1 mM NaCl solution was removed from the electrochemical
cell using a specially bent syringe needle and the cell was
Electroanalysis 19, 2007, No. 2-3, 191 199

rinsed repeatedly with a solution containing 10 mM NaCl in


10 mM HNO3 carefully avoiding any physical contact to the
sample, the measuring cell, or the SECM tip. Finally, a
solution of 10 mM NaCl in 10 mM HNO3 was added to the
cell and the SM-SECM measurement was started.

3. Results and Discussion


Visualization of local corrosion using either feedback mode
SECM or generator collector experiments is difficult due to
the active potential modification of the surface, reactions
with the redox mediator or the applied potential. As a
matter of fact, it would be advantageous to investigate
sample surfaces without perturbing the corrosion processes
themselves. However, this is only possible if the measurements can be performed without the application of an
external potential to the sample and additionally the
presence of any redox couples in the electrolyte can be
avoided. As pointed out above, AC-SECM in low ionic
strength electrolytes allows visualizing local electrochemical activity on surfaces. However, it fails to reveal the nature
of the released ions. On the other hand, stripping voltammetry integrated with SECM should be a suitable tool for
the elucidation of the nature and the local fluxes of specific
metal ions.
For the quantification of Cu2-ions underpotential anodic
stripping voltammetry (UP-ASV) was chosen. In order to
sequentially perform AC-SECM to localize active surface
sites, followed by stripping-mode SECM (SM-SECM) for
the visualization of variations in local Cu2-ion concentrations, a suitable microelectrode has to be developed. For the
detection of Cu2-ions using UP-ASV Au working electrodes are appropriate. Due to the fact, that Au is not forming a
stable surface oxide, it is well known that glass coated Au
wires often suffer from insufficient stability. To circumvent
this phenomenon we suggest using Pt microelectrodes as
basis for the preparation of Au microelectrodes. Direct
galvanic plating of the Pt surface of a 15 mm diameter
polished Pt-disk electrode sealed in glass using a commercial
Au plating solution, did not lead to stable Au films and
repeated potential sweeps clearly led to a loss of the Au
coating (data not shown). Due to the fact, that the Au was
deposited during a sequence of potentiostatic pulses to
potentials as low as  1.1 V vs. Ag/AgCl we assumed that the
formation of molecular H2 most likely causes the poor
adhesion of the Au coating. To overcome this problem, a
first layer of Bi was deposited on the Pt surface to increase
the H2 overpotential. Subsequent reductive deposition of
Au led to an Au-coated Pt microelectrode. Cyclic voltammetry in 1 M H2SO4 reveals a mixed surface of Au and Pt
indicated by the superimposed redox waves typical for Pt
and Au (Figure 1a). Control experiments performed with a
Bi-modified Pt microelectrode clearly show the quantitative
stripping of the Bi film in a potential range between 0 and
0.15 V (data not shown). Thus, it can be concluded that at the
applied positive potentials no Bi remains exposed at the
electrode surface. The AFM image of the Au-coated Pt-Bi-

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

194

D. Ruhlig, W. Schuhmann

Fig. 1. A) Cyclic voltammograms (CV) of a 250 mm diameter Au-coated Pt-Bi disk electrode sealed in glass. (1 M H2SO4; potential
range from 0 to 1.4 V vs. Ag/AgCl reference electrode; scan rate of 100 mV s1). The first scan (dotted lines), the 10th scan (dashed lines)
and the 20th scan (solid line) are shown. B) Contact-mode AFM image of an area of 50  50 mm2 of a 250 mm Pt disk microelectrode
surface before (left) and after (right) deposition of a Bi layer covered subsequently with Au. (scan rate of 33.3 mm s1). The average
surface roughness increased from 32 nm for bare Pt to 826 nm for the Au-modified Pt-Bi surface.

surface (Figure 1b left image) suggests a homogeneous


covering of the Pt surface and a highly increased surface
roughness. The significantly increased accessible surface
area during the Bi- and Au-plating should allow to perform
UP-ASV of Cu with high sensitivity.
In order to demonstrate the suitability of the developed
Au-coated Pt-Bi microelectrodes for UP-ASV of Cu, UPASV was performed in bulk solution with the electrode far
away from any surface. UP-ASV was carried out using
square wave voltammetry (SWV) at different Cu2-ion
concentrations and accumulation times. At an accumulation
potential of  0.4 V vs. a Ag/AgCl pseudo reference
electrode and using accumulation times of 15 s and 45 s,
respectively, followed by SWV from 0 to 0.6 V (scan rate
25 mV s1, frequency 25 Hz, step potential 2.5 mV, amplitude 10 mV) well defined stripping peaks for the dissolution
of accumulated Cu were obtained (Figure 2a). After each
UP-ASV, a potential of 0.6 V vs. Ag/AgCl pseudo reference
electrode was applied to the Au-coated microelectrode for
10 to 35 s to assure the complete removal of any remaining
Electroanalysis 19, 2007, No. 2-3, 191 199

adsorbed Cu from the electrode surface. The peak current


obtained in the background electrolyte (0 nM Cu2; see
Figure 2a) may be caused by a contamination of the
electrolyte with a small but constant Cu2-ion concentration
leading to a constant current offset. Since the aim of the
study was to visualize local differences in Cu2-ion concentrations a constant current offset is not relevant. Related
calibration graphs (Figure 2b) for accumulation times of
15 s and 45 s are demonstrating the linear relation between
peak current and Cu2-ion concentration and the high
sensitivity of this method.
At rather short accumulation times of only 15 s a
detection limit of 0.17 nM (N 4, RSD 7%) was derived.
Increasing the accumulation time to 45 s yielded significantly higher peak currents and a detection limit of 0.15 nM
(N 4, RSD 6%). However, although the calibration
characteristics should in principle allow quantifying the
released amount of Cu2-ions in the planned SM-SECM
experiments, the close distance between tip and sample in
unstirred solution may severely modulate the diffusional

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Spatial Imaging of Cu2-Ion Release

195

Fig. 2. A) Square wave voltammograms (SWVs) obtained using


UP-ASV for quantifying Cu2-ions concentrations at 15 mm
diameter Au-coated Pt-Bi disk microelectrodes (10 mM HNO3
containing 10 mM NaCl; accumulation time 45 s at a potential of
 0.4 V vs. Ag/AgCl). SWV was performed by applying a
potential scan from 0 to 0.6 V vs. Ag/AgCl at a scan rate of
25 mV s1, a frequency of 25 Hz, a step potential of 2.5 mV, and an
amplitude of 10 mV. B) Calibration graphs derived from the peak
currents (Ipeak) of the SWV at increasing Cu2 concentrations. The
detection limits are 0.17 nM (R 0.976) for an accumulation time
of 15 s (triangles) and 0.15 nM (R 0.988) for an accumulation
time of 45 s (squares).

mass transport of the Cu2-ions to the SECM tip. In contrast


to the bulk experiment, the Au-coated microelectrode will
be positioned in close distance from the sample forming a
small thin-layer electrochemical cell. Any Cu2-ions released from the surface into this small gap are assumed to
diffuse preferentially towards the SECM tip to be reductively accumulated there. Thus, a diffusion gradient is
formed in direction to the SECM tip surface leading to an
increased probability for a released Cu2-ion to reach the tip
rather than to escape from the thin-layer cell formed by the
sample and the tip. Thus, the situation severely differs from
the situation in bulk and the peak currents are rather
reflecting a rate of Cu2-ion release from the surface than a
concentration of Cu2-ions. Thus, we did not convert the
peak currents obtained during SECM scanning to Cu2-ion
concentrations using the calibration graphs obtained in bulk
solution.
SM-SECM is slow since an entire UP-ASV cycle is
performed at each grid point. Thus, our strategy was first to
Electroanalysis 19, 2007, No. 2-3, 191 199

find possible sites of local high electrochemical activity


using AC-SECM and subsequently use SM-SECM for
visualizing local Cu2-ion release from these sites. This
allows restricting the scan area of the SM-SECM experiments to already identified sites of potential interest. As a
matter of fact, an overall sequence of AC-SECM visualization of sites of increased electrochemical activity followed by SM-SECM imaging has to be performed with the
same SECM tip and without changing the relative position
of tip and sample. Thus, it was necessary to evaluate if the
Au-coated Pt-Bi microelectrode is suitable for AC-SECM
imaging.
A test structure consisting of an array of 36 individually
addressable Au microelectrodes of 50 mm diameter was used
(see inset Figure 3 for a microscopic picture of a single
microelectrode). The electrodes were electrically connected
by Au lines to contact pads at the outer side of the glass chip.
By means of a lithographically structured insulating top
layer only the Au electrodes and a short part of the contact
line were exposed to the electrolyte solution. One of the
electrodes was modified with metallic Cu by reductive
deposition from an aqueous Cu2 solution yielding brittle
Cu-deposits on the microelectrode surface (see inset Figure 3 for a microscopic picture of a modified electrode).
After the deposition step, the insulating layer on top of the
associated gold contact line was partially removed by careful
polishing under microscopic control in order to expose a
conductive, however, non-modified Au area to the electrolyte solution. In doing so, it was expected to see a clear
contrast in the AC-SECM image between conducting,
insulating, and actively corroding surface areas. An AC
image using the Au-coated Pt-Bi microelectrode as SECM
tip was performed in an area of 650  300 mm2 at 40 kHz and
200 mV rms vs. Ag/AgCl pseudo-reference with a grid
increment for scanning of 15 mm and a tip to sample distance
of about 3 mm. The AC-SECM image displayed in Figure 3
clearly reveals a high-contrast in the current magnitude R,
allowing for visualizing the local electrochemical activity of
the Cu-modified microelectrode. In addition, the noninsulated part of the contact line which was not modified
with Cu can still be clearly distinguished from the insulated
areas. Obviously, the developed Au-modified Pt-Bi microelectrodes can be successfully employed for discriminating
between corroding, conducting and insulating areas using
the AC mode of SECM.
After imaging the sample surface using AC-SECM, the
SECM tip was kept at the same distance to the sample
surface while the electrolyte was replaced by a solution of
10 mM HNO3 containing 10 mM NaCl. From the ACSECM image a suitable scanning area can be derived which
was subsequently scanned using the same grid increment as
in the AC-SECM image. At each pre-defined grid point, an
UP-ASV was performed for determination of locally
released Cu2-ions. The potential was set to  0.4 V during
the accumulation time of 15 s, followed by SWV in the
potential range from 0 to 0.6 V, a scan rate of 25 mV s1, a
frequency of 25 Hz, a step potential of 2.5 mV, and an
amplitude of 10 mV. For excluding any effects caused by

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

196

D. Ruhlig, W. Schuhmann

Fig. 3. AC-SECM visualization of a Cu-coated 50 mm gold disk microelectrode (see inset for microscopic pictures before and after
coating). The AC current magnitude R is plotted against the x-, y-position of the SECM tip. Higher R values represent the Cu-coated
microelectrode surface and the non-isolated Au contact line. The large contrast is caused by changes in the local electrochemical activity
due to local release of Cu2-ions. (40 kHz; 200 mV rms; in 1 mM NaCl; scan rate 3.2 mm s1).

increasing background concentration of Cu2-ions or


changes of the SECM tip properties, a second SWV was
recorded with the same parameters immediately after the
first one, however, at 0 s accumulation time. As proposed
previously [26, 40], this second voltammogram was subtracted from the first one for background correction and for
improving the detection limit. The overall scan area was
215  160 mm2 converting to 153 SWV from which the peak
currents were extracted and plotted over the x-, y-coordinates of the SECM tip position. By this, an image
representing the local Cu2-ion concentration over the
scanned area was obtained (Figure 4a).
A central area of high Cu2-ion concentrations corresponds well with the diameter of the Cu-modified Au
microelectrode, which was obviously actively corroding
although being unbiased. As expected, the local Cu2-ion
concentration decreases with increasing distance from the
Cu-modified area. Due to diffusion, the Cu2-ions released
from the structure a slow continuous increase of the
background concentration is obtained leading to a slightly
increasing background current in the course of the experiment. The sequence of SWValong one x-line scan of the SMSECM image shown in Figure 4a is plotted in Figure 4b and
the related peak currents are displayed over the corresponding x-position as inset. Nevertheless, even after
several hours of scanning the contrast between the surrounding and the corroding spot remained high.
After demonstrating the principle feasibility of sequentially combining AC-SECM and SM-SECM a more complex
test sample was investigated. This test sample was fabricated
by coating a copper metal block with nail varnish. The
protected surface was pinched with a needle to locally create
small holes with dimensions of 100 200 mm diameter down
Electroanalysis 19, 2007, No. 2-3, 191 199

to the Cu surface. A microscopic picture of the test surface is


shown in Figure 5c. The white circles indicate the area with
the deliberately pinched holes in the varnish. This sample
was investigated by AC-SECM to localize spots of increased
electrochemical activity and subsequently using SM-SECM
for revealing localized increased Cu2-ion concentrations.
The obtained images are shown in Figure 5a and 5b. Due to
the size of the pinched holes of 100 200 mm and the
expected significantly smaller Cu2-ion concentrations, the
distance between the grid points was increased to 30 mm and
the accumulation time in UP-ASV was prolonged to 45 s.
The AC-SECM image (Figure 5a) shows the local change
of the phase shift q. Smaller values of q are converted to
higher brightness representing sites with increased local
electrochemical activity. The SM-SECM image (Figure 5b)
was obtained after exchange of the electrolyte solution from
only a part of the area scanned in the AC-SECM mode. As
before, the image was constructed from the peak current of
the SWV as z-coordinate over the x-, y-position of the
SECM tip. The white circles represent the areas of the
pinched holes and correspond with those in the optical
image. The AC-SECM image clearly reveals a number of
bright areas representing either increased electrochemical
activity or variations in the tip-to-sample distance during
constant-height scanning. Nevertheless, at the coordinates
of the expected pinholes the AC-SECM image unequivocally shows increased electrochemical activity. Due to the
increased grid size of 30 mm, the resolution is reduced
causing the rather sharp and square-like representations of
the active areas.
The combination of AC-SECM and SM-SECM should
help to clarify, if the bright areas in Figure 5a are indeed
caused by locally increased electrochemical activity. As the

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Spatial Imaging of Cu2-Ion Release

197

Fig. 4. A) Visualization of local Cu2-ion concentrations over a 50 mm Cu-coated gold disk microelectrode. The image is obtained by
plotting the peak current (Ipeak) of local UP-ASV against the x-, y-position of the 15 mm Au-coated SECM tip. The Cu-deposits are
clearly resolved. (10 mM HNO3 containing 10 mM NaCl; accumulation potential of  0.4 V vs. Ag/AgCl for 15 s, followed by a SWV
from 0 to 0.6 V at a scan rate of 25 mV s1, a frequency of 25 Hz, a step potential of 2.5 mV, and an amplitude of 10 mV). B) SWVs
obtained during one SM-SECM line scan in x-direction over the structure shown in Figure 3. The peak currents of the SWVs are plotted
against the scan direction (see inset) for each line, leading to a three dimensional visualization of a corroding spot as shown in Figure 4a.

pinched holes are due to the manual fabrication of varying


sizes and diameters, the uncoated areas at the bottom of the
holes are of different sizes. Moreover, assuming that the
coating is damaged by the pinching needle but the underlying Cu surface is still completely insulated, the AC-SECM
picture may already visualize a change in local electrochemical activity due to variations in the resistivity while the
SM-SECM will not be able to detect locally increased Cu2ion concentrations. Obviously, if in both the AC-SECM and
the SM-SECM images an overlaying local increased activity
and Cu2 concentration is detected a surface damage down
to the metal surface must have occurred. This is clearly seen
for all three pinched holes.
Electroanalysis 19, 2007, No. 2-3, 191 199

4. Conclusions
The feasibility of integrating the UP-ASV into the SECM
(SM-SECM) has been successfully shown. Using a specifically developed Au-coated Pt-Bi microelectrode as SECM
tip local release of Cu2-ions can be detected in the low nM
concentration range using UP-ASV. AC-SECM as a tool for
localizing surface areas with increased electrochemical
activity without disturbing the surface processes in combination with subsequently performed SM-SECM allows
visualizing active surface sites and local Cu2-ion concentrations. This study is seen as a demonstration of the
feasibility and will be the starting point for more detailed
investigations of the corrosion phenomena of relevant

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

198

D. Ruhlig, W. Schuhmann

Fig. 5. Visualization of a lacquered copper surface with artificially created pinholes. A) AC-SECM image. B) SM-SECM image. The
AC image was obtained at a frequency of 20 kHz and 200 mV rms amplitude by plotting the local differences in the phase shift (q)
against the x-, y-position of the SECM tip. Bright areas correspond to more negative q values and represent increased electrochemical
activity. In the SM-SECM image the brighter areas correspond to higher peak currents in the SWV and hence represent increased Cu2ion concentrations. The white circles represent the locations at which the lacquered surface was pinched with a needle. C) shows an
optical micrograph of the investigated sample surface.

samples, such as ternary copper containing shape memory


alloys (NiTiCu). Of special interest are the effects of
pretreatment and microscopic inclusions on the stability of
the passive films and the concomitant ion release especially
with respect to the biocompatibility of these materials.

5. Acknowledgements
The authors gratefully acknowledge the financial support of
this work by the Deutsche Forschungsgemeinschaft (DFG)
in the framework of the special research program Formgedchtnistechnik (SFB 459-A5). The authors are grateful to Andrea Puschhof for her contribution to this study in
the framework of her Bachelor work.

6. References
[1] N. B. Morgan, Mater. Sci. Eng. A 2004, 378, 16.
[2] C. D. J. Barras, K. A. Myers, Eur. J. Vasc. Endovasc. Surg.
2000, 19, 564.
[3] N. Schiff, B. Grosgogeat, M. Lissac, F. Dalard, Biomaterials
2004, 25, 4535.
[4] G. Riepe, C. Heintz, E. Kaiser, Eur. J. Vasc. Endovasc. Surg.
2002, 24, 117.
[5] G. Heintz, G. Riepe, L. Birken, J. Endovasc. Ther. 2001, 8,
248.
[6] S. A. Shabalovskaya, Biomed. Mat. Eng. 2002, 12, 69.
[7] F. Zhou, D. Thierry, H. S. Isaac, J. Electrochem. Soc. 1997,
144, 1957.
[8] J. W. H. de Wit, D. H. van der Weijde, A. de Jong, F.
Blekkenhorst, S. D. Meijers, Electrochemical Methods in
Corrosion Research VI (Eds: P. L. Bonora, F. Deflorian),
Trans Tech Publications, ZSrich 1998, 28 92, pp. 69.
Electroanalysis 19, 2007, No. 2-3, 191 199

[9] M. Stratmann, R. Feser, A. Leng, Electrochim. Acta 1994, 39,


1207.
[10] T. Suter, H. Bhni, Electrochim. Acta. 1997, 42, 3275.
[11] M. M. Lohrengel, A. Mhrig, M. Pilaski, Electrochim. Acta.
2001, 47, 137.
[12] Scanning Electrochemical Microscopy (Eds: A. J. Bard, M. V.
Mirkin), Marcel Dekker, New York 2001.
[13] M. V. Mirkin, B. R. Horrocks, Anal. Chim. Acta 2000, 2,
119.
[14] D. O. Wipf, Colloids Surf. A 1994, 93, 251.
[15] J. W. Still, D. O. Wipf, J. Electrochem. Soc. 1997, 144, 2657.
[16] N. Casillas, S. Charlebois, W. H. Smyrl, H. S. White, J.
Electrochem. Soc. 1993, 140, L142.
[17] C. H. Paik, H. S. White, R. C. Alkire, J. Electrochem. Soc.
2000, 147, 4120.
[18] J. C. Seegmiller, D. A. Buttry, J. Electrochem. Soc. 2003, 150,
B413.
[19] K. Mansikkamaeki, P. Ahonen, G. Fabricius, L. Murtomaki,
K. Kontturi, J. Electrochem. Soc. 2005, 152, B12.
[20] E. Vlker, C. Gonzalez Inchauspe, E. J. Calvo, Electrochem.
Commun. 2006, 8, 179.
[21] B. R. Horrocks, D. Schmidtke, A. Heller, A. J. Bard, Anal.
Chem. 1993, 65, 3605.
[22] M. A. Alpucho-Aviles, D. O. Wipf, Anal. Chem. 2001, 73,
4873.
[23] B. Ballesteros Katemann, A. Schulte, E. J. Calvo, M.
Koudelka-Hep, W. Schuhmann, Electrochem. Commun.
2002, 4, 134.
[24] B. Ballesteros Katemann, C. Gonzales Inchauspe, P. A.
Castro, A. Schulte, E. J. Calvo, W. Schuhmann. Electrochim.
Acta 2003, 48, 1115.
[25] B. R. Horrocks, in Encyclopedia of Electrochemistry, Vol. 3
(Eds: A. J. Bard, M. Stratmann), Wiley-VCH, Weinheim
2003, p. 444.
[26] M. Etienne, A. Schulte, W. Schuhmann, Electrochem.
Commun. 2004, 6, 288.
[27] A. Schulte, S. Belger, M. Etienne, W. Schuhmann, Mat. Sci.
Eng. A 2004, 378, 523.

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Spatial Imaging of Cu2-Ion Release

199

[28] A. S. Baranski, P. M. Diakowski, J. Solid State Electrochem.


2004, 8, 683.
[29] E. N. Ervin, H. S. White, L. A. Baker, C. R. Martin, Anal.
Chem. 2006, 78, 6535.
[30] F. D. Munteanu, M. Mosbach, A. Schulte, W. Schuhmann, L.
Gorton, Electroanalysis 2002, 14, 1479.
[31] S. Daniele, I. Ciani, C. Bragato, M. A. Baldo, J. Phys. IV
France 2003, 107, 353.
[32] M. Janotta, D. Rudolph, A. Kueng, C. Kranz, H.-S. Voraberger, W. Waldhauser, B. Mizaikoff, Langmuir 2004, 20,
8634.
[33] D. Rudolph, S. Neuhuber, C. Kranz, M. Taillefer, B. Mizaikoff, Analyst 2004, 129, 443.

Electroanalysis 19, 2007, No. 2-3, 191 199

[34] M. Brand, I. Eshkenazi, E. Kirowa-Eisner, Anal. Chem. 1997,


69, 4660.
[35] G. Herzog, D. W.M. Arrigan, Trends Anal. Chem. 2005, 3,
208.
[36] M. Etienne, J. Oni, A. Schulte, G. Hartwich, W. Schuhmann,
Electrochim. Acta 2005, 50, 5001.
[37] C. Kranz, M. Ludwig, H. E. Gaub, W.Schuhmann, Adv.
Mater. 1995, 7, 38.
[38] A. Krolicka, A. Borowski, Electrochem. Comm. 2004, 6, 99.
[39] A. Hengstenberg, C. Kranz, W. Schuhmann, Chem. Eur. J.
2000, 6, 1547.
[40] Y. Bonfil, M. Brand, E. Kirowa-Eisner, Anal. Chim. Acta.
2002, 464, 99.

www.electroanalysis.wiley-vch.de

H 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Das könnte Ihnen auch gefallen