You are on page 1of 11

12362 J. Phys. Chem.

C 2009, 113, 12362–12372

Comprehensive Study of the Application of a Pb Underpotential Deposition-Assisted


Method for Surface Area Measurement of Metallic Nanoporous Materials

Y. Liu,† S. Bliznakov,‡ and N. Dimitrov*,†,‡


Materials Science and Engineering Program, State UniVersity of New York at Binghamton,
P.O. Box 6000, Binghamton, New York 13902-6000, and Department of Chemistry, State UniVersity of New
York at Binghamton, P.O. Box 6000, Binghamton, New York 13902-6000
ReceiVed: February 19, 2009; ReVised Manuscript ReceiVed: May 15, 2009

An inexpensive, fast, selective, and sensitive technique for surface area measurement of metallic nanoporous
materials (MNPM) is developed, systematically tested, and validated. The approach employed is based on
underpotential deposition (UPD) of metals on foreign substrates. In this work, Pb UPD on Au is chosen to
illustrate the applicability of and reveal the advantages and limitations of the proposed method. Experiments
are designed for surface area measurement of nanoporous gold (NPG) electrodes with pore sizes in the range
of 5-15 nm, prepared by electrochemical dealloying of single phase AupAg1-p (atomic fraction p ) 0.1, 0.2,
and 0.3). Dealloying is performed galvanostatically at a current density of 1 mA cm-2 in a AgClO4 solution,
acidified to pH 1. The experimental results suggest a linearly increasing charge in the Pb UPD layer with
NPG thickness. This finding hints at (i) uniformity of the NPG structure and (ii) the general ability of this
method to work for analysis of bulk materials. The proposed approach is tested by studying the dependence
of the NPG surface area upon the original alloy composition and correlating the results with the NPG structure
and morphology imaged by high-resolution scanning electron microscopy. An anomalously high surface area
is registered in dealloyed Au0.1Ag0.9 samples and is attributed to the lack of a pre-existing percolation backbone.
Unlike the instantaneous Pb UPD process on a flat metal surface, the slow and thickness-dependent kinetics
of Pb layer formation on NPG is associated with hindered mass transport through pores. Further validation
of the Pb UPD method is made by experimental monitoring of heat treatment-enforced coarsening and the
basic modeling of the correlation between surface area and ligament size in NPG. Finally, a critical comparison
with Brunauer-Emmett-Teller (BET) analysis reveals important advantages of the developed method for
surface area measurement in MNPM specimens.

1. Introduction for surface area measurement of MNPM would not only enable
efficient surface area measurement but would also help to better
Metallic nanoporous materials (MNPM) can be described as understand the kinetics of the structural evolution in MNPM
metals having a three-dimensional (3D) interconnected solid-void
and to assess the factors and limitations associated with mass
structure in nanoscale. To date, besides nanoporous Au (NPG)
transport through interconnected nanosized channels.
that can be easily prepared by electrochemical dealloying
(selective Ag dissolution) of single phase AupAg1-p alloys,1-4 The first choice for the surface area determination of a fine
successfully synthesized MNPM include nanoporous Pt,5-7 Cu,8,9 structure like MNPM would be today’s area characterization
Ag,9-11 Pd,12,13 and Ni.14,15 The substantial surface area develop- “work horse”, the Brunauer, Emmett and Teller (BET) method.
ment in MNPM manifested by a several orders of magnitude BET is based on a physical absorption of gas molecules onto
increase versus the geometric area of the bulk material renders the substrate under characterization. The area can be obtained
these structures attractive for a variety of upcoming technologies on the basis of a pressure measurement of the physically
associated with catalysis,16-18 fuel cells,19,20 membranes,21 adsorbed gas isotherms in the chamber.25 It has been widely
actuators,22 and electrochemical sensing.23,24 At present, the used to determine the surface area of hydrogen storage materi-
characterization of nanoporous metals relies mostly on field als,26 nanoparticle assemblies,27 nanotubes,28 nanorods,29 and
emission scanning electron microscope (FESEM) and/or trans- nanoporous materials.30 While BET is deemed accurate and
mission electron microscope (TEM), when applicable.1-23 When reliable for many of the above listed systems, there are
the determination of surface area is of interest, FESEM and TEM limitations and shortcomings associated with its efficiency for
could provide only a qualitative picture based on measurements characterization of MNPM. For instance, a large amount of
of average pore and ligament sizes and the assumption of a sample is needed (i.e., more than one gram of NPG) for the
structural isotropy of the substrate of interest. However, when BET surface area measurement because the pressure measure-
a quantitative result is needed, the microscopy techniques fail ment is not sensitive enough. In addition to that, an elevated
to accurately provide details about the actual surface area. Thus, temperature treatment ranging between 100 and 350 °C,
an inexpensive, fast, accurate, selective, and sensitive technique recommended generally as an effective way for water/solvent
and contaminant removal, could substantially enhance the
* To whom correspondence should be addressed. E-mail: dimitrov@ diffusivity of surface atoms and in turn trigger a massive
binghamton.edu. Telephone: +1 607 777 4271.

Materials Science and Engineering Program. coarsening of the sample2,21,31,32 due to the Gibbs-Thomson

Department of Chemistry. effect.33 This would ultimately result in a significant decrease
10.1021/jp901536f CCC: $40.75  2009 American Chemical Society
Published on Web 06/19/2009
Pb UPD-Assisted Method for Surface Area Measurement J. Phys. Chem. C, Vol. 113, No. 28, 2009 12363

in the surface area of MNPM consisting originally of very fine has been increasingly used for surface modification47,48 and thin
ligaments and pores. film growth49,50 on a variety of important substrates would
A way to address the above limitations is to employ undoubtedly be affected by limitations considered earlier in this
electrochemical techniques, which can analyze nanosized layers section. Thus, the work reported herein is aimed at (i) revealing
and do not require heat treatment of the specimens. The double- key details and limitations associated with the mechanism and
layer capacitance and charge of the underpotentially deposited kinetics of UPD formation in architectures featuring intercon-
(UPD) monolayer of metal on foreign substrates are two nected porosity metal structures and (ii) developing an inex-
electrochemical characteristics that are directly related to the pensive, fast, accurate, selective, and sensitive technique for
roughness of the electrode and can thereby be used for sur- surface area measurements of MNPM.
face area measurements of MNPM. The double-layer capaci- In the present paper, Pb UPD on Au as a well understood
tance could be derived by the analysis of electrochemical system39,40 is chosen to illustrate the applicability of and reveal
impedance spectroscopy (EIS). The ratio between the double- the advantages and limitations of the proposed electrochemical
layer capacitance measured on porous and flat electrodes from method. Additional justification for the choice of Pb is provided
the same metal correlates directly with the surface area by its ability to form a UPD layer on a variety of other noble
development. EIS is a powerful technique that also provides metals (Ag,49 Cu,50 Pt,51 and Rh52) that could be processed as
additional information on the shape of the pores,34-38 but as nanoporous substrates. Last but not least, Pb UPD would still
Cattarin et al.34 reported, the recording of impedance spectra is be confined within the limits of a monolayer, even if surface
complicated with NPG samples. In addition, the EI spectra alloying were to take place as no UPD system with Pb
analysis is a complex procedure and is often dependent on participation has been shown to feature mixing that goes beyond
critical assumptions (choice of right model, proper fitting a site exchange between the UPD layer and topmost layer of
procedures, and sensitive software). Unlike EIS, the UPD the substrate.39 Experiments are designed for surface area
measurement is independent of the shape and uniformity of the measurement of NPG electrodes prepared by electrochemical
pores, and the charge corresponding to the UPD monolayer dealloying (selective Ag dissolution) of single phase AupAg1-p,
formation or stripping could be easily determined by standard where the atomic fraction p ) 0.1, 0.2, and 0.3. These feature
electrochemical measurements such as chronoamperometry (CA) stable pore and ligament sizes in the range of 5-15 nm and a
and cyclic voltammetry (CV). Thus for MNPM, the UPD controlled dealloying depth to 5 µm. Instead of cyclic voltam-
measurement appears to be a more suitable and straightforward metry (CV) conventionally used in similar cases,39,40 in this work
method than EIS for surface area assessment. UPD has been UPD measurements were carried out using chronoamperometry
widely used to characterize the surface area and/or surface (CA). First, the surface area development is determined by
roughness of generally flat single and polycrystalline metal measuring the roughness factor, Rf, as a function of the alloy
surfaces.39,40 Most recently, UPD-based protocols have also been composition and dealloying depth. Second, important aspects
of the UPD mechanism and kinetics are discussed in view of
employed in the study of MNPM substrates.7,10,24 The UPD
the duration of UPD at different dealloying depths, the difference
phenomenon is strictly limited to the formation of one mono-
between CA and CV measurements, and the impact of Pb2+
layer and rarely two complete monolayers perfectly reproducing
ion concentration on the formation of a UPD layer. Finally,
the morphology of the underlying substrate.39,40 This way, the
sample coarsening caused by heat treatment, surface area
charge of a UPD monolayer is directly proportional to the
modeling, and BET measurements are conducted to ascertain
substrate surface area. Therefore, measurement of the roughness
and validate the applicability of the UPD-assisted method for
factor, Rf, obtained as a ratio of accurately measured UPD layer
surface area measurement in MNPM.
charges on developed and flat metal surfaces, would guarantee
an accurate determination of the surface area development. 2. Experimental Section
Applications of UPD as a characterization tool emphasize the 2.1. AupAg1-p Alloy Synthesis. Single phase AupAg1-p
assessment of surface activity of electrochemical sensor alloys (p ) 0.1, 0.2, and 0.3) were synthesized by high-
substrates41,42 and the surface area measurement of catalytically temperature melting in inert atmosphere and were characterized
active substrates.43-46 In the second group of applications, by X-ray diffraction (Philips X’Pert MRD system with Cu KR
hydrogen UPD on carbon-supported Pt has been used for surface radiation) and energy dispersive X-ray spectroscopy (EDXS)
area assessment of fuel cell catalysts.43-45 Adding to that some (Zeiss Supra 55 VP system with EDAX detector) to confirm
random attempts for surface area characterization of electrodes the mixing and alloy composition. Precalculated amounts of Au
including MNPM,7,10,24 one finds no systematic and compre- and Ag (both 99.99%, Surepure Chemetals) were placed in a
hensive study on the kinetics and limitations of the UPD process graphite crucible inside a quartz reactor, deoxygenated by ultra
in MNPM published so far in the literature. Another missing high purity (UHP) N2 gas (less than 1 ppb oxygen, moisture
piece of information relates to the mechanism of UPD formation content, CO, and CO2) for at least 2 h prior to the melting.
on MNPM. Slow UPD formation onto MNPM substrates lasting After 20 min melting at 1200 °C, the samples were slowly
for a couple hundred seconds24 has been reported in contrast cooled to room temperature, rolled into 50 µm thick foil, and
with only a few seconds on a metal surface.39,40 It has been cut into small pieces. This procedure was repeated two more
speculated that the UPD might not be under pure kinetic times to ensure maximum homogeneity and best mixing. Finally,
control.5,24 If limitations are contributed to by mass transport the alloys were rolled into a 3 mm thick sheet, and cylindrical
(diffusion) control, it is not clear whether these limitations are samples with a diameter of 6 mm were punched out by a “punch
due to a narrow pore size and/or to a low UPD metal ion and die” press (Roper Whitney of Rockford, Inc.). These
concentration. samples were then annealed in inert atmosphere at 800 °C for
Understanding the UPD process on MNPM is fundamentally 20 h. Finally, linear scanning voltammetry (LSV) at a scan rate
and technologically important. For instance, the UPD process of 1 mV s-1 (NPG Preparation) was conducted on each alloy
on MNPM could serve as a model for the faradaic process sample in an acidic solution in a three-electrode configuration
operating in typical catalytic and/or sensing scenarios. Also, to confirm their composition by measuring the dealloying critical
surface limited redox replacement (SLRR) of UPD layers that potential.
12364 J. Phys. Chem. C, Vol. 113, No. 28, 2009 Liu et al.

2.2. NPG Preparation. For processing of NPG samples the


above-described alloys were first mechanically polished on
alumina slurry (Buehler) to 0.05 µm. The samples were then
rinsed with Barnstead Nanopure water (R > 18.2 Ω cm),
subjected to annealing at 800 °C in N2 atmosphere, and cooled
to room temperature. The dealloying to selectively remove Ag
from the alloys was carried out in a 1 × 10-3 M AgClO4
(99.995+% metal basis, Sigma-Aldrich) and 1 × 10-1 M HClO4
(double distilled, GFS Chemicals) solution at a constant current
density i ) 1 mA cm-2. The selective dissolution was conducted
in a three-electrode cell controlled by a PC-based potentiostat/
galvanostat (EC Epsilon, BASi). Ag wires pretreated by etching
in 8 M HNO3 were used for counter and reference electrodes.
The time, t, required for the penetration of the dissolution front Figure 1. Current density response of AupAg1-p alloys to an anodic
at a certain depth, d, was calculated by eq 1 represented by the potential scan (sweep rate of 1 mV s-1) using LSV. Inset: Potential
Faraday law for electrolysis, including the atomic fraction (also transient of AupAg1-p alloys with a dissolution depth of 1 µm at a
volume fraction in the case of AupAg(1-p)) of Ag, (1-p) to reflect current density of 1 mA cm-2 using chronopotentiometry (CP).
the complex nature of the dissolving sample (alloy).
lens detector at an accelerating voltage of 10 kV and a working
(1 - p)zdFAgF distance of 2 mm was used to characterize the NPG samples
t) (1) after dealloying. Also, BET measurements (Micromeritics ASAP
iaAg
2020) were conducted to validate the developed UPD-assisted
method and to assess the sensitivity of both approaches to heat-
Here, z is the number of electrons transferred for the oxidation treated samples. Specific details concerning each of these
of one Ag atom, F is the Faraday’s constant, FAg is the mass experimental procedures are described in Results and Discus-
density of Ag and aAg is its atomic weight. The analysis of eq sion.
1 suggests that after dissolving, for instance, 90% of the volume
occupied by Ag atoms in a layer of Au0.1Ag0.9 with thickness 3. Results and Discussion
d, 10% will be left as a spongy nanoporous Au. Synthesized 3.1. Dealloying. 3.1.1. Alloy Characterization and Elec-
that way, the NPG samples were rinsed with Barnstead trochemical Results. Alloy specimens AupAg1-p used as testing
Nanopure water immediately after the dealloying and were then vehicles in this work, were analyzed for homogeneity and
immersed in water and kept for an hour to make sure all Ag composition. XRD work done on all three compositions (not
ions diffused out of the pores. shown) ascertained the presence of features characteristic for
2.3. UPD Measurements. UPD-assisted surface area mea- the Au-Ag system.53 Given the close structural similarity of
surements were carried out by chronoamperometry and cyclic Au and Ag (fcc metals with only 0.4% difference in the lattice
voltammetry in a solution containing 1 × 10-2 M Pb(ClO4)2 parameters) that limits the quantitative power of XRD, we used
(99.995+% metal basis, Sigma Aldrich), 1 × 10-1 M NaClO4 EDXS measurements for compositional analysis. Results from
(99.99% Sigma Aldrich), and 1 × 10-2 M HClO4 (double six point EDXS measurements confirmed the alloy composition
distilled, GFS Chemicals). Background experiments assessing with an error of 0.3% for Au0.1Ag0.9, and 0.4% for Au0.2Ag0.8
the charge contribution of side reactions and/or double-layer and Au0.3Ag0.7. In addition, the alloy composition was verified
charging were performed in the same solution but in the absence by registering the onset of a Ag selective dissolution known as
of Pb2+ ions. The reference electrode was Pb wire that was dealloying critical potential.54 It is known that critical potential
pretreated by etching in a diluted nitric acid (1:1 ratio) solution governed by the delicate balance between Ag dissolution and
heated to 50 °C. A Pt wire, flame annealed prior to the Au surface diffusion in single-phase AupAg1-p alloys depends
experiment, served as counter electrode. All UPD experiments strictly upon alloy composition.1,54 Thus, by analyzing the critical
were carried out in a solution that was deoxygenated for at least potential data, one could derive quantitative results for the alloy
2 h by purging with an UHP N2 gas. CA was used to register composition. Figure 1 shows anodic polarization curves in
the current transients of UPD formation at a constant potential perchlorate solution containing Ag ions for all three alloys. The
20 mV (Pb/Pb2+) and UPD stripping at a constant potential 650 anodic curves undoubtedly feature increasing critical potentials
mV (Pb/Pb2+). The surface area of NPG was obtained simply with the rise of the Au fraction in the alloys. A comparison of
by dividing either UPD formation or stripping charges by the the results in Figure 1 with data reported earlier in the literature53
charge density of a Pb UPD layer on a flat polycrystalline Au clearly confirms the EDXS analysis findings and thereby verifies
electrode, prepared by a procedure described elsewhere.49 CV the alloy composition. Finally, the absence of features associated
with a scanning rate 5 mV s-1 was conducted to assess the with dissolution of “free” (unalloyed) Ag on the polarization
kinetic limitations of UPD for the surface area measurement. curves along with the XRD and EDXS results confirms the
In addition to that, regardless of the reference electrode used in homogeneity of the characterized alloys.
this work, all potentials are reported versus a Pb/Pb2+ pseu- 3.1.2. SEM Results. Further in this work, selective removal
doreference electrode (except for the dealloying where Ag/Ag+ of Ag from the above-described AupAg1-p alloys (dealloying)
was quoted as reference), and all solutions used were prepared is carried out to leave porous Au with ligaments and pores in
with the above listed chemicals as received from the vendors nanodimension. The dealloying is performed at a constant
and Barnstead Nanopure water (R > 18.2 Ω cm). current of 1 mA cm-2, and the potential transients registered as
2.4. SEM and BET Experiments. Alternatively to UPD a result of each run are presented in the inset of Figure 1. Here,
measurements, a high-resolution field emission scanning electron the main advantage of working at constant current is the precise
microscope (FESEM) (Zeiss Supra 55 VP) coupled with an in- control of the dealloying depth enabled by a predefined
Pb UPD-Assisted Method for Surface Area Measurement J. Phys. Chem. C, Vol. 113, No. 28, 2009 12365

Figure 3. CV curves of Pb UPD (solid blue) and background (dashed


red) on polycrystalline Au at a sweep rate of 20 mV s-1.

reasons for the discrepancy observed at low Au fractions will


be proposed later in this paper. Another interesting observation
resulting from the analysis of the SEM images presented in
Figure 2 is associated with the similarity between ligament and
pore size. This observation suggests a massive material rear-
rangement during the dealloying process in which pre-existing
percolation ligaments of the more noble metal serve as
nucleation centers for the structuring and growth of the new
interconnected 3D architecture.
3.2. UPD Measurement. Pb UPD on flat polycrystalline Au
is first carried out by CV to investigate the general system
behavior. Then, after determining the formation and stripping
potentials, Pb UPD on NPG is performed using CV and CA.
Further in this section, only results of NPG0.1 are shown as an
illustration of the applicability of the proposed method. Results
featuring identical qualitative and quantitative trends (not shown)
are also obtained for NPG0.2 and NPG0.3.
3.2.1. Pb UPD on Polycrystalline Au. Figure 3 shows the
reversible Pb UPD on a flat polycrystalline Au electrode. The
Figure 2. Top view FESEM micrographs of dealloyed (a) NPG0.1, registered CV allows for determining the formation and stripping
(b) NPG0.2, and (c) NPG0.3 samples. potential and the UPD charge density that is needed as a
normalization factor for the surface area determination of NPG.
dissolution rate. The dissolution time calculation for alloys with The voltammetry presented in Figure 3 shows a solid curve that
different composition is described in the Experimental Section. is similar to those registered by others on polycrystalline Au.55-57
It is shown in the inset of Figure 1, that a time of 30-50 s is It is noteworthy that the sharp peak registered at 0.200 V is
needed for the potential to reach steady state. This steady state typical for Au (111), and the broader wave appearing at 0.450
then operates for dissolution fronts penetrating to 5 µm, which V is characteristic for the (100) and (110) faces of Au.57 The
is the maximum dealloying depth considered in this work. Also, CV suggests also that at the potential of 0.020 mV the Au
the reaching of a steady-state potential for each alloy composi- surface is covered with a monolayer of Pb. Bulk deposition of
tion implies uniformity and isotropy of the substrate structure Pb becomes apparent at potentials more negative than 0.000 V.
developed by dealloying. Also, no Pb atoms are expected on the Au surface at potentials
Following the dealloying, the NPG layers are first character- higher than 0.650 mV. Given the fact that the potential of 0.650
ized using FESEM. An interconnected solid-void 3D structure mV is negative to the critical potential of Ag dealloying from
typical for dealloyed AupAg1-p alloys is observed with dimen- AupAg1-p alloys,54 the UPD measurement would not affect the
sions ranging from 5 to 20 nm for different compositions of alloy intactness.
interest (Figure 2). It is also clearly seen that the ligament and The best choice for “formation” and “stripping” potentials
pore size depend upon the original alloy composition. Thus, in further experiments would be 0.020 and 0.650 V, respectively.
NPG prepared by the dealloying of the Au0.3Ag0.7 alloy (NPG0.3) The UPD charge density can easily be obtained by simply
features the finest pore and ligament size (Figure 2c). At the integrating the current density over time using either the CV
same time, while the dimensions generally increase with a formation or stripping curves and knowing the potential scan
decrease in Au content in the alloy, no clear trend could be rate. Similar to other literature sources57 and after taking into
elucidated solely by simple morphology comparison of alloy account the correction (∼3.5%) for the background current
samples used in this work. For instance, opposing the above- (Figure 3, dotted curve), our results suggest a charge density
mentioned length scale trend, NPG0.1 appears to feature a smaller of 300 ( 10 µC cm-2 for UPD of Pb on a Au polycrystalline
pore and ligament size than those of NPG0.2. Also, its structure flat surface.
is considerably less uniform than that of NPG0.2 and NPG0.3. 3.2.2. Pb UPD on NPG Using CV. Potential cycling between
Finally, as is shown later in this paper, the surface area of NPG0.1 0.650 and 0.020 V is applied to determine the charge of Pb
is almost twice as high compared to that of NPG0.2. Possible UPD on NPG. As shown in Figure 4a, the formation peaks are
12366 J. Phys. Chem. C, Vol. 113, No. 28, 2009 Liu et al.

Figure 4. (a) Pb UPD CV curves of NPG0.1 samples featuring dealloying depths from 0.5 to 5 µm at the scanning rate of 5 mV s-1. (b) Surface
area development, calculated from the corresponded charges of the formation and stripping CV peaks in Figure 4a, as a function of dealloyed depth.

less distinctive for Pb UPD on NPG0.1 (Figure 4a) as compared


to those for polycrystalline Au (Figure 3). At the same time,
two relatively separate stripping peaks at lower dealloying depths
(less than 1 µm), merge to form a dominant peak at the higher
depths (Figure 4a). Indeed, a careful analysis of the results
presented in Figure 4a reveals many aspects of a typical
diffusion-limited deposition scenario operating for the Pb UPD
process on NPG, when a CV experiment is carried out. For
instance, there is a relatively well-shaped peak registered for
the UPD formation when the dealloying depth is 0.5 µm. For
other depths, the cathodic current density (about -1 mA cm-2)
remains relatively constant upon potential scans from 0.250 to
0.020 V, and no peak could be registered. Then, in the reversed
potential scan (anodic direction), the deposited Pb monolayer
begins stripping off immediately only for the sample dealloyed
to the depth of 0.5 µm.
Unlike the CV run for a 0.5 µm thick layer, the UPD
formation process for the rest of the dealloying depths is still
proceeding even when the reverse potential scan is initiated. A
further manifestation of this behavior is represented by negative
currents registered to higher potentials (as high as 0.220 V) at
which a sharp (anodic) stripping peak would occur if a Pb UPD
layer were stripped from a flat Au electrode (compare Figures
3 and 4). These observations undoubtedly suggest a slow
formation process of the Pb UPD layer that becomes even slower
with an increase in dealloying depth.
Mass transport (diffusion) limitations are also hinted at by Figure 5. FESEM images of mechanically cross-sectioned (a) NPG0.1,
the results in Figure 4b, where the surface area obtained by CV (b) NPG0.2, and (c) NPG0.3 samples after dealloying.
formation and stripping charges is plotted as a function of the
depth of dealloying (Figure 4b). Apparently, the insufficient 5 validate the overall uniformity of the dealloyed layers. Thus,
amount of Pb becomes more and more obvious as the dealloying considering the SEM micrographs in Figures 2 and 5, a rough
depth increases, thereby, leading to a leveling of the roughness estimate suggests ligament and pore sizes of 7-10, 12-15, and
factor Rf, defined as the ratio between the actual and geometric 5-6 nm for NPG0.1, NPG0.2, and NPG0.3, respectively. Overall,
surface area. It is to be noted that no difference in the surface the absence of structural anisotropy in dealloyed layers con-
area measurements done either way is generally observed, sidered in this work confirms mass transport limitations as a
suggesting that the time of formation of a complete Pb compelling reason for saturation of the roughness factor, Rf,
monolayer exceeds substantially the time of registration using with the depth of dealloying in the CV experiment. That is why
CV at the scan rate used in our experiments. To ascertain the an accurate charge-based surface area measurement in any
correlation between the leveling of the Rf and structural aspects MNPM structure would require sufficient time. Thus, a viable
of NPG warranting slow diffusion of Pb2+ ions, we registered way to address this issue would be an infinitely slow CV or
SEM micrographs on cross-sectioned NPGp, where p ) 0.1, even a current measurement at constant potential (CA). The
0.2, and 0.3 atomic fraction. For the sake of clarity, only a second option will be described in detail in the next section.
portion of the cross section of interest is presented in Figure 5. 3.2.3. Pb UPD on NPG Using CA. The main benefit to using
While a careful morphology comparison reveals some length CA instead of CV for studying UPD on NPG is that the process
scale differences (reported also by Cattarin, et al.34 for NPG0.25) can be monitored until a complete Pb monolayer is formed or
between the topmost surface of NPG (thickness less than 50 stripped. In these experiments, the potential is pulsed from the
nm) and the cross-sectioned one, the results presented in Figure open circuit potential (OCP, normally higher than 0.650 V) to
Pb UPD-Assisted Method for Surface Area Measurement J. Phys. Chem. C, Vol. 113, No. 28, 2009 12367

Figure 6. (a) Current density transients of Pb UPD formation on NPG0.1 samples after dealloying to different depths (to 5 µm) at a potential of
0.020 V. (b) Surface area development versus dealloying depth.

Figure 7. (a) Current density transients of Pb UPD stripping on NPG0.1 with dealloying depth (to 5 µm) at the potential of 0.650 V. (b) Surface
area development versus dealloying depth.

0.020 V, and the current transient is registered until the


background current level is reached. In this way, the entire
process of Pb UPD formation on NPG can be precisely
monitored regardless of the time needed for the completion of
the monolayer at different dealloying depths (Figure 6a). By
integrating the current density over time and normalizing by
division to the monolayer charge of UPD Pb on a plain Au
polycrystalline electrode, we obtained a linear relationship of
the Rf versus dealloying depth as shown in Figure 6b.
In another run, a stripping experiment presented in Figure 7
is carried out in the following way. After a full monolayer of
Pb is formed (Figure 6a), the potential is pulsed from 0.020 to
0.650 V for the UPD stripping process to take place. While
taking longer than the stripping from a flat surface, the Pb layer Figure 8. Current transients of Pb UPD stripping on a flat polycrys-
stripping (Figure 7a) is faster than its formation counterpart. talline Au electrode at a potential of 0.650 V. Solid blue and red dashed
The overall stripping time also depends upon the depth of curves represnet Pb UPD stripping carried out at 30 and 630 s after
dealloying. In general, trends for UPD stripping are quite similar the layer formation, respectively. Double-layer charging in a background
to those observed in the UPD formation process. It should be solution is represented by the green dotted curve.
noted that prior to the normalization procedure, the charge
obtained from the results in Figures 6a and 7a is reduced by a 6b and 7b is an anticipated result given the overall uniformity
factor of 0.120. This correction takes into account double-layer and isotropy of the dealloyed samples shown in Figures 2 and
charging effects (∼0.085) manifested by the comparison of 5 and reported numerous times in the literature (see ref 1). That
charges under blue and green curves in Figure 8 and background is why the linearity here is regarded as proof of the applicability
charges (∼0.035) seen best through the comparison of blue and of the proposed approach for surface area measurements. Also,
red curves in Figure 3. Also, the close match between the blue the detailed analysis of the results presented in Figures 6-8
and red curves in Figure 8 clearly rules out any possibility for suggests that the CA procedure successfully addresses issues
additional charge accumulation caused by the possible surface with the incomplete Pb monolayer seen earlier with the CV runs.
alloying known to take place in this system as a site exchange It is now interesting to analyze the effect of the Au content
process strictly confined within the topmost substrate layer.58,59 on the Rf and accuracy of the chronoamperometric data obtained
Continuing the analysis of Figures 6b and 7b one finds an separately using formation and stripping of the UPD layer. The
indisputably linear relationship between the roughness factor relationship of the Rf versus dealloying depth is linear for all
and dealloying depth. The linear dependence shown in Figures alloy compositions considered in this analysis as shown in Figure
12368 J. Phys. Chem. C, Vol. 113, No. 28, 2009 Liu et al.

Figure 9. Surface area development at different dealloying thicknesses obtained from (a) UPD formation and (b) UPD stripping using CA.

TABLE 1: Results of Linear Regression Analysis of Data Presented in Figure 9


UPD formation UPD stripping area development
-1 -1
p% Au slope (µm ) correlation coefficient slope (µm ) correlation coefficient surface area density (m2 g-1)
10 233 0.993 173 0.999 65-70
20 157 0.992 116 0.999 20-25
30 277 0.998 224 0.999 30-35

9. In general, qualitatively identical trends of the surface area of that and for the sake of accurate data presentation, only ranges
development are registered regardless of whether formation or of surface area density are shown in Table 1. Presenting ranges
stripping is considered (Figure 9a,b). It can also be seen that instead of values is warranted by the fact that the roughness
the developed surface area at a given dealloying thickness factor, Rf, is measured for a calculated depth of penetration,
decreases in the order of NPG0.3 > NPG0.1 > NPG0.2. Generally, assuming dissolution of the entire Ag during dealloying (eq 1).
without considering structural aspects of the alloys prior to At the same time, work of others61 and our own EDS results
dealloying, the reason for the above trend can be attributed to suggest that up to a third of that Ag could remain trapped in
the argument that smaller pore size is associated with larger the NPG, depending upon the temperature and/or completion
surface area. This way, the order determined by these measure- of the dealloying process. Thus, an additional depth that is
ments is in agreement with the morphology length scale of the unaccounted for in our estimate develops, leading to a reduction
SEM micrographs in Figure 5. In the case of the deep dealloying of the surface area density that is not always trivial to quantify.
considered here, cross-sectioned samples are preferred as more Taking into account the semiquantitative area density ranges
relevant than top view samples (Figure 2). Indeed, SEM images provided in Table 1, we see that higher surface area does not
(not shown) suggest that the topmost surface morphology is necessarily mean a higher area density. Thus, NPG prepared
confined within a layer as thin as 50 nm. While no qualitative from a 10% Au alloy has a lower Rf per 1 µm dealloying depth
difference in the charge trends is found (Figures 9a,b), there than NPG0.3, but it has a higher surface area density. The area
certainly is a quantitative difference between the area measure- density ranges of dealloyed 20% and 30% Au alloys are in
ments using formation and stripping transients. The fitting slopes reasonably good agreement with the results of others on NPM
associated with the measurements in Figure 9 are summarized prepared by dealloying. The most recently reported data
in Table 1. Theoretically, the charges obtained using the UPD measured by BET, impedance, and/or UPD-based approaches
formation and stripping should be equal as an identical amount range from 5 to 41 m2 g-1.7,34,62 Values that are well within the
of atoms participate in both processes. Practically, however, above range could also be indirectly derived by structural
about 22% more surface area is registered on the basis of analysis of recently reported NPG structures with stable ligament
formation in comparison with stripping of the UPD layer. In and pore sizes on the order of 3-6 nm.61,63 Given the above
addition to that, the straight lines obtained from the UPD comparison, the surface area density obtained for NPG0.1 (Table
stripping data feature better fitting intercepts that are closer to 1) appears to be unexpectedly high. While no data is available
the theoretical value of 1. One way of realizing the charge in the literature for dealloyed 10% Au alloys, we believe that
discrepancy in the “formation-stripping” comparison in Table the lack of a pre-existing percolation Au structure at this low
1 is to attribute the additional charge accumulation to side pro- Au fraction is critical for explanation of this anomaly. Indeed,
cesses taking place concomitantly with the UPD. A side process the lowest Au fraction at which such structure exists in this
taking place in that potential range could be the oxygen alloy is believed to be 0.198.64,65 The absence of a percolation
reduction reaction (ORR) on a Au surface.60 Apparently, ORR Au structure leaves no pre-existing nucleation skeleton for
would affect the charge accumulated more during the slower development of the interconnected porous structure resulting
UPD formation process. Because of that argument, UPD from the dealloying process in NPG0.1. This, in turn, would lead
stripping appears to provide more accurate information about to nonuniformity of the structure and to more volume shrinkage
the real surface area development. due to the plastic deformation and/or possible collapse of the
The surface area development could also be analyzed through finest ligament ensembles.66 As a result, the accordingly
the surface area density defined as area developed per gram of processed material would have poorer mechanical properties
Au in the alloy. While this parameter is not in the focus of our compared to those of a standard bicontinuous solid-void
present work, its importance is justified by the immediate structure (like NPG0.3) and would rather represent a more open
association with general criteria for cost effectiveness. Because structure with weakly connected nanoligaments (nanoparticles)
Pb UPD-Assisted Method for Surface Area Measurement J. Phys. Chem. C, Vol. 113, No. 28, 2009 12369

Figure 10. Surface area development at different dealloying depths,


obtained from UPD stripping using CA and CV. Figure 12. Minimum concentration of Pb2+ required for having
instantaneous UPD formation at different pore sizes.

on NPG. The diffusion limitations are a direct consequence of


the insufficient Pb ion concentration inside the pores for
instantaneous UPD formation. This deficiency can be best
understood by looking into an estimate of the minimum Pb2+
concentration, C, required to form instantaneously a monolayer
in pores with a very high aspect ratio. For this purpose, a simple
model treating the pore of NPG as a cylinder with a radius R
associated with the pore size allows for a derivation of the final
formula for calculating C in the following way. The charge
density, σ, of a Pb UPD monolayer on Au is multiplied by the
area of the cylinder wall to be covered (no bases as the pore is
hollow) and is then divided by the product of a mol of electrons
Figure 11. UPD formation time at different dealloying depths, obtained (z is the state of oxidation of the Pb ions, and F is the Faraday
from UPD stripping using CA. Inset: Effect of Pb concentration on constant) and the volume of the cylinder (eq 2).
UPD formation duration. ∆C is the factor of Pb2+ concentration change
with respect to 0.05 M.
σ × Acyl σ × 2πRh 2σ
C) ) ) (2)
and large macro cracks. Structures of that kind that have been zF × Vcyl zF × πR h
2 zFR
synthesized by agglomeration of supported or unsupported
(mainly) Pt nanoparticles (atomic weight 195 versus 197 for
Au), featuring surface area densities nearing 100 m2 g-1.67,68 Thus, for R < 8 nm, a Pb concentration higher than 4 M is
Additional analysis of the surface area development and needed to provide a sufficient amount of atoms to instanta-
discussion of the anomalously high surface area density obtained neously cover NPG with one layer of Pb. Apparently, when R
by dealloying 10% Au alloy will be presented in Validations is less than 3 nm, even a solution of Pb(ClO4)2 with the highest
of the Proposed Method for Surface Area Measurement as well concentration close to its solubility at ambient temperature (∼11
as in a more detailed study aiming at further correlating the mol/L)69 will be unable to provide ions for instantaneous UPD
alloy composition with the surface area density in the near formation. To further support this diffusion limitations argument,
future. the inset in Figure 11 shows a 5-fold faster UPD layer formation
3.3. Kinetic Aspects of Pb UPD on NPG Substrates. It with a 5-fold increase in Pb2+ ion concentration. For a better
has been demonstrated in the previous section that the UPD idea about the big frame encompassing the mass transport
formation is a slow process. Indeed, a comparison between Pb limitations as a function of pore size, Figure 12 summarizes
UPD-assisted surface area measurements at different dealloying data obtained by eq 2. It could be seen that if R is on the order
depth of NPG0.1, conducted by CA and CV and shown in Figure of 3-4 nm, the concentration needed for instantaneous UPD
10, visualizes best this conclusion. Identical results (not shown) formation exceeds 8 - 12 mol/L. The above discussion depicts
are also observed for NPG0.2 and NPG0.3. It is clearly seen that quantitatively the reason for slowing the Pb UPD process on
there is generally no difference in the surface area measurement NPG given the fact that solutions no higher than 0.01 M are
using either method at dealloying depths less than 1 µm. At most frequently used for studying the UPD systems.39 Appar-
depths of dealloying exceeding 1 µm, the surface area keeps ently, higher concentration solutions could be used to quicken
increasing linearly in the CA-based protocol, and it eventually the UPD-based surface area measurements. It ought to be noted,
levels when a CV measurement is performed. In addition, the however, that other effects of such an approach should not be
time of UPD formation increases linearly with dealloying depth discounted. For instance, counterion shielding effects observed
for all three alloy compositions (Figure 11), with the process at higher electrolyte concentration could also invoke additional
being much slower than the one carried out on a flat surface. limitations in the UPD layer formation process. In addition, if
Taking into account that the reversibility of Pb UPD on a flat the solution has a concentration close to the maximum elec-
Au electrode implies fast charge transfer kinetics (Figure 3), it trolyte solubility prior to the UPD stripping process, the peak
is apparent that mass transport (diffusion) limitations are a amount of dissolved atoms would exceed the solubility limit
critical impeding factor that governs the slow Pb UPD formation and a sudden precipitation before the outward transport manages
12370 J. Phys. Chem. C, Vol. 113, No. 28, 2009 Liu et al.

Figure 14. Pb UPD stripping transients of 3 µm thick NPG0.2 samples,


annealed at different temperatures for 10 min. Inset: Surface area change
for the annealed samples.

Figure 13. FESEM micrographs of NPG0.2, dealloyed to a depth of 3


µm and annealed for 10 min at (a) 200 °C, (b) 300 °C, and (c) 400 °C.
Figure 15. Dependences of the modeled NPG surface area develop-
ment as a function of the pore size. Inset: “Building block” unit cell of
NPG that is left after dealloying.
to lower the local concentration could affect the measurement
quality by introducing noise in the signal.70
3.4. Validations of the Proposed Method for Surface Area the UPD-assisted method proposed herein could be made by
Measurement. 3.4.1. Sample Coarsening by Thermal An- a simple surface area modeling of NPG. The model presented
nealing. The first and most straightforward way to assess the below is developed by assuming a “lattice” type structure of
applicability of UPD measurements for surface area determi- the interconnected Au fraction remaining after dealloying as
nation in MNPM is to apply the proposed approach to a system shown by the “building block” unit cell, depicted in the inset
with well-known and documented behavior. It has been observed of Figure 15. As shown in the drawing, a is the “lattice
by SEM that the structure of NPG becomes coarser by thermal parameter” of the unit cell, x and 2(a-x) represent the
annealing.2,21,31,32 Figure 13 shows the FESEM images of ligament (Au) and pore (void left from the Ag dissolution)
dealloyed NPG0.2 samples that are subjected to annealing for size of the NPG structure, respectively. Similar modeling,
10 min in N2 atmosphere. In agreement with previous reports albeit for cylindrical and not intersecting pores, was reported
in the literature,31,32 the ligaments and pores become larger as previously in the literature.71 While representing a very rough
the annealing temperature increases from ambient (Figure 2b) approximation of the remarkable complexity typical for a
to 400 °C (Figure 13 a-c). To register quantitatively the real solid-void interpenetrating structure resulting from
anticipated reduction in the surface area development, we carried dealloying, the proposed model should still be useful in
out an UPD measurement on the samples shown morphologi- providing basic qualitative trends of the surface area evolution
cally in Figure 13 (top view). The respective UPD stripping in porous structures. A comparison of the NPG structure as
curves are presented in Figure 14. A lower stripping current shown in Figures 3 and 5 with the one assumed by the
density and shorter stripping time are observed with the simplistic model in Figure 15 undoubtedly reveals limitations
temperature increase. that would most likely lead to underestimation of the
The annealing temperature impact on the surface area of NPG predicted surface area. Noting that the perfection of the
is even better presented in the inset of Figure 14. The reason proposed surface area model is not among the top priorities
for the (nearly) linear relationship of the surface area reduction of this work, one registers qualitatively similar trends
with respect to the annealing temperature is not immediately comparing the surface area calculated herein and the one
clear at this point, and no discussion is dedicated to this result. obtained by a previously reported model.71 While no details
It is only concluded that the UPD-assisted method could easily of the presented model have yet been published, the Sup-
register quantitative trends that have been only qualitatively porting Information provided with this work reveals sequen-
studied and documented so far in the literature.31,32 tial steps in the derivation of eq 3 that links the roughness
3.4.2. Surface Area Modeling of NPG. A step forward in factor, Rf, and the ligament size, x of NPG developed by
testing and qualitatively confirming the results obtained by dealloying of AupAg(1-p) alloy
Pb UPD-Assisted Method for Surface Area Measurement J. Phys. Chem. C, Vol. 113, No. 28, 2009 12371

Rf )
1500
×
[2p-0.5cos ( cos-1(-p0.5)
3 ) ]
-1

[ ( )]
(3)
x -1.5 cos-1(-p0.5) 3
p cos
3

Figure 15 represents the roughness factor, Rf, as calculated


according to eq 3 in 1 µm thick NPGp, where p ) 0.1, 0.2, and
0.3. Generally, key points of interest here are the power law
increase in the surface area clearly seen at pore sizes less than
20 nm and the increase of the surface area developed at a given
dealloying depth with the increase in the original Au fraction
in the alloy. On the basis of data in Figure 15 and the average
pore size determined (to the extent it is possible) by SEM for
each composition (using the cross-sectioned layers in Figure Figure 16. Determination of surface area development (roughness
5), a comparison between experimentally determined and factor) at 1 µm dealloying depth of alloys with different Au fractions
using the UPD-assisted experiments and modeling method.
predicted-by-modeling roughness factors for the alloys studied
is presented in Figure 16. The comparison reveals a clear trend
in the ratio between experimental and modeling results going
from nearly 3.3 to about 1.3 with the increase of the original
Au fraction in the alloy. Interestingly, the analysis of the ratio
between pore and ligament size as determined by the model
suggests also a similar trend, from 4.1 to about 1.6, respectively
(Supplemental Information). Given literature reports on the
similarity between pore and ligament size in dealloyed struc-
tures72 and considering our own experience suggesting ratios
between pore and ligament size not exceeding 1.4 regardless
of the original alloy composition (Figures 2 and 5 and
Supporting Information), one could anticipate the better match
between the model and experiment at higher Au fractions. At
the same time, the calculated value would also be less than the Figure 17. Surface area measured by the Pb UPD-assisted method
experimentally measured one because even for NPG0.3 the model (crossed line-filled bars) and by BET (plain black bars).
portrays a more opened structure in comparison with the real
one. In summary, while the proposed model has a certain value
in revealing qualitative trends that correlate roughness factor, almost identical results are obtained by comparing a BET
ligament size, and the original Au fraction, other approximations measurement done after water removal and degassing on one
that take into account the structural evolution during dealloying hand and a subsequent Pb UPD-assisted area determination on
are needed for quantitatively matching the experimental results. the other hand (Figure 17, 200 °C, 1 h). The latter result
3.4.3. BET Measurement. BET is the most widely used demonstrates clearly the key limitations of a standard BET setup
technique for the surface area measurements of materials with for surface area measurement of NPG. Indeed, when the result
a fine structure. While rather expensive, BET measurements in obtained by BET is deemed reliable, the original NPG archi-
UHV could be done with minimum pretreatment steps. The most tecture would have coarsened to yield at least a five times lesser
standard setup for BET would include two stages: first degassing surface area density. This finding, along with the requirement
and then analysis. The main obstacle encountered for the usage for at least 1 g of sample for running an accurate BET analysis,
of BET specifically for surface area measurement of NPG gives a distinct edge to the Pb UPD-assisted method for surface
samples occurs in the degassing step, when high-temperature area measurement of MNPM with a fast coarsening rate at low
heating for a long period of time is required to remove water to moderate temperatures.
and absorbed gas that affect the accuracy of the measurement.
4. Conclusions
For instance, holding the temperature at 90 °C for 1 h is
recommended for water removal, and treatment at 350 °C for In summary, the present work illustrates systematically the
3 h is dedicated to the removal of absorbed gaseous contami- applicability of and reveals in detail the limitations of a Pb UPD-
nants. The holding temperature and time might be changed assisted method for surface area determination of MNPM
depending upon the sample itself and the accuracy of the exemplified by NPG. A comprehensive comparison between CV
measurement. As shown in Figure 17, without a temperature and CA in the formation and stripping of a Pb UPD layer on
treatment, it is clearly seen that the BET measurement yields a NPG (i) identifies the registration of stripping current transients
much lower surface area density for NPG0.2 than either the UPD- as the best approach for measurement of surface area and (ii)
assisted method measurement or the surface area modeling. No reveals distinct diffusion limitations slowing the UPD process
significant change (or a slight increase) is registered by BET on NPG with a thickness in the single digit micrometer range.
after the sample is heated at 90 °C for 1 h. This result contradicts The diffusion limitations governing the Pb UPD on NPG
findings of others2 and the results in the present work (Sample substrates are attributed to the insufficient bulk concentration
Coarsening by Thermal Annealing), suggesting a significant of Pb2+ ions for the formation of a full monolayer at a high
reduction of the surface area upon heat treatment. This behavior surface to volume aspect ratio. An estimate confirming the latter
could be attributed to a dynamic balance between an area statement suggested that less than 10% of the Pb monolayer
increase caused by the removal of water and thermal treatment could be instantaneously formed when a Pb concentration of
driven coarsening. In the presented validation experiments, 1-50 mM is used. The method developed in this work shows
12372 J. Phys. Chem. C, Vol. 113, No. 28, 2009 Liu et al.

an overall trend in the surface area developed by dealloying in (26) Tanabe, K. K.; Wang, Z.; Cohen, S. M. J. Am. Chem. Soc. 2008,
the order of NPG0.3 > NPG0.1 > NPG0.2. While no clear-cut 130, 8508.
(27) Carnes, C. L.; Klabunde, K. J. Langmuir 2000, 16, 3764.
understanding of the surface area trends with the original Au (28) Xie, J.; Varadan, V. K. Mater. Chem. Phys. 2005, 91, 274.
fraction has yet been developed, the anomalously high surface (29) Bi, J.; Wu, L.; Li, J.; Li, Z.; Wang, X.; Fu, X. Acta Mater. 2007,
area density registered in NPG0.1 could have important practical 55, 4699.
(30) Trofymluk, O.; Levchenko, A. A.; Tolbert, S. H.; Navrotsky, A.
implications in the design and synthesis of low metal loading Chem. Mater. 2005, 17, 3772.
catalysts. (31) Li, R.; Sieradzki, K. Phys. ReV. Lett. 1992, 68, 1168.
A qualitative validation of the Pb UPD-assisted method for (32) Seker, E.; Gaskins, J. T.; Bart-Smith, H.; Zhu, J.; Reed, M. L.;
surface area determination by a comparison between the Zangari, G.; Kelly, R.; Begley, M. R. Acta Mater. 2007, 55, 4593.
(33) Porter, D. A.; Easterling, K. E. Phase Transformations in Metals
dealloyed and heat-treated NPG specimens shows reasonable and Alloys; CRC Press: Boca Raton, FL, 2004.
agreement with literature data. A simple surface modeling (34) Cattarin, S.; Kramer, D.; Lui, A.; Musiani, M. M. J. Phys. Chem.
approach provides qualitative trends in surface area development C 2007, 111, 12643.
in NPG and suggests that the closest match with experimental (35) Jurczakowski, R.; Hitz, C.; Lasia, A. J. Electroanal. Chem. 2004,
572, 355.
data could be obtained by model structures featuring similar (36) Jurczakowski, R.; Hitz, C.; Lasia, A. J. Electroanal. Chem. 2005,
pore and ligament size. A coarsening effect taking place during 582, 85.
the removal of water and gaseous contaminations limits the use (37) De Levie, R. Electrochemical Response of Porous and Rough
of standard BET protocols for surface area determination of Electrodes. In AdVances in Electrochemistry and Electrochemical Engineer-
ing; Delahay, P., Ed.; Wiley-Interscience: New York, 1967; Vol. 6, pp 329-
NPG. However, a quantitative agreement is found between BET 397.
and Pb UPD-assisted surface measurements for samples treated (38) Song, H.; Hwang, H.; Lee, K.; Dao, L. Electrochim. Acta 2000,
for water and gas removal. Overall, the Pb UPD-assisted method 45, 2241.
(39) Budevski, E.; Staikov, G.; Lorenz, W. J. Electrochemical Phase
is proven as an inexpensive, fast, selective, and sensitive Formation and Growth: An Introduction to the Initial Stages of Metal
technique for the surface area measurement of MNPM. Deposition; VCH: Weinheim, Germany, 1996.
(40) Herrero, E.; Buller, L. J.; Abruna, H. D. Chem. ReV. 2001, 101,
Acknowledgment. The authors acknowledge the support of 1897.
(41) Viyannalage, L. T.; Bliznakov, S.; Dimitrov, N. Anal. Chem. 2008,
this work by the National Science Foundation, Division of 80, 2042.
Materials Research (DMR-0603019). (42) Choi, H.; Laibinis, P. E. Anal. Chem. 2004, 76, 5911.
(43) Watt-Smith, M. J.; Friedrich, J. M.; Rigby, S. P.; Ralph, T. R.;
Supporting Information Available: Sequential steps in the Walsh, F. C. J. Phys. D: Appl. Phys. 2008, 41, 174004.
(44) Jusys, Z.; Kaiser, J.; Behm, R. J. Langmuir 2003, 19, 6759.
derivation of eq 3 and analysis of the ratio between pore and (45) Jusys, Z.; Behm, R. J. J. Phys. Chem. B 2001, 105, 10874.
ligament size. This material is available free of charge via the (46) Colmenares, L.; Jusys, Z.; Behm, R. J. Langmuir 2006, 22, 10437.
Internet at http://pubs.acs.org. (47) Liu, P.; Ge, X.; Wang, R.; Ma, H.; Ding, Y. Langmuir 2009, 25,
561.
(48) Park, Y.; Yoo, S.; Park, S. Langmuir 2008, 24, 4370.
References and Notes (49) Vasilic, R.; Viyannalage, L. T.; Dimitrov, N. J. Electrochem. Soc.
(1) Erlebacher, J.; Aziz, M. J.; Karma, A.; Dimitrov, N.; Sieradzki, K. 2006, 153, C648.
Nature 2001, 410, 450. (50) Viyannalage, L. T.; Vasilic, R.; Dimitrov, N. J. Phys. Chem. C
(2) Ji, C.; Searson, P. C. J. Phys. Chem. B 2003, 107, 4494. 2007, 111, 4036.
(3) Snyder, J.; Livi, K.; Erlebacher, J. J. Electrochem. Soc. 2008, 155, (51) Wilde, C. P.; Zhang, M. J. Electroanal. Chem. 1992, 327, 307.
C464. (52) Kizhakevariam, N.; Weaver, M. J. Surf. Sci. 1992, 277, 21.
(4) Dong, H.; Cao, X. J. Phys. Chem. C 2009, 113, 603. (53) Wang, A.; Hsieh, Y.; Chen, Y.; Mou, C. J. Catal. 2006, 237, 197.
(5) Elliott, J. M.; Birkin, P. R.; Bartlett, P. N.; Attard, G. S. Langmuir (54) Sieradzki, K.; Dimitrov, N.; Movrin, D.; McCall, C.; Vasiljevic,
1999, 15, 7411. N.; Erlebacher, J. J. Electrochem. Soc. 2002, 149, B370.
(6) Liu, H.; He, P.; Li, Z.; Li, J. Nanotechnology 2006, 17, 2167. (55) Kirowa-Eisner, E.; Bonfil, Y.; Tzur, D.; Gileadi, E. J. Electroanal.
(7) Thorp, J. C.; Sieradzki, K.; Tang, L.; Crozier, P. A.; Misra, A.; Chem. 2003, 552, 171.
Nastasi, M.; Mitlin, D.; Picraux, S. T. Appl. Phys. Lett. 2006, 88, 033110. (56) Bondarenko, A. S.; Ragoisha, G. A.; Osipovich, N. P.; Streltsov,
(8) Lu, H.; Li, Y.; Wang, F. Scr. Mater. 2007, 56, 165. E. A. Electrochem. Commun. 2005, 7, 631.
(9) Jia, F.; Yu, C.; Deng, K.; Zhang, L. J. Phys. Chem. C 2007, 111, (57) Engelsmann, K.; Lorenz, W. J.; Schmidt, E. J. Electroanal. Chem.
8424. 1980, 114, 1.
(10) Viyannalage, L. T.; Liu, Y.; Dimitrov, N. Langmuir 2008, 24, 8332. (58) Green, M. P.; Hanson, K. J. Surf. Sci. Lett. 1991, 259, L743.
(11) Yeh, F.; Tai, C.; Huang, J.; Sun, I. J. Phys. Chem. B 2006, 110, (59) Chen, C.; Washburn, N.; Gewirth, A. A. J. Phys. Chem. 1993, 97,
5215. 9754.
(12) Kim, J.; Dohnalek, Z.; Kay, B. D. Surf. Sci. 2005, 586, 137. (60) Juttner, K. Electrochim. Acta 1984, 29, 1597.
(13) Yi, Q.; Huang, W.; Liu, X.; Xu, G.; Zhou, Z.; Chen, A. J. (61) Qian, L. H.; Chen, M. W. Appl. Phys. Lett. 2007, 91, 083105.
Electroanal. Chem. 2008, 619-620, 197. (62) Ji, C.; Searson, P. C. Appl. Phys. Lett. 2002, 81, 4437.
(14) Chang, J.; Hsu, S.; Sun, I.; Tsai, W. J. Phys. Chem. C 2008, 112, (63) Mathur, A.; Erlebacher, J. Appl. Phys. Lett. 2007, 90, 061910.
1371. (64) Rugolo, J.; Erlebacher, J.; Sieradzki, K. Nat. Mater. 2006, 5, 946.
(15) Fukumizu, T.; Kotani, F.; Yoshida, A.; Katagiri, A. J. Electrochem. (65) Stauffer, D.; Aharony, A. Introduction to Percolation Theory; CRC
Soc. 2006, 153, C629. Press: London, 1991.
(16) Zeis, R.; Lei, T.; Sieradzki, K.; Snyder, J.; Erlebacher, J. J. Catal. (66) Parida, S.; Kramer, D.; Volkert, C. A.; Rosner, H.; Erlebacher, J.;
2008, 253, 132. Weissmuler, J. Phys. ReV. Lett. 2006, 97, 035504.
(17) Xu, C.; Xu, X.; Su, J.; Ding, Y. J. Catal. 2007, 252, 243. (67) Tappan, B. C.; Huynh, M. H.; Hiskey, M. A.; Chavez, D. E.; Son,
(18) Ding, Y.; Chen, M.; Erlebacher, J. J. Am. Chem. Soc. 2004, 126, S. F. J. Am. Chem. Soc. 2006, 128, 6589.
6876. (68) Villers, D.; Sun, S. H.; Serventi, A. M.; Dodelet, J. P.; Désilets, S.
(19) Steele, B. C. H.; Heinzel, A. Nature 2001, 414, 345. J. Phys. Chem. B 2006, 110, 25916.
(20) Zeis, R.; Mathur, A.; Fritz, G.; Lee, J.; Erlebacher, J. J. Power (69) Lide, D. R. Handbook of Chemistry and Physics; CRC Press: Boca
Sources 2007, 165, 65. Raton, FL, 2006.
(21) Ding, Y.; Erlebacher, J. J. Am. Chem. Soc. 2003, 125, 7772. (70) Corcoran, S. G.; Sieradzki, K. J. Electrochem. Soc. 1992, 139, 1568.
(22) Biener, J.; Wittstock, A.; Zepeda-Ruiz, L. A.; Biener, M. M.; (71) Lee, M. H.; Sordelet, D. J. Scripta Materiala 2006, 55, 947.
Zielasek, V.; Kramer, D.; Viswanath, R. N.; Weissmuller, J.; Baumer, M.; (72) Erlebacher, J. Dealloying of Binary Alloys: Evolution of Nanopo-
Hamza, A. V. Nat. Mater. 2009, 8, 47. rosity. In Dekker Encyclopedia of Nanoscience and Nanotechnology;
(23) Jia, F.; Yu, C.; Ai, Z.; Zhang, L. Chem. Mater. 2007, 19, 3648. Schwarz, J. A., Contescu, C. I., Putyera, K., Eds.; CRC Press: Boca Raton,
(24) Sanchez, P. L.; Elliott, J. M. Analyst 2005, 130, 715. FL, 2004.
(25) Brunauer, S.; Emmett, P. H.; Teller, E. J. Am. Chem. Soc. 1938,
60, 309. JP901536F

You might also like