You are on page 1of 9

DOI: 10.1002/celc.

201402353 Articles

Probing Structure Modification of Palladium


Nanomaterials during Chemical Synthesis by using In Situ
X-ray Diffraction: Electrochemical Properties
Yaovi Holade, Teko W. Napporn,* Claudia Morais, Karine Servat, and K. Boniface Kokoh[a]

Experimental evidence for Pd-based structure modification of ide and glucose as probing molecules. High active surface
electrocatalysts with hydrogen during their chemical synthesis areas and catalytic activities were found for the unmodified Pd
is still missing. Therefore, in situ X-ray diffraction (XRD) tech- nanostructures through electrochemical studies. Notably, the
niques were as used to investigate hydrogen insertion into the reducing agent has a significant effect on the size distribution
lattice of Pd nanomaterials during their preparation. It was of the nanoparticles. Its use can precisely and effectively tune
shown that the use of NaBH4 as reducing agent led to PdHx their catalytic properties. Consequently, the remarkable behav-
whereas l-ascorbic acid gave Pd nanoparticles. The subsequent ior uncovered herein provides new strategic routes that can be
influence of hydrogen insertion on the catalytic properties of used for the preparation of Pd-based nanomaterials.
Pd nanostructures was demonstrated by using carbon monox-

1. Introduction

Materials synthesis for energy conversion and storage has source reducing agent such as l-ascorbic acid.[10] Unfortunately,
emerged as a crucial target for the development of new appli- none of these preparative methods took into account the in-
cations. The control of different parameters during the synthe- stability of Pd in a hydrogen environment.[7b, 11] As reported by
sis is essential for obtaining very active and efficient materials. Zoubov and Pourbaix in the early 60s,[11f] hydrogen absorption
Pd-based nanomaterials have received tremendous interest by Pd is accompanied by a change in the crystalline structure
from all areas of science and technology.[1] In electrocatalysis, of the metal, probably due to the formation of not only a,
Pd-based catalysts are very active toward alcohol electrooxida- b and solid solutions, but also of hydrides Pd2H (or Pd4H2) or
tion[2] and the well-known oxygen reduction reaction (ORR).[3] PdH0.5.[11c, f, 12] The affinity of Pd with hydrogen and the subse-
Among them, PdAg,[3a] PdNi,[3a] PdCu,[4] and PdAu[3c, 5] exhibit quent effect on the physical structure of the metal is well-
good kinetics towards ORR. Simes et al.[6] have shown that documented.[13] Another aspect is the effect of the size of the
PdAu electrodes are good catalysts for sodium borohydride Pd particles on hydrogen absorption.[11c, 13f, 14] In the early of
(NaBH4) mediated electroreduction in fuel cells (FCs). In 2011, 2000s, Suleiman et al.[13f] found that lattice parameter expan-
Shim et al.,[5] showed that Pd-Au bimetallic catalysts exhibit en- sion occurs during the hydrogen loading of nanometer-sized
hanced kinetics toward ORR. Several FCs with Pd-based nano- palladium combined with the phase transition. Indeed, clusters
materials have been tested. Recent advances in bioelectro- of 3.8 nm predominantly have an icosahedral structure, where-
chemistry have highlighted the advantage of using abiotic as those of 6.0 nm have a cubic structure.[13f] Based on these
nanomaterials as suitable substitutes for enzymes for electro- experimental results, more attention must be paid to the ex-
chemical reaction in biofuel cells, due to their long-term stabili- perimental conditions during the preparation of Pd-based
ty.[7] In particular, nanosized metal particles can be suitable for nanomaterials. For their electrochemical applications, control
catalyzing electrochemical reactions in membrane-free, of the structure of Pd-based nanomaterials is of considerable
enzyme-free biofuel cells under physiological conditions for ac- interest. It allows an assessment of the fundamental properties
tivating an implantable device as a pacemaker.[8] To this end, of the electrode as well as aids in the design of advanced
several methods have been proposed to prepare Pd nanoma- electrodes.
terials.[1a, 2b, 3a, 6, 9] Most of these methods used NaBH4 as a reduc- Herein, by using in situ X-ray diffraction (XRD) techniques,
ing agent and only a few investigations used a non-hydrogen we provide some new insights on hydrogen insertion into the
structure of Pd nanoparticles during their synthesis when
[a] Y. Holade, Dr. T. W. Napporn, Dr. C. Morais, Dr. K. Servat, using NaBH4 as reducing agent. Although dilatation of the Pd
Prof. Dr. K. B. Kokoh lattice parameter due to the insertion of hydrogen (during the
Department of Chemistry, Universit de Poitiers chemical synthesis) into Pd nanoparticle structures has been
IC2MP CNRS UMR 7285, 4 rue Michel Brunet - B27
suggested before,[2b, 3a, 6] no in situ physical method was used
TSA 51106, 86073 Cedex 9 (France)
E-mail: teko.napporn@univ-poitiers.fr to elucidate the details. These studies aim to determine the
Supporting Information for this article is available on the WWW under structure modification caused by hydrogen absorption that
http://dx.doi.org/10.1002/celc.201402353. occurs during the chemical synthesis when using NaBH4 as a re-

ChemElectroChem 0000, 00, 0 – 0 1  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Articles
ducing agent and to establish the effect of the change on the
electrochemical properties of the resulting material. Pd nano-
particles were prepared according to “Bromide Anion Ex-
change: BAE” method.[2] Herein, the method has been slightly
modified to synthesize two kinds of Pd nanomaterials support-
ed on Vulcan XC 72R carbon (20 wt. %). The material obtained
by using l-ascorbic acid as reducing agent was named Pd/C-
AA and that obtained by using NaBH4 is referred as Pd/C-
NaBH4. After characterization through thermogravimetric analy-
sis (TGA) to determine the real metal loading, transmission
electron microscopy (TEM) was used to analyze the subse-
quent effect of the reducing agent on the size and distribution
of the nanoparticles.

2. Results and Discussion


2.1. TGA and TEM Characterization of the Nanomaterials
Firstly, the expected metal loading was confirmed by using
Figure 2. TEM micrographs of a) Pd/C-AA and b) Pd/C-NaBH4, and (insets)
thermogravimetric analyses: 16 wt. % (Pd/C-NaBH4) and their corresponding HRTEM images. Histograms of metal particle size distri-
18 wt. % (Pd/C-AA), which are close to the target value of bution with log-normal fit: more than 400 isolated particles were counted:
20 wt. % (Figure 1). In this figure, Vulcan XC 72R carbon, which c) Pd/C-AA and d) Pd/C-NaBH4 materials.

histograms of particles size distribution depicted in Figure 2 c


(for Pd/C-AA) and 2 d (for Pd/C-NaBH4). The mean particle size
determined from LogNormal function fit (Dm,p) was found to
be 8 and 3 nm for Pd/C-AA and Pd/C-NaBH4, respectively.
Clearly, at this experimental level, in contrast to NaBH4, l-ascor-
bic acid promotes the formation of large nanoparticles. The
difference in size is explained by the fact that l-ascorbic acid is
a weak reducing agent, whereas NaBH4 has strong reduction
abilities.[18] Actually, the molar ratio between the reducing
agent and Pd is 15 for 0.1 m NaBH4 and 7 for 0.1 m l-ascorbic
acid. Increasing the l-ascorbic acid concentration up to 1 m (or
Figure 1. TGA curves at 100 mL min1 air flow and 5 8C min1 linear tempera- ratio up to 15) could probably have an effect on the particle
ture variation. Vulcan XC 72R carbon (Curve a); Pd/C-AA (Curve b) and Pd/C-
size reduction.
NaBH4 (Curve c) materials.
The decrease in the mean particle size, which was evaluated
from TEM analysis, might be in the range of the results from
was used as support for Pd nanoparticles, was considered as XRD (denoted here as Lv). However, with XRD, we can only
reference. It was reported that platinum-based nanomaterials measure the crystallite size, which is the coherent diffracting
catalyze carbon combustion reaction, which typically takes crystalline domain, which is very different from particle size. It
place at approximately 600 8C.[8, 15] Based on the results pre- should be noticed that one particle can be composed of sever-
sented in Figure 1, this character can be extended to Pd nano- al crystallites. Prior to the comparison of crystallite sizes by
particles. It is worth emphasizing that both Pd/C-AA and Pd/C- XRD, the average particle size determined by TEM analyses
NaBH4 were prepared in high chemical synthesis yield;[2a] that must be normalized with the total volume of the counted par-
is, 97 and 99 %, respectively. Compared with other meth- ticles (Dv, 3D representation). The procedure is explained in
ods,[7a, 16] this confirms that the BAE approach minimizes metal detail in our recent paper.[17] Indeed, XRD measurements con-
loss during the synthesis as recently shown for the preparation cern the 3D structure; so the result from TEM must be in the
of Pt or gold-based catalysts.[2a, 17] This expected higher value same range before any comparison. When these fundamentals
(> 90 %) for the BAE method has been attributed to its simplic- are understood, particle size from TEM (Dv) must always be
ity of implementation.[2a] bigger than crystallite (domain) size from XRD (Lv) or, in case of
Figures 2 a (for Pd/C-AA) and 2 b (for Pd/C-NaBH4) display small nanoparticles, both of them can be the same. The
TEM micrographs of the as-synthesized nanomaterials. Their domain size (evaluated from XRD) should never be larger than
typical high-resolution images (HRTEM) are presented in the the particle size determined by TEM, when the latter value is
inset. The images show a homogeneous dispersion of the normalized with volume.
nanoparticles on the support, and this is demonstrated by the

& ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 2  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Articles
2.2. Structural Characterization by using In Situ XRD above this temperature. A final diffractogram was recorded at
the initial temperature (30 8C) for comparison. The complete
Pd/C-AA and Pd/C-NaBH4 were first characterized by XRD at XRD patterns are reported in Figure S1b and S1c. The lattice
30 8C. We report in Figure S 1a the XRD pattern for Pd/C-AA parameters were calculated by using the Bragg law (aBragg) and
sample. Except for the peak at approximately 258, assigned to also by interplanar spacing (a(hkl)). The obtained values are re-
Vulcan XC 72R carbon[17] and those marked by an asterisk (red) ported in Tables S1 and S2. Figure 3 a–d present the (111) and
due to the sample support [the ribbon in kanthal (Iron-
Chrome-Aluminum alloy)], the others belong to Pd, which ex-
hibit face-centered cubic (fcc) symmetry. For Pd/C-AA, peaks
are situated at 39.9, 46.5, 68.0, 81.6, 86.6, 104.5, 119.7, and
125.48 and correspond, respectively, to the crystallographic
planes (111), (200), (220), (311), (222), (400), (331), and (420);
see Table S1. Parameters for Pd/C-NaBH4 can be found in
Table S2. Surprisingly, peaks shift towards lower diffraction
angles compared with the Pd/C-AA sample. This can be attrib-
uted to changes in the Pd nanoparticles derived from the two
opposite reducing agents: hydrogen content (NaBH4) and free-
hydrogen compound (l-ascorbic acid). To gain further details,
quantitative analyses were performed through analysis of the
lattice parameter. This parameter was determined by two
methods: interplanar spacing of the crystallographic plane
(hkl) method [Eq. (2)] and Bragg’s law,[19] [Eq. (3)], which is the
combination of Equations (1) and (2). The results are summar-
ized in Tables S1 and S2 for different crystallographic planes.

l ¼ 2dsinq ð1Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Figure 3. In situ XRD patterns for (111) and (220) crystallographic planes of
aðhklÞ ¼ dðhklÞ h2 þ k2 þ l2 ð2Þ the catalysts a,b) Pd/C-AA and c,d) Pd-NaBH4 at different temperatures.

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h2 þ k 2 þ l 2 ð3Þ
aðBraggÞ ¼ l (220) diffraction peaks of Pd obtained at different temperatures
2 sin q
for the two synthesized materials. As shown, no shift of the
In these equations, l is the wavelength of X-ray (1.54060 ); peak position is observed when AA is used as reducing agent
a is the lattice parameter; q is the Bragg’s angle and d(hkl) is the (Pd/C-AA). Moreover, the determined lattice parameter did not
interplanar spacing of crystallographic plane (hkl). vary for different temperatures. In contrast to the behavior de-
The evaluation of the lattice parameter by using Equation (3) scribed for Pd/C-AA material, the Pd/C-NaBH4 materials pre-
for different crystallographic planes (hkl) gave 3.89–3.90  for pared by using NaBH4 as reducing agent showed a shift of the
Pd/C-AA and 3.93–3.95  for the Pd/C-NaBH4, whereas Pd bulk peak position and a variation of the interplanar spacing of
has 3.89  as lattice parameter.[11c, 12b] Values obtained for Pd/C- crystallographic plane (hkl), d(hkl). In fact, the peak positions
AA are equal to or slightly greater than that of the bulk. This shift toward higher 2q values when the temperature increased
can be explained by a dilatation of the lattice parameter due (see Figure 3 c,d, and Table S2). As shown in Figure 2 c and 2d,
to the small size of Pd particles, as already reported.[13f] Howev- the shift is 0.38 and 0.78 for (111) and (200) planes, respectively.
er, this slight dilatation is not sufficient to explain the more sig- The complete set of results is summarized in Table S2. This dif-
nificant variation observed for Pd/C-NaBH4 material. In our pre- ference in the XRD patterns recorded for the two kinds of ma-
vious work as well as in other reports,[3a, 6] we assumed that terials results from a potential modification of Pd nanomateri-
there is hydrogen insertion in the Pd crystal lattice.[2b] This in- als by the reducing agent. From this in situ XRD study, it was
sertion occurs into the network of Pd nanoparticles during the found that the peak positions increased toward those of Pd/C-
chemical synthesis, using NaBH4 as reducing agent. To further AA when the temperature was increased. As seen in Table S2
explore this hypothesis, in situ XRD experiments under (Pd/C-NaBH4 material), results at 150 8C and at 30 8C show that
a helium atmosphere were performed at different tempera- Pd regains its initial peak positions beyond 150 8C.
tures to insure the removal of absorbed hydrogen (see Experi- Figure 4 shows aBragg values obtained for the well-defined
mental Section for full details). Indeed, if the presence of hy- peak (220) at different temperatures. The aBragg value remains
drogen in the Pd crystal is effective, the relaxation of the crys- constant for the Pd/C-AA material, whereas for Pd/C-NaBH4
tal lattice should occur at higher temperature; in this case, material, the lattice parameter decreases at increasing temper-
above 60 8C. The maximum temperature was set at 150 8C to atures (see Table S2). For the final temperature (150 8C), the ob-
avoid Vulcan carbon combustion; nevertheless, hydrogen re- tained aBragg value is the same for both materials, indicating
lease from the network of Pd nanoparticles must be effective that the hydrogen absorbed during the synthesis was totally

ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 3  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Articles

Figure 4. Evolution of Bragg’s lattice parameter for (220) crystallographic


plane during in situ XRD analysis: (&) Pd/C-AA; (*) Pd-NaBH4 at different
temperatures. At 30 8C after heat-treatment at 150 8C: (~) Pd/C-AA; (N) Pd-
NaBH4.

desorbed at 150 8C. The average crystallite size evaluated by


the Debye–Scherrer relation (not shown) demonstrated that
there is no increase of the particles size during this thermal
treatment. This treatment can affect the structure and the
properties of the palladium nanoparticles. Indeed, it is well-
known that the lattice distance usually has a notable effect on Figure 5. CO stripping experiments at 20 mV s1 in 0.1 mol L1 NaOH at
25 8C; CO was adsorbed at 0.3 V vs. RHE for 5 min. Full CVs of a) Pd/C-AA
the catalytic properties of the nanomaterials by tuning the d- and b) Pd/C-NaBH4 : CV before oxidation (black solid line), CO oxidation (red
band center of the metal nanoparticles. Corollary, the dilatation dotted line), and first cycle after CO stripping (blue dashed line). c) Compari-
of the Pd lattice parameter will induce a significant change in son of polarization curves.
the catalytic performance of the particles because the latter
are conditioned by the valence electrons of the metal. To this
end, electrochemical tests were undertaken to identify any domains of Pd are present on the CVs that were recorded with
change in the catalytic properties of the material. a lower potential limit set at 0.3 V vs. RHE to avoid any irrever-
sible hydrogen insertion (absorption) in the Pd nanocrys-
tal.[2b, 20] During the forward potential scan, hydroxyl adsorption
2.3. Electrochemical Characterization of the Nanomaterials
occurs before 0.6 V vs. RHE, followed by Pd surface oxidation,
by CO Stripping
which takes place at high potentials. When the upper potential
The catalytic activities of the two as-prepared Pd nanomateri- limit is fixed at 1.45 V vs. RHE, the formed PdII oxide (PdO) is
als were then investigated. For an initial characterization of cat- then reduced during the backward potential scan at approxi-
alytic performance, experiments on carbon monoxide electro- mately 0.63 V vs. RHE.[2b, 6] The red dotted curves in Figure 5 a
oxidation by CO stripping were undertaken in alkaline medium and 5 b illustrate the adsorbed CO electrooxidation. This reac-
to probe the surface of the nanoparticles. To conduct the CO tion is concurrent with the hydroxyl formation at the Pd sur-
stripping experiment, three steps were involved: 1) adsorption face, indicating a Langmuir–Hinshelwood mechanism,[21] which
of CO on the electrode at a fixed potential (herein 0.3 V vs. involves two adsorbed species. The most important aspect
RHE) for approximately 5 min; 2) nitrogen or argon was then concerns the comparison of the performance of the two kinds
bubbled for more than 15 min to completely remove nonad- of catalytic electrodes. For this purpose, we compared the two
sorbed CO in the bulk solution; 3) Finally, cyclic voltammetry polarization curves recorded during the first cycle of the CO
measurements were performed. In this case, only the remain- stripping experiment (Figure 5 c ). The current was normalized
ing CO that was adsorbed at the electrode surface was electro- with the electrochemical active surface area (SA, see the Exper-
oxidized during the forward potential scan. In the reverse scan, imental Section for the method) and also by the Pd weight. By
there were no reactive species that would be oxidized again. It taking account the active sites (via SA) and the metal content,
should be noted that, the adsorption step starts after record- we can obtain real activity values for comparison. For the
ing a steady-state cyclic voltammogram (CV). CVs before CO same Pd nanoparticle structures, because particles from NaBH4
stripping experiments at 20 mV s1 in a supporting electrolyte are very small compared with those from AA, a very large dif-
(0.1 m NaOH) are depicted in Figure 5 a (Pd/C-AA) and 5 d (Pd/ ference in terms of activity was expected. Otherwise, Pd/C-
C-NaBH4) as solid black curves, scaled versus RHE. Rigorously, NaBH4 must be more active than Pd/C-AA. Against all expecta-
the CV for the third step must be superimposed on the normal tions, Pd/C-AA exhibited the best catalytic activity by present-
steady-state CV. This occurs only when CO adsorption does not ing good kinetics at lower potentials than Pd/C-NaBH4. More
induce any irreversible modification on the catalytic surface, as importantly, Pd/C-AA shows a specific current density of
observed on Pt-based electrode materials. All typical specific 18.2 A cm2 g1 at 820 mV vs. RHE; an 18.2 % increase com-

& ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 4  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Articles
pared with Pd/C-NaBH4 (15.4 A cm2 g1 at 830 mV vs. RHE).
This improvement in catalytic activity could be supported by
the formation of PdHx when NaBH4 is used as reducing agent.
Indeed, the presence of the absorbed hydrogen as PdHx can
significantly reduce the chemisorption of CO on Pd and, subse-
quently, the activity of the nanoparticles. These electrochemi-
cal results constitute experimental evidence for modification
by hydrogen insertion of the catalytic properties of palladium
nanoparticles. The specific electrochemical active surface area
obtained by the PdO peak reduction method (see Experimen-
tal Section) is 60 m2 g1 for Pd/C-AA and 81 m2 g1 for Pd/C-
NaBH4, whereas the utilization of the CO stripping charge Figure 6. a) Cyclic voltammograms recorded at 25 8C and 20 mV s1 in N2-sa-
turated 0.1 mol L1 NaOH (dotted curve a) and in the presence of 10 mm
gives values of 51 m2 g1 (Pd/C-AA) and 74 m2 g1 (Pd/C-
glucose (solid curve b) on Pd/C-AA electrode. b) Comparison of polarization
NaBH4). In both cases, the obtained values are high when com- curves on (c) Pd/C-AA and (g) Pd/C-NaBH4.
pared with reported values of 28 m2 g1 obtained under the
same experimental conditions.[6] Furthermore, it should be takes place in this potential range. We then examined the two
stated that the differences obtained by using the two methods electrode materials derived from Pd/C-AA and Pd/C-NaBH4 cat-
were expected and they were attributed to the incomplete CO alysts by comparing their polarization curves (Figure 6 b). As
monolayer.[22] The differences between the studied materials can be observed in Figure 6 b, no significant difference in the
are not unexpected and are consistent with the obtained parti- specific activity was found between the two anode catalysts
cle size distributions, with the particles being smaller for Pd/C- toward the glucose oxidation reaction, despite the size distri-
NaBH4. bution of the Pd particles produced by the reducing agents.
This could be due to the fact that glucose is a larger molecule
than CO. The specific activity thus seems to be unaffected by
2.4. Electrochemical Activity of the Nanomaterials toward
the particle size difference. Therefore, for the glucose oxidation
Glucose Oxidation
reaction, AA could be used to synthesize Pd nanoparticles
Among the carbohydrates that can be used as fuel in biofuel without activity loss.
cells, glucose is an excellent probing molecule for various ap- The peak position depends on the electrode material
plications.[2a, 7a, c, 8, 23] In the case of the hybrid[2a, 7a] or abiotic[7c, 8] (shape, size, etc) and on the method of preparation. The
biofuel cells, abiotic anode materials have been utilized to re- anodic peak potential is situated at 0.84 and 0.82 V vs. RHE on
place some enzymes. More importantly, it was recently report- Pd/C-NaBH4 and Pd/C-AA, respectively. The position was ap-
ed that replacing enzymes in biofuel cells by metal nanoparti- proximately 1.07 V vs. RHE on Pd/C synthesized from micro-
cles that are well-dispersed onto carbon substrate leads to an wave assisted polyol method[26] and 1.15 V vs. RHE on Pd nano-
improvement in the long-term stability of the fuel cell.[8] For wires.[25a] These values show clearly that the glucose oxidation
the first time, the authors have found that a membrane-free shifts toward lower potentials on the Pd/C catalyst prepared
biofuel cell in which abiotic nanoparticles replaced enzymes[24] by using the BAE approach. By using geometrical surface area
can successfully substitute batteries for the activation of a pace- normalization (see Figure S2), the peak current density was ap-
maker under physiological conditions. The as-prepared Pd/C proximately 5.5 mA cm2 on Pd/C-AA electrode, which is an en-
catalysts were characterized electrochemically thorough the hanced activity compared with other reports.[25–27] Additionally,
glucose oxidation reaction. As pointed out above, the investi- it is notable that Pd/C-AA showed good catalytic performance
gations were performed in alkaline medium to measure the in- for glucose electrooxidation in buffered solution at pH 7.4.
trinsic activity of the catalyst in the absence of some blocking When it was used as anode material in a hybrid glucose bio-
anions such as chloride and phosphate in physiological buf- fuel cell, a better power performance was obtained (result
fered solution. Figure 6 a shows the CV in N2-saturated 0.1 m shown herein).
NaOH (dotted curve a) and in the presence of 10 mm of glu-
cose (solid curve b) on Pd/C-AA electrode. The same graphs
3. Conclusions
normalized with the glassy carbon geometrical surface area
(0.07 cm2) is provided in Figure S2. Glucose oxidation starts at We have investigated the effect of hydrogen insertion into the
approximately 0.4 V vs. RHE followed by a peak current at ap- palladium structure and correlated this finding with electro-
proximately 0.82 V vs. RHE associated with the bulk glucose chemical properties by using two different reducing agents. In
oxidation to gluconate.[2a, 7a, 21b] This onset potential is much contrast to NaBH4, the use of ascorbic acid (AA) as reducing
lower than those reported elsewhere[10b, 25] and reflects good ki- agent slowed down the reduction process of the Pd nanoparti-
netics of the Pd/C catalyst synthesized through the BAE syn- cles, whereas borohydride ions led to narrower nanomaterial
thetic route. The decrease in the current is then attributed to size. The latter particles catalyze the decomposition of the re-
catalyst deactivation due to PdO formation. During the nega- ducing agent to hydrogen, which is inserted into the palladi-
tive going scan, the surface becomes active after 0.6 V vs. RHE um crystallographic structure, as evidenced by in situ XRD
because of the reduction of PdO to Pd; a reaction that typically measurements. Transmission electron microscopy analysis has

ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 5  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Articles
shown that NaBH4 promotes the formation of very small nano- bic acid (AA; 0.1 mol L1, 13 mL), was then added dropwise under
particles (1–7 nm), whereas AA leads to larger particles (3– vigorous stirring to synthesize the Pd/C-NaBH4 or Pd/C-AA nano-
18 nm). These trends have been explained in detail by examin- materials, respectively. Finally, the supported materials were fil-
tered, rinsed several times with ultrapure water, and dried in an
ing the difference in their reduction properties, with NaBH4
oven at 40 8C for 12 h.
being a more powerful reducing agent than AA. In the pres-
ence of a small probing molecule such as CO, the catalytic ac-
tivity of Pd/C was affected to a greater or lesser extent accord-
Physicochemical Characterization of the Abiotic Catalysts
ing to the reducing agent used during the synthesis. Actually,
CO oxidation is less sensitive to the Pd/C-NaBH4 surface, be- To confirm the metal loading, differential and thermogravimetric
cause the dilation of the lattice parameter by the hydride indu- analysis (DTA/TGA) were performed under air atmosphere. During
the analysis, TGA provides metal loading whereas DTA gives infor-
ces CO oxidative removal at higher potentials. Inversely, for
mation about the thermodynamics of the reaction. TA Instruments
a larger molecule such as glucose, the beneficial effect of AA
SDT Q600 apparatus equipped with TA Universal Analysis software
on the lattice parameter is canceled by a larger particle size, was used for DTA/TGA analysis. Two alumina crucibles, one con-
which decreases the electrode surface area of Pd/C-AA. Never- taining approximately 7 mg of the sample and a second empty
theless, the present study reinforces the conclusion that suita- (reference) were thermally heated under air atmosphere (flow
ble reducing agents must be used for the synthesis of ad- 100 mL min1) from 25 8C to 900 8C with a temperature rate of 5 8C
vanced metal nanoparticles for heterogeneous catalysis, and min1.
that a clear fundamental comprehension of their structure is The XRD experiments performed in situ at different temperatures
crucial. were conducted with a Bruker D8 Advance (BRUKER AXS) diffrac-
tometer in a Bragg–Brentano (q-q) configuration with a copper
tube powdered at 40 kV and 40 mA (Cu-Ka1 = 1.54060  and Cu-
Experimental Section Ka2 = 1.54443 ). A nickel filter was installed in a secondary optic to
eliminate the kb component. Diffraction angle (in 2q) ranging from
Chemicals
208 to 1308 were investigated by using a step of 0.0998. To further
Potassium tetrachloropalladate(II) (K2PdCl4, 99 %), sodium borohy- explore the possible modification of palladium structure by hydro-
dride (NaBH4, 99 %), potassium bromide (KBr,  99 %), and l-ascor- gen absorption during the reduction step, XRD experiments were
bic acid ( 99 %) were purchased from Sigma–Aldrich. These chem- performed in situ at different temperatures (30, 50, 70, 100, 150 8C)
icals were analytical grade and were used without further purifica- as illustrated in Figure 7. Between two temperature values, the
tion. Vulcan XC 72R carbon black was provided by Cabot and ther-
mally activated (see section below). Ultra-pure water from Milli-Q
Millipore source (18.2 MW cm at 20 8C) was used in all experiments.

Nanocatalyst Synthesis by Bromide Anion Exchange


Prior to its use as a nanoparticle supporting material, Vulcan XC
72R carbon black was treated at 400 8C under a nitrogen atmos-
phere with the primary aim of removing undesired contaminants
(sulfur) coming from its industrial manufacture.[1] For the pretreat-
ment, an amount of commercial material was placed in a Pyrex
crucible and introduced into a closed furnace under a nitrogen at-
mosphere. The temperature was then elevated at rate of 5 8C min1
up to 400 8C and kept at 400 8C for 4 h. Finally, the sample was
cooled. We selected a temperature of 400 8C to avoid carbon com-
bustion at higher temperature. The nitrogen adsorption-desorption
Figure 7. Experimental setup for in situ XRD measurements at control
isotherm measurements (performed with Micrometrics ASAP 2010
temperature.
instruments) highlighted a surprising enhancement of the specific
surface area, calculated from the Brunauer–Emmett–Teller (BET)
theory,[28] which increases from 262 m2 g1 (for pristine Vulcan) to temperature increase ramp was 5 8C min1. The experiments were
322 m2 g1 for our thermal annealed material, meaning a 23 % conducted under a helium atmosphere at a constant flow of
gain.[17] The supported palladium nanoparticles were then pre- 500 mL min1. Before the analysis, a purge was done under helium
pared according to the recently reported method for tailoring of atmosphere for 90 min at 30 8C. The accumulating time was of 5 s
palladium-based nanoscale active materials.[2b] This BAE method for the initial temperature and 2.5 s for the higher temperature
was slightly modified, herein, to synthesize two kinds of palladium
values to avoid long exposure time. Basically, the sample (Pd/AA or
nanomaterials supported on carbon Vulcan XC 72R (Pd/C). Briefly,
Pd/C-NaBH4) was mounted on a ribbon in kanthal (Iron-Chrome-
the BAE method consists of the reduction of PdII in aqueous solu-
tion in the presence of bromide ion. Firstly, K2PdCl4 (61.4 mg) was Aluminum alloy) and heated by Joule effect. In all cases, the diver-
dissolved in ultrapure water (100 mL) in a thermostatted bath at gence slit was of 0.38 and the opening window of the detector
25 8C. KBr (32.7 mg) was then added under vigorous stirring and was of 38. The analysis of the position and crystallite size values
Vulcan XC 72R carbon black (80 mg) was added under constant ul- were both obtained by fitting the experimental angular range of
trasonic homogenization for 45 min to obtain a metal loading of interest with the HighScorePlus software using the pseudo-Voigt
20 wt. %. The reducing agent, NaBH4 (0.1 mol L1, 15 mL) or l-ascor- function.[29]

& ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 6  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Articles
To obtain further information, in particular on particle dispersion
and particle size distribution, the materials were also characterized
by TEM with a TEM/STEM JEOL 2100 UHR (200 kV) equipped with
a LaB6 filament. The mean particle size and particle size distribu-
tion were determined by measuring the diameter of isolated parti-
cles (more than 400 particles) using ImageJ free software.[30]

Electrochemical Measurements
An analogue potentiostat EG&G PARC Model 362 (Princeton Ap-
plied Research) monitored by a computer and a conventional
three-electrode cell served for all electrochemical experiments.[31] Figure 8. Illustration of the methods used for electrochemical active surface
We used a slab of glassy carbon with a geometrical surface area of area determination by cyclic voltammetry on Pd/C-AA. a) The blue solid line
6.48 cm2 as counter electrode and an Hg/HgO (0.1 m NaOH, MMO, represents the complete steady-state CV whereas the shaded curve shows
0.158 V vs. standard hydrogen electrode) as reference electrode. the integrated region. The experiments were recorded at 20 mV s1 in
The working electrode was composed of a catalytic ink (3 mL) de- 0.1 mol L1 NaOH and 25 8C. b) The “weighing method” showing the cut
region; the inset shows the tracing paper used as reference for weighing on
posited on a 3 mm diameter glassy carbon disk used as conductive
an ultra-sensitive balance.
carrier. The catalytic ink was prepared according to our standard
method[7] by ultrasonic mixing ultrapure water (375 mL), Nafion sus-
pension (5 wt. %, 50 mL), and catalytic powder (4 mg). All cyclic vol- active surface area in cm2, whereas Equation (7) gives the SECSA in
tammetry[6, 8] experiments were performed in 0.1 mol L1 NaOH as m2 g1:
supporting electrolyte, provided by Sigma–Aldrich (ReagentPlus,
97 %) and prepared with ultra-pure water. To avoid dissolved
oxygen interference during the electrochemical experiments, the Qex
SAðcm2 Þ ¼ ð4Þ
electrolytic solution was systematically deoxygenated by bubbling Qm
high quality nitrogen (from Air Liquide) for 30 min prior to any Z Eend

measurement. The effect of hydrogen insertion in Pd nanomateri- 1
Qex ðmCÞ ¼ iðE ÞdE ð5Þ
als on their catalytic properties was probed by using carbon mon- v Eonset

oxide (CO, Air Liquide, ultra-pure). Their catalytic activity was then
examined in 10 mm glucose (d-(+)-glucose, Sigma–Aldrich, 99.5) m1 s0 ab
Qex ðmCÞ ¼ ð6Þ
solution. All experiments were performed at controlled tempera- m0 xy v
ture of 25 8C. For CO stripping experiments, CO was adsorbed at
0.30 V vs. RHE for 5 min and the solution was then deaerated for SA
SECSAðm2 g1 Þ ¼ ð7Þ
a minimum of 30 min. Metal weight

where v is the scan rate, and Eonset and Eend are the limits of integra-
SECSA Evaluation tion (see Figure S5).

Various methods can be used to determine the electrochemical


active surface area of metals: adsorption/desorption of hydrogen,
stripping of a probe molecule (CO), under potential deposition
Acknowledgements
(UPD), or oxide reduction peak.[32] It is well-known that Pd (and in
particular Pd NPs) has the ability to absorb hydrogen into its crys- This research was financially supported by the French National
tal lattice.[11b, c, 20b, c, 33] To avoid this phenomenon, we did not use Research Agency (ANR) through the ChemBio-Energy program
the hydrogen adsorption-desorption region method but instead (2012–2015).
used the reduction peak of palladium oxide (PdO) and CO strip-
ping methods to determine the SECSA. A charge density of
424 mC cm2 (Qm) is associated with the reduction of the formed Keywords: crystal engineering · electrochemistry · fuel cells ·
PdO monolayer,[2b, 34] whereas the charge associated with the oxida- nanoparticles · palladium
tion of the CO monolayer is 420 mC cm2 (Qm). Basically, the electro-
chemical active surface area was evaluated by using Equation (4), [1] a) C. Bianchini, P. K. Shen, Chem. Rev. 2009, 109, 4183 – 4206; b) C. Bian-
where the exchange charge (Qex) is experimentally determined as chini, in Interfacial Phenomena in Electrocatalysis, Vol 51 (Ed.: C. G. Vaye-
aforementioned. Figure 8 a shows the integrated region using PdO nas), Springer, New York, 2011, pp. 203 – 253; c) M. Hartings, Nat. Chem.
reduction; Qex was obtained from Equation (5) using Origin8 soft- 2012, 4, 764 – 764.
ware. To determine confirmation, we used the classical so-called [2] a) Y. Holade, A. B. Engel, S. Tingry, K. Servat, T. W. Napporn, K. B. Kokoh,
“weighing method” to obtain this charge. Briefly, this involves cut- ChemElectroChem 2014, 1, 1976 – 1987; b) Y. Holade, C. Morais, S. Arrii-
ting a known surface of tracing paper (surface s0) and weighing it Clacens, K. Servat, T. W. Napporn, K. B. Kokoh, Electrocatalysis 2013, 4,
as reference (weight m0). Then, the part of the CV containing the 167 – 178.
[3] a) D. Diabat, T. W. Napporn, K. Servat, A. Habrioux, S. Arrii-Clacens, A.
area of interest (here PdO region) is printed on any paper and
Trokourey, K. B. Kokoh, J. Electrochem. Soc. 2013, 160, H302 – H308; b) Y.
weighed (weight m1). After that, the assessment of Qex is a straight- Suo, L. Zhuang, J. Lu, Angew. Chem. Int. Ed. 2007, 46, 2862 – 2864;
forward mathematic task that is summarized in Equation (6). The Angew. Chem. 2007, 119, 2920 – 2922; c) V. Mazumder, M. Chi, K. L. More,
variables a, b, x, y and v used in Equation (6) are represented in S. Sun, Angew. Chem. Int. Ed. 2010, 49, 9368 – 9372; Angew. Chem. 2010,
Figure 8 b. It is highly recommended to perform at least three 122, 9558 – 9562; d) C.-H. Cui, S.-H. Yu, Acc. Chem. Res. 2013, 46, 1427 –
tests. In both scenarios, Equation (4) was used to calculate the 1437.

ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 7  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Articles
[4] S. J. Hwang, S. J. Yoo, J. Shin, Y.-H. Cho, J. H. Jang, E. Cho, Y.-E. Sung, sumoto, T. Ogawa, M. Yamada, K. Sekizawa, T. Taniguchi, ChemElectro-
S. W. Nam, T.-H. Lim, S.-C. Lee, S.-K. Kim, Sci. Rep. 2013, 3, 1309. Chem 2014, 1, 366 – 370.
[5] J. H. Shim, J. Kim, C. Lee, Y. Lee, Chem. Mater. 2011, 23, 4694 – 4700. [17] Y. Holade, C. Morais, K. Servat, T. W. Napporn, K. B. Kokoh, Phys. Chem.
[6] M. Simes, S. Baranton, C. Coutanceau, J. Phys. Chem. C 2009, 113, Chem. Phys. 2014, 16, 25609 – 25620.
13369 – 13376. [18] L. Wang, C. Hu, Y. Nemoto, Y. Tateyama, Y. Yamauchi, Cryst. Growth Des.
[7] a) A. Habrioux, K. Servat, S. Tingry, K. B. Kokoh, Electrochem. Commun. 2010, 10, 3454 – 3460.
2009, 11, 111 – 113; b) P. Jena, C. B. Satterwaite, Electronic Structure and [19] M. F. Perutz, Nature 1971, 233, 74 – 76.
Properties of Hydrogen in Metals, Plenum Press, New York, 1983; c) C. [20] a) Y. Holade, C. Morais, K. Servat, T. W. Napporn, K. B. Kokoh, ACS Catal.
Kçhler, M. Frei, R. Zengerle, S. Kerzenmacher, ChemElectroChem 2014, 1, 2013, 3, 2403 – 2411; b) N. Tateishi, K. Yahikozawa, K. Nishimura, Y.
1895 – 1900. Takasu, Electrochim. Acta 1992, 37, 2427 – 2432; c) C.-C. Hu, T.-C. Wen,
[8] Y. Holade, K. MacVittie, T. Conlon, N. Guz, K. Servat, T. W. Napporn, K. B. Electrochim. Acta 1995, 40, 495 – 503.
Kokoh, E. Katz, Electroanalysis 2014, 26, 2445 – 2457. [21] a) A. Habrioux, D. Diabat, J. Rousseau, T. Napporn, K. Servat, L. Gutaz,
[9] Y. Holade, C. Morais, T. W. Napporn, K. Servat, K. B. Kokoh, ECS Trans. A. Trokourey, K. B. Kokoh, Electrocatalysis 2010, 1, 51 – 59; b) P. Tonda-
2014, 58, 25 – 35. Mikiela, T. W. Napporn, C. Morais, K. Servat, A. Chen, K. B. Kokoh, J. Elec-
[10] a) J. Turkevich, G. Kim, Science 1970, 169, 873 – 879; b) B. Lim, M. Jiang, trochem. Soc. 2012, 159, H828 – H833; c) P. Urchaga, S. Baranton, C. Cou-
J. Tao, P. H. C. Camargo, Y. Zhu, Y. Xia, Adv. Funct. Mater. 2009, 19, 189 – tanceau, G. Jerkiewicz, Langmuir 2011, 28, 3658 – 3663.
200. [22] A. Lpez-Cudero, A. Cuesta, C. Gutirrez, J. Electroanal. Chem. 2005,
[11] a) B. Kappesser, U. Stuhr, H. Wipf, J. Weißmller, C. Klos, H. Gleiter, J. 579, 1 – 12.
Alloys Compd. 1995, 231, 337 – 342; b) H. Kobayashi, M. Yamauchi, H. Ki- [23] A. Heller, B. Feldman, Chem. Rev. 2008, 108, 2482 – 2505.
tagawa, Y. Kubota, K. Kato, M. Takata, J. Am. Chem. Soc. 2008, 130, [24] K. MacVittie, J. Halamek, L. Halamkova, M. Southcott, W. D. Jemison, R.
1818 – 1819; c) D. G. Narehood, S. Kishore, H. Goto, J. H. Adair, J. A. Lobel, E. Katz, Energy Environ. Sci. 2013, 6, 81 – 86.
Nelson, H. R. Gutirrez, P. C. Eklund, Int. J. Hydrogen Energy 2009, 34, [25] a) C. Zhu, S. Guo, S. Dong, Adv. Mater. 2012, 24, 2326 – 2331; b) L. Yan,
952 – 960; d) P. Dibandjo, C. Zlotea, R. Gadiou, C. Matei Ghimbeu, F. A. Brouzgou, Y. Meng, M. Xiao, P. Tsiakaras, S. Song, Appl. Catal. B 2014,
Cuevas, M. Latroche, E. Leroy, C. Vix-Guterl, Int. J. Hydrogen Energy 2013, 150 – 151, 268 – 274.
38, 952 – 965; e) M. Suleiman, J. Faupel, C. Borchers, H. U. Krebs, R. Kirch- [26] A. Brouzgou, L. L. Yan, S. Q. Song, P. Tsiakaras, Appl. Catal. B 2014, 147,
heim, A. Pundt, J. Alloys Compd. 2005, 404 – 406, 523 – 528; f) D. N. 481 – 489.
Zoubov, M. Pourbaix, in Atlas d’quilibres lectrochimiques:  25 C (Ed.: [27] D. Basu, S. Basu, Int. J. Hydrogen Energy 2012, 37, 4678 – 4684.
M. Pourbaix), Gauthier-Villars et Cie, Paris, 1963, pp. 358 – 363. [28] R. Yu, L. Chen, Q. Liu, J. Lin, K.-L. Tan, S. C. Ng, H. S. O. Chan, G.-Q. Xu,
[12] a) L. Schlapbach, Nature 2009, 460, 809 – 811; b) T. B. Flanagan, W. A. T. S. A. Hor, Chem. Mater. 1998, 10, 718 – 722.
Oates, Annu. Rev. Mater. Sci. 1991, 21, 269 – 304. [29] J. I. Langford, D. Louer, P. Scardi, J. Appl. Crystallogr. 2000, 33, 964 – 974.
[13] a) C. M. Ghimbeu, C. Zlotea, R. Gadiou, F. Cuevas, E. Leroy, M. Latroche, [30] a) M. D. Abr moff, P. J. Magalh¼es, S. J. Ram, Biophotonics Int. 2004, 11,
C. Vix-Guterl, J. Mater. Chem. 2011, 21, 17765 – 17775; b) G. Li, H. Ko- 36 – 42; b) C. A. Schneider, W. S. Rasband, K. W. Eliceiri, Nat. Methods
bayashi, J. M. Taylor, R. Ikeda, Y. Kubota, K. Kato, M. Takata, T. Yamamoto, 2012, 9, 671 – 675; c) T. J. Collins, Biotechniques 2007, 43, S25 – S30.
S. Toh, S. Matsumura, H. Kitagawa, Nat. Mater. 2014, 13, 802 – 806; c) L. [31] A. J. Bard, L. R. Faulkner, Electrochemical Methods: Fundamentals and Ap-
Schlapbach, A. Zuttel, Nature 2001, 414, 353 – 358; d) J. A. Eastman, L. J. plications, 2nd ed., Wiley, New York, 2001.
Thompson, B. J. Kestel, Phys. Rev. B 1993, 48, 84 – 92; e) F. A. Lewis, Plati- [32] a) S. Trasatti, O. A. Petrii, J. Electroanal. Chem. 1992, 327, 353 – 376; b) E.
num Met. Rev. 1982, 26, 20 – 27; f) M. Suleiman, N. M. Jisrawi, O. Dankert, Herrero, L. J. Buller, H. D. AbruÇa, Chem. Rev. 2001, 101, 1897 – 1930;
M. T. Reetz, C. Bhtz, R. Kirchheim, A. Pundt, J. Alloys Compd. 2003, c) G. Jerkiewicz, Electrocatalysis 2010, 1, 179 – 199; d) D. Chen, Q. Tao, L.
356 – 357, 644 – 648. Liao, S. Liu, Y. Chen, S. Ye, Electrocatalysis 2011, 2, 207 – 219.
[14] C. Zlotea, F. Cuevas, V. Paul-Boncour, E. Leroy, P. Dibandjo, R. Gadiou, C. [33] a) L. L. Jewell, B. H. Davis, Appl. Catal. A 2006, 310, 1 – 15; b) H. Kobaya-
Vix-Guterl, M. Latroche, J. Am. Chem. Soc. 2010, 132, 7720 – 7729. shi, M. Yamauchi, H. Kitagawa, Y. Kubota, K. Kato, M. Takata, J. Am.
[15] R. Sellin, J.-M. Clacens, C. Coutanceau, Carbon 2010, 48, 2244 – 2254. Chem. Soc. 2010, 132, 5576 – 5577.
[16] a) M. Boutonnet, J. Kizling, P. Stenius, G. Maire, Colloids Surf. 1982, 5, [34] M. Grden, M. Lukaszewski, G. Jerkiewicz, A. Czerwinski, Electrochim. Acta
209 – 225; b) C. Roychowdhury, F. Matsumoto, P. F. Mutolo, H. D. AbruÇa, 2008, 53, 7583 – 7598.
F. J. DiSalvo, Chem. Mater. 2005, 17, 5871 – 5876; c) W.-D. Lee, D.-H. Lim,
H.-J. Chun, H.-I. Lee, Int. J. Hydrogen Energy 2012, 37, 12629 – 12638; Received: October 20, 2014
d) D. Wang, H. L. Xin, R. Hovden, H. Wang, Y. Yu, D. A. Muller, F. J. DiSal- Revised: November 17, 2014
vo, H. D. AbruÇa, Nat. Mater. 2013, 12, 81 – 87; e) G. Sakai, T. Arai, T. Mat- Published online on && &&, 2014

& ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 8  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


ARTICLES
The way you make it! Hydrogen inser- Y. Holade, T. W. Napporn,* C. Morais,
tion into palladium nanoparticles during K. Servat, K. B. Kokoh
their chemical synthesis using a hydro-
&& – &&
gen source compound has been high-
lighted by conducting in situ X-ray dif- Probing Structure Modification of
fraction measurements. Subsequent ef- Palladium Nanomaterials during
fects on the catalytic properties of the Chemical Synthesis by using In Situ X-
materials were scrutinized. ray Diffraction: Electrochemical
Properties

ChemElectroChem 0000, 00, 0 – 0 www.chemelectrochem.org 9  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ

You might also like