Sunteți pe pagina 1din 7

ARTICLES

PUBLISHED ONLINE: 24 FEBRUARY 2013 | DOI: 10.1038/NCHEM.1574

Improving the hydrogen oxidation reaction rate by


promotion of hydroxyl adsorption
Dusan Strmcnik1, Masanobu Uchimura1,2, Chao Wang1, Ram Subbaraman1, Nemanja Danilovic1,
Dennis van der Vliet1, Arvydas P. Paulikas1, Vojislav R. Stamenkovic1 and Nenad M. Markovic1 *

The development of hydrogen-based energy sources as viable alternatives to fossil-fuel technologies has revolutionized
clean energy production using fuel cells. However, to date, the slow rate of the hydrogen oxidation reaction (HOR) in
alkaline environments has hindered advances in alkaline fuel cell systems. Here, we address this by studying the trends in
the activity of the HOR in alkaline environments. We demonstrate that it can be enhanced more than fivefold compared to
state-of-the-art platinum catalysts. The maximum activity is found for materials (Ir and Pt0.1Ru0.9) with an optimal balance
between the active sites that are required for the adsorption/dissociation of H2 and for the adsorption of hydroxyl species
(OHad). We propose that the more oxophilic sites on Ir (defects) and PtRu material (Ru atoms) electrodes facilitate
the adsorption of OHad species. Those then react with the hydrogen intermediates (Had) that are adsorbed on more
noble surface sites.

T
he ever-growing concerns about global warming and energy step towards rationalizing these differences has been taken recently
security demand the expansion of renewable energy sources by a proposition that there are both qualitative and quantitative
as viable alternatives to fossil-fuel-based technologies. In all differences between the HER in alkaline and acidic media11. In par-
of these concepts, hydrogen is the key energy carrier and, despite ticular, the rate of the former reaction is governed by a delicate
several hurdles that still need to be overcome, it seems to be the balance between the charge transfer-induced water dissociation
most likely fuel of the future. The two reactions that will inevitably step, the interaction of dissociation products with a surface (Had
reign over the hydrogen economy are the hydrogen oxidation and OHad) and concomitant rates of the Had recombination step
reaction (HOR) and its reverse, the hydrogen evolution reaction (H2 production) and desorption of OHad (refs 11,20). However,
(HER), in aqueous environments. The HOR predominantly finds such a level of understanding for the HOR in alkaline solutions is
application in fuel cells, while the HER features in various still lacking.
electrolysers. The HER and HOR in acid (2Hþ þ 2e2 ↔ H2) and Here, we have used a rational design approach to develop real-
alkaline (2H2O þ 2e2 ↔ H2 þ 2OH2) environments are also elec- world catalysts for the HOR in alkaline environments, by success-
trochemical reactions of fundamental importance because the fully transferring the knowledge gained from single-crystal surfaces
basic laws of electrocatalysis have been traditionally developed to commercial anode catalysts. First, by studying pH variations in
and verified by examining these two reactions. To date, a large the HER and HOR and then electrocatalytic trends on metal sur-
number of experimental and theoretical methods have been faces, we find that the rate of the HOR is controlled by both the
applied to unravel the reaction mechanisms of the hydrogen reac- substrate H2/Had and substrate OHad energetics. We then propose
tion in acid electrolytes1–8. The intrinsic kinetic rate, defined as that the most active materials (Ir and Pt0.1Ru0.9 bulk alloy)
the rate at which a reaction proceeds at the equilibrium potential provide the optimal interaction energies of H2/Had and OHad
(zero net current), varies by several orders of magnitude depending with metal surfaces.
on the electrode material4,7,9–12. These variations in activity are
closely related to variations in the hydrogen adsorption free energies Results and discussion
from one material to another, with the highest rates observed on Pt Role of pH in the HER and HOR: experiments and simulations.
group materials, with an optimal interaction of Had with the catalyst We start by presenting a series of experimentally measured (Fig. 1a–c)
surface (around zero free energy of adsorption)4,7,9,13. and simulated (Fig. 1d–f ) polarization curves for the HER and
Much less fundamental work has been directed towards under- HOR on Au(111), Pt(111) and polycrystalline Ir (Ir-poly) in
standing the HER and the HOR in alkaline solutions, although acidic (pH 1–4), neutral (pH 4–11) and alkaline (pH 11–13)
these two processes are of paramount importance for the develop- environments. For the HER portion of the polarization curves
ment of alkaline electrolysers and alkaline fuel cell (AFC) systems. shown in Fig. 1a–c, five distinct features are noteworthy. First, at
Traditionally, the differences in the kinetic rates for the HER and the same pH values, the current–potential curves for Au(111) are
HOR on various electrode materials in alkaline environments shifted towards higher overpotentials (0.4–0.8 V) relative to
have also been linked to variations in the hydrogen adsorption Pt(111) and Ir-poly, because of the intrinsic differences in the
energy. Although this supposition is thermodynamically viable, it is Au–H2/Had bond strength13,21. Second, because the rates of the
still puzzling as to why the HER and HOR activities on platinum are HER and HOR on Pt and Ir are facile at very low pH values,
two to three times higher in acid than in alkaline electrolytes11,14–17, we mainly measure the contributions from the concentration
or why the reactions are more sensitive to the catalysts’ surface overpotential, defined as the deviation from the equilibrium
structure in alkaline media than in acids9,15,18,19. An important potential due to the difference in redox couple concentrations in

1
Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, USA, 2 Advanced Materials Laboratory, Nissan Research Center,
Kanagawa, 237-8532, Japan. * e-mail: nmmarkovic@anl.gov

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 1

© 2013 Macmillan Publishers Limited. All rights reserved.


ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574

a Au(111) b Pt(111) c Ir-poly


4 4
Current density (mA cm–2) pH = 13–5 pH = 13 pH = 13

Current density (mA cm–2)

Current density (mA cm–2)


0 2 12 11 2 12 11
4 10.5 10.5
–2 0 0 9–5
3 9–5 4
–2 –2
–4 3 3
2.5 –4 –4
2.5
–6 2.5
2 –6 2 –6
–8 2
–8 –8
1
–10 –10 –10 1
1
–12 –12 –12

–1.8 –1.6 –1.4 –1.2 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 0.2 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 0.2 –0.8 –0.6 –0.4 –0.2 0.0 0.2
E (V versus SHE) E (V versus SHE) E (V versus SHE)

d 2
Au type e 4
Pt type f 4
Ir type
pH = 13–5 pH = 13
Current density (mA cm–2)

Current density (mA cm–2)

Current density (mA cm–2)


0 2 pH = 13 12 11 2
12 11
4 0 10.5 0 10.5 9–5
–2
3 4 9–5 4
–2 3 –2
–4 3
2.5 –4 2.5 –4 2.5
–6
–6 2 –6
–8 2 2
–8 –8
–10 1 –10 1 –10 1
–12 –12 –12

–1.8 –1.6 –1.4 –1.2 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 0.2 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 0.2 –0.8 –0.6 –0.4 –0.2 0.0 0.2
E (V versus SHE) E (V versus SHE) E (V versus SHE)

Figure 1 | Determining the role of OH– on the HOR in alkaline solutions. Current–voltage behaviour was compared versus SHE, which, for the Hþ/H2
couple, is by definition E 0 ¼ 0.0 V. a–c, Measured pH-dependent current–potential polarization curves at rotation rates of 1,600 r.p.m. and a sweep rate of
50 mV s21 for Au(111) (a), Pt(111) (b) and Ir-poly (c). d–f, Simulated current–potential polarization curves for the HER/HOR Au-type electrode (K1anodic ¼ 0,
K2anodic ¼ 0, K1cathodic ¼ 1029, K2cathodic ¼ 1026) (d), Pt-type electrode (both anodic and cathodic, K1 ¼ 1024, K2 ¼ 1023) (e) and Ir-type electrode (both
anodic and cathodic, K1 ¼ K2 ¼ 1) (f). Simulation of the polarization curves was used to determine the role of pH-dependent (H3Oþ, OH2) and pH-
independent (H2 , H2O) reactant concentrations in the overall current–potential relationships. The presence of the proton branch and water branch for the
HER was explained. Also, two distinct mass-transport controlled plateaux at intermediate pH values (pH 9) for the HOR are noted for Pt and Ir. From
comparing the experimental and simulation results for the HOR at pH . 9, we determine that the OH2 species plays a critical role in the overall HOR rates.

the bulk and at the surface of the electrode. Thus, for pH 1–4, any considered here. On the other hand, even though there are quanti-
kinetic analysis (reaction mechanism) of the HOR and HER tatively different reaction trends for the HOR between Ir-poly and
on these two surfaces is meaningless within +200 mV (versus Pt(111) (as discussed at the end of this section), the qualitative
standard hydrogen electrode, SHE). In contrast, due to relatively trends are nearly identical. In particular, pure diffusion-limited cur-
slower rates in alkaline solutions (pH . 11) kinetic analysis rents are observed for pH 13–11, suggesting that the reaction is con-
might be possible but, for reasons explained in Supplementary trolled almost entirely by the mass transport of reactants required
Section S1, will not be considered here. Third, in solutions with for the electrochemical conversion of H2 to H2O. For pH 11–9.5,
pH , 4, pure diffusion-limiting currents are observed at higher however, two mass transfer-dependent plateaux are clearly visible
overpotentials (21.5 , E , 20.8 V for Au(111); 20.9 , E , 20.4 V (decreasing in magnitude with decreasing pH), indicating that the
for Pt(111) and 20.7 , E , 20.2 V for Ir-poly), implying that overall current is influenced by two separate processes (species).
under these experimental conditions the HER on all three Finally, Fig. 1b,c shows that for pH , 9 the reaction rate is strongly
surfaces is controlled by the mass transport of reactive hydronium dependent on the applied potential, whereas below 20.2 V negli-
species (H3Oþ) rather than the charge transfer reaction. Similar gible currents are observed and above 20.2 V the total rate of the
experimental results have also been obtained by Auinger et al.22 reaction is controlled by a pure mass transport of H2.
on a polycrystalline Pt electrode (Pt-poly). Fourth, within the As summarized in the Methods and given in detail in
same potential regions, very small yet discernible currents (on the Supplementary Section S1, to gain further insight into the pH-
order of mA cm22) are observed in solutions with higher pH dependent processes that are controlling the polarization curves in
values (pH . 5). Finally, an additional reduction process takes Fig. 1a–c, we performed a simulation of the experimental results
place at negative potentials of 21.5 V, 20.9 V and 20.7 V for as a means of describing how the variations in bulk and near-
Au(111), Pt(111) and Ir-poly, respectively. Figure 1a reveals that surface concentrations of both pH-dependent ([Hþ] and [OH2])
above pH 5, polarization curves for Au(111) are pH-independent. as well as pH-independent ([H2] and [H2O]) components may
In contrast, Fig. 1b,c shows that in the same pH range Pt(111) influence the total current density (i þ j), using a simple set of
and, in particular Ir-poly, do exhibit discernible pH dependence. equations. This is given as the sum of four processes described by
It is also important to note that irrespective of the material type equations (1) to (4):
(Au(111) versus Pt(111) versus Ir-poly) or the surface structure of
the electrodes (single-crystal versus polycrystalline), the qualitative H2 + 2H2 O  2H3 O+ + 2e− (1)
behaviour of the HER on these materials as a function of pH is
nearly identical. H2 + 2OH−  2H2 O + 2e− (2)
We now analyse the observed pH variations in the HOR region of
the polarization curves on Au(111), Pt(111) and Ir-poly surfaces 2H3 O+ + 2e−  H2 + 2H2 O (3)
(Fig. 1a–c). In line with earlier observations23, Au(111) shows no
measurable currents for the HOR in the range of potentials 2H2 O + 2e−  H2 + 2OH− (4)

2 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2013 Macmillan Publishers Limited. All rights reserved.


NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574 ARTICLES
0 1/2 alkaline environments (pH 13–11) pure diffusion-limiting currents
i = 2FK1 [H2 ]x=0 [OH − ]x=0 e(1−a1 )(F/RT(E−E1 )−ln[H2 ] )

(5) are also observed, suggesting that reactant supply (in our simu-
0
− 2FK1 [H2 O]x=0 e−a1 (F/RT(E−E1 )−ln[H2 ]
1/2
) lations H2 and OH2) is sufficient to maximize the electrochemical
rate of H2 conversion to H2O. In line with the experimental results,
0 1/2 the simulations reveal that the rate of the HOR on Ir is significantly
j = 2FK2 [H2 ]x=0 e(1−a2 )(F/RT(E−E2 )−ln[H2 ] )
higher than on Pt. As a consequence, the contribution of the hydro-
0 1/2
(6) gen oxidation currents to the water branch of hydrogen reduction
− 2FK2 [H3 O+ ]x=0 e−a2 (F/RT(E−E2 )−ln[H2 ] )
currents below 20.8 V is more pronounced for Ir than for Pt.
Figure 1a,d also show that in the range pH 11–9.5, a decrease in
where i and j represent current densities for reactions (2) & (4) and the diffusion-limiting currents is obtained on both Pt and Ir,
(1) & (3), respectively, F is the Faraday constant and [H2]x¼0 , scaling well with the decrease in availability of OH2 ions. Finally,
[OH2]x¼0 , [H3Oþ]x¼0 and [H2O]x¼0 are activities/concentrations for pH , 9.5 diffusion-limiting currents for process (2) are observed
of the reactants at the electrode surface (x ¼ 0). In our treatment on neither of these two surfaces, signalling that the concentration of
of the current density versus potential relationship, coverages of OH2 is so small that the measured currents are (as in the case of the
intermediates (Had/OHad) as well as spectators have been included HER) again in the invisible (mA) range. Clearly, then, this provides
in the rate constants (K1,2), which are effective rate constants for the a first strong indication that OH2 plays an important role in the
four elementary steps; that is, they are not intrinsic rate constants. HOR, an often overlooked detail in the alkaline HOR analysis.
E 01 and E02 are standard potentials for reaction pairs (2) & (4) and
(1) & (3), respectively, a is the transfer coefficient, R is the standard Role of substrate–OHad energetics on the HOR in alkaline
gas constant, and T is the temperature (in kelvin). Note that because environments. Up to this point, we have considered OH2 as a
all the experiments were performed at a hydrogen partial pressure of reactant in the HOR. However, because the source of oxygenated
1 atm, the ln[H2]1/2 term was omitted from equations (5) and (6) in species can also be OHad (formed by OH2 ↔ OHad þ e2), based
our simulation (for further details see Supplementary Section S1). solely on the results summarized in Fig. 1, it would be impossible
The values of K1,2 were estimated from the experimentally observed to unambiguously determine if the reaction requires OH2 or
differences in activity of a particular surface (Fig. 1, Supplementary OHad. In what follows, however, we demonstrate that OHad rather
Fig. S1). Finally, mass transfer effects were introduced in overall than OH2 is the key reactant in the HOR in alkaline solutions.
kinetic rates using equation (7): We note, in passing, that the surface coverage of OHad (QOHad) is
strongly dependent on the bulk concentration of OH2, the elec-
i, j = 0.62nFAv1/2 v−1/6 D2/3 ∗
1 – 4 ([C1 – 4 ] − [C1 – 4 ]x=0 ) (7) tronic properties of the substrate and the applied electrode poten-
tial24. It is also well established that in many electrochemical
where v is the rotation rate, n is dynamic viscosity, D1–4 are diffu- processes OHad plays either a catalytic role or an inhibiting
sion coefficients of the reactants in equations (1) to (4), and role20,24. For the HOR, it has always been considered that the
[C1–4]* and [C1–4]x¼0 are the bulk and surface activities of reactants OHad is a spectator, a species that is blocking the active sites for
in equations (1) to (4), respectively. the adsorption of H2 (ref. 25). However, our results (summarized
To test the validity of this method, we first focused on the in Fig. 2) indicate that this might not be true for alkaline environ-
reduction currents under different pH values. As shown in ments where certain oxygenated species, hereafter denoted ‘reactive
Fig. 1d–f, for pH , 2 the rates of HER on Au, Pt and Ir were deter- OHad’, may be an important reactant in the HOR. Unfortunately,
mined by the concentration of H3Oþ (denoted hereafter as the there are no experimental methods capable of ‘seeing’ either reactive
‘proton branch’) as well as by the values of the rate constants K (the or spectator OHad. To overcome this limitation, it has been custom-
intrinsic activity of the substrate for reaction (3)). Furthermore, in ary to probe both QOHad and the substrate–OHad energetics simply
the pH range 2.5–4.0, the simulation effectively captures the diffu- by establishing adsorption/catalytic trends that are known to
sion-limiting currents for the HER on all three surfaces. In agreement depend on the oxophilicity of the surface atoms. Similarly, the
with experimental results, simulations also reveal that at higher pH trends in the substrate–H2/Had energetics also have to be considered
values (pH . 9), the HER currents are mainly controlled by the here, because the rate of the HOR depends on the energy of adsorp-
pH-independent transformation of H2O to H2 (reaction (4)), signify- tion of reactants (H2) and intermediates (Had)4,7,9,13.
ing the existence of a pH-independent process (that is, the ‘water Three single crystals with completely different affinities for Had
branch’, which is governed by a slow charge-transfer-induced water and OHad were selected: Au(111) with an extremely weak inter-
dissociation step). As discussed further in the following paragraph, action, Ru(0001) with a strong interaction, and Pt(111) with
deviations from an ideal pH-independent behaviour for the HER neither too weak nor too strong interaction with these adsorbates13.
on Pt (E , 20.9 V) and Ir (E , 20.7 V) is entirely due to contri- In alkaline environments, the expected adsorption trends for these
butions of currents associated with the HOR. species are clearly observed in the cyclic voltammograms (CVs)
The simulated polarization curves presented in Fig. 1d–f also depicted in Fig. 2. For example, Fig. 2a shows that while underpo-
describe the hydrogen oxidation currents on the three metals. On tentially deposited hydrogen (Hupd , defined as hydrogen adsorbed
Au, given the weak binding energies between Au and H2 (very at a potential that is more positive than the reversible potential
low K value), the rate of the HOR in acid (reaction (1)) and alkaline for the hydrogen reaction) is not observed on Au(111), a small
(reaction (4)) environments is found to be negligible. Consequently, yet clearly discernible adsorption of OH2 is observed above 0.6 V
in the entire pH range the contribution of the hydrogen oxidation (from Supplementary Fig. S3 and refs 26,27) versus RHE (defined
currents to the hydrogen reduction currents (reactions (3) and as SHE 260 mV/pH). In agreement with the literature24, between
(4)) is also negligible, so, below 21.5 V the water branch for Au 0.05 and 0.35 V, Pt(111) is covered by Hupd , while the formation
exhibits an ideal pH-independent behaviour. As expected, a dis- of the OHad layer is visible only above 0.6 V (Fig. 2b). In contrast,
tinctly different situation is encountered for the HOR on Pt and based on CO displacement experiments26 (not shown), we found
Ir, because Pt/Ir–H2 interactions are optimized for these two that Ru(0001) (Fig. 2d) is covered by OHad even at 0.05 V, verifying
systems (medium and high K values, respectively). For example, a strong Ru–OHad interaction. Hence, we have three different
in acidic environments (pH , 3), our simulated curves reveal that systems with the same trends in energy of interaction between
the HOR on these two surfaces proceeds through the direct (pH- the substrate and the OHad/Had; that is, Au(111) ≪ Pt(111) ,
independent) oxidation of H2 to protons (reaction (1)). In strong Ru(0001). As we demonstrate further in the next paragraph, these

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 3

© 2013 Macmillan Publishers Limited. All rights reserved.


ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574

a E (V versus RHE) b E (V versus RHE) c E (V versus RHE)


0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 –0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 –0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Au(111) Pt(111) Ru(0001)

3 mA cm–2 20 µA cm–2

Current density

Current density
Current density

40 µA cm–2

400 µA cm–2
100 µA cm–2

50 µA cm–2

d e Ir-poly f g Pt 3–4 nm
Ir(111) Ir HSAC
3 mA cm–2 3 mA cm–2
Current density

Current density
Current density
3 mA cm–2
20 nm

400 µA cm–2 h Ir 3–4 nm

20 nm
–0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 –0.2 0.0 0.2 0.4 0.6 –0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
E (V versus RHE) E (V versus RHE) E (V versus RHE)

Figure 2 | Determining the role of OHad–metal energetics on the rate of HOR in alkaline solutions. Establishing the trends in HOR activity as a function of
binding energies between the substrate and OHad/Had species led to the identification of the active sites for the HOR in alkaline environments. a–d, HOR
polarization curves (black) and CVs (dashed grey and red) for Au(111) (a), Pt(111) (b), Ru(0001) (c) and Ir(111) (d) recorded in 0.1 M KOH. To minimize the
pseudocapacitive contributions from the CVs of Au(111) and Ru(0001), polarization cures were recorded with a sweep rate of 1 mV s21. For Pt(111) and Ir(111),
CVs and polarization curves were recorded at 50 mV s21. In d, for comparison, the HOR on Pt(111) is shown as a dashed grey and blue line. e, Polarization
curves for the HOR on Pt-poly (dashed grey and blue) and Ir-poly (black) surfaces. f, In a similar manner as in e, polarization curves for the HOR on Pt
(dashed grey and blue) and Ir (black) high surface area catalysts (HSAC) with nanoparticles of size 3–4 nm and cuboctahedral shape. The loading for both
catalysts was 8 mg cm22. g,h, TEM images of Pt (g) and Ir (h) nanoparticle catalysts. All polarization curves were measured at a rotation rate of
1,600 r.p.m. and reported versus RHE. Blue dashed lines represent zero current density.

results will serve as a basis for establishing variations in the rates of the decreasing from highly defected Pt(110) to almost defect-free
HOR as a function of the metal–Had and metal–OHad bond strength. Pt(111), as found in refs 11,19,28. The order of activity in
As already observed by Conway and co-workers23, Fig. 2a shows Supplementary Fig. S4 is, however, slightly different, mainly due
that the HOR on Au(111) in alkaline solutions only occurs above to the differences in the experimental protocols used in this
0.6 V. Although Conway’s work provided clear evidence that the and previous work (for details see Supplementary Section S3).
HOR may occur on Au surfaces, the explanation for such a puzzling Nevertheless, the fact that the HOR on Pt(111) takes place in the
result has been unclear. Here, based on the fact that the onset of the Hupd potential region (Fig. 2b) confirms our previous hypothesis
HOR coincides with adsorption of OHad , we propose that the HOR that in alkaline solution, OHad is present on defect sites29,30 well
on Au(111) is, in fact, controlled by potential-dependent adsorption below 0.6 V (refs 31,32). We propose, therefore, that the structure
of OHad. As depicted in Fig. 2a, a small ‘pre-ignition’ region sensitivity in alkaline solutions is mainly controlled by the struc-
observed above 0.6 V is followed first by an ignition-like increase ture-sensitive adsorption of OHad on low-coordinated Pt atoms.
in the oxidation current at 1.1 V, and then by a rapid decrease in To further support this, we purposely either decreased (by using a
activity above 1.3 V, possibly due to the loss of virgin metal sites well-established CO-annealing protocol32,33) or increased (by depos-
for adsorption of H2. The onset of HOR on Ru(0001) (Fig. 2c) is iting Pt ad-islands11) the number of low-coordinated sites on
observed essentially at zero overpotential, suggesting that in alkaline Pt(111) terraces. As expected, while the HOR was strongly inhibited
solutions Ru(0001) may have a very high intrinsic activity. The on the former surface, on the latter it was highly activated
maximum oxidation currents are, however, only a few mA cm22 (Supplementary Fig. S5). These results also shed some light on
and, as the potential is increased in the positive direction, this the puzzling differences in activity between the HOR in acid and
current very quickly vanishes to zero (E . 0.3 V). Based on this alkaline solutions. Namely, at low pH values, activity is controlled
observation, it appears that the HOR on Ru(0001) is also concomi- solely by the substrate–H2/Had energetics, but the reaction in alka-
tant with the formation of OHad , which at low potentials may act as line solutions is more demanding, also requiring active sites for the
the reactive OHad but above 0.3 V behaves as a spectator species. We formation of reactive OHad. We conclude that fine-tuning of the
suggest, therefore, that the HOR on Ru(0001) in alkaline solution is adsorption energy of OHad (while keeping the Pt-like adsorption
more influenced by the binding energies between Ru and OHad than energy for H2/Had) could be a new way for designing more efficient
the Ru–Had energetics. anode catalysts that can be utilized in AFC systems.
As discussed frequently in the literature6,27, the best catalysts A material that may be uniquely qualified to fulfil this require-
for the HOR are those with an optimal interaction between the ment is Ir, a metal that is known to be slightly more oxophilic
substrate and the hydrogen intermediates, which is the case for than Pt (a stronger interaction with OHad) but with almost the
Pt–Had bonding. Although this is true for acidic environments, same binding energy with Had (ref. 21). As shown in Fig. 2d,
our results indicate that the HOR in alkaline solutions may not be the hydrogen adsorption/desorption on Ir is accompanied by
controlled entirely by the Pt–Had energetics. In particular, the OHad desorption/adsorption, producing a sharp peak centred
HOR is strongly structure-sensitive, with the order in activity at 0.375 V. The presence of OHad was confirmed using a

4 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2013 Macmillan Publishers Limited. All rights reserved.


NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574 ARTICLES
a i E (V versus RHE) b i E (V versus RHE)
–0.2 0.0 0.2 0.4 –0.1 0.0 0.1 0.2 0.3 0.4
3
2
2
Current density (mA cm–2)

Current density (mA cm–2)


0
1
–2 Pt(111)
Platinum 0
–4 Pt(111) + Ni(OH)2
adislands Pt0.5Ru0.5
Pt0.1Ru0.9 –1
–6
Iridium
–8 –2

–10 –3

ii ii 7
H2 Platinum
6
δ+ δ+ Pt0.5Ru0.5

Current density (mA cm–2)


δ– OH– δ+ δ+
H2O Pt0.1Ru0.9 5
– δ–
Iridium
4

Ni(OH)2 3

2
OHad
Had 1

Figure 3 | Bifunctional catalysts for the HOR in alkaline solution. a, (i) HOR/HER polarization curves for Pt(111) (black) and Pt(111) modified with
Ni(OH)2—coverage 20% (ref. 11) (dashed red and grey). (ii) Schematic representation of the HOR on Ni(OH)2/Pt(111); Ni(OH)2 provides the active sites
for adsorption of reactive OHad , and Pt provides the active sites for dissociative adsorption of H2 and production of Had , which then react with reactive OHad.
b, (i) HOR/HER polarization curves for Pt-poly (black curves) and PtRu alloys with 50% Ru (dashed grey) and 90% Ru (dashed red). (ii) Comparison
between activities (expressed as ‘kinetic currents’ @ 0.05 V) for the HER on monometallic and bimetallic surfaces. Note that the activity of Pt0.1Ru0.9 is so
high that, as in the case of Ir (the dashed green curve), all we measure is the concentration overpotential. Thus, for these two surface it is not possible to
measure true kinetic currents. Significantly, the activities of Ir and Pt0.1Ru0.9 in alkaline solutions and Pt in acid solutions are identical. All experiments were
carried out in 0.1 M KOH, with a 50 mV s21 sweep rate and 1,600 r.p.m. rotation rates.

CO-displacement experiment34 (for details see Supplementary the precious metal loading on the anode of the AFC by as much
Fig. S6). More importantly, Fig. 2d shows that at 0.05 V, Ir(111) is as 80%) serves as the best example that the knowledge obtained
25-fold more active for the HOR than the corresponding Pt(111) by studying the HOR on well-defined extended surfaces can be
surface. As in the case of the three other single-crystal surfaces, used for tuning the catalytic activity of real-world catalysts.
the minute amounts of reactive OHad adsorbed on Ir(111) defect On the basis of the foregoing, there is reason to believe that the
sites at low potentials is sufficient to react effectively with the hydro- HOR in alkaline solutions should be enhanced substantially on
gen intermediates. Provided that, at the same electrode potential, bimetallic materials that can simultaneously provide active sites
QOHad is greater on Ir(111) than on Pt(111), the turnover rate of for the dissociative adsorption of H2 and adsorption of reactive
the HOR (defined as the reaction rate per sites per second) is OHad. The importance of such bifunctional effects is well estab-
higher on Ir(111) than on Pt(111). Not surprisingly, an increase lished in electrocatalysis, for example, the HER on Pt modified
in activity on Ir-poly, albeit slightly less pronounced (a factor of with Ni(OH)2. For this system, the edges of Ni(OH)2 promote dis-
6), is also observed when compared to the Pt-poly electrode sociative adsorption of water and the production of hydrogen and
(Fig. 2c). Also noteworthy in this context is that the activity of Ir- hydroxyl intermediates8. The dissociation step is then followed by
poly at 0.05 V is so high that even in alkaline solution the HOR is hydrogen adsorption and recombination into H2 on the nearby
almost entirely controlled by the concentration overpotential. Pt sites and adsorption of OH2 on Ni(OH)2. Provided that its
Having established the enhancements in activities for the reverse, the HOR, requires the same type of active sites, one might
extended Pt and Ir surfaces, we then extended the same approach anticipate that the very same Ni(OH)2/Pt would be more active
to compare the HOR on the corresponding nanoparticle catalysts than a Ni(OH)2-free Pt electrode. Indeed, Fig. 3a(i) shows that
(Fig. 2f ). Transmission electron microscopy (TEM) analysis Pt(111) decorated with Ni(OH)2 clusters is more active than a
revealed that well-dispersed Pt (Fig. 2g) and Ir (Fig. 2h) catalysts bare Pt(111) surface. As depicted schematically in Fig. 3a(ii), the
have nearly identical cubo-octahedral shapes and a particle size reaction proceeds along the perimeter of hydrogen intermediates
distribution centred in the 3–4 nm range. As a consequence, we adsorbed on Pt sites and neighbouring OHad that are adsorbed on
do not expect to find any significant difference in the trends in Ni(OH)2. This result provides the first successful means of design-
activity for the nanoparticle catalysts compared to the extended sur- ing bifunctional catalysts for the HOR in alkaline environments.
faces discussed above. Indeed, the fact that in Fig. 2f we observe an Another strategy for developing materials for efficient conversion
enhancement in activity by a factor of 5 at 0.05 V (this can reduce of H2 and OHad to H2O is by alloying Pt with more oxophilic

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 5

© 2013 Macmillan Publishers Limited. All rights reserved.


ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574

elements, such as Ru. In this case, Ru would provide the sites for for the latter process (H2O/H2, OH2) being the lower of the two for the systems
OHad , which can then effectively remove the hydrogen intermedi- considered here. Finally, the values of the effective rate constants K used in the
simulations were chosen to satisfy the experimentally observed curves at pH 13
ates that are present on the nearby Pt sites. The results for the and pH 1.
HOR on Pt0.1Ru0.9 and Pt0.5Ru0.5 bulk alloys, summarized in
Fig. 3b, are used as the most direct probe of the effects of OHad Received 11 September 2012; accepted 15 January 2013;
on the rate of the HOR. We notice in passing that, due to enhanced published online 24 February 2013
surface segregation (the enrichment of one element at the surface
relative to the bulk), the annealed surfaces of the Pt0.1Ru0.9 and References
Pt0.5Ru0.5 have, in fact, a composition of 50% Pt and 90% Pt, 1. Parsons, R. The rate of electrolytic hydrogen evolution and the heat of
respectively35. Nevertheless, Fig. 3b(i) and 3b(ii) show that at low adsorption of hydrogen. Trans. Faraday Soc. 54, 1053–1063 (1958).
2. Gerischer, H. Mechanism of electrolytic discharge of hydrogen and adsorption
overpotentials (0.05 V) the order of activity increases in the energy of atomic hydrogen. Bull. Soc. Chim. Belg. 67, 506 (1958).
sequence Pt , Pt0.5Ru0.5 ≪ Pt0.1Ru0.9 ¼ Ir-poly, confirming that 3. Conway, B. E. & Bockris, J. O. M. Electrolytic hydrogen evolution kinetics and its
the HOR in alkaline environments can be controlled by introducing relation to the electronic and adsorptive properties of the metal. J. Chem. Phys.
bifunctional effects. Most significantly, the activity of Pt0.1Ru0.9 is so 26, 532–542 (1957).
high that, as in the case of Ir-poly, all we measure is the concen- 4. Trasatti, S. Work function, electronegativity, and electrochemical behaviour of
metals: III. Electrolytic hydrogen evolution in acid solutions. J. Electroanal.
tration overpotential. Overall, then, bimetallic systems offer a Chem. 39, 163–184 (1972).
unique opportunity for designing a new generation of materials 5. Norskov, J. K., Bligaard, T., Rossmeisl, J. & Christensen, C. H. Towards the
for the HOR in alkaline environments. These materials can be computational design of solid catalysts. Nature Chem. 1, 37–46 (2009).
designed to be as active as the best catalysts in acidic media. We 6. Breiter, M. W. in Handbook of Fuel Cells (eds. Vielstich, W., Lamm, A. &
Gasteiger, H.) 361–368 (Wiley, 2010).
anticipate that extending our bifunctional approach to other
7. Schmickler, W. & Santos, E. in Interfacial Electrochemistry 163–175
systems will ultimately lead to the discovery of active as well as (Springer, 2010).
cost-effective anode materials for AFCs. 8. Wolfschmidt, H., Paschos, O. & Stimming, U. in Fuel Cell Science: Theory,
Fundamentals, and Biocatalysis (eds Wieckowski, A. & Nørskov, J. K.)
Methods 1–70 (Wiley, 2010).
Extended surface electrode preparation. Pt(111), Ir(111), Au(111), Pt-poly and 9. Conway, B. E. in Interfacial Electrochemistry: Theory, Experiment, and
Ir-poly electrodes were prepared by inductive heating for 5 min at 1,050 8C for Applications (ed. Wieckowski, A.) 131–150 (CRC Press, 1999).
Pt, 800 8C for Au and 1,200 8C for Ir electrodes in an argon hydrogen flow 10. Savadogo, O. in Interfacial Electrochemistry: Theory, Experiment, and
(3% hydrogen)32. The Ru(0001) sample was prepared by sputtering and annealing in Applications (ed. Wieckowski, A.) 937–954 (CRC Press, 1999).
ultra high vacuum (UHV). The annealed specimens were cooled slowly to room 11. Subbaraman, R. et al. Enhancing hydrogen evolution activity in water splitting
temperature under an inert atmosphere and immediately covered with a droplet of by tailoring Liþ–Ni(OH)2–Pt interfaces. Science 334, 1256–1260 (2011).
deionized water. Electrodes were then assembled into a rotating disk electrode 12. Conway, B. E. & Tilak, B. V. Interfacial processes involving electrocatalytic
(RDE). Voltammograms were recorded in argon-saturated electrolytes. Polarization evolution and oxidation of H2 , and the role of chemisorbed H. Electrochim. Acta
curves were recorded in hydrogen-saturated electrolyte. 47, 3571–3594 (2002).
13. Greeley, J. & Markovic, N. M. The road from animal electricity to green
High-surface-area catalysts. For the synthesis of Ir nanoparticles, iridium energy: combining experiment and theory in electrocatalysis.
acetylacetonate was reduced by 1,2-tetradecanediol in a benzyl ether solution at Energy Environ. Sci. 5, 9246–9256 (2012).
290 8C, with oleylamine and oleic acid as stabilizing ligands. For the synthesis of 14. Petrii, O. A. & Tsirlina, G. A. Electrocatalytic activity prediction for
Pt nanoparticles, platinum acetylacetonate was reduced by borane tributylamine at hydrogen electrode reaction: intuition, art, science. Electrochim. Acta 39,
120 8C in an oleylamine solution. These nanoparticles were transferred onto the 1739–1747 (1994).
glassy carbon disk and the organic surfactants were removed by thermal treatment 15. Markovic, N. M., Sarraf, S. T., Gasteiger, H. A. & Ross, P. N. Hydrogen
(185 8C) in air28. electrochemistry on platinum low-index single-crystal surfaces in alkaline
solution. J. Chem. Soc. Faraday Trans. 92, 3719–3725 (1996).
Chemicals. Solutions of different pH values were prepared by adding 0.1 M KOH or
16. Markovic, N. M., Grgur, B. N. & Ross, P. N. Temperature-dependent hydrogen
0.1 M HClO4 to 0.1 M KClO4 solution. All chemicals used in our experiments
electrochemistry on platinum low-index single-crystal surfaces in acid solutions.
were obtained in the highest purity from Sigma Aldrich. Electrolytes were
J. Phys. Chem. B 101, 5405–5413 (1997).
prepared with Millipore Milli-Q water. All gases (argon, oxygen and hydrogen) were
17. Sheng, W., Gasteiger, H. A. & Shao-Horn, Y. Hydrogen oxidation and evolution
of 5N5 quality and purchased from Airgas.
reaction kinetics on platinum: acid vs alkaline electrolytes. J. Electrochem. Soc.
Electrochemical measurements. A typical three-electrode Teflon cell was used to 157, B1529–B1536 (2010).
avoid contamination from glass components. Experiments were controlled using 18. Barber, J. H. & Conway, B. E. Structural specificity of the kinetics of the hydrogen
an Autolab PGSTAT 302N potentiostat. The crystal electrodes, embedded into the evolution reaction on the low-index surfaces of Pt single-crystal electrodes in
RDE assembly, were transferred into a standard three-compartment electrochemical 0.5 M dm23 NaOH. J. Electroanal. Chem. 461, 80–89 (1999).
cell, where the voltammograms and/or polarization curves were recorded. The 19. Schmidt, T. J., Ross, P. N. & Markovic, N. M. Temperature dependent
nanocatalysts supported on glassy carbon were measured in a hanging meniscus surface electrochemistry on Pt single crystals in alkaline electrolytes: Part 2.
configuration. All reported polarization curves and voltammograms were first cycle The hydrogen evolution/oxidation reaction. J. Electroanal. Chem. 524,
measurements to limit the effects of possible contamination from the electrolyte36. 252–260 (2002).
20. Subbaraman, R. et al. Trends in activity for the water electrolyser reactions on 3d
Simulation of experimental results. As discussed in the main text, we carried M(Ni,Co,Fe,Mn) hydr(oxy)oxide catalysts. Nature Mater. 11, 550–557 (2012).
out simulations of the experimental results to describe and understand how the 21. Danilovic, N. et al. Enhancing the alkaline hydrogen evolution reaction activity
variations in bulk and near-surface concentrations of both pH-dependent ([Hþ] and through the bifunctionality of Ni(OH)2/metal catalysts. Angew. Chem. Int. Ed.
[OH2]) and pH-independent ([H2] and [H2O]) components may influence the 124, 12663–12666 (2012).
total current density (i þ j) of the hydrogen reaction. In particular, the main purpose 22. Auinger, M. et al. Near-surface ion distribution and buffer effects during
of this approach was to incorporate the mass transport effects related to the various electrochemical reactions. Phys. Chem. Chem. Phys. 13, 16384–16394 (2011).
species and how they control the diffusion-limiting currents in different pH 23. Angerstein-Kozlowska, H., Conway, B. E. & Hamelin, A. Electrocatalytic
regions. We also used this treatment to identify the role of [OH2] in the observed mediation of oxidation of H2 at gold by chemisorbed states of anions.
diffusion-limiting currents at intermediate pH, which led to the necessary insight for J. Electroanal. Chem. 277, 233–252 (1990).
tackling the H2 oxidation reactions in an alkaline environment. However, we did not 24. Marković, N. M. & Ross, P. N. Jr. Surface science studies of model fuel cell
use this method to model the surface kinetic processes (intermediate coverage of electrocatalysts. Surf. Sci. 45, 117–229 (2002).
adsorbates) or to determine the rate-limiting steps for the reactions in these 25. Strmcnik, D. et al. The role of non-covalent interactions in electrocatalytic
solutions. As a result, the rate constants in our equations are not intrinsic rate fuel-cell reactions on platinum. Nature Chem. 1, 466–472 (2009).
constants, but instead are simply potential-dependent effective rate constants. 26. Wang, J. X., Marinković, N. S., Zajonz, H., Ocko, B. M. & Adžić, R. R. In situ
All equations used in our simulations are derived in Supplementary Section S1. X-ray reflectivity and voltammetry study of Ru(0001) surface oxidation in
It is also important to note that, to fully describe the experimentally observed electrolyte solutions. J. Phys. Chem. B 105, 2809–2814 (2001).
polarization curves, the sum of two processes (Hþ/H2 and H2O/H2 , OH2) has to 27. Skúlason, E. et al. Modeling the electrochemical hydrogen oxidation and
be considered. These two processes not only have different standard potential values, evolution reactions on the basis of density functional theory calculations.
E 0 (0 and 20.84 V, respectively), but also different effective rate constants, the one J. Phys. Chem. C 114, 18182–18197 (2010).

6 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2013 Macmillan Publishers Limited. All rights reserved.


NATURE CHEMISTRY DOI: 10.1038/NCHEM.1574 ARTICLES
28. Marković, N. M. et al. Effect of temperature on surface processes at the 36. Subbaraman, R. et al. Origin of anomalous activities for electrocatalysts in
Pt(111)2liquid interface: hydrogen adsorption, oxide formation, and CO alkaline electrolytes. J. Phys. Chem. C 116, 7, 22231–22237 (2012).
oxidation. J. Phys. Chem. B 103, 8568–8577 (1999).
29. Marinkovic, N. S. et al. Hydrogen adsorption on single-crystal platinum Acknowledgements
electrodes in alkaline solutions. J. Electroanal. Chem. 330, 433–452 (1992). This work was supported by the Office of Science, Office of Basic Energy Sciences, Division
30. Strmcnik, D. S. et al. Unique activity of platinum adislands in the CO of Materials Sciences, US Department of Energy (contract no. DE-AC02-06CH11357).
electrooxidation reaction. J. Am. Chem. Soc. 130, 15332–15339 (2008).
31. Marković, N. M., Grgur, B. N., Lucas, C. A. & Ross, P. N. Electrooxidation of
CO and H2/CO mixtures on Pt(111) in acid solutions. J. Phys. Chem. B 103,
Author contributions
D.S. and N.M.M. conceived and designed the experiments. D.S., M.U., D.v.D., N.D.
487–495 (1999).
and A.P.P. performed the experiments. C.W. contributed materials (Ir nanoparticles).
32. Strmcnik, D. S. et al. Unique activity of platinum adislands in the CO
D.S., R.S., V.R.S. and N.M.M. discussed the results and co-wrote the paper.
electrooxidation reaction. J. Am. Chem. Soc. 130, 15332–15339 (2008).
33. Arenz, M. et al. The effect of the particle size on the kinetics of CO
electrooxidation on high surface area Pt catalysts. J. Am. Chem. Soc. 127, Additional information
6819–6829 (2005). Supplementary information is available in the online version of the paper. Reprints and
34. Climent, V., Attard, G. A. & Feliu, J. M. Potential of zero charge of platinum permissions information is available online at www.nature.com/reprints. Correspondence and
stepped surfaces: a combined approach of CO charge displacement and N2O requests for materials should be addressed to N.M.M.
reduction. J. Electroanal. Chem. 532, 67–74 (2002).
35. Gasteiger, H. A., Ross, P. N. & Cairns, E. J. LEIS and AES on sputtered and Competing financial interests
annealed polycrystalline Pt–Ru bulk alloys. Surf. Sci. 293, 67–80 (1993). The authors declare no competing financial interests.

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 7

© 2013 Macmillan Publishers Limited. All rights reserved.

S-ar putea să vă placă și