Вы находитесь на странице: 1из 10

PNAS PLUS

Active sites and mechanisms for H2O2 decomposition


over Pd catalysts
Anthony Plaucka, Eric E. Stanglandb, James A. Dumesica,1, and Manos Mavrikakisa,1
a
Department of Chemical and Biological Engineering, University of Wisconsin, Madison, WI 53706; and bCore R&D–Inorganic Materials and Heterogeneous
Catalysis, The Dow Chemical Company, Midland, MI 48674

Contributed by James A. Dumesic, February 24, 2016 (sent for review August 18, 2015; reviewed by Alexis T. Bell and Jens K. Nørskov)

A combination of periodic, self-consistent density functional theory are also thermodynamically favorable reactions. The optimal
(DFT-GGA-PW91) calculations, reaction kinetics experiments on a DSHP catalyst must therefore selectively produce H2O2 at high
SiO2-supported Pd catalyst, and mean-field microkinetic modeling rates and preserve H2O2 from decomposition.
are used to probe key aspects of H2O2 decomposition on Pd in the Pd is widely recognized as the most effective transition metal
absence of cofeeding H2. We conclude that both Pd(111) and for the DSHP; however, Pd generally exhibits poor selectivity in
OH-partially covered Pd(100) surfaces represent the nature of the the absence of promoters and is highly active for the H2O2 de-
active site for H2O2 decomposition on the supported Pd catalyst composition reactions. Experimental strategies to improve se-
reasonably well. Furthermore, all reaction flux in the closed catalytic lectivity of Pd-based catalysts include the following (8, 9): adding
cycle is predicted to flow through an O–O bond scission step in either strongly coordinating anions (e.g., CN−, Cl−, Br−) and acids to
H2O2 or OOH, followed by rapid H-transfer steps to produce the the solvent; alloying Pd with other noble metals, namely Au or
H2O and O2 products. The barrier for O–O bond scission is sensitive Pt; and controlling the nature of the catalyst support material.
to Pd surface structure and is concluded to be the central param- The performance of Pd-based catalysts has also been shown to
eter governing H2O2 decomposition activity.
strongly depend on the oxidation state of Pd (i.e., catalyst pre-

CHEMISTRY
treatment conditions, O2:H2 feed ratio, etc.) (8).
|
catalysis density functional theory | hydrogen peroxide | palladium | A common goal in many of these modifications is the minimi-
microkinetic analysis
zation of the H2O2 decomposition reactions (reactions 3 and 4)
(10). Importantly, H2O2 has been identified as the primary
H ydrogen peroxide (H2O2) is a desirable oxidant because the
only by-product from its reduction is water (1). The largest
demand for H2O2 is in the pulp and paper industry (2). In addition,
product on Au-Pd catalysts at low H2 conversion (11, 12), whereas
the subsequent H2O2 decomposition reactions decrease overall
yield as the reaction progresses. Experiments also suggest that the
there are applications for H2O2 in catalytic oxidations (3) such as the active sites for H2O2 synthesis (reaction 1) and decomposition on
epoxidation of propene to propylene oxide (4). However, the current Pd-based catalysts may be different (12, 13). In particular, Hutchings
production process for H2O2, the anthraquinone process, is complex,
and coworkers (14) demonstrated that the H2O2 decomposition
energy intensive, and only economic for large-scale productions (5).
reactions on a Au-Pd/C catalyst could be suppressed by pretreating
The direct synthesis of hydrogen peroxide (DSHP) from H2
the carbon support with HNO3, and this resulted in a stable catalyst
and O2 is an appealing alternative that has the potential to enable
with >95% selectivity for H2O2 during the DSHP reaction. A de-
small-scale integrated production of H2O2 (5). A commercial tailed understanding of the active site(s) and elementary reaction
DSHP process has not yet been implemented, although Degussa/ mechanisms for these undesired reactions would benefit the identi-
Headwaters announced the successful integration of a pilot plant fication of improved catalysts. Nonetheless, there is still much work
for the DSHP and a pilot plant for the manufacture of propylene to be performed to elucidate the optimal structure and composition
oxide in 2005 (6). García-Serna et al. (7) have discussed the
economic viability of a DSHP process in comparison with the Significance
existing industrial-scale anthraquinone process.
The primary catalytic challenge for the DSHP is identifying cat- The use of hydrogen peroxide (H2O2) for catalytic oxidations is
alysts that can maintain high selectivity for H2O2 at industrially rel- limited by the energy-intensive and wasteful process by which
evant H2O2 concentrations. The complete reduction of O2 to H2O is H2O2 is currently produced—the anthraquinone process. The
more thermodynamically favorable than the partial reduction of O2 direct synthesis of H2O2 (DSHP) is a promising alternative pro-
to H2O2: cess, yet catalysts active for this reaction (Pd being the most
widely studied) are generally hindered by subsequent H2O2 de-
H2 ðgÞ + O2 ðgÞ → H2 O2 ð1Þ ΔG°298K = −120.4  kJ=mol, composition. Through a combined theoretical and experimental
[Reaction 1; approach, our work (i) provides an understanding of the nature
of Pd active sites responsible for H2O2 decomposition and (ii)
H2 ðgÞ + O2 ðgÞ → H2 Oð1Þ + 1/ 2O2 ðgÞ ΔG°298K = −237.1  kJ=mol, identifies a single type of elementary step that controls the rate.
These structural and mechanistic insights are important for de-
[Reaction 2: signing improved DSHP catalysts and for developing transition-
metal–catalyzed oxidations that efficiently use H2O2.
Consequently, H2O2 decomposition,
Author contributions: A.P., E.E.S., J.A.D., and M.M. designed research; A.P. performed
H2 O2 ð1Þ → H2 Oð1Þ + 1/ 2O2 ðgÞ ΔG°298K = −116.7  kJ=mol, research; A.P., J.A.D., and M.M. analyzed data; and A.P., E.E.S., J.A.D., and M.M. wrote
the paper.
[Reaction 3; Reviewers: A.T.B., University of California, Berkeley; and J.K.N., Stanford University.
The authors declare no conflict of interest.
and H2O2 hydrogenation by H2, 1
To whom correspondence may be addressed. Email: jdumesic@wisc.edu or emavrikakis@
wisc.edu.
H2 O2 ð1Þ + H2 ðgÞ → 2H2 Oð1Þ ΔG°298K = −353.8  kJ=mol,
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
[Reaction 4; 1073/pnas.1602172113/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1602172113 PNAS | Published online March 22, 2016 | E1973–E1982


of Pd-based catalysts for the DSHP; some recent important contri- periments at 300 K, using an apparatus and procedure described previously
butions on this front can be found in refs. 11 and 15–18. (43), and applying a surface stoichiometry of 2:3 for CO:surface Pd atom (42).
This paper will focus on the mechanism and the nature of the H2O2 decomposition experiments were performed in a 50-mL Parr In-
dominant active site(s) responsible for H2O2 decomposition strument Company Hastelloy C-276 autoclave containing an overhead
magnetic stirrer (Magnetic Drive A1120HC6CH, Parr Instrument Company,
(reaction 3, with no H2 present) under conditions relevant to the
Moline, IL), a fixed thermocouple, and a pressure gauge. A Teflon liner was
DSHP process. We highlight key factors that will aid in the used in all experiments to minimize contact of H2O2 with metallic compo-
identification of improved DSHP catalysts that exhibit minimal nents in the autoclave. Blank experiments were performed before each re-
H2O2 decomposition activity. These findings are also relevant to action to ensure negligible contributions to H 2O 2 decomposition from the
transition-metal–catalyzed oxidation reactions that use H2O2— stirrer/thermowell/liner; these wetted parts of the reactor were passivated using
whether produced in situ or fed as reactant—but whose effi- 25 vol% HNO3 in cases where significant H2O2 decomposition was measured in
ciency may be limited by catalytic H2O2 decomposition (19–22). the absence of catalyst. The bare spSiO2 support (no Pd loaded) was shown to be
inert toward H2O2 decomposition over all conditions studied.
Materials and Methods In a typical reaction, the autoclave was loaded with catalyst, sealed, and
Density Functional Theory. Periodic Pd(111) and Pd(100) slabs were chosen as purged with Ar. The autoclave was then pressurized to 450 psi with 4% H2 in
representative models for the planar surfaces of the supported Pd nano- Ar (Airgas), and held at 323 K for 1 h to reduce the passivated Pd nano-
particles used in the experiments. Pd(111) and Pd(100) have the lowest particles. After cooling to room temperature, the autoclave was again
surface free energies among the clean Pd facets, and a truncated octa- purged with Ar, pressurized to 115 psi with Ar, and cooled to the desired
hedron minimizes the total surface free energy for Pd particles >3–5 nm reaction temperature using a refrigerated bath circulator (ARCTIC A25;
based on a Wulff construction (23, 24). Therefore, in the absence of strong Thermo Scientific). The H2O2 feed solution (12.5 g of 0.08–0.60 M H2O2 in
particle-support interactions, both of these Pd surfaces are expected to be H2O) was prepared by dilution of a nonstabilized 30 wt% H2O2 solution (<10
in high abundance. ppb Cl−; Gigabit; KMG) in ultrapure water (18 MΩ·cm), cooled to reaction
temperature, and then charged into the autoclave using a HPLC pump
All density functional theory (DFT) calculations were performed using the
(Chrom Tech Series 1). The resulting pressure in the autoclave before re-
DACAPO total energy code (25, 26), using the self-consistent PW91 gener-
action was 150 psi. Stirring (1,200 rpm) was then started. Conversion of H2O2
alized gradient approximation (GGA-PW91) (27, 28) to describe the ex-
was determined by titration of the final solution with 0.05 M Ce(SO4)2 using
change correlation energy and potential, and ultrasoft pseudopotentials
ferroin as indicator.
(29) to describe the ionic cores. Electron density was determined by iterative
diagonalization of the Kohn–Sham Hamiltonian, Fermi population of the Initial reaction rates were calculated by fitting a line through a plot of the
Kohn–Sham states (kBT = 0.1 eV), and Pulay mixing of the resulting electron moles H2O2 consumed versus time for conversions under 15% and normal-
density (30). The total energy was then extrapolated to kBT = 0 eV. The izing to the total number of Pd surface atoms determined by the CO uptake
Kohn–Sham one-electron valence states were expanded using a plane wave experiments; all conversion versus time data points were replicated at least
basis with kinetic energy below 25 Ry. two times. The apparent activation energy barrier was determined over a
temperature range of ∼25 K, and the apparent reaction order with respect
The (111) and (100) metal surfaces were modeled using a slab geometry
to H2O2 was determined by varying the feed concentration of H2O2 with all
with a periodically repeated (2 × 2) unit cell and four atomic layers; this
other reaction parameters constant. The apparent reaction order with re-
corresponds to 1/4 monolayer (ML) coverage of a single adsorbate placed in
spect to the O2 product was determined by varying PO2 in the gas phase
the unit cell. The surface Brillouin zone was sampled using 18 special Chadi–
using a 25% O2 in Ar mixture, with all other conditions invariant. This
Cohen (31) k-points for the (111) slabs, and a (6 × 6 × 1) Monkhorst–Pack (32)
mixture was introduced immediately after charging the autoclave with the
k-point mesh for the (100) slabs. All slab layers were fixed for calculations on
H2O2 feed, before stirring.
the (111) slabs [previous calculations show minimal effect of surface re-
laxation on the calculated energetics for a similar system (33)], whereas the
Microkinetic Model. A mean-field microkinetic model was developed to de-
top two metal layers were allowed to relax in the (100) slabs. A distance of
scribe the experimentally measured reaction rates, reaction orders, and
14 Å of vacuum separated successive slabs in the z direction, and adsorption
apparent activation barrier. The model parameters were defined using a
was only permitted on one of the two available surfaces with the electro-
procedure described in our previous work (44–46), using the ZPE-corrected
static potential adjusted accordingly (34, 35). The equilibrium PW91 bulk Pd
BEs and activation energy barriers determined through DFT as initial guesses;
lattice constant has been calculated previously (33) to be 3.99 Å [experi-
preexponential factors and entropies were derived from the DFT-calculated
mental value is 3.89 Å (36)]. Calculations involving O2 were performed spin-
vibrational frequencies. The maximum adsorbate coverage permitted was
polarized.
1 ML, and adsorption/desorption steps were assumed to be quasiequilibrated. In
Binding energies (BEs) are reported based on the total energy of the metal
the case that the microkinetic model-predicted adsorbate coverage exceeded
slab with the adsorbate on it (Eads) with respect to the total energies of the
the minimum adsorbate coverage in the context of the unit cell used in the
free gas-phase adsorbate (Egas) and the clean slab (Eclean). All reported DFT
DFT calculations (1/4 ML), the DFT calculations were repeated with the appro-
results have been corrected for the zero-point energy (ZPE). The minimum
priate spectator species coadsorbed in the unit cell. Note that, although the
energy paths for elementary steps were calculated using the climbing im-
experimental measurements were performed in a three-phase system using
age–nudged elastic band method (37, 38) with at least seven intermediate
conditions relevant to a DSHP process, no corrections were made to the DFT
images, and the transition state was verified by identification of a single
calculations to reflect potential interaction with the liquid phase. Further details
imaginary frequency along the reaction coordinate. Vibrational frequencies
of the microkinetic model formulation and parameter sets are provided in
were calculated by diagonalization of the mass-weighted Hessian matrix
Supporting Information.
and using the harmonic oscillator approximation (39).
Results
Experiments. A 0.09 wt% Pd/spSiO2 (spSiO2 denotes the spherical silica sup-
port) catalyst was prepared for reaction kinetics experiments. The spSiO2 was The decomposition of H2O2 has been studied both in the vapor
synthesized by a modified Stöber process (40) described in a previous pub- phase (47) and aqueous phase (48) (thermal, noncatalyzed), and
lication (41), resulting in spherical silica particles (∼100–200 nm in diameter) over a variety of materials including metal oxides (49–51) and
with no internal pore structure and a Brunauer–Emmett–Teller surface area metal ions in solution (52, 53). Based on these studies, we have
of 21 m2/g (41). The Pd was loaded onto the spSiO2 by vacuum evaporative compiled an encompassing network of 17 elementary reactions
impregnation (in a rotary evaporator) using a solution of Pd(II) acetate dis- involving four closed-shell species (H2O2, H2O, O2, H2) and four
solved in dichloromethane. The dried material was reduced in a quartz cell surface intermediates (O, H, OH, OOH), which are shown in
following a procedure described in ref. 42 to promote formation of large Pd Table 1. Elementary reactions are classified as follows: adsorp-
particles (average particle size of 5.6 ± 2.4 nm determined from scanning
tion/desorption, O–O bond scission, dehydrogenation, and hy-
transmission electron microscopy images of the Pd/spSiO2 catalyst) that
better compare with the Pd(111) and Pd(100) DFT models. This heat treat-
drogen transfer. Note that the majority of calculations presented
ment procedure involved a temperature ramp to 673 K (10 K/min) and 3-h below on Pd(111) are based on a previous publication (33), and
hold at 673 K in flowing H2 (30 mL/min), followed by cooling to room these results will not be described in detail here aside from
temperature under flow of Ar (30 mL/min) and passivation with 1% O2 in Ar. noting key differences between binding properties and reaction
The Pd surface site density was determined by irreversible CO uptake ex- energetics on Pd(111) and Pd(100). Reference 54 also presents a

E1974 | www.pnas.org/cgi/doi/10.1073/pnas.1602172113 Plauck et al.


PNAS PLUS
Table 1. Calculated BEs of adsorbed species, their preferred of OOH* on Pd(100) is −1.28 eV—stronger than its binding
adsorption sites, and O–O bond lengths (dO–O) on Pd(111) and energy on Pd(111) by 0.34 eV. However, the site preference for
Pd(100) OOH* on Pd(100) and Pd(111) is weak: on Pd(100), OOH* can
Pd(111)† Pd(100) also bind to top and hollow sites with less than 0.12 eV difference
in binding energy from its most stable adsorption site; on Pd(111),
Adsorption Adsorption OOH* can also bind through its nonhydrogenated oxygen atom to
Species site BE, eV dO–O, Ň site BE, eV dO–O, Ň bridge sites with less than 0.03 eV difference in binding energy
from its most stable adsorption site.
H* fcc −2.70 N/A Hollow −2.74 N/A Molecular oxygen (O2*) has the largest disparity in binding
O* fcc −3.64 N/A Hollow −3.90 N/A strength between Pd(111) and Pd(100); the BE on Pd(100) is
OH* Bridge-tilted −2.03 N/A Bridge-tilted −2.43 N/A
−1.27 eV, which is 0.77 eV stronger than that on Pd(111). O2*
OOH* Bent-top −0.94 1.46 Bent-bridge −1.28 1.51
binds flat on Pd(100) centered over a hollow site, whereas on
H2O* Top −0.22 N/A Top −0.30 N/A
Pd(111) O2* binds across a hcp site with one O atom at a bridge
H2O2* Top −0.32 1.48 Top −0.36 1.49
position and the other at a top position. Interestingly, Long et al.
O2* Top-bridge −0.50 1.35 Hollow −1.27 1.41
(57) used probe molecules and electron spin resonance spec-
Reference energy corresponds to the adsorbate in the gas phase far away troscopy to show that Pd(100) can more readily activate O2
from the metal surface. N/A, not applicable. through excitation of ground-state triplet O2 to reactive singlet

Data based on ref. 33. O2. This result is in agreement with our calculations; O2* retains

Calculated gas-phase dO–O for O2, OOH, and H2O2 are 1.24, 1.35, and 1.48 Å, some of its magnetic moment on Pd(111) (33) but a negligible
respectively. magnetic moment on Pd(100). The strong affinity of Pd(100) for
O2* and O* are reflected in the tendency to reconstruct to a ki-
netically stable (√5 × √5)R27° surface oxide phase under moderate
subset of DFT results on Pd(100). The “*” appended to a species chemical potentials of O2 (58). The next best adsorption site for O2
denotes adsorption at a single surface site, or an unoccupied on Pd(100) is a top–top site with a binding energy of −0.85 eV.
surface site if the “*” stands alone.

CHEMISTRY
The binding energies of H2O* and H2O2* are weak (<0.4 eV)
on both Pd(111) and Pd(100). H2O* preferentially binds to top
Thermochemistry and Binding Configurations of Reaction Intermediates
sites on both Pd(111) and Pd(100) with the O–H bonds parallel
on Clean Pd Surfaces. Table 1 summarizes the most stable adsorption
to the Pd surface. The binding energy of H2O* on Pd(100) is
sites and binding energies for all surface species on Pd(111) and
−0.30 eV. H2O2* also preferentially binds to top sites and adopts
Pd(100). Images of the individual adsorbates in their preferred
the trans configuration on both Pd facets; one oxygen atom is
binding geometry on Pd(100) can be viewed in Fig. 1; refer to ref. 33
bound to a top site with its hydrogen atom pointing slightly away
for the corresponding images on Pd(111). The Pd(100) facet is more
from the surface plane, and the other oxygen atom is positioned
open than the Pd(111) one and binds all intermediates more strongly.
over an adjacent (fcc or hollow) site with its hydrogen atom
The BE of atomic hydrogen (H*) on Pd(100) is −2.74 eV, only
pointing toward the surface.
0.04 eV stronger than its BE on Pd(111). H* preferentially binds Other potential intermediates include aquoxyl (OOHH*, an
to the fcc site on Pd(111) and the hollow site on Pd(100). Fur- isomer of H2O2* with both hydrogen atoms on the same oxygen
thermore, H* is expected to be mobile on Pd(100), as the BE of atom) and trihydrogen peroxide (HOOHH*). Similar to our
H* on a bridge site of Pd(100) is only 0.14 eV less stable than findings on Pd(111) (33), neither of these species is stable on
that on the hollow site. Similarly, on Pd(111), the BE of H* con- Pd(100), i.e., adsorption of aquoxyl and trihydrogen peroxide
strained to a bridge site is 0.14 eV less stable than that on the fcc site. structures on Pd(100) results in spontaneous decomposition to
Atomic oxygen (O*) has the same site preferences as H* on (O* + H2O*) and (OH* + H2O*), respectively.
both Pd(111) and Pd(100). The binding strength of O* on Pd(100) Table 1 provides the calculated O–O bond lengths for the
is −3.90 eV, which is 0.26 eV stronger than that on Pd(111). adsorbed dioxygen species (O2*, OOH*, and H2O2*) and the
Moreover, O* has a strong preference for the hollow site on corresponding values calculated in the gas phase. There is sig-
Pd(100), with the next best adsorption site (bridge) being less nificant expansion of the O–O bond in both O2* and OOH*
stable by 0.48 eV. On Pd(111), the next best adsorption site for upon adsorption, whereas the O–O bond length in H2O2* re-
O* is the hcp site, which is 0.15 eV less stable than O* binding to mains within 2% of its calculated gas-phase value. The larger
the fcc site. O–O bond expansion on Pd(100) compared with Pd(111) suggests
Hydroxyl (OH*) is often proposed to be the initial interme- a weaker O–O bond strength on the more open surface for all of
diate generated during H2O2 decomposition, resulting from the dioxygen species.
homolytic O–O bond cleavage in H2O2 (55). OH* binds most
stably to the bridge site on both Pd(111) and Pd(100) with the Activation Energy Barriers of Elementary Steps. The calculated ac-
O–H bond tilted away from the plane perpendicular to the sur- tivation energy barriers (Ea) and reaction energies (ΔE) are
face. The BE of OH* on Pd(100) is −2.43 eV, which is 0.40 eV reported with respect to reactant and product states at infinite
stronger than the BE on Pd(111). OH* binding at the hollow site
of Pd(100) is only 0.05 eV weaker than on the bridge site, which is
similar to the difference in energy between OH* binding at the fcc
site and bridge site on Pd(111) (0.08 eV).
Hydroperoxyl (OOH*) is considered an important intermediate
in the DSHP and was identified spectroscopically during the gas-
phase reaction of H2 and O2 on Au/TiO2 using inelastic neutron
scattering (56). OOH* binds through its nonhydrogenated oxy-
gen atom to a top site on Pd(111) with the hydroxyl group po-
sitioned over an adjacent bridge site and the O–H bond pointing
away from the surface, whereas OOH* binds through its non-
hydrogenated oxygen atom to a bridge site on Pd(100) with the Fig. 1. (A–G) Side and top-down views of the preferred binding sites for all
hydroxyl group positioned over an adjacent hollow site and the adsorbates on Pd(100). Blue spheres are hydrogen, red spheres are oxygen,
O–H bond pointing away from the surface. The binding energy and gray spheres are Pd atoms.

Plauck et al. PNAS | Published online March 22, 2016 | E1975


separation, unless stated otherwise. Table 2 summarizes the results lations show that O* recombination has prohibitively high bar-
for all elementary steps on Pd(111) and Pd(100). Transition-state riers—and is thermodynamically unfavorable—on both Pd(111)
geometries for the elementary steps are shown in Fig. 2. and Pd(100); the activation barrier exceeds 2 eV on Pd(111) and
O–O bond scission. At least one type of O–O bond scission step can 1 eV on Pd(100). These results are in agreement with temper-
be involved in the decomposition mechanism of H2O2. OH* and/ ature programmed desorption experiments for O2 desorption
or O* fragments are the direct products of O–O bond scission. from Pd(111) (62) and Pd(100) (63), in which the evolution of O2
Both Pd(111) and Pd(100) can readily break the O–O bond in from these Pd single crystals after preadsorbing O* at near-
H2O2* and OOH*, but there is a significant difference in the ambient temperatures is only observed at temperatures exceed-
ability of these facets to dissociate O2*. ing 600 K. Furthermore, in the context of the DSHP, Lunsford
H2O2* + * → OH* + OH*. H2O2* decomposes to two OH* on (64) used a mixture of [18O2 + 16O2] with H2 over a Pd/SiO2
Pd(100) with a barrier of 0.05 eV and a reaction energy of −2.29 eV. catalyst and observed that no H216O18O was formed, indicating
The corresponding barrier and reaction energy on Pd(111) are that O*/OH* recombination reactions were not relevant to
0.18 eV and −1.53 eV. This step occurs through a similar mech- H2O2 formation.
anism on both Pd(111) and Pd(100) whereby H2O2* rotates from Dehydrogenation. Because in this study we are only investigating
its most stable position on a top site to the transition state at which the decomposition of H2O2 (reaction 3) in the absence of H2 as
the O–O bond is elongated and both OH groups are bound to reactant, H* can only be derived from dehydrogenation of sur-
adjacent Pd atoms across a bridge site. However, at the transition face species through O–H bond scission. Barriers for O–H bond
state, the O–H bonds are on the same side of H2O2 molecule scission are generally lower on the more open Pd(100) facet
on Pd(100), whereas they are on different sides of the molecule on compared with those on Pd(111).
Pd(111). Following O–O bond scission, the two OH* relax to H2O2* + * → OOH* + H* and OOH* + * → O2* + H*. The O–H
bridge sites in their final coadsorbed state, stabilized through a bonds in H2O2* and OOH* are more difficult to break than the
hydrogen bond. O–O bond, based on the activation barriers in Table 2. On
OOH* + * → O* + OH*. OOH* decomposition to O* and OH* Pd(100), the O–H bond in H2O2* that is pointing toward the
on Pd(100) is nearly spontaneous with a barrier of 0.02 eV and a surface is cleaved over a bridge site. The activation barrier is
reaction energy of −1.83 eV—similar to the energetics on Pd(111). 0.44 eV, and the reaction energy is −0.29 eV.
The O–O bond scission occurs over a hollow site on Pd(100) For OOH* on Pd(100), the O–H bond is also broken over a
and a hcp site on Pd(111). bridge site. This breaking requires rotation of the O–H bond
O2* + * → O* + O*. The dissociation of O2* on Pd(100) occurs toward the surface, starting from the most stable OOH* geometry.
over a hollow site, whereby the O–O bond stretches from 1.41 Å The corresponding activation barrier and reaction energy for O–H
in the initial state to 1.90 Å in the transition state. The reaction bond cleavage in OOH* are 0.52 and −0.67 eV on Pd(100).
energy is −0.98 eV and the barrier is 0.30 eV on Pd(100), which H2O* + * → OH* + H* and OH* + * → O* + H*. Dehydrogenations
is 0.55 eV lower than the corresponding barrier on Pd(111). of H2O* and OH* require a larger activation energy compared
The reverse of O2* dissociation, O* recombination, represents with H2O2* and OOH* dehydrogenations. On Pd(100), the
a potential pathway for formation of the O2 product; this step barrier to break the O–H bond in H2O* is 0.67 eV, and the re-
has been proposed in a number of papers (59–61). Our calcu- action is thermoneutral. OH* dehydrogenation is more difficult

Table 2. Elementary steps considered for the decomposition of H2O2


Pd(111)† Pd(100)

No. Elementary step Ea, eV ΔE, eV Ea, eV ΔE, eV

1 H2O2 + * ↔ H2O2* — −0.32 — −0.36


2 H2O* ↔ H2O + * — 0.22 — 0.30
3 O2* ↔ O2 + * — 0.50 — 1.27
4 H* + H* ↔ H2 + 2* — 1.11 — 1.19
5 H2O2* + * ↔ OH* + OH* 0.18 −1.53 0.05 −2.29
6 OOH* + * ↔ O* + OH* 0.08 −1.50 0.02 −1.83
7 O2* + * ↔ O* + O* 0.85 −1.23 0.30 −0.98
8 OH* + * ↔ O* + H* 1.02 0.07 1.03 0.17
9 H2O* + * ↔ OH* + H* 1.10 0.37 0.67 0.00
10 OOH* + * ↔ O2* + H* 0.59 −0.20 0.52 −0.67
11 H2O2* + * ↔ OOH* + H* 0.62 0.05 0.44 −0.29
12 H2O* + O* ↔ OH* + OH* 0.33 0.33 0.00 −0.51
13 H2O2* + O* ↔ OOH* + OH* 0.04‡ −0.44 0.14‡ −0.87
14 H2O2* + OH* ↔ OOH* + H2O* 0.00 −0.16 0.00 −0.17
15 OOH* + O* ↔ O2* + OH* 0.00 −0.27 0.02 −0.81
16 OOH* + OH* ↔ O2* + H2O* 0.00 −0.38 0.00 −0.13
17 H2O2* + O2* ↔ OOH* + OOH* 0.20 −0.02 0.00 0.00

Energetics are reported with respect to either reactants/products at infinite separation (steps 1–11) or
coadsorbed for H-transfer reactions (steps 12–17) because these reactants/products are generally stabilized
through hydrogen bonding. Elementary steps are classified as follows: adsorption/desorption (steps 1–4);
O–O scission (steps 5–7); dehydrogenation (steps 8–11); and H transfer (steps 12–17). Ea and ΔE represent
the calculated activation energy and reaction energy in the forward direction. —, no activation barriers
are calculated for adsorption/desorption steps.

Data for steps 1–12 on Pd(111) are based on ref. 33.

Activation energy corresponds to breaking Pd–O bonds to lift O* from its preferred binding site (fcc or
fourfold hollow).

E1976 | www.pnas.org/cgi/doi/10.1073/pnas.1602172113 Plauck et al.


PNAS PLUS
Fig. 2. Side and top-down views of the transition-state geometries for O–O bond scission (A–C), dehydrogenation (D–G), and H-transfer (H–M) elementary
steps on Pd(100). Blue spheres are hydrogen, red spheres are oxygen, and gray spheres are Pd atoms. Elementary step numbers are in reference to Table 2.
Bond lengths (dx-y, in Å) refer to the bond being broken in the forward reaction, as written. Note that in I, the transition state for step 13 involves breaking

CHEMISTRY
Pd–O bonds to lift O* from its preferred binding site, followed by spontaneous H transfer from H2O2* to O*.

and has a barrier of 1.03 eV on Pd(100), and the reaction is slightly tential pathways for formation of both the H2O (H transfers to
endothermic. The transition state for O–H cleavage in both H2O* O*/OH*) and O2 (H transfers from H2O2*/OOH*, retaining the
and OH* occurs over a hollow site on Pd(100). O–O bond) products of H2O2 decomposition.
Hydrogen transfer. Formation and cleavage of O–H bonds in which H transfer to O. H2O2*, OOH*, and H2O* can all directly transfer
the Pd surface is directly involved have significant activation a H atom to O*; the activation energy barriers for these steps are
barriers (>0.4 eV). Alternatively, the Pd surface can mediate H 0.04, 0.00, and 0.33 eV on Pd(111), with reaction energies of
transfer between oxygenated intermediates—without involving −0.44, −0.27, and 0.33 eV. The corresponding activation energy
an explicit H* species. These elementary steps involve nearly barriers on Pd(100) are 0.14, 0.02, and 0.00 eV with significantly
spontaneous H transfer in the exothermic direction on both more exothermic reaction energies of −0.87, −0.81, and −0.51 eV.
Pd(111) and Pd(100) (Table 2). The activation energy barriers H transfer to OH. H2O2* and OOH* can also directly transfer a
and reaction energies in this section are reported with respect to H atom to OH*. We calculate that these steps proceed with
coadsorbed reactant and product states, because these states are nearly zero activation energy barrier on Pd(111) and Pd(100).
generally stabilized through hydrogen bonding (∼0.1–0.4 eV per The reaction energy for H transfer from H2O2* to OH* is weakly
hydrogen bond) with respect to the infinitely separated reactants exothermic [−0.16 eV on Pd(111) and −0.17 eV on Pd(100)].
and products. The reaction energy for H transfer from OOH* to OH* is more
The hydrogen atom is always transferred between O atoms exothermic on Pd(111) (−0.38 eV) than that on Pd(100) (−0.13 eV).
involved in hydrogen bonding in the most stable coadsorbed An additional H-transfer step that was explored is H transfer
configuration. Importantly, the H-transfer steps represent po- from H2O2* to O2*. This reaction is nearly thermoneutral on

Fig. 3. Schematic representation of reaction pathways for H2O2 decomposition on clean surfaces. The numbers by the black arrows correspond to the el-
ementary steps from Table 2. The overall reaction for each of the three complete mechanisms described in this figure is as follows: 2 H2O2 → 2 H2O + O2.

Plauck et al. PNAS | Published online March 22, 2016 | E1977


Fig. 4. Potential energy surfaces (thermochemistry only) for reaction pathways from Fig. 3 on clean Pd(111) and Pd(100) based on the DFT-derived ener-
getics. Energies are referenced to two H2O2 molecules in the gas phase. The “j” separating two adsorbates denotes infinite separation from each other. The
“(g)” denotes a gas-phase species. Insets compare O–H and O–O bond scission barriers in H2O2. “TS” denotes transition state.

Pd(111), with a reaction energy of −0.02 eV and an activation studies of H2O2 decomposition on Pd under similar conditions of
energy barrier of 0.20 eV. On Pd(100), the reaction is thermo- temperature and H2O2 concentration (13, 65). We also observed
neutral with negligible barrier. that the addition of O2 to the gas phase did not significantly
affect the decomposition rate of H2O2 up to O2 partial pressures
Catalytic Cycles and Potential Energy Surfaces. Based on the ele- of at least 37 psi, indicating negligible product inhibition over the
mentary steps above, several mechanisms are available to com- conditions studied. This finding is in agreement with the result of
plete the catalytic cycle on clean Pd facets, as summarized in Fig. Choudhary and Samanta (66), who observed only a minor dif-
3: “O*-assisted,” “OH*-assisted,” “O*+O*-recombination,” and ference in the reaction rate for H 2 O 2 decomposition over
“direct dehydrogenation.” A complete mechanism for direct Pd/Al2O3 (in the absence of H2) in a semibatch reactor when
dehydrogenation is not shown for simplicity, as both the DFT flowing either O2 or N2 through the liquid phase.
calculations and microkinetic modeling results suggest that direct Initial estimates for microkinetic model rate parameters are de-
dehydrogenation steps are characterized by much higher barriers rived from the DFT-calculated energetics. The reactor is simulated
and therefore not relevant under the reaction conditions explored as a continuous stirred tank reactor. The turnover frequencies for
in this study. The first step in all other pathways is H2O2 ad- H2O2 decomposition obtained from this model (i.e., the rate of
sorption followed by homolytic O–O bond cleavage to form two H2O2 converted per surface site) are used to calculate reaction
OH* species. The second H2O2* species can also adsorb and orders and apparent activation barrier for comparison with the
directly decompose to two OH*; these OH* species can then experimental data.
disproportionate to form H 2O* and O*—necessitating the The rates, reaction orders, and apparent activation barriers
recombination of two O* to form O2* (O*+O*-recombination predicted by the microkinetic model were initially in poor agree-
mechanism). Alternatively, two channels exist that bypass the ment with the experimental data when using purely DFT-derived
thermodynamically unfavorable and highly activated O* recom- parameters from Pd(111) or Pd(100). We subsequently used
bination step; both involve consecutive H-transfer steps from sensitivity analysis to identify the sensitive DFT-derived BEs of
the second H2O2 molecule to the O*/OH* fragments with re- surface species and transition-state energies. We then fit model-
tention of the original O–O bond in H 2O 2 (O*-assisted and predicted reaction rates to experimental rates (such that the re-
OH*-assisted mechanisms). sidual error is less than 20%) by modifying sensitive parameters,
The potential energy surfaces for all pathways are displayed in
Fig. 4 for both Pd(111) and Pd(100). Based on the DFT calcu-
lations alone, the O*-assisted and OH*-assisted pathways not Table 3. Reaction rates obtained from the kinetics experiments
only provide the most energetically efficient route to form the on Pd/spSiO2
products, but are also mutually competitive on both Pd(111)
Experimental rate,
and Pd(100). However, the deep potential wells associated
Run† Temperature, K y(O2) x(H2O2) mol·molPds−1·s−1
with the strongly bound O*/OH* fragments indicate that there
is a strong thermodynamic driving force to populate the surfaces 1 307 0.00 0.60 71.8
with O*/OH*—especially Pd(100). Therefore, the active surface 2 307 0.00 0.30 31.5
under reaction conditions may be partially covered by O*/OH*. 3 307 0.00 0.15 17.1
Note that on O*/OH*-modified surfaces, there is an increased 4 307 0.00 0.08 10.5
probability of H transfer from H2O2* before O–O bond scission; 5 297 0.00 0.15 7.9
the required O–O bond scission step may then occur in OOH* 6 285 0.00 0.15 3.4
rather than in H2O2*, slightly altering the succession of ele- 7 307 0.25 0.15 16.7
mentary steps proposed in Fig. 3. 8 307 0.13 0.15 16.2
9 307 0.06 0.15 16.9
Kinetics Experiments and Microkinetic Modeling. The results from †
our kinetics experiments are shown in Tables 3 and 4. The ex- y denotes mole fraction in the gas phase (balance Ar), and x denotes molarity
(moles per liter) in the liquid phase at the start of reaction. Pds denotes surface
perimentally determined activation energy barrier of 53.3 ± Pd atoms determined by CO uptake. Total pressure was 150 psi, and the stir rate
3.0 kJ/mol indicates that there is a significant variation in H2O2 was 1,200 rpm for all experiments. Reaction rates reported here correspond to
decomposition rate with reaction temperature under conditions the initial rates measured from conversion versus time data at <15% conversion,
relevant to the DSHP. The nearly first-order dependence on which was approximately linear in this regime. Each reported rate represents
concentration of H2O2 is in agreement with other experimental the average from at least two repeated sets of conversion versus time data.

E1978 | www.pnas.org/cgi/doi/10.1073/pnas.1602172113 Plauck et al.


PNAS PLUS
On the other hand, DFT-derived parameters on Pd(100) de-
viate significantly (>0.35 eV for OH*, OOH*, and O2*) from
this O*-coverage solution parameter set. Pd(100) binds inter-
mediates too strongly, and the low predicted O* coverage is not
expected to destabilize intermediates on Pd(100) sufficiently for
consistency with the O*-coverage solution parameters.
The O*-coverage solution predicts that the dominant reaction
pathway is the O*-assisted pathway shown in Fig. 3 (sequence of
elementary steps from Table 2: 1, 5, −12, 2, 1, 13, 16, 2, 3), with
some reaction flux through the parallel OH*-assisted pathway
(sequence of elementary steps from Table 2: 1, 5, 1, 14, 2, 16, 2, 3).
The rate of O* recombination to form O2* is negligible, and
dehydrogenation reactions are also inactive (atomic H is only
transferred between surface intermediates). The kinetic relevance
of each elementary step was also analyzed using Campbell’s de-
gree of rate control (69, 70):
 
ki ∂r
XRC,i = ,
r ∂ki Ki,eq ,kj

where ki and Ki,eq are the rate constant and equilibrium constant
for step i, and r is the overall reaction rate. O–O bond scission in
Fig. 5. (A and C ) Parity plots of experimental and model-predicted re-
H2O2* (step 5 of Table 2) carries the highest degree of rate
action rates for H2O2 decomposition. Refer to Table 3 for reaction condi-
control over the reaction conditions examined, shown in Table

CHEMISTRY
tions at each of the points in A and C. Pds denotes surface Pd atoms
determined by CO uptake. Red points are varying temperature; blue points 5; the remaining rate control is distributed between the subse-
are varying O2 partial pressure; black points are varying feed concentra- quent H-transfer reactions.
tion of H2O2. (B and D) Microkinetic model-predicted surface coverages of OH*-coverage solution. Using the DFT calculations on clean Pd(100)
the most abundant surface intermediates (0.15 M H2O2 in H2O feed with to derive initial estimates of parameters, a second solution was
150 psi Ar in gas phase). Plots on the Left refer to the O*-coverage solu- identified that also gave agreement with the experimental data set
tion, and plots on the Right refer to the OH*-coverage solution obtained (Fig. 5C and Table 4). In this case, the model-predicted surface
from the microkinetic model. Insets in B and D provide graphical repre- coverage is ∼0.5 ML of OH* (Fig. 5D) and is therefore not self-
sentations of the nature of the active sites as concluded through this study
consistent with the clean Pd(111) and Pd(100) surface models used
[nearly clean Pd(111) and OH*-modified Pd(100)]. Blue spheres are hy-
drogen, red spheres are oxygen, and gray spheres are Pd atoms.
in the DFT calculations. To ensure a solution self-consistent in
coverage, we recalculated the binding energies of surface inter-
mediates and the activation energy barriers for steps carrying sig-
constraining the adjustment of parameters such that (i) the de- nificant reaction flux (as predicted by the OH*-coverage solution)
viation between DFT-derived BEs and activation barriers on a in the presence of 0.5 ML of OH* spectators, i.e., two OH* were
given Pd facet and the corresponding values from the microkinetic added to the unit cell and allowed to relax in the DFT calculations.
model should be within the ∼0.1- to 0.2-eV error bars generally The DFT-derived parameter set for the OH*-modified
attributed to DFT calculations (67), and (ii) the adsorbate cov- Pd(100) surface is found to be in close agreement with the adjusted
erage used in the DFT calculations should be consistent with the parameter set from the OH*-coverage solution (BEs shown in
coverage predicted by the microkinetic model (that is, the solution Table 6, with further details in Supporting Information). The DFT
should be self-consistent with respect to coverage). calculations show that 0.5 ML of OH* destabilizes most inter-
Next, we present the results from two model solutions that mediates and transition states investigated on Pd(100) relative to
satisfy the above criteria, but differ in both the coverage and identity the clean Pd(100) calculations. The binding energies of O*,
of the most abundant surface intermediate, and are denoted as the OH*, O2*, and OOH* are weakened by >0.5 eV, whereas the
“O*-coverage solution” and the “OH*-coverage solution.” binding energies of H*, H2O*, and H2O2* are not significantly
O*-coverage solution. Fig. 5A shows a parity plot comparing the affected. In addition, the activation energy barriers for O–O
experimental H2O2 decomposition rates with the microkinetic bond breaking in OOH* and H2O2* increase by 0.39 and 0.56 eV,
model predictions for the O*-coverage solution [initial estimates respectively. Activation energy barriers for H transfer from
of parameters are derived from DFT calculations on clean H2O2* or OOH* to OH* or O* remain small (<0.2 eV). The
Pd(111), and Supporting Information provides details of the pa- maximum deviation in binding energy or activation barrier be-
rameter adjustments used to obtain this solution]. There is good tween the OH*-coverage solution and DFT calculations on
agreement between model-predicted and experimental reaction OH*-modified Pd(100) is a 0.18-eV decrease in the activation
rates, and the microkinetic model is able to accurately reproduce
the experimental activation barrier and reaction orders (Table 4).
The microkinetic model predictions for the surface coverage Table 4. Experimental and microkinetic model-predicted
for the most abundant intermediate, O*, range from 0.13 to reaction orders and apparent activation energy barriers (Eapp)
0.16 ML (Fig. 5B), with the remaining sites being vacant. Mar- O*-coverage OH*-coverage
ginal changes to the clean surface energetics are expected from Species Experiment† solution solution
adsorbate–adsorbate interactions at such low O* coverage (68),
and furthermore this adjusted parameter set compares well with H2O2 0.92 ± 0.08 1.00 0.97
the DFT-derived parameters on clean Pd(111); the maximum O2 −0.01 ± 0.03 −0.01 0.00
deviation in binding energy or activation barrier is a 0.24 eV Eapp, kJ/mol 53.3 ± 3.0 53.1 56.6
destabilization of the binding energy of O2* on Pd(111). Reported experimental error is the SE from linear regression.
Therefore, a partially O*-covered Pd(111) surface is a plausible †
Subsequent catalyst batches yielded reaction orders and an apparent bar-
representation of the active site for H2O2 decomposition. rier within ∼15% of the values reported here.

Plauck et al. PNAS | Published online March 22, 2016 | E1979


Table 5. Degree of rate control (XRC) calculated for kinetically decomposition activity must focus on tuning surface reactivity to-
relevant reaction steps for reaction condition 3 of Table 3 ward the O–O bond. Retention of the O–O bond in dioxygen
XRC, O*- XRC, OH*- species is generally acknowledged to be a key factor in the se-
coverage coverage lective synthesis of H2O2 by the DSHP both in theoretical (33,
No. Elementary step solution solution 73, 74) and experimental literature (8, 12, 64, 75), and our
results here quantitatively highlight this as the central param-
5 H2O2* + * ↔ OH* + OH* 0.59 0.00 eter governing the subsequent H2O2 decomposition activity on
6 OOH* + * ↔ O* + OH* 0.02 0.55 the Pd surface.
13–16 (H transfers) 0.40 0.43 In view of the aforementioned findings, reduced Pd nano-
XRC is given for both the O*-coverage solution and the OH*-coverage particles would be expected to be an ineffective catalyst for the
solution at this experimental condition. Elementary step numbers (No.) are DSHP due to high activity of Pd for O–O bond breaking [0.18-
in reference to Table 2. XRC for the H transfers is the sum over all steps listed. eV and 0.05-eV barrier to break O–O bond in H2O2 on clean
Pd(111) and Pd(100), respectively]. Extensive surface poisoning
may be necessary to inhibit H2O2 decomposition on Pd, which
barrier for O–O breaking in OOH*. Therefore, OH*-modified our results suggest can readily occur on surface facets of sup-
Pd(100) also appears to be a feasible representation of the active ported Pd nanoparticles that are generally in high abundance
site for H2O2 decomposition on Pd. [the (111) and (100) facets]. Indeed, the experimentally measured
The dominant reaction pathways predicted for the OH*-cover- H2O2 decomposition activity of supported Pd nanoparticles can be
age solution are shown in Fig. 6. At high OH* coverage, immediate effectively quenched upon adding halides (along with acids,
H transfer from H2O2* to OH* is predicted to be nearly quasie- whose role may partly be to facilitate halide adsorption) to the
quilibrated (step 14 of Table 2). The O–O bond breaks in OOH*, reaction medium, often at Pd:halide atomic ratios close to or
and this step carries the highest degree of rate control (Table 5). exceeding 1:1 (66, 75–78). Unfortunately, there are limited fun-
Hydrogen transfers from OOH* to OH* (step 16 of Table 2) and damental studies that examine the halide coverage necessary to
from OOH* to O* (step 15 of Table 2) carry the remaining re- achieve this effect.
action flux to form O2* and H2O*; hydrogen transfer from H2O* Some of the most successful experimental catalysts to-date for
to O* (step 12 of Table 2) is also nearly quasiequilibrated. the DSHP are based on alloys of Pd with Au, on which the sub-
sequent decomposition reactions of H2O2 are partially or com-
Discussion pletely inhibited. DFT calculations indicate that dilution of Pd
The microkinetic modeling results suggest that both the close- surfaces with Au significantly increases barriers for O–O bond
packed Pd(111) and more open Pd(100) facets can contribute to scission (79). However, experiments demonstrate that Au itself is
the total H2O2 decomposition activity. Furthermore, on both Pd generally an ineffective catalyst for the DSHP and gives slow rates
facets, all reaction flux is predicted to go through an O–O bond (14, 80), likely due to the significant activation energy barrier re-
breaking step (in either H2O2* or OOH*), followed by successive quired to dissociate H2 on Au (81) and weak adsorption of O2
H-transfer steps to O*/OH* adsorbates. The relevant surface (33). Promising search directions for improved DSHP catalysts
coverage of O*/OH* is then a function of the ability of the Pd may include bimetallic systems in which an active component (e.g.,
surfaces to generate the O*/OH* fragments through O–O bond Pd) is effectively isolated in a relatively inert component (e.g., Au)
breaking [which can vary strongly with surface coverage, as seen resulting in reduced O–O bond breaking capacity but retention of
from calculations on the OH*-modified Pd(100)], and the H2 dissociation (82) capacity; similar catalysts have proven very
availability of H-donating species (H2O2* and OOH*) to reduce effective for the electrocatalytic synthesis of H2O2 (83, 84).
O*/OH* to H2O* through the rapid H-transfer reactions. This Last, the presence of H2 in the reactor feed (not considered in
mechanism is comparable to the redox mechanism discussed the present work) has been shown to enhance the overall H2O2
in refs. 65 and 71. The direct dehydrogenation and O*+O*- decomposition activity over Pd-based catalysts (66, 85). Choudhary
recombination pathways (Fig. 3) are predicted to be inactive over and Samanta (66) observed that, on unmodified Pd, H2 both in-
all experimental conditions examined. creases the H2O2 decomposition rate and consumes H2O2 through
Although H-transfer steps carry some degree of rate control in complete hydrogenation to H2O (reaction 4). Moreover, although
the microkinetic model solutions, the DFT calculations show adding chloride or bromide to an acidified reaction medium can
that the activation barriers for these steps are nearly insensitive quench H2O2 decomposition on Pd, H2O2 hydrogenation activity
to the surface structure of the Pd substrate. Interestingly, ex- remains (66); this observation may indicate significant differences
perimentally measured activation energy barriers for the gas- in the active site(s) and rate-controlling step(s) responsible for
phase H-transfer reactions of H2O2 or OOH• to OH• or O• H2O2 decomposition versus H2O2 hydrogenation—although
radicals are also readily accessible (<0.2 eV) around room tem- O–O bond breaking is required in both reactions. Tentative
perature (72). The action of the metal substrate is then to generate
the O*/OH* species through O–O bond breaking, and localize the
H-transfer event to the surface. H2O2*, OOH*, and H2O* are Table 6. DFT-calculated BEs of adsorbed species on Pd(100) in
mobile on both Pd(111) and Pd(100) based on the small differ- the presence of 0.5 ML of OH*, and comparison with the BEs in
ences in binding energies among the available binding sites and the OH*-coverage solution
therefore can diffuse across the Pd surface to find—and react Pd(100) + 0.5 ML OH* OH*-coverage solution
with—the O*/OH* fragments. Additionally, we note that O*
strongly prefers the threefold and fourfold hollow sites on Pd(111) Species Adsorption site BE, DFT, eV BE, eV
and Pd(100), respectively; and O* must be lifted slightly from its
H* Hollow −2.56 −2.56
preferred binding site to accept a H atom. This behavior is reflected
O* Hollow −3.16 −2.99
in the low activation energy barrier generally calculated for H
OH* Top-tilted −1.68 −1.83
transfers to O*, compared with the virtually zero barrier calculated
OOH* Hollow-upright −0.61 −0.51
for H transfers to OH*—which is more accessible at its most
H2O* Top −0.26 −0.32
favorable binding site (bridge site) on Pd(111) and Pd(100).
H2O2* Top −0.36 −0.46
Breaking of the O–O bond in either H2O2* or OOH* carries the
O2* Hollow −0.14 −0.14
majority of rate control, suggesting that strategies to reduce H2O2

E1980 | www.pnas.org/cgi/doi/10.1073/pnas.1602172113 Plauck et al.


PNAS PLUS
Fig. 6. Dominant reaction pathways predicted by the microkinetic model for the OH*-coverage solution. The numbers by the black arrows correspond to the
elementary steps numbers given in Table 2.

explanations addressing the role of H2 have been proposed, adjusted parameters are consistent with the DFT-derived pa-
such as maintaining the Pd surface in the reduced state (59) for rameters on a Pd(100) surface with OH* spectators. Therefore,
facile O–O bond breaking. Additionally, direct hydrogenation the microkinetic model suggests that both Pd(111) and Pd(100)
of H2O2 was shown to be a highly activated step using DFT cal- can contribute to H2O2 decomposition activity. Experimental
culations (86). The influence of subsurface hydrogen or even identification of dominant surface species during H2O2 de-
Pd-hydrides may also be relevant (87). The next stage of this work composition on Pd might be realized by in situ X-ray photo-
is to investigate the mechanistic role of H2 in accelerating H2O2 electron spectroscopy measurements in a similar manner to work
decomposition on Pd, in addition to probing the nature of the performed on Pt for the oxygen reduction reaction (88).
active site(s) in the presence of H2. Consistent with the insights from DFT calculations, the domi-
Conclusions nant reaction pathways involve O–O bond breaking in either
H2O2* or OOH* followed by H-transfer reactions between vari-
Both the close-packed (111) and more open (100) facets can
represent the active site for SiO2-supported Pd nanoparticles. The ous reaction intermediates. Breaking of the O–O bond is identi-
DFT results show that O–O bond scission is facile on both Pd fied as the key parameter governing H2O2 decomposition activity,
facets, such that O* and OH* intermediates are readily produced. because this step carries the highest degree of rate control.
Furthermore, H2O2* and OOH* can reduce O* and OH* to H2O
ACKNOWLEDGMENTS. A.P. thanks Assistant Professor Fuat E. Celik for his

CHEMISTRY
through thermodynamically driven H-transfer reactions, liberating initial guidance with the density functional theory calculations, as well as
O2. The alternative step to produce O2 (recombination of O*) is Yunhai Bai and Benjamin Chen for their comments on the article. This material
both thermodynamically and kinetically unfavorable. In addition, is based on work supported as part of a Dow Chemical Company University
steps involving dehydrogenation through direct O–H bond Partner Initiative with the University of Wisconsin–Madison, under Dow Agree-
ment 235744C. Computational time was used at supercomputing resources
cleavage over Pd are less favored than the H-transfer steps. located at Environmental Molecular Sciences Laboratory (EMSL), a national
Microkinetic models based on two parameter sets are able to scientific user facility at Pacific Northwest National Laboratory (PNNL); the
describe the experimental data for a SiO2-supported Pd catalyst: Center for Nanoscale Materials (CNM) at Argonne National Laboratory (ANL);
the first set corresponds to a Pd surface partially covered in <0.2 and the National Energy Research Scientific Computing Center (NERSC). EMSL is
sponsored by the Department of Energy’s Office of Biological and Environmental
ML of O*, and these adjusted parameters are consistent with the Research located at PNNL. CNM and NERSC are supported by the US Department
DFT-derived parameters on clean Pd(111); the second set cor- of Energy, Office of Science, under Contracts DE-AC02-06CH11357 and DE-AC02-
responds to a Pd surface covered in ∼0.5 ML of OH*, and these 05CH11231, respectively.

1. Centi G, Cavani F, Trifirò F (2001) Selective Oxidation by Heterogeneous Catalysis 17. Edwards JK, et al. (2012) The effect of heat treatment on the performance and
(Springer, New York). structure of carbon-supported Au-Pd catalysts for the direct synthesis of hydrogen
2. Walsh PB (1991) Hydrogen peroxide: Innovations in chemical pulp bleaching. Tappi peroxide. J Catal 292:227–238.
Journal 74(1):81–83. 18. Deguchi T, Yamano H, Iwamoto M (2015) Kinetic and mechanistic studies on direct
3. Strukul G (1992) Catalytic Oxidations with Hydrogen Peroxide as Oxidant (Kluwer H2O2 synthesis from H2 and O2 catalyzed by Pd in the presence of H+ and Br− in water:
Academic, Dordrecht, The Netherlands). A comprehensive paper. Catal Today 248:80–90.
4. Nijhuis TA, Makkee M, Moulijn JA, Weckhuysen BM (2006) The production of propene 19. Monnier JR (2001) The direct epoxidation of higher olefins using molecular oxygen.
oxide: Catalytic processes and recent developments. Ind Eng Chem Res 45(10):3447–3459. Appl Catal A Gen 221(1-2):73–91.
5. Centi G, Perathoner S, Abate S (2009) Direct synthesis of hydrogen peroxide: Recent 20. Laufer W, Meiers R, Holderich W (1999) Propylene epoxidation with hydrogen peroxide
advances. Modern Heterogeneous Oxidation Catalysis Design, Reactions and over palladium containing titanium silicalite. J Mol Catal A Chem 141(1-3):215–221.
Characterization, ed Mizuno N (Wiley-VCH, Weinheim, Germany), pp 253–287. 21. Meiers R, Holderich WF (1999) Epoxidation of propylene and direct synthesis of hy-
6. Madden S, Ross T (2005) Degussa/Headwaters Pilot Plant Validates New Tech- drogen peroxide by hydrogen and oxygen. Catal Lett 59(2-4):161–163.
nology (Business Wire, New York). Available at www.businesswire.com/news/home/ 22. Jenzer G, Mallat T, Maciejewski M, Eigenmann F, Baiker A (2001) Continuous epoxi-
20050309005685/en/DegussaHeadwaters-Pilot-Plant-Validates-Technology. Accessed dation of propylene with oxygen and hydrogen on a Pd-Pt/TS-1 catalyst. Appl Catal A
December 3, 2014. Gen 208(1-2):125–133.
7. García-Serna J, et al. (2014) Engineering in direct synthesis of hydrogen peroxide: Tar- 23. Popa C, et al. (2015) Structure of palladium nanoparticles under oxidative conditions.
gets, reactors and guidelines for operational conditions. Green Chem 16(5):2320–2343. Phys Chem Chem Phys 17(3):2268–2273.
8. Samanta C (2008) Direct synthesis of hydrogen peroxide from hydrogen and oxygen: 24. Lim B, et al. (2009) Shape-controlled synthesis of Pd nanocrystals in aqueous solutions.
An overview of recent developments in the process. Appl Catal A Gen 350(2):133–149. Adv Funct Mater 19(2):189–200.
9. Campos-Martin JM, Blanco-Brieva G, Fierro JLG (2006) Hydrogen peroxide synthesis: 25. Hammer B, Hansen LB, Norskov JK (1999) Improved adsorption energetics within
An outlook beyond the anthraquinone process. Angew Chem Int Ed Engl 45(42):6962–6984. density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys Rev
10. Choudhary VR, Samanta C, Choudhary TV (2006) Factors influencing decomposition B 59(11):7413–7421.
of H2O2 over supported Pd catalyst in aqueous medium. J Mol Catal A Chem 260(1-2): 26. Greeley J, Nørskov JK, Mavrikakis M (2002) Electronic structure and catalysis on metal
115–120. surfaces. Annu Rev Phys Chem 53:319–348.
11. Abate S, et al. (2013) On the nature of selective palladium-based nanoparticles on nitrogen- 27. Perdew JP, et al. (1992) Atoms, molecules, solids, and surfaces: Applications of the
doped carbon nanotubes for the direct synthesis of H2O2. ChemCatChem 5(7):1899–1905. generalized gradient approximation for exchange and correlation. Phys Rev B
12. Edwards JK, Freakley SJ, Carley AF, Kiely CJ, Hutchings GJ (2014) Strategies for de- Condens Matter 46(11):6671–6687.
signing supported gold-palladium bimetallic catalysts for the direct synthesis of hy- 28. White JA, Bird DM (1994) Implementation of gradient-corrected exchange-correla-
drogen peroxide. Acc Chem Res 47(3):845–854. tion potentials in Car-Parrinello total-energy calculations. Phys Rev B Condens Matter
13. Pospelova TA, Kobozev NI (1961) Palladium-catalysed synthesis of hydrogen peroxide 50(7):4954–4957.
from the elements. III. Active centres for the catalytic decomposition of hydrogen 29. Vanderbilt D (1990) Soft self-consistent pseudopotentials in a generalized eigenvalue
peroxide on palladium. Russ J Phys Chem 35:584–587. formalism. Phys Rev B Condens Matter 41(11):7892–7895.
14. Edwards JK, et al. (2009) Switching off hydrogen peroxide hydrogenation in the di- 30. Kresse G, Furthmuller J (1996) Efficiency of ab-initio total energy calculations for metals
rect synthesis process. Science 323(5917):1037–1041. and semiconductors using a plane-wave basis set. Comput Mater Sci 6(1):15–50.
15. Ouyang L, et al. (2015) The origin of active sites for direct synthesis of H2O2 on Pd/TiO2 31. Chadi DJ, Cohen ML (1973) Special points in the Brillouin zone. Phys Rev B 8(12):
catalysts: Interfaces of Pd and PdO domains. J Catal 321:70–80. 5747–5753.
16. Arrigo R, et al. (2014) Dynamics of palladium on nanocarbon in the direct synthesis of 32. Monkhorst HJ, Pack JD (1976) Special points for Brillouin-zone integrations. Phys Rev
H2O2. ChemSusChem 7(1):179–194. B 13(12):5188–5192.

Plauck et al. PNAS | Published online March 22, 2016 | E1981


33. Ford DC, Nilekar AU, Xu Y, Mavrikakis M (2010) Partial and complete reduction of O2 62. Zheng G, Altman EI (2000) The oxidation of Pd(111). Surf Sci 462(1-3):151–168.
by hydrogen on transition metal surfaces. Surf Sci 604(19-20):1565–1575. 63. Zheng G, Altman EI (2002) The oxidation mechanism of Pd(100). Surf Sci 504(1-3):
34. Neugebauer J, Scheffler M (1992) Adsorbate-substrate and adsorbate-adsorbate in- 253–270.
teractions of Na and K adlayers on Al(111). Phys Rev B Condens Matter 46(24): 64. Lunsford JH (2003) The direct formation of H2O2 from H2 and O2 over palladium
16067–16080. catalysts. J Catal 216(1-2):455–460.
35. Bengtsson L (1999) Dipole correction for surface supercell calculations. Phys Rev B 65. Bianchi G, Mazza F, Mussini T (1962) Catalytic decomposition of acid hydrogen per-
59(19):12301–12304. oxide solutions on platinum, iridium, palladium and gold surfaces. Electrochim Acta
36. Mermin ND, Ashcroft NW (1976) Solid State Physics (Harcourt College Publishers, Fort 7(4):457–473.
Worth, TX). 66. Choudhary VR, Samanta C (2006) Role of chloride or bromide anions and protons for
37. Henkelman G, Uberuaga BP, Jonsson H (2000) A climbing image nudged elastic band promoting the selective oxidation of H2 by O2 to H2O2 over supported Pd catalysts in
method for finding saddle points and minimum energy paths. J Chem Phys 113(22): an aqueous medium. J Catal 238(1):28–38.
9901–9904. 67. Singh S, et al. (2014) Formic acid decomposition on Au catalysts: DFT, microkinetic
38. Henkelman G, Jonsson H (2000) Improved tangent estimate in the nudged elastic modeling, and reaction kinetics experiments. AIChE J 60(4):1303–1319.
band method for finding minimum energy paths and saddle points. J Chem Phys 68. Grabow LC, Hvolbaek B, Norskov JK (2010) Understanding trends in catalytic activity:
113(22):9978–9985. The effect of adsorbate-adsorbate interactions for CO oxidation over transition
39. Greeley J, Mavrikakis M (2003) A first-principles study of surface and subsurface H on metals. Top Catal 53(5-6):298–310.
and in Ni(111): Diffusional properties and coverage-dependent behavior. Surf Sci 69. Campbell CT (2001) Finding the rate-determining step in a mechanism: Comparing
540(2-3):215–229. DeDonder relations with the “degree of rate control.” J Catal 204(2):520–524.
40. Stober W, Fink A, Bohn E (1968) Controlled growth of monodisperse silica spheres in 70. Dumesic JA (1999) Analyses of reaction schemes using De Donder relations. J Catal
the micron size range. J Colloid Interface Sci 26(1):62–69. 185(2):496–505.
41. Alba-Rubio AC, et al. (2014) Pore structure and bifunctional catalyst activity of 71. Voloshin Y, Manganaro J, Lawal A (2008) Kinetics and mechanism of decomposition
overlayers applied by atomic layer deposition on copper nanoparticles. ACS Catal 4(5): of hydrogen peroxide over Pd/SiO2 catalyst. Ind Eng Chem Res 47(21):8119–8125.
1554–1557. 72. Atkinson R, et al. (2004) Evaluated kinetic and photochemical data for atmospheric
42. Zou WQ, Gonzalez RD (1992) The preparation of silica supported Pd catalysts: The chemistry: Volume I—gas phase reactions of Ox, HOx, NOx and SOx species. Atmos
effect of pretreatment variables on particle size. Catal Lett 12(1-3):73–86. Chem Phys 4:1461–1738.
43. Spiewak BE, Shen J, Dumesic JA (1995) Microcalorimetric studies of CO and H2 ad- 73. Grabow LC, Hvolbaek B, Falsig H, Norskov JK (2012) Search directions for direct H2O2
sorption on nickel powders promoted with potassium and cesium. J Phys Chem synthesis catalysts starting from Au12 nanoclusters. Top Catal 55(5-6):336–344.
99(49):17640–17644. 74. Rankin RB, Greeley J (2012) Trends in selective hydrogen peroxide production on
44. Gokhale AA, Kandoi S, Greeley JP, Mavrikakis M, Dumesic JA (2004) Molecular-level transition metal surfaces from first principles. ACS Catal 2(12):2664–2672.
descriptions of surface chemistry in kinetic models using density functional theory. 75. Burch R, Ellis PR (2003) An investigation of alternative catalytic approaches for the
Chem Eng Sci 59(22-23):4679–4691. direct synthesis of hydrogen peroxide from hydrogen and oxygen. Appl Catal B 42(2):
45. Grabow LC, Mavrikakis M (2011) Mechanism of methanol synthesis on Cu through 203–211.
CO2 and CO hydrogenation. ACS Catal 1(4):365–384. 76. Melada S, Pinna F, Strukul G, Perathoner S, Centi G (2006) Direct synthesis of H2O2 on
46. Grabow LC, Gokhale AA, Evans ST, Dumesic JA, Mavrikakis M (2008) Mechanism of monometallic and bimetallic catalytic membranes using methanol as reaction me-
the water gas shift reaction on Pt: First principles, experiments, and microkinetic dium. J Catal 237(2):213–219.
modeling. J Phys Chem C 112(12):4608–4617. 77. Ntainjua N E, et al. (2009) Effect of halide and acid additives on the direct synthesis of
47. Tessier A, Forst W (1974) Mechanism of hydrogen-peroxide pyrolysis. Can J Chem hydrogen peroxide using supported gold-palladium catalysts. ChemSusChem 2(6):
52(5):794–797. 575–580.
48. Takagi J, Ishigure K (1985) Thermal decomposition of hydrogen peroxide and its ef- 78. Liu Q, Lunsford JH (2006) Controlling factors in the direct formation of H2O2 from H2
fect on reactor water monitoring of boiling water reactors. Nucl Sci Eng 89(2): and O2 over a Pd/SiO2 catalyst in ethanol. Appl Catal A Gen 314(1):94–100.
177–186. 79. Ham HC, Hwang GS, Han J, Nam SW, Lim TH (2009) On the role of Pd ensembles in
49. Lousada CM, Johansson AJ, Brinck T, Jonsson M (2012) Mechanism of H2O2 de- selective H2O2 formation on PdAu alloys. J Phys Chem C 113(30):12943–12945.
composition on transition metal oxide surfaces. J Phys Chem C 116(17):9533–9543. 80. Edwards JK, et al. (2005) Direct synthesis of hydrogen peroxide from H2 and O2 using
50. Hiroki A, Laverne JA (2005) Decomposition of hydrogen peroxide at water-ceramic Au-Pd/Fe2O3 catalysts. J Mater Chem 15(43):4595–4600.
oxide interfaces. J Phys Chem B 109(8):3364–3370. 81. Hammer B, Norskov JK (1995) Why gold is the noblest of all the metals. Nature
51. Lin SS, Gurol MD (1998) Catalytic decomposition of hydrogen peroxide on iron oxide: 376(6537):238–240.
Kinetics, mechanism, and implications. Environ Sci Technol 32(10):1417–1423. 82. Tierney HL, Baber AE, Kitchin JR, Sykes ECH (2009) Hydrogen dissociation and spillover
52. Haber F, Weiss J (1934) The catalytic decomposition of hydrogen peroxide by iron on individual isolated palladium atoms. Phys Rev Lett 103(24):246102.
salts. Proc R Soc Lond A Math Phys Sci 147(861):332–351. 83. Jirkovský JS, et al. (2011) Single atom hot-spots at Au-Pd nanoalloys for electro-
53. Kwan WP, Voelker BM (2002) Decomposition of hydrogen peroxide and organic catalytic H2O2 production. J Am Chem Soc 133(48):19432–19441.
compounds in the presence of dissolved iron and ferrihydrite. Environ Sci Technol 84. Verdaguer-Casadevall A, et al. (2014) Trends in the electrochemical synthesis of H2O2:
36(7):1467–1476. Enhancing activity and selectivity by electrocatalytic site engineering. Nano Lett 14(3):
54. Tian P, Ouyang L, Xu X, Xu J, Han Y-F (2013) Density functional theory study of direct 1603–1608.
synthesis of H2O2 from H2 and O2 on Pd(111), Pd(100), and Pd(110) surfaces. Chin J 85. Landon P, et al. (2003) Direct synthesis of hydrogen peroxide from H2 and O2 using Pd
Catal 34(5):1002–1012. and Au catalysts. Phys Chem Chem Phys 5(9):1917–1923.
55. Lousada CM, Johansson AJ, Brinck T, Jonsson M (2013) Reactivity of metal oxide 86. Li J, Staykov A, Ishihara T, Yoshizawa K (2011) Theoretical study of the decomposition
clusters with hydrogen peroxide and water—a DFT study evaluating the performance and hydrogenation of H2O2 on Pd and Au@Pd surfaces: Understanding toward high
of different exchange-correlation functionals. Phys Chem Chem Phys 15(15): selectivity of H2O2 synthesis. J Phys Chem C 115(15):7392–7398.
5539–5552. 87. Todorovic R, Meyer RJ (2011) A comparative density functional theory study of the
56. Sivadinarayana C, Choudhary TV, Daemen LL, Eckert J, Goodman DW (2004) The direct synthesis of H2O2 on Pd, Pt and Au surfaces. Catal Today 160(1):242–248.
nature of the surface species formed on Au/TiO2 during the reaction of H2 and O2: An 88. Casalongue HS, et al. (2013) Direct observation of the oxygenated species during
inelastic neutron scattering study. J Am Chem Soc 126(1):38–39. oxygen reduction on a platinum fuel cell cathode. Nat Commun 4(2817):1–6.
57. Long R, et al. (2013) Surface facet of palladium nanocrystals: A key parameter to the 89. Madon RJ, et al. (2011) Microkinetic analysis and mechanism of the water gas shift
activation of molecular oxygen for organic catalysis and cancer treatment. J Am Chem reaction over copper catalysts. J Catal 281(1):1–11.
Soc 135(8):3200–3207. 90. Hwang H, Dasgupta PK (1985) Thermodynamics of the hydrogen peroxide-water
58. Simmons GW, Wang YN, Marcos J, Klier K (1991) Oxygen adsorption on Pd(100) system. Environ Sci Technol 19(3):255–258.
surface: Phase transformations and surface reconstruction. J Phys Chem 95(11): 91. Michaelides A (2006) Density functional theory simulations of water-metal interfaces:
4522–4528. Waltzing waters, a novel 2D ice phase, and more. Appl Phys A Mater Sci Process 85(4):
59. Ntainjua NE, et al. (2008) The role of the support in achieving high selectivity in the 415–425.
direct formation of hydrogen peroxide. Green Chem 10(11):1162–1169. 92. Sha Y, Yu TH, Liu Y, Merinov BV, Goddard WA (2010) Theoretical study of solvent effects
60. Choudhary VR, Samanta C, Jana P (2007) Decomposition and/or hydrogenation of on the platinum-catalyzed oxygen reduction reaction. J Phys Chem Lett 1(5):856–861.
hydrogen peroxide over Pd/Al2O3 catalyst in aqueous medium: Factors affecting the 93. Carrasco J, Santra B, Klimes J, Michaelides A (2011) To wet or not to wet? Dispersion
rate of H2O2 destruction in presence of hydrogen. Appl Catal A Gen 332(1):70–78. forces tip the balance for water ice on metals. Phys Rev Lett 106(2):026101.
61. Paunovic V, Ordomsky VV, Sushkevich VL, Schouten JC, Nijhuis TA (2015) Direct syn- 94. Santra B, et al. (2013) On the accuracy of van der Waals inclusive density-functional
thesis of hydrogen peroxide over Au-Pd catalyst: The effect of co-solvent addition. theory exchange-correlation functionals for ice at ambient and high pressures.
ChemCatChem 7(7):1161–1176. J Chem Phys 139(15):154702.

E1982 | www.pnas.org/cgi/doi/10.1073/pnas.1602172113 Plauck et al.

Вам также может понравиться