Performance of DFT+ U Approaches in The Study of Catalytic Materials
Performance of DFT+ U Approaches in The Study of Catalytic Materials
Performance of DFT+ U Approaches in The Study of Catalytic Materials
pubs.acs.org/acscatalysis
units, typically a carrier and an active phase. For many forcing electron occupancy of particular orbitals that, for
reactions, redox-active materials constitute the active phase. instance, improve the prediction of band gaps. When doing so,
Active catalytic phases based on metal oxides include, but are the electrons are more localized instead of being shared by
not limited to, silver, copper, vanadium, molybdenum, iron, several atoms,17 thus breaking symmetries in local config-
urations and generating polarons that for some materials can be
Downloaded via 202.92.144.147 on August 6, 2019 at 04:33:13 (UTC).
manner. To this end, we have taken an example that has been the specific formulation of this term results in the particular
extensively analyzed in our group.1,28,29 We aim at warning implementation of the method: for example, the around mean
potential users about the associated errors that are non- field (AFM)37 or the fully localized limit (FLL)36 methods, the
systematic and affect adsorption and reaction steps in different latter being more widely implemented and used.
manners, depending on the chemical process taking place on The occupation number of localized orbitals is computed by
the surface (acid−base vs redox). The origin of the the projection of the Kohn−Sham states on the localized basis
discrepancies to hybrid methods and potential ways to mitigate set. The implementation of this projection is an important
the problems are discussed. ingredient of the DFT+U method. For the plane wave basis set
representation used for slab models in heterogeneous catalysis,
2. COMPUTATIONAL TECHNIQUES the local basis sets are related to the pseudopotentials. Some
Even though DFT can be perceived as the major computational simplifications are usually included in the implementation to
tool for the description and prediction of heterogeneous allow efficient computations of energy derivatives, like forces on
catalytic processes, it is not free from important shortcomings. atoms. In general, the rotationally invariant implementations
As the exact exchange correlation (xc) functional is unknown, ensure that the occupation numbers are independent with
the approximate expressions for xc fail to capture the ground- respect to the localized atomic orbital basis set rotations.38 The
state properties of the systems where the localization of electron−electron interaction integrals of the Hubbard term
electrons is present. This is related to difficulties in recovering can be factorized with respect to angular and radial
the many-body interactions in the electron gas.10,30 contributions. Such factorizations involve Slater integrals Fk of
In heterogeneous catalysis, the active materials often present the radial part of the atomic wave functions. For d electrons F0,
localized electrons (strongly correlated), that is, transition F2, and F4 are required (F6 for f states). Within this approach,
metal oxides. Their chemical properties are governed by the effective Coulomb and exchange terms can be computed as U =
properties of the valence electrons. These electrons are F0 and J = (F2 + F4)/14. In practice, simplified formulations are
localized on particular d or f orbitals, and the approximate xc used, where only the lowest Slater integrals F0 are considered.
functionals tend to delocalize them, overstabilizing metallic In this formulation, it is customary to introduce an effective
ground states. As a result, DFT might predict metallic Coulomb interaction Ueff = U − J that incorporates the
properties of catalysts that are insulators, as mentioned before exchange correction J. This simplified DFT+U scheme can be
for NiO.9 written as39,40
This intrinsic problem of DFT is related to the unphysical
1
self-interaction of electrons, that is, the charge density of an E DFT + U = E DFT + U ∑ Tr[n I (1 − n I )]
individual electron interacting with itself. This repulsive 2 I (1)
interaction induces a disproportionate delocalization of the
wave functions. From this perspective, the explicit introduction where I stands for the site of the ions, and nI corresponds to the
of the Hartree−Fock approach (the Fock exchange term is free density matrices projected on localized atoms. Unless otherwise
of self-interaction) is often used to improve the accuracy of noted, this is the method discussed throughout this work. The
DFT via the so-called hybrid functionals.23,31 This usually leads value of U can be then obtained by different methods.35,41−48
to insulating ground states, although this approach is still based For the plane wave codes, one of the most extended
on single-electron description. Still, better treatment of these methodologies is the constrained DFT approach that follows
effects requires the improved description of electron interaction the linear response (LR) theory, presented by Cococcioni and
many-body terms that have been formulated in the dynamical Gironcoli.40 The idea behind this methodology is to evaluate
mean field theory32 or reduced density matrix functional the effect of the perturbation of the bare and screened density
theory.33 All these approaches significantly improve the response functions once a perturbation is applied to the local
prediction of the ground-state properties of bulk materials; projector. To ensure that the values are meaningful, these
however, they are computationally much more intensive than authors indicated that the perturbation needs to be performed
standard DFT. Consequently, at present, they are not practical in supercells with increasing size, to guarantee that the values
for the description of complex surfaces or catalytic processes. are not affected by periodic boundary conditions.40 It has to be
DFT+U methods, on the other hand, have only minor noted that the U value shows an important dependence on the
computational cost compared to standard DFT. The functional of choice (LDA vs GGA), the pseudopotential, the
fundamentals of the Hubbard model state that the electronic choice of the fitting properties (if any), or projection operators.
correlations associated with a few localized orbitals are Indeed, important differences have been found, and thus, the U
responsible for the complex electronic structure. These term shall better be reparameterized for each computational
methods are based on model Hamiltonians: the Hubbard setup.49−52 It has been proposed that the use of a unique U
model.16,34−36 Within this approach, the electrons in the value for a metal ion in different environments could lead to
studied system are divided into two categories: (i) the strongly inaccuracies when the local electronic structure changes.41 All
correlated ones, which are usually localized on atomic d or f these procedures to obtain the Hubbard U are particularly
orbitals, are those subject to Hubbard treatment; and (ii) the suited for bulk systems. For the slab models representing
remaining electrons that are treated in the standard DFT catalysts, the U values taken from the bulk are typically used.
manner. However, the changing local environment of the surface atoms
Then the energy of the system can be written as EDFT+U = during a chemical reaction is a potential source of errors.
EDFT + EHubbard[n] − EDC[n], where EHubbard is related to the An exhaustive analysis employing this methodology for
term of the Hubbard Hamiltonian for localized orbitals, EDC is cerium compounds was performed by Lu and Liu.41 These
the double counting term, and EDFT is the total energy of authors illustrated that adequate U values for Ce atoms in
standard DFT approach within the LDA or GGA approx- different configurations ranging from isolated atoms and ions,
imation. The double counting term is not uniquely defined, and to small oxide compounds or extended CeO2 bulk and surface
8371 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
Table 1. Most Typical U Values Suggested for Several Metal Oxides Using GGA+U Approachesa
Mn+ U (eV) method reference
CeOx
Ce3+/Ce4+ 2.0 fitting experimental properties Kresse20
Ce3+/Ce4+ 2.0−3.0 fitting experimental properties Illas19
Ce4+ 2.0−3.0 fitting experimental properties Fabris61
Ce3+ 4.5 linear response Fabris51
Ce4+ 5.0 band gap states Watson62
Ce4+ 5.13 linear response Liu41
Ce3+ 6.70 linear response Liu41
TiOx
Ti3+/Ti4+ 2.0−3.0 fitting experimental properties and band gap states Metiu63
Ti4+ 3.0 fitting experimental properties Nolan64
Ti4+ 3.0 fitting hybrid−dft calculations Selloni65
Ti4+ 3.4 linear response Mattioli66
Ti4+ 4.2 band gap states Watson67
Ti3+ 4.35 fitting experimental properties Wolverton52
Ti4+ 4.76 fitting experimental properties Wolverton52
Ti4+ 4.95b linear response Kitchin68
Ti4+ 10.0 fitting experimental properties Dupuis69
CoOx
Co2+/Co3+ 3.3 fitting experimental properties Ceder70
Co2+/Co3+ 3.52 fitting experimental properties Nørskov71
Co2+ 3.75 fitting experimental properties Wolverton52
Co3+ 4.26 fitting experimental properties Wolverton52
Co2+ 4.4 linear response Selloni72
Co4+ 4.77 fitting experimental properties Wolverton52
Co2+ 4.0−5.0 fitting experimental properties Doublet73
Co3+ 6.7 linear response Selloni72
MnOx
Mn4+ 1.0−1.6 fitting experimental properties Cockayne74
Mn2+ 2.98 fitting experimental properties Wolverton52
Mn4+ 3.19 fitting experimental properties Wolverton52
Mn4+ 3.0−4.0 fitting experimental properties and hybrid-dft calculations Kresse54
Mn3+ 4.54 fitting experimental properties Wolverton52
Mn4+ 6.63 linear response Kitchin68
FeOx
Fe3+ 3.0 fitting experimental properties Rollman75
Fe3+ 3.0 fitting experimental properties Morgan76
Fe2+/Fe3+ 3.61 fitting experimental properties Łodziana77
Fe2+/Fe3+ 3.7 fitting experimental properties De Leeuw78
Fe2+ 4.04 fitting experimental properties Wolverton52
Fe3+ 4.09 fitting experimental properties Wolverton52
MoOx
Mo6+ 2.0 fitting experimental properties Metiu79
Mo6+ 6.0 fitting hybrid−dft calculations Asta80
Mo6+ 6.3 fitting hybrid−dft calculations Willock81
Mo4+/Mo6+ 8.6 fitting experimental properties Bell56
a
Commonly fitted properties are the band gap, the lattice parameter, the bulk modulus, or the ΔH of formation of the oxide. The band gap states
method refers to properly localize these states within the band gap. The linear response method refers to the approach introduced by Cococcioni and
Gironcoli.40. bDifferent functionals, pseudopotentials, and TiO2 polymorphs lead to different values of U.59
models, presented some marked characteristics: (i) the ion and nearest neighbors interactions are needed for Ce in its
charge does not significantly affect the value of U when metallic form.53
properly screened by counterions (U values for Ce atoms, Ce in An extensive compilation of U values for Co, Fe, Ni, Cr, Mn,
Ce3HxO7 clusters, or CeO2 were ranging in the 4.1−5.2 eV V, and Ti ions on different oxide and fluoride environments has
range); and (ii) when ions are isolated the values are much been presented by Aykol and Wolverton on the basis of the
larger (close to 15 and 18 eV for Ce2.5+ and Ce3.5+, experimental ΔH of formation fitting approach.52 These
respectively). Thus, the environment and its polarizability authors aimed at reproducing accurate thermochemistry,54,55
turn out to be crucial to screen the Coulomb interactions on Ce making special emphasis on the role of the oxidation state and
4f orbitals effectively. This observation agrees with Cococcioni ligand contributions. They observed, for instance, that the U
and Gironcoli,40 who found that to obtain Ce 4f values, on-site term is not transferable when going from oxygen to fluoride
8372 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
Figure 1. (a) Adsorption of formaldehyde (A1) on the CeO2(111) regular surface and first hydrogen stripping to CHO and OH (I1). (b) Reaction
scheme for this elementary step. (c) Overlaid structures of TS1 for the different U values. Red stands for oxygen, yellow for cerium, black carbon,
and white hydrogen.
ligands, and is larger when the charge of the cation increases bulk positions. Using the same setup, screened hybrid HSE06
(although exceptions are identified). Similarly, Bell and co- calculations86 have been performed and employed as reference
workers,56 when investigating the reduction energies for to illustrate the dependence of the reactivity with the different
transition (Ti, V, Mo) and rare earth (Ce) metal oxides, values of U. The iron oxide slab contains 18 atomic layers and a
observed that the value of U had to be adjusted for each k-point sampling of 5 × 3 × 1 was employed.
reaction. The same authors in the investigation of the reducible
transition metal oxides, basically Bi2Mo3O12 for hydrogen 3. DISCUSSION
abstraction from propene, observed that both the barrier and
3.1. Formaldehyde Decomposition on CeO2(111).
the final state energy depend on the U parameter used for
Cerium oxide is among the most interesting oxides from a
Mo.57 When compared to their reference functional, M06-L,58
chemical point of view. The coexistence of acid−base, redox,
the thermodynamic and the kinetic parameters were not
and oxygen storage material properties makes it an outstanding
reproduced with a single U value. In a similar study that focused
ingredient in catalytic formulations where it can act as active
on the structure of different TiO2 polymorphs, Kitchin et al.
phase, cooperative or synergetic catalytic phase, or just as
realized that the U value obtained by linear response for the Ti support.28,87 The hydrophobicity of the surface can also be
d orbitals depends on the functional and pseudopotential tuned by the reduction degree.88,89 Applications centered on its
employed, and the particular crystal structure.59 Similar oxidative capacity have been preeminent: ceria is then usually
conclusions were obtained by Greeley et al. for several present in combination with metal phases, although recently it
transition-metal (hydroxy)oxides, where the dependence of has been shown to be active as a single phase in the Deacon
the enthalpy of formation with the U term varies significantly and HBr oxidation reactions,90,91 hydrogenations,92−95 and
when van der Waals effects were considered.60 selective C−C bond breaking.96 Shape-directed synthesis
In Table 1 we present a summary of the most common U protocols enable the synthesis of tailored ceria nano-
values employed together with the GGA formulation. These particles,97,98 in particular, nanocubes exposing {100} facets.
values can range from 2.0 to 10.0 eV for TiOx or from 2.0 to 8.6 The latter, which are polar and present two main
eV for MoOx, depending on a variety of factors. reconstructions,99,100 exhibit unique catalytic activity.93
In the present work, we have employed PBE+U with variable Many of these reactions combine acid−base and redox
U values in order to unravel the robustness of the different elementary steps. In the latter, one of the electrons from
approaches to obtain U and how the values affect the catalytic reactants and/or products can be accommodated at the surface
properties of the materials. To this end, we have used the VASP cations. Therefore, the correct identification of the energy gain/
code82,83 with the PBE functional,84 PAW pseudopotentials,85 loss by the reductive nature of the cations on the surface
and plane waves with a cutoff energy of 500 eV for the valence constitutes a major issue. To illustrate the effect, we have taken
electrons (5s, 5p, 4f, 6s for Ce atoms, 3d, 4s for Fe, 2s, 2p for O as an example the dissociation of formaldehyde on the ceria
and C, and 1s for H). The lattice parameter was optimized (111) surface, for which detailed temperature program
using a dense Γ-centered 7 × 7 × 7 k-point mesh that leads to a desorption experiments exist.101−103 The elementary step for
value of 5.497 Å, in good agreement with the experimental aexp dehydrogenation of formaldehyde is found in Figure 1.
of 5.411 Å. The (111) surface, the most stable termination of The most common values of U for LDA and PBE functionals
ceria, was modeled as a p(2 × 2) supercell with periodically are 5.3 and 4.5 eV,51 which have been extensively used in the
repeated slabs consisting of three O−Ce−O layers (nine single literature. The results of the reaction energy profiles obtained
layers) separated by 15 Å of vacuum space, which was with PBE+U for different values of U are intriguing (Figure 1).
optimized using a Γ-centered (3 × 3 × 1) k-point mesh. The Formaldehyde adsorption remains constant, as it does not
five outermost single layers and the adsorbate(s) were allowed involve any surface reduction, and thus, all points (CH2O + *
to relax, whereas the rest of the atoms were kept fixed to their → CH2O*) lie in the same energy. In contrast, the elementary
8373 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
step for the decomposition of formaldehyde (CH2O* → strength of the effective U parameter. In Figure 2, the problem
CHO* + H*) is always exothermic, but the energy span of electronic structure and how it is dealt with in DFT+U
between the small and large U values is more than 2 eV. The approaches becomes clear. While formaldehyde adsorption is a
reaction is more exothermic as the U increases. A similar constant and thus is not affected by changes in the U
behavior was observed by Bennett and Jones for CO+NO parameter, the following step in the decomposition is strongly
reaction on the ceria surface.104 More importantly, the energy affected. For small U values, the electrons are still delocalized,
of the transition state also presents a wide range of values that and thus, the values for the transition and final states are rather
expands about 1 eV. Considering that the computed activation constant. For higher values of U, the adsorption remains
energy is introduced in the exponential in the transition-state constant, but the transition and final states reduce their energies
theory, this means that at room temperature the difference in according to a linear dependence, yet with different slopes.
the kinetic coefficient for small and large values of U on Ce can At small U values, electrons are not localized even in reduced
account for several orders of magnitude. surface conditions, and thus, for U values lower than a certain
A clearer view of the process can be noticed when the threshold (that depends on the material), all the energy
adsorption-, transition-, and final-state energies are included in parameters for adsorption, transition, and thermodynamics of
the same plot, see Figure 2. The figure clearly illustrates that the the elementary steps are practically constant. Instead, at
balance between thermodynamics and kinetics is broken. The medium to large U values, both the transition and final states
kinetics of the system are wrong for several U values, and the are stabilized, but the size of this stabilization is doubled for the
energy of the transition state with respect to the gas-phase final state. This means that the thermodynamics and the
reference changes from positive to negative depending on the kinetics are no longer synchronous, and the choice of U implies
an error in one or the other. Comparison with experiments for
formaldehyde decomposition sets an experimental constraint to
the transition-state energy, as it should be larger than zero to
explain that desorption is preferred when compared to
dehydrogenation.102,103 This only happens for U values below
5.5 eV, thus constituting the upper bound to this parameter.
The influence of the U parameter on the direct reducibility of
the surface has also been assessed (Figure 3). This reducibility,
Ered, has been taken as the energy difference between the p(2 ×
2) supercell with one extra electron added (where a
background charge is added to preserve charge neutrality)
Figure 2. (Top) Activation energy (Ea, red) and reaction energy (ΔE, and the bare p(2 × 2) surface, as considered in previous
blue) for the first C−H bond breaking of formaldehyde on CeO2(111) work.1,28 The electron localization on a single cerium atom of
as a function of the U parameter. (Bottom) Adsorption energy of the topmost layer is also shown in Figure 3, for each value of U.
formaldehyde (blue), transition step for first hydrogen striping (red), The interpretation of the dependence with the U term in the
and final state (green), all with respect to gas-phase formaldehyde and redox steps and the difference between the kinetics and
the bare CeO2(111) surface, as a function of U. In both cases, the
horizontal lines account for the screened hybrid HSE06 results. PBE0
thermodynamics can be traced back to the local structures and
calculations provide nearly identical results as HSE06 (see text). For the electronic states of these configurations. The different
the redox processes, dependence on the U parameter is observed when slopes observed for the reaction and transition-state energies
electron localization in the Ce centers appears (turquoise background with the U parameter originate from the number of electrons
color). Notably, the dependences of the kinetic and thermodynamic that are localized at the surface. The ratio between the slopes
terms are different. gives the ratio between the electrons at the surface.
8374 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
When compared to more accurate screened hybrid The structurally consistent U suggested by Kulik and Marzari
calculations (HSE06), the U values at which the agreement is is at the origin of the problem. These authors observed the
obtained vary for the kinetic and the thermodynamic need for a smooth potential energy curve incorporating the
parameters (solid lines in Figure 2). This means that to variations of U (structurally consistent U) when the significant
match the HSE06 results a value of 4.27 eV is required for the changes in the hybridization and occupancy of a transition-
thermodynamics of the process, but only of 3.33 eV for the metal center occur along the reaction coordinate. When the
transition state. Similarly, hybrid PBE0 calculations provided authors introduced a polyatomic molecular system, they
almost identical results (4.32 and 3.33 eV, respectively). The recognized that more than one geometric variable affects the
calculation of Ered, Figure 3, where one electron was added to occupation matrix of the d-states and thus the U term depends
the bare surface slab (comparable then with the previous on several variables, U(R). Other structurally consistent
HCHO transition-state calculation), leads to a U value of 2.58 approaches were also indicated in the literature, and the effect
eV to reproduce the HSE06 result. Indeed, Kulik and when changing the coordination by adsorption of different O-
Marzari105 reported that the complete potential-energy surface containing species was reported for Ti.68,110 Kulik and Marzari
cannot be represented by a constant U value, through the study proposed CCSD(T) calculations as an appropriate benchmark
of the dissociation of 4Φ FeO+ by internuclear separation. and suggested that the inclusion of a U(R) in the nudged elastic
Instead they proposed that the geometry plays a major role and band is “trivial”. However, five years after this comment, we are
that, rather than being a parameter, U is a function of the local not aware of such implementation. Unfortunately, although this
geometry. It is worth stressing that although the HSE06 approach seems evident, it could only be applied to very small
functional also contains certain degree of inaccuracy, it typically systems (with one metal atom) that did not present a set of
outperforms LDA(+U) and GGA(+U), especially for the close-lying states.
description of semiconductor and insulator materials.23,106−109 Indeed, the search of transition-state calculations with DFT
Thus, hybrid functional results should not be considered as the and plane waves codes has an important problem when dealing
gold standard but rather as a higher rung in Jacob’s ladder of with the electronic structure. The reason is that in the
DFT. Proper benchmark studies should be performed using algorithms to search for transition states, the files that describe
more accurate RPA or QMC calculations, but these are beyond the electronic structure are taken to be the same (i.e., the U for
the scope of this work. each of the configurations along the band is identical). As we
The origin of the different U to fit the HSE06 values thus can have seen before, the strength of the U parameter seems to
be traced back to the local structures. For ceria, Lu and Liu41 provide unreasonable values for the kinetics. We propose here a
identified the dependence of the U parameter on the geometry methodology that can smooth the errors: (i) to identify acid−
of the environment: the larger the Ce−O distance, the larger base and redox steps; (ii) to determine of the most adequate U
the oxygen screening loss, and thus, the larger the interaction of value by comparison of the thermodynamics of the steps to a
the f electrons shall be (i.e., larger U). These authors illustrated reference calculation (ideally RPA or QMC, otherwise hybrid
that for variations of 0.1 Å, the effect on the U could be as large DFT); (iii) to unravel the dependence of the thermodynamics
as about 0.4 eV. We have compared the elastic modifications with the U (as in Figure 2); (iv) to elaborate on the kinetics an
both for the transition and final states at a given U = 3.5 eV, accurate count of the electron localization at the transition state
Figure 4. and comparison to its corresponding final state is needed, as
this gives the slope for the dependence of the transition state
with U; and (v) to rescale the U value for the transition state
according to the geometric perturbation between transition and
final states. This procedure is more stable than employing
variable U for each initial image in the band, as the initial
guesses both in terms of electronic structure and geometry
might significantly differ from the final solution. This is possible
because, at a difference from molecular calculations, many of
these steps correspond to bound initial and final states.
3.2. Iron Oxides. Iron oxides are among the cheapest
catalysts and present a very versatile structure that accom-
Figure 4. (Top) Most relevant Ce−O distances in Å for the initial, modates different oxygen stoichiometries. The structure of the
transition, and final states of formaldehyde decomposition on CeO2 at material presents certain difficulties in the adequate description
PBE+U (U = 3.5 eV) level. (Bottom) Electron localization in the of its electronic structure that are linked to the metal−insulator
initial, transition, and final states. Same color code as in Figure 1. transition occurring in some of its phases (Verwey transition).77
The use of U, applied to the Fe 3d orbitals, is crucial to obtain a
In contrast, acid−base steps do not present this problem. We proper orbital ordering to understand surface reconstructions in
have performed the first reaction in methanol decomposition a material as common as Fe3O4. In that case, the U value
on ceria(110), in the same manner as presented before. The employed was 3.61 eV. 77 For a completely different
adsorption, transition, and final states (i.e., the thermodynamics stoichiometry, in the study of water interacting with iron
and kinetics) are completely stable and completely independent oxide films, the effective U value was very similar, 3.0 eV. In
of the U parameter. This seems to be a reasonable conclusion both cases the U values were fitted for the PW91 functional
because no surface atoms are reduced during the process. and, as also observed by Lu and Liu for ceria, the charge of the
However, it sets an internal contradiction to generate reaction cation is not the primary term affecting the U value.41 To
energy profiles that maintain the correct balance of the kinetics analyze the extension of the problem, we have investigated the
and thermodynamics in complex reaction networks that contain same reaction (CH2O* + * → CHO* + H*) on Fe2O3. Again,
both acid−base and redox steps. the observed patterns for adsorption, reaction energy for the
8375 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
first abstraction, and the corresponding transition-state energy optimal value, as the local geometric perturbations induced by
(Figure 5) resemble those in Figure 2. However, two single-atom configurations might be very large.
3.4. Other Materials Related to Energy Production
and Storage. The problems observed in understanding the
chemistry of ceria and iron oxides can be further extended to
other materials where an adequate band alignment is needed to
rationalize the chemical processes. Examples of such behavior
can be found in several studies,116−118 which illustrate that the
proper band alignment in titania is required for the photo-
electrochemical properties to be well-balanced, particularly in
complex interfaces. The fitting of U for the different cations is
required. Still, some authors have proposed that the use of
linear response DFT+U in reactions, like the oxygen evolution
activity on rutiles, would only change the activity ordering of
different materials.68 According to these authors, the weakening
of the adsorption energies (taken as descriptors) will not affect
the relative ordering of reaction energies and, at most, the
variation will only affect the top of the activity volcano. In view
Figure 5. Adsorption energy of formaldehyde (blue), transition step
for first hydrogen stripping (red), and final state (green), all with of our results, larger differences shall be expected for doped
respect to gas-phase formaldehyde and the bare Fe2O3 surface, as a systems,119 particularly if chemical and electrochemical steps
function of U. The background color indicates when the electron is are present in the reaction network.120
properly localized in the Fe center (turquoise) and when spin-crossing The problem is also prevalent for other compounds, like
is observed (purple). Notably, the dependences of the kinetic and cobalt oxide, with a wide range of stable compositions seem
thermodynamic terms are different. particularly difficult. Recent studies have employed the fitted
values of U from Wang and Ceder,70 which were adapted to
remarkable points emerge. First, the dependencies are different reproduce thermodynamic properties (ΔH of formation of the
from those in Figure 2, where now the transition-state slope is oxide), but still corrections were needed to obtain the proper
larger than that for the reaction energy. This behavior can be thermodynamics between the different phases.121 Similar
traced back to the occupation of the d shell versus f shell in Fe problems arise in manganese oxides, for which site-dependent
versus Ce. The former is by reduction more than half full, while values were proposed,54 and the most recent use of nickel oxo-
that of Ce is empty or with less than one electron. Second, at hydroxides122−124 with varied valences of the Ni centers and
high U values (6.5 eV, purple region of Figure 5), the electronic mixed Fe−Ni materials for energy-related catalysis125,126 are
state of the initial and final states are not the same as for foreseen. The implication of the U dependence will also affect
intermediate values of U, and therefore, the corresponding other families of materials with applications in Li-ion batteries,
energies do not follow the trends. Yet, this highlights a problem water splitting, or photovoltaics such as Prussian Blue
that is much more acute for less oxidized oxides, like FeO. Our analogues,127,128 perovskites,129,130 or Li-ion battery cathodes.52
preliminary results for that surface show that spin-crossings can
occur at different positions of the intrinsic reaction paths 4. OUTLOOK
connecting reactants and products, and thus the equivalent Computational techniques based on DFT have reached a level
results to those in Figure 2 and 5 will develop in families of of maturity that makes possible the accurate reproduction of
planes for each of the spin states considered. experiments and the prediction of reactivity up to levels that
3.3. Complex Interfaces and Single-Atom Catalysts. seemed unreachable just a decade ago. The statement is
The problem of different U values might be also acute in the particularly true for metals and insulators, but for materials with
case of isolated atoms or monomers. An example that parallels complex electronic structures, multiple spin configurations, and
our investigation on methanol conversion is the same reaction different degrees of electron localization, this is not so
on vanadyl ions and clusters supported on CeO2.111,112 These straightforward. We have shown how the thermodynamics
atoms form a complex interface for which Ce 4f U values were and the kinetics for a simple reaction respond differently to the
employed,111 but then the cooperative effects (as there are two external energy penalty included in the models to account for
different centers that can be reduced competitively) are difficult proper representation of the band gap. The ultimate result is
to unravel. Therefore, identification of the active vanadium that the dispersion leads to a complete disagreement between
containing species might not be trivial.112 the experimental observations and the computed values. In this
The case of the VOx/CeO2 materials shares some common Viewpoint, we presented a revision of what needs to be done to
features with single-atom catalysts, SAC, proposed for isolated achieve a reasonable accuracy at a reasonable computational
atoms in oxide matrices. The term was coined by Zhang and cost, so as to improve the use of DFT for strongly correlated
co-workers113 for the description of isolated Pt atoms systems. Most of the chemistry related to photochemistry,
embedded in FeOx, which were tested in CO oxidation, but energy extraction to suitable energy vectors, conversion, and
follows similar principles to the Au/CeO 2 or Pt/CeO 2 storage are based on the use of materials with complex
chemistry reported by Flytzani-Stephanopoulos.114 In all electronic structures, and thus, the results here are of
fundamental interest to the community.
■
these cases, the theoretical simulations performed to under-
stand this chemistry have been carried out with a U value
derived from the clean host material. For the Pt/FeOx oxide, AUTHOR INFORMATION
the authors took a U = 3 eV from hematite.115 However, as we Corresponding Author
have shown, this approximation might really diverge from the *E-mail: [email protected].
8376 DOI: 10.1021/acscatal.6b01907
ACS Catal. 2016, 6, 8370−8379
ACS Catalysis Viewpoint
ORCID (20) Da Silva, J. L. F.; Ganduglia-Pirovano, M. V.; Sauer, J.; Bayer, V.;
Marçal Capdevila-Cortada: 0000-0003-4391-6580 Kresse, G. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, 45121.
Núria López: 0000-0001-9150-5941 (21) Nolan, M.; Parker, S. C.; Watson, G. W. Surf. Sci. 2005, 595,
223−232.
Notes (22) Perdew, J. P.1995; pages 51−64.
The authors declare no competing financial interest. (23) Paier, J. Catal. Lett. 2016, 146, 861−885.
■ ACKNOWLEDGMENTS
We thank MINECO (CTQ2015-68770-R) for financial support
(24) Bohm, D.; Pines, D. Phys. Rev. 1951, 82, 625−634.
(25) Pines, D.; Bohm, D. Phys. Rev. 1952, 85, 338−353.
(26) Bohm, D.; Pines, D. Phys. Rev. 1953, 92, 609−625.
(27) Erhard, J.; Bleiziffer, P.; Görling, A. Phys. Rev. Lett. 2016, 117,
and BSC-RES for providing generous computer resources. 143002.
M.C.-C. also acknowledges MINECO for a “Juan de la Cierva (28) Capdevila-Cortada, M.; López, N. ACS Catal. 2015, 5, 6473−
− Formación” fellowship (FJCI-2014-20568). Z.L. acknowl- 6480.
edges NSC Poland support through project 2015/17/B/ST3/ (29) Capdevila-Cortada, M.; García-Melchor, M.; López, N. J. Catal.
02478.
■
2015, 327, 58−64.
(30) Mahan, G. D.Many-Particle Physics; Springer U.S.: Boston, MA,
REFERENCES 2000.
(1) Capdevila-Cortada, M.; Vilé, G.; Teschner, D.; Pérez-Ramírez, J.; (31) Becke, A. D. J. Chem. Phys. 1993, 98, 1372.
López, N. Appl. Catal., B 2016, 197, 299−312. (32) Georges, A.; Kotliar, G.; Krauth, W.; Rozenberg, M. J. Rev. Mod.
(2) Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. Chem. Phys. 1996, 68, 13−125.
Rev. 2016, 116, 5987−6041. (33) Yang, W.; Zhang, Y.; Ayers, P. W. Phys. Rev. Lett. 2000, 84,
(3) Sutton, J. E.; Guo, W.; Katsoulakis, M. A.; Vlachos, D. G. Nat. 5172−5175.
Chem. 2016, 8, 331−337. (34) Gutzwiller, M. C. Phys. Rev. Lett. 1963, 10, 159−162.
(4) Nørskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H. Nat. (35) Anisimov, V. I.; Gunnarsson, O. Phys. Rev. B: Condens. Matter
Chem. 2009, 1, 37−46. Mater. Phys. 1991, 43, 7570−7574.
(5) Lejaeghere, K.; Bihlmayer, G.; Bjorkman, T.; Blaha, P.; Blugel, S.; (36) Anisimov, V. I.; Solovyev, I. V.; Korotin, M. A.; Czyżyk, M. T.;
Blum, V.; Caliste, D.; Castelli, I. E.; Clark, S. J.; Dal Corso, A.; de Sawatzky, G. A. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 48,
Gironcoli, S.; Deutsch, T.; Dewhurst, J. K.; Di Marco, I.; Draxl, C.; 16929−16934.
Dulak, M.; Eriksson, O.; Flores-Livas, J. A.; Garrity, K. F.; Genovese, (37) Anisimov, V. I.; Zaanen, J.; Andersen, O. K. Phys. Rev. B:
L.; Giannozzi, P.; Giantomassi, M.; Goedecker, S.; Gonze, X.; Granas, Condens. Matter Mater. Phys. 1991, 44, 943−954.
O.; Gross, E. K. U.; Gulans, A.; Gygi, F.; Hamann, D. R.; Hasnip, P. J.; (38) Liechtenstein, A. I.; Anisimov, V. I.; Zaanen, J. Phys. Rev. B:
Holzwarth, N. A. W.; Iu an, D.; Jochym, D. B.; Jollet, F.; Jones, D.; Condens. Matter Mater. Phys. 1995, 52, R5467−R5470.
Kresse, G.; Koepernik, K.; Kucukbenli, E.; Kvashnin, Y. O.; Locht, I. L. (39) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.;
M.; Lubeck, S.; Marsman, M.; Marzari, N.; Nitzsche, U.; Nordstrom, Sutton, A. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 57,
L.; Ozaki, T.; Paulatto, L.; Pickard, C. J.; Poelmans, W.; Probert, M. I. 1505−1509.
J.; Refson, K.; Richter, M.; Rignanese, G.-M.; Saha, S.; Scheffler, M.; (40) Cococcioni, M.; de Gironcoli, S. Phys. Rev. B: Condens. Matter
Schlipf, M.; Schwarz, K.; Sharma, S.; Tavazza, F.; Thunstrom, P.; Mater. Phys. 2005, 71, 35105.
Tkatchenko, A.; Torrent, M.; Vanderbilt, D.; van Setten, M. J.; Van (41) Lu, D.; Liu, P. J. Chem. Phys. 2014, 140, 084101.
Speybroeck, V.; Wills, J. M.; Yates, J. R.; Zhang, G.-X.; Cottenier, S. (42) Gunnarsson, O.; Andersen, O. K.; Jepsen, O.; Zaanen, J. Phys.
Science 2016, 351, aad3000. Rev. B: Condens. Matter Mater. Phys. 1989, 39, 1708−1722.
(6) Teschner, D.; Novell-Leruth, G.; Farra, R.; Knop-Gericke, A.; (43) Mosey, N. J.; Carter, E. A. Phys. Rev. B: Condens. Matter Mater.
Schlögl, R.; Szentmiklósi, L.; González Hevia, M.; Soerijanto, H.; Phys. 2007, 76, 155123.
Schomäcker, R.; Pérez-Ramírez, J.; López, N. Nat. Chem. 2012, 4, (44) Mosey, N. J.; Liao, P.; Carter, E. A. J. Chem. Phys. 2008, 129,
739−745. 014103.
(7) Pogodin, S.; López, N. ACS Catal. 2014, 4, 2328−2332. (45) Springer, M.; Aryasetiawan, F. Phys. Rev. B: Condens. Matter
(8) Stamatakis, M.; Vlachos, D. G. J. Chem. Phys. 2011, 134, 214115. Mater. Phys. 1998, 57, 4364−4368.
(9) Shen, Z.-X.; List, R. S.; Dessau, D. S.; Wells, B. O.; Jepsen, O.; (46) Shih, B.-C.; Abtew, T. A.; Yuan, X.; Zhang, W.; Zhang, P. Phys.
Arko, A. J.; Barttlet, R.; Shih, C. K.; Parmigiani, F.; Huang, J. C.;
Rev. B: Condens. Matter Mater. Phys. 2012, 86, 165124.
Lindberg, P. A. P. Phys. Rev. B: Condens. Matter Mater. Phys. 1991, 44, (47) Aryasetiawan, F.; Imada, M.; Georges, A.; Kotliar, G.; Biermann,
3604−3626.
S.; Lichtenstein, A. I. Phys. Rev. B: Condens. Matter Mater. Phys. 2004,
(10) Martin, R. M.; Reining, L.; Ceperley, D. M.Interacting Electrons:
Theory and Computational Approaches; Cambridge University Press: 70, 195104.
Cambridge, U.K., 2016. (48) Himmetoglu, B.; Floris, A.; de Gironcoli, S.; Cococcioni, M. Int.
(11) Perdew, J. P.; Zunger, A. Phys. Rev. B: Condens. Matter Mater. J. Quantum Chem. 2014, 114, 14−49.
Phys. 1981, 23, 5048−5079. (49) Fabris, S.; de Gironcoli, S.; Baroni, S.; Vicario, G.; Balducci, G.
(12) Szotek, Z.; Temmerman, W. M.; Winter, H. Phys. Rev. B: Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 71, 41102.
Condens. Matter Mater. Phys. 1993, 47, 4029−4032. (50) Kresse, G.; Blaha, P.; Da Silva, J. L. F.; Ganduglia-Pirovano, M.
(13) Tsuneda, T.; Hirao, K. J. Chem. Phys. 2014, 140, 18A513. V. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 237101.
(14) Zawadzki, P.; Jacobsen, K. W.; Rossmeisl, J. Chem. Phys. Lett. (51) Fabris, S.; de Gironcoli, S.; Baroni, S.; Vicario, G.; Balducci, G.
2011, 506, 42−45. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 237102.
(15) Gudmundsdóttir, H.; Jónsson, E. Ö .; Jónsson, H. New J. Phys. (52) Aykol, M.; Wolverton, C. Phys. Rev. B: Condens. Matter Mater.
2015, 17, 083006. Phys. 2014, 90, 115105.
(16) Hubbard, J. Proc. R. Soc. London, Ser. A 1963, 276, 238−257. (53) Notice that for other metals, like Fe, the same authors indicate
(17) Ganduglia-Pirovano, M.; Da Silva, J.; Sauer, J. Phys. Rev. Lett. that second-nearest-neighbors effects are also important.
2009, 102, 26101. (54) Franchini, C.; Podloucky, R.; Paier, J.; Marsman, M.; Kresse, G.
(18) Kowalski, P. M.; Camellone, M. F.; Nair, N. N.; Meyer, B.; Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, 195128.
Marx, D. Phys. Rev. Lett. 2010, 105, 146405. (55) Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.;
(19) Loschen, C.; Carrasco, J.; Neyman, K. M.; Illas, F. Phys. Rev. B: Persson, K. A.; Ceder, G. Phys. Rev. B: Condens. Matter Mater. Phys.
Condens. Matter Mater. Phys. 2007, 75, 35115. 2011, 84, 45115.
(56) Lutfalla, S.; Shapovalov, V.; Bell, A. T. J. Chem. Theory Comput. (92) Vilé, G.; Bridier, B.; Wichert, J.; Pérez-Ramírez, J. Angew. Chem.,
2011, 7, 2218−2223. Int. Ed. 2012, 51, 8620−8623.
(57) Getsoian, A. “Bean”; Bell, A. T. J. Phys. Chem. C 2013, 117, (93) Vilé, G.; Colussi, S.; Krumeich, F.; Trovarelli, A.; Pérez-Ramírez,
25562−25578. J. Angew. Chem., Int. Ed. 2014, 53, 12069−12072.
(58) Zhao, Y.; Truhlar, D. G. J. Chem. Phys. 2006, 125, 194101. (94) García-Melchor, M.; Bellarosa, L.; López, N. ACS Catal. 2014, 4,
(59) Curnan, M. T.; Kitchin, J. R. J. Phys. Chem. C 2015, 119, 4015−4020.
21060−21071. (95) García-Melchor, M.; López, N. J. Phys. Chem. C 2014, 118,
(60) Zeng, Z.; Chan, M. K. Y.; Zhao, Z.-J.; Kubal, J.; Fan, D.; Greeley, 10921−10926.
J. J. Phys. Chem. C 2015, 119, 18177−18187. (96) Haider, M. H.; Dummer, N. F.; Knight, D. W.; Jenkins, R. L.;
(61) Huang, M.; Fabris, S. J. Phys. Chem. C 2008, 112, 8643−8648. Howard, M.; Moulijn, J.; Taylor, S. H.; Hutchings, G. J. Nat. Chem.
(62) Nolan, M.; Grigoleit, S.; Sayle, D. C.; Parker, S. C.; Watson, G. 2015, 7, 1028−1032.
W. Surf. Sci. 2005, 576, 217−229. (97) Qiao, Z.-A.; Wu, Z.; Dai, S. ChemSusChem 2013, 6, 1821−1833.
(63) Hu, Z.; Metiu, H. J. Phys. Chem. C 2011, 115, 5841−5845. (98) Bruix, A.; Neyman, K. M. Catal. Lett. 2016, 146, 2053−2080.
(64) Nolan, M.; Elliott, S. D.; Mulley, J. S.; Bennett, R. A.; Basham, (99) Lin, Y.; Wu, Z.; Wen, J.; Poeppelmeier, K. R.; Marks, L. D. Nano
M.; Mulheran, P. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 77, Lett. 2014, 14, 191−196.
235424. (100) Capdevila-Cortada, M.; López, N. Nature Mater. 2016,
(65) Finazzi, E.; Di Valentin, C.; Pacchioni, G.; Selloni, A. J. Chem. DOI: 10.1038/nmat4804.
Phys. 2008, 129, 154113. (101) Zhou, J.; Mullins, D. R. Surf. Sci. 2006, 600, 1540−1546.
(66) Mattioli, G.; Filippone, F.; Alippi, P.; Amore Bonapasta, A. Phys. (102) Mullins, D. R.; Robbins, M. D.; Zhou, J. Surf. Sci. 2006, 600,
Rev. B: Condens. Matter Mater. Phys. 2008, 78, 241201. 1547−1558.
(67) Morgan, B. J.; Watson, G. W. Surf. Sci. 2007, 601, 5034−5041. (103) Albrecht, P. M.; Mullins, D. R. Langmuir 2013, 29, 4559−4567.
(68) Xu, Z.; Rossmeisl, J.; Kitchin, J. R. J. Phys. Chem. C 2015, 119, (104) Bennett, L. J.; Jones, G. Phys. Chem. Chem. Phys. 2014, 16,
4827−4833. 21032−21038.
(69) Deskins, N. A.; Dupuis, M. Phys. Rev. B: Condens. Matter Mater. (105) Kulik, H. J.; Marzari, N. J. Chem. Phys. 2011, 135, 194105.
Phys. 2007, 75, 195212. (106) Janesko, B. G.; Henderson, T. M.; Scuseria, G. E. Phys. Chem.
(70) Wang, L.; Maxisch, T.; Ceder, G. Phys. Rev. B: Condens. Matter Chem. Phys. 2009, 11, 443−454.
Mater. Phys. 2006, 73, 195107. (107) Henderson, T. M.; Paier, J.; Scuseria, G. E. Phys. Status Solidi B
(71) García-Mota, M.; Bajdich, M.; Viswanathan, V.; Vojvodic, A.; 2011, 248, 767−774.
Bell, A. T.; Nørskov, J. K. J. Phys. Chem. C 2012, 116, 21077−21082. (108) Freysoldt, C.; Grabowski, B.; Hickel, T.; Neugebauer, J.;
(72) Chen, J.; Wu, X.; Selloni, A. Phys. Rev. B: Condens. Matter Mater. Kresse, G.; Janotti, A.; Van de Walle, C. G. Rev. Mod. Phys. 2014, 86,
Phys. 2011, 83, 245204. 253−305.
(73) Dalverny, A.-L.; Filhol, J.-S.; Lemoigno, F.; Doublet, M.-L. J. (109) Wang, Y.; Cheng, H.-P. J. Phys. Chem. C 2013, 117, 2106−
Phys. Chem. C 2010, 114, 21750−21756. 2112.
(74) Cockayne, E.; Li, L. Chem. Phys. Lett. 2012, 544, 53−58. (110) Hsu, H.; Umemoto, K.; Cococcioni, M.; Wentzcovitch, R.
(75) Rollmann, G.; Entel, P.; Rohrbach, A.; Hafner, J. Phase Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 79, 125124.
Transitions 2005, 78, 251−258. (111) Kropp, T.; Paier, J.; Sauer, J. J. Am. Chem. Soc. 2014, 136,
(76) Pinney, N.; Kubicki, J. D.; Middlemiss, D. S.; Grey, C. P.; 14616−14625.
Morgan, D. Chem. Mater. 2009, 21, 5727−5742. (112) Wu, X.-P.; Gong, X.-Q. J. Am. Chem. Soc. 2015, 137, 13228−
(77) Lodziana, Z. Phys. Rev. Lett. 2007, 99, 206402. 13231.
(78) Santos-Carballal, D.; Roldan, A.; Grau-Crespo, R.; de Leeuw, N. (113) Qiao, B.; Wang, A.; Yang, X.; Allard, L. F.; Jiang, Z.; Cui, Y.;
H. Phys. Chem. Chem. Phys. 2014, 16, 21082−21097. Liu, J.; Li, J.; Zhang, T. Nat. Chem. 2011, 3, 634−641.
(79) Agarwal, V.; Metiu, H. J. Phys. Chem. C 2016, 120, 19252− (114) Fu, Q.; Saltsburg, H.; Flytzani-Stephanopoulos, M. Science
19264. 2003, 301, 935−938.
(80) Ding, H.; Lin, H.; Sadigh, B.; Zhou, F.; Ozoliņs,̌ V.; Asta, M. J. (115) Rollmann, G.; Rohrbach, A.; Entel, P.; Hafner, J. Phys. Rev. B:
Phys. Chem. C 2014, 118, 15565−15572. Condens. Matter Mater. Phys. 2004, 69, 165107.
(81) Coquet, R.; Willock, D. J. Phys. Chem. Chem. Phys. 2005, 7, (116) Scanlon, D. O.; Dunnill, C. W.; Buckeridge, J.; Shevlin, S. A.;
3819. Logsdail, A. J.; Woodley, S. M.; Catlow, C. R. A.; Powell, M. J.;
(82) Kresse, G.; Furthmuller, J. Phys. Rev. B: Condens. Matter Mater. Palgrave, R. G.; Parkin, I. P.; Watson, G. W.; Keal, T. W.; Sherwood,
Phys. 1996, 54, 11169−11186. P.; Walsh, A.; Sokol, A. A. Nat. Mater. 2013, 12, 798−801.
(83) Kresse, G.; Furthmüller, J. Comput. Mater. Sci. 1996, 6, 15−50. (117) El-Sayed, A.; Borghetti, P.; Goiri, E.; Rogero, C.; Floreano, L.;
(84) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, Lovat, G.; Mowbray, D. J.; Cabellos, J. L.; Wakayama, Y.; Rubio, A.;
3865−3868. Ortega, J. E.; de Oteyza, D. G. ACS Nano 2013, 7, 6914−6920.
(85) Blöchl, P. E. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, (118) Selcuk, S.; Selloni, A. Nat. Mater. 2016, 15, 1107−1112.
17953−17979. (119) García-Mota, M.; Vojvodic, A.; Metiu, H.; Man, I. C.; Su, H.-
(86) Krukau, A. V.; Vydrov, O. A.; Izmaylov, A. F.; Scuseria, G. E. J. Y.; Rossmeisl, J.; Nørskov, J. K. ChemCatChem 2011, 3, 1607−1611.
Chem. Phys. 2006, 125, 224106. (120) García-Melchor, M.; Vilella, L.; López, N.; Vojvodic, A.
(87) McCabe, R.; Trovarelli, A. Appl. Catal., B. 2016, in press. ChemCatChem 2016, 8, 1792−1798.
Available at the following: http://dx.doi.org/10.1016/j.apcatb.2016.04. (121) Bajdich, M.; García-Mota, M.; Vojvodic, A.; Nørskov, J. K.;
044. Bell, A. T. J. Am. Chem. Soc. 2013, 135, 13521−13530.
(88) Azimi, G.; Dhiman, R.; Kwon, H.-M.; Paxson, A. T.; Varanasi, K. (122) Hall, D. E. J. Electrochem. Soc. 1983, 130, 317.
K. Nat. Mater. 2013, 12, 315−320. (123) Lyons, M. E. G.; Brandon, M. P. Int. J. Electrochem. Sci. 2008, 3,
(89) Carchini, G.; García-Melchor, M.; Łodziana, Z.; López, N. ACS 1386−1424.
Appl. Mater. Interfaces 2016, 8, 152−160. (124) Bediako, D. K.; Lassalle-Kaiser, B.; Surendranath, Y.; Yano, J.;
(90) Amrute, A. P.; Mondelli, C.; Moser, M.; Novell-Leruth, G.; Yachandra, V. K.; Nocera, D. G. J. Am. Chem. Soc. 2012, 134, 6801−
López, N.; Rosenthal, D.; Farra, R.; Schuster, M. E.; Teschner, D.; 6809.
Schmidt, T.; Pérez-Ramírez, J. J. Catal. 2012, 286, 287−297. (125) Görlin, M.; Chernev, P.; Ferreira de Araújo, J.; Reier, T.;
(91) Moser, M.; Vilé, G.; Colussi, S.; Krumeich, F.; Teschner, D.; Dresp, S.; Paul, B.; Krähnert, R.; Dau, H.; Strasser, P. J. Am. Chem. Soc.
Szentmiklósi, L.; Trovarelli, A.; Pérez-Ramírez, J. J. Catal. 2015, 331, 2016, 138, 5603−5614.
128−137. (126) Dionigi, F.; Strasser, P. Adv. Energy Mater. 2016, 1600621.