scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Scaling properties of adsorption energies for hydrogen-containing molecules on transition-metal surfaces.

06 Jul 2007-Physical Review Letters (American Physical Society)-Vol. 99, Iss: 1, pp 016105-016105
TL;DR: The scaling model is developed into a general framework for estimating the reaction energies for hydrogenation and dehydrogenation reactions and it is found that the adsorption energy of any of the molecules considered scales approximately with the adhesion energy of the central, C, N, O, or S atom.
Abstract: Density functional theory calculations are presented for CHx, x=0,1,2,3, NHx, x=0,1,2, OHx, x=0,1, and SHx, x=0,1 adsorption on a range of close-packed and stepped transition-metal surfaces. We find that the adsorption energy of any of the molecules considered scales approximately with the adsorption energy of the central, C, N, O, or S atom, the scaling constant depending only on x. A model is proposed to understand this behavior. The scaling model is developed into a general framework for estimating the reaction energies for hydrogenation and dehydrogenation reactions.

Summary (1 min read)

Jump to:  and [Citation (APA):]

Citation (APA):

  • The model shows how the adsorbate valency, together with the properties of the d electrons of the surface, determines the adsorption energy.
  • Such a local density approximation for the interaction energy gives a good description of general trends in bonding, including bond lengths of atoms in metals, adsorbates on metal surfaces, and of molecules [22–24].

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright
owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.
Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
You may not further distribute the material or use it for any profit-making activity or commercial gain
You may freely distribute the URL identifying the publication in the public portal
If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately
and investigate your claim.
Downloaded from orbit.dtu.dk on: Aug 09, 2022
Scaling properties of adsorption energies for hydrogen-containing molecules on
transition-metal surfaces
Abild-Pedersen, Frank; Greeley, Jeffrey Philip; Studt, Felix; Rossmeisl, Jan; Fronczek-Munter, Ture
Rønved; Moses, Poul Georg; Skulason, Egill; Bligaard, Thomas; Nørskov, Jens Kehlet
Published in:
Physical Review Letters
Link to article, DOI:
10.1103/PhysRevLett.99.016105
Publication date:
2007
Document Version
Publisher's PDF, also known as Version of record
Link back to DTU Orbit
Citation (APA):
Abild-Pedersen, F., Greeley, J. P., Studt, F., Rossmeisl, J., Fronczek-Munter, T. R., Moses, P. G., Skulason, E.,
Bligaard, T., & Nørskov, J. K. (2007). Scaling properties of adsorption energies for hydrogen-containing
molecules on transition-metal surfaces. Physical Review Letters, 99(1), 016105.
https://doi.org/10.1103/PhysRevLett.99.016105

Scaling Properties of Adsorption Energies for Hydrogen-Containing Molecules
on Transition-Metal Surfaces
F. Abild-Pedersen, J. Greeley, F. Studt, J. Rossmeisl, T. R. Munter, P. G. Moses, E. Sku
´
lason, T. Bligaard, and J. K. Nørskov
Center for Atomic-scale Materials Design, Department of Physics, NanoDTU, Technical University of Denmark,
DK-2800 Lyngby, Denmark
(Received 13 February 2007; published 6 July 2007)
Density functional theory calculations are presented for CH
x
, x 0; 1; 2; 3, NH
x
, x 0; 1; 2, OH
x
, x
0; 1, and SH
x
, x 0; 1 adsorption on a range of close-packed and stepped transition-metal surfaces. We
find that the adsorption energy of any of the molecules considered scales approximately with the
adsorption energy of the central, C, N, O, or S atom, the scaling constant depending only on x. A model
is proposed to understand this behavior. The scaling model is developed into a general framework for
estimating the reaction energies for hydrogenation and dehydrogenation reactions.
DOI: 10.1103/PhysRevLett.99.016105 PACS numbers: 82.45.Jn
The formation of a bond between a molecule and a metal
surface is an important phenomenon in a number of pro-
cesses including heterogeneous catalysis [1], contact for-
mation in molecular electronics [2], and anchoring of
biomolecules to solids for sensors and other biomedical
applications [3]. The adsorption energy is a key quantity
describing the strength of the interaction of molecules with
the surface. The adsorption energy can be measured by
advanced surface science techniques [4 6]. Alternatively,
density functional theory (DFT) offers the possibility of
calculating adsorption energies with reasonable accuracy
[711]. While both experiments and DFT calculations are
feasible for a limited number of systems, they can hardly
be performed in detail for all potentially interesting ad-
sorption systems. There is therefore a need for simple
models with the ability to estimate bond energies in a first
screening of interesting systems. A successful model will
also expose the important factors determining the strength
of an adsorbate-surface bond. In the present Letter we will
develop such a model for hydrogen-containing molecules.
We use DFT calculations to derive a number of correlations
between adsorption energies, and we then present a model
to explain them. The model shows how the adsorbate
valency, together with the properties of the d electrons of
the surface, determines the adsorption energy. We further
develop the scaling model into a method for estimating
hydrogenation or dehydrogenation reaction energies for
organic molecules on transition-metal surfaces. The model
is tested against full DFT calculations for reactions of
hydrocarbons, alcohols, thiols, and amino acids.
First, we present results of extensive DFT calculations of
the adsorption energies of CH
x
, x 0; 1; 2; 3, NH
x
, x
0; 1; 2, OH
x
, x 0; 1, and SH
x
, x 0; 1 on a range of
close-packed and stepped metal surfaces. The study in-
volves the close-packed fcc(111), fcc(100), hcp(0001), and
bcc(110) surfaces, and the stepped fcc(211) and bcc(210)
surfaces. Each of the surfaces is modeled by a (2 2)ora
(1 2) surface unit cell for the close-packed and stepped
surfaces, respectively. Each slab has a thickness of three
layers in the direction perpendicular to the close-packed
surface. These slabs are thick enough to capture the trends
in the chemisorption energetics. The adsorbates and the
topmost layer are allowed to relax fully in all configura-
tions, and in the case of Fe, Ni, and Co, spin polarization is
taken into account. The binding energies of the different
species have been taken for the most stable adsorption sites
on all surfaces. The RPBE functional [12] in the general-
ized gradient approximation is used to describe exchange
and correlation effects. The calculational method and setup
is described in Ref. [13].
Figure 1 summarizes the results of the DFT calculations.
We find for all the molecules studied that the adsorption
energy of molecule AH
x
is linearly correlated with the
adsorption energy of atom A:
E
AH
x
E
A
: (1)
There is some scatter around the linear relations, but we
note that while the adsorption energies vary by several
electron volts over the range of metals considered here,
the mean absolute error (MAE) is only 0.13 eV. Some of
this scatter is related to differences in adsorption sites for
adsorbates with different amounts of hydrogen. CH
3
, for
instance, typically prefers a onefold adsorption site on the
close-packed surfaces while C prefers the threefold site. If
we were to use the adsorption energy of C in the onefold
adsorption site as the reference, the quality of the correla-
tion becomes significantly better (MAE 0:06 eV); see
Fig. 2. We also find that when we use a reference with the
same configuration as the molecule of interest, the scaling
behavior includes alloys with the same accuracy as for the
elemental metals (Fig. 2).
The main observation from Figs. 1 and 2 is that the slope
of the linear relationship in Eq. (1) is given to a good
approximation by the number of H atoms in AH
x
as x
x
max
x=x
max
, where x
max
is the maximum number of H
atoms that can bond to the central atom A (x
max
4 for
PRL 99, 016105 (2007)
PHYSICAL REVIEW LETTERS
week ending
6 JULY 2007
0031-9007=07=99(1)=016105(4) 016105-1 © 2007 The American Physical Society

A C, x
max
3 for A N, and x
max
2 for A O; S).
Since x
max
x is the valency of the AH
x
molecule, we
conclude that for the four families of molecules considered
the slope only depends on the valency of the adsorbate. In
the following we will consider a model that allows us to
understand the origin of this effect.
For some of the considered systems, simple valency or
bond-counting arguments [14] can explain the results:
Comparing CH, CH
2
, and CH
3
on the close-packed sur-
faces, we generally find CH (with a valency of 3) to prefer
threefold adsorption sites, CH
2
(valency of 2) to prefer
twofold adsorption, and CH
3
(valency of 1) to prefer one-
fold adsorption. The implication of these trends is that
unsaturated bonds on the carbon atom form bonds to
surface metal atoms; in effect, each unsaturated sp
3
hybrid
on the central C atom binds independently to the d states of
the nearest neighbor metal atoms, consistent with the
slopes in Fig. 1. However, this picture cannot include
adsorbed atomic C. Adsorbed C also adsorbs in a threefold
site (neglecting long range reconstructions), but it does not
have four bonds as would be needed to explain all the C
data in Fig. 1. We also note that the overall scaling behav-
ior is independent of the adsorption geometry and hence
the details of the bonding; see Fig. 2. The scaling in Figs. 1
and 2 must therefore have a more general explanation that
includes the argument above for CH, CH
2
, and CH
3
as a
special case.
We will base our analysis on the d-band model which
has been used quite successfully to understand trends in
adsorption energies from one transition metal to the next
[8,1519]. According to the d-band model, it is useful to
think of the formation of the adsorbate-surface bond as
taking place in two steps. First, we let the adsorbate states
interact with the transition metal sp states, and then we
include the extra contribution from the coupling to the d
states:
E E
sp
E
d
: (2)
The coupling to the metal sp states usually contributes the
largest part of the bonding and involves considerable hy-
bridization and charge transfer. In terms of variations from
one transition metal to the next it can, however, be consid-
ered to be essentially a constant; the sp bands are broad,
and all the transition metals have one sp electron per metal
atom in the metallic state [20]. According to the d-band
model, the main contribution to the variations in bond
energy from one transition metal to the next comes from
the coupling to the metal d states; the d states form narrow
bands of states close to the Fermi level, and the width and
energy of the d bands vary substantially between transition
metals. According to the d-band model, all the variations
among the metals observed in Fig. 1 should therefore be
given by E
d
. That means that the x dependence of
E
AH
x
x must be given by the d coupling alone: Let us
assume for the moment that the d coupling for AH
x
is
proportional to the valency parameter defined above:
E
AH
x
d
xE
A
d
(3)
Using Eq. (1), this will lead to the kind of relationship in
Fig. 1. We can write the adsorption energy of molecule
AH
x
in terms of the adsorption energy of molecule A as
-6
-5
-4 -3 -2
E
C
(eV)
-2.25
-2
-1.75
-1.5
-1.25
-1
-0.75
-0.5
E
CH
3
(eV)
Fit: y=0.24x-0.02
Fit: y=0.28x-0.27
Ir
Pt
Pd
Ni
Rh
Ru
Cu
Au
Ag
Ir
RuRh
Pt
Pd
Ni
Cu
Au
Ag
Cu
3
Pt
Pd
3
Au
Cu
3
Pt
Pd
3
Au
Ontop adsorption site
Most stable adsorption site
FIG. 2 (color). Binding energies of CH
3
plotted against the
binding energies of C for adsorption in the most stable sites
(triangles) and in the case where both CH
3
and C have been fixed
in the on-top site (squares).
FIG. 1 (color). Adsorption energies of CH
x
intermediates
(crosses: x 1; circles: x 2; triangles: x 3), NH
x
inter-
mediates (circles: x 1; triangles: x 2), OH, and SH inter-
mediates plotted against adsorption energies of C, N, O, and S,
respectively. The adsorption energy of molecule A is defined as
the total energy of A adsorbed in the lowest energy position
outside the surface minus the sum of the total energies of A in
vacuum and the clean surface. The data points represent results
for close-packed (black) and stepped (red) surfaces on various
transition-metal surfaces. In addition, data points for metals in
the fcc(100) structure (blue) have been included for OH
x
.
PRL 99, 016105 (2007)
PHYSICAL REVIEW LETTERS
week ending
6 JULY 2007
016105-2

E
AH
x
E
AH
x
d
E
AH
x
sp
xE
A
d
E
AH
x
sp
xE
A
; (4)
where E
AH
x
sp
xE
A
sp
is independent of the metal
in question. The parameter can be read off Fig. 1 for each
AH
x
=A combination; see Eq. (1). The parameter can be
obtained from calculations on any transition metal. In the
following all model data presented are obtained using
Pt(111) as the reference system.
The basic question remains, why the coupling to the d
states should scale with the valency of the adsorbate as in
Eq. (1). We cannot provide a general rigorous proof of the
scaling, Eq. (1), and most likely no such proof existsthe
scatter in Fig. 1 indicates that the linear relationship is only
approximate. What we will do, however, is show that
Eq. (1) should hold approximately for the kind of systems
investigated in Fig. 1.
The coupling of the adsorbate states to the d band has
two contributions [15,16,21], E E
hyb
d
E
orth
d
. The
first term describes the energy change associated with the
formation of bonding and antibonding states. Since the d
coupling can often be described in second order perturba-
tion theory [8,15,21], we can write E
hyb
d
/ E
orth
d
/ V
2
ad
,
where V
ad
is the Hamiltonian matrix element between the
adsorbate and the metal d states [21]. If there is more than
one adsorbate state, the total interaction energy will scale
with the sum of contributions from the adsorbate states,
V
2
ad
P
i
V
2
a
i
d
. We suggest that V
2
ad
/ for the kind of
systems considered in Fig. 1, and this directly gives the
relation of Eq. (1). The reason is given in the following.
The coupling strength V
2
ad
fr
ai
g is a function of the
number of metal neighbors and their distances to the
adsorbate. Since the d coupling is usually a minor pertur-
bation to the bond energy, it is the sp coupling which
primarily determines the adsorption bond lengths, r
ai
.
As H atoms are added to the central C, N, O, or S atom,
the adsorption bond lengths increase and the coupling
strength decreases. The effective medium theory (EMT)
[22] provides a simple way of quantifying this effect. In the
EMT the bonding of an atom A to other atoms in the
vicinity is approximated by the interaction of A with a
homogeneous electron gas (the effective medium) of a
density given by a spherical average n of the density
provided by the surrounding atoms: E E
hom
n.
Such a local density approximation for the interaction
energy gives a good description of general trends in bond-
ing, including bond lengths of atoms in metals, adsorbates
on metal surfaces, and of molecules [2224]. The energy
of embedding an atom in a homogeneous electron gas,
E
hom
n, generally has a minimum for a particular elec-
tron density, n
0
, and the equilibrium geometry of atom A is
given by the position where A experiences the optimum
electron density, n n
0
.
Consider for instance a C atom outside a metal surface.
The adsorption bond length is given by the distance outside
the surface where the electron density from the surface
around the C atom is n
surf
n
0
. Now add H atoms to the C
atom. Each H atom will provide electron density to the C
atom, and the electron density needed from the surface to
reach n n
0
is smaller. For a fixed adsorption site (one-,
two-, or threefold), the bond length between the surface
atoms and the C atom must therefore increase. Alterna-
tively, the adsorption site can change as in the case of CH,
CH
2
, and CH
3
discussed above. In that case we would then
expect the C-metal bond length to be independent of x,
since the change in metal coordination number corre-
sponds exactly to the increase in the H coordination num-
ber for this sequence of systems. This is precisely what we
find in the calculations. Returning to the general case, the
electron density from the surface n
surf
needed to obtain
n n
0
will continue to decrease as the number x of H
atoms increases. When x 4 the H atoms must contribute
all the electron density needed for the central C atom,
4n
H
n
0
, since a methane molecule does not bind to the
surface at all (neglecting van der Waals interactions). The
density contribution from the surface at the equilibrium
site for CH
x
is therefore
n
surf
x
max
x
x
max
n
0
xn
0
: (5)
The linear dependence on x in Eq. (5) is based on the
reasonable assumption that the contribution to the electron
density is the same for all x H atoms, and that the total
density should add up to n
0
[25]. The electron density n
can be viewed as a generalized bond order [26,27], and the
requirement that n n
0
is then an example of bond order
conservation.
Since we are using EMT to model the sp contribution to
the bonding, n
surf
denotes the sp electron density outside
the surface. The decay length of n
surf
outside the surface is
given asymptotically by the work function (the energy of
the Fermi level relative to vacuum). Since the d states have
energies close to the Fermi level as well, their decay length
is roughly the same. That means that to a first approxima-
tion V
2
ad
scales with n
surf
. We have therefore shown that the
following relations hold approximately,
E
d
x/V
2
ad
x/n
surf
x/
x
max
x
x
max
x; (6)
which implies Eq. (1).
Given the understanding provided above, we can try to
generalize the findings in Fig. 1. For any hydrogenation or
dehydrogenation reaction of molecules bonding to a
transition-metal surface via C, N, O, or S atoms, we should
be able to estimate the reaction energy for all transition
metals given the reaction energy for just one metal. For
each atom A
i
i 1; ...;N bonding to the surface, we
determine the change
i
in the valence parameter during
the reaction, and we can then estimate variation in the bond
energy for the full system from the variations in the bond
energies of the A
i
:
PRL 99, 016105 (2007)
PHYSICAL REVIEW LETTERS
week ending
6 JULY 2007
016105-3

E
X
N
i1
i
E
A
i
i

X
N
i1
i
E
A
i
: (7)
The change in the parameter for the particular reaction
needs to be calculated once and for all by calculating the
reaction energy for one single metal. In Fig. 3, we compare
the model to complete DFT calculations for hydrogenation
or dehydrogenation of a series of hydrocarbons, alcohols,
thiols, and amino acids. In each case, we have calculated
from Pt(111) data. The agreement between the model
and the full DFT calculations indicates that the model has
the power to describe both the absolute magnitude and the
trends in reaction energies for hydrogenation or dehydro-
genation reactions of a number of organic molecules on
transition-metal surfaces. We note that the scaling relations
can easily be generalized so that the adsorption energy of
any hydrogenated species AH
y
is used as the reference
instead of A.
By combining the present model with the Brønsted-
Evans-Polanyi type correlations that have been estab-
lished between activation barriers and reaction energies
for surface reactions [8,10,11], it will be possible to esti-
mate the full potential energy diagram for a surface cata-
lyzed reaction for any transition metal on the basis of the C,
N, O, and S chemisorption energies and a calculation for a
single metal. We suggest that this will be a useful tool in
screening for new catalysts. Such estimates can subse-
quently be followed up by full DFT calculations and ex-
periments for the most interesting systems.
The Center for Atomic-scale Materials Design is spon-
sored by the Lundbeck Foundation. Additional support
from the Danish Research Councils and the Danish
Center for Scientific Computing are also acknowledged.
[1] Z. Ma and F. Zaera, Surf. Sci. Rep. 61, 229 (2006).
[2] C. Joachim and M. Ratner, Proc. Natl. Acad. Sci. U.S.A.
102, 8801 (2005).
[3] B. Kasemo, Surf. Sci. 500, 656 (2002).
[4] G. Somorjai, Introduction to Surface Chemistry and
Catalysis (Wiley, New York, 1994).
[5] W. A. Brown, R. Kose, and D. A. King, Chem. Rev. 98,
797 (1998).
[6] H. Gross, C. Campbell, and D. A. King, Surf. Sci. 572, 179
(2004).
[7] J. K. Nørskov, M. Scheffler, and H. Toulhoat, MRS Bull.
31, 669 (2006).
[8] B. Hammer and J. K. Nørskov, Adv. Catal. 45, 71 (2000).
[9] J. Greeley and M. Mavrikakis, J. Phys. Chem. B 109, 3460
(2005).
[10] V. Pallassana and M. Neurock, J. Catal. 191, 301 (2000).
[11] A. Michaelides et al., J. Am. Chem. Soc. 125, 3704
(2003).
[12] B. Hammer, L. Hansen, and J. K. Nørskov, Phys. Rev. B
59, 7413 (1999).
[13] T. Bligaard et al., J. Catal. 224, 206 (2004).
[14] G. Papoian, J. K. Nørskov, and R. Hoffmann, J. Am.
Chem. Soc. 122, 4129 (2000).
[15] B. Hammer and J. K. Nørskov, Surf. Sci. 343, 211 (1995).
[16] B. Hammer and J. K. Nørskov, Nature (London) 376, 238
(1995).
[17] A. Eichler, F. Mittendorfer, and J. Hafner, Phys. Rev. B 62,
4744 (2000).
[18] J. Greeley and M. Mavrikakis, Nat. Mater. 3, 810 (2004).
[19] A. Roudgar and A. Gross, Phys. Rev. B 67, 033409 (2003).
[20] O. K. Andersen, O. Jepsen, and D. Glo
¨
tzel, Highlights of
Condensed Matter Theory (North-Holland, New York,
1985).
[21] B. Hammer and J. K. Nørskov, Theory of Adsorption and
Surface Reactions (Kluwer, Dordrecht, 1997).
[22] J. K. Nørskov and N. D. Lang, Phys. Rev. B 21, 2131
(1980).
[23] M. Stott and E. Zaremba, Phys. Rev. B 22, 1564 (1980).
[24] M. Puska, R. Nieminen, and M. Manninen, Phys. Rev. B
24, 3037 (1981).
[25] NH
3
, H
2
O, and H
2
S do bind weakly to the surface. This
bonding by lone pairs cannot be described in the simplest
EMT model. The bond lengths to the surface of the fully
hydrogenated species are, however, significantly larger
than for any of the less hydrogenated species, and the
density contribution from the surface is small. This is
therefore a small correction to the scaling picture devel-
oped here.
[26] E. Shustorovich, Surf. Sci. 176, L863 (1986).
[27] A. T. Bell and E. Shustorovich, J. Catal. 121, 1 (1990).
-1.5 -1 -0.5 0 0.5 1 1.5 2
E
reaction
model
(eV)
-1.5
-1
-0.5
0
0.5
1
1.5
2
E
reaction
DFT
(eV)
CH
3
OH(g)+
*
-->CH
3
O
*
+1/2H
2
(g)
C
*
H
2
-CH-C
*
H
2
-->C
*
H-CH-C
*
H
2
+1/2H
2
(g)
CH
3
SH+
*
-->CH
3
S
*
+1/2H
2
(g)
Cysteine+
*
-->S
*
-CH
2
-CH(NH
2
)-COOH+1/2H
2
(g)
C
2
H
4
+H
*
-->C
*
-CH
3
+H
2
(g)
FIG. 3 (color). Calculated reaction energies for a series of
dehydrogenation reactions plotted against the model predictions.
The model data have been generated using calculated Pt(111)
data as reference.
PRL 99, 016105 (2007)
PHYSICAL REVIEW LETTERS
week ending
6 JULY 2007
016105-4
Citations
More filters
Journal ArticleDOI
TL;DR: The emphasis of this review is on the origin of the electrocatalytic activity of nanostructured catalysts toward a series of key clean energy conversion reactions by correlating the apparent electrode performance with their intrinsic electrochemical properties.
Abstract: A fundamental change has been achieved in understanding surface electrochemistry due to the profound knowledge of the nature of electrocatalytic processes accumulated over the past several decades and to the recent technological advances in spectroscopy and high resolution imaging. Nowadays one can preferably design electrocatalysts based on the deep theoretical knowledge of electronic structures, via computer-guided engineering of the surface and (electro)chemical properties of materials, followed by the synthesis of practical materials with high performance for specific reactions. This review provides insights into both theoretical and experimental electrochemistry toward a better understanding of a series of key clean energy conversion reactions including oxygen reduction reaction (ORR), oxygen evolution reaction (OER), and hydrogen evolution reaction (HER). The emphasis of this review is on the origin of the electrocatalytic activity of nanostructured catalysts toward the aforementioned reactions by correlating the apparent electrode performance with their intrinsic electrochemical properties. Also, a rational design of electrocatalysts is proposed starting from the most fundamental aspects of the electronic structure engineering to a more practical level of nanotechnological fabrication.

3,918 citations

Journal ArticleDOI
TL;DR: The first steps towards using computational methods to design new catalysts are reviewed and how, in the future, such methods may be used to engineer the electronic structure of the active surface by changing its composition and structure are discussed.
Abstract: Over the past decade the theoretical description of surface reactions has undergone a radical development. Advances in density functional theory mean it is now possible to describe catalytic reactions at surfaces with the detail and accuracy required for computational results to compare favourably with experiments. Theoretical methods can be used to describe surface chemical reactions in detail and to understand variations in catalytic activity from one catalyst to another. Here, we review the first steps towards using computational methods to design new catalysts. Examples include screening for catalysts with increased activity and catalysts with improved selectivity. We discuss how, in the future, such methods may be used to engineer the electronic structure of the active surface by changing its composition and structure.

3,023 citations


Cites background from "Scaling properties of adsorption en..."

  • ...The number of independent variables is limited strongly by the fact that it has been found that adsorption energies for a number of molecules scale with each othe...

    [...]

Journal ArticleDOI
TL;DR: In this article, a large database of HO* and HOO* adsorption energies on oxide surfaces was used to analyze the reaction free energy diagrams of all the oxides in a general way.
Abstract: Trends in electrocatalytic activity of the oxygen evolution reaction (OER) are investigated on the basis of a large database of HO* and HOO* adsorption energies on oxide surfaces. The theoretical overpotential was calculated by applying standard density functional theory in combination with the computational standard hydrogen electrode (SHE) model. We showed that by the discovery of a universal scaling relation between the adsorption energies of HOO* vs HO*, it is possible to analyze the reaction free energy diagrams of all the oxides in a general way. This gave rise to an activity volcano that was the same for a wide variety of oxide catalyst materials and a universal descriptor for the oxygen evolution activity, which suggests a fundamental limitation on the maximum oxygen evolution activity of planar oxide catalysts.

2,923 citations

Journal ArticleDOI
TL;DR: A new set of ORR electrocatalysts consisting of Pd or Pt alloyed with early transition metals such as Sc or Y, identified using density functional theory calculations as being the most stable Pt- and Pd-based binary alloys with ORR activity likely to be better than Pt.
Abstract: The widespread use of low-temperature polymer electrolyte membrane fuel cells for mobile applications will require significant reductions in the amount of expensive Pt contained within their cathodes, which drive the oxygen reduction reaction (ORR). Although progress has been made in this respect, further reductions through the development of more active and stable electrocatalysts are still necessary. Here we describe a new set of ORR electrocatalysts consisting of Pd or Pt alloyed with early transition metals such as Sc or Y. They were identified using density functional theory calculations as being the most stable Pt- and Pd-based binary alloys with ORR activity likely to be better than Pt. Electrochemical measurements show that the activity of polycrystalline Pt(3)Sc and Pt(3)Y electrodes is enhanced relative to pure Pt by a factor of 1.5-1.8 and 6-10, respectively, in the range 0.9-0.87 V.

2,588 citations

Journal ArticleDOI
TL;DR: A current snapshot of high-throughput computational materials design is provided, and the challenges and opportunities that lie ahead are highlighted.
Abstract: High-throughput computational materials design is an emerging area of materials science. By combining advanced thermodynamic and electronic-structure methods with intelligent data mining and database construction, and exploiting the power of current supercomputer architectures, scientists generate, manage and analyse enormous data repositories for the discovery of novel materials. In this Review we provide a current snapshot of this rapidly evolving field, and highlight the challenges and opportunities that lie ahead.

1,568 citations

References
More filters
Journal ArticleDOI
TL;DR: In this paper, a simple formulation of a generalized gradient approximation for the exchange and correlation energy of electrons has been proposed by Perdew, Burke, and Ernzerhof (PBE), which improves the chemisorption energy of atoms and molecules on transition-metal surfaces.
Abstract: A simple formulation of a generalized gradient approximation for the exchange and correlation energy of electrons has been proposed by Perdew, Burke, and Ernzerhof (PBE) [Phys. Rev. Lett. 77, 3865 (1996)]. Subsequently Zhang and Yang [Phys. Rev. Lett. 80, 890 (1998)] have shown that a slight revision of the PBE functional systematically improves the atomization energies for a large database of small molecules. In the present work, we show that the Zhang and Yang functional (revPBE) also improves the chemisorption energetics of atoms and molecules on transition-metal surfaces. Our test systems comprise atomic and molecular adsorption of oxygen, CO, and NO on Ni(100), Ni(111), Rh(100), Pd(100), and Pd(111) surfaces. As the revPBE functional may locally violate the Lieb-Oxford criterion, we further develop an alternative revision of the PBE functional, RPBE, which gives the same improvement of the chemisorption energies as the revPBE functional at the same time as it fulfills the Lieb-Oxford criterion locally.

5,971 citations

Journal ArticleDOI
20 Jul 1995-Nature
TL;DR: In this paper, a simple surface reaction, the dissociation of H2 on the surface of gold and of three other metals (copper, nickel and platinum) that lie close to it in the periodic table, was studied.
Abstract: THE unique role that gold plays in society is to a large extent related to the fact that it is the most noble of all metals: it is the least reactive metal towards atoms or molecules at the interface with a gas or a liquid. The inertness of gold does not reflect a general inability to form chemical bonds, however—gold forms very stable alloys with many other metals. To understand the nobleness of gold, we have studied a simple surface reaction, the dissociation of H2 on the surface of gold and of three other metals (copper, nickel and platinum) that lie close to it in the periodic table. We present self-consistent density-functional calculations of the activation barriers and chemisorption energies which clearly illustrate that nobleness is related to two factors: the degree of filling of the antibonding states on adsorption, and the degree of orbital overlap with the adsorbate. These two factors, which determine both the strength of the adsorbate-metal interaction and the energy barrier for dissociation, operate together to the maxima] detriment of adsorbate binding and subsequent reactivity on gold.

2,721 citations

Book
01 Jan 1994
TL;DR: The Structure of Surfaces: An Introduction as mentioned in this paper The structure of surfaces and its properties are discussed in detail in Section 5.1.2.3 The Dynamics at Surfaces. 4 Electrical properties of surfaces. 5 Surface Chemical Bond.
Abstract: Preface. Introduction. 1 Surfaces: An Introduction. 2 The Structure of Surfaces. 3 Thermodynamics of Surfaces. 4 Dynamics at Surfaces. 5 Electrical Properties of Surfaces. 6 Surface Chemical Bond. 7 Mechanical Properties of Surfaces. 8 Polymer Surfaces and Biointerfaces. 9 Catalysis by Surfaces. Index.

2,346 citations

Journal ArticleDOI
TL;DR: In this article, the authors present a consistent picture of some key physical properties determining the reactivity of metal and alloy surfaces, and suggest that trends in reactivities can be understood in terms of the hybridization energy between the bonding and anti-bonding adsorbate states and the metal d-bands (when present).

2,008 citations

Journal ArticleDOI
TL;DR: In this paper, it was shown that the reaction rate under given reaction conditions shows a maximum as a function of dissociative adsorption energy of the key reactant, and that for most conditions this maximum is in the same range of reaction energies.

1,218 citations