Você está na página 1de 13

Downloaded by University of Lancaster on 17/01/2015 21:41:48.

Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

CHAPTER 15

Electronic Structure and


Bonding of Water to Noble
Metal Surfaces
HIROHITO OGASAWARA* AND ANDERS NILSSON
SLAC National Accelerator Laboratory, 2575 Sand Hill Rd, Menlo Park,
CA 94025, USA
*Email: hirohito@slac.stanford.edu

15.1 Introduction
Photoelectrochemical (PEC) cells are often designed with semiconductors as
photoelectrodes and noble metals as counter electrodes. The counter reaction
on noble metal surface should be fast to avoid any performance limitations. At
the interface, water interacts both with the noble metal surfaces and via
hydrogen (H-) bonding with other water molecules.1,2 Here, we will focus on
the interaction of water on metal surfaces with a detailed discussion about the
electronic structure and the resulting bonding mechanism. We anticipate that in
particular the electronic structure of the water-surface interaction could have
some similarity for water on semiconductor surfaces.
The interaction of water on metal surfaces has been the center of an extended
debate during the last decade due to the multitude of bonding and overlayer
structures water assumes on dierent single crystal surfaces. On metal surfaces,
water forms two-dimensional hexagonal, or pseudo-hexagonal, H-bond networks in the rst contact layer,1,2 where the H atoms not involved in the twodimensional hydrogen-bond network are either directed toward vacuum (H-up)
RSC Energy and Environment Series No. 9
Photoelectrochemical Water Splitting: Materials, Processes and Architectures
Edited by Hans-Joachim Lewerenz and Laurence Peter
r The Royal Society of Chemistry 2013
Published by the Royal Society of Chemistry, www.rsc.org

406

View Online

Electronic Structure and Bonding of Water to Noble Metal Surfaces

407

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

or towards the surface (H-down). The detailed structure of the rst contacting
layer at a metal surface will aect barriers to dissociation and surface reactivity,
including interaction with additional water layers.3,4

15.2 H-up, H-down and Partially Dissociated Water


Layers
In the traditional structural model proposed by Doering and Madey,5 water
was considered to bind to metal surfaces exclusively via an oxygen lp orbital.
The structure of the internally H-bonded water contact layer on metal surfaces
was consequently considered to be the H-up structure where only every second
water binds to the metal through the metal-oxygen (M-O) bond (H-up,
Figure 15.1) while the other half is displaced towards vacuum and pointing the
non-hydrogen-bonded OH group away from the surface towards vacuum.
Recent work utilizing x-ray photoemission spectroscopy (XPS) and x-ray
absorption spectroscopy (XAS) to investigate the structure of the contact layer68
has, however, showed that an H-down layer, where all water molecules in the
layer bind directly to the surface through alternating either M-O or metalhydrogen (M-HO) bonds, is favored for intact adsorption of water on these
close-packed metal surfaces. However, the details vary for dierent surfaces in
terms of the long-range order in the H-up and H-down congurations.9,10
X-ray photoelectron spectroscopy (XPS) is based on the creation of a core
hole via ionization and provides a method to study the geometric, electronic
and chemical properties of a sample. The XPS binding energies depend
very strongly on the elements involved; even for rst-row elements the corelevel binding energies dier on the order of 100 eV making XPS a highly
H-up

Figure 15.1

Mixed H2O:OH

H-down

Illustrations of three proposed structural models for water adsorption on


metal surfaces. All models contain a at-lying O-bonded water (red/
lighter) but dier in orientation and chemical nature of the second
molecule (blue/darker). H-up: Traditional ice-like bilayer with the
blue/darker water molecule having one of its Hs pointing toward vacuum,
Mixed H2O:OH: partially dissociated water layer with at-lying H2O (red/
lighter) and at-lying OH (blue/darker), and H-down: non-O-bonded
water molecule has one of its Hs pointing toward the metal surface.
Reprinted from reference 8. Copyright 2010, with permission from
Elsevier.

View Online

Chapter 15

element-specic technique. In the case of water, XPS monitors the binding


energy of O1s core-level state. The binding energy of the O1s state is aected by
the valence electrons, which in turn are sensitive to the local environment. We
can therefore expect that the core-levels will be chemically shifted depending on
chemical nature (intact or dissociated), adsorption site, distance to the surface
and molecular orientations. In x-ray absorption spectroscopy (XAS), a core
electron is resonantly excited into unoccupied atomic or molecular orbitals at
or above the Fermi level via a dipole-induced transition.11 XAS provides
element-specic information on the density and the energy level of unoccupied
states, local atomic structure including molecular orientation, the nature,
orientation, and length of chemical bonds (via bonding-antibonding orbital
splitting12) as well as the chemical nature. Using s- and p- polarized x-ray light
eld, XAS measurements determine the orientation of molecular adsorbates
and the directionality of bonding in an adlayer, including proton orientation,
information often unattainable by other spectroscopic probes.11
The work on Pt(111)6 and Ru(0001)7 employed XPS to address the coordination of atoms in the molecularly intact monolayer to the surface (see
Figure 15.2) and XAS to determine the orientation and coordination of the
internal OH-groups in water with respect to the surface. While the H-up
structure (Figure 15.1, H-up) would give only 33% surface coordination
(a)

536

(b)

XPS Ru 3d5/2

534

S =1 M L

S=1 ML

73

72

Figure 15.2

71

S=1/3 ML

70
281
Binding Energy (eV)

S=1/3 ML
B

280

279

Intensity (arb. units)

XPS O 1s
150 K

Intensity (arb. units)

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

408

538

532

530

D2O

Pt(111)

Ru(0001)

536 534 532 530


Binding Energy (eV)

XPS spectra for clean and water-covered Pt(111) and Ru(0001), The
essential results showing low concentration of non-coordinated surface
Pt and Ru atoms (a) and non-dissociated water (b) can be seen directly in
the experimental spectra. A qualitative curve tting analysis (black lines)
in a is indicated to guide the eye. (a): (Top) (Left) Summed Pt 4f7/2
spectra taken at three excitation energies (115, 125 and 135 eV) and
(Right) summed Ru 3d5/2 spectra taken at three excitation energies (380,
390 and 400 eV) to average out photoelectron diraction eects. The
bulk (B) and uncoordinated surface (S) spectral components are indicated. (Bottom) Water coordinated to the surface atoms quenches
intensity of uncoordinated surface atoms (S) and introduces core-level
shifts (new spectral components) compared to the clean surface. The
uncoordinated surface peak (S) is lowered in intensity by more than 60%.
Reprinted from reference 8. Copyright 2010, with permission from
Elsevier.

View Online

409

through M-O bonds, the H-down structure (Figure 15.1, H-down) would give
67% surface coordination through the combination of M-O and M-HO bonds.
The surface coordination number can be experimentally determined by
exploiting the surface core-level shift in XPS.6,7,1315 For clean metals, the lower
coordination of atoms at the surface leads to a dierent core-level binding
energy compared to the bulk.14 On Pt(111), this splitting is 0.4 eV for the Pt 4f
photoemission peak6 (Figure 15.4a, left panel). The introduction of water on
Pt(111) shifts the Pt 4f surface state towards the bulk value for the atoms that
now coordinate to water. The change in XPS intensity of the Pt 4f state for the
adsorbate system compared to the clean surface indicates that more than 60%
of the surface Pt atoms become coordinated to water molecules. This directly
and strongly indicates the H-down layer for Pt(111), where all water molecules
in the contact layer bind to the surface.
On Ru(0001), the non-dissociated water contact layer forms a hexagonal
two-dimensional hydrogen-bond network similar to the case of Pt(111). The
surface coordination number was determined using the surface core level shift
in the Ru 3d photoemission peak7 (Figure 15.2a, right panel), which is sensitive
to changes in local coordination similar to the Pt 4f case. As on Pt(111), more
than 60% of the surface Ru atoms become coordinated, directly showing that
all water molecules in the monolayer bind to atoms at the Ru(0001) surface.
Based on the same coordination of the water layer to Ru(0001) as for the
water layer on Pt(111) and near identical O 1s XPS spectra (see Figure 15.2b),
an H-down model is suggested also for the non-dissociated water contact layer
on Ru(0001).7

IR OD

XAS O1s

Intensity (arb. units)

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

Electronic Structure and Bonding of Water to Noble Metal Surfaces

R/R=0.005

Pt(111)

Cu(110)

ice surface

532 536 540 544


Photon Energy (eV)

Figure 15.3

2800 2600 2400 2200


Wavenumber/cm1

XAS (Reprinted from reference 8. Copyright 2010, with permission from


Elsevier.) and IR spectra24 for D2O ice surface and D2O water monolayer
on Pt(111) and Cu(110). The light eld oscillates in the direction normal
to the surface (p-polarized).

View Online

Chapter 15

The question emerges whether the H-down model is valid in general for
water-metal monolayer adsorbate systems, or does the overlayer structure vary
depending on surface structure? On the open and corrugated (110) surface of
copper, it was found that a mixed monolayer with 2/3 (  15%) of the outerlayer water molecules in H-down conguration, and 1/3 with hydrogen
pointing up toward the vacuum. Thus we nd an H-down:H-up ratio of 2:1 for
the water monolayer on Cu(110).16 The spectroscopic indication of a mixed
H-up/H-down layer is consistent with the large (78) unit cell for monolayer
water on Cu(110),16 indicating that the hexagonal adlayer is rather distorted
with respect to the open substrate which can be expected to lead to a range of
adsorption sites.
The orientation of the uncoordinated OH is conrmed by XAS, which can
selectively probe the local unoccupied orbital structure in dierent directions in
the layer by using s- and p-polarized x-rays, whose x-ray light eld is either
parallel or orthogonal to the surface. The interaction between water and metal
surface, with the possible formation of M-HO bonds, is probed in the outof-plane XAS while in-plane XAS is related to the formation of the twodimensional hydrogen-bond network in the contact layer. There is, indeed, a
strong anisotropy in XAS between aligning the E-vector along the surface
normal (out-of-plane) and parallel to the surface (in-plane) (see reference 17.
for the Pt(111) case). If we had the H-up situation, in which non-hydrogen
bonded OH group of water, free OH, are present on the surface, we should an

XPS O 1s

h=785 eV

D2O/H2O
150 K
OD/OH
Energy

Intensity (arb. units)

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

410

H2O(g)
dissociation

desorption
H2O(a)
dE

180 K
H(a)+OH(a)
538 536 534 532 530
Binding Energy (eV)

Figure 15.4

Reaction Coordinate

XPS for water (H2O and D2O) adsorbed on Ru(0001)7 (left) and schematic
illustration of desorption and dissociation barriers for H2O on Ru(0001)
(right). Chemical species on Ru(0001) are identied through chemical
shifts in the O 1s XPS. The formation of hydroxyl at 180 K is characterized
by the appearance of an O 1s XPS peak at ca. 531 eV, whereas that for
intact water appears at ca. 532533 eV. The decomposition is observed for
H2O but not for D2O due to zero-point energy dierences.
Reprinted from reference 8. Copyright 2010, with permission from
Elsevier.

View Online

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

Electronic Structure and Bonding of Water to Noble Metal Surfaces

411

XAS signature of free OH. There are abundant amount of free OH on the
surface of ice water, which gives rise to the peak at 535 eV assignable to orbitals
localized at free OH groups. Another way to probe the orientation of the nonhydrogen-bonded OH group is via vibrational excitation of the OH vibrational
mode. Several techniques have been utilized to probe the O-H stretch vibration
of adsorbed water: IR absorption, electron energy loss and sum frequency
generation.16,1820 The ngerprint of non-hydrogen-bonded OH groups
pointing towards the vacuum is an isolated high-frequency free OH band at
3680 cm 1 (2730 cm 1 for deuterated water). XAS and IR studies on ammonia
terminated ice surface indicate that these states almost exclusively reside at the
ice surface.21,22 In Figure 15.3, we compare p-polarized XAS and IR results for
the surface of ice water23 and the water monolayer on Pt(111) and
Cu(110);6,16,24 the polarization shown is with the light eld oscillating normal
to the surface (z-direction), i.e. in the direction of the free OH-groups. On
Pt(111), OH groups are no longer uncoordinated resulting in broadened feature
and the loss of intensity. The strong feature at 532 eV is attributed to molecules
binding through oxygen (Pt-O) and the feature at 534 eV to the binding through
hydrogen (Pt-HO), which will be discussed in the next section. On the other
hand, a notable amount of free OH feature remains in XAS for the water
monolayer on Cu(110) with H-down:H-up ratio of 2:1, The IR absorption of
free OH is intense for water adsorbed on Cu(110)16 but negligible on both
Pt(111),18,25,26 which corroborates with the presence of free OH on the mixed
H-down:H-up water layer on Cu(110), but not on the H-down water layer on
Pt(111) .

15.3 Competition Between Thermal Dissociation and


Desorption
Under certain circumstances, the O-H bond of adsorbed water dissociates.
Compilations of studies of water on metal surfaces1,2 show that Ru(0001) and
Cu(110) are on the border between active and inactive metal surfaces with
respect to dissociation of water. The dissociation of water on these surfaces is
supported by the appearance of two dierent O 1s XPS peaks assignable to,
respectively, water and hydroxyl, see Figure 15.4 (left). Although a dissociated
layer is thermodynamically favorable on Ru(0001) and Cu(110), the dissociation must overcome activation barriers. The relative heights among activation
barriers for dierent reaction play in determining the dissociation probability.
The structure of water on Ru(0001) was an issue of debate.7,19,20,2734 Feibelman34 found that a partially dissociated layer consisting of a near-planar
hexagonal mixed network of adsorbed water and hydroxyl (Mixed H2O:OH in
Figure 15.1). There is, however, an activation barrier that impedes the decomposition of water, as depicted in the schematic potential energy diagram in
Figure 15.4 (right). It was found on Ru(0001) that dissociation and desorption
of water occur with very similar barriers, and the probability of dissociation
is thus ruled by the balance between desorption and dissociation kinetics.7

View Online

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

412

Chapter 15

An indication of this delicate balance between dissociation and desorption is


found from the anomalous isotope eect and kinetics in the thermal desorption
spectra of water on Ru(0001).3539 The isotope eect arises from dierences in
the zero-point vibrational energy contribution to the dissociation barrier between H2O and D2O. The dissociation pathway involves elongation of an O-H
or O-D bond. H2O has a 0.1 eV higher zero-point vibrational energy compared
to D2O in the dissociative pathway. The bonding to the surface, on the other
hand, is equivalent for the two isotopes giving similar barriers to desorption.
The zero-point contribution directly aects the barrier to dissociation, which
will be only slightly higher than desorption barrier for H2O but signicantly
more so for D2O. No dissociation is observed on the surfaces of neighboring
elements to the right in the periodic table, e.g. Ni(111), Cu(111), Rh(111) and
Pt(111), for which the barrier to dissociation thus becomes signicantly larger
than the desorption barrier.1,2

15.4 Electronic Structure and Bonding Mechanism


Photoelectron spectroscopy (PES) has been used to probe the valence electronic
structure of water. PES determines the binding energy and character of the
dierent occupied molecular orbitals can be determined. In the PES spectra of
gas phase water,40,41 see Figure 15.5, peaks in the valence region were assigned

h=100eV

PES
1b2

1b1

3a1
x0.2

gas

ice

15
10
5
Binding Energy (eV)

Figure 15.5

PES spectrum of gas phase water measured at a photon energy of


100 eV40 and PES spectra of ice at photon energies of 100 eV.43
Reprinted with permission from reference 43. Copyright 2005, American
Institute of Physics.

View Online

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

Electronic Structure and Bonding of Water to Noble Metal Surfaces

413

to 1b2, 3a1 and 1b1 states of water. The highest occupied state (1b2) has a nonbonding character and highly localized on the oxygen atom, as it has a lobe
pointing away from the two hydrogen atoms. The third lowest occupied state
(1b1) has a lobe pointing toward two hydrogen atoms, which bonds the O and
H atoms (O-H). Though the second lowest occupied state (3a1) also has an O-H
bonding character, it has a lobe pointing away from two hydrogen atoms,
which can be thought as a lp character.
X-ray emission spectroscopy (XES) has also been used to determine the
binding energy and character of the dierent occupied molecular orbitals. One
of the main advantages of XES to PES is that it provides element specicity. In
the case of water the XES process involves the projection of the valence electronic state onto the oxygen 1s core state of water. Since the spatially localized
character of oxygen 1s orbital on an atom, it provides a tool to probe the
molecular orbitals of water selectively and no contribution comes from the
substrate.
In an aqueous electrolyte solution, water makes hydrogen bonds to surrounding water and ions. Though the strength of hydrogen bond is weak,
generally B0.25 eV per bond, the valence states of water is altered upon
hydrogen bonding. The eect of hydrogen bonding to the valence states is
demonstrated by comparing gas phase and ice water PES results in Figure 15.5.
The 3a1 state undergoes the most prominent change of substantial broadening,
which have also seen for PES study in liquid water.40,42 This is due to the
electron re-distribution inside the water molecules to minimize the Pauli repulsion upon hydrogen bonding.43 PES 40,42 and XES 44,45 studies of the liquid
phase have produced similar results.
Metal and semiconductor materials have valence and conduction electrons.
The valence electrons are bound to individual atoms, as opposed to conduction
electrons, which move freely within the material. We now consider the role of
these electrons in the bonding mechanism of water to the metal through XES
studies.
X-ray emission spectroscopy (XES) studies on water on Pt(111) showed the
Pt-O and Pt-HO bonding mechanism of water in a site-specic way.6 For XES
measurements, a core-hole is created in the 1s orbital of water through the x-ray
absorption process. The core-hole will subsequently be lled through the dipole
transition of valence electronic states resulting in emission of x-rays; therefore
the intensity of the x-ray emission peak corresponds to the p component of the
valence electronic states. The s-polarized x-ray emission spectra, where the
x-ray light is emitted in the surface plane, correspond to the oxygen 2p components projected in the surface plane,46 and the electronic states involved in
the two-dimensional hydrogen-bond network of the contact layer are probed.
These spectra8 are very similar to those of bulk ice.47,48 The p-polarized x-ray
emission spectra, whose x-ray light eld is parallel to the surface normal,
corresponds to the oxygen 2p components projected along the surface normal
and a component parallel to the surface plane. On Pt(111), water molecules are
alternately Pt-O and Pt-HO bonding to the surface. By tuning the excitation
photon energy, we can selectively excite either Pt-O or Pt-HO bonding water,

View Online

Chapter 15

projecting the occupied electronic states on the oxygen atom of respectively the
Pt-O and Pt-HO bonding species (Figure 15.6). The bonding mechanism shown
in the insert of Figure 15.6 is proposed based on the analysis of spectral features
in XES combined with electronic structure calculations.6 The interaction of the
O lp orbital (1b1) with the valence electrons in Pt d-orbital form bonding Pt-O
state and anti-bonding Pt-O* states, which appears in the vicinity the Fermi
level. Here the bond strength is predicted by the d-band model;49 in which the
degree of d-band population and the position of d-band center are important.
In the case of water on Pt, a partially unlled nature of Pt 5d-band makes the
Pt-O* state partially unlled. The depopulation is seen as a peak at 532.5 eV in
the out-of-plane XAS spectrum in Figure 15.7.
The closed-shell d10 conguration of Cu surfaces, however, does not aord
this mechanism.50 The decomposition, and comparison with water on
Cu(110)16 (Figure 15.7, highlights the inuence on the bonding by the d-band
position with respect to the Fermi level as depicted in the inset. Let us now
consider the M-O bonding channel. The valence electrons of Cu occupy the 3d
Binding Energy(BE)/eV
15

10

15

Pt-O

XES

10

Pt-HO
OH*

Pt5d

Olp
OH

Int. [arb. units]

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

414

x4

515

520

525

530

x4

515

520

525

530

Photon Energy/eV

Figure 15.6

X-ray emission spectra (p-polarized) from Pt-O and Pt-HO bonding


water showing the occupied orbital structure out-of-plane (pz components).6 The Pt-O bonding water has an x-ray absorption threshold at
lower energy than that for the Pt-HO species bonding through hydrogen.
This allows separated XES spectra for the Pt-O and Pt-HO bonding
water to be obtained by using two excitation energies (532 eV and 538 eV)
and a subtraction procedure. The inset shows schematic diagrams of the
Pt-O and Pt-HO bonds.
Reprinted from reference 6. Copyright 2002 by the American Physical
Society.

View Online

415

Electronic Structure and Bonding of Water to Noble Metal Surfaces


Pt(111)

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

Pt5d-Olp*

M-O
Pt5d-Olp*
Pt-O

EF
Olp

Pt5d
Pt-HO
Pt5d-Olp
Cu(110)

Cu3d-Olp*
Cu-O

EF
Olp

Cu3d
Cu-HO

Cu3d-Olp

532
536
540
544
Photon Energy (eV)

Figure 15.7

X-ray absorption spectrum (p-polarized) for water on Pt(111) (top) and


Cu(110) (bottom) and computed x-ray absorption spectra for M-O and
M-HO bonded water on each surface corresponding to the pz component.50
Reprinted from reference 8. Copyright 2010, with permission from
Elsevier. The inset shows schematic diagrams of the Pt-O and Cu-O
bonds.

orbitals more spatially contracted than the 5d orbitals of Pt. Moreover, high
density of conduction electron in Cu causes the Pauli repulsion with the electrons of water, which will inhibit the approach of O lp to the 3d orbitals of Cu.
These eects give rise to a smaller splitting between the bonding and antibonding states on Cu compared to on Pt as illustrated in Figure 15.7. In the
interaction with the closed-shell water lone pair, both bonding and antibonding states become fully occupied leading to Pauli repulsion. Accordingly,
no such O-bonding related peak appears in the XAS for Cu surfaces resulting in
no net attractive interaction in the M-O bonding channel. The s-electrons are
much more mobile and can easily move away from the bonded metal atom
towards neighboring atoms to minimize the overall repulsion. This can be
described in a simplistic way as that the water lone-pair digs a hole in the
s-band50 as shown schematically in Figure 15.8. Since there is now a partial
positive charge on the metal atom, the lone pair orbital will be stabilized
through electrostatic interaction, often denoted dative bonding. This provides
the main surface bonding mechanism for water and describes general lone-pair

View Online

416

Chapter 15

Olp

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

s-band

dig a hole

Cu3d

Figure 15.8

Schematic illustration of the water O lp orbital digging the hole in the


metal sp-band to open for water-metal dative bonding.

interactions on surfaces.50 While no chemisorbed water layer is observed on


Cu(111),50 it is observed on corrugated Cu(111) in which the depleted density of
the s-electrons at atomic rows on Cu(110) lowers the energy costs to dig the
hole in the s-band.

15.5 Conclusions
The structure of water at noble metal surfaces results from a complex interplay
of a number of eects, including the balance between water-water and watermetal bond strength, which will directly aect barriers to dissociation, desorption and other catalytic reactions. The localized valence electrons and the
mobile conduction electrons play important roles in the bonding. Localized
valence electrons facilitate the covalent bonding to water forming a bonding
and anti-bonding combination. The population and position of valence band
are parameter to tune nature of anti-bonding state in the vicinity of the Fermi
level. The Pauli repulsion by the conduction electron hampers the approach of
water to the valence electrons, which can be reduced by the geometric eects.
The mechanistic picture of bonding of water to noble metals is anticipated to be
applicable also to semiconductors in PEC materials.

Acknowledgements
This work was supported by Oce of Basic Energy Sciences, US Department
of Energy under contract DE-AC02-76SF00515. Portions of this research were
carried out at the Stanford Synchrotron Radiation Lightsource (SSRL), a
division of SLAC National Accelerator Laboratory and an Oce of Science
user facility operated by Stanford University for the U.S. Department of Energy. The results discussed in this review have naturally been obtained in collaboration with a large number of extraordinary scientists, and we like to in
particular thank Klas Andersson, Lars GM Pettersson and Theanne Schiros.

References
1. P. Thiel and T. Madey, Surf. Sci. Rep., 1987, 7, 211.
2. M. Henderson, Surf. Sci. Rep., 2002, 46, 1.

View Online

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

Electronic Structure and Bonding of Water to Noble Metal Surfaces

417

3. G. A. Kimmel, N. G. Petrik, Z. Dohnalek and B. D. Kay, Phys. Rev. Lett.,


2005, 95, 166102.
4. G. A. Kimmel, N. G. Petrik, Z. Dohnalek and B. D. Kay, J. Chem. Phys.,
2007, 126, 114702.
5. D. L. Doering and T. E. Madey, Surf. Sci., 1982, 123, 305.
6. H. Ogasawara, B. Brena, D. Nordlund, M. Nyberg, A. Pelmenschikov,
L. G. M. Pettersson and A. Nilsson, Phys. Rev. Lett., 2002, 89, 276102.
7. K. Andersson, A. Nikitin, L. G. M. Pettersson, A. Nilsson and
H. Ogasawara, Phys. Rev. Lett., 2004, 93, 196101.
8. T. Schiros, K. J. Andersson, L. G. M. Pettersson, A. Nilsson and
H. Ogasawara, J. Electron Spect. Rel. Phen., 2010, 177, 8598.
9. A. Hodgson and S. Haq, Surf. Sci. Rep., 2009, 64, 381451.
10. J. Carrasco, A. Hodgson and A. Michaelides, Nat. Mater., 2012, 11,
667674.
11. J. Stohr, NEXAFS spectroscopy, Springer-Verlag, Berlin Heidelberg,
1992.
12. J. Stohr, F. Sette and A. L. Johnson, Phys. Rev. Lett., 1984, 53, 1684.
13. O. Bjorneholm, A. Nilsson, H. Tillborg, P. Bennich, A. Sandell,
B. Hernnas, C. Puglia and N. Martensson, Surf. Sci., 1994, 315, L983.
14. A. Nilsson, A. Stenborg, H. Tillborg, K. Gunnelin and N. Martensson,
Phys. Rev. B, 1993, 47, 13590.
15. N. Martensson and A. Nilsson, J. Electron Spect. Rel. Phen., 1996, 75, 209.
16. T. Schiros, S. Haq, H. Ogasawara, O. Takahashi, H. Ostrom,
K. Andersson, L. G. M. Pettersson, A. Hodgson and A. Nilsson, Chem.
Phys. Lett., 2006, 429, 415419.
17. T. Schiros, L.-A. Naslund, K. Andersson, J. Gyllenpalm, G. S. Karlberg,
M. Odelius, H. Ogasawara, L. G. M. Pettersson and A. Nilsson, J. Phys.
Chem. C, 2007, 111, 15003.
18. S. Haq, J. Harnett and A. Hodgson, Surf. Sci., 2002, 505, 171182.
19. C. Clay, S. Haq and A. Hodgson, Chem. Phys. Lett., 2004, 388, 8993.
20. D. N. Denzler, C. Hess, R. Dudek, S. Wagner, C. Frischkorn, M. Wolf and
G. Ertl, Chem. Phys. Lett., 2003, 376, 618624.
21. P. Wernet, D. Nordlund, U. Bergmann, M. Cavalleri, M. Odelius,
H. Ogasawara, L. A. Naslund, T. K. Hirsch, L. Ojamae, P. Glatzel,
L. G. M. Pettersson and A. Nilsson, Science, 2004, 304, 995999.
22. H. Ogasawara, N. Horimoto and M. Kawai, J. Chem. Phys., 2000, 112,
82298232.
23. D. Nordlund, H. Ogasawara, P. Wernet, M. Nyberg, M. Odelius,
L. G. M. Pettersson and A. Nilsson, Chem. Phys. Lett., 2004, 395, 161165.
24. H. Ostrom, A. Nillson, M. Kawai and H. Ogasawara, unpublished.
25. H. Ogasawara, J. Yoshinobu and M. Kawai, Chem. Phys. Lett., 1994,
231, 188.
26. M. Nakamura, Y. Singaya and M. Ito, Chem. Phys. Lett., 1999, 309, 123.
27. D. Menzel, Science, 2002, 295, 58.
28. D. N. Denzler, S. Wagner, M. Wolf and G. Ertl, Surf. Sci., 2003,
532535, 113.

View Online

Downloaded by University of Lancaster on 17/01/2015 21:41:48.


Published on 02 October 2013 on http://pubs.rsc.org | doi:10.1039/9781849737739-00406

418

Chapter 15

29. A. Michaelides, A. Alavi and D. A. King, J. Am. Chem. Soc., 2003,


125, 2746.
30. J. Weissenrieder, A. Mikkelsen, J. N. Andersen, P. J. Feibelman and
G. Held, Phys. Rev. Lett., 2004, 93, 196102.
31. S. Meng, E. G. Wang, C. Frischkorn, M. Wolf and S. Gao, Chem. Phys.
Lett., 2005, 402, 384.
32. G. Materzanini, G. F. Tantardini, P. J. Lindan and P. Saalfrank, Phys.
Rev. B, 2005, 71, 155414.
33. N. S. Faradzhev, K. L. Kostov, P. Feulner, T. E. Madey and D. Menzel,
Chem. Phys. Lett., 2005, 415, 165.
34. P. J. Feibelman, Science, 2002, 295, 99.
35. D. N. Denzler, S. Wagner, M. Wolf and G. Ertl, Surf. Sci., 2003, 544,
348348.
36. D. N. Denzler, S. Wagner, M. Wolf and G. Ertl, Surf. Sci., 2003, 532535,
113119.
37. W. Homann and C. Benndorf, Surf. Sci., 1997, 377379, 681686.
38. G. Held and D. Menzel, Surf. Sci., 1995, 327, 301320.
39. P. J. Schmitz, J. A. Polta, S. L. Chang and P. A. Thiel, Surf. Sci. Lett.,
1987, 186, 283284.
40. B. Winter, R. Weber, W. Widdra, M. Dittmar, M. Faubel and I. V. Hertel,
J. Phys. Chem. A, 2004, 108, 26252632.
41. S. Myneni, Y. Luo, L. A. Naslund, M. Cavalleri, L. Ojamae,
H. Ogasawara, A. Pelmenschikov, W. Ph, P. Vaterlein, C. Heske,
Z. Hussain, L. G. M. Pettersson and A. Nilsson, J. Phys.-Condens. Mat.,
2002, 14, L213.
42. M. Faubel, B. Steiner and J. P. Toennies, J. Chem. Phys., 1997, 106,
90139031.
43. A. Nilsson, H. Ogasawara, M. Cavalleri, D. Nordlund, M. Nyberg,
P. Wernet and L. G. M. Pettersson, J. Chem. Phys., 2005, 122, 154505
154509.
44. J. H. Guo, Y. Luo, A. Augustsson, J. E. Rubensson, C. Sathe, H. Agren,
H. Siegbahn and J. Nordgren, Phys. Rev. Lett., 2002, 89, 137402.
45. S. Kashtanov, A. Augustsson, Y. Luo, J. H. Guo, C. Sathe,
J. E. Rubensson, H. Siegbahn, J. Nordgren and H. Agren, Phys. Rev. B,
2004, 69, 024201.
46. A. Nilsson and L. G. M. Pettersson, Surf. Sci. Rep., 2004, 55, 49.
47. B. Brena, D. Nordlund, M. Odelius, H. Ogasawara, A. Nilsson and
L. G. M. Pettersson, Phys. Rev. Lett., 2004, 93, 148302.
48. T. Tokushima, Y. Harada, O. Takahashi, Y. Senba, H. Ohashi,
L. G. M. Pettersson, A. Nilsson and S. Shin, Chem. Phys. Lett., 2008,
460, 387.
49. B. Hammer and J. K. Nrskov, Surf. Sci., 1995, 343, 211220.
50. T. Schiros, O. Takahashi, K. J. Andersson, H. Ostrom, L. G. M. Pettersson,
A. Nilsson and H. Ogasawara, J. Chem. Phys., 2010, 132, 094701.

Você também pode gostar