You are on page 1of 346

Electronic Excitations

at Metal Surfaces
PHYSICS OF SOLIDS AND LIQUIDS
Editorial Board: JozefT. Devreese • University ofAntwerp, Belgium
Roger P. Evrard • University of Liege, Belgium
Stig Lundqvist • Chalmers University of Technology, Sweden
Gerald D. Mahan. University of Tennessee, USA
Norman H. March. University of Oxford. England

Current Volumes in the Series


CHEMICAL BONDS OUTSIDE METAL SURFACES
Norman H. March
CRYSTALLINE SEMICONDUCTING MATERIALS AND DEVICES
Edited by Paul N. Butcher. Norman H. March, and Mario P. Tosi
ELECTRON CORRELATION IN MOLECULES AND CONDENSED PHASES
N. H. March
ELECTRONIC EXCITATIONS AT METAL SURFACES
Ansgar Liebsch
EXCIT ATION ENERGY TRANSFER PROCESSES IN CONDENSED
MATTER: Theory and Applications
Jai Singh
FRACTALS
Jens Feder
INTERACTION OF ATOMS AND MOLECULES WITH SOLID SURFACES
Edited by V. Bortolani, N. H. March, and M. P. Tosi
LOCAL DENSITY THEORY OF POLARIZABILITY
Gerald D. Mahan and K. R. Subbaswamy
MANY-PARTICLE PHYSICS, Second Edition
Gerald D. Mahan
ORDER AND CHAOS IN NONLINEAR PHYSICAL SYSTEMS
Edited by Stig Lundqvist, Norman H. March, and Mario P. Tosi
PHYSICS OF LOW-DIMENSIONAL SEMICONDUCTOR STRUCTURES
Edited by Paul Butcher, Norman H. March, and Mario P. Tosi
QUANTUM TRANSPORT IN SEMICONDUCTORS
Edited by David K. Ferry and Carlo Jacoboni

A Continuation Order Plan is available for this series. A continuation order will bring
delivery of each new volume immediately upon publication. Volumes are billed only
upon actual shipment. For further information please contact the publisher.
Electronic Excitations
at Metal Surfaces

Ansgar Liebsch
Forschungszentrum JUlich
Jiilich, Germany

Springer Science+Business Media, LLC


Library of Congress Cataloging-in-Publication Data

Liebsch, Ansgar.
Electronic excitations at metal surfaces I Ansgar Liebsch.
p. em. -- (Physics of sol ids and 1 iquids)
Includes blb110graphical references and index.
ISBN 978-1-4419-3271-6 ISBN 978-1-4757-5107-9 (eBook)
DOI 10.1007/978-1-4757-5107-9
I. Plasmons (PhYSiCS) 2. Surfaces (Physics) 3. Metals--Surface.
4. Electronic excitation. 5. Nonlinear opt1CS. I. Title.
II. Series.
QCI76.8.P55L54 1997
530.4'16--dc21 97-16144
CIP

ISBN 978-1-4419-3271-6
© 1997 Springer Science+ Business Media N ew York
Originally published by Plenum Press, N ew York in 1997

http://www.plenum.com
All rights reserved
No part of this book may be reproduced, stored in a retrieval system, or transmitted in
any form or by any means, electronic, mechanical, photocopying, microfilming,
recording, or otherwise, without written permission from the Publisher
Preface

Gaminante, no hay camino


sino estelas en la mar.
A. Machado

Electronic excitations at metal surfaces have attracted interest in surface phy-


sics for many years. These excitations playa key role in nearly all spectro-
scopies routinely used today to characterize surfaces. Spectra generated via
incident electrons, photons, atoms, or ions provide basic information on the
electronic properties of the surface region. In addition, in view of the rapidly
growing importance of nanoscale devices, the microscopic understanding of dy-
namical processes involved in the probe-surface interaction has considerable
practical relevance.
The present volume discusses the recent progress achieved by using the
time-dependent local density approximation (TDLDA) in the description of
electronic surface excitations. The focus is on the dynamical response of metal
electrons to several types of incident electromagnetic fields. The virtue of the
TDLDA is twofold: First, it provides a rather accurate method for describing
ground-state properties of electronic systems, including those of metal surfaces.
Second, the dynamical response is treated in a manner consistent with the one
employed for the ground state. Both of these aspects are essential for an
adequate description of surface excitation spectra.
After a general introduction of the time-dependent density functional ap-
proach, the following main topics are discussed: surface plasmons, nonlocal
optics, nonlinear optics, the van der Waals attraction between atoms and
metal surfaces, and low-frequency electronic surface excitations. The latter
processes contribute significantly to a variety of phenomena, such as friction
of ions or atoms near metal surfaces, damping of adsorbate vibrations, surface
resistivity, etc.
Although this book is written from a theorist's perspective, the empha-
sis of the general physical mechanisms underlying spectroscopic observations

v
vi PREFACE

should make it also useful to a broader readership. In addition, the comparison


of many theoretical results with experimental spectra illustrates the present
status of the field.
To dispell somewhat the apparently widespread notion that surface excita-
tions are difficult to calculate properly, several sections provide the necessary
details concerning the theoretical formulations and computational procedures.
These discussions should be of interest to advanced students and researchers
and make the development of computer codes feasible. These more technical
parts form independent units that can be omitted by the general reader.
I wish to take this opportunity to thank Ward Plummer for having en-
couraged so many able graduate students to investigate electronic properties
of simple metal surfaces by using a variety of techniques: B.O. Kim, G. Lee,
K.J. Song, P.T. Sprunger, K.D. Tsuei, J. Wang, G.M. Watson, z.e. Ying, to
mention but a few. Even a cursory glance at the figures in the text shows that
a sizeable part of the recent experimental progress is based on their data. In
retrospect it is evident that simple metals are neither simple nor devoid of ap-
pealing physics. Ward's open ear to surface theorist's abilities and difficulties
played a key role. The contact with him and his group over many years has
been personally and scientifically most rewarding.
Among the theorists with whom I have interacted and collaborated I wish
to thank, in particular, Bill Schaich. Discussions with him over many years
have been very fruitful. I am also indebted to Hiroshi Ishida and Andrei
Petukhov, who made great strides extending the dynamical response theory
to realistic metal surfaces.
The field discussed in this book involves, of course, contributions of many
theorists. The selection of topics reflects my own range of experience and
perspective. Moreover, it was strongly motivated by the desire to discuss
electronic excitations at metal surfaces from a common physical point of view.
Since the TDLDA provides the most accurate results, it was chosen as the
unifying theoretical approach.
I also thank Bill Schaich for a thorough reading of the manuscript and
Pablo Garda-Gonzalez for several useful comments. I am grateful to the
Forschungszentrum Jiilich for providing me with the opportunity to write this
book. Finally, I thank my wife, Teresita, for encouragement and patience.

Ansgar Liebsch
Contents

1 Introduction 1

2 Density Functional Theory of Metal Surfaces 5


2.1 Ground-State Properties . . . . . . . 6
2.1.1 Local Density Approximation 6
2.1.2 Jellium Model . . . . . . . . 9
2.1.3 Stabilized Jellium Model .. . 15
2.1.4 Static Response Properties .. 18
2.2 Time-Dependent Response Properties . 25
2.2.1 Adiabatic Local Density Approximation 26
2.2.2 Dynamical Local Density Approximation . 34
2.3 Computational Procedures . . 37
2.3.1 Eigenfunctions . . . . 37
2.3.2 Ground-State Density 39
2.3.3 Green's Functions .. 40
2.3.4 Response Functions .. 42
2.3.5 Induced Density. . . . 44
2.3.6 Embedding Approach . 46

3 Surface Plasmons 49
3.1 Classical Picture and Hydrodynamic Models 50
3.2 Inelastic Electron Scattering . . . . 55
3.3 Simple Metal Surfaces . . . . . . . . . . . . 58
3.3.1 Surface Excitation Spectra . . . . . . 58
3.3.2 Dispersion of Monopole and Multipole Surface Plasmons 63
3.3.3 Beyond Standard Jellium and Adiabatic TDLDA 66
3.3.4 Lifetime of Surface Plasmons . . . . . . 70
3.3.5 Long-Wavelength Limit. . . . . . . . . . . . . 73
3.3.6 Retardation: Surface Plasmon Polaritons . . . 83
3.3.7 Comparison with Model Potential Predictions 86

VB
viii CONTENTS

3.4 Interband Transitions. . . . . . 90


3.5 Influence of Occupied d Bands. 92
3.5.1 Ag 93
3.5.2 Hg . . . . .104
3.5.3 Pd . . . . .105
3.6 Charged Surfaces . 106
3.6.1 Simple Metals. . 106
3.6.2 Ag....... . 107
3.7 Influence of Dielectric Medium. . 109
3.7.1 Mg - MgO . .111
3.7.2 Ag - Ar . . . . . . . . . . 111
3.8 Overlayers............ . 113
3.8.1 Local Optics Picture and Hydrodynamic Models. . 114
3.8.2 Jellium Model . . . . . . . 116
3.8.3 Alkali Metal Overlayers. . . . . 117
3.8.4 Ag Overlayers . . . . . . . . . . 126
3.8.5 Chemical Interface Damping . . 128
3.9 Small Metal Particles. . . 130
3.9.1 Classical Picture . . . 130
3.9.2 Simple Metals . . . . . 130
3.9.3 Ag and Hg Particles . 138
3.9.4 Chemical Interface Damping. . 140
3.10 Quantum Wells . . . . . . . . . . . . . 141

4 Nonlocal Optics 145


4.1 Classical Picture and Phenomenological Extensions . 146
4.2 Microscopic Theory . . . . . . . . 148
4.2.1 Long-Wavelength Limit. . 150
4.2.2 d.L(w} and dll(w) . . . . . 155
4.2.3 Spectroscopic Quantities . 160
4.2.4 Solution of Response Equation. . 165
4.2.5 Sum Rules . . . . . . . . . . . 171
4.3 Simple Metals . . . . . . . . . . . . . 174
4.3.1 Surface Excitation Spectra. . 174
4.3.2 Emission vs. Absorption . . . 179
4.3.3 Comparison with Measured Spectra. . 180
4.4 Interband Transitions. . . . . . . . . 185
4.5 Influence of Occupied d Bands. . . . 186
4.5.1 Comparison of Approaches. . 186
4.5.2 A g . . . . . . . . . . . . . . . 188
CONTENTS ix

4.6 Charged Surfaces . . . . . . .198


4.7 Overlayers.......... . 199
4.7.1 Local Optics Picture .199
4.7.2 Realistic Alkali Metal Overlayers .200
4.7.3 Ag Overlayers . . . . . . . . . . . .210

5 Nonlinear Optics 213


5.1 Phenomenological Considerations .214
5.1.1 Bulk Polarization .. .214
5.1.2 Surface Polarization . . . .216
5.1.3 Harmonic Radiation .. . . 219
5.1.4 Validity of Long-Wavelength Limit . 227
5.2 Microscopic Theory . . . . . . . . . . . . . . 227
5.2.1 Nonlinear Response. . . . . . . . . . 227
5.2.2 Relation between Linear and Nonlinear Responses. . 229
5.2.3 Surface Polarization . . . . . . . 229
5.2.4 Solution of Response Equation. . 230
5.2.5 Dynamical Force Sum Rule . 232
5.3 Simple Metals . . . . . . . . . . . . . 234
5.3.1 Adiabatic Response. . . . . . 234
5.3.2 Surface Excitation Spectra . . 237
5.3.3 Comparison with Measured Spectra: Al . 240
5.3.4 Alkali Metals . . . 242
5.3.5 Extension to Ag . . 244
5.4 Charged Surfaces . . . . . 246
5.5 Alkali Metal Overlayers . . 250
5.5.1 Jellium Model. . . 252
5.5.2 Realistic Alkali Metal Overlayers . 253
5.6 Surface Anisotropy . 255
5.6.1 AI(111) . . . . . . . . . . . 256
5.6.2 Ag(111) . . . . . . . . . . . 260
5.6.3 Other Crystalline Effects . . 262
5.7 Stepped Metal Surfaces. . . . . . . 263
5.7.1 Anisotropy due to Density Contour . 263
5.7.2 Anisotropy due to Crystal Structure . 265
5.8 Magnetic Surfaces. . . . . . . . . . . . . . . . 268

6 Van der Waals Attraction 271


6.1 Reference Plane Position . 273
6.2 Density Functional Description . 276
x CONTENTS

6.3 Near-Surface Corrections . . . . .278


6.4 Influence of Occupied d Bands . .280

7 Electron-Hole Pair Creation 283


7.1 Inelastic Electron Scattering . . . . . . . . . . .283
7.2 Low-Frequency, Long-Wavelength Excitations .286
7.3 Low-Frequency, Short-Wavelength Excitations .290
7.4 Damping of a Vibrating Dipole . . . . .294
7.5 Sliding Friction of Ions . . . . . . . . . .298
7.6 Sliding Friction of Physisorbed Atoms. .300
7.7 Surface Resistivity . . . . . .303
7.7.1 Classical Picture .. .304
7.7.2 Microscopic Theory . .306

Bibliography 309

Index 331
Electronic Excitations
at Metal Surfaces
Chapter 1

Introduction

Electronic excitations at the boundary of metallic systems are of interest from


a fundamental and practical point of view. These excitations comprise a class
of states of the interacting electron gas that differ greatly from those in purely
two- or three-dimensional systems. The theoretical understanding of these
modes and the identification of those surface parameters that govern their
spectral characteristics have been important topics in surface physics for a
long time. Surface spectroscopies employing electrons, photons, atoms, or ions
invariably involve some type of single-particle or collective electronic excita-
tion. How these external probes couple to the surface modes, and what can
be learned from this interaction about the state of the surface, is therefore
relevant for many investigations in surface physics.
In the present volume we consider electronic excitations induced by a vari-
ety of external electromagnetic fields. The frequencies range from the adiabatic
limit up to or above the bulk plasma frequency. This includes, for example,
low-energy friction processes of atoms and ions near a metal surface as well
as the creation of collective surface modes via incident electrons. The wave-
lengths range from very long, as in the case of photons, to very short, as in
the case of adsorbed species. In most situations, these fields are sufficiently
weak, so that the response is linear. With the aid of incident laser radiation,
however, nonlinear response properties become readily observable. These may
also be used to characterize surfaces.
The principal goal of this book is to discuss these different kinds of elec-
tronic surface excitations within a common theoretical concept. The approach
we take here relies on two essential ingredients: the electronic properties of the
ground state and the dynamical response to external fields. It is now generally
accepted that surface response properties are very sensitive to details of the
electronic density. An accurate treatment of the surface electronic structure is

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
2 CHAPTER 1. INTRODUCTION

therefore a prerequisite for an adequate description of excitations. Response


calculations based on model surface potentials or densities often lead to spectra
that differ qualitatively from those obtained using self-consistent schemes.
The theoretical work discussed in this volume is based on the density func-
tional approach in the local approximation of the exchange-correlation energy
(Hohenberg, Kohn, 1964; Kohn, Sham, 1965). This method has proven to be
very reliable for evaluating ground-state electronic properties of many systems,
including metal surfaces.
Electronic surface excitations are treated using an extension of the den-
sity functional approach to time-dependent external perturbations (Zangwill,
Soven, 1980; Stott, Zaremba, 1980; Mahan, 1980). A key feature of this exten-
sion is that it provides a consistent treatment of electron-electron interactions
in the absence and presence of time-varying fields. The time-dependent den-
sity functional approach therefore represents a more refined response treatment
than the random-phase approximation (Ehrenreich, Cohen, 1959). It includes
single-particle as well as collective excitations; it accounts for the all-important
surface screening effects; and it is based on a fully quantum mechanical for-
mulation of the nonlocal electronic response. Moreover, it yields detailed in-
formation on the spatial distribution of the induced charge density and local
electric field in the surface region.
Starting with the prediction of the surface plasmon by Ritchie in 1957,
research in the area of electronic surface excitations has been carried out for
nearly four decades. During the past few years, however, a variety of experi-
ments on different systems was performed that now provide a much wider and
more systematic picture of spectroscopic observations. The systems include
simple metals of low and high electronic densities, charged surfaces, stepped
surfaces, noble metals, thin adsorbate layers over a wide range of coverages,
and small metal particles. The spectroscopies include electrons, atoms and
photons, the latter in the linear and nonlinear response regime.
Considerable recent progress has also been achieved in the theoretical de-
scription of surface excitations and various spectroscopic quantities. We now
have a clearer understanding of the physical processes involved and a better
appreciation of those electronic properties that are important for an adequate
analysis of the data. In addition, it has become computationally much easier,
at least for simple metals, to carry out surface response calculations at the
relevant frequencies and wave vectors.
The rather large amount of evidence now available shows that surface exci-
tations are reasonably well described within the time-dependent density func-
tional approach. Thus, although this method is not exact (in practice, both
the ground-state and response treatments involve approximations), it seems to
3

capture the main correlation effects that govern electronic excitations at metal
surfaces. The overall consistency of the observations and the fascinating in-
terconnections among a variety of systems and spectroscopic methods suggest
that this important topic in surface science has now matured to a considerable
degree.
Beyond the specific relevance of the phenomena which have been investi-
gated, the main conclusion from this work is significant not only for surface
physics but for condensed matter physics in general. Just as the applica-
bility of the local density approximation for the description of ground state
electronic properties was demonstrated by a large number of comparisons of
experimental and theoretical results, the validity of its extension to the time-
dependent domain is truly astounding. Although the linear and nonlinear
induced densities at metal surfaces (at frequencies up to the bulk plasma fre-
quency and wavelengths down to a few A) do not appear to vary slowly, neither
in time nor space, the agreement between experiment and theory is often semi-
quantitative, sometimes quantitative.
The main aspects of the density functional approach are reviewed in Chap-
ter 2. We define the quantities needed to evaluate various response properties
and discuss important details of the computational procedures. Further theo-
retical aspects concerning surface excitations created in specific spectroscopies
are given in later chapters.
Chapter 3 focuses on the wave vector dispersion of collective excitations at
a variety of metal surfaces. The relation between these modes and collective
excitations in spherical metal particles and quantum wells is also addressed.
The nonlocal optical response of metallic surfaces is the topic of the two
subsequent chapters. Chapter 4 deals with linear excitations that playa role
in the reflection of light, surface photoemission, attenuated total reflection,
and other optical spectroscopies. Chapter 5 describes the nonlinear case, in
particular, second harmonic generation.
In addition, in Chapter 6, we discuss the van der Waals attraction between
atoms and metal surfaces and, finally, in Chapter 7, the creation of low-energy
electron-hole pairs, which influences the damping of adsorbate vibrations, fric-
tion of ions and atoms near metal surfaces, surface resistivity, etc.
For computational reasons, most dynamical response calculations so far
have been restricted to simple metal surfaces and overlayers. Fortunately,
the one-dimensional jellium model allows a fairly accurate description of these
metals. A few results for fully three-dimensional systems are also presented.
In addition, we describe several extensions of the jellium model that deal with
the influence of core levels, occupied d bands, external dielectric media, and
static electric fields that lead to surface charging.
4 CHAPTER 1. INTRODUCTION

In view of the vast range of phenomena related to electronic surface exci-


tations, a complete review is not intended. Instead, we concentrate on those
spectroscopic methods that reveal surface electronic excitations most directly
and those systems for which comparisons with calculations based on the time-
dependent density functional approach are available.
For a discussion of early work in the area of electronic surface excitations,
the reader is referred to the following reviews: Wikborg and Inglesfield (1977),
Apell (1981), Feibelman (1982), Apell et at. (1984), Forstmann and Gerhardts
(1986), Eguiluz (1987), Liebsch (1987), Schaich and Kempa (1987), Raether
(1988); see also Zangwill.(1988). Some of the more recent studies were reviewed
by Mahan and Subbaswamy (1990), Gerhardts (1992), Rocca (1995), Dobson
(1995), and Liebsch (1995).
Atomic units (m = Ii = 1, e = -1) are used throughout unless stated
otherwise. Thus, energies are given in units of 1 Hartree = 27.2 eV, and
distances are measured in units of 1 Bohr radius ao = 0.529 A. Electronic charge
densities are mostly expressed in terms of number densities: p == en = -no
Similarly, electrostatic potentials are written as ~ == </>Ie = -</>. Thus, while
</> actually represents the electrostatic potential energy, we refer to it simply
as potential.
Chapter 2

Density Functional Theory of


Metal Surfaces

One of the main themes of this volume is the sensitivity of electronic excita-
tions at metal surfaces to the ground-state electronic properties in the surface
region and to the nonlocal response of surface electrons to incident electromag-
netic fields. The density functional approach provides an accurate description
of both properties. Since this scheme forms the theoretical basis for nearly all
linear and nonlinear electronic excitations treated in the following chapters, we
describe here the salient aspects of this approach. Section 1 deals with ground-
state properties. We introduce the local density approximation and the jellium
model for which most calculations have so far been performed. We also discuss
the so-called stabilized jellium model that takes the electron-ion interaction
into account in a better way than the standard jellium model. The static re-
sponse properties of jellium surfaces are presented for both models. The time-
dependent extension of the density functional method is the topic of Section 2.
We focus on the adiabatic treatment of the induced exchange-correlation po-
tential, but we also mention recent dynamical generalizations. An important
feature is the description of some of the computational procedures for evalu-
ating surface response properties; these are presented in Section 3. We give
the expressions of relevant quantities and point out the necessary steps that
allow a stable and accurate calculation of induced dipole moments; additional
aspects are pointed out in other chapters.

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
6 CHAPTER 2. DENSITY FUNCTIONAL THEORY

2.1 Ground-State Properties


2.1.1 Local Density Approximation

The central statement of the density functional theory (Hohenberg, Kohn,


1964; Kohn, Sham, 1965) is that the exact ground-state energy of an interact-
ing system of electrons can be expressed as a unique functional of the electronic
density n(f):

E[n]

(2.1)

This ground-state energy is at a minimum with respect to variations of the elec-


tronic density. The first term in this expression represents the kinetic energy of
an equivalent non-interacting electron system with the same ground-state den-
sity as that of the exact many-body electron system. The second term is the
potential energy due to the electron-ion interaction, with Vion(f) representing
the Coulomb potential caused by the positive ionic charges. The third term
denotes the average electron-electron interaction. Finally, the quantity Exc[n]
is the exchange-correlation energy that contains all remaining many-body in-
teractions. In particular, this term also includes part of the true many-body
kinetic energy not contained in T[n].
An important practical advantage of the density functional approach is that
the electronic density can be derived from the solutions of the one-electron
Schrodinger-like equation

(2.2)

where the effective potential Velf is given by

v:elf (rJ;;'\ = TT
Yion
(;;'\
rJ + Id3' r
n(r")~'I
I~
r-r
+ TT
Yxc
(;;'\
rJ (2.3)

with
(2.4)

The electronic density is then

n(f) L h l1Pf(f)12 , (2.5)


f
2.1. GROUND-STATE PROPERTIES 7

where h are Fermi-Dirac occupation factors. The preceding self-consistent


equations are formally exact. They show that the full many-body problem of
the interacting electron gas has been recast in the language of a one-electron
problem. The complexity of the many-body problem is hidden in the func-
tional Exc[n]. In practice, approximations must be introduced to evaluate this
functional and solve for the electronic density of a given system. The most
common procedure is to adopt the local density approximation which we dis-
cuss below.
Since the density functional method provides a scheme for evaluating quan-
tities that depend on the electronic density, the functions 'l/J,Jr') and energies
c,. have no direct physical meaning. These quantities may be viewed as auxil-
iary parameters used to construct the exact density. In particular, the energy
parameters c,. are not identical to the excitation energies observed, for ex-
ample, in photoemission spectra. To analyze such spectra, the quasi-particle
nature of the generated hole states must be taken into account. This includes
the relaxation shift and line broadening due to interaction with the remain-
ing electrons, and, possibly, new spectral features, such as satellites and losses
due to excitation of other elementary modes. Hence, although there is often a
close correspondence between the calculated energy parameters c,. and ener-
gies measured in photoemission or other spectroscopies, this analogy must be
used with caution.
If the density varies slowly, the exchange-correlation energy Exc[n] can be
expanded as

where cxc(n) is the average exchange-correlation energy per electron of a homo-


geneous electron gas. The second term represents the lowest-order correction
due to the spatial variation of the density. The local density approximation
(LDA) is obtained by neglecting these gradient corrections. The exchange-
correlation contribution to the effective potential (2.3) is then given by

TT [
Yxc on I
r, - oncxc(n)
n (;;'1]- . (2.7)
n=n(r)

Thus, in the LDA, the true value of Vxc[n(r')] at the point r of the actual
inhomogeneous density n(t') is replaced by the corresponding value for aho-
mogeneous electron gas of density fi = n(r').
The exchange-correlation energy of the homogeneous electron gas is nu-
merically known for a wide range of metallic densities (Ceperley, Alder, 1980).
8 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Several approximate analytical expressions for this energy as a function of elec-


tron density also exist. The commonly used Wigner formula for the exchange-
correlation energy per electron is (Pines, 1964):

( ) _ 0.458 0.44
(2.8)
Cxc n - - r,(n) - r.(n) + 7.8 '

where the average electron radius r. is defined by the relation

(2.9)

Here, kF is the Fermi wave vector. The first term in (2.8) represents the exact
exchange energy as obtained from a Hartree-Fock calculation. The second
term accounts for the correlation energy. The exchange-correlation potential
(2.7) is then given by

0.611 0.587
Vxc(n) = --(-) - (
r, n
7 )2 (r.
r. + .8
+ 5.85) . (2.10)

Most of the calculations discussed in this volume are based on the Wigner
formula. The quantum mechanical Monte Carlo results of Cepedey and Alder
(1980) for r. > 1 can be represented via the interpolation formula

0.458 'Y
cxc(n) = - r.(n) - , (2.11)
1 + {3IVT. + f32r.
0.611 'Y(1 + ~{3IVT. + ~f32r.)
Vxc(n) = - r.(n) - (2.12)
(1 + {3IVT. + f32r.)2

with'Y = 0.07115, {3I = 1.0529, and {32 = 0.3334 (Perdew, Zunger, 1981). Ex-
tensive discussions of the density functional approach can be found in Lundqvist
and March (1983), Parr and Yang (1989), Jones and Gunnarsson (1989), and
Dreizler and Gross (1990).
The LDA has been used in a large number of practical applications of the
density functional method. It has been found to yield surprisingly accurate
results, even in atoms and solids where the density does not vary slowly. At
metallic surfaces, the LDA is known to give good results for important quanti-
ties, such as the work function and surface energy. This is not obvious because
the equilibrium density decays quite rapidly at the surface. Moreover, the LDA
does not reproduce the exact form of the effective potential far from the sur-
face. Since the density decays exponentially towards the vacuum, it is evident
from (2.7) that the potential also decays exponentially. In contrast, the exact
2.1. GROUND-STATE PROPERTIES 9

nonlocal density functional theory gives an exchange-correlation potential that


approaches the classical image potential far from the surface:
1
Vxc(z) -+ - 4( Z ) (2.13)
- Zl

Here, Zl is the position of the static image plane which we define later. The
positive z-axis specifies the surface normal.
If an electron is moved out of the metal, its exchange-correlation hole re-
mains inside and gradually turns into the image charge of the electron. Since
in the LDA the exchange-correlation hole remains attached to the electron, the
formation of the image charge cannot occur. On the other hand, the image
form of the effective potential is usually reached at distances where the density
is already quite small. Hence, the failure of the LDA to reproduce the image
potential has only a weak influence on the density profile. Accordingly, those
surface quantities that depend only on the ground-state electron density, such
as the work function and the surface energy, are only weakly affected by the
local approximation (Zhang et al., 1990).
There exist now several highly elaborate methods for evaluating ground-
state electronic properties of surfaces and interfaces. The LDA appears to be
sufficiently accurate, so that even subtle phenomena like weak adsorption of
atoms (e.g., rare gas atoms on a metal surface; see Lang, 1981) and surface
magnetism (see, for example, Jepsen et al., 1980) are described realistically.
Because of the complexity of these methods, however, it is difficult to extend
them to finite frequencies. A few such calculations for three-dimensional metal
surfaces are discussed in the following chapters. In the next section, we intro-
duce a model that eliminates the complexity caused by the crystal structure,
but nevertheless retains essential features of the electronic properties of metal
surfaces.

2.1.2 Jellium Model

Most of the systems discussed in this volume involve simple metals. Since
their ionic pseudopotentials are rather weak, the electronic surface properties
can be described quite well within the so-called jellium model. The lattice
of positive ionic charges in this model is replaced by a semi-infinite uniform
background:
(2.14)
where 8(z 2:: 0) = 1 and 8(z < 0) = O. The only free parameter is the volume
density ft, or equivalently, the average electron radius r. defined in (2.9). The
10 CHAPTER 2. DENSITY FUNCTIONAL THEORY

values of rs for various simple metals are: AI: r. = 2.07, Na: r. = 3.99, K:
r. = 4.96, Rb: rs = 5.23, and Cs: rs = 5.63. The ground-state electronic
properties of semi-infinite jellium systems were investigated within the LDA
by Lang and Kohn (1970, 1971, 1973; Lang, 1973).
For the purpose of studying electronic excitations at metal surfaces, the
jellium model within the LDA is highly attractive for several reasons:

• The jellium model provides an accurate, self-consistent description of


the density distribution in the surface region. Since the electronic profile
normal to the surface leads to strong gradients of the normal component
of the electric field, observable quantities related to these field gradients
can also be expected to be represented accurately.

• The unoccupied part of the continuum, which plays an equally important


role for electronic transitions as the occupied states, is also described
qualitatively correctly. The main error is due to the non-image-like shape
of the barrier potential. However, since the image form sets in quite far
from the surface, the contribution to matrix elements from this outer
region is rather small.

• As long as we consider frequencies below the volume plasma frequency,


there are no electronic excitations in the bulk. Thus, the jellium model
allows us to focus on electronic surface excitations without being con-
cerned with complications arising from the interference of bulk and sur-
face excitation processes.

• Since the jellium model is translationally invariant parallel to the sur-


face, the computational effort is much simpler than for three-dimensional
systems.

In realistic systems, of course, interband transitions induced by the lat-


tice potential contribute their own characteristic spectral features and lead to
modifications of intrinsic surface excitations associated with the density pro-
file. Examples of such effects are given in later chapters. The computational
treatment of these effects remains one of the important challenges of future
work in surface physics.
The dynamical surface response properties predicted by the LDA-jellium
model are in good agreement with a large number of experimental observations
on a variety of simple metals. Despite its simplicity, this model is therefore
vastly superior to more approximate models frequently employed in theoretical
studies of electronic surface excitations. Although the ground-state densities
2.1. GROUND-STATE PROPERTIES 11

of finite- or infinite-potential barrier models, for example, may appear decep-


tively reasonable, these models nevertheless often give qualitatively incorrect
results for the finite-frequency response. For low-lying excitations, transitions
between states near the Fermi level are the most important. Thus, the tail
region of the density far in the vacuum must be described well. It is precisely
in this region, however, that densities corresponding to model potentials differ
most severely from the LDA density. In addition, as previously mentioned,
unoccupied continuum states also playa decisive role in electronic excitation
spectra. These states tend to be even more poorly represented in model po-
tentials than the occupied ones.
We may view these approximate, non-self-consistent model potentials as
Hartree potentials corresponding to a semi-infinite ionic background in the
presence of an external electric field. This external field must exhibit very
peculiar and unrealistic features for the total potential to look like the model
potential. It is not surprising, therefore, that the dynamical response in the
presence of such an external field also exhibits unrealistic peculiarities. Since
the numerical effort of calculating the dynamical response of jellium surfaces
within the LDA is not as great as seems to be commonly believed, there is no
need to resort to overly simplified models.
Figure 2.1 shows the electronic density profiles calculated within the LDA
for two jellium systems (Lang, Kohn, 1970). Because of the sharp cutoff of
electronic wave vectors at the Fermi surface, the densities in the interior exhibit
slowly decaying Friedel oscillations with wavelength 7f/kF = 1.64r•. Asymp-
totically, the density has the form

n(z) = n [1 + a COS(2::Z + Q) + ...J ' (2.15)

where the constants a and Q depend on the shape of the surface potential. The
spill-out of the density into the vacuum is caused by the lowering of kinetic
energy. Since states close to the Fermi level decay more slowly than deeper-
lying ones, the outer region of the density plays a particularly important role
for low-frequency surface excitations. The overall decrease of the density from
its bulk value occurs on the scale of an electronic screening length, which is of
the order of a few A at typical metallic densities. An effective location of the
surface can be defined in terms of the quantity

ZII = f dzz d~ n(z) / f dz! n(z)


f dz [n~) - O(-Z)J . (2.16)
12 CHAPTER 2. DENSITY FUNCTIONAL THEORY

1.0

.?:-
.;;;

C
..
<:

<:
.::<J
0

..
L;j
0.5

0.0
-1.0 -0.5 0.0 0.5 1.0
z (Fermi Wavelengths)

Figure 2.1: Ground-state density distributions n(z)/n near jellium


surfaces for T. = 2 and 5. The Fermi wavelengths are AF = 27r /kF =
3.46 A and 8.33 A, respectively. The positive background occupies
the half-space z ::; O. (Lang, Kohn, 1970).

For neutral surfaces, zll = 0, Le., the effective position of the surface coincides
with the jellium edge. In subsequent chapters we will see that zll plays an
important role in the dynamical response of surfaces.
The effective one-electron potential

(2.17)

is shown in Figure 2.2 for T. = 4. In this case, the main contribution to the po-
tential is due to the exchange-correlation potential Vxc(z). The remaining part
arises from the electrostatic potential r/J which satisfies the Poisson equation
d2
dz 2 ¢(z) = -47r [n(z) - n+(z)] (2.18)

I:
This potential may be expressed as

¢(z) = -27r dz' Iz - z'l [n(z') - n+(z')] . (2.19)

As pointed out above, the effective potential decreases exponentially toward


the vacuum instead of approaching the image form as in a nonlocal treatment
2.1. GROUND-STATE PROPERTIES 13

.10 <P
-------------....,
'"':' .05
"
~

..
>.
Er
~
0 f------------/------------
c:
w

-.05
V.1f r.=5

-1.0 -0.5 0.0 0.5 1.0


z (Fermi Wavelengths)

Figure 2.2: Effective one-electron potential v"ff(Z) and electrostatic


part ¢>(z) for jellium surface (rs = 5). (Lang, Kohn, 1970).

of exchange-correlation energy. The transition of Veff(z) toward this image


potential occurs at about z ~ 0.5>'F' where the density is very small. Moreover,
as shown later, the total height of the surface barrier potential is, apart from
the surface dipole, a bulk property. It is therefore only weakly affected by the
local density treatment in the surface region (Lang, 1983).
Because of the spill-out of electrons into the vacuum region, a dipole layer is
formed at the surface. The difference between electrostatic potentials outside
and inside the metal is given by
+00
D = ¢>(oo) - ¢>( -(0) = 41f / -00 dz z [n(z) - n+(z)] (2.20)

This dipole barrier forms an important part of the work function <P, which is
defined as the minimum energy required to remove an electron from the Fermi
level to a point far from the surface:

(2.21)

Here, EN and E N - 1 are the total energies of the Nand N -1 electron systems;
tl. V is the total height of the surface barrier potential, and the Fermi energy
EF = 0.5 ki is measured with respect to the one-electron potential deep inside
14 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Table 2.1: LDA ground-state energies of jellium surfaces. Fermi


energy EF = 0.5 k~, work function <P, bulk exchange-correlation
potential Vxc (-00), dipole barrier D, and height of surface barrier
potential ~V. All energies are given in eV. (Lang, Kohn, 1970).

r. 2 3 4 5 5.6
EF 12.5 5.6 3.1 2.0 1.6
<P 3.9 3.5 3.1 2.7 2.5
Vxc(-oo) -9.6 -6.8 -5.3 -4.4 -4.0
D 6.6 2.3 0.9 0.3 0.1
~V 16.2 9.0 6.2 4.7 4.1

the metal. To show the bulk and surface contributions to the work function
explicitly, we make use of the relations

¢(oo) , (2.22)
¢(-oo) + Vxc(-oo) . (2.23)

(The vacuum level is taken as zero of the energy scale.) Hence,

(2.24)

and
<P = D + IVxc( -00)1- EF . (2.25)
The dipole barrier D represents the surface contribution to the work function,
while the bulk part is given by the exchange-correlation potential Vxc ( -00)
« 0) deep inside the metal. Table 2.1 lists the values of various important
ground-state energies for several jellium surfaces.
Since electronic states near the Fermi level leak farthest into the vacuum,
the exponential decay of the density is directly determined by the work func-
tion:
n(z) ex: exp( -2zm) . (2.26)
The screening response to external electric fields depends crucially on the ex-
citation of electron-hole pairs in the surface region. It is clear, therefore,
that the work function plays an important role in the polarization properties
of the surface density. The LDA calculations for realistic surfaces (including
the full three-dimensional crystal potential) give remarkably accurate values
2.1. GROUND-STATE PROPERTIES 15

of the work function. In the case of the jellium model, neglecting the lattice
potential leads to an underestimate of ip for high-density metals and a slight
overestimate for low-density metals.
Neglecting the lattice potential in the jellium model has an even more deci-
sive effect on the surface energy (J' (the work required per unit area to split the
infinite crystal into two halves along a plane). For r. less than about 2.5, (J' is
negative, so that the metal is mechanically unstable. Using perturbation the-
ory, Lang and Kohn (1970) showed that incorporating ionic pseudopotentials
gives significantly better agreement with measured work functions and surface
energies for a variety of simple metals. The crystal face dependence of both
quantities can also be approximately described in terms of pseudopotentials
(Monnier et al., 1978).

2.1.3 Stabilized Jellium Model

A simple method of including the main effect of the lattice potential in


the jellium model, without destroying its one-dimensional form, was proposed
by Perdew et al. (1990) and Shore and Rose (1991). In these schemes, the
standard jellium model is replaced by the stabilized jellium model in which a
constant (the structureless pseudopotentiaQ is added to the one-electron poten-
tial in the interior of the bulk. In the approach by Perdew et al., this constant
represents the average of the pseudopotential over the bulk Wigner-Seitz cell.
It is chosen to make the metal mechanically stable at its observed valence
electron density.
The treatment by Shore and Rose is slightly different in the bulk but yields
the same surface properties. Since it is somewhat simpler we discuss it here
briefly. Let us consider a homogeneous gas of N electrons with density n.
These electrons are exposed to an average pseudopotential Vo to ensure me-
chanical stability. To determine this constant, let us introduce infinitesimal
gaps between arbitrary pieces of the positive ionic background. If the total
number of electrons is unchanged, these gaps lead to a reduced average den-
sity n' and a change in energy given by El = N[c(n') - c(n)]. Here, c(n) is
the average energy per electron of the homogeneous electron gas:

1.105
c(n) = -2- + cxc(n) . (2.27)
r.

But there is another energy change E2 due to the fact that electrons in the
gaps no longer feel the potential Vo. Thus, E2 = NVo(n' - n)jn. Setting
16 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Table 2.2: LDA ground-state energies of stabilized jellium surfaces.


Height of stabilized jellium dipole potential Vo, work function q>,
dipole barrier D, and height of surface barrier potential AV. All
energies are given in eV.

r. 2 3 4 5 5.6
Yo -2.85 -0.74 -0.1 0.15 0.23
q> 4.3 3.7 3.1 2.6 2.4
D 7.2 2.5 0.9 0.3 0.04
AV i6.8 9.2 6.2 4.7 4.0

El + E2 = 0 and expanding c( n') about the equilibrium density n, we find to


first order in (n' - n):

Vo = -n oc(n) I = r. oc(n) I (2.28)


On Ii 3 or. Ii

At a metallic surface, this term changes the potential in the region of the
positive ionic background relative to the potential in the vacuum. Table 2.2
lists the values of various ground-state energies using the stabilized jellium
model. The parameter Vo is seen to be negative for r. :5 4, thus increasing
the total height of the surface barrier. For r, > 4, Yo is slightly positive.
The calculated work functions and surface energies agree much better with
the experimental values (see Figures 2.3 and 2.4). The attractive feature of
the stabilized jellium model is that it incorporates the electron-ion interaction
in a more realistic manner than the standard jellium model without sacrificing
its one-dimensional simplicity.
By refining the stabilized jellium model, it is also possible to account for
the lateral corrugation of the electronic density along the surface (Perdew et
al., 1990). Since the corrugation affects the strength of the surface dipole, this
effect can be included by adjusting the value of the average pseudopotential
Yo in (2.28). In this way the observed crystal face dependence of the work
function and surface energy can be qualitatively reproduced.
In the following chapters, we discuss several examples illustrating the sen-
sitivity of surface excitation spectra to the ground-state density profile. The
larger work function of high-density metals, for example, that is more correctly
described within the stabilized jellium model, leads to a blue shift of excitation
frequencies and to a reduction of line broadening as a result of the stiffer, less
2.1. GROUND-STATE PROPERTIES 17

5 +

:;-
.......
Q)
4
c
0

-
~ 3
c
::J

~
.... 2
0
~

OL-----~ ______- L______ ~ ______L__ _ _ _~

1 2 3 4 5 6

Figure 2.3: Work function <P calculated within LDA for stabilized
jellium model (solid line). The symbols denote measured work func-
tions for several'simple metals. (Shore, Rose, 1991).

..,...,E 1500 +

u
'-.
"'C>
.... +
~
>-
1000 +
C>
....
Q)
c ++
Q)

Q) + +
u
0 500 t
't +
::J
Vl
+
+ + +
0
1 2 3 4 5 6
r. (co)
Figure 2.4: Surface energy calculated within LDA for stabilized
jellium model (solid line). The symbols denote measured surface
energies for several simple metals. (Shore, Rose, 1991).
18 CHAPTER 2. DENSITY FUNCTIONAL THEORY

polarizable density distribution. Conversely, the reduced work function of low-


density metals causes an increased broadening due to creation of electron-hole
pairs and a red shift of excitation frequencies.

2.1.4 Static Response Properties

The response of surface electrons to a static external electric field is an


important extension of the LDA ground-state calculations discussed above.
The static linear or nonlinear response can be regarded as characteristic of
the dynamical response as long as the excitation frequency is sufficiently far
below the plasma frequency and other collective or single-particle excitations.
Since a static potential can be added to the effective one-electron potential,
the change of the electronic density distribution can be calculated directly
without having to use response functions. Thus, existing electronic structure
codes for slabs or semi-infinite metals can be readily adapted to determine
the static response. Of interest is the adiabatic response to both long- and
short-wavelength potentials.

Long-Wavelength Response
Let us consider first the linear and second-order static response of a semi-
infinite jellium metal to a uniform electric field oriented perpendicular to the
surface. For sufficiently weak fields, the density calculated within the LDA
may be expanded as
(2.29)
Here, no(z) is the ground-state density in the absence of the field and a is the
external charge per unit area. The first- and second-order induced densities
nl(z) and n2(z) may be calculated from the expressions:

(2.30)

(2.31)

where n±(z) are the densities (2.29) for slightly positively and negatively
charged surfaces. The integrated weights of the induced densities are

! dz nl(z) 1, (2.32)

! dz n2(z) 0. (2.33)
2.1. GROUND-STATE PROPERTIES 19

-8 -4 o 4 8
z (u.u.)

Figure 2.5: Linear screening density nl(z) (Lang, Kohn, 1973) and
second-order polarization P2(Z) (Weber, Liebsch, 1987) induced at
jellium surface (r. = 4) by weak static electric field. The centroids
of these distributions are indicated by Zl and Z2. The dot-dashed
curve is the ground-state density no(z) in the absence of the applied
field. The positive background occupies the half-space z :$ O.

Thus, in contrast to nl (z), which consists of a main bump in the surface region
and weak Friedel oscillations in the interior, n2(z) has an extra node, so that
positive and negative charge regions cancel.
Linear screening densities were calculated by Lang and Kohn (1973) for
several jellium surfaces (see Figure 2.5). These authors also proved that the
centroid of nl ( Z ),
(2.34)

represents the position of the static image plane as indicated in (2.13) (see
below). Typically, Zl lies about 1 Bohr radius outside the jellium edge.
Figu:t:e 2.5 also shows the second-order induced polarization P2 (z), which
is related to n2(z) via dP2(z)/dz = n2(z). Hence,

P2(z} = 1:00 dz' n2(Z') . (2.35)

Since the integrated second-order density is zero, P2 (z) vanishes at ±oo. The
main part of this second-order polarization is seen to be located in the region
20 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Table 2.3: Static response properties of jellium surfaces. Position


of static image plane Zl, centroid Z2 of second-order polarization,
and nonlinear induced dipole moment P2' Upper rows: LDA results
for standard jellium model (Weber, Liebsch, 1987); lower rows:
LDA results for stabilized jellium model (Kiejna, 1995; Garcfa-
Gonzalez, Liebsch, 1997). All quantities are given in atomic units.

r. 2 3 4 5 5.6
Zl 1.57 1.35 1.25 1.17 1.14
Z2 2.18 1.92 1.77 1.70 1.68
P2 239 365 579 870 1100
Zl 0.97 1.05 1.19 1.34 1.56
Z2 1.58 1.64 1.71 1.80 1.86
P2 158 304 560 956 1300

where the ground-state density decays rapidly. Weak Friedel oscillations ex-
tend towards the interior. Similar results are obtained for other bulk densities.
The second-order polarization P2(Z) is important for the nonlinear optical

i: I:
response of metal surfaces (see Chapter 5). The centroid

Z2 = dz Z P2(Z) / dz P2(z)

= ~ I: dz Z2 n2(Z) /I: dz zn2(Z) (2.36)

defines the average location of the perpendicular surface current in the low-
frequency limit. As seen in Figure 2.5, z2lies slightly farther out in the vacuum
than Zl' This result indicates that nonlinear screening properties depend even
more sensitively on the tail region of the ground-state density than the linear
response. Table 2.3 lists the values of Zl and Z2 for several jellium surfaces.
Also given are the first moments of n2(z) which are equal to the negative
integrated weight of P2(Z):

P2 = I: dz z n2(Z) = -I: dz P2(z) . (2.37)

This dipole moment determines the amplitude of the nonlinear surface current
in the adiabatic limit.
To illustrate the sensitivity of these quantities to the shape of the density
distribution, all values in Table 2.3 are given for the standard and stabilized
2.1. GROUND-STATE PROPERTIES 21

2r-,------r-----.------.-----~1

N O~~~------~~----------------~

-1

o .01 .02

(J (o.u.)
Figure 2.6: Static image plane position Zl for jellium surface as
a function of net surface charge (solid curve; 1 a.u. = 3.57 e/ A2).
The dashed line indicates the effective edge zil of the density profile.
(Gies, Gerhardts, 1985).

jellium models. In the latter case, the increased work function of high-density
metals makes the surface stiffer, causing an inward shift of the centroid po-
sitions Zl and Z2 and a reduction of P2. The opposite effect is obtained for
low-density metals. The magnitude of these changes is large enough, so that
the variation of Zl and Z2 as functions of r. is reversed: While in the standard
jellium model both positions shift inwards for lower bulk densities, the stabi-
lized jellium model yields the opposite trend (Kiejna, 1995). The value of P2,
on the other hand, increases even more rapidly with increasing r. as a result
of the stabilization correction.
The above procedure for determining the linear and nonlinear static re-
sponse properties of neutral metal surfaces can also be applied to calculate
ground-state properties of positively or negatively charged surfaces. Gies and
Gerhardts (1985) used this method to determine the shift of the static im-
age plane of a jellium surface as a function of net surface charge (see Figure
2.6). As shown by these authors, Zl depends sensitively on the charge state of
the surface. Negative charging lowers the effective barrier height and softens
the density. This leads to an outward shift of the image plane position. The
opposite trend is found for positive charging. In the limit of strong positive
22 CHAPTER 2. DEN~TYFUNCTIONALTHEORY

charging, the image plane approaches the effective edge of the density profile
zlI, defined in (2.16). Negative charging, on the other hand, is limited by the
onset of field emission. Information about electronic excitations of charged
metal surfaces is of particular interest when investigating metal-electrolyte
interfaces.
Linear and nonlinear induced densities are sensitive to the outer regions
of the effective ground-state potential. It is therefore important to study the
influence of nonlocal corrections that reproduce the correct asymptotic image
barrier. This is still a controversial issue since different authors obtain shifts
of the static image plane toward or away from the surface (Ossicini et al.,
1986; Eguiluz, Hanke 1989; Zhang et al., 1990; for a comparison of various
approaches, see Kiejna, 1993). Fortunately, these shifts are in general rather
small because the image form sets in at distances where the density is low.
In view of the computational difficulties involved in calculating dynamical
response properties of realistic metal surfaces, it would be of great interest
to study the influence of lattice effects on the first- and second-order induced
surface charge densities in the static limit. So far, only a few such calculations
have been carried out. Inglesfield (1987) evaluated Zl for AI(OOl) using his
embedding scheme and found Zl R:l 1.1 ao. As the comparison between the
standard and stabilized jellium models shows (see Table 2.3), this inward shift
is mainly related to the higher work function. The corrugation of the electronic
density caused by the crystal structure can lead to an additional modification
of the density profile and to a further shift of Zl. Hence, different crystal faces
exhibit different image plane positions (Perdew et al., 1990).
A similar shift of the static image plane, as well as a reduction of the inte-
grated nonlinear polarizability pz, was found for Ag(OOl). Aers and Inglesfield
(1989) obtained Zl = 0.97 aD and P2 = 110 a.u., while Weinert (1994) found
Zl = 0.91 aD and P2 = 180 a.u. These results demonstrate that virtual in-
terband transitions caused by the lattice potential can have an appreciable
influence on the static screening properties of metal surfaces.

Short-Wavelength Response

Apart from the static response to uniform fields, the surface charge density
induced by an electric potential of the type cPexl(r) rv eiqjl"i'il+qz can also be
derived from an extension of existing ground-state electronic structure codes.
Here, 'VcPext = 0, so that q = Iqj!l. Such potentials represent a periodic
perturbation in the surface region. If qj, is chosen as a simple fraction of
surface reciprocal lattice vectors, it should be relatively straightforward to
find the static induced density nl (f) for realistic metals.
2.1. GROUND-STATE PROPERTIES 23

.4

,'-'v q=O
,.... \
\
r.=2 \
Y.
::J
0
.2 q=k F/2
........
,.... \
\
CT \
N
........ ~

II:
0
Zl

-8 -4 o 4 8

Z (00)

Figure 2.7: Surface screening density fil(Z,q) = e- qZ1 nl(z,q) in-


duced at a jellium surface (T. = 2) by a static electric potential of
the form '" eiqjl"rll+qz. Solid curve: q = kF/2 = 0.9 A-1; dashed
curve: q = O. (Liebsch, 1985a).

To illustrate the q-dependence of the screening density, we show in Figure


2.7 the normalized distribution fil(Z, q) = e- qZ1 nl(z, q) induced by the poten-
tial <Pext(T) = _(271"/q)e iqjl ·r'jI+qZ, where Zl is the image plane position. These
densities were obtained in the static limit using a coupled-channels procedure.
In the small-q limit, fil(Z,q) converges to the induced density calculated by
Lang and Kohn (1973) for uniform fields. At finite q, the main peak shifts
towards the surface since higher-density regions are involved in screening the
applied field. Also, the wave vector of the Friedel oscillation decreases from
a = 2kF to a = (4k~ - q2)1/2 (Dobson, Harris, 1983). The decrease of the
amplitude of fil is caused by the reduced ability of surface electrons to screen
the applied field at finite q.
With the aid of these induced densities, we are able to determine deviations
of the image potential from the asymptotic behavior given in (2.13). Let us
consider a positive unit point charge at a distance d from the surface. The
potential due to this charge and the induced screening density are given by
(the surface area is taken to be unity):

,/, xt(r d) = - -1- / d2q -271" e,qll"rll


"- - eq (z- d) (2.38)
'l'e, (271")2 q ,
24 CHAPTER 2. DENSITY FUNCTIONAL THEORY

-d) =
nl (r, 1 fd 2 qnl ( z,q, d) eiQIl·fjl .
(27r)2 (2.39)

According to the notation used above, we have nl(z, q, d) = e- qd nl(z, q). The
electrostatic interaction energy between the external and induced charges is
then given by

(2.40)

where
g(q} = f dz eqz nl(z,q} . (2.41 )

To first order in q, we find (Lang, Kohn, 1973)

(2.42)

so that
g(q} = 1 + 2qZl . (2.43)
This yields the asymptotic form (2.13). Near-surface corrections are deter-
mined by the behavior of g(q} beyond the linear region. Numerical calculations
show that, for q ~ 0.8 A-I, g(q} is well represented by e2qz , (see Figure 2.8).
Thus, the image form (z == d)

(2.44)

is valid to excellent accuracy as long as the external charge is located outside


the range of the equilibrium density. At shorter distances and beyond the
linear response regime, deviations from (2.44) become important (Appelbaum,
Hamann, 1972).

Extension to Realistic Metals

As we discuss in Chapter 7, induced densities and local potentials asso-


ciated with static short-wavelength perturbations are relevant for a variety of
phenomena involving low-frequency electronic excitations, e.g., sliding friction
of atoms and ions near metal surfaces, surface resistivity, and damping of ad-
sorbate vibrations. In addition, static response properties are important ingre-
dients in calculations of surface phonon spectra (Quong et at., 1991). Because
2.2. TIME-DEPENDENT RESPONSE PROPERTIES 25

,-..
0- 4
'-'
CI

0
0 2

q (.8.- 1)

Figure 2.8: Variation of In g(q) with parallel momentum for several


bulk densities. The dashed lines denote the functions 2qZl' where
Zl are the image plane positions given in Table 2.3. All results are
for the stabilized jellium model. For clarity, the functions for r. == 3
and 4 are displaced upwards.

of the great effort required to evaluate dynamical surface response functions


for realistic metals, static response calculations at finite Qil may well become
the next generation of surface response work. Such studies would significantly
complement the dynamical response of simple metal surfaces that has been of
primary concern in the past.

2.2 Time-Dependent Response Properties

In this section, we describe the extension of the density functional ap-


proach to the presence of time-dependent external potentials. We illustrate
this scheme for a weak potential due to incident electrons far from the metal;
applications are discussed in Chapter 3. Modifications appropriate for other
external probes, such as incident photons or atoms, are treated in other chap-
ters.
26 CHAPTER 2. DENSITY FUNCTIONAL THEORY

2.2.1 Adiabatic Local Density Approximation

The rate of generating electronic surface excitations via an external electric


potential CPext (T, w) can be calculated from Fermi's golden rule:

w(w) = 27r L it;(1- fii') l(k'ICPsCflkW o(cii' - cii - w) , (2.45)


iC,k'

where it; is the Fermi-Dirac distribution function and cii and Ik) = tPii(i') are
the LDA single-particle energies and wave functions discussed in the previous
section. These quantities include all aspects of the ground-state electronic
properties of the semi-infinite metal. The self-consistent potential CPscf(r, w)
that determines the amplitudes of electronic transitions between single-particle
states differs from the applied external potential CPext (f, w) because of screening
processes. The applied potential induces a fluctuating surface charge density
which, in turn, generates an induced potential CPind (f, w). The coherent sum of
these contributions yields the effective or screened complex local potential:

CPscf(f,w) = CPext(f,w) + cPind(f,w) . (2.46)

In the time-dependent extension of the density functional approach (Zang-


will, Soven, 1980; Stott, Zaremba, 1980; Mahan, 1980; See also Ando, 1977;
Peuckert, 1978) this potential is evaluated as follows.
The surface electron density induced by the external potential can be de-
rived exactly from the non local linear response equation:

nl(f,w) = I d r' X(f, f',w) CPexl(T',W) ,


3 (2.47)

where X is the first-order many-body density-density response function (Ma-


han, 1990). Equivalently, the surface density may be calculated from the
expression
(2.48)

In this mean-field treatment, all many-electron correlations are incorporated


into the effective potential CPscf and Xl represents the independent-electron
response function:

(2.49)

where 0 is a positive infinitesimal.


2.2. TIME-DEPENDENT RESPONSE PROPERTIES 27

The basic idea underlying the time-dependent density functional approach


is to write the effective potential tPscf(r, w) as a superposition of external and
induced potentials as indicated in (2.46) and to approximate the induced po-
tential by a sum of electrostatic and exchange-correlation terms. The induced
electrostatic potential satisfies the Poisson equation

(2.50)

The induced exchange-correlation potential is obtained from a Taylor expan-


sion of the ground-state exchange-correlation potential:

A.
'f'xc
(~r, w) = 8Vxc[n] I
~ nl (-)
r, w . (2.51)
no(f)

The induced potential is then given by

tPind(r,w) = tPest(T,w) + tPxc(T,w)


I d r' K(T,r') nl(T',w) ,
3 (2.52)

where
K(T, r') = -1_
1 _'1 + V:c[no (T)] t5(r - r') . (2.53)
r-r
According to (2.51), electron-electron interactions in the presence of the ex-
ternal perturbation are treated in the same manner as in the ground state,
i.e., via the same local exchange-correlation potential Vxc[nJ. This method is
referred to as time-dependent local density approximation (TDLDA). In the
adiabatic limit, this response treatment yields the LDA ground state in the
presence of a static electric field discussed in the previous section.
The response formalism outlined above is quite general and has been ap-
plied to various systems: atoms (Zangwill, Soven, 1980; Nuroh et al., 1982;
Mahan, Subbaswamy, 1990), molecules (Levine, Soven, 1984), small metal
particles (Ekardt, 1985; Puska et al., 1985; Rubio et al., 1992), thin metallic
films (Eguiluz, 1985; Gies, Gerhardts, 1987), and metal surfaces (Liebsch,
1987; Kempa, Schaich, 1988; Dobson, Harris, 1988; Liebsch, Schaich, 1989).
As an example, we show in Figure 2.9 the frequency dependence of the
photoabsorption cross section of aXe atom above the 4d threshold (Zangwill,
Soven, 1980). The agreement between the TDLDA results and the data is
remarkably good. Most of the dynamical response stems from the Coulomb
part of the induced potential. The exchange-correlation contribution weakens
the bare Coulomb interaction and leads to a small but significant downward
shift of spectral weight. This brings the calculated cross section into almost
28 CHAPTER 2. DENSITY FUNCTIONAL THEORY

30
Xe

.......
.c
::IE 20
'-'

'3
b
10

6 7 8 9 10 11
w (Ry)

Figure 2.9: Photoabsorption cross section ofaXe atom as a function


of frequency close to the 4d threshold. Solid curve: time-dependent
LOA; dot-dashed curve: LOA-based RPA (4)xc neglected); dashed
curve: independent-particle result (4)ind neglected). Symbols: ex-
perimental data (Haensel et al., 1969). (Zangwill, Soven, 1980).

perfect agreement with the data. The independent-particle result, on the other
hand, is inadequate because it omits the screening of the incident photon
field. The characteristic effect caused by these screening processes is a shift of
spectral weight to higher frequencies.
The maximum of the cross section near 7.5 Rydb has a physical origin
similar to the Mie plasmon in small metal particles (see Section 3.9). If we
derive from this resonance frequency WM a sphere radius R by using the classical
relation w~ = 47rn/3 = 1O/R3 (n is the average density of the 4d shell), we
obtain R ~ 0.5 A, in rough agreement with the radius of the 4d shell of the Xe
atom (Zangwill, Soven, 1980). The detailed position and shape of absorption
peaks depends, of course, on the microscopic nature of the electronic states
and on the screening properties of the system. The same is true for excitation
spectra of metal surfaces.
Note that, in spite of the consistent treatment of electron interactions in
the ground state and in the presence of the applied field, the TDLOA as
outlined above is not exact: Only correlation effects amounting to a polar-
ization of the electron density are included; higher-order processes, such as
2.2. TIME-DEPENDENT RESPONSE PROPERTIES 29

double excitations or ionization, are neglected. These types of interactions


lead to a frequency dependence of the induced exchange-correlation poten-
tial. The generalization to such potentials is discussed in the next section.
Fortunately, in most cases discussed in this volume, the dynamical surface re-
sponse properties are dominated by the induced Coulomb potential, whereas
the exchange-correlation contribution leads to a weak redistribution of spectral
weight, similar to the example given in Figure 2.9. Thus, in general, correla-
tion effects not included in the TDLDA should not have a major influence on
surface excitation spectra.
Neglecting exchange-correlation contributions to the induced potential is
equivalent to the random-phase approximation (RPA). This does not, however,
amount to the time-dependent Hartree approximation as long as ground-state
properties are treated within the LDAj by RPA we therefore mean 'LDA-based
RPA'. This hybrid electron-electron interaction treatment violates certain sum
rules automatically satisfied by the TDLDA (Liebsch, 1985b). Obviously, this
mixed LDA-RPA response treatment does not give the correct LDA ground
state if the excitation frequency approaches the adiabatic limit. Nevertheless,
since at finite frequencies none of the presently known procedures is exact,
only the comparison with experiment can ultimately tell to what extent a
given physical situation is described adequately.
The solution of the response equations (2.46) - (2.53) represents a self-con-
sistency problem similar to evaluating the ground-state electronic properties.
An important aspect of the numerical solution is the presence of two length
scales: a short-range scale of the order of a metallic screening length, which
characterizes, for example, the density decay at the surface and the correspond-
ing gradients of the local electric fields, and a long-range scale determined by
the slow decay of finite-frequency Friedel oscillations and by the tails of electric
fields in the interior, in particular, in the long-wavelength limit. This compe-
tition of equally important length scales has made the numerical solution of
response equations difficult for many years. In the following section and other
chapters, we discuss the crucial steps that permit an accurate and efficient
solution for several types of external electric fields.

Transition Rate in Terms of Induced Density

The response equations (2.46) - (2.53) suggest that electronic surface ex-
citations may be formulated either in terms of the effective local potential if>scf
appearing in the golden rule (2.45) or in terms of the induced surface charge
density nl. This equivalence can be used to derive a convenient alternative
expression for the transition rate w(w).
30 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Let us consider the quantity

Substituting the explicit form (2.49) of the linear response function Xl yields

(2.55)

where we abbreviate i == k, j == k' and matrix elements are defined as M fi =


UlrP.cfli). Interchanging the dummy variables i and j in the second term, we
easily prove that F(w) can also be written in the form

F(w) "
= L..J Ii IMj;1
ij
2( ci - cf
1 . +
+ w + ZO Ci - cf -
1)
.
w - ZO
(2.56)

Surprisingly, this expression is equivalent to another version of F(w), namely,

F(w) =E
ij
Ii (1- If) IMjil 2 (
Ci - cf +1 w + Z'0 + Ci - 1
cf - w - Z
'0)' (2.57)

Hence, the occupation factor (1 - Ii) is of no consequence for the value of


F(w). This result may be verified by exchanging i and j and expressing F(w)
as an average over both forms. If we now consider the imaginary part of F(w),
it turns out that only the first term in the large parenthesis contributes since
Cj > Ci + w according to the Fermi factors in (2.57).
Comparing this result with the golden rule (2.45), we see that the transition
rate is proportional to the imaginary part of F(w):

w(w) = -21m F(w) . (2.58)

To proceed, we use the self-consistency equation for the local potential

rPscf(r, w) = rPext(r, w) + d3I r'jd r" K(f', r/)


3 Xl (r/, r", w) rPscf(f''', w). (2.59)

Substituting the complex conjugate of this equation into (2.54) and using the
fact that the response kernel function K(r, r') is real-symmetric, the imaginary
part of F (w) may be expressed in the form
2.2. TIME-DEPENDENT RESPONSE PROPERTIES 31

According to the response equation (2.48), we may now write the transition
rate (2.58) as
(2.61)
Surface excitation spectra can therefore be evaluated directly from the spatial
profile of the screened induced surface density. This is analogous to the case of
atomic absorption spectra which may be expressed in terms of the golden rule
(involving the complex potential <Pser) or in terms of the imaginary part of the
complex polarizability (determined by the induced dipole moment) (Zangwill,
Soven, 1980).

Surface Response Function

Let us now consider the case of inelastic electron scattering from the sur-
face of a semi-infinite metal. As long as the incident electJ;on is far from the
metal, the external potential satisfies the Laplace equation \j2<pext = O. This
potential may therefore be expressed as a superposition of evanescent Fourier
components of the form (see (2.38))

(2.62)

The negative sign is chosen so that, in the adiabatic limit, the applied field
induces a negative electronic charge (positive number density) near the surface.
Since we are dealing mainly with simple metals, we use the jellium model
to describe their electronic properties. Because of the two-dimensional trans-
lational invariance of this model, we do not need to introduce reciprocal lattice
vectors parallel to the surface. The induced density is therefore of the form

(2.63)

Substituting this expression and (2.62) into (2.61), we find (Persson, Zaremba,
1985)
411"
w(q,w) = - 1m g(q,w) , (2.64)
q
where the so-called surface response function g( q, w) is the finite-frequency
generalization of g(q) specified in (2.41):

g(q,w) = I dz e qz nl(z,q,W) . (2.65)

Relation (2.64) indicates that surface excitation spectra can be obtained from
the quantity 1m g( q, w), which is referred to as surface loss junction. As will
32 CHAPTER 2. DENSITY FUNCTIONAL THEORY

become evident from the topics discussed in the following chapters, this quan-
tity is of fundamental importance for many phenomena involving electronic
excitations at metal surfaces. For surface excitations, the function g(q,w)
plays the same role as the longitudinal dielectric function cL(q,W) for elec-
tronic excitations in the bulk.

Surface Response Function as Reflection Coefficient

In a system with translational invariance parallel to the surface, the Fourier


components of the induced electrostatic potential satisfy the Poisson equation

(::2 - q2) 4>est(z,q,w) = -471" nl(z,q,W) . (2.66)

Thus, these components may be obtained from the relation

,/,. (z,q,w )
'f'est =q
271" ! d'z e -ql .. - ..'I nl ('
z ,q,w ) . (2.67)

According to (2.53) we may write

4>ind(Z,q,W) = ! dz' K(z,z',q) nl(z',q,w) , (2.68)

where the Fourier components of the static response kernel function are

K(z, z', q) = 271" e-q1z-z'l + V':" [no (z)] o(z - z') . (2.69)
q

The induced density is localized in the surface region (above the volume
plasma frequency, this remains true if damping effects are included). From
(2.68) it then follows that the induced potential has the asymptotic form

4>ind(z,q,w) = -271"q g(q,w) e- qz , (2.70)

== -271" g_ (q,w ) eqz , z«:O, (2.71)


q

where
(2.72)
The surface response function g(q,w) can therefore be interpreted as a gen-
eralized reflection coefficient r(q,w) since it determines the amplitude of the
induced potential in the vacuum (see Figure 2.10). In Chapter 4, we show
2.2. TIME-DEPENDENT RESPONSE PROPERTIES 33

.-'-"_. -
¢l ••t __ - - - - - - - - - - - - - - - - - - - - - .

n,

---
<Pacf
¢I. xt

-10 -5 o 5 10

Z (00)
Figure 2.10: Schematic view of surface screening. Solid curves: in-
duced electronic density nl(z, q, w) and local potential <Pscf(Z, q, w)j
dashed curve: external potential <Pext (z, q, w) j dash-dotted curve:
electrostatic part of induced potential <Pest(z,q,w).

that, at small q, g(q,w) coincides with the reflection coefficient Tp(q,W) of


p-polarized light in the nonretarded limit.

Surface Screening
To illustrate the screening processes included in the TDLDA, it is conve-
nient to write the response equation (2.48) in terms of the external rather than
the local potential. According to the definition (2.68), the Fourier components
of the induced density satisfy the equation:
nl(z, q,w) = j dz' Xl(Z, z', q, w) <Pext(z', q, w)
+ jdz'jdz"Xl(Z,Z",q,w)K(z",z',q)n 1(z',q,w). (2.73)
Because of the exponential form of the external potential and the short-range
nature of the Coulomb kernel, this integral equation may be solved by dis-
cretizing all quantities on a mesh of z points and subsequent matrix inversion.
Schematically, we have (temporarily dropping spatial coordinates and inte-
grals):
(2.74)
34 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Comparing this formula to the exact expression (2.47), it is evident that


the true many-body response function X(q,w) has been approximated by a
screened single-particle response function of the (schematic) form

Xl(q,W)
X(q,w) ~ Xl + X1KXl + ... = 1
-Xl (q,w )K()
q
(2.75)

Hence, as a result of electron correlations in the presence of the time-dependent


external field, the weights of single-electron transitions are redistributed. In
addition, new spectral features may appear due to possible roots of the denom-
inator; these are the collective excitations. The geometric series expression for
X suggests that these collective modes may also be viewed as a coherent su-
perposition of interacting electron-hole pairs.

2.2.2 Dynamical Local Density Approximation

The adiabatic form of the exchange-correlation contribution (2.51) to the


induced potential can be expected to be adequate only if the time-dependence
of the induced density nl is sufficiently slow. Nevertheless, practical applica-
tions give remarkably good agreement with experimental data even at rather
high frequencies. An example is the Xe atomic absorption spectrum shown in
Figure 2.9.
The generalization of the induced exchange-correlation potential that in-
cludes arbitrary dynamical processes is given by the expression (Runge, Gross,
1984; Gross, Kohn, 1985, 1986; Iwamoto, Gross, 1987)

(2.76)

Using (2.47) and (2.48), it can be shown that the response kernel Ixc is formally
related to the inverse of the exact many-body response function via

(- - I )
f xcT,T,W -I) -
T,T,W
=Xl-1 (- - -I) -
X-1 (r,r,w 1 ,
, __ (2.77)
r - r'

In the spirit of the LDA, Ixc is now replaced by the corresponding quantity
of the homogeneous electron gas. If in addition it is assumed that both the
equilibrium and induced densities vary slowly, Ixc is approximated as

Ixc(r,r',w) =8(r-r') Ixc(q=O,w). (2.78)


2.2. TIME-DEPENDENT RESPONSE PROPERTIES 35

Thus, the induced potential is given by

(2.79)

The frequency dependence is now included not only in the induced density but
also in the interaction kernel (2.53).
According to the compressibility sum rule, the static limit of fxc is

(2.80)

This expression coincides with the one employed in the adiabatic version of
the TDLDAj see (2.51).
The high-frequency limit of fxc is determined by the third frequency mo-
ment:

f xc (W -- 00
) -- _i n2/33:.... cxc(n) + 6n1/33:....
d
cxc(n) = f (n ) . (2.81)
5 dn n 5/3 n n 4/3 - 00

Moreover, from perturbation theory one finds the high-frequency behavior

fxc(w) = foo(n) + W~/2 (1 - i) , (2.82)

where c = 237r /15. Interpolating between the low- and high-frequency limits
and using the standard symmetry properties then yields the following imagi-
nary part of f xc :
a(n) w
1m fxc(w) = [1 + b(n) W2J5/4 (2.83)

with

a(n) -c b/C?/3 [foo(n) - fo(n)]5/3 , (2.84)


b(n) = b/C)4/3 [foo(n) - fo(n)]4/3 , (2.85)
'Y [r(I/4)]2/(4y'2;) = 1.311 . (2.86)

The real part of fxc(w) is obtained using the Kramers-Kronig relations. An


extension of the parametrization of fxc(q = O,w) to finite values of q was given
by Dabrowski (1986).
Figure 2.11 shows the real and imaginary parts of fxc derived from (2.83),
for bulk densities corresponding to r. = 2 and r. = 4. The frequency depen-
dence is less pronounced at high densities than at low densities. Since the real
part of Ixc at finite w is smaller than in the adiabatic limit, the excitation
frequencies are shifted upwards. In the RPA response treatment, fxc = 0,
giving still higher excitation frequencies. The dynamical exchange-correlation
36 CHAPTER 2. DENSITY FUNCTIONAL THEORY

,....
:i
~
ci

..
.J
It:
-10

-15
0

,....
:l

~
ci
....=
E

-8
0 2 3
w (a.u.)

Figure 2.11: Real and imaginary parts of fXl!(w) for r. = 2 and 4.


The volume plasma frequencies are wp = 0.61 a.u. = 16.6 eV and
wp = 0.22 a.u. = 6.0 eV, respectively. (Iwamoto, Gross, 1987).

potential should therefore give modes between the adiabatic TDLDA and the
RPA. For the surface collective modes of interest here, the shifts are expected
to be small since the frequencies are rather low on the scale shown in Figure
2.11. The frequency and width of excitations are also modified as a result of
the imaginary part of fxc.
As pointed out by Dobson (1994), the interpolation formula given in (2.83)
is not consistent with the so-called 'harmonic potential theorem' (Kohn, 1961).
In contrast, this theorem is automatically satisfied by the adiabatic local den-
sity approximation discussed in the preceding subsection. For recent develop-
ments of time-dependent exchange-correlation potentials, the reader is referred
to Ullrich et al. (1995) and Vignale and Kohn (1996).
We also note that the interpolation formula (2.83) assumes a smooth fea-
tureless spectral distribution. As a consequence, at low frequencies the real
2.3. COMPUTATIONAL PROCEDURES 37

part of the exchange-correlation potential is a monotonically decreasing func-


tion of w. In realistic systems, this frequency variation should be more com-
plicated once the influence of discrete collective or single-particle excitations
on absorption spectra is taken into consideration. It is conceivable therefore
that 1m fxc(w) exhibits sharp spectral features, so that the real part of fxc(w)
in some frequency range lies below the static value fxc(w = 0) = V:c[n]. Such
an effect could lead to a red shift of spectral weight below the frequencies de-
rived within the adiabatic version of the TDLDA rather than to the blue shift
obtained from the interpolation scheme (Sturm, 1995; B6hm et al., 1996).

2.3 Computational Procedures

In this section, we describe the main steps for evaluating dynamical response
properties of semi-infinite jellium surfaces. The computational procedures are
actually quite simple once important aspects, such as the asymptotic behavior
of electric fields, are properly accounted for. In addition, highly useful sum
rules can be employed. These allow us to circumvent the problem of calculat-
ing dipole moments in the interior where the induced densities exhibit slowly
decaying dynamical Friedel oscillations. Here, we give appropriate expressions
of various quantities and specify important parameters.
We also introduce briefly the embedding approach proposed by Inglesfield
(1981), which permits a convenient formulation of the dynamical surface re-
sponse of semi-infinite three-dimensional metals. Electronic excitations in this
case also form a continuum down to the adiabatic limit. An alternative is to
consider electronic excitations of thin slabs (Eguiluz, 1985; Gies, Gerhardts,
1987) and to generalize these schemes to include the crystal potential. How-
ever, slab methods have the disadvantage that, in the limit of small parallel
wave vectors, coupling between slab surfaces leads to a splitting of collective
modes. In addition, it is non-trivial to deal with quasi-discrete single-particle
transitions below the threshold for emission.

2.3.1 Eigenfunctions

We consider first the evaluation of the ground-state electronic density of a


semi-infinitejellium system. The effective one-electron potential V(z) == Veff(Z)
38 CHAPTER 2. DENSITY FUNCTIONAL THEORY

in this model depends only on the coordinate Z normal to the surface. The
wave functions are therefore of the form

(2.87)

with k == k z . The functions 1/Jk(Z) satisfy the one-dimensional Schrodinger-like


equation
(2.88)

with Ck = 0.5 k 2 + Veff( -00). We choose the vacuum level as zero of the energy
scale. The eigenenergies ci< are given by

(2.89)

Since the potential is constant sufficiently far from the surface, the wave func-
tions 1/Jk(Z) can easily be obtained by numerical integration.
Let us define a mesh of M points Zi in the interval between Zl and Z M. An
adequate choice is Zl = -5 ... - 10 r. and ZM = 5 rs with a mesh size L). =
Zi+l -Zi = 0.05 r •. Since the Fermi wavelength is AF = 27r/k F = 3.274 r., this
mesh size implies about 60 points within one wave function period for states
near the Fermi level. The first point Zl is chosen so that Friedel oscillations of
Ve1f(z) in the interior can be neglected. The outer point ZM is sufficiently far
from the surface so that the one-electron potential is negligibly small. Hence,
M = 201 ... 40l points are usually sufficient.
Because of the exponential form of the bound states in the vacuum region,
it is numerically most stable to integrate them inwards. Since Veff(Z) :::::: 0 for
Z 2 ZM, we may choose

(2.90)

for i = M and i = M -1. Here, '" = (-2ck)1/2. The normalization constant


a is specified below. Wave function values at subsequent mesh points can
be found by using, for example, the Numerov integration procedure (Hartree,
1957):

1/Ji-l = 2 + 10h(V; - Ck) 1/Ji _ 1 - h(V;+1 - ck) 1/J'+1 (2.91)


1 - h(V;-l - Ck) 1 - h(V;-l - Ck) • ,

where h = /),2/12 and Vi == V.ff(Zi)' This method is the most accurate three-
point integration of second-order differential equations with corrections of or-
der L).6.
2.3. COMPUTATIONAL PROCEDURES 39

Close to Zl, the potential is constant: V; = Veff( -oo}. Wave functions


in this region therefore have a sinusoidal shape. We define the normalization
parameter a so that in the interior of the bulk 1{;k(Z} has the asymptotic form

1{;k(Z} = sin(kz -,k} . (2.92)

The phase shifts Ik depend on the shape of the surface barrier potential. They
satisfy the sum rule (Sugiyama, 1960; Langreth, 1972)

2
k~
r
Jo
kF
dk k Ik
7r
(2.93)
4

The ground-state density (2.5) is given by

(2.94)

The factor (k~ - k2 ) in the integrand arises from the analytical integration
over parallel momenta.

2.3.2 Ground-State Density

To achieve self-consistency it is convenient to start with a model density,


for example:

no(z) fi (1 - 0.5 eOZ ) z ~ 0,


= fi 0.5 e- oz , z ~ O. (2.95)

The one-electron potential corresponding to this density is given by

V(z) = -47r 1 dz' (z' - z) [no(z') - n+(z')] + Vxc[no(z)] .


00
(2.96)

The density of the positive background is defined in (2.14) and the exchange-
correlation potential can be taken, for example, from expressions (2.10) or
(2.12).
The new density calculated from this potential does not, in general, satisfy
charge neutrality. To avoid difficulties arising from the long-range Coulomb
potential due to this charge imbalance, it is helpful to scale the density of the
positive background at each iteration step so that neutrality is restored (for a
detailed discussion, see Lang and Kohn, 1970).
40 CHAPTER 2. DENSITY FUNCTIONAL THEORY

A more stable way of dealing with the long-range Coulomb potential is to


convert it into a short-range form by writing the Poisson equation (2.18) as
(Manninen et al., 1975)

ifJ(z) = f dz' e-"lz-z'l {~ [no(z') - n+(z')] + ~ ifJ(z,)} . (2.97)

This expression is valid independently of the parameter /'i,. A reasonable choice


is /'i, ~ kF • The great advantage of this form is that the asymptotic behavior
of the Coulomb potential can be explicitly incorporated as an initial condi-
tion. Oscillations of the surface charge during the iterative procedure remain
localized in the surface region and do not affect the asymptotic form of the
potential. Iteration to self-consistency may then proceed by a small admixture
of the new density to the previous one. Since the Friedel oscillations in the
equilibrium density are more pronounced than in the effective potential, it is
important to take them into account in the neutrality check at each iteration
step. The convergence can be greatly accelerated by using an advanced ver-
sion of the Anderson procedure as suggested by Bliigel (1987). This method
was used to calculate ground-state electronic properties of neutral as well as
charged jellium surfaces.

2.3.3 Green's Functions

The independent-particle susceptibility Xl defined in (2.49) involves sums


over the complete spectrum of occupied and unoccupied single-particle states.
To avoid these spectral representations, it is useful to express Xl in terms of
Green's functions and to evaluate these directly at the appropriate energy.
Green's functions satisfy a differential equation analogous to (2.2):

(2.98)

where c: is an energy parameter. These functions may be expanded in terms


of the single-particle states 1/Jk :

(2.99)

The sign of the imaginary part in the denominator depends on boundary con-
ditions.
2.3. COMPUTATIONAL PROCEDURES 41

In the one-dimensional case, the Green's functions are obtained from the
equation
[-~~2+VeJf(Z)-C] G(Z,Z',C) = -6(z-z'). (2.100)

They have the spectral representation


tfJle( z)tfJi.( z'}
G(z,z',c} = LIe e - Cle ±i6
. (2.101)

These functions may be evaluated directly from the expression


G( , ) = 2 "iL(zd w+(z» (2.102)
z,z ,c [w+, w-l '
where (z<,z» = (z,z') if z < z1 and (z<,z» = (z',z) if z' < z. The
functions w± are solutions of (2.88) at the complex energy c with the following
boundary conditions (Liebsch, 1986):
w_(z) ~ exp(-ik+z), z«:O,
(2.103)
and
w+(z) ~ exp(i,,+z), z» 0 ,
~ cexp(ik+z) +dexp(-ik+z) , z «: O. (2.104)
Here, k+ = [2c - 2V(-ooW/2 and "+ = (2c)1/2. The subscript + of k+
and "+ indicates that complex roots are to be taken so that 1m k+ ~ 0
and 1m/\:+ ~ O. The Wronskian [w_, \11'+1 in (2.102) can be found from the
asymptotic behavior of \II'±:
d d
= w _(z) dz \II' +(z) - \II' +(z) dz \II' _(z)
= 2ik+c = 2i,,+b . (2.105)
The functions w± (at the complex energy c) can be integrated using the
Numerov method as described above. The integration of W_ is started inside
the metal and continued outward until the asymptotic region in the vacuum is
reached. The function \11'+ is integrated inward, starting on the vacuum side,
until the constant-potential region in the interior is reached. The coefficients
a, b, c, d are then found by matching the numerical solutions to the asymptotic
forms specified in (2.103) and (2.104). Inside the metal, \11'_ is regular (the
ingoing solution: propagating or decaying toward the interior), while w+ is
regular outside (the outgoing solution: propagating or decaying away from the
metal).
42 CHAPTER 2. DENSITY FUNCTIONAL THEORY

2.3.4 Response Functions

The Green's function defined in (2.99) may be used to rewrite the linear
response function Xl (2.49) in the form:
-
-, )
Xl (r,r,w = L II< ['¢f(T)'¢I«f') G(f,f',cl<+w)
k
+ '¢k(T),¢'[;(f') G*(f,f',Ck- W)] . (2.106)

Spin factors are included in the summation over occupied states. In the case
of a jellium surface, it is convenient to go to a Fourier representation parallel
to the surface:

XI(z,z',tlil,W) = / d2 rll e-iolil·(r-r')11 XI(f,f',w) . (2.107)

Using (2.99) these Fourier components can be expressed as

XI(Z, z',tlil,w) = L If '¢k(Z)'¢k(Z') [G(z, z',c+) + G(z,z',c)] (2.108)


I<
with k == k z , Ck = V(-oo) + k 2 /2, and
C± = Ck ± (w + i8 - kll . tlil) - Qjf /2 . (2.109)

The functions '¢k(Z) are the real bound states introduced in Section 2.3.1. Ifwe
assume the vector tlil to lie in the x-direction, the sum over ky can be performed
analytically, and we obtain

where we used the symmetry relation G(z,z',c) = G(z',z,c) following from


the definition (2.102).
In the long-wavelength limit Itlill -+ 0 relevant for incident photons, the
integration over k", can be done analytically. The expression for Xl then sim-
plifies to

XI(Z, z',w) 1
2" IokF dk (kF2 - k2) '¢k(Z )'¢k (')
z
7r 0
x [G(z, z', ck + w) + G(z, z', ck - w)] (2.111)

Expressions (2.110) and (2.111) for the linear susceptibility Xl are also
applicable at complex frequencies. Thus, they may be used to describe the
2.3. COMPUTATIONAL PROCEDURES 43

van der Waals attraction between a neutral atom and thejellium surface. This
interaction can be written in terms of a spectral distribution of Xl along the
imaginary frequency axis, i.e., w = iu. The definition of the Green's function
in terms of the states w± remains valid at these frequencies.
The corresponding second-order response function X2, which is used to de-
termine the nonlinear optical response of jellium surfaces to a uniform electric
field normal to the surface, is given by (Zangwill, 1983; Senatore, Subbaswamy,
1986; Liebsch, Schaich, 1989)

X2 (Z, Z,,,
,z ,w, w ) = :2 fok Fdk (k~ - k2)
X [vlk(Z) G(z, z', ck + 2w) G(z', z", ck + w) vlk(Z")
+ vlk(Z)G(Z,Z',Ck - 2w)G(z',Z",ck -W)vlk(Z")
+ vlk(Z") G(z, z", ck - w) G(z, z', ck + w) vlk(Z')] .
(2.112)
The preceding response functions are appropriate for evaluating the electronic
density induced by an external potential. This potential may be generated, for
example, by an incident electron or by the normal component of the electric
field of an incident photon.
For the evaluation of the surface current induced by a general electric field,
we need the conductivity tensor. In linear response, it is given by the expression
(Mahan, 1990)
~~' ,w ) =
Uij ( r,r ut(f,r',w) +uf/i,i',w) , (2.113)

.£w no (f) Oij o(i - i') , (2.114)

= .£ L (A - A,) oi(T)oj(i')
w f,k'
vlf(T) vlf(f') vl,.,(T) vlf,(i')
x ,
w+ c,. - Cf, + iO (2.115)

where the operators Oi( T) are defined as

Oi(T) aO(T) b(T) =~ [aO(T) ob(T) _ oa*(T) b(T)] (2.116)


2z or; ori
Using the spectral representation (2.99), ut may be written in terms of Green's
functions:
ut(i,i',w) = .£w LA
_ Oi (T)Oj (f') [vlf(T) G(i, i', c,. +w) vlf(i')
k

(2.117)
44 CHAPTER 2. DENSITY FUNCTIONAL THEORY

Similarly, in the nonlinear case, we need second-order response functions


of the type (Ishida, Liebsch, 1994)

Xijk ( r, r ,r ,w ) =
~ ~,~"
L J,; oi(T)oj(f')o,,(f")
li
x [1/if(T) G(f, f', cli + 2w) G(f', f", cli + w) 1/iii(f''')
+ 1/ili(T) G*(f, f',cii - 2w) G*(f', f",cli - w) 1/if(f")
+ 1/if(f") G*(f',f''',cli - w) G(r, f",cii+ w) 1/iii(f")] .
(2.118)

2.3.5 Induced Density

The response equations for the induced surface charge densities can be con-
veniently solved in real space by making use of the response functions specified
in the previous section. We outline here the procedure for inelastic electron
scattering (Tsuei et al., 1991). Modifications required for the linear and non-
linear response to incident photon fields are addressed in Chapters 4 and 5.
Step 1 in solving response equation (2.73) is to evaluate the susceptibility
XI(Z, z', q, w) via expression (2.110). Because of the slow decay of the induced
density and complex potential at finite q and finite w, Xl is needed on a wider
range [Zl! ZN] than the range [Zl' ZM] specified in Section 2.3.1 for the bound
states. On the other hand, the induced density does not oscillate as rapidly
as the wave functions. It is therefore sufficient to solve the response equation
on a coarser mesh 6.' = Zi+l - z,. We choose 6.' = 46., ZN = ZM and
Zl = ZN - 2m (ZM - Zl). For parallel momenta q ~ 0.15 A-I, m = 1
is adequate. For smaller values of q, m should be increased to 2 or 4. If
M = 201, this implies N = 1 + 2m (M - 1)6./6.' = 1 + 100m mesh points on
which the response equation is to be solved. The integration range for bound
states and Green's functions, and the region in which the induced density is
evaluated, are shown schematically in Figure 2.12.
To find the bound states 1/ik(Z) and Green's functions G(z,z',c±) out-
side the range [ZI, ZM], we use the asymptotic expressions (2.92), (2.103) and
(2.104). The evaluation of the k and k., integrations in (2.49) must be done
accurately. Typically, we decompose the k (k.,) range into 15 - 50 (10 - 30)
intervals of four Gaussian points.
Step 2 is to evaluate the unscreened induced density

el(Z,q,W) = I dz' XI(Z,z',q,w) ifJext(z',q,w) . (2.119)


2.3. COMPUTATIONAL PROCEDURES 45

z,

asymptotic region .....

z, o
Figure 2.12: Integration ranges [Z!, ZM] for bound states and
Green's functions, and [Zl' ZN] for induced densities. The one-
electron potential is assumed to be constant in the regions Z ~ Zl
and Z 2 ZM. In the evaluation of the unscreened induced density
6, the asymptotic region -00:::; Z :::; Zl is fully included.

In the range z' 2 Zl, Simpson integration is adequate. However, it is crucial


not to neglect the asymptotic range z' ~ Zl' This contribution can easily be
taken into account by using the asymptotic forms of bound states and Green's
functions. By performing first the z' and then the k, k., integrations in Xl , we
can do the z' integration analytically in the entire range z' = -00 ... Zl'
The response equation (2.73) may now be solved via matrix inversion:

(2.120)

where nl and 6 are vectors and Xl and K are matrices on the mesh points
Zl ... ZN' The response kernel K is defined in (2.69) and 1 is the unit matrix.
The Simpson weight factors W ij and Wi; should be chosen so that cusps in
Xl(Z,z',q,w) and K(z,z',q) at z=z' are properly taken care of. To include
the asymptotic region in the internal integration of Xl K, the Coulomb kernel
is written as

e -qlz' -z"l =e q(z' -z") + [-q1z'


e -z"l -eq(z' -z")] . (2.121)

The integral over Xl times the first term on the right-hand side is related to the
unscreened induced density (2.119), i.e., the asymptotic region can be treated
46 CHAPTER 2. DENSITY FUNCTIONAL THEORY

analytically as previously explained. The integral over Xl times the term in the
large brackets is limited to z' > z" since this term vanishes for z' < z". The
accurate handling of the asymptotic region ensures that the surface plasmon
dispersion can be calculated in a stable manner at q values at least as small
as 0.05 A-1. The linear behavior at small q is perfectly consistent with the
coefficient d.l(w.) derived within the TDLDA in the long-wavelength limit.
The procedure just outlined is significantly more accurate than inversions of
the (schematic) form

n1 = [1-~xJ ¢ext
(2.122)

which usually ignore the asymptotic region.


In step 3, the surface loss function Img(q,w) is obtained from the expo-
nential moment defined in (2.65). In the case of inelastic electron scattering,
there is no need to use special procedures to deal with Friedel oscillations in the
interior. The exponential decay of the applied and local potentials is sufficient,
at least for q ~ 0.05 A-1, to achieve accurate integrations if the spatial range is
chosen large enough. The response to incident photons is in this regard much
more subtle. The treatment of the asymptotic fields in this case is discussed
in Chapters 4 and 5.

2.3.6 Embedding Approach

Because of the one-dimensional nature of semi-infinite jellium surfaces, the


real-space version of the dynamical response outlined above is quite straight-
forward and computationally rather simple. The most time-consuming step
is the evaluation of response functions which are represented as matrices of
dimension N x N, where N is typically a few hundred. The response equation
is solved via matrix inversion. More complex electronic properties in the di-
rection normal to the surface do not significantly increase this numerical effort
since summations over parallel momenta of electronic states can still be done
analytically (over k"" ku for tlil = OJ over ku for q", > 0, qu = 0). Thus, matrix
dimensions do not increase (except for possible adjustments of the width of
the surface range).
Realistic systems, of course, require us to consider also the crystal structure
parallel to the surface. It is therefore necessary to introduce parallel reciprocal
lattice vectors, which we denote as §. The real-space formulation perpendic-
ular to the surface is then impractical since the size of matrices representing
2.3. COMPUTATIONAL PROCEDURES 47

,
'E: ",'

embedded region

adatom

jellium

o
Figure 2.13: Geometry used in embedding calculations, here for the
case of adsorption of alkali metal layers on a jellium substrate. The
surface region in which the electronic structure and the dynamical
response are calculated self-consistently, lies between Zl and ZM.
The basis functions have the range [za, Zb].

the Green's functions becomes quickly very large. Evaluating the response
function would be too time-consuming.
This problem can to some extent be alleviated by using the embedding
scheme proposed by Inglesfield (1981). In this procedure, the surface region
[Zl, ZM] is treated explicitly (see Figure 2.13), while the influence of the semi-
infinite interior of the metal (z :::; Zl) and of the vacuum region (z ;:::: ZM) is
taken into account via complex embedding potentials. In the simplest case,
the surface region is fully three-dimensional, while the internal and external
regions are assumed to be one-dimensional. In a more refined version, this
assumption can also be dropped.
To describe the one-electron wave functions and their logarithmic deriva-
tives at the embedding surfaces, the Green's function

(2.123)

is expanded in terms of the non-orthogonal basis set

- ( 2 )
"pn(kll + §, f) = Zba A
1/2 '(k
e'
-)-
11+9 'TII sin(pn(z - Za)] , (2.124)

where Pn = mr I Zba (n > 0) and Zoo = Zb - Za; [za, Zb] specifies the total range
in which the basis functions are defined. Typically, Za = Z1 - 2 a.u. and
Zb = ZM + 2 a.u. The surface area is defined as A. The matrix element of the
48 CHAPTER 2. DENSITY FUNCTIONAL THEORY

energy-dependent embedding potential at Zl is given by

1 [(kll + ff/ - 2cr/2 sin[pn(zl - Za)]


Zoo
x sin[Pn,(zl - za)] Ogg' , (2.125)

where the one-electron energy c is measured from the bottom of the bulk
jellium potential. The matrix element at Zb is obtained by replacing Zl by ZM
and c by c - LlV, where LlV is the height of the surface barrier.
In the embedding region, the induced density has an expansion in terms of
cosine functions of the form

n(f') = :E:E n{jl eig.rll COS[PI(Z - za)] . (2.126)


9 I~O

Analogously, the response functions are expanded as

x(r, r') = L: L: Xgl,g'I' e i (g.f'II-g'·'II) COS[PI(Z - za)] COS[PI'(Z' - Za)]. (2.127)


gg' II'~O

With these expansions, the dimension of the matrices (for the spatial varia-
tion in the z-direction) is reduced from N x N to Lx L, where L specifies the
total number of cosine basis functions. For realistic alkali metal adsorbates,
for example, L ~ 20 ... 40 is adequate (rather than N ~ 101 ... 201). Hence,
coverages down to 0.2 of a monolayer can still be handled. (The number of g
necessary to represent the lateral corrugation increases with decreasing cover-
age.) Results of dynamical response calculations based on this procedure are
discussed in Chapter 4 for overlayers (Ishida, Liebsch, 1992) and in Chapter 5
for stepped metal surfaces (Ishida, Liebsch, 1994).
Chapter 3

Surface Plasmons

Electronic excitations at metal surfaces can be readily observed via inelastic


electron scattering. The most prominent feature in such loss spectra is the
surface plasmon. We discuss first the surface plasmons of several simple met-
als (Mg, AI, Na, K, and Cs) which can be quite well understood in terms of
the eigenmodes of a semi-infinite electron gas. This is true not only for the
ordinary monopole surface plasmon, but also for the weaker multi pole surface
plasmon appearing at slightly higher frequencies. Severe changes due to in-
terference with interband transitions are observed on Li, but the dispersion
of the surface plasmon remains qualitatively similar to that for semi-infinite
jellium. The most striking deviation from this general behavior is seen on
Ag, where the occupied 4d bands not only affect the overall frequency of the
surface plasmon but also its dispersion with wave vector. Noticeable effects
due to shallow core levels are also found for Hg. Other types of modifications
arise if the metal is charged or if it is in contact with a dielectric medium
or metallic adsorbates. These spectra illustrate the remarkable sensitivity of
dynamical response properties to surface conditions. New surface modes and
intriguing interferences among them occur in thin alkali metal overlayers. As
their coverage is increased, the evolution of adsorbate-induced modes to those
at semi-infinite alkali metal surfaces can be observed. The collective modes of
thin adsorbed Ag layers differ from those of alkali metal overlayers because of
the presence of d bands. Finally, the wave vector dispersion of plasmons at
fiat metal surfaces is closely related to the size dependence of the Mie plasmon
in small metal particles and to collective modes in quantum wells.

49

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
50 CHAPTER 3. SURFACE PLASMONS

3.1 Classical Picture and Hydrodynamic


Models

In the classical local optics picture, the metal is treated as a semi-infinite


electron gas with an abruptly terminated electronic density profile. The fluc-
tuating surface charge Pi == enl = -nl then corresponds to a delta function
sheet localized at the boundary of the metal:

(3.1)

where !Jil = (q"" qll) is the wave vector parallel to the surface. Since the system
remains neutral during the charge oscillation, the fluctuating density integrates
to zero:
(3.2)
Thus, along the surface positive and negative charge regions alternate. The
induced density therefore has a dipolar form. The electric field associated with
this density is determined by Gauss's law:

(3.3)

As long as we are concerned with parallel wave vectors that are not too small
(larger than the wave vector of light at optical frequencies, i.e., 1!Ji1i > 0.01
A-1), retardation effects are negligible. We can therefore limit the discussion
to purely longitudinal plasma oscillations and represent E in terms of a scalar
potential <.P == fjJ/e = -fjJ, i.e.,

E(f',w) = VfjJ(f',w) , (3.4)


with
(3.5)
Since V2fjJ = 0 everywhere except at z = 0, this potential must take the form
(3.6)

where q == 1!Ji1i. Away from the surface, this potential decays exponentially
with a range proportional to l/q (see Figure 3.1). From this expression it
follows that the parallel component of the electric field varies continuously
near the surface, whereas the normal component is discontinuous:

(3.7)
3.1. CLASSICAL PICTURE AND HYDRODYNAMIC MODELS 51

~C\r;
----0-
-0-+-+-+-0---

o z
~~~x
Figure 3.1: Profile (normal to surface) of fluctuating surface charge
density nl (z, w) and electrostatic potential ifJ( z, w) in local optics
picture (left). Lateral distribution of surface charge and electric
field lines (right).

If we represent the metal in the long-wavelength limit by the macroscopic


dielectric function c:( w), we have

(3.8)

The discontinuity specified in (3.7) therefore implies

c:(w) = -1 . (3.9)

This condition defines the frequency of the surface plasmon in the tJil = 0 limit.
For a Drude metal with c:(w) = 1 - w;/w2, we have

(3.10)

where wp = (411'fie 2/m)1/2 is the volume plasma frequency. The existence of a


self-sustained electronic surface excitation was predicted in 1957 by Ritchie in
his treatment of characteristic energy losses of fast electrons passing through
thin metallic films. Shortly afterward, Powell and Swan (1959) observed loss
spectra of electrons reflected from thick Al and Mg films, which could be
interpreted as a superposition of volume and surface plasmon excitations (see
Figure 3.2).
For a metal in contact with a thick oxide layer or another dielectric layer
rather than with vacuum, Stern and Ferrel (1960) showed that the condition
(3.8) is replaced by
(3.11)
52 CHAPTER 3. SURFACE PLASMONS

40'

o 40
Energy loss (eV)

Figure 3.2: Energy loss spectra of fast electrons reflected from an Al


surface. Near grazing incidence (8 = 88°), the spectrum consists of
multiple excitations of surface plasmons (w. ~ 10.5 eV). At smaller
incident angles (8 = 40°), mUltiple excitations of both surface and
volume plasmons (wp:=:::: 15.5 eV) are observed. (Powell, 1968).

where co is the macroscopic dielectric function of the overlayer. Thus, instead


of (3.9), we have
e(w) = -co. (3.12)
The surface plasma frequency of a Drude metal is in this case shifted according
to
w. = wp . (3.13)
JeD + 1
This red shift was also observed by Powell and Swan (1960) for oxidized Al
and Mg surfaces.
The classical picture just outlined ignores the microscopic spatial distribu-
tion of the electron density in the surface region. Moreover, it describes the
dynamical surface response in terms of local dielectric properties of the bulk.
The frequency of the surface plasmon in the long-wavelength limit is not a
3.1. CLASSICAL PICTURE AND HYDRODYNAMIC MODELS 53

surface property since it depends only on the bulk dielectric function c(w).
Introducing a q-dependence in the dielectric function of the bulk leads to a
wave vector dispersion of the surface plasma frequency (Ritchie, 1963). This
variation does not, however, reflect the electronic properties of the surface.
Since the wavelength of an electronic surface excitation detected in electron
energy loss spectra can be as small as the decay length of the surface density
profile, it is clear that the microscopic electronic surface properties must be
taken into account to describe the momentum dispersion of surface plasmons.

N on-abrupt Density Profile

In an attempt to incorporate the smoothly decreasing density distribution,


Bennett (1970) considered a linear profile and used hydrodynamic theory to
solve for the electronic response (see Figures 3.3 and 3.4). In the small-q
limit, this model also yields the surface plasma frequency wp /..j2. At finite q,
however, the surface modes exhibit two important features not contained in
the classical picture:
• at small finite q, the plasma frequency w.(q) shifts linearly downwards;
after reaching a minimum near 0.1 to 0.2 A-1, w.(q) increases;

• a second surface mode, wm(q), the so-called multipole surface plasmon,


appears at slightly higher frequencies; it shows a positive wave vector
dispersion even at small q.
Both features are sensitive to the width a of the surface region. For a ~ 3 A,
the lower mode ws(q) agrees approximately with the dispersion observed by
Kunz (1966) in electron energy loss measurements on Mg. The upper mode
wm(q) was observed only recently by Tsuei et al. (1990, 1991) for several alkali
metal surfaces. If the width a is reduced, this multipole surface plasmon ceases
to exist and the ordinary surface plasmon exhibits a positive dispersion also
at small q. In the limit of small a, the dispersion agrees with the one obtained
by Ritchie (1963) for an abruptly terminated electron density. Evidently, the
dispersion of the surface collective excitations is a consequence of the shape of
the inhomogeneous equilibrium electron density.
As we see shortly, the dispersion of surface excitations obtained by Bennett
agrees qualitatively with the microscopic picture derived within the TDLDA.
Incidentally, Bennett's hydrodynamic calculation appeared in the same year
as the LDA ground-state calculations for jellium surfaces by Lang and Kohn
(1970). It seems curious in retrospect that finite-q response calculations (for
54 CHAPTER 3. SURFACE PLASMONS

no

x
z

~
~~---- ~
--- + --- -
----_.:1"___ ---
x
a/2 o -a/2 z

Figure 3.3: Left: Model surface electronic density profile with lin-
ear decay over a region of width a. n+ is the positive ionic back-
ground. The density profiles of the monopole and multipole surface
plasmons are plotted underneath. Right: Oscillatory patterns of
monopole and multipole surface plasmons parallel to the surface.

0.2
o .2 .4 .6
q (A-l)

Figure 3.4: Dispersion of electronic surface modes in linear-density


model. The labels denote the width a (in A) of the surface region.
The lower mode is the usual surface plasmon. The symbols rep-
resent the experimental data for Mg by Kunz (1966). The upper
mode is the multipole surface plasmon. The dashed line indicates
the dispersion of the volume plasmon. (Bennett, 1970).
3.2. INELASTIC ELECTRON SCATTERING 55

instance in the RPA) based on this LDA potential were not carried out until
20 years later.
As a result of the finite width of the ground-state density profile, the fluc-
tuating densities associated with the surface modes also have a finite width.
In the z-direction, the distribution of the ordinary surface plasmon consists
of a simple peak; i.e., it has monopole character. The charge distribution of
the upper mode has a node; i.e., it is of dipole form. Because of this spa-
tial form perpendicular to the surface, this mode is referred to as 'multi pole
surface plasmon'. Parallel to the surface, both modes propagate like plane
waves, with positive and negative surface charge regions alternating periodi-
cally. Consequently, along the surface both modes have a dipolar shape. The
global symmetry of the multipole plasmon is quadrupolar.
It is clear from these results that the distribution of the electronic density
has a decisive influence on the excitation spectra of metal surfaces. This is
especially true for the multipole surface plasmon. Its shorter effective wave-
length normal to the surface implied by the extra node makes it naturally
very sensitive to details of the density profile. Although hydrodynamic models
(Eguiluz, Quinn, 1973) include some of the ingredients of more sophisticated
microscopic treatments in an approximate manner (such as the finite decay
length of the density and nonlocal response features), other aspects are miss-
ing, in particular, the coupling of collective surface modes to electron-hole pair
excitations. As pointed out by Schwartz and Schaich (1982), depending on the
ground-state density profile and assumed boundary conditions, the number
of collective surface excitations and their momentum dispersion greatly vary.
Hence, such models tend not to have sufficient predictive capability.

3.2 Inelastic Electron Scattering

Before describing the results of quantum mechanical calculations of elec-


tronic surface excitation spectra, we briefly discuss the commonly used experi-
mental arrangement for detecting electronic surface modes and the connection
between the measured quantity and calculated spectra.
The most convenient and direct way of measuring electronic surface exci-
tation spectra is via inelastic electron scattering. The dominant contribution
to this process results from elastic reflection of the incident electron beam via
interaction with the crystal potential, combined with small-angle inelastic scat-
tering via excitation of electron-hole pairs or collective modes (Ibach, Mills,
1982). The inelastic event occurs far from the surface, either before or after
56 CHAPTER 3. SURFACE PLASMONS

R' x

Figure 3.5: Dominant inelastic-scattering processes: loss (L) after


elastic reflection (R) and loss before elastic reflection (R') from
metal surface (left). Wave vector diagram for incident and scattered
electrons (right).

elastic scattering (see Figure 3.5). Since the long-range electric field of the
incident electron interacts with the fluctuating dipolar field associated with
the induced surface charges, such processes are referred to as dipole scattering.
The spatial extent of the dipole field increases exponentially with decreasing
q". Accordingly, the inelastic-scattering cross section is strongly pointed in the
forward direction, so that the scattered intensity is strongest at angles close to
the specular direction ('dipole lobe').
In the so-called impact regime, on the other hand, the inelastic-scattering
event occurs in the surface region. Although in this case the angular distribu-
tion of the scattered intensity is usually substantially different, the frequencies
of electronic surface modes should be the same as in the dipole limit since they
are an intrinsic response property of the surface.
Let us denote the energy and polar angle of the incident beam by E and
f),those of the inelastically scattered beam by E' and f)', respectively. Energy
and parallel momentum conservation imply E = E' + nw and kll = kl, + qj"
where nw and qj, are the energy and parallel momentum of the excited surface
mode. There is no conservation of perpendicular momentum. By varying the
energies and angles of the incident and scattered electron beams, a wide range
of excitation energies and wave vectors can be scanned.
In the dipole-scattering limit, the scattering efficiency S per unit solid angle
3.2. INELASTIC ELECTRON SCATTERING 57

df! and per unit energy dliw is given by (Mills, 1975):

IV.lq(R + R') + i(R - R')(w - Vii' <ii1)i2 P(<iil, w)


[Viq2 + (w - Vii' <ii1)2j2 q2
(3.14)
Here, k and k' are the magnitudes of the wave vectors of the incident and
scattered electrons, and V.l and Vii are the normal and parallel velocities of
the incoming electron. As before, we define q = l<iii!. The specular elastic-
scattering amplitude is denoted as R for loss after reflection and as R' for loss
before reflection.
While the first two factors in (3.14) depend purely on the scattering ge-
ometry and elastic reflection conditions, the quantity P( <iiI, w) specifies the
probability of generating electronic losses of frequency wand parallel momen-
tum <iiI' From Fermi's golden rule, we find

(3.15)

where ¢>ext (i, w) is the external potential specified in (2.62), and x(i, i', w)
is the exact many-body density-density response function of the semi-infinite
metal that appeared in response equation (2.47). (The surface area is taken
to be unity.) In the TDLDA, X is replaced by the screened single-particle
response function Xd(1 - KX1), as indicated in (2.75). The effective inter-
action K includes both Coulomb and exchange-correlation terms. Thus P is
approximated as

Using (2.48) and (2.61), we notice that P is related to the transition rate
w(q, w) given by (2.45), or, equivalently, to the imaginary part of the surface
response function g(q,w) defined in (2.65):

q q2
P(<iil,w) = 27T 1m g(q,w) = 87T 2 w(q,w) . (3.17)

In the following sections, we focus on the mode dispersions derived from the
maxima of Img(q,w). Note, however, that the kinematic factors in (3.14)
can vary rapidly with energy and momentum. This effect can lead to spectral
distortions and apparent shifts of the mode frequencies (Gaspar et al., 1991).
58 CHAPTER 3. SURFACE PLASMONS

3.3 Simple Metal Surfaces

The surface plasmon was predicted by Ritchie in 1957. Nevertheless, a


systematic understanding of the wave vector dispersion of electronic surface
modes emerged only a few years ago (Tsuei et al., 1989, 1990, 1991). One rea-
son for this surprisingly long delay is that surface plasmons are most clearly
defined on simple metal surfaces, which are not easy to control experimentally.
Better sample preparation techniques and high-resolution experimental equip-
ment today provide a greatly improved basis for detailed analysis. The second
reason is that the theoretical description of these modes is very sensitive to
the electronic structure of the surface region. Hence, approximate treatments
of the ground-state electronic properties or the response at finite frequencies
can easily give qualitatively erroneous spectra.
There were a few early attempts to measure the surface plasmon dispersion
of simple metals. As shown in Figure 3.4, Kunz (1966) observed a negative
dispersion for Mg to about 0.3 A-1 and a subsequent shift to higher frequencies.
In contrast, Kloos and Raether (1973) found no dispersion for this system
within experimental accuracy. They attributed this discrepancy to surface
contamination. Recent measurements on Mg (Sprunger et al., 1992) show a
negative initial slope and a minimum at about 0.15 A-1 (see below). Duke
et al. (1975) found a negative initial slope on AI(100), while their results on
AI(1l1) were not conclusive. For polycrystalline films of Al and In, Krane and
Raether (1976) excluded a negative slope at small q (see Figure 3.6).

3.3.1 Surface Excitation Spectra

Figure 3.7(a) shows typical electron energy loss spectra for K (Tsuei, 1990).
Close to the specular scattering direction, a main loss peak corresponding to
the usual surface plasmon is observed. At slightly higher frequencies, a weak
shoulder associated with the multipole surface plasmon can be seen. This
loss feature becomes more prominent at angles away from specular scattering.
Excitation of the K volume plasmon then also becomes observable. In the case
of Mg, the multipole mode is much weaker than in K. Figure 3.7(b) shows the
decomposition of the measured loss spectrum into surface and bulk collective
modes (Sprunger et al., 1992). Since Mg has a fairly high bulk density, the
surface density profile is rather stiff, so that the multipole mode is not expected
to be pronounced.
3.3. SIMPLE METAL SURFACES 59

12.0 Al
9.5

>' 9.0
~
~ ~

~ 1l.0 ~
3 -3 8.5
(b)

(0)
8.0.l.--~-~-~-...--""""
0.1 0.2 0.3 0.4 0.5
10.0
0.1 0.2 0.3 0.4 0.5
q (.&.-1)

Figure 3.6: Dispersion of surface plasmon measured on polycrys-


talline films of Al and In. (Krane, Raether, 1976).

Figure 3.8 illustrates loss spectra for K and Al as calculated within the
TDLDA (Tsuei et al., 1991). The numerical evaluation of these spectra is
based on a real-space formulation of response equation (2.48) and a matrix
inversion to find the induced density n1(z, q,w) (for computational details, see
Section 2.3.5). The loss profiles then follow from the imaginary part of the
surface response function:

f
Img(q,w) = 1m dz eqz n1(z,q,W) . (3.18)

All spectra are dominated by the usual surface plasmon, which is seen to first
shift to lower frequencies as q increases and then, from about q = 0.15 A-lon,
towards higher frequencies. The width of this loss peak is caused by decay into
electron-hole pairs in the surface region and strongly increases with q. [Note
that the continuum of bulk electron-hole excitations begins at q ~ qeh(W). For
w = W., qeh = 0.7 A-1 for Al and 0.4 A-1 for K.]
The additional loss feature at higher frequencies in the case of K corre-
sponds to the multi pole surface plasmon. As discussed in Section 3.1, the
existence of such a mode was predicted by Bennett (1970) in hydrodynamic
calculations for model densities with a finite decay length. The TDLDA cal-
culations also yield such modes for Na and es. Unfortunately, in the case
of high-density metals such as Al, identifying the multipole mode is difficult
(in the theoretical and experimental spectra), since it is intrinsically weaker
60 CHAPTER 3. SURFACE PLASMONS

~ 2
C
::J
(b) Io.1g(OOOI)
.e
~
"',
l:-
;;
I:
~
.E 1

OL---L-__L-~~~__~__~
2 3 4 6 8 10 12
'" (eV) '" (eV)

Figure 3.7: (a) Electron energy loss spectra for K. (Tsuei et at.,
1990). (b) Off-specular loss spectrum for Mg(OOOl) with decom-
position into surface plasmon, multipole mode and bulk plasmon.
(Sprunger et al., 1992).

19

45 (a) K 17

15

--:- 35 q = 0.04 A·'


::::> 13
~
~ 11
£:25
OJ

E
q=O.ll A·'
15

q = 0.45 A·'

.5 .6 .7 .8 .9 1.0 .5 .6 .7 .8 .9 1.0

wlwp wlwp

Figure 3.8: Surface loss function Img(q,w) for K and Al at various


q, calculated within TDLDA. (Tsuei et aI., 1991).
3.3. SIMPLE METAL SURFACES 61

than in the alkali metals. Dobson and Harris (1988) investigated the disper-
sion of the Al multipole mode by performing TDLDA calculations at complex
frequencies and extrapolating them to the real w-axis.

Multipole vs. Monopole Surface Plasmon

The distinction between monopole and dipole distribution in the z-direction


applies rigorously only to true surface eigenmodes, not to externally driven
modes generated in response to an applied field. Driven modes at real frequen-
cies ware always dominated by the induced part of the surface charge density,
as illustrated in Figure 3.9. The 'monopole' and 'dipole' characters of these
distributions are partly lost and obscured by pronounced Friedel oscillations.
Instead, eigenmodes correspond to poles of the surface response function in
the lower half of the complex frequency plane (see Figure 3.10). While the
integrated weight of the fluctuating charge density of the monopole mode is
finite, that of the dipole mode vanishes. It is conceivable that higher-order
multipole modes exist at metal surfaces. These would certainly be even more
heavily damped than the dipolar surface plasmon (their poles would lie even
further from the real w-axis), and therefore difficult to observe.

Threshold Excitation

The calculated surface excitation spectra reveal not only monopole and
multi pole surface plasmons but also a spectral feature for w close to the work
function P. Figure 3.11 shows loss spectra near P for Al where P ~ w•.
The spectral weight is seen to increase appreciably as the excitation frequency
passes through the threshold for emission. In the case of alkali metal sur-
faces, this feature is less well resolved because of the proximity to the surface
plasmon. This so-called threshold excitation is caused by a combination of
many-electron surface screening processes and one-electron matrix elements
(Ishida, Liebsch, 1992). In electron energy loss spectra of clean simple met-
als, it has not yet been clearly identified. On the other hand, there seems to
be evidence of this feature in loss spectra of adsorbed alkali metal layers at
low coverages (see Section 3.8), in photoyield measurements from alkali metal
overlayers (Chapter 4), and in the nonlinear optical response of simple metal
surfaces (Chapter 5). The behavior of the surface loss function near w ~ P
should depend on the shape of the surface potential near the vacuum level. In
particular, the image form should give rise to additional fine structure caused
by image states.
62 CHAPTER 3. SURFACE PLASMONS

2 K ,/ \\ (0)
W, , '.
,,
,,
,

-20 -10 o 10
Z (oo)

Figure 3.9: Induced density at real w. Solid (dashed) curves: real


(imaginary) part. (a) monopole mode at 0.67 Wp; (b) dipolar mode
at 0.86wp; q = 0.15 A-l. (Tsuei et al., 1991).

:l 0
~

3.:
[T
-10 0 10

~
" 0

Wm

-20 -10 0 10
Z(A}

Figure 3.10: Induced density at complex w. (a) real part of


monopole mode at (0.67 - iO.027) wp; (b) imaginary part of dipolar
mode at (0.89 - iO.032) Wp; q = 0.13 A-1. (Tsuei et al., 1991).
3.3. SIMPLE METAL SURFACES 63

1.0

---:- AI
=:
~

3' 0.5
q==0.3,!.-1
~
01 0.2
E

0.0
2 3 Sli 4 5 6

w (eV)

Figure 3.11: Surface loss function 1m g(q,w) for Al in the vicinity


of the threshold for emission, calculated within TDLDA. The arrow
denotes the work function P.

3.3.2 Dispersion of Monopole and Multipole Surface


Plasmons

We turn now to the dispersion of collective surface excitations with parallel


wave vector. The theoretical mode frequencies are derived from the maxima
of the surface loss function Img(q,w) and are compared to the peak positions
observed in experimental electron energy loss spectra. Figure 3.12 shows the
calculated and measured dispersion of surface plasmons for K, Na, Cs, and Al
(Tsuei et at., 1990, 1991). Since the theoretical results do not include effects
due to core polarization, all frequencies are normalized to the measured value
of the monopole surface plasma frequency in the q = 0 limit. To illustrate the
importance of exchange-correlation contributions to the local potential, both
the TDLDA and LDA-based RPA results are plotted.
There is good overall agreement between theory and experiment for the
dispersion of the monopole surface plasmon. (The scatter of the Al data is
rather large; see also Figure 3.6). Moreover, the elusive multipole surface
plasmon, which for two decades had been the subject of theoretical speculations
and vain experimental searches, is clearly seen-except on At. The surface
polarizability of Al is so weak that the multipole mode is hidden in the tail
of the ordinary surface plasmon. Once this main loss feature is suppressed by
64 CHAPTER 3. SURFACE PLASMONS

0.96 , - - - . , - - - - , - - - , 0.96,..----.,----,-----,

r···
(a)K (b)Na
0.92 ~ .. 0.92 RPA ..

··r-T·r· 3.5
0.88 1. .. J. . 5.0
°li"
.... y~
LOA

I
LOA
g- m m
§. ~ ~

>- "
U3 '"
U3
~
4.5 '<
e> '<
m
"
c y S C
m
UJ 0.76
3.0 LlJ S
.<
RPA
RPA .. ' 4.0
......

0.6'b'::.0---:'-:--='-::'::---~0.3

0.96 r----,--.-----.----,
(c) eo

(d) AI
0.92
2.50

0.88
m
~

"
U3 3" 0.84
2.25 ~ §.
s" >-
~ 0.80
"c
W
0.76
o
o 16 0 2.00
o
0.72 rfj"
LOA
0.640 0 - - -0-:-'-',:-------,0.L..
2- - - " 0 . 3 0.1 0.2 0.3 0.4 0.5 0.6 0.7
L..

q (A") q (A")

Figure 3.12: Dispersion of surface collective modes, w.(q) and


wm(q), for K, Na, Cs, and Ai. Squares: experimental data; solid
curves: TDLDA results; dashed curves: LDA-based RPA. The cal-
culated frequencies are normalized to the measured value of the
q = 0 limit of the monopole surface plasmon. The vertical bars
indicate the uncertainty of the calculated frequencies. (Tsuei et al.,
1991).
3.3. SIMPLE METAL SURFACES 65

,
,,
8.0 RPA"
,
M9 ,
o
>-
..!. o
,... 7.5 -' LOA
IT
'-' o
-3 o
exp
7.0

0 .1 .2 .3 .4

q (A-')

Figure 3.13: Surface plasmon dispersion for Mg(OOOl). Squares:


fit through experimental data (Sprunger et al., 1992); solid curve:
TDLDA results for stabilized jellium; dashed curve: LDA-based
RPA. The calculated dispersions are normalized to the measured
frequency at qll = O. (Ishida, Liebsch, 1996).

using photons instead of electrons, the Al multipole mode also appears (see
Chapter 4). As shown in Figure 3.13, results similar to those for Na, K, Cs,
and Al are found for Mg (Sprunger et al., 1992). Off-specular loss spectra for
Mg also reveal a multipole mode (see Figure 3.7). This feature is, however,
rather weak and difficult to resolve from the tail of the monopole plasmon.
All simple metals show a negative dispersion of the monopole surface plas-
mon at small q, confirming Feibelman's microscopic calculations (1974) based
on the LDA ground state and RPA response (see below). The modes pass
through a minimum near 0.15 A-1 and exhibit a positive dispersion at larger
q. The multipole plasmon disperses upwards even at small q. It can be traced
only to about 0.15 A-1 in the experiment as well as theory; at larger q, it
becomes too weak compared to the increasingly wider monopole plasmon.
The experimental and theoretical dispersions shown in Figures 3.12 and
3.13 confirm the qualitative picture obtained by Bennett 20 years earlier (see
Figure 3.4). They demonstrate that a proper treatment of the surface den-
sity profile is crucial for the qll-variation of the monopole surface plasmon and
for the existence of the multipole mode. The overall shape of the dispersions
66 CHAPTER 3. SURFACE PLASMONS

is well described by the Coulomb part of the complex local potential. Ad-
ditional exchange-correlation terms tend to improve the agreement with the
measured plasmon frequencies. Since the effective interaction in the TDLDA is
weaker than the bare Coulomb potential, the LDA frequencies lie below those
calculated within the RPA. This red shift is analogous to the one in the Xe
absorption spectrum shown in Figure 2.9. The K and Cs multipole frequencies
are seen to be in excellent agreement with the TDLDA prediction; the Na data
are too uncertain. Also, the dispersion of the monopole mode of Na, K, and
Cs beyond the minimum is well represented within the TDLDA.
The case of Mg is of particular interest since the quality of the Mg single-
crystal surface was much better than that of Al or the evaporated alkali metals.
Thus, Mg can be regarded as a test system for comparing theory and exper-
iment. The interband transition at 0.7 eV is far below the surface plasma
frequency (see Section 3.4), and the intrinsic broadening of the surface plas-
mon is less than, for example, in the case of AI. The data in Figure 3.13 are
compared to theoretical dispersions for the stabilized jellium model (r. = 2.66).
The calculated frequencies were scaled down by 4 % to match the measured
value of w.(q = 0) = 7.38 eV. The TDLDA results for Mg are seen to be in
nearly perfect agreement with the data up to about 0.2 A-i. At larger q, the
calculations slightly overestimate the frequencies. However, this mismatch is
more than an order of magnitude smaller than the intrinsic width of the Mg
surface plasmon which in this range is 1.5 - 2.0 eV (see Section 3.3.4).
We conclude from this analysis that the adiabatic TDLDA provides a rather
accurate description of dynamical correlations. The fact that the initial neg-
ative slope of the Mg surface plasmon is well reproduced within the TDLDA
raises the question whether the mismatch in the linear region in the case of the
alkali metals diminishes with improved surface preparation techniques. Also,
because of the large intrinsic width of the surface plasmon, uncertainties in
the measured dispersion can be appreciable (see, for example, the case of AI).
An additional point that needs further attention is the non-analytic, cusp-
like behavior of the surface plasmon near qll = O. It implies that, as a result
of the finite detector aperture, the observed dispersions appear flatter than
calculated.

3.3.3 Beyond Standard Jellium and Adiabatic TDLDA

The detailed comparison of surface plasmon data of simple metals with


dynamical response calculations based on the jellium model reveals three slight
but nevertheless persistent discrepancies:
3.3. SIMPLE METAL SURFACES 67

• there is a small overestimate of all frequencies;

• the observed dispersions with qll are flatter than calculated;

• the plasmon line widths are considerably larger than calculated.


On the theoretical side, these differences may be related to three types of
approximations inherent in the surface response calculations discussed so far:
• neglect of lattice effects (core polarization, interband transitions);

• the local density approximation (non-image-like surface potential);

• the adiabatic time-dependent local density approximation.


These approximations affect the dispersions in the following qualitative
ways:
• Core polarization causes not only an overall lowering of the frequencies of
surface collective modes but also an upward distortion of the dispersion curve,
thereby making it slightly flatter. The reason is that conduction electrons spill
out further into the vacuum than tightly bound core states. Thus, at finite q,
the effect of core polarization gradually diminishes. This leads to an increasing
blue shift of the surface plasma frequency (see Section 3.5).
• lnterband transitions may appear as separate spectral features and influ-
ence the frequency and width of surface plasmons. The case of Li (see Section
3.4) shows that this interference between single-particle transitions and collec-
tive modes causes a dramatic flattening of the dispersion. The effect of inter-
band transitions on volume plasmons was studied extensively by Sturm (1982)
and, more recently, by Sturm and Oliveira (1989) and Quong and Eguiluz
(1993). In this context, it is interesting to learn that the long-standing puzzle
of the negative dispersion of the Cs volume plasmon (Vom Felde et al., 1989)
appears to be caused by electronic transitions involving unoccupied d bands
(Aryasetiawan, Karlsson, 1994; Fleszar et al., 1997). Exchange-correlation
effects may also playa role (Taut, 1992). (Note that the minimum of the
Cs volume plasmon dispersion occurs at about 0.5 A-1. In the q-range in-
vestigated in surface electron energy loss measurements, these effects should
therefore be less important.)
• Mechanisms that affect the dispersion of volume plasmons should also
influence the dispersion of surface modes. Nevertheless, there is a fundamental
difference since, at simple metal surfaces, electronic screening processes force
the induced charge density to be located outside the surface, giving a negative
initial slope. This geometric constraint is absent in the bulk where the plas-
mon dispersion is governed by a balance of electron-electron interactions and
68 CHAPTER 3. SURFACE PLASMONS

12.0
AI

11.5
--------------- .-~~/

>-
CD
11.0
'-'
,....
~
3"
2.6

2.4

2.2

0 .1 .2 .3

q ($.-1)

Figure 3.14: Dispersion of Al and Cs surface plasmon. Solid curves:


standard jellium model; dashed curves: stabilized jellium model.
Both within adiabatic TDLDA.

interband transitions. Moreover, bulk effects should become less important at


finite q because of the reduced range of the plasmon field.
• Within the stabilized jellium model, the influence of ionic pseudopoten-
tials is included in an average manner. The effect of the average pseudopo-
tentials on the dispersion of surface modes differs for low- and high-density
materials, as illustrated in Figure 3.14 for Al and Cs. The increased work
function for Al (see Tables 2.1 and 2.2) causes a reduction of the slope at
small q and an upward distortion at larger q. The frequency of the multipole
surface plasmon also exhibits a blue shift because of the less polarizable den-
sity profile. Conversely, the lower work function in the case of Cs leads to an
increased polarizability, i.e., an increased negative slope at small q and lower
frequencies at larger q. The multipole mode is also shifted slightly downward.
Evidently, the simple pseudopotential treatment provided by the stabilization
correction does not explain the flattening of the surface plasmon dispersion
observed in the measurements.
3.3. SIMPLE METAL SURFACES 69

1.0 I I I

0.9 __________ _ -
r.=5
~

30.
:::::<T 0.8 l- ,-
E
1

I I I
0.6
o .1 .2 .3

q (A-')

Figure 3.15: Dispersion of surface plasmons w.(q) and wm(q) for


r. = 5. Solid curves: adiabatic TDLDAj dashed curves: RPAj
dotted curves: dynamical TDLDA. (Pehlke, Liebsch, 1991).

• It is not clear at present which effect the image form of the surface
barrier would have on the dispersion of surface modes. We may be tempted
to associate with the image potential an outward shift of the induced charge
because it decays more gradually than the LDA barrier. On the other hand,
the image barrier is density-independent. This effect seems to cause an inward
shift of the image plane (Zhang et al., 1990j Kiejna, 1993). Since surface
excitations at different q imply varying penetration depths of the fields and
different positions of the fluctuating charge density relative to the jellium edge,
it is plausible that nonlocal corrections to the ground-state potential affect the
q-dependence of surface plasmons .
• Dynamical effects on the induced exchange-correlation potential also mod-
ify the surface plasmon dispersion. Figure 3.15 shows the results of response
calculations using the complex f;r;c( q = 0, w) proposed by Gross and Kohn
(1985) (see Section 2.2.2). The corrections to the static V;c[no(z)] lead to a
systematic upward shift of the plasmon frequencies, so that the new disper-
sions lie between the TDLDA and RPA dispersions. Thus, these dynamical
corrections cannot explain the flattening of the observed q-dependence of the
monopole surface plasmon. We emphasize, however, that the interpolation
formula of Gross and Kohn applies only in the long-wavelength limit of the
70 CHAPTER 3. SURFACE PLASMONS

homogeneous electron gas. For a quantitative consideration of dynamical cor-


relation effects on surface plasmons, it is necessary to include their dependence
on 1Ji1. Besides, the interpolation expression of f",c(w) disregards all effects aris-
ing from quasi-discrete electronic excitations. As pointed out in Section 2.2.2,
such improvements of fxc(w) could well lower the plasmon frequencies rather
than increase them as shown in Figure 3.15.
• Typically, the measured full width at half-maximum is about an order of
magnitude larger than the dispersion of the surface plasmon frequency with
1Ji1. Since this width is much larger than calculated within the TDLDA, other
important broadening mechanisms must exist. We may view these processes
in terms of an imaginary part of a frequency- and momentum-dependent self-
energy. The Kramers-Kronig relations then require the presence of a corre-
sponding real part that may cause a q-dependent shift of the plasma frequency.
One important source of line broadening not included in the jellium model
stems from bulk interband transitions. This contribution is discussed in the
following section.

3.3.4 Lifetime of Surface Plasmons

The width of the measured surface plasmon loss peaks of the metals dis-
cussed above tends to be larger than calculated, in particular, at small q. An
example is shown in Figure 3.16 for Mg. While theory predicts the broad-
ening due to decay into electron-hole pairs in the surface region to vanish in
the long-wavelength limit (see below), the data for all simple metals exhibit
a finite width even at q = O. This additional broadening may be caused by
a variety of factors; For instance, interband transitions are known to be the
main source of the intrinsic width of bulk plasmons at small q (Sturm, 1982).
In addition, some part of the observed width is certainly caused by scattering
from defects, steps, phonons, etc., which are present on all real surfaces.
An important indication of intra- and interband contributions to the line
width of surface plasmons is the small-q limit of the surface loss function. As
discussed in the next section, this limit is given by

e(w) - 1
g(q,w) --+ e(W) + 1 ' (3.19)

where e(w) is the bulk dielectric function. At small q, therefore, the width
is dominated by bulk processes. Figure 3.17 shows these spectra for several
metals. It is evident that bulk absorption processes playa very important role
3.3. SIMPLE METAL SURFACES 71

2.5

2.0 Mg
o
,.....,
> 1.5
,.!,.
:3
<l 1.0

0.5

0.0
o .1 .2 .3

q (.&.-1)

Figure 3.16: Dispersion of surface plasmon width I::!.w s for Mg.


Squares; measured width; dot-dashed line: fit through data
(Sprunger et al., 1992); solid line: TDLDA; dashed line: RPA,
both for stabilized jellium. Dotted line: TDLDA with bulk damp-
ing I = 0.5 eV. (Ishida, Liebsch, 1996).

in surface loss spectra. The relative widths I::!.w./w s are:


0.035 (K), 0.027 (Ag), 0.33 (Li), 0.18 (Hg), 0.16 (Mg), 0.035 (AI).
Surface loss measurements at q = 0, on the other hand, give:
0.1 (K), 0.027 (Ag), 0.35 (Li), 0.16 (Hg), 0.16 (Mg), 0.24 (AI).
In the case of K and AI, the width is considerably larger than expected from
bulk data. For Hg, the width is slightly smaller than predicted from (3.19).
Within a two-band model, Beck and Dasgupta (1975) estimated I::!.w./w s ::::::
0.07 for AI.
As long as these bulk and surface broadening mechanisms are not taken
into account, we must be cautious in drawing conclusions from the comparison
between calculated and measured surface plasmon dispersions, in particular,
since the line widths tend to be larger than the frequency variation. Note also
that (3.19) holds in the dipole-scattering regime. Near-surface inelastic events
that contribute to the impact regime may also affect the measured spectra.
For simple metals, a qualitative description of bulk absorption processes
can be achieved by evaluating the surface response at complex frequencies
w + ii, with I derived from a fit of the measured dielectric function to the
72 CHAPTER 3. SURFACE PLASMONS

+., AI

;:; 4

I.,
.....,
.E
2

O~~~~ __ ~ __-L__- L_ _-L__ J-~~ __ ~~

2 6 8 10 12

w (eV)
Figure 3.17: Surface loss spectra in q = 0 limit, lIn (e -l}/(e + 1),
for K, Ag, Li, Hg, Mg and AI, derived from measured bulk dielectric
functions. The K and Al spectra are scaled down by a factor of 4.

Drude form
w2
e{w}=l- p (3.20)
w(w + i2'Y}
Hence,
w2
Ime{w} ~ 21' ~. {3.21}
w
The full width at half-maximum of the surface plasmon is 6.w. ~ 21'w•. As
pointed out by Mermin (1970), this treatment of damping processes fails to
conserve the local electron density. In the case of a three-dimensional homo-
geneous electron gas, the correct response is given by

(- ) x(cl,w + i1'} A. (- )
(3.22)
nl . /( w +')
q, w = 1 - '1' '1' [1 - X{-q, w +'}/
'1' X{-
q, O}] 'f'sci q, W •

Thus, corrections are of the order of l' /w. An extension to semi-infinite metals
would be interesting but difficult since it must address the liil-dependence of
the damping. At larger q = lliill, induced surface fields have less overlap with
the interior, so that bulk damping processes are gradually switched off. It
is not clear whether approximate treatments of this bulk-surface transition
are realistic or whether the q-dependence of the broadening can be properly
3.3. SIMPLE METAL SURFACES 73

described only via a full microscopic theory in which surface and bulk processes
are treated on an equal footing.
The dotted curve in Figure 3.16 shows the effect of q-independent bulk
damping on the calculated width of the surface plasmon of Mg. The constant "'(
is chosen as 0.5 eV to represent the Drude damping derived from the measured
bulk dielectric function. The broadening is increased by approximately 2"'(
independently of q. Comparison with the dispersion of the measured width
shows this treatment to be qualitatively correct only at small qj at larger q,
the bulk damping should be reduced.
A q-dependent transition from bulk to surface damping processes has re-
cently been verified by Ishida and Liebsch (1996). The surface plasmon dis-
persion of Li(llO) was calculated using a one-dimensional crystalline model.
The potential consists of a smooth Lang-Kohn barrier and a sinusoidal bulk
potential in the direction perpendicular to the surface. The surface loss spec-
tra show that, at small q, the width of the monopole surface plasmon is a
sensitive function of the strength of the ion core potential. The large pene-
tration depth of the plasmon field ensures that surface-induced loss processes
play only a minor role. Towards larger q, however, the relative importance
of bulk- and surface-related damping processes is reversed: Since the plasmon
field becomes more localized near the surface, bulk interband transitions cease
to be effective, and electronic excitations generated by the surface barrier dom-
inate. These results support the idea that the penetration depth rv l/qll of
the plasmon field is the natural quantity that controls the relative weight of
bulk and surface broadening mechanisms.
Since bulk and surface absorption processes exhibit their own variation with
momentum, it is even conceivable that the effective width initially decreases
with q. A decreasing width is seen, for example, in the volume plasmon of Li
(for a discussion, see Sturm, 1982). Similar effects may be the origin of the
slight initial decrease of the observed surface plasmon width of Mg (see Figure
3.16) and Hg (see Section 3.5.2).

3.3.5 Long-Wavelength Limit

A remarkable common feature of the measured and calculated dispersions


shown in Figures 3.12 and 3.13 is the negative slope of w.(q) at small parallel
wave vectors. Because of the wide interest this initial slope attracted over the
past 30 years, and its close connection to nonlocal corrections to the optical
response of metal surfaces (see Chapter 4), we discuss the long-wavelength
limit here in detail.
74 CHAPTER 3. SURFACE PLASMONS

As shown first by Harris and Griffin (1971) and by Flores and Garcfa-
Moliner (1972), in the absence of retardation effects, the momentum dispersion
of the usual surface plasmon at small q is given by

(3.23)

Here, dl.(w) is the centroid of the surface charge induced by a uniform electric
field oriented normal to the surface:

(3.24)

This centroid is the finite-frequency analog of the static image plane position
Zl defined in (2.34). The initial slope has an appealing geometric meaning:
The negative dispersion of the surface plasmon implies that the centroid of
the real part of the fluctuating charge associated with the plasmon is located
outside the surface. (The position of the surface is defined here as the edge of
the neutralizing positive background of the semi-infinite jellium.)
The derivation of the linear surface plasmon dispersion by Flores and
Garcfa-Moliner (1972) is particularly interesting since it can be viewed as
a natural extension of the classical model discussed in Section 3.1. The classi-
cal matching condition is avoided and replaced by a more general integration
of the field components Ez and D:r; over the surface region. This approach is
not limited to a particular electron interaction picture and encompasses quan-
tum mechanical surface effects as well as the nonlocality of the dynamical
response. Moreover, it can readily be extended to the reflection of electromag-
netic waves from surfaces (see Section 4.2.1). In the following, we present the
essential steps of the derivation.

Surface Field Integrals

Let us consider the quasi-static Maxwell equations

V· E(f',w) (3.25)
V· D(f',w) (3.26)

If we assume the fields to lie in the x - Z plane and use the translational
symmetry parallel to the surface, we may write (q == 1!Ji1i = q:r;)

E(f',w) = eiq:r; [E.,(z,q,w),O,Ez(z,q,w)]. (3.27)


3.3. SIMPLE METAL SURFACES 75

The above equations imply

:z Ez(z, q, w) + iqE.,(z, q, w) (3.28)


d .
dz Dz(z, q, w) + zqD.,(z, q, w) o. (3.29)

In addition to the Maxwell equations, we need the constitutive relation

Di(z,q,w) = / dz' C;j(z,z',q,w) Ej(z',q,w) , (3.30)

where C;j is the nonlocal dielectric function. Far from the surface and in the
limit of small q, this relation reduces to

D.(z w) = { c(w) Ei(z,q,w) , z~o , (3.31)


• ,q, E·(z
t "
q w) , z~O .

According to (3.4), the Coulomb potential associated with the surface plas-
mon can be written as

¢>(z) = 2; / dz' e-q1z-z'l nl(z', q,w) , (3.32)

which is an obvious generalization of the classical potential (3.6). (For con-


venience, we momentarily drop the momentum and frequency arguments of
the fields and potentials.) Since we focus on a self-sustained eigenmode, no
incident wave needs to be included. The field components are then given by

-27r / dz' e-q1z-z'l sgn(z-z') nl(z',q,w) , (3.33)

27ri / dz' e-q1z-z'l nl(z',q,w). (3.34)

Integrating (3.4) over the surface region, we find

(3.35)

where Zl ~ 0 and Z2 ~ 0 denote the onsets of the asymptotic regions in the


bulk and vacuum, respectively (see Figure 3.18). In other words, the deviation
of the microscopic surface fields from the macroscopic Fresnel fields is assumed
to be confined to the region [Zl' Z2J. Similarly, (3.29) leads to

(3.36)
76 CHAPTER 3. SURFACE PLASMONS

Ez{z)
_--,-=.::.=:;.-- __________________________ , E/{z)
,.......

z, z

Figure 3.18: Schematic representation of electric field Ez near metal


surface. Solid curve: microscopic field; dashed curve: Fresnel field.
Zl and Z2 denote the onsets of the asymptotic regions in the bulk
and in the vacuum, respectively.

Using E z (Zl,2) = ±q ¢(Zl,2) and E",(Zl,2) = iq ¢(Zl,2), the above equations


may be rewritten in the form

(3.37)

(3.38)

These relations provide the fundamental link between the microscopic variation
of the fields in the surface region and the asymptotic macroscopic fields.
The key quantities are the surface field integrals

h (w) = 1 dz - E
Z2

z,
Ez(z)
( )'
z Zl
III(w) = 1 dz - E
z2

%,
( )'
D",(z)
'" Zl
(3.39)

where the fields are evaluated in the q = 0 limit. In Chapter 4, we show that
these integrals also govern the nonlocal corrections of electromagnetic waves
near metal surfaces. The integrals h and III have the dimension of a length
and depend on the (arbitrary) choice of the asymptotic boundaries Zl and Z2.
It is therefore convenient to express h and III in terms of quantities that are
fully localized in the surface region. Integrating by parts we find
h(w) If(w) (c - 1) d.L(w) , (3.40)
III(w) = ~f(w) + (c - 1) dll(w) , (3.41)
3.3. SIMPLE METAL SURFACES 77

where If = CZ2 - Zl and Ilf = Z2 - CZl are the Fresnel limits. The surface
corrections are determined by the functions

d-L(w) = ! dz z d~Ez(z) / ! dz :zEz(Z) , (3.42)

dll(w) =! dz z :zD",(z) /! dz :zD",(z), (3.43)

which are the so-called d-parameters introduced by Feibelman (1982). Since


the integrands in the expressions for d-L and dll are finite only close to the
surface, the integration range can be extended to ±oo. Using Gauss's law
(3.25) in the long-wavelength limit, the identification of d-L(w) in (3.42) as the
centroid of the induced density is obvious (see (3.24)).
As illustrated in Figure 3.18, the main features of microscopic surface fields
are the smooth, non-step-like variation between bulk and vacuum, and the dis-
placement of the effective location of the surface. This may also be seen by
comparing the classical plasmon potential shown in Figure 3.1 with the mi-
croscopic electrostatic potential cf>est shown in Figure 2.10: The cusp of the
classical potential is smoothed, and the maximum is shifted outward. The
complex function d-L(w) is a measure of these effects. The analogous micro-
scopic behavior of D", is characterized by the complex function dll(w).

Linear Dispersion Relation

The relations (3.37) and (3.38) are compatible under the condition

c+1 -q (h + III)
-q (d-L - d ll ) (1 - c) , (3.44)

where we have neglected higher-than-linear terms in q. This compatibility


relation proves that:

• the surface plasma frequency at q = 0 is quite generally determined by


the condition c(w) = -1, just as in the classical picture (Feibelman,
1968);

• the surface integrals over Ez and D", (h, III or equivalently d-L, d ll ),
that appear in the first-order term rv q, are the key quantities that
determine the deviations from the classical surface plasma frequency.
78 CHAPTER 3. SURFACE PLASMONS

For a Drude metal with c(w) = 1 - W;/W2, we obtain the dispersion relation

w.(q) W. {1- ~ Re [h(w.) + III(w.)] + ... }


= w. {1- ~ Re [dol (w.) - dll(w.)] + ... } (3.45)

This dispersion represents a slight generalization of (3.23). It shows that the


linear coefficient depends not only on the centroid of the induced charge density
but also on the variation of the tangential displacement field. For neutral
jellium surfaces dll(w) =;= zll = 0, where zll is defined in (2.16) (Feibelman,
1976). In this case, (3.45) indeed reduces to (3.23).

Surface Response Function


The preceding approach of integrating Maxwell's equations over the sur-
face region may also be used to evaluate the small-q behavior of the surface
response function g(q, w) defined in (2.65) (Persson, Zaremba, 1984). As
shown in (2.70), g(q, w) represents the asymptotic coefficient of the electro-
static potential induced by an applied potential of the form
A.
'f'ext
(
- eqz .
z,q,w ) -_ -211" (3.46)
q
In the vacuum, the total field has the asymptotic behavior
Ez(z) = -211" (e qz + ge- qZ ) , (3.47)
E.,(z) = -211"i (e qz - ge- qZ ) . (3.48)
Multiplying (3.28) by eqz and integrating over the surface region, we find to
first order in q

-411"g 1 dzeqz -d Ez(z) + iq 1 dzeqz E.,(z)


z2
d z2
(3.49)

1
%1 Z %1

d
+q dz z -Ez(z) + iq(Z2 - zl)E.,(Z2) .
Z2
E z(Z2) - Ez(zt)
Z1 dz
From (3.38) we have

e Ez(Zl) = E z(Z2) + iq 1 dz D.,(z) .


%2

Z1
(3.50)

By substituting (3.48) and keeping terms to first order in q, (3.49) and (3.50)
can be expressed in terms of hand lil or, equivalently, dol and dll . Thus,
e-1 c-1
-411" 9 = -211"(1 + g)-- 411"-- q dIP , (3.51)
e e
3.3. SIMPLE METAL SURFACES 79

where
dIP(w) = c(w) h(w) + dll(w) (3.52)
c(w) +1
is the position of the so-called dynamical image plane (Feibelman, 1980). The
small-q behavior of the surface response function g(q,w) may then be written
as

c(w)-l [ ] (3.53)
g(q,w) ~ c(w) + 1 1 + 2qdIP (w) + ...
c(w) - 1
~ (3.54)
c(w) + 1 - 2q [c(w)d.t(w) + dll(w)]
This result proves that the poles of g(q,w) are determined by a condition
equivalent to the compatibility relation (3.44). In Chapter 4, we show that the
small-q behavior of g(q, w) can also be derived from the reflection amplitude
for p-polarized light in the nonretarded limit.
The significance of the position dIP(w) as a dynamical image plane may
be seen by using the expansion (3.53) to derive an expression for the total
electrostatic potential outside the metal that is correct to first order in q:

¢>(z,q,w) = - 2; eqdIP [eq(Z-dIP) _ : ~~ e-q(Z-dIP)] . (3.55)

Far from the surface, this potential is consistent with classical image theory,
provided the effective surface position is taken to be dIP.

Negative Dispersion of Monopole Surface Plasmon

The first microscopic calculations of d.t(w.) using the LDA surface poten-
tials of Lang and Kohn (1970) were carried out by Feibelman (1974) within
the RPA. These results demonstrated that jellium surfaces in the range of
typical bulk densities give Red.t(w.) > 0; i.e., the initial slope of the sur-
face plasmon dispersion is negative. TDLDA calculations of d.t(w) (Liebsch,
1987; Kempa, Schaich, 1988) confirm this picture, giving larger values of the
real parts because of the more attractive potential ¢>scf(Z,W) (see Table 3.1).
The enhancement of Red.t(w.) for Al (T. = 2) is about 50%, while for CS
(T. = 5.6) the TDLDA value is four times larger, indicating the more impor-
tant role of exchange-correlation terms in low-density metals. The imaginary
part of d.t(w s ), which gives the decay of the surface plasmon into electron-
hole pairs, is also larger in the TDLDA than in the RPA since dynamical
correlations make the density profile more polarizable.
80 CHAPTER 3. SURFACE PLASMONS

Table 3.1: Real and imaginary parts of d.dw.) for jellium surfaces.
Upper two rows: TDLDA based on standard jellium model; next
two rows: TDLDA based on stabilized jellium model; lower two
rows: RPA based on standard jellium model. All distances are
given in ao.

r. 2 3 4 5 5.6
Redl.(w.) 1.2 2.3 3.5 4.1 4.3
Imdl.(w.) 2.5 3.2 3.0 2.0 1.3
Redl.(w.) 0.8 1.9 3.3 4.8 5.5
Imdl.(w.) 1.6 2.3 2.6 2.5 2.0
Redl.(w.) 0.8 1.4 1.7 1.4 1.0
Imdl.(w.) 2.1 2.2 1.6 1.0 0.7

The comparison between the TDLDA values for the standard and stabilized
jellium models shows that the increased work function for high bulk densities
leads to an inward shift of the real part of dl.(w.) and to a reduction of its
imaginary component. The opposite effect occurs for low-density alkali metals
where the stabilized jellium model gives a slight reduction of the work function.
The shifts of Redl.(w.) are consistent with those obtained for the static image
plane positions given in Table 2.3.

Geometric Interpretation of Negative Dispersion

The initial negative dispersion of the normal surface plasmon can be under-
stood in terms of a simple physical picture (Forstmann, Stenschke 1979; Tsuei
et al., 1989): Since the centroid of the plasma surface charge is located outside
the jellium edge and the induced electric potential associated with this charge
decays exponentially towards the interior of the metal, this potential extends
over a region of lower average density as q increases (see Figure 3.19). A lower
density implies a lower plasma frequency. On the other hand, if the centroid
were located inside, the frequency would increase because the potential then
extends over a region of increasingly higher density. Towards larger q, the
shorter wavelength parallel to the surface increases the kinetic energy of the
plasma oscillation and ultimately leads to an upturn of the dispersion. This is
indeed observed and calculated for all simple metals.
3.3. SIMPLE METAL SURFACES 81

cI>---
<tJ
no •••....•.••• _
,
",-
, --
• z
Figure 3.19: Schematic illustration of the relation between the po-
sition of the induced charge centroid, Re dl.(w.), and the dispersion
of the surface plasmon. ¢' denotes the potential at larger q. (Tsuei
et al., 1989).

Positive Dispersion of Multipole Surface Plasmon

The full frequency variation of the function dl.(w) is discussed in Chapter


4. Here we merely point out that, for all simple metal surfaces, 1m dl.(w) has
a peak near about Wm ~ 0.8 wp independently of T. (Feibelman, 1982; Liebsch,
1987; Kempa, Schaich, 1988). This feature is the q = 0 limit of the multipole
surface plasmon seen in the loss spectra in Figure 3.8. The strength of this
peak increases for low-density metals since their surface density profiles are
more diffuse, with a tail extending correspondingly farther into the vacuum.
It therefore becomes progressively easier to polarize these surfaces and generate
this higher-order collective surface oscillation. In AI, the density is rather stiff,
and the multipole mode is too weak compared to the main plasmon line for it
to appear as a separate excitation in the loss function 1m g( q, w).
In contrast to the monopole surface plasmon, the dispersion of the multipole
mode at small q is positive (see Figure 3.12). An estimate of this positive slope
may be derived by representing dl.(w) near Wm as
dl.(w) ~ dm Wm (3.56)
Wm-W

and inserting this form into (3.44) (Kempa, Gerhardts, 1985). In this case,
two solutions are obtained for positive dm : one below w., corresponding to the
negative dispersion of the usual surface plasmon, and a second one above W m ,
corresponding to the multipole surface plasmon. At small q, the dispersion of
the latter is given by

wm(q) ~ Wm (1 + qd m 2 W; 2
Wm -W.
+ ...) . (3.57)
82 CHAPTER 3. SURFACE PLASMONS

Since dm > 0, the dispersion of the multipole mode is positive. This is consis-
tent with Redol(w ~ wm ) ~ OJ Le., in contrast to the ordinary surface plasmon,
the charge centroid of the multipole mode is located inside the metal.

Connection between Monopole and Multipole Surface Plasmons

The above discussion shows that the negative slope of the monopole surface
plasmon is closely related to the frequency and strength of the multipole surface
plasmon. In fact, we can view the latter mode as the fundamental quantity on
which Redol(w.) depends via the Kramers-Kronig relation

R e dol (w. ) -- ~Iooodw wImdol(w)


2 2· (3.58)
11" 0 W - W.

Hence, a strong multipole peak with Wm not far above W. implies a large
negative slope of the monopole plasmon. Any weakening or broadening of the
multipole mode reduces the slope of w.(q).
It is important to keep this connection in mind because of the rapid, pole-
like frequency dependence of dol(w) near Wm . A relatively minor uncertainty in
W m , due to approximations inherent in the TDLDA or other processes omitted
so far, implies a much larger uncertainty in the slope of w.(q). In addition, it
seems plausible that interband transitions, surface imperfections, etc., decrease
rather than enhance the surface polarizability. Thus, they tend to weaken the
strength of the multipole mode, thereby leading to an apparent shift of the
slope of w.(q) in the direction of the RPA results. Ultimately, of course, the
dispersion of both modes reflects the efficiency with which electron-hole pairs
can be created in the surface region.

Linear Dispersion vs. Density Spill-out

Note that the linear slopes of the monopole and multipole surface plas-
mons are genuine dynamical quantities that cannot be represented in terms
of ground-state properties. Their complex nature reflects the finite lifetime
due to decay into electron-hole pairs. For instance, the linear slope of the
monopole plasmon is not related to the static image plane Zl (cf. Tables 2.3
and 3.1). In Chapter 4, we see that Imdol(w) exhibits characteristic spectral
features. Naturally, these features do not contribute to Zl = dol(O) with the
same weight as to the slope Redol(w.) given by (3.58).
The linear coefficient of the surface plasmon also differs from the so-called
spill-out parameter A, which specifies the ground-state density extending be-
yond the jellium edge (see (4.118)). This coefficient defines the distance above
3.3. SIMPLE METAL SURFACES 83

the surface up to which the spill-out density would extend if it had uniform
density n. For r. = 2 ... 5, we find), = 0.44, 0.50, 0.54, 0.56 ao, respectively
(Liebsch, 1985).
Interestingly, the surface f-sum rule (see Section 4.2.5) tells us that ). n
equals the number of electrons (per unit area) taking part in the dynamical
response in the q = 0 limit. It can also be shown that A determines the
linear slope of the moment ratio W31(q) = [m3(q)!ml(q)]1/2 ~ ws(l- q).),
where mi is the ith moment of the surface excitation spectrum (Lipparini,
Pederiva, 1993; Barberan et at., 1993; Kiejna, Peisert, 1994). However, since
moments are frequency-integrated properties, they provide no information on
the dynamical response at specific frequencies. Thus, ). is not representative
of the linear dispersion of the surface plasmon. Similar arguments hold with
respect to the dispersion of the moment ratio W31(q) at arbitrary wave vectors.
The increasing width of the surface plasmon at finite q makes this ratio and
the actual peak position of 1m g( q, w) even more disparate.

3.3.6 Retardation: Surface Plasmon Polaritons

The electronic surface excitations discussed until now correspond to parallel


wave vectors that are large compared to the wavelength of light at optical fre-
quencies. In a typical electron energy loss measurement, the finite aperture of
the detector guarantees integration over a sufficiently large momentum range,
so that the dispersion of light plays a negligible role. However, if we focus
on the region qll ~ w! c, the electronic excitations can no longer be treated
independently of the modes of the electromagnetic radiation.

Classical Dispersion of Surface Plasmon Polaritons

To go beyond the electrostatic limit, it is necessary to consider the full


set of Maxwell equations (see Section 4.2). The two curl equations may be
combined to form the wave equation:
~ 2 ~ w2 _
V'(V'. E) - V' E = "2 D . (3.59)
c
If we assume the electric field to lie in the x - z plane, transversality yields
Ez = -(q,)qz) E", for z > 0 and Ez = (q",!q~) E", for z < 0, where if =
(q""O,qz) and if' = (q""O,-q~) are the wave vectors in the vacuum and bulk,
respectively. According to the wave equation, they are related via if,2 =
84 CHAPTER 3. SURFACE PLASMONS

e(W) if2. If we consider a self-sustained mode in the absence of an incident


field, the spatial variation must be of the form
ei(qa .. -wt) e -"z , z > 0, (3.60)
ei(q... -wt) e"'z , z < 0, (3.61)

where qz = il'i. and q~ = il'i.', i.e.:

w2
I'i.'(w) = q~ - e(W)--z . (3.62)
c
From the continuity of Dz and E .. across the surface, we find

(3.63)

Hence,
I'i.'(w) + I'i.(w) e(W) = 0 . (3.64)
This condition may be expressed as

2 w2 e(W)
(3.65)
q.. = ""iJ e(W) + 1 '

provided that e(W) < -1. Inserting this expression into (3.62), we can easily
prove that (3.64) is equivalent to

(3.66)

The solution of (3.65) gives the dispersion relation of the surface plasmon
in the long-wavelength region. Because of the coupling to the radiation field,
the surface plasmon in this region is also called surface plasmon polariton. In
the special case of a Drude metal with e(w) = 1_w;/w2, we find the dispersion

w.(q.. ) = Jw~ + q~c2 - VW! + q:c 4 • (3.67)

As shown in Figure 3.20, a characteristic feature of this dispersion is the gap


in the frequency range between w. = wp/ v'2 and wp as a result of the avoided
crossing of the light line and the surface plasmon.
For ordinary reflection of light, the radiation dispersion is w+ = q;, + q~. cV
In the limit of grazing incidence, this reduces to w = cq... There is no inter-
action with the monopole surface plasmon, since its frequency lies below the
light line at all values of q... The dispersion of surface plasmons in the long-
wavelength region can, however, be observed using attenuated total reflection
3.3. SIMPLE METAL SURFACES 85

,
,
/''' w=cqx
I ,,' W_
,
, ,, ,,
,
, ,,
,/
, I'

"
/ II /",'"

,,'
,,//

"'"-------------------

Figure 3.20: Dispersion of surface plasmon polariton (lower solid


curve). The radiation dispersions at grazing incidence w = cq""
non-grazing incidence w+ and for attenuated total reflection w_
are indicated by the dashed lines. The upper solid curve represents
the dispersion of light in the metal.

(Figure 3.21; Otto, 1968). In this spectroscopy, light is totally reflected inside
the prism, so that only evanescent tails reach the metal surface. The light
dispersion for this geometry is given by the relation w_ = cJq; - q;.
An indication of retardation effects may be the small-q behavior of the
surface plasmons of Al and In shown in Figure 3.6. The downward shift of
w8 (q) in these transmission inelastic electron scattering data extends, however,
to surprisingly large values of q. The wave vector of light in this frequency
range is less than 0.005 A-1.

Nonlocal Corrections to Surface Plasmon Polariton Dispersion

The above derivation applies to the classical description of fields. A more


general dispersion relation for the surface plasmon polariton can be deduced
by avoiding the classical matching condition and integrating instead over the
surface region. The procedure parallels the one employed earlier for the dis-
persion of the nonretarded surface plasmon. Since this problem is similar to
the reflection of p-polarized light, we defer the derivation to Chapter 4 and
merely quote here the result. From the poles of the reflection coefficient, we
obtain the relation (Sipe, 1980; Apell, 1981; see (4.75)):

(3.68)
86 CHAPTER 3. SURFACE PLASMONS

/ /
Figure 3.21: Illustration of attenuated total reflection geometry.

The second identity follows from using (3.66) in the term specifying the correc-
tions to the classical dispersion. As in the nonretarded limit, these corrections
are given by the difference (dl. - dll ). For not too small values of l/lel and
Ie + 11, the dispersion relation (3.65) is, to lowest order in nonlocal effects,
replaced by
(3.69)

where q", is defined in (3.65). At fixed w, the wave vector of the dispersion
curve is enhanced if Re (dl. - d ll ) > O. Accordingly, the polariton frequency
is shifted below the classical limit (see Figure 3.22). Thus, nonlocal effects in
the retarded and nonretarded region should be consistent. If they lead to a
red or blue shift, they must do so in both momentum regions.
In the vicinity of the multipole plasmon, relation (3.68) may have an ad-
ditional solution. The radiation line intersects this mode even under ordinary
reflection conditions. This is simply another statement of the fact that dl.
appears as the lowest-order nonlocal correction to the Fresnel reflection am-
plitudes (see Chapter 4).

3.3.7 Comparison with Model Potential Predictions

The initial slope of the monopole surface plasmon and the existence of
the multipole excitation are extremely sensitive to the ground-state electronic
properties. Feibelman (1973) calculated the linear coefficient for several smooth
surface potentials and found a remarkable variation with barrier shape.
3.3. SIMPLE METAL SURFACES 87

.9

Re(d)<O
.8

30.
~ .7 - ---
~
3
.6

.5
0 .1 .2 .3 .4

qx (A-I)
Figure 3.22: Schematic dispersion of surface plasmon polariton with
and without retardation. Solid curves: Red = 0; dashed (dotted)
curves: nonlocal dispersions for Red> 0 « 0). (d == d.L - d ll )·
The retardation region is greatly magnified.

Figure 3.23 shows a comparison of surface plasmon dispersions for various


models (Schaich, 1997). The TDLDA results for the Lang-Kohn barrier are
consistent with those shown in Figures 3.12 and 3.13. In the limit of positive
charging, the ionic potential increases quadratically, giving rise to a much
stiffer density profile. This leads to a significant blue shift of the plasmon
dispersion and to the disappearance of the minimum. (Excitations at charged
surfaces are discussed in more detail in Section 3.6.) Even larger blue shifts
are obtained for the infinite-barrier model and the hydrodynamic response of
a single-step density.
In the infinite-barrier model, the centroid of the screening charge at w. is
located inside the jellium edge, giving a positive linear slope (Kempa, Schaich,
1985). Moreover, this model does not support the multipole mode at ordinary
metallic densities, presumably because the density profile is not sufficiently
polarizable (Kempa, Gerhardts, 1985). Curiously, the finite-step potential,
which gives a much softer density tail due to states near the Fermi level,
exhibits no multipole (Schaich, Kempa, 1987) or far too weak a multipole
mode (Gies, Gerhardts, 1987). As pointed out in Section 3.3.1, hydrodynamic
models can give one or more higher-order surface modes as long as the ground-
88 CHAPTER 3. SURFACE PLASMONS

.9

a. .8
3
"-
3

.7

o .1 .2 .3

q / kr

Figure 3.23: Dispersion of surface plasmon for several surface mod-


els (r. = 2.15). Solid curve: TDLDA for neutral Lang-Kohn bar-
rier; dashed curve: TDLDA for strongly charged surface; dotted
curve: RPA for infinite-barrier potential; dash-dotted curve: hy-
drodynamic response for single-step density. (Schaich, 1997).

state density profile is sufficiently smooth (Bennett, 1970; Schwartz, Schaich,


1982).
The sensitivity of surface excitation spectra has been more widely appreci-
ated only during recent years. This explains the long history of contradictory
results obtained for various models of the ground state and dynamical response
(see Table I in Tsuei et al., 1991). It is clearly not sufficient to use any den-
sity profile, even if it is quite similar to those derived from the Lang-Kohn
potentials. As an example, we compare in Figure 3.24 two LDA ground-state
densities with that of the infinite-barrier model. In the latter case, the self-
consistent surface potential is replaced by an infinitely high step. The barrier
is located at a distance Zib = 3.AF/16 above the edge of the positive back-
ground to ensure charge neutrality. The density associated with this potential
is given by the expression

n(z) = n [1 + ;3 (ZcosZ - sinZ)] , (3.70)


3.3. SIMPLE METAL SURFACES 89

1.0
~
'iii
c:
CD

.,
"0

.0
0
~ 0.5
0
CD
Q)

0.0
-.5 0 .5

z (Fermi wavelengths)
Figure 3.24: Ground-state density profiles of jellium surfaces. Solid
curves: LDA results for r. = 5 and 2; dashed curve: infinite-barrier
model. The bar denotes the location of the infinite barrier Zib'

Although the infinite-barrier distribution superficially looks similar to the


LDA densities, it nevertheless gives completely different excitation spectra.
It is not so much the overall smoothness of n(z} that matters most, but the
shape of the low-density tail in the vacuum. This tail arises from electronic
states near the Fermi level that are of crucial importance for surface screening
processes. In the case of the LDA barrier, n(z} extends much farther into
the vacuum. In fact, the infinite-barrier density is even stiffer than that of a
strongly positively charged surface and therefore leads to a greatly underesti-
mated surface polarizability.
The spectral weight of unoccupied electronic states in the infinite-barrier
model differs, of course, entirely from that of the LDA potential. Besides, since
the infinite barrier amounts to an unknown electron-electron interaction pic-
ture, the RPA response treatment implies a severe violation of self-consistency.
The main message from the many model calculations that have been per-
formed in the past is that, even though a certain density profile may be rea-
sonable for the ground state, it can utterly fail for the dynamical response.
Hence, the spectrum of occupied and unoccupied states and the overall consis-
tency between ground-state and response treatments are crucial ingredients of
an adequate representation of surface excitation spectra.
90 CHAPTER 3. SURFACE PLASMONS

(a)
IB

(b)

10 12

w (aV)

Figure 3.25: Electron energy loss spectra for (a) Mg(OOOl) and (b)
Li(llO) at a primary electron energy of 30 eV. The main feature
corresponds to the surface plasmon w., IB denotes the interband
transition and wp the bulk plasmon. (Sprunger et al., 1992).

3.4 Interband Transitions

As we saw in the preceding section, the electronic surface modes of sev-


eral simple metals can be understood fairly well in terms of the semi-infinite
jellium model. This is so because interband transitions induced by the weak
lattice potential are dominated by collective excitations. Nevertheless, certain
influences of the crystal potential exist even in these metals: for instance, the
shift of the surface plasma frequency due to core polarization, or contributions
to the line width due to intra- and interband transitions.
Direct excitation of inter band transitions was recently observed by Sprunger
et al. (1992) in electron energy loss measurements on Mg and Li (see Figure
3.25). In Mg, this transition occurs at 0.7 eV, i.e., far below the surface plasma
frequency. Thus, there is little mutual influence, and, as illustrated in Figure
3.13, the dispersion with parallel wave vector agrees very well with the TDLDA
predictions.
3.4. INTERBAND TRANSITIONS 91

4.4 Li

00
,..., o
>., --... D

"'-'
,..., 4.2
--- ~ ... -e- - 0- - -""IJ'- _ ... _ ... _ ... _ _ _ _ - - - - - i:t --
~ o 0

:3 o

4.0

o .1 .2 .3

q (.&.-1)

Figure 3.26: Surface plasmon dispersion for Li(llO). Squares: ex-


perimental data; dashed curve: fit through data (Sprunger et al.,
1992). Solid curve: TDLDA for stabilized jellium, normalized to
the measured frequency at qll = O. This amounts to a reduction of
the calculated frequencies by 25 %. (Ishida, Liebsch, 1996).

The case of Li is of particular interest since the onset of interband transi-


tions at 3.2 eV lies only slightly below the measured surface plasma frequency,
w.(q = 0) = 4.3 eV. The pseudopotential of Li is much stronger than that of
other simple metals. Accordingly, the dielectric function shows large deviations
from Drude behavior and the surface plasma frequency differs appreciably from
the nominal value wp/..f2. For a homogeneous electron gas with r. = 3.25,
wp = 8.03 eV, while optical data give wp = 6.7 eVj the measured ratio w./wp
is 0.64. It is not surprising therefore that the near-degeneracy of bulk single-
particle transition and collective surface excitation has a strong effect on the
surface plasmon dispersion.
The comparison with the TDLDA calculations in Figure 3.26 shows this
quite convincingly: Aside from the large overall reduction of w.(q) by about
25 %, the measured initial slope is much smaller, and the q-dependence is
considerably flatter than for a jellium surface. Since the surface plasmon lies
just above the onset of interband transitions, the imaginary part of the bulk
dielectric function is larger than for the simple metals discussed earlier (see
Figure 3.17). For this reason, the interband transition also contributes to the
92 CHAPTER 3. SURFACE PLASMONS

unusually large surface plasmon line width at small q (1.5 eV). The Li multi pole
plasmon is not seen since it is hidden in the tail of the monopole mode. The
multipole plasmon can, however, be observed in photoyield spectra of thin Li
overlayers (see Section 4.7.2).
The dispersion of the Li surface plasmon may be qualitatively understood
as follows: At q = 0, the frequency is determined by the condition e(w) = -1;
i.e., the strong lowering relative to the jellium value is caused by bulk interband
transitions. At finite q, the plasmon field is short-ranged and couples less well
to bulk excitations. The frequency is then blue shifted towards the dispersion
for jellium surfaces, giving a reduced slope at small q. In the next section,
we show that a similar mechanism can explain the positive dispersion of the
surface plasmons of Ag.
From bulk electron energy loss spectra, we know that the lattice potential
can give rise to new types of collective modes: the so-called zone boundary
collective states (Foo, Hopfield, 1968; Sturm, Oliveira, 1989). The origin of
these modes is the opening of gaps in the imaginary part of the dielectric
function due to transitions between nearly parallel bands. The real part of
e (ij, w) can then pass through zero in these gaps, causing sharp spectral features
in the bulk loss function -Ime(q,w)-l. For Al and Li, such modes were
observed in the low-q region at about 2 eV and 4 eV, respectively (Petri, Otto,
1975; SchUlke et al., 1986). It remains to be investigated to what extent these
modes influence surface loss spectra near the interband transitions (see also
Section 4.4).
The comparison of the measured dispersions for Mg and Li with the calcu-
lations shown in Figures 3.13 and 3.26 suggests that the TDLDA results for the
jellium model are semi-quantitatively correct as long as lattice effects are weak.
Once interband transitions become important, the surface plasmon dispersion
and width change considerably. Thus, at present the theoretical treatment of
lattice effects appears to be more important than many-body corrections to
the induced exchange-correlation potential.

3.5 Influence of Occupied d Bands

The most striking deviations from the dispersion of surface collective modes
at jellium surfaces occur on Ag as a result of the filled 4d bands. An intermedi-
ate case is Hg whose 5d core levels also have a noticeable effect on the plasmon
dispersion. In this section, we focus on the influence of electronic transitions
whose onset lies above the frequencies of the surface collective modes. This
3.5. INFLUENCE OF OCCUPIED D BANDS 93

applies to Ag and Hg, but not to eu and Au. In the latter cases, the d bands
lie only about 2 eV below E F , so that the surface modes are obscured by
interband excitations.

3.5.1 Ag

A unique property of the Ag energy bands is their nearly-free-electron char-


acter down to almost 4 eV below the Fermi energy. Interband transitions in-
volving occupied d bands or from s-p states near EF to higher-lying unoccupied
s-p bands have a well-defined onset at about 3.8 eV. Nevertheless, as shown
in Figure 3.27, the measured dispersion of the Ag surface plasmon differs in
three fundamental ways from the one observed on simple metals:

• the surface plasma frequency detected in the q = 0 limit (w; = 3.7 eV)
lies far below the value calculated from the s-p electron part of the density
(w s = wp/v12 = 6.5 eV, with wp = 9.2 eV);

• the overall slope of the dispersion is positive, even at small q, in contrast


to the negative initial slope seen on simple metals;

• the magnitude of this positive slope differs for the three low-index faces
and it is anisotropic on the (110) face.

Evidently, the 4d bands form an intimate part of the collective surface excita-
tions of Ag. In principle, we can distinguish two main effects caused by the
presence of d bands:

• the s-d hybridization modifies the single-particle wave functions and en-
ergies, so that the nonlocal susceptibility exhibits bandstructure effects;
these include all microscopic electronic properties near the surface, the
d band shift and narrowing, the presence of surface states, etc.;

• the effective time-varying fields are modified due to the mutual polariza-
tion of sand d electron densities.

A full numerical treatment of these two effects would be computationally very


demanding. For this reason, we take advantage of the fact that the Ag surface
plasmon lies below the region of inter band transitions involving the filled d
states. These transitions contribute only as virtual excitations. Hence, for
a qualitative representation of the Ag surface modes, it appears justified to
94 CHAPTER 3. SURFACE PLASMONS

4.0

.
'>
,....
'-'
3.9

~
:3
3.8

3.7
0 .1 .2 .3

q (l-')

Figure 3.27: Experimental dispersions of Ag surface plasmons for


the three low-index faces. With increasing slope, these dispersions
correspond to: (110) - [110], (111), (110) - [001], and (001).
(Contini, Layet, 1987; Suto et al., 1989; Rocca et at., 1990, 1991,
1992; Lee et al., 1991, 1993). (Rocca, 1995).

neglect bandstructure effects and focus instead on the mutual s-d polarization
in the presence of the time-dependent external electric field (Liebsch, 1993a).

s-d Polarization Model

To account for the s-d polarization, we replace the d states by a polarizable


medium characterized by the local dielectric function cd(W). We obtain this
function by expressing the measured bulk dielectric function as

(3.71)

Here, c.(w) is the Drude function (3.20) representing the s-p electron density,
with wp = 9.2 eV (Ehrenreich, Phillip, 1962; Pines, 1964). The effective bulk
plasma frequency is given by the condition c(w) = 0 which yields the value

w; ~ .::
w
VCd
~ 3.8 eV . (3.72)
3.5. INFLUENCE OF OCCUPIED D BANDS 95

(In the region of the collective bulk and surface modes of Ag, cdC w) is real
and has a value of about 5 - 6.} The condition c(w} = -1 gives the observed
surface plasma frequency in the long-wavelength limit:

w.* ~
~
wp
~~3.7e
,. . . . , V . (3.73)
vI + Cd
Thus, the s-d polarization explains the strong reduction of the surface plasma
frequency below the value obtained without considering the d bands. The in-
teresting questions now are how this s-d polarization influences the q-dispersion
of the Ag surface plasmon and whether the essential features of the measured
Ag loss spectra be understood in terms of this simple physical picture.
Consider first the effect of the bound term CdCw) on the small-q behavior of
the surface plasmon if the bulk dielectric function (3.71) is used in the surface
response function g(q, w). It can easily be shown that (3.54) then implies

w;(q) = ~
1 + Cd
[1- -q-
1+
Red.L(w;) + ...J .
Cd
(3.74)

(We assume dll = 0.) The q = 0 limit is given correctly, and the linear term
is rv q/ (1 + Cd) ~ q/6 instead of rv q/2. If we take the s electron response
to be the same as on simple metal surfaces, i.e., with Re d.L(w.) > 0, the Ag
dispersion is negative. The problem is that d.L(w.) must be recalculated in
the presence of the polarizable medium representing the d band to obtain a
consistent picture in the long-wavelength limit. Such a calculation is discussed
in Chapter 4.
To estimate the influence of the s-d polarization on the plasmon dispersion
at arbitrary q, we describe the 5s electrons'by the nonlocal response function
X(z, z', q, w) of a semi-infinite jellium system with Ts = 3. The bulk dielectric
function of Ag is taken from measurements of Hagemann et al. (1975) which
give w; = 3.63 eV, or from Johnson and Christy (1972). The latter data yield
a slightly higher frequency: w; = 3.68 eV.
The only parameter in our problem is the boundary Zd of the polarizable
medium representing the d electrons (see Figure 3.28). Since this distance
depends on optical excitations involving d bands, it is not a ground-state
quantity and therefore not linked in any simple manner to the position of the
first lattice plane. In fact, as shown by Zaremba and Kohn (1976), in a point-
dipole model of the filled d shells, the dynamic image plane lies approximately
half a lattice spacing above the first plane of nuclei, Le., Zd ~ 0 for all three
low-index faces of the fcc crystal. As a result of the finite size of the d orbitals,
and due to the microscopic electronic structure near the surface, deviations
from this condition can be expected to arise.
96 CHAPTER 3. SURFACE PLASMONS

8
(0) (b) /" \
1.0 .... :....: ... ....,.-; .-: .... --.~ - - I
,, 6 / ",
,, Re Ed I \

,,
. ,,
'.

,
4 - --
0.5 2

'd 0

0.0 -2

-4
-5 0 5 2 3 4 5

z (00) w (eV)

Figure 3.28: (a) Schematic view of dynamical response in s-d po-


larization model. Solid curve: induced s electron density; dotted
curve: ground-state density. The polarizable medium representing
the d bands extends up to Zd. (b) Frequency dependence of Ag
bulk dielectric function c(w) (Hagemann et al., 1975) and of bound
contribution Cd(W). (Liebsch, 1993a).

Presumably, a single Ag plane should have a less sharp onset of interband


transitions than a bulk crystal; the surface can be viewed as an intermediate
case. A less sharp d band onset implies a weaker Cd. To simulate this finite-size
effect, we may choose Zd < 0, which amounts to a reduction of the full Cd to
unity in the region Zd < Z < O. However, because of the simplicity of the
model, we do not select a particular value and merely assume Zd to be located
near the edge of the positive background.

Surface Response Function

To evaluate the electronic excitations in the s-d polarization model, we


again use the surface response function g(q,w) defined in (2.65). The induced
s electron density nl is determined by the TDLDA response equation (2.48).
The total electrostatic potential ifJ = ifJext + ifJest is related to the electric field
via E = V'ifJ. From Gauss's law we have V'. E = -41T (nl + nd), where nl and
nd = V' • Pd are the induced sand d electron densities, respectively. Pd is the
3.5. INFLUENCE OF OCCUPIED D BANDS 97

polarization of the d electron medium. Thus,

(3.75)

The modified Poisson equation for <l>est reads (z 1= Zd)


Cd(Z, w) [<I>:st (z, q, w) - q2<1>est(z, q, w)] = -47rnl (z, q, w) , (3.76)

where
(3.77)

The total electrostatic potential satisfies the condition (zi = Zd ± 0+)

Cd(W) <I>'(z;;,q,w) = <I>'(zt,q,w) . (3.78)

To implement this condition, we express the induced Coulomb potential as

A. (
'f'est Z, q, W
) _
-
27r!d'
- z e
-qlz-z'l nl(z',q,w) 27r
() + - ae
-qIZ-Zdl
. (3.79)
q Cd z',w q
The second term accounts for the d electron screening charge at the boundary
of the polarizable medium. From (3.78), it follows that

a=ud(w) [e qZd + !dz' e-q1za-z'l sgn(zd- z') n~~(;,~~~)] , (3.80)

with
_ Cd- 1
(jd= - - . (3.81)
Cd +1
The total Coulomb potential may now be written as

<I>(z,q,w) = ¢ext(z,q,w) + ¢est(z,q,w) , (3.82)

where

¢ext(z, q, w) _ 27r [eqZ _ Ud(W) e-qlz-zdl eqZd] (3.83)


q

¢est (z, q, w) 27r !dz' nl(z',q,w) [e-qlz-z'l


q Cd(Z',W)
+ e-qlz-zdl Ud(W) e-q1z'-Zdl Sgn(Zd - z')] (3.84)

The new contributions to these potentials arise from the sheet of d electron
screening charge at Zd. The coefficient Ud characterizing these terms is about
2/3 for w ::::; w;. Obviously, the screening effect of the d electron medium
98 CHAPTER 3. SURFACE PLASMONS

Figure 3.29: Surface excitation spectra of Ag obtained within s-d


polarization model. The parallel wave vectors are indicated. The
boundary of the d electron medium is located at Zd = -0.8 A.

causes a significant modification of the external potential and the Coulomb


interaction. Apart from these changes, the structure of the response equation is
the same as for the simple metals discussed in Section 3.3. According to (2.70),
the surface response function can be derived from the asymptotic behavior of
the induced electrostatic potential in the vacuum. From (3.79) we obtain

( ) jdze
gq,w =
qz nl(z,q,W)
()
Cd z,W
+ qZtl
ae, (3.85)

where a is defined in (3.80).

Comparison with Measured Dispersion

The TDLDA excitation spectra obtained within this model are shown in
Figure 3.29. The calculations are carried out at complex frequencies W + i"
with, determined from, = 0.5 (w 3 jw;) Imc(w) (see (3.21» and c(w) from
Johnson and Christy (1972). In the small-q limit, this treatment is consistent
with the measured Ag surface loss function shown in Figure 3.17. These spectra
show that
3.5. INFLUENCE OF OCCUPIED D BANDS 99

6.8 -.---, 4.1

~/
6.6 (b)
4.0
, ,
6.4 ' r =3 '
"""" t ~,,~' 3.9
>
~
6.2
~ 3.8
~ 6.0
:3

r<~~0
3.7
3.8

3.6

-~~ J~L~
3.6
- ___
3.4 3.5
0 .1 .2 .3 0 .1 .2 .3

q (A-') q (.!,-')

Figure 3.30: (a) Dispersion of Ag surface plasmon for Zd = 0 (lower


curves) and of jellium with T. = 3 (upper curves). Solid curves:
RPAj dashed curves: TDLDA. (b) Solid curves: dispersion of Ag
surface plasmon for Zd = 0 and ±0.8 A calculated within RPA.
Dashed and dotted lines: measured dispersions for Ag(lll} and
(100) (see Figure 3.27). The data have been shifted down by 0.06 eV
in order to match the value of w;(q = O} derived from the bulk e:(w}
(Hagemann et al., 1975). Symbols: measured dispersion of the
volume plasmon (Zacharias, Kliewer, 1976). (Liebsch, 1993a) .

• the s-d polarization model correctly describes the downward shift of the
surface plasmon frequency from the un screened value Ws = 6.5 eVj

• the plasmon peak exhibits a blue shift with increasing values of q, in


agreement with the data.

Figure 3.30(a} shows the theoretical dispersion of the Ag surface plasmon


(Liebsch, 1993a). Also plotted are the corresponding curves in the absence of
the s-d interaction, i.e., for a semi-infinite electron gas with T. = 3. In the latter
case, the surface plasmon exhibits the behavior typical for all simple metals,
with a negative initial slope given by Re d.dw.}, the centroid of the screening
charge in the q = 0 limit. The s-d interaction causes not only an overall
lowering of the plasma frequency by about 3 eV but also an upward distortion
of the dispersion. As usual, the LDA frequencies lie slightly below the RPA
dispersion because the more attractive induced potential in the surface region
100 CHAPTER 3. SURFACE PLASMONS

pulls the surface charge somewhat farther into the vacuum.


As shown in Figure 3.30(b), these calculations are in qualitative agreement
with the measured dispersions. For Zd < 0, the agreement is slightly better.
This is reasonable since, as argued above, surface-induced finite-size effects
should lead to a weaker inter band onset and a reduction of the real part of
Cd near the surface. As a result of the s-d polarization, the minimum of the
plasmon dispersion curve is shifted to smaller q. As Zd is moved inside, the
minimum disappears altogether and the dispersion becomes positive even at
small q. This suggests that, within the model considered here, the main part
of the experimentally observed dispersion (roughly the region q ~ 0.05 A-1)
is not related to the linear behavior at small q.
The above results show that the red shift from the bare plasmon frequency
w.(q) to the screened w:(q) depends on q. It is largest at small q because the
plasmon potential decays very slowly into the solid. With increasing q, this
potential decays more rapidly and the s-d interaction is gradually 'switched off'.
This q-dependent reduction of the s-d polarization shifts the mode frequencies
up in the direction of the unscreened frequencies and therefore leads to an
upward skewing of the surface plasmon dispersion.
To obtain more physical insight into this behavior, recall the origin of the
negative dispersion of the surface plasmon for simple metals: Since screening
at metal surfaces is rather efficient, the fluctuating density is located in the
tails of the ground-state density. At finite q, the plasmon field samples a lower
average density and the frequency initially decreases. In the case of Ag, the
situation is more complicated since, in the long-wavelength limit, the frequency
w; = wp ! y'1 + Cd is determined not only by the density but also by Cd. Hence,
at finite q, there is a competition between two opposite trends: the usual red
shift as on simple metals, and a blue shift due to the apparent reduction of
Cd· The balance between these mechanisms depends on the location of the
boundary Zd of the polarizable medium. The TDLDA results indicate that,
for reasonable choices of Zd, the blue shift dominates.
The apparent reduction of Cd is a consequence of the spill-out of the 5s elec-
trons (see Figure 3.31). The 'tail' electrons involved in the plasma oscillation
are not as fully exposed to the d-screening effect as s electrons in the bulk. In
fact, the exponential decay of the plasmon field ensures a transition from full
s-d screening in the q = 0 limit to partial s-d screening at finite q. In Section
3.9, we see that the 5s electron spill-out can also explain the blue shift of the
Ag Mie plasmon with decreasing particle size.
Notice that the s-d polarization model is not entirely consistent since we did
not include the effect of the polarizable d electron medium on the ground-state
s electron density, nor on the induced exchange-correlation potential. Consis-
3.5. INFLUENCE OF OCCUPIED D BANDS 101

no

/~j
/0--0/<:-0
//0.:
z

Figure 3.31: Illustration of reduced s-d polarization at Ag surface.


Inside the bulk, the Coulomb interaction between two s electrons is
fully screened by the medium representing the 4d states. Because of
the s electron spill-out at the surface, the s-d screening is reduced.

tent TDLDA calculations, which treat electron-electron interactions in the


presence and absence of the applied field on the same footing, have recently
been carried out by Liebsch and Schaich (1995). While these refinements
change details of the Ag surface plasmon dispersion, the general picture de-
w;
scribed above is confirmed: At finite q, the variation of (q) is dominated by a
reduction of s-d polarization, giving a blue shift towards the unscreened w.(q).
Thus, the overall positive dispersion of the Ag surface plasmon seems to be
a consequence of electrodynamic effects, and not of the microscopic electronic
structure near the surface.
The preceding results have important consequences for the nonlocal optical
response of Ag surfaces. For instance, as illustrated in Figure 3.22, the positive
slope of the surface plasmon at small q implies that the surface polariton
frequency in the retardation region should lie above rather than below the
dispersion expected from Fresnel theory. The small-q behavior is discussed in
Section 4.5.

Crystal Face Dependence

Apart from the overall lowering of the surface plasma frequency and the
positive dispersion, the s-d interaction also causes a crystal face dependence
102 CHAPTER 3. SURFACE PLASMONS

and an anisotropy on Ag(110) (see Figure 3.27). These observations cannot


easily be addressed within the s-d polarization model since it is based on an
abruptly terminated dielectric medium (Liebsch, 1994). If the surface weakens
the interband onset, the reduction of Cd should be more prominent on the less
dense (100) face than on Ag(l11). Thus, Zd should be more negative on (100),
giving rise to a stronger blue shift of the frequencies than on (111). This trend
is indeed observed (see Figure 3.27). But according to this argument, Zd should
be even more negative on Ag(110) whose lattice planes are the least dense. In
conflict with the data, this would imply even higher frequencies.
Obviously, anyone-parameter model cannot explain the full complexity
of the experimental dispersions. Besides, the anisotropy seen on Ag(110) is
evidence of the importance of d ll . It is plausible that this parameter differs for
the three low-index faces. A complete understanding of the face dependence
of the dispersion can therefore not be achieved solely in terms of dl..
To introduce crystallinity effects, the smooth polarizable medium could be
replaced by an fcc lattice of polarizable 4d shells. This improvement would
include local field effects due to the periodic modulation of the induced density
caused by the s-d interaction. A similar model was employed by Sturm et al.
(1990) to determine the influence of core polarization on the volume plasma
frequency of a variety of metals.
A polarizable lattice model was proposed by Tarriba and Mochan (1992)
with the aim of analyzing the face dependence of the Ag surface plasmon.
These authors considered a semi-infinite lattice of point dipoles immersed in
a local Drude system with cavities at the lattice points ('swiss cheese' model).
The polarizability of these cavities accounts for the occupied 4d levels and the
s-p states close to the nuclei. The interstitial region represents the remaining s-
p electron density which is abruptly terminated at the surface. The single-site
polarizability is obtained by fitting the measured Ag bulk dielectric function
c(w) to the Clausius-Mossotti relation
3 C - Cs
a=-----. (3.86)
47rn C + 2c s
Here, Cs is the Drude dielectric function of the s electron density. The in-
duced dipole moments at the lattice sites are calculated from a self-consistent
response equation. The solution yields surface conductivities closely related
to dl.(w) and dll(w). Finally, the surface response function is derived from the
p-polarized reflection coefficient in the nonretarded limit.
This model has recently been extended to finite q by Lopez and Mochan
(1997). As illustrated in Figure 3.32, the surface plasmon peaks are blue
shifted with respect to w:(q = 0) ~ 3.7 eV. Moreover, the spectra reveal a
3.5. INFLUENCE OF OCCUPIED D BANDS 103

........
~ 4
~

3:
0-
o 2

3.70 3.75 3.80 3.85

w (eV)

Figure 3.32: Surface excitation spectra for low-index faces of Ag at


q = 0.2 A-I. The point dipoles are immersed in a Drude electron
gas terminated 0.7 A above the nominal jellium edge. Solid, dashed,
dotted, dash-dotted curves: (100), (111), (110)-[110], (110)-[001]
face. (Lopez, Mochan, 1997).

face dependence and an anisotropy on Ag(110). The model, however, is still


too crude to be able to reproduce the exp~rimental results. It would be very
interesting to replace the Drude description of the s electron response by a
nonlocal q-dependent response as for semi-infinite jellium. This generalization
would be equivalent to the s-d polarization model discussed above, provided
the d electron medium is replaced by a lattice of polarizable dipoles.
Feibelman (1993, 1994) investigated the face dependence in terms of the
influence of s-d matrix elements on dJ.. by adding an extra potential well to the
Lang-Kohn barrier. Such a potential perturbation gives an increased electron
density in the surface region. Since high-density metals have a less polarizable
density profile, the magnitude of dJ.. is reduced. Accordingly, the slope of
the surface plasmon remains negative, albeit with a smaller coefficient than
in the unperturbed jellium model. In the q = 0 limit, this model yields the
unscreened plasma frequency (6.5 eV in the case of Ag) because the long-range
s-d polarization is absent. Nevertheless, s-d matrix elements should become
stronger with decreasing separation between fluctuating s charge and d states.
This mechanism predicts successively less negative slopes for the (111), (100),
104 CHAPTER 3. SURFACE PLASMONS

8.0 1.5

(b)
a
,,, a a
a
" 00 a
a
7.5 1.0
>- '" - __I:L ____ E____
>-
~ ~
.-.. .-..
IT

1
IT

'1
7.0 0.5
---

6.5 0.0
a .1 .2 .3 a .1 .2 .3

q (A-I) q (A-I)

Figure 3.33: Dispersion of (a) frequency and (b) width of Hg surface


plasmon. Squares: experimental data. Upper curves in (a) and
lower curves in (b): semi-infinite jellium. Lower curves in (a) and
upper curves in (b): jellium in the presence of polarizable medium
representing the 5d core levels and with q-dependent bulk damping.
Solid curves: TDLDAj dashed curves: RPA. (Kim et al., 1995).

and (110) faces of Ag. The data show a more positive dispersion on (100) than
on (111), but the two dispersions on (110) lie below those on (100) or (111)
(see Figure 3.27). A microscopic treatment should, of course, include the s-d
polarization as well as matrix element effects.

3.5.2 Hg

Recent electron energy loss measurements on Hg by Kim et al. (1995)


indicate that the 5d core levels have an appreciable influence on the dispersion
of the surface plasmon. These effects are, however, less pronounced than on
Ag. Owing to the polarization of the 5d states (8 eV below E F ), the surface
plasma frequency at q = 0 lies about 1 eV below the value expected for the 6s
electron density alone. Thus, w; = wp/v'1 + ed = 6.8 eV, where wp = 10.9 eV
is the volume plasma frequency of an electron gas at T. = 2.65. The bound
contribution to the dielectric function is ed ;:::: 1.6.
3.5. INFLUENCE OF OCCUPIED D BANDS 105

The Hg loss measurements show that the dispersion of the surface plasmon
is negative at small q, but with a much smaller slope than in the case of
simple metals (see Figure 3.33). To analyze these data, TDLDA calculations
were performed using the s-d polarization model discussed above. The results
show that the s-d screening causes not only an overall red shift of the surface
plasma frequency by about 1 eV, but also a distortion of the dispersion: With
increasing q, the s-d polarization is reduced because of the spill-out of the 6s
electrons. This mechanism leads to an upward skewing of the dispersion curve.
The minimum is accordingly shifted to smaller q and occurs at about 0.1 A-1.
A major difference between Hg and Ag is the presence of strong bulk inter-
band transitions within the 6s bands (see Figure 3.17). In the q = 0 limit, the
measured width of the Hg surface plasmon is about 1 eV, in contrast to 0.1 eV
in the case of Ag. To include these broadening mechanisms in a qualitative
manner, the surface response function was evaluated at complex frequencies
w + iry. The imaginary part, was derived from the measured bulk dielectric
function using (3.21). At finite q, however, the role of bulk absorption pro-

well. This effect was included in an ad-hoc manner by writing ,( q) = ,e-


cesses should decrease since the induced electric field penetrates the metal less

with a = 3 A. As shown in Figure 3.33(b), this procedure gives a surface plas-


qa

mon width of about 1 eV independently of q, in approximate agreement with


the experimental data.

3.5.3 Pd

Electronic excitations of Pd are more complicated than those of Ag and


Hg because of strong interband transitions involving partially occupied 4d
bands. These span a large frequency range and cannot be easily distin-
guished from transitions of predominantly s-p character. Nevertheless, the
bulk and surface loss functions in the long-wavelength limit, i.e., -lm1/c(w)
and -1m 1/[c(w )+1], show broad maxima at about 8 eV and 7 eV, respectively.
These spectral features could be associated with quasi-collective excitations.
Interestingly, Rocca et al. (1995) observed a loss feature with a negative
dispersion, w.(q) = w.(1 - aq/2), where w. ~ 7.4 eV and a = 2 A. Obvi-
ously, this kind of behavior cannot be interpreted in terms of collective surface
excitations of Drude-like metals, where the d electrons merely provide a polar-
izable background. Presumably, a detailed analysis should include an explicit
description of s-p and d electron transitions, with bulk and surface treated on
an equal footing. A theoretical scheme capable of handling these aspects, with
a full account of surface screening processes, is not yet available.
106 CHAPTER 3. SURFACE PLASMONS

r.=3
~ 1.0
·iii
I:
CD
."

·c
(J

.g
(J
0.5
CD
a;

0.0
-5 o 5
Z (00)
Figure 3.34: Comparison of electronic density profiles for neutral
and strongly positively charged jellium surface (Ts = 3). zil = 0
for both profiles. The edge of the positive background for (J > 0 is
indicated by the vertical dashed line. (Gies, Gerhardts, 1985).

3.6 Charged Surfaces


3.6.1 Simple Metals

In Section 3.3, we emphasized that the dispersion of surface plasmons is


sensitive to the electronic properties in the surface region. An example illus-
trating this point rather well is the effect of static surface charges induced at
metal--electrolyte interfaces. By tuning the electric potential, it is possible to
modify the density profile in a systematic manner.
Figure 3.34 compares the density distribution for a neutral jellium surface
with the profile obtained for strong positive surface charging. To compensate
for the overall inward shift of the density due to the presence of the static field,
the origin is chosen so that the effective jellium edge zil (2.16) coincides with
Z = O. Positive charging is seen to cause a stiffening of the density profile, so
that the tail extends less into the vacuum than in the neutral case. This is
consistent with the nearly parabolic shape of the surface barrier, which leads
to a much more rapid decay of electronic states.
Schaich (1994) showed that, in the case of parabolic one-electron potentials,
dl.(w) == o. (The origin corresponds to zil = 0.) This means that, in the limit of
positive charging, single-particle surface excitations are forbidden. According
3.6. CHARGED SURFACES 107

to the long-wavelength behavior discussed in Section 3.3.5, this result implies


that the monopole plasmon begins with a vanishing slope, to linear order
in q it is undamped, and the multipole plasmon is suppressed. Hence, the
modification of the density profile displayed in Figure 3.34, which at first glance
does not seem important, has a surprisingly strong qualitative influence on
the surface collective modes. In fact, as shown by Schaich (1997), the plasmon
dispersion is nearly perfectly quadratic up to large q (see Figure 3.23). Plasmon
decay into electron-hole pairs becomes possible beyond the linear q region, but
it is much weaker than in the case of the neutral surface.
In this context, work by Zaremba and Tso (1994) is also of interest. These
authors calculated collective modes of strongly charged slabs (parabolic quan-
tum wells) within the Thomas-Fermi-von Weizsacker (TFW) approach. For
thick slabs, a surface plasmon with zero initial slope is found. Also, there is
no evidence of a multipole surface plasmon. These results are in agreement
with the prediction by Schaich (1994). However, because of the hydrodynamic
response description in the TFW theory, single-particle excitations are not
included. It is therefore not clear from this work how important quantum
mechanical corrections to the dispersion are outside the linear q region.

3.6.2 Ag

To develop a qualitative picture of the effect of static fields on the surface


modes of Ag, let us use again the s-d polarization model discussed in the
preceding section. For simplicity, we assume that charging alters the 5s electron
distribution but leaves the dielectric medium representing the 4d electrons
unchanged. As pointed out above, positive charging modifies the s electron
density in two ways: it leads (i) to a steepening and (ii) to an inward shift
of the profile. The steeper density is less polarizable, so that higher-density
regions must be involved in the surface screening processes. This effect causes
a blue shift of the plasma frequency. The inward shift of the s electron density
profile, on the other hand, pushes the fluctuating surface charge towards the
4d states and thereby enhances the s-d polarization interaction. Thus, the
second effect leads to a red shift of the Ag surface plasma frequency.
As shown in Figure 3.35, TDLDA calculations for charged Ag surfaces
indicate that the red shift due to the inward shift of the 5s electron density
profile is much stronger than the blue shift due to the increased stiffening of
the density. In fact, whereas the blue shift quickly saturates with increasing
positive charging, the red shift continues to grow much longer. This trend
108 CHAPTER 3. SURFACE PLASMONS

3.9 -~---r I I

3.8

:;-
~
3.7
'""'
S
:3 -, -~ ~.~ _-~--------- clj Ag

-----~
3.6

3.5
o .1 .2

q (A-I)

Figure 3.35: Dispersion of Ag surface plasmon as calculated within


TDLDA. Upper solid curve: clean, neutral Ag; dashed curve:
charged Ag (u = 0.02 a.u.); dotted curve: neutral Ag with dielec-
tric overlayer (co = 2 corresponding to thin 01- layer); dot-dashed
curve: u = 0.02 a.u. and co = 2. Lower solid curve: neutral Ag
with dielectric overlayer (co = 4 corresponding to thin AgOllayer).
(Garcia-Gonzalez, Liebsch, 1997).

stops only when the fluctuating plasmon charge is shifted deep inside the range
of the d electron medium. Thus, the effect of positive charging on the Ag
surface plasmon dispersion is opposite to that found for the simple metals (see
Figure 3.23). We also show the influence due to a thin dielectric overlayer
corresponding to a plane of 01- atoms (see next section). While the q = 0
mode remains unchanged, the frequency at finite q is red shifted because of the
extra screening in the surface region. The combined red shift due to positive
charging and adsorbate-induced dielectric screening can be large enough to
give a minimum in the dispersion curve.
These results are in qualitative agreement with electron energy loss mea-
surements for Cl layers adsorbed on Ag(l11) (Kim et al., 1997b). With in-
creasing coverage, the spectra reveal a dramatic red shift of the Ag surface
plasmon at finite qll' At one monolayer, the dispersion exhibits a minimum,
just like for the simple metals. As a result of the 01 adsorption, there is an
appreciable charge transfer to the overlayer, implying that the Ag 5s density
3.7. INFLUENCE OF DIELECTRIC MEDIUM 109

profile is pushed inwards (Kramar et al., 1995). Hence, the fluctuating plas-
mon density overlaps more with the d states than on the neutral surface. The
effect of Cl-induced positive charging and dielectric screening seems to reduce
the plasma frequency to such an extent that the dispersion exhibits a mini-
mum. Since the data were taken at room temperature, dissolution of Cl into
the substrate might also occur. Additional screening effects could therefore
arise due to the formation of an AgCI surface alloy.

3.7 Influence of Dielectric Medium

The approach discussed in Section 3.5 to describe the effect of occupied d


states on the surface plasmon dispersion can easily be generalized to include
the presence of a dielectric medium in contact with the metal. This case is of
interest since it provides insight into electronic excitations at metal-electrolyte
interfaces. In principle, the metal may be positively or negatively charged,
but we consider here only neutral surfaces. We denote the boundary of the
dielectric medium as Zo and its dielectric constant as co.
In the case of a Drude metal in contact with an external medium, the
Poisson equation is given by (3.76) and cd(Z,W) is now replaced by

Z < Zo , (3.87)
Zo S z.
In the more general situation, the metal itself is also immersed in a dielectric
medium (representing core levels or filled d bands). We then have

Z S Zd ,
Zd<Z<ZO, (3.88)
Zo S Z •

We assume here that the internal and external media are separated by a gap
(Zd < Z < zo) of dielectric constant Ca' (In the example discussed below we set
Ca = 1, but in the following section we consider a case with Ca #1.)

Surface Response Function

The dynamical response of this system can be derived in a similar manner


as for clean Ag. The boundary conditions that must be satisfied by the total
110 CHAPTER 3. SURFACE PLASMONS

potential at the planes Z = Zd and Z = Zo are given by


Cd(W) ¢'(Zi,q,w) Ca ¢'(zt,q,w) , (3.89)
Ca ¢'(zQ,q,w) co ¢'(zt, q, w) . (3.90)
In analogy with (3.79), we write the induced Coulomb potential as

A-.
'l'est
( ) _ 211"jd ' -qlz-z'l n1(z',q,w)
z,q,w - - Ze (' ) + -211" 0
~ai e
-qIZ-Zil
, (3.91)
q Cd Z ,w q i=1
with Z1 == Zd and Z2 == zoo The terms involving the coefficients ai account for
screening charges at the boundaries of the polarizable media. According to
conditions (3.89) and (3.90), these coefficients are given by
2
ai = L aij (e qZj + bj ) , (3.92)
ij=1

where

jd z' e -qlzi-z'l sgn (Zi - Z


') n1(z',q,w)
( ) (3.93)
Cd Z',w
(3.94)
1 + a1a2e~1 '
Cd - lOa lOa - Co
a2= - - - , (3.95)
Cd + lOa ' lOa + cO
e- q (Z2- Z ,) •
(3.96)
The total Coulomb potential can now be written as in (3.82), with

¢ext(Z,q,w) - 211" [eqZ _


q
t.
ij=1
e-q1z-z;I aij e qZj ] (3.97)

¢est(Z, q, W) 211"jdz' n1(z',q,w) [e-qIZ-Z'1


q cd(Z',W)
L
2
+ e-qlz-zil aij e-q1z'-zjl sgn(Zj - z')]. (3.98)
ij=1

With these changes, the structure of the response equation is the same as for
simple metals. In the limit lOa = co = 1, we recover the case of clean Ag. Since
the surface response function is determined by the asymptotic behavior of the
induced electrostatic potential, we obtain from (3.91)

g(q,w) = jdz e qz n1(z,q,W)


ciz,w)
+ t.
i=1
ai e qzi . (3.99)
3.7. INFLUENCE OF DIELECTRIC MEDIUM 111

3.7.1 Mg - MgO

Figure 3.36 shows the surface plasmon dispersion for Mg in contact with an
oxide layer characterized by a dielectric constant co = 3.5 and located in the
region Zo ;::: 1.4 A. The ground state is derived from the stabilized jellium model
and the response is treated within the TDLDA. In the long-wavelength limit,
the surface plasma frequency is given by w; = wpj~. The comparison
with the dispersion at a free Mg surface proves that the oxide layer not only
changes the overall frequency of the surface plasmon but also the dispersion.
The initial slope is positive since, as in the case of Ag, the screening action
of the polarizable medium diminishes with increasing q. Consequently, the
plasmon frequency shifts up in the direction of the unscreened Mg surface
plasmon. The detailed form of the dispersion depends on the location of the
boundary Zoo

3.7.2 Ag - Ar

Figure 3.37 illustrates the dispersion of the surface plasmon of Ag in contact


with Ar. The latter is represented via a dielectric medium, with co = 1.75
in the region Z ;::: 1.6 A. The Ag substrate is described in terms of the two-
component s-d electron model discussed in Section 3.5. We choose Zd = -0.8 A
since this boundary of the d electron medium is in approximate agreement with
the clean Ag surface plasmon dispersion. The Ar overlayer is seen to cause
a q-dependent lowering of the Ag surface plasma frequency. In the q = 0
limit, w; ~ wpj y'co + Cd ~ 3.54 eV instead of w; = 3.68 eV for clean Ag.
At finite q, this reduction is less effective because of the short range of the
plasmon field. The effect of the Ar layer is gradually switched off, giving an
increasing blue shift in the direction of the un screened plasma frequency. This
behavior is consistent with the frequency shift of the Mie plasmon observed
for Ag particles embedded in an Ar matrix (see Section 3.9.3).
In the presence of an external dielectric medium, the long-wavelength be-
havior of the surface plasmon is determined by a condition that represents a
generalization of (3.44):
(3.100)
If the metal dielectric function consists of a Drude term plus a bound contri-
bution, i.e., c(w) = Cd - w;jw 2 , we find the dispersion relation

w;(q) = ~ [1- _c_o- qRe (dJ.. - d ll ) + ...J . (3.101)


cO+cd CO+C:d
112 CHAPTER 3. SURFACE PLASMONS

8
Mg

>-
-!.
7

~
3 6

Mg/MgO
5

0 .1 .2 .3

q (..\-1)
Figure 3.36: Surface plasmon dispersion of Mg in contact with a
thick oxide layer (co = 3.5) (solid curve). Dashed curve: dispersion
for clean Mg.

3.9

3.8 Ag

>-
-!.
,...., 3.7
~
--
3
Ag/Ar
3.6

3.5
0 .1 .2

q (1.- 1 )

Figure 3.37: Surface plasmon dispersion of Ag in contact with a


thick Ar layer (co = 1.75) (solid curve). Dashed curve: dispersion
for clean Ag.
3.8. OVERLA YERS 113

The coefficient co/ (co + Cd) of the linear term is seen to depend on the relative
size of co and Cd. For co > Cd, the coefficient is larger than 0.5, while, for
co < cd, it is diminished. If the internal and external dielectric constants are
equal, the value 0.5 is regained. The results in Figure 3.36 and 3.37 show,
however, that the sign and magnitude of Re (dJ.. - dll) are more important
than this coefficient.
The dielectric medium considered so far is taken to be semi-infinite. For
this reason, the surface plasma frequency is reduced even in the limit qll = O.
For metals in contact with a dielectric layer of finite thickness, on the other
hand, ws(q = 0) coincides with the value for the clean surface. At finite q, the
plasma frequency then shifts downwards so that the dispersion might exhibit
a minimum. An example is shown in Figure 3.35. Because of the exponential
decay of the plasmon field at finite q, the detailed form of the dispersion
depends on the thickness and position of the overlayer.
We also point out that the effects illustrated in Figures 3.36 and 3.37 are of a
purely electrodynamic nature since we assume the equilibrium density profile
to be unaffected by the presence of the neighboring dielectric medium. For
weakly adsorbed systems, this assumption is reasonable. The only important
parameter then is the boundary of the dielectric medium. For more strongly
adsorbed overlayers, the modification of the surface density profile due to the
formation of chemical bonds must be taken into account. Examples of this
type are the topic of the following section.

3.8 Overlayers

Electronic excitations in thin adsorbed overlayers have attracted attention


for more than 25 years. These systems represent quasi-two-dimensional elec-
tron gases whose average density and thickness can be varied over a wide range.
The adsorption of alkali metals is of particular interest because of phenomena
such as large work function changes, surface reconstruction, catalytic promo-
tion and metal-insulator transitions (MacRae et at., 1969; Andersson, Jostell,
1974, 1975; Cousty et at., 1985; Aruga, Murata, 1989; Bonzel, 1987; Bonzel et
at., 1989; Heskett et at., 1988).
At coverages of about one or two monolayers of adsorbed alkali metals,
the adatom-induced electronic states have a pronounced delocalized charac-
ter along the surface. The excitation spectra should therefore be dominated
by collective modes with a characteristic parallel momentum dispersion. At
114 CHAPTER 3. SURFACE PLASMONS

submonolayer coverages, the direct wavefunction overlap between adatoms dis-


appears and quasi-atomic electronic transitions become important.

3.S.1 Local Optics Picture and Hydrodynamic Models

In the classical model, it is assumed that the electronic response of the


semi-infinite substrate and the overlayer are given by their respective local
bulk dielectric functions (Stern, Ferrel, 1960; Gadzuk, 1970). We denote these
functions by c(w) = 1- O~/W2 for the bulk (z S; 0) and by Ca(w) = 1-w;/w2
for the adsorbate (0 S; z S; a). The electrostatic potential may be written in
the form
Aeqz z S; 0 ,
¢(z, q, w) = { B eqz ~ C e- qz , O<zS;a, (3.102)
_eq(z-a) + 9 e-q(z-a) , a < z.

The coefficients can be eliminated by using the continuity of Ex and Dz at the


substrate-adsorbate and adsorbate-vacuum interfaces. The amplitude of the
reflected wave is given by
~(q,w) - 1
g(q,w) = ~(q,w) +1 ' (3.103)

where the effective dielectric function of the adsorbate-substrate system is


defined as
_(
10 q, W
)
=
( ) 10
lOa W
+ lOa + (10 - ca)e- 2qa
• (3.104)
10 + lOa - (10 - ca)e- 2qa

In the limit qa « I, ~(q,w) reduces to c(w), i.e., the response becomes in-
dependent of the presence of the overlayer. The substrate surface plasmon is
then the only important excitation. Conversely, in the opposite limit qa ~ I,
~(q, w) reduces to lOa (w). The electric field then decays so rapidly across the
overlayer that the response becomes independent of the substrate properties
and the overlayer surface plasmon is the dominant mode.

Dispersion of Overlayer Modes

Figure 3.38(a) illustrates the collective surface modes for Drude systems
corresponding to a Na overlayer on AI. At q = 0, there is only one mode at
the surface plasma frequency of the substrate, Os = Op/ v'2. With increasing
q, this mode shifts upwards and, near q ::::; l/a, it approaches the interface
3.8. OVERLAYERS 115

1.0
(a) c=I/8
WI

0.8 WI
wp
' >"
,,
--
n. n. ' ",

c
. 0.6
wp , ,
--- ---
'-
3
w.
w.

0.4
wp
(b) c=I/2 (c) c=0.64
w.
0.2
q q q

Figure 3.38: Dispersion of collective adsorbate-substrate modes in


local optics picture (schematic). The ratio of adsorbate and sub-
strate densities is (a) 1/8; (b) 1/2 and (c) 0.64. np and n. are the
substrate bulk and surface plasma frequencies, Wp and w. are the
overlayer frequencies and Wi is the frequency of the interface mode.
In (c), a hybridzation gap is opened to avoid the crossing of the
substrate-adsorbate modes indicated by the dashed lines.

mode given by Wi = (n~ + W~)1/2 = 0.75 np. Its weight diminishes approxi-
mately like exp( -2qa). The overlayer mode starts at the adsorbate volume
plasma frequency Wp and disperses downwards to w. = wp /v'2. Its weight
increases approximately like [1 - exp( -2qa)]. This excitation is the analog of
the antisymmetric slab mode resulting from the splitting of surface plasmons
due to Coulomb coupling between the slab surfaces.
The two-step density model described above may be viewed as a rough
approximation to the smoothly decreasing density profile of the actual sub-
strate metal. The low-density 'adlayer' then represents some average of the
tail region near the surface. If this density is taken to be just one-half of the
bulk density (this corresponds to the average of the triangular profile shown
in Figure 3.3), the 'adsorbate' volume plasma frequency wp coincides with the
surface plasma frequency f!. of the 'substrate'. The dispersion of these modes
is shown in Figure 3.38(b). The lower mode is now the most intense at all q.
It evolves gradually from f!. = wp to the 'overlayer' surface plasma frequency
116 CHAPTER 3. SURFACE PLASMONS

W. = wp / J2. If this mode is interpreted as the surface plasmon of the two-step


system, the negative slope at small q is a consequence of the reduced electron
density in the surface region. The higher-lying mode has a positive dispersion
and approaches the interface plasma frequency at Wi == v'3np /2. Its weight
diminishes with q.
From the TDLDA calculations we know that simple metal surfaces exhibit
a multipole surface plasmon at nm ~ 0.8np • We can generate such a mode
in the local optics picture by choosing an 'overlayer' density na == 0.64 n,
where n is the volume density of the substrate. The dispersions for this case
are plotted in Figure 3.38(c}. Instead of modes dispersing from n. to Wi and
from Wp to w., we find one mode dispersing from n. to W. and the second
from wp == nm == 0.8np to the interface plasma frequency Wi = 0.91np •
Remarkably, by opening a hybridization gap between the modes, their crossing
is avoided. The upper mode may be viewed as the analog of the multipole
surface plasmon found by Bennett (1970) for a linear density profile, while the
lower mode corresponds to the ordinary surface plasmon (see Figure 3.4).
Overlayer-induced electronic excitations were also investigated using hydro-
dynamic models. Schwartz and Schaich (1982, 1984) determined the collective
modes for a variety of smooth as well as step-like equilibrium density distri-
butions. The number of modes and their dispersion with q were found to be
sensitive to the diffuseness of these profiles. The boundary conditions imposed
on the hydrodynamic equations also play an important role.
Other studies were carried out using a variety of different approaches, such
as step barrier potentials, tight-binding models and thin jellium slabs (Newns,
1972; Inglesfield, Wikborg, 1975; Nakayama et al., 1984; Eguiluz, Campbell,
1985; Ishida, Tsukada, 1986). However, in these calculations the occupied and
unoccupied electronic states of the semi-infinite substrate-adsorbate system
are generally not treated with sufficient accuracy. Accordingly, the overlayer
modes exhibit appreciable variability. In the remainder of this section, we use
the TDLDA since it provides a fully microscopic description of both ground-
state properties and nonlocal dynamical response.

3.8.2 Jellium Model

Electronic excitations of alkali metal adsorbates can be conveniently stud-


ied using an extension of the jellium model (Lang, 1971). In this model, the
ionic charges of the adatoms are spread out to form a thin slab of uniform
density. The thickness a of one monolayer corresponds to the spacing between
3.B. OVERLA YERS 117

close-packed planes in the bulk. At lower coverages, the average ionic density
is proportionally reduced while the thickness is kept constant. For two mono-
layers, the thickness is doubled and the ionic density is the same as in the
bulk.
Obviously, atomic-like or interband transitions are not included in this
model. On the other hand, the average profile of the equilibrium density is
represented rather well. This is discussed in Chapter 4 when we compare re-
sults of fully three-dimensional calculations (Ishida, Liebsch, 1992) with those
for the jellium model. Surprisingly, even at a coverage of half a monolayer,
where the wave function overlap between alkali atoms is quite weak, the planar
average of the corrugated density is nearly the same as in the jellium model.
The same behavior is found for the density induced by uniform static or time-
dependent electric fields oriented normal to the surface. These results imply
that quantities depending only on lateral averages, like the work function or
the induced dynamic dipole moment, are described rather accurately within
the jellium model. We therefore can expect this model to be qualitatively
correct for excitations at parallel wave vectors that are small compared to the
reciprocal lattice vectors of the overlayer.
The ground-state density profiles of Na layers adsorbed on Al are shown
in Figure 3.39 for several coverages. For c = 1/2, the density decreases mono-
tonically from the substrate across the overlayer region towards the vacuum.
The Na valence charge only amplifies the tail of the substrate density. Even
at one monolayer, it is barely possible to distinguish the adsorbate-vacuum
and adsorbate-substrate interfaces. Once two monolayers are adsorbed, the
density in the overlayer region is nearly constant (apart from the usual Friedel
oscillations) and the two interfaces are well separated. Since the work function
varies strongly with coverage, the distribution of unoccupied states depends
on the coverage as well. For these reasons, it is not surprising that different
overlayer coverages reveal very different excitation spectra.

3.8.3 Alkali Metal Overlayers

Electronic excitations of adsorbed alkali metal layers were calculated within


the TDLDA by Gaspar et al. (1991) (in a slab geometry) and Liebsch (1991)
(for overlayers on a semi-infinite substrate). The formalism is the same as for
clean surfaces, except that the single-particle wave functions and energies used
for the construction of Xl(Z,z',q,w} are generated from the LDA potential of
the adsorbate-substrate system. Once the induced electron density nl(z,q,W}
118 CHAPTER 3. SURFACE PLASMONS

1.0

~
'0;
c
CD
"0
<.>
'c0 0.5
~
<.>
CD
a;

0.0
-5 0 5 10 15

Z (00)
Figure 3.39: Ground-state electronic density profiles for Na layers
adsorbed on AI, calculated within the jellium model. The coverages
are c = 1/2, 1 and 2 monolayers. The density distribution of the
bare Al surface is shown by the dotted curve. The dashed lines
denote the profiles of the positive ionic background.

is calculated, the loss spectra are obtained from the imaginary part of the
surface response function g( q, w}.
Figure 3.40 illustrates several overlayer excitation spectra for K adsorbed
on Al (Gaspar et al., 1991). To achieve convergence, the slab thickness had to
be taken at least 25 Al (111) layers. The spectra are seen to vary strongly with
coverage. Below one monolayer, there is only a broad feature at frequencies
below the alkali metal collective modes. With increasing coverage, the spectral
weight shifts upwards until, at 1.5 - 2 monolayers, two peaks are found. Figure
3.41 shows the momentum variation of the K/ Al spectra for c = 2 and c = 1/2.
Three coverage regions exhibiting distinctly different electronic excitations
can be identified (the mode frequencies wP ' w., and Wm used below refer to the
alkali metal):
• For c = 2, two overlayer modes are found at small q: one near wp , cor-
responding to the antisymmetric slab excitation in the overlayer. This mode
is the same as in the local optics model at small q (see Figure 3.38(a)}. As
a result of quantum mechanical effects and the nonlocal nature of the surface
response, an additional mode appears near Wm ~ 0.8 wp. This mode corre-
3.8. OVERLA YERS 119

l
20 I

,...., K/AI
~
c:
:J

.ci
...
~ c=2
,...., I
10

~]
~ 1.5
50>
1.25
E

-----
1.0
0.8
0.6
0
1 2 3 4
w (eY)
Figure 3.40: Surface excitation spectra for K layers on Al at dif-
ferent coverages, calculated within the TDLDA (qll = 0.05 A-I).
(Gaspar et at., 1991).

sponds to the multipole surface plasmon at the adsorbate-vacuum interface.


The momentum dispersion of these overlayer-induced modes is rather interest-
ing, as shown in Figure 3.42. As q increases, the modes undergo a transition
towards the collective modes of the clean alkali metal surface: The overlayer
multipole mode at small q disperses towards the usual monopole mode w.(q)
of the clean surface, whereas the volume plasmon at small q disperses towards
the multipole mode wm(q) of the bare metal. The transition occurs roughly
when exp( -2qa), which characterizes the electrostatic coupling between the
adsorbate-vacuum and adsorbate-substrate interfaces, becomes small.
This peculiar behavior can be explained as follows (Liebsch, 1991): If the
electrostatic mode and the multipole plasmon were independent overlayer ex-
citations, their dispersions would cross, as indicated in Figure 3.43. However,
since both modes may hybridize due to coupling to electron-hole pairs, they
avoid this crossing by opening a hybridization gap in the excitation spectrum.
As a result, the multi pole mode does not disperse upwards as on the clean
surface, but downwards. At large q, it becomes the ordinary overlayer surface
plasmon. The overlayer volume plasmon, on the other hand, disperses towards
the clean surface multipole plasmon.
The avoided crossing in the overlayer excitation spectrum is analogous to
120 CHAPTER 3. SURFACE PLASMONS

56
K·AI 12 K· AI
48 (a) c=2 (b) c=1/2
0.30
10

3 4 5 2 3 4
ro(eV) ro(eV)
Figure 3.41: Excitation spectra for K on AI. (a) c = 2, (b) c = 1/2.
The spectra in (a) are differently normalized. (Liebsch, 1991).

(a) K· AI

~_s_~_~~.~.~::.~_"~_~~---~­
' j

---

0.0 0.1 0.2 0.30.0 0.1 0.2 0.3


q (A -1) 0-1
q (A )
Figure 3.42: Calculated dispersions of K and Na overlayer modes
for various coverages. Solid curves: c = 2, dotted curves: c = 1;
dash-dotted curves: c < 1; dashed curves: c = 00 (i.e., clean K and
Na modes). Arrows: work functions for c < 1. (Liebsch, 1991).
3.8. OVERLAYERS 121

wp ,
,,

w.(q)

Figure 3.43: Schematic dispersion of overlayer modes for c = 2.


Dashed curves: dispersion of anti symmetric local optics mode (from
wp to w.) and of multipole surface plasmon (from Wm upwards as
for clean alkali metal surface) before 'hybridization'. Solid curves:
the same modes after 'hybridization' interaction is switched on. To
avoid the crossing, a gap in the mode spectrum is opened.

the one we encountered in the discussion of the local optics model of adsor-
bate excitations (see Figure 3.38(c)). The same phenomenon also arises in the
case of the surface plasmon polariton: At small q, the surface plasmon cal-
culated within the electrostatic limit crosses the light line. Once retardation
is taken into account, both modes are electromagnetically coupled. A gap is
then opened, so that the mode crossing is avoided (see Figure 3.20).
From the behavior of the overlayer modes described above, it is now straight-
forward to predict the evolution of these modes with increasing coverage: The
dispersions must be similar to those at c = 2, except that the transition be-
tween the small and large q regimes occurs at smaller q (roughly near q ~ a-I).
In the limit of large adsorbate thickness, this transition occurs at q ~ O. Thus,
the clean alkali metal surface modes evolve naturally within this concept. As
shown in Figure 3.44, this kind of variation with overlayer coverage was indeed
found by Eguiluz and Gaspar (1991) in TDLDA calculations for Na on Ai.
Recent electron energy loss measurements for K layers on Al by Kim et
al. (1997a) are in excellent agreement with these theoretical predictions. The
spectra reveal a main overlayer excitation whose dispersion is shown in Figure
3.45 for various coverages. A second feature is seen at slightly higher frequen-
122 CHAPTER 3. SURFACE PLASMONS

5.5
Na/AI

5.0

>
~
....... 4.5
~
:3
4.0

3.5
o .1 .2

q (.£.-1)

Figure 3.44: Dispersion of Na overlayer modes for coverages be-


tween 2 and 50 monolayers, calculated within TDLDA. The upper
modes begin at the Na bulk plasma frequency wp and disperse to-
wards wm(q). The lower modes begin at Wm ~ 0.8wp , and disperse
towards w.(q). For large coverages, the clean Na surface modes are
approached. (Eguiluz, Gaspar, 1991).

cies. At first glance, the frequency of the main mode near q = 0 seems to
contradict the prediction, namely, that it should begin at Wm regardless of
coverage. For K, the measured multipole frequency is Wm ~ 3.1 eV (see Chap-
ter 4). To resolve this discrepancy, we must recognize that it is, in practice,
impossile using inelastic electron scattering to detect the true q = 0 overlayer
mode frequency because of the non-analytical, cusp-like dispersion.
This may be seen as follows: According to (3.53), the only collective surface
mode in the long-wavelength limit is the substrate surface plasmon determined
by the condition c:(w) = -1. The overlayer modes at q = 0 have vanishing
weight. However, due to the finite aperture of the detector (~0.05 A-I), the
exact q = 0 limit can never be reached. Even if the detector is nominally set
at q = 0, spectral weight from finite q is always present. Thus, the observed
overlayer frequency must lie below the calculated dispersion. The apparent
red shift is larger at higher coverages since the transition from Wm towards
w.(q) occurs at progressively smaller q (see Figure 3.44). In the limit of thick
overlayers, the peak frequency detected at q = 0 converges towards w., the
3.8. OVERLA YERS 123

2.8

--->..
'-'
2.7

---~
:3
2.6

o .05 .10 .15 .20 .25

q (.$.-1)

Figure 3.45: Measured dispersion of main overlayer excitation of


K on Al(111) for coverages ranging from 3 to 44 monolayers. The
frequency of the clean K monopole surface plasmon at q = 0 is
w. = 2.6 eV. (Kim et al., 1997a).

long-wavelength plasma frequency of the semi-infinite alkali metal. Hence, the


main overlayer mode should exhibit a gradual change towards the dispersion
of the K monopole surface plasmon ws(q). This is precisely the trend observed
in Figure 3.45.
The second overlayer mode appearing at slightly higher frequencies is also
red shifted because of the finite aperture of the detector. Thus, at q = 0, it
begins below the nominal value, wp ~ 3.6 eV. With increasing coverage, this
mode gradually approaches the multipole surface plasmon dispersion wm(q)
of clean K. These observations are consistent with recent photoyield measure-
ments for K overlayers on Al which are discussed in Section 4.7.2.
We point out that the electron energy loss spectra by Andersson and Jostell
(1975) for Na double-layers on Ni are in agreement with the theoretical picture
discussed above. In the range q = 0.11 ... 0.75 A-1, they show an upward
dispersion of the main adsorbate-induced loss peak from 3.9 eV to 4.3 eV (see
Figure 3.46) .
• For one monolayer, the electronic density in the overlayer region is not
plateau-like (see Figure 3.39). Accordingly, it is not possible to distinguish two
separate adsorbate interfaces. Both the volume plasmon and the multi pole
124 CHAPTER 3. SURFACE PLASMONS

mode are heavily broadened and mixed. The calculated excitation spectra
then reveal only one very broad feature whose maximum lies slightly below
the double-layer multipole surface plasmon. This trend is consistent with the
experimental data (see Figure 3.47) which show a striking decrease of the
overlayer excitation frequency as the coverage is reduced from two layers to
one monolayer.
• Below one monolayer, the overlayer spectra show a maximum near the
work function (see Figure 3.41(b)). Thus, transitions from the Fermi energy to
the vacuum level have the largest weight. This threshold excitation is in qual-
itative agreement with experimental data on several alkali metal adsorption
systems (Andersson, Jostell, 1975; Aruga, Murata, 1989; Heskett et al., 1988).
Below the work function minimum, all of these spectra exhibit a feature whose
frequency and coverage dependence correspond roughly to the work function.
The data shown in Figure 3.47 support this picture.
The results discussed above demonstrate that, as a result of nonlocality,
the electronic excitations in thin chemisorbed alkali metal layers reveal several
stages: at low coverage, a quasi-collective mode (the threshold excitation)
related to the work function; near about one monolayer, a rapid increase of the
excitation frequency, with overlayer multipole and volume plasmons merged
into a single broad peak; finally, near two monolayers, both modes become
separate spectral features. Eventually, in the limit of thick overlayers, these
two modes evolve towards the monopole and multipole surface plasmons of the
semi-infinite alkali metal.

Overlayer Volume Plasmons

Notice that the appearance of the overlayer volume plasmon at a thickness


of only two monolayers is not incompatible with the dispersion law of bulk
plasmons. If we assume the thickness to correspond to half a plasmon wave-
length, we have qz I':::l 7rla. For two K monolayers, qz I':::l 0.4 A-1. The measured
bulk plasma frequency at this wave vector lies less than 0.1 eV above the value
at q = 0 (Vom Felde et al., 1989). In the case of Na, the difference is about
0.4 eV. Because of the non-abrupt adsorbate-vacuum and adsorbate-substrate
interfaces, the frequency of the overlayer mode should be lower than wp(qz),
i.e., closer to wp(O). Of course, with increasing thickness, the wavelength of
the overlayer excitation increases. The frequency therefore shifts even more
towards the volume plasma frequency in the long-wavelength limit.
3.8. OVERLAYERS 125

10 5 o
Enorgy Loss Cov)

Figure 3.46: Measured electron energy loss spectra for two mono-
layers of Na on Ni. a denotes the deflection angle away from the
specular direction; qll = 0.11 A-1 for a = 3.50 and 0.75 A-1 for
a = 18.9°. (Andersson, Jostell, 1975).

4 I 4

/
(a) (b)
.......
>GI
....., '"''
2-
1/1
1/1 3f-0 ·1 - 3
>- <"3"';3)

0\1;
...m '\ 0

..
R30'
GI
E: '
GI
• 0
Na/ AI(111) Na/Ni(001)
2 f- • - 2
I
0 2 0 2
coverage coverage
Figure 3.47: Measured overlayer-induced loss energies as a function
of coverage. (a) Na on Al (Heskett et al., 1988); (b) Na on Ni
(Andersson, Jostell, 1974).
126 CHAPTER 3. SURFACE PLASMONS

3.8.4 Ag Over layers

We now explore how nonlocal effects influence the collective excitations in


thin Ag layers adsorbed on a metallic substrate. According to the simple
picture discussed in Section 3.8.1, a local model yields an overlayer mode
dispersing from the Ag volume plasma frequency w;
= 3.8 eV towards the
surface plasma frequency w; = 3.7 eV.
To incorporate quantum mechanical and nonlocal effects qualitatively, we
extend the s-d polarization model outlined in Section 3.5 in the following man-
ner. The s electron response is assumed to be given by a jellium-on-jellium
system, as in the case of adsorbed alkali metal layers. The substrate density
is taken to be that of Al (Ts = 2), while the s electron density of the Ag layer
corresponds to Ts = 3. The occupied d bands of the Ag overlayer are simulated
by the same polarizable medium as for semi-infinite Ag. The s-d screening is
present within the region Z = O... a, where a is the overlayer thickness. The
thickness of a single Ag layer (a = 2.4 A) corresponds to the spacing between
close-packed (111) layers in the bulk. For this system, the Poisson equation is
given by (3.76), where cd(Z,W) now has the form

Z < 0,
0:::; z:::; a, (3.105)
a < z.
This represents a special case of (3.88). The boundary conditions for the total
electrostatic potential at the planes Z = 0 and Z = a are

</J'(O-,q,w) Cd(W) </J'(O+,q,w) , (3.106)


Cd(W) </J'(a-, q, w) </J'(a+,q,w) . (3.107)

The dynamical response of this overlayer-substrate system can be solved within


the TDLDA by using the procedure described in Section 3.7.
Figure 3.48 shows calculated excitation spectra for Ag overlayers on AI.
For less than two monolayers, the spectrum is extremely broad, without a
well-defined peak. The overlayer-induced excitations merely amplify the low-
frequency tail of the substrate surface plasmon. Genuine Ag modes appear for
about two monolayers. However, their frequencies lie considerably higher than
the bulk or surface plasmons of semi-infinite Ag. As the coverage is increased,
the overlayer mode becomes sharper and the frequency is reduced. In the limit
of large coverages, the surface plasma frequency of clean Ag is approached.
The dispersion of these Ag overlayer modes as a function of parallel wave
vector is illustrated in Figure 3.49. The modes are seen to be strongly blue
3.8. OVERLAYERS 127

Ag/AI
........ 2
:i c=3
ci
'-'

a:
C"
'-'
Cl 2
.E 1.5

0
3.6 3.8 4.0 4.2

w.(q) (aV)

Figure 3.48: Surface excitation spectra for Ag layers adsorbed on


AI, calculated within TDLDA (q = 0.05 A-1). C denotes the num-
ber of Ag monolayers.

c=1.5 Ag/AI
4.0

>
~
2
r-.
~ 3.8
3
Ag

3.6

0 .1 .2 .3

q (A-I)

Figure 3.49: Calculated dispersion of overlayer plasmons for Ag


adsorbed on Al (solid curves). c denotes the number of Ag mono-
layers. Dashed curve: surface plasmon dispersion of semi-infinite
Ag. Squares: bulk and surface plasma frequencies.
128 CHAPTER 3. SURFACE PLASMONS

shifted relative to the clean Ag surface plasmon dispersion. This shift becomes
smaller with increasing coverage. At large coverages, the clean surface mode
is approached, except at very small q, where the overlayer mode must coincide
with w;. This is so because, as long as retardation is neglected, the only q = 0
surface collective mode is the surface plasmon of the substrate. The overlayer
mode at q = 0 has vanishing weight. Once retardation is taken into account,
the overlayer mode approaches zero frequency in the long-wavelength limit. At
the small but finite q vectors relevant for electron energy loss measurements, it
approaches w;. At large coverages, therefore, the surface plasmon dispersion
of clean Ag is recovered.
The reason for the blue shift of the Ag overlayer mode (relative to the
surface plasmon of clean Ag) seems to be the finite (rather than semi-infinite)
range in which the s-d polarization is active. As the coverage is reduced, the
overall screening of the plasmon charge via the polarizable d electron medium
diminishes, giving a correspondingly higher frequency. Moreover, the spill-out
region not affected by the s-d polarization becomes relatively more important,
causing a further blue shift of the plasma frequency.
In the above model, we assumed that the dielectric response of the d states
in thin Ag layers can be described in terms of the bulk dielectric function. How-
ever, as a result of the microscopic electronic structure, the onset of transitions
involving d states should be less sharp than in bulk Ag. Consequently, the real
part of cd(W) below the onset should be reduced as the overlayer thickness is
decreased. This effect would enhance the blue shift.
Electron energy loss measurements on thin Ag overlayers, which could ver-
ify the thickness dependence of the s-d screening, have not yet been performed.
However, the same finite-size effect influences the non local optical response of
adsorbed Ag layers. This was observed in recent differential reflection spectra
which are discussed in Chapter 4.

3.8.5 Chemical Interface Damping

In Section 3.7, we focused on the surface plasmon dispersion in the pres-


ence of a dielectric medium, such as a thick oxide or noble gas layer. More
pronounced changes can occur if the surface is exposed to chemically reactive
atoms or molecules. The distribution of occupied and unoccupied electronic
states in the surface region is then very different from that on the clean metal.
The probability of creating electron-hole pairs and surface collective modes
changes accordingly. Rather than considering the new intrinsic overlayer ex-
3.8. OVERLAYERS 129

6
'""'
::j
d
.......
'""'
3 4
.;
0;

2

0
3.6 3.7 3.8 3.9

w (eV)
Figure 3.50: Surface excitation spectra for thin Os layers adsorbed
on Ag, calculated within TDLDA (q = 0.05 A-1). The parameter
c denotes the Os coverage.

citations, we investigate here the mode at the substrate-adsorbate interface,


Le., the overlayer-induced modification of the substrate plasmon.
Figure 3.50 shows Ag surface excitation spectra for increasing amounts
of adsorbed Os. The equilibrium electronic density is calculated within the
jellium-on-jellium model described in 3.8.2. The Ag substrate is treated within
the s-d polarization model introduced in Section 3.5. The boundary of the
polarizable d electron medium is Zd = -0.8 A. At low Os coverages, the Ag
surface plasmon peak is only weakly affected since the effective Os plasma
frequency is too small. For instance, at c = 1/4, w;(Os) = wp(Os)/2 ~ 1.8 eV.
Thus, the overlayer is transparent. As the coverage approaches one monolayer,
however, the Os collective modes become nearly degenerate with the Ag surface
plasma frequency. The interference of these modes causes an increasing blue
shift of the Ag surface plasmon. This should then be more correctly described
as OslAg interface plasmon.
The calculated spectra also reveal a striking plasmon broadening due to new
channels for decay into electron-hole pairs at the interface. These channels are
related to occupied and unoccupied adsorbate-induced electronic states. This
broadening is referred to as chemical interface damping and was also observed
in absorption spectra of small metal particles.
130 CHAPTER 3. SURFACE PLASMONS

3.9 Small Metal Particles


3.9.1 Classical Picture

The electronic excitations at metal surfaces discussed in the preceding sec-


tions have a close relationship to those in small metal particles. We assume the
particle radius R to be much smaller than the wavelength of the radiation, so
that retardation effects can be ignored. On the other hand, the particles should
be large enough, so that the discrete electronic level structure is unimportant.
Hence, R is taken to be about 10 - 100 A. If the dielectric properties of the
spherical particle are described in terms of the local bulk dielectric function,
the dynamic dipole polarizability in the classical picture is given by

a(w} = R3 c(w} - 1 . (3.108)


c(w} + 2
The absorption cross section is a(w} = (41TWjC) Ima(w}. For simple metals
represented by the Drude function, the condition c(w} = -2 yields the Mie
resonance frequency WM = wpj v'3.
Just as in the case of surface plasmons, it is worthwhile to study the effect of
the microscopic surface properties on the electronic excitations of the particle.
Because of the smooth decay of the equilibrium density at the surface and the
nonlocality of the response to the radiation, the Mie resonance is shifted and
broadened. (For a detailed review, see Fuchs, 1992.)

3.9.2 Simple Metal Particles

The finite-size effects on the particle polarizability were calculated within


the TDLDA as described in Chapter 2 (Ekardt, 1985a; Puska et al., 1985).
Such a theory naturally includes the finite width of the Mie resonance due
to decay into electron-hole pairs. If damping in the interior of the particle
is neglected, the width results purely from surface scattering and is therefore
proportional to R- 1 • This is analogous to the damping of surface plasmons
in the small-q limit. In the absence of bulk absorption, the plasmon width is
proportional to q since the surface plays a negligible role in the long-wavelength
limit. Similar arguments apply to the shift of the Mie frequency related to the
location of the fluctuating screening charge.
To illustrate the similarity between electronic excitations in small metal
particles and at flat metal surfaces, we show in Figure 3.51 the density induced
3.9. SMALL METAL PARTICLES 131

.2
~ rs=4
:J

~
...
'7{ 0

o 10 20 30

r (00)

Figure 3.51: Radial distribution of screening density induced in Na


particle (R = 12.3 A; N=198 electrons) by a static electric field.
Plotted is the quantity a(r) = (47rr 2 /3)nl(r). (Ekardt,1985a).

in a Na jellium particle. Screening in the outer regions of the equilibrium


density is very efficient, so that the centroid of the induced density is located
outside the positive ionic background. Only weak Friedel oscillations extend
into the interior of the particle. These distributions are strikingly similar to
those at a flat jellium surface (see Figure 2.5). The centroid positions relative
to the jellium edge are nearly the same (about 0.7 A).

Self-Energy Formulation of Surface Response Function

To exploit the analogy between electronic excitations in small metal parti-


cles and at flat metal surfaces further, we do not discuss here the microscopic
TDLDA calculations but follow a different, more qualitative path (Liebsch,
1993b). For this purpose, we express the response function g(q,w) for a flat
metal surface in terms of a modified dielectric function

w2
c(q,w) = 1 - w2 + {(q,w) (3.109)

The self-energy is given by

(3.110)
132 CHAPTER 3. SURFACE PLASMONS

(We take dll = D.) The effective dielectric function accounts for nonlocal effects
associated with scattering processes at the surface. Using these definitions, the
small-q expansion of g(q,w} given in (3.54) can be written as:

c(q, w} - 1 w:
9 (q,w ) = = (3.111)
c(q,w} + 1 w~ - w2 - ~(q,w)
The true nonlocal surface response at small q has been reformulated in terms
of a self-energy representing the red shift and damping of the surface plasmon.
Insertion of (3.110) into (3.111) yields the dispersion given in (3.23).
Figure 3.52(a} illustrates the frequency dependence of the self-energy for a
jellium surface corresponding to Na. The main spectral feature results from the
multipole surface plasmon near Wm = 0.8 wP' but there is appreciable particle-
hole broadening over a wide frequency range. The real part of the self-energy
is directly related to the centroid of the induced screening charge. Up to about
wm , it is located outside the metal since screening in the tails of the density
profile is very efficient. Above wm , higher-density regions must take part in the
screening and Red.l(w} shifts inside. As shown in Figure 3.52(b}, these surface
scattering processes give rise to the red shift and broadening of the ordinary
surface plasmon and to the appearance of the multipole surface plasmon.

Self-Energy Formulation of Particle Polarizability

In analogy to the formulation of the surface response function, we now


write the particle polarizability as

(R } =R3C(R,w}-1 (3.112)
a ,w c(R,w) + 2 '

with the effective dielectric function


w2
c(R,w} = 1 - w2 + ~(R,w) . (3.113)

The particle polarizability then takes the form:

3 w~
a(R,w}=R 2 2 ~(R ,w ) (3.114)
WM-W -

The complex self-energy ~(R, w) accounts for scattering processes and nonlocal
effects at the surface of the particle. As a result of these finite-size effects, the
Mie resonance is shifted and broadened.
3.9. SMALL METAL PARTICLES 133

," ,
5 ,, ,,
~
(0) ,
:J ,,
~ ,,
~

3 ,
~ 0
w

3.:
CT
Cb)
0;
0
E w.
0>
Wm
~ -1

..,
'"
'-..
(c)
3.:
'"
'S' 0
E
0> -1
~

2 3 4 5

W CeV)
Figure 3.52: (a) Frequency dependence of self-energy I;(q, w) of
Na surface plasmon for q = 0.05 a.u. Solid (dashed) curve: real
(imaginary) part. (b) Logarithm of Img(q,w) for q = 0, 0.025
and 0.05 a.u. (solid, dashed and dotted curves); (c) Logarithm of
Ima(R,w) for R = 00, 40 and 20 a.u. (solid, dashed and dotted
curves). Ws and Wm: monopole and dipole surface plasmonsj WM:
classical Mie plasmon. The spectra for q = 0 and R = 00 are
artificially broadened. (Liebsch, 1993b).

Since we are concerned here with a qualitative discussion of the absorption


spectra of small metal particles, we do not attempt to calculate the self-energy
by using the TDLDA. Instead, we make the assumption that, in the limit
of large R, the response properties at the surface of a spherical particle be-
come similar to those at a flat metal surface in the long-wavelength limit. For
134 CHAPTER 3. SURFACE PLASMONS

.... __ + _0#

A=21T/q A=21TR

Figure 3.53: Illustration of fluctuating charge associated with sur-


face plasmon and with Mie responance. The parallel wave vector q
implies a wavelength A = 21r/q. In the case of a spherical particle,
the wavelength corresponds to the circumference A = 21r R. This
analogy yields q rv I/R.

frequencies not too close to the volume plasma frequency, this assumption is
justified since quantum mechanical scattering processes contributing to the
self-energy occur within a very narrow region near the surface.
The variable 1/ R is a measure of the surface to volume ratio. It therefore
plays the same role as q in the case of the flat surface. This may also be seen
by comparing the 'tangential' wavelengths of these collective oscillations (see
Figure 3.53): At the flat surface, it is given by A = 21r/q, while, in the case
of a particle, it corresponds to the circumference A = 21rR (Apell, Ljungbert,
1982; Ekardt, 1985b). In analogy to (3.110), we approximateI:(R,w) as
I:(R, w) ~ R- 1 d.L(w) (w; - w2 ) , (3.115)
where d.L(w) should have a similar magnitude and spectral dependence as the
corresponding centroid for the flat surface. A different self-energy formulation
was given by Zaremba and Persson (1987), who write ImI:(R, w) in terms of
the golden rule formula and derive an approximate expression for the effective
local potential.
Substituting (3.115) into (3.114), it is easily verified that the above choice of
I:(R, w) is consistent with the static limit of the particle polarizability (Snider,
Sorbello, 1983):
(3.116)
This expression indicates that the effective particle radius is larger than R
since the centroid of the polarization charge is located outside the positive
background.
3.9. SMALL METAL PARTICLES 135

Size Dependence of Mie Plasmon: N a and K Particles


From the above ansatz for the particle self-energy, we obtain the following
size dependence of the Mie resonance in the limit of large R:
wM(R) = [w~ - Re~(R,wM)t2
~ WM [1-R-1Red.l(WM) +O(R- 2 )] (3.117)
This dependence agrees with the one derived by Apell and Ljungbert (1982).
The 'dispersion' relation of the Mie plasmon can be directly compared to
the behavior of the surface plasmon in the small-q limit (see (3.23)). Since
Red.l(WM) > 0 for simple metals, surface scattering processes cause a red shift
with decreasing radius. The linear coefficient is given by the centroid of the
dynamical screening charge in the large R limit. Figure 3.52(c} illustrates the
spectral distribution of 1m a(R, w} for Na particles at different radii. The main
peak corresponds to the shifted and broadened Mie resonance, while the up-
per feature is the counterpart of the multipole surface plasmon. In the particle
case, this excitation has dipolar angular symmetry, just as the Mie resonance,
but the fluctuating charge exhibits an extra node in the radial direction. This
mode exhibits a blue shift with decreasing particle size.
The comparison of Ima(R,w) as calculated from (3.114) with the micro-
scopic TDLDA results of Ekardt (1985a) for a Na particle shows excellent
qualitative agreement (see Figure 3.54). Although the fine structure caused
by interlevel transitions is, of course, absent in the self-energy approach, the
red shift of the Mie resonance and its intrinsic width are very similar in both
theories. In addition, as a result of the multipole surface plasmon, the particle
polarizability shows a weak second feature at frequencies slightly above the Mie
resonance. The spectral weight exhibited by the TDLDA calculations in this
frequency range seems to have the same physical origin. Near wp , the analogy
between flat metal surfaces and small metal particles breaks down since the
metal becomes transparent. It is then no longer valid to consider only surface
excitations.
The self-energy calculations of Zaremba and Persson (1987) are also in qual-
itative agreement with the microscopic results, but there are some important
differences: The Mie resonance is mainly unshifted, with only a weak shoulder
on the low-frequency side. In addition, the spectral weight in the region of the
multipole plasmon is absent. The reason for these differences is presumably
related to the approximate nature of the local potential used in the golden
rule.
Figure 3.55 illustrates the size dependence of the Mie plasmon for Na and
K clusters. The resonance frequencies are seen to vary approximately linearly
136 CHAPTER 3. SURFACE PLASMONS

2,---,----,---,----,---,----,---,----,--,

No

.;
'" -2
~
.,
<•
,

_4L-__ ~ ___ L_ _ ~ _ _ _ _L __ _~_ __ L_ _~_ _ _ _L-~

o 2 4 6 8

w (eV)

Figure 3.54: Logarithm of polarizability of Na particle (R =


12.3 A). Dashed curve: microscopic TDLDA calculations of Ekardt
(1985a); solid curve: spectrum based on self-energy (3.114); dot-
dashed curve: self-energy approach of Zaremba and Persson (1987).
The arrows indicate the frequencies of the classical Mie resonance
WM, bulk plasmon wp and multipole surface plasmon Wm ~ 0.8wp.
{Liebsch, 1993b}.

with 11R, with a slope similar to that of the surface plasmon at small q. For
the K clusters and the largest Na clusters, Red.1.{wM} ~ 1.3 A. This value
agrees amazingly well with the TDLDA predictions for jellium surfaces: Ac-
cording to Figure 4.5, Red.1.{wp /v'3} ~ 1.3 A for T. = 4 and T. = 5. Keep
in mind, however, that the comparison between excitations at flat metal sur-
faces and in small particles is nontrivial since different preparation techniques,
temperature, etc., also influence the frequencies and widths of the collective
modes.
Experimental evidence of the existence of a radial and angular dipolar
plasma oscillation in small metal particles was found in measurements of
the absorption cross section of large K clusters containing 500 - 900 atoms
{Brechignac et al., 1992}. These data show weak spectral weight near 2.9 eV,
i.e., exactly where this higher-order collective particle resonance is expected to
occur.
3.9. SMALL METAL PARTICLES 137

I I
1.0 -
~,
,,
,,
K
2::z 900'-,,_
,,
3 ,,
500 ,,
'-..
---.. 0.9 - ,,
,,
""
'-.. ,,
,, Na
~
,. ,,
3 93 0
,
,, 9 0
59 0 ,, 0
,, 21 -
0.8 - 41 0 ,,
I I
0 .1 .2

l/R (.A.-')
Figure 3.55: Normalized Mie frequency for K and Na clusters
as a function of inverse particle radius. Solid squares: K data
(Brechignac et al., 1992); open squares: Na data (Reiners et al.,
1995). The numbers denote the number of atoms per cluster. The
dashed line indicates the slope expected from TDLDA calculations
for flat jellium surfaces with r. = 4 and 5 (Liebsch, 1987).

Li Particles

While the Mie plasma frequencies of Na and K clusters are within a few
percent of the values expected for nearly-free-electron systems, recent pho-
toabsorption measurements on Li clusters (Brechignac et al., 1993) reveal fre-
quencies about 25 % below the nominal value for a metal with r. = 3.25
(WM = 4.6 eV). However, the extrapolation of the experimental peak positions
to the R = 00 limit (3.55 eV) agrees with the maximum of 1m (c -l)/(c + 2),
where c(w) is the measured bulk dielectric function. As in the case of the
Li surface plasmon (see Section 3.4), the large ionic pseudopotential causes a
significant reduction of the Mie resonance frequency.
According to the electron energy loss data shown in Figure 3.25, the im-
portant interband transition at 3.2 eV nearly coincides with the Mie plasmon.
It is therefore not surprising that the line width of the Mie resonance is much
broader than for comparably large K clusters. On the other hand, the negative
linear slope of the normalized Li Mie plasmon WM( R) /WM( 00) with inverse ra-
138 CHAPTER 3. SURFACE PLASMONS

4.5

P
Ag
P
>
...!. 4.0 pppp,p~
f-t~
(2 +
'-- +
+
'-'
+ +
3'" +
*
+-11-
3.5

o .1 .2 .3 .4
1/R (A-')

Figure 3.56: Ag Mie frequency as a function of inverse particle ra-


dius. Solid dots: free Ag+ clusters (9 to 70 atoms); empty dots:
free Ag- clusters; crosses: Ag particles (10 A ::; R ::; 50 A) em-
bedded in Ar matrix, after correction for the dielectric constant of
the matrix (Charle et al., 1989). The straight line is a fit through
the classical Ag Mie resonance at WM = 3.5 eV and the Ag+ data.
(Tiggesbaumker et al., 1993).

dius (about -1.5 A) is considerably larger than that of the Li surface plasmon
(roughly -0.3 A; see Figure 3.26). The origin of this different behavior is at
present not understood (Serra et al., 1993).

3.9.3 Ag and Hg Particles

In striking contrast to the red shift of the Mie resonance observed for simple
metal particles, the main absorption peak of Ag particles exhibits a blue shift as
a function of decreasing particle radius (Kreibig, Genzel, 1985; Tiggesbaumker
et al., 1993) (see Figure 3.56). As in the case of the Ag surface plasmon disper-
sion, the presence of the filled 4d states alters the sign of the size dependence
of the particle resonance frequency.
The s-d polarization model discussed in Section 3.5 is also useful for cal-
culating absorption spectra of Ag particles (Liebsch, 1993b; Kresin, 1995). If
3.9. SMALL METAL PARTICLES 139

we represent the d electron medium in the limit of large R by the dielectric


function Cd(W}, the condition c(w} = -2 implies that the Mie frequency is
reduced from its unscreened value WM = wp /v'3 = 5.3 eV to
W
w~~ ~ ~3.5 eV. (3.118)
2 +Cd
This screened value agrees with the extrapolation of the measured frequencies
to the limit of large particle radii.
Since the Ag 4d levels are more localized than the s states, the s-d screen-
ing is absent in the surface region where the 5s electrons spill into the vac-
uum. Thus, part of the density fluctuation associated with the Mie resonance
oscillates at the un screened plasma frequency. This mechanism increases the
resonance frequency. With decreasing particle radius, this effect becomes more
pronounced because of the larger surface to volume ratio. The Mie resonance
should therefore exhibit a blue shift with decreasing particle size.
The self-energy arguments given above can be extended by taking into
account the presence of the polarizable medium. It then follows that the
linear coefficient of the Ag Mie resonance should be similar to the slope of the
Ag surface plasmon. The data support this picture: The slope of the solid line
in Figure 3.56 is 0.9 A, while the slopes of the Ag surface plasmon dispersions
in Figure 3.27 range from 0.4 to 0.8 A. Similar results are also obtained within
microscopic TDLDA calculations for Ag clusters in which the s electrons are
represented via a jellium sphere and the d states via an array of dipoles (Serra,
Rubio, 1997).
Figure 3.56 also illustrates the effect of an Ar matrix on the Mie resonance
(Charie et al., 1989). Interestingly, the data exhibit an even larger blue shift
with decreasing particle radius than Ag particles in vacuum. This observation
is consistent with the extra blue shift of the Ag surface plasmon induced by
an external dielectric medium, as discussed in Section 3.7 (see Figure 3.37).
The physical reason for this behavior is that, at large R, the effect of the Ar
matrix is strongest, lowering the classical resonance from w~ ~ wp / y'2 + Cd
to wp / y'2co + Cd. Here, co is the dielectric constant of the Ar matrix. At
decreasing radius, this reduction becomes less effective because of the increas-
ing surface-to-volume ratio: The polarization red shift due to both Cd and co
diminishes with decreasing particle size.
Notice that the observed resonance frequencies of small Ag particles may be
influenced by various other effects: particle temperature, defect concentration,
deviations from Ag bulk interatomic spacing, etc. All of these produce shifts of
the resonance frequency, in particular, in very small Ag clusters whose discrete
atomic geometry must also be taken into consideration. Transitions between
140 CHAPTER 3. SURFACE PLASMONS

cluster levels then become important, and the description of absorption spectra
in terms of macroscopic response quantities becomes inappropriate. A detailed
discussion of these effects was given by Kreibig and Genzel (1985).

Hg Particles

An interesting intermediate case between simple metal and Ag particles are


Hg clusters. Absorption spectra for clusters of up to 100 atoms (Haberland et
al., 1992) indicate no dispersion or a weak positive dispersion of the Mie plas-
mon in the range 1jR ~ 0.12 ... 0.16 A-i. This behavior is compatible with
the Hg surface plasmon data discussed in Section 3.5.2. These show a weak
negative dispersion for q less than about 0.12 A-1 and a positive dispersion for
larger q (see Figure 3.33). Hence, as a result of the polarization of the 5d core
levels, the size dependence of the Hg Mie plasmon is much less pronounced
than in alkali metal particles. On the other hand, since the 5d levels lie much
deeper than the Ag 4d states, this deviation is not as large as in the case of
Ag particles.
The cluster data by Haberland et al. (1992) reveal a smaller line width
of the Hg Mie plasmon than would be expected from the classical limit. This
result appears at first surprising since it is generally assumed that the width
increases roughly like r = roo + AVF j R. Here, roo is the width obtained from
the classical polarizability (3.108), VF the Fermi velocity and A a parameter of
order unity (Kreibig, Genzel, 1985). It is clear, however, that 'bulk' absorption
processes must diminish with decreasing radius. If this reduction occurs faster
than the increase in surface-related broadening, it is indeed plausible that the
Mie line width passes through a minimum. This phenomenon is similar to the
one discussed in Section 3.3.4 for the broadening of surface plasmons: In the
small-q limit, the width is dominated by bulk processes because of the slow
decay of the plasmon field. With increasing q, this penetration depth into the
bulk diminishes, giving a smaller bulk-induced broadening. If this reduction is
faster than the increase of surface-induced broadening, the total width should
exhibit a minimum.

3.9.4 Chemical Interface Damping

As discussed in Section 3.7, the frequency and dispersion of surface plas-


mons changes appreciably if the metal is in contact with a dielectric medium.
These changes become even stronger if the adsorbed species forms chemical
3.10. QUANTUM WELLS 141

bonds with the metallic surface electrons. Such a 'chemical interface' can give
rise to new states near the Fermi energy and, consequently, to new channels
for electron-hole pair creation in the surface region. These processes greatly
enhance the broadening of surface plasmons.
Similar phenomena have been observed on small metal particles (Charie
et al., 1989; Hovel et al., 1993). For example, Ag particles in contact with
CO molecules exhibit much larger widths of the Mie plasmon than those im-
mersed in inert noble gas matrices. As shown by Persson (1993), this enhanced
broadening is closely related to the increased density of states near E F •

3.10 Quantum Wells

In this section, we briefly call attention to the close similarity between


electronic excitations at metal surfaces and those in semiconductor quantum
wells (Dobson, 1993). With the aid of molecular beam epitaxy, it is now
possible to grow heterostructures of a wide range of electronic densities and
layer spacings. These systems form nearly ideal jellium systems whose profile
can be designed for specific purposes (Yuh, Gwinn, 1994). This capability
makes it feasible to study the dependence of the electronic modes on the system
parameters in a controlled manner. Although most experimental studies so far
were carried out on charged quantum wells (Pinsukanjana et al., 1992), we
consider here for simplicity a single neutral well to illustrate the key aspects
of the electronic excitations.
A popular heterostructure is Gal_",AI",As which can be grown from GaAs
crystals by adding Al. Structures containing arbitrary Al concentrations x
can be produced (Gwinn et al., 1989). The band gap and conduction band
minimum of this semiconductor then acquire a specific profile as a function of
the growth direction. We take this direction to be the z-axis, assuming per-
fect crystalline symmetry in the x - y plane. Conduction electrons supplied
via remote donors move in the wells with an effective mass m*. The semi-
conductor provides an average dielectric constant defined as co. In the case of
Gal_",AI",As, one typically has m* /m = 0.069 and co = 13. The effective Bohr
radius is then ali = n2co/m*e2 :=:::; 100 A. The lattice structure can therefore be
neglected.
Before we discuss the TDLDA excitation spectra for such a quantum well,
let us consider the modes derived within the local optics picture. In the
classical model, the dielectric function is c = co(1 - w;/w2), where wp =
142 CHAPTER 3. SURFACE PLASMONS

(47rne 2/com*)1/2 is the bulk plasma frequency. The dispersion of the electro-
static slab modes with parallel wave vector is given by

( -qL) 1/2
W±(q)=w.1=r=e , (3.119)

with w. = wp / J2. In the limit of large thickness L, the two slab modes
approach the surface plasma frequency. At finite L, the degeneracy is lifted due
to the electrostatic coupling between the two slab surfaces. The fluctuating
charge density associated with the low-frequency mode w+(q) is symmetric
with respect to the slab center ('breathing mode'); that of the upper mode
w_(q) is antisymmetric (the electronic density 'sloshes' across the slab). The
latter is analogous to the electrostatic overlayer mode dispersing from wp to
w. illustrated in Figure 3.38(a).
The local optics picture, of course, ignores several aspects of the electronic
properties of the quantum well that ought to manifest themselves in the exci-
tation spectrum. Owing to the finite width of the slab, electronic transitions
between discrete subbands should appear as spectral features. In addition,
as a result of the smooth density distribution at slab surfaces, all modes are
shifted and broadened. New collective surface modes may also appear if the
surface polarizability is sufficiently high.
These kinds of spectral modifications due to quantum mechanical and non-
local effects have indeed been found by Schaich and Dobson (1994) in TDLDA
calculations for a single quantum well (see Figure 3.57). The slab excitations
were determined by considering the response to an external potential of the
type (2.62) and solving the real-space response equation (2.48). The spectral
distribution is derived from the imaginary part of the surface response function
g(q,w).
At the density and thickness shown, there are nine bound levels below the
Fermi level. As a result, intersubband transitions are clearly visible. Nonethe-
less, the spectrum is dominated by collective modes. Apart from the symmetric
and antisymmetric plasmon modes, a multi pole surface plasmon is found at
about 0.8 wp , just as on the surfaces of simple metals. The induced density
of this mode is odd, like that of the upper surface plasmon. However, the
monopole and dipole characters of these modes are not well resolved since
both densities specify the driven response to the applied potential rather than
true eigenmodes (see also Section 3.3.1).
In analogy to the effect of surface charging discussed in Section 3.6, the
excitation spectra of charged quantum wells differ greatly from those of neu-
tral wells. The stiffening of the density profile upon positive charging reduces
the surface polarizability and leads to the suppression of the multipole surface
3.10. QUANTUM WELLS 143

(a)

W_

E j\.f- iJ "
°0~---1~0~~~2~0~~3~0~

W/21T (em-I)
--:-
:J
8
~ Wm (b) (e)
~
4
~
0"
0
~
C -4
E -8
-40 40 -40 -20 20 40

Figure 3.57: (a) Electronic excitation spectrum for neutral quan-


tum well as calculated within the TDLDA. The slab thickness is
L = 40a~, the bulk density corresponds to rs = 3 and the parallel
wave vector is q = O.016/ao. The main spectral features corre-
spond to the symmetric slab mode w+, the multi pole mode Wm
and the antisymmetric slab mode w_. The induced densities of the
multipole surface mode and antisymmetric slab mode are shown in
panels (b) and (c), respectively. The dashed curve in (c) gives the
induced density obtained at w_ from a single-step hydrodynamic
calculation. (Schaich, Dobson, 1994).

plasmon (Dobson, 1992). These conditions can be achieved, for instance, by


immersing the electron gas in a wide positive background, so that the bare
electrostatic potential is essentially parabolic. The experimentally observed
modes of such systems are in qualitative agreement with the coupled plasmons
predicted by the local optics model (Pinsukanjana et al., 1992) and hydrody-
namic theory (Zaremba, Tso, 1994; see also Section 3.6.1).
Chapter 4

Nonlocal Optics

The microscopic electronic properties of metal surfaces can be probed with


a variety of optical spectroscopies: ellipsometry, differential reflection spec-
troscopy, electroreflection, and attenuated total reflection. Surface photoemis-
sion and inverse photoemission are also intimately related to these microscopic
properties since the amplitude and spatial variation of the local fields in the
surface region determine the frequency dependence of the measured intensity.
The proper treatment of the surface electronic structure and of its dynamical
response characteristics require a generalization of classical Fresnel theory. In
particular, the finite width of the surface density profile and the induced sur-
face charge must be taken into account. Moreover, since these distributions
vary rapidly on the scale of the wavelength, the nonlocal nature of the surface
response cannot be omitted. Hence, near the surface the microscopic fields
deviate appreciably from the standard Fresnel fields. At optical frequencies,
the effect of these microscopic properties on observable quantities can be de-
scribed in terms of two complex functions: d1-(w) and dll(w). The real parts of
these quantities specify the location of the normal and parallel induced surface
currents; the imaginary parts represent the corresponding surface excitation
spectra. The most prominent spectral feature in d1-(w) is the multipole surface
plasmon, which has been identified on several simple metals. In the case of Ag,
the surface excitations are strongly affected by d bands. The great sensitivity
of surface optics with respect to the microscopic electronic structure is evi-
dent from the changes induced on surface charging. Optical spectroscopies are
also well suited for investigating surface excitations localized in thin metallic
overlayers.

145

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
146 CHAPTER 4. NONLOCAL OPTICS

,
,,Z

x
,
, --

Figure 4.1: Reflection and transmission of a p-polarized light wave


at a surface.

4.1 Classical Picture and Phenomenological


Extensions

In standard Fresnel theory a metal is described in terms of the macroscopic


local dielectric function c(w) and only transverse solutions to Maxwell's equa-
tions are admitted. The presence of the surface enters via boundary conditions
for these transverse waves. Let us consider an incident plane wave with wave
vector if = (<]iI, -qz)' which is reflected from a semi-infinite medium charac-
terized by c(w) (see Figure 4.1). Part of this wave is transmitted with wave
vector if' = (<]iI, -q~), where q' = vrcq, q = wlc, and, according to Snell's
law, q~2 = q; + q2(c - 1). The usual boundary conditions imply that the par-
allel components of E and if and the perpendicular components of jj and jj
are continuous at the surface. In the case of p-polarized light (the electric field
vector lies in the plane of incidence), these conditions give the usual Fresnel
expression for the reflection coefficient (Jackson, 1962):

(4.1)

For s-polarized light (the electric field vector is perpendicular to the plane of
incidence), we find instead

(4.2)
4.1. CLASSICAL PICTURE 147

<:( (J)

o a z
Figure 4.2: Phenomenological three-layer model of dielectric re-
sponse at surfaces (McIntyre, Aspnes, 1971). In the extension by
Plieth and Naegele (1977), Ca is allowed to differ for normal and
tangential fields.

According to the above assumptions, these expressions are independent of


surface properties. In particular, these formulas ignore quantum mechanical
effects that lead to the smooth decay of the ground-state density profile at the
surface and to the finite spatial extent of the induced surface charge density.
Attempts to go beyond the simple Fresnel picture were already made by
Drude (1891, 1959) who was concerned with the influence of surface condi-
tions on the reflectivity. To account for surface contributions to the reflection
amplitudes, McIntyre and Aspnes (1971) extended the classical model by in-
troducing a thin layer characterized by a local dielectric function Ca(W) (see
Figure 4.2). Thus,

c(w) , z:SO,
c(z,w) = { ca(W) , O<z:Sa, (4.3)
1, a < z,

where a is the thickness of the surface layer. Plieth and Naegele (1977) gener-
alized this model by allowing for the anisotropic response of the surface layer
to parallel and normal field components. In this so-called three-layer model,
the thickness of the surface layer and its dielectric properties can be modeled
to fit observed spectra. For example, in the case of not-too-thin adsorbed lay-
ers with a known dielectric function ca(w), the three-layer model can be used
to determine the overlayer thickness from reflectivity data. Even if the model
parameters are adjusted to achieve a representation of experimental data, it
is nevertheless clear that they provide little insight into the microscopic elec-
tronic properties of the surface region and the physical origin of the optical
148 CHAPTER 4. NONLOCAL OPTICS

response.
Surface optical data were also analyzed in terms of hydrodynamic theory
(Forstmann, Gerhardts, 1986). Unfortunately, the excitation of electron-hole
pairs, which contributes crucially to the microscopic fields near the surface,
is usually treated in an empirical manner. For this reason, hydrodynamic
results tend not to have sufficient predictive accuracy. To go beyond the clas-
sical models and other phenomenological schemes, it is necessary to develop a
microscopic description of the interaction of light with matter.

4.2 Microscopic Theory

The microscopic theory for the interaction of electromagnetic waves with


matter is based on Maxwell's equations. In the absence of external sources
and for nonmagnetic systems, these are

1 oE 471" -t
V·E 471" p , Vxii --+-J,
c {)f; c
(4.4)
1 off
v·ii = 0, VxE -;;7Jt , (4.5)

where p is the induced charge density and 1the induced current density. Equa-
tions (4.4) can also be expressed as

V·V = 0, vxii (4.6)

The displacement field V is given by

av
at =
7
471" J +
oE
7ft . (4.7)

With this definition, the induced charge and current densities satisfy the con-
tinuity equation
~ op
Voj+{)f;=O. (4.8)

The induced polarization P is introduced via the relation


(4.9)
4.2. MICROSCOPIC THEORY 149

Gauss's law and the continuity equation imply

V·P= oP - (4.1O)
-p, fjt=i.

The two curl equations in (4.5) and (4.6) can be combined into the wave
equation
2
V(V· E) - V2E = w2 fj . (4.11)
c
Within linear response, the displacement field fj and the electric field E are
related via the nonlocal constitutive equation (i, i denote Cartesian coordinates
and repeated indices imply summation)

D;(f,t} = f dV loo dt' c;j(f,f',t-t') EiU'·',t'). (4.12)

All information concerning the electronic properties of the semi-infinite metal


is contained in the microscopic dielectric tensor C;j. Performing a Fourier
transformation with respect to time, this relation becomes

(4.13)

Analogously, the induced current density and the polarization are given by

i;(f,w) = f d r' a;j(f,f',w) Ej(f',w} ,


3 (4.14)

Pi(f,w} f d r' Xij(f,f',w} Ej(f',w) ,


3 (4.15)

where the conductivity aij and the polarizability X;j are related to the dielectric
function via

--,
Cij (r,r ,w ) J: UJ:( r_ - r_'} + -4?ri a" (-
U"
'J
-, )
w 'J r , r , w (4.16)
6ij 6(f - f') + 4?r Xij(f, f', w} . (4.17)

These equations include the nonlocal, anisotropic nature of the response


and the rapid spatial variation of the electronic density. Anisotropy obviously
exists between directions normal and parallel to the surface, but for certain
crystal faces orthogonal parallel directions may also respond differently. Nonla-
cality is a consequence of the rapid spatial variation of certain field components,
which therefore cannot be extracted from the constitutive relations.
As in Chapter 3, we use here a self-consistent field approach. Hence, the
fields in the above equations are effective local fields that incorporate all screen-
ing processes. Since electrons respond to these fields as independent particles,
150 CHAPTER 4. NONLOCAL OPTICS

all response functions are bare, independent-electron response functions. The


single-particle interaction Hamiltonian has the standard form

H' = __e_
2mc
[po A(f,w) + A(f,w). p] + e<I>(f,w) , (4.18)

where A is the vector potential and <I> the scalar potential. The local electric
field is related to these time-dependent potentials via B = (iw / c) A - V <I>. In
the Coulomb gauge, v· A = 0, and <I> can be taken as zero. The Hamiltonian
H' then reduces to H' = -(e/mc) A.p= (ie/mw) B.p. For purely longitudinal
fields, we can set A = 0 and use instead H' = e<I>(f,w) == <Pscf(f,w).

4.2.1 Long-Wavelength Limit

The general solution of the above response equations is exceedingly com-


plex. At optical frequencies, we can use the fact that the radiation wavelength
is much larger than the typical length scale on which electronic surface phe-
nomena take place. The decay of the density profile at the surface and the
width of the induced density occur on the scale of a metallic screening length,
i.e., on the order of a few A. The longitudinal surface field associated with the
induced density varies on a similar scale. Friedel oscillations of the induced
charge and current densities decay more slowly, but their range is also much
shorter than the radiation wavelength.
Let us denote the boundaries of the surface region where the microscopic
fields differ from Fresnel fields by Zl ~ 0 and Z2 » 0 (Figure 3.18). At
optical frequencies, qL ~ 1, where L = Z2 - Zl. It is therefore justified to
expand the fields appearing in Maxwell's equations and to keep terms up to
first order in this small parameter. Following the pioneering work of Feibel-
man (1975a, 1982), the long-wavelength limit was discussed by many authors
(Mukhopadhyay, Lundqvist, 1977; Bagchi et al., 1979; Brodskii, Urbakh, 1980;
Sipe, 1980; Abeles, L6pez-Rios, 1980; Apell, 1981; Maniv, Metiu, 1980; Ger-
hardts, Kempa, 1984; Langreth, 1989; Schaich, Chen, 1989).
In this section, we derive the reflection coefficients for p- and s-polarized
light incident on a jellium surface. The essential idea is to replace the classical
matching conditions by integrating Maxwell's equations over the surface region
(Apell, 1981; Feibelman, 1982). In the long-wavelength limit, we find that the
surface corrections to the classical reflection amplitudes are given by the same
surface parameters d.L(w) and dll(w), which we encountered in the discussion
of the small-q region of the surface plasmon dispersion. We assume at first
4.2. MICROSCOPIC THEORY 151

the frequency to lie below the bulk plasma frequency. The extension to higher
frequencies will be addressed later.

P-Polarization

Let us define the x - Z plane as plane of incidence. Taking into account the
translational invariance parallel to the surface, all field components propagate
like eiqzx . Far from the surface, the z-dependence of a p-polarized wave is given
by (for brevity we omit the frequency and momentum arguments of the fields)

Ez(z) qx (e-iqzz + r eiqzz) , z ~ Z2 (4.19)


q
Ex(z) _ z
q · _ r e,qzZ)
(e-zqz;Z . , "
q
Ez(z) qx t e-iq~z z ::; Zl (4.20)
q' '
I
qz t e-iq~z
Ex(z)
q'
"

The amplitude of the incident wave is taken to be unity. The quantities rand
t are the reflection and transmission coefficients. The transverseness of the
incident, reflected, and transmitted waves determines the relative weights of
the x- and z-components. For small q, the electric field components at the
asymptotic points Zl and Z2 may be expanded as
qx
Ez(Z2) [1 +r iqzZ2 (1 - r)] (4.21)
q
qz
E x(Z2)
q
[1- r iqzz2 (1 + r)]
Ez (Zl) q", t (1 - iq~Zl) , (4.22)
q'
Ex (Zl) q~ t (1 - iq~Zl) .
q'
To determine the reflection and transmission amplitudes, we now derive two
conditions by integrating Ez(z) and Dx(z) over the surface region.
• First, from (4.6) we obtain

( 4.23)

which leads to
(4.24)
152 CHAPTER 4. NONLOCAL OPTICS

This expression replaces the usual boundary condition 'Dz continuous' of Fres-
nel optics. We may rewrite this condition as

(4.25)

The integral over D", is identical to the surface field integral III which we
encountered in the derivation of the long-wavelength behavior of the surface
plasmon (see (3.41)). Substituting expressions (4.21) and (4.22) and using the
long-wavelength relation qz(l- r) = q~tlve (see below), we find
c-1
1 + r = y€ t (1 - a) , a == iq~-- dll ' (4.26)
c
where d ll is defined in (3.43).
• The second condition needed to determine rand t can be obtained by
integrating the z-component of the wave equation (4.11):

(4.27)

This relation replaces the standard boundary condition 'Ex continuous'. The
integral over Ez corresponds to h, the other important surface field integral
determining the long-wavelength behavior of the surface plasmon. Substituting
(4.21) and (4.22) and using the long-wavelength relation 1 + r = vet, (4.27)
reduces to: , 2
qz (1- r) = qz t - i q", (c -l)d.L t , (4.28)
q q' q'
so that:
q~ t
1 - r = - - (1 - b) , (4.29)
qz y€
with d.L defined in (3.42) .
• Eliminatingt from (4.26) and (4.29), we obtain the reflection amplitude
for p-polarized light:
cqz(l - a) - q~(l - b)
cqz(l - a) + q~(l - b)

~ r:(w) {1+2i qz [(d.L- dll ) 2/2_


cqz qx 1
-dill} (4.30)

where r:(w) is the Fresnel amplitude specified in (4.1). At the bulk plasma
frequency, c = 0, indicating that the surface corrections associated with d.L
become negligible.
4.2. MICROSCOPIC THEORY 153

S-Polarization

The reflection amplitude for s-polarized light can be derived in an analogous


manner. Far from the surface, the fields are given by

Ey(z) = e- iqzz +r
eiqzz , (4.31)
H.,(z) _q z · _ r e,qzZ)
(e-,qzZ . , "
q
t e-'qzz ,
• f

Ey(z) (4.32)
,
H.,(z) qz t e-iq~z
"
q

Here, we used the relation

d iw
-d Ey(z) = --H.,(z) , (4.33)
z c
which follows from (4.5). In the long-wavelength limit, these field components
may be expanded at the asymptotic points Zl and Z2 as

Ey(Z2) l+r - iqzZ2 (l-r), (4.34)


H,,(Z2) qz [l-r - iqz z2(1+r)]
q
Ey(zt} t (1 - iq~Zl) , (4.35)

H.,(Zl) q~ t (1 - iq~Zl) .
q

To derive the two conditions needed to determine rand t, we now integrate


H., and Dy over the surface region.
• First, integrating (4.33) we find

(4.36)

which replaces the usual boundary condition 'Ey continuous' of Fresnel optics.
Substituting (4.34) and (4.35), this condition takes the form

(4.37)

• The second condition is found from the z-component of 'V x E and the
v-component of 'V x il which may be written as

d iq~ .
-d H"(z) - -Ey(z) = -zqDy(z) . (4.38)
z q
154 CHAPTER 4. NONLOCAL OPTICS

Integration over the surface region gives

Hz(z2} = Hz(Zl} + E,,(Zl} [i:~ (Z2 - zd - iq f.~2 dz ~:(~~] , (4.39)

instead of the standard boundary condition 'Hz continuous'. Substituting


(4.34) and (4.35), we obtain

(4.40)

Here, we used the definition of dll given in (3.43) and the fact that, for jellium
surfaces, dz = d" = dll (Bagchi, 1977) .
• Comparing (4.37) and (4.40), we notice that, to lowest order in q, the
terms in the square brackets vanish. To first order we have
1 +r t, (4.41)
iq2
1- r = q~ t (1 - a') , a' == q~ (e - 1) dll . (4.42)
qz
Eliminating t we obtain the reflection amplitude for s-polarized light:
qz - q~ (1 - a') F( } ( . d)
r. ()
W = ( ) ~ r. W 1 - 2zqz II , (4.43)
qz + q~ 1- a'
where r;(w} is the Fresnel amplitude defined in (4.2).
The expressions (4.30) and (4.43) are generalizations of the classical Fresnel
formulas. They show that, in the long-wavelength limit, the influence of the
electron distribution at the surface and its nonlocal response properties are
fully characterized by the functions d.L(w} and dll(w} (Feibelman, 1975a).

Above the Bulk Plasma Frequency


So far, we have considered frequencies below the bulk plasma frequency.
In this range, there are no asymptotic longitudinal fields, i.e., the ansatz of
only transverse waves far from the surface is correct. Above Wp , this ansatz
remains valid for s-polarized light. For p-polarized light, on the other hand,
the asymptotic expression (4.20) must be generalized since the surface can give
rise to ail. interference between transverse and longitudinal waves. For z ~ 0
we may write
qz t e-iq~z
Ez(z}
q'
+ , e -iQ • z , (4.44)
,
Ez(z} = qz t e-iq~z , ~z e- iQ• z . (4.45)
q'
4.2. MICROSCOPIC THEORY 155

Here, Q= (q"" 0, -Qz) is the wave vector of the internal longitudinal wave. It
is determined by the condition for the existence of bulk plasmons:

(4.46)

where CL is the longitudinal bulk dielectric function. The relation between


the x- and z-amplitudes of the longitudinal part of E(z) follows from the
longitudinal character of the wave.
Except near wp , the component Qz of the plasmon wave vector is much
larger than that of the transverse wave. If we include the damping of the bulk
plasmon, Qz is complex and the plasmon wave decays towards the interior of
the metal. We may therefore choose the boundary Zl of the asymptotic region
sufficiently far from the surface so that only the transverse wave survives.
Thus, e- iQ ,Zl -+ 0 for Zl «: o. The expansions appropriate in the long-
wavelength limit, (4.21) and (4.22), remain valid, and the reflection amplitude
for p-polarized light is again given by (4.30). The functions dJ..(w) and dll(w)
are defined as before, except that the integration limit Zl must be chosen far
enough inside the metal so that the longitudinal plasmon is fully included.
The decay of the bulk plasmon, which is present in all real situations, ensures
that dJ.. is well defined.
The numerical evaluation of dJ..(w) above wp is nevertheless non-trivial since
the plasmon field decays rather slowly. As discussed later, this problem can
be avoided by using a sum rule that relates the internal contribution to dJ..
(the integration range Z ::; 0) to the external part (z ;:: 0). Hence, the difficult
evaluation of the moments of the slowly decaying field components (Friedel
oscillations below wp , Friedel oscillations and bulk plasmons above wp ) is cir-
cumvented.

4.2.2 d~(w) and dll(w)

Because of the general importance of the d-functions for long-wavelength


excitations at metal surfaces, we summarize here various equivalent expressions
to illustrate their physical meaning. As we saw in Chapter 3 and the previous
section, the fundamental quantities describing surface effects on macroscopic
observables are the surface field integrals

Z2 Ez(z)
h(w) = /. dz - (-) , (4.47)
%1 Ez Zl
156 CHAPTER 4. NONLOCAL OPTICS

In the Fresnel picture, Ez and D., are constant, except for the abrupt jump at
the surface, z = O. From the relations

constant = eEz(zr) = E z (Z2) , (4.48)


= constant = D.,(zl)/e = D.,(Z2) , (4.49)

it follows that

(4.50)

d.l.(w) (4.51)

(4.52)

Since the integrands in the expressions for d.l. and dll are finite only close to
the surface, the integration range can be taken as ±oo. According to these
relations, d.l.(w) and dll(w) specify the corrections to the classical field integrals
over E z and D., due to surface effects. These include the coupling between
transverse and longitudinal fields induced by the surface.
• From Gauss's law, it is evident that d.l.(w) can also be viewed as the
centroid of the charge density induced by a uniform electric field oriented
normal to the surface:

d.l.(w) = I dzznl(Z,W) / I dznl(Z,W) . (4.53)

Furthermore, using (4.7) one can show that

(4.54)

(4.55)

Thus, the real parts of the d-functions measure the location of the charge and
current densities induced at the surface. As seen later, the imaginary parts
specify the power absorption .
• It is also possible to express the d-parameters in terms of tensor compo-
nents of the nonlocal dielectric function:

(4.56)

(4.57)
4.2. MICROSCOPIC THEORY 157

where the functions E:-;}(z,w) and E:",,,,(z,w) are given by

Ez(z)/Dz(Zl) = Jdz' E:-;}(z, z',w) , (4.58)

E:",,,,(z,w) D"(z)/E.,(Zl) = Jdz' E:",.,(z, z',w) . (4.59)

The inverse dielectric function E:-;zl(Z,Z',w) is defined by the relation

Jd Z /I -l( z, z /I ,w ) E: zz (1/'
E: zz Z ,z, w) = u1:( Z - Z') . (4.60)

• Other useful relations for d.J..(w) and dll(w) may be derived in terms of
the polarization induced by the external field. In the long-wavelength limit,

Dz(z) constant = Ez(z) + 47r Pz(z) , (4.61)


E",(z) constant = D",(z) - 47r P",(z) . (4.62)

The polarization components induced in the bulk are


_ F E:-1
P;(Zl) = Pi = - - E;(Zl) , i = x,z. (4.63)
47r
Inserting these relations into (4.51) and (4.52), we find

(4.64)

(4.65)

According to these expressions, the d-functions specify the normal and paral-
lel components of the integrated surface polarization (normalized by the bulk
polarization). In the classical picture, the polarization has the constant com-
ponents P; and P: up to the surface plane, so that d-L(w) = dll(w) = O.
Notice that some of the expressions given above, e.g., (4.53) and (4.56),
are formally divergent at frequencies above wp if the bulk plasmon is assumed
to be undamped. These expressions, however, remain applicable once bulk
damping is included. The plasmon may then be regarded as part of the surface
excitation spectrum if the boundary of the asymptotic bulk region is placed
sufficiently deep inside the metal.
The expression for d-L in terms of the induced surface polarization is il-
lustrated schematically in Figure 4.3. Note that Red-L is not a measure of
the finite spatial extent of the surface polarization nor of the width of the
induced charge density caused by the diffuseness of the ground-state density
158 CHAPTER 4. NONLOCAL OPTICS

~
1.0
~
rL'
'"'-
3'
0.5
~
rL'

0.0
-40 -20 o
z (co)
Figure 4.3: Schematic illustration of microscopic and classical po-
larization P. near metal surface. Plotted is the normalized surface
polarization Pz(z,w)/pnw) (solid curve). The Fresnel limit is in-
dicated by the dashed curve. The induced density nl(z, w) is related
to Pz(z,w) via P: = nl. The parameter d.L(w), i.e., the centroid of
nl, is determined by the difference between the areas A and B.

profile. Instead, Red.L specifies the effective location of the fluctuating sur-
face charge. It provides information on the efficiency of screening at metal
surfaces. Re d.L > 0 tells us that fields are well screened in the low-density
tail of the ground-state electron distribution. Conversely, Re d.L < 0 indi-
cates that high-density regions are involved in the screening processes. The
diffuseness of the density profile, on the other hand, influences the probabil-
ity of generating electron-hole pairs in the surface region. These properties
determine the imaginary part of d.L.
Analogous considerations can be made for the function dll(w). For jellium
surfaces, dll becomes particularly simple since, in the long-wavelength limit,
(4.59) simplifies to (Feibelman, 1982)

41T no(z)
czz(z,w) = 1- 2 (4.66)
W

This result can easily be proven by integrating IY:r;:r; in (2.115) over f" and using
4.2. MICROSCOPIC THEORY 159

the orthogonality of the single-particle states. We then have

dll(W) = zil = f dz [no(z) - 6( -z)] , (4.67)

where zll is defined in (2.16). This result shows that dll (w) is real and frequency-
independent. It specifies the surface position as the centroid of the derivative
of the ground-state density profile. If the location of the surface plane is
taken to be the jellium edge, then dll(w) = 0 because of charge neutrality.
A frequency dependence in dll(w) can arise only from a modification of the
parallel displacement field due to lattice effects.

Extension to N on-J ellium Systems

Until now, we discussed the interaction of light with jellium systems. De-
viations from the macroscopic Fresnel fields are confined to the near-surface
region and the direction normal to the surface. As shown by several authors
(Del Sole, 1981; Mochan et al., 1983; Del Sole, Fiorino, 1984; Mochan, Bar-
rera, 1985; Langreth, 1989; Schaich, Chen, 1989), the d-parameter concept
can be suitably generalized to realistic crystal surfaces. The important result
is that the expressions for the p- and s-polarized reflection coefficients, (4.30)
and (4.43), remain valid, except that dll in Tp and T. is replaced by d", and dy ,
respectively.
In general, the crystal structure can give also rise to non-diagonal tensor
elements of the dielectric function. Inspection of formulas (4.56) and (4.57)
suggests the appearance of analogous non-diagonal d-functions of the type d",y,
d",z, etc. Assuming time-reversal invariance, it can be shown that such terms
lead to corrections to Tp and T. that are second order in qL (Del Sole, 1981;
Schaich, Chen, 1989). In the long-wavelength limit, it is therefore adequate
to consider the diagonal functions di(w} (i = x, y, z) as long as a pure p- or
s-polarized wave is incident. However, if mixed p- and s-polarizations are em-
ployed, as, for example, in ellipsometric measurements, first-order corrections
involving non-diagonal d;j elements must also be taken into account.
Langreth (1989) and Schaich and Chen (1989) derived expressions for the
d; that involve spatial averages of the induced surface polarization and that can
be viewed as generalizations of the formulas (4.64) and (4.65). The important
new feature is the average over the three-dimensional unit cell which smooths
the rapid oscillations of the microscopic polarization induced by the lattice.
Calculations of the optical response of realistic systems so far have been
performed mainly for semiconductors (for an excellent recent review, see Del
Sole, 1995). Since screening in these systems is much less important than in
160 CHAPTER 4. NONLOCAL OPTICS

metals, it is usually sufficient to remain within the one-electron approxima-


tion. Exceptions are many-body effects, such as excitons and local field effects
related to fluctuations of the electric field on the scale of the lattice. In addi-
tion, electron-hole interactions may distort the line shape above the band gap
(Hanke, Sham, 1980).
Surface local field effects in semiconductors have also been studied within
a point-dipole lattice model (Mochan, Barrera, 1985; Chen, Schaich, 1989;
Wijers, Poppe, 1992; Schaich, Wijers, 1995). Recent extensions of this model
to noble metal surfaces are discussed in Section 4.5. Microscopic calculations of
the optical response of metal surfaces, including lattice effects, were performed
for simple metals (Lee, Schaich, 1991a; Burke, Schaich, 1993, 1994; see Section
4.4) and for alkali metal overlayers (Ishida, Liebsch, 1992; see Section 4.7).

4.2.3 Spectroscopic Quantities

The expressions for the reflection amplitudes of p- and s-polarized light are
important for several optical surface spectroscopies. Moreover, the functions
dJ.(w} and dU(w} playa key role in other long-wavelength phenomena at sur-
faces, such as the dispersion of surface plasmons at small q (see Chapter 3), the
atom-metal van der Waals attraction (Chapter 6) and several low-frequency
excitation processes (Chapter 7).

Differential Reflection Spectroscopy


To separate the surface contribution to the reflectivity, it is useful to extract
the Fresnel amplitudes from (4.30) and (4.43). Thus,
IrpI2 Ir:12 [l-4q z ImDJ.(w}] , (4.68)
Ir.1 2 = Ir;12 [1 + 4qz Imdu(w}] , (4.69)
where r: and r; are given in (4.1) and (4.2). The quantity DJ. is defined as

DJ.(w} = (dJ. - du) 2ce 1 - dU ,


ccot -
(4.70)

with e the polar angle of incidence. The differential reflectivities may be writ-
ten as
D..Rp IrpI2 - Ir:12
= = -4qz ImDJ.(w} , (4.71)
~ Ir:12
D..R. Ir.1 2 - Ir;12
RF = = 4qz ImdU(w} . (4.72)
• Irfl2
4.2. MICROSCOPIC THEORY 161

For clean metal surfaces, corrections to the classical Fresnel reflectivities are
usually very small. Besides, it is difficult to identify them because of the in-
evitable uncertainties in the bulk dielectric function. Surface contributions
can, however, be observed in anisotropy spectra (see Section 4.5) and elec-
troreflectance measurements. In the latter case, the surface electronic profile
is modified by the strong static electric field at the electrode-electrolyte inter-
face (Section 4.6). Surface contributions induced upon adsorption of metallic
species are also readily observable (Section 4.7).

Ellipsometry

The ratio p = rp/rs can be measured directly in surface ellipsometry. From


(4.30) and (4.43) we find

p = pF [1 + 2iqz (dl. - dll) c


2 (J 1]' (4.73)
ccot -
where pF = r: /r; is the classical value obtained from (4.1) and (4.2). This
expression is equivalent to the result derived by Abeles and Lopez-Rios (1980).
As pointed out by these authors, only the difference dl. - dll can be extracted
from the data. At the plasma frequency, c(w) -+ 0, so that p -+ pF.

Surface Plasmon Polaritons

As mentioned in Chapter 3, the poles of the reflection amplitude for p-


polarized light determine the wave vector dispersion of surface plasmon po-
laritons. Such poles indicate the existence of self-sustained modes. From the
roots of the denominator in the first expression for rp in (4.30), we find the
condition (Sipe, 1980; Apell, 1981)

(4.74)

Since this mode is localized at the surface, the associated fields are evanescent
in the z-direction. We may therefore write qz = i", and q~ = i",', where", and
",' are defined in (3.62). Substituting these parameters, we obtain

(4.75)

In the classical limit, this condition reduces to ",' + c'" = 0, which yields the
standard surface plasmon polariton dispersion given in (3.65). As illustrated
162 CHAPTER 4. NONLOCAL OPTICS

in Figure 3.20, nonlocal corrections lead to a red (blue) shift of the frequency
relative to the c1assicallimit if Re (d.L - dll) > 0 « 0).

Surface Response Function

Consider the reflection amplitude for p-polarized light in the nonretarded


limit by letting c --* 00. The condition q = q' = 0 can be satisfied only if
qz = q~ = iqx. From (4.26) and (4.29) it follows that

a -qx(e - 1) dille, (4.76)


b q",(c - 1) dl. . (4.77)

Inserting these expressions into (4.30) we find

+ q;z;(c - + dl.)
( 1 + 2q", edl. +1dll )
c -1 l)(d ll e-1
e + 1 + q;z;(c - l)(dll - dl.) c+1 c+
g(q""w) . (4.78)

Hence, in the nonretarded limit, the reflection coefficient for p-polarized light
reduces to the surface response function g(q""w) at small q"" (3.54).

Surface Photoyield

Absorption of photons at metal surfaces via excitation of electron-hole


pairs is another quantity that can be formulated in terms of d-parameters
(Feibelman, 1982). In the absence of bulk absorption processes, Le., for real
c(w), the surface power absorption is given by (Jackson, 1962)

P(w) = -1
2
Re 12,
Z2
dz j(z)
- . E(z)*
- . (4.79)

Using (4.13) - (4.17), we can express the current as

- iw - - (4.80)
j(z) = 41T (E - D) ,

so that
P(w) = -W 1m 122 dz D(z)
- . E(z)*
- . (4.81)
81T z,
4.2. MICROSCOPIC THEORY 163

In the long-wavelength limit, E., and D" can be regarded as independent of z


and may be removed from the integral:

ImlZ2 dz D",(z)E",(z)* 1m III IE.,( zt} 12 , (4.82)


'"
1m 1"2%, dz Dz(z)Ez(z)* Imh IDz (Zl)1 2 /6" . (4.83)

Thus, the power absorption is determined by the surface integrals h and III.
Using (3.43) and (3.42), we may express these in terms of dll(w) and d.LCw).
The absorption then ta.l,{es the form

Dividing by the incident flux

F(w) = ~ IEil2 cos () , (4.85)


871"
we obtain for the total yield Yew) = P(w)/F(w):

Yew) = ~() (6 -1) [1~",12 Imdll(w) + ~ l~zl2 Imd.L(W)] (4.86)


eGOS IE.12 6" IEil2

Here, () is the polar angle of incidence and Es the incident field vector. At
() = 45° and for 1m dll = 0, this expression reduces to

(4.87)

The yield is seen to vanish at the bulk plasma frequency, indicating that
the surface becomes invisible when a volume plasmon of infinite wavelength
is excited. Of course, yew) goes smoothly through the region of the surface
plasmon. Once retardation is included, the surface plasmon dispersion always
lies below the light line, as shown in Figure 3.20. Surface plasmon excitation
via photons is therefore forbidden. However, coupling to surface plasmons
can be induced via new sources of parallel momentum, for example, through
gratings or surface roughness.
Note that the total surface absorption consists of two contributions: exter-
nal absorption, which can, for example, be measured in surface photoemission,
and internal absorption. The former exists only at frequencies above the vac-
uum threshold and corresponds to final electron states that propagate to the
vacuum. The internal part exists at all frequencies and corresponds to final
164 CHAPTER 4. NONLOCAL OPTICS

states that propagate towards the interior of the metal. The expression (4.87)
represents the sum of both contributions and cannot be compared directly to
the photoemitted yield.

Golden Rule Expression

For the purpose of determining explicitly the excitation channels that con-
tribute to the total power absorption at metal surfaces, it is useful to rewrite
Imd1.(w) in terms of the golden rule formula (Lee, Schaich, 1991b; Liebsch
et al., 1994). We start with the expression (2.45) for the rate of generating
electronic surface excitations. In the limit of small q, this formula simplifies
considerably because of the two-dimensional translational symmetry parallel
to the surface. The three-dimensional integrals over occupied and unoccupied
states can be reduced to a single one-dimensional integral over occupied states:

(4.88)

where k~ = (k~ + 2W)1/2 and h(k%) = min (2w,ki - k~). The sum over f
denotes possible final-state channels: Below the vacuum level, only excitation
to standing waves bound by the surface barrier is allowed. Above the photo-
emission threshold, however, transitions to states that propagate towards the
vacuum and the interior of the metal must be included.
With the small-q expansion of the surface response function g(q,w) given
in (3.53), and using (2.64), we can express the rate in terms of Imd1.(w):

210(10 - 1)
w(w) = 47r ( )2 Imd1.(w). (4.89)
10+1

From (4.88) we finally obtain

1m d1. (w) = 2(10(+_1)2


ce: 1)
~3
7r
"lo
L.J
f 0
kF
dk% h(k%)
k'%
1(k'Z I'" Ik) 12
'I'scf Z •
(4.90)

This formula allows a convenient identification of electronic transitions con-


tributing to the external and internal excitation channels. Notice that this
expression does not vanish at the surface plasma frequency since ifJ"cf rv (10 -
1) / (10 + 1). This term cancels the corresponding prefactor.
4.2. MICROSCOPIC THEORY 165

4.2.4 Solution of Response Equation

For jellium surfaces, dll is trivially related to the equilibrium density profile
(see (4.67)). On the other hand, the numerical evaluation of d.L(w) is more
subtle than that of the surface response function g(q, w) because of the long-
range Coulomb potential in the q = 0 limit. There are now three schemes that
give nearly identical results for d.L (w) in those frequency ranges where they
overlap.
• Feibelman (1982) solved the nonlocal real-space response equation for
the electric field

(4.91)

which follows from the constitutive relation (4.13) for q = O. Z2» 0 de-
notes the boundary of the asymptotic region in the vacuum. The ground-state
electronic properties were evaluated within the LDA using the Lang-Kohn
model (1970) for semi-infinite jellium surfaces. The finite-frequency response
was treated within the RPA. To deal with the slowly decaying Friedel oscil-
lations of the induced field, a fitting procedure of these tails was employed.
Stable solutions were found in the frequency range 0.6 -1.2wp for several bulk
densities.
• Liebsch (1987) solved directly for the induced density via the real-space
response equation

nl(Z'w) = / dz' Xl(Z,Z',w) ¢>scf(Z',W) . (4.92)

Here, ¢>scf(Z,W) is the self-consistent potential generated by a uniform electric


field oriented perpendicular to the surface. Xl (z, z', w) is the q = 0 single-
particle susceptibility specified in (2.111). The ground state of semi-infinite
jellium surfaces was treated within the LDA and the dynamical response within
the TDLDA. The numerical problem caused by Friedel oscillations was circum-
vented via a sum rule for d.L(w). This procedure yields stable solutions down
to the adiabatic limit.
• Kempa and Schaich (1988) also performed TDLDA calculations of
nl(z,w) for semi-infinite jellium surfaces. Instead of working in real space,
they solved the response equation (4.92) in Fourier space. To ensure stability,
singularities of the response kernel caused by Friedel oscillations were dealt
with analytically before final matrix inversion. This method is a refinement
of the Fourier transform approach developed by Gies et al. (1987) for jellium
slabs. These authors introduced a damping parameter to avoid the problem of
166 CHAPTER 4. NONLOCAL OPTICS

oct 1

'"....
I

..
'"
a:
o~----------------~~---;

-1

oct 2

-0
I

"'.... 1

...E

W/Wp

Figure 4.4: Frequency dependence of d.l(w) for r. = 3 as calculated


within the RPA using different computational procedures. + sym-
bols: Feibelman (1982); solid curves: Liebsch (1987); dots: Kempa
and Schaich (1988); dashed curves: Gies and Gerhardts (1987).
Note the sign of Red.l. (Kempa et al., 1988).

Friedel oscillations and to smooth out discrete transitions between slab levels
below the emission threshold.
An example of the frequency dependence of d.l(w), obtained within these
schemes, is given in Figure 4.4. Although the computational procedures differ
entirely, the agreement is excellent. The sole exception are the results by
Gies et al.. Presumably, these differ from the others because of the damping
paramet~r (, = 0.05 wp ) and the slab geometry.

Reformulation of Response Equation

The real-space formulation by Liebsch (1987) is rather straightforward and


computationally not demanding. Moreover, it can easily be applied to more
4.2. MICROSCOPIC THEORY 167

complicated surface density profiles. An outline of the main steps is given


below. Since we are interested in the density response to a time-varying uni-
form electric field normal to the surface, we may use as applied potential
¢>ext = -27rz. This normalization corresponds to the q -+ 0 limit of the prob-
lem of inelastic electron scattering treated in Chapter 3. According to (2.46)
and (2.52), the response equation (4.92) takes the form

nl (z, w) !
= dz' Xl(Z, z', w) [¢>ext(z', w) + ¢>est(z', w) + ¢>xc(z', w)] (4.93)

The induced electrostatic potential is given by

(4.94)

In the adiabatic version of the TDLDA, the induced exchange-correlation con-


tribution is derived from the ground-state exchange-correlation functional via

(4.95)

With these definitions, we may rewrite (4.93) as

nl(z, w) = ! dz' (z, z', w) (z', w)


Xl ¢>ext

+ !dz' !dz" Xl(Z, z", w) K(z", z') nl(z', w), (4.96)

where
K(z, z') = -27rIZ - z'l + V:c[no(z)] o(z - z') . (4.97)
A direct solution of this equation via matrix inversion, as for finite q, is not
feasible because of the long-range Coulomb interaction. In order to achieve
stable results, three key steps are taken:
• First, it is important to account explicitly for the asymptotic behavior
of the Coulomb potential before inverting the response equation. Let us define
the integrated weight of the induced density as a(w). From the continuity
of the displacement field D z, we obtain c Ez(Zl) = Ez(Z2)' where Ez(Zl) =
-27r (1 - a) and E z (Z2) = -27r (1 + a). Hence,

a(w) = !dz nl(z,w) = c(w)-l


()
+ .
c w 1
(4.98)

We now separate from the total Coulomb potential the long-range bulk part
¢>b(Z) = -27r z(l - a) = -47r zj(c + 1) and write
(4.99)
168 CHAPTER 4. NONLOCAL OPTICS

The surface component is given by

rPl(Z,W) = -411" too dz' (z - z') nl(z',w) . (4.100)

According to this definition, rPl vanishes for Z « 0, Le., this potential is finite
only in the surface region. In the vacuum, rPl has the asymptotic form

rPl(Z,W) = -411" u(w) [z - d.L(w)] , (4.101)

where d.L(w) is the centroid of the induced charge density. The response
equation (4.96) can now be reformulated as

nl(z, w) = nb(z,w) + Jdz' Jdz" Xl(Z, z", w) k(z", z') nl(z', w), (4.102)

with

nb(Z,w) (4.103)
~l(Z,W) Jdz' Xl(Z,Z',w) rPext(z',w) , (4.104)
K(z, z') = -411"(z - z') (J(z - z') + V:C[no(z)] c5(z - z'). (4.105)
This interaction kernel vanishes towards the interior of the metal. The long-
range bulk response is fully accounted for via the function nb (z, w).
• Second, the numerical evaluation of the induced dipole moment is
greatly facilitated by using the dynamical force sum rule. This rule establishes
a relation between the internal and external contributions to the dipole mo-
ment. This implies that the centroid d.L is fully determined by the tail of the
induced density extending beyond the jellium edge:

d.L(w) = - (1-)
u w
1 dzznl(Z,W)
00

-00
=
e~)+lLoo
()
ew 0
dzznl(Z,W) . (4.106)

This sum rule is derived in the next section.


• Third, although it is possible to solve (4.102) by direct matrix inversion
(Schaich, 1994a), greater numerical stability is achieved by eliminating the
long-range Coulomb response kernel altogether and introducing a short-range
interaction. This can be done by rewriting the Poisson equation

(4.107)

in the form

rPl(Z,W) = Jdz' e-"lz-z'l [2; nl(z',w) + i rPl(Z',W)] . (4.108)


4.2. MICROSCOPIC THEORY 169

Differentiation shows that this equation is equivalent to (4.107), independently


of the value of K, (Manninen et al., 1975). As in the evaluation of the ground-
state density (see Section 2.3.2), this conversion is extremely useful for two
reasons: It provides a numerically well-behaved response kernel and permits
explicit incorporation of the asymptotic behavior of the electric potential in
the vacuum and in the bulk.

Computational Procedure

To implement the asymptotic properties, it is convenient to define a trial


surface density n~ with integrated weight a and centroid do. The corresponding
potential is 4>~. For example, we may define

n~(z,w) f+(w)F+(z) + f_(w}F_(z) , (4.109)


4>~(z, w) -411" iZoodz'(z-z'}n~(z"w}. (4.110)

The weights are given by J± = !a(1 ± do/z o), and F±(z} are normalized
Gaussian distributions centered at ±zo:

(4.111)

Separating n~ from (4.102), we obtain the following response equation for the
remaining density nt = nl - n~ :

f
nt(z,w) = A1(z,w) + dz' Xl(Z,Z',w) [4>t(z',w) + V:c(z') nt(z',w)] , (4.112)
where the 'driving' term is

A1(z,w} = nb(z, w} f
+ dz' Xl(Z, z', w) [4>~(z' ,w) + V:c(z') n~(z' ,w}]
- n~(z,w} . (4.113)

The functions nt and 4>t satisfy a Poisson equation like (4.108). At self-
consistency, the trial density n~ has the correct centroid dl.; by construction, it
has the exact integrated weight a. Since the potential 4>~ then has the correct
asymptotic form, the remaining piece 4>t is finite only close to the surface. We
now have a response problem for two functions, nt and 4>t, that are indeed
fully localized at the surface. We can therefore solve (4.112) and (4.108) (for
170 CHAPTER 4. NONLOCAL OPTICS

nt and ¢>D via matrix inversion. Discretizing all functions on a mesh of points
Zl'" ZN, we combine these equations into a 2N x 2N matrix equation:

(4.114)

Here, 1 is the unit matrix, Wij are integration weight factors, Xl denotes the
matrix Xl(Z;,Zj,w), and K is defined as K;; = ~ e-I<lz,-zjl. The remaining
quantities are vectors such as ill(z;,w). The weights Wi; should be chosen so
that cusps in Xl and K at Zi = Zj are treated accurately.
Let us now define

dl(w) = f dz Z n~(z,w)/a(w) . (4.115)

The total centroid is dJ. = do + d l . Since, in general, this will not agree with
the centroid do of the trial density n~, we iterate the response equation (4.114)
until self-consistency is achieved, Le., until dJ. = do. In this case, the centroid
d l ofnt vanishes, so that the assumption of a fully surface-localized potential
¢>t is justified. The self-consistent density nl and its centroid dJ. are, as they
should be, independent of the parameter K. used in the short-range version of
the Poisson equation (4.108) and of the parameters r and Zo that specify the
width and position of the trial density n~.
Liebsch (1987) calculated d l for four trial densities n~, corresponding to
two real and two imaginary parts of do. For example: do = (±1, ±i). The four
output d l were then interpolated to find do such that dl = O. This value of do
specifies the self-consistent centroid dJ.. Both procedures, the iteration scheme
and the interpolation method, give stable solutions down to the adiabatic limit.
At w ~ 0.01 wP' the static image plane positions calculated by Lang and Kohn
(1973) were reproduced.
The surface range Zl ... ZN in which the induced density is evaluated is the
same as the one discussed in Section 2.3.5 for finite q. Thus, N = lOOn + 1,
with n = 1, is usually sufficient. Occasionally, n must be increased to 2 or 3 to
achieve stable solutions. As before, the wave functions and Green's functions
used to construct Xl(Z,Z',w) are evaluated on a finer mesh than the induced
density (see Section 2.3.1).
A crucial point is again the accurate evaluation of the driving term ill (Z, w)
defined in (4.113), in particular, the evaluation of the asymptotic contribution
to the unscreened density 6(z,w}, (4.104). Near the surface, Simpson inte-
gration is sufficient. In the interior, however, it is necessary to use (2.111)
and first carry out the z' integral. This can be done analytically, exploiting
4.2. MICROSCOPIC THEORY 171

the known asymptotic behavior of the bound states Wk(Z/) and Green's func-
tions G(Z,Z',ck ±w). The integration over occupied states is then performed
numerically.
The method just described is also applicable above wp since using the dy-
namical force sum rule eliminates the evaluation of the dipole moment in the
interior. Of course, the singular behavior of dJ. near wp can be handled only
by introducing a damping parameter.

4.2.5 Sum Rules

The function dJ.(w) satisfies several sum rules. First, according to the
Kramers-Kronig relations:

(4.116)

If ImdJ.(w) consisted of a single narrow peak at a resonance frequency W r ,


dJ.(O) and Wr would be inversely related. The Kramers-Kronig relation is useful
since it establishes a link between the frequency and strength of the multipole
surface plasmon and the initial slope of the monopole surface plasmon (see
(3.58)).
The so-called surface f-sum rule (Persson, Apell, 1983; Persson, Zaremba,
1984) specifies the first frequency moment as:

21
-
7r 0
00
dw W ImdJ.(w) = AW; , (4.117)

where A is the normalized ground-state density outside the jellium edge:

( 4.118)

This 'spill-out' parameter defines the distance up to which the external den-
sity would extend if it had the constant value n. It is remarkable that the
effective number of electrons contributing to the surface excitation spectrum
is determined by this spill-out density.
Kempa and Schaich (1989) showed that both sum rules are well satisfied
by TDLDA calculations of dJ.(w) below and above wp- Interestingly, nearly
independently of T., about 120 % of the spectral weight of ImdJ.(w)/w and
160 % of the weight of W 1m dJ. (w) lie below wp (Liebsch, 1987) [above wp,
ImdJ. < OJ. For this reason, the spill-out parameter A cannot be used to
represent the slope of the surface plasmon RedJ.(w.) (see also Section 3.3.5).
172 CHAPTER 4. NONLOCAL OPTICS

Dynamical Force Sum Rule

The dynamical force sum rule is very useful since it establishes a relation
between the total induced dipole moment and the external part of it that
depends only on the induced density outside the jellium edge. It is therefore
not necessary to calculate the moment in the interior where the induced density
exhibits slowly decaying Friedel oscillations. Because of the importance of this
sum rule for practical calculations of surface excitation spectra, we outline here
the derivation for semi-infinite systems (Liebsch, 1987).
The dynamical force sum rule (Sorbello, 1985; Epstein, Johnson, 1969)
makes use of the equation of motion for the electrons. The force due to the
total electric field acting on the electrons is given by

F(t) = Lm ( -ddt2z.)
i
'
2
= m-
d
dt
2
!dzzn(z,t)
2

!dz e n(z, t) E(z, t) , (4.119)

where Zi is the z-coordinate of the ith electron and n(z, t) is the electronic
density at time t. For clarity, we specify the electron's mass m and charge
e < O. In the case of a jellium surface, E(z, t) arises from the applied field
and the field due to the positive ions of the jellium background. The force
of the electronic density acting on itself vanishes. Converting again to atomic
units, we find to first order in the applied field

(4.120)

To keep track of the various bulk and surface contributions to this equation,
it is convenient to first apply the sum rule to a slab geometry and then go to
the limit of large thickness. Consider a slab placed symmetrically between two
oppositely charged capacitor plates. The slab center is taken to be at z = O.
The positive background density n+(z) = fi is assumed to extend from -L+
to L+ and the electronic density satisfies

10 00
dz no(z) = fi L . (4.121)

In general, the surface may be charged. Neutrality requires L). == L+ - L = o.


The induced electrostatic potential is given by

¢>est(z,w) = -27r ! dz' Iz - z'l nl(z',w) . (4.122)


4.2. MICROSCOPIC THEORY 173

The slab is assumed to be thick enough so that the total electrostatic potential
satisfies the boundary condition

c:(w) r//(Z = O,w) = r//(Z = oo,w) . (4.123)

If we take the external field to be Eext(z,w) = -4'][(To, with (To == c:/(c: + 1),
the integrated surface charge

(4.124)

has the same form (4.98) as for the semi-infinite surface.


To evaluate (4.120), we separate the surface terms from the volume terms
rv L. Since all integrands are even functions of z, it suffices to integrate over

the right half of the slab. The external-field contribution is

The contribution due to the positive ions may be written as

10 00
dz nl(z,w) E+(z) Io'XJ dz n+(z) rjJ~st(z,w)
= 4'][fi [p(w) - p+(w) + (T(w)L] (4.126)

where

p(w) 10 00
dz (z - L) nl(z,w) , (4.127)

p+(w) (00 dz (z - L+) nl(z, w) . (4.128)


lL+
The dipole moment on the left-hand side of (4.120) is separated as

(4.129)

It is easily shown that the volume contributions to (4.120) cancel. The


surface terms satisfy the condition

c:(w)-l
p(w) = c:(w) p+(w). (4.130)

This implies that the total dipole moment is fully determined by the density
in the region outside the positive background!
174 CHAPTER 4. NONLOCAL OPTICS

We now assume that, in the limit of large slab thickness, the density induced
at each slab side approaches that at a semi-infinite metal. With the change of

L:
variables from Z to Z - L, the moments in (4.130) are

p(w) dz znl(Z,W) , (4.131)

p+(w) £.00 dz (z -~) nl(z,w) . (4.132)

For neutral surfaces, ~ = 0, so that (4.130) yields the sum rule (4.106). For
positively charged surfaces, ~ > 0, giving a reduction of p+. If the induced
density is located fully within the positive charge background, p+ vanishes
and d.l.. == O. This result agrees with the one obtained by Schaich (1994b) for
parabolic quantum wells appropriate for strongly charged jellium slabs. The
relation (4.130) can easily be generalized to other profiles of the positive charge
background, for example, to the case of adsorbed alkali metal layers.
We conclude by pointing out that the sum rule (4.120) holds only if electron-
electron interactions in the ground state and in the presence of the applied field
are treated on the same footing. In the TDLDA, this condition is satisfied.
The LDA-based RPA, on the other hand, violates this sum rule since it cor-
responds to a time-dependent Hartree approach in which the ground-state
exchange-correlation potential acts like a rigid external potential. The force
due to this potential gives rise to an additional term

Pxc 1
= -2'
Wp
1-00. dz nl(Z,W) V:c[no(z)] ,
00
(4.133)

which must be added to p+ on the right-hand side of (4.130). The sum rule
(4.106) is then replaced by the modified version (Liebsch, 1987)

[10
d.l..(w) = c(w)( )+ 1
00
1
dz znl(Z,W) - 2' 1 00
dznl(Z,W) v,.c[no(z)]
, ]
. (4.134)
cwo wp -00

In this case, the density is determined from the response equation (4.93), but
the potential ifJxc is omitted.

4.3 Simple Metals


4.3.1 Surface Excitation Spectra

Figure 4.5 shows the frequency dependence of d.l..(w) for various bulk den-
sities. The following spectral features can be distinguished:
4.3. SIMPLE METALS 175

14
8 -LOA
LOA 12 ---- RPA
7 10
(b)

4
2
..., 0
~ 8
E
..., 2 6
"0
4

]
0
4
3
2
1
0
2

0
0 wlwp 1.0 0 wlwp 1.0

Figure 4.5: Frequency dependence of (a) real and (b) imaginary


parts of dl.(w) for several bulk densities. Solid curves: TDLDA;
dashed curves in (b): LDA-based RPA. The arrows denote the
work functions. (Liebsch, 1987).

• At low frequencies, the real part of dl.(w) approaches the static image
plane position as calculated within the LDA by Lang and Kohn (1973). As a
result of the efficient screening at metal surfaces, the image plane lies about 1 ao
outside the positive ionic background. The imaginary part of dl.(w) increases
linearly at low w:
~ w
Imdl.(w) = - - . (4.135)
kF wp
Since this function is proportional to the rate of exciting electron-hole pairs,
the linear slope ~ is relevant for a variety of phenomena involving low-frequency
excitations (see Chapter 7).
176 CHAPTER 4. NONLOCAL OPTICS

• As W approaches the work function, 1m d1. (w) increases appreciably.


This is not a density of states effect related to the flattening of the surface
barrier potential near the vacuum level. Instead, this threshold excitation is
caused by the screened dynamical potential that determines the strength of
matrix elements in the surface region (Ishida, Liebsch, 1992). This potential
is large only in the outer tails of the electronic density. Hence, transitions
involving occupied states near the Fermi energy and unoccupied states near
the vacuum level have the largest amplitude. In the case of low-density metals,
this spectral feature is dominated by the multipole surface plasmon. However,
the threshold excitation is quite pronounced in spectra of adsorbed alkali layers
(Section 4.6) and in nonlinear optical spectra of jellium surfaces (Section 5.3).
• As discussed in Chapter 3, the real and imaginary parts of d1.(w s )
determine the linear slopes of the frequency and width of the surface plasmon
at small q. Since Ws lies just below the multipole peak, i.e., in a region where
d1.(w) exhibits a near-singular behavior, it is clear that d1.(w.) should be rather
sensitive to the assumptions underlying the theoretical model.
• The main spectral feature in 1m d1. (w) corresponds to the multipole
surface plasmon near Wm ~ 0.8 wp. This excitation, which was first derived
microscopically within the LDA by Feibelman (1982) for several jellium sur-
faces, is the q = 0 counterpart of the multi pole mode predicted by Bennett
(1970) (see Figure 3.4). Qualitative arguments for the appearance of such
a feature in surface photo emission spectra were already given by Mackinson
(1937) who, like Bennett, considered a linear surface density profile. Assuming
a local dielectric function of the form e(Z, w) = 1 - 47rno(z )/w 2 and introduc-
ing nonlocal effects via V'. E "I 0, the surface photoyield was shown to exhibit
a maximum below wp.
Why the multi pole mode occurs for all jellium surfaces at nearly the same
relative frequency is not fully understood. Converting Wm into an effective local
density nm via w~ = 47rnm , it turns out that this density corresponds to the
point of the largest gradient of the ground-state density profile: Z ~ -0.3 ao
for r. = 2 and Z ~ -1.1 ao for r. = 5 (Liebsch, 1987). This 'inside' location
is consistent with the positive dispersion of the multipole surface plasmon,
in contrast to the negative dispersion of the ordinary surface plasmon whose
charge centroid is located outside the jellium edge.
In the RPA response treatment, the multipole mode frequency lies higher
than in the TDLDA. The reason for this shift is that the bare Coulomb po-
tential makes the density less polarizable (see also Figure 3.12).
• As the volume plasma frequency is approached, Red1.(w) diverges to
-00. This is to be expected since wp denotes the threshold at which the metal
becomes transparent to incident electromagnetic fields. Above wP ' Imd1.(w)
4.3. SIMPLE METALS 177

0.02 rs :4 (bl
0.2 W/Wp =0.1
0.00
0.0
0.2 0.02 OJ
0.00
0.0
0.04 0.5
0.2 0.00

0.0 0.04 0.6


~ 0.2 ~ 0.00
b
0
"- "-
"3 0.0 "3 0.04
..; N-

0.00
..
co co
>0 0.2 >0

~
C<
0.20
0.0
0.00
0.2

0.0 0.08

0.2 0.00

0.0 -0.08

-40 -30 -20 -10 -40 -30 -20 -10


Z (o.u.l Z (o.u.l

Figure 4.6: Spatial distribution of (a) real and (b) imaginary


parts of density induced by uniform electric field calculated within
TDLDA (T. = 4). The main short- and long-range Friedel oscilla-
tions are ).1 and ).2. Note the different scales in (b). The positive
background occupies the half-space z :s: o. (Liebsch, 1987).

corresponds to excitation of electron-hole pairs in the surface region and of


longitudinal bulk plasmons. Coupling to bulk plasmons is, however, consid-
erably weaker than to the multipole surface plasmon near 0.8 wp (Feibelman,
1982; Kempa, Schaich, 1989; see also Figure 4.10).

Induced Density

The spectral features seen in dot( w) are also manifested in the spatial dis-
tributions of the induced density. Figure 4.6 illustrates the real and imaginary
parts of nl(z, w)ja(w) at several frequencies. At low w, Renl(z, w) converges
to the static screening distribution obtained by Lang and Kohn (1973). Near
178 CHAPTER 4. NONLOCAL OPTICS

Wp , the first maximum of the real part is shifted inside, indicating that the
surface electrons are no longer able to scre!:ln the applied field. Aside from the
main surface peak, the induced density exhibits similar frequency-dependent
Friedel oscillations as the electric field (Feibelman, 1975a; see below). At low
w, the dominant period is given by Al = 21T/[k F + (k; + 2W)l/2] which in the
static limit coincides with the familiar value 1T/kF. Towards higher w, Al is
gradually replaced by A2 = 21TkF/W.
As shown by Kempa and Schaich (1988), close to the multipole frequency,
an important contribution to the oscillatory behavior of the induced density
stems from the wavelength A3 = 21T/[(2~V)1/2 - (2~V - 2W)l/2] , where ~V
is the height of the one-electron surface barrier. This period is related to
transitions from states below the Fermi energy to unoccupied states near the
vacuum level. Near 0.8wp , A2 and A3 are quite close. The amplitude of these
oscillations is then very large and the decay is rather slow. The multipole
charge oscillations are particularly prominent in the imaginary component of
the induced density.
Notice that the densities shown in Figure 4.6 are driven by a field of strength
-21T. Multiplying by the normalization factor u(w) = (1- w2/w:)-l, it is evi-
dent that nl diverges at w•. Thus the distribution at w/wp ~ 0.7 represents the
density fluctuation associated with the monopole surface plasmon in the long-
wavelength limit. The distribution of the multipole plasmon at Wm ~ 0.8 wp is
partly obscured by the driven response to the applied field. To determine the
density associated with the genuine dipolar eigenoscillation, we must find the
pole of the surface response function in the complex plane (see Section 3.3.1).
At real frequencies, there is little qualitative difference between induced den-
sities at w. and wm .

Local Electric Field

The spatial distribution of the induced density is closely related to that of


the perpendicular component of the local electric field. The microscopic behav-
ior of this field is illustrated in Figure 4.7 (Feibelman, 1982). Below the bulk
plasma frequency, the multipole surface plasmon appears as a surface feature
in the imaginary part of the field. The efficient creation of electron-hole pairs
near the multipole frequency is also referred to as 'local field enhancement'.
The real part of the field exhibits a more gradual variation with frequency.
Above wp , excitation of a longitudinal plasma wave appears as an extra sinu-
soidal (weakly evanescent) contribution.
4.3. SIMPLE METALS 179

-.
' ~
....
------\~

---.J W/Wp = 1.26

/\ ,'\ "
,-..,1,/:\'_/: \\ ../
'~ ~
------/\./ '. .. /

W/Wp = Ll8

o-----:'~

-.J W/Wp = 0_55

-10 o 10 20 :10 ~~ b I
20
z(A)

Figure 4.7: Spatial distribution of normal component of electric


field at r. = 2 jellium surface for various frequencies. Solid (dashed)
curves: real (imaginary) parts. The positive background occupies
the half-space z ~ O. Plotted is the quantity [Ez(z)/E; -ll/(l-c),
where E; is the Fresnel field in the bulk and Ez(z) is calculated
within the LDA-based RPA. (Feibelman, 1982).

4.3.2 Emission vs. Absorption

As shown in (4.87), ImdJ.(w) is proportional to the total photoabsorption.


The golden rule expression (4.90) permits an explicit decomposition into the
two possible final states that exist for W > <P: one corresponding to excitation
180 CHAPTER 4. NONLOCAL OPTICS

10 1 1 1

2 (a) r,=2 81- (b) rs==5

6f-

41- -

~/\S
2f-

1 J- 1_'
0
15 0 2 3 4

w (eY) w (eV)

Figure 4.8: Frequency dependence of Imd-L(w) for (a) r. = 2 and


(b) r. = 5, calculated within TDLDA. Solid curves: sum of internal
and external channels (see also Figure 4.5); dashed curves: exter-
nal channel. The arrows indicate the threshold for photoemission.
(Liebsch et al., 1994).

into the vacuum and the other giving an excited electron propagating towards
the back of the sample. Since surface photo emission probes only the first
channel, it is important to check the relative weights and spectral distributions
of these two processes.
Figure 4.8 compares the total Imd-L(w) with the emitted part (finite for
W > <II) for two jellium surfaces. The multipole surface plasmon represents the

main feature in all of these spectra, since it is an intrinsic surface excitation


whose frequency does not depend on the excitation channel. The relative
weight of this mode, however, varies considerably with r •. For AI, the emission
channel near Wm amounts to about two-thirds of the total absorption. Hence,
Imd-L(w) is a reasonable representation of the emitted yield. On the other
hand, in the case of K, the emitted part is only a small fraction of the total
absorption.

4.3.3 Comparison with Measured Spectra

The parameters d-L (w) and dll (w) are typically of the order of an electronic
screening length, i.e., much shorter than the wavelength oflight. Consequently,
4.3. SIMPLE METALS 181

300 400 500


" (nm)
Figure 4.9: Photo emission yield spectra of thick K films for p-
polarized (C:II) and s-polarized (c:1.) light. Dashed curves: yield for
rough samples. Ap denotes the bulk plasmon. (Monin, 1973).

non local corrections to the Fresnel formulas are usually much smaller than the
bulk contribution to the reflectivity. In the case of simple metals, it does
not seem possible to identify these corrections, in particular, in view of the
appreciable uncertainties in the bulk dielectric function c:( w). Surface excita-
tions at simple metal surfaces can, however, be observed directly in photoyield
measurements.
The first spectra that clearly revealed surface absorption are those on alkali
metals (Monin, 1973; Monin and Boutry, 1974) (see Figure 4.9). The yield for
K using p-polarized light shows a pronounced maximum near 400 nm (3.1 eV)
corresponding to about 0.82 wp. The data also exhibit a minimum at wp
(328 nm or 3.8 eV). The remaining intensity is presumably caused by bulk
absorption. The photoyield detected for s-polarized light is very small and
featureless. Similar spectra were observed on Rb and es.
Analogous surface absorption features were also observed on Al (Flod-
strom, Endriz, 1973; Levinson et al., 1979; Levinson, Plummer, 1981; Petersen,
Hagstrom, 1978), Mg (Gesell et al., 1973; Flodstrom, Endriz, 1975), In (Jeze-
quel, 1980), Be (Bartynski et al., 1985), and on adsorbed Na layers (Wallden,
1985). For a recent review on nonlocal surface corrections, the reader is re-
ferred to Lopez-Rlos (1992).
As shown by Feibelman (1982), these observations can be understood
182 CHAPTER 4. NONLOCAL OPTICS

, 4,----,---,---.---.---.---.---.---r----,
RPA Surfoce POloler Absorption vs r, .8, =45·
12
"0

'"
D
~

0
10
VI
D
0
VI 0.8
c
E
0
.c;
a. 0.6
~
C 04
'"
~

'"
D-
02

0
04 08 w/wp 10 12

Figure 4.10: Power absorption as a function of frequency for several


jellium surfaces, calculated within LDA-based RPA. (Feibelman,
1982).

within the dynamical response properties of jellium surfaces. Figure 4.10 illus-
trates the frequency dependence of the calculated power absorption for several
bulk densities. According to (4.87), these distributions are proportional to
Imd.L(w). Thus, the maxima near Wm ~ 0.8wp correspond to the excitation
of the multipole surface plasmon. The vanishing absorption near wp indi-
cates that, at the excitation of a bulk plasmon of infinite wavelength, surface
photoabsorption is negligible. The weak spectral weight above wp has contri-
butions from excitation of electron-hole pairs near the surface and creation of
bulk plasmons.
Figure 4.11 shows the comparison of Feibelman's calculations with mea-
sured surface photoyield data for Al (Levinson et al., 1979; Levinson, Plummer,
1981). Since bulk absorption below the bulk plasma frequency and above the
interband transition near 1.5 eV is weak, the power absorption is dominated
by surface excitations. Evidently, the feature near 12 eV is the multipole sur-
face plasmon. The dashed line illustrates the calculated power absorption in
the absence of surface screening processes, i.e., the photon field is assumed to
be spatially constant. The intensity then rises monotonically and neither the
peak near Wm nor the minimum at wp appear. An adequate understanding of
4.3. SIMPLE METALS 183

'"E ....
o

SURFACE STATE AT r
.
::J o
>, 20 &. 0 o CRYSTAL 1

~ .~
~ • CRYSTAL 2

.c
~
<l
10
.0
..
o

w
co
~ 0

o 9 10 11 12 13 14 15 16 17 16 19 20 21 22 23

PHOTON ENERGY (eV)


Figure 4.11: P-polarized normal-emission cross-section for Al at
45° angle of incidence. Circles: measured cross-section from Fermi
level and from surface state. Solid curve: cross-section calculated
within LDA-based RPA, normalized to measured intensity at 13 eV.
Dashed curve: single-particle cross section for a spatially constant
photon field. (Levinson, Plummer, 1981).

photoemission intensities from simple metals clearly requires proper treatment


of the local field in the surface region.

Inverse Photoemission

Inverse photo emission is another spectroscopy in which local surface fields


can playa crucial role. The radiation generated due to the downward transi-
tion of the incident electron induces surface charges whose longitudinal fields
interfere with the normal component of the bare field. Hence, this process in-
volves similar matrix elements as ordinary surface photoemission. Both cross
sections should therefore exhibit a similar frequency dependence. This has
been demonstrated for Al by Drube et al. (1988) (see Figure 4.12). The in-
verse photoemission spectra reveal a local field enhancement near W m , similar
to the one shown in Figure 4.11 for photoabsorption. In addition, the spectra
exhibit the same characteristic minimum near wp-
184 CHAPTER 4. NONLOCAL OPTICS

A1(100)

... ..
9 11 13 15 17 19 21
w (eV)
Figure 4.12: Inverse photo emission spectrum for Al(100). Solid
dots: experiment; open circles: experiment after subtraction of
a Gaussian distribution centered at 10.4 eV to account for surface
plasmon excitation due to surface roughness; solid curve: calculated
cross section. (Drube et al., 1988).

Beyond Standard Jellium and Adiabatic TDLDA

The excitation spectra discussed in this section are based on the standard
jellium model. Lattice effects, which lead to additional bulk and surface spec-
tral features, are not taken into account. As pointed out in Chapter 2, the
average effect of ionic pseudopotentials can be included by using the stabilized
jellium model. For high-density metals such as AI, the larger work function
causes a stiffening of the surface density profile, giving an upward shift of the
multipole plasmon. Conversely, the lower work function of the heavy alkali
metals makes the density profile more diffuse, giving slightly lower excitation
frequencies.
The shape of 1m d-L (w) in the vicinity of the threshold excitation should
depend on the form of the surface potential near the vacuum level. In par-
ticular, the image barrier gives rise to fine structure related to excitation into
image states. Such spectral features should in principle be observable in inverse
surface photo emission (Schaich, Lee, 1991).
The observed multipole frequency of K is Wm = 3.1 eV and satisfies
wm/wp = 0.82, in agreement with the adiabatic TDLDA. The electron en-
ergy loss data shown in Figure 3.12 are consistent with this frequency (see
also the overlayer photoyield spectra discussed in Section 4.7.2). If dynamical
4.4. INTERBAND TRANSITIONS 185

corrections are included as suggested by Gross and Kohn (1985), the strength
of the interaction is enhanced, shifting Wm to 3.2 eV (0.85wp) (Pehlke, Lieb-
sch, 1991). In the RPA, the frequency is 3.35 eV (0.89wp). In the case of AI,
dynamical corrections are rather small; in the TDLDA, Wm lies about 0.5 eV
below the RPA value.
For a more detailed discussion of these and other mechanisms that influence
surface excitations, see Section 3.3.3.

4.4 Interband Transitions

The theoretical treatment of lattice effects on nonlocal optical properties of


metal surfaces is still in its infancy. The general framework for incorporating
such effects was recently discussed by Schaich and Chen (1989) and Langreth
(1989). Fully three-dimensional calculations for realistic metals have so far
been carried out only for alkali metal overlayers (see Section 4.7.2). Experi-
mental information on the influence of interband transitions on optical surface
spectra is also very scarce. In Section 3.4, we saw that the strong pseudopo-
tential of Li leads to a 25 % lowering of the surface plasma frequency and to
a pronounced flattening of the surface plasmon dispersion with parallel wave
vector. It would clearly be of interest to investigate how these effects man-
ifest themselves in the long-wavelength spectra, in particular, how interband
transitions interfere with the multipole surface plasmon.
To estimate the influence of the crystal potential on the d-functions, Burke
and Schaich (1993, 1994) considered a one-dimensional model with parameters
appropriate for Li(llO). The bulk potential was represented by a single Fourier
component Va = -1.44 eV, while the surface potential was approximated by an
infinite barrier. The response was treated in the RPA, with a small damping
included (-y = 0.25 eV) to account for Drude absorption. Figure 4.13 shows
d.l(w) for this model. The analogous jellium results for Vo = 0 yield a single
spectral feature near wp. (The multi pole surface plasmon in this model is
suppressed because of the infinite barrier.)
The crystal potential is seen to introduce new spectral features near the
onset of interband transitions and at higher frequencies. Part of these new
features may be related to the so-called zone boundary collective states which
are familiar from bulk excitation spectra of Li (Sturm, Oliveira, 1989). In the
vicinity of these lattice-induced spectral peaks, the fluctuating charge density
exhibits long-range oscillations. Their periodicity is incommensurate with the
lattice since they involve dynamical Friedel oscillations and collective behavior.
186 CHAPTER 4. NONLOCAL OPTICS

2
~

~
3'
'-'..,
""0

-1
0 2 3
w / ~F

Figure 4.13: Frequency dependence of dJ.(w) for Li(llO). Solid


(dotted) curves: real (imaginary) part. The bulk plasma frequency
is wp = 1.62 CF = 7.65 eV; the onset of interband transitions occurs
at 0.63 CF. The range in which zone-boundary collective states may
exist is marked by the horizontal bar. (Burke, Schaich, 1993).

As mentioned in Section 3.4, such features may have been observed in electron
energy loss measurements on Li (Sprunger et al., 1992), but so far they have
not yet been identified in surface optical spectroscopies.
A one-dimensional lattice model was recently proposed also by Samuelsen
et al. (1993) for interpreting photoemission spectra from the layer compound
1T-TiS2 ·

4.5 Influence of Occupied d Bands


4.5.1 Comparison of Approaches

The influence of bound states on the optical properties of metal surfaces


has received attention for many years. In their work on the van der Waals
attraction between neutral atoms and noble metals, Zaremba and Kohn (1976;
see Chapter 6) used a two-component s-d electron model similar to that in
our discussion of the Ag surface plasmon in Section 3.5. Since the dielectric
4.5. INFLUENCE OF OCCUPIED D BANDS 187

function of noble metals can be represented as a sum of free- and bound-


electron contributions (see (3.71)), the total induced surface charge can be
expressed as
0" = O"s + O"a • (4.136)

Here, o"s is the integrated s electron surface charge and O"a is the surface po-
larization charge of the dielectric representing the d states. The centroid d1.
of the total induced density is then simply a weighted average of the sand d
contributions, ds and da. Thus,

(4.137)

Zaremba and Kohn approximated ds by using the result for jellium surfaces
and modeled the d states as a semi-infinite lattice of point dipoles. They
showed that da is located approximately half a layer spacing above the first
plane of nuclei, i.e., it roughly coincides with the edge of the positive jellium
background. Hence, da ~ 0 for all three low-index faces.
In the case of the van der Waals interaction, the assumption of independent
sand d electron responses is reasonable since, at imaginary frequencies, d1.
remains outside the metal surface. In the optical case, however, this picture
breaks down since the fields penetrate the metal and the screening charge
eventually shifts inside. To include the coupling between sand d states, Apell
and Holmberg (1984) replaced (4.137) via

d1- .lOs -1 10 d
- - - - s· (4.138)
10 - 1 lOs

This charge centroid approaches the jellium surface at the Ag bulk plasma
frequency. However, since the bulk plasmon also defines the transparency
threshold, d1. should not remain localized at the surface but shift to -00.
The models of Tarriba and Mochan (1992) and Feibelman (1993) were
already discussed in Section 3.5. The dipole lattice model of Tarriba and
Mocha.n has recently also been used to analyze the crystal face dependence of
the optical response and the anisotropy observed on Ag(llO) (see below).
Liebsch and Schaich (1995) evaluated d1.(w) for Ag within the TDLDA by
treating the s-d polarization in the same spirit as discussed in Section 3.5 for
the dispersion of the Ag surface plasmon. This approach is described in the
following section.
188 CHAPTER 4. NONLOCAL OPTICS

4.5.2 Ag

The evaluation of the surface d-functions is relevant for the optical response
of Ag and for the dispersion of Ag surface plasmon polaritons. In the s-d
polarization model, s and d electron contributions to the net d parameter
are calculated in a self-consistent manner. While the s electron response is
described microscopically within the semi-infinite jellium model, the occupied
d states are simulated via a uniform dielectric medium. The motivation for
this representation is that, as long as we are concerned with frequencies below
the interband onset, transitions involving d states are virtual. Because of its
one-dimensional nature, this model describes the overall frequency variation
of the optical response, but it cannot address more intricate features related
to the crystal structure. These aspects are discussed later.

Response Equation

Consider the density response to a uniform electric field oriented normal


to the surface. The applied electric potential is taken to be ¢lext (z, w) = - 2'1l' z.
Within the TDLDA, the induced s electron density is determined by the re-
sponse equation

nl(Z, w) = / dz' Xl (z, z', w) [¢lext(z', w) + ¢lest (z', w) + ¢lxc(z', w)]. (4.139)

Because of the presence of the polarizable medium in the region z ::; Zd, the in-
duced electrostatic potential satisfies the modified Poisson equation (see (3.76»

Cd(Z,W) ¢l~8t(Z,W) = -4'1l'nl(z,w) , Z =1= Zd , (4.140)


with cd( z, w) defined in (3.77). If we neglect the influence of d electrons on the
induced exchange-correlation potential, ¢lxc is given by (4.95).
To bring (4.139) into a form similar to that for jellium systems, we refor-
mulate the problem in the following manner. First, we remark that the total
electrostatic potential ¢l = ¢lext + ¢lest satisfies the boundary conditions

Cd(W) ¢l'(zi, w) ¢l'(zt,w) , (4.141)


c(w) ¢l'( -00, w) = ¢l'(oo,w), (4.142)

where zi == Zd ± 0+. The first condition may be implemented by writing


/, I ' I nl((z'' , w))
A..
'I'est ()
Z, W = - 2'1l' dz z - Z - 2'1l' IZ - Zd
Ia . (4.143)
Cd Z,W
4.5. INFLUENCE OF OCCUPIED D BANDS 189

The extra term involving the coefficient a accounts for the d screening charge
at the plane Z = Zd. According to (4.141), a is given by

' ( ') nl (z' , W)] (4.144)


a=ad (w) [ 1+ / dz sgnzd- z Cd(Z',W) ,

with ad given by (3.81). Far from the surface, if>est has the linear behavior

if>est(Z,w) -+ =f27Ta(w) [z - d.dw)] , Z -+ ±oo . (4.145)

The total charge and its centroid are given by

nl(z,w)
a(w) a+ /d Z Cd Z,W
( )'
(4.146)

d.dw) 1 [ / n1(Z,w)] (4.147)


a(w) aZd + dz Z Cd(Z, w)

These expressions show that the effective induced density consists of an s


electron part that is screened in the region Z :s; Zd, and a d electron charge
sheet at the boundary Zd. Formally, we may also write the total induced
density as a sum of sand d contributions: nl + nd, where nd is given by
1- Cd
nd(Z'w) = ac5(z - Zd) + nl(Z,w) --(J(Zd - z) . (4.148)
Cd

From (4.142), we find


C - 1
a = (4.149)
c+1
If we express the total induced charge as a = a. + ad, with
a.(w) = / dz nl(z,w) , (4.150)

then (4.146) implies


C. -1 Cd -1
a ---
'-10+1' ad=-- . (4.151)
10+1
In the next step, we account for the asymptotic behavior of the Coulomb
potential by separating the long-range bulk part. We write if> = if>b + if>l, where

-27T [z - (z - dl.) a] , (4.152)


( ) ( ) l
-41T Z - Zd a (J Z - Zd - 47T
z
-00
'(
dz Z - Z
,)n1(z',w)
(').
Cd Z, W
(4.153)
190 CHAPTER 4. NONLOCAL OPTICS

The surface potential ¢It vanishes for z «: O. The response equation (4.139)
now takes the form

(4.154)

with
(4.155)

The structure of this equation is analogous to the one for simple metals (see
Section 4.2.4); we therefore can use the same procedure to solve for the induced
density.

d-Parameters

Consider first dl.. Defining the centroid of the induced s density as

d.(w) = / dz z nl(z,w)/u.(w) , (4.156)

and using the two versions for dl. given in (4.137) and (4.147), we deduce the
following expressions for dd:

(4.157)

As the first line shows, dd is fully determined by the induced s density in the
region of the polarizable medium. The second form indicates that dd can be
written in terms of d. and a correction term depending on nl in the region
z ~ Zd.
To evaluate dll, we use (4.52). In the present s-d polarization model, this
implies
d 1 00 d

II = -00 z
c(z) - cF(z)
c- 1 '
(4.158)

where the microscopic local dielectric function is given by

(4.159)

In the Fresnel picture, we have instead cF(z) = 1 + (c -1) B( -z). From these
expressions we find
(4.160)
4.5. INFLUENCE OF OCCUPIED D BANDS 191

Close to the surface plasmon of Ag, c ~ -1 and Cd ~ +5, i.e., d ll ~ -2Zd·


This result implies that, as Zd is moved into the metal, dll shifts into the
vacuum.

Dynamical Force Sum Rule

The form of dJ.. in (4.137), with d. and dd given by (4.156) and (4.157), is
convenient for numerical evaluations since the troublesome region of the Friedel
oscillations appears only in the moment d.. The latter can be simplified by
using the dynamical forte sum rule (see Section 4.2.5). The great advantage
of this rule is that the total induced dipole moment is fully determined by the
part outside the jellium edge.
Because of the polarizable background representing the d electrons, an
additional first-order force must be included in the equation of motion for the
electronic density (Liebsch, Schaich, 1995). Thus, (4.120) is modified as

f
W2 dz znl(Z,W} = f dz {no(z} [Eext(z,w} + Ed(Z,W}] + nl(z,w} E+(z)} .
(4.161)
The field Ed is related to nd via E~ = -47T nd. Evaluating these terms is
straightforward, but tedious. We finally obtain:

(4.162)

where the full moment P.(w} and the external moment P.+(w} of the induced
s electron density are defined as in (4.131) and (4.132), respectively. For Zd ~ 0,
the new correction term is given by

The integrations involve only quantities in the surface region. Dividing (4.162)
by (J'., we obtain the desired sum rule for d. == P./(J'•.

Frequency Dependence of dJ..(w)

Figure 4.14 illustrates the frequency dependence of the total dJ..(w} and the
s and d contributions, d.(w} and dd(W}. Since a Drude damping was included
in the calculations, the divergence of dJ..(w} near the bulk plasma frequency is
192 CHAPTER 4. NONLOCAL OPTICS

3.5 3.6 3.7 3.8 3.5 3.6 3.7 3.8

w (eY) w (eY)

Figure 4.14: Frequency dependence of (a) real and (b) imaginary


parts of d1-(w) calculated within TDLDA for two-component s-d
electron system (Zd = 0). Solid curves: total d1- ( W ), dashed curves:
d.(w), dash-dotted curves: dd(W). The dotted curves indicate d1-(w)
in the absence of damping. (Liebsch, Schaich, 1995).

smoothed out. If'Y is taken to be zero, Red1-(w) as well as Red. and Redd
approach -00 for w --+ w;.
The real part of d1- is seen to remain positive up to just below w;. For
instance, for Zd = 0, we find Red1-(w;) ~ 1 ao. This value is only slightly less
than the static value 1.35 ao. In contrast, in normaljellium systems, Red1- be-
comes negative near the multipole plasma frequency, Wm ~ 0.8wp (see Figure
4.5). This demonstrates that screening at the surface of the two-component
s-d electron system remains very efficient nearly to the transparency threshold.
As Zd is shifted inside, Re d1- turns negative at progressively lower frequen-
cies, implying that the initial slope of the Ag surface plasmon becomes positive.
This is plausible since, at finite q, the shorter range of the plasmon field makes
the s-d polarization less effective, giving an upward shift of the frequency. We
recall that d ll ~ -2Zd enhances the blue shift for Zd < O.
These results indicate that it is incorrect to approximate d1- for Ag by the
corresponding jellium function and simply scale down the frequency axis by
w;/wp ~ 0.4. Instead, the s-d polarization reduces the window between w; and
w; w;
to a tiny fraction of (less than 3 %), while in ordinary jellium this range
comprises 30 %.
4.5. INFLUENCE OF OCCUPIED D BANDS 193

Notice that Re dl. remains positive up to higher frequencies than d. and


dd. The reason for this behavior is that the weights c. = (c. - 1)/(10 - 1) and
Cd = (Cd -1)/(10 -1) = 1- c. in the expression dl. = c.d. + Cddd (see (4.137))
may have opposite signs. Thus, there is no simple relationship between the
total dl. and the s electron part alone. In particular, the real parts of both d.
and dd can be negative, although the total Redl.(w) is positive.
Figure 4.14(b) shows Imdl. as well as the sand d contributions. In the
absence of damping, these functions remain positive up to w;. Once Drude
damping is included, Imdl. becomes negative at about w;. As in the case of
the real part, the imaginary parts of d. and dd may be negative while the total
1m dl. is positive.
The excitation spectrum given in Figure 4.14 does not reveal any evidence
of a multipole surface plasmon. Because of the extreme proximity of the sur-
face and volume plasma frequencies, an additional spectral feature between
these modes would be difficult to discern in electron energy loss spectra. The
extra broadening caused by bulk damping also implies that such a mode would
probably be swamped by the tail of the monopole surface plasmon. Photoyield
spectra of Ag overlayers might provide a more promising alternative (see Sec-
tion 4.7.2).

Induced Polarization

To compare the sand d contributions to the dynamic surface response, it


is instructive to consider the induced polarizations. These are given by the
expressions

p.(Z,w) (4.164)

(4.165)

and Pl. = p. + Pd. Figure 4.15 shows these polarizations, normalized to


Pl. = -a, e.g., p. = p./ Pl.' etc. In the bulk, these functlOns approach
F - F .

the asymptotic values F. = a./a = c. and Fd = ad/a = Cd, so that the


normalized total polarization in the interior is c. + Cd = 1. The polarization of
the d states has the opposite sign of the s electron polarization and therefore
causes a drastic cancellation. The main screening effect takes place within a
few A from the surface; deeper inside, all polarizations exhibit characteristic
Friedel oscillations about their mean asymptotic values. The discontinuities in
Fd(Z) and Fl.(z) are caused by the sheet of d electron screening charge located
at the edge of the polarizable medium.
194 CHAPTER 4. NONLOCAL OPTICS

4 I

'-'
\, .5
,
2 - (a) Re
3'
'\
(b) 1m
~ 0
,''--
QI 0
"" •

-.5
-2t-- -

-20 0 -20 0

z (ao) z (ao)

Figure 4.15: Spatial distributions of (a) real and (b) imaginary parts
of normalized induced polarizations for w = 3.63 eV (Zd = O). Solid
curves: total polarization P1. (z, w), dashed curves: s electron po-
larization p. (z, w), dotted curves: d electron polarization Pd ( z, w}.
(Liebsch, Schaich, 1995).

Surface Plasmon Polaritons

According to our discussion in Section 3.3.6, nonlocal optical effects should


influence retarded and nonretarded responses in a consistent manner. As
shown in Section 3.5, the measured Ag surface plasmons exhibit a positive
slope on all three low-index faces. In the s-d polarization model, these data
can be qualitatively reproduced. The linear behavior of w;(q} at small q is
given by
(4.166)

where Re (d1. -d rr ) < 0 for Zd < O. This implies that, in the retardation region,
the Ag surface plasmon polariton frequencies should lie above the classical
Fresnel dispersion.
Surprisingly, attenuated reflection data on Ag by Tadjeddine et al. (1980)
show the opposite trend. As illustrated in Figure 4.16, the measured fre-
quencies lie below the Fresnel limit (3.65) derived from optical data of both
Hagemann et al. (1975) and Johnson and Christy (1972). This discrepancy
could be due to the fact that the measurements by Tadjeddine et al. were not
4.5. INFLUENCE OF OCCUPIED D BANDS 195

3.0

r--. 2.5
>
-!. I
3

2.0
~
~1;fh,f0li2 j
1.5 I I I I _LLLL~LI -LI....JL--L-L
. ....JL_.l.LJ.....JI----'J
1.05 1.10 1.15 1.20 1.25

q./qo
Figure 4.16: Surface plasmon polariton dispersions for Ag single-
crystal surfaces (Kolb, 1982; Tadjeddine et al., 1980). The dots
(F) denote the classical Fresnel dispersion derived from the bulk
dielectric function of Johnson and Christy (1972) (qO == w/ c).

done under UHV conditions. Small amounts of surface oxide may also lower
the polariton frequencies. Notice, however, that this comparison is sensitive to
the accuracy of the bulk dielectric function. For example, the bulk c quoted
by Tadjeddine et al. (1980) gives a Fresnel dispersion below rather than above
the measured polariton curves.
Another reason for the different behavior might be the frequency depen-
dence of the d-parameters. In fact, according to Figure 4.14, Redl. > 0 for
w < 3.5 eV. On the other hand, (4.160) implies Redll > 0 for Zd < O. The
sign of Re dl. - dll therefore depends on the balance of both contributions.
Measurements at polariton frequencies and near w:,
under identical surface
conditions, could reveal the true nature of the nonlocal corrections.
The microscopic understanding of the surface plasmon polariton dispersion
is crucially important for the study of chemical processes at metal-electrolyte
interfaces (Dzhavakhidze et al., 1989). Surface charging, adsorption, and
dielectric screening have a significant influence on the polariton frequency
(see also Sections 3.6, 4.6, and 5.4). In a technique called 'surface plas-
196 CHAPTER 4. NONLOCAL OPTICS

mon microscopy', this phenomenon has recently been exploited to obtain two-
dimensional images of potential waves in electrochemical systems (FUi.tgen et
al., 1995).

Crystal Face Dependence

The crystal face dependence of the Ag surface plasmon polaritons does not
follow the same trend as the surface plasmons at finite q. According to Figure
3.27, the positive slope of the Ag surface plasmon increases in the following
sequence: (110)-[HO], (111), (110)-[001], and (100). On the other hand, Fig-
ure 4.16 shows that the slopes of the Ag surface plasmon polaritons increase
in the order: (110)-[001], (110)-[110], (100), and (111). Again, this discrep-
ancy may be caused by different experimental surface conditions, or by the fact
that these measurements are carried out at different frequencies. On the other
hand, the difference may also indicate that the overall positive dispersion seen
in electron energy loss spectra does not hold in the long-wavelength limit: As
pointed out above, the results for the s-d polarization model suggest that the
true linear region is very small and governed by other physical processes than
the large-q region.

Anisotropy

The optical anisotropy of Ag(110) is another manifestation of nonlocal


surface effects. A suitable detection method is refiectance spectroscopy which
yields the real and imaginary parts of the quantity

l:!.T == T[lIO] _ 1, (4.167)


T T[OOl]

where T[lIO] and T[OOl] are the complex refiectances for the electric field aligned
in the [HO] direction (along atomic rows) and [001] direction (across atomic
rows). Near normal incidence, (4.30) and (4.43) give

~T ~ -2iqz [d ll [110] (W) - dll[OOl](W)] . (4.168)

Figure 4.17 shows the frequency dependence of this anisotropy in the vicin-
ity of the Ag surface plasmon (Borensztein et al., 1993). Both real and imagi-
nary parts are remarkably large. The experimental data are for Ag in air and
are compared to results derived within the point-dipole lattice model of Tar-
riba and Mochan (1992) (see Section 3.5). Despite the simplicity of this model,
4.5. INFLUENCE OF OCCUPIED D BANDS 197

.1
... ...
'c
<J
'c
<J
., E
a::: o --

-.1
3.6 3.8 4.0 4.2 3.6 3.8 4.0 4.2

w (eY) w (aY)

Figure 4.17: Optical reflectance anisotropy i:J.r Ir for Ag(llO): (a)


real part, (b) imaginary part. Dashed curves: experiment; solid
curves: calculations based on lattice of point dipoles immersed in
a local Drude medium. (Borensztein et al., 1993).

the calculations reproduce the measured reflectance anisotropy qualitatively.


In particular, experiment and theory indicate that

(4.169)

is mainly positive, i.e., Red ll [lIo](w) < Redll[OOl](W). Since, at small q, the
(110) anisotropy of the surface plasmon is determined by Redll(w), this relation
implies that the linear slope in the [110] direction is smaller than in the [001]
direction. This is indeed observed in the electron energy loss measurements
shown in Figure 3.27. We point out, however, that the anisotropy is rather
sensitive to surface conditions. Recent reflectance anisotropy measurements
on Ag(llO) in ultrahigh vacuum by Fernandez et al. (1997) also exhibit a
resonance near 3.7 eV, but the line shape differs appreciably from that shown
in Figure 4.17.
The frequency variation of i:J.rlr may be rationalized as follows. If the d
electrons are represented in terms of a homogeneous polarizable medium, dll
is simply given by (4.160). For Zd < 0, this function exhibits spectral features
close to the interband onset near 3.9 eV. Microscopic effects should make these
198 CHAPTER 4. NONLOCAL OPTICS

features appear at slightly different frequencies if the field is aligned along or


across the atomic rows of the (110) face. In the dipole model, the surface
conductivities in the [110] and [Om] directions show maxima near 4.1 and 3.8
eV, respectively. If we take the difference between these functions, we find the
characteristic spectral behavior shown in Figure 4.17. It would be interesting
to refine these calculations by taking into account the nonlocal response of the
s electron density.

4.6 Charged Surfaces

The optical response of charged metal surfaces is relevant for the investi-
gation of metal--electrolyte interfaces. The ions in the electrolyte generate a
static electric field that may be varied over a wide range. Apart from the prac-
tical importance of electrochemical cells, the study of these interfaces is also
interesting from a fundamental point of view since the electronic density profile
can be modified in a controlled manner. Such studies provide a unique oppor-
tunity for investigating the sensitivity of the optical response with respect to
ground-state electronic properties.
As shown in Figure 3.34, positively charged surfaces tend to have a steeper,
less polarizable density profile. Conversely, negatively charged surfaces have a
softer profile and electrons leak out farther into the vacuum. Although these
modifications appear rather small, they nevertheless have significant conse-
quences for the electronic excitation spectra and the nonlocal corrections to
the Fresnel reflectivity.
The optical response of charged jellium surfaces was studied by Gies and
Gerhardts (1986) within the TDLDA (see Figure 4.18). Since the differential
reflecti vity (4.71) was derived for a slab geometry, a damping was included
to smooth the discrete transitions between slab states (see also Figure 4.4).
The frequency of the multi pole surface plasmon is seen to be strongly affected:
The softer density profile of negatively charged surfaces (E > 0) causes a
downward shift, whereas positive charging (E < 0) leads to a blue shift and
gradual suppression. The latter trend is consistent with the result dl.(w) --+ 0
expected from the dynamical force sum rule (see (4.130)) and the general
arguments presented by Schaich (1994b). In Chapter 5, we see that static
charges also have a strong influence on the nonlinear optical properties of
metal surfaces.
4.7. OVERLAYERS 199

10

C1 b
08

0.6
i.
0:
~

0: 04
<J
I

02

00
05 06 07 0.8 09 1.0

W/Wp

Figure 4.18: Nonlocal correction flRpIR: to Fresnel reflectivity


for p-polarized light incident on a jellium surface (T. = 3, () = 45°).
The labels denote static external electric fields E in VIA: (a) 0.4,
(b) 0.0, (c) -0.8, (d) -2.1, (e) -4.3. (Gies, Gerhardts, 1986).

4.7 Over layers


4.1.1 Local Optics Picture

In the three-layer model mentioned in Section 4.1, the functions d.L(w) and
dll(w) take on the simple forms (see relations (3.41) and (3.40))
I: a - 1 I:
d.L(w) a--- (4.170)
I: - 1 I: a '
I: a -1
a-- (4.171)
I: -1 '

where I:a(w) is the local dielectric function of the overlayer and a the thickness.
For Drude metals, (I:a - 1)/(1: - 1) = nalnb, where na and nb are the bulk
densities of the adsorbate and substrate. Thus,
I: -11:-1:
d.L(w) - dll(w) = a _a_ _ >, _a • (4.172)
I: -1 I: a
At the volume plasma frequency of the surface layer, this difference becomes
large since I:a(w) vanishes. Obviously, this model neglects the microscopic
200 CHAPTER 4. NONLOCAL OPTICS

properties of the adsorbate-vacuum and adsorbate-substrate interfaces. In


Section 3.8, we discussed the influence of these properties on the collective
excitations of metallic overlayers, in particular, the dispersion of surface- and
bulk-like modes with parallel wave vector. In the following section, we examine
overlayer excitations in the long-wavelength limit appropriate for studies using
incident photons.

4.7.2 Realistic Alkali Metal Overlayers

The nonlocal optical response of alkali metal overlayers was evaluated within
the TDLDA for the jellium model (Liebsch et al., 1990) and for realistic, three-
dimensional systems (Ishida, Liebsch, 1990, 1992, 1997). With the aid of these
calculations, it is possible to determine the transition from the low-coverage
regime, where intra-atomic transitions between the quasi-localized adatom lev-
els dominate, to coverages in the monolayer range, where the metallic character
within the overlayer gives rise to collective excitations.

Ground-State Electronic Properties

To illustrate the atomic character of the electronic density of realistic alkali


metal overlayers, we show in Figure 4.19 contour density maps for Na overlay-
ers and the corresponding lateral averages of the Na-induced valence density.
The alkali ion cores in these calculations are represented by norm-conserving
pseudopotentials (Bachelet et al., 1982), and the three-dimensional electronic
structure is evaluated using the embedding approach (see Section 2.3.6).
Surprisingly, despite the considerable corrugation of the density, the planar
averages are very similar to the profiles obtained for equivalent jellium overlay-
ers. At a full monolayer, the Na-induced valence charge is almost symmetric
about the Na ion core. As the coverage is reduced, the valence charge is polar-
ized towards the substrate, indicating a partial charge transfer. The valence
charge then begins to look like the image charge induced at the bare surface
by a uniform external electric field.
Since the surface contribution to the work function depends only on the
planar average of the equilibrium density, the results in Figure 4.19 imply that
the jellium model yields rather accurate values down to about half a monolayer.
At lower coverages, the work function calculated using the jellium model is too
low since the positive jellium slab binds the valence electron less well than the
true ionic potential of the alkali atom.
4.7. OVERLAYERS 201

3
o 0.2 9=1
·3 ....,
4 =l
9=~ .g
<D 0
o -;s,. 02
9=1
C?
<D
~ -4
oS N
,?
x 6 0
I
'@
~ 02 B=t
o Ie
w

-6
5 5 10
Z (o.u.) Z(o.u.)

Figure 4.19: (a) Contour maps of ground-state density of hexago-


nal Na layers on a jellium substrate (r. = 3). The coverages are
c = 1, 1/2 and 1/5. The solid dots denote the Na cores. (b) Planar
averages of the Na-induced valence density. Solid curves: realistic
overlayers (the arrows denote the centroids); dashed curves: equiv-
alent jellium results. The dot-dashed curve for c = 1/5 shows the
image density of the bare metal. The substrate occupies the half-
space z ::; O. (Ishida, Liebsch, 1990).

Similar results are obtained for the density induced by a static electric field
oriented normal to the surface. The induced densities of realistic Na overlayers
are even more corrugated than the ground-state densities since the screening of
the applied field takes place in the region of the adsorbate-vacuum interface.
Nevertheless, the planar averages of these distributions closely resemble the
induced densities derived for jellium overlayers. For c = 1/5, the differences
are more pronounced. These results imply that the static image plane of
the adsorbate-substrate system is also quite accurately represented within the
jellium model as long as c ::::: 1/2.

Overlayer Excitations

Figure 4.20 illustrates contour maps of the dynamical induced density for a
202 CHAPTER 4. NONLOCAL OPTICS

w=O.5 eV w=2.5 eV w=4.5 eV


10

:::J
d 5
N

-5 0 5 -5 o -5
X(ou.l

Figure 4.20: Contour maps of real part of induced density for Na


on Al (c = 1/4). The frequencies range from 1 to 4 eV. The dots
denote the positions of the Na ions. The substrate is located in the
region z ::; O. The horizontal scale extends over a unit cell. Solid,
dashed and dot-dashed lines correspond to positive, negative and
zero values. (Ishida, Liebsch, 1992).

low-coverage Na layer on Al. Because of the weak overlap between neighboring


overlayer sites, the atomic character is very pronounced. At low frequencies,
screening of the applied field by the outer tails of the Na valence electrons is
quite efficient and the centroid of the induced density lies near the adsorbate-
vacuum interface. With increasing frequency, the low-density Na overlayer is
less able to screen the external field. For c = 1/4, the effective Na plasma
frequency is only half of the Na bulk plasma frequency. Near 3 eV, therefore,
the overlayer becomes transparent. The calculations show this effect quite
nicely: The induced density shifts towards the surface (avoiding the Na core
region), until it is located at the adsorbate-substrate interface. At higher
coverages, this inward shift takes place at accordingly higher frequencies.
Despite the atomic character of the induced densities, the surface excitation
spectra reveal no trace of atomic-like transitions. This is illustrated in Figure
4.21, which shows the frequency dependence of Imdl.(w) for several Na layers
on Al. The centroid is derived from the expression

(4.173)

where nl is the density induced in the three-dimensional overlayer by a uniform


electric field oriented normal to the surface. The integrated charge density a(w)
4.7. OVERLA YERS 203

Na/AI
30
::i
.,;

...
10

4 6 8 10
w (eV)

Figure 4.21: Surface excitation spectra for Na layers on Al at several


coverages. Symbols: TDLDA results for realistic Na layersj solid
curves: spectra for equivalent jellium overlayers. The vertical bars
denote the work functions <P. The bulk plasma frequency of Na
is 5.9 eVj the multipole surface plasma frequency is about 4.8 eV.
(Ishida, Liebsch, 1992).

is given by (4.98) and dJ. is measured with respect to the edge of the substrate
positive background. For the sake of comparison, analogous spectra for the
equivalent jellium model are also plotted. Both descriptions are seen to give
nearly identical spectra, even at a coverage as low as c = 1/4.
At two monolayers, the adsorbate spectrum shows two peaks: one near
the volume plasma frequency wp of the overlayer, and another close to the
Na multipole surface plasma frequency Wm ~ 0.8wp • These excitations are
analogous to the q = 0 limit of the spectra shown in Figure 3.41(a} for K
on AI. While the induced density near Wm is localized in the region of the
adsorbate-vacuum interface, the density associated with wp extends across the
entire overlayer.
At one monolayer, the two modes are strongly broadened since the over-
layer density does not exhibit a plateau and one cannot distinguish separate
adsorbate~vacuum and adsorbate-substrate interfaces (see Figure 3.39). As a
result, only a rather broad spectral feature remains. Below one monolayer, the
spectra show a broad feature close to W ~ <P, the threshold for emission. Since
the multipole mode is suppressed, here in the overlayer case, this threshold
excitation is much more pronounced than in spectra of clean metal surfaces
204 CHAPTER 4. NONLOCAL OPTICS

(see Figure 4.5).


The formation of collective modes at coverages in the one- and two-mono-
layer regime is not surprising because of the extended character of electron
states parallel to the surface. However, at low coverages, the absence of atomic-
like spectral peaks is not at all obvious. In the present case, it is a consequence
of the strong hybridization between adatom and substrate states. Occupied
and unoccupied local densities of states are rather broad so that electronic
transitions form a smooth, structureless continuum. Weaker chemisorption
bonds may be more favorable for the appearance of adatom spectral features,
e.g., s-p transitions.
The reason for the maximum of Imdl,(w) near the threshold of emission
is rather subtle since it is not caused by the larger density of one-electron
states near the vacuum level. A detailed analysis of this effect suggests that
it depends on the surface screening processes which strongly reduce the self-
consistent potential in the substrate region. As the calculations show, matrix
elements of this potential have their largest amplitude when the initial state
is near EF and the final state near the vacuum level.
In principle, the lattice structure within the overlayer also alters the induced
currents parallel to the surface. The function dll is a measure of such effects.
For three-dimensional adlayers, the centroid dll is defined as

(4.174)

where the tangential surface current is given by (4.14). The nonlocal conduc-
tivity tensor is specified in (2.113). In the jellium model, dll = constant =
a na/nb' Calculations for realistic Na and K layers on Al give negligible devi-
ations from this value for the real part of dll and very small imaginary parts
(Ishida, Liebsch, 1992). Hence, the dynamical response of these overlayers
is dominated by the perpendicular component. Larger dll should occur, for
example, for Li overlayers as a result of stronger pseudopotentials.

Comparison with Measured Spectra

• The above spectra are relevant for reflectivity measurements from ad-
sorbed alkali metal layers. According to (4.68) and (4.69), nonlocal corrections
to the reflectivity of p- and s-polarized light are determined by the quantities
ImDl, and Imdll , respectively, where Dl, essentially measures the difference
(dl, - dll ) (see (4.70)). Denoting clean-surface quantities by the superscript s
and those in the presence of the adsorbate by a, we find for the change in the
4.7. OVERLA YERS 205

.10 I I I
++
+ + •• +
+ +
c = 2 ++

...
2cr." .05 e- +
+
+
+..
... -
<l

2 3 4
w (aY)
Figure 4.22: Measured adsorbate-induced change of p-polarized re-
flectivity for three coverages of Cs on Ag. (Liebsch et al., 1990).

reflectivity as a result of adsorption:

Ir;12 - Ir;12 = Ir:12 4qz 1m (D~ - DD , (4.175)


Ir!12 - Ir~12 = Ir;12 4qz 1m (dil- dIT) . (4.176)

Since dll is nearly real, we have !1R. :::::; O. Furthermore, as long as c(w) is
real, !1Rp is proportional to 1m (d'l. - djJ. This means that, for weak bulk
and surface absorption due to the substrate, !1Rp gives the reduction of the
reflectivity due to adsorbate-induced excitations. Apart from the prefactor
c /(c cot 2 (1 -1) ((1 is the polar angle of incidence), this reduction is determined
by the adsorbate spectral distribution 1m d'l.( w).
Figure 4.22 shows measured reflectivity spectra by Hincelin and Lopez-Rios
(see Liebsch et al., 1990) for Cs on Ag. At a half-monolayer coverage, a broad
distribution is observed using p-polarized light. This may correspond to the
calculated threshold excitation. At one monolayer, a maximum near 2.3 eV
develops, which is presumably related to the Cs multipole surface plasmon. At
two monolayers, the spectrum becomes stronger and broader because it now
involves the overlayer multipole and volume plasmon. For s-polarized light, the
reflectivity remains negligible at all coverages. These data cannot, of course,
be understood within the local optics picture. The results are, however, in
206 CHAPTER 4. NONLOCAL OPTICS

qualitative agreement with the trend predicted by the TDLDA calculations


shown in Figure 4.21.
• Adsorbate-induced excitations can also be detected in angle-resolved
photo emission measurements. Since these excitations influence mainly the
p-polarized emission, the ratio Yp(w}/y'(w} is greatly enhanced due to the
presence of the adsorbate, in particular, just above threshold. Qualitatively,
this result agrees with experimental data by Benemanskaya and Lapushkin for
Cs on Ag (see Liebsch et al., 1994). The ratio shows characteristic maxima
about a few tenths of an eV above threshold for 0.2 - 1.0 monolayer. Notice
that, close to the emission threshold, the external contribution to the total
photoabsorption is much smaller than the internal one (see Section 4.3.3).
Thus, Imd1-(w} cannot be used to analyze the emitted photoyield in this
frequency range.
• The clearest evidence for the overlayer excitations predicted by the
TDLDA comes from photoyield measurements for alkali metal layers on Al
by Kim et al. (1997) and Barman et at. (1997). Figures 4.23 and 4.24 show
the evolution of the K and Na spectral features with coverage. For less than
one monolayer, there is only a broad peak corresponding to a mixture of the
threshold excitation and the alkali metal multi pole surface plasmon. The lat-
ter mode is still rather weak since the adsorbate-vacuum interface is not yet
well defined. With increasing coverage, the multi pole mode becomes notice-
ably stronger. At the same time, a second peak begins to appear at higher
frequencies. This mode is associated with the alkali metal volume plasmon
(the antisymmetric slab mode) whose strength grows as the electronic density
within the overlayer looks more plateau-like. Towards two monolayers, the
adsorbate-vacuum and adsorbate-substrate interfaces are well separated, so
that the multi pole mode and volume plasmon can be clearly resolved. These
observations are in excellent agreement with the trend exhibited by the theo-
retical spectra shown in Figure 4.21.
The K multipole frequencies measured by Kim et al. (1997) and Barman et
al. (1997) satisfy wm/wp = 0.82, in agreement with the K yield data shown in
Figure 4.9 and the loss data shown in Figure 3.12. The TDLDA predicts the
same ratio, while the RPA response gives wm/wp = 0.89 (see Figure 4.5), im-
plying a 0.25 eV higher frequency. A slightly larger shift is found for Na. These
differences indicate that exchange-correlation contributions to the complex lo-
cal potential are important. Their treatment within the adiabatic version of
the TDLDA provides a remarkably accurate representation of the observed
multipole frequencies.
• Interestingly, Barman et al. (1997) also observed the multipole surface
plasmon for thin Li layers on AI. Figure 4.25 shows a comparison of the data
4.7. OVERLAYERS 207

30

K/AI

20 c=2

.,
'>' 10
o
(;
.r:
c..

o~~~----~--~~~~--~~~~
2 3 4 5

W CeY)
Figure 4.23: Measured p-polarized photoyield spectra for thin K
layers adsorbed on Al at different coverages. (Kim et al., 1997).

15
---'c!! Na/AI
:J

.ciL-
a
'-'
10
;Ni / Wm wp

/___1
1J
-.; c=2
'>'
0
(; 5
.r:
c..
'.,, 0.5 "-'-_

0
3 4 5 6 7 8
w (eY)
Figure 4.24: Measured p-polarized photoyield spectra for thin Na
layers adsorbed on Al at different coverages. (Barman et al., 1997).
208 CHAPTER 4. NONLOCAL OPTICS

with results of TDLDA calculations by Ishida and Liebsch for realistic, three-
dimensional Li overlayers on a jellium substrate. While the multipole peak
near Wm = 5.2 eV is well resolved, the Li volume plasmon near wp = 6.7 eV is
very broad. In the equivalent jellium overlayer model, both excitations yield
distinct spectral features, just like in the Na spectra shown in Figure 4.21. The
r. = 3.25 jellium mode frequencies are wp ~ 8 eV and Wm ~ 0.8wp = 6.4 eV.
The Li lattice potential lowers the frequencies of both modes. In addition,
it gives rise to strong interband transitions within the adsorbate, so that the
bulk-like mode is heavily damped.
The origin of this strikingly different lattice effect on the width of the
Li overlayer modes is the location of the fluctuating charge relative to the
crystal potential. As shown in Figure 4.26, the bulk-like mode corresponds to a
standing wave in the overlayer and is fully exposed to the ionic potentials. The
broadening of this mode is consistent with the large intrinsic width (2.5 eV)
of the Li bulk plasmon (Sturm, 1982). The multipole charge, however, is
concentrated near the adsorbate-vacuum interface. Decay via lattice-induced
excitations is therefore much weaker.
The theoretical spectrum shown in Figure 4.25 is based on the TDLDA.
The position of the calculated multipole frequency agrees very well with the
photoyield data. Surprisingly, an RPA response calculation gives a qualita-
tively different spectral distribution (Ishida, Liebsch, 1997). Since the neglect
of exchange-correlations contributions to the local potential makes the density
profile effectively less polarizable, the frequency is shifted upward. For the
same reason, the fluctuating density is located further inside, i.e., it overlaps
more with the ionic potential, enhancing the decay via interband transitions
within the overlayer. The calculations show that this extra broadening is so
strong that the multi pole mode does not appear as a well-defined spectral fea-
ture; instead, it merely becomes a shoulder on the low-frequency side of the
bulk-like Li overlayer plasmon, in disagreement with the data. This example
shows that surface excitations can depend on a delicate balance of single-
particle transitions and many-body screening effects. A proper treatment of
both features can be essential for an accurate representation of observed spec-
tra.
We recall that, in the loss measurements on clean Li, it was not possible
to detect the multipole surface plasmon (see Section 3.4) because of the large
width of the monopole mode. Photoyield spectra of thin overlayers might
therefore be an excellent method of studying multipole modes: The use of
photons suppresses the ordinary surface plasmon and the finite film thickness
has the tendency to weaken the bulk-like overlayer plasmon. Hence, the multi-
pole mode becomes the main overlayer-induced excitation. Following this idea,
4.7. OVERLA YERS 209

,....,
~--,-----

,,
, - ,,
,,
:::l
(b) ,
~ 20 ,,
,,
2:... ,
""0

E
0

""0
OJ 5 wp
'>' Li/AI
0
0
.c
a..

0
4 6 8 10
w (eV)

Figure 4.25: (a) Measured p-polarized photoyield spectra for Li on


Al (c = 2.3). (b) TDLDA spectra for Li on Al jellium substrate
(c = 2). Solid curve: realistic Li overlayerj dashed curve: equivalent
jellium overlayer. (Barman et al., 1997).

---------1,
,
,,r--------------------

-5 o 5 10 15
z (ao)
Figure 4.26: Laterally averaged fluctuating charge density for Li
on Al jellium substrate (c = 2). The squares denote the positions
of the Li planes. Upper curve: bulk-like Li overlayer modej lower
curve: Li multipole surface plasmon. (Barman et al., 1997).
210 CHAPTER 4. NONLOCAL OPTICS

one might speculate whether the Ag multipole surface plasmon could be de-
tected in yield spectra of thin Ag overlayers. Presumably it would lie between
w; = 3.8 eV and 0.8 x 9 = 7.2 eV, where 9 eV is the plasma frequency of the
5s electrons.

Overlayer Volume Plasmons

As pointed out in Section 3.8.3, the appearance of the bulk-like overlayer


plasmon at coverages of only two monolayers is perfectly compatible with the
dispersion relation of volume plasmons. The physical origin of the excitation
of this mode is the coupling between the incident transverse wave and longi-
tudinal fields in the overlayer-substrate region. According to Sauter (1967),
this type of coupling should occur quite generally at interfaces between me-
dia. Forstmann (1967) calculated the corrections to the Fresnel reflectivity
due to this effect for metal surfaces; Melnyk and Harrison (1968) considered
the coupling to standing plasma waves in thin films. The first experimental
observation of this effect was the excitation of plasma waves via transmission
of light through Ag films (Lindau, Nilsson, 1971).
Another example is photoemission from K films on quartz (Anderegg et al.,
1971.) For thicknesses a = 27 ... 100 A, standing plasma waves were generated.
The normal component of the plasmon wave vector satisfied the condition
qz = mr / a. The sequences of spectral features were found to follow a quadratic
dispersion of the plasma frequency with qz: Wn = wp + Q q; where Q =
0.6cF/wp . As pointed out by Feibelman (1975b), because of the nonsymmetric
geometry of the K films (bounded by vacuum and a quartz substrate), the
indices should be n = 1, 2, 3 ... rather than n = 1, 3, 5 ... as expected for a
symmetric free-standing film.

4.7.3 Ag Over layers

Nonlocal effects were also observed by L6pez-Rfos et al. (1979) in reflectiv-


ity spectra from thin Ag overlayers on Al (see Figure 4.27). With decreasing
coverage, the adsorbate plasmon peak is seen to shift to higher frequencies and
become significantly broader. Similar behavior of the Ag plasma resonance was
seen for Ag layers grown on eu in differential reflectivity and electroreflectance
measurements (Katz, Kolb, 1980) and in surface photo emission spectra (Sass
et al., 1977). These effects cannot be understood within the local optics pic-
ture which yields spectral peaks of constant frequency and width regardless of
4.7. OVERLA YERS 211

I I
.3 f- -

a
39
.2 r -

...
II:: 23
<J
18
-
13
10
6

4000 3000 4000 3000


A(.l) A(.l)
Figure 4.27: (a) Measured reflectivity change for Ag layers on AI. a
denotes the Ag layer thickness in A. (b) Reflectivity change calcu-
lated within a nonlocal hydrodynamic model. (L6pez-Rfos et al.,
1979).

thickness (see (4.172». On the other hand, allowing for the excitation of longi-
tudinal waves within the Ag overlayer, the observed trend can be qualitatively
reproduced using a hydrodynamic model (for a discussion, see L6pez-Rfos,
1992).
According to the model described in Section 3.8.4, a more complete expla-
nation of these observations includes the reduced range of the s-d polarization
as the coverage is decreased. The screening of the Ag overlayer plasmon via
the occupied d bands is then less effective, giving a blue shift of the frequency.
In addition, the relative s electron spill-out is enhanced, causing a further up-
ward shift of the frequency. The excitation of longitudinal waves within the Ag
overlayer is included in this picture since it is based on a nonlocal response of
the locally screened s electron density. As shown in Figure 3.49, the s-d polar-
ization gives a blue shift with decreasing coverage, in analogy to the blue shift
of the Ag surface plasmon as a function of qll and of the Ag Mie resonance as
a function of 1/R. At the same time, the broadening of the plasmon increases
strongly because of the greater overlap with interband transitions involving
the d bands (see also Figure 3.49).
Evidence of this behavior was recently also observed in differential reflec-
212 CHAPTER 4. NONLOCAL OPTICS

4.2

,-..
>CD

:z
'-'

4.0

Ag/Si

,"
3.8

o .1 .2
1/0 (A-1)

Figure 4.28: Plasma resonance observed in thin Ag films on Si as


a function of inverse film thickness in the region from about 2 to
5 monolayers. The dashed line indicates the resonance obtained in
the local optics picture. (Borensztein et al., 1995).

tivity spectra for thin Ag films on Si (Borensztein et al., 1995). As shown


in Figure 4.28, below three monolayers of Ag only a broad spectral feature is
seen (the thickness of one monolayer is about 2.4 A, corresponding to a close-
packed (111) plane). At higher coverages, the Ag plasma resonance begins to
form, but the frequency is blue shifted by several tenths of an eV. This blue
shift diminishes linearly with inverse film thickness. The slope (0.8 A) agrees
quite well with those observed for Ag surface plasmons as a function of qll
(0.4 - 0.8 A; see Section 3.5) and for the Ag Mie resonance as a function of
1/ R (0.9 Ai see Section 3.9). The striking similarity of these slopes suggests
that all of these phenomena are governed by the same physical mechanism.
Chapter 5

Nonlinear Optics

Optical second harmonic generation (SHG) and sum frequency generation


(SFG) provide a unique opportunity for studying surface and interface proper-
ties. After introducing the phenomenological theory of SHG at metal surfaces,
we discuss recent microscopic evaluations of the nonlinear surface polarizability
components. As before, we focus on the nonlinear response at simple metal sur-
faces. In these systems, the isotropic perpendicular component of the surface
polarization is particularly important since it is determined by the gradients
of the normal component of the electric field. A striking result of the TDLDA
calculations is that, as in the linear case, the main spectral features are the
threshold excitation and multipole surface plasmon. These excitations play an
important role in SHG from Al and adsorbed alkali metal layers. The remark-
able surface sensitivity of the perpendicular polarization can be seen if the
metal is charged. We also discuss surfaces at which the inversion symmetry is
broken in the parallel direction, for example, the (111) face of fcc crystals and
stepped metal surfaces. The anisotropic component of the surface polarization
in these systems can be as large as the main isotropic component.
Nonlinear optical methods offer several advantages over conventional sur-
face spectroscopies. Since they do not rely on the short electronic mean free
path to achieve surface sensitivity, they can be used to probe buried interfaces
that are of increasing technological relevance. Moreover, the time resolution
makes it feasible to study dynamical processes and reactions over a wide range
of time scales. Molecular selectivity can be achieved by using sum frequency
generation for surface vibrational spectroscopy. The microscopic understand-
ing of nonlinear optical processes at simple metal surfaces should serve as a
qualitative guide for interpreting spectra in more complex situations.

213

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
214 CHAPTER 5. NONLINEAR OPTICS

5.1 Phenomenological Considerations

Nonlinear optical processes at metal surfaces are considerably more intri-


cate than the linear response discussed in Chapter 4. Before describing the
evaluation of the nonlinear surface polarization within the TDLDA, we dis-
cuss the leading contributions to the harmonic polarization and the resulting
radiation.

5.1.1 Bulk Polarization

The nonlinear second harmonic polarization induced by incident radiation


may quite generally be expanded as (i = x, y, z) (Adler, 1964; Bloembergen et
al., 1968; Shen, 1984)
P2i(f,w) = Xiik(W) E1i(r,W) Elk (f,w)
+ Qiikl(W) Elj(f',W) V k ElI(f',w) + ... , (5.1)
where Eli are components of the linear electric field, Xiik(W) are elements of
the nonlinear dipole polarizability, and Qiikl(W) those of the quadrupole po-
larizability. (We use the convention that all fields and response functions refer
to the fundamental frequency w. Linear and quadratic terms are distinguished
by their indices.)
If the system has inversion symmetry, the incident fields ±E1(w) must
induce polarizations ±P2(W), respectively. This condition is satisfied only if
Xiik(W) == O. In the bulk of a centrosymmetric solid, SHG is therefore dipole-
forbidden. To lowest order, only magnetic dipole and electric quadrupole terms
contribute. In the case of a single beam interacting with a cubic crystal, the
bulk polarization reduces to (Tom et al., 1983)
P2iB (f', w) = -y(w) Vi(E- I . E- 1) + ((w) E1.viEli , (5.2)
where -y(w) and ((w) are the coefficients of the isotropic and anisotropic con-
tributions, respectively. It is evident from this definition that the bulk polar-
ization has the same spatial range as the macroscopic linear field.
In principle, -y(w) and ((w) must be calculated from a microscopic response
theory. The single-particle interaction Hamiltonian associated with the radia-
tion field may be written as H' = HI + H2 , where
e (p·A+A.p
-- - - ) , (5.3)
2mc
e2 ..... .....
-A·A. (5.4)
2mc2
5.1. PHENOMENOLOGICAL CONSIDERATIONS 215

The gauge is chosen so that the scalar potential vanishes. In the absence of
interband transitions, the induced nonlinear polarization can be derived clas-
sically by considering the force acting on a single electron in the homogeneous
medium:
(5.5)

where we have used the relation fA = (iw/c)A. Identifying f~ with the mo-
mentum change (d/dt)mv2, and using J2 = env2, we find for the nonlinear
bulk polarization

(5.6)

The same result is obtained by considering the equation of motion of the


electron fluid (Sipe et al., 1980; Corvi, Schaich, 1986):

The quantum pressure p accounts for the internal properties of the electron
gas. In the Thomas~Fermi approximation, it has the form

p(r, t) = (0 [n(r, t)]5/3 , (5.8)

where (0 = (3rr 2)2/3fj,2/5m follows from the kinetic energy of the homogeneous
electron gas (see (2.27)). From Faraday's law and the identity

(5.9)

we find for the total second-order polarization


-B -B
P2! + P2 , (5.10)
x(w) E2 , (5.11)
'Y(w) V(El . E 1 ) . (5.12)

Here, Pi{ is the linear harmonic polarization and Pf the nonlinear contribu-
tion. The coefficients are given by

e2 fi &-1
x(w) - - -2 (5.13)
4mw 47r
e3 fi e &-1
'Y(w) --- - -2 - - (5.14)
8m 2 w4 2mw 4rr
216 CHAPTER 5. NONLINEAR OPTICS

with e == f(2w). The second term on the right hand side of (5.9) vanishes since
we are considering interaction with a single transverse wave. The nonlinear
bulk polarization is seen to be longitudinal. Hence, it does not radiate unless
it can couple to transverse waves at surfaces or interfaces.
The above derivation applies to a homogeneous electron gas. If lattice ef-
fects are taken into account, a quantum mechanical treatment based on second-
order perturbation theory yields the following modifications of the classical
result (Cheng, Miller, 1964; Jha, Warke, 1967):
• The isotropic part of the nonlinear bulk polarization has the same form
as in (5.12), except that ,(w) is modified. Far from resonances caused by
interband transitions, ,(w) :::::; e(e - 1)/(87rm*w2 ), where e is now the actual
dielectric function and m* the optical mass. Close to an interband transition,
a correction term must be added that exhibits spectral features whenever w or
2w become resonant. (For Drude systems, (e -1)/47r may also be written as
[c(w) - 1]/167r. The derivation by Jha and Warke shows that the expression
(5.14) is the preferred one if band effects are present.)
• An anisotropic bulk polarization appears as indicated by the second term
in (5.2). The coefficient ((w) may also become resonant when w or 2w are near
an interband transition.

5.1.2 Surface Polarization

At a surface, the inversion symmetry is broken. Thus, both dipolar and


higher-order multipole contributions to the nonlinear polarization are allowed.
The surface polarization may in general be obtained from the expression

P~(r,w) = fd3r'ld3r" Xijk(r',r",w) E1j(r',w) Elk(r",w)

+ fd 3 r' Xij(r',2w) E2j (r',w). (5.15)

Contributions to this polarization arise from the rapid variation of electronic


states and electric fields. The first term on the right-hand side represents
the direct polarization generated by the self-consistent linear electric field,
while the second term describes additional linear screening processes at the
harmonic frequency. Because of the rapid variation of the linear and nonlinear
fields in the surface region, an explicit multipole decomposition of the surface
polarization in terms of field gradients as in (5.1) is not practical. The full
microscopic spatial variation used here implicitly includes these multipoles to
arbitrary order.
5.1. PHENOMENOLOGICAL CONSIDERATIONS 217

In the case of a semi-infinite homogeneous electron gas, there are only two
finite surface polarizability elements giving rise to parallel and perpendicular
polarization components. Along the surface, linear fields vary like exp(ic]jl . Til)
and nonlinear ones like exp(i2c]j1 . Til)' We do not show this dependence explic-
itly.
Consider first the parallel surface polarization. The x - z plane is taken
as the plane of incidence. Since a parallel field does not induce charges, we
again use the equation of motion (5.7). The parallel surface polarization can
be derived from the current component j2", = enl VI"" where nl is the linear
charge density induced by E lz and VI", is the parallel component of the linear
velocity field. Using Gauss's law and (d/dt)mvI'" = eEl"" we find (for brevity,
we omit momentum and frequency arguments)

s
P2",(Z) e
= --8 d ()
- - 2 EI",(z) -d Elz z . (5.16)
7rmw z
The gradient of E lz is localized near the surface. Integrating over the surface
region Zl ... Z2 (Zl ~ 0 and Zz » 0) yields

(5.17)

The fields here are microscopic. Outside the surface region, they coincide with
the classical Fresnel fields. In the long-wavelength limit, we may replace Zl by
0- and Z2 by 0+ with the convention that all fields with arguments o± are the
classical fields just above and below the surface. Thus,

(5.18)

This expression shows that the integrated parallel surface polarization of the
homogeneous electron gas depends on purely macroscopic quantities. As in
the case of the bulk polarization, lattice effects of real metals not only modify
e(W) and the effective electronic mass, but also give rise to new resonances
when w or 2w are close to an interband transition.
In contrast to the parallel surface polarization, the perpendicular compo-
nent cannot be reduced to macroscopic quantities, even for the semi-infinite
electron gas. This may be seen as follows. In analogy to (5.16), we may use
the equation of motion (5.7) to write

sed
Pz'(z) = -87rmw 2 Elz(z) dZEIZ(z) + ... , (5.19)
218 CHAPTER 5. NONLINEAR OPTICS

where the additional terms depend on details of the equilibrium density and
the dynamical screening processes. A hydrodynamic description of these mi-
croscopic properties, in general, is not sufficiently accurate. Besides, if we
integrate (5.19) over the surface region, it is clear that the result is not related
in any simple manner to macroscopic response properties. In Section 5.2, we
use the TDLDA to determine the actual spatial distribution of piz (z, w) as a
function of frequency. In analogy to (5.18), we express the integrated surface
polarization in the form

(5.20)

All microscopic effects are implicitly contained in the polarizability element


xzzz(w).
A frequently used alternative parametrization of the surface polarization
was introduced by Rudnick and Stern (1971). Extracting the isotropic bulk
coefficient ,(w) '" n/ w\ we define dimensionless surface coefficients via

2,a(w) , (5.21)
2,b(w) . (5.22)

According to these definitions, the surface polarization is given by

2,a(w) Eiz(O-) , (5.23)


2,b(w) 2 E1",(0-) E1z(0-) . (5.24)

Using this notation, we express the isotropic bulk polarization (5.12) as

(5.25)

The understanding here is that, == -e 3 n/8m 2 w 4 (see (5.14)) and all lattice
effects are included in the coefficients a(w), b(w) and d(w). We emphasize,
however, that this is only a formal way of rewriting the surface polarizability
elements XZZZ) X"''''Z and the true isotropic bulk polarization coefficient ,(w).
These are rigorously determined by the spatial distribution of the nonlinear
surface and bulk polarization sources. (The nonlinear bulk coefficient d( w)
should not be confused with the linear surface parameters d.L(w) and dll(w)
discussed in Chapter 4.)
For a Drude metal, it follows from (5.12) and (5.18) that, independently
of w, b(w) == -1 and d(w) == 1. On the basis of estimates for simple surface
electronic structure models, the magnitude of the dimensionless coefficient a( w)
was for a long time thought to be also of order unity (Rudnick, Stern, 1971; Sipe
5.1. PHENOMENOLOGICAL CONSIDERATIONS 219

w~i/ I

/ / / / / : \ pS/// X
'~
pB
Figure 5.1: Schematic illustration of nonlinear bulk and surface
polarization sources and of generated harmonic radiation.

et al., 1980). Experiments for Al and Ag in contact with quartz (Quail, Simon,
1985) seemed to support these estimates. In contrast, microscopic TDLDA
calculations show that la(w)1 is typically 10 - 50 for simple metals (Weber,
Liebsch 1987aj Liebsch, 1988j Liebsch, Schaich, 1989). Recent measurements
on clean Al surfaces (Murphy et al., 1989j Janz et al., 1991a) agree with these
quantum mechanical results. For alkali metal adsorbates, the magnitude of
a(w) can be as large as 102 - 103 (Liebsch, 1989). In addition, both real
and imaginary parts of a(w) exhibit spectral features reflecting the dynamical
screening properties of the surface. The relative size of this term compared
to the bulk and parallel surface polarizations, and its sensitivity to surface
conditions, indicate that understanding the perpendicular surface polarization
is particularly important for using SHG as a surface diagnostic tool.

5.1.3 Harmonic Radiation

As previously mentioned, the nonlinear bulk polarization has the same spa-
tial extent in the metal as the linear electric field. On the other hand, the
surface polarization reflects the rapid variation of the electronic states and the
normal components of the electric field near the surface. A remarkable aspect
of second harmonic generation is that, although the long-range bulk and short-
range surface polarizations have fundamentally different physical origins, both
contribute to the emitted radiation to the same order in q, where q is the wave
vector of the incident light.
The harmonic radiation generated by the bulk and surface polarizations
(see Figure 5.1) can be derived from the boundary conditions following from
220 CHAPTER 5. NONLINEAR OPTICS

Maxwell's equations. The procedure is the same as for the linear reflectivity
discussed in Section 4.2.1, except that we ignore here small corrections of the
relative order of the linear d-parameters. For simplicity, we consider the case
of incident and emitted p-polarized radiation. The wave vector of the reflected
beam is Q = (Q""O,Qz) = 2q, with Q = 2q = 2w/c. Inside the metal, the
wave vector is Q' = (Q"" 0, -Q~), with Q' = QY'£.
From Ga.uss's law we have

(5.26)

Moreover, the z-component of the wave equation (4.11) reads

(5.27)

Integrating these relations over the surface region, we find

D 2z (Z2) = D2z (zt) - iQ", (". dz D 2",(z) , (5.28)


JZ1
E 2",(Z2) E2",(Zt} + iQ", (". dz [E2Z (z) - (Q/Q",)2 D 2Z (z)] (5.29)
JZ 1
The fields appearing in these relations are related to the total polarization via

(5.30)

Radiation Induced by Surface Polarization

We consider first the fields associated with the nonlinear surface polariza-
tion which we assume to be localized in the surface region Zl ... Z2. Thus, the
displacement field is given by

(5.31)

Eliminating D 2", from (5.28), and E 2z from (5.29), we find

D2z (Z2) = D2z (zt) - iQ", 1~2 dz [411P~(z) + E 2",(z)]


~ D 2z (Zl) - 47riQ", f>~ , (5.32)
E2",(Z2) E2",(Zl) - iQ", 1~2 dz [47rP2~(Z) + (Qz/Q",)2D 2Z (z)]
(5.33)
5.1. PHENOMENOLOGICAL CONSIDERATIONS 221

We have assumed here that the main contributions to [D 2z (Z2) - D2z (zdl and
[E2",(z2) - E2",(zdl are caused by the large but rapidly varying surface po-
larizations P:f.,(z) and ptz(z), respectively. Higher-order terms due to the
small and slowly varying fields E2"'(Z) and D2z(Z) are neglected. The rela-
tions (5.32) and (5.33) show that, in the long-wavelength limit, the asymptotic
behavior of the harmonic fields is fully determined by the integrated weight
of the harmonic surface polarization. ptz and P:f., are analogous to the linear
parameters dl. and dll which, according to (4.64) and (4.65), can be viewed as
normalized integrated weights of the linear surface polarization.
The fields D 2z (Zi) and E 2",(Zi) in (5.32) and (5.33) are microscopic fields.
Outside the surface region, they coincide with the classical Fresnel fields. In
the long-wavelength limit, these relations can therefore be written as

E2z(0+) (5.34)
E2"'(0+) = (5.35)

Radiation Induced by Bulk Polarization

The displacement field related to the total bulk polarization is given by

(5.36)

To satisfy the wave equation

(5.37)

we now make the ansatz


(5.38)
where subscripts T and L refer to transverse and longitudinal waves, respec-
tively. Since P2B is longitudinal (the explicit expression is given in (5.56)), we
can easily verify, by substituting this ansatz into (5.37), that the longitudinal
part of the electric field has the form

(5.39)

This implies
(5.40)
Hence, the harmonic displacement field associated with the bulk polarization
is purely transverse.
222 CHAPTER 5. NONLINEAR OPTICS

In the long-wavelength limit, we obtain from (5.28) and (5.29)

D2z(Zt} , (5.41)
E 2",(Zl} . (5.42)

Using (5.38) - (5.40), and replacing the microscopic fields at Zi by classical


Fresnel fields just above and below the surface, we find

(5.43)
(5.44)

These relations show that only the parallel component of the nonlinear bulk
polarization contributes to the harmonic radiation.

Total Radiation

If the combined effect of surface and bulk polarizations is considered, the


total harmonic radiation is determined by the following equations

(5.45)
(5.46)

For convenience, we omitted the label T since all electric fields appearing in
these relations are transverse. In the case of emitted p polarization, the electric
field has the form

E2 (z} Rp ( - Qz, 0, Q)
Q x e iQ.z , Z» 0, (5.4 7)

Tp
QI
(Q'z" 0 QX) e-iQ~z , z~O , (5.48)

where Rp and Tp are the amplitudes of the reflected and transmitted waves.
Substituting these expressions into (5.45) and (5.46) and eliminating Tp , we
find
(5.49)

To proceed we write the surface polarization in terms of macroscopic linear


internal fields, as indicated in (5.24) and (5.23). According to (4.22), for
p-polarized incident light, the linear field inside the metal is given by

E1 () t (,
Z = -; qz' 0, q", ) e -iq'• z , (5.50)
q
5.1. PHENOMENOLOGICAL CONSIDERATIONS 223

where
t - 1 + r _ 2qz..fi 2..fi (5.51)
- ..fi - cqz + q~ c+s
is the linear transmission coefficient. The quantity s(w) is defined as

s(w) = q~ = [c(w) - sin 2 0]1/2 , (5.52)


qz cos 0
where 0 is the polar angle of incidence. With these definitions, the internal
field components may be expressed as

(5.53)

Hence, the integrated surface polarization components are

4 sin2 0
a(w) 2, (c+s )2 ' (5.54)

b() 8s sin 0 cos 0


(5.55)
w 2, (
c+s )2.
Using the linear internal electric field specified in (5.50), the nonlinear bulk
polarization defined in (5.25) takes the form

d(w)r [E;",(O-) + E;z(O-)] Ve2i(qz",-q~z)


d(w)rt2 2i (q"" 0, -q~) e2i(qz",-q~z) . (5.56)

The amplitude of the x-component just inside the surface is

(5.57)

We now collect bulk and surface terms and substitute (5.54), (5.55), and
(5.57) into (5.49). The reflection amplitude of the harmonic radiation is then
given by the expression

Rp = 647riq,c tan 0 A (5.58)


(c+s)2(e+S) P'
where
. 2n17 _ b(w)2sScos2n17
AP = a (w ) -£ Sin + d(w) (5.59)
c c 2
and
S(w) = Q~/Qz = s(2w) . (5.60)
224 CHAPTER 5. NONLINEAR OPTICS

Second Harmonic Generation Efficiency

It is customary to evaluate the second harmonic generation efficiency 12w /1;


(Sipe et at., 1980), where
c ~ + 2
Izw = 2 471' IE2 (O )1 , (5.61)

are the intensities of the reflected second harmonic and incident fundamen-
tal fields. According to the notation introduced above, IEin(O+) 1 = 1 and
IE2 (O+)1 = IHpI. Thus,
12w _871'e 2 1 c(c - 1) tan 8 A 12 (5.62)
13 - m 2w2 c3 (c + 8)2(& + S) p
At 1 eV, the prefactor 871'e 2 /m 2w2c3 has the value 113 x 10- 20 cm 2/W =
113 x 10- 27 cm 2 s/erg. We used here the relation I = (c: - 1) e/(3271'mwZ) (see
(5.14)), which is valid for Drude metals.
The parametrization of the second harmonic efficiency in terms of the di-
mensionless coefficients a, band d is convenient since these quantities depend
only on frequency. The microscopic optical properties of the system are repre-
sented via these coefficients. All polarization and angle effects are contained in
the remaining Fresnel factors. The above expression applies to the case of in-
cident and reflected p-polarized light. The derivation can easily be generalized
to other incident and reflected polarization configurations. In the case of inci-
dent mixed s-p polarization, the harmonic generation efficiency of p-polarized
reflected radiation is given by

12w 871'e z 1 c:(c-1}tan8 ( 2 ·2 )1 2


13 = m 2w2 c3 (c: + 8)2(& + S) Ap cos 'IjJ + A. sm 'IjJ , (5.63)

where 'IjJ specifies the angle of the incident polarization vector with respect to
the plane of incidence and A. is defined as

A _ d(w) (c: + 8)2


(5.64)
•- 2 c(l + 8)2 .
In the limit of low frequencies, Ap and A. reduce to

Ap ~ a~w) sin28 _ b(w) + d~} , (5.65)

A. ~ d~W) cos 2 B . (5.66)


5.1. PHENOMENOLOGICAL CONSIDERATIONS 225

2.0 40

~ AI 0=1
"- 1.5 30
E 0
," ,
,, ,,
0
:;: (0) (b)
I
1.0 ,,, ,, 20 -30
,, ,,
0
,
.-'
.::::..
,
,,
.,,
I

.,.,
,,
,,
I
,,
"- 0.5 , -1 ,
10
3 /
/ -20 \
/
/
/

0.0 0
40 60 80 40 60 80

angle of incidence angle of incidence

Figure 5.2: Second harmonic generation efficiency 12w / as a func- I;


tion of incident polar angle. (a) a(w} = 1,0, -1; (b) a(w} =
-40, -30, -20. In all cases, w = 1.17 eV, b = -1 and d = 1.
The dielectric functions are for a Drude metal corresponding to AI.

This adiabatic expression of Ap indicates that surface and bulk polarization


sources contribute with roughly the same weight to the radiation amplitude.
Towards grazing incidence, the a-term dominates.
To illustrate the variation of the harmonic intensity with polar incidence
angle, we show in Figure 5.2 the p-polarized efficiency for several cases: (a)
the perpendicular surface polarizability is assumed to be small, with a(w} of
order unity; and (b) a(w} is of the order of 20 - 40, corresponding to the actual
magnitude for AI. A more detailed discussion of SHG from Al is given later.
Here we merely point out that, even in the absence of perpendicular surface
polarization (see the curve for a = O), the intensity is strongly peaked at large
angles, since the emitted radiation couples only to the parallel component of
the bulk polarization (see (5.49)). Besides, the parallel surface polarization
gives nearly the same angular dependence as the bulk term. The normal
surface polarization leads to a variation that is only slightly more pointed at
large angles, since the factor sin 2 () does not change much for () > 60°. Figure
5.2 shows that the true surface polarizability of Al gives intensities that are
more than an order of magnitude larger than those obtained for a(w) of order
unity.
226 CHAPTER 5. NONLINEAR OPTICS

Beyond Jellium Model

Until now, we dealt with the nonlinear response of a semi-infinite ho-


mogeneous electron gas. For real metals, new surface and bulk polarizabil-
ity elements appear as a result of interband transitions. A complete listing
of these terms for cubic centrosymmetric crystals was given by Sipe et al.
(1987). In the case of the (111) face of fcc crystals, the two additional surface
components are the isotropic element Xz"",,(w) (Guyot-Sionnest et al., 1986),
and the anisotropic element X",,,,,,,(w). In analogy to Xzzz(w) and X",,,,z(w), we
parametrize these components as

Xz"",,(w) = 2, f(w) , (5.67)


X",,,,,,,(w) 2,€(w) , (5.68)

where, is defined in (5.14). The radiation amplitude arising from these ele-
ments can easily be obtained from the above formalism by realizing that the
new contributions to the integrated surface polarization (for p-polarized inci-
dent light) are similar to expressions (5.54) and (5.55), except that Pfz must
be multiplied by
E2 (0-)
1", = S2 cot 2 () (5.69)
Elz(O-) ,
whereas Pf., must be multiplied by

(5.70)

The amplitude Ap defined in (5.59) is then generalized to Ap + Bp cos(3¢1),


where the isotropic part is replaced by
e.
Ap = a(w)- sm 2 () -
2sS
b(w)- cos 2 ()
d(w)
+ -2- + e
f(W)-S2 cos 2 (), (5.71)
e: e: e:
and the anisotropic amplitude is

B =_€(W)s2SC~S3(). (5.72)
p e: sm ()
The angle ¢I denotes rotation of the crystal about the surface normal. The bulk
magnetic dipole term d(w)/2 is also augmented by isotropic and anisotropic
electric quadrupolar contributions.
Note that the second harmonic efficiency is proportional to the coefficient
5.2. MICROSCOPIC THEORY 227

at the harmonic frequency. As a consequence, 12w exhibits a maximum near


the minimum of Ic(2w)l. This behavior depends only on the linear response
properties of the metal; it was observed in SHG measurements from Ag that
exhibit a pronounced maximum near the bulk plasma frequency (see Section
5.3.4). Hence, the frequency dependence of 12w caused by the macroscopic pa-
rameters in expression (5.62) can be significant and may outweigh the structure
associated with nonlinear response parameters.

5.1.4 Validity of Long-Wavelength Limit

In the preceding derivation of the emitted harmonic radiation, we used the


long-wavelength limit at various stages. For a realistic metal surface, a com-
plete solution of Maxwell's equations beyond the long-wavelength limit would
be rather complicated and has not yet been attempted. The only fully retarded
solution was obtained by Corvi and Schaich (1986) who applied hydrodynamic
theory to model density distributions consisting of a single and a double step.
The intensities were found to be in excellent agreement with the parametrized
expression (5.62) in which the coefficients a, b, and d were also derived from
a retarded calculation. In addition, the validity of b ~ -1 and d ~ 1 was
confirmed.
Subsequently, Schaich and Liebsch (1988) proved that, to very good accu-
racy, a(w) can be derived by neglecting retardation effects. This finding is of
considerable practical importance. It suggests that, in analogy to d.l(w), the
perpendicular surface polarization can be evaluated from the density response
to a uniform field instead of the density and current responses to arbitrary
fields. Although this result was established only for the hydrodynamic model,
it seems plausible that the electrostatic limit is valid more generally because
retardation effects do not depend on quantum mechanical corrections.

5.2 Microscopic Theory


5.2.1 Nonlinear Response

In this section, we focus on the frequency dependence of the perpendicular


surface polarization Pf.(w). As argued above, we may evaluate this component
by neglecting retardation effects. The quantity of interest is therefore the
harmonic density induced at the surface by a uniform electric field oriented
228 CHAPTER 5. NONLINEAR OPTICS

normal to the surface. In the spirit of the surface excitations discussed in the
preceding chapters, we determine the nonlinear surface charge density within
the TDLDA. This approach ensures that electron-electron interactions in the
ground state and linear and nonlinear response regimes are treated on the same
footing.
From standard second-order perturbation theory (Langhoff et al., 1972), it
follows that the induced density is given by (Zangwill, 1983; Senatore, Sub-
baswamy, 1987)

b"n(z,w) fa;:; f dZ'XI(Z,z',W')cf>scf(Z',w')27rb"(w-w')

+ f ~~ f d2~' f dz'f dz" X2(Z, z', z", w', w")


X cf>scf(Z', w') cf>scf(Z", w") 27rb"(w - w' - w") , (5.74)
where Xl and X2 are the first- and second-order independent-particle suscep-
tibilities of the semi-infinite electron gas specified in (2.111) and (2.112). All
screening effects are incorporated into the complex local potential cf>scf.
The first-order density nl(z,w) == b"n(z,w) is determined by the following
response equations (see Section 4.2.5)

nl(z,w) f dz' XI(Z, z' ,w) cf>I,scf(Z', w) (5.75)


cf>l,scf( z, w) cf>ext(z, w) + cf>l,ind (z, w) (5.76)
cf>I,ind(Z,W) f dz' K(z,z') nl(z',w) (5.77)
K(z, z') -27rIZ - z'l + V:c[no(z)]b"(z - z') . (5.78)
The applied potential is defined as cf>ext(z,w) = -27rz.
To find the second-order density n2(z,w) == b"n(z,2w), we consider those
terms in (5.74) that oscillate at frequency 2w. This density obeys the response
equation

n2(z, w) = f dz'f dz" X2(Z, z', z" ,w, w) cf>I,scf(Z', w) cf>l,scf(Z", w)

+ f dz' Xl(Z, z', 2w) {0.5V:~[no(z')]ni(z' ,w) + cf>2,ind(Z' ,w)}


(5.79)
The second-order induced potential is given by

cf>2,ind(Z,W) = f dz' K(z, z') n2(z',w) , (5.80)

and V;~[no(z)] is the second derivative of the exchange-correlation potential


Vxc with respect to the ground-state density.
5.2. MICROSCOPIC THEORY 229

5.2.2 Relation between Linear and Nonlinear Responses

It is evident that the linear and nonlinear response equations (5.75) and
(5.79) have a similar structure. In fact, both may be written in the form
(v=1,2)

!
n... (z, w) = ~. . (z,w) + dZ,/ dz" XI(Z, Zll, vw) K(Z", z') nv(z', w) , (5.81)

where ~... are the first- and second-order bare or unscreened induced densities

6(z,w) ! dZ'XI(Z,Z',W) <Pext(z',w) , (5.82)

~2(Z,W) ! dz' ! dz" X2(Z, z', Zll, W, w) <PI,scf(Z', w) <PI,scC(Z", w)


+ ! dz' XI(Z, z', 2w) 0.5V~~[no(z')1 n~(z', w) . (5.83)

Schematically, both response equations can be expressed as (we temporarily


omit the spatial coordinates and integrations):

n.,(w) = [1 - XI(VW) Kti ~v(w) . (5.84)

Thus, whenever there is a resonance at Wr in the linear excitation spectrum,


there is an analogous resonance in the quadratic response at 0.5 wr • Two
photons are now required to generate this resonance. Of course, the spectral
shape of these features differs because of the different driving terms defined
by the unscreened densities ~.,(w). Later we see that the linear and nonlinear
surface excitation spectra of simple metals can indeed be understood in terms
of the same physical processes.

5.2.3 Surface Polarization

The linear and quadratic dipole moments associated with the induced elec-
tronic surface densities are defined as (v = 1,2)

P... (w) = i: dz z n.,(z,w) . (5.85)

According to (4.53), PI (w) is proportional to the centroid of the linear screen-


ing charge density dl.(w). From the relation dPiz(z,w)/dz = n2(z,w), it
230 CHAPTER 5. NONLINEAR OPTICS

follows that the second-order moment P2(W) is proportional to the integrated


perpendicular surface polarization:

(5.86)

Here, we used the fact that the surface polarization is confined to the surface
region, since the integrated weight of the second-order density n2(z, w) van-
ishes. According to our definition of the applied field, the classical linear field
just below the surface is given by Elz(O-) = -27f(1- a) = -47f/(c + 1). From
the definition of the parameter a(w), (5.23), we then find

a(w) = -4ft P2(w)/a(w? . (5.87)

5.2.4 Solution of Response Equation

As in the linear case, the numerical solution of the second-order response


equation (5.79) is nontrivial because of the long-range induced Coulomb po-
tential. Its asymptotic behavior must be treated analytically. We are then left
with quantities that are truly localized in the surface region. This procedure
leads to a new response equation that can be solved in real space by matrix
inversion (Liebsch, Schaich, 1989).
Since the integrated weight of the second-order induced density n2 vanishes,
no long-range fields are generated. To ensure that this is satisfied even if self-
consistency has not yet been achieved, it is convenient to write the induced
potential in the form

(5.88)

where, in analogy to (4.108), <P2 is determined by

(5.89)

The asymptotic behavior of the induced potential can be accounted for


explicitly by introducing a trial density ng. For instance, we write

ng(z,w) f+(w)F+(z) + f_(w)F_(z) , (5.90)


<pg(z, w) -47f foo dz' (z - z') ng(z,w) . (5.91)
5.2. MICROSCOPIC THEORY 231

The weights are defined as f±(w) = ±p2(w)/2zo , and F± are the Gaussian
distributions specified in (4.111). The moment P2 is to be determined self-
consistently. Separating n~ from (5.79), we obtain the following response equa-
tion for the remaining density n~ = n2 - n~ :

n~(z,w) = Ll 2(z,w) + / dz' Xl(Z,Z', 2w) [¢>~(z',w) + V:c(z') n~(z',w)] ,


(5.92)

where the 'driving' term is given by

Ll2(Z,W) = 6(z,w)+ /dZ'Xl(Z, z', 2w) [¢>~(z',w)+V:c(z')n~(z',w)]


- n~(z,w) . (5.93)

The functions n~ and ¢>~ satisfy a Poisson equation like (5.89).


Discretizing all functions on the mesh points Zl ... ZN, we can combine
these equations into a 2N x 2N matrix equation:

-X1W ) (n~) (Ll2) (5.94)


1- Kw ¢>~ 0

Here, Xl denotes the matrix Xl(z;,zj,2w) and Ll2 the vector Ll 2(Zi,W). The
remaining quantities are the same as in the corresponding first-order response
equation (4.114). Since (5.94) and (4.114) have the same structure, we can use
the same computational procedure to obtain the solution.
It is very important to evaluate accurately the unscreened induced density
6(z,w), (5.83). In the asymptotic range z' ~ Zl, the ±(2w,w) and (w,-w)
contributions to X2 in (2.112) lead to near-cancellations of very large terms.
This problem can be handled by first carrying out the z'-integral, which can
be done analytically using the exact asymptotic behavior of the bound states
and Green's functions. The integration over occupied states is then performed
numerically.
The final issue related to the numerical evaluation of the surface parameter
a(w) concerns the nonlinear induced dipole moment P2(W) defined in (5.85).
Because of the slowly decaying Friedel oscillations of the induced density in
the interior, this definition of P2(W) is computationally impractical. As in the
linear case, this difficulty can be overcome by using the dynamical force sum
rule. The extension of this rule to the quadratic response is outlined in the
following section.
The procedure just presented is stable even at low frequencies. Close to the
adiabatic limit, the results coincide with those obtained by Weber and Liebsch
232 CHAPTER 5. NONLINEAR OPTICS

(1987a), who determined the nonlinear response by solving the Schrodinger


equation in the presence of a uniform static electric field.

5.2.5 Dynamical Force Sum Rule

In Section 4.2.5, we saw that the dynamical force sum rule can be used to
relate the total induced dipole moment to the moment of the density outside
the jellium background. From the equation of motion (4.119), we obtain in
second order (Liebsch, Schaich, 1989)

This relationship holds as long as the total number of electrons in the metal
is kept constant. In the case of a neutral slab, however, n2(z,w) is symmetric
with respect to the slab center, so that all terms in this equation vanish for
symmetry reasons. It is therefore necessary to consider a grounded slab to
which the induced surface charge density is supplied externally. This implies
that the number of electrons in the metal oscillates with time. Since this
in- and outflow of electrons does not require any force, the expression (5.95)
should be corrected (Schaich, 1995).
The flux of momentum flowing into the system may be expressed as a
product of the momentum mVl and the flux nVl at some position deep inside

i:
the metal. The flux is given by

nVl = -iw dznl(Z,W) = -iwO'(w) , (5.96)

where 0' is the integrated linear surface charge density specified in (4.98). This
relation follows from jl = envl = -iwPl and P l = eO' far from the surface.
The convective momentum flux to be subtracted from the left-hand side of
(5.95) is then given by m Vl n Vl = -m w20'2 In. Converting again to atomic
units, the second-order sum rule becomes

(5.97)

Turning now to the terms on the right-hand side of (5.97), we can use our
definition E ext = -211' and express the first contribution as

! dz nl(z,w) Eext(z,w) = -211' O'(w) . (5.98)


5.2. MICROSCOPIC THEORY 233

The second term involving the field due to the positive background may be
reformulated as

1 dz n+(z) cP~(z,w)
47rn [P2(W) - P2+(W)] (5.99)

where PH(W) is the external moment

PH(W) = I: dz (z -~) n2(z,w) . (5.100)

We assumed here that the positive background is abruptly terminated at the


distance z =~. In the case of a neutral jellium system, ~ = o. For strong
positive charging, ~ » 0, so that P2+(W) = o. The above relation can easily
be generalized to other positive background profiles such as those appropriate
for jellium overlayers.
Inserting these expressions into (5.97), and using the Drude formula for the
dielectric function, we may rewrite this relation in the form
c:(2w) - 1 a 2(w)
p2(W) = c:(2w) PH(W) + 2n . (5.101)

Finally, according to (5.87), the a-parameter is given in terms of PH via


4n c:(2w)-1
a(w) = -~() (5.102)
a w c: (2w) P2+(W) - 2.
In the limit of strong positive charging, P2+(W) = 0 and a(w) = -2 in-
dependently of w. This limiting behavior agrees with the result found by
Schaich (1994) for highly charged metallic slabs. Without the flux correc-
tion, the a-parameter has the same first term, but the final -2 is replaced by
-[c:(w) + 1]/[2c:(2w)] (Liebsch, Schaich, 1989). The difference between these
two versions is anew - aold = -1/c:(2w). This difference is negligible at low
frequencies but becomes large as 2w approaches the plasma frequency. Near
the multipole surface plasmon (2w ~ Wm ~ 0.8wp ), the difference is about 2,
which is much less than the actual values of a in this frequency range (see
below). The flux correction has therefore only a weak influence on the main
part of the nonlinear surface excitation spectrum.
We note here that, in the adiabatic limit, the sum rule (5.101) implies

(5.103)

This relation agrees with the result derived by Budd and Vannimenus (1975)
for the static second-order induced dipole moment.
234 CHAPTER 5. NONLINEAR OPTICS

The derivation of (5.102) relies on a consistent treatment of electron-


electron interactions in the ground state and in the presence of the time-
dependent field. The TDLDA provides such a consistent formulation. How-
ever, if the response is described within the RPA while the LDA is used to
evaluate the ground state, the sum rule is violated. The exchange-correlation
potential Vxc[no(z)] then acts like a rigid external zero-order potential that
must be added to the right-hand side of (5.97). Instead of (5.102), we find the
modified sum rule

4fi c(2w) - 1 {
a(w) = -~()
<7 W
()
c 2w
1
P2+(W) - 2'
wp
/00 dzn2(Z,W) Vxc[no(z)]
-00
'} - 2.
(5.104)

5.3 Simple Metals


5.3.1 Adiabatic Response

If the laser frequency employed in SHG studies is low compared to the


single-particle or collective electronic excitations of the sample, the dynamical
response is dominated by near-adiabatic behavior. For this reason, it is useful
to investigate the dipole moments induced by a static electric field oriented
normal to the surface. In Section 2.1.4, we mentioned this case since the
relevant electronic density in the adiabatic limit can be calculated directly from
the Schri:idinger equation in the presence of an external potential. Existing
ground-state electronic structure codes can therefore be easily generalized to
evaluate the coefficient a(O). Such extensions would be valuable for estimating
the influence of lattice effects. Their role in nonlinear optical processes is so
far poorly understood.
For simple metals, static calculations of the a-parameter were carried out
by Weber and Liebsch (1987a) within the standard jellium model. According
to (5.87), at low frequencies, a(w) approaches the value

a(O) = -4fiP2 , (5.105)

since <7(w) -+ 1. Here, fi is the bulk density, and P2 is the static nonlinear
induced dipole moment. Microscopic calculations showed that the magnitude
of a for realistic equilibrium density profiles is much larger than unity. Table
5.1 lists the values of a(O) for several jellium surfaces and different electronic
structure treatments. The large range of values underscores the remarkable
sensitivity of a to details of the equilibrium density profile. Non-self-consistent
5.3. SIMPLE METALS 235

Table 5.1: -a(O) for various surface response treatments: LDA-


j: standard jellium model (Weber, Liebsch, 1987a); LDA-sj: sta-
bilized jellium model (Kienja, 1995); LDA-RPA: ground state
LDA standard jellium, response RPA; FBM-RPA: finite-barrier
model in RPA; IBM-RPA: infinite-barrier model in RPA (Liebsch,
Schaich, 1989); one-step: single-step density, hydrodynamic re-
sponse (Corvi, Schaich, 1986); quadratic: quadratic-density, hy-
drodynamic response (Maytorena et al., 1995); TFW(A): Thomas-
Fermi-von Weizsacker approach with A = 1/4 and A = 1
(Chizmeshya, Zaremba, 1988).

r. 2 3 4 5
LDA-j 28.4 12.9 8.6 6.6
LDA - sj 18.9 11.3 8.9 7.5
LDA-RPA 22.1 10.0 6.6 5.0
FBM-RPA 7.0 4.0 3.2 2.8
IBM-RPA 0.8 1.0 1.2 1.5
One - step 2/9 2/9 2/9 2/9
Quadratic 9 9 9 9
TFW(1/4) 31.8 18.2 12.7 9.8
TFW(l) 34.7 14.9 8.9 6.2

densities based on finite- or infinite-barrier models tend to give significantly


smaller values, since they underestimate the polarizability of the surface elec-
tronic density. Of particular interest is the comparison between LDA results
for the standard and stabilized jellium models (Kiejna, 1995). In the case of
Al (r. = 2), the average pseudopotential increases the work function by about
0.5 eV (see Tables 2.1 and 2.2), so that the density profile becomes less polariz-
able. Accordingly, the magnitude of a(O) is 1/3 less than for standard jellium.
The change in the case of the alkali metals is much smaller because of weaker
pseudopotentials. For r. > 4, stabilization leads to smaller work functions,
and a(O) is slightly enhanced.
We also draw attention to the results obtained using the Thomas-Fermi-
von Weizsacker (TFW) approach (Chizmeshya, Zaremba 1988). In this scheme,
the total energy of the semi-infinite metal is expressed as a functional of the
electronic density. However, in contrast to the LDA, a gradient expansion is
236 CHAPTER 5. NONLINEAR OPTICS

used for the kinetic energy (atomic units):

T[nol = !dz {Cdno(z)l5/3 + n~:) [dn;;z)t} , (5.106)

where 0 1 = (3/10) (311"2)2/3 and C2 = >./8. The remaining Coulomb and


exchange-correlation contributions to the total energy are the same as in
expression (2.1) for the LDA. The first term in (5.106) corresponds to the
standard Thomas-Fermi theory while the second is the von Weizsacker gradi-
ent correction. The coefficient 0 1 is related to (0 in expression (5.8) for the
quantum pressure via 0 1 = 3(0/2 a.u. The parameter >., which specifies the
importance of the gradients terms, is taken to be adjustable. Incorporating
this correction is a prerequisite for a sufficiently accurate description of elec-
tronic surface properties. For example, the gradient correction ensures that
the density exhibits proper exponential decay. This improvement is particu-
larly important for the surface polarizability. For metallic surfaces, ).. = 1/4
was found to provide good results for the work function, surface energy, and
other properties.
As shown in Table 5.1, the TFW values of a(O) compare quite well with
the LDA values (both for the standard jellium model) because the equilibrium
density distributions are also well reproduced. In contrast, the hydrodynamic
model applied to the single-step density profile (Corvi, Schaich, 1986) gives
a(O) = -2/9 independently of r.. Slightly larger values are obtained for
two-step profiles. The comparison with the TFW results suggests that this
underestimate is a consequence of the inadequate assumed density, not of the
hydrodynamic treatment of the nonlinear response.
Support for this conclusion also comes from recent hydrodynamic calcula-
tions by Maytorena et al. (1995) for quadratic density profiles. If the width of
the surface region over which the density decays is adjusted to fit the frequency
of the multipole surface plasmon, these authors find a(O) ~ -9 independently
of r •. Less polarizable linear profiles yield a much smaller value: a(O) ~ -3.
The important message from these different response treatments is that the
nonlinear polarizability is very sensitive to the shape of the electronic density
and the screening properties in the surface region.
An example of the static second-order induced polarization P2(Z) is shown
in Figure 2.5. According to Table 2.3, the centroid Z2 of P2(z) lies even further
out in the vacuum than the static image plane Z1. Hence, as a result of the
efficiency of metallic screening, the perpendicular component of the surface
polarization is located mainly above the positive ionic background. It is not
surprising, therefore, that this component is particularly sensitive to the tail
of the equilibrium electron density.
5.3. SIMPLE METALS 237

60
r, =2
40
(\Im
Re I \ r, =4
\
\

" I O~~~~_\p'~~~~
\
-20
-20

40
40
/--_),m 20 r, =5 Re
20 Re /1 ........ _ . . -\

] O~~-~-+/-+~~~~\~
"I -20
-20
-40
-40

o 0.1 0.2 OJ 0.4 OS -600 0.1 0.2 03 OS


w/wp w/wp

Figure 5.3: Frequency dependence of nonlinear parameter a(w) for


several bulk densities, calculated within TDLDA. Solid curves: real
parts; dashed curves: imaginary parts. The arrows indicate <1>(2,
where <1> is work function. (Liebsch, Schaich, 1989).

5.3.2 Surface Excitation Spectra

The frequency dependence of the parameter a(w) was calculated by Liebsch


and Schaich (1989) within the TDLDA using the procedure described in the
preceding section. As shown in Figure 5.3, two main spectral features can be
distinguished. Comparison with Figure 4.5 suggests that these excitations are
closely related to those characterizing the linear response function dol(w):
• The first feature occurs when 2w approaches the work function <1>. This
peak is the analog of the threshold excitation that we encountered on
several occasions in the linear response. As argued in Section 5.2.2, first-
and second-order responses are governed by the same screening processes.
In the nonlinear case, two photons of w ,::::; <1>(2 are required to generate
this excitation. While this feature is rather weak in dol (w), it is the
dominant nonlinear excitation for high bulk densities.
238 CHAPTER 5. NONLINEAR OPTICS

1.0

"-,

30-

'-... 0.5
3 "": "-""",,,

'''-''

0.0
0.0 0.5 1.0

W, / Wp

Figure 5.4: Map of spectral features observable in sum frequency


generation. Wi and W2 are the fundamental frequencies. The vertical
and horizontal dotted lines correspond to Wi or W2 equal to Wm =
0.8 wp. The dashed lines denote pairs of constant Wi + W2 = Wm
and Wi + W2 = wp. The diagonal dotted line corresponds to second
harmonic generation. The squares denote particularly pronounced
resonances in the nonlinear response .

• The second spectral feature occurs when 2w reaches the multipole plasma
frequency Wm ~ 0.8wp • Evidently, two photons are needed to excite this
collective mode, which is observed in electron energy loss spectra and
photoyield measurements. For low-density metals, the multipole surface
plasmon is more prominent than the threshold excitation.

Additional spectral features should occur at frequencies above wp /2, in par-


ticular, when the fundamental frequency coincides with the multipole surface
plasmon (see Figure 5.4). Such excitations were found by Maytorena et al.
(1997) in hydrodynamic calculations for linear and quadratic density profiles.
Apart from the peak at W = wm /2, these authors obtained new resonances at
W = wp /2 and at W = W m • The latter peak is caused by the multipole exci-
tation at the fundamental frequency. It is more than an order of magnitude
stronger than the one at wm /2. It would be interesting to explore the spectral
region above wp /2 within the TDLDA.
Notice that the frequency dependence of a(w) is not caused by transitions
involving surface states or other bands. Occupied and unoccupied states of
5.3. SIMPLE METALS 239

a semi-infinite electron gas form a smooth continuum, so that only the dy-
namical surface screening processes contribute to the above spectral features.
In real materials, such effects presumably also exist, but they are mixed with
interband transitions.
In the case of AI, which was studied most extensively, laser frequencies
commonly used are much smaller than the plasma frequency, i.e., one is close
to the static limit. One can, however, investigate the region near the threshold
excitation (2w :'S <P :::::; 4.5 eV). In the case of alkali metals, available laser fre-
quencies lie in the range where the harmonic light can couple to the multipole
surface plasmon.

Sum Frequency Generation

The generalization of the TDLDA to sum frequency generation from jel-


lium surfaces has not yet been carried out. Hydrodynamic calculations by
Maytorena et al. (1997) yield an interesting map of spectral features related
to special excitations at the input or output frequencies. As indicated in Fig-
ure 5.4, sum frequency spectra show peaks when one of the fundamentals is at
Wm and the other near zero, W m , or wp - W m . Ridges of large intensity are found
for WI + Wz equal to Wm or wp' At WI = wz, these conditions yield maxima in
second harmonic generation. The rather large weight observed for WI = Wm
and Wz near zero or wp - Wm should make these spectral features ideally suited
for investigating multipole surface plasmons.

Nonlinear Induced Density

The spatial distribution of the nonlinear surface charge exhibits a striking


variation with frequency (see Figure 5.5). Far below the plasma frequency,
the density is localized within a few A of the surface. Only weak Friedel
oscillations extend into the interior. As the harmonic frequency approaches
wp , the amplitude of these oscillations becomes very large, and the density
reaches several tens of A into the solid. Similar behavior is found for the linear
density as W approaches wp.
The preceding results illustrate that, contrary to common belief, the effec-
tive region ZI'" Z2, where the surface polarization is localized, does not only
depend on the ground-state properties. As a result of dynamical screening pro-
cesses, the active range can be much larger than the typical depth up to which
the surface electronic density differs from the bulk density.
240 CHAPTER 5. NONLINEAR OPTICS

-4
-8

-12\---~_ _ _----c~-----i

20 0.2 w,

-20

'"
-I,()

0.4w,
80
40

-80
-40
-120
-80
-60 15 -6O~---4~5---~~--~15~~~1-5
z(a.uJ

Figure 5.5: Spatial distribution of second-order density n2(z,w)


induced by uniform electric field oriented normal to the surface.
(a) real part, (b) imaginary part. The bulk plasma frequency is
9.2 eV (To = 3). (Liebsch, Schaich, 1989).

5.3.3 Comparison with Measured Spectra: Al

Murphy et al. (1989) measured the second harmonic intensity from Al sur-
faces at 1.17 eV as a function of polar incident angle. As shown in Figure 5.6,
their results for AI(111) in the p-in, p-out configuration are in good agreement
with TDLDA calculations obtained from (5.62), with a(w) = -36 - 9i (see
Figure 5.3), b(w) = -1 and d(w) = 1. Data for AI(lOO) and polycrystalline
Al give intensities that are about three times lower.
In comparisons with experimental data, it is important to recall that har-
5.3. SIMPLE METALS 241

,," ,,
~ ,,
;:-.- 40 AI(ll1) ,
E
lilI
0
. AI(100)
0
.::.
N
..1! 20 * polycrystalline AI

"-
.!<

polar angle of incidence

Figure 5.6: Second harmonic efficiency for Al as a function of polar


angle of incidence. w = 1.17 eVj p-in, p-out polarization. Sym-
bols: experimental data (Murphy et at., 1989)j solid curve: TDLDA
results for standard jellium, Drude c(w)j dashed (dotted) curve:
TDLDA results for standard (stabilized) jellium, measured c:(w).

102 r----------------,

-
AI (111)
B=67S

~--
-,-----
e=45 0
""W"

" B=22.5°

"f<t~
1O-~00 600 700 800 900
WAVELENGTH (nm)

Figure 5.7: Second harmonic efficiency for AI(1l1) as a function


of wavelength for three polar angles of incidence and p-in, p-out
polarization (Janz et al., 1991a). Dashed curves: LDA-based RPA
results for standard jellium (Liebsch, Schaich, 1989).
242 CHAPTER 5. NONLINEAR OPTICS

monic intensities are sensitive to the dielectric functions used in the Fresnel
coefficients. If the Drude values [c(w) ~ -180, c(2w) ~ -44] are replaced
by measured values [c(w) = -95 + 33i, c(2w) = -33 + lOi], the intensity
increases by about 50 %. However, to remain consistent, we should then also
replace the jellium a-parameter by the corresponding value for real AI. If the
more realistic surface polarizability derived for the stabilized jellium model is
used, the intensity is lowered appreciably and the qualitative agreement with
the data is restored. (According to Table 5.1, stabilization reduces the static
value of a by about one-third. We assume here the same reduction at 1.17 eV.
Real Al would, in addition, require us to consider damping processes and the
interband transition near 1.5 eV.)
The dependence of the second harmonic intensity on the crystal face indi-
cates that band structure effects in Al are not negligible. Evidence for such
effects was also observed by Janz et al. (1991a). Figure 5.7 shows the har-
monic intensity for AI(l11) in the interval 565 - 860 nm (1.4 - 2.2 eV). The
data are compared with calculated intensities based on LDA-RPA values of
a(w). According to Table 5.1, these parameters are in fact close to the LDA
results for the more realistic stabilized jellium model. The measured AI(lOO)
intensities are again considerably weaker.
If a(w) is assumed to be constant in the wavelength window shown in
Figure 5.7, the calculated intensities also increase towards larger w as a result
of the frequency dependence of the dielectric function. However, this increase
is weaker than the one plotted. Part of this behavior is therefore caused by
the increase of la(w)1 as 2w approaches the threshold for emission (see Figure
5.3).
For AI(111), Janz et at. (1991a) also detected an appreciable azimuthal
anisotropy arising from interband transitions (see Section 5.6). Neverthe-
less, the comparison of the calculated harmonic intensities with the measure-
ments shown in Figures 5.6 and 5.7 suggests that the perpendicular surface
polarization is a very important term in Al and that the TDLDA (prefer-
ably for the stabilized jellium model) provides a good representation of the
a-parameter. The earlier SHG experiment on Al by Quail and Simon (1985),
giving a(w) = +1.5±0.3 at 1.17 eV, was performed for an AI-quartz interface
and cannot be compared with data on clean Al under UHV conditions.

5.3.4 Alkali Metals

As we saw in Figure 5.3, nonlinear excitation spectra of alkali metals are


dominated by the multi pole surface plasmon near 2w = Wm ~ 0.8 wp. For
5.3. SIMPLE METALS 243

, .... --,
~ 100
'- , ,,
Eo K , ,
,
~
1
o ,, ,
,, ",,,
50 ,,
,
Na

,
,,
,,
oL-~~d=~L-~ __- L_ _~~_ _-L~
o 20 40 60 80

polar angle of incidence

Figure 5.8: Theoretical second harmonic efficiency for Na and K


as a function of polar angle of incidence. w = 1.79 eV, p-in, p-out
polarization. Solid curve: Na with a = (-10 - 20i)j dashed curve:
K with a = (20 + 20i) (see Figure 5.3). In both cases, d = -b = 1.

Na, K, Rb, and Cs, the required fundamental frequencies are 2.3, 1.5, 1.4,
and 1.2 eV, respectively. With existing tunable lasers it should therefore be
possible to verify the strong resonance-like variation of a(w). Unfortunately,
such studies have not yet been undertaken. Measurements at only one or two
wavelengths would not be sufficient because of the considerable uncertainties
in the dielectric constants.
In the early theoretical work of Wang et al. (1973) based on hydrodynamic
calculations for the free-electron model, good agreement was claimed with SHG
measurements on Na and K by Krivoshchekov et al. (1969, 1970) at 1.79 eV.
According to the TDLDA calculations, at this frequency a(w) should be about
(-10 - 20i) for Na and (+20 + 20i) for K. The reason for the sign change is
that, for Na, this frequency lies below the multipole plasmon, but above for
K. Figure 5.8 illustrates the angular variation of the SHG efficiency obtained
within the TDLDA. The dielectric functions are: Na: c(w) = -8.7 + 0.29i,
c(2w) = -1.5+0.14ij K: c(w) = -3.8+0.19i, c(2w) = -0.17+0.05i (Smith,
1970). Although we use the same Fresnel factors as Wang et al. (1973), the
maxima of their SHG efficiencies are about 2000 x 10- 20 cm 2 /W for both
metals: significantly larger than predicted by the TDLDA. There is clearly a
need for improved SHG measurements on alkali metal surfaces.
Resonance behavior similar to that on clean alkali metal surfaces should
244 CHAPTER 5. NONLINEAR OPTICS

also be observable on adsorbed alkali metal layers. Because of the efficiency


of metallic screening, the density profile at the adsorbate-vacuum interface re-
sembles that at the semi-infinite surface once the coverage exceeds one mono-
layer. The nonlinear response of alkali metal overlayers is the topic of Section
5.5.

5.3.5 Extension to Ag

At low frequencies, the dielectric response of Ag is free-electron-like. We


might therefore expect the nonlinear surface response properties to be similar
to those of simple metals. When Quail and Simon (1985) performed SHG
measurements on Ag-quartz interfaces, they found a(w) = +0.9±0.2 at 1.17 eV
This value is presumably too small for the same reason as for Ai. Electronic
properties at the interface apparently differ strongly from those at free Ag and
Al surfaces in vacuum. Perhaps this is not surprising, since the tail of the
conduction electron density is modified due to formation of chemical bonds
with the quartz overlayer. Surface electrons are then much less free to take
part in the screening of the applied electromagnetic field. Hence, surface dipole
moments should be reduced.
To determine the isotropic surface polarization of Ag in the region far below
the interband onset, Wong and Richmond (1993) carried out SHG measure-
ments at 0.81 and 1.17 eV (see Figure 5.9). By analyzing measurements for
different in-out polarizations, the components of the nonlinear surface polar-
izability were determined. At 1.17 eV, the magnitude of X"zz was found to
be 230 x 10- 14 esu, implying la(w)1 ~ 10. Surprisingly, the magnitude of
X.,., .. was only 14 x 10-14 esu, giving Ibl ~ 0.36 instead of unity, as predicted
by all models based on the uniform electron gas. The third isotropic surface
polarizability element Xu., was found to be negligibly small.
The TDLDA value for jellium with TB = 3 is a = (-15 - 3i) (see Figure
5.3), but, as pointed in Section 2.1.4, this value should be reduced to account
for hybridization between sand d states. Figure 5.9 illustrates the theoretical
intensity derived from (5.62) for a = (-7.5 -1.5i) = constant, b = -1 and d =
1. The lowering of a relative to the TDLDA value for jellium corresponds to the
reduction obtained by Weinert (1994) in the adiabatic limit. The low-frequency
behavior agrees quite well with the experimental data. The measurements at
2w = 4.68 eV cannot be compared to the theoretical results, since at present
we have no information about the nonlinear bulk and surface responses in the
range of interband transitions. (Anisotropic intensities are discussed in Section
5.6.)
5.3. SIMPLE METALS 245

*
6
~
"-
NE Ag
" 4
~
0
.:::;,
N
.3
2
"-
~

0
2 3 4 5
2w (eV)

Figure 5.9: Second harmonic efficiency for Ag(1l1) as a function


of harmonic frequency, () = 31°; p-in, p-out polarization. Symbols:
isotropic (stars) and anisotropic (squares) intensities measured by
Wong and Richmond (1993); solid curve: theoretical isotropic in-
tensity for a = (-7.5 - LSi), b = -1 and d = 1.

1.0

1.0~
0.8
0.8 3 0.6 I
.5:'
0.4
0.2

0.0 '--::'2o:----'---:':6o:--'--:17coo,-'--:14...,o-L..,.-18~0-L
Temperature (C)

200 Wavelength (nm)

Figure 5.10: Harmonic efficiency for Ag(llO) as a function of wave-


length from 260 nm (4.8 eV) to 340 nm (3.6 eV); () = 60°; p-in,
p-out polarization. Filled (open) circles: 94 K (573 K). Inset: tem-
perature dependence of intensity at 314 nm. (Hicks et al., 1988).
246 CHAPTER 5. NONLINEAR OPTICS

The calculated intensity in Figure 5.9 is based on frequency-independent


coefficients a, b, and d. Thus, the peak near 3.8 eV is entirely a consequence
of Fresnel factors, in particular, of the coefficient (5.73) in the vicinity of the
Ag bulk plasmon. This behavior was observed in SHG studies on Ag(llO)
by Hicks et al. (1988) (see Figure 5.10). The variation of this peak with
temperature agrees qualitatively with the behavior deduced from (5.73). As
the temperature is increased, the expansion of the lattice causes a broadening
of the Ag bulk plasmon and a slight red shift of the frequency (Liljenvall,
Mathewson, 1970). As shown in the inset, the SHG intensity at 314 nm can
be used as a signal of the Ag surface temperature.
Hicks et al. (1988) used this behavior in a pump-probe experiment to study
the transfer of heat from the surface to the bulk. Since the relevant length
scale for heat diffusion is 105 A, the optical penetration depth of the first laser
pulse at 1.17 eV (about 125 A) is negligible. The probe beam at 628 nm was
then used to monitor the temperature evolution of the Ag surface within a
few nanoseconds after initial exposure. The measurements were found to be in
good agreement with predictions based on the classical heat diffusion model.
According to the linear optical response of Ag discussed in Chapter 4, the
surface parameters dl.(w} and dll(w) exhibit a rapid frequency dependence near
the bulk plasmon as a result of the mutual polarization between the sand d
electron densities. It is to be expected that the nonlinear surface coefficient
a( w} (and possibly other nonlinear parameters as well) show a similar behavior,
since analogous screening processes should occur at the harmonic frequency.
To obtain a qualitative understanding of this effect, it would be instructive to
extend the s-d polarization model used in Sections 3.5 and 4.5 to the nonlinear
optical response.
The microscopic electronic structure at the Ag surface should lead to ad-
ditional features in SHG spectra. The importance of transitions involving
surface states and resonances was pointed out by Jiang et al. (1991) and Hu
(1996). Experimental evidence for such phenomena was observed by Urbach
et al. (1992) on Ag(llO).

5.4 Charged Surfaces

In view of the surface specificity of nonlinear optical methods, there is


considerable interest in applying SHG and SFG to metal--electrolyte interfaces
to characterize structural and chemical surface properties and to investigate
how these vary with the electric potential (Richmond et al., 1988). Linear
5.4. CHARGED SURFACES 247

(eV)

0
o
-2 -a
E
-4

-6

-8

-10 o 10
z (u.u)

Figure 5.11: One-electron potential V(z) for jellium (r. = 3) in the


presence of a static electric field. The net surface charge densities
are ±O.0002 and ±O.0004 a.u. (Liebsch, 1991).

optical methods have the disadvantage that most of the signal originates in
the bulk of the sample. Interface properties can, however, be accentuated by
varying the chemical potential and focusing on features in difference spectra.
In this section, we limit the discussion to the influence of the electric po-
tential on the perpendicular component of the nonlinear surface polarizability.
The presence of the electrolyte is simulated via a static electric field. As we
saw in Section 4.6, linear excitation spectra are very sensitive to the charge
state of the surface. This applies even more so to the nonlinear response, since
the centroid of the harmonic surface polarization is located even further in the
tails of the equilibrium density.
For negatively charged electrodes, the electronic density is more diffuse
and more polarizable than that of the neutral surface. Positively charged
electrodes, on the other hand, have a stiffer, less polarizable profile. Figure 5.11
illustrates the one-electron potential for several electric fields. The effective
'work function' (the difference between the potential near the tail of the density
and the Fermi energy) is seen to be very sensitive to the static field. Since the
one-electron states change quite rapidly with the applied field, we expect the
spectral features of a(w) and its overall magnitude also to vary with the net
surface charge.
248 CHAPTER 5. NONLINEAR OPTICS

40
30
20
] 10
.,
d

0: 0
I

-10
-20
40
30
20
] 10
d
0
~
I
-10

-20
-300
3 4
w(eV)

Figure 5.12: Frequency dependence of (a) real and (b) imaginary


parts of a(w) for neutral surface (symbol 0) and for charged surfaces
(±) calculated within the TDLDA (r. = 3). The surface charges
are ±0.0004 a.u. (Guyot-Sionnest et al., 1990).

Figure 5.12 shows the frequency dependence of a(w) as calculated within


the TDLDA. The spectrum for the neutral surface is the same as in Figure
5.3(b). Because of the lower barrier height of a negatively charged surface, the
threshold excitation shifts downwards, causing an increase in both real and
imaginary parts of a(w) at low frequencies. This change is consistent with
the more diffuse shape of the density profile and its enhanced polarizability.
Positive charging has the opposite effect, giving a reduction of the real and
imaginary parts of a(w) at low w.
Evidence of this behavior was found in SHG spectra by Guyot-Sionnest et
al. (1990) for Ag(l11) in contact with an aqueous solution of KCI0 4 . Figure
5.13 illustrates the harmonic efficiency for various electric potentials. Using
(5.62), the a-parameter was derived from these data under the assumption
5.4. CHARGED SURFACES 249

iJZl
4

·1 -0.5 0
ElSCEM

~2
E
"
'0

O~~-L~~~~-L~~~~~~~~
o 10 20 30 40 50 60 70 80 90
ANGLE OF INCIDENCE (degrees)

Figure 5.13: Square root of second harmonic efficiency R = I'}.wl I~


for Ag(111)-electrolyte interface as a function of polar angle of
incidence. w = 1.17 eV; p-in, p-out polarization. The symbols
denote potentials between 0.0 and -1.3 V ISCE. Inset: variation of
Rea(w) with potential. (Guyot-Sionnest et at., 1990).

20

'"
'", 10
•....
I'·
""''j' .
I . i··...... "
I I

-0.005 O.OOS 0.010


charge (a.u.1

Figure 5.14: Rea(w) for Ag-electrolyte interface as a function


of surface charge. Symbols: experimental values. Solid (dashed)
curve: TDLDA results at 1.17 eV (in static limit); dotted curve:
solid curve reduced by factor of 0.3 to account for lattice effects.
(Guyot-Sionnest et at., 1990).
250 CHAPTER 5. NONLINEAR OPTICS

that d = -b = 1 and f = O. The inset shows that the real part of a varies
between -10 and +0.7 as a function of the applied potential. In accordance
with the preceding arguments, a(w) becomes more negative with negative po-
tential since the electronic density profile is more polarizable. Towards positive
potentials, TDLDA jellium calculations give a(w) = -2 as a limiting value
(see (5.102) and Schaich (1994)). The data, on the other hand, indicate a
sign change towards positive a. This behavior is presumably related to the
interaction of negative ions, in particular OH-, with the Ag surface.
To compare the measured a-parameter with the TDLDA results, we show
in Figure 5.14 the real part of a(w) as a function of the net surface charge.
Although the jellium results reproduce the overall trend, namely, a decrease
(increase) of la(w)1 with positive (negative) charging, the magnitude of lal is
too large. This is not surprising since the 4d electrons of Ag tend to reduce
the surface polarizability by about a factor of 2 - 3 (see also Section 5.3.5).
While the inclusion of this effect leads to qualitative agreement with the data,
important questions remain. These concern, in particular, the importance
of other nonlinear isotropic and anisotropic surface polarization terms and
their dependence on the presence of the electrolyte. More theoretical and
experimental work in this area is needed to understand the chemical and optical
properties of the metal-electrolyte interface.

5.5 Alkali Metal Overlayers

SHG is known to be highly sensitive to the presence of adsorbed atoms


on the surface. One of the most striking examples of this phenomenon is the
adsorption of alkali metals on metal surfaces. For thin Na, K, and Cs layers
on Rh, Tom et al. (1986) observed enhancements of the second harmonic
intensity of clean Rh by up to four orders of magnitude (see Figure 5.15).
Similar enhancements were observed for Rb on Ag (Song et al., 1988; see Figure
5.16), Cs on Cu (Lindgren, Wallden, 1992), Cs on Ag (Reiff et al., 1994), and
K on Al (Wang et al., 1995). Qualitatively, these effects are related to the
large polarizability of loosely bound valence states of alkali atoms. The rich
frequency and coverage dependence of the enhancement indicates, however,
that the microscopic nature of occupied and unoccupied chemisorption states
plays an important role in this phenomenon. In this section, we show that the
intrinsic dynamical response of the alkali metal layers is the main origin of the
observed enhancements.
5.5. ALKALI METAL OVERLA YERS 251

AEXC = 0532,..,.
c.
c.

No

No

-,
'OL-----~------~------­ 10 O~---~'--------;\----
ALKALI METAL COVERAGE (Hep MONOLAYER)
ALKALI METAL COVERAGE (HCP MONOLAYER)

Figure 5.15: Harmonic intensity as a function coverage for alkali


metals on Rh at 1.17 eV (left) and 2.34 eV (right); p-in, p-out
polarization. The intensity of clean Rh is taken to be unity. (Tom
et al., 1986).

1500
Rb/Ag(110)

>- 1000
u
c

~"
w
C)
:t:
(j)

500

l064nm, s-pol

20 40 60 80 100 120

Thickness (A)

Figure 5.16: Harmonic intensity for Rb layers on Ag as a function


of overlayer thickness at 1.17 eV and 2.34 eV. (Song et al., 1988).
252 CHAPTER 5. NONLINEAR OPTICS

5.5.1 JelliuID Model

As long as the direct overlap between adatom orbitals is negligible, the


observed increase of the harmonic signal can be approximately explained in
terms of optical transitions between adatom levels that are broadened due to
hybridization with substrate conduction bands (Persson, Dubois 1989). At
higher coverages (roughly beyond the work function minimum), the extended
character of the overlayer states must be taken into account.
In the adiabatic limit, the nonlinear surface polarization of adsorbed alkali
metal layers was calculated by Weber and Liebsch (1987b) as a function of
coverage. For Na and K on AI, we find that the harmonic intensity is about
one order of magnitue larger than for clean AI. It is clear from these estimates
that the static polarizability of alkali metal layers is far too small to explain
the observed enhancements of the harmonic intensity.
In contrast to the near-adiabatic response of clean Al discussed earlier, the
laser frequencies employed in the experiments (1.17 and 2.34 eV) are not small
compared to the alkali metal volume and multipole surface plasma frequencies.
In fact, from Figure 5.3 we can anticipate that the nonlinear surface excitations
exhibit resonances when the fundamental or harmonic frequency coincides with
the multi pole plasmon of the alkali metal. This is indeed the case, as illustrated
in Figure 5.17, which shows the frequency dependence of a(w) for monolayers
of Na, K, and Cs on Al as calculated within the TDLDA. The imaginary part
of a(w) is seen to exhibit strong resonances near 0.5wm ~ OAwp , Le., when
two photons generate the overlayer-induced multipole surface plasmon. The
real part of a(w) exhibits the typical resonance shape as expected from the
Kramers-Kronig relations.
At 1.17 eV, we obtain the values: Na-AI: a = (-132 - 24i), K-AI:
a = (-400 - 175i), Cs-AI: a = (-800 - 715i). The enhancement of the
harmonic intensity caused by the alkali metal layer is 17, 195 and 1200, for
Na, K and Cs, respectively. Here, we assumed p-polarized light incident at a
polar angle of 60° and a = (-36 - 9i) for the Al substrate. In the stabilized
jellium model, a for Al is about one-third smaller, while the alkali metal values
remain nearly the same (see Table 2.3). This means that the adsorbate-induced
enhancement of the intensity increases by a factor of 9/4, giving 38, 440, and
2700, for Na, K, and Cs, respectively.
As illustrated in Figure 5.18, these TDLDA predictions are in qualitative
agreement with experimental data for alkali metal layers on Al by Wang et al.
(1994). In the case of Ag and Rh substrates, a is certainly smaller than for AI,
so that we would obtain even larger enhancements. These results show that
5.5. ALKALI METAL OVERLAYERS 253

No-Al K-Al
'600 1200 1200

400

200
2 Re
0
I
0

-200

-400

2 3 0 2 3 a 2 4
w(eV) w(eV) w(eV)

Figure 5.17: Frequency dependence of a(w) for Na, K, and


Cs monolayers on AI, calculated within TDLDA for adsorbate-
substrate jellium model. The arrows indicate 0.5wm . The dots
denote the values of a(w) at w = 1.17 eV. (Liebsch, 1989).

the intrinsic dynamical response of the overlayer contributes significantly to the


second harmonic intensity as a result of the close proximity of the harmonic
frequency to the multipole surface plasma frequency of the alkali metals.

5.5.2 Realistic Alkali Metal Overlayers

As mentioned before, the jellium model accounts for the metallic char-
acter in the overlayer, but it omits interband transitions between the two-
dimensional adsorbate states. The three-dimensional linear response calcula-
tions discussed in Section 4.7.2 indicate that these transitions have negligible
importance as long as c 2: 1/2 (see Figure 4.20). Since a(w) represents a lateral
average similar to d.L(w), we can expect the jellium model to be appropriate
also for the nonlinear response to the perpendicular field component.
254 CHAPTER 5. NONLINEAR OPTICS

4.--.-----,-----,-----,-----,--,

Rb/AI c
CS/AI

oil 2 K/AI
"-il
'6;
.2 Nc/AI

0
AI

2 4 6

r. (Co)
Figure 5.18: Logarithm of enhancement of second harmonic inten-
sity for alkali metal layers on Al (w = 1.17 eV). 1!L: intensity
of bare Al surface, 12",: intensity after adsorption of a monolayer.
Open squares: measurements (Wang et al., 1994); solid squares:
TDLDA results for adsorbate-substrate jellium model.

In the static limit, this was verified by Ishida and Liebsch (1990). As
illustrated in Figure 5.19, the atomic corrugation of the nonlinear induced
density n2(T) is even more pronounced than that of the equilibrium density
and the linear induced density (see Figures 4.19 and 4.20). The reason for
this effect is that the screening occurs even farther in the tails of the ground-
state density. Nevertheless, the planar average n2(z) of the induced density for
c ~ 1/2 is nearly the same as the induced density obtained within the jellium
model.
A similar conclusion applies to the planar average of the second-order in-
duced polarization. This quantity is related to the density via dP2z(Z)/dz =
n2(z). As long as c ~ 1/2, the nonlinear dipole moment

(5.107)

is therefore well represented within the jellium model. Thus, the frequency
dependence of a(w) shown in Figure 5.17 should be qualitatively correct for
realistic alkali metal overlayers. At very low coverages, it is more appropriate
to consider the response of single adsorbed atoms. In the static limit, this was
investigated by Kuchler and Rebentrost (1993, 1994).
5.6. SURFACE ANISOTROPY 255

9=1

2 9=~
~ 1
2
Ia.N 0 p.....~"'<--+---"""t
'12
3 9=i

-5 10

Figure 5.19: (a) Contour maps of the nonlinear density n2(T) in-
duced in hexagonal Na overlayers for several coverages. The dots
denote the position of the Na ions. The substrate (Ts = 3) occu-
pies the half-space z :::; O. The vertical scale extends over one unit
cell. The solid, dashed and dot-dashed lines correspond to positive,
negative and zero values. (b) Planar-average polarization F2(Z) for
hexagonal Na overlayers (solid curves) and for equivalent jellium
layers (dashed curves). (Ishida, Lie bsch, 1990).

5.6 Surface Anisotropy

Until now we have concentrated on the nonlinear response of metal surfaces


to the normal component of the electric field. The symmetry breaking caused
by the surface and the strong gradients of this field component make the Xzzz
element of the harmonic polarizability tensor a particularly important one.
In this section and the following one, we consider systems in which the
inversion symmetry is also broken in the direction parallel to the surface. For
example, on the (111) face of fcc crystals, the second and third atomic planes
are displaced in a non-symmetric fashion with respect to the first one. Only
the fourth plane has the same lateral origin as the first plane. This implies
that, although the parallel component of the electric field is nearly constant
256 CHAPTER 5. NONLINEAR OPTICS

in the surface region, a nonlinear parallel polarization is nevertheless induced


due to the non-symmetric stacking of lattice planes. Clearly, the physical
processes giving rise to this anisotropic component differ fundamentally from
those contributing to the isotropic perpendicular polarization. While, in the
latter case, metallic screening plays the crucial role, the anisotropic parallel
surface current is caused by interband transitions. We may then ask which
length scale governs the effective penetration depth of this current and what is
the origin of its surface sensitivity.
In the case of the (111) face offcc crystals, the second harmonic efficiency
for p-polarized incident and reflected radiation takes the form

12w/I~ '" lAp + B" cos(3¢W , (5.108)

where ¢ is the azimuthal angle of rotation. The overall prefactor is the same
as in (5.62). The coefficients A" and B" denote the isotropic and anisotropic
amplitudes defined in (5.71) and (5.72), respectively. Interband transitions
giving rise to the surface anisotropy also modify the bulk polarization. As
shown by Petukhov and Liebsch (1993), the bulk effect can be taken into
account by replacing the coefficient d(w) in (5.71) via

d(w) ---+ diso(O, w) + daniso(O, w) cos(3¢) , (5.109)

where diso is related via Fresnel factors to the isotropic and anisotropic bulk
coefficients 'Y(w) and ((w) (see (5.2», while daniso is proportional to ((w).

5.6.1 AI(111)

Figure 5.20 shows measured rotational patterns of the second harmonic effi-
ciency for AI(I11) (Janz et al., 1991a; Ying et al., 1993). Since the laser fre-
quencies range from 1.17 to 2.0 eV, the fundamental or harmonic frequencies
are close to the main interband transition at 1.5 eV. For p-in, p-out polariza-
tion, these patterns may be fitted to an expression like (5.108). The observed
ratio IB"/A,, I has the value 0.9 at 1.17 eV and decreases from 0.33 to 0.14
to 0.08 as the frequency is increased from 1.5 to 1.7 to 2.0 eV. Although the
isotropic response of Al in this frequency range is large, the anisotropy caused
by interband transitions can be just as large. Furthermore, it is strongly
frequency dependent. In part, the anisotropy may be induced by electric
quadrupole fields in the bulk. However, these sources are too weak to explain
the experimental observations (Petukhov, Liebsch, 1993).
5.6. SURFACE ANISOTROPY 257

~r---r---r---r--,
g
'"o
20r---------------~ '"
N

AI (111)

o 0

A=630 nm

A=730 nm

'" 5
A=820 nm

00 10 20 30 40 50 60 -60 0
'" (deg) Azimuthal Angle '"

Figure 5.20: Measured second harmonic efficiency from Al(111)


as a function of azimuthal incidence angle. (a) Wavelengths: 630,
730 and 820 nm (2.0, 1.7 and 1.5 eV). Polar angle of incidence:
67.5°; p-in, p-out polarization. (Janz et at., 1991a). (b) Differ-
ent incident and reflected polarizations. Wavelength: 1064 nm
(1.17 eV), polar angle of incidence: 60°. (Ying et at., 1993). The
solid curves are fits to the data.

To determine the anisotropic part of the surface polarization for AI(111),


Petukhov and Liebsch (1994, 1995b) used the quantum mechanical approach of
Jha and Warke (1967). The relevant energy bands were represented within the
two-band model which is known to give a good description of linear dielectric
properties of Al (Sturm, Oliveira, 1984). The Lang-Kohn potential was used
as a surface barrier. The nonlinear parallel current was evaluated within the
independent-particle approximation. Lateral screening effects giving rise to
local field corrections (nonzero Fourier components in the parallel electric field)
were neglected.
The spatial distribution of the anisotropic parallel surface current is plotted
in Figure 5.21. The anisotropic polarizability element X"''''''' is related to this
current via
(5.110)

The important point of these results is that the penetration depth of the par-
258 CHAPTER 5. NONLINEAR OPTICS

,,
"
:; -20
-':. 5
"
, "

-40 -20 o

Z (A)
Figure 5.21: Spatial distribution of parallel anisotropic surface cur-
rent hx(z, w)/ Elx for Al(111) at two photon energies. Solid curves:
real part; dashed curves: imaginary part. (Petukhov, Liebsch,
1995b).

allel current is much larger than that of the perpendicular isotropic current.
The patterns consist of a superposition of short- and long-wavelength oscilla-
tions whose period varies with frequency. In a homogeneous electron gas, the
longest Friedel oscillation has a wavelength ..\(w) = 27rkF/w (see Section 4.3.1).
For Al, we find at 1.17 eV: ..\(w) = 70 A and "\(2w) = 35 A.
To achieve a better understanding of the physical origin of the lateral cur-
rent, calculations were carried out for an increasing number of lattice planes.
As long as less than six planes are included (the remaining ones replaced by
jellium) ,the amplitude of the current is almost an order of magnitude smaller
and its penetration depth is much shorter. Only beyond about 12 layers does
the current acquire its full strength and spatial range. These results demon-
strate that the non-symmetric stacking of many atomic planes contributes to
the surface anisotropy. Obviously, there is no physical mechanism confining
the tangential current to the first atomic layer.
5.6. SURFACE ANISOTROPY 259

111

Too

10

Too

(k •. k.) (k .. -k.)
Figure 5.22: Lattice geometry and band structure for Al(l11). Top:
intersection of surface with (liD) plane; bottom: energy bands in
[110] and [122] directions. Dashed line: Fermi level; arrows: waves
travelling to surface along [122] and reflected along [110] where they
encounter the main band gap. (Petukhov, Liebsch, 1995b).

Experimental observations on many systems show that the anisotropy is


sensitive to surface conditions. At first sight, this seems to contradict the large
penetration depth of the tangential current. This puzzle is resolved if we recall
that electronic waves must be reflected from the surface to contribute to this
current. A modification of reflection conditions, for example, via adsorption
of foreign atoms or surface reconstruction, may lead to the admixture of new
parallel momentum components and to new combinations of incident and re-
flected electronic waves. Such surface modifications should therefore directly
affect the anisotropy.
This point is further illustrated in Figure 5.22, which shows the atomic
geometry and bandstructure corresponding to Al(l11). Propagating electronic
waves may be reflected at the surface in the direction of a bulk band gap, giving
rise to superpositions of oscillatory and evanescent states. The spatial range
260 CHAPTER 5. NONLINEAR OPTICS

of these decaying components is determined by the size of the band gap. The
main gap of Al is about 1.5 eV, implying decay lengths 2:: 16 A. These states
enter the optical matrix elements contributing to the anisotropic current. The
finite range of the evanescent states is consistent with the width of the surface
region of the currents shown in Figure 5.21.
The total nonlinear anisotropy has bulk as well as surface sources. In the
case of AI, we can introduce an effective anisotropy parameter

eel/(W) = e(w) + ((w)h(w) , (5.111)

where ((w) denotes the bulk contribution and hew) is a Fresnel factor. The
frequency dependence of e(w) and ((w)h(w) is plotted in Figure 5.23. The
surface anisotropy is seen to be large when w or 2w corresponds to the interband
transition near 1.5 eV. The bulk quadrupole term ((w)h(w) also has spectral
features at these frequencies but the overall magnitude is much smaller than
that of the surface term.
The isotropic and anisotropic amplitudes Ap and Bp in (5.108) are displayed
in Figure 5.23(c), assuming b = -1 and f = O. The results indicate that, in the
region of the Al interband transition, the anisotropic and isotropic amplitudes
are of comparable magnitude. Below 1 eV, the anisotropy should become even
larger than the isotropic terms. Above 1.5 eV, the anisotropy decreases rapidly.
This behavior is in qualitative agreement with data by Janz et al. (1991a) and
Ying et al. (1993) shown in Figure 5.20. More complete frequency-dependent
experimental studies would be desirable to test the predicted spectral distri-
bution of the anisotropy of Al and to determine the influence of lattice effects
on the isotropic coefficient a(w).

5.6.2 Ag(lll)

Measurements on Ag(l11) by Wong and Richmond (1993) (see Figure 5.9)


reveal a large lateral anisotropy not only in the inter band region but also far
below the d band onset. This finding seems at first sight surprising, since the
energy bands near the Fermi level have s-p character.
The s-p bands of Ag are similar to those of Al (see Figure 5.22), except
that the Fermi level lies very close to the lower edge of the main band gap at
the L-point. Since this gap is nearly 3 eV wide, lattice effects clearly produce
appreciable deviation from isotropic free-electron behavior. To estimate the in-
fluence of this nearly-free-electron behavior on the nonlinear surface anisotropy,
Petukhov (1996) applied the two-band model to Ag(111). Direct transitions
5.6. SURFACE ANISOTROPY 261

.,
~
-'
<>:: 0 ----------

~
~

E
0

CD
---c
VI
.0
15

/',,~(W)
10 ,,-
~
---
OJ
.0
C
5

0
0.5 1.0 1.5 2.0

W (eY)

Figure 5.23: Frequency dependence of (a) real and (b) imagi-


nary parts of nonlinear anisotropy parameters for Al(111). Solid
curves: surface contribution ~(w); dashed curves: bulk contribution
((w)h(w). (c) Comparison of calculated total anisotropic amplitude
Bp (solid curve) with experimental results by Janz et al. (1991a)
(squares) and by Ying et al. (1993) (circle). The total calculated
isotropic amplitude Ap is indicated by the dashed curve. The polar
angle of incidence is 60 0 • (Petukhov, Liebsch, 1995b).

involving d bands are neglected in this model. At low w, this is justified, since
d band transitions contribute only as virtual excitations. The results suggest
that the measured anisotropy can indeed be qualitatively understood within
this model. Thus, the parallel surface current is generated mainly by tran-
sitions within the s-p band near the Fermi level and close to the gap at the
L-point.
The remarkable feature of the anisotropic nonlinear current is that, as
in the case of AI(111), many lattice planes contribute to it. The parallel
component of the incident electric field decays very slowly and gives rise to
interband transitions deep in the solid. The surface anisotropy is caused by
262 CHAPTER 5. NONLINEAR OPTICS

the admixture of propagating and evanescent waves via the surface barrier
potential. The penetration depth of the lateral current is therefore determined
by the nature of these electronic states in the vicinity of the Ag band gap. Since
the existence of this current depends crucially on the condition for the reflection
of waves from the surface barrier, it is plausible that a large penetration depth
does not imply a lack of surface sensitivity.

5.6.3 Other Crystalline Effects

The lateral anisotropy on AI(I11) and Ag(I11) are examples of the struc-
tural sensitivity of second harmonic generation, which was experimentally
demonstrated on many surfaces and interfaces (for recent overviews, see: Janz,
van Driel, 1993; Shen, 1989, 1994; Reider, Heinz, 1995). The microscopic de-
scription of these effects requires in general the full treatment of energy bands.
Calculations of this kind are exceedingly complex and time-consuming. For
this reason, theoretical studies have been limited so far to semiconductors and
insulators in which many-body screening effects are usually much less impor-
tant than in metals. An independent-particle treatment is then a reasonable
starting point (Ghahramani et al., 1990, 1991; Akulin et al., 1992; Patterson
et al., 1992; Cini et al., 1993; Reining et al., 1994; Westin, Rosen, 1991).
Because of the complexity of the nonlinear optical response of realistic
crystals, there is clearly a need for useful models that allow the investigation
of at least certain qualitative aspects. A simple model, that focuses entirely
on the role of localized bound charges and allows consideration of surface
screening processes, is the so-called 'dipolium' consisting of an ordered semi-
infinite lattice of polarizable shells. As discussed in Chapter 4, refined versions
of this model were useful in analyzing the linear nonlocal optical response of
metals and semiconductors.
The nonlinear optical response of a dipole lattice was studied by Schaich
and Mendoza (1992). Local field effects were calculated for several low-index
faces, with a full account of all bulk and surface contributions to the harmonic
polarization. This model was also used by Wijers et al. (1993) to analyze
the second harmonic generation from semiconductor surfaces. An extension to
noble metals was presented by Mochan and Mendoza (1993). These authors
assume the dipole lattice to be immersed in a local Drude medium, in analogy
to the model by Tarriba and Mochan (1992) described in Section 3.5. It would
be interesting to improve this model by taking into account the nonlocallinear
and nonlinear response properties of the s-p electron density. A continuum
5.7. STEPPED METAL SURFACES 263

version of the dipolium model was recently proposed by Mendoza and Mochan
(1996). In this scheme, the nonlinear a parameter is expressed solely in terms
of the macroscopic bulk dielectric function. It is therefore independent of the
density profile n(z). Surface effects are included via the rapid spatial variation
of the normal electric field.
Because of the localized nature of the shells included in the dipole model,
and due to the short range of the dipole-dipole interaction between neighboring
lattice planes, the penetration depth of surface currents is much shorter than
in Al(11l) and Ag(l11). This illustrates the diversity of phenomena that may
arise. Evidently, the active surface range in SHG depends not only on the
polarizability components probed by certain input and output polarizations; it
also reflects the nature of electronic states taking part in optical transitions and
the screening processes at the fundamental and harmonic frequencies.

5.7 Stepped Metal Surfaces

A particularly striking example illustrating the importance of anisotropic


polarization terms are SHG measurements on vicinal Al(lOO) surfaces by Janz
et al. (199lb) (see Figure 5.24). According to these data, the anisotropic
harmonic response generated by the steps can be as large as the main isotropic
terrace contribution. For the example shown, the observed intensity can be
fitted to the expression:
(5.112)
where ± refers to incidence up or down the step direction. For small ter-
race angles a relative to the macroscopic surface, the ratio of anisotropic and
isotropic amplitudes may be written as BpiAp = a R = -a 0.32 exp(i0.46).
For a::::; 5° (corresponding to 10-atom terrace length and I-atom step height),
IBplApl = 1.6, i.e., a much larger anisotropy than on Al(11l) at the same
frequency. However, as on Al(lll), if the frequency is increased to 2 eV, the
anisotropy becomes negligible.

5.7.1 Anisotropy due to Density Contour

It is tempting to associate the step-induced anisotropy with the lateral


polarizability of electronic states in the vicinity of steps. At low terrace angles,
electrons are quite free to move along the surface. The symmetry breaking in
264 CHAPTER 5. NONLINEAR OPTICS

1500
...-- 6 6
!!
·c:J
.a. 1000
0
'--'
~
·iii
c
«II 500 - 6 6
C
::r:
VI

0
-5 0 5
a. (degrees)
Figure 5.24: Harmonic intensity for vicinal AI(100) surfaces as a
function of the terrace angle a with respect to the macroscopic
surface plane. The fundamental wavelength is 820 nm (1.5 eV);
(J = 67.5°; p-in, p-out polarization. Solid curve: fit discussed in the
text. (Janz et al., 1991b).

the direction normal to the steps may then give rise to strong parallel surface
currents.
To determine the magnitude of this effect, Ishida and Liebsch (1994) per-
formed nonlinear response calculations for stepped jellium surfaces. The ionic
charges in this model are replaced by a smooth positive background assumed
to have a regular sawtooth profile along the surface. The electronic proper-
ties of these two-dimensional systems (they are translationally invariant in the
third direction) were evaluated within the embedding scheme. For simplic-
ity, the response normal to the surface was calculated in the adiabatic limit.
The parallel response at finite w was determined in the independent-particle
approximation, assuming E iz to be spatially uniform.
The isotropic and anisotropic contributions to the amplitude of the har-
monic field in the present case are given by (Sipe et al., 1987)

e (Xzzz f: + X.zz I;) - ~: (Xu. + 2,b}21c!. +,d , (5.113)

£ X.z. 2/el. - Fe ( 2 2)
F. XZ •• I. + Xzz", Ie , (5.114)
5.7. STEPPED METAL SURFACES 265

where £ = c(2w) and "f is defined in (5.14). The b term represents the standard
parallel surface term arising from the action of the parallel linear electric field
on the linearly induced surface charge. The coefficient X",,,,Z accounts for the
additional contribution caused by the corrugated density. The d term is the
usual bulk magnetic dipole term. The remaining factors are

f. sinO/.fi , fe = S cos O/.fi , (5.115)


F. sinO/V[ , Fe = S cos (J / V[ , (5.116)

where sand S are defined in (5.52) and (5.60).


Figure 5.25 shows contour maps of the equilibrium density no(T), of the
nonlinear density n2(r) induced by a uniform static field oriented normal to the
surface, and of the parallel anisotropic current h", (T, w) generated by a uniform
parallel field at w = 1 eV. All contours are nearly confined to the region where
the density is corrugated. Only Friedel oscillations extend towards the interior.
The ground-state density shows the typical fiow of electrons from the top of
the step towards the lower corner (Smoluchowski smoothing). This charge
rearrangement reduces the surface dipole and lowers the work function.
Because of the greater electronic mobility in the normal direction, the linear
and nonlinear surface polarizabilities are enhanced. Surprisingly, this effect
increases the adiabatic a-parameter by only 10 % relative to its value for the
fiat jellium surface. Hence, the step-induced normal polarizability is rather
small. We can assume this behavior is representative of the dynamic response
at 1.5 eV, since the harmonic frequency is below the threshold excitation.
The parallel field is seen to induce a lateral current whose dipolar spatial
distribution consists of well-separated maxima and minima. Similar results are
obtained for other anisotropic surface polarization terms. The unexpected re-
sult of these calculations is that, despite the large density corrugation, the total
anisotropic surface polarization is much smaller than the isotropic component.
For the particular experimental conditions just mentioned, the calculated value
of the ratio Bp/ Ap at 1.5 eV is more than one order of magnitude smaller than
the measured one. Obviously, the lateral polarizability of electronic states near
steps is far too small to explain the experimental observations.

5.7.2 Anisotropy due to Crystal Structure

A possible source of the discrepancy between the second harmonic data on


stepped Al and calculations based on the stepped-jellium model is the neglect
of interband transitions. An indication of the importance of such effects is the
266 CHAPTER 5. NONLINEAR OPTICS

(0)

~------.L_---,--_~,~~:~t--'~_~_
20 o 10 20
X(o.u)

-10~--'!O~~~~~10:-"-~~-'--";2;;;:O~
X(OU)

Figure 5.25: Contour maps of electronic density for stepped jel-


lium surface (Ts = 2). The terrace length corresponds to 5 atoms,
the step height to one atom. (a) Equilibrium density no(r). (b)
Second-order density n2(f) induced by static field normal to the
surface. (c) Nonlinear tangential current h",(f,w) induced by uni-
form parallel field at w = 1 eV. The solid, dashed and dot-dashed
lines denote, positive, negative, and zero values. The edge of the
positive background is indicated by the heavy solid line. (Ishida,
Liebsch, 1994).

observed frequency variation of the anisotropy: In the interval 1.5 - 2.2 eV,
the coefficient BpiAp drops from 1.6 to negligible values. In contrast, the
stepped jellium model yields a rather weak frequency dependence. This differ-
ence points to a resonance phenomenon presumably caused by the electronic
transition at 1.5 eV. As discussed in the preceding section, this transition is the
origin of the large nonlinear anisotropy observed on Al(1l1) at low frequencies.
Lattice planes parallel to the macroscopic surface of vicinal Al surfaces are
5.7. STEPPED METAL SURFACES 267

20 , /\
,, ,

,-..
:J
-20
~
0

.
M
5
1(1 = 0.38

-40 -20 o

z (.A.)
Figure 5.26: Spatial distributions of parallel anisotropic surface cur-
rent h,,(z,w) at 1.5 eV fundamental frequency. (a) vicinal AI(100)
surface with 3.80 terrace angle. (b) AI(111). Solid (dashed) curves:
real (imaginary) parts. (Petukhov, Liebsch, 1995a).

stacked in a much more non-symmetric fashion than in the case of AI(111).


One must now go far deeper into the crystal to find an atomic plane with
the same lateral origin as the first plane. The effective unit cell is very large
and contains many more atoms that contribute to the lack of symmetry in the
parallel direction.
To determine the surface anisotropy induced by the lattice mechanism,
Petukhov and Liebsch (1995a) applied the two-band model discussed in Sec-
tion 5.6.1. Figure 5.26 illustrates the spatial distribution of the anisotropic
tangential current j2:c( z, w) for a terrace angle a = 3.80 • The comparison
with the corresponding current for AI(l11) shows that the current amplitude
for vicinal AI(lOO) is much larger because of the larger number of atoms per
unit cell. On the other hand, the penetration depth is similar, since in both
cases evanescent states near the Al band gap determine the spatial decay.
268 CHAPTER 5. NONLINEAR OPTICS

Similar results are found in more detailed microscopic calculations based on


the embedding method (Ishida et al., 1995). The calculated ratio BpjAp is in
qualitative agreement with the measurements by Janz et al. (1991b).
Comparing these results to the anisotropy induced by the corrugation of
the outer density profile, we can conclude that the non-symmetric stacking
of lattice planes provides a much more efficient source of anisotropic surface
polarization. The preceding analysis of the anisotropy of stepped Al surfaces
underlines the importance of a microscopic understanding of nonlinear optical
processes. Because of their complexity, it is non-trivial to associate observed
spectral features with certain electronic transitions.

5.8 Magnetic Surfaces

As mentioned before, second harmonic generation is particularly useful for


studying buried interfaces not accessible via standard surface spectroscopies.
Section 5.4 illustrates this point for metal-liquid interfaces. Another example
of considerable practical importance are magnetic interfaces. The recent search
for high-density storage systems has greatly stimulated the investigation of
magnetic materials using magneto-optical techniques (Falicov, 1992; Reim,
Schoenes, 1990). In contrast to linear methods, which probe mainly magnetic
bulk properties, nonlinear spectroscopies can provide important information
on the magnetization profile at surfaces and interfaces.
Consider first the linear magneto-optical Kerr effect. When polarized light
is reflected from a sample, spin-orbit coupling acts like a magnetic field that
rotates the polarization vector of the light. In paramagnetic systems, equal
up- and down-spin distributions lead to a cancellation. In ferromagnets, how-
ever, the exchange splitting of occupied and unoccupied spin bands causes a
net rotation, so that the polarization of the reflected beam depends on the
orientation of the magnetization (Kittel, 1951; Argyres, 1955). For normal
incidence and magnetization normal to the surface, the complex linear Kerr
angle is given by

vI +
..... 1 ( ) _ _ O':z;y
""K W - , (5.117)
O':z;:z; 47fiO':z;:z;jw

where O'ij(W) are components of the macroscopic linear conductivity tensor.


The frequency variation of these quantities is sensitive to the band structure
and optical matrix elements. Detailed evaluations within the LDA (Oppeneer
et at., 1992) show that, for Fe, the Kerr angle is about 0.5°; the measured
5.8. MAGNETIC SURFACES 269

,; «
$ 60 0.5
i':'
.~ ~ O.c+-...k:;----f~~-..,
~ 40 E
.9 >-
~ -0.5

20 .1.00L......o~40~~.....,..,.10":20~..,,16C::"0-'

analyzer angle a

30 60 90 120 150 180

analyzer angle a (deg.)

Figure 5.27: Left: Second harmonic intensity I 2w from an Fe/Cr


multilayer on Si as a function of output polarization for s-polarized
input. Filled dots: Mllx; crosses: Mil - x (longitudinal Kerr ef-
fect) (Koopmans et al., 1995). Right: Asymmetry A 2w of second
harmonic intensity for Fe film on Cu as a function of output polar-
ization for p-polarized input. The magnetization is normal to the
surface (polar Kerr effect). (Straub et al., 1996).

frequency dependence can be well reproduced. In the case of Ni, the angle is
only 0.1 0 • Agreement with the data is less satisfactory because of the failure
of the LDA to yield the correct exchange splitting. The quasi-particle nature
of the correlated states probed by optical spectroscopies must be taken into
account (Liebsch, 1979).
If we now turn to second harmonic generation, a crucial point is that the
magnetization does not break the inversion symmetry of the bulk. At the
surface, however, new polarizability elements arise that are odd under a change
of magnetization direction (Pan et al., 1989). Figure 5.27 shows two examples
that beautifully illustrate the potential of studying magnetic systems using
nonlinear optical methods. The asymmetry parameter is defined as

A It., Iiw
(5.118)
2w = It., + I 2w '

where ± denotes the orientation of the magnetic field. Of course, I 2w and A2w
depend on the analyzer angle a (a = 0 and a = 90 0 correspond to p- and
270 CHAPTER 5. NONLINEAR OPTICS

s-polarized output, respectively). In both experiments, the p-polarized output


is even in the magnetic field, while the s-polarized output is odd. The complex
nonlinear Kerr angle is defined as the ratio

(5.119)

In the case of Fe/Cr (20 A Cr on 20 A Fe deposited on a Si wafer), the


magnetic polarizability elements are believed to originate mainly in the Fe layer
because the Cr film is antiferromagnetic. The magnitude of the nonlinear Kerr
angle is about 17°. In the case of Fe (7 monolayers on Cu), 1il>~1 ~ 4°. Both
measurements use input light at 770 nm (1.62 eV) at nearly the same angle
of incidence (rv 45°). However, the input polarizations in Figure 5.27 differ,
so that quite different nonlinear surface polarizability elements are sampled.
Moreover, the magnetization is in the plane of the sample for Fe/Cr/Si and
normal to the sample for Fe/Cu. Similar large nonlinear Kerr angles were
observed on thin Co/Cu films (Wierenga et al., 1995) and the Heussler alloy
PtMnSb (Reif et al., 1993).
Nonlinear dynamical response calculations for magnetic systems using the
TDLDA would be exceedingly complex and are not yet feasable. Hiibner
and Bennemann (1989; Hiibner, 1990) calculated the nonlinear susceptibility
Xijk(W) by starting from a three-dimensional empirical ferromagnetic band
structure and projecting out two-dimensional components. Recent evaluations
of the nonlinear Kerr angle for Fe (Pustogowa et ai., 1994) are qualitatively
consistent with the measurements shown in Figure 5.27(a).
For a detailed understanding of the magnetic part of the nonlinear surface
polarization, in particular, of the spatial range contributing to the harmonic
intensity, it would be instructive to investigate several issues. On the one hand,
because of the localized character of d states, the deviation from inversion
symmetry caused by the surface should be more short-ranged than in the case
of simple metals. On the other hand, the electromagnetic coupling between
layers is not as localized since linear and nonlinear screening effects in transition
metals are less pronounced. As can be expected, these properties depend
on the frequencies and the polarization vectors of the incident and reflected
light. Since the nonlinear magneto-optical Kerr angle involves a superposition
of several complex surface polarizability elements, it would be interesting to
determine those dominating in a given configuration and find the effective
range over which magnetic surface properties are probed.
Chapter 6

Van der Waals Attraction

Understanding the interaction between neutral atoms and metal surfaces is


important for a variety of phenomena that can be used to probe static and dy-
namical properties of surfaces. For example, elastic scattering of helium beams
provides information on structural properties of ordered surfaces (Rieder, 1985)
and on properties of surface defects (Poelsema, Comsa, 1989). Furthermore,
inelastic scattering can be used to investigate surface phonons, sticking prob-
abilities and thermal desorption processes (Toennies, 1984; Engel, 1978).
In this chapter, we consider the interaction between a neutral atom at rest
above a metal surface. Dynamical corrections to this interaction that become
important when the atom is moving relative to the surface are addressed in
Chapter 7. The most common description of the atom-surface interaction is
based on a separation of short-range repulsive and long-range attractive con-
tributions to the physisorption potential (Kleiman, Landman, 1973; Zaremba,
Kohn, 1976, 1977). The motivation for this decomposition is that the repul-
sion is finite at distances where the electronic densities of metal and atom
overlap. The van der Waals attraction exists even when these densities are
well separated. The key contribution to the repulsive interaction is related
to the Pauli principle: The deformation of wave functions required to sat-
isfy the orthogonalization in the overlap region increases the kinetic energy.
Other contributions to the short-range potential arising from electrostatic and
exchange-correlation interactions can, however, reduce the Pauli repulsion to
such an extent that the net potential exhibits a minimum.
LDA calculations of the atom-surface interaction were performed by Lang
(1981) for several rare gas atoms on a jellium substrate. While the depth of
the physisorption wells agrees remarkably well with experimental data, the
asymptotic exponential form of the atom-surface potential is in conflict with

271

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
272 CHAPTER 6. VAN DER WAALS ATTRACTION

the law
(6.1)
which must be satisfied outside the region of wave function overlap. d is the
distance of the atom from the surface. This implies that correlation effects be-
yond those included in a local density formulation must be taken into account
to reproduce the long-range attraction. (We consider distances smaller than
the wavelength of light at optical frequencies so that retardation effects can be
ignored.)
The physical origin of the asymptotic -lids behavior is the polarization
interaction between quantum mechanical charge fluctuations in the atom and
solid. Accordingly, the strength C of the van der Waals attraction is deter-
mined by the atomic dipole polarizability a(w) and the dielectric function of
the metal e(W). As shown by Lifshitz (1956), C may be expressed as
1 1000 e(iu) - 1
C = -4
7r 0
du a(iu) e C)
tu +
1· (6.2)

The appearance of purely imaginary frequencies indicates that only virtual


electronic excitations are involved in the polarization interaction, i.e., no en-
ergy transfer occurs between atom and solid.
A general formulation of the van der Waals attraction between atoms and
surfaces was given by Zaremba and Kohn (1976). As shown by these authors,
the second-order contribution of the quantum mechanical interaction energy
takes the form

EvdW(d) = 1
--
27r
10
0
00
du ! d ! d T4
S Tl . .. S
1T1 -
1 ~ 1
1'3 + dl lfi - f4 + dl
~

x Xa(f1,fi,iu) X(1'3,f4,iu), (6.3)


where d = (0,0, d) and Xa, X are the exact density-ciensity response functions
of the atom and semi-infinite metal, respectively.
In the absence of lattice effects, (6.3) simplifies to (q == IQi!i)

_2- roo du L
27r Jo _
4:2 !dZ1 ...!dZ
q
4 e-q1z,-Z3+dl e-qlz2-Z4+dl
qll

X Xa(Zl,Z2,q,iu) X(zs,z4,q,iu). (6.4)


Evidently, E vdW remains finite even if the densities overlap. If d is large enough,
the overlap is negligible and the absolute-value signs in the exponents can be
dropped. We then find

EvdW(d) = --
1 100 00
du 10 00 dq e- 2qd a(q, iu) g(q, iu) , (6.5)
27r 0
6.1. REFERENCE PLANE POSITION 273

where
a(q,iu) = -ldz1/dZ2 e- q (Zl+ Z2) Xa(Zl,z2,q,iu), (6.6)

g(q,iu) - 2; I dZ 1 /dZ 2 eQ (Zl+ Z2) X(Zl,Z2,q,iu). (6.7)

Since the exact response functions Xa and X are not known, approximations
must be sought. For the metal surface, we use again the TDLDA. Thus, g(q, iu)
may be written as
(6.8)
The induced density nl(z, q, iu) is determined by the response equation (2.48)
with w = iu. This definition of the surface response function g(q, iu) is anal-
ogous to the one introduced in (2.65) and characterizes the dynamical surface
response to an external potential of the form eiqjl"rll+qz. According to (3.53),
g(q, iu) has the small-q expansion

g(q, iu) = e:~~u~ - 1 [1 + 2qdIP (iu)] , (6.9)


e: m +1
where dIP (iu) is the position of the dynamic image plane.
The atomic polarizability could in principle also be evaluated within the
TDLDA (Zangwill, Soven, 1980). The light rare gas atoms, however, are more
accurately represented within the time-dependent Hartree-Fock approxima-
tion or other more refined treatments. In general, a(q, iu) has the power series
expansion (Hutson et aZ., 1986)
. . (2q)21
a(q, zu) = L
1>0
a,(m) - 1)1
('
2 .
(6.10)

where
I
a,(iu) = 214: 11 dr dr' (rr')'+2 Xa,l(r, r', iu) . (6.11)
The asymptotic behavior given in (6.1) follows from (6.5) if only the lowest-
order terms are retained, i.e., g(q, iu) = [e:(iu) - 1]j[e:(iu) + 1] and a(q, iu) =
qal(iu).

6.1 Reference Plane Position

According to the definition in (6.2), the magnitude of the van der Waals
constant C does not depend on the microscopic electronic properties of the
274 CHAPTER 6. VAN DER WAALS ATTRACTION

surface. These properties become important at shorter atom-surface distances


where the potential begins to deviate from the asymptotic form (6.1). The
fourth-order term rv 1/d4 may in fact be used to determine the reference
position do from which the van der Waals attraction should be measured:

C
(6.12)
(d - dO)3

As shown by Zaremba and Kohn (1976), do follows from the linear term in the
expansion of g(q, iu):

1 roo c(iu) - 1
do = 47rC 10 du a(iu) c(iu) + 1 dIP(iu) . (6.13)

Thus, the van der Waals reference plane position may be interpreted as a
suitably weighted average of dIP (iu). This result is another example of the
fact that the effective surface location depends on how the metal electrons are
able to respond to electromagnetic fields involved in a particular probe-surface
interaction.
For jellium surfaces, the dynamic image plane position is given by (see
(3.52) )
. c(iu) .
dIP(zu) = C) 1 dl.(zu) , (6.14)
c zu +
since dll = o. To evaluate the reference position do, we must therefore deter-
mine the centroid of the screening charge induced by a uniform electric field
oriented normal to the surface. In the work by Zaremba and Kohn (1976),
dl.(iu) was approximated as

(6.15)

where u = u/wp and dl.(O) = Zl is the static image plane position defined
in (2.34). According to the Kramers-Kronig relation (see below), this expres-
sion is equivalent to assuming that all of the spectral weight of 1m dl. (w) is
concentrated at the surface plasma frequency wp /,;2.
Persson and Zaremba (1984) improved this approximation by account-
ing for the high-frequency behavior that follows from the surface f-sum rule
(4.117). According to this rule,

u~1, (6.16)
6.1. REFERENCE PLANE POSITION 275

1.5
r.=2 r.=3 r.=4

.~\
r-.. .~\
:J 1.0
~ ~\
r-..
2, "\'\,
"0
,
0.5 ,,
,,

0.0
a a a

u/w. u/w. u/w.


Figure 6.1: Centroid d1.(iu) of induced surface density as a function
of imaginary frequency for several jellium surfaces. Dashed curves:
expression (6.15) used by Zaremba and Kohn (1976); dot-dashed
curves: sum rule expression (6.17) used by Persson and Zaremba
(1984); solid curves: TDLDA (Liebsch, 1986a).

where the spill-out parameter A defines the ground-state density extending


beyond the jellium edge (see (4.118)). Hence, a better representation of d1.(iu)
is
d (iu) = d1.(O) (6.17)
1. 1 + 'fJ ii,2 '
with 'fJ = d1.(O)/A. Within the LDA, 'fJ = 3.6, 2.7, and 2.3 for jellium surfaces
with T. = 2, 3, and 4, respectively. The sum rule shifts the image plane
position inwards, giving rise to a corresponding inward shift of do. To improve
the description of dIP(iu) at intermediate frequencies, Persson and Zaremba
also used the Kramers-Kronig relation

d (. ) = ~Ioood w Imd1.(w) (6.18)


1.lU 7r 0
w
W +u
2 2 '

employing an approximate representation of d1.(w), The frequency dependence


of the centroid d1.(iu) obtained from (6.15) and (6.17) is shown in Figure 6.1.
276 CHAPTER 6. VAN DER WAALS ATTRACTION

6.2 Density Functional Description

The centroid d.L(iu) can be calculated directly within the TDLDA by con-
sidering the density response to a uniform electric field oriented normal to the
surface and varying in time like"" eut (Liebsch, 1986a). The induced density
may be found from the following response equations:

nl(z, iu) !dZ'Xl(Z,Z',iU)¢>srf(Z',iU) , (6.19)


¢>SCf( z, iu) = ¢>ext(z, iu) + ¢>ind(Z, iu) , (6.20)
Qr.nd(Z, iu) = ! dz' K(z,z') nl{z',iu) , (6.21)
K(z,z') -21rlz - z'l + v,:c[no(z)J6(z - z') . (6.22)

The external potential is taken to be ¢>ezt(z, iu) = -21rz. Since, at imaginary


frequencies, Xl(Z, z, iu) is a real quantity, nl and ¢>srf are also real.
To solve these equations with sufficient accuracy, the procedure discussed in
Chapter 4 for evaluating dl..(w) can be followed. In particular, the long-range
Coulomb interaction may be replaced by a more stable short-range version.
Also, the dynamical force sum rule can be used to avoid the numerical evalu-
ation of the induced dipole moment in the interior where the density exhibits
slowly decaying Friedel oscillations. At purely imaginary frequencies, there
is only one such oscillation that coincides with the period 1r/kF in the static
limit. Details of the solution procedure for finding dl..{iu) were given by Liebsch
(1986b).
Figure 6.2 illustrates the density induced at a jellium surface for various
imaginary frequencies. The integrated weight of these distributions diminishes
like
. c:(iu) - 1 1
a(zu) = c:(iu) + 1 = 1 + 2u 2 '
(6.23)

and the centroid gradually shifts towards the jellium edge because of less effi-
cient screening at increasing u. In the adiabatic limit, nl(z, iu) coincides with
the image charge density calculated by Lang and Kohn (1973).
The frequency dependence of the centroid dl..{iu), calculated within the
TDLDA, is also plotted in Figure 6.1. The formula by Zaremba and Kohn tends
to overestimate the value of dl..(iu) mainly at low r•. The sum rule expression
gives a remarkably accurate representation of the frequency variation except
at low values of u, where it also lies consistently above the TDLDA results.
In general, however, these differences are rather small and lead to only minor
shifts of the van der Waals reference position.
6.2. DENSITY FUNCTIONAL DESCRIPTION 277

0.4
rs = 3

0.2
u = 0.2 wp

---
::J
00

i. 0.2

---
N
::J
0.1
~

c
00

0.1

u= wp

00

-10 -5
z (aulD

Figure 6.2: Surface density nl(z, iu) induced by uniform electric


field oriented normal to the surface and varying in time like eut .
The bulk density corresponds to r. = 3. The imaginary frequen-
cies are indicated. The arrows denote the centroid position d.diu).
(Liebsch, 1986b).

For helium interacting with jellium surfaces (r. = 2, 3, 4), the TDLDA
gives do = 0.736, 0.638, and 0.592 ao, while the Zaremba-Kohn formula (6.15)
gives do = 1.018,0.803, and 0.708 ao, respectively. Additional inward shifts of
dIP(iu) and do should occur for r. < 4 if the standardjellium model is replaced
by the stabilized jellium model, since the latter leads to less polarizable density
profiles (see Section 2.1.3).
These values of do may now be used to construct the total atom-surface
potential. In the case of helium and neon, the asymptotic form (6.12) of
the van der Waals attraction is adequate, since the singular behavior near do
occurs deep in the repulsive region. The shape of the physisorption well is
therefore not affected by this singularity. For heavier rare gas atoms, however,
the attraction is much stronger so that the asymptotic form must be damped
near do to obtain reasonable physisorption wells. So far, there is no well-
defined procedure for introducing such a damping. A rigorous solution to this
278 CHAPTER 6. VAN DER WAALS ATTRACTION

problem is difficult, in particular, since one must avoid the double-counting of


correlation terms in the repulsive and attractive potential contributions in the
overlap region. For a discussion of the repulsive interaction and how to deal
with the divergence of the van der Waals attraction, the reader is referred to
the work by Chizmeshya and Zaremba (1989, 1992).
Andersson et al. (1996) recently determined physisorption potentials of H2
and D2 on Al from an analysis of selective-absorption structures in elastic-
backscattering experiments. This rather elegant method makes use of res-
onances between the reflected state and a state temporarily trapped in the
surface well. The overall agreement with theoretical predictions is good, but
the data are sufficiently accurate to detect a sizeable crystal face dependence
of the well depth. In part, such a dependence stems from the influence of the
density corrugation on the repulsive interaction potential. In addition, it ap-
pears that the face dependence of the reference plane position do of the van der
Waals attraction must be taken into account. Since this quantity is a surface
property, it is natural that do should vary for different faces. According to the
sum rule expression (6.17), do should roughly vary like the static image plane
position. The face dependence of the spill-out parameter >., which determines
the high-frequency behavior of dJ..(iu), might also playa role. The overall
face dependence of the net potential then results from a delicate balance of
repulsive and attractive contributions.

6.3 Near-Surface Corrections

Near-surface corrections to the van der Waals attraction may also arise
due to higher-order terms '" 1/dn >4 in the interaction. Such terms lead to
deviations from the asymptotic form given in (6.12). To determine the van der
Waals attraction at short atom-surface distances, it is useful to return to (6.5),
which holds as long as the overlap between metallic and atomic densities can be
ignored. As far as the metal is concerned, the key quantity is again the surface
response function g(q, iu). Within the TDLDA, 9 can be evaluated following
the prescription discussed in Section 2.3 and replacing w by iu (Liebsch, 1987).
The numerical task is simpler than at real w due to the shorter penetration
depth of the effective potential.
Figure 6.3 shows the variation of the surface response function with parallel
wave vector. At small q, g(q, iu) is well represented by the expression

go(q, iu) = u(iu) exp[2qdIP (iu)] , (6.24)


6.3. NEAR-SURFACE CORRECTIONS 279

6
5
4
"3 ./
-. 3 u =0 ./
g ./
5'2 V'
./
u = OSwp

0 os 1.0 0 0.5 1.0


q (a.u.1 q (a.u.1

Figure 6.3: Logarithm of surface response function g(q, iu) as a


function of q for u = 0 and u = 0.5 wp (r s = 2). Solid curves:
TDLDA results derived from (6.19); dashed lines: approximate ex-
pression (6.20). (Liebsch,1987).

which has the same linear behavior as the exact g(q,iu). At larger q, the full
g(q, iu) increases beyond go. We point out, however, that q should not exceed
the decay constant of wave functions at the Fermi level. Hence, q should
remain below about 2~ ~ 1 a.u., where <I> = 4.5 eV is a typical work
function value. According to (6.5), larger values of q are relevant only at short
atom-surface separations where the wave function overlap must be taken into
account. The original form of EVdW(d) in (6.4) then ensures that the van der
Waals attraction remains finite. In the limit u -t 0, g(q,iu) converges to the
function g( q) shown in Figure 2.8.
Using the power series expansion of a(q, iu) presented in (6.10), and ap-
proximating the full g(q, iu) by go(q, iu), the van der Waals energy may be
written as
1 1000 a/(iu) cr(iu)
EYdW(d) = --4:E du [d - d C WI+! .
7r 1>0 0
(6.25)
IP lU

The 1 = 1 contribution agrees to order 1/d4 with the asymptotic expression


(6.12). Outside the overlap region, the higher-order terms included in (6.25)
make the potential in the case of helium about 10 % more attractive than the
asymptotic form (6.12).
280 CHAPTER 6. VAN DER WAALS ATTRACTION

6.4 Influence of Occupied d Bands

The above discussion applies to jellium systems. Zaremba and Kohn (1976)
extended their formulation of the van der Waals attraction to noble metals by
treating the total valence electron density as a two-component system whose
sand d contributions respond independently to the neutral atom above the
surface. The net induced surface charge density (J = (J. + (Jd is given by
. c(iu)-l
(J(zu) = c(iu) + 1 (6.26)

The individual sand d densities are


. )_c.(iu)-l . ) _ cd(iu) - 1
(J. (zu - (.) , (Jd (zu - (.) 1. (6.27)
c zu + 1 cZU +
The bulk dielectric function c is separated into a Drude-like s electron contribu-
tion and a bound d electron term as indicated in (3.71). The s-d decomposition
is analogous to the one we used in discussing the Ag surface plasmon dispersion
(Section 3.5) and the nonlocal optical response of Ag surfaces (Section 4.5).
It can easily be shown that, within the two-component s-d electron system,
the effective centroid of the total screening charge is of the form (4.137), Le.:

. )_c.(iu)-l d (·)
d.L (zu cd(iu)-l d (·)
zu - 1 • zu
- c (.) + c (.)
zu - 1 d ZU ,
(6.28)

where d.(iu) is the centroid of the induced s electron density. To estimate the
centroid of the induced d charge, Zaremba and Kohn simulated the d bands by
a semi-infinite lattice of point dipoles. In this limit, dd(iu) is located about half
a lattice spacing above the outermost atomic plane, implying dd( iu) :::::; O. Chen
and Schaich (1989) showed that, as a result of local field corrections (induced
dipole moments near the surface differ from those deep inside because of the
reduced number of neighbors), dd(iu) shifts away from the jellium edge. The
sign of dd depends on the structure of the dipole lattice.
Figure 6.4 shows physisorption potentials for helium on Cu, Ag, and Au
calculated by Chizmeshya and Zaremba (1992). The well depth increases from
Cu to Au, and the position of the minimum shifts towards the surface. This
trend is primarily caused by the more important d band contribution Cd(iU)
to the dielectric function. This effect increases the van der Waals constant C
and shifts the reference plane position do inwards. The bound-state energies
for helium on Cu agree very well with experiment, while those calculated for
Ag are about 20 % deeper than the measured ones.
6.4. INFLUENCE OF OCCUPIED D BANDS 281

T I
5 - -

:;-

>
CD
E
'-"
..---
"0
'-"
lO
i
0

-5 f-
tV
\
I
,
I

\,
I
1:

\ \,._. //1
I

I ,
//
,
,
,~/
/ '
/
.-,,/--

I ,
I
,,
\',/, ,,
-10 f- -
I I

o 5 10 15

d (aD)

Figure 6.4: Physisorption potentials for helium on Cu (solid line),


Ag (short dashed line) and Au (long dashed line). (Chizmeshya,
Zaremba, 1992).

The main difference between the preceding approach and the s-d polariza-
tion model discussed in Sections 3.5 and 4.5 is that the latter scheme includes
the mutual interaction between sand d electron densities. Thus, sand d elec-
trons are not assumed to respond independently to the atom. Instead, the s
and d charge centroids are calculated simultaneously by taking into account
the screening effect of both density components. It would be interesting to
apply the s-d polarization model to the van der Waals attraction to check how
much the reference plane position and the minimum of the physisorption well
are shifted as a result of the s-d interaction.
As in the case of AI, H2 and D2 physisorption potentials on Cu reveal a
significant crystal face variation (Andersson, Persson, 1993). Both repulsive
and attractive terms contribute to this effect. As far as the d band contribution
to the van der Waals attraction is concerned, we could try to estimate the
role of the lattice structure by generalizing the s-d polarization model. If the
homogeneous dielectric background representing the d states is replaced by a
point-dipole lattice, both sand d electron contributions to the dynamic image
plane position (6.28) depend on the crystal face.
We conclude by pointing out that, in some systems, there still exists a
282 CHAPTER 6. VAN DER WAALS ATTRACTION

surprisingly large mismatch between calculated and measured values of the


van der Waals constant C. A detailed discussion of possible sources of this
discrepancy was given by Mehl and Schaich (1980) and Marvin and Toigo
(1982). Also, as first noted by Casimir and Polder (1948), at atom-surface
distances comparable to the wavelength of the main electronic transitions, e.g.,
d", c/wp '" 1000 A, retardation effects are no longer negligible. These change
the 1/d3 law associated with the electrostatic interaction to 1/d4 • For recent
measurements of the van der Waals attraction at intermediate distances, we
refer the reader to the work by Landragin et al. (1996).
Chapter 7

Electron-Hole Pair Creation

In the previous chapters, we focused on electronic excitations at energies where


collective surface modes play the dominant role. Electron-hole pair creation
of course also occurs, giving rise, for instance, to the width of the monopole
and multipole surface plasmons. There exist, however, various other physical
phenomena in which low-energy electron-hole pairs, say, up to a few tenths of
an eV, represent the main electronic excitation mechanism. Examples are the
width of the quasi-elastic peak in electron energy loss measurements, damping
of adsorbate vibrations, friction of ions and atoms moving near a metal surface,
surface resistivity, etc. Chapter 7 is devoted to these kinds of electronic surface
excitations.

7.1 Inelastic Electron Scattering

Consider first the low-frequency surface response to incident electrons. In


the dipole-scattering regime, the electron is far from the surface and interacts
with the metal electrons via the long-range Coulomb potential. The exter-
nal potential is of the form in (2.62), and the surface excitation spectra are
determined by the surface response function g(q,w) defined in (2.65).
Persson and Zaremba (1985) used this approach to analyze inelastic elec-
tron scattering data on Cu(100) (Andersson, Persson, 1983). Figure 7.1 com-
pares calculated and measured scattering probabilities as a function of incident
electron energy. The theoretical results are obtained by integrating the scatter-
ing probability (3.14) over the solid angle of detection. The total loss function
appearing in this expression is decomposed as
Img = Imgv + Img. + Imgb + Img; , (7.1)

283

A. Liebsch, Electronic Excitations at Metal Surfaces


© Springer Science+Business Media New York 1997
284 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

30
Cu (100)
nw 00.1 eV
, 20
>
~
D-
<J
~
10

Electron energy (eV)

Figure 7.1: Inelastic scattering probability t1P for Cu(100) as a


function of incident electron energy. The open and solid circles
represent the experimental data at T = 293 K and T = 80 K,
respectively, for 0.1 eV loss energy (Andersson, Persson, 1983). The
solid curves are the theoretical results (Persson, Zaremba, 1985).

where the Drude contribution


c:(W) -1 4 WF iii !::!.-
ImgD = 1m ( ) (7.2)
c: W + 1 wp kFl wp

accounts for the effect of phonon scattering in the bulk. Here, l = VFT is the
phonon mean free path, WF is the Fermi frequency, and iii = mopt/m = 1.5 is
the relative optical mass. At room temperature, the mean free path in Cu is
l = 221 A.
The remaining terms in (7.1) are calculated using the golden rule expression
for the transition rate (2.45) and separating the effective local potential into
surface and bulk contributions:

(7.3)

The surface potential 4>.(z,q,w) is replaced by the corresponding potential


in the limit W = 0, q = 0, i.e., by the induced potential derived by Lang
and Kohn (1973). This 'quasi-static' long-wavelength approximation of 4>. is
reasonable since the parallel momentum transfer is roughly given by qll '" lid.
Here, d '" (2E;/liw)/k; is that distance between incident electron and surface
at which the interaction in the dipole-scattering regime begins to take place.
E; = n?kl/2m is the kinetic energy of the incident electron. For E; '" 2 eV
7.1. INELASTIC ELECTRON SCATTERING 285

rRAS EELS

Figure 7.2: Relative strengths of the normal and parallel electric


field components at a metal surface in infrared reflection-absorption
spectroscopy (IRAS) and electron energy loss spectroscopy (EELS).
(Persson, 1991).

and nw rv 0.1 eV, d rv 50 A and qll rv 0.02 A-1. This q value is indeed
rather small; as a consequence, the bulk potential can be approximated as
¢Jb(Z,q,W) = _2; eqz (1- g_), where (2.72) implies g_ :::::; (J' = (c: -1)/(c: + 1).
Taking into account the optical mass of Cu, we obtain (J' :::::; 1 + 2m w2 / w~. The
interference term 1m gi results from the cross term between bulk and surface
matrix elements.
The agreement between theory and experiment is seen to be excellent,
with respect to the variation with incident energy as well as temperature.
This close agreement is particularly remarkable in view of the fact that the
interference between long-range bulk and short-range surface terms leads to a
sizeable cancellation. We point out, however, that only the electrostatic part
of the surface potential was used to evaluate the golden rule matrix elements
in (2.45). In a consistent description of surface excitations, the exchange-
correlation contribution to the effective local potential ¢Jecf should also be taken
into account (see Section 7.2).
Note that, at low frequencies and small q, electron energy loss spectroscopy
(EELS) is mainly sensitive to the perpendicular surface response measured by
d.l(w) (see Figure 7.2). This is evident from the expansion of g(q,w) given in
(3.53). For c: «: 0, we find
2 I cd.l(w)+dll(w)
Img(q,w) :::::; q m 1
c+

:::::; 2qlm [d.l(W) - :; dll(W)] (7.4)

Hence, in contrast to the region of collective excitations near c:::::; -1, where
dll and d.l playa similar role, at lowW the contribution due to dll is reduced
286 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

2.---------------'1------/-/'/--------.

/
/

plasmons /
/

11- / -
/

/
electron-hole pairs
/

O~------------~I------------~
o 0.5

q/k r

Figure 7.3: Electron-hole continuum of the three-dimensional


electron gas. The boundary is determined by the function q/k F =
(W/WF + 1)1/2 - 1 which approaches W/2WF for q ~ kF .

by a factor l/c ~ w2 /w;. The reason for this suppression is that, according to
(3.47) and (3.48), the external field components at small q are E. '" c/(c + 1)
and Ez '" l/(c + 1). For z > 0, we then find E./Ez '" c. In other words,
outside the metal, E. is greatly enhanced relative to Ez if c ~ o. On the other
hand, inside the metal, E. '" Ez '" l/(c + 1); i.e., both components are much
smaller than the external normal field E.(z > 0). Thus, the efficient screening
at metal surfaces makes EELS at low W extremely sensitive to the perpendicular
dynamical response in the surface region. In infrared reflection measurements,
on the other hand, the relationship between the field components is quite
different from that in EELS, so that parallel modes can also be observed (see
Section 7.4).

7.2 Low-Frequency, Long-Wavelength Excita-


tions

Evaluating electronic excitations at low frequencies is non-trivial as can be


seen from the following arguments. Figure 7.3 shows schematically the w-vs.-
7.2. LOW-FREQUENCY, LONG-WAVELENGTH EXCITATIONS 287

.06

,.-...
r.=2
:::J
ci
....., .04

,
: " ... _---------

O~--~~~~--~L----L----~--~

0 2 3

w (eV)

Figure 7.4: Surface loss spectrum 1m g( q, w) at low frequencies for


q = 0.1 A-l (Ts = 2). The dashed curve represents the spectrum
generated by the bulk potential ¢>b '" eqz • The vertical line denotes
the frequency w = 2qWF/kF = 1.38 eV up to which electron-hole
pair excitation in the bulk is allowed.

q region where single-particle excitations in the bulk are allowed. For small q
and low w, this continuum is bound by the line W/WF ~ 2 q/k F. The parameter

1 kF W W
TJ == - - - (7.5)
2 q WF

is > 1 above this line and < 1 below. At the surface, the electron-hole contin-
uum does not have a sharp boundary, since the surface potential can provide
sufficient perpendicular momentum so that single-particle excitations are pos-
sible at arbitrary wand qll. Nevertheless, the distinction between the regions
TJ > 1 and TJ < 1 is also important in surface response problems because long-
wavelength expansions are possible only if qll «: kF and qll «: W/VF, i.e.,
TJ » 1. At a fixed frequency, these conditions can be satisfied if qll is suf-
ficiently small. In the loss measurement shown in Figure 7.1, TJ ~ 1/2. A
long-wavelength expansion of the loss function would be inappropriate under
these scattering conditions.
Figure 7.4 shows the behavior of Img(q,w) at low frequencies as calculated
within the TDLDA. For illustrative purposes, we also plot the bulk contribu-
288 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

tion which is given by the analytical expression (Persson, Zaremba, 1985):

1] <1,
(7.6)
1] >1.
This expression is valid in the small-q, low-w region. The bulk contribution ex-
tends beyond the formal bulk boundary w = qVF, since three-dimensional bulk
excitations are projected onto the two-dimensional surface. Hence, at a fixed
value of qll, a continuum of bulk excitations of different q.l can be generated.
(Recall that the 'bulk' potential in (7.3) is rv e qz rather than rv eiq.L z .) Figure
7.4 shows that the total scattering probability results from a complicated su-
perposition ofterms generated by long- and short-range potential components.
To illustrate further the relative importance of surface and bulk contribu-
tions and to demonstrate the sensitivity of g(q,w) to model assumptions, let
us consider the low-w, small-q region such that 1] ~ 1. Because of phase-space
reasons, the variation of 1m 9 must be linear in w. If q is small enough, we
may write
q w
Img(q,w) = 2 -k - C (7.7)
F wp

According to the expansion of g(q, w) given in (3.53), ~ is identical to the linear


coefficient of Imd.l(w) introduced in (4.135). In Section 4.2.1, we showed that
Imd.l(w) can be expressed in terms of the golden rule. In the limit of small w,
(4.90) yields
kF3wp
~ = - -
7r
.
hm
o
w .....
Io
0
kF
1
dkz -k
z
I(k'Iz ¢J.e! I)
kz 12 , (7.8)

where k~ = (k~+2w )1/2 and ¢J.cf(Z, w) is the complex local potential calculated
within the TDLDA.
Table 7.1 lists the values of ~ for several jellium surfaces. The comparison of
the TDLDA and LDA-based RPA results demonstrates that the low-frequency
electron-hole pair production is sensitive to the metallic screening in the sur-
face region. At low bulk densities, exchange-correlation contributions to ¢J.e!
are seen to be very important: They shift the induced density profile further
into the more polarizable tails of the equilibrium density. For K (r. = 5),
this effect makes ~ one order of magnitude larger than in the RPA. This dis-
crepancy illustrates the severe consequences of the inconsistent treatment of
electron-electron interactions in the LDA-based RPA.
It is instructive to evaluate the coefficient ~ by using the decomposition
of the local potential given in (7.3). This separation is analogous to the one
employed in Section 4.2.4 in the calculation of d.l(w). In the long-wavelength
limit, the bulk potential becomes ¢Jb(Z,W) = -47rz/(c+l). At low frequencies,
7.2. LOW-FREQUENCY, LONG-WAVELENGTH EXCITATIONS 289

Table 7.1: Low-frequency coefficient { of 1m d.L(w) for jellium


surfaces. Upper two rows: TDLDA and LDA-based RPA. Middle
rows: surface, bulk, and interference contributions to total {t in
TDLDA for quasi-static approximation (Liebsch, 1987). Last row:
e.
surface contribution obtained by including only the electrostatic
part of the surface potential in the evaluation of the golden rule
formula.

r. 2 3 4 5
{LOA 0.82 0.20 0.08 0.04
{RPA 0.53 0.09 0.02 0.004
{. 2.28 1.32 1.00 0.84
{b 1.13 0.92 0.80 0.70
ei -2.62 -2.04 -1.71 -1.50
{t 0.79 0.20 0.09 0.04
~81e8t 0.85 0.31 0.09 0.03

¢>. can be approximated by its static limit. Inserting these potentials into (7.8),
it is evident that { can be expressed as

(7.9)

where the bulk term has the analytical form {b = 3wF/(2wp) = 1.6r;1/2 (see
(7.6) for 1/ ~ 1). The surface coefficient is given by

kFWp
{. = - 3- IokF dk", 1
-k I( k", 1¢>. 1k", )1 2, (7.10)
7r 0 '"

with ¢>. = ¢>.cf(Z, q = 0, W = 0). The interference contribution {i results from


the cross term between surface and bulk matrix elements.
According to the results given in Table 7.1, the total {t obtained within
the quasi-static approximation agrees very well with the low-frequency slope
{LDA of 1m d.L(w). The individual surface and bulk contributions are, however,
considerably larger than the total { . Although bulk and surface potentials
are spatially well separated, electron-hole pairs generated by them interfere
strongly and lead to large cancellations. For a consistent description of these
excitations, it is therefore essential to include both Coulomb and exchange-
correlation terms in the evaluation of the golden rule.
290 CHAPTER 7. ELECTRON~HOLE PAIR CREATION

This point is also illustrated by comparing the values of ~. with ~.,e.t in


the last row. The latter values are obtained if the surface part of 4>scf in
the golden rule is approximated by the electrostatic potential associated with
nl(z,w). (These values differ from ~RPA, where exchange-correlation terms are
also neglected in the evaluation of nd The large differences with respect to
the TDLDA results underline again the tremendous sensitivity of dynamical
surface response properties to model assumptions. The consistent treatment of
electron~electron interactions is particularly important for the near-adiabatic
excitations treated in this chapter.

7.3 Low-Frequency, Short-Wavelength Excita-


tions

In the remainder of Chapter 7, we consider perturbations of the electron


distribution via oscillating species only a few A from the surface: vibrating
dipoles, charges and atoms. In these phenomena, the dynamical response at
intermediate parallel wave vectors dominates: q rv lid rv 0.1 ... 1.0 A, where
d is the particle~surface distance. Because of the general importance of the
low-frequency surface response to short-range perturbations, we examine this
region here in more detail.
If excitation frequencies are small on the scale of surface collective modes or
the threshold excitation, the surface loss function can be written as (Persson,
Zaremba, 1985)
q w
Img(q,w) = 2 -k - ~(q) , (7.11)
F wp

where the coefficient ~(q) depends only on static response properties. An ex-
plicit expression of ~(q) in terms of the local potential 4>.cf may be derived from
the golden rule formula (2.45) as follows. Taking advantage of the translational
symmetry of jellium systems, the transition rate can be simplified to

(7.12)

where k' == (k.,+q, ky, [k~+2w-2k.,q_q2P/2). These relations follow from the
conservation of parallel momentum and single-particle energy, i.e., ck' = ck+w.
(We take !Jil along the x-direction.)
As a result of the occupation factors, the three-dimensional integral over
k may be reduced further. First, we make use of the relation K;~ :::; k~ :::; K;~,
7.3. LOW-FREQUENCY, SHORT-WAVELENGTH EXCITATIONS 291

where "'0 = (k~ - k; - k; - 2W)I/2 and "'1 = (k~ - k; - k;)1/2. Defining


'" = (k~ - k;)1/2, we have

(7.13)

The last identity follows from the expansion of "'1 - "'0 for small w. Let us
now introduce the variable <p via k", == '" cos <po Thus, dk", = -"'1 d<p. The
transition rate takes the form
811" W
w(q,w) = -k - ~(q) , (7.14)
F wp

where
(7.15)

The relation (2.64) then yields (7.11). The preceding derivation shows that
the linearity Img(q,w) w arises exclusively from phase space factors. The
f'.J

remaining terms may therefore be evaluated in the static limit. Thus, k~ =


(k; - 2"'qcos<p - q2)1/2 and 4>scf == 4>scf(Z,q,W = 0).
Notice that this derivation holds only if the local potential 4>scf involves
purely short-range surface contributions. This does not apply in the small-q
limit, where 4>scf acquires also long-range bulk-like behavior because of the
slow decay of eqz (see Section 7.2). Denoting the short-range surface part of
4>scf in the small-q limit by 4>. = 4>scf(Z, q = 0, w = 0) and setting k~ = kz for
q = 0, it follows from (7.15) that the linear surface coefficient coincides with
~s defined in (7.10).
Figure 7.5 illustrates the function ~(q) for several values of r •. These results
are obtained from the slopes of Img(q,w) calculated within the TDLDA for
w cv 0.1 eV. At small q, the functions ~(q) are seen to extrapolate very well to
the coefficient ~. defined in (7.10) (see Table 7.1). This is so because here we
first take the small-wand then the small-q limit, so that T} ~ 1. On the other
hand, owing to the discontinuity of bulk-like excitations at T} = 1, the total ~
defined in (7.9) corresponds to first taking the small-q limit and subsequently
letting w become small. This procedure implies T} ~ 1. Since the results in
Figure 7.5 are evaluated at a very low frequency, the bulk discontinuity occurs
at q = 0.01 A-I, i.e., in a range that is irrelevant for the damping and friction
phenomena considered in this chapter.
At large q, the surface response is seen to become very small when q ~
1.5 k F • Such a cut-off is to be expected since the metal electrons are no longer
292 CHAPTER 7. ELECTRON-HOLE PAm CREATION

6r----,r----,-----.-----,-----,----~

/
r.=2
2
-----/
,, ,,
, ..-.3 ,
4
OL---~-- __- L_ _ _ _ ~ _ _ _ _L-__~~__~

o 2 3

q (.&.-1)

Figure 7.5: Variation of ~(q) with parallel momentum for several


bulk densities. Solid curves: standardjellium model; dashed curves:
stabilized jellium model. The symbols at q = 0 denote the values
of ~. given in Table 7.1. The Fermi wave vectors are kF = 1.8, 1.2
and 0.9 A-I for r. = 2, 3 and 4, respectively. (Liebsch, 1997).

able to screen the rapidly varying external potential. Of course, q should not
exceed the decay constant of electronic states and Green's functions in the
vacuum. Large q vectors are important only at very short distances, i.e., when
the assumption of negligible overlap with metal states ceases to be valid.
The maximum of ~(q) at intermediate values of q is a consequence of the
increasing amplitude of ¢>ext near the centroid of the induced density. If this
effect is accounted for by multiplying~(q) by e- 2qz " where Zl is the position of
the static image plane in the q = 0 limit, the product is indeed a monotonically
decreasing function of q, just as we would expect on physical grounds.
To illustrate the sensitivity of these coefficients to the shape of the density
profile, Figure 7.5 also shows ~(q) for the stabilized jellium model for r. = 2
and r. = 3 (the results for r. = 4 are nearly unchanged). Since the equilib-
rium density of these surfaces is less polarizable than for standard jellium, the
probability of exciting electron-hole pairs is reduced.
Figure 7.6 shows the normalized density Til (Z, q, w) = e- qz , nl (z, q, w) in-
duced by the electric potential ¢>ext(f',w) = -(2rr/q) ei<lil·rll+qz e- iwt . Because
of the rather low frequency, these distributions are almost indistinguishable
7.3. LOW-FREQUENCY, SHORT-WAVELENGTH EXCITATIONS 293

.4

r.=2
,......
:i
~ .2
,......
3
.;
~
C
II> 0
0:::

-10 -5 o 5

z (aD)
Figure 7.6: Real part of normalized induced surface density
fi 1(z,q,w) at 0.1 eV as calculated within the TDLDA (r. = 2).
Solid curve: q = 0.1 A-I j dashed curve: q = 0.9 A-i.

from the corresponding static induced densities plotted in Figure (2.7). In


principle, at finite q and finite w, nl(z,q,W) also exhibits a propagating 'bulk'
contribution that vanishes in the adiabatic limit. This term has, however, a
very weak amplitude. In addition, in the evaluation of the surface response
function g(q,w), the bulk-like induced density is further suppressed by the
exponential weight factor eqz •

Extension to Realistic Metals

The TDLDA calculations at small w and finite q show that, to excellent


numerical accuracy, the golden rule expression (7.15) for e(q) gives the same re-
sults as those derived from Img(q,w), i.e., from the induced density n1(z, q,w).
This equivalence is of considerable practical importance for future evaluations
of e(<<iil) for realistic metals. Since the local potential ifJ.ef in the golden rule
expression can be taken in the static limit, it can be derived from existing
ground-state electronic structure codes by applying a weak static potential of
the form'" ei!ljl·i'jl+qz. As pointed out in Chapter 2, static response calcula-
tions at finite «iiI should be highly valuable for analyzing several important
phenomena involving low-frequency excitations at metal surfaces.
294 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

7.4 Damping of a Vibrating Dipole

The decay of vibrationally excited species adsorbed on metal surfaces can


occur via radiative or non-radiative processes. In the case of low-lying vibra-
tions of physisorbed or chemisorbed molecules, radiative decay is negligible.
Damping via excitation of phonons may also occur, but this process is effective
only at frequencies not far from the highest bulk phonon frequency. Here, we
focus on contributions to the lifetime arising from the excitation of electron-
hole pairs inside the metal. If the overlap between molecular and metallic
electronic states is small, the lifetime is due to direct Coulomb coupling of the
vibrating dipole to electronic excitations of the metal.
The typical width of molecular resonances is less than 1 meV. Inelastic
electron scattering therefore cannot be used since the resolution of electron
spectrometers (1 - 5 meV) is not sufficiently high. The preferred method is
infrared reflection-absorption spectroscopy (IRAS) which has a resolution of
0.1 meV or better. As discussed in Section 4.2.3, the reflectivity change of p-
polarized light due to surface excitations is directly related to the parameters
d.l and dll· For c ~ 0, we obtain from (4.71)

llRp 4w [. ]
R F = ---(} S1ll2(} Imd.l(w) - Imdll(w) ,
pecos
(7.16)

where () is the polar angle of incidence and R% the Fresnel reflectivity. While
EELS at low frequencies is largely determined by d.l (see (7.4)), in IRAS the
parameters d.l and dll are equally important.
According to the field expressions derived in Section 4.2.1, we have
E:r; 1 W
Z > 0, (7.17)
Ez ..fi sin () IV wp sin () ,
..fi IV
wp
Z < o. (7.18)
sin () W sin () ,
From the continuity of E." it then follows that Ez(z < 0) (w/w p )2Ez(z > 0).
IV

This shows that the external parallel field in IRAS is much smaller than the
perpendicular component, but it is not as much suppressed as in EELS (see
Figure 7.2 and Langreth, 1985). Near normal incidence, excitation via Ez
vanishes, and only electronic excitations induced parallel to the surface survive.
The field penetration depth in IRAS is therefore c/(wlcI 1 / 2 ), while in EELS
IV

it is approximately IV v/(wlcI / ). Here, v = w/qll is the velocity of the


1 2

incident electron in a typical inelastic electron scattering experiment. Thus,


IRAS is less surface sensitive than EELS by a factor of about v / c .
7.4. DAMPING OF A VIBRATING DIPOLE 295

Let us denote the frequency of the oscillating dipole by wand its position
relative to the surface by l = (0,0, d). The dipole is assumed to be oriented
normal to the surface. From Fermi's golden rule we find that the damping
rate is given by (Chance et al., 1978; Persson, 1978; Persson, Schaich, 1981;
Volokitin, Persson, 1993):

(7.19)

where X(f', f", w) is the exact many-body density-density response function of


the semi-infinite metal and
1
V(f') = Vz--~ (7.20)
If' - dl
is the effective potential acting on the electronic density; p, = I(altLIO) I is
the dynamic dipole moment, tL the dipole moment operator, and 10), la)
denote the ground state and a vibrationally excited state of the molecule.
Contributions to the damping rate due to multi pole fields of the molecule are
neglected.
In the case of jellium systems, the damping rate simplifies to

7- 1 -4np2 1000 dq q e- 2qd ! !


dz dz' eqz 1m X(z, z', q, w) eqz'

2p,2 1000 dqq2 e- 2qd lmg(q,w) . (7.21)

We assumed that the dipole does not overlap with the equilibrium density. If
the dipole is not entirely outside the range of the density, the preceding result
should be generalized to

7- 1 = -47rp,2 10 00 dqq! dz! dz' e-qld-zl sgn(d - z) e-q1d-z'l sgn(d - z')


x ImX(z,z',q,w) . (7.22)

This expression is valid only approximately since, as a result of the finite size
of the molecule, the interaction with the metal cannot be treated within the
point-dipole model.
Suppose the metal is described via a Drude dielectric function. For w ~ wp
we then have
c(w) - 1 WWF
1m g( q, w) = 1m () = 4 -lk (7.23)
c W +1
2'
F wp

where l is the electronic mean free path. Thus,


-1 2p,2 wWF
7 =---.
d3 kFl w; (7.24)
296 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

This contribution to the damping rate accounts for bulk electron-hole pair
creation due to impurities, phonons, and the lattice potential. These processes
yield a '" 1/d3 dependence and determine the vibrational lifetime at large
dipole-metal separations. At short distances, surface contributions to 7- 1
arising from the nonlocal response of the metal become important. For a
given d, (7.21) indicates that the surface loss function is required at q values
at least up to l/d.

CO on Cll

Infrared reflection-absorption measurements for CO adsorbed on Cu by


Ryberg (1982) showed that the total lifetime of the C-O stretch vibration is
about 7 = 1.3 X 10- 12 s. This corresponds to a full width at half-maximum of
0.5 meV. The mode frequency is 0.25 eV and the adsorbate-surface separation
is d ~ 1.6 A. The dynamic dipole moment is f.l = 0.25 D = 0.1 eao.
To analyze these data, Eguiluz (1984) performed dynamical surface re-
sponse calculations within a slab geometry. Since the dipole is not quite out-
side the electron density of the metal, (7.22) was used to evaluate the damping
rate. Within the RPA, the lifetime was found to be 7 = 9.1 X 10- 12 s. A sub-
sequent TDLDA response calculation gave 7 = 5.6 X 10- 12 s (Eguiluz, 1987).
A decrease is to be expected because exchange-correlation contributions to the
effective potential make the surface density profile more polarizable.
TDLDA calculations by Liebsch (1985) for semi-infinite jellium, with the
dipole assumed to be outside the electronic density, gave 7 = 5 X 10- 12 s.
This result was derived from a quasi-static treatment and included only the
electrostatic part of the local surface potential in the golden rule formula.
According to Table 7.1, this approximation leads to an overestimate of the
lifetime. A new TDLDA calculation for Cu (stabilized jellium, r. = 2.67)
gives 7 ~ 3.2 X 10- 12 s for the dipole outside the density (using (7.21)) and
7 ~ 6.0 X 10- 12 s for the dipole inside the density (using (7.22)).
Evidently, these dynamical response calculations for jellium surfaces yield
lifetimes of roughly the same magnitude as the experimental value. The over-
estimate is presumably related to the presence of another damping mechanism
caused by the overlap between molecular and metallic wave functions. In the
case of CO on Cu, the partial filling of the anti bonding27r* level and the associ-
ated oscillatory charge transfer to the substrate during the molecular vibration
seem to provide the most important decay channel (Persson, Persson, 1980).
In addition, we must be cautious about using the jellium model for noble met-
als, since the hybridization between sand d states reduces the polarizability
of the s electron density (see Section 2.1.4).
7.4. DAMPING OF A VIBRATING DIPOLE 297

clean

~
~
a:
~ T
<I 0.25%
..L

200 250 300 350 400 450 500


Frequency (cm-1)

Figure 7.7: Frequency variation of refiectance of a clean and adsor-


bate covered Cu(lOO) surface. The CO coverage is (J = 0.15 and
the temperature is T = 90 K. The features near 337 cm- 1 and
285 cm- 1 correspond to the perpendicular stretch vibration and to
the frustrated rotation, respectively. (Hirschmugl et al., 1990).

Damping of Parallel Vibrations

As we argued above, in infrared spectroscopy, the parallel electric field


outside the metal surface is not as strongly suppressed as in EELS. In fact,
according to (7.16), the adsorbate-induced refiectivity change at normal inci-
dence is given by
f).Rp 4w
Rj = ~ Imdll(w) . (7.25)

Although direct excitation of parallel molecular adsorbate vibrations is usually


forbidden because of symmetry selection rules, such modes have been observed
in lRAS on several systems (Chabal, 1988; Hirschmugl et al., 1990).
Figure 7.7 illustrates the refiectance change induced on Cu(100) by a low
coverage of CO. While the dip at 337 cm- 1 corresponds to absorption caused
by the C-O stretch vibration, the asymmetric spectral feature near 285 cm- 1
is due to the frustrated rotation of the CO molecule. Since direct excitation of
this mode is dipole-forbidden, the coupling is believed to take place indirectly,
via the tangential current induced in the metal by the parallel electric field.
The relative motion between this current and the adsorbate vibration gives
rise to a friction force that contributes to the vibrational lifetime. Persson
(1991) estimated the damping of parallel adsorbate modes via this mechanism
298 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

for several types of adsorbate-substrate bonds. In the next two sections, we


discuss the interaction of ions and atoms with metal surfaces for cases in which
the overlap with the surface electronic density can be neglected.

7.5 Sliding Friction of Ions

Low-frequency electronic excitations are important for the damping of the


motion of ions near a metal surface. The induced screening charge does not
follow the ion's motion adiabatically because of the finite response time of the
semi-infinite electron gas. As a result of this time lag, the force acting on the
ion at a particular time differs from its adiabatic value. Electronic excitations
in the metal therefore give rise to dynamical corrections to the image potential
(Harris, Jones, 1973; Mahan, 1973; Schaich, 1974; Ferrel et al., 1979; Sols,
Flores, 1982).
Let M and Q be the mass and charge of the ion and d its distance from
the surface. The friction force may be written as

f= -M (7711 Vii + TJ.L'ih) . (7.26)

Here, if is the particle velocity, and TJII,.L are the friction coefficients for the
motion parallel and normal to the surface. If the charge is located outside the
electronic density of the metal, it can be shown that TJII = TJ.L/2 (Harris, Jones,
1974).
The ionic friction coefficient is closely related to the vibrational lifetime
discussed in Section 7.2. This may be seen by expressing the dynamic dipole
moment as 11 = Q8, where 8 = 1/v'2Mw is the normal-mode coordinate
of the oscillation. Thus, TJ.L takes the form (D'Agliano et al., 1975; Persson,
Schaich, 1981)

2
TJ.L =-Q
M wlim !:.!d3 r!d 3 r'
.... o W
V(f') ImX(f,f',w) V(f') , (7.27)

where X( f, f', w) is again the exact many-body density-density response func-


tion of the semi-infinite metal and V(f') is given by (7.20). The limit w ~ 0
applies to slowly moving ions. For jellium systems, (7.27) simplifies to

1 1000
TJ.L = Q2
M lim - dqq 2 e -2q d Img(q,w) . (7.28)
w-+o w 0
7.5. SLIDING FRICTION OF IONS 299

3,-----,------,------,-----,------,

2 r.=2

~ 3
"-

0
0 2 4

d (A)
Figure 7.8: Ionic friction integral F(d} as a function of ion-surface
separation for several bulk densities. The metal is described within
the stabilized jellium model and the dynamical response is treated
within the TDLDA. (Liebsch, 1997).

At large ion-surface separations, bulk processes dominate. In this case, the


surface loss function can be approximated as in (7.23). The friction parameter
is then
Q2 WF 1
'fJJ. = M kFlw; d3 •
(7.29)

At short distances, electronic excitations in the surface region become impor-


tant, so that the full q-variation of Img(q,w} must be taken into account.
Using the low-frequency form of g(q,w} given in (7.11), the friction coefficient
can be written as
Q2 2 3!
'fJJ. = M kFWp (2d}4 F(d) ,
(7.30)

where F(d} is the dimensionless ionic friction integral defined via

(2d}4
F(d) = - -
1000 dqq3e-2qd~(q} . (7.31)
3! 0

The TDLDA values of this function are displayed in Figure 7.8 for several r •.
At large ion-surface separations, F(d} approaches ~s specified in (7.10).
300 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

K on eu
As an application, we consider the damping of the parallel frustrated trans-
lation of a charged alkali atom adsorbed on a metal surface. The adatom is
treated as a point charge of magnitude Q that oscillates parallel to the surface
at frequency w. In the case of a low coverage of potassium atoms on Cu(100),
the distance of the K nucleus from the effective jellium edge (half a lattice
constant above the first plane of nuclei) is about 5 ao. From the measured
work function change at low coverage, we can deduce the effective charge as
Q ~ 0.88 Ie!- With these parameters we find for stabilized jellium (r. = 2.67,
F(d) = 1.65)
7711 = 0.5 771. ~ 1.6 X 109 S-1 , (7.32)
so that T = 1/7711 ~ 0.6 x 10-9 s (Liebsch, 1997). T is therefore similar
to the lifetime estimated by Persson (1991) for a covalent bond model (T ~
0.7 X 10-9 s). In the latter case, the K 4s level is broadened into a resonance due
to hybridization with the electronic states of the Cu substrate. This interaction
leads to a damping since the phase of the effective potential changes during
the oscillatory motion of the adatom.

7.6 Sliding Friction of Physisorbed Atoms


In Chapter 6, we saw that the instantaneous mutual polarization between
neutral species gives rise to a long-range attractive interaction. If the atom
moves relative to the metal, the induced surface charge lags behind. Low-
frequency electronic surface excitations are therefore also relevant for the fric-
tion of neutral atoms and molecules (Schaich, Harris, 1981; Sols et at., 1982,
1984; Sokoloff, 1995; Persson, Volokitin, 1995). Dynamical corrections to the
physisorption potential can be observed, for example, in quartz-crystal mi-
crobalance measurements that detect the frequency shift and broadening of
the lateral oscillation due to adsorption (Krim et at., 1991). In principle, the
short-range repulsive and long-range attractive interactions contribute, but
we consider here only the damping caused by the van der Waals attraction.
Accordingly, we ignore all effects related to the overlap between the charge
densities of metal and adsorbate.
As recently shown by Persson and Volokitin (1995), the friction coefficient
for physisorbed atoms can be expressed in a similar manner as the ionic coef-
ficient:

(7.33)
7.6. SLIDING FRICTION OF PHYSISORBED ATOMS 301

The index i = x, z refers to parallel or perpendicular motion of the atom, and


X is the many-body response function of the metal. The effective interaction
V;(r') is approximately given by

V;( r') =

(7.34)

Here, a(O) is the static polarizability of the adatom and l = (0,0, d) is the
location relative to the jellium edge. Since V;(r') does not satisfy the Laplace
equation, it can only be approximately represented in terms of a superposition
of evanescent plane waves of the form

V;(f') = L V;(Qi,) eiqjl·rll+qZ . (7.35)


qll

Inserting this expansion into (7.33), we can express TJi in terms of the surface
loss function:

TJi e2
=M 1 lim.!.
-2
7r w-tO W
L
_
1V;(Qi,W q Img(q,w) . (7.36)
qll

If we now use the low-frequency behavior of Img(q,w) given in (7.11), the


friction coefficient takes the form

TJi =
qjl

e2 [k~a(OW m WF
ao (kFd)lO M wp kFao Ii(d) . (7.37)

The dimensionless atomic friction integrals Ii (d) are defined as

(7.38)

and ij" = ij - ij', Q", = q~/q/l, Qz = 1. The TDLDA results for these
functions are shown in Figure 7.9. At large atom-surface separations, these
integrals scale like the coefficients ~s specified in (7.10).
302 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

50
r,=2
,-...
'0
~
... 3

0
0 2 4

10

OL-__ ~~~ __ ~ ______L -_ _ _ _ ~ ____ ~

o 2 4

d (A)

Figure 7.9: Atomic friction integrals J.L(d) and III(d) as functions


of atom-surface separation for several bulk densities. The metal
is described within the stabilized jellium model and the dynamical
response is treated within the TDLDA. (Liebsch, 1997).

Xe on Ag

In the case of Xe atoms physisorbed on Ag(111), one has d = 2.4 A. The


static polarizability of Xe is a(O) = 4.0 A3. Using the TDLDA results for {(q)
shown in Figure 7.5, we find for stabilized jellium (r. = 3) 1711 rv 3.4 X 108 S-l
and 171. rv 17.4 X 108 S-l (Liebsch, 1997).
The calculated parallel friction coefficient 1711 is in qualitative agreement
with the values obtained from recent experiments for Xe on Ag: Surface resis-
tivity data yield 1711 rv 3 X 108 S-l (Holzapfel et al., 1990), while quartz-crystal
7.7. SURFACE RESISTIVITY 303

microbalance measurements give 1]11 rv 8 X 108 S-l (Daly, Krim, 1996, 1997).
For phonon-related friction of isolated Xe atoms on Ag, Persson and Nitzan
(1996) estimated a significantly smaller value: 1]11 rv 0.6 X 108 S-l. This value
should also be a reasonable approximation for fluid adsorbate layers. For
incommensurate solid adsorbate layers, on the other hand, the phonon con-
tribution to the friction is expected to vanish (Sokoloff, 1990). The friction
coefficient arising from the van der Waals attraction is also significantly larger
than the contribution due to the Pauli repulsion. The latter has been estimated
~t about 1]11 '" 0.6 x 108 S-l (Persson, 1991).
The theoretical value of 1]11 given above presumably represents a slight over-
estimation since the Xe atom is not completely outside the range of the elec-
tronic density profile. Moreover, the derivation of (7.33) implies some uncer-
tainty. Also, as a result of the s-d hybridization, the surface polarizability of
real Ag may be slightly smaller than that of the corresponding jellium model.
We note, in addition, that for Xe there may exist some friction due to chemi-
cal effects resulting from the broadening of the Xe 6s level. Estimates of this
mechanism yield 1]11 rv 1.5 X 108 S-l (Persson, 1991). For the lighter rare gas
atoms, this effect should be negligible, since the lowest unoccupied s level does
not extend to the Fermi energy.

7.7 Surface Resistivity

The electrical resistivity of thin metal films is known to be highly sensitive


to surface conditions. (For an excellent review of this topic, see Schumacher,
1993.) Overlayer coverages of 0.1 % of a monolayer can easily be detected.
As previously mentioned, the resistivity increase Pa of a thin metal film on
adsorption can be used to measure the friction coefficient 1]11 of the parallel
vibrational motion of adsorbed species. This remarkable relation was recently
derived by Persson (1991) who showed that

(7.39)

Here, T is the lifetime of the vibration, M the adsorbate mass, na the areal
density of the adsorbed species, d the thickness of the metal film, and n the
conduction electron density in the metal.
304 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

o
Figure 7.10: Schematic illustration of adsorbate-induced resistivity
(left) and of damping of parallel vibration of adsorbed atom (right).

7.7.1 Classical Picture

A simple way of proving (7.39) is to consider the ohmic heating generated


by the adsorbate if a homogeneous current flows along a metal film:

(7.40)

where A is the surface area. The current density j", = E",/ Pa induced by the
parallel electric field E", is related to the drift velocity v'" of the conduction
electrons via j", = env",. Hence,

(7.41)

Another way of expressing P is to change the reference frame and consider


the adatom moving with velocity -v", parallel to the metal surface (see Figure
7.1O). The atom feels a friction force given by

F", = Mil", = 7- 1 Mv", , (7.42)

indicating that the velocity diminishes according to v", rv e- t / r , where 7 is the


slip time due to excitation of electron-hole pairs in the metal. Contributions to
7 due to phonons are neglected. The dissipation associated with N adsorbed
atoms is given by

P = N F", v", = na A 7- 1 M v~ , (7.43)

where na = N / A is the number of adatoms per unit area. Equating (7.41) and
(7.43), we obtain the identity (7.39). The above derivation is closely related
7.7. SURFACE RESISTIVITY 305

to the energy loss concept employed by Gerlach (1984, 1986) to calculate the
resistivity induced via impurities in semiconductors and adsorbates on metal
films.
Note that the friction force F", may be interpreted as the wind force acting
on the adatoms as a result of the applied electric field E",. In contrast to the
direct force, which is proportional to the ionic charge of the adsorbate nuclei,
the wind force or indirect force is caused by scattering of current-carrying
conduction electrons from the adatoms. Using (7.39), F", may be written as
ne .
F", = - dPaJ",. (7.44)
na
A similar relation has long been known to hold in the bulk (Schaich, 1976;
Sorbello, 1981; Lodder, 1984).
The presence of adsorbed atoms also affects the reflectivity of the metal.
According to (7.16), at infrared frequencies the reflectivity change for normal
incidence is
(7.45)

where R: is the Fresnel reflectivity. To determine d ll , let us describe the


:s :s
dielectric response of the surface region (-d z 0) by the Drude function

(7.46)

The lifetime Ta is related to the dc adsorbate-induced resistivity via

(7.47)

In the interior of the metal (z :s -d), absorption processes are ignored and the
dielectric function is simply c = 1- w;/w2. Using the relations for d ll derived
in Section 4.2.2, it can easily be shown that

(7.48)

From (7.39) and (7.47), it follows that for large d

(7.49)

The adsorbed layer therefore reduces the reflectivity according to

-4 M na ~. (7.50)
m n CT
306 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

This result proves that the adsorbate-induced reflectivity change provides in-
formation on the lifetime T of the parallel frustrated translation of the adsor-
bate. Persson (1991) has argued that, for CO on Cu, the lifetime obtained
from the reflectivity change on adsorption (Hirschmugl et al., 1990; see Figure
7.7) is roughly consistent with dc resistivity data.

7.7.2 Microscopic Theory

The picture outlined above was recently formulated microscopically by


Ishida (1994, 1995) who employed the TDLDA to evaluate the dynamical
response of realistic overlayers to a parallel electric field. The semi-infinite
substrate is assumed to be a jellium metal. The surface region may consist
of an ordered layer of chemisorbed atoms at an arbitrary coverage, or of sev-
eral surface layers of substrate material. In the latter geometry, the resistivity
change due to surface reconstruction or formation of steps can be investigated.
The ground-state properties are determined using the embedding scheme (see
Sections 2.3.6 and 4.7.2).
The effective Hamiltonian describing the interaction with the parallel field
has the form

H' -
- -Wi E.,p., + jd3 r , nl(T',w} avxc[n11
1_r - r-'I + -a-- nl
(-)
r, W , (7.51)
n no

where p., is the momentum operator. The first term corresponds to the bare
applied field, while the second and third terms account for the Coulomb and
exchange-correlation screening contributions, respectively. The solution of
the self-consistency equation yields the induced density distribution nl (1", w).
From the linear change of the Hellmann-Feynman force, it is then possible to
derive the friction force F., acting on the adsorbate layer as a result of E",.
Using the conductivity tensor defined in (2.113), we may also determine
the current density j.,(r, w} induced by E.,:

. (-)
J.,r,w= e2 jd3ra",,,,r,r,w
ie2no(T) E"'+m ' P (- -, ) E'" (7.52)
mw
(non-diagonal elements are neglected). The first term gives the macroscopic
current flowing in the interior of the metal

(7.53)
7.7. SURFACE RESISTIVITY 307

while the second term specifies the microscopic current localized in the surface
region.
The energy dissipation per unit time is given by

From (7.40), it then follows that the adsorbate-induced resistivity may be


expressed in terms of the nonlocal conductivity as

d Pa ()
W = mwn
Ae
2

2 2
Jd r Jd r'R
3 3 P (~~'
e (J"'''' r, r ,W ) . (7.55)

Ishida (1994) also proved that the microscopically derived friction force is
related to the nonlocal conductivity tensor via

(7.56)

Combining this expression with (7.55), we obtain

. F", ne )
hm --:- = - dpa(w -+ 0 . (7.57)
w--+O J", na

Hence, the microscopic approach fully confirms the classical expression (7.44)
for the friction force acting on the adatoms. Note that (7.52) applies in
the independent-electron picture. In the TDLDA, Coulomb and exchange-
correlation terms should also be taken into account, as indicated in (7.51).
Nevertheless, since the adsorbate-induced current density remains localized
near the surface, these corrections to the friction force vanish in the static
limit. Thus, (7.56) also remains valid in the presence of electron interactions.
Using (4.55) for dll(w}, it can also be shown that

dIlw( )-- -An-1 Jd3 r zdnodz-(T)- iw


An Jd r Jd r
3 3 ' P
(J",,,, (~~'
r, r ,W ) . (7.58)

Thus,
(7.59)
The microscopic theory of surface resistivity therefore yields the same relation
between Pa and dll as the macroscopic picture discussed earlier.
Ishida evaluated the current density induced by a uniform electric field E",
for various kinds of overlayers on a jellium substrate. As an example, Figure
7.11 illustrates the friction force F"" normalized by the macroscopic lateral
308 CHAPTER 7. ELECTRON-HOLE PAIR CREATION

20
IJ. _ - 8 - ____ -0- _______ - - - - - - - - - - - - - - - - --e

Re
10

.
~, __._--&--_ _ .__ ,_ _ _ ,_ _ _ ~-----------EI

...,
o
-------"-£1
1m
----a
-10

o .01 .02 .03

w (a.u.)

Figure 7.11: Normalized friction force F",/i", = dpa(w)ne/n a for


Si layer on jellium (r. = 2) as a function of frequency. The upper
(lower) two lines denote the real (imaginary) parts. Solid lines: self-
consistent TDLDA response; dashed lines: independent-particle re-
sponse. The symbol on the vertical axis indicates F",/ j", in the static
limit. (Ishida, 1994).

current j." for a hexagonal Si layer as a function of frequency. The interatomic


spacing is 12 ao, corresponding to a very low coverage with negligible wave
function overlap within the overlayer. The distance above the jellium substrate
is 2.3 ao, giving rise to a strong chemisorption bond. The results show that
the independent-particle picture overestimates the friction force about a factor
of two. Hence, when evaluating the adsorbate-induced resistivity, a proper
treatment of surface screening processes is clearly important. The results also
prove that the TDLDA values obtained at finite frequencies extrapolate very
well to the friction force derived from (7.56) in the static limit.
In an extension of this work, Ishida (1995) recently calculated the micro-
scopic surface resistivity for stepped Al surfaces. The resistivity in the direction
perpendicular to the steps was found to be about twice as large as along the
steps. This is plausible, since the steps act as scattering centers for electronic
waves propagating orthogonally to the steps.
Bibliography

Chapter 1

Apell, P., Phys. Scr. 24, 795 (1981).

Apell, P., A. Ljungbert, and S. Lundqvist, Phys. Scr. 30, 367 (1984).
Dobson, J.F., in Density Functional Theory, K.U. Gross and R.M. Dreizler,
eds. (Plenum, New York, 1995).
Eguiluz, A.G., Phys. Scr. 36, 651 (1987).
Ehrenreich, H., and M.H. Cohen, Phys. Rev. 115, 786 (1959).
Feibelman, P.J., Prog. Surf. Sci. 12, 287 (1982).
Forstmann, F., and R.R. Gerhardts, Metal Optics Near the Plasma Frequency,
Springer Tracts in Modern Physics 109 (Springer, Berlin, 1986).
Gerhardts, R.R., in Electromagnetic Waves: Recent Developments in Re-
search; vol. 1: Spatial Dispersion in Solids and Plasmas, P. Halevi, ed.
(Elsevier, Amsterdam, 1992), p. 109.
Hohenberg, P., and W. Kohn, Phys. Rev. 136, B864 (1964).
Kohn, W., and L.J. Sham, Phys. Rev. 140, A1l33 (1965).
Liebsch, A., Phys. Scr. 35, 354 (1987)
Liebsch, A., in Electromagnetic Waves: Recent Developments in Research;
vol. 2: Photonic Probes of Surfaces, P. Halevi, ed. (Elsevier, Amsterdam,
1995), p. 479.
Mahan, G.D., Phys. Rev. A 22, 1780 (1980).
Mahan, G.D., and K.R. Subbaswamy, Local Density Theory of Polarizability
(Plenum, New York, 1990).
Raether, H., Surface Plasmons, Springer Tracts in Modern Physics 111 (Sprin-
ger, Berlin, 1988).

309
310 BIBLIOGRAPHY

Ritchie, R.H., Phys. Rev. 106, 874 (1957).


Rocca, M., Surf. Sci. Rep. 22, 1 (1995).
Schaich, W.L., and K. Kempa, Phys. Scr. 35,204 (1987).
Stott, M.J., and E. Zaremba, Phys. Rev. A 21, 121 (1980).
Wikborg, E., and J.E. Inglesfield, Phys. Scr. 15, 37 (1977).
Zangwill, A., and P. Soven, Phys. Rev. A 21, 1561 (1980).
Zangwill, A., Physics at Surfaces (Cambridge University, Cambridge, 1988).

Chapter 2

Aers, G.C., and J.E. Inglesfield, Surf. Sci. 217, 367 (1989).
Ando, T., Solid State Commun. 21, 133 (1977).
Appelbaum, J.A., and D.R. Hamann, Phys. Rev. B 6, 1122 (1972).
Bliigel, S., doctoral thesis, RWTH Aachen (1987).
B6hm, H.M., S. Conti, and M.P. Tosi, J. Phys. CM 8, 781 (1996).
Ceperley, D.M., and B.J. Alder, Phys. Rev. Lett. 45, 566 (1980).
Dabrowski, B., Phys. Rev. B 34,4989 (1986).
Dobson, J.F., Phys. Rev. Lett. 73,2244 (1994).
Dobson, J.F., and G.H. Harris, Phys. Rev. B 27, 6542 (1983).
Dobson, J.F., and G.H. Harris, J. Phys. C 21, L729 (1988).
Dreizler, R.M., and E.K.U. Gross, Density Functional Theory (Springer,
Berlin, 1990).
Eguiluz, A.G., Phys. Rev. B 31, 3303 (1985).
Eguiluz, A.G., and W. Hanke, Phys. Rev. B 39, 10433 (1989).
Ekardt, W., Phys. Rev. B 31, 6360 (1985).
Garda-Gonzalez, P., and A. Liebsch, to be published (1997).
Gies, P., and R.R. Gerhardts, Phys. Rev. B 31, RC6843 (1985).
Gies, P., and R.R. Gerhardts, Phys. Rev. B 36, 4422 (1987).
Gross, E.K.U., and W. Kohn, Phys. Rev. Lett. 55, 2850 (1985); Erratum:
ibid. 57, 923 (1986).
BIBLIOGRAPHY 311

Haensel, R., G. Keitel, P. Schreiber, and C. Kunz, Phys. Rev. 188, 1375
(1969).
Hartree, D.R., The Calculation of Atomic Structures, Sec. 5.3.3 (Wiley, New
York, 1957).
Hohenberg, P., and W. Kohn, Phys. Ret}. 136, B864 (1964).
Inglesfield, J.E., J. Phys. C 14,3795 (1981).
Inglesfield, J.E., Surf. Sci. 188, L701 (1987).
Ishida, H., and A. Liebsch, Phys. Rev. B 45, 6171 (1992).
Ishida, H., and A. Liebsch, Phys. Rev. B 50, 4834 (1994).
Iwamoto, N., and E.K.U. Gross, Phys. Rev. B 35, 3003 (1987).
Jepsen, 0., M. Madsen, and O.K. Andersen, J. Magn. Magn. Mat. 15-18,
867 (1980).
Jones, R.O., and O. Gunnarsson, Rev. Mod. Phys. 61, 689 (1989).
Kempa, K., and W.L. Schaich, Phys. Rev. B 37, 6711 (1988).
Kiejna, A., in Inelastic Energy Transfer in Interactions with Surfaces and Ad-
sorbates, B. Gumhalter, A.C. Levi, and F. Flores, eds. (World Scientific,
Singapore, 1993) p. 129.
Kiejna, A., Surf. Sci. 331-333, 1167 (1995).
Kohn, W., Phys. Rev. 123, 1242 (1961).
Kohn, W., and L.J. Sham, Phys. Rev. 140, A1133 (1965).
Lang, N.D., Solid State Phys. 28, 225 (1973).
Lang, N.D., Phys. Rev. Lett. 46, 842 (1981).
Lang, N.D., in Theory of the Inhomogeneous Electron Gas, S. Lundqvist and
N.H. March, eds. (Plenum, New York, 1983).
Lang, N.D., and W. Kohn, Phys. Rev. B 1, 4555 (1970).
Lang, N.D., and W. Kohn, Phys. Rev. B 3, 1215 (1971).
Lang, N.D., and W. Kohn, Phys. Rev. B 7, 3541 (1973).
Langreth, D.C., Phys. Rev. B 5, 2842 (1972).
Levine, Z. H., and P. Soven, Phys. Rev. A 29, 625 (1984).
Liebsch, A., Phys. Rev. Lett. 54, 67 (1985a).
312 BIBLIOGRAPHY

Liebsch, A., Phys. Rev. B 32, 6255 (1985b).


Liebsch, A., J. Phys. C 19, 5025 (1986).
Liebsch, A., Phys. Rev. B 36, 7378 (1987).
Liebsch, A., and W.L. Schaich, Phys. Rev. B 40, 5401 (1989).
Lundqvist, S., and N.H. March, eds. Theory of the Inhomogeneous Electron
Gas (Plenum, New York, 1983).
Mahan, G.D., Phys. Rev. A 22,1780 (1980).
Mahan, G.D., Many-Particle Physics (Plenum, New York, 1990).
Mahan, G.D., and K.R. Subbaswamy, Local Density Theory of Polarizability
(Plenum, New York, 1990).
Manninen, M., R.M. Nieminen, P. Hautojarvi, and J. Arponen, Phys. Rev.
B 12,4012 (1975).
Monnier, R., J.P. Perdew, D.C. Langreth, and J.W. Wilkins, Phys. Rev. B
18, 656 (1978).
Nuroh, K., M.J. Stott, and E. Zaremba, Phys. Rev. Lett. 49, 862 (1982).
Ossicini, S., C.M. Bertoni, and P. Gies, Surf. Sci. 178, 244 (1986).
Parr, R.G., and W. Yang, Density Functional Theory of Atoms and Molecules
(Oxford University, Oxford, 1989).
Perdew, J.P., H.Q. Tran, and E.D. Smith, Phys. Rev. B 42, 11627 (1990).
Perdew, J.P., and A .. Zunger, Phys. Rev. B 23, 5048 (1981).
Persson, B.N.J., and E. Zaremba, Phys. Rev. B 31, 1864 (1985).
Peuckert, V., J. Phys. C 11,4945 (1978).
Pines, D., Elementary Excitations in Solids (Benjamin, New York, 1964).
Puska, M.J., R.M. Nieminen, and M. Manninen, Phys. Rev. B 31, 3486
(1985).
Quong, A.A., A.A. Maradudin, R.F. Wallis, J.A. Gaspar, A.G. Eguiluz and
G.P. Alldredge, Phys. Rev. Lett. 66, 743 (1991).
Rubio, A., L.C. Balbas, and J.A. Alonso, Phys. Rev. B 46, 4891 (1992).
Runge, J.H., and E.K.U. Gross, Phys. Rev. Lett. 52,997 (1984).
Senatore, G., and K.R. Subbaswamy, Phys. Rev. B 35, 2440 (1987).
Shore, H.B., and J.H. Rose, Phys. Rev. Lett. 66, 2519 (1991).
BIBLIOGRAPHY 313

Stott, M.J., and E. Zaremba, Phys. Rev. A 21, 121 (1980).


Sturm, K., private communication (1995).
Sugiyama, A., J. Phys. Soc. Japan 15, 965 (1960).
Tsuei, K.D., E.W. Plummer, A. Liebsch, E. Pehlke, K. Kempa, and P. Bakshi,
Surf. Sci. 247,302 (1991).
Ullrich, C.A., U.J. Gossmann, and E.K.U. Gross, Phys. Rev. Lett. 74, 872
(1995).
Vignale, G., and W. Kohn, Phys. Rev. Lett. 77,2037 (1996).
Weber, M., and A. Liebsch, Phys. Rev. B 35, 7411 (1987).
Weinert, M., private communication (1994).
Zangwill, A., and P. Soven, Phys. Rev. A 21, 1561 (1980).
Zangwill, A., J. Chem. Phys. 78, 5926 (1983).
Zhang, Z.Y., D.C. Langreth, and J.P. Perdew, Phys. Rev. B 41, 5674 (1990).

Chapter 3

Andersson, S., and U. Jostell, Surf. Sci. 46, 625 (1974).


Andersson, S., and U. Jostell, Faraday Discuss. Chem. Soc. 60, 255 (1975).
Apell, P., Phys. Scr. 24, 795 (1981).

Apell, P., and A. Ljungbert, Solid State Commun. 44, 1367 (1982).

Aruga, T., and Y. Murata, Prog. Surf. Sci. 31, 61 (1989).


Aryasetiawan, F., and K. Karlsson, Phys. Rev. Lett. 73, 1679 (1994).
Barberan, N., J. Sellares, and J. Bausells, Surf. Sci. 292, 195 (1993).
Beck, D.E., and B.B. Dasgupta, Phys. Rev. B 12, 1995 (1975).
Bennett, A.J., Phys. Rev. B 1, 203 (1970).
Bonzel, H.P., Surf. Sci. Rep. 8, 43 (1987).
Bonzel, H.P., A.M. Bradshaw, and G. Ertl, eds. Alkali Adsorption on Metals
and Semiconductors (Elsevier, Amsterdam, 1989).
Brechignac, C., Ph. Cahuzac, N. Kebaili, J. Leygnier, and A. Sarfati, Phys.
Rev. Lett. 68,3916 (1992).
314 BIBLIOGRAPHY

Bnkhignac, C., Ph. Cahuzac, J. Leygnier, and A. Sarfati, Phys. Rev. Lett.
70,2036 (1993).
Charle, K.P., W. Schulze, and B. Winter, Z. Phys. D 12, 471 (1989).
Contini, R., and J.M. Layet, Solid State Commun. 64, 1179 (1987).
Cousty, J., R. Riwan, and P. Soukiassian, J. Phys. {Paris} 46, 1693 (1985).
Dobson, J.F., J. Phys. B 46, 10163 (1992).
Dobson, J.F., Aust. J. Phys. 46, 391 (1993).
Dobson, J.F., and G.H. Harris, J. Phys. C 21, L729 (1988).
Duke, C.B., L. Pietronero, J.O. Porteus, and F. Wendelken, Phys. Rev. B
12, 4059 (1975).
Eguiluz, A.G., and J.J. Quinn, Phys. Lett. A 53, 151 (1973).
Eguiluz, A.G., and A. Campbell, Phys. Rev. B 31, 7572 (1985).
Eguiluz, A.G., and J.A. Gaspar, Springer Proc. Phys. 62, M. Cardona and
F. Ponce, eds. (Springer, Heidelberg, 1991) p. 23.
Ehrenreich, H., and H.R. Phillip, Phys. Rev. 128, 1622 (1962).
Ekardt, W., Phys. Rev. B 31, 6360 (1985a).
Ekardt, W., Phys. Rev. B 32, 1961 (1985b).
Feibelman, P.J., Phys. Rev. 178, 551 (1968).
Feibelman, P.J., Phys. Rev. Lett. 30, 975 (1973).
Feibelman, P.J., Phys. Rev. B 9, 5077 (1974).
Feibelman, P.J., Phys. Rev. B 14, 762 (1976).
Feibelman, P.J., Phys. Rev. B 22, 3654 (1980).
Feibelman, P.J., Prog. Surf. Sci. 12, 287 (1982).
Feibelman, P.J., Surf. Sci. 282, 129 (1993).
Feibelman, P.J., Phys. Rev. Lett. 72, C788 (1994).
Fleszar, A., R. Stumpf, and A.G. Eguiluz, Phys. Rev. B 55, 2068 (1997).
Flores, F., and F. Garcfa-Moliner, Solid State Commun. 11, 1295 (1972).
Foo, E.Ni, and J.J. Hopfield, Phys. Rev. 173, 635 (1968).
Forstmann, F., and H. Stenschke, Phys. Rev. B 17, 1489 (1979).
BIBLIOGRAPHY 315

Fuchs, R., in Electromagnetic Waves: Recent Developments in Research; vol.


1: Spatial Dispersion in Solids and Plasmas, P. Halevi, ed. (Elsevier,
Amsterdam, 1992), p. 289.
Gadzuk, J.W., Phys. Rev. B 1, 1267 (1970).
Garcia-Gonzalez, P., and A. Liebsch, to be published (1997).
Gaspar, J.A., A.G. Eguiluz, K.D. Tsuei, and E.W. Plummer, Phys. Rev.
Lett. 67, 2854 (1991).
Gies, P., and R.R. Gerhardts, Phys. Rev. B 31, 6843 (1985).
Gies, P., and RR. Gerhardts, Phys. Rev. B 36, 4422 (1987).
Gross, E.K.U., and W. Kohn, Phys. Rev. Lett. 55, 2850 (1985); Erratum:
ibid. 57, 923 (1986).
Gwinn, E., R.M. Westervelt, P.F. Hopkins, RM. Rimberg, M. Sundaram,
and A.C. Gossard, Phys. Rev. B 39, 6260 (1989).
Haberland, H., B. v. Issendorf, Ji Yufeng, and T. Kolar, Phys. Rev. Lett.
69, 3212 (1992).
Hagemann, H.J., W. Gudat, and C. Kunz, J. Opt. Soc. Am. 65, 742 (1975).
Harris, J., and A. Griffin, Phys. Lett. 34A, 51 (1971).
Heskett, D., K.H. Frank, K. Horn, E.E. Koch, H.J. Freund, A. Baddorf, K.D.
Tsuei, and E.W. Plummer, Phys. Rev. B 37, 10387 (1988).
Hovel, H., S. Fritz, A. Hilger, U. Kreibig, and M. Vollmer, Phys. Rev. B 48,
18178 (1993).
Ibach, H., and D.L. Mills, Electron Energy Loss Spectroscopy and Surface
Vibrations (Academic, New York, 1982).
Inglesfield, J., and E. Wikborg, J. Phys. F 5, 1706 (1975).
Ishida, H., and A. Liebsch, Phys. Rev. B 45, 6171 (1992).
Ishida, H., and A. Liebsch, Phys. Rev. B 54, 14127 (1996).
Ishida, H., and M. Tsukada, Surf. Sci. 169, 225 (1986).
Johnson, P.B., and R.W. Christy, Phys. Rev. B 6, 4370 (1972).
Kempa, K., and RR. Gerhardts, Solid State Commun. 53, 579 (1985).
Kempa, K., and W.L. Schaich, Phys. Rev. B 32, 8375 (1985).
Kempa, K., and W.L. Schaich, Phys. Rev. B 37, 6711 (1988).
316 BIBLIOGRAPHY

Kiejna, A., in Inelastic Energy Transfer in Interactions with Surfaces and Ad-
sorbates, B. Gumhalter, A.C. Levi, and F. Flores, eds. (World Scientific,
Singapore, 1993), p. 129.
Kiejna, A., and J. Peisert, Surf. Sci. 320, 355 (1994).
Kim, B.O., G. Lee, E.W. Plummer, P.A. Dowben, and A. Liebsch, Phys. Rev.
B 52, 6057 (1995).
Kim, B.O., E.W. Plummer, and A. Liebsch, to be published (1997a).
Kim, J.S., L. Chen, L.L. Kesmodel, P. Garcia-Gonzalez, and A. Liebsch, to
be published (1997b).
Kloos, T., and H. Raether, Phys. Lett. 44A, 157 (1973).
Kramar, T., D. Vogtenhuber, R. Podloucky, and A. Neckel, Electrochimica
Acta 40,43 (1995).
Krane, K.J., and H. Raether, Phys. Rev. Lett. 37, 1355 (1976).
Kreibig, U., and L. Genzel, Surf. Sci. 156, 678 (1985).
Kresin, V.V., Phys. Rev. B 51, 1844 (1995).
Kunz, C., Z. Phys. 196, 311 (1966).
Lang, N.D., Phys. Rev. B 4, 4234 (1971).
Lang, N.D., and W. Kohn, Phys. Rev. B 1, 4555 (1970).
Lee, G., P.T. Sprunger, E.W. Plummer, and S. Suto, Phys. Rev. Lett. 67,
3198 (1991).
Lee, G., P.T. Sprunger, and E.W. Plummer, Surf. Sci. 286, L547 (1993).
Liebsch, A., Phys. Rev. B 32, 6255 (1985).
Liebsch, A., Phys. Rev. B 36, 7378 (1987).
Liebsch, A., Phys. Rev. Lett. 67, 2858 (1991).
Liebsch, A., Phys. Rev. Lett. 71, 145 (1993a).
Liebsch, A., Phys. Rev. B 48,11317 (1993b).
Liebsch, A., Phys. Rev. Lett. 72, C789 (1994).
Liebsch, A., and W.L. Schaich, Phys. Rev. B 52, 14219 (1995).
Lipparini, N., and F. Pederiva, Z. Phys. D 22, 553 (1992).
Lopez, C., and L. Mochan, to be published (1997).
BIBLIOGRAPHY 317

MacRae, A.U., K. Miiller, J.J. Lander, J. Morrison, and J.C. Phillips, Phys.
Rev. Lett. 22, 1048 (1969).
Mermin, N.D., Phys. Rev. B 1, 2362 (1970).
Mills, D.L., Surf. Sci. 48, 59 (1975).
Nakayama, M., T. Kato, and K. Othomi, Solid State Commun. 50, 409
(1984).
Newns, D.M., Phys. Lett. A 39, 341 (1972).
Otto, A., Z. Phys. 216, 398 (1968).
Pehlke, E., and A. Liebsch, unpublished (1991).
Persson, B.N.J., Surf. Sci. 281, 153 (1993).
Persson, B.N.J., and E. Zaremba, Phys. Rev. B 30, 5669 (1984).
Petri, E., and A. Otto, Phys. Rev. Lett. 34, 1283 (1975).
Pines, D., Elementary Excitations in Solids (Benjamin, New York, 1964).
Pinsukanjana, P.R., E.G. Gwinn, J.F. Dobson, E.L. Yuh, and N.G. Asmar,
Phys. Rev. B 46, 7284 (1992).
Powell, C.J., and J.B. Swan, Phys. Rev. 115,869 (1959); 116,81 (1959).
Powell, C.J., and J.B. Swan, Phys. Rev. 118, 640 (1960).
Powell, C.J., Phys. Rev. 175, 972 (1968).
Puska, M.J., R.M. Nieminen, and M. Manninen, Phys. Rev. B 31, 3486
(1985).
Quang, A.A., and A. Eguiluz, Phys. Rev. Lett. 70, 3955 (1993).
Reiners, Th., C. Ellert, M. Schmidt, and H. Haberland, Phys. Rev. Lett. 74,
1558 (1995).
Ritchie, R.H., Phys. Rev. 106,874 (1957).
Ritchie, R.H., Prog. Theor. Phys. 29, 607 (1963).
Rocca, M., Surf. Sci. Rep. 22, 1 (1995).
Rocca, M., and U. Valbusa, Phys. Rev. Lett. 64,2398 (1990).
Rocca, M., M. Lazzarina, and U. Valbusa, Phys. Rev. Lett. 67, 3197 (1991).
Rocca, M., M. Lazzarino, and U. Valbusa, Phys. Rev. Lett. 69, 2122 (1992).
Rocca, M., F. Moresco, and U. Valbusa, Phys. Rev. B 45, 1399 (1992).
318 BIBLIOGRAPHY

Rocca, M., S. Lizzit, B. Brena, G. Cautero, G. Comelli, and G. Paolucci, J.


Phys. CM 7, L611 (1995).
Schaich, W.L., SUTf. Sci. 318, L1157 (1994).
Schaich, W.L., Phys. Rev. B 55, 9379 (1997).
Schaich, W.L., and J.F. Dobson, Phys. Rev. B 49, 14700 (1994).
Schaich, W.L., and K. Kempa, Phys. SCT. 35,204 (1987).
SchUlke, W., N. Nagasawa, and S. Mourikis, Phys. Rev. Lett. 33, 6744
(1986).
Schwartz, C., and W.L. Schaich, Phys. Rev. B 26, 7008 (1982).
Schwartz, C., and W.L. Schaich, J. Phys. C 17,537 (1984).
Serra, Ll., G.B. Bachelet, N.V. Giai, and E. Lipparini, Phys. Rev. B 48,
14708 (1993).
Serra, Ll., and A. Rubio, Phys. Rev. Lett. 78, 1428 (1997).
Sipe, J.E., Phys. Rev. B 22, 1589 (1980).
Snider, D.R., and R.S. Sorbello, Phys. Rev. A 28, 5702 (1983).
Sprunger, P.T., G.M. Watson, and E.W. Plummer, SUTf. Sci. 269/270,551
(1992).
Stern, E.A., and R.A. Ferrel, Phys. Rev. 120, 130 (1960).
Sturm, K., Advances in Physics 31, 1 (1982).
Sturm, K., and L.E. Oliveira, Phys. Rev. B 40, 3672 (1989).
Sturm, K., E. Zaremba, and K. Nuroh, Phys. Rev. B 42, 6973 (1990).
Suto, S., K.D. Tsuei, E.W. Plummer, and E. Burstein, Phys. Rev. Lett. 63,
2590 (1989).
Tarriba, J., and W.L. Mochan, Phys. Rev. B 46, RC 12902 (1992).
Taut, M., J. Phys. CM 4, 9595 (1992).
Tiggesbaumker, J., L. Koller, K.H. Meiwes-Broer, and A. Liebsch, Phys.
Rev. A 48,1749 (1993).
Tsuei, K.D., Ph.D. thesis, University of Pennsylvania (1990).
Tsuei, K.D., E.W. Plummer, and P.J. Feibelman, Phys. Rev. Lett. 63,2256
(1989).
BIBLIOGRAPHY 319

Tsuei, K.D., E.W. Plummer, A. Liebsch, K. Kempa, and P. Bakshi, Phys.


Rev. Lett. 64,44 (1990).
Tsuei, K.D., E.W. Plummer, A. Liebsch, E. Pehlke, K. Kempa, and P. Bakshi,
Surf. Sci. 241, 302 (1991).
Yom Felde, A., J. Sprosser-Prou, and J. Fink, Phys. Rev. B 40, 10181
(1989).
Yuh, E.L., and E.G. Gwinn, Surf. Sci. 305, 202 (1994).
Zacharias, P., and K.L. Kliewer, Solid State Commun. 18, 23 (1976).
Zaremba, E., and W. Kohn, Phys. Rev. B 13, 2270 (1976).
Zaremba, E., and B.N.J. Persson, Phys. Rev. B 35, 596 (1987).
Zaremba, E., and H.C. Tso, Phys. Rev. B 49, 8147 (1994).
Zhang, Z.Y., D.C. Langreth, and J.P. Perdew, Phys. Rev. B 41, 5674 (1990).

Chapter 4

Abeles, F., and T. L6pez-Rfos, Surf. Sci. 96, 32 (1980).


Anderegg, M., B. Feuerbacher, and B. Fitton, Phys. Rev. Lett. 21, 1565
(1971).
Apell, P., Phys. Scr. 24, 795 (1981).
Apell, P., and C. Holmberg, Solid State Commun. 49,693 (1984).
Bachelet, G.B., D.R. Hamann, and M. Schluter, Phys. Rev. B 26, 4199
(1982).
Bagchi, A., Phys. Rev. B 15,3060 (1977).
Bagchi, A., RG. Barrera, and A.K. Rajagopal, Phys. Rev. B 20, 4824 (1979).
Barman, S.R, K. Horn, P. Haberle, H. Ishida, and A. Liebsch, to be published
(1997).
Bartynski, R.A., E. Jensen, T. Gustafsson, and E.W. Plummer, Phys. Rev.
B 32, 1921 (1985).
Bennett, A.J., Phys. Rev. B 1, 203 (1970).
Borensztein, Y., W.L. Mochan, J. Tarriba, RG. Barrera, and A. Tadjeddine,
Phys. Rev. Lett. 11, 2334 (1993).
Borensztein, Y., M. Roy, and R. Alameh, Europhys. Lett. 31, 5 (1995).
320 BIBLIOGRAPHY

Brodskii, A.M., and M.1. Urbakh, Sur/. Sci. 94, 369 (1980).
Burke, K., and W.L. Schaich, Phys. Rev. B 48, 14599 (1993).
Burke, K., and W.L. Schaich, Phys. Rev. B 49, 11397 (1994).
Del Sole, R., Solid State Commun. 31, 537 (1981).
Del Sole, R., in Electromagnetic Waves: Recent Developments in Research;
vol. 2: Photonic Probes of Surfaces, P. Halevi, ed. (Elsevier, Amsterdam,
1995), p. 13l.
Del Sole, R., and E. Fiorino, Phys. Rev. B 29, 4631 (1984).
Drube, W., F.J. Himpsel, and P.J. Feibelman, Phys. Rev. Lett. 60, 2070
(1988).
Drude, P., Ann. Phys. 43, 126 (1891).
Drude, P., The Theory of Optics (Dover, New York, 1959), p. 290.
Dzhavakhidze, P.G., A.A. Kornyshev, A. Tadjeddine, and M.1. Urbakh, Phys.
Rev. B 39, 13106 (1989).
Epstein, S.T., and R.E. Johnson, J. Chem. Phys. 51, 188 (1969).
Feibelman, P.J., Phys. Rev. B 12, 1319 (1975a).
Feibelman, P.J., Phys. Rev. Lett. 35,617 (1975b).
Feibelman, P.J., Prog. Surf. Sci. 12, 287 (1982).
Feibelman, P.J., Sur/. Sci. 282, 129 (1993).
Fernandez, V., D. Pahlke, N. Esser, K. Stahrenberg, O. Hunderi, A.M. Brad-
shaw, and W. Richter, to be published (1997).
FUitgen, G., K. Krischer, B. Pettinger, K. Doblhofer, H. Junkes, and G. Ertl,
Science 269, 668 (1995).
Flodstrom, S.A., and J.G. Endriz, Phys. Rev. Lett. 31, 893 (1973).
Flodstrom, S.A., and J.G. Endriz, Phys. Rev. B 12,1252 (1975).
Forstmann, F., Z. Phys. 203,495 (1967).
Forstmann, F., and R.R. Gerhardts, Metal Optics Near the Plasma Frequency,
Springer Tracts in Modern Physics 109 (Springer, Berlin, 1986).
Forstmann, F., and H. Stenschke, Phys. Rev. B 11, 1489 (1978).
Gerhardts, R.R., and K. Kempa, Phys. Rev. B 30, 5704 (1984).
BIBLIOGRAPHY 321

Gesell, T.F., E.T. Arakawa, M.W. Williams, and R.N. Hamm, Phys. Rev. B
7, 5141 (1973).
Gies, P., and R.R. Gerhardts, Europhys. Lett. 1, 513 (1986).
Gies, P., and R.R. Gerhardts, Phys. Rev. B 36, 4422 (1987).
Gies, P., R.R. Gerhardts, and T. Maniv, Phys. Rev. B 35, 458 (1987).
Gross, E.K.U., and W. Kohn, Phys. Rev. Lett. 55, 2850 (1985); Erratum:
ibid. 57, 923 (1986).
Hagemann, H.J., W. Gudat, and C. Kunz, J. Opt. Soc. Am. 65, 742 (1975).
Hanke, W., and L.J. Sham, Phys. Rev. B 21 4656 (1980).
Ishida, H., and A. Liebsch, Phys. Rev. B 42, 5505 (1990).
Ishida, H., and A. Liebsch, Phys. Rev. B 45, 6171 (1992).
Ishida, H., and A. Liebsch, to be published (1997).
Jackson, J.D., Classical Electrodynamics (Wiley, New York, 1962).
Jezequel, G., Phys. Rev. Lett. 45, 1963 (1980).
Johnson, P.B., and R.W. Christy, Phys. Rev. B 6, 4370 (1972).
Kempa, K., A. Liebsch, and W.L. Schaich, Phys. Rev. B 38, 12645 (1988).
Kempa, K., and W.L. Schaich, Phys. Rev. B 37, 6711 (1988).
Kempa, K., and W.L. Schaich, Phys. Rev. B 39, 13139 (1989).
Kim, B.O., E.W. Plummer, and A. Liebsch, to be published (1997).
Kolb, D.M., in Surface Polaritons, V.M. Agranovich and D. L. Mills, eds.
(North Holland, Amsterdam, 1982), p. 299.
Kotz, R., and D.M. Kolb, Surf. Sci. 97, 575 (1980).
Lang, N.D., and W. Kohn, Phys. Rev. B 1, 4555 (1970).
Lang, N.D., and W. Kohn, Phys. Rev. B 7, 3541 (1973).
Langreth, D.C., Phys. Rev. B 39, 10020 (1989).
Lee, J.T., and W.L. Schaich, Phys. Rev. B 43, 4629 (1991a).
Lee, J.T., and W.L. Schaich, Phys. Rev. B 44, 13010 (1991b).
Levinson, H.J., E.W. Plummer, and P.J. Feibelman, Phys. Rev. Lett. 43,
952 (1979).
322 BIBLIOGRAPHY

Levinson, H.J., and E.W. Plummer, Phys. Rev. B 24,628 (1981).


Liebsch, A., Phys. Rev. B 36, 7378 (1987).
Liebsch, A., G. Benemanskaya, and M. Lapushkin, Surf. Sci. 302, 303
(1994).
Liebsch, A., G. Hincelin, and T. L6pez~Rios, Phys. Rev. B 41, 10463 (1990).
Liebsch, A., and W.L. Schaich, Phys. Rev. B 52, 14219 (1995).
Lindau, I., and P.O. Nilsson, Phys. Scr. 3, 87 (1971).
L6pez~Rios, T., M. De Crescenzi, and Y. Borensztein, Solid State Gommun.
30, 755 (1979).
L6pez~Rios, T., in Electromagnetic Waves: Recent Developments in Re-
search; vol. 1: Spatial Dispersion in Solids and Plasmas, P. Halevi,
ed. (Elsevier, Amsterdam, 1992), p. 215.
Mackinson, R.E.B., Proc. Roy. Soc. A 162, 367 (1937).
Maniv, T., and H. Metiu, Phys. Rev. B 22, 4731 (1980).
Manninen, M., R. Nieminen, P. Hautojarvi, and J. Arponen, Phys. Rev. B
12, 4012 (1975).
McIntyre, J.D.E., and D.E. Aspnes, Surf. Sci. 24,417 (1971).
Melnyk, A.R., and M.J. Harrison, Phys. Rev. Lett. 21, 85 (1968).
Mochan, W.L., E. Fuchs, and R.G. Barrera, Phys. Rev. B 27, 771 (1983).
Mochan, W.L., and R.G. Barrera, Phys. Rev. Lett. 55, 1192 (1985); Phys.
Rev. B 32, 4984 (1985).
Monin, J., Acta Electron. 16, 139 (1973).
Monin, J., and S.G.A. Boutry, Phys. Rev. B 9, 1309 (1974).
Mukhopadhyay, G., and S. Lundqvist, Phys. Scr. 17,69 (1977).
Pehlke, E., and A. Liebsch, unpublished (1991).
Persson, B.N.J., and A. Apell, Phys. Rev. B 27, 6058 (1983).
Persson, B.N.J., and E. Zaremba, Phys. Rev. B 30, 5669 (1984).
Petersen, H., and S.B.M. Hagstrom, Phys. Rev. B 41, 1314 (1978).
Plieth, W.J., and K. Naegele, Surf. Sci. 64, 484 (1977).
Samuelsen, D., A. Yang, and W. Schattke, Surf. Sci. 281/288, 676 (1993).
BIBLIOGRAPHY 323

Sass, J.K., S. Stucki, and H.J. Lewerenz, Surf. Sci. 68, 429 (1977).
Sauter, F., Z. Phys. 203,488 (1967).
Schaich, W.L., Phys. Rev. B 50, 17587 (1994a).
Schaich, W.L., Surf. Sci. 318, L1157 (1994b).
Schaich, W.L., and W. Chen, Phys. Rev. B 39, 10714 (1989).
Schaich, W.L., and J.T. Lee, Phys. Rev. B 44, 5973 (1991).
Schaich, W.L., and C.M.J. Wijers, Phys. Rev. B 51, 10189 (1995).
Sipe, J.E., Phys. Rev. B 22, 1589 (1980).
Sorbello, R.S., Solid State Commun. 56,821 (1985).
Sprunger, P.T., G.M. Watson, and E.W. Plummer, Surf. Sci. 269/270,551
(1992).
Sturm, K., Advances in Physics 31, 1 (1982).
Sturm, K., and L.E. Oliveira, Phys. Rev. B 40, 3672 (1989).
Tadjeddine, A., D.M. Kolb, and R. Kotz, Surf. Sci. 101, 277 (1980).
Tarriba, J., and W.L. Mochan, Phys. Rev. B 46, RC 12902 (1992).
Wallden, L., Phys. Rev. Lett. 54, 943 (1985).
Wijers, C.M.J., and G.P.M. Poppe, Phys. Rev. B 46, 7605 (1992).
Zaremba, E., and W. Kohn, Phys. Rev. B 13, 2270 (1976).

Chapter 5

Adler, E., Phys. Rev. 134, A728 (1964).


Akulin, V.M., S. Goller, G.J.G. DaCosta, and F. Rebentrost, J. Phys. CM
4, 3857 (1992).
Argyres, P., Phys. Rev. 97, 334 (1955).
Bloembergen, N., R.K. Chang, S.S. Jha, and C.H. Lee, Phys. Rev. 174, 813
(1968).
Budd, H.F., and J. Vannimenus, Phys. Rev. B 12, 509 (1975).
Cheng, H., and P.B. Miller, Phys. Rev. 134, A683 (1964).
Chizmeshya, A., and E. Zaremba, Phys. Rev. B 37, 2805 (1988).
324 BIBLIOGRAPHY

Cini, M., R. Del Sale, and L. Reining, Surf. Sci. 281/288, 693 (1993).
Corvi, M., and W.L. Schaich, Phys. Rev. B 33, 3688 (1986).
Falicov, L.M., Phys. Today 45, 46 (1992).
Ghahramani, E., D.J. Moss, and J.E. Sipe, Phys. Rev. Lett. 642815 (1990);
Phys. Rev. B 43, 9700 (1991).
Guyot-Sionnest, P., W. Chen, and Y.R. Shen, Phys. Rev. B 33, 8254 (1986).
Guyot-Sionnest, P., A. Tadjedinne, and A. Liebsch, Phys. Rev. Lett. 64,
1678 (1990).
Hicks, J.M., L.E. Urbach, E.W. Plummer, and H.L. Dai, Phys. Rev. Lett.
61, 2588 (1988).
Hu, C.D., J. Phys. CM 8, 6629 (1996).
Hiibner, W., Phys. Rev. B 42, 11553 (1990).
Hiibner, W., and K.H. Bennemann, Phys. Rev. B 40, 5973 (1989).
Ishida, H., and A. Liebsch, Phys. Rev. B 42, 5505 (1990).
Ishida, H., and A. Liebsch, Phys. Rev. B 50, 4834 (1994).
Ishida, H., A.V. Petukhov, and A. Liebsch, Surf. Sci. 340, 1 (1995).
Janz, S., K. Pedersen, and H.M. van Driel, Phys. Rev. B 44, 3943 (1991a).
Janz, S., D.J. Bottomley, H.M. van Driel, and R.S. Timsit, Phys. Rev. Lett.
66, 1201 (1991b).
Janz, S., and H.M. van Driel, Int. J. Nonl. Opt. Phys. 2, 1 (1993).
Jha, S.S., and C.S. Warke, Phys. Rev. 153, 751 (1967).
Jiang, M.Y., G. Pajer, and E. Burstein, Surf. Sci. 242, 306 (1991).
Johnson, P.B., and R.W. Christy, Phys. Rev. B 6, 4370 (1972).
Kiejna, A., Surf. Sci. 331-333, 1167 (1995).
Kittel, C., Phys. Rev. 83, A208 (1951).
Koopmans, B., M.G. Koerkamp, T. Rasing, and H. van den Berg, Phys. Rev.
Lett. 14, 3692 (1995).
Krivoshchekov, G.V., and V.1. Stroganov, Sov. Phys. Solid State 11, 89
(1969); 11, 2151 (1970).
Kuchler, M., and F. Rebentrost, Phys. Rev. Lett. 11, 2662 (1993); Phys.
Rev. B 50, 5651 (1994).
BIBLIOGRAPHY 325

Langhoff, P.W., S.T. Epstein, and M. Karplus, Rev. Mod. Phys. 44, 602
(1972).
Liebsch, A., Phys. Rev. Lett. 43, 1431 (1979).
Liebsch, A., Phys. Rev. Lett. 61, 1233 (1988).
Liebsch, A., Phys. Rev. B 40, RC 3421 (1989).
Liebsch, A., and W.L. Schaich, Phys. Rev. B 40, 5401 (1989).
Liebsch, A., in Condensed Matter Physics Aspects of Electrochemistry, M.P.
Tosi and A.A. Kornyshev, eds. (World Scientific, Singapore, 1991), p.
274.
Liljenvall, H.G., and A.G. Mathewson, J. Phys. C Metal Phys. Suppl. 3,
5341 (1970).

Lindgren, s.A., and L. Wallden, Phys. Rev. B 45, 6345 (1992).


Maytorena, J.A., W.L. Mochan, and B.S. Mendoza, Phys. Rev. B 51, 2556
(1995).
Maytorena, J.A., B.S. Mendoza, and W.L. Mochan, to be published (1997).
Mendoza, B.S., and W.L. Mochan, Phys. Rev. B 53, 4999 (1996).
Mochan, W.L., and B.S. Mendoza, J. Phys. CM 5, A183 (1993).
Murphy, R., M. Yeganeh, K. J. Song, and E.W. Plummer, Phys. Rev. Lett.
63, 318 (1989).
Oppeneer, P.M., T. Maurer, J. Sticht, and J. Kubler, Phys. Rev. B 45,10924
(1992).
Pan, R.P., H.D. Wei, and Y.R. Shen, Phys. Rev. B 39, 1229 (1989).
Patterson, C.H., D. Weaire, and J.F. McGilp, J. Phys. CM 4, 4017 (1992).
Persson, B.N.J., and L.H. Dubois, Phys. Rev. B 39, 8220 (1989).
Petukhov, A.V., Surf. Sci. 347,143 (1996).
Petukhov, A.V., and A. Liebsch, Surf. Sci. 294, 381 (1993).
Petukhov, A.V., and A. Liebsch, Surf. Sci. 320, L51 (1994).
Petukhov, A.V., and A. Liebsch, Surf. Sci. 331/333, 1335 (1995a).
Petukhov, A.V., and A. Liebsch, Surf. Sci. 334, 195 (1995b).
Pustogowa, U., W. Hubner, and K.H. Bennemann, Phys. Rev. B 49, 10031
(1994).
326 BIBLIOGRAPHY

Quail, J.C., and H.J. Simon, Phys. Rev. B 31, 4900 (1985).
Reider, G.A., and T.F. Heiz, in Electromagnetic Waves: Recent Developments
in Research; vol. 2: Photonic Probes of Surfaces, P. Halevi, ed. (Elsevier,
Amsterdam, 1995), p. 413.
Reif, J., C. Rau, and E. Matthias, Phys. Rev. Lett. 71, 1931 (1993).
Reiff, S., W. Drachsel, and J.H. Block, Surf. Sci. 304, L420 (1994).
Reim, W., and J. Schoenes, in Ferromagnetic Materials, E.P. Wohlfahrt and
K.H.J. Buschow, eds. (North-Holland, Amsterdam, 1990), vol. 5, p.133.
Reining, L., R. Del Sole, M. Cini, and J.G. Ping, Phys. Rev. 50,8411 (1994).
Richmond, G.L., J.M. Robinson, and V.L. Shanon, Prog. Surf. Sci. 28, 1
(1988).
Rudnick, J., and E.A. Stern, Phys. Rev. B 4, 4272 (1971).
Schaich, W.L., Surf. Sci. 318, L1157 (1994).
Schaich, W.L., unpublished (1995).
Schaich, W.L., and A. Liebsch, Phys. Rev. B 37, 6187 (1988).
Schaich, W.L., and B.S. Mendoza, Phys. Rev. B 45, 14279 (1992).
Senatore, G., and K.R. Subbaswamy, Phys. Rev. A 35, 2440 (1987).
Shen, Y.R., The Principles of Nonlinear Optics (Wiley, New York, 1984).
Shen, Y.R., Nature 337, 519 (1989).
Shen, Y.R., Surf. Sci. 299-300, 551 (1994).
Sipe, J.E., V.C.Y. So, M. Fukui, and G.!. Stegeman, Phys. Rev. B 21, 4389
(1980).
Sipe, J.E., D.J. Moss, and H.M. van Driel, Phys. Rev. B 35, 1129 (1987).
Smith, N.V., Phys. Rev. B 2, 2840 (1970).
Song, K.J., D. Heskett, H.L. Dai, A. Liebsch, and E.W. Plummer, Phys. Rev.
Lett. 61, 1380 (1988).
Straub, M., R. Vollmer, and J. Kirschner, Phys. Rev. Lett. 77, 743 (1996).
Sturm, K., and L.E. Oliveira, Phys. Rev. B 30, 4352 (1984).
Tarriba, J., and W.L. Mochan, Phys. Rev. B 46, RC 12902 (1992).
Tom, H.W.K., T.F. Heinz, and Y.R. Shen, Phys. Rev. Lett. 51, 1983 (1983).
BIBLIOGRAPHY 327

Tom, H.W.K., C.M. Mate, X.D. Zhu, J.E. Crowell, T.F. Heinz, G.A. Somorjai
and Y.R. Shen, Su'f. Sci. 172,466 (1986).
Urbach, L.E., K.L. Percival, J. Hicks, E.W. Plummer, and H.L. Dai, Phys.
Rev. B 45, 3769 (1992).
Wang, C.S., J.M. Chen, and J.R. Bower, Opt. Commun. 8,275 (1973).
Wang, J., Z.C. Ying, and E.W. Plummer, unpublished (1994).
Wang, J., Z.C. Ying, and E.W. Plummer, Phys. Rev. B 51, 5590 (1995).
Weber, M., and A. Liebsch, Phys. Rev. B 35, 7411 (1987a).
Weber, M., and A. Liebsch, Phys. Rev. B 36, 6411 (1987b).
Weinert, M., private communication (1994).
Westin, E., and A. Rosen, Su'f. Sci. 269/270,77 (1991).
Wierenga, H.A., W. de Jong, M.W.J. Prins, T. Rasing, R. Vollmer, A. Kiri-
lyuk, H. Schwabe, and J. Kirschner, Phys. Rev. Lett. 74,1462 (1995).
Wijers, C.M.J., Th. Rasing, and R.W.J. Hollering, Solid State Commun. 85,
233 (1993).
Wong, E.K.L., and G.L. Richmond, 1. Chern. Phys. 99, 5500 (1993).
Ying, Z.C., J. Wang, G. Andronica, J.Q. Yao, and E.W. Plummer, 1. Vac.
Sci. Technol. A 11, 2255 (1993).
Zangwill, A., 1. Chern. Phys. 78, 5926 (1983).

Chapter 6

Andersson, S., and M. Persson, Phys. Rev. B 48, 5685 (1993).


Andersson, S., M. Persson, and J. Harris, Su'f. Sci. 360, L499 (1996).
Casimir, H.B.G., and D. Polder, Phys. Rev. 73, 360 (1948).
Chen, W., and W.L. Schaich, Su'f. Sci. 218, 580 (1989).
Chizmeshya, A., and E. Zaremba, Su.f. Sci. 220, 443 (1989).
Chizmeshya, A., and E. Zaremba, Su'f. Sci. 268, 432 (1992).
Engel, T., 1. Chern. Phys. 69, 373 (1978).
Hutson, J.M., P.W. Fowler, and E. Zaremba, Su'f. Sci. 175, L775 (1986).
Kleiman, G.G., and U. Landman, Phys. Rev. B 8, 5484 (1973).
328 BIBLIOGRAPHY

Landragin, A., J.Y. Courtois, G. Labeyrie, N. Vansteenkiste, C.1. Westbrook


and A. Aspect, Phys. Rev. Lett. 77, 1464 (1996).
Lang, N.D., Phys. Rev. Lett. 46, 842 (1981).
Lang, N.D., and W. Kohn, Phys. Rev. B 7, 3541 (1973).
Liebsch, A., Phys. Rev. B 33, 7249 (1986a).
Liebsch, A., J. Phys. C 19, 5025 (1986b).
Liebsch, A., Phys. Rev. B 35, 9030 (1987).
Lifshitz, E.M., Sov. Phys. JETP 2, 73 (1956).
Marvin, A.M., and F. Toigo, Phys. Rev. A 25, 803 (1982).
Mehl, M.J., and W.L. Schaich, Surf. Sci. 99, 553 (1980).
Persson, B.N.J., and E. Zaremba, Phys. Rev. B 30, 5669 (1984).
Poelsema, B., and G. Comsa, Scattering of Thermal Energy Atoms, Springer
Tracts in Modern Physics 115 (Springer, Berlin, 1989).
Rieder, K.H., Contemp. Phys. 26,559 (1985).
Toennies, J.P., J. Vac. Sci. Technol. A 2, 1055 (1984).
Zangwill, A., and P. Soven, Phys. Rev. A 21, 1561 (1980).
Zaremba, E., and W. Kohn, Phys. Rev. B 13, 2270 (1976).
Zaremba, E., and W. Kohn, Phys. Rev. B 15, 1769 (1977).

Chapter 7
Andersson, S., and B.N.J. Persson, Phys. Rev. Lett. 50, 2028 (1983).
Chabal, Y.J., Surf. Sci. Rep. 8, 211 (1988).
Chance, R.R., A. Prock, and R. Silbey, Adv. Chern. Phys. 37, 1 (1978).
D'Agliano, E.G., O. Kumar, W.L. Schaich, and H. Suhl, Phys. Rev. B 11,
2122 (1975).
Daly, C., and J. Krim, Phys. Rev. Lett. 76, 803 (1996).
Daly, C., and J. Krim, in Micro/Nanotribology and Its Applications, B. Bhu-
shan, ed. (Kluwer, Dordrecht, 1997), p. 31l.
Eguiluz, A.G., Phys. Rev. B 30, 4366 (1984).
Eguiluz, A.G., Phys. Scr. 36,651 (1987).
BIBLIOGRAPHY 329

Ferrel, T.L., P.M. Echenique, and R.H. Ritchie, Solid State Commun. 32,
419 (1979).
Gerlach, E., Phys. Stat. Sol. (b) 121, 757 (1984).
Gerlach, E., J. Phys. C 19,4585 (1986).
Harris, J., and R.O. Jones, J. Phys. C 6, 3585 (1973).
Harris, J., and R.O. Jones, J. Phys. C 7, 3751 (1974).
Hirschmugl, C.J., G.P. Williams, F.M. Hoffmann, and Y.J. Chabal, Phys.
Rev. Lett. 65, 480 (1990).
Holzapfel, C., W. Akemann, and D. Schumacher, Surf. Sci. 227, 123 (1990).
Ishida, H., Phys. Rev. B 49, 14610 (1994).
Ishida, H., Phys. Rev. B 52, 10819 (1995).
Krim, J., D.H. Solina, and R. Chiarello, Phys. Rev. Lett. 66, 181 (1991).
Lang, N.D., and W. Kohn, Phys. Rev. B 7, 3541 (1973).
Langreth, D.C., Phys. Rev. Lett. 54, 126 (1985).
Liebsch, A., Phys. Rev. Lett. 54, 67 (1985).
Liebsch, A., Phys. Rev. B 36, 7378 (1987).
Liebsch, A., Phys. Rev. B 55, 13263 (1997).
Lodder, A., J. Phys. F 14,2943 (1984).
Mahan, G.D., in Collective Properties of Physical Systems, S. Lundqvist, ed.
(Nobel symposion 24, 1973), p. 164.
Persson, B.N.J., J. Phys. C 11,4251 (1978).
Persson, B.N.J., Phys. Rev. B 44, 3277 (1991).
Persson, B.N.J., and M. Persson, Solid State Commun. 36,175 (1980).
Persson, B.N.J., and W.L. Schaich, J. Phys. C 14,5583 (1981).
Persson, B.N.J., and E. Zaremba, Phys. Rev. B 31, 1863 (1985).
Persson, B.N.J., and A. Nitzan, Surf. Sci. 367,261 (1996).
Persson, B.N.J., and A.1. Volokitin, J. Chem. Phys. 103, 8679 (1995).
Ryberg, R., Surf. Sci. 114,627 (1982).
Schaich, W.L., Solid State Commun. 15, 357 (1974).
330 BIBLIOGRAPHY

Schaich, W.L., Phys. Rev. B 13, 3350 (1976).


Schaich, W.L., and J. Harris, J. Phys. C 11, 65 (1981).
Schumacher, D., Surface Scattering Experiments with Conduction Electrons,
Springer Tracts in Modern Physics 128 (Springer, Berlin, 1993).
Sokoloff, J.B., Phys. Rev. B 42, 760 (1990).
Sokoloff, J.B., Phys. Rev. B 52, 5318 (1995).
Sols, F., and F. Flores, Solid State Commun. 42,687 (1982).
Sols, F., F. Flores, and N. Garcia, Surf. Sci. 137, 167 (1984).
Sorbello, R.S., J. Chem. Phys. Solids 42, 309 (1981).
Volokitin, A.I., and B.N.J. Persson, in Inelastic Energy Transfer in Inter-
actions with Surfaces and Adsorbates, B. Gumhalter, A.C. Levi, and F.
Flores, eds. (World Scientific, Singapore, 1993), p. 217.
Index

adiabatic limit, 18, 29, 35, 165, 175, charged surface, 18, 21,87,106,142,
231, 234, 252, 265 172, 198, 233, 246
adsorbate vibration, 25, 294, 297, chemical interface damping, 128, 140
300-306 Cl on Ag, 108
Ag, 22, 71, 92, 107, lll, 188, 219, classical matching, 74, 150
227, 244, 248 classical optics, see local optics
Ag(lll),260 Clausius-Mossotti relation, 102
overlayer, 126, 210 CO on Cu, 296, 306
particle, 138 collective excitations, 34
Al, 22, 51, 58, 61, 63, 68, 71, 85, compressibility sum rule, 35
181-183,219,240 computational procedure, 37-48, 169,
Al(ll1), 256 231
vicinal, 263, 308 conductivity tensor, 43, 149, 268,
alkali metals, 53, 242 306
overlayers, ll7, 250 constitutive relation, 75, 149
realistic, 200, 253 continuity equation, 148
particles, 135 core polarization, 67, 102, 104
anisotropy, 147 correlation energy
Ar overlayer, III Ceperley-Alder, 8
asymptotic region, 45, 75, 150, 170, Wigner, 8
231 Coulomb potential
atoms, 27 asymptotic, 167, 189
attenuated total reflection, 84, 194 long-range, 40, 167, 230
short-range, 40, 168, 230
avoided crossing, 84, ll6, ll9
cross section, 27
crystal face dependence, 15, 22, 187,
Be, 181 196, 242, 262
bulk dielectric function, 32, 53, 70, Cs, 61, 63, 67, 181
105, 137, 155, 161, 181, 194 overlayer, 129, 205
bulk excitations, 10, 71, 92, 105 Cu,283
bulk plasmon, 51, 58, 67, 70, 154,
157,177, 178 d-parameter, 77, 79, 86, 103, 132,
bulk polarization, 157 135,150,155-164,171,174,

331
332 INDEX

190,202,221,246,275,280, equation of motion, 172, 191, 215


307 evanescent states, 260, 262, 267
crystal face dependence, 102 exchange-correlation
non-diagonal, 159 energy, 6
density contour, 200, 254, 265 hole, 9
density functional theory, 6 potential, 8, 12, 27, 34
density profile, 11, 39, 130 nonlocal, 22
linear, 53, 116, 176, 236, 238 external potential, 26, 167, 188, 276
quadratic, 236, 238
spill-out, 11,67,105,171,211 Fe, 268
Fermi wave vector, 8
dielectric medium, 95, 104, 109, 188
finite-barrier potential, 11, 87, 235
dipole barrier, 13-16
finite-size effects, 96, 130, 210
dipole lattice, 95, 102, 160, 187, 262,
Fresnel fields, 75, 221
280
Fresnel reflectivity, 294
dipole moment, 20, 172, 229, 232
Fresnel theory, 146
dipole scattering, 55, 71, 283
friction force, 298-304
displacement field, 148
friction integral
drift velocity, 304
atomic, 301
driven modes, 61
ionic, 299
dynamical image plane, 79, 273, 274
Friedel oscillations, 11, 19, 23, 29,
dynamical image potential, 298
46, 117, 131, 172, 178, 186,
dynamical processes, 34
191, 231, 239, 258, 265, 276
EELS, see electron energy loss spec- frustrated parallel translation, 303
troscopy gap in excitation spectrum, 84, 116,
effective potential, 6, 12, 38 119
eigenmodes, 61 Gauss's law, 50, 96, 217, 220
electron energy loss spectroscopy, 285, golden rule, 26, 29, 57, 164, 284,
294, 297 288, 290, 295
electron radius r., 9 Green's functions, 40
electron-electron interactions, 27 ground-state properties, 6-25
electron-hole continuum, 287
electron-hole pairs, 14, 59, 130, 141, harmonic potential theorem, 36
283 heat diffusion model, 246
electrostatic limit, 74, 227 Hg, 71, 104
ellipsometry, 159, 161 particle, 140
embedding scheme, 22, 46-48, 264, high-frequency limit, 35
268, 306 homogeneous electron gas, 7
energy dissipation, 307 hydrodynamic theory, 53, 55, 87, 116,
energy loss concept, 305 148,210,218,227,236,238,
INDEX 333

239, 243 lattice potential, 15, 22, 102, 216,


242
image charge, 9 LDA, see local density approxima-
image potential, 9, 13, 23, 61, 69 tion
impact scattering, 55, 71 LDA-based RPA, see random-phase
In, 58, 85, 181 approximation
induced current density, 148, 306 length scales, 29
induced density, 18, 22, 44, 50, 61, Li, 67, 71, 90, 185
131,148,177,202,228,276, overlayer, 206
292, 306 particle, 137
centroid, see d-parameter linear Kerr angle, 268
unscreened, 44, 170, 229, 231 local density approximation, 6-8, 69
induced potential, 26, 228 local fields, 102, 160, 178, 262, 280
asymptotic, 32 local optics, 50, 114, 130, 146, 199,
electrostatic, 27, 167 205, 210, 305
exchange-correlation, 27, 34,167 local potential, 26, 165, 183, 228,
inelastic electron scattering, 31, 44, 288
55, 283 long-wavelength limit, 73-82, 150,
infinite-barrier potential, 11, 235 158, 221, 222, 227
infrared reflection-absorption spec- low-frequency excitations, 11, 25, 286,
troscopy, 294, 297 290
integration of surface fields, 74, 150
interaction Hamiltonian, 150, 214 magnetic surfaces, 268
interband transitions, 10,22,66,67, magneto-optical Kerr effect, 268
90, 137, 185, 226, 242, 256, many-body response function, 26, 34,
265 57, 272, 295, 298
inverSIon symmetry, 214, 216, 269 matrix equation, 33, 45, 170, 231
broken, 255, 263, 267 Maxwell equations, 74, 148, 220
lRAS, see infrared reflection-absorp- mean-field treatment, 26
tion spectroscopy mechanical stability, 15
metal particles, 27, 130-141
jellium model, 9-15, 80 polarizability, 130, 132
overlayer, 116, 252 metal-electrolyte interface, 22, 106,
109, 195, 198, 246
K, 58, 63, 71, 181, 288 Mg, 51, 53, 58, 63, 70, 71, 90, 111,
on Cu, 300 181
overlayer, 118-124,206 MgO overlayer, 111
particle, 135 microscopic fields, 76
Kramers-Kronig relation, 35, 82,171, Mie plasmon, 28, 111, 130
252, 274 blue shift, 138
334 INDEX

lifetime, 130, 132, 135, 140, 141 oxide layer, 51, 111
red shift, 135, 138
self-energy, 132 Pauli repulsion, 271
size dependence, 135 Pd,105
model potentials, 11, 86, 116, 235 penetration depth, 239, 256, 262,
molecules, 27 263,270
phonon scattering, 284
Na, 61, 63 photoabsorption, 179
overlayer, 117, 121, 181, 200, 206 external, 163
particle, 132, 135 internal, 163
Ni,269 photoemission, 176, 179, 206
noble metals, 93, 186, 262, 280 inverse, 183
non-symmetric stacking, 256, 258, photoyield, 162, 206
267, 268 physisorption potential, 271, 280
nonlinear Poisson equation, 12, 27, 32, 40, 97,
a-parameter, 218, 230,233, 237, 168, 188
248,252 polarizability tensor, 149
b-parameter, 218 pseudopotential, 9, 15, 68, 91, 137,
d-parameter, 218 184, 200, 204, 235
I-parameter, 226 quantum wells, 141
amplitude, 226, 256 quartz crystal microbalance, 300
anisotropy, 226, 255-257, 260, quasi-static approximation, 284, 289
262, 263, 265
bulk polarization, 214, 256, 260 random-phase approximation, 29, 35,
dipole moment, 20, 254 63, 79, 100, 165, 174, 176,
dipole polarizability, 214 234, 288, 296
induced density, 239, 265 Rb,181
Kerr angle, 270 realistic metals, 22, 24, 46, 159, 293
quadrupole polarizability, 214 reflectance anisotropy, 196
radiation, 219-224 reflection coefficient, 32, 85, 146, 152,
response function, 43, 228 154, 159, 162
surface polarization, 20, 216, 229, reflectivity, 160, 198, 205, 305
236 response
nonlocal dielectric function, 75, 156 long-wavelength, 18, 287
nonlocal response, 26, 85, 130 short-wavelength, 22, 24, 290
Numerov method, 38, 41 response equation, 26, 98, 110, 165,
188, 228, 231, 276
occupied d bands, 92, 186, 280 solution, 44, 165, 230
ohmic heating, 304 response functions, 42-44
overlayer plasmon, 113, 210 retardation, 83, 282
INDEX 335

RPA, see random-phase approxima- surface energy, 8, 15-16


tion surface excitation spectra, 31, 58,
98, 118, 126, 174,237
s-d hybridization, 93, 244 surface field integral, 76, 152, 155
s-d polarization model, 94, 126, 138, surface loss function, 32, 46
188, 211, 246, 281 surface phonons, 25
scattering probability, 57, 283 surface plasmon, 51
screening length, 11, 29 anisotropy, 93, 102
second harmonic generation, 224, 240, crystal face dependence, 93, 102
243, 246, 248, 251 dispersion, 53, 63
self-consistency, 29, 40, 170 linear, 77, 171, 176
self-consistent field approach, 26, 149 negative, 65, 73, 79, 105
semi-infinite uniform background, 9 positive, 65, 93, 111, 212
semiconductors, 141, 159, 262 lifetime, 67, 70, 132
SHG, see second harmonic genera- monopole, 55, 61
tion multipole, 53, 55, 58, 61, 81,
Si on jellium, 308 116,119,132,135,142,171,
simple metals, 58, 174, 234 176, 182,236,238,242,252
particles, 130 polariton, 83, 161, 194
single-particle response function, 26, self-energy, 131
34, 57 surface polarization, 148, 157, 193
sliding friction, 25 surface resistivity, 25, 303, 306
atoms, 300 surface response function, 31, 57, 78,
ions, 298 96, 162, 278, 283, 290
Smoluchowski smoothing, 265 surface screening, 33, 158
stabilized jellium model, 15-16, 68, surface-bulk interference, 285, 289
80, 184, 235, 242, 277
static electric field, 18, 27, 247 TDLDA, see time-dependent local
static image plane, 9, 19-24, 170, density approximation
175,292 thin films, 27, 37, 116
static response properties, 18-25 Thomas-Fermi-von Weizsacker ap-
stepped metal surface, 263 proach, 107, 235
sum frequency generation, 238 three-layer model, 147
sum rule, 171, 232 threshold excitation, 61, 124, 176,
dynamical force, 168, 172, 191, 204, 237, 242, 248, 265
198, 232, 276 time-dependent Hartree approxima-
violation, 29, 174, 234 tion,29
surface f-sum rule, 171, 274 time-dependent local density approx-
surface barrier height, 13-15 imation, 27, 63, 79, 90, 98,
surface conditions, 82, 195, 259 107,111,130,165,171,174,
336 INDEX

228,240,243,244,248,250,
252,273,276,288,296,302,
306
adiabatic, 26
dynamical, 34, 69, 185
time-dependent response properties,
25-37
transition rate, 26, 57, 284
two-band model, 257, 260, 267

unoccupied states, 10, 89, 117, 129,


176

van der Waals


attraction, 186, 271-282, 300
constant, 272, 282
reference position, 274, 276
virtual excitations, 93, 188, 261

wave equation, 83, 149, 220, 221


wind force, 305
work function, 8, 13-16, 61, 117,
124, 176, 237, 279

Xe,27
on Ag, 302

zone boundary collective states, 92,


185

You might also like