Sie sind auf Seite 1von 200

Phys 48W

Physics and Chemistry at Surfaces

An electronic course o↵ered at the Physics Department of the


Boğaziçi University to 4th-year and Masters-Degree students

Mehmet Erbudak

Physics Department, Boğaziçi University


and
Laboratorium für Festkörperphysik, ETHZ, CH-8093 Zurich

mehmet.erbudak@boun.edu.tr
erbudak@phys.ethz.ch

September 2012
Phys 48W
Physics and Chemistry at Surfaces

2012/2013 Fall Semester


An electronic course o↵ered for the Bachelor- and Masters-Degree students

Mehmet Erbudak
Physics Department, Boğaziçi University
mehmet.erbudak@boun.edu.tr
and
Laboratorium für Festkörperphysik, ETHZ, CH-8093 Zurich
erbudak@phys.ethz.ch

Several di↵erent phenomena are observed at surfaces that do not have a counterpart
in bulk materials. Corrosion, epitaxial growth, heterogenous catalysis, or tribology
are just few of these. While all bulk processes can be accounted for on equal footing
owing to the universal description of electronic states, at surfaces symmetry is
broken, and we need to redefine the electronic and crystal structure. In this course,
we deal with the electronic structure of the bulk and surfaces, study the geometric
structure and get acquainted with appropriate experimental tools to observe surface-
specific processes. A chemical analysis is part of the complete characterization
of surfaces. We realize that the atomic structure, the electronic properties and
chemistry of surfaces are all interrelated.
Every week, students obtain the script for the week, exercises, and a short video
clip. I will be present for a few lectures personally including the exam at the end
of the semester.

Prerequisite: Modern Physics.

i
Preamble

I will place the learning material as well as exercises to your disposal in internet
in the pdf format every week. The learning material is planned to occupy your
attention during about 3 hours per week to justify the 3 credit hours. I will men-
tion to you some books as supporting material if needed, and will present relevant
publications. With some basic knowledge on Quantum Mechanics and Solid State
Physics, I assume you will appreciate the presented material as an introduction
to several directions of Surface Physics and Chemistry as well as modern Materi-
als Science. Similarly, the concepts you will be introduced correspond to those of
low-dimensional phenomena. Thus, this course is thought to be as an introduction
to your future research in many fields. I will mostly emphasize the experimental
achievements. For any question please do consult me per mail. During the semester,
you may use my Skype address erbudak if you wish a personal contact.
The presented material may be too extensive. My intention is to trigger your
interest on this subject. Interest and curiosity are required for innovative research
and progress. In the following you will find a comprehensive introduction followed
by a chapter on the electronic structure of the bulk and the surface as well as a
chapter on photoemission. The next chapter is on the atomic geometry, likewise
of the bulk and the surface. Following chapters are devoted on the chemical com-
position and adsorption of foreign atom on the surface as well as their behavior. I
appreciate any suggestions or corrections on this material.
As additional reading material, I suggest you follow some professional journals,
such as Surface Science, Physical Review Letters, Science, or Nature. I recommend
to you few books like:
AZ - A. Zangwill, Physics at Surfaces, CUP
D.P. Woodru↵ and T.A. Decker, Modern Techniques of Surface Science, CUP
G. Ertl and J. Küppers, Low-Energy Electrons and Surface Chemistry, Verlag
Chemie, Weinheim
JSB - J.S. Blakemore, Solid State Physics, W.B. Sounders Co., Philadelphia
CK - C. Kittel, Introduction to Solid State Physics, John Wiley & Sons, New York
AM - N.W. Ashcroft and N.D. Mermin, Solid State Physics, Saunders College Pub-
lishing, Fort Worth

I hope you will find the course useful and enjoy it!

ii
Contents

Phys 48W
Physics and Chemistry at Surfaces i

Preamble ii

1 Introduction: Why surfaces? 1


1.1 Surface Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Surface-Induced Chemical Reactions . . . . . . . . . . . . . . . . . 4
1.3 Epitaxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Surface Melting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Carbon-Based Structures . . . . . . . . . . . . . . . . . . . . . . . . 10
1.6 Scattering Cross Section and Mean Free Path . . . . . . . . . . . . 12
1.7 Vacuum Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2 Electronic Structure 16
2.1 Free Electrons in Metals . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.1 Free Electrons in 3DIM . . . . . . . . . . . . . . . . . . . . . 17
2.1.2 2DIM Electron Gas . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.3 Fermi-Dirac Distribution Law . . . . . . . . . . . . . . . . . 20
2.2 Band Theory of Metals . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.1 Periodic Lattice . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.2 Motion of Electrons in a Periodic Potential . . . . . . . . . . 23
2.2.3 Band Gaps . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2.4 Bloch Functions . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.5 Reduced-Zone Scheme . . . . . . . . . . . . . . . . . . . . . 26
2.3 Tight-Binding Approximation . . . . . . . . . . . . . . . . . . . . . 28
2.4 Surface Electronic Structure . . . . . . . . . . . . . . . . . . . . . . 29

3 Photoelectric Emission 36
3.1 Photoemission Process . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.1.1 Optical Excitation: Conservation rules . . . . . . . . . . . . 36
3.1.2 Energy Conservation . . . . . . . . . . . . . . . . . . . . . . 38
3.1.3 Conservation of Momentum . . . . . . . . . . . . . . . . . . 39
3.1.4 Three-Step Model . . . . . . . . . . . . . . . . . . . . . . . . 40

iii
CONTENTS iv

3.2 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2.1 Band Structure . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2.2 Surface States . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2.3 Spin Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.4 Localized States . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2.5 Resonant Photoemission . . . . . . . . . . . . . . . . . . . . 47

4 Crystal Structure 49
4.1 Chemical Bonding . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.1.1 The van-der-Waals Bond . . . . . . . . . . . . . . . . . . . . 51
4.1.2 The Covalent Bond . . . . . . . . . . . . . . . . . . . . . . . 53
4.1.3 Covalent–van-der-Waals Structures . . . . . . . . . . . . . . 54
4.1.4 The Ionic Bond . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.1.5 The Hydrogen Bond . . . . . . . . . . . . . . . . . . . . . . 56
4.1.6 The Metallic Bond . . . . . . . . . . . . . . . . . . . . . . . 56
4.2 Symmetry Operations . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.2.1 Bulk Structure . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3 Determination of Bulk Structure . . . . . . . . . . . . . . . . . . . . 61
4.3.1 X-Ray Di↵raction . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3.2 X-Ray Absorption Fine Structures . . . . . . . . . . . . . . 66

5 Surface Structure 74
5.1 2DIM Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.1.1 Ideal Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.1.2 Deviation from the Ideal Case . . . . . . . . . . . . . . . . . 76
5.1.3 Classification of Reconstructions . . . . . . . . . . . . . . . . 77
5.2 Determination of Surface Structure - Real-Space Techniques . . . . 78
5.2.1 Confocal Microscopy . . . . . . . . . . . . . . . . . . . . . . 79
5.2.2 Photoelectron Spectromicroscopy . . . . . . . . . . . . . . . 80
5.2.3 Scanning Electron Microscopy . . . . . . . . . . . . . . . . . 83
5.2.4 Field Emission Microscopy . . . . . . . . . . . . . . . . . . . 84
5.2.5 Field Ion Microscopy . . . . . . . . . . . . . . . . . . . . . . 88
5.2.6 Secondary-Electron Imaging . . . . . . . . . . . . . . . . . . 92
5.2.7 X-Ray Photoelectron Di↵raction . . . . . . . . . . . . . . . . 97
5.2.8 Scanning Microscopes . . . . . . . . . . . . . . . . . . . . . 98
5.2.8.1 Scanning Tunnelling Microscope . . . . . . . . . . 100
5.2.8.2 Atomic Force Microscopy . . . . . . . . . . . . . . 103
5.3 Determination of Surface Structure -
Reciprocal-Space Techniques . . . . . . . . . . . . . . . . . . . . . . 107
5.3.1 Low-Energy Electron Di↵raction . . . . . . . . . . . . . . . . 108
5.3.1.1 Intensity of Di↵racted Beams . . . . . . . . . . . . 115
5.3.1.2 Deviation from the Ideal Case . . . . . . . . . . . . 117
5.3.1.3 Calculation of Di↵racted Intensities . . . . . . . . . 132
CONTENTS v

6 Inner Shells 135


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.1.1 Filling the shells with electrons . . . . . . . . . . . . . . . . 135
6.1.2 Occupation of the shells . . . . . . . . . . . . . . . . . . . . 136
6.1.3 Shell Binding Energy . . . . . . . . . . . . . . . . . . . . . . 137
6.2 Ionization of Inner Shells . . . . . . . . . . . . . . . . . . . . . . . . 138
6.3 X-ray Absorption Spectroscopy . . . . . . . . . . . . . . . . . . . . 139
6.4 Excitation of an Atom by Electrons . . . . . . . . . . . . . . . . . . 141
6.4.1 Electron Energy-Loss Spectroscopy . . . . . . . . . . . . . . 145
6.5 X-Ray Photoelectron Spectroscopy . . . . . . . . . . . . . . . . . . 148
6.6 Relaxation Processes . . . . . . . . . . . . . . . . . . . . . . . . . . 153
6.6.1 Dipole Radiation . . . . . . . . . . . . . . . . . . . . . . . . 153
6.6.2 Auger Electron Emission . . . . . . . . . . . . . . . . . . . . 156

7 Vibrational Spectroscopies 165


7.1 Infrared Absorption Spectroscopy . . . . . . . . . . . . . . . . . . . 167
7.2 Inelastic Electron Scattering . . . . . . . . . . . . . . . . . . . . . . 171
7.3 Raman Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
7.3.1 A Classical Consideration for the Raman Process . . . . . . 174
7.3.1.1 An Isolated Molecule . . . . . . . . . . . . . . . . . 174
7.3.1.2 Selection Rules . . . . . . . . . . . . . . . . . . . . 175
7.3.1.3 An Example: CO2 Molecule . . . . . . . . . . . . . 176
7.3.1.4 Extended Systems . . . . . . . . . . . . . . . . . . 177
7.3.1.5 Raman Tensor . . . . . . . . . . . . . . . . . . . . 178
7.3.2 Quantum Mechanical Features of the Raman Process . . . . 178
7.3.3 Variants of the Raman Process . . . . . . . . . . . . . . . . 180

8 Surface Reactions 183


8.1 Adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
8.2 Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
8.3 Desorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

Index 193
Chapter 1

Introduction: Why surfaces?

The majority of processes, that played a crucial role in the development of our tech-
nological society, is based on physical and chemical properties of surfaces. Catalyt-
ical reactions and semiconductor (SC) structures are the most important examples.
A pertinent question is whether we can describe the elementary surface processes
of model catalysts on atomic scale and during the chemical reaction. Can we un-
derstand the basics of SC technology on microscopic scale? In all these issues, our
goal is the insight into the connection between the microscopic properties of matter
and its macroscopic behavior. What are the relevant concepts that help us reach
this goal?
The last few atomic layers of a solid constitute the interface with its environ-
ment. On this interface, there is a multiple of atomic and molecular processes that
take place in the quasi 2 dimensional (2DIM) stage. These processes are the basis
of our present-day technology. For example, without a detailed knowledge in the
production of SC devices, no progress could have been achieved in the information
and telecommunication technology. We also have access to nanostructured materi-
als with extraordinary functional properties, such as SC quantum dots and carbon
nanotubes. We have a growing understanding of how these structural features con-
trol the electronic properties. Throughout the years we have learned and mastered
the crystal growth. A real revolution was the invention of epitaxy. It allows the
fabrication of almost any material at will and makes possible the creation of any
alloy in 2DIM which otherwise does not exist according to the 3DIM phase diagram.
Most of the chemical reactions take place at the surface and heterogeneous catalysis
is a surface reaction, while the catalytic substance does not take part in the reac-
tion. A Ni-Fe-Cr alloy is called stainless steel because of its resistance to oxidation
and corrosion. In fact, owing to adsorption-induced segregation, Cr di↵uses to the
surface and binds to oxygen forming a thin oxide layer. The cromiumoxide cover at
the surface of the alloy acts as a protection and prevents further oxidation. Similar
surface passivation processes are successfully used in SC devices, like an atomic
layer of Gd2 O3 on GaAs surface. Internal di↵usion of impurities to the surface
results in segregation. In the worst case the grain boundary segregation of sulfur in

1
CHAPTER 1. INTRODUCTION: WHY SURFACES? 2

stainless steel is responsible for its brittleness. We lose so much energy due to the
friction, yet without friction we cannot even walk. Tribology deals with this surface
e↵ect.
The atoms of the bulk material are arranged in a symmetrical way. This sym-
metry allows many simplifications. The ion cores constitute a periodic potential.
Under its influence the electrons of the material can be described as Bloch waves.
There is a universal behavior of the bulk owing to the symmetry in all the phase
transitions such that we can speak of universality. At the surface the symmetry
is broken, no regularities can be found analogous to the bulk. The 3DIM phase
diagram cannot be applied. Any observation has to be dealt with separately. These
points are in fact responsible why Surface Physics or Surface Chemistry have ad-
vanced not before the last 50 years.
I will now mention some processes specific of the surface.

1.1 Surface Processes


At a general surface, physical and chemical modifications can take place not known
of for the bulk.

Figure 1.1: When a SC crystal is cleaved the top atomic layer may relax
by x in either direction, as seen in the panel on the right-hand side.

One can cleave a SC perpendicular to a crystallographic direction and expose a


surface for which the least amount of bonds are broken. No charge separation takes
place as a result of cleavage. The surface atoms thereby have a reduced coordination
and may react to this change in order to reduce the total energy. The shift normal
to the surface of the top atomic layer is called relaxation as depicted in Fig. 1.1.
If atoms are shifted pairwise lateral to the surface to form surface dimers
the atomic symmetry of the surface will change and the periodicity is doubled
[Fig. 1.2(left)]. This is a simple example for reconstruction. Buckling is illustrated
in the Fig. 1.2(right).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 3

Figure 1.2: The top atomic layer may show reconstruction (left) or buckling (right).

There may also occur chemical modifications at a general surface. Foreign atoms
may arrive at the surface and stick to it. Consider the potential formed between
the foreign atom, the adsorbate, and the surface, the substrate, shown in (Fig. 1.3).
In the case of physisorption there is a weak binding between the adsorbate and
the substrate, as is in the noble-gas adsorption on surfaces or adsorption of gases
on noble-metal surfaces. By a moderate heating the foreign atoms will be desorbed .
The binding is strong for chemisorption, while the adsorbates are trapped by the
strong attractive potential well (see Fig. 1.3).

Figure 1.3: The potential formed by the adsorbate and the substrate.

In chemisorption there is an electron transfer between the substrate and the


adsorbate, while physisorption is typically formed by some van-der-Waals forces. If
chemisorption proceeds, a new compound, an oxide, can be formed that di↵ers from
the bulk by chemical composition and atomic structure. Also surface segregation
leads to a similar situation. Thus, we need to fully characterize the surface from
scratch for its chemical composition, atomic geometry, and electronic properties.
CHAPTER 1. INTRODUCTION: WHY SURFACES? 4

1.2 Surface-Induced Chemical Reactions


In a heterogenous catalysis, the catalyzer does not take part in the reaction, it
just triggers the reaction. It provides the electron wave functions with the required
symmetry in order to combine the reaction components. An extremely important
example is the Haber-Bosch reaction for the synthesis of ammonia which dates back
to a time prior to the advent of surface physics or chemistry. It is an exothermal
reaction:
3H2 + N2 ! 2NH3 + 22.1 kcal (1.1)
We need small Fe crystals for the reaction to proceed at 200 atu and 475
600 C. This reaction is used to produce the artificial fertilizer without which a
great proportion of mankind would have starved during the 20. century. Fritz
Haber received the Nobel Prize in 1918 for his achievement and Carl Bosch in 1931.
The microscopic description of the reaction came as late as in 1975 by Gerhard Ertl.
He is a surface physicist from Munich and could explain the process with proper
wave functions upon which he received a professorship in Berlin at the Fritz-Haber-
Institute. Later he was awarded with the Nobel Prize 2007 in Chemistry.
Similarly, Fischer-Tropsch synthesis, is a collection of chemical reactions that
converts a mixture of carbon monoxide and hydrogen into liquid hydrocarbons.
The process produces synthetic fuel typically by burning low-cost coal, natural gas,
or biomass.
(2n + 1)H2 + nCO ! Cn H(2n+2) + nH2 O (1.2)
The Fischer-Tropsch process operates in the temperature range of 150 300 C and
uses Ni or Co catalysts.

1.3 Epitaxy
The thermodynamics of 3DIM structures are governed by their phase diagram.
This limitation does not apply to 2DIM systems, and therefore a wealth of di↵erent
materials can be fabricated at the surface with tailored properties. The growth
method is called epitaxy (epi = ‘top’ and taxis = ‘order’ in Greek).

Figure 1.4: Wetting of the surface by the adsorbate and island formation.

In epitaxy1 the surface is exposed to a gas, e.g., metal vapor, which condenses
on the surface. This way the surface becomes a contact place between two solids
1
J.R. Arthur, Surface Sci. 500, 189 (2002).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 5

which is called an interface. The fundamental question in epitaxy is whether the gas
atoms adsorbed on the surface will wet the surface or form islands. Figure 1.4(left)
schematically shows a mono atomic layer of adsorbate. This case occurs as a result
of strong forces between adsorbate and surface atoms at T = 0. This is a typical
case of adhesion. If, on the other hand, the adsorbate-adsorbate interaction is
stronger than adsorbate-surface interactions, island form on the surface which are
termed clusters. Hence, the wetting properties of a gas upon a specify surface are
the necessary conditions for the epitaxial growth.

Figure 1.5: Schematic illustration of the three epitaxial growth modes. From
R. Kern et al., In Current Topics in Materials Science, ed. E. Kaldis, Vol. 3,
Chapter 3, North-Holland, Amsterdam, 1979.

Epitaxial growth can basically be classifies in three modes, illustrated in Fig. 1.5.
In the simplest case, we may assume that the growth proceeds in a 2DIM fashion,
one layer after the next, up to some required film thickness. This is called layer-by-
layer growth. This is Frank-Van der Merwe (FV) growth, named after the investi-
gators first described the process. However, this is not always the case. One often
finds that the deposited material coagulates into clusters which at a stage may form
a polycrystalline layer. This is Volmer-Weber type growth; 3DIM crystallites form
upon deposition and some surface area remains uncovered at the initial stages of
deposition. Stranski-Krastanov growth is in-between, few layers may grow in FV
fashion before 3DIM clusters begin to form.
We can in fact estimate in advance which growth mode is more probably for
a given adsorbate-substrate system. We need to know three macroscopic quanti-
ties, namely the three surface tensions: a , i , s , the free energy per unit area at
the adsorbate-vacuum interface, the adsorbate-substrate interface, and substrate-
CHAPTER 1. INTRODUCTION: WHY SURFACES? 6

vacuum interface, respectively. We expect an ideal wetting of the substrate for


= a+ i s < 0, FV growth, VW growth for > 0, and SK growth for
= 0. For thicker films, i contains the contribution of the strained adsorbate
layer and therefore depends on the film thickness. This is the case for pseudomorphic
growth.
In epitaxy, atoms or molecules are deposited on the substrate and some struc-
tures evolve as a result of a multitude of processes. This is a non-equilibrium
phenomenon and any growth scenario is governed by the competition between ki-
netics and thermodynamics. Self assembly and self organization are modes through
which desired nanometer-size structures grow on the surface.
Microscopically, the primary mechanism in the growth of surface nanostructures
from adsorbate species is the transport of these species on a flat terrace, involving
random hopping processes at the substrate atomic lattice. This surface di↵usion
is thermally activated. This means that di↵usion barriers need to be surmounted
when moving from one stable (or metastable) adsorption site to another. The
di↵usivity D, which is the mean square distance travelled by an adsorbate per
unit time, obeys an Arrhenius law. If the deposition rate F of atoms in a growth
experiment is kept constant, then the ratio D/F determines the average distance
that an adsorbate species has to travel to meet another adsorbate for nucleation.
Thus, the ratio of D/F is a key parameter characterizing the growth kinetics. If the
deposition is slow (large D/F at the high-temperature limit), growth occurs close to
equilibrium conditions: the adsorbates have sufficient time to explore the potential
energy surface so that the system reaches a minimum energy configuration. If the
deposition is fast (small D/F ), then the pattern of growth is essentially determined
by kinetics; individual processes leading to metastable structures are important2 .
SC nanostructures are usually grown at intermediate D/F values and their
morphology is determined by the complex interplay between kinetics and ther-
modynamics. Strain e↵ects are particularly important and can be used to active
mesoscopic ordering.
Low-temperature growth of metal nanostructures on metal surfaces is the pro-
totype of kinetically controlled growth methods. Metal bonds have essentially no
directionality that can be used to direct interatomic interactions. Indeed, kinetic
control provides an elegant way to manipulate the structure and morphology of
metallic nanostructures. On homogenous surfaces, their shape and size are largely
determined by the competition between di↵erent displacements the atoms can make
along the surface, such as di↵usion on terraces, over and along step edges. Each of
these displacement modes has a characteristic energy barrier, related to the local
coordination of the di↵using atom. It is the natural hierarchy of di↵usion bar-
riers that determines the details of the growth process. Terrace smoothening by
step-flow growth is one example. Supermolecular self assembly is achieved at the
high-temperature limit close to equilibrium.
2
J.V. Barth et al., Nature 437, 671 (2005).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 7

Figure 1.6: Schematic energy band of a SC. The conduction band edge
is Ec and valence band edge is Ev . Analogous to metals, is defined as
E1 EF , where EF is the Fermi level.

Figure 1.7: Energy bands of (a) narrow gap GaAs and (b) large band gap
(AlGa)As. If GaAs is placed between two (AlGa)As layers by molecular
beam epitaxy, a quantum-well structure is developed. Ref. [3].

Epitaxy is based on the revolutionary ideas of Leo Esaki and Raphael Tsu back
in 1960’s and today it is extensively used in research and development as well as
CHAPTER 1. INTRODUCTION: WHY SURFACES? 8

in technology. The fabrication of superlattice diodes is a prominent example. First


let us look at the energy-level diagram of a SC shown in Fig. 1.6.

Figure 1.8: An electron micrograph of a superlattice structure. Ref. [4].

It is essential to note that the vertical axis is the energy of electrons. E1 Ev


is called the ionization potential , and E1 Ec the electron affinity. For an intrinsic
SC, EF is in the middle of the gap.
Superlattices are manufactured by alternate epitaxial deposition of GaAs and
(AlGa)As layers.3 GaAs quantum wells with small band gaps are found successively
between (AlGa)As layers that possess electrons confined in 2DIM, as displayed in
Fig. 1.7. Figure 1.8 shows a cross section observed in transmission electron micro-
scope, TEM, across a laser diode consisting of a superlattice structure.4 Observe
the precision in the production of numerous layers.

1.4 Surface Melting


Surface melting is a classical example for a surface-specific phenomenon. The slip-
periness of ice is widely referred to as premelting, which is the existence of liquid at
temperatures and pressures below the normal phase boundary5 . The atoms at the
surface are loosely bound compared to those in the bulk. As a result, the amplitude
3
L.J. Challis, Contemp. Phys. 33, 111 (1992).
4
D.D. Vvedensky, In Low-Dimensional Semiconductor Structures, ed. K. Barnham and D.D.
Vvedensky, CUP, Cambridge, 2008.
5
J.G. Dash et al., Rep. Prog. Phys. 58, 115 (1995).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 9

of surface-atom vibrations is larger and hence the surface softer. There are super-
cooled liquids (like glass), but no superheated solids, possibly because the surface
melts at a lower temperature than the bulk does. In the experimental terminology
we may speak of lower Debye temperature.
We characterize a phase with an appropriate order parameter (OP). In a phase
transition, e.g., solid/liquid, OP is best chosen in such a way that it is zero in one
phase and finite in the other. An abrupt change in OP at the critical temperature,
Tc , is characteristic of a first-order phase transition. In this case the two independent
curves of free energy cross each other and the system jumps from one state to the
other one, like in the case of nucleation and growth. In this type of the transition,
a seed is required to trigger the transition. In a continuous phase transition, two
equivalent phases coexist and become indistinguishable. OP changes continuously
with temperature, and near Tc it behaves like (T Tc ) . Figure 1.9 illustrates these
two kinds of phase transitions schematically.

OP OP

β
t

Tc Tc
T T

Figure 1.9: A typical first-order (left) and second-order (right) phase


transition. OP is plotted as a function of temperature, where t is the
reduced temperature and the critical exponent.

The numerical value of , the critical exponent, only depends on some physical
properties, like symmetry of the system or dimensionality of the order parameter.
The property that the phase transition behaves similarly for all systems with the
same dimensionality is called universality and suggests that unexpected phenomena
might take place at the surface (2DIM) in contrast to the bulk (3DIM).
Melting is a first-order phase transition for which the surface acts as a 2DIM
seed. In all investigations so far, the OP for the surface behaves like that in a
second-order phase transition so that we may say that the surface at temperatures
much lower than Tc anticipates the bulk melting. Depending on crystallographic
orientation, di↵erent melting temperatures have been observed for some metals6 .
As yet, there is no universal microscopic theory for surface melting.
6
J.F. van der Veen et al., Phys. Rev. Lett 59, 2678 (1987).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 10

Figure 1.10: Conventional cubic cell of the diamond lattice. A sixfold-


symmetric planar honey-comb unit is highlighted.

1.5 Carbon-Based Structures


Graphene7 is a flat monolayer of carbon atoms tightly packed into a 2DIM hon-
eycomb lattice. It is the basic building block for graphitic materials of any DIM.
Carbon is the first element of the 4th group. It has the 1s2 2s2 2p2 electronic config-
uration. In the case of 3DIM diamond structure, the outer 2s2 2p2 electrons form an
sp3 hybrid which has a tetrahedral symmetry with the extremely stable 109.47 the
bond angle. It is a compact structure, macroscopically the hardest lattices. Dia-
mond is an insulator with a large energy band gap. The diamond structure is shown
in Fig. 1.10. Nearest-neighbor bonds are drawn in. The four nearest neighbors of
each point form the vertices of a regular tetrahedron. In the 3DIM structure, a
(111) plane is highlighted in order to emphasize the relation to graphene. Hence,
diamond structure can be thought of a special way of stacking graphene layers along
the [111] direction under consideration of the tetrahedral symmetry.

Figure 1.11: Graphite is formed by periodic stacking of individual graphene layers.

7
A.K. Geim and K.S. Novoselov, Nature Mat. 6, 183 (2007).
CHAPTER 1. INTRODUCTION: WHY SURFACES? 11

Figure 1.12: Fullerenes with 60, 70, 76, and 78 carbon atoms. Ref. [8].

Another 3DIM carbon-atom morphology is the graphite. In graphite, carbon


atoms occupy a 2DIM hexagonal lattice with 120 bond angles formed by the sp2
hybridization. The additional p electron fixes the hexagonal layers within a loose ⇡
bonding. The distance between the planes is almost 4.8 times the nearest-neighbor
distance within planes. The graphite material is soft, black, and shows a metallic
conduction along the graphene sheets. The graphite structure is schematically
illustrated in Fig. 1.11. On the right-hand side, the sp2 electron orbitals are shown
where the orbitals are 120 apart on a plane, while the ⇡ orbitals are perpendicular
to this plane.
Fullerene is a 0DIM cluster of carbon atoms arranged on the vertices of its typical
dome-like structure in icosahedral symmetry, as shown in Fig. 1.12. Fullerenes are
found as a by-product of carbon burning. They show spectacular properties when
doped with metallic species from being magnetic to superconducting.8
Nanotubes9 are 1DIM cylindrical structures based on the hexagonal lattice of
carbon atoms that forms crystalline graphite. By rolling up the graphene sheet a
chiral vector C is defined by C = na1 + ma2 (chiral, cheir = ‘hand’ in Greek). The
chiral angle is defined between C and a1 . For the chiral nanotube, is between
0 and 30 . The nanotube is termed armchair if n = m and = 30 , zigzag for
m or n = 0 and = 0. This situation is illustrated in Fig. 1.13. The electronic
properties of nanotubes are determined by their diameter and the chiral angle. For
the motion of electrons, a nanotube is metallic if n m = 3q with q an integer.
Thus, all armchair nanotubes are metallic, so are 1/3 of zigzag nanotubes; the
rest is semiconducting. The conductivity does nor depend on the length L of the
nanotube.
8
G. Sun and M. Kertesz, J. Chem. Phys. A 104, 7398 (2000).
9
http://physicsweb.org/articles/world/11/1/9
CHAPTER 1. INTRODUCTION: WHY SURFACES? 12

Figure 1.13: Nanotube is a graphene sheet rolled into a 1DIM tube. Ref. [9].

1.6 Scattering Cross Section and Mean Free Path


According to the classical description, atoms consist of a positively charged nu-
cleus which is enclosed by an electron cloud. This model is also referred to as the
Rutherford model. Around 1910 experiments have been conducted with ↵ par-
ticles to investigate the classical ideas about the electron cloud. Not much was
known about electron orbits microscopically, wave mechanics had not emerged yet.
Nevertheless, some ideas about scattering were developed which are still used suc-
cessfully today, as is done for scattering. So we first deal with elastic cross section
in scattering.
A scattering event is elastic if the energy of the system is not changed. So if
a small particle with a mass m and initial velocity ~vo and initial momentum p~o
collides with a larger mass M initially at rest, we have after the collision ~v1 and p~1
for the small mass and V~2 and p~2 for the larger mass. Consider central collision for
simplicity.
In elastic scattering momentum and energy are both conserved. Hence p~o =
p~1 + p~2 and Eo = p2o /(2m) = p21 /(2m) + p22 /(2M ). For an energy transfer E during
the collision we obtain
4mM
E= Eo (1.3)
(m + M )2
CHAPTER 1. INTRODUCTION: WHY SURFACES? 13

This expression is smaller for a noncentral collision. For M m we can write


E ' 4m/M Eo . For scattering of electrons at isolated atoms (m/M  10 4 ), we
obtain E  10 4 Eo . For scattering at a solid with 1023 atoms/cm3 the energy
transfer is even smaller E ' 10 27 Eo . We realize that there is practically no
energy transfer if a small particle collides with a larger one.

The di↵erential cross section @ /@⌦ is defined as the e↵ective area per atom
scattering into the solid angle ⌦. The number of scattered particles is given by
@
Ns (⌦, ⌦) = Io ⌦N, (1.4)
@⌦
where Io is the intensity of incoming particles, N number of target particles. The
target area is R2 ⌦. Then the intensity of the scattered beam is given by

Is (⌦) · R2 ⌦ = Ns (⌦, ⌦). (1.5)

which leads to
@ Is (⌦) R2
= (1.6)
@⌦ Io N
R2 and N are known quantities.
The scattering probability is
@W @ N @
= = nd (1.7)
@⌦ @⌦ Ao @⌦
with Ao the area, n the density of target, and d the path length of scattering
particles. These ideas are valid for dilute targets where multiple scattering can be
neglected. Integration over ⌦ results in W = nd. For W = 1 we obtain d = 1/n .
This quantity is called the mean free path or escape depth, ⇤.
Actually, this derivation is valid for thin targets with d < ⇤. Otherwise multiple
scattering will dominate. For an infinitesimally thin layer we may write @W =
n@d. If we scale with intensity, we obtain
@I(d)
= Io (d) n. (1.8)
@d
Considering the conservation of particles, i.e., Io (0) = I(d) + Io (d), leads to

nd d/⇤
Io (d) = Io (0)e = Io (0)e (1.9)
In all experiments we use in our investigations, there are electrons, photons,
atoms, or ions involved. These particles interact with the solid with di↵erent inten-
sities. As a result, the mean free path is limited, and particles are either strongly
attenuated when they enter the solid or during escape. In any case, experiments are
more surface sensitive the shorter the mean free path is. Generally, ⇤ for photons
is quite long, while for electrons in the energy range 30 < E < 300 eV as short
as few-atomic distances. Thus, experiments involving such electrons are the basic
tools in Surface Science.
CHAPTER 1. INTRODUCTION: WHY SURFACES? 14

1.7 Vacuum Technique


There are several reasons why Surface Chemistry and Surface Physics have devel-
oped relatively late. The most important one is the question how to prepare a
clean surface and how to keep it clean during the measuring time. In general, a
SC surface is exposed by cleavage. Unfortunately, only a few crystallographically
defined surfaces are accessible. Others, like all the metals, are first oriented along
the desired direction by x-ray methods and subsequently cut by spark erosion to
expose the net plane. Prior to introducing into the vacuum chamber, the surface is
polished with appropriate powders with decreasing grain size down to 0.3 0.1 µm.
In vacuum, surfaces are cleaned by bombardment with Ar+ ions of 500 2000 eV
and heated to elevated temperatures to restore the crystalline structure. Once one
has an appropriate surface, several systems, metals, alloys, SC’s, can be generated
by epitaxy.
A thus cleaned surface does not remain clean during long periods. The time a
one-monolayer (ML) contaminant reaches the surface is used as a measure for the
quality of the experiment and depends on the vacuum conditions. So the question
is, for given vacuum conditions, how long does a surface remain clean?
For an ideal gas, N = pV /kT is the number of particles in a volume V (l),
at a pressure p (Torr) and Temperature T (K) with the Boltzmannqconstant k.
Consider that the Maxwell distribution relates the average speed v̄ = 8kT /⇡m of
the particles with their mass m and the temperature T . Air molecules, CO or CO2 ,
at 20 C have an average speed of v̄ ⇡ 500 m/s.
p
Mean free path ⇤ of the particles is given by ⇤ = 1/( 2⇡d2 )(N/V ) 1 , while
d is the particle diameter. Thus, for air at p = 1 Torr, we have ⇤ = 4.5 µm at
room temperature. These particles move in the experimental chamber and hit all
exposed surfaces. The number of particles n that hit a surface of area F in a
time t is given by n/ F / t = 1/4 v̄ (N/V ) = v̄/4 (p/kT ), where the number
4 considers di↵erent directions.
Now we define the contamination time ⌧ (s) as an interval during which an
originally clean surface is covered by 1 ML of adsorbed particles:
1015 /( n/ F / t) ⇡ 3x10 6 /p (Torr).
(1 ML ⇡ 1015 atoms/cm2 ; 1 Pa = 10 5
bar; 1 Torr = 1.33 mbar = 1.33 x 10 2
Pa)
Now we have a stringent condition that we have to perform the experiments in
vacuum with the best possible conditions. For a pump, one defines the pumping
power Q = kT Ṅ (Torr·l/s) and pumping speed S = Q/p (l/s).
For the evacuation of an experimental chamber, we can write:
Ṅ = dN/dt = (V /kT )dp/dt and Ṅ = Q/kT = pS/kT , which results in:
p(t) = po exp ( S/V t). Hence, the decrease of pressure is limited.
Consider a chamber of cubic volume of V = 1 m3 at a pressure of p = 10 5 Torr.
It contains N = 1017 atoms. The same chamber has ⇡1020 atoms (6 x 104 cm2 x
CHAPTER 1. INTRODUCTION: WHY SURFACES? 15

1015 atoms/cm2 ) sticking at its inner walls if only 1 ML of adsorbates is present.


Hence, the number of atoms and molecules adsorbed at surfaces is much higher
than those present in the volume. Therefore, we have to get rid of the adsorbates
by making them desorb at elevated temperatures. So, we bake out the chamber to
attain good vacuo. This fact limits the choice of materials used to construct the
experimental chamber to those with high vapor pressure.
In this course we will deal with spectroscopic experiments and results. Every
spectroscopy has three ingredients. First, there is an initial disturbance, an excita-
tion. As a result, the system undergoes a transition from the ground state to an
excited state. This transition costs some energy and has a probability to occur.
We measure both, considering that the probability of the transition is the intensity
which is experimentally accessible. The excited state has some limited life time that
determines the accuracy of the observations. In the last step, the system relaxes
back to the ground state by emitting some energy. Also this time, we observe the
process. Our hope is that the measured quantities constitute the dominant part of
the transition.
The major lesson is that we cannot ever measure the ground state – observable
are the excitations.
Chapter 2

Electronic Structure

The electronic properties of matter determines its macroscopic behavior. The mag-
netic phenomenon or the superconducting behavior of a metal has its roots in the
electronic structure. With the expression electronic structure we mean the ener-
getic and spatial distribution of electrons and how their energy is related to their
motion in a solid. The electronic structure of each element is di↵erent. Yet, there
are close resemblance between elements of the same group, i.e., that lie in the same
column of the periodic table. The reason is given by the fact that the electronic
shells are filled by one electron as we go from left to right in the periodic table, and
elements with the same shell occupancy show similar behavior. Instead of deal-
ing with di↵erent elements1 one considers elements in some characteristic groups.
Here, we first analyze the behavior of free electrons. In this free-electron model
the weekly bound electrons of the atoms move about freely in the metal without
the influence of the attractive potential of ion cores. Thus the potential energy is
neglected, the total energy is in the form of kinetic energy. The free-electron model
gives us considerable information about several electronic properties of the so-called
simple metals. The alkali metals or nobel metals can be regarded as simple metals.
The free-electron model cannot explain the reasons why some elements are metals
and others insulators. Besides, in the free-electron model electrons can travel long
distances without scattering. In the next step, we place the electrons in the periodic
potential of the ions and see the formation of energy bands, band gaps, e↵ective
masses, etc. At the end, we mention the other extreme, the localized states. We
need this approach to account for the nature of the 4f electrons of the Rare-Earth
elements.
These subjects will be a repetition for you with your Solid State Physics back-
ground.
Subsequently, we apply these ideas to surfaces. Since the symmetry is broken in
one direction (it is customary to define this as the z-direction) we expect a severe
response in the electronic behavior because the electronic motion is restricted and
1
V.L. Moruzzi, J.F. Janak, and A.R. Williams, Calculated electronic properties of metals, Perg-
amon, New York, 1978.

16
CHAPTER 2. ELECTRONIC STRUCTURE 17

new states may be formed that do not exist in the bulk.


In the following chapter, we will learn some experimental techniques that are
currently used to investigate the electronic properties of the bulk and the surface.

2.1 Free Electrons in Metals


2.1.1 Free Electrons in 3DIM
An unconfined electron in free space is described by the Schrödinger equation
✓ ◆
h̄2 2 h̄2 @ 2 ' @ 2 ' @ 2 '
r '= + 2 + 2 = E', (2.1)
2m 2m @x2 @y @z
where m is the free-electron mass. The solutions of this equation,
1
'k (r) = 3
eik·r (2.2)
(2⇡)

are plane waves labelled by the wave vector

k = (kx , ky , kz ) (2.3)

and correspond to the energy

h̄2 k 2 h̄2 2
E= = (k + ky2 + kz2 ). (2.4)
2m 2m x
The vector components of k are the quantum numbers for the free motion of the
electron, one for each of the classical degrees of freedom.
The number of states in a volume dk = dkx dky dkz of k-space is

2
N (k) dk = dk (2.5)
(2⇡)3

with the factor of two accounting for the spin-degeneracy of the electrons. To
express this density of states in terms of energy states, we use the fact that the
energy (Eq. 2.4) depends only on the magnitude of k. Thus, by using spherical
polar coordinates in k-space,

dk = k 2 sin ✓ dk d✓ d , (2.6)

where the variables have their usual ranges (0  k < 1, 0  < 2⇡, and 0  ✓  ⇡)
and integrating over the polar and azimuthal angles, we are left with an expression
that depends only on the magnitude k:
2 1
N (k) dk = 3
dk = 2 k 2 dk (2.7)
(2⇡) ⇡
CHAPTER 2. ELECTRONIC STRUCTURE 18

By invoking Eq. 2.4, we can perform a change of variables to bring the right-hand
side of this equation into a form involving the di↵erential of the energy:
✓ ◆ ✓ ◆3/2 p
1 2 1 2mE dk 1 2m
k dk = 2 2 dE = 2 E dE (2.8)
⇡ 2 ⇡ h̄ dE 2⇡ h̄2
From this equation, we deduce the well-known density of states g(E) of a free
electron gas in three dimensions:
✓ ◆3/2 p
1 2m
N (E) = E (2.9)
2⇡ 2 h̄2
Notice the characteristic square-root dependence on the energy (Fig. 2.1). This
results from the fact
p that in 3DIM, the surfaces of constant energy in k-space are
spheres of radius 2mE/h̄.

Figure 2.1: The densities of states g(E) of an ideal 3DIM electron gas.

The density of states weighted by the amplitude square of the wave function is
denoted as the local density of states.
X
N (E, r) = |'i (r)|2 (E Ei ), (2.10)
i
R
where density of states is obtained by integration: N (E) = g(E, r)dr.

2.1.2 2DIM Electron Gas


If we restrict electron motion in one dimension, for instance in the z-direction, and
let the electrons move freely in the other two directions, we encounter a quantum
CHAPTER 2. ELECTRONIC STRUCTURE 19

well in the z-direction. The Schrödinger equation is the same as that in Eq. 2.1 for
free electrons in 3DIM
h̄2 2
r ' = E', (2.11)
2m
but the boundary conditions on ' are di↵erent. In the direction of confinement
along the z-direction we locate the confining planes with infinite potentials at z = 0
and z = L. The boundary conditions for ' are

'(x, y, 0) = '(x, y, L) = 0 (2.12)

Thus, a natural way to solve the Schrödinger equation is by the method of separation
of variables. By writing the wave function ' as

'(x, y, z) = (x, y) (z) (2.13)

and substituting into Eq. 2.2 we obtain the Schrödinger equations for
✓ ◆
h̄2 @ 2 @2
+ = Ek (2.14)
2m @x2 @y 2
and for ,
h̄2 d2
= En (2.15)
2m dz 2
with the boundary condition

(0) = (L) = 0 (2.16)

Notice that the equation and boundary condition for are precisely those for
a particle in a 1DIM box with infinite barriers at the edges. The solution for the
not-normalized eigenfunction is

n (z) = sin(kn z), (2.17)

where kn = n⇡/L and n a positive integer. The energy associated with these
functions are
h̄2 kn2
En = (2.18)
2m
By solving the Schrödinger equations in Eq. 2.14 and Eq. 2.15, we find that the
energy eigenvalues for the quantum well in all three directions are given by

h̄2 2
En,k = (k + kx2 + ky2 ), n = 1, 2, . . . (2.19)
2m n
This shows that this energy dispersion has features of both the 2DIM electron gas
(associated with the motion in the x-y plane) and the 1DIM electron gas (associated
with the confinement along the z direction).
In the ground state of a system of N free electrons with an electron concentration
of N/V , where V is the volume, the occupied states may be represented as points
CHAPTER 2. ELECTRONIC STRUCTURE 20

Figure 2.2: First three energy levels and wave functions of electron con-
fined to z-direction within a sheet of thickness L. The wavelengths are
indicated on the wave functions. From CK.

inside a sphere in k space. The energy at the surface of the sphere is the Fermi
energy EF with the wave vector kF such that
h̄2 2
EF = (k ). (2.20)
2m F
Considering that there is one allowed wave vector for the volume element in k space
and two allowed values of ms , the spin quantum number, for each allowed value of
k, we find that EF depends on the mass and electron concentration:
h̄2
EF = (3⇡ 2 N/V )2/3 . (2.21)
2m

2.1.3 Fermi-Dirac Distribution Law


The ground state is the state of the system at absolute zero. The situation that
happens as the temperature is increased is given by the Fermi-Dirac distribution
function. The kinetic energy of the electron gas increases as the temperature is
increased, and some energy levels are occupied which were vacant at absolute zero,
and some levels are vacant which were occupied at absolute zero. The situation is
illustrated in Fig. 2.3, where the plotted curve is of the function

f (E) = (e(E µ)/kB T


+ 1) 1 . (2.22)
CHAPTER 2. ELECTRONIC STRUCTURE 21

Figure 2.3: Plot of Fermi-Dirac distribution function f (E) versus E/µ


for zero temperature and for a temperature kB T = 15 µ. The value of
f (E) gives the fraction of levels at a given energy which are occupied
when the system is in thermal equilibrium. From CK.

The Fermi-Dirac distribution function gives the probability that a state at energy
E will be occupied in thermal equilibrium. The quantity µ is also a function of
temperature. The quantity µ is called the chemical potential, which is equal to EF
at absolute zero. At all temperatures f (E) is equal to 12 when E = µ. As usual,
EF is defined as the topmost filled energy state at absolute zero.

2.2 Band Theory of Metals


Every solid contains electrons which are arranged in energy bands separated by
regions for which no electron energy states are allowed. Such forbidden regions are
called energy gaps. If the number of electrons in a crystal is such that the allowed
energy bands are either filled or empty, then no electrons can move in an electric
field and the crystal will behave as an insulator. If the bands are partially filled,
the crystal will act as a metal. To understand the di↵erence between insulators
and conductors the free-electron model must be extended to take account of the
periodic lattice of the solid.

2.2.1 Periodic Lattice


A crystal consists of a periodic repetition of a set of atoms in space. Periodic
structures have long-range translational order and can be described by a lattice and
CHAPTER 2. ELECTRONIC STRUCTURE 22

a basis (unit cell). A unit cell is a collection of atoms at each point; this collection
is identical at each lattice point. In simple solids, the basis consists of one single
atom, complicated organic structures may have thousands of atoms. We need three
basis vectors a. b, and c, which define the unit cell. With the basis and the lattice,
all structures are uniquely defined.
3DIM structures can be described by symmetry operations which map the struc-
ture in itself. Translation is a parallel shift of the structure by T = pa + qb + r c,
which is called translational invariance. p, q, and r are integers, analogous to the
Miller indices. Any point in the crystal can be reached from the origin using T.
The basis vectors a, b, c span the unit cell. The sum of all translations for any p.
q, and r is called the translational group of the structure. The translational group
defines the 3D periodicity of the structure.
Rotation and mirror reflection are called point operations. They leave 1 point or
1 line unchanged. Point group consists of operations which leave 1 point unchanged
and the structure invariant. In periodic structures there are two possible operations:
mirror reflection around a line (mirror plane) and rotation through 2⇡/n (n =
1, 2, 3, 4, 6) around a point. The numbers give the n-fold rotation around a point.
Only these are compatible with translational properties.
In general, there are only some selected translation symmetries that are com-
patible with a given point group. There are fourteen di↵erent lattice types, Bravais
lattices, which fulfill this requirement. More precisely, we can state that the choice
of the unit cell in a crystal is not unique, the basis has to be defined accordingly.
If the smallest possible unit cell contains just one atom, such a lattice is called a
Bravais lattice.
It is functional and appropriate, that any periodic function f (r) is represented
by its Fourier transform. Then, we deal with Fourier components of the function
rather than with the function itself in real space. Any function defined for a crystal,
such as the electron density, is periodic with the same translation vector T as the
basis vectors. We can write
f (r + T) = f (r). (2.23)
The crystal itself transforms into the reciprocal lattice with the vectors r⇤ , such that

G = ha⇤ + k b⇤ + l c⇤ . (2.24)

The function f ⇤ (r⇤ ) in the reciprocal space has the property

f ⇤ (r⇤ + G) = f ⇤ (r⇤ ) (2.25)

We define a unit cell in the reciprocal space beyond which f (r) repeats itself. The
smallest such unit cell is called a Brillouin zone. It is defined by the area surrounded
by the planes that are perpendicular bisectors of the vectors from the origin to the
reciprocal lattice points.
CHAPTER 2. ELECTRONIC STRUCTURE 23

It is remarkable that the value of k enters into the conservation laws for collision
processes of electrons in crystals. For this reason k is called the crystal momentum
of the electron. In a process involving a momentum transfer K, we can write

k + K = k0 + G (2.26)

with G being the reciprocal lattice vector.

Figure 2.4: Plot of energy E versus wave vector k (a) for a free electron
and (b) for an electron in a monatomic linear chain with a lattice constant
a. The energy gap Eg is associated with the first Bragg reflection at
k = ±⇡/a. From CK.

2.2.2 Motion of Electrons in a Periodic Potential


In many situations the band structure of a crystal can be accounted for by the
nearly-free electron model for which the band electrons are treated as perturbed
only weakly by the periodic potential of the ion cores. Often the gross overall
aspects of the band structure can be explained on this model. We consider, for the
sake of simplicity, a linear chain of atoms with a lattice constant a. In Fig. 2.4 the
band structure is shown (a) for free electrons and (b) nearly-free electrons with an
energy gap at k = ±⇡/a. The Bragg condition (k + G)2 = k2 for di↵raction of a
wave with the wave vector k becomes in 1DIM
1
k = ± G = ±n⇡/a, (2.27)
2
where G = ±2n⇡/a and n is an integer. The first reflection and the first energy
gap occur at k = ±⇡/a. The reflections at these particular values of k arise because
the wave reflected from one atom in the linear lattices interferes constructively with
the wave reflected from a nearest-neighbor atom. The di↵erence in phase between
the two reflected waves is just ±2⇡ for these two values of k. The region in k space
between ⇡/a and ⇡/a is called the first Brillouin zone of this lattice.
CHAPTER 2. ELECTRONIC STRUCTURE 24

Figure 2.5: (a) Variation of potential energy of a conduction elec-


tron in the field of the ion cores in a linear lattice. (b) Distribution
of probability density in the linear lattice for | (+)|2 / cos2 (⇡x/a),
| ( )|2 / sin2 (⇡x/a), and for a traveling wave. The wave function (+)
piles up electronic charge on the cores of positive ions, thereby lowering
the potential energy, while (+) piles up charge in the region between
the ions removing it from the ion cores and thereby raising the potential
energy. From CK.

2.2.3 Band Gaps


At k = ±⇡/a, when the Bragg condition is satisfied, a wave traveling in one direction
is Bragg-reflected and then travels in the opposite direction. Thus, at k = ±⇡/a,
the wave functions are made up equally of waves traveling to the right and to the
left and form two di↵erent standing waves. If the traveling waves are in the form
ei⇡x/a and e i⇡x/a , the standing waves are:

(+) / (ei⇡x/a + e i⇡x/a


) = 2 cos(⇡x/a);
( ) / (ei⇡x/a e i⇡x/a
) = 2 cos(⇡x/a). (2.28)

The standing waves (+) and ( ) are even and odd, respectively, when x is
substituted for x.
Figure 2.5(a) indicates schematically the variation of potential energy of a con-
duction electron in the field of the positive ion cores of a monatomic linear chain.
The potential energy of an electron in the field of a positive ion is negative. In
Fig. 2.5(b) the distribution of electron density corresponding to the standing waves
(+), ( ), and to a traveling wave is sketched. The traveling wave / eikx dis-
tributes electrons uniformly with | |2 = 1, while standing waves distribute electrons
CHAPTER 2. ELECTRONIC STRUCTURE 25

preferentially either midway between ion cores [ ( )] or on the ion cores [ (+)].
The potential energy of the three charge distributions is di↵erent in such a way
that it is higher for ( ) than for a traveling wave and lower for (+) compared
to a traveling wave. If the potential energies of ( ) and (+) di↵er by Eg , there
is an energy gap of width Eg between the two solutions at k = ⇡/a in Fig. 2.4 or
between the two solutions at k = ⇡/a. The wave functions at points A in Fig. 2.4
will be (+), and the wave functions above the energy gap at points B will be
( ).

2.2.4 Bloch Functions


Bloch2 has proved that the solutions of the Schrödinger equation with a periodic
potential are of the form
= uk (r)eik·r , (2.29)
where u is a function, depending in general on k,which is periodic in x, y, z with the
periodicity of the potential (that is, with the period a of the lattice). This amounts
to that the plane wave eik·r is modulated with the period of the lattice. In order
to justify the Bloch theorem, we can follow some simple arguments. Consider N
lattice points on a ring of length N a, and suppose that the potential is periodic in
a, so that
V (x) = V (x + ga), (2.30)
where g is an integer. Because of the symmetry of the ring we look for eigenfunctions
such that
(x + a) = C (x), (2.31)
where C is a constant. Then

(x + ga) = C g (x); (2.32)

and, if eigenfunction is to be single-valued.

(x + N a) = (x) = C N (x), (2.33)

so that C is one of the N roots of unity, or

C = ei2⇡g/N ; g = 0, 1, 2, . . . , N 1. (2.34)

We have then
(x) = ei2⇡g/N a ug (x) (2.35)
as a satisfactory solution, where ug (x) has periodicity a. Letting

k = 2⇡g/N a, (2.36)
2
F. Bloch, Z. Physik 52, 555 (1928).
CHAPTER 2. ELECTRONIC STRUCTURE 26

we have
= eikx uk (x), (2.37)
which is the Bloch result. A function in the form of the form Eq. 2.37 is known as a
Bloch function. All one-electron wave functions in an ideal crystal are of the Bloch
form.

2.2.5 Reduced-Zone Scheme


It is often convenient to select the wave vector k of the Bloch function so that it
always lies within the first Brillouin zone. This procedure is known as the reduced-
zone scheme. If we encounter a Bloch function written as
0
k0 (r) = eik ·r uk0 (r), (2.38)

with k0 outside the first zone, we may always find a suitable reciprocal lattice vector
G0 such that
k = k 0 G0 (2.39)
lies within the first Brillouin zone. Then Eq. 2.38 may be written as
0
⇣ 0

k0 (r) = eik ·r uk0 (r) = eik·r eiG ·r uk0 (r) ⌘ eik·r uk (r) = k (r), (2.40)

where we have defined


0
uk (r) ⌘ eiG ·r uk0 (r). (2.41)

Figure 2.6: (a) The result of the perturbation associated with the nuclear poten-
tials on the free-electron levels. Gaps are opened up at ±n⇡/a. (b) The bands
using the extended-zone scheme may be folded into the first zone of (a). From
J.K. Burdett, Chemical Bonding in Solids, Oxford University Press, Oxford, 1995.
CHAPTER 2. ELECTRONIC STRUCTURE 27

0
Both eiG ·r and uk0 (r) are periodic in the crystal lattice so uk (r) is also, hence
k (r) is of the Bloch form. Even with free electrons it may be useful to work in the
reduced-zone scheme, as is seen in Fig. 2.6. It follows also that any energy Ek0 for k 0
outside the first zone is equal to an Ek in the first zone, where k is related to k 0 by
Eq. 2.39. Thus we need solve for the energy only in the first Brillouin zone, for each
band. The opening of the band gap is illustrated in Fig. 2.7 for the reduced-zone
scheme.
Near the top or bottom of a band the energy is generally a quadratic function
of the wave number, so that by analogy with the expression E = (h2 /2m)k 2 for free
electrons we may define an e↵ective mass m⇤ such that @ 2 E/@k 2 = h̄2 /m⇤ . The
motion of the electron is characterized by this e↵ective mass, m⇤ .

Figure 2.7: (a) Ek relationship drawn in the reduced-zone scheme (a) for
free electrons. The branch AC is reflected in the vertical line at k = ⇡/a
gives the usual free-electron parabola for Ek vs k for positive k. (b) A
crystal potential U introduces band gaps at the edges of the zone, but
the overall features of the band structure remain. The dashed portions
are due to free-electron parabola. Two energy bands ↵ and are shown
separated by Eg at k = ±⇡a. From CK.
CHAPTER 2. ELECTRONIC STRUCTURE 28

2.3 Tight-Binding Approximation


A wide question relates to the relative importance of band structure and atomic
correlation e↵ects in solids. The conduction electrons of free-electron like metals,
such as alkali metals or Al, are shared between atoms for conduction, and may
be treated by the above methods. The potential in which they move is rather
smooth, and they can be well represented by plane waves. In band calculations,
available electrons are successively filled into the calculated bands with di↵erent `-
components and the Fermi level is obtained by the state filled by the last electron.
In contrast, if electrons, or pairs of electrons, are localized in covalent bonds,
they are in a state associated with a specific atom. In the extreme case of the inner-
shell electrons that are localized at atomic sites, the dominant process of conduction
is the motion of electron from one atomic site to other. This motion is governed by
correlation e↵ects came about by the interplay between repulsive electron-electron
interaction and wave function hybridization. Usually, this interplay is in favor of
charge interaction, and in the extreme case of heavy Fermions the system shows
Fermi-liquid behavior with a large m⇤ .
The 4f -electrons of most rare-earth metals and narrow-band d-electrons of tran-
sition metals behave as if they were unfilled core levels and atomic interactions
dominate their behavior. In this case, the binding energy of electrons is a strong
function of the band occupancy. Such bands show predominantly occupations by
an integer number of electrons. For these localized-electron systems, interatomic
type charge fluctuations may be responsible for electronic conduction in the form
dnA dnB ! dnA 1 dn+1
B , where A and B are atomic sites and n the electron occupan-
cies. Roughly, for the conduction to occur, the energy required for this process
should be less than the d-band width w for transition metals. If the d d Coulomb
and exchange interaction U is larger than the dispersional band width w than it
is impossible for the above charge fluctuation to occur and the material becomes
an insulator in spite of its unfilled d shell. These are the so-called Mott-Hubbard
insulators. If U < w, the system will be a metal as predicted in the elementary
band theory.
The localized-electron behavior is well accounted for by the approximation which
starts out from the wave functions of the free atoms which is known as the tight-
binding approximation. It is quite good for the inner electrons of atoms and used
to describe the d-bands of some of the transition metals.
Consider an atomic orbital (r ~`) with a well-defined character, e.g., 1s, 2s, 2p
orbitals centered on an atom at position ~` in the crystal. In this approximation,
the wave functions are constructed using such orbitals and obey Bloch theorem
X ~
k (r) = eik·` (r ~`). (2.42)
~
`

In the orthogonalized plane waves method the localized and extended characters
of the wave function are combined. An atomic region around each atom is defined
CHAPTER 2. ELECTRONIC STRUCTURE 29

where the wave function is described in terms of atomic orbitals. Outside this region,
the relatively smooth parts of wave functions are expanded in terms of plane waves.
Also in this method, one constructs a set of Bloch functions using occupied core
states of the ions in each atom.

2.4 Surface Electronic Structure


Fundamental aspects of Surface Science have their roots in electronic properties.
These include the charge density in the neighborhood of the vacuum interface, the
di↵erence of the electron states near the surface compared to those in the bulk,
chemical bonding states in the first few atomic planes, the electrostatic potential
felt by surface atoms. Surface states that are present at the surface and not in
the bulk give us mostly the clue about the behavior of surfaces. The macroscopic
behavior of surfaces and interfaces, like oxidation, heterogeneous catalysis, or crystal
growth, strongly depend on its electronic properties.

Figure 2.8: Electron density profile with a positively charged uniform


background n̄ for z  0. From N.D. Lang and W. Kohn, Phys. Rev. B1,
4555 (1970).

In order to study the electronic structure of surfaces, one starts with a semi-
infinite crystal which has total number of electrons N at positions R. We have to
consider the kinetic energy of all the electrons, the electron-ion interaction for all
N and R, and the electron-electron repulsion for all electrons. The final term is
not straightforward to handle, it is namely the exchange-correlation term, and since
1940’s several approaches have been used to satisfactorily solve the problem. In the
CHAPTER 2. ELECTRONIC STRUCTURE 30

so-called jellium model the discrete ion potential of a semiinfinite lattice is approxi-
mated by an averaged uniform positive charge density, n̄. Inverse charge density of
the background is often related to a spherical volume: (4⇡/3)rs3 = 1/n̄. Figure 2.8
displays the electron density profile n(z) for two choices of the background density
rs . As a result of uncertainty principle electron density may not abruptly change
from zero to its finite value n̄ as we enter the solid from the vacuum side, this
means that there is no sharp edge to the electron distribution. As a consequence,
there is an exponentially decaying probability to find electrons outside the solid.
In other words, electrons spill out into the vacuum region for z > 0 and thereby
create an electrostatic dipole layer at the surface. We also notice that n(z) oscillates
as it approaches an asymptotic value that exactly compensates the uniform bulk
background charge. The wavelength of these Friedel oscillations is ⇡/kFr , where
kFr = (3⇡ 2 n̄)1/3 . The oscillations arise because the electrons try to screen out the
positive background charge distribution which includes a step at z = 0.
The formation of a surface dipole layer is a result that the electrostatic potential
in vacuum is larger than the mean value in the crystal. This potential step keeps
the electrons within the crystal. The exchange and correlation is a bulk e↵ect
which makes neighboring electrons stay away from each other and thus lowers the
potential energy of each electron. The work function, , which is the minimum
energy required to remove an electron from the bulk to a point away from crystal,
is given by the dipole layer. Whenever the atomic density at the surface is large, the
spilling out is similarly large, and the work function has a smaller value. Because
of this surface contribution, depends sensitively on the exposed crystalline plane
as well as on the impurity e↵ects at the surface. Surface geometric e↵ects, like
reconstruction or relaxation, also modify surface dipole and consequently the work
function, as expected. In calculations a reasonably accurate polycrystalline work
function of some metal surface can be obtained using a uniform positive background
jellium model.
The jellium model of a metal surface neglects the electron-ion interaction and
emphasizes the smooth surface potential barrier. Tamm3 has investigated a linear
chain of 1DIM atoms possessing delta-function like positive potentials. For the bulk,
he obtained solutions for the Schrödinger equation in the form of (z) = uk (z)eikz ,
where uk (z + na) = uk (z) reflects the periodicity of the linear chain. For the real
values of k(z) the solutions for (z) are the common Bloch waves, extended over
the entire chain. There are also complex values of k(z) which are associated with
surface states, present at the surface and decay exponentially inside the surface for
the z < 0.
Shockley4 has similarly considered atomic potentials arranged in a linear chain
with the interatomic distance a. For large a the energy values resemble that of free
atoms, and there are no surface states. As a is deceased, energy values come together
3
J.E. Tamm, Z. Physik 76, 849 (1932).
4
W. Shockley, Phys. Rev. 56, 317 (1939).
CHAPTER 2. ELECTRONIC STRUCTURE 31

and broaden to form bands. For adequately small nearest-neighbor distances, two
discrete states move away from the bands to form surface states.
In contrast to the jellium model, these models emphasize the lattice aspects of a
linear chain of atoms and simplify the surface barrier. In fact, in any proper model,
it is the surface barrier that makes the electrons reflect at z = 0 and leads to forma-
tion of surface states. Lets now consider a linear 1DIM chain of atoms periodically
spaced in the z-direction starting from the surface. We make the assumption that
the formation of the surface has no e↵ect on the interatomic distance. In the light of
formation of dimmers at the surface as a result of reduced coordination and surface
reconstruction, we know that this assumption is unrealistic. The second simplifica-
tion is the modelling of the potential. We consider the step function at z = 0, but
for the chain of ion cores, we consider a weak and smoothened periodic potential:
V (z) = Vo + 2Vg cos gz, where g = 2⇡/a. Hence g is the reciprocal lattice vector
of the chain. The solution of the Schrödinger equation
✓ ◆
h̄2 d2
· + V (z) ' = E', (2.43)
2m dz 2
using the screened ion-core potentials V (z) and neglecting the electron-electron
interactions,
q
leads to 'vac,1 = e+kz and 'vac,2 = e kz on the vacuum side with k =
2m(V0 E)/h̄. The solutions must be finite, and we remain with 'vac,2 = e kz ,
because otherwise ' ! 1 for large values of z.

Figure 2.9: The periodic potential V (z) for 1DIM semi infinite lattice
with a step at the surface, z = 0. The dashed curve is more realistic.

In the bulk the solutions have the Bloch form 'k (z) = uk eikz because the po-
tential is periodic: V (z + a) = V (z). We may assume for metals with nearly-free
electrons that the potential is weak, and we expand V (z) around its average value
V0 : X
V (z) Vo = Vg eigz (2.44)
g

Away from the zone boundaries we have in units of h̄/2m:


Vg ei(k g)z
'k = eikz + (2.45)
k 2 (k g)2
CHAPTER 2. ELECTRONIC STRUCTURE 32

and
|Vg |2
Ek = k 2 + , (2.46)
k2 (k g)2
where Ek is measured relative to Vo .
Near the zone boundary, k ⇠ |k g|, and above equations are not valid any
more. One has to use degenerate perturbation theory with

'k = ↵eikz + ei(k g)z


. (2.47)

The coefficients can be found as


⇣ ⌘
(k 2 E)↵ + Vg = 0 and Vg ↵ + (k g)2 E = 0. (2.48)

Eq. 2.48 leads to


r !
1 2 ⇣ ⌘2
2
Ek = k + (k g) ± k2 (k g)2 + 4|Vg |2 . (2.49)
2
with the wave functions near the band gap:
E k2
'k = eikx + ei(k g)z
. (2.50)
Vg

At the zone boundary, k = 12 g, thus,


1 g V i g2 z
E± = ( g)2 ± |V | and '± = e i 2 z ± e . (2.51)
2 |V |
The wave functions are standing waves determined by the sign of V :
Energy V >0 V <0
E+ cos 12 gx sin 12 gx
E sin 12 gx cos 12 gx
In Eq. 2.49 we introduce k 0 which is measured relative to the zone boundary: k =
1
2
g + k 0 and we obtain
1 q
2 2
E = E(k 0 ) = ( g)2 + k 0 ± Vg2 + g 2 k 0 2 (2.52)
2
For positive values of k 0 2 this equation is compatible with the standard solution.
The novelty about this equation is that it has real solutions E even for k 0 2 < 0,
that is there are real solutions for Vg2 /g 2 < k 0 2 < 0, namely in the energy gap.
Figure 2.10 illustrates this situation, where the Brillouin zone boundary k 0 = 0
corresponds to the energy gap. Schrödinger equation has acceptable solutions also
for imaginary values of k 0 : 0 < |k 0 | < |Vgg | . This region of imaginary k 0 is shown
in the figure as dashed line. These solutions are not valid for the bulk, because for
|z| ! 1 the values diverge5 .
5
V. Heine, Proc. Phys. Soc. 81, 300 (1962).
CHAPTER 2. ELECTRONIC STRUCTURE 33

Figure 2.10: Relevant energies in the energy gap.

E(k 0 2 ) is a continuous function of k 0 2 . For negative values there is one real


function E, and we introduce a complex wave vector k = g2 + iµ. Thus we have
results in the complex k-plane as shown in Fig. 2.11. In order to determine the
wave functions for the solutions in the band gap, we introduce Eq. 2.52 in Eq. 2.50
and use k 0 = iµ. It follows:
⇣ gx ⌘
'k = e µz cos + (2.53)
2
with |µ|max = | Vgg |.
Now, we have to perform wave matching for coupling the wave functions in
order to find acceptable solutions. This means that ', '0 and '0 /' (logarithmic
derivative) must both be continuous.
'0 g ⇣ gz ⌘
= µ tan + (2.54)
' 2 2
kz
At the vacuum side we had ' = 'vac,2 = e , and we find
'0 q
= k= 2m(Vo E)/h̄. (2.55)
'
We put the origin, zo = a/2 on a surface atom and obtain:
'0 g ⇣⇡ ⌘ q
= |µ| tan a + = Vo E (2.56)
' 2 2
The Eq. 2.56 must be solved graphically and yields solutions for Vg > 0.
CHAPTER 2. ELECTRONIC STRUCTURE 34

Figure 2.11: Energy bands in complex k -space. Ref. [5].

Figure 2.12 shows the surface state decaying both towards vacuum and bulk.
The decay towards bulk is determined by eµz (z < 0). At the edge of the gap µ ! 0,
and the surface state reaches relatively deep into the bulk. At the mid gap µ = µmax
and the surface state decays within a few atomic distances.

Figure 2.12: The atomic potential of the 1DIM linear chain and the surface
state at the surface. The evanescent wave for z < 0 is called the Shockley state.

Even the simple model, we have used, has shown the existence of electronic
states localized at the surface. In the real case the problem has to be considered
CHAPTER 2. ELECTRONIC STRUCTURE 35

in 3DIM. We introduce the notion of surface-projected band structure and surface


Brillouin zone. The wave vector can be split into the components k? perpendicular
to the surface, k? = kz , and kk in the surface, kk = kxy . In Fig. 2.13 we see on the
right-hand side a surface that represents E(kk ). The dashed area is the projected
bulk band structure at the surface. Every energy can be identified with a bulk
state in 3DIM k-space. The lines at the not-dashed areas stand for the surface
states, showing some dispersion in the k? -space, which is termed a band of surface
states. In the figure, the lower band of surface states mixes with a bulk state,
depicted in the shaded circle, and produces an unusually large intensity observable
in experiments. These states are called surface resonances. On the left-hand side
of the figure, we have the E(k? ) plane. At the lower energy gap far right, there is
one point in k? where the surface state exists. The encircled shaded area highlights
the formation of another gap, namely a hybridization gap. Hybridization of wave
functions occurs in order to avoid band crossing. Also in the hybridization gap
surface states may exist as indicated in the upper gap of E(kk ) plane.

Figure 2.13: Projected bulk band structure at the surface of a metal.


From AZ.

In the following chapter, we will introduce photoelectric emission as a standard


tool to investigate the bulk and surface electronic structure of matter. There has
been two Nobel prizes for photoelectric emission work, A. Einstein (1921) and K.
Siegbahn (1981), which shows the impact of the method on our todays understand-
ing of the electronic structure of matter.
Chapter 3

Photoelectric Emission

There are few emission mechanisms that promote bound electrons of solids into
free space, like the thermionic emission, secondary-electron emission, Auger elec-
tron emission, or field emission. An important tool is the photoelectric emission,
that provides the basic, direct, and most relevant information about the material
and the physics of the emission process. Photoemission is one of the most elaborate
and costly techniques, because one needs a synchrotron-radiation source, compli-
cated monochromator, ultra-high vacuum, etc. Still the method is widely employed
because of the wealth of information we can extract from the results. The first
question is why we would prefer photoemission over other spectroscopic tools. The
answer is that in a photoemission experiment we detect a photoelectron, and we
know its exact history from excitation to detection process. A photon with energy
h⌫ is absorbed leading to an electronic excitation from an initial state i with an
energy Ei to a final state f with an energy Ef . This process may help us learn, e.g.,
detailed information about the band structure of the solid. Electrons are excited
somewhere in the solid and move to the surface. The analysis contains information
about electron scattering processes. We may also learn about the structure of atoms
at the surface. Electrons which arrive at the surface escape into vacuum if their
energy overcomes the potential barrier, the work function. This process gives the
opportunity to study surface chemical processes.

3.1 Photoemission Process


3.1.1 Optical Excitation: Conservation rules
Photoemission process starts with optical excitations. A quantum of light h⌫,
photons, is absorbed by an electron such that Ef Ei = h⌫. It is essential that
the energy transfer is definite and the photon transmits all its energy. This is
not the case in excitation processes using electrons or ions. The ground state is
associated with N electrons, and the final state with a system of (N 1) electrons
and a photoelectron. We observe two di↵erent cases. Often, h⌫ is absorbed by

36
CHAPTER 3. PHOTOELECTRIC EMISSION 37

one electron and other (N 1) electrons remain in their initial states which they
have occupied prior to the excitation. We may call these spectator electrons. This
case is the subject of one-electron approximation. If the spectator electrons are
perturbed during optical excitation process, some of the excitation energy remains
in the (N 1) electrons. We have correlated electrons.
The absorption process is illustrated in Fig. 3.1 in an E(~k) diagram where the
energy conservation is evident. The photon energy is E = h⌫ and the wavelength
is related to the frequency ⌫ over speed of light c = ⌫ . Using appropriate units,
we find
12400
E(eV ) = (3.1)
(Å)
Thus 10 eV photons have a wavelength of 1240 Å. The momentum of these photons
is p = h̄k = h/ = h/1240. For a conduction electron, on the other hand, we may
write pe = h̄kB = h/a = h/2, using the lattice constant a. kB is defined by the
Brillouin zone.

Figure 3.1: Plot of energyE versus wave vector k for the excitation of the
initial state i to f by absorbing a photon h⌫. Since the momentum of
the photon is negligibly small optical transitions are virtually vertical.

These numbers show that photons have energy, but no momentum, and for a
photoexcitation process using h⌫ = 10 eV momentum transfer is negligible: p~ =
h̄ ~k ⇡ 0. This momentum conservation is expressed by vertical transitions in
Fig. 3.1.
Not all energy and momentum conserving transitions take place. There are
other restrictions of optical transitions due to selection rules. The simplest case is
illustrated in an atom with its spherically symmetric wave functions. The matrix
element for optical transitions is given by M = h f |V | i i. For linearly polarized
~ is parallel to, e.g., x. Then, V / x, which means that
light the electric field vector E
CHAPTER 3. PHOTOELECTRIC EMISSION 38

it has negative parity and changes sign for inversion (x ! x, etc.). Depending on
the state of the wave function, i.e., s, p, d, . . . the spherical functions may have either
R
parity: (~r) = ( ~r) or (~r) = ( ~r). Then, the matrix element M = fx i
gives zero, if f and i have the same parity, and it is nonzero, if f and i have
di↵erent parities. As a result, s p transition is allowed, but s d or s f transition
is forbidden. Practically, optical transitions take place if ` = ±1.
In solids, we have mixed states, and their symmetry is not necessarily spheric.
Yet, there are strict selection rules for optical transitions that allow us to draw
conclusions on the symmetry of the wave functions i and f . This information is
relevant for surface chemistry and catalysis.
In summary, optical excitations are selective with respect to energy, momentum,
and selection rules. This is a general issue valid for optics. The advantage of
photoemission is that emitted photoelectrons are detected and thus the final state
is observed and can be localized in the Brillouin zone.

3.1.2 Energy Conservation


In the one-electron approximation only one electron changes its state as a result of
photon absorption, and the other (N 1) electrons keep their states. Thus, we can
write EB = h⌫ Ekin .

Figure 3.2: Plot of electron energy for the photoemission process in one-
electron approximation. EB is the binding energy and Ekin the kinetic
energy of the photoemitted electron, the work function of the metal.

The binding energy EB of the electron is defined in the absolute energy scale and
measurable in the ground state. This is only possible if the photoelectron does not
undergo any energy dissipating collisions during the emission process. In summary,
CHAPTER 3. PHOTOELECTRIC EMISSION 39

the one-electron approximation can be applied for cases where the hole state does
not interact with other electrons of the material. This is true for the conduction
electrons with extended character of several metals. One-electron approximation
cannot be applied if the hole state is only partially screened by the electron gas,
e.g., for core states in atoms and molecules.

3.1.3 Conservation of Momentum


In a photoemission experiment one measures the kinetic energy of electrons along
the direction of emission, the measurement mode is called angle- and energy-resolved
measurement.

Figure 3.3: We measure Ekin of the photoemitted electron under the


appropriate emission angle. The relationship between the measured
2
quantity and the vacuum momentum p~vac is Ekin = p~vac /(2m).

Now we need a relationship between the vacuum and crystal momentum of


the electron in the solid before the excitation process. We assume that optical
absorption causes transitions between Bloch states. The momentum of the excited
~
electron in the crystal is given by (~r) = u~k (~r)eik·~r with u~k (~r + ~a) = u~k (~r) owing to
the translational invariance (periodicity) of the crystal. Fourier expansion yields
X ~ ~
(~r) = aG~ i ei(k+Gi )·~r (3.2)
~i
G

with G~ i the reciprocal lattice vector. If K ~ i is the total k-vector and ~k the k-vector
in the reduced scheme, then K ~ i = ~k + G
~ i . Since there is only invariance with respect
to translations by a the lattice vector, K ~ k is conserved. Hence it is a good quantum
number.
The conserved quantities can be expressed as
~ kcrystal = ~kkcrystal + G
p~kvac = K ~k (3.3)
CHAPTER 3. PHOTOELECTRIC EMISSION 40

The translational symmetry is broken normal to the surface and hence momen-
tum is not conserved
~ ?crystal
p~?vac 6= K (3.4)
~2
During the photoemission process, electrons have a kinetic energy h̄2 K ?crystal /(2m)
normal to the surface, while the kinetic energy after the escape process reduces to
h̄2 p~?crystal
2
/(2m). During escape, forces act on the excited electron and it slows
down; the perpendicular component of momentum is not conserved. In summary,
in an angle-resolved photoemission process we can determine the binding energy EB
of the electron before the process in its ground state and its K ~ k in the crystal. So,
just using the conservation rules we can have a detailed picture of photoemission
without a microscopic description. This is sufficient in order to experimentally map
the dispersion relation, the band structure, of solids.

3.1.4 Three-Step Model


For complicated quantum-mechanical problems, as it is the case in photoemission,
it is customary to divide the process in smaller, more manageable subprocesses.
Accordingly photoemission is imagined as a three-step process.1

Figure 3.4: Photoelectrons are excited in the solid. Subsequently,


they travel to the surface. In the third step, they overcome the
surface barrier and are emitted into vacuum. Ref. [1].

1
W.E. Spicer, Phys. Rev. 112, 114 (1958).
CHAPTER 3. PHOTOELECTRIC EMISSION 41

Figure 3.4 depicts these processes. We have dealt with optical excitations and
energy conserving processes. This is the first step. The second step is the motion
of photoexcited electrons to the surface. The absorption coefficient of photons are
in the order of µm. Yet the photoexcited electrons travel much shorter distances
without inelastic collisions. This e↵ect makes photoemission spectroscopy a surface
sensitive one. It also introduces into the measured spectrum electrons that have
lost energy. These are undesirable, because we can apply the conservation rules
only on “primary” electrons which are emitted without energy loss. The third step
is the escape of electrons into vacuum overcoming the potential barrier, the work
function . For the “primary” electrons the momentum parallel to the surface is
conserved and we can track back the electrons in order to obtain information on
the initial states.

3.2 Applications
3.2.1 Band Structure
Now we show a historical experiment. Recall that in angle-resolved photoemission
experiments we know the initial- and final-state energies exactly. The emission
angle helps us find the crystal momentum. Hence, we can locate the transition in
the Brillouin zone in the reduced-zone scheme.

Figure 3.5: Valence-band dispersions E(~k) of GaAs along major symmetry direc-
tions. Dashed curves are theoretical dispersion curves for valence bands. Ref. [2].
CHAPTER 3. PHOTOELECTRIC EMISSION 42

Figure 3.5 shows one of the first results investigating the band structure. Data
have been obtained in a wide range of ~k so we can easily compare them with com-
putations.2 The parabolic shape of the sp states at the point are well reproduced.
It is worth noting that the ideas of Bloch states and energy-band dispersions have
been used since the advent of solid state physics. Yet the experimental verification
of these ideas had to await until about 1977.

3.2.2 Surface States

Figure 3.6: Experimental energy-distribution curves for the Cu(111) sp


surface state for several angles near normal emission. Ref. [3].

Surface states exist in SC’s and metals in the bulk band gap. For a Cu crystal
there exists a band gap along the L line. The binding energy of these states
at the point, the center of the 2DIM Brillouin zone, is 0.4 eV relative to EF ,
and they disperse parabolically upward as the momentum parallel to the surface is
2
T.-C. Chiang et al., Phys. Rev. B 21, 3513 (1980).
CHAPTER 3. PHOTOELECTRIC EMISSION 43

increased.3 Some typical high-resolution spectra are shown in Fig. 3.6.

Figure 3.7: Energy-dispersion relation for the surface state. The solid
curve is a parabolic least-square fit. The shaded region is the projected
bulk continuum of states. Ref. [3].

Figure 3.7 shows the dispersion relation E(~kk ) of the surface state along with
the projected bulk continuum. The surface state enters the bulk continuum just
above EF . A least-squares analysis of the E(~kk ) points produces the fit E(kk ) =
8.25kk2 0.389, with E in eV and kk in Å 1 . This way we can find the e↵ective
mass of surface-state electrons.

3.2.3 Spin Spectroscopy


In a ferromagnet the spin moments are aligned in the initial state along a quan-
tization axis which we may define. We measure in photoemission electron spin
polarization (ESP) of photoemitted electron, which is defined as ESP = (N "
N #)(N " +N #) 1 where N " (N #) represents the number of electrons with
spin parallel (antiparallel) to the quantization axis. The photoemission process is
faster than the spin-flip process. Therefore, the spin direction is another conserved
quantity in photoemission, and a measurement of ESP gives information about the
alignment of spins in the initial state.
Figure 3.8 shows on the left-hand side the spin-resolved density of states in Ni.
Below the Curie temperature (TC ), the number of electrons with the spin moment
parallel to magnetization direction is larger than those with antiparallel spins. This
is because of the exchange energy J, namely electrons with parallel spins avoid each
3
S.D. Kevan, Phys. Rev. Lett. 50, 526 (1983).
CHAPTER 3. PHOTOELECTRIC EMISSION 44

other. The exchange energy itself is in competition with the kinetic energy, while
the potential energy is reduced because of correlation. If J is sufficiently large and
similarly D(EF ) the density of states at EF , then the so-called Stoner criterium
for ferromagnetic order J · D(EF ) > 1 is fulfilled. This means that the cost in
kinetic energy is small for transporting an electron from "band into #band. Thus
Ni is a strong ferromagnet. D(EF ) is apparently high and consists of minority-spin
electrons.

Figure 3.8: (Left) The exchange-split 3d bands of ferromagnetic Ni. The density of
states consists of minority-spin electrons at EF . (Right) Energy-distribution curves
of Ni proving the existence of the exchange splitting in the 3d bands. Ref. [4].

The photoemission results near the EF provide a test for these above ideas. On
the right-hand side of the figure we have spin-resolved energy distribution curves
obtained from Ni(110) at di↵erent temperatures below TC . Indeed, the electrons
with antiparallel spins dominate at EF , and the two exchange-split bands are well
resolved. The splitting is more pronounced as the temperature is lowered.4
ESP is determined in a so-called Mott scattering of electrons at energies above
100 keV. It makes use of the relativistic e↵ect that there is a left-right intensity
asymmetry for large scattering angles in the scattering plane proportional to ESP.
To avoid multiple scattering, the experiments are done either at atomic beams or
at extremely thin films of heavy metals, preferably Au. Thus, the efficiency of
4
E. Kisker et al., Phys. Rev. Lett. 43, 966 (1979).
CHAPTER 3. PHOTOELECTRIC EMISSION 45

Mott scattering is around 10 4 , which presents a drawback of the spin-resolved


experiments.

3.2.4 Localized States


Nearly-free electron states are well accounted for by the one-electron approximation.
Localized electrons, on the other hand, are confined near the atomic core where a
positive charge due to a core state is e↵ectively screened by the conduction electrons.
As a results, localized electrons feel the hole state in photoemission as an additional
positive charge, and the orbitals adjust themselves depending on the interaction
with the hole state. Most often there are several holes for a given h⌫, and we
observe several di↵erent final states, called the multiplets.

Figure 3.9: Photoemission spectrum from Nd near EF and a calcu-


lation of the multiplet structure of the final state. Ref. [5].

The 4f electrons of Rare Earth metals are near EF and show an atomic behavior.
They do not participate in chemical bonding. Neodymium has in the ground state
4f 3 configuration. To find the symmetry in the ground state we apply the Hund’s
rules. Three electrons result in a total orbital moment of L = 6 as the boxes
show. Total spin moment is 3/2. Since the f band is less than half full, the total
J = L S = 9/2. The resulting term symbol is then 4 I9/2 .
CHAPTER 3. PHOTOELECTRIC EMISSION 46

3 2 1 0 -1 -2 -3

" " "

Photoemission leaves behind the f band with one hole, in the 4f 2 configuration.
As seen from Figure 3.9 there are 13 final-state multiplets. This multiplet structure
dominates the photoemission spectrum. We observe a huge emission intensity due
to the 4f electrons at an energy which corresponds to a binding energy of about 5
eV. This is in fact not the binding energy, but the correlation energy we have to
pay additionally to remove an electron from the 4f band.

Figure 3.10: The periodic table for the Rare Earth elements and photoe-
mission spectra of 4f n 1 state in Gd, Tb, Dy, and Sm. Ref. [5].

Figure 3.10 shows on the left-hand side the multiplet manifolds of the 4f 7 , 4f 8 ,
and 4f 9 final states in Gd, Tb, and Dy, respectively.5
In localized systems, the occupation is always an integer number, like in atoms,
because these states do not mix with others. The binding energy of the localized
states is a function of the occupation, as it is known from atoms. In SmB6 , Sm exists
in two di↵erent valencies, as Sm3+ and Sm2+ , because interestingly the energies of
5
J.K. Lang, Y. Baer, and P.A. Cox, J. Phys. F 11. 121 (1981).
CHAPTER 3. PHOTOELECTRIC EMISSION 47

4f n+1 and 4f n 5d state (n = 5) are almost the same. As a result the system
fluctuates between the two states. Photoemission is faster than the fluctuation
frequency and hence records both states in one single spectrum. The multiplets of
Sm ion with fluctuating valency are displayed on the right-hand side of the figure.

3.2.5 Resonant Photoemission


For the 3p 3d excitation in 3d transition metals there is a large overlap between
the 3p hole and the excited 3d electrons, and the Coulomb interaction between the
initial and final state is large. As a result, the independent electron description
breaks down, and the 3p 3d transition spectrum resembles that encountered in
isolated atoms. Thus, the observed structure in the emission spectrum is due to
the final-state multiplets arising from the 3p6 3dN ! 3p5 3dN +1 excitation, broadened
asymmetrically by resonant interaction between the di↵erent discrete configurations
and continua.6 The excited state rapidly decays via 3p5 3dN +1 ! 3p6 3dN 1 + e .
The 3p6 3dN 1 state can directly be reached by the excitation of a 3d electron.
Whenever there are two coherent channels leading to the same final state, resonances
are encountered between the discrete channels and continua. The emitted electron
carries information about the detailed processes.

Figure 3.11: Resonant photoemission in a gadolinium film grown on copper. Ref. [7].

These are atomic-like processes because a) d bands are narrow, b) there is a


large overlap between the states involved, c) the time scale for the decay of the
6
L.C. Davies and L.A. Feldkamp, Solid State Commun. 19, 413 (1976).
CHAPTER 3. PHOTOELECTRIC EMISSION 48

self-screened state is extremely short. These processes are also encountered in the
4d transition metals and 4f Rare Earth metals.
In Gd the 4d 4f transition energy is about 140 eV leading to the transition
4d 5s 5p6 4f 7 ! 4d9 5s2 5p6 4f 8 . By analyzing the photoemitted electron from dif-
10 2

ferent states of Gd at the photon energy of about 140 eV, we investigate di↵erent
decay channels of the excited state.7
Figure 3.11 displays photoemission spectra recorded at di↵erent photon energies
below and above the 4d 4f transition energy. The spectra are decomposed into
three regions, where we observe the 4f , 5p, and 5s resonances. The former is
displayed at the right-hand side at low binding energies up to 20 eV. In this energy
region we observe the direct recombination 4d9 5s2 5p6 4f 8 ! 4d10 5s2 5p6 4f 6 + e .
The strength of resonance emission is evident at the correct d f transition energy
as h⌫ is scanned.
Figure 3.11 shows in the middle panel spectra in the EB region of Gd 5p elec-
trons. Di↵erent spectra are obtained at various h⌫ scanning the d f transition
energy. Now we focus on the recombination 4d9 5s2 5p6 4f 8 ! 4d10 5s2 5p5 4f 7 + e as
observed in the enhancement of emission for the 5p states.
We observe finally the recombination of 5s states in the same figure on the left-
hand side panel. EB is tuned for the appropriate energy. The decay channel is
4d9 5s2 5p6 4f 8 ! 4d10 5s1 5p6 4f 7 + e , and the emitted electron carries information
about the intriguing processes. Note that the system behaves like an atom far from
the independent-electron approximation.

7
J.A. Scarfe et al., phys. stat. sol. (b) 171, 377 (1992).
Chapter 4

Crystal Structure

Until the discovery of the quasicrystalline state in 1984, solids were generally clas-
sified according to their structure as amorphous and crystalline. In amorphous
structures, the range over which translational and orientational correlations decay
to zero is finite. Hence the atomic structure is random. The window glass has an
amorphous, vitreous, structure. Glass is in fact a supercooled liquid in which the
viscosity is too large to permit atomic rearrangement towards a more ordered form.
An amorphous solid yet exhibits a considerable degree of short-range order (SRO)
in its nearest-neighbor bonds, but not the long-range order (LRO) of a periodic
atomic lattice.
Quasicrystals, also called quasiperiodic crystals, are a relatively new ordering
state of condensed matter. For their discovery,1 Shechtman received the Nobel
Prize for Chemistry in 2011. Shechtman and coworkers observed electron di↵rac-
tion patterns typical of icosahedral point-group symmetry from a rapidly solidified
Al6 Mn alloy. The existence of a di↵raction pattern indicates LRO in atomic posi-
tions, but the presence of the fivefold-symmetry axes, forbidden in crystalline solids,
represents the paradox presented by quasicrystals.
This paradox can very simply be transformed to a tiling problem in 2DIM,
namely it can be reduced to the question how to completely cover a surface with
fivefold-symmetric tiles. M. Gardner proposed several possibilities upon which the
so-called Penrose tiles become popular,2 Penrose tilings have many amazing proper-
ties: There are infinitely many Penrose tilings, i.e., there are infinitely many distinct
tilings admitted by oblate and prolate rhombi. Not only are Penrose tilings ape-
riodic, they have no translational symmetry. Any finite region of a Penrose tiling
occurs infinitely many times in that tiling. In fact, any finite region that occurs in
a Penrose tiling appears infinitely many times in every Penrose tiling. Yet, no part
of the tiling repeats itself in the sense of crystallography. Every Penrose tiling by
thick and thin rhombi as shown in Fig. 4.1 can be obtained by projecting a 5DIM
1
D. Shechtman, I. Blech, D. Gratias, and J.W. Cahn, Phys. Rev. Lett. 53, 1951 (1984).
2
Martin Gardner, Extraordinary nonperiodic tiling that enriches the theory of tiles, Mathemat-
ical Games, Scientific American, January, 1977, p. 110 – 121.

49
CHAPTER 4. CRYSTAL STRUCTURE 50

Figure 4.1: A Penrose tiling consisting of oblate and prolate rhombi


(diamonds) that cover the surface area completely. Ref. [2].

cubic structure onto a 2DIM plane cutting through 5DIM space at an irrational
angle.3
The translation order in quasicrystals is quasiperiodic rather than periodic, but
imposing LRO which makes possible the existence of di↵raction patterns. These
patterns consist of a set of Bragg peaks which densely fill the reciprocal space
without a minimum spacing between peaks due to missing periodicity.
One year later Bendersky4 discovered decagonal quasicrystals. In this class
of quasicrystals the atomic distribution is quasiperiodic within planes, while the
planes are stacked periodically along the tenfold-symmetry axis. They are 2DIM
quasicrystals and represent an intermediate state between crystalline and quasicrys-
talline materials.
Quasicrystals are all binary, ternary, or quaternary alloys whose great majority
is Al based. The origin of the existence of stable quasicrystal phases remains in
question. No proven explanation clarifies why a material favors crystallographically
forbidden rotational symmetry and translational quasiperiodicity when at nearby
chemical compositions it forms more conventional crystal structures. Besides their
extraordinary structures and symmetries, quasicrystals possess unusual physical
properties.
3
M. Arık and M. Sancak, Pentaplex Kaplamalar , TÜBİTAK Popüler Bilim Kitapları 254,
Ankara, 2006.
4
L. Bendersky, Phys. Rev. Lett. 55, 1461 (1985).
CHAPTER 4. CRYSTAL STRUCTURE 51

4.1 Chemical Bonding


All of the mechanisms which cause bonding between atoms derive from electrical
attraction and repulsion. The di↵erent strengths and di↵erent types of bond are
determined by the particular electronic structures of the atoms involved. The weak
van-der-Waals bond provides a universal weak attraction between closely spaced
atoms and its influence is overridden when the conditions necessary for ionic, cova-
lent, or metallic bonding are also present.
The energy gained by forming a stable bonding arrangement compared to iso-
lated atoms is known as the cohesive energy, and ranges in value from 0.1 eV/atom
for solids which is typical only to the weak van-der-Waals bond to 7 eV/atom or
more in some covalent and ionic compounds and some metals. The cohesive energy
constitutes the reduction in potential energy of the bonded system compared to
separate atoms minus the additional kinetic energy which the Heisenberg uncer-
tainty principle tells us must result from localization of the nuclei and outer shell
electrons.
In covalent bonding the angular placement of bonds is very important, while in
some other types of bonding a premium is placed upon securing the largest possible
coordination number (= number of nearest neighbors). Such factors are clearly
important in controlling the most favorable 3DIM structure. For some solids, two
or more quite di↵erent structures would result in nearly the same energy, and a
change in temperature or hydrostatic pressure can then provoke a change from one
allotropic form of the solid to another.

4.1.1 The van-der-Waals Bond


Van-der-Waals bonding occurs universally between closely spaced atoms, but is
important only when the conditions for stronger bonding mechanisms fail. It is a
weak bond, with a typical strength of 0.2 eV/atom, and occurs between neutral
atoms and between molecules. The weak attractive forces between molecules in a
gas lead to an equation of state which represents the properties of real gases.
A neutral atom has zero permanent electric dipole moment, as do many mole-
cules; yet such atoms and molecules are attracted to others by electrical forces. The
zero-point motion, which is a consequence of the Heisenberg uncertainty principle,
gives any neutral atom a fluctuating dipole moment whose amplitude and orienta-
tion vary rapidly. The field induced by a dipole falls o↵ as the cube of the distance.
Thus if the nuclei of two atoms are separated by a distance r, the instantaneous
dipole of each atom creates an instantaneous field proportional to r 3 at the other.
The potential energy of the coupling between the dipoles which is attractive is then
Eatt = Ar 6 .
Eatt would reach 10 eV if r could be as small as 1 Å. However, a spacing this small
is impossible because of overlap repulsion. As the interatomic distance decreases,
the attractive tendency begins to be o↵set by a repulsive mechanism when the
CHAPTER 4. CRYSTAL STRUCTURE 52

Figure 4.2: Sum of the attractive and repulsive terms give the total potential energy
in a stable van-der-Waals bond at an internuclear distance of ro . From JSB.

electron clouds of the atoms begin to overlap. This can be understood in terms of
the Pauli exclusion principle, that two or more electrons may not occupy the same
quantum state.
The variation of repulsive energy with interatomic spacing can be simulated in
terms of a characteristic length ⇢. The total energy can be written as
6 r/⇢
E= Ar + Be (4.1)

which is shown in Fig. 4.2. The strength of the bond formed and the equilibrium
distance ro between the atoms so bonded depend on the magnitudes of the pa-
rameters A, B, and ⇢. Since the characteristic length ⇢ is small compared to the
interatomic spacing, the equilibrium arrangement of minimum E occurs with the
repulsive term making a rather small reduction in the binding energy.
There are no restrictions on bond angles, and solids bound by van-der-Waals
forces tend to form in the (close-packed) crystal structures for which an atom has
the largest possible number of nearest neighbors. This is the case, for example,
in the crystals of the inert gases Ne, Ar, Kr, and Xe, all face-centered-cubic (fcc)
structures, in which each atom has twelve nearest neighbors. The rapid decrease of
van-der-Waals attraction with distance makes atoms beyond the nearest neighbors
of very little importance.
Besides the solid inert gases, crystals of many saturated organic compounds and
also for solid H2 , N2 , O2 , F2 , Cl2 , Br2 , and I2 are examples of solids which are
bound solely by van-der-Waals forces. This group of solids have low melting and
boiling points. They are electrical insulators and are transparent for visible to far
ultraviolet light.
CHAPTER 4. CRYSTAL STRUCTURE 53

4.1.2 The Covalent Bond


The covalent bond is an electron-pair bond in which two atoms share two elec-
trons. The result of this sharing is that the electron charge density is high in the
region between the two atoms. An atom is limited in the number of covalent bonds
it can make depending on how much the number of outer electrons di↵ers from a
closed-shell configuration, and there is a marked directionality in the bonding. Thus
carbon can be involved in four bonds at tetrahedral angles (109.5), and the char-
acteristic tetrahedral arrangement is seen in crystalline diamond. Other examples
of characteristic angles between adjacent covalent bonds are 105 in plastic sulphur
and 102.6 in tellurium.
The hydrogen molecule, H2 , serves as a simple example of the covalent bond.
Two isolated hydrogen atoms have separate 1s states for their respective electrons.
When they are brought together, the interaction between the atoms splits the 1s
state into two states of di↵ering energy, as sketched in Fig. 4.3. When the two
nuclei are very close together, the total energy is increased for both kinds of states
by internuclear electrostatic repulsion; but for the 1s state marked g , which has an
symmetric orbital wave-function, the energy is lowered, i.e., there is an attractive
tendency, for a moderate spacing.

Figure 4.3: Variation of energy with r for the neutral hydrogen molecule.
The bonding ( g ) and antibonding ( u ) orbitals, accommodate two elec-
trons with antiparallel and parallel spins, respectively. g and u represent
gerade and ungerade which stand for even and odd. From JSB.

This symmetric g solution requires that the electron charge density e 2 be


concentrated in the region between the two nuclei. The requirement of the Pauli
principle that total wave functions combine in an antisymmetric manner is satisfied
if the g 1s state is occupied by two electrons with antiparallel spins.
CHAPTER 4. CRYSTAL STRUCTURE 54

The alternative u state would have to be occupied by two electrons with parallel
spins in order to conform with the Pauli principle, but as observed in Fig. 4.3, this
state is an antibonding (repulsive) one at all distances. This is unimportant for
H2 , since the g state can accommodate the only two electrons in the system and a
strong bond results.
Some of the classes of covalently bonded materials are: 1. Most bonds within
organic compounds. 2. Bonds between pairs of halogen atoms (and between pairs
of atoms of hydrogen, nitrogen, or oxygen) in the solid and fluid forms of these
media. 3. Elements of Group VI (such as the spiral chains of tellurium), Group
V (such as in the crinkled hexagons of arsenic), and Group IV (such as diamond,
Si, Ge, ↵-Sn). 4. Compounds obeying the 8 N rule (such as InSb) when the
horizontal separation in the Periodic Table is not too large.
It is often found that valence-bonded solids can crystallize in several di↵erent
structures for almost the same cohesive energy. The energetically most favored
structure can be displaced from its prime position by a change of temperature or
pressure, resulting in the situation known as allotropy or polymorphism. Thus ZnS
can exist either in a cubic form (zinc blende) or as a hexagonal structure (wurtzite).
The coordination of nearest neighbors is the same for zinc blende and wurtzite; it
is the arrangement of second-nearest neighbors which creates a very slight energy
di↵erence between the two structures.
Allotropic conversion is provided by tin, which is stable as a gray semimetal
(↵-Sn) below 17 C, crystallizing in the diamond lattice with four tetrahedrally-
located bonds. Temperatures above 17 C (or application of pressure even below
that temperature), cause a conversion to a much more dense white metallic form
( -Sn) with a tetragonal structure in which each atom has six nearest neighbors.
Covalently bonded materials are hard and have high melting points. They are
electrical insulators or SC’s. They absorb light above a characteristic threshold.

4.1.3 Covalent–van-der-Waals Structures


This combination of bonding mechanisms is found in materials such as solid hydro-
gen, in which each pair of atoms is internally covalently bonded and van-der-Waals
bonds create a “molecular crystal”. The same principles apply to most organic
solids.
An example of another kind of covalent-residual bonding is provided by tel-
lurium, in which successive atoms in each spiral chain are covalently bonded. The
forces between chains are much weaker and are probably little more than van-der-
Waals attraction. Consequently, tellurium has a low structural strength and is
anisotropic in all its mechanical, thermal, and electronic properties.
Similarly, in graphite (Fig. 1.11) carbon atoms are arranged in hexagons in each
layer, so that three of the four outer shell electrons from each atom are used in
valence bonds within the layer. The fourth electron is free. The interlayer spacing
CHAPTER 4. CRYSTAL STRUCTURE 55

is large, with essentially only van-der-Waals attraction. Thus the planes can slide
over each other very easily, the property which makes graphite useful as a “solid
lubricant.” The same considerations apply in MoS2 .

4.1.4 The Ionic Bond


An ionic crystal is made up of positive and negative ions arranged so that the
Coulomb repulsion between ions of the same sign is more than compensated for by
the Coulomb attraction of ions of opposite sign. Thus, the main contribution to
the binding energy of ionic crystals is electrostatic and called the Madelung energy.
The alkali halides such as NaCl are typical members of the class of ionic solids;
NaCl crystallizes as Na+ Cl . Electron transfer from Na to Cl occurs to such a
major extent because the ionization potential Ie of the alkali metal is small (work
eIe must be done to convert Na into the cation Na+ with a closed electronic shell
configuration), whereas the electron affinity Ea of the halogen is large. (Energy Ea
is provided when Cl receives an electron and becomes the anion Cl , also with a
closed shell configuration.)

Figure 4.4: The energy of a Na+ Cl molecule as a function of internuclear


spacing r. From JSB.

When a Na+ ion and a Cl ion approach each other in the absence of any
other atoms, the attractive Coulomb energy at internuclear separation r is ECoul =
e2 (4⇡✏o r) 1 since the closed-shell electronic charge distributions are spherically
symmetrical. The approach distance is limited by repulsion when the closed-shell
electron clouds of anion and cation overlap, in consequence of the Pauli principle.
The energy associated with repulsion varies rapidly with separation, as noted in
CHAPTER 4. CRYSTAL STRUCTURE 56

connection with van-der-Waals bonding; two approximate ways of describing it are


Erep = Ar n with (n ⇡ 12) or Erep = Be r/⇢ . The stable bond length between
Na+ and Cl will be the value for which the total energy is a minimum. Figure 4.4
shows this minimum.
Materials with ionic bonding often dissociate upon heating. They are electrical
insulators at low temperatures, while ionic conduction is observed at high temper-
atures. They absorb light above an intrinsic photon energy.

4.1.5 The Hydrogen Bond


A hydrogen atom, having one electron, can be covalently bonded to only one atom.
However, the hydrogen atom can involve itself in an additional electrostatic bond
with a second atom of highly electronegative character, such as fluorine, oxygen,
or to a smaller extent with nitrogen. This second bond permits a hydrogen bond
between two atoms or structures. The hydrogen bond is found with strengths
varying from 0.1 to 0.5 eV per bond.
Hydrogen bonds connect the H2 0 molecules in ordinary ice, a structure similar
to wurtzite in which there is a spacing of 2.76 Å between the oxygen atoms of
adjacent molecules. This is much more than twice the “ordinary” O-H spacing of
0.96 Å for an isolated water molecule. The molecules in ice can flip into a variety
of arrangements, the equilibrium one depending on pressure and temperature, and
numerous high pressure allotropic modifications of ice are known. Materials ex-
hibiting hydrogen bonding are electrical insulators and optically transparent. They
are easily polarizable.

4.1.6 The Metallic Bond


Metallic structures have typically large internuclear spacings and prefer lattice ar-
rangements in which each atom has many nearest neighbors. In many metals only
one electron per atom is involved in bonding. So, in a metallic solid we have a
widely spaced array of positively charged ion cores with a superposed electron gas
to give macroscopic charge neutrality. The wave functions of the electrons compris-
ing this gas overlap strongly and are therefore delocalized. The electrons which are
used for binding have di↵erent energies; the average energy per binding electron
is smaller than that of an isolated atom. Incomplete inner shells and correlation
e↵ects within the electron gas also contribute to binding.
In summary, there is large spacing between the atoms and large coordination
number. Metals are conductors for electricity and heat. They are opaque and highly
reflecting the visible light.5
5
J.K. Burdett, Chemical Bonding in Solids, Oxford University Press, New York.
CHAPTER 4. CRYSTAL STRUCTURE 57

4.2 Symmetry Operations


A crystal is an infinite 3DIM repetition of identical blocks each with the same
orientation. These building blocks, called the basis can be an atom, a molecule
or a group of atoms or molecules. The basis is located in the unit cell, a 3DIM
parallelepiped, which is translated in three directions to fill all the space. This
operation is called translational symmetry

T = pa + qb + r c (4.2)

with the translation vectors a, b, c that lie along three adjacent edges of the unit
cell and integers p, q, r. Translational symmetry means that the local arrangement
of atoms at the point r must be the same at any other point r0 : r0 = r + T. The
set of operations T defines a Bravais lattice. The points in a Bravais lattice that
are closest to a given point are called its nearest neighbors. Because of the periodic
nature of a Bravais lattice, each point has the same number of nearest neighbors.
This number is a property of the lattice, and is called coordination number.
The primitive basis is the smallest unit cell that is sufficient to characterize the
crystal structure. Figure 4.5 displays the conventional cell and the primitive unit
cell for the body-centered cubic (bcc) and fcc Bravais lattices. The volume of the
conventional cells are 4⇥ and 2⇥ larger than the corresponding primitive unit cells
for the fcc and the bcc lattices, respectively.

Figure 4.5: The conventional (large) and primitive (shaded) unit cells
for the a) face-centered and b) body-centered cubic Bravais lattice. The
primitive cell unit cell is defined by the vectors a, b, and c. From AM.

The most common choice of a primitive cell with the full symmetry of the Bravais
lattice is the Wigner-Seitz cell. It is a region of space around a lattice point that is
closer to that point than to any other lattice point. The Wigner-Seitz unit cell about
a lattice point is constructed by drawing planes as bisectors to lines that connect
the lattice point with others and taking the smallest polygon thus generated around
CHAPTER 4. CRYSTAL STRUCTURE 58

the lattice point. Figure 4.6 shows Wigner-Seitz cells for the bcc and fcc Bravais
lattices.

Figure 4.6: The Wigner-Seitz cell for the (a) bcc structure is a truncated
octahedron and for the (b) fcc structure a rhombic dodecahedron. From AM.

4.2.1 Bulk Structure


A lattice is invariant with respect to:
- Translational symmetry
- Reflection at a plane
- Rotation about an axis by 2⇡/n with n = 1, 2, 3, 4, or 6
- Inversion through a point (= rotation by 180 + reflection)
- Glide (= reflection + translation)
- Screw (= rotation + translation).
The possible ways in which the symmetry of the basis of atoms can be related to
the symmetry of the lattice result in the point groups. Further there are some plane
groups. In total, there are 230 3DIM space groups or structures. There are 14 3DIM
Bravais lattices, illustrated in Fig. 4.7. These can further be grouped into seven
crystal systems, including triclinic, monoclinic, orthorhombic, tetragonal, cubic,
trigonal , and hexagonal . The conventional unit cells for these seven systems arise
as a result of progressing distortions of cubic symmetry.
Any function defined for a crystal, such as the electron density, is bound to be
periodic, repeating itself with the same translation vectors as those that span the
lattice. Thus, f (~r + T~ ) = f (~r). Such periodic functions lend themselves easily
to Fourier transform. Under certain conditions, it becomes more advantageous to
deal with the Fourier components of such systems rather dealing with them in real
space. Besides the conventional definition of the Fourier transformation, we can
consider a plane wave eik·r . The set of all wave vectors k that yield plane waves
with the periodicity of a given Bravais lattice is known as its reciprocal lattice. We
can then characterize the reciprocal lattice with a set of k satisfying eik·r = 1. It is
appropriate to define primitive vectors a⇤ , b⇤ , and c⇤ for the reciprocal lattice such
CHAPTER 4. CRYSTAL STRUCTURE 59

that a · a⇤ = b · b⇤ = c · c⇤ = 2⇡ and a · b⇤ = a · c⇤ = c · b⇤ = 0. The reciprocal


lattice is spanned by the reciprocal lattice vector:
G = ha⇤ + k b⇤ + l c⇤ . (4.3)

Figure 4.7: Conventional unit cells for the fourteen possible 3DIM
Bravais lattices. From JSB.

We can construct the primitive vectors of reciprocal space such that they obey
2⇡b ⇥ c 2⇡c ⇥ a 2⇡a ⇥ b
a⇤ = ; b⇤ = ; c⇤ = (4.4)
a · (b ⇥ c) a · (b ⇥ c) a · (b ⇥ c)
The term on the denominator is numerically equal to the volume of the unit cell in
real space. For each G in Eq. 4.3, the Fourier transform of the periodic function
CHAPTER 4. CRYSTAL STRUCTURE 60

can then be written as


1 Z
f (G) = drf (r)e iG·r (4.5)
V V
where V is the cell volume.
The crystallographic directions are fictitious lines linking atoms, ions, or mole-
cules. Similarly, the crystallographic planes are fictitious planes linking nodes of a
crystal. Some directions and planes have a higher density of nodes. The behavior
of the crystal depends on the fact how dense the planes are. In particular, optical
properties, adsorption and reactivity, surface tension, and mechanical properties
closely depend on the number of atoms on a plane. For all these reasons, it is
important to determine the planes for which we first need a notation system.
A family of lattice planes is determined by three integers h, k, and l, the Miller
indices. They are written (hkl), and each index denotes a plane orthogonal to
a direction (h, k, l) in the basis of the reciprocal lattice vectors. By convention,
negative integers are written with a bar, as in 3̄. The integers are written in lowest
terms, i.e., their greatest common divisor should be 1. Miller index (100) represents
a plane orthogonal to direction h; index (010) represents a plane orthogonal to
direction k, and index (001) represents a plane orthogonal to l. The notation {hkl}
denotes the set of all planes that are equivalent to (hkl) by the symmetry of the
lattice. The corresponding notation [hkl] denotes a direction in the basis of the
direct lattice vectors instead of the reciprocal lattice; and similarly, the notation
< hkl > denotes the set of all directions that are equivalent to [hkl] by symmetry.

Figure 4.8: Three lattice planes and their Miller indices in a simple cubic
Bravais lattice. From AM.

The direction [hkl] is not generally normal to the (hkl) plane, except in a cubic
lattice. For this case, the lattice vectors are orthogonal and of equal length, similar
to the reciprocal lattice. Thus, in this common case, the Miller indices (hkl) and
[hkl] both simply denote normals/directions in Cartesian coordinates. For cubic
crystals with lattice constant a, the spacing d between adjacent (hkl) lattice planes
is:
a
dhkl = p 2 (4.6)
h + k 2 + l2
CHAPTER 4. CRYSTAL STRUCTURE 61

4.3 Determination of Bulk Structure


If we seek to obtain useful information about the a crystal structure we have to
use a tool with appropriate properties. Direct-space information is ruled out by
the long wavelength of radiation used, because we cannot resolve details finer than
the wavelength. Hence, we work with di↵raction techniques, using radiation of a
wavelength comparable with atomic dimensions. Di↵raction gives us information in
Fourier form, which can be analyzed in terms of the average spacings of lines and
planes of atoms in a solid, the angles between lines and planes, the symmetries of
point groups, and the location of particular species of atoms. There are three types
of excitations one uses for crystallographic investigations which have quite di↵erent
energies for wavelengths useful for di↵raction.
For di↵raction of neutrons at a crystal, we have to consider the wavelength
h h
= =p (4.7)
Mv 2M E
Thus neutrons with a wavelength of 1 Å move at a speed of 4000 m/s, and have a
kinetic energy of 0.08 eV. This energy is comparable with lattice vibrational energy,
and neutrons do not interact with the solid intensively.
Electrons with a wavelength of 1 Å have an energy of about 150 eV (Eq. 5.1).
An electron of this energy moves at a speed about 7 ⇥ 106 m/s, but interacts
with the electrons of the solid so excessively that it cannot penetrate the solid at all
without losing energy. Electrons in this energy range are used in low-energy electron
di↵raction (LEED). For bulk structural work, one uses high-energy electrons in a
transmission electron microscope (TEM).
The photon energy is E = h⌫ and the wavelength is related to the frequency
⌫ over speed of light c = ⌫ . Using appropriate units, we find for photons E(eV ) =
12400/ (Å), as derived in Eq. 3.1. For a wavelength of 1 Å the energy is about 12.4
keV easily accessible in laboratories. The suggestion to use x-rays for structural
studies first came from von Laue in 1912. He regarded the crystal as a 3DIM
di↵raction grating, and a di↵raction pattern provides information about the regular
arrangements of atoms. X-rays are up to date the principal source of information
about the crystallography of solids.

4.3.1 X-Ray Di↵raction


W.L. Bragg in 1913 proposed an expression for the geometric condition which must
be satisfied if waves are to be di↵racted by a parallel set of planes. The Bragg
di↵raction condition specifies that the angles of incidence and reflection are equal,
i.e., that reflection is specular. Bragg’s Law is not concerned with the arrangement
of atoms within a single reflecting plane.
We can see from Fig. 4.9, that the x-rays are reflected from planes, and con-
structive interference for specular reflection occurs for monochromatic radiation
CHAPTER 4. CRYSTAL STRUCTURE 62

whenever the Bragg condition is satisfied:

n = 2d sin ✓. (4.8)

Here, d is the spacing of the planes of atoms and n is an integer. All specularly
reflected components will be able to combine constructively in phase to form the
intensity of reflected radiation. This formulation is independent of the atomic ar-
rangement within each plane.

Figure 4.9: A Bragg (specular) reflection from a particular family of


lattice planes, separated by a distance d. Incident and reflected rays are
shown for the two neighboring planes. The path di↵erence is 2d sin ✓.

We note that Bragg’s results, presented in Eq. 4.8, are identical to those of von
Laue. If we analyze di↵raction in terms of initial and final wave vectors k and k0
in reciprocal space, we end up with Laue equations:

a· k = 2⇡h
b· k = 2⇡k
c· k = 2⇡l (4.9)

Eq. 4.9 ties together real- and reciprocal-space entities. It can be expressed more
elegantly as
k = G, (4.10)
which led Ewald in 1921 to suggest a geometric interpretation of the di↵raction
requirement. This is illustrated in Fig. 4.10. Given the incident wave vector ko , a
sphere of radius |k| is drawn to intersect the origin. The condition for di↵raction,
given in Eq. 4.10, is fulfilled whenever the sphere goes through a reciprocal lattice
point.
In considering the Bragg di↵raction of an x-ray beam by a single crystal, we must
note that Eq. 4.8 requires a suitable combination of d, ✓, and for constructive
CHAPTER 4. CRYSTAL STRUCTURE 63

Figure 4.10: The Ewald construction.

interference. Thus monochromatic x-rays directed on a crystal at an arbitrary angle


will usually not be reflected. Either x-rays of many wavelengths (white light) must
be used at a single angle of incidence (the Laue method ) , or monochromatic rays
must be allowed to encounter the crystal at a variety of angles. This is accomplished
mechanically in the rotating crystal , oscillating crystal , and Weissenberg methods,
while the Debye-Scherrer method uses a powdered sample so that every angle of
incidence is encountered for some of the crystallites.
The Laue method is very simple in concept and operation. A narrow beam of
unmonochromatized x-rays fall on a single crystal, as illustrated in Fig. 4.11. For
any wavelength satisfying the Bragg condition, a di↵racted beam will emerge. The
figure shows two positions in which a photographic plate can be placed to record
a set of di↵raction spots which is characteristic of the structure and orientation of
the crystal. The Laue method is rarely used for investigation of new structures.
Its main use lies in the rapid and convenient orientation of known crystals which
is achieved by tuning the angles necessary for perfect alignment of a major crystal
plane with the axis of the Laue camera.
The rotating-crystal method is a technique in which the crystal is rotated
through the angle ✓ about a fixed axis, while monochromatic x-rays are presented
CHAPTER 4. CRYSTAL STRUCTURE 64

Figure 4.11: A Laue flat plane camera. The Laue spots can be recorded
in either forward or backward direction in order to orient the crystal.
The crystal can be rotated about three orthogonal axes. From JSB.

to a single crystal. The variation in the angle brings di↵erent atomic planes into
position for reflection. The essential aspects of the experimental arrangement are
sketched in Fig. 4.12. The sample is rotated and a cylinder of photographic film,
placed coaxially with the rotation axis, records a spot whenever the Bragg condition
is fulfilled. The inclination of these spots to the direction of the incident beam has
a component of ✓ in the plane perpendicular to the cylinder axis.
Elaborations of the rotating crystal experiment to give much more information
about the crystal symmetry include the oscillating crystal method, the Weissenberg
method.
Instead of using a single x-ray wavelength and a time-dependent angle of inci-
dence, we may present a crystalline sample with every ✓ simultaneously. This is
the Debye-Scherrer technique of using a finely powdered crystalline sample in which
the crystalline orientations are random. Rays which for one crystallite or another
satisfy the Bragg condition emerge from the sample as a series of cones concentric
with the incident beam direction. Thus a photographic plate records a series of
concentric circles.
The scattering amplitude F of x-rays is generally given by
Z
i k·r
F = dV n(r)e (4.11)

If the di↵raction condition k = G is met,


Z
iG·r
FG = N dV n(r)e = N · SG . (4.12)
CHAPTER 4. CRYSTAL STRUCTURE 65

Figure 4.12: The crystal is rotated through ✓, while di↵raction intensity


is recorded at 2✓. From JSB.

The relative intensities of the reflections depend on the number, position, and elec-
tronic distribution of the atoms in the unit cell. The function SG is called the
structure factor . X
SG = fj e iG·rj (4.13)
j

is obtained by summation over the cell. This finally results in


X
S(hkl) = fj ei2⇡(pj h+qj k+rj l) (4.14)
j

with rj = pj a + qj b + rj c and G = ha⇤ + kb⇤ + lc⇤


Considering the basis of the bcc structure we obtain for the structure factor of
the bcc lattice
S = 0 when h + k + l = odd integer
S = 2f when h + k + l = even integer
As a consequence, di↵raction spectrum does not contain lines such as (100), (300),
(111), or (221).
Similarly, we obtain the structure factor for the fcc lattice
i⇡(k+l) i⇡(h+l) i⇡(h+k)
S(hkl) = f [1 + e +e +e ]

If all indices are even or odd integers, S = 4f . In an fcc lattice, no reflections can
occur for which the indices are partly even and partly odd. We thus have a tool to
di↵erentiate between bcc and fcc if we just know that the sample is cubic.
CHAPTER 4. CRYSTAL STRUCTURE 66

4.3.2 X-Ray Absorption Fine Structures


When a specimen is presented to electromagnetic radiation a portion of intensity is
absorbed causing interband transitions in the material. The probability Wif for a
transition of an electron from an initial state i with the energy Ei to a final state f
of the energy Ef as a result of absorption of a photon with the energy h⌫ is given
by the Fermi Golden Rule:

4⇡ 2
Wif = |hf |V |ii|2 (Ei EF + h⌫) (4.15)
h

The interaction with the electromagnetic field is described by the operator:


e e2
V = pA + A2 , (4.16)
mc 2mc2
with the vector potential A and the momentum p. In the dipole approximation
(k · r ⌧ 1) the absorption cross section is expressed as:
X
= 2⇡↵h! |hf |r~✏|ii|2 (Ei EF + h⌫) (4.17)
f

Here, ↵ is the fine structure constant, ~✏ the polarization vector and r is the position
of the particle. Using the density of the particles nc we obtain the absorption coef-
ficient µ = nc which is an experimentally observable quantity. The measurement
of µ and the interpretation of data demonstrate that geometric, electronic, and
chemical properties of the material under investigation are interrelated, and just
one experimental result contains all this information.
The shape of the absorption cross section is characteristic to the absorbing
atom. The sharp absorption edge reveals the binding energy of the particular inner
shell. X-ray absorption spectrum thus carries information about the identity of
atoms constituting the sample. If the atom is bound in a molecule, a solid, or even
placed in a liquid (cf. Fig. 4.13), thus having neighbors, one observes modulations
of the absorption coefficient beyond the absorption edge extending to a wide energy
range.6
Two regions in energy are arbitrarily di↵erentiated (cf. Fig. 4.13):
1. A range as narrow as about few 10 eV near the edge. Here, the structures are
due to the creation of localized excitations, variations of the density of states, as
well as many-body excitations, referred to as x-ray absorption near-edge structure
(XANES).
2. The higher energy range. The absorption coefficient shows a more or less simple
modulation, especially from atoms having a symmetric environment, called extended
x-ray absorption fine structure (EXAFS).
A convincing example is the absorption spectrum from the ferrocene molecule,
Fe(C5 H5 )2 , as shown in Fig. 4.14. In the molecule, the central Fe atom is surrounded
6
A. Di Cicco et al., Phys. Rev. B 54, 9086 (1996).
CHAPTER 4. CRYSTAL STRUCTURE 67

Figure 4.13: K -edge spectra of gaseous (lower curve), liquid (0.1 – 0.75 GPa), and
solid Kr as a function of pressure at room temperature. Ref. [6]. The vertical line
indicates the arbitrary separation of the spectra in XAN ES and EXAF S regions.

by 10 carbon atoms at exactly the same distance. Owing to identical contributions


from 10 neighbors, absorption measurements from the ferrocene molecule show en-
hanced sinusoidal oscillations.7
In crystalline materials the shape of the oscillations in the extended region can
look complicated because it consists of superposition of oscillations with di↵erent
periodicities (frequencies), as observed in Fig. 4.15 for Cu and Ni.
Additional observations include:
1. The energy distance between extrema decreases with growing distance between
the absorbing atom and its neighbors.
2. With increasing temperature thermal amplitude of atoms grow around their zero-
temperature position, and the EXAFS oscillations are e↵ectively damped. This
situation is shown for a Cu crystal in Fig. 4.15 (left) and clearly demonstrate the
link between the absorption of photons, an electronic process, and the position of
the absorbing atoms, a geometric e↵ect.
7
T.K. Sham and R.A. Holroyd, J. Chem. Phys. 80, 1026 (1983).
CHAPTER 4. CRYSTAL STRUCTURE 68

Figure 4.14: A ball-and-stick model of the ferrocene molecule and the absorption
coefficient µ. Ref. [7].

Historically, there are two di↵erent explanations for the observed oscillations:
a. Near-order theory claims that the observed structures originate from the matrix
element for absorption |Mif |2 , where |Mif | = hf |r · ✏|ii. This case describes the sit-
uation where excited electrons do not travel long distances in the crystal. Through
multiple scattering they are confined near the absorbing atom.
b. Far-order theory places the origin of the structures in the density of states
P
N (E) = f (Ei Ef + h⌫). If electrons are allowed to travel long distances after
excitation in the crystal, they sample information in a larger volume around the
absorbing atom.

Figure 4.15: The absorption coefficient of single-crystalline copper and nickel.


P. Aebi, unpublished
CHAPTER 4. CRYSTAL STRUCTURE 69

Today we know that both theories do correctly apply and the absorption spectra
are influenced by the matrix elements as well as the density of states that are
unoccupied prior to the excitation. The only crucial issue is that for kinetic energies
little over the Fermi energy, excited electrons travel long distances in the crystal.
They have a long mean free path, as displayed in Fig. 5.7. Hence the near-order
theory better describes the oscillations in the absorption coefficient. This is the
XANES regime. For excitations with higher photon energies, kinetic energy of
excited electrons are proportionally higher, the mean free path is shorter and the
absorption coefficient is rather influenced by the proximity of the absorbing atom.
This is the EXAFS regime.

Figure 4.16: Three possible high-symmetry positions of the absorbed


atoms on an fcc (100) surface.

The determination of site-specific information about adsorbed species on oth-


erwise clean surfaces is always a challenge due to the extremely small signal from
the adsorbates. Figure 4.16 shows schematically the ordered adsorption of foreign
atoms on a (100) surface of an fcc crystal. The amount of adsorbates are referred to
as 0.5 ML, that means that there are 1/2 foreign atoms per surface host atom. The
ordering of adsorbates shows that the periodicity in both directions is doubled with
an addition adsorbate atom in the middle of the adsorbate square thus formed. If
we take an adsorbate unit cell aligned with the substrate cell, this adsorbate con-
figuration is written as c(2x2), c designating centered . If we seek the smallest unit
p p
of adsorbates, than 2 ⇥ 2 45 is appropriate to describe the adsorbate order.
Nevertheless, the adsorbate structure has three di↵erent high-symmetry positions
on the surface not accounted for by the simple description of adsorbate geometry.8
The top position is connected with the simplest binding geometry. In the bridge
configuration, an adsorbate atom has two substrate neighbors, while in the hollow
configuration, there are four substrate neighbors, as illustrated in Fig. 4.16.
8
D.D. Vvedensky et al., Phys. Rev. B 35, 7756 (1987).
CHAPTER 4. CRYSTAL STRUCTURE 70

Figure 4.17: Comparison between computations and XANES measurements


indicate that the hollow-site positions for oxygen atoms on Cu(100) are more
probable. Ref. [8].

Figure 4.17 shows on the left-hand side the experimentally determined absorp-
tion spectrum of oxygen on Cu(100). Also shown are calculated absorption spectra
for three di↵erent adsorption sites, the top, bridge, and hollow site. The similarity
between the computed and measured curves indicates that adsorption prefers the
hollow site for oxygen atoms on Cu(100). Calculated curves for di↵erent heights of
O on the Cu surface indicates a value of 0.5 Å as the most probable position.

These are results of multiple scattering calculations that are very cumbersome.
The EXAFS regime, on the other hand, involves the close proximity of the absorb-
ing atom, and a simple scattering procedure delivers successful results. For the
interpretation of EXAFS data, one considers the absorption coefficient
µ = µo [1 + (k)]. µo is the atomic contribution without the neighbors and hence
without the oscillations. The contribution from the neighbors is contained in (k)
which can be represented as
m X Ni 2Ri /⇤ 2k2 2
(k) = ti (2k) e sin[2kRi + 2 i (k)] e i (4.18)
4⇡h2 k i Ri2
CHAPTER 4. CRYSTAL STRUCTURE 71

Figure 4.18: Data processing for EXAFS measurements on Ni. P. Aebi, unpublished.

Here, the following structural and electronic parameters are used. Their proper
values can be extracted from a successful experiment:
Ni number of atoms in the i-th coordination shell
Ri average interatomic distance to the i-th coordination shell
ti (2k) electron backscattering factor
⇤ electron mean free path
i phase shift for the scattering at the potential of the i-th atom
2
i Debye-Waller factor to consider deviations in Ri
Figure 4.18 presents an example of EXAFS data-evaluation process. An absorp-
tion spectrum is obtained from a crystalline Ni sample at 77 K, as shown in the
figure. First, a smooth background is subtracted from data to eliminate the atomic
contribution µo of Ni and obtain just the oscillatory part (E). The next step is to
convert the abscissa from E to k using

k = [2m/h̄2 (h̄! Eo )]1/2 (4.19)

with Eo the inner potential. Subsequently, the resulting spectrum is truncated


at kmin (⇡ 2Å 1 ) in order to eliminate the multiple-scattering XANES signal
CHAPTER 4. CRYSTAL STRUCTURE 72

and kmax (⇡ 10 Å 1 ) in order to account for more energetic absorption edges.


Fourier transformation converts data to real space where the atom-atom distance
can directly be obtained but a phase shift . Phase shifts have to be calcu-
lated. One also uses known values from comparable measurements with related
substances. The truncation of kmax and kmin limits the resolution scale in real
space by R ⇡ ⇡[2(kmax kmin )] 1 (Å).

Figure 4.19: EXAFS oscillations for (a) crystalline and (b) amorphous germanium.
Only the oscillatory part of the absorption edge is shown. Ref. [9].

An example how one obtains useful structural information without an elaborate


phase-shift calculation is illustrated for germanium in Fig. 4.19. One measures
a crystalline sample for which the interatomic distance is well known from x-ray
di↵raction data. One accurately obtains herewith the phase shift. This value is
then used in in amorphous Ge. As seen in Fig. 4.19, no data analysis is necessary
to extract usable information from measured data. Visual inspection shows that
there are strong harmonics in the crystalline substance, whereas in the amorphous
material the fundamental frequency dominates. This comparison can be interpreted
that in the amorphous material every atom has its nearest-neighbors in an ordered
coordination sphere and SRO prevails. In the crystalline sample, on the other hand,
the contribution of the coordination shells beyond the first one is equally important
producing higher harmonics in the spectrum.9
Figure 4.20 displays the x-ray absorption spectrum obtained from a polycrys-
talline sample of stainless steel containing predominantly Cr, Fe, and Ni. This
example demonstrates that the EXAFS method contains element-specific spacial
9
D.E. Sayers et al., Phys. Rev. Lett. 27, 1204 (1971).
CHAPTER 4. CRYSTAL STRUCTURE 73

Figure 4.20: EXAFS following the K edges in an Fe-based steel


containing 30 % Ni and 20 % Cr. Unknown report.

information from a sample which is not necessarily crystalline. Hence, we can obtain
from multicomponent systems crucial information about the identity and geome-
try of each component. We learn about bond lengths and coordination numbers.
There is no requirement on the crystallinity of the samples. All these properties
show that the method suits in an excellent way to study biomolecules for which
the large number of atoms often hinders the application of conventional di↵raction
methods.
Chapter 5

Surface Structure

The structure of any surface, clean or adsorbate covered, is the geometric framework
of any macroscopic process that takes place at the surface. Therefore, the knowledge
of the surface structure is inevitable in characterizing the material. How do we
determine where the surface atoms are located? With the bare eye we resolve
details as large as about 0.1 mm. A magnifying glass allows ⇥100 enlargement,
and the common light microscopes ⇥500. It is a principle question how much we
can enlarge. The di↵raction at the circular hole help us find the answer. Using
the Rayleigh criterium and Bessel function zeroth order, = 1.22 · /D, where
is the distance between two objects that we should be able to resolve, D is the
aperture, and = 0.4 µm the wavelength of the light we use. This basic idea
results in 0.1 mm for our eyes which have an opening of D = 6 mm. So, instead of
visual inspection or microscopic views, we use a scanning electron microscope that
has a variable enlargement up to about ⇥10 000. We can study grain boundaries,
single-crystal regions with di↵erent orientations, crystal defects, etc. The structure
analysis which interests us deals with the positions of surface atoms.
Scanning methods, scanning tunnelling microscopy or atomic force microscopy
rely on the precision with which one can steer a probe on the surface. They do
practically have atomic resolution and map the surfaces in real space. In LEED
electrons are used in the energy range of 30 300 eV. These electrons interact with
those of the solid so intensively that they cannot propagate large distances without
losing energy. Their wavelength is comparable with the interatomic distance at
the surface that results in interference e↵ects. Both of these e↵ects result in the
projection of atomic positions in the reciprocal space where we observe a di↵raction
pattern from the surface. Electrons have a kinetic energy
1 1
E = mv 2 = p2 /m (5.1)
2 2
Using the de Broglie wavelength = h/mv for electrons, one obtains a relationship
between the wavelength and energy in appropriate units:
q
1/2
= h(2mE) and (Å) ⇡ 150/E (eV) (5.2)

74
CHAPTER 5. SURFACE STRUCTURE 75

5.1 2DIM Structure


5.1.1 Ideal Geometry
Compared to the bulk material, the surface region has a finite thickness and no
periodicity in the direction normal to the surface. In 2DIM we need two basis
vectors a and b, which define the unit cell. Any point on the surface can be
reached from the origin using T = p a + q b. The basis vectors a and b span
the unit cell. The sum of all translations for any p and q is called the translational
group of the structure. The translational group defines the 2DIM periodicity of the
structure.

Figure 5.1: (a) Graphical representation of the ten 2DIM point groups. Points
denote positions of equivalent points. (b) Symbols for the various symmetry opera-
tions. From L.J. Clarke, Surface Crystallography, John Wiley, New York, 1985.

Like in the bulk, at surfaces rotation and mirror reflection leave 1 point or 1 line
unchanged. Point group consists of operations which leave one point unchanged and
the structure invariant. For periodic structures there are two possible operations:
mirror reflection around a line (mirror plane) and rotation through 2⇡/n (n =
1, 2, 3, 4, 6) around a point. Only these are compatible with translational properties.
If we combine all the permissible rotations and mirror reflections, we obtain 10
di↵erent point groups instead of 14 in 3DIM.
In Fig. 5.1 we see the ten possible point groups. The numbers give the n-fold ro-
tation around a point, while the first m is a mirror line, e.g., normal to the line that
CHAPTER 5. SURFACE STRUCTURE 76

connects two atoms. The second m in the point group representation denotes other
mirror planes that come about combinations of even-numbered rotations. There
are only 5 di↵erent lattice types, Bravais lattices, which fulfil this requirement in
2DIM, whereas there are 14 types in 3DIM. Below are some properties of these
lattices , including d which is the distance between hk-planes.

A) Oblique lattice — lattice symbol p (primitive) — point group 1,2


|a| =
6 |b| 6= 90
1 h 2 k2 2hk cos
2
d (hk)
= 2
a sin 2 + 2
b sin 2 + ab sin2

B) Square lattice — lattice symbol p — point group 4, 4mm


|a| = |b| = 90
1 h +k 2
2
2
d (hk)
= a2

C) Hexagonal (60 Rhombus) — lattice symbol p — point group 3, 3m, 6, 6mm


|a| = |b| = 120 a and b are two shortest vectors, separated by 120 .
1
d2 (hk) = 43 [(h2 + hk + k 2 )/a2 ]

D) Primitive rectangle — lattice symbol p — point group 1m, 2m


|a| =
6 |b| = 90
1
d2 (hk)
= ( ha )2 + ( bk )2
E) Centered rectangle — lattice symbol c (centered)
|a| =
6 |b| = 90
1
d2 (hk)
= ( a ) + ( bk )2
h 2

5.1.2 Deviation from the Ideal Case


For the most clean metal surfaces the surface structure is identical to the bulk
structure (bulk terminated), whereas at clean SC surfaces considerable deviations
may occur. In particular, variations of the surface periodicity are observed after the
formation of ordered adsorbate structures. The classification of surface structures
frequently follows a nomenclature proposed by Wood based on x-ray crystallogra-
phy.1
Generally, we observe:
Surface relaxation: Upper lattice constant normal to the surface changes.
Reconstruction: Change in the surface structure. A new surface periodicity is
formed in 2DIM which is energetically favorable. We di↵erentiate: a. sponta-
neous surface reconstruction, as shown in Fig. 1.2, b. adsorbate-induced surface
reconstruction, c. structure of the adsorbates.
1
Elizabeth A. Wood, Vocabulary of Surface Crystallography, J. Appl. Phys. 35, 1306 (1964).
CHAPTER 5. SURFACE STRUCTURE 77

Figure 5.2: Surface structures of common cubic crystal types.

The advantage of Wood’s nomenclature is its simplicity, but it cannot describe


all types of periodic surface structures adequately. For a universally applicable
description, we define at the surface new basis vectors as und bs (s for surface),
with Ts = p as + q bs . We can express the new reconstructed basis at the surface
as a linear combination of not reconstructed vectors:
a s = p 1 a + q1 b bs = p2 a + q2 b or
! ! !
as a p 1 q1
=M using M= . (5.3)
bs b p 2 q2
M is the definition matrix for the reconstruction.
Ideally, as = a and bs = b and M is identity.
The area of the unreconstructed cell: |a ⇥ b| = F
The area of the reconstructed cell: |as ⇥ bs | = F·Det M
In nature the periodicity of reconstruction is always larger than that of the bulk-
terminated surface.

5.1.3 Classification of Reconstructions


A) Simple surface structures
p and q are integers, Det |M| is a whole number.
CHAPTER 5. SURFACE STRUCTURE 78

B) Coincidence structures
p and q are rational numbers, Det |M| is a rational number.
C) Incoherent surface structures
p and q are irrational numbers, no relationship between the surface and the bulk.

Examples:
Simple case: as = p1 a and bs = q2 b, i.e., as and bs are parallel to a and b:
!
p1 0
M=
0 q2
The notation is S(hkl)p ⇥ q – A; S for the substrate and A for the adsorbate.
Si(111) 7⇥7; Au(110) 2⇥1; Al(111) 3⇥3 – Si
Less simpler case:
a— centered (001) 2⇥2
p p
b— (001) 2 ⇥ 2 45

S(hkl) |a|a|s| ⇥ |b|b|s| ↵ A


↵ is the rotation of the surface lattice relative to the substrate.

Further examples:
p p
Ni(111) 3 ⇥ 3 30 O
p p
Ni(100) 2 ⇥ 2 45 S or identical to Ni(100) c(2⇥2) S
p p
Cu(100) 2 ⇥ 2 45 O

Figure 5.3: Examples of overlayer structures. a) 2 ⇥ 2, b) c(2 ⇥ 2),


p p
c) 3 ⇥ 3 - R 30 .

5.2 Determination of Surface Structure - Real-


Space Techniques
The observation of surface structures from a distance X far away with respect to
interatomic distances is called far-field microscopy, while techniques investigating
CHAPTER 5. SURFACE STRUCTURE 79

the surface structure from atomic proximity is termed near-field microscopy. Basi-
cally, the wavelength of radiation used is ⌧ X. Far-field methods have limited
resolution given by the wavelength of the radiation. They find a remedy by leaving
the direct space and operating in reciprocal space, i.e., di↵raction space. The best
known method is LEED for surface studies. Near-field techniques are capable to
resolve each surface atom, yet an overall surface information remains desirable. It
is arbitrary how one places experimental techniques in di↵erent groups. Real-space
methods present the advantage that the structural information is directly available
without the necessity of transforming the data back to direct space.

5.2.1 Confocal Microscopy


This technique is derived from conventional optical microscopy. It is designed to
overcome some limitations of traditional wide-field microscopes. In a conventional
microscope, the entire specimen is flooded evenly in light from a light source. In
a fluorescence microscope, all parts of the specimen in the optical path are excited
at the same time and the resulting fluorescence is detected by the microscope’s
photodetector or camera including a large unfocused background part. A confocal
microscope uses point illumination on the sample at the focal plane of the objective
lens. Reflected light from the illuminated volume of the specimen is collected by the
objective and reflected by a beam splitter towards a pinhole arranged in front of the
detector. Light from the focal plane is focused on the detector pinhole and registered
by the detector, whereas information which does not originate from the focus level
of the microscope objective, is faded out by this arrangement, as can be seen in
Fig. 5.4. The advantage of out-fading information from above or below the focal
plane enables the confocal microscope to perform depth-dependent measurements
and optical tomography becomes possible.

Figure 5.4: Schematic drawing of a confocal microscope showing its


principles of operation. Form Wikipedia.org.
CHAPTER 5. SURFACE STRUCTURE 80

Figure 5.5: A confocal micrograph obtained from the organ Corti.


From Wikipedia.org.

The image above is obtained from the sensory cells of the ear, the organ of Corti.
The organ is located in the inner ear of all mammals and contains auditory sensory
cells. The hair-like parts vibrate and transmit stimuli when sound waves are picked
up by the ear canal.

5.2.2 Photoelectron Spectromicroscopy


The photoemission process contains information primarily on the electronic prop-
erties of the sample. In conventional experiment one measures the number of pho-
toemitted electrons as a function of their kinetic energy. The inset of Fig. 5.6
schematically shows the absorption of x-rays with an energy h⌫ causing the excita-
tion of a K-shell electron into continuum with a kinetic energy Ekin = (h⌫ EB ).
EB is the binding energy of the K-shell electron before the excitation. EB is
found by subtracting Ekin from h⌫, as seen by the abscissa of Fig. 5.6. The fig-
ure shows the spectrum of electrons in a kinetic-energy range that corresponds to
0 < EB < 100 eV. At the high-energy end of the spectrum, i.e., near EF (EB = 0),
the signal originates from 5d electrons. The striking feature of the spectrum is the
the presence of intensity at a binding energy 80 90 eV. These strong emission
lines are due to the spin-orbit split 4f levels that have sharp binding energies. This
definite binding energy is a fingerprint for Au.2
2
K. Siegbahn et al., ESCA: Atomic, Molecular and Solid State Structure by mans of Electron
Spectroscopy, Almqvist and Wiksells, Uppsala, 1967.
CHAPTER 5. SURFACE STRUCTURE 81

Figure 5.6: X-ray photoemission spectrum from polycrystalline Au. Ref. [16].

Since the photoelectrons due to the 4f levels of Au have a kinetic energy of


about 1150 eV, inspecting Fig. 5.7, we estimate their escape depth to be around
15 Å. Hence, the photoexcited 4f electrons mostly originate from a near-surface
region.
All photoelectron spectromicroscopes fall in one of two groups: scanning-focusing
instruments, displayed in Fig. 5.8, which achieve high lateral resolution by focusing
the photon beam on the sample, and electron-imaging instruments, which use a
specially developed electron-optics device.3 The first type of instruments primarily
implements photoelectron energy spectroscopy, whereas electron-imaging instru-
ments primarily perform photon-energy absorption spectroscopy. Therefore, the
two types of instruments are largely complementary rather than in competition
with each other. The focusing of x-rays is achieved by Fresnel zone plates. One
usually obtains excellent chemical and a spatial resolution in the order of 1 µ.4

5.2.3 Scanning Electron Microscopy


Scanning Electron Microscopy (SEM) consists of scanning the specimen under vac-
uum with a highly focussed beam of electrons of variable energy and imaging the
reflected electrons. These carry information about the sample’s surface topogra-
3
M. Kiskinova et al., Surface Rev. Lett. 6, 265 (1999) and references therein.
4
B.P. Tonner et al., Rev. Sci. Instrum. 63, 564 (1992).
CHAPTER 5. SURFACE STRUCTURE 82

Figure 5.7: The 2DIM imaging of surface structures is achieved either (a) using
an imaging electron optics or (b) scanning the photons on the surface. Ref. [3].

phy. Thus, SEM can produce high-resolution images of a sample surface, revealing
details less than 1 nm in size. Due to the extremely narrow electron beam, SEM
micrographs have a large depth of field yielding a characteristic 3DIM appearance
useful for understanding the surface structure of a sample. Magnification in a SEM
can be controlled over a range of up to 6 orders of magnitude from about 10 to
500,000 times.
The spatial resolution of the SEM (somewhere between less than 1 nm and
20 nm) depends on the size of the electron spot, which in turn depends on both
the wavelength of the electrons and the electron-optical system that produces the
scanning beam. The resolution is also limited by the size of the interaction volume
in the specimen. The spot size and the interaction volume are both large compared
to the distances between atoms, so the resolution of the SEM is not high enough
to image individual atoms, as is possible in the shorter wavelength, i.e., higher
energy TEM. The SEM has compensating advantages, though, including the ability
to image a comparatively large area of the specimen; the ability to image bulk
CHAPTER 5. SURFACE STRUCTURE 83

Figure 5.8: X-ray scanning photoelectron micrographs obtained from two cell
cultures. The labels (a) and (b) refer to two microscopic areas with distinctly
di↵erent chemical composition. Ref. [3].

materials (not just thin films or foils); and the variety of analytical modes available
for measuring the composition and properties of the specimen.
If the SEM is equipped with a cold stage for cryo-microscopy, the cryogenically
fixed specimens can successfully be imaged. These specimens may be cryo-fractured
under vacuum to reveal the internal structure. Low-temperature SEM is also appli-
cable to the imaging of temperature-sensitive materials such as ice and fats. In the
environmental SEM, samples with high vapor pressure can be kept during measure-
ment under pressure in a protective gas ambient in a di↵erentially-pumped sample
stage.
Characteristic x-rays, which are produced by the interaction of electrons with
the sample, are routinely detected in a SEM equipped for energy-dispersive x-ray
spectroscopy (EDX or EDAX) or wavelength dispersive x-ray spectroscopy to iden-
tify the composition and measure the abundance of elements in the sample (see
Section 6.6.1). This tool renders possible the combination of the geometric infor-
mation with the chemistry of the sample.

5.2.4 Field Emission Microscopy


The field emission process belongs to the group of processes, like thermionic emis-
sion, Auger electron emission, or secondary electron emission that describes extrac-
tion of electrons from a solid. In the case of field emission strong electric fields make
electrons tunnel into vacuum, a process also called the Shottky e↵ect. In order to
achieve strong electric fields we shape the emitting surface into a very fine tip with
a radius of curvature r = 100 nm or smaller and apply to it a negative voltage as
low as U = 100 V. This situation results in a field of F = 107 V/cm at the tip. The
schematics of the experiment is given in Fig. 5.11.
In Fig. 5.12 an energy diagram of a metal is displayed on the left-hand side.
CHAPTER 5. SURFACE STRUCTURE 84

Figure 5.9: A SEM image of quasicrystalline Al-Cu-Fe, revealing the pentagonal


dodecahedral shape of the grain. From F.W. Gayle, J. Metals 40, 5 (1988).

Conduction electrons fill the states up to EF . They can be promoted to vacuum


if they can overcome the work function . Field emission is realized by shaping
the metal to an extremely fine tip and applying a negative voltage to it to have a
large electric field F at the tip. Classically, we expect an emission current at F ⇠
108 V/cm. By virtue of tunnelling emission occurs even at lower fields. Consider an
electron just outside the surface. It is additionally subject to its own image force
[ e2 /(2z)2 ], as shown in the figure.
Finally, the image potential together with the applied electric field lower the
work function by and field emission occurs via tunnelling already at fields 100⇥
lower than expected. We have to consider that image potential can be obtained by
integration because it is work done by promoting an electron from z to infinity
Z 1
Wimage = e2 /(2z)2 dz = e2 /4z. (5.4)
z

It is then a quantum mechanical task to derive (analytically) the Fowler-Nordheim


equation for the current density J using Wentzel-Kramers-Brillouin (WKB) method:
3/2 /F
J = A(F 2 / )e c
(5.5)

using A = e3 /8⇡h̄ and c = 4(2m)1/2 /(3h̄e). J depends only on two quantities, the
work function and the applied field F . A moderate voltage of 100 V at a fine
tip with a radius of curvature r = 100 nm we have a field strength of 107 V/cm.
For a work function of = 5 eV, typical of many metals, we obtain a current
CHAPTER 5. SURFACE STRUCTURE 85

Figure 5.10: A tip radius of ⇡ 100 nm results in very strong electric fields that
make electrons emit from the tip surface and fly towards the concentric collector.

Figure 5.11: Energy diagram of a metallic tip. The applied field F lowers the work
function and helps electrons tunnel through the reduced barrier into vacuum.

density of 103 A/cm3 . This is a very large current density, having additionally
an extremely high brilliance (= total current/emitting surface area). Thus, field
emission cathode is preferably used in many applications as an electron source, e.g.,
in electron microcopy. In the laboratory, we can plot ln(J) as a function of 1/F
and determine the work function of the emitter.
Fig. 5.11 helps us realize that the sample in the form of a fine tip and the large
collector render a magnification of M = R/r which is in the order of 106 in the
present case. If the human eye resolves structures of 0.1 mm, the field emission
CHAPTER 5. SURFACE STRUCTURE 86

Figure 5.12: Erwin Müller realized the field-emission microscope back in 1930. On
the left, the tungsten tip is shown enlarged. The field-emission micrographs show
the tungsten surface with di↵erent amounts of adsorbed barium atoms. Ref. [5].

experiment can be used as a microscope to resolve atomic details. This is the


principle of the field-emission microscope (FEM) invented by Müller.5
As seen in the Fig. 5.13, the vacuum system for FEM is relatively simple. Con-
centric with the sharp tip (see inset) there is a spherical fluorescent screen. There
is additionally a Ba source for adsorption studies. Initially, the tip is cleaned by
reversing the polarity of the voltage, i.e., applying a positive voltage to the tip. On
the tip, the molecules or atoms making up the contamination, are also charged pos-
itive and accelerated away from the surface, thus causing desorption and cleaning
the surface. This process is called field desorption. For imaging by field emission,
negative voltage is applied to the tip. Four FEM images from a W tip are shown on
the right-hand side with di↵erent amounts of Ba, either a thick layer of Ba at di↵er-
ent stages of field desorption or di↵erent amounts of Ba after successive evaporation
from the Ba source.
There are two limitations to FEM. Cleaning the surface by field desorption
requires that the specimen has a high melting point. Otherwise, the sample itself
is desorbed away. The second limitation requires that the sample has to be in a
special form, namely a very fine tip.
Observing signals originating from various crystallographic directions, the sam-
ple is mounted within some deflection plates in order to steer di↵erent portions of
the field emission signal onto the screen, as shown on the left-hand side of Fig. 5.14.
Additionally, by placing a hole in the middle of the imaging screen, one can direct
5
E.W. Müller, Z. Phys. 108, 668 (1938).
CHAPTER 5. SURFACE STRUCTURE 87

Figure 5.13: (left) A modern FEM instrument with deflectors around the tip
and hole in the collector screen to allow electrons from preselected directions
into the energy analyzer. (right) Total energy distribution of field emitted
electrons from (100) plane of tungsten. The dashed lines indicate free-electron
energy distribution. Ref. [6].

parts of the field emitted electrons into an energy analyzer. Figure 5.14 shows on
the right-hand side an energy distribution curve from a clean W(100) surface.6
Intuitively one expects to record the density of states if one measures the en-
ergy distribution curves in FEM. Using the deflection plates one can additionally
steer di↵erent directions into the energy analyzer and study the dependence of the
electronic properties on the directions. Unfortunately, no such dependence could
convincingly be found. The tunnelling current actually depends on the density of
states, but also on the Fermi-Dirac distribution, tunnelling probability, and on the
rate of electrons hitting the surface. The last item is the group velocity of elec-
trons / dE/dk. Density of states, on the other hand, is proportional to (dE/dk) 1 .
Hence, these terms cancel out, and the measurement of the energy distribution of
field emitted electrons essentially reflects the energy dependence of the tunnelling
probability, illustrated in Fig. 5.14 (dashed curve) . The arrows mark energy posi-
tions of enhanced intensity above the dashed curve. These are additional emission
channels beyond the bulk density of states. In fact, Fig. 5.14 is the first documenta-
tion of the existence of surface states in a metal surface. Relying on this insight into
the field emission process, we also realize that electrons have a velocity component
parallel to the surface. This component blurs FEM images.
Finally, we realize that the WKB method delivers realistic solutions for quasifree
sp electrons. On the other hand, d electrons with l = 2 or f electrons with l = 3
are subject to centrifugal barrier in addition to Coulomb-type potential for which
6
E.W. Plummer and J.W. Gadzuk, Phys. Rev. Lett. 25, 1493 (1970).
CHAPTER 5. SURFACE STRUCTURE 88

the radial part can be written as

Ve↵ (r) = V (r) + l(l + 1)h̄2 /(2mr2 ) (5.6)

This potential e↵ectively traps the d or f electrons in its potential well. They have
to overcome the centrifugal barrier while tunnelling into vacuum reducing thereby
the tunnelling probability.

5.2.5 Field Ion Microscopy


Field desorption was used to clean the tip in a FEM experiment and at the same time
was taken as a limitation because it removed substantial portions of the material.
Müller has realized that the field desorbed particles fly away from the tip radially
to the collector and they can be used for imaging the surface. He followed even
a better idea and introduced an imaging gas, preferably He, into the experimental
chamber where the tip had a positive polarity.7 The noble gas atoms are polarized
under the influence of the inhomogeneous field between the tip and the collector
and attracted to the tip. Once arrived onto the tip, surface the atoms jump around
on the surface until they are ionized at an atomic site with increased electric field
(Fig. 5.15). Such sites are located at protruding atoms or steps and edges formed by
the crystallographic topology of the tip. The ionized gas particles fly away radially
from the tip and imaged on the collector screen and produce a bright spot signalling
the location on the tip where they are ionized. Since the He atoms are ionized at
sites with small radius of curvature that produce highest electric fields, such places
are imaged on the screen.
Figure 5.16 shows a field-ion microscope (FIM) image obtained from W(011),
cooled down to 4 K in order to attenuate the lateral motion of surface atoms and
obtain a sharp image. In order to visualize the crystal structure we have to con-
sider that the imaging gas ions have straight trajectories from the sample tip to
the collector screen. This is called gnomonic projection. Thus we observe in the
middle of the pattern the principal orientation of the crystal, while other directions
appear under an angle characteristic to the inclination of these directions relative
to crystal orientation. Hence crystallographic directions are conserved in FIM, but
no information is available about interatomic distances.8
A seminal experiment on surface di↵usion for adsorbates on refractory metal
surfaces has successfully been performed with FIM as early as 1976.9 Consider an
atom in a chemisorbed state. Adsorption places it at the bottom of a potential
well in thermal equilibrium with the substrate. Thermal fluctuations tend to drive
the particle back into the gas phase at a rate proportional to e Ed /kT . This rate
7
E.W. Müller, In Advances in Electronics and Electron Physics, Vol. 13, Academic, New York,
1960.
8
see for a review T.T. Tsong, Prog. Surf. Sci. 10, 165 (1980).
9
K. Stolt et al., J. Chem. Phys. 65, 3206 (1976).
CHAPTER 5. SURFACE STRUCTURE 89

Figure 5.14: Diagram showing schematically the field-ion image


formation. Refs. [7,8].

Figure 5.15: FIM image from a W crystal oriented along [011]. Ref. [7,8].
CHAPTER 5. SURFACE STRUCTURE 90

is quite small since typical values of the heat of desorption Ed are 1 5 eV. The
adsorbate also can jump laterally along the surface from one well to the next. The
energy barrier to this motion, Em , is less than Ed because the particle always is in
contact with the substrate. The direction of each jump is completely random and
uncorrelated from jump to jump. This kind of behavior is called a random walk.

Figure 5.16: FIM image of a rhenium atom on W(211). Ref. [9].

Figure 5.17: Potential energy surface felt by a di↵using atom over a 1DIM cluster.
From T.R. Linderoth et al., Phys. Rev. Lett. 82, 1494 (1999).

For an uneven path, as shown in Fig. 5.18, the end atom has to surmount the
activation barrier Eup . On the upper plane the adatom is in a metastable state and
can move to another site if it overcomes Edown . The displacement rate would be
proportional to a Boltzmann factor e Eup /kT .
In a 2DIM square lattice with lattice constant l, P (x, y, t) is the probability that
the particle is at lattice site (x, y) at time t. The Einstein di↵usion relationship
states h(R Ro )2 i = N l2 = 2D⌧ , where Ro is the initial position, N the number
of steps, and D the di↵usion constant. If each jump occurs after a time ⌧ , we
can write D = l2 /(4⌧ ). Figure 5.17 shows that FIM directly images the di↵usive
motion of individual rhenium adatoms deposited onto clean W(211) surface. The
CHAPTER 5. SURFACE STRUCTURE 91

(211) surface provides a trough in which the adatom is trapped and its motion is
confined only in one direction. The position of the adatom has been determined for
di↵erent temperatures which helps us compute values of D using an Arrhenius plot
Em /kT
D = Do e . (5.7)

Figure 5.18: Arrhenius plots of rhenium atoms and dimers on W(211). Ref. [9]

The investigation of di↵usion of Re/W(211) reveals that the activation barrier


Em is of the order of 5-20 % of the chemisorption energy and the pre-exponential
factor Do is of the order of 10 3 cm2 /s. If we take Do = ⌫l2 /4 for a simple random
walk model, the choice of l = 3 Å as a typical nearest-neighbor jump distance, we
obtain ⌫ = 1013 /s which is a reasonable estimate of the oscillation frequency at the
bottom of a chemisorption well.
Di↵usion proceeds rapidly on a close-packed (111) surface and rather more slowly
on the (atomically) rougher surfaces of the crystal. This type of data is directly
relevant to the question whether an elemental material can achieve its equilibrium
crystal shape. This is so because surface di↵usion is the rate-limiting step to the
dissolution of high Miller index crystal faces into facets of low index planes with
lower surface tension.
CHAPTER 5. SURFACE STRUCTURE 92

5.2.6 Secondary-Electron Imaging


Secondary-electron imaging (SEI) is a method to investigate the arrangement of
atoms in real space. This method involves the excitation of the surface with elec-
trons of energy near 2000 eV and 2DIM imaging of the quasi-elastically backscat-
tered electrons. SEI probes the immediate environment of a particular atomic site.
Individual atomic species need not possess LRO, but the local environment must
be in the same orientation around each atom. Therefore, SEI is ideally suited for
the study of structures that only possess orientational LRO, e.g., quasicrystals.

Figure 5.19: Atomic scattering amplitude |f (✓)| of electrons at Ni atoms at


di↵erent energies. With increasing energy, electron intensity is concentrated
predominantly in the forward-scattering direction, defined as ✓ = 0. From M.
Fink and A.C. Yates, At. Data 1. 385 (1970).

The processes involved in imaging the surface atomic structure of solids by SEI
can be visualized in a two-step model: (1) the excitation of the surface by primary
electrons having an energy near 2000 eV, and (2) subsequent 2DIM spherical record-
ing of the backscattered secondary electrons. In the first step, primary electrons
penetrate the solid and are di↵racted at atoms in a surface layer. This process lo-
calizes the electron wave field at atoms below the surface. In the second step, these
atoms are considered to be point sources of secondary electrons which emanate as
spherical waves from localized atomic sites, and are subsequently scattered by the
atoms surrounding the source atoms.
At energies above a few hundred eV, elastic forward scattering is the dominant
process and the scattered intensity is focussed in the forward direction, as shown
in the Fig. 5.20. Therefore, the intensity of secondary electrons are concentrated
into internuclear directions with respect to the source atoms. Electrons which can
escape the solid are predominantly those scattered by atoms around and near the
source. Figure 5.21 shows schematically one such source of secondary electrons
CHAPTER 5. SURFACE STRUCTURE 93

Figure 5.20: Secondary electrons are emitted from an atomic source localized at a
lattice site. The scattering intensity is concentrated in forward-scattering directions.

below the surface. By virtue of this process, a spherical recording of these electrons
represents a central projection of the internuclear directions in a near-surface region
of the sample.
Since energy-loss processes are involved in the production of secondary electrons,
scattering events around di↵erent atomic sites contribute incoherently to the total
emitted current, and, therefore, successful imaging is possible even in the absence
of LRO. This is not a coherent process. Electrons at around 2 keV have longer
mean free path, and therefore penetrate several atomic layers into the solid. Hence,
one does not only observe the arrangement of atoms at the surface, but multiple
atomic layers can be analyzed simultaneously. An additional advantage of the
SEI technique compared to other methods is that images can be recorded, while
the sample is rotated through an appropriate axis. These images provide 3DIM
views of the structure. Consequently, all symmetry elements of the sample can
be observed with equal precision, and the angles between the symmetry axes can
exactly be determined. Further, since the patterns are mostly produced by forward
scattering along atomic chains and contain real-space features, time-consuming data
transformation is no longer needed. This advantage together with the high signal-to
noise ratio markedly accelerates the acquisition of structural information.
Figure 5.22 shows an SEI pattern obtained from an AlPdMn alloy. Five bright
patches positioned on a small circle with a radius of ✓ = 32 form an equilateral
CHAPTER 5. SURFACE STRUCTURE 94

Figure 5.21: (Top panel) SEI image from the pentagonal surface of the icosahedral
quasicrystal Al-Pd-Mn. The electron gun used for the excitation of secondary elec-
trons and its holder appear in the middle of the figure. Structural details become
visible as the sample is tilted in one direction. (Lower panel) A wire-frame model
of an icosidodecahedron in the same orientation. Ref. [10]

pentagon, which clearly reveals the presence of fivefold symmetry in the structure.
There is a bright band of increased electron intensity which connects each adjacent
bright patch. Other bright spots appearing halfway between the two main patches
are probably generated by additional atoms in the quasicrystal structure. By rotat-
ing the sample through polar and azimuthal angles we can experimentally determine
the angles between the major symmetry axes and prove that the specimen possesses
icosahedral (53m) point-group symmetry.10
Some electromagnetic forces try to establish order between particles, while ther-
mal motion tends the opposite. The state of order is a universal phenomenon in
3DIM and is described by phase transitions. There are phenomenological methods
to describe phase transitions. They rely on concepts like symmetry or order. Almost
no microscopic ideas have been used so far. Phase transitions occur because all sys-
tems in thermodynamic equilibrium try to minimize their free energy F = U T S.
U is the internal energy and S the entropy. A certain phase dominates the other
at a given temperature because di↵erent states distribute their energy di↵erently
between U (T ) and S(T ).
For 3DIM, first-order phase transition is characterized by an abrupt change
10
Y. Weisskopf, Growth of CsCl-type domains on icosahedral quasicrystal Al-Pd-Mn, Logos
Verlag, Berlin, 2006.
CHAPTER 5. SURFACE STRUCTURE 95

in the order parameter at the critical temperature, while in a second-order phase


transition, two equivalent phases do coexist and become indistinguishable. The
order parameter changes continuously with temperature and near Tc as (T Tc ) .
These cases are schematically shown in Fig. 1.9.

Figure 5.22: Stacking sequence in the fcc and hcp structures..

Cobalt is known to exist in two allotropic phases, a low-temperature hexag-


onal close-packed (hcp) phase and a high-temperature fcc phase. The hcp – fcc
transformation is reversible and weakly first order. The hcp and fcc phases are in
equilibrium at Teq ' 422 C, but there is considerable thermal hysteresis between
the heating and cooling transformations. Because of these characteristics, together
with the di↵usionless nature of the phase change, this transformation is designated
as martensitic. There is no di↵erence in the corresponding coordination numbers
between the hcp and fcc phases, there is a substantial change in the stacking se-
quence, as seen in Fig. 5.23. Both the fcc and hcp structures can be represented by
stacking planes with atoms in sixfold symmetry, the (111) planes. The di↵erence
in stacking sequence leads to fcc or hcp structures. Layers with alternating AB
planes result in the hcp structure, while a sequence alternating ABC gives the fcc
structure. Figure 5.23 shows this structural assembly in top and side view. For
the fcc structure [111] is normal to the sixfold-symmetry planes, it is the [0001]
for the hcp structure. Thus, characteristic modifications in microscopic properties,
e.g., the electronic structure, and macroscopic properties, e.g., the electrical resis-
tivity and the work function, have been observed to accompany the martensitic
transformation.
The change in the stacking sequence in the near-surface region is observed di-
CHAPTER 5. SURFACE STRUCTURE 96

Figure 5.23: The electrical resistivity ⇢ and the surface spot intensity A as a function
of T . ⇢ and A are both normalized to unity for the hcp phase and zero for the fcc.
The resistivity is used to record the bulk phase transitions. The inset shows ⇢(T )
/⇢(RT ). A is the intensity of backscattered electrons along an hcp-characteristic
direction divided by the intensity along an fcc-characteristic direction. Ref. [11].

rectly with SEI because the depth of the imaging is given by the mean free path
of 1750-eV electrons (' 15 Å), which corresponds approximately to 5 6 layers.
The resistance of the specimen was measured continuously for temperatures be-
tween 300 and 870 K with the SEI patterns recorded simultaneously. Since the
lateral arrangement of atoms on the (0001) surface of hcp Co is the same as that
on a (111) surface of fcc Co, observing the hcp!fcc phase change on the surface is
not straightforward. The change in the structure can be detected only by surface
methods that are sensitive to at least the top three layers of atoms.11
Figure 5.24 shows the temperature dependence of the ratio A = Ihcp /If cc . A is
normalized to unity at room temperature and to zero at the high-temperature end.
Each A point represents an individual SEI measurement such that large (low) values
of A are associated with the hcp (fcc) structure at the surface. The arrows show
the direction of the temperature change. Also shown is the bulk resistivity ⇢ of the
11
M. Erbudak et al., Phys. Rev. Lett. 79, 1893 (1997).
CHAPTER 5. SURFACE STRUCTURE 97

Co crystal, which has been transformed by subtracting a parabolic background and


normalized to unity for the hcp phase and zero for the fcc phase. The inset shows
the raw resistivity data. The temperature dependence of A shows a hysteresis, but
one that is distinctly broader in temperature than that of the bulk.
The origins of the transformation occurring at di↵erent temperatures on the
surface than in the bulk remain unclear. One possibility is that the thermal equi-
librium transition temperature of the surface is e↵ectively higher than that in the
bulk. A second possibility is that the sliding of the planes is better pinned at the
surface than in bulk, giving rise to a more stable metastable state.

5.2.7 X-Ray Photoelectron Di↵raction


X-ray photoelectron di↵raction (XPD) relies on the interaction of photoexcited
electrons with the atoms of the sample during the escape process. The heart of the
experiment is the fact that the core levels of an atom have a well-defined binding
energy EB , as seen for the 4f electrons of a gold sample in Fig. 5.6, and the x-
ray photoemission process leads to excited electrons with a similarly well-defined
kinetic energy Ekin = h⌫ EB . Thus the photoemission process produces
sources of electrons below the surface localized at atomic sites. These electrons
scatter elastically at atoms during the escape process. For large values of h⌫ and
shallow core levels, Ekin is sufficiently large that scattering of emitted electrons can
be taken as being in forward-scattering direction, as shown in Fig. 5.21. Like it
is done in SEI, one collects these electrons for a real-space image of the surface
atomic symmetry. Compared to SEI, XPD has the advantage that the core levels
are element specific and one maps the structural symmetry around each atom in a
multi-component alloy, like it is done in EXAFS.12
Data are obtained by fixing the location and tuning the energy of the energy
analyzer on the core level and rotating the sample through polar and azimuthal
angles, while recording the intensity. The entire pattern is constructed by an angular
collection of discrete measurement points. Thus, the disadvantage of XPD is that
data collection time is too long that the sample surface has to be cleaned frequently;
one only obtains static pictures, while in SEI video recording of the symmetry
properties allows observations during structural phase transitions.

5.2.8 Scanning Microscopes


In far-field microscopy, the distance between the specimen and the radiation source
is large. 2DIM spatial information is obtained either by scanning the source beam
on the surface or the emitted particles are imaged with spatial resolution. In either
case, physical laws prevent atomic resolution. This limitation is circumvented in
near-field microscopy by keeping the distance between the object and the light
12
T. Greber et al., Phys. Rev. Lett. 81, 1654 (1998).
CHAPTER 5. SURFACE STRUCTURE 98

Figure 5.24: (left) The sample holder allowing for polar and azimuthal rotations of
the sample. (middle) The energy analyzed electrons are recorded as a function of
the polar and azimuthal angles and are plotted in an arbitrary fashion. Here, the
pattern represents a stereographic projection. Schematic drawing of the principal
crystallographic properties of a cubic lattice in the same representation. Ref. [12].

source much smaller than the wavelength . The ability to meet this condition
marks the advent of scanning techniques with atomic resolution.
Historically, stylus profilometer represents the principles of near-field microscopy.
The idea is sensing some surface details while scanning it. The precision of scanning
and the method of sensing both are the main factors that limit the spatial resolution.
The idea of imaging with microscopic resolution came with topografiner as shown
in Fig. 5.26.13
The topografiner is based on the revolutionary idea of using piezo-electricity
to attain picometer resolution in three directions. Quartz crystal is piezo-electric,
namely it shows charge separation under mechanical stress. If the crystal is cut in
the proper direction, we observe that the polarity changes for tension or compression
and its magnitude depends linearly on mechanical stress. Conversely, if an external
voltage U is applied, e.g., in x-direction the dimensions of the crystal change by
x = 11 U in the x-direction and y = 11 U ly /lx in the y-direction keeping z = 0.
This situation is illustrated in Fig. 5.27. The polarity of U changes the sign of
x and y . The piezo-electric coefficient 11 is as large as 2.3 pm/V allowing the
required atomic resolution by applying voltages in a practical range. 11 is linear
and has a large dynamical range. These properties are also found in inexpensive
piezo-ceramic material. The principle of moving the probe and the specimen using
the piezo-electric e↵ect applies to all scanning techniques.
13
R. Young et al., Rev. Sci. Instrum. 43, 999 (1972)
CHAPTER 5. SURFACE STRUCTURE 99

Figure 5.25: The experiment is performed in vacuum by scanning a fine


tip in x- and y-directions and sensing structural details in the z-direction
using piezo-electric components. Ref. [13].

5.2.8.1 Scanning Tunnelling Microscope


The principles of imaging in the scanning tunnelling microscope (STM) is based on
a fine metallic tip, like in FEM or FIM, through which an imaging current tunnels
to the specimen. For tunneling the required separation s between the sample and
the tip is adjusted by an additional piezo-ceramic unit. When s is made sufficiently
small there is an overlap of the wave functions of the tip and the sample. If a voltage
V is applied across the tunnel junction, there is a tunnelling current through vacuum

Figure 5.26: Piezo-electric e↵ect in a quartz crystal showing polarization


charges produces by stress and strain.
CHAPTER 5. SURFACE STRUCTURE 100

2
JT / Vs e A s . This is the Fowler-Nordheim equation with an average work function
of the sample and the tip. A variation s in the tunnelling gap by 1 Å results in
a change of current JT by ⇥10. The tip is scanned across the surface at a fixed bias
voltage V and a piezo-electric feedback mechanism regulates the vertical motion
of the tip so that the tunnelling current is kept constant. Thus the tip traces the
contours of constant wave function overlap, namely the surface topology. In the
topografiner the tip was used as an electron source. The distance s was so large
that the experiment was not performed in the tunnelling mode and the current did
not depend critically on s. The success of STM has been achieved by bringing the
tip near to the sample (⇡ 5 Å) in a tunnelling position (G. Binning and H. Rohrer,
Nobel Prize 1986).

Figure 5.27: The working principle of STM showing electron tunnelling


at the tip. From Wikipedia.org.

In STM, the tip can be represented by a spherical potential well, for which we
have to solve the Schrd̈ingier equation. If we only keep the spherically symmetric
solutions, the expression for the tunnelling current reduces to I / eV ⇢(ro , EF ) for
low applied voltages. ⇢(ro , EF ) is the so-called local density of states at the Fermi
energy EF and at the site ro of the surface. Therefore, the surface is scanned while
the current is kept constant, so that the contours of the constant charge density
at EF can be plotted. This reasoning shows that STM is sensitive to local charge
density and not on atomic positions.
One of the first experiments was on the Si(111) surface, because this surface
provides the most interesting reconstruction, namely 7 ⇥ 7 that was discovered
many years ago in LEED experiments, but the real-space visualization of the atomic
arrangement was not complete yet. This surface shows metallic behavior.
On Fig. 5.29 we see the line scans of the STM in two surface directions.14 The
14
G. Binnig and H. Rohrer, Surface Sci. 126, 236 (1983).
CHAPTER 5. SURFACE STRUCTURE 101

tunnelling current, i.e., surface topology, is represented in the third direction. We


can make up rhombic unit cells on the surface, each edge of which is approximately
27 Å-wide. This cell is called the 7 ⇥ 7, because each edge is as long as 7 atomic
distances. Thus, the surface unit cell is ⇥49 larger than the not reconstructed cell.
We further see 12 hills which can be grouped into 2 each with 6 hills. Probably
these hills represent individual atoms.

Figure 5.28: One of the first applications of STM was on the Si(111) 7 ⇥ 7. Ref. [14].

The revolutionary idea in STM is the possibility of spectroscopy at the surface,


which comes about by scanning the surface using di↵erent applied potentials, as
illustrated in Fig. 5.30. If the tip is biased positively current flows from the occu-
pied states of the sample to unoccupied states of the tip (left panel). This provides
information on the occupied states below EF , these are the bonding orbitals. In
contrast, if we bias the tip negatively, current will flow into the unoccupied anti-
bonding orbital states of the sample (right panel). Thus all the states below and
above EF can be recorded.

Figure 5.29: The states involved in tunnelling depending on the polarity of the tip.
CHAPTER 5. SURFACE STRUCTURE 102

2
The lowest intensity of the wave functions for the antibonding states | A B|
is located just between the atoms A and B in agreement with the STM image on
the left-hand side of the Fig. 5.31. This fact renders the best atomic resolution.15
In contrast, she shape of the bonding orbitals does not allow to resolve individual
atoms, because | A + B |2 represents just a big intensity.

Figure 5.30: On the left we see the image of empty states, on the right it is the
image of the filled states of the reconstructed Si(001) surface. Ref. [15].

STM is limited to metallic samples. For others, the local density of states
near EF is negligibly small and we have vanishing tunnelling current. Further, one
has to work with one magnification value. There are some instances where lower
magnification would help to better visualize the surface geometry. This possibility
is not applicable in STM.

5.2.8.2 Atomic Force Microscopy


If the tip is brought near the surface, the surface atoms exert a force on the tip.
This force is / 1/s2 and constitutes the principle of atomic force microscopy (AFM)
operation. The tip is mounted on a flexible cantilever and scanned over surface.
The force acting on the cantilever will be directly related to the deflections of the
cantilever. This situation is illustrated schematically in Fig. 5.32. The deflections
can be registered by shining a sensitive laser on the cantilever and looking for
changes in the reflected beam as the cantilever changes its position slightly. This
technique is known as AFM and allows lateral resolution on the atomic level. The
vertical resolution is below 0.1 nm.
AFM is a good complementation to STM. AFM is sensitive to all electrons of the
solid, the surface need not to be conductive. The resolution of AFM is improved if
s corresponds to just a few atomic distances. This mode is called the contact mode.
For larger tip-to-sample distances we have the lower resolution non-contact mode.
15
R.J. Hamers and Y. Wang, Chem. Rev. 96, 1261 (1996).
CHAPTER 5. SURFACE STRUCTURE 103

Figure 5.31: Operating principle of the atomic force microscope.


From Wikipedia.org.

The first known, but not recognized as such, self-organization phenomenon is


the surface reconstruction. Si(111) is probably the most expensive surface so far,
because the amount of research to explore the exact structure has been enormous.
Figure 5.33 shows on the left-hand side an atomically resolved AFM image of the
Si(111) (7 ⇥ 7) surface. On the right-hand side, the remarkable structural model of
this reconstruction is shown. The key structural features of the model are: (a) 12
top layer adatoms, (b) a stacking fault in one of the two triangular subunits of the
second layer, (c) nine dimers that border the triangular subunits in the third layer,
and (d) a deep vacancy at each apex of the unit cell.16

Figure 5.32: An AFM image of Si(111) (7 ⇥ 7) surface and the so-called


dimer-adatom-stacking fault model for the reconstruction. Ref. [16].

16
K. Takayanagi et al., Surface Sci. 164, 367 (1985).
CHAPTER 5. SURFACE STRUCTURE 104

We identify a force between two objects as friction. This force can be in any
direction, but usually act to impede a movement. AFM senses forces normal to the
surface. Lateral forces, on the other hand, give a clue on the friction between the tip
and sample. Torsion at the cantilever, as seen on the left-hand side of the Fig. 5.34,
can be plotted over the entire surface for performing tribology on nanometer scale.
The polymer imaged here is a mixture of polyallylaminehydrochliride and polydial-
lyldimethylammoniumchloride.17 The micrograph suggest a cluster growth on the
even matrix. Similarly, biomolecules have successfully been detected, analyzed, and
manipulated using AFM.18

Figure 5.33: (left) Principle of the friction-force microscopy. (right)


Friction-force image of polymer macromolecules. Ref. [17].

The strength of the scanning probes resides in the fact that they can operate
under any ambient condition. Polarization-force microscope is just one example.
For a liquid with the dielectric constant ✏, the image charge is reduced by (✏ 1)(✏+
1) 1 . Therefore, the Coulomb force between a tip and the liquid will depend on the
local changes of ✏. In this mode of operation, the electric field between the tip and
the surface induces a polarization charge at the surface and hence a polarization
force, which is measured in Fig. 5.35
" 1 R2
F ⇡ "0 (Vtip + )2 (5.8)
" + 1 z2
Wetting of mica (= aluminum silicate) is technologically important because mica
is widely used in cosmetic industry in facial powder, lip stick, or nail polish, as well
as in electronic industry. It can withstand electric fields as large as 1500 V/mm.
Fig. 5.36 shows microscopically the wetting process of a mica surface. Each frame
is 5 µ ⇥ 5 µ large.19
17
H.-U. Krotil et al., Surf. Interface Anal. 27, 341 (1999).
18
J. Fritz et al., Science 288, 316 (2000).
19
L. Xu et al., Phys. Rev. Lett. 84, 1519 (2000).
CHAPTER 5. SURFACE STRUCTURE 105

Figure 5.34: Principle of the polarization-force microscopy. The induced


polarization force is measured with the microscope. Ref. [19].

Figure 5.35: (left) Time-resolved wetting of mica surface (mica is visible on dark ar-
eas, I, water on light regions, II) observed by polarization-force microscopy. Ref. [19]

The magnetic force microscope (MFM) is a variation of AFM, where a sharp


magnetized tip scans on a magnetic sample. The tip-sample magnetic interactions
are detected and used to reconstruct the magnetic structure of the sample surface.
Many kinds of magnetic interactions are measured by MFM, including magnetic
CHAPTER 5. SURFACE STRUCTURE 106

dipole–dipole interaction. MFM scanning often uses non-contact AFM mode.

Figure 5.36: Fe-atoms placed on Cu(111). Ref. [20].

Scanning techniques have also been used to manipulate surface atoms and
molecules by simply dragging them on a flat surface. The success depends on
the force between the tip and the atom pushed away which must be smaller than
the force between the atom and the surface. Otherwise, the atom will stick to the
surface. Figure 5.37 displays progressive STM images of Fe atoms assembled in a
circular ring on a Cu(111) surface.20 When the cage is closed, we observe interesting
details like concentric water waves in a circular well. These details are produced by
the conduction electrons, which are trapped in the cage, interfere and form standing
waves. They move otherwise as Bloch waves.

5.3 Determination of Surface Structure -


Reciprocal-Space Techniques
Far-field microscopy cannot image surface details finer than the wavelength of radi-
ation used. This fact makes us image the surface in reciprocal space (Frauenhofer
di↵raction) instead of in the real space. Following the recording process the in-
formation is subsequently transferred into the real space. We now transport the
Bravais lattice into reciprocal space, as we did for 3DIM (cf. Eq. 4.4).
20
M.F. Crommie et al., Science 262, 218 (1993).
CHAPTER 5. SURFACE STRUCTURE 107

If a⇤ and b⇤ are the basis vectors⇣in⌘ reciprocal space:


a
a⇤ ·a = 2⇡ b⇤ ·b = 2⇡ b
(a⇤ b⇤ ) = 4⇡
a⇤ ·b = 0 b⇤ ·a = 0 d.h. a ? b⇤ and b ? a⇤
We find an example for the relationship of basis vectors in real and reciprocal lattice
in Fig. 5.38. Translational invariance: Each point of the lattice can be reached from
the origin:
Real lattice T=pa+qb
Reciprocal lattice G = h a⇤ + k b⇤
For the superstructures we had expressed the reconstructed surface as a linear
combination of the vectors of the unreconstructed lattice:
! ! !
as a p1 q1
=M ; with M=
bs b p2 q2
M is the definition matrix for the superstructure.

In reciprocal space: ! !
a⇤ s ⇤ a

=M
b⇤ s b⇤
M⇤ = M̃ 1 , or M = (M̃⇤ ) 1 , where ˜stands for the transposed matrix.

In nature |as | |a|, i.e., the periodicity is multiplied with the presence of additional
di↵raction spots.

Figure 5.37: Note that a ? b⇤ and b ? a⇤ .

Recall that:
! ! ! !
as a a⇤ s ⇤ a

=M = M
bs b b⇤ s b⇤
! ! ! !
a as a⇤ a⇤ s
=M 1
= (M⇤ ) 1
b bs b⇤ b⇤ s
CHAPTER 5. SURFACE STRUCTURE 108
⇣ ⌘ ⇣ ⌘
a 1 0
b
(a⇤ b⇤ ) = 4⇡ 0 1
= 4⇡
!
as
M 1
(a⇤ s b⇤ s ) ((M⇤ ) 1 )T = 4⇡
bs
| {z }
2⇡

((M⇤ ) 1 )T = M or M⇤ = (MT) 1

5.3.1 Low-Energy Electron Di↵raction


De Broglie has postulated the wave nature of particles in his historical work.21 Ac-
cording to him, the nonrelativistic electrons have a wave length = h/p. Only one
year later Davisson and Germer22 proved this postulate for electrons in scattering
experiments with Ni(110).

IT , kr
Io , ki

Figure 5.38: A schematic view of the transmission experiment.

Consider a transmission experiment through a thin specimen thickness d as


shown in Fig. 5.39. We write IT = Io e n d , with the scattering cross section,
n density of particles, and n = µ, the absorption coefficient. 1/µ = ⇤(E) is
called the penetration depth or mean free path. X-rays are used to determine the
3DIM structure of the bulk, because 1/µ ⇡ 1000 10000 Å is (at approximately
1.5 keV). Electrons scatter at valence and core electrons of the target material, they
excite bulk and surface plasmons. These are inelastic interactions with large energy
transfer with almost no momentum transfer, i.e., strong in the forward-scattering
direction. Electrons that lose energy do not take part in interference. For elastically
scattered electrons, the mean free path is very short. Information is limited to a
surface region. For electron energies 30 300 eV the mean free path ⇤(E) is limited
to atomic distances as seen in Fig. 5.7.
21
L. de Broglie, J. Physique 7, 1 and 321 (1926).
22
C.J. Davisson and L.H. Germer, Phys. Rev. 30, 705 (1927).
CHAPTER 5. SURFACE STRUCTURE 109

Elastic scattering is strong and occurs at the Coulomb potential of the target
atoms. There is no energy transfer only a momentum transfer (This was briefly
mentioned in the classical limit and given in Eq. 1.3). Since Coulomb potentials
are located where the atoms are, the resulting electron patterns carry information
on the location of target atoms at the surface. This is the working principle of
LEED.23

Figure 5.39: The surface sensitivity is enhanced by a factor of 2 in the exponent in


case of elastic reflection.

How surface sensitive is a LEED experiment? The intensity of electrons prior


to an elastic scattering at a depth z can be written as I(z) = Io e z/⇤ with Io the
intensity of the primary electrons. The intensity of electrons that can leave the
surface after an elastic scattering is IR = Io e 2d/⇤ . The doubling of the exponent
is the additional advantage of LEED. For several other surface-sensitive techniques
the sampling depth is given by e d/⇤ .
Another description divides the intensity of electrons IR reflected from the sur-
face in two component, if Io is the intensity of the impinging electrons. Id orig-
inates from a depth between 0 and d. IV originates from 1 to d (Fig. 5.40).
dIR = Io e z/⇤ · (SN dz) · e z/⇤ = Ae 2z/⇤ dz, where S is the elastic scattering factor,
N the number of scattering centers pro unit area parallel to the surface and N dz
number of atoms per volume. A = Io SN is a constant.
Z d
2z/⇤ 2d/⇤
Id = A e dz = A⇤/2 (1 e ) (5.9)
o
Z 1
2z/⇤ 2d/⇤
IV = A e dz = A⇤/2 e (5.10)
d

Important are relative values: The ratio of surface-to-volume contribution is


23
see for a review E. Bauer, In Interactions on Metal Surfaces, ed. by R. Gomer, Springer Series
Topics in Applied Physics , Vol. 4, New York, 1975, p. 225.
CHAPTER 5. SURFACE STRUCTURE 110

2d/⇤
1 e
2d/⇤
= e2d/⇤ 1. (5.11)
e
When normalized, the surface contribution reads e 2d/⇤ with the corresponding
bulk value of (1 e 2d/⇤ ).
This consideration holds true for most of the metals and for normal incidence at
E = 30 300 eV. Varying the angle of incidence facilitates better surface sensitivity.
If we are interested for a slab d = 3 Å at the surface and ⇤ ⇡ 5Å we have
Id /IV = 2.3, which means that 70 % of the signal originates from the near-surface
region (approximately one single atomic layer) and 30 % from the bulk. Variation
of the angle of incidence may facilitate better surface sensitivity.
If we chose to work between E = 30 300 eV to utilize the extreme surface
sensitivity as seen in Fig. 5.7, we obtain ⇡ 2.2 0.7 eV for the wavelength of elec-
trons, using Eq. 5.2. The wavelength is comparable with the interatomic distance
of most of the elemental solids, which leads to interference e↵ects. Now we deal
with the di↵raction of plane waves at 2DIM lattices. We make three assumptions:
a - only surface scattering (not realistic, because we have seen that about 30% in-
tensity originates from the bulk)
b - no multiple scattering (Born approximation!)
c - elastic scattering, i.e., |ki | = |kr |

First we draw a geometric 1DIM picture (Fig. 5.41). In order to have con-
structive interference in the direction ↵, the path di↵erence must be a b =
d(sin ↵ sin ). It follows

d(sin ↵ sin ) = n . (5.12)

The di↵erence between Eq. 4.8 and Eq. 5.12 exhibits the crucial di↵erence between
the 3DIM bulk and 2DIM surface di↵raction.

Figure 5.40: The path di↵erence between the rays is given by a b.

Following a more physical consideration (Fig. 5.42) we realize that


a) kk is conserved.
b) crystal momentum Gk can always be added parallel to the surface
CHAPTER 5. SURFACE STRUCTURE 111

k
kk = kkr - ki = nGk = n2⇡/d. It follows: |kr | sin ↵ - |ki | sin = n/d ko or
with |kr | = |ki | = |ko

d(sin ↵ sin ) = n .

Figure 5.41: Electron momentum parallel to the surface is conserved.

Here is 2⇡/d = |G| the reciprocal lattice vector or the unit vector of the recip-
rocal space. We have constructive interference whenever kk is a multiple of the
reciprocal lattice vector Gk . In 1DIM kk = nGk , where |Gk | = 2⇡ a
and in an
k k k 2⇡
additional dimension k = mG , where |G | = b . n and m represent the order
of the interference. Zeroth order di↵raction means, = ↵, n = m = 0, specular (=
mirror) reflection, which is encountered for the 00 beam.
The conditions for the constructive interference is best represented by the Ewald -
construction in the reciprocal space. The 3DIM case relevant for x-ray di↵raction
was shown in Fig. 4.10 where the Ewald sphere goes through a reciprocal point in
order to fulfill the condition for constructive interference. In general (3DIM), G =
h a⇤ + k b⇤ + l c⇤ , where h, k, and l are integers. In 2DIM, there is no condition to
be obeyed in the third dimension, and, therefore, l is arbitrary and c⇤ is any vector
perpendicular to the surface. This results in rods in the reciprocal space instead of
points as shown in Fig. 5.43.
Thus in 2DIM the Ewald sphere always intersects the rods and the condition for
constructive interference is always fulfilled for a set of n and m as seen in Fig. 5.44.
In 3DIM the radius of the Ewald sphere, i.e., the electron energy, must exactly be
tuned for the condition to be met.
In Fig. 5.44, G = 2⇡/d, (00) is the origin of the reciprocal space, ki the incident
and kr the reflected beam, and the Ewald sphere has the radius 2⇡/ . Wherever the
sphere cuts the rods, the di↵raction condition is met. This results, at all energies, to
bright spots on the fluorescent screen. Therefore, we realize that the LEED pattern
represents a projection of the unit cell of the surface reciprocal lattice. From the
CHAPTER 5. SURFACE STRUCTURE 112

Figure 5.42: Rods in the reciprocal space of 2DIM di↵raction.

pattern, we extract the size G of the unit cell and its symmetry (nm). Fig. 5.45
represents a typical pattern with a superstructure.24
Figure 5.46 schematically shows the experimental details. Primary electrons
hit the surface exactly at a point which is the center of the concentric and spher-
ical multiple-grid system. The first grid is on the same potential as the sample
to produce a field-free region not to influence the path of the scattered electrons.
The second grid suppresses inelasticly scattered electrons and transmits elastically
scattered ones onto a spherical fluorescent collector screen. Thus electrons backscat-
tered from the sample surface produce di↵raction spots which can be viewed without
distortion.
We have to consider three issues:
1. The condition for constructive interference is 2kmin |G| = 2⇡/d or  2d.
Otherwise the Ewald sphere is to small to cut the rods. In that case we have
evanescent waves propagating into the crystal with imaginary k and exponentially
diminishing intensity, as it is known for surface states.
2. In LEED we have a case of surface di↵raction where the 00 beam always exists
because the Ewald sphere always cuts the reciprocal space rods.
3. In 3DIM we have bulk di↵raction. Like in x-ray scattering, the energy must be
varied considerably in order to find the condition for constructive interference, i.e.,
for the Ewald sphere to go through a reciprocal-space spot (Fig. 5.47).
In 3DIM it is almost impossible to fulfill the condition for constructive inter-
ference. In 2DIM, however, we always have constructive interference. Figure 5.47
shows on the left-hand side that the cross section for elastic scattering decreases
constantly with increasing energy and scattering angle. Its dependence on Z is
rather complicated that makes an atomic identification impossible.
Figure 5.48 displays di↵raction pattern obtained from the Si(111) surface with
di↵erent surface phases: (a) impurity stabilized (1 ⇥ 1) pattern, (b) (7 ⇥ 7) pattern
p p
of the clean surface, (c) ( 19 ⇥ 19) R(23.56 ) - Ni pattern, (d) (5 ⇥ 5) - Cu
24
L.J. Clarke and L.M. de la Garza, Surface Sci. 121, 32 (1982).
CHAPTER 5. SURFACE STRUCTURE 113

Figure 5.43: Ewald construction for di↵raction in 2DIM.

pattern, (e) (6 ⇥ 6) - Au pattern, (f) (7 ⇥ 7) pattern.25

5.3.1.1 Intensity of Di↵racted Beams


We have dealt with the geometry and the symmetry of the atomic distribution at
the surface. Equally useful information can be extracted from the investigation of
the intensity of di↵racted beams.
In general terms, we consider a plane wave o = o eiko ·r incident onto the
surface in the direction so with the momentum ko = 2⇡ so .
rj – position of the primitive cell j in the 2D lattice. In Fig. 5.49 we identify
rj = pa + qb
rl – position of the scattering center l in the primitive cell with total L atoms.
There are M1 ⇥ M2 primitive cells aligned periodically , i.e., 1  p  M1 and
1  q  M2 . M1 ⇥ M2 or aM1 ⇥ bM2 is called the region of coherence. We have to
consider all atoms in the cell and all the cells.
The total scattered wave at X
X
(k) / fl (|kr ko |)ei(kr ko )·(rj +rl )

l,j

fl is the scattering amplitude of the atom l. More elegantly:


X X
(k) / fl (|kr ko |)ei(kr ko )·rl )
· ei(kr ko )·rj )

l j

25
J.C. Phillips, Surface Sci. 40, 459 (1973).
CHAPTER 5. SURFACE STRUCTURE 114

Figure 5.44: LEED pattern observed from Mo(110) with Cl


atoms adsorbed on the surface. Ref. [24].

The first factor is equal for each cell and depends only on the atomic species.
L
X M1
X M2
X
i(kr ko )·rl i(kr ko )·ap
(k) / fl (|kr ko |)e · e · ei(kr ko )·bq
(5.13)
l=1 p=1 q=1

The first term is the scattering factor F for the primitive cell. It also depends on
the phase factor: |kr ko | = 2|k| sin(✓/2).
The product of the other two terms is called the interference function or the lattice
contribution G. It depends on the 2DIM periodicity as well as on (kr ko ). Hence

/F·G

In the experiment, only the intensity is accessible but not the amplitude:

I = | |2 / |F|2 · |G|2

It is because of the lattice contribution that the intensity varies if we move on a


circle of radius R. So far, this description holds for the dynamic as well as the
kinematic theory.
The di↵erence between the two approaches lies in the treatment of the scattering
factor F. The dynamic approach considers all the scattering paths into the unit cell
and among unit cells. It includes multiple-scattering e↵ects. It correctly treats F
as dependent on both ko und kr . In the kinematic approach one only considers
a week interaction of the plane wave with matter. Only simple single-scattering
events take place. This approximation works well for x-ray or high-energy electron
scattering. Its validity is doubtful for slow electrons. Yet it is very popular, because
CHAPTER 5. SURFACE STRUCTURE 115

Figure 5.45: Schematic representation of the LEED experiment.

- the calculations are relatively simple,


- the spatial distribution of the spots are correct,
- relative magnitudes are often correct.

5.3.1.2 Deviation from the Ideal Case


1. Lattice of finite size
How does the interference function look like after summing up I / |G|2 ?
Consider:
MX1 1 1
X X 1 xM
xm = xm xm =
m=0 m=0 m=M 1 x 1 x

sin2 [ 12 M1 a · (kR ko )] sin2 [ 12 M2 b · (kR ko )]


I/ · (5.14)
sin2 [ 12 a · (kR ko )] sin2 [ 12 b · (kR ko )]

Maxima of the intensity is expected each time when the argument of the denomi-
nator is multiples of ⇡. Thus 12 a · (kR ko ) = h⇡ and 12 b · (kR ko ) = k⇡, since
kR = 2⇡ sR and ko = 2⇡ so (both s are unit vectors).
We obtain a·(sR so ) = h and b·(sR so ) = k . These are the Laue conditions
for constructive interference at a 2DIM lattice.
CHAPTER 5. SURFACE STRUCTURE 116

I 2 DIM I 3 DIM

E E

Figure 5.46: The intensity I vs. energy of the di↵racted beams in 2D and 3D.

Figure 5.47: Selected LEED patterns from Si(111) with di↵erent


reconstructions, see text. Ref. [23].

rl ko

rj
θ
O
a kr
b
R

X
Figure 5.48: A plane wave is incident onto the surface with the unit cell that contains
L atoms. We receive the scattered wave at X in a distance R from the surface.
CHAPTER 5. SURFACE STRUCTURE 117

Figure 5.49: Modulations in the di↵raction intensity for finite M . From CK.

The equation for I implies that the height of the interference maxima is pro-
portional to (M1 M2 )2 and occurs whenever the denominator is zero, as depicted in
Fig. 5.50.
In order to find the half width of the di↵raction spot we have to move way from
the intensity maximum by ✏ to arrive at I = 0, i.e., the nominator becomes zero.
At this place sin[ 12 M |a||k| sin(↵ + ✏)] = 0 or 12 M |a||k| sin(↵ + ✏) = ⇡. This leads
to, with ✏ the full-width at half-maximum (FWHM) of the peak (see Fig. 5.51).

✏ = /(M |a| cos ↵) (5.15)

This relationship between the crystallite size, namely the domain size, and the width
of the di↵raction spot is known as Scherrer equation for 3DIM case. It just follows
from the relationship between the real space and reciprocal space.

Hence, the width of a di↵raction spot is proportional to 1/M and the area (=
width ⇥ hight) proportional to M . a · M1 or b · M2 is the radius of an area on
the surface in which the periodicity prevails, it is called the region of coherence.
Figure 5.51 shows the reciprocal-space rods broadened according to the limited
region of coherence.
The size of the di↵raction spot is inversely proportional to the number of scat-
terers in a periodic region. Actually, 25 to 100 atoms already give a di↵raction
CHAPTER 5. SURFACE STRUCTURE 118

Figure 5.50: For a finite domain size M , reciprocal-lattice rods become dashed
regions, which represent the FWHM of the intensity distribution 2⇡/M . There
are (M 2) additional maxima between the main intensities. Ref. [26].

pattern. Theoretically, the spot size goes to zero and the spot intensity diverges
as the region of coherence grows. Actually, electrons used in the experiment have
a coherence length as well. For the derivation of the di↵raction intensity we have
assumed electrons be plane waves, i.e., the coherence length is not limited. In prac-
tice, it is about 10 15 nm that puts an upper limit on the intensity of beams
di↵racted at a perfect surface.26
Coherence is a definite phase relationship. In reality, the incident wave bares
an uncertainty in its wave vector, in magnitude and in direction. It is | k| ⇠ U
in magnitude and k (= divergence) in direction. One speaks about temporal and
spatial coherence. The energy uncertainty U is due to the high temperature of the
filament required for thermionic emission. Thus wave chains are produced which are
only able to interfere within a short distance of the coherence length. Within this
length scale, they keep their phase relationship. The path di↵erence between two
waves may not be larger than this distance for interference. The angular divergence
( ) is related to the size of the emitting surface. If we call L is the width of a
perfect, monoenergetic beam then the coherence length CL is given by

CL = L
2 (1 + U/U )

Example:
26
M. Henzler, Appl. Surface Sci. 11/12, 450 (1982).
CHAPTER 5. SURFACE STRUCTURE 119

Source width = 0.5 mm, distance source - sample = 10 cm, U = 0 und = 1 Å


results in CL = 100 Å.
U ⇡ 0.4 eV for 2000 C W cathodes ( 32 kT )
U ⇡ 0.2 eV for 800 C LaB6 cathodes.
CL is a measure for the deviation of electrons from plane waves. If the average
coherence region (a · M1 ⇥ b · M2 ) on the crystal surface is smaller than CL, than the
form of the LEED spots is given by M1 M2 , otherwise by CL. Hence the information
is always limited.

Figure 5.51: LEED pattern of an Au(110)-2 ⇥ 1 surface at di↵erent primary


energies. Ref. [27].

2. Phase Transitions

The (110) surfaces of fcc metals, in particular, those of the noble metals show
a 2 ⇥ 1 reconstruction.27 Today thanks to several STM investigations a detailed
insight into the reconstruction has been gained. However, LEED is still most help-
ful describing the reconstruction and, if applicable, the nature of structural phase
transitions.28
Figure 5.52 shows on the Au(110) surface the integer- and half-order spots which
are all elongated in the 001̄ direction along which the half-order spots are visible.
This elongation is a measure for some existing disorder in that direction. Figure 5.53
displays on the right-hand side the real-space view into the fcc lattice. We recognize
27
D. Wolf et al., Surface Sci. 77, 265 (1978).
28
R. Feder et al., Z. Physik B 28, 265 (1977).
CHAPTER 5. SURFACE STRUCTURE 120

Figure 5.52: The bulk-derived (110) surface of Au is shown on the right-hand


side top. The real-space views of the 2 ⇥ 1 reconstructed Au(110) is shown on
the left-hans side. The reciprocal space view, as observed in LEED, consists
of integer-order (•) and half-order ( ) spots. Ref. [28].

the atomic chains in the 11̄0 direction. This surface would give a LEED pattern
consisting of the integer-order spots (•) as seen below. In reality, however, we also
observe half-order spots ( ). Such spots signal the doubling of the periodicity in
real space. One possible mechanism for the increased periodicity is referred to as
the missing-row model , as seen on the left-hand side of the figure. Every second
chain of atoms running in the 11̄0 direction would be missing. The transformation
of information from the Fourier space back to the real space is not unique, and
several other real-space structures are possible.
Figure 5.54 shows some models that are capable to account for the observed 2⇥1
reconstruction at the (110) surface of Au. Above the critical temperature of 650 K
both the half-order spots and the disorder disappear reversibly. Investigations of the
spot intensity indicate a second-order-like transition at the surface with a certain
critical exponent, as indicated on the right-hand side of in Fig. 1.9. A fit of the
temperature dependence of the order parameter allows the determination of critical
exponents and hence the universality class of the transition. The Au(110) 2 ⇥ 1
case is unique because the 2 ⇥ 1 spots due to reconstruction originates from a very
shallow surface sheet.29 Here, there is no counterpart of the transition in the bulk.
29
M.S. Daw and S.M. Foiles, Phys. Rev. Lett. 59, 2756 (1987) and references therein.
CHAPTER 5. SURFACE STRUCTURE 121

Figure 5.53: Some models of 2 ⇥ 1 reconstruction at the (110) surface of noble


metals. Each structure is shown both from the side and from the top. In all the
structures the nearest-neighbor distances are the same as in the bulk. Ref. [29].

As another example we mention c(2 ⇥ 2)-to-1 ⇥ 1 order disorder transition in Cu3 Au


single crystal.
In the case of Cu3 Au the structural phase transition takes place at the surface
as well as in the bulk. In this crystal, below 390 C the basal plane contains both
Cu and Au atoms, which occupy well-defined sites in the fcc lattice, depicted in
Fig. 5.55(b).30 Above 390 C, the unit cell is an average Cu/Au atom (one atom
per cell), i.e., there is a statistical distribution of Au and Cu atoms among all the
lattice sites, as illustrated in Fig. 5.55(a). In LEED the ( 12 12 ) spots disappear above
the transition temperature. We can assign the intensity of the fractional-order spots
to to an order parameter such that t to describe the transition.

Figure 5.54: Cu3 Au in its disordered and ordered phase. Ref. [30].

Detailed work has later shown that at temperatures moderately away from the
30
V.S. Sundaram et al., Surface Sci. 46, 653 (1974).
CHAPTER 5. SURFACE STRUCTURE 122

critical point this approach may have its validity. Near the critical point, however,
multiple-scattering contribution renders the di↵raction intensity not conserved and
modifies the values of critical exponents sufficiently to shift assignment from one
universality class to another. In short, a simple kinematic analysis of peak intensities
may lead to errors in the determination of critical values and one has to resort to
multiple scattering.

3. Influence of the Third Dimension


In the simplest case, we consider surface di↵raction by single scattering. This
simplification of 2DIM case requires two Laue conditions: a · (sR so ) = h and b ·
(sR so ) = k , with kR = 2⇡ sR and ko = 2⇡ so . We know that also the bulk
has a non-zero contribution on di↵raction. In 2DIM, the Laue condition for the
third dimension c · (sR so ) = l is not satisfied simultaneously for all energies.
Actually, due to strong inelastic scattering, only few surface layers take part in
di↵raction, and as a consequence, the influence of the third dimension is actually
only a modulation.

Figure 5.55: The normalized intensity of the specular reflection at Ni(100)


as a function of primary-electron energy. The indices for di↵raction orders
should be taken omitting the first zero. Ref. [31].

In Fig. 5.56 the intensity of the (00) beam is shown as a function of electron
CHAPTER 5. SURFACE STRUCTURE 123

energy. We observe strong modulation of intensity including secondary Bragg max-


ima owing to double scattering. Such results demonstrate that the kinematic theory
is not sufficient to fully describe di↵raction features from single crystal surfaces.31
The unknown value of the inner potential is responsible for shift of the forecasted
energy from the observed value.
The influence of scattering properties can also be realized considering the time-
reversal invariance, i.e., the theorem of reciprocity (Fig. 5.57). The reciprocity
theorem for scattering processes states that the amplitude for scattering from k to
k0 is equal to that for scattering from the reversed final direction k0 to the reversed
initial direction k.32 This is analogous to the fact that the reciprocal space has
an inversion symmetry. For LEED, this implies that the intensity of the di↵racted
beam does not change under time reversal I(k, k0 ) = I( k0 , k). As a consequence
of the time-reversal invariance, the di↵raction experiment doubles the symmetry of
the (00) beam and we observe, owing to the time-reversal invariance, symmetries
that do not exist in the crystal. The threefold-symmetric (111) surface of an fcc
structure and the (0001) surface of a hcp lattice both give rise to a sixfold-symmetric
rotation axis for the (00) beam. Similarly, the intensity of specular reflection from
the ideal (111) surface, from a twinned surface, and from a faceted but not twinned
surface all result in the same rotation diagrams.

k' k -k' -k Ik k' = I -k' -k

Figure 5.56: Time-reversal invariance.

4. Influence of the Temperature


When the temperature is raised, the atoms vibrate around their assumed positions
rj . The atomic positions are then given by

r(t) = rj + uj (t) (5.16)

for a cell with one single atom. r(t) is the variation in time of rj , the position of
the atom at T = 0. The interference factor
X
F= ei k·rj
(5.17)

and the intensity become


X ⇣ i k·r ⌘2
I / |hFi|2 = e j · hei k·u(t) 2
i. (5.18)
31
K. Christmann et al., Surface Sci. 40, 61 (1973).
32
H.E. Farnsworth, Phys. Rev. 49, 598 (1936); Phys. Lett. 36 A, 56 (1971).
CHAPTER 5. SURFACE STRUCTURE 124

In the expression for I the first factor is for a rigid lattice, i.e., at T = 0, and is
called Io . The second factor is a time average. We have then:
( k)2 ·hu2 (t)i
I = Io e , (5.19)

where
3h2 T
hu2 i = 2
· 2 (5.20)
4⇡ mkB ✓D
with ✓D , the Debye temperature, and m mass of the atoms.

Figure 5.57: The change in ✓D with the primary energy for di↵rac-
tion at Ni(100). Ref. [33].

( k)2 · hu2 (T )i is the Debye-Waller factor (DW). This consideration shows that
the intensity of di↵racted beams decreases exponentially with temperature, while
the spot size does not change.
Already in 1962 Germer and McRae have shown that the ✓D for the surface is
lower than that of the bulk.33 They have compared results of di↵raction experi-
ments from a Ni(100) surface at di↵erent electron energies (Fig. 5.58). Electrons at
di↵erent energies have di↵erent penetration depths and therefore carry information
from di↵erent depths below the surface of the sample. That way ✓D can be studied
as a function of sample depth. The bulk value is obtained from x-ray scattering.
Fig. 5.58 clearly shows that the Debye temperature for the surface is appreciably
lower than that of the bulk.
We also note that the DW factor depends on k. This implies that it is direc-
tional, and we have only access on the component u that is parallel to k. Hence,
by varying the experimental geometry we can measure di↵erent components of u.
33
A.U. MacRae and L.H. Germer, Phys. Rev. Lett. 8, 489 (1962); Ann. N.Y. Acad. Sci. 101,
627 (1963).
CHAPTER 5. SURFACE STRUCTURE 125

It has been determined that


hu2 i ? to the surface ⇡ 2hu2 i for the bulk.
hu2 i k to the surface ⇡ 1, 2 to 1,5 hu2 i for the bulk.
This means that the atoms move more e↵ectively normal to the surface compared
to their lateral motion (Fig. 5.59). Challenging exercise: anisotropic melting.

Figure 5.58: Intensity of electrons elastically scattered from a Ni(110) surface


close to the [11̄0] direction with = 0 (open circles) and = 70 (filled
circles) as a function of temperature. Electron energy is 40 eV. Di↵erent
components of u have di↵erent temperature dependences. Ref. [33].

5. Surface Melting
LEED provides a very convenient and readily accessible tool for studying melting
(Section 1.4) at crystal surfaces. It is extreme surface sensitive, and the information
is available over di↵racted intensities. Figure 5.60 illustrates schematically scatter-
ing of primary electrons with an intensity of Io at a clean surface on the left-hand
side with the di↵racted beam IG . The surface displayed on the right-hand side is
additionally covered with a thin structureless film of thickness d which represents
the already melted layers. The reflected intensity is I. The di↵racted intensity IGo ,
which is IG at T = 0, is attenuated due to several factors:

k)2 ·hu2 (T )i
I G = I Go e ( (5.21)
and
k)2 ·hu2 (T )i
I = I Go e ( e 2d/⇤
. (5.22)
CHAPTER 5. SURFACE STRUCTURE 126

I0
I0

IG I

Figure 5.59: Electron scattering at a clean and a homogeneously covered surface.

We assume that melting is homogenous over the entire surface and proceeds in a
layer-by-layer fashion. The logarithm of both sides yields
ln I = ln IGo DW 2d/⇤. (5.23)
Experimenting with Pb is practical, because Pb melts at conveniently attainable
temperatures. Frenken and van der Veen have observed, as shown in Fig. 5.61, that
there is a linear behavior of the DW factor up to 500 K and a strong increase above
this temperature.34 IGo is independent of temperature. Their results include:
a. The (111) surface is a bulk-like and closed surface. It melts at TM , the bulk
melting point.
b. The (110) surface is an open surface. This surface shows depending on the
direction di↵erent melting behavior: (10) beam is observable as high as the bulk
melting point TM = 600 K. Thus the [001] direction remains ordered up to TM . The
(01) beam shows, however, that there is a complete disorder in the [110] direction
at already 560 K (Fig. 5.61). Here we have anisotropic melting.
Theories about melting imply that d diverges if the temperature goes towards
TM . For long-range interactions
r
d / (TM T) (5.24)
with a critical exponent r = 13 .35
For short-range forces36
d / | ln(TM T )|. (5.25)

6. Stepped Surfaces
We often encounter terms like facetted or vicinal surfaces. These are surfaces with
uniform steps. The chemical activity of atoms at step edges is strongly enhanced,
that makes this subject interesting for catalysis. If these steps or terraces are peri-
odic, as displayed in Fig. 5.62, we have an additional condition for the interference
in form of modulation of the Bragg peak.37
34
J.W.M. Frenken and J.F. van der Veen, Phys. Rev. Lett 54, 134 (1985).
35
J.K. Kristensen and R.M.J. Cotterill, Phil. Mag. 36, 437 (1977).
36
J.Q. Broughton and G.H. Gilmer, Acta Metall. 31, 845 (1983).
37
B. Lang et al., Surface Sci. 30, 440 (1972).
CHAPTER 5. SURFACE STRUCTURE 127

Figure 5.60: Melting of surface layers on Pb(110). The inset is an expanded view of
the highest 10-K interval. The shaded band corresponds to the calibration uncer-
tainties in TM . The arrow indicates the surface melting point. Ref. [34].

Imagine N steps perpendicular to a. Each terrace has a coherence region of M


unit cells (Fig. 5.62). If the step hight is d, which is the basis vector for the steps:
M
X N
X
G= eip k·a
· eiq k·d

p=1 q=1

sin2 [ 12 M a · k] sin2 [ 12 N k · d]
|G|2 = · (5.26)
sin2 [ 12 a · k] sin2 [ 12 a · k]
In Eq. 5.26 the first term represents the ideal surface, while the second term de-
scribes the modulation within the di↵raction maxima as a function of electron
energy.
Figure 5.63(a) is a side view of the stepped surface shown in Fig. 5.62. The
di↵raction condition in real and reciprocal space is displayed Fig. 5.63(b). A typical
LEED pattern with split spots is shown in Fig. 5.64.

7. Domains
A partial region on the surface with a well-defined structure is called a domain. If
CHAPTER 5. SURFACE STRUCTURE 128

Figure 5.61: Schematic representation of the atomic structure of a


stepped Pt surface with (111) terraces. Ref. [37].

Figure 5.62: (a) Vicinal surface with monotonically increasing steps and constant
terrace size. (b) The sharp rods represent a lattice inclined to the average surface
with each terrace representing one lattice point. Their separation depends on the
terrace size. The broad rods represent the FWHM 2⇡/N a of the terrace structure
factor, which has (N 2) subsidiary maxima. Reflections occur where the product of
these two reciprocal lattices is nonzero. The terrace size is five-atoms wide. Ref. [26].

there is one single domain on the surface surrounded by the bare surface we speak
of an island . Islands form if the interaction between adsorbates are larger than that
between adsorbate and the surface as a result of adhesion forces.
There are three distinct types of domains:
a) If a domain is larger than the spot size of the electrons, we obtain an individual
di↵raction pattern from each domain.
b) If a domain is smaller than the beam diameter, but larger than the coherence
length of electrons, the resulting LEED pattern is a superposition of individual
pattern (not a coherent addition) from domains that are illuminated by the electron
beam.
c) If the typical domain sizes are smaller than the coherence length of electrons, we
CHAPTER 5. SURFACE STRUCTURE 129

Figure 5.63: Di↵raction at a periodic array of steps on a Pt surface. Ref. [37].

have interference of beams scattered from di↵erent domains located in a region as


large as the coherence length. The resulting LEED pattern is a coherent addition
the beams from individual domains.

Figure 5.64: An antiphase domain structure formed from adjacent occupancy of the
two equivalent adsorption sublattices of an ordered 2 ⇥ 1 overlayer. From AZ.

Any commensurately adsorbed monolayer that has a superlattice unit mesh


larger than the substrate will form translational and possibly also rotational (de-
pending on the symmetry of the unit mesh) antiphase domains. Figure 5.65 is a
schematic representation of the former, while the latter is presented in Fig. 5.67,
both schematically. In each region a 2 ⇥ 1 reconstruction prevails, while the super-
structure of each adjacent region is shifted by half a unit in one direction. Thus,
a fractional monolayer may exist in the form of antiphase islands at some finite
temperature if there is a net attractive interaction between the adsorbate atoms.
CHAPTER 5. SURFACE STRUCTURE 130

There may be several reasons why a fractional monolayer might exist as islands, the
most important being kinetic limitations and point or line defects on the substrate
that act as nucleation sites.

Figure 5.65: (Left) clean Cu(100) surface. Observation of the antiphase


domains due to successive oxygen adsorption. P. Aebi et al., unpublished.

Figure 5.66 presents LEED patterns obtained from Cu(100) surface. The clean
surface delivers the pattern on the left-hand side. With 1 ML oxygen, a reconstruc-
tion is observed typical of c(2 ⇥ 2) structure. Further oxygen adsorption introduces
still additional fractional-order spots. The new surface phase is accounted for by a
rotational antiphase-domain structure, as illustrated in Fig. 5.67.

Figure 5.66: A model showing the c(2 ⇥ 2) antiphase domains of oxygen on Cu(100).

Quasicrystals lack periodicity yet display characteristics of long-range orienta-


tional order. The ternary alloy Al-Pd-Mn has icosahedral symmetry in the bulk,
and its pentagonal surface shows bulk-terminated behavior, namely it is not recon-
structed and appears as expected from the bulk structure. Figure 5.68 shows on the
left-hand side the di↵raction pattern from the clean surface. Compared to crystals,
where one observes di↵raction spots due to only few Brillouin zones, quasicrystal
surfaces display enumerable number of spots. In fact, the longer one chooses the
CHAPTER 5. SURFACE STRUCTURE 131

recording time for di↵raction intensities, the more spots appear on the collector
screen. We notice that the azimuth distribution of the pattern has 2⇡/5 symmetry.
p
The spot distribution on radial directions obeys the golden ratio, ( 5 + 1)/2, so
does the intensity distribution of all the spots.

Figure 5.67: LEED patterns obtained at 55 eV from (left) a clean


pentagonal surface of icosahedral Al-Pd-Mn, (right) after exposure
to O2 at 700 K and subsequent annealing for 2 h. Ref. [38].

When the quasicrystal surface is exposed to oxygen at elevated temperatures,


Al segregates from the bulk and enriches the surface.38 In contact with oxygen, Al
readily oxidizes and the LEED pattern transforms to one observed on the right-
hand side. This pattern mainly shows 30 equally bright and equally distributed
spots. They represent five distinct groups of hexagonal spots, typical of the (111)
surface of Al2 O3 . This observation is interpreted in terms of five distinct crystalline
Al2 O3 domains formed on the pentagonal surface of Al-Pd-Mn. Using Eq. 5.15 we
estimate the Al2 O3 islands as large as 3 nm.

5.3.1.3 Calculation of Di↵racted Intensities


The symmetry of the atomic order at the surface is studied by observing the di↵rac-
tion spots in 2DIM; we study the projection of the unit cell in the reciprocal space.
Additional relevant information can be extracted investigating the scattered inten-
sity, I(E), I(✓), I( ). A comparison of these measurements with detailed scattering
calculations on model structures facilitates detailed knowledge on interatomic dis-
tances, vibration amplitudes, and other structural features not accessible to simple
inspection of the di↵raction pattern.
38
J.-N. Longchamp et al., Phys. Rev. B 76, 094203 (2007).
CHAPTER 5. SURFACE STRUCTURE 132

The process consists of the assumption of a model structure and calculation of


the di↵raction intensity in order to find some agreement. The structure parame-
ters are varied and calculations are repeated until some agreement is found. This
agreement is a necessary but not sufficient condition for the validity of the model
structure and all the assumed ingredients of the computation. In the computations
one solves the scattering problem. The steps include
- using an e↵ective potential usually from a band theory
- slicing the sample in slabs parallel to the surface
- calculating multiple scattering in each slab
- adding up all the contributions considering multiple scattering between the slabs
and attenuating the signal from deeper slabs by the mean free path ⇤ appropriate
for the energy.

!"#$$% !"#$%
) )*)'
&(
θ(
θ'

&'

Figure 5.68: Refraction of electrons while entering the solid.

One has thereby to consider that electrons are faster in the solid than in vacuum
owing to the inner potential (Figs. 5.69 und 5.70). Consequently, the normal com-
ponent of the electron velocity changes and the tangential component is conserved
such that vo sin ✓o = v1 sin ✓1 . Energy conservation dictates:
1
e(U + Uo ) = mv12 (5.27)
2
1
eU = mvo2 (5.28)
2
Snell’s law gives the index of refraction:
s
sin ✓0 v1 U + U0
n= = = . (5.29)
sin ✓1 v0 U
Note that U0 > 0, U + U0 > U .
After considering all the details, e.g., the inner potential, mean free path of
the electrons involved, the Debye temperature of the surface and bulk, the crystal
CHAPTER 5. SURFACE STRUCTURE 133

θ'

Figure 5.69: The apparent scattering angle due to the refraction of electrons.

structure and the reconstruction at the surface, one uses realistic potentials and
calculates the measured intensity curves using multiple scattering. In the case of
adsorbates on the surface, one additionally has to take into account where exactly
these species sit on the surface. The computed curves are compared with the
experimental ones, and the input parameters to computations are modified until
the results converge to an acceptable agreement. The input to calculations are then
taken as most probable values representing the real case.
CHAPTER 5. SURFACE STRUCTURE 134

Figure 5.70: (Left) Measured di↵raction intensities of the 00 beam at a W(100)


surface as a function of energy. (Right) Comparison of calculated and experimental
di↵raction intensities for a Ni(100) surface covered with oxygen (right). Ref. [23].
Chapter 6

Inner Shells

6.1 Introduction
The electronic states near the Fermi level, EF , are responsible for the chemical
and electronic, i.e., macroscopic, behavior of elements and compounds, while the
binding energy EB of inner-shell states, the core levels, is a finger print of the iden-
tity and chemical state of the atom. The Koopmans’ theorem postulates that the
binding energy of the core levels of the atoms corresponds to eigenvalues of one-
electron Hamiltonian ✏i in the ground state.1 Because each atom has a di↵erent
nuclear charge Z and a di↵erent number of electrons, the resulting binging energy
of a particular shell is also di↵erent for each atom. An assessment of EB provides
a chemical analysis of the material, it is a finger print of the element. We can
additionally determine the chemical state of the atom by considering the shift in
EB relative to value in the neutral atom. In most of the cases, valence electrons
are also involved in these measurements and thus we obtain simultaneously infor-
mation on the electronic structure. This is the basis of the spectroscopic methods
which investigate the transition between the inner shells of the atoms, while atomic
species are identified by the binding energy of the inner shells. We may call such
experiments with the title Inner-Shell Spectroscopies. We will deal with
1. the order in which the shells are filled with electrons,
2. the occupation of the shells with electrons,
3. the binding energy of the shells.

6.1.1 Filling the shells with electrons


The filling order of the electronic states is determined in principle by relativistic
e↵ects. Figure 6.1 illustrates a classical recipe for the filling order of the shells that
delivers acceptable results. Following this recipe, cesium with its 55 electrons has
the configuration:
1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p6 5s2 4d10 5p6 6s1
1
T. Koopmans, Physica 1, 104 (1934).

135
CHAPTER 6. INNER SHELLS 136

Figure 6.1: The gradual filling of the electronic states with an additional electron as
the nuclear charge Z increases. The rows represent the shells with the main quantum
number n and the columns stand for the orbital quantum number `. The additional
electron occupies shells along the diagonal (constant n+`) with the smallest possible
n + `. The numbers at the opposite axes represent the occupation number of each
shell.

6.1.2 Occupation of the shells


The number of electrons in a shell is called the multiplicity, viz. degeneracy, of each
shell. A state with |n, ` > has the total multiplicity 2(2` + 1). The factor 2 outside
(. . . ) takes care of the spins.

`=0 s 2 degenerate states `=2 d 10 degenerate states


`=1 p 6 degenerate states `=3 f 14 degenerate states

Here, s and ` are good quantum numbers.


There are two mechanisms that partially lift the degeneracy and lead to a splitting:
1. Spin-orbit coupling which is a relativistic e↵ect. For the ith electron, it is:
Zie↵ (ri )
2
↵ (` · s), (6.1)
ri3
where ↵ is the fine structure constant, ` and s are orbital and spin momentum
operators, Zie↵ is the e↵ective nuclear charge felt by the ith electron. Spin-orbit
CHAPTER 6. INNER SHELLS 137

coupling is large for large Z, large `, and small r. Now, ` und s are no longer good
quantum numbers, but the total orbital momentum j = ` + s is. For this case, the
multiplicity is given by (2j + 1).

`=0 j = 1/2 K-shell (K, LI , MI ) s multiplicity 2


`=1 j = 1/2 LII , MII -shell p1/2 multiplicity 2
`=1 j = 3/2 LIII , MIII -shell p3/2 multiplicity 4
`=2 j = 3/2 MIV , NIV -shell d3/2 multiplicity 4
`=2 j = 5/2 MV , NV -shell d5/2 multiplicity 6
`=3 j = 5/2 NV I , OV I -shell f5/2 multiplicity 6
`=3 j = 7/2 NV II , OV II -shell f7/2 multiplicity 8

Let us now recall Hund’s rules that put the energies into correct order:
a. The state with the maximum total spin has the smallest energy.
b. The state with the maximum orbital momentum has the smallest energy.
c. If a shell is less than half full, J = |L S| and J = |L + S|, if it is more than
half full.
2S+1
It is customary to denote a state with its term symbol in the form of: LJ
2 3
Carbon atom with its p electrons in the unfilled outer shell has the P0 ground
state, while the ground state for oxygen with its p4 electrons is 3 P2 .
2. Exchange splitting between a singlet and a triplet configuration. This splitting
does not originate from magnetic dipole forces between the spins, but depends on
the charge distribution that is modified by the Pauli principle according to the spin
direction.
In summary, the nuclear charge Z is responsible for the chemical identity. Di↵erent
Z values change the binding energy of electrons and the structure of the electronic
shells. The L S-coupling determines the fine structure of the shells.

6.1.3 Shell Binding Energy


The Koopmans’ theorem provides the binding energy of the core levels of the atoms
as a result of one-electron eigenvalues in the ground state. This theorem is difficult
to test in an experiment because only excitations are accessible in experiments not
the ground state. At most we can examine how close the measured value agrees
with the calculations. The discrepancy between these values is a measure of the
response function of the system to our perturbation. Generally, this is referred to as
relaxation energy and its magnitude depends on electron correlation in the excited
state. There is no general recipe, but related materials can roughly be grouped into
similar categories. The easiest case is the nearly-free electron systems which follow
the Koopmans’ theorem where the so-called one-electron approximation prevails.
Correlated electron systems with d or f electrons fall into the opposite category.
CHAPTER 6. INNER SHELLS 138

6.2 Ionization of Inner Shells


We can produce a hole in an inner shell by exciting an electron into a state above
EF using energetic electrons, ions, or photons. We add energy to the system and
promote it to an excited state.

Figure 6.2: Experimental electron binding energies. See e.g. Photoemission in


Solids I, ed. M. Cardona and L. Ley, Springer, New York, 1978, p. 265.
CHAPTER 6. INNER SHELLS 139

6.3 X-ray Absorption Spectroscopy


An atom is excited by x-rays, but not ionized, e.g., 3p6 3dN ! 3p5 3dN+1 . In
a typical absorption experiment, photons with the intensity Io traverse the speci-
men of thickness d. Some of the photons are absorbed, the rest, IT , is detected,
as shown in the Fig. 5.39. The absorption coefficient µ is a function of h⌫ and
depends on the density of the particles, nc , in the specimen and the ionization
cross section, (h⌫). The optical excitation is selective with respect to the energy

(h⌫ = EN EN ), momentum ( ` = ±1), and the selection rules (symmetry of
the wave functions, polarization of light, etc.). The method is referred to as x-ray
absorption spectroscopy (XAS).

Figure 6.3: The dependence of the x-ray absorption coefficient on photon


energy, and a typical experiment.

For low energies, µ has a monotonous dependence on photon energy h⌫. When-
ever h⌫ is sufficiently high to excite an electron from an occupied shell to a state
above EF , µ shows an abrupt increase. This is called an absorption edge. For still
higher energies, µ drops monotonously with h⌫ until EB of the next shell is reached
where µ shows another abrupt increase. Thus, the absorption coefficient displays
a saw-tooth like energy dependence. The experiment, absorption process and a
typical result are illustrated schematically in Fig. 6.3.
Herewith we learn that:
1. The absorption edges correspond to EB of the shells. This quantity is defined
with respect to EF , illustrated in Fig. 3.2.
2. Depending on the chemical environment of the absorbing atom, EB can be shifted
up to a few eV due to the change of the screening of the nuclear charge as a result
of electron transfer at the atomic site – this is called the chemical shift.
CHAPTER 6. INNER SHELLS 140

Figure 6.4: Experimental setup and data-acquisition method in XAS. Ref. [2].

Figure 6.4 shows on the top panel an experimental setup for XAS with a spatial
resolution on the specimen in the nanometer range.2 X-rays are energy selected
and focussed using a Fresnel plate onto the sample surface. The sample (lower
left) is mounted on a 50-µm thin SiN holder that allows heating in a selected gas
ambient. The sample holder is mounted on piezo-electric drivers that allow accurate
positioning of the sample in the photon beam. The photomultiplier detects the
transmitted light IT . In the experiment, for a preset photon energy, IT is determined
at di↵erent areas A, B, or C of the sample. Finally, complete XAS spectra in the
measured photon energy range is assembled using the acquired data for the areas
A, B, or C. This method allows us to observe catalytic processes on di↵erent regions
2
E. de Smit et al., Nature 456, 222 (2008).
CHAPTER 6. INNER SHELLS 141

of the specimen with an unprecedented spatial resolution.

Figure 6.5: Chemical environment of Fe and O during the Fischer-


Tropsch process. Ref. [2].

Figure 6.5 depicts some results obtained on an iron oxide surface. On the left-
hand side the photon absorption around the iron site is analyzed by observing the
L2 and L3 core-level regions. On the right-hand side XAS around the oxygen site is
observed in order to identify the chemical processes. Finally the maps are assembled
in the middle panel. Thus, nonoscale chemical imaging of a working catalyst is
achieved. Here, Fischer-Tropsch synthesis (Eq. 1.2) is observed on nanometer-size,
active ion-oxide components.

6.4 Excitation of an Atom by Electrons


An excitation experiment performed with electrons displays several advantages.
Electrons are easy to focus and their energy can be changed readily. Experiments
are cheaper and can easily be combined with others. Data analysis can be am-
biguous because there are two electrons in the final state that share the excess
energy arbitrarily. It is a further problem that we do not exactly know whether the
momentum-selection rules apply. Still, electron-excitation experiments present the
advantage that they can be performed in the “home” laboratory.
CHAPTER 6. INNER SHELLS 142

An electron with the momentum p (= h̄k) loses energy Ee = (p2i p2f )/2m in an
atomic collision. The atom gains the energy Ea = Ef Ei .3
We apply the 1st Born Approximation where V (r) is a week perturbation and i
and f are plane waves. The probability for the transition to an excited state is
given by the Fermi Golden Rule
2⇡ Ei ,pi 2
d! = |MEf ,pf | (Ee Ea )dpfx dpfy dpfz . (6.2)

with M = h f f |V (r)| i i i.

f, iqare the wave functions of the atom before and after the scattering.
m iki ·r 1 ikf ·r
i = pi
e and f = (2⇡h̄) 3/2 e are wave functions of the incident and reflected
electrons, while the scattering angle is between ki and kf .
Coulomb interaction between the incident electrons and the atom is given by:
Z
Ze2 X e2
V (r) = (6.3)
r a=1 |r ra |
The first term is the attraction between the e and the nucleus, while the second
term standsZ for the e-e repulsion.
d
We use d! = to obtain the scattering cross section.
dE
p2f dpf d⌦ = 12 pf d(p2f )d⌦ (in spherical coordinates).
!
d ⇡Z p2i p2f
= |MEEi,p i 2
f ,pf
| Ea pf d(p2f )d⌦
dE h̄ 2m
2m⇡ Ei,pi 2
= |MEf ,pf | pf d! (6.4)

after integration over p2f
q
with MEEi,p i
f ,pf
= 1
(2⇡h̄)3/2
m
pi
h fe
ikf ·r
|V (r)| eiki ·r i i
| {z }
M0
d m2 pf
= 2 4 |M 0 |2 d⌦ (6.5)
dE 4⇡ h̄ pi
R
where M 0 = h f | e iq·r V (r)dV | i i with q = ki kf with q the momentum
transfer. Using Eq. 6.3
Z 2 Z Z
0 iq·r Ze iq·r
X e2
M =h f | e dV | ii h f | e dV | i i. (6.6)
r a=1 |r ra |
The first expression is the Fourier transform of the Coulomb potential 1/r. Owing
to the orthogonality of f and i it is zero. The second expression is the Fourier
transform of 1/r with a delay of ra .
! ✓ ◆
1 iq·ra 1 4⇡ iq·ra
FT =e FT = 2e
|r ra | r q
3
H.A. Bethe, Ann. Physik (Leipzig) 5, 325 (1930).
CHAPTER 6. INNER SHELLS 143

4⇡ 2 PZ
Then M 0 = q2
eh f | a=1 e iq·ra
| i i and therefore
!2
d e2 m 4ps X
iq·ra 2
= |h f | e | i i| d⌦. (6.7)
dE h̄2 q 4 pinc

q = ki kf results in q 2 = ki2 + kf2 2ki kf cos and qdq = ki kf sin d = ki kf d⌦


2⇡
.

Transformation d⌦ ! dq allows us to write for one electron


!2
d2 e 2 m2 1 iq·ra 2
= 8⇡ |h f |e | i i| (6.8)
dEdq h̄2 q 3 ki2
!2
e2 1 iq·ra 2
= 8⇡ |h f |e | i i| (6.9)
h̄2 vi q3

with vi = h̄ki /m.


If q · ra ⌧ 1, we can expand e iq·ra = 1 iq · ra 12 (q · ra )2 + ...
Since i and f are orthogonal, the first expression has no contribution. Using
q = ✏ˆ · |q|, the second expression reads
!2
d2 e2 1 2
= 8⇡ |h f |ˆ✏ · ra | i i| . (6.10)
dEdq h̄2 vi q

This expression is proportional to Y1 ( , ) and describes the electrical dipole tran-


sitions (`0 = ` ± 1).
d 2
Using the wave functions f and i , dEdq can be calculated.
If the experiment is performed in a particular interval [qmin , qmax ]:
⇣ ⌘
N (E) / dE d
/ log qqmax
min
|h f |ˆ✏ · ra | i i|2 .
In analogy to the excitation with photons (ê corresponds to the direction of light
polarization) the absorption coefficient is given as µ(E) / ↵ (E) / |h f |ê·ra | i i|2 .
The third term of the series expansion
1 1 2 2 1 2 2 q2
2
(q · ra ) 2 = 2
q z = 6
q r 6
(3z 2 r2 )
contains terms that are proportional to Y0 ( , ) and Y2 ( , ) and hence describe
monopole (`0 = `) and quadrupole (`0 = ` ± 2) transitions. These are generally
forbidden in optics. Higher-order terms describe multipole transitions.
We can define a generalized oscillator strength, GOS , as
d2
GOS = q12 |h f |e iq·ra | i i|2 and dEdq / 1q GOS
If Ei and Ef are known, ki and kf are determined, while q changes with the scat-
tering angle . In fact, q varies between two extreme values:
For min = 0 ! qmin
p q ⇣p p ⌘ p
qmin = ki kf = c Ei c Ef = c Ei Ei Ee with c = k/ E.
1 kf
For max = cos ki
! qmax
CHAPTER 6. INNER SHELLS 144

Figure 6.6: Minimum and maximum values of momentum transfer q.

Figure 6.7: Maximum value of the momentum transfer qmax is only


related to the energy transfer Ee during a collision, while the minimum
value qmin is limited also by the primary energy Ei .

q q p
qmax = ki2 kf2 = c Ei Ef = c Ee .
As summarized in Fig. 6.7, qmax only depends on the absolute value of the energy
loss, while qmin is a↵ected by the primary energy as well. GOS describes to a great
extend those transitions that occur at the smallest values of q. Therefore, results
of experiments performed at higher primary energies Ei are expected to compare
agreeably with those obtained in XAS.
CHAPTER 6. INNER SHELLS 145

6.4.1 Electron Energy-Loss Spectroscopy


GOS is the basis of electron energy-loss spectroscopy, EELS , that contains several
characteristics analogous to XAS in the energy region of the core states.

Figure 6.8: On the left, x-rays excite a core level to states above EF . For constant
matrix elements, the absorption coefficient µ maps the unoccupied density of states.
On the right, an electron with a kinetic energy Ei loses E by promoting core-
level electrons into states above EF , just into the same states as in XAS. Hence the
measured GOS is closely related to µ, shown as shaded area. Ref. [4].

Figure 6.9: EELS from selected compounds. The similarity at the


nitrogen edge for ZrN and NbN is due to the local density of unoccupied
states at the nitrogen site. The comparison between NbC and NbN
reveals details of Nb-derived density of states above EF . Ref. [5].
CHAPTER 6. INNER SHELLS 146

The close analogy between XAS and EELS is indicated schematically in Fig. 6.8.
Roughly, both experiments carry information about the density of unoccupied states
at the site of the absorbing atom.4 While the absorption coefficient is clearly defined
by the `-projected density of states, no clear evidence is presently available for
the electron excitations. At sufficiently high electron energies (Fig. 6.7), electron
excitations can be described within the dipole approximation, yet some mixing from
non-dipole channels cannot absolutely be excluded. Nevertheless, for the sake of a
rough inspection of the density of states, EELS is a very versatile tool. In Fig. 6.9
the measured EELS K-edge spectra of the nonmetals are compared with the total
density of states above EF .5

Figure 6.10: XAS spectra at Fe 2p- and O 1s-core levels of di↵erent


Fe-compounds. Ref. [6].

The fingerprinting of atoms in di↵erent coordinations is illustrated in Fig. 6.10.


XAS spectra at Fe 2p- and O 1s-core levels show salient features that allow us to
4
M. De Crescenzi and G. Chiarello, J. Phys. C 18, 3595 (1985).
5
J. Pflüger et al., Solid State Commun. 55, 675 (1985)
CHAPTER 6. INNER SHELLS 147

uniquely identify the Fe compound. On the left-hand panel we observe that both
2p3/2 - and 2p1/2 -components are measured. With additional consideration of O
1s-core levels even the ↵- and - phases of Fe(OH)O become distinguishable.6

Figure 6.11: The cubic spinel structure of Fe3 O4 with the two di↵erent sublattice
sites. EELS data obtained in TEM from either sublattice use the density of states
above EF as a fingerprint for di↵erentiating between two Fe ions. Ref. [7].

As a further example for EELS, Fig. 6.11 illustrates the structure of magnetite in
the cubic spinel structure. It consists of tetrahedral and octahedral sublattices. The
tetrahedral sites are occupied with Fe3+ ions and the octahedral sites with Fe2+ ions.
In TEM standing-wave fields are created that have maxima either on tetrahedral
or octahedral sites, depending on the tilt angle of the sample. Thus, di↵raction
geometries are used to set up standing-wave fields with maxima at di↵erent sites.
In Fig. 6.11(a) octahedral sites are selected and in Fig. 6.11(b) tetrahedral sites.
With this site selection, the EELS data, shown on the right-hand side, originate
either from (a) Fe2+ ions or (b) Fe3+ ions.7
Inclusion of di↵erent metals in carbon structures results in a wealth of intriguing
phenomena. These include intercalated graphite, when metal ions are introduced
between the carbon sheets. Similarly, metals in nanotube structures show ballistic
conduction. Superconductivity is observed in metallofulleranes.
6
S.-Y. Chen et al., Phys. Rev. B 79, 104103 (2009).
7
J. Taftø and O.L. Krivanek, Phys. Rev. Lett. 48, 560 (1982).
CHAPTER 6. INNER SHELLS 148

Figure 6.12: (Left) A high-resolution TEM image and schematic illustration of the
Gd-metallofullerane in a nanotube. The scale bars are 3 nm. (Right) An EELS
spectrum of the same specimen measured in tandem in TEM within 35 ms. Ref. [8].

Figure 6.12 shows on the left-hand side (A) a conventional high-resolution


TEM image and (B) a schematic representation of the structure. A chain of Gd-
metallofulleranes is encapsulated in a single-walled nanotube. On the right-hand
side of the figure, a typical EELS spectrum is shown acquired within 35 ms.8 We
observe the Gd N -edge and C K-edge. This experiment proves that in a modern
TEM one can detect single atoms within ms and identify them uniquely.

6.5 X-Ray Photoelectron Spectroscopy


We have introduced photoemission spectroscopy in Chapter 3 in detail in order
to extract electronic information on the sample. In short, the process involves
the excitation of electron states to a final state using photons with a known photon
energy h⌫ and measuring the kinetic energy of emitted electrons outside the sample.
Figure 6.13 shows schematically on the left-hand side the energy diagram of
band states and some core levels in a metal in the initial state. According to the
one-electron approximation, the excitation involves the promotion of every electron
state by h⌫, while the response of the rest of the electron gas is neglected. The
schematic energy distribution of the photoelectrons on the right-hand side just
implies some broadening of the core levels and smoothening of the distribution at
EF due to Fermi-Dirac function (Section 2.1.3).
The electronic states near EF are responsible for the macroscopic properties of
the sample, while the binding energy EB is taken as a fingerprint for the chemical
identity of the atom. In the following we will observe how band states become
core levels when we move from left to right in the periodic table by introducing
each time one electron into the system. Subsequently, we will give examples for the
utilization of core levels in chemical analysis. This is the reason why photoemission
experiments are referred to as Electron Spectroscopy for Chemical Analysis, ESCA,
8
K. Suenaga et al., Science 290, 2280 (2000).
CHAPTER 6. INNER SHELLS 149

Figure 6.13: The photoexcited electrons escape into vacuum, and their energy dis-
tribution is taken as a replica of electron states in the metal prior to excitation.
This method is also referred to as X-ray Photoelectron Spectroscopy, XPS .

in some laboratories.
Palladium atom has 4d10 configuration. In the metallic form it has the highest
density of states at EF , so far known, as seen in Fig. 6.14 on the top spectrum.
Hence, it can be used to easily determine the position of EF and calibrate the
experimental energy scale. As one additional electron is introduced to the system,
we obtain Ag with a completely filled 4d band, which appears at about 5 eV below
EF . A splitting starts to already develop in the band. Ag is a good conductor owing
to its 5s electrons at EF . As we go gradually to the right in the periodic table,9 we
observe (a) how the 4d band further moves to higher EB , (b) that the spin-orbit
splitting in the 4d band increases in energy making the splitting more pronounced,
(c) how the sp states develop at and near EF .
Figure 6.15 illustrates Al 2p spectra on the left as a function of oxygen exposure
on the (100) crystal face measured at an excitation energy of 130 eV. Similar ex-
9
R.A. Pollak et al., Phys. Rev. Lett. 29, 274 (1972).
CHAPTER 6. INNER SHELLS 150

Figure 6.14: The behavior of electron states in the transition metals Pd to Te.
Ref. [9].

periments on the Al(111) surface is shown in the middle. The choice of the photon
energy results in a kinetic energy of the photoexcited electrons with a very shallow
escape depth. Hence, the spectra are extremely surface sensitive. Note the evo-
lution of the 2.6-eV feature on both surfaces due to the partial oxidation of the
surface. The additional chemical shift by 1.4 eV corresponds to the chemisorbed
phase on the (111) surface as shown in (a) on the right-hand side in contrast to
the (b) chemisorbed phase.10 This example shows that, besides determining the
chemical identity of surface atoms, XPS reveals their chemical state. The spectral
intensities are used as a measure for the elemental concentration. Hence, XPS is
well suited for rapid chemical fingerprinting of surface species.
Ytterbium is the last element of the rear-earth series. The completely occupied
4f states are located few eV below EF and show a typical spin-orbit splitting into 5/2
and 7/2 components. A Yb film, grown on a clean and flat Mo substrate, consists
of atoms which have di↵erent neighbors. Yb-atoms at the Mo substrate have both
Yb and Mo neighbors, while those at the surface have missing neighbors. Other Yb
atoms, the “bulk” atoms, have a complete Yb coordination. Figure 6.16 displays
10
P.S. Bagus et al., Phys. Rev. B 44, 9025 (1991).
CHAPTER 6. INNER SHELLS 151

Figure 6.15: Oxygen in the chemisorbed and oxide state produces di↵erent core-level
shifts on Al(100) and (111) faces, while it is di↵erently incorporated at the surface.
Ref. [10].

XPS spectra from a Mo substrate with 2, 3, and 4 ML of Yb. Each spectrum is


decomposed into three components, the “interface”, the “bulk”, and the “surface”
components for both spin-orbit components.11 A 2-ML film has no “bulk” atoms
and consists of “interface” and “surface” atoms. The amount of “surface” atoms
does not change for thicker films, so the spectral intensity remains constant. The
amount of “interface” atoms does also not change as the film gets thicker, but
its intensity decreases due to electron attenuation through the thicker film. The
number of “bulk” atoms increases linearly with the film thickness, so we observe
a marked increase in the “bulk” component. This example shows that di↵erent
atomic coordinations are well resolved in XPS.
Zn is a 3d transition metal with completely occupied 3d band. The photoelectron
spectrum shows the energy region near the 3s and 3p regions for ZnF2 , as depicted
in Fig. 6.17. A similar XPS spectrum is shown on the right-hand side for FeF2 .12
We observe a clearly-resolved splitting of the 3s level of Fe. There is no mechanism
that splits the s level in the initial state as documented by th spectra for ZnF2 .
Hence the observed splitting of the 3s level is a final-state e↵ect as long as the atom
possesses an unfilled d shell. The splitting is generated by the alignment of either
spin direction of the s state left behind in the photoemission process. The spin of
11
N. Mårtensson et al., Phys. Rev. Lett. 60, 1731 (1988).
12
L. Ley et al., Phys. Rev. B 8, 2392 (1973).
CHAPTER 6. INNER SHELLS 152

Figure 6.16: Di↵erent neighborhoods of Yb induce di↵erent core-


level shifts on the 4f levels. Ref. [11].

this electron can namely be aligned parallel or antiparallel with the spins of the 3d
CHAPTER 6. INNER SHELLS 153

Figure 6.17: Splitting of the 3s state in Fe which is missing in Zn. Refs. [12,13].

electrons of iron.13 Thus we observe the s-d exchange coupling of the final state.
The splitting is a measure of the local magnetic field of d electrons influencing the
s level. This process is in fact the smallest magnetometer we can think of. The
material does not have to be ferromagnetic, the only requirement is an unfilled d
shell. Hence, XPS can be used to measure the local magnetic field.

6.6 Relaxation Processes


The ionized core level has a finite life time after which it relaxes back to its ground
state. The life time ( = h̄/⌧ ) is determined by the sum of all mechanisms that
makes the excited state relax. The life time is very long for atoms, and relatively
short for systems with overlapping electron states. Essentially, there are two di↵er-
ent relaxation processes, the dipole radiation and electronic recombination, called
Auger electron emission.

6.6.1 Dipole Radiation


An electron from a higher state fills the hole, and the energy gained is emitted as
a quantum of light. Since this process is a dipole radiation, appropriate selection
rules are obeyed:
h⌫ = EB1 EB2
` = ±1
j = 0, ±1
The emission of x-rays is used in research as a spectroscopic tool, in electron mi-
croscopy for elemental analysis, and in crystallography for structural analysis.
Figure 6.18 illustrates schematically di↵erent x-ray emission processes. For the
K series of emission, the initial hole in the K shell is transferred to higher states
13
B.V. Veal and A.P. Paulikas, Phys. Rev. Lett. 51, 1995 (1983).
CHAPTER 6. INNER SHELLS 154

Figure 6.18: Principal structure of x-ray K- and L-series emission. D.S. Urch, In
Electron Spectroscopy: Theory, Techniques and Applications, ed. C.R. Brundle and
A.D. Baker, Vol. 3, Academic, New York, 1979, p. 3.

observing the selection rules. The emitted photons are designated as K↵ and K
series. Using the table for binding energies of elements in Fig. 6.2 we can estimate
some frequently-used x-ray transitions. In laboratory XPS measurements mostly Al
or Mg anodes are used. The corresponding photon energies can easily be calculated
for AlK↵1,2 1560 74 = 1486 eV and for MgK↵1,2 1305 52 = 1253 eV. In XRD
experiments mostly a Cu anode is used, because copper conducts the power load in
high-intensity sources e↵ectively. The Cu lines are well separated by 20 eV: K↵1,2
8979 951 = 8028 eV and 8979 931 = 8048 eV. Using Eq. 3.1 we find that these
energies correspond to 1.575 Å and 1.541 Å wavelength which are comparable with
interatomic distances and hence well suited for XRD work.
The dipole radiation is called the fluorescence. While most of the observed
fluorescence lines are normal, certain lines may also occur in x-ray spectra that,
at first sight, do not abide to the basic selection rules. These lines are called
forbidden lines; they arise from outer orbital levels where there is no sharp energy
distinction between orbitals. As an example, in the transition elements, where
the 3d level is only partially filled and strongly interacting with the 3p levels, a
weak forbidden transition (the 5) is observed. A third type are satellite lines
CHAPTER 6. INNER SHELLS 155

arising from dual ionization. Following the ejection of the initial electron in the
photoelectric process, a short, but finite, period of time elapses before the vacancy
is filled. This time period is called the lifetime of the excited state. For the lower
atomic number elements, this lifetime increases to such an extent that there is a
significant probability that a second electron can be ejected from the atom before
the first vacancy is filled. The loss of the second electron modifies the energies of
the electrons in the surrounding subshells, and thus x-ray emission lines with other
energies are produced. For example, instead of the K↵1,2 line pair, a double-ionized
atom will give rise to the emission of satellite lines such as the K↵3,4 and the K↵5,6
pairs. Since they are relatively weak, neither forbidden transitions nor satellite lines
have great analytical significance; however, they may cause some confusion in the
qualitative interpretation of spectra and may sometimes be misinterpreted as being
analytical lines of trace elements.

Figure 6.19: XRF spectra from an Al-Pd-Mn alloy reveals the local density of
occupied states at the site of each alloy component.

The intensity of fluorescence lines depends on the number of holes in the initial
state, number of electrons that contribute to the dipole transition, and the dipole
matrix elements. For certain transitions, the electrons involved may originate from
bands near EF , and their number is the density of states. Consequently, an x-
ray fluorescence spectrum (XRF) reveals the `-projected density of states at the
atomic site. In an alloy, by tuning the photon energy to the core level energy
CHAPTER 6. INNER SHELLS 156

of the alloy constituents, we can determine the density of electronic states in the
neighborhood of these atoms. This situation is presented for the quasicrystal Al-
Pd-Mn in Fig. 6.19.
Figure 6.19 illustrates schematically the emission of x-rays originating from Al,
Pd, and Mn atoms in an Al-Pd-Mn alloy. A spectrum around h⌫ = 118 eV will
reflect the p-projected density of states of Al atoms, around h⌫ = 531 and 559 eV
the s- and d-projected density of states at Pd site, and around h⌫ = 641 and 652
eV s- and d-projected density of states at Mn site.
X-rays are not e↵ectively absorbed at air. Hence XRF experiments do not
require very high vacuum conditions especially so if the initial hole is created by
x-rays. Therefore, XRF is well suited for elemental analysis of technical alloys and
widely used in industrial applications. We have to bare in mind that the method is
not surface sensitive.
The emission of characteristic x-rays is also used in SEM as described in Sec-
tion 5.2.3 in order to identify the elements in the specimen which is currently imaged.
The method is widely referred to as EDX or EDAX .

6.6.2 Auger Electron Emission


The intensity of electromagnetic radiation depends on the dipole interaction be-
tween the hole state and the state of the electron that fills the hole. This quantity
depends on the energy di↵erence between the two states and hence on the atomic
number Z square. It follows that the probability !R for radiation depends on Z 4 .

Figure 6.20: Dependence of the fluorescence yield (highlighted area)


and Auger yield (white area) on atomic number. Ref. [14].
CHAPTER 6. INNER SHELLS 157

Another process following a core excitation is of purely electronic nature. It is


a redistribution of electrons and holes resulting in emission of an electron. This
radiationless process is called Auger electron emission after Pierre Auger (1899
1993). He observed electrons with constant kinetic energy emitted from Ar+ ions.
This process occurs between electronic shells and depends on the overlap of the shells
and hence governed purely by Coulomb interaction. This interaction is e↵ective at
the atomic site where the hole is located and depends on e2 /r. The probability for
Auger electron emission is !A . The Auger process transfers the holes from states
with higher EB to those with lower energy. In the radiation process the initial and
final states both are a hole state, while in the Auger electron emission the initial
state is a hole state and the final state has two holes.
The initial hole state is bound to relax and hence the sum of the probabilities
!A and !R is unity. !R depends on the nuclear charge Z in the fourth power.14
Figure 6.20 shows the Z dependence of the fluorescence yield. It is obvious that
for lighter elements like C or O, !R is negligibly small which makes detection us-
ing EDAX extremely challenging. For the light elements Auger electron emission
dominates, as indicated by the white area.

Figure 6.21: Estimated lifetime broadening of the ionization energies of


the K, L, and M core levels. J.C. Fuggle, In Electron Spectroscopy:
Theory, Techniques and Applications, ed. C.R. Brundle and A.D. Baker,
Vol. 4, Academic , New York, 1981, p. 85.

14
R.W. Fink et al., Rev. Mod. Phys. 38, 513 (1966).
CHAPTER 6. INNER SHELLS 158

If the system survives in a quantum state for a time ⌧ , in this case the initial
hole state, the energy of the quantum mechanical system in principle cannot be
determined with accuracy better than the spontaneous decay ⇡ h̄/⌧ . This is
fundamental uncertainty relation for energy. In principle, no excited state has
infinite lifetime ⌧ , thus all excited states are subject of the lifetime broadening ,
and the shorter the lifetimes of the states involved in a transition, the broader
the corresponding spectral lines. As seen in Fig. 6.21, K shell broadening of some
heavy elements is so large that the ionization of these levels cannot be detected. In
general, the lifetime of an ionized state is short if there are several electronic states
with smaller EB which can fill the state. These channels are numerous for heavy
elements.
For slow Auger processes, i.e., long life time of the ionized state, we can treat
the ionization and relaxation process separately. If, on the other hand, the hole
state is filled by an electron of the same shell, with the same principal quantum
number n, the probability !A becomes very high owing to the strong overlap of the
wave functions. The relaxation process is then very fast and is called the Coster-
Kronig (CK) transition. Thus, CK transition instantly transfers the hole state to
subshells with lower EB and displays a very broad spectral width. If all three states
involved in the relaxation process originate from the same shell, we have a super-
Coster-Kronig transition with a still higher probability, and we cannot sort out the
ionization and relaxation processes from each other because of the fast time scale.

Figure 6.22: Auger and Coster-Kronig transition in phosphorus. Unknown report.

Figure 6.22 shows the intensity of the L2,3 relaxation as a function of the ex-
citation energy in phosphorus. When the energy is sufficiently high to ionize the
L1 shell around 190 eV, the L1 hole state is transferred to L2,3 shells via a CK
process, as shown the energy diagram on the right-hand side, and the probability
CHAPTER 6. INNER SHELLS 159

!A grows markedly. In order a CK process can take place, the energy di↵erence
between the L1 and L2,3 shells must be larger than the energy necessary to excite
an electron from the valence band to vacuum. Hence, the CK process is not always
energetically possible.
The kinetic energy of Auger electrons is characteristic of the emitting atom and
is well suited for elemental analysis. Yet, the exact energy value of the transition
depends on several relaxation mechanisms. The intraatomic relaxation is a term
associated with the redistribution of charge before the Auger process takes place.
So, it is the response of the system to the creation of the hole state. In cases where
the transitions are so fast that the spectator electrons cannot react, sudden approxi-
mation (= frozen-orbital approximation) is applicable and Koopmans’ theorem may
be used: EB = ✏i . Relaxation of the system is taken negligible.
On the contrary, if the hole state lives long, the spectator electrons can reor-
ganize in energy. This situation is referred to as equivalent-core approximation (=
Z + 1 approximation). This is the adiabatic limit with EB = ✏i Erelax . The
estimation of Erelax is not straightforward, it changes from element to element,
compound to compound. One usually considers besides the intraatomic relaxation
the extraatomic relaxation which describes the screening of the final-state holes.
The Auger transition takes place between electron shells with well-defined bind-
ing energies. As a result the Auger electrons have a definite kinetic energy charac-
teristic to the element in which the hole state is created. Thus, the measurement
of the kinetic energy of Auger electrons serves as a tool for chemical analysis of
surfaces. The method is surface sensitive owing to the limited mean free path of
the emitted electrons, given in Fig. 5.7. The emission intensity, on the other hand,
is a measure for the concentration of the element in a near-surface region.

Figure 6.23: Auger transitions in the quasicrystal Al-Pd-Mn. M. Erbudak,


unpublished.
CHAPTER 6. INNER SHELLS 160

Figure 6.23 shows an Auger spectrum from a clean Al70 Pd20 Mn10 quasicrystal
surface. The sample is excited by 2.4-eV electrons, and the backscattered electrons
are energy analyzed. The signal is electronically di↵erentiated in order to detect
the Auger signal without excessive noise due to the strong background. The main
signal around 70 eV reveals the Al component, while the structure around 360 eV
is due to Pd. Mn signal is less intense and is distributed between 600 700 eV.
Similar spectra are used during the cleaning process of samples that show around
280 eV carbon and 520 eV oxygen signals due to contamination. One continues the
cleaning process until these signals disappear.

Figure 6.24: The film B is vacuum deposited on the substrate A. If the growth
is homogenous, known as FV growth, we expect an exponential dependence on
the thickness d of the Auger signal for both elements. The broken lines point
to a layer-by-layer growth. The Auger signal increases linearly within a film
until the film is completed.

Epitaxial growth can be monitored using Auger electron spectroscopy, as illus-


trated schematically in Fig. 6.24. In homogenous growth of the layer B on the sub-
strate A, the Auger signal from the substrate decreases exponentially, IA / e d/⇤B ,
and the signal from the film material grows as IB / (1 e d/⇤B ) as the film thickness
d increases. ⇤B the mean free path of electrons in the film B.
Because of the high energy and high brightness of the incident electrons, high-
quality Auger electron spectra can be acquired with extremely high signal-to-noise
ratios. Figure 6.25(a) shows a high energy resolution Auger electron spectrum
of clean silver nanoparticles supported on a small MgO crystal; the silver MNN
doublet is clearly resolved. Figure 6.25(b) shows the corresponding oxygen KLL
Auger peak from the same specimen area.15 Surface compositional analysis of in-
dividual nanoparticles is essential for understanding the activity and selectivity of
industrial bimetallic or multi-component catalysts used in a variety of chemical pro-
cesses. Because of the high-surface sensitivity of Auger electrons, it is possible to
determine qualitatively and, in some cases, quantitatively, the surface composition
of nanoparticles consisting of multiple components. High spatial resolution Auger
15
J. Liu et al., Surface Sci. 262, L111 (1992).
CHAPTER 6. INNER SHELLS 161

electron spectra can provide information about the surface enrichment of specific
elements and information about how this enrichment varies with the size of the
nanoparticles.

Figure 6.25: Auger electron spectra of (a) Ag MNN and (b) O


KLL peaks of an Ag/MgO model catalyst. Auger maps of silver
and oxygen are shown in (c) and (d), respectively. Ref. [15].

The primary-electron beam used to induce Auger transitions can be scanned


on the surface to produce spatially resolved chemical information. The resolution
is determined by the size of the focused beam and to a certain extent by the in-
teraction volume of electrons in a near-surface region of the sample. Yet, for some
samples, an image resolution < 1 nm can be achieved in scanning Auger microscopy.
Such Auger maps have received great popularity because the experiment is easily
combined with several others. Silver nanoparticles < 1 nm in diameter and con-
taining as few as 15 silver atoms have been detected. Figures 6.25(c) and (d) show,
respectively, Ag and O maps of an Ag/MgO model catalyst, clearly revealing the
high-spatial resolution of Auger elemental maps. The resolution in images depends
on several sample- and instrument-related e↵ects. The sample-related e↵ects in-
clude: (i) surface topography, (ii) escape depth of the collected Auger electrons,
CHAPTER 6. INNER SHELLS 162

(iii) contribution from backscattered electrons and (iv) localization of the Auger
electron generation processes. The last factor sets the ultimate resolution limit
that will be achievable in images. Since the primary inelastic scattering processes
involve excitation of inner-shell electrons, the generation of Auger electrons is highly
localized. With thin specimens and high-energy incident electrons, the contribu-
tion from backscattered electrons should be negligible; it may, however, degrade the
image resolution and a↵ect the image contrast of bulk samples. The instrument-
related e↵ects include: (i) the intensity distribution of high-energy electron probes,
(ii) the collection efficiency of the emitted Auger electrons and (iii) the instability
of the microscopes. In a modern instrument, the instrument-related factors set the
limits of obtainable resolution to ⇠ 1 nm in Auger peak images of thin specimens.
The minimum detectable mass in high spatial resolution images is < 3 ⇥ 10 21 g.
The Auger transition is a local process, because the driving force is the hole state
localized at the atom. The process leaves behind the atom in a two-hole state. The
interaction of these holes determines the line shape of the emission spectrum. For
quasi-free Bloch states participating in the Auger transition, the Coulomb repulsion
can well be neglected. As a result, the line shape for such a CV V transition, where
C stands for a core state with the initial hole and V for valence band, corresponds
to a self-convolution of the density of states in the valence band. Further, the initial
hole is well screened in metals with Bloch states and lowers the total energy of the
excited state. This is not the case in atoms. Surfaces behave in between.

Figure 6.26: Progressive symmetry-breaking for the final L shell


configurations in KLL transitions. Ref. [16].
CHAPTER 6. INNER SHELLS 163

On the contrary, the line shape of a CCC Auger transition is determined by the
type of interaction of the two-hole state. We di↵erentiate three principal cases:16
a. In the light elements the Russel-Sounders coupling (Coulomb) dominates.
b. For the heavy elements, mainly spin-orbit (j-j) coupling is observed.
c. Intermediate coupling is relevant for elements in between.
The relative energy splitting of each multiplet caused by these mechanisms is
illustrated in Fig. 6.26. Far left, for light elements, energies are completely degen-
erate with nuclear Coulomb potential. There is no electrostatic interaction between
electrons and no spin-orbit interaction, n, l, and s are good quantum numbers.
As elements become heavier, electrostatic interaction has to be considered between
electrons, resulting in states of di↵erent total orbital angular momentum L. Split-
ting into triplets and singlets due to exchange interaction occurs, and the e↵ect of
spin-orbit interaction results in five allowed final states in pure LS-coupling (the
triplet 3 P0,1,2 belonging to the (2s)(2p)5 configuration is degenerate in this limit).
In the intermediate coupling we have nine allowed final states. Far right, for the
heavy elements in the limit of jj-coupling, six allowed final states are presented.

Figure 6.27: Spectrum of the M4,5 N N Auger transition in Kr gas. Ref. [17].

16
K. Siegbahn et al., ESCA: Atomic, Molecular and Solid State Structure by mans of Electron
Spectroscopy, Almqvist and Wiksells, Uppsala, 1967.
CHAPTER 6. INNER SHELLS 164

Krypton atom falls into the intermediate region, as illustrated in Fig. 6.26. The
Auger spectrum recorded from Kr gas is depicted in Fig. 6.27 for the relevant energy
region. In gases the electronic interactions between the excited state and the rest of
electrons are extremely week and the final state lives very long. Hence the lifetime
broadening is negligible and each final-state multiplet is well resolved as seen in the
figure.17

Figure 6.28: Experimental LM M Auger transition spectrum from Cu.


F. Vanini, unpublished.

Some filled shells in common metals may also behave like core levels, such as
the 3d10 electrons of copper located near EF . Fig. 6.28 shows the Auger LM M
transitions in Cu metal. We observe that the Auger spectrum is dominated by
final-state multiplets, this time the final 3d8 state. The L3 M4,5 M4,5 components are
located near 920 eV, while the L2 M4,5 M4,5 multiplets are observed near 940 eV. By
inspection, we can state that among the possible multiplets the 3 F4 is the Hund’s
ground state. This means that the emitted electrons leave the Cu atom in the least
excited state and hence they have the highest kinetic energy. This transition is
denoted with an arrow for the L3 M4,5 M4,5 group in the figure.

17
L.O. Werme et al., Physica Scripta 6, 141 (1971).
Chapter 7

Vibrational Spectroscopies

Characterization of chemisorption has been the major goal of several studies of solid
surfaces because of its obvious relevance to catalysis and corrosion. The study of
kinetics, e.g., thermal desorption and flash desorption, directly yields binding states,
relative saturation densities, and desorption rate parameters. Condensation kinetics
provides an indirect means of characterizing adsorption because sticking coefficient
depends on the state being populated, and the coverage dependence in a state is
controlled by the mechanism of condensation. Surface di↵usion yields information
on the potential experienced by an adsorbate as it moves laterally along a surface.
A surface di↵usion process governs annealing and order-disorder transitions and
may be rate-limiting steps in desorption of a dissociated species.
We seek information related to adsorbate density n, fractional saturation cov-
erage ✓, work function change , activation energy of desorption Ed , sticking
coefficient or probability of condensation s, the number of binding states and their
saturation densities, and symmetry of surface structure. The adsorbate structure,
i.e., the location of adsorbate atoms on a surface, requires the knowledge of the
positions of the substrate atoms. Most of these data are available in experiments
we have reviewed like LEED, XPS, Auger Electron Spectroscopy, etc.
Chemical-specific adsorbate site-symmetry information is contained in vibration
properties of the species at the surface. The geometry around an atom or molecule
adsorbed on a surface defines a certain point-group symmetry. This symmetry
determines the number of observable vibrations. A single atom on the (100) surface
of a cubic crystal can be adsorbed at an ontop, bridge, or fourfold position each
with di↵erent symmetries. The same symmetries apply to an adsorbed dimer. The
number of observable vibrational modes depends on the adsorbed species and the
site symmetry. The observation of a single mode after adsorption of a molecule of
an element is strong evidence for dissociative adsorption.
Further information about the chemical nature and the position of the adsor-
bates is contained in the energy of vibration itself. For a molecular adsorption,
Prof. Dr. Ceyhun Bulutay (bulutay@fen.bilkent.edu.tr), Department of Physics, Bilkent
University, Ankara, has generously contributed the Section 7.3 on Raman Scattering.

165
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 166

a vibration energy close to the vibration energy of the free molecule should be
observed. The energy of vibration also contains some information about the ad-
sorption site: For an adsorbate atom in a position on top of a substrate atom,
the bond is concentrated on a single atom and the force constant e↵ective for the
vibration normal to the surface will be comparatively high. For an atom in the
bridge position or in a fourfold-coordination site the bond is shared between two
and four atoms, respectively. For a vibration normal to the surface the bond forces
however are only partially e↵ective. If one assumes the sum of the bond forces to
be roughly equivalent for the three sites (which to some extent is equivalent to the
assumption of the same desorption energy for the di↵erent sites), one may argue
that the relation !top > !bridge > !f ourf old should hold. This relation is especially
useful if the same adsorbate can be observed in di↵erent sites.
Essential to any investigation of the vibrational properties of an adsorbed mole-
cule is a classification of the localized vibrational modes. An N -atomic molecule
has 3N degrees of freedom, of which 3 are translational, 3 rotational, and 3N 6
vibrational. If this molecule is adsorbed onto a surface and is not mobile, the
translational and rotational degrees of freedom become vibrational degrees of free-
dom. These six “frustrated” translations and rotations are termed external modes,
while others are internal modes. Although there will always be 3N localized vibra-
tional modes associated with an isolated adsorbate molecule, it is not necessarily
the case that 3N bands will appear in a vibrational spectrum. This will depend
on the symmetry that is dictated by the molecular orientation and the surface site,
i.e., on the molecular point group of the adsorbate complex (the surface molecule
group). We examine two simple cases. The adsorption of a hydrogen atom results
in the three translational degrees of freedom being converted into three vibrational
modes. If the H atom is adsorbed on a threefold hollow site only two bands could be
observed in an experimental spectrum: the “frustrated” translation perpendicular
to the surface and the degenerate (due to the symmetry of the triangle) pair of
“frustrated” translations parallel to the surface.1 This is shown in Fig. 7.1. The
degeneracy is lifted if the adatom occupies a bridge site since the bridging atoms
define a preferred direction in the plane, giving rise to a possible three bands in the
experimental spectrum. It should be emphasized that this discussion has been re-
stricted to the number of observable bands in a hypothetical spectrum. The number
of bands actually observed will depend on the appropriate selection rules.
These arguments confirm that vibrational spectroscopy is well qualified for de-
tailed analysis of adsorbate structures. For chemical identification, the character-
istic internal vibrational mode frequencies of gas phase molecules serve as ideal
“fingerprints” that can be found in the signal derived from a solid surface covered
with unknown chemisorbed species.
Vibration of adsorbates can be excited using infrared (IR) radiation.2 Well
1
A.M. Bradshaw, Appl. Surface Sci. 11/12, 712 (1982).
2
C. Backx et al., Surface Sci. 68, 516 (1977).
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 167

Figure 7.1: Vibrational modes of adsorbed hydrogen in a threefold hollow


site and a bridging site on a tungsten surface. Ref. [1].

monochromatized electron beams can alternatively be used to induce adsorbate


vibrations. In both, the sensitivity of the measurement is ⇠ 0.005 ML. Surface-
enhanced Raman spectroscopy3 or inelastic electron tunnelling spectroscopy4 as well
as inelastic He scattering5 are similar techniques to investigate adsorbate vibrations.
In the following, we will limit our discussion to the first three of these.

7.1 Infrared Absorption Spectroscopy


Electromagnetic radiation travels with the speed of light, c = 3 ⇥ 108 m/s. The
photon energy E = h⌫ and c = ⌫. For the energy of visible and x-ray radiation,
used for the investigation of electronic properties of matter, the units are given in
eV such that soft x-rays with a wavelength = 100 Å have an energy of 124 eV
(Eq. 3.1). IR refers to that part of the electromagnetic spectrum between the visible
and microwave regions. In the mid IR region, the wavelength is 2.5 25 µm, which
corresponds to 500 50 meV. Researchers in the field of IR prefer to work with
‘spectroscopic’ wavenumber6 and the range becomes 4000 400 cm 1 . In the IR
spectroscopy, the excitation of adsorbate vibrations are measured in the reflected
beam. Energy is extracted from the radiation field when the frequency of the
light matches the eigenfrequency of the dipole-active molecule on a metal surface.
The interaction between the radiation and the vibrating dipole is produced by
the electric field of the light exerting a force on the charge of the oscillator. The
wavelength of light is long compared to atomic distances and the excitation will
therefore be almost completely in phase for neighboring dipoles. For metals the
dielectric constant is rather high. Therefore the tangential component of the electric
field is practically zero, while a substantial field normal to the surface exists for light
3
M.S. Dresselhaus et al., Group Theory, Springer, New York, 2010.
4
P.K. Hansma, Physics Rep. 30, 145 (1977).
5
A.M. Lahee et al., Surface Sci. 177, 371 (1986).
6
What is called the “wavenumber” is nothing but the inverse of the wavelength. Hence 1 meV
corresponds to 8 cm 1 .
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 168

reflected at grazing incidence and polarized in the plane of incidence. Thus in IR


spectroscopy only the perpendicular component of a surface vibration is excited.
An important feature of IR spectroscopy is its broad-band capability. If a suit-
able source, window material, and detector are available spectra over a wide range
can be measured. The second feature is high resolution; even in most of the com-
mercial devices a resolution better than 1 cm 1 is routinely attained. The third
important feature is the ability to polarize the radiation. Vibrations parallel and
perpendicular to the surface can easily be distinguished, thus making it possible to
determine the symmetry of the normal modes.

Figure 7.2: The electric field vectors for the two light polarizations. Ref. [8].

Figure 7.2 shows the field distribution for the two di↵erent polarizations. While
p-polarized radiation can probe modes both parallel (x direction) and perpendicular
(z direction) to the surface, s-polarized radiation is only sensitive to modes parallel
(y direction) to the surface.
The golden rule expression for the intensity of a vibrational band in IR spec-
troscopy is
~ i i|2 ,
I ⇠ |h'f |~µi · E|' (7.1)
where 'i and 'f are vibrational wave functions of the i -th mode, E ~ is the electric
field acting on the oscillator and µ ~ i is the dipole moment operator associated with
the i -th mode. µ ~ i is often referred to as the dynamic dipole moment, i.e., the
dipole moment associated with the mode when the nuclei are displaced from their
equilibrium position. The excitation normally involved is from the ground state to
the first vibrationally excited state. Using Group Theory a simple selection rule
can be derived which states that the integral in Eq. 7.1 is only non-zero when the
i-th mode belongs to the same irreducible representation of the molecular point
group as one or more of the Cartesian components of the dynamic dipole. If this
condition is fulfilled one speaks of a dipole-active mode.
It is clear from Eq. 7.1 that maximum absorption occurs when E is parallel
to the dynamic dipole. Important for the IR experiment is thus the behavior of
~ 2 upon reflection at a typical metal surface. Additional selection rules can be
|E|
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 169

Figure 7.3: The image dipole model. From AZ.

“induced” if specific adsorption conditions fix the direction of E.~ In the same way
that the response of the metal valence electrons to the incident electromagnetic wave
screens out the electric field components parallel to the surface, the dynamic dipole
of the vibrational mode itself does not remain una↵ected. This phenomenon is best
discussed in terms of a “fictitious” image dipole. Figure 7.3 shows two instantaneous
dipoles, lying perpendicular and parallel to a metal surface. The response of the
metal valence electrons to the dipole fields may be represented in the terms of an
induced image dipole, which in turn interacts with the real dipole. Since for the
dipole oriented normal to the surface the image dipole is in the same sense, this
interaction corresponds to a reinforcement. For the parallel dipole the interaction
e↵ectively gives rise to an electric quadrupole, for which excitation probability in
IR spectroscopy is usually a factor 1000 lower. Hence, the combination of the two
e↵ects - the screening of both the electric vector and the oscillating dipole - means
that the chances of observing a dipole-active mode where the change in dipole
moment is exclusively parallel to the surface are practically zero.
Hydrogen adsorption on a Si(100) surface is a classical example for the IR spec-
troscopy, where its strength is evident.7 Hydrogen molecules dissociate upon ad-
sorption on Si(100)(2 ⇥ 1) and form the monohydride phase. There are two local
vibration modes separated by 12 cm 1 , illustrated in Fig. 7.4. The inset shows the
directions of hydrogen-atom motion responsible for these modes. The s-polarized
light only excites the antisymmetric stretch vibration (2087 cm 1 ), while the p-
polarized light excites both the strong symmetric stretch vibration (2099 cm 1 )
and the antisymmetric vibration. This example shows that the IR spectroscopy has
an unprecedented resolution to distinguish between these modes and exploits the
polarizability of the excitation in distinguishing between the modes.
In any absorption spectroscopy we measure how e↵ective a sample absorbs light
at each wavelength. The classical and most straightforward way to do this is the
technique of dispersive spectroscopy where a monochromatic light beam is directed
at a sample and measure the intensity of the absorbed light. We repeat this proce-
dure for each di↵erent wavelength.
The Fourier-transform spectroscopy is a less intuitive way to obtain the same
information. Rather than shining a monochromatic beam of light at the sample, this
technique shines a beam containing many frequencies of light at once, and measures
how much of that beam is absorbed by the sample. Next, the beam is modified
7
Y.J. Chabal, Surface Sci. 168, 594 (1986).
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 170

Figure 7.4: IR spectrum on a Si(100)(2 ⇥ 1)-H surface. The solid line


corresponds to p-polarized spectrum and the dashed line to s-polarized
spectrum. The inset shows the symmetric and antisymmetric vibrations
of the adsorbed hydrogen atoms. Ref. [8].

to contain a di↵erent combination of frequencies, giving a second data point. This


process is repeated many times. Afterwards, a computer takes all these data and
works backwards to infer what the absorption is at each wavelength. The beam is
generated by starting with a broadband light source, containing the full spectrum of
wavelengths to be measured. The light shines into a Michelson interferometer. As
this mirror moves, each wavelength of light in the beam is periodically blocked and
transmitted by the interferometer due to the interference. Di↵erent wavelengths
are modulated at di↵erent rates, so that at each moment, the beam coming out of
the interferometer has a di↵erent spectrum. Subsequently, a computer processing
is required to turn the raw data, i.e., light absorption for each mirror position,
into the desired result, i.e., light absorption for each wavelength. The processing
required turns out to be a common algorithm of the Fourier transform, hence the
name, Fourier transform IR spectroscopy or FTIR.8
Since an FTIR spectrometer simultaneously collects spectral data in a wide
spectral range, which confers a significant advantage over a dispersive spectrometer,
which measures intensity over a narrow range of wavelengths at a time, FTIR has
made dispersive IR spectrometers all but obsolete, opening up new applications of
IR spectroscopy.
8
Y.J. Chabal, Surf. Sci. Rep. 8, 211 (1988).
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 171

7.2 Inelastic Electron Scattering


When an electron approaches the surface the electric field of the electron exerts a
force on dipole-active oscillators located at the surface. Just as in IR spectroscopy
and for the same reason for metals the field is practically normal to the surface.
Therefore, the same selection rule with respect to the orientation of the dipole
oscillator applies. As a consequence of the long-range nature of the Coulomb field
the most significant contributions to the total interaction arise during the time
where the electron is still many lattice constants away from the surface. The field is
then nearly homogeneous on the atomic scale and therefore mostly long-wavelength
surface waves are excited. The lateral extension of the electric field, however, being
a function of the distance of the electron shrinks up as the electron approaches the
surface. This is the reason, why unlike to IR spectroscopy, a continuous distribution
of surface wave vectors is excited with electrons.
In IR spectroscopy the field is periodic in time. Therefore for harmonic oscilla-
tors only the fundamental frequency is excited. This is di↵erent with electrons. The
total interaction time of the electron is of the order of an oscillator period. From
the standpoint of the oscillator the external force therefore contains all frequencies.
As a result, multiple excitations may occur.
The operator which causes excitations in inelastic electron scattering (IES) is
iq·r
e where r is the position of the scattered electron and q the transferred mo-
mentum, as we had in Eq. 6.9. In the case of a localized initial state, this operator
may be expanded and the linear term dominates for q sufficiently small and causes
dipole transitions; the quadratic term, which gives rise to monopole, quadruple,
and higher-order cross terms in the matrix element, becomes important for larger
q. Thus, when q is small, we observe the same transitions as are seen in optical
absorption studies; “forbidden” transitions may be observed at larger q.
In the case of electron reflection from an ideal conducting metal surface, the
conduction electrons respond to the incident radiation producing a dipole image
charge potential normal to the vacuum-metal interface. The time response of this
long-range Coulombic interaction between the incoming charged particle and the
surface electric field extends out into the vacuum. This crude representation of
a complex scattering process serves to illustrate the point that the electron beam
acts as a source of wide-band radiation. The optical absorption properties of an
adsorbate will determine at which frequencies ⌫ “photons” are adsorbed from this
radiation source, corresponding to an energy loss h⌫ of an electron scattered in the
specular direction, i.e., relatively small net momentum transfer q associated with
the loss.
In addition to this optical dipole selection rule, the electric field normal to the
surface, associated with the induced image charge, provides a second selection rule,
which is important in the interpretation of energy-loss spectra, as it is in the IR
spectra. In this case of atomic and molecular adsorbate vibrational excitations at IR
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 172

frequencies, only vibrations which give dipole changes perpendicular to the surface
will absorb radiation strongly. This “surface-normal dipole selection rule” is equally
true whether the incident radiation is an electron or an IR photon. It can therefore
be readily seen in Fig. 7.3 that, for example, a bond-stretching vibration, which
would give rise to a dipole change parallel to the surface, will produce an equal and
opposite change in the induced “image” dipole in the metal substrate; the net dipole
change is e↵ectively zero. On the other hand, the oscillating dipole perpendicular
to the surface will be reinforced by the oscillating image dipole. Another important
point is that the surface carries a permanent normal dipole field (see, Section 2.4,
the origin of the work function) that e↵ectively polarizes otherwise IR-inactive
species when adsorbed. For this reason, the normal vibration modes of chemisorbed
hydrogen produce inelastic scattering of the incident electrons.

Figure 7.5: Electron energy-loss spectra of hydrogen chemisorbed on


W(100) at an impact energy of E = 9.65 eV: (a) specular beam direction;
(b) 17 o↵ the specular direction towards the surface. The fundamental
vibrational modes in the inset correspond to bridge-site bonding. Ref. [9].

In the IES experiment low-energy electrons of typically 2 5 eV energy are


monochromatized (ideally E < 10 meV) with an electrostatic energy analyzer
and reflected from the surface. The low resolution of IES makes it inferior to IR
spectroscopy. A second electrostatic energy analyzer then discriminates the scat-
tered electrons according to their energy. Those electrons which have excited vibra-
tional modes in the adsorbate will appear as discrete loss peaks in the spectrum.
Two excitation mechanisms appear to be primarily involved: dipole scattering and
impact scattering. Dipole scattering is a long-range interaction between the inci-
dent or departing electron via its associated electric field and the oscillating dipole.
Because of the response of the metal electrons to the approaching electron, which
can be thought of in terms of the production of an image electron in the substrate,
the lines of electric field strength terminate perpendicularly at the surface. Further-
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 173

more, at distances far enough from the surface electrons cannot distinguish between
the dipole and its image and so the simple dipole screening model of Fig. 7.3 is main-
tained. The discussion of the IR excitation mechanism based on Eq. 7.1 can thus
in a first approximation be taken over and applied to dipole scattering in IES. In
particular, we only expect to observe those modes which have a component of the
dynamic dipole normal to the surface.
Figure 7.5(a) shows energy losses from H adsorbed on W(100) in the specular
beam. Only the dipole-active vibration mode is observed at 130 meV the same
way as in the IR spectra. The fact that there is only one single loss peak is safely
interpreted as dissociative adsorption of hydrogen and the loss energy corresponds
to the symmetric stretch vibration normal to the surface.
IES possesses one significant advantage towards IR spectroscopy: short-range
impact scattering of the incoming electron from the local adsorbate potential can
excite non-dipole active modes. Here, we are dealing with a short-range interaction
between the electron and the atomic potentials of the adsorbate when the electron
is at a distance from the surface comparable to the molecular dimensions and where
the picture of classical image screening breaks down.9 Figure 7.5(b) displays energy-
loss spectra recorded in an o↵-specular direction from the same W(100)-H system.
The 260 eV loss peak is the overtone (⌫ = 0 ! 2) of the fundamental dipole-
active vibration, while 80 and 160 eV losses are non-dipole excitations. A simple
spring model predicts that the 160 meV mode is the asymmetric stretch and the
80 meV loss is then associated with an out-of-plane bending. The main feature of
Fig. 7.5(b) is that all of the fundamental modes of the surface-molecule complex
can be excited and observed when the measurement is made o↵ the specular beam
direction. Whereas in dipole scattering the loss electrons are to be found in a
narrow cone around the specular beam, in impact scattering they form a more
isotropic distribution. The loss peak intensities will also be considerably weaker
due to cross sections which are several orders of magnitude lower. Also important
is the fact that impact scattering can occur not only from perpendicular modes
but also from parallel modes and from modes that are not dipole-active. Thus by
measuring scattering intensities as a function of angle it is possible to distinguish
the perpendicular modes from the rest and to perform a structural analysis.
IES experiments require ultrahigh-vacuum conditions, while IR spectroscopy
can be performed on real surfaces.

9
W. Ho et al., Phys. Rev. Lett. 40, 1463 (1978).
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 174

7.3 Raman Scattering


When a monochromatic electromagnetic wave is incident on a surface, we get the
reflected and refracted beams as governed by Snell’s law. Even though this picture
is quantitatively quite accurate, qualitatively it misses some important phenomena.
Real materials also contain static and dynamic scatterers. The former cause an
elastic scattering (i.e., at the same wavelength as the incoming wave) in all direc-
tions, which is about 1/10,000 of the incident intensity, and is known as Rayleigh
scattering. Moreover, about 1/1000 of the scattered wave gets inelastically scat-
tered (i.e., at a shifted wavelength with respect to incident wave) that arises from
dynamic processes which is known as the Raman scattering. This Raman signal,
though usually a tiny fraction (⇠ 1/107 ) of the incident intensity, plays a vital role
in identifying the atomic constituents of the sample through their vibrational and
rotational fingerprints.
Hence, both Raman and IR spectroscopy are tools primarily for exploring vi-
brational properties of molecules, surfaces, and solids (crystalline or amorphous).
They usually act as complementary techniques, as some vibrational modes are only
IR active, and some are only Raman active. There are also cases where a particular
mode can be active or silent for either one. Among the other di↵erences between the
two techniques, Raman spectroscopy has a simpler spectrum than IR as often only
the fundamental (lowest-order) processes are visible. Furthermore, in the Raman
spectroscopy a monochromatic (usually a laser) source is used, whereas IR utilizes
broadband sources.

7.3.1 A Classical Consideration for the Raman Process


First, we shall see that some essential features of the Raman scattering can be
captured by a basic classical treatment.10,11,12

7.3.1.1 An Isolated Molecule


We start with a single molecule that may in general have a permanent dipole mo-
ment µ~ , and an induced dipole moment through a polarizability tensor ↵ij , where i, j
are the Cartesian indices, so that under an incident monochromatic electromagnetic
~ =E
field E ~ 0 cos(!t), the resultant dipole moment becomes

pi = µi + ↵ij Ej ,

with the convention that repeated Cartesian indices are implicitly summed over.
Note that both µ
~ and ↵ij are functions of the coordinates of the nuclei and electrons.
10
W. Demtröder, Atoms, Molecules and Photons, Springer, New York, 2006.
11
P.Y. Yu and M. Cardona, Fundamentals of Semiconductors, 4th Ed., Springer, New York,
2010.
12
D.A. Long, The Raman E↵ect, John Wiley, New York, 2002.
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 175

We shall assume that electronic cloud can instantly respond to any change in the
nuclear vibrations. Therefore, essentially these quantities are controlled by the
nuclear displacements, for which we shall use their so-called normal coordinates as
denoted by Q ~ k , where k labels any one of the vibrational degrees of freedom. As
mentioned earlier, for an N -atom molecule, these are in total 3N 6. If we further
assume that the incident light is o↵-resonant with both electronic and vibrational
transitions, then such small-amplitude oscillations for the nuclear normal modes
can be governed by the first two terms of the Taylor expansion around equilibrium
coordinates Q~ k = 0 yielding

3N
X6 @µi
µi ' µi (0) + Qkj , (7.2)
k=1 @Qkj 0
3N
X6 @↵ij ~k .
↵ij ' ↵ij (0) + Q (7.3)
k=1
~k
@Q 0

Each normal mode of vibration k will be oscillating at its eigenfrequency !k , there-


~ k (t) = Q
fore we can write Q ~ k0 cos(!k t). With this form the dipole moment of the
molecule becomes
3N
X6 @µi
pi = µi (0) + Qk0j cos(!k t) + ↵ij (0)E0j cos(!t)
k=1 @Qkj 0
| {z }
| {z } Rayleigh scat.
IR spectrum
8 9
>
> >
>
1 3N
X6 @↵ij < =
+ E0j ~ k0 cos[(! + !k )t] + cos[(!
Q !k )t] . (7.4)
2 ~
k=1 @ Qk 0
>
:|
> {z } | {z }>
>
;
anti-Stokes Stokes

According to this expression, there is a component at the same frequency as the


incident field, called the Rayleigh scattering; additionally we have a red- and blue-
shifted frequency parts which constitute the Raman scattering, where the former
(latter) is called the Stokes13 (anti-Stokes) scattered light. This derivation suggests
us that Raman scattering can be seen as a frequency mixing process just like the
amplitude modulation concept used in electronic communication: incident field,
acting as the carrier signal is modulated by the nuclear vibrations, here playing
the role of the information signal. Fig. 7.6 illustrates the basic spectrum and the
transitions corresponding to Stokes and anti-Stokes lines.

7.3.1.2 Selection Rules


As a result of the symmetries of the medium and of the vibrational modes involved
in the scattering, some requirements are imposed. These are generally termed as
13
This terminology historically originates from a somewhat similar observation by Stokes on
fluorescence where the frequency of the fluorescent radiation is always less than that of the incident.
Ref. [12].
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 176

Figure 7.6: The spectrum showing Rayleigh and Raman signals, together
with the associated transition schemes. Ref. [10].

selection rules, which determine whether a specific perturbation V yields a non-


zero transition from an initial state |vi to a final state |v 0 i as quantified by the
matrix element hv 0 |V|vi. The study of such selection rules falls in the realm of
so-called Group Theory, from which we shall state a few useful remarks without
any derivation.14 For the IR and Raman spectroscopy, if we refer to Eq. (7.4), the
operators that govern these transitions are the dipole moment µi and polarizability
↵ij operators, respectively. The former has components which transform as x, y
and z, whereas the latter transforms as the binary products of x2 , y 2 , z 2 , xy, xz, yz,
or equally their linear combinations such as x2 y 2 . According to Group Theory,
if the direct product [v] ⇥ [µ] ⇥ [v] contains the totally symmetric irreducible
representation of the point group, the transition v ! v 0 is IR active. Likewise,
if the direct product [v] ⇥ [↵] ⇥ [v] contains the totally symmetric irreducible
representation of the point group, the transition v ! v 0 becomes Raman active.

7.3.1.3 An Example: CO2 Molecule


In some cases, IR and Raman activity of certain vibrational modes can be extracted
simply by observation, without the use of rigorous Group Theory machinery. One
such example is a linear symmetric molecule ABA, such as the CO2 molecule. It
has four modes of vibration: a symmetric stretching mode Q1 , an antisymmetric
stretching mode Q2 , and two bending modes Q3a and Q3b which form a degenerate
pair and have the same frequency of vibration. Three of these vibrations are shown
in Fig. 7.7. We observe that symmetric stretching mode (on the left panel) has non-
zero polarizability derivative at the equilibrium position (@↵/@Q 6= 0) therefore this
vibration will give rise to an induced polarizability contribution under an incident
field, hence it is Raman-active. However, as the A-B and B-A dipoles are of equal
strength but pointing in opposite directions, this mode of vibration produces no net
14
D.C. Harris and M.D. Bertolucci, Symmetry and Spectroscopy, Dover, Mineola, 1989.
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 177

dipole moment derivative at the center (@µ/@Q 6= 0), therefore it will be IR-silent.
The situation is just the opposite for the center and right panels, as they are both
IR-active but Raman-inactive.

Figure 7.7: Polarizability and dipole moment variations in the neighbor-


hood of the equilibrium position for a linear ABA molecule. Ref. [10].

7.3.1.4 Extended Systems


Next, we move from a single N -atom molecule to an extended system (such as a
solid) having n molecules per unit volume. The relation between the molecular
quantities µ ~ , ↵ij , and those of the extended system, the permanent i and induced
ij electric susceptibilities are i = nµi /✏0 , and ij = n↵ij ✏0 , where ✏0 is the
permittivity of free-space. Similarly the polarization field of the medium is related to
molecular dipole moment via P~ = n~p. Hence, for an incoming electromagnetic wave,
this time including its wave vector dependence ~k as well, E(~ ~ r, t) = E~ 0 cos(~k ·~r !t),
and expanding into Taylor series under nuclear vibrations with a form Q ~ k (~r, t) =
~ k0 (~q , !k ) cos(~q ·~r !k t) yields the following terms, where we only show the Raman
Q
contributions
8
>
>
1 @ ij <
Pi = E0j ~ k0 (~q , !k ) cos[(~k + ~q ) · ~r
Q (! + !k )t]
2 ~k
@Q 0 >>
:| {z }
anti-Stokes
9
>
>
=
+ cos[(~k ~q ) · ~r (! !k )t]> . (7.5)
| {z }>
;
Stokes
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 178

For an isolated molecule the Raman scattering occurs in all directions with no
angular preference. However, in an extended system we see that the Stokes-shifted
wave has ~kS = ~k ~q , while the anti-Stokes is ~kAS = ~k + ~q . This means that in
this case, the observation direction for the scattered wave inherently selects the
participating phonon mode through the above momentum-conservation relations.
Furthermore, note that for the visible light which is commonly used in Raman
spectroscopy, its spectroscopic wavenumber is of the order of 105 cm 1 which makes
up only a tiny proportion with respect to the extend of a typical Brillouin zone
⇠ 107 cm 1 . Therefore, in the case of crystals, because of the above momentum
conservation requirement only the zone center phonons q ! 0 participate in a
Raman process.
As the nuclear vibrations attain larger amplitudes one may have to retain higher-
order terms in the Taylor series expansion. These bring cross terms which will give
rise to Raman shifts as ±!n ± !m for the second-order. When these modes are
identical, the resultant two-phonon peak is called an overtone.

7.3.1.5 Raman Tensor


The intensity for the Raman scattered wave polarized along the unit vector direction
ês in response to an incident polarization along êi is proportional to
2
@ ij ~ k0 (q = 0, !k ) · êi
IS / ês · Q . (7.6)
~k
@Q 0
~ k , where i, j run
Therefore the scattering is governed by a third-rank tensor @ ij /@ Q
over Cartesian directions and k labels any of the optical phonon modes at the zone
center. By introducing a unit vector along every available normal mode phonon
~ Q|,
displacements Q̂ = Q/| ~ we can essentially introduce a second-rank (with the
same components as ij ) for each specific phonon polarization, called the Raman
$
tensor as15 R= (@ /@ Q)~ 0 Q̂(!k ) , so that IS becomes
!S4 $ 2
IS (!S ) /
ê ·
s R ·ê i .
c4
Note that, we explicitly show the k 4 = ! 4 /c4 wavelength dependence which is
ubiquitous in any form of sub-wavelength scattering like the Rayleigh scattering as
popularized by the explanation of the blueness of the sky. If we neglect the small
frequency di↵erence between the incident and scattered waves, the Raman tensor
can be accepted as symmetric with respect to its Cartesian indices, just like the
electric susceptibility ij .

7.3.2 Quantum Mechanical Features of the Raman Process


Next, we discuss features which necessitate a quantum mechanical treatment.11
First, we notice from Eq. (7.6) that the scattered intensity is proportional to the
15
Here, we suppress the phonon polarization subscript on the Raman tensor.
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 179

vibration amplitude Q ~ squared. Hence, according to the classical treatment there


would be no Stokes scattering if no atomic vibration is present. However, once we
quantize the vibrational modes into phonons, in Stokes scattering where a phonon is
excited in the medium by the incident radiation, the intensity becomes proportional
to Nq + 1, while the anti-Stokes will be proportional to Nq . Here, Nq is the phonon
occupancy, which is given in the case of equilibrium by the Bose-Einstein distri-
bution. For low temperatures Nq ⌧ 1, therefore the ‘+1’ contribution in Stokes
scattering intensity dominates. As one can recall, it arises from zero-point oscil-
lations, a hallmark of quantum mechanics, and gives rise to spontaneous phonon
emission even at zero temperature having Nq ! 0.
Based on these facts, we can write the intensity ratio, anti-Stokes over the Stokes
as ✓ ◆
! + !k 4 Nq
,
! !k Nq + 1
where the first term is again from the k 4 dependence for each frequency, while the
second term contains the temperature dependence which simplifies to e h̄!q /kB T . A
fringe benefit of this is that the intensity ratio of the two lines can be used to extract
the lattice temperature. In the light of this discussion, Stokes lines will be more
intense as they originate from the v = 0 zero-phonon level, whereas anti-Stokes
lines originate from the v = 1 one-phonon level with much less population.
Another important feature is that the Raman lines have non-zero widths which
further increase with temperature. Primarily it is due to the fact that the optical
phonons taking part in the Raman process themselves are not of infinite lifetime
because of the inherent anharmonicity of lattice vibrations causing phonon-phonon
interactions. Therefore, they decay into two longitudinal acoustic phonons with
large and opposite momenta. This finite phonon lifetime broadens the Raman
resonances into a Lorentzian or a Gaussian line shape. The thermal variation of
the width of the Raman line is given by16
✓ ◆
2
(!k , T ) = (!k , 0) 1 + . (7.7)
eh̄!k /2kB T 1
A further important deficiency of the classical picture is that it does not reflect
the microscopic mechanism of how Raman scattering actually takes place. As a
matter of fact, the direct excitation of phonons via photons is extremely weak.
Therefore, this scattering proceeds quite di↵erently in at least three steps.
1. An incident photon with an energy h̄!i interacts with the charges within the
sample through Hrad X , exciting the medium into an intermediate state |ni
by creating a so-called exciton which is the bound state of an electron and
hole pair.

2. This exciton being composed of an electron and a hole interacts strongly with
the environment (lattice) via the electron-phonon (or hole-phonon) interaction
16
H. Kuzmany, Solid-State Spectroscopy, Springer, New York, 1998.
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 180

(He ion ) and gets scattered into another state by emitting a phonon (consid-
ering the Stokes case) of energy h̄!k . This intermediate excitonic state will
be denoted as |n0 i.

3. The exciton in state |n0 i spontaneously recombines radiatively via Hrad X


with the emission of a so-called scattered photon of energy h̄!S .

So, this reveals that the electrons mediate the Raman scattering of phonons
although they remain unchanged after the process. Since the transitions involving
the electrons are virtual they do not have to conserve energy, although they still have
to conserve (crystal) momentum. The corresponding Feynman diagram describing
this process is shown in Fig. 7.8. Note that there are higher-order virtual processes
other than the one above which also contribute Raman scattering. However, if we
limit ourselves to only this lowest-order mechanism, then the Raman scattering rate
can be calculated using Fermi Golden Rule as
2
2⇡ X hi|Hrad X (!S )|n0 ihn0 |He ion (!k )|nihn|Hrad X (!i )|ii
WR (!S ) =
h̄ n,n0 [h̄!i (En Ei )] [h̄!i h̄!k (En0 Ei )]
⇥ (h̄!i h̄!k h̄!S ) . (7.8)

Figure 7.8: The Feynman diagram corresponding to the basic Raman


process. The left and right wiggly lines represent incident and scattered
photon propagators, solid and dashed lines correspond to electron and
hole propagators, and the spiral line is that of the phonon.

7.3.3 Variants of the Raman Process


The basic Raman process discussed above in practice comes with quite a few vari-
ations, with each one having a di↵erent utility and/or novelty. Some of these are
very briefly mentioned below.12,10,11

• Brillouin Scattering – Inelastic scattering of light by acoustic waves was


first proposed by Brillouin. As a result, this kind of light scattering spec-
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 181

troscopy is known as Brillouin scattering. There is very little di↵erence be-


tween Raman scattering and Brillouin scattering. In solids the main di↵er-
ence between them arises from the di↵erence in dispersion between optical
and acoustic phonons.

• Hyper-Raman – The intensity of Raman scattering is directly proportional


to the irradiance of the incident radiation and so such scattering can be de-
scribed as a linear process. When the intensity of the incident light wave
becomes sufficiently large, the induced oscillation of the electron cloud sur-
passes the linear regime. This implies that the induced dipole moment p of
the molecules are no longer proportional to the electric field E, but the E 2
and E 3 terms in the Taylor expansion of p(E) need to be retained giving rise
to hyper-polarizability and second-hyper-polarizabilities, respectively. Such
nonlinear optical e↵ects are named as hyper-Raman processes having quite
di↵erent selection rules compared to linear one. The overtones of the main
Raman signal as mentioned previously, are nothing but the manifestations of
the hyper-Raman processes.

• Electronic Raman – If the frequency di↵erence !i !k corresponds to an


electronic transition of the system, we speak of electronic Raman scattering,
which gives complementary information to electronic-absorption spectroscopy.
This is because the initial and final states must have the same parity, and
therefore a direct dipole-allowed electronic transition |ii ! |f i is not possible.

• Resonant Raman – The enhancement of the Raman cross section near an


electronic resonance is also known as resonant Raman scattering. Obviously,
resonant Brillouin scattering is defined similarly. The enhancement in the
Raman cross section at resonance is only two orders of magnitude relative
to the nonresonant background. On the other hand, both free excitons and
bound excitons have been shown to enhance the Raman cross section by sev-
eral orders of magnitude because of their small damping constants (with a
typical width of 3 meV) at low temperatures. Such strong resonance e↵ects
have made possible the observation of new phenomena, such as wavevector-
dependent electron–LO phonon interaction, electric-dipole forbidden transi-
tions, higher-order Raman scattering involving more than three phonons, and
the determination of exciton dispersions.

• Fourier-transform Raman – The combination of Raman spectroscopy with


Fourier-transform spectroscopy allows the simultaneous detection of larger
spectral ranges in the Raman spectra.

• Surface-Enhanced Raman Scattering – Raman scattering cross-section


is typically 14-15 orders of magnitude smaller over that of the fluorescence
of efficient dye molecules. The intensity of Raman scattered light may be
CHAPTER 7. VIBRATIONAL SPECTROSCOPIES 182

enhanced by several orders of magnitude if the molecules are adsorbed on


a surface. A number of mechanisms contribute to this enhancement. Since
the amplitude of the scattered radiation is proportional to the induced dipole
moment, pi = ↵ij Ej , the increase of the polarizability ↵ by the interaction of
the molecule with the surface is one of the causes for the enhancement. In the
case of metal surfaces due to the presence of plasmonic e↵ects, the electric field
E at the surface may also be much larger than that of the incident radiation,
which also leads to an increase of the induced dipole moment. Both e↵ects
depend on the orientation of the molecule relative to the surface normal,
on its distance from the surface, and on the morphology, in particular the
roughness of the surface. Small metal clusters on the surface increase the
intensity of the molecular Raman lines. The frequency of the exciting light
also has a large influence on the enhancement factor. In the case of metal
surfaces it becomes maximum if it is close to the plasma frequency of the
metal. Giant enhancement factors as high as 1014 have been reported on
single-molecule studies.17 Because of these dependencies, surface-enhanced
Raman spectroscopy has been successfully applied for surface analysis and
also for tracing small concentrations of adsorbed molecules.

• Coherent anti-Stokes Raman Scattering – Ordinary Raman scattering


is an incoherent spontaneous process and as a result the intensity of scatter-
ing from a material system of N non-interacting molecules is simply N times
that from one molecule. There are also non-spontaneous Raman processes,
the three important ones are the stimulated Stokes Raman scattering, stim-
ulated anti-Stokes Raman scattering and coherent anti-Stokes Raman scat-
tering (CARS). The most popular is the last one where two incident laser
beams are used with an energy di↵erence deliberately chosen to match a
Raman-active vibrational mode of the sample. This produces highly direc-
tional beams of scattered radiation with small divergences. In particular, the
intensity of the anti-Stokes signal is by far larger than in spontaneous Ra-
man spectroscopy. The scattered intensity is proportional (a) to the square
of the number of scattering molecules and (b) to the square of the irradiance
of the incident radiations. With pulsed lasers time-dependent processes in
molecules and their influence on the change of level populations can be stud-
ied by CARS; two examples being photosynthesis and the visual processes in
our eyes.

17
K. Kneipp et al., Phys. Rev. Lett. 78, 1667 (1997).
Chapter 8

Surface Reactions

In most of the investigations of solid surfaces the emphasis is placed on the charac-
terization of clean and adsorbate covered surfaces because of the evident relevance
to catalysis and corrosion. A reaction on the surface, as we have in heterogenous
catalysis, proceeds at least in three steps. Each step is controlled by the reaction
rate determined by the temperature and gas pressure:
1. Adsorption or coadsorption of two or several gases at the surface.
2. Reaction at the surface.
3. Desorption of reaction products from the surface.
We will deal with adsorption, reaction, and desorption phenomena without going
into much detail. Several spectroscopies we have studied so far had obvious con-
nections to these subjects so that we can say that most of the time chemistry is one
of the driving forces of detailed investigations in surface science.

8.1 Adsorption
Adsorption takes place when an attractive interaction between a particle and a
surface is sufficiently strong to overcome the disordering e↵ect of thermal motion.
When the attractive interaction is essentially the result of van-der-Waals forces
then physisorption takes place. Physisorptive bonds are characterized by binding
energies below approximately 1 eV. Chemisorption occurs when the overlap between
the molecular orbitals of the adsorbed particle and the surface atoms permit the
formation of chemical bonds, which are characterized by binding energy typically
exceeding 1 eV. The chemisorption is always an activated process, i.e., the formation
of a chemisorptive bond requires that an activation barrier is overcome. A common
feature of molecular chemisorption is the weakening of intramolecular bonds that
often lead to the dissociation of the adsorbed molecule. An important example for
activated, dissociative chemisorption is the adsorption of oxygen molecules on most
metal surfaces at room temperature.
A descriptive view of chemisorption considers an adsorbate atom with a filled
level of energy Ea that approaches a metal surface. As the atom nears the metal Ea

183
CHAPTER 8. SURFACE REACTIONS 184

will broaden by interaction with the metal. The closer Ea to EF the larger is the
broadening. In the extreme case the overlap may lead to bonding and antibonding
states. This interaction lowers the energy of the metal states below EF thus forming
a strong bond. As found in all bonding the essential feature is that two electrons
can simultaneously be on the adsorbate. Then the intraatomic Coulomb repulsion
U of these electrons presents a crucial aspect of adsorption. If the broadening of
the levels or the separation of bonding and antibonding orbitals exceeds U the
chemical bond is stable. Fortunately, the value of U is considerably reduced near
a metal surface by screening e↵ects which can be understood in terms of image
interaction. Thus, the level Ea is pushed up by the image force (= e2 /4z) and the
resulting ion interacts attractively with the metal through its image charge.
The simplest approach to quantitative analysis of adsorption is to follow that
what is already postulated by Langmuir. He made several assumptions:
1. A surface contains enumerable sites on which an adsorbate can be bound. The
fraction ✓ of the occupied sites also called surface coverage, is 0  ✓  1. When all
adsorbed sites are occupied, the substrate surface is saturated.
2. Adsorption can only take place at sites that are not yet occupied.

Figure 8.1: Adsorption is possible at a previously not occupied site.

Since the fraction of the sites that are not occupied is given by (1 ✓), this number
gives the fraction of species that are adsorbed at the surface upon arrival and is
called the sticking probability, S(✓).

Figure 8.2: In the simplest case, the sticking coefficient S(✓) is linear
with surface coverage, while real surfaces follow the power law.
CHAPTER 8. SURFACE REACTIONS 185

3. There is no interaction between the adsorbates. This means that the adsorbates
“wet” the substrate and there is no island formation. like in an idealized FV-type
growth (see Figs. 1.4 and 1.5). This assumption requires that the adsorbed particles
are immobile.
4. Adsorption proceeds in equilibrium. With the number of adsorption sites N ,
dN dN
( )ads = ( )desorb (8.1)
dt dt
Using the constants k 0 , k 00 , and b = k 0 /k 00 and the gas pressure p, we can write

dN dN
( )ads = k 0 p(1 ✓) and ( )desorb = k 00 ✓. (8.2)
dt dt
This leads to adsorption isotherms describing the surface coverage as a function of
gas pressure over the sample:
bP
✓= (8.3)
1 + bP
Actually, b = f (T ).

Figure 8.3: Adsorption isotherm of Langmuir.

We recognize two limiting cases:


a. Low-pressure regime leads to weak adsorption, where ✓ = bp
b. High-pressure regime leads to saturation with ✓ = 1.
This case is a first-order rate equation. There is also a second-order process,
which corresponds to dissociative adsorption and recombinative desorption of di-
atomic molecules, relevant for catalysis. Note that the dissociation of molecules
can often be suppressed by lowering the temperature of the surface. An example
is the adsorption of O2 on Ag(110), which is dissociative at room temperature, but
non-dissociative at 150 K. For the second-order process, the coverage is given by
p
bP
✓= p (8.4)
1 + bP
CHAPTER 8. SURFACE REACTIONS 186

A modern approach to investigate the adsorbate structure, which simple means


the location of the adsorbate atoms on the surface, requires the knowledge of the
positions of the substrate atoms. We have seen several real-space and reciprocal-
space techniques that deliver information on the clean solid surfaces. Di↵erent
crystallographic orientations of bcc metals, like W, Mo, Ta, Nb, are frequently
studied in chemisorption because of ease of their preparation. Most of the clean
surfaces give only (1 ⇥ 1) LEED patterns, implying one deals with bulk-terminated
surfaces. The fcc metals include the best catalysts, like Pt, Ni, Pd, and Rh and
the noble metals Cu, Ag, and Au. LEED indicates that (111) planes have (1 ⇥ 1)
symmetry for all the metals. On Ni, Pd, and Cu the (100) planes also have (1 ⇥ 1)
symmetry. For other orientations, the situation becomes complex. The clean (100)
planes of Pt, Au, and Ir exhibit a complicated (5 ⇥ 1) symmetry. The (110) surfaces
of Au and Pt show a (2 ⇥ 1) LEED pattern. Electronic information on clean and
adsorbate covered surfaces is gained by photoemission experiments, while Auger
electron spectroscopy is widely used for chemical characterization of the adsorption
process at any stage. Our knowledge on di↵erent adsorption sites is based on
desorption spectroscopy (Section 8.3).

Figure 8.4: Flash desorption spectra for hydrogen on


W(110) and W(111). Ref. [1].

Hydrogen on tungsten is one of the classical examples; clean tungsten surfaces


have routinely been prepared for chemisorption studies. Desorption spectra have
revealed two major adsorption sites on the (110) and at least four sites on the
(111) plane.1 Figure 8.4 shows the relevant desorption data for H2 on W(110) and
W(111). The pressure caused by desorbing hydrogen is plotted as a function of
sample temperature, while maxima are related to individual adsorption/desorption
energies.
LEED data on W(110) indicate that there are two atomic states of equal density,
the 1 and 2 phases, with the existence of (1⇥1) and (2⇥1) periodicity, respectively.
1
P.W. Tamm and L.D. Schmidt, J. Chem. Phys. 51, 5352 (1969); 54, 4775 (1971).
CHAPTER 8. SURFACE REACTIONS 187

The saturation density is approximately one H atom per surface W atom, as shown
in Fig. 8.5. The 2 state occupies every other surface site and the 1 state occupies
the remaining sites.

Figure 8.5: Proposed adsorbate structures for H2 on the W(110) surface


with two states of equal density which fill sequentially. Ref. [1].

8.2 Reaction
A single catalytic reaction consists of a closed sequence of elementary processes or
steps. Summation of these steps, multiplied each by an appropriate stoichiometric
number reproduces the stoichiometric equation for the reaction. In the first step,
a species called the catalyst enters as a reactant whereas it appears as a product in
the last step of the sequence.
The catalyst may be a gaseous molecule or a grouping of atoms at the surface
of a solid called the active site and denoted by an asterisk “⇤”. In the latter case,
catalysis is called heterogeneous. It is only a special case of a general phenomenon.
The great technological advantage of heterogeneous catalysis is the easy separation
between the solid catalyst and the fluid reaction medium. The two basic mecha-

Table 8.1: Dissociative and associative mechanisms in ammonia synthesis. Ref. [2].

Dissociative Associative
2⇤ + N2 )* 2⇤N 1 * ⇤N2
⇤ + N2 ) 1
2⇤ + H2 *) 2⇤H 3 ⇤N2 + H2 *) ⇤N2 H2 1
⇤N + ⇤H * ) ⇤NH + ⇤ 2 ⇤N2 H2 + H2 *
) ⇤N2 H4 1
⇤NH + ⇤H * ) ⇤NH2 + ⇤ 2 ⇤N2 H4 + H2 *
) 2NH3 + ⇤ 1
⇤NH2 + ⇤H * ) ⇤NH3 + ⇤ 2
⇤NH3 *) NH3 + ⇤ 2
N2 + 3H2 * ) 2NH3 N2 + 3H2 *
) 2NH3
CHAPTER 8. SURFACE REACTIONS 188

nisms of catalysis are dissociative and associative. As an example, let us postulate


two possible sequences of steps in ammonia synthesis (Table 8.1).2 Qualitative con-
siderations suggest that the steps in a dissociative mechanism will have a higher
probability (entropy factor) but also higher energy barrier (energy factor) than their
associative counterparts.3 Hence a dissociative mechanism may be favored at high
temperatures, while at low temperatures associate mechanisms might prevail.
The first quantitative measure of the catalytic act is the rate of the single reac-
tion called the activity of the catalyst. The best measure of the rate, which lends
itself to comparison of the activity of various catalysts is the turnover number de-
fined as the rate per mole of site, that rate itself being defined as the rate of change
of the extent of reaction with time. In most cases, only an average or nominal
number may be obtained, as the number and types of sites may not be known.
The activity of a catalyst is rarely its most important characteristic. Selectivity,
which is defined as a ratio of rates, that of a desired reaction over the sum of rates
of other undesirable side reactions, is often more important to achieve and more
subtile to understand.
The habit of catalytic materials is dictated by technological usage and by the
frequent need for a very large specific surface area. Thus, porous materials with
average pore dimensions around 10 nm are often used as the catalyst itself or as
a support or carrier for the catalytic material. Clearly, the texture of a porous
catalyst, i.e., the shape of a replica of the material is of great importance in deter-
mining activity and selectivity because of the unavoidable gradients of temperature
and concentration in the porous medium.
Quantitative measures of the texture of catalytic materials, namely the total
specific surface area and the pore size distribution, can be obtained from isotherms
of physisorbed nitrogen, according to standard methods which have received uni-
versal acceptance.
In catalysis there are promoters and poisons which must be considered together:
they consist of additives introduced with the reactants or during the catalyst prepa-
ration and they a↵ect catalyst activity or selectivity. A first kind of promoter, said
to be textural , prevents loss of surface area of the catalyst. A typical example is
alumina Al2 O3 introduced in small quantities during the preparation of ammonia
synthesis iron catalysts. After reduction, the catalyst consists of metallic iron par-
ticles with a mean diameter of about 35 nm. The alumina is unreduced and covers
about half of the iron surface, preventing sintering of metallic particles.4
Another kind of promoter is structural or chemical . An example is chlorine
in the selective oxidation of ethylene to ethylene oxide on silver catalysts. It ap-
2
M. Boudart, In Interactions on Metal Surfaces, ed. by R. Gomer, Springer Series Topics in
Applied Physics , Vol. 4, New York, 1975, p. 275.
3
G.K. Boreskov, New Approach to Catalysis, Catalysis Society of Japan, Tokyo, 1973.
4
P.H. Emmet, Structure and Properties of Solid Surfaces, ed. R. Gomer and C.S. Smith, The
University of Chicago Press, Chicago, 1955.
CHAPTER 8. SURFACE REACTIONS 189

pears that chemisorbed chlorine inhibits the activated or non-activated dissociative


chemisorption of oxygen on the metal. Since oxygen adatoms are responsible for the
non-selective oxidation of ethylene, the role of the promoter is to enhance selectiv-
ity. Since the promoter occupies part of the surface, its role may also be conceived
as that of a selective poison.
There are as many types of poisons as there are types of sites. Some are re-
versibly held, i.e., they can be removed under reaction conditions as is the case for
surface oxygen during ammonia synthesis. Some are irreversibly held as are most
types of carbonaceous residues accumulating on catalytic surfaces during reactions
involving hydrocarbons. These residues must be removed in a special step of cata-
lyst regeneration. In fact, catalyst deactivation during use is the rule rather than
the exception: it is the central problem in the transfer of any catalyst from the
laboratory to the plant.5
Reversible poisons are called also inhibitors especially when they are participants
in the reaction. Thus reaction products are often inhibitors as is the case for
ammonia in ammonia synthesis on iron.
The fraction of sites which is active in a given reaction depends on the catalyst
and on the reaction, which is called active centers and are the site or group of sites
taking part in the reaction. The identification of the active centers and the structure
of their complexes with the reactive intermediates under reaction conditions is the
central goal of research in catalysis. These are some classical ideas of catalytic
reactions.

8.3 Desorption
Breaking the bonds of surface species and their liberation from the surface is called
desorption. This can be accomplished in di↵erent ways. If the temperature of the
system is high enough that a sizeable fraction of the adsorbate complexes has ener-
gies above the desorption energy in the Maxwellian distribution and can therefore
leave the surface, one speaks of thermal desorption. Electron impact can lead to
transitions to excited states of the adsorbate whose potential energy at the equilib-
rium distance of the ground state is higher than that of the respective free particle,
so that ions or neutrals can be desorbed. This process is called electron-impact des-
orption or electron-stimulated desorption. Similar processes can be brought about
by excitation by light; we speak of photodesorption. The impact of ions or neutrals
can cause the removal of adsorbates in di↵erent ways; these processes are called ion-
impact desorption. Finally, a very high electric field can bend down the ionic curve
of the adsorbate so far that rapid tunneling occurs from ground state to the ionic
state, and the resulting adsorbate ion carries away from the surface immediately;
these processes are termed field desorption.
5
J.B. Butt, Advan. Chem. Series 109, 259 (1972).
CHAPTER 8. SURFACE REACTIONS 190

Measurement of desorption process, especially that of thermal desorption, have


widely been used to define adsorbate states (Fig. 8.4) and to measure their popula-
tions. Thermal desorption is an important step in heterogenous catalysis, and the
detailed analysis of the process is the basis of thermal desorption spectroscopy which
is still used with success. It was the Swedish scientist Arrhenius who established
the necessary formalism to calculate the desorption rate rd around 1920.
dNA d✓ Ed /kT
rd = = NA = NA Ce (8.5)
dt dt
Here, NA is number of adsorption sites per unit area, k is the Boltzmann constant,
Ed the desorption energy, and C a constant. The last part of the Eq. 8.5 is the
ansatz of Polanyi-Wigner.
If we let the temperature change linearly as T = To + t, then
NA C Ed /kT
rd (T ) = e (8.6)

dT dN dN dT dN
with = dt
. Recall that dt
= dT
· dt
= dT
.

Figure 8.6: The change of desorption rate rd with increasing temperature. The
desorption rate has a maximum value at TM , which helps us find the desorption
energy Ed . The area under the curve corresponds to the number of particles
liberated from the surface.

drd
The maximum value for rd requires dT
= 0, which leads to

Ed C Ed /kT
2
= e . (8.7)
kTM
In most cases there are more than one maximum value for rd , indicating more
than one desorption energy and, consequently, more than one di↵erent adsorption
site at the surface.
We di↵erentiate three main regions depending on the heating rate, :

1. Rapid heating implies dT


dt
= ! 1 and called flash desorption. It gives us
information on the total surface coverage.
CHAPTER 8. SURFACE REACTIONS 191

dT
2. dt
= 0 is isothermal tion and gives us information about the average time an
adsorbate spends on the surface.

3. 0 < < 1 is the region interesting for thermal desorption spectroscopy.

In a previous section we have seen a classical example of flash desorption spectra


of hydrogen on tungsten surfaces (Fig. 8.4) as well as model adsorbate surface
structures (Fig. 8.5). Palladium (110) surface is yet another prominent example,
because palladium is a promising candidate for hydrogen storage. Fig. 8.7 presents
a serious of thermal-desorption spectra of H2 liberated from Pd(110) as a function
of T . Prior to desorption experiments, Pd(110) was loaded with H2 at 100K.6

Figure 8.7: Series of thermal-desorption spectra of H2 /Pd(110). Ref. [6].

At each stage of adsorption, LEED is used to characterize the surface structure.


Especially the 1 and 2 phases are interesting. The 2 phase is reached with
1 ML of adsorption at 100 K which displays (2 ⇥ 1) surface structure. Additional
hydrogen with a total ✓ = 1.5 ML results in the 1 phase which shows (1⇥2) surface
reconstruction. There is a reversible transformation between (1 ⇥ 2) and (2 ⇥ 1)
phase as temperature and hydrogen gas pressure are modified.7
6
M.G. Cattania et al., Surface Sci. 126, 382 (1983).
7
K.H. Rieder et al., Phys. Rev. Lett. 51, 1799 (1983).
CHAPTER 8. SURFACE REACTIONS 192

Figure 8.8: Top view of hard-sphere model of Pd(110) surface with 1.0 (2 ⇥ 1
2 phase) and 1.5 ML (1 ⇥ 2 1 phase) of hydrogen adsorbed on the surface.
The surface unit cell is indicated. The large circles denote Pd atoms and
the full small circles H atoms. The hatched small circles indicate hydrogen
chemisorption sites on the second Pd layer which can be occupied because of
the reconstruction of the metal substrate (pairing of close-packed metal rows).
Metal atoms in the second layer are shown as dotted large circles. Ref. [7].
Index

antiphase domains, 129–131 EDAX, 84, 156, 157


Arrhenius, 6, 91, 92, 190 EELS, 145–147
Auger maps, 160, 161 energy gap, 7, 8, 10, 16, 21, 23, 25, 27,
32–35
basis, 22, 57, 58, 65, 75, 77, 107, 128 epitaxy, 1, 4–8, 14, 160
Bloch states, 2, 25–31, 39, 42, 107, 162 ESCA, 80, 148
Bragg condition, 23, 24, 61, 62, 122, 127 Ewald sphere, 62, 63, 111–113
Bravais lattice, 22, 57–60, 76, 107
Brillouin scattering, 181 far field, 78, 98, 107
Brillouin zone, 22, 23, 26, 27, 32, 35, 37, Fermi Golden Rule, 66, 142, 168, 180
38, 41, 42, 178 Fermi-Dirac distribution, 20, 21, 87, 148
Fischer-Tropsch, 4, 141
catalysis, 1, 4, 29, 38, 127, 165, 183, 185, fluorescence, 154
187–190 Fowler-Nordheim equation, 85, 100
central projection, 93 friction force, 104
chemical shift, 139, 150 FTIR, 170
chemisorption, 3, 91, 165, 172, 183, 186, fullerene, 11
189, 192
cleavage, 2, 14 gnomonic projection, 89
coherence length, 118, 119, 129 graphene, 10–12
Coster-Kronig transition, 158 graphite, 10, 11, 54, 55, 147
covalent bond, 28, 51, 53, 54
Haber-Bosch, 4
de Broglie, 74, 108 Hund’s rules, 45, 137, 164
Debye temperature, 9, 124, 125, 133
Debye-Scherrer, 63, 64 inner potential, 71, 122, 133
Debye-Waller factor, 71, 124, 126 interference function, 116, 117, 124
density of states, 17, 18, 43, 44, 66, 68, ionic bond, 55, 56
69, 87, 88, 101, 102, 145–147, 149, Koopmans’ theorem, 135, 137, 159
155, 156, 162
desorption, 86, 88, 90, 183, 185, 186, 189– Langmuir, 184, 185
191 Laue condition, 61–64, 117, 122
diamond, 10, 53, 54 lifetime, 155, 157, 158, 164, 179
di↵usion, 1, 6, 89, 91, 96, 165
mean free path, 13, 69, 81, 94, 97, 108,
dipole, 30, 51, 105, 137, 143, 146, 153–
109, 132, 133, 159, 160
156, 167–169, 171–177, 181, 182
Miller indices, 22, 60, 91

193
INDEX 194

momentum conservation, 12, 37, 39, 178, translation, 21, 22, 39, 40, 49, 50, 57, 58,
180 75, 129, 166
monolayer, 10, 14, 129, 130
unit cell, 22, 57–59, 65, 69, 75, 101, 103,
nanotubes, 1, 11, 148 113, 115, 116, 121, 128, 132, 192
near field, 79, 98
van-der-Waals bond, 51, 52, 54–56, 183
phase transition, 2, 9, 95–98, 120, 121 vicinal surface, 127
physisorption, 3, 183
point group, 22, 58, 61, 75, 76, 166, 168, work function, 30, 36, 38, 41, 85, 86, 96,
176 100, 165, 172
polarization, 66, 99, 105, 106, 139, 143, XAS, 139–141, 144–146
168–170, 172, 174, 176–178, 181, XPS, 149–151, 153, 154
182 XRD, 154
quantum well, 7, 8, 19 XRF, 155, 156
quasicrystals, 49, 50, 84, 92, 94, 95, 130,
131, 156, 159, 160

Rayleigh criterium, 74
Rayleigh scattering, 174–176, 178
reciprocity theorem, 122
reconstruction, 2, 3, 30, 31, 76, 77, 101–
105, 107, 115, 120, 121, 129, 130,
133, 191, 192
reduced-zone scheme, 26, 27, 39, 41
region of coherence, 116, 118, 119, 128
relaxation, 2, 30, 76
resonant photoemission, 47

segregation, 1, 3, 131
spin polarization, 43–45
stereographic projection, 98
sticking probability, 165, 184
Stokes scattering, 175, 178–180, 182
sudden approximation, 159
surface melting, 8, 9, 125–127
surface states, 29–31, 34, 35, 42, 88, 113
surface tension, 5

TDS, 191
TEM, 8, 61, 83, 147, 148
time reversal, 123
topografiner, 98, 100

Das könnte Ihnen auch gefallen