+ All Categories
Home > Documents > Lattice Dynamics Simulation of Ionic Crystal Surfaces: I...

Lattice Dynamics Simulation of Ionic Crystal Surfaces: I...

Date post: 05-Aug-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
41
Lattice Dynamics Simulation of Ionic Crystal Surfaces: I.~Criteria for Vibrational Mode Localisation and Surface Modes Andreas Markmann, 1 Jacob L. Gavartin, 2 and Alexander L. Shluger 2 1 Theoretische Chemie, Technische Universität München, Lichtenbergstraße 4, 85 747 Garching, Germany 2 Condensed Matter and Materials Physics Group, Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK Abstract A simple function, called attenuation parameter, is presented that can be used to extract quantitative information on the degree of localisation of normal modes at locations of arbitrary dimensionality from eigenvector components calculated in lattice dynamics simulations. Normal modes are identified as lo- calised if their attenuation parameter values exceed a certain threshold value. As examples for localisation at two-dimensional locations, modes localised at the KBr(001), MgO(001) and CaF (111) surfaces mod- elled by periodic shell model slabs are determined using the attenuation parameter for surface lo- calisation. This allows to gain an overview of the frequencies of the crystalline vibrations localised at the surface which is important for the vibrational interaction with adsorbed molecules. Surface modes exist at frequencies unoccupied by the bulk spectrum (in lagoons of MgO and CaF and in the vibrational gap in KBr) but also overlapping with it (resonant surface modes). Non-resonant normal modes are generally localised more strongly than resonant local modes. Ambiguities exist due to the finite size of the supercell model but physically relevant criterion cutoffs can be extracted from histogrammatic considerations. Exper- imentally measured surface modes that were not identified by previous theoretical studies are seen to have intermediate surface values. 1
Transcript
Page 1: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Lattice Dynamics Simulation of Ionic Crystal Surfaces: I.~Criteria for

Vibrational Mode Localisation and Surface Modes

Andreas Markmann,1 Jacob L. Gavartin,2 and Alexander L. Shluger2

1Theoretische Chemie, Technische Universität München,

Lichtenbergstraße 4, 85 747 Garching, Germany�2Condensed Matter and Materials Physics Group,

Department of Physics and Astronomy, University College London,

Gower Street, London WC1E 6BT, UK

Abstract

A simple function, called attenuation parameter, is presented that can be used to extract quantitative

information on the degree of localisation of normal modes atlocations of arbitrary dimensionality from

eigenvector components calculated in lattice dynamics simulations. Normal modes are identified as lo-

calised if their attenuation parameter values exceed a certain threshold value. As examples for localisation

at two-dimensional locations, modes localised at the KBr(001), MgO(001) and CaF2(111) surfaces mod-

elled by periodic shell model slabs are determined using theattenuation parameteraplane for surface lo-

calisation. This allows to gain an overview of the frequencies of the crystalline vibrations localised at the

surface which is important for the vibrational interactionwith adsorbed molecules. Surface modes exist

at frequencies unoccupied by the bulk spectrum (in lagoons of MgO and CaF2 and in the vibrational gap

in KBr) but also overlapping with it (resonant surface modes). Non-resonant normal modes are generally

localised more strongly than resonant local modes. Ambiguities exist due to the finite size of the supercell

model but physically relevant criterion cutoffs can be extracted from histogrammatic considerations. Exper-

imentally measured surface modes that were not identified byprevious theoretical studies are seen to have

intermediate surfaceaplane values.

1

Page 2: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

I. INTRODUCTION

Molecules adsorbed on a crystal surface can exchange energywith it by vibrational interaction.

The natural coordinates for describing vibrations in crystals are vibrational normal modes, the

modes of collective motion of atoms that do not interact witheach other in the harmonic approx-

imation. These vibrational normal modes can be classified bytheir frequency (or energy), their

wave vector and by the degree to which certain atoms participate in the mode. The aim of this

article is to use shell model lattice dynamics to calculate vibrational modes of three ionic crys-

tal surfaces – MgO(001), KBr(001) and CaF2(111) – and analyse them. In contrast tomolecular

dynamics calculations, atoms are not displaced from their minimum energy positions inlattice

dynamics and the dynamical properties of the crystal are described well only if the harmonic ap-

proximation is valid and the system is near the equilibrium geometry.

A well-known example for a surface mode is the Rayleigh wave1 which was predicted by

Rayleigh in 1885 as a special solution to the problem of vibrations in an continuous, semi-infinite

elastic medium. It is characterised by a rapid decay of vibration into the medium, such that 90%

of the vibrational energy is found up to a depth of one wave length. First microscopic calculations

of surface modes of atomic lattices, using Green’s functions, were carried out by Lifshitz and

Rosenzweig in 19482. The first calculations of surface phonon dispersion curvesusing an atomistic

model were done by de Wette and co-workers in 19713–5. The lowest energy mode in atomistic

simulations using a periodic slab model has Rayleigh-mode character. However, the participation

of atoms in this mode often decays more slowly than the Rayleigh wave with distance from the

surface. It is then called “generalised Rayleigh wave”6.

The presence of a surface introduces two perturbations to the model of the infinite bulk crystal:

A change of the Madelung field due to the truncation of the crystal and a change in the short-range

forces acting on the surface atoms due to their reduced coordination. In purely ionic crystals,

this often causes the force constants near the surface to decrease, i.e. the frequency of surface

modes is lowered with respect to the unperturbed7. This is in contrast to predominantly covalent

systems like silicon or germanium, where the force constants frequently increase near the surface.

In mixed ionic/covalent systems, surface modes of increased frequency may also exist alongside

such of reduced frequency (see for example ref. 8).

Sangster and Strauch7 considered the normal mode structures of diatomic linear chains with

single defects using Green’s function methods. They found that normal modes with energies split

2

Page 3: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

from the main band appeared only when the changes were strongenough and did not act to cancel

each other out (by, for example, increasing a mass but increasing the force constants acting on it

as well).

Most importantly, they found that when a mode has a frequencyoutside the bands of the per-

fect crystal, it is necessarily localised at the defect. Note that the converse is not true, i.e. there

may be normal modes that are localised at a defect but not split from the main band. These are

called resonant local modes or “resonances”. This means that localised vibrational modes cannot

be detected purely by analysing the spectral properties of acrystal. It may even be difficult to

detect spectrally separated local modes in computational applications, as there the model system

is always finite and hence the spectrum is discrete.

Slab calculations of ionic crystal surfaces (reviewed in ref. 6) have predicted that modes lo-

calised at the surface plane have lower frequency than modeslocalised at the first sub-surface

plane and so on up to a depth that is determined by the strengthof the perturbation caused by the

introduction of the surface (depending on the crystal).

Experimental measurements of surface phonon dispersion curves have been developed by Il-

bach and Mills9 using Electron Energy Loss Spectroscopy (EELS) and by Toennies and co-

workers10 using Inelastic Helium Atom Scattering (HAS). The HAS technique is widely used,

mostly to measure low-energy mode excitations, due to its favourable properties such as high in-

tensity, low energy, monochromatism, sensitivity only to the surface, charge neutrality (avoiding

the charging problems of EELS), chemical inertness and highresolution. It has recently been

further developed in order to find higher energy modes such asshear horizontal11 and optical sur-

face modes12 in the NaCl(001) surface and high frequency surface modes such as in CaO(001)13.

Furthermore, all-optical techniques using time-resolvedsecond-harmonic generation have been

developed by Tom and co-workers, which allow to probe modes at buried interfaces14.

We would like to perform a comparative study of surface dynamics of three strongly ionic

crystals with similar (cubic) structure with stable and nonreconstructing surfaces but with distinct

bulk vibration properties - MgO(001), KBr(001) and CaF2(111). Ideally, we would like to make

a connection between their bulk and surface dynamic properties. More specifically, we compare

localization (spatial extent) of the surface modes.

As indicated above, spatial localization of the phonons is related to spectral localization7. It

is important to distinguish between the density of states (DOS) view and the dispersion spectrum

view. To determine the density of states, the vibrational spectral density is integrated over the

3

Page 4: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Brillouin zone, so that vibrational modes non-resonant to the DOS are either above the Debye

frequency (ie. spectrally above the entire bulk spectrum) or in a density gap, if one exists. For

the dispersion spectrum, each k-point in the Brillouin zone(or at the boundary of its irreducible

element) is considered separated from the others. There maybe regions in k space where no bulk

modes exists over a certain frequency range that is, however, bulk-resonant at other k-points so

that this frequency range does not show up as gap in the DOS. Such frequency and k-point regions

are called lagoons because they show up as such in the dispersion spectrum. Vibrational modes

of a surface that reside in lagoons of the bulk spectrum are also necessarily surface-localised. The

Rayleigh mode is one such lagoon mode, as it exists below the bulk at all k-points but at non-zero

frequencies and hence is resonant with the bulk DOS. Note, however, that in finite model systems

(such as the ones we used) the bulk spectrum is discrete and hence the terms “gap” and “lagoon”

are ill-defined and are based only on visual inspection of thedispersion spectra. Correspondingly,

it is also difficult to distinguish bulk-resonant modes neara lagoon edge from modes residing

inside the lagoon.

On the other hand, surface-localised modes may exist that are bulk-resonant. These modes

cannot be identified by their position in the dispersion spectrum but only due to the decay with

distance from the surface of their atomic participation. Wewill employ a method to classify

vibrational modes based on their atomic participation and hence need not rely on the identification

of gaps or lagoons. It may be argued that gap and lagoon modes need to be localised more strongly

than bulk-resonant modes. In fact, it will turn out that for the crystals considered, the bulk-resonant

modes exist whose localisation is comparable to gap and lagoon modes.

In essence, it can be said that spectral localisation is a sufficient but not necessary condition for

surface localisation and that spectrally localised mode are among the most strongly localised but

bulk-resonant modes exist that feature the same degree of surface-localisation.

Dispersion spectra of thin slabs modelling the MgO(001) surface15 predict normal modes up

to a frequency of 21.5 THz, in good agreement with experiment. Dispersion spectra from rigid

ion calculations in ref. 15, however, show the typical overestimate of optical frequencies (up to

32 THz). This supports our view that we need to take polarisability into account to obtain useful

results. For both models, the overall shapes of the dispersion spectra agree, with lagoons predicted

around the� and X points.

A comparison of shell model slab calculation and HAS resultsfor the KBr(001) surface in

ref. 15 shows a good quantitative agreement between theory16 and experiment17, especially for the

4

Page 5: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

modes that are split from the bulk bands. The maximum vibrational frequency in KBr is 4.9 THz.

Due to the larger anion-cation mass ratio, KBr has a gap between the acoustic and optical bands,

at 2.8 THz. Lagoons are reported in the optical band at the X point and, to a small extent, the�point.

The CaF2(111) surface phonon Density Of States (DOS) based on a shellmodel has been

calculated and compared to the bulk DOS by Allan and Mackrodt18. Theoretical CaF2 surface

dispersion curves have been compared with experimental results by Jockisch et al.19. They predict

vibrational frequencies ranging up to about 13.8 THz using the shell model and an unrelaxed slab.

Lagoons are predicted at the K point and between the M and the� points.

We have performed periodic slab lattice dynamics calculations using the setup described in

section II, using shell model parameters that are speciallygeared towards reproducing crystalline

elastic properties. In section III, we present a simple function, called attenuation parameter, that

can be used to extract quantitative information on the degree of localisation of normal modes at

locations of arbitrary dimensionality from eigenvector components calculated in lattice dynamics

simulations. Normal modes can be identified as localised if their attenuation parameter values

exceed a certain threshold value. Our lattice dynamics results and their analysis are presented in

section IV and further discussed in section V.

II. MODEL

To model the interionic interactions, the shell model was used. This is a useful compromise

between the rigid ion model, which suffers from overestimated optical frequencies, and higher

accuracy methods such as the breathing shell model or ab initio methods which would prove too

demanding for our purposes. The parameters for MgO are due toStoneham and Sangster20 and

have been used in a number of studies, for example to model an AFM tip21. The KBr parameters

are due to Catlow et al.22. Table I reviews the shell model parameters used to model CaF2. They

were modified by Foster23 from a set developed by Binks24. The parameter sets used are specifi-

cally fitted to reproduce specifically to reproduce well the experimental bulk elastic and dielectric

constants. These are properties that are among the most relevant for lattice dynamics calculations.

A crystal surface can be modelled in one of several ways. Leaving aperiodic approaches aside,

a supercell may be used to model the system where, to conservecomputational effort, lattice

dynamics are to be performed only on the ions near the surfaceof the film. This would have the

5

Page 6: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

advantage that the “fixed” ions in the supercell can be placedat their bulk coordinates, so that a

semi-infinite surface could be approximated.

We have performed calculations of the MgO(001) surface withthis setup. Among others,

modes localised at the interface between the lattice dynamics region and the fixed region resulted.

This implies that the interface acts as an open end to vibrations, a behaviour that is clearly unphys-

ical as the interface is artificial. This is in accordance with a shear instability in MgO identified

from first principles calculations25. Therefore, this model for a semi-infinite crystal has to be dis-

carded for the purpose of dynamical calculations of ionic crystals. Note that for some covalent

systems such as carbon nanotubes such vibrational modes localised at the fixed/lattice dynamics

interface are not observed26, so that in these systems such an approach may be more useful.We

have used a supercell whereall ions take part in the lattice dynamics calculation.

The interest of this article lies with the normal modes of theslab and their frequencies. These

properties are very sensitive to stress, so the translationvectors of the unit cells parallel to the

surface were relaxed to zero stress. Strictly speaking, theresults presented below correspond to

the vibrations in a thin film (which may be deposited on a substrate with which it interacts very

weakly). It may be hoped, however, that at a sufficient thickness the results approach the surface

vibrations of a semi-infinite crystal. It must be stressed again that the model system is finite in the

direction of the surface vector, so the problem of surface-localisation is ill-defined. However, nor-

mal modes whose ionic participation decays towards the slabmiddle will be identified as surface

modes.

To build up the supercell used for the (001) surface, an odd number of atomic layers consisting

of one cation and one anion each are stacked in alternating positions (as illustrated in Fig. 1a). An

odd number of layers is used in order to avoid a macroscopic total dipole-moment due to different

surface relaxation of anions and cations which may lead to supercell interaction errors in a periodic

calculation. Three-dimensional periodic boundary conditions (PBC) are used, i.e. an alternation

of crystal slabs with finite vacuum gaps is used to model surfaces.

The vacuum gap between slabs was set to twice the slab thickness which is more than enough

to converge the results. The side view of Fig. 1a will be used for the representation of particular

normal modes by displacement diagrams below. With the exception of �-point and bulk calcula-

tions, all calculations in this article were performed using eleven sampling k-points per segment

connecting the corners of the irreducible part of the Brillouin zone.

The bulk crystal ofCaF2 may be built up by the supercell shown in Fig. 2a – with the coordinate

6

Page 7: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

system aligned such that the (111) surface normal is in z-direction. This unit cell consists of three

layers with one Ca and two F ions each. Note that each layer normal is at an angle to the (111)

vector. More layers can be added by continuing the sequence A–B–C periodically. A supercell

with 15 ions, i.e. five layers, was chosen for calculations ofthe (111) surface. This allowed fast

calculations and was sufficiently extended to observe surface localisation. A view onto the surface

is shown in Fig. 2b. Note that the fluoride ion 2 protrudes fromthe crystal, while the fluoride ion

1 is below the calcium cation in that layer. The surface lattice vectors are indicated by arrows.

The (111) surface is the most stable in CaF2. This can be rationalised, looking at Fig. 2, as

follows: a crystal cleaves along the planes where the layer distance is largest and with it the

interlayer cohesion is weakest. In the MgO and KBr lattices,this is the (001) plane. Figure

2a demonstrates the large interlayer distanced along the (111) vector. The surface consists of

stoichiometric layers and hence is also stoichiometric. The ions at the boundary are fluoride ions.

III. DETECTION OF SURFACE MODES

If a surface is present, there may be vibrational modes in which atoms near the surface take

part to a higher degree than atoms far away from the surface. If the degree of participation decays

(ideally exponentially) with the distance from the surface, these modes are called localised at the

surface or surface (phonon) modes. All other modes are called bulk vibrational modes.

Surface modes are particularly interesting for studying the interaction of an adsorbed molecule

with a surface. Surface vibrational modes are relevant for other processes as well, such as dissi-

pation of energy from the tip of a scanning probe microscope,for example a scanning tunnelling

microscope (STM) or a non-contact atomic force microscope (AFM). This is because they con-

tribute most to the motion of substrate atoms at the surface.Conversely, a displacement of a surface

atom from its equilibrium position has a higher overlap withsurface modes than with bulk modes.

This can be seen by recalling that the normal mode vectorspn are orthonormal (jpnj2 = 1), i.e. a

surface modepn has larger componentspn(i�) at surface ionsi (with Cartesian directions�) and

by looking at the equation bx2i��n = ~2�i ��pn(i�)��2!n �Nn + 12� ;giving the expectation value of the quantum mechanical displacement operator of ioni in the

Cartesian direction�, whereNn is the population number of the vibrational moden, !n is its

7

Page 8: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

frequency and�i the ion mass. The equation connects the mean square displacement of ioni with

the population of vibrational moden. Particularlyhbx2i�in / ��pn(i�)��2.The participation of atomj in the motion of moden that we have used tentatively above is

strictly defined as Patom (n; j) = 3X�=1 p2n(j�): (1)

Note that, due to the normality of the normal mode vector, this has to be seen in relation to

unity. Note also that, away from the�-point (wherek=0 andD is symmetric rather than generally

Hermitean), the components of the vectorspnm are generally complex, allowing for non-zero

phases between the mode components.

We will confirm in this article that in MgO, KBr and CaF2 normal modes localised at surfaces

exist, and will demonstrate that these localised modes may be split from or resonant with the

main bands. To detect these localised modes we have used the analysis methods presented in the

following.

It is important to establish a well-defined criterion to decide whether a mode is localised at a

chosen point, line or surface, in order to be able to analyse the results of large calculations auto-

matically. Such a criterion in the form of an algorithm wouldideally be usable for any localisation

problem and useful to decide unambiguously about the localisation property of any mode. At least

it should offer a means of measuring localisationat a chosen location, in order to characterise

modes and postulate a meaningful threshold value beyond which a mode would be viewed as lo-

calised. An algorithm based on such a criterion allows to decide cases which are ambiguous to

the naked eye and to analyse results of calculations with very many ionic coordinates (and hence

equally many normal modes).

In order to get a general overview of the degree to which a moden is localised, the value of the

participation function, also called participation ratio can be used. It was initially introduced for

the analysis of wave functions27. For the time being, it can be defined intuitively aseP (n) = NXj=1 Patom (n; j)2 = NXj=1 3X�=1 p2n(j�)!2 : (2)

In order to interpret the participation function, two extremal cases can be considered:

1. All atoms participate in the mode to the same degree, i.e.Patom (n; j) = 1N for all j due to

the normality of the normal mode vector. In this case, the mode is fully delocalised and the

8

Page 9: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

participation function is eP (n) = NXj=1 1N2 = N � 1N2 = 1N � 1: (3)

2. The mode is localised on a single atom, in which case it is obvious thateP (n) = 1:The participation function measures to which degree a mode is localised. Theinverse participation

function P (n) = eP (n)�1 (4)

then gives a measure of the number of atoms that participate in the mode. Note that this approach

replies on a finite number of atoms in the model, as otherwise for delocalised modeseP (n) = 0.

In practical applications, information about the localisation of a mode is usually given in this

form and therefore the inverse participation function is often referred to as participation function.

The form used can be recognised by the limit compared with, which is unity for the participation

function and the total number of atoms in the simulation cellfor its inverse.

In theory, the participation function can be used to find normal modes localised at a defect in

a periodic lattice.30 Unfortunately, some of the normal modes detected as localised by the partic-

ipation function may in fact be delocalised, as a few ions perunit cell participating in a certain

normal mode may form sublattices extending over the whole crystal. Normal modes avoiding the

defect that are in fact extended only involve a subset of the supercell ions. In a calculation where

the size of the system is of the same order as the localisationlength of localised normal modes (as

is the case in the calculations presented here), these are also falsely detected as localised by~P .

In an infinite or semi-infinite periodic crystal, the sublattice modes and bulk modes avoiding

the defect would not be detected as localised by the participation function~P but since our model

has a finite size, this distinction cannot be made.

Additionally, neither participation function nor localisation length give information as towhere

a normal mode is localised, for example in a problem where twoor more defects exist. These

may be a surface (two-dimensional defect) with an admolecule (zero-dimensional defect for small

molecules). The information that the participation function can provide is therefore used as a

supplemental information measuring the degree of localisation of a mode that has already been

identified as localised at the location of interest.

Due to these problems, we propose a simple criterion for normal mode localisation that is

applicable to finite model systems. Note that in a finite model, the localisation problem is ill-

9

Page 10: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

defined, as only a finite number of ions exist that may participate in any given normal mode.

However, the decay of ionic participation with growing distance from the defect can be measured.

This is the basic idea used in the construction of our criterion.

We will use the example of surface modes to exemplify the proposed criterion but the method

is also transferable to special locations of any dimensionality. To model a crystal surface, we have

used a slab of2b� 1 crystallographic planes with vacuum gaps above and below. This means that

the slab has two surfaces which have to be taken into account.

Just looking at the participation of the atoms in the surfacemonolayers is not sufficient to de-

tect surface modes, for example, as these may be large for an extended mode that only involves

a sublattice. Instead, one should look for modes in which theparticipation of atoms decays sys-

tematically with the distance from the surface. Previously6, so-called “surface attenuation curves”

were analysed:

The surface attenuation curve is formed by summing up the atomic participations of all atoms

in the crystallographic planes. Because the numbering of the crystallographic planes from the

surface is a discrete parameter, it would be more appropriate to refer to the attenuation curve as

attenuation profile. Let n be the number of the phonon being analysed and = 1; : : : ; 2b � 1 the

index enumerating the planes, let} ( ) be the set of indices of all atoms contained in planec, such

that the disjoint union of all sets} ( ) is the set of all atoms in the lattice dynamics calculation:2b�1℄p=1 } ( ) = f1; : : : ; Ng :Then, using (1), the attenuation profile is given byPplane (n; ) = Xj2}( )Patom (n; j) = Xj2}( ) 3X�=1 p2n(j�):Note that for a surface mode of a slab model with thickness2b � 1, the attenuation profile will

generally decrease from = 1 to = b; a range which we will refer to as the upper side of the

slab, and then it will increase again from = b to = 2b� 1; a range which we will refer to as the

lower side of the slab.

Fig. 3a shows the surface attenuation profile of a fictitious sub-surface mode. Note that, due to

its systematic decay away from the surface, this mode shouldbe detected as a surface mode. Due

to the small participation of the surface atoms, however, itwould not be detected by an algorithm

that does not take any subsurface atoms into account.

10

Page 11: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

The often complicated behaviour of attenuation profiles is described in great detail in ref. 6 but

no mention of an automated analysis of the attenuation profile is made there. We must therefore

assume that attenuation curves were analysed visually whenthe surface character of a mode could

not be determined by its frequency or symmetry. However, systems with steps or with adsorbed

molecules, which we are also interested in involve very manyatoms, resulting in very many normal

modes to be analysed. A visual analysis then becomes impracticable for our purposes which is

why we have developed criteria in the form of algorithms for the detection of surface modes. It is

also often difficult to decide on the localisation of modes ofintermediate character. An algorithm

will help to make such decisions in a uniform way for all normal modes.

Instead of summing the squares of the atomic particiations as is done to obtain the participation

function, we propose a weighted sum of the atomic participations geared towards detection of

localisation at some location (here the surface). We first consider the part of the slab enumerated

from 1 tob. The other part can then be treated symmetrically. Letwb ( ) be a weighting function

defined by wb ( ) = 12 � 1� ar tan 12 � � b3�b ! : (5)

This function, as shown in Fig. 3b forb = 12; has the following numerical properties for everyb > 0: 1: limx!1wb (x) = 0; limx!�1wb (x) = 1;2: wb � b3� = 12 ; wb � b4� = 34 ; hen ewb � 5b12� = 14 ;3: wb (0) � 0:92; bR0 wb (x) dx � 0:35 � b; wb (b) = 0:04:A change of the parameterb converts corresponds to a stretching of the abscissa. Summation

of thewb�weighted attenuation profile values yields a parameter thatcan be used for deciding on

the degree of surface localisation of a certain mode:aplane (n) = bX =1 wb ( � 1) � Pplane (n; ) + 2b�1X =b+1wp (2b� 1� ) � Pplane (n; ) : (6)

The second sum in (6) takes the lower half of the slab into account (those layers with indices

from p = b to p = 2b � 1). Note that summation starts atp = b + 1 to avoid double counting of

the middle layer. Due to the normality of the normal mode vector, the behaviour of this parameter

can now easily be connected to the degree of localisation at the surface:

The displacements of atoms near the surface are weighted higher than the displacements of

atoms inside the slab. If and only if moden is a surface mode, atoms at the surface are participating

11

Page 12: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

most, resulting in a value ofaplane near unity. If all planes participate equally in moden thenPplane (n; ) = 12b�1 for all c and a value ofaplane of around20:35�b2b�1 � 0:35 results (the factor of

two stems from the sum over the two sides of the slab). The modecould even be localised at the

centre of the slab, in which caseaplane almost vanishes.

We call the functionaplane the plane attenuation parameter (surface attenuation parameter if

the plane of interest is the surface) and propose the following criterion for detection of a surface

mode:

Moden is considered a surface mode if the functionaplane exceeds a certain critical

value0:35 < a rit � 0:92 (also called criterion cutoff or threshold):aplane (n) > a rit: (7)

Note that the weighting function is not perfectly flat anywhere. If it were nearly flat near the

surface, surface and sub-surface modes could not be distinguished. Due to the larger derivative

(modulus) of the weighting function in the intermediate area, however, surface and subsurface

modeaplane values are still sufficiently separated from bulk mode values.

Properties of the Proposed Criterion

1. It is smoothly scaleable.It can be applied to slabs of all thicknesses by adapting the param-

eterb in eq. (5), yielding comparable results.

2. The cutoff valuea rit is arbitrary. This is both a disadvantage and an advantage. A

parameter-free criterion would yield incontestable results but will remain elusive since the

notion of localisation in a finite system is, strictly speaking, ill-defined. On the upside,

the cutoff parameter can be set to a value that leaves a minimum number of normal modes

ambiguous. This can easily be done based on a histogram of thevaluesaplane (n) over alln.

3. The criterion is applicable to other localisation problems. Instead of using the numberp of

the plane starting from the surface as an argument for the weighting function, the distance

from any defect of interest can be used. The parameterb in eq. (5) will then be the maximum

distance.

4. It can be generalised.Different weighting functions may be used to change the selectivity

of the criterion.

12

Page 13: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Using the approach indicated in point 3, localisation at an arbitrary plane, a line (e.g. a step at the

surface) or at a point (e.g. a point defect, kink or adsorbed molecule) can be analysed. We will

call the resulting attenuation functionsaplane, aline, andapoint (plane, step and point attenuation

parameters). This is a considerable extension of applicability compared to attenuation profile

approaches that rely on a layered structure in the supercell.

With the weight function presented above, the surface-localisation criterion can be heuristically

envisaged as a mode being identified as a surface mode iff it extends into the surface by less than

one third of the way to the slab centre.

IV. RESULTS

Lattice dynamics results presented here were obtained withthe GULP software by J. Gale28.

The vibrational mode vectors were analysed using a custom-made program.

A. Bulk Densities of States

To demonstrate the different vibrational properties of thebulk crystals, we have calculated the

bulk Densities Of State (DsOS) of the three compounds considered in this study with a k-point

sampling of15� 15� 15 k-points in the Brillouin zone.

The Monkhorst-Pack scheme29 is used by the GULP software to optimise the sampling to avoid

calculating mutually equivalent k-points twice. The resulting vibrational DsOS are shown in the

solid lines of figs. 4–6. It can be seen that MgO features much higher vibrational frequencies than

the other two compounds. Previous predictions of the frequency range made from models at the

same level (i.e. involving polarisabilities but not breathing shells) are reproduced.

MgO and CaF2 have no vibrational band gap, while KBr does due to the largermass ratio be-

tween anion and cation. It can therefore be expected that, ifsurface-localised normal modes are

found, they are resonant for MgO(001) and CaF2(111) and resonant or in the gap for KBr(001).

The maximum in the middle of the CaF2 DOS is due to transverse optical (TO) normal modes

involving only the fluorine sublattice of the crystal due to symmetry reasons. An analogous char-

acter is also expected for surface modes of the CaF2(111) slab in this frequency region. These

expectations will be reviewed at the end of this article whenthe surface-localised modes of the

atomistic model systems are known.

13

Page 14: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

B. Relaxation and Changes in the Vibrational Densities of State due to Formation of a Surface

None of the surfaces reconstruct due to surface formation. Relaxation occurs only perpen-

dicular to the surface (rumpling). Rumpling at the MgO(001)and the KBr(001) surfaces is, at

0.05 Åand 0.1 Å, respectively, within the range of experimental measurements. Rumpling in the

CaF2(111) surface is, at below 0.01 Å, almost absent in ab initio calculations, which was taken

into account in the parameterisation of the shell model23.

In order to avoid a macroscopic total dipole-moment in the slab which would, due to periodicity,

lead to errors in the calculation, the number of layers in allslab calculations is odd, making the

slab strictly symmetric. This means, however, that bulk calculations corresponding to the slab

supercells cannot be performed just by setting the vacuum gap to zero since that would cause ions

of the same charge to touch at the interface.

Instead, the DOS resulting from lattice dynamics of the bulkunit cell with the Monkhorst-

Pack29 k-point sampling scheme is used for comparison. The surfaceDOS is formed by sampling

the boundary of the irreducible part of the Surface Brillouin Zone (SBZ). For the (001) surface of

MgO and KBr, this is simply the isosceles triangle� = (0; 0; 0), M = (0:5; 0; 0), X = (0:5; 0:5; 0)in reciprocal space, as illustrated in Fig. 4b.31 This sampling was chosen because the results of this

calculation can immediately be used to produce dispersion spectra.

Fig. 4a shows a comparison between the MgO surface DOS of a slab 15 layers thick and the

bulk DOS. MgO is a comparatively hard crystal, with a maximumfrequency of around 21.5 THz,

and has no vibrational band gap. It is purely ionic, i.e. surface-localised normal modes are expected

to be lowered in frequency with respect to their bulk counterparts. As no phonon band gap exists,

high-frequency surface-localised normal modes must be resonant with the bulk spectrum.

Note that the resolution of the surface DOS (dotted line in Fig. 4a) is higher than that of the

bulk DOS. Only the main structures should be considered. Thesurface DOS does not feature

frequencies absent in the bulk DOS, i.e. surface modes are expected to be mostly resonant with

the bulk DOS. This is in accordance with previous work15.

Fig. 5 shows a comparison between the KBr surface DOS of a slab15 layers thick and the bulk

DOS. As for MgO, the surface DOS is formed by sampling the boundary of the irreducible part of

the quadratic SBZ.

The surface DOS adds a small amount of density in the vibrational gap of the bulk spectrum but

also introduces considerable changes resonant with the bulk spectrum. Surface modes are therefore

14

Page 15: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

expected to be found in the gap as well as resonant with the bulk DOS. This is in accordance with

previous work15.

Fig. 6 shows a comparison between the CaF2 surface DOS of a slab 15 atoms thick and the bulk

DOS. The surface DOS is formed by sampling the boundary of theirreducible part of the SBZ. For

the (111) surface of fluoride structure crystals, this is thetriangle� = (0; 0; 0), M = (0:5; 0; 0),andK = �23 ; 13 ; 0�, as illustrated in Fig. 6b.32

The surface DOS does not feature frequencies absent in the bulk DOS, i.e. surface modes

are expected to be resonant with the bulk DOS. This and the DOSchange are qualitatively in

accordance with a the calculation of the surface DOS by Allanand Mackrodt18, where, however,

a different parameterisation was used. The frequency rangepredicted here is somewhat smaller

(maximum frequency of�12 THz vs.�14 THz).

C. Surface Modes Detected with the aplane Attenuation Criterion

Histograms of the surface localisation surface attenuation parameteraplane and the participation

function ~P (eq. 2) for 15 layer slab calculations of MgO, KBr and CaF2 are shown in Fig. 7.The~Phistograms (solid) have peaks at or below 0.2. These stand for the majority of normal modes that

involve almost all ions in the supercell. The tails at high values are made up of modes involving

only few ions in the supercell. Some of these normal modes may, however, be delocalised, as

the few ions participating may form sublattices extending over the whole supercell. Among the

normal modes in the~P histogram tail are also normal modes localised at the centreof the slab that

are present due to the finite slab thickness.

In a semi-infinite periodic crystal, the sublattice modes and bulk modes avoiding the surface

would not be detected as localised by the participation function ~P but since the model has a finite

width, this distinction cannot be made.

Theaplane criterion can distinguish between surface modes, normal modes localised at the slab

centre and sublattice modes. Theaplane histograms (dotted) have a broad peak around the medium

value of 0.35 which was predicted as average value for a bulk phonon (the integral under the

attenuation weighting function, as shown in Fig. 3b). This demonstrates bulk type characteristics

for the majority of vibrational modes. The tails of the histograms at higher values contain the most

localised vibrational modes. Criterion threshold values,indicated by vertical dotted lines, can be

set such that only a few normal modes with the highestaplane values are identified as surface modes

15

Page 16: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

and the number of ambiguous modes at the cutoff is minimal. Due to the more diffuse nature of the

CaF2 histogram, the threshold there has to be set lower to obtain acomparable number of surface

modes.

1. Surface Modes of the MgO(001) Surface

Fig. 8 shows the dispersion spectrum of a 15 layer MgO slab with surface modes marked. It

agrees well with previous publications where the shell model was used15. Lines in Fig. 8 represent

the dispersion relations for the 90 modes and points mark modes that were labelled as surface mode

with different cutoffs of theaplane criterion. Fig. 8 demonstrates that the generalised Rayleigh

wave is clearly split from the bulk and accordingly yields a comparatively highaplane attenuation

parameter value. The MgO dispersion spectrum features lagoons, i.e. gaps in the spectrum that are

present only at limited ranges of k points. However, since finite supercells are used, the spectrum

is discrete and hence the term “gap” is ill-defined. Statements on gaps and lagoons made here are

based solely on visual inspection of the dispersion spectra.

Optical surface modes can be found at the top end of lagoons aroundk = (0:5; 0:5; 0) and

around the�-point – here the lagoon seen at about 10 THz is not very wide and was not recognised

in previous work due to either omitted surface relaxation oruse of the rigid ion model6. Note that,

at a threshold value of 0.6, another surface mode is detectedat the bottom of the same lagoon.

This means that it has split off from thetop of its band, a sign that for this mode, the introduction

of the surfaceincreasesthe force constant. All other surface modes are resonant with bulk modes.

A comparison with the earlier results shows good agreement,however more detailed information

about the degree of localisation at the surface can be extracted from Fig. 8.

Resonant surface modes exist at the centre of the dispersiondiagram and on the left below

10 THz (circled in the middle panel of Fig. 8). These were not classified as surface localised in

previous theoretical studies6. This demonstrates that the surface attenuation parameteraplane is

useful for comparing surface modes of different character on an equal footing, preventing such

oversights.

The average participation function (over all k-points) of the 15 layer MgO slab is 14.3 (out of

30 ions). The peaks of the surface modes are, however, higherthan the peaks belonging to modes

localised elsewhere, so with a carefully chosen threshold value it may be possible to distinguish

surface modes and modes localised elsewhere. However, use of the participation function alone

16

Page 17: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

does not provide any help with the choice of this threshold value. It can only be defined a posteriori

with the knowledge of all surface modes in mind.

Fig. 9 illustrates the properties of some modes in MgO in the form of displacement diagrams.

Note that the lengths of the vectors shown correspond to mass-weighted eigenvectors as resulting

from the diagonalisation of the dynamical matrix rather than actual displacements. The different

directions of the arrows in Fig. 9 have to be explained: Away from the�-point, where the dy-

namical matrix is Hermitean but not symmetric, the vibrational mode displacements have non-real

eigenvalues. Let p1 = jp1j � ei'1 and p2 = jp2j � ei'2be two eigenvector components, with an eigenvalue!. The atomic displacements at timet from

the equilibrium position then arex1 = jp1jp�1 ei(!t+'1) andx2 = jp2jp�2 ei(!t+'2);so the relative phase between the eigenvector components isequal to the relative phase between the

two different ion displacements. If there is a non-zero phase between two coordinate components

of the displacement of the same ion, this ion will vibrate on an elliptical path rather than a linear

one. Such a path cannot be shown, however, in a diagram of thiskind.

What is shown, are the phases between ions. E.g. in Fig. 9b thearrows pointingtowardsions

signify that at!t = 2�n the ion is displaced and at!t = 2�n + �2 it is at its equilibrium position,

contrary to the intuitive notion that is valid for ions with arrows pointing away from them, where

the ion is at its equilibrium position at!t = 2�n and displaced at!t = 2�n+ �2 .

The surface mode shown in Fig. 9a is of Rayleigh type. The modes shown in Fig. 9b and (c)

are surface modes. Fig. 9d shows a mode with displacements increasing towards the middle of the

slab, to contrast the surface modes.

2. Surface Modes of the KBr(001) Surface

Our dispersion spectrum for KBr (Fig. 10) agrees well with previous publications where the

shell model was used16. It shows a marked contrast to that of MgO (Fig. 8): MgO features

much higher vibrational frequencies than KBr, (maximum frequencies – MgO:�2112 THz, KBr:�6 THz). KBr has a phonon band gap, so that modes appearing in the phonon band gap can

immediately be identified as surface modes.

17

Page 18: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

However, we demonstrate here that resonant surface modes exist in KBr as well (as before,

surface-localised modes are marked in Fig. 10). Unlike in MgO, the generalised Rayleigh wave

is not so clearly split from the bulk spectrum and accordingly yields a loweraplane attenuation

parameter value.

The normal mode of highest frequency is a longitudinal optical mode around 4.9 THz, in agree-

ment with ref. 16. Overall, the results presented here agreevery well with the previous results.

Two normal modes between 2.5 THz to 3.5 THz exist at all k-points, showing up as uninter-

rupted strings of dots in the dispersion spectrum. The higher of these is resonant with the bulk

bands and would not be detected without inspection of the ionic displacement in the modes.

The lowest frequency surface mode is the generalised Rayleigh wave. Due to its extended

nature, it is not detected at all k-points with higher threshold values. The number of atomic layers

is, however, comparatively small at 15, so detecting these modes with the threshold value 0.6 (in

the bottom panel of Fig. 10 is satisfactory.

Several other surface modes do not span the whole k-space, most notably modes around (0.5,

0.5, 0). The lowest of these is a bulk resonant mode, while thetwo others split from the bottom of

sub-bands into lagoons in the dispersion spectrum. At the threshold value of 0.6, a further mode

is identified as a surface mode at the bottom of the optical band around (0.5, 0, 0).

By relaxing the criterion threshold from 0.8 through to 0.5,more and more modes are des-

ignated as surface modes. The bulk resonant modes around the�-point from 2.3 THz downward

(circled in bottom panel of Fig. 10) are only detected at threshold values below 0.6. In previous ex-

periments, this surface mode branch is detected but it is notpredicted by some previous theoretical

studies15. The reason for this, it can now be said, is the weak localisation of these modes.

Fig. 10 demonstrates that among the most surface-localisednormal modes (top panel) there

are gap as well as bulk-resonant modes. As the criterion cutoff is relaxed, modes residing at the

top of lagoons as well as resonant modes are added to those identified as surface-localised. This

means that, while gap and lagoon modes are among the most localised, bulk-resonant modes at

the same level of localisation exist. As for surface localisation in general, a frequency outside the

bulk bands is a sufficient but not necessary condition for a high degree of localisation.

The diagrams in Fig. 11 illustrate the atomic displacementscorresponding to some modes in

KBr. The surface mode shown in Fig. 11a is an example for an optical surface mode featuring

strong localisation. The mode shown in Fig. 11b shows the gapmode, an optical surface mode

that involves only a sublattice, i.e. has nodes on the ions tothe right. (This immediately means

18

Page 19: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

that it is degenerate with a mode that involves the other sublattice.)

To contrast the surface modes to a bulk mode, which form the overwhelming majority of modes,

we present in Fig. 11c an optical mode which is actually localised around the middle of the slab,

as the displacements become bigger with distance from the surface.

3. Surface Modes of the CaF2(111) Surface

Monolayers at the CaF2(111) surface consist of three ionic layers. For the purposeof defining

theaplane attenuation parameter, the ionic layer furthest from the middle of the slab is postulated

as location of the surface. The weighting functionwR (r) assigns different weights to different

ionic layers within one monolayer.wR (r) is, however, flat enough near the surface that this will

not affect the physical relevance of the result.

Fig. 12 shows the dispersion spectrum of CaF2 calculated with the present model. The overall

shape of our dispersion spectrum agrees well with the previous shell model publication by Jockisch

et al.19 but we find lagoons around the K and� points, not between the M and� points. We

attribute this discrepancy to the fact that the results of ref. 19 are based on a slab model fixed at the

bulk structure. This means not only that rumping is absent (which is small in CaF2) but that also

that changes in the lattice constant parallel and normal to the surface (of the order of 1%) were

previously not taken into account.

Marked in Fig. 12 are the frequencies of surface modes detected for the threshold values 0.85,

0.75 and 0.65 in the framework of the dispersion spectrum. These values were used as fewer

normal modes reach the cutoff of 0.8 that was used for MgO and KBr. As it is possible to fill

in the dispersion spectrum with more and more surface modes detected according to decreasing

cutoff values, some previously unassigned experimental data points19 around the K point�23 ; 13 ; 0�

at about 7 THz and around the�-point from 6.7 THz downward can now be assigned (circled in

Fig. 12).

Other features of the previous study are qualitatively reproduced, although the frequency range

is different due to the different shell model parameter set.In terms of the range of vibrational fre-

quencies,CaF2 is intermediate between MgO and KBr, with a maximum frequency of 11.4 THz.

Lagoons between the LO band and the TO band open up around the�, M and K points. Spectrally

localised surface modes are seen in these areas, at 9 to 10 THz. Resonant surface modes can be

found at frequencies between 5 and 8 THz, mainly near the�-point.

19

Page 20: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Fig. 13 shows displacement diagrams of some normal modes of afifteen layer CaF2 slab that

are detected as localised at the surface with a cutoff value of a rit = 0:75.

Rayleigh type surface modes are omitted as their displacement diagrams look as expected.

Fig. 13a shows a doubly degenerate�-point surface mode at 6.68 THz which is situated at the

bottom end of a lagoon. A resonant surface mode split about 1 THz above, at 7.8 THz, involving

motion perpendicular to the surface is shown in Fig. 13b. This normal mode is almost degenerate

with a similar surface mode at 7.8 THz that involves oppositemotion of the subsurface layer (not

shown). Note that these normal modes involve the surface andsub-surface fluoride ions more

strongly than the surface calcium ions. The small participation of the calcium ions seen in the

surface modes is symmetry-allowed due to the introduction of the surface that breaks the bulk

symmetry.

Away from the�-point, the displacements become more complicated, so thatthey can some-

times only be summarised incompletely in displacement diagrams. The surface mode shown in

Fig. 13c, atk = �23 ; 13 ; 0�, f = 9:3 THz, has a complicated ionic displacement that involves

circular motion around the equilibrium geometry.

Overall, surface modes are found in CaF2 at isolated spots in the dispersion spectrum, similar to

the situation in MgO. Hence the smaller rumpling in CaF2 does not seem to imply a smaller number

of surface modes. The difference to MgO lies in the existenceof surface-localised modes that are

dominated by the surface anions, corresponding to strictlysymmetry-forbidden participation of

the calcium ions in the bulk TO modes. Although to a lesser degree, the cationsdo take part in

these modes, as the surface breaks the bulk symmetry.

V. DISCUSSION

Table II shows the frequency range of the compounds discussed with the frequencies and k-

points of their highest-frequency surface modes. It can be seen that in the ionic crystals studied the

surface modes never occupy the highest-frequency parts of the bulk vibrational spectra nor are they

above them. Some experimental points above the bulk spectrum are reported for CaF2 by Jockisch

et al.19 which are not reproduced by our nor their theoretical models. All other classes of surface-

localised modes, i.e. gap, lagoon and resonant modes were found where the corresponding features

were present (gap only in KBr). While gap and lagoon modes areamong the most localised, bulk-

resonant modes at the same level of localisation exist. As for surface localisation in general, a

20

Page 21: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

frequency outside the bulk bands is a sufficient but not necessary condition for a high degree of

localisation.

Lattice dynamics provide information about the frequencies and spatial extent (localisation) of

normal modes. However, as these considerations are based onthe harmonic approximation, no

specific statements about the coupling between normal modes, i.e. vibrational energy dissipation,

can be made. Speculations about dissipation from moleculesadsorbed on the surface can be

obtained from the spectral properties and the localisationof the normal modes. We are interested in

small molecules, i.e. the surface vibrations are assumed tobe transmitted directly to the molecule

as opposed to modulation of optical modes due to admoleculesbridgeing neighbouring surface

ions.

If a molecule adsorbed on a surface has molecular normal modes of vibration at frequencies

where the crystal possesses surface-localised modes, vibrational coupling is likely to be enhanced.

Similarly, if the surface-localisation of a lattice mode isparticularly high, the amplitude of ionic

motion at the surface at a given temperature will be particularly high.

As expected from the comparison of the bulk densities of state of the three compounds with

the surface DsOS (sections IV A and IV B), surface-localisedmodes are mostly resonant with the

bulk spectrum in MgO and resonant or in the vibrational gap inKBr. As the surface attenuation

function scales with the slab thickness, a comparison of thedifferent materials can be made:

The gap mode in KBr is the most strongly localised among the compounds studied, however,

a resonant surface mode with comparable localisation exists in KBr. In CaF2, the most surface-

localised modes are dominated by the fluoride sublattice, analogous to the bulk, where the cations

are excluded from TO due to symmetry reasons. Calcium ions participate in the TO surface modes

to a small degree, however, as the surface breaks the bulk symmetry.

All crystals feature generalised Rayleigh modes at the bottom of the dispersion spectrum which

are, however, not necessarily localised very strongly. This may, however, be an artifact of the thin

film model. The surface-localised modes in CaF2 exist at isolated spots in the dispersion spectrum

(Fig. 12), similarly to MgO. The difference between the two crystals lies in the preference for anion

participation in CaF2 that appears to stem from the strictly symmetry-forbidden participation of

the anions in the CaF2 TO modes.

Some of the MgO surface modes are found in lagoons (upper panel of Fig. 8). One surface

mode between 10 THz and 15 THz almost spans the entire dispersion spectrum. In KBr, the

dispersion spectrum (upper panel of Fig. 10) two surface modes span the whole boundary of the

21

Page 22: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

irreducible part of the SBZ, while some more lagoon and resonant surface modes exist at lower

density around the k-pointX.

On the basis of the above, it can be speculated that enhanced dissipation into surface modes is

likely for molecules adsorbed on the surfaces if they feature molecular vibrations up the frequen-

cies shown in table II. Depending on the crystal, these may bevery low (i.e. easy to avoid). Of

the three crystals, KBr has the most pronounced surface localisation, hence admolecule vibrations

would dissipate most quickly into surface modes on this crystal.

On the other hand, surface modes are more likely to give up their vibrational energy to the bulk

if they are resonant. So it can be speculated that the crystaloffering longer dissipation lifetimes

for dissipation from the adsorption site into the surface would have shorter dissipation lifetimes

for dissipation from the surface into the bulk.

VI. CONCLUSION

We have used shell model lattice dynamics to predict the normal mode structure in the (001) sur-

faces of MgO and KBr and the (111) surface of CaF2. The localisation of modes can be analysed

quantitatively by the introduction of a well-defined measure for localisation at a chosen location,

as is done by theaplane attenuation parameter. If the measure for localisation lies beyond a certain

threshold value, the mode is said to be localised at that location. The threshold values can be

based on a histogrammatic evaluation of the attenuation parameter over all normal modes. Using

this criterion, vibrational modes localised at the surfaces of MgO, KBr and CaF2 are identified

among the modes resulting from shell model lattice dynamicscalculations.

The surface phonon calculations we have performed qualitatively reproduce earlier results that

were obtained with other methods. The fact that some surfacemode branches have been over-

looked by some previous theoretical studies is assigned to the weaker localisation of these surface

modes, as quantified by the surface attenuation parameter. We attribute our dissenting result for

the location of the lagoons in the CaF2 dispersion spectrum to the different model and lack of

geometry optimisation if the previous work19. Additionally to this, we present here the first quan-

titative study of the surface localisation of different modes, making it possible to analyse normal

mode localisation reproducibly.

KBr displays the strongest surface localisation of the three compounds. Four classes of modes

are found that consistently exceed anaplane value of 0.8 over all sampled k-points. At the same

22

Page 23: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

cutoff value, the normal modes identified as surface modes inMgO are scattered as short branches

over the dispersion spectrum. This means that, for most modes, the property of being a surface

mode only depends on the symmetry of the mode in KBr, whereas in MgO and CaF2 it is strongly

dependent on the wave vector as well.

When the criterion cutoff is relaxed to 0.6, more weakly localised normal modes are picked up

which, for example in the case of the generalised Rayleigh modes in MgO and CaF2, now appear

at all sampled k-points. Additional branches of surface-localised modes are detected as well, such

as those starting at medium frequencies around the� point. These surface modes show up in

the experimental data but were not predicted in some previous theoretical studies of KBr15. The

problems in these studies can now be ascribed to the low degree of localisation of these modes.

One surface mode is detected in MgO at thelower end of a lagoon, which runs counter to the

notion that surface-localised modes in ionic crystals generally appear with lowered frequency,

i.e. at the upper end of lagoons in the dispersion spectrum. This surprising fact necessitates further

analysis.

Spectral localisation outside the bulk band turns out to be asufficient but not necessary con-

dition for a high degree of localisation in all compounds. Alow anharmonicityand spectral sep-

aration between admolecule and substrate justify the selection of theproton desorption in HCl

adsorbed on the MgO(001) surface as example reactionfor our study of coherent control of a

molecule adsorbed on an insulator surface.

Acknowledgements

The authors are indebted to J. Harding and A.H. Harker at University College London for fruit-

ful discussions on shell model lattice dynamics. A. Markmann would like to thank the Physics and

Astronomy Department of University College London for his studentship and the Delbrücksche

Familienstiftung for further financial support. J. L. Gavartin would like to thank the Leverhulme

trust for funding.

� Electronic address:[email protected]

1 J. W. S. Rayleigh,On waves propagated along the plane surface of an elastic solid, Proc. London Math.

Soc.17 (1885) 4.

23

Page 24: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

2 I. M. Lifshitz, L. M. Rosenzweig, Dinamika kristalliqesko$i rexetki, zapol-n��we$i poluprostranstvo (Dynamics of semi-infinite crystal lattices), �. Eksp. Teor.Fiz. (J. Eksp. Teor. Fis.)18 (1948) 1012.

3 R. E. Allen, G. P. Aldredge, F. W. de Wette,Studies of vibrational surface modes. I. General formulation,

Phys. Rev. B4 (1971) 1648.

4 R. E. Allen, G. P. Aldredge, F. W. de Wette,Studies of vibrational surface modes. II. Monatomic fcc

crystals, Phys. Rev. B4 (1971) 1661.

5 G. P. Aldredge, R. E. Allen, F. W. de Wette,Studies of vibrational surface modes. III. Effect of an

adsorbed layer, Phys. Rev. B4 (1971) 1682.

6 F. W. de Wette,Study of surface phonons by the slab method, in: W. Kress, F. W. de Wette (Eds.),Surface

Phonons, Vol. 21 ofSpringer Series in Surface Sciences, Springer Verlag, 1991, p. 67.

7 M. J. L. Sangster, D. Strauch,Localised modes associated with defects in ionic crystals, J. Phys. Chem.

Solids51 (1990) 609.

8 T. Aizawa, W. Hayami, S. Otani,Surface phonon dispersion of ZrB2(0001) and NbB2(0001), Phys. Rev.

B 65 (2001) 024303.

9 H. Ilbach, D. L. Mills, Electron Energy Loss Spectroscopy and Surface Vibrations, Academic Press,

1982.

10 J. P. Toennies,Phonon inelastic scattering of He atoms from single crystalsurfaces, J. Vac. Sci. Technol.

A 2 (1984) 1055.

11 A. Glebov, W. Silvestri, J. P. Toennies, G. Benedek, J. G. Skofronik, Evidence for the shear horizontal

phonon mode on the NaCl(001) surface, Phys. Rev. B54 (1996) 17866.

12 G. Benedek, A. Glebov, W. Silvestri, J. G. Skofronik,Evidence for a surface optical phonon mode on

NaCl(001), J. Vac. Sci. Technol. A14 (1996) 1522.

13 K. Wolter, D. Scarano, J. Fritsch, H. Kuhlenbeck, A. Zecchina, H.-J. Freund,Observation of a localized

surface phonon mode on an oxide surface, Chem. Phys. Lett.320 (2000) 206.

14 Y.-M. Chang, L. Xu, H. W. K. Tom,Coherent phonon spectroscopy of GaAs surfaces using time-resolved

second-harmonic generation, Chem. Phys.251 (2000) 283.

15 W. Kress, F. W. de Wette (Eds.),Surface Phonons, Vol. 21 of Springer Series in Surface Sciences,

Springer Verlag, 1991.

16 W. Kress, F. W. de Wette, A. D. Kulkarni, U. Schröder,Surface dynamics of relaxed (001) slabs of alkali

halides and MgO, Phys. Rev. B35 (1987) 5783.

24

Page 25: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

17 W. P. B. G. Chern, J. G. Skofronik, S. A. Safran,Surface-phonon dispersion curves of KBr(001) via

helium-atom scattering: Comparison with calculations, Phys. Rev. B39 (1989) 12828.

18 N. L. Allan, W. C. Mackrodt,Calculated surface phonon densities of states of ionic oxides and fluorides,

J. Phys. Cond. Mat.1 (1989) SB189.

19 A. Jockisch, U. Schröder, F. W. de Wette, W. Kress,Relaxation and dynamics of the (111) surface of the

fluorides CaF2 and SrF2, J. Phys. Cond. Mat.5 (1993) 5401.

20 M. J. L. Sangster, A. M. Stoneham,The diffusion of ions with multiple valence - The oxidation of

transition-metal alloys, Phil. Mag. B52 (1985) 717.

21 A. I. Livshits, A. L. Shluger, A. L. Rohl, A. S. Foster,Model of noncontact scanning force microscopy

on ionic surfaces, Phys. Rev. B59 (1999) 2436.

22 C. R. A. Catlow, K. M. Diller, M. J. Norgett,Interionic potentials for alkali halides, J. Phys. C10 (1977)

1395.

23 A. S. Foster, Theoretical Modelling of Non-contact Atomic Force Microscopy on Insulators, Ph.D. the-

sis, University College London (2001).

24 D. J. Binks, Computer Modelling of Zinc Oxide and Related Oxide Ceramics, Department of Chemistry,

Ph.D. thesis, University of Surrey (1994).

25 B. Karki, G. Ackland, J. Craine,Elastic instabilities in crystals from ab initio stress strain relations, J.

Phys. Cond. Mat.9 (1997) 8579.

26 G. Cuniberti, private communication.

27 J. M. Ziman,Models of disorder: The theoretical physics of homogeneously disordered systems, 1st

Edition, Cambr. Univ. Press, 1971.

28 J. D. Gale,GULP: A computer program for the symmetry-adapted simulation of solids, J. Chem. Soc.,

Faraday Trans.93 (1997) 629.

29 H. Monkhorst, J. Pack,Special points for Brillouin-zone integrations, Phys. Rev. B13 (1976) 5188.

30 Such as surfaces or point defects.

31 Neglecting the lattice vector perpendicular to the surfacewhich plays no role, the surface lattice vectorsa1 anda2 span a 90Æangle, hence the reciprocal basis vectors areb1 = 2�a2 andb2 = 2�a1.32 Neglecting the lattice vector perpendicular to the surfacewhich plays no role, the surface lattice vectorsa1 anda2 span a 60Æangle, i.e.a1 = a0x, a2 = a0 �12x+ p32 y�, wherea0 is the length of the surface

lattice vectors andx andy are the cartesian unit vectors. Then the reciprocal basis vectors areb1 =25

Page 26: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

2�a0 �x� 1p3y� andb2 = 4�p3a0y which enclose an angle of 120Æ. From this the hexagonal shape of the

SBZ and the vectors M and K follow.

26

Page 27: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

ion charge [e]core-shell spring constant

Ca core 0.719

Ca shell 1.28134.05 eV/Å

F core 0.378

F shell -1.37824.36 eV/Å

ion1 ion2 A [eV] � [Å] C [eV�Å6] cutoff [Å]

Ca shell F shell 1140.0 0.303 0 40.0

F shell F shell 911.690.2707 13.8 40.0

Table I: Shell model parameters for the simulation of MgO andKBr. Top: shell model charges and force

constants; bottom: Buckingham potentialb (r) = Ae� r� � Cr6 parameters from ref. 23.

27

Page 28: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

material top frequencytop surface frequency k-point

MgO 21.5 18.3 (0:5; 0:5; 0)KBr 6.0 4.1 (0:5; 0:5; 0)

CaF2 11.5 10.0 (0; 0; 0)Table II: Maximum frequencies and maximum surface mode frequencies of crystals discussed. The column

on the right lists the k-points of the highest-frequency surface modes. The lowest frequencies are zero,

i.e. the frequency of the generalised Rayleigh mode at the�-point. Frequencies in THz.

28

Page 29: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Figure captions:

Figure 1: Setup of surface mode calculations in MgO and KBr.(a) Side view of the unit cell

with an odd number of atomic layers and two atoms in the surface unit cell. (b) Top view at the

surface with the unit cell atoms in black. The surface is built up by translating the unit cell along

multiples of the lattice vectors (shown as arrows). Images of the unit cell under this operation are

shown in gray to illustrate the orientation of the crystal relative to the surface lattice vectors.

Figure 2: (a) Side view and(b) Top view of the CaF$_2$ surface with surface unit cell (in

black) and surface lattice vectors. An equivalent, straight shape of the surface unit cell is indicated

by a thick black line linking the corresponding ions.

Figure 3: (a) Attenuation profile of a fictitious subsurface mode in a 17 layer slab(b = 9).The participation of atoms in this mode decays systematically with distance from the surface and

therefore the mode should be detected as a mode localised near the surface.\\

(b) Weighting function for the attenuation profile for a 23 layerslab(b = 12). Maximum(� 0:92),minimum(� 0:04), value 0.5~atb3 and integral from 0 tob (� 0:35 � b) are shown. This weighting

function multiplied with the attenuation profile shown in (a) will give a large number near 0.92,

while the attenuation profile of a bulk mode (with small displacement in the surface layers) will

give a small number near 0.04.

Figure 4: (a) Comparison of MgO bulk and 15 layer surface DsOS.(b) Diagram illustrating

the irreducible part of the surface Brillouin zone for the (001) surface of MgO and KBr at the

boundary of which k-points are sampled.

Figure 5: Comparison of KBr bulk and 15 layer surface DsOS.

Figure 6: (a) Comparison of CaF$_2$ bulk and surface DsOS. This result is qualitatively

comparable with Allan and Mackrodt18 although the frequency range is different due to the use of

the more realistic shell model.(b) Diagram illustrating the irreducible part of the Surface Brillouin

Zone for the (111) surface of fluoride structure crystals at the boundary of which k-points are

sampled.

Figure 7: Histograms for 15 layer slab calculations of some ionic surfaces of the~P function of

eq.~(2) and theaplane function from section III. From top to bottom MgO, KBr andCaF2: Cutoff

values to be used to identify surface modes are shown as dotted lines. It can be seen that the CaF2aplane histogram is more diffuse, while the~P histogram is more sharply peaked at low values than

for MgO and KBr.

29

Page 30: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

Figure 8: Dispersion spectrum of 15 layer MgO slab with surface modes. From top to bottom,

the threshold value is relaxed from 0.8 through 0.7 to 0.5–0.6.

Figure 9: Displacement diagrams of some MgO slab modes. The dash-dotted line at the bottom

shows the middle of the slab, i.e. the surface is drawn at the top. (a) Rayleigh type surface mode

at k = (0; 0; 0), f = 1.0 THz, (b) surface mode (lagoon mode) atk = �12 ; 12 ; 0�, f = 18.2 THz,

(c) surface mode atk = �12 ; 0; 0�, f = 12.7 THz,(c) bulk mode atk = �12 ; 12 ; 0�, f = 18.7 THz,

participation of ions in the middle is bigger than at the surface.

Figure 10: Dispersion spectrum of 15 layer KBr slab with surface modes. From top to bottom,

the threshold value is relaxed from 0.8 through 0.7 to 0.5–0.6.

Figure 11: Displacement diagrams of some KBr slab modes, allat k = �12 ; 0; 0�. The dash-

dotted line at the bottom shows the middle of the slab, i.e. the surface is drawn at the top.(a)

Optical surface mode at f = 3.6 THz,(b) sublattice optical surface mode (gap mode) at f = 3.0 THz,

(c) bulk mode at f = 2.2 THz, participation of ions in the middle isbigger than at the surface.

Figure 12: Dispersion spectrum of 15 layerCaF2 slab with surface modes for criterion thresh-

old values 0.85, 0.75 and 0.65.

Figure 13: Displacement diagrams of some CaF2 slab modes. The ions at the bottom are in the

middle of the slab as before, i.e. the surface is drawn at the top. Rayleigh type surface modes are

not shown.(a) k = (0; 0; 0), f = 6:7~THz, (b) k = (0; 0; 0), f = 7:8~THz, (c) k = �23 ; 13 ; 0�,f = 9:3~THz.

30

Page 31: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

odd

num

ber

of la

yers

(a) (b)

Figure 1: A. Markmann et al., ???

31

Page 32: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

x

y

z

C

B

A

laye

rs:

d

z =

(11

1)

(a)

z x

y

I

II2

1

(b)

Figure 2: A. Markmann et al., ???

32

Page 33: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

of a (sub)−surface mode

attenuation profile

layer from surface2 3 4 5 6 7 8 91lo

g di

spla

cem

ent i

n la

yer

(a)

0

0.25

0.5

0.75

1

0 1 2 3 4 5 6 7 8 9 10 11 12

weighting function

��������������������������������

��������������������������������

������������

������������0.35 b.

bulksurface (scaled accordingly)

wei

ght o

f dis

plac

emen

ts

0.04

0.92

(b)

Figure 3: A. Markmann et al., ???

������������������������������������������������������������������������������������������������������������������������������������������������������������

������������������������������������������������������������������������������������������������������������������������������������������������������������

b

a2

a1 b2

1

SBZ

M

X

Γ

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

0 5 10 15 20

frequency [THz]

MgO bulk vibrational DOSMgO surface DOS

(a) (b)

Figure 4: A. Markmann et al., ???

33

Page 34: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 1 2 3 4 5

frequency [THz]

KBr bulk vibrational DOSKBr surface DOS

Figure 5: A. Markmann et al., ???

��������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������

�������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������������� a2

b2

b

a1

1

SBZ

Γ

M

K

0

0.05

0.1

0.15

0.2

0.25

0.3

0 2 4 6 8 10 12

frequency [THz]

CaF2 bulk vibrational DOSCaF2 surface DOS

(a) (b)

Figure 6: A. Markmann et al., ???

34

Page 35: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

0

1

2

3

4 MgO P~ histogram

MgO asurf histogram

0

1

2

3

4

5

6

KBr P~ histogram

KBr asurf histogram

0

1

2

3

4

5

6

0 0.2 0.4 0.6 0.8 1

CaF2 P~ histogram

CaF2 asurf histogram

0

0

Figure 7: A. Markmann et al., ???

35

Page 36: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

0

5

10

15

20

25

freq

uenc

y [T

Hz]

mgo 15 layer dispersionthreshold .80

0

5

10

15

20

25

freq

uenc

y [T

Hz]

mgo 15 layer dispersionthreshold .70

0

5

10

15

20

25

freq

uenc

y [T

Hz]

0 0 0 .5 0 0 .5 .5 0 0 0 0

mgo 15 layer dispersionthreshold 0.60

" 0.55" 0.50

0

0

Figure 8: A. Markmann et al., ???

36

Page 37: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

template is mgorayleigh.jpg

(a)

template is mgosurf.jpg

(b)

template isflickk10p0035.jpg

(c)

template is mgonosurf.jpg

(d)

Figure 9: A. Markmann et al., ???

37

Page 38: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

0

1

2

3

4

5

6

freq

uenc

y [T

Hz]

0 0 0 .5 0 0 .5 .5 0 0 0 0

KBr 15 layer dispersionthreshold 0.60

" 0.55" 0.50

0

1

2

3

4

5

6

freq

uenc

y [T

Hz]

kbr 15 layer dispersionthreshold .80

0

1

2

3

4

5

6

freq

uenc

y [T

Hz]

kbr 15 layer dispersionthreshold .70

0

0

Figure 10: A. Markmann et al., ???

38

Page 39: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

template is kbrsurf.jpg

(a)

template iskbralsosurf.jpg

(b)

template iskbrnosurf.jpg

(c)

Figure 11: A. Markmann et al., ???

39

Page 40: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

0

2

4

6

8

10

12

14

freq

uenc

y [T

Hz]

0 0 0 1/2 0 0 2/3 1/3 0 0 0 0

CaF2 15 Layers - Comparison of Threshold Levels

dispersionthreshold 0.65threshold 0.75threshold 0.85

Figure 12: A. Markmann et al., ???

40

Page 41: Lattice Dynamics Simulation of Ionic Crystal Surfaces: I ...ursula.chem.yale.edu/.../refs/A1-AMarkmannPre-1.pdf · (111) – and analyse them. In contrast to molecular dynamics calculations,

(a) (b) (c)

Figure 13: A. Markmann et al., ???

41


Recommended