intermediate size have received a great deal of attention for the promising technological
applications in optoelectronics.
The present dissertation is focused on studies of physical properties of size-selected carbon
and silicon clusters. The research work is experimental - the electronic structure of small clusters
has been investigated employing photoelectron spectroscopy. Fragmentation spectroscopy has
been performed as well for a better comprehension of the photoelectron results.
1.1 The systems: Carbon and silicon clusters
Although carbon and silicon belong to the same group (IV) of the periodic table of the
elements, they represent a particularly interesting sequence. They show slightly different properties
which are attributed to the decreasing importance of pi bonding with increasing atomic
dimensions. This leads to the formation of a large variety of single and multiple bonding structures
in the case of carbon while mostly single bonding structures for silicon.
In crystals, the nature of their bonding is intermediate between metals and insulator – carbon
and silicon crystals are covalent materials. In chemical terminology this means that each atom of
the covalent material participates in four covalent bonds by sharing two electrons with each of its
four neighbours in the typical crystal structure of diamond. Significant covalent bond strength
arises if hybrid bond orbitals are formed from linear combinations of low-lying atomic s- and p-
orbitals. The highly directional sp3 hybrid bonds that result determine the diamond and zincblende
crystal structures of most common semiconductors. Overlapping hybrid orbitals on neighboring
tetrahedral coordinated sites produce bonding and antibonding levels which ultimately broaden
into the semiconductor valence and conduction band, respectively, of the band model theory.
Although the common nature of the bonds, the physical properties of carbon and silicon bulk
show some differences generally related to the resistivity:
• The most covalent carbon crystal - diamond - is an insulator;
• The hexagonal lattice form of bulk carbon - graphite - is a semimetal;
• Silicon, with the same crystal structure of diamond, is a semiconductor.
• There is no evidence of the existence of a graphitic form of silicon.
The atomic arrangement on the surface of covalent bonded materials is often different from
that in the bulk. The modification of the crystal lattice at the border is known as surface
reconstruction. In the electronic density of states (DOS) the surface states are defects of the band
structure which can be clearly distinguished from the other states only in presence of a band gap.
The study of the electronic properties of carbon and silicon particles provides a clue to the
origin of the band structure and surface reconstruction.
The information that can be gained by spectroscopic investigation is explained in the chapter
“Experimental Setup”. There the techniques treated are exclusively those employed in our
experimental work.
2 STATE OF THE ART: CARBON
Carbon clusters are fascinating examples of the richness and variety of carbon chemistry. Due
to the enormous bonding flexibility of carbon, viz. its unique ability to form stable single, double,
or triple bond, carbon clusters appear in a wide range of structural forms that are synthesised
spontaneously in hot carbon plasmas produced during the energetic processing of carbon-rich
materials. Such natural environments include the atmospheres surrounding carbon stars,
interstellar dust clouds irradiated by intense UV radiation, and terrestrial sooting flames [Ger87,
Her83]. The most widely used laboratory technique for producing carbon clusters is laser ablation
or arc-discharge vaporisation of graphite followed by supersonic expansion into an inert carrier
gas, usually helium or argon. Neutral, anionic, and cationic carbon clusters grow directly from the
plasma with the sizes ranging from one to hundreds of atoms. The size distribution produced in
this way can be controlled by varying such experimental parameters as laser power, gas pressure,
and the geometry of the supersonic nozzle. Elucidating the evolution of carbon cluster structure,
from linear chains to rings to closed spheroidal cages to nanotubes, that takes place as the cluster
size increases, constitutes a major scientific challenge and requires an intimate interplay of state of
the art experimental and theoretical techniques. In this chapter the literature on carbon clusters is
reviewed for two groups of clusters: the fullerenes and the other geometrical structures of carbon
clusters. This classification is based on the peculiarity of the cage clusters among all the possible
single-element clusters - fullerenes are the only atomic clusters which crystallise in new solids.
Low mass fullerenes are of particular interest because their high curvature and increased strain
energy owing to adjacent pentagonal rings could lead to solids with unusual intermolecular
bonding and electronic properties. The aim of this research project is to demonstrate the
existence of the smallest possible fullerene C20.
2.1 Particular class of clusters: Fullerenes
In 1985 Kroto and Smalley [Kro85], producing carbon clusters by a laser graphite-
vaporisation source have observed, for the first time that, under certain source conditions, the
intensity of C60 could be made to fully dominate the mass spectrum. The recognition of the unique
stability of the cluster with 60 carbon atoms has induced the researchers to propose a highly
symmetric truncated icosahedral structure (“football or soccerball”) for this molecule. It was
christened with the name “Buckminster fullerene”, or more commonly fullerene, in recognition of
the structure of Buckminster Fuller’s geodesic domes. Such a geometrical arrangement involves
no dangling bonds and has been calculated to be very stable with pi-electrons delocalised over
the outer and inner surfaces of the spheroidal structure.
An sp3-hybridized array of carbons is three-dimensional (e.g. diamond) but has a surface with
attendant (destabilising) dangling bonds. Similarly an sp2-hybridized planar fragment of a graphite
sheet has a reactive edge with dangling bonds. In C60, however, the incorporation of pentagons
into the hexagonal graphite planes allows the sp2-hybridized structure to bend around and form a
cage, thus removing any dangling bonds and being further stabilised due to the overlap of adjacent
pi-electrons. Each carbon atom has all valences satisfied by two single bonds and one double
bond. A C60 has many resonance structures, and appears to be aromatic. This kind of structure
has, in fact, been predicted on a number of occasions before the Smalley/Kroto experiments
were carried out. Simple geometrical considerations show that it is possible in principle to
envisage similar cage structures for all of the even carbon clusters with 20n ≥ , with the
exception of n = 22. The structural peculiarity is the presence of an even number of three-
coordinated sp2 carbon atoms that arrange themselves into 12 pentagonal faces and any number
(except 1) of hexagonal faces. Odd-numbered clusters will always have at least one atom with a
remaining dangling bond and will thus be very reactive and unstable. The cage should be stable if
the curvature related strain is symmetrically distributed (geodesic structure) and if the pentagons
are isolated as much as possible by the hexagons in order to avoid the inherent instability of
fused-pentagon configurations (see “isolated pentagon rule” in Kro87). The unique stability of C60
can thus be understood since it is the only 5/6-ring cage for which all the atoms are equivalent (i.e.
the strain is perfectly distributed) and it is the smallest fullerene for which all pentagons can be
isolated. The next fullerene which is able to avoid abutting pentagons is C70, which is observed
also to be particularly stable.
While great progress has been made in recent years in understanding the physical and
chemical properties of this new class of clusters, many unanswered questions still remain. For
example:
• The mechanism of formation of fullerenes is not yet well understood since their formation
does not require extreme plasma conditions – it accompanies soot formation by burning
candles [Sma90]. A pentagon road accretion as been proposed as mechanism for
fullerene formation [Kro88].
• The stability of small cage structures like C20, which is the smallest carbon cluster that can
exist as fullerene, is still an open question since different studies show conflicting
predictions concerning the relative energies of the possible isomers.
• The whole question of energy storage and conversion within C60 and the other fullerenes
is topical. Thermionic emission has been observed to occur from neutral fullerenes
[Cam91], giant (n = 150 – 600) positively charged fullerenes and negatively charged C60-
[Sma91] on a time scale of microseconds.
2.1.1 C20: Cage, bowl and ring structure
A dodecahedron consisting of 12 pentagons whose vertexes are occupied by 20 carbon
atoms is topologically the smallest possible fullerene. Its structure is shown in the left side of figure
2.1. Kroto [Kro87] has underlined that the existence of such a structure with adjacent pentagons
would be unfavorable for a cage structure. His theory is known as the “isolated pentagon rule”.
He formulated the hypothesis that curved graphitic networks, like bowl C20 (see the relative
structure in the middle of figure 2.1), are intermediates in the growth mechanism which produces
buckminsterfullerene and larger closed cages [Kro88]. This further theory is known as the
“pentagon road” in the formation of fullerenes. In laser ablation experiments of graphite under
conditions that produce fullerenes for clusters bigger than C30
-
[Han95a], only a monocyclic ring
isomer (see the right picture in figure 2.1) is produced containing 20 carbon atoms. The structure
has been assigned on the basis of the agreement between experimental and theoretical electronic
structures [Sai00].
In calculation, a bowl-shaped isomer which minimizes the number of dangling bonds in a
graphite-type structure, has been found to be competitive in energy with the fullerene and ring
isomers [Gro95, Jon97, Scu96]. In figure 2.1 the structures of the three candidates for the ground
state geometry are shown: cage (three dimensional), bowl (two dimensional) and ring (one
dimensional).
Figure 2.1: The three most probable isomeric form of C20. From the left to the right are
shown cage, bowl and ring.
Empirical and semiempirical previsions are of little help in predicting the most stable structure
of C20. The ring, bowl, and fullerene isomers contain different types, as well as different number
(20, 25, and 30 respectively) of bonds.
Entropic factors greatly favor the ring isomer over the bowl and fullerene isomers at high
temperature.
The relatively low energy of the bowl isomer is consistent with an accretion mechanism
(pentagon road) of fullerene formation [Kro88] in which the growing carbon sheets have enough
time to anneal to their lowest energy forms, simultaneously minimizing the number of dangling
bonds. The absence of the bowl isomer in mobility experiments may simply be related to entropic
and kinetic (high-reactivity) factors.
Figure 2.2: The relative energy differences for the ring, bowl, and cage C20 isomers. For
each theory the lowest energy structure is taken as a reference. LDA (Local Density
Approximation), HF (Hartree-Fock), DMC (Diffusion Monte Carlo), BLYP (Becke-Lee-
Yang-Parr functional) [Mit00].
The striking feature about this problem is the enormous discrepancy between the various
theoretical methods. The graph below shows the relative energies of these three clusters
calculated following four different methods: Local Density Approximation, Hartree-Fock,
Diffusion Monte Carlo, Becke-Lee-Yang-Parr functional (figure 2.2). The data are obtained by
[Gro95] and are corrected with respect to the performed last calculation [Mit00]. Therefore, a
relative energy scale of the different structures cannot be established and thus the ground state
geometry cannot be predicted. Furthermore, stability alone does not determine which isomer
actually is formed preferentially.
Galli et al. [Gal98] have calculated the density of the occupied valence states of C20 for the
three isomeric structures. The results for the anion and neutral samples are shown in figure 2.3.
The dotted line is relative to the anions, the straight line is relative to the neutrals.
Figure 2.3: Density of valence occupied states (arbitrary units)
of the three C20 Isomers. Solid and dotted lines indicate the results
for neutral and negatively charged clusters, respectively. The
bottom of the valence bands of the neutral and charged clusters has
been arbitrarily aligned for each isomer [Gal98].
The features at binding energies higher than 10 eV distinguish the
three isomers. Large intervals between the peaks characterises the
DOS of the cage isomers. Conversely, for electron binding energy
lower than 10 eV, the electronic structure does not allow to
distinguish the three isomers from each other. The peaks are just of
different intensities. Therefore the theory predicts difficulties in the
assignation of the structures by performing PES with conventional
either visible or near-ultraviolet excitation sources.
In the particular case of the cage structure, the energy
thresholds of neutral and anion is the same. Therefore, Galli’s
calculation predicts no band gap between HOMO and LUMO of
C20 fullerene, which is in contrast with the observed behaviour of
larger fullerenes.
The investigation of allowed vibrational modes is more promising for the identification of the
atomic structures. In experiments involving transitions between electronic states, oscillation
frequencies peculiar of each structure can be excited which are related to the nature and
symmetry of the involved molecular orbitals.
The vibrational frequencies of C20 cage, bowl and ring have been investigated by Seifert
[Sei99]. His molecular dynamic calculations are relative to the electron emission process from the
anion to the neutral C20 and, in particular, the results predict the vibrational modes excited during
the emission process for samples at 0 K.
In a private communication, Seifert refereed us the results of the calculated vibrational spectra:
• The spectrum of the ring isomer shows exclusively a vibrational mode at 2000 cm-1. It is
attributed to the stretching of the acetylenic bond.
• The spectrum of the bowl is more complex than the ring. Several vibrational modes have
been predicted. However, the most intensive peak in the spectra corresponds to 2200
cm-1. The relative vibrational mode is attributed to an acetylenic stretching as well.
• No high-energy vibrational modes have been predicted for the cage isomer. This result is
expected since no triple bond are present in this structure. The spectrum relative to the
cage shows a large number of peaks. The most intensive ones appear at 700 cm-1 and
500 cm-1 wave numbers. In this case the attribution is not simple since the relative
vibrations can be due to different kinds of stretching and bending of carbon bonds.
2.1.2 Chemical synthesis to generate structural precursors
of the smallest fullerene: C20
To investigate the existence of the smallest possible fullerene, consisting of 20 carbon atoms
posed at the vertices of 12 pentagons forming the surface of the cage structure, a dodecahedral
parent is synthesised in Prof. H. Prinzbach’s laboratory [Pri94]. The research project is aimed to
verify the existence of the dodecahedral fullerene by studying the fragmentation pattern of the
structural parent, as well as the spectroscopic properties of the C20 fragment. The parent is
dodecahedrane with molecular formula C20H20. It was firstly synthesised by Paquette and co-
workers [Paq83]. Prinzbach has started out the synthesis from the former insecticide C12-isodrin
and followed the multi-step sequence ultimately based on C5 (cyclopentadiene) and C2
(acetylene) building blocks. The synthesis has been highly optimised and is known as “isodrin-
pagodane” route [Pri94].
The highly symmetric structure of the dodecahedrane has been unequivocally established by
1H NMR and X-ray spectroscopies [Ber97, Paq83]. The carbon atoms are all tetra-coordinated
in a slightly distorted tetrahedron geometry, the cage structure is stabilised by the presence of the
outer-cage carbon-hydrogen bonds and therefore the strain energy is minimised. However, the
dodecahedrane has been proved to be a very stable molecule. In order to make the
transformation of the precursor into the fullerene possible, the strongly bound hydrogen had to be
replaced by more weakly bound atoms. The research group of Prof. Prinzbach has produced two
halogen-functionalised dodecahedrane: A chlorinated and a brominated derivative. As is
extensively discussed in the next paragraph, mass spectroscopic results have induced to opt for
the bromide [Ber97, Sch98]. For an extensive treatment of the bromination process see [Lan00].
An oxygen-insensitive product of average [C20HBr13] elemental composition was obtained
and was spectroscopically and chemically identified as a multitude of primarily isomeric C20H0-
3Br14-11 trienes. Catalytic reduction back to dodecahedrane proved the cage skeleton to have
survived the bromination [Ber97].
In order to exclude the isomerisation of the cage into the bowl structure during the
debromination, the latter generated independently was pursued. This was likewise achieved by
gas phase debromination of a brominated precursor with average [C20HBr9] elemental
composition. The bowl precursor C20H10, better known as corannulene, was synthesised and
brominated by Prof. Scott and co-workers [Sco97].
3 STATE OF THE ART: SILICON
Silicon is currently the preferred material for electronic devices. Due to its superior material
and processing properties it is the most useful and exploited material in modern technology.
However, silicon is an extremely inefficient light emitter and for this reason its optical applications
are few. This optical behaviour can be microscopically explained: Silicon is a semiconductor with
an indirect band gap in the electronic structure due to the symmetry of its crystal lattice. The
indirect band gap energy is Eg = 1.17 eV. The presence of this indirect gap makes the optical
properties of silicon different from those of direct gap semiconductors. In fact, each electron
transition between the bottom of the conduction band and the top of the valence band in direct
gap semiconductors requires an electron-photon interaction. Instead, in an indirect gap
semiconductor the interaction of at least three particles is necessary to conserve the momentum: a
phonon, a photon and an electron. Therefore, radiative recombination (i.e., luminescence) is a
very improbable process and the time scale for the photon emission (~ 10 µs) is slow if
compared with that of direct gap semiconductors. Other relaxation processes requiring no photon
emission are therefore faster and more probable than luminiscence. For this reason, silicon is not
used in the generation of optical signals.
Due to the reduced dimension, it is expected that silicon nanocrystals have completely
different optical and electronic properties. The luminescence should become more allowed in
small crystallites as translational symmetry is broken.
One experimental evidence is the measured visible fluorescence of porous silicon [Ram91].
The energy of the emitted radiation is higher than the energy gap of bulk silicon. This effect is due
to the presence of fused nanometer-diameter silicon wires or particles in the sample. This
enhancement of the energy of the emitted photon in small systems can be explained as follows: the
electron and hole acquire in small crystallites (quantum crystallites) a significant “quantum energy”
of localisation which increases the recombination energy. The effect known as quantum
confinement is extensively discussed in the literature. For a deeper comprehension of this effect,
we suggest to read the explanation of Brus in [Bru94] and the review of Yoffe on the behaviour
of semiconductor microcrystallites [Yof93].
Surface effects can affect as well the electronic structure of small silicon clusters since
covalent bonds favour the reconstruction on the surface of the particles.
The luminescence of silicon nanocrystals is a topical research field. A lot of studies, both
theoretical and experimental, have tried to explain it [Ali96, Bru84, Xia99, Bru91, Wil93,
Wol99]. For example, the results of tight-binding cluster model calculations qualitatively unify the
quantum confinement model and the surface-localised-state model for the luminescence
mechanism [Xia99]. In accordance with the quantum confinement effect, Xia’s calculations
predict the increase of the energy gaps for decreasing diameter. Furthermore, the results of the
calculations indicate that the energy gap becomes smaller in case of incomplete saturation of
surface bonds.
A systematic experimental investigation of the electronic structure of silicon clusters is missing.
The main goal of this research project is to investigate the ground and excited electronic structure
of silicon clusters. We have started with very small particles, but following experiments will be
performed in the mesoscopic range (quantum dots). The question we hope to answer in the
present work is whether the properties of small silicon clusters are extensively determined by
quantum size effects or by the formation of covalent bonds on the surface.
3.1 Ground electronic states
Cheshnovsky et al. [Che87] measured anion photoelectron spectra of Sin- ( 13n ≤ ) clusters,
yielding electron affinities and a qualitative picture of the electronic states of the neutral clusters.
The spectra are shown in the right side panel of figure 3.1. The experimental resolution was
insufficient to observe any vibrational structure, and no assignments of the electronic states were
attempted. Neumark et al. [Kit90, Kit91] later measured vibrationally resolved anion
photoelectron spectra of Sin
-
in the range 3 < n < 7 and zero electron kinetic energy (ZEKE)
spectra. These spectra show well-resolved vibrational progressions for several electronic states,
yielding vibrational frequencies and electronic term values for some of the low-lying electronic
states.
Chelikowsky et al. [Bin95] calculated the electron binding energies for the highest occupied
orbitals of small silicon clusters whose geometries are shown in the left panel of figure 3.1. They
investigated Sin
-
, 7n4 ≤≤ , at finite temperature (500 K and 1500 K), via ab initio Langevin
molecular dynamics simulations. The results of their calculation are shown in the left side panel of
figure 3.1. Furthermore, they compared the calculated Density of States (DOS) with
photoelectron spectra on cluster anions. The comparison demonstrates that the atomic
environment in small silicon clusters and the coexistence of more stable isomers have considerable
effects on the distribution of the electron binding energies.
Figure 3.1: Calculated density of states for the Sin- n = 4 to 7 clusters as obtained from
Langevin simulations at T = 500 K (left side panel) and T = 1500 K (central panel). The
spectrum indicated by the dashed curve and shaded area corresponds to the distorted
octahedron isomer of Si6-. The experimental photoelectron spectra are shown on the right
side panel. The calculated spectra for Si6 have been reported in the right side for
comparison with experiment. Data from [Bin95, Che87].
Experimental results on the electronic structure of bigger silicon clusters are missing.
Theoretical predictions on the development of the band gap were formulated.
The electronic properties of semiconductor particles with perfect crystal structure (quantum
dots), were studied by Brus [Bru83, Bru84, Wil93, Bru94]. The predictions are the following:
Due to a pure quantum confinement effect, the splitting of the valence band and conduction band
becomes larger and the electron affinity decreases for decreasing cluster size.
For studying metal clusters the simple classical model of a spherical metal droplet is used
extensively. This simple model predicts the decrease of the electron affinity for decreasing cluster
size. This behaviour was observed for copper [Pet88, Che90] and silver clusters [Gan88] which
are elements with s electrons in the valence shell. However, open shell transition metal clusters
with valence d electrons show electron affinity values lower than those predicted by the model of
the metal droplet.
Since silicon has more delocalised p valence electrons, it can be expected that this model
does not predict the electron affinity values of silicon clusters.
Recently, Delerue et al. [Wol99] have investigated the electronic state in silicon nanocrystals
as a function of cluster size and surface passivation. The results of their calculation are plotted in
figure 3.2. There, the energy of the electronic states is as a function of the diameter of the
particles. The figure shows a behaviour of the band gap as predicted on the basis of the quantum
confinement effect.
The effect of passivation due to oxygen bound to the cluster surface is shown as well. The
related electronic states are inside the band gap.
Figure 3.2 Dependence of the energy of the electronic states of silicon particles on the
diameter of the clusters. Electronic states of the bottom of the conduction band (o) and of
the top of the valence band () are shown. The effect of passivation due to oxygen bonding
is shown as well. The symbol H is relative to electrons trapped in a p-state localised on the
Si atom of the Si=O bond. The trapped hole states � are p-states localised on the oxygen
atom [Wol99].
Recent calculations of tight-binding type have been performed by Xia et al. for understanding
the size effect and the effect of surface-bond saturation [Xia99]. Once more the prediction due to
the quantum confinement turned out to be correct. In addition, they have found that the energy
gap becomes smaller in the case of incomplete saturation of surface bonds.
3.2 Motivation for the experiments
Small silicon clusters display a rich variety of structures. These structures are neither open
(chains, rings, and fullerene), like those of small carbon clusters, nor fragments of the most stable
bulk (diamond) silicon phase, like those of metal clusters. Instead, the most stable Sin are usually
compact, but they are not always quasispherical and/or high-symmetric structures.
Related to these peculiar symmetries and to the reduced dimension, the electronic properties
of silicon clusters are different to those of the bulk, as demonstrated by the fluorescence of
porous silicon.
The modifications of the electronic structure due to the surface reconstruction and quantum
confinement have been theoretically predicted.
Experimental investigations are thus necessary to test the theories.
CONCLUSIONS OF OUR EXPERIMENTS
The present research work has investigated the electronic properties of size-selected silicon
and carbon clusters.
The experiments demonstrate that electronic structures of small clusters of these two elements
cannot be understood taking the bulk as the reference and explained by the band model theory.
Basically, the strong type of atomic bond – covalent bond – is assumed to determine the unique,
molecule-like structures of these small clusters in order to minimise the number of dangling bonds.
The reached results on carbon and silicon clusters can be summarised as follows:
• For the first time, the existence of the fullerene at 20 carbon atoms has been
demonstrated by comparing PE spectra of C20- isomers produced in different ways. It is
not easy to achieve this fullerene. However, an opportune precursor can be synthesised
whose fragmentation does not destroy the carbon cage structure. We suggest that the
electronic properties of this fullerene are different than those of larger fullerenes. The
existence of the bowl isomer has been proved as well. The vibrational frequencies and the
electron affinities of these two isomers have been measured and compared with those of
the ring isomer. The experimental results are in line with theoretical predictions. The
lifetime of these C20- carbon isomers has been estimated to be longer than 0.4 ms.
• For negatively charged silicon clusters in the size range 60n3 ≤≤ , the electronic
structure does not show the band gap predicted by quantum confinement. We explain this
as due to the occupation of surface states. The existence of excited states of different
nature has been demonstrated. Their lifetimes have been found to vary in the range from
few hundreds of fs to ns. The involved relaxation processes are supposed of at least three
types.
Furthermore, a new experimental method to investigate photofragmentation has been
proposed. For the first time, pump-probe-type photoelectron spectroscopy has been used to
study the fragmentation pattern. This technique allows one to make a time-resolved measurement
of the fragmentation process. The achieved information are of static and dynamic type -
fragmentation behaviour and velocity of the process.