12.02.2013 Views

Book of Abstracts Book of Abstracts - Universität Konstanz

Book of Abstracts Book of Abstracts - Universität Konstanz

Book of Abstracts Book of Abstracts - Universität Konstanz

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Symposium on<br />

Size Selected Clusters<br />

<strong>Book</strong> <strong>of</strong><br />

<strong>Abstracts</strong><br />

28 February - 3 March 2005 in Brand / Austria<br />

http://www.s3c.uni-konstanz.de/


The S 3 C –<br />

Symposium on Size Selected<br />

Clusters<br />

was generously supported by:<br />

• German Science Foundation<br />

• Varian


Table <strong>of</strong> Contents<br />

FULL LENGTH TALKS AND HOT TOPICS......................................................................... 1<br />

POSTERS.................................................................................................................................. 33<br />

CARBON 35<br />

CHEMISTRY 43<br />

DYNAMICS 63<br />

MAGNETISM 77<br />

MATERIALS 83<br />

METHODS 91<br />

METALS 101<br />

MOLECULAR CLUSTERS 113<br />

PHASE TRANSITIONS 121<br />

RARE GASES 127<br />

SEMICONDUCTORS 135<br />

SOLVATIONS 141<br />

SURFACE 147<br />

AUTHOR INDEX.................................................................................................................... 167<br />

� Poster Session A (Monday evening): Metals, Molecular Clusters, Phase<br />

Transitions, Rare Gases, Semiconductors, Solvations, Surface<br />

� Poster Session B (Wednesday evening): Carbon, Chemistry, Dynamics,<br />

Magnetism, Materials, Methods<br />

� The poster numbers are given at the top <strong>of</strong> the abstracts which can be found in<br />

the author index at the end <strong>of</strong> this book.


Full Length Talks<br />

and Hot Topics<br />

1


Metallic clusters and oxygen<br />

Catherine Bréchignac<br />

Laboratoire Aimé Cotton, CNRS, Bâtiment 505, Université Paris Sud, 91405 Orsay cedex, France<br />

The interaction between metallic substance and oxygen is one <strong>of</strong> the most important<br />

chemical and biological process in nature. This talk will discuss two aspects <strong>of</strong> oxygen<br />

interaction with matter at nanometer-scale.<br />

The first part <strong>of</strong> the talk deals with the interaction between metallic clusters and oxygen. Here<br />

cluster acts as finite reservoir <strong>of</strong> electrons and it is shown how the reactivity strongly depends<br />

on cluster size.<br />

In a second part the cluster acts as a vehicle to bring oxide molecule locally on supported<br />

islands. Using fractal islands as a test case for non-equilibrium morphologies, we show evidence<br />

<strong>of</strong> oxide driven elongated pearled shapes similar to those observed in pearling instability <strong>of</strong><br />

membrane tubes with anchored polymers. Such a similarity indicates that the relaxation <strong>of</strong><br />

fractals shares the same universality as the shape transition in neurons and membrane tubes.<br />

3


4<br />

The Disapperarance <strong>of</strong> Nobility: Reactivity Studies on free Gold and<br />

Silver Clusters<br />

Ludger Wöste<br />

Freie <strong>Universität</strong> Berlin, Institut für Experimentalphysik, Arnimallee 14, D-14195 Berlin, Germany<br />

Contrary to the bulk metal the clusters <strong>of</strong> noble metals are very reactive. The process <strong>of</strong><br />

gold sensitisation in silver photography is a prominent example. Most exciting in this regard is<br />

the non-scalable size regime below about 100 atoms per particle [1], where the physical<br />

properties <strong>of</strong> noble metal clusters show very pronounced size and charge effects. Their<br />

comprehension requires a clear identification <strong>of</strong> their geometries and electronic structures. We<br />

investigated chemical reactions <strong>of</strong> mass-selected gold and silver clusters in a temperature<br />

controlled rf-ion trap setup. Product ion concentrations as a function <strong>of</strong> storage time enable the<br />

determination <strong>of</strong> reaction kinetics and reaction mechanisms on free clusters. As an example, the<br />

CO combustion reaction on small gold clusters is studied. It is known that supported gold<br />

clusters with few atoms up to nm size exhibit relevant catalytic properties. Our ion trap<br />

measurements reveal catalytic activity already for negatively charged gold dimers. When the<br />

reaction kinetics are investigated as a function <strong>of</strong> temperature, intermediate products with CO<br />

and O2 co-adsorbed can be isolated [2]. In contrast, small atomic silver clusters have not been<br />

found relevant for catalytic oxidation processes so far. Our investigations show for the first time<br />

evidence for a strongly size dependent catalytic activity <strong>of</strong> Agn- (n=1-13) [3]. In order to<br />

elucidate the details <strong>of</strong> their reaction dynamics, femtosecond pump-probe experiments were<br />

performed on the educts, products and intermediates.<br />

References<br />

[1] U. Landman, Int. J. Mod. Phys. B 6, 3623 (1992).<br />

[2] L. D. Socaciu, J. Hagen, T. M. Bernhardt, L. Wöste, U. Heiz, H. Häkkinen, U. Landman:<br />

“Catalytic CO oxidation by free Au2-: Experiment and theory”, J. Am. Chem. Soc. 125, 10437<br />

(2003).<br />

[3] L. D. Socaciu, J. Hagen, J. Le Roux, D. Popolan, T. M. Bernhardt, L. Wöste,<br />

S. Vajda: “Strongly cluster size dependent reaction behavior <strong>of</strong> CO with O2 on free silver cluster<br />

anions”, J. Chem. Phys. 120, 2078 (2004).


Infrared Spectroscopy <strong>of</strong> Metal Ion Complexes: Models for Metal<br />

Ligand Interactions and Solvation<br />

Michael A. Duncan<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Georgia, Athens, GA 30602<br />

maduncan@uga.edu<br />

http://www.arches.uga.edu/~maduncan<br />

Weakly bound complexes <strong>of</strong> the form M + -Lx (M=Fe, Ni, Co, etc.; L=CO2, C2H2, H2O,<br />

benzene, N2) are prepared in supersonic molecular beams by laser vaporization in a pulsednozzle<br />

cluster source. These species are mass analyzed and size-selected in a reflectron time-<strong>of</strong>flight<br />

mass spectrometer. Clusters are photodissociated at infrared wavelengths with a Nd:YAG<br />

pumped infrared optical parametric oscillator/amplifier (OPO/OPA) laser or with a tunable<br />

infrared free-electron laser. M + -(CO2)x complexes absorb near the free CO2 asymmetric stretch<br />

near 2349 cm -1 but with an interesting size dependent variation in the resonances. Small clusters<br />

have blue-shifted resonances, while larger complexes have additional bands due to surface CO2<br />

molecules not attached to the metal. M + (C2H2)n complexes absorb near the C-H stretches in<br />

acetylene, but resonances in metal complexes are red-shifted with repect to the isolated<br />

molecule. Ni + and Co + complexes with acetylene undergo intracluster cyclization reactions to<br />

form cyclobutadiene. Transition metal water complexes are studied in the O-H stretch region,<br />

and partial rotational structure can be measured. M + (benzene) and M + (benzene)2 ions (M=V, Ti,<br />

Al) represent half-sandwich and sandwich species, whose spectra are measured near the free<br />

benzene modes. These new IR spectra and their assignments will be discussed as well as other<br />

new IR spectra for similar complexes.<br />

5


6<br />

Diffuse Electron States<br />

Kit H. Bowen, Jr.<br />

Department <strong>of</strong> Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, USA<br />

The extra electron in valence (conventional) anions resides in an orbital which is firmly<br />

grounded on the nuclear framework <strong>of</strong> the system, i.e., the extra electron is localized. Likewise,<br />

when such anions are solvated and become anion-molecule complexes, the extra electron<br />

usually remains localized with the anion. During photodetachment <strong>of</strong> the excess electron from<br />

such a cluster anion, this sub-anion acts as the chromophore, and the photoelectron spectrum<br />

reflects the characteristics <strong>of</strong> that anion, albeit stabilized and spectrally shifted to higher electron<br />

binding energy.<br />

In other cluster anions, however, the excess electrons are more delocalized, even though<br />

they remain closely associated with their nuclear frameworks. There is, however, still another<br />

class <strong>of</strong> molecular and cluster anions in which the excess electrons are neither localized nor<br />

closely associated with their nuclear frameworks, and these are anions with highly diffuse,<br />

excess electron distributions. Examples include dipole bound anions, quadrupole bound anions,<br />

double Rydberg anions, and some sizes <strong>of</strong> water cluster anions. While these species share the<br />

characteristic <strong>of</strong> having spatially diffuse excess electrons, they are otherwise bound by different,<br />

albeit idealized interactions. The excess electron in dipole bound anions is chiefly bound by the<br />

dipolar field <strong>of</strong> the corresponding neutral system, with the excess electron ballooning out into<br />

space at the positive end <strong>of</strong> the molecular or cluster system. Similarly, the excess electron in<br />

quadrupole bound anions is primarily bound by the quadrupolar field <strong>of</strong> the corresponding<br />

neutral system. Double Rydberg anions are the negative ions <strong>of</strong> Rydberg radicals. Each one can<br />

be thought <strong>of</strong> as a closed shell, cationic core which is enveloped by two Rydberg-like, highly<br />

diffuse electrons. Also, many water cluster anion systems appear to exhibit diffuse excess<br />

electron states <strong>of</strong> at least two types, with a third type being associated with internalizing<br />

electron states.<br />

In this talk, we will survey these different types <strong>of</strong> diffuse electron states, but we will focus<br />

on double Rydberg anions, in part because they are the least well-known diffuse, excess electron<br />

species. In particular, we will discuss the ammonia-based double Rydberg anions, (NnH3n+1) - ,<br />

where n = 1- 7. Using anion photoelectron spectroscopy, we have measured electron affinities<br />

<strong>of</strong> these species, found examples <strong>of</strong> solvated double Rydberg anions, and accessed the first<br />

electronic states <strong>of</strong> several neutral Rydberg radicals by photodetaching electrons from their<br />

corresponding double Rydberg anions. Insight into the nature <strong>of</strong> these species was further aided<br />

by comparisons with the absorption measurements <strong>of</strong> Fuke on neutral Rydberg cluster-like<br />

radicals and by several theoretical studies, especially those <strong>of</strong> Ortiz.


Model Gold Catalysts by Size-Selected Cluster Deposition<br />

Sungsik Lee, Chaoyang Fan, Tianpin Wu, and Scott L. Anderson<br />

Dept. <strong>of</strong> Chemistry, University <strong>of</strong> Utah, 315 S. 1400 E. Rm 2020, Salt Lake City UT 84112, USA<br />

Planar model gold catalysts were prepared by deposition <strong>of</strong> size- and energy-selected gold<br />

cluster cations, on well characterized supports in UHV. The CO oxidation reaction was used to<br />

characterize the chemical behavior <strong>of</strong> the resulting samples. As shown in the figure, strongly<br />

size-dependent activity is observed for room temperature CO oxidation on Aun/TiO2, with<br />

significant activity observed for n ≥ 3. No activity is observed for Au/Al2O3. Detailed studies<br />

over a wide temperature range, using x-ray photoelectron spectroscopy (XPS), ion scattering<br />

spectroscopy (ISS), and CO thermal adsorption and desorption were used to probe the nature <strong>of</strong><br />

the samples resulting from cluster deposition. It is found that on titania, the gold migrates to<br />

surface oxygen vacancy sites between 200 and 300K, and appears to be stably trapped at the<br />

vacancies at room temperature. XPS shows that binding at the vacancies injects electron density<br />

into the clusters, and absence <strong>of</strong> this effect on alumina is probably responsible for the observed<br />

lack <strong>of</strong> activity. At T ≥ 450 K, significant agglomeration occurs. Correlations <strong>of</strong> activity with<br />

the cluster morphology and ability to bind CO and O2, both on the clusters, and in adjacent<br />

substrate sites, will be discussed.<br />

C 16 O 18 O+ intensity (arb.)<br />

7e+5<br />

6e+5<br />

5e+5<br />

4e+5<br />

3e+5<br />

2e+5<br />

1e+5<br />

0<br />

CO oxidation<br />

TiO2 Au1 Au2 Au3 Au4 Au5 Au6 Au7<br />

7


8<br />

Structural Order in Metal Clusters: Dependence on Size and Charge<br />

State<br />

Joel H Parks<br />

Rowland Institute at Harvard, Cambridge, Massachusetts 02142<br />

Trapped ion electron diffraction (TIED) has been applied to Ag and Au cluster ions<br />

produced by a sputter discharge aggregation source. A specific cluster size is selectively stored<br />

in an rf ion trap in which the trapped ions are annealed and brought into thermal equilibrium<br />

with a background He gas. The stored ions are then exposed to a high energy electron beam and<br />

electron scattering is imaged by a multichannel plate detector onto a phosphor screen.<br />

Diffraction patterns are collected by a CCD camera and analyzed to extract the molecular<br />

interference directly related to the arrangement <strong>of</strong> atoms in the mass selected cluster.<br />

Following a brief introduction to TIED experiment and analysis, this talk will present recent<br />

measurements for Agn + cations, and Aun -/+ cation and anion clusters. Two principal results will<br />

be discussed: the transition from local to global five-point order in Agn + clusters over the size<br />

range 36≤n≤55, and the dependence <strong>of</strong> Aun -/+ structures on charge state for n=20,32,38 and 55.<br />

Diffraction data will be compared with isomer structures predicted by density functional<br />

calculations and molecular dynamic simulations. Experimental opportunities for diffraction<br />

measurements will be briefly considered.


Ultra-stable Hetero-nuclear Microclusters <strong>of</strong> (CdSe)33 and (CdSe)34<br />

Atsuo Kasuya 1 , Rajaratnam Sivamohan 1 , Yurii A. Barnakov 1 , Igor M. Dmitruk 1,6 , Takashi Nirasawa 2,3 ,<br />

Grzegorz Milczarek 1 , Sergiy V. Mamykin 1 , Volodymyr R. Romanyuk 1 , Kazuyuki Tohji 2 , Valachandran<br />

Jeyadevan 2 , Kozo Shinoda 2 , Toshiji Kudo 3 , Osamu Terasaki 4 , Zheng Liu 4 , Tetsu Ohsuna 5 , Rodion V.<br />

Belosludov 5 , Vijay Kumar 1,5 , Vijayaraghavan Sundararajan 1 , Yoshiyuki Kawazoe 5<br />

1 Center for Interdisciplinary Research, Tohoku University, Sendai, 980-8578, Japan<br />

2 Department <strong>of</strong> Geoscience and Technology, Tohoku University, Sendai, 980-8579, Japan<br />

3 Bruker Daltonics K.K., Kanagawa-ku, Yokohama, 221-0022, Japan<br />

4 Department <strong>of</strong> Physics, Tohoku University, Sendai, 980-8578, Japan<br />

5 Institute for Material Research, Tohoku University, Sendai, 980-8577, Japan<br />

6 Kyiv National Taras Shevchenko University, Kyiv, 03022, Ukraine<br />

Time-<strong>of</strong>-Flight mass spectra reveal that nanoparticles, (CdSe)n, grown in solution can be mass<br />

selected at atomic level by reverse micelle method. This method allows us to produce<br />

macroscopic quantity <strong>of</strong> (CdSe)33 and (CdSe)34 by making use <strong>of</strong> their highly selective<br />

stabilities in reverse micelles [1]. This preferential growth at particular n’s depicted by mass<br />

spectra suggests that the structure <strong>of</strong> these particles is quite different from a bulk fragment and<br />

is high-symmetric (CdSe)28 cage incorporating 5 or 6 CdSe pairs inside and forming sp 3<br />

network. Similar selective stabilities have been found in CdS and other A II B VI compounds,<br />

demonstrating that such highly stabilied clusters may be found in a variety <strong>of</strong> compound<br />

systems.<br />

Figure 1. Time-<strong>of</strong>-flight mass spectra <strong>of</strong> positive ions from (CdSe) n prepared in toluene (curve 1), and from powders<br />

<strong>of</strong> bulk CdSe (curve 2) and CdS (curve 3). Suggested structure <strong>of</strong> (CdSe) 34 cluster.<br />

References<br />

[1] Kasuya A., Sivamohan R., Barnakov Yu., Dmitruk I., et al, Nature Materials 3, 99 (2004).<br />

9


Dynamics and spectroscopy <strong>of</strong> anion clusters and helium nanodroplets<br />

10<br />

Daniel M. Neumark<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> California, Berkeley, CA. USA 94720, and Chemical Sciences<br />

Division, Lawrence Berkeley National Laboratory, Berkeley, CA USA 94720<br />

Recent results on the spectroscopy and dynamics <strong>of</strong> water cluster anions and helium<br />

nanodroplets will be presented. Mass-selected (H2O)n - and (D2O)n - anions were studied using<br />

one-photon and time-resolved photoelectron imaging. The one-photon studies established the<br />

existence <strong>of</strong> both internal and surface states <strong>of</strong> water clusters anions, and showed that one can<br />

switch between isomers by varying ion source conditions. The time-resolved studies yielded<br />

excited state lifetimes for both isomers, showing dramatically different dependence on cluster<br />

size. Extrapolation <strong>of</strong> the internal conversion rates for the internal states to the bulk limit<br />

strongly supports the “non-adiabatic” relaxation mechanism proposed for the hydrated electron<br />

in aqueous solution.<br />

The photoionization and photoelectron spectroscopy <strong>of</strong> He nanodroplets was investigated using<br />

synchrotron radiation at the Advanced Light Source at Lawrence Berkeley National Laboratory.<br />

Photoelectron images were obtained at photon energies below and above the ionization potential<br />

<strong>of</strong> atomic He (24.6 eV). The images below the IP were dominated by extremely slow electrons,<br />

with an average energy <strong>of</strong> < 1 meV, believed to originate from excitation and autoionization <strong>of</strong><br />

a Hen subunit within the cluster followed by thermalization <strong>of</strong> the electron before it leaves the<br />

cluster. The images above the IP are very different, showing photoelectron signal at higher<br />

energy than that seen from atomic He. This high energy signal is attributed to direct detachment<br />

to the attractive region <strong>of</strong> Hen+ potential energy surfaces.


Tailoring functionality <strong>of</strong> clusters by size, structures and lasers<br />

Vlasta Bonačić-Koutecký<br />

Humboldt <strong>Universität</strong> zu Berlin, Institut für Chemie, Brook-Taylor-Strasse 2,<br />

D-12489 Berlin, Germany, e-mail: vbk@chemie.hu-berlin.de<br />

Theoretical exploration <strong>of</strong> cluster properties for which each atom counts will be presented<br />

for the following topics:<br />

1. Chemical reactivity <strong>of</strong> nobel-metal oxides as foundation for design <strong>of</strong> nanoscale<br />

catalytic materials.<br />

2. Emissive properties <strong>of</strong> silver particles at silver oxide defects as attractive candidates for<br />

optical data storage.<br />

3. Ultrafast dynamics and its control by tailored laser fields in clusters and in their<br />

complexes with biomolecules for selection <strong>of</strong> reaction channels and <strong>of</strong> photochemical<br />

processes.<br />

4. Importance <strong>of</strong> mutual interaction between theory and experiment for all three subjects<br />

will be illustrated.<br />

11


A new infrared spectroscopy technique for structural studies <strong>of</strong> massselected<br />

neutral polar molecular species without chromophore<br />

12<br />

B. Lucas, F. Lecomte, G. Grégoire, Y. Bouteiller, J.-P. Schermann, C. Desfrançois<br />

Lab. de Physique des Lasers, CNRS-Univ. Paris-Nord, 93430 Villetaneuse, France<br />

We will present a new infrared spectroscopy technique for structural studies <strong>of</strong> massselected<br />

neutral polar molecules or complexes without chromophore [1]. In these experiments,<br />

we use dipole-bound anion formation, as a totally non-perturbative ionization method. Anion<br />

signal depletion allows us to monitor IR absorption, in between 2700 and 3800 cm -1 , on massselected<br />

neutral species. For weakly bound complexes, with typical H-bond dissociation<br />

energies about 1500 cm -1 , IR absorption indeed leads to fast vibrational predissociation prior to<br />

ionization. For both isolated molecules and complexes, the very weak excess electron binding<br />

energy (typically 100 cm -1 ) in dipole-bound anions also allows for vibrational autodetachment<br />

due to IR absorption. In both cases, the signature <strong>of</strong> IR absorption by the neutral species is a<br />

loss in the dipole-bound anion signal at the corresponding mass. This is an alternative technique<br />

to the IR-REMPI technique when the neutral species do not possess any chromophore.<br />

We will discuss results on small molecules and clusters containing water and formamide:<br />

water dimer, formamide-water complex, isolated formamide and formamide dimer. These four<br />

test-cases exemplify the capabilities and limitations <strong>of</strong> this technique and allow us to plan for<br />

future technical improvements in order to gain sensitivity for further experiments on molecules<br />

and complexes <strong>of</strong> biological interest. Experimental IR spectra will be interpreted with the help<br />

<strong>of</strong> quantum chemistry calculations. In particular, the use <strong>of</strong> anharmonic frequency calculations<br />

will be discussed.<br />

relative anion signal depletion<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

free CH<br />

2850 2900 2950<br />

bonded<br />

NH / OH<br />

free NH<br />

?<br />

3400 3450 3500 3550<br />

frequency (cm -1 )<br />

free OH<br />

3700 3750<br />

IR spectrum <strong>of</strong> mass-selected neutral formamide-water complexes obtained from dipole-bound anion signal<br />

depletion. Small squares correspond to relative depletion values averaged over 3 independent measurements while<br />

full thick lines correspond to a smooth over 3 adjacent points. Experimental line positions are indicated by thin dash<br />

lines and theoretical frequencies are located by full (anharmonic values) or dash (scaled harmonic values) bars. The<br />

question mark corresponds to a line that is not assigned.<br />

References<br />

[1] B. Lucas, F. Lecomte, B. Reimann, H.-D. Barth, Gilles Grégoire, Y. Bouteiller, J.-P. Schermann,<br />

C. Desfrançois, Phys. Chem. Chem. Phys. 6, 2600 (2004).


The hydrogen transfer reaction: from molecular clusters to proteins<br />

C. Jouvet 1 , H. Kang 1 , C. Dedonder-Lardeux 1 , G. Grégoire 2 , C. Desfrançois 2 , J-P. Schermann 2 , M. Barat 3<br />

and J. A. Fayeton 3<br />

1 Laboratoire de Photophysique Moléculaire du CNRS et Laboratoire ELYSE, Bat349, Université Paris-<br />

Sud, 91405 Orsay, France.<br />

2 Laboratoire de Physique des Lasers du CNRS, Institut Galilée, Université Paris-Nord, 93430<br />

Villetaneuse, France.<br />

3 Laboratoire des Collisions Atomiques et Moléculaires du CNRS, Université Paris-Sud, 91405 Orsay,<br />

France.<br />

Four years ago we have demonstrated that the excited state reaction <strong>of</strong> phenol-amonnia<br />

clusters is an H atom transfer leading to the formation <strong>of</strong> hypervalent NH4(NH3)n clusters 1, 2 , and<br />

not by a proton transfer as suggested by the liquid phase experiments. This first observation lead<br />

to the developement <strong>of</strong> a more general model 3 where a surprisingly simple and general<br />

mechanistic picture for the nonradiative decay <strong>of</strong> aromatic biomolecules such as nucleic bases<br />

and aromatic amino acids has been suggested.<br />

The key role in this model is played by an excited singlet state <strong>of</strong> πσ* character, which has<br />

a repulsive potential-energy function with respect to the stretching <strong>of</strong> OH or NH bonds. The<br />

1 πσ∗ potential-energy function intersects not only the bound potential-energy functions <strong>of</strong> the<br />

1 ππ∗ excited states, but also that <strong>of</strong> the electronic ground state 3 . Via predissociation <strong>of</strong> the 1 ππ∗<br />

states and a conical intersection with the ground state, the � πσ* state triggers an internalconversion<br />

(IC) process. The lifetime <strong>of</strong> the optically excited � ππ∗ state is governed by the first<br />

intersection which determines the barrier for IC process, and varies from one molecule to the<br />

other depending largely on the energy gap between the 1 πσ∗ and the 1 ππ* states.<br />

In clusters <strong>of</strong> phenol or indole 4 and ammonia, calculations predict that the H transfer leads<br />

to the formation <strong>of</strong> hydrogenated ammonia clusters NH4(NH3)n through an avoided crossing<br />

between the 1 ππ∗ and the 1 πσ∗ states 5 . In this system the reaction seems to proceed via<br />

tunneling, since the process is fast in the hydrogenated species and considerably slower for<br />

deuterated species 6-7 . We recently demonstrated that this model also applies in protonated<br />

aromatic amino acid (tryptophan, tyrosine) and explains why the excited state lifetime <strong>of</strong><br />

protonated Tryptophan (400fs) 8 is ten times shorter than the lifetime <strong>of</strong> protonated Tyrosine.<br />

References<br />

[1] G. Pino, G. Gregoire, C. Dedonder-Lardeux, C. Jouvet, S. Martrenchard and D. Solgadi, Physical<br />

Chemistry Chemical Physics, 2000, 2, 893-900.<br />

[2] G. A. Pino, C. Dedonder-Lardeux, G. Gregoire, C. Jouvet, S. Martrenchard and D. Solgadi, J.<br />

Chem. Phys., 1999, 111, 10747.<br />

[3] A. L. Sobolewski, W. Domcke, C. Dedonder-Lardeux and C. Jouvet, Phys. Chem. Chem. Phys.,<br />

2002, 4, 1093-1100.<br />

[4] C. Dedonder-Lardeux, D. Grosswasser, C. Jouvet and S. Martrenchard, Physchemcomm, 2001,<br />

1-3.<br />

[5] A. L. Sobolewski and W. Domcke, J. Phys. Chem. A, 2001, 105, 9275.<br />

[6] G. Gregoire, C. Dedonder-Lardeux, C. Jouvet, S. Martrenchard, A. Peremans and D. Solgadi,<br />

Journal <strong>of</strong> Physical Chemistry A, 2000, 104, 9087-9090.<br />

[7] S. Ishiuchi, K. Daigoku, M. Saeki, K. Hashimoto, M. Sakai and M. Fujii, J. Chem. Phys., 2002,<br />

117, 7077.<br />

[8] H. Kang, C. Jouvet, C. Dedonder-Lardeux, S. Martrenchard, G. Grégoire, C. Desfrançois, J.-P.<br />

Schermann, M. Barat and J. A. Fayeton, P.C.C.P., 2004, to be published.<br />

13


14<br />

Immobilisation <strong>of</strong> proteins by size-selected nanoclusters on surfaces<br />

Richard E. Palmer<br />

Nanoscale Physics Research Laboratory, School <strong>of</strong> Physics and Astronomy, The University <strong>of</strong><br />

Birmingham, Birmingham B15 2TT, UK<br />

r.e.palmer@bham.ac.u<br />

www.nprl.bham.ac.uk<br />

The controlled deposition <strong>of</strong> size-selected nanoclusters represents a novel route to the<br />

fabrication <strong>of</strong> nanostructured surfaces, generating lateral features <strong>of</strong> size 1-10 nm. This is<br />

precisely the size scale <strong>of</strong> biological molecules and provides a new method to immobilize<br />

individual proteins, with potential applications in structural biology, protein complex formation<br />

and high throughput medical diagnostics (microarrays).<br />

Scaling relations which describe both the implantation [1] and pinning [2] <strong>of</strong> the clusters<br />

enable the controlled preparation <strong>of</strong> 3D, nanoscale surface features, stable to temperatures as<br />

high as 400°C. In particular, we report the pinning <strong>of</strong> size-selected AuN clusters (N = 1 – 100) to<br />

the hydrophobic graphite surface to create films <strong>of</strong> arbitrary, sub-monolayer density. Gold<br />

presents an attractive binding site for sulphur and thus potentially for the cysteine amino acid<br />

residues in proteins (which contain a thiol, SH, group), suggesting the possibility <strong>of</strong> residuespecific<br />

immobilization <strong>of</strong> oriented protein molecules.<br />

AFM measurements in liquid (buffer) solution show that GroEL chaperonin molecules<br />

(ring-shaped molecules with diameter 15 nm), which contain free cysteine residues, bind to the<br />

clusters and are immobilised [3]. Both peroxidase molecules [4], in which the cysteine residues<br />

pair up to form disulphide bonds, and oncostatin molecules can similarly be immobilised. In<br />

both cases protein clusters are also formed. By contrast, green fluorescent protein (GFP) and<br />

luciferase molecules do not bind to the nanoclusters. The behaviour <strong>of</strong> the different protein<br />

molecules is predicted by a model, based on molecular surface area calculations, which<br />

considers the “accessibility” <strong>of</strong> cysteine (and other) residues at the outer surface <strong>of</strong> the folded<br />

protein. The results demonstrate the ground rules for, and generality <strong>of</strong>, protein immobilisation<br />

by metal cluster films.<br />

Finally, I will discuss ongoing experiments with human IgG molecules which seem to<br />

present the first evidence that protein immobilization depends on cluster size.<br />

References<br />

[1] S. Pratontep, P. Preece, C. Xirouchaki, R.E. Palmer, C.F. Sanz-Navarro, S.D. Kenny and R. Smith,<br />

Phys. Rev. Lett. 90 055503 (2003).<br />

[2] S.J. Carroll, S. Pratontep, M. Streun, R.E. Palmer, S. Hobday and R. Smith, J. Chem. Phys. 113<br />

7723 (2000); also Nature (News & Views) 408 531 (2000).<br />

[3] R.E. Palmer, S. Pratontep and H.-G. Boyen, Nature Materials 2 443 (2003).<br />

[4] C. Leung, C. Xirouchaki, N. Berovic and R.E. Palmer, Adv. Mater. 16 223 (2004).


Atomic and Molecular Architecture at Surfaces<br />

Klaus Kern<br />

Max-Planck-Institut für Festkörperforschung, Heisenbergstr. 1, D-70569 Stuttgart<br />

klaus.kern@fkf.mpg.de<br />

A promising route toward the realization <strong>of</strong> functional nanosystems is the exploitation <strong>of</strong><br />

nature’s inherent drive to generate complexity. The transcription <strong>of</strong> the corresponding<br />

organization principles to artificial compounds and environments comprises intriguing<br />

perspectives. We emphasize the conductance <strong>of</strong> self-organized growth processes at well-defined<br />

surfaces. The atomistic insight gained into the underlying mechanisms and interactions is used<br />

to control the formation <strong>of</strong> low-dimensional atomic and molecular architectures. This knowhow<br />

opens up new avenues in engineering nanomaterials <strong>of</strong> well-defined shape, composition<br />

and functionality to be harnessed for future technological applications.<br />

15


Size-Selected Aun and Agn Clusters on Rutile TiO2 (110) (1x1) Surfaces<br />

Probed by UHV-STM<br />

16<br />

Steven K. Buratto, Xiao Tong, Lauren Benz, Andrei Kolmakov, Paul Kemper, Steeve Chrétien, Horia<br />

Metiu, and Michael T. Bowers<br />

Department <strong>of</strong> Chemistry and Biochemistry, University <strong>of</strong> California, Santa Barbara<br />

Santa Barbara, CA 93106-9510, United States <strong>of</strong> America<br />

Catalysis <strong>of</strong> the oxidation <strong>of</strong> CO and small olefins by nanoclusters <strong>of</strong> Au and Ag on oxide<br />

supports is known to be strongly dependent on the size <strong>of</strong> the cluster and its interaction with the<br />

oxide support. In our group we have probed the size dependence by depositing size-selected<br />

clusters <strong>of</strong> Agn + and Aun + (n = 1-7) from the gas phase onto single crystal rutile TiO2 (110) (1x1)<br />

surfaces at room temperature under s<strong>of</strong>t-landing (< 2 eV/atom) conditions. We analyze the<br />

clusters on the surface using ultra-high vacuum scanning tunneling microscopy (UHV-STM)<br />

and compare the resulting structures with theory. In the case <strong>of</strong> Ag + and Ag2 + clusters deposited<br />

under s<strong>of</strong>t-landing conditions (Figure 1), we observe large, sintered clusters indicating high<br />

mobility for these species on the surface. For Agn + (n ≥ 3) clusters (also shown in Fig. 1)<br />

deposited under s<strong>of</strong>t-landing conditions, however, we observe a high density <strong>of</strong> intact clusters<br />

bound to the surface and no large, sintered clusters indicating that these species have very<br />

limited mobility on the surface.<br />

Figure 1. STM Images <strong>of</strong> the titania surface after the deposition <strong>of</strong> Ag 1, Ag 2, and Ag 3, respectively, from left to right.<br />

All images are 100 nm x 100 nm, insets are ~ 7 nm x 7 nm.<br />

In the case <strong>of</strong> Aun + clusters, we observe large, sintered clusters only from the deposition <strong>of</strong> Au +<br />

and a high density <strong>of</strong> intact clusters from the deposition <strong>of</strong> Aun + (n ≥ 2) as depicted in Figure 2.<br />

In cases where we observe intact clusters we can observe the binding site and geometry in the<br />

STM image and compare these with structures calculated using density functional theory (DFT)<br />

as well as structures observed in the gas phase.<br />

Figure 2. STM Images <strong>of</strong> the titania surface after the deposition <strong>of</strong> Au 1, Au 2, Au 3, Au 4 and Au 7, respectively, from<br />

left to right. All images are 100 nm x 100 nm, insets are ~ 7 nm x 7 nm.


Magnetic Properties <strong>of</strong> Deposited Transition Metal Atoms and<br />

Clusters<br />

Matthias Reif, Leif Glaser, Michael Martins, Wilfried Wurth<br />

Insitut für Experimentalphysik, <strong>Universität</strong> Hamburg, Luruper Chaussee 149,<br />

Hamburg, D-22761, Germany<br />

From a fundamental point <strong>of</strong> view size-selected transition metal clusters allow to study the<br />

transition from the magnetic properties <strong>of</strong> atoms to those <strong>of</strong> the respective solids as a function <strong>of</strong><br />

cluster size. For possible future applications these clusters have to be supported on a substrate <strong>of</strong><br />

embedded in a matrix where cluster-support interaction will influence theier physical properties<br />

as well. We have therefore started to investigate the magnetic properties <strong>of</strong> small transition<br />

metal clusters in contact with ferromagnetic substrates as a function <strong>of</strong> the number n <strong>of</strong> cluster<br />

atoms in the size region from n=2 to n=20.<br />

To achive thos goal we have developed a UHV-compatible cluster source, which enables us to<br />

investigate size selected transition metal clusters deposited in-situ under UHV conditions on<br />

solid surfaces using s<strong>of</strong>t X-ray spectroscopy at third generation synchrotron radiation sources.<br />

This contribution will discuss especailly x-ray magnetic circular dichroism (XMCD) studies <strong>of</strong><br />

size selected deposited clusters and recent experimental results <strong>of</strong> supported 3d and 4d transition<br />

metal clusters will be presented. For iron clusters [1,2] on nickel a strong size dependence <strong>of</strong> the<br />

spin- and orbital moment, which could be measured independently with the XMCD method, are<br />

found. In particular the orbital moment per iron atom is strongly enhanced as compared to the<br />

values <strong>of</strong> bulk solid or ultrathin films.<br />

Recent experimental and theoretical studies on chromium clusters in the size range from n=1 to<br />

n=13 deposited on ultrathin ferromagnetic nickel and iron surfaces demonstrate the importance<br />

<strong>of</strong> the substrate. Whereas for Cr clusters on Ni films no XMCD signal is found, for Cr clusters<br />

on Fe films a strong decrease <strong>of</strong> the magnetic moment by a factor <strong>of</strong> 4 is found with increasing<br />

cluster size. This effect can be interpreted in terms <strong>of</strong> a non-collinear coupling <strong>of</strong> the magnetic<br />

moments within the cluster and relative to the surface [3]. Finally, first results on the magnetic<br />

properties <strong>of</strong> deposited 4d transition metal atoms and dimers (Ru, Mo) will be given .<br />

References<br />

[1] J.T. Lau , A. Föhlisch, R. Nietubyc, M. Reif, and W. Wurth, Phys. Rev. Lettl. 89, 057202 (2002).<br />

[2] J.T. Lau , A. Föhlisch, M. Martins, R. Nietubyc, M. Reif, and W. Wurth, New J. Of Phys. 4, 98.1<br />

(2002)<br />

[3] M. Reif, L. Glaser, M. Martins, W. Wurth, S. Lounis, and S. Blügel, in preparation.<br />

17


Emergent physics and chemistry <strong>of</strong> clusters: Fermion/boson clusters in<br />

quantum dots/traps, and nanocatalysis<br />

18<br />

U. Landman<br />

University School <strong>of</strong> Physics, Georgia Institute <strong>of</strong> Technology, Atlanta, Georgia 30332-0430, USA<br />

We discuss several physical and chemical cluster phenomena that are emergent in nature.<br />

Focus is placed on emergent symmetry-breaking physical processes resulting in formation <strong>of</strong><br />

electron crystalline clusters in 2D quantum dots, and on the breaking <strong>of</strong> trapped Bose-Einstein<br />

condensates into Bose quantum crystallites, occurring as the inter-atomic repulsion is increased.<br />

Emergent chemical cluster phenomena are discussed in the context <strong>of</strong> nanocatalysis by gold<br />

nanoclusters, with an emphasis on clusters supported on metal-oxide surfaces (specifically<br />

MgO), and the role <strong>of</strong> surface defects and cluster charging effects.


Cluster ferroelectricity<br />

Walt de Heer<br />

Georgia Institute <strong>of</strong> Technology<br />

Experiments on beams <strong>of</strong> isolated Nb, Ta, V, Al clusters with up to 200 atoms, show that<br />

these particles attain relatively large electric dipole moments at cryogenic temperatures. This<br />

ferroelectric property cannot be reconciled with normal metallic behavior since the ferroelectric<br />

state implies large internal electric fields. However various aspects <strong>of</strong> this state <strong>of</strong> matter appear<br />

to be closely related to bulk superconductivity. For example, the effect only occurs in metals<br />

that are superconductors in the bulk and a strong dependence even-odd alternation is observed.<br />

The ferroelectric transition temperature is similar to the bulk Tc. A survey <strong>of</strong> the electronic<br />

properties <strong>of</strong> these and other metallic clusters reinforce the superconducting pairing hypotheses<br />

in these clusters.<br />

19


20<br />

Molecular Magnets in the Gas Phase: Stern-Gerlach Beam Deflection<br />

Studies <strong>of</strong> Metal Clusters<br />

Mark B. Knickelbein<br />

Chemistry Division, CHM 200, Argonne National Laboratory, Argonne, Illinois 60439, USA<br />

The study <strong>of</strong> transition metal clusters in the gas phase <strong>of</strong>fers the opportunity to study the<br />

emergence <strong>of</strong> novel magnetic behavior atom-by-atom, in the size range that bridges atomic and<br />

bulk behavior. In this contribution, the results <strong>of</strong> Stern-Gerlach molecular beam deflection<br />

studies <strong>of</strong> transition metal clusters will be presented. Bare manganese clusters (Mn5-99) exhibit<br />

spatial deflections or broadening <strong>of</strong> magnitudes far in excess <strong>of</strong> those expected based on the<br />

susceptibility <strong>of</strong> bulk manganese, indicating that clusters in this size range are magnetically<br />

ordered. The magnitude <strong>of</strong> the magnetic moments (Figure 1), interpreted in light <strong>of</strong> recent<br />

density functional theory calculations, suggest that Mn clusters in this size range are molecular<br />

ferrimagnets. Recent magnetic deflection results for clusters composed <strong>of</strong> the Group IIIB<br />

metals, Scn, Yn, and Lan, indicate that they are also molecular magnets. Several Group IIIB<br />

clusters stand out as bona fide high-moment molecular magnets: Sc13 (6.0±0.2µb), Y8<br />

(5.2±0.1 µb), Y13 (9.6±0.1 µb), and La6 (4.8±0.2 µb).<br />

References<br />

moment per atom (µ b )<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

12<br />

13<br />

10 20 30 40 50 60 70 80 90 100<br />

Figure 1. Magnetic moments per atom for manganese clusters produced at 68K.<br />

[1] M. Knickelbein, Phys. Rev. Lett. 86, 5255 (2001).<br />

[2] M. Knickelbein, Phys. Rev. B 70, 014424 (2004).<br />

19<br />

n<br />

57<br />

Mn n


Oxidation Effects <strong>of</strong> the Small Chromium and Manganese Cluster<br />

Ions<br />

K. Tono, 1,2 A. Terasaki, 2 T. Ohta, 1 and T. Kondow 2<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Tokyo, 7-3-1 Hongo, Bunkyo, Tokyo 113-0033, Japan<br />

2 Cluster Research Laboratory, Toyota Technological Institute in East Tokyo Laboratory, Genesis<br />

Research Institute, Inc., 717-86 Futamata, Ichikawa, Chiba 272-0001, Japan<br />

Properties <strong>of</strong> transition-metal clusters are affected by reaction with foreign elements such as<br />

nitrogen and oxygen. For example, it has been predicted that the magnetic property <strong>of</strong> a<br />

chromium dimer is drastically changed by either nitridation or oxidation [1,2]. Motivated by<br />

these theoretical studies, we investigated the electronic structures <strong>of</strong> the chromium- and<br />

manganese-oxide cluster anions through measurements <strong>of</strong> photoelectron spectra and analyses<br />

with the aid <strong>of</strong> the density functional theory (DFT). It was found that the spin magnetic moment<br />

<strong>of</strong> Cr2O – is as large as 9 µB, while that <strong>of</strong> Cr2 – is 1 µB. This result shows that the spin coupling<br />

between the chromium sites becomes ferromagnetic upon oxidation <strong>of</strong> Cr2 – whose chromium<br />

sites are coupled antiferromagnetically [3,4]. The mechanism <strong>of</strong> the ferromagnetic spin coupling<br />

is explained on the basis <strong>of</strong> the geometric and electronic structures obtained by the DFT<br />

calculations. The oxygen atom at the bridge site weakens the strong bond between the<br />

chromium atoms, which is responsible for the antiferromagnetic coupling in Cr2 – . In addition,<br />

significant hybridization between chromium 3d and oxygen 2p orbitals allows electrons in the<br />

same spin state to be transferred from the two chromium atoms to the oxygen atom.<br />

Consequently the local spins on the chromium sites are bound to form a ferromagnetic coupling<br />

in a similar scheme <strong>of</strong> the superexchange interaction encountered in solids. An oxidation effect<br />

<strong>of</strong> a similar kind was found for Mn2O – , in which a ferromagnetic coupling between the local<br />

spins (about 5.5 µB) at the manganese sites stabilizes its electronic structure [5,6].<br />

Manganese clusters differ distinctly from chromium clusters; the manganese atoms are very<br />

weakly bound because <strong>of</strong> the closed shell <strong>of</strong> 4s electrons. This is also true to singly charged<br />

manganese cluster cations, MnN + (N=3–7), where the binding energy is found to be less than 1<br />

eV/atom as reported in Ref. [7] and as shown further by the present measurements. The<br />

structural specificity results from a weak van-der-Waals like interaction between the constituent<br />

atoms, and is likely to change greatly if one introduces an oxygen atom into MnN + , because the<br />

oxygen could glue manganese atoms in its vicinity. Actually we elucidated how the introduced<br />

oxygen changes the geometrical and electronic structures <strong>of</strong> MnN + in photodissociation<br />

experiments <strong>of</strong> MnNO + (N=3–5). In practice, the mass spectra <strong>of</strong> photodissociation products<br />

from MnNO + were measured as a function <strong>of</strong> the photon energy <strong>of</strong> an excitation laser. The<br />

results show that the oxygen atom and two manganese atoms form a Mn2O + core ion and the<br />

other manganese atoms are weakly bound to the core ion. Analyses <strong>of</strong> the photodissociation<br />

action spectra <strong>of</strong> MnNO + are now in progress in order to obtain detailed information on the<br />

electronic and geometric structures.<br />

References<br />

[1] S. E. Weber, B. V. Reddy, B. K. Rao, and P. Jena, Chem. Phys. Lett. 295, 175 (1998).<br />

[2] B. V. Reddy and S. N. Khanna, Phys. Rev. Lett. 83, 3170 (1999).<br />

[3] K. Tono, A. Terasaki, T. Ohta, and T. Kondow, Phys. Rev. Lett. 90, 133402 (2003).<br />

[4] K. Tono, A. Terasaki, T. Ohta, and T. Kondow, J. Chem. Phys. 119, 11221 (2003).<br />

[5] K. Tono, A. Terasaki, T. Ohta, and T. Kondow, Chem. Phys. Lett. 388, 374 (2004).<br />

[6] S. N. Khanna, P. Jena, W.-J. Zheng, J. M. Nilles, and K. H. Bowen, Phys. Rev. B 69, 144418<br />

(2004).<br />

[7] A. Terasaki, S. Minemoto, and T. Kondow, J. Chem. Phys. 117, 7520 (2002).<br />

21


22<br />

Electronic structure and spin polarization <strong>of</strong> dilute magnetic<br />

semiconducting nanocrystals: a first principles approach<br />

Xiangyang Huang 1 , Adi Makmal 2 , James R. Chelikowsky 1 , and Leeor Kronik 2<br />

1 Department <strong>of</strong> Chemical Engineering and Materials Science and the Institute for the Advanced Theory<br />

<strong>of</strong> Information Materials, University <strong>of</strong> Minnesota, Minneapolis 55455, USA<br />

2 Department <strong>of</strong> Materials and Interfaces, Weizmann Institute <strong>of</strong> Science, Rehovoth 76100, Israel.<br />

Bulk dilute magnetic semiconductors have attracted considerable attention in recent years<br />

because they exhibit semiconducting and magnetic properties at the same time, potentially<br />

enabling novel electronic devices that manipulate the electron spin in addition to its charge,<br />

commonly known as “spintronic” devices. 1 Novel spintronic effects are expected at<br />

nanocrystalline dilute magnetic semiconductors because the carrier confinement greatly<br />

enhances spin-spin interactions.<br />

Here, we present what we believe to be the first ab initio study <strong>of</strong> the electronic structure <strong>of</strong> Mncontaining<br />

Ge, GaAs, and ZnSe nanocrystals, using a real-space pseudopotential-density<br />

functional theory approach. We find that in all cases significant spin-polarization is formed and<br />

that the magnetic moment distribution around the Mn atom displays clear chemical trends,<br />

becoming more localized with increasing bond ionicity, as shown in Fig. 1. A detailed analysis<br />

<strong>of</strong> the electronic structure reveals different patterns <strong>of</strong> level filling and ferromagnetic or antiferromagnetic<br />

couplings. Importantly, the electronic structure exhibits considerable quantum<br />

size effects. The electronic and magnetic properties <strong>of</strong> the dilute magnetic nanocrystals are<br />

compared and contrasted with the properties <strong>of</strong> the corresponding bulk systems. Potential device<br />

implications are discussed.<br />

References<br />

Figure 1. Distribution <strong>of</strong> the net magnetic moment for (from left to right)<br />

passivated MnGe81,MnGa40As41, and MnZn40Se41 nanocrystals.<br />

[1] S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnar, M. L. Roukes A.<br />

Y. Chtchelkanova, and D. M. Treger, Science 294, 1488 (2001).


Electronic structure and bonding in gold coated silica cluster: a nano<br />

bullet for tumor<br />

Q. Sun, Q. Wang, and P. Jena<br />

Physics Department, Virginia Commonwealth University, Richmond, VA 23284, USA.<br />

Recently an exciting application <strong>of</strong> gold coated silica nano-shells has been found in treating<br />

tumor and cancer. The nano-shell consists <strong>of</strong> silica (SiO2) core <strong>of</strong> order 100 nm coated with<br />

about 20 nm <strong>of</strong> gold. It absorbs near infrared light (NIR) and converts that light to heat, causing<br />

irreversible thermal cellular destruction. When the nano-shells were incorporated into human<br />

breast cancer cells in a test tube, and then exposed to NIR, 100 percent <strong>of</strong> the cancer cells were<br />

killed. Motivated by this experiment we have investigated the coating <strong>of</strong> a small silica cluster<br />

(<strong>of</strong> the order <strong>of</strong> a few nano-meters) with Au to understand (1) if such a small cluster can absorb<br />

infrared radiation and (2) if so, does the mechanism for such absorption have electronic origin?<br />

(3) How do Au atoms bind with silica core? We show for the first time that gold atoms bind to<br />

silicon atoms with dangling bonds and serve as seeds for the growth <strong>of</strong> Au islands (Fig. 1a). The<br />

large electron affinity <strong>of</strong> gold causes significant change in the electronic structure <strong>of</strong> silica<br />

resulting in a substantial reduction in the HOMO-LUMO gap (Fig. 1 b), thus allowing it to<br />

absorb near infrared radiation (Fig. 1 c). Our study suggests that a small cluster can have similar<br />

functionality in the treatment <strong>of</strong> cancer as the large size nano-shell, but for a different<br />

mechanism. The advantage <strong>of</strong> such small nano-bullet for targeting tumor and cancer is that it<br />

can easily penetrate the crowded environments such as the biological milieu <strong>of</strong> cells and live<br />

tissues for effective drug delivery.<br />

Au<br />

Au<br />

References<br />

Geometry<br />

O<br />

Si<br />

Au<br />

Energy (eV)<br />

(a) (b) (c)<br />

Figure 1. (a) Geometry, (b) HOMO-LUMO gap, and (c) optical spectra <strong>of</strong> Au 3-(SiO 2) 3.<br />

[1] L.R. Hirsch, et al. PNAS 100 , 13549 (2003).<br />

[2] Q. Sun, Q. Wang, B.K. Rao, and P. Jena, Phys. Rev. Lett. 93, 186803 (2004).<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

-7<br />

-8<br />

-9<br />

-10<br />

-11<br />

(SiO 2 ) 3<br />

Au 3 -(SiO 2 ) 3<br />

Energy spectra<br />

Optical absorption (arb.units)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Optic spectra<br />

0<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

Wave length (nm)<br />

23


24<br />

Low-temperature Oxidation <strong>of</strong> N2 on Supported Tungsten<br />

Nanoclusters<br />

Junichi Murakami 1 , Wataru Yamaguchi 2<br />

1 Nanotechnology Research Institute, National Institute <strong>of</strong> Advanced Industrial<br />

Science and Technology, Central 4, 1-1-1 Higashi, Tsukuba, 305-8562, Japan<br />

2 Materials Research Institute for Sustainable Development, National Institute <strong>of</strong><br />

Advanced Industrial Science and Technology, 2266-98 Anagahora, Shidami,<br />

Moriyama-ku, Nagoya, 463-8560, Japan<br />

It is known that enzymes in organisms <strong>of</strong>ten contain nanoclusters composed <strong>of</strong> several<br />

transition-metal atoms and sulfur or oxygen atoms. One <strong>of</strong> the famous examples <strong>of</strong> such<br />

enzymes is the nitrogenase that carry out nitrogen fixation (conversion <strong>of</strong> N2 in air into<br />

ammonia) at room temperature and 0.8 atm. Nitrogenase is known to carry several kinds <strong>of</strong><br />

nanoclusters, among which FeMo cluster (MoFe7S9) is believed to be most important for<br />

activation and reduction <strong>of</strong> N2 at room temperature. The fact that the nanocluster-containing<br />

enzymes carry out otherwise difficult chemical reactions suggests the nanoclusters are exotic<br />

materials that can catalyze the reactions under mild conditions.<br />

We recently started studying reactions <strong>of</strong> N2 on deposited transition-metal nanoclusters to<br />

investigate whether they also have such catalytic activities. Size-selected tungsten nanoclusters<br />

(Wn;n=2-6) are s<strong>of</strong>tlanded at room temperature on graphite (HOPG) surfaces that were<br />

bombarded by Ar+ ions before the cluster deposition. Defects created by the ion bombardment<br />

work as pinning centers for the clusters. The deposited nanoclusters are exposed to various<br />

gases and chemical reactions <strong>of</strong> N2 on the clusters were investigated.<br />

When the clusters at 140K were exposed simultaneously to N2 and H2O, it was found that N2O<br />

forms on the clusters. This was observed by X-ray photoelectron spectroscopy(XPS) and<br />

confirmed by thermal desorption spectroscopy(TDS). TDS measurements using isotopemers <strong>of</strong><br />

N2 and H2O revealed that N2, activated on the clusters, reacted with an O atom from H2O to<br />

form N2O at a temperature as low as 140K. The details including cluster-size dependence <strong>of</strong> the<br />

reaction will be presented at the symposium.<br />

References<br />

[1] W.Yamaguchi and J.Murakami, Chem. Phys. Lett. 378, 521 (2003).


Catalytic CO oxidation on ionic platinum clusters<br />

O. Petru Balaj, Iulia Balteanu, Tobias Roßteuscher, Vladimir E. Bondybey, Martin K. Beyer<br />

Department Chemie, Physikalische Chemie 2, Technische <strong>Universität</strong> München, Lichtenbergstraße 4,<br />

85747 Garching, Germany<br />

Cationic and anionic platinum clusters with up to 25 atoms are produced from isotopically<br />

highly enriched platinum-195 ( 195 Pt, 97.28%) in a laser vaporization source, and their reactions<br />

are investigated by Fourier transform ion cyclotron resonance mass spectrometry. For certain<br />

clusters, depending on both charge state and size, efficient catalytic conversion <strong>of</strong> CO and N2O<br />

to CO2 and N2 is observed. If N2O is present in sufficient excess, the catalytic cycle shown in<br />

Figure 1 for Pt7 + is running continuously. Poisoning occurs by adsorption <strong>of</strong> several CO<br />

molecules, and can be suppressed by increasing the N2O pressure in the ICR cell. All clusters<br />

containing more than five platinum atoms readily adsorb CO, while the reactivity with N2O is<br />

strongly size-dependent [1]. Efficient catalytic conversion <strong>of</strong> CO is observed only for those<br />

clusters which react efficiently with N2O [2].<br />

Figure 1. Catalytic cycle <strong>of</strong> CO oxidation with N 2O on Pt 7 + as a catalyst. Pt7 + , Pt7O + , Pt 7O 2 + as well as Pt7CO + are<br />

active species in the catalytic cycles. Upon addition <strong>of</strong> a second CO molecule to Pt 7CO + , the efficiency <strong>of</strong> the<br />

oxidation reaction is significantly reduced, additional CO molecules poison the cluster. CO adsorption stops when<br />

Pt 7CO 10 + is reached.<br />

References<br />

[1] I. Balteanu, O. P. Balaj, M. K. Beyer, V. E. Bondybey, Phys. Chem. Chem. Phys. 6, 2910 (2004).<br />

[2] O. P. Balaj, I. Balteanu, T. T. J. Roßteuscher, M. K. Beyer, V. E. Bondybey, Angew. Chem. in<br />

print.<br />

25


26<br />

Cluster Reactions and Properties: Laying the Foundation for Cluster<br />

Assembled Materials<br />

A. W. Castleman Jr.<br />

Departments <strong>of</strong> Chemistry and Physics, Penn State University, University Park, PA 16802<br />

Currently, there is extensive interest in systems <strong>of</strong> finite size as they <strong>of</strong>ten give rise to<br />

unique properties that differ from those <strong>of</strong> an extended solid or the individual molecular<br />

constituents <strong>of</strong> which they are comprised. Particularly interesting are nanoscale systems whose<br />

composition can be selectively chosen, and ones whose individual characteristics can be<br />

retained when assembled as an extended material. A major long-term goal <strong>of</strong> our research in this<br />

area is to develop ways <strong>of</strong> tailoring the design and formation <strong>of</strong> new nanoscale materials <strong>of</strong><br />

chosen electronic and catalytic properties. Recent success has been obtained in producing alloy<br />

cluster species <strong>of</strong> specific size and composition that simulate various elements <strong>of</strong> the periodic<br />

table.<br />

As a complementary approach to conventional surface studies widely used in the field <strong>of</strong><br />

catalysis, it is becoming increasingly recognized that cluster science can help elucidate the<br />

physical and chemical properties <strong>of</strong> condensed phase catalysts and, can provide detailed<br />

information on the mechanisms <strong>of</strong> reactions and the nature <strong>of</strong> various reaction sites that enables<br />

certain catalytic materials to be especially effective. In addition, the possible direct use <strong>of</strong><br />

clusters as catalysts has aroused interest due to the difference in reactivities <strong>of</strong>ten observed for<br />

nanoscale materials compared to conventional bulk catalysts. Our latest findings on mass<br />

selected transition metal oxides, carbon containing, and metallic clusters will be presented. In<br />

related studies, we are employing femtosecond pump-probe techniques to explore the electronic<br />

properties <strong>of</strong> the clusters, including their excitation and relaxation behavior, which is providing<br />

insights into their evolving band structure. Findings pertaining to the electronic characteristics<br />

<strong>of</strong> metal-carbon systems and efforts to obtain deposits <strong>of</strong> mass selected species will also be<br />

discussed.


Melting, Pre-melting, and Glass Transitions in Gallium and Aluminum<br />

Clusters<br />

G.A. Breaux, B. Cao, C.M. Neal, and M. F. Jarrold<br />

Department <strong>of</strong> Chemistry, Indiana University, 800 East Kirkwood Avenue, Bloomington, Indiana 47405-<br />

7102, USA<br />

The melting <strong>of</strong> gallium and aluminum clusters with between 10 and 70 atoms has been<br />

examined using multicollision induced dissociation (MCID) calorimetry measurements and high<br />

temperature ion mobility measurements. A first order melting transition is indicated by a spike<br />

in the heat capacity due to the latent heat and an abrupt change in the average collision cross<br />

section due to a change in the volume or shape <strong>of</strong> the cluster when it melts. For some aluminum<br />

clusters (e.g. Al51 + and Al52 + ) the peaks in the heat capacity are bimodal suggesting the presence<br />

<strong>of</strong> a pre-melting transition where the surface <strong>of</strong> the cluster melts before the core. While premelting<br />

transitions are present in many simulations <strong>of</strong> cluster melting, this is the first time they<br />

have been observed experimentally. For some gallium clusters, the collision cross sections<br />

suggest that a melting transition occurs while there is no peak in the heat capacity. This is the<br />

signature for a glass transition which has been predicted to occur for some cluster sizes in<br />

several recent simulations. A glass transition occurs for clusters which are intrinsically<br />

amorphous - clusters that do not have a well-defined geometry.<br />

27


Phase Transition and Phase Coexistence in Atomic Clusters and Other<br />

Mesoscopic Systems<br />

28<br />

Mihai Horoi, Koblar Alan Jackson<br />

Department <strong>of</strong> Physics, Central Michigan University, Mount Pleasant, Michigan 48859, USA<br />

We investigate signatures <strong>of</strong> shape transition and shape coexistence phases in small atomic<br />

clusters and other mesoscopic systems, such as the atomic nuclei. One <strong>of</strong> the best know example<br />

is the prolate-compact shape transition/coexistence phenomena observed in the mobility<br />

experiments <strong>of</strong> Si clusters.[2] Using a hierarchical strategy that features an extensive tightbinding-based<br />

search <strong>of</strong> the energy surface, followed by a full density functional theory<br />

investigation <strong>of</strong> the most stable structures, we have determined the lowest-energy clusters across<br />

the range n=2 to 28.[3] The calculated properties <strong>of</strong> these clusters are in very good agreement<br />

with available measurements <strong>of</strong> dissociation energies, ionization energies and ion mobilities,<br />

providing strong evidence that these structures are the ones found in experiments. The<br />

calculations clearly exhibit a transition in the relative stability <strong>of</strong> prolate and compact clusters<br />

between n=25 and 26, coinciding exactly with the experimental behavior. The lowest energy<br />

prolate and compact structures are found almost degenerate, justifying the coexistence <strong>of</strong> shapes<br />

observed in the experiment. The binding energy per atom <strong>of</strong> the ground states vs N -1/3 exhibits a<br />

long plateau in the transition/coexistence region (see Fig.), suggesting that this phenomena may<br />

be related to a relative s<strong>of</strong>t surface tension. Similar behavior was recently found in the isotope<br />

152 <strong>of</strong> the nucleus <strong>of</strong> the element Sm, for which the ground sate, J π = 0 + , is known to be very<br />

deformed (oblate), and the first excited J π = 0 + , is compact. A similar signature in the binding<br />

energy per atom <strong>of</strong> the Sm isotopes can also be extracted from the available experimental data.<br />

Figure 1. Shape transition and shape coexistence signature in the binding energy per atom <strong>of</strong> Si clusters. Figure taken<br />

from Ref. [1], but red dots are the results <strong>of</strong> our simulations (Ref. [3]).<br />

References<br />

[1] T. Bachels and R. Schaefer, Chem. Phys. Lett. 324, 356 (2000).<br />

[2] R.R. Hudgins, I. Motoharu, and M.F. Jarrold, J. Chem. Phys. 111, 7865 (1999).<br />

[3] A.K. Jackson, M. Horoi, I. Chaudhuri, Th. Fraunheim, and A. A. Shvartsburg, Phys. Rev. Lett. 93,<br />

013401 (2004).


Vibrational Spectroscopy <strong>of</strong> Large Sodium Doped Water Clusters:<br />

Size Selection by Ionization-Vibration Coupling<br />

Christ<strong>of</strong> Steinbach, Udo Buck<br />

Max-Planck-Institut für Strömungsforschung, Bunsenstrasse 10, 37073 Göttingen, Germany<br />

One way to solve the problem <strong>of</strong> measuring reliable size distributions <strong>of</strong> weakly bound<br />

clusters is to dope the clusters with one sodium atom and to ionize the system by a single<br />

photon close to the threshold. 1,2 In this contribution we have extended this method to obtain<br />

complete size selection. This is achieved by coupling the UV radiation <strong>of</strong> a dye laser below the<br />

ionization threshold with the tunable infrared radiation <strong>of</strong> an optical parametric oscillator. The<br />

procedure works provided that there is sufficient coupling between the vibrational and the<br />

electronic motion <strong>of</strong> the ionization.<br />

We have applied this method to the measurement <strong>of</strong> the vibrational OH-stretch spectra <strong>of</strong><br />

Na(H2O)n clusters in the size range from n=8 to 80. The spectra are dominated by intensity<br />

peaks around 3400 cm -1 which we attribute to an increased transition dipole moment <strong>of</strong><br />

delocalized electrons which have been observed in calculations for this type <strong>of</strong> clusters. 3 The<br />

spectral features are discussed and compared with those obtained for pure water clusters 4 and<br />

water clusters doped with sodium ions. 5<br />

References<br />

[1] S. Schütte and U. Buck, Int. J. Mass. Spectrom. 220, 183 (2002).<br />

[2] C. Bobbert, S. Schütte, C. Steinbach, and U. Buck, Eur. Phys. J. D Journal 19, 183 (2002).<br />

[3] C. J. Mundy, J. Hutter, and M. Parrinello, J. Am. Chem. Soc. 122, 4837 (2000).<br />

[4] C. Steinbach, P. Andersson, J. K. Kazimirski, U. Buck, V. Buch, and T.A. Beu,<br />

J. Phys. Chem. A 108, 6165 (2004).<br />

[5] F. Schulz and B. Hartke, Phys. Chem. Chem. Phys. 5, 5021 (2004).<br />

29


30<br />

Novel elongated Chevrel-type (Mo3S3)nS2 clusters<br />

Gotthard Seifert and Sibylle Gemming<br />

Institut f. Physikalische Chemie, Technische <strong>Universität</strong>Dresden, D-01062 Dresden, Germany<br />

Clusters <strong>of</strong> the composition Mo3nS3n+2 can be regarded as building blocks <strong>of</strong> the<br />

corresponding sulfur-based Chevrel phases [1]. These compounds are generally built from equal<br />

amounts <strong>of</strong> two Mo3nS3n+2 species with different values <strong>of</strong> n, thus the overall formula unit <strong>of</strong> the<br />

network is Mo3nS3n+4. For the alkali metals as counter ions the charge state <strong>of</strong> the single cluster<br />

building block is -1. Thus, the clusters in these compounds occur in the same charge state as the<br />

free clusters recently observed by experimental studies [2].<br />

The cluster series up to n = 9 was investigated by density-functional-based calculations. Both<br />

the neutral and the anionic species were investigated. All structures up to Mo21S23 exhibit a<br />

HOMO-LUMO gap <strong>of</strong> about 0.5 eV. The relative stability <strong>of</strong> the neutral clusters shows a weak<br />

even-odd-alternation, which roughly correlates with the HOMO-LUMO spacings. The relative<br />

stabilities <strong>of</strong> the different members <strong>of</strong> the cluster sequence are highly dependent on the charge<br />

state, i.e. on the number and type <strong>of</strong> counter ions in a solid phase. For comparison the infinitely<br />

long [Mo6S6] n chain was investigated by DFT band structure calculations, employing periodic<br />

boundary conditions.<br />

Calculated I-V curves fur such clusters show that the quasi one-dimensional Chevrel-type<br />

structures are promising candidates for nanoelectronic devices.<br />

Figure 1. Chevrel type structures <strong>of</strong> the Mo6S8 , Mo9S11 , Mo18S20 clusters and part <strong>of</strong> an infinite chain <strong>of</strong> a Chevrel<br />

structure (left to right, black Mo atoms ).<br />

References<br />

[1] D. Salloum, R. Gautier, P. Gougeon, M. Potel, J. Solid. State Chem. 177, 1672 (2004).<br />

[2] N. Bertram , Y.D. Kim, G. Ganteför, to be published


Novel Properties <strong>of</strong> S<strong>of</strong>t-Nano-Molecular Clusters<br />

Koji Kaya 1 , Atsushi Nakajima and Masaaki Mitsui 2<br />

1 Discovery Institute, RIKEN, Wako, 351-0198 Saitama, Japan<br />

2 Department <strong>of</strong> Chemistry, Keio Uinversity, Hiyoshi, Yokohama, Japan<br />

Over the past 20 years, our group at Keio University has devoted to the development <strong>of</strong><br />

new material science in nao-size regime on the basis <strong>of</strong> small molecules and metal atoms.<br />

Several new findings have been already reported. However , these reports have been restricted<br />

to relatively small size clusters with mass number less than several hundreds. Recently we have<br />

succeeded in the synthesis <strong>of</strong> relatively large sized molecular clusters <strong>of</strong> aromatic molecules<br />

such as benzene, naphthalene and anthracene. From the photoelectron spectra <strong>of</strong> anion clusters<br />

<strong>of</strong> these molecules as shown in the figure, we found peculiar behavior <strong>of</strong> the clusters in the mid<br />

size range where ordered and disordered electronic properties coexists. We will discuss this<br />

subject.<br />

Then, discussin will extends to the post nanoscience project which will be conducted under the<br />

collaboration <strong>of</strong> several national organizations including RIKEN. The project focuses attention<br />

on the synthesis and understanding the mechanism <strong>of</strong> novel bio-miemetic materials by the<br />

collaboration among chemists, bio-scientists and information physicists.<br />

Photoelectron Spectra <strong>of</strong> Size Selected Anthracene Cluster<br />

Anions<br />

II<br />

I<br />

II<br />

I<br />

31


32<br />

Melting <strong>of</strong> Sodium Clusters: Where Do the Magic Numbers Come<br />

from?<br />

H. Haberland, T. Hippler, J. Donges, O. Kostko, M. Schmidt, and B. v. Issendorff<br />

Fakultät für Physik, <strong>Universität</strong> Freiburg, H.Herderstr. 3, 79104 Freiburg, Germany<br />

Melting temperatures <strong>of</strong> Na clusters show size-dependent fluctuations that have resisted<br />

interpretation so far. It will be discussed that these temperatures, in fact, cannot be expected to<br />

exhibit an easily understandable behavior. The energy and entropy differences between the<br />

liquid and the solid clusters turn out to be much more relevant parameters. They exhibit<br />

pronounced maxima that correlate well with geometrical shell closings, demonstrating the<br />

importance <strong>of</strong> geometric structure for the melting process. Icosahedral symmetry dominates, a<br />

conclusion corroborated by new photoelectron spectra measured on cold cluster anions. In the<br />

vicinity <strong>of</strong> the geometrical shell closings the measured entropy change upon melting is in good<br />

agreement with a simple combinatorial model.


Posters<br />

33


Carbon<br />

35


Structure and Bonding in Carbon Clusters C14 to C60<br />

Jeongeon Park, Jihye Shim, and Eok Kyun Lee<br />

Department <strong>of</strong> Chemistry, Korea Advanced Institute <strong>of</strong> Science and Technology, 373-1, Guseong-dong,<br />

Yuseong-gu, Deajeon, 305-701, Republic <strong>of</strong> Korea<br />

B - 1<br />

Density functional calculations have been applied to study various isomers <strong>of</strong> neutral carbon<br />

clusters C2n( 7 ≤ n ≤ 30 ) using both local spin density and generalized gradient approximations<br />

for the exchange-correlation energy. The structures <strong>of</strong> stable isomers include chains, rings,<br />

cages, and graphitic structures. We examined the most stable structure for each cluster <strong>of</strong><br />

various size and found a specific tendency as the size changes. Cage structure is the most stable<br />

structure for n≥16. We observed a fourfold periodicity in the monocyclic ring structures. Small<br />

effect <strong>of</strong> spin-polarization was observed in the monocyclic rings and the graphitic isomers with<br />

n = 4k( k = 1, 2, 3, ....). We calculated the ionization energy and the cohesive energy for the<br />

cage structures, and compared with previous results. All structures have been fully relaxed by ab<br />

initio molecular dynamics combined with simulated annealing, and the change <strong>of</strong> the<br />

trajectories <strong>of</strong> each atomic coordinate during the structure relaxation has been recorded . We<br />

further examined the characteristic features <strong>of</strong> the electronic structure <strong>of</strong> each isomer.<br />

References<br />

[1] R.O. Jones and G.Seifert, Phys. Rev. Lett. 79, 443 (1997).<br />

[2] G. Seifert, K. Vietze and R. Schmidt, eJ. Phys. B: At. Mol. Opt. Phys. 29, 5183 (1996).<br />

37


38<br />

B - 2<br />

High Resolution Measurements <strong>of</strong> Fullerenes and Endohedrals<br />

K. Głuch 2 , S. Matt-Leubner 1 , S. Feil 1 , O. Echt 3 , C. Lifshitz 4 , S. Denifl 1 , S. Ptasinska 1 , B. Concina 5 , P.<br />

Scheier 1 , and T. D. Märk 1<br />

1 Institut für Ionenphysik, Leopold Franzens <strong>Universität</strong>, A-6020 Innsbruck Austria<br />

2 Institute <strong>of</strong> Mathematics, Physics and Informatics Marie Curie-Sklodowska University, Lublin, Poland<br />

3 University <strong>of</strong> New Hampshire, Durham, New Hampshire, USA<br />

4 The Hebrew University <strong>of</strong> Jerusalem, Jerusalem, Israel<br />

5 Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Aarhus, Denmark<br />

The stability <strong>of</strong> fullerenes has been a controversial topic for some time. The (adiabatic)<br />

dissociation energies for the preferred dissociation reactions <strong>of</strong> isolated (gas-phase) fullerenes<br />

Cn → Cn-2 +C2 cannot be derived from measured thermodynamic quantities, not even for n = 60<br />

or 70. There is also a growing interest in fullerenes <strong>of</strong> sizes larger and smaller than C60 and C70.<br />

We have measured the kinetic energy released in the unimolecular dissociation <strong>of</strong> fullerene ions,<br />

Cn+ → Cn-2+ + C2, for sizes 42 ≤ n ≤ 90 [1]. A three sector field mass spectrometer equipped<br />

with two electrostatic sectors has been used in order to ensure that contributions from<br />

isotopomers (same n but containing 13C) do not distort the experimental kinetic energy release<br />

distributions. We apply the concept <strong>of</strong> microcanonical temperature to derive from these data the<br />

dissociation energies <strong>of</strong> fullerene cations. They are converted to dissociation energies <strong>of</strong> neutral<br />

fullerenes with the help <strong>of</strong> published adiabatic ionization energies. The results are compared<br />

with literature values.<br />

The fascinating possibility <strong>of</strong> caging atoms was noticed immediately by Kroto et al when they<br />

proposed the hollow icosahedral structure <strong>of</strong> C60. The kinetic stability <strong>of</strong> endohedrals under<br />

ambient conditions has originally presented some puzzles. We have measured the kinetic energy<br />

release distributions for unimolecular C2 loss from singly and multiply charged Sc3N@C78z+<br />

(z=1,2) and Sc3N@C80z+ (z=1,2,3) [2]. Using finite heat bath theory we have deduced the<br />

dissociation energies <strong>of</strong> these endohedral ions toward loss <strong>of</strong> C2. The data show that the<br />

complexation energies i. e. the adiabatic binding energies between Sc3N and the fullerene cage<br />

Cnz+ are, for a given charge state and within the experimental uncertainty, identical for n = 76,<br />

78 and 80.<br />

Formation <strong>of</strong> negative fullerene ions, especially <strong>of</strong> C60–, has been a contentious issue over the<br />

last decade. We try to clarify the mystery <strong>of</strong> the zero energy peak <strong>of</strong> the negative C60 with the<br />

help <strong>of</strong> a specially dedicated electron monochromator instrument. Moreover we compare for the<br />

first time experimental data <strong>of</strong> negative empty cage fullerenes and endohedrals.<br />

References<br />

[1] K. Gluch, S. Matt Leubner, O. Echt, B. Concina, P. Scheier, and T.D. Märk, High-resolution<br />

kinetic energy release distributions and dissociation energies for fullerene ions Cn+, 42


Photoelectron Spectroscopy <strong>of</strong> Fullerene Dianions C76 2- , C78 2- and C84 2-<br />

Oli T. Ehrler 1 , Filipp Furche 2 , J. Mathias Weber 1 , and Manfred M. Kappes 1<br />

B - 3<br />

1 Institut für Physikalische Chemie mikroskopischer Systeme, <strong>Universität</strong> Karlsruhe (TH), Kaiserstrasse<br />

12, Karlsruhe, D-76128, Germany<br />

2 Institut für Physikalische Chemie, Theoretische Chemie, <strong>Universität</strong> Karlsruhe (TH), Kaiserstrasse 12,<br />

Karlsruhe, D-76128, Germany<br />

In the past years, investigations <strong>of</strong> multiply negatively charged anions (MCAs) have<br />

become a research field <strong>of</strong> their own. MCAs should obey a particular high degree <strong>of</strong> electron<br />

correlation to minimize Coulomb repulsion between the excess charges, which are no longer<br />

screened due to the overall neutral core and are thus an interesting benchmark system for<br />

theoretical chemists. Moreover, long-range Coulomb interaction between an outgoing electron<br />

and the still negatively charged core leads to a significantly different photodetachment threshold<br />

law than Wigner’s law for monoanions due to the formation <strong>of</strong> the repulsive Coulomb barrier<br />

(RCB), which can kinetically hinder the decay <strong>of</strong> excited MCAs and even stabilize metastable<br />

MCAs with respect to electron emission.<br />

Due to their high symmetry and available wide size distribution maintaining a similar<br />

morphologic pattern, multiply negatively charged fullerenes are a particularly suitable target to<br />

study size evolution effects in MCAs. In the present work [1,2], we report laser photoelectron<br />

spectra <strong>of</strong> the doubly negatively charged fullerenes C76 2- , C78 2- and C84 2- at 2.33 eV, 3.49 eV, and<br />

4.66 eV photon energy. From these spectra, second electron affinities and vertical detachment<br />

energies, as well as estimates for the RCBs are obtained. These results are discussed in the<br />

context <strong>of</strong> electrostatic models. They reveal that fullerenes are similar to conducting spheres,<br />

with electronic properties scaling with their size. The experimental spectra are compared with<br />

the accessible excited states <strong>of</strong> the respective singly charged product ions calculated in the<br />

framework <strong>of</strong> time dependent density functional theory (TDDFT).<br />

Figure 1. Size dependence <strong>of</strong> AEA 2 values for fullerenes.<br />

Black squares: Spectroscopic data with experimental uncertainty. Open squares/circles: upper/lower limit for AEA 2<br />

(fully correlated point charges, dielectrically shielded vs. uncorrelated delocalized charges, w/o dielectric shielding);<br />

open triangles: AEA2 values obtained from a scalable model (see [1]).<br />

References<br />

[3] O.T. Ehrler, J.M. Weber, F. Furche, and M.M. Kappes, Phys. Rev. Lett. 91, art. no. 113006 (2003).<br />

[4] O.T. Ehrler, F. Furche, J.M. Weber, and M.M. Kappes, J. Chem. Phys. — accepted.<br />

39


40<br />

B - 4<br />

Radiative Cooling <strong>of</strong> Fullerenes and Endohedral Fullerenes<br />

Martin Hedén, Fredrik Jonsson, Elin Rönnow, Andrei Gromov, Klavs Hansen, Eleanor E. B. Campbell<br />

Department <strong>of</strong> Physics, Göteborg University, SE-41296 Göteborg, Sweden<br />

One <strong>of</strong> the interesting properties <strong>of</strong> the fullerenes is that, due to the similarity between their<br />

binding energies and ionisation potentials, excited molecules have decay channels in which the<br />

neutral fragmentation competes with delayed ionisation. An additional competing channel is<br />

that <strong>of</strong> radiative decay. In order to be able to extract reliable information from experiments<br />

concerning the binding energies <strong>of</strong> the molecules it is essential to understand the various<br />

competing mechanisms and have a consistent picture <strong>of</strong> the behaviour <strong>of</strong> these highly excited<br />

molecules. In this presentation we return to the determination <strong>of</strong> the radiative cooling <strong>of</strong><br />

fullerene ions as determined by time-<strong>of</strong>-flight experiments [1]. We compare the extracted<br />

product <strong>of</strong> emissivity and dissociation energy from experiments in which the starting molecule<br />

is different. In particular, we were interested in comparing the emissivity <strong>of</strong> an endohedral<br />

fullerene (in this case La@C82) with empty fullerenes. The data is displayed in Fig. 1 and<br />

compared with earlier data from C60 and more recent data from the Århus group [2]. Rather<br />

good agreement is obtained between the different data sets for the small fullerenes with a<br />

substantially higher scatter in the points for the larger molecules. However, we note that there is<br />

no significant difference in the emissivity <strong>of</strong> the endohedral fullerene compared with that <strong>of</strong> the<br />

empty fullerenes. The experimental details and assumptions will be presented and results will be<br />

discussed in the context <strong>of</strong> a simple model to describe the radiative cooling.<br />

εD 3<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

40 44 48 52 56 60 64 68 72 76 80 84<br />

Figure 1. Product <strong>of</strong> emissivity and the cube <strong>of</strong> the dissociation energy, extracted from time-<strong>of</strong>-flight data for<br />

different parent molecules. The full squares are data obtained from C60, C70 or C84 starting molecules. The empty<br />

squares are starting from La@C82. Full circles are early data from [1]. Empty circles are extracted from data given<br />

by [2].<br />

References<br />

[1] K. Hansen and E.E.B. Campbell, J. Chem. Phys., 104 5012 (1996).<br />

[2] S. Tomita et al., Phys. Rev. Lett. 87 073401 (2001)<br />

N


Nucleation <strong>of</strong> single-walled carbon nanotubes on catalyst particles:<br />

MD and electronic structure calculations<br />

Arne Rosén 1, 1, 2<br />

, Feng Ding and Kim Bolton<br />

1 Experimental Physics, School <strong>of</strong> Physics and Engineering Physics, Göteborg University and Chalmers<br />

University <strong>of</strong> Technology, SE-412 96, Göteborg, Sweden<br />

2 School <strong>of</strong> Engineering, University College <strong>of</strong> Borås, SE-50190, Borås Sweden<br />

B - 5<br />

Metal catalyzed single-walled carbon nanotube (SWNT) nucleation was studied by classical<br />

molecular dynamics (MD) and electronic structure theory. The simulations revealed the atomiclevel<br />

mechanism for SWNT nucleation on metal catalyst particles. The SWNTs nucleate<br />

between 800 - 1400 K, [1] which is the same temperature interval used for chemical vapor<br />

deposition (CCVD) experiments. Also, in agreement with experimental results the nucleated<br />

SWNT has the same diameter as the catalyst particle [2]. The studies also show that a highly<br />

supersaturated carbon concentration in the catalyst particle is needed to initiate the nucleation<br />

process. [1] Based on the simulations a detailed Vapor-Liquid-Solid (VLS) growth model was<br />

developed for both liquid and solid catalyst particles. [1, 3] Furthermore, MD and electronic<br />

structure theory studies indicate that the catalyst particle must be able to maintain an open end<br />

<strong>of</strong> the growing SWNT in order to be suitable for growth.<br />

References:<br />

[1] F. Ding, K. Bolton and A. Rosén, J. Phys. Chem. B, 108, 17369-17377 (2004).<br />

[2] F. Ding, A. Rosén, and K. Bolton, J. Chem. Phys. 121, 2775-2779 (2004).<br />

[3] F. Ding, A. Rosen, and K. Bolton, Chem. Phys. Lett. 393, 309-313 (2004).<br />

41


42<br />

B - 6<br />

Sphere currents in fullerenes<br />

Mikael P. Johansson, Dage Sundholm, Jonas Jusélius, Juha Vaara<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Helsinki, P. O. Box 55, FI-00014 Helsinki, Finland<br />

We investigate the magnetically induced electric currents in fullerenes, using the newly<br />

developed GIMIC methodology [1]. With GIMIC, a quantitative measure <strong>of</strong> the current<br />

strengths can, for the first time, be obtained. A remarkable difference between neutral<br />

Buckminsterfullerene and its spherically aromatic [2] +10 cation is revealed. In both, uniform<br />

three-dimensional currents, circling the fullerenes are induced. In C60, these sphere currents are<br />

oppositely directed; the exohedral diamagnetic and endohedral paramagnetic currents to a large<br />

extent cancel, resulting in global non-aromaticity. C60 10+ , on the other hand, supports a<br />

unidirectional, global diamagnetic current <strong>of</strong> significant strength, consistent with spherical<br />

aromaticity. These global sphere currents should be considered the defining feature <strong>of</strong><br />

aromaticity in fullerenes [3]. The stabilising effect <strong>of</strong> spherical aromaticity is discussed in<br />

connection with the recently postulated all-golden fullerene, Au32 [4]. Spherical aromaticity is<br />

found to be <strong>of</strong> crucial importance for the stability <strong>of</strong> the species.<br />

References<br />

Figure 1. From left to right: Au 32; sphere currents; C 60.<br />

[1] J. Jusélius, D. Sundholm, J. Gauss, "Calculation <strong>of</strong> current densities using gauge-including atomic<br />

orbitals", J. Chem. Phys. 121, 3952 (2004).<br />

[2] A. Hirsch, Z. Chen, H. Jiao, "Spherical Aromaticity in Ih Symmetrical Fullerenes: The 2(N+1) 2<br />

Rule", Angew. Chem. Int. Ed. 39, 3915 (2000).<br />

[3] M. P. Johansson, J. Jusélius, D. Sundholm, "Sphere Currents <strong>of</strong> Buckminsterfullerene", Angew.<br />

Chem. (in press).<br />

[4] M. P. Johansson, D. Sundholm, J. Vaara, "Au32: A 24-Carat Golden Fullerene", Angew. Chem. Int.<br />

Ed. 43, 2678 (2004).


Chemistry<br />

43


Gold cluster molecular surface chemistry<br />

André Fielicke 1 , Gert von Helden 1 , Gerard Meijer 1 , David B. Pedersen 2 , Benoit Simard 2 and<br />

David M. Rayner 2<br />

B - 7<br />

1 Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, D-14195 Berlin, Germany<br />

2 Steacie Institute for Molecular Sciences, National Research Council, 100 Sussex Drive, Ottawa, Ontario<br />

K1A 0R6, Canada<br />

The chemistry <strong>of</strong> gold nanoparticles is an important example <strong>of</strong> the fundamental difference<br />

between material in its bulk and in its cluster forms. Whereas bulk gold is chemically inert,<br />

deposited gold nanoparticles and clusters have a highly size-specific catalytic activity. A<br />

dramatic size-dependence in reactivity and catalytic turnover has even been found for isolated<br />

gold clusters containing only a few atoms.<br />

We are working to characterize prototypical reactions on free metal clusters in the gas phase.<br />

The intention is to provide a basis for understanding the chemistry <strong>of</strong> supported nanoparticles.<br />

Our approach is to monitor the independent adsorption, co-adsorption and reaction <strong>of</strong> relevant<br />

reaction partner molecules on size-selected clusters using infrared multi-photon depletion<br />

spectroscopy with a free electron laser. Our ultimate goal is to obtain thermodynamic<br />

information on cluster surface reactions through variable temperature studies.<br />

Here we discuss progress towards this goal, concentrating on gold clusters. We report on the<br />

interaction <strong>of</strong> carbon monoxide, CO, and nitric oxide, NO, with cationic and anionic gold<br />

clusters. Successive adsorption <strong>of</strong> CO molecules on the Aun+ clusters proceeds until a cluster<br />

size specific saturation coverage is reached. Structural information for the bare gold clusters is<br />

obtained by comparing the saturation composition with the number <strong>of</strong> available equivalent sites<br />

presented by candidate structures <strong>of</strong> Aun+. Our findings are in agreement with the planar<br />

structures <strong>of</strong> the Aun + cluster cations with n < �7 suggested by ion mobility experiments<br />

combined with DFT calculations [1]. By inference we also establish the structure <strong>of</strong> the<br />

saturated Aun(CO)m + complexes. In certain cases we find evidence that successive adsorption <strong>of</strong><br />

CO can distort the metal cluster framework. The vibrational spectra <strong>of</strong> the Aun(CO)m+<br />

complexes in both the CO stretching region and in the region <strong>of</strong> the Au-C stretch and the Au-C-<br />

O bend add further support to aid in the Aun+ structure determination, provide information on<br />

the structure <strong>of</strong> the CO complexes and allow comparison with CO adsorbates on deposited<br />

clusters or surfaces.<br />

For Aun- anions we find the same CO uptake patterns reported previously [2]. There is no clear<br />

correlation with geometric structure. Vibrational spectroscopy in the CO stretching region<br />

confirms molecular adsorption <strong>of</strong> CO, discounting the proposal that CO molecules might react<br />

together to form “glyoxylates” on the surface <strong>of</strong> Aun - [2].<br />

Nitric oxide absorption on Aun + shows similar uptake patterns to CO for small clusters, but NO<br />

absorbs less readily on clusters with n > 8. Vibrational spectroscopy shows that NO absorbs<br />

molecularly on Aun+ at 300 K. Aun+ is the first cluster system on which we have been able to<br />

find this direct evidence for molecular absorption <strong>of</strong> NO. The NO stretching frequency, ν(NO),<br />

is consistent with atop NO bonding in a bent configuration but shows a pronounced odd-even<br />

shift that we reproduce by DFT calculations. Finally, preliminary experiments on co-adsorbed<br />

CO and NO show that there is no reaction between CO and NO on Aun + clusters at 300 K.<br />

References<br />

[1] S. Gilb, P. Weis, F. Furche, R. Ahlrichs and M.M. Kappes, J. Chem. Phys. 116, 4094 (2001).<br />

[2] W.T. Wallace and R.L.Whetten, J. Phys. Chem. B 104, 10964 (2000).<br />

45


46<br />

B - 8<br />

Triplatinum carbonyl complexes: preference for terminal sites and<br />

chiral structures<br />

Timo Santa-Nokki and Hannu Häkkinen<br />

Department <strong>of</strong> Physics, Nanoscience Center, University <strong>of</strong> Jyväskylä, Finland<br />

We present a systematic density functional theory (DFT) study on the structures <strong>of</strong><br />

triplatinum carbonyls Pt3(CO)x q with x=1-6 and q=0,-1. The calculations are able to fully<br />

reproduce and explain the experimentally observed anomalous behaviour in the binding energy<br />

<strong>of</strong> the ligands in anionic cluster carbonyls by Grushow and Ervin, 1 with the three first ligands<br />

strongly bound (by 220-275 kJ/mol per ligand), the fourth and fifth ligand weakly bound (by<br />

135 kJ/mol), and the sixth ligand again stronger bound (by about 200 kJ/mol). The first three<br />

ligands bind in a tilted configuration at terminal sites in the plane <strong>of</strong> the triangular platinum<br />

core, making the half-saturated Pt3(CO)3 - a chiral complex. We discuss the relative preference <strong>of</strong><br />

terminal and bridge adsorption sites for CO in Pd, Pt, and Au cluster carbonyls in terms <strong>of</strong> the<br />

strength <strong>of</strong> the relativistic effects in the metal-carbon bond. We also discuss the implications <strong>of</strong><br />

the current work regarding design <strong>of</strong> bimetallic Au-Pt cluster nanocatalysts for CO oxidation. 2<br />

Figure 1. Sequential binding energy <strong>of</strong> CO ligands to anionic platinum cluster carbonyls Pt 3(CO) x - for x=1-6. The<br />

DFT results <strong>of</strong> this work are compared to TCID and PD data from Ref. 1. A conservative estimate for the accuracy <strong>of</strong><br />

the theoretical values is ±10 kJ/mol. For each x we also show the ground state structure <strong>of</strong> the anionic cluster<br />

carbonyl.<br />

References<br />

[1] A. Grushow and K.M. Ervin, J. Am. Chem. Soc. 117, 11612 (1995); A. Grushow and K.M. Ervin,<br />

J. Chem. Phys. 106, 9580 (1997); K.M. Ervin, Int. Reviews in Physical Chemistry 20, 127 (2001).<br />

[2] T. Santa-Nokki and H. Häkkinen, submitted.


Who probes whom: Early transition metal cluster ions meet<br />

(non)aromatic hydrocarbon molecules<br />

Britta Pfeffer, Stephanie Wies, Anita Lagutschenkov, Jens Brück, Matthias Vetter, Tobias Pankewitz,<br />

Gereon Niedner-Schatteburg<br />

TU Kaiserslautern, Fachbereich Chemie, Erwin-Schrödinger-Str. 52, 67663 Kaiserslautern, Germany<br />

B - 9<br />

As already known form former studies [1-4] benzene as well as common olefines react with<br />

niobium cluster cations through complete dehydrogenation. In contrast to other cluster sizes the<br />

Nb19 + cluster is the only one which exclusively binds the aromatic hydrocarbon molecule by<br />

intact adsorption. This exception can also be observed in the reaction with naphthalene (also a<br />

homoaromatic molecule).<br />

In the present work we focus on the difference in reactivity <strong>of</strong> homoaromatic and <strong>of</strong><br />

heteroaromatic systems (benzene, quinoline, naphthaline, pyridine, furan and thiophene).<br />

The data allow for tentative conclusions on the prevailing coordination geometries (parallel or<br />

perpendicular to “cluster suface”). The correspondence <strong>of</strong> cluster to bulk surfaces will be<br />

discussed.<br />

References<br />

[1] J. Chem. Phys. 102 (12), 4870-4884 (1995).<br />

[2] .J. Phys. Chem. 99 (42), 15497-15501 (1995).<br />

[3] J. Chem. Phys. 108 (13), 5398-5403 (1998).<br />

[4] J. Chem. Phys. 262 (1), 143-149 (2000).<br />

47


B - 10<br />

48<br />

Structural and Electronic Properties <strong>of</strong> (HAlO)n Clusters<br />

Yi Dong and Michael Springborg<br />

Physical and Theoretical Chemistry, University <strong>of</strong> Saarland, 66123 Saarbrücken, Germany<br />

The structure <strong>of</strong> (HAlO)n clusters was optimized for n up to 26 using our own 'Aufbau'<br />

approach and for n up to 18 using a genetic-algorithms approach together with a parameterized<br />

density-functional method. The two set <strong>of</strong> results for (HAlO)n from the 'Aufbau' and the geneticalgorithms<br />

approaches are very close, proving that our two unbiased approaches are reliable.<br />

The shape <strong>of</strong> the clusters as well as so-called similarity functions, stability functions are<br />

presented in order to analysing the structural properites <strong>of</strong> the (HAlO)n clusters. Subsequently,<br />

we also study the properties <strong>of</strong> two interacting clusters that were optimized by the 'Aufbau'<br />

approaches in order to study the marcroscopic, nanoscaled HAlO material that is produced in<br />

experiment.


Efficient Low-Temperature Oxidation <strong>of</strong> Carbon-Cluster Anions by<br />

SO2: Atmospheric Soot and Health Implications<br />

B - 11<br />

Andrew J. Leavitt 1,2 , Richard B. Wyrwas 2 , William T. Wallace 2+ , Daniel Serrano 1 , Melissa G. Arredondo 2 ,<br />

Farooq A. Khan 1 , and Robert L. Whetten 2*<br />

1 Department <strong>of</strong> Chemistry, State University <strong>of</strong> West Georgia, Carrollton, GA 30118.<br />

2 School <strong>of</strong> Chemistry and Biochemistry, Georgia Institute <strong>of</strong> Technology, Atlanta, GA 30332-0400.<br />

+Current address: Department <strong>of</strong> Chemistry, Texas A&M University, College Station, TX 77842-30012<br />

*Author to whom correspondences should be addressed.<br />

Carbon-cluster anions, CN - , are very active toward SO2 (sticking probability <strong>of</strong> 0.023 ±<br />

0.005 for C27 - at 25 o C), as compared to their inertness toward other common atmospheric gases<br />

and pollutants. In atmospheric-pressure, fast-flow-reactor experiments at ambient temperature,<br />

primary adsorption <strong>of</strong> SO2 by the carbon cluster anions, N = 4 – 60, yields CNSO2 - (αN) or CN-<br />

1S - (βN), with neutral CO2, as also detected in collision-induced dissociation. At higher<br />

temperatures, the reaction <strong>of</strong> SO2 with nascent carbon clusters yields the αN, βN, and CN-1SO -<br />

(γN) along with undetected CO2 and CO. The size-dependent initial reactivity reflects the<br />

previously established cluster structural transitions, i.e. from linear-chain to cyclic to cage<br />

structures. Secondary adsorption is also observed, with higher intrinsic reactivity, suggesting an<br />

accelerating oxidation process. Such carbon clusters serve as model compounds for blackcarbon<br />

soot, as they are formed in sooting flames and can act as nuclei for the formation <strong>of</strong><br />

primary soot particles, which also share the local structural features <strong>of</strong> active soot particle sites.<br />

These findings may therefore have implications for understanding the health and environmental<br />

effects attributed to the coincidence <strong>of</strong> soot and SO2.<br />

49


B - 12<br />

50<br />

Size Dependent Carbon Monoxide Adsorption on<br />

Neutral Gold Clusters<br />

Nele Veldeman 1 , Peter Lievens 1 , Mats Andersson 2<br />

1 Laboratorium voor Vaste-St<strong>of</strong>fysica en Magnetisme, K.U.Leuven,<br />

Celestijnenlaan 200D, B-3001 Leuven, Belgium<br />

2 Department <strong>of</strong> Experimental Physics, Chalmers University <strong>of</strong> Technology and<br />

Göteborg University, SE-412 96 Göteborg, Sweden<br />

In spite <strong>of</strong> gold being chemically inert as bulk material, supported nanosize gold clusters are<br />

found to show a remarkable catalytic activity for many reactions. Among these are CO<br />

combustion, NOx reduction, methanol synthesis and water-gas shift [1]. In order to clarify the<br />

catalytic behaviour shown by small gold particles supported on metal oxides, a number <strong>of</strong><br />

studies on deposited clusters, mainly focussing on the size and support dependence, have been<br />

performed [2,3]. Although the substrate <strong>of</strong>ten plays an important role, also the investigation <strong>of</strong><br />

chemical reactivity and catalytic activity <strong>of</strong> gas phase clusters is <strong>of</strong> great importance, as they<br />

may provide deeper insight in mechanisms <strong>of</strong> reaction processes. In this respect attention so far<br />

is especially drawn to reactivity and catalytic activity [4,5,6,7,8] <strong>of</strong> ionic gas phase gold<br />

clusters. As a complete understanding <strong>of</strong> the mechanisms involved in heterogeneous catalysis<br />

by nanoscale gold particles still remains a challenge, extended knowledge gained by reactivity<br />

experiments on neutral gas phase gold clusters towards reactants <strong>of</strong> interest could be very<br />

helpful.<br />

We report on experiments probing the reactivity <strong>of</strong> neutral Aun clusters, n=10-65, with CO gas.<br />

The gold clusters are produced in a pulsed laser vaporization cluster source [9], operated at<br />

room temperature (RT) or at liquid-nitrogen temperature (LNT), pass through a low-pressure<br />

reaction cell containing CO gas, and are subsequently laser ionised. The reaction probabilities<br />

are determined by recording mass abundance spectra with time <strong>of</strong> flight mass spectrometry.<br />

The main observations can be summarized as follows. (i) Upon cooling <strong>of</strong> the cluster source to<br />

LNT, the reactivity increases substantially. At LNT, the reaction probabilities for Aun with the<br />

first CO molecule are about a factor ten higher than at RT. Moreover, adsorption <strong>of</strong> two, three<br />

and even four CO molecules is observed, in contrast to RT clusters which at most adsorb one<br />

CO molecule. (ii) The reactivity is found to exhibit strong variations with cluster size, similar at<br />

both temperatures. The observed reaction probabilities yield evidence for electronic properties<br />

governing the size dependence: counting the number <strong>of</strong> delocalised electrons in reacted species<br />

reveals a correlation between reactivity and cluster electronic shell closings, and odd-even<br />

staggering in the reaction probabilities as a function <strong>of</strong> the number <strong>of</strong> atoms is observed.<br />

References<br />

[1] G.C. Bond and D.T. Thompson, Catal. Rev.-Sci. Eng. 41, 319 (1999)<br />

[2] M. Haruta, Catal. Today 36, 153 (1997)<br />

[3] A. Sanchez et al., J. Phys. Chem. A 103, 9573 (1999)<br />

[4] H. Häkkinen and U. Landman, J. Am. Chem. Soc. 123, 9704 (2001)<br />

[5] W.T. Wallace and R.L. Whetten, J. Am. Chem. Soc. 124, 7499 (2002)<br />

[6] J. Hagen et al., Phys. Chem. Chem. Phys. 4, 1707 (2002)<br />

[7] I. Balteanu et al., Phys. Chem. Chem. Phys. 5, 1213 (2003)<br />

[8] D. Stolcic et al. J. Am. Chem. Soc. 125, 2848 (2003)<br />

[9] M. Andersson et al., J. Phys. Chem. 100, 12222 (1996)


B - 13<br />

Photoelectron spectra <strong>of</strong> lanthanum oxide clusters: decreasing electron<br />

affinities with increasing number <strong>of</strong> oxygen atoms<br />

R. Klingeler 1 , G. Lüttgens 1 , P. S. Bechthold 1 , N. Pontius 2 , M. Neeb 2 , W. Eberhardt 2<br />

1 Institut fürFestkörperforschung, Forschungszentrum Jülich, 52425 Jülich, Germany<br />

2 Bessy GmbH, Albert-Einstein-Str. 15, 12489 Berlin, Germany<br />

We present a systematic investigation <strong>of</strong> the electronic structure <strong>of</strong> La2On (n=1-6) and<br />

La3Om (m=0-4) using photoelectron spectroscopy <strong>of</strong> anions (hν = 3.495 eV). For La2O and<br />

La2O2 the experiments are supported by density functional and configuration interaction<br />

calculations. The results for the C2v and the D2h ground state isomers, respectively, are in good<br />

agreement with the experimental data. In contrast to some other metal dimers and trimers we<br />

find as a general trend a decreasing electron affinity with an increasing number <strong>of</strong> chemisorbed<br />

oxygen atoms.<br />

References<br />

[1] R. Klingeler, G. Lüttgens, N. Pontius, R. Rochow, P.S. Bechthold, M. Neeb, and W. Eberhardt,<br />

Eur. Phys. J. D 9, 263 (1999).<br />

51


B - 14<br />

52<br />

Binding energies <strong>of</strong> CO and H2O on gold cluster cations<br />

Aun+ (n=1-65): A FT-ICR kinetics study<br />

Marco Neumaier 1 , Florian Weigend 1 , Oliver Hampe 1 , and Manfred M. Kappes 1,2<br />

1 Institut für Nanotechnologie, Forschungszentrum Karlsruhe, D-76021 Karlsruhe, Germany<br />

2 Institut für Physikalische Chemie, <strong>Universität</strong> Karlsruhe, D-76128 Karlsruhe, Germany<br />

Room temperature CO adsorption on isolated gold cluster cations is studied over a wide size<br />

range (Aun + , 1


The Bonding <strong>of</strong> CO with Au8: Evidence <strong>of</strong> Cluster Charging<br />

M. A. Röttgen 1 , A. S. Wörz 1 , K. Judai 1,3 , S. Abbet 1 , J. M. Antonietti 1 , U. Heiz 1 , B. Yoon 2 , and<br />

U. Landman 2<br />

1 Institute <strong>of</strong> Surface Chemistry and Catalysis, <strong>Universität</strong> Ulm, 89069 Ulm, Germany<br />

2 Georgia Institute <strong>of</strong> Technology, School <strong>of</strong> Physics, Atlanta, Georgia 30332-0430, USA<br />

3 Institute for Molecular Science, Myodaiji, Okazaki 444-8585, Japan<br />

B - 15<br />

While Gold is inert as bulk material, it becomes catalytically active as nanoscale particles or<br />

clusters supported onto oxide support materials. In the last few years, effort was made to get a<br />

better insight into the relevant reaction mechanisms and to understand this astonishing<br />

phenomenon on a molecular level. For large gold particles a few nanometers in size, it is<br />

discussed on the one hand whether the catalytic activity can be explained by the presence <strong>of</strong><br />

active sites on the cluster-support interface or by low-coordinated, highly reactive sites. On the<br />

other hand, for small clusters, the gold octamers bound to F-centers <strong>of</strong> an MgO surface are the<br />

smallest known gold heterogeneous catalysts able to transform CO into CO2. The key for the<br />

low-temperature conversion on Au8 is the binding <strong>of</strong> the molecular oxygen and the concomitant<br />

O-O bond activation to a peroxo-like adsorbate state. This is possible as the cluster’s electronic<br />

states show resonance with the 2π* molecular states <strong>of</strong> oxygen and as these antibonding states<br />

are pushed below the Fermi level <strong>of</strong> the system upon interaction with the cluster. Interestingly,<br />

the same cluster bound to a terrace site <strong>of</strong> an MgO surface is inert for the combustion reaction.<br />

Extensive ab initio calculations predicted that charging from the F-center into the cluster is<br />

essential for the activation <strong>of</strong> gold octamers.<br />

In this contribution, we present new findings showing experimental evidence for this prediction.<br />

The measured CO stretch frequencies, ν(CO), under reaction conditions are red-shifted by about<br />

30-50 cm-1 for the octamer bound to the F-center (Au8/CO/O2/MgO(FC)) in comparison to the<br />

octamer bound to the five-coordinated terrace sites (Au8/CO/O2/MgO(O5c)). Extensive ab<br />

initio calculations reveal that this shift is caused by enhanced backdonation into the antibonding<br />

2π* orbital <strong>of</strong> CO adsorbed on the Au8/O2/MgO(FC) complex and reveal important details <strong>of</strong><br />

the bonding <strong>of</strong> CO for understanding the backdonation. In addition, calculations on free<br />

Au8/O2/CO complexes indeed show this effect to be related to a substantial change in charging<br />

<strong>of</strong> the cluster [1].<br />

References<br />

[1] B. Yoon et al., Accepted for Publication in Science (2004).<br />

53


B - 16<br />

54<br />

Reactivity <strong>of</strong> Size-Selected Gas-Phase Transition Metal Carbide and<br />

Sulfide Clusters<br />

James Lightstone 1 , Melissa Patterson 1 , Robert J. Beuhler 2 1, 2<br />

, Michael G. White<br />

1 Department <strong>of</strong> Chemistry, Stony Brook University, Stony Brook, NY 11974, USA<br />

2 Chemistry Department, Brookhaven National Laboratory, Upton, NY 11973, USA<br />

We have recently constructed a cluster deposition apparatus which employs a magnetron<br />

sputtering source for generating gas-phase cation clusters <strong>of</strong> pure metals and metallic<br />

compounds. The focus is on generating clusters <strong>of</strong> the early transition metal compounds<br />

(carbide and sulfides), which are known in their bulk form to be active catalysts for a wide<br />

range <strong>of</strong> heterogeneous reactions [1,2]. In many cases, metal carbides <strong>of</strong>fer distinct advantages<br />

over the parent metal in selectivity and resistance to poisoning, and for certain reactions their<br />

catalytic behavior is similar to that <strong>of</strong> the Group 8-10 noble metals [2,3]. Very recent theoretical<br />

calculations also suggest that small carbide clusters, i.e., the Ti8C12 Met-Car and Ti14C13<br />

nanocrystallite, are more reactive than bulk TiC surfaces [4]. The work reported here examines<br />

the gas-phase reactivity <strong>of</strong> small transition metal carbide and sulfide clusters as a first step<br />

towards investigations <strong>of</strong> model catalysts prepared by size-selected deposition.<br />

Ion beams containing a wide array <strong>of</strong> transition metal compound clusters, TxMy + (T ≡ Ti,<br />

Mo, Zr, Nb; M ≡ C, S), were generated using a magnetron sputtering source, including “magic”<br />

number species such as the Ti8C12 + Met-Car. The gas-phase reactivity <strong>of</strong> mass-selected TxMy +<br />

clusters was studied using a hexapole collision cell, into which reactive gases were introduced,<br />

and a second quadrupole mass filter for product detection. Results will be presented for the<br />

adduct formation and reactions <strong>of</strong> the Ti8C12 + Met-Car with a variety <strong>of</strong> sulfur containing<br />

molecules, such as OCS shown in the accompanying figure below. In addition, we will also<br />

present results on the reactivity <strong>of</strong> MoxSy + clusters and future plans for size-selected cluster<br />

deposition. This work was supported by the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy<br />

Sciences, Division <strong>of</strong> Chemical Sciences under contract No. DE-AC02--98CH10886<br />

Figure 1. Mass spectrum <strong>of</strong> the reaction<br />

products resulting from collisions between<br />

Ti8C 12 + and OCS at 300K. .<br />

References<br />

500 600 700 800 900 1000<br />

[1] J.G. Speight, The Chemistry and Technology <strong>of</strong> Petroleum, second ed., New York, McGraw-Hill,<br />

1991.<br />

[2] Chen, J. G., Chem. Rev. 1996, 96, 1477; Oyama, S. T., Catal. Today 15, 179 (1992).<br />

[3] Levy, R. L.; Boudart, M., Science 181, 547 (1973).<br />

[4] H. Hou, J. T. Muckerman, P. Liu and J. A. Rodriguez, J. Phys. Chem. 107, 9344 (2003); P. Liu and<br />

J. A. Rodriguez, J. Chem. Phys. 119, 10895 (2003).<br />

Intensity<br />

1.20E-008<br />

1.00E-008<br />

8.00E-009<br />

6.00E-009<br />

4.00E-009<br />

2.00E-009<br />

0.00E+000<br />

Ti 8 C 12<br />

Ti 8 C 12 + S<br />

Ti 8 C 12 + SCO<br />

Ti 8 C 12 + S(SCO) 3<br />

Mass<br />

Ti 8 C 12 + S(SCO) 5


B - 17<br />

Probing Strong Hydrogen Bonds and Cold Metal Oxide Clusters with<br />

Infrared Photodissociation Spectroscopy<br />

Knut R. Asmis<br />

Abteilung Molekülphysik, Fritz-Haber-Institut der Max-Planck-Gesellschaft<br />

Faradayweg 4-6, D 14195 Berlin<br />

In recent years much effort has been aimed at improving the sensitivity and selectivity <strong>of</strong><br />

experimental methods to study the vibrational spectroscopy <strong>of</strong> gas phase ions. Due to the lack <strong>of</strong><br />

widely tunable infrared (IR) light sources most <strong>of</strong> these studies had been limited to the region<br />

above 2400 cm -1 , i.e., the spectral region <strong>of</strong> hydrogen-stretching motions and <strong>of</strong> IR-active<br />

combination bands. The application <strong>of</strong> the infrared free-electron laser FELIX, which generates<br />

tunable radiation in the 40 – 2200 cm -1 region, to molecular spectroscopy by Meijer, von Helden<br />

and coworkers has bridged this gap [1]. We have recently extended their technique to study the<br />

vibrational spectroscopy <strong>of</strong> mass-selected gas-phase cluster cations and anions at variable<br />

temperature. Application <strong>of</strong> this technique to two important compound classes will be discussed<br />

and compared to the results <strong>of</strong> complimentary experiments on mass-selected neutral clusters<br />

using femtosecond pump-probe spectroscopy.<br />

The hydrogen bond interaction is key to understanding the structure and properties <strong>of</strong> water<br />

and biomolecules. However, our understanding <strong>of</strong> strong, low-barrier hydrogen bonds and their<br />

central role in enzyme catalysis, biomolecular recognition, proton transfer across biomembranes<br />

and proton transport in aqueous media remains sketchy. We recently measured the first IR<br />

spectra <strong>of</strong> several prototypical systems, directly probing the shared proton region <strong>of</strong> the<br />

potential energy surface [2,3]. Our experiments demonstrate that the theoretical description <strong>of</strong><br />

the vibrations <strong>of</strong> strong hydrogen bonds is considerably more complex than originally<br />

anticipated and not yet solved satisfactorily, even for small systems and at the highest levels <strong>of</strong><br />

theory currently available.<br />

Transition metal oxides play an increasingly important role in heterogeneous catalysis.<br />

Recent gas-phase reactivity studies on vanadium oxide cluster ions were aimed at understanding<br />

the nature <strong>of</strong> the reactive sites. The interpretation <strong>of</strong> the reactivity data requires information on<br />

the structure <strong>of</strong> the cluster ions. We employ IR photodissociation spectroscopy in combination<br />

with electronic structure calculations to identify the geometric structure as well as structural<br />

trends as a function <strong>of</strong> cluster size in vanadium oxide cluster cations and anions [4]. For the<br />

cluster anions we present the first experimental evidence for the formation <strong>of</strong> polyhedral<br />

vanadium oxide cages in the gas phase. The measured IR spectra are sensitive probes for charge<br />

localization <strong>of</strong> unpaired d-electrons and show a strong resemblance with the IR spectrum <strong>of</strong> a<br />

V2O5 surface already at medium cluster sizes.<br />

References<br />

[1] G. von Helden, I. Holleman, G. M. H. Knippels, A. F. G. van der Meer, and G. Meijer, Phys. Rev.<br />

Lett. 79, 5234 (1997).<br />

[2] K. R. Asmis, N. L. Pivonka, G. Santambrogio, M. Brümmer, C. Kaposta, D. M. Neumark, and L.<br />

Wöste, Science 299, 1375 (2003).<br />

[3] M. Nee, C. Kaposta, A. Osterwalder, C. Cibrian Uhalte, T. Xie, A. Kaledin, S. Carter, J. M.<br />

Bowman, G. Meijer, D. M. Neumark, and K. R. Asmis, J. Chem. Phys. 121, 7259 (2004).<br />

[4] K. R. Asmis, G. Meijer, M. Brümmer, C. Kaposta, G. Santambrogio, L. Wöste, and J. Sauer, J.<br />

Chem. Phys. 120, 6461 (2004).<br />

55


B - 18<br />

56<br />

Oxygen adsorption at anionic free and supported Au clusters<br />

L. M. Molina 1,2 and B. Hammer 1<br />

1 iNANO and Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Aarhus, DK-8000 Aarhus C, Denmark<br />

2 Departamento de Física Teórica, Universidad de Valladolid, E-47011 Valladolid, Spain<br />

Recently, much attention has been paid to the catalytic activity <strong>of</strong> supported Au<br />

nanoparticles and clusters [1]. Both small free unsupported anions [2] and MgO-supported<br />

clusters [3] have been found very active towards CO oxidation. However, sizable differences in<br />

the size-dependent activity are found on each case, which are difficult to explain. In this work,<br />

the structure, stability and O2 adsorption properties <strong>of</strong> free anionic Aun - (n=1-11) clusters and<br />

negatively charged Aun clusters supported at defected MgO(100) surfaces are investigated using<br />

density functional theory. The free anionic clusters with an even number <strong>of</strong> atoms readily<br />

adsorb O2 since the excess unpaired electron can be easily transferred to the adsorbate. On the<br />

contrary, Au clusters supported at F + -centers show a much more complex behaviour. They attain<br />

similar electronic structure as the anionic free clusters, but the Madelung potential pins one<br />

electronic orbital to the defect. For large band gap clusters (mainly small clusters, which prefer<br />

2D geometries, see Figure) this is the orbital with the excess electron, which therefore cannot be<br />

transferred to an adsorbate, meaning that O2 adsorbs only weakly. Larger clusters become threedimensional<br />

and more metalic and hence capable <strong>of</strong> screening the support potential and binding<br />

O2 with charge transfer. This helps to understand why the supported clusters only become active<br />

starting at Au8 [3], whereas free Au6 - has been found to be extremely active [2].<br />

Figure 1. Equilibrium structures for O 2 adsorption at Au n (n=2-11) clusters attached to an F + -center. For n=4-11, the<br />

most representative 2D and 3D Au n isomers are shown. At Au 7-Au 8, a transition from 2D to 3D is found.<br />

References<br />

[1] M. Haruta, Catal. Today 36, 153 (1997).<br />

[2] W. T. Wallace and R. L. Whetten, J. Am. Chem. Soc. 124, 7499 (2002).<br />

[3] A. Sanchez et al., J. Phys. Chem A 103, 9573 (1999).


B - 19<br />

Density Functional Studies Of Noble Metal Clusters. Adsorption Of O2<br />

And CO On Gold And Silver Clusters<br />

Eva M. Fernández 1 , María B. Torres 2 , and Luis C. Balbás 1<br />

1 Dpto. de Física Teórica, Atómica y Óptica, Universidad de Valladolid E-4701 Valladolid, Spain<br />

2 Dpto. de Matemáticas y Computación, Universidad de Burgos E-090006 Burgos, Spain<br />

By means <strong>of</strong> first-principles density functional calculations, based on norm-conserving<br />

pseudo-potentials and numerical atomic basis sets, we study different noble metal cluster<br />

systems and properties. First, we show that, together with relativistic effects, the type <strong>of</strong><br />

exchange-correlation functional (GGA or LDA) is more critic for Au than for Ag or Cu clusters<br />

in order to determine the planar or three dimensional cluster geometry 1 . Second, we find cagelike<br />

stable structures for neutral Au18, Au20, Au32, Au50, and Au162, see fig.1. However, after<br />

adding an extra electron only Au20 - remains cage-like. On the other hand, Ag20 and Cu20 clusters<br />

adopt compact amorphous-like Cs structures. Third, we investigate the element- and sizedependent<br />

electron stability <strong>of</strong> gold clusters cations doped with a transition metal impurity 2 ,<br />

AunTM + (TM = Sc, Ti, V, Cr, Mn, Fe, Au; n ≤ 9), and we obtain a clear explanation <strong>of</strong> the<br />

cluster abundance peaks observed recently in photo-fragmentation experiments. Fourth, we<br />

study the size dependent adsorption <strong>of</strong> Om (m=2,4) molecules on small Agn - anion clusters (3≤ n<br />

≤9), as well as the adsorption <strong>of</strong> O2 on neutral Aun (5 ≤ n ≤ 10) and anions Aun - (n ≤ 7) clusters.<br />

The adsorption energy <strong>of</strong> O2 show marked odd even effects in all the cases. For Agn - anions<br />

with odd (even) n, the adsorption <strong>of</strong> a second O2 molecule increase (decrease) the adsorption<br />

energy <strong>of</strong> O2. The atomic O on Aun - O is preferable for n ≥ 4 than O2 adsorption. The<br />

dissociative adsorption is spontaneous for neutral gold clusters with n=6,7. For Aun, clusters<br />

with odd number <strong>of</strong> atoms are more active toward O2 adsorption. The O2 adsorbs preferably on<br />

top <strong>of</strong> Au atoms, except for Au5, for which a bridge site is favorable, see fig. 2. Fifth, we study<br />

also the size dependent adsorption <strong>of</strong> CO on small Aun clusters (5 ≤ n ≤ 10). The binding energy<br />

<strong>of</strong> CO adsorption on Au does not present odd even effects with clusters size. Top position is<br />

preferable for CO adsorption except for n = 5,7. Cluster with n = 5,7 present the biggest binding<br />

energy.<br />

We study also the trend in the bond distances, the vibrational frequencies and Mulliken<br />

population <strong>of</strong> adsorbed species on gold clusters.<br />

Figure 1. Cage-like and isomeric compact structures for Aun with n = 18, 20, 32, 50 and 162.<br />

Figure 2. O2 adsorption on Aun clusters. * dissociative adsorption. + molecular adsorption.<br />

References<br />

[1] E. M. Fernández, J. M. Soler, I. L. Garzón and L. C. Balbás, Phys. Rev. B 70, 165403 (2004).<br />

[2] M. B. Torres, E. M. Fernández and L. C. Balbás, Phys. Rev. B (submitted).<br />

57


B - 20<br />

Gold Clusters, CO-Chemisorbed Au Clusters, and Binary Au Clusters<br />

58<br />

Xi Li, B. Kiran, H. J. Zhai, Jun Li, Hai-Feng Zhang, and Lai-Sheng Wang<br />

Department <strong>of</strong> Physics, Washington State University, 2710 University Drive, Richland, WA 99352, USA<br />

and W. R. Wiley Environmental Molecular Sciences Laboratory and Chemical Science Division, Pacific<br />

Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, USA.<br />

Gold is a unique element with unusual chemistry due to the strong relativistic effect [1]. We<br />

will present our recent work on small gold clusters [2-4], CO-chemisorbed gold clusters [5], and<br />

novel gold-containing molecules [6-8]. We found previously that Au20 is a highly stable<br />

tetrahedral cluster [2]. Recently, we showed that we can also synthesize the tetrahedral Au20<br />

cluster in solution with phosphine ligands [4]. In experiments aimed at providing insight into the<br />

catalytic effect <strong>of</strong> nanogold, we studied CO-chemisorbed Au clusters, Aum(CO)n - (m = 2-5; n =<br />

1-7) [5]. The photoelectron data suggest that CO acts as an electron donor to the Au clusters and<br />

provide information about the CO-chemisorption site. This work provides direct experimental<br />

insight into the cooperative chemisorption <strong>of</strong> CO and O2 onto Au clusters.<br />

Au13 does not have the high symmetry Ih structure [3]. However, a class <strong>of</strong> stable Ih clusters<br />

containing 12 Au atoms and an impurity atom (M), M@Au12, was predicted to have Ih structures<br />

[9]. We will report our experimental observation and theoretical confirmation <strong>of</strong> these highly<br />

stable and symmetric gold clusters with M = Mo, W, V, Na, and Ta [6,7]. Finally we will<br />

present our discovery <strong>of</strong> the novel H-like chemistry <strong>of</strong> Au in forming the highly stable<br />

aurosilane SiAu4 and other Si-Au clusters in analogy to Si-hydride molecules [8].<br />

References<br />

Td Au20 Au20(PH3)4 Ih M@Au12 Td SiAu4<br />

[1] P. Pyykko, Angew. Chem. Int. Ed. 43, 4412 (2004).<br />

[2] J. Li, X. Li, H. J. Zhai, and L. S. Wang, Science 299, 864-867 (2003).<br />

[3] H. Häkkinen, B. Yoon, U. Landman, X. Li, H. J. Zhai, and L. S. Wang, J. Phys. Chem. A 107,<br />

6168-6175 (2003).<br />

[4] H. F. Zhang, M. Stender, R. Zhang, C. M. Wang, J. Li, and L. S. Wang, J. Phys. Chem. B 108,<br />

12259-12263 (2004).<br />

[5] H. J. Zhai and L. S. Wang, J. Chem. Phys., in press.<br />

[6] X. Li, K. Boggavarapu, J. Li, H. J. Zhai, and L. S. Wang, Angew. Chem. Int. Ed. 41, 4786-4789<br />

(2002).<br />

[7] H. J. Zhai, J. Li, and L. S. Wang, J. Chem. Phys. 121, 8369-8374 (2004).<br />

[8] B. Kiran, X. Li, H. J. Zhai, L. F. Cui, and L. S. Wang, Angew. Chem. Int. Ed. 43, 2125-2129<br />

(2004).<br />

[9] P. Pyykko and N. Runeberg, Angew. Chem. Int. Ed. 41, 2174-2176 (2000).


B - 21<br />

Mass-Selected Infrared Photodissociation Spectroscopy <strong>of</strong> Vanadium<br />

Oxide Cluster Anions<br />

G. Santambrogio 1 , M. Brümmer 1 , S. Fontanella 1 , L. Wöste 1 , J. Sauer 2 , G. Meijer 3 , and K. R. Asmis 3<br />

1 Institut für Experimentalphysik, Freie <strong>Universität</strong> Berlin, Arnimallee 14, Berlin, 14195, Germany<br />

2 Institut für Chemie, Humboldt-<strong>Universität</strong> Berlin, Unter den Linden 6, Berlin, 10099, Germany<br />

3 Abteilung Molekülphysik, Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, 14195,<br />

Berlin, Germany<br />

We present infrared spectra for gas phase vanadium oxide cluster anions. Structures are<br />

assigned on the basis <strong>of</strong> comparison between measured spectra and DFT calculations. We<br />

- -<br />

concentrate on the cluster size between V2O6 and V8O20 . Cage structures, which were originally<br />

predicted, [1] have now been spectroscopically identified for the first time. The spectra show<br />

three characteristic absorption regions <strong>of</strong> the IR spectrum which correspond to three different<br />

vibrational modes, namely the peroxo-modes at highest energies around 1100 cm -1 , the vanadylmodes<br />

in the intermediate region from 920 to 1020 cm -1 , and the V-O-V modes below 900 and<br />

to about 600 cm -1 . The smallest clusters studied, the divanadium oxide anions, can be classified<br />

-<br />

as “quasi-planar”, characterized by nearly planar V-O-V-O ring. V3O8 is an intermediate case; it<br />

-<br />

is the first anion that shows a three dimensional backbone structure. V4O10 is the first cage<br />

structure, in which each vanadium atom forms three V-O-V bonds and one vanadyl bond. Cages<br />

-<br />

are observed for all larger clusters up to V8O20 , indicating that these are particular stable<br />

-<br />

structures. The spectrum <strong>of</strong> V4O20 (see figure) reveals some striking similarities with the<br />

properties <strong>of</strong> a vanadium oxide single crystal surface[2] suggesting it is an interesting candidate<br />

for studying surface adsorption and surface reactivity on a model system in the gas phase.<br />

References<br />

Fragment Ion Signal<br />

Intensity<br />

0<br />

0<br />

Photon Energy (cm -1 )<br />

600 700 800 900 1000 1100 1200 1300 1400<br />

V-O(3)<br />

O-V-O<br />

V-O(2)<br />

V=O<br />

V-O(1)<br />

IR-PD (gas phase)<br />

V 8 O 20 ¯→ V 4 O 10 ¯ + V 4 O 10<br />

HREELS (surface)<br />

600 700 800 900 1000 1100 1200 1300 1400<br />

Electron Energy Loss (cm -1 )<br />

[1] S. F. Vyboishchikov and J. Sauer, J. Chem. Phys. A 105, 8588 (2000).<br />

[2] B. Tepper, B. Richter, A.-C. Dupuis, H. Kuhlenbeck, C. Hucho, P. Schilbe, M. A. bin Yarmo, and<br />

H.-J. Freund, Surface Science 496, 64 (2002).<br />

59


B - 22<br />

60<br />

Probing the Electronic Structure and Chemical Bonding <strong>of</strong> Metal<br />

Oxide Clusters Using Photoelectron Spectroscopy<br />

Hua-Jin Zhai, B. Kiran, Xi Li, and Lai-Sheng Wang<br />

Department <strong>of</strong> Physics, Washington State University, 2710 University Drive, Richland, WA 99352, USA<br />

and W. R. Wiley Environmental Molecular Sciences Laboratory and Chemical Science Division, Pacific<br />

Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, USA<br />

We report investigations <strong>of</strong> the electronic structure and chemical bonding <strong>of</strong> metal oxide<br />

clusters using photoelectron spectroscopy and theoretical calculations. One key advantage <strong>of</strong><br />

cluster studies is the flexibility to create clusters with any stoichiometry. It was shown [1] that<br />

metal oxide clusters with fixed numbers <strong>of</strong> metal atoms exhibit a behavior <strong>of</strong> sequential<br />

oxidation up to the maximum oxidation state <strong>of</strong> the metal atoms. For oxygen-rich clusters, O2<br />

unit begins to appear as per- or super-oxide. In this work, we show our recent investigation <strong>of</strong><br />

Mo and W oxide clusters (MoOx - and WOx - , x = 3-5) [2]. Comparisons between photoelectron<br />

spectra and theoretical calculations are used to provide detailed information about the structures<br />

and chemical bonding <strong>of</strong> the oxide clusters. The MO3 molecules are stoichiometric species and<br />

are observed to be closed shell with large energy gaps. However, we observe that the O2 moiety<br />

does not appear in the O-rich MO4 species, which exhibit a distorted tetrahedral structure with<br />

diradical character because <strong>of</strong> the strong M-O bonding. WO5 is found to be an unusual chargetransfer<br />

complex, (O2 - )WO3 + , also with diradical character.<br />

We will also report recent studies on the electronic structure <strong>of</strong> polyoxometalate anions (PMAs)<br />

using electrospray and photoelectron spectroscopy [3]. PMAs are a class <strong>of</strong> well-defined and<br />

negatively charged nanoclusters that exist in solution. The electrospray technique allows these<br />

species to be studied in the gas phase, facilitating direct comparisons between theory and<br />

experiment.<br />

References<br />

[1] L. S. Wang, H. Wu, and S. R. Desai, Phys. Rev. Lett. 76, 4853-4856 (1996).<br />

[2] H. J. Zhai, B. Kiran, L. F. Cui, X Li, D. A. Dixon, and L. S. Wang, J. Am. Chem. Soc. 126, 16134-<br />

16141 (2004).<br />

[3] X. Yang, T. Waters, X. B. Wang, R. A. J. O’Hair, A. G. Wedd, D. A. Dixon, J. Li, and L. S. Wang,<br />

J. Phys. Chem. A 108, 10089-10093 (2004).


Interaction <strong>of</strong> oxygen with mass-selected Ag cluster anions and<br />

nanoparticles<br />

Ignacio Lopez-Salido, Dong Chan Lim, Gerd Ganteför, and Young Dok Kim<br />

Department <strong>of</strong> Physics, University <strong>of</strong> <strong>Konstanz</strong>, D-78457 <strong>Konstanz</strong>, Germany<br />

http://www.clusterphysik.uni-konstanz.de/<br />

B - 23<br />

Adsorption <strong>of</strong> oxygen on free negatively charged Ag clusters were studied using Time-<strong>of</strong>-<br />

Flight mass spectrometry and photoelectron spectroscopy. In contrast to the results <strong>of</strong> Au cluster<br />

anions, which become inert towards oxygen chemisorption for Au n - with n > 20, Ag clusters<br />

consisting <strong>of</strong> more than 30 atoms are still reactive towards O 2 chemisorption. The even-odd<br />

alteration <strong>of</strong> the O 2 chemisorption reactivity was found, which has been also observed for the<br />

Au cluster anions. Photoelectron spectroscopy was used to study the nature <strong>of</strong> oxygen species<br />

on the Ag cluster anions. To shed light on the chemical properties <strong>of</strong> larger Ag nanoparticles,<br />

we deposited Ag on Highly Ordered Pyrolytic Graphite (HOPG) surfaces, leading to the<br />

formation <strong>of</strong> Ag nanoparticles with particle diameters up to 10 nm. Scanning Tunneling<br />

Microscopy (STM) and X-ray Photoelectron Spectroscopy (XPS) results suggest quite narrow<br />

size distributions <strong>of</strong> Ag islands on HOPG, allowing investigations on size dependent changes <strong>of</strong><br />

electronic and chemical properties <strong>of</strong> Ag nanoparticles. For the Ag nanoparticles smaller than<br />

10 nm, completely different oxidation patterns were observed compared to the Ag bulk crystal.<br />

We found Ag oxide species, which cannot be observed by oxidation <strong>of</strong> Ag bulk crystal.<br />

Implication <strong>of</strong> this result in heterogeneous catalysis is discussed in connection with the size<br />

selectivity in ethylene epoxidation.<br />

61


B - 24<br />

62<br />

Joint theoretical and experimental investigations <strong>of</strong> reactivity <strong>of</strong><br />

anionic and cationic gold oxide clusters with carbon monoxide<br />

Christian Bürgel 1 , Roland Mitrić 1 , Vlasta Bonačić-Koutecký 1 , Michele L. Kimble 2 , Nelly A. Moore 2 , A.<br />

Welford Castleman Jr. 2<br />

1 Humboldt <strong>Universität</strong> zu Berlin, Institut für Chemie, Brook-Taylor-Strasse 2, D-12489 Berlin, Germany<br />

2 Department <strong>of</strong> Chemistry and Physics, Pennsylvania State University, University Park, Pennsylvania<br />

16802, USA<br />

Recent literature shows that gold oxides, containing both atomic and molecular oxygen may<br />

play a role as the active site in supported gold catalysts. Therefore, a fundamental question can<br />

be raised as to whether it is sufficient to have an atomic oxygen in order for the oxidation<br />

reaction to proceed. A second fundamental question concerns the role <strong>of</strong> a charge. We will<br />

address both aspects by presenting the oxidation <strong>of</strong> CO in the presence <strong>of</strong> anionic and cationic<br />

gas phase gold oxide clusters. The influence <strong>of</strong> the charge on the association and replacement<br />

reaction has been also investigated. The experimental studies were carried out utilizing a metal<br />

ion flow tube in conjunction with a guided ion beam apparatus where gold oxides with both<br />

atomized and molecular oxygen are produced. Based on DFT and CCSD calculations we<br />

identified the reactive center and proposed reaction mechanisms. It will be shown that a<br />

peripheral O-atom is the most effective center for the oxidation reaction, while cooperative<br />

effects are mandatory for a reaction to proceed if molecular oxygen is involved. However, the<br />

presence <strong>of</strong> atomic oxygen does not guarantee that the oxidation reaction will proceed. We also<br />

wish to show the importance <strong>of</strong> cationic clusters for the association and replacement reactions.<br />

References<br />

[1] M. L. Kimble, A. W. Castleman, Jr., R. Mitrić, C. Bürgel, V. Bonačić-Koutecký, J. Am. Chem.<br />

Soc., 126, 2526 (2004).<br />

[2] M. L. Kimble, C. Bürgel, Nelly A. Moore A. W. Castleman, Jr., R. Mitrić,<br />

V. Bonačić-Koutecký, J. Am. Chem. Soc., in prep.


Dynamics<br />

63


Photoionization Dynamics <strong>of</strong> Pure Helium Droplets<br />

Darcy S. Peterka 1,2 , Jeong Hyun Kim 2 , Chia Wang 1 , Musahid Ahmed 2 , Daniel M. Neumark 1,2<br />

B - 25<br />

1 College <strong>of</strong> Chemistry, University <strong>of</strong> California, Berkeley, Address, Berkeley, 94720, USA<br />

2 Chemical Sciences Division, Lawrence Berkeley National Laboratory, 1 Cyclotron Rd, Berkeley, Ca<br />

94720, USA<br />

Helium nanodroplets have been demonstrated to provide a unique environment for<br />

spectroscopic studies. Their ability to capture and cool foreign dopants in a controlled way has<br />

led to their description as an "ultracold nanolaboratory". Many groups have used this property to<br />

their advantage, performing rotational and vibrational spectroscopic studies in great detail, with<br />

slightly less work being done by probing electronic transitions. However, there is almost no<br />

work on the photoionization <strong>of</strong> pure clusters. We have studied the single photon vacuum<br />

ultraviolet ionization <strong>of</strong> pure helium droplets below the atomic helium ionization threshold<br />

using velocity-mapped photoelectron imaging. We see wavelength dependant changes in the<br />

ionization dynamics <strong>of</strong> the droplet. Below the atomic threshold, indirect processes are the<br />

dominant source <strong>of</strong> electrons, leading to electrons <strong>of</strong> very low kinetic energy 1 . With photon<br />

energies larger than the Helium atom ionization energy, direct ionization to Hen + takes place,<br />

giving fast photoelectrons.<br />

References<br />

[1] D. S. Peterka, A. Lindinger, L. Poisson, M. Ahmed, and D.M. Neumark, Phys. Rev. Lett. 91,<br />

043401 (2003).<br />

65


B - 26<br />

66<br />

Excited-state relaxation dynamics in Au cluster anions<br />

J. Stanzel, F. Burmeister, M. Neeb, W. Eberhardt<br />

BESSY GmbH, Albert-Einstein-Str. 15, 12489 Berlin<br />

We report on time-resolved two-photon photoemission measurements on small gold cluster<br />

anions (Aun - , n = 6, 7, 8, 9, 14, 20). These measurements were performed to monitor the<br />

relaxation dynamics <strong>of</strong> optically excited states. As reported recently optically excited electronic<br />

states in Au3 - and Au6 - are long lived (> 1 ns) [1] in comparison to open d-shell metal clusters as<br />

Pt, Pd, Ni [2] and W [3]. This is mainly attributed to the large HOMO-LUMO band gap <strong>of</strong> gold<br />

clusters, which minimizes the probability <strong>of</strong> fast electron-electron-scattering. As we will present<br />

here, larger Au-clusters, e.g. Au8 - (fig. 1) can show relaxation processes in the lower picosecond<br />

range.<br />

Figure 1. Time resolved two-photon photoemission spectra <strong>of</strong> Au 8 - : The peak at 2.9 eV binding energy is recorded<br />

with the probe pulse (3.12 eV) and is assigned to the HOMO. In the pump & probe spectra (pump: 1.56 eV) a peak<br />

appears above the HOMO at 1.4 eV with its maximum at a delay <strong>of</strong> 0.1 ps. At longer delay times this peak broadens<br />

and shifts to higher binding energies until it completely vanishes.<br />

References<br />

[1] M. Niemietz, P. Gerhardt, G. Ganteför, Y.D. Kim, Chem. Phys Lett., 380, 99 (2003).<br />

[2] N. Pontius, M. Neeb, W. Eberhardt, G. Lüttgens, P. S. Bechthold, Phys. Rev. B 67, 354251 (2003)<br />

and references therein.<br />

[3] to be published.


Fast Electronic Relaxation in Metal Clusters via Excitation <strong>of</strong><br />

Coherent Shape Deformations: Slipping Through a Bottleneck<br />

V. V. Kresin 1 , Yu. N. Ovchinnikov 2 , V. Z. Kresin 3<br />

1 Dept. <strong>of</strong> Physics, University <strong>of</strong> Southern California, Los Angeles, California 90089-0484, USA<br />

2 Landau Institute for Theoretical Physics, Russian Academy <strong>of</strong> Sciences, Moscow 117332, Russia<br />

3 Lawrence Berkeley Laboratory, University <strong>of</strong> California, Berkeley, California 94720, USA<br />

B - 27<br />

We introduce and analyze a channel <strong>of</strong> electronic relaxation in metallic clusters which<br />

enables large energy transfer to the ionic subsystem and, correspondingly, fast electron<br />

relaxation dynamics. The discreteness <strong>of</strong> electronic levels in finite size-quantized systems such<br />

as metal nanoparticles, quantum dots, etc., represents a challenge for understanding the<br />

relaxation behavior <strong>of</strong> excited electrons. As is known, cases when the spacing between<br />

electronic energy levels exceeds the vibrational frequencies raise the prospect <strong>of</strong> so-called<br />

relaxation bottleneck phenomena. That is to say, if in order to drop to a lower-lying level an<br />

excited electron needs to emit multiple simultaneous vibrational quanta, the probability <strong>of</strong> such<br />

a transition should be strongly suppressed: multiphonon processes normally occur only as high<br />

order interaction corrections.<br />

The proposed relaxation mechanism which overcomes the bottleneck involves the special ability<br />

<strong>of</strong> free clusters to deform. A cluster with an electron in a previously unoccupied orbital (e.g.,<br />

excited by light or externally injected) will proceed to deform from its original shape.<br />

Microscopically, this deformation invalidates the selection rule which in the lowest order allows<br />

only single-phonon processes: whereas the rule assumes that the ionic equilibrium position is<br />

fixed during the electronic transition, a shift in the cluster’s ionic framework can produce a<br />

coherent multiphonon transition without any additional smallness. We have developed a method<br />

which allows us to describe the evolution <strong>of</strong> the deformation parameter up to the level crossing<br />

point, followed by electronic intershell transitions. The process is analogous to internal<br />

conversion in polyatomic molecules, but generally involves more than a single pair <strong>of</strong> electronic<br />

orbitals. It is found that the corresponding relaxation time indeed can be quite short.<br />

As an application, we have performed a model calculation for aluminum clusters. This was<br />

motivated by recent experimental work [1] which used time-resolved two-photon photoemission<br />

<strong>of</strong> Aln¯ (n=6-15) to demonstrate that the intermediate electronic state decays within a few<br />

hundred femtoseconds in all <strong>of</strong> the above cluster sizes, even in the closed-shell Al13¯. Our<br />

results are in agreement with the observed lifetime.<br />

This work was supported by the NSF (V.V.K.), a NATO Collaborative Linkage Grant (V.V.K.<br />

and Yu.N.O.), and DARPA (V.Z.K.).<br />

References<br />

[1] P. Gerhardt, M. Niemietz, Y. D. Kim, G. Ganteför, Chem. Phys. Lett. 382, 454 (2003).<br />

67


B - 28<br />

68<br />

Femtosecond Laser-Induced Fusion <strong>of</strong> Fullerenes Inside Clusters<br />

Martin Hedén, Klavs Hansen, Eleanor E. B. Campbell<br />

Department <strong>of</strong> Physics, Göteborg University, SE-41296 Göteborg, Sweden<br />

The interaction <strong>of</strong> fullerene molecules with femtosecond laser radiation has been the subject<br />

<strong>of</strong> much interest in recent years. One interesting observation is the clear separation <strong>of</strong> different<br />

ionisation mechanisms that can be observed as a function <strong>of</strong> the timescale for the excitation [1].<br />

We have shown that for intermediate timescales (ca. 70-300 fs) the ionisation can be largely<br />

regarded as a thermal electron emission from the equilibrated hot electron system before energy<br />

is transferred to the nuclear degrees <strong>of</strong> freedom [2]. Interesting dynamics have also been<br />

observed in collisions between fullerenes leading to molecular fusion <strong>of</strong> the collision partners<br />

and the formation <strong>of</strong> a metastable larger fullerene that subsequently decays by the emission <strong>of</strong><br />

C2 molecules [3].<br />

In this contribution we combine these two topics <strong>of</strong> recent research interest and show that it is<br />

possible to induce molecular fusion <strong>of</strong> fullerenes that are weakly (van der Waals) bonded in<br />

molecular clusters. Fig. 1. Shows a typical mass spectrum <strong>of</strong> the clusters after laser irradiation.<br />

The fragmentation pattern is clearly reminiscent <strong>of</strong> C2 emission from a large fullerene-like<br />

structure. We show that this observation is consistent with our knowledge <strong>of</strong> both the ionisation<br />

mechanisms <strong>of</strong> C60 on irradiation with fs laser pulses and with our knowledge <strong>of</strong> the energetic<br />

threshold for fullerene fusion and the C2 decay <strong>of</strong> large fullerenes. It is shown that substantial<br />

atomic rearrangement can take place on a timescale shorter than that needed for the complete<br />

destruction <strong>of</strong> the highly excited cluster. This is different for the observations made in<br />

experiments where fullerene clusters are collided with highly charged ions [4]. Reasons for the<br />

difference will be discussed.<br />

N<br />

Figure 1. Time-<strong>of</strong>-flight mass spectrum from fullerene clusters after excitation with a 200 fs laser pulse.A distribution<br />

<strong>of</strong> fragment masses is seen corresponding to each cluster size, with a mass separation <strong>of</strong> the fragments in each<br />

distribution <strong>of</strong> C2. This is a clear indication that the clusters have ionised and then had time for extensive atomic<br />

rearrangement to form a single large fullerene-like structure before the original clusters completely disintegrate.<br />

References<br />

counts<br />

10000<br />

1000<br />

60 80 100 120 140 160 180 200 220 240 260 280 300<br />

[1] E.E.B. Campbell, K. Hansen, K. H<strong>of</strong>fmann, G. Korn, M. Tchaplyguine, M. Wittmann, I.V. Hertel,<br />

Phys. Rev. Lett., 84 2128 (2000)<br />

[2] K. Hansen, K. H<strong>of</strong>fmann, E.E.B. Campbell, J. Chem. Phys. 119 2513 (2003).<br />

[3] E.E.B. Campbell, A.V. Glotov, A. Lassesson, R.D. Levine, C.R. Physique 3 (2002) 341<br />

[4] B. Manil, L. Maunoury, B. A. Huber, J. Jensen, H. T. Schmidt, H. Zettergren, H. Cederquist, S.<br />

Tomita and P. Hvelplund, Phys. Rev. Lett. 21 (2003) 215504


Femtosecond photoelectron spectroscopy <strong>of</strong> sodium clusters<br />

Abdollah Malakzadeh, Christian Hock, Christ<strong>of</strong> Bartels, Raphael Kuhnen and Bernd v. Issendorff<br />

Fakultät für Physik, <strong>Universität</strong> Freiburg, H.Herderstr. 3, 79104 Freiburg, Germany<br />

B - 29<br />

Recently we have shown that the electron gas in a free sodium cluster can be heated by a<br />

single fs-laser pulse to temperatures high enough for thermal electron emission to occur [1,2].<br />

This emission can be described by a simple model: the laser excites the collective motion <strong>of</strong> the<br />

electrons (the plasmon), which after a very short time (a few fs) decays into single electron-hole<br />

excitations. Due to the strong electron-electron interaction within a short time this nonthermal<br />

energy distribution <strong>of</strong> the electrons is transformed into a thermal one. The electron gas then<br />

cools by electron-phonon coupling.<br />

Time-resolved pump-probe photoelectron and phot<strong>of</strong>ragmentation spectroscopy has been<br />

applied to measure the electron-phonon-coupling constant as a function <strong>of</strong> cluster size. This has<br />

been done using relatively long laser pulses (τ ≈ 200 fs). In order to obtain more information<br />

about the early stages <strong>of</strong> the excitation, shorter laser pulses are necessary. Therefore a NOPA<br />

(noncolinear optical parametric amplifier) has been constructed, which allows the production <strong>of</strong><br />

wavelength tunable, short (τ < 50 fs) laser pulses, as well as a hollow capillary compressor.<br />

The performance <strong>of</strong> these devices and first results for sodium clusters are discussed.<br />

References<br />

[1] R.Schlipper, R. Kusche, B.v.Issendorff and H.Haberland, App. Phys. A. 72, 255 (2001)<br />

[2] M.Maier, M. Astruc H<strong>of</strong>fmann, and B.v. Issendorff, NJP 5, 3 (2003)<br />

69


B - 30<br />

70<br />

Photoelectron Spectroscopy on Small, Size Selected, Neutral Silver<br />

Clusters Embedded in Helium Droplets<br />

Andreas Przystawik, Paul Radcliffe, Josef Tiggesbäumker, Karl-Heinz Meiwes-Broer<br />

Institut für Physik, <strong>Universität</strong> Rostock, <strong>Universität</strong>splatz 3, 18051 Rostock, Germany<br />

Experiments on neutral clusters are <strong>of</strong>ten hampered by a broad distribution <strong>of</strong> cluster sizes.<br />

Therefore most <strong>of</strong> the work concerning the electronic structure <strong>of</strong> clusters is performed with<br />

charged particles, where one can prepare a molecular beam <strong>of</strong> species with a certain number <strong>of</strong><br />

atoms. We present a possibility to exploit absorption resonances to select a cluster size for<br />

excitation and ionization from the neutral beam.<br />

This technique is used to record two-photon photoelectron spectra <strong>of</strong> small silver clusters grown<br />

in helium nanodroplets by the pickup method. We derive information on both, the cluster and its<br />

electronic structure, and the interaction with the helium environment. An analysis <strong>of</strong> the spectra<br />

recorded at different wavelengths shows that all investigated systems undergo a rapid relaxation<br />

on a picosecond timescale to the lower edge <strong>of</strong> the broad excitation, see figure 1.<br />

One possibility <strong>of</strong> helium pickup sources is to prepare clusters in high spin states as already<br />

demonstrated with alkali clusters on the surface <strong>of</strong> droplets [1]. Despite the lack <strong>of</strong> selectivity in<br />

favor <strong>of</strong> high spin states when growing the clusters inside the droplet we can clearly detect<br />

silver dimers in triplet states not present in gas phase spectroscopy.<br />

Figure 1. Schematic diagram <strong>of</strong> the ionization dynamics <strong>of</strong> Ag 8 [2]. After excitation to the unoccupied band E *<br />

roughly 4.0 eV above the ground state, the cluster quickly relaxes to the lower edge EL. In a R2PI<br />

experiment, ionization occurs from this long-living level.<br />

References<br />

[1] P. Claas, D. Schumacher, and F. Stienkemeier, Phys. Rev. Lett. 92, 013401 (2004).<br />

[2] P. Radcliffe et al., Phys. Rev. Lett. 92, 173403 (2004).


Signatures <strong>of</strong> the Giant Dipole Resonance in the Response <strong>of</strong> Metal<br />

Clusters for Strong-Field Dual-Pulse Excitation<br />

Thomas Fennel, Tilo Döppner, Josef Tiggesbäumker and Karl-Heinz Meiwes-Broer<br />

Institut für Physik, <strong>Universität</strong> Rostock, <strong>Universität</strong>splatz 3, 18051 Rostock , Germany<br />

B - 31<br />

The response <strong>of</strong> metal clusters irradiated with a sequence <strong>of</strong> two femtosecond laser pulses is<br />

strongly dependent on the optical delay between the pulses. Recent experiments have shown<br />

that the adjustment <strong>of</strong> the delay provides a powerful way to control the coupling <strong>of</strong> strong laser<br />

radiation to finite systems. Both the yield <strong>of</strong> highly charged ions [1] as well as the kinetic<br />

energy <strong>of</strong> emitted electrons is strongly enhanced for a particular optimal delay. The dynamics <strong>of</strong><br />

the signals is much more pronounced compared to earlier measurements using stretched<br />

pulses [2].<br />

For the investigation <strong>of</strong> the time-dependent interaction <strong>of</strong> dual pulses with simple metal clusters<br />

numerical simulations have been performed using Thomas-Fermi-Vlasov molecular dynamics<br />

(TFV-MD) [3]. The semiclassical treatment allows a fully three-dimensional microscopic<br />

simulation <strong>of</strong> the laser-cluster interaction and provides a reasonable description <strong>of</strong> delocalized<br />

electrons in the cluster groundstate as well as the violent nonlinear response to a strong field<br />

close to TDLDA.<br />

Simulations on Na55 and Ag55 show that also in the strong field regime the most effective<br />

absorption is caused by resonant excitation <strong>of</strong> multi-plasmon modes. The calculations are in<br />

qualitative agreement with the measurements and clearly reproduce an optimal optical delay for<br />

enhanced ionization <strong>of</strong> the cluster [1], see Fig. 1, as well as a strong energy enhancement in the<br />

electron emission.<br />

Figure 1. (a) Calculated average charge <strong>of</strong> emitted ions after dual-pulse excitation <strong>of</strong> Na55 as a function <strong>of</strong><br />

the optical delay. Both pulses have a width <strong>of</strong> 50fs at a wavelength <strong>of</strong> 800nm. (b) Measured normalized<br />

dual-pulse signals <strong>of</strong> Ag 5+ from the explosion <strong>of</strong> silver clusters in helium droplets, resulting from the<br />

excitation with 130 fs pulses. Taken from Ref. [1].<br />

References<br />

[1] Döppner, Fennel et al., Phys. Rev. Lett, in press<br />

[2] Köller, Schumacher, Köhn et al., Phys. Rev. Lett. 82:3783, 1999<br />

[3] Fennel, Bertsch and Meiwes-Broer et al., Eur. Phys. J. D 29:367, 2004<br />

71


B - 32<br />

Fragmentation <strong>of</strong> small alkali cluster attached to helium nanodroplets<br />

72<br />

Patrick Claas 1 , Georg Droppelmann 1 , Claus-Peter Schulz 2 and Frank Stienkemeier 1<br />

1 Fakultät für Physik, <strong>Universität</strong> Bielefeld, <strong>Universität</strong>sstr. 25, D-33615 Bielefeld, Germany<br />

email: franks@physik.uni-bielefeld.de<br />

2 Max-Born-Institut, Max-Born-Str. 2a, D-12489 Berlin, Germany<br />

We use helium nanodroplet isolation (HENDI) to form cold small alkali clusters. In<br />

particular, potassium atoms are picked up successively by superfluid helium droplets in order to<br />

aggregate to clusters at temperatures <strong>of</strong> 380 mK. In a femtosecond pump-probe setup we<br />

determine fragmentation times <strong>of</strong> the clusters upon electronic excitation. In contrast to earlier<br />

experiments on free clusters (cf. Figure 1) we find substantially longer times in the<br />

fragmentation processes. The following aspects have to be considered: (a) population <strong>of</strong> highspin<br />

states [3]; (b) missing thermal activation because <strong>of</strong> the low temperature; (c) cage effects<br />

by the helium environment. Studies performed at different photon energies, droplet sizes, etc.<br />

allow a closer look at the fragmentation mechanism and the items given above.<br />

Furthermore first results are presented on the formation <strong>of</strong> complexes <strong>of</strong> alkali clusters with<br />

foreign atoms or molecules (e.g. H2O) attached to helium droplets.<br />

decay time [ps]<br />

Figure 1.: Fragmentation times upon femtosecond laser excitation <strong>of</strong> potassium clusters K n attached to helium<br />

nanodroplets as a function <strong>of</strong> cluster size. As a comparison results <strong>of</strong> gas-phase clusters by Wöste et al. are also<br />

included.<br />

References<br />

10<br />

1<br />

2 4 6 8 10<br />

cluster size n<br />

[1] H. Kühling et al., J. Phys. Chem. 98, 6679 (1994).<br />

[2] A. Ruff et al., Z. Phys. D 37, 175 (1996).<br />

[3] C. P. Schulz, et al., Phys. Rev. Lett. 92(1), 013401 (2004) .<br />

this work, 1.55eV, K<br />

this work, 1.47eV, K<br />

[1], 1.47 eV, K<br />

[1], 2.94 eV, K<br />

[2], 2.92 eV Na


Embedding <strong>of</strong> Na clusters in raregas matrices<br />

F. Fehrer 1 , P.-G. Reinhard 1 , E. Suraud 2 , M. Dinh 2 , G. Bousquet 2<br />

1 Institut für theoretische Physik, <strong>Universität</strong> Erlangen, Staudtstrasse 7, D-91058 Erlangen, Germany<br />

2 Laboratoire de Physique Quantique, Universite Paul Sabatier, 118 Route de Narbonne,<br />

F-31062 Toulouse, cedex, France<br />

B - 33<br />

We have developed a microscopic model for the description <strong>of</strong> a Na-cluster embedded in a<br />

raregas (Rg) substrate. The cluster valence electrons are treated at the level <strong>of</strong> TDLDA. The Rgatoms<br />

are considered as classical point-particles interacting with the cluster-electrons via local<br />

pseudopotentials. In order to describe the polarization <strong>of</strong> the raregas, we ascribe each Rg-atom a<br />

time-dependent dipolemoment. A short range repulsive Rg-core-potential together with a term<br />

taking the vdW-interaction into account is added. The model is calibrated to reproduce bindingproperties<br />

<strong>of</strong> the NaRg molecule and bulk Rg.<br />

We first study the influence <strong>of</strong> the Rg on the optical response <strong>of</strong> the cluster. We then investigate<br />

the coupled nonlinear dynamics <strong>of</strong> cluster and matrix following a strong laser excitation.<br />

73


B - 34<br />

Relaxation Dynamics <strong>of</strong> Optically Excited States <strong>of</strong> Ag Cluster Anions<br />

74<br />

Markus Engelke, Marco Niemietz, Young Dok Kim, Gerd Ganteför<br />

Fachbereich Physik, <strong>Universität</strong> <strong>Konstanz</strong>, 78457 <strong>Konstanz</strong>, Germany<br />

http://www.clusterphysik.uni-konstanz.de/<br />

For bulk metals, optically excited electronic states can relax within few tens <strong>of</strong><br />

femtoseconds due to very fast Auger-like electron-electron scattering processes. As the size <strong>of</strong> a<br />

particle decreases, changes in relaxation dynamics <strong>of</strong> optically excited states are observed [1,2].<br />

Using Time-Resolved Two-Photon Photoemission (TR-PPE) spectroscopy, relaxation times <strong>of</strong><br />

optically excited states <strong>of</strong> Ag cluster anions consisting <strong>of</strong> less than 22 atoms (pump photon<br />

energy <strong>of</strong> 1.5 eV) were determined. For Ag18 - , Ag19 - and Ag21 - , relaxation times amount to about<br />

500 femtoseconds, which are significantly longer than those <strong>of</strong> d-metal clusters consisting <strong>of</strong><br />

less than 8 atoms [3]. The long life times <strong>of</strong> the excited states <strong>of</strong> these Ag cluster anions can be<br />

explained by lower density <strong>of</strong> states above the upper levels <strong>of</strong> the non-excited electrons within<br />

the pump photon energy in these clusters, caused by a large gap between 2s and 1f electronic<br />

shells, i.e. electronic shell configurations significantly affect the electron relaxation dynamics.<br />

-<br />

Ag18 Intensity (a.u.)<br />

References<br />

0 fs<br />

133 fs<br />

266 fs<br />

932 fs<br />

1,9 ps<br />

12,9 ps<br />

48,6 ps<br />

0 1 2<br />

Binding Energy (eV)<br />

Figure 1. TR-PPE spectra <strong>of</strong> Ag 18 - . A distinct peak can<br />

be observed at a binding energy <strong>of</strong> about 1.3 eV,<br />

accompanying broad states at higher binding energies.<br />

The state at 1.3 eV can be attributed to the excitation <strong>of</strong><br />

an electron in the HOMO to an unoccupied state. As the<br />

delay between pump and probe photons increases, the<br />

signals between 1.2eV - 2.5 eV attenuate. By assuming<br />

an exponential decay <strong>of</strong> the excited state, the lifetime is<br />

estimated to be about 420 fs.<br />

[1] J.R.R. Verlet, A.E. Bragg, A. Kammrath, O. Cheshnovsky, and D.M. Neumark,<br />

J. Chem. Phys. 121, 10015 (2004).<br />

[2] M. Niemietz, P. Gerhardt, G. Ganteför and Y. D. Kim, Chem. Phys. Lett. 380, 99 (2003).<br />

[3] N. Pontius, M. Neeb, W. Eberhardt, G. Lüttgens, P. S. Bechthold, Phys. Rev. B 67, 354251 (2003)<br />

and references therein


Theoretical studies <strong>of</strong> ultrafast processes in clusters and metalbiomolecule<br />

complexes<br />

Roland Mitrić, Vlasta Bonačić-Koutecký<br />

B - 35<br />

Humboldt <strong>Universität</strong> zu Berlin, Institut für Chemie, Brook-Taylor-Strasse 2, D-12489 Berlin, Germany<br />

Our theoretical approach for exploration <strong>of</strong> ultrafast dynamics is based on combination <strong>of</strong><br />

the adiabatic and nonadiabatic ab initio MD ’on the fly’ with the Wigner distribution approach<br />

and permits to include all degrees <strong>of</strong> freedom. We show that it allows to accurately simulate<br />

femtosecond spectra in complex systems involving both ground and excited electronic states<br />

and to reveal the underlying dynamical processes [1].<br />

This will be illustrated on the following examples:<br />

i) Real time investigation <strong>of</strong> the ultrafast processes in the ground and excited states <strong>of</strong> noble<br />

metal clusters can provide insights relevant for understanding <strong>of</strong> their reaction dynamics. We<br />

present the simulation <strong>of</strong> the fs signals <strong>of</strong> the Ag2Au − /Ag2Au/Ag2Au + [2] system in the<br />

framework <strong>of</strong> the NeNePo spectroscopy. This allowed us to establish conditions under which<br />

the geometric relaxation can be distinguished from the IVR process which plays a crucial role<br />

for the stabilization <strong>of</strong> the complexes between the cluster and reacting molecule (e. g. O2 or<br />

CO). Study <strong>of</strong> cluster size effects on the dynamics in the excited electronic states <strong>of</strong> anionic<br />

clusters have been explored on the example <strong>of</strong> the Au5 − and Au7 − with the perspective to<br />

identify the dynamical processes in the time resolved photoelectron spectra. Based on these<br />

studies we propose the resonant two photon NeNePo technique, involving anionic excited<br />

states, which may serve for control <strong>of</strong> the reactivity <strong>of</strong> clusters.<br />

ii) Interplay between nonradiative and radiative relaxation processes play an important role in<br />

the design <strong>of</strong> efficient light-emitting materials. Influence <strong>of</strong> nonadiabaticity on the ultrafast<br />

excited state dynamics in the third 1 1 B1 electronic state <strong>of</strong> Na3F cluster has been investigated<br />

employing the ab initio MD ’on the fly’ in the framework <strong>of</strong> TDDFT [1,3] and CI methods. The<br />

simulated pump-probe signals exhibit characteristic oscillations at early times which are in good<br />

agreement with experimental results [3]. At later times simulations <strong>of</strong> the nonadiabatic<br />

dynamics in the framework <strong>of</strong> the CI method allow us to estimate the nonradiative lifetime.<br />

iii) Formation <strong>of</strong> complexes between silver clusters and tryptophan can possibly lead to<br />

enhancement <strong>of</strong> the light absorption and to fluorescence. Therefore, the influence <strong>of</strong> Ag + cation<br />

on the conformation and optical properties <strong>of</strong> tryptophan has been investigated. For this purpose<br />

we employ the combination between semi-empirical CI methods with the MD ”on the fly”.<br />

These studies serve as a starting point for the exploration <strong>of</strong> emissive properties and ultrafast<br />

dynamics.<br />

References<br />

[4] V. Bonačić-Koutecký, R. Mitrić, Chem. Rev., 105, 11 (2005).<br />

[5] T. M. Bernhardt, J. Hagen, L. D. Socaciu, J. Le Roux, D. Popolan, M. Vaida, L.Wöste, R.<br />

Mitrić, V. Bonačić-Koutecký , A. Heidenreich, J. Jortner, ChemPhysChem, 6, 243 (2005).<br />

[6] M. C. Heitz, G. Durand, F. Spiegelman, C. Meier, R. Mitrić, V. Bonačić-Koutecký, J. Chem.<br />

Phys., 121, 9906 (2004).<br />

75


B - 36<br />

76<br />

Rapid Alloying in binary Clusters: A Molecular Dynamics Study<br />

Y. Shimizu, T. Kobayashi, K. S. Ikeda and S. Sawada<br />

Ritsumeikan University, Kwansei Gakuin University<br />

Dynamics <strong>of</strong> surface atoms penetrating into microclusters is investigated in connection with<br />

the very rapid alloying (RA) in metal microclusters. RA was experimentally discovered by<br />

Yasuda and Mori for various types <strong>of</strong> nano-sized binary metal clusters. In short, RA is a<br />

manifestation <strong>of</strong> an unexpectedly fast diffusion <strong>of</strong> solute atoms into a host cluster, which is<br />

controlled by the magnitude <strong>of</strong> heat <strong>of</strong> solution and size <strong>of</strong> a cluster. Isothermal molecular<br />

dynamics simulation is done to investigate the mechanism <strong>of</strong> RA.<br />

In particular, we put our focus upon a cooperative interplay between atomic rearrangement<br />

along a cluster surface and that along radial direction <strong>of</strong> a cluster.<br />

Our results are summarized in the following two points. The dependence <strong>of</strong> the alloying and<br />

that <strong>of</strong> radial diffusion upon negative heat <strong>of</strong> solution are elucidated. It is found that dependency<br />

<strong>of</strong> the “alloying rate” upon the negative heat <strong>of</strong> solution and the size <strong>of</strong> clusters reproduced the<br />

qualitative features observed in the experiments. To quantify the dependence <strong>of</strong> atomic<br />

rearrangement dynamics upon temperature, the activation process <strong>of</strong> diffusive motion <strong>of</strong> atoms<br />

is analyzed.<br />

Although the activation energy <strong>of</strong> radial diffusion is about two times as large as that <strong>of</strong><br />

surface diffusion, there are circumstantial evidences that the latter process controls the former<br />

one. In that sense the rapid radial diffusion <strong>of</strong> a surface atom into a cluster is caused by the<br />

surface diffusion unlike the usual diffusion in bulk solid. In order to clarify the mechanism<br />

converting the active surface motion into the rapid radial diffusion, we numerically enumerate<br />

reaction paths. We found some candidates <strong>of</strong> characteristic collective motion which would<br />

effectively induce the rapid radial diffusion in terms <strong>of</strong> the surface activity.<br />

These two results suggest that surface atoms diffuse into the inside <strong>of</strong> the cluster due to the<br />

accumulation <strong>of</strong> active surface motion during long-time dynamics beyond a microsecond, even<br />

though the cluster is almost in solid-phase. Whilst diffusion due to the mechanism is crucial for<br />

the onset <strong>of</strong> RA, the same atomistic process is similarly expected not only for binary clusters,<br />

but also for monotonous ones. It leads us to an idea that a microcluster can be viewed as a<br />

dynamic material in which a transport <strong>of</strong> atoms between the surface area and the core region is<br />

continually induced in the course <strong>of</strong> long time scale.


Magnetism<br />

77


Electronic and Magnetic Properties <strong>of</strong><br />

Lanthanide Organometallic Clusters<br />

Atsushi Nakajima<br />

Department <strong>of</strong> Chemistry, Faculty <strong>of</strong> Science and Technology, Keio University & JST-CREST,<br />

3-14-1 Hiyoshi, Kohoku-ku, Yokohama 223-8522, Japan<br />

B - 37<br />

Binary systems are very important to create functionality <strong>of</strong> materials, and binary clusters<br />

consisting <strong>of</strong> a few to hundreds atoms provide microscopic viewpoints toward chemical<br />

bonding, geometry, and electronic structures including spin states. For metal and organic<br />

ligands, a variety <strong>of</strong> 1D, 2D, and 3D structures are formed depending on their combinations, and<br />

they exhibit novel electronic properties, such as electron delocalization and multiple charge<br />

transfers.<br />

With the aid <strong>of</strong> proper organic ligands, 1D structures are preferably formed in gas phase<br />

reaction. For example, Eum(C8H8)n sandwich clusters were produced in gas phase by mixing the<br />

vapor <strong>of</strong> europium (Eu) atoms and 1, 3, 5, 7-cyclooctatetraene (C8H8) molecules and 18-layered<br />

clusters were produced at maximum, which is about 8 nm length. With photoionization and<br />

photoelectron spectroscopies, formation mechanism and electronic structures will be presented<br />

in the viewpoint <strong>of</strong> sequential reactions with harpoon mechanism.<br />

In the design <strong>of</strong> molecular magnetic materials, furthermore, a subject to conquer is how to<br />

align the spin in the cluster. One-dimensional sandwich clusters are an interesting target since<br />

their magnetism should reflects geometric anisotropy and the condition <strong>of</strong> electronic spins on<br />

each layer. Stern-Gerlach type magnetic deflection measurements were performed for two types<br />

<strong>of</strong> multiple sandwich clusters: vanadium-benzene Vn(C6H6)n+1 and terbium-cyclooctatetraene<br />

Tbn(C8H8)n+1. Beams <strong>of</strong> Vn(C6H6)n+1 clusters (n = 1-4) showed symmetric broadening induced<br />

by the inhomogeneous field, indicating free spin behavior similar to that displayed by isolated<br />

paramagnetic atoms. By contrast, beams <strong>of</strong> Tbn(C8H8)n+1 clusters displayed one-sided deflection,<br />

indicating that fast spin relaxation occurs within the clusters. The difference in the magnetic<br />

deflection behavior exhibited by these two systems is explained by their electronic structures,<br />

specifically the bonding characteristics between metal atoms and ligand molecules.<br />

References<br />

[1] A. Nakajima, K. Kaya “A novel network structure <strong>of</strong> organometallic clusters”<br />

JPC A 104, 176 (2000).<br />

[2] S. Nagao, A. Nakajima, K. Kaya “Multiple-decker sandwich poly-ferrocene clusters”<br />

JACS 122, 4221 (2000).<br />

[3] K. Judai, K. Yagi, S. Yabushita, A. Nakajima, K. Kaya “A s<strong>of</strong>t-landing experiment <strong>of</strong> organometallic<br />

cluster ions: infrared spectroscopy <strong>of</strong> V(benzene)2 in Ar matrix” CPL 334, 277 (2001).<br />

[4] K. Miyajima, T. Yasuike, S. Yabushita, A. Nakajima, K. Kaya “The quasi-band electronic structure<br />

<strong>of</strong> Vn(Benzene)n+1 clusters” JPC A 106, 10777 (2002).<br />

[5] K. Miyajima, A. Nakajima, S. Yabushita, M. B. Knickelbein, and K. Kaya “Ferromagnetism in onedimensional<br />

vanadium-benzene sandwich clusters” JACS 126, 13202 (2004).<br />

[6] N. Hosoya, R. Takegami, K. Miyajima, M. B. Knickelbein, S. Yabushita, A. Nakajima “Lanthanide<br />

organometallic sandwich nanowires: formation mechanism” JPC A (Letter) in press.<br />

[7] K. Miyajima, M. B. Knickelbein, A. Nakajima “Stern-Gerlach studies <strong>of</strong> organometallic sandwich<br />

lusters” EPJD in press.<br />

79


B - 38<br />

80<br />

Mn clusters: a nanoscale magnetic transition<br />

Sudha Srinivas and Koblar Alan Jackson<br />

Department <strong>of</strong> Physics, Central Michigan University, Mount Pleasant, Michigan 48859, USA<br />

Small Mn clusters exhibit remarkable magnetic behavior. Early ESR experiments[1] found<br />

the smallest clusters (n=2-5) to be ferromagnetic (FM), with total net spin corresponding to 5 µB<br />

per atom, the Hund’s rule value for an isolated Mn atom. Later Stern-Gerlach measurements[2]<br />

found larger clusters (n>12) to have very small net moments. For example, Mn13 was found to<br />

have a net moment <strong>of</strong> only 3 µB for the entire cluster. Our recent calculations[3,4] explain that<br />

these data reflect a transition in magnetic ordering as a function <strong>of</strong> cluster size, occurring at n=7<br />

atoms. Specifically, the FM arrangements <strong>of</strong> the atomic spins favored in smaller clusters give<br />

way to antiferromagnetic (AF) arrangements in larger clusters. The calculations indicate that the<br />

FM → AF transition occurs at n=7, in agreement with experimental data, and is driven by a<br />

large change in the relative energies <strong>of</strong> the FM and AF structures. For n


The magnetic moments <strong>of</strong> isolated metal clusters<br />

Gereon Niedner-Schatteburg 1 and Wilfried Wurth 2<br />

1 Fachbereich Chemie, TU Kaiserslautern, Germany (gns@chemie.uni-kl.de)<br />

2 Institut für Experimentalphysik, <strong>Universität</strong> Hamburg, Germany (wilfried.wurth@desy.de)<br />

B - 39<br />

Within the last fiveteen years X-ray Magnetic Circular Dichroism (XMCD) has made itself<br />

useful in solid state physics and in surface science [2] since it allows for the elucidation <strong>of</strong><br />

magnetic properties that would be unavailable otherwise. For chemists XMCD has become a<br />

valuable tool when investigating paramagnetic organometalic and coordination compounds [2].<br />

In a few cases even paramagnetic metalloproteins have been studied successfully [3]. Only<br />

recently size selected iron clusters became subject <strong>of</strong> XMCD measurements [4]. All <strong>of</strong> these<br />

studies, however, have utilized condensed phase samples. Substrates and solvents may exert<br />

decisive but unknown influence on the magnetic properties <strong>of</strong> deposited and desolved samples,<br />

respectively. It is mandatory to identify suitable model systems for direct investigation and void<br />

<strong>of</strong> such an interference.<br />

We propose to apply XMCD to isolated gas phase metal clusters <strong>of</strong> selected sizes. This novel<br />

experimental scheme shall allow for determination <strong>of</strong> total magnetic moments as arising<br />

through spin and orbit contributions. Temperature tuning shall enable the recording <strong>of</strong><br />

magnetization curves.<br />

This shall be enabled through application <strong>of</strong> circularly polarized x-ray synchrotron radiation<br />

(600eV) to isolated gas phase transition metal clusters which have been size selected and stored<br />

in an ion trap [5,6] under high vacuum.<br />

The talk shall outline the basic features <strong>of</strong> the proposed scheme and it is going to suggest a<br />

particular way to carry out actual experiments.<br />

References<br />

[1] G. Schütz et al., Phys. Rev. Lett. 58, 737 (1987); C. T. Chen, N.V. Smith, and F. Sette, Phys. Rev.<br />

B 43, 6785 (1991); B. T. Thole et al., Phys. Rev. Lett. 68, 1943 (1992); M. Altarelli, Phys. Rev. B<br />

47, 597 (1993); C. T. Chen et al., Phys. Rev. B 48, 642 (1993); Y. U. Idzerda et al., Surf. Science<br />

287/288, 741 (1993)<br />

[2] S. P. Cramer in: ACS Symposium series Vol. 858 (2003)<br />

[3] J. Van Elp. et al., Proc. Nat. Acad. Sci. U.S.A. 90, 9664 (1993)<br />

[4] S. T. Lau, A. Föhlisch, R. Nietubyc, M. Reif, and W. Wurth, Phys. Rev. Lett. 89, 057201 (2002)<br />

[5] G. Niedner-Schatteburg and V. E. Bondybey, Chem. Rev. 100, 4059 (2000)<br />

[6] A. Lagutschenkov et al., Poster, S 3 C, Brand 2005<br />

81


Materials<br />

83


B - 39<br />

84


Intensity<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

Stable Clusters <strong>of</strong> Layered Semiconductors Prepared by Laser<br />

Ablation<br />

M. Galak 1 , O. Yeschenko 1 , I. Dmitruk 1 , V. Romanyuk 2 , A. Kasuya 2<br />

540<br />

500 600 700 800 900<br />

wavelength, nm<br />

B - 40<br />

1 Department <strong>of</strong> Physics, National Taras Shevchenko University<strong>of</strong> Kyiv, prosp. akad. Glushkov 2, Kyiv,<br />

01022, Ukraine<br />

2 Center for Interdisciplinary Research, Tohoku University, Sendai, 980-8578, Japan<br />

The aim <strong>of</strong> the research presented is to investigate cluster formation under pulsed laser<br />

ablation <strong>of</strong> layered semiconductors. Rapid evaporation <strong>of</strong> material provoked by high intensity<br />

laser beam creates favorable conditions for cluster formation. Recent success in investigations<br />

<strong>of</strong> new ultra-stable material compositions [1] led us to search for clusters with elevated stability<br />

<strong>of</strong> other materials.<br />

Films consisting <strong>of</strong> HgI2 and PbI2 nanoparticles (NP) have been prepared by pulsed N2 laser<br />

ablation. Absorption spectrum shows one broad peak <strong>of</strong> HgI 2 NP and one peak <strong>of</strong> PbI 2 NP<br />

corresponding to wide size distribution. Photoetching technique was used to decrease number <strong>of</strong><br />

large particles. Strong blue shift <strong>of</strong> PL peak <strong>of</strong> NP film caused by quantum size effect is<br />

observed (Fig.1). Semi-empirical structure calculations were performed using the MNDO basis<br />

set in the GAMESS package. The calculation shows that there are few stable clusters <strong>of</strong> HgI2<br />

and PbI2 with atomic masses less than 4000 a.m.u. To confirm theoretical calculations mass<br />

spectra <strong>of</strong> samples <strong>of</strong> HgI2 (Fig. 2) and PbI2 are obtained. Samples for mass-spectroscopy were<br />

prepared at similar conditions and in the same manner as for optical investigations. Peak<br />

positions correspond to stable clusters <strong>of</strong> both materials HgnIm and PbnIm consisting <strong>of</strong> n = 3, m<br />

= 5; n = 4, m = 7; n = 4, m = 9 atoms. There is a good agreement <strong>of</strong> theoretical prediction and<br />

observed peaks in mass spectra.<br />

Figure 1. Photoluminescence spectrum <strong>of</strong> HgI2<br />

stable clasters. Narrow line at 540nm<br />

corresponds to emmision <strong>of</strong> nanoparticles.<br />

References<br />

Figure 2. Mass-spectrum <strong>of</strong> HgnIm clusters<br />

prepared by laser ablation. Numbers in brackets<br />

corresponds to assigned number <strong>of</strong> atoms <strong>of</strong> Hg<br />

and I, respectively.<br />

[1] A.Kasuya, R.Sivamohan, Yu.Barnakov, I.Dmitruk, et al, Nature Materials, No.3 (2004).<br />

85


B - 41<br />

Stable Cluster Motifs For Nanoscale Materials<br />

S. N. Khanna<br />

Physics Department, Virginia Commonwealth University, Richmond, Va. 23284-2000, USA<br />

A major step in creating nanoscale materials with desirable electronic and magnetic<br />

properties is the identification <strong>of</strong> structural motifs that can simultaneously display the stability<br />

requisite for implementation as building blocks and the preservation <strong>of</strong> electronic and magnetic<br />

properties within this motif. Using three different examples (marking different bonding<br />

characteristics), we present our recent results on how by controlling size, composition and<br />

formation conditions, it may be possible to design very stable clusters with desirable traits that<br />

could serve as the building blocks for extended materials.<br />

The first part <strong>of</strong> the work concerns our recent investigations on chromium and iron oxide<br />

clusters. We have shown that by controlling the formation conditions, it is possible to generate<br />

different classes <strong>of</strong> metal-oxide clusters, each with unique electronic and magnetic properties. In<br />

particular, size-selected iron-oxide clusters can oxidize CO and reduce NO and <strong>of</strong>fer<br />

tremendous potential for applications [1].<br />

The second part <strong>of</strong> the work concerns our recent investigations on aluminum halogen clusters. It<br />

is shown that some <strong>of</strong> the acid etching reactions that can lead to very stable aluminum halogen<br />

mixed clusters. The studies provide evidence that Al13 - can act as a superhalogen when<br />

combined with conventional halogens. What is truly remarkable is that a completely new class<br />

<strong>of</strong> poly-halides can be formed by using Al13 - in combination with conventional halogens [2].<br />

Finally, we have just extended our previous work on semiconductor metal clusters [3]. We have<br />

re-examined some <strong>of</strong> rules that govern the stability <strong>of</strong> this important class <strong>of</strong> clusters and found<br />

that the metal-semiconductor bonding obeys simple electron counting rules. Needless to say,<br />

these rules can guide the search <strong>of</strong> new stable clusters.<br />

References<br />

[1] B. V. Reddy and S. N. Khanna, Phys. Rev. Lett. 93, 068301 (2004).<br />

[2] D. E. Bergeron, A. W. Castleman, Jr., T. Morisato and S. N. Khanna, SCIENCE 304, 84 (2004); D.<br />

E. Bergeron, P. J. Roach, A. W. Castleman, Jr., N. O. Jones, and S. N. Khanna, SCIENCE<br />

(Submitted).<br />

[3] S. N. Khanna, B. K. Rao, and P. Jena, Phys. Rev. Lett. 89, 016803 (2002).<br />

86


Passivation, charging and optical properties <strong>of</strong> gold nanoclusters<br />

Hannu Häkkinen<br />

Department <strong>of</strong> Physics, Nanoscience Center, University <strong>of</strong> Jyväskylä, Finland<br />

B - 42<br />

Nanoclusters and nanoparticles are currently a subject <strong>of</strong> intense experimental and<br />

theoretical research efforts due to their unique physical and chemical properties that might be <strong>of</strong><br />

potential use in future nanotechnologies. The properties <strong>of</strong> Gold nanoclusters are extraordinary<br />

sensitive to a number <strong>of</strong> factors, including size, charge state, passivation by inorganic ligands,<br />

and the type and amount <strong>of</strong> impurity atoms. Varying these factors allows one to "tune" or<br />

"functionalize" the electrical, optical and chemical properties in a desired way. This contribution<br />

will discuss phosphine- and thiol-passivated WAu12, Au39 and Au55 clusters in light <strong>of</strong> recent<br />

large-scale density functional simulations. It is shown that the properties <strong>of</strong> the gold core can<br />

easily be controlled by changing the ligand from π-acidic to σ-donating, i.e., by “chemical<br />

charging”. Exchange <strong>of</strong> phosphine ligands to CO molecules is energetically possible for some<br />

<strong>of</strong> the clusters, making them interesting candidates for nanocatalysts for CO oxidation.<br />

Figure 1. Spacefill visualization <strong>of</strong> trimethylphosphine-passivated Au 55 chloride Au 55(PMe 3) 12Cl 6 – a model for the<br />

famous “Schmid cluster” (ref. 1).<br />

References<br />

[1] G. Schmid, Chem. Rev. 92, 1709 (1992).<br />

87


B - 43<br />

88<br />

Deposition <strong>of</strong> Magic Silicon Clusters<br />

Rainer Dietsche, Felix von Gynz-Rekowski, Dong Chan Lim, Nils Bertram, Tim Fischer,<br />

Ignacio Lopez-Salido, Young Dok Kim, and Gerd Ganteför<br />

Fachbereich Physik, <strong>Universität</strong> <strong>Konstanz</strong>, 78457 <strong>Konstanz</strong>, Germany.<br />

http://www.clusterphysik.uni-konstanz.de/<br />

Since the discovery <strong>of</strong> the “supermagic” cluster C60 [1] researchers are fascinated by the<br />

possibility to synthesize new materials consisting <strong>of</strong> stable clusters. Aiming this question we are<br />

investigating small Silicon clusters and evaluating whether these clusters might be suitable as<br />

building blocks for cluster materials. Previous studies on mass-selected cluster anions <strong>of</strong> Silicon<br />

suggests, that clusters containing 4, 6, 7, and 10 Silicon atoms are closed-shell species with<br />

band gaps <strong>of</strong> 1 to 1.5 eV [2][3] indicating relatively inert and stable properties <strong>of</strong> these clusters.<br />

In our experiment “magic” Silicon clusters Si4 and Si7 are produced by a magnetron sputter<br />

source, mass selected and s<strong>of</strong>t landed (Ekin ≤ 0.3 eV/atom) on Highly Oriented Pyrolytic<br />

Graphite (HOPG), amorphous Carbon and Silver. The deposited clusters are studied via<br />

Ultraviolet Photoelectron Spectroscopy (UPS), X-Ray Photoelectron Spectroscopy (XPS), Low<br />

Energy Electron Diffraction (LEED), Auger Electron Spectroscopy (AES) and High-Resolution<br />

Electron Energy Loss Spectroscopy (HREELS). The results are compared to those <strong>of</strong> “nonmagic”<br />

clusters like Si8 and Silicon bulk.<br />

The experimental data indicates that “magic” Silicon clusters do not form bulk material upon<br />

deposition whereas Silicon monomers show bulk like electronic structures. Furthermore the<br />

deposited clusters show no significant reactivity towards Oxygen. However, additional<br />

measurements have to be made.<br />

References<br />

[1] H. W. Kroto, J. R. Heath, S. C. O’Brien, R. F. Curl, and R. E. Smalley, Nature 318, 162 (1985).<br />

[2] O. Cheshnovsky, S. H. Yang, C. L. Pettiette, M. J. Craycraft, Y. Liu, and R. E. Smalley, Chem.<br />

Phys. Lett. 138, 119 (1987).<br />

[3] J. Müller, B. Liu, A. A. Shvartsburg, S. Ogut, J. R. Chelikowsky, K. W. M. Siu, K.-M. Ho, and G.<br />

Ganteför, Phys. Rev. Lett. 85, 1666 (2000).


WS and MoS: Magic Clusters and Nanoplatelets studied by PES<br />

N. Bertram, J. Cordes, Young Dok Kim, G. Ganteför<br />

Department <strong>of</strong> Physics, University <strong>of</strong> <strong>Konstanz</strong>, 78457 <strong>Konstanz</strong>, Germany.<br />

http://www.clusterphysik.uni-konstanz.de/<br />

B - 44<br />

MoS and WS are very similar to carbon in the bulk. Thus it is interesting to see whether<br />

layered structures or even fullerenes can be formed out <strong>of</strong> these materials in the nano-regime.<br />

For WS clusters, a series <strong>of</strong> stable triangular platelet structures has been predicted by theory<br />

[1] consisting <strong>of</strong> 6, 10, 15 etc. W atoms forming the plane <strong>of</strong> the triangle and a corresponding<br />

number <strong>of</strong> S atoms located on both sides. WS clusters have now been generated in the PACIS<br />

source and studied by means <strong>of</strong> mass spectroscopy and photoelectron spectroscopy (PES),<br />

supporting the platelet hypothesis. The platelets show an electron affinity around 4 eV and, in<br />

agreement with theory, no electronic gap.<br />

In addition to that, several smaller magic MoS and WS cluster have been studied and very<br />

large electronic gaps (~ 2 eV) have been found for Mo4S6 and W4S6 for example, indicative <strong>of</strong> a<br />

high stability. For MoS, Chevrel phases have been observed, consisting <strong>of</strong> Mo6S8 building<br />

blocks. Due to their high stability, the aforementioned magic clusters are very promising<br />

building blocks for deposition on surfaces and the formation <strong>of</strong> cluster materials.<br />

References<br />

Figure 1. Measured photoelectron spectrum <strong>of</strong> W 4S 6 - and calculated structure <strong>of</strong> Mo4S 6 - by G. Seifert.<br />

Figure 2. Calculated structure <strong>of</strong> MoS-platelet by G. Seifert and Chevrel phase consisting <strong>of</strong> Mo 6S 8.<br />

[1] Gotthard Seifert, Institute for Physical Chemistry, Technical University Dresden, 01062 Dresden,<br />

Germany.<br />

89


Methods<br />

91


PEACE - a new photoelectron action spectroscopy<br />

Shai Ronen, Israel Wolf, Rina Giniger and Ori Cheshnovsky<br />

School <strong>of</strong> Chemistry, the Sackler Faculty <strong>of</strong> Exact Sciences,<br />

Tel Aviv University, Tel Aviv 69978 Israel<br />

B - 45<br />

Neumark and coworkers [1] have pioneered ZEKE spectroscopy on anionic molecules and<br />

clusters. Unfortunately, due to the low cross-section for electron detachment <strong>of</strong> anions at<br />

threshold, this method is very difficult to apply.<br />

We present a new experimental approach, in which the action spectroscopy is recorded with<br />

electrons <strong>of</strong> fixed finite kinetic energy. This approach circumvents some shortcomings <strong>of</strong> the<br />

anionic ZEKE. By working above threshold the cross-section is higher, allowing for higher<br />

signal, and the detection <strong>of</strong> electrons with l>s angular momentum. Neumark and coworkers<br />

have recently developed an equivalent technique based on the velocity map imaging.[2]<br />

The method is based on a modified Magnetic Bottle Photoelectron spectrometer (MBPES). A<br />

tunable laser is used to detach electrons from mass selected anions, which move collinearly with<br />

the 50 cm MBPES drift tube. To avoid Doppler broadening, a low voltage pulse removes the<br />

velocity component <strong>of</strong> anions from the detached electrons[3]. 20 cm downstream, after the<br />

parallelization <strong>of</strong> the electrons in the drift tube, another pulse is applied to reduce their velocity<br />

by a fixed amount. The spectrum is collected by recording the wavelength dependence <strong>of</strong><br />

electron-signal at a predetermined TOF window. This arrival time corresponds to a specific<br />

electron-kinetic energy. We have chosen to call this approach PEACE, denoting PhotoElectron<br />

Action spectroscopy at Constant kinetic Energy.<br />

Our best resolution so far is 1 meV for 10 meV electrons. Below we demonstrate a PEACE<br />

spectrum <strong>of</strong> HgCl- together with the corresponding simulated theoretical spectrum.<br />

References:<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0-12<br />

1-13<br />

Simulated Spectrum<br />

Observed Spectrum<br />

27000 27500 28000 28500 29000 29500 30000 30500<br />

Energy [cm -1 ]<br />

[1] T.N. Kitsopoulos, I. M. Walker, J.G. Loser, D.M. Neumark, Chem. Phys. Letters 159 300 (1989)<br />

[2] Osterwalder, M. J. Nee, J. Zhou, and D. M. Neumark, J. Chem. Phys. 121, 6317 (2004).<br />

[3] R. Giniger T. Hippler, S. Ronen, O. Cheshnovsky, Rev. Sci. Instrum. 72, 2543 (2001)<br />

93


B - 46<br />

94<br />

The effect <strong>of</strong> cluster-gas collisions on the averages <strong>of</strong> charge, energy<br />

and mass <strong>of</strong> an Ar gas cluster ion beam and implications for cluster<br />

induced surface modifications<br />

D. R. Swenson<br />

Epion Corporation, Billerica, MA 01833 USA<br />

Cluster-gas collisions have been the subject <strong>of</strong> several recent papers, but there have been no<br />

studies for large, multi-charged clusters that are typical <strong>of</strong> commercial Gas Cluster Ion Beams<br />

(GCIB). For such large clusters the standard techniques are not adequate because they measure<br />

m/q ratios leaving the mass and charge ambiguous. In a recent paper I reported the first<br />

measurements <strong>of</strong> charge state q, energy E, and mass m for very large Ar clusters as are<br />

commonly used in GCIB processing <strong>of</strong> surfaces [1]. In that paper the average charge <strong>of</strong> the<br />

I<br />

clusters was determined by q = , where I is the electrical current measured by a Faraday<br />

αeΓ<br />

cup, Γ is the flow rate <strong>of</strong> particles in the beam, as measured using a modified Daly detector and<br />

applying particle counting techniques, and α was the measured relative detection efficiency.<br />

The experimental apparatus also included TOF and electrostatic spectrometer measurements, all<br />

for the same pico-amphere-level sample <strong>of</strong> a high intensity Ar GCIB that was accelerated with<br />

30 kV accel potential. This so named “QEM” technique was used to look for the presence <strong>of</strong><br />

multiply charged clusters in the beam and to study the effect <strong>of</strong> cluster-gas collisions. The data<br />

from the experiment consists <strong>of</strong> measurements <strong>of</strong> q , and spectra <strong>of</strong> velocity v and E/q that were<br />

measured for various thicknesses <strong>of</strong> Ar gas that was introduced into a cell immediately<br />

upstream <strong>of</strong> the instruments. The averages v and { E q}<br />

ave are determined from the spectra,<br />

2E<br />

and averages <strong>of</strong> cluster mass and energy are given by E = q{<br />

E q}<br />

ave and by m = . The 2<br />

v<br />

derived averages are strictly valid if the various distributions are uncorrelated. In this paper I<br />

further analyze the data from this experiment and show that a simple analytical model can<br />

predict the main features <strong>of</strong> the data. A more detailed Monte-Carlo simulation is also compared<br />

to the results and to the analytical model, and it is used investigate the limitations <strong>of</strong> the<br />

experimental technique, particularly to what degree correlations affect the average values<br />

derived using the QEM technique. The clusters studied had averages <strong>of</strong> +3.2 charges, 64 keV<br />

energy, and 10,400 atoms at the lowest gas target thickness and were progressively abraded as<br />

the target thickness was increased. It is shown that the mass loss is consistent with a simple<br />

theory that assumes thermalization <strong>of</strong> the collision energy followed by evaporation.<br />

Surprisingly, the clusters lose mass at a greater rate than charge and the N/q ratio decreases as a<br />

result <strong>of</strong> gas collisions. The measurements are important for cluster-surface interactions because<br />

they allow cluster energy and dose to be measured for multi-charged clusters. The theory<br />

developed can be used to understand, predict and optimize the effect <strong>of</strong> gas-cluster collisions on<br />

cluster-surface processing.<br />

References<br />

[1] D.R. Swenson, Nucl. Instr. and Meth. B 222 (2004) 61


B - 47<br />

Computational Electron Spectroscopy – A Powerful Tool for Studies <strong>of</strong><br />

Clusters<br />

Julius Jellinek<br />

Chemistry Division, Argonne National Laboratory, Argonne, Illinois 60439, USA<br />

A new highly accurate scheme for computation <strong>of</strong> the electron energy spectra within the<br />

density functional theory [1] will be sketched, and the results <strong>of</strong> its application to one- and twocomponent<br />

metal clusters will be presented and analyzed. The analysis will include the<br />

evolution <strong>of</strong> the spectra with the cluster size [2-4], structure [4-5], charge state [2-4], and<br />

composition [6]. The computed results will be compared with the available experimental data<br />

[7,8], and their role in the interpretation <strong>of</strong> these data will be illustrated. The discussion will<br />

include analysis <strong>of</strong> the phenomenon <strong>of</strong> size-induced nonmetal-to-metal transition, which defines<br />

the limits <strong>of</strong> miniaturization in nanoelectronics, and the possible relationship between this<br />

phenomenon and the photonic and catalytic properties <strong>of</strong> clusters/nanoparticles.<br />

References<br />

[1] J. Jellinek and P. H. Acioli, J. Chem. Phys. 118, 7783 (2003).<br />

[2] P. H. Acioli and J. Jellinek, Phys. Rev. Lett. 89, 213402 (2002).<br />

[3] J. Jellinek and P. H. Acioli, J. Phys. Chem. A 106, 10919 (2002); 107, 1670 (2003).<br />

[4] J. Jellinek and P. H. Acioli, in Metal Ligand Interaction in Molecular-, Nano- and Macro-Systems<br />

in Complex Environments, N. Russo, D. R. Salahub, and M. Witko (Eds.), Kluwer Academic<br />

Publishers, Dordrecht, 2003, p. 121.<br />

[5] P. H. Acioli and J. Jellinek, Eur. Phys. J. D 24, 27 (2003).<br />

[6] P. H. Acioli and J. Jellinek, to be published.<br />

[7] O. C. Thomas, W. Zheng, S. Xu, and K. Bowen, Phys. Rev. Lett. 89, 213403 (2002).<br />

[8] L.-S. Wang et al., to be published.<br />

95


B - 48<br />

96<br />

Carbon clusters for the calibration in mass spectrometry<br />

Alexander Herlert 1,2 , Sudarshan Baruah 1 , Klaus Blaum 3,4 , Michael Block 4 ,<br />

Ankur Chaudhuri 1 , Frank Herfurth 4 , Alban Kellerbauer 2 , H.-Jürgen Kluge 4 ,<br />

Gerrit Marx 1 , Mannas Mukherjee 4 , Lutz Schweikhard 1 , and Chabouh Yazidjian 2,4<br />

for the ISOLTRAP and SHIPTRAP collaboration<br />

1 Institut für Physik, Ernst-Moritz-Arndt-<strong>Universität</strong> Greifswald, 17487 Greifswald, Germany<br />

2 CERN, Department <strong>of</strong> Physics, 1211 Geneva 23, Switzerland<br />

3 Institut für Physik, Johannes Gutenberg-<strong>Universität</strong> Mainz, 55099 Mainz, Germany<br />

4 GSI, Planckstr. 1, 64291 Darmstadt, Germany<br />

The masses <strong>of</strong> radionuclides are <strong>of</strong> importance in a wide range <strong>of</strong> physics branches [1]:<br />

From nuclear structure studies to tests <strong>of</strong> the Standard model. Penning traps have proven to<br />

allow the mass determination <strong>of</strong> short-lived nuclides with a relative mass uncertainty <strong>of</strong> less<br />

than δm/m=1·10 -8 [2,3] and in case <strong>of</strong> stable nuclides even below 2·10 -11 [4,5].<br />

The principle <strong>of</strong> accurate mass measurements is based on the precise determination <strong>of</strong> the<br />

cyclotron frequency νc=qB/(2πm) <strong>of</strong> the stored ions in the trap. The experimental scheme<br />

monitors the gain in kinetic energy <strong>of</strong> the radial ion motion after excitation with a quadrupolar<br />

radi<strong>of</strong>requency field [6]. For the determination <strong>of</strong> the ion mass, the magnetic field needs to be<br />

calibrated. This is achieved with the measurement <strong>of</strong> the cyclotron frequency <strong>of</strong> an ion with<br />

well-known mass. Here, carbon clusters are the reference particles <strong>of</strong> choice. First, the mass <strong>of</strong> a<br />

12 C atom defines the atomic mass unit. Second, with respect to their mass, carbon clusters Cn are<br />

distributed like a grid on the chart <strong>of</strong> nuclides, where the maximum distance <strong>of</strong> any nuclide to a<br />

carbon cluster is six atomic masses.<br />

In addition, cross-reference measurements allow consistency checks. Thus, carbon clusters<br />

have been used to determine the systematic uncertainties in the mass determination <strong>of</strong> very<br />

short-lived nuclides [7] at the triple trap mass spectrometer ISOLTRAP [8] at ISOLDE/CERN.<br />

For these studies the clusters were produced by laser ablation from a Fullerene target [9]. A new<br />

carbon cluster ion source has recently been installed and will be tested for the upcoming beamtime<br />

period at ISOLDE/CERN. A similar cluster ion source is presently being installed at the<br />

mass spectrometer SHIPTRAP [10] at GSI, which is devoted to mass spectrometry <strong>of</strong> superheavy<br />

nuclides that are delivered from the mass separator SHIP. First results are presented.<br />

References<br />

[1] D. Lunney et al., Rev. Mod. Phys. 75, 1021 (2003).<br />

[2] G. Bollen, Eur. Phys. J. A 15, 237 (2002).<br />

[3] M. Mukherjee et al., Phys. Rev. Lett. 93, 150801 (2004).<br />

[4] R.S. van Dyck et al., Phys. Rev. Lett. 92, 220802 (2004).<br />

[5] S. Rainville et al., Science 303, 334 (2004).<br />

[6] M. König et al., Int. J. Mass Spectrom.Ion Proc. 142, 95 (1995).<br />

[7] A. Kellerbauer et al., Eur. Phys. J. D 22, 53 (2003).<br />

[8] F. Herfurth et al., J. Phys. B: At. Mol. Opt. Phys. 36, 391 (2002).<br />

[9] K. Blaum et al., Anal. Bioanal. Chem. 377, 1133 (2003).<br />

[10] G. Marx et al., Hyperfine Interact. 146/147, 245 (2003).


Photoionisation using Kohn-Sham wave functions<br />

Michael Walter and Hannu Häkkinen<br />

Department <strong>of</strong> Physics, Nanoscience Center, University <strong>of</strong> Jyväskylä, Finland<br />

B - 49<br />

The determination <strong>of</strong> photoionisation cross sections using the Kohn-Sham wave functions<br />

from accurate DFT calculations is studied. The continuum electron is described by an analytical<br />

wave function and the matrix elements are calculated in both in length and velocity form <strong>of</strong> the<br />

dipole operator. The test case <strong>of</strong> water shows excellent agreement between the calculations and<br />

experiment in the ionisation energies if a vertical transition is considered. The cross sections<br />

agree well with the experiment in particular in case <strong>of</strong> the velocity form results. Differences to<br />

the length form and deviations from the experimental asymmetry parameters indicate the<br />

limitations in an analytical description <strong>of</strong> the continuum [3].<br />

Figure 1. State resolved photoionisation cross sections in velocity (full line) and length form (dotted line) for the two<br />

states with the lowest binding energy <strong>of</strong> water. Experimental data from [1] (boxes and triangles) and [2] (circles).<br />

References<br />

[1] M. S. Banna et al, J. Chem. Phys. 84, 4789 (1986).<br />

[2] C. M. Truesdale et al, J. Chem. Phys. 76, 860 (1982).<br />

[3] M.Walter and H. Häkkinen, submitted to Phys. Rev. A<br />

97


B - 50<br />

98<br />

Investigation <strong>of</strong> cluster ions in a Paul trap<br />

Franklin Martinez 1 , Alexander Herlert 1,2 , Gerrit Marx 1 , Hagen Ritter 1 , Lutz Schweikhard 1<br />

1 Institut für Physik, Ernst-Moritz-Arndt-<strong>Universität</strong> Greifswald, 17487 Greifswald, Germany<br />

2 CERN, Department <strong>of</strong> Physics, 1211 Geneva 23, Switzerland<br />

The storage <strong>of</strong> cluster ions in a Penning trap has proven to allow various possibilities <strong>of</strong><br />

experimental investigation [1,2,3,4]. The advantage <strong>of</strong> only an upper but no lower mass limit<br />

and thus trapping <strong>of</strong> particles within a large mass range (even electrons and comparatively<br />

heavy cluster ions) has to be paid by the limitation to only one charge-state polarity and a<br />

reduced access to the stored ions since the Penning trap is mounted inside the bore <strong>of</strong> a<br />

superconducting magnet.<br />

The Paul trap or radi<strong>of</strong>requency trap does not need the strong magnetic field but uses a timevarying<br />

quadrupolar electric field. This allows the storage <strong>of</strong> positively and negatively charged<br />

particles at the same time. In addition the rf trap can be constructed very compact and with<br />

relatively open trap electrodes and therefore axial as well as radial access to the trap is relatively<br />

easy. As an example, a Paul trap has been successfully applied by Parks and coworkers for<br />

electron-diffraction measurements [5].<br />

In the framework <strong>of</strong> experiments on stored clusters similar to the ClusterTrap measurements [1],<br />

the application <strong>of</strong> a Paul trap for future cluster studies is investigated. In a first study the<br />

storage, manipulation and detection <strong>of</strong> atomic clusters are tested. The trapping <strong>of</strong> fullerenes <strong>of</strong><br />

either charge state and their interaction with electrons, ions, neutral atoms as well as photons<br />

will be investigated. The new experimental setup and first results will be presented.<br />

References<br />

[1] L. Schweikhard et al., Eur. Phys. J. D 24, 137 (2003).<br />

[2] C. Walther et al., Phys. Rev. Lett. 83, 3816 (1999).<br />

[3] C. Yannouleas et al., Phys. Rev. Lett. 86, 2996 (2001).<br />

[4] M. Vogel et al., Phys. Rev. Lett. 87, 013401 (2001).<br />

[5] S. Krückeberg et al., Phys. Rev. Lett. 85, 4494 (2000).


Structural characterization <strong>of</strong> large molecules and clusters by<br />

phot<strong>of</strong>ragmentation spectroscopy and scattering experiments.<br />

Michalis Velegrakis<br />

Foundation for Research & Technology – Hellas (FORTH),<br />

Institute <strong>of</strong> Electronic Structure and Laser (IESL)<br />

P.O. Box 1527, Heraklion, GR 71110, Crete<br />

B - 51<br />

The rapid progress in molecular beam techniques, in combination with lasers in the last<br />

years, allowed the production <strong>of</strong> novel molecules in the gas phase such as atomic clusters, metal<br />

oxides and complexes between metal ions and biomolecules. There is growing interest in<br />

obtaining structural information for these molecules.<br />

In this contribution, recent results obtained in the Molecular Dynamics & Clusters laboratory at<br />

IESL are presented. Using a molecular beam apparatus equipped with a time–<strong>of</strong>-flight massspectrometer,<br />

charged clusters are produced in an ion-source which is based on the combination<br />

<strong>of</strong> laser ablation and supersonic gas-expansion.<br />

Mass-spectroscopy, phot<strong>of</strong>ragmentation spectroscopy <strong>of</strong> mass-selected ions and molecular<br />

beam scattering techniques with the aid <strong>of</strong> theoretical models are employed in order to infer the<br />

stability and the structure <strong>of</strong> these molecules.<br />

99


Metals<br />

101


A - 0<br />

102


Size, charge, and isomer specific vibrational spectroscopy <strong>of</strong> isolated<br />

metal clusters<br />

André Fielicke 1 , Christian Ratsch 1,2 , Gert von Helden 1 , and Gerard Meijer 1<br />

1 Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, D-14195 Berlin, Germany<br />

2 Department <strong>of</strong> Mathematics, UCLA, Los Angeles, CA 90095-1555, USA<br />

A - 1<br />

We report on the vibrational spectra <strong>of</strong> neutral and charged metal clusters in the far infrared.<br />

These spectra are obtained via far infrared resonance enhanced multiple photon dissociation<br />

(FIR-MPD) <strong>of</strong> the complexes <strong>of</strong> metal clusters with rare gas atoms. The experiments make use<br />

<strong>of</strong> the Free Electron Laser for Infrared eXperiments FELIX in Nieuwegein, The Netherlands, as<br />

an intense and widely tunable far-infrared radiation source.<br />

The measured FIR-MPD spectra <strong>of</strong> the complexes represent the infrared absorption spectra <strong>of</strong><br />

the bare metal clusters. These spectra are unique for each cluster size and are true fingerprints <strong>of</strong><br />

the cluster’s structure. This FIR-MPD technique has been applied to cationic vanadium clusters<br />

[1] and cationic and neutral niobium clusters containing 3 to more than 20 atoms. Interestingly,<br />

for some cluster sizes, the vibrational spectra <strong>of</strong> neutral and cationic niobium clusters differ<br />

rather strongly, indicating different geometric structures. For smaller sized clusters (n


A - 2<br />

104<br />

Gold Clusters with Density-Functional Based Tight-Binding Method<br />

Pekka Koskinen 1 , Hannu Häkkinen 1 , Gotthard Seifert 2 , and Michael Moseler 3<br />

1 Department <strong>of</strong> Physics, NanoScience Center, 40014 University <strong>of</strong> Jyväskylä, Finland<br />

2 Institut für Physikalische Chemie, TU Dresden, Mommsenstrasse. 13, 01062 Dresden, Germany<br />

3 Fraunh<strong>of</strong>er Institut für Werkst<strong>of</strong>fmechanik, Wöhlerstrasse 11, 79108 Freiburg, Germany<br />

In this work we employ the density-functional based tight-binding method (DFTB)[1] in the<br />

study <strong>of</strong> structures and energies <strong>of</strong> neutral and charged gold clusters. These clusters have been<br />

studied extensively with the density-functional approach (DFT), but for clusters <strong>of</strong> larger size or<br />

for molecular dynamics simulations <strong>of</strong> more complex structures one should have a more<br />

efficient yet accurate computational tool. In this work we demonstrate that this efficient method<br />

might indeed be a useful tool in the study <strong>of</strong> gold clusters.<br />

As an example, Fig. 1 shows the reproducibility <strong>of</strong> the relative energies <strong>of</strong> the isomers <strong>of</strong> Au - 7clusters,<br />

as compared with previous DFT studies [2]. Even better than the energy, the method<br />

gives mainly the same stable cluster geometries as density-functional calculations. The<br />

electronic structure is well described, as can be seen in the vertical detachment energies in Fig 2.<br />

The DFTB-method is able to show that the structure <strong>of</strong> the ground state <strong>of</strong> small gold anions<br />

transforms from two-dimensional to three-dimensional at approximately correct cluster size. We<br />

studied larger Au - 55-cluster isomers and found the same stable structures compared to previous<br />

studies, as well as reasonable isomer energies [3].<br />

Fig.1 Energies and structures <strong>of</strong> a few Au - 7-clusters.<br />

The energies are calculated with self-consistent<br />

charge (SCC) DFTB as well as with DFT and plotted<br />

relative to the corresponding ground state.<br />

References<br />

Fig. 2 The vertical detachment energies <strong>of</strong> the Au - 7clusters<br />

shown. The values are consistently lower<br />

than gradient-corrected DFT (GGA), and the error is<br />

<strong>of</strong> the same order than with local-density DFT<br />

(LDA).<br />

[1] M. Elstner, D. Porezag, G. Jungnickel, J. Elstner, M Haug, Th. Frauenheim, S. Suhai, and G.<br />

Seifert, Phys. Rev. B 58, 7260 (1998).<br />

[2] H. Häkkinen, M. Moseler, and U. Landman, Phys. Rev. Lett. 89, 033401 (2002).<br />

[3] H. Häkkinen, M. Moseler, O. Kostko, N. Morgner, M.A. H<strong>of</strong>fmann, and B.v.Issendorff, Phys. Rev.<br />

Lett. 93, 093401 (2004)


A - 3<br />

Structure, energetics, and thermodynamics <strong>of</strong> copper, nickel, and gold<br />

clusters: N = 2–150<br />

V. G. Grigoryan, D. Alamanova, and M. Springborg<br />

<strong>Universität</strong> des Saarlandes, Physikalische und Theoretische Chemie, Postfach 151150<br />

D-66041 Saarbrücken, Germany<br />

The three most stable structures <strong>of</strong> CuN, NiN, and AuN clusters with N between 2 and 150<br />

have been determined using a combination <strong>of</strong> the Embedded-Atom (EAM), the quasi-Newton,<br />

and our own [1-3] Aufbau/Abbau methods for calculations <strong>of</strong> the total energies, local and global<br />

minima, accordingly. We have employed two well-known versions <strong>of</strong> the EAM: (1) the bulk<br />

version <strong>of</strong> Daw, Baskes and Foiles and (2) the Voter-Chen version which takes into account<br />

properties <strong>of</strong> the dimer. The lower-energy structures (also for the smallest ones) <strong>of</strong> CuN and NiN<br />

clusters (structural details, symmetry and the ordering <strong>of</strong> the isomers) obtained with the two<br />

versions are almost the same. Thus our study points to the universality <strong>of</strong> the bulk embedding<br />

functions and potentials for copper and nickel at least for computations <strong>of</strong> structural properties.<br />

But the bulk EAM fails completely to describe the structures <strong>of</strong> the smallest Au clusters. Hence<br />

for the gold clusters only the Voter-Chen version <strong>of</strong> the EAM has been used. Copper and nickel<br />

grow qualitatively similarly. Generally, we find a multilayered icosahedral or layer-by-layer<br />

growth. But the cluster growth pattern between two closed-shell icosahedra is not simple. It is<br />

predominantly icosahedral with islands <strong>of</strong> fcc, tetrahedral, and decahedral growth. Further, we<br />

have observed an enhanced ability <strong>of</strong> fcc clusters to compete with icosahedral and decahedral<br />

structures in the vicinity <strong>of</strong> N=79. The truncated centered octahedron at N=79 is the first and the<br />

second lower-energy structures for Cu and Ni clusters, respectively. The derived results are<br />

detailed discussed and compared with those <strong>of</strong> available ab initio and semiempirical studies.<br />

Agreement with experimental studies is very good in many cases. The situation for gold is more<br />

complicated. We could not detect any favored structural motif. For example, Au13 is the first<br />

Mackay icosahedron, Au38 is the truncated octahedron, but all the three isomers for Au55 are<br />

low-symmetric. Our result for 38-atom gold cluster agrees with the many-body Sutton-Chen and<br />

Murrell-Mottram calculations, but it is in contradiction with the 'disordered' result <strong>of</strong> the n-body<br />

Gupta potential. Hence more studies are needed. Furthermore, we determine the full vibrational<br />

spectrum and the vibrational heat capacities (contributions from rotations are negligible) <strong>of</strong><br />

clusters for each cluster size. It is found, that the 23-atom Cu or Ni cluster (triple icosahedron)<br />

possesses the highest vibrational frequency. Further it is shown that the highest vibrational<br />

frequency is lower in the fcc and decahedral clusters and higher in the icosahedral ones. The<br />

heat-capacity plots for low temperatures show very interesting fine structures with minima at<br />

magic numbers <strong>of</strong> 13 and 55 and with a broad maximum in the vicinity <strong>of</strong> N=35.<br />

References<br />

[1] V. G. Grigoryan and M. Springborg, Phys. Chem. Chem. Phys. 3, 5125 (2001).<br />

[2] V. G. Grigoryan and M. Springborg, Chem. Phys. Lett. 375, 219 (2003).<br />

[3] V. G. Grigoryan and M. Springborg, Phys. Rev. B 70, 205415 (2004).<br />

105


A - 4<br />

Structure Determination <strong>of</strong> Small Silver Cluster Ions by Trapped Ion<br />

Electron Diffraction<br />

106<br />

Detlef Schooss 1 , Martine N. Blom 1 , Lars Walter 1 , Mattias Kordel 1 , Jason Stairs 1 , Joel H. Parks 2 and<br />

Manfred M. Kappes 1,3<br />

1 Institute <strong>of</strong> Nanotechnology, Forschungszentrum Karlsruhe,D-76021 Karlsruhe, Germany<br />

2 The Rowland Institute at Harvard, Cambridge, MA 02142, USA<br />

3 Institut für Physikalische Chemie, <strong>Universität</strong> Karlsruhe, D-76128 Karlsruhe, Germany<br />

The structure <strong>of</strong> silver clusters ions is studied by the recently developed technique <strong>of</strong><br />

trapped ion electron diffraction (TIED) [1]. In brief, cluster ions are generated by a magnetron<br />

sputter source [2] and injected into a Paul ion trap. After mass selection and thermalization the<br />

trapped ions are irradiated with a 40 keV electron beam and the resulting diffraction pattern is<br />

integrated with a CCD detector. The sensitivity and resolution <strong>of</strong> the apparatus is documented<br />

by figure 1 which shows the molecular scattering function obtained for Ag55 + at 90K.<br />

+<br />

Figure 1: Experimental diffraction pattern (open squares) <strong>of</strong> mass selected Ag55 at 90K. The measurement is<br />

consistent with an icosahedral structure as calculated from RIDFT calculations (line).<br />

Comparison to RIDFT structure calculations for various trial geometries reveals that the cluster<br />

ensemble which was probed comprises predominantly icosahedral Ag55 + . TIED measurements<br />

at different cluster sizes and charge states (Ag19 +/- to Ag79 +/- ) at a variety <strong>of</strong> temperatures ranging<br />

from 100 - 650 K are presently ongoing.<br />

References<br />

[1] M. Maier-Borst, D. C. Cameron, M. Rokni and J. H. Parks, Phys. Rev. A, 59, R3162 (1998), S.<br />

Krückeberg, D. Schooss, M. Maier-Borst, J. H. Parks, Phys. Rev. Lett. 85(21), 4494-4497 (2000)<br />

[2] H. Haberland, M. Mall, M. Moseler, Y. Qiang, T. Reiners and Y. Thurner, J. Vac, Sci. Technol.<br />

A 12(5), 2925 (1994)


Metal Cluster Anions Produced by Attachment <strong>of</strong> Slow Electrons:<br />

Readjustment and Blurring <strong>of</strong> the Magic Numbers<br />

R. Rabinovitch, R. Moro, C. Xia, V. V. Kresin<br />

Dept. <strong>of</strong> Physics, University <strong>of</strong> Southern California, Los Angeles, California 90089-0484, USA<br />

A - 5<br />

The high electric polarizabilities <strong>of</strong> metal clusters enable them to attach low-energy<br />

electrons with very large cross sections [1]: a passing electron is captured by a strong longrange<br />

polarization potential. But very little is understood about the last stage <strong>of</strong> the collision<br />

process: where and how fast is the captured electron’s energy deposited? The questions are<br />

closely related to the general problem <strong>of</strong> electron relaxation in size-quantized nanoscale<br />

systems.<br />

To explore this issue, we are carrying out measurements <strong>of</strong> the mass spectra <strong>of</strong> negative sodium<br />

cluster ions born in the electron-cluster interaction region. If the energy deposited by the<br />

captured electron is quickly thermalized and is sufficient to cause rapid cluster evaporation,<br />

there should result - a rearrangement <strong>of</strong> the cluster abundances and a shift <strong>of</strong> the magic numbers<br />

Na<br />

from Nan to n-1 .<br />

In our experiment such a shift is clearly observed at shell closings near n=20 and n=40.<br />

However, near n=58 and n=92 the shell closings become completely blurred (see Figure 1). This<br />

interesting change may be due to a bottleneck in the electron relaxation and/or to insufficiently<br />

fast evaporation. We will present the current data and discuss their implications.<br />

This work was supported by the U. S. National Science Foundation.<br />

Counts<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

References<br />

Masscan <strong>of</strong> clusters Na18-Na21<br />

Sm oothing<br />

50<br />

380 400 420 440 460 480 500<br />

Mass, AMU<br />

-<br />

Counts<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

Masscan <strong>of</strong> clusters Na56-Na59<br />

Sm oothing<br />

1260 1280 1300 1320 1340 1360 1380<br />

Mass, AMU<br />

Figure 1. Mass spectra <strong>of</strong> Na (left) and Na (right) formed by electron attachment.<br />

18-21<br />

56-59<br />

The shell edge is visible only in the first spectrum.<br />

[1] V.Kasperovich, G.Tikhonov, K.Wong, and V.V.Kresin, Phys. Rev. A 62, 063201 (2000); Phys.<br />

Rev. Lett. 85, 2729 (2000).<br />

-<br />

107


A - 6<br />

108<br />

Mass Spectrometric Investigations <strong>of</strong> Binary Clusters:<br />

Evidence for Endohedral Clusters with Enhanced Stability<br />

Peter Lievens, Ewald Janssens, Sven Neukermans, Xin Wang, Nele Veldeman, Roger E. Silverans<br />

Laboratorium voor Vaste-St<strong>of</strong>fysica en Magnetisme, K.U.Leuven,<br />

Celestijnenlaan 200D, B-3001 Leuven, Belgium<br />

While clusters composed <strong>of</strong> rare gas atoms exhibit enhanced stabilities when they have<br />

magic numbers enabling them to adopt high symmetry geometries ("shells <strong>of</strong> atoms"), the magic<br />

numbers in alkali and other simple metal clusters are determined by the number <strong>of</strong> delocalized<br />

valence electrons occupying a one-particle state in a (centro-)symmetric potential well ("shells<br />

<strong>of</strong> electrons"). The quest for such electronically and geometrically closed shells, which goes<br />

along with enhanced cluster stability, has been a leading thread in cluster research for many<br />

years. One route has been the investigation <strong>of</strong> binary clusters, which allows tailoring both the<br />

cluster geometry (number <strong>of</strong> atoms) and electronic properties (number <strong>of</strong> delocalized electrons)<br />

independently.<br />

We produce beams <strong>of</strong> binary clusters with a dual-target dual-laser vaporization cluster source,<br />

and investigate the clusters in the gas phase with mass spectrometric and laser spectroscopic<br />

techniques. In particular, size and composition dependent stability fluctuations are investigated<br />

with phot<strong>of</strong>ragmentation and mass spectrometry, and size dependent ionization energies are<br />

measured with threshold laser ionization spectroscopy.<br />

Lately we focussed on investigations <strong>of</strong> two types <strong>of</strong> bimetallic systems: clusters consisting <strong>of</strong><br />

noble metals doped with transition metal atoms, and clusters <strong>of</strong> group IVa elements doped with<br />

metal atoms [1-3]. In this contribution we will present evidence for the existence <strong>of</strong> combined<br />

closures <strong>of</strong> shells <strong>of</strong> atoms and shells <strong>of</strong> electrons for specific binary cluster species.<br />

Phenomenological interpretations <strong>of</strong> new electronic shell closures enhanced by size specific<br />

geometry considerations will be given based on comparisons with density functional theory<br />

calculations. In particular the formation <strong>of</strong> stable icosahedral 13 atom systems with the dopant<br />

atom occupying a central position will be highlighted. Also, size dependent magnetic properties<br />

<strong>of</strong> endohedral clusters with a magnetic atom occupying a central position will be discussed.<br />

This work was supported by the Fund for Scientific Research – Flanders (FWO), the Flemish<br />

Concerted Action (GOA) program, and the Belgian Interuniversity Poles <strong>of</strong> Attraction (IAP)<br />

program. EJ and SN are Postdoctoral Researchers <strong>of</strong> the FWO.<br />

References<br />

[1] S. Neukermans, E. Janssens, H. Tanaka, R.E. Silverans, P. Lievens, Phys. Rev. Lett. 90 (2003)<br />

#033401.<br />

[2] E. Janssens, H. Tanaka, S. Neukermans, R.E. Silverans, P. Lievens, New J. Phys. 5 (2003) #46.<br />

[3] S. Neukermans, E. Janssens, Z. Chen, R.E. Silverans, P.v.R. Schleyer, P. Lievens, Phys. Rev. Lett.<br />

92 (2004) #163401.


Photoelectron spectrometry <strong>of</strong> big sodium clusters<br />

Oleg Kostko and Bernd von Issendorff<br />

Fakultät für Physik, <strong>Universität</strong> Freiburg, H.Herderstr. 3, 79104 Freiburg, Germany<br />

A - 7<br />

We have measured photoelectron spectra (PES) <strong>of</strong> Na – n clusters with n=5-4000 at a photon<br />

energy <strong>of</strong> 4.02 eV. For the smaller sizes (n=5-480) a liquid nitrogen cooled rf-octupole trap was<br />

used for thermalization and intensity bunching. Bigger clusters were examined without<br />

additional thermalization. The estimated temperature <strong>of</strong> the clusters in both cases was about<br />

100 K. Comparison <strong>of</strong> the PES obtained this work with data published earlier [1,2] shows that<br />

the lower temperature leads to significantly better spectral resolution and in some cases to a<br />

change <strong>of</strong> the shape <strong>of</strong> the spectra.<br />

In the PES <strong>of</strong> the bigger clusters a periodic vanishing and appearing <strong>of</strong> structure was observed.<br />

For instance Na – 730 and Na – 1250 exhibit almost unstructured spectra, while Na – 670 and Na – 1350<br />

exhibit rather structured spectra. The reason for this behaviour seems to be a combination <strong>of</strong><br />

geometrical shell closing and electronic supershell modulation.<br />

From the measured photoelectron spectra the electron affinities <strong>of</strong> the anions and the ionization<br />

potentials <strong>of</strong> the neutrals were extracted. These values are plotted in Fig.1 as a function <strong>of</strong> N -1/3 .<br />

Their difference yields the cluster charging energy, which can be used to determine the Wigner-<br />

Seitz radius <strong>of</strong> sodium (Rws) and the electron spillout (δ). In our case values <strong>of</strong> 2.07 Å and<br />

0.97 Å were obtained, respectively. The literature value for Rws <strong>of</strong> bulk sodium at 100 K is<br />

2.08 Å. Fitting the electron affinity and ionization potential curves according to the charged<br />

droplet model<br />

2<br />

e<br />

IP = WF + α 1/<br />

3<br />

R N + δ<br />

additionally yields the parameter α=0.377 and the work function WF=2.84 eV, which is slightly<br />

higher then the literature value <strong>of</strong> 2.75 eV.<br />

EA and IP, eV<br />

4<br />

3<br />

2<br />

1<br />

IP<br />

EA<br />

0.0 0.1 0.2 0.3<br />

N<br />

0.4 0.5 0.6<br />

-1/3<br />

Figure 1. Electron affinity and ionization potential as a function <strong>of</strong> the inverse <strong>of</strong> the cubic root <strong>of</strong> the cluster size.<br />

References<br />

2,84 eV<br />

α=0.377<br />

δ=0.974Å<br />

R =2.072Å<br />

ws<br />

[1] G.Wrigge, M.Astruc H<strong>of</strong>fmann and B.v.Issendorff, Phys. Rev. A 65, 063201 (2002).<br />

[2] M.Moseler, B.Huber, H.Häkkinen, U.Landman, G.Wrigge, M.Astruc H<strong>of</strong>fmann and<br />

B.v.Issendorff, Phys. Rev. B 68, 165413 (2003).<br />

WS<br />

109


A - 8<br />

110<br />

Probing the properties <strong>of</strong> polyanionic clusters in a Penning trap<br />

Alexander Herlert 1,2 , Andreas Lassesson 1 , Franklin Martinez 1 , Gerrit Marx 1 ,<br />

Lutz Schweikhard 1 , Noelle Walsh 1<br />

1 Institut für Physik, Ernst-Moritz-Arndt-<strong>Universität</strong> Greifswald, 17487 Greifswald, Germany<br />

2 CERN, Department <strong>of</strong> Physics, 1211 Geneva 23, Switzerland<br />

Highly negatively charged clusters have attracted much interest within the last ten years [1]<br />

and their study has led to the development <strong>of</strong> a new field in cluster research. The addition <strong>of</strong><br />

further electrons to an already negatively charged system is in itself an experimental challenge,<br />

especially with respect to the stability <strong>of</strong> the polyanionic products. Furthermore, the additional<br />

negative charges may lead to a change in the dissociation processes after excitation <strong>of</strong> the highly<br />

charged clusters.<br />

With the application <strong>of</strong> a Penning trap for the simultaneous storage <strong>of</strong> low-energy electrons and<br />

singly-charged metal-cluster anions, the production <strong>of</strong> polyanionic metal clusters is reproducible<br />

under controlled conditions [2]. Size-selected metal cluster dianions as well as trianions are<br />

readily available for further investigation. First experiments concentrated on the relative yield <strong>of</strong><br />

the polyanionic metal clusters as a function <strong>of</strong> cluster size. The results are related to the second<br />

electron affinity <strong>of</strong> the clusters including electronic shell effects [3].<br />

Further investigations involved the excitation <strong>of</strong> the metal cluster dianions with photons and<br />

also by collisions with inert gas atoms. Both methods were already established at the<br />

ClusterTrap experiment [4]. In the case <strong>of</strong> collisional activation, the dianions are brought to a<br />

larger cyclotron radius to increase their kinetic energy. After the collision with argon gas atoms<br />

the emission <strong>of</strong> the surplus electron from the cluster was observed, but the change <strong>of</strong> the massover-charge<br />

ratio led to a change <strong>of</strong> the ion motion and thus to the loss <strong>of</strong> the product ions [5].<br />

For photoexcitation, the dianions were centered in the Penning trap and electron emission as<br />

well as dissociation was observed in case <strong>of</strong> gold-cluster dianions [6]. The data is currently<br />

being analyzed with respect to direct and thermionic electron emission.<br />

In addition to the experiments on polyanionic metal clusters, the investigation <strong>of</strong> highly charged<br />

fullerene anions has started. A new ion source for the ClusterTrap experiment has been<br />

constructed that allows the production <strong>of</strong> both positively and negatively charged fullerenes.<br />

Attachment <strong>of</strong> additional electrons to the singly charged fullerene anions has been successful<br />

and dianions Cn 2- in the size range n ≥ 70 were observed. First results <strong>of</strong> the production and<br />

further investigation <strong>of</strong> the size- and charge-state selected fullerene dianions will be presented.<br />

References<br />

[1] A. Dreuw and L.S. Cederbaum, Chem. Rev. 102, 181 (2002).<br />

[2] A. Herlert and L. Schweikhard, Int. J. Mass Spectrom. 229, 19 (2003).<br />

[3] C. Yannouleas et al., Phys. Rev. Lett. 86, 2996 (2001).<br />

[4] L. Schweikhard et al., Eur. Phys. J. D 24, 137 (2003).<br />

[5] A. Herlert and L. Schweikhard, Int. J. Mass Spectrom. 234, 161 (2004).<br />

[6] L. Schweikhard et al., Int. J. Mass Spectrom. 219, 363 (2002).


Geometric and electronic tructure <strong>of</strong> large sodium cluster anions<br />

Michael Moseler 1,2 , Bernd Huber 2 ,B. von Issendorff 3<br />

1 Fraunh<strong>of</strong>er Institut für Werkst<strong>of</strong>fmechanik, Wöhlerstrasse 11, 79108 Freiburg, Germany<br />

2 Freiburger Materialforschungszentrum , Stefan-Meier-Strasse 21, 79104 Freiburg, Germany<br />

3 Fakultät für Physik, <strong>Universität</strong> Freiburg, Stefan-Meier-Str. 21, 79104 Freiburg, Germany<br />

A - 9<br />

Photoelectron spectroscopy is a frequently used experimental method to extract electronic<br />

binding energies from atomic, molecular, and condensed-matter systems. For molecules and<br />

clusters at low temperature, a comparison <strong>of</strong> the measured photoelectron spectra (PES) with the<br />

electronic density <strong>of</strong> states (DOS) obtained from density functional calculations for T=0 optimal<br />

structures, provides information about the underlying electronic structure and may lead to<br />

assignments <strong>of</strong> pertinent structures.<br />

For the range <strong>of</strong> N=4-19 the PES <strong>of</strong> NaN - clusters were compared with the DOS recorded in our<br />

finite-temperature ab initio simulations [1]. Although the main features <strong>of</strong> the PES can be<br />

understood through the use <strong>of</strong> the jellium model, we are able to detect and explain several<br />

spectral details caused by the finite temperature dynamics. Meanwhile experimental spectra<br />

from very cold cluster beams are available and a comparison with T=0 static spectra reveals<br />

new information concerning the ground states <strong>of</strong> the clusters.<br />

Also larger sodium cluster anions have been measured and simulated by us. For Na55 -, Na147 -<br />

and Na309 - the DOS <strong>of</strong> Mackay-icosahedral clusters was in good agreement with the<br />

corresponding PES (see Fig. 1) and for Na71 - a clear difference between the DOS <strong>of</strong> an anti-<br />

Mackay and an Mackay over layer has been detected.<br />

References<br />

Figure 1. Comparision <strong>of</strong> the Na309 - PES and the DOS <strong>of</strong> the Mackay-icosahedron.<br />

[1] M. Moseler, B. Huber, H. Häkkinen, U. Landman, G. Wrigge, M. Astruc H<strong>of</strong>fmann and B. von<br />

Issendorff, Phys. Rev. B 68, 165413 (2003).<br />

111


A - 10<br />

112<br />

Planar Boron Clusters and 2D to 3D Transitions<br />

Hua-Jin Zhai, 1,2 A. N. Alexandrova, 3 B. Kiran, 1,2 S. Bulusu, 4 Jun Li, 2 Xiao Cheng Zeng, 3<br />

A. I. Boldyrev, 3 and Lai-Sheng Wang 1<br />

1 Department <strong>of</strong> Physics, Washington State University, 2710 University Drive, Richland, WA 99352, USA<br />

and<br />

2 W. R. Wiley Environmental Molecular Sciences Laboratory andChemical Science Division, Pacific<br />

Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, USA.<br />

3 Department <strong>of</strong> Chemistry and Biochemistry, Utah State University, Logan, Utah 84322-0300, USA<br />

4 Department <strong>of</strong> Chemistry and Center for Materials Research and Analysis, University <strong>of</strong> Nebraska-<br />

Lincoln, Lincoln, NE 68588, USA.<br />

One <strong>of</strong> the most interesting features <strong>of</strong> elemental boron and many bimetallic boron<br />

compounds is the occurrence <strong>of</strong> highly symmetric icosahedral clusters. The rich chemistry <strong>of</strong><br />

boron is also dominated by three-dimensional (3D) cage structures. We report our recent<br />

investigations <strong>of</strong> small boron clusters in the size range from 3 to 20 atoms using photoelectron<br />

spectroscopy and ab initio calculations [1-8]. We will present experimental and theoretical<br />

evidence that small boron clusters prefer planar structures and exhibit aromaticity and<br />

antiaromaticity according to the Hückel rules, akin to planar hydrocarbons up to the size <strong>of</strong> 19<br />

atoms. Aromatic boron clusters possess more circular shapes whereas antiaromatic boron<br />

clusters are elongated, analogous to structural distortions <strong>of</strong> antiaromatic hydrocarbons. In a<br />

very recent study [8], we show that 3D structures begin at B20, which has a double ring global<br />

minimum structure. The 2D to 3D structural transition observed at B20 suggests that it may be<br />

considered as the embryo <strong>of</strong> the thinnest single-wall boron nanotubes.<br />

References<br />

B12 B13 B20<br />

[1] H. J. Zhai, L. S. Wang, A. N. Alexandrova, and A. I. Boldyrev, J. Chem. Phys. 117, 7917-7924<br />

(2002).<br />

[2] A. N. Alexandrova, A. I. Boldyrev, H. J. Zhai, L. S. Wang, E. Steiner, and P. W. Fowler, J. Phys.<br />

Chem. A 107, 1359-1369 (2003).<br />

[3] H. J. Zhai, A. N. Alexandrova, K. A. Birch, A. I. Boldyrev, and L. S. Wang, Angew. Chem. Int. Ed.<br />

42, 6004-6008 (2003).<br />

[4] H. J. Zhai, B. Kiran, J. Li, and L. S. Wang, Nature Materials 2, 827-833 (2003).<br />

[5] H. J. Zhai, L. S. Wang, A. N. Alexandrova, A. I. Boldyrev, and V. G. Zakrzewski, J. Phys. Chem.<br />

A 107, 9319-9328 (2003)<br />

[6] A. N. Alexandrova, A. I Boldyrev, H. J. Zhai, and L. S. Wang, J. Phys. Chem. A 108, 3509-3517<br />

(2004).<br />

[7] “Boron Flat Out” (S. K. Ritter), Chem. & Eng. News 82 (9), 28-32 (March 1 (2004).<br />

[8] B. Kiran, S. Bulusu, H. J. Zhai, S. Yoo, X. C. Zeng, and L. S. Wang, Proc. Natl. Acad. Sci. (USA),<br />

in press.


Molecular Clusters<br />

113


A - 10<br />

114


Molecular Beams <strong>of</strong> Organic Molecules<br />

Tim Krause, Adnan Sarfraz, Wolfgang Christen<br />

Institut für Chemie, Humboldt-<strong>Universität</strong> zu Berlin, 12489 Berlin, Germany<br />

A - 11<br />

The transfer <strong>of</strong> non-volatile or thermally unstable molecules into the gas phase is <strong>of</strong> utmost<br />

importance for a variety <strong>of</strong> applications such as analytical mass spectrometry, optical spectroscopy,<br />

or the epitaxial growth <strong>of</strong> thin films. One possible method is the expansion <strong>of</strong> supercritical<br />

solutions into a molecular beam. Here the key idea is to dissolve the substance at high<br />

pressure, exploiting the supercritical fluid's solvent power, and precipitate it by expansion.<br />

Extending our previous work on caffeine [1] and small molecules (vitamin K3, phenylalanine<br />

tertbutylester) [2] we report the successful expansion <strong>of</strong> several molecules <strong>of</strong> biological interest<br />

in supercritical CO2 at temperatures <strong>of</strong> only 310-315 K, and accompanying detection using mass<br />

spectrometry.<br />

References<br />

[1] W. Christen, S. Geggier, S. Grigorenko, K. Rademann, Rev. Sci. Instrum. 75, 5048 (2004).<br />

[2] W. Christen, T. Krause, H. Noack, A. Sarfraz, J. Chem. Phys., to be submitted.<br />

115


A - 12<br />

116<br />

Infrared Spectroscopy <strong>of</strong> CO2 Containing Cluster Anions<br />

Holger Schneider 1 , A. Daniel Boese 2 1, *<br />

, J. Mathias Weber<br />

1 Institut für Physikalische Chemie, <strong>Universität</strong> Karlsruhe, Kaiserstr. 12, 76229 Karlsruhe, Germany<br />

* email: jmathias.weber@chemie.uni-karlsruhe.de<br />

2 Institut für Nanotechnologie, Forschungszentrum Karlsruhe, Pstfach 3640, 76021 Karlsruhe, Germany<br />

Molecular clusters are relevant model systems for the study <strong>of</strong> intermolecular interaction<br />

potentials governing the structures <strong>of</strong> molecular complexes. Structural information on clusters<br />

can be gained from infrared photodissociation spectroscopy in molecular beams [1, 2], and we<br />

have applied this technique to several species, probing the interaction <strong>of</strong> various anions with<br />

CO2. The experimental results are interpreted in the framework <strong>of</strong> density functional theory with<br />

anharmonic calculations <strong>of</strong> vibrational spectra.<br />

Halide-CO2 complexes afford the study <strong>of</strong> the interaction <strong>of</strong> CO2 with closed shell anions <strong>of</strong> low<br />

reactivity [3]. The results are compared with those for Au - ·CO2 [4], which is relevant for the<br />

study <strong>of</strong> catalytic oxidation <strong>of</strong> CO by gold clusters on surfaces [5, 6]. This complex shows<br />

strong interaction between CO2 and the Au - anion, with a strongly deformed CO2 unit.<br />

Complexes <strong>of</strong> O2 - with CO2 show very interesting motifs <strong>of</strong> ion solvation [7]. The π* orbital <strong>of</strong><br />

O2 - <strong>of</strong>fers a template with four a priori equivalent binding sites for the structure <strong>of</strong> the<br />

microsolvation environment. However, only one CO2 ligand is strongly bound, with additional<br />

molecules weakly attached to the resulting CO4 - anion.<br />

The general motif <strong>of</strong> anions binding to CO2 is binding to CO2 via the carbon atom, in contrast to<br />

cations, which generally dock to one <strong>of</strong> the oxygen atoms. The geometry <strong>of</strong> the CO2 molecule is<br />

changed upon complexation with the anion by electrostatic and charge transfer effects. Both<br />

contributions deform the molecule away from the linear configuration <strong>of</strong> neutral CO2 towards<br />

the bent anionic shape.<br />

References<br />

[1] W.H. Robertson and M.A. Johnson, Annu. Rev. Phys. Chem. 54 (2003) 173.<br />

[2] M.A. Duncan, Int. Rev. Phys. Chem. 22 (2003) 407.<br />

[3] J.M. Weber and H. Schneider, J. Chem. Phys. 120 (2004) 10056.<br />

[4] A.D. Boese, H. Schneider, A.N. Gloess, and J.M. Weber, in preparation (2004).<br />

[5] A. Sanchez, et al., J. Phys. Chem. A 103 (1999) 9573.<br />

[6] L.D. Socaciu, et al., J. Am. Chem. Soc. 125 (2003) 10437.<br />

[7] H. Schneider, A.D. Boese, and J.M. Weber, in preparation (2004).


Infrared Spectroscopy <strong>of</strong> XY - ·(C2H2)n Cluster Ions (XY = O2, C2H)<br />

Holger Schneider, J. Mathias Weber *<br />

Institut für Physikalische Chemie, <strong>Universität</strong> Karlsruhe, Kaiserstr. 12, 76229 Karlsruhe, Germany<br />

* corresponding author; email: jmathias.weber@chemie.uni-karlsruhe.de<br />

A - 13<br />

Proton or hydrogen transfer is an important process in chemistry. The intermolecular<br />

interaction potentials governing reactive processes can be studied by vibrational spectroscopy <strong>of</strong><br />

complexes containing the reactants.[1, 2] We report the infrared photodissociation spectra <strong>of</strong><br />

ion-acetylene clusters, with O2 - and C2H - as proton acceptors. The experimental results are<br />

interpreted in the framework <strong>of</strong> density functional theory.<br />

Complexes <strong>of</strong> strong ionic proton acceptors with acetylene molecules can be studied as models<br />

for proton acceptors [3]. All studied complexes are formed by C2H2 molecules docking to the<br />

ions by forming an H-bond. In the case <strong>of</strong> superoxide, each ligand binds to a lobe <strong>of</strong> the π*<br />

orbital <strong>of</strong> O2 - , affording a similar binding motif for all acetylene molecules. In contrast, HC2 - has<br />

only one ionic binding site, and the one <strong>of</strong> the ligands is favoured, while all others are less<br />

strongly bound.<br />

References<br />

[1] W.H. Robertson and M.A. Johnson, Annu. Rev. Phys. Chem. 54 (2003) 173.<br />

[2] M.A. Duncan, Int. Rev. Phys. Chem. 22 (2003) 407.<br />

[3] P. Botschwina, et al., Faraday Discuss. 118 (2001) 433.<br />

117


A - 14<br />

118<br />

Computational Study <strong>of</strong> (H2S) 2≤ n ≤ 10 Clusters<br />

Mehmet Bahat and Ziya Kantarci<br />

Gazi University, Physics Department, 06500, Teknikokullar, Ankara, Turkey.<br />

Molecular clusters are the subject <strong>of</strong> continuing investigations both experimentally and<br />

theoretically in recent years. One <strong>of</strong> them water clusters have been the subject <strong>of</strong> significant<br />

research activity. However, little information for the H2S clusters is available from theoretical<br />

and experimental point <strong>of</strong> view at present. Most calculations have focused on the dimer<br />

structure. Experimentally photoionization studies for the H2S clusters through heptamer have<br />

been studied.<br />

We have studied structures,energies and vibrational spectra <strong>of</strong> (H2S) 2≤ n ≤ 6 clusters<br />

computationally, will be send elsewhere.<br />

In this study, we will present structure, energy, polarizability and hyperpolarizability properties<br />

<strong>of</strong> (H2S) 2≤ n ≤ 10 clusters which calculated using B3LYP/cc-pVTZ level <strong>of</strong> theory.


Structural and spectral features <strong>of</strong> size selected water clusters in the<br />

n=7-21 regime: Results from electronic structure calculations and<br />

empirical potentials<br />

A - 15<br />

George S Fanourgakis 1 , Anita Lagutschenkov 2 , Gereon Niedner-Schatteburg 2 and Sotiris S. Xantheas 1<br />

1 Chemical Sciences Division, Pacific Northwest National Laboratory, 902 Battelle Boulevard, PO Box<br />

999, MS K1-83, Richland, WA 99352, USA<br />

2 Department <strong>of</strong> Chemistry, Chemie, University <strong>of</strong> Kaiserslautern, Erwin Schroedinger Strasse, 67663<br />

Kaiserslautern, Germany<br />

The structural patterns adopted by the clusters <strong>of</strong> the first few molecules <strong>of</strong> water are<br />

determined from the maximization <strong>of</strong> hydrogen bonding, the minimization <strong>of</strong> dangling bonds<br />

and the formation <strong>of</strong> local networks, which exhibit large cooperative effects. To this end, the<br />

global minima <strong>of</strong> the first few water clusters are characterized by structures in which all<br />

molecules are on the surface <strong>of</strong> the cluster. The structural transition between “all-surface” and<br />

“internally solvated” water cluster structures is investigated using high-level ab-initio electronic<br />

structure methods and the results are compared to the ones with empirical potentials. An<br />

informative probing <strong>of</strong> these hydrogen bonding networks is achieved by infrared (IR)<br />

spectroscopy in the 3000 – 4000 cm -1 range, a part <strong>of</strong> the spectra that corresponds to the<br />

“fingerprint” <strong>of</strong> the underlying network and its connectivity. We present the spectral features<br />

that are associated with the structural transition from “all-surface” to “internally solvated”<br />

structures in water clusters. We furthermore rely on the use <strong>of</strong> Schlegel diagrams in order to<br />

identify the most intense, red-shifted OH stretching vibrations which are used in order to probe<br />

the underlying hydrogen bonding network.<br />

119


Phase Transitions<br />

121


A - 15<br />

122


Structural changes in small and medium-size rare-gas cluster cations.<br />

Aleš Vítek, František Karlický, and René Kalus<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Ostrava,30. dubna 22, 701 03 Ostrava, Czech Republic<br />

A - 16<br />

Constant-energy and constant-temperature Monte Carlo simulations have been performed to<br />

localize temperature-dependent structural changes in selected krypton cluster cations, Kr3 + , Kr4 + ,<br />

and Kr12-14 + . The trimer and tetramer cations are used to analyze possible structural changes in<br />

the ionic cores <strong>of</strong> larger KrN + clusters, the Kr12-14 + cations are included as representatives <strong>of</strong><br />

these larger clusters to assess the role <strong>of</strong> solvation effects and to study the phase changes<br />

occurring in neutral solvation shells. The intra-cluster interactions <strong>of</strong> KrN + are described by<br />

means <strong>of</strong> previously developed models based on the diatomics-in-molecules approach with the<br />

inclusion <strong>of</strong> the spin-orbit coupling and the most important three-body polarization interactions.<br />

123


A - 17<br />

124<br />

Experimental study <strong>of</strong> the phase transformations<br />

in molecular clusters <strong>of</strong> Ar and N2<br />

O. G. Danylchenko, S. I. Kovalenko, V. N. Samovarov<br />

B. Verkin Institute for Low Temperature Physics and Engineering, Natl. Academy <strong>of</strong> Sciences <strong>of</strong> Ukraine,<br />

47 Lenin Ave., Kharkov, 61103, Ukraine<br />

We use an electron diffraction technique to study the structure <strong>of</strong> substrate-free Ar and N2<br />

clusters and nanoparticles formed in the process <strong>of</strong> homogeneous nucleation in a supersonic jet<br />

isentropically expanding into vacuum. The average size N <strong>of</strong> aggregations varied between 5⋅10 2<br />

and 1.5⋅10 5 atoms (molecules) per cluster. The temperature <strong>of</strong> clusters and nanoparticles was<br />

T=(38±5) K in the diffraction zone.<br />

Analysis <strong>of</strong> the obtained electron-diffraction patterns has shown for the first time the existence<br />

<strong>of</strong> an amorphous phase in large rare-gas clusters (N≈600 atoms/cluster). As N increased in a<br />

relatively narrow size interval a transition occurred from the amorphous phase to the multilayer<br />

icosahedral one observed for N=1300 atoms/cluster. Formation <strong>of</strong> a crystalline face-centered<br />

cubic (fcc) structure <strong>of</strong> argon with stacking faults took place for N=3⋅10 3 atoms/cluster which is<br />

in good accordance with the previous observations [1,2]. The electron-diffraction patterns <strong>of</strong><br />

bigger clusters (N∼10 4 atoms/cluster) are shown to be characterized by hexagonal close-packed<br />

(hcp) maxima along with the fcc peaks. The intensity <strong>of</strong> the hcp peaks grew with increasing<br />

aggregation size, and for nanoparticles with N∼10 5 atoms/cluster we clearly detected (100),<br />

(101), (103), and (202) peaks characteristic <strong>of</strong> an hcp structure with intensity comparable with<br />

that <strong>of</strong> the fcc peaks.<br />

In case <strong>of</strong> linear molecules like those <strong>of</strong> nitrogen, besides the central interaction between<br />

particles there is a considerable contribution <strong>of</strong> non-central electrostatic forces which are<br />

responsible for the orientational ordering <strong>of</strong> molecules at low temperatures. The structure <strong>of</strong><br />

such clusters can differ significantly from the structure <strong>of</strong> rare-gas clusters depending on the<br />

anisotropic contribution to the total lattice energy. In contrast to amorphous argon clusters, for<br />

nitrogen beams with N≈500 molecules/cluster we first observed the formation <strong>of</strong> aggregations<br />

with multilayer icosahedral packing <strong>of</strong> molecules. The ‘wings’ <strong>of</strong> the most prominent<br />

icosahedral-phase maximum (in the region corresponding to the (100) and (101) peaks <strong>of</strong> the<br />

high-temperature β phase <strong>of</strong> N2, α-β phase transition in a bulk at T=35.6 K) are found to be<br />

curved thus indicating the presence <strong>of</strong> traces <strong>of</strong> the β phase in these clusters. As the cluster size<br />

grows at the same T in the diffraction zone, the diffraction pattern undergoes transformations<br />

and for N∼10 3 molecules/cluster it contains maxima <strong>of</strong> both cubic (α) and hcp (β) phases.<br />

Further cluster growing is accompanied by a decrease in the intensity <strong>of</strong> the cubic-phase peaks<br />

and strenghening <strong>of</strong> the hcp peaks. Clusters with N∼10 4 actually have only the high-temperature<br />

β phase.<br />

We propose and discuss a mechanism underlying the observed structural transformations in Ar<br />

and N2 which supposes cluster formation from a liquid droplet and takes into account the<br />

kinetics <strong>of</strong> phase growth when there are energy barriers between phases.<br />

References<br />

[1] I. Farges, M.F. de Feraudy, B. Roult, G. Torchet, Adv. Chem. Phys. 70, 45 (1988).<br />

[2] O.G. Danylchenko, S.I. Kovalenko, V.N. Samovarov, Low Temp. Phys. 30, 166 (2004).


Molecular Dynamics Simulations <strong>of</strong> the Melting-like Transition in<br />

Homogeneous and Heterogeneous Alkali Clusters<br />

Andrés Aguado and José M. López<br />

A - 18<br />

Department <strong>of</strong> Theoretical Physics, University <strong>of</strong> Valladolid, C/ Real de Burgos s/n, Valladolid, 47011,<br />

Spain<br />

A theoretical analysis <strong>of</strong> the equilibrium geometries and thermal behaviour <strong>of</strong> the impuritydoped<br />

A1Na54 (with A=Li,K,Rb,Cs) clusters, <strong>of</strong> binary Li13Na42 and Na13Cs42 and <strong>of</strong> ternary<br />

Li13Na32Cs12 nanoalloys is presented. The calculations are based on the orbital-free approach to<br />

density functional theory and the classical newtonian equations to deal with the electronic and<br />

ionic subsystems, respectively. The icosahedral symmetry <strong>of</strong> homogeneous Na55 is preserved in<br />

the impurity-doped clusters, with the Li impurity located at the center <strong>of</strong> the icosahedron and<br />

K,Rb and Cs impurities occupying a surface site. The ground state<br />

isomer <strong>of</strong> Na13Cs42 and Li13Na32Cs12 nanoalloys is an onion-like polyicosahedral structure, with<br />

a core shell formed by the element <strong>of</strong> highest surface tension and smallest atomic size, and a<br />

surface shell formed by the atomic species <strong>of</strong> largest size and lowest surface tension. Li13Na42<br />

adopts an amorphous-like structure, albeit with significant polyicosahedral local order, due to<br />

partial mixing <strong>of</strong> Li and Na species in the core-shell. Regarding the thermal properties, the<br />

impurity-doped clusters A1Na54 are found to melt in two well-defined steps upon heating, each<br />

one associated to the activation <strong>of</strong> diffusive motion for atoms <strong>of</strong> a given species. Regarding the<br />

thermal behavior <strong>of</strong> nanoalloys, we have not studied yet explicitly a representative number <strong>of</strong><br />

examples and can not draw very general conclusions. However, our partial results already<br />

indicate that: (1) when the polyicosahedral ordering is not perfect and/or the structure is not<br />

highly compact (as is the case for Li13Na42 and Na13Cs42), premelting effects are more relevant<br />

than in corresponding homogeneous clusters; (2) if<br />

polyicosahedral ordering is strictly preserved (as it<br />

happens for Li13Na32Cs12) the relative stability <strong>of</strong> the<br />

solid-like phase can be significantly enhanced. For<br />

example, the surface <strong>of</strong> this ternary cluster (formed by<br />

Cs atoms) melts at approximately 140 K, which is 50<br />

K above typical surface melting temperatures <strong>of</strong><br />

homogeneous Cs clusters.<br />

If possible, we will complement the results mentioned<br />

in the previous paragraph with those <strong>of</strong> additional<br />

molecular dynamics simulations on homogeneous and<br />

heterogeneous alkali clusters which are still on the<br />

way at the date <strong>of</strong> submission <strong>of</strong> this abstract.<br />

References<br />

Figure 1. Ground state structure <strong>of</strong> Li 13Na 30Cs 12.<br />

[1] A. Aguado, L. E. González, and J. M. López, J. Phys. Chem. B 108, 11722 (2004).<br />

[2] A. Aguado, J. M. López, and S. Núñez, Comp. Mat. Sci. to be published.<br />

[3] A. Aguado and J. M. López, unpublished works.<br />

125


A - 19<br />

126<br />

Phase Changes in Clusters: Targets for Experiments<br />

R. Stephen Berry 1 , and Ana Proykova 2<br />

1 Department <strong>of</strong> Chemistry and The James Franck Institute, The University <strong>of</strong> Chicago, Chicago,<br />

Illinois 60637, USA<br />

2 University <strong>of</strong> S<strong>of</strong>ia, Faculty <strong>of</strong> Physics, 5 James Bourchier Blvd, S<strong>of</strong>ia 1126, Bulgaria<br />

It is now well established that some clusters exhibit the finite-system counterparts <strong>of</strong> firstorder<br />

transitions, and that these differ from bulk phase transitions ins<strong>of</strong>ar as clusters may exist<br />

in two or more phases in observable amounts, in thermal equilibrium. It is also apparent from<br />

simulations that bulk second-order transitions may have small-system analogues that have either<br />

one or two local free-energy minima, corresponding to second-order-like or first-order-like<br />

behavior. These findings raise several questions that await experimental study. One is the<br />

observation <strong>of</strong> sharp upper or (more likely) lower temperature bounds for the local stability <strong>of</strong><br />

the unfavored phase. A second is the possibility <strong>of</strong> observing and distinguishing between cases<br />

in which phases persist long enough to exhibit well-defined, measurable characteristics and<br />

cases in which the dynamic equilibrium between the phases is so labile that no well-defined<br />

distinguishable phases can be observed. Still a third challenge is the possibility <strong>of</strong> observing<br />

phase changes or phase coexistence that look first-order-like in small clusters but become<br />

second-order in the large-system limit. With this challenge is the alternative <strong>of</strong> true secondorder-like<br />

behavior in which multiple phases do not coexist because the free energy has only a<br />

single minimum. All <strong>of</strong> these are experimental investigations that will almost certainly require<br />

monodisperse cluster samples, because the phase behavior <strong>of</strong> small clusters is highly variable<br />

and not at all monotonic with cluster size.


Rare Gases<br />

127


A - 19<br />

128


A - 20<br />

Decay behaviour <strong>of</strong> dimer ions: kinetic energy distributions and 1/t law<br />

J. Fedor 1 , K. Głuch 2 , S. Matt-Leubner 1 , O. Echt 3 , P. Scheier 1 , F. Hagelberg 4 , K. Hansen 5 , J. U. Andersen 6 ,<br />

T. D. Märk 1<br />

1 Institute for Ion Physics, Innsbruck University, Technikerstrasse 25, A-6020 Innsbruck, Austria<br />

2 Institute <strong>of</strong> Mathematics, Physics and Informatics, Maria Curie-Skłodowska University, Lublin 20-031,<br />

Poland<br />

3 Dept. <strong>of</strong> Physics, University <strong>of</strong> New Hampshire, Durham, NH 03824, USA<br />

4 Dept. <strong>of</strong> Physics, Atmospheric Sciences, and General Science, Jackson State University, Jackson, MS<br />

39217, USA<br />

5 Department <strong>of</strong> Physics, Gothenburg University, SE-41296 Gothenburg, Sweden<br />

6 Institute <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Aarhus, DK-8000 Aarhus, Denmark.<br />

There exists a plethora <strong>of</strong> data on metastable decay <strong>of</strong> large cluster ions. For instance, the<br />

analysis <strong>of</strong> the kinetic energy distributions <strong>of</strong> fragment ions has led to the evaluation <strong>of</strong> their<br />

binding energies [1] and the time dependence <strong>of</strong> the decay intensity has been shown to obey a<br />

powerlaw [2]. These decays have been successfully described with statistical theories.<br />

Dimers represent the system where the degree <strong>of</strong> randomization required for thermal statistical<br />

treatment is not possible. Experimental findings on dimer ions however show that they can have<br />

decay characteristics resembling those <strong>of</strong> the large clusters.<br />

Here we present the measurements <strong>of</strong> the kinetic energy release distributions <strong>of</strong> the rare gas<br />

dimer anions (Ne2 + , Ar2 + , Kr2 + , Xe2 + ) using a three sector field mass spectrometer, as well as the<br />

measurements <strong>of</strong> decay intensity time dependence for the dissociation <strong>of</strong> the metal dimer anions<br />

(Cu2 - , Ag2 - ) obtained on the ELISA storage ring in Aarhus. The decay behaviour has been<br />

successfully explained either by the metastable electronic transitions in case <strong>of</strong> rare gases or by<br />

the tunneling through the high angular momentum barrier in the metal anions. The quasicontinuum<br />

<strong>of</strong> rotational states in the later case is sufficiently dense to lead to a 1/t law.<br />

References<br />

[1] K. Gluch, S. Matt-Leubner, O. Echt, R. Deng, J.U. Andersen, P. Scheier and T.D. Märk, Chem.<br />

Phys. Lett. 385 449 (2004)<br />

[2] K.Hansen, J.U.Andersen, P.Hvelplund, S.P.Møller, U.V.Pedersen and V.V.Petrunin,<br />

Phys.Rev.Lett. 87 123401 (2001)<br />

129


A - 21<br />

130<br />

Fragmentation dynamics <strong>of</strong> rare-gas trimer cations.<br />

Ivan Janeček, Daniel Hrivňák, and René Kalus<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Ostrava,30. dubna 22, 701 03 Ostrava, Czech Republic<br />

Hemiquantal, mean-field dynamics and recently developed extended diatomics-inmolecules<br />

models <strong>of</strong> intra-cluster interactions with the inclusion <strong>of</strong> the spin-orbit coupling and<br />

the most important three-body forces are used to study fragmentation <strong>of</strong> argon, krypton and<br />

xenon trimer cations after a sudden ionization <strong>of</strong> respective vibrationally excited neutral trimers.<br />

Both fragmentation from adiabatic and diabatic states is included and the computational results<br />

are compared, where possible, with data from experiments as well as from other theoretical<br />

studies. The main focus is on the role the spin-orbit coupling plays in the relaxation <strong>of</strong> the initial<br />

electronic excitation and during its conversion into cluster vibrational energy. Kinetics <strong>of</strong> the<br />

cluster fragmentation is thoroughly analyzed and recognized to consist <strong>of</strong> two (or even more)<br />

consecutive first-order processes most probably involving the electronic relaxation and the<br />

subsequent cluster decay.


High resolution electron ionization study <strong>of</strong> helium-clusters<br />

S. Denifl 1 , S. Ptasińska 1 , F. Martinez 2 , K. Głuch 3 , S. Feil 1 , P. Scheier 1 , T. D. Märk 1<br />

A - 22<br />

1 Insitut für Ionenphysi,, <strong>Universität</strong> Innsbruck, Technikerstrasse 25, 6020 Innsbruck, Austria<br />

2 Institut für Physik, Ernst Moritz Arndt <strong>Universität</strong> Greifswald, Domstrasse 10a,17487 Greifswald,<br />

Germany<br />

3 Institute <strong>of</strong> Physics, Maria Curie-Sklodowska University, Pl. Marii Curie-Sklodowskiej 1, 20-031 Lublin,<br />

Poland<br />

After many years <strong>of</strong> research the determination and interpretation <strong>of</strong> appearance energies<br />

(AE’s) for clusters using electron impact ionization is still difficult and involves large error bars.<br />

The reasons for this are a number <strong>of</strong> technical obstacles to prepare electrons with high-energy<br />

resolution. In addition we are confronted with a complicated physical situation <strong>of</strong> a reaction<br />

complex involving a quantum mechanical many body system. The use <strong>of</strong> photoionization<br />

sometimes appears to be less difficult but nevertheless the AE values obtained are not directly<br />

comparable to those obtained by electron impact studies. Both processes are intrinsically<br />

different, because the transition complexes in the ionization process are not the same.<br />

This work presents the high resolution study <strong>of</strong> the threshold behavior for small helium cluster<br />

ions. The crossed electron/cluster beams apparatus used for these measurements consists <strong>of</strong> a<br />

cluster source, a hemispherical electron monochromator and a quadrupole mass spectrometer. A<br />

standard home-built hemispherical electron analyzer produces electrons with a typical energy<br />

distribution <strong>of</strong> 150 meV. The helium clusters are produced by supersonic expansion in a cluster<br />

source which can be cooled down to 8.5 K. Previously already all other rare gases (Ne, Ar, Kr,<br />

Xe), H2-, D2-, N2- and N2O-clusters have been studied with the present experimental setup<br />

(except a different cluster source was used). The data evaluation is basing on an extension <strong>of</strong><br />

Wannier’s law for the ionization <strong>of</strong> atoms towards the case <strong>of</strong> clusters. The present results for<br />

helium cluster up to size n = 10 are compared with those <strong>of</strong> previous experiments with small<br />

helium clusters using electron impact [1] or photon impact [2], respectively.<br />

This work was supported by FWF, Wien, Austria, and the European Commission,<br />

Brussels.<br />

References<br />

[1] R. Fröchtenicht, U. Henne, J. P. Toennis, A. Ding, M. Fieber-Erdmann, T. Drewello, J. Chem.<br />

Phys. 104, 2548 (1996).<br />

[2] B. E. Callicoat, K. Förde, L. F. Jung, T. Ruchti, K. C. Janda, J. Chem. Phys. 109, 10195 (1998).<br />

131


A - 23<br />

132<br />

Surface Entropy <strong>of</strong> Argon Clusters<br />

Mikael Kjellberg, Sergey Prasalovich, Vladimir Popok, Klavs Hansen, Eleanor E. B. Campbell<br />

Department <strong>of</strong> Physics, Göteborg University, SE-41296 Göteborg, Sweden<br />

The enhanced stability <strong>of</strong> certain cluster sizes is traditionally inferred from the enhanced<br />

abundance <strong>of</strong> these species in mass spectra. For rare gas clusters, the stability pattern can be<br />

described as due to a close packing <strong>of</strong> atoms into closed structures (shells) with the symmetry <strong>of</strong><br />

Mackay icosahedra. The mere presence <strong>of</strong> shell closings does not give any information on the<br />

magnitude <strong>of</strong> the energies involved. A quantitative comparison requires knowledge <strong>of</strong> the<br />

binding energies <strong>of</strong> the clusters. It is possible to extract this information from the mass spectra<br />

by making reasonable general assumptions [1] with the restriction that only relative values can<br />

be found. The flat abundance spectrum (after background subtraction) is related to relative<br />

I ∝ D + C D − D G.<br />

energies as ( )<br />

N N+ 1 v N N+<br />

1 /<br />

This has been done in Fig 1(A) for masses around the N = 55 “magic number”. The amplitude<br />

in the relative energy variations does not follow the same trend as the local abundance<br />

variations. In particular the maximum in the relative energy is shifted to smaller N. The reason<br />

for this is that the relative energies obtained in this procedure are not the dissociation energies,<br />

DN, <strong>of</strong> the clusters but the free energies, FN = DN – TSN, where SN is identified with the surface<br />

entropy (the natural logarithm <strong>of</strong> the number <strong>of</strong> ways the configurational ground state can be<br />

contructed). This can be estimated from the structures calculated by Northby [2] and used to<br />

extract the dissociation energies. This is shown in Fig. 1(B) and compared with the calculated<br />

dissociation energies, where the absolute scale was obtained by scaling with the bulk and<br />

surface energy. The agreement <strong>of</strong> the trend in the values is very satisfactory.<br />

∆F (meV)<br />

80<br />

75<br />

70<br />

65<br />

60<br />

Figure 1. A, Left:Integrated peak intensities from Ar N + mass spectrum filled squares, (right y-axis) and the<br />

dissociation energies extracted from the spectrum (free energies) (open squares, left y-axis). B, Right: Extracted<br />

dissociation energies (free energies) (open circles), dissociation energies after subtraction <strong>of</strong> the entropy term (open<br />

squares), calculated energies from Northby [2] (filled squares).<br />

References<br />

50 51 52 53 54 55 56 57 58 59 60<br />

0<br />

61<br />

N<br />

[1] K. Hansen and U. Näher, Phys. Rev. A , 60 1240 (1999).<br />

[2] J.A. Northby, J. Xie, D.L. Freeman, J.D. Doll, Z. Phys. D 12 69 (1989)<br />

∆F<br />

I N<br />

3<br />

2<br />

1<br />

I N<br />

Energies [meV]<br />

85<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

45<br />

∆F N,exp<br />

D N,exp<br />

D N,Northby<br />

50 52 54 56 58 60<br />

N


Photodissociation <strong>of</strong> rare-gas trimer cations.<br />

Daniel Hrivňák 1 , René Kalus 1 , Florent X. Gadéa 2<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Ostrava, Czech Republic<br />

2 Groupe NanoScience, CEMES, France<br />

A - 24<br />

Hemiquantal, mean-field dynamics and recently developed extended diatomics-inmolecules<br />

models <strong>of</strong> intra-cluster interactions and electronic transitions in rare-gas cluster<br />

cations are used to study the dissociation <strong>of</strong> argon, krypton and xenon trimer cations after<br />

absorption <strong>of</strong> a photon. The main focus is on the importance <strong>of</strong> the spin-orbit coupling for the<br />

photodissociation dynamics. The theoretical data on fragmentation channels, the kinetic energy<br />

<strong>of</strong> phot<strong>of</strong>ragments and the ratios for the symmetric and asymmetric fragmentation are<br />

thoroughly compared with the experimental data by Haberland’s group. A very good agreement<br />

is obtained. The importance <strong>of</strong> initial vibrational excitations <strong>of</strong> the trimer cations is also<br />

analyzed using a constant-temperature Monte Carlo sampling method. It has been found that the<br />

kinetic energy <strong>of</strong> phot<strong>of</strong>ragments is only negligibly influenced by the initial heating <strong>of</strong> the<br />

trimer, the ratios <strong>of</strong> the symmetric fragmentation change significantly with temperature<br />

however, particularly in the temperature region corresponding to structural changes <strong>of</strong> the<br />

trimers.<br />

133


Semiconductors<br />

135


A - 24<br />

136


A - 25<br />

Theoretical Investigation <strong>of</strong> the Influence <strong>of</strong> Ligands on Structural and<br />

Electronic Properties <strong>of</strong> Indium Phosphide Clusters<br />

Sudip Roy, Michael Springborg<br />

Physikalische Chemie, <strong>Universität</strong> des Saarlandes, 66123 Saarbrücken, Germany<br />

Results <strong>of</strong> a theoretical study <strong>of</strong> the effects <strong>of</strong> including ligands on stoichiometric InnPn<br />

clusters are presented. We apply a parameterized density-functional method 1 and consider<br />

clusters with n up to above 70. As ligands we consider H atoms and CH3 groups, and the results<br />

are compared with our earlier ones for the naked clusters 2 . We find that the ligands lead to only<br />

smaller structural changes, but that an enhanced In-to-P electron transfer in the outermost parts<br />

<strong>of</strong> the clusters, that we observed for the naked clusters, is largely suppressed, so that there is a<br />

more homogeneous In-to-P transfer throughout the whole cluster. Adding the ligands leads in<br />

most cases to an increase in the HOMO-LUMO gap and, therefore, also to an increase in the<br />

stability <strong>of</strong> the clusters. However, we find also that the HOMO-LUMO gap depends critically<br />

on the type, sites, and number <strong>of</strong> the ligands that are added.<br />

References<br />

[1] G. Seifert, D. Porezag, and Th. Frauenheim, Int. J. Quant. Chem., 58, 185 (1996).<br />

[2] S. Roy and M. Springborg, J. Phys. Chem. B , 107, 2771 (2003)<br />

137


A - 26<br />

138<br />

Sizedependence <strong>of</strong> Electronic and Structural Properties <strong>of</strong><br />

Semiconductor Nanoparticles<br />

Michael Springborg, Jan-Ole Joswig, Sudip Roy, Pranab Sarkar<br />

Physical and Theoretical Chemistry, University <strong>of</strong> Saarland, 66123 Saarbrücken, Germany<br />

Using a parameterized density-functional method we have calculated the electronic and<br />

structural properties <strong>of</strong> naked AB semiconductor nanoparticles as a function <strong>of</strong> size with up to<br />

around 200 atoms. We assumed that the structure is related to that <strong>of</strong> a spherical cutout <strong>of</strong> either<br />

the zincblende or the wurtzite crystal structure, which subsequently is allowed to relax to its<br />

closest total-energy minimum. In addition, we have also studied core/shell nanoparticles where<br />

one semiconductor forms the shell on the core <strong>of</strong> another semiconductor. The properties that we<br />

shall discuss include structure, optical gap, charge transfers and stability.


Structural, Electronic and Optical Properties <strong>of</strong><br />

Surface saturated Cadmium Sulfide Clusters<br />

Johannes Frenzel 1 , Gotthard Seifert 1 , Jan-Ole Joswig 2 , Michael Springborg 2<br />

1 Institut für Physikalische Chemie, Technische <strong>Universität</strong> Dresden, D-01062 Dresden, Germany<br />

2 Institut für Physikalische Chemie, <strong>Universität</strong> des Saarlandes, D-66123 Saarbrücken, Germany<br />

A - 27<br />

Due to the quantum-confinement effects the optical properties <strong>of</strong> cadmium sulfide clusters<br />

tune with respect to their size. In this theoretical work clusters up to 4 nm in diameter, derived<br />

from the electronically almost degenerated crystal structures <strong>of</strong> cadmium sulfide, wurtzite and<br />

zinc blende (sphalerite), were studied systematically using a simplified LCAO-DFT-LDA<br />

method (DFTB) 1,2 . With a time-dependent extension (TD-DFRT-TB) 2 <strong>of</strong> this method the optical<br />

spectra <strong>of</strong> bare and surface saturated cadmium sulfide clusters were calculated. Since there is a<br />

strong dependency <strong>of</strong> the lowest unoccupied molecular orbitals on structural properties,<br />

basically on the number single-bonded surface atoms the HOMO/LUMO gap trend to zero for<br />

clusters having a large number <strong>of</strong> single bonded Cadmium atoms 3 . It is shown that effect,<br />

caused by low lying cadmium 5s states, disappears when to these clusters surfactants are added,<br />

namely hydrogen atoms and thiol-groups and their spectra show an asymptotical decreasing <strong>of</strong><br />

the HOMO/LUMO gap toward the bulk band gap <strong>of</strong> CdS while increasing their size.<br />

Figure 1. Structurual shape (left side) and TD-DFRT-TB spectra (right side) <strong>of</strong> a spherical cadmium sulfide clusters,<br />

unsaturated (dashed curve) and saturated (solid curve).The curves are broadened with Gausians (0.1 eV).<br />

References<br />

[1] Porezag, D. et al., Phys. Rev. B 51, 12947 (1995).<br />

[2] Seifert, G. et al., Int. J. Quantum Chem. 58,185 (1996).<br />

[3] Niehaus, T. et al., Phys. Rev. B 63, 085108 (2001).<br />

[4] Joswig, J.-O. et al., J. Phys. Chem B 107, 2897 (2003).<br />

139


Solvations<br />

141


A - 27<br />

142


Ionization Potentials <strong>of</strong> Large Sodium Doped Ammonia Clusters<br />

Christ<strong>of</strong> Steinbach, Udo Buck<br />

Max-Planck-Institut für Strömungsforschung, Bunsenstrasse 10, 37073 Göttingen, Germany<br />

A - 28<br />

In a continuous neat supersonic expansion ammonia clusters are generated and doped with<br />

sodium atoms in a pickup cell. In this way Na(NH3)n clusters are produced and photoionized by<br />

a tunable dye laser system. The ions are mass analyzed in a reflectron time-<strong>of</strong>-flight mass<br />

spectrometer, and the wavelength dependent ion signals serve for the determination <strong>of</strong> the<br />

ionization potentials (IP) <strong>of</strong> the different clusters in the size range 10 < n < 500. For clusters n <<br />

18 the results agree with those published previously. 1 In particular, the plateau for 10 < n


A - 29<br />

Charge and Energy Transfer Processes Induced by Core Ionization <strong>of</strong><br />

Anion-Molecule Microsolvated Clusters<br />

144<br />

Nikolai V. Kryzhevoi, Lorenz S. Cederbaum<br />

Theoretical Chemistry, Institute <strong>of</strong> Physical Chemistry, Heidelberg University,<br />

Im Neuenheimer Feld 229, 69120 Heidelberg, Germany<br />

Cluster physics pertains to exploration <strong>of</strong> various active topics <strong>of</strong> chemical science.<br />

Solvation is among them. Studying microsolvated clusters provides an opportunity to gain<br />

insight into the solvation phenomenon at the microscopic level. Anion-molecule microsolvated<br />

clusters are <strong>of</strong> wide interest due to their unique properties. As well as being a subject <strong>of</strong> diverse<br />

experimental studies, small clusters are also amenable to accurate theoretical treatments from<br />

first principles. The first core level ab initio investigation <strong>of</strong> selected clusters <strong>of</strong> this type are<br />

reported here. The localized character <strong>of</strong> core electrons allows one to probe local properties <strong>of</strong><br />

selected units <strong>of</strong> clusters providing information on the chemical state and interactions <strong>of</strong> each<br />

unit with its local environment. Ionization <strong>of</strong> molecular core levels in anion-molecule clusters<br />

induces, besides local excitations in the molecule itself, a variety <strong>of</strong> charge and energy transfer<br />

processes which are nonlocal in nature. Low-lying satellites appearing in core level spectra due<br />

to these nonlocal processes are apparently not present in spectra <strong>of</strong> free molecules and are a<br />

peculiarity <strong>of</strong> clusters only. Energies and intensities <strong>of</strong> these satellites are found to strongly<br />

depend on the chemical type <strong>of</strong> the anion as well as on the type <strong>of</strong> the ionized solvent molecule.<br />

It is shown that valuable information on properties <strong>of</strong> anion-molecule clusters and their<br />

monomers like the strength <strong>of</strong> anion-molecule interaction and the geometry <strong>of</strong> the clusters can<br />

be obtain by studying charge and energy transfer satellites in core level spectra <strong>of</strong> clusters.


First principles studies on the reaction mechanisms for the intracluster<br />

reactions in Mg + (H2O)n and Al + (H2O)n<br />

Zhifeng Liu<br />

Department <strong>of</strong> Chemistry<br />

Chinese University <strong>of</strong> Hong Kong<br />

Shatin, Hong Kong, China<br />

A - 30<br />

Size dependent reactions in solvation clusters provide invaluable clues to the link<br />

between the microsolvation environment and the reaction mechanisms. The loss <strong>of</strong> a<br />

hydrogen atom in an Mg + (H2O)n and the loss <strong>of</strong> a hydrogen molecule in an Al + (H2O)n<br />

cluster are the classical examples <strong>of</strong> such reactions, which are turned on at a certain size<br />

(5 for Mg + (H2O)n and 12 for Al + (H2O)n) and then turned <strong>of</strong>f at a larger size (17 for<br />

Mg + (H2O)n and 24 for Al + (H2O)n). The properties and structures <strong>of</strong> these clusters have<br />

attracted much attention over the years. In this presentation, I will report the recent<br />

progresses made in my group to understand the mechanisms for these size dependent<br />

reactions by first principles calculations [1, 2, 3]. In the case <strong>of</strong> Al + (H2O)n, the reactivity<br />

is actually due to its isomer (HAlOH) + (H2O)n-1 and is controlled by the cage-like cluster<br />

structure, which is essential to bring a proton near the H--Al bond. In contrast, the<br />

crucial factor in the case <strong>of</strong> Mg + (H2O)n is the solvation <strong>of</strong> the unpaired electron, as the<br />

relative distance between the electron and the Mg 2+ center determines the barrier for the<br />

production <strong>of</strong> the H atom. The details for both clusters provide interesting examples on<br />

the subtle and intricate effects <strong>of</strong> solvation on chemical reactions.<br />

References:<br />

[1] C.K. Siu, Z.F. Liu, and J.S. Tse, “Ab Initio Studies on Al + (H2O)n, HAlOH + (H2O)n-1, and the Size<br />

Dependent H2 Elimination Reaction”, J. Am. Chem. Soc. 124 (2002) 10846-60<br />

[2] C.K. Siu, and Z.F. Liu, “Ab initio Studies on the Mechanism <strong>of</strong> the Size Dependent H Elimination<br />

Reaction in Mg + (H2O)n”, Chem. Eu. J. 8 (2002) 3177-86<br />

[3] C.K. Siu, and Z.F. Liu, “Reaction mechanisms for the size-dependent H loss in Mg + (H2O)n:<br />

solvation controlled electron transfer”, submitted to Phys. Chem. Chem. Phys.<br />

145


Surface<br />

147


148


Alkali Metal Atoms and Clusters on Graphite:<br />

a Density Functional Study<br />

K. Rytkönen 1 , J. Akola 2 , M. Manninen 1<br />

1 Nanoscience Center, P.O. Box 35, FIN-40014 University <strong>of</strong> Jyväskylä, Finland<br />

2 Institut fϋr Festkörperforschung, Forschungszentrum Jϋlich, D-52425 Jϋlich, Germany<br />

A - 31<br />

Sodium atoms and clusters (≤ 5) on graphite (0001) are studied using density functional<br />

theory, pseudopotentials and periodic boundary conditions [1]. A single Na atom is observed to<br />

bind at a hollow site 2.45 Å above the surface with an adsorption energy <strong>of</strong> 0.51 eV. The small<br />

diffusion barrier <strong>of</strong> 0.06 eV indicates a flat potential energy surface. Increased Na coverage<br />

results in a weak adsorbate-substrate interaction, which is evident in the larger separation from<br />

the surface in the cases <strong>of</strong> Na3, Na4, Na5, and the (2×2) Na overlayer. The binding is weak for<br />

Na2, which has a full valence electron shell. The presence <strong>of</strong> substrate modifies the structures <strong>of</strong><br />

Na3, Na4, and Na5 significantly, and both Na4 and Na5 are distorted from planarity. The<br />

calculated formation energies suggest that clustering <strong>of</strong> atoms is energetically favorable, and<br />

that the open shell clusters (e.g. Na3 and Na5) can be more abundant on graphite than in the gas<br />

phase. Analysis <strong>of</strong> the lateral charge density distributions <strong>of</strong> Na and Na3 shows a charge transfer<br />

<strong>of</strong> ~0.5 electrons in both cases. Results for other alkali metals (Li, K, Rb, Cs) will be presented<br />

too.<br />

References<br />

[1] K. Rytkönen, J. Akola, and M. Manninen, Phys. Rev. B 69, 205404 (2004).<br />

149


A - 32<br />

150<br />

Single Atom Dispersion <strong>of</strong> Platinum on Carbon Nanotubes<br />

Yong-Tae Kim 1 , Kazuyoshi Ohshima 1 , Koichi Higashimine 2 , Kazuo Kato 3 , Tomoya Uruga 3 , Keiichi<br />

Osaka 3 , Kenichi Kato 3,4 , Masaki Takada 3,4 , Hiroyoshi Suematsu 3 , Tadaoki Mitani 1<br />

1 School <strong>of</strong> Materials Science and Center for Nano Materials and Technology, Japan Advanced<br />

2 Institute <strong>of</strong> Science and Technology (JAIST), 1-1 Asahidai, Tatsunokuchi, Ishikawa 923-1292, Japan<br />

3 Japan Synchrotron Radiation Research Institute (JASRI)/SPring-8, 1-1-1 Kouto, Mikazuki,<br />

Hyogo 679-5198, Japan<br />

4 Core Research for Evolutional Science and Technology, Japan Science Technology Cooperation<br />

(CREST/JST), 4-1-8 Honcho, Kawaguchi, Saitama 332-0012, Japan<br />

Electrocatalyst is the key factor to determine the performance <strong>of</strong> fuel cell. Generally, it is<br />

prepared by the supporting electrocatalytical active material such as platinum on carbon<br />

materials in order to elevate an activity and reduce the usage <strong>of</strong> novel metal. Carbon supports<br />

therefore affect largely the properties <strong>of</strong> electrocatalyst such as dispersity, size distribution,<br />

particle shape and intrinsic conductivity. Since the discovery <strong>of</strong> carbon nanotubes (CNT) in<br />

1991 by Iijima 1 , the application to electrocatalyst supports has been attempted in order to utilize<br />

their superior properties such as good intrinsic conductivity, adequate specific surface area,<br />

accessible surface texture and electrochemical stability. However, there have been also a<br />

limitation that it is difficult to disperse highly and uniformly electrocatalytical active material on<br />

CNT, because they have the inert and plane surface which is difficult to wet in precursor<br />

solution and easy to cause the agglomeration 2 . In order to solve this problem, we adopted the<br />

introduction <strong>of</strong> thiol groups known to be the functional groups having a good affinity to novel<br />

metal on CNT surfaces. It was expected that high dispersity could be obtained by preventing the<br />

agglomeration among Pt clusters, since surface thiol groups formed the strong bonding with Pt<br />

clusters.<br />

Surprisingly, it was revealed that the surface <strong>of</strong> thiolated carbon nanotubes was covered with the<br />

mono-layer <strong>of</strong> single platinum atoms bonded with the surface thiol groups over our expectation.<br />

This phenomenon was confirmed with several analysis tools, such as X-ray Absorption Fine<br />

Structure analysis (EXAFS, XANES), X-Ray Diffractometry (XRD) with synchrotron radiation<br />

<strong>of</strong> SPring-8, X-ray Photoemission Spectrometry (XPS) and Transmission Electron Microscopy<br />

(TEM).<br />

Here we report the single atom dispersion as the extreme dispersed state <strong>of</strong> metal and the<br />

possibility to control the cluster size uniformly from single atom.<br />

References<br />

O<br />

Pt<br />

S<br />

NH<br />

C<br />

O<br />

Pt<br />

S<br />

NH<br />

C<br />

Figure 1. Schematic diagram <strong>of</strong> single Pt atom dispersion on CNT.<br />

[1] S. Iijima, Nature 354, 56 (1991).<br />

[2] E. Dujardin, T. W. Ebbesen, H. Hiura, K. Tanigaki, Science 265, 1850 (1994).<br />

O<br />

Pt<br />

S<br />

NH<br />

C<br />

O<br />

Pt<br />

S<br />

NH<br />

C<br />

O<br />

Pt<br />

S<br />

NH<br />

C


Self-Assembly <strong>of</strong> Metal Nanoclusters on Oxide Surfaces<br />

N. Berdunov 1 , G. Mariotto 1 , S. Murphy 1 , K. Balakrishnan 1 , Y. M. Mukovskii 2 , I. V. Shvets 1<br />

1 SFI Laboratories, Physics Dept., Trinity College Dublin, Ireland<br />

2 MISIS, Leninsky Prospect 4, Moscow 119991, Russia<br />

A - 33<br />

The self-assembly <strong>of</strong> ordered arrays <strong>of</strong> metal nanoclusters is a fascinating scientific<br />

phenomenon as well as a particularly promising subject for technological applications, such as<br />

microelectronics, ultra-high density recording, corrosion and catalysis. For example, magnetite<br />

and iron play an important role in industrial processes such as the production <strong>of</strong> hydrogen and<br />

the synthesis <strong>of</strong> ammonia. The ability <strong>of</strong> controlling the size and arrangement <strong>of</strong> Fe<br />

nanostructures may influence the way catalytic processes take place.<br />

We have studied the formation <strong>of</strong> Fe nanostructures on the magnetite (111) surface <strong>of</strong> single<br />

crystals and thin films. The Fe3O4(111) surface exhibits an hexagonal 42 Å superstructure, when<br />

annealed in oxygen atmosphere [1]. We have shown that this highly regular pattern is useful as<br />

a template for the self-assembly <strong>of</strong> nanostructures.<br />

We have deposited Fe films <strong>of</strong> 0.2, 0.5, 1 and 2 Å thickness at room temperature by means <strong>of</strong><br />

electron beam evaporation. STM images prove that ordered nucleation <strong>of</strong> nanoclusters takes<br />

place only on the patterned regions <strong>of</strong> the surface, while random nucleation takes place on the<br />

unpatterned regions (see Fig. 1). These results have been reproduced in the case <strong>of</strong> a 0.5 Å Cr<br />

film. The metal nanostructures are characterized by a very regular size distribution over a length<br />

<strong>of</strong> several hundreds nanometers.<br />

Our results demonstrate that the self-assembly <strong>of</strong> crystalline Fe and Cr nanostructures takes<br />

place on the preferential nucleation sites provided by the nanopatterned Fe3O4(111) surface. We<br />

suggest that ordered arrays <strong>of</strong> nanostructures could be grown on different oxides displaying<br />

long-range surface reconstruction, as recently demonstrated by the growth <strong>of</strong> ordered Pd<br />

clusters on Al2O3 [2].<br />

Figure 1. 80 nm × 80 nm STM image <strong>of</strong> 0.5 Å Fe deposited onto the Fe3O4(111) surface. Self-assembly <strong>of</strong> Fe<br />

nanostructures takes place only on the patterned areas <strong>of</strong> the crystal (terrace A), while Fe nucleates randomly onto the<br />

unpatterned areas (terrace B).<br />

References<br />

[1] N. Berdunov, S. Murphy, G. Mariotto and I.V. Shvets, Phys. Rev. B 70, 085404 (2004).<br />

[2] S. Degen, C. Becker and K. Wandelt, Faraday Discuss. 125, 343 (2004).<br />

151


A - 34<br />

152<br />

Supported Gold Clusters and Cluster-Based Nanoparticles:<br />

Characterization, Stability and Growth Studies by In Situ Grazing<br />

Incidence Small Angle X-ray Scattering Technique<br />

Stefan Vajda 1 , Jeffrey W. Elam 2 , Byeongdu Lee 3 , Michael J. Pellin 4 , Sönke Seifert 3 , George Y.<br />

Tikhonov 1 and Nancy A. Tomczyk 1 , Randall E. Winans 1,3<br />

1 Chemistry Division, 2 Energy Systems Division, 3 Experimental Facility Division, and 4 Materials Science<br />

Division<br />

Argonne National Laboratory, 9700 South Cass Avenue, Argonne, Illinois 6043, USA<br />

Supported clusters possess unique, strongly size-dependent properties. However, the<br />

Achilles heal <strong>of</strong> supported particles remains their sintering at elevated temperatures or when<br />

exposed to mixtures <strong>of</strong> reactive gases. In this paper, the issue <strong>of</strong> thermal stability and<br />

agglomeration <strong>of</strong> atomic gold clusters and nanoparticles produced by cluster deposition on<br />

oxide surfaces is addressed as a function <strong>of</strong> deposited cluster size, level <strong>of</strong> surface coverage,<br />

temperature and time <strong>of</strong> heat treatment by synchrotron X-ray radiation at the Advanced Photon<br />

Source.<br />

The continuous beam <strong>of</strong> gold clusters was generated in a laser vaporization source. The Aun +<br />

clusters were deflected into setup with an incorporated mass spectrometer for the pre-selection<br />

<strong>of</strong> narrow cluster distributions in the size range Aun + , n=2,…10, with 1 to 5 dominant sizes. As<br />

support, silica as well as Al2O3 and TiO2 films grown by atomic layer deposition on<br />

SiO2/Si(111) were used. Coverages up to 45% <strong>of</strong> atomic monolayer equivalents were employed.<br />

The temperature region <strong>of</strong> aggregation was determined by gradually heating up the samples and<br />

recording 2-D X-ray scattering images during the heat treatment. The analysis <strong>of</strong> the data<br />

provides information about the average size and shape <strong>of</strong> supported particles [1].<br />

The size <strong>of</strong> the nanoparticles formed on the surface upon cluster deposition and their thermal<br />

stability exhibited a very strong dependence on the initial size <strong>of</strong> the clusters, level <strong>of</strong> surface<br />

coverage and material <strong>of</strong> the support. When applying surface coverages higher than 5% ML,<br />

aggregation <strong>of</strong> the clusters into nanometer-size particles with narrow size distribution prevailed.<br />

With decreasing coverage, the fraction <strong>of</strong> not-aggregated clusters significantly increased. Heat<br />

treatment led to final particle sizes determined by initial cluster size, coverage and length <strong>of</strong><br />

treatment. The most striking results were observed when depositing 2.5 % ML <strong>of</strong> Au7-Au10<br />

clusters on SiO2/Si(111): particles <strong>of</strong> one size, corresponding to Au20 were formed. Moreover,<br />

these particles exhibited extraordinary and unexpected thermal stability as well, not undergoing<br />

aggregation until 400 °C, when an onset <strong>of</strong> a slow agglomeration takes place [2]. Work on<br />

supported clusters <strong>of</strong> single size and in reactive gas atmosphere is currently under progress.<br />

This work and use <strong>of</strong> APS was supported by the U.S. Department <strong>of</strong> Energy, Office <strong>of</strong> Basic<br />

Energy Sciences, Division <strong>of</strong> Chemical Sciences, Geosciences, and Biosciences under contract<br />

number W-31-109-ENG-38.<br />

References<br />

[1] R. E. Winans, S. Vajda, B. Lee, S. J. Riley, S. Seifert, G.Y. Tikhonov, N. Tomczyk, J. Phys.<br />

Chem. B, 108, 18105 (2004).<br />

[2] S. Vajda, R. E. Winans, J.W. Elam, B. Lee, M.J. Pellin, S.J. Riley, S. Seifert, G.Y. Tikhonov and<br />

N. A. Tomczyk, invited paper, submitted to ACS.


Oxidation <strong>of</strong> free and supported Pd clusters<br />

Bernd Huber 1 , Michael Moseler 1,2<br />

1 Freiburger Materialforschungszentrum , Stefan-Meier-Strasse 21, 79104 Freiburg, Germany<br />

2 Fraunh<strong>of</strong>er Institut für Werkst<strong>of</strong>fmechanik, Wöhlerstrasse 11, 79108 Freiburg, Germany<br />

A - 35<br />

The adsorption sites <strong>of</strong> O2 on neutral PdN clusters (N=1-5) were studied using spin density<br />

functional theory [1]. Only for Pd1O2 molecular adsorption is found to be favorable. For Pd2-5O2<br />

dissociative adsorption with the oxygen sitting on Pd bridge sites is preferred. For N=4,5 the<br />

cluster increases its spin state from triplet to a quintet state in comparision to the pure gas phase<br />

Pd cluster. For molecular adsorption the O-O bond gets activated to a superoxolike state.<br />

While for the gas phase clusters the adsorption-induced relaxation <strong>of</strong> the PdN cluster was small,<br />

we see for the oxidation on MgO supported clusters the breaking <strong>of</strong> Pd-Pd bonds and the<br />

formation <strong>of</strong> nano-oxides reflecting the ionic crystal structure <strong>of</strong> the MgO underlayer (Fig. 1).<br />

Figure 1. The formation <strong>of</strong> Pd4 nano-oxides on a MgO underlayer. While the oxidized gas phase Pd 4 cluster (a)<br />

remains in the same structure as the pure cluster, supported Pd 4O 2 deforms into a structure reflecting the ionic crystal<br />

structure <strong>of</strong> the MgO underlayer (b + c).<br />

References<br />

[1] B. Huber, H. Häkkinen, U. Landman and M. Moseler, Comp. Mat. Sci. (accepted).<br />

153


A - 36<br />

The Size-Evolution <strong>of</strong> the Optical Properties <strong>of</strong> Supported Gold<br />

Clusters and Nanocrystals Studied by Cavity Ringdown Spectroscopy:<br />

From the Atom to the Bulk<br />

154<br />

J. M. Antonietti, M. Michalski, H. Jones, and U. Heiz<br />

Institute <strong>of</strong> Surface Chemistry and Catalysis, <strong>Universität</strong> Ulm, 89069 Ulm, Germany<br />

Measuring the size-dependent electronic structure <strong>of</strong> supported clusters near the Fermi level<br />

is <strong>of</strong> critical importance in order to understand the catalytic activities <strong>of</strong> supported clusters or to<br />

tune the optical properties <strong>of</strong> cluster assembled materials with the number <strong>of</strong> atoms. Up to now,<br />

extensive studies on the electronic structure <strong>of</strong> size-selected, supported clusters are sparse,<br />

because <strong>of</strong> the experimental challenges which must be overcome: preparation <strong>of</strong> monodispersed<br />

samples, high sensitivity required for recording the signal arising from a highly dilute sample,<br />

detection <strong>of</strong> the signal due to the clusters from the intense background signal arising from the<br />

substrate.<br />

In order to overcome these problems, a new experimental setup has been designed and built to<br />

perfom Cavity Ringdown Spectroscopy (CRDS) in the visible spectral range on size-selected<br />

clusters deposited on silica in UHV conditions. The clusters are produced in a laser vaporization<br />

source, size-selected in the gas phase, and s<strong>of</strong>t-landed onto the substrate.<br />

In this work, the ability to prepare monodispersed cluster samples is demonstrated. The ultimate<br />

sensitivity <strong>of</strong> the setup is such that absorption cross-sections <strong>of</strong> 0.02 Å 2 with cluster coverage<br />

down to 0.05% <strong>of</strong> a monolayer (~1 10 12 clusters/cm 2 ) can be detected. The absorption spectra <strong>of</strong><br />

gold clusters, Aun (n = 1, 4, 8, and 20) deposited on silica are presented. Additionally, the<br />

optical properties <strong>of</strong> self-organized gold nanocrystals with well-defined diameters (1 to 20 nm),<br />

prepared by chemical means and deposited on silica have been measured. The spectrum <strong>of</strong> the<br />

atom exhibits sharp peaks which are attributed to atoms attached at various defect sites on the<br />

surface. A dramatic size-sensitivity <strong>of</strong> the absorption properties <strong>of</strong> clusters is observed.<br />

However, Au8 shows features which are associated with nanocrystals, rather than small clusters.


A - 37<br />

The Influence <strong>of</strong> Growth Kinetics on Island Morphology for Antimony<br />

and Bismuth Diffusion and Aggregation on Graphite<br />

Bernhard Kaiser 1 , Bert Stegemann 1 , Shelley Scott 2;3 , Simon Brown 2;3<br />

1 Institute for Chemistry, Humboldt-University, Berlin, Germany<br />

2 The MacDiarmid Institute for Advanced Materials and Nanotechnology, New Zealand<br />

3 Department <strong>of</strong> Physics and Astronomy, University <strong>of</strong> Canterbury, Christchurch, New Zealand<br />

The morphology <strong>of</strong> antimony and bismuth islands has been studied with scanning electron<br />

microscopy (SEM), transmission electron microscopy (TEM) and atomic force microscopy<br />

(AFM). Altering the deposition parameters (particle flux and coverage) shifts the balance<br />

between thermodynamics and kinetics. For antimony islands, we observe that for low flux and<br />

coverage the islands are compact, but undergo a transition to dendritic morphologies for higher<br />

flux and coverage regimes, consistent with previous investigations [1]. The islands also become<br />

flatter with increasing flux. For bismuth islands, we find a morphology transition from<br />

hexagonal to six-pointed star with increasing kinetic influence. Also, the island height is<br />

independent <strong>of</strong> flux, and vertical growth proceeds via an unusual striped layer growth.<br />

Furthermore, for bismuth it is possible to grow nanorods with a well-defined diameter <strong>of</strong> only a<br />

few nanometers and a length <strong>of</strong> several micrometers.<br />

References<br />

[1] B. Kaiser, B. Stegemann, H. Kaukel, K . Rademann, Surf. Sci. 496, L18 (2002)<br />

155


A - 38<br />

156<br />

Transition From One- to Two-Dimensional Island Growth During<br />

Molecular Beam Epitaxy: A Monte Carlo Study<br />

S. P. Shrestha 1 and C. Y. Park 2<br />

1 Department <strong>of</strong> Physics, Patan M. Campus, Patan Dhola, Nepal<br />

2 Department <strong>of</strong> Physics, Sung Kyun Kwan University, Suwon 440-746, South Korea<br />

We present the Monte Carlo results for submonolayer island growth during Molecular<br />

Beam Epitaxy using irreversible nucleation and growth model. We have monitored island<br />

morphology and island anisotropy parameter for different diffusional and sticking anisotropy<br />

case. We find that in anisotropic sticking case, when diffusion is isotropic, increase in D/F leads<br />

highly elongated islands where as when diffusion is anisotropic, only slight elongation <strong>of</strong><br />

islands are observed. In this case, for all values <strong>of</strong> D/F ratio, increase in DA is observed to cause<br />

decrease in island elongation. Thus, when sticking is anisotropic, diffusional anisotropy<br />

produces adverse effect on island elongation. When sticking is isotropic islands always show 2d<br />

nature and when it is highly anisotropic the islands show 1d nature. At intermediate sticking<br />

ratio, the D/F ratio results in transition <strong>of</strong> island morphology from 1d nature at low D/F ratio to<br />

2d nature at high D/F ratio.


Spectroscopy <strong>of</strong> free clusters and clusters on surfaces<br />

Chunrong Yin 1 , I. Barke 2 , D. Böcker 2 , Th. Irawan 2 , H. Hövel 2 , and B. v. Issendorff 1<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Freiburg, Stefan-Meier Str. 21, D-79104 Freiburg, Germany<br />

2 University <strong>of</strong> Dortmund, Experimentelle Physik I, D-44221 Dortmund, Germany<br />

A - 39<br />

Understanding the evolution <strong>of</strong> physical properties from free clusters to clusters on surfaces<br />

is a prerequisite for the use <strong>of</strong> clusters in technical applications. Photoelectron spectroscopy<br />

(PES) is a powerful tool to study the electronic properties <strong>of</strong> free clusters in vacuum as well as<br />

<strong>of</strong> supported clusters. Scanning tunneling spectroscopy (STS) is able to probe the electronic<br />

structure <strong>of</strong> individual clusters. Up to now, however, for metal clusters on surfaces these three<br />

techniques seemed to yield rather different results. A direct comparison <strong>of</strong> the results obtained<br />

on free clusters as well as on the same clusters deposited on well defined surfaces will hopefully<br />

clarify this situation, and yield definite informations about the electronic structure <strong>of</strong> the<br />

cluster/surface system and the nature <strong>of</strong> charge transfer processes between the cluster and the<br />

surface. Furthermore this comparison is expected to improve the fundamental understanding <strong>of</strong><br />

the results <strong>of</strong> tunneling spectroscopy on metal particles.<br />

Currently, a new setup is being built which will allow the deposition <strong>of</strong> size selected clusters on<br />

surfaces within an existing surface science apparatus. Scanning tunneling microscopy (STM)<br />

and ultraviolet photoelectron spectroscopy (UPS) will be performed on the deposited clusters<br />

and can be compared to the results <strong>of</strong> photoelectron spectroscopy obtained on the free clusters.<br />

Preliminary experiments have already been done on noble metal islands produced by deposition<br />

<strong>of</strong> metal atoms on rare gas films. First results show a dependence <strong>of</strong> the island Fermi edge<br />

position on the thickness <strong>of</strong> the isolating rare gas layer. Further measurements are underway.<br />

157


A - 40<br />

158<br />

Multiple Size-Selected Cluster Deposition:<br />

Epitaxial Growth at Thermal Energies<br />

Krist<strong>of</strong>fer Meinander, Kai Nordlund, and Juhani Keinonen<br />

Accelerator Laboratory, P.O.Box 43, FIN-00014 University <strong>of</strong> Helsinki, Finland<br />

It has been known for some time that when a small enough cluster lands on a singlecrystalline<br />

substrate it will align completely epitaxially upon impact [1]. In a previous study,<br />

using molecular dynamics simulations, it was noted that the upper size limit <strong>of</strong> Cu nanoclusters<br />

that will align epitaxially with a smooth (100) Cu substrate upon deposition at thermal energies<br />

is linearly dependent on temperature [2]. Deposition <strong>of</strong> clusters with sizes below the upper size<br />

limit will initially result in the formation <strong>of</strong> epitaxial structures on the surface <strong>of</strong> the substrate,<br />

however, as deposition continues clusters may impact on previously deposited clusters. The<br />

result <strong>of</strong> this may be a change in the likelihood <strong>of</strong> epitaxial alignment, as the deposition no<br />

longer takes place on a smooth surface, and hence a lowering in the upper size limit <strong>of</strong> clusters<br />

that can be used in the deposition <strong>of</strong> thick epitaxial films.<br />

Molecular dynamics simulations <strong>of</strong> multiple Cu cluster deposition were carried for different<br />

cluster sizes, over a temperature range <strong>of</strong> 0-750 K, in order to investigate the effect <strong>of</strong> cluster on<br />

cluster impacts to the upper size limit <strong>of</strong> epitaxial alignment. For each cluster size, several<br />

clusters were deposited in sequence, with a 2 ns interval between each cluster, on the same<br />

initially smooth, (100) Cu substrate. Through analysis <strong>of</strong> the degree <strong>of</strong> epitaxy between each<br />

deposition event, the maximum number <strong>of</strong> deposited clusters resulting in an epitaxial structure,<br />

for each specific cluster size, could be determined as a function <strong>of</strong> temperature.<br />

It was found that there is a substantial decrease in the upper size limit for epitaxial alignment<br />

when two clusters are deposited, as compared to that for the deposition <strong>of</strong> one cluster, for cases<br />

where the second cluster impacts directly on top <strong>of</strong> the one previously deposited. As the amount<br />

<strong>of</strong> deposited clusters increases, this decrease in the upper size limit is lowered, and ceases<br />

approximately after the fifth or sixth cluster.<br />

References<br />

[1] M. Yeadon et al., J. Elect. Microsc. 48 (Suppl.S), 1075 (1999).<br />

[2] K. Meinander, J. Frantz, K. Nordlund, and J. Keinonen, Thin Solid Films 45/1-2, 297-303 (2003).


C58 on HOPG<br />

A. Böttcher, P. Weis, S. Jester, A. Bihlmeier, D. Löffler, M. M. Kappes<br />

A - 41<br />

Institut für Physikalische Chemie, <strong>Universität</strong> Karlsruhe, Fritz-Haber-Weg 4, 76131 Karlsruhe, Germany<br />

The process <strong>of</strong> electron-impact induced fragmentation <strong>of</strong> C60 molecules has been exploited<br />

as an intense source <strong>of</strong> C58 + ions. The ionic beam was directed toward a HOPG substrate by a<br />

system <strong>of</strong> electrostatic lenses. A new solid material grows on the substrate as result <strong>of</strong> gentle<br />

deposition <strong>of</strong> C58 + ions (kinetic energy < 0.1 eV/atom). Thermal desorption spectra <strong>of</strong> the films<br />

created indicate that neither substrate-mediated decomposition nor dimerization <strong>of</strong> C58 occurs.<br />

The activation energy for desorption <strong>of</strong> C58 molecules from the HOPG surface is about 1 eV<br />

higher than found for C60 molecules on the same substrate. The apparent binding energy varies<br />

with increasing adsorbate coverage from 2 to 2.2 eV. The morphology <strong>of</strong> the deposited films<br />

has been investigated by means <strong>of</strong> AFM and STM. A fractal growth <strong>of</strong> 2D islands has been<br />

found to be the dominating deposition channel. The growth kinetics depends on the impact<br />

energy and surface temperature.<br />

159


A - 42<br />

160<br />

A comparative simulation <strong>of</strong> Al interaction with Ge and Si<br />

nanoclusters<br />

Ol’ga Ananyina, Olexandr Yanovs’ky<br />

SSEM Department, Zaporozhye State University, Zhukovsky Str. 66, Zaporozhye, 69063 Ukraine, e-mail:<br />

ananyina@zsu.zp.ua.<br />

Results <strong>of</strong> quantum-chemical calculations <strong>of</strong> Al-clusters (Al2, Al3, Al4) interaction with Si<br />

and Ge(100) surfaces are represented in this work. Al adsorption on clean semiconductor<br />

surfaces has an important technological significance as one <strong>of</strong> the ways <strong>of</strong> receiving metalsemiconductor<br />

contacts. Our calculations were carried out with the help <strong>of</strong> semi-empirical<br />

MNDO method (modified neglect <strong>of</strong> differential overlap) for Si33(Ge33) and Si67(Ge67) clusters,<br />

which are modeling the clean ordered Si(100)-(2×4) and Ge(100)-(2×1) surfaces. Cluster<br />

geometry and total energy, atom bonds orders, value <strong>of</strong> the electron density, atom orbital<br />

populations and molecular localized orbitals were calculated. Simulation <strong>of</strong> possible structures,<br />

formed by Al-clusters on Si(Ge) surfaces is the goal <strong>of</strong> this work, because forming <strong>of</strong> the certain<br />

adsorbates structure defines further growth <strong>of</strong> Al film.<br />

Two possible states <strong>of</strong> surface covered with Al2 clusters were received: the Al-Al “vertical” and<br />

“lateral” dimer models. Electronic states <strong>of</strong> Si(Ge)-(100) surface for different dimer models are<br />

compared.<br />

Fig.1. A result <strong>of</strong> Al2 - and Al3 adsorption on Ge(100)-(2×1) surface: a) a part <strong>of</strong> a cluster modelling clean<br />

Ge(100) surface with adsorbed Al2; “lateral” Al dimers forming; b) a part <strong>of</strong> a cluster modelling clean<br />

Ge(100) surface with adsorbed Al2; “vertical” Al dimers forming; c) cluster Ge63 with adsorbed Al3.<br />

Al3 adsorption on the semiconductors surfaces can be accompanied by the dissociation <strong>of</strong> Al3<br />

cluster on the Al atoms, which saturates bonds <strong>of</strong> surface atoms <strong>of</strong> Si-Si (Ge-Ge) dimers. But<br />

minimal total system energy corresponds to the state when Al3 cluster adsorbed on the surface<br />

without dissociation. An increase in aluminum cluster size causes the increase <strong>of</strong> cluster<br />

dissociation probability.<br />

The influence <strong>of</strong> surface defects on the processes <strong>of</strong> “surface - cluster” interaction and the<br />

possible reaction mechanisms at different coverage <strong>of</strong> such surfaces by dihydride and<br />

monohydride chemisorption phases are discussed. Comparative analysis <strong>of</strong> results <strong>of</strong> simulation<br />

for clusters <strong>of</strong> different size is carried out.


Characterization and Catalysis on a Well-defined Pt Cluster<br />

in NaY Zeolite<br />

Xiong Liu 1 , Rüdiger-A. Eichel 2 , Emil Roduner 1<br />

1 Institut für Physikalische Chemie, <strong>Universität</strong> Stuttgart, Pfaffenwaldring 55<br />

D-70569 Stuttgart, Germany<br />

2 Institute <strong>of</strong> Physical Chemistry III, Darmstadt University <strong>of</strong> Technology, Petersenstraße 20<br />

D-64287 Darmstadt, Germany<br />

A - 43<br />

Size effects <strong>of</strong> metal clusters play an important role in heterogeneous catalysis since the<br />

electronic properties, reactivity and selectivity <strong>of</strong> catalysts with a few atoms are strongly sizedependent.<br />

Platinum clusters dispersed in zeolites or on porous oxide supports are highly active<br />

catalysts for the oxidation <strong>of</strong> CO and residual hydrocarbons in automotive exhaust catalysis and<br />

for hydrogenation in petrochemical reactions. It is found that the catalyst performance increases<br />

with decreasing Pt cluster size. Usually, in supported metal catalysts the cluster size, shape and<br />

morphology are not uniform, which make the evaluation <strong>of</strong> catalytic properties complicated.<br />

Therefore, the preparation and characterization <strong>of</strong> homogeneous size clusters for understanding<br />

catalytic elementary steps and realizing catalysis on well-defined clusters is very desirable.<br />

Recently, we have demonstrated that under carefully controlled conditions it is possible to<br />

observe a single, well-defined small Pt cluster in zeolites by EPR spectroscopy. A highly<br />

symmetric cluster with 12 paramagnetic equivalent Pt atoms was characterized by continuous<br />

wave and pulse EPR. ENDOR and HYSCORE experiments reveal the proton hyperfine<br />

couplings <strong>of</strong> the adsorbed hydrogen on the surface. The observed species may be Pt13H12 n+ [1]<br />

or Pt13H20 n+ (n = +1 or +3) cluster with an icosahedral structure.<br />

Hydrogen or deuterium adsorbed on the Pt cluster exchanges and is desorbed and re-adsorbed<br />

reversibly. The desorption energy amounts to 0.69 eV per atom, in good agreement with<br />

theoretical values based on DFT calculations. Exposure to CO results in the decomposition <strong>of</strong><br />

the cluster and growth <strong>of</strong> new paramagnetic Pt-carbonyl clusters. Amazingly, reaction with O2<br />

is rather slow. SQUID measurements reveal that there is a large fraction <strong>of</strong> EPR silent<br />

paramagnetic and diamagnetic Pt species. They are partly in electron transfer equilibrium with<br />

the observed cluster.<br />

2600 2800 3000 3200<br />

Magnetic Field (G)<br />

6% Pt/NaY; reduction in H 2<br />

then exchanged by D 2<br />

6% Pt/NaY; reduction in H 2<br />

Figure 1. EPR spectrum <strong>of</strong> 6% Pt/NaY recorded at 20 K, demonstrating reversible exchange <strong>of</strong> hydrogen isotopes.<br />

References<br />

[1] T. Schmauke, R-A. Eichel, A. Schweiger, E. Roduner; Phys. Chem. Chem. Phys. 5, 3076 (2003).<br />

161


A - 44<br />

DFT-Investigations <strong>of</strong> coalescence behaviour <strong>of</strong> Si4 and Si7 clusters on<br />

surfaces<br />

162<br />

Wolfram Quester 1 , Dominik Fischer 1,2 and Peter Nielaba 1<br />

1 Fachbereich Physik, <strong>Universität</strong> <strong>Konstanz</strong>, 78457 <strong>Konstanz</strong><br />

2 IBM Zürich Research Laboratory, Säumerstraße 4, 8803 Rüschlikon<br />

Experimental results show that Si4 and Si7 clusters do not form islands <strong>of</strong> bulk Si on weakly<br />

interacting surfaces (HOPG). We investigated the coalescence behaviour using Density<br />

Functional Theory implemented in the CPMD code (www.cpmd.org).<br />

We calculated potential energy curves <strong>of</strong> two approaching Si4 clusters on two reaction channels.<br />

It could be shown that there exists a fusion barrier which is higher than room temperature.<br />

These calculations will be extended to Si7 clusters. To get informations about the influence <strong>of</strong><br />

the substrate a calculation <strong>of</strong> a Si4 cluster on a gold surface was perfomed. This simulation<br />

revealed that gold is not suited as substrate for depositing Si clusters.


A - 45<br />

Morphologies Studies <strong>of</strong> Mass Selected Au Clusters on Rutile TiO2 as a<br />

Function <strong>of</strong> Impact Energy and Surface Temperature<br />

P. Convers, R. Vallotton, R. Monot, W. Harbich<br />

Institut de Physique des Nanostructures,<br />

Ecole polytechnique Fédérale de Lausanne (EPFL),<br />

CH-1015 Lausanne<br />

In 1989 Haruta and coworkers have shown that small gold particles (diameter < 4 nm) are<br />

catalytically active [1]. Gold was found to be even more efficient for temperatures below 400°C<br />

than Pt, usually used for CO oxydation [2, 3]. More recent studies by Heiz et al. (Mass selected<br />

Aun deposited on MgO) and Anderson et al. (Mass selected Aun deposited on TiO2) report a<br />

pronounced size dependence <strong>of</strong> the catalytic activity [4, 5]. However informations on the<br />

morphology <strong>of</strong> the clusters is sparse. Wahlström et al. showed that atomic Au thermally<br />

deposited on a surface <strong>of</strong> TiO2(110) form clusters that are stabilized by oxygen vacancies at the<br />

surface [6]. The clusters are small enough for catalysis, however a temperature increase causes a<br />

sintering <strong>of</strong> the clusters and probably decreases the catalytic efficiency.<br />

To understand the clusters morphology as a function <strong>of</strong> temperature, we deposit mass-selected<br />

gold clusters with defined kinetic energy on a TiO2(110). The morphology is then studied by<br />

STM after different heating cycles up to 1000 K. We observe that large cluster sizes (n = 7) and<br />

high deposition energy (up to 4500 eV ) stabilize gold on the surface. Figure 1 is a typical STM<br />

image <strong>of</strong> a deposit <strong>of</strong> Au7 on TiO2 at 20 eV after a heating cycle at 800 K.<br />

References<br />

Figure 1. Au 7/TiO 2 at 20eV after annealing at 800K<br />

[1] M. Haruta, N. Yamada, T. Kobayashi, and S. Iijima, Journal <strong>of</strong> Catalysis 115, 301 (1989).<br />

[2] M. Haruta, Catalysis Today 36, 153 (1997).<br />

[3] M. Valden, X. Lai, and D. W. Goodman, Science 281, 1647 (1998).<br />

[4] A. Sanchez, S. Abbet, U. Heiz, W.-D. Schneider, H. Häkkinen, R. N. Barnett, and U. Landman, J.<br />

Phys. Chem. A 103, 9573 (1999).<br />

[5] S. Lee, C. Fan, T. Wu, and S. L. Anderson, J. Am. Chem. Soc. 126, 5682 (2004).<br />

[6] E. Wahlstrom, N. Lopez, R. Schaub, P. Thostrup, A. Ronnau, C. Africh, E. Laegsgaard, J.<br />

Norskov, and F. Besenbacher, Phys. Rev. Lett. 90, 026101 (2003).<br />

163


A - 46<br />

164<br />

Mass-selected Nanocrystal Superlattices<br />

J. T. Lau 1 , H.Weller 2 , T. Möller 1<br />

1 Technische <strong>Universität</strong> Berlin, IAPF PN 3-1, Hardenbergstraße 36, D-10623 Berlin<br />

2 <strong>Universität</strong> Hamburg, Institut für Physikalische Chemie, Grindelallee 117, D-20146 Hamburg<br />

Nanocrystal superlattices built <strong>of</strong> identical particles are expected to exhibit novel electronic and<br />

optical properties that can be tuned by particle size and interparticle distance. Below a distance<br />

<strong>of</strong> approx. 2–5 Å, for example, coherent tunneling coupling or exchange coupling between<br />

nanocrystals is expected because <strong>of</strong> wave function overlap.<br />

We will present the design <strong>of</strong> our experiment and discuss how superlattice properties can be<br />

probed by X-ray absorption, X-ray photoelectron spectroscopy, X-ray scattering and<br />

photoluminescence spectroscopy.


165


Author Index<br />

167


168


Abbet, S.......................................................... 53<br />

Aguado, A. ................................................... 125<br />

Ahmed, M....................................................... 65<br />

Akola, J......................................................... 149<br />

Alamanova, D............................................... 105<br />

Alexandrova, A. N........................................ 112<br />

Ananyina, O. ................................................ 160<br />

Andersen, J. U. ............................................. 129<br />

Anderson, S. L.................................................. 7<br />

Andersson, M. ................................................ 50<br />

Antonietti, J. M....................................... 53, 154<br />

Arredondo, M. G. ........................................... 49<br />

Asmis, K. R. ............................................. 55, 59<br />

Bahat, M....................................................... 118<br />

Balaj, O. P. ..................................................... 25<br />

Balakrishnan, K............................................ 151<br />

Balbás, L. C.................................................... 57<br />

Balteanu, I. ..................................................... 25<br />

Barat, M.......................................................... 13<br />

Barke, I......................................................... 157<br />

Barnakov, Y. A................................................. 9<br />

Bartels, Chr..................................................... 69<br />

Baruah, S. ....................................................... 96<br />

Bechthold, P. S. .............................................. 51<br />

Belosludov, R. V. ............................................. 9<br />

Benz, L. .......................................................... 16<br />

Berdunov, N. ................................................ 151<br />

Berry, R. St................................................... 126<br />

Bertram, N................................................ 88, 89<br />

Beuhler, R. J. .................................................. 54<br />

Beyer, M. K.................................................... 25<br />

Bihlmeier, A. ................................................ 159<br />

Blaum, K. ....................................................... 96<br />

Block, M......................................................... 96<br />

Blom, M. N................................................... 106<br />

Böcker, D. .................................................... 157<br />

Boese, A. D. ................................................. 116<br />

Boldyrev, A. I............................................... 112<br />

Bolton, K. ....................................................... 41<br />

Bonačić-Koutecký, V. ........................ 11, 62, 75<br />

Bondybey, V. E. ............................................. 25<br />

Böttcher, A. .................................................. 159<br />

Bousquet, G.................................................... 73<br />

Bouteiller, Y. .................................................. 12<br />

Bowen, Jr., K. H............................................... 6<br />

Bowers, M. T.................................................. 16<br />

Bréchignac, C. .................................................. 3<br />

Brown, S....................................................... 155<br />

Brück, J........................................................... 47<br />

Brümmer, M. .................................................. 59<br />

Buck, U................................................... 29, 143<br />

Bulusu, S. ..................................................... 112<br />

Buratto, S. K................................................... 16<br />

Bürgel, Chr..................................................... 62<br />

Burmeister, F.................................................. 66<br />

Campbell, E. E. B............................. 40, 68, 132<br />

Castleman Jr., A. W.................................. 26, 62<br />

Cederbaum, L. S........................................... 144<br />

Chaudhuri, A...................................................96<br />

Chelikowsky, J. R. ..........................................22<br />

Cheshnovsky, O. .............................................93<br />

Chrétien, St. ....................................................16<br />

Christen, W. ..................................................115<br />

Claas, P. ..........................................................72<br />

Concina, B. .....................................................38<br />

Convers, P.....................................................163<br />

Cordes, J. ........................................................89<br />

Danylchenko, O. G. ......................................124<br />

de Heer, W.......................................................19<br />

Dedonder-Lardeux, C. ....................................13<br />

Denifl, S..................................................38, 131<br />

Desfrançois, C...........................................12, 13<br />

Dietsche, R......................................................88<br />

Ding, F. ...........................................................41<br />

Dinh, M...........................................................73<br />

Dmitruk, I....................................................9, 85<br />

Dong, Y...........................................................48<br />

Donges, J.........................................................32<br />

Döppner, T......................................................71<br />

Droppelmann, G..............................................72<br />

Duncan, M. A....................................................5<br />

Eberhardt, W.............................................51, 66<br />

Echt, O. ...................................................38, 129<br />

Ehrler, O. T. ....................................................39<br />

Eichel, R.-A. .................................................161<br />

Elam, J. W.....................................................152<br />

Engelke, M......................................................74<br />

Fan, Ch..............................................................7<br />

Fanourgakis, G. S..........................................119<br />

Fayeton, J. A. ..................................................13<br />

Fedor, J. ........................................................129<br />

Fehrer, F..........................................................73<br />

Feil, S......................................................38, 131<br />

Fennel, Th. ......................................................71<br />

Fernández, E. M..............................................57<br />

Fielicke, A...............................................45, 103<br />

Fischer, D......................................................162<br />

Fischer, T. .......................................................88<br />

Fontanella, S. ..................................................59<br />

Frenzel, J.......................................................139<br />

Furche, F. ........................................................39<br />

Gadéa, F. X. ..................................................133<br />

Galak, M. ........................................................85<br />

Ganteför, G. ..................................61, 74, 88, 89<br />

Gemming, S. ...................................................30<br />

Giniger, R........................................................93<br />

Glaser, L. ........................................................17<br />

Głuch, K..........................................38, 129, 131<br />

Grégoire, G. ..............................................12, 13<br />

Grigoryan, V. G. ...........................................105<br />

Gromov, A. .....................................................40<br />

Haberland, H...................................................32<br />

Hagelberg, F..................................................129<br />

Häkkinen, H................................46, 87, 97, 104<br />

Hammer, B......................................................56<br />

Hampe, O........................................................52<br />

169


Hansen, K................................. 40, 68, 129, 132<br />

Harbich, W. .................................................. 163<br />

Hedén, M.................................................. 40, 68<br />

Heiz, U. .................................................. 53, 154<br />

Herfurth, F...................................................... 96<br />

Herlert, A.......................................... 96, 98, 110<br />

Higashimine, K............................................. 150<br />

Hippler, T. ...................................................... 32<br />

Hock, Chr. ...................................................... 69<br />

Horoi, M......................................................... 28<br />

Hövel, H. ...................................................... 157<br />

Hrivňák, D............................................ 130, 133<br />

Huang, X. ....................................................... 22<br />

Huber, B. .............................................. 111, 153<br />

Ikeda, K. S...................................................... 76<br />

Irawan, Th. ................................................... 157<br />

Jackson, K. A. .......................................... 28, 80<br />

Janeček, I...................................................... 130<br />

Janssens, E.................................................... 108<br />

Jena, P............................................................. 23<br />

Jester, S......................................................... 159<br />

Jeyadevan, V. ................................................... 9<br />

Johansson, M. P.............................................. 42<br />

Jones, H. ....................................................... 154<br />

Jonsson, F....................................................... 40<br />

Joswig, J.-O. ......................................... 138, 139<br />

Jouvet, C......................................................... 13<br />

Judai, K........................................................... 53<br />

Jusélius, J........................................................ 42<br />

Kaiser, B....................................................... 155<br />

Kalus, R........................................ 123, 130, 133<br />

Kang, H. ......................................................... 13<br />

Kantarci, Z.................................................... 118<br />

Kappes, M. M........................... 39, 52, 106, 159<br />

Karlický, F.................................................... 123<br />

Kasuya, A................................................... 9, 85<br />

Kato, K. ........................................................ 150<br />

Kawazoe, Y. ..................................................... 9<br />

Kaya, K........................................................... 31<br />

Keinonen, J................................................... 158<br />

Kellerbauer, A. ............................................... 96<br />

Kemper, P....................................................... 16<br />

Kern, K........................................................... 15<br />

Khan, F. A. ..................................................... 49<br />

Khanna, S. N. ................................................. 86<br />

Kim, J. H. ....................................................... 65<br />

Kim, Y. D..................................... 61, 74, 88, 89<br />

Kim, Y.-T. .................................................... 150<br />

Kimble, M. L.................................................. 62<br />

Kiran, B. ........................................... 58, 60, 112<br />

Kjellberg, M. ................................................ 132<br />

Klingeler, R. ................................................... 51<br />

Kluge, H.-J. .................................................... 96<br />

Knickelbein, M. B. ......................................... 20<br />

Kobayashi, T. ................................................. 76<br />

Kolmakov, A. ................................................. 16<br />

Kondow, T...................................................... 21<br />

Kordel, M. .................................................... 106<br />

Koskinen, P. ................................................. 104<br />

170<br />

Kostko, O................................................32, 109<br />

Kovalenko, S. I. ............................................124<br />

Krause, T.......................................................115<br />

Kresin, V. V............................................67, 107<br />

Kresin, V. Z. ...................................................67<br />

Kronik, L.........................................................22<br />

Kryzhevoi, N. V............................................144<br />

Kudo, T. ............................................................9<br />

Kuhnen, R. ......................................................69<br />

Kumar, V. .........................................................9<br />

kyun Lee, E.....................................................37<br />

Lagutschenkov, A. ..................................47, 119<br />

Landman, U. ...................................................53<br />

Lassesson, A. ................................................110<br />

Lau, J. T. .......................................................164<br />

Leavitt, A. J.....................................................49<br />

Lecomte, F. .....................................................12<br />

Lee, B............................................................152<br />

Lee, S. ...............................................................7<br />

Li, J. ........................................................58, 112<br />

Li, X..........................................................58, 60<br />

Lievens, P................................................50, 108<br />

Lifshitz, C. ......................................................38<br />

Lightstone, J....................................................54<br />

Lim, D. C. .................................................61, 88<br />

Liu, X............................................................161<br />

Liu, Z. .......................................................9, 145<br />

Löffler, D. .....................................................159<br />

López, J. M. ..................................................125<br />

Lopez-Salido, I..........................................61, 88<br />

Lucas, B. .........................................................12<br />

Lüttgens, G......................................................51<br />

Makmal, A. .....................................................22<br />

Malakzadeh, A. ...............................................69<br />

Mamykin, S. V..................................................9<br />

Manninen, M.................................................149<br />

Mariotto, G....................................................151<br />

Märk, T. D. .....................................38, 129, 131<br />

Martinez, F......................................98, 110, 131<br />

Martins, M. .....................................................17<br />

Marx, G.............................................96, 98, 110<br />

Matt-Leubner, S. .....................................38, 129<br />

Meijer, G...........................................45, 59, 103<br />

Meinander, K. ...............................................158<br />

Meiwes-Broer, K.-H. ................................70, 71<br />

Metiu, H..........................................................16<br />

Michalski, M.................................................154<br />

Milczarek, G. ....................................................9<br />

Mitani, T. ......................................................150<br />

Mitrić, R....................................................62, 75<br />

Mitsui, M. .......................................................31<br />

Molina, L. M...................................................56<br />

Möller, T. ......................................................164<br />

Monot, R. ......................................................163<br />

Moore, N. A....................................................62<br />

Moro, R.........................................................107<br />

Moseler, M....................................104, 111, 153<br />

Mukherjee, M..................................................96<br />

Mukovskii, Y. M...........................................151


Murakami, J.................................................... 24<br />

Murphy, S..................................................... 151<br />

Nakajima, A. ............................................ 31, 79<br />

Neeb, M.................................................... 51, 66<br />

Neukermans, S.............................................. 108<br />

Neumaier, M................................................... 52<br />

Neumark, D. M......................................... 10, 65<br />

Niedner-Schatteburg, G.................... 47, 81, 119<br />

Nielaba, P. .................................................... 162<br />

Niemietz, M.................................................... 74<br />

Nirasawa, T. ..................................................... 9<br />

Nordlund, K.................................................. 158<br />

Ohshima, K................................................... 150<br />

Ohsuna, T. ........................................................ 9<br />

Ohta, T............................................................ 21<br />

Osaka, K....................................................... 150<br />

Ovchinnikov, Y. N. ........................................ 67<br />

Palmer, R. E. .................................................. 14<br />

Pankewitz, T................................................... 47<br />

Park, C. Y..................................................... 156<br />

Park, J............................................................. 37<br />

Parks, J. H................................................. 8, 106<br />

Patterson, M. .................................................. 54<br />

Pedersen, D. B................................................ 45<br />

Pellin, M. J. .................................................. 152<br />

Peterka, D. S................................................... 65<br />

Pfeffer, B. ....................................................... 47<br />

Pontius, N....................................................... 51<br />

Popok, V....................................................... 132<br />

Prasalovich, S. .............................................. 132<br />

Proykova, A.................................................. 126<br />

Przystawik, A. ................................................ 70<br />

Ptasińska, S............................................. 38, 131<br />

Quester, W.................................................... 162<br />

Rabinovitch, R.............................................. 107<br />

Radcliffe, P..................................................... 70<br />

Ratsch, Chr................................................... 103<br />

Rayner, D. M.................................................. 45<br />

Reif, M. .......................................................... 17<br />

Reinhard, P.-G................................................ 73<br />

Ritter, H.......................................................... 98<br />

Roduner, E.................................................... 161<br />

Romanyuk, V. ............................................ 9, 85<br />

Ronen, S. ........................................................ 93<br />

Rönnow, E...................................................... 40<br />

Rosén, A......................................................... 41<br />

Roßteuscher, T................................................ 25<br />

Röttgen, M. A................................................. 53<br />

Roy, S................................................... 137, 138<br />

Rytkönen, K. ................................................ 149<br />

Samovarov, V. N.......................................... 124<br />

Santambrogio, G............................................. 59<br />

Santa-Nokki, T. .............................................. 46<br />

Sarfraz, A. .................................................... 115<br />

Sarkar, P. ...................................................... 138<br />

Sauer, J. .......................................................... 59<br />

Sawada, S. ...................................................... 76<br />

Scheier, P........................................ 38, 129, 131<br />

Schermann, J.-P........................................ 12, 13<br />

Schmidt, M......................................................32<br />

Schneider, H..........................................116, 117<br />

Schooss, D. ...................................................106<br />

Schulz, C.-P. ...................................................72<br />

Schweikhard, L. ................................96, 98, 110<br />

Scott, S..........................................................155<br />

Seifert, G.........................................30, 104, 139<br />

Seifert, S. ......................................................152<br />

Serrano, D. ......................................................49<br />

Shim, J. ...........................................................37<br />

Shimizu, Y. .....................................................76<br />

Shinoda, K. .......................................................9<br />

Shrestha, S. P. ...............................................156<br />

Shvets, I. V....................................................151<br />

Silverans, R. E. .............................................108<br />

Simard, B. .......................................................45<br />

Sivamohan, R....................................................9<br />

Springborg, M................. 48, 105, 137, 138, 139<br />

Srinivas, S. ......................................................80<br />

Stairs, J..........................................................106<br />

Stanzel, J. ........................................................66<br />

Stegemann, B................................................155<br />

Steinbach, Chr.........................................29, 143<br />

Stienkemeier, F. ..............................................72<br />

Suematsu, H..................................................150<br />

Sun, Q. ............................................................23<br />

Sundararajan, V. ...............................................9<br />

Sundholm, D. ..................................................42<br />

Suraud, E.........................................................73<br />

Swenson, D. R. ...............................................94<br />

Takada, M. ....................................................150<br />

Terasaki, A......................................................21<br />

Terasaki, O........................................................9<br />

Tiggesbäumker, J. .....................................70, 71<br />

Tikhonov, G. Y. ............................................152<br />

Tohji, K.............................................................9<br />

Tomczyk, N. A..............................................152<br />

Tong, X. ..........................................................16<br />

Tono, K. ..........................................................21<br />

Torres, M. B....................................................57<br />

Uruga, T........................................................150<br />

Vaara, J. ..........................................................42<br />

Vajda, St. ......................................................152<br />

Vallotton, R...................................................163<br />

Veldeman, N. ..........................................50, 108<br />

Velegrakis, M..................................................99<br />

Vetter, M.........................................................47<br />

Vítek, A.........................................................123<br />

von Gynz-Rekowski, F. ..................................88<br />

von Helden, G.........................................45, 103<br />

von Issendorff, B............... 32, 69, 109, 111, 157<br />

Wallace, W. T. ................................................49<br />

Walsh, N. ......................................................110<br />

Walter, L. ......................................................106<br />

Walter, M........................................................97<br />

Wang, C. .........................................................65<br />

Wang, L.-S........................................58, 60, 112<br />

Wang, Q..........................................................23<br />

Wang, X........................................................108<br />

171


Weber, J. M. ................................... 39, 116, 117<br />

Weigend, F. .................................................... 52<br />

Weis, P. ........................................................ 159<br />

Weller, H. ..................................................... 164<br />

Whetten, R. L. ................................................ 49<br />

White, M. G.................................................... 54<br />

Wies, St. ......................................................... 47<br />

Winans, R. E................................................. 152<br />

Wolf, I. ........................................................... 93<br />

Wörz, A. S...................................................... 53<br />

Wöste, L. .................................................... 4, 59<br />

Wu, T................................................................ 7<br />

Wurth, W.................................................. 17, 81<br />

172<br />

Wyrwas, R. B..................................................49<br />

Xantheas, S. S. ..............................................119<br />

Xia, C............................................................107<br />

Yamaguchi, W. ...............................................24<br />

Yanovs’ky, O................................................160<br />

Yazidjian, Ch. .................................................96<br />

Yeschenko, O..................................................85<br />

Yin, Ch..........................................................157<br />

Yoon, B...........................................................53<br />

Zeng, X. Ch...................................................112<br />

Zhai, H. J...........................................58, 60, 112<br />

Zhang, H.-F.....................................................58


Time<br />

8.30 a.m.<br />

9.15 a.m.<br />

10.00 a.m.<br />

10.30 a.m.<br />

11.15 a.m.<br />

12.00 p.m.<br />

2.00 p.m.<br />

-<br />

4.00 p.m.<br />

5.00 p.m.<br />

5.45 p.m.<br />

Sunday Monday<br />

Tuesday Wednesday Thursday Friday<br />

28 February 05<br />

27 February 05<br />

8:20 Opening<br />

1 March 05 2 March 05 3 March 05 4 March 05<br />

Arrival<br />

C. Bréchignac<br />

Metal Clusters and Oxygen<br />

L. Wöste<br />

Reactivity Ag+Au Clusters<br />

M. A. Duncan<br />

IR <strong>of</strong> Metal Complexes<br />

K. H. Bowen, Jr.<br />

Diffuse Electron States<br />

S. L. Anderson<br />

Model Au Catalysts<br />

J. H. Parks<br />

Structural Order in Metal Clusters<br />

D. M. Neumark<br />

Dynamics and Droplets<br />

V. Bonačić-Koutecký<br />

Tailoring functionality by size<br />

C. Desfrancois<br />

New Infrared Spectroscopy Tech.<br />

C. Jouvet<br />

The H-Transfer Reaction<br />

R. E. Palmer<br />

Immobilisation <strong>of</strong> Proteins<br />

K. Kern<br />

Architecture at surfaces<br />

W. A. de Heer<br />

Cluster Ferroelectricity<br />

M. B. Knickelbein<br />

Molecular Magnets<br />

A. W. Castleman, Jr.<br />

Cluster Assembled Materials<br />

M. F. Jarrold<br />

Al/Ga Cluster Melting<br />

HT 3 HT 8<br />

HT 5 HT 10<br />

HT 7<br />

K. Kaya<br />

S<strong>of</strong>t-Nano-Molecular Clusters<br />

H. Haberland<br />

Melting and Magic Numbers<br />

6.30 p.m. A. Kasuya (CdSe) n Clusters S. K. Buratto Ag n+Au n on TiO Concluding Remarks (P. Jena)<br />

*) Please see Table <strong>of</strong> Contents for the Poster Sessions<br />

Schedule<br />

Fermion/Boson Clusters+Catalysis<br />

C<strong>of</strong>fee<br />

Lunch<br />

Discussions<br />

K. Tono Oxidation <strong>of</strong> Cr n/Mn n<br />

L. Kronik Mag. Semiconductors<br />

Q. Sun Nanobullet for Tumor<br />

J. Murakami N 2 on supported W n<br />

M. Beyer CO-Oxidation on Pt n<br />

Special P. Leiderer Talk 15<br />

Potentials+Prospects <strong>of</strong> Nanoscience<br />

M. Horoi Phase Transitons<br />

U. Buck Na-doped H 2O Clusters<br />

G. Seifert MoS Clusters<br />

7.00 p.m.<br />

Dinner Conference<br />

8.30 p.m.<br />

Poster<br />

M. Martins Magnetic Clusters<br />

Poster<br />

Dinner<br />

9.15 p.m.<br />

Session A *)<br />

U. Landman<br />

Session B *)<br />

Post Conference Session<br />

Sketch: Scientific Dialogology - four<br />

techniques for asking proper questions<br />

Departure after<br />

breakfast

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!