Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

REVIEWS OF MODERN PHYSICS, VOLUME 77, JANUARY 2005

Structural properties of nanoclusters: Energetic, thermodynamic,


and kinetic effects
Francesca Baletto* and Riccardo Ferrando†
INFM and Dipartimento di Fisica, Università di Genova, via Dodecaneso 33, 16146
Genova, Italy
共Published 24 May 2005兲

The structural properties of free nanoclusters are reviewed. Special attention is paid to the interplay
of energetic, thermodynamic, and kinetic factors in the explanation of cluster structures that are
actually observed in experiments. The review starts with a brief summary of the experimental methods
for the production of free nanoclusters and then considers theoretical and simulation issues, always
discussed in close connection with the experimental results. The energetic properties are treated first,
along with methods for modeling elementary constituent interactions and for global optimization on
the cluster potential-energy surface. After that, a section on cluster thermodynamics follows. The
discussion includes the analysis of solid-solid structural transitions and of melting, with its size
dependence. The last section is devoted to the growth kinetics of free nanoclusters and treats the
growth of isolated clusters and their coalescence. Several specific systems are analyzed.

CONTENTS 2. Computational methods 398


3. Size dependence of the melting point 400
C. Studies of selected systems 403
I. Introduction 371
1. Lennard-Jones clusters 403
II. Experimental Methods for Free-Nanocluster
2. Sodium clusters 403
Production 373
3. Noble-metal and transition-metal clusters 404
III. Energetics of Free Nanoclusters 374
4. Silicon clusters 405
A. Geometric shells: Structural motifs and general
V. Kinetic Effects in the Formation of Nanoclusters 405
trends in energetics 375
A. Freezing of liquid nanodroplets 406
B. Electronic shells 378
1. Lennard-Jones clusters 406
C. Calculation of the total energy of nanoclusters 378
2. Silver clusters 406
D. Global optimization methods 380
3. Gold clusters 407
E. Studies of selected systems 383
4. Entropic effects and kinetic trapping in Pt55 408
1. Noble-gas clusters 384
5. Copper, nickel, and lead clusters 409
2. Alkali-metal clusters 385
B. Solid-state growth 409
3. Noble-metal and quasi-noble-metal clusters 385
1. Universal mechanism of the solid-state
a. Gold clusters 385
growth of large icosahedra 409
b. Silver clusters 387
2. Growth of small noble-metal and
c. Copper clusters 388
transition-metal clusters 410
d. Platinum clusters 388
3. Growth of intermediate- and large-size Ag
e. Palladium clusters 389 and Au
f. Nickel clusters 389 clusters 412
4. General structural properties of noble-metal 4. Aluminum clusters 413
and quasi-noble-metal clusters 390 5. Clusters of C60 molecules 414
5. Silicon clusters 391 C. Coalescence of nanoclusters 415
6. Clusters of fullerene molecules 393 VI. Conclusions and Perspectives 416
IV. Thermodynamics of Free Nanoclusters 393 Acknowledgments 417
A. Entropic effects and solid-solid transitions 394 References 417
1. Structural transitions in the harmonic
approximation 395
2. Anharmonic and quantum corrections 396
I. INTRODUCTION
B. Melting of nanoclusters 397
In the last decade, we have seen the explosive devel-
1. Experimental methods 397
opment of a new field, now commonly known as nano-
science 共Nalwa, 2004兲. This field extends through phys-
ics, chemistry, and engineering and addresses a huge
*Present address: The Abdus Salam International Center for number of important issues, ranging from basic science
Theoretical Physics, Strada Costiera 11, I34014 Trieste, Italy. to a variety of technological applications 共in the latter
Electronic address: baletto@ictp.trieste.it case, the word nanotechnology is often employed兲. The

Electronic address: ferrando@fisica.unige.it purpose of nanoscience and nanotechnology is to under-

0034-6861/2005/77共1兲/371共53兲/$50.00 371 ©2005 The American Physical Society


372 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 2. High-resolution electron microscopy image of a 8.6


⫻ 6.3-nm2 truncated decahedral gold particle deposited on
amorphous carbon. The particle was produced in an inert-gas
FIG. 1. High-resolution electron microscopy image of Ag clus- aggregation experiment and then deposited and observed.
ters deposited on an inert substrate after being produced in an Adapted from Koga and Sugawara, 2003.
inert-gas aggregation experiment. The clusters indicated by ar-
rows are identified as being icosahedra. Adapted from Rein-
hard, Hall, Ugarte, and Monot, 1997. geneous. They may be neutral or charged. They may be
held together by very different kinds of forces: strong
attraction between oppositely charged ions 共as in NaCl
stand, control, and manipulate objects of a few nano- clusters兲, van der Waals attraction 共as in He and Ar clus-
meters in size 共say, 1 – 100 nm兲. These nano-objects are ters兲, covalent chemical bonds 共as in Si clusters兲, or a
thus intermediate between single atoms and molecules metallic bond 共as in Na and Cu clusters兲. Small clusters
and bulk matter. Their properties are often peculiar, be- of metal atoms are held together by forces more like
ing qualitatively different from those of their constituent those of covalent bonds than like the forces exerted by
parts 共either atoms or molecules兲 and from those of mac- the nearly free electrons of bulk metals.
roscopic pieces of matter. In particular, nano-objects can Clusters containing no more than a few hundred par-
present properties that vary dramatically with size. This ticles 共diameters of 1 – 3 nm兲 are expected to have
opens the possibility of controlling these properties by strongly size-dependent properties 共for example, geo-
controlling precisely their formation process. metric and electronic structure, binding energy, melting
Among nano-objects, nanoclusters occupy a very im- temperature兲. Larger clusters, with many thousands of
portant place, since they are the building blocks of nano- atoms and diameters in the range of 10 nm and more,
science. Nanoclusters are aggregates of atoms or mol- have a smoothly varying behavior, which tends to the
ecules of nanometric size, containing a number of bulk limit as size increases.
constituent particles ranging from ⬃10 to 106 共Castle- Nanoclusters have peculiar properties because they
man and Bowen, 1996; Johnston, 2002; Wales, 2003兲. are finite small objects. To finite objects, the constraint
In contrast to molecules, nanoclusters do not have a of translational invariance on a lattice does not apply.
fixed size or composition. For example, the water mol- For this reason, clusters can present noncrystalline struc-
ecule contains one oxygen and two hydrogen atoms, tures, icosahedra and decahedra being the best known.
which are placed at a well-defined angle to each other. Of course, it is possible also to build up crystalline clus-
On the other hand, silver, gold 共see Figs. 1 and 2兲, or ters, which are simply pieces of bulk matter. An impor-
even water clusters may contain any number of constitu- tant issue in cluster science is to understand whether
ent particles and, for a given size, present a variety of crystalline or noncrystalline structures prevail for a
morphologies. There are borderline cases that are not given size and composition. Being small objects, nano-
easily classifiable unambiguously either as clusters or clusters have a very high surface/volume ratio. Thus the
molecules, the fullerene buckyball 共C60兲 being an ex- surface energy contribution 共including terms from facets,
ample. In the following, we shall classify C60 as a mol- edges, and vertices兲 is not negligible and usually strongly
ecule, since clusters and solids of buckyballs have been size dependent.
produced in which each C60 constituent preserves its in- Nanoclusters are well suited for several applications,
dividuality. Clusters can be homogeneous, that is, com- whose number is rapidly increasing. There has been a
posed of only one type of atom or molecule, or hetero- traditional interest in applications to catalysis 共see, for

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 373

example, Henry, 1998兲, because of the very favorable out by some kind of simulation. In the words of Marks
surface/volume ratio of nanoclusters. More recently, 共1994兲 “small particle structures cannot be understood
there have been developments towards biological uses. purely from experimental data, and it is necessary to
For example, gold nanoparticles studded with short seg- simultaneously use theoretical or other modeling.” The
ments of DNA 共Alivisatos et al., 1996; Mirkin et al., discussion of theoretical and simulation results will
1996兲 could form the basis of an easy-to-read test to make frequent reference to the experiments.
single out genetic sequences 共Alivisatos, 2001兲. Some ap- We cannot pretend to be exhaustive on the subjects of
plications of nanoclusters have a much older history. It our review, since thousands of articles about nanoclus-
has been recently discovered 共Borgia et al., 2002; ters have been produced in the last decade. Some fasci-
Padeletti and Fermo, 2003兲 that the Renaissance masters nating fields, for example, the field of binary clusters and
in Umbria, Italy, used nanoparticles in the decoration of nanoalloys, whose properties depend crucially not only
majolicas with lustre. Lustre consists of a thin film con- on size but also on composition, are completely left out
taining silver and copper clusters with diameters up to a for space reasons. But even on the topics that we explic-
few tens of nanometers, often of noncrystalline struc- itly treat we do not claim to be exhaustive. If some con-
ture. Due to the inclusion of these nanoparticles, lustre tributions have been left out, we apologize in advance.
gives beautiful iridescent reflections of different colors. The review is structured as follows. In Sec. II we give
The starting point for an understanding of cluster a brief description of the methods for free-cluster pro-
properties is the study of their structure. With this goal duction. In Sec. III we deal with the energetics of free
in mind, the first question to answer is the following: nanoclusters, describing general trends, modeling the el-
Given size and composition, what is the most stable clus- ementary interactions, and finding the global minima of
ter structure from an energetic point of view? To answer the PES. In Sec. IV we consider the thermodynamics of
to this question we need to find the global minimum on nanoclusters, focusing on the possibility of solid-solid
the potential-energy surface 共PES兲 of the cluster 共Wales, transitions at finite temperature and on the melting tran-
2003兲. The PES is the product of the elementary inter- sition. Finally Sec. V treats the growth kinetics of nano-
actions among the cluster constituents. clusters, which may take place either in the liquid or in
Once the answer to this first question is given, a sec- the solid state, and the coalescence of nanoclusters. In
ond one arises: What is the effect of raising the tempera- each part, specific systems are treated in details.
ture on the structural properties of a cluster? The study
of this question introduces us to a fascinating field, the II. EXPERIMENTAL METHODS FOR FREE-NANOCLUSTER
thermodynamics of finite systems. PRODUCTION
Finally, the experimental time scales of cluster pro-
duction are often short with respect to the time scales of The experimental methods for producing free nano-
morphology transitions. This leads to a third important clusters have been reviewed by de Heer 共1993兲 and more
question: What are the kinetic effects in the formation recently by Milani and Iannotta 共1999兲 and Binns 共2001兲.
of nanoclusters? To answer to this question, the cluster Here we present a brief summary, focusing on those as-
growth process must be analyzed. pects 共such as the time scale of free-nanocluster growth兲
In this review, we try to summarize and discuss the most closely related to the subjects treated below.
present knowledge about these three points and, to the Sources producing free beams of nanoscale clusters
best of our ability, to answer the above questions con- were made available about 30 years ago for the produc-
cerning the behavior of free clusters. Two general results tion of noble-gas clusters 共Raoult and Farges, 1973兲 and
can, however, be anticipated. more than 20 years ago for producing metallic clusters
First, what clearly emerges from the comparison of 共Sattler et al., 1980; Dietz et al., 1981兲; more recent de-
experimental and theoretical results is that all three fac- velopments have made available sources that are ultra-
tors, energetics, thermodynamics, and growth kinetics of high vacuum compatible, that can produce binary clus-
nanoclusters, must be taken into account when analyz- ters 共Rousset et al., 1996; Cottancin et al., 2000兲, and that
ing realistic situations. Second, interactions among the incorporate extremely efficient mass-selection tech-
constituent particles must be understood in terms of the niques 共see Binns, 2001, and references therein兲.
basic concepts of range and type. The interaction range In most cases, the heart of a cluster source is a region
rules the behavior of systems with pair potentials 共see, where a supersaturated vapor of the material forming
for example, Wales, 2003兲. In systems characterized by the clusters is produced. The first step in cluster produc-
interactions with strong many-body character, metals, tion is the heating of the material to obtain a hot vapor.
for example, other factors come into play together with This can be done in different ways, for example, by heat-
the interaction range, namely, the bond-order/bond- ing up a piece of bulk material in a crucible 共as in seeded
length correlation and the directionality of the interac- supersonic nozzle and inert-gas aggregation sources兲, or
tions. by hitting a target with a laser pulse or an ion beam
The focus of this review article is on theoretical and 共laser evaporation and ion sputtering sources兲. To obtain
simulation issues in free-cluster science. We remark that supersaturation, the hot vapor must be cooled down.
cluster science is a field in which the interplay among There are essentially two ways to achieve this. The first
experiments, theory, and simulations is very active: in is by means of a supersonic expansion 共as in seeded su-
fact, the analysis of an experiment is very often carried personic nozzle sources, where the material is mixed

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


374 F. Baletto and R. Ferrando: Structural properties of nanoclusters

with a high-pressure inert gas and then expanded兲, squared modulus of the Fourier transform of the scatter-
which causes an adiabatic cooling 共Anderson and Fenn, ers with respect to the momentum transfer q, one builds
1985兲. Hot clusters are usually produced at temperatures a theoretical diffraction pattern by weighting contribu-
close to the evaporation limit 共Bjørnholm et al., 1991兲. In tions from clusters of different structures. In fact, one
the second way, the hot vapor is mixed with a cold inert- can expect that, for example, an icosahedron has a dif-
gas flow, which acts as a collisional thermostat. This ferent Fourier transform than a part of an fcc lattice.
method is used in inert-gas aggregation sources to pro- The geometry is obtained directly by fitting the weights
duce very cold metallic clusters. Here, due to the low of the different structures to the experimental data. On
temperature of the inert gas, cluster production pro- the other hand, information on the size can be obtained
ceeds mainly by the addition of single atoms, and re- from the width of the diffraction rings, because the
evaporation is negligible 共de Heer, 1993兲. When growth larger the cluster, the narrower the diffraction rings. The
ceases, the clusters can be further reheated or cooled more scatterers add their contributions coherently, the
down by subsequent stages of contact with gas at differ- sharper the resulting pattern. Finally, information about
ent temperatures. the cluster temperature is contained in the temperature
What are the typical time scales of nanocluster pro- dependence of the scattered intensity
duction? This question is extremely important, because,
as we see in the following, the finite lifetime of free clus- I共T兲 = I共T=0兲e−2W . 共2兲
ters in flight may cause their trapping into metastable Here W is the Debye-Waller factor,
structures. Let us consider an inert-gas aggregation
source for noble-metal clusters 共Reinhard, Hall, Ugarte, 1
W = 具u2典q2 , 共3兲
and Monot, 1997; Reinhard et al., 1998; Koga and Sug- 3
awara, 2003兲. There, the typical metal vapor tempera-
where 具u2典 is the mean-square vibrational amplitude of
ture Tv and pressure pv are of the order of 1000– 1500 K
the cluster atoms. For harmonic vibrations, 具u2典 ⬀ T.
and 1 – 10 mbar, respectively. Consider now a growing
nanocluster, say with a radius R of 1 nm, which can cor-
respond to 100–200 atoms. If we assume a spherical clus- III. ENERGETICS OF FREE NANOCLUSTERS
ter, the kinetic theory of gases gives for the atomic flux
⌽v 共i.e., the number of metal atoms per second hitting At low temperatures, the most favored structure of a
the cluster surface兲 cluster of N particles is the one that minimizes its total
energy. For example, in atomic clusters, the most fa-
pvAeff vored structure is the global minimum of the potential
⌽v = 共1兲
冑2 ␲ m v k B T v , energy as a function of the coordinates of the atomic
cores 共the potential-energy surface or PES兲. Tradition-
where Aeff = 4␲R2, and mv is the mass of the atoms. This ally, great effort has been devoted to finding reliable
gives ⌽v ⬃ 107 s−1, corresponding to an interval between methods for calculating the total energy, and searching
depositions ␶dep = ⌽−1 for local and global minima. This task implies the fol-
v ⬃ 10 ns. Thus a cluster of 10 at-
2 3

oms is grown on a time scale of a fraction of a millisec- lowing two steps:


ond. 共a兲 Construction of a model for the interactions be-
Once clusters are produced, they have to be detected tween the elementary constituents of the cluster;
in some way, possibly while they are still in their beam. this can be accomplished either by trying to solve
Indeed, the detection of slow neutral clusters is a diffi- directly the Schrödinger equation 共ab initio meth-
cult task; however, the cluster structures can be investi- ods兲 or by constructing semiempirical interparticle
gated by diffraction methods. They can then be ionized potentials 共see Sec. III.C兲.
for an efficient mass-selective detection and finally de-
posited and observed by several microscopy techniques 共b兲 A search for the most favored isomers by some
共Henry, 1998; José-Yacamán, Ascencio, et al., 2001兲. global optimization algorithm 共see Sec. III.D兲.
In a diffraction experiment, a well-collimated electron Depending on material and size, both 共a兲 and 共b兲 may
beam with an energy of 30 to 50 keV crosses a cluster present enormous difficulties. Therefore, as a prelimi-
beam. The fast electrons are scattered from the cluster nary step, it is extremely important to find general
atoms, and the diffraction pattern is recorded. A series trends that help to single out sequences of favorable
of diffraction rings around the position of the primary structures in different size ranges. This can be done ei-
electron beam is recorded. The interpretation of diffrac- ther on the basis of geometric considerations, as in Sec.
tion profiles is not straightforward 共Hall et al., 1991; III.A 共where the construction of families of highly sym-
Reinhard, Hall, Berthoud, et al., 1997, 1998; Reinhard, metric structures, the structural motifs, is treated兲, or
Hall, Ugarte, and Monot, 1997兲, being based on a fit to a based on electronic shell effects, as in Sec. III.B. The
theoretical profile whose construction assumes the pres- sizes of the most energetically stable structures are often
ence of some selected cluster structures. However, infor- called magic sizes. Magic sizes may 共tentatively兲 corre-
mation on the geometry, the average size, and the tem- spond either to the completion of a geometrically per-
perature of the clusters in the beam can be extracted as fect structure 共geometric magic sizes兲 or to the closing of
follows. Remembering that the scattered intensity is the an electronic shell 共electronic magic sizes兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 375

FIG. 4. Face-centered-cubic clusters: 共a兲 octahedron; 共b兲 trun-


cated octadedron; 共c兲 cuboctahedron. Each cluster is shown in
FIG. 3. Qualitative behavior of ⌬ 关Eq. 共5兲兴 for crystalline, four views. 共a兲 An octahedron is made up of two square pyra-
icosahedron, and decahedron clusters. mids sharing a common basis. Its surface consists of eight tri-
angular close-packed 共111兲 facets, but the structure has a high
surface/volume ratio. Polyhedra with a lower surface/volume
Section III.E is devoted to the study of selected sys- ratio, are obtained by truncating symmetrically the six vertices
tems of special interest. of an octahedron, thus obtaining square and hexagonal 共or tri-
angular, see below兲 facets. A truncated octahedron can be
characterized by two indexes: nl is the length of the edges of
A. Geometric shells: Structural motifs and general trends the complete octahedron; ncut is the number of layers cut at
in energetics each vertex. In the figure, for the octahedron in 共a兲 共nl , Ncut兲
= 共7 , 0兲, the truncated octahedron in 共b兲 共nl , Ncut兲 = 共7 , 2兲, and
In general, the binding energy Eb of a cluster of size N the cuboctahedron in 共c兲 共nl , Ncut兲 = 共7 , 3兲. A perfect truncated
with a given structure can be written in the form1 octahedron has thus a number of atoms, NTO共nl , ncut兲 = 31 共2n3l
3 2
+ nl兲 − 2ncut − 3ncut − ncut. This equation defines the series of
Eb = aN + bN2/3 + cN1/3 + d, 共4兲 magic numbers for truncated octahedron structures. The
square facets have a 共100兲 symmetry and edges of ncut + 1 at-
where the first term corresponds to a volume contribu-
oms. The 共111兲 facets are not in general regular hexagons. In
tion, while the others represent surface contributions
fact, three edges of the hexagons are in common with square
from facets, edges, and vertices. Volume and surface facets, thus having ncut + 1 atoms, while the remaining three
contributions are in competition. Clusters with low sur- edges have nl − 2ncut atoms. Regular hexagons are thus possible
face energy must have quasispherical shapes 共thus opti- if nl = 3ncut + 1; truncated octahedra with regular hexagonal fac-
mizing the surface/volume ratio兲, and close-packed fac- ets are referred to as regular truncated octahedra. When nl
ets. On the other hand, it is not possible to build up = 2ncut + 1 the hexagonal facets degenerate to triangles and the
clusters of spherical shape without internal strain, which cuboctahedron is obtained, which is usually not energetically
gives a volume contribution. favored because of its large 共100兲 facets.
A useful parameter for comparing the stability of clus-
ters in different size ranges is ⌬共N兲,
Let us now build up structural motifs by trying to op-
Eb共N兲 − N␧coh timize either volume or surface energy contributions.
⌬共N兲 = , 共5兲
N2/3 The easiest way to minimize volume contributions is to
where ␧coh is the cohesive energy per particle in the bulk cut a piece of bulk matter so that interparticle distances
solid and ⌬ is the excess energy 共that is, the energy in inside the cluster are automatically optimized. For such
excess with respect to N atoms in a perfect bulk crystal兲 clusters of crystalline structure the parameter a in Eq. 共4兲
divided approximately by the number of surface atoms. is simply ␧coh, so that limN→⬁ ⌬ = b. As we shall see in the
Other indicators of structural stability are the binding following, nanoclusters can be also of noncrystalline
energy per atom, Eb共N兲 / N, and the first and second dif- structures; for these clusters a is larger than ␧coh, and ⌬
ferences ⌬1共N兲 and ⌬2共N兲 in the binding energy, diverges at large sizes 共see Fig. 3兲.
Consider now fcc crystalline structures. Try to cut a
⌬1共N兲 = Eb共N − 1兲 − Eb共N兲, cluster from a bulk fcc crystal in such a way that its
surface has only close-packed facets. A possible result-
⌬2共N兲 = Eb共N − 1兲 + Eb共N + 1兲 − 2Eb共N兲. 共6兲 ing shape is the octahedron 共see Fig. 4兲, that is, two
⌬1 and ⌬2 measure the relative stability of clusters of square pyramids that share a basis. Even if the whole
nearby sizes. Peaks in ⌬2共N兲 were found to be well cor- surface of the octahedron is close packed, its shape does
related to peaks in the mass spectra 共Clemenger, 1985兲. not optimize the surface energy because of its high
surface/volume ratio. Clusters with more spherical
shapes are obtained by cutting the vertices, thus produc-
1
See, for example, Hill 共1964兲, Northby et al. 共1989兲, Xie et al. ing a truncated octahedron. Its surface has eight close-
共1989兲, Cleveland and Landman 共1991兲, Jortner 共1992兲, Uppen- packed 共111兲 and six square 共100兲 facets; the latter have
brink and Wales 共1992兲, Baletto, Ferrando, et al. 共2002兲. a higher surface energy in most materials. A deeper cut

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


376 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 5. Icosahedral clusters. The Mackay icosahedron is a


noncrystalline structure organized in shells. An icosahedron
with k shells has NIh共k兲 = 10 2 1
3 k − 5k + 3 k − 1 atoms 共so that the
3

series of magic numbers is 1, 13, 55, 147,…兲. The icosahedron


in the figure has k = 4 shells. An icosahedron with k shells has
the same number of particles as a cuboctahedron with ncut = k
− 1. An icosahedron has 20 triangular facets of side k and 12 FIG. 6. Decahedral clusters: 共a兲 regular decahedra; 共b兲 Ino
vertices. Each pair of opposite vertices lie along a fivefold sym- truncated decahedra 共Ino, 1969兲; 共c兲 Marks truncated decahe-
metry axis. An icosahedron can be thought of as composed of dra. A decahedron is made up of two pentagonal pyramids
20 fcc tetrahedra sharing a common vertex 共in the central site兲. sharing a common basis. It has a single fivefold axis and is
When 20 regular tetrahedra are packed around a common ver- formed by five tetrahedra sharing a common edge along the
tex, large interstices remain. To fill these spaces the tetrahedra fivefold axis. When five regular tetrahedra are packed, gaps
must be distorted, thus generating a huge strain on the struc- remain, but they are smaller than in the case of the icosahedra.
ture. Intershell distances are compressed, while intrashell dis- These gaps are filled by distorting the tetrahedra, thus intro-
tances are expanded. The facets of an icosahedron are of dis- ducing some strain. Regular decahedra 共first row兲 are limited
torted 共111兲 symmetry. Adatoms deposited on the facets of a by ten close-packed 共111兲-like facets, but have a large surface/
Mackay icosahedron can be placed either on sites of fcc or hcp volume ratio, which can be lowered by truncating the edges
stacking. Islands of fcc stacking are part of the next Mackay around the common basis, thus obtaining the Ino decahedron
shell, while islands of hcp stacking form a so-called anti- with five 共100兲-like facets. An even better structure is the
Mackay overlayer. Marks decahedron 共Marks, 1984兲, obtained by introducing re-
entrances that separate the 共100兲-like facets 共see the third row兲.
A decahedron is characterized by three integer indices
gives a more compact shape having, however, larger 共m , n , p兲, where m and n are the lengths of the sides of the
square facets. If size is sufficiently large, the optimal cut 共100兲 facets, perpendicular and parallel to the axis, respec-
is given by the Wulff construction. This was developed tively; p is the depth of the Marks reentrance. A regular deca-
to find the equilibrium shape of macroscopic crystals by hedron has indices of the form 共m , 1 , 1兲 关the 共5,1,1兲 decahedron
minimizing the surface energy at fixed volume 共see, for is shown in the top row兴; Ino decahedra have indices 共m , n , 1兲,
example, Pimpinelli and Villain, 1998兲. From the Wulff with n ⬎ 1 关the 共4,2,1兲 Ino decahedron is shown in the second
construction, the best truncated octahedron structure row兴, and Marks decahedra have 共m , n , p兲 with n , p ⬎ 1 关the
should fulfill the condition 共2,2,2兲 Marks decahedron is shown in the third row兴. A Marks
decahedron has h = m + n + 2p − 3 atoms along its symmetry axis
␥共100兲 d共100兲 and a total number of atoms given by NM−Dh = 共30p3 − 135p2
= , 共7兲 + 207p − 102兲 / 6 + 兵5 m3 + 共30 p − 45兲m2 + 关 60 共 p2 − 3 p兲 + 136兴 m其 / 6
␥共111兲 d共111兲
+ 兵n关15m2 + 共60p − 75兲m + 3共10p2 − 30p兲 + 66兴其 / 6 − 1. From this
where ␥共100兲 and ␥共111兲 are the 共100兲 and 共111兲 surface formula it follows that a Ino decahedron has NIno = 关5m3
energies, respectively, whereas d共100兲 and d共111兲 are the − 15m2 + 16m + n共15m2 − 15m + 6兲兴 / 6 − 1 atoms. For n = m and p
= 1 关square 共100兲-like facets兴 a decahadron has thus the same
distances of the facets from the center of the cluster. At
number of atoms as an icosahedron of m shells and as a cub-
large size, the introduction of higher-order facets can
octahedron with ncut = m − 1. Finally, a regular decahedron has
make fcc clusters more spherical 共see, for example, NDh = 共5m3 + m兲 / 6 atoms.
Raoult et al., 1989a兲. Different groups 共Cleveland and
Landman, 1991; Valkealahti and Manninen, 1998; Bal-
etto, Ferrando, et al., 2002兲 have shown that the Wulff are expanded. Therefore Mackay icosahedra are highly
construction is a reliable tool for identifying the best strained structures, and their ⌬ is proportional to N1/3 as
crystalline clusters for nanometric sizes. In any case, N → ⬁. This indicates that icosahedra could be expected
even the optimal Wulff shapes are quite far from being to be the most favorable structures only at small sizes.
spherical and are expected to be the most favorable clus- Icosahedra are not the only possible noncrystalline
ters at large sizes. structures. Another noncrystalline motif is represented
A better solution to the problem of building up com- by the decahedra 共see Fig. 6兲. A decahedron is formed
pact quasispherical shapes was found by Mackay 共1962; by two pentagonal pyramids with a shared base; its sur-
see also Martin, 1996兲, who constructed the Mackay face has only close-packed facets, but its shape is very
icosahedron 共see Fig. 5兲. This is a noncrystalline struc- far from being spherical, so that truncations are advan-
ture, with fivefold rotational axes. Icosahedral clusters tageous also in this case. Ino 共1969兲 proposed a trunca-
are limited by 共111兲-like close-packed facets only, thus tion in which the five edges limiting the common basis of
optimizing the surface energy well. However, this is ob- the pyramids are cut to expose 共100兲-like facets. This
tained at the expense of a volume contribution, since improves the surface/volume ratio, but usually does not
interatomic distances are not the ideal ones: radial 共in- produce the best possible decahedra in a given size
tershell兲 bonds are compressed, while intrashell bonds range 共see, for example, Baletto, Ferrando, et al., 2002兲

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 377

because it creates large 共100兲-like facets. Marks 共1984,


1994兲 proposed a more efficient truncation scheme, with
reentrances exposing further close-packed facets which
separate neighboring 共100兲-like facets. Marks decahedra
can achieve a better optimization of the surface energy
than truncated octahedron structures. On the other
hand, decahedra are also strained structures, with a vol-
ume contribution to the excess energy giving ⌬ ⬀ N1/3 at
large N. The strain, however, is much smaller than for
icosahedra.
In summary, the icosahedral motif should be the most
favored at small sizes, while truncated octahedron clus-
ters are expected for large sizes; truncated decahedra
could be favored in intermediate ranges. This trend has
been verified in experiments. Farges et al. 共1986兲 found a
transition from icosahedron to close-packed 关not neces- FIG. 7. 共Color in online edition兲 Morse potential for ␳0 values:
sarily fcc, see van de Waal et al. 共2000兲兴 structures at N dash-dotted line, sodium atoms; dashed line, Lennard-Jones
⬃ 750 in Ar clusters obtained in a free-jet expansion; molecules; and solid line, C60 molecules. The last is a very
Reinhard, Hall, Berthoud, et al. 共1997, 1998兲 were able sticky interaction.
to identify small icosahedra, intermediate-size decahe-
dra, and large fcc clusters in inert-gas aggregation ex-
periments on Cu. Several systems have been investi- UM = ⑀ 兺
i⬍j
e␳ 共1−r /r 兲关e␳ 共1−r /r 兲 − 2兴,
0 ij 0 0 ij 0 共9兲
gated theoretically, showing that the icosahedral
→ decahedral and decahedral→ fcc crossover sizes are
where the rij are the interatomic distances, r0 is the equi-
strongly material dependent.2
librium separation, and ⑀ is the well depth. Thus by vary-
Doye et al. 共1995兲 proposed a quite simple principle
ing the parameter ␳0 one can adjust the width of the
for a qualitative understanding of crossover sizes. This
potential well without changing the position or depth of
principle states that soft interactions, with wide potential
wells, stabilize strained structures, while sticky interac- the minimum. Large values of ␳0 give short-ranged at-
tions, with narrow potential wells, cannot easily accom- tractions with a steep repulsive part, that is, a narrow
modate the strain and thus favor crystalline structures. well and a sticky interaction 共see Fig. 7兲. Small ␳0 corre-
Doye et al. 共1995兲 considered a pair potential and de- spond to soft potentials. As can be seen in Fig. 8, the
composed the cluster energy into three parts, most favored structure changes from icosahedral at
small ␳0 to decahedral in the intermediate range, and
Eb = − nNN␧NN + Estrain + ENNN , 共8兲 finally to close-packed clusters at high values of ␳0. A
qualitative idea of the trends among different materials
where nNN is the number of nearest-neighbor pairs, ␧NN can be obtained by fitting the parameters of the Morse
is their bond strength at the optimal distance, Estrain is potential. This fit gives large values of ␳0 for the interac-
the strain contribution due to the fact that some nearest- tion potentials between fullerene molecules 关␳0 = 13.62
neighbor pairs can be at nonoptimal distances, and ENNN and 11.92 for Girifalco 共1992兲 and Pacheco and Prates-
is the contribution from further neighbors, which is, to a
first approximation, negligible. The usual competition is
between the first and the second term in Eq. 共8兲; icosa-
hedral structures, which have the largest number of
nearest-neighbor bonds, optimize the first term at the
expense of the second, while the opposite holds for fcc
clusters. The relative weights of these two terms depend
on the range of the potential. Decreasing the range has
the effect of destabilizing strained structures, because
the potential wells narrow so that the distortion of the
nearest-neighbor distance with respect to its ideal value
becomes more costly. In the case of a Morse 共1929兲 in-
teraction potential, Doye et al. 共1995兲 were able to con-
struct a structural phase diagram. The Morse potential
UM can be written as

2
See Raoult et al. 共1989b兲, Cleveland and Landman 共1991兲,
Uppenbrink and Wales 共1992兲, Turner et al. 共2000兲, Doye et al. FIG. 8. Phase diagram for Morse clusters: The lines separate
共2001兲, Baletto, Ferrando, et al. 共2002兲, Doye and Hendy 共2003兲 domains pertaining to different structural motifs. Figure cour-
and Sec. III.E for details on some specific systems. tesy of Jonathan Doye.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


378 F. Baletto and R. Ferrando: Structural properties of nanoclusters

Ramalho 共1997兲 potentials, respectively兴, ␳0 = 6 for a


Lennard-Jones crystal, ␳0 = 3.96 for Ni 共Stave and De-
Pristo, 1992兲, and even smaller values for alkali metals
关␳0 = 3.15 and 3.17 for sodium and potassium 共Girifalco
and Weizer, 1959兲兴.
Baletto, Ferrando, et al. 共2002兲 applied a similar crite-
rion, based on the stickiness of the interactions, to dis-
cuss crossover sizes in noble-metal and quasi-noble-
metal clusters modeled by many-body semiempirical
potentials. This point is discussed in detail in Sec.
III.E.3, where it is shown that there are other factors,
besides the interaction range, that determine the struc-
ture of metallic clusters 共Soler et al., 2000兲. These factors
arise from the many-body character of the metallic in-
teractions, which causes a strong correlation between
bond order and bond length, whose effects tend to make
icosahedral structures less stable than what follows from
the analysis of the interaction range. Moreover, bond
directionality effects can be important. Finally, the pres- FIG. 9. Mass spectrum of Nan clusters of size n, photoionized
ence of multiple minima or of secondary maxima in the with 3.02-eV photons. Closed-shell clusters are more difficult
two-body part of the interaction 共Doye and Wales, 2001; to ionize, so that they correspond to minima in the spectrum.
Doye, 2003兲 can have strong effects on the preferred Two sequences of minima appear in the spectrum. These se-
cluster structures. quences are at equally spaced n1/3 intervals on the size scale
and correspond to an electronic shell sequence and a structural
shell sequence. Adapted from Martin, 2000.
B. Electronic shells

Electronic shell closing has been extremely successful IV.B.3兲, plays a crucial role. The major evidence for elec-
in explaining the observed experimental abundances of tronic shell closing is for small alkali-metal clusters
alkali-metal clusters, like those found in the seminal ex- 共Knight et al., 1984; Bjørnholm et al., 1990; Nishioka et
periments of Knight et al. 共1984兲 on sodium clusters. In al., 1990兲. For these systems, several calculations 共Röth-
this framework, a cluster is modeled as a super-atom; lisberger and Andreoni, 1991; Spiegelman et al., 1998;
valence electrons are delocalized in the cluster volume Solov’yov et al., 2002兲 indicate that electronic shell clos-
and fill discrete energy levels. There are several degrees ing is a better criterion than atomic packing for deter-
of sophistication of this model. Detailed accounts can be mining the most stable clusters. For larger alkali clusters
found in the reviews by Brack 共1993兲 and de Heer the situation is more complicated. In fact, Martin et al.
共1993兲; a simple and very clear discussion is found in 共1991a兲 and Martin 共1996, 2000兲 have shown that Na
Johnston 共2002兲. Here we briefly sketch only the spheri- clusters, after presenting electronic magic numbers up to
cal jellium model. 2000 atoms, reveal a series of geometric magic numbers
The spherical jellium model assumes a uniform back- at larger sizes 共see Fig. 9兲. This would correspond to a
ground of positive charge, in which electrons move and transition from liquid to solid clusters at increasing size.
are subjected to an external potential. The simplest Moreover, when solid Na clusters are heated and
forms of the potential are the infinitely deep spherical melted, geometric magic numbers disappear to the ad-
well and the harmonic well. The solution of the single- vantage of electronic magic numbers. The same kind of
electron Schrödinger equation for the spherical well behavior was also found for aluminum clusters 共Martin
gives the following series of magic numbers: 2, 8, 18, 20, et al., 1992; Baguenard et al., 1994兲. Therefore the indi-
34, 40, 58,…, etc. On the other hand, the harmonic well cation is that electronic shells are seen when hot liquid
gives the series: 2, 8, 20, 40, 70,…, etc. The experimental clusters are produced, while geometric shells are exhib-
spectra of Knight et al. 共1984兲 reveal high peaks at 8, 20, ited by cold, solid clusters 共Johnston, 2002兲. In small
40 in agreement with both models. There are, however, clusters, both electronic and geometric effects can play
less evident peaks at 18 and 58, which appear only in the important roles, as Zhao et al. 共2001兲 have shown in their
spherical well. Experiments on larger sodium clusters tight-binding global optimization study of Ag clusters at
have revealed electronic shells in these clusters, up to N 艋 20. In conclusion, the interplay of electronic and
about 2000 atoms 共Martin et al., 1991a; Martin, 1996, geometric shell effects, depending on material and size,
2000兲, and in other metals; see, for example, Johnston remains to be fully understood.
共2002兲.
Candidates for observing electronic shell effects are C. Calculation of the total energy of nanoclusters
the metals with weakly bound valence electrons, prima-
rily the alkali and then the noble metals. It seems also A key point in the theoretical study of clusters is the
that the temperature T, which determines whether the choice of an appropriate energetic model. This depends
cluster is solid or liquid depending on N 共see Sec. on the material and the size of the cluster, as well as on

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 379

the physical and chemical properties one wishes to in- quantum Monte Carlo calculations in ordering the dif-
vestigate. To cover and illustrate in detail all the meth- ferent isomers of the Si20 cluster, even though there are
ods used in theoretical cluster science would require an still some quantitative differences.
entire book, and it is well beyond the scope of our re- Density-functional calculations can be inserted into a
view. Here we mention only the main methods, trying to molecular-dynamics procedure to give the ab initio mo-
give an idea of their underlying philosophy and range of lecular dynamics. The best known example is the Car-
applicability. Parrinello method 共Car and Parrinello, 1985兲, which
As pointed out in the review by Bonacić-Koutecký et could be well suited to the investigation of thermody-
al. 共1991兲, even clusters of a few atoms are very compli- namic and kinetic properties at T ⬎ 0 K.
cated systems. Indeed, the complexity of quantum me- Even though density-functional methods are nowa-
chanics forces one to employ approximate methods. days reliable and efficient for a large variety of systems,
The ab initio methods of quantum chemistry 关Hartree- there is still a great interest in developing methods re-
Fock and post Hartree-Fock; see Bonacić-Koutecký et quiring a smaller computational effort. In fact, global
al. 共1991兲, and references therein兴 were extensively ap- optimization 共see Sec. III.D兲 of clusters in the frame-
plied to the study of small clusters about 20 years ago. work of ab initio calculations is not feasible at present
For example, small-size 共2 艋 N 艋 9兲 Li and Na clusters except for a few systems and at very small sizes. More-
contain relatively few electrons, so that all-electron cal- over, ab initio molecular dynamics is limited to small
culations were possible 共Boustani et al., 1987兲. When ei- systems on short time scales, so that both the accurate
ther the size of the cluster or the nuclearity of the atoms sampling of thermodynamic properties and the simula-
increase, these methods become cumbersome, and at tion of kinetic processes 共diffusion, structural transfor-
present they are less commonly employed than in the mations, growth兲 are far beyond present capabilities.
past. Therefore several approximate energetic models for
Methods based on density-functional theory 共Hohen- clusters have been developed, often on semiempirical
berg and Kohn, 1964; Kohn and Sham, 1965兲, when ad- grounds. At an intermediate degree of computational
equately tested, can be of very high accuracy and less effort, there is the tight-binding model for semiconduc-
cumbersome from the computational point of view, tors 共Ho et al., 1998兲 and metals 共Barreteau, Guirado-
making it possible to treat a wide variety of systems and Lopez, et al., 2000; Bobadova-Parvanova et al., 2002兲.
somewhat larger sizes. Calculations on metals up to a This allows global optimization searches for clusters of a
few hundred atoms are present in the literature, even for few tens of atoms, and molecular-dynamics simulations
difficult systems such as the transition and noble metals even for clusters of 102 atoms 共Yu et al., 2002兲, even
共see, for example, Häberlen et al., 1996; Jennison et al., though on rather short time scales.
1997; Garzón, Michaelian, et al., 1998; Fortunelli and Larger sizes and longer time scales can be now treated
Aprà, 2003; Nava et al., 2003兲. The weak point in by classical atom-atom 共or even molecule-molecule兲 po-
density-functional calculations is often the exchange and tentials, which are built up on the basis of approximate
correlation term, which is treated in an approximate quantum models. These potentials contain parameters
way. The validity of the treatment depends on the sys- fitted to experimental material properties 共semiempirical
tems and has to be checked each time. The simplest ap- potentials兲 or to density-functional calculations 关ab-
proximation is the local-density approximation; more so- initio-based potentials 共Kallinteris et al., 1997; Garzón,
phisticated approaches include gradient corrections Kaplan, et al., 1998兲兴. For metallic systems, several atom-
共very important in transition and noble metals兲, which atom potentials have been developed, such as
are often called the generalized gradient approximation. embedded-atom 共Daw and Baskes, 1984; Voter, 1993兲,
At this level, different exchange and correlation func- glue 共Ercolessi et al., 1988兲, second-moment tight-
tionals are available 共see Perdew and Wang, 1986; binding 共Gupta, 1981; Rosato et al., 1989兲, Sutton-Chen
Becke, 1988, 1996; Perdew et al., 1992, 1996兲. In order to 共Sutton and Chen, 1990兲, and effective-medium 共Jacob-
check the validity of gradient-corrected density- sen et al., 1987兲 potentials. A discussion of the criteria
functional calculations, Mitás et al. 共2000兲 performed for fitting the parameters in Gupta 共1981兲 potentials is
quantum Monte Carlo simulations 共Ceperley, 1994兲 on found in López and Jellinek 共1999兲. For metals, these
small-size silicon clusters. In quantum Monte Carlo, potentials must contain a many-body term, which is re-
which is computationally very demanding, many-body sponsible for the correct surface relaxations 共Desjon-
correlations are directly taken into account by an ex- queres and Spanjaard, 1998兲. The advantage of this ap-
plicit correlation in the trial wave function. It turned out proach is that it allows full global optimization to sizes
that there were significant discrepancies between the of the order of ⬃200 atoms, local relaxation of clusters
density-functional 共with different types of exchange and of ⬃105 atoms, and molecular-dynamics simulations on
correlation functionals兲 and the quantum Monte Carlo time scales of 10 ␮s and more for clusters of 102 atoms,
results. Mitás et al. 共2000兲 confirmed the already known quite close to the growth time scales of free clusters in
bias of density-functional calculations towards compact inert-gas aggregation sources. The disadvantage is that
structures; this bias is very strong at the level of the even the qualitative accuracy of these potentials is often
local-density approximation, and considerably reduced questionable, and they must be tested carefully before
if gradient corrections are included. There are some being used 共see, for example, Ala-Nissila et al., 2002兲. As
functionals that give the same qualitative results as the we shall see in the following, there are several systems

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


380 F. Baletto and R. Ferrando: Structural properties of nanoclusters

for which reliable potentials have been developed. In


any case, the use of semiempirical potentials is in prac-
tice a necessary tool for the study of medium- or large-
size clusters. In contrast to bulk materials and crystal
surfaces, where one usually knows where to place the
atoms, the best structures of clusters are not known in
principle, so that a semiempirical modeling is the start-
ing point for more sophisticated approaches. For ex-
ample, in the work of Garzón, Michaelian, et al. 共1998兲,
global optimization by a semiempirical potential selects
the most promising candidates for a further relaxation
study by ab initio methods.
Semiempirical interaction potentials for molecular FIG. 10. 共Color in online edition兲 Transformation of the
clusters have been built up too. Here we note the Giri- potential-energy surface 共PES兲 to a staircase. Figure courtesy
falco 共1992兲 and Pacheco and Prates-Ramalho 共1997兲 po- of Giulia Rossi.
tentials for fullerene molecules. Finally we mention the
oldest semiempirical potentials, Lennard-Jones and
Morse potentials, which are the usual benchmarks for necessary at this stage: no global optimization technique
testing new theoretical tools. The Lennard-Jones poten- can warrant that the lowest minimum is really reached;
tial is also a popular model for noble gases. the only way to reach the global minimum with prob-
ability one is to sample all minima, compare them, and
choose the lowest one. This can be done only in essen-
D. Global optimization methods tially infinite time in cases of practical interest. For ex-
ample, while good putative global minima for Lennard-
Given the potential-energy surface of the cluster, that Jones clusters have been obtained at sizes well above
is, the potential energy U共兵r其兲 关where 兵r其 = 共r1 , r2 , . . . , rN兲兴 100 atoms, only for much smaller sizes have these
as obtained by methods like those of the previous sec- minima proven to be global 共Maranas and Floudas,
tion, one is confronted with the formidable task of find- 1992兲.
ing its deepest minimum. Indeed, Wille and Vennik Let us now analyze the good features of a PES which
共1985兲 demonstrated that this problem is NP-hard by a allows fast convergence to its global minimum and cor-
mapping to the traveling-salesman problem.3 The num- respondingly how an efficient global optimization algo-
ber of minima increases more than polynomially with rithm must be constructed to exploit these features
the size 关there are indications of a proportionality to 共Doye, 2004兲. To this end, a few definitions are needed
exp共N兲兴: a Lennard-Jones cluster of 13 atoms has about in the framework of what was proposed by Stillinger and
103 local minima 共Hoare and McInnes, 1976; Tsai and Weber 共1982兲 introducing the inherent structure of liq-
Jordan, 1993兲, but this number is at least 1012 for a 55- uids. Here below we follow the very clear exposition of
atom cluster 共Doye and Wales, 1995兲. Clearly, a com- Becker and Karplus 共1997兲. A thorough account is given
plete sampling of all these minima would be simply im- in the excellent book by Wales 共2003兲.
possible, and the ability of a given system to reach its We introduce a mapping from the continuum configu-
energy global minimum 共or at least one of the usually rational space of the cluster into the discrete set of its
few good local minima兲 should reside in some special local minima. The mapping associates each point 兵r其
features of its PES. This point has also been extensively with its closest minimum, i.e., the one reached by a
debated in the field of protein folding 共Wales, 2003兲, steepest-descent 共or a quenching兲 minimization starting
since proteins efficiently fold to their native state 共which at 兵r其. This amounts effectively to the following transfor-
may not coincide with the lowest minimum in the PES兲 mation of the PES 共see Fig. 10兲:
even if their PES presents a huge number of local
minima. A good search algorithm should exploit the fea- Ũ共兵r其兲 = min关U共兵r其兲兴, 共10兲
tures of the PES to ensure a fast convergence to low- where min means that the minimization is started from
lying minima. However, as we show below, depending on
兵r其. Ũ共兵r其兲 is a multidimensional staircase potential 共Li
the interaction potential, size, and composition of the
and Scheraga, 1987; Doye and Wales, 1998a兲.
cluster, there are easy potential-energy surfaces where
most algorithms converge fast to some good putative The set of all points associated with a minimum s con-
global minima, and others where there are low-lying stitutes its basin. All points of a given basin are con-
minima 共often of very high symmetry兲 that are ex- nected by definition. This mapping is a partition of the
tremely difficult to reach, so that most algorithms get 共3N − 6兲-dimensional space of the internal coordinates of
stuck in some less favorable configuration. A remark is the cluster into disjoint sets, which are indeed the attrac-
tion basins of the different minima. The boundaries be-
tween the basins constitute a network of 3N − 7 dimen-
3
NP stands for nonpolynomial. The computational complex- sions, where the mapping is not defined. Nearby basins
ity of this optimization increases more than polynomially with 共usually of the same structural family兲 can be grouped
cluster size. into metabasins, at different level of complexity. A very

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 381

Correspondingly, such systems are much easier to attack


by good global optimization algorithms than those with
rough PES or with multiple-funnel PES. As an example,
for a cluster of given size, icosahedral structures are usu-
ally grouped in a wide metabasin with a funnel structure,
which is easily accessible from liquidlike configurations
共Doye and Wales, 1998a兲, because the latter have in
common with the icosahedra a pronounced polytetrahe-
dral character 共Nelson and Spaepen, 1989; Doye and
Wales, 1996a兲, being formed by tetrahedral units packed
together. In contrast, decahedral and truncated octahe-
dral funnels are much narrower, so that global minima
pertaining to these motifs are not easily reached from
liquidlike configurations. The best-known example of
this kind is the Lennard-Jones cluster of 38 atoms, which
presents a double-funnel PES 共see Fig. 11兲, with a wide
icosahedral funnel and a narrower but deeper close-
packed funnel, so that the trapping of search algorithms
in the former is very likely. Moreover, one can expect
that systems with short-range 共sticky兲 potentials will
present a larger number of minima and a rougher PES
than systems with soft potentials 共Wales et al., 2000;
Doye, 2004兲. Multicomponent systems will be harder to
optimize because of their much smaller number of
equivalent permutational isomers with respect to pure
systems 共Darby et al., 2002兲.
From the above considerations, it is clear that an effi-
cient search algorithm must be able to perform and in-
tegrate the following tasks:
共a兲 From any given point on the PES, to find the local
minimum with which it is associated; this is simply
what is needed to construct the above-described
FIG. 11. 共Color in online edition兲 Disconnectivity diagrams:
mapping, or equivalently to transform the PES into
Top panel, schematic of a single-minimum potential-energy
a multidimensional staircase.
surface 共PES兲 with weak noise; center panel, a single-funnel
PES. The disconnectivity diagrams of these two panels show at 共b兲 To make transitions possible from a given basin to
which energetic level the different local minima of a PES can another and, more important, from one metabasin
be considered connected. Adapted from Becker and Karplus, to another. In this way, the algorithm would also be
1997. The lowest panel gives the disconnectivity diagram for able to explore a multiple-funnel PES.
the complete double-funnel PES of the Lennard-Jones cluster
of size 38. Figure courtesy of Jonathan Doye. The search by global optimization algorithms can be
unbiased when the starting configuration is randomly
chosen, or seeded when a set of 共supposedly兲 good struc-
convenient representation of the connections between tures is used to begin the optimization procedure.
basins, metabasins, etc. is given by the disconnectivity Seeded searches are often faster, since they use prior
diagram 共see Fig. 11兲, which allows one to understand knowledge about the system under study, but they have
pictorially which basins are connected at different values the disadvantage of making difficult the finding of unex-
of the total energy of the cluster 共two basins are con- pected low-lying minima. Almost every algorithm men-
nected when the highest point on the minimum-energy tioned below can be used either for unbiased or for
path between them lies below that total energy兲. seeded searches; obviously, when comparing different al-
Portions of the PES can be classified into three differ- gorithms, the unbiased search is more significant.
ent types according to the features of the connections Genetic algorithms have applications in a large variety
among the basins that they contain: rough PES, single- of fields; they are based on the analogy of evolution
minimum PES with weak noise, and funnel-like PES through natural 共fitness-based兲 selection. The fitness is
共see Fig. 11兲. It can be easily understood that the abso- the parameter to be optimized, here the potential en-
lute minimum is reached quickly in both a single- ergy. The coordinates of each cluster are encoded in a
minimum or a steep funnel-like portion of the PES, string of bits, called the chromosome. At each step, from
while a rough PES will resist global minimization. the present generation of clusters, a new generation is
Therefore we expect that in systems whose entire PES is built up. Sons are built up by mixing the chromosomes
given by a single minimum with weak noise or by a of parent clusters, or simply by inserting some mutation
single funnel the global minimum can be found quickly. in the present chromosomes. The individuals of the old

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


382 F. Baletto and R. Ferrando: Structural properties of nanoclusters

generation are compared to the sons via their fitness, Another popular algorithm is simulated annealing
and a new generation is formed from the old individuals 共Kirkpatrick et al., 1983; Biswas and Hamann, 1986;
and the sons, by some rule, which always includes the Freeman and Doll, 1996兲. In thermal simulated anneal-
best-fit individual in the new generation. Very often, sev- ing, the system is evolved at constant high T on the un-
eral populations are evolved in parallel, and individuals transformed PES by either Monte Carlo or molecular
are exchanged between them from time to time. In a dynamics, then slowly cooled down. Simulated anneal-
genetic algorithm, task 共b兲 is accomplished by chromo- ing has the advantage of being pretty much physical: one
some mixing and mutations and by exchanges of indi- tries to mimic the procedure of cooling a sample which
viduals between subpopulations, while task 共a兲 corre- hopefully will reach its most stable configuration if cool-
sponds to comparing the fitness of the individuals after a ing is sufficiently slow. This algorithm is easily incorpo-
local minimization on them, i.e., comparing Ũ共sons兲 to rated into standard Monte Carlo and molecular-
dynamics codes 共including ab initio molecular dynamics兲,
Ũ共parents兲. Recent developments include similarity
and the analysis of its output is easy and intuitive. For
checking among cluster structures to keep the diversity
these reasons, this algorithm has been used for a large
of the population as the genetic optimization goes on, as
in Cheng et al. 共2004兲. The use of genetic algorithms in variety of systems.4 There is, however, a major drawback
cluster optimization was pioneered by Hartke 共1993兲 and to simulated annealing: it utilizes effectively a single lo-
by Xiao and Williams 共1993兲, who made applications to cal optimization, and if the system fails to be confined to
Si4 and various molecular clusters, respectively. Their the basin of attraction of the global minimum as the
work was followed by applications to a wide variety of temperature is decreased, the algorithm may fail, even if
systems. Michaelian et al. 共1999兲 optimized transition it passed through the basin of the global minimum when
and noble-metal clusters by a symbiotic variant of a ge- T was high. As a result, simulated annealing is less effi-
netic algorithm 共Michaelian, 1998兲. Hartke 共2000, 2003兲 cient than genetic and basin-hopping algorithms
also treated water clusters; Deaven and Ho 共1995兲, 共Hartke, 1993; Zeiri, 1995兲. To avoid trapping in a single
Deaven et al. 共1996兲, and Ho et al. 共1998兲 optimized basin, one can make long high-T Monte Carlo or
Lennard-Jones and silicon clusters; Rata et al. 共2000兲 and molecular-dynamics runs, collecting a large quantity of
Bobadova-Parvanova et al. 共2002兲 considered silicon and different snapshots and quenching them down 共Sebetci
iron clusters; Darby et al. 共2002兲, Massen et al. 共2002兲, and Güvenc, 2003; Baletto et al., 2004兲, monitoring the
Bailey et al. 共2003兲, and Lloyd et al. 共2004兲 optimized short-term average of kinetic energy 共Jellinek and
Pd-Pt, Au-Cu, and Al-Ni nanoalloy clusters; finally Garzón, 1991; Garzón and Posada-Amarillas, 1996兲. T
Rossi et al. 共2004兲 searched for the global minima of must be chosen carefully. If it is too high, the probability
Ag-Cu, Ag-Ni, and Ag-Pd clusters. Recent reviews are of catching the global minimum is very small; if T is too
those of Hartke 共2002兲 and Johnston 共2003兲. low, the cluster is likely to remain trapped in the basin it
The basin-hopping algorithm 共Li and Scheraga, 1987; starts from.
Doye and Wales, 1998a; Doye, 2004兲 differs from genetic Another family of annealing algorithms is known as
algorithms because it accomplishes task 共b兲 simply by a quantum annealing 共Amara et al., 1993; Finnila et al.,
canonical Monte Carlo simulation at constant T on the 1994; Freeman and Doll, 1996兲. In quantum annealing,
transformed PES. In this framework, the transformation the system is first collapsed into its quantum ground
of the PES amounts to lowering the barriers between state by using diffusion or Green’s-function Monte Carlo
basins to the maximum possible extent, while keeping techniques 共Ceperley, 1994兲 and then quantum mechan-
the levels of the minima unchanged. Starting from a ics is slowly turned off. This algorithm utilizes delocal-
given basin, a move with random displacements is tried, ization and tunneling as the primary means for avoiding
and the energy difference between the new and the old trapping in metastable states. Amara et al. 共1993兲 and
position ⌬Ũ is calculated. If ⌬Ũ ⬍ 0 the move is ac- Finnila et al. 共1994兲 applied quantum annealing to
cepted with probability 1, otherwise with probability Lennard-Jones clusters up to N = 19. Lee and Berne
共2000兲 coupled quantum and thermal annealing, being
exp关−⌬Ũ / 共kBT兲兴. On Ũ, transitions can occur in any di- thus able to find the fcc global minimum of the Lennard-
rection, not only through saddle points, and transitions Jones cluster of size 38, but this approach failed for
to lower-lying minima are always accepted. Basin hop-
other difficult cases at larger size, such as the Marks
ping has been applied successfully to the optimization of
decahedron at N = 75. Recently, Liu and Berne 共2003兲
a large variety of systems, from Lennard-Jones clusters
have developed a new quantum annealing procedure,
共Wales and Doye, 1997兲 to clusters of fullerene mol-
based on quantum staging of path-integral Monte Carlo
ecules 共Doye et al., 2001兲, of transition and noble metals
sampling and local minimization of individual imaginary
共Doye and Wales, 1998b兲, and of aluminum 共Doye,
time slices. This method is able to locate all Lennard-
2003兲. Basin hopping was able to find the difficult puta-
tive global minima in Lennard-Jones clusters at sizes 38,
75, 76, 77, and 98, starting from random initial condi- 4
To cite but a few, Jones 共1991兲, Kumar and Car 共1991兲, Röth-
tions. Recently, Lai et al. 共2002兲 optimized sodium clus- lisberger and Andreoni 共1991兲, Jones and Seifert 共1992兲, Stave
ters by both genetic and basin-hopping algorithms, find- and DePristo 共1992兲, Vlachos et al. 共1992兲, Cheng et al. 共1993兲,
ing the same set of minima. This could indicate that both Christensen and Cohen 共1993兲, Jones 共1993兲, Ma and Straub
algorithms have comparable efficiency. 共1994兲, Ahlrichs and Elliott 共1999兲, Bacelo et al. 共1999兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 383

Jones minima for N 艋 100 except at N = 76, 77, 98, being creases with increasing potential energy, thus stabilizing
thus comparable to basin-hopping and genetic algo- the lower energy states and broadening the thermody-
rithms. This form of quantum annealing uses local mini- namics. The transformation to a staircase PES is not the
mization to accomplish task 共a兲, and mutations 共in which only useful transformation. Other possibilities have been
coordinates of randomly selected atoms are reset兲 to fa- suggested by Pillardy and Piela 共1995兲, Locatelli and
cilitate task 共b兲. Schoen 共2003兲, and Shao et al. 共2004兲.
An algorithm that tries to combine the advantages of What are the maximum sizes presently tractable by
simulated-annealing, genetic, and basin-hopping ap- global optimization? This answer depends on the system
proaches has been recently developed by Lee et al. under study and on the degree of sophistication of the
共2003兲. Their algorithm was able to optimize Lennard- interaction potential. In the Cambridge Cluster Data-
Jones clusters up to size 201. base 共see www-wales.ch.cam.ac.uk/CCD.html兲, putative
Finally, a very recent proposal is to search for global global minima for several systems are reported, ranging
minima by means of a growth procedure. Solov’yov et al. from Lennard-Jones to Morse, metallic, and water clus-
共2003, 2004兲 applied this method to Lennard-Jones clus- ters. Even for the simplest potentials, the largest sizes
ters, starting from a tetrahedron, and adding one atom are up to N = 190 关for Al 共Doye, 2003兲 and Pb clusters
at a time according to different procedures 共adding an 共Doye and Hendy, 2003兲 modeled by the glue potential
atom on the surface, inside the surface, or at the center 共Lim et al., 1992兲兴, and sizes decrease for more complex
of mass of the cluster兲. After adding a new atom, the interactions. Very recent seeded optimizations of
cluster is quenched down to search for the closest local Lennard-Jones clusters 共Shao et al., 2004兲 reached N
minimum, and the results of the different addition pro- = 330. In order to treat large sizes, Krivov 共2002兲 has
cedures are compared to select the best minimum. This proposed a hierarchical method for optimizing qua-
procedure exploits the fact that clusters of similar sizes siseparable systems. In these systems, distant parts are
are likely to have global minima with some resemblance, independent, so that a perturbation to the coordinates of
and in this respect it is a seeded algorithm. Solov’yov et a given part has very little influence on distant atoms.
al. 共2004兲 were able to find all global minima up to N According to Krivov 共2002兲, one can exploit this prop-
erty to build up a hierarchical procedure that can treat
= 150. Again, their procedure works on Ũ, since clusters Lennard-Jones clusters of several hundreds of atoms.
are compared for selection after local minimization. More complex systems such as water clusters have been
This last observation lead us to a summarizing com- optimized up to N ⯝ 30 共Wales and Hodges, 1998;
ment. The most successful algorithms make extensive Hartke, 2003兲. Ahlrichs and Elliott 共1999兲 optimized alu-
use of local minimization, comparing the structures for minum clusters within a density-functional approach up
selection after having performed that minimization. This to N = 15 by simulated annealing. Tekin and Hartke
amounts effectively to working on Ũ. To use the words 共2004兲 optimized Si clusters by a combination of an em-
of Doye 共2004兲, “the method for searching Ũ is of sec- pirical global search and density-functional local optimi-
ondary importance to the use of the transformation it- zation up to N = 16. In the case of tight-binding modeling
self.” The searching method may be based on thermal of the interactions, the sizes amenable by global optimi-
Monte Carlo, quantum tunneling, or genetic operations, zation approaches extend to a few tens of atoms. See,
but in the end the efficiencies are comparable, given that for example, Zhao et al. 共2001兲, who optimized Ag clus-
the algorithms are well constructed. ters up to N = 20 by a genetic algorithm.
Doye and Wales 共1998a兲 have shown that the transfor- How do the best algorithms scale with cluster size?
This might be an ill-posed question, since there is no
mation to Ũ modifies the thermodynamics of the system demonstration that the minima found by the different
in such a way that the temperature range in which tran- algorithms are really global. However, Liu and Berne
sitions among different funnels are possible is enlarged, 共2003兲 find that their quantum annealing algorithm
and the probability of occupying the basin of the global scales as N3.2 for Lennard-Jones clusters in the size
minimum is increased. This can be understood by ana- range 11–55, and Hartke 共1999兲 finds that his genetic
lyzing the occupation probability of a minimum s. For an algorithm scales approximately as N3, again for
untransformed PES within the harmonic approximation Lennard-Jones clusters up to N = 150.
共see Sec. IV兲 ps ⬀ exp共−␤Es兲 / ⍀s3N−6, where Es is the
potential energy of minimum s, ⍀s is the geometric
mean vibrational frequency, and ␤ = 共kBT兲−1, with kB the E. Studies of selected systems
Boltzmann constant. For a transformed PES, ps
⬀ As exp共−␤Es兲, where As is the hyperarea of the basin In this section we review the energetics of atomic clus-
of attraction of minimum s. The differences between ters of some selected materials, chosen both for their
these expressions, the vibrational frequency and hyper- importance from the point of view of basic science 共for
area terms, have opposite effects on the thermodynam- example, noble-gas and alkali-metal clusters, which have
ics. The higher energy minima are generally less rigid been the benchmark for testing theoretical models兲 and
and have a favorable vibrational entropy, so that the for their practical applications 共silicon clusters, transition
transitions are pushed down to lower temperature and and noble-metal clusters兲. Finally, we treat a specific
sharpened. By contrast, the hyperarea of the minima de- kind of molecular cluster, those of fullerene molecules,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


384 F. Baletto and R. Ferrando: Structural properties of nanoclusters

whose interpretation in energetic terms has been a long- 共2,2,2兲 and 共2,3,2兲 Marks decahedra, respectively. Later
standing puzzle. on, Leary and Doye 共1999兲 discovered that the global
minimum at N = 98 is tetrathedral. Romero et al. 共1999兲
1. Noble-gas clusters optimized clusters in the range 148艋 N 艋 309 on icosa-
hedral and decahedral lattices, finding again a preva-
Noble-gas clusters have been thoroughly studied since lence of icosahedron putative global minima with only
the early 1980s. Excellent accounts of their properties 11 decahedron exceptions.
are those of Haberland 共1994兲 and Johnston 共2002兲. At larger sizes, Raoult et al. 共1989b兲 compared perfect
Here we briefly review their energetics, focusing on the icosahedron and decahedron structures and concluded
transition sizes among structural motifs and comparing that Marks decahedra finally prevail over icosahedra at
theory with experiment. A special place among noble-
N ⬎ 1600. Concerning the transition to fcc structures,
gas clusters is occupied by He clusters, which are the
Xie et al. 共1989兲 compared icosahedral and cuboctahe-
best known example of quantum clusters. Recently
dral clusters, finding a crossover at N ⬃ 104. However,
there has been a noticeable experimental interest in He
cuboctahedra are not the most favorable fcc clusters.
clusters, which can be formed by 4He, 3He, or a mixture Therefore the crossover should be at much lower sizes
of the two.5 In the following, we treat “classical” clusters 共Raoult et al., 1989a; van de Waal, 1989兲, comparable to
first and then He clusters. the crossover size with Marks decahedra.
In a classical series of electron-diffraction experi- The combination of these energy-minimization results
ments, Farges et al. 共1983, 1986兲 investigated the struc- gives a reasonable interpretation of the experiments,
ture of neutral Ar clusters produced in a free-jet expan- even though there are some points that probably cannot
sion. They compared the experimental diffraction be explained in terms of the energetics alone. At small
patterns with those obtained by freezing Lennard-Jones sizes, the agreement between the global optimization
droplets of different sizes in molecular-dynamics simula- calculations and the experiments is good, even though
tions and were thus able to identify the experimental the decahedron global minima are not yet observed.
products up to N ⬃ 103 atoms. Harris et al. 共1984兲 mea- This may indicate the presence of entropic effects, of the
sured the mass spectra of charged Ar clusters at N kind discussed in Sec. IV.A, or that even at small sizes
⬍ 100. At 20⬍ N ⬍ 50, both experiments found polyi- some kind of kinetic trapping is taking place 共see Sec.
cosahedral structures, composed by interpenetrating 13- V.A兲. At larger sizes, the comparison of theory and ex-
atom icosahedra. From N = 50 to N ⬃ 750 multishell periment is much more difficult, and there has been con-
structures based on the Mackay icosahedron were siderable debate in the interpretation of the diffraction
found. Farges et al. 共1986兲 were also able to single out data in terms of known structures 共van de Waal, 1996兲.
features from diffraction by fcc planes in their spectra of For example, fcc clusters with twin faults give electron
clusters with N ⬎ 600 atoms. Thus they came to the con- diffraction patterns very similar to those of Marks deca-
clusion that the icosahedron→ fcc transition is placed at hedra. In any case, a coarse-grained picture is in agree-
N ⬃ 750. However, in a more recent analysis, van de ment with the general trend of a transition from the
Waal et al. 共2000兲 have shown that the transition is to a icosahedron motif to other structures. A more detailed
mixture of fcc, hcp, and random close-packed regions, analysis reveals that several kinds of clusters can be
with no significant preference for the fcc bulk structure. present in the same size range, so that structural transi-
Let us see how the energetics calculations compare tions are not really sharp. Quantitative agreement on
with these experimental results. Most of the calculations the transition size between the Lennard-Jones calcula-
are based on the use of the Lennard-Jones potential, tions and the experiments has not been reached, prob-
which is a very popular model for noble-gas systems ably due to the limited accuracy of the Lennard-Jones
even though not completely accurate. Several global op- potential for Ar. At present, it is very difficult to ascer-
timization studies have shown that the icosahedron mo- tain whether the diffraction data from large clusters sup-
tif is dominant at small sizes,6 with few exceptions. Poly- port the existence of either entropic or kinetic effects.
icosahedral clusters are formed at sizes 19, 23, and 26 Some interesting results addressing this point 共Ikeshoji
共Farges et al., 1988; Ikeshoji et al., 1996兲. Indeed, Wales et al., 2001兲 are discussed in Sec. V.A.
and Doye 共1997兲 found that there are only seven excep- Let us now consider 4He clusters. While the dimer is
tions in the range 13艋 N 艋 150: at N = 38 the global mini- stable but very weakly bound 共Luo et al., 1993; Grisenti
mum is a truncated octahedron; at N = 75, 76, 77 and N et al., 2000兲, the trimer 关which should have a noticeable
= 102, 103, 104 the best structures are based on the contribution from linear configurations according to the
calculations by Lewerenz 共1996兲兴 and larger clusters are
5
much more stable 共Toennies and Vilesov, 1998兲. Chin
Recent discussions of He clusters are those of Whaley and Krotscheck 共1992兲 calculated the ground-state prop-
共1994兲, Toennies and Vilesov 共1998兲, Scoles and Lehmann
erties of various 4He clusters modeled by the Aziz et al.
共2000兲, Callegari et al. 共2001兲, Northby 共2001兲, Johnston 共2002兲,
Jortner 共2003兲. 共1987兲 potential, considering several sizes up to N = 112,
6
See Hoare and Pal 共1975兲, Farges et al. 共1985兲, Freeman and by diffusion Monte Carlo simulations. They found evi-
Doll 共1985兲, Northby 共1987兲, Wille 共1987兲, Coleman and Shal- dence of density oscillations indicating a possible shell
loway 共1994兲, Xue 共1994兲, Pillardy and Piela 共1995兲, Deaven et structure. Chin and Krotscheck 共1995兲 confirmed the ex-
al. 共1996兲, Wales and Doye 共1997兲, Romero et al. 共1999兲. istence of these oscillations for larger clusters, as well,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 385

and were also able to recover the bulk limit of the exci- 共2003兲. Ishikawa et al. 共2001兲 reanalyze the problem of
tation spectrum. In a very recent experiment Brühl et al. Li6, in which planar and 3D structures were found to be
共2004兲 observed magic sizes at N = 10, 11, 14, 22, 26, 27, in competition, by a global optimization procedure em-
and 44 atoms, in 4He clusters produced in a free-jet ex- ploying the replica exchange method 共Swendsen and
pansion. By comparing the experimental results with dif- Wang, 1986兲 applied to ab initio calculations. They find a
fusion Monte Carlo calculations, Brühl et al. 共2004兲 3D structure of D4h symmetry group, which is slightly
showed that these magic sizes are not related to en- lower in energy than the planar triangle and the pen-
hanced binding energies at specific values of N, but to tagonal pyramid. Poteau and Spiegelmann 共1993兲 per-
the sizes at which excited levels cross the chemical- formed a search for the best Na isomers up to N = 34 by
potential curve and become stabilized. a growth Monte Carlo algorithm in the framework of a
Clusters of 3He are much less stable, because the con- distance-dependent tight-binding 共Hückel兲 model. They
stituent atoms are fermions, so that the minimum num- found that at N = 34 the best isomer is the double icosa-
ber of atoms Nmin needed to form a stable cluster is hedron surrounded by a ring of 15 atoms. Sung et al.
relatively large. A configuration-interaction calculation 共1994兲 optimized Li clusters up to N = 147 by simulated
based on a phenomenological density functional gave annealing within the local spin-density approximation to
Nmin = 29 共Barranco et al., 1997兲, while subsequent varia- the density-functional theory, finding that for 55艋 N
tional Monte Carlo calculations with the Aziz potential 艋 147 the structures based on the Mackay icosahedron
gave 34–35 as an upper bound to Nmin 共Guardiola, 2000; are the lowest in energy. This has been recently con-
Guardiola and Navarro, 2000兲. The same kind of calcu- firmed by Reyes-Nava et al. 共2002兲, who performed a
lations 共Guardiola and Navarro, 2002兲 have recently global genetic optimization of sodium modeled by
been applied to mixed 4He-3He clusters, finding that Gupta 共1981兲 potentials. A thorough global optimization
their stability has a nontrivial dependence on size and study 共by both genetic and basin-hopping algorithms兲 of
composition. Na, K, Rb, and Cs clusters in the framework of the
Gupta potential 共as developed by Li et al., 1998兲 has
2. Alkali-metal clusters
been performed by Lai et al. 共2002兲 up to N = 56. They
found a sequence based on icosahedron clusters except
The structural properties of small Li and Na clusters for N = 36 and N = 38. At N = 38 the global minimum is a
have been the subject of extensive theoretical activity, truncated octahedron, with distorted facets in the cases
following the seminal experiments of Knight et al. of Na and Cs; at N = 36 the structure is a distorted in-
共1984兲. Bonacić-Koutecký et al. 共1991兲 gave an excellent complete truncated octahedron.
account of the earlier developments, which are very The transition from electronic to geometric magic
briefly summarized here. Systematic ab initio numbers in Na clusters has already been discussed in
configuration-interaction studies on Li and Na clusters Sec. III.B. Here we simply recall that for 2000⬍ N
共Boustani et al., 1987; Bonacić-Koutecký et al., 1988兲 for ⬍ 20 000, there is experimental evidence 共Martin et al.,
N 艋 9 have shown a complete analogy between the two 1991a; Martin, 1996, 2000兲 of peaks in the abundances of
elements. The lowest isomers are planar up to the pen- Na clusters at sizes corresponding to the completion of
tamer; at N = 6 a pentagonal pyramid and a triangular perfect icosahedron, Ino decahedron 共Ino, 1969兲, or cub-
planar structure are in very close competition; larger octahedron structures, which have exactly the same
clusters are 3D. In agreement with the experimental magic numbers. From these magic numbers, one cannot
magic numbers and with the electronic shell closing ar- directly infer the actual cluster structures. However,
guments, the cluster of 8 atoms, of tetrahedral symmetry there is at least one indirect argument in favor of icosa-
共Jellinek et al., 1994兲, is especially stable compared to hedral structures: in fact, while complete icosahedra are
neighboring sizes. Röthlisberger and Andreoni 共1991兲 surely more favorable than incomplete ones, Ino deca-
and Röthlisberger et al. 共1992兲 performed Car-Parrinello hedra are not usually the best decahedral structures, nor
calculations on Na at slightly larger sizes. At N = 13, they are cuboctahedra the best among the possible truncated
found that the most stable isomer is neither an icosahe- octahedron clusters.
dron nor a cuboctahedron, but a capped pentagonal bi-
pyramid. At N = 18 the most stable isomer was the
double icosahedron minus one vertex, and at N = 20 the 3. Noble-metal and quasi-noble-metal clusters
isomers based on pentagonal symmetries 共which are Here we treat the energetics of pure clusters of the
closely related to the double icosahedron兲 were found to elements of the last two columns of the transition series:
be more stable than structures with tetrahedral symme- Ni, Cu, Pd, Ag, Pt, and Au. We restrict our treatment to
try. On the other hand, Bonacić-Koutecký et al. 共1991兲 neutral clusters unless otherwise specified.
found that two tetrahedron structures are the lowest in
energy for Li20 and Na20 共Bonacić-Koutecký, Fantucci, et
al., 1993兲. Spiegelman et al. 共1998兲 found results in rea- a. Gold clusters
sonable agreement with the previous calculations by a Small Au clusters 共N 艋 10兲 have been thoroughly
tight-binding approach. treated by ab initio methods. Bravo-Pérez et al. 共1999a兲
Recent developments are discussed by Ishikawa et al. studied sizes 3 艋 N 艋 6, finding that the best structures
共2001兲, Solov’yov et al. 共2002兲, and Matveentsev et al. are planar 共Bravo-Pérez et al., 1999b兲. These results have

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


386 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 13. 共Color in online edition兲 Two views of the lowest-


energy isomer of Au at N = 55 as found by Garzón, Michaelian,
et al. 共1998兲. This cluster is a strongly rearranged and distorted
55-atom icosahedron, similar to the double-rosette structure
FIG. 12. Lowest-energy isomers of Au clusters according to discussed in the case of Pt55 in Sec. V.A.4. Figure courtesy of
Bonacić-Koutecký et al. 共2002兲. All clusters are planar up to Ignacio Garzón.
N = 10 at least. The binding energy per atom Eb / N 共in eV兲 is
also shown. For isomers with energy difference smaller than
close competition, and the global minimum is sensitive
0.1 eV, both competing structures are shown 共see sizes 3, 4,
and 7兲. From Bonacić-Koutecký et al., 2002.
to the fine details of the potential parametrization. An
argument in favor of the low-symmetry structures fol-
lows from the density-functional relaxation at N = 75
been confirmed and extended by others.7 Au clusters are seen by Michaelian et al. 共1999兲, who found that a low-
planar up to N = 10 at least, according to Bonacić- symmetry structure is lower in energy than the 共2,2,2兲
Koutecký et al. 共2002; see Fig. 12兲, and only to N = 6 Marks decahedron, in contrast with all semiempirical re-
according to Wang et al. 共2002兲, who however found that sults. The physical origin of low-symmetry structures is
flat 共but nonplanar兲 structures are the lowest in energy discussed in Sec. III.E.4.
up to N = 14. In the experiments on Au cations by Gilb et For larger sizes, Cleveland et al. 共1997兲 and Baletto,
al. 共2002兲 there is evidence for planar clusters at least up Ferrando, et al. 共2002兲 performed semiempirical calcula-
to N = 7. The physical origin of the preference for planar tions to compare structures of the different motifs, find-
structures is discussed in Sec. III.E.4. On the other hand, ing that Marks decahedra are the most favorable, the
Li et al. 共2003兲 found experimental and computational crossover with fcc structures being at N ⬃ 500.
evidence that Au20 is a tetrahedron, well separated from The comparison of these theoretical predictions with
higher isomers. This finding has been confirmed by the the experiments is complicated by the fact that Au clus-
calculations of Wang, Wang, and Zhao 共2003兲. ters are very often passivated during or after the forma-
Global optimization methods 共Wilson and Johnston, tion process 共Whetten et al., 1996; Schaaff and Whetten,
2000兲 have been applied to larger clusters modeled by 2000兲, and it is difficult to determine to what extent the
semiempirical potentials. Garzón and co-workers action of the passivating agents can modify the struc-
共Garzón, Michaelian, et al., 1998; Garzón, Beltrán, et al., tures 共Alvarez et al., 1997兲. With this in mind, we try,
2003兲 and Michaelian et al. 共1999兲 modeled Au by the however, to determine how the calculated crossover
Gupta 共1981兲 potential and searched the best isomers by sizes compare with the experimental results. To this pur-
a genetic algorithm, to use them as the starting point of pose, we must single out those experiments in which the
a local relaxation by the density-functional method. observed clusters were able to reach their equilibrium
They found that Au presents low-symmetry structures at structure. This was very likely the case in the work of
some geometric magic numbers, such as N = 38 and 55, Patil et al. 共1993兲, who produced free Au clusters in
which therefore are not true magic numbers for Au clus- inert-gas aggregation sources. The clusters were subse-
ters. These structures are often called “amorphous” in quently slowly heated up above their melting tempera-
the literature, even though the most appropriate termi- ture and then slowly cooled down 共for about one sec-
nology is low-symmetry structures. The low-symmetry ond兲. Finally, the clusters were deposited and observed.
cluster at N = 55 is a strongly rearranged and distorted These clusters were identified as being fcc, even at the
icosahedron, which conserves some fivefold vertexes smallest sizes 共N ⯝ 400兲, in good agreement with the
共see Fig. 13兲. Low-symmetry structures are found in sev- crossover sizes calculated by Baletto, Ferrando, et al.
eral semiempirical calculations,8 the only exception be- 共2002; see Table I and Fig. 14兲. On the other hand, there
ing the Murrell-Mottram potential 共Cox et al., 1999兲, have been several observations of large icosahedral
which gives high-symmetry clusters. All these results in- structures,9 at N Ⰷ 400. For example, Ascencio et al.
dicate that high- and low-symmetry structures are in 共1998, 2000兲 have observed by high-resolution electron
microscopy a variety of structures 共decahedron, trun-
cated octahedron, icosahedron, and amorphous兲 for pas-
7
See Grönbeck and Andreoni 共2000兲, Häkkinen and Land-
man 共2000兲, Bonacić-Koutecký et al. 共2002兲, Gilb et al. 共2002兲,
9
Häkkinen et al. 共2002兲, Wang et al. 共2002兲, Zhao et al. 共2003兲. See Buffat et al. 共1991兲, Marks 共1994兲, Martin 共1996兲, Ascen-
8
See, for example, Doye and Wales 共1998b兲, Li, Cao, and cio et al. 共1998兲, Ascencio et al. 共2000兲, Koga and Sugawara
Jiang, 共2000兲, Darby et al. 共2002兲, Garzón et al. 共2002兲. 共2003兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 387

TABLE I. Potential parameters and cluster crossover sizes for several metals. Parameters p , q of the Rosato et al. 共1989兲 potential;
parameters ␴1 = pq / 2, ␴2, ␴3, and ␴A 关see Eqs. 共16兲 and 共17兲兴; sizes where ⌬ is minimum 共NIh ⌬ for icosahedra and N⌬ for
Dh

decahedra兲, and crossover sizes 共NIh→Dh, NDh→fcc, and NIh→fcc兲. in the case of the different noble and quasi-noble metals. The
parameters of the potentials are taken from Baletto, Ferrando, et al. 共2002兲 except for those of Ni, which are found in Meunier
共2001兲 and Baletto, Mottet, and Ferrando 共2003兲. The results on the crossover sizes are taken from Baletto, Ferrando, et al. 共2002兲,
except those for Ni which are published here for the first time. The results for ␴A are taken from Soler et al. 共2000兲.

Metal p q ␴1 ␴2 ␴3 ␴A NIh
⌬ NDh
⌬ NIh→Dh NDh→fcc NIh→fcc

Cu 10.55 2.43 13.1 0.062 0.35 0.24 309 20000 1000 53000 1500
Ag 10.85 3.18 17.2 0.065 0.29 0.54 147 14000 ⬍300 20000 400
Au 10.53 4.30 22.6 0.082 0.16 1.86 147 1300 ⬍100 500 ⬍100
Ni 11.34 2.27 12.9 0.055 0.37 0.05 561 17000 1200 60000 2000
Pd 11.00 3.79 20.9 0.062 0.21 0.78 147 5300 ⬍100 6500 ⬍100
Pt 10.71 3.85 20.6 0.073 0.22 1.40 147 5300 ⬍100 6500 ⬍100

sivated particles of a few nanometer’s diameter, with a Sutton and Chen 共1990兲 potential and optimized Ag
prevalence, however, of 共possibly defected兲 Marks and clusters at N 艋 80, finding the same lowest-energy struc-
Ino decahedra. Defected clusters 2 – 4 nm in diameter tures as Bonacić-Koutecký, Cespiva, et al. 共1993兲 at N
have been detected by x-ray diffraction 共Zanchet et al., = 7 , 8 , 9, and icosahedral global minima at N = 13 and 55.
2000兲. Moreover, in a recent experiment, Koga and Sug- Thirty of the global minima were icosahedral in charac-
awara 共2003兲 have analyzed a large sample of free Au ter, but several decahedron and fcc clusters were found,
clusters obtained in an inert-gas aggregation source, the most stable being at N = 38 共truncated octahedron兲
with diameters of 3 – 18 nm and found that icosahedra
were the most frequent and decahedra second-most fre-
quent, fcc clusters being absent. These results are again
an indication in favor of kinetic effects and are discussed
in Sec. V.B.3.

b. Silver clusters
As in the case of Au, the energetics of Ag clusters
have been the subject of intensive study in all size
ranges. For N 艋 9, Bonacić-Koutecký, Cespiva, et al.
共1993, 1994兲 searched for the lowest-energy clusters by a
self-consistent Hartree-Fock procedure which was able
to produce structures in good agreement with the ex-
periments 共Ganteför et al., 1990; Ho et al., 1990;
Alameddin et al., 1992; Jackschath et al., 1992兲. They
found that the best isomers are planar up to the pen-
tamer, as confirmed by Matulis et al. 共2003兲 by density-
functional calculations. Santamaria et al. 共1994兲 found a
planar structure also at N = 6. Recent density-functional
studies have considered N 艋 12 共Fournier, 2001兲, and N
= 13 共Oviedo and Palmer, 2002兲. The latter study sug-
gested that the best isomer of 13 atoms is of low sym-
metry, and that even the cuboctahedron is more favor-
able than the icosahedron. By contrast, previous density-
functional studies on larger Ag clusters 共Jennison et al.,
1997兲 found that at N = 55 the icosahedron is lower in
energy than the cuboctahedron.
Larger clusters have been studied mainly by semi-
empirical interatomic potentials. This approach has al-
lowed global-minimum searches for N up to 100 atoms FIG. 14. ⌬ as a function of N for different cluster motifs in Ni,
共Doye and Wales, 1998b; Erkoç and Yilmaz, 1999兲 and Cu, Pd, Ag, Pt, and Au: 䊊, iscosahedra; 䊐, decahedra; 䉭,
the comparison of selected magic structures belonging to trunacted octahedra. Data are taken from Baletto, Ferrando,
the icosahedral, decahedral, and truncated octahedral et al. 共2002兲, except for the data concerning Ni, which are origi-
motifs for N up to 40 000 atoms 共Baletto, Ferrando, et nal. For decahedra and truncated octahedra only the most fa-
al., 2002兲. Doye and Wales 共1998b兲 modeled Ag by the vorable clusters are shown in the different size ranges.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


388 F. Baletto and R. Ferrando: Structural properties of nanoclusters

and N = 71, 75 共decahedron兲. Mottet et al. 共1997兲 mod- atoms and that decahedron and truncated octahedron
eled Ag by the Rosato et al. 共1989兲 semiempirical poten- structures are in close competition up to 30 000 atoms at
tial, reproducing essentially the same behavior as the least 共see Table I兲. Contrary to the Ag case, this behav-
Sutton and Chen 共1990兲 potential at small sizes. They ior is in agreement with the inert-gas aggregation experi-
showed that clusters of high stability can be obtained by ments of Reinhard, Hall, Berthoud, et al. 共1997, 1998兲,
removing the central atom in perfect icosahedra. In fact, who were able to identify a prevalence of small icosahe-
a central vacancy allows neighboring atoms to better re- dra, intermediate-size decahedra, and large fcc clusters,
lax, expanding their intrashell distance. The same effect with a wide size interval in which decahedra and fcc
was also found in Cu and Au clusters. We note that, at clusters coexisted. The crossover size between icosahe-
variance with Au, there is no indication in Ag in favor of dron and decahedron structures is in agreement with the
disordered structures at geometric magic numbers 共38, calculations.
55, 75兲. Finally Baletto, Ferrando, et al. 共2002兲 and Mot-
tet et al. 共2004兲 looked at the crossover sizes 共see Table I兲
d. Platinum clusters
by an extensive comparison of clusters of the different
motifs. The energetic stability of small Pt clusters has been
All the previous theoretical results agree in predicting addressed by fewer studies than the other metals treated
that Ag follows the general trend outlined in Sec. III.A, here. Grönbeck and Andreoni 共2000兲 performed a
with small icosahedra, intermediate-size decahedra, and density-functional study in the size range 2 艋 N 艋 5, find-
large truncated octahedra clusters. This is in clear con- ing that the lowest-lying isomers are planar for both the
trast with the inert-gas aggregation experiments by tetramer and the pentamer. The preference for planar
Reinhard, Hull, Ugarte, and Monot 共1997兲, in which structures is in agreement with previous density-
small clusters 共⬃2 nm of diameter兲 were mainly decahe- functional calculations by Yang et al. 共1997兲, who also
dral 共or fcc兲 while large clusters 共above 5 nm兲 were found planar structures for N = 6 by a dynamical quench-
mostly icosahedral. This discrepancy can only be ex- ing procedure. In contrast, other ab initio calculations
plained by taking into account kinetic trapping effects 共Dai and Balasubramanian, 1995兲 on the tetramer gave
leading to the growth of metastable structures, as dis- preference to the tetrahedal structure, as in the density-
cussed in Sec. V.B.3. functional results of Fortunelli 共1999兲, who found that
the lowest-lying isomer is indeed the tetrahedron, with a
c. Copper clusters planar rhombic structure that was slightly higher in en-
ergy.
The energetics of small Cu clusters has been exten- Sachdev et al. 共1992兲 modeled Pt at larger sizes, by an
sively treated in a recent review by Alonso 共2000兲; here embedded atom potential and made simulated anneal-
we give a brief summary of the results for small sizes and ing calculations, finding that both at 13 and at 55 atoms
then focus on larger clusters. low-symmetry isomers are the lowest in energy. On the
Several ab initio calculations10 have been performed other hand, further global optimization results based on
for N 艋 10, again comparing a set of reasonably good semiempirical modeling 共Doye and Wales, 1998b; Bal-
structures. There is evidence of planar structures up to etto, Ferrando, et al., 2002; Massen et al., 2002; Sebetci
N = 5 or N = 6 maximum, as suggested by recent density- and Güvenc, 2003兲 gave preference to the icosahedral
functional calculations 共Jug, Zimmerman, Calaminici, et structure at N = 13. Doye and Wales 共1998b兲 and Massen
al., 2002兲. For N = 13, density-functional calculations et al. 共2002兲 found a low-symmetry structure at N = 55,
共Fujima and Yamaguchi, 1989兲 indicate that the icosahe- while Baletto, Ferrando, et al. 共2002兲 found the icosahe-
dron is favored over the cuboctahedron. dral structure also at this size, though with small cross-
For larger size, up to N ⬃ 100, global optimization over sizes among icosahedral, decahedral, and truncated
studies by semiempirical potentials have been per- octahedral structural motifs 共see Table I兲. Ab initio re-
formed. Doye and Wales 共1989b兲 used the Sutton and sults are even fewer. Yang et al. 共1997兲 found that at N
Chen 共1990兲 potential, while Darby et al. 共2002兲 used the = 13 several low-symmetry structures are lower in energy
Gupta 共1981兲 potential. In both cases, the best structure than the icosahedron and the cuboctahedron. Watari
at N = 13 is the icosahedron, in agreement with density- and Ohnishi 共1998兲 found that the cuboctahedron is
functional calculations, and there is no indication in fa- more stable than the icosahedron at N = 13 by density-
vor of disordered structures at geometric magic num- functional calculations. Recently, Fortunelli and Aprà
bers. Cu thus behaves like Ag, and very differently from 共2003兲 extended their calculations to N = 13, 38, and 55.
Au. Baletto, Ferrando, et al. 共2002兲 compared the ener- At N = 13, they found that the Ino decahedron is lower
gies of icosahedron, decahedron, and truncated octahe- in energy than the icosahedron and the cuboctahedron,
dron clusters at larger sizes, finding that the crossover
but a structure of D4h symmetry group originating from
from icosahedral to decahedral structures is around 1000
the symmetry breaking of the cuboctahedron is even
lower. At N = 55 the order of the structures is inverted,
10
See Bauschlicher 共1989兲, Bauschlicher et al. 共1988, 1989, and the icosahedron becomes lower than the Ino deca-
1990兲, Fujima and Yamaguchi 共1989兲, Massobrio et al. 共1995兲, hedron and the cuboctahedron. Fortunelli and Aprà
Jug, Zimmerman, Calaminici, et al. 共2002兲, Jug, Zimmerman, 共2003兲 attributed the results at N = 13 to the small size of
and Köster 共2002b兲. the molecule, while the behavior at N = 55 is intermedi-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 389

ate between finite molecules and fully metallic systems. dron at N = 309, but not at N = 561. However, it must be
Very recent density-functional calculations by Aprà et al. kept in mind that the cuboctahedron and the Ino deca-
共2004兲 show that rosette structures, originating from the hedron are not usually favorable fcc and decahedron
disordering of one or two vertices in icosahedra of 55 clusters. This observation is in agreement with previous
atoms, are considerably lower in energy than the icosa- density-functional calculations by Jennison et al. 共1997兲,
hedron itself. This result would support the preference who showed that for N = 140, the truncated octahedron
of Pt for low-symmetry structures at icosahedral magic is favored over the icosahedron structure obtained by
numbers. removing seven vertex atom from the Ih147, while in Ag
the opposite happens.
e. Palladium clusters Global optimization studies up to N ⯝ 100 by the Sut-
ton and Chen 共1990兲 potential give for Pd the same re-
The energetics of small Pd clusters have been investi-
sults as for Ag, with icosahedron structure at N = 55, as
gated by several groups employing different methods.11
confirmed also in the optimization of the Gupta 共1981兲
Besides their catalytic properties, these clusters are very
potential by Massen et al. 共2002兲. At larger sizes, a com-
interesting because their lowest-lying isomers could
parison of the different structural motifs shows that Pd
have nonzero spin 共Reddy et al., 1993; Watari and
behaves in a similar way to Pt 关with rather small cross-
Ohnishi, 1998兲, being thus magnetic. Zhang et al. 共2003兲
over sizes 共see Table I兲, and fcc clusters already in close
have recently performed an extensive study of the sta-
competition with the other motifs at N ⬃ 100兴 when
bility of Pd clusters within the density-functional ap-
modeled by the Rosato et al. 共1989兲 potential, while it
proach at N 艋 13. In this size range they compared sev-
behaves very similarly to Ag when treated by means of
eral 1D, 2D, and 3D isomers; moreover, they considered
embedded atom potentials 共Baletto, Ferrando, et al.,
a few selected structures at N = 19 and 55. Their results 2002兲. Calculations by Jennison et al. 共1997兲 better sup-
indicate that the lowest-lying isomers are 3D starting port the results from the Rosato et al. 共1989兲 potential,
with the tetramer 共Zacarias et al., 1999; Moseler et al., finding that fcc structures are more favored in Pd than in
2001兲, and that icosahedral structures are favored over Ag. However, this point would need further investiga-
both decahedral and cuboctahedral structures for N tion, and it is not resolved by the analysis of the avail-
= 13 and 55. These results are in good agreement with able experimental data. José-Yacamán, Marín-Almazo,
previous calculations by Kumar and Kawazoe 共2002兲. and Ascencio 共2001兲 observed by transmission electron
On the other hand, Watari and Ohnishi 共1998兲 found microscopy the thiol-passivated Pd nanoparticles in the
that the cuboctahedron is more stable than the icosahe- range of 1 – 5 nm diameter, seeing a variety of structures,
dron at N = 13 according to density-functional calcula- ranging from fcc 共simple and twinned兲, to icosahedral, to
tions. Moseler et al. 共2001兲 were able to compare their decahedral, to amorphous structures. They were able to
results on vertical electron detachment energies for an- observe rather large icosahedral clusters, explaining
ionic clusters with the experimental data 共Ervin et al., their presence as being probably due to kinetic trapping
1988; Ho et al., 1991; Ganteför and Eberhardt, 1996兲, effects.
obtaining a good agreement.
At larger sizes, ab initio results are few, being re-
stricted to the comparison of a small set of selected f. Nickel clusters
structures at some special sizes. Kumar and Kawazoe As in the case of Cu, small Ni clusters have been re-
共2002兲 have compared icosahedron with cuboctahedron cently reviewed by Alonso 共2000兲. Reuse and Khanna
and Ino decahedron clusters at N = 55 and 147, finding 共1995兲 and Reuse et al. 共1995兲 performed several density-
that slightly distorted icosahedral structures are the low- functional studies of small Ni clusters, concluding that
est in energy in both cases. Nava et al. 共2003兲 also found the lowest-lying isomers are nonplanar for N 艌 4. At N
that a Jahn-Teller distorted icosahedron is lower in en- = 7, two isomers, the pentagonal bipyramid and the
ergy than a perfect icosahedron at N = 55. Moreover, capped octahedron, are in close competition 共Nayak et
Nava et al. 共2003兲 found that icosahedra and cuboctahe- al., 1996兲. The former is favored by density-functional
dra are very close in energy at both N = 147 and N 共Nygren et al., 1992兲 and embedded-atom 共Boyakata et
= 309, with the icosahedra prevailing at 147 and cuboc- al., 2001兲 calculations, while the latter is most likely to
tahedra at 309. By tight-binding calculations, Barreteau be observed in the experiments 共Parks et al., 1994兲, al-
et al. 共Barreteau, Desjonquères, and Spanjaard, 2000; though the experimental data do not completely rule out
Barrateau, Guirado-Lopez, et al., 2000兲 found that the the pentagonal bipyramid 共Nayak et al., 1996兲. Nayak et
icosahedron is still lower in energy than the cuboctahe- al. 共1996兲 performed ab initio calculations, finding that
these structures are almost degenerate in energy; then
11 they performed molecular-dynamics simulations by the
See Valerio and Toulhoat 共1996, 1997兲, Jennison et al.
Finnis and Sinclair 共1984兲 semiempirical potential, find-
共1997兲, Zacarias et al. 共1999兲, Barreteau, Desjonquères, and
Spanjaard 共2000兲, Barreteau, Guirado-López, et al. 共2000兲, ing that the capped octahedron has a wider catchment
Efremenko and Sheintuck 共2000兲, Guirado-López et al. 共2000兲, basin, becoming thus much more favorable at high tem-
Efremenko 共2001兲, Krüger et al. 共2001兲, Moseler et al. 共2001兲, peratures. This is clear evidence of an entropic effect
Roques et al. 共2001兲, Kumar and Kawazoe 共2002兲, Nava et al. 共see Sec. IV.A兲. At N = 13, Parks et al. 共1994兲 experimen-
共2003兲, Zhang et al. 共2003兲. tally identified the icosahedron as the most stable struc-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


390 F. Baletto and R. Ferrando: Structural properties of nanoclusters

ture. Ab initio calculations at the same size 共Reuse and Soler et al. 共2000兲. What follows is a summary of these
Khanna, 1995; Reuse et al., 1995兲 have shown that the works. The energetic model is the Gupta 共1981兲 or
icosahedron is favored over the cuboctahedron, but the Rosato et al. 共1989兲 potential, which predicts the correct
most stable isomer is obtained by distorting the icosahe- surface reconstructions 共Guillopé and Legrand, 1989兲
dron slightly to obtain a cluster with D3d symmetry and reproduces quite accurately the diffusion barriers
group. The experimental observations 共Parks et al., 1995, 共Montalenti and Ferrando, 1999a, 1999b; Ala-Nissila et
1997, 1998兲 indicate a truncated octahedron structure at al., 2002兲 for these metals. In this framework, the poten-
N = 38, a capped decahedron structure at N = 39, and tial energy E of a cluster of N atoms is E = 兺i=1
N
Ei, where
共possibly defected兲 icosahedral structures at N = 55. The- the energy of atom i is given by

冑兺
oretical results are in reasonable agreement with these nv nv
experimental findings. Different global optimization re-
sults within semiempirical schemes 共Doye and Wales, Ei = A 兺
j=1
e−p共rij/r0−1兲 − ␰
j=1
e−2q共rij/r0−1兲 , 共11兲
1998b; Grygorian and Springborg, 2003兲 and tight-
binding calculations 共Lathiotakis et al., 1996兲 predict fcc where nv is the number of atoms within an appropriate
and icosahedral structures at N = 38 and 55, respectively. cutoff distance 共in the following we include only the first
Moreover, capped decahedral structures just above size neighbors兲, and A , ␰ , p , q are parameters fitted to the
38 are found 共Wetzel and DePristo, 1996; Andriotis and bulk properties of the element. The first term in Eq. 共11兲
Menon, 2004兲. At larger sizes, the comparison of the is a repulsive Born-Mayer energy, while the second is
different structural motifs by different semiempirical ap- the attractive band energy. Following Tománek et al.
proaches 共Cleveland and Landman, 1991; Mottet et al., 共1985兲, one can eliminate the parameters A and ␰, in
2004兲 shows that the crossovers icosahedral→decahedral order to have a function of 共p , q兲 and of the cohesive
and decahedral→ fcc take place at large sizes as in the energy per atom Ecoh. This is achieved by requiring that
case of Cu 共see Table I兲. in the bulk crystal at equilibrium Ei = Ecoh and ⳵Ei / ⳵r
= 0 for r = r0. The result is
4. General structural properties of noble-metal and
quasi-noble-metal clusters Ei =
兩Ecoh兩 v

冋兺 n

q e−p共rij/r0−1兲
12共p − q兲 j=1


Here we sketch general trends among the six metals
of the previous section with respect to three features:
formation of planar clusters, magnitude of crossover
sizes among the motifs, and preference for low-
− 冑12p 冑兺 nv

j=1
e−2q共rij/r0−1兲 . 共12兲

symmetry global minima at geometric magic sizes. This explicitly shows that the properties of the potential
Concerning the preference for formation of planar depend only on the parameters 共p , q兲, Ecoh and r0 play-
structures, there is a clear indication of an increasing ing the role of scale factors on energy and distance.
tendency going down from the 3d to the 5d series and Since atomic distances rij in noncrystalline structures are
from left to right in the Periodic Table. Thus Au has the not optimal, one can expect a metal that strongly in-
strongest tendency, followed by Ag, Pd, and Pt, all three
creases its energy for a change in rij 共i.e., that has a sticky
being at the same level 共however, there is still some de-
interatomic potential兲 to have small crossover sizes
bate about planar Pt clusters兲; Cu and Ni have no pref-
共Doye et al., 1995兲. In a solid at equilibrium, all 12 first
erence at all for planar structures. Several groups have
neighbors are at rij = r0. When one changes all rij by a
investigated the origin of this trend. Bravo-Pérez et al.
common factor rij → 共1 + ␧兲rij, developing the crystal en-
共1999b兲 correlated the preference for planar structures
ergy per atom Ei共␧兲 in Eq. 共12兲 共with nv = 12兲 to the sec-
in Au to the fact that nonadditive many-body interac-
tions are stronger than additive two-body ones in this ond order and dividing it by the equilibrium value 兩Ecoh兩,
element. Bonacić-Koutecký et al. 共2002兲 noticed that in one obtains 共Baletto, Ferrando, et al., 2002兲
planar Au clusters, d electrons contribute more to the Ei共␧兲 − Ei共0兲 1 2
bonding than in 3D structures; this would also qualita- ␳共␧兲 = = ␧ pq. 共13兲
兩Ei共0兲兩 2
tively explain the weaker tendency to planar clusters for
Ag, because in Ag the bonding is more of s type. Finally, Here ␴1 = ␳共␧兲 / ␧2 = pq / 2 is essentially the product of the
Häkkinen et al. 共2002兲 investigated Au7−, Ag7−, and Cu7− bulk modulus and the atomic volume divided by the co-
by density-functional calculations and demonstrated that hesive energy per atom of the bulk crystal. Thus the
the propensity of Au clusters to favor planar structures larger the pq, the smaller the crossover sizes from icosa-
is correlated with the strong hybridization of the atomic hedra to decahedra NIh→Dh and from decahedra to fcc
5d and 6s orbitals due to relativistic effects. crystallites NDh→fcc, as can be seen in Table I.
The tendency to form strained structures like icosahe- Let us now discuss the tendency to form low-
dra and decahedra, leading to large crossover sizes, and symmetry structures at geometric magic sizes 共referred
the tendency to present low-symmetry structures at geo- to as the tendency to amorphization in the following兲.
metric magic numbers have been discussed, in the This tendency is due to a specific feature of the metallic
framework of the same well-defined and quite reliable bonding, which derives from its many-body character
energetic model, by Baletto, Ferrando, et al. 共2002兲 and and relates the optimal distance of the bonding to the

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 391

coordination of the atom 共bond-order/bond-length cor- difference being the indication that Pt should have a
relation兲. This can be easily seen by taking a number of stronger amorphization tendency than Pd, due essen-
neighbors 1 艋 nv 艋 12 in Eq. 共12兲, all at the same distance tially to the comparatively weaker enthalpy of melting
r. Minimization with respect to r leads to the optimum of Pt 共the ratio ⌬Hmelt / 兩Ecoh兩 is much smaller in Pt than
distance r*共nv兲, in Pd兲.

冉 冊
Finally, let us compare metals with Lennard-Jones
r*共nv兲 1 nv clusters, in order to sketch a qualitative picture of the
=1+ ln , 共14兲
r0 2共p − q兲 12 effects of many-body forces. If one applies the criterion
of Eq. 共13兲 to Lennard-Jones clusters, one finds a value
which is more and more contracted for decreasing nv. It
also leads to the energy of ␴1, which is even larger than those for Au, Pt, and Pd.

冉 冊
This would suggest that Lennard-Jones systems have a
共p−2q兲/2共p−q兲
nv weaker tendency to form icosahedral clusters than these
E*共nv兲 = − 兩Ecoh兩 . 共15兲 metals, but this is not the case, since crossover sizes
12
icosahedron→ decahedron and icosahedron→ fcc are
The atoms on the surface of the cluster try to contract in much larger for Lennard-Jones clusters than for Au, Pt,
order to minimize their elastic energy 共Soler et al., 2001兲. or Pd. This happens because the bond-order/bond-
In the case of highly coordinated structures like the length correlation of metallic elements tends to destabi-
icosahedra 共which have intrashell expanded distances兲, lize icosahedral structures. In fact, metals would prefer
the optimization of the bond lengths is made at the ex- contracted interatomic distances for the low-
pense of the number of bonds. Therefore one expects coordination surface atoms, and bulk distances for the
amorphization to be easy in the elements with high elas- highly coordinated inner atoms. This is exactly the op-
tic energy 共i.e., large pq兲 and strong contraction of the posite of what happens in perfect icosahedral structures,
bonds with decreasing coordination, the latter leading to which have contracted internal distances and expanded
a weak dependence of E* on nv. The contraction of the surface distances. Moreover, in some metals such as Pt,
bonds and the dependence of E* on the coordination bond directionality effects 共which are not included in the
can be quantified by the following dimensionless param- potentials considered in Table I兲 can be important 共For-
eters: tunelli and Velasco, 2004兲 and favor the appearance of
nv dr* 1 共111兲-like hexagonal facets on the cluster surface 共Aprà
␴2 = = , et al., 2004兲, with a further destabilizing effect on icosa-
r0 dnv 2共p − q兲 hedral clusters.

nv dE* p − 2q
␴3 = * = . 共16兲 5. Silicon clusters
E dnv 2共p − q兲
Silicon nanoclusters are of great practical interest be-
As is evident from Table I, all parameters agree in indi- cause of their intense photoluminescence at room tem-
cating that Au has by far the strongest tendency to perature and the presence of quantum size effects.
amorphization, having the largest ␴1 and ␴2 and the There is an ongoing debate about the size at which the
smallest ␴3, followed by Pt and Pd; Ag and especially Cu most favorable clusters adopt the bulklike diamond
and Ni have the weakest tendency. structure. Bachels and Schäfer 共2000兲 produced neutral
On the same line of reasoning, Soler et al. 共2000兲 de- Si clusters in a laser vaporization source and measured
fined a parameter based solely on experimental quanti- their binding energy in samples having average sizes N̄
ties to extend the estimate of the tendency to amor- from 65 up to 890 atoms. They showed that the binding
phization to other systems. This parameter is ␴A and is energy per atom Eb in this size range scales as N1/3, this
defined as the elastic energy ␦Eel that is needed to form being characteristic of approximately spherical clusters
an ordered surface with contracted atomic distances, di- 共Kaxiras and Jackson, 1993a兲. At smaller sizes, different
vided by the amorphization energy ␦Eam which is re- regimes in the behavior of Eb with N were found, as can
quired to form a scattered distribution of bond lengths. be seen from Fig. 15. There, Bachels and Schäfer 共2000兲
␦Eel is expressed by means of the 共Voigt-averaged兲 bulk plotted, besides their results at large sizes, Eb for smaller
and shear moduli B and G. ⌬Eam is approximated by the clusters 共N 艋 7兲 obtained from the results of other
enthalpy of melting ⌬Hmelt, which is an estimate of the groups 共Jarrold and Honea, 1991; Schmude et al., 1993,
energetic cost to form amorphous structures resembling 1995; Fuke et al., 1994兲. Three regimes are clearly
those found in liquids. This leads to shown. For N ⬍ 10 the binding energy increases rapidly
␦Eel v共3B − 5G兲2 with the cluster sizes, and compact elementary units are
␴A = ⯝ , 共17兲 built up. For 10⬍ N ⬍ 25, Eb is practically independent
⌬Eam 320B⌬Hmelt
of N, as would happen for prolate structures 共Kaxiras
where v is the atomic volume. Equation 共17兲 is valid for and Jackson, 1993a兲. This agrees with the previous ob-
small clusters having almost all atoms on their surface. servation of small prolate Si clusters by Jarrold and Con-
Elements with large ␴A are expected to have a stronger stant 共1991兲. Finally, for larger clusters the N1/3 behavior
tendency to amorphization. As can be seen in Table I, is recovered. Bachels and Schäfer 共2000兲 were also able
the inspection of ␴A confirms the above trends, the only to produce metastable prolate structures at sizes up to

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


392 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 15. Size dependence of the binding energies of neutral


silicon clusters: 䊐, binding energies of the two groups of clus-
FIG. 16. The best isomers of Si20 according to Mitás et al.
ter isomers found by calorimetric measurements; 쎲, data for
共2000兲. Structure E is the lowest in energy according to quan-
neutral silicon clusters obtained from the collision-induced dis-
tum Monte Carlo calculations. From Mitás et al., 2000.
sociation experiments of Jarrold and Honea 共1991兲 on the cor-
responding silicon cluster cations, taking the experimentally
determined photoionization potentials into account as from
Fuke et al. 共1994兲; 䉱, the Knudsen mass spectrometric mea- that Si20 is made of six- and eight-atom subunits. They
surements of Schmude et al. 共1993, 1995兲. The crossover from also found that the dissociation energies and inverse mo-
elongated to the spherical neutral silicon structures can be es- bilities calculated from their global minima were in ex-
timated from the binding energies to be located around N cellent agreement with the experiments 共Jarrold and
= 25. The structure for an elongated Si26 cluster is taken from Honea, 1991; Hudgins et al., 1999兲 up to N = 18, conclud-
Grossman and Mitás 共1995b兲. Adapted from Bachels and ing that cluster formation is dominated by the energetics
Schäfer, 2000. up to these sizes.
In summary, experiments and calculations agree in
finding prolate structures of a few tens of atoms which
about 170 atoms. These structures were mostly present
are built up by small subunits in the range 6–10 atoms.
in the part of the molecular beam with short dwell times.
How do these results compare to calculations? And is This is also consistent with fragmentation experiments
the transition size to quasispherical structures the same of clusters of about 150 atoms by Ehbrecht and Huisken
as the transition size to bulklike diamond structures? In 共1999兲, yielding Si6+ – Si11+ products.
the following we address these questions. Clusters are quasispherical for about N ⬎ 25 共Hudgins
Theoretical results on the energetics of small silicon et al., 1999兲, even if they do not form crystalline struc-
clusters have been obtained by quantum Monte Carlo, tures at these sizes. Indeed, the experiments of Ehbrecht
density-functional, and tight-binding calculations. The and Huisken 共1999兲 could be interpreted as assuming
latter also permitted the application of global optimiza- that compact shapes at N ⬃ 150 are built up by small
tion methods such as simulated annealing 共Yu et al., subunits 关the prolate metastable structures observed by
2002兲 and genetic algorithms 共Ho et al., 1998; Rata et al., Bachels and Schäfer 共2000兲 being possibly unfolded ver-
2000; Wang, Wang, et al., 2001兲. sions of these clusters兴, and this agrees with the simu-
For very small clusters, Raghavachari and Logovinsky lated annealing calculations by Yu et al. 共2002兲. But this
共1985兲 found that the best clusters are planar only up to is not the only possibility, since Kaxiras and Jackson
N = 4; several calculations 共Fournier et al., 1992; Li et al., 共1993a兲, Ho et al. 共1998兲, and Mitás et al. 共2000兲 have
1999; Zickfeld, et al., 1999; Yu et al., 2002兲 found a gen- shown that quasispherical noncrystalline structures,
erally good agreement with the above results. An impor- which are not built up by stacking smaller subunits, be-
tant result at small sizes is that Si10 is a very stable clus- come favorable at N 艌 20. These sizes are large enough
ter 共Raghavachari and McMichael-Rohlfing, 1988; to allow the formation of cages containing at least one Si
Grossman and Mitás, 1995a; Ramakrishna and Bahel, atom inside 共Mitás et al., 2000兲. These results compare
1996; Li, Yin, et al., 2000兲 so that it can act as a subunit rather well with the experiments of Bachels and Schäfer
of larger ones 共see Fig. 16兲. For example, quantum 共2000兲. However, the theoretical determination of the
Monte Carlo, density-functional 共Mitás et al., 2000兲, and most stable structures in medium-size Si clusters is a
tight-binding calculations 共Yu et al., 2002兲 find that the very complex task: see, for example, the debate about
best structures at N = 20 are simply formed by linking the structure of Si36 in Sun et al. 共2003兲 and Bazterra et
two Si10 subunits. Indeed, the most reliable explanation al. 共2004兲.
for the observation of prolate clusters is that they are Mélinon et al. 共1997, 1998兲 demonstrated that the
built up by stacked subunits, whose structure is still un- models based on quantum confinement in crystalline sili-
der debate. Kaxiras and Jackson 共1993a, 1993b兲 pro- con clusters fail for films containing grains of less than
posed that the subunits are sixfold rings; Raghavachari 2 nm diameter 共say N ⯝ 200兲. They were not able to ob-
and McMichael-Rohlfing 共1988兲, Jarrold and Bower serve any crystallization by transmission electron mi-
共1992兲, and Ho et al. 共1998兲 were in favor of tricapped croscopy. On the other hand, if the grains are of 3 nm
trigonal prisms of nine atoms; Rata et al. 共2000兲 found diameter, there is evidence of crystalline ordering 共Eh-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 393

brecht et al., 1997; Ehbrecht and Huisken, 1999; Ledoux isomers are found at N = 38 共truncated octahedron兲, N
et al., 2000兲. This would indicate that the transition to = 75 关共2,2,2兲 Marks decahedron兴 and 101 关共2,3,2兲 Marks
crystalline structures is around N = 400 atoms, as Yu et al. decahedron兴, and that N = 55 is not a magic number. At
共2002兲 found by simulated annealing tight-binding calcu- that size, the icosahedron is higher in energy than the
lations. best decahedron structure by 0.3 and 2 eV according to
the Pacheco and Prates-Ramalho 共1997兲 and Girifalco
6. Clusters of fullerene molecules 共1992兲 potentials, respectively.
These results completely disagree with the experimen-
The condensed-phase properties of C60 molecules tally observed structures. In fact, Martin et al. 共1993兲 and
have been the subject of considerable interest because Branz et al. 共2000, 2002兲 found that clusters grown at low
of the unusual character of molecular interactions. In T and annealed at ⬃500 K 共to allow the less bound sur-
fact, above room temperature, C60 molecules can be face molecules to evaporate兲 present a mass spectrum
considered as large spherical pseudoatoms which are with a clear sequence of icosahedral magic numbers up
free to rotate; the effective interaction between these to very large sizes, well above N = 100 共see Fig. 17兲. The
pseudoatoms is extremely short-ranged relative to the peak at N = 55 is very evident, and no peaks are found at
large equilibrium pair separation. This is at the origin of N = 38 and 75. The mass spectra are qualitatively similar
peculiar properties. In fact, such sticky interactions 共see for neutral, positively and negatively charged clusters
Fig. 7兲 disfavor all strained configurations, like those oc- 共Branz et al., 2002兲. Recently, Rey et al. 共2004兲 compared
curring in icosahedra or in the liquid phase. Indeed, the the neutral and singly-ionized cluster structures in a
existence of a liquid phase for bulk C60 共Hagen et al., model including the Girifalco 共1992兲 potential plus a
1993; Caccamo et al., 1997; Ferreira et al., 2000兲 is still point polarizable dipole electrostatic model. They found
debated. Therefore clusters of C60 molecules are on the that the structures of ionized clusters are very similar to
opposite side with respect to clusters of alkali metals, those of neutral clusters, thus confirming the observa-
and we expect noncrystalline structures to be the global tions. Branz et al. 共2002兲 found decahedra and close-
minima only for aggregates of a small number N of mol- packed structures only after further annealing at higher
ecules. T. This result indicates the possible existence of kinetic
This qualitative prediction has been confirmed by all trapping effects in the formation of the clusters of C60
existing calculations. We can divide these studies into molecules 共Baletto, Doye, and Ferrando, 2002兲, as we
two classes. The first uses an all-atom potential, with discuss in Sec. V.B.5. Trapping effects are overcome only
each carbon atom interacting with atoms of other mol- after a strong annealing.
ecules by a Lennard-Jones potential. This approach in- The resulting high-T experimental structures, how-
cludes deviations from spherical symmetry, but at ever, still do not coincide with those predicted by the
present allows global optimization only for small N. The Girifalco 共1992兲 and Pacheco and Prates-Ramalho
second class uses spherically averaged potentials, such as 共1997兲 potentials. In fact, there is a sequence of experi-
the Girifalco 共1992兲 potential, in which Lennard-Jones mental magic numbers that can be attributed to the
centers are continuously and uniformly distributed on Leary tetrahedron at N = 98 and to its fragments down to
the surface of a sphere, or the potential developed by N = 48, while the calculations do not attribute special sta-
Pacheco and Prates-Ramalho 共1997兲, which is derived by bility to these structures. This discrepancy is probably
fitting density-functional calculations. The Girifalco due to the fact that even the Pacheco and Prates-
共1992兲 potential is purely two-body, whereas the Pacheco Ramalho 共1997兲 potential is too sticky; a slightly less
and Prates-Ramalho 共1997兲 potential includes two-body sticky Morse potential gives the Leary tetrahedron as a
and three-body terms 共the latter, however, being of magic structure 共Doye et al., 2001兲.
rather small importance兲.
All these approaches agree in predicting that icosahe-
dral structures are favored only at very small N. All-
atom potentials favor icosahedral structures up to N IV. THERMODYNAMICS OF FREE NANOCLUSTERS
= 16 共Doye et al., 1997; Garcia-Rodeja et al., 1997; Rey et
al., 1997兲, while calculations with the Girifalco 共1992兲 In this section, we analyze the effects of raising the
potential, which is the stickiest, indicate that icosahedra temperature on cluster structures. These effects may in-
are the lowest in energy only up to N = 13 共Rey et al., clude solid-solid structural transitions and the melting of
1994; Wales, 1994a; Doye and Wales, 1996b兲. Finally, the the cluster if the temperature is raised enough. Both
Pacheco and Prates-Ramalho 共1997兲 potential favors solid-solid transitions and melting are cases that show
icosahedral structures up to N = 15. Larger clusters are the peculiar thermodynamic behavior of small finite sys-
either based on decahedra or on close-packed struc- tems such as clusters. Indeed, the thermodynamics of
tures, as shown by Doye et al. 共2001兲, who performed an finite systems is a fascinating and complex field, involv-
exhaustive basin-hopping global optimization study up ing many subtleties. Here we have no intention of being
to N = 105 on the basis of both Pacheco and Prates- exhaustive and refer readers who need more complete
Ramalho 共1997兲 and Girifalco 共1992兲 potentials. Close- treatments to existing monographs, from the classical
packed structures become more frequent as the size in- book of Hill 共1964兲 to the excellent book by Wales
creases. Both potentials indicate that the most stable 共2003兲. In the following we treat only those issues that

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


394 F. Baletto and R. Ferrando: Structural properties of nanoclusters

structures may coexist. This coexistence is dynamic, like


that of coexisting chemical isomers.12 Second, thermody-
namic properties 共like the melting point兲 can be strongly
size dependent, in analogy to what happens to the
global-minimum structures. General trends for thermo-
dynamic properties may be deduced on the basis of the
following expression for the Gibbs free energy G 共Hill,
1964兲. In a bulk system, G = Ng共p , T兲 where g共p , T兲 is the
specific Gibbs potential. For a small system, we have
also to consider contributions coming from surfaces,
edges, etc. Thus we write G as
G = Ng共p,T兲 + b共p,T兲N2/3 + c共p,T兲N1/3 + d共p,T兲,
共18兲
2/3
where the term in N is a surface free energy, the term
in N1/3 is due to edge contributions, and the last term
might be due, among other things, to the rotation of the
cluster. In the limit N → ⬁, one has G → Ng, the macro-
scopic relationship. But when the system is small all
terms are important. The finite size of the clusters also
implies that treatments in different thermodynamic en-
sembles 共microcanonical and canonical兲 may give differ-
ent behaviors for thermodynamic quantities such as the
heat capacity 共Bixon and Jortner, 1989; Jortner, 1992兲.
The thermodynamics of clusters have been studied by
a variety of theoretical and simulation tools. These in-
clude Monte Carlo and molecular-dynamics simulations
and analytical methods.13
In Sec. IV.A we focus on the role of entropic contri-
butions to the free energy at increasing temperature,
which may cause solid-solid structural transitions when
the structure corresponding to the global energy mini-
mum ceases to be the most likely at high T, so that other
structures prevail. In Sec. IV.B we deal with the melting
transition. There, we first discuss what is meant by melt-
ing 共and premelting兲 in clusters, briefly reviewing experi-
mental and simulation methods. Then we focus on the
size dependence of the melting point, which may be very
complex, as in the case shown in Fig. 18. We treat phe-
nomenological theories for the 共average兲 monotonic de-
pendence of the melting point on size, and discuss the
origin of its nonmonotonic variations. Finally, in Sec.
IV.C we focus on some systems of particular interest.

A. Entropic effects and solid-solid transitions


FIG. 17. Mass spectra of C60 clusters: panel 共a兲, clusters pro-
duced at low T do not reveal any magic number; panel 共b兲, Structural transitions can take place upon increasing
after mild annealing to evaporate the less bound molecules, a the temperature for a given size because of entropic ef-
clear series of icosahedral magic numbers emerges; panels 共c兲
and 共d兲, decahedra and close-packed magic numbers appear
only after a high-T annealing. Adapted from Branz et al., 2002. 12
See, for example, Honeycutt and Andersen 共1987兲, Berry et
al. 共1988兲, Labastie and Whetten 共1990兲, Bartell 共1992兲, Mat-
suoka et al. 共1992兲, Cleveland and Landman 共1994兲, Schmidt et
have a direct bearing on temperature-dependent
al. 共1998兲, Jellinek 共1999兲.
changes in the cluster structures. 13
Early studies are those of Lee et al. 共1973兲, Briant and Bur-
Generally speaking, two points must be kept in mind ton 共1975兲, Imry 共1980兲, Nautchil and Petsin 共1980兲, Natanson
about cluster thermodynamics. First, phase transitions in et al. 共1983兲, Berry et al. 共1984兲, Quirke and Sheng 共1984兲,
small systems are gradual, not sharp 共Hill, 1964兲. A con- Amar and Berry 共1986兲, Davis et al. 共1987兲, Luo et al. 共1987兲,
sequence of this fact is that there are bands of tempera- Beck and Berry 共1988兲, Reiss et al. 共1988兲, Bixon and Jortner
ture and pressure within which two or more cluster 共1989兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 395

fects. When the minimum-free-energy structure is differ-


ent from the minimum-energy structure, we can define a
temperature Tss at which a solid-solid transformation oc-
curs. In other words, for T ⬎ Tss the minimum-energy
structure 共which is always the most likely for T → 0兲
ceases to be the most probable to the advantage of some
other structure, which prevails because of entropic ef-
fects. Structural changes from fcc to decahedral and
icosahedral structures as T increases have been pre-
dicted theoretically and seen in simulations of several
systems 关small Lennard-Jones 共Doye and Calvo, 2001兲,
Au 共Cleveland et al., 1998, 1999兲 and Cu clusters 共Bal-
etto et al., 2004兲兴. Moreover, solid-solid entropy-driven
structural transformations in Lennard-Jones clusters of
about 200 atoms and with different structures have been
observed in Monte Carlo simulations by Polak and Pa-
trykiejew 共2003兲. Recent experiments by Koga et al.
共2004兲 support the existence of entropy-driven solid-
solid transitions. Here we follow the theoretical treat-
ment of Doye and Calvo 共2001, 2002兲, dealing first with
the classical harmonic approximation, and then intro-
ducing anharmonic and quantum corrections.

1. Structural transitions in the harmonic approximation


Let H be the Hamiltonian of a cluster of size N: FIG. 18. Size-dependence of the measured melting tempera-
tures and latent heat of melting for sodium clusters. These
H = T共兵p其兲 + U共兵r其兲, 共19兲 quantities are compared to the mass spectrum of the top panel,
which shows electronic magic numbers. The correlations be-
where T共兵p其兲 = 兺i=1 N 2
pi / 2m is the kinetic energy and tween the electronic magic sizes and the peaks in the melting
U共兵r其兲 = U共r1 , . . . , rN兲 is the potential energy. Assume temperature are rather weak, indicating that the melting be-
now that the cluster has a 共nonlinear兲 structure, corre- havior of small sodium clusters cannot be explained by simple
sponding to a given minimum s of energy Es0. If the cou- models. Adapted from Schmidt et al., 1998. Reprinted with
pling of rotational and vibrational motions can be ne- permission from Nature http://www.nature.com/
glected, Zs can be factored as

where
Zs = ZstrZsrotZsvib ,
Zstr and Zsrot
are related to the center-of-mass
translation and to rigid rotational motions, respectively,
共20兲
Zstr = 冉
V
MkBT
2␲ប2
冊 3/2
, Zsrot = 冉 2␲kBTĪs
ប2
冊 3/2

, 共23兲

while Zsvib is related to the vibrational motion around s.


Zsvib depends on the internal coordinates ␰s,i where i where V is the volume of the box in which the cluster is
ranges from 1 to ␬ = 3N − 6. For small oscillations, the contained, M is the cluster mass, and Īs is the average
vibrational motion can be treated within the harmonic inertial moment in s 关Īs = 共IsxxIsyyIszz兲1/3, with Isxx, Isyy, and
approximation. In this case, it is convenient to choose
Isxx the principal moments of inertia兴.
the ␰s,i as the normal-mode coordinates, in order to write
To estimate the temperature dependence of the prob-
U as
ability of finding the cluster in s, we have to compute the
␬ total partition function Z. Before doing so, we note that
1
U = Es0 + 兺 ␻2 ␰2 ,
2 i=1 s,i s,i
共21兲 each minimum s has a number ns of equivalent permu-
tational isomers, of equivalent minima which are ob-
where the ␻s,i are the normal-mode frequencies. The tained by exchanging the coordinates of atoms of the
transformation to normal modes allows an easy evalua- same species. For homogeneous clusters, ns = 2N! / hs,
tion of Zsvib, which, in the classical case, is given by where hs is the order of the symmetry group of mini-
mum s 共Wales, 2003兲. If, at a given T, there are Mmin

0 ␬ ⬁
e −␤Es minima with a non-negligible probability of being occu-

2 2 ˙2
Zsvib ⯝ d␰s,id␰˙ s,ie−共␤/2兲共␻s,i␰s,i+␰s,i兲
共2␲ប兲␬ i=i −⬁
pied, the probability ps that the cluster is in the en-
semble of the permutational isomers of minimum s can

冉 冊

k BT be evaluated by the superposition approximation 共Doye

0
= e −␤Es . 共22兲 and Wales, 1995兲. Within this approximation, Z is ob-
i=i ប␻s,i
tained by summing up the contributions from all signifi-
The classical expressions for Zstr and Zsrot are cant minima. This gives

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


396 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 19. 共Color in online edition兲 Schematic representation of FIG. 20. Structural phase diagram in the N,T plane for
the probabilities pS and pS⬘ of the local energy minima S and Lennard-Jones clusters in the harmonic superposition approxi-
S⬘, with S⬘ having higher energy than S vs temperature. In the mation. From Doye and Calvo, 2002.
temperature range close to the transition temperature TSS⬘ the
shift upwards through icosahedral→ decahedral and
two structures coexist, being almost equally likely.
decahedral→ fcc crossover sizes with increasing tem-
perature, as can be seen in the 共coarse-grained兲 struc-
0 tural phase diagram of Fig. 20. This happens because
n sZ s n sZ s nsĪs3/2e−␤Es ⍀s−␬ icosahedral structures have on average smaller vibra-
ps = ⯝ Mmin ⯝ Mmin , 共24兲
Z tional frequencies than decahedral ones, the latter hav-
兺 兺
0
n ␴Z ␴ 3/2 −␤E␴
n␴Ī␴ e ⍀␴−␬ ing softer vibrations than fcc clusters 共Doye and Calvo,
␴=1 ␴=1
2002兲. Metals modeled by Sutton and Chen 共1990兲 po-
where ⍀s = 共⌸␬i ␻i,s兲1/␬ is the geometric average of the vi- tentials behave in a qualitatively similar way 共Doye and
brational frequencies of s. Calvo, 2001兲. A discussion of the observation of solid-
Let us now consider the simplest case of a potential- solid transitions in experiments and simulations on spe-
energy surface having two minima, s and s⬘ 共with Es0 cific systems is given in Sec. IV.C. Here we only remark
that solid-solid transitions may be extremely slow in
⬍ Es0 兲, and calculate the ratio ps / ps⬘ between their occu-
⬘ typical situations, so that observing them may require a
pation probabilities as a function of T. Assuming that
substantial overheating of the initial structure 共Koga et
Īs ⯝ Īs⬘, one obtains al., 2004兲, to surmount the energy barriers leading to
ps
=
Z sn s
p s⬘ Z s⬘n s⬘
= e−␤⌬E
0
冋 冉 冊册
n s ⍀ s⬘
n s⬘ ⍀ s

, 共25兲
escape from the initial basin.
2. Anharmonic and quantum corrections
The harmonic superposition approximation is often
where ⌬E0 = Es0 − Es0 . Since ⌬E0 ⬎ 0 we have two cases:
⬘ accurate for cluster thermodynamics up to temperatures
共i兲 if ns / ⍀s␬ ⬎ ns⬘ / ⍀s␬⬘, s is favored over s⬘ for all tem- close to the melting range, as shown for Ag38 and Cu38
peratures; by Baletto et al. 共2004兲. However, there are cases in
which this approximation is not sufficient, and anhar-
共ii兲 if ns / ⍀s␬ ⬍ ns⬘ / ⍀s␬⬘, for T ⬎ Tss, s⬘ becomes more monic effects must be included 共Doye and Wales, 1995兲.
favored than s and thus a solid-solid transition is Solid-solid transitions can be close to the melting tem-
possible; see Fig. 19. perature, in which case anharmonic effects are not neg-
From the above equations, we estimate Tss as ligible. Moreover, the harmonic approximation may
strongly overestimate the probability of minima that
⌬E0 have low-barrier escape paths. An example of failure of
Tss = . 共26兲 the harmonic approximation is found by Wang, Blastein-
kB关ln共ns⬘/ns兲 + ␬ ln共⍀s/⍀s⬘兲兴
Barojas, et al. 共2001兲 and Doye and Calvo 共2003兲.
To find a solid-solid transition, not only must case 共ii兲 One can introduce anharmonic effects via
hold, but Tss must be below the melting range of the T-dependent frequencies, i.e., ⍀̃s共T兲 = ⍀s共1 − ␤s0 / ␤兲,
cluster. where ␤0 is a measure of the anharmonicity. This then
How do entropic contributions behave in typical gives for the transition temperature, neglecting the per-
cases? Doye and Calvo 共2002兲 throughly analyzed mutational contribution,

冋 冉 冊册
Lennard-Jones clusters within the harmonic superposi-
tion approximation. They grouped together the minima ␬ ␤ − ␤s0
␤ss
anHA
= ln共⍀ /⍀ s⬘ 兲 + ln , 共27兲
pertaining to each motif 共icosahedron, decahedron, and ⌬E0
s
␤ − ␤s0⬘
fcc兲, thus being able to estimate transition temperatures
between structural motifs and not simply between pairs where the second term represents the first order ap-
of minima. They found that the entropic contributions proximation to the anharmonic correction to ␤ss anHA

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 397

anHA
= 1 / kBTss . To apply the previous equation it is neces- 共1998兲. However, there are several differences between
sary to estimate ␤0 but the available methods do not the solid-liquid transition of clusters and that of its bulk
allow one to treat large sizes. counterpart:
Quantum corrections can be inserted both in Zsrot and 共i兲 The melting point is reduced 共with a few known
Zsvib. Since quantum corrections to Zsrot should become exceptions兲 with a complex dependence on size.
important only at very low temperatures, we now con-
共ii兲 The latent heat is smaller; this can be understood,
sider only the corrections to Zsvib. These corrections are
for example, by noting that disordering the sur-
very easily introduced within the harmonic approxima-
face costs less than disordering inner atoms.
tion. The quantum partition function of a set of oscilla-
tors is given by 共iii兲 The transition does not take place sharply at one
definite temperature, but smoothly over a finite

temperature range. There, solid and liquid phases

0
Zsvib = e−␤Es 关2 sinh共␤ប␻i,s/2兲兴−1 . 共28兲 may coexist dynamically in time 共Berry et al.,
i=1 1988; Lynden-Bell and Wales, 1994兲. When one
observes the cluster over a long time interval,
The classical limit is valid when it is possible to approxi-
there will be subintervals in which the cluster ap-
mate the sinh共␤ប␻i,s / 2兲 with its argument; thus it is valid
pears to be solid and others in which the cluster is
for high temperatures. The introduction of quantum os-
fully liquid.
cillators modifies the expression of the ratio between the
occupation probabilities of two minima s and s⬘ in the 共iv兲 The heat capacity can become negative in micro-
following way: canonical environments 共Schmidt et al., 2001兲.
This means that the microcanonical average ki-

冤 冥
␬ netic energy may be a nonmonotonically increas-

ps ns
兿i sinh共␤ប␻i,s⬘/2兲 ing function of the total energy in the range of the
= e−␤⌬E
0
. 共29兲 transition 共Bixon and Jortner, 1989兲.

p s⬘ n s⬘
共v兲
兿i sinh共␤ប␻i,s/2兲 The melting transition depends on cluster struc-
ture and chemical ordering 共this being, indeed, a
nonequilibrium effect兲.
In particular, we note that here the temperature depen- Often the melting transition is preceded by premelting
dence is also in the vibrational contribution and not only phenomena. To quote Calvo and Spiegelman 共2000兲,
in the Boltzmann factor. To understand when quantum “premelting phenomena are characteristic of isomeriza-
corrections are needed we may refer to the Debye tem- tions taking place in a limited part of the configuration
perature ⌰D. If the temperature of the system is above space,” for example, isomerizations involving surface at-
0.70⌰D the system can be treated as classical to a good oms only 共in this case the term surface melting is used兲.
approximation, while at lower temperature we also need Premelting phenomena are often singled out by addi-
to take quantum effects into account 共Ashcroft and Mer- tional peaks in the specific heat vs temperature.
min, 1976兲. Calvo et al. 共2003兲 satisfactorily explained the
finite-temperature spectroscopic properties of CaArN
clusters including quantum and anharmonic corrections 1. Experimental methods
in their calculations.
Following Haberland 共2002兲, experiments studying
cluster melting can be divided into two classes:
B. Melting of nanoclusters 共i兲 study of the change of some physical property
across the melting point 共for example, changes in
The concepts of solid and liquid states, which are photon or x-ray diffraction patterns兲
commonly employed when discussing extended systems,
共ii兲 measurement of the caloric curve E = E共T兲, that is,
can be applied to clusters. In fact, at low temperature,
the particles of a cluster spend most of the time making the cluster’s internal energy E as a function of T.
small-amplitude vibrations around the global minimum, Takagi 共1954兲 made the first observation of the melt-
in analogy to what happens in bulk solids. If the tem- ing point depression by transmission electron micros-
perature is increased, other minima begin to be popu- copy. This is a standard technique for studying the size-
lated, and this is associated with the onset of some dif- dependent melting point of small particles by
fusive motion. Finally, if the temperature becomes high monitoring the changes in the diffraction pattern associ-
enough, the cluster explores the basins of a huge num- ated with the disordering of the structure. Electron and
ber of minima, with fast rearrangements, thus behaving x-ray diffraction and nanocalorimetry techniques have
like a liquid droplet. Within this description, cluster been used to study melting in deposited clusters 共Lai et
melting is seen as an isomerization transition, with the al., 1996; Peters et al., 1998; Efremov et al., 2000兲. At
number of probable isomers increasing drastically after present these methods are, however, not applicable to
some threshold temperature. A nice example of this be- free, mass-selected clusters in vacuum for two reasons:
havior for Ag6 clusters is given by Garzón, Kaplan, et al. first, no method of temperature measurement is known

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


398 F. Baletto and R. Ferrando: Structural properties of nanoclusters

in this case, and the density of mass-selected clusters is


so small that it is extremely difficult to collect a diffrac-
tion signal. In any case, as the size decreases, the diffrac-
tion techniques become increasingly inaccurate due to
line broadening. Several experiments have tried to mea-
sure the melting behavior of free cluster, mainly by
methods belonging to class 共i兲. Even et al. 共1989兲 and
Buck and Ettischer 共1994兲 looked for some spectro-
scopic evidence. Electron diffraction from a 共not mass-
selected兲 supersonic expansion gives Debye-Scherrer
like diffraction rings, whose intensity is a measure of
FIG. 21. Schematic representation of the experimental method
cluster temperature. This method was pioneered by
for measuring E共T兲 for free clusters. Adapted from Schmidt et
Farges et al. 共1981兲 and later intensively studied by Hov- al., 1998. Reprinted with permission from Nature http://
ick and Bartell 共1997兲. Martin et al. 共1994兲 were the first www.nature.com/
to publish a size dependence of the melting temperature
of free clusters. They showed that the structure on mass
spectra of large sodium clusters can depend sensitively havior as the laser-heated ones. A second mass spec-
on the temperature. The disappearance of structure was trometer measures the distribution of the fragment ions
interpreted as being due to melting. Another method produced. Different numbers of absorbed photons lead
has been proposed by Shvartsburg and Jarrold 共2000兲 to to clearly separated groups of fragments in the mass
measure the melting temperature Tm for small tin clus- spectrum, with the distance between two groups corre-
ters, with the surprising result that they melt higher than sponding to exactly the energy of one photon. This al-
bulk tin. Here, cluster ions are injected into a helium-gas lows one to calibrate the mass scale in terms of energy. If
atmosphere, and are pulled by an electric field through the temperature of the heat bath is varied, the internal
the gas. The collisions with the gas produce an effective energy of the selected clusters changes and thus also the
friction force. Clusters having a small collision cross sec- number of evaporated atoms 共see Fig. 21兲.
tion experience a smaller friction force and arrive first. In the method of Bachels et al. 共2000兲 Sn clusters are
A new approach to the study of the melting processes in produced by a laser ablation source using a pulsed
gallium nanoparticles embedded in a matrix has been nozzle whose temperature is variable. Cluster tempera-
recently developed by Parravicini et al. 共2003a, 2003b兲. tures should deviate from nozzle temperatures only by
This method is based on capacitance measurements 10– 20 K. Neutral clusters are studied, so that there is no
through the derivative of the dielectric constant with re- mass selection; the width of the distribution of the clus-
spect to T. ter sizes can be of the order of 60% of the average size.
Calorimetry experiments have been performed on de- Energy is measured by a sensitive pyroelectric foil in
posited clusters 共for example, Lai et al., 1996, considered which the clusters impinge, causing a temperature in-
tin clusters on a SiN substrate兲 and free clusters. Free- crease which leads to a measurable voltage jump.
cluster calorimetry has been applied by Haberland’s
group 共Schmidt et al., 1997, 1998, 2001; Kusche et al.,
1999兲 to Na clusters and by Bachels et al. 共2000兲 to Sn
clusters. Haberland’s method consists of two steps, the 2. Computational methods
preparation of size-selected clusters of known tempera-
ture and the determination of their energy. A beam of Here we first discuss the quantities that may be used
cluster ions is produced and thermalized in a heat bath to single out the melting transition, and then we briefly
at temperature T1; a mass spectrometer is used to select review the most common computational methods.
a single cluster size. Then the clusters are irradiated by a The most common method for studying the melting
laser beam and absorb several photons. The basic idea transition is the calculation of the caloric curve, that is,
of the experiment is that absorbtion of photons of en- the total cluster energy E as a function of T, in a simu-
ergy ␦U = h␯ will raise the temperature beyond T1. The lation where the cluster is heated up from a low-T solid
photon energy quickly relaxes into vibrations and heats configuration. E共T兲 may show a smooth jump in the
the cluster to a temperature T2 at which the clusters do melting region 共see Fig. 22兲, corresponding to a peak in
not emit atoms on the time scale of the experiment 共sev- the heat capacity c共T兲 = ⳵E / ⳵T, as in Fig. 23. However,
eral microseconds兲. Only the adsorption of more pho- the caloric curve is not always an efficient indicator, be-
tons from the laser pulse raises the temperature above cause the jump may be small and difficult to find, or
Tevap, the temperature needed for evaporation of atoms even absent. When properly calculated, peaks in the
from the cluster. The size distribution of the remaining heat capacity can indicate the melting temperature Tm.
cluster ions is measured, and this is a very sensitive mea- However, when multiple peaks are present, the assign-
sure of the cluster internal energy. The increased tem- ment of the melting point may become difficult. One
perature of the cluster, T2, is then identified by increas- may identify Tm with the temperature of the highest
ing the cluster source temperature until the thermally peak, but sometimes this criterion fails 共see the discus-
heated clusters show the same photofragmentation be- sion in Frantz, 2001兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 399

tures. Then two distinct phases can be identified and can


be said to coexist 共Wales, 2003, and references therein兲.
The Landau free energy is defined as Fl共Q兲
= kBT ln P共Q兲 + const, where P共Q兲 is the probability of
observing a value Q of the order parameter for the tran-
sition in the simulation run. Thus, in an equivalent way,
one can find that the probability distribution of the or-
der parameter has two minima in the canonical en-
semble. As N → ⬁, the free-energy barrier between these
minima is expected to increase and the phase transition
tends to the first-order bulk phase transition.
The practical problem is now to find an adequate set
of order parameters. A possible order parameter is the
total potential energy Ep or just the difference between
FIG. 22. 共Color in online edition兲 Caloric curves from a the configurational energy and the global-minimum en-
molecular-dynamics melting simulation: 쎲, Ag38; 䊏, Cu38. The ergy 共Lynden-Bell and Wales, 1994; Shah et al., 2003兲.
heating rate is of 0.1 K / ns. The clusters are modeled by the
Other parameters are related to the local ordering, such
Rosato et al. 共1989兲 potential. E* is the average total energy E
of the cluster after the subtraction of the global-minimum en- as the orientational bond order parameters QL 共Stein-
ergy and of the kinetic and harmonic contributions: E* = E hardt et al., 1983兲, or the signatures14 of the common-
− EGM − 3共N − 1兲kBT. The jump in the caloric curve is clear for neighbor analysis 共Faken and Jónsson, 1994兲. The radial
Ag38, and less evident for Cu38. Figure courtesy of Christine distribution function is also routinely monitored. This is
Mottet. useful also to single out surface disordering phenomena
共Qi et al., 2001; Huang and Balbuena, 2002兲.
In general, criteria for distinguishing solidlike and liq- Solidlike and liquidlike phases may also be distin-
uidlike phases and sensitive indicators of the melting guished by their different mobility. In bulk systems, the
transition are needed. Lindemann criterion is commonly used. The atomic vi-
Solidlike and liquidlike phases can be distinguished if brational amplitude 具⌬r2典1/2 is defined as a disorder pa-
an order parameter Q can be found such that the Lan- rameter ⌬L; experiments and simulations show that its
dau free energy Fl共Q兲 is bistable for a range of tempera-
critical value is around 0.10–0.15 in units of the atomic
spacing 共Bilgram, 1987; Lowen, 1994兲. For irregular fi-
nite systems, however, the pure Lindemann criterion is
not appropriate; a better idea is to introduce the
distance-fluctuation measure ⌬DF, as proposed by Etters
and Kaelberer 共1977兲 and Berry et al. 共1988兲:

14
In the common-neighbor analysis a signature is assigned to
each pair of neighbors. This signature is a triplet of integers
共r , s , t兲, where r is the number of common nearest neighbors of
two atoms of the pair, s is the number of nearest-neighbor
bonds among the r common neighbors, and t is the length of
the longest chain that can be formed with the s bonds. We have
found that the monitoring of the signatures 共r , s , t兲 = 共5 , 5 , 5兲,
共4,2,1兲, 共4,2,2兲 is sufficient to distinguish icosohedral, decahe-
dral, and fcc structures over a wide range of sizes 共Baletto,
Mottet, and Ferrando, 2001a; Baletto, 2003兲 during growth
simulations. The 共5,5,5兲 signature indicates pairs located along
a 共locally兲 fivefold symmetry axis. The 共4,2,1兲 signature is asso-
FIG. 23. Heat capacity and bond-length fluctuations in the 57- ciated with pairs with fcc neighborhood. The 共4,2,2兲 signature
atom Lennard-Jones cluster. Left panel, heat capacity curves. is associated with pairs comprising an atom of a 共locally兲 five-
Open circles and solid line represent the results of parallel fold axis and an atom outside the axis. In perfect fcc clusters
tempering simulations, while the dotted line and the thin solid one expects to find a large percentage of 共4,2,1兲 signatures and
line are the results of Metropolis Monte Carlo runs of different no 共5,5,5兲 and 共4,2,2兲 signatures. Comparing icosahedral and
length. Right panel, ⌬DF calculated in two Monte Carlo simu- decahedral clusters in the same size range, one finds that the
lations of different length 共the longer one corresponds to the percentage of 共5,5,5兲 signatures is much larger in icosahedra,
open circles兲. The temperature at which ⌬DF starts its strong while the percentage of 共4,2,1兲 signatures is larger in decahe-
increase corresponds well to the temperature of the heat ca- dra. Tables with the values of the percentages for these signa-
pacity peak. The heat-capacity c and ⌬DF are given in reduced tures are found in the dissertation of Baletto 共2003; see also
units 共see Frantz, 2001兲. Adapted from Frantz, 2001. Baletto, Mottet, and Ferrando, 2001a兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


400 F. Baletto and R. Ferrando: Structural properties of nanoclusters

2 冑具⌬r2ij典 from ergodic higher-T simulations to the quasiergodic


⌬DF = 兺
N共N − 1兲 i⬍j 具rij典
, 共30兲 lower-T simulations. Monte Carlo J-walking methods,
for example, couple the usual small-scale Metropolis
where rij is the distance between atoms i and j, 具⌬r2ij典 moves made by a lower-T random walker with occa-
= 具共rij − 具rij典兲2典. The key difference between the Linde- sional large-scale jumps that move the random walker
mann and distance-fluctuation criteria is that the latter is out of confined regions of configurational space. These
based on the fluctuation of the distance between pairs of large-scale jumps are to configurations that are obtained
atoms while the former is based on the fluctuation of in higher-T simulations. Calvo et al. 共2000兲 and Neirotti
individual atoms relative to their average position. The et al. 共2000兲 showed that parallel tempering is remark-
critical value of ⌬DF for a solid-liquid transition has been ably successful in overcoming quasiergodicity and in cal-
suggested to be around 0.03–0.05 共Zhou et al., 2002兲. culating accurate heat-capacity curves. In parallel tem-
Some caveats are necessary when using ⌬DF. As pointed pering, several simulations are run in parallel at
different temperatures. Periodically, exchange moves are
out by Frantz 共1995, 2001兲, ⌬DF depends on the length of
attempted between configurations at nearby tempera-
the simulation run. For very long simulations, a value
tures, and accepted or rejected according to the Me-
⌬DF ⬃ 0.1 can be obtained even at low temperatures, at
tropolis rule.
which the cluster is simply making isomerizations among
permutational isomers in its solid form. However, com-
puting ⌬DF can be a very sensitive method for monitor- 3. Size dependence of the melting point
ing qualitatively the melting transition from below
共Frantz, 2001兲. The size dependence of the cluster melting point for a
In order to have quantitative information on the mo- given material usually shows a monotonic decrease with
bility of the cluster particles, while avoiding the short- decreasing size, and has irregular variations on a fine
comings of ⌬DF, one can consider quantities that do not scale. Here we first deal with the justification for the
depend on the simulation length. Rytkönen, Valkealahti, monotonic trend and then treat the fine-scale nonmono-
and Manninen 共1998兲 monitored the average rate of tonic variations.
change in the nearest neighbors of the cluster atoms. The average dependence of a melting point with size
Another possibility is to compute the number of distinct N has been derived by means of a few phenomenologi-
basins visited per unit time. cal models. The classical calculation by Pawlow 共1909兲
The melting of nanoclusters has largely been studied has been further extended and modified by several
by standard simulation methods such as molecular dy- groups. A recent account of some of these developments
namics and Monte Carlo 共see, for example, Frenkel and is given by Chushak and Bartell 共2001b兲. Here we derive
Smit, 1996兲. Unlike the Monte Carlo method, which is Pawlow’s formula following the approach of Buffat and
based on a fictitious dynamics, molecular dynamics Borel 共1976兲. We consider a cluster of size N and of
closely mimics the true dynamics of the system. This al- spherical shape. At a given pressure p, its melting tem-
lows the calculation of equilibrium time-dependent cor- perature is Tm共N兲, which has to be compared with the
relation functions, like those related to the diffusion of bulk melting temperature Tm共⬁兲. In analogy with bulk
the cluster particles. Molecular dynamics also permits melting, one identifies the solid-liquid transition by
the realistic simulation of melting and freezing pro- equating the chemical potentials ␮s and ␮l of the solid
cesses, which take place on finite time scales and thus and of the liquid, so that Tm共N兲 at a given pressure p
may include kinetic effects. On the other hand, Monte follows from the solution of this equation:
Carlo can be faster in sampling the configuration space,
being thus more appropriate if one is interested in static ␮s共p,T兲 = ␮l共p,T兲. 共31兲
quantities only.
In fact, a major problem when calculating thermody- This equation means that the chemical potential of a
namic properties by simulation is quasiergodicity, that is, fully liquid and of a fully solid cluster are equal at melt-
the incomplete sampling of the configurational space ing. The chemical potential can be expanded around its
that may occur in the phase-change region. Quasiergod- value at the triple point; we retain first-order terms only:
icity can lead to overestimated transition temperatures
in molecular-dynamics caloric curves 共Calvo and Guet, ⳵␮ ⳵␮
␮共p,T兲 = ␮共p0,T0兲 + 共T − T0兲 + 共p − p0兲. 共32兲
2000兲 if the heating rate is too fast. ⳵T ⳵p
Various methods have been employed to reduce the
systematic errors resulting from quasiergodicity, includ- From the Gibbs-Duhem equation 共−Vdp + SdT + Nd␮
ing the histogram, jump-walking, smart-walking, and = 0兲 it follows that
parallel tempering methods.15 Many of these methods
⳵␮ ⳵␮ 1
are based on the coupling of configurations obtained = − s, = , 共33兲
⳵T ⳵p ␳
15
See, for example, Frantz et al. 共1990兲, Labastie and Whetten where s = S / N is the entropy per particle and ␳ = V / N is
共1990兲, Calvo et al. 共2000兲, Ghayal and Curotto 共2000兲, Neirotti the number density. From Eqs. 共31兲–共33兲, and taking into
et al. 共2000兲, Frantz 共2001兲, Fthenakis et al. 共2003兲. account that ␮s共p0 , T0兲 = ␮l共p0 , T0兲, one obtains

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 401

1 1
− sl共T − T0兲 + 共pl − p0兲 = − ss共T − T0兲 + 共ps − p0兲.
␳l ␳s
共34兲
Here, one must distinguish between the pressure ps of a
solid cluster and the pressure pl of a liquid cluster. In
fact, the pressure inside a small object of radius r is
larger than the external pressure because of the Laplace
contribution pLap = 2␥ / r to the pressure, where ␥ is the
interface tension of the cluster and r is its radius. Since a
liquid and a solid cluster of N atoms can differ both in
interface tension 共␥lv for the liquid-vapor interface, ␥sv
for the solid-vapor interface兲 and in radius, they can
have different Laplace pressure terms,
FIG. 24. A particle with a solid core of radius r − ␦ and a liquid
2␥lv 2␥sv shell of thickness ␦.
pl = pext + , ps = pext + . 共35兲
rl rs
For a cluster of r ⬃ 10 nm, typical interface tensions are the case of metallic particles was given by Kofman et al.
of the order of 103 erg/ cm2, so that the Laplace pressure 共1994兲 and Vanfleet and Mochel 共1995兲; here we briefly
is much larger than pext under the usual conditions. Thus sketch their approach. The extension of this approach to
pext can be neglected in Eq. 共35兲. Moreover, for spherical molecular clusters 共by including van der Waals forces兲
clusters, can be found in Levi and Mazzarello 共2001兲.

冉冊
Let us consider a particle like that in Fig. 24, contain-
rs ␳l 1/3
ing Nl particles in the liquid shell. Its free energy G is
= . 共36兲
rl ␳s given by
Substituting Eqs. 共35兲 and 共36兲 into Eq. 共34兲, neglecting G = 共N − Nl兲␮s + Nl␮l

冋冉 冊 册
pext, and taking into account that L = T0共sl − ss兲 is the la-
tent heat of melting per particle, one obtains r−␦ 2
+ 4␲r2 ␥sl + ␥lv + S⬘e−␦/␨ , 共39兲

1−
Tm共N兲
=
2
Tm共⬁兲 ␳srsL

␥sv − ␥lv s
␳l
冋 冉冊册 2/3
. 共37兲
where
r

Since rl ⬀ N , one finds that Eq. 共37兲 can be rear-


冋 冉 冊册
1/3
r−␦ 2
ranged to the form 共Reiss et al., 1988兲 S⬘ = ␥sv − ␥lv + ␥sl . 共40兲

冋 册
r
C
Tm共N兲 = Tm共⬁兲 1 − , 共38兲 Here ␨ is the characteristic length of the interaction
N1/3
among atoms in liquid metals, and the term S⬘e␦/␨ takes
where C is a constant. This formula gives a very simple into account the effective interaction between the solid-
dependence for Tm共N兲 which can thus be considerably liquid and liquid-vapor interfaces. This effective interac-
lower than Tm共⬁兲 共Buffat and Borel, 1976; Lewis et al., tion is repulsive and favors the formation of a liquid
1997; Rytkönen, Valkealahti, and Manninen, 1998兲. An shell between the solid core and the vapor. For T not too
expression for C easily follows from Eq. 共37兲, but the far from Tm共⬁兲 one may approximate Nl共␮l − ␮s兲
quantities involved in that expression may not be easy to ⯝ Vl␳L关Tm共⬁兲 − T兴 / Tm共⬁兲, where Vl = 4␲关r3 − 共r − ␦兲3兴 / 3 is
evaluate. Moreover, several rather crude approxima- the volume of the liquid layer, and the density difference
tions are involved in Eq. 共37兲, so that usually C is con- between liquid and solid is neglected 共␳l = ␳s = ␳兲. Mini-
sidered as a fitting parameter.
mizing G with respect to ␦, one finds the following solu-
Several attempts have been made to improve
tion for Tm共N兲:
Pawlow’s theory. Buffat and Borel 共1976兲 included
higher-order terms in Eq. 共32兲, and solved numerically
Tm共N兲 2␥sl S ⬘r 2
the resulting equation. Including higher-order terms is in 1− = 共1 − e−␦/␨兲 + e −␦/␨ .
principle more accurate, but these terms increase the Tm共⬁兲 ␳L共r − ␦兲 ␳L␨共r − ␦兲2
number of unknown parameters to be evaluated. Reiss 共41兲
and Wilson 共1948兲, Hanszen 共1960兲, and finally Sambles
共1971兲 refined Pawlow’s model by including the possibil- Pawlow’s result of Eq. 共37兲 is recovered in the limit ␨
ity of surface melting, that is, of having clusters 共of total → 0, which means a sharp interface, or equivalently, no
radius r兲 made of an inner core of radius r − ␦ and an effective repulsive interaction between the solid-liquid
external liquid shell of thickness ␦. The melting tem- and the liquid-vapor interfaces. For molecular clusters,
perature is found by imposing the equilibrium condition this effective interaction is proportional to 1 / ␦2 instead
on this solid-core/liquid-shell particle. A derivation for of e−␦/␨ 共Levi and Mazzarello, 2001兲.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


402 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 26. Comparison of theoretical and experimental melting


points of supported tin clusters: 쎲, experiment; solid line, the
fitting by means of Eq. 共42兲. From Lai et al., 1996.
FIG. 25. Melting point of gold clusters from molecular-
dynamics simulations 共black dots兲, and comparison with differ- metal clusters 共Rytkönen, Valkealahti, and Manninen,
ent theoretical results: solid line, Pawlow’s theory; ⫻, second- 1988; Valkealahti and Manninen, 1993; Frantz, 2001兲兴.
order corrections from Buffat and Borel 共1976兲; heavy dashed However, the situation can be more complicated.
curves, the liquid-shell model by Sambles 共1971兲, Eq. 共42兲; thin Schmidt et al. 共1998兲 measured the melting point of Na
dashed curves, second-order corrections to Sambles’s formula. clusters up to 200 atoms. They found an irregular behav-
Reprinted with permission from Chushak and Bartell 共2001b兲.
ior of Tm, with few well-defined peaks, as shown in Fig.
Copyright 2001 American Chemical Society.
18, and L showed the same kind of behavior. The peaks
in Tm共N兲 and L were not well correlated with those in
Following a somewhat different line of reasoning, the mass spectrum, corresponding to the closing of elec-
Chushak and Bartell 共2001b兲 derived a similar expres- tronic shells. These results showed that the most abun-
sion, which is close to those found by Hanszen 共1960兲 dant clusters and the highest-melting-point clusters do
and Sambles 共1971兲: not always coincide. In a recent experiment, Schmidt et

1−
Tm共N兲
=
2 ␥sl

Tm共⬁兲 ␳lL r − ␦
+
␥lv
r
1−
␳s
␳l
冋 冉 冊 册冎 2/3
. 共42兲
al. 共2003兲 were able to measure the energy and the en-
tropy change in the melting of Na clusters. They found
that the peaks in Tm共N兲 are driven by the energy differ-
If ␥sl is evaluated as ␥sv − ␥lv 共that is, if the liquid shell ence between the liquid and the solid phases. The en-
perfectly wets the solid core兲, and ␦ = 0, Eq. 共42兲 reduces tropy difference is closely correlated to the energy dif-
to Eq. 共37兲. ference and simply causes a damping of the energetic
Both Eqs. 共41兲 and 共42兲 are more accurate than effects. This would show that the main indicator for a
Pawlow’s result down to small sizes 共see Fig. 25兲. Peters high Tm is the energetic separation of the global mini-
et al. 共1998兲 found that the solid-core/liquid-shell model mum from higher isomers. However, it is not yet clear
gives a better fit of the experimental melting tempera- whether entropic differences are always correlated to
tures of supported lead clusters. Chushak and Bartell energetic differences, or whether this is a trait of Na
共2001b兲 and Wang, Zhang, et al. 共2003兲 melted a series of clusters.
selected fcc clusters of gold and copper, respectively, by The irregular variations of the melting point in
molecular-dynamics simulations within an embedded- Lennard-Jones clusters were studied by Frantz 共2001兲,
atom energetic model. They found that Eq. 共38兲 is accu- who thoroughly analyzed the range 26艋 N 艋 60 by the
rate down to N ⬃ 103 and 5 ⫻ 102 for Au and Cu, respec- parallel tempering technique. Frantz 共2001兲 found that
tively. At smaller sizes, Eq. 共42兲 is in better agreement the peaks in Tm共N兲 were generally well correlated with
with the simulation data in the case of Au. Other simu- those of other stability indicators, such as the energetic
lation results are given by Ercolessi et al. 共1991兲, Lewis et separation of the second isomer from the global mini-
al. 共1997兲, and Qi et al. 共2001兲. Lai et al. 共1996兲 obtained mum ESM共N兲 − EGM共N兲, ⌬1共N兲, and ⌬2共N兲 关see Eq. 共6兲兴.
an excellent fit to their experimental data on the melting An exception was N = 58, a magic size for the stability
of tin clusters 共see Fig. 26兲 by the formula of Hanszen indicators, but not corresponding to a peak of Tm共N兲.
共1960兲; the same data were successfully refitted by means Finally, even though the melting point in nanoclusters
of Eq. 共41兲 共Bachels et al., 2000兲. is usually depressed, some evidence of exceptions to this
When the size dependence of the melting point is ex- rule is beginning to accumulate. The first experimental
amined on a fine scale, irregular variations are found, evidence of an exception is due to Shvartsburg and Jar-
especially at small sizes, where the addition or the re- rold 共2000兲, who studied the melting of Sn clusters. Small
moval of a single atom can have dramatic effects. There Sn clusters have a rather elongated structure, which
is a good agreement among several simulation results in should change to nearly spherical upon melting. No sig-
predicting that clusters of special stability, such as icosa- nature of this change was observed by Shvartsburg and
hedral clusters at magic sizes 共55, 147, etc.兲, melt at Jarrold 共2000兲, so they concluded that Sn cluster ions
higher temperatures than predicted by Eq. 共38兲 关see, for containing 10–30 atoms have a melting point at least
example, the cases of Lennard-Jones, Na, and noble- 50 K above the bulk one. This behavior has been con-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 403

firmed by Joshi et al. 共2002兲, who performed ab initio the melting takes place, Rytkönen et al. calculated the
molecular-dynamics simulations of the melting of Sn10. radial distribution function, finding that, at nonmagic
They found that binding in such a small cluster is indeed icosahedral sizes, there is a clear tendency to surface
covalent, and that the specific heat of Sn10 shows a melting while the cores preserve an icosahedral mor-
shoulder around 500 K due to a permutational rear- phology.
rangement of atoms that preserves the trigonal prism Frantz 共1995, 2001兲 made a detailed systematic study
core of the ground state. Only at much higher tempera- of small Lennard-Jones clusters by Monte Carlo simula-
tures, T ⯝ 1500 K, does this core distort and break up, tions. He found that the smallest clusters presented an
yielding a peak in the specific heat around 2300 K. Joshi irregular dependence of their thermodynamic properties
et al. 共2003兲 also simulated the melting of Sn20, finding on size. For N ⬎ 25 some trends were pronounced. Icosa-
that it has a lower melting point than Sn10, but one that hedral packing was dominant 共with the exception of the
is still much higher than that of bulk Sn. Chuang et al. truncated octahedron at N = 38兲 and the heat capacity
共2004兲 have found by ab initio Langevin molecular dy- peak parameters formed two overlapping sequences as a
namics that Sn6, Sn7, and Sn13 also melt at a higher T function of N, depending on whether the overlayer had
than bulk Sn. Mackay or anti-Mackay stacking. In clusters with an
Small Sn clusters seem not to be the only ones to melt anti-Mackay overlayer, the heat-capacity peak shifted
at higher temperature than the bulk crystal. There is towards higher T and became smaller at increasing N,
recent experimental and theoretical evidence for gallium while clusters with Mackay rearrangements had small,
clusters 共Breaux et al., 2003; Chacko et al., 2004兲, whose low-T peaks that generally shifted to higher T and grew
high melting point is attributed to the fact that, in con- in size as N increased. There was a sequence of magic
trast to bulk Ga, binding in small clusters is fully cova- numbers, N = 36, 39, 43, 46, 49, 55, at which the heat-
lent. Finally, in their simulations, Akola and Manninen capacity peak was stronger. As already discussed in Sec.
共2001兲 observed a bulklike behavior of Al13− above the IV.B.3, this sequence correlates well with that extracted
Al bulk melting temperature. from the binding energy differences.
As a final remark, we note that thermodynamic effects
on structural stability and the melting of nanoclusters 2. Sodium clusters
suggest a more general definition of the stability, includ-
ing temperature effects 共Baletto, Rapallo, et al., 2004兲. Experiments by Martin et al. 共1994兲 and Schmidt et al.
共1998兲 have revealed a complex dependence of Tm共N兲
for Na clusters. Its irregular small-size behavior has al-
C. Studies of selected systems ready been discussed in Secs. IV.B.3. Liu et al. 共2002兲
performed a systematic simulation study of the melting
1. Lennard-Jones clusters
of different morphologies over a wide size range. They
There are several computational studies of melting in found that the melting points for all sizes and structural
Lennard-Jones clusters, especially at magic icosahedral types were in a narrow T range 共200– 300 K兲 and all
sizes. Early molecular-dynamics studies on the melting clusters presented a liquid-gas transition around 1000 K.
of Ar clusters by Briant and Burton 共1975兲 indicated a Another interesting point about the melting of Na
relatively sharp first-order-like transition at T below the clusters, which is of great general interest, is the pres-
bulk melting temperature. This result was later con- ence of premelting effects. A detailed study of this topic
firmed. At N = 55 and N = 147 all simulations indicate at 8 艋 N 艋 147 atoms has been performed by Calvo and
relatively sharp jumps in the caloric curves at T ⯝ 35 K Spiegelman 共2000兲 by a Monte Carlo thermodynamical
and T ⯝ 42 K, and the radial distributions confirm that analysis close to the solid-liquid transition, within a
whole clusters melted during the transitions 共Etters and semiempirical modeling by Gupta 共1981兲 potentials. Up
Kaelberer, 1975; Honeycutt and Andersen, 1987; Mat- to 147 atoms, the thermodynamics appears to be directly
suoka et al., 1992兲. At N = 13 the cluster fluctuates back related to the lowest-energy structures, and melting by
and forth between the two phases and a consistent de- steps is favored by the presence of surface defects. For
termination of Tm is very difficult. The breadth of the N ⬍ 75 the Tm presents a strong nonmonotonic behavior
solid-liquid transition is also revealed for Ar13 adsorbed with N, which is typical of geometric size effects. For
on a surface 共Blaisten-Barojas et al., 1987兲, where the larger sizes the transition becomes more and more simi-
cluster changes from an icosahedral, solidlike structure lar to the bulk case. Calvo and Spiegelman 共2000兲 find
at low T to a set of liquidlike structures at high T. Clus- evidence of premelting phenomena at small sizes, with
ters containing 309 or more atoms are observed to des- the caloric curves presenting multimodal behavior with
orb atoms at temperatures where the core is still solid increasing T. Simple isomerization between a few struc-
共Rytkönen et al., 1997兲. Thus it is very complicated to tures, surface melting, or competition between several
define their melting point, if the heating rate is slow. Fast funnels on the energy landscape may be the causes of
heating rates can, however, cause the superheating of this premelting. Premelting seems to be the rule for
the cluster. Rytkönen, Valkealahti, and Manninen 共1998兲 small Na clusters, the only exceptions being the very
found a rather good agreement with Eq. 共38兲, except stable magic icosahedral structures. The secondary
that Tm at small sizes and for icosahedral structures is peaks in the heat capacity become less pronounced at
higher than predicted analytically. To investigate how N ⬎ 100, indicating that premelting becomes less impor-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


404 F. Baletto and R. Ferrando: Structural properties of nanoclusters

tant. Calculations by Gupta potentials and ab initio Li, Ji, et al. 共2000兲 made a thorough analysis at N
schemes 共Rytkönen, Häkkinen, and Manninen 1998, = 55 for Au, Ag, and Cu by means of molecular-
1999; Reyes-Nava et al., 2002兲 agree reasonably well dynamics simulations. Ag55 and Cu55 showed abrupt
with the experiments, but several aspects of the melting changes during the meltinglike transitions, while the
behavior of Na clusters are still to be understood. Calvo transition of Au55 seemed to proceed for a relatively
and Spiegelman 共2004兲 have recently reexamined the broader interval, with ⌬DF increasing gently from 300 up
melting behavior of Gupta Na clusters in the size range to 600 K and with a small, ladderlike energy jump in the
around 130 atoms, finding that the potential is somewhat caloric curve. The degrees of reduction for Tm were dif-
inadequate because it predicts strong premelting effects ferent among the three metals. Compared to Ag and Cu,
that are not observed in the experiments. These pre- Au exhibited the largest size-induced drop of Tm. This
melting effects are related to the surface melting of the
confirms the trends of Table I, since Cu also gives the
outer icosahedral shell.
most stable icosahedron cluster from the thermody-
namic point of view.
At N = 38 a complex melting behavior for Cu was
3. Noble-metal and transition-metal clusters found by Baletto et al. 共2004兲, using molecular-dynamics
simulations. A solid-solid transition took place from the
In the case of metallic clusters, the literature is exten- truncated octahedron at the global minimum. Then the
sive, especially for Au clusters.16 To begin, we remark cluster rearranged itself into defected decahedron struc-
that, although the melting point is depressed, its reduc- tures before melting, in analogy with the behavior of the
tion is smaller than that of Lennard-Jones clusters, and Lennard-Jones cluster of the same size 共Doye et al.,
it is strongly material dependent 共Jellinek et al., 1986; 1998兲. On the other hand, Ag38, which has the same
Garzón and Jellinek, 1991; Jellinek and Garzón, 1991兲. global-minimum structure as Cu38, does not exhibit any
Usually melting is accompanied by a peak in the specific solid-solid transition to defected decahedral structures.
heat, and by a substantial rearrangement of the cluster. At this size Au presents a coexistence phase region
But there are exceptions. For example, the simulations among different isomers already at T ⬃ 250 K 共Garzón et
by Westergren et al. 共2003兲 revealed that Pd34 melts with- al., 1999兲.
out an accompanying peak in the heat capacity, and the Molecular-dynamics simulations of the melting transi-
atoms become mobile without any significant change in tions for crystalline nickel clusters above N ⬃ 750 show
geometric structure. García-Rodeja et al. 共1994兲 and Lee
that Tm共N兲 closely follows Eq. 共38兲 共Qi et al., 2001兲,
et al. 共2001兲 studied several transition and noble metals,
while, for N ⬍ 500, icosahedral structures present higher
at very small sizes, around 13 atoms, modeled by Gupta
Tm and large latent heat. Similar results are found also
potentials. Their simulation results were in agreement
with those obtained by Güvenc and Jellinek 共1992兲. The for Cu 共Wang, Zhang, et al., 2003兲. For smaller N, the
salient result is that the behavior of Ni, Cu, Pd, Ag, and icosahedral packing is more stable and presents higher
melting points than those derived from Eq. 共38兲
Pt is similar, the main difference being that Pd13 presents
a more complex anharmonic behavior than Ag 共Wester- 共Valkealahti and Manninen, 1993兲.
gren and Nordholm, 2003兲. The 13-atom clusters un- Gold clusters have been studied intensively in recent
dergo a structural transition from a rigid, solidlike icosa- years 共see Cleveland et al., 1998, 1999; Lee et al., 2001; Li
hedral structure to a nonrigid, liquidlike one via an et al., 2002兲. All these studies agree that a solid-solid
intermediate temperature range in which both forms co- structural transformation from the low-T optimal struc-
exist. The caloric curves do not present any sharp tran- tures to icosahedral structures takes place below the
sition, and Tm is found by monitoring ⌬DF 关Eq. 共30兲兴 and melting temperature. Detailed analysis of the atomic
looking for a maximum in the specific heat. The instabil- trajectories and of the structural evolution indicates that
ity of some specific atom can play a key role in melting, this solid-to-solid transition is essentially without diffu-
as in the case of the capped atoms at N = 14. The atom sive motion, occurring quickly and involving many
added to the 13-atom icosahedron can diffuse rapidly at 共small兲 cooperative displacements of the atoms. The
T below the melting region, and this causes a small peak structural transformation is driven by the vibrational
in the specific heat 共García-Rodeja et al., 1994; Lee et al., and configurational entropy at elevated T 共Luo et al.,
2001兲 so that Tm is smaller than for the 13-atom cluster. 1987; Ajayan and Marks, 1988兲. A thorough molecular-
Similar behavior is found at N = 20 for the capped dynamics simulation study was made by Liu et al. 共2001兲.
double icosahedron. In contrast, when the double icosa- Different morphologies were compared up to N
hedron is not the global minimum, the heat capacity has ⬃ 25 000. At intermediate sizes, for all type of mor-
an unexpected behavior. For example, Pd19 shows a dis- phologies, Liu et al. 共2001兲 found that the melting pro-
tinct abrupt change at T ⬃ 400 K and a rounded-off cess occurs in three stages: a relatively long time of sur-
broad peak at higher T 共Lee et al., 2001兲. face disordering and reordering, a relatively short time
of surface melting, and finally a rapid overall melting.
Concerning the differences among the structural motifs,
16
See Buffat and Borel 共1976兲, Garzón and Jellinek 共1993兲, starlike decahedra are the hardest to melt, while icosa-
Cleveland et al. 共1997, 1998, 1999兲, Li, Ji, et al. 共2000; Li, Lee, et hedron clusters are the easiest, with regular decahedra
al., 2002兲, Lee et al. 共2001兲, and Koga et al. 共2004兲. being in between. Small cuboctahedra, up to N = 309 at

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 405

least, transform into icosahedra before melting, while cluster melts at temperatures comparable with other
large ones do not. Contrary to the other motifs, Tm of nearly spherical structures.
icosahedra saturates at N ⬎ 12 000. Liu et al. 共2001兲 also
compared different structures at identical sizes, finding
V. KINETIC EFFECTS IN THE FORMATION OF
that transition temperatures could differ by as much as NANOCLUSTERS
75 K. This nonequilibrium effect indicates the differ-
ences in kinetic stability of the different structures From the results reported in the previous sections,
against heating. one can see that there are discrepancies between the
Very recently Koga et al. 共2004兲 were able to heat up results of energetic and/or thermodynamic calculations
metastable gold icosahedra produced in an inert-gas ag- and the real outcomes of the experiments, viz., the struc-
gregation source 共see Sec. V.B.3兲. Upon heating to T tures that are actually observed in the production of free
= 1273 K, which is only 64 K below the bulk melting solid clusters. In some cases, the disagreement is of such
temperature of Au, they were able to produce a clearly a qualitative nature that it is unlikely to be due to the
dominant fraction of decahedron structures in the range failure of the energetic or of the thermodynamic model-
6 – 12 nm diameter. A considerable fraction of fcc clus- ing. To cite just the most striking cases, we mention the
ters was obtained only heating above the bulk melting observation of small decahedra and large icosahedra in
temperature, and only for diameters above 10 nm. Since the inert-gas aggregation experiments on the production
all calculated crossover sizes decahedron→ fcc at 0 K of Ag clusters 共Reinhard, Hall, Ugarte, and Monot,
are much smaller 共Cleveland et al., 1997; Baletto, Fer- 1997兲, the production of large icosahedral clusters of C60
rando, et al., 2002兲, these experimental results give sup- molecules 共Martin et al., 1993; Branz et al., 2000, 2002兲,
port to the hypothesis of an upward displacement of the and the detection of octahedral Al clusters presenting
crossover size with increasing temperature, in agreement only 共111兲 facets 共Martin et al., 1992兲. Moreover, there is
with the predictions of the harmonic theory 共Doye and still quantitative disagreement between the theoretical
Calvo, 2001兲, as can be seen in Fig. 20. estimates 共based on the total energy optimization兲 and
the experimental data about the crossover sizes for Ar
4. Silicon clusters
clusters 共Ikeshoji et al., 2001兲. These results indicate that
kinetic effects must be taken into account to explain the
Wang, Wang, et al. 共2001兲 performed a systematic actual free-cluster formation in the experiments. As dis-
simulation study of structural transitions and thermody- cussed in Sec. II, the time scale of nanocluster formation
namic properties of small Si clusters within a tight- in typical sources ranges from a fraction of a millisecond
binding molecular-dynamics scheme. They confirmed to a few milliseconds. On this time scale, clusters may
previous results on the thermodynamics of very small not be able to reach the minimum free-energy structure,
clusters found by means of empirical many-body poten- thus remaining trapped in some metastable configura-
tials 共Stillinger and Weber, 1985; Blaisten-Barojas and tion that can have a very long lifetime, especially when
Levesque, 1986; Tchofo-Dinda et al., 1995兲, and also con- the clusters are further cooled down after their solidifi-
sidered larger sizes. Their canonical Monte Carlo simu- cation.
lations revealed that the melting point of Si clusters In studying the formation process of solid clusters in
changes dramatically when the global-minimum struc- contact with a thermal bath, such as the inert-gas atmo-
tures change from prolate and cagelike 共for N = 17 in this sphere in inert-gas aggregation sources, we can think of
study兲 to an atom-centered and nearly spherical mor- two alternative models. In both models, clusters grow
phology. The nearly spherical clusters present a much mainly by the addition of single atoms 共de Heer, 1993兲.
broadened melting region, extending from 750 to In the first model 共Sec. V.A兲, which is suited for high
1300 K. This is because, in nearly-spherical structures, growth temperatures, the cluster solidifies after its
the cluster core is more tightly bonded than the surface growth is completed, that is, at a further cooling stage
atoms, and a much higher temperature is needed to dis- taking place outside the growth chamber. In this case,
order it. In the range 650⬍ T ⬍ 1050 K, the nearly the final cluster structure does not depend on the kinet-
spherical clusters keep their structure, although there is ics of the growth process, because the cluster remains a
a noticeable surface diffusion. After that, melting can liquid droplet while growing, but rather depends on the
take place via two pathways. The first possibility is that, kinetics of the cooling after growth, which may take
as T increases, the clusters develop to the prolate mor- place on time scales like those discussed in Sec. V.A,
phology up to 1200 K, then break into subunits, which namely, in the range of 1 – 10−2 K / ns. This model will be
finally become less and less stable until the clusters dis- referred to as the liquid-state growth model. It is simu-
order completely. The other possibility is that the clus- lated by freezing a liquid droplet until it solidifies, usu-
ters develop from quasispherical to prolate to molten ally at constant N and decreasing T. In fact, there are
oblate structures. Atoms in prolate cagelike structures systems 共such as the noble gases兲 in which evaporation
are more stable than the surface atoms of the nearly of atoms is non-negligible at temperatures close to the
spherical structures. The melting of cagelike structures melting point, because melting and boiling temperatures
often accompanies the overall deformation and frag- are close to each other. Evaporating atoms can thus play
mentation of the cage framework. Thus a much higher an important role even in the cooling of solid clusters. In
temperature is needed to break the cage, but the whole contrast, metals present huge differences between melt-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


406 F. Baletto and R. Ferrando: Structural properties of nanoclusters

ing and boiling points, so that evaporative cooling is


negligible for solid metallic clusters.
In the second model 共Sec. V.B兲, which is suited for
relatively low growth temperatures, clusters solidify
while they are still growing.17 This model will be re-
ferred to as the solid-state growth model. In this model,
the final outcome is determined by the kinetics of the
growth process itself. Solid-state growth is simulated by
adding single atoms to a small initial seed at constant
temperature 共Baletto, Mottet, and Ferrando, 2000a,
2000b兲.
Finally 共see Sec. V.C兲, there are cases in which growth
proceeds not only by the addition of single atoms, but FIG. 27. 共Color in online edition兲 Percentages of the different
also by the collision and coalescence of already formed structures obtained in molecular-dynamics simulations of Ag
clusters. cluster freezing at rc = 1 K / ns for 130⬍ N ⬍ 310: 쎲, icosahedral
structures; 䊏, decahedra; 䉱, fcc clusters. Adapted from Bal-
etto, Mottet, and Ferrando, 2002.
A. Freezing of liquid nanodroplets

When simulating cluster freezing, the relevant param- ration method let the cluster evaporate atoms in a simu-
lation at constant energy. They considered 380 clusters
eter is the rate rc at which the temperature T is rescaled,
in order to mimic a thermal contact with a cold atmo- in the range 160⬍ N ⬍ 2200, finding that there was a
sphere. For example, the cluster can be cooled down by transition at increasing size from icosahedra to a mixture
of structures 共decahedron, fcc, hcp, and icosahedron兲.
small steps ␦T at each time interval ␦t so that rc
The transition did not depend on the cooling method,
= ␦T / ␦t. In an inert-gas atmosphere, one can estimate
and took place for N ⯝ 450. This is lower than the tran-
that a cluster of radius R and area Aeff = 4␲R2 will collide
sition size 共750 atoms兲 observed in experiments on argon
with gas atoms with a frequency ␾exp G
given by clusters 共Farges et al., 1986兲, but it is closer than any
PGAeff other estimate based on total energy optimization, thus
␾exp
G
⬃ 共43兲
冑2 ␲ m G k B T G , indicating the possible presence of kinetic trapping ef-
fects. Moreover, calculated and experimental diffraction
where PG and TG are the pressure and the temperature, patterns were in good agreement. Besides pure clusters,
respectively, of the inert gas of mass mG. In the har- Ikeshoji et al. 共2001兲 considered binary Lennard-Jones
monic approximation, the energy loss is given by systems, finding that the formation of large icosahedra
was favored by the size mismatch. This finding may fur-
␦T ␦E nish a qualitative explanation of the experimental obser-
⬃ ␾exp
G
, 共44兲
␦t 3NkB vation of large icosahedron clusters in binary 共metallic兲
where ␦E is the energy transfer at each collision, and systems 共Saha et al., 1997, 1999兲.
since Aeff ⬀ N2 and N ⬀ R3 we obtain ␦T / ␦t ⬀ 1 / R. Follow-
ing Westergren et al. 共1998兲, we can estimate that the loss 2. Silver clusters
for each collision 共for example with a helium atom兲 ␦E is
Baletto, Mottet, and Ferrando 共2002兲 studied the
1 – 10 meV. Using typical parameters for the gas, PG
freezing of Ag liquid nanodroplets by molecular-
⬃ 1 – 100 mbar, TG ⬃ 300 K 共Reinhard, Hall, Ugarte, and
dynamics simulations with realistic cooling rates rc in the
Monot, 1997; Koga and Sugawara, 2003兲 and considering
range 0.1– 5 K / ns. They made a systematic study of Ag
R ⬃ 1 nm, we have ␦T / ␦t in the range 0.01– 1 K / ns.
freezing at 130艋 N 艋 310, and in addition considered the
freezing at the icosahedral magic numbers 147, 309, 561,
1. Lennard-Jones clusters and 923. The latter simulations were made to investigate
Ikeshoji et al. 共2001兲 studied the freezing of Lennard- the possibility of obtaining large metastable Ag icosahe-
Jones clusters by molecular-dynamics simulations. Con- dra by freezing liquid droplets, in order to ascertain
trary to what happens for metals, the melting and boil- whether the liquid-state growth model could explain the
ing points of these clusters are rather close in experimental observation of metastable icosahedra by
temperature, so that the evaporation of atoms during Reinhard, Hall, Ugarte, and Monot 共1997兲.
the solidification process is non-negligible. Because of The results for 130艋 N 艋 310 are summarized in Fig.
that, Ikeshoji et al. 共2001兲 considered two ways of cool- 27. On a coarse-grained description, the energetic calcu-
ing clusters down. The thermostat method decreased lations by Baletto, Ferrando, et al. 共2002兲 showed that
temperature in a canonical simulation, while the evapo- icosahedral clusters are the best up to N ⯝ 170; then the
best structures are decahedral, except for the icosahe-
dral magic number N = 309, while fcc clusters become
17
Clusters may solidify during growth at constant tempera- competitive with the decahedra for N ⬎ 600. The freez-
ture after reaching a critical size. ing simulations of Baletto, Mottet, and Ferrando 共2002兲

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 407

TABLE II. Numbers of icosahedron 共Ih兲, decahedron 共Dh兲,


and fcc structures at magic icosahedral sizes. The results are
obtained from five molecular-dynamics simulations for each
cooling rate rc and for each size. Adapted from Baletto, Mot-
tet, and Ferrando, 2002.

rc 共K / ns兲 N NIh NDh Nfcc

1 147 5
309 4 1
561 2 2 1
923 1 4

5 147 5
309 3 2 FIG. 28. 共Color in online edition兲 Molecular-dynamics simula-
561 2 3 tion of silver cluster freezing. On the left, we show the initial
923 1 4 configurations of Ag liquid nanodroplets at magic icosahedral
numbers 共147, 309, 561, and 923兲, while on the right, there are
20 147 5 typical final structures obtained by cooling at 1 K / ns: from top
309 2 1 2 to bottom, 147-atom icosahedron; 309-atom icosahedron; a
561 1 3 1 decahedron at 561 关strongly asymmetric and with an island on
hcp stacking above 共Baletto and Ferrando, 2001兲兴; a fcc poly-
923 1 3 1
hedron at 923.

gave the following results. At rc = 1 K / ns, particles so- Reinhard, Hall, Ugarte, and Monot 共1997兲, who did not
lidified as icosahedra in the range 135⬍ N ⬍ 165. Around find strong evidence of small icosahedra, while they ob-
N = 165 there was a transition to decahedron clusters, served a dominant percentage of icosahedra at large
which were the most frequent for 170⬍ N ⬍ 245. About sizes. Therefore, one can rule out the liquid-state growth
20% of the runs gave fcc structures in this range, and a model as an explanation for the outcome of this experi-
ment.
few icosahedron clusters were formed. For 245⬍ N
⬍ 310, the dominant structures at the end of the freezing
were icosahedra, but fcc clusters were also common, es- 3. Gold clusters
pecially in the range N ⬃ 250– 260 and N ⬃ 280– 300, in Chushak and Bartell 共2001a, 2001b兲 studied the solidi-
which they made up more than 50%. At slower cooling fication of Au nanoclusters by molecular-dynamics simu-
rates, rc = 0.12 K / ns, the percentage of fcc and icosahe- lations. However, their method is rather different from
dron particles in the range 170⬍ N ⬍ 245 dropped prac- that applied to Ag clusters by Baletto, Mottet, and Fer-
tically to zero, indicating that decahedron clusters are rando 共2002兲. Chushak and Bartell 共2001a, 2001b兲
likely to be the most favored from the thermodynamic started from high-T liquid clusters of rather large sizes
point of view in this size range. On the other hand, at 共459, 1157, and 3943 atoms兲. These clusters were then
faster cooling rates, rc = 5 K / ns, the percentage of fcc cooled down very quickly, at a rate rc = 5 ⫻ 104 K / ns, to a
clusters in the range 245⬍ N ⬍ 310 decreased in favor of temperature of about 700 K, at which their most stable
decahedron clusters, while the dominant proportion of form is solid. At this temperature, production runs of
icosahedra remained practically constant. 1 ns were performed. After that, clusters were further
From the freezing results at the icosahedral magic quenched down to 300 K, again with a fast rate
numbers 147, 309, 561, and 923, it was possible to extract 共300 K / ns兲. As a result of this cooling procedure
the following tendency 共see Table II and Fig. 28兲. The Chushak and Bartell found that icosahedron clusters
percentage of icosahedral structures at the end of the were preferentially produced, even though they should
freezing process decreased with cluster size for all cool- not be the lowest-energy structures at these sizes 共see
ing rates, passing from 100% at N = 147 to less than 20% Baletto, Ferrando, et al., 2002兲. In this case, icosahedron
at N = 923. There was no evidence of changes in the re- clusters were produced with high probability during
sults when rc ranged within 1 – 20 K / ns, probably indi- freezing because of the width of the icosahedron funnel
cating that these rates are sufficient to mimic a freezing 共Doye, 2004兲 with respect to the funnels leading to ei-
process taking place close to equilibrium. ther close-packed or decahedral structures. This is a ki-
In summary, the simulations showed that it is not pos- netic trapping effect, which is stronger at rapid cooling
sible to avoid the formation of a large percentage of rates, and agrees with the results on Ag freezing by Bal-
small icosahedra 共2 – 3 nm diameter兲 if freezing takes etto, Mottet, and Ferrando 共2002兲.
place after the growth is completed. Moreover, the prob- Nam et al. 共2002兲 have studied the mechanism by
ability of forming large icosahedra by freezing liquid Ag which metastable gold icosahedron clusters are frozen
droplets with realistic cooling rates is small. Both find- out of nanodroplets using molecular-dynamics simula-
ings are in contrast with the experimental results of tions within the embedded-atom model. Their simula-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


408 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 29. 共Color in online edition兲 Platinum cluster structures


at N = 55. From left to right icosahedron 共Ih兲, single rosette
共Sr兲, and double rosette 共Dr兲. The Ih is the lowest in energy,
the Sr and the Dr are, respectively, 0.48 and 0.42 eV higher in
energy.

tions started from a liquid droplet at 1500 K, which was


cooled down at a rapid rate of 100 K / ns. Again, the ma-
jority of clusters solidified in the icosahedral symmetry,
with the solidification starting at the cluster surface, in FIG. 30. Melting and freezing molecular-dynamics simulations
such a way that a well-ordered, close-packed surface of Pt55 clusters. In the upper panel we show the melting caloric
with fivefold symmetry points was formed while the curves for for three structures: dashed line, icosahedral 共Ih兲;
cluster was still amorphous inside. The formation of solid line, single rosette 共Sr兲; dash-dotted line, double rosette
such a surface triggered the solidification process to- 共Dr兲. For 600⬍ T ⬍ 700 K, all melting curves exhibit several
wards the formation of icosahedra: Nam et al. 共2002兲 oscillations among the three structures. The lower panel shows
noticed that when ordering began at the surface, the fi- two freezing curves. The dashed line ends with an icosahedral
nal result was an icosahedron cluster. Conversely, if the structure, while the solid line ends with a single-rosette struc-
surface did not order first, the final cluster was either ture. This is an example of the interplay of entropic and kinetic
decahedral or fcc. effects. Due to entropic effects, the Sr structure is the most
likely in the temperature range in which the cluster begins to
solidify. This structure may then be preserved down to low
4. Entropic effects and kinetic trapping in Pt55
temperatures by a trapping effect.
As a paradigm showing the peculiar interplay between
entropic and kinetic trapping effects, we report on a
molecular-dynamics study of the melting of Pt55 共Bal- the double, but it is quantitatively poor because it un-
etto, 2003兲, modeled by the Rosato et al. 共1989兲 poten- derestimates the probability of both with respect to the
tial. Within this model, Pt55 presents the competing probability of the icosahedron 共see Fig. 31兲. Another
structures of Fig. 29: the icosahedron, the single rosette entropy-driven effect concerns decahedral structures.
共Sr兲, and the double rosette 共Dr兲, which are, respectively, Even though some decahedral structures are lower in
the minimum-energy structure, the high-T minimum- energy than the single and double rosette, we find that
free-energy structure, and a structure frequently found the entropic contribution favors single rosettes and
in growth simulations 共Baletto, 2003兲. The single and
double rosettes are obtained by modifying the external
shell of the Ih55. In the single rosette, a single vertex
atom is displaced to form a hexagonal ring 共the rosette兲
together with its nearest-neighbor atoms on the surface.
The two rosettes of the double rosette are at nearby
vertices, giving a close resemblance to the global mini-
mum of Au55 found by Garzón, Michaelian, et al. 共1998兲
and shown in Fig. 13. In the melting simulations 共see Fig.
30兲, the single rosette becomes the most favored struc-
ture for 600⬍ T ⬍ 700 K. In this range, the cluster still
oscillates among icosahedral, single-rosette, and double-
rosette structures. Finally, it melts above 750 K. This is
FIG. 31. Probabilities of the different structures of Pt55 vs T in
due to entropic effects. In fact, calculations within the
the harmonic superposition approximation: solid line, the
harmonic approximation 共see Sec. IV.A.1兲 show that the
probability of the single rosette; dashed line, probability of the
probability of finding the single rosette becomes consid-
icosahedron; dash-dotted line, probability of the double ro-
erable above 600 K, while the probability of the double sette; dotted line 共which coincides with the T axis on this
rosette always remains much smaller. Constant-T simu- scale兲, probability of the decahedron. The last always have a
lations on long time scales 共several ␮s兲 of the evolution very small probability, even if they are energetically favored
of Pt55 above 600 K confirm this scenario, showing also with respect to the single rosette, which is the best structure
that the harmonic approximation is qualitatively right in from the entropic point of view and becomes the most likely at
predicting that the single rosette is more probable than high T.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 409

double rosettes over decahedral structures even at low B. Solid-state growth


temperatures.
The freezing simulations in Fig. 30 show an interplay Even though the liquid-state growth model can ex-
of kinetic and entropic effects. The cluster starts to so- plain some kinetic effects in the growth of nanoclusters
lidify at T ⯝ 700 K, a temperature range in which the 共see, for instance, the cases of lead and argon clusters
single rosette is the most probable structure. Then the treated in Sec. V.A兲, there are still several experimental
cluster is likely to remain kinetically trapped in this results that cannot be explained by it: the growth of
structure down to low temperatures. large icosahedra in clusters of Ag and of C60 molecules,
and the growth of octahedra in Al clusters. For these
systems, it turns out that the solid-state growth model is
5. Copper, nickel, and lead clusters much more appropriate.
Valkealahti and Manninen 共1997兲 performed a system- In the following, we deal first with a very general
atic molecular-dynamics study of the freezing of Cu mechanism for the solid-state growth of metastable
nanodroplets within the effective-medium model. They icosahedra 共Sec. V.B.1兲; then we treat 共Sec. V.B.2兲 the
varied rc in the range 1000– 10 K / ns, thus also consider- molecular-dynamics simulation results for noble-metal
ing rates that are rather close to the experimental ones, and quasi-noble-metal clusters at small sizes 共Baletto,
and analyzed a wide range of sizes up to about 4000 Mottet, and Ferrando, 2001b兲. These simulations were
atoms. In their simulations, the clusters solidified as run with very slow fluxes, in order to test the solid-state
icosahedra at small icosahedral magic numbers 共up to growth model against the energetics and thermodynam-
N = 147兲; small clusters with other sizes and large clusters ics results. In Sec. V.B.3, results on the growth of Ag and
could solidify as twinned fcc structures. No evidence of Au clusters of intermediate and large size are reported.
unfaulted truncated octahedron or decahedron clusters Finally Secs. V.B.4 and V.B.5 are devoted to the growth
was found. The fact that icosahedron clusters are domi- of aluminum and C60 clusters, respectively.
nant at small sizes but not at larger ones is in agreement
with the experimental findings of Reinhard, Hall, Ber- 1. Universal mechanism of the solid-state growth of
thoud, et al. 共1997兲, who found, however, some evidence large icosahedra
in favor of intermediate-size decahedron clusters. Before focusing on specific systems, we treat a very
Qi et al. 共2001兲 have studied the freezing of several Ni general mechanism which we believe to be responsible
clusters at sizes from 336 to 8007 atoms by molecular- for the observation of large metastable icosahedra in
dynamics simulations within the embedded-atom poten- several systems 共see Ag, Au, and C60 below兲. This
tial. In their simulations, they performed heating and mechanism consists of two steps. The first is the solid-
cooling cycles at very fast rates 共4000 K / ns兲: starting solid transformation of a decahedron into an icosahe-
from a solid cluster they increased T to melt the struc- dron, the second the shell-by-shell growth of the icosa-
ture, and then cooled it down again. As a result, they hedron to reach large sizes. This is a very peculiar
found icosahedron clusters after freezing for N ⬍ 500, example of structural transformation, because it has a
and fcc clusters for larger sizes. All these studies confirm clear kinetic origin, occurs only through solid states, and
the tendency to form icosahedral clusters only after does not depend on the interparticle potential, but is
freezing small liquid droplets, in agreement with the pre- essentially due to geometric reasons. First of all, we note
vious cases, whereas at larger sizes other structures are that this transformation is possible because a decahe-
dominant. dron is a fragment of a larger icosahedron, so that by
The first study of the solidification of Pb clusters was proper addition of atoms, a decahedron can grow to-
made by Lim et al. 共1994兲; they considered a single clus- wards an icosahedron. However, this is only a possibility,
ter of about 8000 atoms, obtaining an icosahedron after and a further ingredient is necessary to render this trans-
fast quenching, in disagreement with their own energetic formation kinetically favored. This key ingredient is the
calculations 共Lim et al., 1992兲, which predicted that the fact that on the 共111兲-like facets of a decahedron are
cuboctahedron would be lower in energy than the icosa- found a larger number of stable adsorption sites of hcp
hedron. This topic has been recently analyzed in a more stacking than of fcc stacking 共Baletto and Ferrando,
systematic way by Hendy and Hall 共2002兲, who 2001兲.
quenched a large variety of liquid Pb droplets in the size Islands of fcc stacking preserve the decahedral rear-
range between 600 and 6000 atoms. In their simulations, rangement in columns, around the fivefold symmetry
the liquid droplet was quenched suddenly below the axis, while hcp islands break the decahedral symmetry
cluster melting point, and then equilibrated for 10 ns. 共see Fig. 32兲 and can lead to the transformation towards
Up to 4000 atoms, they found mostly icosahedra with a a larger icosahedron. A decahedron plus a hcp island is a
reconstructed surface, showing that they are lower in part of a larger icosahedron. The displacement of a hcp
energy than any other known structure in the same size island to the fcc stacking costs a considerable amount of
range. The freezing of clusters of about 6000 atoms pro- energy, which increases with the stickiness of the poten-
duced faulted fcc structures. The production of icosahe- tial. While the growth of islands of fcc stacking leads
dron clusters agrees with the experimental observation simply to a larger decahedron 关this is the umbrella
of icosahedron particles of 3 – 6 nm diameter by Hyslop growth model 共Martin et al., 1991b兲兴, the nucleation of
et al. 共2001兲 in inert-gas aggregation experiments. hcp islands can be the starting point for transforming a

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


410 F. Baletto and R. Ferrando: Structural properties of nanoclusters

FIG. 32. A 146-atom decahedron plus an hcp island above a


cap. The black dots indicate fcc adsorption sites, the gray dots
the hcp sites. The fcc sites on the edges between facets are not FIG. 34. Growth simulations of small metallic clusters. ⌬ as a
stable for single adatoms. Compared to adsorption on icosahe- function of cluster size at different temperatures: low tempera-
dral clusters, the fcc and hcp sites correspond to Mackay and tures 共200 K for Au and Ag, 300 K for Ni and Pt, 400 K for Cu
anti-Mackay sites, respectively. and Pd兲 on the left and high temperatures 共400 K for Au and
Ag, 600 K for Ni, 700 K for Pd, Pt, and Cu兲. Minima in ⌬
single out the most stable structures.
decahedron into a larger icosahedron. The island can
be either one or two layers thick, leading to trans-
formations to icosahedra of different sizes, through dif- which can then grow further in a shell-by-shell mode
ferent rearrangement pathways 共Baletto, Mottet, and preserving their symmetry.
Ferrando, 2001a兲. The size at which a solid-state The above considerations lead to a general observa-
decahedron→icosahedron transformation can take place tion about the possibility of building up noncrystalline
depends strongly on temperature. A larger starting 共especially icosahedral兲 structures in nanoclusters. In
decahedron requires a higher temperature. fact, even though a sticky potential disfavors these struc-
While a growing decahedron can transform into an tures, at the same time it enhances kinetic trapping ef-
icosahedron, there is no natural growth sequence from fects, because diffusion barriers are high and rearrange-
icosahedron to decahedron structures, because an icosa- ments involving many particles become very difficult
hedron is not a fragment of a larger decahedron. To 共Wales, 1994b; Baletto, Doye, et al., 2003兲. These kinetic
transform a growing icosahedron into a decahedron, a effects very likely lead to the growth of icosahedra 共see
complete rearrangement of the cluster is necessary. This Fig. 33兲.
rearrangement becomes likely only at high tempera-
tures, close to the melting point, as proved by the ex- 2. Growth of small noble-metal and transition-metal
periments of Koga et al. 共2004兲. Therefore we can well clusters
understand that it is rather common to find wide tem-
perature ranges in which kinetic trapping into icosahe- Here we analyze simulations of the growth of small
dral structures dominates the growth sequence. The Cu, Ag, Au, Ni, Pd, and Pt clusters modeled by the Ro-
sato et al. 共1989兲 potential, in order to single out quali-
solid-state decahedron→ icosahedron transformation
tative differences in the behavior of these elements. To
thus allows us to explain several experimental results:
this purpose, we show in Fig. 34 the quantity ⌬ 关Eq. 共5兲兴
from small decahedra, larger icosahedra are grown,
as a function of the cluster size for growth simulations at
different temperatures, depending on the metals, with
fixed ␶dep = 98 ns. In this case, ⌬ is defined by putting into
Eq. 共5兲 the temperature-dependent total energy E of the
cluster, averaged at each size over several snapshots, in-
stead of the zero-temperature binding energy Eb. Stable
structures are singled out by minima in ⌬. Finally we
focus on the growth of Ag and Cu clusters at N = 38, to
show that in this case kinetic trapping effects are over-
FIG. 33. The solid-state decahedron→ icosahedron transfor- whelming. In fact, the global-minimum structure, which
mation obtained in growth simulation of clusters of C60 mol- is a truncated octahedron according to this energetic
ecules. The same kind of transformation is also found in simu- model, is grown with a non-negligible probability only in
lations of Ag cluster growth. Ag and C60 present completely a narrow temperature range for a given flux.
different interactions, soft and sticky, respectively. Adapted Growth sequences for Cu and Ni clusters are clearly
from Baletto, 2003. dominated by icosahedral or polyicosahedral structures,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 411

FIG. 35. 共Color in online edition兲 Ag38 growth simulation re- FIG. 36. 共Color in online edition兲 Cu38 growth simulation re-
sults. In the top row, snapshots from a simulation at T sults. Symbols as in Fig. 35. In the top row, snapshots from a
= 200 K are shown at different sizes. The structures of the Cu simulation at T = 400 K are reported. Again, the snapshots
snapshots pertain to the global minima up to N = 33; then reproduce the global minima up to N = 33. On the contrary, the
growth is kinetically trapped into defected decahedron struc- growth structure at N = 38 is not the global minimum but a Dh,
tures, which are neither energetically nor entropically favored. which is entropically favored at this temperature. Around
The graph shows the structural frequencies fTO, fDh, fLS of 200 K, there is a strong kinetic trapping into Dh structures
different structures at N = 38 and at different growth tempera- formed at smaller sizes. At intermediate temperatures the TO
tures: open stars, truncated octahedra; solid stars, decahedra; structure is preferentially grown; at high T, Dh structures again
triangles, low-symmetry structures. The lines are only guides to prevail due to entropic effects. Adapted from Baletto et al.,
the eyes. At low temperatures, strong kinetic trapping into 2004
decahedral structures occurs. This is followed by an interme-
diate temperature regime where the three structural motifs are
essentially equally likely. At high temperatures, entropic ef- interesting behavior takes place around for Pt55, as an-
fects favor decahedral and low-sym-metry structures at the ex- ticipated in Sec. V.A.4, with the probable growth of ro-
pense of the truncated octahedron. Adapted from Baletto et sette structures, either because of kinetic trapping in in-
al., 2004. complete icosahedral structures formed at lower sizes,
or because of the onset of entropic effects at high tem-
peratures.
as can be seen in Fig. 34. The magic sizes are 13, 19, 25, A complex interplay of thermodynamic and kinetic
40, 43, 46, 49, and 55, with 13, 19, and 55 being of special effects takes place in the growth simulations of Ag38 and
stability. The sequences are well reproducible over a Cu38 共Baletto, Rapallo, et al., 2004兲. At this size, the glo-
wide range of growth temperatures. The main difference bal minimum is a truncated octahedron, which is in com-
between Ni and Cu is that in Ni it is easier to obtain petition with a series of defected decahedron structures
truncated octahedra of 38 atoms, even at relatively low 共mostly based on decahedra of 23 atoms covered by is-
temperatures 共400 K兲. In this case, growth continues lands兲, and in the case of Ag, with a low-symmetry struc-
with the cluster trapped in non icosahedral structures. ture that is neither a defected decahedron nor a close-
Growth sequences for Ag show the same kind of struc- packed cluster. In the case of copper, these decahedron
tures, but they are reproducible over a narrower tem- structures become favored over the truncated octahedra
perature range. In contrast, the growth sequences of Au at temperatures around 350 K, because of entropic ef-
reveal completely different magic sizes. Icosahedral fects. Moreover, for sizes just below N = 38, the global
structures are completely unfavored. The minima at N minima are defected decahedra in both Ag and Cu. For
= 13 and 19 are absent, replaced by peaks at N = 16 and these reasons, the truncated octahedron is not likely to
22. At N = 30 we find a decahedral minimum, which sur- be grown in the simulations. Performing several simula-
vives up to T ⬃ 300 K. At higher temperatures the clus- tions at T = 200, 250, 300, 350, and 400 K, and observing
ters oscillate among different structures over a time the structures grown at N = 38, one obtains the results
scale shorter than ␶dep so that they can be considered reported in Figs. 35 and 36. At 200– 250 K, kinetic trap-
either melted or quasimelted. There is no evidence of ping into decahedron structures is the fate of essentially
entropic effects favoring the transition to icosahedral all simulations, so that the magic structure never coin-
structures. Finally, Pt and Pd show an intermediate be- cides with the global minimum, which is also the free-
havior between Cu, Ni, and Ag on the one side, and Au energy minimum in this temperature range. This kinetic
on the other. For Pd, there is a minimum at N = 13, but trapping is due to the fact that the global minima in the
there is no evidence of any other magic size up to the range 30–37 atoms are decahedral 共with the single ex-
Ih55, which is likely to be grown. In Pt, minima are found ception of Cu37兲. In the interval 300– 350 K, kinetic trap-
at N = 13 共as in Cu兲 and 22 共as in Au兲. At high T, it is ping becomes less effective, even if it is still somewhat
possible to grow truncated octahedra of 38 atoms. Very present. In this range, the truncated structures are still

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


412 F. Baletto and R. Ferrando: Structural properties of nanoclusters

quite likely at equilibrium, and they are indeed ob-


served, together with decahedra and low-symmetry
structures 共the latter in Ag only兲. Above 350 K, again
decahedral structures prevail in Cu38, as at low tempera-
ture. However, this is not due to kinetic trapping, but to
an entropic effect, because at these temperatures deca-
hedral structures are the most likely at equilibrium.

3. Growth of intermediate- and large-size Ag and Au


clusters
The experiments of production of Ag clusters in inert-
gas aggregation sources 共Reinhard, Hall, Ugarte, and
Monot, 1997兲 show that the abundances of small clusters
共2 nm diameter兲 are not dominated by icosahedral struc- FIG. 37. 共Color in online edition兲 Snapshots of silver cluster
tures, while icosahedra are the most abundant structures growth at ␶dep = 7 ns from three simulations at three different
for much larger sizes. Here we discuss how these results temperatures: left column, 400 K, middle column, 500 K, right
column, 600 K. In each column, the snapshots are taken at
can be explained in the framework of a solid-state
sizes of 55, 105, and 147 atoms from top to bottom. At 75
growth model 共Baletto, Mottet, and Ferrando, 2000b,
decahedra are preferentially grown up to 500 K. Icosahedral
2001a兲.
structures of 147 atoms are obtained in the simulations at 400
Let us consider first the sizes around 150 atoms, that and 600 K, and a decahedron at 500 K. At 55 atoms, we obtain
is, diameters of about 2 nm. In this range, the molecular- an icosahedron at 400 and 500 K 共the structure at 600 K is
dynamics growth simulations show that a reentrant mor- rapidly fluctuating兲.
phology transition takes place. This means that, at a
given deposition flux, there is an intermediate tempera-
ture window 共400– 500 K for fluxes of experimental in- The reentrant morphology transition is due to kinetic
terest兲 in which decahedral structures are preferentially trapping and not to entropic effects. In fact, on the basis
grown, while at low and high temperatures icosahedra of the considerations reported in Sec. IV.A, from Eq.
are obtained. 共25兲 it follows that, for N = 147, pDh / pIh monotonically
In fact, at intermediate temperatures, Ag clusters are increases with T.
able to optimize their shape up to N ⬃ 100 关the best Let us now consider a growth sequence leading to the
structure is the 共2,3,2兲 decahedron at 101 atoms兴 and formation of small decahedra and of large metastable
then remain trapped in decahedral structures. This leads icosahedra, in agreement with the results of the experi-
to the formation of a metastable decahedron cluster ments by Reinhard, Hall, Ugarte, and Monot 共1997兲.
around N = 150 关the 共3,2,2兲 decahedron cluster at 146 at- This sequence is obtained for 400⬍ T ⬍ 500 K and is es-
oms is the best in this range兴 instead of the lower-energy sentially a solid-state decahedron→ icosahedron trans-
icosahedra related to the N = 147 icosahedron. At higher formation followed by a shell-by-shell trapping in icosa-
temperatures 共T ⬎ 550 K兲, the clusters are liquid up to hedral structures, as reported in Fig. 38. In fact,
N ⯝ 130, and the final outcomes at N ⯝ 150 are icosahe- continuing the deposition on a metastable decahedron
dra. Finally, at low temperatures 共T ⬍ 450 K兲, the cluster
has sufficient kinetic energy to optimize its shape only
up to 75 atoms. Then it remains trapped in this structure
and evolves towards the 共6,1,1兲 decahedron at N = 100.
When this decahedron is almost completed, the addition
of further atoms causes the nucleation of hcp islands on
the facets and starts the formation of larger icosahedra.
This is an example of a solid-state decahedron
→ icosahedron transformation 共Baletto and Ferrando,
2001兲. Typical simulation results are summarized by the
snapshots shown in Fig. 37.
The metastable decahedra at N ⯝ 150 can have very
long lifetimes. Baletto, Mottet, and Ferrando 共2000b兲
calculated the lifetime ␶Dh of the 共3,2,2兲 decahedron in
the range 550艋 T 艋 650 K, finding an activated behavior
of the kind ␶Dh = ␶Dh0
exp关⌬E / 共kBT兲兴. After estimating
the prefactor ␶Dh and the activation barrier ⌬E, they
0
FIG. 38. Growth sequences of intermediate-size Ag clusters at
were able to extrapolate a lifetime of several millisec- different T: from left to right, 400, 450, and 600 K. In the top
onds at 450 K. This should indicate that this structures is row, N ⯝ 150, in the middle row N ⯝ 200, and in the bottom row
likely to survive on the experimental time scale. N ⯝ 300.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 413

FIG. 39. High-resolution electron microscopy image of a large


icosahedral Au cluster grown in an inert-gas aggregation
source and then deposited on an inert substrate. Adapted from
Koga and Sugawara, 2003.

of ⬃150 atoms, an island of hcp stacking is nucleated.


This island can be either one or two layers thick, leading
to the formation of icosahedra with N = 309 or 561, re- FIG. 40. Experimental population distributions of icosahedral,
spectively 共Baletto, Mottet, and Ferrando 2001a兲. For decahedral, and fcc Au clusters grown in an inert-gas aggrega-
T ⬎ 600 K, icosahedral structures are never obtained for tion experiment. Adapted from Koga and Sugawara, 2003.
N ⬎ 180. The deposition of a few tens of atoms above
N = 147 causes a sharp transition towards decahedra of icosahedra to a kinetic trapping effect 共Baletto, Mottet,
192 atoms. This grows passing through different decahe- and Ferrando, 2000a, 2001a兲, caused by the shell-by-
dral structures, towards one with N = 318 by the nucle- shell growth of small icosahedral clusters. At the mo-
ation of islands of fcc stacking. For T ⬎ 650 K, fcc struc- ment, simulations on Au cluster growth are not available
tures are preferentially grown. This happens because to rule out the possibility of a liquid-state growth pro-
around N = 201 and N = 314, decahedral and fcc clusters cess, but the formation of such large icosahedra by the
are in close competition from the energetic point of freezing of liquid droplets seems unlikely in the light of
view. Because of that, when temperature is high enough the results reported in Sec. V.A.2 and by Baletto, Mot-
and the energy differences become less important, the tet, and Ferrando 共2002兲.
growing cluster can pass through different faulted fcc
and asymmetric decahedral structures, growing finally as 4. Aluminum clusters
a fcc structure. At low temperatures, around 400 K,
icosahedra are preferentially grown around N = 150, and The formation of Al clusters was studied experimen-
these icosahedra continue growing in a shell-by-shell tally by different groups 共Lermé et al., 1992; Martin et
mode. al., 1992兲. The mass spectra presented regular oscilla-
The growth of Au clusters has also been investigated, tions, whose maxima were equally spaced by a quantity
and there is evidence for kinetic trapping effects of the ␦N1/3 when the cluster abundance was plotted as a func-
same kind as those found in Ag. In a recent experiment, tion of N1/3 共see Fig. 41兲. The numerical value of the
Koga and Sugawara 共2003兲 generated and analyzed a spacing turned out to be ␦N1/3 ⯝ 0.22. Valkealahti et al.
huge population of Au clusters with diameters in the 共1995兲 nicely explained these features as being of geo-
range 3 – 18 nm. The clusters were produced in an inert- metric origin within the octahedral growth model. In this
gas aggregation source with carrier helium gas, then de- model, each maximum corresponds to the addition of a
posited on an amorphous carbon substrate and analyzed single close-packed layer on one of the eight facets of
by high-resolution electron microscopy. Some of these the octahedron. Let us consider an octahedron with n1
clusters are shown in Figs. 2 and 39. Analysis of the close-packed layers in the direction of one of its eight
cluster population revealed a striking feature: for all facets. This octahedron has Noct共n1兲 atoms. When we
sizes, most of the clusters were icosahedral 共see Fig. 40兲,
with a proportion of about 90% at small sizes, slowly
decreasing to 60–70 % at large sizes. The remaining
clusters were mostly of decahedral symmetry, while a
very few fcc clusters were observed. This result is in dis-
agreement both with the energetic calculations 共Cleve-
land et al., 1997; Baletto, Ferrando, et al., 2002兲 and with
the experimental observation of Patil et al. 共1993兲 indi-
cating very low crossover sizes from the icosahedron to
the other motifs. Patil et al. 共1993兲 observed Au clusters
after melting and very slow refreezing; they were thus
very likely observing equilibrated clusters. Koga and FIG. 41. Experimental mass spectrum of aluminum clusters.
Sugawara 共2003兲 attributed the observation of large Adapted from Valkealahti et al., 1995.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


414 F. Baletto and R. Ferrando: Structural properties of nanoclusters

add a further layer above one facet, the number of at-


oms becomes Noct共n1 + 1兲. Valkealahti et al. 共1995兲 dem-
onstrated that

␦N1/3 = 关Noct共n1 + 1兲 − Noct共n1兲兴1/3 = 冉冊


1 3
6 2
2/3
共45兲

in very good agreement with the experiments.


In a subsequent molecular-dynamics study,
Valkealahti and Manninen 共1998兲 analyzed the possibil-
ity of growing octahedral clusters starting from trun-
cated octahedra. As a first step towards understanding
the growth mechanism, they calculated diffusion barriers
for adatoms and dimers, finding very interesting mecha-
nisms such as chain diffusion, which allows direct mass
transport between two nonadjacent 共111兲 facets through
an intermediate 共100兲 facet. Then, in their growth simu-
lations, they took a truncated octahedron of 586 atoms
as a starting seed, and deposited atoms one by one at FIG. 42. Growth sequences of C60 clusters for ␶dep = 100 ns and
constant temperature 共T = 400 K兲 with a rather fast different temperatures. From top to bottom snapshots at N
deposition rate 共␶dep = 0.5 ns兲 up to a total of 201 depos- = 13, 25, 38, 45, and 55 particles are shown. Adapted from
ited atoms. The results demonstrated the possibility of Baletto, Doye, and Ferrando, 2002.
transforming a truncated octahedron into an octahedron
by the following mechanism. When depositing on the
truncated octahedron, atoms are very likely to fall on usually resembles the global minimum, although this is
共111兲 facets, where they can diffuse fast. When an ada- not necessarily the case in between these sizes. Only at
tom reaches the border with a 共100兲 facet, it can ex- very low T 共⬍300 K兲 does the structure remain trapped
change easily with an edge atom, which is then trapped in the icosahedral shape after 13 molecules. The 25-
on the 共100兲 facet. The reverse process 关exchange from a molecule structure is decahedral but with an island on
共100兲 to a 共111兲 facet兴 is very difficult because the ad- the bottom 共111兲 faces of the cluster. This structure plays
sorption energy is much more favorable on 共100兲 than on a key role, because growth is dominated by kinetic trap-
共111兲 facets. The accumulation of atoms on 共100兲 facets ping effects beyond this size, since a complete solid-state
leads naturally to the formation of an octahedron expos- decahedron→ icosahedron transformation takes place.
ing only 共111兲 facets. This mechanism was confirmed Furthermore, the structure is a fragment of the icosahe-
later by molecular-dynamics growth simulations of Ag dron with N = 55, and continued growth around the bot-
and Au clusters 共Baletto, Mottet, and Ferrando, 2000a兲 tom apex of the decahedron provides a pathway to this
that were performed starting from a truncated octahe- structure, with its apex ending up at the center of the
dron of N = 201 but with a much slower deposition rate. resulting icosahedron 共Baletto et al., 2001a兲.
The growth simulations of truncated octahedron clusters This growth pattern bears no resemblance to the se-
have revealed the formation of stacking faults quence of global minima, which develops through either
共Valkealahti and Manninen, 1998; Baletto, Mottet, and close-packed or decahedral structures. In particular, the
Ferrando, 2000a兲. Manninen and Manninen 共2002兲 and 共C60兲38 global minimum is a truncated octahedron, while
Manninen et al. 共2003兲 showed that clusters with stacking the growth simulations always give structures with five-
faults are also obtained in global optimization on an fcc fold symmetries. The same happens at N = 45, and finally
lattice with all possible 共111兲 stacking faults allowed, in a Mackay icosahedron results at N = 55. If the growth is
the case of different model potentials. continued for N ⬎ 55, a well-ordered anti-Mackay over-
layer 共Martin, 1996兲 develops on the surface of the icosa-
5. Clusters of C60 molecules hedron, in agreement with the sequence of magic num-
bers observed experimentally for 55⬍ N ⬍ 100 共Branz et
In the case of C60 clusters, growth takes place in the al., 2000, 2002兲. Decahedral and fcc structures are only
solid state. Baletto, Doye, and Ferrando 共2002兲 simu- obtained in a few cases at rather high growth tempera-
lated their growth with ␶dep between 100 and 200 ns and tures 共525– 550 K兲, as found by Baletto, Doye, and Fer-
T ⬍ 600 K, modeling the interactions by the Girifalco rando 共2002兲. If one uses the Pacheco and Prates-
共1992兲 potential. Some results are shown in Fig. 42, Ramalho 共1997兲 potential, the growth of nonicosahedral
where snapshots from typical growth simulations at dif- structures becomes more likely, but icosahedra are still
ferent T and ␶dep = 100 ns are reported at some signifi- obtained in most cases.
cant sizes. All the sequences T ⬍ 525 K develop along Let us discuss how these simulation results compare
the same line and lead to the formation of a well- to the experiments by Branz et al. 共2000, 2002兲 discussed
ordered icosahedron of 55 atoms. At N = 13 the structure in Sec. III.E.6. The experimental data concerning the
is always icosahedral, in agreement with the global opti- formation of icosahedral clusters are easily explained by
mization results. Similarly, around N = 25 the structure the outcome of the simulations by Baletto, Doye, and

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 415

Ferrando 共2002兲: further annealing of the growth struc-


tures indicates that it is indeed quite easy to eliminate
the less bound molecules from the surfaces of the clus-
ters, thus leaving only icosahedral magic clusters.
On the other hand, the experimental results obtained
after annealing at higher T 共close to 600 K兲 are more
difficult to explain: close-packed or decahedral clusters
were mostly observed in the experiments 共Branz et al., FIG. 43. Molecular-dynamics simulation of the coalescence of
2000, 2002兲. In principle, these clusters could be formed two 565-atom lead icosahedra initially at 300 K. The sequence
of snapshots at 3.75-ps intervals shows the early growth of the
in two ways:
neck after the initial contact. Adapted from Hendy et al., 2003.
共i兲 the high-T annealing allows some of the icosahe-
dral clusters obtained at lower temperatures to re- perature constant during the coalescence process, and
arrange in such a way that they are able to reach therefore their results are probably better applied to
the structure with minimum free energy; supported clusters, which can exchange energy with the
共ii兲 the high-T annealing causes the evaporation of al- substrate at a rapid rate. They considered the coales-
most all icosahedral clusters, and only the few 共al- cence between two solid clusters, a liquid and a solid
ready existing兲 nonicosahedral clusters, which are cluster, and two liquid clusters. The coalescence of two
more stable, survive and are observed. liquid clusters takes place rapidly. A single spherical
cluster is formed by the deformation of the two clusters
Moelcular-dynamics simulations of the annealing of
in such a way as to optimize the contact surface, without
icosahedral clusters of different sizes at high tempera-
interdiffusion. Later on, interdiffusion takes place, but
tures 共650– 750 K兲 have never produced a rearrange-
the spherical shape is reached over much shorter time
ment of the cluster to either decahedral or close-packed
scales by a collective rearrangement phenomenon. The
structures 共Baletto, Doye, and Ferrando, 2002兲. Instead,
coalescence of a solid and a liquid cluster proceeds in
dissolution by desorption of molecules 共with some local
two stages. At first the contact surface is maximized rap-
rearrangement within the structure兲 occurs. These re-
idly, on the same time scale as the coalescence of liquid
sults are consistent with mechanism 共ii兲 but we cannot
clusters. At this stage the cluster is far from being
draw a firm conclusion. Some structural transformations
spherical but has a faceted ovoidal shape. After that the
from icosahedral to decahedral structures have been ob-
spherical shape is reached by a slow process driven
served in extremely long 共0.1 ms兲 simulations by Branz
mainly by surface diffusion. The rapid changes seen at
共2001兲 and Branz et al. 共2002兲 at T ⬃ 700 K, where, how- short times are due to elastic and plastic deformations.
ever, the number of atoms was forced to remain constant At long times the presence of facets slows down the
in a closed simulation box, and this forbade cluster dis- diffusion 共Baletto and Ferrando, 2001兲, so that coales-
solution, which is, at this temperature, dominant. Unfor- cence times are much longer 共Mazzone, 2000兲 than pre-
tunately, at the moment it is not possible to simulate dicted by the macroscopic theory of sintering 共Nichols
dissolution at 600 K, because the time scales involved
are too long.

C. Coalescence of nanoclusters

The coalescence of supported clusters is of great im-


portance in the field of surface nanostructuring. This
topic has been the subject of an excellent review paper
by Jensen 共1999兲. Here we intend to focus on the coales-
cence of free clusters, a mechanism that can be impor-
tant for free-cluster formation, especially in the late
stages of the growth process, when already formed clus-
ters can collide and join together. Some experimental
evidence in favor of this mechanism may be inferred
from the results of Patil et al. 共1993兲, who found the
FIG. 44. Evolution of the temperature and of the aspect ratio
formation of polycrystalline Au clusters in an inert-gas during the molecular-dynamics simulation of the coalescence
aggregation source and attributed them to the encounter of two 565-atom lead icosahedra at the initial temperature of
and coalescence of different smaller units. However, we 300 K. At the point of first contact between clusters 共approxi-
remark that polycrystalline clusters may also result as mately 30 ps after the beginning of the simulation兲 the tem-
the outcome of the freezing of single liquid droplets perature rises sharply due to the release of surface energy. The
共Valkealahti and Manninen, 1997兲. inset shows the final cluster structure, which is almost spheri-
Lewis et al. 共1997兲 studied the coalescence of free Au cal. The spherical shape is quickly reached, since the aspect
clusters by molecular-dynamics simulations. They ratio is very close to one after 1 ns. Adapted from Hendy et al.,
coupled the clusters to a thermostat, to keep the tem- 2003.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


416 F. Baletto and R. Ferrando: Structural properties of nanoclusters

and Mullins, 1965; Jensen, 1999兲 via surface diffusion. structures, such as the icosahedra, and favoring the for-
Finally, the coalescence of two solid clusters 共simulated mation of low-symmetry structures. Finally, bond direc-
also by Zhu, 1996 for Cu clusters兲 is a complex phenom- tionality is crucial in semiconductor clusters and in some
enon, which takes place on a slow time scale and may metals, too.
involve either the formation of a single domain cluster Several examples have shown that the interplay of en-
or of complicated structures presenting grains. This de- ergetics, thermodynamics, and kinetics is crucial, and
pends on size and structure of the initial clusters, at vari- that a satisfactory explanation of the experimental out-
ance with the previous cases. Moreover, the nucleation comes is very often impossible on the basis of energetic
of a critical island on the surface can govern the rear- considerations alone. This can have deep consequences
rangement kinetics 共Combe et al., 2000兲. for the very important issue of controlling the shapes of
Recently Hendy et al. 共2003兲 simulated the coales- the produced nanoclusters, which is of great technologi-
cence of free lead clusters in the microcanonical en- cal importance. Given that the interplay of the three
semble, without coupling the clusters to a thermostat. factors is crucial, each one is extremely interesting by
This seems to be the most appropriate method when itself and poses stimulating theoretical challenges.
dealing with free clusters produced in inert-gas aggrega- The development of reliable methods for modeling
tion sources. In fact, when two clusters come into con- the energetics of nanoclusters is a rapidly developing
tact, many new bonds form, causing a considerable re- field, both from the point of view of ab initio calcula-
lease of surface energy. This can cause a noticeable tions, well suited for precise calculations on small sys-
temperature increase 共more than 100 K for clusters of tems, and from the point of view of semiempirical mod-
N ⯝ 500兲. Since cooling rates are less than 1 K / ns 关see eling, which is needed to treat larger systems and to
Eq. 共44兲兴, and the initial stage of coalescence develops in simulate long time scales. In cluster science, these ap-
a few ns 共see Figs. 43 and 44兲, the inert gas is not likely proaches are complementary and are both necessary.
to affect the process strongly by taking energy away. The Since the best structure of a cluster of given size and
temperature increase makes the coalescence process composition is generally unknown, the semiempirical
much faster, by enhancing surface diffusion. Hendy et al. modeling, which has a much lower computational cost, is
共2003兲 considered the coalescence of two solid surface- the starting point for selecting good candidates for sub-
reconstructed icosahedral clusters 共Hendy and Hall, sequent ab initio local structural optimization. We have
2001兲 of 565 atoms. They found that, depending on the seen that there are already several examples in the lit-
initial temperature and the sizes of the two clusters, the erature showing that this approach is extremely fruitful,
final aggregate could be either solid or liquid. When the and now the challenge is to extend it to large sizes and
final temperature Tf of the resulting cluster of 1130 at- complex systems 共for example, nanoalloys兲. The search
oms was below Tm of the Ih565, the coalescence took for good structural candidates is performed by global
place through solid states. On the other hand, when Tf optimization methods. The literature on these methods
was above Tm of the resulting cluster of 1130 atoms, the and their applications to clusters has exploded in the last
final cluster was liquid. Finally, when Tf was in between few years; several different procedures have been pro-
the two melting temperatures, the aggregate melted at posed, with important progress towards the understand-
first and solidified again at a later stage. ing of the requirements for building up efficient global
optimization algorithms. The main point here is that the
VI. CONCLUSIONS AND PERSPECTIVES most efficient algorithms work after transforming the
original PES to a multidimensional staircase. In the field
In this review we have tried to present an overview of of global optimization, there is still the need to develop
the physical properties of nanoclusters, with the aim of algorithms for complicated systems and large sizes. The
showing the interplay of energetic, thermodynamic, and latter task is, however, limited by the intrinsic NP-hard
kinetic factors in building up the structures that are pro- nature of the global optimization problem itself 共see
duced in the experiments or observed in the simulations. footnote 3 and Wille and Vennik, 1985兲. Thus there is no
We hope that we have demonstrated that the physics of hope in optimizing very large clusters, although good
nanoclusters can be understood only by taking into ac- putative global minima are found by the present algo-
count all three factors and relating them to the features rithms by exploring only a very small fraction of the
of the interparticle potential. Among these features, the huge number of local minima. The reason why these
potential range is always a guideline for understanding algorithms work lies in the features of the PES, which
the qualitative features of structural properties and may present funnels in which optimization is fast, and in
transformations. For example, we have seen that soft the favorable transformation of the thermodynamics of
interparticle interactions admit noncrystalline clusters as the system when working on the staircase PES.
minimum-energy isomers. In contrast, for sticky inter- The thermodynamics of nanoclusters is a very active
particle potentials, crystalline structures are energeti- research field, with several problems under debate. For
cally favored, but at the same time kinetic trapping phe- example, the relation between elementary interactions
nomena are enhanced. The latter often cause the growth and the type of phase changes in clusters is still to be
of noncrystalline structures in the actual experiments. In understood to a large extent for realistic model poten-
metallic systems, the bond-order/bond-length correla- tials. Moreover, the effect of chemical composition on
tion plays a crucial role in destabilizing some cluster melting and on structural transitions has been the sub-

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 417

ject of very few studies. Finally, the approach of the bulk Aziz, R. A., F. R. W. McCourt, and C. C. K. Wong, 1987, Mol.
limit is not yet well understood. Each of these problems Phys. 61, 1487.
poses complex theoretical and computational chal- Bacelo, D. E., R. C. Binning, and Y. Ishikawa, 1999, J. Phys.
lenges. Chem. A 103, 4631.
The study of the growth kinetics of nanoclusters is a Bachels, T., H. J. Güntherodt, and R. Schäfer, 2000, Phys. Rev.
field that is just starting to develop. Here, a systematic Lett. 85, 1250.
study of small systems by the available methods is still to Bachels, T., and R. Schäfer, 2000, Chem. Phys. Lett. 324, 365.
be performed. Moreover, new methods should be devel- Baguenard, B., M. Pellarin, J. Lermé, J. L. Vialle, and M.
oped, first of all to extend the size of the simulated sys- Broyer, 1994, J. Chem. Phys. 100, 754.
tems and the time scale of the simulations, and then to Bailey, M. S., N. T. Wilson, C. Roberts, and R. L. Johnston,
treat the formation of clusters in complex environments, 2003, Eur. Phys. J. D 25, 41.
Baletto, F., 2003, Ph.D. thesis 共Universita di Genova兲.
such as formation in liquid solutions or on surfaces, in-
Baletto, F., J. P. K. Doye, and R. Ferrando, 2002, Phys. Rev.
teraction with passivating agents and adsorbed mol-
Lett. 88, 075503.
ecules, and so on.
Baletto, F., J. P. K. Doye, C. Mottet, and R. Ferrando, 2003,
In conclusion, we hope that our review article has Surf. Sci. 532, 898.
given at least some idea of the present development of Baletto, F., and R. Ferrando, 2001, Surf. Sci. 490, 361.
the fascinating and lively subject of cluster science. Baletto, F., R. Ferrando, A. Fortunelli, F. Montalenti, and C.
Mottet, 2002, J. Chem. Phys. 116, 3856.
ACKNOWLEDGMENTS Baletto, F., C. Mottet, and R. Ferrando, 2000a, Surf. Sci. 446,
31.
The authors are grateful to Jonathan Doye, Alessan- Baletto, F., C. Mottet, and R. Ferrando, 2000b, Phys. Rev. Lett.
dro Fortunelli, Ignacio Garzón, Hellmut Haberland, 84, 5544.
Roy Johnston, Andrea Levi, Christine Mottet, Arnaldo Baletto, F., C. Mottet, and R. Ferrando, 2001a, Phys. Rev. B
Rapallo, and Giulia Rossi for their help and suggestions. 63, 155408.
Baletto, F., C. Mottet, and R. Ferrando, 2001b, Eur. Phys. J. D
The authors acknowledge financial support from the
16, 25.
Italian MIUR under the project 2001021133. R.F. ac-
Baletto, F., C. Mottet, and R. Ferrando, 2002, Chem. Phys.
knowledges financial support from CNR for the project
Lett. 354, 82.
SSA-TMN in the framework of ESF EUROCORES- Baletto, F., C. Mottet, and R. Ferrando, 2003, Phys. Rev. Lett.
SONS. 90, 135504.
Baletto, F., A. Rapallo, G. Rossi, and R. Ferrando, 2004, Phys.
REFERENCES Rev. B 69, 235421.
Barranco, M., J. Navarro, and A. Poves, 1997, Phys. Rev. Lett.
Ahlrichs, R., and S. D. Elliott, 1999, Phys. Chem. Chem. Phys. 78, 4729.
1, 13. Barreteau, C., M. C. Desjonquères, and D. Spanjaard, 2000,
Ajayan, P. M., and L. D. Marks, 1988, Phys. Rev. Lett. 60, 585. Eur. Phys. J. D 11, 395.
Akola, J., and M. Manninen, 2001, Phys. Rev. B 63, 193410. Barreteau, C., R. Guirado-López, D. Spanjaard, M. C. Desjon-
Ala-Nissila, T., R. Ferrando, and S. C. Ying, 2002, Adv. Phys. quères, and A. M. Oleś, 2000, Phys. Rev. B 61, 7781.
51, 949. Bartell, L. S., 1992, J. Phys. Chem. 96, 108.
Alameddin, G., J. Hunter, D. Cameron, and M. M. Kappes, Bauschlicher, C. W., 1989, Chem. Phys. Lett. 156, 91.
1992, Chem. Phys. Lett. 192, 122. Bauschlicher, C. W., S. R. Langhoff, and H. Partridge, 1989, J.
Alivisatos, A. P., 2001, Sci. Am. 共Int. Ed.兲 285, 66. Chem. Phys. 91, 2412.
Alivisatos, A. P., K. P. Johnsson, X. G. Peng, T. E. Wilson, C. J. Bauschlicher, C. W., S. R. Langhoff, and H. Partridge, 1990, J.
Loweth, M. P. Bruchez, and P. G. Schultz, 1996, Nature Chem. Phys. 93, 8133.
共London兲 382, 609. Bauschlicher, C. W., S. R. Langhoff, and P. R. Taylor, 1988, J.
Alonso, J. A., 2000, Chem. Rev. 共Washington, D.C.兲 100, 637. Chem. Phys. 88, 1041.
Alvarez, M. M., J. T. Khoury, T. G. Schaaff, M. Shafigullin, I. Bazterra, V. E., O. Oña, M. C. Caputo, M. B. Ferraro, P. Fu-
Vezmar, and R. L. Whetten, 1997, Chem. Phys. Lett. 266, 91. entealba, and J. C. Facelli, 2004, Phys. Rev. A 69, 053202.
Amar, F. G., and R. S. Berry, 1986, J. Chem. Phys. 85, 5943. Beck, T. L., and R. S. Berry, 1988, J. Chem. Phys. 88, 3910.
Amara, P., D. Hsu, and J. E. Straub, 1993, J. Phys. Chem. 97, Becke, A. D., 1988, Phys. Rev. A 38, 3098.
6715. Becke, A. D., 1996, J. Chem. Phys. 98, 5648.
Anderson, J. B., and J. B. Fenn, 1985, Phys. Fluids 8, 780. Becker, O. M., and M. Karplus, 1997, J. Chem. Phys. 106, 1495.
Andriotis, A. N., and M. Menon, 2004, J. Chem. Phys. 120, 230. Berry, R. S., T. L. Beck, H. L. Davis, and J. Jellinek, 1988,
Aprà, E., F. Baletto, R. Ferrando, and A. Fortunelli, 2004, Evolution of Size Effects in Chemical Dynamics 共Wiley, New
Phys. Rev. Lett. 93, 065502. York兲, p. 75.
Ascencio, J. A., C. Gutiérrez-Wing, M. E. Espinosa, M. Marín, Berry, R. S., J. Jellinek, and G. Natanson, 1984, Phys. Rev. A
S. Tehuacanero, C. Zorrilla, and M. José-Yacamán, 1998, 30, 919.
Surf. Sci. 396, 349. Bilgram, J. H., 1987, Phys. Rep. 153, 1.
Ascencio, J. A., M. Pérez, and M. José-Yacamán, 2000, Surf. Binns, C., 2001, Surf. Sci. Rep. 44, 1.
Sci. 447, 73. Biswas, R., and D. R. Hamann, 1986, Phys. Rev. B 34, 895.
Ashcroft, N. W., and N. D. Mermin, 1976, Solid State Physics Bixon, M., and J. Jortner, 1989, J. Chem. Phys. 91, 1631.
共Saunders College, Orlando兲. Bjørnholm, S., J. Borggreen, O. Echt, K. Hansen, J. Pedersen,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


418 F. Baletto and R. Ferrando: Structural properties of nanoclusters

and H. D. Rasmussen, 1990, Phys. Rev. Lett. 65, 1627. Chacko, S., K. Joshi, D. G. Kanhere, and S. A. Blundell, 2004,
Bjørnholm, S., J. Borggreen, O. Echt, K. Hansen, J. Pedersen, Phys. Rev. Lett. 92, 135 506.
and H. D. Rasmussen, 1991, Z. Phys. D: At., Mol. Clusters 19, Cheng, H.-P., R. N. Barnett, and U. Landman, 1993, Phys. Rev.
47. B 48, 1820.
Blaisten-Barojas, E., I. L. Garzón, and M. Avalos-Borja, 1987, Cheng, L., W. Cai, and X. Shao, 2004, Chem. Phys. Lett. 389,
Phys. Rev. B 36, 8447. 309.
Blaisten-Barojas, E., and D. Levesque, 1986, Phys. Rev. B 34, Chin, S. A., and E. Krotscheck, 1992, Phys. Rev. B 45, 852.
3910. Chin, S. A., and E. Krotscheck, 1995, Phys. Rev. B 52, 10 405.
Bobadova-Parvanova, P., K. A. Jackson, S. Srinivas, M. Horoi, Christensen, O. B., and M. L. Cohen, 1993, Phys. Rev. B 47,
C. Köhler, and G. Seifert, 2002, J. Chem. Phys. 116, 3576. 13 643.
Bonacić-Koutecký, V., J. Burda, R. Mitrić, M. Ge, G. Chuang, F.-C., C. Z. Wang, S. Ögüt, J. R. Chelikowsky, and K.
Zampella, and P. Fantucci, 2002, J. Chem. Phys. 117, 3120. M. Ho, 2004, Phys. Rev. B 69, 165 408.
Bonacić-Koutecký, V., L. Cespiva, P. Fantucci, and J. Chushak, Y., and L. S. Bartell, 2001a, Eur. Phys. J. D 16, 43.
Koutecký, 1993, J. Chem. Phys. 98, 7981. Chushak, Y. G., and L. S. Bartell, 2001b, J. Phys. Chem. B 105,
Bonacić-Koutecký, V., L. Cespiva, P. Fantucci, J. Pittner, and J. 11 605.
Koutecký, 1994, J. Chem. Phys. 100, 490. Clemenger, K., 1985, Phys. Rev. B 32, 1359.
Bonacić-Koutecký, V., P. Fantucci, C. Fuchs, C. Gatti, J. Pitt- Cleveland, C. L., and U. Landman, 1991, J. Chem. Phys. 94,
ner, and S. Polezzo, 1993, Chem. Phys. Lett. 213, 522. 7376.
Bonacić-Koutecký, V., P. Fantucci, and J. Koutecký, 1988, Cleveland, C. L., and U. Landman, 1994, J. Phys. Chem. 98,
Phys. Rev. B 37, 4369. 6272.
Bonacić-Koutecký, V., P. Fantucci, and J. Koutecký, 1991, Cleveland, C. L., U. Landman, T. G. Schaaff, M. N. Shafigullin,
Chem. Rev. 共Washington, D.C.兲 91, 1035. P. W. Stephens, and R. L. Whetten, 1997, Phys. Rev. Lett. 79,
Borgia, I., B. Brunetti, I. Mariani, A. Sgamellotti, F. Cariati, P. 1873.
Fermo, M. Mellini, C. Viti, and G. Padeletti, 2002, Appl. Surf. Cleveland, C. L., W. D. Luedtke, and U. Landman, 1998, Phys.
Sci. 185, 206. Rev. Lett. 81, 2036.
Boustani, I., W. Pewestorf, P. Fantucci, V. Bonacić-Koutecký, Cleveland, C. L., W. D. Luedtke, and U. Landman, 1999, Phys.
and J. Koutecký, 1987, Phys. Rev. B 35, 9437. Rev. B 60, 5065.
Boyukata, M., Z. B. Güvenc, S. Ozcelik, P. Durmus, and J. Coleman, T., and D. Shalloway, 1994, J. Global Optim. 4, 171.
Jellinek, 2001, Int. J. Quantum Chem. 84, 208. Combe, N., P. Jensen, and A. Pimpinelli, 2000, Phys. Rev. Lett.
Brack, M., 1993, Rev. Mod. Phys. 65, 677. 85, 110.
Branz, W., 2001, Ph.D. thesis 共University of Stuttgart兲. Cottancin, E., J. Lermé, M. Gaudry, M. Pellarin, J.-L. Vialle,
Branz, W., N. Malinowski, A. Enders, and T. P. Martin, 2002, M. Broyer, B. Prével, M. Treilleux, and P. Mélinon, 2000,
Phys. Rev. B 66, 094107. Phys. Rev. B 62, 5179.
Branz, W., N. Malinowski, H. Schaber, and T. P. Martin, 2000, Cox, H., R. L. Johnston, and J. N. Murrell, 1999, J. Solid State
Chem. Phys. Lett. 328, 245. Chem. 145, 517.
Bravo-Pérez, G., I. L. Garzón, and O. Novaro, 1999a, J. Mol. Dai, D., and K. Balasubramanian, 1995, J. Chem. Phys. 103,
Struct.: THEOCHEM 493, 225. 648.
Bravo-Pérez, G., I. L. Garzón, and O. Novaro, 1999b, Chem. Darby, S., T. V. Mortimer-Jones, R. L. Johnston, and C. Rob-
Phys. Lett. 313, 655. erts, 2002, J. Chem. Phys. 116, 1536.
Breaux, G. A., R. C. Benirschke, T. Sugai, B. S. Kinnear, and Davis, H. L., J. Jellinek, and R. S. Berry, 1987, J. Chem. Phys.
M. F. Jarrold, 2003, Phys. Rev. Lett. 91, 215508. 86, 6456.
Briant, C. L., and J. J. Burton, 1975, J. Chem. Phys. 63, 2045. Daw, M. S., and M. I. Baskes, 1984, Phys. Rev. B 29, 6443.
Brühl, R., R. Guardiola, A. Kalinin, O. Kornilov, J. Navarro, Deaven, D. M., and K. M. Ho, 1995, Phys. Rev. Lett. 75, 288.
T. Savas, and J. P. Toennies, 2004, Phys. Rev. Lett. 92, 185301. Deaven, D. M., M. Tit, J. R. Morris, and K. M. Ho, 1996,
Buck, U., and I. Ettischer, 1994, J. Chem. Phys. 100, 6974. Chem. Phys. Lett. 256, 195.
Buffat, P., and J.-P. Borel, 1976, Phys. Rev. A 13, 2287. de Heer, W. A., 1993, Rev. Mod. Phys. 65, 611.
Buffat, P.-A., M. Flueli, R. Spycher, P. Stadelmann, and J.-P. Desjonqueres, M. C., and D. Spanjaard, 1998, Concepts in Sur-
Borel, 1991, Faraday Discuss. 92, 173. face Physics 共Springer, Berlin兲.
Caccamo, C., D. Costa, and A. Fucile, 1997, J. Chem. Phys. Dietz, T. G., M. A. Duncan, D. E. Powers, and R. E. Smalley,
106, 255. 1981, J. Chem. Phys. 74, 6511.
Callegari, C., K. K. Lehmann, R. Schmied, and G. Scoles, 2001, Doye, J. P. K., 2003, J. Chem. Phys. 119, 1136.
J. Chem. Phys. 115, 10 090. Doye, J. P. K., 2004, unpublished.
Calvo, F., and C. Guet, 2000, J. Chem. Phys. 113, 1315. Doye, J. P. K., and F. Calvo, 2001, Phys. Rev. Lett. 86, 3570.
Calvo, F., J. P. Neirotti, D. L. Freeman, and J. D. Doll, 2000, J. Doye, J. P. K., and F. Calvo, 2002, J. Chem. Phys. 116, 8307.
Chem. Phys. 112, 10 350. Doye, J. P. K., and F. Calvo, 2003, J. Chem. Phys. 119, 12 680.
Calvo, F., and F. Spiegelman, 2000, J. Chem. Phys. 112, 2888. Doye, J. P. K., A. Dullweber, and D. J. Wales, 1997, Chem.
Calvo, F., and F. Spiegelman, 2004, J. Chem. Phys. 120, 9684. Phys. Lett. 269, 408.
Calvo, F., F. Spiegelman, and M.-C. Heitz, 2003, J. Chem. Phys. Doye, J. P. K., and S. C. Hendy, 2003, Eur. Phys. J. D 22, 99.
118, 8739. Doye, J. P. K., and D. J. Wales, 1995, J. Chem. Phys. 112, 9659.
Car, R., and M. Parrinello, 1985, Phys. Rev. Lett. 55, 2471. Doye, J. P. K., and D. J. Wales, 1996a, Science 271, 484.
Castleman, A. W., and K. H. Bowen, 1996, J. Phys. Chem. 100, Doye, J. P. K., and D. J. Wales, 1996b, Chem. Phys. Lett. 262,
12 911. 167.
Ceperley, D. M., 1994, Rev. Mod. Phys. 67, 279. Doye, J. P. K., and D. J. Wales, 1998a, Phys. Rev. Lett. 80,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 419

1357. Fujima, N., and T. Yamaguchi, 1989, J. Phys. Soc. Jpn. 58, 1334.
Doye, J. P. K., and D. J. Wales, 1998b, New J. Chem. 22, 733. Fuke, K., K. Tsukamoto, F. Misaizu, and M. Sanekata, 1994, J.
Doye, J. P. K., and D. J. Wales, 2001, Phys. Rev. Lett. 86, 5719. Chem. Phys. 99, 7807.
Doye, J. P. K., D. J. Wales, and R. S. Berry, 1995, J. Chem. Ganteför, G., and W. Eberhardt, 1996, Phys. Rev. Lett. 76,
Phys. 103, 4234. 4975.
Doye, J. P. K., D. J. Wales, W. Branz, and F. Calvo, 2001, Phys. Ganteför, G., M. Gauss, K. H. Meiwes-Broer, and H. Lutz,
Rev. B 64, 235409. 1990, J. Chem. Soc., Faraday Trans. 86, 2483.
Doye, J. P. K., D. J. Wales, and M. A. Miller, 1998, J. Chem. Garcia-Rodeja, J., C. Rey, and L. J. Gallego, 1997, Phys. Rev.
Phys. 109, 8143. B 56, 6466.
Efremenko, I., 2001, J. Mol. Catal. A: Chem. 173, 19. Garcia-Rodeja, J., C. Rey, L. J. Gallego, and J. A. Alonso,
Efremenko, I., and M. Sheintuck, 2000, J. Mol. Catal. A: 1994, Phys. Rev. B 49, 8495.
Chem. 160, 445. Garzón, I. L., M. R. Beltrán, G. Gonzalez, I. Gutierrez-
Efremov, M. Y., F. Schiettekatte, M. Zhang, E. A. Olson, A. T. Gonzalez, K. Michaelian, J. A. Reyes-Nava, and J. I.
Kwan, R. S. Berry, and L. H. Allen, 2000, Phys. Rev. Lett. 85, Rodriguez-Hernandez, 2003, Eur. Phys. J. D 24, 105.
3560. Garzón, I. L., and J. Jellinek, 1991, Z. Phys. D: At., Mol. Clus-
Ehbrecht, M., and F. Huisken, 1999, Phys. Rev. B 59, 2975. ters 20, 235.
Ehbrecht, M., B. Kohn, F. Huisken, M. A. Laguna, and V. Garzón, I. L., and J. Jellinek, 1993, Z. Phys. D: At., Mol. Clus-
Paillard, 1997, Phys. Rev. B 56, 6958. ters 26, 316.
Ercolessi, F., W. Andreoni, and E. Tosatti, 1991, Phys. Rev. Garzón, I. L., I. G. Kaplan, R. Santamaria, and O. Novaro,
Lett. 66, 911. 1998, J. Chem. Phys. 109, 2176.
Ercolessi, F., M. Parrinello, and E. Tosatti, 1988, Philos. Mag. Garzón, I. L., K. Michaelian, M. R. Beltrán, A. Posada-
A 58, 213. Amarillas, P. Ordejón, E. Artacho, D. Sánchez-Portal, and J.
Erkoç, S., and T. Yilmaz, 1999, Physica E 共Amsterdam兲 5, 1. M. Soler, 1998, Phys. Rev. Lett. 81, 1600.
Ervin, K. M., J. Ho, and W. C. Lineberger, 1988, J. Chem. Garzón, I. L., K. Michaelian, M. R. Beltrán, A. Posada-
Phys. 89, 4514. Amarillas, P. Ordejón, E. Artacho, D. Sánchez-Portal, and J.
Etters, R. D., and J. Kaelberer, 1975, Phys. Rev. A 11, 1068. M. Soler, 1999, Eur. Phys. J. D 9, 211.
Etters, R. D., and J. Kaelberer, 1977, J. Chem. Phys. 66, 5112. Garzón, I. L., and A. Posada-Amarillas, 1996, Phys. Rev. B 54,
Even, U., N. Ben-Horin, and J. Jortner, 1989, Phys. Rev. Lett. 11 796.
62, 140. Garzón, I. L., J. A. Reyes-Nava, J. J. Rodriguez-Hernandez, I.
Faken, D., and H. Jónsson, 1994, Comput. Mater. Sci. 2, 279. Sigal, M. R. Beltrán, and K. Michaelian, 2002, Phys. Rev. B
Farges, J., M.-F. de Feraudy, B. Raoult, and G. Torchet, 1981, 66, 073403.
Surf. Sci. 106, 95. Ghayal, M. R., and E. Curotto, 2000, J. Chem. Phys. 113, 4298.
Farges, J., M. F. de Feraudy, B. Raoult, and G. Torchet, 1983, Gilb, S., P. Weis, F. Furche, R. Ahlrichs, and M. M. Kappes,
J. Chem. Phys. 78, 5067. 2002, J. Chem. Phys. 116, 4094.
Farges, J., M. F. de Feraudy, B. Raoult, and G. Torchet, 1985, Girifalco, L. A., 1992, J. Phys. Chem. 96, 858.
Surf. Sci. 156, 370. Girifalco, L. A., and V. G. Weizer, 1959, Phys. Rev. 114, 687.
Farges, J., M. F. de Feraudy, B. Raoult, and G. Torchet, 1986, Grisenti, R. E., W. Schöllkopf, J. P. Toennies, G. C. Hegerfeldt,
J. Chem. Phys. 84, 3491. T. Köhler, and M. Stoll, 2000, Phys. Rev. Lett. 85, 2284.
Farges, J., M.-F. de Feraudy, B. Raoult, and G. Torchet, 1988, Grönbeck, H., and W. Andreoni, 2000, Chem. Phys. 262, 1.
Adv. Chem. Phys. 70, 45. Grossman, J. C., and L. Mitás, 1995a, Phys. Rev. Lett. 74, 1323.
Ferreira, A. L. C., J. M. Pacheco, and J. P. Prates-Ramalho, Grossman, J. C., and L. Mitás, 1995b, Phys. Rev. B 52, 16735.
2000, J. Chem. Phys. 113, 738. Grygorian, V. G., and M. Springborg, 2003, Chem. Phys. Lett.
Finnila, A. B., M. A. Gomez, C. Sebenik, C. Stenson, and J. D. 375, 219.
Doll, 1994, Chem. Phys. Lett. 219, 343. Guardiola, R., 2000, Phys. Rev. B 62, 3416.
Finnis, M. W., and J. E. Sinclair, 1984, Philos. Mag. A 50, 45. Guardiola, R., and J. Navarro, 2000, Phys. Rev. Lett. 84, 1144.
Fortunelli, A., 1999, J. Mol. Struct.: THEOCHEM 493, 233. Guardiola, R., and J. Navarro, 2002, Phys. Rev. Lett. 89,
Fortunelli, A., and E. Aprà, 2003, J. Phys. Chem. 107, 2934. 193401.
Fortunelli, A., and A. M. Velasco, 2004, Int. J. Quantum Guillopé, M., and B. Legrand, 1989, Surf. Sci. 215, 577.
Chem. 99, 654. Guirado-López, R., M. C. Desjonquères, and D. Spanjaard,
Fournier, R., 2001, J. Chem. Phys. 115, 2165. 2000, Phys. Rev. B 62, 13 188.
Fournier, R., S. B. Sinnott, and A. E. DePristo, 1992, J. Chem. Gupta, R. P., 1981, Phys. Rev. B 23, 6265.
Phys. 97, 4149. Güvenc, Z. B., J. Jellinek, and A. F. Voter, 1992, in Physics and
Frantz, D. D., 1995, J. Chem. Phys. 102, 3747. Chemistry of Finite Systems: From Clusters to Crystals, edited
Frantz, D. D., 2001, J. Chem. Phys. 115, 6136. by P. Jena, S. N. Khanna, and B. K. Rao 共Kluwer Academic,
Frantz, D. D., D. L. Freeman, and J. D. Doll, 1990, J. Chem. Dordrecht兲, p. 411.
Phys. 93, 2769. Haberland, H., 1994, in Clusters of Atoms and Molecules, ed-
Freeman, D. L., and J. D. Doll, 1985, J. Chem. Phys. 82, 462. ited by H. Haberland 共Springer, Berlin兲, p. 374.
Freeman, D. L., and J. D. Doll, 1996, Annu. Rev. Phys. Chem. Haberland, H., 2002, Atomic Clusters and Nanoparticles: Les
47, 43. Houches Session LXXIII, 2-28 July 2000 共Springer, Berlin兲, p.
Frenkel, D., and B. Smit, 1996, Understanding Molecular Simu- 31.
lations 共Academic Press, San Diego兲. Häberlen, O. D., S.-C. Chung, M. Stener, and N. Rösch, 1996,
Fthenakis, Z., A. N. Andriotis, and M. Menon, 2003, J. Chem. J. Chem. Phys. 106, 5189.
Phys. 119, 10 911. Hagen, M. H. J., E. J. Meijer, G. C. A. M. Mooij, D. Frenkel,

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


420 F. Baletto and R. Ferrando: Structural properties of nanoclusters

and H. N. W. Lekkerkerker, 1993, Nature 共London兲 365, 425. ters 20, 239.
Häkkinen, H., and U. Landman, 2000, Phys. Rev. B 62, R2287. Jennison, D. R., P. A. Schultz, and M. P. Sears, 1997, J. Chem.
Häkkinen, H., M. Moseler, and U. Landman, 2002, Phys. Rev. Phys. 106, 1856.
Lett. 89, 033401. Jensen, P., 1999, Rev. Mod. Phys. 71, 1695.
Hall, B. D., M. Flüeli, R. Monot, and J.-P. Borel, 1991, Phys. Johnston, R. L., 2002, Atomic and Molecular Clusters 共Taylor
Rev. B 43, 3906. & Francis, London兲.
Hanszen, K. J., 1960, Z. Phys. 157, 523. Johnston, R. L., 2003, J. Chem. Soc. Dalton Trans. 2003, 4193.
Harris, I. A., R. S. Kidwell, and J. A. Northby, 1984, Phys. Rev. Jones, R. O., 1991, Phys. Rev. Lett. 67, 224.
Lett. 53, 2390. Jones, R. O., 1993, J. Chem. Phys. 99, 1194.
Hartke, B., 1993, J. Phys. Chem. 97, 9973. Jones, R. O., and G. Seifert, 1992, J. Chem. Phys. 96, 7564.
Hartke, B., 1999, J. Comput. Chem. 20, 1752. Jortner, J., 1992, Z. Phys. D: At., Mol. Clusters 24, 247.
Hartke, B., 2000, Z. Phys. Chem. 共Munich兲 214, 1251. Jortner, J., 2003, J. Chem. Phys. 119, 11 335.
Hartke, B., 2002, Angew. Chem., Int. Ed. 41, 1468. José-Yacamán, M., J. A. Ascencio, H. B. Liu, and J. Gardea-
Hartke, B., 2003, Phys. Chem. Chem. Phys. 5, 275. Torresdey, 2001, J. Vac. Sci. Technol. B 19, 1091.
Hendy, S., S. A. Brown, and B. D. Hall, 2003, Phys. Rev. B 68, José-Yacamán, M., M. Marín-Almazo, and J. A. Ascencio,
241403. 2001, J. Mol. Catal. A: Chem. 173, 61.
Hendy, S. C., and B. D. Hall, 2001, Phys. Rev. B 64, 085425. Joshi, K., D. G. Kanhere, and S. A. Blundell, 2002, Phys. Rev.
Henry, C. R., 1998, Surf. Sci. Rep. 31, 235. B 66, 155329.
Hill, T. L., 1964, Thermodynamics of Small Systems, Parts I Joshi, K., D. G. Kanhere, and S. A. Blundell, 2003, Phys. Rev.
and II 共Benjamin, Amsterdam兲. B 67, 235413.
Ho, J., K. M. Ervin, and W. C. Lineberger, 1990, J. Chem. Jug, K., B. Zimmermann, P. Calaminici, and A. M. Köster,
Phys. 93, 6987. 2002, J. Chem. Phys. 116, 4497.
Ho, J., K. M. Ervin, M. L. Polak, M. K. Gilles, and W. C. Jug, K., B. Zimmermann, and A. M. Köster, 2002, Int. J. Quan-
Lineberger, 1991, J. Chem. Phys. 95, 4845. tum Chem. 90, 594.
Ho, K. M., A. A. Shvartsburg, B. G. Pan, Z. Y. Lu, C. Z. Wang, Kallinteris, G. C., N. I. Papanicolaou, G. A. Evangelakis, and
J. G. Wacker, J. L. Fye, and M. F. Jarrold, 1998, Nature D. A. Papaconstantopoulos, 1997, Phys. Rev. B 55, 2150.
共London兲 392, 582. Kaxiras, E., and K. Jackson, 1993a, Phys. Rev. Lett. 71, 727.
Hoare, M. R., and J. McInnes, 1976, Faraday Discuss. Chem. Kaxiras, E., and K. Jackson, 1993b, Z. Phys. D: At., Mol. Clus-
Soc. 61, 12. ters 26, 346.
Hoare, M. R., and P. Pal, 1975, Adv. Phys. 24, 645. Kirkpatrick, S., C. D. Gelatt, and M. P. Vecchi, 1983, Science
Hohenberg, P., and W. Kohn, 1964, Phys. Rev. 136, B864. 220, 671.
Honeycutt, J. D., and H. C. Andersen, 1987, J. Phys. Chem. 91, Knight, W. D., K. Clemenger, W. A. de Heer, W. A. Saunders,
4950. M. Y. Chou, and M. L. Cohen, 1984, Phys. Rev. Lett. 52, 2141.
Hovick, J. W., and L. S. Bartell, 1997, J. Mol. Struct. 413, 615. Kofman, R., P. Cheyssac, A. Aouaj, Y. Lereah, G. Deutscher,
Huang, S. P., and P. B. Balbuena, 2002, J. Phys. Chem. B 106, T. Ben-David, J. M. Penisson, and A. Bourret, 1994, Surf. Sci.
7225. 303, 231.
Hudgins, R. R., M. Imai, M. F. Jarrold, and P. Dugourd, 1999, Koga, K., T. Ikeshoji, and K. Sugawara, 2004, Phys. Rev. Lett.
J. Chem. Phys. 111, 7865. 92, 115507.
Hyslop, M., A. Wurl, S. A. Brown, B. D. Hall, and R. Monot, Koga, K., and K. Sugawara, 2003, Surf. Sci. 529, 23.
2001, Eur. Phys. J. D 16, 233. Kohn, W., and L. J. Sham, 1965, Phys. Rev. 140, A1133.
Ikeshoji, T., B. Hafskjold, Y. Hashi, and Y. Kawazoe, 1996, Krivov, S. V., 2002, Phys. Rev. E 66, 025701.
Phys. Rev. Lett. 76, 1792. Kruger, S., S. Vent, F. Nörtemann, M. Staufer, and N. Rösch,
Ikeshoji, T., G. Torchet, M.-F. de Feraudy, and K. Koga, 2001, 2001, J. Chem. Phys. 115, 2082.
Phys. Rev. E 63, 031101. Kumar, V., and R. Car, 1991, Phys. Rev. B 44, 8243.
Imry, Y., 1980, Phys. Rev. B 21, 2042. Kumar, V., and Y. Kawazoe, 2002, Phys. Rev. B 66, 144413.
Ino, S., 1969, J. Phys. Soc. Jpn. 27, 941. Kusche, R., T. Hippler, M. Schmidt, B. von Issendorff, and H.
Ishikawa, Y., Y. Sugita, T. Nishikawa, and Y. Okamoto, 2001, Haberland, 1999, Eur. Phys. J. D 9, 1.
Phys. Lett. A 333, 199. Labastie, P., and R. L. Whetten, 1990, Phys. Rev. Lett. 65, 1567.
Jackschath, C., I. Rabin, and W. Schulze, 1992, Z. Phys. D: At., Lai, S. K., P. J. Hsu, K. L. Wu, W. K. Liu, and M. Iwamatsu,
Mol. Clusters 22, 517. 2002, J. Chem. Phys. 117, 10 715.
Jacobsen, K. W., J. K. Nørskov, and M. J. Puska, 1987, Phys. Lai, S. L., J. Y. Guo, V. Petrova, G. Ramanath, and L. H.
Rev. B 35, 7423. Allen, 1996, Phys. Rev. Lett. 77, 99.
Jarrold, M. F., and J. E. Bower, 1992, J. Chem. Phys. 96, 9180. Lathiotakis, N. N., A. N. Andriotis, M. Menon, and J. Con-
Jarrold, M. F., and V. A. Constant, 1991, Phys. Rev. Lett. 67, nolly, 1996, J. Chem. Phys. 104, 992.
2994. Leary, R. H., and J. P. K. Doye, 1999, Phys. Rev. E 60, R6320.
Jarrold, M. F., and E. C. Honea, 1991, J. Phys. Chem. 95, 9181. Ledoux, G., O. Guillois, D. Porterat, C. Reynaud, F. Huisken,
Jellinek, J., 1999, Theory of Atomic and Molecular Clusters B. Kohn, and V. Paillard, 2000, Phys. Rev. B 62, 15 942.
共Springer, Berlin兲. Lee, J., I.-H. Lee, and J. Lee, 2003, Phys. Rev. Lett. 91, 080201.
Jellinek, J., T. L. Beck, and R. S. Berry, 1986, J. Chem. Phys. Lee, J. K., J. A. Barker, and F. F. Abraham, 1973, J. Chem.
84, 2783. Phys. 58, 3166.
Jellinek, J., V. Bonacić-Koutecký, P. Fantucci, and M. Lee, Y. H., and B. J. Berne, 2000, J. Phys. Chem. A 104, 86.
Wiechert, 1994, J. Chem. Phys. 101, 10 092. Lee, Y. J., R. M. Nieminen, E.-K. Lee, and S. Kim, 2001, Com-
Jellinek, J., and I. L. Garzón, 1991, Z. Phys. D: At., Mol. Clus- put. Phys. Commun. 142, 201.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 421

Lermé, J., M. Pellarin, J. L. Vialle, B. Baguenard, and M. 1994, J. Chem. Phys. 100, 2322.
Broyer, 1992, Phys. Rev. Lett. 68, 2818. Massen, C., T. V. Mortimer-Jones, and R. L. Johnston, 2002, J.
Levi, A. C., and R. Mazzarello, 2001, J. Low Temp. Phys. 122, Chem. Soc. Dalton Trans. 2002, 4375.
75. Massobrio, C., A. Pasquarello, and R. Car, 1995, Chem. Phys.
Lewerenz, M., 1996, J. Chem. Phys. 106, 4596. Lett. 238, 215.
Lewis, L. J., P. Jensen, and J.-L. Barrat, 1997, Phys. Rev. B 56, Matsuoka, H., T. Hirokawa, M. Matsui, and M. Doyama, 1992,
2248. Phys. Rev. Lett. 69, 297.
Li, B. X., P. L. Cao, and M. Jiang, 2000, Phys. Status Solidi B Matulis, V. E., O. A. Ivashkevich, and V. S. Gurin, 2003, J. Mol.
218, 399. Struct.: THEOCHEM 664-665, 291.
Li, B. X., M. Qiu, and P. L. Cao, 1999, Phys. Lett. A 256, 386. Matveentsev, A., A. Lyalin, I. A. Solov’yov, A. V. Solov’yov,
Li, J., X. Li, H.-J. Zhai, and L.-S. Wang, 2003, Science 299, 864. and W. Greiner, 2003, Int. J. Mod. Phys. E 12, 81.
Li, T. X., Y. J. Ji, S. W. Yu, and G. H. Wang, 2000, Solid State Mazzone, A. M., 2000, Philos. Mag. B 80, 95.
Commun. 116, 547. Mélinon, P., P. Keghelian, B. Prével, V. Dupuis, A. Perez, B.
Li, T. X., S. M. Lee, S. J. Han, and G. H. Wang, 2002, Phys. Champagnon, Y. Guyot, M. Pellarin, J. Lermé, M. Broyer, J.
Lett. A 300, 86. L. Rousset, and P. Délichere, 1998, J. Chem. Phys. 108, 4607.
Li, T. X., S. Y. Yin, Y. L. Ji, B. L. Wang, G. H. Wang, and J. J. Mélinon, P., P. Keghelian, B. Prével, A. Perez, G. Guiraud, J.
Zhao, 2000, Phys. Lett. A 267, 403. LeBrusq, J. Lermé, M. Pellarin, and M. Broyer, 1997, J.
Li, Y., E. Blaisten-Barojas, and D. A. Papaconstantopoulos, Chem. Phys. 107, 10 278.
1998, Phys. Rev. B 57, 15 519. Meunier, I., 2001, Ph.D. thesis 共University of Aix-Marseille II兲.
Li, Z., and H. A. Scheraga, 1987, Proc. Natl. Acad. Sci. U.S.A. Michaelian, K., 1998, Chem. Phys. Lett. 293, 202.
84, 6611. Michaelian, K., N. Rendón, and I. L. Garzón, 1999, Phys. Rev.
Lim, H. S., C. K. Ong, and F. Ercolessi, 1992, Surf. Sci. 270, B 60, 2000.
1109. Milani, P., and S. Iannotta, 1999, Cluster Beam Synthesis of
Lim, H. S., C. K. Ong, and F. Ercolessi, 1994, Comput. Mater. Nanostructured Materials 共Springer, Berlin兲.
Sci. 2, 495. Mirkin, C. A., R. L. Letsinger, R. C. Mucic, and J. J. Storhoff,
Liu, H. B., J. A. Ascencio, M. Pérez-Alvarez, and M. José- 1996, Nature 共London兲 382, 607.
Yacamán, 2001, Surf. Sci. 491, 88. Mitás, L., J. C. Grossman, I. Stich, and J. Tobik, 2000, Phys.
Liu, H. B., R. Perez, G. Canizal, and J. A. Ascencio, 2002, Rev. Lett. 84, 1479.
Surf. Sci. 518, 14. Montalenti, F., and R. Ferrando, 1999a, Phys. Rev. B 59, 5881.
Liu, P., and B. J. Berne, 2003, J. Chem. Phys. 118, 2999. Montalenti, F., and R. Ferrando, 1999b, Phys. Rev. Lett. 82,
Lloyd, L. D., R. L. Johnston, S. Sahli, and N. T. Wilson, 2004, 1498.
J. Mater. Chem. 2004, 1691. Morse, P. M., 1929, Phys. Rev. 34, 57.
Locatelli, M., and F. Schoen, 2003, Comput. Optim. Appl. 26, Moseler, M., H. Häkkinen, R. N. Barnett, and U. Landman,
173. 2001, Phys. Rev. Lett. 86, 2545.
López, M. J., and J. Jellinek, 1999, J. Chem. Phys. 110, 8899. Mottet, C., J. Goniakowski, F. Baletto, R. Ferrando, and G.
Lowen, H., 1994, Phys. Rep. 237, 249. Tréglia, 2004, Phase Transitions 77, 101.
Luo, F., G. C. McBane, G. Kim, G. F. Giese, and W. R. Gentry, Mottet, C., G. Tréglia, and B. Legrand, 1997, Surf. Sci. 383,
1993, J. Chem. Phys. 98, 3564. L719.
Luo, J., U. Landman, and J. Jortner, 1987, in Physics and Nalwa, H. S., 2004, Ed., Encyclopedia of Nanoscience and
Chemical of Small Clusters, edited by P. Jena, B. K. Rao, and Nanotechnology 共American Scientific, New York兲.
S. N. Khanna 共Plenum, New York兲, p. 201. Nam, H.-S., N. M. Hwang, B. D. Yu, and J.-K. Yoon, 2002,
Lynden-Bell, R. M., and D. J. Wales, 1994, J. Chem. Phys. 101, Phys. Rev. Lett. 89, 275502.
1460. Natanson, G., F. Amar, and R. S. Berry, 1983, J. Chem. Phys.
Ma, J. P., and J. E. Straub, 1994, J. Chem. Phys. 101, 533. 78, 399.
Mackay, A. L., 1962, Acta Crystallogr. 15, 916. Nautchil, V. V., and A. J. Petsin, 1980, Mol. Phys. 40, 1341.
Manninen, K., J. Akola, and M. Manninen, 2003, Phys. Rev. B Nava, P., M. Sierka, and R. Ahlrichs, 2003, Phys. Chem. Chem.
68, 235412. Phys. 5, 3372.
Manninen, K., and M. Manninen, 2002, Eur. Phys. J. D 20, 243. Nayak, S. K., B. Reddy, B. K. Rao, S. N. Khanna, and P. Jena,
Maranas, C. D., and C. A. Floudas, 1992, J. Chem. Phys. 97, 1996, Chem. Phys. Lett. 253, 390.
7667. Neirotti, J. P., F. Calvo, D. L. Freeman, and J. D. Doll, 2000, J.
Marks, L. D., 1984, Philos. Mag. A 49, 81. Chem. Phys. 112, 10 340.
Marks, L. D., 1994, Rep. Prog. Phys. 57, 603. Nelson, D. R., and F. Spaepen, 1989, Solid State Phys. 42, 1.
Martin, T. P., 1996, Phys. Rep. 273, 199. Nichols, F. A., and W. W. Mullins, 1965, J. Appl. Phys. 36, 1826.
Martin, T. P., 2000, Solid State Ionics 131, 3. Nishioka, H., K. Hansen, and B. R. Mottelson, 1990, Phys.
Martin, T. P., T. Bergmann, H. Gohlich, and T. Lange, 1991a, Rev. B 42, 9377.
Z. Phys. D: At., Mol. Clusters 19, 25. Northby, J. A., 1987, J. Chem. Phys. 87, 6166.
Martin, T. P., T. Bergmann, H. Gohlich, and T. Lange, 1991b, J. Northby, J. A., 2001, J. Chem. Phys. 115, 10 065.
Phys. Chem. 95, 6421. Northby, J. A., J. Xie, D. L. Freeman, and J. D. Doll, 1989, Z.
Martin, T. P., U. Näher, and H. Schaber, 1992, Chem. Phys. Phys. D: At., Mol. Clusters 12, 69.
Lett. 199, 470. Nygren, M. A., P. E. M. Siegbahn, U. Wahlgren, and H. Akeby,
Martin, T. P., U. Näher, H. Schaber, and U. Zimmermann, 1992, J. Phys. Chem. 96, 3633.
1993, Phys. Rev. Lett. 70, 3079. Oviedo, J., and R. E. Palmer, 2002, J. Chem. Phys. 117, 9548.
Martin, T. P., U. Näher, H. Schaber, and U. Zimmermann, Pacheco, J. M., and J. P. Prates-Ramalho, 1997, Phys. Rev.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


422 F. Baletto and R. Ferrando: Structural properties of nanoclusters

Lett. 79, 3873. Reuse, F. A., S. N. Khanna, and S. Bernel, 1995, Phys. Rev. B
Padeletti, G., and P. Fermo, 2003, Appl. Phys. A: Mater. Sci. 52, R11 650.
Process. 76, 515. Rey, C., L. J. Gallego, and J. A. Alonso, 1994, Phys. Rev. B 49,
Parks, E. K., K. P. Kerns, and S. J. Riley, 1998, J. Chem. Phys. 8491.
109, 10 207. Rey, C., G. García-Rodeja, and L. J. Gallego, 2004, Phys. Rev.
Parks, E. K., G. C. Nieman, K. P. Kerns, and S. J. Riley, 1997, B 69, 073404.
J. Chem. Phys. 107, 1861. Rey, C., J. García-Rodeja, and L. J. Gallego, 1997, Z. Phys. D:
Parks, E. K., L. Zhu, J. Ho, and S. J. Riley, 1994, J. Chem. At., Mol. Clusters 40, 395.
Phys. 100, 7206. Reyes-Nava, J. A., I. L. Garzón, M. R. Beltrán, and K. Michae-
Parks, E. K., L. Zhu, J. Ho, and S. J. Riley, 1995, J. Chem. lian, 2002, Rev. Mex. Fis. 48, 450.
Phys. 102, 7377. Romero, G., C. Barrón, and S. Gómez, 1999, Comput. Phys.
Parravicini, G. B., A. Stella, P. Tognini, P. G. Merli, A. Commun. 123, 87.
Migliori, P. Cheyssac, and R. Kofman, 2003a, Appl. Phys. Roques, J., C. Lacaze-Dufaure, and C. Mijoule, 2001, Surf. Sci.
Lett. 82, 1461. 479, 231.
Parravicini, G. B., A. Stella, M. C. Ungureanu, P. G. Merli, A. Rosato, V., M. Guillopé, and B. Legrand, 1989, Philos. Mag. A
Migliori, and R. Kofman, 2003b, Phys. Status Solidi B 237, 59, 321.
374. Rossi, G., A. Rapallo, C. Mottet, A. Fortunelli, F. Baletto, and
Patil, A. N., D. Y. Paithankar, N. Otsuka, and R. P. Andres, R. Ferrando, 2004, Phys. Rev. Lett. 93, 105503.
1993, Z. Phys. D: At., Mol. Clusters 26, 135. Röthlisberger, U., and W. Andreoni, 1991, J. Chem. Phys. 94,
Pawlow, P., 1909, Z. Phys. Chem., Stoechiom. Ver- 8129.
wandtschaftsl. 65, 1. Röthlisberger, U., W. Andreoni, and P. Giannozzi, 1992, J.
Perdew, J. P., K. Burke, and Y. Wang, 1996, Phys. Rev. B 54, Chem. Phys. 96, 1248.
16 533. Rousset, J. L., A. M. Cadrot, F. S. Aires, A. Renouprez, P.
Perdew, J. P., J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Melinon, A. Pérez, M. Pellarin, J. L. Vialle, and M. Broyer,
Pederson, D. J. Singh, and C. Fiolhais, 1992, Phys. Rev. B 46, 1996, Surf. Rev. Lett. 3, 1171.
6671. Rytkönen, A., H. Häkkinen, and M. Manninen, 1998, Phys.
Perdew, J. P., and Y. Wang, 1986, Phys. Rev. B 33, 8800. Rev. Lett. 80, 3940.
Peters, K. F., J. B. Cohen, and Y.-W. Chung, 1998, Phys. Rev. B Rytkönen, A., H. Häkkinen, and M. Manninen, 1999, Eur.
57, 13 430. Phys. J. D 9, 451.
Pillardy, J., and L. Piela, 1995, J. Phys. Chem. 99, 11 805. Rytkönen, A., S. Valkealahti, and M. Manninen, 1997, J.
Pimpinelli, A., and J. Villain, 1998, Physics of Crystal Growth Chem. Phys. 106, 1888.
共Cambridge University, Cambridge兲. Rytkönen, A., S. Valkealahti, and M. Manninen, 1998, J.
Polak, W., and A. Patrykiejew, 2003, Phys. Rev. B 67, 115402. Chem. Phys. 108, 5826.
Poteau, R., and F. Spiegelmann, 1993, J. Chem. Phys. 98, 6540. Sachdev, A., R. I. Masel, and J. B. Adams, 1992, Catal. Lett.
Qi, Y., T. Cagin, W. L. Johnson, and W. A. Goddard III, 2001, 15, 57.
J. Chem. Phys. 115, 385. Saha, D. K., K. Koga, and H. Takeo, 1997, Nanostruct. Mater.
Quirke, N., and P. Sheng, 1984, Chem. Phys. Lett. 110, 63. 8, 1139.
Raghavachari, K., and V. Logovinsky, 1985, Phys. Rev. Lett. Saha, D. K., K. Koga, and H. Takeo, 1999, Eur. Phys. J. D 9,
55, 2853. 539.
Raghavachari, K., and C. McMichael-Rohlfing, 1988, J. Chem. Sambles, J. R., 1971, Proc. R. Soc. London, Ser. A 324, 339.
Phys. 89, 2219. Santamaria, R., I. G. Kaplan, and O. Novaro, 1994, Chem.
Ramakrishna, M. V., and A. Bahel, 1996, J. Chem. Phys. 104, Phys. Lett. 218, 395.
9833. Sattler, K., J. Mühlback, and E. Recknagel, 1980, Phys. Rev.
Raoult, B., and J. Farges, 1973, Rev. Sci. Instrum. 44, 430. Lett. 45, 821.
Raoult, B., J. Farges, M. F. de Feraudy, and G. Torchet, 1989a, Schaaff, T. G., and R. L. Whetten, 2000, J. Phys. Chem. B 104,
Philos. Mag. B 60, 5067. 2630.
Raoult, B., J. Farges, M. F. de Feraudy, and G. Torchet, 1989b, Schmidt, M., J. Donges, T. Hippler, and H. Haberland, 2003,
Z. Phys. D: At., Mol. Clusters 12, 85. Phys. Rev. Lett. 90, 103401.
Rata, I., A. A. Shvartsburg, M. Horoi, T. Frauenheim, K. W. Schmidt, M., R. Kusche, T. Hippler, J. Donges, W. Kronmüller,
M. Siu, and K. A. Jackson, 2000, Phys. Rev. Lett. 85, 546. B. von Issendorff, and H. Haberland, 2001, Phys. Rev. Lett.
Reddy, B. V., S. N. Khanna, and B. I. Dunlap, 1993, Phys. Rev. 86, 1191.
Lett. 70, 3323. Schmidt, M., R. Kusche, W. Kronmüller, B. von Issendorff, and
Reinhard, D., B. D. Hall, P. Berthoud, S. Valkealahti, and R. H. Haberland, 1997, Phys. Rev. Lett. 79, 99.
Monot, 1997, Phys. Rev. Lett. 79, 1459. Schmidt, M., R. Kusche, B. von Issendorff, and H. Haberland,
Reinhard, D., B. D. Hall, P. Berthoud, S. Valkealahti, and R. 1998, Nature 共London兲 393, 238.
Monot, 1998, Phys. Rev. B 58, 4917. Schmude, R. W., Q. Ran, and K. A. Gingerich, 1993, J. Chem.
Reinhard, D., B. D. Hall, D. Ugarte, and R. Monot, 1997, Phys. 99, 7998.
Phys. Rev. B 55, 7868. Schmude, R. W., Q. Ran, K. A. Gingerich, and J. E. Kingcade,
Reiss, H., P. Mirabel, and R. L. Whetten, 1988, J. Phys. Chem. 1995, J. Chem. Phys. 102, 2574.
92, 7241. Scoles, G., and K. K. Lehmann, 2000, Science 287, 2429.
Reiss, H., and I. B. Wilson, 1948, J. Colloid Sci. 3, 551. Sebetci, A., and Z. B. Güvenc, 2003, Surf. Sci. 525, 66.
Reuse, F. A., and S. N. Khanna, 1995, Chem. Phys. Lett. 234, Shah, P., S. Roy, and C. Chakravarty, 2003, J. Chem. Phys. 118,
77. 10 671.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005


F. Baletto and R. Ferrando: Structural properties of nanoclusters 423

Shao, X., H. Jiang, and W. Cai, 2004, J. Chem. Inf. Comput. Wales, D. J., 1994b, J. Chem. Phys. 101, 3750.
Sci. 44, 193. Wales, D. J., 2003, Energy Landscapes with Applications to
Shvartsburg, A. A., and M. F. Jarrold, 2000, Phys. Rev. Lett. Clusters, Biomolecules and Glasses 共Cambridge University,
85, 2530. Cambridge, England兲.
Soler, J. M., M. R. Beltrán, K. Michaelian, I. L. Garzón, P. Wales, D. J., and J. P. K. Doye, 1997, J. Phys. Chem. A 101,
Ordejón, D. Sánchez-Portal, and E. Artacho, 2000, Phys. Rev. 5111.
B 61, 5771. Wales, D. J., J. P. K. Doye, M. A. Miller, P. N. Mortenson, and
Soler, J. M., I. L. Garzón, and J. D. Joannopoulos, 2001, Solid T. R. Walsh, 2000, Adv. Chem. Phys. 111, 1.
State Commun. 117, 621. Wales, D. J., and M. P. Hodges, 1998, Chem. Phys. Lett. 286,
Solov’yov, I. A., A. V. Solov’yov, and W. Greiner, 2002, Phys. 65.
Rev. A 65, 053203. Wang, G. M., E. Blastein-Barojas, A. E. Roitberg, and T. P.
Solov’yov, I. A., A. V. Solov’yov, and W. Greiner, 2004, Int. J. Martin, 2001, J. Chem. Phys. 115, 3640.
Mod. Phys. E 13, 697. Wang, J., G. Wang, and J. Zhao, 2003, Chem. Phys. Lett. 380,
Solov’yov, I. A., A. V. Solov’yov, W. Greiner, A. Koshelev, and 716.
A. Shutovich, 2003, Phys. Rev. Lett. 90, 053401. Wang, J. L., G. H. Wang, F. Ding, H. Lee, W. F. Shen, and J. J.
Spiegelman, F., R. Poteau, B. Montag, and P.-G. Reinhard, Zhao, 2001, Chem. Phys. Lett. 341, 529.
1998, Phys. Lett. A 242, 163. Wang, J. L., G. H. Wang, and J. J. Zhao, 2002, Phys. Rev. B 66,
Stave, M. S., and A. E. DePristo, 1992, J. Chem. Phys. 97, 3386. 035418.
Steinhardt, P. J., D. R. Nelson, and M. Ronchetti, 1983, Phys. Wang, L., Y. Zhang, X. Bian, and Y. Chen, 2003, Phys. Lett. A
Rev. B 28, 784. 310, 197.
Stillinger, F. H., and T. A. Weber, 1982, Phys. Rev. A 25, 978. Watari, N., and S. Ohnishi, 1998, Phys. Rev. B 58, 1665.
Stillinger, F. H., and T. A. Weber, 1985, Phys. Rev. B 31, 5262. Westergren, J., H. Grönbeck, A. Rosen, and S. Nordholm,
Sun, Q., Q. Wang, P. Jena, S. Waterman, and Y. Kawazoe, 1998, J. Chem. Phys. 109, 9848.
2003, Phys. Rev. A 67, 063201.
Westergren, J., and S. Nordholm, 2003, Chem. Phys. 290,
Sung, M.-W., R. Kawai, and J. H. Weare, 1994, Phys. Rev. Lett.
189.
73, 3552.
Westergren, J., S. Nordholm, and A. Rosen, 2003, Phys. Chem.
Sutton, A. P., and J. Chen, 1990, Philos. Mag. Lett. 61, 139.
Chem. Phys. 5, 136.
Swendsen, R. H., and J.-S. Wang, 1986, Phys. Rev. Lett. 57,
Wetzel, T. L., and A. E. DePristo, 1996, J. Chem. Phys. 105,
2607.
572.
Takagi, M., 1954, J. Phys. Soc. Jpn. 9, 359.
Whaley, K. B., 1994, Int. Rev. Phys. Chem. 13, 41.
Tchofo-Dinda, P., G. Vlastou-Tsinganos, N. Flyzanis, and A. D.
Whetten, R. L., J. T. Khoury, M. M. Alvarez, S. Murthy, I.
Mistriotis, 1995, Phys. Rev. B 51, 13 697.
Vezmar, Z. L. Wang, P. W. Stephens, C. L. Cleveland, W. D.
Tekin, A., and B. Hartke, 2004, Phys. Chem. Chem. Phys. 6,
Luedtke, and U. Landman, 1996, Adv. Mater. 共Weinheim,
503.
Toennies, J. P., and A. F. Vilesov, 1998, Annu. Rev. Phys. Ger.兲 5, 428.
Chem. 49, 1. Wille, L. T., 1987, Chem. Phys. Lett. 133, 405.
Tomanek, D., A. A. Aligia, and C. A. Balseiro, 1985, Phys. Wille, L. T., and J. Vennik, 1985, J. Phys. A 18, L419.
Rev. B 32, 5051. Wilson, N. T., and R. L. Johnston, 2000, Eur. Phys. J. D 12, 161.
Tsai, C. J., and K. D. Jordan, 1993, J. Phys. Chem. 97, 11 227. Xiao, Y., and D. E. Williams, 1993, Chem. Phys. Lett. 215, 17.
Turner, G. W., R. L. Johnston, and N. T. Wilson, 2000, J. Xie, J., J. A. Northby, D. L. Freeman, and J. D. Doll, 1989, J.
Chem. Phys. 112, 4773. Chem. Phys. 91, 612.
Uppenbrink, J., and D. J. Wales, 1992, J. Chem. Phys. 96, 8520. Xue, G. L., 1994, J. Global Optim. 4, 425.
Valerio, G., and H. Toulhoat, 1996, J. Phys. Chem. 100, 10 827. Yang, S. H., D. A. Drabold, J. B. Adams, P. Ordejón, and K.
Valerio, G., and H. Toulhoat, 1997, J. Phys. Chem. A 101, 1969. Glassford, 1997, J. Phys.: Condens. Matter 9, L39.
Valkealahti, S., and M. Manninen, 1993, Z. Phys. D: At., Mol. Yu, D. K., R. Q. Zhang, and S. T. Lee, 2002, Phys. Rev. B 65,
Clusters 26, 255. 245417.
Valkealahti, S., and M. Manninen, 1997, J. Phys.: Condens. Zacarias, A. G., M. Castro, J. M. Tour, and J. M. Seminario,
Matter 9, 4041. 1999, J. Phys. Chem. A 103, 7692.
Valkealahti, S., and M. Manninen, 1998, Phys. Rev. B 57, Zanchet, D., B. D. Hall, and D. Ugarte, 2000, J. Phys. Chem. B
15 533. 104, 11 013.
Valkealahti, S., U. Näher, and M. Manninen, 1995, Phys. Rev. Zeiri, Y., 1995, Phys. Rev. E 51, R2769.
B 51, 11 039. Zhang, W., Q. Ge, and L. Wang, 2003, J. Chem. Phys. 118,
van de Waal, B. W., 1989, J. Chem. Phys. 90, 3407. 5793.
van de Waal, B. W., 1996, Phys. Rev. Lett. 76, 1083. Zhao, J., Y. Luo, and G. Wang, 2001, Eur. Phys. J. D 14, 309.
van de Waal, B. W., G. Torchet, and M.-F. de Feraudy, 2000, Zhao, J., J. L. Yang, and J. G. Hou, 2003, Phys. Rev. B 67,
Chem. Phys. Lett. 331, 57. 085404.
Vanfleet, R. R., and J. M. Mochel, 1995, Surf. Sci. 341, 40. Zhou, Y., M. Karplus, K. D. Ball, and R. S. Berry, 2002, J.
Vlachos, D. G., L. D. Schmidt, and R. Aris, 1992, J. Chem. Chem. Phys. 116, 2323.
Phys. 96, 6880. Zhu, H., 1996, Philos. Mag. Lett. 73, 27.
Voter, A. F., 1993, Los Alamos Unclassified Technical Report Zickfeld, K., M. E. García, and K. H. Bennemann, 1999, Phys.
LA-UR 93, 3901. Rev. B 59, 13 422.
Wales, D. J., 1994a, J. Chem. Soc., Faraday Trans. 90, 1061.

Rev. Mod. Phys., Vol. 77, No. 1, January 2005

You might also like