Sie sind auf Seite 1von 23

www.aspbs.

com/enn
Encyclopedia of
Nanoscience and
Nanotechnology
Self-Organization of Colloidal
Nanoparticles
Joydeep Dutta and Heinrich Hofmann
Powder Technology Laboratory, Materials Institute, Swiss Federal Institute of Technology,
1015-Lausanne, Switzerland
CONTENTS
1. Introduction
2. Self-assembly and Self-Organization: Clarication
3. Self-Organization of Nanoparticles
Into Superlattice
4. Conclusions and Perspective
Glossary
Acknowledgments
References
1. INTRODUCTION
The emerging elds of nanoscience and nanoengineering
are leading to unprecedented understanding and control
over the fundamental building blocks of all physical matter.
Today, nanophase engineering expands in a rapidly growing
number of structural and functional materials, both inor-
ganic and organic, allowing the manipulation of mechan-
ical, catalytic, electrical, magnetic, optical, and electronic
functions. The synthesis of nanophase or cluster-assembled
materials is usually based upon the creation of separated
small clusters which are then fused into a bulk-like material
or as a thin lm or even by embedding into a solid matrix,
materials like polymers or glass either in a disordered man-
ner or in an ordered array.
There is a need to learn how to manipulate nanosized
particles and form the ensemble into periodically ordered
particulate materials for application into devices making use
of the special physical properties that arise due to the nano-
metric crystallite sizes, like the quantum effects, for example.
From recent developments in the cluster source technolo-
gies (thermal, laser vaporization, aerosol techniques, chemi-
cal vapor deposition, and sputtering) [14], it is now possible
to produce intense cluster beams of any materials, even the
most refractory or complex systems, for a wide range of sizes
from a few atoms to a few thousand atoms [5].
Signicant progress in the organization of semiconduc-
tor quantum dots (QD) has been made by employing
a variety of microelectronic fabrication techniques [610],
but the most widely used strategy employs the deposi-
tion of a narrower bandgap, compressively strained epi-
layer over a wider bandgap substrate, which results in
the spontaneous nucleation (self-assembly) of coherently
strained three-dimensional islands on top of a thin (1 nm),
two-dimensional wetting layer. Subsequent deposition of
substrate material to encapsulate the QD produces a fully
three-dimensional connement potential. This approach
however, is, outside the scope of the present review where
we will concentrate on the fabrication of ordered layers or
structures from colloidal particles.
An elegant method to organize nanometric particles
is by the so-called self-assembly methods often inspired
by nature (biomimetic processes). When self-assembly is
used to organize nanomaterial into a useful form, irre-
spective of whether the structures so formed are points,
lines, layers, or composites, they are going to be peri-
odic in arrangements, since such arrangements minimize the
free energy of the systems. Methods developed to produce
three-dimensional, bulk-like hierarchical structures include
biomimetic methods, which use polypeptides as building
blocks, and amphiphile and colloidal templating, which use
amphiphilic or colloidal mesophases as templates for inor-
ganic mesoporous materials. Designing nite mesostruc-
tures with a given geometry is the future challenge to
make use of the special properties arising due to the scal-
ing of the particle sizes for applications. Ordered assemblies
of nanometer-sized particles are thus an interesting class of
nanomaterials that provide exceptional potentials to achieve
one-two-and three-dimensional structures for a wide variety
of applications ranging from photonic devices to memory
devices, two-dimensional arrays of magnetic nanoparticles
[1113], or single-electron microelectronic devices, amongst
others [1416]. In order to make full use of this potential
for materials engineering, it is important to identify and gain
control over the relevant growth and ordering parameters
both during the formation of ordered structures as well as
after the growth process, so that the structures so formed
do not get distorted.
ISBN: 1-58883-001-2/$35.00
Copyright 2003 by American Scientic Publishers
All rights of reproduction in any form reserved.
Encyclopedia of Nanoscience and Nanotechnology
Edited by H. S. Nalwa
Volume X: Pages (123)
2 Self-Organization of Colloidal Nanoparticles
This is a eld in its nascent stage and experimentalists
are still struggling to build up on the ideas propagated by
the theorists to develop the complex technology required to
give two- or three-dimensional structures to a material on
the nano or micron scale (in this case the primary materials
are of nanoscopic length scales). More and more interesting
structures, prepared by a variety of methods twinned around
self-assembly methods, are being reported regularly in the
literature. Self-assembly and self-organization are concepts
that are broadly used over a whole range of elds includ-
ing physics, chemistry, biology, and engineering. Nanoscale
self-assembly is a concept that nature has been making use
of since the beginning of life, and it is only recently that
we have started realizing the potential of utilizing this phe-
nomenon for the articial control of matter [17].
Letting aside the self-assembly or self-organization pro-
cesses, even nanotechnology did not emerge as an exper-
imental science before the 1980s due to the unavailability
of suitable tools to allow researchers to observelet alone
manipulateindividual atoms. It evolved dramatically with
the invention of two new microscopy techniques: atomic
force microscopy (AFM) and scanning tunneling microscopy
(STM). Both these techniques were a radical departure from
previous types of microscopy, which worked by reecting
either light (in the case of optical microscopes) or an elec-
tron beam (in the case of electron microscopes) off a sur-
face and onto a lens. But no reective microscope, not even
the most powerful, could image an individual atom. The
discovery of the STM got scientists and engineers to look
into the self-organization processes of nanoscopic matter
more closely, as STM brought in the possibility of visual-
izing at the nanoscopic levels and also the possibility of
moving atoms to structure them into useful forms. Direct
writing of nanoscopic features with atoms is a very slow
process that is normally achieved, for example, by utilizing
tools like the scanning probe microscope (SPM). Thus, a
more practical method is albeitthe use of natural forces to
structure nanoscopic features. With the discovery of other
microscopic tools like AFM, and the improvement of the
electron microscopy tools further developed the possibilities
of visualizing and implementing the experiments designed
by researchers which has given boosted efforts to explore
the immense possibilities this eld has in reserve.
Two approaches for the building up of nanoscopic fea-
tures have been envisaged: the so-called top-down method
and bottom-up methods. In the top down methods, the
features are written directly onto a substrate, for exam-
ple, by electron beams, and then by applying appropriate
etching and deposition processes, the nanoscopic features
are engraved. In the bottom-up approach, nanocompo-
nents are made from precursors in the liquid, solid, or gas
phase employing either chemical or physical deposition pro-
cesses that are integrated into building blocks within the
nal material structure. The bottom-up approach, or sim-
ply the assembly from atoms, include wide-ranging meth-
ods encompassing supramolecular chemistry to biomimetic
processes to template synthesis, amongst others. The aim
of this article is to consider the bottom-up approach for
building up nanostructures focusing on some of the self-
organization methods that are used in structuring colloidal
nanoparticles into regular lms, arrays, or three-dimensional
structures. This article will attempt to distinguish between
the self-assembly processes (that is often misconceived in
some experiments reported in the literature) from the self-
organization of colloidal nanoparticles.
2. SELF-ASSEMBLY AND
SELF-ORGANIZATION: CLARIFICATION
Organization during self-assembly processes is driven mainly
by competing molecular interactions like hydrophobic ver-
sus hydrophilic components, gravitational, van der Waals, or
Coulombic interactions. Self-assembly is an assembly pro-
cess in which only the constituents of the nal structure take
part, that is, get incorporated into the resulting structure
[18]. Self-organization, in retrospect, is dened as a mech-
anism for building patterns, processes, and structures at a
higher level through multiple interactions among the com-
ponents at the lower level(s), where the components interact
through local rules that do not explicitly code for the pat-
tern [1920]. In colloidal materials, self-assembly process has
often been related to describe the growth of nanoparticles,
for example, of controlling the size of the particles by self-
organization of polymers on these growing particles and has
to be clearly differentiated from the self-organization pro-
cesses of the colloidal particles into either two- or three-
dimensional structures.
The phenomena of pattern formation and self-organization
occur as natural responses of complex systems to strong
external stimulation. The collective interaction between
system components under external driving forces, which
drive the system far from equilibrium, results in the self-
organization of its constituents [2122]. Some of the salient
features of this viewpoint can be found in the behavior
of chemical [2324], liquid crystal nematics [2526], and
uid systems [27]. It has also been recognized that con-
densed matter systems show a rich variety of patterns and
self-assembled microstructures under conditions as diverse
as solidication [28], electro-chemical deposition [29], plas-
tic deformation [3031], surface modications [32], rapid
laser heating [33], irradiation by energetic particles [34],
magnetic domains [35], and more recently, atom deposi-
tion into self-assembled monolayers [36]. All these phenom-
ena are currently being revisited, as self-assembly is nothing
new; for example, lustre decoration of medieval and renais-
sance pottery consists of self-assembled silver and copper
nanocrystals within the glassy phase of the ceramic glaze.
A peculiar property of this lustre was enumerated to arise
due to the high density of pseudo-spherical nanocrystals
and their locally homogeneous distribution in the ceramic
[3739]. Another historic example of self-assembly (or even
self-organization) of nanoparticles is found in the Maya
blue paints [40]. Studies of authentic samples from Jaina
Island show that the material is made of needle-shaped paly-
gorskite (a clay) crystals that form a superlattice with a
period of 1.4 nm, with intercalates of amorphous silicate
substrate containing inclusions of metal (Mg) nanoparticles.
The beautiful tone of the blue color is obtained only when
both these nanoparticles and the superlattice are present
together, as has been shown by the fabrication of synthetic
samples.
Self-Organization of Colloidal Nanoparticles 3
2.1. The Self-Assembly Process
As explained earlier, self-assembly in a general sense, might
be dened as the spontaneous formation of complex hier-
archical structures from predesigned building blocks, typi-
cally involving multiple energy scales and multiple degrees
of freedom. It is also a very general principle in nature, as
seen in the formation for example, of membranes from lipid
molecules or the living cell as probably the most important
paradigm [41]. Self-assembly has been dened by Kuhn and
Uhlman as a process in which supramolecular hierarchical
organization is established in a complex system of interlock-
ing components [42]. Self-assembled monolayers (SAMs)
have been studied in great details by surface scientists by
using different types of molecules on a host of different sub-
strates. Detailed description and discussion on the SAMs
are outside the scope of this article, but a comprehensive
report of SAMs as described by surface scientists have been
recently published [43]. As an example, the self-assembly can
be achieved through an alkane chain typically with 1020
methylene units that can be given a head group with a strong
preferential adsorption to a chosen substrate. For example,
thiol (S-H) groups adsorb readily from a solution onto a
gold surface creating a dense monolayer with the tail group
pointing outwards from the surface. Numerous work, both
calculations [44] and experiment [45], has been reported
describing the self-assembly of thiol molecules to control
gold nanoparticle sizes during synthesis [46] and/or during
the assembly of these particles on a substrate including a
milestone review recently published by Brust and Kiely [47].
2.2. Surface Modication of Nanoparticles
Particles in the nanometer-size range have a strong tendency
to agglomerate due to van der Waals interactions. It is there-
fore important to develop synthetic methods by which the
particles can be stabilized, that is, where repulsive forces
between the particles can be provided to balance this attrac-
tion (Fig. 1(A)) [48]. Generally two types of stabilization
are used to prevent agglomeration of nanoparticles, namely,
electrostatic stabilization and steric stabilization by adsorbed
molecules or steric hindrance [4950]. Electrostatic stabi-
lization involves the creation of an electrical double layer
arising from ions adsorbed on the surface and associated
counterions that surround the particle. Thus, if the electric
potential associated with the double layer is sufciently high,
the Coulombic repulsion between the particles will prevent
their agglomeration (Fig. 1(B) and Fig. 2) [5152].
(A) (B)
Figure 1. Particles in a colloid (A) uncharged particles are free to col-
lide and agglomerate and (B) charged particles repel each other.
Figure 2. Schematic representation of the stabilization of a nano-
particle by charging the surface with counter ions.
Steric stabilization can be achieved by the adsorption of
large molecules such as polymers at the surface of the par-
ticles (Fig. 3). Indeed, the coil dimensions of polymers are
usually larger than the range over which the attraction forces
between colloidal particles are active. Two distinct effects
describe this type of stabilization (the volume restriction
contribution and the osmotic diffusion), and they both con-
tribute to the interaction free energy [46, 5354]. First, the
fact that the adsorbed molecules are restricted in motion
causes a decrease in the congurational entropic contribu-
tion to the free energy (Fig. 3). Second, the local increase
in concentration of polymer chains between approaching
particles results in an osmotic repulsion, since the solvent
reestablishes equilibrium by diluting the polymer molecules
and separating the particles. This can be described by the
energy of free mixing of polymer segments and solvent
molecules, calculated by the FloryKrigbaum theory [55].
The conformation of adsorbed polymers tends to be con-
trolled by the strength of segment/surface interactions [56],
which may be described by the classical loop-train-tail model
[5758]. For an effective particle stabilization, it is impor-
tant that the polymer form a complete, dense layer around
the particle [59]. Then, the polymeric stabilizer must have
sufcient tail length and adsorb uniformly enough to screen
the attractive interaction between the particles.
Another requirement for good particle stabilization is the
use of appropriate solvents. The stabilizing polymer has to
possess high afnity with the solvent in order to solvate
the particles and form an extended layer for screening the
(A) (B)
Figure 3. Steric stabilization of metal colloid particles by polymers:
(A) entropic effect, (B) osmotic effect.
Proof's Only
4 Self-Organization of Colloidal Nanoparticles
van der Waals attraction between the particles. It is well
known, however, that adsorption is generally stronger when
the afnity of the polymer to the solvent is low [6061].
Therefore, block or graft copolymer micelles, in the pres-
ence of a selective solvent that solubilizes one of the blocks,
lead to expanded chains which occupy more surface area. In
the case of polymers with strong anchoring groups, sponta-
neous adsorption occurs from good solvent [46]. De Gennes
has discussed the density prole of polymers attached to
the surface by one end and in contact with an appropri-
ate solvent [62]. Another important advantage of covalently
attached polymers is that the resulting interaction with the
particle surface is stronger and desorption cannot occur by
solvent effects or variations in temperature [6364].
As previously mentioned, alkanethiols spontaneously self-
assemble onto colloidal gold particles and this capping of
gold particles by alkanethiols are sometimes termed three-
dimensional self-assembly, which is often differentiated from
the so-called two-dimensional self-assembly on gold thin
lms [6566] or alkylsilanes on silica lms. To differenti-
ate between the arrangement of nanoparticles into ordered
structures and the passivation of nanoparticles by polymeric
entities [67], it is our belief the stabilization of nanoparticle
growth or agglomeration by polymers is a self-assembly pro-
cess. For example, the self-assembly on gold particles occurs
through a strong thiolate bond between the surfactant and
the gold surface which leads to the assembly of these sur-
factant molecules on the surface of gold nanoparticles [68].
There are numerous reports on the self-assembly of alka-
nethiols, aromatic thiols, and alkylamines onto gold and sil-
ver colloidal particles, which has been reviewed in great
detail [47, 6970]. Figure 4 illustrates an example of sur-
face derivatization of gold and silver particles with three-
dimensional SAMs of w-functionalized thiol molecules.
Analogous to the attachment of organic stabilizers, bio-
logical components have also been used for stabilization of
these colloids. Typically, wet-chemical preparation of the
nanoparticles is carried out in the presence of stabilizing
agents (often citrates, phosphanes, or thiols) which bind to
the atoms exposed at the surface of the nanoparticles. This
capping leads to a stabilization and prevents uncontrolled
growth and aggregation of the nanoparticles. In the case of
a labile capping layer such as citrate, biomolecules can be
linked directly with a metal particle by exchange reactions
with stronger binding ligands. This method is usually applied
in the coating of colloidal gold with thiol-containing pro-
teins, for example, immunoglobulins (IgG) and serum albu-
mins, which have cysteine residues that are accessible for
the heterogeneous interphase coupling. If no such residues
-functionalized alkanethiol
Figure 4. Schematic representation of the surface functionalization of
a gold or silver nanoparticle by an alkanethiol.
are available in the native proteins, thiol groups can be
incorporated by chemical means [71] or by genetic engi-
neering. For example, DNA molecules can be synthesized
with alkylthiol groups, and n-alkylthiolated DNA has been
used extensively in the preparation of DNA-functionalized
gold and semiconductor nanoparticles [7277]. As an inter-
esting alternative, DNA oligonucleotides, which contain sev-
eral adenosyl phosphothioate residues at their ends, have
also been used to directly interact with the metal surface
of nanoparticles [78]. The use of cyclic disulde linkers
[7778] leads to nanoparticle capping, which are more sta-
ble towards ligand exchange than the corresponding conju-
gates prepared with the conventional reagents that contain
a single thiol group or acyclic disulde units. The greater
stability is likely a result of the anchoring of the ligands to
the nanoparticles through the two sulfur atoms [77].
Surface functionalized metal nanoparticles have been
used for a variety of applications. Terms like nanofactory
has been often cited, which basically study the reactivitys
of functional groups anchored to gold colloidal particles
[47, 79]. Another approach to use these self-assembled
metal clusters has arisen due to the local eld enhancement
and nanoscale resonant behavior of metal clusters excited
by an electromagnetic radiation. In metal particles, absorp-
tion in the UV-visible spectra (with a peak at 520 nm for
10 nm gold particles, for instance) is due to the excitation
of a plasmon resonance which corresponds to a collective
oscillation frequency of conduction electrons [80]. These
absorption spectra can be theoretically explained by the Mie
theory [8182], in which the theoretical absorption spectrum
of dilute spherical particles (surface plasmon resonance) is
related to their size, and their frequency-dependant dielec-
tric constant, compared to the surrounding medium [8384].
Immunogold labeling and immunogold chromatography as
well as surface plasmon resonance transduction of metal
cluster binding, which has been successfully commercial-
ized using these surface stabilized nanoparticles, is outside
the scope of this review [85]. Numerous other applica-
tions that include the deposition of these self-assembled
nanoparticles have been reported, for example, local plas-
mon sensor [86], colorimetric DNA analysis [87], optically
anisotropic windows [8889], surface enhanced raman spec-
troscopy (SERS) substrates [90], amongst others. Several
authoritative reviews are available dealing with the self-
assembly and self-organization of gold nanoparticles [47, 91].
Self-assembled monolayer formation provides an easy route
towards surface functionalization by organic molecules con-
taining suitable functional groups like -SH, -CN, -COOH,
-NH
2
, and silanes on selected metallic (Au, Cu, Ag, Pd,
Pt, Hg, and C) [92], as well as semiconducting surfaces
(Si, GaAs, ZnS, CdS, etc.) [9398]. Self-assembly inside a
nanoscale cavity has also been reported where, for exam-
ple, the reaction between divalent cadmium or zinc with sul-
de or selenide is performed in a restricted environment in
the nanometer scale. Examples of nanoscale cavities include
porous glasses and xerogels [99], reverse micelles [100],
zeolites [101], membranes [102103], and hollow proteins,
amongst others [104]. Yeast and tomato virus, and likely
other organisms, produce CdS quantum dots as a detoxica-
tion response to an overload of cadmium [105106]; the CdS
Proof's Only
Self-Organization of Colloidal Nanoparticles 5
thus produced is coated with a particular peptide, which was
used as the stabilizing thiol to make CdS quantum dots [107].
3. SELF-ORGANIZATION
OF NANOPARTICLES
INTO SUPERLATTICE
The general interest in the self-organization of spherical
and non-spherical nanoparticles, of late, has arisen due
to the possible applications in a wide range of domains
[91, 108110]. Periodic arrays are structures where nanopar-
ticles are arranged in a regular fashion in either two or three
dimensions [111]. These structures would be useful for dis-
plays, sensors, data storage, or photonic bandgap materials
[91]. Monosized spheres of silica and polystyrene have most
often been used for studying the self-organization processes,
due to their easy availability. On the other side, QDs of
semiconducting and metallic materials, due to their poten-
tial and already realized applications, as well as fundamental
chemical and physical interest, have been studied in great
details with some activities on the self-organization phenom-
ena reported [112].
Nonlithographic bottom-up approaches based on self-
assembly and self-organization are especially appealing
because of their intrinsically low overhead for large-scale
production. This approach has been useful in the self-
organization of monolayer-protected metal nanoparticles
into periodic two-dimensional (2D) arrays, with many of
these assemblies demonstrating novel optical or electronic
properties as a function of particle size or interparticle spac-
ing [68, 113116]. Numerous examples of 2D arrays com-
prised of small (-10 nm) gold nanoparticles have been
reported, but details on well-ordered 2D arrays of larger
gold nanoparticles are not available in the literature. Parti-
cles beyond a certain size (about 15 nm) tend to agglomerate
into multilayer or three-dimensional aggregates rather than
form 2D monoparticulate lms [117]. This can be attributed
to the large Hamaker constant for gold [118119] and the
rapid increase in van der Waals attraction between particles
as a function of size [120], as well as the loss of surfactant
chain mobility on the planar facets of the nanoparticles.
Polystyrene (PS) nanoparticles are commercially avail-
able and have been used for a wide range of applications
ranging from applications in elds of photonics [121125],
nanotechnology [126132], and life science [133140] to fun-
damental studies in the behavior of colloidal suspensions,
like the role of hydrodynamic [140141] and entropic [142]
forces and studies on phase transition [143] and crystalliza-
tion [112] of model atoms in two or three dimensions.
Irrespective of this plurality of applications, at the focus
of these activities is the adsorption of the particles as a
2D layer onto a surface, where, depending on the specic
aims of the subject, close-packed ordered arrays or irreg-
ular, less densely packed layers are selectively used. The
schemes proposed so far for patterning PS particles range
from purely physical concepts, such as micro-molding in cap-
illaries (MIMIC) [144] and electrophoresis(to be discussed
in a later section) to those combining physical interaction
mechanisms like electrostatic or capillary forces with the
selectivity of chemically patterned surfaces [145146]. The
Assembly of periodic structure
Self-assembly Self-organization Template-directed assesmbly
External field assisted Chemically enhanced
Figure 5. Diagram showing various self-organization techniques
reported in the literature.
self-organization techniques that have been used to build 2D
or three-dimensional (3D)-ordered structures are schemat-
ically represented in Figure 5 and are discussed in greater
details in the next few sections.
3.1. Self-Organization Through
Capillary Forces
A comprehensive review of various types of capillary forces
on particles has been reported by Kralchevsky and Denkov,
which is summarized in Figure 6 [147]. If a droplet of the
colloid suspension is dried slowly on an unpatterned polar
surface (such as silica), the particles aggregate at the rim
of the droplet because of attractive capillary forces between
the particles when the water lm thickness is in the dimen-
sion of the particle diameter, which is schematically shown
in Figure 6 [147148]. The mobility of the electrostatically
attached particles is lost after complete drying, suggesting
a three-step mechanism for the particle adsorption. First,
positioning and adhesion in the suspension liquid are con-
trolled by charge and polar interactions between the sub-
strate and particle surfaces. In the second step, capillary
forces between the particles and the surface laterally dis-
place the particles during drying. After complete evapora-
tion in the third stage, an irreversible reorganization of the
particlesubstrate interface occurs, which prevents resuspen-
sion or displacement when resubmerging the assembly into
the adsorption liquid. A submonolayer of electrostatically
adsorbed particles forms on the charged pattern surface in
the bulk phase of the suspension. The submonolayer cover-
age is caused by Coulomb repulsion of the like-charged par-
ticles, which prevents dense packing in the aqueous medium.
During evaporation of the suspension, the liquid front moves
very slowly over the adsorbed particle layer, and a thin liq-
uid meniscus remains over the particle assembly, which gen-
erates attractive capillary forces between the particles, as
shown in Figure 6(c).
Because of the slow-moving, droplet front, particles from
the suspension do have time to migrate into the thin liq-
uid lm at the drying front and aggregate into multilayer
over the colloid assembly pattern driven by capillary forces
(Fig. 7). Many more factors inuencing the particlesurface
interactions, such as surface topography and corrugation
[149151] or even nanoscopic air bubbles [152] might have
to be taken into account in order to completely understand
the ordering phenomena of colloidal nanoparticles through
capillary forces. Another important factor that should be
considered is that during the drying process, the ionic
strength of the colloid changes affecting the double-layer
thickness and hence the nal organization process.
Proof's Only
6 Self-Organization of Colloidal Nanoparticles
CAPILARY FORCES
Figure 6. Types of capillary forces: (a) The normal capillary forces can be due to either liquid-in-gas or gas-in-liquid capillary bridges, which
lead to particle-particle and particle-wall interactions, the force is directed normally to the contact line. In the case of lateral capillary forces
(b, c, d, e), the force is parallel to the contact line. The interaction is due to the overlap of interfacial deformations created by the separate
particles. (b) In the case of otation force, the deformations are caused by the particle weight and buoyancy. In the case of immersion forces
(c, d, e) the deformations are related to the wetting properties of the particle surface: position and shape of the contact line, and magnitude of
the contact angle. When the deformation around an isolated particle is axisymmetric, we deal with capillary charges, one can distinguish cases of
innite (c) and nite (d) menisci. (e) The forces between particles of undulated or irregular contact line can be described as interactions between
capillary multipoles, in analogy with electrostatics. Reprinted from [148], P. A. Kralchevsky and N. D. Denkov, Curr. Opin. Coll. Inter. Sci., 6, 383
(2001). 2001, Elsevier Science.
3.2. Template-Assisted
In this section, we review only the approaches that com-
bine templating with attractive capillary forces to assemble
particles into complex aggregates with well-controlled sizes,
shapes, and internal structures [153154]. Surface conne-
ment provided by liquid droplets has been used to assemble
colloidal particles or microfabricated building blocks into
spherical objects [155157]. In this process, patterned arrays
Figure 7. Schematic representation of a droplet of a colloid suspension
dried slowly on an unpatterned polar surface (such as silica), the parti-
cles aggregate at the rim of the droplet because of attractive capillary
forces between the particles when the water lm thickness is in the
dimension of the particle diameter.
of templates are fabricated in thin lms of photoresists
(for example, spin-coated on glass, silicon wafer, or quartz
substrates), using conventional photolithography techniques
used in microelectronics or printing processes. Control over
the number of particles in each such template is governed by
the size of the particles and the size and shape of the tem-
plates [153]. Patterned arrays of relieves on solid substrates
have also been exploited to grow colloidal crystals with
unusual crystalline order and orientations [149, 158159].
Templating against opaline arrays of colloidal spheres
has also been successfully applied to the fabrication of
3D macroporous structures from a wide variety of mate-
rials, including organic polymers, ceramic materials, inor-
ganic semiconductors, and metals (for example, as in Fig. 8)
[161176]. Fabrication based on this approach is remarkable
for its simplicity, and for its delity in transferring the struc-
ture from the template to the replica. The size of the pores
and the periodicity of the porous structures can be precisely
controlled and readily tuned by changing the size of the col-
loidal spheres. There is no doubt that a similar approach is
extendible to other materials. The only requirement seems
to be the availability of a precursor that can inltrate into
the void spaces among colloidal spheres without signi-
cantly swelling or dissolving the template (usually made of
Proof's Only
Self-Organization of Colloidal Nanoparticles 7
Figure 8. (A) Scanning electron micrograph (SEM) of a PS colloidal
crystal. (B) Low magnication SEM of a single particle of 3DOM
titania. (C) High magnication SEM of 3DOM silica showing close-
packed macropores, which are interconnected through smaller windows.
The white regions are walls of the rst layer, the gray regions are walls
of the second layer, and the black regions are voids. (D) Transmis-
sion electron micrograph (TEM) of 3DOM silica, which has amorphous
walls. The dark regions are the silica walls. Reprinted, with permission,
from [160], A. Stein and R. C. Schroeden, Curr. Opin. Solid State Mater.
Sci. 5, 553 (2001). 2001, Elsevier Science.
polystyrene beads or silica spheres). Some gaseous precur-
sors have also been employed in this process, albeit the
initial product deposited on the surfaces of samples might
block the ow in of gaseous precursors [175]. Presently, the
smallest colloidal spheres that have been successfully used in
this method are 35 nm in diameter [167]; the lower limit to
the particle size that can be incorporated into this technique
has not been completely established.
In another article, fabrication of germanium inverse opals
by a combination of wet-chemical and gas-phase synthe-
sis has been reported by Meseguer et al. [177]. Ge is
transparent to infrared radiation (below 0.67 eV at room
temperature) and it presents an extremely high refractive
index contrast (n = 4). This makes Ge inverse opals one
of the most promising PBG material made out of a semi-
conductor. It was reported that the photonic bandgap mate-
rial formation by the classical inltration method, using
hydrolysis followed by a reduction and the use of chemi-
cal vapor deposition (CVD) technique, increased the homo-
geneity of the germanium layer. In the hydrolysis route,
a germanium precursor (Ge(OCH
3
)
4
) was completely inl-
trated in the opal voids (in a template made with 870 nm
silica particles), followed by a hydrolysis process to obtain
GeO
2
. Figure 9(A) shows opal voids impregnated with small
(100 nm) GeO
2
microcrystallites. Ge was obtained by a
reduction process at 550

C. To increase the inltration per-
centage, the opals are subjected to multiple cycles of hydrol-
ysis followed by a reduction process. The Ge inltrated opals
were then chemically etched to remove silica particles with
the same method used for the other inverse opals. In this
way, we seek to remove the SiO
2
spheres from the compos-
ite and obtain Ge inverse opals. Figure 9(C) shows SEM
images of the cleft edge of a Ge inverse lattice. In a fur-
ther improvement of this classical method, the authors also
reported the use of CVD for the inltration of silica tem-
plates made from slightly larger spheres (1200 nm). The ger-
manium precursor, Ge
2
H
6
gas, was at rst inltrated into
the template placed in a high vacuum cell (510
6
torr), and
then, the sample cell was cooled by using liquid nitrogen,
(A) (B) (C)
Figure 9. SEM images of an opal at the different stages of Ge inl-
tration. (A) shows a GeO
2
-silica composite after hydrolysis reaction.
(B) is the Ge-silica that results from the reduction process. After the
etching process, one obtains the inverse lattice (C). The opal template
is formed from 870 nm silica particles and then sintered at 950

C for
3 h. Reprinted, with permission, from [177], F. Meseguer et al., Coll.
Surf. A. 2002, Elsevier Science.
which gave rise to the condensation of the precursor into
the opal void volume. Thereafter, upon heating the reactor
up to the decomposition temperature of Ge
2
H
6
, microcrys-
talline germanium were formed in the pores, as shown in
Figure 9. By this method, the authors claim that the Ge ll-
ing fraction can be controlled up to 100% of the template
pore volume. The inltrated samples were later chemically
etched to obtain a germanium inverse opal with long-range
FCC order of the original template.
Recently, Yang et al. demonstrated a simple and conve-
nient method for fabricating hierarchically porous materials
by combining colloidal arrays and block copolymers into a
single system [176]. In summary, dewetting of colloidal dis-
persions from contoured surfaces, attractive capillary forces,
and geometric templating can be combined to provide an
attractive method of assembling colloidal particles into com-
plex aggregates. The success of this organization approach
depends on the control of a variety of parameters like the
meniscus of the curvature of the rear edge of the liquid slug,
concentration, and surface charges on the colloidal particles
and the templates. However, though it has not been conclu-
sively proved yet, this method may be limited to a few tens
of nanometer regions due to the interaction of the Brownian
motion and the capillary forces. One such example of the
use of physical templating and capillary forces to assemble
monodispersed spherical colloids into uniform aggregates
with well-controlled sizes, shapes, and structures is shown in
Figure 10 (and also Table 1). When an aqueous dispersion of
colloidal particles is allowed to dewet from a patterned solid
surface with appropriate relief structures, the particles get
trapped and assemble into aggregates whose structures are
determined by the geometric connement provided by the
templates. By this method, the capability and feasibility of
assembling polystyrene beads and silica colloids (150 nm in
diameter) into complex aggregates which include polygonal
or polyhedral clusters, linear or zigzag chains, and circular
rings, have been demonstrated and a couple of examples are
shown in Figures 11 and 12.
Immobilization of metal colloid particles on a suitable
structured matrix using a self-assembly technique is one of
the promising strategies for the construction of colloidal
nanostructures [92, 179]. Assembling a homogeneous dis-
persion in a well-dened 3D nanostructure offers an alter-
native route to these stepwise 3D structure formations
[180181]. Zeolites and porous membranes (like porous
silicon, anodized alumina, etc.) have also been used as
Proof's Only
8 Self-Organization of Colloidal Nanoparticles
Figure 10. SEM images of polygonal aggregates formed by templat-
ing PS beads against 2D arrays of cylindrical holes (2 m diameter).
(A) 2D array of dimers formed from 1.0 m PS beads; (B) 2D array of
trimers formed from 0.9 m PS beads; (C) 2D array of square tetramers
formed from 0.8 m PS beads; and (D) 2D array of pentagons formed
from 0.7 m PS beads. Reprinted, with permission, from Y. Yin et al.
J. Amer. Chem. Soc., 2001, American Chemical Society.
templates for the growth of nanoparticles inside the pores
[182]. One such example of three dimensionally arranged
gold colloids has been prepared in the nanoscale pores
of alumina membranes [181]. Immobilization of the col-
loid particles performed by vacuum incorporation were
achieved by a modication of the inner walls of the porous
alumina substrate with alkoxysilanes Y-(CH
2
)
x
Si(OR)
3
and Y-(CH
2
)
x
SiR(OR)
2
bearing suitable functional groups
(Y = NH
2
, SH), followed by anchoring of the colloids by
self-assembly onto the modied surface. Well-washed mem-
branes are dark-red in color upon incorporation of gold
colloids, and a schematic description of an alumina mem-
brane containing immobilized gold is shown in Figure 13.
A part of the sample showing the particles located only near
Table 1. Control of the Self-Assembled Particles By Using Cylindrical
Templates of Different Structure of Clusters.
Geometric shape of the template Structures of the cluster D/d
1.00-2.00
d
D
2.00-2.15
2.15-2.41
2.41-2.70
2.70-3.00
3.00-3.30
Reprinted, with permission, from [178], Y. Yin et al. J. Amer. Chem. Soc. 123,
8718 (2001). 2001, American Chemical Society.
Figure 11. SEM images of 2D arrays of colloidal aggregates which were
assembled under the connement of templates etched in the surfaces of
Si (100) substrates; (A) 800-nm PS beads in square pyramidal cavities
1.2 m wide at the base; (B) 1.0 m silica colloids in square pyramidal
cavities 2.2 m wide at the base; (C) 0.8 m PS beads in V-shaped
grooves 2.5 m wide at the top; and (D) 1.6 m PS beads in V-shaped
grooves 10 m wide at the top. Note that the use of V-shaped grooves as
the templates also allowed one to control the orientation of the colloidal
crystals. In (C) and (D), the face-center-cubic structures have a (100)
orientation rather than (111), the one that is most commonly observed
when spherical colloids are crystallized into 3D lattices. The arrows
indicate defects, where one can also see the colloidal beads underneath
the rst layer of the structure. Reprinted, with permission, from [178],
Y. Yin et al., J. Amer. Chem. Soc. 123, 8718 (2001). 2001, American
Chemical Society.
Figure 12. (A) AFM image of the template (a parallel 2D-array of
trenches that were 150 and 150 nm in width and depth, respectively),
which was fabricated using near-eld optical lithography. (B) SEM
image of the linear chains that were formed by templating 150-nm PS
beads against the trenches shown in (A). These PS beads represent the
smallest objects that have been successfully incorporated in the TASA
process. Reprinted, with permission, from [178], Y. Yin et al., J. Amer.
Chem. Soc. 123, 8718 (2001). 2001, American Chemical Society.
Proof's Only
Self-Organization of Colloidal Nanoparticles 9
Immobilized colloidal
metal particle
Aluminum oxide
membrane
Figure 13. Model structure of a colloid-charged porous alumina mem-
brane. Reprinted, with permission, from [181], T. Hanaoka et al., Eur.
J. Inorg. Chem. 807 (1998). 1998, Wiley-VCH.
the pore walls is shown in Figure 14(A), which is schemat-
ically explained in Figure 14(B). The particles are closely
packed, vertically overlapped, and generally observed along
the walls.
3.3. Electrostatic Complexation
of Nanoparticles with Charged
Langmuir Monolayers
The air-water interface has been recognized to be an
excellent media for organization of inorganic cations using
charged amphiphilic monolayers (Langmuir lms) [183].
This approach has been extended to assemble large inorganic
anions [184185] and biological macromolecules [186187].
Colloidal particles can apparently be trapped at the liq-
uid surface as a result of the electrostatic and surface
tension forces [188191]. The interaction between the
ions/macromolecules in solution and charged lipid Langmuir
monolayers drives the organization at the air-water interface.
The nanoparticles organized in this fashion by the Langmuir
Blodgett (LB) method can then be transferred into appro-
priate substrates, for example, carboxylic-derivatized silver
nanoparticles can be immobilized at octadecylamine (ODA)
Langmuir monolayer surfaces [192194].
The behavior of particle monolayers at liquid surfaces has
also received considerable attention of late but was already
studied some 30 years ago [195196]. Kumaki reported
an interesting work on monomolecular PS nanoparticles
(A) (B)
Figure 14. (A) TEM image of gold colloids in the pores. Both the upper
and lower walls were removed by sectioning; (B) schematic represen-
tation illustrating the situation. Reprinted, with permission, from [181],
T. Hanaoka et al., Eur. J. Inorg. Chem. 807 (1998). 1998, Wiley-VCH.
(50 nm diameter) monolayers and discussed the possi-
ble signicance of the surface pressures on the organiza-
tion of these particles [197]. However, application of the
LB technique to colloidal particles led to submonolayers
with coverage less than 80% [198]. Only recently, Bardosova
et al. [199] reported on the successful LangmuirSchaefer
transfer of monolayers of silica particles modied with
3-aminopropyl groups.
The formation of 2D arrays by self-assembly on solid
supports make use of numerous strategies [200203], for
example, a 2D array of colloidal spheres can be formed
at the air-liquid interface which can then be subsequently
transferred onto a surface of a solid substrate. The surfaces
of these colloidal spheres have to be modied such that
they will only be partially immersed into the surface of a
liquid after they have been spread onto the air-liquid inter-
face through a spreading agent (usually an alcohol) [189].
It is the strong attractive interactions (e.g., those between
dipoles induced by the asymmetric interface) among the col-
loidal spheres that lead to the spontaneous formation of a
2D aggregate at the interface. The morphology of the aggre-
gate usually exhibits fractal characteristics, but it can also
be changed by varying a number of parameters such as the
size, the number concentration, the surface hydrophobicity,
or the charge density on the colloids, and the electrolytic
properties of the underlying liquid [189]. In a recent demon-
stration, for example, Kondo et al. were able to generate 2D
arrays of silica colloids (1 m in diameter) with relatively
large domain sizes by controlling the degree to which the
silica colloids were immersed into the liquid surface [204].
Deckman et al. [205], Lenzmann et al. [206], and Fulda and
Tieke [207] applied the LB lm technique in this fashion
to obtain polycrystalline 2D arrays of polymer latexes over
areas as large as several square centimeters.More recently,
Burmeister et al. also demonstrated a similar technique
that was capable of forming ordered 2D arrays of colloidal
spheres on various types of substrates [208].
Another method that has been largely explored by
Nagayawa, Picard, Rakers, et al. [202203, 209213], uses
the attractive capillary forces among colloidal spheres to
organize them into a hexagonal 2D array in a thin lm of
liquid supported on a at substrate. In a typical experiment,
a liquid dispersion of colloidal spheres is spread onto the
surface of a solid substrate. When the solvent slowly evap-
orates under a controlled condition, these colloidal spheres
are self-assembled into a closely packed, hexagonal array,
which has also been followed experimentally by using an
optical microscope [202, 214]. They found that a nucleus
an ordered region that consists of a number of colloidal
sphereswas rst formed when the thickness of the liq-
uid layer approached the diameter of the colloids. More
colloids were driven toward this nucleus by a convective
transport, and eventually organized around the nucleus due
to the attractive capillary forces. A at, clean, and chemi-
cally homogeneous surface has to be used in order to gen-
erate a highly ordered array with relatively large domain
sizes. Solid substrates, such as glass slides or silicon wafers,
have been used in this technique. Lazarov et al. have also
explored the use of liquids such as preuorinated oil (F-oil)
or mercury as the substrates in forming highly ordered 2D
arrays of colloidal spheres [215]. At the air-water interface
Proof's Only
10 Self-Organization of Colloidal Nanoparticles
under appropriate passivation conditions, mid-nanometer-
sized gold particles can undergo self-organization into
densely packed monoparticulate lms [216]. These lms,
when transferred onto hydrophilic Formvar-coated copper
grids by vertical retraction of immersed substrates, pro-
duced monoparticulate lms with variable extinction and
reectance properties, related to the hexagonal close-packed
arrays that form in a micron length scale. The extinction
bands of these arrays shifted by hundreds of nanometers
to near-IR wavelengths and broadened enormously with
increasing periodicity. Large particle arrays also demon-
strated extremely high SERS, with enhancement factors
>10
7
, which were correlated with increasing periodicity.
A slight modication of this technique has been reported
by Goldenberg et al. [217] following ideas of Aveyard et al.
[218], which seems an interesting method for practical appli-
cation in building up 2- and 3D lattices of spheres. This
simple and fast technique utilizes the nonmiscibility of liq-
uids and trapping particles between two liquids of differ-
ent viscosity so as to make a monolayer of the lm in
the liquid interface, which can be suitably transferred onto
an appropriate substrate. Goldenberg et al. utilized this
technique to organize hexagonally ordered monolayers of
hydrophilic micrometer size PS and silica particles at water/
alkane interfaces, and to transfer them onto solid substrates.
These monolayers showed good diffraction properties as
observed by optical microscopy with a Bertrand lens [217].
A schematic representation of this method is shown in
Figure 15(A), while a photo of actual latex particles oat-
ing on hexane is shown in Figure 15(B); a typical lm so
formed is shown in Figures 15(C) and 15(D) and the col-
lected particles on the glass in Figures 16(A) and 16(B). The
hexane
water
ethanolic
suspension
glass substrate
monolayer
monolayer on glass
substrate
A
C D
B
Figure 15. Scheme 1. Schematic description of the particle monolayer
formation at the water-hexane interface: (A, B) spread of particle from
ethanol dispersion on the water-alkane interface; (C) formation of the
oating hexagonal arrays of particles: (D) transfer of particle onto solid
substrate. Reprinted, with permission, from [217], L. M. Goldenberg
et al., Langmuir, 18, 5627 (2002). 2002, American Chemical Society.
A
B
Figure 16. Optical microscopic image of a monolayer of 1.4 m
polystyrene-2,3-epoxypropyl methacrylate (PS-EPMA) (A), and (B) sil-
ica (1.7 m) particles transferred from the water/hexane interface.
Inset: corresponding diffraction images with Bertrand lens (right insert)
and 2DFFT transformation (left insert, 512 512 pixels). Reprinted,
with permission, from [217], L. M. Goldenberg et al., Langmuir 18,
5627 (2002). 2002, American Chemical Society.
actual experimental condition of the PS particles oating in
the water-hexane interface is shown in Figure 17(A), while
17(B) shows the long-range order in these monolayer lms.
These methods or a supplement of other methods has
been used to make 3D arrays of colloidal nanoparticles.
Sedimentation in a gravitational eld is the simplest
approach that has been used for the formation of 3D crys-
talline arrays from colloidal spheres [219222]. Although
it looks simple, the method involves several complex
A B
Figure 17. (A) Polystyrene particles spread on a water hexane interface,
and (B) the successive lm that can be collected from this partially
covered layer on a glass slide.
Proof's Only
Self-Organization of Colloidal Nanoparticles 11
processes such as gravitational settling, translational diffu-
sion (or Brownian motion), and crystallization (nucleation
and growth). In addition, these force parameters need to be
properly controlled as well as the size and density of the col-
loidal spheres and the sedimentation rate to succeed in mak-
ing 3D colloidal lattices. The colloidal spheres can always
settle completely to the bottom of a container as long as the
size and density of these spheres are sufciently high. Only
when the sedimentation process is slow enough, the colloidal
spheres concentrated at the bottom of the container will
undergo a hard-sphere disorder-to-order phase transition to
form a three dimensionally ordered lattice [223224].
Monodispersed silica colloids are most commonly
employed in sedimentation due to the high density of amor-
phous silica. Opalescent structures (usually referred to as
synthetic or articial opals) have been obtained from these
colloidal materials under carefully controlled conditions
[225229]. Two types of crystalline arrays of colloidal spheres
(or colloidal arrays) have been extensively studied: the rst
type includes fcc lattices formed from highly charged col-
loidal spheres and their volume fractions of colloids are
often less than 10%; the second type is a Cubic-close-packed
(ccp) structure (also a Face-centered-cubic (fcc) lattice) and
its volume fraction of colloids is always close to 74%. It is
generally accepted that the 3D crystalline arrays produced
by this method have a ccp structure (or a fcc lattice with
a packing density of 74%) similar to that of a natural
opal [229230]. The preference of a ccp structure over a
hexagonal-close-packed (hcp) one has been suggested to be
a result of the difference in entropy between these two struc-
tures [231]. Recently, van Blaaderen et al. demonstrated
the use of lithographically dened surfaces as templates to
grow 3D crystalline arrays with desired spatial structures
[232]. This process is the mesoscopic equivalent of epitaxial
growth: highly ordered and well-controlled arrays as large
as 1 cm
3
could be generated. Jiang et al. also developed
a layer-by-layer sedimentation method for fabricating ccp
arrays of silica colloids; these arrays have a tightly controlled
number of layers along the [111] direction [233].
3.4. Chemically Patterned Substrates
Stripes and channels with alternating wetability can also
be produced by using wetting instabilities during the LB
lm transfer that has long been used for building up thin
organic coatings of monomolecular layers (monolayers) onto
solid substrates [234]. For example, Gleiche et al. [235] used
monolayers of L-o-dipalmitoyl-phosphatidycholine (DPPC)
on mica to generate a structured surface with a channel lat-
tice exhibiting a high wetability contrast forming a lateral
structure, which can be obtained by rapidly withdrawing a
mica substrate (1,000 ms
1
) at a low monolayer surface
pressure and constant temperature. Under these conditions,
the lm adsorption becomes unstable, leading to periodic
interruptions in the molecular deposition which generate
regular hydrophilic channels of about 200 nm in width and
are separated by hydrophobic stripes of monomolecular
height and latitude of about 800 nm. This structured sur-
face can be used as a template to deposit materials along
the channels with high adsorption selectivity by making use
of the different wetting behavior of the hydrophilic chan-
nels and the hydrophobic stripes, for example, by either
an anisotropic wetting/dewetting process (determined by the
surface structure) or by using capillary (capillary lling) and
electrostatic forces.
Gold clusters (Au55) stabilized by an organic ligand shell
dissolved in 1-phenyloctane were dropped on the structured
mica surface which lled the channels with Au55 clusters as
shown in Figure 18. Mostly single, large-cluster aggregates
(almost 4 nm in height) were lined up in the channels, but
some two-layers were also noticed (see spike structure in
Figure 18(B)). In contrast, the DPPC monolayer is virtually
not wetted and only a few cluster aggregates are located on
the top of the stripe region, resulting in a nearly perfect
selective adsorption.
Self-organization on chemically patterned surfaces can
be achieved in a rather straightforward way by the use
of charged surfaces, as the nanoparticles in suspension
are most often charged to prevent coagulation. Substrates
are patterned by molecular layers in mesoscale dimen-
sions. Substrates for such chemisorption are usually gold
A
B
Figure 18. Dynamic scanning force microscopy (SFM) image of liquid-
deposited Au55 metal cluster. (A) topography of the selectively
adsorbed organic-ligand-stabilized metal cluster. Cluster aggregates
(bright spots) were aligned along the channels even when scanning the
previously wetted area. (B) the spikes depicted in the three dimension-
ally rendered selected area indicate the large height difference (image
size 6.5 6.5 m
2
) Reprinted, with permission, from M. Gleiche et al.,
Nature. 403, 173 (2000). 2000, Macmillan Magazines Ltd.
Proof's Only
12 Self-Organization of Colloidal Nanoparticles
surfaces regiospecically modied with functionalized thiol
molecules by microcontact printing [146] or silica surfaces
that have been photo-patterned by a preadsorbed cationic
monomer (as shown in Figures 19 and 20) [236]. In these
cases, the contrast between covered and uncovered areas,
as well as the packing density within the covered regions,
depends strongly on the electrochemical parameters of
the system, such as ionic strength, pH, and effective sur-
face charge. Moreover, only moderate packing density is
achieved due to the repulsive forces between the particles
[237]. One way to overcome this problem is the applica-
tion of cationic surfactants that increase the charge shielding
and, due to adsorption onto the particle surface, introduce
attractive forces such as van der Waals interaction. Recently,
Chen et al. demonstrated the validity of this concept [145].
The site-selective assembly of colloidal polymer particles
onto laterally patterned silane layers has also been exten-
sively studied for the assembly of colloidal nanoparticles at
mesoscale dimensions [145146, 237239]. The structured
silane monolayers on silicon-oxide substrates are either fab-
ricated by liquid- or gas-phase deposition or a combination
of both methods, using different trialkoxysilanes with a pho-
tolithographic patterning technique. By using this method,
various types of surface functionalization, such as regions
A
B
C
D
Figure 19. Schematic representation of a chemically induced patterning of PS particles via microcontact printing: (A) rst, octadecanethiol (C18)
is stamped onto a native gold surface via an elastomer stamp, which had been previously dipped into ethanolic C18 solution for a few seconds and
dried in air; (B) next, the surface is covered with an aqueous thioglycolate (TG) solution for 1 h, to adsorb TG on the native gold areas; (C) after
rinsing and drying of the chemical pattern on the gold surface, it is exposed to a suspension of PS nanoparticles, which had been prepared according
to the standard procedure. After 1 h, the suspension is washed off with DI water and a patterned PS particle layer is obtained. (D) for analysis
the layer was coated with 20 nm-thick gold lm deposited via thermal evaporation. Reproduced, with permission, from [236], M. Himmelhoos and
H. Takei, Phys. Chem. Chem. Phys. 4, 496 (2002). 2002, Royal Society of Chemistry.
with amino functions next to areas of the bare silica sur-
face or positively charged regions of a quaternary ammo-
nium silane surrounded by a hydrophobic octadecylsilane
lm have been reported (as shown in Figures 21 and 22).
Often, photo-protected amino groups were, used which
allowed direct photo-patterning after monolayer prepara-
tion, leading to free NH
2
groups at the irradiated regions.
Depending on the assembly conditions (different surface
functionalization, pH, and drying conditions), the particles
can be directed selectively onto a specic surface region.
The regiospecic assembly of colloidal particles onto lat-
erally structured silane monolayers can generate complex
structures. The required monolayer patterns can be obtained
by a combination of the industrially well-established photo-
lithography process with the deposition of silanes from the
vapor and liquid phase onto technologically and highly rele-
vant substrates such as silicon with an oxide layer, glass, and
quartz. The particle adsorption process driven essentially
by electrostatic and polar interactions in combination with
adhesion inhibition at hydrophobic surface layers is a self-
controllable and highly parallel process which should allow
the simultaneous fabrication of many devices in one assem-
bly step with essentially no size limitations from nanometer
to the millimeter length scales.
Proof's Only
Self-Organization of Colloidal Nanoparticles 13
A
B
Figure 20. SEM micrographs of a patterning experiment by utilization
of an elastomer stamp with a square pattern as structure for print-
ing. (A) micrograph gives a survey, while that on (B) demonstrates the
particle density within one square. Note the onset of regular close pack-
ing at several locations within the square. Dimensions of the pattern:
squares (24 m)
2
, gaps 15 m. Reproduced, with permission, from
[236], M. Himmelhous and H. Takeir, Phys. Chem. Chem. Phys. 4, 496
(2002). 2002, Royal Society of Chemistry.
3.5. Optically Directed Self-Organization
At the current stage of development, all of the above-
mentioned methods are only capable of generating colloidal
arrays built up by small domains, and the largest single
domain usually contains fewer than 10,000 colloidal spheres
[240]. Most of these methods can only form 2D hexagonal
arrays in which the colloidal spheres are in physical con-
tact. As a result, it is very hard to independently vary the
lattice constant and the particle size. The approach based
on optical forces seems to have the potential to overcome
these difculties [241242]. In this method, colloidal spheres
are organized into a highly ordered 2D structure in a liquid
by creating an optical standing wave pattern having a reg-
A
C
D
B
Figure 21. Optical micrographs (difference inter ference contrast) of
the drying front of a latex suspension droplet on an ammonium silane
(squares)-octadecylsilane monolayer pattern before (A) and after (B)
removal of the liquid (inset) Schematic side view of a resting droplet on
the substrate. The SEM images show a particle submonolayer on the
ammonium pattern formed by electrostatic assembly in suspension (C
after rapid removal of the suspension liquid) and a colloid multilayer on
a similar pattern generated by capillary forces at the slow moving drying
from (A) and schematics in C and D show the individual particles of
the submonolayer and multilayer, respectively. octadecyltriethoxysilane
(OTE). Reprinted, with permission, from [237], U. Jonas et al., PNAS,
99, 5034 (2002). 2002, National Academy of Sciences, U.S.A.
ular array of intensity antinodes. The colloidal spheres are
then driven to the antinodes maxima by the optical forces.
Depending on how many laser beams are used to create the
standing wave, patterns as complex as a quasicrystal have
also been produced. As demonstrated by Misawa et al. and
by Mio and Marr, the method based on optical forces was
also capable of generating an arbitrary 2D or 3D pattern by
adding individual colloidal spheres to an array one at a time
[243244].
Several years ago, Ashkin et al. experimentally demon-
strated that small particles (25 nm to 10m) in a suspen-
sion could be trapped by a single-focused laser beam [245].
Following their work, many researchers have used opti-
cal trapping techniques (also known as optical tweezers)
to manipulate micrometer and submicrometer-sized objects
[246247]. Because of their nondestructive, sterile nature,
optical tweezers have found great success in manipulating
biological systems, including bacteria, viruses, chromosomes,
and DNA [248251]. More recently, Hayward et al. has
shown the usefulness of light-induced orientation in what
they called optically tunable micropatterns using ultravio-
let light illumination [252]. Though the mechanism is largely
unexplained, it could result from the light-induced changes
Proof's Only
14 Self-Organization of Colloidal Nanoparticles
A
B
Figure 22. (A) Optical micrograph of latex particles assembled onto an
NVoc silane monolayer after photo-deprotection with an argon laser
(364 nm) through a gold mask on quartz (dark squares were protected
from light). (B) Particle assembly (pH 4.5) on a mixed monolayer of
the NVoc and ammonium silane after irradiation (similar to A). Bright
squares were protected from light and show a lower particle density in
the SEM. (C) Fluorescence micrograph of a mixed monolayer (NVoc
and ammonium silane, similar to B) after irradiation and staining with
a uorescent probe (Alexa Fluor 488). Reproduced, with permission,
from [237], U. Jonas et al., PNAS 99, 5034 (2002). 2002, National
Academy of Science.
on the wetting behavior of the substratea mixture of opti-
cal illumination for chemical patterning [253].
3.6. Electric or Magnetic Field-Assisted
Self-Organization
Application of an electric eld results in structural transi-
tions in the colloidal suspension because the interparticle
electrostatic interactions due to polarization are stronger
than Brownian forces. The tendency of particles in suspen-
sion to form structures such as chains upon application of an
electric eld was reported centuries ago by scientists such as
Priestly and Pohl [254255]. Quantitative experiments on the
electrorheological effect were rst performed by Winslow in
1949, when he reported that suspensions of silica gel parti-
cles in low-viscosity oils tend to brillate upon application
of electric elds, with bers forming parallel to the eld
[256]. It was reported that at elds larger than 3 kV/mm,
the suspensions behaved like a solid, which owed like a
viscous uid above a yield stress that was proportional to
the square of the applied electric eld. A recent and com-
prehensive survey of electrorheological (ER) uids, where
the issue of particle aggregation is addressed, is provided by
Parthasarathy and Klingenberg [257]. At low elds, below
100 V/mm, where the interparticle electrostatic interac-
tion energies are low compared to thermal energies, no sig-
nicant particle association are generally observed and the
nanoparticles tend to sediment to the bottom electrode, if
their density is larger than the solvent density. However, at
eld strengths 1000 V/mm, eld-induced structures, such
as chains of touching particles, are formed that do not break
up due to thermal uctuations, implying that the electro-
static energy at contact was many times kT [258].
As an example of the ordering of nanoparticles onto
structured surface, Kumacheva et al. [259] has reported
an experiment with poly(methylmethacrylate) (PMMA)
spheres electrodeposited onto Indium tin oxide (ITO)-
covered substrates. The speed of colloid crystal growth can
be controlled via electrodeposition parameters such as volt-
age and time, and electrophoretic mobility of the parti-
cles, thus control over layer-by-layer deposition is possible.
They further reported that the particles get into ordered
phase after reaching the electrode, which they attributed to
the reorganization of the microspheres in the grooves by
squeezing the newly arriving spheres between the already
deposited particles, and followed by synergistic particle rear-
rangement. This resulted in large-scale particle ordering,
as is shown in Figure 23(D) for assembly of the colloidal
spheres in 4.2 m-thick grooves.
The use of nonuniform electric elds for the manipula-
tion of m-sized particles are reasonably well-documented
A B C D
Figure 23. SEM images of the colloidal arrays of latex spheres of
PMMA electrodeposited on nonpatterned (A) and patterned (BD)
ITO-coated surfaces. The widths of the grooves are 5.5 m (B) and
4.2 m (C, D). The SEM images of the colloidal arrays with the width
4.2 m were cropped in (A) and (B). Scale bars are 1 m (AC) and
10 m (D). Reprinted, with permission, from [259], E. Kumacheva
et al., Adv. Mater., 14, 221 (2002). 2002, Wiley-VCH.
Proof's Only
Self-Organization of Colloidal Nanoparticles 15
[260268]. A variety of biological applications including
collection, fusion, and separation of biological cells [269],
immobilization of DNA [270], and a collection of viral
particles [271] have been demonstrated. Dielectrophoresis
(DEP) has been used in the assembly of a biosensor by con-
centrating o-protein-coated latex microspheres in the area
between two oppositely charged electrodes [272]. Alignment
of nanospheres [273] and dielectrophoresis for the trapping
of submicron latex spheres inside a 3-D, multielectrode array
[263], and alignment of metallic needle-shaped particles with
diameters between 70350 nm and a few m in length
using interdigitated electrodes has also been reported; an
example is shown in Figure 24 [274]. The possibility of,
contact-free handling for assembly and mechanical pro-
cessing of microbeads has been demonstrated [275]. Today
DEP has not yet been sufciently exploited for the self-
organization of nanoparticles, though the fundamental prin-
ciples of DEP are reasonably well-understood. This is an
attractive method for the self-organization of nanoparticles
and will almost certainly gain more importance in the future
works for the self-organization of nanoparticles.
The electric eld forces on particles can be controlled by
monitoring the charge on the colloidal particles (through the
control of the double layer). Recent work within this area,
has shown that the electric eld is more than just a substitute
for gravity, and the formation of highly ordered colloidal
Figure 24. (a) Schematic presentation of the 3D particle organization
achieved via DEP. The rst particles on the plane are pushed to the cen-
ter of the electrode setup (negative DEP) and form an ordered struc-
ture. Particles that later arrive in the vicinity of the electrodes deposit on
the top of the 3D structure, creating additional layers. (b) A freestand-
ing, pyramid-shaped structure of silica particles is shown as a result of
dielectrophoretic organization. (c) The structure collapses immediately
after the electric eld is switched off. (d) SEM image of a 3D, free-
standing formation of silica particles in the area between two long and
parallel electrodes. Reprinted, with permission, from [216], A. Docoslis
and P. Alexandris, Electrophoresis. 2002, Wiley-VCH.
layers has been reported for electrophoretic deposition of
micrometer-sized silica and polystyrene particles [277279],
and nanometer-sized gold particles [277, 280281]. The self-
organization phenomenon was explained by electro-osmotic
or electro-hydrodynamic microows around the particles,
which are induced by the electric eld and lead to attractive
forces between the particles at the surface [277, 282]. While
the exact mechanism of ordering is still open to debate, it is
clear that improved ordering results from the fact that the
colloidal particles are only weakly bound (physisorbed) to
the substrate and remain fairly mobile. Only after an addi-
tional xing step (e.g., application of a high eld), the col-
loidal lm is anchored to the substrate. Based on these ideas,
it has been proposed that highly ordered multilayers of col-
loidal particles could be grown by using dc or low frequency
ac elds (one such example is shown in Figure 25 [277]).
Magnetic elds can, in principle, be used to direct parti-
cles onto a substrate, making use of the motion of magnetic
particles in magnetic eld gradients [283], and there have
been reports where magnetic eld forces have been used
to direct the colloidal crystal formation [284]. In a recent
publication, Bizdoaca and Lin et al. have reported on the
fabrication of grids of micrometer-sized core-shell particles,
utilizing magnetophoretic deposition techniques in a water-
based colloidal suspension with core-shell type spherical
polystyrene particles of 640 nm diameter core, covered with
a shell of ve layers of 12 nm diameter Fe
3
O
4
nanocrystals.
The separation and length of the individual chains were
tuned by the magnetic eld [285286]. Magnetic eld forces
may play an important role in the organization of colloidal
nanoparticles into ordered arrays most likely in conjunction
with electric eld forces.
3.7. Biologically Assisted
Nature allows structural building blocks to hierarchically
organize into an ensemble with an atomic order resolution,
which are formed spontaneously with high accuracy and
minimum energy similar to the bottom-up self-organization
A B
Figure 25. (A) Large single domain of the opal made of 300 nm silica
spheres. The long-range ordering can be observed over 100 m. The
sample for the SEM observation was derived by carefully removing one
substrate of the sandwich-like cell. (B) Hexagonal arrangement of the
silica spheres with excellent arrangement was observed. It was found
that the domain size exceeds 100 mm. The hexagonal arrangement was
assigned to the close-packed surface of a face-centered cubic or hexag-
onal close-packed structure. It is interesting to note that opaline mate-
rials synthesized by conventional sedimentation methods show both the
hexagonal and tetragonal arrangements. Reproduced, with permission,
from [230], H. Miguez et al., Langmuir 13, 6009 (1997). 2002, Amer-
ican Institute of Physics.
Proof's Only
16 Self-Organization of Colloidal Nanoparticles
process [287]. On the contrary, it is quite difcult for
chemists to achieve the total synthesis and structural control
of macromolecules by covalent chemical synthesis methods
only [288]. Nobel Laureate Jean-Marie Lehn used the words
of Leonardo da Vinci Where nature nishes producing its
own species, man begins, using natural things and with the
help of this nature, to create an innity of species, while,
writing about the perspectives on the future and perspec-
tives of supramolecular chemistry [289]. Natural evolution
has led to highly functional assemblies of proteins, nucleic
acids, and other (macro)molecules which perform compli-
cated tasks that are still difcult to emulate. For example,
the 20-nm ribosome particle is an effective supramolecu-
lar machine which spontaneously self-assembles from more
than 50 individual protein and nucleic acid building blocks,
thereby impressively demonstrating the power of biologically
programmed molecular recognition [290292]. The bacte-
rial agellar motor, which is 30 nm in diameter, resembles
the electric rotary motor [293295], while muscle proteins
resemble a linear motor, but the engine size is approxi-
mately 30 nm [296]. Assemblies of these proteins are self-
assembled molecular machines that are small in size with
highly efcient energy transduction properties. These char-
acteristics make proteins good solutions for constructing
sophisticated devices on the nanometric scale.
Protein molecules have the ability to build nanometer-
sized supramolecules by self-assembly, which can them-
selves create higher order structures like cell components.
Some of the schemes that attach protein molecules are
illustrated in Table 2. Mirkin et al. [75] and Alivisatos
et al. [297] have showed that complementary DNA oligonu-
cleotides offer possibilities for self-assembled nanoparticles.
Protein-based conjugation offers various advantages like a
large number of complementary systems with a wide range
of free-association energies are synthetically available, and
the systems are well known in biochemistry and molecu-
lar immunology which can be handled by standard pro-
cedures [298299]. From a materials science point-of-view,
Table 2. Various Methods to Attach Protein to Self-Assembled Surfaces
Functional Side Group Available Surface
on Peptide Derivatization Type of Binding
Natural
-COOH (carboxylic acid) Amino Electrostatic covalent
Asp amide (after carboxy
activation)
NH
2
(amino) Carboxylic acid, Electrostatic
Lys, Gln, Arg active ester, covalent amide
epoxy
-SH (thiol) Maleimide Covalent thio-ether
Cys
-OH Epoxy Covalent ether
Ser, Thr
Synthetic
His-Tag Ni-NTA Complex Coordination complex
Strep-Tag Strep-Tactin, Supramolecular
Streptavidin complex
Biotin Streptavidin Supramolecular
complex
polypeptides offer many advantages over conventional syn-
thetic polymers due to their inherent ability to adopt sta-
ble conformations and self-assemble into precisely dened
structures that allow unprecedented control over materi-
als morphology and properties. Synthetic strategies for the
preparation of polypeptides can be divided essentially into
three classes: ring-opening polymerization, solid-phase syn-
thesis, and protein engineering. Recent advances in these
methods have allowed the preparation of (poly)peptides and
(poly)peptide hybrids, which can be assembled in a con-
trolled fashion into supramolecular architectures and mate-
rials that mimic the structure and function of proteins [300].
Right from the discovery of the double-helix structure
of DNA, biology has evolved from a purely descriptive
and phenomenological discipline to exact science in the
molecular regime, and recombinant DNA technology has
brought insights into the basic principles of many biochem-
ical processes and is also shaping developments in modern
biotechnology. Today, we can genetically engineer relatively
simple bacterial cells, and hopefully, in the future, we will be
able to tailor complex organisms. In view of such revolution-
ary developments, scientists and researchers have reported
wide-ranging developments in the fusion of biotechnology
with nanoparticles, since biomolecular components have typ-
ical size dimensions in the range of about 5 to 200 nm [301].
The concept of DNA hybridization-based self-organization
of molecular compounds has been applied to the assem-
bly of DNA-derivatized gold colloids [75, 297, 302]. Such
dened arrangements of nanocrystal metal clusters (quan-
tum dots) have been investigated for their novel physical
properties and possible applications in the eld of opto-
electronic technology [303306]. For this, superlattices or
quantum dot molecules are necessary, in which crystallites
from different materials are spatially assembled to tailor the
properties. Recently, Takahagi et al. [307] described the fab-
rication of 3D particle structures based on DNA hybridiza-
tion. The results indicate that DNA hybridization is a very
useful method for assembling nanoparticles into a 3D nano-
structure. To fabricate this structure, two colloidal gold sus-
pensions containing particles of 9 nm and 20 nm in diameter,
and oligonucleotides that are functionalized by heptanethiol
at their 5

termini were used by the authors. The thiol


group at the end of the oligonucleotides was adsorbed to
the particle surfaces by a chemical bond upon mixing with
the colloidal gold suspension, resulting in oligonucleotide
monolayers on the particles. The 9-nm particle suspensions
were mixed with thiol-5

-GGTCGGCACA-3

, and the 20 nm
particle suspension was mixed with thiol-5

-TGTGCCGACC-
3

. After keeping for 48 h at room temperature, the two mix-


ture suspensions were blended and diluted at pH 7. A 3D
structure has formed with alternating superposition of the
big and small particles. Though the internal structure can-
not be observed with SEM, the 20 nm gold particles can be
seen to exclusively connect with the 9 nm particles on their
surfaces. These results suggest that the internal structure of
DNA-linked particles might consist of alternating small and
large particles, and 3D particle networks had formed.
The highly specic recognition properties of antibod-
ies and antigens can be used to assemble a wide range
of nanoparticle-based structures with specic cross-linking,
compositions, and macroscopic architectures; for example,
Self-Organization of Colloidal Nanoparticles 17
IgE and IgG antibodies have been used to attach on sur-
faces followed by interparticle conjunction in the presence
of bivalent antigens with appropriate double-headed func-
tionalities [297]. Similarly, streptavidin/biotin cross-linking
has been used to bind nanocrystals onto a substrate [299].
Some protein-supramolecules have the ability to accom-
modate inorganic materials and this process is called bio-
mineralization. Their surfaces are thus designed to sequester
inorganic materials into composites like seashells or bone
[308]. This biomineralization process has been used to make
proteins that accommodate nanometer-size inorganic quan-
tum dots [300]. As proteins are more easily damaged or
destroyed compared to inorganic materials, there are sev-
eral potential means to eliminate the protein moiety from
an assembled array by heat treatment and UV irradiation,
or through other chemical means.
Combining the abilities of proteins to self-assemble, orga-
nize materials, and then be selectively eliminated, it is
expected that nanometer-size inorganic materials can be
positioned by protein supramolecules into nanostructures.
The protein, therefore, acts as a temporary scaffold, which
determines the nature of the nal structure without being
part of it. As a feasibility study of this method, the ferritin
molecule assembly was employed to make an array of inor-
ganic nano-dots suitable for a quantum electronic device.
Several methods have been reported to obtain arrays of fer-
ritin molecules and it has become possible to make large
2D crystals or arrays of ferritin at an air-water interface
[309313].
Another process has been proposed which exploits
protein molecules as scaffolds for producing inorganic, func-
tional nanostructures on a at surface. An array of con-
ductive FeO dots with a diameter of approximately 6 nm
was fabricated and a hexagonally packed array of ferritin
molecules was formed at an air-water interface by self-
assembly. Following this, the ordered ferritin was transferred
onto a modied Si surface and then the protein shell of the
ferritin molecule was eliminated by 450

C heat treatment
under nitrogen, which resulted in an ordered array of con-
ducting nano-dots bound to the Si surface [314]. Combining
this result with the fact that ferritin molecules have the abil-
ity to accommodate various metals and metal complexes, it
is possible to organize nano-dot arrays suitable for quantum
electronic devices by the so-called bio-mimetic Bio Nano
Process method [315321]. As the ferritin molecules are
produced from the same gene, the cavity size of each apo-
ferritin molecule is identical and hence, the nanoparticles
so produced are expected to be identical in sizea key
requirement in producing quantum opto-electronic devices.
Combined with the self-assembly of the protein, it may
be possible to build more complex functional inorganic
nanostructures on a at surface than simple arrays.
Nanoscale-ordered arrays of metal and semiconductor
quantum dots have been fabricated by binding presynthe-
sized nanoparticles onto crystalline protein templates made
from genetically engineered hollow double-ring structures
called chaperonins [322]. Using structural information as a
guide, a thermostable recombinant chaperonin subunit was
modied to assemble into chaperonins with either 3 nm or
9 nm apical pores surrounded by reactive thiols. These were
crystallized into 2D templates up to 20 m in diameter. The
periodic solvent-exposed thiols within these crystalline tem-
plates were used to size-selectively bind and organize either
gold or CdSe-ZnS quantum dots into arrays.
The cell membrane is another example of a self-
assembled system. Simplied models of cell membranes
consider solid-supported membranes as model systems
for fundamental biophysical research, biosensors, and the
design of phantom cells exhibiting well-dened adhesive
properties and receptor densities [323]. Several strategies for
the assembly of biomembranes onto various solid supports
have been reviewed in great detail [324328]. The major
techniques are direct vesicle fusion [329], the LB technique
[330], and molecular self-assembly from dilute organic solu-
tions [331]. Three types of supported membranes can be
assembled [332]:
(a) supported, lipid bilayer membranes prepared by vesi-
cle spreading or LB deposition at hydrophilic sur-
faces,
(b) covalently tethered SAMs of functionalized lipid
derivatives, prepared by coadsorption of membrane-
forming components,
(c) asymmetric or hybrid bilayer membranes, composed
of an outer lipid layer and an inner SAM. The most
natural membrane bilayer environment is the rst
approach.
Another innovative approach has been reported by
Walboomers et al. [333] and Britland et al. [334], where
they have demonstrated that cells can be contact-guided
along deep grooves on a patterned surface without any
chemical treatments to the surface itself. Britland et al.
[334] modied glass substrates with both chemical and topo-
graphical cues. The chemical adhesive tracks were patterned
both parallel and perpendicular to the topographical grooves
with laminin, an ECM protein carrying the RGD peptide
sequence. The cells were found to orient along chemical
adhesive track when the grooves on the substrate are less
than 500 nm in depth. For deeper grooves, the contact
guidance by the topography dominates the cellular behav-
ior (alignment). If nanoparticles are attached to the cells,
then by virtue of the cell alignment, the particles will get
aligned into the structure that has been engraved, analogous
to reports on decorating such structures by capillary forces
that are quite difcult to control.
In another recent article, biodirected epitaxial nano-
deposition of polymers was achieved on a template with an
oriented molecular surface. Acetobacter xylinum synthesized
a ribbon of cellulose I microbrils onto a xed, nematic,
ordered substrate of glucan chains with unique surface char-
acteristics. The substrate directed the orientation of the
motion due to the inverse force of the secretion during
biosynthesis, and the microbrils were aligned along the
orientation of the molecular template. This approach also
promises some outstanding possibility of self-organization
of nanoparticles [335]. Bacteriophages offer another way
to form large-dimension, patterned assemblies by engineer-
ing the phage to specically bind functional components, as
has been done for a variety of inorganic materials [336].
Because the phage is strongly bound to the inorganic com-
ponents, the phages are participants rather than bystanders
18 Self-Organization of Colloidal Nanoparticles
in the assembly process. Relatively small (20 nm in diame-
ter) particles are easily organized into the layered structure
formed by the viral rods, leading to well-populated lines only
a few tens of nanometers across, spaced by the length of the
bacteriophages. In a recent article recently published, Lin
et al. has mannose-encapsulated gold nanoparticles to type 1
pili in Escherichia coli [337]. Type 1 pili are lamentous, pro-
teinaceous appendages that extend from the surface of many
gram-negative organisms and are composed of FimA, FimF,
FimG, and FimH proteins. FimA accounts for more than
98% of the pilus protein, and FimH is uniquely responsible
for the binding to D-mannose (Fig. 26).
4. CONCLUSIONS AND PERSPECTIVE
The synthesis of nanostructured materials with tunable
properties is central to the development of varying appli-
cations in nanoscale science and technology. The con-
vergence of physics, chemistry, and biology will actually
lead to an explosive increase in possibilities of R&D
directions. Choices will be difcult and sometimes limited
by the breadth of expertise available. Encouragement of
multidisciplinary and multilaboratory collaborations would
accelerate progress to the benet of all. Through devices
such as quantum well lasers and their enabling roles in
applications, ranging from telecommunications to home
entertainment, it is evident that nanotechnology has and is
having a major impact on our daily lives which will only mul-
tiply over the decades ahead. Self-assembly of nanometric
particles will denitely play an important role in this. Non-
lithographic approaches based on thermodynamically driven
self-organization processes are especially appealing, because
of their potential for large-scale production with very small
infrastructure investments. The spontaneous organization of
monolayer-protected metal nanoparticles into periodic 2D
arrays is archetypal of this approach, with many of these
arrays demonstrating novel optical or electronic properties
as a function of particle size or interparticle spacing. Stabi-
lized particles beyond a nanometric-size threshold become
increasingly prone to multilayer or 3D aggregate formation,
which can be attributed to the rapid increase in van der
Waals attraction and the loss of surfactant chain mobility on
the planar facets of the nanoparticles as a function of size.
Figure 26. Typical TEM images of sectioned areas of (A) pili of the
E. coli ORN178 strain bound with m-AuNP; (B) the E. coli ORN208
strain decient of the mH gene without m-AuNP binding. The exper-
iments were performed in LB at room temperature. Reprinted, with
permission, from [337], C. C. Lin et al., J. Amer. Chem. Soc. 127, 3508
(2002). 2002, American Chemical Society.
The other approach of positioning molecules ranging
from small alkanethiols to larger biosystems, such as pro-
teins or DNA on surfaces that can be used as a generic
approach to position other molecules such as enzymes, colla-
gens, bronectin, or synthetic moieties, is another interesting
possibility to assemble ordered structures. As the interdisci-
plinarity of this eld develops further, we will see the birth
of innovative techniques aimed at mimicking nature. An
interesting article on the biomimetic approach to produce
materials has been published recently [338]. The interface
between biological systems and functional semiconductors
remains a relatively unexplored area that requires research
in the near future. The ability of electronic/optical devices to
sense minute samples of biological/medical materials in situ,
and to relay this information to a central location opens up
possibilities for quick diagnostics and the remote monitoring
of patients.
A research eld, which is not yet treated in detail, is the
inuence of size and ordered arrangement of particles in
layers on the heat conductivity. The phonon heat conduc-
tion mechanism in nanostructures differs signicantly from
the behavior in bulk material. Phonon size effects include
increased phonon scattering at grain boundaries, modica-
tion of the phonon dispersion relation, and phonon rarefac-
tion surrounding small structures have been predicted and
also observed experimentally [339340]. It also seems that
change of the phonon spectrum, due to connement effects,
could lead to phonon-phonon part of the thermal conduc-
tivity, like it occurs in superlattices. The fact that thermal
transport properties in nanostructured materials are affected
by the size opens up new possibilities for material scien-
tists to engineer structures adapted to the desired proper-
ties depending on its applications. Provided we understand
the phenomena in detail, one can engineer the structure of
the materials by assembly of particles in such a way that the
phonon conduction (phononics or phonon engineering) in
different directions could be possible. For microelectronics,
phonon engineering can provide improved device design to
minimize the impact of adverse size effects on the device
temperature [341].
The extension of Moores law is expected to continue until
the next decade (as shown in Fig. 27) with devices in their
present form, which will ultimately give way to disruptive
technologies involving nanoparticles, single molecule elec-
tronics, and single electron transistors [342]. Again, there
will be a need for the design of complex novel architectures
to accommodate these new devices with novel materials and
gate designs and interconnects. Today, the major shortfall of
nanoscience and nanotechnology is the lack of understand-
ing of the physics of the interaction of objects (surfaces,
particles, individual molecules) at the nanoscale. Questions
need to be answered regarding how nanoparticles can be
stabilized, and in what media, how nanoparticles interact
and inuence each other, what proportions in a hybrid
system make a critical difference, and how the ordered
structures can be created and retained. How can these
characteristics be predicted? This area of Extreme Nano-
technology requires vast amounts of imagination and intu-
ition if nanoscience research is to have any practical impact
on human lives. Its potential is enormous, and the need
Proof's Only
Self-Organization of Colloidal Nanoparticles 19
Figure 27. Moores law predictions. Adapted, with permission, from
[342], V. V. Khirnov and D. J. C. Herr, Computer, January 2001, p. 34.
2001, IEEE.
to understand underpinning nanoscience should be prop-
erly supported so that the bottom-up nanoscience and nano-
technology is not overshadowed by the reluctance to invest
in revolutionary technologies, since the benets from this
area will be considerable in the long run.
GLOSSARY
ACKNOWLEDGMENTS
The authors would like to thank the Swiss National program
in Nanotechnology (TOP NANO 21; 5971.2 TNS) for nan-
cial support. They would also like to thank colleagues and
students for assisting in the preparation of this article. Spe-
cial thanks to Dr. P. Bowen, Dr. R. Houriet, Dr. Y. Houst,
and Mr. F. Juillerat for their support and encouragement.
REFERENCES
1. V. Paillard, P. Melinon, J. P. Perez, V. Dupuis, A. Perez, and
B. Champagnon, Phys. Rev. Lett. 71, 4170 (1993).
2. P. Melinon, V. Paillard, V. Dupuis, A. Perez, P. Jensen, A. Hoareau,
M. Broyer, J.-L. Vialle, M. Pellarin, B. Baguenard, and J. Lerme,
Internat. J. Mod. Phys. B 9, 339 (1995).
3. H. Haberland M. Moseler, Y. Qiang, O. Rattunde, T. Reiners, and
Y. Thurner, Surf. Rev. Lett. 3, 887 (1993).
4. J. Dutta, H. Hofmann, Ch. Hollenstein, and H. Hofmeister,
Nanoparticles in Solids and Solutions: Preparation, Characteriza-
tion and Utilization, J. H. Fendler, Ed. Wiley-VCH, Weinheim,
Germany, 1998, Chapter 8 p. 173.
5. H. Hofmann, J. Dutta, S. Scholz-Odermatt, and J.-Ch. Valmalette,
Ceramic Engineering and Science Proceedings, 18(4B) (1997) 687.
6. S. Fafard, Z. R. Wasilewski, and M. Spanner, Appl. Phys. Lett. 75,
1866 (1999).
7. N. Carlsson, T. Junno, L. Montelius, M.-E. Pistol, L. Samuelson,
and W. Seifert, J. Cryst. Growth 191, 347 (1998).
8. V. I. Klimov, D. W. McBranch, C. A. Leatherdale, and M. G.
Bawendi, Phys. Rev. B 60, 13740 (1999).
9. F.-Y. Tsai and C. P. Lee, J. Appl. Phys. 84, 2624 (1998).
10. D. Gammon, E. S. Snow, B. V. Shanobrook, D. S. Katzer, and
D. Park, Science 273, 87 (1996).
11. Z. Y. Zhong, B. Gates, Y. N. Xia, and D. Qin, Langmuir 16, 10369
(2000).
12. M. Zahn, J. Nanoparticle Res. 3, 73 (2001).
13. J. Shi, S. Gider, K. Babcock, and D. D. Awschalom, Science 271,
937 (1996).
14. C. Lebreton, C. Vieu, A. Ppin, M. Mejias, F. Carcenac, Y. Jin,
and H. Launois, Microelectr. Eng. 41/42, 507 (1998).
15. A. D. Yoffe, Adv. Phys. 50, 1 (2001).
16. L. P. Kouwenhoven, D. G. Austing, and S. Tarucha, Rep. Prog.
Phys. 64, 701 (2001).
17. P. Ball, The Self-Made Tapestry: Pattern Formation in Nature
Oxford University Press, New York, 1999.
18. A. Miller, The Developmental Biology of Plants and Animals
(C. F. Graham and P. F. Wareing, Eds.), pp. 249269, Oxford:
Blackwell Scientic Publications, 1976.
19. S. Camazine and J.-L. Deneubourg, in Les Insectes Sociaux 12th
Congress of the I.U.S.S.I. Paris, Sorbonne, (A. Lenoir, G. Arnold,
and M. Lepage), p. 228, Universit Paris Nord, Paris 1994.
20. Y. Schiffmann, Prog. Bio. Mol. Biol., 68, 145 (1997).
21. G. Nicolis, G. Deweland, and J. Turner (Eds.), Order and Fluctu-
ations in Equilibrium and Nonequilibrium Statistical Mechanics,
Wiley, New York, 1981.
22. M. C. Cross and P. C. Hohenberg, Rev. Mod. Phys. 65, 851 (1993).
23. P. Borckmans, G. Dewel, De A. Wit, and D. Walgraef, in Chem-
ical Waves and Patterns, (R. Kapral and K. Showalter, Eds.),
Kluwer, Dordrecht 1994, pp. 323363.
24. R. Vacassy, S. M. Scholz, J. Dutta, C. J. G. Plummer, R. Houriet,
and H. Hofmann, J. Amer. Cer. Soc. 81, 2699 (1998).
25. E. Bodenschatz, W. Zimmermann, and L. Kramer, J. Phys., 49
(1988) 875.
26. I. Rehberg, B. L. Winkler, M. de la Torre Juarez, S. Rasenatand,
and W. Schopf, Festkorperprobleme/Adv. Solid State Phys. 29, 35
(1989).
27. F. H. Busse, Rep. Prog. Phys. 41, 1929 (1978).
28. G. I. Sivashinsky, Physica D 8, 243 (1983).
29. M. Q. Lpez-Salvans, F. Sagusand, and J. Claret, Proc. 1st Latin
American School on Materials Instabilities, Valparaiso, Chile,
Kluwer Publication, Dordrecht 2000.
30. H. Neuhauser, Plastic Instabilities and the Deformation of Met-
als, in: Patterns, Defects and Materials Instabilities (D. Walgraef,
and N. M. Ghoniem, Eds.). Kluwer Academic Publishers, Dor-
drecht, 1990.
31. J. Kratochvil, Rev. Phys. Appl. 23, 419 (1988).
32. R. Kossowsky and S. C. Singhal, Surface Engineering, Surface
Modication of Materials, Martinus Nijhoff, Dordrecht, 1984.
33. D. Bauerle, Laser Processing and Chemistry, Springer-Verlag,
New York, 1996.
34. J. H. Evans, Nature 29, 403 (1971).
35. M. Seul and R. Wolfe, Phys. Rev. A 46, 7534 (1992).
36. M. C. Roco, S. Williams, and P. Alivisatos (Eds.), Nanotech-
nology Research Directions: IWGN Workshop Report Vision for
Nanotechnology Research and Development in the Next Decade,
WTEC, Loyola College, Maryland, September, 1999.
37. I. Borgia, B. Brunetti, I. Mariani, A. Sgamellotti, F. Cariati,
P. Fermo, M. Mellini, C. Viti, and G. Padeletti, Appl. Surf. Sci. 185,
206 (2002).
38. G. Decher, Science 277, 1232 (1997).
39. J.-M. Lehn, PNAS 4763 (2002).
40. M. Jose-Yacaman, L. Rendon, J. Arenas, and M. C. S. Puche,
Science 273, 223 (1996).
41. http://www.liv.ac.uk/Chemistry/Links/selfassembly.html
42. E. Hu and D. Shaw, http://www.wtec.org/loyola/nano/nal/ch2.pdf,
p. 25
43. F. Schreiber, Prog. Surf. Sci. 65, 151 (2000).
Proof's Only
20 Self-Organization of Colloidal Nanoparticles
44. J. A. Larson, M. Nolan, and J. C. Greer, J. Phys. Chem. B 106,
593 (2002).
45. J. Flink, C. J. Kiely, D. Bethell, and D. Schiffrin, Chem Mater. 10,
922 (1998).
46. G. Carrot, J. C. Valmalette, C. J. G. Plummer, S. M. Scholz,
J. Dutta, H. Hofmann, and H. Hilborn, Coll. Poly. Sci. 276, 853
(1998).
47. M. Brust and C. J. Kiely, Coll Surf. A202, 175 (2002).
48. D. H. Napper, Polymeric Stabilization of Colloidal Dispersions,
Academic Press, London 1983.
49. J. T. G. Overbeek, Colloidal Dispersions, Royal Society of
Chemistry, London, 1981.
50. C. Bossel, J. Dutta, R. Houriet, J. Hilborn, and H. Hofmann,
Mater. Sci. Eng. A 204, 107 (1995).
51. P. C. Hiemenz, Principles of Colloid and Surface Chemistry,
Marcel Dekker, New York, 1986.
52. D. Myers, Surfaces, Interfaces and Colloids, VCH, Weinheim,
1991.
53. F. T. Hesselink, A. Vrij, and J. T. G. Overbeek, J. Phys. Chem. 75,
2094 (1971).
54. D. J. Meier, J. Phys. Chem. 71, 1861 (1967).
55. P. J. Flory and W. R. Krigbaum, J. Chem. Phys. 18, 1086 (1950).
56. F. R. Eirich, J. Coll. Internat. Sci. 58, 423 (1977).
57. A. J. Silberberg, J. Chem. Phys. 46, 1105 (1967).
58. G. Steinberg, J. Phys. Chem. 71, 292 (1967).
59. T. F. Tadros, J. Polym. 23, 683 (1991).
60. J. M. Stouffer and T. J. McCarthy, Macromolecules 21, 1204 (1988).
61. E. Killman, J. Eisenlauer, and M. Korn, J. Polym. Sci. Polym. Symp.
61, 413 (1977).
62. P.-G. de Gennes, Macromolecules 13, 1069 (1980).
63. R. Zerushalmi-Royen, J. Klein, and L. Fetters, Science 263, 793
(1994).
64. G. Reiter, Europhys. Lett. 33, 29 (1996).
65. R. G. Nuzzo and D. L. Allara, J. Am. Chem. Soc. 105, 4481 (1983).
66. T. L. Freeman, S. D. Evans, and S. D. Ulman, Langmuir 11, 4411
(1995).
67. R. Vacassy, L. Lemaire, J.-C Valmalette, J. Dutta, and H. Hof-
mann, J. Mater. Sci. Lett. 17, 1665 (1998).
68. M. Brust, M. Walker, D. Bethell, D. J. Schiffrin, and R. Whyman,
J. Chem. Soc., Chem. Commun. 7, 801 (1994).
69. C. P. Collier, T. Vossmeyer, and J. R. Heath, Ann. Rev. Phys. Chem.
49, 371 (1998).
70. D. J. Lavrich, S. M. Wetterer, S. L. Bernasek, and G. Scoles, J.
Phys. Chem. B 102, 3456 (1998).
71. S. S. Ghosh, P. M. Kao, A. W. McCue, and H. L. Chappelle, Bio-
conjugate Chem. 1, 71 (1990).
72. X. Yang, L. A. Wenzler, J. Qi, X. Li, and N. C. Seeman, J. Am.
Chem. Soc. 120, 9779 (1998).
73. E. Winfree, F. Liu, L. A. Wenzler, and N. C. Seeman, Nature 394,
539 (1998).
74. W.-L. Shaiu, D. D. Larson, J. Vesenka, and E. Henderson, Nucleic
Acids Res. 21, 99 (1993).
75. C. A. Mirkin, R. L. Letsinger, R. C. Mucic, and J. J. Storhoff,
Nature 382, 607 (1996).
76. S.-J. Park, A. A. Lazarides, C. A. Mirkin, P. W. Brazis, C. R. Kan-
newurf, and R. L. Letsinger, Angew. Chem. 112, 4003 (2000).
77. R. L. Letsinger, R. Elghanian, G. Viswanadham, and C. A. Mirkin,
Bioconjugate Chem. 11, 289 (2000).
78. H. Mattoussi, J. M. Mauro, E. R. Goldman, G. P. Anderson, V.
C. Sundar, F. V. Mikulec, and M. G. Bawendi, J. Am. Chem. Soc.
122, 12142 (2000).
79. R. S. Ingram, M. J. Hostetler, and R. W. Murray, J. Am. Chem.
Soc. 119, 9175 (1997).
80. G. Schmid, Clusters and Colloids, VCH, Weinheim, 1994.
81. M. Kerker, The Scattering of Light and Other Electromagnetic
Radiation, Academic Press, New York, 1969.
82. U. Kreibig and P. Z. Zacharias, Physics 231, 128 (1970).
83. R. H. Doremus and P. J. Rao, Mater. Res. 11, 2834 (1996).
84. G. L. Hornyak, C. J. Patrissi, C. R. Martin, J.-C Valmalette,
J. Dutta, and H. Hofmann, Nano. Mater. 9, 575 (1997).
85. T. Schalkhammer, Monatshefte fr Chemie 129, 1067 (1998).
86. T. Okamoto and I. Yamaguchi, Opt. Lett. 25, 372 (2000).
87. J. J. Storhoff, R. Elghanian, R. C. Mucic, C. A. Mirkin, and R. L.
Letsinger, J. Am. Chem. Soc. 120, 1959 (1998).
88. Y. Dirix, C. Bastiaansen, W. Caseri, and P. Smith, Adv. Mater. 11,
223 (1999).
89. A. H. Lu, G. H. Lu, A. M. Kessinger, and C. A. Foss Jr., J. Phys.
Chem. B 101, 9139 (1997).
90. L. G. Olson, Y. Lo, T. P. Beebe, Jr., and J. M. Harris, Anal. Chem.
73, 4268 (2001).
91. A. N. Shipway, E. Katz, and I. Willner, Chemphyschem. 1, 18
(2000).
92. A. Ulman, Chem. Rev. 96, 1533 (1996).
93. X. Michalet, F. Pinaud, T. D. Lacoste, M. Dahan, M. P. Bruchez,
A. P. Alivisatos, and S. Weiss, Single Mol. 2, 261 (2001).
94. A. L. Elfros and M. Rosen, Ann. Rev. Mater. Sci. 30, 475 (2000).
95. R. Vacassy, S. M. Scholz, J. Dutta, C. J. G. Plummer, R. Houriet,
and H. Hofmann, J. Am. Cer. Soc. 81, 2699 (1998).
96. M. L. Steigerwald, A. P. Alivisatos, J. M. Gibson, T. D. Harris,
R. Kortan, A. Muller, A. M. Thayer, T. M. Duncan, D. C. Dou-
glass, and L. E. Brus, J. Am. Chem. Soc. 110, 3046 (1988).
97. N. Herron, Y. Wangand, and H. Eckert, J. Am. Chem. Soc. 112,
1322 (1990).
98. S. M. Scholz, R. Vacassy, L. Lemaire, J. Dutta, and H. Hofmann,
App. Organo. Chem. 12, 327 (1998).
99. K. M. Choi and K. J. Shea, J. Phys. Chem. 98, 3207 (1994).
100. M. P. Pileni, L. Motte, and C. Petit, Chem. Mater. 4, 338 (1992).
101. X. K. Zhao, S. Baral, R. Rolandi, and J. H. Fendler, J. Am. Chem.
Soc. 110, 1012 (1988).
102. R. S. Urquhart, D. N. Furlong, T. Gengenbach, N. J. Geddes, and
F. Grieser, Langmuir 11, 1127 (1995).
103. G. L. Hornyak, C. J. Patrissi, C. R. Martin, J. C. Valmalette,
L. Lemaire, J. Dutta, and H. Hofmann, Nano. Mater. 9, 571 (1997).
104. K. K. W. Wong and S. Mann, Adv. Mater. 8, 928 (1996).
105. C. T. Dameron, R. N. Reese, R. K. Mehra, A. R. Kortan, P. J.
Carroll, M. L. Steigerwald, L. E. Brus, and D. R. Winge, Nature
338, 596 (1989).
106. R. N. Reese, C. A. White, and D. R. Winge, Plant Physiol. 98, 225
(1992).
107. C. T. Dameron and D. R. Winge, Inorg. Chem. 29, 1343 (1990).
108. I. Soten and G. A. Ozin, Curr. Opin. Coll. Interf. Sci. 4, 325 (1999).
109. N. Malikova, I. Pastoriza-Santos, M. Schierhorn, N. A. Kotov, and
L. M. Liz-Marzan, Langmuir 18, 3694 (2002).
110. P. C. Andersen and K. L. Rowlen, Appl. Spectro. 56, 124A (2002).
111. X. M. Lin, H. M. Jaeger, C. M. Sorensen, and K. J. Klabunde,
J. Phys. Chem. B 105, 3353 (2001).
112. C. P. Collier, T. Vossmeyer, and J. R. Heath, Ann. Rev. Phys. Chem.
49, 371 (1998).
113. C. B. Murray, C. R. Kagan, and M. G. Bawendi, Science 270, 1335
(1995).
114. R. P. Andres, J. D. Bielefeld, J. I. Henderson, D. B. Janes,
V. R. Kolagunta, C. P. Kubiak, W. J. Mahoney, and R. G. Osifchin,
Science 273, 1690 (1996).
115. R. L. Whetten, J. T. Khoury, M. M. Alvarez, S. Murthy, I. Vezmar,
Z. L. Wang, P. W. Stephens, C. L. Cleveland, W. D. Luedtke, and
U. Landman, Adv. Mater. 8, 428 (1996).
116. V. M. Shalaev and M. Moskovits, Nanostructured Materials: Clus-
ters, Composites and Thin Films, ACS Symposium Series,Vol.
679, Washington DC, American Chemical Society, 1997.
117. C. S. Weisbecker, M. V. Merritt, and G. M. Whitesides, Langmuir
12, 3763 (1996).
118. S. Biggs and P. Mulvaney, J. Chem. Phys. 100, 8501 (1994).
119. V. Kane and P. Mulvaney, Langmuir 14, 3303 (1998).
Self-Organization of Colloidal Nanoparticles 21
120. J. Israelachvili, Intermolecular and Surface Forces, 2nd ed., Aca-
demic Press, New York 1992, Chapter 10.
121. T. Fujimura, T. Itoh, A. Imada, R. Shimada, T. Koda, N. Chiba,
H. Muramatsu, H. Miyazaki, and K. Ohtaka, J. Lumin. 8789, 954
(2000).
122. A. Rogach, A. Susha, F. Caruso, G. Sukhorukov, A. Kornowski,
S. Kershaw, H. Mhwald, A. Eychmuller, and H. Weller, Adv.
Mater. 12, 333 (2000).
123. R. Shimada, Y. Komori, T. Koda, T. Fujimura, T. Itoh, and
K. Ohtaka, Mol. Cryst. Liq. Cryst. Sci. Technol. Sect. A 349, 5 (2000).
124. G. Subramania, K. Constant, R. Biswas, M. M. Sigalas, and K. M.
Ho, J. Lightwave Technol. 17, 1970 (1999).
125. Y. Xia, B. Gates, and S. H. Park, J. Lightwave Technol. 17, 1956
(1999).
126. A. K. Boal, F. Ilhan, J. E. DeRouchey, T. Thurn-Albrecht, T. P.
Russell, and V. M. Rotello, Nature 404, 746 (2000).
127. F. Auer, M. Scotti, A. Ulman, R. Jordan, B. Sellergren, J. Garno,
and G.-Y. Liu, Langmuir, 16, 7554 (2000).
128. R. Meallet-Renault, P. Denjean, and R. B. Pansu, Sens. Actuators
B 59, 108 (1999).
129. F. Burmeister, W. Badowsky, T. Braun, S. Wieprich, J. Boneberg,
and P. Leiderer, Appl. Surf. Sci. 144145, 461 (1999).
130. C. P. Collier, T. Vossmeyer, and J. R. Heath, Ann. Rev. Phys. Chem.
49, 371 (1998).
131. J. Boneberg, F. Burmeister, C. Schae.e, P. Leiderer, D. Reim,
A. Fery, and S. Herminghaus, Langmuir 13, 7080 (1997).
132. J. C. Halteen and R. P. Van Duyne, J. Vac. Sci. Technol. A 13, 1553
(1995).
133. M. H. Charles, M. T. Charreyre, T. Delair, A. Elaissari, and
C. Pichot, S. T. P. Pharm. Sci., 11, 251 (2001).
134. O. D. Velev and E. W. Kaler, Langmuir 15, 3693 (1999).
135. G. Bauer, F. Pittner, and T. Schalkhammer, Mikrochim. Acta 131,
107 (1999).
136. H. Takei, in Microuidic Devices and Systems (A. B. Frazier
and C. H. Ahn, Eds.) SPIE-Internat. Soc. Opt. Eng., Santa Clara,
CA, 3515, p. 278 (1998).
137. C. G. J. Koopal and R. J. M. Nolte, Enzyme Microb. Technol. 16,
402 (1994).
138. F. Ganachaud, A. Elaiessari, C. Pichot, A. Laayoun, and P. Cros,
Langmuir 13, 701 (1997).
139. S.-C. Huang, H. Swerdlow, and K. D. Caldwell, Anal. Biochem.
222, 441 (1994).
140. W. Hrtl, C. Beck, and R. Hempelmann, J. Chem. Phys. 110, 7070
(1998).
141. Y. Solomentsev, M. Bhmer, and J. L. Anderson, Langmuir 13,
6058 (1997).
142. D. Rudhardt, C. Bechinger, and P. Leiderer, Polymers 112, 163
(1999).
143. C. Bechinger, M. Brunner, and P. Leiderer, Phys. Rev. Lett. 86, 930
(2001).
144. E. Kim, Y. Xia, and G. M. Whitesides, Nature 376, 581 (1995).
145. K. M. Chen, X. Jiang, L. C. Kimerling, and P. T. Hammond, Lang-
muir 16, 7825 (2000).
146. J. Aizenberg, P. V. Braun, and P. Wiltzius, Phys. Rev. Lett. 84, 2997
(2000).
147. P. A. Kralchevsky and N. D. Denkov, Curr. Opin. Coll. Interf. Sci.
6, 383 (2001).
148. C. D. Dushkin, G. S. Lazarov, S. N. Kotsev, H. Yoshimura, and
K. Nagayama, Coll. Polym. Sci. 227, 914 (1999).
149. K.-H. Lin, J. C. Crocker, V. Prasad, A. Schoeld, D. A. Weitz,
T. C. Lubensky, and A. G. Yodh, Phys. Rev. Lett. 85, 1770 (2000).
150. M. Heniand and H. Lwen, Phys. Rev. Lett. 85, 3668 (2000).
151. P. V. Braun, R. W. Zehner, C. A. White, M. K. Weldon, C. Kloc,
S. S. Patel, and P. Wiltzius, Adv. Mater. 13, 721 (2001).
152. J. W. G. Tyrrell and P. Attard, Phys. Rev. Lett. 87, 176104 (2001).
153. Y. Lu, Y. Yin, and Y. Xia, Adv. Mater. 13, 34 (2001).
154. Y. Yin, Y. Lu, and Y. Xia, J. Am. Chem. Soc. 123, 771 (2000).
155. O. D. Velev, K. Furusawa, and K. Nagayama, Langmuir 12, 2374
(1996).
156. W. T. S. Huck, J. Tien, and G. M. Whitesides, J. Am. Chem. Soc.
120, 8267 (1998).
157. O. D. Velev, A. M. Lenhoff, and E. Kaler, Science 87, 2240 (2000).
158. G. A. Ozin and M. Y. Yang, Adv. Func. Mater. 11, 95 (2001).
159. E. Kim, Y. Xia, and G. M. Whitesides, Adv. Mater. 8, 245 (1996).
160. A. Stein and R. C. Schroeden, Curr. Opin. Solid State Mater. Sci.
5, 553 (2001).
161. O. D. Velev, T. A. Jede, R. F. Lobo, and A. M. Lenhoff, Nature
389, 447 (1997).
162. O. D. Velev, T. A. Jede, R. F. Lobo, and A. M. Lenhoff, Chem.
Mater. 10, 3597 (1998).
163. S. H. Park and Y. Xia, Adv. Mater. 10, 1045 (1998).
164. B. Gates, Y. Yin, and Y. Xia, Chem. Mater. 11, 2827 (1999).
165. B. T. Holland, C. F. Blanford, and A. Stein, Science 281, 538 (1998).
166. B. T. Holland, C. F. Blanford, T. Do, and A. Stein, Chem. Mater.
11, 795 (1999).
167. S. A. Johnson, P. J. Ollivier, and T. E. Mallouk, Science 283, 963
(1999).
168. J. S. Yin and Z. L. Wang, Adv. Mater. 11, 469 (1999).
169. H. Yan, C. F. Blanford, B. T. Holland, M. Parent, W. H. Smyrl,
and A. Stein, Adv. Mater. 11, 1003 (1999).
170. Y. A. Vlasov, N. Yao, and D. J. Norris, Adv. Mater. 11, 165 (1999).
171. G. Subramanian, V. N. Manoharan, J. D. Thorne, and D. J. Pine,
Adv. Mater. 11, 1261 (1999).
172. O. D. Velev, P. M. Tessier, A. M. Lenhoff, and E. W. Kaler, Nature
401, 548 (1999).
173. P. Jiang, J. Cizeron, J. F. Bertone, and V. L. Colvin, J. Am. Chem.
Soc. 121, 7957 (1999).
174. P. V. Braun and P. Wiltzius, Nature 402, 603 (1999).
175. A. A. Zakhidov, R. H. Baughman, Z. Iqbal, C. Cui, I. Khayrullin,
S. O. Dantas, J. Marti, and V. G. Ralchenko, Science 282, 897
(1998).
176. P. Yang, T. Deng, D. Zhao, P. Feng, D. Pine, B. F. Chmelka, G. M.
Whitesides, and G. D. Stucky, Science 282, 2244 (1998).
177. F. Meseguer, A. Blanco, H. Miguez, F. Garcia-Santamaria,
M. Ibisate, and C. Lopez, Coll. Surf. A: 202 (2002) 281.
178. Y. Yin, Y. Lu, B. Gates, and Y. Xia, J. Am. Chem. Soc. 123, 8718
(2001).
179. A. Badia, S. Singh, L. Demers, L. Cuccia, G. R. Brown, and R. B.
Lennox, Eur. J. Chem. 2, 359 (1996).
180. J. Schmitt, G. Decher,W. J. Dressick, S. L. Brandow, R. E. Geer,
R. Shashidhar, and J. M. Calvert, Adv. Mater. 9, 61 (1997).
181. T. Hanaoka, H. P. Kormann, M. Krll, T. Sawitowski, and
G. Schmid, Eur. J. Inorg. Chem. 807 (1998).
182. M. E. Davis, Nature 417, 813 (2002).
183. A. Ulman, An Introduction to Ultrathin Organic Films: From
LangmuirBlodgett to Self-Assembly, Academic Press, San
Diego, CA, 1991.
184. P. Ganguly, D. V. Paranjape, K. R. Patil, M. Sastry, and F. Ron-
delez, Langmuir 13, 5440 (1997).
185. M. Clemente-Leon, C. Mingotaud, B. Agricole, C. J. Gomez-
Garcia, E. Coronado, and P. Delhaes, Angew. Chem. Internat. Ed.
Engl. 36, 1114 (1997).
186. A. Riccio, M. Lanzi, F. Antolini, C. De Nitti, C. Tavani, and
C. Nicolini, Langmuir 12, 1545 (1996).
187. U. Raedler, C. Heiz, P. Luigi, and R. Tampe, Langmuir 14, 6620
(1998).
188. P. Pieranski, Phys. Rev. Lett. 45, 569 (1980).
189. A. J. Hurd and D. W. Schaefer, Phys. Rev. Lett. 54, 1043 (1985).
190. A. J. Armstrong, R. C. Mockler, and W. J. OSullivan, J. Phys.
Condens. Matt. 1, 1707 (1989).
191. J. C. Earnshaw, J. Phys. D 19, 1863 (1986).
192. A. Ulman, Adv. Mater. 3, 298 (1991).
193. J. H. Fendler, Chem. Mater. 8, 1616 (1996).
22 Self-Organization of Colloidal Nanoparticles
194. K. S. Mayya, V. Patil, M. Kumar, and M. Shastry, Thin Solid Films
312, 308 (1998).
195. H. Schuller, Kolloid-Z. 216217, 380 (1967).
196. E. Sheppard and N. Tcheurekdjian, J. Coll. Interf. Sci. 28, 48 (1968).
197. J. Kumaki, Macromolecules, 19, 2258 (1996).
198. K.-U. Fulda, D. Pieche, B. Tieke, and H. Yarmohammadipour,
Prog. Coll. Polym. Sci. 101, 178 (1996).
199. M. Bardosova, P. Hodge, L. Pach, V. Smatko, and R. H Tred-
gold, 8th European Conference on Organised Films, Book of
Abstracts, 2001.
200. C. A. Murray and D. H. V. Winkle, Phys. Rev. Lett. 58, 1200 (1987).
201. A. T. Skjeltorp and P. Meakin, Nature 335, 424 (1988).
202. N. D. Denkov, O. D. Velev, P. A. Kralchevsky, I. B. Ivanov,
H. Yoshimura, and K. Nagayama, Nature 361, 26 (1993).
203. G. Picard, Langmuir 14, 3710 (1998).
204. M. Kondo, K. Shinozaki, L. Bergstrom, and N. Mizutani, Langmuir
11, 394 (1995).
205. H. W. Deckman, J. H. Dunsmuir, and S. M. Gruner, J. Vac. Sci.
Technol. B7, 1832 (1989).
206. F. Lenzmann, K. Li, A. H. Kitai, and H. D. H. Stover, Chem. Mater.
6, 156 (1994).
207. K. U. Fulda and B. Tieke, Adv. Mater. 6, 288 (1994).
208. F. Burmeister, C. Schae, T. Matthes, M. Bohmisch, J. Boneberg,
and P. Leiderer, Langmuir 13, 2983 (1997).
209. N. D. Denkov, O. D. Velev, P. A. Kralchevsky, I. B. Ivanov,
H. Yoshimura, and K. Nagayama, Langmuir 8, 3183 (1992).
210. A. S. Dimitrov and K. Nagayama, Langmuir 12, 1303 (1996).
211. O. D. Velev, N. D. Denkov, V. N. Paunov, P. A. Kralchevsky, and
K. Nagayama, Langmuir 9, 3702 (1993).
212. S. Rakers, L. F. Chi, and H. Fuchs, Langmuir 13, 7121 (1997).
213. A. S. Dimitrov, T. Miwa, and K. Nagayama, Langmuir 15, 5257
(1999).
214. A. S. Dimitrov, C. D. Dushkin, H. Yoshimura, and K. Nagayama,
Langmuir 10, 432 (1994).
215. G. S. Lazarov, N. D. Denkov, O. D. Velev, P. A. Kralchevsky, and
K. Nagayama, J. Chem. Soc. Faraday Trans. 90, 2077 (1994).
216. B. Kim, S. L. Tripp, and A. Wei, Materials Research Society
Symposium Proceedings, Purdue University, 2002, 676 (Synthe-
sis, Functional Properties and Applications of Nanostructures),
Y6.1.1Y6.1.7.
217. L. M. Goldenberg, J. Wagner, J. Stumpe, B.-R. Paulke, and
E. Grnitz, Langmuir 18, 5627 (2002).
218. R. Aveyard, J. H. Clint, D. Nees, and V. N. Paunov, Langmuir 16,
8820 (2000).
219. D. H. Everett, Basic Principles of Colloid Science, Royal Society
of Chemistry, London 1988.
220. W. B. Russel, D. A. Saville, and W. R. Schowalter, Colloidal Dis-
persions, Cambridge University Press, Cambridge, 1989.
221. R. J. Hunter, Introduction to Modern Colloid Science, Oxford
University Press, Oxford, 1993.
222. A. K. Arora and B. V. R. Tata (Eds.), Ordering and Phase Tran-
sitions in Colloidal Systems, VCH, Weinheim, 1996.
223. K. E. Davis, W. B. Russel, and W. J. Glantschnig, Science 245, 507
(1989).
224. P. N. Pusey and W. van Megen, Nature 320, 340 (1986).
225. P. J. Darragh, A. J. Gaskin, and J. V. Sanders, Aust. Gemmol.,
November, 109 (1977).
226. T. C. Simonton, R. Roy, S. Komarneni, and E. Breval, J. Mater.
Res. 1, 667 (1986).
227. A. P. Philipse, J. Mater. Sci. Lett. 8, 1371 (1989).
228. V. N. Bogomolov, S. V. Gaponenko, I. N. Germanenko, A. M.
Kapitonov, E. P. Petrov, N. V. Gaponenko, A. V. Prokoev, A. N.
Ponyavina, N. I. Silvanovich, and S. M. Samoilovich, Phys. Rev.
E55, 7619 (1997).
229. J. V. Sanders, Nature 204, 1151 (1964).
230. H. Miguez, F. Meseguer, C. Lopez, A. Mifsud, J. S. Moya, and
L. Vazquez, Langmuir 13, 6009 (1997).
231. L. V. Woodcock, Nature 388, 235 (1997).
232. A. van Blaaderen, R. Ruel, and P. Wiltzius, Nature 385, 321 (1997).
233. P. Jiang, J. F. Bertone, K. S. Hwang, and V. L. Colvin, Chem. Mater.
11, 2132 (1999).
234. H. Kuhn, D. Mbius, and H. Bcher, in Physical Methods of
Chemistry (A. Weisenberger and B. Rossiter Eds.) Vol. 1 (3b),
577. Wiley, New York, 1972.
235. M. Gleiche, L. F. Chi, and H. Fuchs, Nature 403, 173 (2000).
236. M. Himmelhaus and H. Takei, Phys. Chem. Chem. Phys. 4, 496
(2002).
237. U. Jonas, A. del Campo, C. Krger, G. Glasser, and D. Boos,
PNAS 99, 5034 (2002).
238. S. Friebel, J. Aizenberg, S. Abad, and P. Wiltzius, Appl. Phys. Lett.
77, 2406 (2000).
239. M. Nakagawa, S.-K. Oh, and K. Ichimura, Adv. Mater. 12, 403
(2000).
240. G. S. Lazarov, N. D. Denkov, O. D. Velev, P. A. Kralchevsky, and
K. Nagayama, J. Chem. Soc. Faraday Trans. 90, 2077 (1994).
241. M. M. Burns, J. M. Fournier, and J. A. Golovchenko, Science 249,
749 (1990).
242. W. Hu, H. Li, B. Chang, J. Yang, Z. Li, J. Xu, and D. Zhang, Opt.
Lett. 20, 964 (1995).
243. H. Misawa, K. Sasaki, M. Koshioka, N. Kitamura, and
H. Masuhara, Appl. Phys. Lett. 60, 310 (1992).
244. C. Mio and M. D. W. Marr, Langmuir 15, 8565 (1999).
245. A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, Opt. Lett.
11, 288 (1986).
246. M. M. Burns, J. M. Fournier, and J. A. Golovchenko, Science 249,
749 (1990).
247. S. P. Smith, S. R. Bhalotra, A. L. Brody, B. L. Brown, K. Boyda,
and M. Prentiss, Am. J. Phys. 67, 26 (1999).
248. A. Ashkin, J. M. Dziedzic, and T. Yamane, Nature 330, 769 (1987).
249. A. Ashkin and J. M. Dziedzic, Science 235, 1517 (1987).
250. M. D. Wang, H. Yin, R. Landick, J. Gelles, and S. M. Block,
Biophys. J. 72, 1335 (1997).
251. S. Sato and H. Inaba, Opt. Quantum Electron. 28, 1 (1996).
252. R. C. Hayward, D. A. Saville, and I. A. Aksay, Nature 404, 56
(2000).
253. N. Sakai, A. Fujishima, T. Watanabe, and K. Hashimoto, J. Phys.
Chem. B105, 3023 (2001).
254. J. Priestley, The History and Present State of Electricity with
Original Experiments, 2nd ed, J. Dodsley, London 1769.
255. H. A. Pohl, Dielectrophoresis, Cambridge University Press,
Cambridge, UK, 1978.
256. W. M. Winslow, J. Appl. Phys. 20, 1138 (1949).
257. M. Parthasarathy and D. Klingenberg, Mater. Sci. Eng. R.17, 57
(1996).
258. U. Dassanayake, S. Fraden, and A. van Blaaderen, J. Chem. Phys.
112, 3851 (2000).
259. E. Kumacheva, R. K. Golding, M. Allard, and E. H. Sargent, Adv.
Mater. 14, 221 (2002).
260. J. S. Crane and H. A. Pohl, Electrostatics 5, 11 (1978).
261. Y. Huang and R. Pethig, Meas. Sci. Technol. 2, 1142 (1991).
262. T. L. Mahaworashilpa, H. G. L. Coster, and E. P. George, Biophys.
Acta 1193, 118 (1994).
263. T. Schnelle, T. Mller, G. Gradl, S. G. Shirley, and G. Fuhr, Elec-
trophoresis 21, 66 (2000).
264. T. B. Jones, Electromechanics of Particles, Cambridge University
Press, New York, 1995.
265. A. Docoslis, N. Kalogerakis, L. A. Behie, and K. V. I. S. Kaler,
Biotechnol. Bioeng. 54, 239 (1997).
266. M. P. Hughes, M. F. Flynn, and H. Morgan, Electrostatics 163, 81
(1999).
267. N. G. Green and H. Morgan, J. Phys. Chem. B103, 41 (1999).
Self-Organization of Colloidal Nanoparticles 23
268. A. Docoslis and P. Alexandridis, Proc. 2001 Annual Meeting
American Electrophoresis Society, AIChE Pub. No. 148, 91
(2001).
269. C. Reichle, T. Schnelle, T. Mller, T. Leya, and G. Fuhr, Biochim.
Biophys. Acta 1459, 218 (2000).
270. T. Yamamoto, O. Kurosawa, H. Kabata, N. Shimamoto, and
M. Washizu, IEEE Trans. Ind. Appl. 36, 1010 (2000).
271. H. Morgan, M. P. Hughes, and N. G. Green, Biophys. J. 77, 516
(1999).
272. O. D. Velev and E. W. Kaler, Langmuir 15, 3693 (1999).
273. K. D. Hermanson, S. O. Lumsdon, J. P. Williams, E. W. Kaler, and
O. D. Velev, Science 294, 1082 (2001).
274. P. A. Smith, C. D. Nordquist, T. N. Jackson, T. S. Mayer, B. R.
Martin, J. Mbindyoand, and T. E. Mallouk, Appl. Phys. Lett. 77,
1399 (2000).
275. G. Fuhr, C. Rechle, T. Mller, K. Kahlke, K. Schutze, and
M. Stuke, Appl. Phys. A Mater. Sci. Process. 69, 611 (1999).
276. A. Docoslis and P. Alexandris, Electrophoresis 23, 2174 (2002).
277. M. Trau, D. A. Saville, and I. A. Aksay, Langmuir 13, 6375 (1997).
278. M. Trau, D. A. Saville, and I. A. Aksay, Science 272, 706 (1996).
279. M. Bohmer, Langmuir 12, 5747 (1996).
280. M. Giersig and P. Mulvaney, J. Phys. Chem. 97, 6334 (1993).
281. M. Giersig and P. Mulvaney, Langmuir 9, 3408 (1993).
282. Y. Solomentsev, M. Bhmer, and J. L. Anderson, Langmuir 13,
6058 (1997).
283. R. E. Rosensweig, Ferrohydrodynamics, Dover Publications,
Mineola, 1997.
284. M. Golosovsky, Y. Saado, and D. Davidov, Appl. Phys. Lett. 75,
4168 (1999).
285. E. L. Bizdoaca, M. Spasova, M. Farle, M. Hilgendor, and
F. Caruso, J. Magnetism Magnetic Mater. 240, 44 (2002).
286. J. Lin, J. Solid State Chem. 159, 26 (2001).
287. E. Baer, A. Hiltner, and H. D. Keith, Science 235, 1015 (1987).
288. T. Shimizu, Macromol. Rapid Comm. 23, 311 (2002).
289. J.-M. Lehn, Supramolecular Chemistry Concepts and Perspec-
tives, VCH, Weinheim, 1995.
290. T. Moore and A. Steitz, Science 289, 905 (2000).
291. E. Westhof and N. Leontis, Angew. Chem. 112, 1651 (2000).
292. D. M. J. Lilley, Chem. Bio. Chem. 2, 31 (2001).
293. M. L. DePamphilis and J. Adler, J. Bacteriol. 105, 376 (1971).
294. R. M. Macnab, Ann. Rev. Genet. 26, 131 (1992).
295. N. R. Francis, G. E. Sosinsky, D. Thomas, and D. J. DeRosier, J.
Mol. Biol. 235, 1261 (1994).
296. A. Houdusse, V. N. Kalabokis, D. Himmel, A. G. Szent-Gyorgyi,
and C. Cohen, Cell 97, 459 (1999).
297. P. Alivisatos, K. P. Johnsson, X. Peng, T. E. Wilson, C. J. Loweth,
M. Bruchez, and P. G. Schultz, Nature 382, 609 (1996).
298. W. Shenton, S. A. Davis, and S. Mann, Adv. Mater. 11, 449 (1999).
299. S. Connolly and D. Fitzmaurice, Adv. Mater. 11, 1202 (1999).
300. H. A. Klok, Angew. Chem. Int. Ed. 41, 1509 (2002).
301. D. Philip and J. F. Stoddart, Angew. Chem. Internat. Ed. Engl. 35,
1154 (1996).
302. R. Elghanian, J. J. Storhoff, R. C. Mucic, R. L. Letsinger, and
C. A. Mirkin, Science 277, 1078 (1997).
303. H. Weller, Angew. Chem. 108, 1159 (1996).
304. D. Bethell and D. J. Schiffrin, Nature 382, 581 (1996).
305. A. P. Alivisatos, Science 271, 933 (1996).
306. R. F. Service, Science 271, 920 (1996).
307. T. Takahagi, S. Huang, G. Tsutsui, H. Sakaueand, and S. Shin-
gubara, Mater. Res. Soc. Symp. Proc. 707, 87 (2002).
308. K. K. W. Wong and S. Mann, Curr. Opin. Coll. Inter. Sci. 3, 63
(1998).
309. T. Furuno, H. Sasabe, and K. M. Ulmer, Thin Solid Films 180, 23
(1989).
310. H. Yoshimura, M. Matsumoto, S. Endo, and K. Nagayama, Ultra-
Microscopy 32, 265 (1990).
311. H. Yoshimura, T. Scheybani, W. Baumeister, and K. Nagayama,
Langmuir 10, 3290 (1994).
312. S. Takeda, H. Yoshimura, S. Endo, T. Takahashi, and
K. Nagayama, Proteins 23, 548 (1995).
313. T. Scheybani, H. Yoshimura, W. Baumeister, and K. Nagayama,
Langmuir 12, 431 (1996).
314. I. Yamashita, Thin Solid Films 393, 12 (2001).
315. K. M. Towe, in Origin, Evolution and Modern Aspects of Biomin-
eralization in Plants and Animals, (R. E. Crick, Ed.), Plenum
Press, New York, p. 265, 1990.
316. F. C. Meldrum, V. J. Wade, D. L. Nimmo, B. R. Heywood, and
S. Mann, Nature 349, 684 (1991).
317. F. C. Meldrum, B. R. Heywood and S. Mann, Science 257, 522
(1 1992).
318. F. C. Meldrum, T. Douglas, S. Lei, P. Arosio, and S. Mann, J. Inorg.
Biochem. 58, 59 (1995).
319. T. Douglas, D. P. E. Dickson, S. Betteridge, J. Charnock, C. D.
Garner, and S. Mann, Science 269, 54 (1995).
320. S. R. Sczekan and J. G. Joshi, Biochim. Biophys. Acta 990, 8
(1998).
321. K. K. W. Wong and S. Mann, Adv. Mater. 8, 928 (1996).
322. A. McMillian, C. D. PMVOLA, J. Howard, S. L. Chaw, N. J.
Zaluzec, and J. D. Trent, Nature Mater. 1, 247 (2002).
323. E. Sackmann, Science 271, 43 (1996).
324. C. A. Steinem, A. Janshoff, W. P. Ulrich, M. Sieber, and H. J.
Galla, Biochim. Biophys. Acta-Biomembr. 1279, 169 (1996).
325. G. Puu and I. Gustafson, Biochim. Biophys. Acta 1327, 149
(1997).
326. A. L. Plant, Langmuir 15, 5128 (1999).
327. E. Sackmann and M. Tanaka, Trends Biotechnol. 18, 58 (2000).
328. S. G. Boxer, Curr. Opin. Chem. Biol. 4, 704 (2000).
329. R. G. Horn, Biochim. Biophys. Acta 778, 224 (1984).
330. L. K. Tamm and H. M. McConnell, Biophys. J. 47, 105 (1986).
331. C. Duschl, M. Liley, H. Lang, A. Ghandi, S. M. Zakeeruddin,
H. Stahlberg, A. Nemetz, W. Knoll, and H. Vogel, Mater. Sci. Eng.
C4, 7 (1996).
332. A. N. Parikh, J. D. Beers, A. P. Shreve, and B. I. Swanson, Lang-
muir 15, 5369 (1999).
333. X. F. Walboomers, W. Monaghan, A. S. G. Curtis, and J. A. Jansen,
J. Biomed. Mater. Res. 46, 212 (1999).
334. S. Britland, C. Perridge, M. Denyer, H. Morgan, A. Curtis, and
C. Wilkinson, Exp. Biol. Online 1, 2 (1996).
335. T. Kondo, M. Nojiri, Y. Hishikawa, E. Togawa, D. Romanovicz,
and R. M. Brown, Jr., PNAS, 99, 14008 (2002).
336. S. R. Whaley, D. S. English, E. L. Hu, P. F. Barbara, A. M. Belcher,
Nature, 405, 665 (2000).
337. C. C. Lin, Y. C. Yeh, C. Y. Yang, C. L. Chen, G. F. Chen, C. C.
Chen, and Y. C. Wu, J. Am. Chem. Soc. 124, 3508 (2002).
338. J. F. V. Vincent, Materials Today, Dec. 2002, p. 28.
339. G. Chen, Internat. J. Therm. Sci. 39, 471 (2000).
340. L. Braginsky, N. Lukzen, V. Shklover, and H. Hofmann, Phys. Rev.
B 66, 134203 (2002).
341. L. B. Kish, Phys. Lett. A 305, 144 (2002).
342. V. V. Zhirnov and D. J. C. Herr, Computer, January 2001, p. 34.

Das könnte Ihnen auch gefallen