Sunteți pe pagina 1din 14

Nanotechnology: The Top-Down and

Bottom-Up Approaches
Parvez Iqbal,1 Jon A. Preece,2 and Paula M. Mendes1
1
2

School of Chemical Engineering, The University of Birmingham, Birmingham, UK


School of Chemistry, The University of Birmingham, Birmingham, UK

1 Introduction
2 Nanofabrication
3 Conclusion and Outlook
Acknowledgments
References

1
5
11
12
12

INTRODUCTION

Nanotechnology involves the study, imaging, measuring,


modeling, or manipulation of matter at scales falling in
the range of 1100 nanometers (nm). It is a highly multidisciplinary field, drawing from fields such as chemistry, materials science, colloidal science, applied physics,
engineering, and biology. In a relatively short span of
about 30 years of research, nanotechnology is already having an impact on society and several industrial sectors,
and such impact will increasingly be felt as its products increase in number and become more commercialized.1 Nanotechnology-based substances are now found
in a wide range of household products and products
intended for professional use, including sports gear, cosmetics, sunscreen lotions, food and food packaging material,
clothing, household appliances, electronic devices, disinfectants, paints, furniture varnishes, building materials, and
medicines. Nanotechnology is predicted to touch nearly

every industry and every part of our lives and become the
basis for remarkably powerful and inexpensive computers,
fundamentally new diagnostic and therapeutic technologies
(see Supramolecular Nanoparticles for Molecular Diagnostics and Therapeutics, Nanotechnology) that could
enhance human health and longevity,2 advanced sensors3
for military applications and environmental protection, and
new zero-pollution transportation technology.4, 5
This article provides an overview of nanotechnology,
describing the origins of the field, present technology,
ongoing research, and future aspirations. In addition, the
two possible methodologies of fabricationthe top-down
and bottom-up approachesare discussed, covering the
merits and drawbacks of each approach.

1.1

Brief history of nanotechnology

Historically, the concept of nanotechnology was first proposed by the Nobel laureate Richard Feynman, when he
gave a now-famous talk called Theres Plenty of Room at
the Bottom at an American Physical Society meeting at
Caltech in 1959.6 With this visionary talk, Feynman discussed both top-down and bottom-up possibilities of working at the molecular level, most of which are still relevant
today. Extrapolating from known physical laws, he argued
the possibility of molecular writing, seeing and rearranging
individual atoms, the prospect of designing molecules one
atom at a time, and the challenges involved in developing
nanometer-scale devices. In his talk, Feynman made several
references to examples in nature such as cells, which are
very tiny, but they are very active; they manufacture various substances; they walk around; they wiggle; and they do

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology

all kinds of marvelous thingsall on a very small scale.


Nonetheless, Feynman never used the term Nanotechnology to describe this new scientific field.
Several years later, in 1974, Professor Norio Taniguchi
coined the term Nanotechnology in order to define topdown ultraprecision machining, describing the term as
mainly consisting of the processing of separation, consolidation, and deformation of materials by one atom or a
molecule.7 Later, in the 1980s, Eric K. Drexler popularized the word nanotechnology through his biologically
inspired bottom-up visions of building molecular machines,
the so-called molecular assemblers, which could guide
chemical reactions by positioning reactive molecules with
atomic precision.8, 9 However, Drexlers sweeping visions
and theories of grey goo have proved highly controversial,10 and nanotechnology as a scientific field was
established in a way that diverged from Drexlers original vision of molecular manufacturing. Today, the goal for
nanotechnology research is not to immediately create synthetic molecular assemblers but rather to understand the
unique properties of the nanoscale and use that knowledge to create new, high-performance materials, devices,
and processes.
Although nanotechnology was first theorized by Feynman,6 then coined by Taniguchi,7 and then later popularized
by Drexlers controversial vision of molecular manufacturing,810 a flurry of activity in the field was spurred by the
invention of the scanning tunneling microscope (STM), the
atomic force microscope (AFM), and the first manipulation of atoms. In 1981, Gerd Binnig and Heinrich Rohrer
developed the STM at IBMs laboratories in Switzerland.11
This microscope enabled atomic scale characterization of
conducting surfaces and opened the possibility of imaging

and mapping nanoscale materials. The next leg of the standard story jumps us to 1986 with the development of
the AFM, which enabled mapping on nonconducting surfaces.12 In 1990, Don Eigler and Erhard Schweizer invented
a technique for picking up individual atoms using the tip
of an STM and depositing them in patterns onto a surface.13 They used the technique to position 35 individual
xenon atoms on a nickel metal surface to spell out their
corporate logo IBM, demonstrating how atoms could be
moved and positioned. The technique has since been used
to create a variety of structures out of many different
atoms.
In addition, the advances in supramolecular chemistry, the constant pressure in device miniaturization in
the electronic industry, and the development of materials
with nanoscale dimensions, such as fullerene,14 metallic
nanoparticles,15 graphene,16 carbon nanotubes (CNTs),17
have contributed to the rapid growth of the field in the
past 30 years.

1.2

Current researchnanotools,
nanostructured materials, and nanodevices

Current research into nanotechnology may be divided into


three broad categories: nanotools, nanostructured materials,
and nanodevices. The various components of these categories are schematically illustrated in Figure 1.
Nanotools are a collection of methods and techniques
employed to produce and evaluate nanostructured materials
and nanodevices. The fabrication of nanostructured materials has been led by the development of new synthetic
methods and major advances in supramolecular chemistry.

Information
technology

Medicine and
health
Nanotools
Supramolecular chemistry

Food and
nutrition

Synthetic methods
Surface science
Nanolithography
Analytic tools
Computer simulations

Biotechnology
and agriculture

Textiles and
clothing

Figure 1
world.

Nanomaterials

Nanodevices

Nanoelectronics
Nanoparticles
Spintronics
Nanowires
Nano-optoelectronics
Fullerene
Nanosensors
Graphene
Drug delivery systems
Carbon nanotubes
Nanocomposites
Thin solid films
Nanopatterned surface
Supramolecular systems

Transportation
and aerospace

Energy and
environment

National security
and defense

Schematic illustration showing how nanotechnology and its nanotools, nanomaterials and nanodevices are impacting our

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches


With advancements in synthetics methods, nanostructured
materials with exquisite control over size, morphology, and
functionality have been achieved.18 Remarkable progress in
the field of supramolecular chemistry has enabled the design
of molecular components to interact favorably with each
other in such a way that they can self-assemble, through
noncovalent interactions, into larger, well-defined entities
on the nanoscale with tailored properties.19 Supramolecular
approaches have been successfully employed to construct,
among other supramolecular systems, molecular devices
and machines (see Photochemically Driven Molecular
Devices and Machines, Nanotechnology), functional, biologically derived supramolecular systems (see Biologically
Derived Supramolecular Materials, Nanotechnology),
polymeric nanomaterials (see Self-Assembled Nanoparticles, Nanotechnology), supramolecular hybrid nanomaterials (see Supramolecular Hybrid Nanomaterials as
Prospective Sensing Platforms, Nanotechnology), and
supramolecular nanoparticles (see Supramolecular Nanoparticles for Molecular Diagnostics and Therapeutics, Nanotechnology).
Furthermore, advances in surface science have led to the
creation of techniques to fabricate nanoscale molecularassembly structures onto a variety of substrate surfaces,
including self-assembled monolayers (SAMs),20 LangmuirBlodgett technique,21 and two-dimensional supramolecular assembly (see Two-Dimensional Supramolecular Chemistry, Nanotechnology). Many lithographic techniques have emerged for patterning surfaces with nanometer resolution (see Nanolithography, Nanotechnology).
Photolithography, electron-beam lithography, soft lithography, nanoimprinting, colloidal lithography, and dip-pen
nanolithography are but a few examples of such techniques. Analytic tools, such as the AFM (see Atomic Force
Microscopy Measurements of Supramolecular Interactions, Nanotechnology), STM,11 and the near-field scanning optical microscope,22 have provided revolutionary
improvements in our ability to investigate the structures
and functions of nanostructured materials and nanodevices.
High-performance computer simulations based on advanced
mathematical and physical modeling are at present a necessary tool in the development, design, and understanding of
nanoscopic systems.23 With these powerful contemporary
toolsadvanced synthetic methods, supramolecular chemistry, surface science, nanolithography, new or improved
analytical techniques, and high-performance computer simulationsmany novel nanostructured materials and nanodevices have been constructed and characterized. The
improvement of existing and the development of new nanotools is an unrenounceable condition for further progress
in nanotechnology.
Nanostructured materials can be defined as any material
that has structured components with at least one dimension

<100 nm. Encompassed by this class of materials are nanoparticles, nanowires, nanorods, nanocapsules, nanofibers,
nanotubes, nanocomposites, nanostructrured surfaces, and
thin solid films with nanoscale thickness.18 In many cases,
the properties of such nanostructured materials can be very
different from those of corresponding bulk materials, and
desirable novel electrical, mechanical, chemical, optical,
magnetic, thermal, chemical, and/or biological properties
may be obtained. For instance, quantum effects can begin to
dominate the behavior of matter at the nanoscale, affecting
the optical, electrical, and magnetic properties of materials.18 In tandem with their small size, nanostructured materials may pass biological barriers, which are inaccessible to
larger materials,24 or acquire superior chemical properties
such as enhanced or new reactivities.25
Nanoparticles are one of the most prominent groups
of nanostructured materials, and examples include carbonbased nanoparticles such as fullerenes14 and metallic15
(e.g., Au, Ag, Pt), semiconducting26 (e.g., CdSe, CdS,
ZnS, GaAs), magnetic27 (e.g., Fe3 O4 ), polymeric (see SelfAssembled Nanoparticles, Nanotechnology), and hybrid28
(e.g., coreshell) nanoparticles. Synthetic methods have
also been developed to achieve other morphologies such
as nanowires, nanorods, nanocapsules, nanofibers, and nanotubes.18 Among the different nanotubes, CNTs, which
were discovered in 1991,17 have received much attention due to their unique physical/mechanical, electronic,
chemical, optical, and other properties (see Advances in
Supramolecular Chemistry of Carbon Nanotubes, Nanotechnology).
Nanocomposites are a class of hybrid materials that have
at least one component with nanoscale dimensions, with
the most common involving polymers with either inorganic nanoparticles (see Magnetically Responsive SelfAssembled Composite Materials, Nanotechnology) or
CNTs.29 These materials can exhibit markedly enhanced
mechanical and other properties compared to conventional
composite and noncomposite materials. Nanopatterned surfaces (see Nanolithography, Nanotechnology) and thin
solid films, such as LangmuirBlodgett films (LBFs),21
SAMs,20 and two-dimensional supramolecular assemblies
(see Two-Dimensional Supramolecular Chemistry, Nanotechnology), are two other groups of nanostructured materials. SAMs and LBFs will be discussed in greater detail
later in this article. An extreme limit of nanoscaling is
monolayer graphene (see One-Dimensional Nanostructures of Molecular Graphenes, Nanotechnology), which
is only one atomic layer thin. Since its discovery in 2004, it
has attracted great interest because of its outstanding electrical and mechanical properties.16 Undoubtedly, the main
driving force behind research and development in nanostructured materials is the expectation of accessing novel
and unique material properties and functionalities, with the

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology

potential to add great value to a wide range of everyday


products.
Nanodevices are systems with nanostructured materials
that carry out specific functions with either improved performance or new attributes. Recent years have witnessed
the emergence of new device paradigms based on nanostructured materials, including nanoelectronic devices, nanooptoelectronic devices, spintronic devices, nanosensors, and
drug and gene delivery systems.
Nanoelectronics (see Nanoelectronics, Nanotechnology)
have already revolutionized the semiconductor device
industry, in the form of integrated circuits with nanoscale
transistors that pack more and more functionality into compact devices. Since the development of the first integrated
circuits over 60 years ago, the semiconductor industry has
seen the size of transistor devices decrease by a factor of 2
in every 18 months,30 a trend that was first pointed out by
Gordon Moore in the 1960s and is referred to as Moores
law.31 The miniaturization has primarily been achieved by
optimizing the photographic technique such as going from
using visible light sources to UV and currently stands at
22 nm; however, limits of the process are fast approaching,
where further miniaturization will no longer be possible
with the current setup. Therefore, to continue the trend
set by Moores law, fairly dramatic changes in the way
transistors are designed and operate are required. Emerging nanoelectronic devices, such as Si nanowire field-effect
transistors (FETs),32 carbon-nanotube FETs,33 and graphene
nanoribbon FETs,34 are providing new opportunities for
very-large-scale integration circuits in order to achieve continuing cost minimization and performance improvement,
while simultaneously enabling the extension of Moores
law well into the next decade and beyond.35, 36 Another
extremely important area of research in nanoelectronics is
molecular electronics, which is the utilization of a single
molecule or group of molecules as key active components
in electronic devices.37 Molecular electronic devices are
expected to not only address the ultimate limits of possible miniaturization but also offer unlimited possibilities for
technological development due to the potentially diverse
electronic functions of the component molecules, which can
be tailored by chemical design and synthesis.38
Advances in spintronics have already made their way into
magnetic hard discs, allowing for a huge increase in their
storage capacity.39 In this rapidly growing field, researchers
are paving the way for spin computers, which will use
the electrons spin state to store and process vast amounts
of information more quickly while requiring less energy
and generating less heat. Nanostructured materials, such as
graphene40 and lithographically nanopatterned surfaces,41
are two promising candidates for use in spin computers.
Progresses in the synthesis of semiconducting nanostructures, such as nanowires and nanoparticles, are leading to

the development of ultracompact and power-efficient optoelectronic devices such as photodetectors42 and lasers.43
Nanowires have also been harnessed as building blocks
for the construction of nano-light emitting diodes (LEDs)
with the ability of emitting in many different wavelengths
depending on nanowire composition.44, 45 The conventional,
low efficient photovoltaic cells, which have restricted largescale production of electrical energy, can be now replaced
by nanosolar cells with much higher efficiencies and lower
costs.46
Nanosensor devices incorporating nanostructured materials as sensing probes (see Supramolecular Hybrid Nanomaterials as Prospective Sensing Platforms, Nanotechnology) continue to advance toward commercialization for a
number of different applications, including, but not limited
to, medical diagnostics, food safety, environmental protection, national security, and aerospace.3 While in the realm
of medicine nanosensors can detect the onset of disease, in
the area of national security they could be used to detect
radioactive materials or biological warfare agents (such as
anthrax and smallpox). The detection principle in nanosensors is based on measuring the physical and chemical property changes, such as electrical, optical, magnetic, mass, and
pH value, derived from the interaction of the target analyte
with the nanodetection device. Today, numerous gas, chemical, and biosensors are being developed with substantially
smaller size, lower weight, more modest power requirements, greater sensitivity, better specificity, and, in some
cases, with the ability to detect multiple analytes at the same
time through high-density nanoarrays.47 In order to achieve
such powerful nanosensor capabilities, researchers have
been exploiting numerous nanostructured materials, including nanoparticles,24 CNTs,47 nanowires,48 , nanoscaled thin
films,49 and nanocantilevers.50 For instance, over the past
two decades, the evolution of fluorescent semiconductor
nanoparticles known as quantum dots (QDs) has helped to
usher in a new era in biomedical research and applications
(i.e., as labels for the detection of DNA and immunosensing
of disease biomarkers, and to improve biomedical imaging).
The defining features of QDs are their exceptional photostability and size-dependent tunable photoluminescence.
QDs emit different colors depending on their size, allowing
them to be used to color-code and track different cell processes, thereby providing high-resolution cellular imaging,
long-term observation of individual molecules, and their
movement within cells.24
Liposomal drug delivery systems and delivery systems
based on drug conjugates are two classes of nanotechnology
therapeutic products that are used in clinical practices.51
Present research on nanotechnology-based drug and gene
delivery systems is focused on achieving targeted delivery
of drugs to specific cells or tissues, improved delivery of
poorly water-soluble drugs, multiple drug administration,

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches


monitoring drug delivery by combining therapeutic agents
with imaging biomarkers, and real-time read on the in vivo
efficacy of a therapeutic agent.51, 52 Nanodevices are also
expected to be developed for both diagnosis and therapy,
and such theragnostic devices hold great promises in personalized medicine (see Supramolecular Nanoparticles
for Molecular Diagnostics and Therapeutics, Nanotechnology).
Nanodevices are in some ways the most complicated
nanotechnological systems. They require the understanding
of fundamental phenomena, the synthesis of appropriate
nanostructured materials, the use of those materials to
fabricate functioning devices, and the integration of these
devices into working systems. Nevertheless, significant
progress is currently being reported, and development and
availability of increasingly sophisticated nanodevices are
greatly anticipated.

1.3

Possible applications of nanotechnology

The huge scope of nanotechnology is summed up in a


quote from a report by the National Science Foundation
in 1999,53 which quoted, nanoscience and technology will
change the nature of almost every human-made object in the
21st century. The fascinating and often unrivaled properties
of nanostructured materials and devices have been, and will
continue, to open new and sometimes unexpected fields of
application. Today, the widespread applications range from
information and communication technology, transportation
and aerospace, energy and environment, national security
and defence to healthcare and medicine, food and nutrition,
biotechnology and agriculture, textiles and clothing, and
many more (Figure 1).
Over the years, advances in the field of nanoelectronics,
which deals with the miniaturization of electronic devices,
have enabled the appearance of new products in a range
of areas, including consumer electronics (e.g., computers,
mobile phones, televisions, etc.), the automotive industry,
healthcare and environmental management.53 Nanoelectronics are expected to have an impact in many areas of our
lives, as more and more functions are integrated into everyday products. For instance, nanoelectronic devices could
be used to regulate energy use in buildings, while in cars
additional built-in electronics could allow for more assisted
driving.
Renewable energy technologies are often regarded as
clean or green energy since they are far less harmful to
the environment than conventional fossil fuel technologies.
Nanotechnology is offering a range of new opportunities,
such as nanosolar cells that would be energy-intensive
and far less expensive to make,46 solar panels capable of
tapping not only the visible light from the sun but also

from infrared light as well, thus significantly increasing


energy output.54 Wind, wave, and geothermal energy is
also expected to be harnessed more effectively using new
nanostructured materials and stored or delivered more
efficiently through nanotechnological advances in batteries
and hydrogen fuel cells.55 Nanotechnology is playing a
major role in the development of hydrogen fuel cells, which
is considered to be one of the most promising clean energy
conversion devices for a wide variety of power applications
ranging from portable and stationary power supplies to
transportation.4, 5
More recently, subfields of nanotechnology, such as
nanobiotechnology and nanomedicine, are contributing
toward the development of highly accurate and sensitive
early-stage diagnostic devices. For instance, biosensors and
molecular probes are capable of directly interacting with
the biological molecule and converting the interactions
into directly transduced or significantly amplified electrical
or electromagnetic signals.2, 56 Dramatic breakthroughs are
also expected in life sciences research that could contribute
to the treatment of a number of human diseases, including
cancer and neurodegenerative diseases, such as Alzheimers
and Parkinsons. Therapeutic fields, such as gene and drug
delivery, tissue engineering, and drug discovery, will also
benefit greatly from advances in nanobiotechnology and
nanomedicine.57, 58 For instance, one of the current challenges is developing multifunctional nanodevices that are
capable of targeting specific malignant cells, visualizing
their location in the body, killing primarily the cancer cells,
with minimal side effects on the bodys normal cells and
tissues, and monitoring treatment effects in real time.59

NANOFABRICATION

Nanofabrication methods can be categorized into two


groups: the top-down and bottom-up approaches. The
top-down approach revolves around fabrication via etching away bulk material to achieve the required smaller
structural architectures and this is generally achieved by
lithographic processes (see Nanolithography, Nanotechnology).60, 61 This process could be likened to sculpting
a block of stone to the required image. In contrast, the
bottom-up approach involves the structures being crafted
atom by atom or molecule by molecule through covalent
or supramolecular interactions6063 in a similar manner to
how a house is built brick by brick. Both approaches have
their merits and drawbacks, which will be discussed later.

2.1

Top-down approach

Presently, the top-down approach is dominantly used in


industry for the fabrication of many man-made materials,

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology

a prime example being the semiconductor industry,64, 65


where features of metal oxide semiconductor field effect
transistor (MOSFET) are imprinted onto a silica wafer
via a lithography-based procedure termed photolithography.64, 65 The technique is based on a projection printing
system, which is done in a device called a stepper, where
the features of the transistors are projected through a photomask onto a silicon wafer that has been prespinned with
a photoresist (light-sensitive) material using UV light. If
the photoresist is positive, the regions of the photoresist
that are exposed to UV light become soluble to a particular developing solvent and are washed away during the
developing step, leaving a pattern of raised features on the
wafer identical to the dark regions on the mask. Conversely,
for a negative resist, the regions of the photoresist that are
exposed to UV light become insoluble to a particular developing solvent and only the unexposed regions are washed
away during the developing step, leaving a pattern of raised
features on the wafer identical to the clear regions on the
mask. For more details on the fabrication of MOSFET, refer
to article Nanolithography, Nanotechnology.

2.1.1 Limitations of the top-down approach


Semiconductor technology is beginning to reach the limits of miniaturization, with Intel announcing in May 2011
another milestone as they demonstrated a production-ready
3D transistor technology for 22 nm called Tri-Gate. Currently, the photolithography process uses 193 nm wavelength of light to pattern the wafers, but both technical
and material limitations are envisaged to be approached
as smaller features are to be obtained.60 For instance,
quantum effects and defects formed during the patterning
process (see Nanoelectronics, Nanotechnology) will play
more dominant roles as smaller features are fabricated.60 If
smaller features are to be generated, new fluids, lens materials, and resist material with high index will be required
or a new generation of lithography techniques such as
extreme UV lithography (EUVL), which uses light with
wavelengths in the range 1050 nm, needs to be introduced
(see Nanoelectronics, Nanotechnology). If smaller wavelengths of light and hence higher energy photons are used,
it becomes deleterious to the materials used, such as the
focussing lenses, and resists layers.60

2.2

The bottom-up approach

Scientists curiosity to understand and mimic how biological architectures are preprogrammed to self-assemble
and self-organize into ordered, yet dynamic and functional,
structures through supramolecular interactions (i.e., hydrogen bonding, van der Waals, electrostatic, interactions,

hydrophilichydrophilic, and hydrophobichydrophobic


interactions), in nature has been a major inspiration toward
the development of the bottom-up approach. In fact,
supramolecular chemistry provides an exciting tool that
combines the concepts of self-assembly and molecular
recognition for the fabrication of three-dimensional intelligent nanodevices. Similar to nature, artificial systems are
either responsive to external stimuli, such as electrical,
chemical/biochemical, temperature, and photons, or can
partake through intermolecular interactions with other isolated components to form functional materials. There are
a number of two-dimensional and three-dimensional selfassemblies as shown in Figure 2, which can be utilized
as the fundamentals for building novel nanotechnological devices. The nanofabrication and applications of the
three-dimensional self-assemblies in fields such as molecular diagnostics, therapy, and electronics are discussed
in the articles to follow (see Self-Assembled Nanoparticles, Magnetically Responsive Self-Assembled Composite Materials, Supramolecular Nanoparticles for
Molecular Diagnostics and Therapeutics, Advances in
Supramolecular Chemistry of Carbon Nanotubes, Nanotechnology) and hence will not be discussed further in this
article.

2.2.1 Two-dimensional thin solid films


Self-assembly of well-ordered two-dimensional ultrathin
films on conducting, semiconducting, and insulating surfaces provides a simple, cheap, and reproducible method of
obtaining nanoscale films with a wide variety of functional
groups, which can be chemically manipulated.23, 6669 These
surfaces are increasingly being utilized as foundations for
building nanodevices such as sensors and for electronic
applications.23, 66, 67 Three types of ultrathin films used for
such potential applications are LBFs,21, 70, 71 SAMs,20, 72 and
two-dimensional supramolecular assemblies. The following discussion excludes two-dimensional supramolecular
assemblies since they are covered in great detail in article Two-Dimensional Supramolecular Chemistry, Nanotechnology.
LangmuirBlodgett films (LBFs)
LBFs are the first practical example of ordered molecular
assemblies.21, 70, 71 The approach involves the transfer of
monolayers from a liquidair interface, which is denoted
as a Langmuir film onto a solid substrate. The Langmuir
films are produced by amphiphiles, which are molecules
that have a hydrophobic end and a hydrophilic end.70, 71 The
amphiphile molecule is deposited onto a water subphase
avoiding the formation of multilayers in the Langmuir film.
Initially, the distances between the molecules in the phase
are large relative to the molecular dimensions and the film

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches

Self-assembly entities

Two-dimensional
ultrathin films

Self-assembled
monolayers

LangmuirBlodgett
films

Two-dimensional supramolecular assemblies

Figure 2

Three-dimensional
assemblies

Metallic,
semiconducting,
magnetic, polymeric,
supramolecular and
hybrid nanoparticles

Nanorods

Vesicles

Micelles

Fullerenes

Carbon
nanotubes

Schematic representation of examples of two-dimensional and three-dimensional self-assemblies.

is disordered and known as two-dimensional gas phase.


The film is compressed to bring the molecules closer to
each other. The area per molecule decreases and the surface
pressure increases when the distance between the molecules
approaches molecular dimensions. The profile between
the surface pressure against molecular area is known as
the pressure-area isotherm.70, 71 The pressure-area isotherm
provides information on the stability of the monolayer
formed, at the liquidair interface, the orientation of the
molecules in the two-dimensional system, phase transitions,
and conformational transformation.
There are two possible methods of transferring the monolayers from the liquidair interface onto a solid substrate. The most conventional method that is used is the
vertical deposition of the substrate (LangmuirBlodgett
method).70, 71 The second method is the horizontal deposition of the substrate onto the Langmuir film, which is
known as the LangmuirSchaefer method.70, 71 In the LangmuirBlodgett method, the monolayer is transferred onto
the substrate, as the substrate is either emersed (retraction
or upstroke) or immersed (dipping or downstroke) into the
Langmuir film, as shown in Figure 3(a).70, 71 When the substrate surface is hydrophilic, the monolayer is transferred
as the substrate is retracted. However, if the substrate is
hydrophobic the monolayer is transferred as the substrate
is immersed. The speed at which the substrate is dipped
and retracted is important for a quality film to be produced. One of the unique benefits of LBFs is that multilayers can be formed in a controlled manner. The number

of layers formed depends on the number of immersions


and emersions. There are three possible multilayer structures that can be formed, which are X-, Y-, and Z-type
modes (Figure 3b). X and Z modes are formed only in the
downstroke and upstroke, respectively, whereas Y mode is
formed both in the upstroke and downstroke and is the most
stable and commonly formed mode.
In the LangmuirSchaefer method, a flat substrate is
placed horizontally onto a compressed monolayer on the
liquidair interface. When the substrate is lifted horizontally and separated from the water subphase, the monolayer
is transferred onto the substrate (Figure 4). The method is
useful to transfer viscous films as well as monolayers of
lipids and proteins.73
Self-assembled monolayers
SAMs are two-dimensional quasi-ordered molecular assemblies, which are formed via adsorption of molecules from
solution. SAMs have been increasingly used over LBFs
as they offer a number of advantages.20, 72 First and foremost, SAMs are more stable than the LBFs stemming
from the molecules being chemisorbed onto the substrate,
whereas in LBFs the molecules are physisorbed. Secondly,
SAMs provide more flexibility in the molecular design
of the molecules because the molecules that form LBFs
need to be amphiphiles, whereas in SAMs such a need is
not required. Hence, the surfaces of the monolayers are
easily tunable, and a wide range of terminal functionalities have been studied (e.g., carboxylic acid, amine, nitro,

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology
Barrier

Substrate holder

Trough
(a)
Substrate

(a)

Direction of the
barrier movement

Hydrophilic substrate

X-type film
Hydrophobic
hydrophilic
interaction

(b)
Raising of the
solid substrate

Y-type film
Hydrophilic
hydrophilic
interaction

Z-type film
Hydrophilic
hydrophobic
interaction
(b)

Figure 3 Schematic representation of (a) vertical deposition of


the Langmuir film onto the solid substrate and (b) possible
multifilms that can be formed by the vertical deposition process:
X-type, Y-type, and Z-type.

hydroxyl, and methyl groups) and exploited to provide


control over surface properties such as wettability23, 67 and
adhesion.68, 69
The molecular structure of the adsorbate or surfactant
molecules can be divided into three components, head
group, backbone, and terminal group (Figure 5). The head
group is the anchor, which binds the adsorbate to the
substrate. The choice of head group depends on the
substrate used, as different groups have varied affinity for
particular substrates. The most common head groups are
thiols (SH) on gold and silanes on silica substrates.20, 72 The
backbone takes a major role in the molecular ordering and
the thermal stability of the SAM formed.70 The backbone
connects the head group with the terminal group and is
generally made out of an aliphatic chain and/or aromatic
components. Each molecule in the SAM interacts with
neighboring molecules through the backbone. Depending
on the groups in the backbone, the molecules can interact
by van der Waals or interactions, leading to relative
well-ordered molecular layers. The terminal group is the
surface group, which plays a crucial role in the properties of

(c)

Figure 4 Schematic representation of the LangmuirSchaefer


method. (a) Langmuir film, (b) the solid substrate is placed
horizontally on the Langmuir film, and (c) the solid substrate is
lifted with a LangmuirSchaefer film.

Surface group or
terminal group
Backbone
Head group

Figure 5 Cartoon representation of the molecular structure of


an adsorbate and how they are typically tilted at an angle of in
the monolayer.

the surface such as wettability and corrosion. The terminal


group also has an influence on the packing density in the
SAM.71
Organosulfur-based SAMs
Organosulfur-based SAMs
have been extensively studied and are well understood.
These types of molecules have shown to self-assemble on
a number of metal surfaces including Au, Ag, Cu, Pt, and
Fe,74 the preferred surface being clean and hydrophilic. Au
is the most commonly used for such type of SAMs, as
sulfur has a relative strong bond to Au. In addition, Au
does not have a stable oxide, hence is easy to handle in

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches

Adsorbates
physisorb

(a)
Adsorbates chemisorb
onto the substrate
S
S

(a)

S
S

Si
Si
Si
O Si O
O O O OO
Si
Si
Si
Si OO
O
O
O
O
O
O OO

S
S

(b)
Formation
of islands

(b)

Figure 6 Representation of the structures of (a) alkanethiolate


SAM on Au and (b) alkanesilane SAM on silica substrate, also
showing the cross-linking network between the molecules through
SiOSi bonds.

van der Waals


interaction
(c)
Self-assembled monolayer

ambient conditions. Several different types of organosulfur


molecules have been shown to form well-ordered monolayers on Au: alkanethiol,20, 72 dialkyl disulfide,75 dialkyl
sulfide,76 thioctic acid,77 and thiophene.78, 79 AlkanethiolAu SAMs (Figure 6a) are the most studied and understood.
The organosulfur-Au SAMs are generally deposited from
1 mM solution of the adsorbate molecule in an organic
solvent such as EtOH75, 76, 78 and CHCl3 .79 The deposition
process of alkanethiol-Au SAMs are well understood and is
thought to go through four stages. The first stage involves
the adsorbates physisorbing onto the surface (Figure 7a),
followed by chemical bonds forming between the adsorbate
and the substrate (Figure 7b). As the number of adsorbates
bound to the substrate increases, the adsorbates reorganize
through van der Waals and other interactions to form islands
(Figure 7c), eventually leading to the formation of a wellordered SAM (Figure 7d). The molecules in the SAM are
always slightly tilted from the normal of the substrate
(Figure 5) due to the optimization of the intermolecular
interactions between the molecules (i.e., van der Waals
interactions) and the trans conformation in alkyl chains in
the backbone.
Organosilicon-based SAMs
There are two types of
silicon-based adsorbates studied, alkyltrioxysilanes
(RSi(OR)3 )80 and alkyltrichlorosilanes (RSiCl3 ).20 These
are especially attractive for electronic applications because
the SAMs can be formed on silica or silicon wafers
(Si/SiO2 ). SAMs are formed on hydroxylated surfaces
(usually the native oxide). The SAM is afforded via a
condensation reaction between the organosilane, hydroxylated surface and the neighboring silane,80 due to the
extensive cross-linking between the molecules (Figure 6b).
The organosilane-Si/SiO2 SAMs are thermally more stable than organosulfur-Au SAMs. The major drawback for

(d)

Figure 7 Schematic representation of the deposition process of alkanethiol on Au. (a) Physisorption, (b) chemisorption,
(c) formation of islands, and (d) formation of SAM.

organosilane-based SAMs is their susceptibility to hydrolysis even in mild conditions.80, 81


Mixed SAMs
Molecular level control over the density
and spatial distribution of functional groups on surfaces
is important in a wide variety of applications, and, in
particular, for biomedical applications. In many areas of
medical and biological research, the functional groups
mediate the immobilization of biomolecules to surfaces
and their efficient immobilization not only requires that
biomolecules preserve their activity after immobilization
but also that biomolecules are presented on the surface at an
optimal density and spatial distribution, such that efficient
binding can occur between the immobilized biomolecules
and target species in solution. At present, mixed SAMs
offer the best option for controlling the density and spatial
distribution of the biomolecules on surfaces.
There have been a number of different approaches used
for obtaining the mixed SAMs. The simplest approach is
the formation of a monolayer via deposition of a substrate
in a mixed solution that contains an active (containing the
binding functional group) and inert surfactant (acts as the
spacer to distribute the functional group in the first surfactant).82 This method is simple, but has some limitations
such as the fact that the ratio of active surfactants in the
monolayer is rarely identical to the ratio of active surfactants in the solution due to the preferential adsorption
of one of the components, thus affecting control over the

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

10

Nanotechnology

density of the functional group on the mixed monolayers.83 Further, the formation of two-component monolayers
has been reported by some authors8486 to lead to phasesegregated mixed SAMs, that is, formation of two distinct
local domains, each of which was mostly composed of one
constituent surfactant. In order to address some of these
issues, asymmetric dialkyl sulfide87 and dialkyl disulfide88
SAMs have been prepared, which consist of the active
functional moiety on only one of the chains bound to
the sulfur. Although this approach gives a more homogeneous monolayer compared to the mixed SAMs mentioned above, the spacing between the active molecules
is restricted. Recently, Tokuhisa et al.89 proposed using
dendrons as spacers. The dendrons were bound to lipoic
acid through an ester linkage.89 After the formation of the
monolayer, the dendrons were removed by hydrolysis under
basic conditions and subsequently, a second surfactant was
deposited onto the vacant areas to form a mixed monolayer.
The spacing between the active sites was controlled by the
size or generation of the dendron used. Novel approaches
for molecular level control over the functional groups on
surfaces will continue to aid in the development of more
advanced surface materials.

switch between bioinert and bioactive states, under an external thermal-, photo-, chemical/biochemical-, or electricalinduced stimulus, to trigger capture or release of biological entities.56 However, existing switchable SAM surfaces
rely mostly on nonspecific interactions (i.e., hydrophobic/hydrophilic and electrostatic) of the biomolecules with
the active surface, thus lacking biospecificity and selectivity and substantially limiting the application potentialities of such surface systems. There are relatively few
reported examples in which specific biomolecular interactions have been dynamically controlled in response to
applied stimuli.9092 Electro-switchable oligopeptide SAM
surfaces have been successfully used to reversibly control
biomolecular interactions upon application of an electrical stimulus.92 In another example, a thermoresponsive
oligo(ethylene glycol) derivative has been exploited to control the affinity binding between surface-tethered biotin
groups and streptavidin.91 Albeit the progress and scientific advances in the field, exciting future developments
are ahead of us. One of the major challenges in the
field of switchable biological surfaces today is the design
of new and more versatile surfaces with tunable biospecific interactions. For a detailed discussion on the progress
made on stimuli-responsive surfaces, refer to the review by
Mendes.56

Switchable SAMs
Recently, there has been an increasing activity in fabricating stimuli-responsive SAM surfaces,
where the surface properties are manipulated through external stimulus providing an on and off switch for regulating the immobilization of biological and chemical particulates (Figure 8).56 Such SAM surfaces potentially provide a wide range of applications to many areas in science and technology, especially in the life sciences. These
SAM surfaces enable modulation of biomolecule activity, protein immobilization, and cell adhesion at the liquidsolid interface for applications including biofouling,
cell culture, regenerative medicine, and tissue engineering.56 This field is in its infancy and early examples of
switchable biological SAM surfaces include surfaces that

2.2.2 Hybrid approach: Combination of the


top-down and bottom-up approaches
Currently, a significant challenge in nanotechnology is
to spatially self-organize self-assembled nanoscale components (such as nanoparticles, CNTs, proteins, cells, organic
molecules, polymers, etc.) onto surfaces to fabricate functional nanostructured systems for electronic, optoelectronic,
or sensing applications.93, 94 Precise control over the relative position and orientation of the nanocomponents is
frequently required in such systems to obtain useful properties. Moreover, the integration and the stability of interfaces to these nanostructures from the micron-length and
macroscopic scales are key to the success of future applications.
Stimulus

Chemical / Biochemical
Thermal
Electric
Optical

Inactive surface

Active surface

Figure 8 Schematic representation depicting the range of stimuli and how they can be used to modulate the binding of particulates
on stimuli responsive-based self-assembled monolayers.
Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches


Nanostructuring surfaces of macroscopic materials by
top-down nanolithography techniques, and subsequently
building into the third dimension utilizing bottom-up
self-organization of self-assembled nanoentities, is an
attractive approach for creating such a bridge between
macroscopic systems and the nanoscale dimensions that
many modern technologies demand. With this in mind,
different hybrid top-down and bottom-up fabrication approaches have been investigated to create three-dimensional
nanostructured surfaces.93, 95, 96 Several strategies rely on
the formation of an ultrathin film (e.g., SAMs) on the surface material of interest, followed by chemical transformation/damage on the film caused by a lithographic technique
(see Nanolithography, Nanotechnology), and finally the
immobilization of nanoscale components on the modified
surfaces through either covalent or supramolecular interactions to form functionalized three-dimensional nanostructures. SAMs are attractive ultrathin organic films for
patterning high-resolution features on a number of technologically relevant surface substrates. The attraction of
these systems as ultrathin resists is driven primarily because
SAMs eliminate depth of focus, transparency issues and
can be prepared with a discrete number of well-defined
chemical functional groups to permit further nanoscale
materials attachment. Currently, patterning SAMs can be
achieved by several different types of lithographic techniques, which include (i) stamping or moulding methods,
that is, soft lithography,64 (ii) scanning probe lithography (SPL)-based techniques such as dip-pen nanolithography,98, 99 nanografting,100, 101 and nanoshaving,102 and (iii)
radiative techniques that include ultraviolet/visible (UV/vis)
light, X-rays, and electron-beam. Generation of threedimensional nanostructures by self-organization of various self-assembled nanoscale components onto nanopatterned SAM surface templates has also been demonstrated.93, 9597 The integration of the top-down and bottomup methodologies is representing a new paradigm for
Table 1

11

creating nanostructured materials with high degrees of


control, and is being actively pursued by a number of
groups within the scientific and engineering communities.9397

CONCLUSION AND OUTLOOK

Nanotechnology is going to play an increasingly important


role in a number of sectors, addressing important issues
such as health, energy, environment, transportation, water,
food, and security. For instance, in medicine, nanotechnology has a role to play in developing novel, highly
accurate and sensitive early-stage diagnostic devices, as
well as providing novel methodologies for the treatment
of chronic (such as diabetes) or life threatening diseases
such as cancer through gene therapy or drug delivery.
Both the top-down, which relies on dimensional reduction
through selective etching and various nanoimprinting techniques, and bottom-up methods, which assemble atoms or
molecules into nanostructured materials, in several cases
through use of supramolecular chemistry, are at the heart
of such developments. As a result of the numerous fundamental breakthroughs made in the past three decades,
neither the top-down nor the bottom-up approach is superior
at the moment; each has its advantages and disadvantages
(Table 1).
The top-down nanofabrication has been used with obvious success by the semiconductor industry for several
decades now, with physicists and engineers manipulating
progressively smaller pieces of matter by photolithography and related techniques, but the top-down approach
is quickly reaching its physical and economic limits. On
the other hand, the bottom-up nanofabrication offers ultimate limits of miniaturization, opens virtually unlimited
possibilities concerning the design and construction of functional nanostructured materials, and has the potential to be

Summary of the merits and drawbacks for the top-down and bottom-up approaches.

Advantages

Top-down approach

Bottom-up approach

Already well understood and have established techniques

Self-assembly provides a simple, fast, and low-cost


method for producing nanostructured materials
Offer ultimate limits of miniaturization

Provides control and precision when patterning the surfaces


through lithography
Procedures reproducible

Disadvantages

More sensitive to defects as features become smaller

Opportunities open to the fabrication of a wider range


of functional nanostructured materials by chemical
synthesis
At present, the mastery of self-assembly is limited
to fairly simple nanostructured materials, not being
able, for example, to create integrated devices

Tighter tolerance as features become smaller


More expensive as compared to the self-assembly methods
Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

12

Nanotechnology

more cost-effective than top-down nanofabrication. However, at present our ability to build nanostructured materials
from the bottom-up approach is fairly limited in scope,
even though chemical synthesis has been developed to a
breathtaking level of sophistication. The integration of the
sophisticated techniques of the top-down and bottom-up
approaches is an exciting development, which is offering
a unique opportunity to fabricate complex nanostructured
nanomaterials with high degrees of control and significantly expanding the possibilities of nanofabrication and
functions.
We can peer into the future with reasonable confidence. We can be confident that we will witness many
breakthroughs based on bottom-up approaches in the next
decades, leading to nanostructured materials with novel
and unique material properties and functionalities, and to
increasingly sophisticated nanodevices. While current indications are that bottom-up nanofabrication methods will not
completely replace top-down nanofabrication techniques, in
the decades to come we will see more applications originating either from bottom-up techniques alone or from hybrid
approaches combining the strengths of bottom-up and topdown methods.

ACKNOWLEDGMENTS
The authors thank the Leverhulme Trust (F/00094/AW) for
their support.

12. G. Binnig, C. F. Quate, and C. Gerber, Phys. Rev. Lett.,


1986, 56, 930.
13. D. M. Eigler and E. K. Schweizer, Nature, 1990, 344, 524.
14. H. W. Kroto, J. R. Heath, S. C. Obrien, et al., Nature, 1985,
318, 162.
15. M. Brust, M. Walker, D. Bethell, et al., J. Chem. Soc. Chem.
Commun., 1994, 801.
16. K. S. Novoselov, A. K. Geim, S. V. Morozov, et al., Science, 2004, 306, 666.
17. S. Iijima, Nature, 1991, 354, 56.
18. C. N. R. Rao, A. Muller, and A. K. Cheetham, eds, The
Chemistry of Nanomaterials: Synthesis, Properties and Applications, Wiley-VCH Verlag GmbH, Germany, 2004.
19. B. H. Northrop, A. B. Braunschweig, P. M. Mendes, et al.,
in Handbook of Nanoscience, Engineering, and Technology,
CRC Press, 2007.
20. A. Ulman, Chem. Rev., 1996, 96, 1533.
21. J. Sagiv, J. Am. Chem. Soc., 1980, 102, 92.
22. D. W. Pohl, W. Denk, and M. Lanz, Appl. Phys. Lett., 1984,
44, 651.
23. W. K. Liu, E. G. Karpov, S. Zhang, and H. S. Park, Comput. Meth. Appl. Mech. Eng., 2004, 193, 1529.
24. X. Michalet, F. F. Pinaud, L. A. Bentolila, et al., Science,
2005, 307, 538.
25. K. J. Klabunde and R. S. Mulukutla, in Nanoscale Materials
in Chemistry, ed. K. J. Klabunde, John Wiley & Sons, Inc.,
New York, 2002.
26. N. Peyghambarian, B. Fluegel, D. Hulin, et al., IEEE J.
Quantum Electron., 1989, 25, 2516.
27. F. Yang, M. A. Zhang, W. He, et al., Small, 2011, 7, 902.
28. U. Banin, M. Bruchez, A. P. Alivisatos, et al., J. Chem.
Phys., 1999, 110, 1195.

REFERENCES

29. T. T. Tung, T. Y. Kim, and K. S. Suh, Org. Electron., 2011,


12, 22.

1. E. Tasciotti, J. Sakamoto, and M. Ferrari, Nanomedicine,


2009, 4, 619.

30. P. Ball, Nature, 2000, 406, 118.

2. C. M. Niemeyer and C. A. Mirkin, eds, Nanobiotechnology,


Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2004.

32. Y. Cui, Z. H. Zhong, D. L. Wang, et al., Nano Lett., 2003,


3, 149.

3. S. Shelley, Chem. Eng. Prog., 2008, 104, 8.

33. S. Fregonese, C. Maneux, and T. Zimmer, IEEE Trans. Electron Devices, 2011, 58, 206.

4. N. Cele and S. S. Ray, Macromol. Mater. Eng., 2009, 294,


719.

31. G. E. Moore, Electronics, 1965, 38, 114.

5. R. Devanathan, Energy Environ. Sci., 2008, 1, 101.

34. X. R. Wang, Y. J. Ouyang, X. L. Li, et al., Phys. Rev. Lett.,


2008, 100 (Art No.: 206803).

6. R. P. Feynman, Eng. Sci., 1960, 23, 22.

35. P. Chatterjee, EDN, 2011, 56, 18.

7. N. Taniguchi, Proc. Intl. Conf. Prod. Eng. Tokyo, Part II,


Japan Soc. Precision Eng. 1974, 18.

36. W. J. Rao, C. M. Yang, R. Karri, and A. Orailoglu, Computer, 2011, 44, 46.

8. K. E. Drexler, Proc. Natl. Acad. Sci. U. S. A., 1981, 78, 5275.

37. A. H. Flood, J. F. Stoddart, D. W. Steuerman, and J. R.


Heath, Science, 2004, 306, 2055.

9. K. E. Drexler, Engines of Creation: The Coming Era of


Nanotechnology, 1st edn, Anchor Press/Doubleday, Garden
City, New York, 1986.

38. H. Song, M. A. Reed, and T. Lee, Adv. Mater., 2011, 23,


1583.

10. R. Baum, Chem. Eng. News, 2003, 81, 37.

39. D. Grundler, Phys. World , 2002, 15, 39.

11. G. Binnig, H. Rohrer, C. Gerber, and E. Weibel, Phys. Rev.


Lett., 1982, 49, 57.

40. W. Han, K. Pi, K. M. McCreary, et al., Phys. Rev. Lett.,


2010, 105, 167202.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

Nanotechnology: the top-down and bottom-up approaches

13

41. M. V. Costache and S. O. Valenzuela, Science, 2010, 330,


1645.

69. K. M. Roth, J. S. Lindsey, D. F. Bocian, and W. G. Kuhr,


Langmuir, 2002, 18, 4030.

42. O. Lupan, L. Chow, G. Y. Chai, et al., Phys. Status Solidi


A-Appl. Mater., 2008, 205, 2673.

70. D. R. Talham, Chem. Rev., 2004, 1004, 5479.

43. S. Strauf, K. Hennessy, M. T. Rakher, et al., Phys. Rev.


Lett., 2006, 96 (Art. No.: 127404).
44. F. Qian, S. Gradecak, Y. Li, et al., Nano Lett., 2005, 5, 2287.
45. Y. Huang, X. F. Duan, and C. M. Lieber, Small , 2005, 1,
142.
46. L. Tsakalakos, J. Balch, J. Fronheiser, et al., Appl. Phys.
Lett. 2007, 91 (Art. No.: 233117).

71. B. D. Gates, Q. Xu, M. Stewart, et al., Chem. Rev., 2005,


105, 1171.
72. J. C. Love, L. A. Estroff, J. K. Kriebel, et al., Chem. Rev.,
2005, 105, 1103.
73. A. Ulman, in An Introduction to Ultrathin Films: From
Langmuir-Blodgett to Self-Assembly, Academic Press, San
Diego, 1991, p. 101.

47. N. Sinha, J. Z. Ma, and J. T. W. Yeow, J. Nanosci. Nanotechnol., 2006, 6, 573.

74. A. Ulman, in An Introduction to Ultrathin Films: From


Langmuir-Blodgett to Self-Assembly, Academic Press, San
Diego, 1991, p. 237.

48. L. De Vico, M. H. Sorensen, L. Iversen, et al., Nanoscale,


2011, 3, 706.

75. C. Jung, O. Dannenberger, Y. Xu, et al., Langmuir, 1998,


14, 1103.

49. M. K. Kumar, L. K. Tan, N. N. Gosvami, and H. Gao, J.


Phys. Chem. C , 2009, 113, 6381.

76. E. B. Troughton, C. D. Bain,


Langmuir 1988, 4, 365.

50. P. G. Datskos and T. Thundat, J. Nanosci. Nanotechnol.,


2002, 2, 369.

77. J. Madoz, B. A. Kuznetzov, F. J. Mederana, et al., J. Am.


Chem. Soc., 1997, 119, 1043.

51. O. C. Farokhzad and R. Langer, ACS Nano, 2009, 3, 16.

78. J. Noh, E. Ito, K. Nakajima, et al., J. Phys. Chem. B , 2002,


106, 7139.

52. M. Solomon and G. G. M. DSouza, Curr. Opin. Pediatr.,


2011, 23, 215.
53. M. C. Roco, S. Williams, and P. Alivisatos, eds, Nanotechnology Research Directions: Vision for Nanotechnology in the
Next Decade, U.S. National Science and Technology Council,
Washington, DC, 1999.
54. S. A. McDonald, G. Konstantatos, S. G. Zhang, et al., Nat.
Mater., 2005, 4, 138.
55. A. S. Arico, P. Bruce, B. Scrosati, et al., Nat. Mater., 2005,
4, 366.

G. M. Whitesides,

et al.,

79. P. Iqbal, K. Critchley, S. Bequm, et al., J. Exp. Nanosci.,


2006, 1, 143.
80. N. Choi, T. Ishida, A. Inoue, et al., Appl. Surf. Sci., 1999,
144145, 445.
81. D. Appelhans, D. Ferse, H.-J. Adler, et al., Colloids Surf. A
Physicochem. Eng. Asp., 2000, 161, 203.
82. E. Ostuni, B. A. Grzybowski, M. Mrksich, et al., Langmuir,
2003, 19, 1861.

56. P. M. Mendes, Chem. Soc. Rev., 2008, 37, 2512.

83. P. E. Laibinis, M. A. Fox, J. P. Folkers,


Whitesides, Langmuir, 1991, 7, 3167.

57. M. Youns, J. D. Hoheisel, and T. Efferth, Curr. Drug Targets, 2011, 12, 357.

84. K. Tamada, M. Hara, H. Sasabe, and W. Knoll, Langmuir,


1997, 13, 1558.

58. J. A. Barreto, W. OMalley, M. Kubeil, et al., Adv. Mater.,


2011, 23, H18.

85. D. Hobara, M. Ota, S. Imabayashi, et al., J. Electroanal.


Chem., 1998, 444, 113.

59. K. J. Cho, X. Wang, S. M. Nie, et al., Clin. Cancer Res.,


2008, 14, 1310.

86. T. Sawaguchi, Y. Sato, and F. Mizutani, J. Electroanal.


Chem., 2001, 496, 50.

60. B. K. Teo and X. H. Sun, J. Cluster Sci., 2006, 17, 529.


61. N. A. Ochekpe, P. O. Olorounfemi, and N. C. Ngwieluka,
Trop. J. Pharm. Res., 2009, 8, 265.
62. H.-J. van Manen, T. Autetta, B. Bordi, et al., Adv. Funct.
Mater., 2002, 12, 811.
63. V. Lavayer, S. B. Newcomb, and C. M. S. Torres, Nat. Nanotechnol., 2009, 4, 239.
64. Y. Xia and G. M. Whitesides, Angew. Chem. Int. Ed., 2006,
16, 3997.
65. H.-J. Kim, K.-J. Kim, and D.-S. Kwok, Qual. Reliab. Eng.
Int., 2010, 26, 765.
66. P. M. Mendes, S. Jacke, K. Critchley, et al., Langmuir,
2004, 20, 3766.
67. R. K. Smith, P. A. Lewis, and P. S. Weiss, Prog. Surf. Sci.,
2004, 75, 1.
68. H. Imahori, M. Arimura, T. Hanada, et al., J. Am. Chem.
Soc. 2001, 123, 335.

and

G. M.

87. F. L. Callari and S. Sortino, J. Mater. Chem., 2007, 17, 4184.


88. H. Akiyama, K. Tamada, J. Nagasawa, et al., J. Phys. Chem.
B , 2003, 107, 130.
89. H. Tokuhisa, J. Liu, K. Omori, et al., Langmuir, 2009, 25,
1633.
90. D. B. Liu, Y. Y. Xie, H. W. Shao, and X. Y. Jiang, Angew.
Chem. Int. Ed., 2009, 48, 4406.
91. H. M. Zareie, C. Boyer, V. Bulmus, et al., ACS Nano, 2008,
2, 757.
92. C. L. Yeung, P. Iqbal, M. Allan, et al., Adv. Funct. Mater.,
2010, 20, 2657.
93. P. M. Mendes and J. A. Preece, Curr. Opin. Colloid Interface Sci., 2004, 9, 236.
94. W. Lu and C. M. Lieber, Nat. Mater., 2007, 6, 841.
95. D. Mijatovic, J. C. T. Eijkel, and A. van den Berg, Lab
Chip, 2005, 5, 492.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

14

Nanotechnology

96. J. Y. Cheng, C. A. Ross, H. I. Smith, and E. L. Thomas,


Adv. Mater., 2006, 18, 2505.
97. A. Dhawan, Y. Du, D. Batchelor, et al., Small , 2011, 7, 727.
98. J. C. Huie, Smart Mater. Struct., 2003, 13, 264.
99. A. A. Tseng, A. Notargiacoma, and T. P. Chen, J. Vac. Sci.
Technol. B, 2005, 23, 877.

100. S. Xu and G.-Y. Liu, Langmuir, 1997, 13, 127.


101. J.-F. Liu, J. S. Cruchon-Dupeyrat, J. C. Garno, et al., Nano
Lett., 2002, 2, 937.
102. G.-Y. Liu, S. Xu, and Y. Qian, Accounts Chem. Res., 2000,
33, 457.

Supramolecular Chemistry: From Molecules to Nanomaterials, Online 2012 John Wiley & Sons, Ltd.
This article is 2012 John Wiley & Sons, Ltd.
This article was published in the Supramolecular Chemistry: From Molecules to Nanomaterials in 2012 by John Wiley & Sons, Ltd.
DOI: 10.1002/9780470661345.smc195

S-ar putea să vă placă și