Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Nanostructures
Nanostructures
Nanostructures
Ebook566 pages6 hours

Nanostructures

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Nanostructures covers the main concepts and fundamentals of nanoscience emphasizing characteristics and properties of numerous nanostructures.

This book offers a clear explanation of nanostructured materials via several examples of synthesis/processing methodologies and materials characterization. In particular, this book is targeted to a range of scientific backgrounds, with some chapters written at an introductory level and others with the in-depth coverage required for a seasoned professional.

Nanostructures is an important reference source for early-career researchers and practicing materials scientists and engineers seeking a focused overview of the science of nanostructures and nanostructured systems, and their industrial applications.

  • Presents an accessible overview of the science behind, and industrial uses of, nanostructures. Gives materials scientists and engineers an understanding of how using nanostructures may increase material performance
  • Targeted to a wide audience, including graduate and postgraduate study with a didactic approach to aid fluid learning
  • Features an analysis of different nanostructured systems, explaining their properties and industrial applications
LanguageEnglish
Release dateOct 21, 2016
ISBN9780323497831
Nanostructures

Related to Nanostructures

Titles in the series (97)

View More

Related ebooks

Electrical Engineering & Electronics For You

View More

Related articles

Reviews for Nanostructures

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Nanostructures - Osvaldo de Oliveira Jr

    Nanostructures

    Edited by

    Alessandra L. Da Róz

    Federal Institute of Education, Science and Technology of São Paulo Itapetininga, São Paulo, Brazil

    Marystela Ferreira

    Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    Fábio de Lima Leite

    Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    Osvaldo N. Oliveira Jr.

    São Carlos Institute of Physics

    University of São Paulo (USP)

    São Carlos, São Paulo, Brazil

    Table of Contents

    Cover

    Title page

    Copyright

    List of Contributors

    1: Basic Concepts and Principles

    Abstract

    1.1. Introduction

    1.2. Final Considerations

    List of Symbols

    2: Supramolecular Systems

    Abstract

    2.1. General Concepts of Supramolecular Systems

    2.2. Molecular Recognition

    2.3. Self-Organized Systems

    2.4. Multicyclic Supramolecular Systems

    3: Electrochemical Synthesis of Nanostructured Materials

    Abstract

    3.1. Introduction

    3.2. Fundamental Aspects of Electrochemistry

    3.3. Synthesis of Nanostructured Films by Electrodeposition

    3.4. Oxide Formation by Anodization of Valve Metals

    3.5. Conclusions

    List of Symbols

    4: Nanostructured Films: Langmuir–Blodgett (LB) and Layer-by-Layer (LbL) Techniques

    Abstract

    4.1. Introduction

    4.2. Langmuir–Blodgett Technique

    4.3. Layer-by-Layer Technique

    4.4. Final Considerations

    Acknowledgment

    5: Low-Dimensional Systems: Nanoparticles

    Abstract

    5.1. Introduction

    5.2. Synthesis Methods

    5.3. Properties

    5.4. Characterization Methods

    5.5. Applications

    5.6. Final Considerations

    List of Symbols

    6: Magnetic Nanomaterials

    Abstract

    6.1. Introduction

    6.2. Basic Concepts of Magnetism

    6.3. SPIOs

    6.4. The Structure and Physico-Chemical Properties of the SPIO Systems

    6.5. Biomedical Applications

    6.6. Conclusions and Perspectives

    List of Symbols

    7: Nanocomposites of Polymer Matrices and Lamellar Clays

    Abstract

    7.1. Polymer Nanocomposites

    7.2. Structures of Lamellar Clays

    7.3. Structure of Polymer Nanocomposites

    7.4. Methods for Obtaining Polymer Nanocomposites

    7.5. Compatibilization in Nanocomposites with Nonpolar Matrices

    7.6. Nanocomposites with Polar Matrices

    7.7. Nanocomposites with Nonpolar Matrices

    7.8. Final Considerations

    8: Mathematical Fundamentals of Nanotechnology

    Abstract

    8.1. Introduction

    8.2. Classical Mechanics

    8.3. Quantum Mechanics

    List of Symbols

    9: Carbon-Based Nanomaterials

    Abstract

    9.1. Introduction

    9.2. Fullerenes

    9.3. Carbon Nanotubes

    9.4. Graphene

    9.5. Conclusions and Perspectives

    Acknowledgments

    Index

    Copyright

    William Andrew is an imprint of Elsevier

    The Boulevard, Langford Lane, Kidlington, Oxford, OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    Copyright © 2017 Elsevier Inc. All rights reserved.

    This English edition of Nanostructures by Osvaldo Novais de Oliveira, Jr, Marystela Ferreira, Alessandra Luzia Da Róz, Fabio Leite is published by arrangement with Elsevier Editora Ltda.

    Originally published in the Portuguese language as Nanoestruturas 1st edition (ISBN 978-85-352-8089-0) © Copyright 2015 Elsevier Editora Ltda.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    ISBN: 978-0-323-49782-4

    For information on all William Andrew publications visit our website at https://www.elsevier.com/

    Publisher: Matthew Deans

    Acquisition Editor: Simon Holt

    Editorial Project Manager: Sabrina Webber

    Production Project Manager: Jason Mitchell

    Designer: Greg Harris

    Typeset by Thomson Digital

    List of Contributors

    L. Caseli,     Institute of Environmental, Chemical and Pharmaceutical Sciences, Federal University of São Paulo, São Paulo, São Paulo, Brazil

    A.L. Da Róz,     Federal Institute of Education, Science and Technology of São Paulo Itapetininga, São Paulo, Brazil

    A. de Barros,     State University of Campinas, Institute of Chemistry, Campinas, São Paulo, Brazil

    M.L. de Moraes,     Institute of Science and Technology, Federal University of São Paulo, São Paulo, São Paulo, Brazil

    R.F. de Oliveira,     Brazilian Center for Research in Energy and Materials, Brazilian Nanotechnology National Laboratory, Campinas, São Paulo, Brazil

    M. Ferreira,     Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    F. de Lima Leite,     Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    R. Marchiori,     Interdisciplinary Department of Science and Technology, Federal University of Rondônia, Ariquemes, Rondônia, Brazil

    C.M. Miyazaki,     Federal University of São Carlos, Center of Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    Osvaldo N. Oliveira, Jr.,     São Carlos Institute of Physics, University of São Paulo (USP), São Carlos, São Paulo, Brazil

    F.R. Passador,     Institute of Science and Technology, Federal University of São Paulo, São Paulo, São Paulo, Brazil

    L.G. Paterno,     Institute of Chemistry, University of Brasilia, University Campus Darcy Ribeiro, Brasilia, Brazil

    L.A. Pessan,     Department of Materials Engineering, Federal University of São Carlos, São Carlos, São Paulo, Brazil

    A. Riul, Jr.,     Gleb Wataghin Institute of Physics, State University of Campinas, Campinas, São Paulo, Brazil

    A. Ruvolo-Filho,     Department of Materials Engineering, Federal University of São Carlos, São Carlos, São Paulo, Brazil

    F.R. Simões,     Institute of Marine Sciences, Federal University of São Paulo, Santos, SP, Brazil

    J.R. Siqueira, Jr.,     Institute of Exact Sciences, Natural and Education, Federal University of Triângulo Mineiro (UFTM), Uberaba, Minas Gerais, Brazil

    J.S. Santos,     Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    M.A.G. Soler,     Institute of Physics, University of Brasília, University Campus Darcy Ribeiro, Brasilia, Brazil

    M. Souza Sikora,     Federal Technological University of Paraná, Pato Branco, Paraná, Brazil

    H.H. Takeda,     Department of Interdisciplinary Sciences and Technology, Federal University of Rondônia, Ariquemes, Rondônia, Brazil

    F. Trivinho-Strixino,     Federal University of São Carlos, Center for Sciences and Technology for Sustainability, Sorocaba, São Paulo, Brazil

    1

    Basic Concepts and Principles

    F.R. Simões*

    H.H. Takeda**

    *    Institute of Marine Sciences, Federal University of São Paulo, Santos, SP, Brazil

    **    Department of Interdisciplinary Sciences and Technology, Federal University of Rondônia, Ariquemes, Rondônia, Brazil

    Abstract

    Many scientific and technological advances can be credited to nanoscience and nanotechnology. Indeed, the scientific technological breakthroughs in these fields have not only affected specific sectors, such as equipment for research or manufacturing certain products, but also many products that are relevant to society as a whole. This chapter presents important descriptions of the key concepts and basic principles necessary for understanding nanoscience and nanotechnology, including the notion of scale; the history of nanoscience and nanotechnology development; the advent of techniques facilitating the fabrication of nanotechnology products; the relationships among politics, economics and research; and some commercial applications.

    Keywords

    nanoscience

    nanotechnology

    nanoscale

    nanomaterials

    science

    economics

    policy

    nanotechnology products

    technology

    innovation

    Chapter Outline

    1.1 Introduction

    1.1.1 Understanding the Nanoscale and Nanotechnology

    1.1.2 Nanoscience: History, Concepts, and Principles

    1.1.3 Investments, Strategies, Actions, and Research in Nanotechnology

    1.1.4 Commercial Products Involving Nanotechnology

    1.2 Final Considerations

    List of Symbols

    References

    1.1. Introduction

    1.1.1. Understanding the Nanoscale and Nanotechnology

    To understand nanoscience and nanotechnology, it is necessary to know the origin of the prefix nano, which is Greek and means dwarf. One nanometer (nm) is simply 1 billionth of 1 m (1 nm = 1 × 10−9 m). For comparison, the ratio between the size of a soccer ball and the Earth is approximately the same as that between a soccer ball and a sphere of 60 carbon atoms known as a C-60 fullerene (Fig. 1.1). The Earth is approximately 100 million times larger than a soccer ball, and in turn, the ball is approximately 100 million times larger than the fullerene [1].

    Figure 1.1   Illustration of the diameter ratios between the Earth and a ball, and between a ball and a C-60 fullerene.

    Several other common examples can be used to understand the nano scale. Fig. 1.2 compares different nanoscale materials. A human hair is approximately 100,000 nm wide, whereas a red blood cell is approximately 7,000 nm in diameter. Even smaller are typical viruses, which are between 45 and 200 nm in size. On the atomic scale, the length of a typical bond between carbon atoms and the spaces between atoms in a molecule are on the order of 0.12–0.15 nm [1–4].

    Figure 1.2   Illustrations of various materials ranging from millimeter (mm) to nanometer (nm) scales.

    Thus, structures of nanoscale materials (called nanostructures) are intermediate between the smallest structure that can be produced by man and the largest molecules of living systems. Humans’ abilities to control and manipulate nanostructures, therefore, facilitate exploring novel physical, biological, and chemical properties of systems that are of intermediate size between atoms and molecules, such as, nanoscale materials.

    Two standard definitions exist for the term nanotechnology: one given by the International Organization for Standardization–Technical Committee (ISO–TC) and the other given by the National Nanotechnology Initiative of the US (NNI). According to the ISO–TC, (i) Understanding and control of matter and processes at the nanoscale, typically, but not exclusively, below 100 nanometres in one or more dimensions where the onset of size-dependent phenomena usually enables novel applications; and (ii) Utilizing the properties of nanoscale materials that differ from the properties of individual atoms, molecules, and bulk matter, to create improved materials, devices, and systems that exploit these new properties. Therefore, for a device to be considered nanotechnological, in addition to nanometric dimensions, it must also have unique properties associated with the nanoscale [2–5]. In contrast, as defined by the NNI, nanotechnology must fall between 1 and 100 nm in size [2]. The lower limit is defined by the size of atoms, as this branch of science must construct devices from atoms and molecules. For example, one hydrogen atom has a diameter of approximately one quarter of a nanometer (d = 0.25 nm). The upper limit was established based on our ability to modulate properties on scales up to 100 nm and observe the resulting phenomena in larger structures that be used to generate specific devices [2]. These phenomena differentiate truly nanoscale devices from those that are simply miniaturized versions of an equivalent macroscopic device. Thus, such larger-scale devices should be considered as microtechnologies [3].

    Thus, nanotechnology is used to describe molecular-scale engineering systems. More specifically, this term refers to the ability to design, construct or manipulate devices, materials and functional systems on the nanometric scale [4,5].

    1.1.2. Nanoscience: History, Concepts, and Principles

    Between the discovery of fire by man and the present day, tremendous advancements have occurred in science and technology, accompanied by remarkable development in the research and manufacture of new products, including new materials, pharmaceuticals, and foods. For example, very light materials with mechanical resistances exceeding that of steel have been produced as a result of advances in nanoscience, which can be defined as the science that governs the study of nanotechnology for the development or improvement of materials based on the possibility of manipulating atoms and/or molecules with observed effects that are closely linked to the nanoscale and that have attractive physical, chemical and/or biological properties [6].

    Although a large number and variety of technologies and articles have recently been published on the topic of nanotechnology, this science has been applied and studied for a long time, albeit without knowledge of the relationship among the scale, the product and the resulting properties. In other words, man has manipulated materials at the nanometer level for a long time without understanding that the effects obtained were related to their nanoscale natures. One example is medieval glass-blowers who, using mixtures of gold nanoparticles of various sizes, produced differently colored stains for the fabrication of stained glass windows. A study by a research team at the University of Queensland [7] found that, in addition to the staining produced by gold nanoparticles, these nanoparticles also functioned as photocatalytic air purifiers: for example, when sunlight shone on the stained glass, air purification occurred. Another example is the experiment conducted by Michael Faraday (late 19th century), who synthesized gold nanoparticles [8] but did not understand their properties.

    Regarding the manipulation of particles at the atomic level, scientists have investigated the nanometric world to find explanations and rationales for their theories, such as the atomic theory first proposed by Democritus in 400 BC, which was refined in 1913 by Ernest Rutherford and Niels Bohr [9]. Additionally, in 1867, James Clerk Maxwell performed an experiment known as Maxwell’s demon, confirming that the second law of thermodynamics has only one statistical certainty. Briefly, in this experiment, Maxwell used a chamber containing a gas at equilibrium that was divided into two parts by a wall containing a door. When the door was opened, only the particles with higher and lower velocities could change sides, resulting in the heating of one side of the chamber and the cooling of the other [10]. Thomson (1906) and Lewis (1916) [9] developed the theory of chemical bonds (ionic and covalent bonding) to describe the formation of molecules. From 1934 to 1938, Lise Maitner, Otto Frisch, Otto Hahn, and Fritz Strassman studied the radioactive isotopes produced by bombarding uranium with neutrons (the experiment was conducted by Enrico Fermi). Based on their results, they discovered the phenomenon of nuclear fission, which releases 200 MeV of energy. This discovery led to the development of atomic bombs and nuclear power plants [11]. Another example is the nuclear fusion that occur in the Sun, n (Fig. 1.3), in which four hydrogen atoms fuse to form a helium atom, generating all the energy we observe and feel on Earth [12].

    Figure 1.3   The Sun, a star undergoing constant nuclear fusion [12].

    Many scientists have confirmed their theories or answered chemical and/or physical questions by simply observing the phenomena resulting from the atomic or molecular behaviors of materials, and thus, it can be concluded that a scientific field and its foundation develop via studies and the formulations of theories, concepts and principles based on experimental observations. Thus, defining the basic concepts and principles of nanoscience and nanotechnology would be much more difficult without the theoretical knowledge of other branches of sciences. Indeed, most of the laws, concepts and principles that govern nanoscience are essentially the same ones that govern physics and chemistry. To illustrate this, we simply review the portion of the definition of nanotechnology that states that nanotechnology is a novel branch of science responsible for the study and development of materials at the atomic and molecular levels that have unique characteristics associated with the nanoscale. This concept confirms that the manipulation of atoms or molecules first requires theoretical knowledge of atomic theory and chemical bonds, as indicated previously.

    Nanoscience is not only related to chemical and physical knowledge. There is also a great demand for biological knowledge. Biology and biochemistry also have much to gain from nanoscientific advances because DNA, viruses, and organelles are considered nanostructures [13]. For example, the National Aeronautics and Space Administration (NASA) has studied the development of nanoparticles containing DNA repair enzymes and ligands for the recognition of damaged cells [14].

    In 1959, the American physicist Richard Feynman (Fig. 1.4), in his lecture There’s plenty of room at the bottom [15], introduced the first nanoscientific approach. Feynman explained that the entire area on the head of a pin (1/16 in.), if amplified 25,000 times, would have an area capable of housing all the pages of the Encyclopedia Britannica. He explained that the resolving power of the human eye (approximately 1,120 points per inch) corresponds to approximately the diameter of one of the tiny dots in the high-quality half-tone reproductions in the Encyclopedia. If this dot could be demagnified by over 25,000 times, it would have a diameter of 80 Å, which is sufficient space for 32 atoms of a common metal. In other words, Feynman explained that one such dot (1,120 of which can be seen by the human eye in 1 in. is large enough to contain approximately 1,000 atoms, and therefore, the size of each dot can be easily adjusted, as required by photoengraving. As a result, the entire contents of the Encyclopedia could fit on the head of a pin.

    Figure 1.4   Richard Feynman [16].

    The term "Nanotechnology" was coined in 1974 by Norio Taniguchi of Tokyo University and used to describe the ability to create nanoscale materials. Until the term was formally defined, nanotechnology had not undergone major developments because it is a field that manipulates atoms and molecules, but no methods to observe and, thus, manipulate material in a controlled manner existed. However, since the invention of the first microscope, scientists have sought to amplify their ability to observe matter. Using a typical microscope with optical lenses, objects invisible to the naked eye and smaller than the wavelength of light can be observed. In contrast, with an electron microscope, it is possible to observe smaller particles with better definition, although individual atoms cannot be clearly distinguished. Then, in 1981, in the International Business Machines (IBM) laboratories in Zurich, Switzerland, Gerd Binning and Heinrich Rohrer developed a microscope known as the scanning tunneling microscope (STM), which won the Nobel Prize in Physics in 1986 and opened the door to nanotechnology and nanoscience. Fig. 1.5 shows the first commercially offered STM.

    Figure 1.5   First scanning tunneling microscope (STM) produced by W.A. Technology of Cambridge in 1986. Free license Image, Creative Commons License, provided by Science Museum London, Flickr. Available from: https://www.flickr.com; https://www.flickr.com/photos/sciencemuseum/9669013645 [17].

    Briefly, this microscope is equipped with a very fine probe that very closely scans the sample, removing electrons and generating an image of the atomic topography on the sample surface. The Binnig and Rohrer’s STM gave rise to an entire family of instruments and techniques that revolutionized our ability to visualize surfaces and materials that previously could not be observed. Atomic force microscopy (AFM) is one example of a technique derived from STM that allowed visualizing materials that do not conduct electricity. Indeed, these novel microscopes permitted not only visualization but also manipulation of matter on the nanoscale.

    One example of such manipulation is the experiment conducted by Donald M. Eigler and Erhard Schweinzer in 1989 at IBM. In their work, they manipulated 35 xenon atoms on a nickel substrate to spell out the company’s initials [18] (Fig. 1.6).

    Figure 1.6   An STM image of the International Business Machines (IBM) initials [18].

    Based on this experiment, many other researchers demonstrated the possibility of manipulating matter on the nanoscale. For example, researchers at the Brazilian Agricultural Research Corporation (EMBRAPA), in the Agricultural Instrumentation division, developed a method for the nanomanipulation of a compact disk’s (CD’s) polycarbonate surface involving mechanical modification via nanolithography using a phosphorus-doped silicon tip. They used this method to draw the EMBRAPA symbol and the Brazilian flag on the polycarbonate substrate in a controlled manner in an area of 10 μm × 10 μm [19] (Fig. 1.7).

    Figure 1.7   EMBRAPA and an image of the Brazilian national flag scratched onto the surface of a polycarbonate compact disk (CD) with an atomic force microscopy (AFM) tip [19].

    More recently, in 2009, researchers at Stanford University led by Hari Manoharan wrote the initials of Stanford University (SU) in letters smaller than atoms by encoding 35 bits of information per electron [20] (Fig. 1.8).

    Figure 1.8   Stanford University’s initials (SU) written using electron waves in a piece of copper and used to design a tiny hologram [20].

    Given these advances in atomic-scale microscopy, interest in nanoscience and nanotechnology has been steadily growing. According to Whitesides [13], there are six reasons to study nanoscience:

    1. Many properties remain mysterious, such as, the operation of the flagellar motor of E. coli bacteria and how electrons move through organometallic nanowires.

    2. Nanomaterials are relatively difficult to obtain. Unlike colloids, micelles and crystal nuclei, molecules are easily obtained and characterized. The development of chemical syntheses of colloids that are as accurate as those of molecules remains highly challenging.

    3. Many nanostructures are still inaccessible, and their study may lead to the observation of new phenomena.

    4. Nanostructures have many sizes in which quantum phenomena (especially quantum entanglement and other reflections of the material’s wave character) are expected to occur. Observing such quantum phenomena will contribute to explaining the behaviors and properties of atoms and molecules, but they are typically masked by the classical behaviors of matter and macroscopic structures. For example, quantum dots and nanowires have been produced and found to exhibit unique electronic properties.

    5. The nanometric and functional structures responsible for the primary functions of a cell represent the frontier of biology. For example, ribosomes, histones, chromatin, the Golgi apparatus, the interior structure of the mitochondria, the flagellar micromotor, the centers of photosynthetic reactions, and ATPases of cells are nanostructures that must be characterized and understood.

    6. Nanoscience is the basis for the development of nanoelectronics and photonics.

    The word nanotechnology is relatively new, but this field is not. It is estimated that nature has evolved on Earth for approximately 3.8 billion years, and nature includes many materials, objects and processes that function on the macroscale to the nanoscale [4]. Thus, understanding the behaviors and properties of these materials and processes may facilitate the production of nanomaterials and nanodevices. Biomimicry, a term derived from the Greek word biomimesis, was defined by Otto Schmitt in 1957 and denotes the development of biologically inspired designs that are derived or adapted from nature [4]. The term biomimicry is relatively new, but our ancestors have looked to nature for the inspiration and know-how to develop various devices for many centuries [21,22]. Indeed, throughout history, many objects and beings, including bacteria, plants, soil, aquatic animals, shells and spider webs, have been found to have commercially interesting properties.

    Bacterial flagella rotate at more than 10,000 rpm [23] and constitute an example of a molecular biological machine. The flagella motor is driven by proton flow caused by electrochemical potential differences across the cell membrane. The diameter of the bearings is approximately 20–30 nm, and the gap is approximately ≈ 1 nm [4].

    Many billions of years ago, molecules began to organize themselves into the complex structures that gave rise to life. Photosynthesis uses solar energy to support plant life. The molecular assemblies present in the leaves of plants (such as chlorophyll) take the energy from sunlight and transform it into chemical energy to power the biochemical processes of plant cells, which have processes ranging from the nanometric to the micrometric scale. This technology has been exploited and developed for solar energy applications [4].

    Some natural surfaces, including plant leaves with water repellents, are known to be superhydrophobic and self-cleaning because of their roughness (arising from nanostructures) and the presence of a wax coating [24]. Using roughness to imbue surfaces with superhydrophobicity and self-cleaning properties is of interest for many applications, including windows, windshields, exterior paints, ships, kitchenware, tiles, and textiles. Superhydrophobic surfaces can also be used for energy conversion and storage [25], whereas surfaces with low wettability can reduce the friction of contacting surfaces at machine interfaces [26].

    The fixation structures present on the feet of various creatures, including many insects (e.g., beetles and flies), spiders and lizards, can adhere to a variety of surfaces and be used for locomotion. These structures cling to and detach from different types of surfaces [27,28]. The dynamic adhesion capacity is called reversible adherence or smart adhesion. Common adhesives leave residues and are not reversible. Thus, replicating the characteristics of gecko feet would facilitate the development of a super adhesive polymer tape capable of clean, dry, and reversible adhesion [4]. Such a tape would have potential applications in everyday objects and high-tech applications, such as, microelectronics.

    Many aquatic animals can move at high speeds through water with low drag energy. For example, most shark species move through water with high efficiency. Shark skin is fundamental for this behavior, reducing friction, and exhibiting a self-cleaning effect that removes ectoparasites from its surface [4]. These characteristics are attributable to very small structures present in shark skin, called dermal denticles, which are ridges with longitudinal grooves that result in very effective mobility through water and minimize the adherence of barnacles and algae [4].

    Speedo developed a full-body, shark skin-based swimsuit, called Fastskin, for elite swimming. Furthermore, the builders of boats, ships, and aircraft have also attempted to mimic shark skin to reduce drag and minimize the fixation of organisms on the surfaces of these craft. The mucus on the skin of aquatic animals, including sharks, acts as a barrier against the osmotic salinity of sea water, protects against parasites and infections, and functions as a friction-reducing agent. Artificial fish-derived mucus products are currently used to propel crude oil through the Alaskan pipeline [4].

    Shells are natural nanocomposites with laminated structures and superior mechanical properties. Spider webs are made of silk fiber with high tensile strength. The materials and structures used in these objects have led to the development of various materials and fibers with high mechanical resistance [4]. Moths have eyes with multifaceted surfaces on the nanoscale and are structured to reduce the reflection of light. Their antireflective structure inspired the development of antireflective surfaces [29].

    Biological systems’ self-healing abilities are highly interesting. For example, the chemical signals originating from the site of a fracture initiate a systemic response that sends agents to repair the injury. Inspired by these activities, various artificial self-structuring materials have been developed [30]. Human skin, for example, is sensitive to impact, which leads to purple discoloration of affected areas. This behavior led to the development of coatings that indicate impact-related damage [21].

    Sensor arrays mimicking human senses, such as smell [31] and taste [32–35], consist of a set of sensors based on nanostructured materials and have been widely used in various applications, such as gas and liquid sensing, respectively [31,36–40].

    Nanostructured materials are typically named according to their shapes and sizes and may take the forms of particles, tubes, wires, films, flakes and reservoirs, provided they have at least one nanoscale dimension [41,42]. One material that has been widely studied in nanotechnology is carbon nanotubes (Fig. 1.9), which are so named because their diameters are between 1 and 100 nm, although their lengths are typically on the order of hundreds of nanometers.

    Figure 1.9   Image of carbon nanotubes immobilized onto a film of poly(allylamine) hydrochloride obtained with a high-efficiency electron microscope [a scanning electron microscope coupled with a field emission gun (SEM-FEG)] in the Materials Engineering Department of the Federal University of São Carlos (DEMA-UFSCar) [43].

    There are two main modes of developing nanotechnological materials. In the bottom-up approach, materials and devices are built from molecular components that are chemically organized according to the principles of molecular recognition. In contrast, in the top-down approach, nanoscale objects are built from other, larger scale objects, without control at the atomic level [44]. Many methods using these two modes have been reported.

    In the bottom-up approach, a DNA molecule may, for example, be used to build other larger and well-defined structures using DNA and other nucleic acids [4]. Another example is the self-assembly technique, which can create self-organized molecular layer films [45,46]. Additionally, as previously discussed, an AFM tip can be used as a nanoscale recording head [19].

    Due to top-down approach, many solid-state technologies used to create silicon-based microprocessors are now able to use resources on a scale smaller than 100 nm. Other techniques can also be applied to create devices known as Nano Electro Mechanical Systems (NEMS) derived from Micro Electro Mechanical Systems (MEMS) [4,5].

    Examples of NEMS include microcantilevers with integrated nanotips for STM and AFM, AFM tips for nanolithography, molecular gears used to attach benzene molecules to the outer walls of carbon nanotubes, magnetic media used in hard disk drives, magnetic tape units [4], and ion beams that can directly remove or deposit materials in the presence of precursor gases. AFM can also be used in the top-down approach for the deposition of resistive films on a substrate that is subsequently subjected to an etching process

    Enjoying the preview?
    Page 1 of 1