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

Only $11.99/month after trial. Cancel anytime.

Functional Materials: Preparation, Processing and Applications
Functional Materials: Preparation, Processing and Applications
Functional Materials: Preparation, Processing and Applications
Ebook1,309 pages15 hours

Functional Materials: Preparation, Processing and Applications

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Functional materials have assumed a very prominent position in several high-tech areas. Such materials are not being classified on the basis of their origin, nature of bonding or processing techniques but are classified on the basis of the functions they can perform. This is a significant departure from the earlier schemes in which materials were described as metals, alloys, ceramics, polymers, glass materials etc. Several new processing techniques have also evolved in the recent past. Because of the diversity of materials and their functions it has become extremely difficult to obtain information from single source. Functional Materials: Preparation, Processing and Applications provides a comprehensive review of the latest developments.

  • Serves as a ready reference for Chemistry, Physics and Materials Science researchers by covering a wide range of functional materials in one book
  • Aids in the design of new materials by emphasizing structure or microstructure – property correlation
  • Covers the processing of functional materials in detail, which helps in conceptualizing the applications of them
LanguageEnglish
Release dateDec 9, 2011
ISBN9780123851437
Functional Materials: Preparation, Processing and Applications

Related to Functional Materials

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Functional Materials

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Functional Materials - S. Banerjee

    Table of Contents

    Cover image

    Front-matter

    Copyright

    Preface

    About the Editors

    Contributors

    1. Soft Materials — Properties and Applications

    1.1. Introduction to Soft Matter

    1.2. Intermolecular Interactions in Soft Materials

    1.3. Colloids

    1.4. Surfactant Assemblies

    1.5. Polymer Solutions

    1.6. Experimental Techniques in Soft Matter

    1.7. Applications of Soft Matter

    2. Conducting Polymer Sensors, Actuators and Field-Effect Transistors

    2.1. Introduction

    2.2. Synthesis of Conducting Polymers

    2.3. Conducting Polymer Gas Sensors

    2.4. Electrochemical Actuators

    2.5. Conducting Polymer FETs

    2.6. Summary

    3. Functional Magnetic Materials

    3.1. Introduction

    3.2. Magnetocaloric Effect

    3.3. Molecular Magnetic Materials

    3.4. Magnetic Nanoparticles

    3.5. CMR Manganites

    3.6. Summary and Conclusion

    4. Multiferroic Materials

    4.1. Introduction

    4.2. Origin of Ferro- and Antiferromagnetism

    4.3. Origin of Ferroelectricity

    4.4. Mutually Exclusive Reason for Multiferroicity

    4.5. Types of Multiferroic Material

    4.6. Observation of Multiferroic Properties

    4.7. Examples

    4.8. Applications

    4.9. Summary

    5. Spintronic Materials, Synthesis, Processing and Applications

    5.1. Introduction

    5.2. Ferromagnetic Semiconductors or Dilute Magnetic Semiconductors

    5.3. Spintronics

    5.4. Overview of some Major Spintronic Materials

    5.5. Oxide Semiconductors

    5.6. Material Synthesis, Processing and Characterization

    5.7. Characterization

    5.8. Recent Results

    5.9. One-Dimensional Structures of ZnO-Based Materials

    5.10. Applications (Spintronic Devices)

    6. Functionalized Magnetic Nanoparticles

    Nomenclature

    6.1. Introduction

    6.2. Methods of Preparation of Nanoparticles

    6.3. Characterization of Magnetic Nanoparticles

    6.4. Magnetic Properties of Nanoparticles

    6.5. Induction Heating Behaviour of Particles

    6.6. Therapeutic Efficacy of Magnetic Nanoparticles in Human Cancer Cells

    6.7. Future Perspectives

    7. Functional Superconducting Materials

    7.1. Background

    7.2. Niobium Titanium (NbTi)

    7.3. A15 Superconductors and Nb3Sn

    7.4. Chevrel-Phase Superconductors

    7.5. High-Tc Superconductors

    7.6. MgB2

    7.7. Borocarbides

    7.8. Iron Arsenide Superconductors

    7.9. Conclusions

    8. Optical Materials

    8.1. Introduction

    8.2. Origin of Different Types of Optical Material and their Applications

    8.3. Optical Parameters

    8.4. Optical Properties of Metals

    8.5. Optical Properties of Insulators

    8.6. Optical Properties of Nanomaterials

    8.7. Nonlinear Optical Materials

    8.8. Organic Optical Materials

    8.9. Photonic Band-Gap Materials

    9. Glass and Glass-Ceramics

    9.1. Introduction

    9.2. Glasses

    9.3. Glass-Ceramics

    9.4. Preparation of Glass and Glass-Ceramics

    9.5. Characterization

    9.6. Mechanical Properties

    9.7. Wetting Property

    9.8. Some Useful Properties

    9.9. Some Important Functionalities

    9.10. Transparency

    9.11. Conclusion

    10. Nuclear Fuels

    10.1. Introduction

    10.2. Types of Fuel Material

    10.3. Phase Diagrams

    10.4. Fission Gas Release

    10.5. Vapourisation of the Fuel

    10.6. Swelling Due to Gas Bubbles

    10.7. Swelling Due to Solid Fission Products

    10.8. Pore Migration

    10.9. Restructuring

    10.10. Mechanical Phenomenon

    10.11. Temperature Distribution

    10.12. Fuel Modelling

    10.13. Conclusions

    11. Super-Strong, Super-Modulus Materials

    11.1. Introduction

    11.2. Origin of Modulus

    11.3. Strength of Materials

    11.4. Ultra-strength

    11.5. Summary

    12. Corrosion-Resistant Materials

    12.1. Introduction

    12.2. Materials Resistant to Uniform Corrosion

    12.3. Materials Resistant to Localized Corrosion

    13. Nafion Perfluorosulphonate Membrane

    13.1. Introduction

    13.2. Synthesis and Characterization

    13.3. Properties of Nafion Membranes

    13.4. Applications

    13.5. Conclusions

    14. Fundamentals and Applications of the Photocatalytic Water Splitting Reaction

    14.1. Introduction

    14.2. Basis of Photocatalytic Water Splitting

    14.3. Experimental Method for Water Splitting

    14.4. Some Heterogeneous Photocatalyst Materials Used for Water Splitting

    14.5. Conclusions

    15. Hydrogen Storage Materials

    15.1. Introduction

    15.2. Experimental Techniques

    15.3. Examples of Hydrogen Storage Materials and Their Properties

    15.4. Applications

    15.5. Conclusions

    16. Electroceramics for Fuel Cells, Batteries and Sensors

    16.1. Introduction

    16.2. Preparation and Processing of Electroceramics

    16.3. Electrochemical and Microstructural Characterization

    16.4. Applications

    17. Nanocrystalline and Disordered Carbon Materials

    17.1. Introduction

    17.2. Fullerene

    17.3. CNTs

    17.4. Graphene: The Slimmest Carbon

    17.5. Nano-Diamond

    17.6. Carbon Nanofoam

    17.7. Amorphous Carbon

    Front-matter

    Functional Materials

    Functional Materials

    Preparation, Processing and Applications

    S. Banerjee

    Department of Atomic Energy,

    Mumbai,

    India

    A. K. Tyagi

    Chemistry Division,

    Bhabha Atomic Research Centre,

    Mumbai,

    India

    AMSTERDAM • BOSTON • HEIDELBERG • LONDON • NEW YORK • OXFORD • PARIS • SAN DIEGO • SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO

    Copyright

    Elsevier

    32 Jamestown Road London NW1 7BY

    225 Wyman Street, Waltham, MA 02451, USA

    First edition 2012

    Copyright © 2012 Elsevier Inc. All rights reserved

    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 arrangement 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.

    British Library Cataloguing-in-Publication Data

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

    Library of Congress Cataloging-in-Publication Data

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

    ISBN: 978-0-12-385142-0

    For information on all Elsevier publications visit our website at elsevierdirect.com

    This book has been manufactured using Print On Demand technology. Each copy is produced to order and is limited to black ink. The online version of this book will show color figures where appropriate.

    Preface

    S. Banerjee and A.K. Tyagi

    Solid state materials have long been conventionally classified into metals, semiconductors and insulators based on their transport properties. They can be either crystalline or amorphous depending on the range of crystallographic ordering in them. Again, based on their surface finishing and dielectric properties, they are termed as transparent, opaque or reflecting materials. Processing techniques, on the other hand, entail them to be subdivided into ingots, powders, coatings, slurries and aerosols. Above all, they are classified into organics and inorganics based on their origin. However, such traditional classification of materials failed badly when a member of an apparently insulating oxide family exhibited the very high electrical conductivity after the discovery of a completely new phenomenon called superconductivity. Subsequently, conducting polymers were discovered and single-layer graphene came into existence against all theoretical predictions. People are now looking for materials that simultaneously exhibit two apparently antagonistic properties. For example, multi-ferroics are a class of materials that are supposed to be ferromagnetic and ferroelectric at the same time. Further, restriction in dimensionality has opened up an all-new world of nanotechnology. New molecules like fullerenes and carbon nanotubes, with exotic morphology and excellent properties, have been synthesized. This called for a more practical classification of materials. In view of this, a new scheme has been proposed wherein materials (more appropriately now termed functional materials) comprising representatives of the different existing classes are categorized solely on their functional properties and their prospective applications. Functional materials have assumed very prominent positions in several high-tech areas. Such materials are not classified on the basis of their origin, nature of bonding or processing techniques but on the basis of functions that they can perform. This is a significant departure from the earlier schemes in which materials were described as metals, alloys, ceramics, polymers, glass and so on. For example, a material will be termed as a gas sensor irrespective of whether it is of organic or inorganic origin, and a sensing mechanism can be either resistive or luminescent. Immediately, the new nomenclature became very popular with the scientific community, as is evident from the many symposia and conferences being held, and articles and reports being published on functional materials. However, unfortunately, until now, there has been no appropriate book encompassing the wide variety of functional materials.

    In view of such lean resources, and to provide an easy single-source access to the data being generated on this upcoming topic, we compiled this book, entitled ‘Functional Materials: Preparation, Processing and Applications’. It encompasses a wide panorama of functional materials starting from the superstrong materials at one end of the spectrum to the soft materials at the other. The book is subdivided into 17 chapters, each of which is dedicated to a particular class of functional materials. Separate chapters are provided for materials with magnetic, optical and catalytic properties. The energy crisis is a burning global issue today, and special emphasis is given to those materials with immense prospects for nuclear and other non-conventional energy applications. There are also chapters that deal with functional materials for societal applications like bio-medical applications and advanced electronics. The chapters are written by researchers who have extensive research experience on their respective topics. The chapters have been appropriately structured to provide an essence of the different techniques leading to the synthesis of these functional materials and their subsequent processing for targeted applications. Each chapter broadly discusses physical basis of the functionality, materials synthesis and processing, characterization, properties and applications. New concepts have been introduced and the most recent literature cited. We are immensely thankful to all the authors for their rich contributions, without which this book would not have been possible. Due efforts have been taken to make the book as error-free as possible. Despite that, some errors might have crept in. We shall be thankful to vigilant readers for bringing such unintentional errors to our notice. Finally, we sincerely hope that our efforts will be of use both to new and experienced researchers in the field.

    About the Editors

    Dr. S. Banerjee, currently Chairman, Atomic Energy Commission, India, obtained his BTech (metallurgical engineering) from the Indian Institute of Technology (IIT), Kharagpur, in 1967. He joined the erstwhile Metallurgy Division, Bhabha Atomic Research Centre (BARC), Mumbai, in 1968, after graduating from the 11 batch of BARC Training School. He obtained his PhD in 1974 from IIT, Kharagpur. He occupied several important positions such as Head, Metallurgy Division, BARC; Director, Materials Group, BARC; and Director, BARC; he subsequently became Chairman, Atomic Energy Commission in November 2009. He has done pioneering work in the field of martensitic transformations, rapid solidification, omega transformation, quasi-crystalline solids and shape memory alloys.

    He was a senior visiting Fellow at the University of Sussex, UK; Humboldt Foundation Fellow at Max-Planck Institute for Metallforschung, Stuttgart, Germany; and a visiting Professor at Ohio State University, Columbus, USA.

    In recognition of his seminal contribution to the field of materials science, he has been conferred with numerous national and international prestigious awards such as Indian National Science Academy (INSA) Young Scientist Award (1976); National Metallurgists’ Day Award (1981); Acta Metallurgica Outstanding Paper Award (1984); Shanti Swaroop Bhatnagar Prize in Engineering Sciences from the Council of Scientific and Industrial Research (1989); GD Birla Gold Medal of The Indian Institute of Metals (1997); INSA Prize for Materials Science; Materials Research Society of India (MRSI)-Superconductivity and Material Science Prize (2003); Indian Nuclear Society Award (2003); Alexander von Humboldt Research Award (2004); Professor Brahm Prakash Memorial Medal (2004) from INSA; Padma Shri from Government of India (2005); Distinguished Materials Scientist Award from MRSI (2008); Indian Science Congress Association’s Excellence in Science and Technology Award (2009); Ram Mohun Puraskar of Ram Mohun Mission for an Outstanding Contribution to Nuclear Science (2010); and CNR Rao Prize Lecture in Advanced Materials (2011). He has been conferred with several Doctor of Science (Honoris Causa) degrees from various universities and institutions such as Sathyabama University, Chennai; Bengal Engineering and Science University, Shibpur; Indian Institute of Technology, Kharagpur; Guru Ghasidas University, Chattisgarh; and the University of Calcutta. He is an elected Fellow of Indian Academy of Sciences; National Academy of Sciences, India; Indian National Science Academy; Indian National Academy of Engineering; and Third World Academy of Sciences.

    Dr. A.K. Tyagi obtained his MSc (Chemistry) degree in 1985 from Meerut University and joined the 29th batch of Bhabha Atomic Research Centre (BARC) Training School in the same year. After completing the one year orientation course, he joined the Chemistry Division of BARC, Mumbai, in 1986. Since then he has been working on various aspects of solid state chemistry, for example oxide superconductors, rare-earth based mixed fluorides, nuclear materials, framework solids, nanomaterials, electro-ceramics and multi-functional materials.

    He was awarded his PhD by Mumbai University in 1991. He was at Max-Planck Institute for Solid State Research, Stuttgart, Germany, for postdoctoral research during 1995–1996 on a Max-Planck Fellowship. In addition, he has also visited the Institute of Superior Technology, Portugal; the National Research Council, Ottawa, Canada; Dalhousie University, Halifax, Canada; Moscow State University, Moscow, Russia; Institute for Materials, Nantes, France; University of Malay, Malaysia; National Institute of Materials Science, Tsukuba, Japan; National University of Singapore, Singapore; Royal Institute of Technology, Stockholm, Sweden; Rice University, Houstan, USA; Tongi University, Shanghai, China and University of Valencia, Spain.

    In recognition of his significant contributions to the field of solid state chemistry, he has been conferred with several awards, such as the Dr. Laxmi Award by the Indian Association of Solid State Chemists and Allied Scientists (2001); Rheometric Scientific-Indian Thermal Analysis Society Award (2002); Gold Medal of the Indian Nuclear Society (2004); Materials Research Society of India Medal (2005); Chemical Research Society of India Medal (2006); Homi Bhabha Science and Technology Award (2006); IANCAS-Dr. Tarun Datta Memorial Award (2007); Rajib Goyal Prize in Chemical Sciences (2007); RD Desai Memorial Award from the Indian Chemical Society (2009); and Department of Atomic Energy's Science Research Council (DAE-SRC) Outstanding Researcher Award (2010). He is a Fellow of the National Academy of Sciences (FNASc), India; Maharashtra Academy of Sciences (FMASc); and Royal Society of Chemistry (FRSC), UK. Recently, he has been selected for the CNR Rao National Prize for Chemical Sciences. At present, he is heading the Solid State Chemistry Section of the Chemistry Division, BARC, Mumbai. He is also a Professor (Chemistry) at Homi Bhabha National Institute (HBNI), Mumbai.

    Contributors

    S.N. Achary

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Arvind Ananthanarayanan

    Glass and Advanced Ceramics Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    D.K. Aswal

    Technical Physics Division, Bhabha Atomic Research Center, Trombay, Mumbai, Maharashtra, India

    A.M. Banerjee

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    S. Banerjee

    Depatment of Atomic Energy, Government of India, Anushakti Bhavan, Mumbai, Maharashtra, India

    Shyamala R. Bharadwaj

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    J.K. Chakravartty

    Mechanical Metallurgy Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    G.K. Dey

    Materials Science Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    R. Ganguly

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    T.R. Govindan Kutty

    Radiometallurgy Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    P.A. Hassan

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    O.D. Jayakumar

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Vivekanand Kain

    Corrosion Science Section, Materials Science Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    R. Kapoor

    Mechanical Metallurgy Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    G.P. Kothiyal

    Glass and Advanced Ceramics Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Amit Kumar

    Radiation Biology & Health Sciences Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    R.S. Ningthoujam

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Mrinal R. Pai

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    B.N. Pandey

    Radiation Biology & Health Sciences Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Jayshree Ramkumar

    Analytical Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Gurazada Ravikumar

    Technical Physics Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Mainak Roy

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Vibha Saxena

    Technical Physics Division, Bhabha Atomic Research Center, Trombay, Mumbai, Maharashtra, India

    K. Shashikala

    Formerly at Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    J.B. Singh

    Mechanical Metallurgy Division, Bhabha Atomic Research Centre, Mumbai, Maharashtra, India

    V. Sudarsan

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    A.K. Tripathi

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    A.K. Tyagi

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    R.K. Vatsa

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Gunjan Verma

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    J.V. Yakhmi

    Formerly at Technical Physics Division, Bhabha Atomic Research Center, Trombay, Mumbai, Maharashtra, India

    S.M. Yusuf

    Solid State Physics Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    B.N. Wani

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    1. Soft Materials — Properties and Applications

    P.A. Hassan, Gunjan Verma and R. Ganguly

    Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai, Maharashtra, India

    Soft materials comprise a very large class whose common feature is that they are composed of mesoscopic entities, i.e. typical sizes of 1–1000nm, held together by interaction energies comparable to thermal energy. Colloids, polymers, liquid crystals, surfactants, etc. are typical examples that form soft materials under appropriate conditions. As the name implies, they can be easily deformed by external forces and have shear modulus three to six orders of magnitude smaller than atomic solids. The existence of a finite, yet small value of shear modulus puts them at a state intermediate between that of true solids and liquids and thus they can be visualized as ‘soft’ solids. In recent years, this field has emerged as an identifiable area in materials science that possesses substantial scientific challenges and industrial applications. This chapter will discuss the basic principles governing the generation of soft materials, their characterization and typical applications with special reference to colloids, polymers and surfactants.

    Keywords

    Soft matter, micelle, polymers, light scattering, small-angle neutron scattering, rheology, drug delivery

    1.1. Introduction to Soft Matter

    1.1.1. Introduction

    The macroscopic properties of materials are largely dependent on the strength and length scale of intermolecular interactions between the moieties that constitute the matter. Gas molecules condense into a liquid when the attractive force that operates between the molecules overcomes the effect of thermal motion. If the liquid is cooled below its freezing point, it forms a solid where each molecule is locked into a definite location in the solid. In a liquid, the positions of the molecules are not fixed, though they are held together by intermolecular forces. In the liquid state, the attractive energy of interaction between the molecules is comparable to that of the kinetic energy of motion. This interplay of attractive forces and the thermal motion leads to widely differing mechanical properties to the condensed phase materials. In a classical definition, liquids take the shape of a container in which they reside, whereas solids maintain their shape indefinitely. In other words, liquids flow and solids are elastic in nature when subjected to external shear forces. These behaviours are specified in terms of the shear viscosities for liquids and shear modulus for solids. ‘Soft condensed matter’ or simply ‘soft matter’ is a subfield of condensed matter which deals with materials that can be easily deformed by external forces. The magnitude of the shear modulus of soft matter is much smaller than that observed for true atomic crystals. Moreover, its mechanical behaviour is intermediate between a solid and a liquid, in the sense that these materials maintain their shape for a time, but they eventually flow. They appear as elastic solids for short times and viscous liquids over long times; hence they are also called viscoelastic materials. The characteristic time required for them to change from solid to liquid varies over a wide range, depending on the material.

    Many materials that we encounter in our everyday life belong to this class. For example, many food products such as mayonnaise, ice cream, chocolate, jelly, ketchup and cornflour paste are soft materials. Mayonnaise is an emulsion comprising vegetable oil, water and egg yolk. The natural surfactant lecithin present in the egg yolk acts as the emulsifier. Similarly, the mechanical properties of many personal-care products such as shampoos, toothpaste, lotions and skin creams are tailored to suit their application. The flow behaviour of shampoo is controlled by the addition of high-molecular weight polymers and careful control of surfactant concentration. Toothpaste is a complex mixture of colloidal particles, surfactants and polymers in which the required flow behaviour and softness is achieved by judicious combination of the ingredients. High-molecular weight polymers such as polyacrylic acid (PAAc), commercially known as carbomers, are used very often to control the flow behaviour of aqueous formulations. Gelatin, a biopolymer derived from collagen, in aqueous media forms a gel with a high volume fraction of water, which is again a classic example of soft material. Amphiphilic molecules or surfactants can self-assemble in selective solvents to form a variety of structures such as micelles, vesicles and lyotropic liquid crystals. Such self-assembled structures can act as the structural unit to form soft matter under appropriate conditions. One common feature in all the materials discussed above is that they are composed of units or structures at the mesoscopic length scale. Colloidal particles, polymers and surfactant assemblies have dimensions in the range of a few nanometres to a few micrometres. This length scale of the constituent units makes the materials indeed soft, i.e. their rigidity against mechanical deformation is many orders of magnitude smaller than their atomic counterparts. Thus, colloids, polymers and surfactants can be visualized as some of the building blocks for creating soft materials.

    1.1.2. Soft Matter: A Viscoelastic Fluid

    The principal quantity that tells how ‘hard’ or ‘soft’ a material is its shear modulus. Rheological measurements provide information about this property, which depends on the timescale at which it is probed. Rheology is altogether a different branch of science which deals with the deformation and flow of matter. It deals with the response of materials when they are subjected to external forces. In general, the two limiting cases of rheological properties of a liquid and a solid can be generalized in terms of flow and elasticity, respectively. Flow is an irreversible deformation which is the characteristic of a liquid, whereas elasticity is a reversible deformation characteristic of a solid. These are two ideal cases and their behaviour can be best illustrated by a simple shear experiment as shown in Figure 1.1.

    Consider a material contained between two parallel plates each of which having an area, A (Figure 1.1). If a force F is applied to the upper plate such that it moves with a velocity, v, relative to the lower plate and if the displacement of a given element located at dx is dy, then the shear strain γ are defined as

    (1.1)

    The force F acting at the upper plate produces a shear stress, σ=F/A.

    , is proportional to the shear stress σ as shown by the Newton’s equation,

    (1.2)

    The proportionality constant η is called the viscosity of the liquid. The above relationship is not true in the case of an elastic solid in which the shear stress, σ, is proportional to the strain, γ, as given by the Hooke’s law,

    (1.3)

    The proportionality constant, G, is called the rigidity modulus or shear modulus. In practice, Newton’s law and the Hooke’s law are two limiting cases of the response of a material under shear. The behaviour of many materials, especially soft matter, cannot be described by either Newton’s law or Hooke’s law but they exhibit both viscous and elastic responses. Such materials are designated as viscoelastic materials. The term ‘linear viscoelasticity’ is applied to the study of viscoelastic properties under small deformations where the strain varies linearly with the stress. This implies that the stress, strain and their time derivatives are related by a linear differential equation with constant coefficients. The stress–strain relationship for a general viscoelastic fluid can be described by a linear differential equation

    (1.4)

    where the coefficients αi and βi are material parameters.

    Another point to note here is that the response of a viscoelastic material is time dependent. For a viscoelastic material, the internal stresses are a function of both the instantaneous deformation and the past history of deformation. The difference in the mechanical behaviour of viscous, elastic and viscoelastic materials can be best understood from time- or frequency-dependent measurements. If we apply stress to a soft matter on a slow timescale, it flows like a viscous liquid; if one observes at a fast timescale, it behaves like a solid. For each material, there is a characteristic time known as the relaxation time (t) of the material such that at observation times less than t, it behaves like a solid, whereas at times longer than t, the material flows like a liquid. The ratio of the characteristic relaxation time to the observation time is called the Deborah number (De). When De≫1, the behaviour of material is ‘solid like’; when De≪1, it behaves like a liquid.

    One simple mathematical model for a viscoelastic material is the Maxwell fluid, in which one spring (elastic) and a dashpot (viscous) are connected in series as shown in Figure 1.2.

    The stress–strain relationship for such a Maxwell fluid can be written as

    (1.5)

    Rearranging the above equation, we get

    (1.6)

    where τ=η/G is the relaxation time. Thus, for a Maxwell fluid the characteristic relaxation time is the ratio of viscosity to shear modulus.

    The deviation of a fluid from Newtonian behaviour can be observed in different ways. Based on the variation of shear stress versus shear rate, non-Newtonian fluids are placed in different classes. Some commonly observed non-Newtonian behaviours are Bingham fluid (viscoplastic), shear thinning (pseudoplastic) and shear thickening (dilatant). The variation of shear stress and apparent viscosity with shear rate for the above-mentioned classes is compared with the Newtonian liquid in Figure 1.3.

    1.1.3. Shear Modulus and the Energy Density

    The elastic properties of an ideal solid can be understood if one considers the intermolecular interactions in the solid. Let us consider a perfect cubic crystal having a lattice constant a, as shown in Figure 1.4A. This crystal is sheared by an external force, F, in the x-direction such that every lattice plane is displaced parallel to itself by an amount dx (Figure 1.4B).

    Assuming that the interaction between the atoms can be represented by elastic springs with a spring constant, k, and each atom occupies an area a², the force per unit area of the plane (shear stress) can be approximated as (F/A)∝(k·dx/a²) and the shear strain is given by dx/a.

    Substituting the above values of shear stress and shear strain in Eq. (1.3), the shear modulus of the crystal can be obtained as G≈(k/a).

    Without going into the details of the interaction potential, if one assumes that ε is the bond energy and this much energy is spent in moving the atom over a distance a, the spring constant k can be approximated as ε/a².

    Substituting this value of k in the shear modulus, we get G~ε/a³. The above calculation shows that the shear modulus of a solid is related to the energy of the bond between the neighbouring atoms and is inversely proportional to the volume occupied by one such bond. This is nothing but the total bond energy per unit volume of the sample. Thus, the modulus can be thought of as an energy density of the material. A material with a high density of strong bonds is ‘stiff’, whereas that with a low density of weak bonds is ‘soft’.

    The bond energy of atomic systems ranges from 0.1eV for noble gases to about 10eV for metals [1]. The interaction energy in colloidal systems typically ranges from 1 kBT to about 100kBT, where kB is Boltzmann’s constant, and T is the absolute temperature. Because kBT~0.025eV, the energy scales for atomic and colloidal systems are about the same order of magnitude. Thus, the major difference in the shear modulus of atomic solids and soft materials arises from the interparticle distance. As the typical length scales in soft materials vary from a few nanometres to micrometres, the shear modulus can decrease by 3–12 orders of magnitude. For example, if we calculate the shear modulus of a solid where the bond energy is ~100kJ/mole and the interparticle distance is 1Å, the shear modulus, G~(10⁵J/mol)/(6.023×10²³×10−30m³)≈10¹²N/m².

    If we consider a material with similar magnitude of interaction energy, but with an interparticle distance of 100nm, the modulus is of the order of 10³N/m² only.

    1.2. Intermolecular Interactions in Soft Materials

    The intermolecular forces that operate between atoms or molecules exist between the constituents of soft matter as well, though their magnitude and length scale can be affected by the size of the particles and the nature of the medium in which they are dispersed. There are different classes of intermolecular forces between atoms or molecules. First, there is the electrostatic interaction arising from the Coulomb forces between charged bodies. The interactions between charges, permanent dipoles, quadrupoles, etc. come in this category. The polarization forces that arise from the induced dipole moments of molecules created by the electric fields of charges or dipoles represent another class. Finally, there are forces that are quantum mechanical in nature such as covalent bonds, hydrogen bonds and charge transfer interactions. In addition to the above, other modes of interaction such as depletion, hydration, steric and hydrophobic forces exist in soft materials. A detailed description of the intermolecular forces between atoms, molecules and large particles can be found elsewhere [2].

    1.2.1. Charge–Charge Interaction

    Coulomb interaction between two charged bodies is by far the strongest of the physical forces. The interaction energy for the electrostatic interaction between two charges Q1 and Q2 is given by

    (1.7)

    where εr is the relative permittivity or dielectric constant of the medium, and r is the distance between the two charges. To get an estimate of the magnitude of the Coulomb interaction energy in terms of the thermal energy, let us take the example of Na+ and Cl− ions in contact. The sum of the two ionic radii is 2.76Å. Using this value as the distance between the two ions, we get the interaction energy as −8.4×10−19J. At 300K, kBT is 4.1×10−21J. Thus the Coulomb energy turns out to be of the order of 200 kBT per ion pair in vacuum. When the distance between the two ions in vacuum is around 56nm, this energy will be in the range of kBT.

    1.2.2. Ion–Dipole Interactions

    Ion–dipole interaction exists between a charged ion and a polar molecule. The magnitude of this interaction depends on the charge Q on the ion, the dipole moment μ and the orientation angle θ between the dipole and the line joining the two molecules. When the distance between the ion and dipole is longer than the length of the dipole, the interaction energy can be written as

    (1.8)

    Simple calculations show that at typical interatomic separations, i.e. 2–4Å, the ion–dipole interaction is much stronger than kBT and strong enough to bind ions to polar molecules. The orientation angle dependence of the interaction energy suggests mutual alignment of ion and dipole. When ion–dipole interaction takes place in water, the above energies will be reduced by a factor of 80, the dielectric constant of water. Nevertheless, the strength of this interaction will exceed kBT for small divalent and multivalent ions and is by no means negligible for monovalent ions.

    1.2.3. Dipole–Dipole Interactions

    This interaction exists between two polar molecules when they approach each other. This depends on the magnitude of the dipole moments μ1 and μ2 and orientation of the dipoles with respect to the line joining the two dipoles. The maximum attraction occurs when the two dipoles are lying in line. The interaction energy for such an orientation is given by

    (1.9)

    This interaction is not as strong as the previous two electrostatic interactions and for dipoles of the order of 1Debye, it is already weaker than kBT at distances of about 3.5Å. At large separations or in a medium or high dielectric constant, the angle dependence of the interaction energy falls below the thermal energy and the dipoles can now rotate more or less freely. However, even though all orientations are possible, the angle-averaged potentials are not zero because there is always a Boltzmann weighting factor that gives more weight to those orientations that have a lower energy. The Boltzmann averaging of the interaction energy, overall orientations, leads to an angle-averaged interaction energy,

    (1.10)

    This angle-averaged interaction is usually referred to as the orientation or Keesom interaction. It is one of three important interactions that contribute to the van der Waals interaction between atoms and molecules. The interaction energy varies with the inverse sixth power of distance, as is the case with van der Waals interaction, which will be discussed later.

    1.2.4. Ion-Induced Dipole Interactions

    When a non-polar molecule is placed near an ion, the electric field of the ion can induce a dipole moment in the non-polar molecule, the strength of which depends on the polarizability (α) of the molecule. For a non-polar molecule, the polarizability arises from the displacement of its negatively charged electron cloud relative to the positively charged nucleus under the influence of an external electric field. For polar molecules, there are other contributions to the polarizability. The ion-induced dipole interaction energy between an ion of charge Q and an induced dipole of polarizability α can be written as

    (1.11)

    1.2.5. Dipole-Induced Dipole Interaction

    This represents the interaction between a polar and a non-polar molecule. This is similar to the ion-induced dipole interaction, but the polarizing field arises from the permanent dipole rather than from an ion. This depends on dipole moment μ, polarizability α and the angle of orientation of the dipole. The strength of this interaction is not sufficient to orient the molecules mutually, and the effective interaction is therefore angle-averaged energy. The angle-averaged interaction energy is

    (1.12)

    This is known as the Debye or induction interaction. It constitutes the second of the three contributions to the total van der Waals interaction energy between molecules.

    1.2.6. Induced Dipole–Induced Dipole Interactions

    This interaction exists between non-polar molecules. Even for a non-polar molecule, there exists an instantaneous dipole moment whose field will polarize a nearby neutral atom, giving rise to an attractive interaction analogous to the dipole-induced dipole interaction. The strength of this interaction depends on the polarizabilities (α1 and α2) and ionization potentials (I1 and I2) of the two neutral atoms. The interaction energy can be written as

    (1.13)

    This interaction is known as the London or dispersion interaction. This constitutes the third and most important contribution to the van der Waals force between atoms and molecules.

    1.2.7. Hydrogen Bonds

    This is a special kind of dipole–dipole interaction, which occurs when a hydrogen atom is covalently bonded to a highly electronegative atom such as oxygen or nitrogen. The unusually small size of the hydrogen atom and the large dipole moment associated with the NH or OH bond make it possible for other electronegative atoms to come closer to the hydrogen atom. This results in a strong attractive force that can align the neighbouring dipoles in both liquid and solid states. This type of interaction, which involves a directional dipole–dipole interaction mediated by a hydrogen atom, is known as hydrogen bonding interaction. The strength of this interaction lies between 10 and 40kJ/mol (25–100 kBT), i.e. intermediate in magnitude between a covalent bond and van der Waals interaction.

    1.2.8. Hydrophobic Interactions

    The ability of water molecules to form strong hydrogen bonds influences their interaction with other non-polar molecules that are incapable of forming hydrogen bonds. The presence of molecules that cannot take part in hydrogen bonds with water perturb the local structure of liquid water around them. The best configuration of water molecules around the non-polar molecule will be such that least number of tetrahedral charges point towards the non-polar species. The water molecules are locally made more ordered by the presence of a foreign molecule, which results in a decrease in the entropy of water molecules. The extra ordering imposed by two foreign molecules can be reduced if the two molecules are brought closer together. This leads to an effective attractive interaction between the foreign molecules and is known as hydrophobic interaction. This is an entropy-driven interaction, the magnitude of which is of the order of 10−20J.

    1.2.9. Depletion Interactions

    Depletion interaction arises in a mixture of colloidal particles of different sizes and is again an example of entropy-driven interaction. The most common case occurs when a colloidal solution contains a dissolved polymer that does not adsorb onto the surface of the colloidal particles. This situation is depicted in Figure 1.5.

    The polymer molecules, shown here as small spheres of radius R, are excluded from a region of thickness R away from the surface of the large particles – the depletion zone. At comparable volume fractions there will be far more small particles than large particles, and the entropy of the small particles (polymers) will dominate. The volume excluded to small spheres in the presence of large particles will be smaller if the excluded volumes of the two large spheres overlap. This leads to an increase in the entropy of the small particles. The large spheres will prefer to adopt those configurations that maximize the entropy of the small particles. Thus, there is an entropically driven attractive force between the large particles in the presence of large number of small particles. This is called a depletion force. Alternatively, this can be visualized as the attraction between large particles owing to the osmotic pressure difference of the solution in the bulk compared with the depletion layer. Thus, the interaction energy due to depletion can be written as

    where Vdep is the volume of the overlapping depletion zone between two colloidal particles, and Posm is the osmotic pressure of the solution, given by NkBT/V if there are N small spheres (polymers) in volume V of the solution.

    This summarizes some of the important forces that operate between atoms, molecules or even macroscopic bodies that are relevant to soft materials. It is worth mentioning that some of the intermolecular interactions will be further modified because of the mesoscopic structure of the system and can be different depending on the constituent units. This will be dealt separately in the forthcoming sections, with special reference to the individual building blocks. Three main classes of soft matter systems that have been studied classically are colloids, amphiphiles (surfactants) and polymers. We will discuss them separately in the following sections.

    1.3. Colloids

    A colloidal system consists of uniformly distributed dispersion of one phase in a dispersion medium (continuous phase). The colloidal regime (~1–1000nm) deals with an intermediate length scale between true solutions and bulk materials. This length scale comes from the requirement that they are significantly larger than the molecules of the dispersion medium, and the upper limit ensures that the particles’ Brownian motion is dominated by external forces such as gravitational settling. One important aspect of colloids compared with the bulk form of materials is that their surface to volume ratio is very large. For example, when a spherical particle of radius 1cm is divided into several particles of radius 100nm, there is a 10⁵ times increase in the total surface area. The large interfacial area associated with colloidal particles puts them in a thermodynamically unfavourable condition, as the interface always possesses a positive contribution to free energy. Owing to the high interfacial free energy associated with it, colloids are susceptible to agglomeration and growth and hence need to be stabilized by other means (kinetically). Some of the most common methods involve electrostatic (charge) stabilization or steric stabilization achieved by the addition of various additives such as surfactants, organic ligands or polymers. These stabilizing agents provide an energetic barrier to the agglomeration of particles through either Coulombic or steric repulsion between the particles. To understand this energy barrier, we need to consider the interparticle interaction between colloidal particles.

    1.3.1. Interactions Between Colloidal Particles

    The interparticle forces that operate between colloidal systems are primarily electrostatic in origin. However, their manifestations can be widely different, depending on the geometry, and are influenced by the presence of solvent, electrolytes, etc. We will consider each of these aspects separately in the forthcoming sections.

    van der Waals Interaction

    As we saw earlier, between any two bodies there exists an interaction due to fluctuating electromagnetic fields associated with their polarizabilities. Though these forces are not as strong as Coulombic or hydrogen bonding interactions, they play a central role in colloidal interactions because they are always present. To calculate the van der Waals interaction between two macroscopic bodies, we assume that the interaction is additive. First, let us consider the interaction between an atom and a flat surface consisting of the same atoms. Let us use the expression for attractive pair potential between two atoms as U(r)=−C/r⁶, where C is a proportionality constant that depends on the polarizability and ionization energy of the atoms according to the London-dispersion equation. Net interaction energy between an atom and a planar surface of solid made up of the same atoms can be obtained by adding up the interaction energies of all the atoms in the solid and the external atom. For specific geometries, the integral of the interaction energies over the volume of the solid can be solved analytically. The resulting expression for the variation of interaction energy between an atom and a flat surface, as a function of the distance (D) between the atom and the solid surface, can be written as

    (1.14)

    where ρ is the number density of atoms in the solid. Note that the interaction energy decays as 1/D³, much slower than 1/r⁶ dependence of interatomic interaction. By a similar approach, we can solve for the interaction energy between two flat surfaces of unit area. The resulting expression is

    (1.15)

    Conventionally, this expression has been used to represent the van der Waals interaction energy per unit area of two flat surfaces, in the form

    (1.16)

    where the constant A=πρ²C is known as the Hamaker constant. This constant is a material property and has the dimension of energy. Typical values of the Hamaker constant are of the order of 10−19J, for interactions in vacuum.

    Electrostatic Forces Between Surfaces

    We found that the van der Waals force between two similar particles in a medium is always attractive. This implies that two similar colloidal particles will coagulate in a medium, if only van der Waals force exists. However, this does not happen in many cases owing to the presence of other repulsive forces. The repulsive forces between two particles can arise from the charges on the surface of the particles, steric forces or salvation forces. Here, we consider the electrostatic repulsive forces arising from the surface charges. The origin of the surface charge can be through ionization of functional groups on the surface, specific adsorption of charged molecules or ions and differential solubility of the constituent ions in the crystal. The surface charge of the particle is balanced by an equal but oppositely charged region of counterions. The distribution of counterions around the surface of the particles is such that it creates a double layer of bound (Helmholtz) and diffused counterions. The counterion distribution and the resulting electrostatic potential around the surface can be derived from the Poisson–Boltzmann equation, under different approximations. The repulsive interaction energy per unit area due to the electrical double layers of two flat surfaces can be written as

    (1.17)

    where B is a constant that depends on the surface charge density, and κ is a constant that depends on the ionic strength of the medium. 1/κ is known as the Debye length and is given by

    (1.18)

    ni being the number density of ions of charge zi.

    1.3.2. DLVO Theory of Colloid Stability

    DLVO theory, named after the scientists Derjaguin, Landau, Verwey and Overbeek, states that the total interaction energy between two colloidal particles can be obtained by summing up the attractive van der Waals interaction and the repulsive electrostatic interaction. The net interaction energy between two flat charged surfaces can be obtained as

    (1.19)

    The form of the above equation is such that both at small and large separations, van der Waals attraction dominates, whereas at intermediate separation repulsive interaction dominates which depends on the ionic concentration in solution. Depending on the relative strengths of the two forces, an energy barrier may form between the two surfaces and this leads to kinetic stabilization of the particles. The height of the barrier is sensitive to the electrolyte concentration in the medium and can explain the salt-induced coagulation of colloidal particles. For a given surface charge density, the critical coagulation concentration can be estimated, which is found to vary inversely with the sixth power of the valency of the ions present. This result is consistent with the empirical Hardy–Schulz rule for coagulation of colloids.

    1.4. Surfactant Assemblies

    The name surfactant arises as an acronym for ‘surface active agent’. Surfactants are molecules that possess two groups of opposing solubility tendencies, typically a water-soluble polar group commonly called the ‘head group’ and an oil-soluble hydrocarbon or fluorocarbon group designated as the ‘tail group’. This hydrophilic–hydrophobic combination makes the molecules unique in their properties, hence they are often called amphiphilic molecules. One general way of classifying surfactants is based on the nature of the hydrophilic group. Depending on the chemical structure of the head group, surfactants are classified as anionic, cationic, non-ionic or zwitterionic.

    Anionic surfactants carry a negative charge on the head group. This category includes the traditional long-chain carboxylate soaps and the early synthetic detergents, the sulphonates and the sulphates. The major advantage of the sulphonates and sulphates over the carboxylates is their greater tolerance of divalent metal ions in hard water. Anionic surfactants are used extensively in cleaning formulations.

    The hydrophilic group of cationic surfactants carries a positive charge. They are usually quaternary ammonium, imidazonium or alkyl pyridinium compounds. Owing to the positive charge on the head group, they have strong affinity to negatively charged fibres such as cotton and hair, and hence they are used as fabric and hair conditioners.

    Non-ionic surfactants do not carry any charge on the head group. Their water solubility is derived from polar groups like alkyl ethoxylates. This class of surfactants includes several semi-polar compounds such as amine oxides, sulphoxides, phosphine oxides, pyrrolidones and alkanolamides. Non-ionic surfactants are extensively used in low-temperature detergents and as emulsifiers.

    Zwitterionic surfactants possess both positively and negatively charged groups in their hydrophilic part. Thus they can act as either cationic or anionic, depending on the pH of the solution. Betaines, sulphobetaines, naturally occurring surfactants of the class lecithin and phosphatidyl cholines, etc. come under this class of surfactants. These compounds are milder on the skin and have low eye irritation, which leads to their use in toiletries and baby shampoos.

    1.4.1. Surface Tension and Surface Activity

    The surface tension of a liquid is equal to the free energy change required to create a unit area of surface. For a liquid–vapour interface, if w is the interaction energy per pair of molecules (which is negative for attraction), and a is the effective area occupied by one molecule at the surface, the surface energy can be obtained as

    (1.20)

    where ns and nb represent the number of nearest neighbours at the surface and bulk, respectively. Because there are fewer nearest neighbours for molecules at the surface compared with those in the bulk and w is negative for condensed phases, the surface energy is always positive. A surfactant solution has a surface populated by adsorbed molecules, which reduces the surface energy. Adsorption of surfactants at the air–liquid interface plays a key role in many commercial applications such as foaming and emulsification.

    Adsorption of surfactants at an interface is usually investigated by measuring the interfacial tension of the surfactant solution, although there are direct methods for obtaining the surface concentrations. Measurement of interfacial tension allows one to obtain the amount of the surfactant adsorbed at the interface by an indirect method using the famous Gibbs adsorption equation.

    A relationship between the surface tension and the surface excess concentration of the surfactant is usually obtained by developing a surface equation of state, treating the adsorbed surface as an isolated two-dimensional gas or liquid. This has been done thermodynamically by the treatment of Gibbs, and for a single-component non-ionic surfactant the Gibbs equation takes the form

    (1.21)

    where Γ is the surface excess concentration of the surfactant, and c is the bulk concentration. The maximum surface excess concentration Γmax, expressed in units of moles per unit area, can be used to calculate the area occupied by one molecule at the interface, assuming a complete monolayer type of adsorption at the saturation surface excess concentration. The area per molecule at the liquid–gas interface (Ao) can be obtained as

    (1.22)

    where NA is Avogadro’s number.

    To understand the adsorption ability of the mixture much better, one can combine the Gibbs equation with the Langmuir adsorption isotherm such that the surface coverage at any concentration can be defined as the ratio Γ/Γmax. According to the Langmuir isotherm for a monolayer adsorption, this can be written as

    (1.23)

    where θ is the surface coverage, and K is an empirical constant known as the adsorption coefficient. The higher the value of K, the better is the ability of the adsorbent to adsorb on a surface.

    1.4.2. Surfactant Aggregation and Hydrophobic Effect

    Above a narrow range of concentration that is characteristic of each solvent–solute system, the amphiphilic molecules associate themselves to form aggregates called micelles. This concentration above which appreciable amounts of micelles are formed is termed the ‘critical micelle concentration’ (CMC). There is a sudden, well-defined change in some of the physico-chemical properties of aqueous surfactant solutions above the CMC. Some of the important physical properties that have been found to exhibit this behaviour are the interfacial tension, equivalent conductivity, turbidity, osmotic pressure, self-diffusion coefficient, solubilization and viscosity.

    Figure 1.6 illustrates the variation in concentration dependence of a wide range of physico-chemical quantities of aqueous surfactant solutions around the CMC. One can clearly observe that over a narrow range of concentrations almost all physical properties of the solutions suffer a discontinuity in their variation with concentration. This sudden change in their bulk properties corresponds to the formation of micellar aggregates and hence is used to determine the CMC of surfactant solutions. However, it is worth mentioning that this change occurs over a narrow concentration range rather than at a precise point, and the magnitude of the CMC obtained depends on the property being measured.

    In principle any of the physical properties illustrated in Figure 1.6 could be used to determine the CMC, but one of the widely used techniques is based on surface tension. Beyond the CMC, the concentration of the surfactant in the bulk remains constant and results in a nearly constant value of surface tension once the micelle starts forming. Thus a plot of surface tension versus log (concentration) shows a break corresponding to the CMC.

    The standard free energy of transfer of a single hydrocarbon molecule from aqueous phase into oil is large and negative, reflecting the obvious fact that non-polar oils have extremely small solubilities in water. Similar behaviour would be expected for the hydrophobic tails of surfactant molecules. The thermodynamics of micelle formation shows that the enthalpy of micellization in aqueous solution is probably positive: that is, they are endothermic. That they do form micelles above the CMC indicates that their free energy of formation, ΔG, must be negative. Because ΔGHTΔS and the enthalpy of formation, ΔH, is positive, its entropy change, ΔS, should be positive. The positive entropy change associated with micellization, even though the molecules are clustering together, indicates a contribution to the entropy from the solvent. When a hydrocarbon molecule is surrounded by water, the water molecules form a clathrate cage. As a result of thisacquisition of structure, the entropy of water decreases. Once the surfactant molecules have been herded into small clusters, individual molecules no longer have to be held in solvent cages and hence they are less constrained. This grouping together of hydrophobic sections of the molecules results in an increase in the entropy of the solvent molecules, which is the origin of the hydrophobic interaction.

    1.4.3. Thermodynamics of Micelle Formation

    A knowledge of the changes in thermodynamic quantities upon micellization is important not only for understanding the forces at play during micellization but also for predicting the behaviour of micellar solutions upon changes in thermodynamic parameters such as temperature and pressure. Several models exist in the literature for describing the micellization phenomenon, but a very useful tool for the description of micelle formation is the mass action model [3]. In this model the micelle is treated as a dynamic species which is in equilibrium with its monomers. In the case of an ionic surfactant, the molecule is considered as a 1:1 electrolyte which undergoes complete dissociation, whereas in the micelle the dissociation is not complete. Hence the micellar system is composed of free surfactant molecules and counterions which are in equilibrium with the micelles. For a cationic surfactant AB, with A+ as the hydrophobic part and B− as the counterion, the micelle formation is assumed to take place by a single step reaction represented as

    (1.24)

    where n is the aggregation number, and m/n is the micelle ionization degree. Applying the law of mass action to the above equilibrium, we get the equilibrium constant, K, as

    (1.25)

    The standard free energy of micellization per monomer is given as

    (1.26)

    (1.27)

    (1.28)

    At the CMC, [A]≈[B]≈CMC. It is found in practice that for surfactants with alkyl chain of C8 or longer, n becomes sufficiently large so that the third term in Eq. (1.28) can be neglected. Hence Eq. (1.28) approximates to

    (1.29)

    In the case of non-ionic micelles the mass law treatment has been applied more successfully and the equation is simplified to

    (1.30)

    The enthalpy change accompanying micellization can be calculated using the well-known Gibbs–Helmholtz equation, which is given as

    (1.31)

    By using this expression one can easily show that the enthalpy of micellization can be obtained as

    (1.32)

    In general, but not always, the micelle formation is found to be an exothermic process, favoured by a decrease in temperature. The enthalpy of micellization may therefore be either positive or negative, depending on the system and conditions. The process, however, always has a substantial positive entropic contribution to overcome any positive enthalpy term, so that micelle formation is primarily an entropy-driven process.

    1.4.4. Dynamics of Micelle Formation

    Micelles are known to be dynamic species in which the monomer rapidly joins and leaves the micelle whose aggregation number represents only an average over time. Different methods are used to study the kinetics of such dynamic processes in micellar solutions and they usually involve relaxation techniques. The most commonly used are stopped flow, temperature jump, pressure jump, ultrasonic relaxation, nuclear magnetic resonance (NMR), electron paramagnetic resonance (EPR), etc. All those studies have revealed that two major relaxation processes exist in micellar dynamics: one occurring in the microsecond range, the other in the millisecond range. It is now well established that the faster relaxation process which is in the microsecond timescale is due to the release of a single surfactant molecule from the micelle and its subsequent incorporation into the micelle. The slow relaxation process occurring in the millisecond range represents the total dissolution of the micelle into its monomers and its subsequent re-association.

    1.4.5. Phase Behaviour of Surfactants

    The phase behaviour of surfactants and its variation with temperature depends largely on the nature of the hydrophilic group. With ionic surfactants the solubility increases slowly initially with an increase in temperature until a value is reached at which there is a sudden increase in the solubility. This temperature is often referred as the Kraft temperature or simply the Kraft point. If both solubility and the CMC are plotted as a function of temperature, one finds an intersection point between the solubility and CMC curves. This temperature at which the intersection occurs is the Kraft point, and it is the temperature at which the solubility becomes equal to the CMC. At the Kraft point, there is an equilibrium between the micelles, monomers and solid surfactant.

    Non-ionic surfactants, unlike ionic surfactants, are very sensitive to temperature changes. If a dilute solution of non-ionic surfactant is heated, then above a critical temperature strong light scattering (LS) is observed and the solution becomes cloudy. This temperature at which this cloudiness occurs is designated as the cloud point of the surfactant solution at this concentration. The cloud point of a surfactant solution obviously depends on its concentration.

    Below the cloud point curve, various phases may be distinguished depending on the surfactant concentration. In dilute solutions, when concentration is not too far from the CMC, the micelles remain more or less spherical in shape and are randomly distributed in solution phase and the solution is isotropic. As the surfactant concentration is increased, it simply forces the micelles to come closer, thereby increasing the extent of intermicellar interaction. The formation of a series of regular geometries, as volume fraction is increased, is a way in which the surfaces can be allowed to maximize their separation. Initially there is a transition from more or less spherical to cylindrical or rod-like micelles. At surfactant concentrations of perhaps 20–30% by weight, a new phase appears which is birefringent and quite viscous. This phase consists of many long, parallel rod-like micelles arranged in a hexagonal array. The micelle interiors are apparently rather fluid, resembling a liquid hydrocarbon in many respects. This phase is a liquid crystalline phase and is normally called the normal hexagonal (H1) or middle phase. At even higher surfactant concentrations the arrangement of surfactant molecules into bilayers becomes favourable and another liquid crystalline phase known as lamellar () or neat phase appears. This phase is built up from flexible bilayer sheets of indefinite area but arranged parallel to each other. Lamellar phases are less viscous than hexagonal phase though they contain less water. This is because of the ease with which the parallel layers can slide over each other during shear.

    Another important category of supramolecular aggregates that are often encountered in phospholipids are the mesophases of closed bilayers capable of entrapping ions in their aqueous interiors. These dispersions came to be known as vesicles or liposomes, and they attained considerable attention as membrane models. There have been reports of spontaneous vesicle formation in certain aqueous mixtures of commercially available single-tailed surfactants with oppositely charged head groups [4]. This way of vesicle preparation offered a remarkably simple way of tailoring vesicle properties and their surface charges. This allows efficient encapsulation to take place without mechanical or chemical perturbation of the final vesicle composition or structure. Figure 1.7 depicts, pictorially, some of the major structures often encountered in surfactant–water systems.

    1.4.6. Packing Parameter and Bending Rigidity

    As mentioned in the previous section, there are different types of aggregate structures

    Enjoying the preview?
    Page 1 of 1