You are on page 1of 97

CHEMICAL SENSORS

FUNDAMENTALS OF SENSING MATERIALS


VOLUME 3: POLYMERS AND OTHER MATERIALS

CHEMICAL SENSORS
FUNDAMENTALS OF SENSING MATERIALS
VOLUME 3: POLYMERS AND OTHER MATERIALS
EDITED BY

GHENADII KOROTCENKOV
GWANGJU INSTITUTE OF SCIENCE AND TECHNOLOGY GWANGJU, REPUBLIC OF KOREA

MOMENTUM PRESS, LLC, NEW YORK

Chemical Sensors: Fundamentals of Sensing Materials. Volume 3: Polymers and Other Materials Copyright Momentum Press, LLC, 2010 All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or by any meanselectronic, mechanical, photocopy, recording or any other except for brief quotations, not to exceed 400 words, without the prior permission of the publisher. First published in 2010 by Momentum Press, LLC 222 East 46th Street, New York, NY 10017 www.momentumpress.net ISBN-13: 978-1-60650-230-3 (hard back, case bound) ISBN-10: 1-60650-2309-1 (hard back, case bound) ISBN-13: 978-1-60650-232-7 (e-book) ISBN-10: 1-60650-232-8 (e-book) DOI forthcoming Cover design by Jonathan Pennell Interior design by Derryfield Publishing, LLC First Edition: December 2010 10 9 8 7 6 5 4 3 2 1 Printed in the United States of America

CONTENTS
PREFACE TO CHEMICAL SENSORS: FUNDAMENTALS OF SENSING MATERIALS PREFACE TO VOLUME 3: POLYMERS AND OTHER MATERIALS ABOUT THE EDITOR CONTRIBUTORS 1 POLYMERS IN CHEMICAL SENSORS B. Adhikari P. Kar 1 Introduction 2 What Are Polymers? 3 Parameters of Polymers Promising for Chemical Sensor Application 4 Synthesis of Polymers 5 Deposition of Polymers 6 Functionalization of Polymers 6.1 Structure Modification 6.2 Surface Modification 6.3 Composition Modification 7 Polymers in Chemical Sensors 7.1 7.2 7.3 7.4 7.5 Optical and Fiber Optic Polymer-Based Sensors Conductometric Gas Sensors SAW and QCM Polymer-Based Sensors Electrochemical Polymer-Based Sensors Chemically Sensitive FET-Based Sensors xi xiii xv xvii 1

1 3 4 7 9 13 13 14 15 16 20 27 40 45 57 61 61 62

8 Outlook 9 Acknowledgments References

vi CONTENTS

MOLECULAR IMPRINTING (TEMPLATING)A PROMISING APPROACH FOR DESIGN OF POLYMER-BASED CHEMICAL SENSORS G. Korotcenkov B. K. Cho 1 Introduction 2 General Principles of Molecular Imprinting (Templating) 3 Methods of Imprinting (Templating) 3.1 In-Block Imprinted Polymers 3.2 In Situ Imprinted Polymers 3.3 Polymer-Imprinted Beads 4 Components of Imprinting Technology 4.1 4.2 4.3 4.4 4.5 Target Molecules The Imprinting Matrix Cross-Linkers Solvents (Porogens) Initiators

77

77 79 80 80 82 82 83 83 85 86 88 89 89 91 92 92 92 93 99 101 103 106 108 109 117

5 MIP Preparation Methods 6 Combination of MIPs and Monomolecular Host Molecules 7 Control of the Imprinting Effect 8 Application of Imprinting Polymers in Chemical Sensors 8.1 8.2 8.3 8.4 Advantages of MIP-Based Chemical Sensors Detection Principles Used in MIP Chemical Sensors Interfacing the MIP with the Transducer Factors Controlling the Sensing Characteristics of MIPs-Based Chemical Sensors 8.5 Micro- and Nanofabricated MIPs 9 Outlook 10 Acknowledgments References 3 CALIXARENE-BASED MATERIALS FOR CHEMICAL SENSORS H. M. Chawla N. Pant S. Kumar D. StC. Black N. Kumar 1 Introduction

117

CONTENTS vii

2 Molecular Receptors and Generation of Signal for Sensing Target Species 3 Calixarenes and Thiacalixarenes 4 Synthesis of Calix[n]arenes 4.1 Base-Catalyzed Condensation Reactions 4.2 Acid-Catalyzed Condensation Reactions 5 Synthesis of Thiacalix[n]arenes (Sulfur-Bridged Calixarenes) 6 Physical Properties of Calixarenes and Tetrathiacalixarenes 6.1 Melting points 6.2 Solubilities and pKa Values 7 Spectral Properties and Characterization of Calixarenes 7.1 Infrared Spectra 7.2 Ultraviolet Spectra 7.3 NMR Spectra 8 Conformational Structures of Calixarenes and Thiacalixarenes 9 Conformational Characterization of Calix[n]arenes 10 Calixarenes as Materials for Chemical Sensors 11 Calixarene-Based Materials for Recognition of Alkali and Alkaline Earth Metal Ions 12 Calixarene-Based Materials for Recognition of Transition and HeavyMetal Ions 13 Calixarene-Based Materials as Dual Probes for Sensing and Extraction 14 Calixarene-Based Materials for Sensing Lanthanides and Actinides 15 Sensor Materials Based on Polymeric Calixarenes 16 Naked-Eye Sensing: Calixarene-Based Chromogenic Materials for Sensing Ions and Molecules 17 Calixarene-Based Electroactive Sensing Materials 18 Calixarene-Based Materials for Sensing Anions 18.1 Calixarene-Based Electron-Deficient or Positively Charged Anion Receptors 18.2 Calixarene-Based Neutral Anion Receptors 18.3 Calixarene-Based Ditopic Molecular Receptors 19 Calixarene-Based Sensor Materials for Neutral Molecules and Biological Amines 20 Calixarene-Based Materials for Gas Sensors 21 Outlook

119 120 123 123 124 124 127 127 127 128 128 128 129 129 132 132 133 137 139 139 143 143 156 161 161 165 175 176 179 181

viii CONTENTS

22 Acknowledgments References 4 BIOLOGICAL AND BIOMIMETIC SYSTEMS IN CHEMICAL SENSORS R. Jelinek S. Kolusheva 1 Introduction 2 Polymers and Polymer/Biomolecule Assemblies 2.1 Conductive Polymers 2.2 Luminescent Conjugated Polymers 3 Membranes in Chemical Sensors 3.1 Chemical Membranes 3.2 Biological Membranes 4 Biomimetic Systems for Molecular and Ionic Recognition 4.1 4.2 4.3 4.4 4.5 Biological Receptors and Channels Synthetic Receptors Biomimetic Enzyme-Based Sensors Nanobiosensors Other Biomimetic Sensors

181 182 201

201 202 202 205 210 210 213 219 219 221 223 225 232 242 242 244 247 248 248 263

5 Monolayers and Films 5.1 Self-Assembled Monolayers 5.2 Langmuir-Blodgett Films 6 Challenges and Limitations of Biosensors 7 Conclusions and Outlook References 5 NOVEL SEMICONDUCTOR MATERIALS FOR THE DEVELOPMENT OF CHEMICAL SENSORS N. Chaniotakis N. Sofikiti V. Vamvakaki 1 Introduction 2 The Silicon EraClassical Semiconductors in Chemical Sensing 3 Fundamentals of Sensor Development 4 Surface Chemistry of Semiconductors in Chemical Sensing 5 Band Gap Theory and Its Relationship to Sensor Design 6 Pinning of the Surface Fermi Level

263 265 269 269 271 272

CONTENTS ix

7 New Semiconductor Substrates 7.1 Diamond 7.2 Silicon Carbide 7.3 Gallium Nitride and III-Nitrides 8 Nanosemiconductor Structures in Chemical Sensors 9 Forecasting the Future References 6 ION CONDUCTORS AND THEIR APPLICATIONS IN CHEMICAL SENSORS R. V. Kumar C. Schwandt 1 Introduction 1.1 Solid Electrolytes 1.2 Chemical Sensors 2 Ionic Conduction in Solids 3 Oxygen IonConducting Solid Electrolytes 3.1 3.2 3.3 3.4 4.1 4.2 4.3 4.4 4.5 Zirconia-Based Solid Electrolytes Defect Chemistry of Stabilized Zirconia Preparation of Stabilized Zirconia Oxygen Sensors Based on Stabilized Zirconia High-Temperature Proton-Conducting Solid Electrolytes Defect Chemistry of Substituted Perovskites Preparation of Substituted Perovskites Hydrogen Sensors Based on Substituted Perovskites Low-Temperature Proton-Conducting Solid Electrolytes

273 274 277 279 281 283 284 291

291 292 293 294 296 297 304 305 308 319 319 320 323 324 330 332 332 337 343 344 351 351 358

4 Proton-Conducting Solid Electrolytes

5 Metal IonConducting Solid Electrolytes 5.1 Defect Chemistry and Preparation of -Aluminas 5.2 Sensors Based on -Aluminas 6 Outlook and Future Trends References 7 SENSOR MATERIALS: SELECTION GUIDE G. Korotcenkov 1 Acceptable Materials for Chemical Sensors 2 Which Metal Oxides Are Better for Gas Sensors?

x CONTENTS

3 Choosing a Polymer for a Chemical Sensor Application 4 Technological Limitations in Sensing Material Applications 5 Future Trends 6 Toward a Theory of Chemical Sensors 7 Summary 8 Acknowledgments References INDEX

362 362 363 368 370 370 370 375

PREFACE TO CHEMICAL SENSORS: FUNDAMENTALS OF SENSING MATERIALS


Sensing materials play a key role in the successful implementation of chemical and biological sensors. The multidimensional nature of the interactions between function and composition, preparation method, and end-use conditions of sensing materials often makes their rational design for real-world applications very challenging. The world of sensing materials is very broad. Practically all well-known materials could be used for the elaboration of chemical sensors. Therefore, in this series we have tried to include the widest possible number of materials for these purposes and to evaluate their real advantages and shortcomings. Our main idea was to create a really useful encyclopedia or handbook of chemical sensing materials, which could combine in compact editions the basic principles of chemical sensing, the main properties of sensing materials, the particulars of their synthesis and deposition, and their present or potential applications in chemical sensors. Thus, most of the materials used in chemical sensors are considered in the various chapters of these volumes. It is necessary to note that, notwithstanding the wide interest and use of chemical sensors, at the time the idea to develop these volumes was conceived, there was no recent comprehensive review or any general summing up of the fundamentals of sensing materials The majority of books published in the field of chemical sensors were dedicated mainly to analysis of particular types of devices. This threevolume review series is therefore timely. This series, Chemical Sensors: Fundamentals of Sensing Materials, offers the most recent advances in all key aspects of development and applications of various materials for design of chemical sensors. Regarding the division of this series into three parts, our choice was to devote the first volume to the fundamentals of chemical sensing materials and processes and to devote the second and third volumes to properties and applications of individual types of sensing materials. This explains why, in Volume 1: General Approaches, we provide a brief description of chemical sensors, and then detailed discussion of desired properties for sensing materials, followed by chapters devoted to methods of synthesis, deposition, and modification of sensing materials. The first volume also provides general background information about processes that participate in chemical sensing. Thus the aim of this volume, although not exhaustive, is to provide basic knowledge about sensing materials, technologies used for their preparation, and then a general overview of their application in the development of chemical sensors.

xi

xii PREFACE TO CHEMICAL SENSORS: FUNDAMENTALS OF SENSING MATERIALS

Considering the importance of nanostructured materials for further development of chemical sensors, we have selected and collected information about those materials in Volume 2: Nanostructured Materials. In this volume, materials such as one-dimension metal oxide nanostructures, carbon nanotubes, fullerenes, metal nanoparticles, and nanoclusters are considered. Nanocomposites, porous semiconductors, ordered mesoporous materials, and zeolites also are among materials of this type. Volume 3: Polymers and Other Materials, is a compilation of review chapters detailing applications of chemical sensor materials such as polymers, calixarenes, biological and biomimetric systems, novel semiconductor materials, and ionic conductors. Chemical sensors based on these materials comprise a large part of the chemical sensors market. Of course, not all materials are covered equally. In many cases, the level of detailed elaboration was determined by their significance and interest shown in that class of materials for chemical sensor design. While the title of this series suggests that the work is aimed mainly at materials scientists, this is not so. Many of those who should find this book useful will be chemists, physicists, or engineers who are dealing with chemical sensors, analytical chemistry, metal oxides, polymers, and other materials and devices. In fact, some readers may have only a superficial background in chemistry and physics. These volumes are addressed to the rapidly growing number of active practitioners and those who are interested in starting research in the field of materials for chemical sensors and biosensors, directors of industrial and government research centers, laboratory supervisors and managers, students and lecturers. We believe that this series will be of interest to readers because of its several innovative aspects. First, it provides a detailed description and analysis of strategies for setting up successful processes for screening sensing materials for chemical sensors. Second, it summarizes the advances and the remaining challenges, and then goes on to suggest opportunities for research on chemical sensors based on polymeric, inorganic, and biological sensing materials. Third, it provides insight into how to improve the efficiency of chemical sensing through optimization of sensing material parameters, including composition, structure, electrophysical, chemical, electronic, and catalytic properties. We express our gratitude to the contributing authors for their efforts in preparing their chapters. We also express our gratitude to Momentum Press for giving us the opportunity to publish this series. We especially thank Joel Stein at Momentum Press for his patience during the development of this project and for encouraging us during the various stages of preparation. Ghenadii Korotcenkov

PREFACE TO VOLUME 3: POLYMERS AND OTHER MATERIALS


This volume covers a variety of topics in the rapidly developing field of chemical sensors. The purpose of this volume is to explain and illustrate the use of multifunctional materials such as polymers, calixarenes, ion conductors, biological systems, and novel semiconductors in chemical sensors. These materials differ fundamentally from standard metal oxides and metals, so their application provides opportunities to design sensors based on entirely different mechanisms of sensing. As a result, new trends in the elaboration of sensors with different functional attributes and for building instruments with previously unavailable capabilities demanded by new applications have opened up. Therefore, this book is intended to be a primary source for both fundamental and practical information related to these multifunctional materials, which will be necessary for future development. This volume comprises seven chapters written by active researchers who are well-known experts in their fields and who have made significant contributions to the field over the past several years. Thus, this book presents the most recent advances in all the key aspects of development and application of polymers and other multifunctional materials for chemical and biological analysis. The chapters in this book have been written to give the reader the big picture, from the design phase to implementation of chemical sensors of various types. Every chapter addresses the particulars of multifunctional materials synthesis and characterization. In every chapter you will also find descriptions of a very wide range of devices that may be designed using such multifunctional materials.. These chapters thus highlight the materials, the physics, the devices, and even key fabrication issues. We hope that the information presented in this volume will help the reader understand the details of sensing material design for specific applications and establish quantitative structurefunction relationships. The intended audience is scientists, researchers, and engineers in industries and research laboratories. With its many references to the vast resources of recently published literature on the subject, this book serves as a significant and insightful source of valuable information pertaining to the ongoing scientific debates, the current state of understanding, and future directions. Students will also find the book to be very useful in their research and understanding of chemical sensors and multifunctional materials. The structure of this book offers a basis for a high-throughput instrumentation course at the

xiii

xiv PREFACE TO VOLUME 3: POLYMERS & OTHER MATERIALS

advanced undergraduate or graduate level. As such, it should be very useful to university post-docs and professors as well. Ghenadii Korotcenkov

ABOUT THE EDITOR

Ghenadii Korotcenkov received his Ph.D. in Physics and Technology of Semiconductor Materials and Devices in 1976, and his Habilitate Degree (Dr.Sci.) in Physics and Mathematics of Semiconductors and Dielectrics in 1990. For a long time he was a leader of the scientific Gas Sensor Group and manager of various national and international scientific and engineering projects carried out in the Laboratory of Micro- and Optoelectronics, Technical University of Moldova. Currently, he is a research professor at Gwangju Institute of Science and Technology, Gwangju, Republic of Korea. Specialists from the former Soviet Union know G. Korotcenkovs research results in the study of Schottky barriers, MOS structures, native oxides, and photoreceivers based on Group IIIV compounds very well. His current research interests include materials science and surface science, focused on metal oxides and solid-state gas sensor design. He is the author of five books and special publications, nine invited review papers, several book chapters, and more than 180 peer-reviewed articles. He holds 16 patents. He has presented more than 200 reports at national and international conferences. His articles are cited more than 150 times per year. His research activities have been honored by the Award of the Supreme Council of Science and Advanced Technology of the Republic of Moldova (2004), The Prize of the Presidents of Academies of Sciences of Ukraine, Belarus and Moldova (2003), the Senior Research Excellence Award of Technical University of Moldova (2001, 2003, 2005), a Fellowship from the International Research Exchange Board (1998), and the National Youth Prize of the Republic of Moldova (1980), among others.

xv

CONTRIBUTORS
Basudam Adhikari (Chapter 1) Materials Science Centre Indian Institute of Technology Kharagpur 721302, India David St. Clair Black (Chapter 3) School of Chemistry University of New South Wales Sydney 2052, New South Wales, Australia Nikos Chaniotakis (Chapter 5) Laboratory of Analytical Chemistry Department of Chemistry University of Crete Voutes 71003 Iraklion, Crete, Greece Har Mohindra Chawla (Chapter 3) Department of Chemistry Indian Institute of Technology Delhi New Delhi 110016, India Beongki Cho (Chapter 2) Department of Material Science and Engineering Gwangju Institute of Science and Technology Gwangju 500-712, Republic of Korea Raz Jelinek (Chapter 4) Department of Chemistry and Ilse Katz Institute for Nanotechnology Ben Gurion University Beer Sheva 84105, Israel

xvii

xviii CONTRIBUTORS

Pradip Kar (Chapter 1) Polymer Engineering Department Birla Institute of Technology, Mesra Ranchi 835215, India Sofiya Kolusheva (Chapter 4) Department of Chemistry and Ilse Katz Institute for Nanotechnology Ben Gurion University Beer Sheva 84105, Israel Ghenadii Korotcenkov (Chapters 2 and 7) Department of Material Science and Engineering Gwangju Institute of Science and Technology Gwangju 500-712, Republic of Korea and Technical University of Moldova Chisinau, Republic of Moldova Naresh Kumar (Chapter 3) School of Chemistry University of New South Wales Sydney 2052, New South Wales, Australia Ramachandran Vasant Kumar (Chapter 6) Department of Materials Science and Metallurgy University of Cambridge Cambridge CB2 3QZ, United Kingdom S. Kumar (Chapter 3) Department of Chemistry Indian Institute of Technology Delhi New Delhi 110016, India Nalin Pant (Chapter 3) Department of Chemistry Indian Institute of Technology Delhi New Delhi 110016, India Carsten Schwandt (Chapter 6) Department of Materials Science and Metallurgy University of Cambridge Cambridge CB2 3QZ, United Kingdom

CONTRIBUTORS xix

Nikoletta Sofikiti (Chapter 5) Laboratory of Analytical Chemistry Department of Chemistry University of Crete Voutes 71003 Iraklion, Crete, Greece Vicky Vamvakaki (Chapter 5) Laboratory of Analytical Chemistry Department of Chemistry University of Crete, Voutes 71003 Iraklion, Crete, Greece

CHAPTER 1

POLYMERS IN CHEMICAL SENSORS

B. Adhikari P. Kar 1. INTRODUCTION


Plants and animals have built-in natural sensor devices targeted either for detection of external agencies in their surroundings or for performing some specific function. Plants have electromagnetic detectors for sensing external attacks. Animals have devices for sensing through five different sense organsthe tongue, skin, eye, ear, and nose. They use these organs to perform normal activities as well as to remain away from unfavorable situations. To help them remain away from deadly poisonous chemical substances, for instance, animals are alerted by smell and by irritation of the eyes and skin. Both the feeling of irritation and detection of abnormal odors occur via their built-in sensor system. These sensor systems, formed through biosynthesis from organic molecules, are embedded in an aqueous environment containing soluble electrolytic salts. Thus the sensing network in living systems is organic as well as polymeric in nature. Gathering knowledge by studying biological sensing systems and their functioning can help in mimicking such sensor systems using synthetic macromolecules. Such sensor networks consist of a chemical detection or recognition element, a transducing element, and a signal-processing element, which appear to be extremely complex even though their function and response appear to be very simple and spontaneous (see Figure 1.1). The first element is a chemical detection element of some sensing material, which interacts with the environment and generates a response. The second element is a transducer, which reads the response from the chemical detection element and converts it into an interpretable and quantifiable term for the third element, the signal processor. The chemical detection element is the

2 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Analyte or substrate Chemical detection element (polymer) Transducer Signal processor

Figure 1.1. Schematic representation of simplied sensor setup.

heart of the sensor system and can be considered the primary part of the sensor. The response, recovery, selectivity, and sensitivity of a chemical sensor depend on the chemical detection element used. Sensor technology depends on progress in materials science and technology for this chemical detection element layer. The choice of a particular interactive material is based on the sensitivity, selectivity, reliability, and the reversibility of the related sensing mechanisms. Various chemical detection materials are available, including metal and metal oxide semiconductors, solid electrolytes, insulators, catalytic materials, polymers, composites, and others. Exposure to toxic and hazardous chemicals may cause serious problems to mammals, such as irritation, vomiting, suffocation, or illness; the chemical may even be deadly poisonous. Thus, as a measure of protection to living bodies, it is necessary to detect hazardous and toxic chemicals using artificial sensor devices. Toward that goal, sensing of toxic and hazardous chemicals is an emerging field, in which modern research tries to develop more efficient sensor devices than those in living bodies. The chemical sensors field is one of the fastest-growing areas in both research and commercial application. During the last 25 years, global research and development in the field of sensors has expanded exponentially in terms of financial investment, published literature, and number of active researchers. Most of the research work in this field is concentrated toward reducing the size of sensors and enabling identification and quantification of multiple species. Since easy handling, quick response, good reversibility and reproducibility, sensitivity, and selectivity are qualities expected of an excellent sensor, there is a need for further research. Exploring the present state of the art of materials used in chemical sensors is our primary goal in this chapter, with special reference to the role of organic polymers as chemical detecting elements in sensors. Since chemical sensors need to either detect or estimate chemical analytes, the sensing element should be selective as well as interactive with the analyte. In terms of general operating principles, there are three major categories of chemical sensors: electrochemical, optical, and mass sensors. Therefore, the detecting element, which is the key element of a sensor device, must respond to electrochemical or optical stimulation or undergo structural changes due to a change in its mass as a result of absorption of a minute quantity of an analyte. From this point of view, polymers represent a class of highly tailorable materials, which have already been qualified as detecting elements that respond at ambient temperature to chemical or electrochemical stimulation, to optical stimulation, and that express piezoelectric behavior after absorbing a small mass. In addition, polymers are easy to synthesize and process to develop a suitable device for sensor application. The properties of polymers, such as their polyelectrolytic nature, intrinsic conductivity, electrochromism, etc., have made them todays materials of choice in modern sensor devices, gradually replacing the metal oxides and inorganic semiconductors that earlier dominated the field.

POLYMERS IN CHEMICAL SENSORS 3

2. WHAT ARE POLYMERS?


In contrast to discrete small molecular compounds, polymers are macromolecules. Except for a few, polymers are organic macromolecules made of carbon and hydrogen atoms in major percentage with some heteroatoms such as nitrogen, oxygen, sulfur, phosphorous, halogens, etc., as minor constituents. A polymer molecule is formed by the repetitive union of a large number of reactive small molecules in a regular sequence (see Figure 1.2). The repeated unit in the backbone of the polymer molecule is known as the mer unit, and the reactive small molecule from which the polymer is formed is called the monomer. The simplest example of a polymer is polyethylene, in which the ethylene moiety is the mer unit.

H N
n

S
n

N H

n
Polypyrrole

R
Poly(3 alkylthiophene)

Polyacetylene

Polyaniline

S n
Polythiophene

O
n
Polyfuran

S
n

Polyphenylene

Polyphenylene sulfide

S
n

n
Polyisothianapthene

Polyphenylenevinylene

Polythienylenevinylene

N H

Polyazulene

Polycarbazole

Polyfluorene

Figure 1.2. Repeating unit structures of some common conducting polymers.

4 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

As a class, polymers are unique over other materials with respect to their tailorability and broad range of properties as well as versatility. In terms of size and molecular weight, in general, polymers are more than a million times bigger than small molecular compounds. Properties of polymers, in general, depend on their chemical composition, molecular structure, molecular weight, molecular-weight distribution, and morphology. Morphologically, polymers are quasi-crystalline in nature, having small crystallites dispersed in an amorphous matrix in which a single molecule may extend from one amorphous region to a distant amorphous region while passing through several crystallite regions. Both the bulk and the surface of a polymer sample may or may not contain active functional groups which can respond to a stimulus in chemical sensing. Although primary covalent bonds predominate in polymers, secondary bonding influences both their processing and their functional performance. Due consideration should be given to the role of secondary bonding on the interaction of a polymer with the analyte during the sensing function. Extensive secondary bonding interaction can be a major cause of insolubility of a polymer in solvents, which may restrict its processability to produce a suitable device for chemical sensing. Interpretation of sensing response and recovery is easier if it can be correlated with chemical bonding in the polymer. On the other hand, organic macromolecules are known which can exhibit ionic or electronic conduction properties. These are polymers that have ionizable functional groups, viz., polyelectrolytes, and extended -electron conjugation, viz., intrinsically conducting polymers. In general, polymers are electrically insulating in nature due to the nonavailability of free electrons, since the four valence electrons of carbon are fully saturated. However, some organic polymer molecules are semiconducting in nature, by virtue of their extended -electron conjugation along the backbone chain of the macromolecule. The ability of conducting polymers to conduct electricity depends on the alternating double bondsingle bond structure in the polymer backbone, coupled with the formation of some charged centers on the chain by partial oxidation. The introduction of such extra charges on the polymer by doping (in analogy to inorganic semiconductors), alters the conductivity of such polymers from almost insulators to something approaching a metallic conductor. Due to the more reactive nature of the ionic groups in polyelectrolytes and -electron conjugation in conducting polymers, these are the important sensing sites for corresponding analyte compounds.

3. PARAMETERS OF POLYMERS PROMISING FOR CHEMICAL SENSOR APPLICATION


It has been pointed out that built-in sensing devices in biological systems are polymeric in nature, although they are complex. The actual structure and chemistry of the polymers in the biological sensing devices of various sense organs are not known. However, by virtue of their light weight, ease of synthesis, good processability, stability, nonhazardous nature, and low cost, polymers offer a lot of advantages as sensor materials over other materials in the majority of sensor technologies. The sensing ability of many synthetic polymers in various artificial sensor devices has been established by their excellent tailorability and easy processability to form very-thin-layer devices. Polymers have high tailorability of their molecular structure and composition for better processability or to improve specific behavior in the bulk material or on its surface. As a result, a broad spectrum of properties can be easily obtained with polymers. In addition, the ease of creation of new functional groups

POLYMERS IN CHEMICAL SENSORS 5

on the polymer backbone, their ability to respond to redox systems, their charge-carrying capability, and their ability to respond to optical stimulation are other advantages of using them in sensor devices. Due to the flexible nature of the polymer chains, compact and miniaturized sensor arrays can be easily fabricated. Unlike metal and metal-oxide semiconductor materials, there is no need for special clean room, high temperature, or special high-cost process techniques to fabricate sensor devices using polymers. Good selective sensing of a specific analyte in a mixture is also possible with a polymer because of its high structural tailorability. Polymers have operational/functional advantages at ambient temperature. The suitability of a polymer in a sensor device for use at a particular temperature may be suggested by the glass transition temperature (Tg) of the polymer. A polymer in a sensor device may provide proper function if the temperature of the sensing measurement is kept close to its Tg but well below its melting temperature (Tm). As a basic principle, a polymer in a sensor device can function by absorption/adsorption of an analyte in the form of gas or liquid, reversibly or irreversibly, followed by interaction and change in properties of the polymer, which is transduced electronically, electrochemically, or optically. Processing of the transduced signal provides the sensor output. Large numbers of articles and reviews have been published on chemical sensors, in which many polymers have been utilized to fabricate sensor devices (Armstrong and Horvai 1990; Bidan 1992; Adhikari and Majumdar 2004; Persaud 2005). Although few such polymer-based sensor technologies have been commercialized as yet, many are in the exploration stage, and understanding the structure and properties of suitable polymers for use in sensor devices may lead to more advanced devices for chemical sensing. Polymers have unique characteristics that have been proved to influence the operating parameters of sensor devices. Properties of polymers that influence the operating parameters of sensors can be physicochemical, electrical (conductivity, resistivity), chemical, optical (photo- and electroluminescence, optoelectronic), redox, hydrophobic/hydrophilic, piezoelectric/pyroelectric, etc. Being organic in nature, polymers provide an inherent affinity to chemical species that need to be detected or estimated. Comparing the solubility parameters of both the polymer as sensing element and the analyte to be sensed can lead to correlation of this affinity. Because of its long backbone chain and flexible nature, a polymer can accommodate a large quantity of a foreign substance in the form of fine particle (as filler) or fiber or even highly viscous liquid. Thus, a polymer in its solid state contains ample free volume in the bulk. Depending on its affinity toward a foreign agent, the polymer can hold it for quite a long period. This can happen in two ways: by simple physical entrapment provided the foreign agent is compatible with the host polymer; or by chemical bonding between them. The latter is well known to provide covalent immobilization of the foreign agent as the recognition element. A polymer to be used as either a solid carrier or directly as the sensing element can be best selected based on the solubility parameters of both the polymer and the analyte chemical compound. Another important parameter to be considered in selecting a polymer for sensor application is diffusion. Better sensing response in terms of sensitivity and sensor recovery can be obtained from a polymerbased sensor if the polymer offers a lower diffusion barrier to the analyte compound. This way, short response time as well as short recovery time with good sensor repeatability can be available from a particular polymeranalyte system. Exploring the nature of interactions between polymer and analyte can also help in a major way to obtain a good sensing device. Such interactions can be judged by looking at the chemical structure of the polymer chain and its pendant groups vis--vis the chemical structure of the

6 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

incoming analyte compound. Adsorption/desorption, stability, morphological features (crystalline/amorphous), surface area, and the population of the active sensing site also influence sensing characteristics. The stability of the polymers in the sensor device is another issue to be considered before selecting a polymer (Durst et al. 1997). In polymers, some percentages of heteroatoms such as nitrogen, oxygen, sulfur, phosphorous, and halogens are present together with a long chain of carbon and hydrogen. The carboncarbon or carbonhydrogen bonds are comparatively more stable than the bonds between carbon and the heteroatoms. A polymer that has very good electrical conductivity may not have good stability in the ambient environment, or its stability may decrease when it is in contact with the analyte compound. This occurs because the unsaturated bonds in conducting polymers are often very reactive when exposed to environmental agents such as oxygen or moisture. Apart from the inherent stability of the polymers, the stability of some foreign chemical compounds, which are used as dopants in conducting polymers, are also important. These dopants may not be very stable within the polymer matrix and may also react with environmental agents. The polymer selected should maintain its intrinsic sensing characteristics across a wide temperature range. It must possess adequate mechanical strength to sustain handling and other stresses. Stability and degradation of the polymer is also important when it may be exposed to chemical environments during sensing. The polymer should be resistant to degradation or dissolution in organic solvents but should interact with the reactive sensing element. The permeability of the polymer film used as the detecting layer is also important, because this property affects the transport properties of the other components. Therefore, criteria for polymers to be used in chemical sensors need attention, and some information about the stability and degradation of such polymers is essential. Experimental and theoretical work on polymeric and supramolecular compounds helps in designing highly selective chemical sensors. Different transducer principles are used for sensing molecules in air and water by monitoring changes in mass, temperature, capacitance, and thickness. Such changes in physical and chemical properties are monitored using resonator, calorimetric, impedance, or fiber optic sensors. Supramolecular chemistry aims at the preparation of molecules with specific binding sites inside their cavities. Self-organized layers with a well-defined architecture make it possible to design highly specific chemical sensors utilizing very fast adsorption processes between recognition sites at the surface of monolayers and molecules to be detected (Schierbaum 1992). Permselective polymeric membranes, used to separate different gaseous and liquid constituents, can also be used in analytical chemistry and the field of sensors. Practical examples include membranes in gas sensors that have improved selectivity to certain gas constituents, membranes in ion-selective electrodes and ion-sensitive field-effect transistors (ISFETs), and diffusion membranes in amperometric electrochemical cells. Polyethylene, polytetrafluoroethylene (PTFE), polyfluoroethylene-propylene (PFEP), cellulose acetate, silicone rubber, plasticized polyvinyl chloride (PVC), polydimethylsiloxane (PDMS), and other polymers are used as membranes. Development of miniaturized sensor arrays using nanostructured polymers and composites has come to the forefront of chemical sensor research in order to obtain specific selectivity of one particular chemical analyte in a mixture. By virtue of their inherently flexible nature, the backbones of polymer chains can undergo conformational change to provide reversible adsorption/desorption in sensing response, or they can undergo oxidation/reduction when exposed to an analyte with an attendant change in electrical conductivity as a sensing signal. The major functions of polymers in sensor devices are to serve as a solid support for a chemical recognition element, as a selective agent for a specific analyte, or as the recognition element itself.

POLYMERS IN CHEMICAL SENSORS 7

4. SYNTHESIS OF POLYMERS
Using some processing and fabrication techniques, pure polymers of adequate molecular weight and properties can be employed in sensor devices. In many cases, however, sensor device fabrication is difficult because many solubility and processing techniques are not compatible with all polymers. These polymers either possess in-built sensing sites or are modified to attach sensing functionalities. This indicates that some polymers have inherent properties that offer sensing function, whereas other polymers are used as carriers of sensing elements. So, depending on the need for a specific sensing function, a polymer should be selected before attempting its synthesis. Most carrier polymers are electrical insulators in which sensing elements are either physically immobilized or chemically attached. In these cases the polymer of choice should be procured from market or synthesized as per the requirement. Polymers used in sensor devices are either formed in situ or processed from a fully grown polymer prepared by conventional techniques. Polymers are synthesized from their respective monomers either by condensation of reactive functional groups in the monomer or by addition chain reaction through olefinic unsaturation in the monomer followed by isolation and purification. Flory (1953) described briefly these two types of polymerization, chain-growth polymerization and step-growth polymerization, by which most of the polymer molecule may be built up. The procedures generally followed are bulk, solution, suspension, emulsion, or interfacial polymerization (Flory 1953; Odian 2004). In chain-growth polymerization, an initiator reacts with a monomer molecule to create a reactive site, and the reactive site then reacts with successive monomer molecules to yield the polymer. A homopolymer that forms via chain-growth polymerization usually forms from one monomer; copolymers result from chain-growth polymerization of two or more monomers with the same type of reactive functional site (olefinic unsaturation). This type of polymerization is popular for the monomers that have double bonds, which can act as the reactive functional site during polymerization. The formation of polyethylene from ethylene in the presence of an initiator is an example of chain-growth polymerization. Other examples are styrene, butadiene, propylene, vinylene monomer, and acryl monomer. Initiators that are commonly used for polymerization include peroxide (ROOR), azo compounds (RN=NR), and redox compounds [FeSO4, FeCl3, K2S2O8, (NH4)2S2O8, etc.] by thermal or photochemical pathways (Odian 2004). Step-growth polymerization begins when one monomer with two reactive functional groups reacts with another monomer containing functional groups of another type such that a small by-product molecule leaves the chain. Polymerization usually proceeds by reactions between two different reactive functional groups, e.g., hydroxyl and carboxyl groups, isocyanate and hydroxyl groups, amine and acid groups, etc. So, according to the pair of functional groups in the monomers, a number of different chemical reactions may be used to synthesize polymeric materials by step polymerization, e.g., esterification, amidation, the formation of urethanes, aromatic substitution, etc. (Odian 2004). The synthesis of condensation polymers by ring-opening polymerization is also possible. For the general principle behind the synthesis of polymers, readers are referred to standard textbooks (Flory 1953; Odian 2004). Intrinsically conducting polymers are widely used in various electrochemical sensing devices. Presently used conducting polymers for these applications are mostly polyheterocycles such as polypyrrole, polythiophene, polyfuran, polyisothionapthalene, polyindole, polyaniline, polycarbazole, etc.,

8 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

and polyaromatics such as polyazulene, poly-p-phenylene (PPP), poly p-phenylene vinylene (PPV), polypyrene, etc. (see Figure 1.2) (Gurunathan et al. 1999). Those conducting polymers are usually synthesized by one of two popular methods, chemical or electrochemical oxidation of the corresponding monomers. Thiophene, furan, carbazole, aniline, indole, azulene, and their derivatives are the major monomers used for synthesis of conducting polymers. Synthesis of conducting polymers by oxidative coupling polymerization is a very easy and simple method and, therefore, this method is suitable for producing bulk quantities of polymer in a conducting oxidized state with associated counterions from the polymerization medium. The procedure involves simple mixing of monomer and oxidant in aqueous or organic protonic acid solution. Commonly used oxidants are ammonium persulfate, ferric chloride, hydrogen peroxide, potassium dichromate, cerium sulfate, etc. The majority of the redox polymers are synthesized by chemical polymerization. The oxidative polymerization of inexpensive, simple aromatic benzenoid or nonbenzenoid (mostly amines, e.g., aniline, o-phenylenediamine), and heterocyclic compounds (e.g., pyrroles, thiophenes, indoles, azines, etc.) is of greatest interest (MacDiarmid and Epstein 1989; Syed and Dinesan 1991; Martin et al. 1993). Polythiophene and its derivatives are synthesized via oxidative coupling reactions (McCullough 1998). Quite a large number of published reports are available in the literature on the chemical synthesis by oxidative coupling reaction of aniline, pyrrole, thiophene, etc. A comprehensive picture may be obtained from some review reports (Toshima and Hara 1995; Feast et al. 1996; Smith 1998). Different electrochemical principles are followed to synthesize intrinsically conducting polymers, viz., galvanostatic, potentiostatic, cyclic voltammetry, and other potentiodynamic methods (Toshima and Hara 1995; Smith 1998; Feast et al. 1996). These techniques utilize a three-electrode system: a working electrode, a counter electrode, and a reference electrode. During electrochemical synthesis, the conducting polymers are electrochemically deposited on the working electrode, which is made of materials such as platinum, stainless steel, gold, indium tin oxide (ITO), or glass. The polymer-deposited electrodes are either used directly or the polymers deposited on the electrode surface are peeled off as self-standing films for a particular application. In electrochemical polymerization, the electrochemically active groups are either built into the polymer structure or are added as a pendant group. These groups can also be incorporated into the polymer during polymerization, or attached to the polymer network in an additional step after the coating procedure (postcoating functionalization) to obtain polymer film electrodes (Bruce 1995; Inzelt 2000). Sato et al. (1986) showed that the electrochemical polymerization of long-chain alkyl-substituted thiophene and pyrrole yields highly conducting films, some of which are soluble in common organic solvents in their conducting state. Although chemical oxidation of pyrrole by Fe(ClO4)3 leads to conducting polypyrrole (MacDiarmid and Epstein 1989; Syed and Dinesan 1991; Martin et al. 1993), electrochemical polymerization is a preferred technique for polymer film electrodes, thin-layer sensors, in microtechnology, etc., because the control of potential is a prerequisite for polymer film deposition on the anode during synthesis. The oxidation state of the polymer can be varied electrochemically by cycling the potential between oxidized, conducting, and the neutral, insulating state, or by using suitable redox compounds. Varying the composition of the polymerization medium leads to a change in the conductivity of the polymer. For example, increasing the pH of the polymerization medium (MacDiarmid and Epstein 1989; Paul et al. 1985) or including an electron-donor molecule (e.g., NH3) in the gas phase decreases the conductivity of polyaniline (PANI) or polypyrrole (PPy) films (Miasik et al. 1986; Pei and Inganas 1993).

POLYMERS IN CHEMICAL SENSORS 9

Figure 1.3. Schematic plasma polymerization experimental setup. A gaseous monomer or monomer vapor such as acetylene or aniline and a gaseous dopant such as iodine, chlorine, or HCl gas are used within the chamber. (Reproduced with permission from Saravanan et al. 2004. Copyright 2004 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft.)

Many conducting polymers and their derivatives are usually synthesized through well-known chemical routes rather than by oxidative polymerization. These chemical routes include various well-known reactions, such as the Wittig reaction (Diaz et al. 1979), the Heck reaction (Kobayashi et al. 1985), and Gilch polymerization (Malhotra et al. 1986). Plasma polymerization or electropolymerization of monomers are preferred when the polymers are difficult to process because of insolubility or infusibility. Plasma-polymerized thin films from various monomer precursors have been prepared by employing radio-frequency plasma polymerization techniques (Saravanan et al. 2004). The apparatus consisted of a 50-cm glass tube of 8 cm diameter, with provision for charging monomer vapors by evacuation. A schematic of the experimental setup is shown in Figure 1.3. Chemically and ultrasonically cleaned glass substrates were placed inside the glass tube exactly under the space separated by the aluminum foil electrodes, which were capacitively coupled and wrapped around the glass tube, separated by a distance of 5 cm.

5. DEPOSITION OF POLYMERS
The deposition of polymer as a sensing element on an electrode surface is a very tricky process, since the sensor sensitivity depends on the thickness, chemical composition, crystallinity, conductivity of the coated polymer, etc. The present state of the art of polymer-coated electrode preparation is based on extensive research and development, and numerous research articles are available in various sensor-related journals and books. In a chemical sensor, a properly functionalized polymer acts as a chemoreceptor, which is a selective receiving site for analyte recognition and reaction. In the case of a biologically derived receptor, the more specific term biochemical receptor or bioreceptor may be used. The bioreceptor might be an enzyme, tissue organelles, antigen/antibodies, etc. These biorecognition elements might be immobilized in a polymer by covalent attachment, physical entrapment, or cross-linking. Electrodes are usually fabricated by chemical modification or by deposition of polymer on the electrode surface by one of the following techniques.

10 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

1. The chemisorption-adsorption technique utilizes valence forces of the same kind as those operating in the formation of chemical compounds, where the chemical polymeric film is strongly and, ideally, irreversibly adsorbed (chemisorbed) onto the electrode surface to provide a monolayer film (Gold et al. 1987). This modification creates substrate-coupled self-assembled monolayers (SAMs) in which uncorrelated molecules spontaneously chemisorb at specific sites on the surface of the electrode to form a superlattice (Allara 1995). 2. Covalent attachment of one to several monomolecular layers of the chemical modifier to the electrode surface involves some combination of chemisorption and low solubility in the contacting solution, or physical anchoring on a porous electrode. 3. Formation of a polymer film using a chemical modifier is also done. The polymer film can be organic, organometallic, or inorganic as long as it contains the desired chemical modifier. Other forms of possible modification are substrate-decoupled SAMs, in which adsorbate molecules are arranged on the electrode surface independent of any substrate structure (Allara 1995). In order to get a polymer-coated ion-selective electrode, two conditions must be met. First, the cations must be freely mobile within the polymer matrix, which can be achieved by saturating the polymer with a solvent. The solvent helps to dissociate the cations associated with the pendant acidic groups and allows them to move within the polymer. Second, the membrane surface must be made conductive by depositing a thin metal electrode on both surfaces of the membrane. Typically, the ion-exchange property of the polymer is exploited to facilitate metal deposition. The polymer surface needs to be pretreated by sand blasting, hydrating, and cleaning with a strong acid, thus ensuring that the polymer is fully saturated with protons. The membrane is then placed in an aqueous solution containing ions of the metal to be plated. These ions are allowed to exchange with the protons in the polymer for a predetermined amount of time and are then reduced to their neutral state at the surface of the polymer by a reducing agent (typically NaBH4 or LiBH4) in the outer solution. In this solvated and electroded form, a Nafion membrane can be made to bend toward the anode side when a small voltage (15 V) is applied across its thickness, thus making it a soft, distributed actuator. Membranes in this form can also be used as distributed sensors. Several researchers have shown that the transient voltage generated across the membrane is correlated to the quasi-static displacement of the membrane (Sadeghipour et al. 1992; Shahinpoor et al. 1998; Keshavarzi et al. 1999). The active polymer layer may be in free-standing film form or on a suitable substrate and forms the heart of the sensor. Polymer film or polymer coated on a substrate can be obtained using the following methods. 1. Dip coating. This process consists of immersing the electrode material in a solution of the polymer in a suitable solvent for a sufficient period to allow spontaneous film formation onto the substrate by adsorption. The film thickness may be controlled by adjusting the polymer solution viscosity and the speed of withdrawing the electrode from the solution, followed by solvent evaporation to form the polymer film on the electrode. 2. Solvent evaporation. Applying a drop of polymer solution of the required consistency on to the electrode surface, followed by solvent evaporation, creates a film. The quality and thickness of a polymer film formed by this manual technique depends on the personal skill of the research

POLYMERS IN CHEMICAL SENSORS 11

3.

4.

5.

6.

worker, but the method is advantageous because the thickness of the polymer-coated film on the electrode can be known from the original concentration of polymer solution and droplet volume. This is the oldest and still popular method for free-standing polymer film casting. Spin coating. In spin coating or spin casting, a dilute polymer solution droplet is placed on the surface of a rotating electrode. Because of the rotation, excess solution is spun off the surface and a very thin polymer film is formed. Following the same procedure, multiple layers can be formed until the desired thickness is obtained. This procedure provides pinhole-free thin films. Layer-by-layer (LBL) self-assembly. A composite of two polymeric electrolytes on some suitable substrate can be fabricated (Ram et al. 2005a; Nohria et al. 2006). In this method the substrate is alternatively immersed into a polymeric anion solution and a polymeric cation solution. Insoluble doped conducting polymers, e.g., polyaniline with positive charge on the backbone, can be deposited on a polymeric anion layer. The thickness of the LBL film depends on the number of times the process is repeated (Nohria et al. 2006). Langmuir-Blodgett film casting. In Langmuir-Blodgett (LB) film casting, polymer molecules with hydrophilic heads and hydrophobic tails are spread on a water surface and then the molecules are compressed using a barrier to align the molecules (see Figure 1.4). Here, the polymer itself contains hydrophobic and hydrophilic groups within the backbone, or monomers having such groups are polymerized, which causes the polymer to orient itself during film casting. A singlelayer film is cast on a substrate which is drawn up from beneath the water surface. Generally, LB films form by balancing the interactions at the polymerwater, airwater, and polymerair interfaces (Wegner and Remmers 1995). The resulting film is very well ordered, single-layered, and in the range of molecular thickness. Electrochemical polymerization. A monomer solution is oxidized or reduced to an activated form that leads to a polymer film formed directly on the electrode surface. This procedure results in few pinholes, since polymerization would be accentuated at exposed (pinhole) sites at the electrode surface. Unless the polymer film itself is redox-active, electrode passivation occurs and further film growth is prevented. Presently, polymers are being used in layer or film form as an active part of inorganic solid-state sensing devices. Polymers can be easily deposited on various substrates by simple techniques. Easy

Polymer layer

Substrate (hydrophobic) Barrier

Polymer layer in substrate

Water
Figure 1.4. LB lm deposition mechanism.

12 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

film casting is also possible for fabrication of different sensor systems. In the case of electrode preparation by direct electrochemical deposition of the conducting polymer, knowledge of the kinetics of the electrodeposition process is also of utmost importance in order to obtain proper sensing function of the electrode. Along with the above-mentioned influencing parameters on polymer growth during electrodeposition, the effect of the electrode material and its surface properties need attention. For example, in the autocatalytic oxidation of aniline over a platinum electrode, specific interactions and wetting may determine the nucleation and dimensionality of the growth process. Two or more stages of the polymerization process have been distinguished in the case of PANI: a compact layer (200 nm) formed on the electrode surface via a potential-independent nucleation and a twodimensional (2-D, lateral) growth of PANI islands followed by 1-D growth of the polymer chain with continuous branching in the advanced stage leading to an open structure (Bade et al. 1992; Cruz and Ticianelli 1997). It was established that the formation of the aniline-based polymer involves two electrons in the formation of one monomermonomer bond (Linford 1987; Lyons 1994, 1996). The growth rate is proportional, except for the early induction period, to the 0.5 power with respect to the film volume, and it is first-order with respect to aniline concentration (Stilwell and Park 1988). To avoid hydrolytic degradation of the oxidized PANI (pernigraniline form), the positive potential limit of cycling can be decreased after 210 cycles because of the autocatalytic nature of the electropolymerization (Horanyi and Inzelt 1989; Stilwell and Park 1989). In addition to the head-to-tail coupling, formation of p-aminodiphenylamine by tailto-tail dimerization (benzidine) also occurs as a minor intermediate, as evidenced by the rate constant of dimerization for radicalcation coupling to produce the former product, which is about 2.5 times higher than that for the tail-to-tail dimer (Yang and Bard 1992). Generally, a mixed material is deposited on the surface, containing electrochemically active and conducting as well as inactive and insulating parts, if the polymerization conditions are not carefully optimized (Otero and Rodriguez 1994). Since polythiophene does not adhere on a Au or Ti surface, it prefers electrochemical formation and precipitation from the medium. Therefore, a suitable approach is the deposition of a thin polypyrrole layer on Ti or Au that ensures the deposition of polythiophene on these substrates (e.g. Ti, Au) (Gratzl et al. 1990). 7. Deposition by radio-frequency polymerization of a suitable monomer. In this method, a monomer vapor is exposed to a radio-frequency (RF) plasma discharge to form a thin polymer film on the electrode surface. Chemical damage of the polymer film producing unknown functionalities and structural modifications due to high energetics of the RF discharge is a disadvantage of this technique. 8. Polymer deposition followed by cross-linking. In this technique, chemical components of a polymer film are bonded with the electrode to impart stability, decreased permeability, or altered electrontransport characteristics to the polymer film. Cross-linked films are often formed by polymerization of bifunctional and polyfunctional monomers or by chemical, electrochemical, photolytic, radiolytic, or thermal activation. 9. Pellet preparation. In this technique the well-dried polymer powder is made into a thin pellet of particular thickness using a steel die in a hydraulic press under a certain constant pressure. The pellet preparation technique is very useful for polymers which are not soluble in organic solvents, such as conducting polymersespecially, doped forms of conducting polymers.

POLYMERS IN CHEMICAL SENSORS 13

6. FUNCTIONALIZATION OF POLYMERS 6.1. STRUCTURE MODIFICATION


The polymer film may also be functionalized subsequent to film application, because immobilizing the film is easier than working with monolayers, such modified films are usually more stable, and stronger electrochemical responses can be available from multiple layers of redox sites. Some questions remain, however, as to how the electrochemical reactions of multimolecular layers of electroactive sites in a polymer matrix occur, for example, the mass-transport and electron-transfer processes by which the multilayers exchange electrons with the electrode and with reactive species such as molecules and ions in the contacting solution (Murray 1984; Murray et al. 1987; Inzelt 1994). Lack of sufficient knowledge about the structure and properties of polymer films, as well as morphological changes arising out of various chemical, electrochemical, and physical processes during use, may lead to uncertainty in the ultimate sensing performance. Electrocatalysis on a modified electrode is usually an electron-transfer reaction between the electrode and some solution substrate, which, when mediated by an immobilized redox couple, proceeds at a lower overpotential than would otherwise occur at the bare electrode (Durst et al. 1997). After about four decades of research on conducting polymers and their utilization in sensor devices, there is sufficient scope and many avenues for enhancing the conductivity as well as improving sensing ability. From the literature on polymer-based sensor research, it has become apparent that efficiency in sensor output depends on parameters such as polymer structure, composition, morphology, and active functional sites. Therefore, to achieve a specific sensor device, it is usually necessary to modify the structure of the conducting polymer and perhaps substitute for one of the common conducting polymers. Not only the sensor output but also the stability of the polymer in the device needs special attention. Polymers having excellent sensing characteristics have been found to be environmentally unstable. Processing and fabrication of the synthesized conducting polymers into a suitable device is a challenging problem today. Many researchers have tried to solve this problem by chemical substitution in the monomer stage or by changing the polymerization conditions. Although both electrically insulating as well as conducting polymers have been successfully utilized in the fabrication of chemical sensor devices, conducting polymers have clear advantages in sensing performance over insulating polymers. By virtue of tailor-made characteristics, structural as well as functional modification of conventional conducting polymers can add a new dimension to polymers as materials for chemical sensors. Some of the common modification strategies being practiced for conducting polymers intended for use in sensor devices are described in the following paragraphs, together with some research outcomes. 1. Chemical group substitution on monomers and polymers. Appropriate substitutions in the monomer can improve the stability in air of the electrochemically produced polymers. Although prepared under identical electrochemical conditions, the substitution of a methyl group, e.g., poly(3methyl thiophene) has shown better air stability than that of polythiophene (Tourillon and Garnies 1982; Bryce et al. 1987). In addition to its increased stability in air, poly(3-methyl thiophene) has also shown improved conductivity over that of the parent polythiophene (Waltman

14 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

et al. 1983). Unfortunately, such an increase in electrical conductivity has not been observed in a pyrrole system after alkyl-group substitution. The methyl-substituted polypyrroles had lower conductivities than the parent polypyrrole (Diaz et al. 1982). It appears that a balance between electronic and steric effects in substituted polymers of five-membered heterocycles might account for either more or less conductivity than in the parent polymers (Waltman 1984; Waltman and Diaz 1985). Several polyaniline derivatives have been developed by polymerizing various alkyl- and alkoxysubstituted polyanilines with the objective of improving conductivity as well as achieving better-processable conducting polyanilines. Common derivatives have included ortho- and metasubstituted anilines with NH2 (Gouette and Leclerc 1995; Levin et al. 2005), OH (Rivas et al. 2002; Salavagione et al. 2004; Levin et al. 2005; Kar et al. 2008) OCH3 (Dao et al. 1989; Park et al. 1989, Cataldo and Maltese 2002), CH3 (Leclerc et al. 1989; Falcou et al. 2005), SO3H (Yue and Epstein 1990; Kuo et al. 1998), Cl, F, I (Kang et al. 1990), CH2CH3 (Dao et al. 1989; Leclerc et al. 1989; Yue and Epstein 1990), etc., giving rise to similar products. A few of these derivatives had improved processability, but very little or no increase in electrical conductivity was achieved. 2. Copolymerization. Electrochemically synthesized conducting polymers suffer from the drawback of poor mechanical strength, which restricts their use in commercial products. Since copolymerization is one of the most effective methods for improving the mechanical strength of brittle polymers, this approach may be followed for conducting polymers without compromising their conductivity. In this respect, copolymerization of polyheterocycles with poly(p-phenylene sulfide) (PPS) can lead to better mechanical strength, less pronounced O2 sensitivity, and increased conductivity of the copolymer by doping of PPS. However, copolymerization by the electrochemical technique is a challenging task due to the different electrochemical oxidation potentials of individual monomers. Graft copolymerization and blending are other tailoring approaches to improve atmospheric stability and mechanical strength of conducting polyheterocycles.

6.2. SURFACE MODIFICATION


A wide variety of modifications to the surface morphology of polymers is possible for suiting specific sensor architecture, e.g., lithographic and soft lithographic techniques, replication from masters, pattern formation using self-assembly and controlled deposition, nanomachining, emulsion templating, and wetting-assisted templating (Xia et al. 1999; Hamley 2003; Helt et al. 2004; Drain and Batteas 2004; Barbetta et al. 2005). Surface modification is also possible using plasma polymerization techniques. For example, one existing plasma polymerization setup (see Figure 1.3) was slightly modified for the surface modification of inorganic materials by organic coating (Sunny et al. 2006). The controlled flow of monomer vapor in the radio-frequency plasma polymerization chamber creates a thin deposition on the existing ceramic or semiconductor surface. The monomer is plasma-polymerized inside the chamber due to the RF source and is deposited on substrate surfaces placed inside the bottom of the discharge tube. In this method, conducting polymers such as polyaniline, polypyrrole, and polythiophene can be deposited on the surface of conventional semiconducting materials.

POLYMERS IN CHEMICAL SENSORS 15

Coating of polymers is also possible by thermal spraying of polymer powders onto a wide variety of materials. In this method, polymer powder is injected into a heat source (hot flame or plasma) and transported to a preheated substrate. This is an effective method of producing protective barrier coatings. Polymers that have been sprayed include polyethylene, polymethyl methacrylate, polyether ether ketone, polyphenylene sulfide, nylon, phenolic, epoxy, Tefzel (modified ethylene-tetrafluoroethylene polymer), and postconsumer commingled polymers. The thickness of the coating depends on the number of repeated passes of the spray gun across the substrate. However, polymers with large particle size or higher molecular weight may form more heterogeneous microstructures within the coating, creating voids, trapped gasses, unmelted particles, splats, and pyrolized material. Direct surface modification of polymer substrates such as the polyimide Nafion, a sulfonated ionic fluoropolymer, which is commonly used in polymer sensor-actuator devices, can be done by the plasma technique. Here the polymer surface is treated with a suitable gas plasma to obtain reactive functional groups on the surface of the treated polymer. For example, Kapton, a polyimide, was treated in argon plasma for 2 min at 35 W and 0.2 torr, and Nafion was treated in oxygen plasma for 10 min at 50 W power and 0.2 torr (Supriya and Claus 2004). The plasma treatment activates the polymer surface to facilitate the deposition of other polymers having active functionality, or the polymer can be used directly as the detecting layer for a chemical sensing device. Other important surface modification methods include grafting, surface coupling reactions, electron beam, glow discharge, corona discharge, UV treatment, etc. (Uyama et al. 1998).

6.3. COMPOSITION MODIFICATION


It has been established that -electron conjugation along the backbone of a polymer chain is one of the criteria for a polymer to exhibit good electrical (conducting and semiconducting) behavior. Additionally, it is also known that the presence of heteroatoms in polymers can lead to improved electrical (conducting/semiconducting) performance. We have already mentioned that, for chemical sensor detection by the electrochemical principle, it is important to have inherent electrical properties which can alter during analyte interaction with the polymer detection element. As is known, the electrical properties depend on the mobility of free or loosely bounded electrons. For polymers to be used as detection elements, the inherent electrical conduction may be obtained in two ways: by metal or metal compound dispersion within the polymer matrix; or by -electron conjugation along the polymer chain. The free electrons or charges on the metal are responsible for the electrical conductivity of the metal or metal compound dispersed polymer detection system. In this system the polymer does not play a direct role in the alteration of electrical properties. However, the polymer can increase the performance of metal or metal compounds during sensing by chemical interaction of some particular groups. On the other hand, a particular group of polymers may aid interaction with the detection element. Since the discovery of intrinsically conducting polymers (ICPs), which show inherent electrical properties through the movement of a loosely bounded -electron cloud, these have been used successfully as chemical detection elements in electrochemical transducer systems. Metal or metal compounds can also be dispersed within the ICP matrix to increase the sensing performance of the polymer. One of the easiest routes to composition modification is composite formation.

16 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Gurunathan and Trivedi (2000) incorporated TiO2 in polyaniline by chemical and electrochemical techniques to study the effect of photoconducting inorganic semiconductors on thermal stability and application of a new composite. The structure and conductivity of the polymer can also be altered by further chemical reactions. Active sensing polymers are used in sheet or film form as an integral part of inorganic solid-state devices. Conducting polymer composite sheets/films contain conductive filler loaded in an electrically insulating polymer matrix. The change in resistivity as a function of filler concentration in such composites can be understood according to the percolation concept, which describes conduction through electrically conducting paths between two filler particles. The number of these paths dramatically decreases below a critical volume fraction of filler (Lundberg and Sundqvist 1986). Common filler materials are metals (Cu, Pd, Au, Pt), carbon black, and semiconducting metal oxides (V2O3, TiO2, etc.). The insulating matrix polymers have included polyethylene, polyimides, polyesters, poly(vinyl acetate) (PVAc), PTFE, polyurethane, poly(vinyl alcohol) (PVA), epoxies, acrylics, poly(methyl methacrylate) (PMMA), and others. These composites have been used successfully in positive temperature coefficient (PTC) thermistors and in piezoresistive pressure, tactile, humidity, and gas sensors (Lundberg and Sundqvist 1986). The ionic conductivity of polyelectrolytes is modulated by several environmental parameters, which is the basis of their application in sensors. The conductivity of polyelectrolyte films can be increased by increasing the number of ionic carriers through addition of ions from the environment and modifying the degree of dissociation of the polymer electrolyte (Sadaoka et al. 1986; Sakai et al. 1989). Ionconducting polymers, viz., Nafion, polyhydroxyethyl methacrylate (polyHEMA) and its copolymers, are widely used as solid electrolytes in electrochemical cells for the detection of various gases or ionic components. Alkali saltpolyether complexes, such as polypropylene oxide (PPO) and polyethylene oxide (PEO) with LiClO4, LiCl, LiCF3SO3, LiSCN (Watanabe et al. 1985; Sadaoka et al. 1986), polystyrene sulphonate, and quaternized polyvinyl pyridine (PVPy), were used in impedance-type or semiconductor-based humidity sensors. The structures of -electron conjugated conducting polymers, such as polyacetylene, polyaromatics, and polyheterocycles contain a one-dimensional organic backbone, which enables formation of a superorbital for electronic conduction. Conduction through these polymers takes place by charge hopping both along the polymer chains and also between the macromolecules that make up individual fibers as well as between the fibers themselves. However, in the neutral (undoped) state, these materials behave like semiconductors. The electronic conductivity in these materials is obtained by doping, i.e., by injecting electrons or holes into the superorbital (Bidan 1992).

7. POLYMERS IN CHEMICAL SENSORS


As we have indicated, many types of chemical sensors have been designed for different analytical tasks, such as detection of hazardous gas concentration, control of groundwater pollution, medical analysis and diagnostics, industrial quality and process controls, environmental safety, and so on. Some toxic chemicals and their common sources are listed in Table 1.1. All these various sensor types work by different principles and require sensing materials with different properties. However, polymers, due to their unique physical and chemical properties, can be used in all types of chemical sensors (see Table 1.2).

Table 1.1. Sources of toxic and hazardous chemicals in the environment


COMMON SOURCES Petrochemical industries and motor fuels Solvents and cleaning agents used in chemical industries

CATEGORY

TOXIC CHEMICALS

Toxic industrial chemicals

Methane, ethylene, benzene, toluene, xylene, formaldehyde, carbon monoxide, carbon dioxide, oxides of sulfur and nitrogen, heavy metals and their oxides, etc.

Methyl ethyl ketone, acetone, methanol, ethanol, chloroform, carbon tetrachloride, tetrahydrofuran, isopropanol, methyl chloride, trichloroethylene, benzene, toluene, xylene, ammonia, inorganic acids, hydrogen sulfide, oxides of sulfur and nitrogen, isocyanides, halogens, etc.

Freon-22, chlorofluorocarbons, methyl chloride, methanol, etc.

Refrigerants and cooling systems Polymer, rubber, and textile industries

Vinyl chloride, methyl ethyl ketone, acetone, methanol, ethanol, chloroform, carbon tetrachloride, tetrahydrofuran, isopropanol, methyl chloride, trichloroethylene, benzene, toluene, xylene, inorganic acids, etc.

Toxic household chemicals

Methane, kerosene, phenol, cresol, inorganic acids, sodium hydroxide, bleach, isopropanol, 2-butoxyethanol, naphthalene, etc.

House cleaning Kitchens Medicinal and human habits Dissolved gases in water Rainwater Groundwater Marine pollution

Methane, ethylene, benzene, toluene, xylene, formaldehyde, carbon monoxide, carbon dioxide, oxides of sulfur and nitrogen, etc.

Morpholine, methanol, ethanol, carbon dioxide, carbon monoxide, etc.

Water-polluting chemicals

Carbon monoxide, carbon dioxide, oxides of sulfur and nitrogen, etc.

Ammonia, inorganic acids, hydrogen sulfide, oxides of sulfur and nitrogen, isocyanides, halogens, etc.

Ammonia, inorganic acids, hydrogen sulfide, oxides of sulfur and nitrogen, isocyanides, halogens, arsenides, toxic metals, etc.

Methane, ethylene, benzene, toluene, xylene, formaldehyde, carbon monoxide, carbon dioxide, oxides of sulfur and nitrogen, petroleum products, heavy metals and their oxides, etc.

Chemical weapons

Chlorine, chloropicrin, hydrogen cyanide, arsines, psychotomimetic agents, toxins, carbon monoxide, carbon dioxide, phosgene, mustard gas, tear gas, lewisite, G-series nerve agents, V-series nerve agents, etc.

Chemical gas weapons

Mercury fulminate, lead styphnae, lead azide, dynamite, TNT, RDX, PETN, HMX, ammonium nitrate, etc.

Explosives

Table 1.2. Polymers used in gas sensors


USES Detection of terpene in atmosphere Identification of gases and gas mixtures Polymercarbon black composite film Zee and Judy 2001 Piezoelectric sensor coated with molecular imprinted polymer Percival et al. 2001 SPECIAL FEATURES REFERENCES

POLYMER

Copolymers of poly (EDMA-co-MAA)

Polyethylmethacrylate, chlorinated polyisoprene, polypropylene (isotactic, chlorinated), styrene/ butadiene, ABA block copolymer, styrene/ ethylene/butylene ABA block copolymer, polyepichlorohydrin Detection of ethanol gas concentration Detection of NO2 Layers of polymer films formed by Langmuir-Blodgett (LB) and self-assembly techniques Films of polymer prepared via LB deposition and casting Solid polymer electrode of 10% PVC Nanocomposite of iron oxide polypyrrole prepared by simultaneous gelation and polymerization Polymer filmcoated quartz resonator balance Amperometric sensor Electrical property measurement Electrical property measurement Electrical property measurement Fuel cell with polymer electrolyte membrane Kim et al. 2000 Xie et al. 2002

Nafion

Polyaniline (PANI), PANIacetic acid mixed film, PANIpolystyrenesulfonic acid (PSSA) composite film Gas sensor Detection of NO2 in air Detection of CO2, N2, CH4 gases at varying pressures Detection of toluene, xylene O2 and NO NH3 NH3 NH3 and NO2 NH3 NH3

Poly[3-(butylthio)thiophene]

Rella et al. 2000 Hrncirova et al. 2000 Suri et al. 2002

Polyvinyl chloride (PVC)

Polypyrrole nanocomposite

Propylenebutyl copolymer

Nanto et al. 2000 Mizutani et al. 2001 Nylander et al. 1983 Hirata and Sun 1994 Li et al. 2000 Chabukswar et al. 2001 Yadong et al. 2000

Polydimethylsiloxane (PDMS) membrane

Polypyrrole-impregnated filter paper

Polyaniline (PANI)

Nanocomposite ultrathin films of PANI and isopolymolybdic acid (PMA)

Acrylic aciddoped PANI

Polypyrrole

Polypyrrolepoly(vinyl alcohol) (PVA) composite H2S H2S Electrochemical detection Electrochemical sensor Shi et al. 2001 Schiavon et al. 1995 SO2 Cathodic stripping voltammetry

NH3

Gangopadhyay and De 2001 Opekar and Bruckenstein 1984

Ag-metalized porous Teflon membrane

Solid polymer electrolytic (SPE) ion-exchange membrane

Poly(4-vinyl pyridine) on palladium and iridium oxide (PVP/Pd/IrO2)coated platinum microelectrode NO Amperometric sensor

Nitrocellulose membrane and a silicone rubber outer layer, phenylene diamine-modified carbon fiber NO2 Irreversible increase of conductivity due to self-ionization of N2O4 to NO+NO3 Amperometric sensor Conductivity increase Electrical conductivity Electrochemical Electrical conductivity Electrical conductivity

Ichimori et al. 1994; Friedemann et al. 1996 Christensen et al. 1993

Polystyrene film

Pt/Nafion electrode NO2 NO2, NH3 CO CO2 CO and H2

NO2

Ho and Hung 2001 Xie et al. 2002 Rella et al. 2000 Otagawa et al. 1990 Sakai et al. 1995 Torsi et al. 1998

Multilayer films of PANI and PSSA

Poly[3-(butylthio)thiophene] deposited on alumina substrates

Nafion as solid polymer electrolyte

Polyethylene glycolalkali carbonate complex film

Cu/Pd-doped PPy film

20 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

7.1. OPTICAL AND FIBER OPTIC POLYMER-BASED SENSORS


Sensors based on optical and fiber optic properties are a real alternative to electronic ones in electrically noisy environments where electronic sensors do not operate correctly. In fact, optic and fiber optic chemical sensor technology is a field of growing importance, with diversified applications in oceanography, chemical process, clinical diagnosis, and environmental monitoring. Optical sensors offer advantages such as light weight, a passive nature, low attenuation, and the possibility of multiplexing, among others. In a recent review, Elosua et al. (2006) classified these devices according to the sensing mechanism, taking into account the sensing materials and the different methods of fabrication.

7.1.1. Requirements for Sensing Materials


A variety of polymers and polymer composites [e.g., PMMA, poly(2-hydroxyethyl methacrylate) (PHEMA), PTFE, cellulose, Nafion, PVC, polyacrylamide, polyvinyl pyrrolidone, etc.] together with indicator dyes are widely used in chemical sensors, including ion, gas, humidity, and enzyme sensors (Wolfbeis 1991; Gauglitz and Reichert 1992). Polymer materials must fulfill various requirements to enable optical sensing. First of all, the indicator dye and all additives need to be dissolved well in the polymer (and must not be washed out). The polymer should be of such chemical nature that it can dissolve the analyte and allow it to diffuse easily within the polymer. For practical applications, the polymer must be chemically and physically stable in order to achieve good operational lifetime and shelf life. Furthermore, no crystallization/migration/ reorientation of the indicator molecules must occur within the polymer. These phenomena can occur even after weeks or months if the indicator lacks solubility as expected. The polymer must be thermally stable so that it can sustain steam sterilization without degradation. The polymer should be stable against sunlight, acids, bases, oxidants, etc., and nontoxic as well as biocompatible for clinical and biochemical applications. The polymer should not have any intrinsic colour/luminescence, and it should be optically transparent in the spectral range where measurements will be performed. Finally, the polymer should possess adequate stability to mechanical stresses (Mohr 2006). If it has apparent optical properties (AOP), the polymer can be used as the analyte-sensitive layer for various analytes. Polymers that have good mechanical properties, homogeneity, preparation simplicity, and optical transparency, are very good materials for the AOP sensing layer. They also allow reproducibility of film production and therefore are excellent for sensor fabrication. Polymers with very high molecular weights and capable of forming a cagelike structure may be helpful when trying to strengthen the required properties. When sensing analytes, in some polymers a mechanism of analyte dissolution is accompanied by diffusion, in others the analyte simply diffuses through the pores of the polymers. So, here the bulk free volume, which can be monitored from the film thickness or by tailoring the polymer structure, is useful for the sensing of the particular analytes. Hydrophobic polymers which can be penetrated by gases and hydrophilic polymers which can be penetrated by specific ionic species are the most important polymers for selective sensing of analytes. Films of these polymers can also be attached directly to an optical fiber and can be deposited along the length of the fiber before curing. The intrinsic fluorescence of some polymers caused by their backbone

POLYMERS IN CHEMICAL SENSORS 21

units [e.g., PPy, PPV, polyphenylene, polyfluorene, and phenylthiohydantoin (PTH) derivatives] or fluorescence of fluorophores introduced into the polymer can be used.

7.1.2. Mechanisms of Polymer Sensing


Optical sensing is based on colorimetric, fluorescence, or luminescence effects and on changes in light refraction/propagation. Reseach has shown that the change in optical properties of a polymer can provide good sensing response (Wolfbeis 1991; Gauglitz and Reichert 1992). It has been established that an interaction with an analyte that oxidizes/reduces or protonates/deprotonates a conducting polymer influences not only electrical properties but also optical properties. For example, in a polyaniline system, as a result of oxidation/reduction, a wide range of reversible visible color changes can be observed (see Table 1.3). Using this phenomenon, a polyaniline-based optical pH sensor as well as ozone, NO2, and H2S gas sensors have been developed (Ge et al. 1993; Agbor et al. 1997; Hu et al. 2002; Jin et al. 2001; Ando et al. 2002). In intrinsically conducting polymers, the color of the polymer depends on the band gap of the polymer. Changes in the redox or protonation state of an intrinsically conducting polymer lead to strong modification of its electronic band structure. Therefore, certain conjugated conducting polymers are extremely sensitive to structural perturbations as a result of the absorption of external analyte. These structural perturbations in the electronic network of the conducting polymer causes a self-amplifying fluorescence quenching response upon binding of analytes. Depending on the system, these polymers can have strong luminescence properties, which can be used as good fluorescent sensor material. Figure 1.5 shows the mechanism, which is known as photo-induced electron-transfer fluorescence quenching. For conjugated polymers it amplifies the molecular recognition signal via migration of electrons along the polymer chain (Desmonts et al. 2007). Photon irradiation of the polymer causes promotion of an electron to the lowest unoccupied molecular orbital (LUMO) from the highest occupied molecular orbital (HOMO). Now, due to analyte binding, trapping of the electron occurs to the quenchers

Table 1.3. Oxidation states of polyaniline (PANI) with visible colors Structure of PANI
N H N H N N

1-y n

Value of y Form Name Color

1 Reduced

0.5 Half-oxidized

0 Oxidized

01 Protonated salt Polyemeraldine salt Green

Polyleucoemeraldine or Polyemeraldine or Polypernigraniline polyprotoemeraldine base polynigraniline base base Transparent Blue Purple

22 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Figure 1.5. Electron-transfer uorescence quenching mechanism for a conjugated polymer upon interaction with the analyte. (Reproduced with permission from Desmonts et al. 2007. Copyright 2007 The Royal Society of Chemistry.)

LUMO (here, analyte) and effectively deactivates the electron transfer (see Figure 1.5). This destroys the polymer-based excited state, and the polymer can no longer fluoresce. Interaction of an analyte which oxidizes/reduces or protonates/deprotonates an intrinsically conducting polymer can also be measured by UV-vis or photoluminescence spectroscopy. This principle is especially applicable for detection of gaseous analytes. Using the above-discussed techniques, various metal ions have been detected by fluorescent conjugated polymers, e.g., 2,2,9-bipyridyl-phenylene-vinylenebased polymers (Wang and Wasielewski 1997), terpyridine-based poly(p-phenylene-ethynylene)-alt-(thienylene-ethylene) polymers (Zhang et al. 2002), and poly(p-phenylene-ethynylene)based polymers (Chen et al. 2004; Kim and Bunz 2006). Successful detection of fluoride ions as quencher in some fluorescent conjugated polymers has also been reported (Kim and Swager 2003; Tong et al. 2003; Saxena et al. 2004; Zhou et al. 2005). The most successful use of these types of fluorescent polymers has been in detection of vapors of nitroaromatic explosives as quencher, such as trinitrotoluene (TNT) and dinitrotoluene (DNT) (Mcquade et al. 2000; Cumming et al. 2001, 2004; Sohn et al. 2003) by photochemical quenching. Another important mechanistic approach for the optical sensing of analyte by dye-solvated polymer matrices is the polarity change of the polymer. By incorporating solvochromic dye in polymer matrices of varying polarity, hydrophobicity, pore size, flexibility, and swelling tendency, unique sensing regions are created with different fluorescence responses for different organic vapors (Dickinson et al. 1996; Albert et al. 2000). Here the permeation of analyte for some polymers is due to dissolution by diffusion; for some other polymers it is due to diffusion only. In general, hydrophobic polymers have the ability to permeate gases, and hydrophilic polymers have the ability to permeate ionic species (Brook and Narayanaswamy 1998). This mechanism is utilized with silicone, Nafion, and PVC films for the determination of gases (chlorine, nitrogen dioxide, oxygen, etc.) and moisture (Brook and Narayanaswamy 1998). Nafion used as an ion-exchange matrix for crystal violet for the determination of relative humidity is a suitable matrix in an optical fiberbased moisture sensor (Sadaoka et al. 1992).

POLYMERS IN CHEMICAL SENSORS 23

The optical path length in a given medium is determined by the refractive index and the geometric path length. Changes in both parameters result in a change in the phase shift, which can be detected by interferometry. Based on the same effects, in optical waveguides the waveguide effective refractive index or transmission effectiveness variation can also be measured by the change in refractive index of the cladding. By this means, film materials that formerly could not be used may now be used for fiber optic or optical integrated chemical sensors. The refractive index of the polymeric film and/or cladding varies with the permittivity when the vapor to be detected is absorbed in it. Another frequently used method is to detect small changes in thickness of a transparent layer using optical reflection mode interferrometry, which can also be used in fiber optic sensors. Polydimethylsiloxane (PDMS) was examined in several studies and seems to be a good candidate for applications in optical sensors. It shows both swelling and refractive index shift when exposed to organic solvent vapors (Gauglitz et al. 1993). In solution, binding with analytes can also change the conformation of polymers, which leads to a corresponding modification of optical adsorption. This principle was applied to detection of cations in the solution phase (Ewbank et al. 2004). More sensitive detection of optical changes in ICPs can be realized by application of surface plasmon resonance (SPR) or Raman spectroscopy. Other optics-based sensors such as optodes utilize colorimetric changes as well or fluorescence quenching of indicator dyes using excited light for detection (Wolfbeis 1991). It is also possible to fabricate chemical and biochemical sensors based on interferrometry at thin (multi-) layers (Gauglitz et al. 1993). In some cases, the addition of indicators such as dyes or fluorescent materials to the film is not necessary, because the physical-chemical changes take place with the direct participation of the sensing polymer material. Polysiloxane polymers seem to present these properties and can be used successfully in chemical sensor applications. A chemical vapor sensor working with a monochromatic light source and based on optical fiber coated with a thin siloxane polymer film has been developed for in situ monitoring of volatile organic compounds, such as ethylbenzene, o-xylene, heptane, octane, chloroform, carbon tetrachloride, ethanol, and butanol in indoor atmospheres and confined areas of industrial environments. The sensor consists of a monomode optical fiber with an end surface coated with a thin polymeric film by dip coating. The light source utilized was a stabilized laser diode at 1550 nm, and the light power changes were measured with a photodiode. The sensor was tested for different volatile organic compounds and for different individual concentrations in terms of stability, sensitivity, repeatability, and reversibility of the analytical signal. The response and desorption times were found to be 30 s, and good reproducibility and accuracy were also obtained. Finally, the analytical performance of the sensor was also evaluated and found to be adequate for actual monitoring in indoor atmospheres (Silva et al. 2008).

7.1.3. Examples of Polymer-Based Optical Sensors


7.1.3.1. HCL SENSORS
Hydrochloric acid is a source of dioxin produced during the incineration of plants and in acid rain. It is also known as a workplace hazard with a short-term exposure limit of 5 ppm. Composites of alkoxy-

24 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

substituted tetraphenylporphyrinpolymer composite films were developed by Nakagawa et al. (2001) for the detection of HCl at sub-ppm levels. The alkoxy group imparts basicity to the material and hence increases its sensitivity to HCl. Nakagawa et al. achieved high selectivity to sub-ppm levels of HCl gas using 5,10,15,20-tetra (4-butoxyphenyl)porphyrin-butylmethacrylate [TP (OC4H9)PH2-BuMA] composite film. Supriyatno (2001) detected HCl gas optochemically using a mono-substituted tetraphenylporphinpolymer composite film. In this system a higher and preferred sensitivity to sub-ppm levels of HCl was achieved using a polyhexylmethacrylate matrix in the composite.

7.1.3.2. SENSORS OF GASES AND VOLATILE ORGANIC COMPOUNDS


Changes in the optical and electrical properties of -conjugated polyaniline occurs as a result of the interaction of the emeraldine salt (ES) with NH3 gas (Nicho et al. 2001). The interaction of polyaniline with gas molecules decreases the polaron density in the band gap of the polymer. It was observed that PANIPMMA composite coatings are sensitive to very low concentrations of NH3 gas (10 ppm). Optical fiberbased sensors for volatile organic compounds (VOCs) offer new and interesting properties, which can overcome some of the limitations of traditional gas sensors. Horiuchi et al. (2006), for instance, developed a portable gas sensor using fiber optic microfluidic devices (concentration and detection cells) for detecting and identifying gaseous aromatic VOCs such as benzene, toluene, xylene, styrene, and ethylbenzene. They used precise control of gas transfer from the concentration cell to the detection cell and optimized the spectrum measurement conditions in order to improve the sensitivity of the sensor. A detection level of 10 ppb for benzene gas (an aromatic VOC) in a sampling time of 50 min was reported. Good linearity was shown in the calibration curve in the range of 10100 ppb of benzene. The same authors (Horiuchi et al. 2006) carried out gas monitoring experiments in a garage. They also showed the applicability of the aromatic VOC sensor for field monitoring. Blum et al. (2001) designed an alcohol sensor in which two lipophilic derivatives of Reichardts phenolbetaine were dissolved in thin layers of plasticized poly(ethylene vinylacetate) copolymer coated with microporous white PTFE in order to facilitate reflectance (transflectance) measurements. The sensor layers were found to respond to aqueous ethanol with a change of color from green to blue with increasing ethanol content. The highest signal changes were observed at a wavelength of 750 nm, with a linear calibration function up to 20% ethanol (v/v) and a detection limit of 0.1% (v/v). This sensor also responded to acetic acid, thus affecting measurements on beverages, but that limitation was overcome by adjusting the pH of the sample solution. Patra and Mishra (2001) developed an optical sensor for detecting nitro aromatic compounds such as nitrobenzene, m-dinitrobenzene, o-nitrotoluene, m-nitrotoluene, p-nitrobromobenzene, o-nitroaniline, p-nitrophenol, etc., by fluorescence quenching of benzo[k]fluoranthene (BkF) in poly(vinyl alcohol) film. The fluorescence spectra of BkF-doped PVA films in various solvents are shown in Figure 1.6. Figure 1.6 demonstrates that, due to the enhanced swelling of the film in methanol, the sensor film showed good fluorescence quantum yield in methanol compared to other solvents. However, less fluorescence was observed with the same PVA film due to the lower quantum yield of BkF in water than in methanol, although PVA swells more in water. Polypyrrolenitrotoluene copolymer provides selective response to aromatic hydrocarbons (Josowicz et al. 1987).

POLYMERS IN CHEMICAL SENSORS 25

Figure 1.6. Emission spectra of PVA lm doped with benzo[k]uoranthene in various solvent media (ex = 308 nm). (Reproduced with permission from Patra and Mishra 2001. Copyright 2001 Elsevier.)

Several polymers have been used to detect nitro aromatic explosives by a variety of transduction schemes. Detection relies on both electronic and structural interactions between the sensing material and the analyte. Quenching of luminescent polymers by electron-deficient nitro aromatic explosives, such as trinitrotoluene, may be monitored to detect explosives. Luminescent polymetallocenes have recently been investigated for sensing explosives in aqueous-based solutions and for improved visual detection of trace particulates on surfaces (Toal and Trogler 2006). Photoluminescence quenching of the various metallocene polymers and their copolymers for picric acid, TNT, DNT, and nitrobenzene were reported by Sohn et al. (2003). The quenching of photoluminescence spectra by the nitro aromatics was better for the substituted silolesilane copolymer (see Figure 1.7) in the order picric acid > TNT > DNT> nitrobenzene at various concentrations.

Ph

Ph

Ph MeO

Si Si H

Ph

OMe n

Ph

Figure 1.7. Structure of substituted silolesilane copolymer.

26 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

7.1.3.3. OXYGEN SENSORS


Luminescent sensors made of composites of transition-metal complexes dispersed in polymer matrices were developed for sensing oxygen in biomedical and barometric applications. Phosphorescent dyes were dispersed in a polymer matrix that had high gas permeability. Pang et al. (1996) controlled the sensitivity of the phosphorescent oxygen sensors over a wide range by using SNP polymers as a novel matrix material. Miura et al. (1984) developed a concentration celltype O2 sensor by using Nafion membrane as a proton conductor. The concept of using oxygen-quenchable luminescent dyes in chemically homogenous polymer layers may lead to promising applications in oxygen sensing. The oxygen-sensing behavior was studied by exploiting the luminescence quenching behavior of tris(4,7-diphenyl-1,10-phenanthroline) Ru(II) perchlorate dissolved in a polystyrene layer (Hartmann et al. 1995). The photophysical and photochemical properties of aluminum phthalocyanine polystyrene film were measured for use in a fluorescence-quenching oxygen sensor (Amao et al. 2000). Next they developed an optical oxygen sensor based on the luminescence intensity changes of tris(2phenylpyridine anion) iridium(III) complex ([Ir(ppy)3]) immobilized in a fluoropolymer, poly(styreneco-2,2,2-trifluoroethyl methacrylate) [poly (styrene-co-TFEM)] film (Amao et al. 2001). They observed a decrease of luminescence intensity of [Ir(ppy)3] in poly(styrene-co-TFEM) film with increasing oxygen concentration.

7.1.3.4. OPTICAL pH SENSORS


Munkholm et al. (1986) prepared a pH sensor device using photochemically polymerized copolymer of acrylamide-methylenebis(acrylamide) containing fluoresceinamine covalently attached to an optical fiber surface (core diameter 100 mm). The need for organic dyes has been eliminated by the use of conducting polymers in the preparation of optical pH sensors. Demarcos and Wolfbeis (1996) developed an optical pH sensor with polypyrrole by oxidative polymerization. This polymer film, having suitable optical properties for optical pH sensing, has eliminated the immobilization step for an organic dye during preparation of the sensor layer. The first commercial optical blood pH sensor was developed in 1995 by Leiner. Likewise, several groups of researchers (Ge et al. 1993; Pringsheim et al. 1997; Grummt et al. 1997) have developed optical pH sensors from polyaniline for measurement of pH in the range 212 and reported that the polyaniline films, synthesized within a time span of 30 min, are very stable in water. Jin et al. (2000) reported an optical pH sensor based on polyaniline film prepared by chemical oxidation at room temperature. They significantly improved the stability of the polyaniline film by increasing the reaction time up to 12 h. The film showed a rapid, reversible color change as a result of a change of pH. The solution pH was determined by monitoring either the absorption at a fixed wavelength or the maximum absorption wavelength of the film. The change in the electronic spectrum of polyaniline was explained by the different degrees of protonation of the imine nitrogen atoms in the polymer chain as a result of the change in pH (Chiang and MacDiarmid 1986). That these optical pH sensors were environmentally stable was verified by keeping the sensors exposed in air for over 1 month without any deterioration in sensor performance.

POLYMERS IN CHEMICAL SENSORS 27

Ferguson et al. (1997) developed a pH sensor by incorporating acryloyl fluorescein as pH indicator in poly(hydroxyethyl methacrylate) hydrogel. Shakhsher et al. (1994) developed a fiber optic pH sensor based on changes accompanying the swelling of a small drop of aminated polystyrene (quaternized) on the tip of the single optical fiber. A low-cost optical pH sensor was developed by immobilizing a direct indicator dye, e.g., naphthyl red or alizarine yellow, to a porous and transparent acetylcellulose film. The membranes were durable (>12 months) and had short response times (<5 s) (Ensafi and Kazemzadeh 1999). Janowiak et al. (2001) developed a technology to support specific analytical reagents on polymeric membranes for fiber opticbased pH sensors. A synthesis strategy resulting in a three-layered probe was designed. The technique involves the growth of a hydrophilic polymer matrix from a fiber optic and incorporates specific pH indicator dyes. Wolfbeis (2006) has reviewed developments in optical sensors for gases, vapors, and humidity. A review by Korostynska et al. (2007) describes current state-of-the-art methods of measuring pH levels that are based on polymer materials, viz., polymer-coated fiber optic sensors, devices with electrodes modified with pH-sensitive polymers, fluorescence pH indicators, potentiometric pH sensors, as well as sensors that use a combinatory approach for ion concentration monitoring.

7.1.4. Future Trends in Optical Sensor Design


Optical sensing devices should ideally have high sensitivity, selectivity, and reliability, and they should be able to perform measurements in real time, in a site-specific fashion. However, although optical sensors are important because they are both very simple and cost-effective to manufacture and they enable rather sophisticated multisensor applications, they also suffer from some serious disadvantages. Existing optical sensing principles can be applied to a huge number of applications, but doing so requires interdisciplinary understanding of the detection principles, sensitive layers, kinetics, and thermodynamics of the interaction processes and of the fluidics. Thus, fundamental research must be performed on these problems to characterize the layers and the interaction processes. For better prediction of sensor performance, characterization of (plasticized) polymers with respect to lipophilicity, dielectric constant, polarity, water uptake, etc., under reproducible conditions is needed.

7.2. CONDUCTOMETRIC GAS SENSORS 7.2.1. Conducting Polymers in Gas Sensors


A change in conductivity upon exposure to an analyte is the main requirement for a sensing material intended for use in a conductometric gas sensor. Conducting polymers having acidbase or oxidizing/ reducing characteristics fulfill this requirement and therefore they are widely used in various conductometric sensing devices. It has been demonstrated (Heeger 2001) that the molecular arrangement in a conducting polymer must contain alternating single and double bonds in order to allow the formation of delocalized electronic states. The driving force for the delocalization of these states is associated with the resonance-stabilized structure of the polymer. Electrically conducting polymers display electrical

28 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

conductivities that are dependent on the concentration of dopant ions incorporated in the material. As we have shown before, most of the conducting polymers are doped/undoped by redox reactions; therefore, their doping level can be altered by transferring electrons from or to the analyte. Electron transfer causes the changes in resistance and work function of the sensing material. The conductivity, usually measured using direct current (DC) techniques, may be modulated reversibly and rapidly at ambient temperature by adsorption and desorption of volatile chemicals (Amrani et al. 1996; Bissell et al. 2002). As the conducting polymer undergoes a reversible redox reaction during interaction with various analytes, sensing by the conducting polymer is possible due to interaction with the polymer chain or the dopant, either chemically or electronically. Conducting polymers that are currently being used in various gas sensors are listed in Table 1.2. Composites of conducting polymers with PVC, PMMA, etc., with active functional groups and solid polymer electrolytes (SPEs) are also used to detect such gases. Nowadays, a very simple electrochemical sensor setup is used for conducting polymers (see Figure 1.8). Either two probes or four probes are connected to the polymer-detecting substrate to measure directly the resistivity (for two probes) or voltage at a particular applied current or vice versa (for four probes). A special feature of semiconducting electronic devices is their efficient performance at ambient temperature with low power consumption. From this point of view, conducting polymers are superior over inorganic semiconductors. Reasons for choosing conducting polymers as sensing elements in conductometric gas sensors are as follows (Persaud 2005): (1) The sensors show rapid adsorption and desorption

(a)

(b)

Figure 1.8. Conductometric gas sensor setup: (a) two-probe; (b) four-probe.

POLYMERS IN CHEMICAL SENSORS 29

kinetics at room temperature; (2) the sensor elements feature low power consumption (of the order of microwatts), because no heater element is required; (3) the polymer structure can be correlated to specificity toward particular classes of chemical compounds; and (4) the sensors are resilient to poisoning by compounds that would normally inactivate some inorganic semiconductor sensors. In addition, one can expect good sensitivity from a higher surface area of deposited conducting polymer.

7.2.2. Mechanism of Conducting Polymer Sensing


7.2.2.1. CHEMICAL REACTIONS
The sensing mechanism of conducting polymers depends on the detecting analyte as well as the detecting polymer layer. The Lewis acid and base characteristics of these compounds have led to the concept called secondary doping: Electron donation or withdrawal by the analyte vapors leads to conductivity changes in the sensor films, in addition to polymer doping by counter ions before gas exposure (primary doping) (Persaud 2005). A similar idea has been applied to the more common analytes of volatile organics, leading to a model of partial charge transfer between the polymer and the analytes, although the Lewis acid and base features of these compounds are much less prominent. Partial electron transfer may increase or reduce the concentration of the charge carriers (polarons and bipolarons) in the polymer backbone and, hence, polymer conductivity; the direction is determined by the relative magnitude of the electronegativity of the vapor and the work function of the polymer (Janata and Josowicz 2003). In addition to this model based on the band theory of semiconductors, Charlesworth et al. (1993) have presented a model of a dielectric effect on electron hopping. Polymer conductivity is considered to be determined by electron hopping between polymer chains or over intrachain defects, and the hopping rate is affected by the dielectric property of the vapor. Others have reported physical effects on conductivity as a result of polymer swelling by the organic vapors. It has also been recognized that variation in the extent of sorption of different vapors by the polymer may lead to significant differences in sensor performance. So, conducting polymers interact with a gas or vapor analyte mainly in two ways: chemically or physically. Such interactions of various gases with conducting polymers are represented in Table 1.4 (Lange et al. 2008). As the conducting polymer interacts with gaseous species, it can act either as an electron donor or as an electron acceptor. If a p-type conducting polymer donates electrons to the gas, its hole conductivity increases. Conversely, when the same polymer acts as an electron acceptor, its conductivity decreases. In addition to the change in the number of carriers, there may also be a change in bulk mobility. This is usually due to conformational changes of the polymer backbone (Zheng et al. 1997). It has been observed that nucleophilic gases (H2S, NH3, N2H4, methanol, ethanol, etc.) cause a decrease in conductivity in most widely used conducting polymers in gas-sensing applications, such as polythiophene, polypyrroles, polyaniline, and their derivatives or composites. For example, ammonia (Mohammad 1998; Sakurai et al. 2002) and H2S (Hanawa et al. 1989) gases were found to decrease the conductivity of conducting polythiophene due to the reduction of the conducting polymer. Reduction of polypyrrole (Torsi et al. 1998; Roy et al. 2003) and polyaniline (Roy et al. 2003) also resulting in a decrease in conductivity during hydrogen sensing.

Table 1.4. Interaction of gas molecules with conducting polymers and their effects

TYPE OF INTERACTION EXAMPLES CHCl3, CH2Cl2, alcohols, acetone, acetonitrile, alkanes, cyclohexane, benzene, toluene Aliphatic alcohols, acetone NO2, SO2, O3 Increase in electrical conductivity Decrease in electrical conductivity Increase in electrical conductivity Decrease or increase in electrical conductivity SENSING PROPERTIES REFERENCES Athawale and Kulkarni 2000; Tan and Blackwood 2000; Vercelli et al. 2002; Virji et al. 2004; Reemts et al. 2004; Torsi et al. 2004; Ruangchuay et al. 2004; Segal et al. 2005; Li et al. 2007 Ruangchuay et al. 2004; Kar et al. 2009 Prissanaroon et al. 2000; Ando et al. 2002, 2005; Xie et al. 2002; Ram et al. 2005; Yan et al. 2007 Ratcliffe 1990; Ellis et al. 1996; Mohammad et al. 1998; Torsi et al. 1998; Sakurai et al. 2002; Virji et al. 2004; Yang et al. 2006; Thomas and Swager 2006 Decrease or increase in electrical conductivity Hao et al. 2003; Reemts et al. 2004; Virji et al. 2004; Hao et al. 2005; Krondak et al. 2006; Sengupta et al. 2006, 2009

WITH CONDUCTING POLYMER

Physical interaction

Diffusion

Weak hydrogen bonding

Chemical interaction H2, N2H4, NH3, H2S

Oxidation

Reduction

Protonation/deprotonation

HCl, NH3

POLYMERS IN CHEMICAL SENSORS 31

Ammonia sensors based on protonic aciddoped polyaniline films have been extensively studied (Koul et al. 2001; Debarnot and Epaillard 2003; Sengupta et al. 2009). The mechanism of ammonia sensing by HCl-doped conducting PANI is somewhat different (Sengupta et al. 2006). Here, the analyte ammonia is attached to the dopant H+ and, due to the weak doping effect of the resulting NH4+ ion, the conductivity of the film is decreased (see Figure 1.9). This deprotonation of PANI by ammonia is a cause of increased resistance. Polypyrrole shows the same effect as polyaniline at low ammonia concentrations, but at higher concentrations, in the presence of humidity, the deprotonation process becomes reversible (Gustafsson et al. 1989). The resistance of the film changes through the formation of a positively charged electric barrier of NH4+ ions in the submicrometer film. The electron lone pair of the NH3 gas acts as a donor to the p-type polypyrrole semiconductor, thereby reducing the number of holes in the polypyrrole and hence causing the resistivity to increase. The reduction of conducting polymers such as polypyrrole (Ratcliffe 1990), poly-3-hexylthiophene (P3HTH) (Ellis et al. 1996; Yang et al. 2006), and PANI (Virji et al. 2004) by hydrazine vapors has also been reported. On the other hand, electrophilic gases with higher electron affinity than the conducting polymer (NOx, PCl3, SO2, O2, etc.) show the opposite effect (Slater and Watt 1991), by increasing the number of charge carriers in the conducting polymer through oxidative doping. For this reason, NO2 gas was found to decrease the resistance in polyaniline (Xie et al. 2002) and P3HTH (Ram et al. 2005b); however, an increase in resistance was observed for emeraldine salt nanofiber, due to oxidation by NO2 to pernigraniline base (Yan et al. 2007). SO2 also decreases the resistance of PPY by a similar mechanism as proposed for the NO2P3HTH system (Prissanaroon et al. 2000). A decrease in resistance of a PANI film was observed when it was exposed to CO gas (Misra et al. 2004; Dixit et al. 2005; Densakulprasert et al. 2005; Watcharaphalakorn et al. 2005). According to Densakulprasert et al. (2005), no difference was observed in UV-vis and x-ray diffraction patterns of the film before and after CO exposure. According to their explanation, the stable resonance structure of CO with +C=O withdraws electrons to form amine nitrogen (NH), which results in a net increase of positive charge carriers on the polymer backbones and therefore an increase in conductivity. A welldefined scheme has been drawn by Bai and Shi (2007). For the case of chlorinated hydrocarbon sensing by PANI, similar phenomena were observed by Anitha and Subramanian (2005).

NH

NH

A NH3
N

NH3
NH

NH4A
Figure 1.9. Sensing mechanism of ammonia by HCl-doped polyaniline lm.

32 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Another explanation for the response of PANI to CO is also based on the redox reaction. The decrease in resistance was interpreted as reduction of the barrier height between grains (Misra et al. 2004; Dixit et al. 2005). The conductance of the sensing film is governed by potential barriers between polymer grains. The oxidation that occurs at the grain surfaces in the presence of CO gas causes the surface coverage of adsorbed oxygen to decrease. Thus, the surface potential, barrier height, and depletion length are reduced, which leads to a decrease in resistance.

7.2.2.2. WEAK INTERACTIONS


Many important organic analytes are not reactive at room temperature and under mild conditions. Therefore, it is difficult to detect them by their chemical reactions with conducting polymers. However, they may have weak physical interactions with the sensing polymers, involving absorbing or swelling of the polymer matrixes. Like other polymers, conducting polymers can swell in many organic solvents. This is controlled by the vapor molecular volume, the affinity of the vapor to the sensing polymer, and the physical state of the polymer (Segal et al. 2002). It is necessary to note that the swelling is the main mechanism of the conductivity response of sensors based on nanocomposites, including isolated polymer and carbon black. Weak physical interactions with conducting polymers are observed for nonreactive volatile organic compounds such as chloroform, acetone, aliphatic alcohols, benzene, toluene, and some other VOCs. These interactions do not change the oxidation levels of the conducting polymers, but they can influence the properties of the sensing materials and thus make these gases detectable (Ruangchuay et al. 2003; Virji et al. 2004). For example, the diffusion of those gases to the intermolecular space or surface void space of the polymer can lead to some effect on the conductivity of the polymer. It has been found that for a pure conducting polymer, inserting analyte molecules into the polymer matrix generically increases interchain distance, which affects electron hopping between different polymer chains. Experiments have shown that this type of diffusion of chloroform, acetone, ethanol, acetonitrile, toluene, and hexane reduced the conductivity of PANI, PPY, PTH, and polythiophene derivatives (Torsi et al. 1998; Vercelli et al. 2002; Ruangchuay et al. 2004; Virji et al. 2004; Li et al. 2007). The swelling process in a composite conducting polymer is complicated. One or more components can swell to different extents, which results in various changes in overall conductivity. In some cases, the analyte dissolves conducting polymer better than the other component, and it will swell first. Segal et al. (2005) synthesized polyaniline/polystyrene (PS) composite films and tested their response to alcohols. Because PANI has a higher solubility in polar alcohols, it swelled much more than PS, which in fact increased the effective volume of the conducting PANI. This resulted in increasing the conductivity of the PANI. In some cases, other components than the conducting polymers in the composite swelled more. For example, when PPy/PMMA composite film was exposed in acetone, PMMA swelled much more than PPy and separated conducting PPy. Thus, the conductivity of the composite film was decreased (Ruangchuay et al. 2003). Similar results were also obtained in a PANI/PVA composite sensor for humidity (McGovern et al. 2005) and in PPy/polyvinyl acetate (PVAc), PPy/PS, and PPy/polyvinyl chloride (PVC) for some toxic gases (Hosseini and Entezami, 2003). In some cases, a catalyst incorporated in the conducting polymer film can help in detecting some inert analytes. Athawale et al. (2006) prepared a nanocomposite of Pd/PANI and found that its electrical

POLYMERS IN CHEMICAL SENSORS 33

O O S O N H

H
O

Od H O R

H W h e re , R = C H 3 : M e th a n o l = C H 2 C H 3 : E th a n o l
Figure 1.10. Aliphatic alcohol sensing mechanism by sulfuric aciddoped conducting poly(m-aminophenol). (Reproduced with permission from Kar et al. 2009.Copyright 2009 Elsevier.)

resistance responded rapidly and reversibly in the presence of methanol. Athawale et al. assumed that the effective positive charges on the imine nitrogen atoms were reduced by the methanol molecules in the presence of Pd nanoparticles. Experimental results also demonstrated that some analyte gases, especially alcohols (Svetlicic et al. 1998; Athawale et al. 2000) and ketones (Ruangchuay et al. 2004), could change the crystallinity of conducting polymers. For example, the increase in crystallinity is explained as the reason for the conductivity increase during aliphatic alcohol sensing by polyaniline and its derivatives (Athawale and Kulkarni 2000). Recently, Kar et al. (2009) reported the effect of functional groups on aliphatic alcohol vapor sensing by poly(m-aminophenol). The OH groups of methanol or alcohol molecules were hydrogen-bonded with the phenolic OH groups present in the polymer molecule (see Figure 1.10), and ultimately a week electron flow throughout the polymer chain increased. Adsorption of ethanol and hexanol on dipentoxy-substituted polyterthiophene changes the potential barrier at the boundaries between polymer grains and is the reason for the change in conductivity (Torsi et al. 2004). Hydrogen bonding between acetone and polypyrrole makes electron hopping difficult, and so the conductivity decreases in acetone sensing by polypyrrole (Ruangchuay et al. 2004).

7.2.3. Examples of Polymer-Based Conductometric Gas Sensors


Conducting polymers such as polypyrrole, polythiophene, polyaniline, polyindole, and their derivatives and composites are the most widely used materials as transduction matrices sensitive to gases, vapors, ions, and biomolecular systems, resulting in a straightforward conductance, impedance, or redox potential change via modulation of their doping level (Bidan 1992). They are also very popularly used as sensing materials for gas and chemical odors in sensor arrays (Persaud and Pelosi 1985; Amrani et al. 1993).

34 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

An electronic nose, which is an array-based system that attempts to mimic the mammalian olfactory system, can recognize complex, distinct, and diverse odors of cheeses, beers, olive oil, explosives, and pathogenic bacteria (Gardner and Bartlett 1999). Polypyrrole, the first conducting polymer used in gas sensors, showed low sensitivity, long response time, and incomplete reversibility of the sensor response. During the last decade, the properties of polypyrrole-based polymers were considerably improved. However, due to easy synthesizability, better processability, easy dopability, and better stability, polyaniline (PANI), its derivatives, and their composites have become the important conducting polymer materials for gas sensors (Bai and Shi 2007).

7.2.3.1. AMMONIA GAS SENSORS


Nylander et al. (1983) reported ammonia gas sensing using polypyrrole-impregnated filter paper and obtained linear performance of the sensor at room temperature with higher concentrations (0.55%) with around 1.0 min response time. Composites of electrically conducting polyacrylonitrile (PAN)/polypyrrole (Park and Han 1992), polythiophene/polystyrene, polythiophene/polycarbonate, polypyrrole/polystyrene, and polypyrrole/polycarbonate (Wang et al. 1990) have been prepared by electropolymerization of the conducting polymers within the matrix of the insulating polymers PAN, polystyrene, and polycarbonate, respectively. Ammonia has been detected by electroactive nanocomposite ultrathin films of PANI and isopolymolybdic acid (PMA). The device was prepared by alternate deposition of PANI and PMA following Langmuir-Blodgett and self-assembly techniques (Li et al. 2000). The process was based on a dopinginduced deposition effect of the emeraldine base. Chabukswar et al. (2001) used acrylic acid as dopant in polyaniline for sensing of ammonia vapor over a broad range of concentrations, viz., 1600 ppm. The change in DC electrical resistance of the polymer was found to increase linearly upon exposure to ammonia up to 58 ppm concentration and to saturate thereafter. They explained the decrease in resistance on the basis of removal of a proton from the acrylic acid dopant by the ammonia molecules, thereby rendering free conduction sites in the polymer matrix. A plot of the variation of relative response of the ammonia gas sensor with increase in the concentration of ammonia gas is shown in Figure 1.11. For an acrylic acid doped polyaniline sensor, a sharp increase in relative response was obtained for around 10 ppm ammonia, which subsequently remained constant beyond 500 ppm. On the other hand, the nanocomposite of polyaniline and PMA showed a decrease of relative response with the increase in ammonia concentration. NH3 sensitivity was also detected by the change in resistance of a submicrometer polypyrrole film (Yadong et al. 2000). A polypyrrolepoly(vinyl alcohol) composite was prepared by electropolymerizing pyrrole in a cross-linked matrix of PVA for the purpose of sensing NH3 gas (Gangopadhyay and De 2001). Linsey and Street (1984) studied gas-sensing behavior of PPy-PVA films prepared by electrochemical polymerization on to a precoated PVA matrix. These studies contrasted the advantages of mechanical properties of the host polymer with the electrical properties of PPy-PVA composite films. To obtain good response and reproducibility in gas sensing, a thin film is better than a pellet, because absorption and desorption of gas will be faster in a film than in a pellet, due to molecular order and compactness inside the film. The drawback of previously reported results (Bidan 1992; Hirata and Sun 1994) is related

POLYMERS IN CHEMICAL SENSORS 35

Figure 1.11. Variation of relative response of ammonia gas by acrylic aciddoped polyaniline sensor with increased concentration. (Reproduced with permission from Chabukswar et al. 2001. Copyright 2001 Elsevier.)

to insufficient gas sensitivity and irreversibility for ammonia sensing by HCl-doped polyaniline. This is due to the deprotonation interaction of ammonia with the doped polymer (see Figure 1.9). The same effect has been observed in ammonia sensing by polypyrrole. Although the response is somewhat reversible at low ammonia concentration, it becomes irreversible at higher ammonia concentration, especially in the presence of humidity (Gustafsson et al. 1989).

7.2.3.2. NITROGEN OXIDE GAS SENSORS


Although sensor development for acidicbasic gases (e.g., CO2, NH3) and oxygen have a long history, there is a challenge in the need for rapid, sensitive detection of nitric oxide (NO). There is increasing interest in determination of NO, primarily because of its role in intra- and intercellular signal transduction in tissues (Cunningham 1998). Christensen et al. (1993) developed a NO2 sensing device using a polystyrene film. Upon exposing the film to a 1:10 v/v mixture of NO2/N2, the conductivity of the film increased irreversibly and rapidly, by several orders of magnitude. It was believed that the increase in conductivity of the film was due to self-ionization of N2O4 to NO+NO3.

36 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Xie et al. (2002) reported a polyaniline-based gas sensor made by ultrathin-film technology. They prepared multilayer films of PANI, PANI and acetic acid, and PANI and polystyrenesulfonic acid (PSSA) composite by Langmuir-Blodgett (LB) and self-assembly (SA) techniques. The pure PANI films prepared by the LB technique showed good sensitivity to NO2, while the SA films exhibited faster recovery. PANI is oxidized in contact with NO2. Contact of NO2 with the -electron network of polyaniline results in the transfer of an electron from the polymer to the gas, making the polymer positively charged, which gives rise to increased conductivity. The PANIacetic acid films showed lower sensitivity, due to the occupation of the chemically blocked sensitive sites responsive to NO2 by acetic acid molecules. Films of poly [(3-butylthio) thiophene] were prepared by LB deposition and casting techniques for applications in gas sensor devices (Rella et al. 2000). The sensing layer was prepared by the two techniques separately: the LB deposition of the polymer in a mixture with arachidic acid and direct casting from a solution of the polymer in chloroform. In both cases the deposition was done on alumina substrates equipped with gold interdigitated electrodes. The deposited devices showed changes in electrical conductivity when exposed to a mixture of NO2-oxidizing and NH3-reducing agents at about 100C (see Figure 1.12). According to Rella et al. (2000), the first drop of current is due to the reducing interaction

Figure 1.12. Response curve for poly-3-(butylthio)thiophene LangmuirBlodgett lms in NH3 and NO2 mixture with dry air at 100C. (Reproduced with permission from Rella et al. 2000. Copyright 2000 Elsevier.)

POLYMERS IN CHEMICAL SENSORS 37

of 1000 ppm ammonia, while the next rise of current is because of the p-type doping by 10-ppm NO2 (see Figure 1.12).

7.2.3.3. CO2 GAS SENSORS


A CO2 gas sensor, consisting of K2CO3polyethylene glycol solution supported on porous alumina ceramics, was developed by Wu et al. (1989). Upon exposure to CO2 and under an applied voltage, the resistance of the device increased. A linear relationship between the sensitivity (the ratio of resistance in CO2 to that in air) and the CO2 concentration from 1% to 9% was reported. An improvement of this sensor was reported by solidifying the sensing layer using a solid polyethylene glycol of high molecular weight doped with a solution comprised of liquid polyethylene glycol and K2CO3 (Sakai et al. 1995). The change in resistance was attributed to the change in concentration of the charge carrier K+ ion.

7.2.3.4. H2 AND CO SENSORS


Doping of electrochemically synthesized conducting polymers, such as polypyrrole and poly-3-methylthiophene, with copper and palladium creates a good material for gas sensing (Torsi et al. 1998). These metals were deposited potentiostatically, either on pristine conducting films or on partially reduced samples. The resistance of the PPy and Cu-doped PPy film sensors after exposure to H2 and CO reducing gases was enhanced. On the other hand, there was a drastic drop in resistivity of the PdPPy sensor to H2 and CO, while a resistivity enhancement was shown upon exposure to ammonia. Moreover, the responses of the PdPPy sensor to CO and H2 were highly reversible and reproducible. Roy et al. (2003) reported the hydrogen gassensing characteristics of doped polyaniline and polypyrrole films.

7.2.3.5. SENSORS OF VOCS


Environmental pollution arising out of volatile organic compounds from chemicals and fertilizers, pesticides, and waste streams is a global concern today. For detection as well as assessment of levels of pollution from those volatile organic compounds, chemical sensors play a major role. Detection of VOCs is a topic of growing interest, with applications in diverse fields ranging from environmental uses to food and chemical industries. Recently, highly crystalline nanostructures of regioregular polythiophenebased conductive copolymers have been shown to hold considerable promise as an active layer in VOC chemiresistor sensors (Li et al. 2006). The regioregular polythiophene polymer chain provides a charge conduction path, and its chemical sensing selectivity and sensitivity are altered either by incorporating a second polymer to form a block copolymer or by making a random copolymer of polythiophene with different alkyl side chains. These copolymers were exposed to different VOC vapors. The electrical conductivity of the copolymers either increased or decreased depending on the polymer composition and the specific analytes. Measuring at room temperature, the responses were found to be fast and appeared to be completely reversible. Various copolymers of polythiophene in a sensor array provided much better discrimination to various analytes compared to existing solid-state sensors.

38 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Since conducting polymers have gained popularity as competent sensor materials for organic vapors, a few reports are available describing the use of polyaniline as a sensor for alcohol vapors, such as methanol, ethanol, and propanol (Sukeerthi and Contractor 1994; Hatfield et al. 1994). Polyaniline doped with camphor sulfonic acid (CSA) also showed a good response to alcohol vapors (Xia et al. 1995; MacDiarmid and Epstein 1994, 1995; Svetlicic et al. 1998). These reports discussed the sensing mechanism on the basis of the crystallinity of polyaniline. Polyaniline and its substituted derivatives, such as poly(o-toluidine), poly(o-anisidine), poly(N-methyl aniline), poly(N-ethyl aniline), poly(2,3dimethyl aniline), poly(2,5-dimethyl aniline), and poly(diphenyl amine) were reported to be sensitive to various alcohols such as methanol, ethanol, propanol, butanol, and heptanol vapors by Athawale and Kulkarni (2000). All the polymers were found to respond to saturated alcohol vapors by undergoing a change in resistance. While the resistance decreased in the presence of short-chain alcohols, viz., methanol, ethanol, and propanol, an opposite trend in the change of resistance was observed with butanol and heptanol vapors. The authors attributed the change in resistance of the polymers upon exposure to different alcohol vapors to their chemical structure, chain length, and dielectric nature. These polymers showed a sensitivity of 60% for short-chain alcohols, at up to 3000 ppm concentration, but none of them were suitable for long-chain alcohols. As a reason, the authors indicated a vapor-induced change in the crystallinity of the polymer. Recently we have reported sulfuric aciddoped poly(m-aminophenol) as a very good aliphatic alcohol vapor sensor material (Kar et al. 2009). The polymer shows very good response for methanol and

Figure 1.13. Variation of response for sulfuric aciddoped poly(m-aminophenol) with te methanol and ethanol vapor concentration (the polynomial equations are given in the corner, and R2 is the correlation coefcient). (Reproduced with permission from Kar et al. 2009. Copyright 2009 Elsevier.)

POLYMERS IN CHEMICAL SENSORS 39

ethanol but is a very poor sensor for other higher alcohols such as isopropanol. The variation of response with methanol and ethanol vapor concentration is shown in Figure 1.13 with the equations for the polynomial fit. From the equation on the plot, one can calculate the lower limits for alcohol sensing. These were obtained as 80 ppm for ethanol and 100 ppm for methanol. Polypyrrole has also been studied as a sensing layer for alcohols. Jun et al. (2003) incorporated dodecyl benzene sulfonic acid (DBSA) and ammonium persulfate (APS) in polypyrrole in a sensor, which showed a linear change in resistance upon exposure to methanol vapor in the range 875000 ppm. Bartlett and Ling-Chung (1989) also detected methanol vapor by measuring the change in resistance of a polypyrrole film and obtained a rapid and reversible response at room temperature. The effects of methanol concentration, operating temperature, and film thickness on the response were investigated. A liquidphase alcohol sensor, based on a reflection hologram distributed within a poly(hydroxyethyl methacrylate) film, was reported by Mayes et al. (1999) for measuring alcohol-induced thickness changes. With advantages of high selectivity and operation at ambient temperatures over inorganic gas sensor materials, polyaniline and polypyrrole have found use in multicomponent gas sensing (Harsnyi et al. 1999).

7.2.4. Problems of Polymer-Based Conductometric Gas Sensors


Although conducting polymers as sensor materials have advantages over inorganic semiconductors, no commercial systems have yet been developed. The reasons can be attributed to the following problems. 1. The response times of conductometric sensors are long. Usually, response times exceed hundreds of seconds, and only for some ultrathin-film sensors can this time be as short as about several seconds. For example, the response time for ammonia sensing by polyaniline is around 4 min, and the response of the polyaniline film generally depends on the concentration of ammonia vapor (Kukla et al. 1996). Using polyacrylic acidblended polyaniline, Athawale and Chabukswar (2001) reported good response for ammonia at very low concentration with moderate response time but long recovery time. 2. These sensors possess long-time instability. The performance of this kind of sensors decreased dramatically when they were stored in air for a relatively long time. This phenomenon can be explained as de-doping of the conducting polymers. Many conducting polymers such as PPy and PTh are easily de-doped when they are exposed to air. In addition, oxygen may cause degeneration of some conducting polymers. However, it was reported that PPy doped with big anions can retain conductivity for 20 years (Ricks-Laskoski and Buckley 2006), and PPy doped with an amphiphilic anion can reduce the influence of water and oxygen (Bay et al. 2002). Polymer sensors used for environmental control are also inclined to degradation due to their sensitivity to UV radiation and the presence of oxidizing gases. It has been reported that ozone and other oxidizing components of the polluted atmosphere of industrial centers may be initiators or accelerators of photochemical destruction of polymers (Razumovskii and Zaikov 1982; Heeg et al. 2001). As a result, gas sensors based on these materials have short life spans, especially in normal atmospheres containing water and active gases. It was found that among other polymers, undoped PPy, as a semiconducting polymer, is rather stable toward UV irradiation, which

40 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

can actually increase its conductivity (Fang et al. 2002). However, the stability of PPy against UV irradiation depends on the type of dopant present in the polymer and the power density of the UV irradiation (Rabek 1995). 3. High sensitivity to air humidity is other important disadvantage of polymer-based gas sensors (Sandier and Karo 1974; Kumar and Sharma 1998). This means that humidity must be considered when detecting other gases in air. 4. Conducting polymers have low selectivity. Not only sensors based on conducting polymers, but also other sensors have to face this problem. A single sensor cannot distinguish different analytes, and the response can be easily influenced by the presence of other gases. Nearly all the conducting polymers are sensitive to redox-active gases, such as NH3 and NO2, and to organic vapors (Bai and Shi 2007). 5. Irreversibility is another disadvantage of many polymer-based sensors. It was found that the response of sensors, especially ammonia sesnors, gradually fell during sensing cycles, or the signal did not return to the original value after exposure to analytes (Hao et al. 2005; Kim et al. 2005). The irreversibility of PPy in ammonia may be caused by nucleophilic attack on the carbon backbone (Kemp et al. 2006), but the mechanism of irreversibility is still not clear. It is necessary to note that all the disadvantages of conductometric gas sensors mentioned above are problems of other polymer-based sensors as well.

7.3. SAW AND QCM POLYMER-BASED SENSORS


Mass sensing is a popular method for chemical analysis, since the mass is a fundamental physical property of any matter. Mass-sensitive devices transform the change in mass after interaction with analyte into a change in a property of the sensor detection element. Quartz crystal microbalance (QCM) and surface acoustic wave (SAW) sensors are very stable devices, capable of measuring the change of an extremely small mass (Forster 1998). Piezoelectric devices are based on the measurement of the frequency change of the quartz oscillator plate caused by adsorption of a mass of the analyte at the oscillator. In SAW sensors a change in velocity of surface acoustic waves, caused by a change in mass of the coating on the sensor due to absorption of an analyte species, alters the resonant frequency of the wave (Buff 1992). To achieve the necessary sensitivity, QCM and SAW devices contain a special coating layer with adsorptive properties. This coating determines the selectivity of QCM and SAW sensors. Oscillations are applied to the sensor through a set of metallic electrodes formed on the piezoelectric surface, over which an adsorptive coating is deposited (Forster 1998). For example, Figure 1.14 shows that the formation of an acoustic wave by an AC voltage applied to a set of interdigited electrodes at one end of the device. SAW propagates toward the acoustic aperture of the crystal by distortion of the piezoelectric material under an electric field. When the wave arrives at the other end, a duplicate set of interdigited electrodes generates an AC signal as the acoustic wave passes underneath them. This AC signal can be monitored in terms of amplitude, frequency, and phase shift. Most SAW sensors were developed for continuous operation in situ for detection of volatile organic compounds (Slater and Paynter 1994; Grate et al. 1997; Ho et al. 2003). Special packages that encase

POLYMERS IN CHEMICAL SENSORS 41

Figure 1.14. Layout of a single acoustic aperture surface acoustic wave (SAW) device. (Reproduced with permission from Bai and Shi 2007. Copyright 2007 MDPI.)

the SAW device and integrated circuit board allow the sensor to be used in air, soil, and even water. These SAW devices operate well at ultrahigh frequencies and sense as little as 1 pg of material.

7.3.1. Requirements to Sensing Materials


QCM-, SAW-, and cantilever-based sensors are adsorption-type sensors, in which the change in weight of the sensing element is the determining factor (Houser et al. 2001). Therefore, the role of the sensing material in such devices is selectively and reversibly to sorb an analyte of interest from sampled air or liquid, and to concentrate it so that lower concentrations can be detected. Maximum and reversible sorption of specific analytes or classes of analytes, with rapid sorption kinetics and minimal sorption of interferents, are key aims in the development of a successful chemoselective coating for SAW-, QCM-, and cantilever-based sensors (Thompson and Stone 1997). This means that the selective coating should be such that it interacts well, either physically or chemically, with the chemical species to be sensed. Polymer films are normally chosen to coat the surface of QCM and SAW sensors. Phthalocyanine (Zhou et al. 1993), polymerceramic composites (Dias et al. 1993), epoxy resin (Ema et al. 1989), cellulose (Nakamoto et al. 1990), and many other polymers have been tested. As we have shown before, polymers can have different chemical properties, allowing sensing of a variety of organic chemical classes such as hydrocarbons, alcohols, ketones, and oxygenated, chlorinated, and nitrogenated compounds. Therefore, by choosing the proper polymer films with suitable active free functional groups, each chemical vapor of interest can be determined. Just this property of polymers is used in arrays of QCM or SAW devices for use as an electronic nose. Every sensor in the array has a different polymer film. Each film is chosen to have chemical absorption characteristics different from the others. The sensitivity and selectivity of an analyte for a polymer-based SAW or QCM sensor can be correlated from the relation of the solubility parameters of the polymer and the analyte (Gopel et al.

42 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

1998). The closer the solubility parameters of the polymer layer are to those of the analyte, the better the sensitivity and selectivity, because of the better affinity. For example, in a QCM device, a polybutyl methacrylate layer showed better sensitivity and selectivity for carbon tetrachloride, benzene, and chlorobenzene than for ethanol, and the response was maximal for carbon tetrachloride vapor (Koshets et al. 2003) (see Figure 1.15a). A copolymer of propylenebutyl film coated on a quartz resonator microbalance was used as a material for sensing harmful gases such as toluene, xylene, diethyl ether, chloroform, and acetone, the solubility parameters of which are close to those of the polymer (Nanto et al. 2000). The quartz resonator microbalance sensor coated with this copolymer exhibited high sensitivity and excellent selectivity for these harmful gases, especially for toluene and xylene gas, suggesting that

Figure 1.15. QCM sensitivity of sensors coated with (a) polybutyl methacrylate and (b) Polyvinyl formal ethylal toward various volatile organic vapors. (Reproduced with permission from Koshets et al. 2003. Copyright 2003 V. Lashkaryov Institute of Semiconductor Physics, National Academy of Sciences of Ukraine.)

POLYMERS IN CHEMICAL SENSORS 43

the solubility parameter is an effective parameter for use in functional design of the sensing membrane for quartz resonator gas sensors. The dipoledipole interaction between the polymer layer and the analyte can also be taken into account as a sensing mechanism for some analytes. For instance, polyvinyl formal ethylal copolymers show a high sensitivity toward alcohol but low sensitivity toward chlorine organic vapors (see Figure 1.15b) (Koshets et al. 2003).

7.3.2. Examples of QCM and SAW Polymer-Based Sensors


A QCM-type SO2 gas sensor was fabricated by Matsuguchi et al. (2001) with amino-functional poly (styrene-co-chloromethylstyrene) derivative on a quartz surface. They used three different diamine compounds, N,N-dimethyl ethylene diamine (DMEDA), N,N-dimethyl propane diamine (DMPDA), and N,N-dimethyl-p-phenylene diamine (DPEDA), to attach an amine group onto the copolymer backbone. The functioning of this sensor is based on the absorption of the basic amino group by SO2, which is a strong Lewis acid gas. Sensing characteristics are influenced by many factors, such as the mole fraction of chloromethyl styrene in the copolymers, the structure of the diamine compound attached, measurement temperature, and addition of organically modified siloxane oligomer. The sensor containing DPEDA functional copolymer showed the shortest response time (t100 = 11 min), with complete reversibility even at 50C. Figure 1.16 shows the response characteristics of a SO2 gas sensor using various amino-functional copolymers (DMEDA, DMPDA, DPEDA) measured for 50 ppm SO2 gas at 30C (Matsuguchi et al.

Figure 1.16. Response characteristics of sensor using aminofunctional copolymers measured for 50 ppm of SO2 at 30C; () DMEDA; () DMPDA; () DPEDA. (Reproduced with permission from Matsuguchi et al. 2001. Copyright 2001 Elsevier.)

44 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

2001). A thin film of 1,4-polybutadiene was used to construct a small and very sensitive (<10 ppb) ozone sensor (Fog and Rietz 1985). As in QCM sensors, the nature of the coating on the SAW sensor also determines the selectivity of the device; for example, LiNbO3 (Wohltjen and Dessy 1979) and fluoropolymers selectively sense pollutant organophosphorus gas (Dejous et al. 1995), and commercially available gas chromatographic phases as coatings sense toluene in dry air (Arn et al. 1991) with a response time of the order of 1 s. Micropatterned polymeric diffraction gratings were developed (Bailey and Hupp 2003) for sensing volatile organic chemicals, operating under nonresonant conditions. The sensor elements were found to respond in a rapid (response time 515 s) and reproducible fashion to each analyte investigated. The detection limits of micropatterned polyepichlorohydrin, polyisobutylene, and polybutadiene gratings were found, respectively, to be 8, 11, and 7 ppm for toluene; 25, 258, and 72 ppm for methyl ethyl ketone; 41, 102, and 34 ppm for chloroform; and 460, 60, and 59 ppm for hexane. While generally less than one order of magnitude higher than those observed for identical polymer/analyte combinations in SAW studies, the observed limits of detection were at or below governmental standards for each analyte evaluated with these polymeric gratings. SAW sensors using three different kinds of polymers to detect selectively three different volatiles, dimethyl methylphosphonate, CH3CN, and CH2Cl2, have been reported by Joo et al. (2005). In an interdigital transducer (IDT) SAW device the polymers used as the sensing material were polyisobutylene (PIB), polyepichlorohydrin (PECH), and polydimethylsiloxane (PDMS) on an aluminum substrate. The thin films were coated on quartz substrate by spin coating.

Figure 1.17. Calibration curves for frequency shift using polystyrene-coated quartz crystal electrode of benzene (), toluene (), ethylbenzene (), and xylene (). (Reproduced with permission from Mirmohseni and Rostamizadeh 2006. Copyright 2006 MDPI.)

POLYMERS IN CHEMICAL SENSORS 45

Koshets et al. (2003) reported a quartz crystal microbalance sensor coated with polybutyl methacrylate (PBMA) and polyvinyl formal ethylal (PVFE). It was found that the PBMA filmcoated sensor showed good sensitivity as well as selectivity for chlorobenzene, although it also responded to chloroform, chloromethylene, and toluene. However, the film did not show any significant response for alcohols. On the other hand, the responses of the PVFE filmcoated device show high sensitivity toward alcohol but low sensitivity toward chlorine organic vapors. A QCM sensor was developed (Mirmohseni and Hassanzadeh 2001, 2006) for the detection of volatile BTEX vapors (benzene, toluene, ethylbenzene, and xylenes). The adsorption of BTEX organic vapors on a thin layer of polydimethylsiloxane (PDMS) (Mirmohseni and Hassanzadeh 2001) or polystyrene (Mirmohseni and Rostamizadeh 2006) coated on AT-cut quartz crystals with gold electrodes was studied. Calibration graphs were constructed for the frequency shifts due to the sorption of BTEX compounds at the ppm level. As an example, calibration curves for the frequency shift using a polystyrene-coated quartz crystal electrode of benzene, toluene, ethylbenzene, and xylene is shown in Figure 1.17.

7.4. ELECTROCHEMICAL POLYMER-BASED SENSORS


In sensor technology, the electrochemical transducer principle is the oldest and most widely used strategy. Generally, electrochemical methods are based on the transformation of chemical information (chemical reaction) into an analytically useful signal. As a result of electron transfer, the change of electrical properties of the detection element can be measured with the help of a well-defined electrical circuit by various modes of measurements, e.g., voltammetric (measurement of voltage change), impediometric (measurement of impedance change), conductometric (measurement of conductivity or resistance change), amperometric (measurement of current change), or potentiometric (measurement of potential change) (Bakker 2004). The last two techniques are the most utilized ones. In potentiometric sensors, a local equilibrium is established at the sensor interface, and information about the composition of a sample is obtained from the potential difference between two electrodes. Amperometric sensors exploit the use of a potential applied between a reference and a working electrode, to cause the oxidation or reduction of an electroactive species; the resulting current is measured. The most widely used potentiometric device is the pH electrode, which has been used for several decades. More detailed description of these methods can be found in reviews (e.g., Bakker 2004) and in other chapters of this series. Any sensor used in electroanalytical determinations contains two basic functional units: a receptor, which transforms the chemical information into a form of energy; and a transducer, which transforms the energy, bearing chemical information, into a useful signal. Given the multifunctionality and the large variety of possible structures, polymers in electrochemical sensors can act as elements with different functions, i.e., as solid electrolytes, membranes, or solid contacts (Bobacka 2006).

7.4.1. Ion-Selective Electrodes


Potentiometric ion sensors (Uhlmann et al. 1998), a subgroup of electrochemical sensors (Bakker 2004), are attractive for practical applications because they are characterized by small size, portability, low

46 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

energy consumption, and low cost compared to many other analytical techniques. Typical constructions of ion-selective electrodes (ISEs) are shown in Figure 1.18. Ion-selective electrodes based on polymeric membranes containing neutral or charged carriers (ionophores) are among the most successful electrochemical sensors in routine use today. In the classic configuration, the electrodes are usually based on porous membranes, consisting of permselective membrane for the comparable size of the ions which separate the sample from the inside of the electrode. When such systems come in contact with analytes to be sensed, some ionic exchange/interaction occurs. Due to this interaction and the movement of the selected ions, a potential difference is generated, which is measured in the sensor device. So, the ISE is an indicator electrode capable of selectively measuring the activity of a particular ionic species. ISEs are suitable for determination of specific ions in a solu-

Figure 1.18. Construction principles for various ion-selective electrodes: (a) conventional ISE with an internal reference electrode and internal lling solution; (b) coated-wire ISE; (c) ISE with a hydrogel contact; (d) ISE with a conducting polymer contact; (e) ISE with a conducting polymer dissolved in the ion-selective membrane; (f) ISE with a (functionalized) conducting polymer as sensing membrane. 1, electronic conductor; 2, internal reference electrode; 3, inner lling solution; 4, ionselective membrane; 5, hydrogel; 6, electronic conductor with a high work function; 7, conducting polymer; 8, ion-selective membrane containing a conducting polymer; 9, conducting polymer containing ion-recognition sites (ionophores). (Reprinted with permission from Bobacka 2006. Copyright 2006 Wiley-VCH Verlag GmbH&Co.)

POLYMERS IN CHEMICAL SENSORS 47

tion in the presence of other ions. The quantitative analysis of ions in solutions by ISEs is a widely used analytical method. Ion sensors find wide application in medical, environmental, and industrial analysis (Bakker and Meyerhoff 2000). According to Bobacka (2006), conducting polymers are useful for application in solid-state ISEs for several reasons: 1. Conducting polymers are electronically conducting materials that can form an ohmic contact to materials with high work function, such as carbon, gold, and platinum. 2. Conducting polymers can be conveniently electrodeposited on the electronic conductor by electrochemical polymerization of a large variety of monomers. 3. Alternatively, several conducting polymers are soluble and can therefore be deposited from solution. 4. Conducting polymers are electroactive materials with mixed electronic and ionic conductivity, which means that they can transduce an ionic signal into an electronic one in the solid state. These multifunctional properties of conducting polymers are excellent for ion-to-electron transduction in the solid state. Ion selectivity can be introduced in two different ways (Bobacka 2006). One possibility is to use the conducting polymer only as an ion-to-electron transducer (solid contact) in combination with a classical ion-selective membrane (Cadogan et al. 1992). In these solid-contact ISEs (see Figure 1.18d), the ion selectivity is determined mainly by the ion-selective membrane, and this methodology allows full utilization of a large number of existing ionophore-based polymeric membrane formulations (Uhlmann et al. 1998). The operating principle is shown schematically in Figure 1.19a. By using a conducting polymer that is soluble in the same solvent as the components of the ion-selective membrane, it is possible to construct a so-called single-piece ISE, in which the conducting polymer, acting as an ion-to-electron transducer, is integrated into the ion-selective membrane (Figure 1.18e). Another possibility for inducing selectivity is to incorporate the ion-recognition sites directly into the conducting polymer matrix, e.g., by doping the conjugated polymer with counter ions containing ion-complexing groups or by covalent binding of ion-recognition sites to the conjugated polymer chain (Figure 1.18f ) (Garnier 1989). The operating principle is shown schematically in Figure 1.19b. This approach requires precise control of both the electronic and the ionic transport properties of the membrane in order to optimize the ion-recognition and transduction processes (Bobacka 2006). The main challenge is to obtain a selective ionic response while minimizing redox interference. However, the possible benefits are huge, because covalent binding of ion-recognition sites to the conducting polymer backbone allows integration of the ion-recognition sites and ion-to-electron transducer even within the same macromolecule. This approach is still in an early stage of development, but one can expect that it will become of great importance for the construction of durable micro- and nano-sized ion sensors in the future (Bobacka 2006). Polymers that are usually applied for design of ISEs are listed in Tables 1.5 and 1.6. As follows from previous discussion, membrane materials for ISEs must possess adequate porosity. Initially, porous polymeric ISE membranes were obtained by soaking a viscous solution of a water-

48 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

(a)

(b)

Figure 1.19. Operating principles of (a) solid-contact ISE based on an oxidized (p-doped) conducting polymer as ion-to-electron transducer and (b) solid-state ISE based on an oxidized (p-doped) conducting polymer as sensing membrane. EC, electronic conductor; CP, conducting polymer; ISM, ion-selective membrane; S, solution; e, electron; +, hole (oxidized CP); L, ion-recognition site (mobile/xed). (Reproduced with permission from Bobacka 2006. Copyright 2006 Wiley-VCH Verlag GmbH&Co.)

immiscible, nonvolatile polymer containing the dissolved ionophore. Recently, homogeneous polymer membrane matrices having ionic functional groups have replaced these materials. For preparation of a flexible sensing membrane, the polymer is mixed with a compatible plasticizer. In such membranes, the polymers must have some special physical properties, e.g., elasticity and mechanical stability.

7.4.2. Electrochemical Polymer-Based Gas Sensors


One approach to designing a room-temperature gas sensor is to use solid polymer electrolytes (SPEs). Solid polymer electrolytes became important during the mid-1970s because of the inefficiencies and maintenance requirements of liquid electrolytes then used in amperometric and potentiometric electrochemical sensors (Korotcenkov et al. 2009). The simplest setup for these technologies is shown in Figure 1.20. Originally, the SPE was a solid plastic sheet of perfluorinated sulfocationite polymer that, when saturated with water, became an excellent ionic conductor. Ionic polymers in contact with a conductive medium such as a metal allow electrochemical reactions at the interface. (It is necessary to note that SPEs are not electronic conductors.)

POLYMERS IN CHEMICAL SENSORS 49

Table 1.5. Solid-state ion-selective electrodes using polymer-based materials as solid contact (ion-to-electron transducer) SOLID CONTACT SUBSTRATE Glassy carbon Graphite Pt Pt Ag/AgCl + Pt Glassy carbon Au EPD + graphite PEDOT Glassy carbon Ag/AgCl + Pt Au or Pt Au Au Pt Glassy carbon Pt Pt
MATERIAL

ION-SELECTIVE
MEMBRANE

PRIMARY ION Cl; Ca2+; K+ Cl K+; pH pH CO32 N-MePy+; Bupivacaine; Ag+; K+ K+ K+ K+; Ca2+; Cu2+ Ca2+ CO32 Cl Ca2+; Pb2+ Cl pH K+ Dimedrol Chlordiazepoxide

PPy PPy PPy PPy (+ Nafion) PHDP PEDOT PEDOT PEDOT PEDOT PMT POT POT PVC POT PANI PANI PANI PNA PAP

PVC PVC PVC Glass SR PVC PVC PVC PVC PVC SR PUR MMA/DMA PVC , PUR PVC PVC PVC PVC

PPy = polypyrrole; PVC = ion-selective membrane based on plasticized poly(vinyl chloride); PHDP = poly(1-hexyl-3,4-dimethylpyrrole); SR = ion-selective membrane based on plasticized silicone rubber; PEDOT = poly(3,4-ethylenedioxythiophene); N-MePy+ = N-methyl pyridinium; bupivacaine = 1-butyl-N-[2,6-dimethylphenyl]2-piperidinecarboxamide; EPD = poly(ethylene-co-propylene-co-5-methylene-2norbornene); PMT = poly(3-methylthiophene); POT = poly(3-octylthiophene); PUR = ion-selective membrane based on plasticized polyurethane (Tecoflex); MMA/DMA = ion-selective membrane based on poly(methyl methacrylate)/poly(decyl methacrylate) copolymer; PANI = polyaniline; PNA = poly(-naphthylamine); PAP = poly(oaminophenol). Source: Data from Bobacka 2008.

Nafion, a typical solid polymer electrolyte, is a hydrated copolymer of polytetrafluoroethylene (PTFE) and polysulfonyl fluoride vinyl ether containing pendant sulfonic acid groups. It is a cation exchanger that contains hydrophilic SO3 radicals firmly bound to the hydrocarbon backbone, whose charge is compensated by counter ions (mostly H+). The counter ions are dissociated and solubilized by water present within the polymer structure and give rise to the ionic (proton) conductivity of the polymer. The water

50 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Table 1.6. Examples of solid-state ion-selective electrodes using conducting polymers as ion-selective membranes CONDUCTING
POLYMER

SUBSTRATE Pt, glassy carbon Pt, glassy carbon Graphite Pt Pt Graphite-epoxy Pencil lead Glassy carbon Glassy carbon Glassy carbon

DOPANT Arsenazo-I Calcon; SSA; Tiron; Calcion; ATP Heparin Co-bis(dicarb) HCO3 TPB DS NO3 ClO4 Fe(CN)63/4 Sulfonated calixarenes Heparin PSS Cl; Fe(CN)63/4 Sulfonated calixarenes; methylsulfonated resorcarenes Cl SO42 Cl DS DBS ClO4

PRIMARY ION Ca2+; Mg2+; Cu2+ Ca2+; Mg2+ pH pH Zn2+ DS NO3 ClO4 K+ Ag+

PPy

PEDOT

Pt, glassy carbon Glassy carbon Glassy carbon Glassy carbon

Ca2+; Mg2+ K+ Cl Ag+

PANI

Carbon fiber Au Glassy carbon Pt Pt Pt

pH pH pH DS DBS Aniline

PPy = polypyrrole; TPB = tetraphenylborate; Arsenazo-I = 2-(o-arsenophenylazo)-1,8dihydroxynaphtalene-3,6-disulfonic sodium salt; Calcon =[1-(1-hydroxy-2-naphthylazo)-2naphtol-4-sulfonic]sodium salt; SSA = 1-hydroxy-4-sulfobenzoic acid; Tiron = 1,2-dihydroxybenzene-3,5-disulfonic disodium salt; Calcion (or calcichrome) = C3OH14O22N4S6Na5; heparin = highly sulfonated glycosaminoglycan constituted by disaccharic repeating units of D-glucosamine and L-iduronic acid; co-bis(dicarb)jcobaltbis(dicarbollide) [3,3-Co(1,2C2B9H11)]; DS = dodecylsulfate; PEDOT = poly(3,4-ethylenedioxythiophene); DBS = dodecylbenzene sulfonate. Source: Data from Bobacka 2008. required for proton solubility is bound in the hydration mantles of the ions present, so the polymer is a solid which contains no macroscopic liquid phase unless excess water is present (Opekar and Stulik 1999). Nafion has good proton conductivity, high gas permeability, outstanding chemical stability, and good mechanical strength, and it has been widely used in fuel-cell and sensor applications.

POLYMERS IN CHEMICAL SENSORS 51

However, the geometric dimensions of Nafion and its electrical properties (primarily its ionic conductivity) are strongly dependent on the amount of water in the polymer. The maximum water content, corresponding to 22 water molecules per single sulfo group of the polymer, is attained by boiling Nafion in water, and this number decreases to 14 for the polymer in contact with a gaseous phase saturated with water vapor; the water content fluctuates with the relative humidity (RH) of the surrounding medium (Zawodzinski et al. 1993). In general, perfluorinated polymer membranes show high ionic conductivities at high water vapor pressure (Anantaraman and Gardner 1996). Nafion electrolyte gas sensors, because Nafion conductivity is a function of RH, typically produce a gas response that depends on the RH (Yan and Liu 1993; Samec et al. 1995). This RH response is not desired for an ambient-air sensor, because the RH can change over wide limits and so it is typically either eliminated or compensated for. Nafion and polymer electrolytes such as sulfonated polybenzimidazole (PBI), sulfonated polyether ether ketone (S-PEEK) (Bouchet et al. 2002; Sundmacher et al. 2005), and PVA/H3PO4 (Ramesh et al. 2003) can be used in H2 sensors. Some of these solid polymer electrolytes have excellent mechanical and thermal properties and good protonic conductivity even in dry atmospheres (Rosini and Siebert 2005). The remarkable properties of these polymers lie in the combination of the high hydrophobicity of the perfluorinated polymer backbone and the high hydrophilicity of the sulfonic acid branch. The hydrophilic branches act as a plasticizer, and the backbone retains strong mechanical properties (Colomban 1999). More detailed information about polymers used in electrochemical gas sensors can be found in various reports in the literature (Colomban 1999; Maksymiuk 2006; Korotcenkov et al. 2009). In some cases, hydrogels or an electrolyte inside a porous matrix is used instead of a free liquid electrolyte in order to raise the viscosity, lower evaporation rates, and resist leakage of the electrolyte from sensor devices. The polymers or hydrogels can prevent the evaporation of electrolyte during sensor fabrication, especially in microsensor devices in which very small amounts of electrolyte are used. Using polymer electrolytes provides opportunities for the design of planar sensors and the applications of standard microelectronic fabrication technologies (Shi andAnson 1996). Polymers also allow decreasing both the size and weight of electrochemical sensors. In addition, polymer electrolytes allow a larger range of operating temperatures for the electrochemical sensor. Polymer-based gas sensors can operate successfully in the range from room temperature to approximately 100C (Sakthivel and Weppner 2006). Further, compared to liquid electrolytes, solid polymer electrolytes can be used as separators in electrochemical cells, do not dissolve impurities from the gas as easily, and permit the construction of miniaturized devices that are leakproof to help avoid premature sensor failure.

Polymer-coated sensing electrode Reference electrode Counter electrode Electrolyte


Figure 1.20. Electrode setup for an electrochemical sensor.

52 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

7.4.3. Examples of Polymer-Based Electrochemical Sensors


7.4.3.1. ION-SELECTIVE ELECTRODES 7.4.3.1.1. Polyvinyl ChlorideBased Sensors
Polyvinyl chloride (PVC) has a long history as a polymeric matrix in various electrochemical sensors. Ion-selective electrodes based on plasticized PVC with valinomycin (an ionophore) as neutral carrier are now widely used (Janata and Bezegh 1988; Armstrong and Horvai 1990). The ionophore forms a complex with the primary ion being sensed rather than with any other similar ions of the membrane. Ionophores, in most cases small molecules, must be lipophilic to remain in the polymer phase without entering into aqueous phase in contact. However, research has shown that PVC sometimes shows deviation due to leaching of plasticizer and ionophores. Therefore, considerable efforts have been made to fabricate electrodes with chemically modified PVC. A PVC-based membrane electrode was prepared by Shamsipur et al. (2001) using 3,4-di[2-(2-tetrahydro-2H-pyranoxy)]ethoxy styrenestyrene copolymer as ionophore to successfully determine beryllium in a mineral sample. The membrane was composed of oleic acid (OA) and sodium tetraphenylborate (STB) as anionic additives, and dibutyl phthalate (DBP), dioctyl phthalate (DOP), acetophenone (AP), and nitrobenzene (NB) as plasticizing solvent mediators. A membrane made of PVC:NB:I:OA in the ratio 3%:55%:10%:5% gave the best performance. Liu et al. (2000) reported the selective determination of silver ions in electroplating wastewater using PVC membrane electrodes with 5% bis(diethyldithiophosphate) as ionophore and 65% 2-nitrophenyl octyl ether (o-NPOE) as plasticizer. A 7-ethylthio-4-oxa-3-phenyl-2-thioxa-1,2-dihydropyrimido[4,5-d]pyrimidine (ETPTP) ionophore incorporating a PVC matrix membrane (Saleh et al. 2001) showed good potentiometric response for Al3+ over a wide concentration range (105101 M). The sensor provided a stable response with a slope of 19.5 mV per decade for at least 1 month, and was also selective for Al3+ in the presence of alkali, alkaline earth, transition, and heavy metal ions with minimal interference from Hg2+ and Pb2+. Zhang et al. (2000) prepared a PVC membrane electrode based on chloro[tetra(m-aminophenyl) porphinato]-manganese [T(m-NH2)PPMnCl] and 2-nitrophenyl octyl ether (o-NPOE) in the composition 3:65:32 [T(m-NH2)PPMnCl:o-NPOE:PVC] to determine fluoroborate in electroplating solution. A Hg(II) ion-selective PVC membrane sensor based on ethyl-2-benzoyl-2-phenylcarbamoyl acetate (EBPCA) was developed by Hassan et al. (2000), which showed selectivity for Hg(II) ion in comparison with alkali, alkaline earth, transition, and heavy metal ions. The sensor was applied for the determination of Hg(II) content in some amalgam alloys. High selectivity for Ag+ ions over Na+, K+, Ca2+, Sr2+, Pb2+, and Hg2+ was observed with a PVC membrane containing bis-pyridine tetramide macrocycle (Mahajan and Parkash 2000). Hassan et al. (2001) described a novel uranyl ionresponsive sensor based on a uranyl PVC matrix membrane containing tris(2-ethylhexyl) phosphate (TEHP) as both the electroactive material and plasticizer, and sodium tetraphenylborate (NaTPB) as an ion discriminator. The sensor displayed a rapid and linear response for UO22+ ions over a wide concentration range in a pH range of 2.83.6 with a life span

POLYMERS IN CHEMICAL SENSORS 53

of 4 weeks. They also reported another PVC-based uranyl sensor containing o-(1,2-dihydro-2-oxo-1pyridyl)-N,N,N,N-bis(tetramethylene) uranium hexafluorophosphate (TPTU) as a sensing material, sodium tetraphenyl borate as an ion discriminator, and dioctyl phenylphosphonate (DOPP) as a plasticizer. Linear and stable response was obtained with the sensor in a working pH range of 2.53.5 and over a life span of 6 weeks. Gorton and Fiedler (1977) developed a zinc-sensitive polymeric membrane electrode to detect the end point during potentiometric titration of Zn2+ against ethylenediamminetetraacetic acid (EDTA). The membrane, having (by weight) 8% ligand [zinc salt of di-n-octylphenylphosphoric acid (HDOPP)], 62% solvent (di-octylphenylphosphonate (DOPP-n), and 30% polymer (PVC) composition, was found to have at least a 3-month lifetime. A potentiometric method was described by Abbas et al. (2000) for the determination of cetylpyridinium (CP) cation using a PVC powder membrane sensor as an end-point indicator electrode in potentiometric titration of some anions, which was applied for the determination of anionic surfactants in some commercial detergents and waste water. Mousavi et al. (2000) constructed a PVC membrane nickel(II) ISE using 1,10-dibenzyl-1,10-diaza-18-crown-6 (DbzDA18C6) as a neutral carrier that exhibited relatively good selectivity for Ni(II) over a wide variety of other metal ions, and could be used in a pH range of 4.08.0. Wroblewski et al. (2000) investigated the influence of the membrane components on phosphate selectivity by anion-selective PVC plasticized membranes containing uranyl salophene derivatives. The highest H2PO4 selectivity over other anions tested was obtained for lipophilic uranyl salophene III in PVC/o-nitrophenyl octylether membrane containing 20 mol% of tetradecylammonium bromide.

7.4.3.1.2. Other Polymers


Other polymers besides PVC can also be used to design effective ion-selective electrodes. For example, a calcium ion-sensitive electrochemical sensor device (Artigas et al. 2001) was fabricated with a photocurable polymer membrane based on aliphatic diacrylated polyurethane instead of PVC to measure calcium activity in water samples extracted from agricultural soils. The authors claimed their results to be well correlated with those obtained by standard methods. Torres et al. (2001) developed five different types of membranes by solubilizing poly(ethyleneco-vinyl-acetate) copolymer (EVA) and tri-caprylyl-trimethyl-ammonium chloride (Aliquat-336S) in chloroform without using any plasticizer, followed by film casting, for the detection of iodide, periodate, perchlorate, salicylate, and nitrate ions. The membrane performance was in the concentration range 105101 mol L1 at steady state. Gupta and Mujawamariya (2000) found a significant dependence of sensitivity, working range, response time, and metal-ion interference on the concentration of ionophore, plasticizer, and molecular weight of cyanocopolymers for a Cd(II) ISE based on a cyanocopolymer matrix and 8-hydroxyquinoline as ionophore. The cyano groups of the copolymers contributed significantly to enhance the Cd(II) selectivity of the electrode for Cd2+ ions in the presence of alkali and alkaline-earth metal ions in the pH range 2.56.5. The selectivity coefficients for cesium ion over alkali, alkaline-earth, and ammonium ions were also determined.

54 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Polysiloxane membranes and membranes with polyHEMA hydrogel intermediate film were designed by Sudhlter et al. (1989). A modified silicone rubberbased membrane was used for detection of Na+ in body fluids (Tsujimura et al. 1995). Teixeira et al. (2000) used a MnO2-based graphite epoxy electrode for determination of lithium ions by the potentiometric method. In that study the best potentiometric response of 30-s response time and 6 months lifetime was obtained for an electrode composition of 35% -MnO2, 15% graphite, and 50% epoxy resin. A lipophilic acrylate resin as a matrix for the sensing membrane with long-term stability was developed by Numata et al. (2001) to determine water hardness by analyzing Ca2+ and Mg2+ with equal selectivity. The performance of the electrode was maintained during a test of the electrode conducted in tap water at a continuous flow rate of 4 ml min1 with long-term stability of the electrode for 1 year due to the strong affinity of 1-decylalcohol for the lipophilic acrylate resin. A new Ca2+-selective polyaniline (PANI)-based membrane was also developed (Lindfors and Ivaska 2001) for all-solid-state sensor applications. The membrane contained electrically conducting PANI as the matrix polymer, bis[4-(1,1,3,3-tetramethylbutyl) phenyl] phosphoric acid (DTMBP-PO4H), dioctyl phenylphosphonate (DOPP), and cationic (tridodecylmethylammonium chloride, TDMACl) or anionic [potassium tetrakis(4-chlorophenyl) borate, KTpClPB] as lipophilic additives.

7.4.3.2. pH SENSORS
Since the solution pH has a significant effect on chemical reactions, the measurement and control of pH is very important in chemistry, biochemistry, clinical chemistry, and environmental science. The pH, of course, indicates the amount of hydrogen ions in a solution. A pH-sensitive layer was obtained by reacting aminoethylcellulose fibers with 1-hydroxy-pyrene-3,6,8-trisulfochloride, followed by attachment of the sensitive layer to the surface of a polyester foil, and embedding the composite in an ion-permeable polyurethane (PU)based hydrogel material. Hydrogen ion-selective solid-contact electrodes based on N,N-dialkylbenzylethylenediamine (alkyl = butyl, hexyl, octyl, decyl) were also prepared. Solid-contact electrodes and coated-wire electrodes were fabricated from polymer cocktail solutions based on N,N-dialkylbenzylethylenediamine (alkyl = butyl, hexyl, octyl, decyl). The response ranges and slopes were influenced by the alkyl chain length. The solid-contact electrodes showed linear selectivity to hydrogen ion in the pH ranges 4.513.0, 4.213.1, 3.4 13.0, and 3.013.2. Stability was also improved in comparison to coated-wire electrodes (Han et al. 2001). Pandey et al. (2000) developed a solid-state poly(3-cyclohexyl)thiophene-treated electrode as a pH sensor and, subsequently, a urea sensor. Later, Pandey and Singh (2001) reported the pH-sensing function of a polymer-modified electrode (a novel pH sensor) in both aqueous and nonaqueous media. The sensor was prepared with a polymer-modified electrode by electrochemical deposition of PANI in dry acetonitrile containing 0.5 M tetraphenyl borate at 2.0 V versus Ag/AgCl. Han et al. (2001) have shown that N,N-didecylbenzylethylenediaminebased solid-contact electrodes can also be successfully used for design selectivity and a reproducible potentiometric pH sensor.

POLYMERS IN CHEMICAL SENSORS 55

7.4.3.3. ELECTROCHEMICAL GAS SENSORS 7.4.3.3.1. Nitric Oxide Sensors


Ichimori et al. (1994) introduced a commercial amperometric NO-selective electrode. The Pt/Ir (0.2) electrode was modified with a NO-selective nitrocellulose membrane and a silicone rubber outer layer. The electrode showed linear response at nanomolar concentration with a time constant of 1.5 s. A threefold increase in sensitivity was achieved by raising the temperature from 26C to the physiological value of 37C. In an in vivo application, the electrode was used for the measurement of NO in rat aortic rings under acetylcholine stimulation. Friedemann et al. (1996) used a carbon-fiber electrode modified with an electrodeposited ophenylenediamine coating. A Nafion underlayer with this modified electrode provided good sensitivity to NO, and a three-layer overcoat of the Nafion optimized selectivity against nitrite. Using a polystyrene film, Christensen et al. (1993) developed a NO2-sensing device. Upon exposing the film to a 1:10 v/v mixture of NO2/N2, the conductivity of the film increased irreversibly and rapidly by several orders of magnitude. It was believed that the increase in conductivity of the film was due to self-ionization of N2O4 to NO+NO3. Ho and Hung (2001) prepared a Pt/Nafion electrodebased amperometric gas sensor for NO2 detection with a concentration range from 0 to 485 ppm. Reticulated vitreous carbon (RVC) (Hrncirova et al. 2000) was used as the indicator electrode in solid-state gas sensors. A typical sensor contained a RVC indicator, a platinum auxiliary, and a Pt/air reference electrode, with a SPE of 10% PVC, 3% tetrabutylammonium hexafluorophosphate (TBAHFP), and 87% 2-nitrophenyloctyl ether (NPOE). The gaseous nitrogen dioxide analyte in air was monitored by reduction at 500 mV versus a Pt/air electrode. The RVC was found to successfully replace noble metals in solid-state gas sensors. With hydrophobic SPEs, the sensitivity decreases with increasing humidity, while with hydrophilic ones (e.g., Nafion), it increases. Due to the extraordinary chemical inertness, the signal stability of RVC is affected not only for detection in solution, but also in sensors in which RVC remains in contact with a SPE. Mizutani et al. (2001) fabricated amperometric gas sensors for the measurement of dissolved oxygen and nitric oxide at lower concentrations using a permselective polydimethylsiloxane (PDMS) membrane at room temperature.

7.4.3.3.2. Other Gases


Ammonia and carbon dioxide gas-sensing systems in both static and continuous flow conditions were prepared using polymer-membrane pH electrodes as internal sensing elements (Opdycke et al. 1983). The pH-responsive polymer membranes were prepared with plasticized PVC by incorporating tridodecylamine as the neutral carrier. It was suggested that for miniature static gas sensors, the internal polymer pH electrode could be made with or without an internal reference solution. In the latter case, the polymeric membrane was coated directly onto a graphite substrate. Response times and reproducibility of these new gas-sensing systems were evaluated using optimized internal electrolytes, flow rates, and gas-permeable membrane materials.

56 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Opekar and Bruckenstein (1984) determined gaseous hydrogen sulfide by cathodic stripping voltammetry after preconcentration on a silver metalized porous Teflon membrane electrode. The reproducibility of the determination, expressed in terms of the relative standard deviation, was 3.2%. SPE ion-exchange membranesupported porous silver electrodes were used in an electrochemical sensor for the detection of trace hydrogen sulfide in gaseous samples (Schiavon et al. 1995). The high sensitivity and fast response of this sensor was due to the elimination of oxygen interference. One side of this sensor faced the sample and the other side of the membrane faced an internal electrolyte solution containing the counter and reference electrodes. The sensor functioned by electroanalysis of H2S by amperometric monitoring, cathodic stripping measurements, and flow injection analysis. Shi et al. (2001) reported an electrochemical sensor for the detection of sulfur dioxide in both gas and solution. The sensor was constructed by chemical modification of the electrode by polymerizing 4-vinyl pyridine (4-VP) onto a palladium and iridium oxide (PVP/Pd/IrO2)coated platinum microelectrode, which exhibited excellent catalytic activity toward sulfite with an oxidation potential of +0.50 V. In this SO2 gas sensor, the PVP/Pd/IrO2-modified electrode functioned as the detecting electrode, a Ag/AgCl electrode as reference electrode, Pt as counter electrode, and a porous film in direct contact with the gas-containing atmosphere. The sensor was found to have high current sensitivity, a short response time, and good reproducibility for the detection of SO2, and it showed good potential for use in the field of environmental monitoring and control. Otagawa et al. (1990) fabricated a planar miniaturized electrochemical CO sensor that was comprised of three Pt electrodes (sensing, counter, and reference) and a solution-cast Nafion as solid polymer

Figure 1.21. Induced voltage signals at different ethanol concentrations using Naon membrane on Pt/C electrode. (Reproduced with permission from Kim et al. 2000. Copyright 2000 Elsevier.)

POLYMERS IN CHEMICAL SENSORS 57

electrolyte. The sensor showed a linear response to the CO concentration in air with a sensitivity of 8 Pa/ppm and 70 s response time. Methane gas was determined via pre-adsorption on a dispersed Pt electrode backed by a SPE membrane (Nafion) in contact with 10 M sulfuric acid (Jacquinot et al. 2001). The adsorption process was strongly temperature-dependent, with an activation energy of 8.7 kcal mol1. The measurement of other hydrocarbons, viz., ethane, propane, and butane, was also shown to be possible using this sensor, with a significant reduction of cross-sensitivity to carbon monoxide and hydrogen by means of suitable chemical adsorption filters. Fuel cells using a polymer electrolyte membrane were fabricated and tested by Kim et al. (2000) for the measurement of ethanol gas concentration. The polymer electrolyte (Nafion 115 membrane) and 10% Pt/C sheets with 0.5 mg/cm2 Pt loading were used as catalyst electrodes. In Figure 1.21, the voltage responses of the fuel cell for various ethanol concentrations are shown at a load resistance of 100 . Kim et al. (2000) correlated the ethanol concentration with the peak height of the signals instead of the area because the shape of voltage signals is the same even at different concentrations (see Figure 1.21). Analytical performance of a potentiometric pCO2 (partial pressure of CO2) sensing catheter was reported by Opdycke and Meyerhoff (1986). The sensing catheter consisted of an inner tubular PVC pH electrode in conjunction with an outer gas-permeable silicone rubber tube. The results of conventional blood-gas instruments correlated well with the continuous pCO2 values obtained by the sensor during 6 h of in vitro blood pump studies. Based on the preliminary results of study with this sensor, the suitability of this catheter was demonstrated by intravascular implantation in a dog for continuous in vivo monitoring of pCO2.

7.5. CHEMICALLY SENSITIVE FET-BASED SENSORS


As we have shown, conductometric polymer-based sensors possess good sensitivity to various gases and organic vapors, but they can suffer from temperature and humidity dependence. Work using electrochemically deposited polypyrrole as a field-effect transistor (FET) gate material has demonstrated devices which can possess better operating characteristics. Such a combination of technologies could lead to a new generation of low-cost, low-power gas sensors for hand-held monitors (Persaud 2005). Polymers used in chemically sensitive FET (CHEMFET)based sensors are listed in Table 1.7.

7.5.1. Insulated-Gate FETs


Chemically sensitive insulated-gate field-effect transistors (IGFETs) are a class of sensors which contain a polymer gate as detecting element on a silicon base (Janata and Josowicz 2003). In IGFETs, the gate contact is on the surface of the device and is the sensing part that is exposed to the analyte. The properties of the polymer determine the selectivity of such devices. The buried inorganic semiconductor, where the field-effect transport takes place, normally is not exposed to the analyte. A simplified IGFET setup is shown in Figure 1.22 (Reese et al. 2004).

58 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Table 1.7. Polymers used in CHEMFETs


POLYMER Polythiophene and its derivatives RECOGNITION MECHANISM Alkyl or alkoxy side chains CHEMICAL ANALYTE Ammonia, water vapor, chloroform, alcohols, ketones, thiols, nitriles, esters, ring compounds, alkanes Volatile chiral molecules H+ H+ H , IrCl6, O2, H2
+

REFERENCES Assadi et al. 1990; Ohmori et al. 1991; Crone et al. 2001; Torsi et al. 2003a, 2003b Tanese et al. 2004 Bartic et al. 2002, 2003; Loi et al. 2005 Nishizawa et al. 1992 Thackeray et al. 1985; Thackeray and Wrighton 1986 Paul et al. 1985; Chao and Wrighton 1987; Hoa et al. 1992; Dabke et al. 1997; Nilsson et al. 2002

Poly(phenylene ethynylene) Poly(3-hexylthiophene) Polypyrrole Poly(3-methylthiophene)

Enantioselective pendant groups Proton-sensitive dielectric layer Membrane Platinum particles

Polyaniline

Moisture-sensitive solid-state electrolyte, crown ether

H+, Ru(NH3)63+/2+, Fe(CN)6, water vapor, SO2, K+

Source: Data from Mabeck and Malliaras 2006.

The IGFET consists of a pair of electrodes forming contacts (top or bottom contact) to the conducting polymer, deposited on an insulating substrate; under a constant current, the resulting potential difference at the electrodes becomes the response signal. By interacting with the analyte, the detecting element changes the semiconductor work functions (M S) of the gate system. The transduction principle is based on the linear relationship between the transistor threshold voltage (VtIGFET) and the gate-metal/semiconductor work-function offset (Bergveld 2003): Vt IGFET (f M - fS ) The sensitivity and selectivity of the device depends on the nature of the gate materials and the operating temperature (usually between 50 and 200C). Incorporation of small analyte receptor molecules into the polymeric semiconductor layer by chemical and physical methods causes a change in the gate of IGFETs. Several intrinsically conducting polymers (ICPs) have been successfully employed as IGFET gate layers (Janata and Josowicz 1998). For example, PANI-based FET devices have been demonstrated by Potje-Kamloth et al. (2002). The morphology of the polymeric semiconductor layer is important to the sensing behavior, so it is necessary to find an optimum semiconductor morphology that will interact well with the analyte.

POLYMERS IN CHEMICAL SENSORS 59

Figure 1.22. Schematic presentation of a simplied CHEMFET. (Reproduced with permission from Reese et al. 2004. Copyright 2004 Elsevier.)

A wide variation in the work function of ICPs is possible due to the tailorability and porous morphology of ICPs, which helps easy permeation of analyte molecules and alters the work function of the polymer/gate dielectric system. One of the major advantages of ICP-based IGFET sensors is the ability to detect changes in voltage (Vt) by passing a negligible current. It has also been established that ICP-based IGFETs are more stable than chemiresistors in terms of signal-to-noise ratio and very high detection limits (ppm level) (Janata and Josowicz 2003).

7.5.2. Polymeric FETs


Another variant of FET-based devices which is interesting for chemical sensor design is polymeric field effect transistors (PolyFETs) (Covington et al. 2004; Dodabalapur 2006). In this type of CHEMFET device, all the elements are fabricated from polymeric materials. The interaction of an analyte with a polymer layer of the PolyFET directly affects the conductive channel, where sensing events occur at the gate or gate/insulator boundary and indirectly modulate the current by capacitive coupling.

60 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Polymer-based FET devices are promising on account of their lower power dissipation and ease of circuit design. The first polymer large-scale integrated circuits were implemented using this circuit approach. PolyFET backplanes are ideally suited for electronic paper applications and other display schemes due to their low cost and processing advantages. Polymer materials require no epitaxial templating, and most PolyFET processes are simple, with low thermal energy requirements, and are compatible with a range of substrates. In fact, among polymeric semiconductors with higher charge carrier mobilities, -conjugated ladder polymers have the best prospects for application in PolyFETs (Torsi et al. 2000; Babel and Jenekhe 2003). In particular, Torsi et al. (2003, 2004) have developed gas-sensitive organic thin-film transistors in which alkyl- and alkoxy-substituted regioregular polythiophenes, such as dipentoxy-substituted polyterthiophene, are used as active layers, and a set of volatile organic compounds carrying different chemical functionalities is employed as analytes. It was shown that rapid and reversible responses of PolyFETs with remarkable response repeatability are obtainable for a series of alcohol vapors (Crone et al. 2001; Torsi et al. 2000, 2003, 2004). An introduction to PolyFET principles and history, as well as state-of-the-art organic semiconductor structure and performance of PolyFETs, is available in reviews (Facchetti 2007; Kymissis 2009). The operation mechanism of the transistors is explained in terms of bulk conductivities and field-effect mobilities along with temperature dependencies on the basis of variable-range hopping for heavily doped systems and polaronic thermally activated transport for lightly doped systems. This is a consequence of the density of states of conjugated systems, which changes dynamically upon introduction of charge either by a field effect or through doping (Brown et al. 1997). Therefore, the charge transport inside the detecting layer in terms of both holes and electrons in conjugated polymer semiconductors is important for PolyFET device applications. When the detecting element, i.e., the conducting polymer, interacts with gaseous species, it can act as either an electron donor or an electron acceptor, depending on the oxidizing or reducing character of the gas on that particular conducting polymer (see Table 1.3). As the charge carriers in conjugated polymers are a measure of the mobility of holes and electrons, their injection by the analyte may cause the change in the overall voltage of the device. Because of this interaction, both the electron mobility and the carrier number varies, and their relative contribution to the overall change of conductivity is observed in the PolyFET device. This change in conductivity at the electrode/conducting polymer contacts can be attributed to modulation of the height of the Schottky barrier, which is determined by the difference in work function or potential of the organic semiconductor and the metal.

7.5.3. Future Trends in Design of Polymer-Based Chemically Sensitive FET Sensors


The first PolyFET sensors were proposed in 19861987 (Tsumura et al. 1986; Laurs and Heiland 1987). However, despite the availability of a wide range of sensing materials and sensor technologies, commercial implementation of a sensitive, selective, reliable, and inexpensive portable system based on PolyFETs and IGFETs for detecting volatile analytes is still a major issue today. Although many different types of gas sensors have been developed, only a few of them seem to satisfy the minimum requirements

POLYMERS IN CHEMICAL SENSORS 61

for portable sensing instruments capable of performing fast and in situ detection of volatile analytes. In addition to sensitivity and selectivity robustness, low power consumption and compact size of sensor devices have yet to be achieved. Facchetti (2007) believes that the problems lie with the stability and processability of the conducting polymer. To achieve the required parameters in PolyFETs, Facchetti (2007) proposes the following activities. 1. Realization of high carrier mobility. The field-effect mobilities of electrons or holes could be increased by doping, due to the underlying hopping transport of charge, which results in an increase in bulk conductivity. 2. Design of environmentally stable semiconductors. It is also well known that the surface conductivity of the conducting polymer is affected by water vapor or moisture, so moisture is the most common interferant for polymer-based devices operating at room temperature; 3. Achievement of high FET performance from solution-deposited semiconducting films. 4. Achievement of high performance using inexpensive processes for polymer circuit fabrication, such as those employed in the graphic arts industry. 5. Addressing the fundamental FET operational stability parameters related to charge transport and trapping.

8. OUTLOOK
This chapter has shown that polymers possess unique properties that make them interesting and important materials for chemical sensing. A large number of various types of chemical sensing devices have been explored through the tireless efforts of scientists and researchers. Researchers are trying hard to make polymers as sensor materials to replace the commercially available inorganic semiconductors. So far, however, commercial polymer-based sensors are not yet available, because there are still some serious unresolved problems in using polymers as sensing materials in sensor devices. Recent results, however, indicate that many such sensors are now moving from the research stage to potentially commercial phases. It can be predicted that the impact of conducting polymer materials on chemical sensing will depend on the cost of the sensors themselves and on the ability to fabricate sensors that are repeatable, robust, and able to operate reliably in real environments. Understanding sensor devices, sensor materials, sensing analytes, and mechanisms of detection can be helpful for variety of academic areas. Such knowledge might result from more collaboration among researchers in different fields, e.g., physics, chemistry, engineering, biology and biochemistry, materials science, and others. The interdisciplinary nature of sensor technology makes progress difficult for any single discipline and requires interdisciplinary scientists and engineers to work together on complex goals to develop useful physical and chemical/biochemical sensors.

9. ACKNOWLEDGMENTS
We would like to express our gratitude to the Ghenadii Korotcenkov for his valuable scientific and technical suggestions and his contribution in preparing this chapter.

62 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

REFERENCES
Abbas M.N., Mostafa G.A.E., and Homoda A.M.A. (2000) Cetylpyridiniumiodomercurate PVC membrane ion selective electrode for the determination of cetylpyridinium cation in Ezafluor mouth wash and as a detector for some potentiometric titrations. Talanta 53, 425432. Adhikari B. and Majumdar S. (2004) Polymers in sensor applications. Prog. Polym. Sci. 29, 699766. Agbor N.E., Creswell J.P., Petty M.C., and Monkman A.P. (1997) An optical gas sensor based on polyaniline Langmuir-Blodgett films. Sens. Actuators B 41, 137145. Albert K. J., Lewis N.S., Schauer C.L., Sotzing G.A., Stitzel S.E., Vaid T.P., and Walt D.R. (2000) Cross-reactive chemical sensor arrays. Chem. Rev. 100(7), 25952626. Allara D.L. (1995) Critical issues in applications of self-assembled monolayers. Biosens. Bioelectron. 10(910), 771783. Amao Y., Asai K., and Okura I. (2000) Fluorescence quenching oxygen sensor using an aluminum phthalocyanine polystyrene film. Anal. Chim. Acta 407, 4144. Amao Y., Ishikawa Y., and Okura I. (2001) Green luminescent iridium(III) complex immobilized in fluoropolymer film as optical oxygen-sensing material. Anal. Chim. Acta 445, 177182. Amrani M.E.H., Payne P.A., and Persaud K.C. (1996) Multi-frequency measurements of organic conducting polymers for sensing of gases vapours. Sens. Actuators B 33, 137141. Amrani M.E.H., Ibrahim S., and Persaud K.C. (1993) Synthesis, chemical characterisation and multifrequency measurements of poly N-(2-pyridyl) pyrrole for sensing volatile chemicals. Mater. Sci. Eng. C 1, 1722. Anantaraman A.V. and Gardner C L. (1996) Studies on ion-exchange membranes. Part 1. Effect of humidity on the conductivity of Nafion. J. Electroanal. Chem. 414, 115120. Ando M., Swart C., Pringsheim E., Mirsky V.M., and Wolfbeis O.S. (2002) Optical ozone detection by use of polyaniline film. Solid State Ionics 152153, 819822. Ando M., Swart C., Pringsheim E., Mirsky V.M., and Wolfbeis O.S. (2005) Optical ozone-sensing properties of poly(2-chloroaniline), poly(N-methylaniline) and polyaniline films. Sens. Actuators B 108, 528534. Anitha G. and Subramanian E. (2005) Recognition and exposition of intermolecular interaction between CH2Cl2 and CHCl3 by conducting polyaniline materials. Sens. Actuators B 107, 605615. Armstrong R.D. and Horvai G. (1990) Review article: Properties of PVC based membranes used in ion-selective electrodes. Electrochim. Acta 35(1), 17. Arn D., Blom N., Dubler-Studle K., Graber N., and Widmer H.M. (1991) Surface acoustic wave gas sensors: Applications in the chemical industry. Sens. Actuators A 26, 395397. Artigas J., Beltran A., Jimenez C, Bartroli J., and Alonso J. (2001) Development of a photopolymerisable membrane for calcium ion sensors. Application to soil drainage waters. Anal. Chim. Acta 426, 310. Assadi A., Gustafsson G., Willander M., Svensson C., and Inganas O. (1990) Determination of field-effect mobility of poly(3-hexylthiophene) upon exposure to NH3 gas. Synth. Met. 37(13), 123130. Athawale A.A. and Chabukswar V.V. (2001) Acrylic acid-doped polyaniline sensitive to ammonia vapors. J. Appl. Polym. Sci. 79, 19941998. Athawale A.A. and Kulkarni M.V. (2000) Polyaniline and its substituted derivatives as sensors for aliphatic alcohols. Sens. Actuators B 67, 173177. Athawale A.A., Bhagwat S.V., and Katre P.P. (2006) Nanocomposite of Pd-polyaniline as a selective methanol sensor. Sens. Actuators B 114, 263267. Babel A. and Jenekhe S.A. (2003) High electron mobility in ladder polymer field-effect transistors. J. Am. Chem. Soc. 125, 1365613657. Bade K., Tsakova V., and Schultze J.W. (1992) Nucleation, growth and branching of polyaniline from microelectrode experiments. Electrochim. Acta 37(12), 22552261.

POLYMERS IN CHEMICAL SENSORS 63 Bai H. and Shi G. (2007) Gas sensors based on conducting polymers. Sensors 7, 267307. Bailey R.C. and Hupp J.T. (2003) Micropatterned polymeric gratings as chemoresponsive volatile organic compound sensors: Implications for analyte detection and identification via diffraction-based sensor arrays. Anal. Chem. 75, 23922398. Bakker E. and Meyerhoff M.E. (2000) Ionophore-based membrane electrodes: New analytical concepts and nonclassical response mechanisms. Anal. Chim. Acta 416, 121137. Bakker E. (2004) Electrochemical sensors. Anal. Chem. 76, 32853298. Barbetta A., Carnachan R.J., Smith K.H., Zhao C.T., Cameron N.R., Kataky R., Hayman M., Przyborski S.A., and Swan M. (2005) Porous polymers by emulsion templating. Macromol. Symp. 226(1), 203212. Bartic C., Palan B., Campitelli A., and Borghs G. (2002) Monitoring pH with organic-based field-effect transistors. Sens. Actuators B 83, 115122. Bartic C., Campitelli A., and Borghs S. (2003) Field-effect detection of chemical species with hybrid organic/inorganic transistors. Appl. Phys. Lett. 82, 475477. Bartlett P.N. and Ling-Chung S.K. (1989) Conducting polymer gas sensors part II: Response of polypyrrole to methanol vapour. Sens. Actuators 19, 141150. Bay L., Mogensen N., Skaarup S., Sommer-Larsen P., Jorgensen M., and West K. (2002) Polypyrrole doped with alkyl benzenesulfonates. Macromolecules 35, 93459351. Bergveld P. (2003) Thirty years of ISFETOLOGY: What happened in the past 30 years and what may happen in the next 30 years? Sens. Actuators B 88, 116. Bidan G. (1992) Electroconducting conjugated polymers: New sensitive matrices to build up chemical or electrochemical sensors: A review. Sens. Actuators B 6, 4556. Bissell R.A., Persaud K.C., and Travers P. (2002) The influence of nonspecific molecular partitioning of analytes on the electrical responses of conducting organic polymer gas sensors. Phys. Chem. Chem. Phys. 4, 34823491. Blum P., Mohr G.J., Matern K., Reichert J., and Spichiger-Keller U.E. (2001) Optical alcohol sensor using lipophilic Reichardts dyes in polymer membranes. Anal. Chim. Acta 432, 269275. Bobacka J. (2006) Conducting polymer-based solid-state ion-selective electrodes. Electroanalysis 18(1), 718. Bouchet R., Rosini S., Vitter G., and Siebert E.A. (2002) Solid-State potentiometric sensor based on polybenzimidazole for hydrogen determination in air. J. Electrochem. Soc. 149, H119H122. Brook T.E. and Narayanaswamy R. (1998) Polymeric films in optical gas sensors. Sens. Actuators B 51, 7783. Brown A.R., Jarrett C.P., de Leeuw D.M., and Matters M. (1997) Field-effect transistors made from solutionprocessed organic semiconductors. Synth. Met. 88, 3755. Bruce P.G. (1995) Solid State Electrochemistry. Cambridge University Press, Cambridge, UK. Bryce M.R., Chissel A., Kathirgamanathan P., Parker D., and Smith N.R.M. (1987) Soluble, conducting polymers from 3-substituted thiophenes and pyrroles. J. Chem. Soc. Chem. Commun. 6, 466467. Buff W. (1992) SAW sensors. Sens. Actuators A 30, 117121. Burgi L., Sirringhaus H., and Friend R.H. (2002) Noncontact potentiometry of polymer field-effect transistors. Appl. Phys. Lett. 80, 29132915. Cadogan A., Gao Z., Lewenstam A., Ivaska A., and Diamond D. (1992) All-solid-state sodium-selective electrode based on a calixarene ionophore in a poly(vinyl chloride) membrane with a polypyrrole solid contact. Anal. Chem. 64(21), 24962501. Cataldo F. and Maltese P. (2002) Synthesis of alkyl and N-alkyl-substituted polyanilines: A study on their spectral properties and thermal stability. Eur. Polym. J. 38, 17911803. Chabukswar V.V., Pethkar S., and Athawale A.A. (2001) Acrylic acid doped polyaniline as an ammonia sensor. Sens. Actuators B 77, 657663. Chao S. and Wrighton M.S. (1987) Characterization of a solid-state polyaniline-based transistor: Water vapor

64 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS dependent characteristics of a device employing a poly(vinyl alcohol)/phosphoric acid solid-state electrolyte. J.Am. Chem. Soc. 109(22), 66276631. Charlesworth J. M., Partridge A.C., and Garrard N. (1993) Mechanistic studies on the interactions between poly(pyrrole) and organic vapors. J. Phys. Chem. 97(20), 54185423. Chen Z., Xue C.H., Shi W., Luo F.T., Green S., Chen J., and Liu H.Y. (2004) Selective and sensitive fluorescent sensors for metal ions based on manipulation of side-chain compositions of poly(p-phenyleneethynylene)s. Anal. Chem. 76(21), 65136518. Chiang J.C. and MacDiarmid A.G. (1986) Polyaniline: Protonic acid doping of the emeraldine form to the metallic regime. Synth. Met. 13, 193205. Christensen W.H., Sinha D.N., and Agnew S.F. (1993) Conductivity of polystyrene film upon exposure to nitrogen dioxide: A novel NO2 sensor. Sens. Actuators B 10, 149153. Colomban Ph. (1999) Latest developments in proton conductors. Ann. Chim. Sci. Mater. 24, 118. Covington J.A., Gardner J.W., Bartlett P.N., and Toh C.-S. (2004) Conductive polymer gate FET devices for vapour sensing. In: IEE Proc.Circuits Devices Syst. 151(4), 326334. Crone B., Dodabalapur A., Gelperin A., Torsi L., Katz H.E., Lovinger A.J., and Bao Z. (2001) Electronic sensing of vapors with organic transistors. Appl. Phys. Lett. 78, 22292231. Cruz C.M.G.S. and Ticianelli E.A. (1997) Electrochemical and ellipsometric studies of polyaniline films grown under cycling conditions. J. Electroanal. Chem. 428(12), 185192. Cumming C.J., Fisher M., and Sikes J. (2004) Amplifying fluorescent polymer arrays for chemical detection of explosives. In: Gardner W. and Yinon J. (eds.), Electronic Noses and Sensors for the Detection of Explosives. Kluwer Academic Publishers, Dordrecht, The Netherlands, 5370. Cumming C.J., Aker C., Fisher M., Fox M., La Grone M.J., Reust D., Rockley M.G., Swager T.M., Towers E., and Williams V. (2001) Using novel fluorescent polymers as sensory materials for above-ground sensing of chemical signature compounds emanating from buried landmines. IEEE Trans. Geosci. Remote Sensing 39(6), 11191128. Cunningham A.J. (1998) Introduction to Bioanalytical Sensors. John Wiley, New York, 95. Dabke R.B., Singh G.D., Dhanabalan A., Lal R., and Contractor A.Q. (1997) An ion-activated molecular electronic device. Anal. Chem. 69(4), 724727. Dao L.H., Leclerc M., Guay J., and Chevalier J.W. (1989) Synthesis and characterization of substituted poly(anilines). Synth. Met. 29, 377382. Debarnot D.N. and F.P. Epaillard (2003) Polyaniline as a new sensitive layer for gas sensors. Anal. Chim. Acta 475, 115. Dejous C., Rebiere D., Pistre J., Tiret C., and Planade R. (1995) A surface acoustic wave gas sensor: Detection of organophosphorus compounds. Sens. Actuators B 24, 5861. Demarcos S. and Wolfbeis O.S. (1996) Optical sensing of pH based on polypyrrole films. Anal. Chim. Acta 334, 149153. Densakulprasert N., Wannatong L., Chotpattananont D., Hiamtup P., Sirivat, A., and Schwank, J. (2005) Electrical conductivity of polyaniline/zeolite composites and synergetic interaction with CO. Mater. Sci. Eng. BSolid State Mater. Adv. Technol. 117, 276282. Desmonts L.B., Reinhoudt D.N., and Calama M.C. (2007) Design of fluorescent materials for chemical sensing. Chem. Soc. Rev. 36, 9931017. Dias C., Das-Gupta D.K., Hinton Y., and Shuford R.J. (1993) Polymer/ceramic composites for piezoelectric sensors. Sens. Actuators A 37, 343347. Diaz A.F., Kanazawa K.K., and Gardini G.P. (1979) Electrochemical polymerization of pyrrole. J. Chem. Soc., Chem. Commun. 14, 635636. Diaz A.F., Castillo J., Kanazawa K.K., Logan J.A., Salmon M., and Fajardo O. (1982) Conducting poly-N-alkylpyrrole polymer films. J. Electroanal. Chem. 133(2), 233239.

POLYMERS IN CHEMICAL SENSORS 65 Dickinson T.A., White J., Kauer J.S., and Walt D.R. (1996) A chemical-detecting system based on a cross-reactive optical sensor array. Nature, 382, 697700. Dixit V., Misra S.C.K., and Sharma B.S. (2005) Carbon monoxide sensitivity of vacuum deposited polyaniline semiconducting thin films. Sens. Actuators B, 104, 9093. Dodabalapur A. (2006) Organic and polymer transistors for electronics. Mater. Today 9(4), 2430. Drain C.M. and Batteas J.D. (2004) A benchtop method for the fabrication and patterning of nanoscale structures on polymers. J. Am. Chem. Soc. 126(2), 628634. Durst R.A., Baumner A.J., Murray R.W., Buck R.P., and Andrieux C.P. (1997) Chemically modified electrodes: Recommended terminology and definitions. Pure Appl. Chem. 69(6), 13171323. Ellis D.L., Zakin M.R., Bernstein L.S., and Rubner M.F. (1996) Conductive polymer films as ultrasensitive chemical sensors for hydrazine and monomethylhydrazine vapor. Anal. Chem. 68(5), 817822. Elosua C., Matias I.R., Bariain C., and Arregui F.J. (2006) Volatile organic compound optical fiber sensors: A review. Sensors 6, 14401465. Ema K., Yokoyama M., Nakamoto T., and Moriizumi T. (1989) Odour sensing system using a quartz-resonator sensor array and neural-network pattern recognition. Sens. Actuators 18, 291296. Ensafi A.A. and Kazemzadeh A. (1999) Optical pH sensor based on chemical modification of polymer film. Microchemical J. 63, 381388. Ewbank P.C., Loewe R.S., Zhai L., Reddinger J., Sauve G., and McCullough R.D. (2004) Regioregular poly(thiophene-3-alkanoic acid)s: Water soluble conducting polymers suitable for chromatic chemosensing in solution and solid state. Tetrahedron 60, 1126911275. Facchetti A. (2007) Semiconductors for organic transistors. Mater. Today 10(3), 2937. Falcou A., Dhcuene A., Hourquebie P., Marsaeq D., and Balland-Longeall A. (2005) A new chemical polymerization process for substituted anilines: Application to the synthesis of poly(N-alkylanilines) and poly(o-alkylanilines) and comparison of their respective properties. Synth. Met. 149, 115122. Fang Q., Chetwynd D.G., and Gardner J.W. (2002) Conducting polymer films by UV-photo processing. Sens. Actuators A 99, 7477. Feast W.J., Tsibouklis J., Pouwer K.L., Groenendaal L., and Meijer E.W. (1996) Synthesis, processing and material properties of conjugated polymers. Polymer 37(22), 50175047. Ferguson J.A., Healey B.G., Bronk K.S., Barnard S.M., and Walt D.R. (1997) Simultaneous monitoring of pH, CO2 and O2 using an optical imaging fiber. Anal. Chim. Acta 340, 123131. Flory P.J. (1953) Principles of Polymer Chemistry. Cornell University Press, Ithaca, NY, chap. 2. Fog H.M. and Rietz B. (1985) Piezoelectric crystal detector for the monitoring of ozone in working environments. Anal. Chem. 57, 26342638. Forster R.J. (1998) Miniaturized chemical sensors. In: Diamond D. (ed.), Principles of Chemical and Biological Sensors. John Wiley, New York, 239. Friedemann M.N., Robinson S.W., and Gerhardt G.A. (1996) o-Phenylene diamine-modified carbon fiber electrodes for the detection of nitric oxide. Anal. Chem. 68, 26212628. Gangopadhyay R. and De A. (2001) Conducting polymer composites: noble materials for gas sensing. Sens. Actuators B 77, 326329. Gardner J.W. and Bartlett P.N. (1999) Electronic Noses: Principles and Application. Oxford Science, Oxford, UK. Garnier F. (1989) Functionalized conducting polymersTowards intelligent materials. Angew. Chem. 101(4), 529533. Gauglitz G. and Reichert M. (1992) Spectral investigation and optimization of pH and urea sensors. Sens. Actuators B 6, 8386. Gauglitz G., Brecht G., Kraus G., and Nahm W. (1993) Chemical and biochemical sensors based on interferometry at thin (multi-) layers. Sens. Actuators B 11, 2127.

66 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS Ge Z., Brown C.W., Sun L., and Yang S.C. (1993) Fiber-optic pH sensor based on evanescent wave absorption spectroscopy. Anal. Chem. 65, 23352338. Gold V., Loening K.L., McNaught A.D., and Sehmi P. (1987) Compendium of Chemical Terminology, IUPAC Recommendations. Blackwell Scientific, Oxford, UK. Gpel W., Ziegler C., Breer H., Schild D., Apfelbach R., Joerges J., and Malaka R. (1998) Bioelectronic noses: A status report part I. Biosens. Bioelec. 13(34), 479493. Gorton L. and Fiedler U. (1977) A zinc-sensitive polymeric membrane electrode. Anal. Chim. Acta 90, 233236. Gouette M.A. and Leclerc M. (1995) Structure-property relationships in poly(o-phenylenediamine) derivatives. J.Electroanal. Chem. 382(12), 1723. Grate J.W., Abraham M.H., and McGill R.A. (1997) Sorbent polymer materials for chemical sensors and arrays. In: Kress-Rogers E. (ed.), Handbook of Biosensors and Electronic Noses: Medicine, Food and the Environment. CRC Press, Boca Raton, FL. Gratzl M., Hsu D.F., Riley A.M., and Janata J. (1990) Electrochemically deposited polythiophene. 1. Ohmic drop compensation and the polythiophene paradox. J. Phys. Chem. 94(15), 59735981. Grummt U.W., Pron A., Zagorska M., and Lefrant S. (1997) Polyaniline based optical pH sensor. Anal. Chim. Acta 357, 253259. Gupta K.C. and Mujawamariya J.D. (2000) Effect of concentration of ion exchanger, plasticizer and molecular weight of cyanocopolymers on selectivity and sensitivity of Cd(II) ion selective electrode. Talanta 52, 10871103. Gurunathan K. and Trivedi D.C. (2000) Studies on polyaniline and colloidal TiO2 composites. Mater. Lett. 45(5), 262268. Gurunathan K., Murugan A.V., Marimuthu R., Mulik U.P., and Amalnerkar D.P. (1999) Electrochemically synthesised conducting polymeric materials for applications towards technology in electronics, optoelectronics and energy storage devices. Mater. Chem. Phys. 61(3), 173191. Gustafsson G., Lundstrom I., Liedberg B., Wu C.R., Inganas O., and Wennerstrom O. (1989) The interaction between ammonia and poly(pyrrole). Synth. Met. 31(2), 163179. Hamley I.W. (2003) Nanotechnologie mit weichen Materialien. Angew. Chem. 115(15), 17301752. Han W.S., Park M.Y., Chung K.C., Cho D.H., and Hong T.K. (2001) Potentiometric sensor for hydrogen ion based on N,N(-dialkylbenzylethylenediamine neutral carrier in a poly(vinyl chloride) membrane with polyaniline solid contact. Talanta 54, 153159. Hanawa T., Kuwabata S., Hashimoto H., and Yoneyama H. (1989) Gas sensitivities of electropolymerized polythiophene films. Synth. Met. 30(2), 173181. Hao Q., Kulikov V., and Mirsky V.M. (2003) Investigation of contact and bulk resistance of conducting polymers by simultaneous two- and four-point technique. Sens. Actuators B 94, 352357. Hao Q., Wang X., Lu L., Yang X., and Mirsky V.M. (2005) Electropolymerized multilayer conducting polymers with response to gaseous hydrogen chloride. Macromol. Rapid Commun. 26(13), 10991103. Harsnyi G., Rczey M., Dobay R., Lepsnyi I., Illyefalvi-Vitz Z., Steen V.J., Vervaet A., Reinert W., Urbancik J., Guljajev A., Visy C., Inzelt G., and Brsony, I. (1999) Combining inorganic and organic gas sensor elements: A new approach for multicomponent sensing. Sens. Rev. 19(2), 128134. Hartmann P., Leiner J.P., and Lippitsch M.E. (1995) Lumunescence quenching behavior of an oxygen sensor based on a Ru(II) complex dissolved in polystyrene. Anal. Chem. 67, 8893. Hassan S.S.M., Saleh M.B., Abdel Gaber A.A., Mekheimer R.A.H., and Abdel Kream N A. (2000) Novel mercury (II) ion-selective polymeric membrane sensor based on ethyl-2-benzoyl-2-phenylcarbamoyl acetate. Talanta 53, 285293. Hassan S.S.M., Ali M.M., and Attawiya A.M.Y. (2001) PVC membrane based potentiometric sensors for uranium determination. Talanta 54, 11531161.

POLYMERS IN CHEMICAL SENSORS 67 Hatfield V., Neaves P., Hicks P.J., Persaud K., and Travers P. (1994) Towards an integrated electronic nose using conducting polymer sensors. Sens. Actuators B 1819, 221228. Heeg J., Kramer C., Wolter M., Michaelis S., Plieth W., and Fisher W.J. (2001) Polythiophene-O3 surface reactions studied by XPS. Appl. Surf. Sci. 180, 3641. Heeger A.J. (2001) Semiconducting and metallic polymers: The fourth generation of polymeric materials (Nobel Lecture). Angew. Chem. Int. Ed. 40, 25912611. Hirata M. and Sun L. (1994) Characteristics of an organic semiconductor polyaniline film as a sensor for NH3 gas. Sens. Actuators A 40, 159163. Ho K.C. and Hung W.T. (2001) An amperometric NO2 gas sensor based on Pt/Nafion electrode. Sens. Actuators B 79, 1116. Ho C.K., Lindgren E.R., Rawlinson K.S., McGrath L.K., and Wright J.L. (2003) Development of a surface acoustic wave sensor for in-situ monitoring of volatile organic compounds. Sensors 3, 236247. Hoa D.T., Kumar T.N.S., Punekar N.S., Srinivasa R.S., Lal R., and Contractor A.Q. (1992) A biosensor based on conducting polymers. Anal. Chem. 64(21), 26452646. Horanyi G. and Inzelt G. (1989) In situ radiotracer study of the formation and behaviour of polyaniline film electrodes using labelled aniline. J. Electroanal. Chem. 264(12), 259272. Horiuchi T., Ueno Y., Camou S., Haga T., and Tate A. (2006) Portable aromatic VOC gas sensor for onsite continuous air monitoring with 10-ppb benzene detection capability. NTT Tech. Rev. 4(1), 3037. Hosseini S.H. and Entezami A.A. (2003) Conducting polymer blends of polypyrrole with polyvinyl acetate, polystyrene, and polyvinyl chloride based toxic gas sensors. J. Appl. Polym. Sci. 90, 4962. Houser E.J., Mlsna T.E., Nguyen V.K., Chung R., Mowery E.L., and McGill R.A. (2001) Rational materials design of sorbent coatings for explosives: applications with chemical sensors. Talanta 54, 469485. Hrncirova P., Opekar F., and Stulik K. (2000) An amperometric solidstate NO2 sensor with a solid polymer electrolyte and a reticulated vitreous carbon indicator electrode. Sens. Actuators B 69, 199204. Hu H., Trejo M., Nicho M.E., Saniger J.M., and Garcia-Valenzuela A. (2002) Adsorption kinetics of optochemical NH3 gas sensing with semiconductor polyaniline films. Sens. Actuators B 82, 1423. Ichimori K., Ishida H., Fukahori M., Nakazawa H., and Murakami E. (1994) Practical nitric oxide measurement employing a nitric oxide selective electrode. Rev. Sci. Instrum. 65, 15. Inzelt G. (1994) Mechanism of charge transport in polymer-modified electrodes. In: Bard A.J. (ed.), Electroanalytical Chemistry, Vol. 18. Marcel Dekker, New York, 89. Jacquinot P., Muller B., Wehrli B., and Hauser P.C. (2001) Determination of methane and other small hydrocarbons with a platinum-Nafion electrode by stripping voltammetry. Anal. Chim. Acta 432, 110. Janata J. and Bezegh A. (1988) Chemical sensors. Anal. Chem. 60, 62R74R. Janata J. and Josowicz M. (2003) Conducting polymers in electronic chemical sensors. Nat. Mater. 2, 1930. Janata J. and Josowicz M. (1998) Chemical modulation of work function as a transduction mechanism for chemical sensors. Acc. Chem. Res. 31, 241248. Janowiak M., Huang H., Chang S., and Garcia-Rubio L.H. (2001) Development of a fiber optic pH sensor for online control. ACS Symp. Ser. 795, 195210. Jin Z., Su Y., and Duan Y. (2000) An improved optical pH sensor based on polyaniline. Sens. Actuators B 71, 118122. Jin Z., Su Y., and Duan Y. (2001) Development of a polyaniline-based optical ammonia sensor. Sens. Actuators B 72, 7579. Joo B.S., Lee J.H., Lee E.W., Song K.D., and Lee D.D. (2005) Polymer film SAW sensors for chemical agent detection. In: Proceedings of 1st International Conference on Sensing Technology, November 2123, Palmerston North, New Zealand, 307310.

68 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS

Josowicz M., Janata J., Ashley K., and Pons S. (1987) Electrochemical and ultraviolet-visible spectroelectrochemical investigation of selectivity of poltentiometric gas sensors based on polypyrrole. Anal Chem. 59, 253258. Jun H.K., Hoh Y.S., Lee B.S., Lee S.T., Lim J.O., Lee D.D., and Huh J.S. (2003) Electrical properties of polypyrrole gas sensors fabricated under various pretreatment conditions. Sens. Actuators B 96, 576581. Kang E.T., Neoh K.G., Tan K.L., and Tan B.T.G. (1990) Structural studies of halogen-substituted polyanilines by X-ray photoelectron spectroscopy. Synth. Met. 35, 345355. Kar P., Pradhan N.C., and Adhikari B. (2008) A novel route for the synthesis of processable conducting poly(maminophenol). Mater. Chem. Phys. 111(1), 5964. Kar P., Pradhan N.C., and Adhikari B. (2009) Application of sulfuric acid doped poly (m-aminophenol) as aliphatic alcohol vapor sensor material. Sens. Actuators B 140, 525531. Kemp N.T., Kaiser A.B., Trodahl H.J., Chapman B., Buckley R.G., Partridge A.C., and Foot P.J.S. (2006) Effect of ammonia on the temperature-dependent conductivity and thermopower of polypyrrole. J. Polym. Sci. B 44, 13311338. Keshavarzi A., Shahinpoor M., Kim K.J., and Lantz J. (1999) Blood pressure, pulse rate, and rhythm measurement using ionic polymermetal composites sensors. In: Bar-Cohen Y. (ed.), Proc. SPIEInt. Soc.Opt. Eng. 3669, 369376. Kim K.C., Cho S.M., and Choi H.G. (2000) Detection of ethanol gas concentration by fuel cell sensors fabricated using a solid polymer electrolyte. Sens. Actuators B 67, 194198. Kim T.H. and Swager T.M. (2003) A fluorescent self-amplifying wavelength-responsive sensory polymer for fluoride ions. Angew. Chem. Int. Ed. 42(39), 48034806. Kim B.K., Kim Y.H., Won K., Chang H.J., Chi Y.M., Kong K.J., Rhyu B.W., Kim J.J., and Lee J.O. (2005) Electrical properties of polyaniline nanofibre synthesized with biocatalyst. Nanotechnology 16, 11771181. Kim I.B. and Bunz U.H.F. (2006) Modulating the sensory response of a conjugated polymer by proteins: An agglutination assay for mercury ions in water. J. Am. Chem. Soc. 128, 28182819. Kobayashi M., Colaneri N., Boysd M., Wudl F., and Heegh A.J. (1985) The electronic and electrochemical properties of poly(isothianaphthene). J. Chem. Phys. 82(12), 57175723. Korostynska O., Arshak K., Gill E., and Arshak A. (2007) Review on state-of-the-art in polymer based pH sensors. Sensors 7, 30273042. Korotcenkov G., Han S.-D., and Stetter J.R. (2009) Review of electrochemical hydrogen sensors. Chem. Rev. 109(3), 14021433. Koshets I.A., Kazantseva Z.I., and Shirshov Y.M. (2003) Polymer films as sensitive coatings for quartz crystal microbalance sensors array. Semicond. Phys. Quan. Electron. Optoelectron. 6(4), 505507. Koul S., Chandra R., and Dhawan S.K. (2001) Conducting polyaniline composite: A reusable sensor material for aqueous ammonia. Sens. Actuators B 75, 151159. Krondak M., Broncova G., Anikin S., Merz A., and Mirsky V. (2006) Chemosensitive properties of poly-4,4dialkoxy-2,2-bipyrroles. J. Solid State Electrochem. 10(3), 185191. Kukla A.L., Shirshov Y.M., and Piletsky S.A. (1996) Ammonia sensors based on sensitive polyaniline films. Sens. Actuators B 37, 135140. Kumar D. and Sharma R.C. (1998) Advances in conductive polymers. Eur. Polym. J. 34(8), 10531060. Kymissis I. (2009) Organic Field Effect Transistors: Theory, Fabrication and Characterization. Series on Integrated Circuits and Systems (Chandrakasan A., ed.). Springer-Verlag, New York. Lange U., Roznyatovskaya N.V., and Mirsky V.M. (2008) Conducting polymers in chemical sensors and arrays. Anal. Chim. Acta 614, 126. Laurs H. and Heiland G. (1987) Electrical and optical properties of phthalocyanine films. Thin Solid Films 149, 129135.

POLYMERS IN CHEMICAL SENSORS 69 Leclerc M. Guay J., and Dao L.H. (1989) Synthesis and characterization of poly(alkylanilines). Macromolecules 22, 649653. Leiner J.P. (1995) Optical sensors for in vitro blood gas analysis. Sens. Actuators B 29, 169173. Levin O., Kondratiev V., and Malev V. (2005) Charge transfer processes at poly-o-phenylenediamine and poly-oaminophenol films. Electrochim. Acta 50(78), 15731585. Li B., Sauve G., Iovu M.C., Jeffries-EL M., Zhang R., Cooper J., Santhanam S., Schultz L., Revelli J.C., Kusne A.G., Kowalewski T., Snyder J.L., Weiss L.E., Fedder G.K., McCullough R.D., and Lambeth D.N. (2006) Volatile organic compound detection using nanostructured copolymers. Nano Lett. 6(8), 15981602. Li B., Santhanam S., Schultz L., Jeffries-EL M., Iovu M.C., Sauve G., Cooper J., Zhang R., Revelli J.C., and Kusne A.G. (2007) Inkjet printed chemical sensor array based on polythiophene conductive polymers. Sens. Actuators B 123, 651660. Li D., Jiang Y., Wu Z., Chen X., and Li Y. (2000) Self-assembly of polyaniline ultrathin films based on dopinginduced deposition effect and applications for chemical sensors. Sens. Actuators B 66, 125127. Linford R.G. (ed.) (1987) Electrochemical Science and Technology of Polymers, Vol. 1. Elsevier, London. Lindfors T. and Ivaska A. (2001) Calcium-selective electrode based on polyaniline functionalized with bis[4(1,1,3,3-tetramethylbutyl) phenyl]phosphate. Anal. Chim. Acta 437, 171183. Lindsey S.E. and Street G.B. (1984) Conductive composites from polyvinyl alcohol and polypyrrole. Synth. Met. 10, 6769. Liu D., Liu J., Tian D., Hong W., Zhou X., and Yu J.C. (2000) Polymeric membrane silver-ion selective electrodes based on bis(dialkyldithiophosphates). Anal. Chim. Acta 416, 139144. Loi A., Manunza I., and Bonfiglio A. (2005) Flexible, organic, ion-sensitive field-effect transistor. Appl. Phys. Lett. 86(10), 103512103514. Lundberg B. and Sundqvist B. (1986) Resistivity of a composite conducting polymer as a function of temperature, pressure, and environment: Applications as a pressure and gas concentration transducer. J. Appl. Phys. 60(3), 10741079. Lyons M.E.G. (ed.) (1994) Electroactive Polymer Electrochemistry, Part 1. Plenum Press, New York. Lyons M.E.G. (ed.) (1996) Electroactive Polymer Electrochemistry, Part 2. Plenum Press, New York. Mabeck J.T. and Malliaras G.G. (2006) Chemical and biological sensors based on organic thin-film transistors. Anal. Bioanal. Chem. 384, 343353. MacDiarmid A.G. and Epstein A.J. (1989) Polyanilines: A novel class of conducting polymers. Faraday Discuss. Chem. Soc. 88, 317332. MacDiarmid A.G. and Epstein A.J. (1994) The concept of secondary doping as applied to polyaniline. Synth. Met. 65, 103116. MacDiarmid A.G. and Epstein A.J. (1995) Secondary doping in polyaniline. Synth. Met. 69, 8592. Mahajan R.K. and Parkash O. (2000) Silver (I) ion selective PVC membrane based on bis-pyridine tetramide macrocycle. Talanta 52, 691693. Maksymiuk K. (2006) Chemical reactivity of polypyrrole and its relevance to polypyrrole based electrochemical sensors. Electroanalysis 18, 15371551. Malhotra, B.D., Kumar N., and Chandra S. (1986) Recent studies of heterocyclic and aromatic conducting polymers. Prog. Polym. Sci. 12(3), 179218. Martin C.R., Parthasarathy R., and Menon V. (1993) Template synthesis of electronically conductive polymersA new route for achieving higher electronic conductivities. Synth. Met. 55(23), 11651170. Matsuguchi M., Tamai K., and Sakai Y. (2001) SO2 gas sensors using polymers with different amino groups. Sens. Actuators B 77, 363367. Mayes A.G., Blyth J., Kyrolainen-Reay M., Millington R.B, and Lowe C.R. (1999) A holographic alcohol sensor. Anal Chem. 71, 33903396.

70 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS McCullough R.D. (1998) The chemistry of conducting polythiophenes. Adv. Mater. 10, 93116. McGovern S.T., Spinks G.M., and Wallace G.G. (2005) Micro-humidity sensors based on a processable polyaniline blend. Sens. Actuators B 107, 657665. Mcquade D.T., Pullen A.E., and Swager T.M. (2000) Conjugated polymer-based chemical sensors. Chem. Rev. 100(7), 25372574. Miasik J., Hooper A., and Tofield B. (1986) Conducting polymer gas sensors. J. Chem. Soc. Faraday Trans. 1 82(4), 11171127. Mirmohseni A. and Hassanzadeh V. (2001) Application of polymer-coated quart crystal microbalance (QCM) as a sensor for BTEX compounds vapors. J. Appl. Polym. Sci. 79, 10621066. Mirmohseni A. and Rostamizadeh K. (2006) Quartz crystal nanobalance in conjunction with principal component analysis for identification of volatile organic compounds. Sensors 6, 324334. Misra S.C.K., Mathur P., and Srivastava B.K. (2004) Vacuum-deposited nanocrystalline polyaniline thin film sensors for detection of carbon monoxide. Sens. Actuators A 114, 3035. Miura N., Kato H., Yamazoe N., and Seiyama T. (1984) A proton conductor oxygen sensor operative at lower temperature. Denki Kagaku 52, 376377. Mizutani F., Yabuki S., Sawaguchi T., Hirata Y., Sato Y., and Iijima S. (2001) Use of a siloxane polymer for the preparation of amperometric sensors: O2 and NO sensors and enzyme sensors. Sens. Actuators B 76, 489493. Mohammad F. (1998) Compensation behaviour of electrically conductive polythiophene and polypyrrole. J. Phys. D: Appl. Phys. 31(8), 951960. Mohr G.J. (2006) Polymers for optical sensors. In: Baldini F., Chester A.N., Homola J., Martellucci S. (eds.), Optical Chemical Sensors. Springer-Vrlag, Dordrecht, The Netherlands, 297321. Mousavi M.F., Alizadeh N., Shamsipur M., and Zohari N. (2000) A new PVC-based 1,10-dibenzyl-1,10-diaza-18crown-6 selective electrode for detecting nickel II ion. Sens. Actuators B 66, 98100. Munkholm C., Walt D.R., Milanovich F.P., and Klainer S.A. (1986) Polymer modification of fiber optic chemical sensors as a method of enhancing fluorescence signal for pH measurement. Anal. Chem. 58, 14271430. Murray R.W. (1984) Chemically modified electrodes. In: Bard A.J (ed.), Electroanalytical Chemistry, Vol. 13. Marcel Dekker, New York, 191. Murray R.W., Ewing A.G., and Durst R.A. (1987) Chemically modified electrodes. Molecular design for electroanalysis. Anal. Chem. 59(5), 379A390A. Nakagawa K., Sadaoka Y., Supriyatno H., Kubo A., Tsutsumi C., and Tabuchi K. (2001) Optochemical HCl gas detection using alkoxy substituted tetraphenyl porphyrinpolymer composite films. Effects of alkoxy chain length on sensing characteristics. Sens. Actuators B 76, 4246. Nakamoto T., Fukunishi K., and Moriizumi T. (1990) Identification capability of odor sensor using quartz-resonator array and neural-network pattern recognition. Sens. Actuators B 1, 473476. Nanto H., Dougami N., Mukai T., Habara M., Kusano E., Kinbara A., Ogawa T., and Oyabu T. (2000) A smart gas sensor using polymer-film-coated quartz resonator microbalance. Sens. Actuators B 66, 1618. Nicho M.E., Trejo M., Garca-Valenzuela A., Saniger J.M., Palacios J., and Hu H. (2001) Polyaniline composite coatings interrogated by nulling optical-transmittance bridge for sensing low concentrations of ammonia gas. Sens. Actuators B 76, 1824. Nohria R., Khillan R.K., Su Y., Dikshit R., Lvov Y., and Varahramyan K. (2006) Humidity sensor based on ultrathin polyaniline film deposited using layer-by-layer nano-assembly. Sens. Actuators B 114, 218222. Nilsson D., Kugler T., Svensson P.O., and Berggren M. (2002) An all-organic sensortransistor based on a novel electrochemical transducer concept printed electrochemical sensors on paper. Sens. Actuators B 86, 193197. Nishizawa M., Matsue T., and Uchida I. (1992) Penicillin sensor based on a microarray electrode coated with pHresponsive polypyrrole. Anal. Chem. 64(21), 26422644.

POLYMERS IN CHEMICAL SENSORS 71 Numata M., Baba K., Hemmi A., Hachiya H., Ito S., Masadome T., Asano Y., Ohkubo S., Gomi T., Imato T., and Hobo T. (2001) Determination of hardness in tapwater and upland soil extracts using a long-term stable divalent cation selective electrode based on a lipophilic acrylate resin as a membrane matrix. Talanta 55, 449457. Nylander C., Armgarth M., and Lundstrom I. (1983) An ammonia detector based on a conducting polymer. Anal. Chem. Symp. Ser. 17, 203207. Odian G. (2004) Principles of Polymerization, 4th ed. John Wiley, Hoboken, NJ. Ohmori Y., Takahashi H., Muro K., Uchida M., Kawai T., and Yoshino K. (1991) Gas-sensitive Schottky gated field effect transistors utilizing poly(3-alkylthiophene) films. Jpn. J. Appl. Phys. 30(7B), L1247L1249. Opdycke W.N., Parks S.J., and Meyerhoff M.E. (1983) Polymer-membrane pH electrodes as internal elements for potentiometric gas-sensing systems. Anal. Chim. Acta 155, 1120. Opdycke W.N. and Meyerhoff M.E. (1986) Development and analytical performance of tubular polymer membrane electrode based carbon dioxide catheters. Anal. Chem. 58, 950956. Opekar F. and Bruckenstein S. (1984) Determination of gaseous hydrogen sulfide by cathodic stripping voltammetry after preconcentration on a silver metalized porous membrane electrode. Anal. Chem. 56, 12061209. Opekar F. and Stulik K. (1999) Electrochemical sensors with solid polymer electrolytes. Anal. Chim. Acta 385, 151162 Otagawa T., Madou M., Wing S., Rich-Alexander J., Kusanagi S., Fujioka T., and Yasuda A. (1990) Planar microelectrochemical carbon monoxide sensors. Sens. Actuators B 1, 319325. Otero T.F. and Rodriguez J. (1994) Parallel kinetic studies of the electrogeneration of conducting polymers: Mixed materials, composition and properties control. Electrochim. Acta 39(2), 245253. Pandey P.C., Upadhyay S., Singh G., Prakash R., Srivastava R.C., and Seth P.K. (2000) A new solid-state pH sensor and its application in the construction of all solid-state urea biosensor. Electroanalysis12, 517521. Pandey P.C. and Singh G. (2001) Tetraphenylborate doped polyaniline based novel pH sensor and solid-state urea biosensor. Talanta 55, 773182. Pang Z., Gu X., Yekta A., Masoumi Z., Coll J.B., Winnik M.A., and Manners I. (1996) Phosphorescent oxygen sensors utilizing sulfurnitrogenphosphorus polymer matrixes. Adv. Mater. 8, 768771. Park Y.H. and Han M.H. (1992) Preparation of conducting polyacrylonitrile/polypyrrole composite films by electrochemical synthesis and their electrical properties. J. Appl. Polym. Sci. 45, 19731982. Park Y.W., Moon J.S., Bak M.K., and Jin J.I. (1989) Electrical properties of polyaniline and substituted polyaniline derivatives. Synth. Met. 29, 389394. Patra D. and Mishra A.K. (2001) Fluorescence quenching of benzo[k] fluoranthene in poly(vinyl alcohol) film: A possible optical sensor for nitro aromatic compounds. Sens. Actuators B 80, 278282. Paul E.W., Ricco A.J., and Wrighton M.S. (1985) Resistance of polyaniline films as a function of electrochemical potential and the fabrication of polyaniline-based microelectronic devices. J. Phys. Chem. 89(8), 14411447. Pei Q. and Inganas O. (1993) Conjugated polymers as smart materials, gas sensors and actuators using bending beams. Synth. Met. 57(1), 37303735. Percival C.J., Stanley S., Galle M., Braithwaite A., Newton M.I., McHale G., and Hayes W. (2001) Molecularimprinted, polymer-coated quartz crystal microbalances for the detection of terpenes. Anal. Chem. 73, 42254228. Persaud K.C. and Pelosi P. (1985) An approach to an artificial nose. Trans. Am. Soc. Artif. Int. Organs 31, 297300. Persaud K.C. (2005) Polymers for chemical sensing. Mater. Today 8, 3845. Potje-Kamloth K., Polk B.J., Josowicz M., and Janata J. (2002) Doping of polyaniline in solid-state with photogenerated triflic acid. Chem. Mater. 14, 27822787. Prissanaroon W., Ruangchuay L., Sirivat A., and Schwank J. (2000) Electrical conductivity response of dodecylbenzene sulfonic acid-doped polypyrrole films to SO2N2 mixtures. Synth. Met. 114(1), 6572.

72 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS Pringsheim E., Terpetschnig E., and Wolfbeis O.S. (1997) Optical sensing of pH using thin films of substituted polyanilines. Anal. Chim. Acta 357, 247252. Rabek J.F. (1995) Polymer Photodegradation: Mechanism and Experimental Methods. Chapman & Hall, London. Ram M.K., Yavuz O., Lahsangah V., and Aldissi M. (2005a) CO gas sensing from ultrathin nano-composite conducting polymer film. Sens. Actuators B 106, 750757. Ram M.K., Yavuz O., and Aldissi M. (2005b) NO2 gas sensing based on ordered ultrathin films of conducting polymer and its nanocomposite. Synth. Met. 151(1), 7784. Ramesh C., Velayutham G., Murugesan N., Ganesan V., Dhathathreyan K.S., and Periaswami G. (2003) An improved polymer electrolyte-based amperometric hydrogen sensor. J. Solid State Electrochem. 8, 511516. Ratcliffe N.M. (1990) Polypyrrole-based sensor for hydrazine and ammonia. Anal. Chim. Acta 239, 257262. Razumovskii S.D. and Zaikov G.Y. (1982) Effect of ozone on saturated polymers. Polym. Sci. U.S.S.R. 24(10), 28052325. Reemts J., Parisi J., and Schlettwein D. (2004) Electrochemical growth of gas-sensitive polyaniline thin films across an insulating gap. Thin Solid Films 466(12), 320325. Reese C., Roberts M., Ling M., and Bao Z. (2004) Organic thin film transistors. Mater. Today 7(9), 2027. Rella R., Siciliano P., Quaranta F., Primo T., Valli L., Schenetti L., Mucci A., and Iarossi D. (2000) Gas sensing measurements and analysis of the optical properties of poly[3-(butylthio)thiophene] LangmuirBlodgett films. Sens. Actuators B 68, 203209. Ricks-Laskoski H.L. and Buckley L.J. (2006) Twenty-year aging study of electrically conductive polypyrrole films. Synth. Met. 156, 417419. Rivas B.L., Sanchez C.O., Bernede J.C., and Mollinie P. (2002) Synthesis, characterization, and properties of poly(2- and 3-aminophenol) and poly(2- and 3-aminophenol)-Cu(II) materials. Polym. Bull. 49(4), 257264. Rosini S. and Siebert E. (2005) Electrochemical sensors for detection of hydrogen in air: model of the non-Nernstian potentiometric response of platinum gas diffusion electrodes. Electrochim. Acta 50, 29432953. Roy S., Sana S., Adhikari B., and Basu S. (2003) Preparation of doped polyaniline and polypyrrole films and applications for hydrogen gas sensors. J. Polym. Mater. 20, 173180. Ruangchuay L., Sirivat A., and Schwank J. (2003) Polypyrrole/poly(methylmethacrylate) blend as selective sensor for acetone in lacquer. Talanta 60, 2530. Ruangchuay L., Sirivat A., and Schwank J. (2004) Electrical conductivity response of polypyrrole to acetone vapor: Effect of dopant anions and interaction mechanisms. Synth. Met. 140(1), 1521. Sadaoka Y., Matsuguchi M., Sakai Y., and Murata Y. (1992) Optical humidity sensing characteristics of Nafion-dyes composite thin films. Sens. Actuators B 7, 443446. Sadaoka Y., Sakai Y., and Akiyama H. (1986) A humidity sensor using alkali salt-poly(ethylene oxide) hybrid films. J. Mater. Sci. 21(1), 235240. Sadeghipour K., Salomon R.. and Neogi S. (1992) Development of a novel electrochemically active membrane and smart material-based vibration sensor/damper. Smart Mater. Struct. 1, 172179. Sakai Y., SadaokaY., and Matsuguchi M. (1989) A humidity sensor using cross-linked quaternized polyvinylpyridine. J. Electrochem. Soc. 136(1), 171174. Sakai Y., Sadaoka Y., Matsuguchi M., Yokouchi H., and Tamai K. (1995) Effect of carbon dioxide on the electrical conductivity of polyethylene glycol-alkali carbonate complex film. Mater. Chem. Phys. 42, 7376. Sakthivel M. and Weppner W. (2006) Development of a hydrogen sensor based on solid polymer electrolyte membranes. Sens. Actuators B 113, 9981004. Sakurai Y., Jung H.S., Shimanouchi T., Inoguchi T., Morita S., Kuboi R., and Natsukawa K. (2002) Novel arraytype gas sensors using conducting polymers, and their performance for gas identification. Sens. Actuators B 83, 270275.

POLYMERS IN CHEMICAL SENSORS 73 Salavagione H.J., Arias J., Garces P., Morallon E., Barbero C., and Vazquez J.L. (2004) Spectroelectrochemical study of the oxidation of aminophenols on platinum electrode in acid medium. J. Electroanal. Chem. 565(2), 375383. Saleh M.B., Hassan S.S.M., Abdel Gaber A.A., and Abdel Kream N.A. (2001) Novel potentiometric membrane sensor for selective determination of aluminum(III) ions. Anal. Chim. Acta 434, 247253. Samec Z., Opekar F., and Crijns G.J.E.F. (1995) Solid-state hydrogen sensor based on a solid-polymer electrolyte. Electroanalysis 7, 10541058. Sandler S.R. and Karo W. (1974) Polymer Synthesis. Academic Press, New York. Saravanan S., Mathai C.J., Venkatachalam S., and Anantharaman M.R. (2004) Low K thin films based on RF plasma-polymerized aniline. New J. Phys. 6, 6468. Sato, M., Tanaka S., and Kacriyama K. (1986) Electrochemical preparation of conducting poly(3-methylthiophene): comparison with polythiophene and poly(3-ethylthiophene). Synth. Met. 14(4), 279288. Saxena A., Fujiki M., Rai R., Kim S.Y., and Kwak G. (2004) Highly sensitive and selective fluoride ion chemosensing, fluoroalkylated polysilane. Macromol. Rapid Commun. 25(20), 17711775. Schierbaum K.D., Gerlach A., Haug M., and Gpel W. (1992) Selective detection of organic molecules with polymers and supramolecular compounds: Applications of capacitance, quartz microbalance, and calorimetric transducers. Sens. Actuators A 31, 130137. Segal E., Tchoudakov R., Narkis M., and Siegmann A. (2002) Thermoplastic polyurethane-carbon black compounds: Structure, electrical conductivity and sensing of liquids. Polym. Eng. Sci. 42, 24302439. Segal E., Tchoudakov R., Narkis M., Siegmann A., and Yen W. (2005) Polystyrene/polyaniline nanoblends for sensing of aliphatic alcohols. Sens. Actuators B 104, 140150. Schiavon G., Zotti G., Toniolo R., and Bontempelli G. (1995) Electrochemical detection of trace hydrogen sulfide in gaseous samples by porous silver electrodes supported on ion exchange membranes (solid polymer electrolytes). Anal. Chem. 67, 318323. Sengupta P.P., Kar P., and Adhikari B. (2009) Influence of dopant in the synthesis, characteristics and ammonia sensing behavior of processable polyaniline. Thin Solid Films 517, 37703775. Sengupta P.P., Barik S., and Adhikari B. (2006) Polyaniline as a gas-sensor material. Mater. Manuf. Proc. 21, 263270. Shahinpoor M., Bar-Cohen Y., Simpson J., and Smith J. (1998) Ionic polymer-metal composites (IPMCs) as biomimetic sensors, actuators and artificial muscles: a review. Smart Mater. Struct. 7(6), R15R30. Shakhsher Z., Seitz W.R., and Legg K.D. (1994) Single fiber-optic pH sensor based on changes accompanying polymer swelling. Anal. Chem. 66, 17311735. Shamsipur M., Ganjali M.R., Rouhollahi A., and Moghimi A. (2001) Beryllium-selective membrane sensor based on 3,4-di[2-(2-tetrahydro-2H-pyranoxy)]ethoxy styrene-styrene copolymer. Anal Chim Acta 434, 2327. Shi G., Luo M., Xue J., Xian Y., Jin L., and Jin J.Y. (2001) The study of PVP/Pd/IrO2 modified sensor for amperometric determination of sulfur dioxide. Talanta 55, 241247. Shi M. and Anson F.C. (1996) Effects of hydration on the resistances and electrochemical responses of nafioncoatings on electrodes. J. Electroanal. Chem. 415, 4146. Silva L.I.B., Rocha-Santos T.A.P., and Duarte A.C. (2008) Sensing of volatile organic compounds in indoor atmosphere and confined areas of industrial environments. Global NEST J. 10(2), 217225. Slater J.M. and Paynter J. (1994) Prediction of gas sensor response using basic molecular parameters. Analyst 119, 191195. Slater J.M. and Watt E.J. (1991) Examination of ammonia-poly(pyrrole).interactions by piezoelectric and conductivity measurements. Analyst 116, 11251130. Smith J.D.S. (1998) Intrinsically electrically conducting polymers. Synthesis, characterization, and their applications. Prog. Polym. Sci. 23, 5779.

74 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS Sohn H., Sailor M.J., Magde D., and Trogler W.C. (2003) Detection of nitroaromatic explosives based on photoluminescent polymers containing metalloles. J. Am. Chem. Soc. 125(13), 38213830. Stilwell D.E. and Park S.M. (1988) Electrochemistry of conductive polymers. J. Electrochem. Soc. 135(9), 22542262. Sudhlter E.J.R., Van der Wal P.D., Skowronska-Ptasinska M., Van den Berg A., and Reinhoudt D.N. (1989) Ionsensing using chemically-modified ISFETs. Sens. Actuators 17, 189194. Supriyatno H., Nakagawa K., and Sadaoka Y. (2001) Optochemical HCl gas detection using mono-substituted tetraphenyl porphinpolymer composite films. Sens. Actuators B 76, 3641. Sukeerthi S. and Contractor A.Q. (1994) Applications of conducting polymers as sensors. Indian J. Chem. Sect. A 33, 565571. Sundmacher K., Rihko-Struckmann L.K., and Galvita V. (2005) Solid electrolyte membrane reactors: Status and trends. Catal. Today 104, 185199. Sunny V., Narayanan T.N., Sajeev U.S., Joy P.A., Kumar D.S., Yoshida Y., and Anantharaman M.R. (2006) Evidence for intergranular tunnelling in polyaniline passivated -Fe nanoparticles. Nanotechnology 17, 47654772. Suri K., Annapurni S., Sarkar A.K., and Tandon R.P. (2002) Gas and humidity sensors based on iron oxide polypyrrole nanocomposites. Sens. Actuators B 81, 277282. Supriya L. and Claus R.O. (2004) Fabrication of electrodes for polymer actuators and sensors via self-assembly. In: Proceedings of IEEE Sensor Conference, October 2427, Vienna, 2, 619622. Svetlicic V., Schmidt A.J., and Miller L.L. (1998) Conductometric sensors based on the hypersensitive response of plasticized polyaniline films to organic vapors. Chem. Mater. 10, 33053307. Syed A.A. and Dinesan M.K. (1991) Review: PolyanilineA novel polymeric material. Talanta 38(8), 815837. Tan C.K. and Blackwood D.J. (2000) Interactions between polyaniline and methanol vapour. Sens. Actuators B 71, 184191. Tanese M.C., Torsi L., Cioffi N., Zotti L.A., Colangiuli D., Farinola G.M., Babudri F., Naso F., Giangregorio M.M., Sabbatini L., and Zambonin P.G. (2004) Poly(phenyleneethynylene) polymers bearing glucose substituents as promising active layers in enantioselective chemiresistors. Sens. Actuators B 100, 1721. Teixeira M.F.S, Fatibello-Filho O., Ferracin L.C., Rocha-Filho R.C., and Bocchi N. (2000) A l-MnO2-based graphite-epoxy electrode as lithium ion sensor. Sens. Actuators B 67, 96100. Thackeray J.W., White H.S., and Wrighton M.S. (1985) Poly(3-methylthiophene)-coated electrodes: Optical and electrical properties as a function of redox potential and amplification of electrical and chemical signals using poly(3-methylthiophene)-based microelectrochemical transistors. J. Phys. Chem. 89(23), 51335140. Thackeray J.W. and Wrighton M.S. (1986) Chemically responsive microelectrochemical devices based on platinized poly(3-methylthiophene): Variation in conductivity with variation in hydrogen, oxygen, or pH in aqueous solution. J. Phys. Chem. 90(25), 66746679. Thomas S.W. III and Swager T.M. (2006) Trace hydrazine ddetection with fluorescent conjugated polymers: A turnon sensory mechanism. Adv. Mater. 18(8), 10471050. Thompson M. and Stone D.C. (1997) Surface-Launched Acoustic Wave Sensors: Chemical Sebsing and Thin-Film Characterization. John Wiley, New York. Toal S.J. and Trogler W.C. (2006) Polymer sensors for nitroaromatic explosives detection. J. Mater. Chem. 16, 28712883. Tong H., Wang L.X., Jing X.B., and Wang F.S. (2003) Turn-on conjugated polymer fluorescent chemosensor for fluoride. Macromolecules 36(8), 25842586. Torres K.Y.C., Garcia C.A.B., Fernandes J.C.B., de Oliveira Neto G., and Kubota L.T. (2001) Use of self-plasticizing EVA membrane for potentiometric anion detection. Talanta 53, 807814. Torsi L., Pezzuto M., Siciliano P., Rella R., Sabbatini L., Valli L., and Zambonin P.G. (1998) Conducting polymers doped with metallic inclusions: New materials for gas sensors. Sens. Actuators B 48, 362367.

POLYMERS IN CHEMICAL SENSORS 75 Torsi L., Dodabalapur A., Sabbatini L., and Zambonin P.G. (2000) Multiparameter gas sensors based on organic thin-film-transistors. Sens. Actuators B 67, 312316. Torsi L., Tanese M., Cioffi N., Gallazzi M., Sabbatini L., Zambonin P., Raos G., Meille S., and Giangregorio M. (2003a) Side-chain role in chemically sensing conducting polymer field-effect transistors. J. Phys. Chem. B 107(31), 75897594. Torsi L., Tafuri A., Cioffi N., Gallazzi M.C., Sassella A., Sabbatini L., and Zambonin P.G. (2003b) Regioregular polythiophene field-effect transistors employed as chemical sensors. Sens. Actuators B 93, 257262. Torsi L., Tanese M.C., Cioffi N., Gallazzi M., Sabbatini L., and Zambonin P.G. (2004) Alkoxy-substituted polyterthiophene thin-film-transistors as alcohol sensors. Sens. Actuators B 98, 204207. Torsi L., Tanese M.C., Crone B., Wang L., and Dodabalapur A. (2007) Organic transistor chemical sensors. In: BaoZ. and Locklin J. (eds.), Organic Field-Effect Transistors. CRC Press, Boca Raton, FL. Toshima N. and Hara S. (1995) Direct synthesis of conducting polymers from simple monomers. Prog. Polym. Sci. 20, 155183. Tourillon G. and Garnies F. (1982) New electrochemically generated organic conducting polymers. J. Electroanal. Chem. 135(1), 173178. Tsujimura Y., Yokoyama M., and Kimura K. (1995) Practical applicability of silicone rubber membrane sodiumselective electrode based on oligosiloxane-modified calix[4]arene neutral carrier. Anal. Chem. 67, 24012404. Tsumura A., Koezuka H., and Ando T. (1986) Macromolecular electronic device: Field effect transistor with a polythiophene thin film. Appl. Phys. Lett. 49, 12101219. Uhlmann P.B., Pretsch E., and Bakker E. (1998) Carrier-based ion-selective electrodes and bulk optodes. 2. Ionophores for potentiometric and optical sensors. Chem. Rev. 98, 15931688. Uyama Y., Kato K., and Ikada Y. (1998) Surface modification of polymers by grafting. Adv. Polym. Sci. 137, 139. Vercelli B., Zecchin S., Comisso N., Zotti G., Berlin A., Dalcanale E., and Groenendaal L.B. (2002) Solvoconductivity of polyconjugated polymers: The roles of polymer oxidation degree and solvent electrical permittivity. Chem. Mater. 14(11), 47684774. Virji S., Huang J., Kaner R.B., and Weiller B.H. (2004) Polyaniline nanofiber gas sensors: Examination of response mechanisms. Nano Lett. 4(3), 491496. Waltman J., Bargon J., and Diaz A.F. (1983) Electrochemical studies of some conducting polythiophene films. J.Phys. Chem. 87(8), 14591463. Waltman J., Diaz A.F., and Bargon J. (1984) Substituent effects in the electropolymerization of aromatic heterocyclic compounds. J. Phys. Chem. 88(19), 43434346. Waltman J. and Diaz A.F. (1985) The electrochemical oxidation and polymerization of polycyclic hydrocarbons. J.Electrochem. Soc. 132(3), 631634. Wang B. and Wasielewski M.R. (1997) Design and synthesis of metal ion-recognition-induced conjugated polymers: An approach to metal ion sensory materials. J. Am. Chem. Soc. 119(1), 1221. Wang H.L., Toppare L., and Fernandez J.E. (1990) Conducting polymer blends: Polythiophene and polypyrrole blends with polystyrene and poly(bisphenol A carbonate). Macromolecules 23, 10531059. Watanabe M., Sanui K., Ogata N., Kobayashi T., and Ohtaki Z. (1985) Ionic conductivity and mobility in network polymers from poly(propylene oxide) containing lithium perchlorate. J. Appl. Phys. 57(1), 123128. Watcharaphalakorn S., Ruangchuay L., Chotpattahanont D., Sirivat A., and Schwank J. (2005) Polyaniline/polyimide blends as gas sensors and electrical conductivity response to CO-N2 mixtures. Polym. Int. 54, 11261133. Wegner G. and Remmers M. (1995) Thin films. In: Ullmanns Encyclopedia of Industrial Chemistry, A26 Thin Films. VCH, Weinheim, Germany, 103. Wolfbeis O.S. (1991) Biomedical applications of fibre optic chemical sensors. Int. J. Optoelectron. 6(5), 425441. Wolfbeis O.S. (2006) Fiber-optic chemical sensors and biosensors. Anal. Chem. 78, 38593874.

76 CHEMICAL SENSORS: FUNDAMENTALS. VOLUME 3: POLYMERS & OTHER MATERIALS Wohltjen H. and Dessy R.E. (1979) Surface acoustic wave probe for chemical analysis. I. Introduction and instrument description. Anal. Chem. 51, 14581464. Wroblewski W., Wojciechowski K., Dybko A., Brzozka Z., Egberink R.J.M, Snellink-Ruel B.H.M., and Reinhoudt D.N. (2000) Uranyl salophenes as ionophores for phosphate-selective electrodes. Sens. Actuators B 68, 313318. Wu X.Q., Shimizu Y., and Egashira M. (1989) Carbon dioxide sensor consisting of potassium carbonate-polyethylene glycol solution supported on porous ceramics. II. Study of carbon dioxide-sensing mechanism by employing cyclic voltammetry. J. Electrochem. Soc. 136, 28922895. Xie D., Jiang Y., Pan W., Li D., Wu Z., and Li Y. (2002) Fabrication and characterization of polyaniline-based gas sensor by ultra-thin film technology. Sens. Actuators B 81, 158164. Xia Y., Rogers J.A., Paul K.E., and Whitesides G.M. (1999) Unconventional methods for fabricating and patterning nanostructures. Chem. Rev. 99(7), 18231848. Xia Y., Wiesinger J.M., MacDiarmid A.G., and Epstein A.J. (1995) Camphorsulfonic acid fully doped polyaniline emeraldine salt: Conformations in different solvents studied by an ultraviolet/visible/near-infrared spectroscopic method. Chem. Mater. 7, 443445. Xie D., Jiang Y., Pan W., Li D., Wu Z., and Li Y. (2002) Fabrication and characterization of polyaniline-based gas sensor by ultra-thin film technology. Sens. Actuators B 81, 158164. Yadong J., Tao W., Zhiming W., Dan L., Xiangdong C., and Dan X. (2000) Study on the NH3-gas sensitive properties and sensitive mechanism of polypyrrole. Sens. Actuators B 66, 280282. Yan H. and Liu C. (1993) Humidity effects on the stability of a solid polymer electrolyte oxygen sensor. Sens. Actuators B 10, 133136. Yan X.B., Han Z.J., Yang Y., and Tay B.K. (2007) NO2 gas sensing with polyaniline nanofibers synthesized by a facile aqueous/organic interfacial polymerization. Sens. Actuators B 123, 107113. Yang H. and Bard A.J. (1992) The application of fast scan cyclic voltammetry. Mechanistic study of the initial stage of electropolymerization of aniline in aqueous solutions. J. Electroanal. Chem. 339(12), 423449. Yang H., Wan J., Shu H., Liu X., Lakshmanan R.S., Guntupalli R., Hu J., Howard W., and Chin B.A. (2006) Hydrazine leak detection using poly (3-hexylthiophene) thin-film micro-sensor. Proc. SPIEInt. Soc. Opt. Eng. 6222(1), 62220S-8. Yue J. and Epstein A.J. (1990) Synthesis of self-doped conducting polyaniline. J. Am. Chem. Soc. 112, 28002801. Zawodzinski T.A., Springer T.E., Uribe F., and Gottesfeld S. (1993) Characterization of polymer electrolytes for fuel cell applications. Solid State Ionics 60, 199211. Zee F. and Judy J.W. (2001) Micromachined polymer-based chemical gas sensor array. Sens. Actuators B 72, 120128. Zhang X.B., Guo C.C., Jian L.X., Shen G.L., and Yu R.Q. (2000) Fluoroborate ion sensitive PVC membrane electrode based on chloro[tetra(m-amino-phenyl)porphinato]-manganese as neutral carrier. Anal. Chim. Acta 419, 227233. Zhang Y., Murphy C.B., and Jones W.E. (2002) Poly[p-(phenyleneethynylene)-alt-(thienyleneethynylene)] polymers with oligopyridine pendant groups: Highly sensitive chemosensors for transition metal ions. Macromolecules 35(3), 630636. Zheng W., Min Y., MacDiarmid A.G., Angelopoulos M., Liao Y.H., and Epstein A.J. (1997) Effect of organic vapors on the molecular conformation of nondoped polyaniline. Synth. Met. 84, 6364. Zhou R., Haug M., Geckeler K.E., and Gopel W. (1993) NOx sensitivity of monomeric and polymeric N-macrocyclic compounds. Sens. Actuators B 16, 312316. Zhou G., Cheng Y.X., Wang L.X., Jing X.B., and Wang F. (2005) Novel polyphenylenes containing phenol-substituted oxadiazole moieties as fluorescent chemosensors for fluoride ion. Macromolecules 38(6), 21482153.

You might also like