Chemical Sensors: Simulation and Modeling Volume 2: Conductometric-Type Sensors

You might also like

Download as pdf
Download as pdf
You are on page 1of 69

Type Sensors

Phenomenological
Approachesused
the surroundings
for chemical sensing.
About the e
Hisresearchactivities
MoldovanAcademies
CHEMICAL
Avolumeinthe
v
CONTENTS
Preface xiii
about the editor xvii
contributors xix
1 numerical simulation of electrical resPonses to Gases in
advanced structures 1
A. etkus
1 Introduction 1
2 Analytic and Numeric Modeling 3
2.1 Basic Equations 3
2.2 Analytical Approaches 7
2.3 Numerical Simulations 15
2.4 Verifcation of Models 22
3 Resistive Sensors 28
3.1 Introductory Remarks 28
3.2 Polycrystalline Films 29
3.3 Nanostructured Films 34
3.4 Conductive Polymer Layers 38
3.5 Molecular Structures 40
4 Concluding Comments 43
References 44
2 co-adsorPtion Processes and Quantum mechanical modelinG of
Gas-sensinG effects 51
J.-J. Velasco-Vlez
1 Introduction 51
2 SolidGas Interaction 55
vi contents
2.1 Adsorption 55
2.2 Chemisorption 57
2.3 Electronic Transitions in Chemisorption 65
2.4 Chemisorption in Equilibrium, Wolkenstein Isotherm 67
2.5 Reaction Time 71
2.6 Charge Transfer Model (CTM) 73
3 Co-adsorption 79
3.1 Quantum Model 79
3.2 Statistical Model 81
3.3 Adsorption Time 83
4 Discussion 84
5 Summary 90
6 Nomenclature 92
Dedication 93
Acknowledgment 93
References 94
3 nanosensors: a Platform to model the sensinG mechanisms in
metal oxides 97
F. Hernandez-Ramirez
J. D. Prades
A. Cirera
1 Introduction 97
2 Toward a Better Description of Gas-Sensing Mechanisms in
Metal Oxides: Oxygen Diffusion in Tin Dioxide Nanowires 100
2.1 Description of Oxygen Sensing Using Diffusion 105
2.2 Summary 107
3 Toward a Systematic Understanding of Photo-Activated Gas
Sensors 108
3.1 Experimental Background 109
3.2 Theoretical Model of the Photo-Activated Response to
Oxidizing Gases (NO
2
) 112
3.3 Comparison with Experiments 116
3.4 Other Target Gases 119
3.5 Summary 119
4 Conclusions 120
Acknowledgments 120
References 120
contents vii
4 surface state models for conductance resPonse of metal
oxide Gas sensors durinG thermal transients 127
A. Fort
M. Mugnaini
S. Rocchi
V. Vignoli
1 Introduction 127
2 Surface-StateBased Models of Resistive Chemical Sensors 129
2.1 Depleted Surface 135
2.2 Enhanced Surface 141
3 Building a Chemical-Physical Sensor Model: From the Chemistry
to the Resistance Variations 143
3.1 The Mechanism of Conduction in the Film: Effect of the
Film Structure 143
3.2 Selection of a Model for Surface Potential Barrier Height as
a Function of the Surface Charge: Solution of the Poisson
Equation 146
3.3 Selection of a Model for the Evolution of the Surface Charge
as a Function of the Surface Chemical Reactions 151
4 Surface StateBased Models for Chemical Resistive Sensors:
Different Assumptions and Points of View 155
5 Developing a Treatable Gray Model from the Physical-Chemical
Model 156
5.1 The Intrinsic Model 157
5.2 The Extrinsic Model: Contributions from Oxygen and
Reducing Gas 159
5.3 Effects of Water Vapor 161
6 Conclusions 168
Nomenclature 169
References 170
5 conductance transient analyses of metal oxide Gas sensors
on the examPle of sPinel ferrite Gas sensors 177
K. Mukherjee
S. B. Majumder
1 Introduction 177
2 Salient Features of GasSolid Interaction during Gas Sensing 178
3 Experimental 181
viii contents
4 Modeling the Conductance Transients during Response and
Recovery 182
4.1 Derivation of Response and Recovery Conductance
Transients Based on Langmuir Adsorption Isotherm 184
4.2 Nonlinear Fitting of Response and Recovery Transients 187
4.3 Variation of Response and Recovery Time Constants with
Sensor Operating Temperature 189
4.4 Variation of the Estimated Fitted Parameters with Test Gas
Concentration: Addressing the Selectivity Issue 193
5 Characteristic Features Observed in Resistance Transients 196
5.1 Investigations on Irreversible and Reversible Gas Sensing in
Oxide Gas Sensors 198
5.2 Periodic Undulation of the Resistance Transients during
Response and Recovery 211
5.3 Spikelike Features in Resistance Transients 218
6 Summary and Conclusions 222
7 Appendix 223
7.1 Solution of Eq. (5.36) 223
7.2 Solution of Eq. (5.42) 224
8 Nomenclature 226
Acknowledgment 228
References 228
6 model of thermal transient resPonse of semiconductor
Gas sensors 233
Akira Fujimoto
1 Introduction 233
2 Improvement in Selectivity of the Semiconductor Gas Sensor
Using Transient Response 234
3 Model of Thermal Transient Response of Semiconductor Gas
Sensors 236
3.1 Transient Response of Semiconductor Gas Sensors 236
3.2 Thermal Transient Response of Semiconductor Gas Sensors 237
3.3 Physical and Chemical Processes in the Semiconductor
Gas Sensor Under Transient Response 238
4 Modeling of Gas Sensor Processes 239
4.1 Heat Conduction Processes 239
4.2 Chemical Reaction Processes 242
4.3 Diffusion Processes 243
4.4 Sensor Output 244
contents ix
5 Calculation Methods 245
5.1 Heat Conduction 245
5.2 Gas Concentrations on the Sensor Surface 246
6 Calculated Transient Responses of Gas Sensors 247
6.1 Temperature Change on the Sensor Surface 247
6.2 Concentration Change of Substance in the Vicinity of the
Sensor Surface 249
6.3 Comparison to Experimental Results 250
7 Application of the Model of Transient Response 251
7.1 Transient Responses Under Heating with Various Waveforms 251
7.2 Activation Energy Dependence of Transient Response 253
8 Conclusions 257
References 258
7 exPerimental investiGation and modelinG of Gas-sensinG effect
in mixed metal oxide nanocomPosites 261
L. I. Trakhtenberg
G. N. Gerasimov
V. F. Gromov
M. A. Kozhushner
O. J. Ilegbusi
1 Introduction 261
2 Types of Mixed Metal Oxides 263
3 Synthesis of Metal Oxide Nanocomposites 264
4 Charge Transfer Processes and Conductivity 265
5 Conductivity Mechanism 267
6 Sensor Properties 269
7 Mechanism of Sensor Effect 272
7.1 Sensors Based on Single Nanofbers 272
7.2 Polycrystalline Sensors 275
8 Modeling of the Sensory Effect for Reduced Gases 283
8.1 Qualitative Discussion of the Sensory Mechanism 283
8.2 Equilibrium Electronic Characteristics Of SnO
2
284
8.3 Sensor Response 285
9 Conclusions 290
Acknowledgment 291
References 291
x contents
8 the influence of Water vaPor on the Gas-sensinG Phenomenon
of tin dioxidebased Gas sensors 297
R. G. Pavelko
1 Introduction 297
2 Direct Water Effects on Tin DioxideBased Gas Sensors 299
2.1 Undoped SnO
2
299
2.2 Doped SnO
2
307
3 Indirect Water Effects on Tin DioxideBased Gas Sensors 310
3.1 Reducing Gases 311
3.2 Oxidizing Gases 322
4 Phenomenological Model 323
5 Conclusions 330
Acknowledgments 330
References 330
9 comPutational desiGn of chemical nanosensors:
transition metaldoPed sinGle-Walled carbon nanotubes 339
Duncan J. Mowbray
Juan Mara Garca-Lastra
Iker Larraza Arocena
ngel Rubio
Kristian S. Thygesen
Karsten W. Jacobsen
1 Introduction 339
2 TM-Doped SWNTs as Nanosensors 342
3 Density Functional Theory 346
4 Kinetic Modeling 351
5 Nonequilibrium Greens Function Methodology 355
5.1 Divacancy II 358
5.2 Divacancy I 361
5.3 Monovacancy 363
5.4 Target and Background Molecules 364
6 Sensing Property 369
7 Conclusions 372
Acknowledgments 373
References 373
contents xi
10 al-doPed GraPhene for ultrasensitive Gas detection 379
Z. M. Ao
Q. Jiang
S. Li
1 Emerging Graphene-Based Gas Sensors 379
1.1 The Role of Aluminum Doping in Sensing Applications 380
2 Aluminum-Doped Graphene for CO Detection 381
2.1 Sensitivity Enhancement of CO Detection in Aluminum-
Doped Graphene 381
2.2 Effect of Electric Field on CO Detection 387
2.3 Effect of Temperature on CO Detection 393
3 Aluminum-Doped Graphene for Formaldehyde Detection 399
3.1 Adsorption Enhancement with Aluminum Doping 399
3.2 Variation of Electronic Properties Induced by Adsorption 402
4 Aluminum-Doped Graphene for Detection of HF Molecules 404
4.1 Adsorption Enhancement of Aluminum-Doped Graphene 406
4.2 Adsorption Enhancement Mechanism 410
4.3 Effect of Electric Field on Adsorption 410
5 Conclusion and Future Challenges 411
Acknowledgments 413
References 413
11 Physics-based modelinG of sno
2
Gas sensors With field-effect
transistor structure 419
P. Andrei
L. L. Fields
A. J. Soares
R. J. Perry
Y. Cheng
P. Xiong
J. P. Zheng
1 Introduction 419
2 Physics-Based Modeling of the Nanobelts 421
3 Model Calibration 425
4 Analytical Model for Nanobelt Sensors 430
4.1 Case 1: Nanobelt with Ohmic Contacts in the Presence of
Hydrogen 431
4.2 Case 2: Nanobelt with Ohmic Contacts in the Presence of
Oxygen 435
xii contents
4.3 Case 3: Nanobelt with Schottky Contacts in the Presence of
Oxygen 439
5 Conclusion 440
Appendix: Fabrication and Experimental Data 441
References 443
12 modelinG and simulation of nanoWire-based field-effect
biosensors 447
S. Baumgartner
M. Vasicek
C. Heitzinger
1 Introduction 447
2 Homogenization 450
3 The Biofunctionalized Boundary Layer 452
3.1 The Site-Dissociation Model 453
3.2 Screening and Biomolecules 454
3.3 Summary 460
4 The Current Through the Nanowire Transducer 461
4.1 The Drift-Diffusion-Poisson System 461
4.2 Self-Consistent Simulations of Sensor Systems 462
5 Summary 464
Acknowledgment 465
References 465
index 471
xiii
PREFACE
This series, Chemical Sensors: Simulation and Modeling, is the perfect comple-
ment to Momentum Presss six-volume reference series, Chemical Sensors:
Fundamentals of Sensing Materials and Chemical Sensors: Comprehensive Sensor
Technologies, which present detailed information about materials, technologies,
fabrication, and applications of various devices for chemical sensing. Chemical
sensors are integral to the automation of myriad industrial processes and every-
day monitoring of such activities as public safety, engine performance, medical
therapeutics, and many more.
Despite the large number of chemical sensors already on the market, selec-
tion and design of a suitable sensor for a new application is a diffcult task for
the design engineer. Careful selection of the sensing material, sensor platform,
technology of synthesis or deposition of sensitive materials, appropriate coatings
and membranes, and the sampling system is very important, because those deci-
sions can determine the specifcity, sensitivity, response time, and stability of the
fnal device. Selective functionalization of the sensor is also critical to achieving
the required operating parameters. Therefore, in designing a chemical sensor, de-
velopers have to answer the enormous questions related to properties of sensing
materials and their functioning in various environments. This fve-volume com-
prehensive reference work analyzes approaches used for computer simulation and
modeling in various felds of chemical sensing and discusses various phenomena
important for chemical sensing, such as surface diffusion, adsorption, surface
reactions, sintering, conductivity, mass transport, interphase inter actions, etc.
In these volumes it is shown that theoretical modeling and simulation of the pro-
cesses, being a basic for chemical sensor operation, can provide considerable
assistance in choosing both optimal materials and optimal confgurations of
sensing elements for use in chemical sensors. The theoretical simulation and
model ing of sensing material behavior during interactions with gases and liquid
surroundings can promote understanding of the nature of effects responsible for
high effectiveness of chemical sensors operation as well. Nevertheless, we have to
understand that only very a few aspects of chemistry can be computed exactly.
xiv preface
However, just as not all spectra are perfectly resolved, often a qualitative or ap-
proximate computation can give useful insight into the chemistry of studied phe-
nomena. For example, the modeling of surface-molecule interactions, which can
lead to changes in the basic properties of sensing materials, can show how these
steps are linked with the macroscopic parameters describing the sensor response.
Using quantum mechanics calculations, it is possible to determine parameters
of the energetic (electronic) levels of the surface, both inherent ones and those
introduced by adsorbed species, adsorption complexes, the precursor state, etc.
Statistical thermodynamics and kinetics can allow one to link those calculated
surface parameters with surface coverage of adsorbed species corresponding to
real experimental conditions (dependent on temperature, pressure, etc.). Finally,
phenomenological modeling can tie together theoretically calculated characteris-
tics with real sensor parameters. This modeling may include modeling of hot plat-
forms, modern approaches to the study of sensing effects, modeling of processes
responsible for chemical sensing, phenomenological modeling of operating char-
acteristics of chemical sensors, etc.. In addition, it is necessary to recognize that
in many cases researchers are in urgent need of theory, since many experimental
observations, particularly in such felds as optical and electron spectroscopy, can
hardly be interpreted correctly without applying detailed theoretical calculations.
Each modeling and simulation volume in the present series reviews model-
ing principles and approaches particular to specifc groups of materials and de-
vices applied for chemical sensing. Volume 1: Microstructural Characterization and
Modeling of Metal Oxides covers microstructural characterization using scanning
electron microscopy (SEM), transmission electron spectroscopy (TEM), Raman
spectroscopy, in-situ high-temperature SEM, and multiscale atomistic simulation
and modeling of metal oxides, including surface state, stability, and metal oxide
interactions with gas molecules, water, and metals. Volume 2: Conductometric-
Type Sensors covers phenomenological modeling and computational design of
conductometric chemical sensors based on nanostructured materials such as
metal oxides, carbon nanotubes, and graphenes. This volume includes an over-
view of the approaches used to quantitatively evaluate characteristics of sensitive
structures in which electric charge transport depends on the interaction between
the surfaces of the structures and chemical compounds in the surroundings.
Volume 3: Solid-State Devices covers phenomenological and molecular model-
ing of processes which control sensing characteristics and parameters of various
solid-state chemical sensors, including surface acoustic wave, metal-insulator-
semiconductor (MIS), microcantilever, thermoelectric-based devices, and sensor
arrays intended for electronic nose design. Modeling of nanomaterials and nano-
systems that show promise for solid-state chemical sensor design is analyzed as
well. Volume 4: Optical Sensors covers approaches used for modeling and simu-
lation of various types of optical sensors such as fber optic, surface plasmon
resonance, Fabry-Prot interferometers, transmittance in the mid-infrared region,
preface xv
luminescence-based devices, etc. Approaches used for design and optimization
of optical systems aimed for both remote gas sensing and gas analysis cham-
bers for the nondispersive infrared (NDIR) spectral range are discussed as well.
A description of multiscale atomistic simulation of hierarchical nanostructured
materials for optical chemical sensing is also included in this volume. Volume 5:
Electrochemical Sensors covers modeling and simulation of electrochemical pro-
cesses in both solid and liquid electrolytes, including charge separation and
transport (gas diffusion, ion diffusion) in membranes, protonelectron transfers,
electrode reactions, etc. Various models used to describe electrochemical sensors
such as potentiometric, amperometric, conductometric, impedimetric, and ion-
sensitive FET sensors are discussed as well.
I believe that this series will be of interest of all who work or plan to work in
the feld of chemical sensor design. The chapters in this series have been prepared
by well-known persons with high qualifcation in their felds and therefore should
be a signifcant and insightful source of valuable information for engineers and
researchers who are either entering these felds for the frst time, or who are al-
ready conducting research in these areas but wish to extend their knowledge in
the feld of chemical sensors and computational chemistry. This series will also be
interesting for university students, post-docs, and professors in material science,
analytical chemistry, computational chemistry, physics of semiconductor devices,
chemical engineering, etc. I believe that all of them will fnd useful information in
these volumes.
G. Korotcenkov
xvii
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 scientifc Gas Sensor Group and manager of
various national and international scientifc and engineering projects carried out
in the Laboratory of Micro- and Optoelectronics, Technical University of Moldova.
Currently, Dr. Korotcenkov is a research professor at the Gwangju Institute of
Science and Technology, Republic of Korea.
Specialists from the former Soviet Union know Dr. Korotcenkovs research
results in the feld of 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
nanostructured metal oxides and solid-state gas
sensor design. Dr. Korotcenkov is the author or
editor of 11 books and special issues, 11 invited
review papers, 17 book chapters, and more than
190 peer-reviewed articles. He holds 18 patents,
and he has presented more than 200 reports at
national and international conferences.
Dr. Korotcenkovs research activities have
been honored by an Award of the Supreme
Council of Science and Advanced Technology
of the Republic of Moldova (2004), The Prize of
the Presidents of the Ukrainian, Belarus, and
Moldovan Academies of Sciences (2003), Senior
Research Excellence Awards from the 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.
xix
CONTRIBUTORS
Arnas etkus (Chapter 1)
Department of Physical Technologies
Center for Physical Sciences and Technology
Vilnius LT01108, Lithuania
Juan-Jess Velasco-Vlez (Chapter 2)
Materials Sciences Division
Large Lawrence Berkeley National Laboratory
Berkeley, California 94720, USA
Francisco Hernandez-Ramirez (Chapter 3)
Institut de Recerca en Energia de Catalunya (IREC)
Barcelona, Spain
and
Departament dElectrnica
Universitat de Barcelona
Barcelona, Spain
J. Daniel Prades (Chapter 3)
Departament dElectrnica
Universitat de Barcelona
Barcelona, Spain
Albert Cirera (Chapter 3)
Departament dElectrnica
Universitat de Barcelona
Barcelona, Spain
Ada Fort (Chapter 4)
Information Engineering Department
University of Siena
53100 Siena, Italy
xx contributors
Marco Mugnaini (Chapter 4)
Information Engineering Department
University of Siena
53100 Siena, Italy
Santina Rocchi (Chapter 4)
Information Engineering Department
University of Siena
53100 Siena, Italy
Valerio Vignoli (Chapter 4)
Information Engineering Department
University of Siena
53100 Siena, Italy
Kalisadhan Mukherjee (Chapter 5)
Materials Science Centre
Indian Institute of Technology
Kharagpur 721302, India
Subhasish Basu Majumder (Chapter 5)
Materials Science Centre
Indian Institute of Technology
Kharagpur 721302, India
Akira Fujimoto (Chapter 6)
Wakayama National College of Technology
Nadacho, Gobo-shi 644-0023 Japan
Leonid I. Trakhtenberg (Chapter 7)
Semenov Institute of Chemical Physics
Russian Academia of Sciences
Moscow 119991, Russia
Genrikh N. Gerasimov (Chapter 7)
Semenov Institute of Chemical Physics
Russian Academia of Sciences
Moscow 119991, Russia
Vladimir F. Gromov (Chapter 7)
Semenov Institute of Chemical Physics
Russian Academia of Sciences
Moscow 119991, Russia
contributors xxi
Mortko A. Kozhushner (Chapter 7)
Semenov Institute of Chemical Physics
Russian Academia of Sciences
Moscow 119991, Russia
Olusegun J. Ilegbusi (Chapter 7)
University of Central Florida
Orlando, Florida 32816-2450, USA
Roman G. Pavelko (Chapter 8)
Department of Energy and Material Sciences
Faculty of Engineering Sciences
Kyushu University
Kasuga-shi, Fukuoka 816-8580, Japan
Duncan J. Mowbray (Chapter 9)
NanoBio Spectroscopy Group and ETSF Scientifc Development Centre
Departamento de Fsica de Materiales
Universidad del Pas Vasco UPV/EHU and DIPC
E20018 San Sebastin, Spain
Juan Mara Garca-Lastra (Chapter 9)
NanoBio Spectroscopy Group and ETSF Scientifc Development Centre
Departamento de Fsica de Materiales, Centro de Fsica de Materiales CSICUPV/
EHUMPC and DIPC
Universidad del Pas Vasco UPV/EHU
E20018 San Sebastin, Spain
and
Center for Atomic-Scale Materials Design, Department of Physics
Technical University of Denmark
DK2800 Kgs. Lyngby, Denmark
Iker Larraza Arocena (Chapter 9)
NanoBio Spectroscopy Group and ETSF Scientifc Development Centre
Departamento de Fsica de Materiales
Universidad del Pas Vasco UPV/EHU
E20018 San Sebastin, Spain
ngel Rubio (Chapter 9)
NanoBio Spectroscopy Group and ETSF Scientifc Development Centre
Departamento de Fsica de Materiales, Centro de Fsica de Materiales CSICUPV/
EHUMPC and DIPC
Universidad del Pas Vasco UPV/EHU
E20018 San Sebastin, Spain
xxii contributors
Kristian S. Thygesen (Chapter 9)
Center for Atomic-Scale Materials Design, Department of Physics
Technical University of Denmark
DK2800 Kgs. Lyngby, Denmark
Karsten W. Jacobsen (Chapter 9)
Center for Atomic-Scale Materials Design, Department of Physics
Technical University of Denmark
DK2800 Kgs. Lyngby, Denmark
Zhimin Ao (Chapter 10)
School of Materials Science and Engineering
The University of New South Wales
Sydney, New South Wales 2052, Australia
Qing Jiang (Chapter 10)
Key Laboratory of Automobile Materials, Ministry of Education, and School of
Materials Science and Engineering
Jilin University
Changchun 130022, Peoples Republic of China
Sean Li (Chapter 10)
School of Materials Science and Engineering
The University of New South Wales
Sydney, New South Wales 2052, Australia
Petru Andrei (Chapter 11)
Department of Electric and Computer Engineering
Florida A&M UniversityFlorida State University College of Engineering
Tallahassee, Florida 32310, USA
Leonard L. Fields (Chapter 11)
Corning Inc.
Optical Physics and Networks Technology
Corning, New York 14831, USA
Antonio J. Soares (Chapter 11)
Department of Electronic Engineering Technology
Florida A&M University
Tallahassee, Florida 32301, USA
Reginald J. Perry (Chapter 11)
Department of Electric and Computer Engineering
Florida A&M UniversityFlorida State University College of Engineering
Tallahassee, Florida 32310, USA
contributors xxiii
Yi Cheng (Chapter 11)
Institute for Systems Research (ISR)
University of Maryland
College Park, Maryland 20742, USA
Peng Xiong (Chapter 11)
Department of Physics and Integrative NanoScience Institute (INSI)
Florida State University
Tallahassee, Florida 32306, USA
Jianping Zheng (Chapter 11)
Department of Electric and Computer Engineering
Florida A&M UniversityFlorida State University College of Engineering
Tallahassee, Florida 32310, USA
Stefan Baumgartner (Chapter 12)
Department of Mathematics
University of Vienna
1010 Vienna, Austria
and
AIT Austrian Institute of Technology
Vienna, Austria
Martin Vasicek (Chapter 12)
Department of Mathematics
University of Vienna
1010 Vienna, Austria
and
Wolfgang Pauli Institute c/o Department of Mathematics
University of Vienna
1010 Vienna, Austria
Clemens Heitzinger (Chapter 12)
Department of Applied Mathematics and Theoretical Physics (DAMTP)
University of Cambridge
Cambridge CB2 1TN, United Kingdom
and
Department of Mathematics
University of Vienna
1010 Vienna, Austria
and
AIT Austrian Institute of Technology
Vienna, Austria
1 DOI: 10.5643/9781606503140/ch1
Chapter 1
NUMERICAL SIMULATION OF ELECTRICAL
RESPONSES TO GASES IN ADVANCED
STRUCTURES
a. etkus
1. IntroductIon
Research and development of gas sensors stands mainly on known technologies
that implement fundamental principles of the conversion of chemical interaction
into a change of physical properties. For most practical applications it is essential
to produce well-functioning and stable devices, but detailed models seem hardly
necessary, because empirical approaches allow one to accomplish this task suc-
cessfully. However, it seems evident that understanding the key mechanisms of
the response, and having a fundamental description of the processes involved,
will make it possible to better defne the targets of research and development work
as well as to evaluate expectable progress in the modifcation, improvement, and
optimization of gas sensors.
Depending on the varying physical properties, proposed devices can be divided
into a few main classes, namely, electrical, optical, and mechanical gas-sensitive
solid structures. Analysis of the processes and models of the mechanisms require
different fundamental approaches and tools of description for these three classes
of gas sensors, which exceeds the limits of this survey. This chapter is devoted
only to sensors in which the electrical properties of sensitive materials depend on
2 ChemiCal sensors moDelinG anD simUlation: VolUme 2
the infuence of the gas. Metal oxides are frequently used as the sensitive material
in these sensors.
In general, metal oxide gas sensors should be assumed to be partly electronic
conductors and partly ionic conductors. Depending on the dominating component
of conductivity, the oxides are accepted as being typical ionic conductors (Y
2
O
3
,
ZrO
2
, etc.) or typical electronic conductors (SnO
2
, In
2
O
3
, WO
3
, etc.). For certain
ceramics that are typically based on transition metals, both the electronic and
ionic conductivities have to be considered in the analysis of the electrical prop-
erties. From the viewpoint of practical applications, ionic and mixed conductors
are typically used in the development of oxygen gas sensors, while metal oxides
with dominant electronic conductance are widely used in the development of vari-
ous gas sensors for diverse odor-detection systems. Though general principles of
simulation are analogous for all the metal oxides, in the present review, only the
metal oxides with dominating electronic conductance are discussed. For readers
who are interested in ionic conductors and applications of these oxides, it may be
useful to start with recent publications such as those of Fergus (2008), Zhulykov
(2008), Rder-Roith et al. (2009), Hubert et al. (2011), and Schonauer et al. (2011).
In metal oxides with dominating electronic conductance, the response to gas
is determined by the changes in the electrical charge transport produced by the
interactions between the surfaces and gases. In scientifc publications, these sen-
sors are typically called conductive, conductometric, and resistive sensors. In this
chapter, this type of sensor is preferably called a resistive gas sensor.
For the classical resistive sensors, the response to gas is determined by a se-
ries of interrelated processes occurring in the heterogeneous system that includes
the gas medium, the interface region, and the semiconducting material. The com-
plete model of the system must combine the descriptions of diverse mechanisms
that are highly specifc to the individual parts of the system. Well-defned and
justifed connections among these parts are crucial for development of simulation
models for gas sensors. This chapter represents an attempt to overview the publi-
cations containing suggestions about the simulation of electrical responses to gas
and to arrange the known approaches in some overall picture that explains the
fundamentals of functioning of these advanced gas sensors. In this chapter, we
frst discuss the basic equations defning the processes of sensor functioning. It is
shown that using special simplifcations, analytical solutions can be obtained for
these equations and adapted for simulation of the sensor response. In more com-
plicated situations the sensor response can be simulated by numerical methods.
The ways used to verify the simulation models are also reviewed in this chapter.
This chapter also includes several sections in which some specifc aspects of the
response simulations are discussed for the polycrystalline metal oxide sensors,
nanostructured flms, conductive polymers, and molecular sensors. Finally, some
general concluding comments about possibilities to simulate both existing and
emerging resistive gas sensors are presented.
nUmeriCal simUlation of eleCtriCal responses to Gases 3
2. AnAlytIc And numerIc modelIng
2.1. Basic Equations
Quantitative description of electrical properties of solid-state chemical sensors
requires deep understanding of a complete picture of the physical processes that
determine the response of these sensors to external chemical infuence. Details
about diverse aspects of the response mechanism make it possible to develop the
most acceptable model for analysis of experimental data. Theoretical description
of the response model gives the basis for numerical evaluation of the properties
under investigation. Consequently, one needs either to develop an original ap-
proach or choose an already-known approach to theoretical modeling of the func-
tioning of chemical sensors, aiming to reveal the most important factors in the
sensor technology and the methods of application. This overview is focused on the
electrical properties of semiconducting gas sensors, though general approaches
to the numerical evaluation of the parameters may be acceptable for other classes
of sensors.
It must be noted here that, to date, there were no detailed studies analyzing
proportions between the ionic and electronic components in metal oxide gas sen-
sors. On the other hand, the known experimental facts have been successfully
explained and described by models based on only the electronic conduction com-
ponent. This approach to charge transport is suffciently good for the synthesized
ceramic metal oxide sensors or even for thin-flm gas sensors with thicknesses
greater than about 100 nm. On the other hand, there is experimental data about
the specifc role of ionic transport in nanostructured materials, which, conse-
quently, reduces the ability to rely on numerical simulation of electrical properties
based on classical approaches in these nanostructured metal oxide gas sensors.
These special aspects will be discussed later in this chapter.
Classical models of metal oxide gas sensors are based on the semiconduct-
ing properties of nonstoichiometric metal oxides. The oxygen vacancies are typi-
cally associated with donor-type point defects in these oxides. The shallow donor
levels are completely oxidized at temperatures above 300 K and provide the free
electrons in the conduction band in these materials. In general, in the presence
of an electric feld

E , the electrical charge transport is determined by the drift of


carriers and the diffusion due to nonuniform distribution of carriers in semicon-
ducting materials. Thus, the current density

j is equal to
= +


q
j q n E q D n (1.1)

In Eq. (1.1), the frst term describes the drift component of conductivity while
the second one describes the diffusion component. These two components in-
clude specifc parameters: q, the electrical charge of a single carrier; n, the carrier
4 ChemiCal sensors moDelinG anD simUlation: VolUme 2
concentration; m, the mobility; and D
q
, the diffusion coeffcient. Typically, the
diffusion term is omitted in the models describing the response mechanisms in
metal oxide gas sensors. However, this term can be important at least for those
sensors in which the layer controlled by the surface potential is comparatively
thick with respect to the dimensions of the conductive channel. Therefore, the
Eq. (1.1) can be considered in metal oxide sensors in which the Debye screen-
ing length is comparable to the dimensions of structural elements (e.g., grains in
polycrystalline oxides).
By omitting the diffusion term in Eq. (1.1), the simulation models are signif-
cantly simplifed. As a result, the theoretical description acceptable for numerical
evaluation of the parameters considers only the conductance s, which is defned
by the factor from the frst term of Eq. (1.1) as follows:
= q n (1.2)
In fact, Eq. (1.2) is typically the basis of simulation models for gas sensors that
have been presented in the literature to date. In Eq. (1.2), both the carrier con-
centration and the mobility can be dependent on the surface properties of the
semiconducting constructions. The exact form of the dependence is defned by the
individual model of gas sensor. However, two notes can be added here. First, in the
sensor models, the resistance R of gas-sensitive structures is frequently evaluated
instead of the conductance (R = l/Ss
1
, with length l and cross-sectional area S as
the geometric factors). Second, gas sensor models do not include an explicit analy-
sis of scattering mechanisms that determine the mobility in a homogeneous semi-
conductor. For example, the mobility is inversely proportional to the concentration
of ionized impurities (m ~ N
ii
1
) if the ionized impurities are the dominant scattering
mechanism in the transport of charge carriers. The effect of this scattering can be
illustrated quantitatively by well-known facts about crystalline semiconductors.
According to the experiments of Prince (1953), in p-Ge an increase in concentra-
tion of ionized impurities from 10
14
cm
3
to 10
17
cm
3
results in a decrease in the
electron mobility from about 3900 cm
2
V
1
s
1
to about 1500 cm
2
V
1
s
1
. An analo-
gous effect of the scattering on the mobility of charge carriers was also obtained for
doped silicon (Prince 1954). Since the mobilities of the charge carriers are signif-
cantly less studied in metal oxides, it is not possible to evaluate quantitatively the
infuence of the scattering effects on the mobility in homogeneous parts of metal
oxide gas sensors. Typically, it is assumed that the effect is negligible compared to
other mechanisms determining an electrical response to gas.
In metal oxide gas sensors the surface potential defnes both the charge trans-
port and the spatial variation of free carriers as well as ionized point defects. Most
simulation models includes Poissons equation, the solution of which describes a
dependence of the potential on the space charge density. In one dimensional form,
Poissons equation is written as
nUmeriCal simUlation of eleCtriCal responses to Gases 5

2
2
0
4
x
(1.3)
In Eq. (1.3), f is the electrostatic potential, r is the electrical charge density,
and e and e
0
are the relative permittivity of the material and the vacuum permit-
tivity (e
0
= 8.85 10
12
F/m), respectively. The exact defnition of the charge r
depends on the individual simulation model but, in general, the free carriers and
ionized impurities have to be considered as follows:

( )

+ -
= - - +
1 1
d a
q N N n p (1.4)
where N
d
and N
a
are the concentrations of donors and acceptors, respectively,
while n and p are the concentration of free charges carriers, namely, electrons
and holes, respectively. Simplifying the simulation of gas sensors, it is typically
assumed that (1) only one type of carriers (n or p) is present and (2) the concentra-
tion and distribution of ionized impurities is constant for all analyzed processes
in the sensors. It is reasonable to expect that acceptability of these simplifcations
should be carefully considered because it can result in crucial deviations in the
calculations of sensor characteristics.
In gas sensors at nonequilibrium conditions, the carrier densities within
a given unit of volume of the material varies as a function of time due to the
carrier transport, capture by the appearing surface states and release from the
disappearing states. A change of the surface states is produced by adsorption
chemisorption, desorption, and chemical reactions on the surfaces. The changes
in carrier densities with time are describe by the current continuity equations.
Considering the electronic-type metal oxides, in this work only n is assumed to
be important. Assuming analogy between optical generationrecombination pro-
cesses and the trappingreleasing phenomena for the surface states, the continu-
ity equation for the current density can be written as


= - + -

n
n n
j n
d a
t x
(1.5)
In Eq. (1.5), n is the concentration of electrons in the conduction band, j
n
is the
electron current, and d
n
and a
n
are the rates of release (delocalization) and capture
of electrons from/to the surface states, respectively. In general, it follows from the
continuity equation that if charge is moving out of a differential volume (i.e., di-
vergence of the current density is positive), then the amount of charge within that
volume will decrease, so that the rate of change of the charge density is negative.
Therefore the continuity equation amounts to a conservation of charge.
6 ChemiCal sensors moDelinG anD simUlation: VolUme 2
A change of free carriers in solid chemical sensors is typically related to the
chemical interaction between the surfaces and the surrounding particles. For
exam ple, the chemisorption of the atmosphere oxygen on metal oxide is associ-
ated to the localization of conductive electrons at the chemisorption sites on the
solid surface. Formation and the properties of the point surface defects depend
on the materials, adsorbed particles, and specifc processes. Therefore, the exact
description of the rates for released d
n
and captured electrons a
n
in Eq. (1.5) de-
pends on the individual model. Some general considerations about the capture
of free electrons by the surface chemisorption sites can be found in the study by
Wolkenstein (1991) but, in simplifed simulations, it can frequently be assumed
that a
n
is equal to the oxygen chemisorption rate and d
n
is equal to the sum of the
oxygen desorption and surface reaction rates.
In general, the kinetics of the surface coverage by the chemisorbed species
of gases can be described by a modifed rate equation based on a Langmuir ap-
proach as follows:
( )


= - - - -




des
ads 0 des OG G
0
exp
E N N
S N N N
t kT N
(1.6)
where N and N
0
are the densities of the chemisorbed gas species and the chemi-
sorption surface sites, respectively, F is the fux of particles hitting the solid sur-
faces, S
ads
is the sticking probability of gas particles, n
des
is the desorption rate,
E
des
is the desorption activation energy, n
OG
is the rate of chemical reaction be-
tween chemisorbed particles and the gas particles hitting the surface, and Q
G
is
the fux of gas particles hitting the solid surface. The third term in Eq. (1.6) de-
scribes a chemical reaction between two particles of different origin. One particle
is chemisorbed on the surface, while another hits the surface area close to the
chemisorption site. The Eq. (1.6) form without this third term is the most well
known in the surface science models, and it describes the dynamics of the surface
coverage determined by the adsorption and desorption processes.
The dynamic picture of adsorption and desorption for solid surfaces has been
thoroughly studied in surface science for a number of years. Therefore, the simu-
lation models for solid surface coverage with particles defne infuences of diverse
processes (see, e.g., Kreuzer 1990; Zhdanov 2002). The diffculties in simulation
of adsorptiondesorption kinetics were discussed by Zhdanov (2001). A theory
that incorporates the solid surface reconstruction phenomena occurring during
adsorption and desorption was proposed in terms of the Langmuir approach in
work by Cerofolini (2003). Based on statistical rate theory, simulation models
for adsorptiondesorption processes on heterogeneous surfaces were proposed
by Rudzinski et al. (2005). Panczyk and Rudzinski (2004) showed that applica-
tion of statistical rate theory to describe the kinetics of dissociative adsorption
leads to very fexible expressions which may account for the variety of physical
nUmeriCal simUlation of eleCtriCal responses to Gases 7
situations found in these systems. The desorption kinetics and a variation of the
solid surface coverage with gaseous species produced by desorption was explicitly
described by Payne et al. (2006).
The basic equations are used to defne the characteristics that have to be in-
cluded in consistent simulation models, allowing one to calculate the responses of
semiconducting sensors to gas and to evaluate the key characteristics of materi-
als and processes producing the most signifcant infuence on the parameters of
sensors. Using special approaches, the equations can be modifed and solved to
provide explicit relationships among the characteristics of the sensors.
2.2. analytical approachEs
The simulation models that provide analytical descriptions of the response to gas
are the most valuable results of studies on fundamental processes in gas sensors.
Based on these models, a quantitative description of gas sensor characteristics
can be obtained. However, in general, it is impossible to obtain an analytical
solution for a set of the equations that defnes the relationship between the gas
surface interaction and a change in the conductance of a solidstate construc-
tion. Typically, a series of specifc simplifcations is made in the mechanism and
processes, aiming to obtain an analytical form for description of parameters for
gas sensors. These simplifcations are dependent on the mechanisms selected
to describe the conversion of the gassurface interaction into the response and,
consequently, are justifed by the limitations that determine the acceptability of
the simulation model.
The complete model of the response mechanisms in metal oxide gas sensors
is still under discussion. In spite of this, the core aspects in the understanding of
the response of metal oxides to gases are commonly accepted and have been used
by sensor researchers and developers for more than 20 years. These classical ap-
proaches have been nicely presented in previous papers (e.g., Barsan et al. 2001,
2007; Oprea et al. 2009). We will use this classical fundamental understanding
of the response mechanisms in metal oxides for defnition of simulation models
resulting in analytical descriptions of the characteristics and a quantitative evalu-
ation of the dependencies between the sensor parameters and the core factors in
the technology as well as in functioning.
The classical approaches in the explanation of sensor functioning are based
on the polycrystalline structure of typical gas sensors. Supposing grains similar
to spheres with the same diameter, the sticking points between the grains are
accepted as the contacts through which electrical current fows. The electrical
charge on the surfaces of the grains can be changed by chemical interaction be-
tween the surface and gas. A variation of the surface charge produces changes in
the electron transport in the polycrystalline metal oxide. There are at least three
8 ChemiCal sensors moDelinG anD simUlation: VolUme 2
models explaining the relationship between the surface charge and the electron
transport, depending on which Eqs. (1.1), (1.2), (1.3), and (1.4) can be presented
in diverse forms acceptable for the simulation of the conductance. The most-
exploited approach proposes the contacts between the grains being similar to
the double Schottky junction. Thermally activated fow of electrons above the po-
tential barrier of the grain boundary is the core assumption in this approach.
Another approach includes an assumption about the coalescent grains. It follows
from this approach that the electrical charge fows in a channel with varying cross
section. The largest parts in the channel are equal to the diameter of grains, while
the necks in the channel appear at the junctions between the grains. The width
of the channel can be changed by the surface electrical charge. In this approach,
the basic equations have to be used to defne the relationship between the surface
charge and the width of the conductive channel. The third approach rests on the
assumption about the straightforward relationship between the free electron con-
centration in the bulk of the semiconductor and the adsorbed gas species on the
surfaces of sensor. Defnition of this relationship leads to comparatively simple
modifcation of the basic equations acceptable for quantitative description of the
sensor characteristics.
We limit ourselves to the double Schottky barrier approach in this chapter
because this approach is used more frequently in simulations than the other two.
Without deep analysis, a few remarks can be added here about the choice of ap-
proach. First, development of constructions with a continuous conductive channel
(without potential barriers at the junctions) requires special growth technologies
that allow one to obtain continuous metal oxide layers with the monocrystal-
line structure. To date, most sensor technologies used have been acceptable for
formation of polycrystalline materials. Second, supposing 1000 ppm for a gas,
the maximum density of chemisorbed gas species is about 10
14
cm
2
. Assuming
one extra free electron per each chemisorbed species, an increase in the free
electron density n
extra
can be estimated as 10
14
cm
3
. The concentration of free
electrons in gas-sensitive metal oxides n is typically estimated to be about 10
17

10
18
cm
3
. Consequently, an electron concentration response to gas can be about
n
extra
/n <10
3
and even less.
It must be also noted here that there are signifcant uncertainties that limit
acceptability of an analytical solution for the simulation of sensor properties in
the double Schottky barrier approach. There are ongoing discussions about the
surface properties of metal oxides. It has been shown by various studies that the
surface properties are dependent not only on the material but also on the struc-
ture of the solid. The details about these dependencies can be found in recent
publications, namely, for tin oxide in Batzill and Diebold (2005), for titanium
oxide in Diebold (2003), and for indium oxide in King et al. (2009) and ONeil et al.
(2010). Summarizing the studies on the surface properties, one can conclude that
there is no straightforward relationship between the density of surface oxygen and
nUmeriCal simUlation of eleCtriCal responses to Gases 9
an increase in the density of surface negative charge. Both depletion and accumu-
lation layers can be obtained on the surfaces of metal oxides. The type of surface
charge layer depends on the crystallographic planes of the oxides and the oxygen
adsorption modes [see, e.g., for SnO
2
, Sensato et al. (2002), and for In
2
O
3
, Walsh
(2011)]. Since the details of simulation depend on the exact model, classical prin-
ciples in the depletion-layer approach are accepted as the basis for the analysis of
the response of metal oxides to gas in this section.
In our approach, Eq. (1.1) must be modifed to describe electron transport
in conductive channels with potential barriers. Typically, it is supposed that the
electric charge is transported only by the conduction-band electrons with energies
higher than the height of the potential barrier. The thermally activated electron
transport can be described by two analogous forms, however, whose origin is dif-
ferent. The most typical approach is based on general understanding of electron
distribution with energies. Using the Boltzmann distribution, the concentration
of electrons with energies exceeding the height of the potential barrier V
b
can be
defned as



=




0
exp
b
c
eV
n n
kT
(1.7)
In Eq. (1.7), k is Boltzmanns constant, T is the temperature of the sensor,
and n
0
is the free electron concentration in the conductance band in the bulk of
the semiconductor. Neglecting the diffusion in Eq. (1.1), the density of current
through the intergranular potential barrier can be defned by



= =




0 0 0
exp
b
eV
j E qn
kT
(1.8)
In general, the characteristics of the junction barrier depend on various
para meters such as the position relative to the junction, ionized impurities, sur-
face charge, external electric feld, and so on. Therefore, an analytical descrip-
tion of V
b
can be obtained for some simplifed double Schottky model. Supposing
a one-dimensional approach with the position of the junction at x = 0 and com-
pletely ionized donor impurities outside the junction zone to the left (x < x
L
) and
the right (x < x
R
), the charge density in the barrier region can be described by the
following formula:
( ) ( ) ( ) ( )

= + - - -

D L R I
x eN x x x x Q x (1.9)
where N
D
defnes the donor concentration in the metal oxide bulk, (x) is the
Heaviside step function, x
L
and x
R
are the lengths of the left and the right depletion
10 ChemiCal sensors moDelinG anD simUlation: VolUme 2
region, respectively, and d(x) is the Dirac d function, Q
I
defnes the density of sur-
face charges and can be described by
( ) ( )

I
I I I
E
Q q dE N E f E (1.10)
In (1.10), N
I
(E) and E
I
are the density and the lowest energy of the electronic
traps in the gas chemisorption centers on the surface, respectively, and f
I
(E) is the
electron distribution function. Supposing fat bands outside the depletion region
and constant concentration of ionized impurities in the region and replacing (1.4)
with (1.9), it is possible to obtain an analytical solution for Eq. (1.3) similar to that
of Blatter and Greuter (1986a, 1986b). The solution is as follows:



= -



2
0
0
1
1
4
b b
b
V
V (1.11)
Here the barrier height without the external electrical feld is
( )

= =
2
0 0
0
2
I
b b
q D
Q
V V
qN
(1.12)
V is the drop of external electrical feld across the junction V = (x
L
+ x
R
)E. The ap-
plied voltage V
c
= fb
0
is often interpreted as the critical voltage for electrical break-
down in the single junction.
It must be noted here that the simulation of the response in metal oxides is
much easier if an external electric feld E is assumed comparatively low, so that
an infuence on the barrier height V
b
can be neglected. In the classical simulation
model this assumption is justifed by comparatively low measurement voltages
(~10 V), long distances between the contacts (~10
3
m), and high number of the
junctions between the contacts (~10
3
). Assuming the junction thickness is about
10
7
m, the electric feld across the single junction is about 10
5
V/m and practi-
cally negligible. However, for higher felds, the infuence of the applied external
voltage on the properties of the gas sensor can be considered as in Varpula et al.
(2008). More signifcant infuence of the feld can be detected in the nanostruc-
tured metal oxide gas sensors. Some aspects of this effect will be discussed below.
Neglecting the infuence of the external electric feld on the characteristics of
the junctions, Eq. (1.12) can be accepted as a defnition of the relationship be-
tween the surface charge and the barrier height that defnes the electron trans-
port in metal oxide gas sensors in (1.8). Here we can suppose that Q
I
is determined
only by the chemisorbed gas species on the surfaces of the sensor and Q
I
= qN
A
,
nUmeriCal simUlation of eleCtriCal responses to Gases 11
with N
A
the density of the chemisorbed species. Using the description of the resis-
tance R = L/(S) for a sample with length L and the cross section S, from (1.8) and
(1.12) one can obtain for the sensor resistance



2
2
0
0
exp
8
A
d
q
R R n
kTN
(1.13)
The defnition (1.13) is actually obtained for the static situation, after the
transient processes are fnished. Assuming the density of the chemisorbed gas
species to be proportional to the gas partial pressure in atmosphere, N
A
~ p
g
, a
dependence of the resistance response on the amount of gas in the air follows
from (1.13) if the adsorption is the core mechanism in the response. However, the
resistance response to gas is typically determined by the surface chemical reac-
tions between the chemisorbed oxygen species and the gas. Then, a relationship
between the gas and the surface charge has to be obtained and included in (1.13)
supposing that N
A
represents the density of chemisorbed oxygen species on the
surfaces of the metal oxide gas sensor.
In this simulation model we assume that a single type of oxygen species is
dominant on the surfaces of the metal oxide and all allowable chemisorption sites
can be occupied only by oxygen. Specifc sites exist on these surfaces for adsorp-
tion of gas that, however, do not create surface traps for electronsi.e., the surface
charge does not depend directly on adsorption of gas. The density of chemi sorbed
oxygen is determined by the equilibrium of the chemisorptiondesorption pro-
cesses under constant conditions. This initial state of the sensor surfaces defnes
the electrical properties of the metal oxide gas sensor before exposure to gas as it
was proposed by etkus (2002). The system sensoratmosphere is accepted as
being stable under equilibrium conditions.
Consider a response of a metal oxide sensor to a comparatively low amount
of impurity gas (reducing) in the atmosphere. We suppose here that the partial
pressure of the impurity gas P
gas
<< P
O2
in the atmosphere and a steplike change
in the atmosphere composition occurs. After the change in the composition of the
atmosphere, the amount of oxygen remains unchanged, but a fxed amount of the
target gas P
gas
is mixed into the atmosphere. Heterogeneous catalytic reactions
typically take place on the solid surfaces of the sensor. These reactions can be de-
scribed as a sequence of a few elementary steps, namely, the adsorption (chemi-
sorption), desorption, and Langmuir-Hinshelwood (LH) bimolecular interaction,
including adsorbed species of oxygen and gas. A solution of the rate equations
for these steps results in a relationship between the surface electrical charge and
current in the metal oxide and, consequently, defnes the simulation model of the
sensor response to gas.
On real surfaces, these equations are typically complicated by terms related
to adsorbateadsorbate lateral interaction (including electrostatic interaction),
12 ChemiCal sensors moDelinG anD simUlation: VolUme 2
surface heterogeneity, spontaneous and adsorbate-induced changes in the sur-
faces, and/or limited mobility of adsorbed species. In spite of this, the basic prop-
erties can frequently be simulated by a model of an ideal adsorbed layer, when the
surface is uniform and stable and there is no adsorbateadsorbate lateral inter-
action. These simplifcations make it possible to obtain analytical solutions for the
model equations.
Considering our simplifcations, the surface reactions that produce a response
by the sensor to gas can formally be split into the following steps:

- -
+
1 1
gas ads
O O e (1.14)

gas ads
G G (1.15)
( )
- -
+ +
1 1
ads ads
gas
O G OG e (1.16)
Step 1 (1.14) represents the chemisorption of oxygen, while the interaction
between the target gas (G
gas
) and the surface is described in step 2 (1.15). Step 3
of the bimolecular interaction (1.16) includes the surface species of oxygen and
gas. It is typically assumed that the product (OG)
gas
is removed from the surface
to the atmosphere at once.
According to the common understanding of heterogeneous catalysis (see, e.g.,
Kreuzer 1990; Xu and Koel 1994; Carlsson and Madix 2001; Busse et al. 2001;
Zhdanov 2002), the gaseous species are adsorbed on the surfaces prior to the
LH reaction in step 3. Usually, as in Lantto and Romppainen (1987), Clifford and
Tuma (1982/1983), Gardner (1990), Rantala et al. (1993), Simon et al. (2001),
Sakai et al. (2001), and Nakata et al. (2002), it is supposed for the sensor response
that the surface coverage with gases is determined by the adsorptiondesorption
equilibrium prior to step 3 (1.16). Based on this assumption, the simulations of
the sensor response are crucially simplifed, excluding a series of processes except
for step 3 (1.16). Such simplifcations allow one to limit the response description
to a comparatively simple rate equation and to obtain a solution that typically is
a pure exponential transient for the response signal.
In a more realistic approach, an injection of impurity gas is associated with
changes in the coverage of the sensor surfaces by both the oxygen and the impu-
rity gas species. Taking into account the steps of the surface chemical reactions in
(1.14)(1.16), two rate equations can be written for the oxygen coverage (Q
O
) and
the impurity gas coverage (Q
G
) in a form analogous to that of Nakata et al. (2002),
Carlsson and Madix (2001), Busse et al. (2001), and Kreuzer (1990). These rate
equations are as follows:

( )
( ) ( ) ( ) ( )



= - - -

O
O O O0 O O O OG O G
1
d t
F c S t t t t
dt
(1.17)
nUmeriCal simUlation of eleCtriCal responses to Gases 13

( )
( ) ( ) ( ) ( )



= - - -

G
G G G0 G G G OG O G
1
d t
F c S t t t t
dt
(1.18)
In Eqs. (1.17) and (1.18) the constant parameters are as follows: F represents
the fux of gas particles hitting a unit area of the surface from the atmosphere with
constant P, where P is the partial pressure of gas particles in the atmosphere, c is
the area of a single adsorption site, b is the probability for the desorption from a
site per time unit, and n
OG
is the probability of step 3 (1.16) occurring at a site per
time unit. The index O is for oxygen, while G is for the impurity gas. For general
understanding, it is acceptable to replace F with P.
The parameters S
O0
and S
G0
represent the sticking probabilities for the oxygen
and impurity gas on the clean surfaces, respectively. In general, the sticking prob-
ability is not a constant but should be dependent on various specifc factors such
as the surface structure, parameters of chemisorption sites, the surface band
bending, etc. According to the studies of co-adsorbed layers and multicomponent
surfaces of Carlsson and Madix (2001), Busse et al. (2001), Kreuzer (1990), and
Xu and Koel (1994), the sticking probability is infuenced by the surface adatom
modifers because the modifer precursor state can be created on the solid sur-
faces. In some studies, e.g., Zhdanov (2002) and Persson (1992), it was demon-
strated that the different gas species can occupy the individual adsorption sites
on the heterogeneous surfaces. In our presentation, simplifying the understand-
ing of chemisorption, it is supposed that the oxygen is adsorbed on the surface
sites with density N
maxO
, while the target gas is adsorbed on some precursor states
with constant density N
maxG
.
In the rate equations (1.17) and (1.18) the frst terms on the right-hand sides
of the equations describe the adsorption rate, while the second terms represent
the desorption rates. The third terms defne the rates of the LH bimolecular step
in which adsorbed species of oxygen and impurity gas are involved as defned in
Eq. (1.15).
The rate equations (1.16) and (1.17) have to be solved numerically if accurate
descriptions are desired for the fractional coverage Q
O
and Q
G
in the response
simulation. However, an explicit defnition of these parameters provides better
understanding about infuences of various factors on the sensor response. An
analytical solution can be obtained for the rate equations (1.17) and (1.18) if the
surface coverage Q
G
with the impurity gas is supposed small with respect to the
oxygen surface species (Q
G
<< Q
O
). Based on this assumption, the total oxygen
coverage Q
O
is supposed to be represented by almost constant initial coverage Q
O0

and a small correction dQ
O
produced by the surface chemical reaction between the
impurity gas and oxygen. Therefore, the solution for the rate equation (1.17) can
be written in the form
= -
O O0 O
(1.19)
14 ChemiCal sensors moDelinG anD simUlation: VolUme 2
In (1.19) the fractional coverage Q
O0
is defned by the equilibrium of the oxygen
adsorptiondesorption processes in clean air [step 1 (1.14)] and dQ
O
is a correction
to the coverage produced by the surface chemical reaction between the oxygen
and the impurity gas in the LH bimolecular step (1.16). It is accepted for all condi-
tions dQ
O
<< Q
O0
.
For nondissociative chemisorption and simple bimolecular reaction on
the sensor surfaces defned by (1.16), it is acceptable to suppose that dQ
O
= Q
G
.
Consequently, only a solution of (1.18) is required for a simulation of the sensor
response to gas defned in (1.13). For this, the coverage Q
O
in (1.18) has to be re-
placed by the following expression
( ) ( ) = -
O O0 G
t t (1.20)
After the substitution, Eq. (1.18) is written in the following form:
( )

= - + + +
2 G
OG G G G G0 G OG O0 G G G G0
d
F c S F c S
dt
(1.21)
The initial condition can accepted as Q
G
(t 0) = 0 for the injection of impurity
gas into the atmosphere. Depending on the parameters, the solution of Eq. (1.21)
can be obtained in the following two forms:
( ) ( ) ( )




= - - + <



1 2 1 2
2 2 2
1G G OG G OG G OG
OG
1
4 4 if 4
2 2
t
t tg (1.22)
( )
( )
( ) ( )

-
= >
-
2
2G G OG
1 exp
if 4
1 exp
t
t p
p q t
(1.23)
where
=
G G G G0
F c S (1.24)
= + +
G G OG O0
(1.25)

( ) ( )
-
-

= + -


1 2
1
2
G OG OG
1 1 4 2 p (1.26)

( ) ( )
-
-

= - -


1 2
1
2
G OG OG
1 1 4 2 q (1.27)

( )

-
= -
1 2
2
G OG
1
1 4 (1.28)
nUmeriCal simUlation of eleCtriCal responses to Gases 15
According to the defnition, fractional coverage of the surfaces with oxygen Q
O

represents a ratio between the density of chemisorbed oxygen N
O
and the density
of the chemisorption sites for oxygen N
Omax
(Q
O
= N
O
/N
Omax
). Replacing N
A
in (1.13)
with N
O
and using (1.20), one can obtain a simulation formula for the resistance
response of a metal oxide sensor to the injection of impurity gas as follows:
( ) ( )
{ }


= - =

2
0 O0 G
exp 1,2,...
i
R t R Z t i (1.29)
where

= =
2 2
Omax max
0
8
b
d
q N qV
Z
kTN kT
(1.30)
The formula (1.29) can be rewritten in more convenient form for the simula-
tion of sensor response. To do this, a defnition of R(0) = R(t 0) can be obtained
from (1.29) and, consequently, a normalized resistance response can be described
as follows:

( )
( )
( )
{ }

-


= - -




2
2 1
O0 O0 G
exp 1 1
0
R t
Z t
R
(1.31)
The analytical expression (1.31) is acceptable for quantitative calculation of
transient signal for the resistance response of a metal oxide sensor to a single im-
purity gas in air. Time dependencies of the response after exposure of the sensor to
gas are explicitly defned by the solutions (1.22) and (1.23). The magnitude of the
resistance response can be obtained for the sensors from the saturated transient
response signals evaluated under the assumption of extremely long exposure time
using the limit t in (1.31). The acceptability of the response simulation using
the analytical expression (1.31) was discussed by etkus (2002, 2004).
2.3. numErical simulations
Analytical solutions of gas sensor problems typically describe the dependencies
between the response to gas and the factors important in the sensor technology,
the measurement methodology, and the response mechanisms. A special form of
the solution is typically generated in such studies that explicitly reveals the rela-
tionships among various parameters. The analytical solutions actually represent
a theoretical simulation model for the sensor response and the related pheno-
mena in the sensors. This simulation model allows one to calculate the response
16 ChemiCal sensors moDelinG anD simUlation: VolUme 2
curves and ft these curves with experimental ones, which, consequently, gives a
quantitative description of the parameters in the model. Such an approach gives
quite detailed insight into the processes that determine the mechanisms and pa-
rameters of the sensor. Analogous results can also be obtained without defning
the analytical solutions. There are numerical methods acceptable for solving the
basic equations of the simulation model. However, this approach makes the ft-
ting to the experimental data rather complicated. Therefore, the approaches based
on analytical solutions are much frequently used for the simulation of sensor re-
sponses than those based on pure numerical methods.
Numerical representation of the analytical model makes it possible to use the
transient responses for defnition of a series highly important parameters. The dy-
namics of a change in electrical charge transport produced by exposure of a sen-
sor to gas contains specifc information about both the physical and the chemical
processes in the bulk and on the surfaces of semiconducting metal oxides. At
least three independent methods have been proposed for obtaining the transients
of the electrical signals in metal oxides exposed to gas. Techniques based on the
controlled change in the sensor temperature, the bias voltage, and the composi-
tion of gas atmosphere have been used. An interesting simulation model is pro-
posed by Varpula et al. (2011) for the transients generated by small step changes
in either temperature, bias voltage, or concentration of reducing gas. Varpula et
al. (2011) showed that model-based analysis of experimental data allows calcula-
tion of several important quantities, such as the time constant of the electronic
trapping process, the height of the grain boundary (GB) potential barrier, the
relative change of the occupied GB states, the resistance coeffcient, and the ef-
fective number of GBs between the electrodes of the sensor. The parameter values
obtained can be used for further study of the underlying physical and chemi-
cal phenomena in gas-sensitive metal oxides, including determining the electron
transport over the grain boundaries.
Thermally activated processes of surface chemical reactions were analyzed
by Ahlers et al. (2005) aiming to simulate the bell-shaped variation of the sensor
response to gas with the sensor operation temperature T. In this study, it was
supposed that, at typical sensor operation temperatures above 425 K, the prevail-
ing adsorbate is the oxygen species O

. Adsorption and desorption of oxygen is


defned by the transition

- -
+
2( ) ( )
O 2 2O
g S
e (1.32)
An infuence of temperature appears in this process because the chemisorp-
tion requires two electrons to be thermally emitted from the Fermi energy E
F
across
the surface barrier (E
C
+ qV
s
E
F_bulk
) to become trapped at a pair of adsorbed oxy-
gen atoms O
(S)

. E
C
is the energy of the conduction band. Desorption of the sur-
face oxygen ions requires re-emission of two electrons across an energy barrier of
height (E
C
E
O_minus
) with respect to the energy level E
O_minus
of chemisorbed oxygen
nUmeriCal simUlation of eleCtriCal responses to Gases 17
species O
(S)

. If reducing analyte gas molecules are present in the atmosphere, the


reactions between the gas and surface oxygen take place. These reactions reduce
the density of surface oxygen ions, N
O
. It is supposed that for this reaction a kine-
tic barrier, represented by activation energy E
a
, needs to be overcome. The reac-
tion products then leave the surface and the electrons formerly trapped at surface
oxygen ions are released to the conduction band.
Based on these ideas, the simulation model of Ahlers et al. (2005) includes
the solution (1.12) of Poissons equation (1.3) for the intergranular junction with
the double Schottky barrier and the solution of the rate equation for the surface
chemical reaction (1.6) that was specifcally modifed by including the parameters
of thermally activated processes as follows:
( ) ( )


+ - -

= - - - -



_ bulk O_ minus
2 2 O
0 O2 0 O
2 2
exp exp
C S F C
f C r
E qV E E E
dN
p N N G
dt kT kT
(1.33)
where p
O2
is the oxygen partial pressure,
f0
and
r0
are the kinetic parameters for
adsorption and desorption, and G is the term describing the reaction between the
gas and chemisorbed oxygen as follows:



= -




O 0
exp
a
E
G N k
kT
(1.34)
The term G includes a Langmuir relative coverage Q defned by
=
+
gas
gas 0
p
p p
(1.35)
In (1.35), p
gas
is the partial pressure of reducing gas and p
0
is



= -




ads
0
exp
Q
E kT
p
V kT
(1.36)

( )
-
=
3 2
*
gas 0 gas Q
V h M M kT (1.37)
Here T
gas
is the gas temperature (e.g., 300 K), T is the sensor operation tempera-
ture, E
ads
is the binding energy of the adsorbate, M
0
is the atomic mass unit (1.67
10
27
kg), and M
gas
is the relative atomic mass of the reducing gas (e.g., 28 for
CO and 2 for H
2
). Considering the descriptions in (1.34)(1.37) and simplifcations
analogous to that in Section 1.2, the solution for N
O
is obtained from (1.33) by
Ahlers et al. (2005).
18 ChemiCal sensors moDelinG anD simUlation: VolUme 2
The fnal formula acceptable for the simulation of the response versus T
was obtained by defning the relationship between N
O
and the width of the space
charge layer W
SCR
in the double Schottky barrier and the resistance of the sensor.
In the work of Ahlers et al. (2005), the analytical description of the simulation
model was presented in the following form:

( )
-
= - = -
-
SCR_gas
air
gas
gas SCR_air
, 1 1
S
S
D W
R
S T c
R D W
(1.38)
Here R
air
and R
gas
are the sensor resistances in clean air and air with gas, respec-
tively, W
SCR_air
is the space-charge region width in clean air, W
SCR_gas
is the space-
charge region width in the presence of reducing gas, and D
S
is the metal oxide
layer thickness. The function S(T, c
gas
) in (1.38) defnes the relationship between
the normalized resistance response to gas and two parameters, temperature T and
amount of tested gas c
gas
, and was used for the calculation of theoretical depen-
dencies. This simulation model of Ahlers et al. (2005) reproduces the particular
form of the relationship between the gas response and sensor temperature, with
the sensitivity maximum S
M
occurring at the temperature T
M
. The solution also
demonstrated the sublinear variation of S with the analyte gas concentration c
gas
.
Details of the comparison between the simulation model and the experiment can
be found in the work of Ahlers et al. (2005).
In studies with strictly defned conditions, the simulation models can be ex-
tremely simplifed using empirical descriptions of some parameters and limiting
the analysis to specifc tasks, as proposed by Vilanova et al. (1998). In this work,
the infuence of electrode position and geometry on the semiconductor gas sen-
sor response was studied. It was assumed by Vilanova et al. (1998) that the dif-
fusion of gas into a thick-layer sensor is the only process limiting the kinetics of
the sensor response. Adsorption and chemical kinetics are assumed to be faster
than physical diffusion of gas. At equilibrium, the relation between free (C) and
adsorbed gas concentration (N) can be defned by the frequently used formula

= N BC (1.39)
where B is a constant. Assuming one-dimensional diffusion, the rate equation for
the surface chemical reaction can be written as


- + + =

2
2
0
a
C C N
D KN
t t y
(1.40)
Here K is the reaction rate constant and a is the reaction order, assumed to be a = 1
by Vilanova et al. (1998). The frst two terms in (1.40) represent the diffusion of
nUmeriCal simUlation of eleCtriCal responses to Gases 19
gas into the thick sensor, while the terms with N defne the variation of the density
of the gas on the sensor surfaces. The problem can be solved by computational
methods, and the carrier concentrations can be derived assuming proportionality
between absorbed gas concentration and conduction electron concentration.
The numerical simulation model of Vilanova et al. (1998) demonstrated a cor-
relation between the contacts and the sensor response. It was shown by this
simulation that when the sensor has low catalytic activity for a gas, the variables
electrode position, active-layer thickness (y
O
), and electrode gap (W) have no rele-
vance. For higher catalytic activity, when electrodes are placed on the bottom, the
sensor sensitivity is lower, although it can be increased slightly using wider elec-
trodes. For a sensor with a very wide electrode gap, the relative response to gas is
obtained as being almost similar for all electrode positions in the sensor.
As follows from the above paragraph, the solutions of equations can be
obtained in numerical form using special computational methods without the
need to accept at least some of simplifcations required for an analytical solution.
Numerical methods frequently seem less restricting for sensor analysis than spe-
cifc requirements for the conditions of tests; however, these methods do not re-
sult in explicit relationships between the response and probed factors. Numerical
simulation of gas sensor characteristics is typically based on similar models that
are used for obtaining analytical description of sensor parameters, but the results
of numerical simulations are frequently limited to a quantitative representation of
a relationship between a selected factor and the sensor response.
Rothschild and Komem (2003), proposed a computational method for numeri-
cal calculations of the amount of chemisorbed species and the electrostatic po-
tential barrier, which are crucial for simulation of the sensor response. In the
proposed method, Wolkensteins suggestions about the relationship between the
gas chemisorption on semiconductors and the flling of the surface electronic lev-
els were used (Wolkenstein 1963). This proposition actually complicates the rate
equations because the chemisorbed species are split into two groups: neutral che-
misorbed species with empty electronic levels and ionized chemisorbed species
with occupied electronic levels. Excluding the detailed analysis of the occupation
mechanism, a simplifed description of the occupation probability of the chemi-
sorption-induced surface electronic states is proposed in this approach. The de-
scription of the occupation is actually based on a trivial use of the Fermi-Dirac
distribution function as follows:

-
-
-

-


= + -



+

1
0
1 exp
F SS
E E N
kT N N
(1.41)
Here N
0
is the number of neutral adsorbates per unit surface area, N

is the
number of chemisorbed adions per unit surface area, and E
SS
is the energy of
20 ChemiCal sensors moDelinG anD simUlation: VolUme 2
chemisorption-induced electronic levels on the surfaces. Defning the coverage of
the surfaces with the chemisorbed particles by

-
+
=
0
partial
*
N N
p
N
(1.42)
where N
*
is the number of adsorption sites per unit surface area, one can use
the rate equation (1.6) for the defnition of the equilibrium problem with explicitly
included factors defning the sensor properties and the surroundings. Thus, as-
suming that the desorption of the neutral chemisorbed species is determined by
specifc desorption energy x
0
and omitting the bimolecular interaction, it is pos-
sible to obtain the equation for the equilibrium coverage of the surfaces produced
by nondissociative chemisorption as in Rothschild and Komem (2003):
( )

- -

+ -

- = - + -




0 0
* 0 0 *
0
1 exp exp
2
b
C SS
E E p
S N N
kT kT
mkT
(1.43)
Here S
0
is the adsorption probability, n
0
and n

are the desorption probabilities


for neutral and ionized chemisorbed species, respectively, Q
0
= N
0
/N
*
and Q

=
N

/N
*
, and E
C
b
is the conduction-band edge in the bulk of the sensor. Since the
capturing of the electrons by the chemisorption-induced surface levels depend on
the surface potential V
S
, Poissons equation has to be solved before calculating the
surface coverage Q with the chemisorbed species.
Rothschild and Komem (2003) proposed a method for calculation of the
equili brium coverage of chemisorbed species and the height of the chemisorption-
induced potential barrier. This method actually enables numerical simulations of
the sensor response to gases and quantitative evaluation of the effects of various
parameters such as the operating temperature or the doping level on the sensor
response. This method was tested by Rothschild and Komem (2003) by simulating
n-type SnO
2
-based sensors.
Wolkensteins proposition (Wolkenstein 1963, 1991) about distinguishing be-
tween weak (neutral chemisorbed species) and strong (ionized chemisorbed spe-
cies) gas chemisorption states for the interactions between the surface and gas
was successfully applied in a simulation model of the dynamic response of sen-
sors to gas by Guerin et al. (2008). Following Wolkensteins approach, the chemi-
sorption was split into two steps by Guerin et al. (2008). During the frst step,
the bond between the adsorbate and the substrate is weak and does not involve
electronic transfer from the bulk to the surface or vice versa. The electrons of the
atom or the molecule remain located in the vicinity of the adsorbate, involving a
simple deformation of the orbitals. The binding energy of the adsorbate is E
ads
and
corresponds to the loss of free energy of the system during the adsorption process.
nUmeriCal simUlation of eleCtriCal responses to Gases 21
This neutral chemisorption does not change the electrical properties of the mate-
rial, but the perturbation created by the adsorbate induces a surface state E
SS
in
the bandgap. This surface state acts as a trap for the electrons.
The second step (strong chemisorption) occurs when an electron of the con-
duction band, whose energy is E
C
, is transferred from the semiconductor to
the adsorbed species. The binding energy of the adsorbate is increased by E
S

= E
C
E
SS
, that is, the loss of free energy of the system during the ionization
process. This process involves the creation of a negative superfcial charge and
a chemisorption-induced surface potential barrier V
S
(V
S
< 0) that is defned by
Poissons equation.
Considering these two steps, the evolution of the surface coverage with chem-
isorbed gas Q is defned by the rate equation corresponding to that in (1.43),
where we assumed dQ/dt = 0 at equilibrium. The rate equation includes both the
weak and strong chemisorption states characterized by the ionized Q

and neutral
Q
0
coverage of the surface. These two parameters are related to the total coverage
Q by the Fermi-Dirac statistics.
The simulation of sensor response requires simultaneous numerical solution
of both the rate equation and Poissons equation. Moreover, some assumption
must be added to obtain the relationship between the solution and the resist-
ance response of the sensor to gas. For this, it is assumed that the mechanisms
of adsorptiondesorption at the grain surface are much slower than the genera-
tion recombination effects in the semiconductor. For the free charge carriers it
must be acceptable that n/t = p/t = 0. The continuity equation (1.5) and the
transport equation (1.1) also have to be solved for the defnite structure of sen-
sor. Therefore, the grains in the polycrystalline sensor are assumed to be quasi-
spherical, identical in size, and single-crystal. They are joined, coupled by a small
contact surface that allows a high porosity. This simplifed model allows one to
easily determine the electrical features of a grain taking into account the mecha-
nisms of the electrical conduction.
The numerical simulation of sensor response needs exact quantitative defni-
tions of all the parameters included in the model. This requirement makes numer-
ical simulation more diffcult than the analytical approach. This is because only
some estimated magnitudes are known for the required parameters. However,
some parameters, such as the energy parameters, are practically unknown but
are particularly critical for the simulation because a small change of these val-
ues can produce large variations in the fnal results. Guerin et al. (2008) made a
particular attempt to quantitatively defne the parameters required for the simu-
lation of WO
3
sensors. Tabulated magnitudes of parameters can be found in this
work. The response of gas sensors to ozone in dry air was calculated based on
the adsorption/desorption mechanisms of the species O
2
, O
2

, O, and O

at the
surface of the grains. The response kinetics based on the Wolkenstein adsorption
theory were simulated numerically for comparatively low temperatures (less than
22 ChemiCal sensors moDelinG anD simUlation: VolUme 2
525 K) and nondissociative adsorption. The model was then extended to dissocia-
tive adsorption at temperatures higher than 525 K.
It is evident from the reported studies that the numerical simulation of sen-
sor response to gas based on any model, including the analytical models and the
computational methods, aims at quantitative evaluation and fundamental under-
standing of sensor functioning. Accurate description of the response becomes
increasingly important for the comprehensive analysis of the complete physical-
chemical picture of the conversion of the interaction between the surfaces and
gases into the detectable signal and, consequently, for the intentional optimiza-
tion of sensor characteristics and measurement methodologies. The numerical
implementation of the model description allows one to compare theory and experi-
ment by ftting the calculations according to the mathematical models with the
measured quantities.
2.4. VErification of modEls
The barrier-limited conduction model accompanied by chemical rate equations
has been very successful in simulating gas-sensing properties in metal oxide
construction. Quantitative comparison between the outputs of simulation mod-
els and experimental results allows one to evaluate both the acceptability of the
model formalism and the correctness of understanding about the core factors in
the analysis. Successful simulation typically results in close match between the
calculated and measured parameters under a comparatively wide spectrum of test
conditions. It is also possible to predict some specifc features from the theoretical
calculations using a simulation model of sensor response.
The analytical description of the response (1.30) with the solution (1.21) pre-
dicts instability of the response of a metal oxide to gas under specifc conditions
defned by relations between the parameters in (1.21). It follows from (1.21) that
a periodic variation of resistance should be detected for the metal oxide sensors
in response to gas. The conditions under which the response oscillations may
occur are dependent on a special combination of parameters, namely, the fux
of particles F, the effective cross section of the chemisorption site c, the sticking
probability S
0
, the desorption probability b, the rate of bimolecular interaction on
the surfaces n
OG
, and the surface coverage with oxygen Q
O0
. In addition, included
in the simulation model is a series of measurable parameters that actually defne
the parameters explicitly (1.21). For example, the fux F is defned by the partial
pressure of the corresponding gas p
gas
in air, the surrounding temperature T, and
the mass of the gas molecule m
G
. The other parameters from (1.21) to (1.27) can
also be defned by special relationships, including the measurable characteristics
of sensor properties. Consequently, the theoretical calculations of the dependen-
cies between the response and various factors in the response simulations are
nUmeriCal simUlation of eleCtriCal responses to Gases 23
extremely complicated because of lack of reliable information about the quantita-
tive defnitions of required parameters. Nevertheless, the simulation model allows
one to reveal the key processes producing the oscillations of the sensor response
to gas. Instability in the response can be obtained for special proportions between
parameters defning the three steps (1.13)(1.15) in the surfacegas interaction.
In general, the simulation model prediction is verifed by a series of experimen-
tal results. Galdikas et al. (2000), for instance, demonstrated experimentally that
for SnO
2-x
-based sensors originally modifed with metallic catalysts, oscillations
can occur in the resistance response of these specifc sensors to gas under specifc
conditions. It was found that, if an odor with a fxed amount of formaldehyde was
tested, the oscillations of the resistance were typically detected by these sensors.
The effect was verifed by comparing the response to the odors of three analogous
solutions, glucoseglycine, glucoseglycineformaldehyde, and glucose glycine
acetaldehyde. There were no other types of sensors with oscillations of the re-
sponse, but the oscillations were reproduced in other similar sensors under the
same conditions. It must be noted here that the oscillations of the response were
detected only within a strictly limited interval of sensor temperatures.
The study of Galdikas et al. (2000) is not the only one that has reported
about the oscillations of the resistance response. The oscillations of the sensor
response are frequently detected for CO gas. It was reported by Nitta et al. (1978)
and Nitta and Haradome (1979) that, in a ThO
2
-doped SnO
2
sensor, a new self-
oscillation phenomenon was found only when the sensor was exposed to CO gas.
Based on experimental results, this phenomenon was related to the environ-
mental CO gas concentration, substrate temperature, and applied voltage. It was
demonstrated that the oscillation is extremely sensitive to the concentration of
CO gas and can be detected especially in the region of 0.20.3%. The effect was
also detected for a SnO
2
gas sensor in a study by Nakata et al. (1996). The simul-
taneous measurement of the temperature on the semiconductor surface and the
conductance suggested that the temperature change was a key variable in the
oscillatory phenomenon. As a preliminary theoretical model, a simulation was
performed using the surface concentration of CO and the temperature as two
independent variables.
In addition, an oscillation has also been detected in the response to H
2
gas in
air (for details, see Kanefusa et al. 1981). Sensors based on sintered SnO
2
with
mixture of additives Pd, Rh, and MgO were used in these tests. It was experimen-
tally demonstrated that this phenomenon depends on the H
2
gas concentration,
the working device temperature, and the voltage applied to the device.
The simulation model (1.30) with the solution (1.22) results in the transient
response defned by monotonically changing signal that approaches constant
magnitude asymptotically as t . This result has been commonly verifed
by numerous experiments for various sensors tested under various conditions.
Actually, it proves that the conditions for the parameters in (1.22) are typically
24 ChemiCal sensors moDelinG anD simUlation: VolUme 2
acceptable for most practical metal oxide sensors. In these cases, the formula
(1.30) with defnition (1.22) seems to be less advantageous compared to other
simulation models proposed for comparatively simple forms of the response tran-
sient in various studies. In general, any relationship containing an exponent can
be acceptable for simulation of the sensor response under typical conditions.
Compared to this, the most outstanding advantage of the simulation model based
on (1.30) with (1.22) is in adequate description of several specifc aspects of the
response kinetics in metal oxide sensors.
The time constant t defned in (1.27) and included in (1.22) actually de-
termines the response time for the metal oxide sensors (see, e.g., etkus 2002).
Considering the commonly known relationship between the response time and
the amount of gas, the most unusual conclusion follows from the simulation
model (1.27) for the time constant in the interval of comparatively low amount of
the impurity gas. The simulation based on (1.27) leads to a singular minimum in
the dependency of the time constant t on the partial gas pressure p
gas
. This result
can be obtained for comparatively low desorption rate b
G
and the high rate for
the LH step n
OG
. The calculated dependencies were proved experimentally for a
SnO
2
-based sensor exposed to H
2
gas by etkus (2002). etkus et al. (2004, 2005)
also demonstrated that the proportions between the rates can be changed by sur-
face modifcations with catalytic metals. These modifcations results in signifcant
change of the dependencies of the time constant t versus the partial pressure of
the impurity gas p
gas
. The rates of the processes, namely chemisorption, desorp-
tion, and bimolecular interaction, can be increased or decreased depending on
the metal and the amount of it if the impurity only partially covers the surfaces of
the basic oxide flm. The comparison between the simulation model and experi-
ment was performed for the SnO
2
-based sensors exposed to pure H
2
and CO gases
in air.
Asymmetry in the rise and the fall of the resistance response to gas also is
given by the simulation model (1.30) with (1.22). The response time of the sensor
is determined by the three steps of the gassurface interaction during injection
of impurity gas into the atmosphere. The time constant t (1.27) depends on the
adsorption fux F
G
, desorption rate b
G
, and the high rate for the LH step n
OG
. After
restoration of clean air conditions, the time constant that defnes the time of the
response fall depends only on the desorption rate b
G
and the high rate for the LH
step n
OG
. Consequently, the transient signals are defned by individual time con-
stants for the rise and fall of the response of the sensor to gas.
In general, a good match between simulation of the response and experimen-
tal results can be obtained only if similar conditions are applied to the simulation
and the experimental measurements. Consequently, it is acceptable to ft the cal-
culations of the response defned by (1.30) and (1.22) only if a steplike change in
gas composition is accomplished during the experiments. However, sometimes it
is possible to modify the simulation model by simple substitution of parameters,
nUmeriCal simUlation of eleCtriCal responses to Gases 25
aiming to evaluate the infuence of specifc conditions. Based on the formalism
in Section 2.2.2, the resistance response of metal oxide sensors can be obtained
for varying amounts of impurity gas, as demonstrated in etkus (2004). In this
study, the transient signals of the response were obtained for periodically varying
amounts of impurity gas in air. In general, various special functions can be used
for the simulation of the response of sensor to varying amounts of impurity gas by
the formalism based on (1.30) and (1.22).
Discrepancies between the theoretical simulation of the sensor response and
the measured results typically occur due to mismatch between the conditions of
theory and those of experiment, and due to lack of specifc aspects in the model.
Though the origin of the discrepancies is not always clear, the situation can fre-
quently be improved by adding some details in the theoretical description of the
response that solve the problem of the mismatch.
An attempt to improve the commonly known approach of thermally stimulated
electron transport in metal oxide sensors with intergranular junctions was made
by introducing the tunneling effect. Lee (2006) assumed that if a gas sensor is
highly doped, up to degeneracy, then the Fermi level lies above the bottom of the
conduction band. Because of heavy doping, the depletion region is very thin and
at low temperatures free electrons with energy close to the Fermi level can tunnel
across the double Schottki barrier between the grains. Though the proposition
was not verifed by reliable results, it was suggested that the tunneling effect can
be important at comparatively low temperatures. At practically important temper-
atures, however, thermoemission transport over the barrier should be dominant.
A much more successful attempt to introduce the tunneling effect in the simu-
lation of sensor resistance response was presented by Malagu et al. (2009). In
this study, it was suggested that after oxygen adsorption, the barrier height and
the depletion width become larger and, as a consequence, the sample resistance
increases. Different factors (such as type of defects, morphology, and additives)
contribute to the electrical response of the gas sensor. The barrier height V
S
of the
intergranular junction was related to the characteristics of the depletion region by
the following solution of Poissons equation:

-


= -


2
2 1
2
1
2 3
d
S
q N
qV R (1.44)
Here R is the radius of the particle, L is the depletion-layer width, and N
d
is the
donor density. After thermal treatment of metal oxide sensors, the donor concen-
tration N
d
defned by the oxygen vacancies in the bulk typically decreases, which
can lead to an enlargement of the depletion layer (increase in L). These effects pro-
duce an increase of the grain boundary resistance. In all the temperature range
for such grain boundary conditions, the tunneling current can be more than an
26 ChemiCal sensors moDelinG anD simUlation: VolUme 2
order of magnitude larger than the thermionic current. Consequently, the electric
current in the metal oxide sensor must be calculated according to the two-term
defnition as follows:
( ) ( )


= + -

2
0
exp
S
V
S
V AT
J f E P E dE AT
k kT
(1.45)
In the defnition of current, the frst term corresponds to the tunneling cur-
rent and the second to the thermionic current. The parameters A and k are the
Richardson and Boltzmann constants, and f(E) is the Fermi-Dirac distribution.
P(E) is the transmission probability for a reverse-biased Schottky barrier, and it
is defned by a special function by Malagu et al. (2009). The contribution of the
tunneling current was verifed by experimental results of impedance spectroscopy
in the study of Malagu et al. (2009).
In metal oxide sensors the electronic processes can also be accompanied by
ionic driftdiffusion processes caused by the driftdiffusion of the particles ad-
sorbed on the surface, as suggested by Liess (2002) and Simakov et al. (2006).
Simulation of the electrical current in the sensor in this approach includes varia-
tion of the surface ions with the position between the electrical contacts of the
sensor. Simakov et al. (2006) proposed modifying the rate equation for the che-
misorbed particles [similar to (1.16) and (1.17)] by including the ion migration on
the surfaces. Simplifying the analysis by the limit of the equilibrium state, the rate
equation was proposed in the following form:
( ) ( )

-


- - - - =



+ +

1
0
A a
d A a i
a S A
a a
N n
N n dI
N N qw
n n n n dx
(1.46)
Here
a
is the gas particle adsorption effective frequency,
d
= 1/
l
is the desorption
effective frequency, and
r
is the effective frequency of the reaction between the
oxidizing and reducing gas at the surface. N
A
is the surface density of the particles
absorbed from the gas, and N
S
is the maximum (saturated) density at the anode of
the sensor [N
S
= N
A
(x = L), with L being the length of the sensor]. The ion current
is defned by the relationship

-
=
i A i
I qN Ew (1.47)
with mobility of ions m
i
, external electric feld E, elementary electrical charge q,
contact width w, density of the chemisorbed gas ions N
A

= N
a
n/(n + n
a
). n is the
concentration of the free bulk electrons and n
a
is the electron concentration cor-
responding to the Fermi level, equal to the energy of the surface acceptor level E
a
.
A solution of (1.46) was obtained for the distribution of electrical charge between
nUmeriCal simUlation of eleCtriCal responses to Gases 27
the contacts of a sensor by Simakov et al. (2006). Unfortunately, the simulation
of the electrical current in the sensor was not supported by adequate experimen-
tal results. The measured dependence of current versus applied voltage (the I V
characteristic) does not directly represent the distribution of the surface electrical
charge (or chemisorbed gas ions) in the sensor. Therefore, an improvement of the
simulation model of the electrical response of sensor to gas by migration of che-
misorbed ions on the surfaces of sensors still needs more detailed studies about
the simulation model and much more reliable verifcation of this model performed
under adequate experimental conditions.
Varying density of chemisorbed gas particles and possible migration of these
particles on the surfaces suggested an original modifcation of the simulation
model for resistive gas sensors. Varpula et al. (2008) assumed that oxygen ion-
ization on the surfaces of sensors due to chemical reactions is equivalent to the
phenomenon of electron trapping in the surface states. Based on this, it was ac-
cepted that free electrons in the gas sensors can be trapped on the surfaces of the
grains or grain boundaries in two kinds of surface states: intrinsic and extrinsic
surface states. The intrinsic states are composed of the states which are created
by the existence of the surface itself and the states which are created by surface
impurities, doping, and surface defects. The extrinsic surface states are created
by adsorbed gas atoms or molecules at the surface. If the electron is trapped in
an adsorbed atom or molecule (extrinsic state), creating an ion, this process is
called ionization. In other words, electron trapping and oxygen ionization is the
same phenomenon. The simulation model of sensor conductance of Varpula et
al. (2008) suggests an infuence of external electric feld on trapping of electrons
into the extrinsic states due to a change in the grain boundary characteristics
produced by the applied electric feld. Poissons equation is solved for the sensor
without and with an applied electric feld, assuming that the complete length of
the depletion region (x
d1
+ x
d2
) at the contact between adjacent grains with the
double Schottky barrier is constant and independent of the feld. By solving the
continuity equation (1.1), the current density through the potential barrier is ob-
tained in the following form:



= -




max 0 bar
0
exp sinh
2
B
d
qV qU
J q N E
kT kT
(1.48)
where N
d
is the donor concentration, E
0
max
is the maximum applied electric feld
producing the potential shift equal to the barrier height, and U
bar
is the applied
voltage. An initial grain boundary potential V
B0
is defned by the following solution
of Poissons equation:

=
2
0
8
B
B
d
qN
V
N
(1.49)
28 ChemiCal sensors moDelinG anD simUlation: VolUme 2
where N
B
is the charge density on the extrinsic states on the grain boundary.
Based on the same equations, the density of trapped electrons on the grain
boundary is defned by Varpula et al. (2008) by



= -




0 bar
exp cosh
2
B
B d
qV qU
n N
kT kT
(1.50)

The simulation model proposed by Varpula et al. (2008) suggests two regions
in the dependence of the response versus applied voltage: a linear regime at low
voltages and a superlinear regime at high voltages. Consequently, the extension
of the conductive model used for simulation of barrier-limited electron transport
in gas sensors allows one to combine the conduction and surface-state models
and to obtain a description of nonlinear I V characteristics in which the linear,
sublinear, and superlinear current regimes are detected. The model simulation
was applied and verifed by response tests on WO
3
-based sensors in the study
of Varpula et al. (2008). A good ft between experimental data and the proposed
model was obtained. According to the model, sensitivity degradation occurs only in
the regime where traps are almost already flled in thermodynamic equilibrium in
clean air. In the opposite regime, where the traps are emptier, the model suggests
constant or slightly increasing sensitivity with the bias voltage. The generality of
the proposed model allows investigation of the fundamental physical-chemical
phenomenon, the trapping process, in all resistive metal oxide sensors.
3. resIstIve sensors
3.1. introductory rEmarks
Resistive gas sensors are still the most-exploited type of sensors for both research
projects and practical applications. Simplicity of the resistance measurements is
the most signifcant advantage of these sensors over the other ones, especially in
commercial devices. Moreover, it has been demonstrated by various studies that
comparatively simple models of electric charge transport can adequately explain
the most essential features of sensor functioning and the most promising develop-
ments in sensor technology. Since these models can typically be described by
standard fundamental equations, it would seem that simulation of the response
would be quite an easy task. However, there are at least two aspects in the simula-
tion models that must be carefully considered.
First, resistive gas sensors are typically based on nonhomogeneous materials.
In general, such sensors can be described as some complicated network of semi-
conducting elements each of which can be individually analyzed in terms of the
basic equations describing the electronic properties. Therefore, the effect of the
nUmeriCal simUlation of eleCtriCal responses to Gases 29
complicated structure of the material is typically excluded from sensor simula-
tion models by limiting the analysis to the single most essential part of sensor
assuming that this selected part completely determines the functioning of the
sensor. This limitation must be suffciently justifed for the sensor simulation
model to be accepted as reliable.
Second, simulation models of the resistance response are frequently verifed
by comparing the calculated data with experimental results obtained for sensors
with two electrodes. These can be highly affected by the electrode resistances caus-
ing a signifcant mismatch between the calculated and measured responses, be-
cause the simulation models typically do not include the effects of the electrodes.
The most systematic studies are currently required for development of re-
sponse simulation models for new resistive gas sensors based on nanostructured
and organic materials. Attempts have been made to adapt the classical models of
electrical properties to describe the processes in constructions with nanometer
dimensions. However, simulation of the resistance response in nanosensors and
molecular sensors is still challenging for sensor researchers, and highly innova-
tive approaches are probably necessary. In this section, an overview of existing
approaches to response simulation is presented for several of the most popular
types of resistive sensors.
3.2. polycrystallinE films
To date, gas sensors based on metal oxide flms with polycrystalline structure have
been the most popular type of sensors to be investigated in various studies and
applied in practical devices. Extensive research and development of these sensors
has been highly stimulated by the commercial successes of Figaro, Inc. (Japan)
since 1969. Basic understanding has been deepened and simulation models of
sensor responses have been extended to explain various aspects of the response
mechanisms in metal oxide gas sensors, as has already been presented in Section
2.2 of this chapter. It is commonly recognized that the structure of the sensor is
important in the development of simulation models of the response. In addition,
it is already proved that some processes determining the response can be related
only to the polycrystalline structure of sensors. One such process is gas diffusion
in and out of the pores in polycrystalline materials. A model for simulation of the
response, including the infuence of gas diffusion, has been developed and thor-
oughly analyzed in a series of investigations (Sakai et al. 2001; Matsunaga et al.
2002, 2003).
The basis of the model rests on the understanding that, at equilibrium, the
gas concentration inside the porous sensing layer will decrease with increasing
depth due to diffusion-limited access. As a result, the gas concentration profle
inside the flm must depend on the rates of gas diffusion and the surface reaction.
30 ChemiCal sensors moDelinG anD simUlation: VolUme 2
In fact, the diffusion and the surface reaction are competing processes, the frst of
which results in an increase in the quantity of gas amount inside the flm, while
the second consumes the gas and, in general, limits the diffusion into the flm.
Assuming that the diffusion of gas is effective only into the depth of the flm,
the process can be accepted as a one-dimensional problem with the main axis
directed toward the flm depth. Then, the access of gas to the surfaces of grains in
a polycrystalline sensitive flm can be characterized by the rate equation

( ) ( )
( )

= -

2
2
, ,
,
C x t C x t
D kC x t
t t
(1.51)
Here C(x, t) is the target gas concentration, D is the diffusion coeffcient, k is the
reaction constant, x is the depth from the surface, and t is time. The frst term in
the right side of Eq. (1.51) represents diffusion of gas, while the second defnes the
surface chemical reaction.
The partial differential equation (1.51) can be solved only if the boundary
and initial conditions are defned. For this, several details about the model must
be known. In the studies of Sakai et al. (2001), Matsunaga et al. (2002), and
Matsunaga et al. (2003), it was supposed that the sensor is like a plate with thick-
ness 2L and the target gas can access both surfaces of the plate. Before the test,
there are no gases inside the sensor and, consequently, C(x, 0) = 0 is the initial
condition for this problem. After injection of the target gas into the atmosphere,
the amount of gas is C
S
at both surfaces of the plate representing the sensor. The
target gas is supplied continually at the same rate, so the amount C
S
is constant
in this analysis. Based on this assumption, the boundary conditions are defned
as C(0, t ) = C(2L, t ) = C
S
. Using these initial and boundary conditions it is pos-
sible to obtain the solution for Eq. (1.51). Details about how to obtain the solution
are precisely described by, e.g., Matsunaga et al. (2003). Here we present only the
fnal solution, which describes the changes in the target gas concentration profle
in a platelike polycrystalline sensor after exposure to a constant amount of gas C
S
.
The gas concentration profle is defned by the following solution of Eq. (1.51):
( ) ( )




= - - +




2
, 1 exp sin
2
S n
n x
C x t C A t B
L
(1.52)

( )

=
- -
=

1
1 1
2
n
n
A
n
(1.53)

( ) ( )
( )



- + -


=

+ - -


2 2
2 2
1 1 exp
1 exp
n n
n n
k t t
B
k t
(1.54)
nUmeriCal simUlation of eleCtriCal responses to Gases 31

=
1 2
2
n
n
D
L
(1.55)
The model developed in the series of studies by Sakai et al. (2001), Matsunaga
et al. (2002), and Matsunaga et al. (2003) gives sophisticated variation of the depth
profle of the amount of gas in the porous flm, as shown by Eqs. (1.52)(1.55). The
profle of the amount of gas C(x, t) is represented by a monotonic function with the
flm depth at some periods of time, but there are time intervals when the profle is
defned by a unique shape. For the defnite time interval, the profle of the amount
of gas has a maximum at some x in the depth of the flm. It was proposed that the
maximum appears as a result of contribution by the surface reaction.
Analogous to (1.52), a profle C
*
(x, t ) can be obtained for the sensor after
the injection of gas into the atmosphere is stopped. The details can be found in
the cited works by Sakai et al. (2001), Matsunaga et al. (2002), and Matsunaga
et al. (2003).
Aiming to simulate the sensor response that is infuenced by gas diffusion,
one can use a commonly known empirical formula for the sensor electrical con-
ductance s, which can be written in the following form:
= +
0
m
C (1.56)
Here s
0
is the clean-air conductance; g and m are the variable constants. The
amount of gas C should be replaced by (1.52)(1.55). It must be noted here that
it is complicated to use the model developed by Sakai et al. (2001), Matsunaga
et al. (2002), and Matsunaga et al. (2003) for the simulation of sensor response
because the defnition (1.56) is acceptable only for a thin piece of sensor plate
defned by dx. To obtain the sensor resistance, integration of parallel slices of the
sensor dx over the thickness of the sensor is required. Quantitative evaluation of
gas diffusion on the sensor response was illustrated by the numerically simulated
detection of sensor signal by Boeker et al. (2002). Supposing a series of repeating
short pulses for the injection of target gas, Boeker et al. (2002) demonstrated that
the sensor signal shows a pseudo-drift over long time compared to the short pe-
riod of a single pulse of injected gas. The simulation proved that the drift resulted
from the analytes remaining within the flm. The period of drift was signifcantly
shorter for a comparatively large amount of the target gas than it was for lower
concentrations, when only very slow drift of the response signal was detected in
the numerical experiment by Boeker et al. (2002).
The structure of polycrystalline gas sensors can be important not only for
gas access to the surfaces of the sensor, but also to the electron transport that
determines the electrical response of the sensor to the gas. Numerous funda-
mental studies have proposed simulation models for the electrical properties in
non homogeneous semiconductors that are based on the theory of solid-state
32 ChemiCal sensors moDelinG anD simUlation: VolUme 2
electronics. However, models for the simulation of metal oxide gas sensors typi-
cally include only some specifc aspects of the electrical properties of polycrystal-
line flms. An infuence of proportions between the length of the depletion region
and the grain size is the most frequently discussed problem for the sensor re-
sponse models. An attempt to include this effect in the simulation of sensor re-
sponse was described in recent reports by Yamazoe and Shimanoe (2010, 2011).
Yamazoe and Shimanoe (2010, 2011) assumed that the occupation of the
electronic states localized on the surfaces of grains is determined by the charac-
teristics of the depletion region if the length of the region defned by the Debye
screening length L
D
is comparable to the radius of the crystals, a. Since the den-
sity of the surface states depends on the amount and type of chemisorbed gas
species, an infuence of the grain size on the sensor response to gas can be in-
cluded in the simulation model. For this, however, a few general assumptions
are implicitly accepted, simplifying the theoretical defnition of the model called
receptor action of grain by Yamazoe and Shimanoe (2010, 2011). These simpli-
fcations extremely restrict the acceptability of the simulation model for analysis
of the resistance response of metal oxide polycrystalline sensors to both reducing
and oxidizing gases.
The frst assumption introduces an imaginary fat zone approach that omits
the fundamental origin of surface electronic levels. It is also assumed that, in an
almost insulating grain, the trapping of electrons into the surface levels created
by chemisorption can be simply described by statistics based on the Fermi distri-
bution, which allow calculation of the occupation of electron levels in the bulk of
semiconductors. Based on this, a relationship between the grain size and electron
transport is derived in the following form (Yamazoe and Shimanoe 2010, 2011):

[ ]





= =





2
0
2
exp exp
6
d F
D
S
N E a
R R
e kT L
(1.57)
where N
d
is the concentration of donors, [e]
S
is the density of conduction electrons
on the surfaces of grains, R
0
is the resistance of the sensor corresponding to the
fat-band assumption, L
D
is the Debye length, and DE
F
is the Fermi energy shift
due to the band bending at the surface.
The second assumption excludes adsorption and desorption of reducing gas
from the response mechanism. It is assumed that the adsorption and desorption
determines only the coverage of the grain surfaces with oxygen species, while the
density of reducing gas particles reacting with the surface oxygen is simply pro-
portional to the partial pressure of the gas in the atmosphere. Therefore, only the
rate equation for oxygen is included in the model in the following form (Yamazoe
and Shimanoe 2010, 2011):
nUmeriCal simUlation of eleCtriCal responses to Gases 33
[ ]
-
- -
-




= - -


2
2
1 O2 1 2 H2
O
O O
S
d
k P e k k P
dt
(1.58)
Here [O

] is the density of the chemisorbed oxygen species O

, P
O2
and P
H2
are the
partial pressures of oxygen and reducing gas in the atmosphere, respectively, k
1

and k
1
are the rate constants of the forward and reverse interactions between
the oxygen and the grain surfaces, respectively, and k
2
is the rate constant of the
reaction between the reducing gas (H
2
in this case) and the surface oxygen. This
simplifcation allows excluding from the model the coupling between the two rate
equations for both the oxygen and the reducing gas that was described by Eqs.
(1.17) and (1.18) in this chapter.
The third general assumption of Yamazoe and Shimanoe (2010, 2011) defnes
the interaction between the grain surfaces and the two types of oxidizing gases. In
the original study, oxygen and NO
2
gases were considered. It was assumed that
the total surface electrical charge Q
SC
was the same in both atmospheres, that is,
the clean air with only [O

] and the air with NO


2
gas. It was also assumed that the
proportion between the amounts of the two chemisorbed gas species was mainly
determined only by the adsorption rates. Therefore, the exposure to NO
2
in air
produced a change of the surface coverage with the species [O

] and [NO
2

] that
can be described by the following rate equations:
[ ]
-
-
-




= -


2
2 NO2 2 2
NO
NO
S
d
k P e k
dt
(1.59)
[ ]
-
-
-




= -


2
1 O2 1
O
O
S
d
k P e k
dt
(1.60)
Here k
2
and k
2
are the rate constants of adsorption and desorption of the oxidiz-
ing gas (NO
2
), respectively, and P
NO2
is the partial pressure of NO
2
gas in air.
Details about deriving the expressions that are acceptable for simulation of rate
transients of sensor response to both reducing and oxidizing gases are thoroughly
described in the reports of Yamazoe and Shimanoe (2010, 2011). In these studies,
problems were solved for the responses of a semiconductor gas sensor to particular
gases, that is, O
2
, H
2
, and NO
2
, however, the simulation model is acceptable for
sensors exposed to other analogous gases. In each case analyzed, the rate con-
stant (reciprocal of time constant) is determined as a combination of three terms,
which refect the rate of surface reactions, the semiconductor properties of the
constituent crystals, and the size (and shape) of the crystals. One of the terms in
34 ChemiCal sensors moDelinG anD simUlation: VolUme 2
the description of the simulation model displays the infuence of the depletion re-
gion on the sensor response. Yamazoe and Shimanoe (2010, 2011) found that the
rate of response is signifcantly reduced if the proportion a/L
D
increases beyond
5 for spherical crystals. In addition, in Yamazoe and Shimanoe (2011) there are
detailed instructions about how to derive an explicit equation for the response to
reducing gases (represented by H
2
). In that study, the simulation model was used
to reveal explicitly how various physicochemical parameters contribute to the gas
response as well as to fnd a way to proper analyses of response data.
It must be noted here that the model of Yamazoe and Shimanoe (2010, 2011)
contains a large number of original parameters. These parameters are hardly re-
placeable by the characteristics commonly used for description of surface pro-
cesses and electronic properties in felds such as surface science. Therefore, the
simulation of sensor responses is actually dependent on a large number of vari-
ables that must be independently defned and that can hardly be quantitatively
evaluated using the results of these specialized studies.
The most recent experimental investigations, namely, those of etkus et al.
(2009) and Bukauskas et al. (2010), demonstrated that a decrease in the grain size
down to nanometer scale creates additional problems in simulation of the sensor
resistance response of metal oxide gas sensors. It was found that the structure of
the flms can be signifcantly altered by a comparatively low external voltage applied
to the flm. For SnO
2-x
and In
2
O
3
flms with thickness < 30 nm and nanometer-size
grains (about 1020 nm), the grains are physically rearranged by applied voltages
exceeding some critical magnitude [about 1 V according to etkus et al. (2009)].
The structural changes are dependent on the surrounding atmosphere and are
comparatively larger in air with CO than in clean air, as was found by Bukauskas
et al. (2010). In metal oxide nanosystems, the electron transport is accompanied
by permanent changes in the surface properties and the intergranular junctions
that seem to originate from lattice oxygens hopping through the vacancies along
the external electric feld. Therefore, the classical approaches for improving the re-
sistance simulations models for nanosized metal oxides can hardly be successful.
Novel methods must be used for these metal oxidebased nanosystems.
3.3. nanostructurEd films
Nanostructured flms can be obtained from polycrystalline metal oxides in which
the grain size is drastically reduced to the dimensions comparable to the nano-
meter scale. It is expected that the properties of the nanostructured materials
may be much more acceptable than those used previously for the development of
gas sensors. According to Di Francia et al. (2009), reducing sensors to nanome-
ter size may be a way to make use of specifc low-dimension properties such as
quantum effects, high surface-to-volume ratio, controlled structure with a specifc
nUmeriCal simUlation of eleCtriCal responses to Gases 35
surface termination, nanoparticle doping, morphology, aggregation, and nanoma-
terial agglomeration state. However, the technology of chemical nanosensors is
still being developed; consequently, the properties of these sensors are not usually
reproducible in practice, and the quantitative defnition of the characteristics is
not very reliable. As a result, effort is still necessary to develop simulation models
for calculation of the response. In this respect, the models proposed so far seem
to be acceptablre as ideas for consideration in further studies.
Straightforward transfer of ideas about gas-sensing mechanisms in metal ox-
ides to nanosensors is the most typical approach in the development of response
simulation models. This approach seems most justifed for flms with nano-
metric grains, as proposed by Malagu et al. (2008). Highly reduced dimensions
of grains are comparable with the length of the depletion region at the surfaces
of the nanograins, similar to the ideas discussed in Section 2.3.2. Malagu et al.
(2008) proposed including both the thermionic and the tunneling current com-
ponents in the simulation of electron transport in flms with junctions between
the nanograins and the corresponding potential barriers. Accepting the Wentzel-
Kramers-Brillouin approximation to a double parabolic barrier at the junction
between the nanograins (Bender and Orszag 1978; Sakurai 1993), the electrical
current in the flms is defned as
= +
1 2
I I I (1.61)
where I
1
is the thermionic current component,

( )

= - +





1
1
00
0
2
exp
S S
kT y eV eV
I d
kT kT E
(1.62)
and I
2
is the tunneling current component,



= -




2
exp
S
eV
I
kT
(1.63)
where
( ) ( )
( )


+ -

= - -



0.5
0.5
0.5
1 1
1 ln y (1.64)
=
S
E
eV
(1.65)
36 ChemiCal sensors moDelinG anD simUlation: VolUme 2




1 2
*
4
d
N eh
E
m
(1.66)
In these defnitions, E is the energy of an electron in the conduction band.
The two-component current model allows one to prove numerically that the
electrical current in flms with nanograins is lower than with comparatively large
grains, because the increase in thermionic current across the junction between
the nanograins is practically compensated by the signifcant decrease in the tun-
neling current in flms with nanograins. Assuming that the barrier height V
S
de-
pends on the gas chemisorption, the response of nanograin-based flms to gas
can be simulated by an electrical current model analogous to that defned by
Eq. (1.61).
An exchange of electrons between the bulk and the surface states that results
in variation of the characteristics of the surface depletion layer is the core ef-
fect in most of the simulation models proposed for nanostructured gas-sensitive
materials. In a series of studies, by Lupan et al. (2010), Hongsith et al. (2010),
and Dmitriev et al. (2007), the simulation of the response of nanowires to gas
was based on variation of the characteristics of the depletion region and corre-
sponding changes in the conductive electrons. Lupan et al. (2010) proposed two
components for description of ZnO nanowire response to H
2
gas: an electron con-
centration term and a term of geometric factors. Accepting

=
2
0 0
r
G qn
l
(1.67)
for the conductance of the nanowire, the change in electrical conductance of the
nanowire exposed to gas (the electron concentration term) was determined by
Lupan et al. (2010) as a change in concentration of charge carriers Dn as follows:


=
S
g g
n G
G n
(1.68)
Here Dn
S
= n
g
n
0
, n
g
and n
0
are the concentrations of charge carriers correspond-
ing to gas and clean air, respectively, and r and l are the radius and length of the
nanowire, respectively.
The gas response term due to the changes in geometric factors includes the
length of the depletion area and, according to Lupan et al. (2010), can be obtained
from the following expression:

( ) ( )


-

= = - = -



1 2
1 2 1 2 0
0
2 2
g a
D
Da Dg Sa Sg
g g
G G
G
V V
G G r r qN
(1.69)
nUmeriCal simUlation of eleCtriCal responses to Gases 37
Here l
Da
and l
Dg
are the Debye radius in clean air and gas, respectively, and V
Sa

and V
Sg
are the band bending at the surfaces of nanowire in the clean air and gas,
respectively.
In general, the response simulation model for a nanowire sensor of Lupan
et al. (2010) was supported by the study of Hongsith et al. (2010), where the re-
sponse of a ZnO nanowire sensor to ethanol gas was described by the model. It
must be noted here that only one term was used in the model by Hongsith et al.
(2010), and it contained a specifc time constant because the rate equations of the
chemical reactions were used for development of the simulation model. In addi-
tion, it was suggested that the expression of the nanowire sensor response to gas
can be used not only for simulation of ZnO-based nanosensors but also for other
sensitive metal oxide nanostructures.
The effect of narrowing of the channel for the electron transport in nanow-
ires can be enhanced by intentional modulation of the nanowire diameter, as
proposed by Dmitriev et al. (2007). In contrast to the work of Lupan et al. (2010)
and Hongsith et al. (2010), the nanowire in the Dmitriev et al. (2007) study was
assumed to be constructed from two segments with equal length L but different
cross sections, namely, a thick segment with radius r
D
and a thin segment with
radius r. The electrical resistance of the nanowire with two segments and the con-
tacts connected in series is simply defned by




= + +


2 2
seg
2 2
2
1
D D
L r r
R
r r r
(1.70)
Here r is the initial resistivity of the segments and b is determined by a geometric
factor a that is independent of the radius of the nanowires (b = pa/2rL). Accepting
an infuence of the ionosorbed oxygen on thickness W of the depletion layer and,
consequently, on the effective diameter of the thin nanowire, r = r
0
+ W, the resis-
tance response of the segmented nanowire to gas can be calculated according to
the following expression:


-


= - +




1
2 2
seg
seg
2
seg 0
2
1 1
D
R
r r
S
R r r
(1.71)
This simulation model was applied to SnO
2
-based segmented nanowires by
Dmitriev et al. (2007). In the calculations of Dmitriev et al. (2007), the thick seg-
ments were accepted as being r
D
= 500 nm, while diameter of thin segments r was
varied from 10 to 60 nm. The simulation by Dmitriev et al. (2007) proved that
the response of segmented nanowires to gas is signifcantly higher than that of
straight (without the necks) nanowires.
38 ChemiCal sensors moDelinG anD simUlation: VolUme 2
Much more sophisticated methods can also be used for simulation of the
resistance response of nanowires to adsorption of gas. A nonequilibrium Greens
function formalism was used for analysis of electron-transport properties in
atomic-chain-scaled Si nanowires with two gold electrodes by Zhang et al. (2009).
The currentvoltage and conductancevoltage dependencies were calculated for
the nanowires without and with adsorbed gas molecules. The simulation demon-
strated that the conductance of Si nanowire depends on the distance between
the adsorption site and the electrode, though the overall transport properties are
hardly dependent on the gas adsorption. In fact, the simulation of Zhang et al.
(2009) suggests that the ionosorption-dependent depletion-layer approach may
not be acceptable for models of the response mechanism in comparatively very
short nanowires. The outcome of Zhang et al. (2009) was supported by analo-
gous results for single-walled carbon nanotubes (SWCNTs) by Hsieh et al. (2011).
Similar dependence of the electronic conductance on the distance between the gas
adsorption location and the electrode was obtained for SWCNTs by Hsieh et al.
(2011). The results were obtained from simulation of the conductance carried out
using the self-consistent nonequilibrium Greens function method for the system
Al electrodeSWCNTAl electrode and a single NO
2
molecule.
In spite of promising results, the simulation models of the response of nano-
structured materials to gas seem hardly acceptable as reliable at the present
state of investigations. Attempts to adapt the well-known approaches approved for
comparatively large sensors still have to be justifed theoretically. Various investi-
gations focused on nanosystems suggest that unique and unexpected phenomena
may exist in the systems at nanometer scale. These phenomena have to be con-
sidered in the development of simulation models for the nanosensors.
3.4. conductiVE polymEr layErs
In numerous studies (see reviews, e.g., Lange et al. 2008; Lu et al. 2011; Nambiar
and Yeow 2011), conductive polymers, have been shown to be acceptable for ap-
plication in chemical sensors and biosensors because of their highly attractive
properties. In spite of extensive investigations and promising application tests,
however, fundamental understanding of the processes and mechanisms in these
materials must still be signifcantly improved. Theoretical descriptions of electri-
cal properties and electrical responses to chemical interactions are only rarely
proposed in scientifc publications. Published theoretical studies frequently rep-
resent attempts to test separated ideas of diverse approaches. In general, a com-
paratively large variety of mechanisms have been proposed in different studies for
the description of electrical properties in the conductive polymers.
Lei et al. (2007) proposed a mechanism based on varying distances be-
tween the conductive carbon particles to describe the response to gas in a sensor
nUmeriCal simUlation of eleCtriCal responses to Gases 39
consisting of insulating polymers and dispersed carbon black particles. Based
on this mechanism, the mathematical description of the response of the sensor
to gas (chemical vapors) combines the dependence of electrical resistance on the
concentration of carbon black with the dependence of the volume fraction of com-
ponents in the composite on the vapor pressure of the vapor.
The resistivity of the sensor, defned as the resistivity of the composite mate-
rial, r
m
, can be derived from the following equation obtained on the basis of gene-
ral effective media:

( ) ( ) ( )


- - - -
- - - -
- - -
+ =
- -
1 1 1 1
1 1 1 1
1
0
k k k k
C m C m
k k k k
C m C m
f f
B B
(1.72)
where

-
=
1
C
C
f
B
f
(1.73)
r
C
and r
P
are the resistivities of the conductive component [carbon black in Lei
et al. (2007)] and the polymer, respectively, k is an exponent, f is the volume
fraction of the conductive component in the sensor, and f
C
is the critical volume
fraction of this conductive component at the percolation threshold. The volume
fraction of the other components in the sensor is defned by (1 f ), which, in
general, includes both the polymer and the vapor particles and consequently de-
pends on the surrounding atmosphere. Supposing that the sensor response is
produced by only one target gas, the volume fraction of the nonconductive com-
ponents (1 f ) is defned by the following relationship:

( )
-
=
+ -
0
0
1
CC A
A p CC
f f f
f f f f
(1.74)
Here f
A
and f
p
are the volume fractions of the absorbed vapor particles and the
polymer in the sensor; f
0CC
is the volume fraction of the conductive component in
the sensor in the clean air. It follows from (1.74) that (1 f ) = f
p
in the sensor in
clean air (f
A
= 0).
Lei et al. (2007), defned the dependence of absorption of the target gas (vapor)
by the sensor material on the vapor pressure in terms of thermodynamic theory.
According to the defnition, the partial pressure of the target gas is related to the
volume fraction of the absorbed gas in the composite structure of the sensor. Lei
et al. (2007) write this relationship as follows:




= + + +





2
sat
1
1
ln ln 1
p A A
P
f f f
m P
(1.75)
40 ChemiCal sensors moDelinG anD simUlation: VolUme 2
Here P and P
sat
are the partial pressure and the saturated partial pressure of the
target gas, respectively, c is the gaspolymer interaction parameter, m = v
p
/v
A
,
and v
p
and v
A
are the molar volumes of the polymer and the gas, respectively.
According to the simulation model of Lei et al. (2007), the resistance response
(r
Am
r
0m
) of a conductive polymer sensor to gas with partial pressure P can be
obtained by calculating r
0m
and r
Am
from (1.72) with (1.73)(1.75) for the clean air
and the air with the target gas, respectively. This resistance response simulation
model is acceptable for evaluation of the sensor responses to the vapor at various
amounts of the vapor.
Absorption of analytes (vapor) from gaseous surroundings can directly change
the electrical properties of conducting polymer sensors, as a result of localiza-
tion/delocalization of electrons at the absorption sites, similar to some extent to
the chemisorption effects in metal oxidebased gas sensors. Transfer of electrons
between the adsorbed analyte and the conductive polymer can change the resis-
tance of the sensor because the doping technology in most conductive polymers
is based on the changes produced by redox reactions. In spite of existing qualita-
tive interpretations of the effect, the mechanisms determining conversion of the
analytepolymer interaction into a response signal still have to be described in
more detail, the absence of which is crucial for development of simulation models
of sensor response. The state of the art of studies of conductive polymer sensors
was reviewed recently by Bai and Shi (2007).
It is commonly accepted now that conductive polymers can be used success-
fully for development of chemical sensors, producing a resistance response to
volatile chemical compounds. However, more information and better understand-
ing is required for development of strictly defned theoretical descriptions of the
conversion of the interaction between the conductive polymers and analytes into
the response signal.
3.5. molEcular structurEs
Molecular-scale electronic devices typically are assumed to be practical imple-
mentations of understanding about the processes and mechanisms specifc to
systems with single molecules. In these systems, the molecules themselves can
emulate the properties of well-known solid-state devices and can stand for novel
nanosystems with unique properties. Fundamental studies of charge transport
through an interface between metals or semiconductors and organic or inorganic
molecules not only opened various ways for innovative developments of molecular
electronics but also revealed serious problems defning new challenges. Research
reviews on this topic include, among others, those of Weiss et al. (2007) and Heath
(2009). It is recognised by researchers that the ultimate aim of studies on electron
transport in single molecules is to defne the infuence of metal electrodes on the
nUmeriCal simUlation of eleCtriCal responses to Gases 41
energy spectrum of the molecule and to understand how the electron-transport
properties of the molecule depend on the strength of the electronic coupling be-
tween it and the electrodes. A variety of phenomena are observed to infuence the
charge transport in these systems, including those acceptable for recognition of
chemical compounds and chemical interactions between the molecular system
and the surroundings.
Depending on the mechanism determining the charge transport through the
molecular systems, it is possible to classify the emerging approaches in the re-
sponse of electrical parameters to chemical surrounding as the bridging-effect
recognizing systems, interaction-dependent tunneling transport devices, and
organic molecule detectors.
The molecular bridging system reviewed by Lindsay et al. (2010) consists of
two metal electrodes separated by a gap L. The electric current through the gap
can be originated by the electrons in the existing highest occupied energy state
close to the Fermi energy, E
F
. The potential barrier that retains electrons within
the metal is V, so the work function, , is given by = V E
F
. In the absence of a
state close to E
F
, in the gap the tunnel conductance is given approximately by

( )
( ) - = -
0 0
exp 1.02 exp G G L G L (1.76)
where G
0
is the quantum of conductance, is the work function in electron volts,
L is the tunnel gap in angstroms, and b is the inverse electronic decay length. A
real molecule spanning the gap will consist of molecular orbitals y
n
, often well
described in terms of linear combinations of the atomic orbitals of the constituent
atoms and the hopping matrix elements between adjacent orbitals. An electron is
propagated from an electrode to the opposite one through all the paths made of
interacting orbitals and connecting both electrodes. Mathematically, the trans-
mission is calculated with a Greens function,
( )

- +

,
0
| |
n n
L R
n
n
L R
G E
E E i
(1.77)
Here, L and R represent states at the energy E on the left and right electrodes, E
0n
is the eigenenergy of the nth eigenstate of the system, and is an infnitesimal.
The presence of G(E) in the transmission includes the energy levels of the mol-
ecule and naturally introduces the possibility of resonances and multiple refec-
tions. Therefore, the electron current through the molecular bridge between the
metal electrodes is individual to the molecule type. This current appears only if
the molecule binds to both electrodes. Based on this simulation model, a possibil-
ity exists to identify molecules and specifc molecular structures, as suggested by
Lindsay et al. (2010).
42 ChemiCal sensors moDelinG anD simUlation: VolUme 2
A simulation model of a molecular wire containing a redox system was pro-
posed by Crus et al. (2010). In this study, interaction of the redox system with a
classical solvent was analyzed, assuming that the solvent state was represented
by a solvent coordinate q in the spirit of the Marcus theory (Marcus 1956). Using
the wide-band approximation, an exact expression for the quantum conductance
of a chain of arbitrary length was derived in matrix form:
Tr
a
L R
g G G



=


(1.78)
Here the subscripts L and R refer to the left and right reservoirs (electrodes), G
denotes the imaginary part of the self-energy, G is the Greens function obtained
from a special Hamiltonian defning the model system. The index r represents the
redox species on the molecular wire, while a means the chain of atoms. Based on
(1.78), an infuence of electrochemical environment on the molecular wire with the
redox system was demonstrated by Crus et al. (2010) by means of explicit calcula-
tions performed for a chain of three atoms.
A classical approach in analysis of a one-dimensional conductivity problem
(molecular wire) was proposed by Koslowski and Wilkening (2010). In this work,
the solution of the problem is based on Kirchhoffs second law applied to a resis-
tor network with the nodes corresponding to individual molecules. The nodes are
connected by channels characterized by local conductance G. It is assumed that
an individual node can only be occupied by a limited number of charge carriers.
Electron transfer between the nodes is defned by a classical Master equation that
relates the occupation probability of individual node p
i
and the transfer rates be-
tween the adjacent nodes, w
ij
.
In general, the equations can be solved only numerically, but analytical ex-
pressions can be obtained for illustration in a highly simplifed system. This sim-
plifed system contains two sites connected to two leads. Conductance between
the two sites is accepted to be G
12
= G, and the conductances between sites and
adjacent electrodes are G
1e
and G
2e
, respectively. Supposing p
1
and p
2
as the occu-
pation probabilities of an individual site, V
1
and V
2
are the site voltages, while V
1e

and V
2e
are the external voltages applied to the system electrodes. For this simpli-
fed system of Koslowski (2010), the current is defned by the following analytical
expression:

( )
( ) ( )
-
=
+ + -
2
1 2 1 1
2
1 2 1 1 2 1
1
1
e e
e e e e
G G Gp p
I
G G G p G G p
(1.79)
This expression gives a general idea about the parameters that are important
for the simulation model of a molecular wire. It also follows from this expression
that the current depends on the structure and the components of the wire and,
consequently, it can be used for identifcation of the molecule in a single-molecule
nUmeriCal simUlation of eleCtriCal responses to Gases 43
wire. Moreover, supposing that the conductivity of the components and occupa-
tion of the sites can be infuenced by an interaction between the molecular wire
and the surrounding media, the theoretical model proposed by Koslowski and
Wilkening (2010) can be adapted for simulation of an electrical response of a mo-
lecular wire to the chemical surroundings.
4. concludIng comments
This chapter has provided an overview of simulation models of resistive gas sen-
sors reported in the literature over a considerable period of research and appli-
cation activities. The systematic search for publications with suggestions about
sensor response simulations has revealed that theoretical calculations of the re-
sponse are frequently based on empirical formulas aiming to quantitatively defne
some specifc characteristics of sensor response to gas. The detailed description
of gas sensor properties acceptable for the basic simulation of sensor responses
to gas actually requires specifc combinations of approaches from diverse sci-
entifc areas such as state electronics, semiconductor physics, surface states,
heterogeneous catalysis, adsorption, transport in gases, etc. This considerably
complicates the simulation of the sensor response and, possibly, explains the
relative lack of reports about such studies in the literature. It must be noted here
that various aspects of the problems have already been analyzed in specifc stud-
ies focused on surface chemical interactions, properties of the surface electronic
states, electrical properties of nanostructured semiconductors, etc., and reported
in numerous specialized scientifc journals. Justifed usage of proposed ideas and
integration of the diverse problem descriptions into a systematic model seem to
be the main diffculties in the development of a well-organized, reliable simulation
model of gas sensor response.
It is obvious from the literature that the signifcant reduction of sensors down
to the nanometer scale and to the molecule size creates new challenges of gas
sensor modeling by introducing unexpected and unique features of the nano-
systems. The methods and the physical theories of the nanosystems that are
intensively discussed in the dedicated publications are hardly refected in the
gas sensor studies. However, the idea is increasingly supported that the pro-
cesses in naonsystems can hardly be simulated by classical physical models with
simply reduced dimensions. It seems that the processes and the mechanisms in
solid nanosystems are frequently highly specifc and cannot be predicted by these
classical models, as demonstrated, for instance, for TiO
2
based nanosystems by
Strukov et al. (2008). These fundamental problems in understanding the response
mechanisms may be the root cause of unexpectedly slow progress in obtaining
considerable improvements in gas-sensing characteristics using the promising
technologies of nanostructured sensors.
44 ChemiCal sensors moDelinG anD simUlation: VolUme 2
The existing simulation models are mainly acceptable for quantitative descrip-
tion of some fundamental aspects in the sensor response to gas and are useful in
explaining the basic mechanisms of sensor functioning. However, the infuence
of the basic parameters of sensing materials on the sensor response to gas is in-
suffciently defned. Attempts to improve the simulation models of the resistance
response to gas are frequently aimed at enhancing the description of the elec-
tronic processes at the surfaces in the semiconducting materials of the sensors.
A distinct lack of explicit description of the relationship between the sensor char-
acteristics and the properties of both the materials and the constructions is likely
discouraging developers from using numerical tests in research and development
projects. Therefore, fundamental studies focused on enhancing the theory of gas
sensors and response simulation models seems to be highly important in accel-
erating further progress in gas sensor technology and applications. It is hoped
that the overview in this chapter about the present state of response simulation
models will be interesting for gas sensor developers and persuade them to carry
out much more systematic research on sensor technologies and functioning.
references
Ahlers S., Muller G., and Doll T. (2005) A rate equation approach to the gas sensitivity
of thin flm metal oxide materials. Sens. Actuators B 107, 587599. DOI: 10.1016/j.
snb.2004.11.020
Bai H. and Shi G. (2007) Gas sensor based on conducting polymers. Review. Sensors 7,
267307. DOI: 10.3390/s7030267
Barsan N., Koziej D., and Weimar U. (2001) Conduction model of metal oxide gas sensors.
J. Electroceram. 7, 143167. DOI: 10.1023/A:1014405811371
Barsan N., Koziej D., and Weimar U. (2007) Metal oxide-based gas sensor research: How
to? Sens. Actuators B 121, 1835. DOI: 10.1016/j.snb.2006.09.047
Batzill M. and Diebold U. (2005) The surface and materials science of tin oxide. Prog. Surf.
Sci. 79, 47154. DOI: 10.1016/j.progsurf.2005.09.002.
Bender C.M. and Orszag S.A. (1978) Advanced Mathematical Methods for Scientists and
Engineers. McGraw-Hill, New York.
Blatter G. and Greuter F. (1986a) Carrier transport through grain boundaries in semicon-
ductors. Phys. Rev. B 33, 39523966. DOI: 10.1103/PhysRevB.33.3952
Blatter G. and Greuter F. (1986b) Electrical breakdown at semiconductor grain bounda-
ries. Phys. Rev. B 34, 85558572. DOI: 10.1103/PhysRevB.34.8555.
Boeker P., Wallenfang O., and Horner G. (2002) Mechanistic model of diffusion and re-
action in thin sensor layersThe DIRMAS model. Sens. Actuators B 83, 202208.
DOI: 10.1016/S0925-4005(01)01041-3.
Bukauskas V., Mironas A., etkus A., and Strazdien V. (2010) Nanostructures pro-
duced by SPM voltage ramping in metal oxide flms. Surf. Interface Anal. 42, 991995.
DOI: 10.1002/sia.3328.
Busse H., Voss M.R., Jerdev D., Koel B.E., and Paffet M.T. (2001) Adsorption and reaction
nUmeriCal simUlation of eleCtriCal responses to Gases 45
of gaseous H(D) atoms with D(H) adatoms on Pt(1 1 1) and Sn/Pt(1 1 1) surface alloys.
Surf. Sci. 490, 133144. DOI: 10.1016/S0039-6028(01)01323-1
Carlsson A.F. and Madix R.J. (2001) The extrinsic precursor kinetics of molecular methane
adsorption onto Pt(1 1 1)p(2 2)-O, n-nutylidyne and isobutylidyne. Surf. Sci. 479,
98109. DOI: 10.1016/S0039-6028(01)00960-8
Cerofolini G.F. (2003) Theory of localized adsorption on surfaces undergoing reversible re-
construction. Phys. Rev. E 67, 041603-17. DOI: 10.1103/PhysRevE.67.041603.
Clifford P.K. and Tuma D.T. (1982/1983) Characteristics of semiconductor gas sen-
sors. II. Transient response of temperature change. Sens. Actuators 3, 255281.
DOI: 10.1016/0250-6874(82)80027-9
Cruz A.V.B., Mishra A.K., and Schmickler W. (2010) Electron tunneling between two elec-
trodes mediated by a molecular wire containing a redox center. Chem. Phys. 371, 10
15. DOI: 10.1016/j.chemphys.2010.01.009
Diebold U. (2003) The surface science of titanium dioxide. Surf. Sci. Rep. 48, 53229.
DOI: 10.1016/S0167-5729(02)00100-0
Di Francia G., Alfano B., and La Ferrara V. (2009) Conductometric gas nanosensors.
J. Sensors 209, ID 659275-118; DOI: 10.1155/2009/659275
Dmitriev S., Lilach Y., Button B., Moskovits M., and Kolmakov A. (2007) Nanoengineered
chemiresistors: The interplay between electron transport and chemisorption proper-
ties of morphologically encoded SnO
2
nanowires. Nanotechnology 18, 055707-16.
DOI: 10.1088/0957-4484/18/5/055707
Fergus J.W. (2008) A review of electrolyte and electrode materials for high tempera-
ture electrochemical CO
2
and SO
2
gas sensos. Sens. Actuators B 134, 10341041.
DOI: 10.1016/j.snb.2008.07.005
Galdikas A., Mironas A., Senulien D., and etkus A. (2000) Specifc set of the time con-
stants for characterisation of organic volatile compounds in the output of metal oxide
sensors. Sens. Actuators B 68, 335343. DOI: 10.1016/S0925-4005(00)00454-8
Gardner J.W. (1990) A non-linear diffusion-reaction model of electrical conduction in semicon-
ductor gas sensors. Sens. Actuators B 1, 166176. DOI: 10.1016/0925-4005(90)80194-5
Guerin J., Bendahan M., and Aguir K. (2008) A dynamic response model for the WO
3
-based
ozone sensors. Sens. Actuators B 128, 462467. DOI: 10.1016/j.snb.2007.07.010
Heath J.R. (2009) Molecular electronics. Annu. Rev. Mater. Res. 39, 123. DOI: 10.1146/
annurev-matsci-082908-145401
Hongsith N., Wongrat E., Kerdcharoen T., and Choopun S. (2010) Sensor response formula
for sensor based on ZnO nanostructures. Sens. Actuators B 144, 6772. DOI: 10.1016/
j.snb.2009.10.037
Hsieh S.-C., Wang S.-M., and Li F.-Y. (2011) A theoretical investigation of the effect of ad-
sorbed NO
2
molecules on electronic transport in semiconducting single-walled carbon
nanotubes. Carbon 49, 955965. DOI: 10.1016/j.carbon.2010.09.062
Hubert T., Boon-Brett L., Black G., and Banach U. (2011) Hydrogen sensorsA review.
Sens. Actuators B 157, 329352; DOI: 10.1016/j.snb.2011.04.070.
Kanefusa S., Nitta M., and Haradome M. (1981) Oscillations in a SnO
2
-based gas sensing
device exposed to H
2
gas. J. Appl. Phys. 52, 498499. DOI: 10.1063/1.328433
King P.D.C., Veal T.D., Fuchs F., Wang Ch.Y., Payne D.J., Bourlange A., Zhang H., Bell
G.R., Cimalla V., Ambacher O., Egdell R.G., Bechstedt F., and McConville C.F. (2009)
46 ChemiCal sensors moDelinG anD simUlation: VolUme 2
Band gap, electronic structure, and surface electron accumulation of cubic and rhombo-
hedral In
2
O
3
. Phys. Rev. B 79, 205211-110. DOI: 10.1103/PhysRevB.79.205211
Koslowski T. and Wilkening A. (2010) A combined KirchhoffMaster equation approach
to electronic transport in one-dimensional molecular conductors. Chem. Phys. 369,
2226. DOI: 10.1016/j.chemphys.2010.02.001
Kreuzer H.J. (1990) Theory of surface processes. Appl. Phys. A 51, 491497; DOI: 10.1007/
BF00324732
Lange U., Roznyatovskaya N.V., and Mirsky V.M. (2008) Conducting polymers in chemical
sensors and arrays. Anal. Chim. Acta 614, 126. DOI: 10.1016/j.aca.2008.02.068
Lantto V. and Romppainen P. (1987) Electrical studies on the reactions of CO with dif-
ferent oxygen species on SnO
2
surfaces. Surf. Sci. 192, 243264. DOI: 10.1016/
S0039-6028(87)81174-3
Lee S.P. (2006) Electrical behaviour in gassolid interface of gas sensors based on
oxide semiconductors. Int. J. Appl. Ceram. Technol. 3, 225229. DOI: 10.1111
/j.1744-7402.2006.02074.x
Lei H., Pitt W.G., McGrath L.K., and Ho C.K. (2007) Modeling carbon black/polymer com-
posite sensors. Sens. Actuators B 125, 396407. DOI: 10.1016/j.snb.2007.02.041
Liess M. (2002) Electric-feld-induced migration of chemisorbed gas molecules on a sen-
sitive flmA new chemical sensor. Thin Solid Films 410, 183187. DOI: 10.1016/
S0040-6090(02)00209-2
Lindsay S., He J., Sankey O., Hapala P., Jelinek P., Zhang P., Chang S., and Huang S.
(2010) Recognition tunnelling. Topical review. Nanotechnnology 21, 262001-112;
DOI: 10.1088/0957-4484/21/26/262001
Lu X., Zhang W., Wang C., Wen T.-C., and Wei Y. (2011) One-dimensional conducting
polymer nanocomposites: Synthesis, properties and applications. Prog. Polym. Sci. 36,
671712. DOI: 10.1016/j.progpolymsci.2010.07.010
Lupan O., Ursaki V.V., Chai G., Chow L., Emelchenko G.A., Tiginyanu I.M., Gruzintsev
A.N., and Redkin A.N. (2010) Selective hydrogen gas nanosensor using individual
ZnO nanowire with fast response at room temperature. Sens. Actuators B 144, 5666.
DOI: 10.1016/j.snb.2009.10.038
Malagu C., Martinelli G., Ponce M.A., and Aldao C.M. (2008) Unpinning of the Fermi level
and tunneling in metal oxide semiconductors. Appl. Phys. Lett. 92(16), 162104-13;
DOI: 10.1063/1.2916709
Malagu C., Carotta M.C., Giberti A., Guidi V., Martinelli G., Ponce M.A., Castro M.S.,
and Aldao C.M. (2009) Two mechanisms of conduction in polycrystalline SnO
2
. Sens.
Actuators B 136, 230234. DOI: 10.1063/1.2916709
Marcus R.A. (1956) On the theory of oxidation-reduction reactions involving electron trans-
fer. I. J. Chem. Phys. 24(5), 966978; DOI: 10.1063/1.1742723.
Matsunaga N., Sakai G., Shimanoe K., and Yamazoe N. (2002) Diffusion equation-based
study of thin flm semiconductor gas sensor-response transient. Sens. Actuators B 83,
216221. DOI: 10.1016/S0925-4005(01)01043-7
Matsunaga N., Sakai G., Shimanoe K., and Yamazoe N. (2003) Formulation of gas diffusion
dynamics for thin flm semiconductor gas sensor based on simple reactiondiffusion
equation. Sens. Actuators B 96, 226233. DOI: 10.1016/S0925-4005(03)00529-X
Nakata S., Kato Y., Kaneda Y., and Yoshikawa K. (1996) Rhythmic chemical reaction of
nUmeriCal simUlation of eleCtriCal responses to Gases 47
CO on the surface of a SnO
2
gas sensor. Appl. Surf. Sci. 103, 369376. DOI: 10.1016/
S0169-4332(96)00551-X
Nakata S., Takemura T., and Neya K. (2002) Non-linear dynamic response of semiconduc-
tor gas sensor. Sens. Actuators B 76, 436441. DOI: 10.1016/S0925-4005(01)00652-9
Nambiar S. and Yeow J.T.W. (2011) Conductive polymer-based sensors for biomedical ap-
plications. Biosens. Bioelectron. 26, 18251832. DOI: 10.1016/j.bios.2010.09.046
Nitta M., Kanefusa S., Taketa Y., and Haradome M. (1978) Oscillation phenomenon in ThO
2
-
doped SnO
2
exposed to CO gas. Appl. Phys. Lett. 32, 590591. DOI: 10.1063/1.90114
Nitta M. and Haradome M. (1979) Oscillation phenomenon in thick-flm CO sensor. IEEE
Trans. Electron. Dev. ED-26, 219223. DOI: 10.1109/T-ED.1979.19408
ONeil D.H., Walsh A., Jacobs R.M.J., Kuznetsov V.L., Egdell R.G., and Edwards P.P. (2010)
Experimental and density-functional study of the electronic structure of In
4
Sn
3
O
12
.
Phys. Rev. B 81, 085110-18; DOI: 10.1103/PhysRevB.81.085110.
Oprea A., Barsan N., and Weimar U. (2009) Work function changes in gas sensitive mate-
rials: Fundamentals and applications. Sens. Actuators B 142, 470493. DOI: 10.1016/
j.snb.2009.06.043
Panczyk T. and Rudzinski W. (2004) A statistical rate theory approach to kinetics of dis-
sociative gas adsorption on solids. J. Phys. Chem. B 108, 28982909. DOI: 10.1021/
jp034786+
Payne S.H., McEwen J.-S., Kreuzer H.J., and Menzel D. (2006) Lateral interactions and
nonequilibrium in adsorption and desorption. Part 2. A kinetic lattice gas model for
(22)-(3O+NO)/Ru(001). Surf. Sci. 600, 46604669. DOI: 10.1016/j.susc.2006.07.019
Persson B.N.J. (1992) Ordered structures and phase transitions in adsorbed layers. Surf.
Sci. Rep. 15, 17. DOI: 10.1016/0167-5729(92)90012-Z
Prince M.B. (1953) Drift mobilities in semiconductors. I. Germanium. Phys. Rev. 92, 681
687. DOI: 10.1103/PhysRev.92.681
Prince M.B. (1954) Drift mobilities in semiconductors. II. Silicon. Phys. Rev. 93, 1204
1206. DOI: 10.1103/PhysRev.93.1204
Rantala T.S., Lantto V., and Rantala T.T. (1993) Rate equation simulation of the height of
Schottky barriers at the surface of oxidic semiconductors. Sens. Actuators B 13/14,
234237. DOI: 10.1016/0925-4005(93)85369-L
Rder-Roith U., Rettig F., Rder T., Janek J., Moos R., and Sahner K. (2009) Thick-flm solid
electrolyte oxygen sensors using the direct ionic thermoelectric effect. Sens. Actuators B
136, 530535. DOI: 10.1016/j.snb.2008.12.024
Rothschild A. and Komem Y. (2003) Numerical computation of chemisorption isotherms
for device modeling of semiconductor gas sensors. Sens. Actuators B 93, 362369.
DOI: 10.1016/S0925-4005(03)00212-0
Rudzinski W., Panczyk T., and Plazinski W. (2005) Kinetics of isothermal gas adsorption
on heterogeneous solid surfaces: Equations based on generalization of the statistical
rate theory of interfacial transport. J. Phys. Chem. B 109, 2186821878. DOI: 10.1021/
jp052671v
Sakai G., Matsunaga N., Shimanoe K., and Yamazoe N. (2001) Theory of gas diffusion
controlled sensitivity for thin flm semiconductor gas sensor. Sens. Actuators B 80,
125131. DOI: 10.1016/S0925-4005(01)00890-5
Sakurai J.J. (1993) Modern Quantum Mechanics. Addison-Wesley, Lebanon, IN, USA.
48 ChemiCal sensors moDelinG anD simUlation: VolUme 2
Schonauer D., Nieder T., Wiesner K., Fleischer M., and Moos R. (2011) Investigation of the
electrode effect in mixed potential type ammonia exhaust gas sensors. Solid State Ionics
192, 3841. DOI: 10.1016/j.ssi.2010.03.028
Sensato F.R., Custodio R., Calatayud M., Beltran A., Andres J., Sambrano J.R., and Longo
E. (2002) Periodic study on the structural and electronic properties of bulk, oxidized
and reduced SnO
2
(1 1 0) surfaces and the interaction with O
2
. Surf. Sci. 511, 408420.
DOI: 10.1016/S0039-6028(02)01542-X
etkus A. (2002) Heterogeneous reaction rate based description of the response kinet-
ics in metal oxide gas sensors. Sens. Actuators B 87, 346357. DOI: 10.1016/
S0925-4005(02)00269-1
etkus A. (2004) Detection of dynamic smell intensity. Characterisation of explosives by
the response of gas sensor array to modulated intensity of the smell. In: Gardner J.,
and Yinon J. (eds.), Proceedings of the NATO Advanced Research Workshop: Electronic
Noses and Sensors for the Detection of Explosives, Sept. 30Oct. 3, 2003, Warwick,
U.K. NATO Science Series II: Mathematics, Physics and Chemistry Vol. 159, Kluwer
Academic Publishers, Dordrecht, The Netherlands, 159179.
etkus A., Baratto C., Comini E., Faglia G., Galdikas A., Kancleris ., Sberveglieri G., and
Senulien D. (2004) Infuence of metallic impurities on response kinetics in metal oxide
thin flm gas sensors. Sens. Actuators B 103, 448456. DOI: 10.1016/j.snb.2004.05.004
etkus A., Kaciulis S., Pandolf L., Senulien D., and Strazdien V. (2005) Tuning of the
response kinetics by the impurity concentration in metal oxide gas sensors. Sens.
Actuators B 111112, 3644. DOI: 10.1016/j.snb.2005.07.021
etkus A., Bukauskas V., Mironas A., Senulien D., and Strazdien V. (2009) Specifc re-
sponse of ultra-thin metal oxide flms to gas. Phys. Status Solidi C 6, 27532755.
Simakov V., Yakusheva O., Grebennikov A., and Kisin V. (2006) IV characteristics of gas-
sensitive structures based on tin oxide thin flms. Sens. Actuators B 116, 221225.
DOI: 10.1016/j.snb.2005.11.077
Simon I., Barsan N., Bauer M., and Weimar U. (2001) Micromachined metal oxide gas
sensors: Opportunities to improve sensor performance. Sens. Actuators B 73, 126.
DOI: 10.1016/S0925-4005(00)00639-0
Strukov D.B., Snider G.S., Stewart D.R., and Williams R.S. (2008) The missing memristor
found. Nature Lett. 453, 8083. DOI: 10.1038/nature06932
Varpula A., Novikov S., Sinkkonen J., and Utriainen M. (2008) Bias dependent sen-
sitivity in metal-oxide gas sensors. Sens. Actuators B 131, 134142. DOI: 10.1016/
j.snb.2007.12.013
Varpula A., Novikov S., Haarahiltunen A., and Kuivalainen P. (2011) Transient characteri-
zation techniques for resistive metal-oxide gas sensors. Sens. Actuators B 159, 1226.
DOI: 10.1016/j.snb.2011.05.059
Vilanova X., Llobet E., Brezmes J., Calderer J., and Correig X. (1998) Numerical simulation
of the electrode geometry and position effects on semiconductor gas sensor response.
Sens. Actuators B 48, 425431. DOI: 10.1016/S0925-4005(98)00080-X
Walsh A. (2011) Surface oxygen vacancy origin of electron accumulation in indium oxide.
Appl. Phys. Lett. 98, 261910-13. DOI: 10.1063/1.3604811
Weiss E.A., Kriebel J.K., Rampi M.-A., and Whitesides G.M. (2007) The study of charge
transport through organic thin flms: Mechanism, tools and applications. Phil. Trans. R.
Soc. A 365, 15091537. DOI: 10.1098/rsta.2007.2029
nUmeriCal simUlation of eleCtriCal responses to Gases 49
Wolkenstein Th. (1963) The Electron Theory of Catalysis on Semiconductors. Macmillan,
New York.
Wolkenstein T. (1991) Electronic Processes on Semiconductor Surfaces During Chemisorption.
Consultants Bureau, New York. DOI: 10.1007/978-1-4615-3656-7
Xu C. and Koel B.E. (1994) Adsorption kinetics on chemically modifed or bimetallic sur-
faces. J. Chem. Phys. 100, 664670. DOI: 10.1063/1.466931
Yamazoe N. and Shimanoe K. (2010) Theoretical approach to the rate of response of semicon-
ductor gas sensor. Sens. Actuators B 150, 132140. DOI: 10.1016/j.snb.2010.07.030
Yamazoe N. and Shimanoe K. (2011) Explicit formulation for the response of neat oxide semi-
conductor gas sensor to reducing gas. Sens. Actuators B 158, 2834. DOI: 10.1016/j.
snb.2011.04.068
Zhang Y.H., Zhang X.Q., Li H., Taft C.A., and Paiva G. (2009) A molecule detector: Adsorbate
induced conductance gap change of ultra-thin silicon nanowire. Surf. Sci. 603, 847
851. DOI: 10.1016/j.susc.2009.01.025
Zhdanov V. P. (2001) Adsorption-desorption kinetics and chemical potential of adsorbed
and gas-phase particles. J. Chem. Phys. 114, 4746-4748. DOI: 10.1063/1.1349178
Zhdanov V. P. (2002) Impact of surface science on the understanding of kinetics of
hetero geneous catalytic reaction. Surf. Sci. 500, 965984. DOI: 10.1016/S0039-
6028(01)01626-0
Zhuiykov S. (2008) Gas sensor applications of oxygen-ionic electrolytes: Development of
their electron model. Sens. Actuators B 130, 488496. DOI: 10.1016/j.snb.2007.09.027
Also from Momentum Press
The Essentials of Finite Element Modeling and Adaptive Refnement: For
Beginning Analysts to Advanced Researchers in Solid Mechanics by John O. Dow
Virtual Engineering by Joe Cecil
Reduce Your Engineering Drawing Errors: Preventing the Most Common Mistakes
by Ronald Hanifan
Biomedical Sensors by Deric P. Jones
Bio-Inspired Engineering by Chris Jenkins
The Six Volume Set of Chemical Sensors Technologies, also edited by Ghenadii
Korotcenkov, which includes Fundamentals of Sensing Materials, Solid State
Devices, Electrochemical and Optical Sensors, Polymers and Other Materials,
Microstructural Characterization and Modeling of Metal Oxides, as well as
General Approaches and Applications.
For more information, please visit www.momentumpress.net
___________________________________________________________________
The Momentum Press Digital Library
45 Engineering Ebooks in One Digital Library or As Selected Topical Collections
Our ebrary platform, our digital collection features
a one-time purchase; not subscription-based,
that is owned forever,
allows for simultaneous readers,
has no restrictions on printing, and
can be downloaded as a pdf.
The Momentum Press digital library is an affordable way to give many readers
simultaneous access to expert content.
For more information, please visit www.momentumpress.net/library.
To set up a trial, please contact adam.chesler@momentumpress.net.
Momentum Press is proud to bring to you Chemical Sensors: Simulation and Modeling Volume 2: Conductometric-
Type Sensors, edited by Ghenadii Korotcenkov. This is the second of a new fve-volume comprehensive refer-
ence work that provides computer simulation and modeling techniques in various felds of chemical sensing
and the important applications for chemical sensing such as bulk and surface diffusion, adsorption, surface
reactions, sintering, conductivity, mass transport, and interphase interactions. In this second volume, you will
fnd background and guidance on:
Phenomenological modeling and computational design of conductometric chemical sensors, based on
nanostructured materials such as metal oxides, carbon nanotubes, and graphenes
Approachesusedtoquantitativelyevaluatecharacteristicsofsensitivestructuresinwhichelectriccharge
transport depends on the interaction between the surfaces of the structures and chemical compounds in
the surroundings
Chemical sensors are integral to the automation of myriad industrial processes and everyday monitoring of
such activities as public safety, engine performance, medical therapeutics, and many more. This fve-volume
reference work covering simulation and modeling will serve as the perfect complement to Momentum Presss
6-volume reference work, Chemical Sensors: Fundamentals of Sensing Materials and Chemical Sensors: Compre-
hensive Sensor Technologies, which present detailed information related to materials, technologies, construction,
and application of various devices for chemical sensing. Each simulation and modeling volume in the present
series reviews modeling principles and approaches peculiar to specifc groups of materials and devices applied
for chemical sensing.
About the editor
Ghenadii Korotcenkov received his Ph.D. in Physics and Technology of Semiconductor Materials and De-
vices in 1976, and his Habilitate Degree (Dr.Sci.) in Physics and Mathematics of Semiconductors and Dielec-
trics in 1990. For many years, he was a leader of the Gas Sensor Group, and manager of various national and
international scientifc and engineering projects carried out in the Laboratory of Micro- and Optoelectronics,
Technical University of Moldova. Currently, Dr. Korotcenkov is a research Professor at the Gwangju Institute
of Science and Technology, Republic of Korea. His research has included signifcant work on Schottky barri-
ers, MOS structures, native oxides, and photo receivers on the base of III-Vs compounds. He continues with
research in various aspects of materials sciences and surface science, with a particular focus on nanostructured
metal oxides and solid state gas sensor design. Dr. Korotcenkov is the author or editor of eleven books and spe-
cial issues, eleven invited review papers, seventeen book chapters, and more than 190 peer-reviewed articles.
HisresearchactivitieshavebeenhonoredwiththeAwardoftheSupremeCouncilofScienceandAdvanced
Technology of the Republic of Moldova (2004) and The Prize of the Presidents of the Ukrainian, Belarus and
MoldovanAcademiesofSciences(2003),amongmanyothers.
ISBN: 978-1-60650-312-6
9 781606 503126
90000
www.momentumpress.net
C
H
E
M
I
C
A
L

S
E
N
S
O
R
S
K
O
R
O
t
c
E
N
K
O
v
CHEMICAL SENSORS VoLuME 2: ConduCtoMEtrIC-typE SEnSorS
Edited by Ghenadii Korotcenkov, ph.d., dr. Sci.
AvolumeintheSensors Technology Series Edited by Joe Watson
Published by Momentum Press

v
O
l
u
m
E

2
C
o
n
d
u
c
t
o
m
e
t
r
i
c
-
t
y
p
e

S
e
n
s
o
r
s

You might also like