You are on page 1of 42

13.

02 Review of Physical Principles of Sensing and Types


of Sensing Materials
E Spain and A Venkatanarayanan, Dublin City University, Dublin, Ireland
Ó 2014 Elsevier Ltd. All rights reserved.

13.02.1 Introduction 5
13.02.1.1 Desired Properties of Sensing Materials 6
13.02.2 Metals as Sensing Material 7
13.02.2.1 Metallic Bonding 8
13.02.3 Optical Fiber Sensing Materials 9
13.02.3.1 Composition of Fiber Optics 12
13.02.4 Semiconductors 12
13.02.4.1 Classification of Semiconductors 13
13.02.4.2 Doping in Semiconductors 13
13.02.4.3 Semiconductors as Sensing Materials 15
13.02.4.4 Conducting Polymers 16
13.02.5 Dielectric Materials 16
13.02.5.1 Polarization 19
13.02.5.2 Frequency Response 19
13.02.5.3 Piezoelectric, Pyroelectric, and Ferroelectric Dielectric Materials 20
13.02.5.3.1 Piezoelectric Sensors 20
13.02.5.3.2 Pyroelectric Sensors 21
13.02.5.3.3 Ferroelectric Sensors 22
13.02.6 Magnetic and Superconducting Materials 23
13.02.6.1 Diamagnetism 24
13.02.6.2 Ferromagnetism 25
13.02.6.3 Paramagnetism 25
13.02.6.4 Superconductors 27
13.02.6.4.1 Type I Superconductors 28
13.02.6.4.2 Type II Superconductors 28
13.02.6.4.3 Superconductors as Sensing Materials 29
13.02.7 Solid Electrolytes 30
13.02.7.1 Point Defect Type 30
13.02.7.2 Cationic Conductors 33
13.02.7.3 Anionic Conductors 33
13.02.7.4 Organic Ionomers 33
13.02.7.5 Solid Electrolytes in Sensors 33
13.02.8 Biological Sensing Materials 34
13.02.8.1 Enzymes 34
13.02.8.2 Nucleic Acids 36
13.02.8.3 Antibodies 38
13.02.8.3.1 Antibody–Antigen Interaction 39
13.02.9 Conclusion 40
References 40

13.02.1 Introduction

Sensing materials as a subject of interest is very broad. Virtually, all well-known materials could be used for the elaboration of
chemical sensors. The world is separated into natural and man-made objects. Natural sensors are found in living organisms, and
they typically respond with electrochemical signals, i.e., their physical nature is based on ion transport. In man-made appliances,
information is transmitted and processed in electrical form, however, this time through the transport of electrons. Therefore, in
this chapter, we have tried to include the widest possible number of sensing materials and to evaluate their general background
information, as well as their real advantages and potential applications in today’s sensors. Materials in this chapter are considered in
a series of sections by category (metals, optical fibers, etc.).
First, a brief introduction is given below on the description of chemical sensors and the desired properties for sensing
materials. Sensors that are employed in artificial systems need to communicate with the devices at which they are interfaced. This

Comprehensive Materials Processing, Volume 13 http://dx.doi.org/10.1016/B978-0-08-096532-1.01302-9 5


6 Review of Physical Principles of Sensing and Types of Sensing Materials

language is electrical in its nature, and a man-made sensor must be competent to respond with signals where the information is
carried by displacement of electrons rather than ions. Hence, a sensor should be capable of connecting to an electronic system
through electrical wires rather than through an electrochemical solution or a nerve fiber. Therefore, a sensor is a transducer or
converter that responds to an input quantity (stimulus) by generating a functionally related output usually in the form of an
electrical or optical signal (1–3). A list of definitions relating to today’s sensors are given below.
Sensor element: The fundamental transduction device (e.g., a material) that converts one form of energy into another. Some
sensors may integrate more than one sensor element (e.g., a compound sensor).
Sensor: A sensor element together with its physical packaging and external connections (e.g., electrical or optical).
Sensor system: A sensor and its assortment of signal processing hardware (analog or digital) with the processing either in or on the
same package or separate from the sensor itself.

13.02.1.1 Desired Properties of Sensing Materials


The number of physical and chemical parameters that characterize sensor properties is large, and some of these parameters are
difficult to control. The problem of optimization is largely empirical and remains a kind of art. There is a search for optimal sensing
materials and designing of adequate theoretical models that may promote optimization of sensors. The forms of energy involved in
the sensing operation can be conveniently characterized into categories of principles, which yielded a matrix of effects (2,4). Figure 1
highlights the structural, electrical, and mechanical properties of materials used in sensing.
The energy forms or signal domains usually encountered with examples of typical properties that are measured using those
energy forms are shown in Table 1. The sensors for four of the five nonelectrical energy forms are shown in Table 2 with the
corresponding human senses.
In order to describe and characterize the performance of a sensor, a large and specific vocabulary is required. The four key
properties of any sensor (2) are described below.
1. Sensitivity
Sensitivity is defined as the ratio of sensor output to the change in the value of the measured. For example, consider a thermistor,
which records change in resistance when temperature changes. An increase in resistance by 10 ohm when the temperature increases
by 5 will lead to a sensitivity of 2 ohm per degree.
S ¼ 10=5 ¼ 2
It is to be noted that the units of sensitivity will depend on the property the sensor is measuring.

Figure 1 Structural, electrical, and mechanical properties of materials used in sensing.

Table 1 Categories of energy involved in sensing processes

Category Sensing materials properties used in sensing

Acoustic Wave amplitude, phase, polarization, spectrum, wave velocity


Mechanical Position (linear, angular), stress, pressure, strain, mass, density, moment, torque, speed of flow, rate of mass transport, shape,
roughness, length, area, volume, all time derivatives such as linear/angular velocity, linear/angular acceleration force, force, torque,
pressure, acoustic wavelength and acoustic intensity orientation, stiffness, compliance, viscosity
Biological Biomass (types, concentration, states)
Chemical Components (identities, concentration, states), reaction rates, redox potential, biological properties, pH, oxidation/reduction potential
Electric Charge, current, potential, voltage, electric field (amplitude, phase, polarization, spectrum), conductivity, permittivity, resistance,
capacitance, frequency, inductance, dielectric constant, polarization, dipole moment
Viscosity Crystallinity, structural integrity
Radiation Type, energy, intensity, phase, wavelength, amplitude, transmittance, polarization, reflectance, refractive index
Magnetic Magnetic field (intensity, amplitude, phase, polarization, spectrum), magnetic flux density, permeability, moment
Thermal Temperature, flux, specific heat, thermal conductivity, entropy, heat flow, state of matter
Optical Wave amplitude, phase, polarization, spectrum, wave velocity, refractive index, emissivity, reflectivity, absorption
Reproduced from Lion, K. S. Transducers: Problems and Prospects. IEEE Trans. Ind. Electron. Control Instrum. 1969, IECI-16, 2–5; Moseley, P. T.; Crocker, A. J. Sensor Materials,
1996; p 227; Fraden, J., Ed. Handbook of Modern Sensors: Physics, Designs, and Applications; Springer: San Diego, CA, 2010; Vol. 4, p 663; White, R. A. Sensor Classification
Scheme. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 1987, 34, 124–126.
Review of Physical Principles of Sensing and Types of Sensing Materials 7

Table 2 Forms of energy sensed by artificial sensors and human senses

Energy form Human senses

Chemical Smell, taste


Magnetic
Mechanical Hearing, touch
Radiant Sight, touch
Thermal Touch
Reproduced from Moseley, P. T.; Crocker, A. J. Sensor Materials, 1996; p 227.

2. Resolution
Resolution is defined as the smallest physical change the sensor can detect. This is usually limited by the accuracy of the
measurement device. In the particular case of measurement from zero, the resolution is the threshold or detection limit. To calculate
the resolution of the sensor, divide the smallest change in physical property by the sensitivity of the system.
3. Response time
Sensors do not change the output state immediately when an input parameter change occurs. Rather it will change to the new
state over a period of time called the response time. The response time is simply the time it takes for the sensor to respond to the
change in the physical property it is sensing.
4. Dynamic range
The dynamic range is defined as the total range of the sensor from minimum to maximum. It is the ratio between the largest and
the smallest possible values of a changeable quantity. It is usually measured as a ratio or as a base-10 or base-2 logarithmic value.
The overall key specifications of any sensor are given below in Table 3.
The field of sensing materials is extremely broad; its future development in sensors will involve the interaction on practically
every scientific and technical field. Modern sensors have incorporated the stated sensing materials for a number of applications such
as automotive, civil engineering, construction, domestic appliances, security, energy, power, information, telecommunication,
health, medicine, marine, manufacturing, distribution, commerce, finance, environment, meteorology, recreation, toys, military,
space, scientific measurement, and transportation. The group of sensing materials used in these sensors has been selected. It is the
purpose of this chapter to provide a guide to the main categories of sensing materials as well as their direct property in the sensor
function.

13.02.2 Metals as Sensing Material

Metals are inorganic minerals that form naturally below the surface of the earth. From building space shuttles to preserving food in
tin cans, man has been using metals in some form or another for more than 5000 years. Almost everything we make or use today
contains some amount of metals. If the object is nonmetal there is good chance it was built using some tools made of metals.
Currently there are 86 known metals. With respect to sensors, metals can be classified into two groups: ferrous metals and
nonferrous metals. Ferrous metals like steel are often used in conjunction with magnetic sensors to measure motion, distance,
magnetic field strength, etc. (5). They are also used as magnetic shields (5). Nonferrous metals on the other hand are permeable to
magnetic fields and are used whenever the magnetic field is of no concern to the measurement (2). However, they offer a wide
variety of mechanical and electrical properties.

Table 3 Specifications of a sensor

Static Dynamic
Accuracy Dynamic error response
Distortion Hysteresis
Hysteresis Instability and drift
Minimum detectable signal Noise
Nonlinearity Operating range
Selectivity/specificity Repeatability
Selectivity Step response
Threshold Stability (short and long terms)
Overload characteristics Speed of response
Operating life Cost, size, and weight
Resolution Environment conditions
Linearity Dead band
Reproduced from Lion, K. S. Transducers: Problems and Prospects. IEEE Trans. Ind. Electron. Control Instrum.
1969, IECI-16, 2–5; Moseley, P. T.; Crocker, A. J. Sensor Materials; 1996; p 227; Fraden, J., Ed. Handbook of
Modern Sensors: Physics, Designs, and Applications; Springer: San Diego, CA, 2010; Vol. 4, p 663.
8 Review of Physical Principles of Sensing and Types of Sensing Materials

The key feature that distinguishes a metal from a nonmetal is metallic bonding. Each metal atom is surrounded by similar atoms,
each with a small number of electrons in its outer shell. The electron clouds of individual atoms overlap, and the outer electrons of
each become universally shared. The remaining part of each atom is localized as an ion on its lattice site while the free electrons form
an ‘electron gas’ that occupies the totality of space between the ions and binds them together. The flow of these free electrons under
the applied electric field gives rise to the high electrical conductivity exhibited by metals. The electron gas and the spherical ions pull
each other together into a compact mass with nondirectional bonding. Metals are opaque and lustrous because the free electrons are
able to absorb most of the light energy that falls on them and radiate it back (2,6).
Metals are crystalline in nature; in order to achieve the state of lowest energy, the atoms are arranged in a regular manner. An
ideal single crystal is defined as a body of atoms or ions stacked to form a three-dimensional (3D) net that is determined by
translational vectors a1, a2, and a3. From the definition, a single crystal is free of lattice defects (5). A characteristic feature of the
translational symmetry of a lattice is that an arrangement of atoms about a given lattice point, determined by the vector R1, is
identical with that observed from any other point that can be reached by the simple transformation

R0 ¼ R1 þ n1 a1 þ n2 a2 þ n3 a3 ;
where a1, a2, and a3 are any noncoplanar vectors and ni ¼ 0, 1, 2, 3, . , i ¼ 1, 2, 3 (2).

13.02.2.1 Metallic Bonding


A 3D lattice produced by all the points whose locations are defined by the vectors R0 is called a Bravais lattice (7). There are total of
14 3D Bravais lattices discussed in solid state physics (7,8). The majority of single crystals of metals are characterized by one of the
following close packed structures, as shown in Figure 2:
a. A1 structure of face-centered cubic (FCC) lattice
b. A2 structure of body-centered cubic (BCC) lattice
c. A3 or hexagonal close packed (HCP) structure
The metal structures that do not fall under these three categories are classified as complex. These close packed crystal structures
enable a metal to withstand tensile stresses well but offer less resistance to shearing forces. Therefore, within certain temperature
ranges the metals exhibit high ductility (7). Because of the free electrons that are universally shared among all the atoms, different
metallic crystals can be bonded well together so that the cohesive strength of grain boundaries in metals is very high. Thus one
finds it extremely difficult to break a cold metal along its grain boundary unless it has been contaminated by some impurity.
Similarly, it is often very easy to join two pieces of metal by simply bringing their clean surfaces together, as in welding or
soldering (8).
Within the simple close packed structure adopted by metals it is possible for the lattice sites to be occupied by atoms of more
than one type provided there is not too large a discrepancy between the metallic radii involved. This is how metal alloys are formed
over a wide range of composition. Also due to the close packed structure, the packaging density of the metal is high, making metallic
elements denser than other materials (6,9). Thus it is clear that the mechanical properties of the metals are largely due to their crystal
structure; these are in turn strongly dependent on the free electron bonding.
The thermal properties of metals also greatly depend on this free electron model. If the free electrons are regarded as particles that
are capable of passing through all parts of the solid, the specific heat of a metal based on the successful prediction of the specific heat
of gases is given as 3Nk, where N is the Avogadro number and k is the Boltzmann constant. Unfortunately, this is an inadequate
description of the specific heat of a solid since it fails to predict temperature dependence. It is only when a quantum theoretical
approach is adopted that the anomalous results are avoided and the measured specific heat can be matched by theory. The high
electronic conductivity of the metals is due to the free motions of significant concentration of electrons. The thermal conductivity (k)
of solids has contributions from both electrons (ke) and phonons (lattice vibrations (kL)), although in metals the former term is
dominant (eqn [1]).

k ¼ ke þ kL [1]

Figure 2 Close packed structures of single crystal of metals with (a) face-centered cubic lattice; (b) body-centered cubic lattice; and (c) hexagonal
close packed structure.
Review of Physical Principles of Sensing and Types of Sensing Materials 9

Hence, there is a close correspondence between the electrical and thermal conductivities of metals (3). At normal temperatures,
the ratio ke/s for most pure metals is close to 7  106 WU K1. This ratio, eqn [2], is known as the Wiedmann-Franz ratio and is
proportional to absolute temperature (T).
ke =s ¼ LT [2]
The proportionality constant, L, is known as Lorenz number. The thermal conductivity of metals, like electrical conductivity,
decreases with the increase in temperature.
The several characteristic properties of metal briefly outlined above find distinct applications in the functions of sensors, ranging
from simple electrical connection as electrodes through to specific functions dependent on high work function or catalytic activity.
While selecting a metal for sensor application one must not only consider its physical properties but also its ease of mechanical
processing. For example, while copper has excellent thermal and electrical properties, it is very difficult to machine, therefore
aluminum is considered instead.
Aluminum has a high strength-to-weight ratio and possesses its own anticorrosion mechanism. When exposed to air it does not
go to oxide progressively like iron. Instead it forms a microscopic oxide coating that seals the bare metal from the environment.
Hundreds of aluminum alloys have been developed, as they are easy to process. As aluminum is a superb reflector of light over
nearly the entire spectrum from UV to radio waves, it is widely used in mirrors and waveguides in the form of thin films. In the mid-
and far-infrared regions, the only superior to aluminum reflector is gold. Beryllium, on the other hand, shows low density, high
specific heat, excellent dimensional stability, and transparency to X-rays. However, it is very expensive.

13.02.3 Optical Fiber Sensing Materials

Fiber optic sensors (FOSs) have been under development since the 1970s. The first experiments reported their use in the form of
low-loss, low transparency optical fibers in early endoscopes for medical and industrial applications contradictory to telecom-
munication, as had been their primary motivation for development (10). Since then various prototype FOSs that respond to
numerous physical and chemical variables have been demonstrated in academic and corporate laboratories around the world
(11,12). These pioneering works quickly led to the growth of fiber optic technology, particularly for sensing and measurement
purposes.
Propagation of light along an optical fiber can be described by two theories. The first theory considers light as a simple ray and is
known as the ‘ray theory’or geometrical optics approach. The ray theory approximates the light acceptance and guiding properties of
optical fibers (11).
The second theory describes light as an electromagnetic wave. This theory is called the ‘mode theory’ or wave representation
approach. The mode theory describes the behavior of light within an optical fiber. The mode theory is useful in describing the
optical fiber properties of absorption, attenuation, and dispersion.
When light travels from one medium (air) to another (water), especially at an angle, it undergoes a change in its phase velocity
(11). There is a change in direction of the wave due to change in its speed, referred to as refraction, Figure 3.
Refraction is described by Snell’s law, eqn [1], which states that for a given pair of media and a wave with a single frequency, the
ratio of the sines of the angle of incidence q1 and the angle of refraction q2 is equivalent to the ratio of phase velocities (v1/v2) in the
two media or the opposite ratio of indices of refraction (n2/n1) (3,11,12).
sin q1 v1 n2
¼ ¼ [3]
sin q2 v2 n1
Therefore, when light travels from a medium of higher refractive index (RI) to a medium of lower RI, the incident wave is
partially reflected and partially refracted, as shown by the green line in Figure 4. Total internal reflection is a special case in which
light strikes the medium boundary at an angle larger than a particular critical angle (qc) with respect to the normal to the surface. If
the refractive index is lower on the other side of the boundary (n2 < n1), then no light can pass through and all light is reflected, as
shown by the blue line in Figure 4.

Figure 3 Schematic showing light ray (red line) undergoing refraction.


10 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 4 Schematic showing refraction path of light ray when it enters a medium of low RI at an angle.

Optic fibers are based entirely on the principle of total internal refraction (TIR). Optic fibers (or fiber optics) are long thin strands
of very pure glass about the diameter of a human hair. They are arranged in bundles called optical cables, and they transmit light
signals over long distances (13). An optic fiber consists of three main parts:
1. Core: a high RI thin glass center of the fiber where light travels.
2. Cladding: a lower RI optical material surrounding the core that reflects the light back into the core.
3. Buffer coating: plastic coating that protects the fiber from damage and moisture.
Generally, hundreds and thousands of these optical fibers are arranged in bundles in optical cables that are protected by an outer
covering called a jacket, as shown in Figure 5.
The RI of the core is always higher than the cladding. As a result, the light beam is refracted toward the inside of the fiber by TIR,
as shown in Figure 6. Successive internal refractions results in propagation of light through the fiber (14).

Figure 5 Schematic diagram of a single-mode optical fiber.

Figure 6 Propagation of light in an optic fiber.


Review of Physical Principles of Sensing and Types of Sensing Materials 11

Figure 7 Graded index and step index optic fibers.

When the RI between the core and the cladding changes abruptly and the RI is uniform within the core, it is called a step index
fiber, as shown in Figure 7. When the RI continuously varies between the core and the cladding or is graded, while the RI of the
cladding remains uniform, it is called a graded index fiber, Figure 7.
When light enters a fiber, it is important to determine the maximum angle of entry that will result in total internal reflection.
If we take that minimum angle of an internal reflection q0 ¼ q3, then the maximum angle q2 can be found using Snell’s law (2):
 
q2max ¼ arc sin root n2  n21 n [4]

For air n z 1, therefore,


sin qin max ¼ n1 sin q2 max [5]
Combining [2] and [3], we obtain the largest angle with the normal to the fiber end for which the TIR will occur in the core

qin max ¼ arc sin (root (n2  n21)).

Propagating light consists of two components:


1. The guided field in the core;
2. The exponentially decaying evanescent field in the cladding.
In a uniform diameter fiber, the evanescent field decays to almost zero within the cladding. Light rays entering the fibers at angles
greater than qin max will pass through to the jacket and will be lost. This light does not interact with the fibers surrounding and is
undesirable for data transmission. However, if the cladding of the fiber is removed or reduced, Figure 8, the evanescent field
interacts with the surrounding, which leads to the use of optical fibers in sensing (2,3,11,12).

Figure 8 Schematic of (a) de-cladded optic fiber; (b) tapered optic fiber; (c) U-shaped optic fiber; (d) tapered tip. Reproduced from Leung, A.; Shankar,
P. M.; Mutharasan, R. A Review of Fiber-Optic Biosensors. Sens. Actuators, B 2007, 125, 688–703.
12 Review of Physical Principles of Sensing and Types of Sensing Materials

As stated, optic fiber sensors require light to interact with the fibers surrounding. The distance to which the evanescent wave
extends beyond the core-cladding interface is described as the penetration depth, which is defined as the distance where the
evanescent field decreases to 1/e of its value at the core-cladding interface (15).
The construction of optical fibers involves the following steps:
1. Construction of a ‘platform’ that is on the scale of centimeters in size.
2. Heating the platform to draw it down to a much smaller diameter (often nearly as small as a human hair), shrinking the platform
cross-section but usually maintaining the same features.
In this way, kilometers of fibers can be produced from a single platform (2). One of the greatest advantages of optic fibers is
that they can be formed into a variety of different geometrical shapes. This makes them very useful in the design of miniature
optical sensors that are responsive to stimuli like pressure temperature, chemical concentration, etc. (3). Additional advantages
include immunity to electromagnetic interference, light weight, remote sensing, and the ability to be embedded into textile
structure.
Optical fibers can be separated into two classes based on their application in sensor devices: one, where the optical fiber itself acts
as the sensing medium (intrinsic sensors); and two, where the fiber acts as a light guide (extrinsic sensors) to connect the light to
a sensor cell and return the light from the sensor cell to an analyzing system that is usually spectroscopic or electronic in nature (2).
From the simplest fiber optic lamps to detecting single molecule interactions in a total internal reflection fluorescence (TIRF)
microscope, we find that extrinsic optic sensors play a major role as light guides. They are also involved in a unique role in enabling
the source and the analyzing system to be separated from a chemically or electrically hazardous environment, which is very useful
in remote sensing applications. However, this chapter highlights intrinsic FOSs (3,11).
An optic fiber is a rotationally symmetric waveguide, and light is an electromagnetic wave. Hence by definition stationary
modes will occur in a resonant cavity with nodes at the walls and antinodes of maximum amplitude at the center of the cavity
diminishing toward the exterior. The number of nodes is the same as the number of antinodes; there can be many hundred
in an optical fiber. If the core is made extremely small, only one more will be able to propagate and the fiber is called
monomode (2,11).
In a straight optical fiber, power decreases exponentially with distance. This attenuation can be represented as a function of
fiber length and is given in decibels by eqn [6].
1 PðzÞ
a ¼ 1 10$log dB=km [6]
z Pð0Þ

where a is called the attenuation coefficient of the optical fiber. Attenuation of an optical fiber mainly depends on the fiber core
diameter, and it increases as the core diameter decreases. Spectral width and numerical aperture (NA) of the light source employed
also affect the value of attenuation. An increase in either one of them increases the attenuation (16).
Monomode fibers have a great advantage in telecommunication applications as they have lower attenuation, greater bandwidth,
and lower transmission bit error rates (10,15).

13.02.3.1 Composition of Fiber Optics


Fiber optics are generally made of glass or polymers like polymethyl methacrylate (3). Silicate glass has been used to manufacture
optical fibers, but due to its high absorption in the near infrared region, special provisions to minimize the amount of absorbing
ions, i.e., transition metals like Cu2þ, Fe2þ, Ni2þ, and lead are made. Silica fibers exhibit low intrinsic losses; however, they have
high softening temperatures (1400  C) compared to conventional glass (700  C) (2). Various techniques have been developed to
produce silica-based optical fibers (10). In order to modulate the refractive index of the core in comparison to the cladding, the core
is generally doped with germanium (GeCl4) to increase the RI and phosphorous oxychloride (POCl3), which decreases the soft-
ening point (10).

13.02.4 Semiconductors

Semiconductor is composed of two words: semi (not completely) and conductor (something that can conduct electricity).
In short, a semiconductor is a material that has electrical conductivity due to flowing electrons. It has distinctive phys-
ical properties somewhere between a conductor like aluminum and an insulator like glass, resulting in conductivity
roughly in the range of 103–108 S cm1 (17). The resistivity scale of metals, semiconductors, and insulators is shown in
Figure 9.
Semiconductors are mainly classified into two categories: intrinsic and extrinsic. An intrinsic semiconductor material is chemically
very pure; however, it possesses poor conductivity. It has equal numbers of negative carriers (electrons) and positive carriers (holes)
(19). At very low temperatures, intrinsic semiconductors act like insulators. At ambient temperature, electrical conductance is much
more improved than an insulator, but less readily than a conductor. At higher temperatures or under light, intrinsic semiconductors
can readily become conductive (17). An extrinsic semiconductor is an enhanced intrinsic semiconductor with a small quantity of
impurities added by a process known as doping, which modifies the electrical properties of the semiconductor and enhances its
conductivity (20,21).
Review of Physical Principles of Sensing and Types of Sensing Materials 13

Figure 9 Resistivity scale of materials found in nature. Reproduced from Tyagi, M. Introduction to Semiconductor Materials and Devices; John Wiley &
Sons: Canada, 1991.

13.02.4.1 Classification of Semiconductors


Semiconductors are classified as elements or compounds. Elemental semiconductors such as silicon (Si) and germanium (Ge)
found in group IV of the Periodic Table serves as the chief component for commercially produced semiconductors (22). Ge was used
for the development of the 1st transistor, by Barbeen, Brattain, and Shockley in 1947. In the early 1960s, Ge was replaced by Si as it
can be readily oxidized to form silicon dioxide (SiO2) insulator. SiO2 can be used as an excellent barrier layer for the selective
diffusion steps needed for integrated circuits. Si has a wider band gap than Ge, and it is inexpensive and abundant in nature.
Therefore, silicon is the material of choice for semiconducting sensing materials (17). Si and Ge both have a crystalline structure
referred to as a diamond lattice. Each atom has its four nearest neighbors at the corners of a regular tetrahedron, with the atom
itself at the center (Figure 10).
Besides pure element semiconductors, many alloys and compounds are semiconductors. Compound semiconductor materials
are formed from special combinations of group III and group V elements. Compound semiconductors include a range of other
materials, such as InSb, InAs, GaP, GaSb, GaAs, SiC, and GaN (22). Compound semiconductors have the advantage of wide range of
energy gaps and nobilities, so sensing materials are available with properties that meet individual requirements. Therefore, some of
these semiconductors are called wide band gap semiconductors (19).

13.02.4.2 Doping in Semiconductors


As already stated, a pure semiconductor is often called an ‘intrinsic’ semiconductor. The ease at which the conductivity can be
modified is one of the main promising properties of semiconductors as sensing materials that makes them most useful for
constructing electronic devices (19).
In a process called ‘doping’ the electronic properties and the conductivity of a semiconductor can be changed in a controlled
manner by introducing impurities called ‘dopants’ into their crystal lattice. By adding impurity to pure semiconductors, the electrical
conductivity may be varied not only by the number of impurity atoms, but also by the type of impurity atom; the changes may be
on the order of thousand fold and million fold. Examples include silicon, the basic material used in the integrated circuits (17,22),
and germanium, the semiconductor used for the first transistors (17).
The materials chosen as suitable dopants depend on the atomic properties of both the dopant and the material to be doped. The
difference in the number of valence electrons between the doping material, or dopant, and host semiconductor results in two groups
of semiconductors: the negative charge conductor (n-type) and the positive charge conductor (p-type). The dopant is known as an

Figure 10 Tetrahedron configuration of Si crystal with the four nearest neighbors. Reproduced from Tyagi, M. Introduction to Semiconductor Materials
and Devices; John Wiley & Sons: Canada, 1991.
14 Review of Physical Principles of Sensing and Types of Sensing Materials

acceptor atom if it ‘accepts’ an electron from the semiconductor atom. It is known as a donor atom if it ‘donates’ a weakly bound
valence electron to the material, creating excess negative charge carriers (i.e., becomes incorporated into the crystal lattice) (18,23).
These weakly bound electrons can freely move about in the crystal lattice and can aid conduction in the presence of an electric field.
The donor atoms introduce some states under but very close to the conduction band edge. The electrons at these states can be readily
excited to the conduction band, becoming free electrons, at room temperature. On the other hand, an activated acceptor produces
a hole. For instance, a silicon atom has four valence electrons, two of which are required to form a covalent bond. In n-type silicon,
donor atoms such as phosphorus (five valence electrons) replace some silicon, providing extra negative electrons. In p-type silicon,
acceptor atoms with three valence electrons such as aluminum (Al) lead to the removal of an electron, or hole, that performs like
a positive electron. These extra electrons, or holes, can conduct electricity (18). The ‘n-type’ and ‘p-type’ labels indicate which charge
carrier acts as the material’s majority carrier. The opposite carrier is called the minority carrier, which exists due to thermal excitation
at a much lower concentration compared to the majority carrier (18).
The concentration of dopant introduced to an intrinsic semiconductor indirectly affects many of its electrical properties. The
vital factor that doping directly influences is the material’s carrier concentration. In an intrinsic semiconductor under thermal
equilibrium (18) the concentration of electrons and holes is equivalent. That is,
n ¼ p ¼ ni [7]
If we have a nonintrinsic semiconductor in thermal equilibrium, the relation becomes
n0 þ p0 ¼ n2i [8]

where n0 is the concentration of conducting electrons, p0 is the electron hole concentration, and ni is the material’s intrinsic carrier
concentration. Intrinsic carrier concentration varies between materials and is dependent on temperature. Silicon’s ni, for example, is
roughly 1.08  1010 cm3 at 300 K (24).
The conductivity of a semiconductor increases with temperature, light, or the addition of impurities, as they increase the number
of conductive valence electrons of the semiconductor. Valence or outer electrons are the carriers of the electrical current. In solid-
state physics, the electronic band structure of a solid describes electrons with energies only within certain bands (i.e., ranges of levels
of energy) called energy bands. Energetically, these bands are located between the energy of the ground state (the state in which
electrons are tightly bound to the atomic nuclei of the material) and the free electron energy (the energy required for an electron to
break completely from the material). The energy bands each relate to a large number of discrete quantum states of the electrons, and
the majority of the states with low energy (closer to the nucleus) are full, up to a particular band called the valence band (22).
Figure 11 illustrates a simplified picture of the bands in a solid that allows the three major types of materials to be identified:
insulators, semiconductors, and metals.
Metals contain a band that is partly empty and partly filled regardless of temperature, therefore possessing very high
conductivity. The lower, almost fully occupied band in an insulator or semiconductor is called the valence band. The upper,
almost unoccupied band is called the conduction band because only when electrons are excited to the conduction band can
current flow in these materials (22,24). In an intrinsic semiconductor such as silicon, the valence electrons of an atom are paired
and shared with other atoms, producing covalent bonds that seize the crystal together. Under such conditions, these valence
electrons are unable to move around as electrical current. Temperature or light excites the valence electrons out of these bonds,
allowing them to conduct current. The vacant positions left behind (holes) by the freed electrons can move around as well,
contributing to the flow of electricity (18,22–24). The holes themselves do not actually move, but a neighboring electron can
move to fill the hole, leaving a hole at the place it has just come from; in this way the holes appear to move, and the holes perform
as if they were actual positively charged particles. Electrons move faster within the holes. For silicon, mobility (m) is 2.5 for
electrons and 2.7 for holes. For germanium, m is equal to 1.66 for electrons and 2.33 for holes. Covalent bonds between adjacent

Figure 11 A two dimensional energy band diagram of insulators, semiconductors, and metals.
Review of Physical Principles of Sensing and Types of Sensing Materials 15

atoms in the solid is 10 times stronger than the binding of the single electron to the atom, so releasing the electron does not
involve destruction of the crystal structure (18,23).
Light can also give this energy boost and create an electron-hole pair (a free electron and a free hole); this phenomenon is called
absorption. Photoconductivity (25) is the increase of current in a semiconductor due to the absorption of photons. Light has a dual
nature: it acts as a wave and as a particle. The particle associated with light is called a photon, which consists of different energies.
When light illuminates a semiconductor, photons are absorbed by the material and the electrons from the valence band have
enough energy to jump to the conduction band. The conductivity increases due to the higher number of electrons in the
conduction band. The conversion of electrical energy into light is called electroluminescence. The electrons in the conduction
band/excited state eventually fall back into the valence band in a lower energy state, releasing the extra energy that they have. This
energy is emitted as a photon, whereas photons emitted by electroluminescence emerge in random directions. This type of light is
called incoherent light (25).
The ease with which electrons in the semiconductor can be excited from the valence band to the conduction band depends on
the band gap between the bands. The size of this energy band gap is roughly 4 eV between semiconductors and insulators. The
energy gap (Eg) is given by Eg ¼ 1:21  3:60  104 T for silicon and Eg ¼ 0:785  2:23  104 T for germanium, where T is
temperature. At room temperature, the Eg for silicon is equal to 0.72 eV and for germanium, it is equal to 1.1 eV (17,22,24).
The differentiation between insulators and semiconductors is that the band gap between the valence band and conduction band
is larger in an insulator, so that fewer electrons originate there and the electrical conductivity is lower. Because one of the main
mechanisms for electrons to be excited to the conduction band is due to thermal energy, the conductivity of semiconductors is
strongly dependent on the temperature of the material. The relation between intrinsic concentration and temperature (17)
is given by
 
Eg
ni 2 ¼ A0 T 2 exp  [9]
KT
where A0 is constant, KT is the Boltzmann constant, and Eg is the energy gap at 0 K.
The equation shows that the intrinsic concentration increases with temperature, which in turn increases the conductivity. The
band gap characterization of semiconductors is one of the most useful aspects, as it strongly influences the electrical and optical
properties of the material. Electrons can transfer from one band to the other by means of carrier generation and recombination
processes. The band gap and defect states created in the band gap by doping can be used to create sophisticated semiconductor
devices (17). For example, when a p-type semiconductor region is placed adjacent to an n-type region, they form a diode; this region
of contact is called a p–n junction. A diode is a two-terminal device that conducts current in one direction only (22). By combining
junctions like these, transistors and other semiconductor devices can be readily made, the electrical behavior of which can be
controlled by the electrical stimuli. Combinations of many transistors and other active components on a single chip of silicon have
resulted in the integrated circuit. This is a complex electronic appliance designed to complete certain functions depending on the
controlling signals (22).

13.02.4.3 Semiconductors as Sensing Materials


Semiconductor materials are the foundation for most modern electronics, which includes radios, computers and telephones (25).
These products consist of the different forms of transistors, solar cells, diodes, including the light-emitting diode (18,23) silicon
controlled rectifiers, and digital and analog integrated circuits (22,26,27). Correspondingly, semiconductor solar photovoltaic
panels directly convert light energy into electrical energy and have become a key focus for many renewable green energy sources
companies.
In its pure state, semiconductor material is an excellent insulator. The heavier the doping, the greater the conductivity/the lower
the resistance. New semiconductor materials play a decisive role in their applicability to biosensors. The chemical synthesis
(growth) of the material, the postmaterial treatment such as doping or ion implantation, and the final chemical surface treatment
are the three most important properties of semiconductors (17). The development and optimization of these semiconductor
technology procedures has allowed for the growth or synthesis of materials with very well controlled and unique physical and
chemical characteristics, down to the atomic level. Organic semiconductors offer unique characteristics such as tunability (28) of
electronic properties via chemical synthesis, compatibility with mechanically flexible substrates, low-cost manufacturing, and facile
integration with chemical and biological functionalities. Organic semiconductors denote semiconducting organic compounds
exhibiting electronic conductivities between about 109 and 103 S cm1 (26). These characteristics have prompted the application
of organic semiconductors and their devices in physical, chemical, and biological sensors. Compared to the organic (such as
phenenthrene and polybenzimidole) and elemental (such as Si, Ge, GaAs, and GaP) semiconductors, semiconductor oxides have
been more successfully employed as sensing materials for the detection of different gases (29), such as CO, CO2, H2, alcohol, H2O,
NH3, O2, and NOx (30–34).
Both n-type (such as SnO2, TiO2, and ZnO) and p-type semiconductor oxides are capable of being used as gas sensor materials
(35). Recently, organic semiconductors have been fabricated with large sensing surface areas and disposable electronics on thin and
flexible substrates (26), creating considerable interest in the fields of electronic and photonic devices (25). They can operate at low
temperature and can also be chemically modified to detect a wide range of gases and vapors at a low cost (35). Their low cost
production is very simple, using recent high vacuum evaporation and spin coating technology.
16 Review of Physical Principles of Sensing and Types of Sensing Materials

Semiconductors are useful as sensing materials as they offer the following advantages:
1. Easier deposition of thin films (27);
2. Higher degree of sensitivity to external agents;
3. Work function or contact potential (36–38);
4. Low cost (39);
5. Greater electronic tenability (28);
6. Higher absorption coefficient (26,40–42);
7. Greater flexibility (26,35);
8. Ability to convert into a direct current (43);
9. Chemical reactions with interfacial electron transfer (44–47);
10. Surface recombination–photoconductivity/photoemission (25);
11. Change in contact potential with light (48–51);
12. Surface conductance–channel effect (52–55);
13. Change in surface conductance with electrostatic field–field effect (35,56);
14. Noise (35).
The properties of these sensing materials can be used as the basis for the development of chemical sensors and biosensors.
Monitoring surface current, potential, or impedance characteristics can be directly related to the chemical stimuli interrogating
the semiconductor sensing surface. The temperature of semiconductors is also another important property. Semiconductors
can function at ambient temperatures higher than 150  C without external cooling. This is useful in important applications,
especially in the automotive, aerospace, and energy production industries (57). Semiconductors are therefore ideal matrices as
sensor elements and transducers for the expansion of chemical sensor and biosensor systems. Semiconductor-based electronic
have successively been studied in area such as surface active potentiometric ion selective electrodes (ISEs), field effect tran-
sistors (FETs) (58), amperometric biosensors, surface acoustic wave (SAW) sensors, and film bulk acoustic resonators (FBARs)
(57).

13.02.4.4 Conducting Polymers


Polyacetylene is the simplest conjugated organic polymer and hence is used extensively to model mechanisms of electrical
conduction for such systems (59). Various models explain electrical properties of polymers in terms of delocalized charge carriers
that are free to propagate along the polymer backbone, namely soliton pairs, polarons, and bipolarons, depending on the polymer,
as illustrated in Figure 12 (60).
Since the discovery of polyacetylene there has been great interest into the research of conducting polymers; many new
polymers have been synthesized (61). Many conjugated polymers with varying degrees of functionality have been produced,
the most common being polyaniline (PANI) (62,63), polythiophene (PTh) (64), and polyrrole (PPy) (61). These structures can
be seen in Figure 13. These remarkable new materials exhibit the electrical and optical properties of metals or semiconductors
while retaining the attractive mechanical properties and processing advantages of organic polymers. The essential structural
characteristic of all conjugated polymers is their quasi-infinite p system extending over a large number of recurring monomer
units. This extended p-conjugated system of the conducting polymers have single and double bonds alternating along the
polymer chain. This feature results in materials with a directional conductivity, which is strongest along the axis of the chain
(65). The resulting higher values of electrical conductivity of the organic polymers have led to the collective name ‘synthetic
metals’ (66) (Figure 13).

13.02.5 Dielectric Materials

In the area of electromagnets, materials are classified into four broad categories:
1. Conductor
2. Semiconductor
3. Dielectric (insulators)
4. Magnetic

Figure 12 Schematic representation of the conduction mechanism in Polyacetylene. Reproduced from Kaiser, A. B.; Park, Y. W. Conduction
Mechanisms in Polyacetylene Nanofibres. Curr. Appl. Phys. 2002, 2, 33–37.
Review of Physical Principles of Sensing and Types of Sensing Materials 17

Figure 13 Structures of some conducting polymers commonly used in biosensors. Reproduced from Wallace, G. G.; Spinks, G. M.; Kane-Maguire, L. A. P.
Conductive Electroactive Polymers: Intelligent Materials Systems, 2nd ed.; 2002; p 224.

Dielectric materials present an attractive media for electric field interactions. For example, in the area of electronics packaging
application, the choice of dielectric materials used is based on the inherent physical properties of that material (67,68). These
properties influence the performance of the package in several key areas: signal speed, power consumption, thermal stability,
interaction with other materials, and wiring density (67,68). The ease with which dielectric materials can be modified to exhibit
certain electrical, mechanical, and thermal characteristics and their ease of processing have made them model materials for use in
electronics packaging (2).
A capacitor is used to store energy as well as charge. The charges in general are stored on two conductive plates: the anode
(positively charged plate) and the cathode (negatively charged plate) (67,68). The ratio of the charge magnitude on each plate to the
electric potential between the plates is known as capacitance. The energy collected in a capacitor is the energy required to move
the stored charge through the potential of the capacitor (67,68). Figure 14 shows an ‘empty capacitor’ with a vacuum between the
plates with capacitance C ¼ Q/V, where C is the standard unit of capacitance which are described by Farads (F), Q is the charge on
each capacitor plate, and V is the voltage between capacitor plates. Therefore, 1 F ¼ 1 CV1.
An electrical insulator is positioned between the anode and cathode, and it has the ability to store energy when an external
electric field is applied. This material is classified as ‘dielectric’ and it has the ability to develop a polarization (Figure 15) in response
to this applied electric field (67,68).
The two plates are positioned so that scarce movement occurs between them; when a DC voltage source is placed across a parallel
plate capacitor, they become charged and the force on the dielectric increases. Polarization of the dielectric material amplifies the
storage capacity of charge on the surfaces that are in contact with the capacitor plates. This partially counteracts the electric field by
neutralizing charges between the plates, which typically would give a decrease in field strength and thus can give a decrease in
voltage, which decreases proportionally with the charge, Figure 16 (68).

Figure 14 “Empty capacitor” with a vacuum between the plates.


18 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 15 Capacitor with a dielectric material between the plates.

Figure 16 Capacitor with a dielectric material between the plates.

The capacitance of the dielectric capacitor is reliant on the plate geometry as well as the nature of the dielectric. The capacitance
with the dielectric material is directly proportional to the dielectric constant and inversely proportional to the thickness of the
dielectric and its geometric area A (67–71). The term ‘dielectric constant’ (k) is often used in conjunction with relative permittivity
(εr), which is the absolute dielectric constant (ε) relative to the permittivity of free space (ε0). The permittivity indicates the ability of
a given material to store energy imparted by an electric field in addition to its efficiency in storing the energy. The two definitions of
the dielectric constant (k) are demonstrated by the illustration in Figure 17. In general, the more available polarization mechanisms
a material possesses, the larger its overall polarization in a given field will be, and hence its dielectric constant will be increased.
Therefore, k ¼ C0 /C and k ¼ ε/ε0 (67–71).

Figure 17 Diagram of the dielectric constant with vacuum and dielectric permittivity (the arrows represent the electric field).
Review of Physical Principles of Sensing and Types of Sensing Materials 19

Permittivity (ε0) can also be expressed as a function of the permittivity of free space. This quantity is known as relative
permittivity (eqn [1]) (2).
ε
k ¼ ε0r ¼ [10]
ε0
The relative complex permittivity (eqn [2]) εr can then be expressed as
εr ¼ ε0r  jε00r [11]

The two terms of the complex permittivity are represented by the dielectric constant, presenting the amount of energy elastically
stored from the external electric field in the material and the dielectric loss factor from the AC signal as the electric field energy is
dissipated by the material. Ideally, a dielectric material would prohibit the flow of electric charge; however, only a displacement of
charge via polarization occurs. Real dielectric materials always have some energy loss when an external electric field is applied (2).
In an ideal model, a capacitor is formed by placing a dielectric between parallel electrodes and applying an alternating electric
field. In this ideal condition, no power would be absorbed by the dielectric, and the current would guide the voltage by a phase of
90 (2). However, when dealing with authentic materials, a loss is constantly detected and the current phase settles d degrees behind
that of an ideal matter. The quantity of dissipated energy is defined as tan d (2). The dissipation factor (eqn [3]), also identified as the
dielectric loss factor (ε00r ) and dielectric constant (ε0r ), is the ratio of the energy dissipated to the energy accumulated in the
material (2), and has been shown to be a useful measure to remove permittivity’s dependence on density (2,67).
ε00r
tanðdÞ ¼ [12]
ε0r

A conversion of the dissipated energy into heat occurs due to conduction of electrons flowing through the material. The
magnitude of this energy loss is stimulated by both the intrinsic and extrinsic factors (72,73). Intrinsic losses are linked with the
crystal structure such as atomic masses, atomic charges, bond strength, and the phonon interaction with the AC electric field.
Extrinsic factors are associated by defects in the crystal such as microstructural defects, porosity, impurities, grain boundaries, oxygen
vacancies, and random crystallite orientation (67–71).

13.02.5.1 Polarization
Dielectrics are known as electrically insulating materials and consequently have high electrical resistivities, of the order of
1015–1018 Um (SM). Nonetheless, after an electric field has been applied, the field generates a shift in charge to form an electrical
dipole, resulting in an overall polarization (67–71) resulting in a polarized material. Polarization can occur in a number of ways,
i.e., the presence of a fixed charge, the movement of mobile charge, the rearrangement of dipole, or the displacement of fixed charge
(SM). There are numerous mechanisms for accumulating energy based on polarization of individual atoms and molecules by the
presence of an electric field for dielectric materials. Mechanisms consist of electronic polarization, resulting from the deformation of
the 3 electron cloud adjacent to individual atoms, and orientation polarization occurring when polar molecules are redirected by
an electric field (67–71,74). Other mechanisms include space charge polarization and atomic or ionic polarization. Atomic and
ionic polarization are the only mechanisms involved in the displacement of atoms or ions within a crystal structure while
demonstrating an extremely high dielectric constant (67–71). Research into dielectric materials has been concentrated on molecular
polar dipoles, which can have moderately large and distinctive dipole moments. The polarity of a dipole is measured in Debyes
(C m), representing the potency of a charge and the displacement distance (67–71).

13.02.5.2 Frequency Response


The complex permittivity of dielectrics is a frequency dependent property (2), whereby the relative permittivity or polarizability
may be expressed by
εs  εN
ε ¼ εN þ [13]
1 þ jus
Equation [13] is known as the Debye equation, where the complex permittivity of a material could be represented as a function
of the response dielectric constant at low frequencies (εs), high frequencies (εN ), a relaxation time constant (s), and the frequency
(u) (2). This equation can be further expressed into real (eqn [14]) and imaginary terms (eqn [15]), dielectric constants, and loss.
εs  εN
ε0r ¼ εN þ [14]
1 þ u2 s2

ðεs  εN Þus
ε00r ¼ εN þ [15]
1 þ u2 s2
The relative permittivity will vary with frequency (2):
At low frequency, a maximum permittivity will be obtained at the static limit (u ¼ 0 Hz) and thus tan d tends to be zero, as does
the energy loss.
20 Review of Physical Principles of Sensing and Types of Sensing Materials

At high frequencies, a monotonic decrease in dielectric constant is observed until the high frequency limit (u/N) is reached.
At the resonant frequency, maximum, located at the midpoint between static and high frequency dielectric constants, is achieved
when the relaxation frequency u and the relaxation constant s are equal.

13.02.5.3 Piezoelectric, Pyroelectric, and Ferroelectric Dielectric Materials


There are 32 crystal classes of solids, 21 of which have noncentrosymmetric structures as illustrated in Figure 18. Of these 21 classes,
20 correspond to piezoelectric materials, materials in which an applied mechanical strain produces an electric field or an electrical
field that produces a strain. The other 432 noncentrosymmetric crystals are nonpiezoelectric due to all the piezoelectric moduli that
cancel out as a result of the operation of the symmetry elements (2). Piezoelectric crystals can be subdivided into nonpolar and
permanent polarization. There are 10 classes of polarization, and they produce an electric field after an imposed temperature has
been applied. These types of materials are called ‘pyroelectric’ materials. A subset of these pyroelectric materials are the ferroelectric
materials, which can be both piezoelectric and pyroelectric. The direction of polarization can be reversed with these materials by
applying a high enough electric field (2).

13.02.5.3.1 Piezoelectric Sensors


The piezoelectric effect was discovered by Jacques and Pierre Curie in 1880 (75). The Curie brothers found that quartz changed its
dimensions when subjected to an electrical field and generated electrical charge dimension caused by an imposed mechanical force
(piezoelectric or generator effect) (76). Piezoelectricity refers to the generation of electricity or an electric polarization (i.e., charge)
exhibited by noncentrosymmetric crystals when subjected to mechanical stress; conversely, the generation of stress in such crystals is
in response to an applied voltage (76). Conversely, an applied electric field will produce a mechanical stress (electrostrictive or
motor effect) that is directly proportional. They transform energy from mechanical to electrical and vice versa (75). The latter
phenomenon is known as the converse effect and is utilized for sensing purposes (e.g., microphone, transducer) (77–80) and for
actuating applications (81–84). The former is called the direct effect and is used in sensing dynamic pressure changes, changes in
acceleration (from shock or vibration), and changes in force (85–87). Since that time, researchers have found piezoelectric prop-
erties in hundreds of ceramic and plastic materials.
The fundamental theory behind piezoelectricity is based on the electrical dipole (76). At the molecular level, the configuration of
a piezoelectric material is characteristically an ionic bonded crystal. In the absence of applied stress, the dipoles formed by the
positive and negative ions cancel each other as a result of the symmetry of the crystal structure, and an electric field is not detected.
Upon stress, the crystal deforms, symmetry is lost, and a net dipole moment is produced. This dipole moment forms an electric field
across the crystal that is proportional to the pressure applied. Dielectric sensors require intimate contact with the material under test
(76). Sensors can be located on the surface or within the material (75). Through suitable design and range of materials, the
frequency range that piezoelectric materials can detect changes in force or motion can range from below 1 Hz to several MHz (88).
Displacements in the millimeter range can be accurately measured as well as force changes from mN to kN (84). Piezoelectricity
exists in a number of naturally occurring crystals such as quartz (75) and rochelle salt, and it can be induced in polymers such as
nylon and copolymers of vinylidene fluoride (89) (VDF) with trifluoroethylene (90,91) (TrFE), or with tetrafluoroethylene (92–94)
(TeFE). The robust solid-state construction of industrial piezoelectric ceramic sensors permits them to operate under harsh envi-
ronmental conditions, such as dirt, oil, and most chemical atmospheres. They perform well over a wide temperature range and resist
damage from severe shock and/or vibration (84).

13.02.5.3.1.1 Typical applications of piezoelectric sensors


1. Flow meters (95–97)
2. Thickness gauges (98,99)
3. Hydrophones (88,100–102)
4. Pressure sensors (85–87)
5. Accelerometers (103–107)
6. Impact-type gas ignitions (108)
7. Precise movement control (82,109–111)
8. Microphones (77–80)

Figure 18 Classes of solids.


Review of Physical Principles of Sensing and Types of Sensing Materials 21

9. Musical pick-ups (112,113)


10. Diagnostic medical ultrasound (114,115)
11. Actuators (81–84)

13.02.5.3.2 Pyroelectric Sensors


Pyroelectric sensors are the bridge between ferroelectrics and piezoelectrics. A pyroelectric material possesses a spontaneous
polarization which is not necessarily switchable by an electric field, but interpreted via the ionic positions (116). The pyroelectric
effect is the event whereby a change in temperature in a polar dielectric will generate charge and induce a change in its electrical
dipole moment (117). The change in temperature alters the positions of the atoms slightly within the crystal structure, such that the
polarization of the material changes. This polarization change gives rise to a temporary voltage across the crystal. It can be used to
create a flow of current in an external circuit connected to a piece of the material (116).
The effect of pyroelectric on infrared detection was first investigated by Ta in 1938 (117). Pyroelectric infrared (PIR) detectors
convert the changes in incoming infrared light to electric signals. Pyroelectric materials are characterized by having spontaneous
electric polarization, which is altered by temperature changes as infrared energy is absorbed in a thin chip of the pyroelectric
material and illuminates the elements. This current can be then be detected in an external circuit. Much research has been carried out
since then, and detectors have been reported from the visible through the infrared (118,119) to submillimeter in addition to
millimeter wavelengths (117). PIR detectors have furthermore been used at radiation modulation frequencies from a few hertz to
several gigahertz (117).
PIR detectors offer an advantage of room temperature operation and wide spectral response. PIR detectors offer an
advantage of ambient temperature operation, leading to low power consumption (118), broad wavelength response, high
stability, and high sensitivity in comparison with other thermal detectors such as thermopiles or resistive bolometers (119).
These devices usually consist of single or dual elements in a simple transistor-type package, incorporating a window that is
transparent over the wavelength range of interest. The compactness and lack of any requirement for logistical support in the
form of cooling fluids or high-pressure gas gives PIR detectors a number of significant advantages over photon detectors used
in these wavebands, such as MCT or InSb (120). These must be cooled to cryogenic temperatures to obtain optimum
performance.
Over the past few years, there has been a remarkable growth of interest in arrays of pyroelectric elements for a variety of
applications. The chief driving force for infrared detection and thermal imaging (121) has been used in areas such as military
night vision, target acquisition, and missile guidance (118). Addition applications include automotive vision enhancement and
prevention, remote temperature measurements, and medical diagnostics (cancer and rheumatism detection) (118). Other appli-
cations of pyroelectric detectors have found a huge range of applications in products ranging from flame and fire detectors (118),
remote light switches to intruder detectors, in low cost spectroscopic instrumentation such as gas analyzers (120) where the window
on the front of the detector is selected and/or coated to make the device sensitive at the wavelength of interest and laser beam
characterization. Below is a list of common pyroelectric applications:
1. Intrusion sensor (121,122)
2. Pollution monitors (121)
3. Light control (117)
4. Temperature measurements (123,124)
5. Flame detector (123,125)
6. Automatic door switch (126,127)
7. Visitor detector (117)
8. Home security (128)
9. Pyrometry (128)
10. Spectrometry (128)
11. Gas analysis (121,128)
The basic design of a pyroelectric detector is illustrated in Figure 19. The pyroelectric component is attached to a high input
impedance amplifier, characteristically a field effect transistor. The pyroelectric current, ip (ip ¼ Apdq/dt, where A is the area of the
pyroelectric element, p is the pyroelectric coefficient, and dq/dt is the rate of change of the element temperature (q) with time)
creates a voltage across the electrode. The dipole moment, when normalized by the amount of material present, yields a polar-
ization. This change in polarization is observed when a stress is applied to the material, yet it will not reverse under the
application of an electric field since it will break down first, i.e., the coercive field exceeds the breakdown field (129). This is only
true for a nonferroelectric pyroelectric material. With a pyroelectric, the polarization will typically decrease when its temperature is
raised. This is because the increasing disorder results in a reduced segregation of charge, so the arising dipoles are lessened in
magnitude.
The circuit shown is a unity-gain voltage amplifier that effectively couples the high impedance source of current (the pyroelectric
element) to a low input impedance following circuit. The ip can be determined in terms of the average IR power (sinusoidal
modulated signal W(t) ¼ Woexp(iut)) occurrence upon the pyroelectric element. This verifies that the PIR detector responds to the
changes in element temperature with time, and that ip is maximized by first of all maximizing the active area of the element. Large
22 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 19 Schematic design of a pyroelectric detector. Reproduced from Whatmore, R. W. Pyroelectric Arrays: Ceramics and Thin Films. J. Elec-
troceram. 2004, 13, 139–147.

area detectors are inclined to be more intrinsically sensitive, and any material with a larger pyroelectric coefficient will produce more
charge and therefore more current (117).

13.02.5.3.3 Ferroelectric Sensors


Ferroelectricity is a class of piezoelectric materials that was discovered in 1921 by J. Valasek, who was investigating the dielectric
properties of Rochelle salt (NaKC4H4O6$4H2O) (2). All ferroelectric materials are pyroelectric (thermoelectric transducers);
however, not all pyroelectric materials are ferroelectric. Ferroelectric and pyroelectric materials are polar and exhibit a spontaneous
polarization or electric dipole moment upon cooling below the Curie point (130,131). Conversely, the polarity with ferroelectric
materials can be reoriented or reversed fully by applying an electric field (132). Complete reversal of the spontaneous polarization is
called ‘switching.’ The nonpolar phase encountered above the Curie temperature is known as the paraelectric phase. The direction of
the spontaneous polarization corresponds to the crystal symmetry of the material. The level of the spontaneous polarization is
greatest at temperatures well below the Curie temperature and approaches zero as it comes in close proximity to the Curie
temperature (130,131).
Since all pyroelectric materials are piezoelectric (electromechanical transducers), this means ferroelectric materials are inherently
piezoelectric. These materials will produce an electric charge proportional to an applied mechanical load (131). Correspondingly,
the material will produce a mechanical deformation in response to an applied voltage. Properties such as the piezoelectric,
dielectric, and electro-optic coefficients can vary by several orders of magnitude when presented around the Curie temperature. The
piezoelectric coefficient is much greater in the region of the Curie temperature. Other properties such as dielectric strength and
electro-optic properties in addition transform more noticeably in the Curie temperature region when compared to other temper-
ature ranges.
Further research on ferroelectric materials in the 1950s led to the discovery of barium titanate (BaTiO3) (132), which is perhaps
the most commonly thought of material when one thinks of ferroelectricity. It is a commonly based ceramic used in capacitor
applications and piezoelectric transducer devices. Since then, many other ferroelectric oxides with high dielectric constants and high
electromechanically coupling coefficients such as perovskite (133–135), tungsten bronze (136), pyrochlore (137,138), and bismuth
titanate layer structures (139–141) have been discovered. While there are some 250þ materials that exhibit ferroelectric properties
(2), some of the more significant ceramic including lead titanate (PbTiO3) (142), lead zirconate titanate (PZT) (132), lead
lanthanum zirconate titanate (PLZT), and lead magnesium niobate (PMN) (142) have been developed and utilized for a variety of
applications. These transition-metal elements are the highly polarizable ‘active’ ions promoting ferroelectricity and the high
permittivity and piezoelectric constants required for sensing (143). With the growth of ceramic processing and thin film technology,
various novel applications have materialized (132).
Ferroelectric materials have found applications in a wide range of sensors and transducers. The sensor materials include bulk
ceramics, multilayer ceramics, single crystals, polymers, and ceramic–polymer composites (131). Ferroelectric sensors are used
to measure force, pressure, flow and motion, temperature, and IR radiation (68). Imaging using ferroelectric transducers
includes ultrasonic medical imaging, underwater transducers, devices for nondestructive evaluation, and testing and thermal
imaging systems (68). The principal use of ferroelectric ceramics has been in the regions such as dielectric ceramics for capacitor
Review of Physical Principles of Sensing and Types of Sensing Materials 23

applications (132) ferroelectric thin films for nonvolatile memories, piezoelectric materials for medical ultrasound imaging and
actuators, and electro-optic materials for data storage and displays (68). A summary of ferroelectric applications are given
below:
1. Capacitors (132)
2. Nonvolatile memory (144–146)
3. Ultrasound imaging and actuators (147–149)
4. Electro-optic materials for data storage applications (150,151)
5. Thermistors (152–154)
6. Electrochemical switches (155–158)
7. Oscillators and filters (159–161)
8. Light deflectors, electro-optical-acoustic modulators and displays (131)
In recent years, extensive effort has been made to develop infrared and imaging devices that do not require a cooling stage.
Consequently, the use of crystalline ferroelectrics, such as triglycine sulfate (TGS), polyvinyldifluoride (PVDF), strontium barium
niobate (SBN), lithium tantalate (LT), and ferroelectric perovskite ceramics have been developed (118). The bulk of the piezoelectric
materials used for commercial sensing applications come from synthetic polycrystalline ferroelectric ceramics, such as PZT (142).
PIR detection that use PZT-based ferroelectric oxide ceramics offer a detector that is cheap to the manufacture by using standard
mixed ceramic oxide processing steps (142). Compared to single crystals, ceramics propose high strength and ease of fabrication,
especially into complex shapes and large area pieces. These detectors are also mechanically and chemically robust, allowing large
area wafers to be made, which can be easily machined into thin sections (118). They have the advantage of being nonhygroscopic, as
they are high pyroelectric performance triglycine sulfate single crystals (118).
The overall characteristic and advantages of ferroelectric materials (68) are summarized below:
1. High sensitivity
2. Flexible selection of IR wavelength filters
3. Ambient temperature operation
4. Low cost
5. Robust even under harsh environmental conditions
6. Stable against temperature and atmospheric changes
7. Stable against electromagnetic interferences

13.02.6 Magnetic and Superconducting Materials

Magnetism was first comprehensively studied by William Gilbert. It was his conclusion that the Earth acts as a huge magnet. The
word ‘magnetism’ comes from the district of Magnesia in Asia, which is one of the places where magnetic stones were found. There is
a strong similarity between the properties of materials under the influence of a magnetic field and under an electric field (162). For
example, two electrically charged rods have like and unlike ends, very much in the same way as two magnets have opposite ends. In
magnets, these ends are called N (north) and S (south) poles, as shown in Figure 20. The like poles repel, whereas the unlike ones
attract. Unlike electric charges, the magnetic poles always come in pairs. When a magnetic pole is placed in a certain space, that space

Figure 20 Test magnet in a magnet field. Reproduced from Frank, R. Handbook of Modern Sensors: Physics, Designs, and Applications (edited by
J. Fraden). Anal. Bioanal. Chem. 2005, 382, 8–9.
24 Review of Physical Principles of Sensing and Types of Sensing Materials

Table 4 SI and cgs units of important magnetic quantities

Quantity Symbol SI unit Cgs unit

Magnetic field B0 T 104 Oe


Magnetic induction B T 104 Oe

Magnetic permeability mr –
Magnetic susceptibility X J T2 kg1 106 erg Oe2 g1
per unit mass 106 emu g1
Magnetic susceptibility k J T2 m3 107 erg Oe2 cm3
per unit volume 107 emu cm3
Magnetization per unit s J T1 kg1 1 erg Oe1 g1
mass 1 emu g1
Magnetization per unit M J T1 m3 103 erg Oe1 cm3
volume 103 emu cm3
Energy product of permanent magnet k J m3 0.1256 MG Oe 104 erg cm3
Magnetic polarization J T 104 G
Reproduced from Moseley, P. T.; Crocker, A. J. Sensor Materials, 1996; p 227.

around the pole appears to have been altered from what it was before. If a piece of iron is brought into that space, it will experience
a force that it will not experience if the magnet is removed. This altered space is called a magnetic field (162).
Magnetic sensing materials are different from other sensing materials in that their detectors do not directly measure the physical
property of interest. Other devices described in this chapter detect properties such as temperature, pressure, strain, or flow, which
provides an output that directly reports the desired parameter (163). Magnetic sensors differ as they detect changes, or disturbances, in
magnetic fields that have been created or modified and obtain information on properties; for example, direction, presence, rotation,
angle, or electrical current. Magnetic detectors are reported to be somewhat difficult to use, yet they do offer accurate and reliable data
without the need for physical contact (163). The units of magnetism are quoted in SI units as H field in amperes per meter (A m1) or
as B field in teslas; or in cgs units as oersted (Oe) or gauss (G) (2). The units of important magnetic quantities in SI and cgs units are
highlighted in Table 4. The field is in B units, where B0 is for free space field and B for magnetic induction (both in teslas).
Magnetic sensors can be categorized by low-, medium-, and high-field sensing ranges. The magnetic induction B in a material
is related to that in vacuum B0 by the relative permeability mr, i.e.,
mr ¼ B=B0 [16]
Every substance reacts to a magnetic field to some degree and there response can be divided into various classes depending on
their mr value (2):
1. Diamagnetic materials (mr  1, independent of B or temperature);
2. Paramagnetic materials (mr  1, independent of B, decreases with temperature);
3. Ferromagnetic materials (large and positive mr, vary with B);
4. Antiferromagnetic materials (mr  1, hysteresis effect dependent on B, Neel temperature similar to Curie temperature);
5. Ferrimagnetic materials (mr similar to ferromagnetics but nonmetallic).
Devices that detect magnetic fields <1 mG (microgauss) are considered low-field sensors; those with a range of 1 mG to 10 G are
Earth’s field sensors; and detectors that sense fields >10 G are identified as bias magnet field sensors. Table 5 lists magnetic sensor
technologies and their sensing ranges (163).

13.02.6.1 Diamagnetism
Diamagnetism was discovered and named in September 1845 by Michael Faraday. This is the weakest form of magnetism that is
displayed in the presence of an external magnetic field. The orbital motion of electrons changes due to the external magnetic field.
This induced magnetic moment is extremely small and in a direction opposite to that of the applied field. When placed within
a strong electromagnet, diamagnetic materials are attracted toward regions where the magnetic field is weak (162).
Diamagnetism is found in all materials; however, because it is so weak it can only be observed in materials that do not exhibit
other forms of magnetism. Diamagnetic substances are composed of atoms that have no net magnetic moments (i.e., all the orbital
shells are filled and there are no unpaired electrons). However, when subjected to a field, a negative magnetization is produced and
thus the susceptibility is negative (2).
An exception to the ‘weak’ nature of diamagnetism occurs when a number of materials become superconducting. Supercon-
ductors are ideal diamagnets; when positioned in an external magnetic field, they expel the field lines from their interiors
(depending on field intensity and temperature). Superconducting magnets are the foremost elements of most magnetic resonance
imaging (MRI) systems and are among the most important applications of diamagnetism. Bismuth, which is used in guns, displays
the strongest diamagnetism. Bismuth can be melted down and molded to efficiently capture any diamagnetic properties. Pyrolytic
graphite is also an unusually strongly diamagnetic material that can be stably floated on a magnetic field (162).
Review of Physical Principles of Sensing and Types of Sensing Materials 25

Table 5 Magnetic sensor technologies and their detectable field ranges

Reproduced from Lenz, J. A. Review of Magnetic Sensors. Proc. IEEE 1990, 78, 973–989.

13.02.6.2 Ferromagnetism
Ferromagnetism comes from the term ‘ferrous’ meaning iron, the first type of metal discovered to exhibit attraction to magnetic
fields. Ferromagnetism is the basic method in which a compound forms a permanent magnet or is attracted to a magnetic field. In
a nonmagnetic compound, permanent magnetic dipoles tend to line up antiparallel in order to cancel each other out. In ferro-
magnetism, however, the opposite is true. Ferromagnetism arises from the spontaneous lining up of permanent dipoles parallel to
each other within a compound. These magnetic dipoles arise from the movement of pairs of electrons within their atomic/
molecular orbital’s (162). Ferromagnetism is normally seen in materials that have partially filled outer valence shells. Ferromagnetic
materials exhibit parallel alignment of moments, resulting in large net magnetization even in the absence of a magnetic field.
However, this aligning of magnetic moments does not mean that a ferromagnetic compound is magnetic itself. Although in
ferromagnetic compounds the permanent dipoles line up with each other, they do so in domains. Throughout the bulk of the
material, the dipoles line up in sections or domains. These line up opposite to each other, meaning that the overall magnetic
moment throughout the material is zero. Elements such as Fe, Ni, and Co and alloys with other elements (titanium, aluminum) that
exhibit relative magnetic permeabilities up to 104 are well known ferromagnetic materials (2).
In ferromagnetic substances within a certain temperature range, there are net atomic magnetic moments that line up in such
a way that magnetization persists after the removal of the applied field. Below a certain temperature, called the Curie point (or
Curie temperature), an increasing magnetic field applied to a ferromagnetic substance will cause increasing magnetization to
a high value called the saturation magnetization. This is because a ferromagnetic substance consists of small magnetized regions
called domains. The total magnetic moment of a sample of the substance is the vector sum of the magnetic moments of the
component domains (2).
In order to induce permanent magnetism, a ferromagnetic compound must be placed within a strong magnetic field. Once in the
magnetic field, the domains line up parallel to generate one strong magnetic moment. It is known for some materials that the
domains remain in this position even when the external magnetic field is removed. The ability of some materials to remember
this ‘magnetic history’ is called hysteresis (2). Since energy is required for the reformation and saturation of the domains, there is
a hysteresis in the magnetization field curve (hysteresis loop), as illustrated in Figure 21. In most cases, not all of the domains will
remain in this configuration. The fraction of domains remaining in this configuration is known as remanence (162).
There are a number of crystalline materials that exhibit ferromagnetism. A list of ferromagnetic materials (2,68), along with their
Curie temperatures (temperature above which they cease to be ferromagnetic), are highlighted in Table 6.

13.02.6.3 Paramagnetism
Paramagnetism is similar to ferromagnetism. The atoms or molecules of the substance have net orbital or spin magnetic moments
that are capable of being aligned in the direction of the applied field. Therefore, they have a positive (but small) susceptibility and
a relative permeability slightly in excess of one. Paramagnetism occurs in all atoms and molecules with unpaired electrons; e.g.,
free atoms, free radicals, and compounds of transition metals containing ions with unfilled electron shells. It also occurs in metals
as a result of the magnetic moments connected with the spins of the conducting electrons (2,68).
26 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 21 A hysteresis curve for ferromagnetic material with remanant magnetization Mr and coercive field Hc. Reproduced from Moseley, P. T.;
Crocker, A. J. Sensor Materials, 1996; p 227.

Like ferromagnetism, paramagnetism is the aligning or permanent dipole moments parallel to each other. Paramagnetic
materials atoms do have permanent dipole moments; however, ferromagnetism is not active. Paramagnetism only occurs when the
material is subjected to an external magnetic field; the dipole moments try to line up with the magnetic field but are prevented from
becoming perfectly aligned by their random thermal motion (162). Since the dipoles try to line up together with the applied field,
the susceptibilities of such materials are positive; but in the absence of the strong ferromagnetic effect, the susceptibilities are rather
small (magnetic susceptibility of the order of 105–103). However, once the strong magnetic field is removed thermal processes
will quickly disrupt the magnetic alignment, randomizing the dipole moments, and the net magnetic alignment is lost as the dipoles
relax back to their normal random motion. This results in an overall zero magnetic moment. Magnetization obeys Curie-Weiss law
(similar to that found for piezoelectric materials). A weak force between a magnetic field and the material (repulsive for diamagnetic
materials and attractive for paramagnetic materials) under relatively low magnetic field saturation results in magnetization
according to Curie-Weiss law (2,162),
 
B
M¼C [17]
T
where M is the resulting magnetization (magnetic dipole moment/unit volume) measured in amp m1, C is a material-specific
Curie constant (different for each different material), B is the applied magnetic field measured in teslas, and T is absolute
temperature measured in Kelvins. Curie’s law says that if B is increased, the magnetization increases (stronger magnetic field aligns
additional dipoles). It also states that if there is an increase in temperature, the magnetization decreases (the increased thermal
agitation helps prevent alignment). Curie’s law only works for samples in which only a comparatively small fraction of the atoms
are aligned, on the average, with the magnetic field. Upon the aligned fraction becoming larger, Curie’s law no longer holds, as it
predicts magnetization just goes up forever with increasing applied magnetic field B. Once the dipoles are 100% aligned, further
increases in the magnetization are impossible as the material is saturated, and further increases in B or decreases in T will no longer
change the magnetization as the atoms are completely aligned (2,162). Paramagnetic materials, similar to ferromagnetic materials,

Table 6 A selection of crystalline ferromagnetic materials, along with their


Curie temperatures in kelvins (K)

Material Curie temperature (K)

Fe 1043
Co 1388
Ni 627
Gd 292
Dy 88
MnAs 318
MnBi 630
MnSb 587
CrO2 386
MnOFe2O3 573
FeOFe2O3 858
NiOFe2O3 858
CuOFe2O3 728
MgOFe2O3 713
EuO 69
Y3Fe5O12 560
Review of Physical Principles of Sensing and Types of Sensing Materials 27

have a positive response to external magnetic fields, i.e., it becomes a magnet. As long as the strong magnetic field is present, it will
attract and repel other magnets in the usual way. However, unlike ferromagnetic materials, paramagnetic materials do not contain
domains. The material reacts as a whole to the field. For this reason only a small fraction of the dipole moments align with the field,
producing only a slight effect. The attraction to the field is so slight that it is unnoticeable in everyday life. A list of paramagnetic
materials is given below (2,68):
l Aluminum
l Barium
l Calcium
l Liquid oxygen
l Platinum
l Sodium
l Strontium
l Uranium

13.02.6.4 Superconductors
The component of an electrical circuit that causes energy loss is described as resistance, which can be defined as a material’s
opposition to current being passed through it. Typically, this resistance results in the production of heat, sound, or another form of
energy. In many cases, this conversion of energy is useful in applications such as toasters, heaters, and light bulbs. However,
resistance can hinder the performance of high-voltage transmission wires, electric motor output, and other cases where internal
system energy losses are unwanted. This is where the phenomenon of superconducting materials comes into play and provides
possible solutions to energy loss problems (164).
Superconductors are materials that have exactly zero electrical resistance and are one of the last great frontiers of scientific
discovery (164). Superconductivity can transport electrical charge without resistance, without ‘paying’ anything for this transport,
and without heating up the wires. The electrons, when cooled below a certain temperature, form a very organized state, giving the
material a number of unusual properties. The first is that electric current is able to flow through the conductor with entirely no loss
of energy. In other words, the resistance goes to zero (165). Superconductivity is also a phenomenon of expulsion of magnetic fields
occurring in certain materials when cooled below a characteristic critical temperature (164). The second unusual property observed
is when a superconductor shaped like a ring is positioned in a magnetic field and a current begins to flow. Faraday’s law of induction
states that altering the magnetic flux in a loop causes a voltage to occur, and this voltage can drive a current. However, since the
resistance in a superconductor is zero, the current does not dissipate and die out; instead it keeps flowing around the loop in
perpetual motion (166). Finally, the third unusual property and the most spectacular exhibition of superconductivity can be
demonstrated if a magnet is put above a sheet of superconductor. The superconductor will force out the field from the magnet,
causing it to levitate in the air (165).
Zero resistance is displayed under certain conditions (167–169). These conditions are called the ‘critical temperature’ and ‘critical
field,’ denoted by Tc and Hc, respectively. The Tc is the highest temperature state the material can attain and remain supercon-
ductive. The Hc is the highest magnetic field the material can be exposed to before reverting to its normal magnetic state (170,171).
The limits of superconductivity have not yet been reached, but the theories that explain superconductor behavior seem to be
constantly under review. In 1911, superconductivity was first discovered when Heike Kamerlingh Onnes’ student reported him the
zero resistance state below 4.1 K (452  F, 269  C) of mercury (166). The experiment was repeated several times before it became
apparent that they were facing a genuine discovery. In 1913, this sudden and fundamental disappearance of resistance was rewarded
by the Nobel Prize in physics for Onnes’ research in this area (164). During the decades new superconductors were discovered,
a wealth of experimental results have built up, which finally led to the understanding of the microscopic mechanism of super-
conductivity by Bardeen, Cooper, and Schrieffer (BCS theory) in 1957 (165). The description of this macroscopic coherent quantum
state by the BCS theory was rewarded by the Nobel Prize in 1972. Since then, many other metals that become superconductors at
very cold temperatures have been found, such as lead, aluminum, tin, and niobium and titanium, which are mixed together and
drawn into wires to make very strong magnets for MRI (172–174).
The zero resistance of superconductors is an attractive property for applications such as power transmission and photon
detectors. Superconductivity arises when electrons in a suitable material bind together in what are identified as ‘Cooper pairs’ which
then flow en masse like a superfluid. The BCS theory suggests that electrons team up in Cooper pairs in order to help each other
overcome molecular obstacles (164). Cooper pairing is initiated by an attractive force between electrons from the exchange of
phonons. The energy spectrum of this Cooper pair fluid possesses an energy gap, meaning that a minimum amount of energy (DE)
must be provided in order to excite the fluid. Therefore, if DE is larger than the thermal energy of the lattice, given by kT (where k is
Boltzmann’s constant and T is the temperature), the fluid will not be scattered by the lattice. The Cooper pair fluid is therefore
a superfluid, meaning it can flow devoid of energy dissipation. Superconductivity only occurs in a material when it is cooled below
an extremely low temperature, termed its critical transition temperature (Tc). Cooling a material decreases the vibrations of its
atoms. However, an increase in temperature above that threshold will cause the thermal energy of the vibrations to break the Cooper
pairs, eliminating the superconductivity occurrence. Many superconducting devices have to be refrigerated to within a few degrees
above absolute zero (0 K) due to their sensitivity to heat. Some materials require temperatures as low as a few hundredths of
28 Review of Physical Principles of Sensing and Types of Sensing Materials

a Kelvin. It is for this reason that scientists have struggled for years to develop materials with more robust superconductivity that
survives at higher temperatures (172–174).

13.02.6.4.1 Type I Superconductors


Within the substances currently known to superconductors, there is a divide between what has come to be called type I and type II
superconductors. Perfect diamagnetism means that a superconducting material does not permit an externally applied magnetic field
to penetrate into its interior. Those superconductors that totally exclude an applied magnetic flux are known as type I supercon-
ductors (166). The type I category are composed of pure substances, usually metals and metalloids that show some conductivity at
room temperature (Figure 22) (170,171).
Type I superconductors were discovered first and require the coldest temperatures to become superconductive. These super-
conductors are often characterized as the ‘soft’ superconductors (174). They need to be cold (Figure 22) to slow down molecular
vibrations sufficiently to facilitate unimpeded electron flow in accordance with the BCS theory. They exhibit a very sharp transition
to a superconducting state and have the ability to repel a magnetic field completely (diamagnetism). Below is a list (Table 7) of Type
1 superconductors alongside their critical transition temperature (Tc) (166).

13.02.6.4.2 Type II Superconductors


Type II superconductors are composite compounds, usually some sort of ceramic. Except for the elements vanadium, technetium,
and niobium, the type II category of superconductors is comprised of metallic compounds and alloys (174). Type II supercon-
ductors are sometimes denoted as ‘hard’ superconductors. They differ from type I superconductors in that their transition from
a normal to a superconducting state is gradual across a region of ‘mixed state’ behavior (174). Further differences between type I and
type II exist, chiefly that type II display superconducting qualities at much higher temperatures (type II can have Tc’s of over 130 K)
and can remain superconductive in the presence of much higher magnetic fields. The difference in magnetic fields between both
types of superconductors is quite large. Type I superconductors withhold fields up to approximately 2000 Gauss (w0.2 T), while
type II can with stand fields of up to several hundred thousand Gauss (w10 T). The magnetic field for any temperature below the Tc
is given by the following equation:

Bc ¼ Bcð0Þ  ½1  ðT=TcÞ  2 [18]


Type II will allow some penetration by an external magnetic field into its surface creating novel mesoscopic incidents like
superconducting ‘stripes’ and ‘flux-lattice vortices’ (166). For example, recently discovered superconducting ‘perovskites’ (metal-
oxide ceramics that normally have a ratio of 2 metal atoms to every 3 oxygen atoms) achieve higher Tc’s than type I superconductors
by a mechanism that is still not completely understood (166). Researchers suggests the holes of hypocharged oxygen in the charge
reservoirs are responsible (167).
The first superconducting type II compound was made in 1930 by W. de Haas and J. Voogd. It was an alloy of lead and
bismuth, yet its property as a superconductor was not recognized until the ‘Meissner’ effect had been discovered (175). The
Meissner effect describes a superconductor when it is placed in a weak external magnetic field H and cooled below its transition

Figure 22 Effect of temperature on Type 1 superconductors. Reproduced from Poole, C. K. Handbook of Superconductivity, 1999; p 540.
Review of Physical Principles of Sensing and Types of Sensing Materials 29

Table 7 Type 1 superconductors alongside their critical transition


temperature (Tc)

Lead (Pb) 7.196 K


Lanthanum (La) 4.88 K
Tantalum (Ta) 4.47 K
Mercury (Hg) 4.15 K
Tin (Sn) 3.72 K
Indium (In) 3.41 K
Palladium (Pd) 3.3 K
Chromium (Cr) 3K
Thallium (Tl) 2.38 K
Rhenium (Re) 1.697 K
Protactinium (Pa) 1.40 K
Thorium (Th) 1.38 K
Aluminum (Al) 1.175 K
Gallium (Ga) 1.083 K
Molybdenum (Mo) 0.915 K
Zinc (Zn) 0.85 K
Osmium (Os) 0.66 K
Zirconium (Zr) 0.61 K
Americium (Am) 0.60 K
Cadmium (Cd) 0.517 K
Ruthenium (Ru) 0.49 K
Titanium (Ti) 0.40 K
Uranium (U) 0.20 K
Hafnium (Hf) 0.128 K
Iridium (Ir) 0.1125 K
Beryllium (Be) 0.023 K
Tungsten (W) 0.0154 K
Platinum (Pt) 0.0019 K
Lithium (Li) 0.0004 K
Rhodium (Rh) 0.000325 K

temperature and the magnetic field is ejected. The Meissner effect does not cause the field to be entirely ejected, but as an
alternative penetrates the field of superconductor to a very small distance, denoted as l. This is called the London penetration
depth, which decays exponentially to zero within the bulk of the material. The Meissner effect is a crucial characteristic of
superconductivity. For most superconductors, the London penetration depth is on the order of 100 nm. A superconductor with
little or no magnetic field within it is said to be in the Meissner state. This latest category of superconductors was identified by L.V.
Shubnikov at the Kharkov Institute of Science and Technology in the Ukraine in 1936, when he found two distinct critical
magnetic fields (known as Hc1 and Hc2) in PbTl2 (166,176).
The first of the oxide superconductors was created in 1973 by DuPont researcher Art Sleight when Ba(Pb,Bi)O3 was found to
have a Tc of 13 K (166,176). A breakthrough was made when scientists found that certain ceramic compounds became super-
conductors at relatively high temperatures (Tc’s that reached 23 K) in 1986. These superconducting cuprate compounds are
layered structures, with the superconductivity usually occurring in a plane of copper and oxygen atoms (166). Such a high
transition temperature is theoretically impossible for a conventional superconductor, leading the materials to be termed high
temperature superconductors. Liquid nitrogen boils at 77 K, assisting numerous experiments and applications that are less
practical at lower temperatures. As of late, the world record holder for the highest Tc attained at ambient pressure for a material
that will form stoichiometrically is the compound Hg0.8Tl0.2Ba2Ca2Cu3O8.33, which becomes a superconductor at 138 K
(211Â F) and nonstoichiometrically has been 18  C (167). Conventional superconductors are used at 39 K for magnesium
diboride (MgB2) (177,178), as well as cuprate superconductors YBa2Cu3O7, with a critical temperature of 92 K. Mercury-based
cuprates have been found with critical temperatures in excess of 130 K (179). However, because ceramics are fragile, the
superconductors cannot be easily drawn into wires like metal ones. Instead, the ceramic is developed on top of a flexible tape in
a thin layer so that it does not crack when bent. This tape can then be wrapped around a tube to carry the liquid nitrogen coolant
and surrounded with insulation to make a superconducting cable (179). These cables are used inside power plants to transmit
electric power.

13.02.6.4.3 Superconductors as Sensing Materials


Superconductors differ fundamentally from conventional materials in the manner by which electrons, or electric currents, move
through the material. It is these differences that give rise to the unique properties of sensing materials from all other known
30 Review of Physical Principles of Sensing and Types of Sensing Materials

conductors. As a novel sensing material, they have an extensive impact on today’s sensors throughout the world. A list of super-
conductor’s unique properties as sensing materials is given below (175,179–181):
1. Zero resistance to direct current (electric power transmission);
2. Exceptionally high current density (powerful magnets for motors, generators, energy storage, medical equipment, and industrial
separations);
3. Particularly low resistance at high frequencies (reduces substantially the challenges inherent to miniaturization brought about
by resistive, or I2R, heating);
4. Extremely low signal scattering (microwave components, communications technology, and several military applications);
5. Extraordinary sensitivity to magnetic field (a unique 1000 sensing capability to today’s best conventional measurement
technology);
6. Exclusion of externally applied magnetic field (multilayer electronic component miniaturization, provides a mechanism for
magnetic levitation and enables magnetic field containment of charged particles);
7. High-speed single flux quantum transfer (digital electronics and high-speed computing);
8. Close to speed of light signal transmission (digital electronics and high-speed computing).
The superconducting sensors available today are 10–100 times more sensitive than conventional sensors operated at room
temperature. These devices are improving measurements in a broad range of fields. Superconductivity forms the basis for new
commercial products that are transforming our economy and daily life (175,179–181), such as
1. Magnetic resonance imaging (MRI)
2. Nuclear magnetic resonance (NMR)
3. High-energy physics accelerators
4. Plasma fusion reactors
5. Industrial magnetic separation of kaolin clay

13.02.7 Solid Electrolytes

An electrolyte is a substance that conducts electricity through the movement of ions. Most electrolytes are liquids or molten salts,
but some electrolytes are solids (182,183). Solid electrolytes allow the conduction of ions but not the conduction of electrons (2).
Such materials often have rather special crystal structures like open tunnels or layers through which the mobile ions may move. The
conductivity values of 103 S cm1 for Naþ ion migration in b-alumina at 25  C are comparable to some strong liquid electrolytes.
There are, however, a few solids like LixCoO2 or Li intercalated graphite that conduct electricity both by ions and electrons (holes).
These are referred to as mixed conductors, and they are particularly important battery electrode materials. Hence, there is currently
great interest in studying the properties of the solid electrolytes, developing new ones, and extending their range of applications in
solid state electrochemical devices. However, this chapter will deal with basics of solid electrolyte materials with respect to their use
in sensors only.
As already stated, solid electrolytes conduct electricity by the passage of ions. Usually, one type of ion (anion or cation) is
predominantly mobile and conducts electricity in the solid (182–184). Many ionic solids exhibit high electrical conductivity above
a certain temperature (AgI) (185). However, there are a few with room temperature (25  C) conductivity greater than 102–
103 S m1. They are called fast ionic conductors or superionic conductors. Ag4RbI5 and Cu4RbCl3I2 are two such superionic
conductors that are known to exhibit excellent conductivity of about 30 S m1 at 25  C (186).
Microscopically, the type of electrical conductivity exhibited by a solid is dictated by the defect structure of the material. A perfect
crystal of an ionic compound would be an insulator. Based on the types of defects, the solid electrolytes are classified as point defect
type and molten sublattice type (184,185).

13.02.7.1 Point Defect Type


The various kinds of point imperfections possible in an ionic crystal taking into account the charge neutrality are as follows:
Vacancies: A missing Mþ ion in a pure binary compound missing from its normal site depicted with an arrow in Figure 23.
Interstitials: An ion Mþ or X occurs in the interstitial site, as denoted in Figure 24. The term vacancy (or interstitial) defect may be
understood to occur if in a volume element of the lattice, a particle is missing or contains an excess particle, with respect to the ideal
lattice, independent of whether at that point or in its immediate vicinity particles deviate from their normal position.
Misplaced atoms: An atom X occupying a normal M site or vice versa, as shown in Figure 25.
Schottky defect: A cation vacancy together with an anion vacancy predominantly present in alkali halides, Figure 26. In this
vacancy mechanism of conduction, positive and negative ions leave their normal sites to jump into vacancies.
Frenkel defect: A cation vacancy together with an interstitial cation as in silver halides, as seen in Figure 27.
The driving force for these Schottky and Frenkel defects is the increase in entropy or degree of system disorder, and they apply
both to fully stoichiometric and nonstoichiometric materials. Solid electrolytes usually have large band gaps and high Schottky and
Frenkel defect formation energies. These thermally induced defects normally contribute very little to ionic conductivity (2).
Review of Physical Principles of Sensing and Types of Sensing Materials 31

Figure 23 Vacant lattice site.

Figure 24 Interstitial defect.

Figure 25 Misplaced atom defect.

Figure 26 Schottky defect.

Stoichiometric variation defect: Defects due to stoichiometric variations such as gain or loss of oxygen by a metal oxide also occur.
These mechanisms are mostly readily available for compounds in which at least one element, e.g., transition metal, can readily take
up more than one oxidation state, and the normal result is a change in electronic carrier density.
Impurities: Defects due to doping of ionic compound with an aliovalent impurity is the most important defect mechanism for
producing conductivity in oxides of metals with only one stable oxidation state. Impurities incorporated at the regular atomic site in
32 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 27 Frenkel defect.

the crystal are called substitutional defects (2). Aliovalent impurity is one type of substitutional defect, where the ion substituted for
the original ion is of a different oxidation state. These substitutions change the overall charge within the ionic compound, but as the
ionic compound must be neutral, charge compensation becomes necessary. Compensation occurs in the form of one of the metals
being partially or fully oxidized or reduced, or ion vacancies being created. The resultant conductivity is ionic; it is associated with
the mobility of the defect species formed (184).
Molten sublattice: It is a case of a liquid-like molten sublattice in which all the ions are available for conduction since the number
of defects or void sites in the sublattice is greater than the number of ions. So the ions can move freely from one position to another
with low activation energy and high conductivity. The number of mobile ionic charge carriers is 1022 cm3. These are often marked
by a channeled or layered structure (185).
It is characteristic of all materials with high ionic conductivity that they contain a high concentration of mobile ions distributed
among a higher concentration of crystallographic sites, having nearly equivalent energies. Therefore, in order that a material might
be useful as a solid electrolyte, its transference number for ions should be very close to unity (2). The ionic conductivity of some
solid electrolytes is as shown in Figure 28 (184). While searching for new materials, one aims for the top right-hand corner of the
diagram.
The conductivity range for solid electrolytes is 103 S cm1 < s < 10 S cm1, and it decreases exponentially as temperature
decreases; whereas it is 10 < s < 105 S cm1 for metals, and it increases linearly as temperature decreases (184). However, for some
solid electrolytes, electronic conductivity mechanisms become significant at high temperatures. For example, zirconia meets the
requirements for a good electrolyte over a wide range of oxygen partial pressures at 1000 K, but the electrolytic domain shrinks
rapidly with increasing temperature; at 2000 K the material is semiconducting across almost the entire range of oxygen partial
pressure (187). Hence it is important to define the region of temperature-oxygen partial pressure space within which the electronic
component of conductivity remains negligible in order to identify the conditions within which such oxide ion conductor may be
useful in an electrochemical sensor.
It is clear that the ionic conductivity is due to the presence of a high density of crystal defects. In order for the material to be
a useful solid electrolyte, however, the material needs to exhibit acceptable physical properties; its overall structure should remain

Figure 28 Ionic conductivity of some solid electrolytes with concentrated H2SO4 for comparison. Reproduced from West, A. R. Solid State Chemistry
and Its Applications, 1984; p 734.
Review of Physical Principles of Sensing and Types of Sensing Materials 33

coherent despite the rapid diffusion of one species within it. In some solid electrolytes the conductivity is due to the mobility of
cations, and in others it is due to the mobility of anions. Materials of both types find application in sensors. Some examples of
different types of solid state conductors are given below (182).

13.02.7.2 Cationic Conductors


Electrical conductivity is imparted by the positive ions.
Agþ ion conductors: Silver ion conductors show the high ionic conductivity at ambient temperatures compared to all other types of
solid conductors (188). Examples: AgI, Ag6I4WO4, KAg4I5, NH4Ag4I5, Ag7I4PO4, and RbAg4I5.
Naþ ion conductors: One of the most extensively studied classes of superionic conductors is the group with the general formula
nA2O3B2O (A ¼ Al3þ, Ca3þ, Fe3þ, B ¼ Naþ, Kþ). In these compounds, the conductivity is due to the motion of Bþ ions in the loosely
packed structural planes of the crystal lattice. They are mostly used in the development of high-energy density battery (189).
Examples: sodium b-alumina (NaAl11O17, Na2Al16O25) and NASICON (Na3Zr2PSi2O12).
Liþ ion conductors: LiCoO2, LiNiO2, and LiMnO2 (190).
Proton conductors: Solid state proton conductors are used in sensors, fuel cells, and electrochromic devices (191,192). Examples:
H8UO2(IO6)2$4H2O, polyamides, and polysulfinimide.

13.02.7.3 Anionic Conductors


Here, negative ions act as the charge carriers. Anionic conductors do not exhibit good ionic conductivity at ambient temperatures
(193). There are two types of anionic conductors (184,185).
O2 ion conductors: Motion of oxygen ion is responsible for the conduction mechanism. Most of them show significant value of
conductivity only at high temperatures (1273 K) and their conductivity strongly depends on the doping of the aliovalent impurity
(Ca2þ, Y3þ, Sr2þ), which controls the number of point defects and their mobility.
Examples: Cubic stabilized ZrO2, d-Bi2O3, and defect perovskites (Ba2In2O5).
F ion conductors: They are more conductive than the oxide ion because the former is univalent even though the ionic radii
of these two ions are almost the same.
Examples: PbF2, KBiF4, LaF3, Zr–Ba–CCs–F, and AF2 (A ¼ Ba, Sr, Ca).

13.02.7.4 Organic Ionomers


Nafion: A copolymer of tetrafluoroethylene and a perfluorosulphonic acid with a molecular structure shown in Figure 29 has been
extensively investigated as a potential electrolyte in solid state gas sensors (2,194).
The perfluorinated ionomers exhibit high ionic conductivity, remarkable permselectivity, and excellent chemical and thermal
stability. The ionic component comprises a minor part of the polymer (10 mol%) and readily takes up water. The mechanism of
ionic conductivity in nafion membranes is quite distinct from those involving the movement of vacancies and interstitial ions, as
stated before. The mobile species in nafions are positively charged counterions and the conductivity depends crucially on the water
content, which provides a hydration shell around the charged species. However, if the water content of the material is decreased, the
ionic conductivity also decreases as an increasing fraction of counterions start to strongly interact with the sulphonate groups of the
polymer (2).

13.02.7.5 Solid Electrolytes in Sensors


Similar to liquid electrolytes, solid electrolytes support the function of electrochemical cells, in which the chemical reactions are
allowed to proceed to completion if separate paths are provided for the flow of ions and electrons, e.g., solid oxide fuel cells
(188,195–197). In chemical sensing applications, the use of solid electrolytes dates back to the discovery of yttria-stabilized zirconia
by Nernst in 1899, which is still in use today. The solid electrolyte separates the two regions of distinct activity of the species to be
monitored and allows for high mobility of an ion of that species between the two regions. However, during the past decade solid
electrolytes have not only been used as materials to maintain a separation between two active electrodes, but they have also played
a significant role as key material in sensing electrodes and in reversible electrodes for high-energy batteries (198). The function of the

Figure 29 Structure of nafion.


34 Review of Physical Principles of Sensing and Types of Sensing Materials

sensor is addressed through the measurement of potential, current, or charge passed, and to this end it is necessary for the ionic
conductivity of the electrolyte to be high and other modes of electrical conductivity to be ideally zero (2).
The range of materials that have been employed as electrolytes in energy storage devices has been broadly matched by parallel
applications in electrochemical cells for sensing purposes (199). Solid electrolytes used in chemical sensors function in one of
a variety of modes: potentiometric, amperometric, or coulometric. However, solid electrolytes are not used in conductometric
mode. The most widespread use of solid electrolytes is in potentiometric gas sensors. A recent chapter by Pasierb et al. (200) gives
a detailed overview of CO2 potentiometric sensors with regard to construction details, working mechanism, and performance (200).
Möbius et al. (201) later added details of solid state O2 sensing to this chapter. Compared to liquid electrolyte sensors, solid
electrolyte sensors exhibit some significant advantages (199).
1. Long-term stability;
2. Lower limit of detection (ppb range);
3. Faster response times (<1 s);
4. Higher selectivity for certain components in a broader matrix of other gases;
5. High operating temperature for applications in combustion control or biotechnology, which involves high temperature steril-
ization techniques.
Solid electrolytes are used in chemical sensors in one of the following modes (2,194,199,200,202,203):
Potentiometric sensors
Amperometric sensors
Impedimetric sensors
Resistive sensors
Mixed potential sensors
Solid electrolytes already find use in a wide variety of sensors. With new materials being invented at a tremendous rate, this range
of applications is likely to increase. In the already well-established area of oxygen sensing future developments of solid electrolytes
are likely to advance toward increasingly miniaturized devices with reduced power requirement and shorter response time.

13.02.8 Biological Sensing Materials

The intrinsic ability of biomolecules (enzymes, antibodies, receptors, nucleic acids, etc.) to selectively bind certain compounds can
be utilized as a core for recognizing and/or detecting compounds (i.e., analytes) in a device for chemical and/or biological analysis.
There are two types of biosensors: biocatalytic and bioaffinity-based biosensors. The biocatalytic biosensor predominantly uses
enzymes as the biological compound, catalyzing and signaling biochemical reaction. The bioaffinity-based biosensor monitors the
binding event itself and uses specific binding proteins, lectins, receptors, nucleic acids, membranes, whole cells, antibodies, or
antibody-related substances for biomolecular recognition. In the latter two cases, these biosensors are often referred to as immu-
nosensors. Bioaffinity sensors based on enzymes, antibodies, and cells can be monitored in a solution based or surface-immobilized
format. The former entails that all components, analyte, and biological elements react in a homogeneous liquid phase, freely
moving in solution. In the latter example, the biological (biosensing) element is reconstituted in an artificial environment, for
example covalently tethered or adsorbed to a solid surface, where it is used as the principal component in recognition and detection
of the analyte. An overview of possible assay systems and the sensing materials is given below.

13.02.8.1 Enzymes
Enzymes, like antibodies, are proteins; however, they convey different properties. Enzymes are low affinity proteins that can catalyze
chemical reactions, unlike antibodies, which are high affinity proteins that bind strongly to an antigen. The most significant
property of these proteins is their natural ability to selectively catalyze (enzymes) or bind (antibodies) certain compounds or
a number of compounds, with absolute specificity for one analyte in some cases. The biological biorecognition is based on the
interaction between the binding site on the enzyme/antibody and the analyte (204). Enzymes in particular are biological catalysts
responsible for the majority of chemical reactions in living organisms, where the mode of action involves oxidation and reduction.
They initiate or accelerate reactions that would otherwise not take place. They also have the ability to slow down reactions as well as
split up reactions into separate steps, to control the heat evolution of exothermic reactions; otherwise the heat evolution could lead
to cell death (204).
Enzymes are generally bigger than the substrate they bind, so only a small portion of substrate is effectively involved in the
enzymatic reaction. The molecular recognition of the substrate is achieved by the well-known lock and key principle between the
respective receptor area and the analyte to be recognized (205). The enzyme structure is generally made up of a single polypeptide
chain, but the active site can be a separate molecule embedded in the polypeptide backbone that can bind itself to the substrate, and
in most cases the detection is done electrochemically. Only certain molecules are allowed access through the protein shell and the
binding site, and not the active site itself. Therefore, the accessibility of the enzyme is very specific. Urease, for example, is an enzyme
that is used as a biocatalyst for one compound (206), whereas other enzymes are specific for a whole group of substrates.
Review of Physical Principles of Sensing and Types of Sensing Materials 35

A reactive intermediate or the enzyme substrate (ES) complex is formed when the substrate (S) binds to the binding site of the
enzyme (E) (204). The complex ES is then converted to E and a product (P) by the active site, as shown in eqn [19]. During
enzymatic reactions, substrates are consumed and products formed. Even though the complex formation is reversible, the product
forming step with rate constant K2 is irreversible due to the enzyme’s affinity toward the product being practically negligible (207).
K1 K2
S þ E 4 ES / P þ E [19]
For example, glucose plays an important role of the metabolic processes in the human body. Glucose biosensors are devices to
be used for determination of glucose level in a biological sample (208). They are now widely used to test the concentration of
glucose for diabetic patients (208). The traditional glucose sensor detects the hydrogen peroxide produced in the oxidation
process catalyzed by a glucose oxidase enzyme (GOx). Glucose biosensors were first described in 1962 by Clark and Lyons (209).
Glucose oxidase, an enzyme that catalyzes the oxidation of glucose to gluconic acid, was attached to the biosensors. Shortly
thereafter, glucose, lactate, and alcohol biosensors were commercialized in the 1970s and have been developed for past six
decades (206).
Flavin adenine dinucleotide (FAD) is part of the GOx structure, where redox reaction takes place (210). This FAD/FADH2 redox
center is insulated deep inside the protein shell, and therefore the reactive enzyme site is some distance away from the electrode
surfaces. To overcome the inefficiency of the direct electron transfer between the FAD center and the electrode surface, the early
glucose biosensors had employed the reduction of glucose by oxygen (eqn [20]). Their first device relied on a thin layer of GOx
entrapped over an oxygen electrode and monitoring the oxygen consumed by the enzyme-catalyzed reaction.
GOx
Glucose þ O2 ƒ! Gluconic acid þ H2 O2 [20]
During the catalytic cycle the enzyme is first reduced and then regenerated by oxidation with the molecular oxygen in the sample
solution. The glucose sensor mechanism is based on the electrochemical oxidation of the hydrogen peroxide to regenerate oxygen
and to complete the cycle at the working electrode (eqn [21]). In 1973, Guilbault and Lubrano (211) described an enzyme electrode
that monitored the liberation of hydrogen peroxide for the determination of blood glucose–based sensor (212).

H2 O2 /O2 þ 2Hþ þ 2e [21]


The most important drawbacks of this sensor are the denaturation of the GOx by H2O2 and the strong dependence of the
electrical signal on the cosubstrate concentration. The concentration of oxygen is an order of magnitude lower than that of glucose;
second generation biosensors make use of an artificial electron carrier, or mediator, to overcome this limitation (212). The electron
involved in the redox process is shuttled back and forth between the active center of the enzyme and the electrode (eqns [22]–[24]).
Both mediator and enzymatic substrate must be in the analytical solution (212).

Glucose þ GOxðFADÞ/gluconic acid þ GOxðFADH2 Þ [22]

GOxðFADH2 Þ þ 2MðoxÞ/GOxðFADÞ þ 2MðredÞþ þ 2e [23]

2MðredÞ/2MðoxÞ þ 2e [24]


where M(ox) and M(red) are the oxidized and reduced forms of the mediator. The electrochemical measurements are independent
of oxygen by using mediators such as ferrocene or ferricyanide (212).
In an analytical process, enzymes are used for specific evaluation of the corresponding substrates, and they provide a consi-
derable amplification for the sensitive detection of a substrate (213). Enzymatic biosensors utilize specific enzymes for the capture
and catalytic generation of the product (213). Generation of the product is directly monitored using a range of transducers; however,
a majority of enzyme-based biosensors employ the amperometric transduction method (214). According to the International
Union of Biochemistry and Molecular Biology (IUBMB), there are six major classes of enzymes according to their function. The
IUBMB nomenclature committee recommendations have attributed a code number (Table 8) to every enzyme to identify the main
category and subcategory as well as the reaction it catalyses.
Various enzymes are used as labels, including horseradish peroxidise (HRP), glucose oxidase, and alkaline phosphatase (AP) (208).
Products of these enzymes can be detected spectrophotometrically and electrochemically. Horseradish peroxidase is a commonly used

Table 8 Six major classes of enzymes with their code number and function

Classification Type of reaction

EC1. Oxidoreductases Redox reaction


EC2. Transferase Functional group migration
EC3. Hydrolases Hydrolysis
EC4. Lyases Elimination to get double bonds
EC5. Isomerases Isomerization
EC6. Ligases Bond formation along ATP hydrolysis
36 Review of Physical Principles of Sensing and Types of Sensing Materials

enzyme. It has a high kinetic rate that maximizes the enzymatic signal amplification (215). By a simple conversion of HRP catalyzed
electron transfer to an amperometric signal, an electrochemical sensor can measure the number of target oligonucleotides on the
sensor surface. The current observed is proportional to the number of molecular targets in the sample. Zhang et al. (216) reported HRP
in the amplification of amperometric detection of DNA. The detection of DNA was as low as 3000 copies of a concentration of 0.5 fM.
Pividori et al. (217) reported a five-step procedure for an enzyme-based amperometric detection of hybridization:
1. DNA target immobilization was absorbed onto a nylon membrane;
2. Hybridization occurred between DNA target and biotin DNA probe;
3. A complexation reaction occurred between the biotin DNA probe and HRP conjugate;
4. The modified membrane was integrated onto electrochemical transducer;
5. Amperometric detection using suitable substrate for enzyme labeled duplex.
Enhanced chemiluminescense is commonly used in detection assays in biosensors. Horseradish peroxidase is attached to
a molecule of interest, which is catalyzed by enhanced chemiluminescense to convert the substrate into a sensitized reagent in the
vicinity of the molecule of interest. The determination of the molecule is caused by the further oxidation by hydrogen peroxide,
producing a triplet carbonyl that emits light when it decays to the singlet carbonyl. Enhanced chemiluminescense allows for
femtomolar concentrations of proteins to be detected (218,219). Crumbliss et al. (220) were the first to show the enzymatic
activity on colloidal Au sols. This biomolecule nanoparticle conjugate was well retained when deposited onto conducting
matrixes. As a result of these findings, gold nanoparticles (AuNPs) have been extensively used in the fabrication of HRP
biosensors to detect hydrogen peroxide. Chen and coworkers (221) were the first to attach gold nanoparticles onto gold
electrodes. They investigated H2O2 reduction in the presence of catechol as a mediator using HRP-labeled Au colloids attached to
a cysteamine monolayer on gold.
Other enzymes such as peroxidise or luciferase deliver products for luminal or luciferin luminescence, allowing optical detection.
Among the commercially available enzymes, oxidases are the most often used in biosensor applications due to their stability and
their ability to work without coenzymes or cofactors (222). Oxidases are also capable of transferring hydrogen from a substrate to
molecular oxygen and can be divided into two groups based on the product formed during catalysis. Water producing (copper
containing) and hydrogen peroxide producing oxidases (222) are widely studied. Copper containing oxidase enzyme lactase is
another example of a widely used enzyme in biosensor applications. The stability of the enzyme biosensor relies on how well the
enzyme bonds to the sensor surface and remains there during use (223). Enzyme immobilization is generally a technique
specifically designed to restrict this freedom of movement. There are five principal methods (224) for immobilizing enzymes onto
the electrode surface: physical adsorption, covalent binding, encapsulation, entrapment, and cross-linking. Even though a wide
range of enzymes are commercially available with well-defined assay characteristics, they are expensive and the cost of extracting,
isolating, and purifying them can be excessive. Enzymes also tend to lose activity and degrade after a relatively short period of time
(225). However, improvements based on enzyme catalyzed microbial biochemical pathways have been made, and the existence of
species specific metabolic pathways have been discovered (206). Specific key marker enzymes participate in such metabolism;
therefore, the direct assaying of marker enzymes can be a viable alternative for indicator organism detection. In short, enzyme
mediated reactions can be rapid and time saving.

13.02.8.2 Nucleic Acids


Biological functions such as genetic information cannot be obtained from the chemical structure of a single strand of DNA.
However, the publication in the journal Nature in 1953 by James Watson and Francis Crick showed that DNA adopts a double
stranded structure (duplex) and the mechanism of DNA replication became obvious. This work was dependent on the research of
Erwin Chargaff, who formulated a rule that molar ratio of purines to pyrimidines is always 1:1 (226). Regardless of the sequence of
nucleotides in the DNA strand, the number of moles of adenine in DNA was always equal to that of thymine, and the same was true
for guanine and cytosine (227). This meant the formation of purine–pyrimidine pairs: A only pairs with T, and G with C
(Figure 30).
The formation of these base pairs provided evidence that the two strands of the double helix are complementary in sequence.
From this research, Watson and Crick were able to deduce that the sequence of one strand of DNA precisely defines the sequence of
the other; the two strands are said to be complementary. The two separate strands are antiparallel, with the 50 -end of one strand next
to the 30 -end of the other. They coil around each other to form a right-handed double helix. At the center of the helix structure lies
the hydrophobic base pairs and the sugars, which are planar and stacked upon each other with the specific A-T and G-C base pairing.
Selective detection by hybrid receptors is based on the detection of a unique sequence of nucleic acid bases through hybrid-
ization. This property of base pairing gives the ability of one single strand to recognize its complementary strand to form a duplex
(228,229). A nucleic acid (NA) biosensor employs oligonucleotides as the sensing element, with a known sequence of bases, or
a complex structure of DNA or RNA. Most NA biosensors are based upon highly specific hybridization of well-defined sequences of
single strands of DNA/RNA as biological receptors onto a solid matrix (230).
According to Keller and Manak (231) in 1989, a probe is a molecule that interacts with a specific target and has a definite means
of detection after such an interaction. Radioactive (32P), fluorometric, or enzymatic labeling of the nucleic acid probes are
commonly used for the DNA receptor probes (231,232). The complementarity between NA base sequences and the existence of
some unique sequences in each species form the basis of NA methods (233). Probes can be used either in the ‘short’ synthetic form
Review of Physical Principles of Sensing and Types of Sensing Materials 37

Figure 30 Hydrogen bonding between the Watson-Crick base pairs in DNA. Reproduced from Polsky, R.; Gill, R.; Kaganovsky, L.; Willner, I. Nucleic
Acid-Functionalized Pt Nanoparticles: Catalytic Labels for the Amplified Electrochemical Detection of Biomolecules. Anal. Chem. 2006, 78, 2268–2271.

or in the ‘long’ cloned form. They have been used to detect cancers, viral infections, and genetic diseases (234) A DNA probe is
hybridized to DNA or RNA from an unknown sample. If the probe combines with the unknown nucleic acid because of pairing of
complementary base recognition, detection and identification are then possible. The DNA-based analytical method is the ideal
method for detecting genetic modifications and is the most responsive approach for detecting microorganisms (222). Krull et al.
(235) were the first to report an optical DNA biosensor. The assay was completely built up on an optical fiber. The transducer surface
was first modified with long chain aliphatic spacers, and the capture probe was then synthesized in an automated solid phase
synthesis technique. The detection of the dsDNA at the fiber surface was achieved by staining the DNA duplex with ethidium
bromide, an intercalating dye. Millan et al. (236) then reported a novel DNA biosensor that involved the covalent immobilization
of DNA onto glassy carbon electrodes. In 1997, Henke et al. (237) covalently immobilized DNA probes on an optical fiber using an
amine terminal moiety. Electrochemical impedance spectroscopy has recently been employed to detect nucleotides that were
covalently immobilized onto aldehyde activated silanized alumina membranes (238).
Genosensors result from integrations of sequence specific probes, typically a short synthetic oligonucleotide and a signal
transducer. A probe immobilized onto the transducer surface acts as a biorecognition molecule and recognizes either the target DNA
or RNA (239–243). The detection of DNA hybridization on microarrays usually involves detecting the signal generated by the
binding of a reporter probe (fluorescent, chemiluminescense, colorimetric) to the target DNA sequence (244). Performing a typical
nucleic acid based method for indicator organism detection involves the fixation of test samples on membranes and then the
introduction of a probe. When samples contain whole cells, they are permeabilized to expose the nucleic acids. Incubation is then
carried out to anneal the probe and target nucleic acids. Nonspecific binding of the probes is then removed followed by tracing of
the probes on the sample using different techniques that include autoradiography, fluorescence, or color monitoring (233,245). The
presence of the target sequence is confirmed by detection of the signal.
Typically, it takes about 20 h or more to finish a traditional filter hybridization procedure. Therefore, the detection of target
biochemical molecules by means of traditional analytical methods can often be thought of as slow and time consuming (246). In
general, target DNA occurs at low concentrations and sometimes in the presence of interfering compounds in the environment.
Saiki et al. (247) proposed a novel amplification method called polymerase chain reaction (PCR), which has widespread
applications in clinical analysis, environmental microbiology, and forensics. Selective amplification of the DNA of interest is
enabled by use of primers for the target sequence, after which the PCR products are separated by electrophoresis and hybridized
with the probe (231).
However, DNA biosensors can give results within an hour (248–250). The application of electrochemical methods provides
significant advantages over optical biosensors (251–253). Specifically, the advantages of electrochemical biosensors include low
cost, reusability, speed (220), as well as the relatively high stability and environmental insensitivity of electroactive labels; wide
dynamic range and the wide range of available electrodes (220). Furthermore, biosensors have been designed to exhibit high
selectivity based on the inherent properties of the immobilized biocomponents. The biological recognition element that is attached
to the transducer surface selectively interacts with the analyte present in a complex sample matrix, thus eliminating the requirement
for separation and/or preconcentration of the analyte. This means that DNA-based biosensors can be simple and cost effective while
providing fast results accurately.
The most common way of attaching DNA capture probes is to use thiol modified probe DNA that interacts with the gold surface
through the sulfur atom of the thiol group. Almost 30 years ago, the self-assembly of sulfur-containing molecules was characterized
by Nuzzo and Allara (254). Disulphides, sulphides, and thiols coordinate very strongly onto a variety of metals, e.g., gold, silver,
platinum, or copper. Nonetheless, gold is the most favored because it is reasonably inert. The assumed reaction (eqn [25]) between
a thiolate compound and a gold substrate is:
RðCH2 Þn SH þ Au/RðCH2 Þn SAu þ e þ 1=2H2 [25]
38 Review of Physical Principles of Sensing and Types of Sensing Materials

The amount of probe DNA immobilization on the electrode surface is an important factor because it directly influences
the sensitivity of a DNA biosensor (255). General features of the direct chemisorption of self-assembled monolayers (SAMs)
of thiol-modified DNA probes onto gold surfaces have been summarized in several papers and reviews (256–260) and are briefly
covered in the following section. Herne et al. (261) gave a comprehensive account of the structure of surface attached probes on
gold disc electrodes and the impact of the probes distribution and orientation on the efficiency of hybridization reactions. Their
studies emphasized the following important points that are the basis of today’s platform for sequence specific DNA biosensors.
Gold surfaces have been used in biosensor applications for a number of years, since they can be easily modified with self-
assembled monolayers of u functionalized thiols (262). Single or double stranded DNA (for certain lengths only) modified with
a thiol is similar to long chain n-alkane thiols. They form uniform and closely packed monolayers on the gold electrode surface.
Thiolation of ssDNA is most commonly used for the attachment of capture DNA followed by the packed self-assembly onto
a surface such as gold, silver, or copper (263). The sulfur donor atoms strongly chemisorb onto the gold (typical bond energy is
145–188 KJ mol1) and the self-assembly process then begins (264). Each of the sulfur atoms is coordinative bound to
three Au atoms (the sulfur is sp3 hybridized). The thiol chain is therefore tilted with respect to the gold surface by approxi-
mately 20 to 40 (262).
Previous reports (265) have discussed the effect of the length and the presence of an anchoring group on the assembly of
DNA single strands on gold surfaces. The immobilization of single-stranded oligonucleotides containing a 50 -hexanethiol anchoring
group for lengths from 8 to 48 bases has been discussed. It has been shown that the oligonucleotide immobilization is enhanced in
the presence of a thiol group. However, the enhancement is reduced for longer DNA strand lengths (265). The surface coverage
begins to decrease with DNA strands longer than 24 bases, as a less ordered arrangement of the DNA is observed. The great
advantage of thiol tethered DNA is that the surface coverage of the immobilized DNA can be controlled by adjusting the immo-
bilization time and the concentration of the short chain alkyl thiols (265). Typically, gold substrates with DNA monolayers are
treated with a dilute ‘backfiller, that is 6-mercapto-1-hexanol (MCH) to improve the quality of the sensor. Coimmobilization of this
OH-terminated thiol ensures the desired orientation of the immobilized ssDNA at the electrode surface, therefore increasing
hybridization efficiency. The strong affinity of the thiol group of the MCH for gold results in the displacement of less strongly,
nonspecifically adsorbed DNA bases. The negatively charged DNA backbone is repelled by the net negative dipole of the alcohol
terminus, thus assisting the DNA strands into the solution (265).

13.02.8.3 Antibodies
Antibodies can be classified into five classes of glycoproteins known as immunoglobulins (Ig): IgG, IgM, IgA, IgE, and IgD (266).
Immunoglobin G (IgG) is the most commonly used antibody in immunoanalytical techniques (267). IgG is a Y-shaped molecule
(Figure 31) based upon two distinct types of polypeptide chains where the two chains is 450 amino acids for the H-chain

Figure 31 A schematic illustrating the “Y”-shaped structure of an antibody. Reproduced from Fowler, J. M.; Wong, D. K. Y.; Brian Halsall, H.;
Heineman, W. R.. Electrochemical Sensors, Biosensors and Their Biomedical Applications In Recent Developments in Electrochemical Immunoassays and
Immunosensors; Academic Press: San Diego, 2008; pp 115–143, (chapter 5).
Review of Physical Principles of Sensing and Types of Sensing Materials 39

(w55 000 Da) and 212 amino acids for the L-chain (w25 000 Da). There are four polypeptide chains, two identical heavy (H)
chains, and two identical light (L) chains (268). In each IgG molecule, the two light and two heavy chains are held together by
interchaindisulfide linkages. The name ‘immunoglobulin’ comes from the globular structure of the protein that is a result of these
interchain bonds (269).
Both heavy and light chains are divided into two subdomains: constant (C) and variable (V). These subdomains are classified
according to the variability of their amino acid sequence. The light chains have a single variable domain (VL) and a single constant
domain (CL). The heavy chain consists of a single variable domain (VH) and three constant domains (CH1, CH2, CH3). In general,
the antibody molecule may be divided into two main fragments: the nonantigen-binding fragment and the antigen-binding
fragment (268). The base of the Y shape (Figure 31) is called the Fc fragment (fragment that crystallizes) and is formed by the
association of the two CH2 and the two CH3 domains. The fragment containing antigen-binding site (FAB) is found on each arm of
the Y shape and is formed by association of CH1 with CL and VH with VL (268).

13.02.8.3.1 Antibody–Antigen Interaction


Antibodies do not possess the same catalytic activity as enzymes; however, antibodies have been used as immunoassays for many
years. They have exhibited higher binding power than enzymes have to their substrates, making them ultrasensitive recognition
elements. This is due to the specificity of antigen–antibody interaction (269). The variable domains in both chain types are the
most important regions with regard to the antibody–antigen binding interaction (270). The interaction between antibodies and
antigens are reversible and noncovalent. The interactions involve hydrogen bonds, van der Waals forces, ionic coulombic inter-
actions, and hydrophobic bonds. The specificity of an antibody toward the binding site (called the epitope) of its antigen is
a function of its amino acid sequence. The particular part of the antigen molecule that acts as an epitope can vary from one antibody
to another for exactly the same molecule. The binding affinity between an antibody (Ab) and an antigen (Ag) can generally be
described by the equilibrium expression (eqn [26]) (270).

½Ab  Ag
K¼ [26]
½Ab ½Ag

where K is the equilibrium constant for the interaction, Ab represents antibody, and Ag antigen. Typical values of K range from
106 to 1012 L mol1 (267). The physiochemical change induced by the antigen–antibody interaction cannot be detected elec-
trochemically. Therefore, the antibody or antigen is labeled with fluorescent compounds and electrochemically active species.
The antibodies used are specific to the epitope on the antigen. A detectable signal can be obtained in the presence of the epitope,
even if the host is alive or dead. The disadvantage of this system is that a positive result can be produced even from nonviable
pathogens (225).
The highly specific ability of antibodies to form complexes with corresponding antigens leads to great sensitivity. Mammals
produce antibodies to fight intruders such as microorganisms, viruses, bacteria, and parasites (271,272). This is called the immune
system, and its main function is to specifically recognize and eliminate pathogens. The most important analytical systems of
antibodies are immunoassays and immunosensors (271). Immunoassays can be distinguished into two large groups, i.e., the
‘limited reagent’ methods (competitive immunoassays) and the ‘reagent excess’ methods (noncompetitive immunoassays). In
competitive assays, the analyte and the labeled analyte/tracer are added to a limited amount of antianalyte antibodies. The analyte
and tracer will then ‘compete’ for the binding sites on the antibodies. After incubation of immunoreagents the antibody-bound or
nonbound fraction of tracer is determined; this will reflect the binding ratio between analyte and tracer. Noncompetitive assays are
based on adding an overload of antibodies to the sample, which ensures that all analyte is bound. The immunocomplex is then
determined, and the total concentration of analyte in the sample is found. The overall sensitivity in competitive assays is discovered
by the affinity constant of analyte and tracer toward the antibody, while in noncompetitive assays, nonspecific binding of labeled
immunoreagents will be limiting (273).
Competitive and noncompetitive systems can be further divided into homogeneous or heterogeneous assays. Homogeneous
assays have both the bound and free fraction of immunoreagents present within the system when the detection is performed. In
contrast, heterogeneous assays separate the bound fraction from the free fraction before detection, and in this case, either of the
two portions can be determined (273). The separation needed in heterogeneous immunoassays is very often accomplished by
using surface-immobilized antibodies or analytes. Thus, after formation of the immunocomplexes, the free fraction can be
removed while keeping the bound fraction on the support material. Enzyme linked immunosorbent assay (ELISA), immuno-
fluorescent assay (IFA), and immunoblotting are just a few of the examples of immunobased methods (266,274). Each of these
techniques take advantage of the specificity of antibody–antigen interactions, and selective detection of indicator organisms is
possible through targeting of species-specific proteins. Antibodies for these targets can also be conjugated to enzymes that will
catalyze the breakdown of chromogenic or fluorogenic substrates. Horseradish peroxidase, alkaline phosphatase, urease, and b-D-
galactosidase (GAL) are a number of enzymes that can be used in conjunction with the antibodies. There are various forms of
ELISA reported in today’s literature (275) such as direct (direct labeled antibody (Ab)/antigen (Ag)), indirect, sandwich (direct/
direct), and competition (direct/indirect Ab competition and direct/indirect Ag competition) assays. E. Williams et al. (276) used
a biosensor that was amplified by an enzyme labeled antibody, as illustrated in Figure 32. Hybridization occurred between
a target DNA to the capture probe and an antigen labeled DNA sequence. Electrochemical detection was achieved when the
enzyme labeled antibody specific to the antigen is added. DNA hybridization in this manner is favored due to the high efficiency
40 Review of Physical Principles of Sensing and Types of Sensing Materials

Figure 32 Schematic representation of the DNA analysis based on an electrochemical streptavidin carbon-polymer biocomposite. (a) Electrode
modified with streptavidin. (b) One step immobilization/ hybridization procedure: target-DNA is hybridized with the capture probe and with the
digoxigenin modified probe. (c) Enzyme labeling based on the immunological reaction between the immobilized dsDNA-Dig with anti-Dig-HRP. (d)
Electrochemical detection of the enzyme (HRP) labeled dsDNA. Reproduced from Williams, E.; Pividori, M. I.; Merkoçi, A.; Forster, R. J.; Alegret, S. Rapid
Electrochemical Genosensor Assay Using a Streptavidin Carbon-Polymer Biocomposite Electrode. Biosens. Bioelectron. 2003, 19, 165–175.

and specificity of the electrochemical biosensor in the presence of a mixture of many different, noncomplementary, nucleic acids
(239,277,278).

13.02.9 Conclusion

This chapter has shown that as a subject of interest, sensing materials are very broad, and that each material possesses unique
properties that make it interesting and important materials for chemical sensing. A large number of various types of sensing
materials, both natural and man-made, have been thoroughly discussed in this chapter. Researchers are trying hard to clearly
understand sensor devices, sensor materials, and sensing analyte as well as alternative mechanisms of detection in many academic
areas. Recent results have indicated that many sensing materials have moved from the research stage of sensors to potentially
commercial phases. In this chapter, we have tried to include the widest possible number of sensing materials and to evaluate their
general background information as well as their real advantages and potential applications in today’s sensors. Such information
might result from more collaboration among interdisciplinary researchers and engineers in different fields, to work together on
new sensing materials to develop functional physical and chemical/biochemical sensors.

See also: Review of Recent Developments in Sensing Materials.

References

1. Lion, K. S. Transducers: Problems and Prospects. IEEE Trans. Ind. Electron. Control Instrum. 1969, IECI-16, 2–5.
2. Moseley, P. T.; Crocker, A. J. Sensor Materials; Institute of Physics Publication: Bristol, PA, 1996; p 227.
3. Fraden, J., Ed. Handbook of Modern Sensors: Physics, Designs, and Applications; Springer: San Diego, CA, 2010; Vol. 4, p 663.
4. White, R. A. Sensor Classification Scheme. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 1987, 34, 124–126.
5. Cullity, B. D. Introduction to Magnetic Materials; John Wiley & Sons: New Jersey, 2009; p 568.
6. Bloxam, C. H. Metals: Their Properties and Treatment; Longmans, Green & Co: London, 2007; p 452.
7. Kittel, C., Ed., Introduction to Solid State Physics, 8th ed.; John Wiley & Sons: New York 2005; Vol. 8, p 704.
Review of Physical Principles of Sensing and Types of Sensing Materials 41

8. Ashcroft, N. W. Solid State Physics; Holt Rinehart & Winston: Philadelphia, 1976.
9. Kiejna, A. W. Metal Surface Electron Physics; Elsevier: Oxford, 1996; p 303.
10. Grattan, K. T. V.; Sun, T. Fiber Optic Sensor Technology: An Overview. Sens. Actuators, A 2000, 82, 40–61.
11. Paschotta, R. In Encyclopedia of Laser Physics and Technology; Wiley-VCH: Germany, 2008; Vol. 1, p 856.
12. Atashbar, M. Z. In General Approaches; Korotcenkov, I. G., Korotcenkov, G., Eds.; Momentum Press: New York, 2010; p 33.
13. Measures, R. In Structural Monitoring with Fiber Optic Technology; Academic Press: San Diego, CA, 2001; Vol. 1, p 716.
14. Bogue, R. Fibre Optic Sensors: A Review of Today’s Applications. Sens. Rev. 2011, 31, 304–309.
15. Leung, A.; Shankar, P. M.; Mutharasan, R. A Review of Fiber-Optic Biosensors. Sens. Actuators, B 2007, 125, 688–703.
16. Zubia, J.; Arrue, J. Plastic Optical Fibers: An Introduction to Their Technological Processes and Applications. Opt. Fiber Technol. 2001, 7, 101–140.
17. May, G. S.; Sze, S. M. Fundamentals of Semiconductor Fabrication; Wiley: New York, 2003; p 352.
18. Tyagi, M. Introduction to Semiconductor Materials and Devices; John Wiley & Sons: Canada, 1991.
19. Schroder, D. K. Semiconductor Materials and Device Characterization, 2nd ed.; John Wiley: New York; Chichester, 1998; p 648.
20. H.F., M. Transistor Engineering and Introduction to Integrated Semiconductor Circuits. Alvin B. Phillips; McGraw-Hill, New York, 1962, 366 pp., $12.00. Solid-State Electron.
1963, 6, 196–197.
21. Henri, A. An Introduction to Semiconductors. By W. Crawford Dunlap, Jr. 417 pages, diagrams, 6  9 in. New York, John Wiley & Sons, Inc., 1957. Price, $11.75. J. Franklin
Inst. 1957, 264, 251.
22. Orton, J. W. The Story of Semiconductors; Oxford University Press: Oxford, 2004; p 400.
23. Ghafouri-Shiraz, H., Ed. The Principles of Semiconductor Laser Diodes and Amplifiers: Analysis and Transmission Line Laser Modelling; 2002; p 450.
24. Sproul, A.; Green, M. Improved Value for the Silicon Intrinsic Carrier Concentration from 275-K to 375-K. J. Appl. Phys. 1991, 70, 846–854.
25. T.P., N. Defect Analysis in Organic Semiconductors. Mater. Sci. Semicond. Process. 2006, 9, 198–203.
26. Leontie, L.; Druta, I.; Danac, R.; Prelipceanu, M.; Rusu, G. I. Electrical Properties of Some New High Resistivity Organic Semiconductors in Thin Films. Prog. Org. Coat. 2005,
54, 175–181.
27. Farag, A. A. M.; Mansour, A. M.; Ammar, A. H.; Rafea, M. A. Characterization of Electrical and Optical Absorption of Organic Based Methyl Orange for Photovoltaic Application.
Synth. Met. 2011, 161, 2135–2143.
28. Celle, C.; Suspène, C.; Simonato, J.; Lenfant, S.; Ternisien, M.; Vuillaume, D. Self-Assembled Monolayers for Electrode Fabrication and Efficient Threshold Voltage Control of
Organic Transistors with Amorphous Semiconductor Layer. Org. Electron. 2009, 10, 119–126.
29. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconducting Metal Oxides as
Sensors for Environmentally Hazardous Gases. Sens. Actuators, B 2011, 160, 580–591.
30. G., K. Practical Aspects in Design of One-Electrode Semiconductor Gas Sensors: Status Report. Sens. Actuators, B 2007, 121, 664–678.
31. Hirata, M.; Sun, L. Characteristics of an Organic Semiconductor Polyaniline Film as a Sensor for NH3 Gas. Sens. Actuators, A 1994, 40, 159–163.
32. Marinelli, F.; Dell’Aquila, A.; Torsi, L.; Tey, J.; Suranna, G. P.; Mastrorilli, P.; Romanazzi, G.; Nobile, C. F.; Mhaisalkar, S. G.; Cioffi, N.; Palmisano, F. An Organic Field Effect
Transistor as a Selective NOx Sensor Operated at Room Temperature. Sens. Actuators, B 2009, 140, 445–450.
33. Takada, T.; Fukunaga, T.; Maekawa, T. New Method for Gas Identification Using a Single Semiconductor Sensor. Sens. Actuators, B 2000, 66, 22–24.
34. Zaitsev, V. B.; Panova, T. V. The Use of Vibronic Phenomena in Adsorption Phase for Developing Semiconductor Gas Sensors. Appl. Surf. Sci. 2000, 167, 184–190.
35. Chaniotakis, N.; Sofikiti, N. Novel Semiconductor Materials for the Development of Chemical Sensors and Biosensors: A Review. Anal. Chim. Acta 2008, 615, 1–9.
36. Gurlo, A.; Sahm, M.; Oprea, A.; Barsan, N.; Weimar, U. A p- to n-Transition on a-Fe2O3-Based Thick Film Sensors Studied by Conductance and Work Function Change
Measurements. Sens. Actuators, B 2004, 102, 291–298.
37. Bergstrom, P. L.; Patel, S. V.; Schwank, J. W.; Wise, K. D. A Micromachined Surface Work-Function Gas Sensor for Low-Pressure Oxygen Detection. Sens. Actuators, B 1997,
42, 195–204.
38. Schierbaum, K. D.; Weimar, U.; Göpel, W.; Kowalkowski, R. Conductance, Work Function and Catalytic Activity of SnO2-Based Gas Sensors. Sens. Actuators, B 1991, 3,
205–214.
39. Shenai, K. High-Power Robust Semiconductor Electronics Technologies in the New Millennium. Microelectron. J. 2001, 32, 397–408.
40. Fenukhin, A. V.; Kazanskii, A. G.; Kolosko, A. G.; Terukov, E. I.; Ziminov, A. V. Absorption Spectra of Organic Semiconductors in IR-Range Measured by Constant Photocurrent
Method. J. Non Cryst. Solids 2006, 352, 1668–1670.
41. Idrish, M. Optical Sensor Protector Using Wide-Bandgap Semiconductors. Optik – Int. J. Light Electron. Optics 2012, 123 (17), 1580–1582.
42. Riede, M.; Lüssem, B.; Leo, K. Organic Semiconductors. In Comprehensive Semiconductor Science and Technology; Bhattacharya, Pallab, Fornari, Roberto,
Kamimura, Hiroshi, Eds.; Elsevier: Amsterdam, 2011; pp 448–507 (chapter 4.13).
43. Liu, D.; Kamat, P. V. Electrochemical Rectification in CdSe þ TiO2 Coupled Semiconductor Films. J. Electroanal. Chem. 1993, 347, 451–456.
44. Elena, G. Linkers for Anchoring Sensitizers to Semiconductor Nanoparticles. Coord. Chem. Rev. 2004, 248, 1283–1297.
45. Gerald, J. M. Molecular Control of Photo-Induced Electron and Energy Transfer at Nanocrystalline Semiconductor Interfaces. J. Photochem. Photobiol. A 2003, 158, 119–124.
46. Heimer, T. A.; Meyer, G. J. Luminescence of Charge Transfer Sensitizers Anchored to Metal Oxide Nanoparticles. J. Lumin. 1996, 70, 468–478.
47. Gundlach, L.; Ernstorfer, R.; Willig, F. Ultrafast Interfacial Electron Transfer from the Excited State of Anchored Molecules into a Semiconductor. Prog. Surf. Sci. 2007, 82,
355–377.
48. Ismail, A. B. M.; Yoshinobu, T.; Iwasaki, H.; Sugihara, H.; Yukimasa, T.; Hirata, I.; Iwata, H. Investigation on Light-Addressable Potentiometric Sensor as a Possible
Cell–Semiconductor Hybrid. Biosens. Bioelectron. 2003, 18, 1509–1514.
49. Stein, B.; George, M.; Gaub, H. E.; Parak, W. J. Extracellular Measurements of Averaged Ionic Currents with the Light-Addressable Potentiometric Sensor [LAPS]. Sens.
Actuators, B 2004, 98, 299–304.
50. Wagner, T.; Schöning, M. J. Comprehensive Analytical Chemistry. In Light-Addressable Potentiometric Sensors [LAPS]: Recent Trends and Applications, Vol. 49; Elsevier:
Amsterdam, 2007; pp 87–128, (chapter 5).
51. Xu, G.; Ye, X.; Qin, L.; Xu, Y.; Li, Y.; Li, R.; Wang, P. Cell-Based Biosensors Based on Light-Addressable Potentiometric Sensors for Single Cell Monitoring. Biosens.
Bioelectron. 2005, 20, 1757–1763.
52. Chen, K.; Li, B.; Chen, Y. Silicon Nanowire Field-Effect Transistor-Based Biosensors for Biomedical Diagnosis and Cellular Recording Investigation. Nano Today 2011, 6,
131–154.
53. Hsu, C.; Lin, K.; Chen, H.; Chen, T.; Huang, C.; Chou, P.; Liu, R.; Liu, W. On a Heterostructure Field-Effect Transistor [HFET] Based Hydrogen Sensing System. Int. J. Hydrogen
Energy 2011, 36, 15906–15912.
54. Müller, R.; Smout, S.; Rolin, C.; Genoe, J.; Heremans, P. High Mobility Short-Channel p-Type Organic Transistors with Reduced Gold Content and Completely Gold-Free
Source/Drain Bottom Contacts. Org. Electron. 2011, 12, 1227–1235.
55. Sant, W.; Temple-Boyer, P.; Chanié, E.; Launay, J.; Martinez, A. On-Line Monitoring of Urea Using Enzymatic Field Effect Transistors. Sens. Actuators, B 2011, 160,
59–64.
56. Boscher, N. D.; Carmalt, C. J.; Palgrave, R. G.; Parkin, I. P. Atmospheric Pressure Chemical Vapour Deposition of SnSe and SnSe2 Thin Films on Glass. Thin Solid Films
2008, 516, 4750–4757.
57. Govindaraju, N.; Singh, R. N. Processing of Nanocrystalline Diamond Thin Films for Thermal Management of Wide-Bandgap Semiconductor Power Electronics. Mater. Sci. Eng.
B 2011, 176, 1058–1072.
42 Review of Physical Principles of Sensing and Types of Sensing Materials

58. Pearton, S. J.; Abernathy, C. R.; Thaler, G. T.; Frazier, R.; Ren, F.; Hebard, A. F.; Park, Y. D.; Norton, D. P.; Tang, W.; Stavola, M.; Zavada, J. M.; Wilson, R. G. Effects of
Defects and Doping on Wide Band Gap Ferromagnetic Semiconductors. Phys. B: Condens. Matter 2003, 340–342, 39–47.
59. Thomas, H. R.; Salaneck, W. R.; Duke, C. B.; Plummer, E. W.; Heeger, A. J.; Macdiarmid, A. G. Photoelectron-Spectra of Conducting Polymers-Molecularly Doped
Polyacetylenes. Polymer 1980, 21, 1238–1246.
60. Kaiser, A. B.; Park, Y. W. Conduction Mechanisms in Polyacetylene Nanofibres. Curr. Appl. Phys. 2002, 2, 33–37.
61. Kanazawa, K. K.; Diaz, A. F.; Geiss, R. H.; Gill, W. D.; Kwak, J. F.; Logan, J. A.; Rabolt, J. F.; Street, G. B. Organic Metals – Polypyrrole, a Stable Synthetic Metallic Polymer.
J. Chem. Soc., Chem. Commun. 1979, 854–855.
62. Diaz, A. F.; Lee, W. Y.; Logan, A.; Green, D. C. Chemical Modification of a Polypyrrole Electrode Surface. J. Electroanal. Chem. 1980, 108, 377–380.
63. Diaz, A. F.; Logan, J. A. Electroactive Polyaniline Films. J. Electroanal. Chem. 1980, 111, 111–114.
64. Diaz, A. F.; Castillo, J. I. A Polymer Electrode with Variable Conductivity – Polypyrrole. J. Chem. Soc., Chem. Commun. 1980, 397–398.
65. Buchner, W.; Garreau, R.; Roncali, J.; Lemaire, M. Synthesis of 3-[Polyfluoroalkyl] Thiophenes. J. Fluorine Chem. 1992, 59, 301–309.
66. Wallace, G. G.; Spinks, G. M.; Kane-Maguire, L. A. P. Conductive Electroactive Polymers: Intelligent Materials Systems, 2nd ed.; CRC press: Boca Raton, FL, 2002; p 224.
67. Baklanov, M., Green, M., Maex, K., Eds. Dielectric Films for Advanced Microelectronics; Wiley Sons: New York, 2007; p 486.
68. Ye, Z., Ed. Handbook of Dielectric, Piezoelectric and Ferroelectric Materials: Synthesis, Properties, and Applications; Woodhead Publishing Ltd: Abingdon, Cambridge, UK,
2008; p 1060.
69. Nalwa, H. S., Ed. Handbook of Low and High Dielectric Constant Materials and Their Applications; Academic Press: New York, 1999; p 569.
70. Nalwa, H. S., Ed. Handbook of Low and High Dielectric Constant Materials and Their Applications; Academic Press: New York, 1999; p 539.
71. Houssa, M. High-K Gate Dielectrics; IOP: Berkshire, 2004; p 601.
72. Penn, S. J.; Alford, N. M.; Templeton, A.; Wang, X.; Xu, M.; Reece, M.; Schrapel, K. Effect of Porosity and Grain Size on the Microwave Dielectric Properties of Sintered
Alumina. J. Am. Ceram. Soc. 1997, 80, 1885–1888.
73. Penn, S.; Alford, N. Ceramic Dielectrics for Microwave Applications. In Handbook of Low and High Dielectric Constant Materials and Their Applications; Nalwa, Hari Singh, Ed.;
Academic Press: Burlington, 1999; pp 493–532 (chapter 10).
74. Von Hippel, A. R. Dielectric Materials and Applications; Wiley: New York, 1954; p 438.
75. Bunde, R. L.; Jarvi, E. J.; Rosentreter, J. J. Piezoelectric Quartz Crystal Biosensors. Talanta 1998, 46, 1223–1236.
76. Minase, J.; Lu, T.; Cazzolato, B.; Grainger, S. A Review, Supported by Experimental Results, of Voltage, Charge and Capacitor Insertion Method for Driving Piezoelectric
Actuators. Precis. Eng. 2010, 34, 692–700.
77. Ko, S. C.; Kim, Y. C.; Lee, S. S.; Choi, S. H.; Kim, S. R. Micromachined Piezoelectric Membrane Acoustic Device. Sens. Actuators, A 2003, 103, 130–134.
78. Lee, W. S.; Lee, S. S. Piezoelectric Microphone Built on Circular Diaphragm. Sens. Actuators, A 2008, 144, 367–373.
79. Scheeper, P. R.; van der Donk, A. G. H.; Olthuis, W.; Bergveld, P. A Review of Silicon Microphones. Sens. Actuators, A 1994, 44, 1–11.
80. Xu, J.; Dapino, M. J.; Gallego-Perez, D.; Hansford, D. Microphone Based on Polyvinylidene Fluoride [PVDF] Micro-Pillars and Patterned Electrodes. Sens. Actuators, A 2009,
153, 24–32.
81. Chao, X.; Yang, Z.; Dong, M.; Li, G. Fabrication, Characteristics and Temperature Stability of Pb[Mg1/3Nb2/3]O3–Pb[Sb1/3Nb2/3]O3–Pb[Ni1/3Nb2/3]O3–Pb[Zr, Ti]O3 Piezoelectric
Actuators. Sens. Actuators, A 2009, 151, 71–76.
82. Ikeda, H.; Morita, T. High-Precision Positioning Using a Self-Sensing Piezoelectric Actuator Control with a Differential Detection Method. Sens. Actuators, A 2011, 170, 147–155.
83. Karpelson, M.; Wei, G.; Wood, R. J. Driving High Voltage Piezoelectric Actuators in Microrobotic Applications. Sens. Actuators, A 2012, 176, 78–89.
84. Watson, B.; Friend, J.; Yeo, L. Piezoelectric Ultrasonic Micro/Milli-Scale Actuators. Sens. Actuators, A 2009, 152, 219–233.
85. Li, X.; Kan, E. C. A Wireless Low-Range Pressure Sensor Based on P[VDF-TrFE] Piezoelectric Resonance. Sens. Actuators, A 2010, 163, 457–463.
86. van den Ende, D. A.; Groen, W. A.; van der Zwaag, S. Development of Temperature Stable Charge Based Piezoelectric Composite Quasi-Static Pressure Sensors. Sens.
Actuators, A 2010, 163, 25–31.
87. Yoo, J.; Hong, J.; Lee, H.; Jeong, Y.; Lee, B.; Song, H.; Kwon, J. Piezoelectric and Dielectric Properties of La2O3 Added Bi[Na, K]TiO3–SrTiO3 Ceramics for Pressure Sensor
Application. Sens. Actuators, A 2006, 126, 41–47.
88. Lewin, P. A.; Lypacewicz, G.; Bautista, R.; Devaraju, V. Sensitivity of Ultrasonic Hydrophone Probes below 1 MHz. Ultrasonics 2000, 38, 135–139.
89. Takashima, K.; Horie, S.; Mukai, T.; Ishida, K.; Matsushige, K. Piezoelectric Properties of Vinylidene Fluoride Oligomer for Use in Medical Tactile Sensor Applications. Sens.
Actuators, A 2008, 144, 90–96.
90. Oh, S. R.; Yao, K.; Chow, C. L.; Tay, F. E. H. Residual Stress in Piezoelectric Poly[Vinylidene-Fluoride-Co-Trifluoroethylene] Thin Films Deposited on Silicon Substrates. Thin
Solid Films 2010, 519, 1441–1444.
91. Xiao, Y.; Liu, Y.; Borg, G.; Li, C. M. Design of a Novel Disposable Piezoelectric Co-polymer Diaphragm Based Biosensor Unit. Mater. Sci. Eng., C 2011, 31, 95–98.
92. Coster, H. G. L.; Farahani, T. D.; Chilcott, T. C. Production and Characterization of Piezo-Electric Membranes. Desalination 2011, 283, 52–57.
93. Kochervinskii, V. V.; Glukhov, V. A.; Kuznetsova, S. Y. Investigation of the Uniaxial Stretching of Vinylidene Fluoride-Tetrafluoroethylene Copolymer Films by an Acoustic
Method. Polym. Sci. U.S.S.R 1987, 29, 1686–1694.
94. Weber, N.; Lee, Y.; Shanmugasundaram, S.; Jaffe, M.; Arinzeh, T. L. Characterization and In Vitro Cytocompatibility of Piezoelectric Electrospun Scaffolds. Acta Biomater.
2010, 6, 3550–3556.
95. Eastman, A.; Kimber, M.; Hirata, A.; Kamitani, G. Thermal Analysis of a Low Flow Piezoelectric Air Pump. Int. J. Heat Mass Transfer 2012, 55, 2461–2471.
96. Manganiello, L.; Ríos, A.; Valcárcel, M.; Ligero, A.; Tena, T. Automatic Determination of Fat in Milk by Use of a Flow Injection System with a Piezoelectric Detector. Anal. Chim.
Acta 2000, 406, 309–315.
97. Zougagh, M.; Ríos, A.; Valcárcel, M. Direct Determination of Total Carbonate Salts in Soil Samples by Continuous-Flow Piezoelectric Detection. Talanta 2005, 65, 29–35.
98. Cueff, M.; Defaÿ, E.; Le Rhun, G.; Rey, P.; Perruchot, F.; Suhm, A.; Aïd, M. Integrated Metallic Gauge in a Piezoelectric Cantilever. Sens. Actuators, A 2011, 172, 148–153.
99. Benes, E.; Gröschl, M.; Burger, W.; Schmid, M. Sensors Based on Piezoelectric Resonators. Sens. Actuators, A 1995, 48, 1–21.
100. Choi, S.; Lee, H.; Moon, W. A Micro-Machined Piezoelectric Hydrophone with Hydrostatically Balanced Air Backing. Sens. Actuators, A 2010, 158, 60–71.
101. Cui, C.; Baughman, R. H.; Iqbal, Z.; Kazmar, T. R.; Dahlstrom, D. K. Improved Piezoelectrics for Hydrophone Applications Based on Calcium-Modified Lead Titanate/Poly
[Vinylidene Fluoride] Composites. Sens. Actuators, A 1998, 65, 76–85.
102. Lau, S. T.; Kwok, K. W.; Chan, H. L. W.; Choy, C. L. Piezoelectric Composite Hydrophone Array. Sens. Actuators, A 2002, 96, 14–20.
103. Kim, K.; Zhang, S.; Salazar, G.; Jiang, X. Design, Fabrication and Characterization of High Temperature Piezoelectric Vibration Sensor Using YCOB Crystals. Sens. Actuators, A
2012, 178, 40–48.
104. Nemirovsky, Y.; Nemirovsky, A.; Muralt, P.; Setter, N. Design of Novel Thin-Film Piezoelectric Accelerometer. Sens. Actuators, A 1996, 56, 239–249.
105. Wlodkowski, P. A.; Deng, K.; Kahn, M. The Development of High-Sensitivity, Low-Noise Accelerometers Utilizing Single Crystal Piezoelectric Materials. Sens. Actuators, A
2001, 90, 125–131.
106. Xie, H.; Fedder, G. K.; Sulouff, R. E. Comprehensive Microsystems. In Accelerometers; Zappe, Hans, Ed.; Elsevier: Oxford, 2008; pp 135–180 (chapter 2.05).
107. Yu, H. G.; Zou, L.; Deng, K.; Wolf, R.; Tadigadapa, S.; Trolier-McKinstry, S. Lead Zirconate Titanate MEMS Accelerometer Using Interdigitated Electrodes. Sens. Actuators, A
2003, 107, 26–35.
108. Jalili, N.; Wagner, J.; Dadfarnia, M. A Piezoelectric Driven Ratchet Actuator Mechanism with Application to Automotive Engine Valves. Mechatronics 2003, 13, 933–956.
109. Jung, S.; Kim, S. Improvement of Scanning Accuracy of PZT Piezoelectric Actuators by Feed-Forward Model-Reference Control. Precis. Eng. 1994, 16, 49–55.
Review of Physical Principles of Sensing and Types of Sensing Materials 43

110. Juuti, J.; Kordás, K.; Lonnakko, R.; Moilanen, V.; Leppävuori, S. Mechanically Amplified Large Displacement Piezoelectric Actuators. Sens. Actuators, A 2005, 120,
225–231.
111. Szufnarowski, F.; Schneider, A. Force Control of a Piezoelectric Actuator Based on a Statistical System Model and Dynamic Compensation. Mech. Mach. Theor. 2011, 46,
1507–1521.
112. Sinclair, I. R. Sound, Infrasound and Ultrasound; Sensors and Transducers, 3rd ed.; Newnes: Oxford, 2001; pp 116–141, (chapter 5).
113. Wilson, S. A.; Jourdain, R. P. J.; Zhang, Q.; Dorey, R. A.; Bowen, C. R.; Willander, M.; Wahab, Q. U.; Willander, M.; Al-hilli, S. M.; Nur, O.; Quandt, E.; Johansson, C.;
Pagounis, E.; Kohl, M.; Matovic, J.; Samel, B.; van der Wijngaart, W.; Jager, E. W. H.; Carlsson, D.; Djinovic, Z.; Wegener, M.; Moldovan, C.; Iosub, R.; Abad, E.;
Wendlandt, M.; Rusu, C.; Persson, K. New Materials for Micro-Scale Sensors and Actuators: An Engineering Review. Mater. Sci. Eng. R: Rep. 2007, 56, 1–129.
114. Fukumoto, A.; Kawabuchi, M.; Sato, J. Design of Ultrasound Transducers Using New Piezoelectric Ceramic Materials. Ultrasound Med. Biol. 1981, 7, 275–284.
115. Kimoto, A.; Maddumapatabendi, J. D. A Proposal of Multi-Imaging System of Electrical and Ultrasonic Properties. Measurement 2011, 44, 2000–2007.
116. Batagiannis, A.; Wübbenhorst, M.; Hulliger, J. Piezo- and Pyroelectric Microscopy. Curr. Opin. Solid State Mater. Sci. 2010, 14, 107–115.
117. Whatmore, R. W. Pyroelectric Arrays: Ceramics and Thin Films. J. Electroceram. 2004, 13, 139–147.
118. Guggilla, P.; Batra, A. K.; Currie, J. R.; Aggarwal, M. D.; Alim, M. A.; Lal, R. B. Pyroelectric Ceramics for Infrared Detection Applications. Mater. Lett. 2006, 60, 1937–1942.
119. Chong, N.; Chan, H. L. W.; Choy, C. L. Pyroelectric Sensor Array for In-line Monitoring of Infrared Laser. Sens. Actuators A 2002, 96, 231–238.
120. Schreiter, M.; Gabl, R.; Lerchner, J.; Hohlfeld, C.; Delan, A.; Wolf, G.; Blüher, A.; Katzschner, B.; Mertig, M.; Pompe, W. Functionalized Pyroelectric Sensors for Gas Detection.
Sens. Actuators B 2006, 119, 255–261.
121. Chang, C. C.; Tang, C. S. An Integrated Pyroelectric Infrared Sensor with a PZT Thin Film. Sens. Actuators A 1998, 65, 171–174.
122. Garcia, M. L. In Interior Intrusion Sensors; Design and Evaluation of Physical Protection Systems, 2nd ed.; Butterworth-Heinemann: Boston, 2008; pp 101–125, (chapter 7).
123. Amon, F.; Pearson, C. Thermal Imaging in Firefighting and Thermography Applications. In Experimental Methods in Physical Sciences, Vol. 43; Academic Press, 2010;
pp 279–331, (chapter 5).
124. Sinclair, I. R. In Temperature Sensors and Thermal Transducers; Sensors and Transducers, 3rd ed.; Newnes: Oxford, 2001; pp 87–115, (chapter 4).
125. Kohli, M.; Huang, Y.; Maeder, T.; Wuethrich, C.; Bell, A.; Muralt, P.; Setter, N.; Ryser, P.; Forster, M. Processing and Properties of Thin Film Pyroelectric Devices.
Microelectron. Eng. 1995, 29, 93–96.
126. Hashimoto, K.; Kawaguchi, C.; Matsueda, S.; Morinaka, K.; Yoshiike, N. People-Counting System Using Multisensing Application. Sens. Actuators, A 1998, 66, 50–55.
127. Hashimoto, K.; Yoshinomoto, M.; Matsueda, S.; Morinaka, K.; Yoshiike, N. Development of People-Counting System with Human-Information Sensor Using Multi-Element
Pyroelectric Infrared Array Detector. Sens. Actuators A 1997, 58, 165–171.
128. Querner, Y.; Norkus, V.; Gerlach, G. High-Sensitive Pyroelectric Detectors with Internal Thermal Amplification. Sens. Actuators, A 2011, 172, 169–174.
129. Lee, M. H.; Guo, R.; Bhalla, A. S. Pyroelectric Sensors. J. Electroceram. 1998, 2, 229–242.
130. Fuentes-Cobas, L. E.; Matutes-Aquino, J. A.; Fuentes-Montero, M. E. Magnetoelectricity. In Handbook of Magnetic Mater, Vol. 19; Elsevier, 2011; pp 129–229, (chapter 3).
131. Samara, G. A. In Ferroelectricity RevisiteddAdvances in Materials and Physics; Solid State Physics; Academic Press, Vol. 56, pp 239–458.
132. Kirova, N.; Brazovskii, S. Ferroelectricity: From Organic Conductors to Conducting Polymers. Phys. B: Condens. Matter 2009, 404, 382–384.
133. Chung, C.; Chang, Y.; Chang, Y.; Chen, G. High Dielectric Permittivity in Ca1xBixTi1xCrxO3 Ferroelectric Perovskite Ceramics. J. Alloys Compd. 2004, 385, 298–303.
134. Gan, B. K.; Yao, K. Structure and Enhanced Properties of Perovskite Ferroelectric PNN–PZN–PMN–PZ–PT Ceramics by Ni and Mg Doping. Ceram. Int. 2009, 35, 2061–2067.
135. Uratani, Y.; Shishidou, T.; Ishii, F.; Oguchi, T. First-Principles Exploration of Ferromagnetic and Ferroelectric Double-Perovskite Transition-Metal Oxides. Phys. B: Condens.
Matter 2006, 383, 9–12.
136. Stennett, M. C.; Miles, G. C.; Sharman, J.; Reaney, I. M.; West, A. R. A New Family of Ferroelectric Tetragonal Tungsten Bronze Phases, Ba2MTi2X3O15. J. Eur. Ceram. Soc.
2005, 25, 2471–2475.
137. Kingon, A. I.; Streiffer, S. K. Ferroelectric Films and Devices. Curr. Opin. Solid State Mater. Sci. 1999, 4, 39–44.
138. Teixeira, Z.; Otubo, L.; Gouveia, R. F.; Alves, O. L. Preparation and Characterization of Powders and Thin Films of Bi2AlNbO7 and Bi2InNbO7 Pyrochlore Oxides. Mater. Chem.
Phys. 2010, 124, 552–557.
139. Chon, U.; Jang, H. M.; Shin, N. S.; Kim, J. S.; Ahn, D. C.; Kim, Y. S.; No, K. Gd-Substituted Bismuth Titanate Film Capacitors Having Ferroelectric Reliability and Large Non-
Volatile Charges. Phys. B: Condens. Matter 2007, 388, 190–194.
140. Duk, A.; Schwarzkopf, J.; Kwasniewski, A.; Schmidbauer, M.; Fornari, R. Impact of the Crystallographic Structure of Epitaxially Grown Strained Sodium–Bismuth–Titanate Thin
Films on Local Piezo- and Ferroelectric Properties. Mater. Res. Bull. August 2012, 47 (8), 2056–2061.
141. Kim, J.; Kim, J. K.; Heo, S.; Lee, H. S. Ferroelectric Properties of Sol-Gel Prepared La- and Nd-Substituted, and Nb-co-Substituted Bismuth Titanate Using Polymeric Additives.
Thin Solid Films 2006, 503, 60–63.
142. Pardo, L.; Ricote, J.; Algueró, M.; Calzada, M. L. Handbook of Low and High Dielectric Constant Materials and Their Applications. In Ferroelectric Materials Based on Lead
Titanate; Nalwa, Hari Singh, Ed.; Academic Press: Burlington, 1999; pp 457–499 (chapter 10).
143. Yoon, J.; Kwon Song, T. Handbook of Thin Films. In Fabrication and Characterization of Ferroelectric Oxide Thin Films; Nalwa, Hari Singh, Ed.; Academic Press: Burlington,
2002; pp 309–367 (chapter 5).
144. Liu, S. Y.; Chua, L.; Tan, K. C.; Valavan, S. E. Novel Ferroelectric Capacitor for Non-Volatile Memory Storage and Biomedical Tactile Sensor Applications. Thin Solid Films
2010, 518, e152–e155.
145. Ng, T. N.; Russo, B.; Krusor, B.; Kist, R.; Arias, A. C. Organic Inkjet-Patterned Memory Array Based on Ferroelectric Field-Effect Transistors. Org. Electron. 2011, 12,
2012–2018.
146. Kim, R. H.; Kang, S. J.; Bae, I.; Choi, Y. S.; Park, Y. J.; Park, C. Thin Ferroelectric Poly[Vinylidene Fluoride-Chlorotrifluoro Ethylene] Films for Thermal History Independent Non-
Volatile Polymer Memory. Org. Electron. 2012, 13, 491–497.
147. Bell, A. J. Ferroelectrics: The Role of Ceramic Science and Engineering. J. Eur. Ceram. Soc. 2008, 28, 1307–1317.
148. Hsu, H.; Benjauthrit, V.; Zheng, F.; Chen, R.; Huang, Y.; Zhou, Q.; Shung, K. K. PMN-PT–PZT Composite Films for High Frequency Ultrasonic Transducer Applications. Sens.
Actuators, A 2012, 179, 121–124.
149. Whatmore, R. W.; Zhang, Q.; Huang, Z.; Dorey, R. A. Ferroelectric Thin and Thick Films for Microsystems. Mater. Sci. Semicond. Process. 2002, 5, 65–76.
150. Gopalan, V.; Sanford, N. A.; Aust, J. A.; Kitamura, K.; Furukawa, Y. Handbook of Advanced Electronic and Photonic Materials and Devices. In Crystal Growth, Characterization,
and Domain Studies in Lithium Niobate and Lithium Tantalate Ferroelectrics; Nalwa, Hari Singh, Ed.; Academic Press: Burlington, 2001; pp 57–114 (chapter 2).
151. Malik, P.; Ahuja, J. K.; Raina, K. K. Effect of Polymer Viscosity on Morphological and Electro-Optic Properties of Aligned Polymer Dispersed Ferroelectric Liquid Crystal
Composite Films. Curr. Appl. Phys. 2003, 3, 325–329.
152. Cai, Z.; Li, X.; Hu, Q.; Zeng, X. Study on Thick-Film PTC Thermistor Fabricated by Micro-Pen Direct Writing. Microelectron. J. 2008, 39, 1452–1456.
153. Pu, Y.; Wu, H.; Wei, J. Influence of Doping BiYO3 and Nb2O5 on PTCR Characteristics of BaTiO3 Thermistor Ceramics. Sens. Actuators, A 2012, 173, 158–162.
154. Yuan, C.; Liu, X.; Liang, M.; Zhou, C.; Wang, H. Electrical Properties of Sr–Bi–Mn–Fe–O Thick-Film NTC Thermistors Prepared by Screen Printing. Sens. Actuators, A 2011,
167, 291–296.
155. Li, W. F.; Weng, G. J. A Micromechanics-Based Thermodynamic Model for the Domain Switch in Ferroelectric Crystals. Acta Mater. 2004, 52, 2489–2496.
156. Sallese, J.; Fazan, P. Switch and rf Ferroelectric MEMS: A New Concept. Sens. Actuators, A 2004, 109, 186–194.
157. Takasu, H. Ferroelectric Memories and Their Applications. Microelectron. Eng. 2001, 59, 237–246.
158. Yen, J. H.; Shu, Y. C.; Shieh, J.; Yeh, J. H. A Study of Electromechanical Switching in Ferroelectric Single Crystals. J. Mech. Phys. Solids 2008, 56, 2117–2135.
159. Su, W. A Novel Method to Suppress Vibration-Induced Phase Noise of Crystal Oscillators. Control Eng. Pract. 2004, 12, 1065–1070.
44 Review of Physical Principles of Sensing and Types of Sensing Materials

160. Uhlmann, D. R.; Teowee, G.; Boulton, J. M.; Motakef, S.; Lee, S. C. Electrical and Optical Properties of Chemically Derived Ferroelectric Films. J. Non Cryst. Solids 1992,
147–148, 409–423.
161. Wersing, W. Microwave Ceramics for Resonators and Filters. Curr. Opin. Solid State Mater. Sci. 1996, 1, 715–731.
162. Frank, R. Handbook of Modern Sensors: Physics, Designs, and Applications (edited by J. Fraden). Anal. Bioanal. Chem. 2005, 382, 8–9.
163. Lenz, J. A. Review of Magnetic Sensors. Proc. IEEE 1990, 78, 973–989.
164. Dalven, R. Introduction to Applied Solid State Physics: Topics in the Applications of Semiconductors, Superconductors, and the Nonlinear Optical Properties of Solids; Plenum
Press: New York, 1980; p 330.
165. Yanase, Y.; Jujo, T.; Nomura, T.; Ikeda, H.; Hotta, T.; Yamada, K. Theory of Superconductivity in Strongly Correlated Electron Systems. Phys. Rep. 2003, 387, 1–149.
166. Poole, C. K. Handbook of Superconductivity; Academic Press: San Diego, CA, 1999; p 540.
167. Kresin, V. Z.; Wolf, S. A. Electron-Lattice Interaction and Its Impact on High Tc Superconductivity. Arch. Condens. Matter 2009, 1–95. arXiv:0904.2038v1 [cond-mat.-
supr-con].
168. Fuller-Mora, W. W.; Wolf, S. A.; Kresin, V. Z. Electron-Phonon Coupling as a Mechanism for High-Temperature Superconductivity. J. Supercond. 1994, 7, 543–545.
169. Kresin, V. Z. Section Title: Electric Phenomena. In Magnetic Scattering; 2000; pp 37–43.
170. Buzea, C.; Yamashita, T. Review of Superconducting Properties of MgB2. Los Alamos Natl. Lab., Prepr. Arch., Condens. Matter 2001, 1–35. arXiv:cond-mat/0108265.
171. Hirsch, J. E.; Marsiglio, F. Hole Superconductivity: Review and Some New Results. Phys. C [Amsterdam] 1989, 162–164, 591–596.
172. Fietz, W. H.; Schauer, W.; Wuh, H.; Meingast, C.; Pintschovius, L.; Schuppler, S.; Winter, H. Basic Research on High-Temperature Superconductivity. Nachr. – For-
schungszent. Karlsruhe 1999, 31, 233–250.
173. Gor’kov, L. P. Section Title: Electric Phenomena. In Foundations of Superconductivity and Extension to Less-Common Systems; 1998; pp 3–23.
174. Mueller, K. A. Recent Achievements in High-Temperature Superconductivity. Phys. C [Amsterdam] 1997, 282–287, 3.
175. Essen, H.; Fiolhais, M. C. N. Meissner Effect, Diamagnetism, and Classical Physics – A Review. Am. J. Phys. 2012, 80, 164–169.
176. Mueller, P., Ustinov, A. V., Eds. The Physics of Superconductors: Introduction to Fundamentals and Applications; Springer: Berlin, New York, 1997; p 225.
177. Boeri, L.; Bachelet, G.; Cappelluti, E.; Pietronero, L. The Origin of Phonon Anharmonicity in MgB2 and Related Compounds. Supercond. Sci. Technol. 2003, 16, 143–146.
178. Kaindl, R. A.; Carnahan, M. A.; Orenstein, J.; Chemla, D. S.; Christen, H. M.; Zhai, H. Y.; Paranthaman, M.; Lowndes, D. H. Far-Infrared Optical Conductivity Gap in
Superconducting MgB2 Films. Phys. Rev. Lett. 2002, 88, 027003.
179. Ardavan, A.; Brown, S.; Kagoshima, S.; Kanoda, K.; Kuroki, K.; Mori, H.; Ogata, M.; Uji, S.; Wosnitza, J. Recent Topics of Organic Superconductors. J. Phys. Soc. Jpn. 2012, 81, 011004.
180. Guo, K. W. Green Nanotechnology of Trends in Future Energy: A Review. Int. J. Energy Res. 2012, 36, 1–17.
181. Hiramatsu, H.; Katase, T.; Kamiya, T.; Hosono, H. Thin Film Growth and Device Fabrication of Iron-Based Superconductors. J. Phys. Soc. Jpn. 2012, 81, 011011.
182. Rao, C. N. R. Novel Materials, Materials Design and Synthetic Strategies: Recent Advances and New Directions. J. Mater. Chem. 1999, 9, 1–14.
183. Manthiram, A.; Kim, J. Low Temperature Synthesis of Insertion Oxides for Lithium Batteries. Chem. Mater. 1998, 10, 2895–2909.
184. West, A. R. Solid State Chemistry and Its Applications; Wiley: Chichester; New York, 1984; p 734.
185. Ratnakumar, B. V.; Narayanan, S. R. Section Title: Electrochemical, Radiational, and Thermal Energy Technology. In Fundamental Aspects of Ion Transport in Solid Elec-
trolytes; 1995; pp 1–40.
186. Babanly, M. B.; Gasanova, Z. T.; Mashadieva, L. F.; Zlomanov, V. P.; Yusibov, Y. A. Thermodynamic Study of the Cu-As-S System by EMF Measurements with Cu4RbCl3I2 as
a Solid Electrolyte. Inorg. Mater. 2012, 48, 225–228.
187. Tuller, H. L. Review of Electrical Properties of Metal Oxides as Applied to Temperature and Chemical Sensing. Sens. Actuators 1983, 4, 679–688.
188. Jacobson, A. J. Materials for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 660–674.
189. Baggetto, L.; Niessen, R. A. H.; Roozeboom, F.; Notten, P. H. L. High Energy Density All-Solid-State Batteries: A Challenging Concept towards 3D Integration. Adv. Funct.
Mater. 2008, 18, 1057–1066.
190. Masquelier, C. Solid Electrolytes Lithium Ions on the Fast Track. Nat. Mater. 2011, 10, 648–650.
191. Larsson, O.; Wang, X.; Berggren, M.; Crispin, X. Proton Motion in a Polyelectrolyte: A Probe for Wireless Humidity Sensors. Sens. Actuators, B 2010, 143, 482–486.
192. Fabbri, E.; Bi, L.; Pergolesi, D.; Traversa, E. Towards the Next Generation of Solid Oxide Fuel Cells Operating below 600 degrees C with Chemically Stable Proton-Conducting
Electrolytes. Adv. Mater. 2012, 24, 195–208.
193. Fergus, J. W. Sensing Mechanism of Non-Equilibrium Solid-Electrolyte-Based Chemical Sensors. J. Solid State Electrochem. 2011, 15, 971–984.
194. Park, C. O.; Fergus, J. W.; Miura, N.; Park, J.; Choi, A. Solid-State Electrochemical Gas Sensors. Ionics 2009, 15, 261–284.
195. Sun, C.; Hui, R.; Roller, J. Cathode Materials for Solid Oxide Fuel Cells: A Review. J. Solid State Electrochem. 2010, 14, 1125–1144.
196. Tsipis, E. V.; Kharton, V. V. Electrode Materials and Reaction Mechanisms in Solid Oxide Fuel Cells: A Brief Review. J. Solid State Electrochem. 2008, 12, 1367–1391.
197. Tucker, M. C. Progress in Metal-Supported Solid Oxide Fuel Cells: A Review. J. Power Sources 2010, 195, 4570–4582.
198. Owens, B. B. Solid State Electrolytes: Overview of Materials and Applications during the Last Third of the Twentieth Century. J. Power Sources 2000, 90, 2–8.
199. Guth, U.; Vonau, W.; Zosel, J. Recent Developments in Electrochemical Sensor Application and Technology – A Review. Meas. Sci. Technol. 2009, 20, 042002.
200. Pasierb, P.; Rekas, M. Solid-State Potentiometric Gas SensorsdCurrent Status and Future Trends. J. Solid State Electrochem. 2009, 13, 3–25.
201. Möbius, H.; Hartung, R. Solid-State Potentiometric Gas Sensors – A Supplement. J. Solid State Electrochem. 2010, 14, 669–673.
202. Garzon, F. H.; Mukundan, R.; Brosha, E. L. Solid-State Mixed Potential Gas Sensors: Theory. Experiments and Challenges. Solid State Ionics 2000, 136, 633–638.
203. Alber, K.; Cox, J.; Kulesza, P. Solid-State Amperometric Sensors for Gas Phase Analytes: A Review of Recent Advances. Electroanalysis 1997, 9, 97–101.
204. AIZAWA, M. Principles and Applications of Electrochemical and Optical Biosensors. Anal. Chim. Acta 1991, 250, 249–256.
205. Tothill, I. Biosensors Developments and Potential Applications in the Agricultural Diagnosis Sector. Comput. Electron. Agric. 2001, 30, 205–218.
206. Luong, J. H. T.; Male, K. B.; Glennon, J. D. Biosensor Technology: Technology Push versus Market Pull. Biotechnol. Adv. 2008, 26, 492–500.
207. Eisenmenger, M. J.; Reyes-De-Corcuera, J. I. High Pressure Enhancement of Enzymes: A Review. Enzyme Microb. Technol. 2009, 45, 331–347.
208. Ferri, T.; Maida, S.; Poscia, A.; Santucci, R. A Glucose Biosensor Based on Electro-Enzyme Catalyzed Oxidation of Glucose Using a HRP-GOD Layered Assembly.
Electroanalysis 2001, 13, 1198–1202.
209. Clark, L. C., Jr.; Lyons, C. Electrode Systems for Continuous Monitoring in Cardiovascular Surgery. Ann. N. Y. Acad. Sci. 1962, 102, 29–45.
210. Wolfgang, S. Amperometric Enzyme Biosensors Based on Optimised Electron-Transfer Pathways and Non-Manual Immobilisation Procedures. Rev. Mol. Biotechnol. 2002, 82,
425–441.
211. Guilbault, G. G.; Lubrano, G. Enzyme Electrode for Amperometric Determination of Glucose. Anal. Chim. Acta 1973, 64, 439–455.
212. Wang, J. Glucose Biosensors: 40 Years of Advances and Challenges RID C-6175-2011. Electroanalysis 2001, 13, 983–988.
213. P.D, P. [Bio]sensors for Measurement of Analytes Implicated in Food Safety: A Review. TrAC – Trends Anal. Chem. 2002, 21, 96–115.
214. Pejcic, B.; De Marco, R. Impedance Spectroscopy: Over 35 Years of Electrochemical Sensor Optimization. Electrochim. Acta 2006, 51, 6217–6229.
215. Gau, V.; Ma, S.; Wang, H.; Tsukuda, J.; Kibler, J.; Haake, D. A. Electrochemical Molecular Analysis without Nucleic Acid Amplification. Methods 2005, 37, 73–83.
216. Zhang, Y. C.; Kim, H. H.; Heller, A. Enzyme-Amplified Amperometric Detection of 3000 Copies of DNA in a 10-mu L Droplet at 0.5 fM Concentration. Anal. Chem. 2003, 75,
3267–3269.
217. Pividori, M. I.; Merkoçi, A.; Alegret, S. Classical Dot–Blot Format Implemented as an Amperometric Hybridisation Genosensor. Biosens. Bioelectron. 2001, 16, 1133–1142.
218. Devadoss, A.; Spehar-Deleze, A.; Tanner, D. A.; Bertoncello, P.; Marthi, R.; Keyes, T. E.; Forster, R. J. Enhanced Electrochemiluminescence and Charge Transport through
Films of Metallopolymer-Gold Nanoparticle Composites. Langmuir 2010, 26, 2130–2135.
219. Li, Z.; Liu, C.; Fan, Y.; Wang, Y.; Duan, X. A Chemiluminescent Metalloimmunoassay Based on Silver Deposition on Colloidal Gold Labels. Anal. Biochem. 2006, 359, 247–252.
Review of Physical Principles of Sensing and Types of Sensing Materials 45

220. Zhao, J. G.; Henkens, R. W.; Stonehuerner, J.; Odaly, J. P.; Crumbliss, A. L. Direct Electron-Transfer at Horseradish-Peroxidase Colloidal Gold Modified Electrodes.
J. Electroanal. Chem. 1992, 327, 109–119.
221. Xiao, Y.; Ju, H.; Chen, H. Hydrogen Peroxide Sensor Based on Horseradish Peroxidase-Labeled Au Colloids Immobilized on Gold Electrode Surface by Cysteamine Monolayer.
Anal. Chim. Acta 1999, 391, 73–82.
222. Mello, L.; Kubota, L. Review of the Use of Biosensors as Analytical Tools in the Food and Drink Industries. Food Chem. 2002, 77, 237–256.
223. Vastarella, W.; Nicastri, R. Enzyme/Semiconductor Nanoclusters Combined Systems for Novel Amperometric Biosensors. Talanta 2005, 66, 627–633.
224. Freire, R.; Pessoa, C.; Mello, L.; Kubota, L. Direct Electron Transfer: An Approach for Electrochemical Biosensors with Higher Selectivity and Sensitivity. J. Braz. Chem. Soc.
2003, 14, 230–243.
225. Thevenot, D.; Toth, K.; Durst, R.; Wilson, G. Electrochemical Biosensors: Recommended Definitions and Classification RID B-6446-2009. Biosens. Bioelectron. 2001, 16, 121–131.
226. Watson, J. D., Ed. Discovering the Double Helix: History of Science; Biotechnology; 2001.
227. Polsky, R.; Gill, R.; Kaganovsky, L.; Willner, I. Nucleic Acid-Functionalized Pt Nanoparticles: Catalytic Labels for the Amplified Electrochemical Detection of Biomolecules.
Anal. Chem. 2006, 78, 2268–2271.
228. Chen, X.; Su, B.; Song, X.; Chen, Q.; Chen, X.; Wang, X. Recent Advances in Electrochemiluminescent Enzyme Biosensors. TrAC – Trends Anal. Chem. 2011, 30, 665–676.
229. Fernando Sérgio Rodrigues Ribeiro, T. Biosensors and Rapid Diagnostic Tests on the Frontier between Analytical and Clinical Chemistry for Biomolecular Diagnosis of Dengue
Disease: A Review. Anal. Chim. Acta 2011, 687, 28–42.
230. Levicky, R.; Horgan, A. Physicochemical Perspectives on DNA Microarray and Biosensor Technologies. Trends Biotechnol. 2005, 23, 143–149.
231. Keller, G.; Huang, D.; Manak, M. Labeling of DNA Probes with a Photoactivatable Hapten. Anal. Biochem. 1989, 177, 392–395.
232. Kohler, S.; Belkin, S.; Schmid, R. Reporter Gene Bioassays in Environmental Analysis RID A-8947-2008. Fresenius J. Anal. Chem. 2000, 366, 769–779.
233. Rompre, A.; Servais, P.; Baudart, J.; de-Roubin, M.; Laurent, P. Detection and Enumeration of Coliforms in Drinking Water: Current Methods and Emerging Approaches.
J. Microbiol. Methods 2002, 49, 31–54.
234. Pandey, P.; Datta, M.; Malhotra, B. D. Prospects of Nanomaterials in Biosensors. Anal. Lett. 2008, 41, 159–209.
235. Piunno, P.; Krull, U.; Hudson, R.; Damha, M.; Cohen, H. Fiberoptic DNA Sensor for Fluorometric Nuclei Acid Determination. Anal. Chem. 1995, 67, 2635–2643.
236. Millan, K.; Spurmanis, A.; Mikkelsen, S. Covalent Immobilization of DNA onto Glassy-Carbon Electrodes. Electroanalysis 1992, 4, 929–932.
237. Henke, L.; Piunno, P.; McClure, A.; Krull, U. Covalent Immobilization of Single-Stranded DNA onto Optical Fibers Using Various Linkers. Anal. Chim. Acta 1997, 344, 201–213.
238. Wang, X.; Smirnov, S. Label-Free DNA Sensor Based on Surface Charge Modulated Ionic Conductance. ACS Nano 2009, 3, 1004–1010.
239. Lucarelli, F.; Marrazza, G.; Turner, A. P. F.; Mascini, M. Carbon and Gold Electrodes as Electrochemical Transducers for DNA Hybridization Sensors. Biosens. Bioelectron.
2004, 19, 515–530.
240. Palecek, E. Past, Present and Future of Nucleic Acids Electrochemistry. Talanta 2002, 56, 809–819.
241. Kerman, K.; Chikae, M.; Yamamura, S.; Tamiya, E. Gold Nanoparticle-Based Electrochemical Detection of Protein Phosphorylation. Anal. Chim. Acta 2007, 588, 26–33.
242. Drummond, T. G.; Hill, M. G.; Barton, J. K. Electrochemical DNA Sensors. Nat. Biotechnol. 2003, 21, 1192–1199.
243. Moeller, R.; Fritzsche, W. Chip-Based Electrical Detection of DNA. IEEE Proc.: Nanobiotechnol. 2005, 152, 47–51.
244. Hernandez-Santos, D.; Diaz-Gonzalez, M.; Gonzalez-Garcia, M. B.; Costa-Garcia, A. Enzymatic Genosensor on Streptavidin-Modified Screen-Printed Carbon Electrodes. Anal.
Chem. 2004, 76, 6887–6893.
245. Feng, P.; Lum, R.; Chang, G. Identification of uida Gene-Sequences in Beta-D-Glucuronidase-Negative Escherichia coli. Appl. Environ. Microbiol. 1991, 57, 320–323.
246. Zhai, J.; Hong, C.; Yang, R. DNA Based Biosensors. Biotechnol. Adv. 1997, 15, 43–58.
247. Saiki, R.; Scharf, S.; Faloona, F.; Mullis, K.; Horn, G.; Erlich, H.; Arnheim, N. Enzymatic Amplification of Beta-Globin Genomic Sequences and Restriction Site Analysis for
Diagnosis of Sickle-Cell Anemia. Science 1985, 230, 1350–1354.
248. Piunno, P. A. E.; Krull, U. J.; Hudson, R. H. E.; Damha, M. J.; Cohen, H. Fiber-Optic DNA Sensor for Fluorometric Nucleic Acid Determination. Anal. Chem. 1995, 67,
2635–2643.
249. Hirschfeld, T. B. Patent Application Country: Application: DE; Patent Country: DE; Priority Application Country: US Patent 3708513, 1987.
250. Watts, H. J.; Yeung, D.; Parkes, H. Real-Time Detection and Quantification of DNA Hybridization by an Optical Biosensor. Anal. Chem. 1995, 67, 4283–4289.
251. Teranishi, T.; Hosoe, M.; Miyake, M. Formation of Monodispersed Ultrafine Platinum Particles and Their Electrophoretic Deposition on Electrodes. Adv. Mater. 1997, 9, 65.
252. Liu, Y.; Lei, J.; Ju, H. Amperometric Sensor for Hydrogen Peroxide Based on Electric Wire Composed of Horseradish Peroxidase and Toluidine Blue-Multiwalled Carbon
Nanotubes Nanocomposite. Talanta 2008, 74, 965–970.
253. Yang, Z. S.; Wu, W. L.; Chen, X.; Liu, Y. C. An Amperometric Horseradish Peroxidase Inhibition Biosensor for the Determination of Phenylhydrazine. Anal. Sci. 2008, 24,
895–899.
254. Nuzzo, R. G.; Allara, D. L. Adsorption of Bifunctional Organic Disulfides on Gold Surfaces. J. Am. Chem. Soc. 1983, 105, 4481–4483.
255. Peterson, A. W.; Heaton, R. J.; Georgiadis, R. M. The Effect of Surface Probe Density on DNA Hybridization. Nucleic Acids Res. 2001, 29, 5163–5168.
256. Palecek, E.; Fojta, M. Detecting DNA Hybridization and Damage. Anal. Chem. 2001, 73, 74A–83A.
257. Lei, C.; Hu, S.; Gao, N.; Shen, G.; Yu, R. An Amperometric Hydrogen Peroxide Biosensor Based on Immobilizing Horseradish Peroxidase to a Nano-Au Monolayer Supported by
Sol–Gel Derived Carbon Ceramic Electrode. Bioelectrochemistry 2004, 65, 33–39.
258. Venkatesh, A. G.; Herth, S.; Becker, A.; Reiss, G. Orientation-Defined Alignment and Immobilization of DNA between Specific Surfaces. Nanotechnology 2011, 22, 145301.
259. Zhang, H.; Lai, G.; Han, D.; Yu, A. An Amperometric Hydrogen Peroxide Biosensor Based on Immobilization of Horseradish Peroxidase on an Electrode Modified with Magnetic
Dextran Microspheres. Anal. Bioanal. Chem. 2008, 390, 971–977.
260. Park, B. W.; Kim, D. S.; Yoon, D. Y. Surface Modification of Gold Electrode with Gold Nanoparticles and Mixed Self-Assembled Monolayers for Enzyme Biosensors. Korean
J. Chem. Eng. 2011, 28, 64–70.
261. Herne, T. M.; Tarlov, M. J. Characterization of DNA Probes Immobilized on Gold Surfaces. J. Am. Chem. Soc. 1997, 119, 8916–8920.
262. Wackerbarth, H.; Grubb, M.; Zhang, J.; Hansen, A.; Ulstrup, J. Long-Range Order of Organized Oligonucleotide Monolayers on Au[111] Electrodes. Langmuir 2004, 20,
1647–1655.
263. Singhal, P.; Kuhr, W. G. Ultrasensitive Voltammetric Detection of Underivatized Oligonucleotides and DNA. Anal. Chem. 1997, 69, 4828–4832.
264. Hu, K.; Lan, D.; Li, X.; Zhang, S. Electrochemical DNA Biosensor Based on Nanoporous Gold Electrode and Multifunctional Encoded DNA-Au Bio Bar Codes. Anal. Chem. 2008,
80, 9124–9130.
265. Steel, A. B.; Levicky, R. L.; Herne, T. M.; Tarlov, M. J. Immobilization of Nucleic Acids at Solid Surfaces: Effect of Oligonucleotide Length on Layer Assembly. Biophys. J. 2000,
79, 975–981.
266. Marcon, L.; Stiévenard, D.; Melnyk, O. Electrical Detection of Human Immunoglobulins G from Human Serum Using a Microbiosensor. Biosens. Bioelectron. 2007, 23, 81–87.
267. Bange, A.; Halsall, H.; Heineman, W. Microfluidic Immunosensor Systems. Biosens. Bioelectron. 2005, 20, 2488–2503.
268. Killard, A.; Deasy, B.; Okennedy, R.; Smyth, M. Antibodies – Production, Functions and Applications in Biosensors Rid A-8676-2008. TrAC – Trends Anal. Chem. 1995, 14,
257–266.
269. Ghindilis, A.; Atanasov, P.; Wilkins, M.; Wilkins, E. Immunosensors: Electrochemical Sensing and other Engineering Approaches. Biosens. Bioelectron. 1998, 13, 113–131.
270. Fowler, J. M.; Wong, D. K. Y.; Brian Halsall, H.; Heineman, W. R.. Electrochemical Sensors, Biosensors and Their Biomedical Applications. In Recent Developments in
Electrochemical Immunoassays and Immunosensors; Academic Press: San Diego, 2008; pp 115–143, (chapter 5).
271. Hock, B.; Dankwardt, A.; Kramer, K.; Marx, A. Immunochemical Techniques – Antibody-Production for Pesticide Analysis – A Review. Anal. Chim. Acta 1995, 311,
393–405.
46 Review of Physical Principles of Sensing and Types of Sensing Materials

272. Crowther, J.; Reckziegel, P.; Prado, J. Quantification of Whole Virus-Particles [146s] of Foot-and-Mouth-Disease Virus in the Presence of Virus Subunits [12s], Using
Monoclonal-Antibodies in a Sandwich Elisa. Vaccine 1995, 13, 1064–1075.
273. Hoffman, B.; Yu, H.; Diamandis, E. Assay of Prostate Specific Antigen from Whole Blood Spotted on Filter Paper and Application to Prostate Cancer Screening. Clin. Chem.
1996, 42, 536–544.
274. Tang, D.; Zhang, D.; Tang, D.; Ai, H. Amplification of the Antigen–Antibody Interaction from Quartz Crystal Microbalance Immunosensors via Back-Filling Immobilization of
Nanogold on Biorecognition Surface. J. Immunol. Methods 2006, 316, 144–152.
275. Wang, J. DNA Based Biosensors. Biosens. Environ. Monit. 2000, 124–135.
276. Williams, E.; Pividori, M. I.; Merkoçi, A.; Forster, R. J.; Alegret, S. Rapid Electrochemical Genosensor Assay Using a Streptavidin Carbon-Polymer Biocomposite Electrode.
Biosens. Bioelectron. 2003, 19, 165–175.
277. Babkina, S. S.; Budnikov, G. K. Electrochemical Biosensors Based on Nucleic Acids and Their Use in Bioaffinity Assays for Determining DNA and Its Effectors. J. Anal. Chem.
2006, 61, 728–739.
278. Braguglia, C. M. Biosensors: An Outline of General Principles and Application. Chem. Biochem. Eng. Q 1998, 12, 183–190.

You might also like