Smith, R.L. “Sensors” The Electrical Engineering Handbook Ed. Richard C. Dorf Boca Raton: CRC Press LLC, 2000 56 Sensors 56.1Introduction 56.2Physical Sensors Temperature Sensors?Displacement and Force?Optical Radiation 56.3Chemical Sensors Ion-Selective Electrode?Gas Chromatograph 56.4Biosensors Immunosensor?Enzyme Sensor 56.5Microsensors 56.1 Introduction Sensors are critical components in all measurement and control systems. The need for computer-compatible sensors closely followed the advent of the microprocessor. Together with the always-present need for sensors in science and medicine, the demand for sensors in automated manufacturing and processing is rapidly growing. In addition, small, inexpensive sensors are finding their way into all sorts of consumer products, from childrens’ toys to dishwashers to automobiles. Because of the vast variety of useful things to be sensed and sensor applications, sensor engineering is a multidisciplinary and interdisciplinary field of endeavor. This chapter introduces some basic definitions, concepts, and features of sensors and illustrates them with several examples. The reader is directed to the references and the sources listed under “Further Information” for more details and examples. There are many terms which are often used synonymously for sensor, including transducer, meter, detector, and gage. Defining the term sensor is not an easy task; however the most widely used definition is that which has been applied to electrical transducers by the Instrument Society of America (ANSI MC6.1, 1975): Trans- ducer—A device which provides a usable output in response to a specified measurand. A transducer is more generally defined as a device which converts energy from one form to another. A usable ouput refers to an optical, electrical, or mechanical signal. In the context of electrical engineering, however, a usable output refers to an electrical output signal. The measurand can be a physical, chemical, or biological property or condition to be measured. Most, but not all, sensors are transducers, employing one or more transduction mechanisms to produce an electrical output signal. Sometimes sensors are classified as direct and indirect sensors according to how many transduction mechanisms are used. For example, a mercury thermometer produces a change in volume of mercury in response to a temperature change via thermal expansion, but the output is a mechanical displace- ment and not an electrical signal. Another transduction mechanism is required. A thermometer is still a useful sensor since humans can read the change in mercury height using their eyes as the second transducing element. However, in order to produce an electrical output for use in a control loop, the height of the mercury would have to be converted to an electrical signal. This could be accomplished using capacitive effects. However, there are more direct temperature sensing methods, i.e., one where an electrical output is produced in response to a change in temperature. An example is given in the next section on physical sensors. Figure 56.1 depicts a Rosemary L. Smith University of California, Davis ? 2000 by CRC Press LLC ? 2000 by CRC Press LLC T ABLE 56.1 Ph ysical and C h emical T r ansduction P r inciples Sec ondar y S i g nal P r imar y S i g nal M e c hanical Ther mal Elect r ical M a g n etic R adiant Chemical M e c hanical (Fluid) mec hanical and F r iction effects (e.g ., Piez oelect r icit y M ag net o-mec hanical Phot oelastic syst ems ac oustic effects (e.g ., fr iction calor imet e r) Piez or esisti v it y effects (e.g ., piez o- (str ess-induc ed diaphr ag m, g r a v it y C o oling effects (e.g ., R esisti v e, capaciti v e, and mag netic effect) b ir efr ingenc e) balanc e, ec ho sounder) ther mal flo w met e rs) inducti v e effects I nt erfer omet ers Sag n ac effect Doppler effect Ther mal Ther mal e xpansion (bimetal str i p , liquid-in-g lass and gas ther m omet ers, r esonant fr equency) Seebec k effect Ther moo p tical effects R eaction acti vation Ther mor esistanc e (e.g ., in liquid cr ystals) (e.g ., ther mal Py r o elect r icit y Radiant emission dissociation) Ther mal ( J ohnson) noise Radiomet e r effect (lig ht mill) Elect r ical Elect r okinetic and elect r o- J oule (r esisti v e) Charge c o llect o rs B iot-Sa v ar t ’ s la w Elect r o optical effects Elect r o lysis mec hanical effects (e.g ., heating Lang m uir pr obe (e.g ., K e r r effect) Elect r o mig r a tion piez oelect r icit y , elect r o- P e ltier effect P o ck el’ s effect met e r , A m per e ’ s la w) Elect r o luminesc enc e M a g n etic M ag net o mec hanical effects Ther momag netic effects Ther momag netic effects M a g n et ooptical effects (e.g ., mag net o r estr i ction, (e.g ., Rig hi-L educ effect) (e.g ., E t tingshausen- (e.g ., F a r a da y effect) mag net o met e r) Gal vanomag netic effects N e r nst effect) C ott on-M out o n effect (e.g ., E t tingshausen Gal vanomag netic effects effect) (e.g ., H a ll effect, mag net o r esistanc e) Radiant Radiation pr essur e B olomet e r ther m opile Phot oelect r ic effects Phot or efr a cti v e effects Phot osynthesis, (e.g ., phot o v oltaic effect, Optical bistabilit y -dissociation phot o c onduc t i v e effec t ) Chemical H y g r omet er Calor imet e r P ot entiomet r y N u clear mag netic (Emission and Elect r o deposition c e ll Ther mal c o nducti v i t y c e ll C o nductimet r y r esonanc e absor p tion) spect r osc op y Phot oac oustic effect A m per o met r y Chemiluminisc enc e Flame ionizationV olta effect Gas-sensiti v e field effect Sour ce: T . G r andk e and J . H esse, I n tr oduction, V ol. 1: F undame ntals and G e ner al A s pects, Se ns ors: A C ompr e he nsive S u r v e y , W . Gopel, J . H esse, and J . H. Z e mel, Eds., W e inheim, Ger man y : V CH, 1989. W ith per mission. sensor block diagram identifying the measurand and associated input signal, the primary and intermediate transduction mechanisms, and the electronic output signal. Active sensors require an external power source in order to produce a usable output signal, e.g., the piezoresistor. Table 56.1 is a 6 ′ 6 matrix of the more commonly employed physical and chemical transduction mechanisms. Many of the effects listed are described in more detail in this handbook (see Chapters 53–58). In choosing a particular sensor for a given application, there are many factors to be considered. These deciding factors or specifications can be divided into three major categories: environmental factors, economic factors, and the sensor characteristics. The most commonly encountered factors are listed in Table 56.2, although not all of these factors may be pertinent to a particular application. Most of the environmental factors determine the packaging of the sensor, with packaging meaning the encapsulation or insulation which provides protection and isolation and the input/output leads or connections and cabling. The economic factors determine the type of manufacturing and materials used in the sensor and to some extent the quality of the materials (with respect to lifetime). For example, a very expensive sensor may be cost effective if it is used repeatedly or for very long periods of time. On the other hand, a disposable sensor, such as is desired in many medical applications, should be inexpensive. The sensor characteristics of the sensor are usually the specifications of primary concern. The most important parameters are sensitivity, stability, and repeatability. Normally, a sensor is only useful if all three of these parameters are tightly specified for a given range of measurand and time of operation. For example, a highly sensitive device is not useful if its output signal drifts greatly during the measurement time and the data obtained is not reliable if the measurement is not repeatable. Other output characteristics, such as selectivity and linearity, can often be compensated for by using additional, independent sensor input or with signal conditioning circuits. In fact, most sensors have a response to temperature, since most tranducing effects are temperature dependent. Sensors are most often classified by the type of measurand, i.e., physical, chemical, or biological. This is a much simpler means of classification than by transduction mechanism or output signal (e.g., digital or analog), since many sensors use multiple transduction mechanisms and the output signal can always be processed, conditioned, or converted by a circuit so as to cloud the definition of output. A description of each class and examples are given in the following sections. The last section introduces microsensors and gives some examples. FIGURE 56.1 Sensor block diagram. Active sensors require input power to accomplish transduction. Many sensors employ multiple transduction mechanisms in order to produce an electronic output in response to the measurand. TABLE 56.2 Environmental Factors Economic Factors Sensor Characteristics Temperature range Cost Sensitivity Humidity effects Availability Range Corrosion Lifetime Stability Size Repeatability Overrange protection Linearity Susceptibility to EM interferences Error Ruggedness Response time Power consumption Frequency response Self-test capability ? 2000 by CRC Press LLC 56.2 Physical Sensors Physical measurands include temperature, strain, force, pressure, displacement, position, velocity, acceleration, optical radiation, sound, flow rate, viscosity, and electromagnetic fields. Referring to Table 56.1, all but those transduction mechanisms listed in the chemical column are used in the design of physical sensors. Clearly, they comprise a very large proportion of all sensors. It is impossible to illustrate all of them, but three measurands stand out in terms of their widespread application: temperature, displacement (or associated force), and optical radiation. Temperature Sensors Temperature is an important parameter in many control systems, most familiarly in environmental control systems. Several distinctly different transduction mechanisms have been employed. The mercury thermometer was mentioned in the Introduction as a nonelectrical sensor. The most commonly used electrical temperature sensors are thermocouples, thermistors, and resistance thermometers. Thermocouples employ the Seebeck effect, which occurs at the junction of two dissimilar metal wires. A voltage difference is generated at the hot junction due to the difference in the energy distribution of thermally energized electrons in each metal. This voltage is measured across the cool ends of the two wires and changes linearly with temperature over a given range, depending on the choice of metals. To minimize measurement error the cool end of the couple must be kept at a constant temperature, and the voltmeter must have a high input impedance. The resistance thermometer relies on the increase in resistance of a metal wire with increasing temperature. As the electrons in the metal gain thermal energy, they move about more rapidly and undergo more frequent collisions with each other and the atomic nuclei. These scattering events reduce the mobility of the electrons, and since resistance is inversely proportional to mobility, the resistance increases. Resistance thermometers consist of a coil of fine metal wire. Platinum wire gives the largest linear range of operation. To determine the resistance indirectly, a constant current is supplied and the voltage is measured. A direct measurement can be made by placing the resistor in the sensing arm of a Wheatstone bridge and adjusting the opposing resistor to “balance” the bridge, which produces a null output. A measure of the sensitivity of a resistance thermometer is its temperature coefficient of resistance: TCR = (DR/R)(1/DT) in units of % resistance per degree of temperature. Thermistors are resistive elements made of semiconductor materials and have a negative coefficient of resistance. The mechanism governing the resistance change of a thermistor is the increase in the number of conducting electrons with an increase in temperature due to thermal generation, i.e., the electrons which are the least tightly bound to the nucleus (valence electrons) gain sufficient thermal energy to break away and become influenced by external fields. Thermistors can be measured in the same manner as resistance thermom- eters, but thermistors have up to 100 times higher TCR values. Displacement and Force Many types of forces are sensed by the displacements they create. For example, the force due to acceleration of a mass at the end of a spring will cause the spring to stretch and the mass to move. Its displacement from the zero acceleration position is governed by the force generated by the acceleration (F = m · a) and the restoring force of the spring. Another example is the displacement of the center of a deformable membrane due to a difference in pressure across it. Both of these examples use multiple transduction mechanisms to produce an electronic output: a primary mechanism which converts force to displacement (mechanical to mechanical) and then an intermediate mechanism to convert displacement to an electrical signal (mechanical to electrical). Displacement can be measured by an associated capacitance. For example, the capacitance associated with a gap which is changing in length is given by C = area ′ dielectric constant/gap length. The gap must be very small compared to the surface area of the capacitor, since most dielectric constants are of the order of 1 ′ 10 –13 farads/cm and with present methods, capacitance is readily resolvable to only about 10 –12 farads. This is because measurement leads and contacts create parasitic capacitances the same order of magnitude. If the capacitance is measured at the generated site by an integrated circuit (see Section III), capacitances as small as 10 –15 farads ? 2000 by CRC Press LLC can be measured. Displacement is also commonly measured by the movement of a ferromagnetic core inside of an inductor coil. The displacement produces a change in inductance which can be measured by placing the inductor in an oscillator circuit and measuring the change in frequency of oscillation. The most commonly used force sensor is the strain gage. It consists of metal wires which are stretched in response to a force. The resistance of the wire changes as it undergoes strain, i.e., a change in length, since the resistance of a wire is R = resistivity ′ length/cross-sectional area. The wire’s resistivity is a bulk property of the metal which is a constant for constant temperature. For example, a strain gage can be used to measure acceleration by attaching both ends of the wire to a cantilever beam, with one end of the wire at the attached beam end and the other at the free end. The cantilever beam free end moves in response to an applied force, such as the force due to acceleration which produces strain in the wire and a subsequent change in resistance. The sensitivity of a strain gage is described by the unitless gage factor, G = (DR/R)/(DL/L). For metal wires, gage factors typically range from 2 to 3. Semiconductors are known to exhibit piezoresistivity, which is a change in resistance in response to strain which involves a large change in resistivity in addition to the change in linear dimension. Piezoresistors have gage factors as high as 130. Piezoresistive strain gages are frequently used in microsensors, described in Section 56.5. Optical Radiation The intensity and frequency of optical radiation are parameters of growing interest and utility in consumer products such as the video camera and home security systems and in optical communications systems. The conversion of optical energy to electronic signals can be accomplished by several mechanisms (see radiant to electronic transduction in Table 56.1); however, the most commonly used is the photogeneration of carriers in semiconductors. The most often-used device is the p-n junction photodiode (Section III). The construction of this device is very similar to the diodes used in electronic circuits as rectifiers. The diode is operated in reverse bias, where very little current normally flows. When light is incident on the structure and is absorbed in the semiconductor, energetic electrons are produced. These electrons flow in response to the electric field sustained internally across the junction, producing an externally measurable current. The current magnitude is propor- tional to the light intensity and also depends on the frequency of the light. Figure 56.2 shows the effects of varying incident optical intensity on the terminal current versus voltage behavior of a p-n junction. Note that for zero applied voltage, a net negative current flows when the junction is illuminated. This device can therefore also be a source of power (a solar cell). 56.3 Chemical Sensors Chemical measurands include ion concentration, chemical composition, rate of reactions, reduction-oxidation potentials, and gas concentration. The last column of Table 56.1 lists some of the transduction mechanisms that have been, or could be, employed in chemical sensing. Two examples of chemical sensors are described FIGURE 56.2 Sketch of the variation of current versus voltage characteristics of a p-n photodiode with incident light intensity. ? 2000 by CRC Press LLC here: the ion-selective electrode (ISE) and the gas chromatograph. They were chosen because of their general use and availability and because they illustrate the use of a primary (ISE) versus a primary plus intermediate (gas chromatograph) transduction mechanism. Ion-Selective Electrode (ISE) As the name implies, ISEs are used to measure the concentration of a specific ion concentration in a solution of many ions. To accomplish this, a membrane which selectively generates a potential which is dependent on the concentration of the ion of interest is used. The generated potential is usually an equilibrium potential, called the Nernst potential, and develops across the interface of the membrane with the solution. This potential is generated by the initial net flow of ions (charge) across the membrane in response to a concentration gradient, and from thence forth the diffusional force is balanced by the generated electric force and equilibrium is established. This is very similar to the so-called built-in potential of a p-n junction diode. The ion-selective membrane acts in such a way as to ensure that the generated potential is dependent mostly on the ion of interest and negligibly on any other ions in solution. This is done by enhancing the exchange rate of the ion of interest across the membrane, so it is the fastest moving and, therefore, the species which generates and maintains the potential. The most familiar ISE is the pH electrode. In this device the membrane is a sodium glass which possesses a high exchange rate for H + . The generated Nernst potential, E, is given by the expression: E = E 0 + (RT/F) ln[H + ], where E 0 is a constant for constant temperature, R is the gas constant, and F is the Faraday constant. pH is defined as the negative of the log[H + ]; therefore pH = (E 0 – E)(log e)F/RT. One pH unit change corresponds to a tenfold change in the molar concentration of H + and a 59 mV change in the Nernst potential at room temperature. Other ISEs have the same type of response, but specific to a different ion, depending on the choice of membrane. Many ISEs employ ionophores trapped inside of a polymeric membrane. An ionophore is a molecule which selectively and reversibly binds with an ion and thereby creates a high exchange rate for that particular ion. The ISE consists of a glass tube with the ion-selective membrane closing that end of the tube which is immersed into the test solution. The Nernst potential is measured by making electrical contact to each side of the membrane. This is done by placing a fixed concentration of conductive filling solution inside of the tube and placing a wire into the solution. The other side of the membrane is contacted by a reference electrode placed inside of the same solution under test. The reference electrode is constructed in the same manner as the ISE but it has a porous membrane which creates a liquid junction between its inner filling solution and the test solution. That junction is designed to have a potential which is invariant with changes in concentration of any ion in the test solution. The reference electrode, solution under test, and the ISE form an electrochemical cell. The reference electrode potential acts like the ground reference in electric circuits, and the ISE potential is measured between the two wires emerging from the respective two electrodes. The details of the mechanisms of transduction in ISEs are beyond the scope of this chapter. The reader is referred to Bard and Faulkner [1980] and Janata [1989]. Gas Chromatograph Molecules in gases have thermal conductivities which are dependent on their masses; therefore, a pure gas can be identified by its thermal conductivity. One way to determine the composition of a gas is to first separate it into its components and then measure the thermal conductivity of each. A gas chromatograph does exactly that. The gas flows through a long narrow column, which is packed with an adsorbant solid (for gas–solid chromatography) wherein the gases are separated according to the retentive properties of the packing material for each gas. As the individual gases exit the end of the tube one at a time, they flow over a heated wire. The amount of heat transferred to the gas depends on its thermal conductivity. The gas temperature is measured a short distance downstream and compared to a known gas flowing in a separate sensing tube. The temperature is related to the amount of heat transferred and can be used to derive the thermal conductivity according to thermodynamic theory and empirical data. This sensor required two transductions: a chemical to thermal energy transduction followed by a thermal to electrical transduction. ? 2000 by CRC Press LLC 56.4 Biosensors Biological measurands are biologically produced substances, such as antibodies, glucose, hormones, and enzymes. Biosensors are not the same as biomedical sensors, which are any sensors used in biomedical appli- cations, such as blood pressure sensors, or electrocardiogram electrodes. Many biosensors are biomedical sensors; however, they are also used in industrial applications, e.g., the monitoring and control of fermentation reactions. Table 56.1 does not include biological signals as a primary signal because they can be classified as either chemical or physical in nature. Biosensors are of special interest because of the very high selectivity of biological reactions and binding. However, the detection of that reaction or binding is often elusive. A very familiar commercial biosensor is the in-home pregnancy test sensor, which detects the presence of human growth factor in urine. That device is a nonelectrical sensor since the output is a color change which the eye senses. In fact, most biosensors require multiple transduction mechanisms to arrive at an electrical output signal. Two examples are given below: an immunosensor and an enzyme sensor. Rather than examine a specific species, the examples describe a general type of sensor and transduction mechanism, since the same principles can be applied to a very large number of biological species of the same type. Immunosensor Commercial techniques for detecting antibody-antigen binding utilize optical or x-radiation detection. An optically fluorescent molecule or radioisotope is nonspecifically attached to the species of interest in solution. The complementary binding species is chemically attached to a glass substrate or glass beads which are packed into a column. The tagged solution containing the species of interest, say the antibody, is passed over the antigen-coated surface, where the two selectively bind. After the specific binding occurs, the nonbound fluo- rescent molecules or radioisotopes are washed away, and the antibody concentration is determined by fluores- cence spectroscopy or with a scintillation counter, respectively. These sensing techniques are quite costly and bulky, and therefore other biosensing mechanisms are rapidly being developed. One experimental technique uses the change in the mechanical properties of the bound antibody-antigen complex in comparison to an unbound surface layer of antigen. It uses a shear mode, surface acoustic wave (SAW) device (see Chapter 51 and [Ballentine et al., 1997]) to sense this change as a change in the propagation time of the wave between the generating electrodes and the pick-up electrodes some distance away on the same piezoelectric substrate. The substrate surface is coated with the antigen and it is theorized that upon selectively binding with the antibody, this layer stiffens, changing the mechanical properties of the interface and therefore the velocity of the wave. The advantages of this device are that the SAW device produces an electrical signal (a change in oscillation frequency when the device is used in the feedback loop of an oscillator circuit) which is dependent on the amount of bound antibody; it requires only a very small amount of the antigen which can be very costly; the entire device is small, robust and portable; and the detection and readout method is inexpensive. However, there are numerous problems which currently preclude its commercial use, specifically a large temperature sensitivity and responses to nonspecific adsorption, i.e., by species other than the desired antibody. Enzyme Sensor Enzymes selectively react with a chemical substance to modify it, usually as the first step in a chain of reactions to release energy (metabolism). A well-known example is the selective reaction of glucose oxidase (enzyme) with glucose to produce gluconic acid and peroxide, according to An enzymatic reaction can be sensed by measuring the rise in temperature associated with the heat of reaction or by the detection and measurement of byproducts. In the glucose example, the reaction can be sensed by measuring the local dissolved peroxide concentration. This is done via an electrochemical analysis technique called amperometry [Bard and Faulkner, 1980]. In this method, a potential is placed across two inert metal CHO O gluconic acid HO kilojoules heat glucose oxidase 6126 2 22 80+? ?????? ++ ? 2000 by CRC Press LLC wire electrodes immersed in the test solution and the current which is generated by the reduction/oxidation reaction of the species of interest is measured. The current is proportional to the concentration of the reduc- ing/oxidizing species. A selective response is obtained if no other available species has a lower redox potential. Because the selectivity of peroxide over oxygen is poor, some glucose sensing schemes employ a second enzyme called catalase which converts peroxide to oxygen and hydroxyl ions. The latter produces a change in the local pH. As described earlier, an ISE can then be used to convert the pH to a measurable voltage. In this latter example, glucose sensing involves two chemical-to-chemical transductions followed by a chemical-to-electrical transduction mechanism. 56.5 Microsensors Microsensors are sensors that are manufactured using integrated circuit fabrication technologies and/or micro- machining. Integrated circuits are fabricated using a series of process steps which are done in batch fashion, meaning that thousands of circuits are processed together at the same time in the same way. The patterns which define the components of the circuit are photolithographically transferred from a template to a semiconducting substrate using a photosensitive organic coating. The coating pattern is then transferred into the substrate or into a solid-state thin film coating through an etching or deposition process. Each template, called a mask, can contain thousands of identical sets of patterns, with each set representing a circuit. This “batch” method of manufacturing is what makes integrated circuits so reproducible and inexpensive. In addition, photoreduction enables one to make extremely small features, on the order of microns, which is why this collection of process steps is referred to as microfabrication. The resulting integrated circuit is contained in only the top few microns of the semiconductor substrate and the submicron thin films on its surface. Hence, integrated circuit technology is said to consist of a set of planar, microfabrication processes. Micromachining refers to the set of processes which produce three-dimensional microstructures using the same photolithographic techniques and batch processing as for integrated circuits. Here, the third dimension refers to the height above the substrate of the deposited layer or the depth into the substrate of an etched structure. Micromachining produces third dimen- sions in the range of 1–500 mm (typically). The use of microfabrication to manufacture sensors produces the same benefits as it does for circuits: low cost per sensor, small size, and highly reproducible behavior. It also enables the integration of signal conditioning, compensation circuits and actuators, i.e., entire sensing and control systems, which can dramatically improve sensor performance for very little increase in cost. For these reasons, there is a great deal of research and development activity in microsensors. The first microsensors were integrated circuit components, such as semiconductor resistors and p-n junction diodes. The piezoresistivity of semiconductors and optical sensing by the photodiode were already discussed. Diodes are also used as temperature-sensing devices. When forward-biased with a constant diode current, the resulting diode voltage increases approximately linearly with increasing temperature. The first micromachined microsensor to be commercially produced was the silicon pressure sensor. It was invented in the mid-to-late 1950s at Bell Labs and commercialized in the 1960s. This device contains a thin silicon diaphragm (?10 mm) which is produced by chemical etching. The diaphragm deforms in response to a pressure difference across it (Fig. 56.3). The deformation produces two effects: a position-dependent displacement which is maximum at the diaphragm center and position-dependent strain which is maximum near the diaphragm edge. Both of these effects have been used in microsensors to produce an electrical output which is proportional to differential pressure. The membrane displacement is sensed capacitively as previously described in one type of pressure sensor. The strain is sensed in another by placing a piezoresistor, fabricated in the same silicon substrate, along one edge of the diaphragm. The two leads of the piezoresistor are connected to a Wheatstone bridge. The latter type of sensor is called a piezoresistive pressure sensor and is the commercially more common type of pressure microsensor. Pressure microsensors constituted about 5% of the total U.S. consumption of pressure sensors in 1991. Most of them are used in the medical industry as disposables due to their low cost and small, rugged construction. Many other types of microsensors are commercially under development, including accelerome- ters, mass flow rate sensors, and biosensors. ? 2000 by CRC Press LLC Defining Terms Micromachining: The set of processes which produce three-dimensional microstructures using sequential photolithographic pattern transfer and etching or deposition in a batch processing method. Microsensor: A sensor which is fabricated using integrated circuit and micromachining technologies. Repeatability: The ability of a sensor to reproduce output readings for the same value of measurand, when applied consecutively and under the same conditions. Sensitivity: The ratio of the change in sensor output to a change in the value of the measurand. Sensor: A device which produces a usable output in response to a specified measurand. Stability: The ability of a sensor to retain its characterisctics over a relatively long period of time. Related Topics 58.6 Smart Sensors ? 114.1 Introduction ? 114.2 Physical Sensors ? 114.3 Chemical Sensors ? 114.4 Bioan- alytical Sensors ? 114.5 Applications References ANSI, “Electrical Transducer Nomenclature and Terminology,” ANSI Standard MC6.1-1975 (ISA S37.1), Research Triangle Park, N.C.: Instrument Society of America, 1975. D. S. Ballentine, Jr. et al., Acoustic Wave Sensors: Theory, Design, and Physico-Chemical Applications, San Diego, Calif.: Academic Press, 1997. A. J. Bard and L. R. Faulkner, Electrochemical Methods: Fundamentals and Applications, New York: John Wiley & Sons, 1980. R. S. C. Cobbold, Transducers for Biomedical Measurements: Principles and Applications, New York: John Wiley & Sons, 1974. W. G?pel, J. Hesse, and J. N. Zemel, Eds., Sensors: A Comprehensive Survey, vol. 1, Fundamentals and General Aspects, T. Grandke and W. H. Ko, Eds., Weinheim, Germany: VCH, 1989. J. Janata, Principles of Chemical Sensors, New York, Plenum Press, 1989. FIGURE 56.3 Schematic cross section of a silicon piezoresistive pressure sensor. A differential pressure deforms the silicon diaphragm, producing strain in the integrated piezoresistor. The change in resistance is measured via a Wheatstone bridge. ? 2000 by CRC Press LLC Further Information Sensors: A Comprehensive Survey, W. Gopel, J. Hesse, and J. N. Zemel, editors. Weinheim, F.R.G. VCH, 1989–1994. Vol. 1: Fundamentals and General Aspects, T. Grandke and W. H. Ko, Eds. Vol. 2, 3, pt. 1–2: Chemical and Biochemical Sensors, W. Gopel et al., Eds. Vol. 4: Thermal Sensors, T. Ricolfi and J. Scholz, Eds. Vol. 5: Magnetic Sensors, R. Boll and K. J. Overshott, Eds. Vol. 6: Optical Sensors, E. Wagner, R. Dandliker, and K. Spenner, Eds. Vol. 7: Mechanical Sensors, H. H. Bau, N. F. deRooij, and B. Kloeck, Eds. J. Carr, Sensors and Circuits: Sensors, Transducers, and Supporting Circuits for Electronic Instrumentation, Measurement, and Control, Englewood Cliffs, N.J.: Prentice-Hall, 1993. J. R. Carstens, Electrical Sensors and Transducers, Englewood Cliffs, N.J.: Regents/Prentice-Hall, 1993. M. J. Usher and D. A. Keating, Sensors and Transducers: Characteristics, Applications, Instrumentation, Inter- facing, 2nd ed., New York: Macmillan, 1996. S. M. Sze, Ed., Semiconductor Sensors, New York: John Wiley & Sons, 1994. D. Tandeske, Pressure Sensors: Selection and Application, New York: Marcel Dekker, 1991. Sensors and Actuators is a technical journal devoted to solid-state sensors and actuators, which is published bimonthly by Elsevier Press in two volumes: Vol. A: Physical Sensors and Vol. B: Chemical Sensors. The International Conference on Solid-State Sensors and Actuators is held every 2 years, hosted in rotation by the U.S., Japan, and Europe. It is sponsored in part by IEEE in the U.S. and a digest of technical papers is published and available through IEEE. ? 2000 by CRC Press LLC