Optical Fiber Technology 9 (2003) 57–79 www.elsevier.com/locate/yofte Invited paper Review of the present status of optical fiber sensors Byoungho Lee School of Electrical Engineering, Seoul National University, Kwanak-Gu Shinlim-Dong, Seoul 151-744, South Korea Received 28 August 2002 Abstract The current status of optical fiber sensors is reviewed. The optical fiber sensors have certain advantages that include immunity to electromagnetic interference, lightweight, small size, high sensitivity, large bandwidth, and ease in implementing multiplexed or distributed sensors. Strain, temperature and pressure are the most widely studied measurands and the fiber grating sensor represents the most widely studied technology for optical fiber sensors. Fiber-optic gyroscopes and fiber-optic current sensors are good examples of rather mature and commercialized optical fiber sensor technologies. In this paper, among the various fiber-optic sensor technologies, especially, technologies such as fiber grating sensors, fiber-optic gyroscopes, and fiber-optic current sensors are discussed with emphasis on the principles and current status. Today, some success has been found in the commercialization of optical fiber sensors. However, in various fields they still suffer from competition with other mature sensor technologies. However, new ideas are being continuously developed and tested not only for the traditional measurands but also for new applications. 2003 Elsevier Science (USA). All rights reserved. Keywords: Optical fiber sensor; Fiber grating; Fiber-optic gyroscope; Optical fiber current sensor 1. Introduction Almost thirty years have passed since the study of optical fiber sensors began. Various ideas have been proposed and various techniques have been developed for various measurands and applications. To date, some types of optical fiber sensors have been commercialized, but it is also true that, among the various techniques that have been studied, only a limited number of techniques and applications have been commercially successful. Optical fiber sensors have advantages such as immunity to electromagnetic interference E-mail address: byoungho@snu.ac.kr. 1068-5200/03/$ – see front matter 2003 Elsevier Science (USA). All rights reserved. doi:10.1016/S1068-5200(02)00527-8 58 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Fig. 1. Distribution of OFS-15 papers according to measurands. Papers not directly related to measurands, such as those that discuss fiber grating sensor interrogators, multiplexing, light sources, etc., are not included in the statistics. If one paper deals with more than one measurand, the counted paper number is equally divided among the measurands. Special session papers on fiber-optic gyroscopes are included in the statistics. (EMI), lightweight, small size, high sensitivity, large bandwidth, and ease in signal light transmission. However, in many fields of application, optical fiber sensors should compete with other rather mature technologies such as electronic measurements. To appeal to users already accustomed to other mature technologies, the superiority of optical fiber sensors over other techniques needs to be clearly demonstrated. Typical users are not interested in specific techniques involved in measurement. They simply desire sensor systems having good performances with reasonable price except for very special uses. Hence, optical fiber sensor systems should be available in the form of complete systems including detecting and signal-processing electronics. In some cases such as electric protection relaying systems, the sensor systems are simply subsystems of rather larger systems. In some cases such as optical gyroscopes and optical current sensors, the optical fiber sensors should compete with other optical bulk sensors as well. Even with these difficulties, considerable efforts have been made to study of optical fiber sensors, and some of them are now nearing maturity. There have been excellent review books and articles on optical fiber sensors such as Refs. [1–5]. It is impossible to review all of the fiber-optic sensor technologies in a journal article with a limited length. In this paper, the current status of some of the actively studied or well-developed optical fiber sensors will be reviewed. Figure 1 shows the distribution of papers presented at the 15th Optical Fiber Sensors Conference (OFS15: IEEE catalog number 02EX533) according to measurands of interest. The conference was held in Portland, Oregon, USA in May 2002, and is a major conference in the field of B. Lee / Optical Fiber Technology 9 (2003) 57–79 59 Fig. 2. Distribution of OFS-15 papers according to technologies. If one paper deals with more than one technology, the counted paper number is equally divided among the technologies. Special session papers on fiber-optic gyroscopes (FOGs) are included in the statistics. optical fiber sensors. Hence the statistics of OFS-15 provide information on recent research trends in the area of optical fiber sensors. The most highly discussed measurands are strain and temperature, the same as in the conference of five years ago [6]. Figure 2 shows the technologies involved in the optical fiber sensors presented at OFS-15. Fiber grating sensors are clearly the most widely studied topic. In Section 2 fiber grating sensor technology, which is the most popular topic in optical fiber sensors, is reviewed. In Sections 3 and 4 two rather mature topics—fiber-optic gyroscopes (FOGs) and fiber-optic current sensors—are reviewed. In Section 5, other sensors are briefly discussed, which is followed by some concluding remarks. 2. Fiber grating sensors Although the formation of fiber gratings had been reported in 1978 [7], intensive study on fiber gratings began after a controllable and effective method for their fabrication was devised in 1989 [8]. Fiber gratings have been applied to add/drop filters, amplifier gain flattening filters, dispersion compensators, fiber lasers and so on for optical communications [4]. Extensive studies have also been performed on fiber grating sensors and some of which have now reached commercialization stages. Figure 3 shows types of fiber gratings. Under phase matching conditions, a fiber Bragg grating (FBG) couples the forward propagating core mode to the backward propagating core mode. A long-period fiber grating (LPG) can couple the forward propagating core mode to one or a few of the forward propagating cladding modes. A chirped fiber grating has a wider reflection spectrum and each wavelength component is reflected at different po- 60 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Fig. 3. Types of fiber gratings. (a) Fiber Bragg grating, (b) long-period fiber grating, (c) chirped fiber grating, (d) tilted fiber grating, (e) sampled fiber grating. sitions, which results in a delay time difference for different reflected wavelengths. A tilted fiber grating can couple the forward propagating core mode to the backward propagating core mode and a backward propagating cladding mode. A sampled fiber grating can reflect several wavelength components with equal wavelength spacing. All these types of gratings have been utilized in various types of fiber grating sensors and wavelength change interrogators. Among them, however, FBGs are the most widely used as sensor heads. In FBGs the Bragg wavelength λB , or the wavelength of the light that is reflected, is given by λB = 2neff Λ, (1) B. Lee / Optical Fiber Technology 9 (2003) 57–79 61 where neff is the effective refractive index of the fiber core and Λ is the grating period. In Eq. (1) it can be seen that the Bragg wavelength is changed with a change in the grating period or the effective refractive index. The former is the case for strain and the latter for temperature variation. The grating period can also be changed with temperature variation, but around room temperature the effect of temperature on refractive index is about one order of magnitude larger than that of thermal expansion (or contraction). The typical response of the Bragg wavelength shift to strain is ∼0.64 pm/µε (µε = microstrain) near the Bragg wavelength of 830 nm, ∼1 pm/µε near 1300 nm, and ∼1.2 pm/µε near 1550 nm [9]. The unit of strain is ε, which is read as “strain”. It is a relative concept, that is, if a 1-mlong fiber is elongated by 1 µm, the strain is 1 µm/1 m = 1 µε. The typical temperature response is ∼6.8 pm/◦ C near 830 nm, ∼10 pm/◦ C near 1300 nm, and ∼13 pm/◦ C near 1550 nm [9] although the values depend on FBG types [10]. (These FBG types should not be confused with fiber grating types of Fig. 3. Detailed information can be found in Ref. [4].) This wavelength-encoded measurand information is a unique characteristic of FBGs. In addition to the common advantages of fiber sensors, this wavelength-encoded characteristic provides robustness to noise or power fluctuation and also enables wavelength division multiplexing (WDM). Hence multi-point sensors can be realized using this technique. Sometimes WDM is combined with spatial division multiplexing (SDM), time division multiplexing (TDM) and code division multiple access (CDMA) techniques. However, it must be admitted that multi-point FBG sensors are not fully-distributed sensors such as backscattering sensors or optical time-domain reflectometers (OTDRs) because physical parameters can be measured only at the FBG positions. As discussed above, FBGs respond to both strain and temperature. Hence their effects need to be separated from each other, in order to measure each physical parameter or to measure both simultaneously. Considerable effort has been focused on this topic and various solutions have been proposed. A straight forward and very practical approach is to use a reference grating [11]. Another FBG (the reference grating), which is isolated from one parameter, e.g., strain, is placed near the sensor FBG. The reference grating can be on the same fiber as the sensor FBG [12]. Another method is to use two FBGs with much different Bragg wavelengths (in general, two light sources are needed), which show different responses to the same measurands [13]. FBGs written on different-diameter fibers have also been proposed, which give different strain responses, while the temperature responses are the same [14,15]. Demodulators or interrogators are required for FBG sensors. Their roles are to extract measurand information from the light signals coming from the sensor heads. The measurand is typically encoded in the form of a Bragg wavelength change, and hence, the interrogators are typically expected to read the wavelength shift and provide measurand data. Optical spectrum analyzers are not suitable for real sensor systems because they are expensive and their wavelength scanning speed is too slow. Various techniques have been developed for the interrogators [16]. Some are quite simple but are more limited in measurement resolution, dynamic range or multiplexing, and some are more complicated and provide better resolution but are more expensive or need stabilization. Not all of them are appropriate for commercialized systems. Table 1 summarizes interrogator types and Table 2 summarizes their performances [17–62]. Many of these are laboratory experiment results. Civic structures such as bridges require the measurement of dynamic strain. When 62 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Table 1 Fiber grating sensor interrogator types Types Technologies References Passive detection schemes Linearly wavelength-dependent device CCD spectrometer Power detection Identical chirped-grating pair [17–21] [22–25] [26–29] [30] Active detection schemes Fabry–Perot filter Unbalanced Mach–Zehnder interferometer Fiber Fourier transform spectrometer Acousto-optic tunable filter Matched FBG pair Michelson interferometer LPG pair interferometer [31–33] [34–43] [44,45] [46–48] [49–52] [53] [54] Other schemes Wavelength-tunable source Mode-locked fiber laser with wavelength-time conversion Optical CDMA correlator Frequency modulation Intra-grating sensing [55–57] [58] [59,60] [61] [62] Table 2 Strain measurement resolution of some fiber grating sensor systems/interrogators Static/quasi-static resolution ±20 µε ∼12 µε 4.12 µε ∼ ±3.5 µε 3.5 µε ∼ ±3 µε 3 µε ±2 µε 1.9 µε ±0.25 µε 40 nε ±8.04 µε Dynamic resolution √ 1.5 µε/√Hz 6 nε/√Hz 0.5 µε/ Hz √ 5 nε/√Hz 7 pε/ Hz √ 42 nε/√Hz 10 nε/√Hz 10 nε/√Hz 2 nε/√Hz 1.5 nε/√Hz ∼7 fε/ Hz References (measurement range) [58] (> 3500 µε) [44] [49] [18] (∼16,000 µε) [40] (2 kHz) [20] (>1050 µε) [50] [32] (±1300 µε) [55] [25] [54] [56] (±1000 µε) [57] [43] [51] (> ±100 µε) [35] (>10 Hz) [42] (10–2000 Hz) [36] the dynamic signal is measured, the minimum detectable strain is determined by the background noise-level. The magnitude of noise changes with the frequency span because the noise power in the detector and amplifying circuit depends on the frequency span. Generally, the amount of noise magnitude in other bandwidths can be approximated by B. Lee / Optical Fiber Technology 9 (2003) 57–79 63 (a) (b) Fig. 4. Some typical interrogators for FBG sensors. (a) Linearly wavelength-dependent device (coupler) method [20]. (b) CCD spectrometer method [24]. (c) Fabry–Perot filter method [32] (FFP: fiber Fabry–Perot filter). (d) Unbalanced Mach–Zehnder interferometer (pseudo-heterodyne detection) method [37] (OPD: optical path length difference, PZT: piezoelectric transducer, BPF: band pass filter). scaling the power spectral density of the noise by the square root of the bandwidth. Thus, √ the normalized minimum detectable dynamic strain is displayed in units of ε/ Hz. Figure 4 shows some typical types of interrogators. Figure 4a is an example of interrogators that use linearly wavelength-dependent devices (couplers or filters) [20]. The Bragg wavelength shift is monitored by the detected power change. This structure is quite simple and has been commercialized, but it is not suitable for multiplexed sensors. In fact, 64 B. Lee / Optical Fiber Technology 9 (2003) 57–79 (c) (d) Fig. 4. (Continued.) this type of power-detection method does not fully use one of the key advantages of FBG sensors—measurand information is coded as wavelength change and the information is not affected by fluctuations in light power. In Fig. 4a to remove the light power (or loss) fluctuation effect, a power-ratio detection technique is used. Figure 4b shows the CCD spectrometer interrogator [24], which takes advantage of the fact that the diffraction angle is dependent on the wavelength of the light incident on a diffraction grating. The structure shown in Fig. 4b adopts both WDM and SDM. Compared with other interrogators such B. Lee / Optical Fiber Technology 9 (2003) 57–79 65 as those adopting scanning filters and TDM, the advantage of the CCD spectrometer interrogator is its light-effectiveness, that is, a high signal-to-noise ratio can be obtained even with low input optical power. A one-dimensional CCD can be used instead of the twodimensional CCD if TDM is used instead of SDM [25]. Figure 4c shows an example of the scanning fiber Fabry–Perot filter interrogator system. U.S. Naval Research Laboratory tested this for a 64-FBG sensor system [32]. The basic principle is to locate the WDM Bragg wavelengths within the free spectral range of the Fabry–Perot filter and monitor their changes by spectrally-scanning the filter. In real systems, generally, the dithering of the filter spectrum is adopted for a fine measurement in the case of a single FBG sensor and the zero-crossing of the time derivative of the received power is monitored in the case of WDM FBG sensors. Figure 4d shows an unbalanced Mach–Zehnder interferometer interrogator system adopting a pseudo-heterodyne detection technique [37]. A detailed description of the principle of operation can be found in Ref. [16]. The unbalanced Mach–Zehnder interferometer provides very high-resolution interrogation. However, this interferometer is quite sensitive to its environment such as temperature change or even air fluctuation. Hence, a reference grating for the interrogator system (this should not be confused with the reference grating for the sensor head to distinguish the temperature and strain effect) is required as shown in Fig. 4d and the interrogator should be well-stabilized. FBG sensors have been tested in various fields such as bridges, dams, mines, composite laminates, airplanes, generators, ship waterjets, geotechnical fields, and railway systems [25,43,63–66]. Some FBG sensors have been commercialized for structural monitoring [67] and oil/gas reservoir monitoring [68]. For example, CiDRA’s FBG pressure/temperature sensor was reported to show good functionality at 150 ◦ C, 5000 psi for over two years with output drift within 0.015% full scale/year [68]. Such sensors have been deployed in oil/gas wells—for example, in a 15,000-feet downhole [68]. 3. Fiber-optic gyroscopes It is generally recognized that the first demonstration of the FOG was achieved 27 years ago [69], though there had been a few previous studies. The basic concept is based on the Sagnac interferometer and is quite simple. It is a good example of an application of special relativity. For a rotating optical fiber coil, as shown in Fig. 5, two lights traveling in opposite directions in the coil experience different lengths, which results in different travel times and a phase difference in the two optical waves. Simple analyses [70,71] show that the phase difference φ is given by 8πN A · , (2) λc where N is the number of coil turns, λ the wavelength in vacuum, c the speed of light in vacuum, A the area vector of the fiber coil (Its magnitude is the area enclosed by the singleturn coil and direction is normal to it.), and the rotating rate (angular frequency) vector. Like the output of typical interferometers, the output response of the FOG is a raised cosine function with respect to the phase difference. In the response curve, if the operating point is at maxima (or minima) such as the point corresponding to φ = 0, the sensitivity φ= 66 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Fig. 5. Principle of the fiber-optic gyroscope. In this figure, the fiber coil is rotating in the clockwise (CW) direction. Hence, for the CW and counter-clockwise (CCW) rotating lights to meet at the same exit point to interfere with each other, they should have entered the coil at different times at different points in the figure. Different entering times mean different entering points because the fiber coil is rotating. is low and the rotational direction of the fiber coil cannot be distinguished due to the symmetric response. Hence, the operating point is shifted to a position where the response slope is not zero. This is accomplished by the use of a phase modulator placed at the end of the fiber coil. There are two types of basic configurations in which a FOG can be constructed; one is an open-loop configuration and the other is a closed-loop configuration. As shown in Fig. 6a, a dynamic phase modulation is done at one end of the fiber coil, such that two lights that start at the coupler at the same time and rotate the fiber coil in opposite directions experience different optical path lengths within the phase modulator (such as a fiber wound around a piezoelectric transducer (PZT) tube) due to their different arrival times at the modulator [72]. This open-loop structure has a limited dynamic range [71]. In the closed-loop configuration a feedback signal is applied to null the output signal and maintain the operation of the system in the linear range [73,74]. Figure 6b shows such an example using an integrated optics chip (IOC) to modulate the phase. The IOC is typically a lithium niobate modulator, which has a larger bandwidth than the PZT and allows a proper modulation for the feedback control. When a FOG is at rest, the output signal is a random function that is the sum of a white noise and a slowly varying function. The photon shot noise is an important source of white noise and limits the fundamental accuracy of the FOG. The white noise is expressed in terms of the standard deviation of the equivalent rotation rate per square root of bandwidth of detection (like the case of the dynamic strain measurement resolution of fiber grating sensors), in other words, the equivalent noise power spectral density in the √ unit of deg /h/ Hz. The white noise causes noise in measured rotation rate output and √ results in an angle random walk (ARW) in the unit of deg / h. The slowly varying noise can be induced by several sources and results in long-term drift or bias instability. B. Lee / Optical Fiber Technology 9 (2003) 57–79 67 (a) (b) Fig. 6. Basic configurations of FOG. (a) Open-loop type (SLD: superluminescent diode. PM: phase modulation). (b) An example of a closed-loop type (IOC: integrated optics chip modulator). The performance of a FOG has been continuously improved and is now quite mature and capable of meeting the most accurate gyroscope requirements. As can be seen in Fig. 7, the FOG has competing technologies such as mechanical gyroscopes and ring laser gyroscopes. The cost of the FOG has been constantly falling in recent years thanks to economies of scale and the dramatic explosion in the optical fiber communications market (component prices are decreasing) [75]. Currently, FOGs are considered to be the most cost-efficient solution for various inertial navigation applications. One important advantage of the FOG is its ruggedness. It contains no moving parts unlike other competitors such as mechanical gyroscopes and ring laser gyroscopes (dithering is involved in ring laser gyroscopes). Due to these unique advantages, it seems likely that the FOG will play a significant role in both military and commercial markets. 68 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Fig. 7. Comparison of gyroscope technologies and applications (MEMS: micro-electro-mechanical system technology, DTG: dynamically tuned gyroscope (mechanical type), FOG: fiber-optic gyroscope, RLG: ring-laser gyroscope, AHRS: attitude heading reference systems). Some recent reports indicate that Honeywell Space Systems currently manufactures high-precision FOGs with a flicker noise below 0.0001 deg/h and an ARW of ∼80 √ µdeg/ h [76] and Ixsea manufactures high-level performance FOGs with a temperature bias stability √ better than 0.01 deg/h (0.003 deg/h at ambient) for a noise level of 0.001 deg/ h [75]. FOGs have been developed by many companies and have been sold or applied in various fields such as commercial aircraft attitude heading reference systems, military helicopters, missile guidance systems, marine, and submarine navigation systems, pipe-mapping (gas, power, and communication cables), compasses for tunnel construction, rockets, and automotive navigation systems [71,75–81]. 4. Fiber-optic current sensors With the growth in the capacity of electric power systems, the role of protection relaying systems is becoming more important. Such a system can immediately recognize any sudden failures, such as a surge, and separate the failure parts from the power systems. These relaying systems require current sensors, referred to as current transformers or current transducers (CTs). Most CTs currently in use are electromagnetic devices that suffer from magnetic saturation effects and residual field effects. Moreover, with the superincrease in voltages (several hundred kV) in power distribution systems, the insulation of the CTs becomes more difficult and expensive. Hence, optical current sensors that do not suffer from electromagnetic interference are good substitutes for conventional CTs. In addition to protection relaying systems, the deregulation and growth of independent power producers and regional transmission companies have created a need for many new B. Lee / Optical Fiber Technology 9 (2003) 57–79 69 high voltage revenue-metering points [82]. In the transfer of power from a generation company to a regional transmission company, a 0.5% uncertainty in metering may result in an uncertainty of millions of dollars per year at a high power metering location [82]. Hence the potential use of optical CTs that provide high accuracy is promising. The concept of the optical CTs is quite old [83,84] and based on the Faraday rotation effect, which states that the polarization of light waves is rotated with the propagation of the light along (or opposite to) a magnetic field inside some material. Optical fiber CTs are not the only optical CTs. Bulk-optic CTs such as flint glass closed-loop type and crystalbased (Faraday cell) CTs have also been extensively studied and tested. Some examples of characteristics of bulk devices are as follows—{material, Verdet constant [deg/Oe cm] at wavelength [µm], temperature coefficient}: {Flint (Pb) glass, 0.84 × 10−3 at 0.85 or 1.1 × 10−3 at 0.63, < ±0.5% (−25–100 ◦ C)}, {BGO, 0.31 × 10−2 at 0.85, < ±1.5% (−25–85 ◦ C)}, {FR-5 glass, 0.21 × 10−2 at 0.85, < ±15% (−25–85 ◦ C)}, {YIG, 0.27 at 1.30, < ±7.0% (−25–85 ◦ C)}, {RIG, 7.0 at 0.83, < ±0.5% (−20–80 ◦ C)}. Here the Verdet constant indicates the rotation angle of the polarization per unit magnetic field per unit propagation length and 1 Oe = 79.6 A/m. Compared with the bulk devices, the Verdet constant of optical fibers is quite small, 1.2 × 10−4 deg/Oe cm (at 850 nm) to 2.13 × 10−4 deg/Oe cm (at 633 nm), but the optical path length can be increased to compensate for it by winding the fiber around a current conducting element a large number of turns. Although the bulk-optic CTs provide better mechanical stability and smaller sizes, optical fiber CTs provide ease in forming a closed-loop (to make the sensor respond only to the enclosed current), adjustability of the sensitivity and dynamic range, less insertion loss, and a higher signal-to-noise ratio. In optical fiber CTs, linearly polarized optical waves are input to the optical fiber coil. The linear polarization can be expressed mathematically as a superposition of two circular polarizations (right-hand and left-hand). The magnetic field induced around a currentcarrying element induces a circular birefringence inside the optical fiber coil. Hence, after passing through the coil, a relative phase difference between two circular polarization components is generated, which results in the rotation of the linear polarization angle in proportion to the enclosed current and the number of fiber turns. Although the concept is quite simple, in actual implementation, several difficulties that limit the resolution of the sensor systems are encountered. Optical fibers have some linear birefringence due to the imperfection of the core shape, which may distort the output signals of optical fiber CTs. This linear birefringence can be reduced considerably by annealing [85]. But, it has a difficulty in practice because the annealing is generally done after the winding process of the fiber coil. An inevitable linear birefringence, induced by bending of optical fiber around the current-carrying element, exists. Moreover, the effects of vibration, mechanical stress or strain and temperature variation on the linear birefringence are also critical. In some applications, high-current-carrying conductors vibrate. The vibration in such a direction to rotate the fiber coil is more critical because the fiber coil acts as a FOG in that case [86]. Several techniques have been proposed to cope with these problems. If a birefringence bias corresponding to 258-degree phase difference is applied, the effect of linear birefringence variation can be minimized [87]. However, the sensitivity is reduced by a factor of 0.217. Tokyo Electric Power Company has been developing current sensors using flint 70 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Fig. 8. Reflection-type fiber-optic current sensor. glass fibers that show a low photoelastic coefficient (4.5×10−10 cm2 /kg) and a high Verdet constant (∼0.065 min/Oe cm) [88], but they suffer from higher loss, difficulty in splicing with other fibers and higher cost. A recent report indicates that the system was successful in hammering and thermal shock tests for a differential current relaying system [86]. The use of twisted optical fiber has also been proposed, in which a high circular birefringence is induced, thus reducing the effect of linear birefringence [89]. A high birefringent spun fiber was also proposed for this application, which is pulled from a rotating silica preform during the fiber draw process [90]. In most of the today’s optical fiber CTs these types of methods of increasing circular birefringence are in use. In practice, the linear birefringence and intentionally induced circular birefringence may vary with time due to vibration and temperature variation. A simple and efficient way to reduce this effect is to use a reflection type structure as shown in Fig. 8. The intentionally induced circular birefringence is reciprocal with the reversal of light propagation direction, while the Faraday rotation is nonreciprocal. Hence if a light traveling along the fiber coil is reflected by a mirror and then travels back along the fiber coil, the Faraday rotation angle is doubled, while the effect of the intentionally induced circular birefringence is cancelled. The structure has another advantage, in that it can be easily mounted around the currentcarrying element because one end of the fiber coil is open. However, if a conventional mirror is used, the effect of linear birefringence is also doubled. However, a Faraday rotator mirror can be utilized to minimize the linear birefringence effect, in which case the two orthogonal linear polarization components are switched at the Faraday rotator mirror and experience the same amount of phase delay (despite the linear birefringence) after the round trip [90,91]. Toshiba reported that, for its reflection type optical CT, the influence of mechanical shock from circuit breaker vibration of 8g (g = gravitational acceleration) was less than the rms value of the electronic noise [92]. Sagnac interferometer type optical B. Lee / Optical Fiber Technology 9 (2003) 57–79 71 Fig. 9. An example [94] of a high-performance fiber-optic current sensor using technology developed for FOG (IOC: integrated optics chip modulator, PC: polarization controller). fiber CTs are also insensitive to reciprocal effects [93]. Several techniques that have been devised for FOGs can be applied in these optical fiber CTs. With the structure shown in Fig. 9, EPFL has achieved ±0.2% accuracy over the ±500 kA range for a temperature range between 0 and 50 ◦ C [94]. The Sagnac structure using a 3 × 3 fiber optic coupler was also tested. This structure enables the interferometer to be optimally biased at all times [95]. Some other recent reports indicate that Siemens has tested an optical fiber CT with an annealed fiber coil for a generator with a measurement accuracy error below 0.25% within the temperature range of 40–90 ◦ C [96] and NxtPhase showed a linearity performance better than ±0.2% from 0.5 to 120% of the rated current [82]. ABB Corporate Research has developed a robust fiber-optic current sensor with inherent temperature compensation of the Faraday rotation effect [97]. Insensitivity of the sensor to temperature within 0.2% was demonstrated between −35 and 85 ◦ C. 5. Others Some other types of optical fiber sensors based on very simple concepts have been commercialized. Some examples include the displacement or pressure sensor based on the light coupling of two fibers, the liquid-level sensor based on frustrated total internal reflection, the pressure sensor using a wiggled (periodically bent) fiber, and the temperature sensor based on the detection of radiation from a heated sensor head (blackbody cavity) [98]. One of the most well-developed and commercialized in-line fiber sensors or diagnostic tools is OTDR, which is based on the monitoring of the backscattering along the fiber. The concept is also quite old [99] and the OTDR has become a standard technique for testing optical fiber links [100]. It typically provides sub-meter spatial resolution but improved techniques can provide mm-order resolution. More complicated methods such as optical frequency domain reflectometry have been studied to achieve mm or sub-millimeter (∼10 µm) spatial resolution [101]. While OTDRs are generally aimed at monitoring optical fiber communication links, active research has also been done on distributed sensors for civil structure monitoring. One 72 B. Lee / Optical Fiber Technology 9 (2003) 57–79 example is a distributed temperature sensor for monitoring concrete setting temperatures of a large dam [102]. The principle is based on the stimulated Brillouin scattering. An acoustic wave couples two counter-propagating beams, which are frequency-shifted by an amount that is dependent on temperature or strain. A commercialized distributed temperature and strain monitoring system shows the measurement range of up to 20 km with a spatial resolution of 0.8 m over 1 km (2–5 m at 10 km and 5–15 m at 20 km), a strain resolution of 20 µε (measurement range of up to 25 mε) and a temperature resolution of 1 ◦ C [103]. Optical fiber acoustic sensors or optical fiber hydrophone systems have also been intensively studied and tested for marine or submarine applications [104]. Fiber-optic lowcoherence interferometry has also been commercialized for civil applications [105]. The dynamic range of the commercialized deformation sensor using the method is of the order of a few tens of mε for elongation (∼5 mε for shortening) [103]. A fiber optic pressure sensor using the fiber Fabry–Perot interferometer method was also commercialized, which shows a dynamic range of 0–3000 psi with ±1.0% (of full range) accuracy and ±0.5% (of full range) linearity [106]. There are other examples of commercialized sensors using Fabry–Perot interferometers (or thin film Fizeau interferometers) and white-light crosscorrelators: a force and load transducer (load capacity: 4500 kg, resolution: 0.01% of full scale); strain gauges (range: ±500 to ±5000 µε, resolution: 0.01% of full scale); a temperature transducer (range: −40 to 350 ◦ C, accuracy: ±0.5 to ±1 ◦ C or 0.5–1% of full scale); a displacement transducer (linear stroke: 20 mm, resolution of up to 2 µm) [107]. Cryogenic temperature sensing is also an attractive field for fiber-optic sensors [108]. Fiber-optic chemical sensors or biosensors have been continuously studied. The principles are based on the monitoring of absorbance, reflectance, luminescence, reflective index change, or light scattering and aimed at the measurement of oxygen, pH, carbon dioxide, ammonia, detergents, biochemical oxygen demand, pesticides, and humidity [5,109]. In many cases optical fibers are simply used to guide light to the measurement point in the specimen. In some cases optical fibers are monitoring the response of a material deposited on the end of the fiber [110]. In some other cases well-developed sensor technologies such as FBG temperature sensors have been used for biological tests [111]. A good example of optical fiber chemical sensors in which the fiber itself plays a key role in the measurement is the use of LPGs [112,113]. As discussed in Section 2, an LPG can couple the core mode to a cladding mode. If the coating or jacket is removed from the optical fiber, the evanescent field of the cladding mode that exists outside the cladding experiences the refractive index change of the outside material. The sensitivity can be adjusted by fiber etching [114]. An example of a commercialized fiber biosensor is the invivo blood pressure sensor that uses the white light interferometry (pressure range: 500 mm Hg to 1060 mm Hg, precision: ±1 mm Hg or ±1% of reading) [107]. 6. Concluding remarks In this paper the current status of optical fiber sensors has been briefly reviewed. As soon as optical fibers were developed, it was concluded that they could also be used for sensors. Hence, the history of research in optical fiber sensors is almost as old as the history of optical fiber communication research. Optical fiber sensors have unique advantages such B. Lee / Optical Fiber Technology 9 (2003) 57–79 73 as high sensitivity, immunity to EMI, small size, lightweight, robustness, and the ability to provide multiplexed or distributed sensing. Although the optical fiber sensors have not experienced the dramatic commercial success of optical fiber communications, they have been continuously and enthusiastically studied. At present, the economic situation is not good. One of the gloomy forecasts predicts that the worldwide optical transport market will decline through the year of 2004 before resuming modest growth in 2005 (by Dell’Oro Group) [115]. It says the 2004 sales will be only 35% of the peak in 2000. Some other analyst groups predict more rapid economic recovery. Strategies Unlimited predicts that the optical component market will start the recovery from 2003. Communications Industry Researchers (CIR) also predicts that the long haul market will start its recovery from 2003. It also predicts that the market of integrated optics for multiplexing, switching, and computing will grow to $400 millions by 2006 [116]. These forecasts are on the market related to optical communications. However, the competitiveness of optical fiber sensors relies on the component cost reduction, which is greatly dependent on the optical communication market. The sensor forecast of 2006– 2011 published in April of 2002 by the Freedonia Group predicts that U.S. market for sensors will grow at a healthy pace, fueled by economic recovery in more mature applications such as process control, industrial machinery, and conventional automotive sensors. It also predicts that the fastest growth will occur in emerging areas such as microelectro-mechanical systems and telematics, and that nearly all types of opto-electronic sensors including fiber optic sensors will exhibit healthy gains [116]. Recently Honeywell Automation and Control Solutions has signed an agreement to acquired Invensys Sensor Systems [117], and General Electric has employed an aggressive strategy for entering the sensor’s arena such as temperature, humidity, pressure, gas, and oxygen sensors [118]. Although the sensors involved here are not confined to optical fiber sensors, these trends are encouraging to the researchers in the field of optical fiber sensors. Fiber grating sensors have been the most widely studied topic among various optical fiber sensor technologies. Some fiber grating sensors have been commercialized for civil health monitoring and oil industries. FBGs have been shipping for WDM devices since 1997 [119]. But for WDM filter applications, FBGs have strong competitors such as arrayed-waveguide gratings and thin-film filters. Several FBG vendors have recently exited the market [119]. However, for sensor applications, FBGs provide a unique characteristic of multi-point EMI-resistant measurement, and they have started to be deployed in civil and oil industry fields. There are other rather mature optical fiber sensor technologies such as OTDRs, FOGs, and optical fiber current sensors. The OTDR has become a standard technique in optical communication line fault monitoring. Today, FOGs are considered as the most costefficient solution for various inertial navigation applications. Optical current and voltage sensors became to be installed in real power industry [120]. New ideas are continuously being proposed and tested for not only various traditional measurands such as strain, temperature, and pressure but also for new applications such as biosensors. Papers on optical fiber sensors have already begun to appear in many fieldapplication journals such as civil engineering journals, sensor journals, power engineering journals, chemical engineering journals, and bioengineering journals. In 2002, the number of papers on optical fiber sensors published in the journals listed in Scientific Citation 74 B. Lee / Optical Fiber Technology 9 (2003) 57–79 Index (SCI) is about 110. Many of the papers are published in the field-application journals. Hence, the research in the optical fiber sensors is more active than optics people feel from the journals of their main interest. The U.S. patent applications on optical fiber sensors during recent two years records several hundreds in number, among which temperature, strain, displacement, refractive index, damage detection, current, vibration, biological and chemical sensors are frequent topics [121]. Biosensors for medical applications will be an important application field of optical fiber sensors. Now, the optical fiber sensors reached the stage at which optics people should collaborate with specialists in other fields such as civil engineering, mechanical engineering, material engineering, chemical engineering, and bioengineering to develop sensor systems useful in real fields and to find new applications. With the continuing effort and enthusiasm, the application fields of optical fiber sensors will be extended and more commercial successes will almost certainly be found. Acknowledgment The author thanks Dr. Minho Song for valuable comments on this paper. References [1] F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002. [2] K.T.V. Grattan, B.T. Meggitt (Eds.), Optical Fiber Sensor Technology, Vol. 2—Devices and Technology, Chapman & Hall, London, 1998. [3] K.T.V. Grattan, B.T. Meggitt (Eds.), Optical Fiber Sensor Technology, Vol. 3—Applications and Systems, Kluwer Academic Publishers, Boston, 1999. [4] A. Othonos, K. Kalli, Fiber Bragg Gratings—Fundamentals and Applications in Telecommunications and Sensing, Artech House, Boston, 1999. [5] M.D. Marazuela, M.C. Moreno-Bondi, Fiber-optic biosensors—an overview, Anal. Bioanal. Chem. 372 (2002) 664–682. [6] Z.Y. Zhang, K.T.V. Grattan, Commercial activity in optical fiber sensors, in: K.T.V. Grattan, B.T. Meggitt (Eds.), Optical Fiber Sensor Technology, Vol. 3—Applications and Systems, Kluwer Academic Publishers, Boston, 1999, Chap. 10, pp. 257–306. [7] K.O. Hill, Y. Fujii, D.C. Johnson, B.S. Kawasaki, Photosensitivity in optical fiber waveguides: application to reflection filter fabrication, Appl. Phys. Lett. 32 (1978) 647–649. [8] G. Meltz, W.W. Morey, W.H. Glenn, Formation of Bragg gratings in optical fibers by a transverse holographic method, Opt. Lett. 14 (1989) 823–825. [9] Y.-J. Rao, Fiber Bragg grating sensors: principles and applications, in: K.T.V. Grattan, B.T. Meggitt (Eds.), Optical Fiber Sensor Technology, Vol. 2, Chapman & Hall, London, 1998, pp. 355–389. [10] X. Shu, Y. Liu, D. Zhao, B. Gwandu, F. Floreani, L. Zhang, I. Bennion, Dependence of temperature and strain coefficients on fiber grating type and its application to simultaneous temperature and strain measurement, Opt. Lett. 27 (9) (2002) 701–703. [11] M.G. Xu, J.L. Archambault, L. Reekie, J.P. Dakin, Thermally-compensated bending gauge using surfacemounted fiber gratings, Int. J. Optoelectron. 9 (1994) 281–283. [12] M. Song, S.B. Lee, S.S. Choi, B. Lee, Simultaneous measurement of temperature and strain using two fiber Bragg gratings embedded in a glass tube, Opt. Fiber Technol. 3 (1997) 194–196. [13] M.G. Xu, J.L. Archambault, L. Reekie, J.P. Dakin, Discrimination between strain and temperature effects using dual-wavelength fiber grating sensors, Electron. Lett. 30 (13) (1994) 1085–1087. B. Lee / Optical Fiber Technology 9 (2003) 57–79 75 [14] S.W. James, M.L. Dockney, R.P. Tatam, Simultaneous independent temperature and strain measurement using in-fibre Bragg grating sensors, Electron. Lett. 32 (12) (1996) 1133–1134. [15] M. Song, B. Lee, S.B. Lee, S.S. Choi, Interferometric temperature-insensitive strain measurement with different diameter fiber Bragg gratings, Opt. Lett. 22 (11) (1997) 790–792. [16] B. Lee, Y. Jeong, Interrogation techniques for fiber grating sensors and the theory of fiber gratings, in: F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002, Chap. 7, pp. 295–381. [17] S.M. Melle, K. Liu, R.M. Measures, A passive wavelength demodulation system for guided-wave Bragg grating sensors, IEEE Photon. Technol. Lett. 4 (5) (1992) 516–518. [18] A.B. Lobo Ribeiro, L.A. Ferreira, M. Tsvetkov, J.L. Santos, All-fibre interrogation technique for fibre Bragg sensors using a biconical fibre filter, Electron. Lett. 32 (4) (1996) 382–383. [19] M. Song, S.B. Lee, S.S. Choi, B. Lee, Fiber laser strain sensor using an LPG (long period grating) Bragg wavelength demodulation filter, in: The 2nd Optoelectronics and Communications Conference (OECC ’97) Technical Digest, Seoul, Korea, 1997, pp. 676–677. [20] M.A. Davis, A.D. Kersey, All-fibre Bragg grating strain-sensor demodulation technique using a wavelength division coupler, Electron. Lett. 30 (1) (1994) 75–77. [21] Q. Zhang, D.A. Brown, H. Kung, J.E. Townsend, M. Chen, L.J. Reinhart, T.F. Morse, Use of highly overcoupled couplers to detect shifts in Bragg wavelength, Electron. Lett. 31 (6) (1995) 480–482. [22] A.D. Kersey, M.A. Davis, H.J. Patrick, M. LeBlanc, K.P. Koo, C.G. Askins, M.A. Putnam, E.J. Friebele, Fiber grating sensors, J. Lightwave Technol. 15 (8) (1997) 1442–1463. [23] C.G. Askins, M.A. Putnam, G.M. Williams, E.J. Friebele, Stepped-wavelength optical fiber Bragg grating arrays fabricated in line on a draw tower, Opt. Lett. 19 (1994) 147–149. [24] Y. Hu, S. Chen, L. Zhang, I. Bennion, Multiplexing Bragg gratings combined wavelength and spatial division techniques with digital resolution enhancement, Electron. Lett. 33 (23) (1997) 1973–1975. [25] R. Willsch, W. Ecke, H. Bartelt, Optical fiber grating sensor networks and their application in electric power facilities, aerospace and geotechnical engineering, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 49–54. [26] S.C. Kang, H. Yoon, S.B. Lee, S.S. Choi, B. Lee, Real-time measurement for static and dynamic strain using a fiber Bragg grating and the ASE profile of EDFA, in: Proceedings of the 13th International Conference on Optical Fiber Sensors (OFS-13), Kyongju, Korea, in: Proc. SPIE, Vol. 3746, 1999, pp. 530– 533. [27] S. Kim, J. Kwon, S. Kim, B. Lee, Temperature-independent strain sensor using a chirped grating partially embedded in a glass tube, IEEE Photon. Technol. Lett. 12 (6) (2000) 678–680. [28] A.D. Kersey, M.A. Davis, T. Tsai, Fiber optic Bragg grating strain sensor with direct reflectometric interrogation, in: Proceedings of the Optical Fiber Sensors Conference (OFS-11), Sapporo, Japan, 1996, pp. 634–637. [29] V. Grubsky, J. Feinberg, Long-period fiber gratings with variable coupling for real-time sensing applications, Opt. Lett. 25 (4) (2000) 203–205. [30] R.W. Fallon, L. Zhang, A. Gloag, I. Bennion, Identical broadband chirped grating interrogation technique for temperature and strain sensing, Electron. Lett. 33 (1997) 705–706. [31] A.D. Kersey, T.A. Berkoff, W.W. Morey, Multiplexed fiber Bragg grating strain-sensor system with a fiber Fabry–Perot wavelength filter, Opt. Lett. 18 (1993) 1370–1372. [32] S.T. Vohra, M.D. Todd, G.A. Johnson, C.C. Chang, B.A. Danver, Fiber Bragg grating sensor system for civil structure monitoring: applications and field tests, in: Proceedings of the 13th International Conference on Optical Fiber Sensors (OFS-13), Kyongju, Korea, in: Proc. SPIE, Vol. 3746, 1999, pp. 32–37. [33] P.J. Henderson, D.J. Webb, D.A. Jackson, L. Zhang, I. Bennion, Highly-multiplexed grating-sensors for temperature-referenced quasi-static measurements of strain in concrete bridges, in: Proceedings of the 13th International Conference on Optical Fiber Sensors (OFS-13), Kyongju, Korea, in: Proc. SPIE, Vol. 3746, 1999, pp. 320–323. [34] A.D. Kersey, T.A. Berkoff, W.W. Morey, High-resolution fibre-grating based strain sensor with interferometric wavelength-shift detection, Electron. Lett. 28 (3) (1992) 236–238. [35] R.S. Weis, A.D. Kersey, T.A. Berkoff, A four-element fiber grating sensor array with phase-sensitive detection, IEEE Photon. Technol. Lett. 6 (12) (1994) 1469–1472. [36] K.P. Koo, A.D. Kersey, Bragg grating based laser sensors systems with interferometric interrogation and wavelength division multiplexing, J. Lightwave Technol. 13 (1995) 1243–1249. 76 B. Lee / Optical Fiber Technology 9 (2003) 57–79 [37] A.D. Kersey, T.A. Berkoff, Fiber-optic Bragg grating differential-temperature sensor, IEEE Photon. Technol. Lett. 4 (10) (1992) 1183–2285. [38] A. Dandridge, A.B. Tveten, T.G. Giallorenzi, Homodyne demodulation scheme for fiber optic sensors using phase generated carrier, IEEE J. Quant. Electron. 18 (1982) 1647–1653. [39] Y.L. Lo, J.S. Sirkis, C.C. Chang, Passive signal processing of in-line fiber etalon sensors for high strain-rate loading, J. Lightwave Technol. 15 (1997) 1578–1585. [40] M. Song, S. Yin, P.B. Ruffin, Fiber Bragg grating strain sensor demodulation with quadrature sampling of a Mach–Zehnder interferometer, Appl. Opt. 39 (7) (2000) 1106–1111. [41] S.C. Kang, S.B. Lee, S.S. Choi, B. Lee, A novel demodulation technique for the wavelength shift of fiber Bragg grating sensors using the I/Q signal processing scheme, in: Proceedings of the Conference on Lasers and Electro-Optics—Pacific Rim (CLEO/Pacific Rim ’99), Seoul, Korea, 1999, pp. 135–136. [42] T.A. Berkoff, A.D. Kersey, Fiber Bragg grating array sensor system using a bandpass wavelength division multiplexer and interferometric detection, IEEE Photon. Technol. Lett. 8 (11) (1996) 1522–1524. [43] G.A. Johnson, S.T. Vohra, B.A. Danver, K. Pran, G.B. Havsgard, G. Wang, Vibration monitoring of a ship waterjet with fiber Bragg gratings, in: Proceedings of the 13th International Conference on Optical Fiber Sensors (OFS-13), Kyongju, Korea, in: Proc. SPIE, Vol. 3746, 1999, pp. 616–619. [44] M.A. Davis, A.D. Kersey, Application of a fiber Fourier transform spectrometer to the detection of wavelength-encoded signals from Bragg grating sensors, J. Lightwave Technol. 13 (7) (1995) 1289–1295. [45] K.B. Rochford, S.D. Dyer, Demultiplexing of interferometrically interrogated fiber Bragg grating sensors using Hilbert transform processing, J. Lightwave Technol. 17 (5) (1999) 831–836. [46] M. Volanthen, H. Geiger, M.G. Xu, J.P. Dakin, Simultaneous monitoring fibre gratings with a single acousto-optic tunable filter, Electron. Lett. 32 (1996) 1228–1229. [47] H. Geiger, M.G. Xu, N.C. Eaton, J.P. Dakin, Electronic tracking system for multiplexed fibre grating sensors, Electron. Lett. 32 (1995) 1006–1007. [48] J.P. Dakin, M. Volanthen, Distributed and multiplexed fibre grating sensors, including discussion of problem areas, IEICE Trans. Electron. E83-C (3) (2000) 391–399. [49] D.A. Jackson, A.B. Lobo Ribeiro, L. Reeckie, J.L. Archambault, Simple multiplexing scheme for a fiberoptic grating sensor network, Opt. Lett. 18 (1993) 1192–1194. [50] G.P. Brady, S. Hope, A.B. Lobo Ribeiro, D.J. Webb, L. Reekie, J.L. Archambault, D.A. Jackson, Demultiplexing of fibre Bragg grating temperature and strain sensors, Opt. Commun. 111 (1994) 51–54. [51] M.A. Davis, A.D. Kersey, Matched-filter interrogation technique for fibre Bragg grating arrays, Electron. Lett. 31 (1995) 822–823. [52] S.C. Kang, S.Y. Kim, S.B. Lee, S.W. Kwon, S.S. Choi, B. Lee, Temperature-independent strain sensor system using a tilted fiber Bragg grating demodulator, IEEE Photon. Technol. Lett. 10 (10) (1998) 1461– 1463. [53] Y.J. Rao, D.A. Jackson, L. Zhang, I. Bennion, Dual-cavity interferometric wavelength-shift detection for in-fiber Bragg grating sensors, Opt. Lett. 21 (19) (1996) 1556–1558. [54] J. Jung, Y.W. Lee, B. Lee, High-resolution interrogation technique for fiber Bragg grating strain sensor using long period grating pair and EDF, in: The 14th International Conference on Optical Fiber Sensors (OFS-14), Venice, Italy, 2000. [55] G.A. Ball, W.W. Morey, P.K. Cheo, Fiber laser source/analyzer for Bragg grating sensor array interrogation, J. Lightwave Technol. 12 (1994) 700–703. [56] T. Coroy, R.M. Measures, Active wavelength demodulation of a Bragg grains fibre optic strain sensor using a quantum well electroabsorption filtering detector, Electron. Lett. 32 (1996) 1811–1812. [57] S.H. Yun, D.J. Richardson, B.Y. Kim, Interrogation of fiber grating sensor arrays with a wavelength-swept fiber laser, Opt. Lett. 23 (11) (1998) 843–845. [58] M.A. Putnam, M.L. Dennis, J.U. Kang, T.-E. Tsai, I.N. Duling, I.E.J. Friebele, Sensor grating array demodulation using a passively mode-locked fiber laser, in: Technical Digest of the Optical Fiber Communication Conference, Dallas, TX, 1997, Paper WJ4, pp. 156–157. [59] A.D. Kersey, A. Dandridge, M.A. Davis, Low-crosstalk code-division multiplexed interferometric array, Electron. Lett. 28 (1992) 351–352. [60] K.P. Koo, A.B. Tveten, S.T. Vohra, Dense wavelength division multiplexing of fibre Bragg grating sensors using CDMA, Electron. Lett. 35 (1999) 165–167. B. Lee / Optical Fiber Technology 9 (2003) 57–79 77 [61] P.K.C. Chan, W. Jin, J.M. Gong, M.S. Demokan, Multiplexing of fiber Bragg grating sensors using an FMCW technique, IEEE Photon. Technol. Lett. 11 (11) (1999) 1470–1472. [62] M. LeBlanc, S.Y. Huang, M. Ohn, R.M. Measures, A. Guemes, A. Othonos, Distributed strain measurement based on a fiber Bragg grating and its reflection spectrum analysis, Opt. Lett. 21 (1996) 1405–1407. [63] Y.J. Rao, S. Huang, Applications of fiber optic sensors, in: F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002, Chap. 10, pp. 449–490. [64] K.-T. Lau, L.-M. Zhou, P.-C. Tse, L.-B. Yuan, Applications of composites, optical fiber sensors and smart composites for concrete rehabilitation: an overview, Appl. Compos. Mater. 9 (2002) 221–247. [65] L. Maurin, J. Boussoir, S. Rougeault, M. Bugaud, P. Ferdinand, A.G. Landrot, Y.-H. Grunevald, T. Chauvin, FBG-based smart composite bogies for railway applications, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 91–94. [66] P.M. Nellen, A. Frank, A. Kenel, High strain and high strain gradients measured with fiber Bragg gratings in structural engineering applications, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 111–114. [67] E. Udd, How to start a small high tech business in Troutdale, Oregon?, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 23–26. [68] D.L. Gysling, F.X. Bostick III, Changing paradimes in oil and gas reservoir monitoring—the introduction and commercialization of in-well optical sensing systems, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 43–46. [69] V. Vali, R.W. Shorthill, Fiber ring interferometer, Appl. Opt. 15 (1976) 1099–1100. [70] J. Blake, Fiber optic gyroscopes, in: K.T.V. Grattan, B.T. Meggitt (Eds.), Optical Fiber Technology, Vol. 2, Chapman and Hall, New York, 1998, pp. 303–328. [71] P.B. Ruffin, Fiber gyroscope sensors, in: F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002, pp. 383–415. [72] R.P. Moeller, W.K. Burns, N.J. Frigo, Open-loop output and scale factor stability in a fiber-optic gyroscope, J. Lightwave Technol. 7 (2) (1989) 262–269. [73] J.L. Davis, S. Ezekiel, Closed-loop, low-noise fiber-optic rotation sensor, Opt. Lett. 6 (1981) 505–507. [74] B.Y. Kim, H.J. Shaw, Gated phase-modulation feedback approach to fiber-optic gyroscopes, Opt. Lett. 9 (1984) 263–265. [75] T. Gaiffe, From R&D brassboards to navigation grade FOG-based INS: the experience of Photonetics/Ixsea, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 1–4. [76] S.J. Sanders, L.K. Strandjord, D. Mead, Fiber optic gyro technology trends—a Honeywell perspective, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 5–8. [77] G.A. Pavlath, Fiber optic gyro based inertial navigation systems at Northrop Grumman, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, p. 9. [78] R. Usui, A. Ohno, Recent progress of fiber optic gyroscopes and applications at JAE, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 11–14. [79] J. Nasu, K. Saito, A. Kurokawa, F. Hayashi, I. Nakatani, Application of fiber gyros at MPC—inertial naigation and guidance system for M–V rocket, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 15–18. [80] R.B. Dyott, S.M. Bennett, D. Allen, J. Brunner, Development and commercialization of open loop fiber gyros at KVH Industries (formerly at Andrew), in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 19–22. [81] T. Kumagai, W. Ohnuki, S. Yamamoto, A. Hongo, I. Sone, Optical fiber sensor applications at Hitachi Cable, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 35–38. [82] G.A. Sanders, J.N. Blake, A.H. Rose, F. Rahmatian, C. Herdman, Commercialization of fiber-optic current and voltage sensors at NxtPhase, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 31–34. [83] U.S. Patent #3,324,393 (1967). [84] U.S. Patent #3,605,013 (1971). [85] A.H. Rose, Z.B. Ren, G.W. Day, Twisting and annealing optical fiber for current sensors, J. Lightwave Technol. 14 (11) (1996) 2492–2498. [86] K. Kurosawa, Y. Tashiro, K. Ohkawara, Differential current relaying system with optical phase detection using flint glass fiber type optical current transformers, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 549–552. 78 B. Lee / Optical Fiber Technology 9 (2003) 57–79 [87] A.J. Rogers, Optical-fibre current measurement, Int. J. Optoelectron. 3 (1988) 391–407. [88] K. Kurosawa, S. Yoshida, K. Sakamoto, Polarization properties of the flint glass fiber, J. Lightwave Technol. 13 (7) (1995) 1378–1384. [89] S.C. Rashleigh, R. Ulrich, Magneto-optic current sensing with birefringent fibers, Appl. Phys. Lett. 34 (11) (1979) 768–770. [90] R.I. Laming, D.N. Payne, Electric current sensors employing spun highly birefringent optical fibers, J. Lightwave Technol. 7 (12) (1989) 2084–2094. [91] N.C. Pistoni, M. Martinelli, Polarization noise suppression in retracing optical fiber circuits, Opt. Lett. 16 (10) (1991) 711–713. [92] M. Takahashi, H. Noda, K. Terai, S. Ikuta, Y. Mizutani, Optical current sensor for gas insulated switchgear using silica optical fiber, IEEE Trans. Power Delivery 12 (4) (1997) 1422–1428. [93] J. Blake, P. Tantaswadi, R.T. de Carvalho, In-line Sagnac interferometer current sensor, IEEE Trans. Power Delivery 11 (1) (1996) 116–121. [94] F. Briffod, D. Alasia, L. Thevenaz, G. Cuenoud, Ph. Robert, Extreme current measurements using a fibre optics current sensor, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, Postdeadline paper PD3. [95] K. Bohnert, H. Brandle, G. Frosio, Field test of interferometric optical fiber high voltage and current sensors, in: Proc. OFS-10, 1994. [96] M. Willsch, T. Bosselmann, Optical current sensor application in the harsh environment of a 120 MVA power generator, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 407–410. [97] K. Bohnert, P. Gabus, J. Nehring, H. Brändle, Temperature and vibration insensitive fiber-optic current sensor, J. Lightwave Technol. 20 (2) (2002) 267–276. [98] E. Udd, Overview of fiber optic sensors, in: F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002, Chap. 1, pp. 1–39. [99] M.K. Barnoski, S.M. Jensen, Fiber waveguides: a novel technique for investigating attenuation characteristics, Appl. Opt. 15 (1976) 2112–2115. [100] J. Beller, OTDRs and backscatter measurements, in: D. Derickson (Ed.), Fiber Optic Test and Measurement, Prentice Hall, Upper Saddle River, NJ, 1998, Chap. 11, pp. 434–474. [101] W.V. Sorin, Optical reflectometry for component characterization, in: D. Derickson (Ed.), Fiber Optic Test and Measurement, Prentice Hall, Upper Saddle River, NJ, 1998, Chap. 10, pp. 383–433. [102] L. Thevenaz, M. Facchini, A. Fellay, P. Robert, D. Inaudi, B. Dardel, Monitoring of large structures using distributed Brillouin fiber sensing, in: Proceedings of the 13th International Conference on Optical Fiber Sensors (OFS-13), Kyongju, Korea, in: Proc. SPIE, Vol. 3746, 1999, pp. 345–348. [103] http://www.smartec.ch. [104] G.D. Peng, P.L. Chu, Optical fiber hydrophone systems, in: F.T.S. Yu, S. Yin (Eds.), Fiber Optic Sensors, Dekker, New York, 2002, Chap. 9, pp. 417–447. [105] D. Inaudi, N. Casanova, SMARTEC: bringing fiber optic sensors into concrete applications, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 27–30. [106] http://www.fiberdynamics.com/index.html. [107] http://www.fiso.com. [108] Y.W. Lee, B. Lee, High resolution cryogenic optical fiber sensor system using erbium-doped fiber, Sensors Actuators A Phys. 96 (1) (2002) 25–27. [109] G. Orellana, M.C. Moreno-Bondi, From molecular engineering of luminescent indicators to environmental analytical chemistry in the field with fiber-optic (bio)sensors, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 115–118. [110] F.J. Arregui, D. Galbarra, I.R. Matias, K.L. Cooper, R.O. Claus, ZrO2 thin films deposited by the electrostatic self-assembly method on optical fibers for ammonia detection, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 265–268. [111] D.J. Webb, M.W. Hathaway, D.A. Jackson, S. Jones, L. Zhang, I. Bennion, First in-vivo trials of a fiber Bragg grating based temperature profiling system, J. Biomed. Opt. 5 (1) (2000) 45–50. [112] S.W. James, N.D. Rees, R.P. Tatam, G.J. Ashwell, Optical fiber long period gratings with thin film overlays, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 119–122. B. Lee / Optical Fiber Technology 9 (2003) 57–79 79 [113] T. Allsop, D.J. Webb, I. Bennion, A high sensitivity long period grating Mach–Zehnder refractometer, in: The 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, 2002, pp. 123–126. [114] S. Kim, Y. Jeong, S. Kim, J. Kwon, N. Park, B. Lee, Control of the characteristics of a long-period grating by cladding etching, Appl. Opt. 39 (13) (2000) 2038–2042. [115] Optical equipment market to decline through 2004, Aug. 1, 2002 (http://wdm.pennnet.com/home.cfm). [116] http://www.giikorea.co.kr. [117] Honeywell to acquire sensor systems business from Invensys, Aug. 19, 2002 (http://lfw.pennnet.com/home. cfm). [118] Frost, Sullivan, Sensor market watch: To buy or to sell? That is the question, Sep. 25, 2002. [119] M. Fuller, Product diversification is key to success in FBG market, Lightwave, May 2002 (http://lw. pennwellnet.com/home.cfm). [120] http://www.nxtphase.com/index.asp. [121] http://www.uspto.gov/patft.