5

Fiber optic sensors for assessing and monitoring civil infrastructures

K.J. Peters,    North Carolina State University, USA

D. Inaudi,    SMARTEC S.A. Switzerland

Abstract:

In this chapter we present a technical description of fiber optic sensors including point, multiplexed, long-base, and distributed sensors, and their advantages. In particular we highlight the sensing capabilities that have no parallel in conventional sensors. To provide a background, this chapter first presents the properties of optical fibers and internal parameters that can be exploited to sense changes in their surroundings such as strain and temperature. Practical considerations in the deployment of such sensors in structural applications are also listed. Next, the sensing mechanisms of conventional fiber optic sensors and their applications to structural monitoring are summarized. Finally, new capabilities for fiber optic sensors enabled by recent advances in optical fiber technology and instrumentation are introduced.

Key words

strain sensing; temperature sensing; sensor multiplexing; distributed sensing

5.1 Introduction

From many points of view, fiber optic sensors (FOSs) are the ideal transducers for structural health monitoring. Being durable, stable, and insensitive to external perturbations, they are especially useful for long-term health assessments of civil structures, geostructures, and aerospace structures. Many different FOS technologies exist and offer a wide range of performances and suitability for different applications. In the last few years, FOSs have made a slow but significant entrance in the sensor panorama. After an initial euphoric phase, when FOSs seemed on the verge of becoming prevalent in the world of sensing, it now appears that this technology is mainly attractive in cases where it offers superior performance compared with the more proven conventional sensors. The additional value can include an improved quality of measurements, a better reliability, the possibility of replacing manual readings and operator judgment with automatic measurements, an easier installation and maintenance, or a lower lifetime cost. Finally, distributed fiber sensors offer new exciting possibilities that have no parallel in conventional sensors.

A great variety of FOSs exist for structural and geotechnical monitoring. Figure 5.1 illustrates the four main types of FOSs:

• Point sensors have a single measurement point at the end of the fiber optic connection cable, similarly to most electrical sensors.

• Multiplexed sensors allow the measurement at multiple points along a single fiber line.

• Long-base sensors integrate the measurement over a long measurement base. They are also known as long-gage sensors.

• Distributed sensors are able to sense at any point along a single fiber line, typically every meter over many kilometers of length.

image
5.1 Fiber optic sensor types. (SOFO surveillance d’ouvrages par senseurs à fibres optiques (structural monitoring by fiber optic sensors).)

The greatest advantages of the FOSs are intrinsically linked to the optical fiber itself, that is either used as a link between the sensor and the signal conditioner, or becomes the sensor itself in the case of long-gage and distributed sensors. In almost all FOS applications, the optical fiber is a thin glass fiber that is protected mechanically by a polymer coating (or a metal coating in extreme cases) and further protected by a multi-layer cable structure designed to protect the fiber from the environment where it will be installed. Since glass is an inert material very resistant to almost all chemicals, even at extreme temperatures, it is an ideal material for use in harsh environments, such as are encountered in geotechnical applications. Chemical resistance is a great advantage for long-term reliable health monitoring of civil engineering structures, making FOSs particularly durable. Since the light confined in the core of the optical fibers used for sensing purposes does not interact with any surrounding electromagnetic field, FOSs are intrinsically immune to any electromagnetic (EM) interference. With such unique advantage over sensors using electrical cables, FOSs are obviously the ideal sensing solution when the presence of EM, radio frequency, or microwaves cannot be avoided. For instance, FOSs will not be affected by any electromagnetic field generated by lightning hitting a monitored bridge or dam, nor from the interference produced by a subway train running near a monitored zone. FOSs are intrinsically safe and naturally explosion-proof, making them particularly suitable for monitoring applications of risky structures such as gas pipelines, coal mines, or chemical plants. But the greatest and most exclusive advantage of such sensors is their ability to offer long range distributed sensing capabilities or large networks of multiplexed sensors.

5.2 Properties of optical fibers

In this section we review basic characteristics of optical fibers and their relevance to sensing parameters. In addition, practical issues of packaging and connecting optical fiber sensors are discussed.

5.2.1 Optical fiber concepts

Optical fibers generally consist of a fused silica (SiO2) core and cladding. During drawing of the optical fiber, dopants are added to the silica to provide an index of refraction distribution throughout the cross-section of the optical fiber,1 with the core index normally higher than the cladding. Typical core diameters are 5–10 μm for single-mode optical fibers and greater than 50–200 μm for multi-mode optical fibers. The cladding diameter for almost all single-mode optical fibers is standardized at 125 μm to permit easy interconnection. Small diameter fibers (80 μm) have also been developed for sensing applications.2 The small diameter makes the fiber sensor less invasive when the fiber is embedded in a host material system, and increases the sensitivity of the sensor to applied loads. It can also allow smaller bending radii.

The cross-section of a step-index optical fiber is shown in Fig. 5.2. The index of refraction distribution is only a function of the radius with

nrθ=n1ranrθ=n2a<rb [5.1]

image [5.1]
image
5.2 Cross-section of step-index optical fiber. Figure not to scale.

In order for propagation of guided lightwaves to occur, n1 > n2. The index of refraction difference between the core and cladding is typically less than 1%.3

Each component of the lightwave propagating along the fiber in the z direction has the form,

ψrθzt=ψ¯rθetωtβz [5.2]

image [5.2]

where ψ¯rθimage is the energy distribution in the plane perpendicular to the propagation direction, ω is the angular frequency (ω = 2πc/λ where c is the speed of light in a vacuum and λ is the free-space wavelength of the light-wave), t is time, and β is the propagation constant of the mode (phase shift per unit length). Each propagating mode has an effective index of refraction, neff = βλ/(). neff corresponds to the index of refraction of an equivalent homogeneous material for which the wave would propagate with the same propagation constant β as through the step-index fiber. For guided modes, n2 < nett < n1. The effective index of refraction neff, or mode propagation constant β, will play a critical role in the response of many optical fiber sensors. Although the step-index fiber is the easiest for which to analyze mode propagation, most commercially available optical fibers have different index distributions as shown in Fig. 5.3.1 While most optical fiber manufacturers do not provide detailed information on the index distributions, they do provide data such as the core radius, design operating wavelength, and neff at common wavelengths (see for example Reference 4). From this information, the user can calculate the response of the optical fiber sensor using the same principles as for the step-index fiber.

image
5.3 Common fiber index of refraction profiles.

The number of guided modes that can propagate through the optical fiber at a particular wavelength depends upon the normalized frequency at that wavelength, V,

V=2πλan12n22 [5.3]

image [5.3]

The first mode, also called the fundamental mode, propagates at all values of V. All other modes have a cutoff value of V below which it cannot propagate. The cutoff value for the second mode is at V = 2.4048. Therefore, for V < 2.4048 the fiber is referred to as a single-mode fiber, and for V > 2.4048 the fiber is referred to as a multi-mode fiber. For several sensing technologies it is important to utilize fibers that are single-mode at the wavelength to be interrogated so that the mode of propagation and coupling are welldefined. Additionally, the fundamental mode is the least affected by bending losses, as the energy is concentrated in the center of the fiber. Multi-mode sensors do exist and are generally easier to practically couple to instrumentation as the fiber core is larger and the modes are more spread out throughout the fiber cross-section.510

Finally, additional modes, called cladding modes, are possible due to the interface condition between the outer diameter of the fiber and the surrounding medium, typically polymer coating or air. The form of these modes depends upon the index of refraction of the surrounding medium, n3, and is present in the case of n3 < n2. The cladding modes can be excited locally in long period grating (LPG) sensors and exploited for the measurement of cure monitoring, chemical, or environmental sensing.11

A second important class of optical fibers applied for sensing is highbirefringence fibers, also referred to as polarization maintaining (PM) fibers.12 PM fibers present an index profile that is not rotationally symmetric and therefore propagate two separate, linearly polarized fundamental modes (LP01) at two separate propagation constants, β1 and β2. The modes are polarized about orthogonal axes, referred to as the fast and slow axes. As the polarization axes are orthogonal, the two modes do not interfere with one another as they propagate along the optical fiber. These orthogonal modes can be used to independently measure multiple parameters. In other cases, only one of the two polarizations propagates, while the other is suppressed.

Standard fused silica fibers are linear elastic until failure for the following material properties: the elastic modulus E = 72 GPa, Poisson’s ratio v = 0.20, and thermal expansion coefficient a = 5.5 × 10− 7/°C.13 Typical elongation at failure is 2–5%. A second common material class used for optical fibers is polymers. Polymer optical fibers (POF) have considerably lower stiffness than silica fibers, with E = 2.4–3.0 GPa and v = 0.34.14 Their stress-strain response is also extremely sensitive to strain rate, temperature, humidity, and hysteresis effects. POF fibers can have elongations at failure of up to 100% or more.

The attenuation spectrum of a silica optical fiber can be seen in Fig. 5.4. Three primary transmission windows are therefore used in telecommunication applications: around (1) 850 nm, (2) 1300 nm, and (3) 1550 nm. The same three transmission windows are typically used for optical fiber sensors due to the relative low-cost, wide selection, and high quality of components including optical fibers, couplers, and laser sources available for telecommunications applications. POFs demonstrate considerably more attenuation at near-infrared wavelengths than fused silica fibers;14 therefore POF sensors most often operate in the visible wavelength ranges.

image
5.4 Material attenuation losses as a function of wavelength for fused silica (SiO2).1

5.2.2 Sensing mechanisms

External sensing parameters are generally encoded into changes within lightwaves propagating through an optical fiber in one of four manners, summarized in Fig. 5.5. The intensity of the lightwave propagating through an optical fiber can be modified through microbending of the optical fiber, a change in coupling from the fundamental mode to other non-guided modes, fracture of the optical fiber, or a change in power coupled into the fiber or from one fiber to another. Measuring the intensity of the propagating lightwave is relatively simple; however, light sources themselves often fluctuate in intensity. Feedback control loops can be applied to reduce these fluctuations, or the illuminating lightwave divided to create a reference intensity, as shown in Fig. 5.6a. Under these conditions, intensitybased sensors can provide absolute measurements, meaning that they are insensitive to power interruptions between data collections. One drawback to intensity-based sensors, however, is that they cannot be multiplexed for sensor networks.

image
5.5 Output signals of optical fiber sensors using four primary sensing mechanisms.
image
5.6 Sensor interrogation schemes for (a) intensity-based sensors and (b) wavelength (frequency)-based sensors. For (b), three different possible schemes are shown. See text for details.

The application of strain or temperature to an optical fiber changes the optical path length traversed by the lightwave as it propagates a distance L through the fiber, neff L. As the phase of the lightwave cannot be measured directly, the lightwave is generally re-combined with a reference lightwave from the same laser source (so that the two are coherent). When the reference lightwave is not exposed to the parameter to be measured, the relative phase shift between the signais can be related to the applied strain or temperature. The phase modulation is extremely sensitive to strain and therefore can provide very accurate measurements;15 however, the measurement of phase shift is not absolute due to the signal periodicity and is therefore affected by power interruptions between data collections. Techniques to alleviate these problems, such as low-coherence interferometry, are discussed later in this chapter. The reference signal can be completely sheltered from external parameters, or used for compensation of unwanted measurands such as temperature, if both fibers are exposed to them.

External sensing parameters can also be converted into spectral information of the lightwave, for example using fiber Bragg gratings (FBG) or Fabry–Pérot interferometers. Such sensors typically act as filters, transmitting certain wavelengths and radiating or reflecting others. Changes in external parameters are therefore converted into a wavelength shift of the transmitted or reflected spectrum of the sensor. A typical signal is shown in the example of Fig. 5.5, for which the reflected wavelength is changing with time. The wavelength encoded signal must then be interpreted, which is commonly performed in using one of three methods: (1) by launching a broadband of light into the fiber and applying a spectrum analyzer to select the reflected wavelength; (2) by applying a tunable laser for which the output wavelength can be scanned while the reflected intensity is measured by a photodetector; or (3) by launching a broadband of light and using a wavelength dependent filter to identify the reflected wavelength (see Fig. 5.6b).

A final method to encode sensing information for transmission through an optical fiber is through the polarization state of the propagating lightwave.1517 For the example shown in Fig. 5.5, the power of a lightwave propagating through a PM fiber is transferred between the mode polarized about the fast axis and the mode polarized about the slow axis. The power in each mode can be determined by applying a polarizing filter or splitter to the output signal. As mentioned above, the two modes propagating through a PM fiber do not normally transfer power as they propagate, since they are orthogonally polarized. However, external stimuli such as pressure or twisting of the optical fiber will induce transfer between the two modes.

5.2.3 Sensor packaging

Traditional fiber optic cable design aims at the best possible protection of the fiber itself from any external influence. In particular, it is necessary to shield the optical fiber from external humidity, side pressures, crushing, and longitudinal strain applied to the cable. These designs have proven very effective in guaranteeing the longevity of optical fibers used for communication and can be used as sensing elements for monitoring temperatures in the − 20 °C to + 60 °C range e.g. in conjunction with Brillouin or Raman distributed sensing systems.

However, many of the measurement techniques described in Section 5.2.2 require that the sensing part of the optical fiber is exposed to the parameter to be measured. This creates a challenge for the sensor designer, to protect the sensor from damage resulting from manipulation and exposure to the environment while at the same time allowing a faithful transfer of the measurand from the object under monitoring to the optical fiber.18,19

Special cable designs using non-standard coatings and buffers are required for sensing temperatures below − 20 °C or above + 60 °C, especially for Brillouin scattering systems, where it is important to guarantee that the optical fiber does not experience any strain that could be misinterpreted as a temperature change due to the cross-sensitivity between strain and temperature.

For strain and deformation sensors, the packaging must faithfully transfer the structural strain to the optical fiber, a goal contradicting all experience from telecommunication cable design where the exact opposite is required. This can be achieved by a continuous mechanical coupling between the sensor and the structure e.g., by continuously gluing the fiber to a steel beam, or by a local coupling at the ends of the sensing zone.20

All sensor packaging designs share common reliability goals independently from the sensing technique and application domain used:

• The optical fibers must be compatible with the selected sensing system: single-mode or multi-mode.

• The fibers must be protected from external mechanical actions during installation and while in use. In particular, the cable design must allow easy manipulation without the risk of fiber damage.

• The cable design must allow sufficient shielding of the optical fibers from chemical aggression by humidity, water, and other harmful substances.

• All optical losses must be kept as low as possible in order not to introduce degradation to the instrument’s performances.

• Installation of connectors and repair of damaged sensors should be compatible with field operations.

5.2.4 Cables, connectors, and splicing

The above considerations on sensor packaging also apply to the interconnection cables, connectors, and splices. In most cases it is however possible to borrow tools and techniques widely used in the telecommunication industry. Cables with different levels of protection are to be used depending on their location: indoor, outdoor, overhead, rodent-resistant, compatible with extreme temperatures, presence of radiation, or chemically aggressive environments.

Standard telecom connectors can be used in junction boxes, but more rugged versions for direct exposure to the elements also exist. Examples include outdoor and waterproof connectors, up to extreme cases such as deep-water remotely operated underwater vehicle (ROV)-matable connectors. Splicing is considered a more reliable way to permanently interconnect optical fibers, but requires tools and expertise that are not always compatible with field conditions. Therefore, an accurate analysis of the project will lead to the appropriate choice of a connector, a splice, or a very long cable.

5.3 Common optical fiber sensors

We now present examples of the most common optical fiber sensors applied to civil infrastructure systems. The sensors are divided into categories based on the mechanism used to convert the sensing parameters into a measurement.

5.3.1 Coherent interferometers

Butter and Hocker21 demonstrated the first in-fiber strain sensor. This ‘fiber optics strain gage’ was based on the measurement of the change in phase shift of a lightwave propagating through an optical fiber. The phase shift was due to applied axial strain through both the elongation of the optical fiber and the material property change of the silica with strain. The phase shift measurement replicated a conventional bulk optic coherent interferometer. Coherent, in-fiber interferometric sensors such as this present an extremely high sensitivity to external parameters. Additionally, as the optical fiber itself is used as the measurement device, such sensors are beneficial for simplicity and cost when compared to other optical fiber-based sensors.

Applying an axial tension to an optical fiber of length L, with a resulting axial strain ε, and a uniform temperature change, ΔT, along the entire optical fiber, we find the change in phase shift, Δφ, of a lightwave propagating through the fiber by:

Δφ=2πλneffLε12εneff2p12vp11+p12+1neffdneffdTΔT+neff22p11+2p12αΔT [5.4]

image [5.4]

where λ is the wavelength of the lightwave.22 The phase shift is therefore linearly proportional to the applied strain and the applied temperature change. Typical material property values for fused silica are p11 = 0.17, p12 = 0.36, a = 0.5 × 10− 6 °C− 1, and dneff/dT= 1.2 × 10− 5 °C− 1.22

The above equations consider a constant strain field applied along the gage length of the optical fiber. In general, the fiber can be mounted in any orientation on a surface, therefore these strain components can also vary along the length of the gage. We find the total phase shift by integrating the local phase shift along the optical fiber

φ=0Ldsds=2πλ0Lneff1pe1+εnds [5.5]

image [5.5]

where εn is the strain component tangent to the fiber path at each location and s is the variable along the path length.23 The transfer of shear stress and transverse stresses are negligible for surface mounted sensors, however, these components could be significant for sensors embedded in materials.22,24

When using an optical fiber as strain sensor, it is not possible to measure the phase shift directly; therefore, we typically measure the interference between the sensor fiber and a second reference fiber that is not exposed to the environmental changes and therefore has a constant phase. Additionally, this second fiber can be exposed to only some of the loading, for example temperature, so as to provide compensation during the measurements. The interferometric measurement of phase shifts for in-fiber sensors parallels that for classical free-space interferometers. The two most commonly applied interferometric arrangements are the Mach-Zehnder and Michelson interferometers.25 Figures 5.7 and 5.8 show each of these classical interferometers and their equivalent for in-fiber sensors. These are typical examples of long-base sensors.

image
5.7 Schematic of Mach-Zehnder interferometer: (a) free-space optics version measures the difference in optical path length through the optical cell and reference cell and (b) in-fiber version measures the difference in optical path length through sensor fiber and reference fiber.
image
5.8 Schematic of Michelson interferometer: (a) free-space optics version and (b) in-fiber version.

Assuming that the light source has a high coherence length, and polarization effects are compensated between the two fibers, the average intensity of the interference pattern varies sinusoidally, as shown in Fig. 5.9. The cyclic form of the intensity measurement presents two challenges when the intensity is near one of the quadrature points i.e., where dI/dφ = 0. The first challenge is that the direction of fringe movement cannot be determined at the quadrature point (directional ambiguity), and the second is that the measurement sensitivity goes to zero at the same point (signal fading). A variety of signal processing techniques have been applied to remove these difficulties, generally categorized into passive and active demodulation.2628 One example of a passive homodyne demodulation is shown in Fig. 5.9. For this example, the sensor lightwave is divided into two channels, with one phase shifted by 90°, and the other interfered with the reference lightwave. When one signal is near a quadrature point, the other is at the point of maximum sensitivity, removing both the signal fading and directional ambiguity issues.

image
5.9 Intensity signals as a function of interferometer phase difference for two sensor signals, phase shifted by 0° and 90°. When one signal is as a quadrature point (dl/ = 0), the other is at the location of maximum sensitivity. (Source: Adapted from Reference 25.)

Furthermore, the measurement of phase shift is not an absolute measurement. Specifically, in order to know the current strain level, the system must be continually operated throughout the lifetime of the structure with a sufficient data acquisition frequency to catch every shift of more than half wavelength. This presents significant challenges for continuous monitoring of structures, since power interruptions lead to a loss of the zero point calibration. One solution to this problem is through low-coherence interferometry, which will be described in the following section. Another challenge is multiplexing interferometric based sensors. Kersey29 reviews multiplexing strategies for interferometric optical fiber sensors, including time-division multiplexing (TDM) and frequency-division multiplexing examples.

5.3.2 Low-coherence interferometers

The light source used in an interferometer sensor is never pure and always contains a spectrum of wavelength distributed around the main wavelength. This causes the sinusoidal interference pattern to fade in intensity with an increasing path difference between the two arms. This fading is faster if the spectral distribution of the source is larger. The path difference for which the interference intensity falls to 1/e is called the coherence length of the source. This negative effect for traditional interferometers can be used advantageously to obtain absolute measurements, while maintaining the high resolution typical of interferometric sensors. If a light source with very low coherence length (i.e. with a large spectral width) is used, the interference pattern will fade very quickly with increasing path imbalance. It is, however, possible to restore the interference if the light is passed through a second interferometer that matches the path imbalance in the sensor. If the path imbalance of the second interferometer can be controlled precisely, it becomes possible to measure the unknown path imbalance in the sensor with the same accuracy, by finding the matching condition that maximizes the fringe intensity. Examples of low-coherence sensors include the surveillance d’ouvrages par senseurs à fibres optiques (SOFO) sensors and Fizeau readout units for Fabry–Pérot Sensors (see next paragraph).

The SOFO interferometric sensors are long-base sensors, integrating the measurement over a long measurement base that can reach 10 m or more. The SOFO system30,31 is a fiber optic displacement sensor with a resolution in the micrometer range and excellent long-term stability. The measurement set-up uses low-coherence interferometry to measure the length difference between two optical fibers installed on the structure to be monitored (Plate I in the color section between pages 294 and 295), by embedding in concrete or surface mounting (see Fig. 5.10a). The measurement fiber is pre-tensioned and mechanically coupled to the structure at two anchorage points in order to follow its deformations, while the reference fiber is free and acts as temperature reference. Both fibers are installed inside the same plastic pipe, and the gage length can be chosen between 200 mm and 10 m. The SOFO readout unit, shown in Fig. 5.10b, measures the length difference between the measurement fiber and the reference fiber, by compensating it with a matching length difference in its internal interferometer. The precision of the system is of ± 2 μm independently from the measurement basis and its accuracy of 0.2% of the measured deformation even over years of operation.

image
5.10 Example of SOFO sensor installed on a (a) rebar and (b) SOFO readout unit.

The SOFO system has been used in particular to monitor civil and geotechnical structures, including bridges,32 tunnels, piles,33 anchored walls, dams, historical monuments,34 and nuclear power plants, as well as laboratory models. The use of long-base SOFO sensors allows the gapless monitoring of the whole length of the structure, and provides average data that are not affected by local features or defects of the construction materials.

5.3.3 Fabry–Pérot interferometers

Fabry–Pérot interferometric sensors35 are a typical example of point sensors and have a single measurement point at the end of the fiber optic connection cable. An extrinsic Fabry–Pérot interferometer (EFPI) consist of a capillary glass tube containing two partially mirrored optical fibers facing each other, but leaving an air cavity of a few microns between them, as shown in Fig. 5.11a. When light is coupled into one of the fibers, a back-reflected interference signal is obtained. This is due to the reflection of the incoming light on the two mirrors. This interference can be demodulated using coherent or low-coherence techniques to reconstruct changes in the fiber spacing. Since the two fibers are attached to the capillary tube near its two extremities (with a typical spacing of 10 mm), the gap change will correspond to the average strain variation between the two attachment points shown in Fig. 5.11a. Based on the same principle, it is possible to design pressure sensors (Fig. 5.11b) and displacement sensors (Fig. 5.11c).

image
5.11 Schematic overview of Fabry–Pérot sensors for (a) strain, (b) pressure, and (c) displacement.

Many sensors based on this principle are currently available as a one-to-one replacement of traditional sensors used in civil and geotechnical monitoring,36 including piezometers, weldable and embedded strain gages, temperature sensors, pressure sensors, and displacement sensors. Examples are shown in Fig. 5.12.

image
5.12 (a) Fabry–Pérot piezometers, (b) embeddable strain sensors, (c) spot-weldable strain sensor, and (d) displacement sensor.

Because they are immune to electromagnetic interference, such as that found in railway applications, or the static electricity and frequent thunderstorms that are found in dams at high altitudes, fiber optic instruments offer an important advantage over the traditional vibrating wire technology for those applications. They are more rugged in such a harsh environment and allow very long cable lengths without the need for any lightning protection.

5.3.4 Fiber Bragg gratings

Fiber Bragg grating (FBG) sensors have many advantages for strain sensing in structural health monitoring applications, including the ability to measure localized strain and temperature and the potential to multiplex hundreds of sensors with a single ingress/egress fiber. They are the most commonly used type of multiplexed sensor. FBG sensors have been applied for a variety of structural health monitoring applications including spacecraft,37,38 bonded aircraft repairs,3941 cryogenic composite tanks,42,43 highway and railway bridges,4447 offshore platforms,48 and nuclear reactors.49

The FBG, shown in Fig. 5.13, is a permanent, periodical perturbation in the index of refraction of the optical fiber core. Hill et al.50 first fabricated permanent Bragg gratings in an optical fiber as a wavelength selective filter for telecommunication applications. When a broad spectrum of wavelengths is passed through the FBG, a narrow bandwidth of wavelengths is reflected, while all others are transmitted (see Fig. 5.13). The wavelength at maximum reflectivity is referred to as the Bragg wavelength, λB and is determined by the condition

λB=2neffΛ [5.6]

image [5.6]
image
5.13 Optical FBG sensor. Reflected and transmitted spectra are shown for a broadband input spectrum. Dashed line is for unstrained FBG, solid line is for strained FBG.

where Λ is the period of the index of refraction variation.

The reflectivity coefficient is defined as the ratio of the intensity of the input lightwave to that of the reflected lightwave, as a function of wavelength, and is given by:

r=sinh2Lκ2σ^2cosh2Lκ2σ^2σ^2/κ2 [5.7]

image [5.7]

where κ and σ^image a are the coupling coefficients defined as:

κ=πλvδneff¯σ=2πλneff+δneff¯πΛ [5.8]

image [5.8]

where L is the grating length, δneff¯image is the amplitude of the index of refraction modulation, and v is the modulation fringe visibility. For sensing applications, FBGs with a narrow bandwidth and high reflectivity produce the largest signal-to-noise ratio. Procedures such as apodization are often applied to the FBG sensor during fabrication to reduce the secondary peaks and narrow the bandwidth.51

As axial strain, ε, is applied to the FBG, the Bragg wavelength shifts to lower wavelengths (compression) or higher wavelengths (tension). The applied strain is thus encoded in the FBG Bragg wavelength shift. For the specific case of pure axial loading we find:

ΔλBλB=1peε [5.9]

image [5.9]

where pe is the effective photo-elastic constant for axial strain. A typical value of pe for silica optical fibers is 0.22–0.25.352 Thus, the shift in Bragg wavelength is linearly related to the applied axial strain.

Similarly, FBG sensors are also sensitive to temperature through

ΔλBλBα+ζΔT [5.10]

image [5.10]

where α is the thermal expansion coefficient and ζ is the thermo-optic coefficient of the optical fiber material. A typical value for fused silica is (α + ζ)= 6.67 × 10− 6 °C− 1.52 It is important to note that the thermal sensitivity of an FBG sensor is considerably higher than that of its electrical strain gage counterpart, increasing the need for thermal compensation for strain measurement applications.

One of the strongest advantages of FBG sensors over other available strain or temperature sensors for structural health monitoring applications is the fact that they can be multiplexed into a large sensor network. For structural health monitoring applications, having only a single or limited ingress/egress points and lead cables is a notable advantage because it can significantly reduce the weight of the lead cables and disruptions to the structure itself.37 Several approaches to multiplexing large FBG sensor networks have been applied in which the sensors are interrogated using wavelength-division and time-division multiplexing (WDM, TDM) and combinations thereof. For WDM applications, reducing the wavelength spacing between each FBG reduces the time required to scan the network, however also increases the cross-talk losses between FBGs. For TDM applications low reflectivity gratings (rmax ≅ 1%) should be used to allow for a large number of multiplexed sensors. By combining WDM with TDM, as shown in Fig. 5.14, one can reduce the power loss per FBG, as well as reducing the total wavelength range that must be scanned to interrogate the entire network.52 Several commercial interrogators are now available for networks of FBG sensors.5355 Interferometric methods have also been applied for high speed interrogation of FBG sensor networks.5658

image
5.14 Combined WDM and TDM for FBG sensor network. (FO refers to fiber optic.)

FBG strain and temperature sensors can also distinguish between multiple strain components and/or temperature changes through the birefringence that occurs in the optical fiber due to applied transverse loads. Figure 5.15 shows an example of three independent loads applied to the optical fiber, along with the resulting three principal strain components at the center of the fiber core. The effect of the axial load (P1) is to shift the reflected peak to higher or lower wavelengths; however, the effect of the transverse loads (P2 and P3) is to create peak-splitting due to the fast and slow axes that develop in the optical fiber. A typical example of peak-splitting behavior is also shown in Fig. 5.15. In general, FBGs written in PM optical fibers exhibit enhanced discrimination between multiple loading components.59 The response of the FBG sensor to the multiple load components can either be predicted numerically5962 or experimentally calibrated.63

image
5.15 Multi-axis loading applied to FBG sensor written into bowtie PM fiber. The graph plots the effects of the multi-axis loading on a single FBG (peak on the left-hand side is the initial peak before load is applied).

Distortion of the grating spectrum due to strain gradients has also been observed in several applications of embedded FBG sensors.6473 An example of experimentally measured spectral distortion due to the highly nonuniform strain field near a notch tip is shown in Fig. 5.16.54 This sensitivity of the sensor response to the form of the strain profile is unique to the optical FBG, since other strain gages (e.g., the classical electrical strain gage) average the applied strain over the gage length. This full-spectral information has been used to measure crack bridging distributions and damage states in laminates. Recently, Vella et al.74 demonstrated a full-spectral interrogator that operates up to 100 kHz for dynamic measurements.

image
5.16 Response of FBG embedded in compact tension specimen. Location of 10 mm grating is shown in sketch of specimen. Experimental grating spectral data in transmission (solid lines) are shown for three load levels (P). The simulated spectra calculated using the transfer matrix method are also plotted (dashed lines).

Finally, while FBGs reflect or transmit lightwaves at certain wavelengths based on coupling between counter-propagating core modes, long period grating (LPG) sensors are based on coupling between the forward propagating core mode and forward propagating cladding modes. The large period of the LPG (100 μm < Λ < 1 mm) results in cladding mode coupling at wavelengths in the near-IR range, appearing as multiple loss peaks in the transmission spectrum (see Fig. 5.17).75 LPG sensors have primarily been applied to the measurement of environmental or chemical parameters due to their strong sensitivity to the index of refraction of the surrounding material system.

image
5.17 Theoretical transmission spectrum for long period FBG. Portion of spectrum in near-IR region is indicated.

5.3.5 Brillouin and Raman scattering distributed sensors

Distributed fiber optic sensing offers the ability to measure temperatures and strains at thousands of points along a single fiber. This is particularly interesting for the monitoring of large structures such as dams, dikes, levees, tunnels, pipelines, and landslides, where it allows the detection and localization of movements and seepage zones with sensitivity and localization accuracy unattainable using conventional measurement techniques.

Unlike electrical sensors and localized FOSs, distributed sensors offer the unique characteristic of being able to measure physical parameters, in particular strain and temperature, along their whole length, allowing the measurements of thousands of points from a single readout unit. The most developed technologies of distributed FOSs are based on Raman76 and Brillouin scattering.77,78 Both systems make use of a nonlinear interaction between the light and the glass material of which the fiber is made. If an intense light at a known wavelength is shone into a fiber, a very small amount of it is scattered back from every location along the fiber itself. Besides the original wavelength (called the Rayleigh component), the scattered light contains components at wavelengths that are higher and lower than the original signal (called the Raman and Brillouin components). These shifted components contain information on the local properties of the fiber, in particular its strain and temperature. Figure 5.18 shows the main scattered wavelengths components for a standard optical fiber. If λ0 is the wavelength of the original laser signal generated by the readout unit, the scattered components appear both at higher and lower wavelengths.

image
5.18 Light scattering in optical fibers and its use for strain and temperature sensing. At every section of fiber, the incoming wavelength λ0 is scattered backward. The backscattered light contains new wavelengths that carry information about the strain and temperature conditions at the location where the scattering occurred.

The two Raman peaks are located symmetrically to the original wavelength. Their position is fixed, but the intensity of the peak at lower wavelength is temperature dependent, while the intensity of the one at higher wavelength is unaffected by temperature changes. Measuring the intensity ratio between the two Raman peaks therefore yields the local temperature in the fiber section where the scattering occurred.

The two Brillouin peaks are also located symmetrically at the same distance from the original wavelength. Their position relative to λ0 is however proportional to the local temperature and strain changes in the fiber section. Brillouin scattering is the result of the interaction between optical and ultrasound waves in optical fibers. The Brillouin wavelength shift is proportional to the acoustic velocity in the fiber that is related to its density. Since the density depends linearly on the strain and the temperature of the optical fiber, we can use the Brillouin shift to measure those parameters.

When light pulses are used to interrogate the fiber it becomes possible, using a technique similar to RADAR, to discriminate different points along the sensing fiber through the different time-of-flight of the scattered light. Combining the radar technique and the spectral analysis of the returned light, one can obtain the complete profile of strain or temperature along the fiber. Typically, it is possible to use a fiber with a length of up to 30 km and obtain strain and temperature readings every meter. In this case we would talk of a distributed sensing system with a range of 30 km and a spatial resolution of 1 m. Figure 5.19 schematically shows an example of distributed strain and temperature sensing.

image
5.19 Schematic example of a distributed strain and temperature measurement.

Systems based on Raman scattering typically exhibit temperature accuracy of the order of ± 0.1 °C and a spatial resolution of 1–3 m over a measurement range up to 30 km. The best Brillouin scattering systems offer a temperature accuracy of ± 0.1 °C, a strain accuracy of ± 20 microstrain, and a measurement range of 50 km, with a spatial resolution of 1–3 m.79 The readout units are portable and can be used for field applications.

Since the Brillouin frequency shift depends on both the local strain and temperature of the fiber, the sensor set-up will determine the actual response of the sensor. For measuring temperatures it is necessary to use a cable designed to shield the optical fibers from an elongation of the cable. The fiber will therefore remain in its unstrained state and the frequency shifts can be unambiguously assigned to temperature variations. Measuring distributed strains also requires a specially designed sensor. A mechanical coupling between the sensor and the host structure along the whole length of the fiber has to be guaranteed. To resolve the cross-sensitivity to temperature variations, it is also necessary to install a reference fiber along the strain sensor. Special cables, containing both free and coupled fibers, allow a simultaneous reading of strain and temperature. Figure 5.20 shows examples of temperature, strain, and combined cables.20

image
5.20 Distributed sensor cables examples: (a) temperature sensor, (b) strain sensor, and (c) combined strain and temperature sensor.

Sensing systems based on Brillouin and Raman scattering are used to detect and localize seepage in dams and dikes, allowing the monitoring of hundreds of kilometers along a structure with a single instrument and the localization of the water path with an accuracy of 1 or 2 m. Distributed strain sensors are also used to detect landslide movements and to detect the onset of cracks in concrete dams and steel bridges.80

Early applications of this technology have demonstrated that the design and production of sensing cables, incorporating and protecting the optical fibers used for the measurement, as well as their optimal locating and installation in the structure under scrutiny, are critical elements for the success of any distributed sensing instrumentation project.

5.4 Future trends

In this section, we describe recent advances in optical fiber and measurement technologies, and new opportunities in optical fiber sensing enabled by these advances.

5.4.1 Multicore fiber sensors

The recent fabrication of multicore optical fibers has led to new sensing capabilities for strain and temperature measurements. An example of a multicore optical fiber is shown in Fig. 5.21. The multiple cores are introduced with spacing sufficient such that there is minimal overlap between modes propagating through each core. These independently propagating modes permit the presence of multiple sensors in the same fiber cross-section. Coupling to the different cores is typically achieved by graded index lenses or fan-outs.81 The most common application of these multicore fiber sensors has been the measurement of curvature of different axes in the fiber cross-section by interfering the different propagating lightwaves or writing FBGs in the different cores.8284 As the cores are located at different locations relative to the neutral bending axes (see Fig. 5.21), the local curvature can be measured at high accuracies with inherent axial strain and temperature compensation. For example, Blanchard et al.83 achieved a bend angle resolution of 100 μrad using a three-core optical fiber. By separating the multiple cores by polarization axes, the axial strain and temperature can also be measured independently.85,86

image
5.21 Cross-section of multicore fiber illuminated with white light and principle of curvature measurement from FBG pairs in different cores67.

Multiplexing FBG sensors along the fiber (in each of the different cores) have also been applied for high resolution shape sensing by integrating the local curvatures (about multiple axes) along the optical fiber length.82,87,88 Flockhart et al.89 and Cranch et al.90 wrote sets of FBG pairs in each core to act as a low-finesse Fabry–Pérot interferometer to enhance the static and dynamic resolution of the distributed shape measurements.

5.4.2 Microstructured optical fiber sensors

The recent invention of microstructured optical fibers has created great potential for both optical fiber communication systems and optical fiber sensors. Such fibers have a complex cross-sectional geometry including both silica and air holes. Confinement of the propagating lightwaves is generated either by an effective index contrast similar to that of conventional optical fibers in which the air holes significantly reduce the cladding index of refraction (‘holey’ fibers) or through a photonic bandgap effect (photonic crystal fibers).91 One example of a microstructure fiber is shown in Fig. 5.22. In either case, the complex geometry of these microstructured fibers provides both new sensing possibilities and enhanced response characteristics for sensors applied to these fibers.

image
5.22 Scanning electron microscope image of the end of a photonic crystal fiber.78

For sensing applications the main potential benefits of these fibers are the following: (1) the reduced amount of silica through which the fundamental mode propagates means that the temperature sensitivity of the sensors can be reduced to near zero;92 (2) the fibers can act in single-mode over a wide range of wavelengths, permitting multiplexing of sensors at significantly different wavelengths;93 (3) for chemical sensing applications, the gas to be sensed can be inserted into the air holes, permitting a long interaction surface between the propagating mode and the gas;94 and (4) PM fibers can be fabricated through the arrangement of the air holes, without the need for induced thermal residual stresses, making the birefringence more stable to temperature variations.95

Nasilowski et al.95 and Bock et al.96 applied microstructured holey fibers for the measurement of temperature and pressure. Several researchers have also written long period FBGs into microstructured fibers through CO2 laser or electric arc etching.9799 These gratings demonstrate an excellent strain sensitivity, with little or no temperature sensitivity.97,99 Finally, researchers have also succeeded in writing FBGs into photonic crystal optical fibers for use as discrete strain sensors.100

5.4.3 Polymer optical fiber sensors

While the previous discussion has been limited to silica optical fibers, many of the same sensing techniques have been applied to polymer optical fibers (POFs) as well. POFs have additional advantages for sensing, such as their high elastic strain limit, high fracture toughness, high flexibility in bending, higher sensitivity to strain than silica, and negative thermo-optic coefficient.101104 Additionally, many polymeric materials have a high biocompatibility. More details on the properties of POFs can be found in References 14 and 105.

On the other hand, a significant challenge to applying POFs as sensors is the difficulties in fabrication of these fibers. Due to the fabrication difficulties, most POF sensors are based on multi-mode POFs.14 Multimode POFs are less expensive and easier to couple than silica singlemode optical fibers; however, they are also larger in diameter. POFs also demonstrate viscoelastic behavior and are extremely sensitive to humidity.102,104,106,107 Typical stress-strain curves for a POF at different applied strain rates are shown in Fig. 5.23. Additionally, many POF sensors operate at lower wavelengths than comparable silica fiber sensors, due to the high attenuation of polymers in the near-infrared range. Multi-mode POF sensors have been demonstrated based on many of the same measurement principles as silica optical fiber sensors, including intensity losses, back scattering, and time-of-flight.108

image
5.23 Measured true stress–strain curves for a single-mode PMMA doped core POF at different applied strain rates.102

Recently, both single-mode solid POFs and single-mode microstructured POFs have been developed that present new capabilities for POF sensors.109,110 The emergence of these optical fibers has created the potential for high precision, large deformation optical fiber sensors for a variety of applications. Kiesel et al.111 recently demonstrated coherent interferometry in a single-mode POF in-fiber Mach-Zehnder interferometer up to 15.8% elongation of the POF (see Plate II in the color section between pages 294 and 295). After calibration of both the mechanical and optomechanical parameters of the single-mode POF,102,103 the authors applied the sensor for strain measurements on tensile coupons.112 Additionally strategies to couple the single-mode POF sensor to single-mode silica optical fibers to remove the attenuation limits were developed.113 Peng et al.114 first demonstrated the writing of FBGs into single-mode POFs for localized strain sensing over large strain ranges. The POF FBG sensors demonstrate a 22% increase in strain sensitivity as compared to silica Bragg gratings115 and have been tuned over a 18 nm wavelength shift by applying a temperature change of 50 °C.116 Finally, van Eijkelenborg et al.117 fabricated microstructured POFs to overcome the high intrinsic attenuation properties of common POFs. These microstructured POFs have since been applied as sensors through inscribed LPGs and FBGs.118120

5.4.4 Rayleigh scattering distributed sensors

A more recent development in distributed fiber optic sensing is the use of Rayleigh backscattering for the interrogation of strain and temperature along an optical fiber. As for the previous distributed sensors, a standard optical fiber can be applied as the sensor, providing distributed measurements over large distances, without expensive individual sensors. Gifford et al.121,122 applied swept-wavelength interferometry (SWI) to measure the backscattered signal in silica and POFs. This interrogation method is fundamentally different from optical time-domain reflectometry measurements, since SWI measures phase shifts rather than amplitudes in the backscattered signal. While the spectrum of the backscattered signal is random, it is deterministic. Local changes in temperature or strain create a wavelength shift in the response, similar to the effect measured by FBG sensors. Interrogating the backscatter spectrum with SWI can provide a high spatial resolution (up to 10 s of microns) with strain resolution up to 1 με and temperature resolution up to 0.1 °C.121 A similar set-up can also be used to produce a distributed acoustic sensor that can be used for intrusion detection and localization.

5.5 Sources for further information and advice

There are a wide variety of available sources of information on optical fiber sensors. For an overall survey of a variety of sensors, the books of Measures,123 Udd,124 Glišić and Inaudi,125 and the Structural Health Monitoring (SHM) encyclopedia126 are all excellent options. Méndez127 also provides information on specialty optical fibers and their potential for sensing applications. Many articles related to the research and development of optical fiber sensors are available in archival journals. The field of optical fiber sensing is highly multi-disciplinary, which means that such articles are distributed among a large number of journals including Smart Materials and Structures, Measurement Science and Technology, Journal of Lightwave Technology, Applied Optics, Optical Fiber Technology, IEEE Sensors, and others.

The past 10 years have seen a rapid growth in the number of commercial businesses developing and installing optical fiber sensors and interrogators for civil infrastructure, aerospace, and energy applications. Their websites provide helpful descriptions of available products and applications. Examples include Roctest-Smartec,128 Micron Optics,53 Technobis,54 FISO,36 and Luna Innovations.129 Additionally, the organization Opticalfibersensors. org130 provides an excellent starting point to search for companies, products, upcoming events, and recent news in the field of optical fiber sensors.

5.6 Conclusions

FOSs are increasingly utilized in structural health monitoring in civil, aerospace, and energy applications. The recent surge in commercial demonstrations of these sensor systems highlights their unique capabilities that are not found in networks of conventional sensors. FOSs can be multiplexed to provide a large number of discrete sensors, or truly distributed measurements can be made using the optical fiber as a measurement device. In addition, a large variety of sensing mechanisms are available to convert the properties of lightwaves transmitted through or reflected from the optical fiber into information on the state of the structure. Their long-term durability, immunity to electromagnetic interference, and ease of integration within structural systems make FOSs ideal candidates for monitoring of full-scale structural systems. At the same time, recent research advances in optical fiber technology and instrumentation point to future capabilities and applications of FOSs that will be realized in the next few years.

5.7 References

1. Ghatak A, Thyagarajan K. Introduction to Fiber Optics. Cambridge, UK: Cambridge University Press; 1998.

2. Takeda N, Okabe Y. Durability analysis and structural health management of smart composite structures using small-diameter fiber optic sensors. Science and Engineering of Composite Materials. 2005;15:1–12.

3. Buck JA. Fundamentals of Optical Fibers. Hoboken, NJ: John Wiley & Sons Inc.; 2004.

4. Corning SMF-28: Optical Fiber Product Information. 2003.

5. Kosaka T, Takeda N, Ichiyama T. Detection of cracks in FRP by using embedded plastic optical fiber. Materials Science Research International. 1999;5:206–209.

6. Xie GP, Keey SL, Asundi A. Optical time-domain reflectormetry for distributed sensing of the structural strain and deformation. Optics and Lasers in Engineering. 1999;32:431–441.

7. Doyle C, Martin A, Liu T, et al. In situ process and conditioning monitoring of advanced fibre-reinforced composite materials using optical fibre sensors. Smart Materials and Structures. 1998;7:145–158.

8. Jiang MZ, Gerhard E. A simple strain sensor using a thin film as a low-finesse fiber-optic Fabry–Perot interferometer. Sensors and Actuators A. 2001;88:41–46.

9. Kuang KSC, Cantwell WJ, Scully PJ. An evaluation of a novel plastic optical fibre sensor for axial strain and bend measurements. Measurement Science and Technology. 2002;13:1523–1534.

10. Han W, Wang AB. Mode power distribution effect in white-light multimode fiber extrinsic Fabry–Perot interferometric sensor systems. Optics Letters. 2006;31:1202–1204.

11. James SW, Tatam RP. Optical fibre long-period grating sensors: characteristics and application. Measurement Science and Technology. 2003;14:R49–R61.

12. Emslie C. Polarization maintaining fibers. In: Méndez A, Morse TF, eds. Specialty Optical Fibers Handbook. Amsterdam, NL: Academic Press; 2001.

13. Carmen GP, Sendeckyj GP. Review of the mechanics of embedded optical sensors. Journal of Composites Technology & Research. 1995;17:183–193.

14. Zubia J, Arrue J. Plastic optical fibers: an introduction to their technological processes and applications. Optical Fiber Technology. 2001;7:101–140.

15. Measures RM. Structural Monitoring with Fiber Optic Technology. San Diego, CA: Academic Press; 2001.

16. Murukeshan VM, Chan PY, Seng OL, Asundi A. On-line health monitoring of smart composite structures using fiber polarimetric sensor. Smart Materials and Structures. 1999;8:544–548.

17. Spillman WB. Multimode polarization sensors. In: Udd E, ed. Fiber Optic Sensors. Hoboken, NJ: John Wiley & Sons, Inc; 2006.

18. Inaudi D. Long-term reliability testing of packaged strain sensors. In: Proceedings of SPIE Smart Structures Conference 2005, San Diego, CA. 2005.

19. Sennhausen U. Nellen Ph Reliability of fiber optic sensors. In: Trends in Nondestructive Testing and Inspection. Kidlington, Oxford, UK: Elsevier; 2000.

20. Inaudi D, Glisic B. Development of distributed strain and temperature sensing cables. In: Proceedings of 17th International Conference on Optical Fibre Sensors 2005, Bruges, Belgium. 2005.

21. Butter CD, Hocker GB. Fiber optics strain-gauge. Applied Optics. 1978;17:2867–2869.

22. Van Steenkiste RJ, Springer GS. Strain and Temperature Measurement with Fiber Optic Sensors. Lancaster, PA: Technomic Publishing Inc; 1997.

23. Haslach HW, Sirkis JS. Surface-mounted optical fiber strain sensor design. Applied Optics. 1991;30:4069–4080.

24. Peters KJ, Washabaugh PD. Balance technique for monitoring in situ structural integrity of prismatic structures. AIAA Journal. 1997;35:869–874.

25. Rogers AJ. Essential optics. In: Dakin J, Culshaw B, eds. Optical Fiber Sensors: Principles and Components. Boston, MA: Artech House; 1988.

26. Dandridge A. Fiber optic sensors based on the Mach-Zehnder and Michelson interferometers. In: Udd E, ed. Fiber Optic Sensors. Hoboken, NJ: John Wiley & Sons Inc; 2006.

27. Koo KP, Tveten AB, Dandridge A. Passive stabilization scheme for fiber interferometers using (3x3) fiber directional couplers. Applied Physics Letters. 1982;41:616–618.

28. Jackson DA, Priest R, Dandridge A, Tveten AB. Elimination of drift in a single-mode optical fiber interferometer using a piezoelectrically stretched coiled fiber. Applied Optics. 1980;19:2926–2929.

29. Kersey AD. Distributed and multiplexed fiber optic sensors. In: Udd E, ed. Fiber Optic Sensors. Hoboken, NJ: John Wiley & Sons Inc; 2006.

30. Inaudi D, Elamari A, Pflug L, Gisin N, Breguet J, Vurpillot S. Lowcoherence deformation sensors for the monitoring of civil-engineering structures. Sensor and Actuators A. 1994;44:125–130.

31. Lloret S, Inaudi D, Glisic B, Kronenberg P, Vurpillot S. Optical set-up development for the monitoring of structural dynamic behavior using SOFO® sensors. In: Proceedings of SPIE, 7th International Symposium on Smart Structures and Materials 2000, Newport Beach, USA. 2000;199–205. SPIE Vol. 3986.

32. Inaudi D, Bolster M, Deblois R, et al. Structural health monitoring system for the new I-35 W St Anthony Falls bridge. In: Proceedings of 4th International Conference on Structural Health Monitoring on Intelligent Infrastructure (SHMII-4) 2009, Zurich, Switzerland. 2009.

33. Glisic B, Inaudi D, Vurpillot S, Bu E, Chen CJ. Piles monitoring using topologies of long-gage fiber optic sensors. In: Proceedings of First International Conference on Structural Health Monitoring and Intelligent Infrastructure 2003, Tokyo, Japan. 2003.

34. Inaudi D, Del Grosso A, Messervey T, Walder R, Fuggini C. Long-Term Structural Health Monitoring Of The Royal Villa Of Monza. In: Proceedings of CSHM-3 International Workshop on Preservation of Heritage Structures 2010, Ottawa, Canada. 2010.

35. Pinet É, Hamel C, Glišic B, Inaudi D, Miron N. Health monitoring with optical fiber sensors: from human body to civil structures. In: 2007; Proceedings of 14th SPIE Annual Symposium on Smart Structures and Materials & Nondestructive Evaluation and Health Monitoring 2007. Vol. 6532 SPIE, San Diego (CA), USA.

36. http://www.fiso.com.

37. Friebele EJ, Askins CG, Bosse AB, et al. Optical fiber sensors for spacecraft applications. Smart Materials and Structures. 1999;8:813–838.

38. Ecke W, Latka I, Willsch R, Reutlinger A, Graue R. Fibre optic sensor network for spacecraft health monitoring. Measurement Science and Technology. 2001;12:914–980.

39. McKenzie I, Jones R, Marshall IH, Galea S. Optical fibre sensors for health monitoring of bonded repair systems. Composite Structures. 2000;50:405–416.

40. Sekine H, Fujimoto SE, Okabe T, Takeda N, Yokobori T. Structural health monitoring of cracked aircraft panels repaired with bonded patches using fiber Bragg grating sensors. Applied Composite Materials. 2006;13:87–98.

41. Li HCH, Beck F, Dupouy O, et al. Strain-based health assessment of bonded composite repairs. Composite Structures. 2006;76:234–242.

42. Mizutani T, Takeda N, Takeya H. On-board strain measurement of a cryogenic composite tank mounted on a reusable rocket using FBG sensors. Structural Health Monitoring – An International Journal. 2006;5:205–214.

43. Kang DH, Kim CU, Kim CG. The embedment of fiber Bragg grating sensors into filament wound pressure tanks considering multiplexing. NDT&E International. 2006;39:109–116.

44. Chan THT, Yu L, Tam HY, Ni YQ, Chung WH, Cheng LK. Fiber Bragg grating sensors for structural health monitoring of Tsing Ma bridge: background and experimental observation. Engineering Structures. 2006;28:648–659.

45. Zhang W, Gao JQ, Shi B, Cui HL, Zhu H. Health monitoring of rehabilitated concrete bridges using distributed optical fiber sensing. ComputerAided Civil and Infrastructure Engineering. 2006;21:411–424.

46. Kister G, Winter D, Badcock RA, et al. Structural health monitoring of a composite bridge using Bragg grating sensors Part 1: evaluation of adhesives and protection systems for the optical sensors. Engineering Structures. 2001;29:440–448.

47. Lau KT, Yuan LB, Zhou LM, Wu JS, Woo CH. Strain monitoring in FRP laminates and concrete beams using FBG sensors. Composite Structures. 2001;51:9–20.

48. Ren L, Li HN, Zhou J, Li DS, Sun L. Health monitoring system for offshore platform with fiber Bragg grating sensors. Optical Engineering. 2006;48 Art. No. 084401.

49. Bin Lin Y, Lin TK, Chen CC, Chiu JC, Chang KC. Online health monitoring and safety evaluation of the relocation of a research reactor using fiber Bragg grating sensors. Smart Materials and Structures. 2006;15:1421–1428.

50. Hill KO, Fujii Y, Johnson DC, Kawasaki BS. Photosensitivity in optical fiber waveguides application to reflection filter fabrication. Applied Physics Letters. 1918;32:641–649.

51. Kashyap R. Fiber Bragg Gratings. San Diego, CA: Academic Press; 1999.

52. Kersey AD, Davis MA, Patrick HJ, et al. Fiber grating sensors. Journal of Lightwave Technology. 1991;15:1442–1463.

53. http://www.micronoptics.com/.

54. http://www.technobis-fibre-technologies.nl/.

55. http://www.smartfibres.com/.

56. Koo KP, Kersey AD. Bragg grating-based laser sensors systems with interferometric interrogation and wavelength-division multiplexing. Journal of Lightwave Technology. 1995;13:1243–1249.

57. Todd MD, Johnson GA, Chang CC. Passive, light intensity-independent interferometric method for fibre Bragg grating interrogation. Electronics Letters. 1999;35:1970–1971.

58. Todd MD, Johnson GA, Althouse BL. A novel Bragg grating sensor interrogation system utilizing a scanning filter, a Mach-Zehnder interferometer and a 3 x 3 coupler. Measurement Science and Technology. 2001;12:771–777.

59. Lawrence CM, Nelson DV, Udd E, Bennett T. A fiber optic sensor for transverse strain measurements. Experimental Mechanics. 1999;39:202–209.

60. Bosia F, Giaccari P, Botsis J, Facchini M, Limberger HG, Salathé R. Characterization of the response of fibre Bragg grating sensors subjected to a two-dimensional strain field. Smart Materials and Structures. 2003;12:925–934.

61. Prabhugoud M, Peters K. Finite element model for embedded fiber Bragg grating sensor. Smart Materials and Structures. 2006;15:550–562.

62. Prabhugoud M, Peters K. Finite element analysis of multi-axis strain sensitivities of Bragg gratings in PM fibers. Journal of Intelligent Material Systems and Structures. 2007;18:861–873.

63. Chehura E, Ye CC, Staines SE, James SW, Tatam RP. Characterization of the response of fibre Bragg gratings fabricated in stress and geometrically induced high birefringence fibres to temperature and transverse load. Smart Materials and Structures. 2004;13:888–895.

64. Peters K, Studer M, Botsis J, Iocco A, Limberger H, Salathé R. Embedded optical fiber Bragg grating sensor in a nonuniform strain field: measurements and simulations. Experimental Mechanics. 2001;41:19–28.

65. Studer M, Peters K, Botsis J. Method for determination of crack bridging parameters using long optical fiber Bragg grating sensors. Composites Part B. 2003;34:347–359.

66. Ling HY, Lau KT, Chen L, Chow KW. Embedded fibre Bragg grating sensors for non-uniform strain sensing in composite structures. Measurement Science and Technology. 2005;16:2415–2424.

67. Ling HY, Lau KT, Su Z, Wong ETT. Monitoring mode II fracture behavior of composite laminates using embedded fiber-optic sensors. Composites Part B. 2007;38:488–497.

68. Kang DH, Park SO, Hong CS, Kim CG. The signal characteristics of reflected spectra of fiber Bragg grating sensors with strain gradients and grating lengths. NDT&E International. 2005;38:712–718.

69. Kuang KSC, Kenny R, Whelan MP, Cantwell WJ, Chalker PR. Embedded fibre Bragg grating sensors in advanced composite materials. Composites Science and Technology. 2001;61:1379–1387.

70. Guemes JA, Menéndez JM. Response of Bragg grating fiber-optic sensors when embedded in composite laminates. Composites Science and Technology. 2002;62:959–966.

71. Takeda N, Yashiro S, Okabe T. Estimation of the damage patterns in notched laminates with embedded FBG sensors. Composites Science and Technology. 2006;66:684–693.

72. Yashiro S, Okabe T, Takeda N. Damage identification in a holed CFRP laminate using a chirped fiber Bragg grating sensor. Composites Science and Technology. 2001;67:286–295.

73. Minakuchi S. Real-time detection of debonding between honeycomb core and facesheet using a small-diameter FBG sensor embedded in adhesive layer. Journal of Sandwich Structures and Materials. 2001;9:9–33.

74. Vella T, Chadderdon S, Selfridge R, et al. Full-spectrum interrogation of fiber Bragg gratings at 100 kHz for detection of impact loading. Measurement Science and Technology. 2010;21:094009.

75. James SW, Tatam RP. Optical fibre long-period grating sensors: characteristics and application. Measurement Science and Technology. 2003;14:R49–R61.

76. Dakin JP, Pratt DJ, Bibby GW, Ross JN. Distributed optical fiber Raman temperature sensor using a semiconductor light source and detectors. Electronics Letters. 1988;21:569–570.

77. Horiguchi T, Kurashima T, Tateda M. Distributed temperature sensing using stimulated Brillouin scattering in optical silica fibers. Optics Letters. 1990;15:1038–1040.

78. Niklès M, Thévenaz L, Robert P. Simple distributed fiber sensor based on Brillouin gain spectrum analysis. Optics Letters. 1995;21:158–160.

79. Niklès M, Briffod F, Burke R, Lyons G. Greatly extended distance pipeline monitoring using fibre optics. In: Proceedings of OMAE05 24th International Conference on Offshore Mechanics and Arctic Engineering, Halkidiki, Greece. 2005.

80. Ravet F, Glisic B, Niklès M, Inaudi D. Sub-millimeter crack detection with Brillouin based fiber optic sensors. IEEE Sensor Journal. 2009;9:1391–1396.

81. MacPherson WN, Flockhart GMH, Maier RRJ, et al. Pitch and roll sensing using fibre Bragg gratings in multicore fibre. Measurement Science and Technology. 2004;15:1642–1646.

82. MacPherson WN, Silva-Lopez M, Barton JS, et al. Tunnel monitoring using multicore fibre displacement sensor. Measurement Science and Technology. 2006;17:1180–1185.

83. Blanchard PM, Burnett JG, Erry GRG, et al. Two-dimensional bend sensing with a single, multi-core optical fibre. Smart Materials and Structures. 2000;9:132–140.

84. MacPherson WN, Gander MJ, McBride R, et al. Remotely addressed optical fibre curvature sensor using multicore photonic crystal fibre. Optics Communications. 2001;193:91–104.

85. Frazão O, Silva SFO, Viegas J, et al. All fiber Mach-Zehnder interferometer based on suspended twincore fiber. IEEE Photonics Technology Letters. 2010;22:1300–1302.

86. Frazão O, Silva SF, Viegas J, Baptista JM, Santos JL, Roy P. A hybrid Fabry–Perot/Michelson interferometer sensor using a dual asymmetric core microstructured fiber. Measurement Science and Technology. 2010;21:025205.

87. Duncan RG, Froggatt ME, Kreger ST, et al. High accuracy fiber-optic shape sensing. In: Peters K, ed. Bellingham, WA: SPIE; 2007;65301S-1. Proceedings of the SPIE Smart Sensor Systems and Networks: Phenomena, Technology and Applications for NDE and Health Monitoring. Vol. 6530 SPIE, 65301S-11.

88. http://www.lunainnovations.com/technologies/shape-sensing.htm.

89. Flockhart GMH, Cranch GA, Kirkendall CK. Differential phase tracking applied to Bragg gratings in multi-core fibre for high accuracy curvature-measurement. Electronics Letters. 2006;42:390–391.

90. Cranch GA, Flockhart GMH, MacPherson WN, Barton JS, Kirkendall CK. Ultra-high sensitivity two-dimensional bend sensor. Electronics Letters. 2006;42:520–522.

91. Zolla F, Renversez G, Nicolet A, Kuhlmey B, Guenneau S, Felbacq D. Foundations of Photonic Crystal Fibres. London, UK: Imperial College Press; 2005.

92. Kim HM, Kim TH, Kim B, Chung Y. Temperature-insensitive torsion sensor with enhanced sensitivity by use of a highly birefringent photonic crystal fiber. IEEE Photonic Technology Letters. 2010;22:1539–1541.

93. Birks TA, Knight JC, Russell PS. Endlessly single-mode photonic crystal fiber. Optics Letters. 1997;22:961–963.

94. Gayraud N, Kornaszewski LW, Stone JM, et al. Mid-infrared gas sensing using a photonic bandgap fiber. Applied Optics. 2008;47:1269–1277.

95. Nasilowski T, Martynkien T, Statkiewicz G, et al. Temperature and pressure sensitivities of the highly birefringent photonic crystal fiber with core asymmetry. Applied Physics B. 2005;81:325–331.

96. Bock WJ, Urbańczyk W, Wojcik J. Measurements of sensitivity of the single-mode photonic crystal holey fibre to temperature, elongation and hydrostatic pressure. Measurement Science and Technology. 2004;15:1496–1500.

97. Wang YP, Xiao L, Wang DN, Jin W. Highly sensitive long-period fiber-grating strain sensor with low temperature sensitivity. Optics Letters. 2006;31:3414–3416.

98. He Z, Zhu Y, Du H. Effect of macro-bending on resonant wavelength and intensity of long-period gratings in photonic crystal fiber. Optics Express. 2007;15:1804–1810.

99. Petrovic JS, Dobb H, Mezentsev VK, Kalli K, Webb DJ, Bennion I. Sensitivity of LPGs in PCFs fabricated by an electric arc to temperature, strain and external refractive index. Journal of Lightwave Technology. 2007;25:1306–1312.

100. Martelli C, Canning J, Groothoff N, Lyytikainen K. Strain and temperature characterization of photonic crystal fiber Bragg gratings. Optics Letters. 2005;30:1785–1787.

101. Silva-López M, Fender A, MacPherson WN, et al. Strain and temperature sensitivity of a single-mode polymer optical fiber. Optics Letters. 2005;30:3129–3131.

102. Kiesel S, Peters K, Hassan T, Kowalsky M. Behaviour of intrinsic polymer optical fibre sensor for large-strain applications. Measurement Science and Technology. 2007;18:3144–3154.

103. Kiesel S, Peters K, Hassan T, Kowalsky M. Calibration of a singlemode polymer optical fiber large-strain sensor. Measurement Science and Technology. 2009;20:034016.

104. Jiang C, Kuzyk MG, Ding JL, Johns WE, Welker DJ. Fabrication and mechanical behavior of dye-doped polymer optical fiber. Journal of Applied Physics. 2002;92:4–12.

105. Kuzyk MG. Polymer Fiber Optics: Materials, Physics and Applications. Boca Raton, FL: CRC/Taylor & Francis; 2001.

106. Large MCJ, Moran J, Ye L. The role of viscoelastic properties in strain testing using microstructured polymer optical fibres (mPOF). Measurement Science and Technology. 2009;20:034014.

107. Zhang C, Zhang W, Webb DJ, Peng GD. Optical fibre temperature and humidity sensor. Electronics Letters. 2010;46:643–644.

108. Peters K. Polymer optical fiber sensors – a review. Smart Materials and Structures. 2011;20:013002.

109. Garvey DW, Zimmerman K, Young P, et al. Single-mode nonlinear-optical polymer fibers. Journal of the Optical Society of America A. 1996;18:2011–2023.

110. Large MCJ, Poladian L, Barton GW, van Eijkelenborg MA. Microstructured Polymer Optical Fibers. Berlin: Springer-Verlag; 2007.

111. Kiesel S, Peters K, Hassan T, Kowalsky M. Large deformation in-fiber polymer optical fiber sensor. IEEE Photonics Technology Letters. 2008;20:416–418.

112. Abdi O, Peters K, Kowalsky M, Hassan T. Validation of a singlemode polymer optical fiber sensor and interrogator for large strain measurements. Measurement Science and Technology. 2011;22 Art. No. 015201.

113. Abdi O, Wong KC, Hassan T, Peters KJ, Kowalsky MJ. Cleaving of single mode polymer optical fiber for strain sensor applications. Optics Communications. 2009;282:856–861.

114. Peng GD, Xiong Z, Chu PL. Photosensitivity and gratings in dyedoped polymer optical fibers. Optical Fiber Technology. 1999;5:242–251.

115. Liu HY, Liu HB, Peng GD. Tensile strain characterization of polymer optical fibre Bragg gratings. Optics Communications. 2005;251:31–43.

116. Liu HY, Peng GD, Chu PL. Thermal tuning of polymer optical fiber Bragg gratings. IEEE Photonics Technology Letters. 2001;13:824–826.

117. van Eijkelenborg MA, Large MCJ, Argyros A, et al. Microstructured polymer optical fibre. Optics Express. 2001;9:319–321.

118. Hiscocks MP, van Eijkelenborg MA, Argyros A, Large MCJ. Stable imprinting of long-period gratings in microstructured polymer optical fibre. Optics Express. 2006;14:4644–4647.

119. Carroll KE, Zhang C, Webb DJ, Kalli K, Argyros A, Large MCJ. Thermal response of Bragg gratings in PMMA microstructured optical fibers. Optics Express. 2001;15:8844–8850.

120. Dobb H, Webb DJ, Kalli K, Argyros A, Large MCJ, van Eijkelenborg MA. Continuous wave ultraviolet light-induced fiber Bragg gratings in few- and single-mode microstructured polymer optical fibers. Optics Letters. 2005;30:3296–3298.

121. Gifford DK, Kreger ST, Sang AK, et al. Swept-wavelength interferometric interrogation of fiber Rayleigh scatter for distributed sensing applications. Proceedings of SPIE Fiber Optic Sensors and Applications V. 2007;6770 6110OF.

122. Kreger S, Sang AK, Gifford DK, Froggatt ME. Distributed strain and temperature sensing in plastic optical fiber using Rayleigh scatter. Proceedings of SPIE Fiber Optic Sensors and Applications VI. 2009;7316:73160A.

123. Measures RM. Structural Monitoring with Fiber Optic Technology. San Diego, CA: Academic Press; 2001.

124. Udd E, ed. Fiber Optic Sensors. Hoboken, NJ: John Wiley & Sons, Inc.; 2006.

125. Glišić B, Inaudi D. Fibre Optic Methods for Structural Health Monitoring. Hoboken, NJ: John Wiley & Sons, Inc.; 2007.

126. Boller C, Chang FK, Fujino Y, eds. Encyclopedia of Structural Health Monitoring. Chichester, UK: John Wiley & Sons, Inc; 2009.

127. Méndez A, Morse TF. Specialty Optical Fibers Handbook. Amsterdam: Academic Press; 2007.

128. www.roctest-group.com.

129. www.lunatechnologies.com.

130. www.opticalfibersensors.org.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset