Next Article in Journal
Distributed Cooperative Control of Flexible Spacecraft Based on PDE-ODE Coupled Dynamics Model
Previous Article in Journal
The Development of a Prototype for Low Altitude Operations of Unmanned Aircraft Flight Plan Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Fiber Optic Sensing Technology in Aerospace: Packaging, Bonding, and Calibration Review

1
School of Agricultural Engineering, Jiangsu University, Zhenjiang 212013, China
2
School of Instrument Science and Engineering, Southeast University, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
Aerospace 2025, 12(9), 827; https://doi.org/10.3390/aerospace12090827 (registering DOI)
Submission received: 21 July 2025 / Revised: 8 September 2025 / Accepted: 12 September 2025 / Published: 15 September 2025
(This article belongs to the Section Aeronautics)

Abstract

With the continuous development of science and technology, aircraft structural health monitoring (SHM) has become increasingly important in the aviation field. As a key component of SHM, wing deformation monitoring is of great significance for ensuring flight safety and reducing maintenance costs. The traditional strain gauge measurement method can no longer meet the needs of modern aeronautical engineering. Fiber Bragg grating (FBG) sensors have been widely used in the engineering field due to their unique advantages, and have shown great potential in aircraft wing deformation monitoring. In the context of SHM in the aircraft field, this article provides an overview of four aspects: classification and principles of fiber optic sensors, packaging forms of FBG sensors, bonding technology, and calibration technology. The packaging forms includes tube-packaged, embedded package and surface-attached package. It then discuss the bonding technology of FBG sensors, and the principle and influencing factors of fiber optic bonding technology are analyzed. Finally, it conducts in-depth research on the calibration technology of FBG sensors. Through comprehensive analysis of these four aspects, the suggestions for optical fiber sensing technology in aircraft wing deformation measurement are summarized and put forward.

1. Introduction

With the development of current aviation technology, large aircraft SHM has become an indispensable part of aviation engineering. As an important part of SHM, wing deformation monitoring is very important to ensure flight safety and reduce maintenance costs [1,2,3]. As shown in Figure 1, in order to measure the structural health of aircraft, a traditional strain gauge is usually used for measurement, but this can form a complex sensor network [4]. The weight of the connection wire associated with the strain gauge is also very large, which will affect the strain measurement response of the structure. Thus, a convenient and high-precision wing deformation monitoring system for flight state is badly needed. FBG sensors have been widely used in the engineering field because of their low weight, anti-electromagnetic interference, low cost and easy mass production [5,6,7].
The NASA Langley Research Center used optical fiber sensors and strain gauges in the test and performed multiple flight maneuvers and the results show that the FBG sensor system performed well throughout entire flight [8]. Researchers of NASA used 2D shape sensing to measure wing deflection and the results verified the excellent characteristics of FOSS in wing shape measurement. In 2018, researchers of NASA published a research report on the use of optical fiber shape sensing [9]. As shown in Figure 2, researchers have applied optical fiber strain sensor networks to the strain and displacement monitoring of wings in a small unmanned aerial system (SUAS), and the results show that the methods and techniques of load redistribution and active shape control proposed by the researchers are effective. Martins et al. used the distributed optical fiber sensor network to measure the dynamic strain of aerospace structures, and the results show that the SHM algorithm based on the distributed optical fiber sensor network can represent the dynamic behavior of the wing model globally [10].
Hyuk et al. designed a full-scale static load detection system for measuring the temperature-compensated strain of the wing structure of a full-scale small aircraft [11]. The researchers applied the load to the tip of the wing structure, and the low-speed interrogator could detect changes in the central wavelength of the FBG sensors at 10 Hz sampling rate. The experiments verify the applicability of the low-speed FBG interrogator in small aircraft. He et al. studied the influence of optical fiber sensors with different embedded depths on shape monitoring of deformed wings, and analyzed the sensitivity of FBG sensors with different embedded depths in polyimide film [12]. The end curvature of the wing is measured by FBG sensors, and then the three-dimensional shape of the skin is reconstructed by integration. The experimental results show that the three-dimensional shape obtained by FBG sensors is in good agreement with the results obtained by visual measurement.
The research progress of many scholars above shows the outstanding effect of fiber optic sensors in detecting aircraft wing deformation. In addition, fiber optic sensors can achieve structural health monitoring of aircraft, strain measurement during flight, environmental temperature measurement during aircraft operation, among others. For various application scenarios, this paper categorizes fiber optic sensors according to their application situations, as shown in Table 1.
As shown in the table above, fiber optic sensors have a variety of application scenarios. However, in practical engineering, the packaging and bonding of fiber optic sensors may be subject to certain interference, which will affect the measurement results. Based on this reason, this paper will introduce the classification and working principle of fiber optic sensors in aircraft application scenarios, analyze the packaging form of fiber optic sensors, and discuss the factors that affect the strain transmission efficiency of fiber optic sensors.
To ensure the rigor and credibility of the research, this paper summarizes the research results of using FBG sensing technology for aircraft wing deformation monitoring through a standardized literature search. The main data sources are Web of Science, Scopus, and IEEEXplore databases, with literature concentrated in academic research literature from the past decade. This review follows the following approach when selecting the literature: (1) the research focuses on the application of FBG sensors for aircraft wing deformation monitoring, rather than their use in civil engineering or marine detection; (2) includes data or engineering validation from wind tunnel or flight experiments, with verifiable engineering data or experimental evidence; and (3) the core technology research direction involved in the article (sensor types, packaging forms, bonding forms, calibration methods) can support the research content of this article. Therefore, the literature selected from the above literature is attached to the reference list for application in this study.

2. Classification of Optical Fiber Sensors

The classification of fiber optic sensors can be based on the characteristics of the materials. This section introduces several main types, including FBG sensors, sensors based on Rayleigh scattering and Brillouin scattering, etc., and analyzes their principles and application characteristics [27].

2.1. FBG Sensor

K.O. et al. discovered the photosensitive effect of germanium-doped fiber, and successfully developed the world’s first fiber grating by standing wave method [28]. Meltz et al. used two coherent ultraviolet beams to write a fiber Bragg grating into a hydrogen-loaded fiber. The fabrication technology makes fiber Bragg grating have potential application value [29]. Hill et al. proposed the phase mask method, which promoted the industrialization of fiber Bragg grating applications, and promoted the application of fiber Bragg grating in the sensing field, and the principle is shown in Figure 3 [30]. When broadband light enters the fiber Bragg grating (FBG) from the left side, due to the periodic refractive index modulation in the fiber core, it will selectively reflect light of a specific wavelength, forming a back reflection spectrum, and sharp reflection peaks appear at the corresponding wavelength. At the same time, the unreflected light is transmitted through, and the transmission spectrum shows troughs at these specific wavelengths. The principle behind a FBG sensor is to precisely measure the Bragg wavelength of the reflected radiation. Since this is sensitive to environmental influences like temperature or strain, any changes in the Bragg wavelength indicate a presence of the disturbance that one wants to measure.

2.2. Rayleigh Scattering

In the optical fiber production process, there is a phenomenon that the material of fiber core and the refractive index inside are not even hooked. When the light wave propagates in the fiber, it produces the scattering in all directions, which is called Rayleigh scattering. Rayleigh scattering belongs to elastic scattering, and the frequency of scattered light is the same as that of incident light wave. The intensity of Rayleigh scattering is directly proportional to the original light intensity, and the ratio of the two is called Rayleigh scattering coefficient [31,32,33,34].
Rayleigh scattering is a naturally occurring phenomenon caused by the elastic collision between incident light and micro-particles in the medium. During the fabrication of optical fibers, the refractive index distribution in the optical fiber has a slight difference and impurity. The defect of this structure causes the fluctuation of dielectric constant, which makes the amplitude and phase of Rayleigh scattering random. When the optical fiber is manufactured, its refractive index distribution is very stable, and Rayleigh scattering can be considered as a long-term stable phenomenon of the optical fiber itself.
Barnosky et al. first proposed OTDR technology based on Rayleigh scattering [35]. OTDR technology can effectively measure the attenuation and continuity of optical fibers. It is very suitable for measuring the length of optical fibers, determining the loss of optical fibers in all areas and locating the fault points in optical fibers. Daichi et al. designed the flight test bed and they used their own OFDR system to demonstrate the distributed sensing flight of the main wing of a medium-sized jet airliner [36]. OFDR-FBG sensing system can measure the strain distribution in FBG in real time with millimeter-level spatial resolution. The results show that the OFDR-FBG sensing system can successfully monitor the variation in strain distribution according to various operations, and the measured strain is in good agreement with the reference strain.

2.3. Brillouin Scattering

Ippen et al. first observed Brillouin scattering spectrum in optical fibers [37]. Since then, researchers have conducted extensive research on stimulated Brillouin scattering (SBS) technology in optical fibers, and Horiguchi et al. proposed Brillouin optical time-domain analysis (BOTDA) in 1989 [38]. Researchers have found that BOTDA has the characteristics of long measuring distance and high measuring accuracy [39,40].
According to the characteristics of the sensor, the distributed sensors also includes Distributed Phase-Encoded Brillouin Optical Time-Domain Analysis (DPP-BODTA), Brillouin Optical Time-Domain Reflectometer (BOTDR), Optical Frequency-Domain Reflectometer (OFDR), Phase Optical Time-Domain Reflectometer (OTDR) [41,42,43]. The different characteristics are listed in Table 2.
Fiber optic sensors applied in the aviation field also cover more technology routes designed for specific scenarios. These technologies are optimized around different usage requirements such as high temperature resistance, miniaturization, and multi-parameter monitoring, further expanding the application boundaries of fiber optic sensing in aircraft systems, as shown in Table 3.

3. Packaging Forms of FBG Sensors

According to the different packaging technology of FBG, FBG sensors can be divided into tube-packaged, embedded packaging and surface-attached packaging. These three monitoring methods have their own characteristics and applicable conditions. In application, different monitoring methods should be selected according to the different mechanical states of the matrix.

3.1. Tube-Packaged FBG Sensors

Researchers have performed extensive research on tube-packaged FBG sensors and have applied them to different engineering fields [58,59,60,61,62]. Generally, a tube-packaged FBG sensor consists of a bare FBG placed inside a tube sleeve with a certain prestressing force applied to keep the FBG straight, and then encapsulation glue poured between the FBG and the fiber, so that the FBG is firmly embedded in the tube sleeve. The encapsulation adhesive has a certain strength and can transfer the strain of the structure to the FBG well.
Ivan et al. introduced a new method to package and install a large number of sensors on metal or composite structures based on the application background of naval ships [63]. The researchers noticed that the long FBG sensors had a high fracture sensitivity in the splicing area. They used Vacuum-Assisted Resin Transfer Molding (VARTM) technology to conduct large-scale test simulation in the laboratory. The results show that this method is suitable for large area ships which need long-term SHM. The method is simple and easy to use, and can avoid complex problems, especially in the field of long FBG sensors management. The method has been proved to be suitable for ships by experiments, but the wing of aircraft has high flexibility and works in totally different harsh environments. Only a limited number of FBG sensors can be used on aircraft wings, and Floris’s research [64] found that the advantages of new shape sensors based on long FBG make up for this limitation. Whether VARTM technology is applicable to aircraft wings can be further verified by experiments in combination with Floris’s research.
Li et al. proposed a capillary encapsulation method. In this method, zinc capillary and different binders were used to package FBG, and the temperature sensing characteristics of FBG in a constant-temperature water bath were studied [65]. The results show that this packaging method can protect the FBG well and improve the temperature sensitivity of FBG effectively. However, it should be noted that the material properties of capillary and adhesive have an effect on the temperature sensitivity of FBG. Different application backgrounds and strength requirements have different requirements for the binder materials.
Sagar et al. proposed a new metal encapsulation method for FBG sensors using stainless steel, tin and high-temperature-resistant samarium cobalt (SMCO) magnets [66]. The encapsulated sensor can be directly contacted with the matrix structure such as iron tube and other ferromagnetic elements without any adhesive, making it easy to disassemble and reuse. The feasibility of using metal-encapsulated FBG sensors for structural health monitoring is verified, which provides a good choice for SHM especially for ferrous pipeline monitoring and for welded structures.
Zhang et al. studied the characteristics of metallic-packaging FBG sensors with coatings and bare fibers, as shown in Figure 4 [67]. Researchers developed metallic-packaging FBG sensors by ultrasonic welding. The good combination of metal alloy with coating and bare optical fibers is clearly shown by cross-sectional scanning electron microscopy (SEM). Compared with bare FBG, polyimide coated FBG for metal packaging has advantages in stability, repeatability, and response.

3.2. Embedded FBG Sensor

Researchers have also performed extensive research on embedded FBG sensors, which has been applied in different engineering fields, and it has been widely used in the field of damage monitoring [68,69,70,71,72]. An embedded FBG sensor is a kind of technology that embeds FBG sensor in the matrix and fills some material to make the fiber and the matrix contact closely. In this way, the stress and strain changes in the matrix can be monitored by the optical fiber. An FBG sensor has the advantages of small size and high sensitivity, so embedded FBG in intelligent composite material not only keeps the FBG’s sensitivity, anti-electromagnetic interference, and other advantages, but also does not affect the performance of composite aeronautical structures [73].
The commonly used embedded FBG sensors are carbon-fiber-reinforced polymer (CFRP) FBG sensors, resin matrix composite FBG sensors, and fiber-reinforced polymer (FRP) FBG sensors, among others [74,75]. Researchers found that the optical fiber sensors directly embedded in the adhesive will lead to a significant reduction in the bonding strength, which directly affects the normal operation of the embedded FBG sensor in the aircraft [76]. Grundmann et al. [77] have performed extensive research on the bond strength of the optical fiber sensors embedded in the CFRP film bonding joints, and results show that the influence of embedding glass fiber with a total diameter less than 100 μm on the fatigue limit of film adhesive joints can be ignored. It is suggested that the P80 optical glass fiber is highly suitable for the application of SHM in aeronautical structures. Murayama et al. installed the embedded optical fiber sensors into the full-scale composite structure of the vehicle, and monitored the temperature and strain distribution of the fuselage during the manufacturing process [78]. The results show that the application of embedded optical fiber sensors in the health monitoring of composite structures is feasible.
Baker et al. developed FBG embedding technology for military aircraft [79]. A new and effective implementation process strategy is proposed, which overcomes the technical constraints of previous problems. The end of the FBG sensor is sealed in a Teflon envelope and then sealed with cement. A frankincense dam is built around the sample to prevent resin flow. The results show that the method is effective and easy to implement in aircraft structural health monitoring. Botsis et al. studied the application of embedded FBG sensors in internal strain measurement [80], and the results show that the reasonable use of embedded FBG sensors combined with appropriate models can be used to characterize the deformation of polymer materials at different scales. Li et al. analyzed the strain transfer characteristics of embedded FBG sensors under non-axial stress [81]. Researchers established a suitable strain transfer model and analyzed the influence of sensor positioning angle deviation on average strain transfer rate. The results show that the angle of embedded FBG sensor plays a vital role in the transfer of strain from the surrounding material to the core of the optical fiber. Takeda et al. of Japan applied embedded optical fiber sensing system to health monitoring of composite structures of smart aircraft [82]. Researchers have developed a real-time impact damage detection system embedded in small-diameter OFSS, and demonstrated it in CFRP fuselage structure with diameter of 1.5 m and length of 3 m. In order to effectively detect impact damage, researchers embedded small-diameter stainless steel in the upper panel.
Liang et al. used embedded FBG sensors to monitor the strain changes during loading, and compared the measured results with the theoretical results [83]. The validity and reliability of embedded FBG sensors for structural health monitoring are studied through experimental data. The experimental results show that the measured results are in good agreement with the theoretical results, and the straightness can reach more than 0.9998. As shown in Figure 5, Maxim et al. used embedded FBG sensors to layout on the wing frame for wind tunnel research and the layout array and method of sensors are distinguished by different colors in the figure [84]. The results show that embedded FBG sensors can provide higher spatial resolution measurement compared with accelerometers and strain gauges. They are well suited for embedded composite wing structures and do not require additional wiring for each sensor. Hyunseok et al. proposed a load monitoring system for a composite wing based on embedded FBG sensors [85]. The embedded FBG sensor is embedded in the main structure of the composite wing, and the strain response of the embedded FBG sensor is used to estimate the wing load. The experimental results show that the average error of this wing load estimation method is 4.19%, which shows that the embedded FBG sensors have certain advantages in wing load monitoring.
He et al. embedded FBG sensors in PVC-enhanced silicon substrate to monitor the three-dimensional shape of the morphing wing [86]. Due to the lack of flexibility of PVC substrate, the adaptability of FBG sensor is reduced. The soft silicone skin can protect the FBG sensor and enhance its adaptability. The researchers combined the advantages of the two materials to greatly improve the sensitivity of the FBG sensor. The researchers also used visual measurement experiments to prove the effectiveness of the FBG sensing method, and the results show that the effect of morphing-wing shape reconstruction based on embedded FBG sensor method is in good agreement with that of visual measurement, and the maximum error is less than 3%. It can be seen that the embedded FBG sensor has a good effect on the SHM of morphing-wing UAV.

3.3. Surface-Attached FBG Sensor

Researchers have performed extensive research on the surface-attached FBG sensor [87,88,89,90]. Compared with the tube-packaged FBG sensor and the embedded FBG sensor, the surface-attached FBG sensor has higher requirements for the bonding process. Some researchers attach FBG sensors directly to the surface of the measured object, and some researchers also cut a small groove in the sensor base and then fixed the bare FBG in the groove [91,92,93]. The main purpose of the groove is to increase the contact area between the substrate and the fiber, so that the strain of the substrate can be effectively transmitted to the FBG.
In order to meet the needs of structural testing and improve the performance of sensors, researchers use different materials to package bare optical fibers. Common matrix materials are copper sheet, aluminum sheet, alloy port sheet, polymer composite materials, etc. The binders used for packaging are epoxide resin, silica epoxide, ceramic binder, among others. Compared with the tube-packaged strain sensor, the surface-attached strain sensor is easy to fabricate, simple in structure and suitable for monitoring the surface strain of the carrier. Sirkis et al. used an integral equation to describe the phase change in light propagating in a single-mode optical fiber when it propagates in a single-mode optical fiber with arbitrary strain field [94]. The hypothesis that strain can be transferred by surface-attached FBG is verified by experiments, which indicates the feasibility of surface-attached FBG for strain measurement. Duncan et al. used distributed sensing systems to monitor structural fatigue of aircraft undergoing full-scale fatigue tests [95]. The researchers used surface-attached optical fiber sensors with high-density on Lockheed Martin’s P-3C Orion fatigue test. The results prove the feasibility of using OFDR technology in distributed optical fiber sensing for aircraft structure monitoring.
Li et al. tested the strain transfer performance of surface-attached FBG sensor with metal welding instead of epoxy [96]. The results show that the average strain transfer efficiency of metal welding surface-attached FBG sensor is kept at 99.2%, which can give full play to the strain measurement ability of surface-attached optical fiber sensors. Sidney et al. designed a special coated surface-attached FBG to evaluate the working quality of the surface-attached FBG sensor by exposing the installed sensor to a full set of real flight conditions [97]. The applied test conditions are based on aerospace standards, including temperature cycle, pressure cycle, exposure to humidity and hydraulic fluid, and fatigue load. The results show that the surface-attached FBG is compatible with flight conditions when it is fully attached on aerospace CFRP specimens.
Sodja et al. introduced the FBG sensing system for aircraft wing deformation reconstruction [98]. According to the position of surface-attached FBG strain sensor on the wing skin, two sensor configurations are formed along the axis and spars, respectively. The researchers used 20 strain rosettes at both aircraft wing skins and 8 FBG sensors per spar. Each skin also has 10 temperature FBG sensors for the correction of thermal strain. The design is very systematic. The compatibility, range, and coast down test of the air brake, engine, and electronic components are also introduced. This study is a good reference for the research of wing deformation sensing technology.
In order to visually compare the differences between tube-packaged, embedded, and surface-attached FBG sensors, and to select the appropriate packaging form according to their needs in practical applications, as shown in Table 4, a systematic comparison is made between them from multiple characteristic dimensions such as core integration methods, target sensing applications, and detection of physical parameters.
Based on the review research in this part, it can be concluded that for the monitoring scenario of aircraft wing deformation, the specific selection should be determined according to the characteristics of different packaging methods. Tube-packaged FBG is suitable for measuring points that require long-term stability, such as the main wing beam, and has strong seismic resistance. Once damaged, the casing needs to be disassembled and replaced, making maintenance more complex. Embedded FBG is preferred for CFRP composite wings, which can monitor internal strain and be compatible with the curing temperature of composite materials. It has outstanding survival ability when integrated with the wing structure, but requires replacement of the overall composite material layer for maintenance. It has high strain transmission efficiency and is not affected by surface environmental interference. Surface-attached FBG is suitable for rapid deployment or modification of existing wings, can be directly replaced and maintained, and are easy to visually inspect, but are susceptible to surface damage that affects their survivability.

4. Bonding Technology of FBG Sensor

4.1. Principle of Optical Fiber Bonding Technology

As shown in Figure 6, the adhesive bonding mechanical transfer model [126,127] consists of FBG sensor, adhesive layer and measured structure. When the matrix is subjected to external forces, stress is transmitted to the grid region through an adhesive layer. During the stress transmission process, there is shear stress inside the adhesive layer, causing the stress to gradually change along the length direction (x-axis) from the matrix side’s. The fiber is connected to the substrate through an adhesive layer. When the substrate undergoes deformation under external force, the adhesive layer transmits the strain of the substrate to the fiber Bragg grating through shear action, thereby achieving sensing measurement.
Researchers have performed extensive research on the strain transfer characteristics of FBG sensors, and achieved differing results. The factors influencing strain transfer characteristics have been analyzed. The relationship between various factors and transfer rate has been provided in [128,129,130,131,132,133]. Due to the effects of the protective coating and adhesive layer, the strain between the optical fiber and the host material will be inconsistent. In studies on strain transfer efficiency, many focus on the theoretical analysis of strain transfer from the host material to the optical fiber. Shiuh-Chuan et al. established a theoretical model for the transmission of strain from the host material to the optical fiber in order to study the measurement accuracy of the optical fiber strain sensor [134]. Then, the theoretical model was used to evaluate the interaction between the host material and the coating. Through the finite element numerical analysis, the correctness of the theoretical prediction is verified by the researchers. Researchers used Mach–Zehnder interferometric optical fiber sensors to measure strain, and the strain difference between the optical fiber strain sensor and the specimen was measured in experiments. The results show that the strain measured on the optical fiber is lower than the true strain of the sample. The percentage of strain actually transferred to the optical fiber in the sample depends on the bonding length of the optical fiber and the protective coating. The general trend of strain transition obtained from experimental tests and theoretical predictions shows that the longer the bonding length is, the harder the coating is, and the more strain is transmitted to the optical fiber. Wang et al. analyzed the strain transfer of six-layer surface-attached FBG sensors under uniform axial stress [135]. A six-layer structure, which is different from the generally recognized four-layer structure, is proposed. It consists of optical fibers, protective coatings, bonding layers, base layers, external bonding layers, and host materials. The theoretical formula of strain transfer rate from the host material to the optical fiber is established, which provides a basis for accurate theoretical prediction. Through parameter analysis, the researchers determined the selection scheme of sensor parameters, which provided a reference for sensor design. The experimental results show that the measured data are in good agreement with the theoretical prediction. The simulation results align well with the theoretical predictions. Finally, the surface-attached FBG model with one substrate layer is extended to a general model with multiple substrate layers, which provides a theoretical basis for future research and design of surface-attached FBG.

4.2. Influencing Factors

4.2.1. Effect of Bonding Thickness

In the bonding process, the function of the bonding layer is to firmly adhere the bare FBG to the surface of the measured substrate, so the bonding layer needs a certain thickness. If the bonding layer is too thin, the FBG sensor is prone to fall off when the matrix is deformed. If the bonding layer is too thick, the strain of the FBG sensor is prone to slow down. Therefore, the influence of the thickness of the bonding layer on the strain transfer of the FBG sensor cannot be ignored.
Kai et al. studied the strain transfer effect of surface-attached FBG sensors [136]. A three-dimensional finite element model of a surface-attached strain sensor was established. The effects of side width, top thickness, bonding length, and bottom thickness on strain transfer rate were studied. The shear lag characteristic is modified by empirical method, which improves the calculation efficiency of strain transfer coefficient. The results show that bonding length and bottom thickness are the main factors affecting strain transfer rate. Wu et al. analyzed the strain transfer mechanism of FBG sensors [137]. Researchers put forward a polynomial shear stress distribution mathematical model, established a strain transfer function, and calculated the strain of the fiber center using ANSYS Workbench. The results show that the strain transfer function derived from polynomial shear stress is very close to the actual situation and meets the accuracy requirements. In addition, the strain values obtained in the middle part of the FBG sensors are accurate, but the strain values at both ends are lower than the actual values. Therefore, FBG sensors need to be long enough. The experimental data also show that the thickness of the viscous layer and the elastic modulus have great influence on the strain transfer coefficient. With the decrease in bonding thickness and the increase in elastic modulus, the strain transfer coefficient and average strain transfer coefficient gradually increase, and the effective length of FBG sensor also increases.
Through analysis, it can be seen the strain transfer coefficient increases with the decreases in the thickness of the bonding layer, that is, the greater the thickness of the bonding layer, the less sufficient the strain transfer. Therefore, in the process of bonding the bare FBGs and the host substrate, under the premise of ensuring that the bare grid and the substrate are firmly bonded, the thinner the thickness of the cement layer, the more sufficient the strain transmission. With the increase in shear modulus of bonding layer, the strain transfer coefficient increases and the strain transfer becomes more sufficient.

4.2.2. Effect of Bonding Width

Besides the influence of the thickness of bonding layer, the width of bonding also affects the strain transfer efficiency. The bonding width is the width perpendicular to the axial direction of the fiber when the bare FBG sensor is bonded to the substrate surface.
Zhao et al. proposed an improved surface-attached FBG sensor connection method for airship hull SHM [138]. The strain transfer relationship between the airship cladding and the core is predicted by theoretical formulas, and then the theoretical prediction is discussed by finite element analysis. On the basis of theoretical analysis and numerical verification, the geometrical and mechanical parameters of the adhesive layer affecting strain transfer rate and effective sensing length ratio were analyzed. The experimental results show that width and Young’s modulus are the key parameters, while bottom thickness and Poisson’s ratio have little effect. Guang et al. proposed a new bonding model to accurately evaluate the measurement accuracy of FBG sensors and correct the strain obtained [139]. The correctness of the bonding model was verified by numerical method, and a theoretical strain correction coefficient was obtained. The sensitivity of the strain correction coefficient to five main parameters was discussed. It was proved that the correction coefficient had significant sensitivity to the coating material and bonding length. The study provides a practical and accurate correction for surface-attached FBG sensors installed with adhesives, and helps to deduce the true strain in the measured object. Finally, the researchers suggested that in practical applications, FBG sensors with shorter length, larger area, and that are thinner improve the correction accuracy.

4.2.3. Effect of Bonding Length

In practical applications, the bonding length of FBG is also a factor to be considered, which directly affects the stress concentration of the sensor on the measured surface and the accuracy of surface measurement. Shiuh-Chuan et al. studied the effect of bonding length on strain transfer efficiency [140]. Researchers used Mach–Zehnder interferometric technology to perform experimental measurements. The experimental results show that the strain transfer increases from 56% to 82% as the bonding length of epoxide resin increases from 5 cm to 12 cm. That is to say, the actual percentage of strain transferred to optical fibers depends on the bonding length of optical fibers and adhesives. The experimental results verify the correctness of the theoretical prediction and reveal the strain difference between the optical fiber strain sensor and the specimen. The general trend of strain transfer obtained by researchers from both experimental and theoretical predictions shows that the longer the bonding length, the higher the bonding strength, and the more strains transmitted to optical fibers. Kwona et al. studied the effect of bonding length on the signal characteristics of surface-attached FBG sensors under different load types [141]. The experimental results show that the strain transfer ratio varies with the loading mode. If the FBG sensors are not equipped with enough bonding length, the strain cannot transfer adequately. Therefore, it is necessary to determine the effective combination length of different load types in order to better obtain stable signals from surface-attached FBG sensors.
However, this can cause the displacement of the measuring point and the inaccurate position of the measuring point if the sensor is too long. If the sensor is too short and the strain transmission is not enough, it is necessary to study the influence of the bonding length of the FBG sensor on the strain transmission of a specific substrate. It can be seen that the strain transfer coefficient decreases with the increase in bonding length, that is, the strain transfer becomes sufficient. Therefore, in the actual strain measurement, the bonding length of the bare grating should be increased as much as possible to meet its good strain transfer and the measurement accuracy will decrease if the bond length is too long. The strain measurement should be carried out according to the usual embedded length of the FBG sensor.

5. Calibration Technology

The calibration measurement method is a vital link in the process of SHM technology. NASA has mature technology to calibrate the wing with strain gauges [142,143,144,145]. FBG sensor has many advantages over traditional strain gauges. Therefore, in addition to the installation form of optical fiber sensors and the design of bonding technology, it is also necessary to calibrate the optical fiber sensors accurately to achieve long-term, effective, and accurate measurement.
Hönisch et al. made a thorough study on the details of strain transfer and the corresponding calibration procedures of resistance strain gauges [146]. Resistance strain gauges are mostly calibrated, verified, and applied according to VDE/VDI 2635 guidelines [147,148,149]. Inaudi et al. used double FBG sensors to measure the mechanical strain on the surface of an aluminum beam [150]. The results show that strain compensation can be quantified by introducing the strain delay factor into an FBG sensor during heating. Wolfgang R Habel proposed a verification method to evaluate the overall quality of sensor system functions to promote the use of optical fiber sensors [151].
As shown in Figure 7, according to the German VDI/VDE 2660 guideline, Roths et al. of Germany developed a strain calibration device, which can provide reference strain conditions with less uncertainty without using electrical strain gauges [152]. Through a series of calibration measurements of two different types of FBG strain sensor components and a specially developed bonding process, a calibration device based on four-point bending structure of stainless steel beams was established. The device can provide a reference strain condition within ±1000 μm/m with a relative uncertainty of about 0.3%. The results show that the relative standard deviation of calibration measurement for the same type of FBG sensor is about 0.16%. and there are significant differences in strain sensitivity between the two FBG sensors. The strain calibration device designed by researchers has a certain reference significance for improving the calibration accuracy.
Habel et al. developed a calibration test device for FBG sensors [153]. The unbiased digital image correlation (DIC) and electronic speckle interferometer (ESPI) are used for calibration measurement, and the resolution of DIC is 2 um and ESPI is 30 nm. The KALFOS (calibration of optical fiber sensors) facility designed by researchers consists of a load-bearing facility and a temperature chamber. In the experiment, the temperature and load changes are controlled by the equipment, and the strains in the materials, samples, adhesives and optical fiber sensors are analyzed objectively by a visual method. The experimental results show that the KALFOS method designed by the researchers is a non-contact method which can measure displacement and strain in all directions. Compared with traditional strain gauges, the KALFOS method can analyze their strain transfer mechanism, and can easily observe the relative motion between sensors and samples. The calibration device helps to validate the definition and quantitative description of component requirements in new or existing standards.
The International Electrotechnical Commission (IEC) adopted a new standard, IEC 61757-1-1:2016, which specifies the most relevant characteristics and characteristics of strain sensors based on FBG and its measuring procedures [154]. The fabrication of FBG sensor packages, the properties of materials and bonded joints can lead to the non-linear behavior of an FBG sensor, which can affect the accuracy of strain measurement. Lazarev et al. designed a nondestructive calibration method for an FBG strain sensor using a mechanical nano-motion sensor [155]. Researchers designed a customized calibration device, which uses a dovetail slideway mechanized by stepping motors. Through the analysis of experimental data, the performance of an FBG strain sensor is studied, the calibration curve of an FBG strain sensor is given, and the feasibility of this method is demonstrated. Święch proposed a calibration measurement system of wing load based on resistance and FBG strain gauge [156]. The data shows that the measurement error is between 12% and 40% according to the different measurement sections. A system of instruments embedded in the wing beam cover can achieve sufficient accuracy for measurement to evaluate the performance of the structure during operation. Święch proposed that because of the heaviness of the interrogator, the FBG sensor has more advantages than the electric strain gauges if a large number of FBG sensors are used. The method proves the feasibility of the calibration method based on FBG sensor, which does not simulate the test effect in a dynamic environment. Whether this method is suitable for the test of flight environment needs further verification.
NASA’s strain gauge calibration technology, the four-point bending calibration device under the German VDI/VDE standard, and the KALFOS non-contact calibration system have achieved differences in calibration principles, equipment, and accuracy for different sensors. FBG sensors based on temperature strain coupling require dual-grating compensation to achieve high calibration accuracy. Traditional strain gauges are influenced by VDI/VDE standards and are susceptible to external electromagnetic fields. Piezoresistive sensors are limited by material dispersion. There has been some research on the calibration methods for the above sensors, but there is a lack of horizontal comparative studies on different calibration techniques for sensors. In order to visually observe the calibration characteristics of various sensors, as shown in Table 5, the calibration methods and accuracy of FBG sensors, traditional strain gauges, and piezoresistive sensors (CNT/CFRP) were compared from the basic principles and methods of calibration.
Based on the review and analysis in this chapter, it can be concluded that onboard calibration is suitable for FBG sensors installed on aircraft wings, and can be completed in situ without disassembly, which is in line with the characteristics of aircraft wings that cannot be easily disassembled. However, due to the interference of flight environment vibration and wide temperature variation, the achievable measurement uncertainty is ±2–5 με. Temperature compensation requires the use of a dual grating layout, which can counteract the impact of temperature fluctuations on the wavelength of the wing surface at −40~85 °C in real time, and optimize the algorithm according to the “Dynamic Environment Compensation” clause in IEC61757-1-1 [171,172]. The laboratory bench calibration based on KALFOS equipment for four-point bending can accurately reproduce the typical force state of the wing, with good environmental controllability and low uncertainty. In addition, a constant temperature box can be used to precisely control the temperature, and temperature compensation can also be achieved through “single-point calibration + temperature coefficient correction”. The calibration process can be carried out according to the “static calibration process” of IEC61757-1-1, and the temperature compensation accuracy should be ensured to meet the requirements of mechanical property matching of materials such as CFRP.
Before calibration, it is necessary to confirm the packaging form of the FBG sensor and prepare a standard strain source according to of IEC61757-1-1. When calibrating the test bench, KALFOS equipment is used to simulate the 0–1000 με bending strain of the wing, record the wavelength strain curve, calculate the sensitivity coefficient, and synchronously test the compensation effect at temperatures of −50~150 °C. On time, the pilot selects non-critical force-bearing areas on the wing (such as wing tip auxiliary structures) as calibration reference points, uses dual grating real-time compensation, and compares the calibration data of the same period test bench to correct deviations. After calibration, according to IEC61757-1-1, “Calibration Effectiveness Determination”, ensure that the deviation of three consecutive measurements is ≤5%, and the wavelength drift after temperature compensation is 0.5 pm/°C, before it can be used for wing deformation monitoring.

6. Conclusions

On the basis of summarizing and analyzing the use of fiber optic sensing technology in aircraft wing deformation monitoring, this article provides a comprehensive introduction to the packaging form, bonding technology, and calibration methods of FBG sensors, and draws the following conclusions. In terms of packaging adaptability, tube-packaged FBGs can play a good role in vibration reduction and have excellent results in the long-term monitoring of the main load-bearing structure of the wing, but their maintenance is difficult. The integration of embedded FBG and CFRP composite wing materials makes internal strain measurement irreplaceable, with a strain transfer efficiency of 90% to 98%. Surface-attached FBG is the best choice for rapid retrofitting of existing wings due to its ability to achieve strain transfer efficiency of up to 99.2% by connecting the structure through metal welding. In terms of bonding technology, the thickness, width, and length parameters of epoxy adhesive directly affect the strain transfer efficiency. Its shear modulus is 3.2 GPa, which can effectively reduce the occurrence of strain hysteresis and has good stability. In terms of standardization of calibration technology, the four-point bending calibration device (VDI/VDE2660) based on the IEC61757-1-1 standard can achieve a static calibration error of <1 με. After calibration using the KALFOS non-contact calibration system, dynamic strain measurements with a relative standard deviation of ≤0.16% can be obtained. The use of dual-grating temperature compensation can offset the temperature fluctuations of −40–85 °C on the wing surface.
Based on the literature review and analysis mentioned above, this article suggests addressing the practical needs of aviation engineering. The CFRP composite wing should first be embedded with FBG with a diameter of 125 μm or less in the prepreg, and then embedded in the grating during the manufacturing process of the laminated board to match the thermal expansion coefficient of the grating with the composite material. It is recommended to use surface-attached FBG for aluminum alloy wings, and use ultrasonic welding technology to firmly weld the sensors on the wing skin with a thickness greater than or equal to 1 mm. The welding point diameter should be greater than or equal to 3 mm to ensure good long-term reliability. Sapphire fiber EFPI sensors can be used in the engine nacelle area, and temperature and pressure can be detected simultaneously through dual chambers. In terms of optimizing the bonding process, it is recommended to use a dispensing process to ensure the uniformity of the adhesive layer when using epoxy adhesive. In the surface treatment of bonding, the wing surface needs to be sandblasted first, and then cured under different temperature conditions to remove internal stress. During static calibration, a temperature strain coupling model can be established using a four-point bending device (VDI/VDE2660) produced in Germany. During dynamic calibration, the KALFOS system can be used to measure the sinusoidal strain excitation applied to the sensor, in order to verify its dynamic response characteristics and accuracy. During on-site calibration, single-point calibration should be conducted within 6 months, and the calibration work should be based on the laser interferometer.
The application of fiber optic sensing technology in the aviation field still needs continuous breakthroughs in the following areas. Due to the harsh flight environment of aircraft, it is recommended to develop new fiber optic materials that are resistant to high temperatures (>1500 °C) and radiation. It is recommended to combine fiber optic sensors with wireless transmission technology to achieve real-time data transmission and remote diagnosis. Sensor arrays for simultaneous measurement of multiple parameters should be researched to improve the integration of monitoring systems.

Author Contributions

Conceptualization, Z.M. and X.W.; methodology, X.C. and B.C.; software, Z.M.; validation, X.W. and Z.M.; formal analysis, Z.M. and B.C.; investigation, Z.M.; resources, Z.M.; data curation, Z.M.; writing—original draft preparation, Z.M. and B.C.; writing—review and editing, Z.M., and X.W.; visualization, Z.M.; supervision, Z.M.; project administration, Z.M.; funding acquisition, X.C. and B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was jointly funded by the Jiangsu Province and Education Ministry Co-sponsored Synergistic Innovation Center of Modern Agricultural Equipment, grant number XTCX2009, the Open Funding from the Key Laboratory of Modern Agricultural Equipment and Technology (Jiangsu University), Ministry of Education, grant number MAET202301, and the National Natural Science Foundation of China, grant number No. 61873064.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, Y. Recent progress in aircraft smart skin for structural health monitoring. Struct. Health Monit. 2022, 21, 2453–2480. [Google Scholar] [CrossRef]
  2. Krishnamoorthy, R.R.; Marius, D. Nondestructive inspection and structural health monitoring of aerospace materials. Aerosp. Mater. 2025, 159–206. [Google Scholar] [CrossRef]
  3. Giannakeas, I.N.; Khodaei, Z.S.; Aliabadi, M.H. Structural health monitoring cost estimation of a piezosensorized aircraft fuselage. Sensors 2022, 22, 1771. [Google Scholar] [CrossRef] [PubMed]
  4. Davis, C.; Tejedor, S.; Grabovac, I. High-strain fiber bragg gratings for structural fatigue testing of military aircraft. Photonic Sens. 2012, 2, 215–224. [Google Scholar]
  5. Miguel, G.C.; Zúñiga, S.J.; Sánchez, G.J. Demonstration and methodology of structural monitoring of stringer runs out composite areas by embedded optical fiber sensors and connectors integrated during production in a composite plant. Sensors 2017, 17, 1683. [Google Scholar] [CrossRef] [PubMed]
  6. Ma, Z.; Chen, X. Fiber Bragg gratings sensors for aircraft wing shape measurement: Recent applications and technical analysis. Sensors 2019, 19, 55. [Google Scholar] [CrossRef]
  7. DAYY Photonics. Fiber Bragg Grating Sensor (FBG); DAYY Photonics. Available online: https://dayyphotonics.com/applications/fiber-bragg-grating-sensor-fbg (accessed on 20 July 2025).
  8. Pena, F.; Richards, L.; Parker, A.R., Jr.; Piazza, A.; Chan, P.; Hamory, P. Fiber Optic Sensing System (FOSS) Technology—A New Sensor Paradigm for Comprehensive Structural Monitoring and Model Validation Throughout the Vehicle Life-Cycle. 2015. Available online: https://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/20180007391.pdf (accessed on 20 July 2025).
  9. Pena, F.; Martins, B.L.; Richards, W.L. Active In-flight Load Redistribution Utilizing Fiber-Optic Shape Sensing and Multiple Control Surfaces; NASA/TM—2018–219741; NASA: Washington, DC, USA, 2018.
  10. Martins, B.L.; Kosmatka, J.B. Health Monitoring of Aerospace Structures via Dynamic Strain Measurements: An Experimental Demonstration. In Proceedings of the AIAA Scitech 2020 Forum, Orlando, FL, USA, 6–10 January 2020; p. 0701. [Google Scholar]
  11. Kim, J.H.; Lee, Y.G.; Park, Y. Temperature-compensated strain measurement of full-scale small aircraft wing structure using low-cost FBG interrogator. Sensors and Smart Structures Technologies for Civil, Mechanical, and Aerospace Systems, Spie Smart Structures and Materials + Nondestructive Evaluation And Health Monitoring, San Diego, CA, USA, 10–14 March 2013. Volume 8692, p. 86922P. Available online: https://www.spiedigitallibrary.org/conference-proceedings-of-spie/8692/86922P/Temperature-compensated-strain-measurement-of-full-scale-small-aircraft-wing/10.1117/12.2011720.short?webSyncID=d49a2853-1ab0-d3a6-b03c-97c009b6bcd7&sessionGUID=46b80d8e-8abe-eaed-6ae0-5c117aac5a91 (accessed on 30 July 2025).
  12. He, Y.; Dong, M.; Sun, G. Shape monitoring of morphing wing using micro optical sensors with different embedded depth. Opt. Fiber Technol. 2019, 48, 179–185. [Google Scholar] [CrossRef]
  13. Ciminello, M.; Sikorski, B.; Galasso, B.; Pellone, L.; Mercurio, U.; Concilio, A.; Apuleo, G.; Cozzolino, A.; Kressel, I.; Shoham, S.; et al. Preliminary Results of a Structural Health Monitoring System Application for Real-Time Debonding Detection on a Full-Scale Composite Spar. Sensors 2023, 23, 455. [Google Scholar] [CrossRef] [PubMed]
  14. Pollock, L.; Wild, G. An Exploration of Structural Health Monitoring for Hypersonic Vehicles. In AIAA Aviation Forum and ASCEND Co-Located Conference Proceedings, Proceedings of the AIAA AVIATION Forum and ASCEND 2024, Las Vegas, NV, USA, 29 July–2 August 2024; American Institute of Aeronautics and Astronautics: Reston, VA, USA, 2024. [Google Scholar]
  15. Apuleo, G.; Ciminello, M.; Pellone, L.; Mercurio, U.; Concilio, A. Application of a Fiber Optic-Based SHM System to a Composite Aircraft Wing and Its Technological Maturity Evaluation. Eng. Proc. 2025, 90, 31. [Google Scholar]
  16. Krenz, A.; Koch, J.; Reimer, V.; Doering, A.; Guehlke, P.; Waltermann, C. Methods for FBG Sensor Integration for RTM Process Monitoring and SHM of the Final CFRP Component. In Proceedings of the 10th ECCOMAS Thematic Conference on Smart Structures and Materials, Patras, Greece, 3–5 July 2023; pp. 1839–1850. [Google Scholar]
  17. Pollock, L.; Kleine, H.; Neely, A.; Wild, G. Optical Fiber Bragg Grating-Based Measurement of Fluid-Structure Interaction on a Cantilever Panel in High-Speed Flow. IEEE Access 2024, 12, 101106–101120. [Google Scholar] [CrossRef]
  18. Ankita, R.; Ghorai, S.K.; Sengupta, S. Vehicle Flow Indication and Identification Using FBG Sensors. Phys. Scr. 2024, 99, 125543. [Google Scholar] [CrossRef]
  19. Aimasso, A.; Dalla Vedova, M.D.L.; Bertone, M.; Maggiore, P. Preliminary Design and Performance Evaluation of Optical Fiber-Based Load Sensor for Aerospace Systems. J. Phys. Conf. Ser. 2024, 2802, 012010. [Google Scholar] [CrossRef]
  20. Reimer, T.; Di Martino, G.; Petkov, I.; Dauth, L.; Baier, L.; Gülhan, A. Design, Manufacturing and Assembly of the STORT Hypersonic Flight Experiment Thermal Protection System. In Proceedings of the 25th AIAA International Space Planes and Hypersonic Systems and Technologies Conference, Bengaluru, India, 28 May 2023–1 June 2024; p. 3089. [Google Scholar]
  21. Parnika, G.; Amit, T.; Xiuyun, H.; Kamil, G.; Kafil, M.R.; Padraic, E.M.; Peter, O.B. Substrate integrated micro-thermoelectric coolers in glass substrate for next-generation photonic packages. J. Opt. Microsyst. 2024, 4, 011006. [Google Scholar]
  22. Lv, Z.; Zhu, B.; Lu, L.; Yuan, P.; Lou, X.; Dong, M.; Zhu, L. Miniature and low-power high-precision FBG interrogator with self-temperature compensation. Opt. Express. 2025, 33, 10289–10301. [Google Scholar]
  23. Luo, X.; Hu, R.; Xie, B. Thermal Management for Opto-Electronics Packaging and Applications; Wiley: Hoboken, NJ, USA, 2024. [Google Scholar]
  24. Rees, A.V.; Tsong, W.; Vlekkert, I.V.D.; Farjana, F.; Lammerink, R.E.M.; Visscher, I.; Roeloffzen, C.G.H.; Musa, S.; Geskus, D. High-Power 150 mW Extended Cavity Si3N4 Tunable Narrow-Linewidth Laser. In Optical Interconnects XXIV; SPIE: Bellingham, WA, USA, 2024; p. 128920J. [Google Scholar]
  25. Zhang, Z.; Huang, B.; Zhang, Z.; Chen, H. On-Chip Reconstructive Spectrometer Based on Parallel Cascaded Micro-Ring Resonators. Appl. Sci. 2024, 14, 4886. [Google Scholar]
  26. Yi, D.; Luan, J.; Wang, Y.; Tsang, H.K. Reconfigurable polarization processor based on coherent four-port micro-ring resonator. Nanophotonics 2023, 12, 4127–4136. [Google Scholar]
  27. Amanzadeh, M.; Aminossadati, S.M.; Kizil, M. Recent developments in fibre optic shape sensing. Measurement 2018, 128, 119–137. [Google Scholar] [CrossRef]
  28. Hill, K.O.; Fujii, Y.; Johnson, D.C. Photosensitivity in optical fiber waveguides: Application to reflection filter fabrication. Appl. Phys. Lett. 1978, 32, 647–649. [Google Scholar] [CrossRef]
  29. Dunphy, J.R.; Meltz, G. Optical Fiber Sensors for Composite Structures. In Optical Fiber Sensors; Optica Publishing Group: Washington, DC, USA, 1988; FCC1. [Google Scholar] [CrossRef]
  30. Malo, B.; Hill, K.O.; Bilodeau, F. Point-by-point fabrication of micro-Bragg gratings in photosensitive fibre using single excimer pulse refractive index modification techniques. Electron. Lett. 1993, 29, 1668–1669. [Google Scholar] [CrossRef]
  31. Hong, C.Y.; Zhang, Y.F.; Zhang, M.X. Application of FBG sensors for geotechnical health monitoring, a review of sensor design, implementation methods and packaging techniques. Sens. Actuators A Phys. 2016, 244, 184–197. [Google Scholar] [CrossRef]
  32. Zhou, T.; Chen, M.Y.; Hu, Z.Z.; Gong, T.Y.; Liang, Y. Research on ultra-wide joint bending angle detection based on transparent tubing structure. IEEE Sens. J. 2023, 24, 1511–1519. [Google Scholar] [CrossRef]
  33. Liang, Y.; Liang, H.; Kong, L.; Li, H.; Dai, Z.; Chen, M. Thin-diameter photonic crystal fiber magnetic field sensor based on controlled hole collapse and infiltration. IEEE Sens. J. 2023, 23, 27337–27342. [Google Scholar] [CrossRef]
  34. Liang, Y. Design of a Mach-Zehnder Interferometric Fiber Sensing System Based on PCA-Assisted OAM Interrogation for Simultaneous Measurement of Refractive Index and Temperature. J. Light. Technol. 2022, 40, 6310–6316. [Google Scholar] [CrossRef]
  35. Barnoski, M.K.; Rourke, M.D.; Jensen, S.M. Optical time domain reflectometer. Appl. Opt. 1977, 16, 2375–2379. [Google Scholar] [CrossRef]
  36. Eickhoff, W.; Ulrich, R. Optical frequency domain reflectometry in single-mode fiber. Appl. Phys. Lett. 1981, 39, 693–695. [Google Scholar] [CrossRef]
  37. Ippen, E.P.; Stolen, R.H. Stimulated Brillouin scattering in optical fibers. Appl. Phys. Lett. 1972, 21, 539–541. [Google Scholar] [CrossRef]
  38. Horiguchi, T.; Tateda, M. BOTDA-nondestructive measurement of single-mode optical fiber attenuation characteristics using Brillouin interaction: Theory. J. Light. Technol. 1989, 7, 1170–1176. [Google Scholar] [CrossRef]
  39. Shi, B.; Xu, X.J.; Wang, D. Study on BOTDR-based distributed optical fiber strain measurement for tunnel health diagnosis. Chin. J. Rock. Mech. Eng. 2005, 24, 2622–2628. [Google Scholar]
  40. Li, X.J.; He, F.; Wang, Y.B. Tunnel structure safety monitoring system based on BOTDA technology. Appl. Mech. Mater. 2014, 621, 281–284. [Google Scholar] [CrossRef]
  41. Li, W.; Bao, X.; Li, Y. Differential pulse-width pair BOTDA for high spatial resolution sensing. Opt. Express 2008, 16, 21616–21625. [Google Scholar] [CrossRef] [PubMed]
  42. Shimizu, K.; Horiguchi, T.; Koyamada, Y. Coherent self-heterodyne detection of spontaneously Brillouin-scattered light waves in a single-mode fiber. Opt. Lett. 1993, 18, 185–187. [Google Scholar] [CrossRef]
  43. Rogers, A.J. Polarization-optical time domain reflectometry: A technique for the measurement of field distributions. Appl. Opt. 1981, 20, 1060–1074. [Google Scholar] [CrossRef] [PubMed]
  44. Ramirez-Hernandez, M.A.; Alonso-Murias, M.; Monzon-Hernandez, D. Polymer-Capped Fiber Fabry-Perot Interferometer for Large Range Displacement Sensing. J. Light. Technol. 2024, 42, 3430–3437. [Google Scholar] [CrossRef]
  45. Siwowski, T.; Rajchel, M.; Howiacki, T.; Sieńko, R.; Bednarski, Ł. Distributed fibre optic sensors in FRP composite bridge monitoring: Validation through proof load tests. Eng. Struct. 2021, 246, 113057. [Google Scholar] [CrossRef]
  46. Zhan, Y.; Han, M.; Wang, Z.; Xu, L.; Song, Z.; Lu, A.; Guo, X.; Deng, W.; Huang, S. Distributed strain monitoring for different composites structures with high resolution based on optical fiber sensing. Optik 2021, 248, 168113. [Google Scholar] [CrossRef]
  47. da Silva Falcão, B.; Giwelli, A.; Kiewiet, M.N.; Banks, S.; Yabesh, G.; Esteban, L.; Kiewiet, L.; Yekeen, N.; Kovalyshen, Y.; Monmusson, L.; et al. Strain measurement with multiplexed FBG sensor arrays: An experimental investigation. Heliyon 2023, 9, e18652. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Jiang, Y.; Yang, S.; Zhang, D. All-sapphire fiber-optic sensor for the simultaneous measurement of ultra-high temperature and high pressure. Opt. Express 2024, 32, 14826–14836. [Google Scholar]
  49. Wang, F.; Zhan, W.; Zhang, X.; Lu, Y. Improvement of spatial resolution for BOTDR by iterative subdivision method. J. Light. Technol. 2013, 31, 3663–3667. [Google Scholar] [CrossRef]
  50. Almoosa, A.S.K.; Hamzah, A.E.; Zan, M.S.D.; Ibrahim, M.F.; Arsad, N.; Elgaud, M.M. Improving the Brillouin frequency shift measurement resolution in the Brillouin optical time domain reflectometry (BOTDR) fiber sensor by artificial neural network (ANN). Opt. Fiber Technol. 2022, 70, 102860. [Google Scholar] [CrossRef]
  51. Rocha, H.; Semprimoschnig, C.; Nunes, J.P. Small-diameter optical fibre sensor embedment for ambient temperature cure monitoring and residual strain evaluation of CFRP composite laminates produced by vacuum-assisted resin infusion. CEAS Space J. 2021, 13, 353–367. [Google Scholar] [CrossRef]
  52. Tserpes, K.I.; Karachalios, V.; Giannopoulos, I.; Prentzias, V.; Ruzek, R. Strain and damage monitoring in CFRP fuselage panels using fiber Bragg grating sensors. Part I: Design, manufacturing and impact testing. Compos. Struct. 2014, 107, 726–736. [Google Scholar] [CrossRef]
  53. Mao, X.; Zhou, X.; Ye, H.; Tan, Y.; Luo, Y. Fast interrogation of dynamic fiber Bragg grating using neighborhood average algorithm. Infrared Phys. Technol. 2023, 128, 104490. [Google Scholar] [CrossRef]
  54. Yi, X.; Chen, X.; Fan, H.; Shi, F.; Cheng, X.; Qian, J. Separation method of bending and torsion in shape sensing based on FBG sensors array. Opt. Express 2020, 28, 9367–9383. [Google Scholar] [CrossRef]
  55. Gong, Y.; Michael, O.L.C.; Hao, J.; Paulose, V. Extension of sensing distance in a ROTDR with an optimized fiber. Opt. Commun. 2007, 280, 91–94. [Google Scholar] [CrossRef]
  56. Yin, J.; Liu, T.; Jiang, J.; Liu, K.; Wang, S.; Wu, F.; Ding, Z. Wavelength-division-multiplexing method of polarized low-coherence interferometry for fiber Fabry–Perot interferometric sensors. Opt. Lett. 2013, 38, 3751–3753. [Google Scholar] [CrossRef]
  57. Kahandawa, G.C.; Epaarachchi, J.; Wang, H.; Lau, K.T. Use of FBG sensors for SHM in aerospace structures. Photonic Sens. 2012, 2, 203–214. [Google Scholar] [CrossRef]
  58. Cusano, A.; Antonello, C.; Jacques, A. Fiber Bragg Grating Sensors: Recent Advancements, Industrial Applications and Market Exploitation; Bentham Science Publishers: Sharjah, United Arab Emirates, 2011. [Google Scholar]
  59. Yang, X. A novel packaging method for FBG temperature sensors based on ultrasonic-assisted soldering technology. Sens. Actuators A Phys. 2024, 373, 115381. [Google Scholar] [CrossRef]
  60. Marco, B. Fiber-bragg-grating-based displacement sensors: Review of recent advances. Materials 2022, 15, 5561. [Google Scholar]
  61. Khan, R.Y.M.; Rahim, U.; Muhammad, F. Design and development of cost-effective fiber Bragg grating temperature sensor package. Meas. Sci. Technol. 2023, 34, 085122. [Google Scholar] [CrossRef]
  62. Huang, L.; Xiong, L.; Guo, Y.; Hu, Z. All-Silicon Packaging Technology for Fiber Bragg Gratings and Its Application in Tilt Sensor. IEEE Sens. J. 2022, 22, 22627–22636. [Google Scholar] [CrossRef]
  63. Grabovac, I.; Nuyens, T.; Davis, C. Packaging and Mounting of In-Fibre Bragg Grating Arrays for Structural Health Monitoring of Large Structures. Defense Science and Technology Organization Victoria (Australia) Maritime Platforms div. 2010. Available online: https://archive.org/details/DTIC_ADA536063/page/30/mode/2up?q=%22richard++muscat%22 (accessed on 20 July 2025).
  64. Floris, I.; Madrigal, J.; Sales, S. Experimental study of the influence of FBG length on optical shape sensor performance. Opt. Lasers Eng. 2020, 126, 105878. [Google Scholar] [CrossRef]
  65. Li, Y.; Wen, C.; Sun, Y. Capillary encapsulating of fiber Bragg grating and the associated sensing model. Opt. Commun. 2014, 333, 92–98. [Google Scholar] [CrossRef]
  66. Jinachandran, S.; Li, H.; Xi, J. Fabrication and characterization of a magnetized metal-encapsulated FBG sensor for structural health monitoring. IEEE Sens. J. 2018, 18, 8739–8746. [Google Scholar] [CrossRef]
  67. Zhang, Y.; Meng, F.; Song, Y. Characterization of metallic-packaging fiber Bragg grating sensors with coated and bare fibers. Appl. Opt. 2019, 58, 1426–1432. [Google Scholar] [CrossRef]
  68. Di Palma, P. Force sensor based on FBG embedded in silicone rubber. IEEE Sens. J. 2022, 23, 1172–1178. [Google Scholar] [CrossRef]
  69. Liao, M.H. The cooperative deformation test of an embedded FBG sensor and strain correction curve verification. Constr. Build. Mater. 2022, 342, 128029. [Google Scholar] [CrossRef]
  70. Zhao, J.R. Embedded Fiber Bragg Grating (FBG) Sensors Fabricated by Ultrasonic Additive Manufacturing for High-Frequency Dynamic Strain Measurements. IEEE Sens. J. 2023, 24, 2853–2862. [Google Scholar] [CrossRef]
  71. Ana, F.M. Embedded fiber Bragg grating sensors for monitoring temperature and thermo-elastic deformations in a carbon fiber optical bench. Sensors 2023, 23, 6499. [Google Scholar] [CrossRef] [PubMed]
  72. Qiang, B. Temperature and external strain sensing with metal-embedded optical fiber sensors for structural health monitoring. Opt. Express 2022, 30, 33449–33464. [Google Scholar] [CrossRef] [PubMed]
  73. Ramakrishnan, M.; Rajan, G.; Semenova, Y. Overview of fiber optic sensor technologies for strain/temperature sensing applications in composite materials. Sensors 2016, 16, 99. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, H.P.; Wu, Y.B.; Chen, C.; Zhang, H.Y.; Jiang, H. Dynamic response of CFRP reinforced steel beams subjected to impact action based on FBG sensing technology. Sensors 2022, 22, 6377. [Google Scholar] [CrossRef] [PubMed]
  75. Shafighfard, T.; Magdalena, M. Experimental and numerical study of the additively manufactured carbon fibre reinforced polymers including fibre Bragg grating sensors. Compos. Struct. 2022, 299, 116027. [Google Scholar] [CrossRef]
  76. Fatemeh, M. Methods for embedding fiber Bragg grating sensors during material extrusion: Relationship between the interfacial bonding and strain transfer. Addit. Manuf. 2023, 68, 103497. [Google Scholar] [CrossRef]
  77. Grundmann, N.; Brüning, H.; Tserpes, K.; Strohbach, T.; Mayer, B. Influence of Embedding Fiber Optical Sensors in CFRP Film Adhesive Joints on Bond Strength. Sensors 2020, 20, 1665. [Google Scholar] [CrossRef]
  78. Murayama, H.; Kageyama, K.; Kamita, T. Structural health monitoring of a full-scale composite structure with fiber-optic sensors. Adv. Compos. Mater. 2002, 11, 287–297. [Google Scholar] [CrossRef]
  79. Baker, W.; McKenzie, I.; Jones, R. Development of life extension strategies for Australian military aircraft, using structural health monitoring of composite repairs and joints. Compos. Struct. 2004, 66, 133–143. [Google Scholar] [CrossRef]
  80. Botsis, J.; Humbert, L.; Colpo, F. Embedded fiber Bragg grating sensor for internal strain measurements in polymeric materials. Opt. Lasers Eng. 2005, 43, 491–510. [Google Scholar] [CrossRef]
  81. Li, H.N.; Zhou, G.D.; Liang, R. Strain transfer analysis of embedded fiber Bragg grating sensor under nonaxial stress. Opt. Eng. 2007, 46, 054402. [Google Scholar] [CrossRef]
  82. Takeda, N.; Minakuchi, S. Smart aircraft composite structures with embedded small-diameter optical fiber sensors. In Proceedings of the Third Asia Pacific Optical Sensors Conference, Sydney, Australia, 31 January–3 February 2012; Volume 8351, p. 83510I. [Google Scholar]
  83. Fang, L.; Chen, T.; Li, R. Application of embedded fiber Bragg grating (FBG) sensors in monitoring health to 3D printing structures. IEEE Sens. J. 2016, 16, 6604–6610. [Google Scholar] [CrossRef]
  84. Freydin, M.; Rattner, M.; Raveh, D.E. Strain-Based Aeroelastic Shape Sensing. In Proceedings of the 2018 AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Kissimmee, FL, USA, 8–12 January 2018; p. 0190. [Google Scholar]
  85. Kwon, H.; Park, Y.; Kim, J.H. Embedded fiber Bragg grating sensor–based wing load monitoring system for composite aircraft. Struct. Health Monit. 2019, 18, 1337–1351. [Google Scholar] [CrossRef]
  86. He, Y.; Zhu, L.; Sun, G.; Meng, F.; Song, Y. Fiber Brag grating monitoring of a morphing wing based on a polyvinyl chloride reinforced silicone substrate. Opt. Fiber Technol. 2019, 50, 145–153. [Google Scholar] [CrossRef]
  87. Qu, X.; Li, J.; Shan, Y.; Yang, Z.; Yang, L.; Xu, H.; Zhao, S. Various static loading condition monitoring of carbon fiber composite cylinder with integrated optical fiber sensors. Opt. Fiber Technol. 2024, 83, 103685. [Google Scholar] [CrossRef]
  88. Waters, D.H.; Kumosa, M.; Hoffman, J. Evaluating strain limits of embedded FBG sensors in pultruded hybrid composite rods. IEEE Trans. Power Deliv. 2022, 37, 5025–5032. [Google Scholar] [CrossRef]
  89. Li, Y.; Wen, C.; Zhang, H. An electroplating method for surface mounting optical fiber sensors on the metal substrate. IEEE Photonics Technol. Lett. 2016, 28, 1811–1814. [Google Scholar] [CrossRef]
  90. Wang, H.P.; Gong, X.S.; Wang, X.Z.; Feng, S.Y.; Yang, T.L.; Guo, Y.X. Discrete curvature-based shape configuration of composite pipes for local buckling detection based on fiber Bragg grating sensors. Measurement 2022, 188, 110603. [Google Scholar] [CrossRef]
  91. Lee, G.; Hossain, O.; Jamalzadegan, S.; Liu, Y.; Wang, H.; Saville, A.C. Abaxial leaf surface-mounted multimodal wearable sensor for continuous plant physiology monitoring. Sci. Adv. 2023, 9, eade2232. [Google Scholar] [CrossRef]
  92. Kaya, B.M.; Esenturk, O.; Asici, C.; Sarac, U.; Dindis, G.; Baykul, M.C. Coating effects of a strain sensor on durability and sensitivity using the fiber loop ringdown spectroscopy technique. Physica Scripta. 2024, 99, 055511. [Google Scholar] [CrossRef]
  93. Hopf, B.; Koch, A.W.; Roths, J. Temperature dependence of glue-induced birefringence in surface-attached FBG strain sensors. J. Light. Technol. 2015, 34, 1220–1227. [Google Scholar] [CrossRef]
  94. Sirkis, J.S.; Haslach, H.W. Interferometric stain measurement by arbitrarily configured surface-mounted, optical fibers. J. Light. Technol. 1990, 8, 1497–1503. [Google Scholar] [CrossRef]
  95. Duncan, R.; Gifford, D.; Childers, B. A Distributed Sensing Technique for Aerospace Applications. In Proceedings of the 42nd AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 5–8 January 2004; p. 544. [Google Scholar]
  96. Li, H.; Zhu, L.; Dong, M. Analysis on strain transfer of surface-bonding FBG on Al 7075-T6 alloy host. Optik 2016, 127, 1233–1236. [Google Scholar] [CrossRef]
  97. Goossens, S.; De Pauw, B.; Geernaert, T. Aerospace-grade surface mounted optical fibre strain sensor for structural health monitoring on composite structures evaluated against in-flight conditions. Smart Mater. Struct. 2019, 28, 065008. [Google Scholar] [CrossRef]
  98. Sodja, J.; De Breuker, R.; Meddaikar, Y.M. Ground Testing of the FLEXOP Demonstrator Aircraft. In Proceedings of the AIAA Scitech 2020 Forum, Orlando, FL, USA, 6–10 January 2020; p. 1968. [Google Scholar]
  99. Boateng, E.K.G.; Schubel, P.; Umer, R. Thermal isolation of FBG optical fibre sensors for composite cure monitoring. Sens. Actuators A Phys. 2019, 287, 158–167. [Google Scholar] [CrossRef]
  100. Guo, Z.S. Strain and temperature monitoring of asymmetric composite laminate using FBG hybrid sensors. Struct. Health Monitor. 2007, 6, 191–197. [Google Scholar] [CrossRef]
  101. Fernández, R.; Gutiérrez, N.; Jiménez, H.; Martín, F.; Rubio, L.; Jiménez-Vicaria, J.D.; Paulotto, C.; Lasagni, F. On the structural testing monitoring of CFRP cockpit and concrete/CFRP pillar by FBG sensors. Adv. Eng. Mater. 2016, 18, 1289–1298. [Google Scholar] [CrossRef]
  102. Xu, L.; Shi, S.; Yan, F.; Huang, Y.; Bao, Y. Experimental study on combined effect of mechanical loads and corrosion using tube-packaged long-gauge fiber Bragg grating sensors. Struct. Health Monit. 2023, 22, 3985–4004. [Google Scholar] [CrossRef]
  103. Kosaka, T.; Kurimoto, H.; Osaka, K.; Nakai, A.; Osada, T.; Hamada, H.; Fukuda, T. Strain monitoring of braided composites by using embedded fiber-optic strain sensors. Adv. Compos. Mater. 2004, 13, 157–170. [Google Scholar] [CrossRef]
  104. Jang, B.-W.; Park, S.-O.; Lee, Y.-G.; Kim, C.-G.; Park, C.-Y. Detection of impact damage in composite structures using high speed FBG interrogator. Adv. Compos. Mater. 2012, 21, 29–44. [Google Scholar] [CrossRef]
  105. Yi, L.; Zhenkai, W.; Jialu, L. Research on health monitor for three-dimensional braided composite material. In Proceedings of the 2011 Third International Conference on Measuring Technology and Mechatronics Automation, Shanghai, China, 6–7 January 2011; pp. 1150–1153. [Google Scholar]
  106. Kim, S.-W.; Kim, E.-H.; Jeong, M.-S.; Lee, I. Damage evaluation and strain monitoring for composite cylinders using tin-coated FBG sensors under low-velocity impacts. Compos. B Eng. 2015, 74, 13–22. [Google Scholar] [CrossRef]
  107. Frazão, O.; Oliveira, R.; Dias, I. A simple smart composite using fiber Bragg grating sensors for strain and temperature discrimination. Microw. Opt. Technol. Lett. 2008, 51, 235–239. [Google Scholar] [CrossRef]
  108. Rao, Y.J.; Yuan, S.F.; Zeng, X.K.; Lian, D.K.; Zhu, Y.; Wang, Y.P.; Huang, S.L.; Liu, T.Y.; Fernando, G.F.; Zhang, L.; et al. Simultaneous strain and temperature measurement of advanced 3-D braided composite materials using an improved EFPI/FBG system. Opt. Lasers Eng. 2002, 38, 557–566. [Google Scholar] [CrossRef]
  109. Zhao, X.; Jin, K.; Yan, M.; Nan, P.; Zhou, F.; Xin, G.; Lim, K.-S.; Ahmad, H.; Zhang, Y.; Yang, H. Inverse heat transfer for real-time thermal evaluation of aircraft thermal protection structure with embedded FBG sensors. Appl. Therm. Eng. 2025, 260, 124869. [Google Scholar] [CrossRef]
  110. Takeda, S.I.; Mizutani, T.; Nishi, T.; Uota, N.; Hirano, Y.; Iwahori, Y.; Nagao, Y.; Takeda, N. Monitoring of a CFRP-stiffened panel manufactured by VaRTM using fiber-optic sensors. Adv. Compos. Mater. 2008, 17, 125–137. [Google Scholar] [CrossRef]
  111. Mizutani, T.; Takeda, N.; Takeya, H. On-board strain measurement of a cryogenic composite tank mounted on a reusable rocket using FBG sensors. Struct. Health Monitor. 2006, 5, 205–214. [Google Scholar] [CrossRef]
  112. Basu, M.; Ghorai, S.K. Strain sensing in fiber-reinforced polymer laminates using embedded fiber Bragg grating sensor. Fiber Integr. Opt. 2014, 33, 279–298. [Google Scholar] [CrossRef]
  113. Sun, L.; Wang, J.; Hu, H.; Fu, C.; Wang, C.; Ni, A. A real-time NDT method based on FBG sensor spectrums to detect the transverse matrix cracks in composite materials. J. Nondestruct. Eval. 2020, 39, 44. [Google Scholar] [CrossRef]
  114. Pran, K.; Farsund, O.; Wang, G. Fibre Bragg grating smart bolt monitoring creep in bolted GRP composite. In Proceedings of the 15th Optical Fiber Sensors Conference Technical Digest, Portland, OR, USA, 10 May 2002; pp. 431–434. [Google Scholar]
  115. Wei, Z.; Song, D.; Zhao, Q.; Cui, H.-L. High pressure sensor based on fiber Bragg grating and carbon fiber laminated composite. IEEE Sensors J. 2008, 8, 1615–1619. [Google Scholar] [CrossRef]
  116. Liu, Y.; Zhao, X.; Wei, X.; Nan, P.; Zhou, F.; Xin, G.; Lim, K.-S.; Zhang, Y.; Yang, H. Real-Time Sensor for Measuring the Surface Temperature of Thermal Protection Structures Based on the Full-Time Domain Temperature Inversion Method. Sensors 2025, 25, 2227. [Google Scholar] [CrossRef]
  117. Papantoniou, A.; Rigas, G.; Alexopoulos, N.D. Assessment of the strain monitoring reliability of fiber Bragg grating sensor (FBGs) in advanced composite structures. Compos. Struct. 2011, 93, 2163–2172. [Google Scholar] [CrossRef]
  118. Kuang, K.S.C.; Kenny, R.; Whelan, M.P.; Cantwell, W.J.; Chalker, P.R. Embedded fibre Bragg grating sensors in advanced composite materials. Compos. Sci. Technol. 2001, 61, 1379–1387. [Google Scholar] [CrossRef]
  119. Minakuchi, S. In situ characterization of direction-dependent cure-induced shrinkage in thermoset composite laminates with fiber-optic sensors embedded in through-thickness and in-plane directions. J. Compos. Mater. 2015, 49, 1021–1034. [Google Scholar] [CrossRef]
  120. .Rajabzadeh, A.; Heusdens, R.; Hendriks, R.C.; Groves, R.M. Characterisation of transverse matrix cracks in composite materials using fibre Bragg grating sensors. J. Lightw. Technol. 2019, 37, 4720–4727. [Google Scholar] [CrossRef]
  121. Geng, X.; Jiang, M.; Gao, L.; Wang, Q.; Jia, Y.; Sui, Q.; Jia, L.; Li, D. Sensing characteristics of FBG sensor embedded in CFRP laminate. Measurement 2017, 98, 199–204. [Google Scholar] [CrossRef]
  122. Prusty, B.G. Failure investigation of top-hat composite stiffened panels. Ships Offshore Struct. 2013, 9, 186–198. [Google Scholar] [CrossRef]
  123. Yu, K.; Peng, Z.; Zhang, Y.; Zhu, P.; Chen, W.; Hao, J. Pipeline Elbow Corrosion Simulation for Strain Monitoring with Fiber Bragg Gratings. Micromachines 2024, 15, 1098. [Google Scholar] [CrossRef]
  124. Rebelo, F.J.; Oliveira, J.R.; Silva, H.M.; Oliveira e Sá, J. Material selection for efficient strain transfer in pavement monitoring using FBG sensors. Int. J. Pavement Eng. 2024, 25, 2440509. [Google Scholar] [CrossRef]
  125. Cheng, C.; Liu, Y.; Li, X.; Yang, J.; Tong, T.; Ke, Q. The strain transfer characteristics of SAW strain sensors and their impact on the sensor performance. Sens. Actuators A Phys. 2025, 389, 116575. [Google Scholar] [CrossRef]
  126. Wahab, M.M.A.; Ashcroft, I.A.; Crocombe, A.D. Prediction of fatigue thresholds in adhesively bonded joints using damage mechanics and fracture mechanics. J. Adhes. Sci. Technol. 2001, 15, 763–781. [Google Scholar] [CrossRef]
  127. Zhou, G.; Li, H.; Ren, L. Influencing parameters analysis of strain transfer in optic fiber Bragg grating sensors, Advanced Sensor Technologies for Nondestructive Evaluation and Structural Health Monitoring II. Int. Soc. Opt. Photonics 2006, 6179, 61790R. [Google Scholar]
  128. Hocker, G.B. Fiber-optic sensing of pressure and temperature. Appl. Opt. 1979, 18, 1445–1448. [Google Scholar] [CrossRef]
  129. Pak, Y.E. Longitudinal shear transfer in fiber optic sensors. Smart Mater. Struct. 1992, 1, 57. [Google Scholar] [CrossRef]
  130. Yuan, L.; Zhou, L. Sensitivity coefficient evaluation of an embedded fiber-optic strain sensor. Sens. Actuators A Phys. 1998, 69, 5–11. [Google Scholar] [CrossRef]
  131. Liang, Y.; Zhang, H.; Lin, W.; Liu, B.; Liu, H.F. Characterization of photo-isomerization-induced refractive index response for azobenzene solution based on capillary-assisted Mach-Zehnder interferometer under 473 nm laser excitation. Opt. Laser Technol. 2022, 151, 108045. [Google Scholar] [CrossRef]
  132. Liang, Y.; Zhang, H.; Huang, B.; Liu, B.; Lin, W.; Sun, J.J.; Wang, D.B. Ultrahigh-sensitivity temperature sensor based on resonance coupling in liquid-infiltrated side-hole microstructured optical fibers. Sens. Actuators A Phys. 2022, 334, 113358. [Google Scholar] [CrossRef]
  133. Sun, Y.Y.; Wang, Y.; Zhang, Z. Analysis and experimental research on the principle of surface bonded FBG strain transfer. J. Funct. Mater. 2016, 47, 07046–07050. [Google Scholar]
  134. Her, S.C.; Huang, C.Y. Effect of coating on the strain transfer of optical fiber sensors. Sensors 2011, 11, 6926–6941. [Google Scholar] [CrossRef]
  135. Wang, Q.; Qiu, Y.; Zhao, H. Analysis of strain transfer of six-layer surface-bonded fiber Bragg gratings. Appl. Opt. 2012, 51, 4129–4138. [Google Scholar] [CrossRef]
  136. Wan, K.T.; Leung, C.K.Y.; Olson, N.G. Investigation of the strain transfer for surface-attached optical fiber strain sensors. Smart Mater. Struct. 2008, 17, 035037. [Google Scholar] [CrossRef]
  137. Wu, R.; Zheng, B.; Liu, Z. Analysis on strain transfer of a pasted FBG strain sensor. Optik 2014, 125, 4924–4928. [Google Scholar] [CrossRef]
  138. Zhao, H.; Wang, Q.; Qiu, Y. Strain transfer of surface-bonded fiber Bragg grating sensors for airship envelope structural health monitoring. J. Zhejiang Univ. Sci. A 2012, 13, 538–545. [Google Scholar] [CrossRef]
  139. Xue, G.; Fang, X.; Hu, X. Measurement accuracy of FBG used as a surface-bonded strain sensor installed by adhesive. Appl. Opt. 2018, 57, 2939–2946. [Google Scholar] [CrossRef] [PubMed]
  140. Her, S.C.; Huang, C.Y. The effects of adhesive and bonding length on the strain transfer of optical fiber sensors. Appl. Sci. 2016, 6, 13. [Google Scholar] [CrossRef]
  141. Kwon, H.; Park, Y.; Shrestha, P. Signal characteristics of the surface bonded fiber Bragg grating sensors by bonding length under different load types. In Proceedings of the 2017 25th Optical Fiber Sensors Conference (OFS), Jeju, Republic of Korea, 24–28 April 2017; IEEE: New York, NY, USA, 2017; pp. 1–4. [Google Scholar]
  142. Eckstorm, C. Flight Loads Measurements Obtained from Calibrated Strain-Gage Bridges Mounted Externally on Skin of Low-Aspect-Ratio Wing; NASA TN D-8349; NASA: Washington, DC, USA, 1976.
  143. Jenkins, J.; DeAngelis, V. A Summary of Numerous Strain-Gauge Load Calibrations on Aircraft Wings and Tails in a Technology Format; NASA Technical Memorandum 4804; NASA: Washington, DC, USA, 1997.
  144. Lokos, W.; Stauf, R. Strain-Gage Calibration Parametric Study; NASA TM 212853; NASA: Washington, DC, USA, 2004.
  145. Monaghan, R.; Fields, R. Experiments to Study Strain-Gage Load Calibrations on Wing Structure at Elevated Temperatures; NASA TN D-7390; NASA: Washington, DC, USA, 1973.
  146. Hönisch, G. Der Einfluß der Dehnungsübertragung zwischen Bauteil und Meßdraht auf den k-Faktor von DMS. Maschinenbautechnik 1966, 15, 61–68. [Google Scholar]
  147. Ko, R. Standard Specification Format Guide and Test Proce-Dure for Single-Axis Interferometric Fiber Optic Gyros; IEEE: New York, NY, USA, 1972. [Google Scholar]
  148. VDI/VDE/GESA 2004 Experimentelle Strukturanalyse—Empfehlung zur Durchf¨uhrung von Dehnungsmessungen bei hohen Temperaturen. In Experimental Structural Analysis—Recommended Practice for High-Temperature Strain Measurements; VDI Richtlinie 2635 Blatt 2; VDI: Berlin, Germany, 2019.
  149. VDI/VDE/GESA 2007 Experimentelle Strukturanalyse—Dehnungsmessstreifen mit metallischem Messgitter—Kenngr¨oßen und Pr¨ufbedingungen. In Experimental Structure Analysis—Metallic Bonded Resistance Strain Gages—Characteristics and Testing Conditions; VDI Richtlinie 2635 Blatt 1; VDI: Berlin, Germany, 2024.
  150. Inaudi, D.; Glisic, B. Reliability and field testing of distributed strain and temperature sensors, Smart Structures and Materials 2006: Smart Sensor Monitoring Systems and Applications. Int. Soc. Opt. Photonics 2006, 6167, 61671D. [Google Scholar]
  151. Habel, W.R.; Schukar, V.G.; Kusche, N. Fibre-optic strain sensors are making the leap from lab to industrial use—Reliability and validation as a precondition for standards. Meas. Sci. Technol. 2013, 24, 094006. [Google Scholar]
  152. Roths, J.; Wilfert, A.; Kratzer, P. Strain calibration of optical FBG-based strain sensors, Fourth European Workshop on Optical Fibre Sensors. Int. Soc. Opt. Photonics 2010, 7653, 76530F. [Google Scholar]
  153. Habel, W.R.; Schukar, V.G.; Kusche, N. Calibration Facility for Quality Certification of Surface-Attached Fiber Optic and Electrical Strain Sensors. In Proceedings of the Sensors, 2011 IEEE, Limerick, Ireland, 28–31 October 2011; IEEE: New York, NY, USA, 2011; pp. 1337–1340. [Google Scholar]
  154. IEC 61757-1-1:2016; Fibre Optic Sensors-Part 1-1: Strain Measurement-Strain Sensors Based on Fibre Bragg Gratings. Available online: http://www.iec.ch/ (accessed on 10 February 2020).
  155. Lazarev, V.A.; Leonov, S.O.; Tarabrin, M.K.; Karasik, V.E.; Lehmann, P.; Osten, W.; Gonçalves, A.A. Fiber Bragg gratings strain measuring system and a sensor calibration setup based on mechanical nanomotion transducer, Optical Measurement Systems for Industrial Inspection X. Int. Soc. Opt. Photonics 2017, 10329, 103292F. [Google Scholar]
  156. Święch, Ł. Calibration of a Load Measurement System for an Unmanned Aircraft Composite Wing Based on Fibre Bragg Gratings and Electrical Strain Gauges. Aerospace 2020, 7, 27. [Google Scholar] [CrossRef]
  157. Ogunleye, R.O.; Rusnáková, S.; Javořík, J.; Žaludek, M.; Kotlánová, B. Advanced sensors and sensing systems for structural health monitoring in aerospace composites. Adv. Eng. Mater. 2024, 26, 2401745. [Google Scholar] [CrossRef]
  158. Pashaie, R.; Shokrieh, M.M.; Vahedi, M.; Mirzaei, A.H.; Akbari, S. A comparative study of residual stress measurement of laminated composites using FBG sensor, DIC technique, and strain gauge. Opt. Quantum Electron. 2024, 56, 1790. [Google Scholar] [CrossRef]
  159. Liu, B.; Pang, J.; Tu, X.; Zhou, Z. Three components strain-gauge type aircraft surface friction resistance sensor: Design, manufacturing, and calibration. Measurement 2023, 218, 113165. [Google Scholar] [CrossRef]
  160. Ibrar Jahan, M.A.; Honnungar, R.V.; Nandhini, V.L.; Malini, V.L.; Vohra, H.; Balaji, V.R.; Royc, S.K. Deciphering the sensory landscape: A comparative analysis of fiber Bragg grating and strain gauge systems in structural health monitoring. J. Opt. 2024, 1–9. [Google Scholar] [CrossRef]
  161. Ling, S.; Chen, X.; Miao, L.; Yin, J.; Jian, J.; Jin, Q. A novel semiconductor piezoresistive thin-film strain gauge with high sensitivity. IEEE Sens. J. 2024, 24, 13914–13924. [Google Scholar] [CrossRef]
  162. Deng, K.; Ompusunggu, A.P.; Xu, Y.; Skote, M.; Zhao, Y. A review of material-related mechanical failures and load monitoring-based structural health monitoring (SHM) technologies in aircraft landing gear. Aerospace 2025, 12, 266. [Google Scholar] [CrossRef]
  163. Güemes, A.; Fernandez-Lopez, A. Recent advances in loads/strain monitoring: A review. Struct. Health Monit. Manag. Aerosp. Struct. 2024, 101–113. [Google Scholar] [CrossRef]
  164. Lu, H.; Feng, Y.; Wang, S.; Liu, J.; Han, Q.; Meng, Q. A high-performance, sensitive, low-cost LIG/PDMS strain sensor for impact damage monitoring and localization in composite structures. Nanotechnology 2024, 35, 355702. [Google Scholar] [CrossRef] [PubMed]
  165. Meher, U.; Sunny, M.R. Localization and quantification of delamination/disbond inside a composite lap-joint using novel cross and drive point mechanical impedance based feature. Mech. Syst. Signal Process. 2024, 220, 111661. [Google Scholar] [CrossRef]
  166. Marinho, N.R.; Loendersloot, R.; Wiegman, J.W.; Grooteman, F.; Tinga, T. Evaluating sensor performance for impact identification in composites: A comprehensive comparison of FBGs with PZTs. Struct. Health Monit. 2024, 14759217241304644. [Google Scholar] [CrossRef]
  167. Xu, Z.E.N.G.; Deshuang, D.E.N.G.; Hongjuan, Y.A.N.G.; Zhengyan, Y.A.N.G.; Shuyi, M.A.; Lei, Y.A.N.G.; Zhanjun, W.U. Research progress in low-velocity impact monitoring technology for aircraft structures. Acta Aeronaut. Astronaut. Sin. 2024, 45, 030368. [Google Scholar]
  168. Verma, P.; Ubaid, J.; Varadarajan, K.M.; Wardle, B.L.; Kumar, S. Synthesis and characterization of carbon nanotube-doped thermoplastic nanocomposites for the additive manufacturing of self-sensing piezoresistive materials. ACS Appl. Mater. Interfaces 2022, 14, 8361–8372. [Google Scholar] [CrossRef]
  169. Chung, D.D.L. A critical review of piezoresistivity and its application in electrical-resistance-based strain sensing. J. Mater. Sci. 2020, 55, 15367–15396. [Google Scholar] [CrossRef]
  170. Carani, L.B.; Humphrey, J.; Rahman, M.M.; Okoli, O.I. Advances in embedded sensor technologies for impact monitoring in composite structures. J. Compos. Sci. 2024, 8, 201. [Google Scholar] [CrossRef]
  171. Krohn, D.; Rannow, R.K.; Westbrook, P.; Rochat, E.; Paul, D.; Arkwright, J. White Paper-Smart Infrastructure Using Distributed Fiber Optic Sensors. In Smart Infrastructure Using Distributed Fiber Optic Sensors; IEEE: Piscataway, NJ, USA, 2025; pp. 1–19. [Google Scholar]
  172. Grating, Sensors—Fiber Bragg, and Interrogator Standard—Terminonolgy. “P2067™/D32.” 2020. Available online: https://ieeexplore.ieee.org/abstract/document/9353446/ (accessed on 20 July 2025).
Figure 1. Application of FBG sensor in aircraft SHM [7].
Figure 1. Application of FBG sensor in aircraft SHM [7].
Aerospace 12 00827 g001
Figure 2. Strain and displacement testing experiments of SUAS [9].
Figure 2. Strain and displacement testing experiments of SUAS [9].
Aerospace 12 00827 g002
Figure 3. Schematic diagram of FBG working principle [30].
Figure 3. Schematic diagram of FBG working principle [30].
Aerospace 12 00827 g003
Figure 4. Diagram of tube-packaged FBG sensor [67].
Figure 4. Diagram of tube-packaged FBG sensor [67].
Aerospace 12 00827 g004
Figure 5. Layout diagram of embedded FBG sensor [84].
Figure 5. Layout diagram of embedded FBG sensor [84].
Aerospace 12 00827 g005
Figure 6. Mechanical transfer model of FBG sensor [6].
Figure 6. Mechanical transfer model of FBG sensor [6].
Aerospace 12 00827 g006
Figure 7. Calibration device based on a four-point bending configuration [152].
Figure 7. Calibration device based on a four-point bending configuration [152].
Aerospace 12 00827 g007
Table 1. Classification of fiber optic sensor applications in the aviation field.
Table 1. Classification of fiber optic sensor applications in the aviation field.
Application Area FunctionExampleRef.
Structural health monitoringReal-time monitoring of structural integrity and detection of damageFBG array in composite beam/wing box[13,14,15,16]
Load/strain/vibration monitoringDetect changes in load and vibration modesFBG array for hypersonic aircraft[17,18,19]
Thermal environment monitoringEvaluate the temperature of the thermal protection system or the surface of the bodySTORT aircraft thermal protection system monitoring[20,21,22]
Photon system supportSupport high-speed optical communication and modulation systemsFBG demodulator, tunable laser, filter[23,24,25,26]
Table 2. Performance chart for distributed sensors [41,42,43].
Table 2. Performance chart for distributed sensors [41,42,43].
Spatial
Resolution
Measurement
Time
Temperature
and Strain
Strain AccuracyDynamic
Measurement
DPP-BOTDA2 m2–5 minYes20 μ ε Yes
BOTDR~1 m1–5 minYes60 μ ε No
OFDR~1 mm(0.01–3) sYes1 μ ε 30 Hz
Phase OTDR~0.5 m<1 msNoNo No
Table 3. Classification by sensor technology route.
Table 3. Classification by sensor technology route.
Sensor TechnologyApplication ScenarioOptimization FocusRef.
Extrinsic Fabry–Perot Interferometer (EFPI)Curing strain monitoring of aviation composite laminates and axial strain detection of wingsExpanding the dynamic range by nine times through a hemispherical polymer cap to achieve sub-nanometer-resolution displacement measurement[44]
OTDRDistributed strain monitoring of composite material bridges and detection of debonding damage to wing bolt linesImprove measurement accuracy and spatial resolution, achieve continuous strain/deflection distribution measurement of the entire length of the structure[45]
BOTDRLong-distance strain/temperature monitoring of large-area aircraft structuresImproving spatial resolution based on iterative subdivision method, up to 0.1 m of 1 km fiber can be achieved under 10 ns pulse[46]
FBGMicro-strain monitoring in the manufacturing process of composite materials and multi-parameter distributed monitoring during service life3D strain mapping achieved through ceramic coated FBG[47]
Sapphire fiber EFPIStrain monitoring of extreme high-temperature environment near the engineCapable of withstanding temperatures exceeding 1000 °C and achieving synchronous measurement of high temperature and high pressure[48]
Iterative subdivision BOTDRFine positioning of local strain in aircraft structure and optimization of long-distance monitoringCan extract sub-Brillouin signals and optimize the ability to capture long-distance local damage[49]
Artificial Neural Network BOTDRStrain monitoring and dynamic data processing of aircraft structures under complex loadsResolution can be increased from 21.13 MHz to 2.88 MHz under 60 ns pulse[50]
Small diameter FBGStrain monitoring of lightweight drone structures and damage detection of thin-walled composite materialsReduce the impact on the mechanical properties of the substrate structure and adapt to the vacuum assisted resin infusion (VARI) process[51]
High pressure vessel compatible with FBG (PTFE protection)Strain monitoring during the manufacturing stage of aviation composite materials and microcrack detection during the curing processCan resist high temperature and high pressure deformation, and can evaluate residual strain[52]
Dynamic FBG Interrogation (FFT)Real-time health monitoring of aircraft, rapid response to fluctuations and captureCan improve wavelength scanning speed, with a resolution of 0.11 pm at a scanning frequency of 40 kHz[53]
Anti torsional deformation FBGMulti-directional load aircraft component strain monitoring and precise measurement of complex stress fieldsCan reduce torsional deformation error, with a relative error of 0.86% in bending direction measurement[54]
Distributed ROTDRMonitoring of temperature field distribution on aircraft wings and thermal strain detection in wide-temperature environmentsCan maintain constant Raman Stokes power reception and extend sensing distance[55]
Wavelength Division Multiplexing FPIMulti-area strain synchronous monitoring of aircraft and integration of multiple sensorsCan demodulate the length of each FPI cavity through a bandpass filter, breaking through the limitations of FPI multiplexing[56]
Prestressed FBGMonitoring of shrinkage strain in thick composite structures and control of deformation in laminated platesIdentifying spectral distortion caused by damage through a fixed FBG filtering decoding system[57]
Table 4. Comparison of the characteristics of FBG sensors in different packaging forms.
Table 4. Comparison of the characteristics of FBG sensors in different packaging forms.
CharacteristicTube-Packaged FBGEmbedded FBGSurface-Attached FBGRef.
Core integration methodPlace the bare FBG in the sleeve, apply pre tension and inject encapsulation adhesive to fix it, so that the FBG is embedded in the sleeveEmbedding reinforcing fiber layers during manufacturing, permanently embedded in composite laminatesPrepare composite laminates, and then use high-performance adhesives to stick FBG onto the surface of the cured laminates[99,100,101]
Target sensing applicationsMonitor structural strain and assist in long-term structural health monitoringMonitor the internal strain and temperature during the curing/post curing stage of composite materials. Detect BVD and locate itDetecting surface damage and locating it. Large-scale structural SHM (such as wind turbine blades)[102,103,104]
Detecting physical parametersStructural strain (dependent on strain transfer between sleeve encapsulation adhesive FBG)Interlayer strain, interlayer temperature, impact, vibration, damageSurface strain, surface temperature, impact, damage, vibration, pressure[105,106]
Manufacturing processTo bare FBG casing, apply pre tightening force, and seal with adhesive. The casing material (metal, capillary, etc.) affects performance and may involve process validation such as VARTMSelect interlayer pre strain fixed FBG, operate before resin impregnation/curing. Preparation of laminated panels including manual layering, hot pressing, VARIM, etc.Curing the surface of laminated panels, adhesive/matrix epoxy bonding pre strain FBG. Preparation of laminated panels including manual layering, hot pressing, VARTM, etc.[107,108]
temperature sensitivityDue to the influence of temperature characteristics of casing and encapsulation adhesive materials, attention should be paid to temperature–strain couplingSensitive to curing/chemical reaction temperature, affected by changes in fiber/matrix temperatureEasy to be affected by external environmental temperature, which may cause inaccurate sensing[109,110,111]
Strain sensitivityDependent on encapsulation adhesive strain transfer, affected by sleeve constraints, it is necessary to optimize encapsulation parameters to ensure transfer efficiencyAxial strain disturbance detection is more sensitiveSurface measurement improves sensitivity, but the value may differ from the actual internal strain[112,113]
Maintain characteristicsReplacement after damage requires removal of sleeves, which is relatively complexDifficult to maintain and replace, requiring replacement of the entire intelligent composite material structureCan be replaced and maintained for easy visual inspection[114,115]
Defects related to laminated panelsIf the compatibility between the encapsulation adhesive and the laminated board is poor, it may affect the local mechanical properties of the laminated boardEasy occurrence of resin-rich areas/bubbles, debonding, decreased mechanical properties, etc.No obvious defects related to laminated panels[116,117,118]
Defects of sensorsThere is uneven strain transmission. Long-term use of encapsulation adhesive aging affects performanceThere are angular offsets, residual strains, etc.Poor bonding process may result in adhesive residue, weak bonding, and affect sensing[119,120]
Long-term stabilityWhen the packaging is good, the stability of the sleeve and packaging adhesive affects long-term performance. If the material is stable, it can work stably for a longer period of timeDue to the influence of internal stability of materials and manufacturing defects, long-term monitoring should pay attention to performance driftDue to the influence of surface environment (humidity, external forces), long-term stability is prone to fluctuations and requires regular calibration[121,122]
Strain transmission efficiencyDue to the influence of mechanical properties of packaging materials, optimize packaging parameters to ensure efficiency is neededTransferred through the matrix material, close to the actual internal strain, but manufacturing defects can easily cause transmission deviationDependent on adhesive transfer, affected by adhesive properties and bonding processes, surface strain transfer is direct but may differ from the interior[123,124,125]
Table 5. Comparison of calibration techniques for FBG sensors, traditional strain gauges, and piezoresistive sensors.
Table 5. Comparison of calibration techniques for FBG sensors, traditional strain gauges, and piezoresistive sensors.
Calibration Technology DimensionFBG SensorTraditional Strain GaugesPiezoresistive Sensor Ref.
Core calibration principleBased on the relationship between Bragg wavelength and strain/temperature variationBased on the relationship between resistance change and strain changeBased on the relationship between resistance changes and strain/damage changes[157,158]
Mainstream calibration methodsStatic calibration, dynamic calibration, temperature–strain coupling calibrationStatic calibration, dynamic calibrationStatic calibration, damage calibration[159,160]
Calibration accuracyHigh (static calibration error ± 1 με, dynamic calibration relative standard deviation ≈ 0.16%)Middle (static error ± 5–10 με, affected by lead resistance)Low (error ± 5–10%, affected by CNT dispersion and interface damage)[161,162]
Environmental interference factorsEncapsulation stress release, fiber eccentricity. Temperature cross-sensitivity requires specialized compensation calibrationVulnerable to electromagnetic interference and changes in lead resistanceAffected by humidity, CNT aggregation, and fiber matrix interface state[163,164]
Calibration equipment complexityHighLow Middle[165,166]
Feasibility of on-site calibrationLowHighMiddle[167,168]
Long term calibration stabilityHighMiddleLow[169,170]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ma, Z.; Chen, X.; Cui, B.; Wang, X. Advanced Fiber Optic Sensing Technology in Aerospace: Packaging, Bonding, and Calibration Review. Aerospace 2025, 12, 827. https://doi.org/10.3390/aerospace12090827

AMA Style

Ma Z, Chen X, Cui B, Wang X. Advanced Fiber Optic Sensing Technology in Aerospace: Packaging, Bonding, and Calibration Review. Aerospace. 2025; 12(9):827. https://doi.org/10.3390/aerospace12090827

Chicago/Turabian Style

Ma, Zhen, Xiyuan Chen, Bingbo Cui, and Xinzhong Wang. 2025. "Advanced Fiber Optic Sensing Technology in Aerospace: Packaging, Bonding, and Calibration Review" Aerospace 12, no. 9: 827. https://doi.org/10.3390/aerospace12090827

APA Style

Ma, Z., Chen, X., Cui, B., & Wang, X. (2025). Advanced Fiber Optic Sensing Technology in Aerospace: Packaging, Bonding, and Calibration Review. Aerospace, 12(9), 827. https://doi.org/10.3390/aerospace12090827

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop