[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,356)

Search Parameters:
Keywords = fiber optics sensors

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 1808 KiB  
Article
Prediction of Thrombus Formation within an Oxygenator via Bioimpedance Analysis
by Jan Korte, Tobias Lauwigi, Lisa Herzog, Alexander Theißen, Kai Suchorski, Lasse J. Strudthoff, Jannis Focke, Sebastian V. Jansen, Thomas Gries, Rolf Rossaint, Christian Bleilevens and Patrick Winnersbach
Biosensors 2024, 14(10), 511; https://doi.org/10.3390/bios14100511 - 18 Oct 2024
Viewed by 249
Abstract
Blood clot formation inside the membrane oxygenator (MO) remains a risk in extracorporeal membrane oxygenation (ECMO). It is associated with thromboembolic complications and normally detectable only at an advanced stage. Established clinical monitoring techniques lack predictive capabilities, emphasizing the need for refinement in [...] Read more.
Blood clot formation inside the membrane oxygenator (MO) remains a risk in extracorporeal membrane oxygenation (ECMO). It is associated with thromboembolic complications and normally detectable only at an advanced stage. Established clinical monitoring techniques lack predictive capabilities, emphasizing the need for refinement in MO monitoring towards an early warning system. In this study, an MO was modified by integrating four sensor fibers in the middle of the hollow fiber mat bundle, allowing for bioimpedance measurement within the MO. The modified MO was perfused with human blood in an in vitro test circuit until fulminant clot formation. The optical analysis of clot residues on the extracted hollow fibers showed a clot deposition area of 51.88% ± 14.25%. This was detectable via an increased bioimpedance signal with a significant increase 5 min in advance to fulminant clot formation inside the MO, which was monitored by the clinical gold standard (pressure difference across the MO (dp-MO)). This study demonstrates the feasibility of detecting clot growth early and effectively by measuring bioimpedance within an MO using integrated sensor fibers. Thus, bioimpedance may even outperform the clinical gold standard of dp-MO as a monitoring method by providing earlier clot detection. Full article
Show Figures

Figure 1

Figure 1
<p>View of modified RatOx: (<b>A</b>) Top view. (<b>B</b>) Lateral view. (<b>C</b>) Axial Micro-CT scan. Hollow fiber mat bundle (+) with integrated sensor fibers (*).</p>
Full article ">Figure 2
<p>Schematic illustration of the experimental setup. dp-MO: pressure difference across the oxygenator. The syringe indicates the location of blood sampling; created with BioRender.com.</p>
Full article ">Figure 3
<p>Exemplary images of stamped-out hollow fiber mats of the RatOx after flushing with sodium chloride solution. (<b>Left image</b>) Blood clot formation adherent to sensor fibers; (<b>right image</b>) no visible blood clot formation. View from inflow direction (1) until the last hollow fiber mat at the outflow region (11); one remaining hollow fiber mat (6) inside the bundle next to the sensor fibers.</p>
Full article ">Figure 4
<p>(<b>A</b>) The bar chart represents the share of hollow fiber mat area covered with blood clot residues for the clotting and control groups; the color contrast ranges from white (no/barely covered with blood clot deposits) to dark red (heavily covered with blood clot deposits) shown within the bars. Mean clot deposit on the hollow fiber mats surface is significantly higher in the clotting vs. control group and within the control group on hollow fiber mat No. 1 and 6 vs. every other hollow fiber mat in the control group. ANOVA: ** <span class="html-italic">p</span> &lt; 0.002 vs. hollow fiber mat No. 2 to 5 and 7 to 11; *** <span class="html-italic">p</span> &lt; 0.001 vs. control. (<b>B</b>) Exemplary scans of the hollow fiber mat layers 1–11, representative of the values shown in (<b>A</b>).</p>
Full article ">Figure 5
<p>Significant increase of impedance signal 5 min prior to fulminant clot formation (green area * <span class="html-italic">p</span> &lt; 0.05) in the clotting group (dots, 0.25 IU/mL heparin, <span class="html-italic">n</span> = 7), in contrast to the control group (squares, 5 IU/mL heparin, <span class="html-italic">n</span> = 5) showing no increase. First indicators for clot formation in hemodynamic parameters (increase in dp-MO) were detectable 3 min later (yellow area). Fulminant clot formation was indicated by an increase in dp-MO in conjunction with doubling of the pump’s speed (&gt;60 RPM) resulting in the termination of the experiments (red area). Mean ± SD, ANOVA; * <span class="html-italic">p</span> &lt; 0.05 vs. baseline (BL).</p>
Full article ">Figure 6
<p>Platelet count (PLTs) and Activated Clotting Time (ACT) over the relative duration of the experiments. (<b>A</b>) PLTs in the clotting group (0.25 IU/mL; <span class="html-italic">n</span> = 7) decreased significantly compared to the those in the control group (5 IU/mL; <span class="html-italic">n</span> = 5). (<b>B</b>) In the control group, ACT increased significantly after initial heparinization, in comparison to the control group. All values within the respective quarter were cumulated and averaged. Mean ± SEM, ANOVA; * <span class="html-italic">p</span> &lt; 0.05, ; ** <span class="html-italic">p</span> &lt; 0.002; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
45 pages, 8541 KiB  
Review
Polymer-Based Optical Guided-Wave Biomedical Sensing: From Principles to Applications
by Malhar A. Nagar and Davide Janner
Photonics 2024, 11(10), 972; https://doi.org/10.3390/photonics11100972 (registering DOI) - 17 Oct 2024
Viewed by 298
Abstract
Polymer-based optical sensors represent a transformative advancement in biomedical diagnostics and monitoring due to their unique properties of flexibility, biocompatibility, and selective responsiveness. This review provides a comprehensive overview of polymer-based optical sensors, covering the fundamental operational principles, key insights of various polymer-based [...] Read more.
Polymer-based optical sensors represent a transformative advancement in biomedical diagnostics and monitoring due to their unique properties of flexibility, biocompatibility, and selective responsiveness. This review provides a comprehensive overview of polymer-based optical sensors, covering the fundamental operational principles, key insights of various polymer-based optical sensors, and the considerable impact of polymer integration on their functional capabilities. Primary attention is given to all-polymer optical fibers and polymer-coated optical fibers, emphasizing their significant role in “enabling” biomedical sensing applications. Unlike existing reviews focused on specific polymer types and optical sensor methods for biomedical use, this review highlights the substantial impact of polymers as functional materials and transducers in enhancing the performance and applicability of various biomedical optical sensing technologies. Various sensor configurations based on waveguides, luminescence, surface plasmon resonance, and diverse types of polymer optical fibers have been discussed, along with pertinent examples, in biomedical applications. This review highlights the use of biocompatible, hydrophilic, stimuli-responsive polymers and other such functional polymers that impart selectivity, sensitivity, and stability, improving interactions with biological parameters. Various fabrication techniques for polymer coatings are also explored, highlighting their advantages and disadvantages. Special emphasis is given to polymer-coated optical fiber sensors for biomedical catheters and guidewires. By synthesizing the latest research, this review aims to provide insights into polymer-based optical sensors’ current capabilities and future potential in improving diagnostic and therapeutic outcomes in the biomedical field. Full article
(This article belongs to the Special Issue Emerging Trends in Optical Fiber Sensors and Sensing Techniques)
Show Figures

Figure 1

Figure 1
<p>Different classes of polymer-based optical fiber sensors.</p>
Full article ">Figure 2
<p>Attenuation characteristics of polymer optical fibers made from different materials: PMMA (polymethyl methacrylate), perfluorinated polymer fiber, and standard silica optical fibers. Reprinted with permission from [<a href="#B34-photonics-11-00972" class="html-bibr">34</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Transmission setup of graded index waveguide for label-free sensing of biomolecules; (<b>b</b>) schematic of guided mode resonance; (<b>c</b>) two-channel microfluidic chip with embedded waveguide; (<b>d</b>) transmission intensity distribution on a charged coupled device; (<b>e</b>) intensity distribution along a specific row. Reproduced from [<a href="#B43-photonics-11-00972" class="html-bibr">43</a>] with permission, under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024); (<b>f</b>) MZI-based slot waveguide for refractive index biosensing. Reproduced with permission [<a href="#B44-photonics-11-00972" class="html-bibr">44</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024). GWT-GMR: gradient waveguide thickness—guided mode resonance, BCB: benzocyclobutene.</p>
Full article ">Figure 4
<p>(<b>a</b>) Stretchable and multifunctional optical sensor integrated with knee pad to monitor knee-related motions; (<b>b</b>) strain response recorded by optical sensor for various knee motions; (<b>c</b>) sensor attached to finger for simultaneous detection of motion and skin temperature; (<b>d</b>) temperature and strain response of sensor during finger wagging. Reprinted with permission from [<a href="#B54-photonics-11-00972" class="html-bibr">54</a>] © Optical Society of America.</p>
Full article ">Figure 5
<p>(<b>a</b>) Basic working principle of SPR sensor; (<b>b</b>) SPR-based planar optical smartphone-assisted bio-chip sensor for lab-on-chip applications. Reprinted from [<a href="#B55-photonics-11-00972" class="html-bibr">55</a>] and [<a href="#B62-photonics-11-00972" class="html-bibr">62</a>], respectively, under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024). I<sub>in</sub>: Input light intensity launched into fiber optical fiber core, I<sub>out</sub>: Output light intensity from optical fiber.</p>
Full article ">Figure 6
<p>Various configurations of optical fiber sensors <b>1</b>. Intensity-based optical sensors; (<b>a</b>) represent simple design for monitoring transmitted light depending on distance d. (<b>b</b>) Transmitted light coupling can be enhanced using differential setup; (<b>c</b>) micro-bend sensing. Reprinted from [<a href="#B74-photonics-11-00972" class="html-bibr">74</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024);. <b>2.</b> FBG-based sensing, before (<b>a</b>) and after (<b>b</b>) changes in effective refractive index (n<sub>eff</sub>) and grating period (Λ) due to external perturbations from various measurands. <b>3</b>. Phase-based sensing; (<b>a</b>) Mach–Zehnder, (<b>b</b>) Michelson, (<b>c</b>) Sagnac, and (<b>d</b>) Fabry–Perot interferometers. Reprinted from [<a href="#B75-photonics-11-00972" class="html-bibr">75</a>], under CC 3.0, <a href="https://creativecommons.org/licenses/by/3.0/" target="_blank">https://creativecommons.org/licenses/by/3.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 7
<p>Various configurations of FBGs and their spectral responses. (<b>a</b>) Uniform, chirped, tilted, etched, and array FBGs. (<b>b</b>) Standard interrogation unit for FBGs (either SLED source coupled to a interrogator or a tunable laser source coupled to a photodetector). Reproduced from [<a href="#B92-photonics-11-00972" class="html-bibr">92</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024). CFBG: chirped FBG, TFBG: tilted FBG, EFBG: etched FBG, ε: strain, T: temperature, RI: refractive index, SLED: superluminescent LED, SP: spectrometer, PD: phototdetector, TIA: transimpedance amplifier, DAQ: data acquisition system.</p>
Full article ">Figure 8
<p><b>1</b>. (<b>a</b>) Poly(L-lactic acid) biopolymer film, (<b>b</b>) polyethylene glycol hydrogel waveguide array guiding green laser light, (<b>c</b>) biopolymer-based waveguide light delivery of photoactivated area at dyed porcine skin. Reprinted from [<a href="#B115-photonics-11-00972" class="html-bibr">115</a>], under CC 4.0 (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>), (accessed on 7 September 2024). <b>2</b>. (<b>a</b>) Citrate-based optical fiber twisted around glass, (<b>b</b>) light delivery using citrate-based fiber, (<b>c</b>) in vivo deep tissue fluorescence sensing using citrate-based fibers, (<b>d</b>) fluorescence spectrum acquired from light collection fiber (denoted as B in <a href="#photonics-11-00972-f008" class="html-fig">Figure 8</a>(2c) inset). Reprinted with permission from [<a href="#B39-photonics-11-00972" class="html-bibr">39</a>].</p>
Full article ">Figure 9
<p><b>1</b>. Hydrogel fibers consisting of poly(acrylamide-co-poly(ethylene glycol) diacrylate) core with functionalized phenylboronic acid for continuous real-time glucose monitoring. <b>2</b>. p(AM-co-PEGDA-co-3-APBA) hydrogel fiber as glucose sensor. (<b>a</b>) Reusability of fiber—fiber expanded by 6% as the conc. of glucose increased from 1.0 to 12.0 mmolL<sup>−1</sup> and decreased back to its initial value when conc. decreased. (<b>b</b>) Refractive index variation with respect to glucose conc. (<b>c</b>) Transmitted light attenuation across fiber as a function of time boronic acid–glucose cis diol binding. (<b>d</b>) Transmitted light intensity with respect to glucose concentration. Reprinted from [<a href="#B29-photonics-11-00972" class="html-bibr">29</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024). <b>3</b>. Microstructured optical fiber and chitosan film-based Fabry–Perot interferometer. 4. (<b>a</b>) Wavelength shift of specific dip of interference pattern with respect to humidity. (<b>b</b>) Stability of humidity sensor. (<b>c</b>) Response of the sensor to human breath containing two cycles of inhalation and exhalation. (<b>d</b>) Analysis of human breath on one cycle. Reprinted from [<a href="#B117-photonics-11-00972" class="html-bibr">117</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 10
<p>(<b>a</b>) Polylactic acid (PLA)/polyethylene terephthalate glycol (PETG) for shape-memory fiber with three configurations; empty central hole, filled with a polystyrene (PS) rod, and filled with acrylonitrile butadiene styrene (ABS); (<b>b</b>) represents diagram; and (<b>c</b>) cross-sectional optical image of hollow sleeve. (<b>d</b>) PMMA fiber integrated with a hollow sleeve. (<b>e</b>) Represents diagram and cross-sectional optical image in (<b>f</b>) reflection and (<b>g</b>) transmission of a PS optical fiber. (<b>h</b>) Diagram and cross-sectional optical image in (<b>i</b>) reflection and (<b>j</b>) transmission of a PETG optical fiber. Reproduced from [<a href="#B122-photonics-11-00972" class="html-bibr">122</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 11
<p>(<b>a</b>) Mechanical flexibility of stretchable polymer optical fiber sensors, (<b>b</b>) visible upconversion emissions at a temperature-sensitive upconversion-nanoparticle (UCNP)-loaded site in stretchable polymer optical fiber (PDMS) when illuminated with 980 nm laser, (<b>c</b>) cycling test of PDMS-UCNP-based temperature sensor, (<b>d</b>) response and recovery behavior of a PDMS-UCNP based temperature sensor. Reproduced with permission from [<a href="#B3-photonics-11-00972" class="html-bibr">3</a>].</p>
Full article ">Figure 12
<p><b>1</b>. (<b>a</b>) Schematic of an OF–waveguide–OF (OFWF) structure with the MIP coating process (<b>b</b>,<b>c</b>) microscopic images of a MIP-coated waveguide when irradiated with 405 nm. (<b>d</b>) Optical loss of MIP coated OFWF structure, SEM images of (<b>e</b>) MIP-OFWF structure, (<b>f</b>) cross-section of a MIP-coated waveguide, and (<b>g</b>) MIP surface layer. Reproduced with permission from [<a href="#B137-photonics-11-00972" class="html-bibr">137</a>]. <b>2</b>. (<b>a</b>) Simulation of a MIP—hybrid probe, (<b>b</b>) Pure polymer waveguide loss in comparison with a hollow quartz nanoparticle (HQNP) hybrid optical sensor, (<b>c</b>) SEM image of joint of MIP-hybrid optical sensor, (<b>d</b>) simulation-based extraction efficiency (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">η</mi> </mrow> <mrow> <mi>extrac</mi> </mrow> </msub> </mrow> </semantics></math>) with respect to MIP layer thickness for evanescent wave only and its integration with light confined in the MIP layer, (<b>e</b>) SEM image of the cross-section of MIP coated waveguide sensing probe and (<b>f</b>) comparison of optical loss between MIP—hybrid OFWF and MIP—polymer OFWF. Reproduced with permission from [<a href="#B138-photonics-11-00972" class="html-bibr">138</a>].</p>
Full article ">Figure 13
<p><b>1</b>. Fabrication process of polymer neural fiber probe. (<b>a</b>) Preform fabrication, (<b>b</b>) thermal drawing of preform into fiber, (<b>c</b>) cross-section of preform consisting of PC core, cyclic olefin copolymer as cladding, and carbon black-doped conductive polyethylene as electrode. (<b>d</b>) Etched neural probe wrapped around pencil. Reproduced from [<a href="#B140-photonics-11-00972" class="html-bibr">140</a>]. <b>2</b>. (<b>a</b>) Thermal drawing of optical fibers from (<b>b</b>,<b>c</b>) PC preforms (<b>b</b>). (<b>d</b>–<b>f</b>) SEM images of cross sections of porous optical fibers; (<b>f</b>,<b>g</b>) represent this fiber embedded inside resin. Reprinted with permission from [<a href="#B141-photonics-11-00972" class="html-bibr">141</a>] © Optical Society of America.</p>
Full article ">Figure 14
<p><b>1</b>. (<b>a</b>) Representation and (<b>b</b>) experimental setup for microstructured polymer optical fiber extrusion; (<b>c</b>) fiber drawn from structured nozzle; (<b>d</b>) microstructured fiber wrapped around 3D printed spool connected to stepper motor. <b>2</b>. (<b>a</b>,<b>b</b>) Microscope images of fiber cross section at different diameters; (<b>c</b>) near field image of transmitted light at 1550 nm for full turn bending of 12.5 mm radius. Reprinted from [<a href="#B146-photonics-11-00972" class="html-bibr">146</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 15
<p>(<b>a</b>) Millisecond grating inscription inside a PMMA OF using 325 nm He-Cd laser. (<b>b</b>) Refection spectra of standard silica FBG. (<b>c</b>) Reflection spectra of 7 ms PFBG. (<b>d</b>) Filtered reflection wavelength shift of silica (in black) and polymer (in blue). Reprinted from [<a href="#B147-photonics-11-00972" class="html-bibr">147</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 16
<p>(<b>a</b>) Illustration of an etching mechanism. (<b>b</b>) Formation of compound parabolic concentrator at different etching times. (<b>c</b>) Spectrum of compound parabolic concentrator and plane-cut tip sensor. Reprinted from [<a href="#B151-photonics-11-00972" class="html-bibr">151</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 17
<p>(<b>a</b>) Injection molding strategy for fabricating PDMS cores in PTFE tube molds. (<b>b</b>) Dip coating of PDMS core inside the THV (terpolymer of tetrafluoroethylene, hexafluoropropylene, and vinylidene fluoride) polymer solution to process the THV clad of 10 µm. (<b>c</b>) PDMS and PDMS/THV coated with Lumogen red dye. Reprinted from [<a href="#B154-photonics-11-00972" class="html-bibr">154</a>], under CC 4.0, <a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>, (accessed on 7 September 2024).</p>
Full article ">Figure 18
<p>(<b>a</b>) Polymer diaphragm-assisted Fabry–Perot-based pressure transducer for coronary physiological pressure assessment. Light signal is sent to pressure transducer via optical fiber, which gets perturbed in response to diaphragm’s deflection changes due to arterial pressure affecting path length of transmitted light. The optical unit identifies interferometry wavelength and maps it to the applied pressure for visualization. (<b>b</b>) Design evolution of the OptoWire (OW) pressure guidewire. Reproduced with permission from [<a href="#B161-photonics-11-00972" class="html-bibr">161</a>].</p>
Full article ">
5 pages, 1128 KiB  
Communication
Modeling a Fully Polarized Optical Fiber Suitable for Photonic Integrated Circuits or Sensors
by Wenbo Sun
Photonics 2024, 11(10), 961; https://doi.org/10.3390/photonics11100961 - 14 Oct 2024
Viewed by 364
Abstract
A method is developed to make an optical fiber that only transmits fully linearly polarized light and maintains the polarization state. The method for efficient ingesting laser into this fiber is also reported. Using an optical fiber with a prism head, we can [...] Read more.
A method is developed to make an optical fiber that only transmits fully linearly polarized light and maintains the polarization state. The method for efficient ingesting laser into this fiber is also reported. Using an optical fiber with a prism head, we can compress a plane wave into the thin rectangular cross-section fiber, and the light intensity within the fiber is much larger than that of the incidence wave. Our finite-difference time-domain (FDTD) simulation results show that the compressed light in the fiber becomes fully polarized and maintains the polarization state, and can be well coupled out by the resonance rings. This method is suitable for developing highly efficient polarization-maintaining optical fibers in a much simpler way, for applications in photonic integrated circuits or optical sensors. Full article
(This article belongs to the Special Issue Optical Sensing Technologies, Devices and Their Data Applications)
Show Figures

Figure 1

Figure 1
<p>Illustration of the optical fiber and its width, thickness, and the half aperture angle of its receiving end.</p>
Full article ">Figure 2
<p>Illustration of the optical fiber, light source implementing plane, uniaxial perfectly matched layer (UPML)-absorbing boundary conditions (ABC) and periodic boundary conditions (PBC) in the computational domain of the FDTD method.</p>
Full article ">Figure 3
<p>The material refractive index profile in the computational domain. The green color denotes fiber material, the red color denotes perfect conductor, and the blue color denotes free space.</p>
Full article ">Figure 4
<p>The finite-difference time-domain (FDTD) result for incidence plane light fully polarized in the y direction (E<sub>y</sub> only). After 300 time steps of simulation, the light intensity inside the fiber is still nearly zero, which means the y-direction polarized light cannot enter the fiber effectively.</p>
Full article ">Figure 5
<p>The finite-difference time-domain (FDTD) result for incidence plane light fully polarized in z direction (E<sub>z</sub> only). After 600 time steps of simulation, the compressed light in the fiber keeps good transfer mode and can be well coupled out by resonance rings. The light intensity inside the fiber is ~4 times of that of incidence.</p>
Full article ">
18 pages, 7980 KiB  
Article
High-Sensitivity Displacement Sensor Using Few-Mode Optical Fibers and the Optical Vernier Effect
by Luis E. Guillen-Ruiz, Gilberto Anzueto-Sánchez, Alejandro Martínez-Rios, Myriam C. Jiménez-Mares and Javier A. Martin-Vela
Appl. Sci. 2024, 14(20), 9300; https://doi.org/10.3390/app14209300 - 12 Oct 2024
Viewed by 517
Abstract
This paper presents a displacement sensor designed to achieve the Optical Vernier Effect (OVE) through a simple yet robust configuration, enhancing sensitivity and precision in small displacement measurements. The sensor structure comprises a few-mode fiber (FMF) placed between two single-mode fibers (SMF) in [...] Read more.
This paper presents a displacement sensor designed to achieve the Optical Vernier Effect (OVE) through a simple yet robust configuration, enhancing sensitivity and precision in small displacement measurements. The sensor structure comprises a few-mode fiber (FMF) placed between two single-mode fibers (SMF) in an SMF-FMF-SMF (SFS) configuration. A series of distinct configurations of concatenated Mach–Zehnder fiber interferometers (MZFI) were examined, with the lengths of the reference FMF (FMFRef) and sensing FMF (FMFSen) adjusted to track the spectral envelope shifts. The results demonstrate that the direction of the spectral shift is governed by the ratio between the FMFRef and FMFSen lengths. The sensor achieved a sensitivity of up to 39.07 nm/mm and a magnification factor (M factor) of up to 50.09, demonstrating exceptional precision and adaptability across a range of applications. The proposed configuration also enhances the overall sensor performance, highlighting its potential for broader use in fields requiring precise displacement monitoring. Full article
(This article belongs to the Special Issue Recent Trends in Fiber Optic Sensor: Technology and Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the cascade MZFIs.</p>
Full article ">Figure 2
<p>Transducer made of TPU that correlated the phase difference induced by the curvature of the optical fiber with the linear displacement.</p>
Full article ">Figure 3
<p>Experimental setup for the in-cascade MZFIs SFS structure to obtain OVE.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>c</b>) show the transmission spectra of <span class="html-italic">MZFI<sub>Ref</sub></span> and <span class="html-italic">MZFI<sub>Sen</sub></span>, respectively. (<b>b</b>,<b>d</b>) represent their corresponding FFT spectra. (<b>e</b>) is the superimposed transmission of two MZFIs, and (<b>f</b>) is the FFT of the superimposed spectrum.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>c</b>) show the transmission spectra of <span class="html-italic">MZFI<sub>Ref</sub></span> and <span class="html-italic">MZFI<sub>Sen</sub></span>, respectively. (<b>b</b>,<b>d</b>) represent their corresponding FFT spectra. (<b>e</b>) is the superimposed transmission of two MZFIs, and (<b>f</b>) is the FFT of the superimposed spectrum.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>c</b>) show the transmission spectra of <span class="html-italic">MZFI<sub>Ref</sub></span> and <span class="html-italic">MZFI<sub>Sen</sub></span>, respectively. (<b>b</b>,<b>d</b>) represent their corresponding FFT spectra. (<b>e</b>) is the superimposed transmission of two MZFIs, and (<b>f</b>) is the FFT of the superimposed spectrum.</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>c</b>) show the transmission spectra of <span class="html-italic">MZFI<sub>Ref</sub></span> and <span class="html-italic">MZFI<sub>Sen</sub></span>, respectively. (<b>b</b>,<b>d</b>) represent their corresponding FFT spectra. (<b>e</b>) is the superimposed transmission of two MZFIs, and (<b>f</b>) is the FFT of the superimposed spectrum.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>c</b>) show the transmission spectra of <span class="html-italic">MZFI<sub>Ref</sub></span> and <span class="html-italic">MZFI<sub>Sen</sub></span>, respectively. (<b>b</b>,<b>d</b>) represent their corresponding FFT spectra. (<b>e</b>) is the superimposed transmission of two MZFIs, and (<b>f</b>) is the FFT of the superimposed spectrum.</p>
Full article ">Figure 9
<p>The spectral shift of the envelopes as a function of linear displacement from table 1: (<b>a</b>) experiment 1, (<b>b</b>) experiment 2, (<b>c</b>) experiment 3, (<b>d</b>) experiment 4 and (<b>e</b>) experiment 5.</p>
Full article ">Figure 10
<p>Slopes of the envelopes from table 1: (<b>a</b>) experiment 1, (<b>b</b>) experiment 2, (<b>c</b>) experiment 3, (<b>d</b>) experiment 4 and (<b>e</b>) experiment 5.</p>
Full article ">Figure 11
<p>Specific characteristics of the first 4 harmonics of the 3 different devices described in table 2: (<b>a</b>) experimental setup 1, (<b>b</b>) experimental setup 2 and (<b>c</b>) experimental setup 3. It is observed that both the peaks and internal envelopes increase in proportion to “<span class="html-italic">i</span> + 1”, where “<span class="html-italic">i</span>” is the harmonic order.</p>
Full article ">Figure 12
<p>Spectral shift due to linear displacement. The red arrow indicates the wavelength shift: (<b>a</b>) first harmonic (one intersection); (<b>b</b>) second harmonic (three intersections); (<b>c</b>) third harmonic (six intersections); (<b>d</b>) fourth harmonic (multiple intersections).</p>
Full article ">Figure 13
<p>Sensitivity of the first four harmonics of the first device: (<b>a</b>) first harmonic; (<b>b</b>) second harmonic; (<b>c</b>) third harmonic; (<b>d</b>) fourth harmonic.</p>
Full article ">Figure 14
<p>Five devices demonstrating Blueshift (red arrows) in panels (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) and Redshift (blue arrows) in panels (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>).</p>
Full article ">
16 pages, 5371 KiB  
Article
Perovskite Nanocrystal-Coated Inorganic Scintillator-Based Fiber-Optic Gamma-ray Sensor with Higher Light Yields
by Seokhyeon Jegal, Siwon Song, Jae Hyung Park, Jinhong Kim, Seunghyeon Kim, Sangjun Lee, Cheol Ho Pyeon, Sin Kim and Bongsoo Lee
Photonics 2024, 11(10), 936; https://doi.org/10.3390/photonics11100936 - 4 Oct 2024
Viewed by 537
Abstract
Radiation possesses inherent physical characteristics, such as penetrability and radionuclide energy, which enable its widespread applicability in fields such as medicine, industry, environment, security, and research. Advancements in scintillator-based radiation detection technology have led to revolutionary changes by ensuring the safe use and [...] Read more.
Radiation possesses inherent physical characteristics, such as penetrability and radionuclide energy, which enable its widespread applicability in fields such as medicine, industry, environment, security, and research. Advancements in scintillator-based radiation detection technology have led to revolutionary changes by ensuring the safe use and precise measurement of radiation. Nevertheless, certain fields require higher scintillation yields to obtain more refined and detailed results. Therefore, in this study, we explored inorganic scintillators coated with perovskite nanomaterials to detect gamma rays with high light yields. By mixing perovskite with a polymer, we improved the intrinsic characteristics of quantum dots, which otherwise failed to maintain their performance over time. On this basis, we investigated the interactions among inorganic scintillators and a mixed material (CsPbBr3 + PMMA) and confirmed an increase in the scintillation yield and measurement trends. Furthermore, optimized scintillation yield measurement experiments facilitated gamma spectroscopy, demonstrating the validity of our approach through the analysis of the peak channel increases in the energy spectra of various gamma sources in relation to the increased scintillation yield. Full article
Show Figures

Figure 1

Figure 1
<p>Transmission electron microscopy (TEM) images of perovskite nanocrystals.</p>
Full article ">Figure 2
<p>Variation of concentration during nucleation-growth processes.</p>
Full article ">Figure 3
<p>Interaction of quantum dots and polymer. (The yellow arrows in the figure represent the absorbed light, while the green arrows indicate the emitted light.)</p>
Full article ">Figure 4
<p>Fabrication of coating material.</p>
Full article ">Figure 5
<p>PMMA dissolution process.</p>
Full article ">Figure 6
<p>Coating process.</p>
Full article ">Figure 7
<p>Fluorescence resonance energy transfer (FRET) mechanism.</p>
Full article ">Figure 8
<p>Configuration of photon counting. (The internal structure of the connector is explained in the below.)</p>
Full article ">Figure 9
<p>Configuration of the gamma-ray spectroscopy.</p>
Full article ">Figure 10
<p>Calculated efficiency of FRET for (<b>a</b>) GAGG, (<b>b</b>) LYSO, and (<b>c</b>) CdWO<sub>4</sub>.</p>
Full article ">Figure 11
<p>Photon-counting results of Cs-137 40 μCi for (<b>a</b>) GAGG, (<b>b</b>) LYSO, and (<b>c</b>) CdWO4.</p>
Full article ">Figure 12
<p>Photon-counting results of Co-60 40 μCi for (<b>a</b>) GAGG, (<b>b</b>) LYSO, and (<b>c</b>) CdWO<sub>4</sub>.</p>
Full article ">Figure 13
<p>Process of Compton scattering.</p>
Full article ">Figure 14
<p>Energy spectrum of Cs-137 measured by GAGG.</p>
Full article ">Figure 15
<p>Energy spectrum of Co-60 measured by GAGG.</p>
Full article ">Figure 16
<p>Performance sustainability achieved by shifting the peak of the Cs-137 source. (<b>a</b>) shows the energy spectrum results, and (<b>b</b>) presents the channel values of the peaks at each time point, normalized to the values at time 0 h.</p>
Full article ">
13 pages, 2371 KiB  
Article
Deflection and Performance Analysis of Shape Memory Alloy-Driven Fiber–Elastomer Composites with Anisotropic Structure
by Anett Endesfelder, Achyuth Ram Annadata, Aline Iobana Acevedo-Velazquez, Markus Koenigsdorff, Gerald Gerlach, Klaus Röbenack, Chokri Cherif and Martina Zimmermann
Materials 2024, 17(19), 4855; https://doi.org/10.3390/ma17194855 - 2 Oct 2024
Viewed by 656
Abstract
Due to their advantageous characteristics, shape memory alloys (SMAs) are prominent representatives in smart materials. They can be used in application fields such as soft robotics, biomimetics, and medicine. Within this work, a fiber–elastomer composite with integrated SMA wire is developed and investigated. [...] Read more.
Due to their advantageous characteristics, shape memory alloys (SMAs) are prominent representatives in smart materials. They can be used in application fields such as soft robotics, biomimetics, and medicine. Within this work, a fiber–elastomer composite with integrated SMA wire is developed and investigated. Bending and torsion occur when the SMA is activated because of the anisotropic structure of the textile. The metrological challenge in characterizing actuators that perform both bending and torsion lies in the large active deformation of the composite and the associated difficulties in fully imaging and analyzing this with optical measurement methods. In this work, a multi-sensor camera system with up to four pairs of cameras connected in parallel is used. The structure to be characterized is recorded from all sides to evaluate the movement in three-dimensional space. The energy input and the time required for an even deflection of the actuator are investigated experimentally. Here, the activation parameters for the practical energy input required for long life with good deflection, i.e., good efficiency, were analyzed. Different parameters and times are considered to minimize the energy input and, thus, to prevent possible overheating and damage to the wire. Thermography is used to evaluate the heating of the SMA wire at different actuation times over a defined time. Full article
Show Figures

Figure 1

Figure 1
<p>SMA fiber elastomer actuator: (<b>a</b>) layout and structure; (<b>b</b>) dimensions of the specimen; (<b>c</b>) braided Ni-Ti SMA core–sheath structure with copper wire and PA yarns.</p>
Full article ">Figure 2
<p>Electrically connected actuator: (<b>a</b>) front side with tracking point pattern; (<b>b</b>) side view of the actuator in initial and deflected state.</p>
Full article ">Figure 3
<p>Top view of the test setup with multi-sensor camera system and IR camera for temperature measurement.</p>
Full article ">Figure 4
<p>Electrical signals measured on wire surface, set with the parameters: (<b>a</b>) 1.5 A; (<b>b</b>) 1.75 A; (<b>c</b>) 2 A.</p>
Full article ">Figure 5
<p>Displacement in Z direction for different actuation modes, 5 s actuation, 45 s switch-off time, time-shifted for better visibility.</p>
Full article ">Figure 6
<p>Measured temperature of wire surface. (<b>a</b>) Overview of the active actuator with highlighted wire measurement detail, shown in (<b>b</b>); (<b>b</b>) SMA wire at two temperatures during activation at the setting of 2 A 10 V; (<b>c</b>) temperature over 15 s at different electrical activation parameters.</p>
Full article ">Figure 7
<p>Displacement of the Z coordinate of the evaluated soft actuator configuration of switch-on and switch-off times at two different electrical activation modes and times: (<b>a</b>) 5 s switch-on and 45 s switch-off time, (<b>b</b>) 10 s activation and 90 s switch-off time.</p>
Full article ">Figure 8
<p>Calculated resistance at different electrical activation parameters over 15 s.</p>
Full article ">Figure 9
<p>Three-dimensional deflection of an actuator, activated with 2 A. (<b>a</b>) Overview of the actuator with characterized details; (<b>b</b>) displacement point A: left down; (<b>c</b>) displacement point B: right down; (<b>d</b>) displacement point C: transition between the textile layers; (<b>e</b>) deflection angle α.</p>
Full article ">Figure 9 Cont.
<p>Three-dimensional deflection of an actuator, activated with 2 A. (<b>a</b>) Overview of the actuator with characterized details; (<b>b</b>) displacement point A: left down; (<b>c</b>) displacement point B: right down; (<b>d</b>) displacement point C: transition between the textile layers; (<b>e</b>) deflection angle α.</p>
Full article ">Figure 10
<p>(<b>a</b>) Image of a damaged actuator and (<b>b</b>) deflection curve of reduced actuation and fatigue of an actuator, activation with activation setting of 3A, 10 V.</p>
Full article ">
11 pages, 1357 KiB  
Article
Application of a Novel Disposable Flow Cell for Spectroscopic Bioprocess Monitoring
by Tobias Steinwedel, Philipp Raithel, Jana Schellenberg, Carlotta Kortmann, Pia Gellermann, Mathias Belz and Dörte Solle
Chemosensors 2024, 12(10), 202; https://doi.org/10.3390/chemosensors12100202 - 1 Oct 2024
Viewed by 506
Abstract
The evaluation of the analytical capabilities of a novel disposable flow cell for spectroscopic bioprocess monitoring is presented. The flow cell is presterilized and can be connected to any kind of bioreactor by weldable tube connections. It is clamped into a reusable holder, [...] Read more.
The evaluation of the analytical capabilities of a novel disposable flow cell for spectroscopic bioprocess monitoring is presented. The flow cell is presterilized and can be connected to any kind of bioreactor by weldable tube connections. It is clamped into a reusable holder, which is equipped with SMA-terminated optical fibers or an integrated light source and detection unit. This modular construction enables spectroscopic techniques like UV-Vis spectroscopy or turbidity measurements by scattered light for modern disposable bioreactors. A NIR scattering module was used for biomass monitoring in different cultivations. A high-cell-density fed-batch cultivation with Komagataella phaffii and a continuous perfusion cultivation with a CHO DG44 cell line were conducted. A high correlation between the sensor signal and biomass or viable cell count was observed. Furthermore, the sensor shows high sensitivity during low turbidity states, as well as a high dynamic range to monitor high turbidity values without saturation effects. In addition to upstream processing, the sensor system was used to monitor the purification process of a monoclonal antibody. The absorption module enables simple and cost-efficient monitoring of downstream processing and quality control measurements. Recorded absorption spectra can be used for antibody aggregate detection, due to an increase in overall optical density. Full article
Show Figures

Figure 1

Figure 1
<p>Monitoring of a fed-batch cultivation with <span class="html-italic">Komagataella phaffii</span> [<a href="#B20-chemosensors-12-00202" class="html-bibr">20</a>]. (<b>A</b>) Scattering in various angles with offline measurements of optical density (OD) and cell dry weight (DCW). (<b>B</b>) the opacity is compared to the 90°-scattered light for optical densities (OD) up to 10.</p>
Full article ">Figure 2
<p>Opacity and viable cell density of a continuous perfusion cultivation of CHO DG44 cells [<a href="#B19-chemosensors-12-00202" class="html-bibr">19</a>]. The non-continuous course of the cell density is due to the process control [<a href="#B19-chemosensors-12-00202" class="html-bibr">19</a>], the online monitoring with the developed sensor shows changes quickly and reliably.</p>
Full article ">Figure 3
<p>Monitoring of an FPLC run of Protein A capture, comparison of different UV detectors.</p>
Full article ">Figure 4
<p>SEC-HPLC results to detect different rates of antibody aggregates by freeze–thaw cycles (FT) during downstream processing. (<b>A</b>) Complete HPLC chromatogram; (<b>B</b>) Focus on aggregate peak at 7.5 min.</p>
Full article ">Figure 5
<p>Absorption spectra for different aggregate rates by freeze–thaw cycles (FT). (<b>A</b>) FPLC chromatogram in normal scale; (<b>B</b>) FPLC chromatogram in logarithmic scale.</p>
Full article ">
16 pages, 1764 KiB  
Article
Optimal Design of a Sensor Network for Guided Wave-Based Structural Health Monitoring Using Acoustically Coupled Optical Fibers
by Rohan Soman, Jee Myung Kim, Alex Boyer and Kara Peters
Sensors 2024, 24(19), 6354; https://doi.org/10.3390/s24196354 - 30 Sep 2024
Viewed by 531
Abstract
Guided waves (GW) allow fast inspection of a large area and hence have received great interest from the structural health monitoring (SHM) community. Fiber Bragg grating (FBG) sensors offer several advantages but their use has been limited for the GW sensing due to [...] Read more.
Guided waves (GW) allow fast inspection of a large area and hence have received great interest from the structural health monitoring (SHM) community. Fiber Bragg grating (FBG) sensors offer several advantages but their use has been limited for the GW sensing due to its limited sensitivity. FBG sensors in the edge-filtering configuration have overcome this issue with sensitivity and there is a renewed interest in their use. Unfortunately, the FBG sensors and the equipment needed for interrogation is quite expensive, and hence their number is restricted. In the previous work by the authors, the number and location of the actuators was optimized for developing a SHM system with a single sensor and multiple actuators. But through the use of the phenomenon of acoustic coupling, multiple locations on the structure may be interrogated with a single FBG sensor. As a result, a sensor network with multiple sensing locations and a few actuators is feasible and cost effective. This paper develops a two-step methodology for the optimization of an actuator–sensor network harnessing the acoustic coupling ability of FBG sensors. In the first stage, the actuator–sensor network is optimized based on the application demands (coverage with at least three actuator–sensor pairs) and the cost of the instrumentation. In the second stage, an acoustic coupler network is designed to ensure high-fidelity measurements with minimal interference from other bond locations (overlap of measurements) as well as interference from features in the acoustically coupled circuit (fiber end, coupler, etc.). The non-sorting genetic algorithm (NSGA-II) is implemented for finding the optimal solution for both problems. The analytical implementation of the cost function is validated experimentally. The results show that the optimization does indeed have the potential to improve the quality of SHM while reducing the instrumentation costs significantly. Full article
Show Figures

Figure 1

Figure 1
<p>Coupling of three Lamb wave signals into a single FBG detector through (<b>a</b>) one 3 × 1 coupler and (<b>b</b>) two 2 × 1 couplers. Ultrasonic waves shown in red.</p>
Full article ">Figure 2
<p>(<b>a</b>) Experimental setup for characterizing the different bonds. (<b>b</b>) Wave travelling to the acoustic coupler. (<b>c</b>) Energy distribution in all the branches after the coupler.</p>
Full article ">Figure 3
<p>Percent of energy reflected, lost and output from a (<b>a</b>) 2 × 1 couplerm (<b>b</b>) a 3 × 1 coupler. Horizontal axis refers to the branch into which the <math display="inline"><semantics> <msub> <mi mathvariant="normal">L</mi> <mn>01</mn> </msub> </semantics></math> is input.</p>
Full article ">Figure 4
<p>Candidate locations and discretization for coverage calculations.</p>
Full article ">Figure 5
<p>Flowchart for the NSGA-II based on [<a href="#B25-sensors-24-06354" class="html-bibr">25</a>].</p>
Full article ">Figure 6
<p>Pareto front for the optimization of AS network.</p>
Full article ">Figure 7
<p>Surface plot for AS network (yellow shows covered areas, ‘X’ are the actuators, ‘rectangles’ are sensors showing orientation).</p>
Full article ">Figure 8
<p>Pareto front for the optimization of acoustic coupler.</p>
Full article ">Figure 9
<p>Pareto front for the optimization of acoustic coupler for better visualization in 2D.</p>
Full article ">Figure 10
<p>Experimental setup for validation (coordinates and dimensions in cm; reprinted with permission from [<a href="#B17-sensors-24-06354" class="html-bibr">17</a>]).</p>
Full article ">Figure 11
<p>Coverage for the AS network (yellow shows covered areas, ‘o’ are the actuators, ‘rectangles’ are sensors, ‘x’ are the damage locations simulated).</p>
Full article ">Figure 12
<p>Localization of damage for Case D5 [<a href="#B17-sensors-24-06354" class="html-bibr">17</a>].</p>
Full article ">Figure 13
<p>Localization of damage for Case D6 [<a href="#B17-sensors-24-06354" class="html-bibr">17</a>].</p>
Full article ">Figure 14
<p>Agreement of analytical and experimental time of arrival.</p>
Full article ">
16 pages, 5561 KiB  
Article
A Hybrid GAN-Inception Deep Learning Approach for Enhanced Coordinate-Based Acoustic Emission Source Localization
by Xuhui Huang, Ming Han and Yiming Deng
Appl. Sci. 2024, 14(19), 8811; https://doi.org/10.3390/app14198811 - 30 Sep 2024
Viewed by 642
Abstract
In this paper, we propose a novel approach to coordinate-based acoustic emission (AE) source localization to address the challenges of limited and imbalanced datasets from fiber-optic AE sensors used for structural health monitoring (SHM). We have developed a hybrid deep learning model combining [...] Read more.
In this paper, we propose a novel approach to coordinate-based acoustic emission (AE) source localization to address the challenges of limited and imbalanced datasets from fiber-optic AE sensors used for structural health monitoring (SHM). We have developed a hybrid deep learning model combining four generative adversarial network (GAN) variants for data augmentation with an adapted inception neural network for regression-based prediction. The experimental setup features a single fiber-optic AE sensor based on a tightly coiled fiber-optic Fabry-Perot interferometer formed by two identical fiber Bragg gratings. AE signals were generated using the Hsu-Nielsen pencil lead break test on a grid-marked thin aluminum plate with 35 distinct locations, simulating real-world structural monitoring conditions in bounded isotropic plate-like structures. It is demonstrated that the single-sensor configuration can achieve precise localization, avoiding the need for a multiple sensor array. The GAN-based signal augmentation expanded the dataset from 900 to 4500 samples, with the Wasserstein distance between the original and synthetic datasets decreasing by 83% after 2000 training epochs, demonstrating the high fidelity of the synthetic data. Among the GAN variants, the standard GAN architecture proved the most effective, outperforming other variants in this specific application. The hybrid model exhibits superior performance compared to non-augmented deep learning approaches, with the median error distribution comparisons revealing a significant 50% reduction in prediction errors, accompanied by substantially improved consistency across various AE source locations. Overall, this developed hybrid approach offers a promising solution for enhancing AE-based SHM in complex infrastructures, improving damage detection accuracy and reliability for more efficient predictive maintenance strategies. Full article
(This article belongs to the Special Issue Advanced Optical-Fiber-Related Technologies)
Show Figures

Figure 1

Figure 1
<p>Schematic of the fiber-optic coil-based acoustic emission sensing system. Inset: Close-up image of the sensor, showing the flexible mounting and dimensions (8 mm outer, 6 mm inner diameter).</p>
Full article ">Figure 2
<p>(<b>a</b>) Aluminum plate with the grid and fiber-optic sensor for AE testing (<b>b</b>) Schematic representation of the aluminum plate detailing the grid layout and test points.</p>
Full article ">Figure 3
<p>Time series augmentation showing the original data (orange) and generated data (green) to ensure each label has a balanced and sufficient number of samples for improved deep learning model performance.</p>
Full article ">Figure 4
<p>(<b>a</b>) Workflow of the hybrid network for AE source localization (<b>b</b>) Architecture of the Inception network for regression.</p>
Full article ">Figure 5
<p>Architecture of the generator and discriminator networks in the GAN for AE signal augmentation.</p>
Full article ">Figure 6
<p>The t-SNE visualization of synthetic and original datasets (<b>a</b>) The training epoch of 1 for GAN (<b>b</b>) The training epoch of 2000 for GAN (<b>c</b>) The training epoch of 2000 for WGAN (<b>d</b>) The training epoch of 2000 for DCGAN (<b>e</b>) The training epoch of 2000 for TSAGAN (<b>f</b>) Augmentation via addition of noise.</p>
Full article ">Figure 7
<p>The comparison of Wasserstein distance convergence across epochs for the four GAN variants (GAN, TSAGAN, WGAN, and DCGAN).</p>
Full article ">Figure 8
<p>The comparison of acoustic emission (AE) source localization performance. (<b>a</b>) Results from the hybrid deep learning model with GAN-based data augmentation and Inception network. (<b>b</b>) Results from the Inception network alone without GAN-based augmentation. Square markers represent actual source locations, star markers show predicted locations, and the large circular marker indicates the sensor position. The x and y axes represent dimensions in inches.</p>
Full article ">Figure 9
<p>The comparison of errors for the different methods.</p>
Full article ">
11 pages, 2269 KiB  
Article
FBG Interrogator Using a Dispersive Waveguide Chip and a CMOS Camera
by Zhenming Ding, Qing Chang, Zeyu Deng, Shijie Ke, Xinhong Jiang and Ziyang Zhang
Micromachines 2024, 15(10), 1206; https://doi.org/10.3390/mi15101206 - 29 Sep 2024
Viewed by 569
Abstract
Optical sensors using fiber Bragg gratings (FBGs) have become an alternative to traditional electronic sensors thanks to their immunity against electromagnetic interference, their applicability in harsh environments, and other advantages. However, the complexity and high cost of the FBG interrogation systems pose a [...] Read more.
Optical sensors using fiber Bragg gratings (FBGs) have become an alternative to traditional electronic sensors thanks to their immunity against electromagnetic interference, their applicability in harsh environments, and other advantages. However, the complexity and high cost of the FBG interrogation systems pose a challenge for the wide deployment of such sensors. Herein, we present a clean and cost-effective method for interrogating an FBG temperature sensor using a micro-chip called the waveguide spectral lens (WSL) and a standard CMOS camera. This interrogation system can project the FBG transmission spectrum onto the camera without any free-space optical components. Based on this system, an FBG temperature sensor is developed, and the results show good agreement with a commercial optical spectrum analyzer (OSA), with the respective wavelength-temperature sensitivity measured as 6.33 pm/°C for the WSL camera system and 6.32 pm/°C for the commercial OSA. Direct data processing on the WSL camera system translates this sensitivity to 0.44 μm/°C in relation to the absolute spatial shift of the FBG spectra on the camera. Furthermore, a deep neural network is developed to train the spectral dataset, achieving a temperature resolution of 0.1 °C from 60 °C to 120 °C, while direct processing on the valley/dark line detection yields a resolution of 7.84 °C. The proposed hardware and the data processing method may lead to the development of a compact, practical, and low-cost FBG interrogator. Full article
(This article belongs to the Special Issue Fiber Optic Sensing Technology: From Materials to Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the FBG interrogation system based on a WSL. SLED: superluminescent light emitting diode; ISO: isolator; FBG: fiber Bragg grating; WSL: waveguide spectral lens; CMOS: complementary metal oxide semiconductor.</p>
Full article ">Figure 2
<p>Schematic of the WSL. Inset (I) shows the cross section of the waveguide. Inset (II) shows the waveguides with output tapers designed for reducing the free-space diffraction loss. BBA: beam broadening area.</p>
Full article ">Figure 3
<p>(<b>a</b>) Photo of the fabricated devices and a matchstick for size comparison. Bottom WSL is used for the FBG interrogation. Wavelength calibration of WSL: (<b>b</b>) captured spectral lines at 779.092 nm, 802.528 nm, and 824.616 nm, respectively. (<b>c</b>) Normalized intensity distributions of the spectral lines. (<b>d</b>) Wavelength calibration result.</p>
Full article ">Figure 4
<p>(<b>a</b>) Comparison between Δλ<sub>B</sub> of the sensing FBG obtained from the OSA and from the WSL-based interrogator. (<b>b</b>) The relation between the spectral line shift and temperature.</p>
Full article ">Figure 5
<p>(<b>a</b>) Image preprocessing and architecture of the neural network. (<b>b</b>) The loss variation in DNN training process. MSE: mean squared error. (<b>c</b>) Scatter plot of actual temperature vs. temperature predicted by DNN.</p>
Full article ">
15 pages, 5338 KiB  
Article
Research on the Fabrication and Parameters of a Flexible Fiber Optic Pressure Sensor with High Sensitivity
by Huixin Zhang, Jing Wu and Chencheng Gao
Photonics 2024, 11(10), 919; https://doi.org/10.3390/photonics11100919 - 28 Sep 2024
Viewed by 602
Abstract
In recent years, flexible pressure sensors have garnered significant attention. However, the development of large-area, low-cost, and easily fabricated flexible pressure sensors remains challenging. We designed a flexible fiber optic pressure sensor for contact force detection based on the principle of backward Rayleigh [...] Read more.
In recent years, flexible pressure sensors have garnered significant attention. However, the development of large-area, low-cost, and easily fabricated flexible pressure sensors remains challenging. We designed a flexible fiber optic pressure sensor for contact force detection based on the principle of backward Rayleigh scattering using a single-mode optical fiber as the sensing element and polymer PDMS as the encapsulation material. To enhance the sensor’s sensitivity and stability, we optimized its structural design, parameters, and fabrication process and measured the fiber strain using an optical frequency domain reflectometer (OFDR). The results showed that the sensor achieved a high sensitivity of 6.93247 με/kPa with a PDMS concentration ratio of 10:1, a curing time of 2 h, and a substrate thickness of 5 mm. The sensor demonstrated excellent linearity and repeatability in static performance tests and was successfully used to monitor the plantar pressure distribution in real time. This flexible fiber optic pressure sensor can be developed via a simple fabrication process, has a low cost, and has high sensitivity, highlighting its potential applications in smart wearables and medical diagnostics. Full article
(This article belongs to the Special Issue Optical Sensors and Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Three-dimensional structure of the sensor; (<b>b</b>) physical drawing of the sensor.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic diagram of the fiber optic structure; (<b>b</b>) physical drawing of the G.65A72.</p>
Full article ">Figure 3
<p>Schematic diagram of the sensor fabrication process.</p>
Full article ">Figure 4
<p>Working principle of OFDR.</p>
Full article ">Figure 5
<p>Test system.</p>
Full article ">Figure 6
<p>Stress–strain diagrams for PDMS substrates of different thicknesses.</p>
Full article ">Figure 7
<p>Stress–strain diagrams for different PDMS ratios.</p>
Full article ">Figure 8
<p>Stress–strain diagrams for different curing times.</p>
Full article ">Figure 9
<p>Relationship between the strain and pressure response of four-cycle loading.</p>
Full article ">Figure 10
<p>Relationship between the strain and pressure response of four loading/unloading cycles. (<b>a</b>) First loading/unloading cycle; (<b>b</b>) second loading/unloading cycle; (<b>c</b>) third loading/unloading cycle; (<b>d</b>) fourth loading/unloading cycle.</p>
Full article ">Figure 11
<p>(<b>a</b>) Pressure zoning of the soles of the feet; (<b>b</b>) insole-type fiber optic sensor structure.</p>
Full article ">Figure 12
<p>Test process.</p>
Full article ">Figure 13
<p>(<b>a</b>) Strain diagram of the output at static stand; (<b>b</b>) cloud view of distribution of plantar pressure during static standing.</p>
Full article ">Figure 14
<p>Gait analysis during walking: (<b>a</b>) heel on the ground; (<b>b</b>) full foot on the ground; (<b>c</b>) foot planted on the ground.</p>
Full article ">Figure 15
<p>The distribution of plantar pressure during walking: (<b>a</b>) heel on the ground; (<b>b</b>) full foot on the ground; (<b>c</b>) foot planted on the ground.</p>
Full article ">
17 pages, 6355 KiB  
Article
Strain Sensing in Cantilever Beams Using a Tapered PMF with Embedded Optical Modulation Region
by Xiaopeng Han, Xiaobin Bi, Yundong Zhang, Fan Wang, Siyu Lin, Wuliji Hasi, Chen Wang and Xueheng Yan
Photonics 2024, 11(10), 911; https://doi.org/10.3390/photonics11100911 - 27 Sep 2024
Viewed by 469
Abstract
This paper presents the design of a strain-sensitive, dual ball-shaped tunable zone (DBT) taper structure for light intensity modulation. Unlike conventional tapered optical fibers, the DBT incorporates a central light field modulation zone within the taper. By precisely controlling the fusion parameters between [...] Read more.
This paper presents the design of a strain-sensitive, dual ball-shaped tunable zone (DBT) taper structure for light intensity modulation. Unlike conventional tapered optical fibers, the DBT incorporates a central light field modulation zone within the taper. By precisely controlling the fusion parameters between single-mode fiber (SMF) and polarization-maintaining fiber (PMF), the ellipticity of the modulation zone can be finely adjusted, thereby optimizing spectral characteristics. Theoretical analysis based on polarization mode interference (PMI) coupling confirms that the DBT structure achieves a more uniform spectral response. In cantilever beam strain tests, the DBT exhibits high sensitivity and a highly linear intensity–strain response (R² = 0.99), with orthogonal linear polarization mode interference yielding sensitivities of 0.049 dB/με and 0.023 dB/με over the 0–244.33 με strain range. Leveraging the DBT’s light intensity sensitivity, a temperature-compensated intensity difference and ratio calculation method is proposed, effectively minimizing the influence of light source fluctuations on sensor performance and enabling high-precision strain measurements with errors as low as ±6 με under minor temperature variations. The DBT fiber device, combined with this innovative demodulation technique, is particularly suitable for precision optical sensing applications. The DBT structure, combined with the novel demodulation method, is particularly well-suited for high-precision and stable measurements in industrial monitoring, aerospace, civil engineering, and precision instruments for micro-deformation sensing. Full article
(This article belongs to the Special Issue Advances in Optical Fiber Sensing Technology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the DBT structure.</p>
Full article ">Figure 2
<p>Optical field transmission simulations: (<b>a</b>) DT structure and (<b>b</b>) DBT structure. (<b>c</b>) Comparative analysis of fiber mode purity. (<b>d</b>) Analysis of the relationship between modulation region geometry and fiber mode purity.</p>
Full article ">Figure 3
<p>(<b>a</b>) DBT structure fabrication process. (<b>b</b>) CCD microscopic image of the DBT structure.</p>
Full article ">Figure 4
<p>Spectral comparison and CCD images of DT and DBT structures with different modulation region shapes.</p>
Full article ">Figure 5
<p>(<b>a</b>) FFT spectral analysis and mode interference analysis of the DBT-S3 structure. (<b>b</b>) IFFT spectral analysis of the DBT-S3 structure.</p>
Full article ">Figure 6
<p>Optical fiber cantilever beam strain measurement system.</p>
Full article ">Figure 7
<p>(<b>a</b>) Strain response of the DBT-S3 structure spectrum with spherical modulation regions. (<b>b</b>) Strain response of the polarization resonance dip in the x-p. (<b>c</b>) Strain response of the polarization resonance dip in the y-p. (<b>d</b>) Statistical analysis of the strain sensing response at the x and y polarization resonance dips.</p>
Full article ">Figure 8
<p>(<b>a</b>) Strain response of the spectrum for the DBT structure with vertical ellipsoidal modulation regions. (<b>b</b>) Strain response of the spectrum for the DBT structure with oblate spheroidal modulation regions. (<b>c</b>) Strain response of the spectrum for the standard concave conical DT structure. (<b>d</b>) Statistical analysis of the intensitystrain sensing response for various parameter structures.</p>
Full article ">Figure 9
<p>(<b>a</b>) Temperature response of the spectrum for the DBT structure with spherical modulation regions. (<b>b</b>) Statistical analysis of the temperature-sensing response at the polarization resonance dip.</p>
Full article ">Figure 10
<p>Schematic diagram of the temperature-compensated polarization-based intensity difference and ratio calculation framework.</p>
Full article ">Figure 11
<p>Statistical analysis of the light intensity and strain sensing after temperature compensation.</p>
Full article ">Figure 12
<p>Comparison of demodulation calculation calibration errors: (<b>a</b>) traditional feature point tracking with linear fittingand (<b>b</b>) temperature-compensated polarization-based intensity difference and ratio.</p>
Full article ">
15 pages, 3150 KiB  
Article
Research on the Conversion Coefficient in Coherent Φ-OTDR and Its Intrinsic Impact on Localization Accuracy
by Zhen Zhong, Ningmu Zou and Xuping Zhang
Photonics 2024, 11(10), 901; https://doi.org/10.3390/photonics11100901 - 25 Sep 2024
Viewed by 423
Abstract
Phase-sensitive optical time domain reflectometry (Φ-OTDR) plays a crucial role in localizing and monitoring seismic waves, underwater structures, etc. Accurate localization of external perturbations along the fiber is essential for addressing these challenges effectively. The conversion coefficient, which links the detected phase signal [...] Read more.
Phase-sensitive optical time domain reflectometry (Φ-OTDR) plays a crucial role in localizing and monitoring seismic waves, underwater structures, etc. Accurate localization of external perturbations along the fiber is essential for addressing these challenges effectively. The conversion coefficient, which links the detected phase signal to the perturbation signal on the fiber, has a significant impact on localization accuracy. This makes the characteristic of parameters relative to the conversion coefficient in Φ-OTDR a subject of deep research. Based on the coherent Φ-OTDR mathematical model, parameters like the modulus, the statistical phase, the phase change, and the peak difference are analyzed with and without the static region, respectively. When perturbations are homogeneously distributed along the fiber, the absence of static region on the phase change-fiber length plane leads to a nonlinear phase change relationship. This deviation from the expected linear relationship in the presence of static region means that the static region is essential for higher localization accuracy. The absence of static region results in a standard deviation of 0.042263 m for the localization deviation value, which could be theoretically reduced by a new sensor design with a static region. These findings underscore the importance of the conversion coefficient and the relevance of the static region in Φ-OTDR to achieving accurate and effective localization. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of traditional coherent Φ-OTDR.</p>
Full article ">Figure 2
<p>Simulation result in coherent Φ-OTDR with the static region (<b>a</b>), the modulus (<b>b</b>), the statistical phase (<b>c</b>), the phase change and (<b>d</b>) the peak difference.</p>
Full article ">Figure 3
<p>(<b>a</b>) Differential value for spatial resolution; relationship between final peak difference and (<b>b</b>) vibration amplitude/(<b>c</b>) vibration length/(<b>d</b>) position of the fiber.</p>
Full article ">Figure 4
<p>(<b>a</b>) The modulus, (<b>b</b>) the statistical phase, (<b>c</b>) the phase change and (<b>d</b>) the peak difference when the external perturbation acts equally on every position of the fiber.</p>
Full article ">Figure 5
<p>Simulation results: (<b>a</b>) the phase change at three different positions; experimental results: (<b>b</b>) the phase change, (<b>c</b>) the local magnification of (<b>b</b>), and (<b>d</b>) the phase change at three different positions.</p>
Full article ">Figure 6
<p>(<b>a</b>) One peak difference-fiber length plot when the perturbation acts on all positions of fiber in traditional coherent Φ-OTDR; (<b>b</b>) localization results for 50 different peak difference-fiber length plots.</p>
Full article ">Figure 7
<p>Scheme of a new fiber sensor using coherent Φ-OTDR with shield regions of fiber and fiber rings.</p>
Full article ">
13 pages, 2132 KiB  
Article
Design and Simulation of High-Performance D-Type Dual-Mode PCF-SPR Refractive Index Sensor Coated with Au-TiO2 Layer
by Xin Ding, Qiao Lin, Mengjie Wang, Shen Liu, Weiguan Zhang, Nan Chen and Yiping Wang
Sensors 2024, 24(18), 6118; https://doi.org/10.3390/s24186118 - 22 Sep 2024
Viewed by 548
Abstract
A novel surface plasmon resonance (SPR) refractive index (RI) sensor based on the D-type dual-mode photonic crystal fiber (PCF) is proposed. The sensor employs a side-polished few-mode PCF that facilitates the transmission of the fundamental and second-order modes, with an integrated microfluidic channel [...] Read more.
A novel surface plasmon resonance (SPR) refractive index (RI) sensor based on the D-type dual-mode photonic crystal fiber (PCF) is proposed. The sensor employs a side-polished few-mode PCF that facilitates the transmission of the fundamental and second-order modes, with an integrated microfluidic channel positioned directly above the fiber core. This design minimizes the distance to the analyte and maximizes the interaction between the optical field and the analyte, thereby enhancing the SPR effect and resonance loss for improved sensing performance. Au-TiO2 dual-layer material was coated on the surface of a microfluidic channel to enhance the penetration depth of the core evanescent field and tune the resonance wavelength to the near-infrared band, meeting the special needs of chemical and biomedical detection fields. The finite element method was utilized to systematically investigate the coupling characteristics between various modes and surface plasmon polariton (SPP) modes, as well as the impact of structural parameters on the sensor performance. The results indicate that the LP11b_y mode exhibits greater wavelength sensitivity than the HE11_y mode, with a maximum sensitivity of 33,000 nm/RIU and an average sensitivity of 8272.7 nm/RIU in the RI sensing range of 1.25–1.36, which is higher than the maximum sensitivity of 16,000 nm/RIU and average sensitivity of 5666.7 nm/RIU for the HE11b_y mode. It is believed that the proposed PCF-SPR sensor features both high sensitivity and high resolution, which will become a critical device for wide RI detection in mid-infrared fields. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Dual-mode PCF physical image; (<b>b</b>) Cross-section of the proposed D-type dual-mode PCF-SPR structure.</p>
Full article ">Figure 2
<p>Core mode field distribution at an operating wavelength <math display="inline"><semantics> <mrow> <mi>λ</mi> </mrow> </semantics></math> = 1.4 μm: (<b>a</b>) HE<sub>11_y</sub> mode; (<b>b</b>) HE<sub>11_x</sub> mode; (<b>c</b>) LP<sub>11a_y</sub> mode; (<b>d</b>) LP<sub>11a_x</sub> mode; (<b>e</b>) LP<sub>11b_y</sub> mode; (<b>f</b>) LP<sub>11b_x</sub> mode. (arrow indicates the electric field direction; color legend refers to the electric filed intensity).</p>
Full article ">Figure 3
<p>Dispersion relationship between the y polarization core modes and the corresponding SPP modes when the RI of measured analyte is 1.34.</p>
Full article ">Figure 4
<p>Mode field distribution when phase matching occurs between the core mode and the SPP mode: (<b>a</b>) HE<sub>11_y</sub> mode; (<b>b</b>) LP<sub>11a_y</sub> mode; (<b>c</b>) LP<sub>11b_y</sub> mode. (arrow indicates the electric field direction; color legend refers to the electric field intensity).</p>
Full article ">Figure 5
<p>Influence of large air hole diameter d<sub>3</sub> in the middle region of the PCF on the loss spectrum.</p>
Full article ">Figure 6
<p>Influence of the gold film thickness t<sub>1</sub> on the loss spectrum.</p>
Full article ">Figure 7
<p>Influence of the TiO<sub>2</sub> thickness t<sub>2</sub> on the loss spectrum.</p>
Full article ">Figure 8
<p>Influence of the microgroove depth d<sub>p</sub> on the loss spectrum.</p>
Full article ">Figure 9
<p>Response to RI under optimized parameters: (<b>a</b>) HE<sub>11_y</sub> mode; (<b>b</b>) LP<sub>11a_y</sub> mode.</p>
Full article ">Figure 10
<p>(<b>a</b>) Relation between the RI and resonance wavelength; (<b>b</b>) <span class="html-italic">FWHM</span> and <span class="html-italic">FOM</span>.</p>
Full article ">
18 pages, 12673 KiB  
Article
Analysis of Field Trial Results for Excavation-Activities Monitoring with φ-OTDR
by Hailiang Zhang, Hui Dong, Dora Juan Juan Hu, Nhu Khue Vuong, Lianlian Jiang, Gen Liang Lim and Jun Hong Ng
Sensors 2024, 24(18), 6081; https://doi.org/10.3390/s24186081 - 20 Sep 2024
Viewed by 515
Abstract
Underground telecommunication cables are highly susceptible to damage from excavation activities. Preventing accidental damage to underground telecommunication cables is critical and necessary. In this study, we present field trial results of monitoring excavation activities near underground fiber cables using an intensity-based phase-sensitive optical [...] Read more.
Underground telecommunication cables are highly susceptible to damage from excavation activities. Preventing accidental damage to underground telecommunication cables is critical and necessary. In this study, we present field trial results of monitoring excavation activities near underground fiber cables using an intensity-based phase-sensitive optical time-domain reflectometer (φ-OTDR). The reasons for choosing intensity-based φ-OTDR for excavation monitoring are presented and analyzed. The vibration signals generated by four typical individual excavation events, i.e., cutting, hammering, digging, and tamping at five different field trial sites, as well as five different mixed events in the fifth field trial site were investigated. The findings indicate that various types of events can generate vibration signals with different features. Typically, fundamental peak frequencies of cutting, hammering and tamping events ranged from 30 to 40 Hz, 11 to 15 Hz, and 30 to 40 Hz, respectively. Digging events, on the other hand, presented a broadband frequency spectrum without a distinct peak frequency. Moreover, due to differences in environmental conditions, even identical excavation events conducted with the same machine may also generate vibration signals with different characteristics. The diverse field trial results presented offer valuable insights for both research and the practical implementation of excavation monitoring techniques for underground cables. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the in-house developed φ-OTDR interrogator; (<b>b</b>) Example of a waterfall plot for 21 km fiber over about 5 min; (<b>c</b>) φ-OTDR interrogator was installed in a sever rack; (<b>d</b>) On-site photo of underground telecommunication fiber in the manhole; (<b>e</b>) Alarm of excavation event in GUI, the green curve represents the monitored fiber route. AOM: acousto-optic modulator, CIR: circulator, PD: photodetector, DAQ: data acquisition card, PC: computer.</p>
Full article ">Figure 2
<p>On-site photos of the five field trials for excavation detection. The yellow curves present the routes of underground fiber cables. Excavation activities were located in the areas marked by red dashed boxes.</p>
Full article ">Figure 3
<p>Field trial of cutting event: (<b>a</b>) at Site I, (<b>b</b>) at Site II, (<b>c</b>) at Site IV; (<b>d</b>) at Site V. (<b>i</b>), (<b>ii</b>), (<b>iii</b>) and (<b>iv</b>) show the cutting event scenes, cutting event signals in time domain, FFT results of the 60 s signals, time-frequency diagrams of the 60 s signal, respectively.</p>
Full article ">Figure 4
<p>Field trial of hammering event: (<b>a</b>) at Site I, (<b>b</b>) at Site II, (<b>c</b>) at Site III, (<b>d</b>) at Site IV; (<b>e</b>) at Site V. (<b>i</b>), (<b>ii</b>), (<b>iii</b>) and (<b>iv</b>) show the hammering event scenes, hammering event signals in time domain, FFT results of the 60 s signal, time-frequency diagrams of the 60 s signal, respectively.</p>
Full article ">Figure 5
<p>Field trials of digging event: (<b>a</b>) at Site I, (<b>b</b>) at Site II, (<b>c</b>) at Site III, (<b>d</b>) at Site IV; (<b>e</b>) at Site V. (<b>i</b>), (<b>ii</b>), (<b>iii</b>) and (<b>iv</b>) show the digging event scenes, digging event signals in time domain, FFT results of the 60 s signal, time-frequency diagrams of the 60 s signal, respectively.</p>
Full article ">Figure 6
<p>Field trial of tamping events: (<b>a</b>) at Site II, (<b>b</b>) at Site III, (<b>c</b>) at Site IV; (<b>d</b>) at Site V. (<b>i</b>), (<b>ii</b>), (<b>iii</b>) and (<b>iv</b>) show the tamping event scenes, tamping event signals in time domain, FFT results of the 60 s signal, time-frequency diagrams of the 60 s signal, respectively.</p>
Full article ">Figure 7
<p>On-site photos of mixed events at Site V: (<b>a</b>) cutting and hammering, (<b>b</b>) cutting and digging, (<b>c</b>) hammering and digging, (<b>d</b>) digging and tamping, (<b>e</b>) digging and digging.</p>
Full article ">Figure 8
<p>Results of mixed events at Site V: (<b>a</b>) cutting and hammering, (<b>b</b>) cutting and digging, (<b>c</b>) hammering and digging, (<b>d</b>) digging and tamping, (<b>e</b>) digging and digging. (<b>i</b>), (<b>ii</b>), and (<b>iii</b>) show the signal in time domain, FFT result of the 60 s signal, time-frequency diagrams of the 60 s signal, respectively.</p>
Full article ">
Back to TopTop