[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (650)

Search Parameters:
Keywords = GFRP

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 7879 KiB  
Article
Mechanical Properties of Concrete Mixes with Selectively Crushed Wind Turbine Blade: Comparison with Raw-Crushing
by Víctor Revilla-Cuesta, Ana B. Espinosa, Roberto Serrano-López, Marta Skaf and Juan M. Manso
Materials 2024, 17(24), 6299; https://doi.org/10.3390/ma17246299 - 23 Dec 2024
Viewed by 311
Abstract
The glass fiber-reinforced polymer (GFRP) materials of wind turbine blades can be recovered and recycled by crushing, thereby solving one of the most perplexing problems facing the wind energy sector. This process yields selectively crushed wind turbine blade (SCWTB), a novel waste that [...] Read more.
The glass fiber-reinforced polymer (GFRP) materials of wind turbine blades can be recovered and recycled by crushing, thereby solving one of the most perplexing problems facing the wind energy sector. This process yields selectively crushed wind turbine blade (SCWTB), a novel waste that is almost exclusively composed of GFRP composite fibers that can be revalued in terms of their use as a raw material in concrete production. In this research, the fresh and mechanical performance of concrete made with 1.5%, 3.0%, 4.5%, and 6.0% SCWTB is studied. Once incorporated into concrete mixes, SCWTB waste slightly reduced slumps due to the large specific surface area of the fibers, and the stitching effect of the fibers on mechanical behavior was generally adequate, as scanning electron microscopy demonstrated good fiber adhesion within the cementitious matrix. Thus, despite the increase in the content of water and plasticizers when adding this waste to preserve workability, the compressive strength only decreased in the long term with the addition of 6.0% SCWTB, a value of 45 MPa always being reached at 28 days; Poisson’s coefficient remained constant from 3.0% SCWTB; splitting tensile strength was maintained at around 4.7 MPa up to additions of 3.0% SCWTB; and the flexural strength of mixes containing 6.0% and 1.5% SCWTB was statistically equal, with a value near 6.1 MPa. Furthermore, all mechanical properties of the concrete except for flexural strength were improved with additions of SCWTB compared to raw crushed wind turbine blade, which apart from GFRP composite fibers contains approximately spherical polymer and balsa wood particles. Flexural strength was conditioned by the proportion of fibers, their dimensions, and their strength, which were almost identical for both waste types. SCWTB would be preferable for applications in which compression stresses predominate. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Selectively crushed wind turbine blade: (<b>a</b>) resulting material; (<b>b</b>) fibers; (<b>c</b>) micro-fibers; (<b>d</b>) other components (balsa wood, polyurethane foam, and other polymers).</p>
Full article ">Figure 2
<p>Illustration of some of the tests conducted: (<b>a</b>) slump test; (<b>b</b>) modulus of elasticity and Poisson’s coefficient; (<b>c</b>) UPV; (<b>d</b>) splitting tensile strength.</p>
Full article ">Figure 3
<p>Hardened density of the concrete mixes.</p>
Full article ">Figure 4
<p>Compressive strength of the concrete mixes.</p>
Full article ">Figure 5
<p>Elastic properties of the concrete mixes: (<b>a</b>) modulus of elasticity; (<b>b</b>) Poisson’s coefficient.</p>
Full article ">Figure 6
<p>Ultrasonic pulse velocity of the concrete mixes.</p>
Full article ">Figure 7
<p>The 28-day bending tensile properties of the concrete mixes.</p>
Full article ">Figure 8
<p>Relationship between 28-day compressive strength and (<b>a</b>) 28-day modulus of elasticity; (<b>b</b>) 28-day splitting tensile strength; (<b>c</b>) 28-day flexural strength.</p>
Full article ">Figure 9
<p>Relationship between 28-day ultrasonic pulse velocity and: (<b>a</b>) 28-day compressive strength; (<b>b</b>) 28-day modulus of elasticity; (<b>c</b>) 28-day splitting tensile strength; (<b>d</b>) 28-day flexural strength.</p>
Full article ">Figure 10
<p>Pearson’s correlation matrix (legend: CS, compressive strength; ME, modulus of elasticity; STS, splitting tensile strength; FS, flexural strength; UPV, ultrasonic pulse velocity).</p>
Full article ">Figure 11
<p>SEM analysis of the interfacial transition zone between a GFRP composite fiber and the cementitious matrix in the MS6.0 mix: (<b>a</b>) 30× magnification; (<b>b</b>) 100× magnification; (<b>c</b>) 450× magnification; (<b>d</b>) 1500× magnification.</p>
Full article ">
24 pages, 11911 KiB  
Article
Development of a Modular Sandwich Panel with a Composite Core of Recycled Material for Application in Sustainable Building
by Juan José Valenzuela Expósito, Elena Picazo Camilo and Francisco Antonio Corpas Iglesias
Polymers 2024, 16(24), 3604; https://doi.org/10.3390/polym16243604 - 23 Dec 2024
Viewed by 307
Abstract
In recent years, the construction industry has faced challenges related to rising material costs, labor shortages and environmental sustainability, resulting in an increased interest in modular construction cores composed of recycled materials, such as XPS, PUR, PLW and GFRP, from waste from the [...] Read more.
In recent years, the construction industry has faced challenges related to rising material costs, labor shortages and environmental sustainability, resulting in an increased interest in modular construction cores composed of recycled materials, such as XPS, PUR, PLW and GFRP, from waste from the truck body industry. Two resins, PUR and polyester, were used to bond these recycled composites. Physical, chemical and mechanical analyses showed that the panels formed with PUR resin had superior workability due to the higher open time of the resin, 11.3% better thermal conductivity than the commercial PLW panel (SP-PLW) and reduced porosity compared to those using polyester resin. The mechanical performance of the panels improved with higher structural reinforcement content (PLW and GFRP). Compared to a commercial panel (SP-PLW), the SP-RCM1 recycled panel showed 4% higher performance, demonstrating its potential for sustainable building applications. Thermal and microscopic characterizations showed good adhesion of the materials in the best performing formulations related to higher thermal stability. Therefore, this research aims to demonstrate the feasibility of using waste from the car industry in the manufacture of sandwich panels for modular construction to address these issues. Full article
(This article belongs to the Section Circular and Green Polymer Science)
Show Figures

Figure 1

Figure 1
<p>Granulometric distribution of GFRP, PLW, PUR and XPS.</p>
Full article ">Figure 2
<p>FTIR patterns: XPS, PUR, PLW and GFRP.</p>
Full article ">Figure 3
<p>TG-DSC: (<b>a</b>) XPS, (<b>b</b>) PUR, (<b>c</b>) PLW and (<b>d</b>) GFRP.</p>
Full article ">Figure 4
<p>SP-RCM manufacturing process (A: resin, B: hardening agent).</p>
Full article ">Figure 5
<p>Dimensional stability (%) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 6
<p>Absorption (%) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 7
<p>Porosity (%) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 8
<p>Nanocomputed images (1.95 µm) of SP-RCM1 and SP-RCM6.</p>
Full article ">Figure 9
<p>Density (kg/m<sup>3</sup>) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 10
<p>Thermal conductivity (W/mK) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 11
<p>FTIR patterns: SP-RCM1–SP-RCM5.</p>
Full article ">Figure 12
<p>FTIR patterns: SP-RCM6–SP-RCM10.</p>
Full article ">Figure 13
<p>Compressive strength (MPa) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 14
<p>Flexural strength (MPa) of SP-PLW and SP-RCM panels.</p>
Full article ">Figure 15
<p>Curves TG-DSC: (<b>a</b>) SP-RCM1 and (<b>b</b>) SP-RCM6.</p>
Full article ">Figure 16
<p>SEM images of SP-PLW and SP-RCM panels at 250X and 700X.</p>
Full article ">Figure 16 Cont.
<p>SEM images of SP-PLW and SP-RCM panels at 250X and 700X.</p>
Full article ">Figure 16 Cont.
<p>SEM images of SP-PLW and SP-RCM panels at 250X and 700X.</p>
Full article ">Figure 17
<p>EDX analysis of SP-RCM1.</p>
Full article ">Figure 18
<p>EDX analysis of SP-RCM6.</p>
Full article ">
23 pages, 4809 KiB  
Article
Experimental Determination of Circumferential Mechanical Properties of GFRP Pipes Using the Split-Disk Method: Evaluating the Impact of Aggressive Environments
by Maria Tănase, Alin Diniță, Gennadiy Lvov and Alexandra Ileana Portoacă
Appl. Sci. 2024, 14(24), 11845; https://doi.org/10.3390/app142411845 - 18 Dec 2024
Viewed by 432
Abstract
This study investigates the mechanical properties of Glass Fiber-Reinforced Plastic (GFRP) pipes in the circumferential direction using the split-disk method, with a focus on understanding the influence of aggressive environmental conditions. The split-disk method was used to determine the key mechanical properties, including [...] Read more.
This study investigates the mechanical properties of Glass Fiber-Reinforced Plastic (GFRP) pipes in the circumferential direction using the split-disk method, with a focus on understanding the influence of aggressive environmental conditions. The split-disk method was used to determine the key mechanical properties, including the hoop tensile strength and modulus, and also the Poisson’s ratio, which are critical for the performance of GFRP pipes under internal pressure. The experiment was conducted under controlled laboratory conditions, and the results were analyzed to assess the effects of exposure to aggressive environments (saltwater and alkaline solutions at 20 °C and 50 °C). The correlations between the UTS, elastic modulus, and Poisson’s ratio highlight how GFRP pipes degrade under environmental exposure. As the UTS decreases, so do the stiffness and lateral deformability, with the most significant reductions occurring in chemically aggressive environments at high temperatures. Exposure to an alkaline solution weakens the GFRP pipes, with the strength dropping more sharply at higher temperatures, with the UTS decreasing by 21%. Saltwater exposure reduces the elastic modulus, especially at higher temperatures, with a 14% decrease, accelerating material degradation and reducing deformation resistance. An alkaline solution further lowers the modulus, with a 21% decrease at 50 °C, showing the lowest stiffness. Air exposure, in contrast, has a less severe effect, with the pipes retaining much of their mechanical integrity. These findings collectively suggest that environmental degradation affects the overall mechanical behavior of GFRP pipes, providing valuable insights for the design and maintenance of GFRP piping systems, particularly in industries where exposure to aggressive environments is common. This study underscores the importance of considering environmental factors in the material selection and design processes to ensure the long-term reliability of GRP pipes. Full article
(This article belongs to the Section Civil Engineering)
Show Figures

Figure 1

Figure 1
<p>Split-disk test specimens: (<b>a</b>) tested samples; (<b>b</b>) dimensional characteristics.</p>
Full article ">Figure 2
<p>Samples immersed at high temperature.</p>
Full article ">Figure 3
<p>Split-disk tensile test system.</p>
Full article ">Figure 4
<p>Circumferential tensile testing.</p>
Full article ">Figure 5
<p>Strain gauge measurements: (<b>a</b>) strain gauge connection; (<b>b</b>) data record.</p>
Full article ">Figure 6
<p>The geometrical model considered in finite element analysis.</p>
Full article ">Figure 7
<p>The finite element model.</p>
Full article ">Figure 8
<p>Specimens after failure.</p>
Full article ">Figure 9
<p>Experimental results: (<b>a</b>) circumferential ultimate tensile strength (UTS), (<b>b</b>) circumferential modulus (E), (<b>c</b>) Poisson’s ratio (<math display="inline"><semantics> <mrow> <msub> <mi>ν</mi> <mrow> <mi>y</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math>). The uncertainty range is represented as error bars, calculated using the standard deviation of the measured values.</p>
Full article ">Figure 9 Cont.
<p>Experimental results: (<b>a</b>) circumferential ultimate tensile strength (UTS), (<b>b</b>) circumferential modulus (E), (<b>c</b>) Poisson’s ratio (<math display="inline"><semantics> <mrow> <msub> <mi>ν</mi> <mrow> <mi>y</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math>). The uncertainty range is represented as error bars, calculated using the standard deviation of the measured values.</p>
Full article ">Figure 10
<p>The distribution of circumferential (<b>a</b>) and axial (<b>b</b>) strains.</p>
Full article ">Figure 11
<p>The distribution of circumferential (<b>a</b>) and axial (<b>b</b>) stresses.</p>
Full article ">Figure 12
<p>Pareto charts for the following: (<b>a</b>) tensile strength; (<b>b</b>) tensile modulus; (<b>c</b>) Poisson’s ratio (factors: A—temperature °C, B—solution type).</p>
Full article ">Figure 12 Cont.
<p>Pareto charts for the following: (<b>a</b>) tensile strength; (<b>b</b>) tensile modulus; (<b>c</b>) Poisson’s ratio (factors: A—temperature °C, B—solution type).</p>
Full article ">Figure 13
<p>Main effect plots for the following: (<b>a</b>) tensile strength; (<b>b</b>) tensile modulus; (<b>c</b>) Poisson’s ratio.</p>
Full article ">Figure 13 Cont.
<p>Main effect plots for the following: (<b>a</b>) tensile strength; (<b>b</b>) tensile modulus; (<b>c</b>) Poisson’s ratio.</p>
Full article ">Figure 14
<p>Interaction plots for the following: (<b>a</b>) tensile strength; (<b>b</b>) tensile modulus; (<b>c</b>) Poisson’s ratio.</p>
Full article ">
23 pages, 15584 KiB  
Article
Comparison of GFRP (Glass Fiber-Reinforced Polymer) and CFRP (Carbon Fiber-Reinforced Polymer) Composite Adhesive-Bonded Single-Lap Joints Used in Marine Environments
by Gurcan Atakok and Dudu Mertgenc Yoldas
Sustainability 2024, 16(24), 11105; https://doi.org/10.3390/su162411105 - 18 Dec 2024
Viewed by 531
Abstract
Macroscopic structures consisting of two or more materials are called composites. The decreasing reserves of the world’s oil reserve and the environmental pollution of existing energy and production resources made the use of recycling methods inevitable. There are mechanical, thermal, and chemical recycling [...] Read more.
Macroscopic structures consisting of two or more materials are called composites. The decreasing reserves of the world’s oil reserve and the environmental pollution of existing energy and production resources made the use of recycling methods inevitable. There are mechanical, thermal, and chemical recycling methods for the recycling of thermosets among composite materials. The recycling of thermoset composite materials economically saves resources and energy in the production of reinforcement and matrix materials. Due to the superior properties such as hardness, strength, lightness, corrosion resistance, design width, and the flexibility of epoxy/vinylester/polyester fibre formation composite materials combined with thermoset resin at the macro level, environmentally friendly sustainable development is happening with the increasing use of composite materials in many fields such as the maritime sector, space technology, wind energy, the manufacturing of medical devices, robot technology, the chemical industry, electrical electronic technology, the construction and building sector, the automotive sector, the defence industry, the aviation sector, the food and agriculture sector, and sports equipment manufacturing. Bonded joint studies in composite materials have generally been investigated at the level of a single composite material and single joint. The uncertainty of the long-term effects of different composite materials and environmental factors in single-lap bonded joints is an important obstacle in applications. The aim of this study is to investigate the effects of single-lap bonded GFRP (glass fibre-reinforced polymer) and CFRP (carbon fibre-reinforced polymer) specimens on the material at the end of seawater exposure. In this study, 0/90 orientation twill weave seven-ply GFRP and eight-ply CFRP composite materials were used in dry conditions (without seawater soaking) and the hand lay-up method. Seawater was taken from the Aegean Sea, İzmir province (Selçuk/Pamucak), in September at 23.5 °C. This seawater was kept in different containers in seawater for 1 month (30 days), 2 months (60 days), and 3 months (90 days) separately for GFRP and CFRP composite samples. They were cut according to ASTM D5868-01 for single-lap joint connections. Moisture retention percentages and axial impact tests were performed. Three-point bending tests were then performed according to ASTM D790. Damage to the material was examined with a ZEISS GEMINESEM 560 scanning electron microscope (SEM). The SEM was used to observe the interface properties and microstructure of the fracture surfaces of the composite samples by scanning images with a focused electron beam. Damage analysis imaging was performed on CFRP and GFRP specimens after sputtering with a gold compound. Moisture retention rates (%), axial impact tests, and three-point bending test specimens were kept in seawater with a seawater salinity of 3.3–3.7% and a seawater temperature of 23.5 °C for 1, 2, and 3 months. Moisture retention rates (%) are 0.66%, 3.43%, and 4.16% for GFRP single-lap bonded joints in a dry environment and joints kept for 1, 2, and 3 months, respectively. In CFRP single-lap bonded joints, it is 0.57%, 0.86%, and 0.87%, respectively. As a result of axial impact tests, under a 30 J impact energy level, the fracture toughness of GFRP single-lap bonded joints kept in a dry environment and seawater for 1, 2, and 3 months are 4.6%, 9.1%, 14.7%, and 11.23%, respectively. At the 30 J impact energy level, the fracture toughness values of CFRP single-lap bonded joints in a dry environment and in seawater for 1, 2, and 3 months were 4.2%, 5.3%, 6.4%, and 6.1%, respectively. As a result of three-point bending tests, GFRP single-lap joints showed a 5.94%, 8.90%, and 12.98% decrease in Young’s modulus compared to dry joints kept in seawater for 1, 2, and 3 months, respectively. CFRP single-lap joints showed that Young’s modulus decreased by 1.28%, 3.39%, and 3.74% compared to dry joints kept in seawater for 1, 2, and 3 months, respectively. Comparing the GFRP and CFRP specimens formed by a single-lap bonded connection, the moisture retention percentages of GFRP specimens and the amount of energy absorbed in axial impact tests increased with the soaking time in seawater, while Young’s modulus was less in three-point bending tests, indicating that CFRP specimens have better mechanical properties. Full article
Show Figures

Figure 1

Figure 1
<p>Formation of specimens from GFRP (<b>a</b>) and CFRP (<b>b</b>) plates using CNC router (<b>c</b>). Geometry model of test specimens. (<b>d</b>) Preparation of specimens before bonding.</p>
Full article ">Figure 2
<p>CFRP (<b>a</b>) and GFRP (<b>b</b>) coded specimen samples.</p>
Full article ">Figure 3
<p>Axial impacts. (<b>a</b>) Schematic view; (<b>b</b>) impact test device with the specimen installed; (<b>c</b>) specimen subjected to impact.</p>
Full article ">Figure 4
<p>Positioning of the sample in the three-point bending testing machine.</p>
Full article ">Figure 5
<p>Dry environment, 1st, 2nd, and 3rd months. Moisture retention rate of samples for comparison of GFRP and CFRP.</p>
Full article ">Figure 6
<p>GFRP dry environment (<b>a</b>), 1st month (<b>b</b>), 2nd month (<b>c</b>), and 3rd month (<b>d</b>) specimens (axial impact).</p>
Full article ">Figure 7
<p>Absorbed energy amount graph of GFRP samples stored in dry environment and seawater for 1, 2, and 3 months.</p>
Full article ">Figure 8
<p>CFRP dry environment (<b>a</b>), 1st month (<b>b</b>), 2nd month (<b>c</b>), and 3rd month (<b>d</b>) specimens (axial impact).</p>
Full article ">Figure 9
<p>Comparing the absorbed energy of composite samples made of GFRP and CFRP.</p>
Full article ">Figure 10
<p>SEM image of GFRP composite material after axial impact test.</p>
Full article ">Figure 11
<p>SEM image of CFRP composite material after axial impact test.</p>
Full article ">Figure 12
<p>GFRP dry environment (<b>a</b>), 1st month (<b>b</b>), 2nd month (<b>c</b>), and 3rd month (<b>d</b>) specimens (three-point bending test).</p>
Full article ">Figure 12 Cont.
<p>GFRP dry environment (<b>a</b>), 1st month (<b>b</b>), 2nd month (<b>c</b>), and 3rd month (<b>d</b>) specimens (three-point bending test).</p>
Full article ">Figure 13
<p>Stress–strain graph of GFRP specimens in dry environment and soaked in seawater for 1, 2, and 3 months.</p>
Full article ">Figure 14
<p>SEM image of GFRP composite material after three-point bending test.</p>
Full article ">Figure 15
<p>CFRP dry environment (<b>a</b>), 1st month (<b>b</b>), 2nd month (<b>c</b>), and 3rd month (<b>d</b>) specimens (three-point bending test).</p>
Full article ">Figure 16
<p>Bending stress–strain graph of CFRP specimens in dry environment and soaked in seawater for 1, 2, and 3 months.</p>
Full article ">Figure 17
<p>SEM image of CFRP composite material after three-point bending test.</p>
Full article ">
38 pages, 9088 KiB  
Article
PET Granule Replacement for Fine Aggregate in Concrete and FRP-Wrapping Effect: Overview of Experimental Data and Model Development
by Omer Fatih Sancak and Muhammet Zeki Ozyurt
Buildings 2024, 14(12), 4009; https://doi.org/10.3390/buildings14124009 - 17 Dec 2024
Viewed by 847
Abstract
In this study, polyethylene terephthalate (PET) was substituted for 10%, 20%, and 30% of the sand volume in concrete. Compressive, splitting tensile, and flexural strength tests were applied to the concrete samples and stress–strain graphs were obtained. It was observed that PET substitution [...] Read more.
In this study, polyethylene terephthalate (PET) was substituted for 10%, 20%, and 30% of the sand volume in concrete. Compressive, splitting tensile, and flexural strength tests were applied to the concrete samples and stress–strain graphs were obtained. It was observed that PET substitution caused a decrease in the mechanical properties of the concrete. For this reason, the concrete with the best PET substitution rate (10%) was reinforced by wrapping it with carbon fiber-reinforced polymer (CFRP) and glass fiber-reinforced polymer (GFRP), and the same experiments were repeated. It was observed that a 10% PET substitution reduced the strength of the reference concrete by about 6%. However, wrapping the PET-substituted concrete with CFRP and GFRP increased the strength by about 1.9 and 1.5 times, respectively, surpassing that of the reference sample. In addition, this study provides a comprehensive database by bringing together experimental data from studies in which PET was used as a substitute by volume or weight instead of fine aggregate in concrete. The models proposed in this study, along with previous models, were tested for applicability. Similarly, the model suggestions in the literature for fiber-reinforced polymer (FRP)-confined concrete were tested with the experimental data in this study, and their suitability for PET-substituted concrete was discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Grading curves of aggregates.</p>
Full article ">Figure 2
<p>PET granules.</p>
Full article ">Figure 3
<p>FRP types: (<b>a</b>) CFRP, (<b>b</b>) GFRP.</p>
Full article ">Figure 4
<p>Experimental setups: (<b>a</b>) compressive, (<b>b</b>) splitting tensile, (<b>c</b>) flexural.</p>
Full article ">Figure 5
<p>Slump–percentage of PET graph.</p>
Full article ">Figure 6
<p>Density–percentage of PET graph.</p>
Full article ">Figure 7
<p>Stress–strain graphs of samples without FRP.</p>
Full article ">Figure 8
<p>Stress–strain graphs of samples with FRP.</p>
Full article ">Figure 9
<p>Flexural stress–flexural strain graph of samples without FRP.</p>
Full article ">Figure 10
<p>Flexural stress–flexural strain graphs of samples with FRP.</p>
Full article ">Figure 11
<p>Compressive strength ratio–PET (%) graph [<a href="#B11-buildings-14-04009" class="html-bibr">11</a>,<a href="#B12-buildings-14-04009" class="html-bibr">12</a>,<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B16-buildings-14-04009" class="html-bibr">16</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B19-buildings-14-04009" class="html-bibr">19</a>,<a href="#B20-buildings-14-04009" class="html-bibr">20</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B22-buildings-14-04009" class="html-bibr">22</a>,<a href="#B23-buildings-14-04009" class="html-bibr">23</a>,<a href="#B24-buildings-14-04009" class="html-bibr">24</a>,<a href="#B25-buildings-14-04009" class="html-bibr">25</a>,<a href="#B26-buildings-14-04009" class="html-bibr">26</a>,<a href="#B27-buildings-14-04009" class="html-bibr">27</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B29-buildings-14-04009" class="html-bibr">29</a>,<a href="#B30-buildings-14-04009" class="html-bibr">30</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B35-buildings-14-04009" class="html-bibr">35</a>].</p>
Full article ">Figure 12
<p>Comparison of R<sup>2</sup> and MABE for compressive strength models [<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B19-buildings-14-04009" class="html-bibr">19</a>,<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B36-buildings-14-04009" class="html-bibr">36</a>,<a href="#B37-buildings-14-04009" class="html-bibr">37</a>,<a href="#B38-buildings-14-04009" class="html-bibr">38</a>,<a href="#B39-buildings-14-04009" class="html-bibr">39</a>].</p>
Full article ">Figure 13
<p>Tensile strength ratio–PET (%) graph [<a href="#B11-buildings-14-04009" class="html-bibr">11</a>,<a href="#B12-buildings-14-04009" class="html-bibr">12</a>,<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B16-buildings-14-04009" class="html-bibr">16</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B22-buildings-14-04009" class="html-bibr">22</a>,<a href="#B26-buildings-14-04009" class="html-bibr">26</a>,<a href="#B27-buildings-14-04009" class="html-bibr">27</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B35-buildings-14-04009" class="html-bibr">35</a>].</p>
Full article ">Figure 14
<p>Comparison of R<sup>2</sup> and MABE for models of the connection between tensile strength and PET (%) [<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B36-buildings-14-04009" class="html-bibr">36</a>,<a href="#B37-buildings-14-04009" class="html-bibr">37</a>,<a href="#B38-buildings-14-04009" class="html-bibr">38</a>].</p>
Full article ">Figure 15
<p>Tensile strength–compressive strength graph [<a href="#B11-buildings-14-04009" class="html-bibr">11</a>,<a href="#B12-buildings-14-04009" class="html-bibr">12</a>,<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B16-buildings-14-04009" class="html-bibr">16</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B22-buildings-14-04009" class="html-bibr">22</a>,<a href="#B26-buildings-14-04009" class="html-bibr">26</a>,<a href="#B27-buildings-14-04009" class="html-bibr">27</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B35-buildings-14-04009" class="html-bibr">35</a>].</p>
Full article ">Figure 16
<p>Comparison of R<sup>2</sup> and MABE for models of the connection between tensile strength and compressive strength [<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B32-buildings-14-04009" class="html-bibr">32</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B36-buildings-14-04009" class="html-bibr">36</a>,<a href="#B37-buildings-14-04009" class="html-bibr">37</a>,<a href="#B40-buildings-14-04009" class="html-bibr">40</a>].</p>
Full article ">Figure 17
<p>Flexural strength ratio–PET (%) graph [<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B25-buildings-14-04009" class="html-bibr">25</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B29-buildings-14-04009" class="html-bibr">29</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B35-buildings-14-04009" class="html-bibr">35</a>].</p>
Full article ">Figure 18
<p>Flexural strength–compressive strength graph [<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B25-buildings-14-04009" class="html-bibr">25</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B29-buildings-14-04009" class="html-bibr">29</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B35-buildings-14-04009" class="html-bibr">35</a>].</p>
Full article ">Figure 19
<p>Comparison of R<sup>2</sup> and MABE for models of the connection between flexural strength and compressive strength [<a href="#B15-buildings-14-04009" class="html-bibr">15</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B31-buildings-14-04009" class="html-bibr">31</a>,<a href="#B34-buildings-14-04009" class="html-bibr">34</a>,<a href="#B40-buildings-14-04009" class="html-bibr">40</a>,<a href="#B41-buildings-14-04009" class="html-bibr">41</a>].</p>
Full article ">Figure 20
<p>Modulus of elasticity ratio–PET (%) graph [<a href="#B11-buildings-14-04009" class="html-bibr">11</a>,<a href="#B12-buildings-14-04009" class="html-bibr">12</a>,<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B16-buildings-14-04009" class="html-bibr">16</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B20-buildings-14-04009" class="html-bibr">20</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B23-buildings-14-04009" class="html-bibr">23</a>,<a href="#B24-buildings-14-04009" class="html-bibr">24</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>].</p>
Full article ">Figure 21
<p>Modulus of elasticity–compressive strength graph [<a href="#B11-buildings-14-04009" class="html-bibr">11</a>,<a href="#B12-buildings-14-04009" class="html-bibr">12</a>,<a href="#B13-buildings-14-04009" class="html-bibr">13</a>,<a href="#B14-buildings-14-04009" class="html-bibr">14</a>,<a href="#B16-buildings-14-04009" class="html-bibr">16</a>,<a href="#B17-buildings-14-04009" class="html-bibr">17</a>,<a href="#B18-buildings-14-04009" class="html-bibr">18</a>,<a href="#B20-buildings-14-04009" class="html-bibr">20</a>,<a href="#B21-buildings-14-04009" class="html-bibr">21</a>,<a href="#B23-buildings-14-04009" class="html-bibr">23</a>,<a href="#B24-buildings-14-04009" class="html-bibr">24</a>,<a href="#B28-buildings-14-04009" class="html-bibr">28</a>,<a href="#B33-buildings-14-04009" class="html-bibr">33</a>].</p>
Full article ">Figure 22
<p>Comparison of R<sup>2</sup> and MABE for models of the connection between modules of elasticity and compressive strength [<a href="#B40-buildings-14-04009" class="html-bibr">40</a>,<a href="#B41-buildings-14-04009" class="html-bibr">41</a>,<a href="#B42-buildings-14-04009" class="html-bibr">42</a>].</p>
Full article ">Figure 23
<p>Experimental data and model predictions for compressive strength under FRP-wrapping effect [<a href="#B50-buildings-14-04009" class="html-bibr">50</a>,<a href="#B51-buildings-14-04009" class="html-bibr">51</a>,<a href="#B52-buildings-14-04009" class="html-bibr">52</a>,<a href="#B53-buildings-14-04009" class="html-bibr">53</a>].</p>
Full article ">Figure 24
<p>Comparison of MABE and MAPE for compressive strength models under FRP-wrapping effect [<a href="#B50-buildings-14-04009" class="html-bibr">50</a>,<a href="#B51-buildings-14-04009" class="html-bibr">51</a>,<a href="#B52-buildings-14-04009" class="html-bibr">52</a>,<a href="#B53-buildings-14-04009" class="html-bibr">53</a>].</p>
Full article ">Figure 25
<p>Experimental data and model predictions for ultimate axial strain under FRP-wrapping effect [<a href="#B50-buildings-14-04009" class="html-bibr">50</a>,<a href="#B51-buildings-14-04009" class="html-bibr">51</a>,<a href="#B52-buildings-14-04009" class="html-bibr">52</a>,<a href="#B53-buildings-14-04009" class="html-bibr">53</a>].</p>
Full article ">Figure 26
<p>Comparison of MABE and MAPE for ultimate axial strain models under FRP-wrapping effect [<a href="#B50-buildings-14-04009" class="html-bibr">50</a>,<a href="#B51-buildings-14-04009" class="html-bibr">51</a>,<a href="#B52-buildings-14-04009" class="html-bibr">52</a>,<a href="#B53-buildings-14-04009" class="html-bibr">53</a>].</p>
Full article ">
35 pages, 12083 KiB  
Review
Flexural Behavior and Failure Modes of Pultruded GFRP Tube Concrete-Filled Composite Beams: A Review of Experimental and Numerical Studies
by Mohammed Jalal Al-Ezzi, Agusril Ayamsir, A. B. M. Supian, Salmia Beddu and Rayeh Nasr Al-Dala’ien
Buildings 2024, 14(12), 3966; https://doi.org/10.3390/buildings14123966 - 13 Dec 2024
Viewed by 558
Abstract
Pultruded glass fiber-reinforced polymer (GFRP) materials are increasingly recognized in civil engineering for their exceptional properties, including a high strength-to-weight ratio, corrosion resistance, and ease of fabrication, making them ideal for composite structural applications. The use of concrete infill enhances the structural integrity [...] Read more.
Pultruded glass fiber-reinforced polymer (GFRP) materials are increasingly recognized in civil engineering for their exceptional properties, including a high strength-to-weight ratio, corrosion resistance, and ease of fabrication, making them ideal for composite structural applications. The use of concrete infill enhances the structural integrity of thin-walled GFRP sections and compensates for the low elastic modulus of hollow profiles. Despite the widespread adoption of concrete-filled pultruded GFRP tubes in composite beams, critical gaps remain in understanding their flexural behavior and failure mechanisms, particularly concerning design optimization and manufacturing strategies to mitigate failure modes. This paper provides a comprehensive review of experimental and numerical studies that investigate the impact of key parameters, such as concrete infill types, reinforcement strategies, bonding levels, and GFRP tube geometries, on the flexural performance and failure behavior of concrete-filled pultruded GFRP tubular members in composite beam applications. The analysis includes full-scale GFRP beam studies, offering a thorough comparison of documented flexural responses, failure modes, and structural performance outcomes. The findings are synthesized to highlight current trends, identify research gaps, and propose strategies to advance the understanding and application of these composite systems. The paper concludes with actionable recommendations for future research, emphasizing the development of innovative material combinations, optimization of structural designs, and refinement of numerical modeling techniques. Full article
(This article belongs to the Section Building Materials, and Repair & Renovation)
Show Figures

Figure 1

Figure 1
<p>Steel corrosion of reinforced concrete beam [<a href="#B8-buildings-14-03966" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Market share of fiber-reinforced polymer (FRP) by application.</p>
Full article ">Figure 3
<p>GFRP pultrusion process.</p>
Full article ">Figure 4
<p>Typical lay-ups of I-section [<a href="#B37-buildings-14-03966" class="html-bibr">37</a>].</p>
Full article ">Figure 5
<p>Longitudinal strain distribution of beams during the loading process [<a href="#B66-buildings-14-03966" class="html-bibr">66</a>].</p>
Full article ">Figure 6
<p>Load–displacement behavior of hollow and concrete-filled GFRP beams: H-0: hollow GFRP beam, H-10: GFRP beam filled grade 10 concrete, H-37: GFRP beam filled grad 37 concrete, and H-43: GFRP beam filled grad 43 concrete [<a href="#B7-buildings-14-03966" class="html-bibr">7</a>].</p>
Full article ">Figure 7
<p>Crack pattern at the failure of hollow and concrete-filled beams [<a href="#B7-buildings-14-03966" class="html-bibr">7</a>].</p>
Full article ">Figure 8
<p>Cross-section and dimensions of the beams (mm) [<a href="#B41-buildings-14-03966" class="html-bibr">41</a>].</p>
Full article ">Figure 9
<p>Failure behavior of GFRP concrete-filled composite beams with different hollow cores [<a href="#B41-buildings-14-03966" class="html-bibr">41</a>].</p>
Full article ">Figure 10
<p>Cross-sections of beam specimens (<b>A</b>) RC, (<b>B</b>) G0C, (<b>C</b>) G0.6A, (<b>D</b>) G1.15A, (<b>E</b>) G1.15B, and (<b>F</b>) G1.15C [<a href="#B68-buildings-14-03966" class="html-bibr">68</a>].</p>
Full article ">Figure 11
<p>Failure behavior of infill concrete [<a href="#B68-buildings-14-03966" class="html-bibr">68</a>].</p>
Full article ">Figure 12
<p>Composed beam steel angles, and penetrating long bolts to prevent concrete slip [<a href="#B40-buildings-14-03966" class="html-bibr">40</a>].</p>
Full article ">Figure 13
<p>Failure behavior of a pultruded GFRP beam (<b>a</b>) Plan view, (<b>b</b>) Transvers section, and (<b>c</b>) Longitudinal [<a href="#B69-buildings-14-03966" class="html-bibr">69</a>].</p>
Full article ">Figure 14
<p>Bending strength of single and multi-cell pultruded GFRP beams [<a href="#B43-buildings-14-03966" class="html-bibr">43</a>].</p>
Full article ">Figure 15
<p>Recommended configurations of the corner of PFRP profiles [<a href="#B72-buildings-14-03966" class="html-bibr">72</a>].</p>
Full article ">Figure 16
<p>Load–deflection curves of square and rectangular GFRP tubes [<a href="#B73-buildings-14-03966" class="html-bibr">73</a>].</p>
Full article ">Figure 17
<p>The effect of bonding on load–deflection behavior of composite beam of (A) hollow GFRP tube, (B) concrete filled GFRP tube, (C) concrete filled GFRP tube with bonded flange, and (D) concrete filled GFRP tube with bonded flange and web [<a href="#B44-buildings-14-03966" class="html-bibr">44</a>].</p>
Full article ">Figure 18
<p>Load–deflection behavior of circular GFRP concrete-filled composite beam with different configurations [<a href="#B16-buildings-14-03966" class="html-bibr">16</a>].</p>
Full article ">Figure 19
<p>(<b>a</b>) GFRP crushing at the control beam’s web-flange junction, and (<b>b</b>) failure of the central plain lightweight concrete core due to tension [<a href="#B81-buildings-14-03966" class="html-bibr">81</a>].</p>
Full article ">Figure 20
<p>Finite element models (FEMs) of midspan cross-sections with mesh for configurations C, D, and E [<a href="#B84-buildings-14-03966" class="html-bibr">84</a>].</p>
Full article ">Figure 21
<p>Typical numerical and experimental cracking patterns of infill concrete in short GFRP beam (<b>a</b>) GRRP longitudinal stress of web and bottom flange at a load of 160 kN, (<b>b</b>) GRRP in-plane shear strain of web and bottom flange at a load of 160 kN, (<b>c</b>) Infill concrete failure behavior of web-bonded beams (<b>d</b>) Infill concrete failure behavior of flange-bonded beams, (<b>e</b>) Infill concrete failure behavior of web, web-bonded beams [<a href="#B84-buildings-14-03966" class="html-bibr">84</a>].</p>
Full article ">Figure 22
<p>Solid finite element model of CFFT beam with tube cuts [<a href="#B87-buildings-14-03966" class="html-bibr">87</a>].</p>
Full article ">Figure 23
<p>(<b>A</b>) Cut length-to-radius ratio (α), and the reduction in moment capacity (ρ), (<b>B</b>) GFRP tubes without circumferential cuts under a 10 kN load, and (<b>C</b>) longitudinal stresses in GFRP tubes with 20% circumferential cuts with a 10 kN load [<a href="#B87-buildings-14-03966" class="html-bibr">87</a>].</p>
Full article ">Figure 24
<p>The failure mode of the GFRP beam with 51% fiber volume fraction, (<b>A</b>) numerical, and (<b>B</b>) experimental [<a href="#B9-buildings-14-03966" class="html-bibr">9</a>].</p>
Full article ">Figure 25
<p>Numerical and experimental results of GFRP composite beam infilled with composite fiber-reinforced polymer; (<b>a</b>) specimen 1, (<b>b</b>) specimen 2, and (<b>c</b>) specimen 3 [<a href="#B94-buildings-14-03966" class="html-bibr">94</a>].</p>
Full article ">
17 pages, 7695 KiB  
Article
Experimental and Numerical Study on the Impact Response of Composite Sandwich Structures with Different Cores
by Guangshuo Feng, Chunlu Xiao, Bo Liu, Haitao Zhang, Peipei Jia and Caizheng Wang
Polymers 2024, 16(23), 3436; https://doi.org/10.3390/polym16233436 - 7 Dec 2024
Viewed by 636
Abstract
This study analyzes the impact mechanical response of sandwich structures with foam and wood cores through experimental and numerical methods. The aim is to determine whether a sustainable core material, such as cork wood, can serve as a reliable alternative to the commonly [...] Read more.
This study analyzes the impact mechanical response of sandwich structures with foam and wood cores through experimental and numerical methods. The aim is to determine whether a sustainable core material, such as cork wood, can serve as a reliable alternative to the commonly used Polystyrene (PS) foam core in sandwich structures. Impact experiments were conducted at varying energy levels using an INSTRON CEAST 9350 drop tower, demonstrating the superiority of sandwich structures compared to single-material alternatives. Numerical models were developed in ABAQUS, where glass fiber reinforced polymer (GFRP) composite panels were represented using solid element C3D8R and the 3D Hashin failure criteria, which were incorporated via the user subroutine VUMAT. The results indicate that the contact force of the sandwich structure with a wood core surpassed that of the foam core counterpart. In both sandwich structures, damage initially occurred at the impact point on the surface, leading to plastic deformation and damage within the core, while the composite panel on the rear surface ultimately failed. These findings provide valuable insights for designers, enabling parametric studies to select appropriate core materials that enhance the impact resistance of sandwich structures. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>The process of preparing sandwich structures.</p>
Full article ">Figure 2
<p>The drop tower equipment (<b>left</b>: falling carriage system; <b>right</b>: impactor and support fixture).</p>
Full article ">Figure 3
<p>Diagram of cohesive zone model.</p>
Full article ">Figure 4
<p>Flowchart of finite element simulation with user subroutine.</p>
Full article ">Figure 5
<p>Finite element model of sandwich structure under impact loading.</p>
Full article ">Figure 6
<p>Contact force histories of the GFRP sample under 20 J, 40 J, 60 J, 80 J, and 100 J.</p>
Full article ">Figure 7
<p>Displacement–time curves of the GFRP sample under 20 J, 40 J, 60 J, 80 J, and 100 J.</p>
Full article ">Figure 8
<p>Contact force–displacement curves of the GFRP sample under different energies and images after impact.</p>
Full article ">Figure 9
<p>Energy absorption histories of the GFRP sample under 20 J, 40 J,60 J, 80 J, and 100 J.</p>
Full article ">Figure 10
<p>(<b>a</b>) Contact force and (<b>b</b>) tip displacement histories of the sandwich structure with different cores under 100 J.</p>
Full article ">Figure 11
<p>Contact force–displacement curves of the sandwich structure with different cores under 100 J and cross-sectional images after impact.</p>
Full article ">Figure 12
<p>Comparison of the initial (1st) and secondary (2nd) flexural stiffness of sandwich panels via the original slope of the force–displacement curve in <a href="#polymers-16-03436-f011" class="html-fig">Figure 11</a>.</p>
Full article ">Figure 13
<p>Comparison of the post-test damage status on both front and back surfaces of all tested specimens.</p>
Full article ">Figure 14
<p>Contact force history from experiment and simulation results: (<b>a</b>) sandwich structure with cork wood core; (<b>b</b>) sandwich structure with PS foam core.</p>
Full article ">Figure 15
<p>Failure process of sandwich structures with (<b>a</b>) cork wood core and (<b>b</b>) PS foam core.</p>
Full article ">Figure 16
<p>Failure pattern of sandwich structures with (<b>a</b>) cork wood core and (<b>b</b>) PS foam core.</p>
Full article ">
23 pages, 6353 KiB  
Article
Effects of Hygrothermal Condition on Water Diffusion and Flexural Properties of Carbon–Glass Hybrid Fiber-Reinforced Epoxy Polymer Winding Pipes
by Ying Zhao, Qiang Li, Guoqiang Zhou, Kehai Zhu, Bo Jing, Kangnan Zhu, Jiajun Shi and Chenggao Li
Polymers 2024, 16(23), 3433; https://doi.org/10.3390/polym16233433 - 6 Dec 2024
Viewed by 541
Abstract
Carbon–glass hybrid fiber-reinforced epoxy polymer (C-GFRP) winding pipes integrated with the advantages of light weight, high strength, corrosion resistance, and cost-effectiveness offer immense potential to mitigate corrosion issues in oil, gas, and water transportation pipelines. In this study, C-GFRP winding pipes underwent accelerated [...] Read more.
Carbon–glass hybrid fiber-reinforced epoxy polymer (C-GFRP) winding pipes integrated with the advantages of light weight, high strength, corrosion resistance, and cost-effectiveness offer immense potential to mitigate corrosion issues in oil, gas, and water transportation pipelines. In this study, C-GFRP winding pipes underwent accelerated aging tests through immersion in distilled water at temperatures of 25 °C, 40 °C, and 60 °C for 146 days. Water absorption tests were conducted to investigate the water absorption behavior of only CFRP- or GFRP-side absorbed water. Bending tests were performed to assess the evolution of the pipes’ flexural properties in two directions (GFRP or CFRP in tension). The results showed that the single-sided water absorption behavior adhered to the two-stage diffusion model. The diffusion coefficient, activation energy, and 146-day water absorption were all higher for the CFRP-side absorbed water compared to the GFRP-side absorbed water. The flexural strength and modulus of C-GFRP pipes were influenced by post-curing and resin hydrolysis/debonding. Initially, the flexural strength of CFRP in tension was higher than that of CFRP in tension. After 146 days of aging, the flexural strength of CFRP in tension was lower than that of CFRP in tension. Utilizing Arrhenius theory, the long-term lives were predicted for the flexural strength at temperatures of 5.4 °C, 12.8 °C, and 17.8 °C. The predicted lives of GFRP in tension were higher than those of CFRP in tension. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>C/GFRP winding pipes.</p>
Full article ">Figure 2
<p>Water absorption sample.</p>
Full article ">Figure 3
<p>Three-point bending tests.</p>
Full article ">Figure 4
<p>The water absorption and fitting curve of a C-GFRP winding pipe.</p>
Full article ">Figure 5
<p>Effect of temperature on the water absorption diffusion coefficient.</p>
Full article ">Figure 6
<p>Flexural failure mode.</p>
Full article ">Figure 7
<p>The effect of temperature and time on the flexural strength.</p>
Full article ">Figure 8
<p>The effect of temperature and time on the flexural modulus.</p>
Full article ">Figure 9
<p>Fitting results of the flexural strength retention of a C-GFRP winding pipe.</p>
Full article ">Figure 10
<p>The fitting parameters of flexural strength retention.</p>
Full article ">Figure 11
<p>The Arrhenius linear fitting of the flexural strength.</p>
Full article ">Figure 12
<p>The relationship between the fitting parameters and flexural strength retention rate.</p>
Full article ">Figure 13
<p>The time shift factors between the given predicted environmental temperatures and accelerated test temperatures.</p>
Full article ">Figure 14
<p>Long-term life prediction curves of C-GFRP winding pipes at given temperatures.</p>
Full article ">Figure 15
<p>The fitting parameters between the given predicted environmental temperatures and accelerated test temperatures.</p>
Full article ">Figure 16
<p>A comparison of life prediction in the B1 and B2 directions.</p>
Full article ">
20 pages, 9659 KiB  
Article
Nondestructive Detection of Osmotic Damage in GFRP Boat Hulls Using Active Infrared Thermography Methods
by Endri Garafulić, Petra Bagavac and Lovre Krstulović-Opara
J. Mar. Sci. Eng. 2024, 12(12), 2247; https://doi.org/10.3390/jmse12122247 - 6 Dec 2024
Viewed by 371
Abstract
This article presents the application of infrared thermography as a nondestructive testing method (NDT) for detecting osmotic damage in glass-fiber-reinforced polymer (GFRP) and glass-reinforced polymer (GRP) boat hull structures. The aim of the conducted experiments is to explore the possibilities of applying active [...] Read more.
This article presents the application of infrared thermography as a nondestructive testing method (NDT) for detecting osmotic damage in glass-fiber-reinforced polymer (GFRP) and glass-reinforced polymer (GRP) boat hull structures. The aim of the conducted experiments is to explore the possibilities of applying active infrared thermography to real structures and to establish a procedure capable of filtering out anomalies caused by various thermal influences, such as thermal reflections from surrounding objects, geometry effects, and heat flow variations on the observed object. The methods used for post-processing IR signals include lock-in thermography (LT), pulse thermography (PT), pulse phase thermography (PPT), and gradient pulse phase thermography (GT). The practical application and advantages and disadvantages of infrared thermography in identifying osmotic damage in GFRP and GRP boat hulls will be illustrated through three case studies. Each case study is based on specific conditions and characteristics of different types of osmotic damage, enabling a thorough analysis of the effectiveness of the method in detecting and assessing the severity of the damage. The post-processed thermal images enable a clearer distinction between damaged and undamaged zones, improving the robustness of detection under realistic field conditions. Full article
(This article belongs to the Section Ocean Engineering)
Show Figures

Figure 1

Figure 1
<p>Formation of osmotic bubbles.</p>
Full article ">Figure 2
<p>(<b>a</b>) Visible blisters and (<b>b</b>) the standard method of detecting the osmotic process.</p>
Full article ">Figure 3
<p>The transition from the (<b>a</b>) time domain to the (<b>b</b>) frequency domain using the FFT algorithm [<a href="#B14-jmse-12-02247" class="html-bibr">14</a>].</p>
Full article ">Figure 4
<p>Signal processing in PPT: (<b>a</b>) thermogram sequence, 3D matrix, and thermal profiles for a defective pixel, red line (Td), a nondefective pixel, blue line (TSa), and the difference between them, green line (TdTSa); (<b>b</b>) amplitudegram sequence and amplitude profiles; (<b>c</b>) phasegram sequence and phase profiles for a defective pixel, red line (Fd), a non-defective pixel, blue line (Fsa), and the difference between them, green line (Fd-Fsa).</p>
Full article ">Figure 5
<p>(<b>a</b>) Setup for nondestructive testing with lock-in thermography, (<b>b</b>) region of interest (ROI) where osmotic damage is circled and marked with labels A, B, C, and D.</p>
Full article ">Figure 6
<p>Sinusoidal response of relay-controlled halogen floodlights.</p>
Full article ">Figure 7
<p>Raw thermal image.</p>
Full article ">Figure 8
<p>Phase delay, P = 24 s, the osmotic damage is circled and marked with labels A, B, C, and D.</p>
Full article ">Figure 9
<p>Phase delay, P = 72 s, the osmotic damage is circled and marked with labels A, B, C, and D.</p>
Full article ">Figure 10
<p>Phase delay, P = 120 s, the osmotic damage is circled and marked with labels A, B, C, and D.</p>
Full article ">Figure 11
<p>Osmotic damage on the boat hull: (<b>a</b>) photo of the boat hull, (<b>b</b>) osmotic blisters A and B.</p>
Full article ">Figure 12
<p>A-scan of the back wall on calibration steel block K1: (<b>a</b>) 4 MHz frequency probe, (<b>b</b>) 1 MHz frequency probe.</p>
Full article ">Figure 13
<p>(<b>a</b>) USM GO device; (<b>b</b>) K1S-C 1 MHz frequency probe with a plexiglass attachment for beam focusing.</p>
Full article ">Figure 14
<p>A-scan: (<b>a</b>) osmotic damage, (<b>b</b>) undamaged material.</p>
Full article ">Figure 15
<p>Phase shift at different excitation frequencies: (<b>a</b>) f = 0.04167 Hz, (<b>b</b>) f = 0.0208 Hz, (<b>c</b>) f = 0.0139 Hz, and (<b>d</b>) f = 0.0083 Hz.</p>
Full article ">Figure 15 Cont.
<p>Phase shift at different excitation frequencies: (<b>a</b>) f = 0.04167 Hz, (<b>b</b>) f = 0.0208 Hz, (<b>c</b>) f = 0.0139 Hz, and (<b>d</b>) f = 0.0083 Hz.</p>
Full article ">Figure 16
<p>Osmotic damage on the hull of the vessel after grinding the anti-fouling paint and protective epoxy coating.</p>
Full article ">Figure 17
<p>The undamaged hull of the vessel: (<b>a</b>) during UT testing, (<b>b</b>) after grinding the anti-fouling paint and protective epoxy coating.</p>
Full article ">Figure 18
<p>Phase shift at different excitation frequencies: (<b>a</b>) f = 0.04167 Hz, (<b>b</b>) f = 0.0208 Hz, (<b>c</b>) f = 0.0139 Hz, and (<b>d</b>) f = 0.0083 Hz.</p>
Full article ">Figure 19
<p>(<b>a</b>) PT applied on a boat’s hull, (<b>b</b>) photography of zones where blisters are detected, and (<b>c</b>) thermogram with location of osmotic blisters.</p>
Full article ">Figure 20
<p>PPT results of blister osmosis detection: (<b>a</b>) selected amplitudegrams—even symmetry; (<b>b</b>) selected phasegrams—odd symmetry, for the following frequencies: ±0.01, 0.02, 0.03, 0.04, 0.05 Hz.</p>
Full article ">Figure 21
<p>Boat’s hull phasegrams, image detail of osmosis damage—2D review for (<b>a</b>) f = 0.01 Hz and (<b>c</b>) f = 0.075 Hz, and 3D review for (<b>b</b>) f = 0.01 Hz and (<b>d</b>) f = 0.075 Hz.</p>
Full article ">Figure 22
<p>(<b>a</b>) Thermogram with the location of the osmotic blisters, (<b>b</b>) and thermal gradient image processing.</p>
Full article ">Figure 23
<p>Gradient of phasegram image shown in <a href="#jmse-12-02247-f022" class="html-fig">Figure 22</a>a.</p>
Full article ">Figure 24
<p>Wet fiberglass and delamination from the acids in zone A.</p>
Full article ">
19 pages, 22250 KiB  
Article
Structural and Mechanical Properties of Recycled HDPE with Milled GFRP as a Filler
by Maciej Jan Spychała, Paulina Latko-Durałek, Danuta Miedzińska, Kamila Sałasińska, Iga Cetnar, Arkadiusz Popławski and Anna Boczkowska
Materials 2024, 17(23), 5875; https://doi.org/10.3390/ma17235875 - 29 Nov 2024
Viewed by 474
Abstract
The increasing complexity and production volume of glass-fiber-reinforced polymers (GFRP) present significant recycling challenges. This paper explores a potential use for mechanically recycled GFRP by blending it with high-density polyethylene (HDPE). This composite could be applied in products such as terrace boards, pipes, [...] Read more.
The increasing complexity and production volume of glass-fiber-reinforced polymers (GFRP) present significant recycling challenges. This paper explores a potential use for mechanically recycled GFRP by blending it with high-density polyethylene (HDPE). This composite could be applied in products such as terrace boards, pipes, or fence posts, or as a substitute filler for wood flour and chalk. Recycled GFRP from post-consumer bus bumpers were ground and then combined with recycled HDPE in a twin-screw extruder at concentrations of 10, 20, 30, and 40 wt%. The study examined the mechanical and structural properties of the resulting composites, including the effects of aging and re-extrusion. The modulus of elasticity increased from 0.878 GPa for pure rHDPE to 1.806 GPa for composites with 40 wt% recycled GFRP, while the tensile strength ranged from 36.5 MPa to 28.7 MPa. Additionally, the porosity increased linearly from 2.65% to 7.44% for composites with 10 wt% and 40 wt% recycled GFRP, respectively. Aging and re-extrusion improved the mechanical properties, with the tensile strength of the 40 wt% GFRP composite reaching 34.1 MPa, attributed to a reduction in porosity by nearly half, reaching 3.43%. Full article
(This article belongs to the Section Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>Tested material preparation process: (<b>a</b>) pellets of rHDPE; (<b>b</b>) bus bumpers waste; (<b>c</b>) rGFRP after grinding; (<b>d</b>) composite pellets 60% rHDPE + 40% rGFRP; (<b>e</b>) dog-bone shaped specimen for mechanical tests; and (<b>f</b>) rGFRP scraps at higher resolution.</p>
Full article ">Figure 2
<p>Scheme of material preparation.</p>
Full article ">Figure 3
<p>FTIR spectrum of (<b>a</b>) rHDPE; (<b>b</b>) resin from rGFRP waste; and (<b>c</b>) white paint from rGFRP waste.</p>
Full article ">Figure 4
<p>Microstructure observations: (<b>a</b>) image of cross-section of rGFRC material before milling; (<b>b</b>) image taken by Keyence VHX-1000 microscope of surface of the rHDPE pellets; (<b>c</b>) example of image of filler contamination analysis of rHDPE/rGFRC 40% sample; (<b>d</b>) example of image with fiber glass contamination analysis of rHDPE/rGFRC 40% sample; (<b>e</b>) surface of cross-section of rHDPE/rGFRC 40% sample; and (<b>f</b>) outer layer of rHDPE material with dimension.</p>
Full article ">Figure 5
<p>Results of CT for composites of rHDPE containing (<b>a</b>) 10 wt%; (<b>b</b>) 20 wt%; (<b>c</b>) 30 wt%; and (<b>d</b>) 40 wt % of rGFRP.</p>
Full article ">Figure 6
<p>Images of rHDPE/rGFRP material: (<b>a</b>) cross-section; and (<b>b</b>) surface of the sample.</p>
Full article ">Figure 7
<p>Images of rHDPE/rGFRC material: (<b>a</b>) macro view on the image; and (<b>b</b>–<b>d</b>) close-up of selected area—brighter areas are elements with higher mass number.</p>
Full article ">Figure 8
<p>Behavior of rHDPE/rGFRP composites during (<b>a</b>) second heating, and (<b>b</b>) cooling, obtained by DSC analysis.</p>
Full article ">Figure 9
<p>Engineering stress–strain curves for tensile test samples: (<b>a</b>) with varying percentages of rGFRP; and (<b>b</b>) with 40 wt% rGFRP after additional processes.</p>
Full article ">Figure 10
<p>Comparison of (<b>a</b>) modulus of elasticity; (<b>b</b>) yield strength; (<b>c</b>) tensile strength; and (<b>d</b>) strain for tensile strength analyzed for all studied composites.</p>
Full article ">Figure 11
<p>Results of CT for composites of rHDPE containing 40 wt% rGFRP after (<b>a</b>) additional extrusion; (<b>b</b>) aging; (<b>c</b>) and extrusion and aging. Measured region: 1 mm × 1 mm × 1 mm region, with resolution 20.90 m.</p>
Full article ">
22 pages, 17612 KiB  
Article
Short-Beam Shear Strength of New-Generation Glass Fiber-Reinforced Polymer Bars Under Harsh Environment: Experimental Study and Artificial Neural Network Prediction Model
by Mesfer M. Al-Zahrani
Polymers 2024, 16(23), 3358; https://doi.org/10.3390/polym16233358 - 29 Nov 2024
Viewed by 525
Abstract
In this study, the short-beam shear strength (SBSS) retention of two types of glass fiber-reinforced polymer (GFRP) bars—sand-coated (SG) and ribbed (RG)—was subjected to alkaline, acidic, and water conditions for up to 12 months under both high-temperature and ambient laboratory conditions. Comparative assessments [...] Read more.
In this study, the short-beam shear strength (SBSS) retention of two types of glass fiber-reinforced polymer (GFRP) bars—sand-coated (SG) and ribbed (RG)—was subjected to alkaline, acidic, and water conditions for up to 12 months under both high-temperature and ambient laboratory conditions. Comparative assessments were also performed on older-generation sand-coated (SG-O) and ribbed (RG-O1 and RG-O2) GFRP bars exposed to identical conditions. The results demonstrate that the new-generation GFRP bars, SG and RG, exhibited significantly better durability in harsh environments and exhibited SBSS retentions varying from 61 to 100% in SG and 90–98% in RG under the harshest conditions compared to 56–69% in SG-O, 71–80% in RG-O1, and 74–88% in RG-O2. Additionally, predictive models using both artificial neural networks (ANNs) and linear regression were developed to estimate the strength retention. The ANN model, with an R2 of 0.95, outperformed the linear regression model (R2 = 0.76), highlighting its greater accuracy and suitability for predicting the SBSS of GFRP bars. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>GFRP bar types used in this study.</p>
Full article ">Figure 2
<p>SEM micrographs of the GFRP bars (near the circumference).</p>
Full article ">Figure 3
<p>Exposure conditions used in this study.</p>
Full article ">Figure 4
<p>SBSS testing as per ASTM D4475-21 [<a href="#B8-polymers-16-03358" class="html-bibr">8</a>].</p>
Full article ">Figure 5
<p>Test matrix (graphical form).</p>
Full article ">Figure 6
<p>ANN architecture.</p>
Full article ">Figure 7
<p>SBSS–crosshead displacement response of control RG and SG bars. Note: multiple lines indicate replicates.</p>
Full article ">Figure 8
<p>Sand-coated SG bars: SBSS–displacement curves (notation: exposure solution–temperature–exposure duration). Note: multiple lines indicate replicates.</p>
Full article ">Figure 8 Cont.
<p>Sand-coated SG bars: SBSS–displacement curves (notation: exposure solution–temperature–exposure duration). Note: multiple lines indicate replicates.</p>
Full article ">Figure 9
<p>Ribbed-type RG bars: SBSS–displacement curves (notation: conditioning solution–temperature–duration). Note: multiple lines indicate replicates.</p>
Full article ">Figure 9 Cont.
<p>Ribbed-type RG bars: SBSS–displacement curves (notation: conditioning solution–temperature–duration). Note: multiple lines indicate replicates.</p>
Full article ">Figure 10
<p>Failure modes.</p>
Full article ">Figure 11
<p>SBSS degradation due to C1 (alkaline) and C2 (alkaline and salt) conditionings.</p>
Full article ">Figure 12
<p>SBSS degradation due to C3 (acidic) and C4 (water) conditionings.</p>
Full article ">Figure 13
<p>Surface features of older-generation GFRP bars.</p>
Full article ">Figure 14
<p>Effect of C1 (alkaline) SBSS retention.</p>
Full article ">Figure 15
<p>Effect of C2 (alkaline and salt) SBSS retention.</p>
Full article ">Figure 16
<p>Effect of C3 (acid) SBSS retention.</p>
Full article ">Figure 17
<p>Effect of C4 (water) SBSS retention.</p>
Full article ">Figure 18
<p>SEM fractography (250× magnification): effect of conditioning C1 on SG bar at 12 months of exposure at 60 °C.</p>
Full article ">Figure 19
<p>SEM fractography (250× magnification): effect of conditioning C1 on RG bar at 12 months of exposure at 60 °C.</p>
Full article ">Figure 20
<p>Linear regression model.</p>
Full article ">Figure 21
<p>Structure of ANN used in MATLAB (w = weight; b = bias).</p>
Full article ">Figure 22
<p>Parity plots—experimental vs. ANN prediction of SBSS.</p>
Full article ">Figure 23
<p>A frequency histogram of the error distribution with 20 bins.</p>
Full article ">
17 pages, 10313 KiB  
Article
Flexural Behavior of Innovative Glass Fiber-Reinforced Composite Beams Reinforced with Gypsum-Based Composites
by Yiwen Liu, Bo Su and Tianyu Zhang
Polymers 2024, 16(23), 3327; https://doi.org/10.3390/polym16233327 - 27 Nov 2024
Viewed by 596
Abstract
Glass Fiber-Reinforced Composite (GFRP) has found widespread use in engineering structures due to its lightweight construction, high strength, and design flexibility. However, pure GFRP beams exhibit weaknesses in terms of stiffness, stability, and local compressive strength, which compromise their bending properties. In addressing [...] Read more.
Glass Fiber-Reinforced Composite (GFRP) has found widespread use in engineering structures due to its lightweight construction, high strength, and design flexibility. However, pure GFRP beams exhibit weaknesses in terms of stiffness, stability, and local compressive strength, which compromise their bending properties. In addressing these limitations, this study introduces innovative square GFRP beams infused with gypsum-based composites (GBIGCs). Comprehensive experiments and theoretical analyses have been conducted to explore their manufacturing process and bending characteristics. Initially, four types of GBIGC—namely, hollow GFRP beams, pure gypsum, steel-reinforced gypsum, and fiber-mixed gypsum-infused beams—were designed and fabricated for comparative analysis. Material tests were conducted to assess the coagulation characteristics of gypsum and its mechanical performance influenced by polyvinyl acetate fibers (PVAs). Subsequently, eight GFRP square beams (length: 1.5 m, section size: 150 mm × 150 mm) infused with different gypsum-based composites underwent four-point bending tests to determine their ultimate bending capacity and deflection patterns. The findings revealed that a 0.12% dosage of protein retarder effectively extends the coagulation time of gypsum, making it suitable for specimen preparation, with initial and final setting times of 113 min and 135 min, respectively. The ultimate bending load of PVA-mixed gypsum-infused GFRP beams is 203.84% higher than that of hollow beams, followed by pure gypsum and steel-reinforced gypsum, with increased values of 136.97% and 186.91%, respectively. The ultimate load values from the theoretical and experimental results showed good agreement, with an error within 7.68%. These three types of GBIGCs with significantly enhanced flexural performance can be filled with different materials to meet specific load-bearing requirements for various scenarios. Their improved flexural strength and lightweight characteristics make GBIGCs well suited for applications such as repairing roof beams, light prefabricated frames, coastal and offshore buildings. Full article
(This article belongs to the Special Issue Application and Characterization of Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>GFRP beams and directions.</p>
Full article ">Figure 2
<p>Samples in two directions of GFRP.</p>
Full article ">Figure 3
<p>Gypsum coagulation tests.</p>
Full article ">Figure 4
<p>Manufacturing process of PVA gypsum blocks.</p>
Full article ">Figure 5
<p>The 50 T self-balancing device and layout of sensors.</p>
Full article ">Figure 6
<p>Relation between the retarder and the setting time of gypsum.</p>
Full article ">Figure 7
<p>Mechanical performance of PVA gypsum blocks.</p>
Full article ">Figure 8
<p>Loading of GFRP beams.</p>
Full article ">Figure 9
<p>Local failure pattern at loading points of hollow GFRP or GBIGC specimens.</p>
Full article ">Figure 10
<p>Mid-span deflection of test beams.</p>
Full article ">Figure 11
<p>Load–strain curves of different gauges.</p>
Full article ">Figure 12
<p>Strain distribution along web height.</p>
Full article ">
23 pages, 12221 KiB  
Article
Experimental Study on Axial Compression Behavior of Molybdenum Tailings Concrete Column Confined by GFRP
by Jian Yuan, Xin Zhao, Lianmin Tian, Zhaolong Hou, Yunfeng Pan and Jun He
Buildings 2024, 14(12), 3779; https://doi.org/10.3390/buildings14123779 - 26 Nov 2024
Viewed by 352
Abstract
To promote the application of molybdenum tailings as the fine aggregate in concrete in construction engineering and verify the feasibility of fiber-reinforced polymer (FRP) material for strengthening molybdenum tailings concrete columns, this study takes a short circular molybdenum tailings concrete column reinforced by [...] Read more.
To promote the application of molybdenum tailings as the fine aggregate in concrete in construction engineering and verify the feasibility of fiber-reinforced polymer (FRP) material for strengthening molybdenum tailings concrete columns, this study takes a short circular molybdenum tailings concrete column reinforced by glass FRP (GFRP) as the research object. The influences of the molybdenum tailings content (0%, 25%, 50%, 75%, and 100%), the concrete grade (C30, C40, and C50), and the layer number (0, 1, and 2) of the GFRP sheet on the axial compressive capacity of the molybdenum tailings concrete column are investigated. The experimental phenomena and failure modes of the unreinforced and GFRP-reinforced columns are analyzed. The axial compressive strengths of the unreinforced and GFRP-reinforced columns are then compared. The load–strain curve and load–displacement curve of typical molybdenum tailings concrete columns are presented. Subsequently, six classical strength models for FRP-reinforced concrete are used to calculate the axial compressive strength of the confined specimens. The results show that the best classical model has a predictive accuracy with an absolute relative deviation (ARD) of 8.5%. To provide a better prediction of the compressive strength of the GFRP-reinforced molybdenum tailings concrete column, the best classical model is further improved, and the ARD of the modified model is only 5.87%. Full article
(This article belongs to the Special Issue Optimal Design of FRP Strengthened/Reinforced Construction Materials)
Show Figures

Figure 1

Figure 1
<p>Raw materials of molybdenum tailings concrete: (<b>a</b>) molybdenum tailings; (<b>b</b>) sand; (<b>c</b>) stone; (<b>d</b>) cement.</p>
Full article ">Figure 2
<p>Axial compression tests of concrete specimens: (<b>a</b>) C30-Mu100; (<b>b</b>) C40-Mu25; (<b>c</b>) C40-Mu75; (<b>d</b>) C50-Mu0; (<b>e</b>) C50-Mu50; (<b>f</b>) C50-Mu100.</p>
Full article ">Figure 3
<p>GFRP cloth and epoxy resin adhesive: (<b>a</b>) GFRP cloth; (<b>b</b>) epoxy resin adhesive.</p>
Full article ">Figure 4
<p>Specimen construction: (<b>a</b>) molds; (<b>b</b>) stirring; (<b>c</b>) specimens.</p>
Full article ">Figure 5
<p>Loading device.</p>
Full article ">Figure 6
<p>Measurement points: (<b>a</b>) unconfined column; (<b>b</b>) GFRP-confined column.</p>
Full article ">Figure 7
<p>Strain measurement diagram: (<b>a</b>) data acquisition system; (<b>b</b>) connection of strain gauge; (<b>c</b>) resistance measurement.</p>
Full article ">Figure 8
<p>Failure process of unconfined concrete column: (<b>a</b>) splitting failure; (<b>b</b>) cone failure.</p>
Full article ">Figure 9
<p>Failure process of GFRP-confined concrete column: (<b>a</b>) failure at upper part; (<b>b</b>) failure at central part; (<b>c</b>) failure at lower part.</p>
Full article ">Figure 9 Cont.
<p>Failure process of GFRP-confined concrete column: (<b>a</b>) failure at upper part; (<b>b</b>) failure at central part; (<b>c</b>) failure at lower part.</p>
Full article ">Figure 10
<p>Axial compressive strength of columns with different concrete grades: (<b>a</b>) C30; (<b>b</b>) C40; (<b>c</b>) C50.</p>
Full article ">Figure 10 Cont.
<p>Axial compressive strength of columns with different concrete grades: (<b>a</b>) C30; (<b>b</b>) C40; (<b>c</b>) C50.</p>
Full article ">Figure 11
<p>Typical load–strain curves: (<b>a</b>) C40-Mu75-0; (<b>b</b>) C40-Mu75-1; (<b>c</b>) C40-Mu75-2.</p>
Full article ">Figure 12
<p>Typical load–displacement curves: (<b>a</b>) C30-Mu50; (<b>b</b>) C40-Mu50; (<b>c</b>) C50-Mu50.</p>
Full article ">Figure 13
<p>Comparisons of calculated strength and tested strength: (<b>a</b>) Lam and Teng [<a href="#B22-buildings-14-03779" class="html-bibr">22</a>]; (<b>b</b>) Plam and Hadi [<a href="#B23-buildings-14-03779" class="html-bibr">23</a>]; (<b>c</b>) Yang and Feng [<a href="#B24-buildings-14-03779" class="html-bibr">24</a>]; (<b>d</b>) Samaan et al. [<a href="#B25-buildings-14-03779" class="html-bibr">25</a>]; (<b>e</b>) Youssef et al. [<a href="#B26-buildings-14-03779" class="html-bibr">26</a>]; (<b>f</b>) Teng et al. [<a href="#B27-buildings-14-03779" class="html-bibr">27</a>]</p>
Full article ">Figure 14
<p>Comparisons of the Samaan et al. [<a href="#B25-buildings-14-03779" class="html-bibr">25</a>] model and the modified model.</p>
Full article ">
17 pages, 6869 KiB  
Article
Analysis and Prediction of Temperature Using an Artificial Neural Network Model for Milling Glass Fiber Reinforced Polymer Composites
by Paulina Spanu, Bogdan Felician Abaza and Teodor Catalin Constantinescu
Polymers 2024, 16(23), 3283; https://doi.org/10.3390/polym16233283 - 25 Nov 2024
Viewed by 636
Abstract
Milling parts made from glass fiber-reinforced polymer (GFRP) composite materials are recommended to achieve the geometric shapes and dimensional tolerances required for large parts manufactured using the spray lay-up technique. The quality of the surfaces machined by milling is significantly influenced by the [...] Read more.
Milling parts made from glass fiber-reinforced polymer (GFRP) composite materials are recommended to achieve the geometric shapes and dimensional tolerances required for large parts manufactured using the spray lay-up technique. The quality of the surfaces machined by milling is significantly influenced by the temperature generated in the cutting zone. This study aims to develop an Artificial Neural Network (ANN) model to predict the temperature generated when milling GFRP. The ANN model for temperature prediction was created using a virtual instrument developed in the graphical programming language LabVIEW. Predicting temperature is crucial because excessive heat during milling can lead to several issues, such as tool wear and thermal degradation in the polymer matrix. The temperature in the tool–workpiece contact surface during the milling process was measured using a thermography technique with a ThermaCAM SC 640 camera (provided by FLIR Systems AB, Danderyd, Sweden), and the data were analyzed using the ThermaCAM Researcher Professional 2.8 SR-2 software. Experimental research shows that the cutting speed has a much more significant effect on the temperature in the cutting zone compared to axial depth of cut and feed speed. The maximum temperature of 85.19 °C was measured in the tool–workpiece contact zone during machining at a cutting speed of 75.39 m/min, a feed rate of 250 mm/min, and an axial depth of cut of 12 mm. This temperature rise occurred due to the larger contact area and heightened friction resulting from the abrasive characteristics of the reinforcement material. Full article
Show Figures

Figure 1

Figure 1
<p>Laminated composite with polymer matrix randomly reinforced with glass fiber.</p>
Full article ">Figure 2
<p>The cutting tool used for milling GFRP.</p>
Full article ">Figure 3
<p>Broken cutting tool.</p>
Full article ">Figure 4
<p>The temperature distribution along the axial direction of the cutting tool.</p>
Full article ">Figure 5
<p>Artificial Neural Network (ANN) architecture. (<b>a</b>) ANN—general architecture; (<b>b</b>) neuron general mathematical representation model.</p>
Full article ">Figure 6
<p>Partial diagram of ANN sub-VI used to create, teach, and test, and ANN.</p>
Full article ">Figure 7
<p>Partial diagram of sub-VI code for teaching and validating data files of ANN.</p>
Full article ">Figure 8
<p>Example of plot of error trend out and key parameters for ANN 3:28:8:1 for GFRP.</p>
Full article ">Figure 9
<p>ANN 3:23:8:1 architecture for GFRP.</p>
Full article ">Figure 10
<p>Comparison of <span class="html-italic">Tmax</span> between experimental data and ANN predictions with 3:23:8:1 architecture for GFRP.</p>
Full article ">Figure 11
<p>Prediction performance of <span class="html-italic">Tmax</span> for an ANN 3:28:8:1 for GFRP: (<b>a</b>) distribution of error per dataset; (<b>b</b>) coefficient of determination <span class="html-italic">R</span><sup>2</sup>.</p>
Full article ">Figure 12
<p>Coefficient of determination variation with ANN architectures.</p>
Full article ">Figure 13
<p>Top 5 ANNs for GFRP for the prediction of <span class="html-italic">Tmax</span> based on a dataset of 35 inputs.</p>
Full article ">
24 pages, 10885 KiB  
Article
Study on Acoustic Emission Characteristics and Damage Mechanism of Wind Turbine Blade Main Spar with Different Defects
by Yanan Zhang, Shaojie Xue, Chuanyong Chen, Tianchang Ma and Bo Zhou
Polymers 2024, 16(23), 3261; https://doi.org/10.3390/polym16233261 - 23 Nov 2024
Viewed by 669
Abstract
This paper aimed to understand the AE signal characteristics and damage mechanism of wind turbine blade main spar materials with different defects during the damage evolution process. According to the typical delamination and wrinkle defects in wind turbine blades, the GFRP composite with [...] Read more.
This paper aimed to understand the AE signal characteristics and damage mechanism of wind turbine blade main spar materials with different defects during the damage evolution process. According to the typical delamination and wrinkle defects in wind turbine blades, the GFRP composite with defects is artificially prefabricated. Through acoustic emission experiments, the mechanical properties and acoustic emission characteristic trends of wind turbine blade main spar composites with different defects under tensile loading conditions were analyzed, and the damage evolution mechanism of different defects was explained according to the microscopic results. The results show that the existence of artificial defects will not only affect the mechanical properties of composite materials but also affect the damage evolution process of the materials. The size and location of delamination defects and the different aspect ratio of the wrinkle defects have a certain influence on the damage mechanism of the material. K-means cluster analysis of AE parameters identified the damage models of GFRP composites. The types of damage modes of delamination defects and wrinkle defects are the same, and the range of characteristic frequency is roughly the same. This study has important reference significance for structural damage monitoring and damage evolution research of wind turbine blade composites. Full article
(This article belongs to the Special Issue Additive Manufacturing of Fibre Reinforced Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Profile structure of wind turbine blade.</p>
Full article ">Figure 2
<p>Typical manufacturing process defects of wind turbine blade main spar structure: (<b>a</b>) wrinkle defects; (<b>b</b>) delamination defects; (<b>c</b>) air bubble defects; (<b>d</b>) lack glue defects; (<b>e</b>) pore defects; (<b>f</b>) inclusion defects [<a href="#B16-polymers-16-03261" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>Diagram of resin introduction layer structure.</p>
Full article ">Figure 4
<p>GFRP composite manufacturing process.</p>
Full article ">Figure 5
<p>Size diagram of delamination defect specimens (units: mm).</p>
Full article ">Figure 6
<p>Size diagram of wrinkle defect specimens (units: mm).</p>
Full article ">Figure 7
<p>Morphological features of defects: (<b>a</b>) artificial wrinkle defects; (<b>b</b>) natural wrinkle defects; (<b>c</b>) artificial delamination defects; (<b>d</b>) natural delamination defects.</p>
Full article ">Figure 8
<p>Acoustic emission testing system.</p>
Full article ">Figure 9
<p>Definition of acoustic emission parameters.</p>
Full article ">Figure 10
<p>Sound velocity calibration for acoustic emission lead break test.</p>
Full article ">Figure 11
<p>The schematic of pencil lead break test.</p>
Full article ">Figure 12
<p>Correlation tree diagram of acoustic emission signal parameters.</p>
Full article ">Figure 13
<p>Evaluation of the number of clusters: specimens A1, A2, A3, and A4.</p>
Full article ">Figure 14
<p>Time domain and frequency domain signals of different damage modes.</p>
Full article ">Figure 15
<p>Frequency and load changes with time for different specimens.</p>
Full article ">Figure 16
<p>Acoustic emission signal spectrum characteristics of specimen A1 at different damage stages.</p>
Full article ">Figure 17
<p>Displacement curve of specimens in tensile test.</p>
Full article ">Figure 18
<p>AE accumulation energy and accumulation count distribution over time: specimens A1, A2, A3, and A4.</p>
Full article ">Figure 19
<p>Tensile force and acoustic emission impact history of delamination defect specimens: specimens A1, A2, A3, and A4.</p>
Full article ">Figure 20
<p>Relationship between load and energy of wrinkle defect.</p>
Full article ">Figure 21
<p>Time history of AE amplitude and accumulated hits of wrinkle defect.</p>
Full article ">Figure 22
<p>Micrograph of delamination defect specimens after fracture.</p>
Full article ">Figure 23
<p>Micrograph of wrinkle defect specimens after fracture.</p>
Full article ">
Back to TopTop