Nothing Special   »   [go: up one dir, main page]

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

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
remove_circle_outline

Search Results (663)

Search Parameters:
Keywords = dry adhesives

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1658 KiB  
Article
Hydrothermal Liquefied Bio-Oil from Municipal Sewage Sludge as a Reactive Filler in Polymeric Diphenylmethane Diisocyanate (p-MDI) Wood Adhesives
by Archana Bansode, Tawsif Rahman, Lucila Carias, Osei Asafu-Adjaye, Sushil Adhikari, Brian K. Via, Ramsis Farag and Maria L. Auad
Sustainability 2025, 17(3), 1318; https://doi.org/10.3390/su17031318 - 6 Feb 2025
Viewed by 584
Abstract
The growing environmental concerns associated with petrochemical-based adhesives have driven interest in sustainable alternatives. This study investigates the use of bio-oil, derived from municipal sewage sludge (MSS) through hydrothermal liquefaction (HTL), as a reactive filler in polymeric methylene diphenyl diisocyanate (pMDI) wood adhesives. [...] Read more.
The growing environmental concerns associated with petrochemical-based adhesives have driven interest in sustainable alternatives. This study investigates the use of bio-oil, derived from municipal sewage sludge (MSS) through hydrothermal liquefaction (HTL), as a reactive filler in polymeric methylene diphenyl diisocyanate (pMDI) wood adhesives. The bio-oil, rich in hydroxyl and carbonyl functional groups, was characterized using FTIR (Fourier transform infrared spectroscopy), elemental analysis, and NMR (nuclear magnetic resonance). These functional groups interact with the isocyanate groups of pMDI, enabling crosslinking and enhancing adhesive performance. Various MSS bio-oil and pMDI formulations were evaluated for tensile shear strength on Southern yellow pine veneers under dry and wet conditions. The formulation with a 1:4 bio-oil to pMDI weight ratio exhibited the best performance, achieving tensile shear strengths of 1.96 MPa (dry) and 1.66 MPa (wet). Higher bio-oil content led to decreased adhesive strength, attributed to reduced crosslinking and increased moisture sensitivity. This study demonstrates the potential of MSS-derived bio-oil as a sustainable additive in pMDI adhesives, offering environmental benefits without significantly compromising adhesive performance and marking a step toward greener wood adhesive solutions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Preparation and application process of MSS bio-oil/pMDI adhesive on wood substrates.</p>
Full article ">Figure 2
<p>FTIR spectra of municipal sewage sludge bio-oil (MSS bio-oil).</p>
Full article ">Figure 3
<p><sup>13</sup>C–<sup>1</sup>H HSQC 2D-NMR spectra of municipal sewage sludge bio-oil (MSS bio-oil).</p>
Full article ">Figure 4
<p>Reaction of (<b>a</b>) pMDI adhesive with wood (<b>b</b>) MSS bio-oil/pMDI adhesive with wood.</p>
Full article ">Figure 5
<p>The tensile shear bonding strength of the wood specimens bonded with different MSS bio-oil/pMDI adhesives. Four wood-bonded specimens were tested and reported for each adhesive measurement in the form of mean and SD (error bars).</p>
Full article ">
10 pages, 2261 KiB  
Brief Report
The Adhesiveness of Chickpea (Cicer arietinum) Seeds Is Conditioned by Their Shape
by Andrey A. Sinjushin, Ana Uhlarik and Irina L. Stepanova
Int. J. Plant Biol. 2025, 16(1), 19; https://doi.org/10.3390/ijpb16010019 - 4 Feb 2025
Viewed by 330
Abstract
The seeds of some chickpea (Cicer arietinum L.) accessions are prone to sticking in twos or threes in a pod in the course of their maturation. Such seeds are usually easy to detach although their coats often become damaged due to forcible [...] Read more.
The seeds of some chickpea (Cicer arietinum L.) accessions are prone to sticking in twos or threes in a pod in the course of their maturation. Such seeds are usually easy to detach although their coats often become damaged due to forcible separation. Sticking is observed both in fields and glasshouses, with frequency potentially increasing in dry hot climates. Our morphometric survey of non-desi seeds (kabuli and intermediate types) suggests that it is seed shape, rather than size or color, that determines seed adhesiveness, with rounder seeds being the most prone to sticking. A similar phenomenon is known in pea (Pisum sativum L.) where it is conditioned by a single rare mutation affecting seed coat features. Unlike pea, numerous chickpea lines and cultivars of different origin have intrinsic susceptibility to seed adhesion, although to a variable extent depending on environment and seed shape, so this feature is multifactorial rather than solely genetic in C. arietinum. Although stuck seeds are mostly detached during mechanical harvesting, the accompanying seed coat lesions may be potentially undesired for seed storage and germination characteristics. Full article
(This article belongs to the Section Plant Reproduction)
Show Figures

Figure 1

Figure 1
<p>Stuck seed morphology in <span class="html-italic">Pisum sativum</span> ((<b>A</b>), line WL1238 homozygous at <span class="html-italic">s</span> mutation) and <span class="html-italic">Cicer arietinum</span> ((<b>B</b>–<b>F</b>), various accessions). In (<b>B</b>–<b>F</b>), the left image is a view from the pod suture (placenta), and the right one is from the opposite side. The seeds of chickpea can be glued together in twos (<b>B</b>,<b>D</b>–<b>F</b>) or threes (<b>C</b>), sometimes deformed (<b>D</b>) and either light- (<b>B</b>–<b>D</b>) or dark-colored (<b>E</b>,<b>F</b>). Scale bar: 1 cm.</p>
Full article ">Figure 2
<p>Damage on the seed coat (arrowheads) of stuck and forcibly detached seeds. Scale bar: 0.5 cm.</p>
Full article ">Figure 3
<p>The correlation between the roundness of seed contour and sticking frequency (<span class="html-italic">n</span> = 18). Each accession is represented with an exemplary seed silhouette with its shape visually classified as angular (red), ‘owl’s head’ (yellow) or rounded (green).</p>
Full article ">
24 pages, 7590 KiB  
Article
The Influence of Roughness of Surfaces on Wear Mechanisms in Metal–Rock Interactions
by Vlad Alexandru Florea, Mihaela Toderaș and Ciprian Danciu
Coatings 2025, 15(2), 150; https://doi.org/10.3390/coatings15020150 - 30 Jan 2025
Viewed by 600
Abstract
The processes of rock excavation and processing involve intense mechanical stresses on cutting, displacing, and transporting tools, inevitably leading to the phenomenon of dry friction wear. The factors influencing the intensity and mechanisms of wear are complex and interdependent, being conditioned by the [...] Read more.
The processes of rock excavation and processing involve intense mechanical stresses on cutting, displacing, and transporting tools, inevitably leading to the phenomenon of dry friction wear. The factors influencing the intensity and mechanisms of wear are complex and interdependent, being conditioned by the physical–mechanical properties of the rocks, the geometric characteristics and materials of the tools, as well as the cutting process parameters (cutting force, feed rate). Previous studies have mainly addressed the global aspect of wear without delving into the microstructural evolution of the contact surfaces during the friction process. In this paper, through controlled tribometric tests, we have investigated in detail the abrasive wear mechanisms of metallic materials in contact with different types of rocks, with an emphasis on the role played by surface roughness and the mineralogical properties of the rocks. Experimentally, we varied the applied forces and the number of friction cycles to simulate different working conditions and evaluate how these parameters influence wear intensity and surface morphology evolution. Microstructural analysis of the samples, combined with roughness measurements, allowed the identification of the predominant degradation mechanisms (abrasion, adhesion, fatigue) and their correlation with the material properties and the friction process parameters. The results have shown a strong correlation between the wear capacity of rocks and their petrographic properties, such as hardness, porosity, and hard mineral content. It was also found that the roughness of the contact surfaces plays an essential role in wear mechanisms, influencing both the initiation and propagation of its effects. Depending on the experimental data, we have developed a classification of rocks based on their abrasive potential and proposed criteria for the optimal adoption of materials and working parameters for the tools of technological equipment depending on the type of rock encountered. The results of this study can contribute to improving the durability of tools, as well as mining equipment, and reducing operating costs. Full article
(This article belongs to the Special Issue Friction and Wear Behaviors in Mechanical Engineering)
Show Figures

Figure 1

Figure 1
<p>The significance of tool geometry in the rock excavation process and the forces involved: (<b>a</b>) interaction between the cutting tool and the rock surface: F<sub>x</sub>—tangential force; F<sub>y</sub>—normal force; v<sub>t</sub>—cutting speed; α—clearance angle; β—rake angle; γ—cutting angle; δ—edge angle; (<b>b</b>) cross-section of the cutting zone: F<sub>z</sub>—lateral force relative to the tool path; h<sub>0</sub>—cutting depth; ψ—chip rupture angle; S<sub>0</sub>—cross-sectional area of the rock chip; 1—tool; 2—rock; 3—chip.</p>
Full article ">Figure 2
<p>Work stand: 1—TRB3 tribometer; 2—Sutronic S128 profilometer; 3—rock sample; 4—static partner (steel ball); 5—loading.</p>
Full article ">Figure 3
<p>Roughness parameters of the rock sample surface measured using a Sutronic S128 profilometer.</p>
Full article ">Figure 4
<p>Test regime parameters for the TRB3 tribometer.</p>
Full article ">Figure 5
<p>Wear assessment of the metal ball: (<b>a</b>) through mass loss measurement; (<b>b</b>) by measuring the wear track.</p>
Full article ">Figure 6
<p>The tested rock sample surface.</p>
Full article ">Figure 7
<p>Variation of the coefficient of friction on the rock sample surface.</p>
Full article ">Figure 8
<p>Rock samples tested using a TRB3 tribometer: 1—limestone; 2—Roșia Poieni andesite; 3—marble; 4—granite; 5—Albini Haneș andesite; 6—gabbro; 7—sandstone.</p>
Full article ">Figure 9
<p>Evolution of rock surface roughness with rock abrasiveness.</p>
Full article ">Figure 10
<p>The wear rate of a ball as a function of rock abrasiveness.</p>
Full article ">Figure 11
<p>The variation of the friction coefficient with rock abrasiveness.</p>
Full article ">
14 pages, 9128 KiB  
Article
Determining Moisture Condition of External Thermal Insulation Composite System (ETICS) of an Existing Building
by Paweł Krause, Iwona Pokorska-Silva and Łukasz Kosobucki
Materials 2025, 18(3), 614; https://doi.org/10.3390/ma18030614 - 29 Jan 2025
Viewed by 461
Abstract
ETICS is a popular external wall insulation system, which is not without possible defects and damages. A frequent cause, direct or indirect, of damage to buildings is the impact of water (moisture). This article presents, among others, the results of tests of the [...] Read more.
ETICS is a popular external wall insulation system, which is not without possible defects and damages. A frequent cause, direct or indirect, of damage to buildings is the impact of water (moisture). This article presents, among others, the results of tests of the moisture content of ETICS layers, the water absorption and capillary absorption of the render by means of the Karsten tube method, numerical thermo-moisture simulations, and tests of interlayer adhesion, in sample residential buildings. Mass moisture content testing of the wall substrate showed acceptable moisture levels (1–4%m) within masonry walls made of silicate blocks, as well as locally elevated moisture levels (4–8%m) in the case of reinforced concrete walls. Moisture testing of the insulation samples showed a predominantly dry condition, and testing of the reinforcement layer showed an acceptable level of moisture. Severe moisture was found in the sample taken in the ground-floor zone at the interface between mineral wool and EPS-P insulation underneath the reinforced layer. Capillary water absorption tests helped classify silicone render as an impermeable and surface hydrophobic coating. Tests of the water absorption of the facade plaster showed that the value declared by the manufacturer (<0.5 kg/m2) was mostly met (not in the ground-floor zone). The simulation calculations gave information that there was no continuous increase in condensation during the assumed analysis time (the influence of interstitial condensation on the observed anomalies was excluded). The tests carried out indicated the occurrence of numerous errors in the implementation of insulation works affecting the moisture content and durability of external partitions. Full article
(This article belongs to the Section Construction and Building Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Extensive detachment of thin-coat plaster from the reinforced layer along the unplastered strip separating plasters of different colour. (<b>b</b>) Fragment of the north elevation of the building in the ground-floor area—detachment and loss of thin-coat plaster. (<b>c</b>) Fragment of the north elevation of the building—spalling and loss of thin-coat plaster with exposed reinforcing mesh.</p>
Full article ">Figure 2
<p>(<b>a</b>) Uncovering cut-out P6: Exposure of wall substrates. (<b>b</b>) Uncovering cut-out P2: Bore P2a in the wall construction substrate.</p>
Full article ">Figure 3
<p>(<b>a</b>) Uncovering cut-out P9: Measurement of the thickness of the air void between the insulation and the wall substrate (ca. 2.5 cm). Thickness of the adhesive mortar exceeding 1 cm. (<b>b</b>) Uncovering cut-out P7: Measuring tape entering under the insulation at the joint of the polystyrene boards (no perimeter distribution of adhesive mortar).</p>
Full article ">Figure 4
<p>(<b>a</b>) Sample P9: Loss of facade render and a section of adhesive mortar exposing fibreglass mesh in the reinforced layer. (<b>b</b>) Sample P10: Measurement of the thickness of the thin-bed plaster and reinforced layer system (insufficient reinforced layer thickness, less than 3 mm).</p>
Full article ">Figure 5
<p>(<b>a</b>) Sample P1: View of the cross-section of the reinforced layer and the thin-coat plaster in the sample. The reinforcing mesh visible in the cross-section is not located in the middle of the thickness of the reinforced layer (deterioration of the mechanical properties of the reinforced layer). (<b>b</b>) Sample P11: View of the cross-section of the reinforced layer and the thin-coat plaster in the sample. The reinforcing mesh visible in the cross-section is not located in the middle of the thickness of the reinforced layer (deterioration of the mechanical properties of the reinforced layer).</p>
Full article ">Figure 6
<p>(<b>a</b>) Sample P1: Visually and organoleptically perceptible dampness of the mineral wool at the interface with the EPS-P insulation underneath the reinforced layer. (<b>b</b>) Uncovering cut-out P1: Lack of adhesion of bituminous membrane edge to the reinforced concrete wall substrate. No proper edge trim of the membrane with a pressure strip.</p>
Full article ">Figure 7
<p>(<b>a</b>) Sample P9: Close-up of the surface of the facade plaster; visible cracks in the plaster; (<b>b</b>) Sample P9: Close-up of the surface of the facade plaster using an optical device; visible micro-cracks in the plaster.</p>
Full article ">Figure 8
<p>(<b>a</b>) Mass moisture content of the construction substrate obtained with the Testo 635-2 moisture meter (green—silicate block substrate, orange—reinforced concrete substrate). (<b>b</b>) Moisture content of the construction substrate, insulation layer, reinforced layer, and plaster, measured with a weighing dryer.</p>
Full article ">Figure 9
<p>(<b>a</b>) Karsten tube water absorption coefficient test for plaster. (<b>b</b>) Water absorption coefficients determined using the Karsten tube method.</p>
Full article ">Figure 10
<p>(<b>a</b>) Water absorption of plaster. (<b>b</b>) Adhesion of the surface layer.</p>
Full article ">Figure 11
<p>(<b>a</b>) Diagram of moisture content [kg/m<sup>3</sup>] in the wall construction layer (silicate blocks) |1|. (<b>b</b>) Graph of moisture content [kg/m<sup>3</sup>] in the adhesive mortar layer fixing the insulation to the wall substrate |2|.</p>
Full article ">Figure 12
<p>(<b>a</b>) Graph of moisture content [kg/m<sup>3</sup>] in the EPS 031 insulation layer |3|. (<b>b</b>) Graph of moisture content [kg/m<sup>3</sup>] in the reinforced layer |4|.</p>
Full article ">Figure 13
<p>(<b>a</b>) Graph of moisture content [kg/m<sup>3</sup>] in the thin-coat plaster layer |5|. (<b>b</b>) Graph of moisture content [%m] in the layers of the envelope |1|–|5|, 6th year.</p>
Full article ">
24 pages, 673 KiB  
Review
The Impact of Fluid Flow on Microbial Growth and Distribution in Food Processing Systems
by Zainab Talib Al-Sharify, Shahad Zuhair Al-Najjar, Zainab A. Naser, Zinah Amer Idrees Alsherfy and Helen Onyeaka
Foods 2025, 14(3), 401; https://doi.org/10.3390/foods14030401 - 26 Jan 2025
Viewed by 728
Abstract
This article examines the impact of fluid flow dynamics on microbial growth, distribution, and control within food processing systems. Fluid flows, specifically laminar and turbulent flows, significantly influence microbial behaviors, such as biofilm development and microbial adhesion. Laminar flow is highly conducive to [...] Read more.
This article examines the impact of fluid flow dynamics on microbial growth, distribution, and control within food processing systems. Fluid flows, specifically laminar and turbulent flows, significantly influence microbial behaviors, such as biofilm development and microbial adhesion. Laminar flow is highly conducive to biofilm formation and microbial attachment because the flow is smooth and steady. This smooth flow makes it much more difficult to sterilize the surface. Turbulent flow, however, due to its chaotic motion and the shear forces that are present, inhibits microbial growth because it disrupts attachment; however, it also has the potential to contaminate surfaces by dispersing microorganisms. Computational fluid dynamics (CFD) is highlighted as an essential component for food processors to predict fluid movement and enhance numerous fluid-dependent operations, including mixing, cooling, spray drying, and heat transfer. This analysis underscores the significance of fluid dynamics in controlling microbial hazards in food settings, and it discusses some interventions, such as antimicrobial surface treatments and properly designed equipment. Each process step from mixing to cooling, which influences heat transfer and microbial control by ensuring uniform heat distribution and optimizing heat removal, presents unique fluid flow requirements affecting microbial distribution, biofilm formation, and contamination control. Food processors can improve microbial management and enhance product safety by adjusting flow rates, types, and equipment configurations. This article helps provide an understanding of fluid–microbe interactions and offers actionable insights to advance food processing practices, ensuring higher standards of food safety and quality control. Full article
(This article belongs to the Section Food Engineering and Technology)
Show Figures

Figure 1

Figure 1
<p>Applications of fluid flow in food processing.</p>
Full article ">
17 pages, 21034 KiB  
Article
Exploring the Effect of Ti on Mechanical and Tribological Properties of an AlCrFe2Ni2Tix High-Entropy Alloy
by Yajuan Shi, Yudong Guo and Yi Wang
Metals 2025, 15(2), 121; https://doi.org/10.3390/met15020121 - 26 Jan 2025
Viewed by 480
Abstract
Low friction and wear constitute a challenge for metallic materials under dry sliding conditions. In the current study, we successfully prepared an AlCrFe2Ni2Tix (x = 0, 0.2, 0.4, 0.6, 0.8, 1.0) high-entropy alloy (HEA) consisting of a body-centered [...] Read more.
Low friction and wear constitute a challenge for metallic materials under dry sliding conditions. In the current study, we successfully prepared an AlCrFe2Ni2Tix (x = 0, 0.2, 0.4, 0.6, 0.8, 1.0) high-entropy alloy (HEA) consisting of a body-centered cubic (BCC) phase and an AlNi2Ti phase that exhibited an outstanding combination of a compression strength of above 3 GPa and a ductility degree of 26% at room temperature. Under a 20 N load, the dry friction tests showed that AlCrFe2Ni2Ti0.4 HEA had the lowest wear volume (1.498 mm3), with a coefficient of friction of 0.3929. It is related to the volume fraction of AlNi2Ti precipitate increasing with increasing Ti content, thus resulting in better wear resistance. Through the strengthening mechanism analysis, it is crucial to manipulate the composition of the AlNi2Ti precipitate to obtain desirable mechanical properties in the AlCrFe2Ni2Tix HEA. The main mechanism of wear friction is identified as adhesion wear. Therefore, the addition of Ti into AlCrFe2Ni2 HEA can effectively improve its mechanical and wear resistance due to the significant improvement in hardness and its inherent solution strengthening. Our study provides a new strategy for designing new BCC HEAs with a combination of high hardness, yield strength, and excellent wear. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> HEAs. (<b>a</b>) Overall diffraction peaks; (<b>b</b>) localized diffraction peaks.</p>
Full article ">Figure 2
<p>Metallography images of the as-cast AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> alloys: (<b>a</b>) Ti<sub>0</sub>, (<b>c</b>) Ti<sub>0.2,</sub> (<b>e</b>) Ti<sub>0.4</sub>, (<b>g</b>) Ti<sub>0.6</sub>, (<b>i</b>) Ti<sub>0.8</sub>, and (<b>k</b>) Ti<sub>1.0</sub>; (<b>b</b>), (<b>d</b>), (<b>f</b>), (<b>h</b>), (<b>j</b>), and (<b>l</b>) are enlarged images of (<b>a</b>), (<b>c</b>), (<b>e</b>), (<b>g</b>), (<b>i</b>), and (<b>k</b>), respectively.</p>
Full article ">Figure 3
<p>SEM images of AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> alloys in the as-cast state. (<b>a</b>) Ti<sub>0</sub>, (<b>c</b>) Ti<sub>0.2</sub>, (<b>e</b>) Ti<sub>0.4,</sub> (<b>g</b>) Ti<sub>0.6</sub>, (<b>i</b>) Ti<sub>0.8</sub>, (<b>k</b>) Ti<sub>1</sub>; (<b>b</b>), (<b>d</b>), (<b>f</b>), (<b>h</b>), (<b>j</b>), and (<b>l</b>) are the corresponding enlarged images of (<b>a</b>), (<b>c</b>), (<b>e</b>), <b>(g</b>), (<b>i</b>), and (<b>k</b>) respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Compressive engineering stress–strain curves of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> alloys. (<b>b</b>) Yield strength and Vickers hardness of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> alloys.</p>
Full article ">Figure 5
<p>TEM microstructure of the as-cast AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> high-entropy alloy. (<b>a</b>) Bright-field image. (<b>b</b>) and (<b>c</b>) are the SAED patterns of the bright and dark regions shown in (<b>a</b>), respectively. (<b>d</b>) TEM-EDS elemental mappings.</p>
Full article ">Figure 6
<p>(<b>a</b>) HRTEM micrograph of the Ti<sub>0</sub>.<sub>1</sub> alloy showing the phase interface between the BCC phase and the AlNi<sub>2</sub>Ti phase. The red dotted line represents the phase boundary; (<b>b</b>,<b>c</b>) corresponding FFT patterns of the BCC phase and the AlNi<sub>2</sub>Ti phase in (<b>a</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic of the wear of AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> and (<b>b</b>) tribological samples.</p>
Full article ">Figure 8
<p>Coefficient of friction of all the alloys at applied wear loads of (<b>a</b>) 5 N, (<b>b</b>) 10 N, and (<b>c</b>) 20 N; (<b>d</b>) overall distribution of the friction coefficient of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> alloys compared with that of 316 steel.</p>
Full article ">Figure 9
<p>Confocal laser three-dimensional (3D) wear surface morphology of AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> (x = 0, 0.4, 0.6, 0.8, 1) under 5 N: (<b>a</b>) Ti<sub>0</sub>, (<b>b</b>) Ti<sub>0.2</sub>, (<b>c</b>) Ti<sub>0.4</sub>, (<b>d</b>) Ti<sub>0.6</sub>, (<b>e</b>) Ti<sub>0.8</sub>, (<b>f</b>) Ti<sub>1</sub>.</p>
Full article ">Figure 10
<p>Confocal laser 3D wear surface morphology of AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> (x = 0, 0.4, 0.6, 0.8, 1) under 10 N: (<b>a</b>) Ti<sub>0</sub>, (<b>b</b>) Ti<sub>0.2</sub>, (<b>c</b>) Ti<sub>0.4</sub>, (<b>d</b>) Ti<sub>0.6</sub>, (<b>e</b>) Ti<sub>0.8</sub>, (<b>f</b>) Ti<sub>1</sub>.</p>
Full article ">Figure 11
<p>Confocal laser 3D wear surface morphology of AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> (x = 0, 0.4, 0.6, 0.8, 1) under 20 N: (<b>a</b>) Ti<sub>0</sub>, (<b>b</b>) Ti<sub>0.2</sub>, (<b>c</b>) Ti<sub>0.4</sub>, (<b>d</b>) Ti<sub>0.6</sub>, (<b>e</b>) Ti<sub>0.8</sub>, (<b>f</b>) Ti<sub>1.0</sub>.</p>
Full article ">Figure 12
<p>Wear volume trends of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> HEAs at loads of 5 N, 10 N, and 20 N.</p>
Full article ">Figure 13
<p>SEM micrographs of the worn surface after the wear tests at a wear load of 5 N. (<b>a</b>) Ti<sub>0</sub>, (<b>c</b>) Ti<sub>0.2</sub>, (<b>e</b>) Ti<sub>0.4</sub>, (<b>g</b>) Ti<sub>0.6</sub>, (<b>i</b>) Ti<sub>0.8</sub>, and (<b>k</b>) Ti<sub>1</sub>; (<b>b</b>), (<b>d</b>), (<b>f</b>), (<b>h</b>), (<b>j</b>), and (<b>l</b>) are enlarged images of the parts marked in red boxes of (<b>a</b>), (<b>c</b>), (<b>e</b>), (<b>g</b>), (<b>i</b>), and (<b>k</b>), respectively.</p>
Full article ">Figure 14
<p>SEM micrographs of the worn surface after wear tests at a wear load of 10 N. (<b>a</b>) Ti<sub>0</sub>, (<b>c</b>) Ti<sub>0.2</sub>, (<b>e</b>) Ti<sub>0.4</sub>, (<b>g</b>) Ti<sub>0.6</sub>, (<b>i</b>) Ti<sub>0.8</sub>, (<b>k</b>) Ti<sub>1.0</sub>; (<b>b</b>), (<b>d</b>), (<b>f</b>), (<b>h</b>), (<b>j</b>), and (<b>l</b>) are enlarged images of the parts marked in red boxes in (<b>a</b>), (<b>c</b>), (<b>e</b>), (<b>g</b>), (<b>i</b>), and (<b>k</b>), respectively.</p>
Full article ">Figure 15
<p>SEM micrographs of the worn surface after wear tests at a wear load of 20 N. (<b>a</b>) Ti<sub>0</sub>, (<b>c</b>) Ti<sub>0.2</sub>, (<b>e</b>) Ti<sub>0.4</sub>, (<b>g</b>) Ti<sub>0.6</sub>, (<b>i</b>) Ti<sub>0.8</sub>, (<b>k</b>) Ti<sub>1</sub>; (<b>b</b>), (<b>d</b>), (<b>f</b>), (<b>h</b>), (<b>j</b>), and (<b>l</b>) are the enlarged images of the parts marked in red boxes in (<b>a</b>), (<b>c</b>), (<b>e</b>), (<b>g</b>), (<b>i</b>), and (<b>k</b>), respectively.</p>
Full article ">Figure 16
<p>Elemental mapping of worn surfaces under 20 N. (<b>a</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>, (<b>b</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>0.2</sub>, (<b>c</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>0.4</sub>, (<b>d</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>0.6</sub>, (<b>e</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>0.8</sub>, (<b>f</b>) AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>1.0</sub>.</p>
Full article ">Figure 17
<p>Schematic diagram of the wear mechanism of the AlCrFe<sub>2</sub>Ni<sub>2</sub>Ti<sub>x</sub> HEAs.</p>
Full article ">
14 pages, 7247 KiB  
Article
Development of Recombinant Human Collagen-Based Porous Scaffolds for Skin Tissue Engineering: Enhanced Mechanical Strength and Biocompatibility
by Yang Yang, Ting Yu, Mengdan Tao, Yong Wang, Xinying Yao, Chenkai Zhu, Fengxue Xin and Min Jiang
Polymers 2025, 17(3), 303; https://doi.org/10.3390/polym17030303 - 23 Jan 2025
Viewed by 479
Abstract
Skin tissue engineering scaffolds should possess key properties such as porosity, degradability, durability, and biocompatibility to effectively facilitate skin cell adhesion and growth. In this study, recombinant human collagen (RHC) was used to fabricate porous scaffolds via freeze-drying, offering an alternative to animal-derived [...] Read more.
Skin tissue engineering scaffolds should possess key properties such as porosity, degradability, durability, and biocompatibility to effectively facilitate skin cell adhesion and growth. In this study, recombinant human collagen (RHC) was used to fabricate porous scaffolds via freeze-drying, offering an alternative to animal-derived collagen where bovine collagen (BC)-based scaffolds were also prepared for comparison. The internal morphology of the RHC scaffolds were characterized by scanning electron microscopy (SEM) and the pore size ranged from 68.39 to 117.52 µm. The results from compression and fatigue tests showed that the mechanical strength and durability of RHC scaffolds could be tailored by adjusting the RHC concentration, and the maximum compressive modulus reached to 0.003 MPa, which is comparable to that of BC scaffolds. The degradation test illustrated that the RHC scaffolds had a slower degradation rate compared to BC scaffolds. Finally, the biocompatibilities of the porous scaffolds were studied by seeding and culturing the human foreskin fibroblasts (HFFs) and human umbilical vein endothelial cells (HUVECs) in samples. The fluorescent images and Cell Counting Kit-8 (CCK-8) assay revealed RHC porous scaffolds were non-cytotoxic and supported the attachment as well as the proliferation of the seeded cells. Overall, the results demonstrated that RHC-based scaffolds exhibited adequate mechanical strength, ideal biodegradability, and exceptional biocompatibility, making them highly suitable for skin-tissue-engineering applications. Full article
(This article belongs to the Special Issue Biopolymers for Drug Delivery and Tissue Engineering)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for designed collagen scaffolds.</p>
Full article ">Figure 2
<p>Fourier infrared spectroscopy analysis of RHC scaffolds with or without EDC crosslinking.</p>
Full article ">Figure 3
<p>The internal morphology of scaffolds. (<b>a</b>) Pore structure of the freeze-dried sample, scale bar = 100 μm; (<b>b</b>) the average pore size of the scaffolds. Significance is indicated with * <span class="html-italic">p</span> ≤ 0.05 and *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 4
<p>Mechanical properties of the samples. (<b>a</b>) Stress–strain curves of the RHC and BC scaffolds; (<b>b</b>) compressive modulus of the RHC and BC scaffolds; and (<b>c</b>) compressive stress of the RHC and BC scaffolds. Significance is indicated with *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 5
<p>The compressive stress ratio of σ<sub>n</sub>/σ<sub>1</sub> of four scaffolds as a function of time under 20 cycles.</p>
Full article ">Figure 6
<p>Degradation rates of the RHC and BC scaffolds at selected time points. Significance is indicated with *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 7
<p>Cytotoxicity test of the samples. (<b>a</b>) Live/dead fluorescence staining of HFFs and HUVECs cells, scale bar = 100 μm; (<b>b</b>) the viability of HFFs; and (<b>c</b>) the viability of HUVECs. Significance is indicated with * <span class="html-italic">p</span> ≤ 0.05, and not statistically significant is indicated with ns.</p>
Full article ">Figure 8
<p>In vitro test of cell seeded scaffolds. (<b>a</b>) Fluorescence staining of HFFs (green) and HUVECs (red) within scaffolds at day 3, scale bar = 500 μm; (<b>b</b>) fluorescent staining of cell proliferation of HFFs and HUVECs within scaffolds at day 7, scale bar = 500 μm; (<b>c</b>) viabilities of HFFs; and (<b>d</b>) viabilities of HUVECs. Significance is indicated with * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, and *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">
26 pages, 5313 KiB  
Review
Interfacial Interlocking of Carbon Fiber-Reinforced Polymer Composites: A Short Review
by Jong-Hyun Joo, Seong-Hwang Kim, Yoon-Ji Yim, Jin-Seok Bae and Min-Kang Seo
Polymers 2025, 17(3), 267; https://doi.org/10.3390/polym17030267 - 21 Jan 2025
Viewed by 706
Abstract
The mechanical properties of the carbon fiber-reinforced polymer composites (CFRPs) are dependent on the interfacial interaction and adhesion between carbon fibers (CFs) and polymer matrices. Therefore, it is crucial to understand how modifying the CFs can influence the properties of these composites. This [...] Read more.
The mechanical properties of the carbon fiber-reinforced polymer composites (CFRPs) are dependent on the interfacial interaction and adhesion between carbon fibers (CFs) and polymer matrices. Therefore, it is crucial to understand how modifying the CFs can influence the properties of these composites. This review outlines recent research progress with a focus on the relationship between the interfacial and mechanical properties of CFRPs and provides a systematic summary of state-of-the-art surface modification techniques. These techniques are divided into four categories: (i) wet, (ii) electrochemical, (iii) dry, and (iv) polymer matrix modifications. Several strategies for enhancing the interfacial interactions and adhesion of CFRPs are discussed, providing insights for future trends. Full article
Show Figures

Figure 1

Figure 1
<p>Statistical survey of the literature based on keywords in the field of CFRPs over a 12-year period (2014–2024). The number of publications related to CFRP interfaces is represented by blue dots. Checked on Web of Science as of June 2024.</p>
Full article ">Figure 2
<p>Schematic 3D cross-sections of CFRP interfaces include (<b>a</b>) a cross-section and (<b>b</b>) the interface and interphases.</p>
Full article ">Figure 3
<p>Interfacial properties of acid-treated CFRPs: (<b>a</b>) Comparison of the simulated and experimentally measured shear strength (<b>left</b>) and the proposed shear mechanism after treatment with a HNO<sub>3</sub>/H<sub>2</sub>SO<sub>4</sub> solution (<b>right</b>); reprinted from Ref. [<a href="#B51-polymers-17-00267" class="html-bibr">51</a>], Copyright 2019, Elsevier. (<b>b</b>) Surface SEM and AFM images of the untreated CFs (C<sub>0</sub>) and the CFs after treatment with HNO<sub>3</sub> for 90 min (C<sub>90</sub>) (<b>left</b>), and a chart showing the correlation between wear resistance and ILSS according to treatment time (<b>right</b>); reprinted from Ref. [<a href="#B52-polymers-17-00267" class="html-bibr">52</a>], Copyright 2012, Elsevier.</p>
Full article ">Figure 4
<p>Interfacial properties of amine-treated CFRPs: (<b>a</b>) Schematic illustration of poly(oxypropylene) diamines grafted to the CF surface in water (<b>left</b>) and a bar chart showing the IFSS values of the CFRPs according to the D400 concentration (<b>right</b>); reprinted from Ref. [<a href="#B57-polymers-17-00267" class="html-bibr">57</a>], Copyright 2017, Elsevier. (<b>b</b>) A schematic diagram (<b>left</b>) showing the functionalization progress and molecular structures of two types of CFs, and a bar chart (<b>right</b>) showing the ILSS values of the corresponding CFRPs; reprinted from Ref. [<a href="#B58-polymers-17-00267" class="html-bibr">58</a>], Copyright 2016, Elsevier.</p>
Full article ">Figure 5
<p>Interfacial properties of silane-treated CFRPs: (<b>a</b>) Schematic illustration (<b>left</b>) of a two-step silane treatment, and a bar chart (<b>right</b>) showing the tensile strengths of the resulting CFRPs; reprinted from Ref. [<a href="#B62-polymers-17-00267" class="html-bibr">62</a>], Copyright 2019, Elsevier. (<b>b</b>) A schematic diagram (<b>left</b>) showing a possible reaction mechanism between CFs and the PU matrix, and a bar chart (<b>right</b>) showing the IFSS of CFRP values of various samples; reprinted from Ref. [<a href="#B63-polymers-17-00267" class="html-bibr">63</a>], Copyright 2015, Elsevier.</p>
Full article ">Figure 6
<p>Interfacial properties of CFRPs when treated with sizing agents: (<b>a</b>) A photographic image and schematic diagram of a continuous fiber sizing tower system used to size the CFs (<b>left</b>), and a bar chart showing the fracture toughness of CFRPs without and with various sizing agents (<b>right</b>); reprinted from Ref. [<a href="#B68-polymers-17-00267" class="html-bibr">68</a>], Copyright 2016, Elsevier. (<b>b</b>) A schematic diagram showing the correlation between surface energy and the distribution of cross-linking points on the CFRPs (<b>left</b>), and a bar chart showing the comparative IFSS values (<b>right</b>); reprinted from Ref. [<a href="#B69-polymers-17-00267" class="html-bibr">69</a>], Copyright 2020, Elsevier. (<b>c</b>) A schematic diagram showing the CFs sizing process (<b>left</b>), and a bar chart showing the ILSS values of the various PEEK-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B70-polymers-17-00267" class="html-bibr">70</a>], Copyright 2020, Elsevier.</p>
Full article ">Figure 7
<p>Interfacial properties of carbon nanotube-coated CFRPs: (<b>a</b>) TEM image of the CNTs grown on the CF surface (<b>left</b>) and a bar chart showing the IFSS values of various CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B14-polymers-17-00267" class="html-bibr">14</a>], Copyright 2022, Elsevier. (<b>b</b>) A schematic diagram showing the surface modification of CFs with GO-PDA (<b>left</b>) and SEM image of the CFs after treatment with GO-PDA for 24 h (<b>right</b>); reprinted from Ref. [<a href="#B76-polymers-17-00267" class="html-bibr">76</a>], Copyright 2021, Elsevier. (<b>c</b>) SEM images and schematic diagrams of the CFs before and after sequential modification with GO and electropolymerization (<b>left</b>), and a bar chart showing the ILSS values of the various CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B77-polymers-17-00267" class="html-bibr">77</a>], Copyright 2020, Elsevier.</p>
Full article ">Figure 8
<p>Interfacial properties of CVD-treated CFRPs: (<b>a</b>) TEM images showing the growth of CNTs on the CF surface (<b>left</b>) and a bar chart showing the ILSS values of the epoxy-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B83-polymers-17-00267" class="html-bibr">83</a>], Copyright 2020, Elsevier. (<b>b</b>) A schematic diagram showing the equipment for the continuous in situ growth of CNTs on the moving CF surface (<b>left</b>) and a bar chart showing the IFSS values of the epoxy-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B84-polymers-17-00267" class="html-bibr">84</a>], Copyright 2021, Elsevier. (<b>c</b>) SEM (<b>upper left</b>) and AFM (<b>lower left</b>) images showing the CFs before and after the growth of CB, and a bar chart showing the tensile and impact strengths of the epoxy-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B85-polymers-17-00267" class="html-bibr">85</a>], Copyright 2017, Elsevier.</p>
Full article ">Figure 9
<p>Interfacial properties of EPD-treated CFRPs: (<b>a</b>) SEM image and schematic diagram of the frame-mounted CFs being lowered into the EPD tank containing an O-CNF solution (<b>left</b>) and a bar chart showing the ILSS values of the various epoxy-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B91-polymers-17-00267" class="html-bibr">91</a>], Copyright 2011, Elsevier. (<b>b</b>) A schematic diagram of the continuous EPD deposition of CNTs onto the CF surface (<b>left</b>) and a bar chart showing the ILSS values of the epoxy-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B92-polymers-17-00267" class="html-bibr">92</a>], Copyright 2015, Elsevier. (<b>c</b>) A schematic diagram showing the alignment of CNTs onto the CFs (<b>left</b>) and a bar chart showing the comparative IFSS values (<b>right</b>); reprinted from Ref. [<a href="#B93-polymers-17-00267" class="html-bibr">93</a>], Copyright 2020, Elsevier.</p>
Full article ">Figure 10
<p>Interfacial properties of plasma-treated CFRPs: (<b>a</b>) SEM images of the CF surfaces before and after various treatment times (<b>left</b>) and a bar chart showing the IFSS values of the bismaleimide-based CFRPs after various treatment times (<b>right</b>); reprinted from Ref. [<a href="#B97-polymers-17-00267" class="html-bibr">97</a>], Copyright 2018, Wiley. (<b>b</b>) A schematic diagram of the APPJ system (<b>left</b>) and a bar chart showing the shear strengths of the formaldehyde-based CFRPs after various treatment times (<b>right</b>); reprinted from Ref. [<a href="#B98-polymers-17-00267" class="html-bibr">98</a>], Copyright 2019, Elsevier. (<b>c</b>) A schematic diagram of CFs in a plasma processor (<b>left</b>) and a plot of the ILSS value against treatment time for the vinyl ester-based CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B99-polymers-17-00267" class="html-bibr">99</a>], Copyright 2020, IOPscience.</p>
Full article ">Figure 11
<p>Interfacial properties of ozone-treated CFRPs: (<b>a</b>) Variations in the surface free energy (<b>left</b>) and fracture toughness (K<sub>IC</sub>) of epoxy-based CFRPs according to the ozone treatment time (<b>right</b>); reprinted from Ref. [<a href="#B103-polymers-17-00267" class="html-bibr">103</a>], Copyright 2005, Elsevier. (<b>b</b>) AFM and fracture surface (<b>left</b>), and a bar chart showing the flexural and compressive strengths of CFRPs under various treatment conditions (<b>right</b>); reprinted from Ref. [<a href="#B104-polymers-17-00267" class="html-bibr">104</a>], Copyright 2006, Elsevier.</p>
Full article ">Figure 12
<p>Interfacial properties of nanofiller-based CFRPs: (<b>a</b>) Schematic diagrams showing the self-assembly of NDI/MWCNT/HMCF (<b>left</b>) and a bar chart showing the IFSS values of the various CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B108-polymers-17-00267" class="html-bibr">108</a>], Copyright 2019, Elsevier. (<b>b</b>) A chart showing the fracture toughness (K<sub>IC</sub>) values (<b>left</b>) and schematic diagrams showing the cracking mechanism of the OCB loaded CFRPs (<b>right</b>); reprinted from Ref. [<a href="#B109-polymers-17-00267" class="html-bibr">109</a>], Copyright 2019, Elsevier. (<b>c</b>) Bar charts showing the ILSS values (<b>left</b>) and IFSS values (<b>right</b>) of CFRPs with various concentrations of SGO; reprinted from Ref. [<a href="#B110-polymers-17-00267" class="html-bibr">110</a>], Copyright 2014, Elsevier.</p>
Full article ">Figure 13
<p>Interfacial properties of nanohybrid-based CFRPs: (<b>a</b>) Var chart showing the K<sub>IC</sub> values at various filler contents (<b>left</b>) and a schematic diagram of the proposed fracture mechanism in the GO-GNF-loaded CFRP (<b>right</b>); reprinted from Ref. [<a href="#B115-polymers-17-00267" class="html-bibr">115</a>], Copyright 2021, Elsevier. (<b>b</b>) Plots of the flexural strength (<b>left</b>) and initial fracture toughness (<b>right</b>) of the CFRPs with various loadings of CNTs/GNPs; reprinted from Ref. [<a href="#B116-polymers-17-00267" class="html-bibr">116</a>], Copyright 2015, Wiley.</p>
Full article ">
32 pages, 8060 KiB  
Article
Study on Robust Path-Tracking Control for an Unmanned Articulated Road Roller Under Low-Adhesion Conditions
by Wei Qiang, Wei Yu, Quanzhi Xu and Hui Xie
Electronics 2025, 14(2), 383; https://doi.org/10.3390/electronics14020383 - 19 Jan 2025
Viewed by 547
Abstract
To enhance the path-tracking accuracy of unmanned articulated road roller (UARR) operating on low-adhesion, slippery surfaces, this paper proposes a hierarchical cascaded control (HCC) architecture integrated with real-time ground adhesion coefficient estimation. Addressing the complex nonlinear dynamics between the two rigid bodies of [...] Read more.
To enhance the path-tracking accuracy of unmanned articulated road roller (UARR) operating on low-adhesion, slippery surfaces, this paper proposes a hierarchical cascaded control (HCC) architecture integrated with real-time ground adhesion coefficient estimation. Addressing the complex nonlinear dynamics between the two rigid bodies of the vehicle and its interaction with the ground, an upper-layer nonlinear model predictive controller (NMPC) is designed. This layer, based on a 4-degree-of-freedom (4-DOF) dynamic model, calculates the required steering torque using position and heading errors. The lower layer employs a second-order sliding mode controller (SOSMC) to precisely track the steering torque and output the corresponding steering wheel angle. To accommodate the anisotropic and time-varying nature of slippery surfaces, a strong-tracking unscented Kalman filter (ST-UKF) observer is introduced for ground adhesion coefficient estimation. By dynamically adjusting the covariance matrix, the observer reduces reliance on historical data while increasing the weight of new data, significantly improving real-time estimation accuracy. The estimated adhesion coefficient is fed back to the upper-layer NMPC, enhancing the control system’s adaptability and robustness under slippery conditions. The HCC is validated through simulation and real-vehicle experiments and compared with LQR and PID controllers. The results demonstrate that HCC achieves the fastest response time and smallest steady-state error on both dry and slippery gravel soil surfaces. Under slippery conditions, while control performance decreases compared to dry surfaces, incorporating ground adhesion coefficient observation reduces steady-state error by 20.62%. Full article
(This article belongs to the Section Electrical and Autonomous Vehicles)
Show Figures

Figure 1

Figure 1
<p>UARR hardware layout.</p>
Full article ">Figure 2
<p>Causality-based modeling simulation platform for road roller.</p>
Full article ">Figure 3
<p>Force analysis of UARR dual bodies: (<b>a</b>) 3D force analysis; (<b>b</b>) planar force analysis and structural parameters.</p>
Full article ">Figure 4
<p>Fitting the Dugoff model to shearing stress–shearing displacement data [<a href="#B39-electronics-14-00383" class="html-bibr">39</a>].</p>
Full article ">Figure 5
<p>Feasibility verification of the Dugoff model for the drum: (<b>a</b>) circular test; (<b>b</b>) X coordinate.</p>
Full article ">Figure 6
<p>Equivalent schematic of the UARR hydraulic steering system.</p>
Full article ">Figure 7
<p>Relative displacement between the valve spool and valve sleeve.</p>
Full article ">Figure 8
<p>Principle diagram of piston rod movement.</p>
Full article ">Figure 9
<p>Validation of the dynamics model on wet dirt road: (<b>a</b>) yaw angle; (<b>b</b>) yaw rate.</p>
Full article ">Figure 10
<p>Validation of the dynamics model on wet gravel road: (<b>a</b>) yaw angle; (<b>b</b>) yaw rate.</p>
Full article ">Figure 11
<p>Validation of the dynamics model on wet dirt road: (<b>a</b>) drum centroid latitude; (<b>b</b>) drum centroid longitude.</p>
Full article ">Figure 12
<p>Validation of the dynamics model on wet gravel road: (<b>a</b>) drum centroid latitude; (<b>b</b>) drum centroid longitude.</p>
Full article ">Figure 13
<p>Hierarchical cascaded framework integrating NMPC and SOSMC with adhesion coefficient estimation.</p>
Full article ">Figure 14
<p>SOSMC tracking performance verification: (<b>a</b>) tracking target steering torque <span class="html-italic">M<sub>j</sub></span>; (<b>b</b>) corresponding steering wheel angle output.</p>
Full article ">Figure 15
<p>(<b>a</b>) Experimental scenario; (<b>b</b>) dry surface; (<b>c</b>) slippery surface.</p>
Full article ">Figure 16
<p>Step-tracking experiment under dry conditions: (<b>a</b>) lateral error; (<b>b</b>) steady-state error distribution.</p>
Full article ">Figure 17
<p>Straight line experiment under dry conditions: (<b>a</b>) lateral error; (<b>b</b>) lateral error distribution.</p>
Full article ">Figure 18
<p>Step-tracking experiment under dry wet and slippery conditions: (<b>a</b>) lateral error; (<b>b</b>) steady-state error distribution.</p>
Full article ">Figure 19
<p>Straight line experiment under wet and slippery conditions: (<b>a</b>) lateral error; (<b>b</b>) lateral error distribution.</p>
Full article ">Figure 20
<p>Ground surface adhesion coefficient estimation based on ST-UKF.</p>
Full article ">Figure 21
<p>Comparison of lateral errors across controllers on roads with varying adhesion coefficients.</p>
Full article ">Figure 22
<p>Lateral error distribution of different controllers under varying adhesion coefficients.</p>
Full article ">
20 pages, 12122 KiB  
Article
Microstructural and Mechanical Characterization of Nb-Doped MoS2 Coatings Deposited on H13 Tool Steel Using Nb-Based Interlayers
by Miguel R. Danelon, Newton K. Fukumasu, Angelo A. Carvalho, Ronnie R. Rego, Izabel F. Machado, Roberto M. Souza and André P. Tschiptschin
Coatings 2025, 15(1), 57; https://doi.org/10.3390/coatings15010057 - 6 Jan 2025
Viewed by 681
Abstract
Molybdenum disulfide is a 2D material with excellent lubricant properties, resulting from weak van der Waals forces between lattice layers and shear-induced crystal orientation. The low forces needed to shear the MoS2 crystal layers grant the tribological system low coefficients of friction [...] Read more.
Molybdenum disulfide is a 2D material with excellent lubricant properties, resulting from weak van der Waals forces between lattice layers and shear-induced crystal orientation. The low forces needed to shear the MoS2 crystal layers grant the tribological system low coefficients of friction (COF). However, film oxidation harms its efficacy in humid atmospheres, leading to an increased COF and poor surface adhesion, making its use preferable in dry or vacuum conditions. To overcome these challenges, doping MoS2 with elements such as Nb, Ti, C, and N emerges as a promising solution. Nevertheless, the adhesion of these coatings to a steel substrate presents challenges and strategies involving the reduction in residual stresses and increased chemical affinity to the substrate by using niobium-based materials as interlayers. In this study, Nb-doped MoS2 films were deposited on H13 steel and silicon wafers using the pulsed direct current balanced magnetron sputtering technique. Different niobium-based interlayers (pure Nb and NbN) were deposited to evaluate the adhesion properties of Nb-doped MoS2 coatings. Unlubricated scratch tests, conducted at room temperature and relative humidity under a progressive load, were performed to analyze the COF and adhesion of the coating. Instrumented indentation tests were conducted to assess the hardness and elastic modulus of the coatings. The microstructure of the coatings was obtained by Scanning Electron Microscopy (SEM), Scanning Transmission Electron Microscopy (STEM), and Transmission Electron Microscopy (TEM), with Energy-Dispersive X-Ray Spectroscopy (EDS). Results indicated that niobium doping on MoS2 coatings changes the structure from crystalline to amorphous. Additionally, the Nb concentration of the Nb:MoS2 coating changed the mechanical properties, leading to different cohesive failures by different loads during the scratch tests. Results have also indicated that an NbN interlayer optimally promoted the adhesion of the film. This result is justified by the increase in hardness led by higher Nb concentrations, enhancing the load-bearing capacity of the coating. It is concluded that niobium-based materials can be used to enhance the adhesion properties of Nb-doped MoS2 films and improve their tribological performance. Full article
(This article belongs to the Special Issue Friction, Wear, Lubrication and Mechanics of Surfaces and Interfaces)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Scheme of deposition and (<b>B</b>) architecture of Nb:MoS<sub>2</sub> coatings.</p>
Full article ">Figure 2
<p>Microstructures of Nb:MoS<sub>2</sub> coatings deposited under different powers applied to the Nb target: (<b>A</b>) 25 W, (<b>B</b>) 30 W, (<b>C</b>) 35 W, and (<b>D</b>) 40 W.</p>
Full article ">Figure 3
<p>Microstructure of Nb25NbN300 sample at SEM.</p>
Full article ">Figure 4
<p>FEG Nb30NbN300 film microstructure. The Pt light gray layer was deposited on the surface to protect the film from being damaged by the focused ion beam.</p>
Full article ">Figure 5
<p>EDS analysis of the cross-section of different layers.</p>
Full article ">Figure 6
<p>TEM image of the transition between Pt, MoS<sub>2</sub>, and Nb:MoS<sub>2</sub> layers.</p>
Full article ">Figure 7
<p>TEM images of the (<b>A</b>) pure MoS<sub>2</sub> layer, (<b>B</b>) Nb:MoS<sub>2</sub> film, and (<b>C</b>) scheme of the crystal structure of a MoS<sub>2</sub> monolayer showing a layer of molybdenum (blue) sandwiched between two layers of sulfur (yellow).</p>
Full article ">Figure 8
<p>TEM images of the (<b>A</b>) MoS<sub>2</sub> and (<b>C</b>) Nb:MoS<sub>2</sub> layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>B</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 8 Cont.
<p>TEM images of the (<b>A</b>) MoS<sub>2</sub> and (<b>C</b>) Nb:MoS<sub>2</sub> layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>B</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 9
<p>The transition between the NbN and Nb layers shows partial epitaxial growth. The blue line was included as an indication of the interface between both layers.</p>
Full article ">Figure 10
<p>TEM images of the (<b>A</b>) NbN and (<b>B</b>) Nb layers. Red circles indicate regions for SAED analysis, with corresponding patterns in (<b>C</b>) and (<b>D</b>), respectively.</p>
Full article ">Figure 11
<p>Raman spectroscopy of niobium nitride from NbN300 sample. The insert was taken from [<a href="#B58-coatings-15-00057" class="html-bibr">58</a>].</p>
Full article ">Figure 12
<p>Scratch images of samples (<b>A</b>) Nb25, (<b>B</b>) Nb30, (<b>C</b>) Nb35, (<b>D</b>) Nb40, and (<b>E</b>) Nb25NbN300.</p>
Full article ">Figure 13
<p>Graphs of COF, tangential force, and normal force as a function of the distance of the scratch for the samples (<b>A</b>) Nb25, (<b>B</b>) Nb30, (<b>C</b>) Nb35, (<b>D</b>) Nb40, and (<b>E</b>) Nb25NbN300.</p>
Full article ">Figure 14
<p>Characterization of cohesive failure of coatings by (<b>A</b>) optical microscope and (<b>B</b>) SEM.</p>
Full article ">Figure 15
<p>Characterization of adhesive failure by (<b>A</b>) optical microscope, (<b>B</b>) SEM and EDS, and (<b>C</b>) combined analysis of COF curve and scratch image.</p>
Full article ">Figure 15 Cont.
<p>Characterization of adhesive failure by (<b>A</b>) optical microscope, (<b>B</b>) SEM and EDS, and (<b>C</b>) combined analysis of COF curve and scratch image.</p>
Full article ">
23 pages, 12546 KiB  
Article
Effects of Beech Wood Surface Treatment with Polyethylenimine Solution Prior to Finishing with Water-Based Coating
by Tanja Palija, Milica Rančić, Daniela Djikanović, Ksenija Radotić, Marko Petrič, Matjaž Pavlič and Milan Jaić
Polymers 2025, 17(1), 77; https://doi.org/10.3390/polym17010077 - 30 Dec 2024
Viewed by 653
Abstract
The surfaces of beech wood samples were treated with polyethylenimine (PEI) solutions at three different concentrations—0.5%, 1% and 2%—and two molecular weights—low molecular weight (LMW) and high molecular weight (HMW). The effects of PEI surface treatment of wood were characterized by FT-IR spectroscopy, [...] Read more.
The surfaces of beech wood samples were treated with polyethylenimine (PEI) solutions at three different concentrations—0.5%, 1% and 2%—and two molecular weights—low molecular weight (LMW) and high molecular weight (HMW). The effects of PEI surface treatment of wood were characterized by FT-IR spectroscopy, the penetration depth of PEI (EPI fluorescence spectroscopy), the bonding position of PEI (by SEM), the wetting and surface energy, and the water uptake. After PEI treatment, the samples were coated with a water-based transparent acrylic coating (WTAC). The dry film thickness, the penetration depth of the coating, the adhesion strength and the surface roughness of the coated wood surface were evaluated. EPI fluorescence and SEM micrographs showed that PEI HMW chains were deposited on the surface, in contrast to PEI LMW, which penetrates deeper into layers of the wood cells. Treatment with a 1% PEI HMW solution resulted in a 72% reduction in water uptake of the wood (compared to untreated samples after 5 min of applying water droplets to the surface) and a 23.2% reduction in surface energy (compared to untreated samples) while maintaining the adhesion strength of the applied WTAC. The lower water uptake of the treated wood samples reduced the roughness of the coated surface, which is particularly important when the wood surface is finished with water-based coatings. Full article
(This article belongs to the Special Issue Advances in Polyelectrolytes and Polyelectrolyte Complexes)
Show Figures

Figure 1

Figure 1
<p>Dialysis of PEI LMW in deionized water: (<b>a</b>) at the beginning of dialysis; (<b>b</b>) 24 h after the start of dialysis.</p>
Full article ">Figure 2
<p>Dry film thickness measurement of water-based transparent acrylic coating.</p>
Full article ">Figure 3
<p>Calculation of coating penetration parameters: (<b>a</b>) snapshot of microtome sample; (<b>b</b>) area of available cell lumens (<span class="html-italic">A<sub>LA</sub></span>) in the penetration surface (<span class="html-italic">PS</span>); (<b>c</b>) area of filled cell lumens (<span class="html-italic">A<sub>LF</sub></span>) in the penetration surface (<span class="html-italic">PS</span>); (<b>d</b>) coating penetration depth at 30 positions and maximum coating penetration depth (<span class="html-italic">D<sub>max</sub></span>) in the penetration surface (<span class="html-italic">PS</span>).</p>
Full article ">Figure 4
<p>Determination of WTAC adhesion strength by the pull-off test.</p>
Full article ">Figure 5
<p>Measuring the roughness of the wood-coated surface.</p>
Full article ">Figure 6
<p>FT-IR spectra of untreated and samples treated with PEI solution in the range of 2000–4000 cm<sup>−1</sup> (<b>a</b>) and from 600 to 2000 cm<sup>−1</sup> (<b>b</b>).</p>
Full article ">Figure 6 Cont.
<p>FT-IR spectra of untreated and samples treated with PEI solution in the range of 2000–4000 cm<sup>−1</sup> (<b>a</b>) and from 600 to 2000 cm<sup>−1</sup> (<b>b</b>).</p>
Full article ">Figure 7
<p>Images of penetration of low- (LMW) and high-molecular-weight (HMW) PEI solutions of different concentrations, stained with rhodamine B, into the surface layer of wood, obtained by overlaying images with DAPI, FAM, and DsRED filters from an epi-fluorescence microscope: (<b>a</b>) 0.5% PEI LMW; (<b>b</b>) 1% PEI LMW; (<b>c</b>) 2% PEI LMW; (<b>d</b>) 0.5% PEI HMW; (<b>e</b>) 1% PEI HMW; (<b>f</b>) 2% PEI HMW.</p>
Full article ">Figure 7 Cont.
<p>Images of penetration of low- (LMW) and high-molecular-weight (HMW) PEI solutions of different concentrations, stained with rhodamine B, into the surface layer of wood, obtained by overlaying images with DAPI, FAM, and DsRED filters from an epi-fluorescence microscope: (<b>a</b>) 0.5% PEI LMW; (<b>b</b>) 1% PEI LMW; (<b>c</b>) 2% PEI LMW; (<b>d</b>) 0.5% PEI HMW; (<b>e</b>) 1% PEI HMW; (<b>f</b>) 2% PEI HMW.</p>
Full article ">Figure 8
<p>SEM images of wood samples protected by a coating layer (900× magnification): (<b>a</b>) untreated wood; (<b>b</b>) samples treated with a solution of 1% PEI LMW; (<b>c</b>) samples treated with a solution of 1% PEI HMW.</p>
Full article ">Figure 9
<p>The contact angle of a water droplet on the surface of untreated samples and samples treated with PEI solution of low (LMW) and high (HMW) molecular weight of different concentrations, during the observation period from 1 to 25 s.</p>
Full article ">Figure 10
<p>The contact angle of the WTAC droplet on the surface of untreated samples and samples treated with PEI solution of low (LMW) and high (HMW) molecular weight of different concentrations, during the observation period from 1 to 25 s.</p>
Full article ">Figure 11
<p>Water uptake of untreated samples and surface treated samples with PEI LMW and PEI HMW solutions of different concentrations (0.5%, 1%, and 2%).</p>
Full article ">
13 pages, 6209 KiB  
Article
Hollow Salt Prepared Through Spray Drying with Alginate Enhances Salinity Perception to Reduce Sodium Intake
by Qian Jiang, Jiayi Yan, Chen Song, Yunning Yang, Guangyuan Chen, Fanhua Kong, Jingfeng Yang and Shuang Song
Foods 2025, 14(1), 19; https://doi.org/10.3390/foods14010019 - 25 Dec 2024
Viewed by 488
Abstract
Currently, high-salt diets have become one of the world’s biggest dietary crisis and long-term high-salt diets are seriously detrimental to human health. In response to this situation, the present study proposed a saltiness enhancement strategy using alginate, which is a dietary fibre from [...] Read more.
Currently, high-salt diets have become one of the world’s biggest dietary crisis and long-term high-salt diets are seriously detrimental to human health. In response to this situation, the present study proposed a saltiness enhancement strategy using alginate, which is a dietary fibre from brown algae and has many health benefits, such as regulating intestinal microbiota, anti-hypertension and anti-obesity. The comparison of alginates with different viscosities showed that alginate of 1000–1500 cps at a concentration of 1.25 g/L could enhance the saltiness of NaCl solution by 11.5%. Then, a solid salt was prepared through spray drying with 4.83% of this alginate, and its structure was characterised by X-Ray diffraction, Fourier transform infrared spectroscopy, scanning electron microscopy and energy dispersive spectroscopy to confirm its hollow structure with a particle size of 6.25 ± 2.26 μm as well as its crystal structure similar to original NaCl. Moreover, the conductivity monition revealed that the hollow salt exhibited a more rapid dissolution in water and its alginate component increased the adhesive retention of sodium ions on the tongue surface, which both effectively enhanced the sensory perception. Finally, as revealed by the sensory evaluation, the prepared hollow salt showed higher saltiness than that of original table salt and it could reduce sodium intake by 29%. Thus, the hollow salt prepared with alginate in the present study has potential for salt reduction. Full article
Show Figures

Figure 1

Figure 1
<p>Saltiness values of NaCl solutions containing alginates of different viscosities obtained by electronic tongue measurement, where different letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05 (<b>A</b>). Saltiness values of solutions containing HV-alginate at different concentrations, and different letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05 (<b>B</b>). Saltiness scores of mixed solutions containing alginates of different viscosities at different concentrations obtained by sensory evaluation, where # indicates the significant difference between each mixed solution and the control (0 g/L alginate), and * indicates the significant difference between the mixed solutions containing alginate with different viscosities at the same concentration. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (<b>C</b>).</p>
Full article ">Figure 2
<p>X-Ray diffraction patterns (<b>A</b>) and FTIR spectra (<b>B</b>) for structural analysis of solid salts.</p>
Full article ">Figure 3
<p>Scanning electron micrographs of different salt particles (<b>A</b>) and particle size distribution of alginate–NaCl (<b>B</b>).</p>
Full article ">Figure 4
<p>Distribution of elements on the surface of alginate–NaCl (<b>A</b>) and elemental signals on the surfaces of alginate–NaCl and NaCl (<b>B</b>).</p>
Full article ">Figure 5
<p>Fluorescence pictures of porcine tongues in vitro (<b>A</b>). Fluorescence intensity obtained using Image J, where * indicates significant differences in fluorescence intensity at the same volume of artificial saliva, ** <span class="html-italic">p</span> &lt; 0.01 (<b>B</b>). The sodium content of the rinsed artificial saliva, where * indicates significant differences, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (<b>C</b>).</p>
Full article ">Figure 6
<p>Dissolution curves of alginate–NaCl and NaCl in water over time, where * indicates a significant difference in conductivity between alginate–NaCl and NaCl at the same time, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (<b>A</b>). Saltiness scores for different solid salts, where different letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05 (<b>B</b>).</p>
Full article ">
16 pages, 9183 KiB  
Article
Effects of Polyol Types on Underwater Curing Properties of Polyurethane
by Cheng Zhang, Yixuan Zhang, Yao Liu, Yiming Cui, Ming Zhao, Shuai Peng, Hecong Wang, Zuobao Song, Qunchao Zhang, Dean Shi and Yuxue Zhu
Polymers 2025, 17(1), 5; https://doi.org/10.3390/polym17010005 - 24 Dec 2024
Cited by 1 | Viewed by 524
Abstract
This study aims to develop castable polyurethane suitable for applications on wet substrates or underwater construction. Polyurethanes were synthesized using various polyols with similar hydroxyl values, including poly(tetrahydrofuran) polyol, polyester polyol, castor oil-modified polyol, soybean oil-modified polyol, and cashew nut shell oil-modified polyol. [...] Read more.
This study aims to develop castable polyurethane suitable for applications on wet substrates or underwater construction. Polyurethanes were synthesized using various polyols with similar hydroxyl values, including poly(tetrahydrofuran) polyol, polyester polyol, castor oil-modified polyol, soybean oil-modified polyol, and cashew nut shell oil-modified polyol. The corresponding polyurethane curing products were evaluated for their underwater curing characteristics by volume expansion ratios and adhesion strength on dry and wet substrates, combined with analyses of reaction exothermic behavior, wetting properties on dry and wet substrates, interfacial tension, and microstructure characterization from the perspectives of reaction activity and water solubility. The results indicate that polyols with higher hydrophobicity and reactivity to isocyanates lead to reduced side reactions during underwater curing, making them more suitable for underwater applications. Soybean oil-based and cashew nut shell oil-based polyurethanes exhibited fast curing (gel times of 1.15 and 1.35 min, respectively), minimal volume change (within 2.5% after 7 days underwater), and strong wet adhesion (1.95 MPa and 2.38 MPa with minimal loss, respectively). The two polyols showed different mechanical properties, providing tailored options for specific underwater engineering applications. Full article
Show Figures

Figure 1

Figure 1
<p>Sample preparation process.</p>
Full article ">Figure 2
<p>Specimens for mechanical tests. (<b>a</b>) Specimens for tensile test; (<b>b</b>) specimens for compression test.</p>
Full article ">Figure 3
<p>FTIR spectra of polyols before and after the reaction. (<b>a</b>) FTIR spectra of polyols; (<b>b</b>) FTIR spectra of the corresponding polyurethanes derived from the polyols.</p>
Full article ">Figure 4
<p>Temperature–time curve of five reaction systems.</p>
Full article ">Figure 5
<p>Gel time of the reaction systems.</p>
Full article ">Figure 6
<p>Rate of temperature change over time for different samples.</p>
Full article ">Figure 7
<p>Reaction formula of isocyanates with water [<a href="#B1-polymers-17-00005" class="html-bibr">1</a>].</p>
Full article ">Figure 8
<p>Volume expansion ratio of samples cured in air and water.</p>
Full article ">Figure 9
<p>Photographic comparison of samples cured in air and water.</p>
Full article ">Figure 10
<p>Microscopic morphology comparison of samples cured in air and water. (<b>a</b>) curing in air; (<b>b</b>) curing in water.</p>
Full article ">Figure 11
<p>Adhesion strengths of samples on wet and dry mortar surfaces.</p>
Full article ">Figure 12
<p>Interfacial tension between polyols and water.</p>
Full article ">Figure 13
<p>Droplet base diameter on dry and wet mortar surfaces over time.</p>
Full article ">Figure 14
<p>Mechanical properties and stress–strain curves of the samples. (<b>a</b>) Tensile properties of the samples; (<b>b</b>) tensile stress–strain curves; (<b>c</b>) compression properties of the samples; (<b>d</b>) compression stress–strain curves.</p>
Full article ">
17 pages, 4437 KiB  
Article
Fabrication of Polytetrafluoroethylene-Reinforced Fluorocarbon Composite Coatings and Tribological Properties Under Multi-Environment Working Conditions
by Changqing Xi, Bochao Zhang, Xiangdong Ye and Honghua Yan
Polymers 2024, 16(24), 3595; https://doi.org/10.3390/polym16243595 - 22 Dec 2024
Viewed by 699
Abstract
Currently, few studies have been conducted on the use of fluorocarbon resin (FEVE) and polytetrafluoroethylene (PTFE) as adhesive substrates and lubricating and anti-corrosion fillers, respectively, for the fabrication of PTFE-reinforced fluorocarbon composite coatings. In this paper, the tribological properties of polytetrafluoroethylene-reinforced fluorocarbon composite [...] Read more.
Currently, few studies have been conducted on the use of fluorocarbon resin (FEVE) and polytetrafluoroethylene (PTFE) as adhesive substrates and lubricating and anti-corrosion fillers, respectively, for the fabrication of PTFE-reinforced fluorocarbon composite coatings. In this paper, the tribological properties of polytetrafluoroethylene-reinforced fluorocarbon composite coatings were investigated through orthogonal tests under various operating conditions. The optimal configuration for coating preparation under dry friction and aqueous lubrication was thus obtained: the optimal filler particle size, mass ratio of FEVE to PTFE, spraying pressure, and curing agent content were 50 μm, 3:4.5, 0.3 MPa, and 0.3, respectively. Under oil lubrication, the corresponding optimal values were 5 μm, 3:4.5, 0.3 MPa, and 0.3, respectively. Tribological tests revealed that the best overall performance of the FEVE/PTFE coating was obtained when the mass ratio of FEVE to PTFE was 3:4.5, and the filler particle size also significantly affected the tribological properties under different environments, including the friction coefficients of the FEVE/50 μm-PTFE coating under both dry friction and aqueous lubrication, as well as the friction coefficient of the FEVE/5 μm-PTFE coating under oil lubrication. These coefficients were 0.067, 0.062, and 0.055, representing decreases of 86%, 92%, and 56%, respectively, compared to those of the pure FEVE coating under the same working conditions. This research was conducted with the goal of expanding the application of fluorocarbon coatings in the field of tribology. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular structure of FEVE.</p>
Full article ">Figure 2
<p>PTFE molecular structure.</p>
Full article ">Figure 3
<p>Multifunctional friction and wear tester.</p>
Full article ">Figure 4
<p>Schematic representation of the friction test principle.</p>
Full article ">Figure 5
<p>Adhesion topographies of different orthogonal test specimens. (<b>a</b>) Specimen no. 1. (<b>b</b>) Specimen no. 2. (<b>c</b>) Specimen no. 3. (<b>d</b>) Specimen no. 4. (<b>e</b>) Specimen no. 5. (<b>f</b>) Specimen no. 6. (<b>g</b>) Specimen no. 7. (<b>h</b>) Specimen no. 8. (<b>i</b>) Specimen no. 9.</p>
Full article ">Figure 6
<p>Variations in the coefficient of friction curves of different orthogonal test specimens. (<b>a</b>) Specimen no. 1. (<b>b</b>) Specimen no. 2. (<b>c</b>) Specimen no. 3. (<b>d</b>) Specimen no. 4. (<b>e</b>) Specimen no. 5. (<b>f</b>) Specimen no. 6. (<b>g</b>) Specimen no. 7. (<b>h</b>) Specimen no. 8. (<b>i</b>) Specimen no. 9.</p>
Full article ">Figure 6 Cont.
<p>Variations in the coefficient of friction curves of different orthogonal test specimens. (<b>a</b>) Specimen no. 1. (<b>b</b>) Specimen no. 2. (<b>c</b>) Specimen no. 3. (<b>d</b>) Specimen no. 4. (<b>e</b>) Specimen no. 5. (<b>f</b>) Specimen no. 6. (<b>g</b>) Specimen no. 7. (<b>h</b>) Specimen no. 8. (<b>i</b>) Specimen no. 9.</p>
Full article ">Figure 7
<p>Plot of the mean values of levels for adhesion factors.</p>
Full article ">Figure 8
<p>Plot of the mean values of levels for friction factors.</p>
Full article ">Figure 9
<p>Friction coefficient variation curves of the composite coatings with different FEVE/PTFE mass ratios under dry friction.</p>
Full article ">Figure 10
<p>Friction coefficient variation curves of the composite coatings with different FEVE/PTFE mass ratios under water lubrication.</p>
Full article ">Figure 11
<p>Friction coefficient variation curves of the composite coatings with different FEVE/PTFE mass ratios under oil lubrication.</p>
Full article ">Figure 12
<p>Friction coefficient variation curves of the FEVE/5 μm-PTFE composite coating under different working conditions.</p>
Full article ">
17 pages, 11720 KiB  
Article
A Worm-like Soft Robot Based on Adhesion-Controlled Electrohydraulic Actuators
by Yangzhuo Wu, Zhe Sun, Yu Xiang and Jieliang Zhao
Biomimetics 2024, 9(12), 776; https://doi.org/10.3390/biomimetics9120776 - 20 Dec 2024
Viewed by 858
Abstract
Worms are organisms characterized by simple structures, low energy consumption, and stable movement. Inspired by these characteristics, worm-like soft robots demonstrate exceptional adaptability to unstructured environments, attracting considerable interest in the field of biomimetic engineering. The primary challenge currently involves improving the motion [...] Read more.
Worms are organisms characterized by simple structures, low energy consumption, and stable movement. Inspired by these characteristics, worm-like soft robots demonstrate exceptional adaptability to unstructured environments, attracting considerable interest in the field of biomimetic engineering. The primary challenge currently involves improving the motion performance of worm-like robots from the perspectives of actuation and anchoring. In this study, a single segment worm-like soft robot driven by electrohydraulic actuators is proposed. The robot consists of a soft actuation module and two symmetrical anchoring modules. The actuation modules enable multi-degree-of-freedom motion of the robot using symmetric dual-electrode electrohydraulic actuators, while the anchoring modules provide active friction control through bistable electrohydraulic actuators. A hierarchical microstructure design is used for the biomimetic adhesive surface, enabling rapid, reversible, and stable attachment to and detachment from different surfaces, thereby improving the robot’s surface anchoring performance. Experimental results show that the designed robot can perform peristaltic and bending motions similar to a worm. It achieves rapid bidirectional propulsion on both dry and wet surfaces, with a maximum speed of 10.36 mm/s (over 6 velocity/length ratio (min−1)). Full article
Show Figures

Figure 1

Figure 1
<p>Physiological structure and locomotion mechanism of typical worms. (<b>a</b>) Earthworms are a typical worm organism. (<b>b</b>) A complete stride period of the peristaltic crawling. (<b>c</b>) Stride period of single-segment robot. (<b>d</b>) Concept of the designed worm-like soft robot.</p>
Full article ">Figure 2
<p>Design and structure of worm-like soft robot. (<b>a</b>) The structure of the robot consists of a central actuation module and two anchoring modules (tail/head module). (<b>b</b>) Hierarchical micro-nano adhesive surface inspired by honeybees and tree frogs.</p>
Full article ">Figure 3
<p>Analytical model of the actuation actuator unit. (<b>a</b>) Activation of electrodes on both sides resulting in linear motion. (<b>b</b>) Activation of electrodes on one side resulting in bending motion. (<b>c</b>) Free energy model of the actuator. (<b>d</b>) Initial-state angle and liquid fill amount of the actuation actuator for different electrode coverage. (<b>e</b>) Strain and bending angle of the actuation actuator unit for different electrode coverage. (<b>f</b>) Force-strain behavior of the actuation actuator unit for different electrode coverage.</p>
Full article ">Figure 4
<p>Fabrication of actuators and adhesive surface. (<b>a</b>) Fabrication process of the actuators. (<b>b</b>) Fabrication process of the biomimetic adhesive surface. (<b>c</b>) Microstructure of the biomimetic adhesive surface.</p>
Full article ">Figure 5
<p>Response performance testing of the actuators. (<b>a</b>) Experimental setup for strain measurement. (<b>b</b>) Voltage–displacement curves. (<b>c</b>) Calculated voltage–strain curves. (<b>d</b>) Displacement and velocity of the Actuators. (<b>e</b>) Voltage–frequency curves. (<b>f</b>) Displacement response. (<b>g</b>) Experimental setup for force measurement. (<b>h</b>) Voltage–strain curves. (<b>i</b>) Frequency–force curves.</p>
Full article ">Figure 6
<p>Bioinspired adhesive surface experiments. (<b>a</b>) Schematic of the configuration used to measure adhesion. (<b>b</b>) Adhesion measurement system. (<b>c</b>) Adhesion forces of FP, TP, and HTP in various surface conditions. (<b>d</b>,<b>e</b>) Preload–adhesion curves in various surface conditions.</p>
Full article ">Figure 7
<p>Design of the robotic actuation system. (<b>a</b>) Periodic control signals for the robot. (<b>b</b>) Locomotion principle of the worm-like robots. (<b>c</b>) The configuration of the control system.</p>
Full article ">Figure 8
<p>Basic motion of the worm-like robot. (<b>a</b>) Rectilinear locomotion over one stride period. (<b>b</b>) Bending locomotion of the robot.</p>
Full article ">Figure 9
<p>Continuous Motion Experiment of the worm-like robot. (<b>a</b>) Crawling locomotion on a PVC flat surface. (<b>b</b>) Crawling locomotion on a PVC flat surface. (<b>c</b>) Bidirectional locomotion on a moist PVC flat surface with a carried load of 100 g.</p>
Full article ">
Back to TopTop