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
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
remove_circle_outline

Search Results (3,917)

Search Parameters:
Keywords = friction coefficient

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 6128 KiB  
Article
Wear Resistance Design of Laser Cladding Ni-Based Self-Fluxing Alloy Coating Using Machine Learning
by Jiabo Fu, Quanling Yang, Oleg Devojno, Marharyta Kardapolava, Iryna Kasiakova and Chenchong Wang
Materials 2024, 17(22), 5651; https://doi.org/10.3390/ma17225651 - 19 Nov 2024
Abstract
To improve the collaborative design of laser cladding Ni-based self-fluxing alloy (SFA) wear-resistant coatings, machine learning methods were applied. A comprehensive database was constructed from the literature, linking alloy composition, processing parameters, testing conditions, and the wear properties of Ni-based SFA coatings. Feature [...] Read more.
To improve the collaborative design of laser cladding Ni-based self-fluxing alloy (SFA) wear-resistant coatings, machine learning methods were applied. A comprehensive database was constructed from the literature, linking alloy composition, processing parameters, testing conditions, and the wear properties of Ni-based SFA coatings. Feature correlation analysis using Pearson’s correlation coefficient and feature importance assessment via the random forest (RF) model highlighted the significant impact of C and B elements. The predictive performance of five classical machine learning algorithms was evaluated using metrics such as the squared correlation coefficient () and mean absolute error (MAE). The RF model, which exhibited the best overall performance, was further combined with a genetic algorithm (GA) to optimize both composition and processing parameters collaboratively. This integrated RF-GA optimization system significantly enhanced efficiency and successfully designed multiple composition and process plans. The optimized alloy demonstrated superior wear resistance with an average friction coefficient of only 0.34, attributed to an enhanced solid solution strengthening effect (110 MPa) and increased hard phase content (52%), such as Ni₃Si, CrB, and NbC. These results provide valuable methodological insights and theoretical support for the preparation of laser cladding coatings and enable efficient process optimization for other laser processing applications. Full article
Show Figures

Figure 1

Figure 1
<p>Framework for the design of laser cladding Ni-based self-fluxing alloys utilizing machine learning methods.</p>
Full article ">Figure 2
<p>Results of feature analysis. (<b>a</b>) PCC between all features; (<b>b</b>) feature importance of input features.</p>
Full article ">Figure 3
<p>Performance of different ML models. (<b>a</b>) Mean <span class="html-italic">R<sup>2</sup></span> and (<b>b</b>) <span class="html-italic">MAE</span>.</p>
Full article ">Figure 4
<p>Experimental values vs. predicted values. Mean results for RF model in (<b>a</b>) training set and (<b>b</b>) testing set, and (<b>c</b>) optimal result for RF model; mean results for MLP model in (<b>d</b>) training set and (<b>e</b>) testing set, and (<b>f</b>) optimal result for MLP model.</p>
Full article ">Figure 5
<p>Optimization and design results using RF model and GA.</p>
Full article ">Figure 6
<p>Comparison of hard phases and undesirable phases at lower temperatures between designed alloys (as indicated by spheres with different colors) and optimal alloys in dataset (as indicated by blocks with different colors) using Thermo-Calc software (2023a) and TCNI12 database.</p>
Full article ">Figure 7
<p>Number of underfitted models under multiple divisions of different machine learning models.</p>
Full article ">Figure 8
<p>Number of overfitted models (<b>a</b>) and better models (<b>b</b>) under multiple divisions of different machine learning models.</p>
Full article ">Figure 9
<p>The change of solution in the calculation process of GA. (<b>a</b>) P<sub>c</sub> = 0.5; (<b>b</b>) P<sub>c</sub> = 0.6; (<b>c</b>) P<sub>c</sub> = 0.7; (<b>d</b>) P<sub>c</sub> = 0.8.</p>
Full article ">Figure 10
<p>Comparison of compositional correlation between designed Alloy D2 and Ni self-fluxing alloys used in dataset using distance function.</p>
Full article ">Figure 11
<p>(<b>a</b>) Elemental content changes in Alloy D2 compared to the Optimal Alloy 1 in the original dataset and (<b>b</b>) phase configurations of Alloy D2.</p>
Full article ">
16 pages, 3956 KiB  
Article
Ball-on-Disk Wear Maps for Bearing Steel–Hard Anodized EN AW-6082 Aluminum Alloy Tribocouple in Dry Sliding Conditions
by Enrico Baroni, Annalisa Fortini, Lorenzo Meo, Chiara Soffritti, Mattia Merlin and Gian Luca Garagnani
Coatings 2024, 14(11), 1469; https://doi.org/10.3390/coatings14111469 - 19 Nov 2024
Abstract
In recent years, Golden Hard Anodizing (G.H.A.®) has been developed as a variant of the traditional hard anodizing process with the addition of Ag+ ions in the nanoporous structure. The tribological properties of this innovative surface treatment are still not [...] Read more.
In recent years, Golden Hard Anodizing (G.H.A.®) has been developed as a variant of the traditional hard anodizing process with the addition of Ag+ ions in the nanoporous structure. The tribological properties of this innovative surface treatment are still not well understood. In this study, ball-on-disk tests were conducted in dry sliding conditions using 100Cr6 (AISI 52100) bearing steel balls as a counterbody and GHA®-anodized EN AW-6082 aluminum alloy disks. The novelty of this work lies in the mapping of the wear properties of the tribocouple under different test conditions for a better comparison of the results. Three different normal loads (equal to 5, 10, and 15 N) and three different reciprocating frequencies (equal to 2, 3, and 4 Hz) were selected to investigate a spectrum of operating conditions for polished and unpolished G.H.A.®-anodized EN AW-6082 aluminum alloy. Quantitative wear maps were built based on the resulting wear rate values to define the critical operating limits of the considered tribocouple. The results suggest that the coefficient of friction (COF) was independent of test conditions, while different wear maps were found for polished and non-polished surfaces. Polishing before anodizing permitted the acquisition of lower wear for the anodized disks and the steel balls. Full article
25 pages, 20805 KiB  
Article
Analysis of Influence of Coating Type on Friction Behaviour and Surface Topography of DC04/1.0338 Steel Sheet in Bending Under Tension Friction Test
by Tomasz Trzepieciński, Krzysztof Szwajka, Marek Szewczyk, Joanna Zielińska-Szwajka, Marek Barlak, Katarzyna Nowakowska-Langier and Sebastian Okrasa
Materials 2024, 17(22), 5650; https://doi.org/10.3390/ma17225650 - 19 Nov 2024
Abstract
The working conditions of tools during plastic working operations are determined by, among other things, temperature, loads, loading method, and processing speed. In sheet metal forming processes, additionally, lubricant and tool surface roughness play a key role in changing the surface topography of [...] Read more.
The working conditions of tools during plastic working operations are determined by, among other things, temperature, loads, loading method, and processing speed. In sheet metal forming processes, additionally, lubricant and tool surface roughness play a key role in changing the surface topography of the drawpieces. This article presents the results of friction analysis on the edge of the punch in a deep drawing process using the bending under tension test. A DC04 steel sheet was used as the test material. The influence of various types of titanium nitride and titanium coatings applied on the surface of countersamples made of 145Cr6 cold-work tool steel was tested by means of high-intensity plasma pulses, magnetron sputtering, and electron pulse irradiation. The influence of the type of tool coating on the evolution of the coefficient of friction, the change in the sheet surface topography, and the temperature in the contact zone is presented in this paper. An increase in the coefficient of friction with sample elongation was observed. Countersamples modified with protective coatings provided a more stable coefficient value during the entire friction test compared to dry friction conditions. The electron pulse irradiated countersample provided the highest stability of the coefficient of friction in the entire range of sample elongation until fracture. The skewness Ssk of the sheet metal tested against the coated countersamples was characterized by negative value, which indicates a plateau-like shape of their surface. The highest temperature in the contact zone during friction with all types of countersamples was observed for the uncoated countersample. Full article
(This article belongs to the Special Issue Advanced Materials and Technologies for Thermal Sprayed Coatings)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Topography and (<b>b</b>) bearing area curve of DC01 steel sheet surface.</p>
Full article ">Figure 2
<p>Test stand.</p>
Full article ">Figure 3
<p>View of samples before and after friction process.</p>
Full article ">Figure 4
<p>A schematic of the BUT test.</p>
Full article ">Figure 5
<p>Flow chart for experimental investigations.</p>
Full article ">Figure 6
<p>The general view of the modified countersamples: (<b>a</b>) Ti-HIPP, (<b>b</b>) TiN-MS, (<b>c</b>) Ti-MS+EPI, and (<b>d</b>) three consecutive (from left to right) orientations of countersample during modification using electron gun.</p>
Full article ">Figure 7
<p>RPI-IBIS device for high-energy plasma generation + view of coaxial rod plasma accelerator with titanium electrodes.</p>
Full article ">Figure 8
<p>The device for layer deposition by the pulsed magnetron sputtering method + view of the magnetron inside the chamber with a titanium target.</p>
Full article ">Figure 9
<p>Electron gun device + countersample inside vacuum chamber of electron gun.</p>
Full article ">Figure 10
<p>Surface topography of cylindrical countersamples: (<b>a</b>) as-received state, (<b>b</b>) Ti-HIPP, (<b>c</b>) TiN-MS, and (<b>d</b>) Ti-MS+EPI.</p>
Full article ">Figure 10 Cont.
<p>Surface topography of cylindrical countersamples: (<b>a</b>) as-received state, (<b>b</b>) Ti-HIPP, (<b>c</b>) TiN-MS, and (<b>d</b>) Ti-MS+EPI.</p>
Full article ">Figure 11
<p>EDS layered images of countersamples: (<b>a</b>) Ti-HIPP, (<b>b</b>) TiN-MS, and (<b>c</b>) Ti-MS+EPI.</p>
Full article ">Figure 12
<p>The EDS spectrum of the following coatings: (<b>a</b>) Ti-HIPP, (<b>b</b>) TiN-MS, and (<b>c</b>) Ti-MS+EPI.</p>
Full article ">Figure 12 Cont.
<p>The EDS spectrum of the following coatings: (<b>a</b>) Ti-HIPP, (<b>b</b>) TiN-MS, and (<b>c</b>) Ti-MS+EPI.</p>
Full article ">Figure 13
<p>EDS elemental mapping of Ti-HIPP coating.</p>
Full article ">Figure 14
<p>EDS elemental mapping of TiN-MS coating.</p>
Full article ">Figure 15
<p>EDS elemental mapping of Ti-MS+EPI coating.</p>
Full article ">Figure 16
<p>Effect of sample elongation on coefficient of friction measured under (a) dry friction and lubricated conditions using (b) S100+ oil and (c) S300 oil.</p>
Full article ">Figure 16 Cont.
<p>Effect of sample elongation on coefficient of friction measured under (a) dry friction and lubricated conditions using (b) S100+ oil and (c) S300 oil.</p>
Full article ">Figure 17
<p>Effect of friction conditions on selected surface roughness parameters: (a) Sa, (b) Sku, (c) Ssk, (d) Sp, (e) Sv, and (f) Sz.</p>
Full article ">Figure 18
<p>Interpretation of parameters (<b>a</b>) Sku and (<b>b</b>) Ssk.</p>
Full article ">Figure 19
<p>The SEM micrographs of the surface of (<b>a</b>) the DC01 sheet in the as-received state and after friction tests under the following conditions: (<b>b</b>) Ti-HIPP, dry friction; (<b>c</b>) Ti-HIPP, lubrication with S100+ oil; (<b>d</b>) TiN-MS, lubrication with S100+ oil; and Ti-MS+EPI, lubrication with S300 oil.</p>
Full article ">Figure 20
<p>View of working surfaces of (<b>a</b>) uncoated countersample and coated countersamples: (<b>b</b>) Ti-HIPP, (<b>c</b>) TiN-MS, and (<b>d</b>) Ti-MS+EPI.</p>
Full article ">Figure 21
<p>Effect of friction conditions on temperature in contact zone.</p>
Full article ">
20 pages, 3272 KiB  
Article
Crosslinking by Click Chemistry of Hyaluronan Graft Copolymers Involving Resorcinol-Based Cinnamate Derivatives Leading to Gel-like Materials
by Mario Saletti, Simone Pepi, Marco Paolino, Jacopo Venditti, Germano Giuliani, Claudia Bonechi, Gemma Leone, Agnese Magnani, Claudio Rossi and Andrea Cappelli
Gels 2024, 10(11), 751; https://doi.org/10.3390/gels10110751 - 19 Nov 2024
Viewed by 70
Abstract
The well-known “click chemistry” reaction copper(I)-catalyzed azide-alkyne 1,3-dipolar cycloaddition (CuAAC) was used to transform under very mild conditions hyaluronan-based graft copolymers HA(270)-FA-Pg into the crosslinked derivatives HA(270)-FA-TEGERA-CL and HA(270)-FA-HEGERA-CL. In particular, medium molecular weight (i.e., 270 kDa) hyaluronic acid (HA) grafted at [...] Read more.
The well-known “click chemistry” reaction copper(I)-catalyzed azide-alkyne 1,3-dipolar cycloaddition (CuAAC) was used to transform under very mild conditions hyaluronan-based graft copolymers HA(270)-FA-Pg into the crosslinked derivatives HA(270)-FA-TEGERA-CL and HA(270)-FA-HEGERA-CL. In particular, medium molecular weight (i.e., 270 kDa) hyaluronic acid (HA) grafted at various extents (i.e., 10, 20, and 40%) with fluorogenic ferulic acid (FA) residue bonding propargyl groups were used in the CuAAC reaction with novel azido-terminated crosslinking agents Tri(Ethylene Glycol) Ethyl Resorcinol Acrylate (TEGERA) and Hexa(Ethylene Glycol) Ethyl Resorcinol Acrylate (HEGERA). The resulting HA(270)-FA-TEGERA-CL and HA(270)-FA-HEGERA-CL materials were characterized from the point of view of their structure by performing NMR studies. Moreover, the swelling behavior and rheological features were assessed employing TGA and DSC analysis to evaluate the potential gel-like properties of the resulting crosslinked materials. Despite the 3D crosslinked structure, HA(270)-FA-TEGERA-CL and HA(270)-FA-HEGERA-CL frameworks showed adequate swelling performance, the required shear thinning behavior, and coefficient of friction values close to those of the main commercial HA solutions used as viscosupplements (i.e., 0.20 at 10 mm/s). Furthermore, the presence of a crosslinked structure guaranteed a longer residence time. Indeed, HA(270)-FA-TEGERA-CL-40 and HA(270)-FA-HEGERA-CL-40 after 48 h showed a four times greater enzymatic resistance than the commercial viscosupplements. Based on the promising obtained results, the crosslinked materials are proposed for their potential applicability as novel viscosupplements. Full article
(This article belongs to the Special Issue Polymeric Hydrogels for Biomedical Application)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic depiction of novel crosslinked gel-like materials <b>HA(270)-FA-TEGERA-CL</b> and <b>HA(270)-FA-HEGERA-CL</b> obtained by CuAAC reaction of <b>HA(270)-FA-Pg</b> graft copolymers, showing different grafting degree values, with clickable agents <b>1a,b</b>, respectively.</p>
Full article ">Figure 2
<p>Comparison of <sup>1</sup>H NMR spectrum recorded with crosslinked material <b>HA(270)-FA-TEGERA-CL-20</b> (D<sub>2</sub>O, 600 MHz) with the one acquired with its corresponding synthetic precursor, the graft copolymer <b>HA(270)-FA-Pg-20</b> (D<sub>2</sub>O, 600 MHz, the peak at 2.98 ppm attributed to alkyne proton showed a reduced intensity because of the exchange with deuterium).</p>
Full article ">Figure 3
<p>Comparison of <sup>1</sup>H NMR spectrum recorded with the crosslinked material <b>HA(270)-FA-HEGERA-CL-20</b> (D<sub>2</sub>O, 600 MHz) with the one acquired with its corresponding synthetic precursor, the graft copolymer <b>HA(270)-FA-Pg-20</b> (D<sub>2</sub>O, 600 MHz, the peak at 2.98 ppm attributed to alkyne proton showed a reduced intensity because of the exchange with deuterium).</p>
Full article ">Figure 4
<p>Elastic (G′) and loss moduli (G″) trend in <b>HA(270)-FA-TEGERA-CL</b> and <b>HA(270)-FA-HEGERA-CL</b> series recorded in shear mode at 37 °C in frequency range 0.01–15 Hz (n = 3).</p>
Full article ">Figure 5
<p>Complex modulus of <b>HA(270)-FA-TEGERA-CL</b>, <b>HA(270)-FA-HEGERA-CL</b>, and HA(270)-FA-HEG-CL series recorded at 37 °C in frequency range 0.01–15 Hz.</p>
Full article ">Figure 6
<p>Elastic (E′) and loss moduli (E″) trend of <b>HA(270)-FA-TEGERA-CL</b> and <b>HA(270)-FA-HEGERA-CL</b> series recorded in compression mode at 37 °C in frequency range 0.01–15 Hz.</p>
Full article ">Figure 7
<p>COF trend of <b>HA(270)-FA-TEGERA-CL</b> and <b>HA(270)-FA-HEGERA-CL</b> series as a function of sliding speed (mm/s).</p>
Full article ">Figure 8
<p>Enzymatic degradation (%) of <b>HA(270)-FA-TEGERA-CL</b> series and <b>HA(270)-FA-HEGERA-CL</b> series in comparison with two commercial viscosupplements.</p>
Full article ">Scheme 1
<p>Preparation of suitable clickable crosslinking agents <b>1a</b> (<b>TEGERA</b>) and <b>1b</b> (<b>HEGERA</b>)<b>. Reagents</b>: (i) CH<sub>3</sub>SO<sub>2</sub>Cl, TEA, CH<sub>2</sub>Cl<sub>2</sub>; (ii) NaN<sub>3</sub>, DMF, CH<sub>3</sub>CN; (iii) CH<sub>3</sub>SO<sub>2</sub>Cl, TEA, CH<sub>2</sub>Cl<sub>2</sub>; (iv) POCl<sub>3</sub>, EtOH; (v) BBr<sub>3</sub>, CH<sub>2</sub>Cl<sub>2</sub>; (vi) Cs<sub>2</sub>CO<sub>3</sub>, NaI, CH<sub>3</sub>CN. <b>a</b>: n = 1; <b>b</b>: n = 4.</p>
Full article ">Scheme 2
<p>Click-crosslinking reaction of <b>HA(270)-FA-Pg</b> graft copolymers exhibiting diverse grafting degree values with resorcinol-based ethyl cinnamate bearing azido-terminated oligo(ethylene glycol) side chains <b>1a (TEGERA)</b> and <b>1b (HEGERA)</b> producing <b>HA(270)-FA-TEGERA-CL</b> and <b>HA(270)-FA-HEGERA-CL</b> gel-like materials, respectively. <b>Reagents</b>: (i) CuSO<sub>4</sub>, sodium ascorbate, <span class="html-italic">tert</span>-butanol, H<sub>2</sub>O.</p>
Full article ">
18 pages, 2514 KiB  
Article
Aloe Vera as a Printed Coating to Mitigate the Wear of Textiles
by Michail Karypidis, Amalia Stalika, Maria Zarkogianni, Apostolos Korlos and Eleftherios G. Andriotis
Coatings 2024, 14(11), 1467; https://doi.org/10.3390/coatings14111467 - 18 Nov 2024
Viewed by 337
Abstract
Aloe vera is well known for its biological properties as a bioflavonoid anti-inflammatory and antibacterial agent. It has been used frequently in the food sector as a food coating due to its hygroscopic properties and as an ingredient in the lucrative cosmetic industry. [...] Read more.
Aloe vera is well known for its biological properties as a bioflavonoid anti-inflammatory and antibacterial agent. It has been used frequently in the food sector as a food coating due to its hygroscopic properties and as an ingredient in the lucrative cosmetic industry. Studies have also included aloe vera as an eco-friendly green solution based on these properties. The current research focuses on the use of aloe vera gel in printing pastes as an alternative sustainable solution to synthetic thickeners, evaluating its wet performance and ease of fabric stitching, and has been inspired by studies that similarly used this substance and measured its effect on the fabric’s coefficient of friction and antimicrobial action. In the current study, printing pastes with natural colourants, such as saffron, curcumin, and annatto, and aloe vera gel thickener derived from natural leaves from Crete increased the fabric’s mechanical resistance to abrasion compared to the untreated pastes. The measured performance did not differ substantially from prints with traditional synthetic pastes, hence tolerating the substitution with the non-contaminant variant. The enhanced resistance to abrasion and wear extends the fabric’s serviceable life and resulting garments, decreasing the need for high industry processing volumes and, as a result, reducing pollution. The resistance to wear was evaluated using the dominant method in textile testing of the Martindale apparatus, which measured the cycles to failure, weight loss, and general appearance deterioration using the official photographic standards. Full article
(This article belongs to the Special Issue Coatings for Antimicrobial Textiles)
Show Figures

Figure 1

Figure 1
<p>Experimental set up of (<b>a</b>) abrasion resistance on Martindale abrasion apparatus, (<b>b</b>) microscope, and (<b>c</b>) test specimen failing point.</p>
Full article ">Figure 2
<p>Abrasion resistance cycles to failure for knitted fabric.</p>
Full article ">Figure 3
<p>Abrasion resistance cycles to failure for woven fabric.</p>
Full article ">Figure 4
<p>Percentage mass loss due to abrasion for knitted fabric.</p>
Full article ">Figure 5
<p>Percentage mass loss due to abrasion for woven fabric.</p>
Full article ">Figure 6
<p>Surface fuzzing of fabric samples comes at a later stage in the coated samples. (<b>a</b>) Untreated knitted fabric; (<b>b</b>) printed knitted fabric; (<b>c</b>) untreated woven fabric; (<b>d</b>) printed woven fabric.</p>
Full article ">Figure 7
<p>Longer fibrils protruding from the surface of the knitted (<b>a</b>) untreated samples compared to (<b>b</b>) aloe-treated pastes coated samples, after 6000 rubbing cycles.</p>
Full article ">Figure 8
<p>Abrasion resistance cycles to failure plot of knitted fabric substrates for stepwise statistical analysis.</p>
Full article ">Figure 9
<p>Abrasion resistance cycles to failure plot of woven fabric substrates for stepwise statistical analysis.</p>
Full article ">Figure A1
<p>The abraded fabric under SEM analysis. (<b>a</b>) Low magnification (<b>b</b>) and high magnification at 5000 cycles.</p>
Full article ">
14 pages, 49185 KiB  
Article
Investigating Influence of Mo Elements on Friction and Wear Performance of Nickel Alloy Matrix Composites in Air from 25 to 800 °C
by Jinming Zhen, Yunxiang Han, Lin Yuan, Zhengfeng Jia and Ran Zhang
Lubricants 2024, 12(11), 396; https://doi.org/10.3390/lubricants12110396 - 18 Nov 2024
Viewed by 289
Abstract
Rapid developments in aerospace and nuclear industries pushed forward the search for high-performance self-lubricating materials with low friction and wear characteristics under severe environment. In this paper, we investigated the influence of the Mo element on the tribological performance of nickel alloy matrix [...] Read more.
Rapid developments in aerospace and nuclear industries pushed forward the search for high-performance self-lubricating materials with low friction and wear characteristics under severe environment. In this paper, we investigated the influence of the Mo element on the tribological performance of nickel alloy matrix composites from room temperature to 800 °C under atmospheric conditions. The results demonstrated that composites exhibited excellent lubricating (with low friction coefficients of 0.19–0.37) and wear resistance properties (with low wear rates of 2.5–28.1 × 10−5 mm3/Nm), especially at a content of elemental Mo of 8 wt. % and 12 wt. %. The presence of soft metal Ag on the sliding surface as solid lubricant resulted in low friction and wear rate in a temperature range from 25 to 400 °C, while at elevated temperatures (600 and 800 °C), the effective lubricant contributed to the formation of a glazed layer rich in NiCr2O4, BaF2/CaF2, and Ag2MoO4. SEM, EDS, and the Raman spectrum indicated that abrasive and fatigue wear were the main wear mechanisms for the studied composites during sliding against the Si3N4 ceramic ball. The obtained results provide an insightful suggestion for future designing and fabricating solid lubricant composites with low friction and wear properties. Full article
(This article belongs to the Special Issue Tribology in Manufacturing Engineering)
Show Figures

Figure 1

Figure 1
<p>COF vs. testing temperature of nickel alloy matrix composites.</p>
Full article ">Figure 2
<p>COF vs. testing temperature of nickel alloy matrix composites.</p>
Full article ">Figure 3
<p>Three-dimensional morphologies of worn surface for 0Mo and 5Mo composites: (<b>a</b>) 0Mo 25 °C, (<b>b</b>) 0Mo 400 °C, (<b>c</b>) 0Mo 600 °C, (<b>d</b>) 0Mo 800 °C; (<b>e</b>) 5Mo 25 °C, (<b>f</b>) 5Mo 400 °C, (<b>g</b>) 5Mo 600 °C, (<b>h</b>) 5Mo 800 °C.</p>
Full article ">Figure 4
<p>Three-dimensional morphologies of worn surface for 8Mo and 12Mo composites: (<b>a</b>) 8Mo 25 °C, (<b>b</b>) 8Mo 400 °C, (<b>c</b>) 8Mo 600 °C, (<b>d</b>) 8Mo 800 °C; (<b>e</b>) 12Mo 25 °C, (<b>f</b>) 12Mo 400 °C, (<b>g</b>) 12Mo 600 °C, (<b>h</b>) 12Mo 800 °C.</p>
Full article ">Figure 5
<p>SEM images of worn surfaces for composites at 25 °C: (<b>a</b>) 0Mo, (<b>b</b>) 5Mo, (<b>c</b>) 12Mo.</p>
Full article ">Figure 6
<p>SEM images of worn surfaces for composites at 400 °C: (<b>a</b>) 0Mo, (<b>b</b>) 5Mo, (<b>c</b>) 8Mo.</p>
Full article ">Figure 7
<p>SEM images of the worn surfaces for the composites at 600 °C: (<b>a</b>) 0Mo, (<b>b</b>) 5Mo, (<b>c</b>) 8Mo.</p>
Full article ">Figure 8
<p>SEM images of the worn surfaces for the composites at 800 °C: (<b>a</b>) 0Mo, (<b>b</b>) 5Mo, (<b>c</b>) 8Mo.</p>
Full article ">Figure 9
<p>SEM images of worn scar for Si<sub>3</sub>N<sub>4</sub> ceramic ball: (<b>a</b>) 0Mo 25 °C, (<b>b</b>) 0Mo 400 °C, (<b>c</b>) 0Mo 800 °C, (<b>d</b>) 5Mo 25 °C, (<b>e</b>) 5Mo 400 °C, (<b>f</b>) 5Mo 800 °C.</p>
Full article ">Figure 10
<p>SEM images and corresponding EDS mapping of worn surface for Si<sub>3</sub>N<sub>4</sub> ceramic ball sliding against 0Mo composite at 800 °C.</p>
Full article ">Figure 11
<p>SEM images and corresponding EDS mapping of worn surface for Si<sub>3</sub>N<sub>4</sub> ceramic ball sliding against 5Mo composite at 800 °C.</p>
Full article ">Figure 12
<p>SEM images of worn scar for Si<sub>3</sub>N<sub>4</sub> ceramic ball: (<b>a</b>) 8Mo 25 °C, (<b>b</b>) 8Mo 400 °C, (<b>c</b>) 8Mo 800 °C, (<b>d</b>) 12Mo 800 °C.</p>
Full article ">Figure 13
<p>SEM images and corresponding EDS mapping of worn scar for Si<sub>3</sub>N<sub>4</sub> ceramic ball sliding against 8Mo composite at 25 °C.</p>
Full article ">Figure 14
<p>SEM images of worn scar for Si<sub>3</sub>N<sub>4</sub> ceramic ball and corresponding EDS mapping coupled with 8Mo composite at 400 °C.</p>
Full article ">Figure 15
<p>SEM images and corresponding EDS mapping of worn surface for Si<sub>3</sub>N<sub>4</sub> ceramic ball coupled with 8Mo composite at 800 °C.</p>
Full article ">Figure 16
<p>SEM images and corresponding EDS mapping of worn surface for Si<sub>3</sub>N<sub>4</sub> ceramic ball sliding against 12Mo composite at 800 °C.</p>
Full article ">Figure 17
<p>SEM images of 8Mo composite and corresponding element distribution at 800 °C.</p>
Full article ">Figure 18
<p>SEM images of sliding surface for 12Mo composite and corresponding element distribution at 800 °C.</p>
Full article ">Figure 19
<p>Raman spectrum for nickel alloy matrix composite: (<b>a</b>) worn surface for four composites at 800 °C, (<b>b</b>) 12Mo composite at 25, 200, 400, and 600 °C.</p>
Full article ">
20 pages, 8837 KiB  
Article
Self-Reinforced Composite Materials: Frictional Analysis and Its Implications for Prosthetic Socket Design
by Yogeshvaran R. Nagarajan, Yasasween Hewavidana, Emrah Demirci, Yong Sun, Farukh Farukh and Karthikeyan Kandan
Materials 2024, 17(22), 5629; https://doi.org/10.3390/ma17225629 - 18 Nov 2024
Viewed by 308
Abstract
Friction and wear characteristics play a critical role in the functionality and durability of prosthetic sockets, which are essential components in lower-limb prostheses. Traditionally, these sockets are manufactured from bulk polymers or composite materials reinforced with advanced carbon, glass, and Kevlar fibres. However, [...] Read more.
Friction and wear characteristics play a critical role in the functionality and durability of prosthetic sockets, which are essential components in lower-limb prostheses. Traditionally, these sockets are manufactured from bulk polymers or composite materials reinforced with advanced carbon, glass, and Kevlar fibres. However, issues of accessibility, affordability, and sustainability remain, particularly in less-resourced regions. This study investigates the potential of self-reinforced polymer composites (SRPCs), including poly-lactic acid (PLA), polyethylene terephthalate (PET), glass fibre (GF), and carbon fibre (CF), as sustainable alternatives for socket manufacturing. The tribological behaviour of these self-reinforced polymers (SrPs) was evaluated through experimental friction tests, comparing their performance to commonly used materials like high-density polyethylene (HDPE) and polypropylene (PP). Under varying loads and rotational speeds, HDPE and PP exhibited lower coefficients of friction (COF) compared to SrPLA, SrPET, SrGF, and SrCF. SrPLA recorded the highest average COF of 0.45 at 5 N and 240 rpm, while SrPET demonstrated the lowest COF of 0.15 under the same conditions. Microscopic analysis revealed significant variations in wear depth, with SrPLA showing the most profound wear, followed by SrCF, SrGF, and SrPET. In all cases, debris from the reinforcement adhered to the steel ball surface, influencing the COF. While these findings are based on friction tests against steel, they provide valuable insights into the durability and wear resistance of SRPCs, a crucial consideration for socket applications. This study highlights the importance of tribological analysis for optimising prosthetic socket design, contributing to enhanced functionality and comfort for amputees. Further research, including friction testing with skin-contact scenarios, is necessary to fully understand the implications of these materials in real-world prosthetic applications. Full article
(This article belongs to the Special Issue Advances in Functional Polymers and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>Schematic of the 2/2 twill weave fabric architecture (<b>a</b>) as received and (<b>b</b>) after vacuum consolidation at elevated temperature. (<b>c</b>) Sketch representing the geometrical details of the specimen used for pin-on-disc friction tests.</p>
Full article ">Figure 2
<p>Working principle of X-ray µCT system methodology.</p>
Full article ">Figure 3
<p>Coefficient of friction response of (<b>a</b>) high-density polyethylene and (<b>b</b>) polypropylene polymer samples recorded at the constant 5 N load under various rotation speeds.</p>
Full article ">Figure 4
<p>Coefficient of friction (COF) curves of (<b>a</b>) neat PLA, (<b>b</b>) srPLA, (<b>c</b>) neat PET, and (<b>d</b>) srPET samples recorded during sliding at a constant load of 5 N under various rotational speeds.</p>
Full article ">Figure 5
<p>Coefficient of friction (COF) curves of (<b>a</b>) srCF and (<b>b</b>) srGF composite samples recorded during sliding at a constant load of 5 N under various rotational speeds.</p>
Full article ">Figure 6
<p>Average coefficient of friction as a function of contact loads at 120 and 240 rpm: (<b>a</b>) srPLA, (<b>b</b>) srPET, (<b>c</b>) srCF, and (<b>d</b>) srGF. In each case, the coefficient of friction values for the neat matrix is also compared.</p>
Full article ">Figure 7
<p>Microscopic image showing the wear track of (<b>a</b>) srPLA; (<b>b</b>) srPET; (<b>c</b>) srGlass Fibre; (<b>d</b>) srCarbon Fibre.</p>
Full article ">Figure 8
<p>Surface roughness at 240 rpm for SrPLA, SrPET, SrGlass fibre, SrCarbon fibre.</p>
Full article ">Figure 9
<p>Wear track at 240 rpm in different load conditions: (<b>a</b>) SrPLA, (<b>b</b>) SrPET, (<b>c</b>) SrGF, and (<b>d</b>) SrCF.</p>
Full article ">Figure 10
<p>CT scans revealing wear effects on polymer composite specimens: (<b>a</b>) SrPLA, (<b>b</b>) SrPET, (<b>c</b>) SrGF, and (<b>d</b>) SrCF (each specimen size is 20 mm × 20 mm).</p>
Full article ">Figure 11
<p>The contacting ball under 5 N at 240 rpm and the depth profile of the track: (<b>a</b>) the steel ball before testing the steel ball’s surface tested against the (<b>b</b>) srPLA, (<b>c</b>) srPET, (<b>d</b>) srGF, and (<b>e</b>) srCF composites.</p>
Full article ">Figure 12
<p>Comparison of wear rate and coefficient of friction of self-reinforced composites with other polymer materials [<a href="#B38-materials-17-05629" class="html-bibr">38</a>,<a href="#B41-materials-17-05629" class="html-bibr">41</a>,<a href="#B42-materials-17-05629" class="html-bibr">42</a>,<a href="#B43-materials-17-05629" class="html-bibr">43</a>,<a href="#B44-materials-17-05629" class="html-bibr">44</a>,<a href="#B45-materials-17-05629" class="html-bibr">45</a>,<a href="#B46-materials-17-05629" class="html-bibr">46</a>,<a href="#B47-materials-17-05629" class="html-bibr">47</a>,<a href="#B48-materials-17-05629" class="html-bibr">48</a>,<a href="#B49-materials-17-05629" class="html-bibr">49</a>].</p>
Full article ">
19 pages, 14171 KiB  
Article
Mechanical, Tribological, and Corrosion Resistance Properties of (TiAlxCrNbY)Ny High-Entropy Coatings Synthesized Through Hybrid Reactive Magnetron Sputtering
by Nicolae C. Zoita, Mihaela Dinu, Anca C. Parau, Iulian Pana and Adrian E. Kiss
Crystals 2024, 14(11), 993; https://doi.org/10.3390/cryst14110993 (registering DOI) - 17 Nov 2024
Viewed by 304
Abstract
This study investigates the effects of aluminum and nitrogen content on the microstructure, mechanical properties, and tribological performance of high-entropy coatings based on (TiCrAlxNbY)Ny systems. Using a hybrid magnetron sputtering technique, both metallic and nitride coatings were synthesized and evaluated. [...] Read more.
This study investigates the effects of aluminum and nitrogen content on the microstructure, mechanical properties, and tribological performance of high-entropy coatings based on (TiCrAlxNbY)Ny systems. Using a hybrid magnetron sputtering technique, both metallic and nitride coatings were synthesized and evaluated. Increasing the aluminum concentration led to a transition from a crystalline to a nanocrystalline and nearly amorphous (NC/A) structure, with the TiAl0.5CrNbY sample (11.8% Al) exhibiting the best balance of hardness (6.8 GPa), elastic modulus (87.1 GPa), and coefficient of friction (0.64). The addition of nitrogen further enhanced these properties, transitioning the coatings to a denser fine-grained FCC structure. The HN2 sample (45.8% nitrogen) displayed the highest hardness (21.8 GPa) but increased brittleness, while the HN1 sample (32.9% nitrogen) provided an optimal balance of hardness (14.3 GPa), elastic modulus (127.5 GPa), coefficient of friction (0.60), and wear resistance (21.2 × 10−6 mm3/Nm). Electrochemical impedance spectroscopy revealed improved corrosion resistance for the HN1 sample due to its dense microstructure. Overall, the (TiAl0.5CrNbY)N0.5 coating achieved the best performance for friction applications, such as break and clutch systems, requiring high coefficients of friction, high wear resistance, and durability. Full article
(This article belongs to the Special Issue Advances of High Entropy Alloys)
Show Figures

Figure 1

Figure 1
<p>2ϴ/ϴ X-ray diffraction patterns corresponding to H1–H4 coatings.</p>
Full article ">Figure 2
<p>AFM surface images (3 × 3 µm<sup>2</sup>) corresponding to (<b>a</b>) H1 and (<b>b</b>) H4 samples. Cross-sectional HR-SEM images corresponding to (<b>c</b>) H1 and (<b>d</b>) H4 samples.</p>
Full article ">Figure 3
<p>Elemental cross-sectional mapping corresponding to H1 coating, 2.15 × 1.54 μm<sup>2</sup>.</p>
Full article ">Figure 4
<p>(<b>a</b>) The averaged values of hardness (H) and Young’s modulus (E). (<b>b</b>) Wear rate. (<b>c</b>) Friction coefficient evolution.</p>
Full article ">Figure 5
<p>(<b>a</b>) 2θ/θ X-ray diffraction patterns and (<b>b</b>) XRR experimental (scattered points) and simulated patterns (continuous lines) corresponding to samples H3, HN1, and HN2; (<b>c</b>) average mass density variation with nitrogen content.</p>
Full article ">Figure 6
<p>AFM surface images (3 × 3 µm<sup>2</sup>) corresponding to (<b>a</b>) H3, (<b>b</b>) HN1, and (<b>c</b>) HN2 samples. Cross-sectional HR-SEM images corresponding to (<b>d</b>) H3, (<b>e</b>) HN1, and (<b>f</b>) HN2 samples.</p>
Full article ">Figure 7
<p>Mechanical and tribological properties of (TiAl<sub>0.5</sub>CrNbY)N<sub>y</sub>/C45 (0 ≤ y ≤ 0.85). (<b>a</b>) Hardness (H) and Young’s modulus (E). (<b>b</b>) Coefficient of friction. (<b>c</b>) Wear rate.</p>
Full article ">Figure 8
<p>SEM micrographs of wear tracks after tribological test corresponding to samples (<b>a</b>) H3 (×300), (<b>b</b>) HN1 (×500), and HN2 (×500). Figures (<b>d</b>) and (<b>e</b>) are magnified views (×1000) of (<b>b</b>) and (<b>c</b>), respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>) Nyquist, (<b>b</b>) Bode magnitude, and (<b>c</b>) phase diagrams.</p>
Full article ">
15 pages, 15671 KiB  
Article
Dry Cold Forging of High Strength AISI316 Wires by Massively Nitrogen Supersaturated CoCrMo Dies
by Tatsuhiko Aizawa, Tatsuya Fukuda and Tomomi Shiratori
Processes 2024, 12(11), 2561; https://doi.org/10.3390/pr12112561 - 16 Nov 2024
Viewed by 212
Abstract
The plasma immersion nitriding system was utilized to make massive nitrogen supersaturation (MNS) to CoCrMo disc and die substrates at 723 K for 21.6 ks. The top layer thickness in the multi-layered MNSed layer was 20 μm. Its nitrogen solute content reached 5 [...] Read more.
The plasma immersion nitriding system was utilized to make massive nitrogen supersaturation (MNS) to CoCrMo disc and die substrates at 723 K for 21.6 ks. The top layer thickness in the multi-layered MNSed layer was 20 μm. Its nitrogen solute content reached 5 mass% on average after SEM-EDX analysis. The surface hardness was 1300 HV1N (HV0.1), which was much higher than the bare CoCrMo with 450 HV1N. The original polycrystalline structure was modified to be a multi-layered microstructure, which consisted of the nanograined MNSed top layer, the buffer layer with a thickness of 5 μm, and the column–granular structured layer with their textured crystallographic orientations. The BOD (ball-on-disc) testing was employed to describe the frictional sliding behavior under the applied loads of 5 N and 10 N and the sliding velocity of 0.1 m/s against the AISI316 ball. The friction coefficient was held constant by 0.68 on average. The CNC (Computer Numerical Control) stamping system was employed to upset the fine-grained 1.0 mm thick AISI316 wire up to 70% in reduction in thickness. The friction coefficient at RT was estimated to be 0.05. A round, fine-grained AISI316 wire was shaped into a thin plate with a thickness of 0.3 mm in cold and dry. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic view of the surface-engineered die to have a multi-layered system by massive nitrogen supersaturation into a standard polycrystalline die material.</p>
Full article ">Figure 2
<p>Nitrogen plasma immersion system for nitrogen supersaturation process. (<b>a</b>) Its schematic view, and (<b>b</b>) its overview.</p>
Full article ">Figure 3
<p>Nitrogen–hydrogen plasma sheath surrounding the CoCrMo die at 723 K.</p>
Full article ">Figure 4
<p>Upsetting experimental system. (<b>a</b>) A schematic view of the upsetting experimental unit, and (<b>b</b>) an overview of the upsetting system.</p>
Full article ">Figure 5
<p>SEM image on the cross-section of bare CoCrMo die materials before plasma nitriding immersion.</p>
Full article ">Figure 6
<p>Comparison of XRD diagram for CoCrMo die after MNS at 723 K for 21.6 ks.</p>
Full article ">Figure 7
<p>SEM image and nitrogen map on the cross-section of MNS-CoCrMo dies. (<b>a</b>) SEM image on the cross-section of MNSed CoCrMo, and (<b>b</b>) nitrogen mapping.</p>
Full article ">Figure 8
<p>Variation of the nitrogen solute content pointwise analyzed by EDX from the surface down to the depth of 120 μm.</p>
Full article ">Figure 9
<p>EBSD analysis on the cross-section of MNS-CoCrMo dies under low magnification. (<b>a</b>) IQ map, (<b>b</b>) inverse pole figure, and (<b>c</b>) kernel angle misorientation (KAM) map.</p>
Full article ">Figure 10
<p>High magnification EBSD analysis on the cross-section of MNS-CoCrMo dies from the top layer to the textured layer across the buffer zone. (<b>a</b>) IQ map, (<b>b</b>) IPF map, and (<b>c</b>) KAM distribution.</p>
Full article ">Figure 11
<p>Hardness depth profile on the cross-section of MNSed CoCrMo from the surface to the depth.</p>
Full article ">Figure 12
<p>Hardness mapping from the vicinity to surface to the depth at d = 1.0 mm. The textured layer in <a href="#processes-12-02561-f009" class="html-fig">Figure 9</a> and <a href="#processes-12-02561-f010" class="html-fig">Figure 10</a> down to d = 120 μm has higher hardness than 550 HV.</p>
Full article ">Figure 13
<p>Variation of the friction coefficient for the MNS-CoCrMo disc against the hard AISI316 ball under W = 5 N and V = 0.1 m/s till the sliding distance of 500 m.</p>
Full article ">Figure 14
<p>Variation in the friction coefficient for the MNS-CoCrMo disc against the hard AISI316 ball under W = 10 N and V = 0.1 m/s till the sliding distance of 100 m.</p>
Full article ">Figure 15
<p>Analysis of the worn-out volume of the AISI316 ball against the MNSed CoCrMo disc after BOD texting till L = 500 m under W = 5 N and V = 0.1 m/s.</p>
Full article ">Figure 16
<p>Variation in the cross-section view for the forged AISI316 works with increasing the reduction in thickness. (<b>a</b>) r = 10%, (<b>b</b>) r = 20%, (<b>c</b>) r = 50%, and (<b>d</b>) r = 70%. White arrows indicate the direction of plastic deformation.</p>
Full article ">Figure 17
<p>Variation in the bar width (W<sub>o</sub>), the contact interface width (W<sub>i</sub>), and the nondimensional bulging displacement (B<sub>g</sub>) with increasing r.</p>
Full article ">
12 pages, 8221 KiB  
Article
PDA Nanoparticle-Induced Lubricating Film Formation in Marine Environments: An Active Approach
by Xinqi Zou, Zhenghao Ge, Chaobao Wang and Yuyang Xi
Machines 2024, 12(11), 817; https://doi.org/10.3390/machines12110817 - 16 Nov 2024
Viewed by 349
Abstract
The low viscosity of water-lubricated films compromises their load-bearing capacity, posing challenges for practical application. Enhancing the lubrication stability of these films under load is critical for the successful use of seawater-lubricated bearings in engineering. Polydopamine (PDA) shows great potential to address this [...] Read more.
The low viscosity of water-lubricated films compromises their load-bearing capacity, posing challenges for practical application. Enhancing the lubrication stability of these films under load is critical for the successful use of seawater-lubricated bearings in engineering. Polydopamine (PDA) shows great potential to address this issue due to its strong bio-inspired adhesion and hydration lubrication properties. Thus, PDA nanoparticles and seawater suspensions were synthesized to promote adhesive lubricating film formation under dynamic friction. The lubrication properties of PDA suspensions were evaluated on Cu ball and ultra-high molecular weight polyethylene (UHMWPE) tribo-pairs, with a detailed comparison to seawater. The results show PDA nanoparticles provide excellent adhesion and lubrication, enhancing the formation of lubricating films during friction with seawater. Under identical conditions, PDA suspensions demonstrated the lowest friction coefficient and minimal wear. At 3 N, friction decreased by 56% and wear by 47% compared to distilled water. These findings suggest a novel strategy for using PDA as a lubricant in seawater for engineering applications. Full article
(This article belongs to the Section Material Processing Technology)
Show Figures

Figure 1

Figure 1
<p>Preparation process of PDA nanoparticles.</p>
Full article ">Figure 2
<p>Schematic illustration of the test apparatus. (<b>a</b>) Schematic diagram of the friction tester (<b>b</b>) Schematic diagram of the copper ball and disk.</p>
Full article ">Figure 3
<p>SEM images of microsized PDA particles.</p>
Full article ">Figure 4
<p>Infrared spectrum of PDA.</p>
Full article ">Figure 5
<p>Self-polymerization process of dopamine hydrochloride.</p>
Full article ">Figure 6
<p>The effects of speed on the particle size: (<b>a</b>) 100 r/min, (<b>b</b>) 200 r/min, (<b>c</b>) 300 r/min, (<b>d</b>) 400 r/min, (<b>e</b>) 500 r/min, (<b>f</b>) 600 r/min.</p>
Full article ">Figure 7
<p>Friction coefficients of polymer disks under various loads: (<b>a</b>) 0.5 N, (<b>b</b>) 1 N, (<b>c</b>) 3 N.</p>
Full article ">Figure 8
<p>Average friction coefficient (<b>a</b>), wear rates (<b>b</b>) of UHMWPE polymer disks at various applied loads.</p>
Full article ">Figure 9
<p>The 3D worn surface topographies after friction tests under 3 N load: (<b>a</b>) Seawater, (<b>b</b>) 0.1 mg/mL-PDA, (<b>c</b>) 0.25 mg/mL-PDA, (<b>d</b>) 0.5 mg/mL-PDA, (<b>e</b>) 0.75 mg/mL-PDA, (<b>f</b>) 1.0 mg/mL-PDA.</p>
Full article ">Figure 10
<p>The SEM and nitrogen EDS images of copper friction pairs after rubbing tests (3 N). (<b>a</b>) Surface morphology of the Cu ball after friction in a seawater environment. (<b>b</b>) N element distribution on the surface of the Cu ball after friction in a seawater environment. (<b>c</b>) Surface morphology of the Cu ball after friction in a PDA &amp; seawater suspension environment. (<b>d</b>) N element distribution on the surface of the Cu ball after friction in a PDA &amp; seawater suspension environment.</p>
Full article ">Figure 11
<p>The SEM and nitrogen EDS images of UHMWPE friction pairs after rubbing tests (3 N). (<b>a</b>) Microscopic surface of UHMWPE after friction in a seawater environment. (<b>b</b>) Microscopic surface of UHMWPE after friction in a PDA &amp; seawater environment.</p>
Full article ">Figure 12
<p>Analysis of the lubrication mechanism of PDA nanoparticles.</p>
Full article ">
26 pages, 14835 KiB  
Article
Mechanical and Tribological Properties of (AlCrNbSiTiMo)N High-Entropy Alloy Films Prepared Using Single Multiple-Element Powder Hot-Pressed Sintered Target and Their Practical Application in Nickel-Based Alloy Milling
by Jeng-Haur Horng, Wen-Hsien Kao, Wei-Chen Lin and Ren-Hao Chang
Lubricants 2024, 12(11), 391; https://doi.org/10.3390/lubricants12110391 - 14 Nov 2024
Viewed by 434
Abstract
(AlCrNbSiTiMo)N high-entropy alloy films with different nitrogen contents were deposited on tungsten carbide substrates using a radio-frequency magnetron sputtering system. Two different types of targets were used in the sputtering process: a hot-pressing sintered AlCrNbSiTi target fabricated using a single powder containing multiple [...] Read more.
(AlCrNbSiTiMo)N high-entropy alloy films with different nitrogen contents were deposited on tungsten carbide substrates using a radio-frequency magnetron sputtering system. Two different types of targets were used in the sputtering process: a hot-pressing sintered AlCrNbSiTi target fabricated using a single powder containing multiple elements and a vacuum arc melting Mo target. The deposited films were denoted as RN0, RN33, RN43, RN50, and RN56, where RN indicates the nitrogen flow ratio relative to the total nitrogen and argon flow rate (RN = (N2/(N2 + Ar)) × 100%). The as-sputtered films were vacuum annealed, with the resulting films denoted as HRN0, HRN33, HRN43, HRN50, and HRN56, respectively. The effects of the nitrogen content on the composition, microstructure, mechanical properties, and tribological properties of the films, in both as-sputtered and annealed states, underwent thorough analysis. The RN0 and RN33 films displayed non-crystalline structures. However, with an increase in nitrogen content, the RN43, RN50, and RN56 films transitioned to FCC structures. Among the as-deposited films, the RN43 film exhibited the best mechanical and tribological properties. All of the annealed films, except for the HRN0 film, displayed an FCC structure. In addition, they all formed an MoO3 solid lubricating phase, which reduced the coefficient of friction and improved the anti-wear performance. The heat treatment HRN43 film displayed the supreme hardness, H/E ratio, and adhesion strength. It also demonstrated excellent thermal stability and the best wear resistance. As a result, in milling tests on Inconel 718, the RN43-coated tool demonstrated a significantly lower flank wear and notch wear, indicating an improved machining performance and extended tool life. Thus, the application of the RN43 film in aerospace manufacturing can effectively reduce the tool replacement cost. Full article
(This article belongs to the Special Issue Recent Advances in Tribological Properties of Machine Tools)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Diagrammatic sketch of sputtering target configuration, and (<b>b</b>) diagrammatic sketch of film design.</p>
Full article ">Figure 2
<p>Morphology and elemental composition of new AlCrNbSiTi powder for hot-pressing sintering targets.</p>
Full article ">Figure 3
<p>Schematic showing tool wear and notch wear measurement positions.</p>
Full article ">Figure 4
<p>Element content of (<b>a</b>) as-sputtered and (<b>b</b>) heat-treated (AlCrNbSiTiMo)N films.</p>
Full article ">Figure 5
<p>X-ray diffraction patterns of (<b>a</b>) as-sputtered and (<b>b</b>) heat-treated (AlCrNbSiTiMo)N films.</p>
Full article ">Figure 6
<p>Cross-section SEM graphics of (AlCrNbSiTiMo)N films: (<b>a</b>) R<sub>N0</sub>, (<b>b</b>) HR<sub>N0</sub>, (<b>c</b>) R<sub>N33</sub>, (<b>d</b>) HR<sub>N33</sub>, (<b>e</b>) R<sub>N43</sub>, (<b>f</b>) HR<sub>N43</sub>, (<b>g</b>) R<sub>N50</sub>, (<b>h</b>) HR<sub>N50</sub>, (<b>i</b>) R<sub>N56</sub>, and (<b>j</b>) HR<sub>N56</sub>.</p>
Full article ">Figure 7
<p>SEM surface graphics of (AlCrNbSiTiMo)N films: (<b>a</b>) R<sub>N0</sub>, (<b>b</b>) HR<sub>N0</sub>, (<b>c</b>) R<sub>N33</sub>, (<b>d</b>) HR<sub>N33</sub>, (<b>e</b>) R<sub>N43</sub>, (<b>f</b>) HR<sub>N43</sub>, (<b>g</b>) R<sub>N50</sub>, (<b>h</b>) HR<sub>N50</sub>, (<b>i</b>) R<sub>N56</sub>, and (<b>j</b>) HR<sub>N56</sub>.</p>
Full article ">Figure 8
<p>OM graphics of scratch track on R<sub>N33</sub> film.</p>
Full article ">Figure 9
<p>Curves of COF for (<b>a</b>) as-deposited (AlCrNbSiTiMo)N films and substrate (WC), and (<b>b</b>) annealed (AlCrNbSiTiMo)N films and substrate (WC).</p>
Full article ">Figure 10
<p>SEM graphics and EDS results of worn traces on (<b>a</b>) R<sub>N0</sub> and (<b>b</b>) R<sub>N43</sub> films and (<b>c</b>) R<sub>N43</sub> film enlarged image and (Note: For comparative analysis, the originally existed elemental compositions of the R<sub>N0</sub>, R<sub>N43</sub> films are also shown).</p>
Full article ">Figure 11
<p>SEM graphics and EDS results for worn traces on (<b>a</b>) HR<sub>N0</sub> film, (<b>b</b>) HR<sub>N43</sub> film (regular magnification), and (<b>c</b>) HR<sub>N43</sub> film (high magnification). (Note: For comparative analysis, the originally existed elemental compositions of the HR<sub>N0</sub> and HR<sub>N43</sub> films are also provided).</p>
Full article ">Figure 12
<p>OM graphics showing the flank wear and notch wear on (<b>a</b>) bare WC cutting tool and (<b>b</b>) R<sub>N43</sub> coated WC cutting tool completing a total processing distance of 18 m.</p>
Full article ">Figure 13
<p>(<b>a</b>) Flank wear and (<b>b</b>) Notch wear of bare WC cutting tool and R<sub>N43</sub> coated WC cutting tool compared with other coated tools after completing three processing distances of 6, 12, and 18 m, individually.</p>
Full article ">
14 pages, 10340 KiB  
Article
Increasing the Wear Resistance of CrWMn Tool Steel Surfaces by Plasma Electrolytic Nitriding and Polishing
by Sergey N. Grigoriev, Tatiana L. Mukhacheva, Ivan V. Tambovskiy, Irina A. Kusmanova, Tatiana M. Golubeva, Pavel A. Podrabinnik, Roman S. Khmyrov, Igor V. Suminov and Sergei A. Kusmanov
Appl. Sci. 2024, 14(22), 10488; https://doi.org/10.3390/app142210488 - 14 Nov 2024
Viewed by 286
Abstract
The positive effect of plasma electrolytic treatment on CrWMn tool steel to increase the wear resistance of its surface is shown. The effect of plasma electrolytic nitriding and subsequent polishing on the structure, phase and elemental composition, microhardness of the surface layer, and [...] Read more.
The positive effect of plasma electrolytic treatment on CrWMn tool steel to increase the wear resistance of its surface is shown. The effect of plasma electrolytic nitriding and subsequent polishing on the structure, phase and elemental composition, microhardness of the surface layer, and surface morphology is established. Steel nitriding leads to the formation of a modified surface layer including Fe2–3N iron nitride and nitrogen martensite, below which hardening martensite is formed, reaching a microhardness value of 1200 HV. Subsequent polishing leads to a decrease in surface roughness by 42–68%. Tribological tests were carried out according to the shaft-bushing scheme. A decrease in the friction coefficient and weight wear of up to 2.6 and 30.1 times, respectively, is shown. The formed structure of the surface layer compensates for the effect of the counter body and determines the destruction of friction bonds by plastic displacement. The wear mechanism has been established and is defined as fatigue wear under dry friction and plastic contact. Full article
(This article belongs to the Section Materials Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>PEN and PEP installation scheme: 1—ventilation duct; 2—protective screen of the working chamber; 3—linear drive; 4—workpiece-electrode (anode); 5—cylindrical cell-electrode (cathode); 6—working chamber; 7—valve with electric drive; 8—flow meter; 9—heat exchanger; 10—pump; 11—water filter.</p>
Full article ">Figure 2
<p>Friction scheme and unit: 1—cylindrical sample; 2—counter body; 3, 4—shaft; 5—crank; 6—pneumatic cylinder; 7—guides; 8—table; 9—strain gauges.</p>
Full article ">Figure 3
<p>X-ray diffraction pattern of the CrWMn steel surface after various types of treatment: (<b>a</b>) PEN; (<b>b</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 1 min; (<b>c</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for and 2 min; (<b>d</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 1 min; (<b>e</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 2 min.</p>
Full article ">Figure 4
<p>SEM image of the cross-section of the surface layer of CrWMn steel after various types of treatment: (<b>a</b>) PEN; (<b>b</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 1 min; (<b>c</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 2 min; (<b>d</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 1 min; (<b>e</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 2 min.</p>
Full article ">Figure 5
<p>Nitrogen concentration distribution in the surface layer of CrWMn steel after various types of treatment. ChG—the solutions of ammonium chloride and glycerol, S—the solutions of ammonium sulfate.</p>
Full article ">Figure 6
<p>Data from EDX analysis of the cross-section of CrWMn steel sample after PEN.</p>
Full article ">Figure 7
<p>Microstructure of the cross-section of the CrWMn steel surface after PEN: 1—outer layer with inclusions of iron nitrides and residual austenite, 2—martensitic structure with inclusions of iron carbides.</p>
Full article ">Figure 8
<p>Microhardness distribution in the surface layer of CrWMn steel after various types of treatment. ChG—the solutions of ammonium chloride and glycerol, S—the solutions of ammonium sulfate.</p>
Full article ">Figure 9
<p>Morphology of the CrWMn steel surface after various types of treatment: (<b>a</b>) Untreated; (<b>b</b>) PEN; (<b>c</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 1 min; (<b>d</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 2 min; (<b>e</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 1 min; (<b>f</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 2 min.</p>
Full article ">Figure 10
<p>Dependence of the friction coefficient of samples made of CrWMn steel after various types of treatment. ChG—the solutions of ammonium chloride and glycerol, S—the solutions of ammonium sulfate.</p>
Full article ">Figure 11
<p>Morphology of friction tracks on the CrWMn steel surface after various types of treatment: (<b>a</b>) Untreated; (<b>b</b>) PEN; (<b>c</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 1 min; (<b>d</b>) PEN and PEP in a solution of NH<sub>4</sub>Cl and C<sub>3</sub>H<sub>8</sub>O<sub>3</sub> for 2 min; (<b>e</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 1 min; (<b>f</b>) PEN and PEP in (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> solution for 2 min.</p>
Full article ">
17 pages, 16140 KiB  
Article
An Investigation on the High-Temperature Stability and Tribological Properties of Impregnated Graphite
by Juying Zhao, Qi Xin, Yunshuang Pang, Xiao Ning, Lingcheng Kong, Guangyang Hu, Ying Liu, Haosheng Chen and Yongjian Li
Lubricants 2024, 12(11), 388; https://doi.org/10.3390/lubricants12110388 - 13 Nov 2024
Viewed by 378
Abstract
Impregnated graphite is a common material for friction pairs in aeroengine seals, especially at high temperatures. For the convenience of the application of graphite materials in aeroengines, an SRV-4 tribometer and a synchronous thermal analyzer are employed to study the tribological properties and [...] Read more.
Impregnated graphite is a common material for friction pairs in aeroengine seals, especially at high temperatures. For the convenience of the application of graphite materials in aeroengines, an SRV-4 tribometer and a synchronous thermal analyzer are employed to study the tribological properties and thermal stability of pure, resin-impregnated, metal-impregnated, and phosphate-impregnated graphite against stainless steel from room temperature to 500 °C. The results indicate that impregnations can improve the wear resistance and thermal stability of graphite at high temperatures. Compared with other impregnated graphite materials, the resin-impregnated graphite shows a good friction coefficient and poor wear rate and thermal stability over 300 °C, due to the degradation and oxidation of the resin-and-graphite matrix. The metal- and phosphate-impregnated graphite materials exhibit excellent wear resistance and thermal stability under 500 °C as a result of the protection of the impregnations, while the average friction coefficient of the metal-impregnated graphite is much greater than the phosphate-impregnated graphite, and even reaches 2.14-fold at 300 °C. The wear rates for the graphite impregnated with resin, metal, and phosphate are 235 × 10−7, 7 × 10−7, and 16 × 10−7 mm3N−1m−1 at 500 °C, respectively. Considering all aspects, the phosphate-impregnated graphite exhibits excellent tribological properties and thermal stability. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of friction test and test samples: (<b>a</b>) friction test and sliding directions; (<b>b</b>) stainless steel sample (left side—top view; right side—bottom view); (<b>c</b>) graphite sample.</p>
Full article ">Figure 2
<p>Friction coefficients of graphite materials under different working conditions. (<b>a</b>) Average friction coefficient; (<b>b</b>) pure graphite; (<b>c</b>) resin-impregnated graphite; (<b>d</b>) metal-impregnated graphite; (<b>e</b>) phosphate-impregnated graphite.</p>
Full article ">Figure 3
<p>The wear rates of the four graphite materials at different temperatures. (“×” indicates that the group did not undergo a friction test for safety reasons).</p>
Full article ">Figure 4
<p>Surface morphologies of four graphite materials under different working conditions. (The ellipsis indicates that the group has not undergone a friction test for safety reasons).</p>
Full article ">Figure 5
<p>Mass loss rates of different graphite materials after 5 h prolonged heating tests. (The blue arrow indicates the correlation between the weight loss percentage of the graphite and the images of the graphite samples after the tests).</p>
Full article ">Figure 6
<p>Test results of thermal stability for different graphite materials. (<b>a</b>) Thermogravimetric curves, (<b>b</b>) DSC curves, (<b>c</b>) FTIR curves.</p>
Full article ">Figure 7
<p>Normalized hardness of different graphite materials under high temperature.</p>
Full article ">Figure 8
<p>The SEM results of the pure graphite and resin-impregnated graphite at R.T. and 500 °C: (<b>a</b>) clean pure graphite at R.T., (<b>b</b>) worn pure graphite at R.T., (<b>c</b>) clean resin–graphite at R.T., (<b>d</b>) worn resin–graphite at R.T., (<b>e</b>) clean resin–graphite at 500 °C, (<b>f</b>) worn resin–graphite at 500 °C (400×), (<b>g</b>) worn resin–graphite at 500 °C (1000×).</p>
Full article ">Figure 9
<p>The SEM results of the metal-impregnated graphite at R.T. and 500 °C: (<b>a</b>) clean metal–graphite at R.T., (<b>b</b>) worn metal–graphite at R.T. (400×), (<b>c</b>) worn metal–graphite at R.T. (1000×), (<b>d</b>) clean metal–graphite at 500 °C, (<b>e</b>) worn metal–graphite at 500 °C (400×), (<b>f</b>) worn metal–graphite at 500 °C (1000×).</p>
Full article ">Figure 10
<p>Energy spectrum distribution of the worn area of the metal-impregnated graphite at R.T.: C (red), Sb (light green), O (dark green), Fe (light yellow), Ni (orange).</p>
Full article ">Figure 11
<p>Element distribution of the unworn area of metal impregnated graphite materials at 500 °C: C (red), Sb (light green), O (dark green), Fe (light yellow), Ni (orange).</p>
Full article ">Figure 12
<p>The SEM results of the phosphate-impregnated graphite at R.T. and 500 °C: (<b>a</b>) clean graphite at R.T., (<b>b</b>) worn graphite at R.T., (<b>c</b>) clean graphite at 500 °C, (<b>d</b>) worn graphite at 500 °C.</p>
Full article ">Figure 13
<p>Tribological properties of graphite materials.</p>
Full article ">
11 pages, 886 KiB  
Article
Energy Loss in Frictional Hertzian Contact Subjected to Two-Dimensional Cyclic Loadings
by Young Ju Ahn
Coatings 2024, 14(11), 1440; https://doi.org/10.3390/coatings14111440 - 13 Nov 2024
Viewed by 291
Abstract
We investigate the effect of three different harmonically varying loads as a function of the friction coefficient on energy loss in a three-dimensional discrete uncoupled frictional contact problem. Three loading cases include (1) a normal force is constant and a tangential force varies, [...] Read more.
We investigate the effect of three different harmonically varying loads as a function of the friction coefficient on energy loss in a three-dimensional discrete uncoupled frictional contact problem. Three loading cases include (1) a normal force is constant and a tangential force varies, (2) normal and tangential forces both vary, but the loading and unloading curves are identical, and (3) normal and tangential forces both vary, but the loading and unloading curves are different. For a higher coefficient of friction, three loading cases show different characteristics. If a normal force is constant and a tangential force varies, there is always some slip, but dissipation tends asymptotically to zero at large coefficient of friction. If normal and tangential forces both vary, but the loading and unloading curves are identical, there is no slip and no dissipation above a critical coefficient of friction. If the loading and unloading curves are different, dissipation occurs for all values of the coefficient of friction, and we expect that the dissipation is asymptotic to the relaxation damping value as the coefficient of friction approaches infinity. For lowering coefficient of friction, the three loading cases show similar behavior. Dissipation increases and reaches a maximum just before a state where gross slip is possible. Full article
(This article belongs to the Special Issue Advanced Wear-Resistant Materials and Coatings)
Show Figures

Figure 1

Figure 1
<p>Frictional contact between two identical elastic spheres.</p>
Full article ">Figure 2
<p>Normalized contact pressure vs. normalized distance for the Hertzian contact without friction: <span class="html-italic">p</span> is contact pressure, <math display="inline"><semantics> <msub> <mi>p</mi> <mn>0</mn> </msub> </semantics></math> is peak pressure in the center, <span class="html-italic">r</span> is the distance from the center, and <span class="html-italic">a</span> is the contact semi-width.</p>
Full article ">Figure 3
<p>Evolution of semi-contact traction for the numerical solution with <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <msub> <mi>P</mi> <mn>0</mn> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>Q</mi> <mn>1</mn> </msub> <mi>sin</mi> <mrow> <mo>(</mo> <mi>w</mi> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, the coefficient of friction <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0.35</mn> </mrow> </semantics></math>, and a period <span class="html-italic">T</span>.</p>
Full article ">Figure 4
<p>Evolution of semi-contact traction for the out-of-phase loading with <math display="inline"><semantics> <mrow> <mi>P</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>P</mi> <mn>0</mn> </msub> <mo>+</mo> <msub> <mi>P</mi> <mn>1</mn> </msub> <mi>cos</mi> <mrow> <mo>(</mo> <mi>w</mi> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>Q</mi> <mn>1</mn> </msub> <mi>sin</mi> <mrow> <mo>(</mo> <mi>w</mi> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, the coefficient of friction <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, and a period <span class="html-italic">T</span>.</p>
Full article ">Figure 5
<p>Evolution of semi-contact traction for the in-phase loading with <math display="inline"><semantics> <mrow> <mi>P</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>P</mi> <mn>0</mn> </msub> <mo>+</mo> <msub> <mi>P</mi> <mn>1</mn> </msub> <mi>cos</mi> <mrow> <mo>(</mo> <mi>w</mi> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>Q</mi> <mn>1</mn> </msub> <mi>cos</mi> <mrow> <mo>(</mo> <mi>w</mi> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, the coefficient of friction <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0.55</mn> </mrow> </semantics></math>, and a period <span class="html-italic">T</span>.</p>
Full article ">Figure 6
<p>Energy dissipation concerning the coefficient of friction <span class="html-italic">f</span> for the three-different loadings.</p>
Full article ">Figure 7
<p>Asymptotic dissipation limit represented by the dotted line for the out-of-phase loading.</p>
Full article ">
29 pages, 7257 KiB  
Article
A New Multi-Axial Functional Stress Analysis Assessing the Longevity of a Ti-6Al-4V Dental Implant Abutment Screw
by Ghada H. Naguib, Ahmed O. Abougazia, Lulwa E. Al-Turki, Hisham A. Mously, Abou Bakr Hossam Hashem, Abdulghani I. Mira, Osama A. Qutub, Abdulelah M. Binmahfooz, Afaf A. Almabadi and Mohamed T. Hamed
Biomimetics 2024, 9(11), 689; https://doi.org/10.3390/biomimetics9110689 - 12 Nov 2024
Viewed by 487
Abstract
This study investigates the impact of tightening torque (preload) and the friction coefficient on stress generation and fatigue resistance of a Ti-6Al-4V abutment screw with an internal hexagonal connection under dynamic multi-axial masticatory loads in high-cycle fatigue (HCF) conditions. A three-dimensional model of [...] Read more.
This study investigates the impact of tightening torque (preload) and the friction coefficient on stress generation and fatigue resistance of a Ti-6Al-4V abutment screw with an internal hexagonal connection under dynamic multi-axial masticatory loads in high-cycle fatigue (HCF) conditions. A three-dimensional model of the implant–abutment assembly was simulated using ANSYS Workbench 16.2 computer aided engineering software with chewing forces ranging from 300 N to 1000 N, evaluated over 1.35 × 107 cycles, simulating 15 years of service. Results indicate that the healthy range of normal to maximal mastication forces (300–550 N) preserved the screw’s structural integrity, while higher loads (≥800 N) exceeded the Ti-6Al-4V alloy’s yield strength, indicating a risk of plastic deformation under extreme conditions. Stress peaked near the end of the occluding phase (206.5 ms), marking a critical temporal point for fatigue accumulation. Optimizing the friction coefficient (0.5 µ) and preload management improved stress distribution, minimized fatigue damage, and ensured joint stability. Masticatory forces up to 550 N were well within the abutment screw’s capacity to sustain extended service life and maintain its elastic behavior. Full article
Show Figures

Figure 1

Figure 1
<p>Meshing of structures: (<b>a</b>) 3D meshing of bone. (<b>b</b>) 3D meshing assembly of implant–abutment and crown complex. (<b>c</b>) 3D meshing of final model within bone. (<b>d</b>) 3D meshing of abutment screw. (<b>e</b>) 3D meshing of implant fixture. (<b>f</b>) 3D meshing of implant–abutment assembly.</p>
Full article ">Figure 2
<p>Implant dimensions and geometry: (<b>a</b>) Implant abutment screw. (<b>b</b>) Implant abutment. (<b>c</b>) Implant fixture.</p>
Full article ">Figure 3
<p>Isometric view of the assembled model: (<b>a</b>) All Components of 3D finite element model: bone, implant fixture, abutment, abutment retaining screw, and crown. (<b>b</b>) Mandible fixation points. (<b>c</b>) Contact areas on occlusal surface of mandibular molar. (<b>d</b>) Cross-sectional view of the assembled model components.</p>
Full article ">Figure 4
<p>Mesh sensitivity analysis.</p>
Full article ">Figure 5
<p>Forces and moments acting on the abutment screw during tightening.</p>
Full article ">Figure 6
<p>Load application and mastication path.</p>
Full article ">Figure 7
<p>Load angulation and path: point “a” is the start of the traveling path of mastication that ends at point “e”.</p>
Full article ">Figure 8
<p>Loading steps invoked in the present idealization. <span class="html-italic">P</span><sub>a</sub> is the axial pressure applied on the lower surface of the screw head to represent the residual existing due to screw tightening and <span class="html-italic">P</span><sub>max</sub> is the maximum pressure on the cells of loading during masticatory contact.</p>
Full article ">Figure 9
<p>Schematic top view of a first mandible molar to show the limiting borders of possible paths of contact with the opposing maxillary molar during the occluding phase of the chewing cycle. LT: laterotrusive movements, MT: mediotrusive movements, and P: protrusive movements.</p>
Full article ">Figure 10
<p>Multi-axial fatigue damage parameter proposed by Kallmeyer et al. [<a href="#B89-biomimetics-09-00689" class="html-bibr">89</a>] to correlate their uniaxial and proportional and non-proportional biaxial experimental fatigue data for Titanium Ti-6Al-4V alloy.</p>
Full article ">Figure 11
<p>Maximum von Mises equivalent stresses (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>e</mi> <mi>q</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> recorded in implant at different intensities of mastication forces: (<b>a</b>) 300 N, (<b>b</b>) 500 N, (<b>c</b>) 800 N, and (<b>d</b>) 1000 N.</p>
Full article ">Figure 12
<p>Maximum von Mises equivalent stresses (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>e</mi> <mi>q</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> recorded in abutment screw at different intensities of mastication forces: (<b>a</b>) 300 N, (<b>b</b>) 500 N, (<b>c</b>) 800 N, and (<b>d</b>) 1000 N.</p>
Full article ">Figure 13
<p>Fatigue damage parameter experienced by the most stressed site within the retaining screw and plotted against contact time during the occluding phase of chewing; maximum fatigue damage invariably takes place at contact time = 206.5 ms.</p>
Full article ">Figure 14
<p>Fatigue damage parameter experienced at contact time = 206.5 by the most stressed site within the retaining screw and plotted against the magnitude of the chewing force.</p>
Full article ">Figure 15
<p>Fatigue damage parameter experienced at contact time = 206.5 ms plotted against the magnitude of the chewing force.</p>
Full article ">Figure 16
<p>Moment Mc for the screw materials of Ti-6Al-4V titanium alloy versus the coefficient of friction μ.</p>
Full article ">Figure 17
<p>Tightening axial pressure, Pa, plotted against the coefficient of friction μ.</p>
Full article ">Figure 18
<p>Moment <span class="html-italic">M</span><sub>c</sub> (frictional resisting moment acting on the lower contact surface of the screw, opposing the screw’s rotation) versus the coefficient of friction μ at a constant tightening torque <span class="html-italic">M</span> of 30 N/cm.</p>
Full article ">
Back to TopTop