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

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (235)

Search Parameters:
Keywords = pin-on-disc

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 5443 KiB  
Article
Carbonaceous Decomposition Products at High Temperatures and Their Interfacial Role in the Friction Behaviour of Composite Brake Material
by Piyush Chandra Verma, Pranesh Aswath, Giovanni Straffelini and Stefano Gialanella
Lubricants 2024, 12(11), 399; https://doi.org/10.3390/lubricants12110399 - 20 Nov 2024
Viewed by 306
Abstract
This study aims to investigate the outcomes of carbonaceous products, derived from the decomposition of the components of vehicular brake materials, under high-temperature wear tests. Pin-on-disc (PoD) wear tests were conducted by using cast iron discs against pins made of commercially available low-steel [...] Read more.
This study aims to investigate the outcomes of carbonaceous products, derived from the decomposition of the components of vehicular brake materials, under high-temperature wear tests. Pin-on-disc (PoD) wear tests were conducted by using cast iron discs against pins made of commercially available low-steel friction material. Tests were carried out at different temperatures: 155 °C, 200 °C, 250 °C, and 300 °C. The characterization of the sliding plateaus on worn pin surfaces was based on X-ray diffraction (XRD), scanning electron microscopy (SEM), and Raman spectroscopy. It was noted that at temperatures above 200 °C, the thermal degradation of the inorganic resin, used as a material binder, occurs. An interesting observation was recorded at 300 °C; the brake pin material’s friction curve showed higher stability despite having an excessive wear rate. However, the brake pin’s specific wear coefficient was higher at this temperature than was observed in the other friction tests. A detailed study of the friction plateaus on the worn-out pins at 300 °C revealed that the decomposed carbon resin product, i.e., the distorted graphite, was widespread over the surface of the pin. Lubricant stabilization can be expected, as established by the observed values of the coefficient of friction (CoF), retaining values within the 0.4–0.6 range, even at high temperatures. Other friction material components may have contributed to the formation of this ubiquitous carbonaceous interface film. Full article
(This article belongs to the Special Issue Recent Advances in High Temperature Tribology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cylindrical shape brake material pin placed on GCI disc. (<b>b</b>) Microstructure of virgin brake material.</p>
Full article ">Figure 2
<p>High-temperature wear test rig.</p>
Full article ">Figure 3
<p>The evolution of the friction coefficient of brake pad friction material at different temperatures.</p>
Full article ">Figure 4
<p>(<b>a</b>) Total wear volume of the worn track on the GCI disc at different elevated temperatures, (<b>b</b>,<b>c</b>) the corresponding 3D profilometer observation of the disc at 155 °C and 300 °C, and (<b>d</b>) the average elemental composition of wear debris on the GCI disc’s surface, as determined using EDXS.</p>
Full article ">Figure 5
<p>(<b>a</b>) Specific wear coefficients for brake pad pins in relation to disc temperature. SEM of worn surfaces of brake pad pins at different elevated temperatures: (<b>b</b>) 155 °C, (<b>c</b>) 200 °C, (<b>d</b>) 250 °C, and (<b>e</b>) 300 °C. (<b>f</b>,<b>g</b>) SEM images of loose wear debris collected at 155 °C and 300 °C friction tests.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Specific wear coefficients for brake pad pins in relation to disc temperature. SEM of worn surfaces of brake pad pins at different elevated temperatures: (<b>b</b>) 155 °C, (<b>c</b>) 200 °C, (<b>d</b>) 250 °C, and (<b>e</b>) 300 °C. (<b>f</b>,<b>g</b>) SEM images of loose wear debris collected at 155 °C and 300 °C friction tests.</p>
Full article ">Figure 6
<p>(<b>a</b>) Comparison of XRD spectra acquired on brake pad pin samples at 155 °C, 200 °C, 250 °C, and 300 °C, respectively. (<b>b</b>) Full-profile Rietveld fit using the MAUD software for pin samples at 250 °C.</p>
Full article ">Figure 7
<p>(<b>a</b>) A Raman spectroscopic comparison at a worn-out pin surface at various working temperatures. (<b>b</b>) A magnified image of the Raman spectra, with Ca<sub>3</sub>P<sub>2</sub>O<sub>8</sub> at 967 cm<sup>−1</sup>, SnO<sub>2</sub> at 776 cm<sup>−1</sup>, Mg<sub>2</sub>SiO<sub>4</sub> at 790 cm<sup>−1</sup>, and with ND as the corresponding wavelength.</p>
Full article ">Figure 8
<p>De-convolution of Raman spectroscopy performed on pin surface at (<b>a</b>) 155 °C, (<b>b</b>) 200 °C, (<b>c</b>) 250 °C, and (<b>d</b>) 300 °C testing temperatures.</p>
Full article ">Figure 9
<p>Wear mechanism for friction tests at (<b>a</b>) lower (155–200 °C) and (<b>b</b>) higher temperatures (250–300 °C).</p>
Full article ">
20 pages, 27328 KiB  
Article
Enhancing Wear Resistance of AA7075/SiC/Fly Ash Composites Through Friction Stir Processing
by Namdev Ashok Patil, Santoshi Pedapati and Srinivasa Rao Pedapati
J. Compos. Sci. 2024, 8(11), 461; https://doi.org/10.3390/jcs8110461 - 7 Nov 2024
Viewed by 469
Abstract
In this study, the wear behavior of AA7075/silicon carbide/fly ash hybrid surface composites processed with a clean and green friction stir processing technique was investigated. The microstructure of the composites was investigated to determine the particle dispersion. Wear tests using a pin-on-disc tribometer [...] Read more.
In this study, the wear behavior of AA7075/silicon carbide/fly ash hybrid surface composites processed with a clean and green friction stir processing technique was investigated. The microstructure of the composites was investigated to determine the particle dispersion. Wear tests using a pin-on-disc tribometer were conducted, and wear tracks and debris analyses were conducted using scanning electron microscopic imaging, EDX, and mapping. The wear rate of the composites was higher in the case of the composites with agglomerated zones, which led to the loose SiC/fly ash particles pulling out during the action of dry sliding. However, on the other hand, the wear resistance was improved in the composites with uniformly distributed SiC/fly ash particles. The hard SiC/fly ash particles acted as optimized load-bearing asperities and induced more wear resistance during the action of dry sliding against the mating plate, which was made of mild steel. In the case of the well-dispersed composites, the wear mechanisms shifted from fretting fatigue and adhesion to abrasion. The presence of a high Fe content in the wear debris was confirmed in the most wear-resistant composite sample, S-20, which was produced with the following parameters: tool rotation (w) of 1000 rpm, tool traverse (v) of 40 mm/min, hybrid ratio (HR) of 75:25, and a volume percentage of reinforcements (vol.%) of 8. Full article
(This article belongs to the Special Issue Welding and Friction Stir Processes for Composite Materials)
Show Figures

Figure 1

Figure 1
<p>SEM image presenting the morphology and particle size of fly ash.</p>
Full article ">Figure 2
<p>The wear rate of aluminum 7075/silicon carbide/fly ash composites (L<sub>27</sub>).</p>
Full article ">Figure 3
<p>Optical microscopy images (<b>a</b>–<b>d</b>) of S-9 AA7075/SiC/fly ash composite stir zone.</p>
Full article ">Figure 4
<p>Optical microscopy images (<b>a</b>–<b>d</b>) of S-20 AA7075/SiC/fly ash composite stir zone.</p>
Full article ">Figure 5
<p>SEM micrographs (<b>a</b>–<b>d</b>) of the S-9 AA7075/SiC/fly ash composite stir zone.</p>
Full article ">Figure 6
<p>FESEM micrographs of S-20 AA7075/SiC/fly ash composite stir zone with magnification (<b>a</b>) 500× (<b>b</b>) 3000× (<b>c</b>) 7000× and (<b>d</b>) 10,000×.</p>
Full article ">Figure 7
<p>FESEM micrographs of S-20 AA7075/SiC/Fly ash composite stir zone with dense particles presence (<b>a</b>) 100× (<b>b</b>) 500× (<b>c</b>) 5000× (<b>d</b>) 10,000× magnification.</p>
Full article ">Figure 8
<p>Micrographs indicating worn-out surfaces on the samples of AA7075-T651 with magnifications of (<b>a</b>) 100× and (<b>b</b>) 1000× captured using SEM.</p>
Full article ">Figure 9
<p>Micrographs indicating worn-out surfaces on the samples of S-20 composite with magnifications of (<b>a</b>) 100× and (<b>b</b>) 300× captured using SEM.</p>
Full article ">Figure 10
<p>Micrographs indicating particles pulling out on the worn-out surface of the S-9 composite with magnifications of (<b>a</b>) 100× and (<b>b</b>) 500× captured using SEM.</p>
Full article ">Figure 11
<p>Worn out surface of the confirmation test composite with magnifications of (<b>a</b>) 500×, (<b>b</b>) 100×, (<b>c</b>) 1000×, and (<b>d</b>) 1000×.</p>
Full article ">Figure 12
<p>SEM EDX and mapping results of the confirmation test AA7075/SiC/fly ash composite.</p>
Full article ">Figure 13
<p>SEM mapping results of the confirmation test AA7075/SiC/fly ash composite.</p>
Full article ">Figure 14
<p>(<b>a</b>–<b>d</b>) Mild steel disc surface after the wear test.</p>
Full article ">Figure 15
<p>(<b>a</b>–<b>d</b>) Base alloy wear debris.</p>
Full article ">Figure 16
<p>SEM images (<b>a</b>–<b>d</b>) of S-20 composite’s worn-out debris.</p>
Full article ">Figure 17
<p>SEM EDX and map of S-20 composite’s worn-out debris.</p>
Full article ">Figure 18
<p>Mapping analysis of S-20 composite’s wear debris.</p>
Full article ">Figure 19
<p>Mapping analysis of S-20 composite’s wear debris at another debris location.</p>
Full article ">Figure 20
<p>SEM micrographs (<b>a</b>–<b>d</b>) of S-9 composite’s wear debris.</p>
Full article ">Figure 21
<p>EDX and elemental mapping of S-9 composite’s wear debris.</p>
Full article ">Figure A1
<p>AA7075/SiC/FA composites produced using FSP.</p>
Full article ">
40 pages, 49163 KiB  
Article
Investigations on Microstructure, Mechanical, and Wear Properties, with Strengthening Mechanisms of Al6061-CuO Composites
by Subrahmanya Ranga Viswanath Mantha, Gonal Basavaraja Veeresh Kumar, Ramakrishna Pramod and Chilakalapalli Surya Prakasha Rao
J. Manuf. Mater. Process. 2024, 8(6), 245; https://doi.org/10.3390/jmmp8060245 - 5 Nov 2024
Viewed by 528
Abstract
Metal matrix composites (MMCs) reinforced with Copper Oxide (CuO) and Aluminum (Al) 6061 (Al6061) alloys are being studied to determine their mechanical, physical, and dry sliding wear properties. The liquid metallurgical stir casting method with ultrasonication was employed for fabricating Al6061-CuO microparticle-reinforced composite [...] Read more.
Metal matrix composites (MMCs) reinforced with Copper Oxide (CuO) and Aluminum (Al) 6061 (Al6061) alloys are being studied to determine their mechanical, physical, and dry sliding wear properties. The liquid metallurgical stir casting method with ultrasonication was employed for fabricating Al6061-CuO microparticle-reinforced composite specimens by incorporating 2–6 weight percent (wt.%) CuO particles into the matrix. Physical, mechanical, and dry sliding wear properties were investigated in Al6061-CuO MMCs, adopting ASTM standards. The experimental results show that adding CuO to an Al6061 alloy increases its density by 7.54%, hardness by 45.78%, and tensile strength by 35.02%, reducing percentage elongation by 40.03%. Dry wear measurements on a pin-on-disc apparatus show that Al6061-CuO MMCs outperform the Al6061 alloy in wear resistance. Al6061-CuO MMCs’ strength has been predicted using many strengthening mechanism models and its elastic modulus through several models. The strengthening of Al6061-CuO MMCs is predominantly influenced by thermal mismatch, more so than by Hall–Petch, Orowan strengthening, and load transfer mechanisms. As the CuO content in the composite increases, the strengthening effects due to dislocation interactions between the matrix and reinforcement particles, the coefficient of thermal expansion (CTE) difference, grain refinement, and load transfer consistently improve. The Al6061-CuO MMCs were also examined using an optical microscope (OM), energy-dispersive spectroscopy (EDS), X-ray diffraction (XRD), and scanning electron microscopy (SEM) before and after fracture and wear tests. The investigation shows that an Al6061-CuO composite material with increased CuO reinforcement showed higher mechanical and tribological characteristics. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Grain size distribution of CuO particles and (<b>b</b>) SEM micrograph of CuO particles. (The symbols represent the dimensions in terms of particle width and they are self explanatory).</p>
Full article ">Figure 2
<p>Schematic representation of fabrication process of Al–CuO MMCs.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) Energy-dispersive spectroscopy of Al6061 alloy and its composites. (<b>a</b>) EDX of Al 6061 base alloy. (<b>b</b>) EDX of Al 6061-2 wt.% CuO. (<b>c</b>) EDX of Al 6061-4 wt.% CuO. (<b>d</b>) EDX of Al 6061-6 wt.% CuO.</p>
Full article ">Figure 4
<p>XRD spectra of Al6061 alloy and Al6061-CuO MMCs.</p>
Full article ">Figure 5
<p>Comparison of experimental and theoretical densities of Al6061-CuO MMCs.</p>
Full article ">Figure 6
<p>Variation in porosity percentage in Al-CuO MMCs.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>d</b>). Optical micrographs of (<b>a</b>) Al6061 alloy and (<b>b</b>) Al6061-2 wt.% CuO, (<b>c</b>) Al6061-4 wt.% CuO, and (<b>d</b>) Al6061-6 wt.% CuO.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>). SEM images of (<b>a</b>) Al6061 alloy and (<b>b</b>) Al6061–2 wt.% CuO, (<b>c</b>) Al6061–4 wt.% CuO, and (<b>d</b>) Al6061–6 wt.% CuO MMCs.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>–<b>d</b>). SEM images of (<b>a</b>) Al6061 alloy and (<b>b</b>) Al6061–2 wt.% CuO, (<b>c</b>) Al6061–4 wt.% CuO, and (<b>d</b>) Al6061–6 wt.% CuO MMCs.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>d</b>). Grain size distribution of (<b>a</b>) Al6061 alloy and (<b>b</b>) Al6061-2 wt.% CuO, (<b>c</b>) Al6061-4 wt.% CuO, and (<b>d</b>) Al6061-6 wt.% CuO MMCs.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>–<b>d</b>). Grain size distribution of (<b>a</b>) Al6061 alloy and (<b>b</b>) Al6061-2 wt.% CuO, (<b>c</b>) Al6061-4 wt.% CuO, and (<b>d</b>) Al6061-6 wt.% CuO MMCs.</p>
Full article ">Figure 10
<p>The hardness of the Al6061 alloy and Al6061-CuO MMCs.</p>
Full article ">Figure 11
<p>UTS, yield strength, and % elongation of Al6061 alloy and its CuO MMCs.</p>
Full article ">Figure 12
<p>Engineering stress–strain diagram of Al6061-Cuo composites.</p>
Full article ">Figure 13
<p>(<b>a</b>) SEM images of Al6061 alloy’s and (<b>b</b>) Al6061-6% CuO MMC’s fractured surfaces.</p>
Full article ">Figure 14
<p>(<b>a</b>–<b>c</b>). Predicted elastic moduli of Al6061-CuO MMCs. (<b>a</b>). Elastic modulus considering porosity. (<b>b</b>). Elastic modulus without considering porosity. (<b>c</b>). Normalized elastic modulus predictions.</p>
Full article ">Figure 14 Cont.
<p>(<b>a</b>–<b>c</b>). Predicted elastic moduli of Al6061-CuO MMCs. (<b>a</b>). Elastic modulus considering porosity. (<b>b</b>). Elastic modulus without considering porosity. (<b>c</b>). Normalized elastic modulus predictions.</p>
Full article ">Figure 15
<p>Variation in dislocation density with the decrease in the CTE in Al6061-CuO MMCs.</p>
Full article ">Figure 16
<p>Contribution of strengthening mechanisms to the Al-Cuo MMCs’ yield strength.</p>
Full article ">Figure 17
<p>Volumetric wear loss and wear mechanism of Al6061-CuO composites. (<b>a</b>). Volumetric wear loss for Al6061 alloy. (<b>b</b>). Volumetric wear loss for Al6061+2% CuO. (<b>c</b>). Volumetric wear loss for Al6061+4% CuO. (<b>d</b>). Volumetric wear loss for Al6061+6% CuO.</p>
Full article ">Figure 18
<p>Comparison of volumetric wear loss in Al6061-CuO MMCs.</p>
Full article ">Figure 19
<p>Sliding wear depth profile.</p>
Full article ">Figure 20
<p>Wear mechanism map.</p>
Full article ">Figure 21
<p>Worn surface morphology tested at 60 N load. (<b>a</b>) Alloy Al6061, (<b>b</b>) Al6061-2 wt.% CuO, (<b>c</b>) Al6061-4 wt.% CuO, and (<b>d</b>) Al6061-6 wt.% CuO.</p>
Full article ">Figure 22
<p>ANN model diagram for predicting volumetric wear loss.</p>
Full article ">Figure 23
<p>Performance, training, and error in the ANN model for volumetric wear loss prediction. (<b>a</b>) Performance Plot. (<b>b</b>) Training State Plot. (<b>c</b>) Error Histogram.</p>
Full article ">Figure 23 Cont.
<p>Performance, training, and error in the ANN model for volumetric wear loss prediction. (<b>a</b>) Performance Plot. (<b>b</b>) Training State Plot. (<b>c</b>) Error Histogram.</p>
Full article ">Figure 24
<p>Regression plot of the trained model for volumetric wear loss prediction.</p>
Full article ">Figure 25
<p>Weight values between the input and the hidden layers for volumetric wear loss.</p>
Full article ">Figure 26
<p>Weight values between the hidden and the output layer for wear.</p>
Full article ">Figure 27
<p>Standard workflow of a tree regression model.</p>
Full article ">Figure 28
<p>Tree regression prediction for volumetric wear loss. (<b>a</b>) Mean Square Error Plot. (<b>b</b>) Residuals Plot. (<b>c</b>) Validation Plot. (<b>d</b>) Response Plot.</p>
Full article ">Figure 28 Cont.
<p>Tree regression prediction for volumetric wear loss. (<b>a</b>) Mean Square Error Plot. (<b>b</b>) Residuals Plot. (<b>c</b>) Validation Plot. (<b>d</b>) Response Plot.</p>
Full article ">
15 pages, 16955 KiB  
Article
Formation and Mechano-Chemical Properties of Chromium Fluorides Originated from the Deposition of Carbon-Chromium Nanocomposite Coatings in the Reactive Atmosphere (Ar + CF4) during Magnetron Sputtering
by Adam Roślak, Józef Doering, Wioletta Strzałka, Marcin Makówka, Anna Jędrzejczak, Łukasz Kołodziejczyk, Jacek Balcerzak, Łukasz Jóźwiak, Ireneusz Piwoński and Wojciech Pawlak
Materials 2024, 17(20), 5034; https://doi.org/10.3390/ma17205034 - 15 Oct 2024
Viewed by 877
Abstract
The literature analysis did not indicate any studies on fluorination tests of carbon nanocomposite coatings doped with transition metals in a form of nanocrystalline metal carbide in amorphous carbon matrix (nc-MeC/a-C). As a model coating to investigate the effect of fluorination in a [...] Read more.
The literature analysis did not indicate any studies on fluorination tests of carbon nanocomposite coatings doped with transition metals in a form of nanocrystalline metal carbide in amorphous carbon matrix (nc-MeC/a-C). As a model coating to investigate the effect of fluorination in a tetrafluoromethane (CF4) atmosphere, a nanocomposite carbon coating doped with chromium-forming nanocrystals of chromium carbides in a-C matrix (nc-CrC/a-C) produced by magnetron sputtering from graphite targets and using a Pulse-DC type medium frequency power supply was chosen. After the deposition of the gradient chromium carbonitride (CrCN) adhesive sublayer, the fluorination of the main coating was conducted in a reactive mode in an (Ar + CF4) atmosphere at various CF4 content. It was observed that the presence of CF4 in the atmosphere resulted in a reduced amount of chromium carbides formed in favor of chromium fluorides. Thus far, this is an observation that seems unnoticed by the carbon coatings researchers. Fluorine was assumed to bond much more readily to carbon than to chromium, due to the stability of tetrafluoromethane (CF4). The opposite seems to be true. The mechanical properties (nano-hardness and Young’s modulus) and tribological properties in the ‘pin-on-disc’ friction pair are presented, along with the analysis of bonds occurring between chromium, carbon, and fluorine by means of X-ray photoelectron spectroscopy (XPS). Full article
(This article belongs to the Section Carbon Materials)
Show Figures

Figure 1

Figure 1
<p>Deposition parameters (magnetron power and gas flow) with process time for fluorinated nanocomposite carbon coatings. The dotted lines refer to the shorter, S4 process.</p>
Full article ">Figure 2
<p>Carbon (■) and fluorine (●) concentration as well as C/F ratio (<span style="color:red">▲</span>) of deposited coatings for different CF<sub>4</sub> partial pressures. The lines are just guides for the eyes.</p>
Full article ">Figure 3
<p>Fracture cross-sectional SEM image of S1 coating with 6.4 at.% of fluorine.</p>
Full article ">Figure 4
<p>EDS line scan of the cross-section of S2 coating with low fluorine content (7.4 at.% F). The inset shows lines for light elements (C, F, and N).</p>
Full article ">Figure 5
<p>EDS line scan of the cross-section of S4 coating with high fluorine content (19.7 at.% F). The inset shows lines for light elements (C, F, and N).</p>
Full article ">Figure 6
<p>XPS C 1s spectra of the coatings with different fluorine content. Please observe the lack of C-F<sub>x</sub> lines.</p>
Full article ">Figure 7
<p>XPS Cr 2p spectra of the coatings with different fluorine content.</p>
Full article ">Figure 8
<p>XPS F 1s spectra of the coatings with different fluorine content. Observe the lack of an F-C line.</p>
Full article ">Figure 9
<p>Carbon bond concentration based on XPS C 1s spectrum fitting.</p>
Full article ">Figure 10
<p>Chromium bonds concentration based on XPS Cr 2p spectrum fitting.</p>
Full article ">Figure 11
<p>Friction coefficients for investigated carbon-based nanocomposite nc-CrC/a-C:F coatings with different amounts of fluorine.</p>
Full article ">Figure 12
<p>S0 (<b>a</b>) and S1 (<b>b</b>) profiles of wear path from optical profilometry. Note the Z-axis scale change.</p>
Full article ">Figure 13
<p>Hardness (<b>a</b>) and elastic modulus (<b>b</b>) curves from continuous stiffness mode of nanoindentation.</p>
Full article ">
16 pages, 8291 KiB  
Article
Mechanical Properties and Tribological Study of Bottom Pouring Stir-Cast A356 Alloy Reinforced with Graphite Solid Lubricant Extracted from Corn Stover
by Vavilada Satya Swamy Venkatesh and Pandu Ranga Vundavilli
Lubricants 2024, 12(10), 341; https://doi.org/10.3390/lubricants12100341 - 2 Oct 2024
Viewed by 651
Abstract
The present work epitomises extracting the graphite (Gr) solid lubricant from the corn stover. The extracted Gr was incorporated as reinforcement in the A356 alloy (Al-7Si), and the effect of the Gr particles on the mechanical and tribological properties was investigated. In spite [...] Read more.
The present work epitomises extracting the graphite (Gr) solid lubricant from the corn stover. The extracted Gr was incorporated as reinforcement in the A356 alloy (Al-7Si), and the effect of the Gr particles on the mechanical and tribological properties was investigated. In spite of this, the input process parameters for the dry sliding wear test at room temperature against the EN31 steel disc were optimised through ANOVA analysis. The fabricated A359—X wt% (X = 0, 2.5, 5, 7.5) composite through bottom pouring stir casting techniques was analysed microstructurally by using XRD and FESEM analysis. The micro Brinell hardness and tensile strength were investigated per ASTME10 and ASTME8M standards. A wear test was performed for the composite pins against the EN31 steel disc according to ASTM G99 specifications. The XRD analysis results depict the presence of carbon (C), aluminium (Al), and silicon (Si) in all the wt% of the Gr reinforcement. However, along with the elements, the Al2Mg peak was confirmed for the A356—7.5 wt% Gr composite and the corresponding cluster element was confirmed in FESEM analysis. The maximum micro Brinell hardness of 92 BHN and U.T.S of 123 MPa and % elongation of 7.11 was attained at 5 wt% Gr reinforcement due to uniform Gr dispersion in the A356 alloy. Based on the ANOVA analysis, the optimal process parameters were obtained at 20 N applied load, 1 m/s sliding velocity, and 1000 m sliding distance for the optimal wear rate of 0.0052386 g/km and 0.364 COF. Full article
(This article belongs to the Special Issue Tribology for Lightweighting)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>b</b>) FESEM and EDX spectra for Gr particles, and (<b>c</b>,<b>d</b>) FESEM and EDX for A356 alloy.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>,<b>b</b>) FESEM and EDX spectra for Gr particles, and (<b>c</b>,<b>d</b>) FESEM and EDX for A356 alloy.</p>
Full article ">Figure 2
<p>(<b>a</b>) Bottom pouring stir casting setup. (<b>b</b>) Fabricated tensile test specimens.</p>
Full article ">Figure 3
<p>Optical microscope images for the (<b>a</b>) A359—2.5 wt% Gr, (<b>b</b>) A359—5 wt% Gr, and (<b>c</b>) A359—7.5 wt% Gr composites.</p>
Full article ">Figure 4
<p>XRD pattern for the A356, C1, C2, and C3 composites.</p>
Full article ">Figure 5
<p>FESEM and EDS spectra for the (<b>a</b>,<b>b</b>) A356—2.5 wt% Gr and (<b>c</b>,<b>d</b>) A356-5 wt% Gr composites.</p>
Full article ">Figure 6
<p>FESEM and EDX for the clusters in the A356—7.5 wt% Gr composite.</p>
Full article ">Figure 7
<p>Hardness of the cast composites.</p>
Full article ">Figure 8
<p>U.T.S and % elongation of the cast composites.</p>
Full article ">Figure 9
<p>(<b>a</b>) Wear track for the tribology test and (<b>b</b>) fabricated composite pins.</p>
Full article ">Figure 10
<p>Variation in (<b>a</b>) COF and (<b>b</b>) wear rate for the cast composites.</p>
Full article ">Figure 11
<p>SEM micrographs for the worn surface at 1000 m sliding distance and 40 N applied load (<b>a</b>) sliding direction (<b>b</b>) Delamination wear (<b>c</b>,<b>d</b>) Fine grooves.</p>
Full article ">Figure 12
<p>Actual and predicted values for the (<b>a</b>) wear rate and (<b>b</b>) COF.</p>
Full article ">Figure 13
<p>Residue vs. run for (<b>a</b>) COF and (<b>b</b>) wear rate.</p>
Full article ">Figure 14
<p>Effect of input process parameters on COF.</p>
Full article ">Figure 14 Cont.
<p>Effect of input process parameters on COF.</p>
Full article ">Figure 15
<p>Effect of input process parameters on the wear rate.</p>
Full article ">
16 pages, 6581 KiB  
Article
Laser Cladding of a Ti–Zr–Mo–Ta–Nb–B Composite Coating on Ti60 Alloy to Improve Wear Resistance
by Kaijin Huang and Xianchao Han
Coatings 2024, 14(10), 1247; https://doi.org/10.3390/coatings14101247 - 30 Sep 2024
Viewed by 654
Abstract
To improve the wear resistance of the Ti60 alloy, laser cladding was used to obtain a composite coating containing a high-entropy (Ti0.2Zr0.2Mo0.2Ta0.2Nb0.2)B2 boride phase, with Ti, Zr, Mo, Ta, Nb, and B [...] Read more.
To improve the wear resistance of the Ti60 alloy, laser cladding was used to obtain a composite coating containing a high-entropy (Ti0.2Zr0.2Mo0.2Ta0.2Nb0.2)B2 boride phase, with Ti, Zr, Mo, Ta, Nb, and B powders as the raw materials. The microstructure and wear characteristics of the coating were studied using XRD, SEM, EDS, and the pin-on-disc friction wear technique. The results show that the coating mainly consists of six phases: (Ti0.2Zr0.2Mo0.2Ta0.2Nb0.2)B2, ZrB2, TiB, TiZr, Ti1.83 Zr0.17, and Ti0.67Zr0.67Nb0.67. The average microhardness of the coating was 1062.9 HV0.1 due to the occurrence of the high-entropy, high-hardness (Ti0.2Zr0.2Mo0.2Ta0.2Nb0.2)B2 boride phase, which was about 2.9 times that of the Ti60 alloy substrate. The coating significantly improved the wear resistance of the Ti60 alloy substrate, and the mass wear rate was about 1/11 that of the Ti60 alloy substrate. The main types of wear affecting the coating were abrasive, adhesive, and oxidation wear, while the main wear affecting the Ti60 alloy matrix was abrasive wear, accompanied by a small amount of adhesive and oxidation wear. Full article
(This article belongs to the Special Issue Laser-Assisted Coating Techniques and Surface Modifications)
Show Figures

Figure 1

Figure 1
<p>XRD pattern of the laser cladding coating.</p>
Full article ">Figure 2
<p>Morphology of a cross-section of the whole laser cladding coating. Where A represents the surface region of the coating, B represents the middle region of the coating, and C represents the interface region between the coating and substrate.</p>
Full article ">Figure 3
<p>Enlarged images of the morphologies at different positions in <a href="#coatings-14-01247-f002" class="html-fig">Figure 2</a> for a cross-section of the laser cladding coating: (<b>a</b>) surface region A, (<b>b</b>) middle region B, and (<b>c</b>) interface region C.</p>
Full article ">Figure 4
<p>Microhardness curve for a cross-section of the laser coating.</p>
Full article ">Figure 5
<p>Columnar comparison of mass wear rates between the coating and Ti60 matrix.</p>
Full article ">Figure 6
<p>SEM image of the Ti60 matrix after wear: (<b>a</b>) whole worn surface, (<b>b</b>) enlarged SEM image at zone A in (<b>a</b>), and (<b>c</b>) enlarged SEM image at zone B in (<b>a</b>).</p>
Full article ">Figure 7
<p>SEM image of coating after wear: (<b>a</b>) whole worn surface, (<b>b</b>) enlarged SEM image at zone A in (<b>a</b>), (<b>c</b>) enlarged SEM image at zone B in (<b>a</b>), and (<b>d</b>) enlarged SEM image at zone C in (<b>a</b>).</p>
Full article ">Figure 7 Cont.
<p>SEM image of coating after wear: (<b>a</b>) whole worn surface, (<b>b</b>) enlarged SEM image at zone A in (<b>a</b>), (<b>c</b>) enlarged SEM image at zone B in (<b>a</b>), and (<b>d</b>) enlarged SEM image at zone C in (<b>a</b>).</p>
Full article ">Figure 8
<p>Schematic diagram of the different phase formation processes in the laser cladding pool of the coating.</p>
Full article ">Figure 9
<p>Gibbs free energy, ΔG, curves of different reactions with temperature T.</p>
Full article ">
19 pages, 5640 KiB  
Article
Tribological Performance of Additive Manufactured PLA-Based Parts
by Moises Batista, Irene Del Sol, Álvaro Gómez-Parra and Juan Manuel Vazquez-Martinez
Polymers 2024, 16(17), 2529; https://doi.org/10.3390/polym16172529 - 6 Sep 2024
Viewed by 697
Abstract
Polymer additive manufacturing has advanced from prototyping to producing essential parts with improved precision and versatility. Despite challenges like surface finish and wear resistance, new materials and metallic reinforcements in polymers have expanded its applications, enabling stronger, more durable parts for demanding industries [...] Read more.
Polymer additive manufacturing has advanced from prototyping to producing essential parts with improved precision and versatility. Despite challenges like surface finish and wear resistance, new materials and metallic reinforcements in polymers have expanded its applications, enabling stronger, more durable parts for demanding industries like aerospace and structural engineering. This research investigates the tribological behaviour of FFF surfaces by integrating copper and aluminium reinforcement particles into a PLA (polylactic acid) matrix. Pin-on-disc tests were conducted to evaluate friction coefficients and wear rates. Statistical analysis was performed to study the correlation of the main process variables. The results confirmed that reinforced materials offer interesting characteristics despite their complex use, with the roughness of the fabricated parts increasing by more than 300%. This leads to an increase in the coefficient of friction, which is related to the variation in the material’s mechanical properties, as the hardness increases by more than 75% for materials reinforced with Al. Despite this, their performance is more stable, and the volume of material lost due to wear is reduced by half. These results highlight the potential of reinforced polymers to improve the performance and durability of components manufactured through additive processes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Pin-on-disc technique scheme.</p>
Full article ">Figure 2
<p>Manufactured specimen surfaces. (<b>a</b>) Using PLA filaments. (<b>b</b>) Using high-performance PLA filaments. (<b>c</b>) Using PLA filaments reinforced with 20% Al particles. (<b>d</b>) Using PLA filaments reinforced with 20% Cu particles.</p>
Full article ">Figure 3
<p>Microhardness of the materials studied.</p>
Full article ">Figure 4
<p>(<b>a</b>) Topological representations using FMV techniques. (<b>b</b>) Comparative surface quality of different test specimens obtained in Sa terms. (<b>c</b>) Comparative surface quality of different test specimens obtained in Sz terms. (<b>d</b>) Comparative surface quality of different test specimens obtained in Sdc terms.</p>
Full article ">Figure 5
<p>Evolution of the coefficient of friction for the different materials studied with a 10 mm of test radius. (<b>a</b>) Using PLA filaments. (<b>b</b>) Using high-performance PLA filaments. (<b>c</b>) Using PLA filaments reinforced with 20% Al particles. (<b>d</b>) Using PLA filaments reinforced with 20% Cu particles.</p>
Full article ">Figure 6
<p>Evolution of the coefficient of friction for the different materials studied with a 20 mm of test radius. (<b>a</b>) Using PLA filaments. (<b>b</b>) Using high-performance PLA filaments. (<b>c</b>) Using PLA filaments reinforced with 20% Al particles. (<b>d</b>) Using PLA filaments reinforced with 20% Cu particles.</p>
Full article ">Figure 7
<p>Evolution of average friction coefficient (<b>a</b>) and amplitude (<b>b</b>).</p>
Full article ">Figure 8
<p>Comparison of the average friction coefficient for each material and its influence on the pin.</p>
Full article ">Figure 9
<p>Comparison of the morphology of the detached material (debris) with a 15 mm of test radius. (<b>a</b>) Using PLA filaments. (<b>b</b>) Using high-performance PLA filaments. (<b>c</b>) Using PLA filaments reinforced with 20% Al particles. (<b>d</b>) Using PLA filaments reinforced with 20% Cu particles.</p>
Full article ">Figure 10
<p>Comparison of the morphology of the grooves with a 10 mm of test radius. (<b>a</b>) Using PLA filaments. (<b>b</b>) Using high-performance PLA filaments. (<b>c</b>) Using PLA filaments reinforced with 20% Al particles. (<b>d</b>) Using PLA filaments reinforced with 20% Cu particles.</p>
Full article ">Figure 11
<p>(<b>a</b>) Evolution of the width of the groove wear. (<b>b</b>) Evolution of the maximum depth of the groove wear.</p>
Full article ">Figure 12
<p>(<b>a</b>) Evolution of the volume of the groove wear. (<b>b</b>) Volume of the groove wear in relation to the function of the material.</p>
Full article ">Figure 13
<p>Correlation matrix of wear test data.</p>
Full article ">
16 pages, 3958 KiB  
Article
The Abrasive Effect of Moon and Mars Regolith Simulants on Stainless Steel Rotating Shaft and Polytetrafluoroethylene Sealing Material Pairs
by Gábor Kalácska, György Barkó, Hailemariam Shegawu, Ádám Kalácska, László Zsidai, Róbert Keresztes and Zoltán Károly
Materials 2024, 17(17), 4240; https://doi.org/10.3390/ma17174240 - 27 Aug 2024
Viewed by 668
Abstract
For space missions to either the Moon or Mars, protecting mechanical moving parts from the abrasive effects of prevailing surface dust is crucial. This paper compares the abrasive effects of two lunar and two Martian simulant regoliths using special pin-on-disc tests on a [...] Read more.
For space missions to either the Moon or Mars, protecting mechanical moving parts from the abrasive effects of prevailing surface dust is crucial. This paper compares the abrasive effects of two lunar and two Martian simulant regoliths using special pin-on-disc tests on a stainless steel/polytetrafluoroethylene (PTFE) sealing material pair. Due to the regolith particles entering the contact zone, a three-body abrasion mechanism took place. We found that friction coefficients stabilised between 0.2 and 0.4 for all simulants. Wear curves, surface roughness measurements, and microscopic images all suggest a significantly lower abrasion effect of the Martian regoliths than that of the lunar ones. It applies not only to steel surfaces but also to the PTFE pins. The dominant abrasive micro-mechanism of the disc surface is micro-ploughing in the case of all tests, while the transformation of the counterface is mixed. The surface of pin material is plastically transformed through micro-ploughing, while the material is removed through micro-cutting due to the slide over hard soil particles. Full article
Show Figures

Figure 1

Figure 1
<p>Theory and practice of modified pin-on-disc measurements: (<b>a</b>) schematic of the layout; (<b>b</b>) the retainer blades cover back the path with regolith.</p>
Full article ">Figure 2
<p>Coefficient of friction of frictional surfaces as a function of travelled distance for lunar (LHS-1, LMS-1) and Martian regoliths (MGS-1, JEZ-1).</p>
Full article ">Figure 3
<p>Wear behaviour of frictional surfaces as a function of travelled distance for lunar (LHS-1, LMS-1) and Martian regoliths (MGS-1, JEZ-1).</p>
Full article ">Figure 4
<p>Light microscopy images of steel discs after 30 min of abrasion test with (<b>a</b>) LHS-1, (<b>b</b>) LMS-1, (<b>c</b>) MGS-1, and (<b>d</b>) JEZ-1 regolith simulants: lunars are significantly covered and embedded while Martians are less damaged.</p>
Full article ">Figure 5
<p>SEM images of steel discs after 30 min of abrasion test with (<b>a</b>) LHS-1, (<b>b</b>) LMS-1, (<b>c</b>) MGS-1, and (<b>d</b>) JEZ-1 regolith simulants.</p>
Full article ">Figure 6
<p>A selected (<b>a</b>) SEM image and EDX elemental mapping for (<b>b</b>) Si, (<b>c</b>) Al, (<b>d</b>) Ca, (<b>e</b>) Mg, and (<b>f</b>) Na elements on the steel disc after 30 min of abrasion test with LHS-1 regolith simulant as abrasive media.</p>
Full article ">Figure 6 Cont.
<p>A selected (<b>a</b>) SEM image and EDX elemental mapping for (<b>b</b>) Si, (<b>c</b>) Al, (<b>d</b>) Ca, (<b>e</b>) Mg, and (<b>f</b>) Na elements on the steel disc after 30 min of abrasion test with LHS-1 regolith simulant as abrasive media.</p>
Full article ">Figure 7
<p>Different modes of abrasive wear on SEM images: cutting mode (<b>a</b>) steel pin on brass plate, wedge-forming mode (<b>b</b>) steel pin on stainless steel plate, ploughing mode (<b>c</b>) steel pin on brass plate [<a href="#B22-materials-17-04240" class="html-bibr">22</a>,<a href="#B23-materials-17-04240" class="html-bibr">23</a>].</p>
Full article ">Figure 8
<p>Extraction of groove data from an abrasion scratch test [<a href="#B29-materials-17-04240" class="html-bibr">29</a>].</p>
Full article ">Figure 9
<p>Photos (<b>a</b>,<b>c</b>) and SEM images (<b>b</b>,<b>d</b>) of PTFE surfaces after 30 min of abrasion test with LHS-1 (upper row) and JEZ-1 (lower row) regolith simulants.</p>
Full article ">Figure 9 Cont.
<p>Photos (<b>a</b>,<b>c</b>) and SEM images (<b>b</b>,<b>d</b>) of PTFE surfaces after 30 min of abrasion test with LHS-1 (upper row) and JEZ-1 (lower row) regolith simulants.</p>
Full article ">Figure A1
<p>Coefficients of friction of the steel–PTFE material pair in the presence of the LHS-1 lunar regolith simulant dust as a function of distance travelled during different test times.</p>
Full article ">Figure A2
<p>Coefficients of friction of the steel–PTFE material pair in the presence of the LMS-1 lunar regolith simulant dust as a function of distance travelled during different test times.</p>
Full article ">Figure A3
<p>Coefficients of friction of the steel–PTFE material pair in the presence of the MGS-1 Martian regolith simulant dust as a function of distance travelled during different test times.</p>
Full article ">Figure A4
<p>Coefficients of friction of the steel–PTFE material pair in the presence of the JEZ-1 Martian regolith simulant dust as a function of distance travelled during different test times.</p>
Full article ">Figure A5
<p>Effect of travelled distance on the wear behaviour of the LHS-1 lunar regolith during different test times.</p>
Full article ">Figure A6
<p>Effect of travelled distance on the wear behaviour of the LMS-1 lunar regolith during different test times.</p>
Full article ">Figure A7
<p>Effect of travelled distance on the wear behaviour of the MGS-1 Martian regolith during different test times.</p>
Full article ">Figure A8
<p>Effect of travelled distance on the wear behaviour of the JEZ-1 Martian regolith during different test times.</p>
Full article ">
18 pages, 14536 KiB  
Article
Structure and Corrosion Behavior of Multiphase Intermetallic ZrCu-Based Alloys
by Rafał Babilas, Katarzyna Młynarek-Żak, Aneta Kania, Akash A. Deshmukh, Tymon Warski and Łukasz Hawełek
Materials 2024, 17(17), 4182; https://doi.org/10.3390/ma17174182 - 23 Aug 2024
Viewed by 533
Abstract
Zirconium-based alloys are highly regarded by the research community for their exceptional corrosion resistance, thermal stability, and mechanical properties. In our work, we investigated two newly developed alloys, Zr42.42Cu41.18Al9.35Ag7.05 and Zr46.81Cu35.44Al10.09 [...] Read more.
Zirconium-based alloys are highly regarded by the research community for their exceptional corrosion resistance, thermal stability, and mechanical properties. In our work, we investigated two newly developed alloys, Zr42.42Cu41.18Al9.35Ag7.05 and Zr46.81Cu35.44Al10.09Ag7.66, in the form of ingots and ribbons. In the course of our investigation, we conducted a comprehensive structural and thermal analysis. In addition, an examination of the corrosion activity encompassing electrochemical studies and an analysis of the corrosion mechanisms was carried out. To further evaluate the performance of the materials, tests of their mechanical properties were performed, including microhardness and resistance to abrasive wear. Structural analysis showed that both alloys studied had a multiphase, crystalline structure with intermetallic phases. The samples in the form of ribbons showed improved corrosion resistance compared to that of the ingots. The ingot containing a higher content of copper Zr42.42Cu41.18Al9.35Ag7.05 was characterized by better corrosion resistance, while showing lower average hardness and a higher degree of abrasive wear based on SEM observations after pin-on-disc tests. Full article
(This article belongs to the Special Issue Structure and Properties of Crystalline and Amorphous Alloys-Part II)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction patterns of ingots of the Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> (<b>a</b>) and Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> (<b>b</b>) alloys.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of ribbons of the Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> (<b>a</b>) and Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> (<b>b</b>) alloys.</p>
Full article ">Figure 3
<p>SEM images of (<b>a</b>) Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub>, and (<b>b</b>) Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> alloys.</p>
Full article ">Figure 4
<p>SEM image of as-cast Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> alloy with EDX maps.</p>
Full article ">Figure 5
<p>SEM image of as-cast Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> alloy with EDX maps.</p>
Full article ">Figure 6
<p>DTA curves of Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> and Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> ingots recorded after heating (<b>a</b>) and cooling (<b>b</b>).</p>
Full article ">Figure 7
<p>Changes in the open-circuit potential with time (<b>a</b>), and polarization curves (<b>b</b>) in Ringer’s solution at 37 °C for samples in the form of ingots and ribbons of Zr-based alloys.</p>
Full article ">Figure 8
<p>Surface morphology of (<b>a</b>,<b>b</b>) Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> and (<b>c</b>,<b>d</b>) Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> ingots after electrochemical tests in Ringer’s solution at 37 °C.</p>
Full article ">Figure 9
<p>Microhardness of Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> and Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> ingots.</p>
Full article ">Figure 10
<p>Friction of coefficient in a function of time curves recorded for studied Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> and Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> ingots during pin-on-disc tests.</p>
Full article ">Figure 11
<p>Wear track after the pin-on-disc test for the Zr<sub>42.42</sub>Cu<sub>41.18</sub>Al<sub>9.35</sub>Ag<sub>7.05</sub> alloy in the SE mode with marked wear mechanisms (<b>a</b>,<b>b</b>) and in the BSD mode with marked points for EDX analysis (<b>c</b>).</p>
Full article ">Figure 12
<p>Wear track after the pin-on-disc test for the Zr<sub>46.81</sub>Cu<sub>35.44</sub>Al<sub>10.09</sub>Ag<sub>7.66</sub> alloy in the SE mode with marked wear mechanisms (<b>a</b>,<b>b</b>) and in the BSD mode (<b>c</b>).</p>
Full article ">
16 pages, 3828 KiB  
Article
Effects of Nitriding and Thermal Processing on Wear and Corrosion Resistance of Vanadis 8 Steel
by Alejandro González-Pociño, Florentino Alvarez-Antolin and Luis Borja Peral-Martinez
Coatings 2024, 14(8), 1066; https://doi.org/10.3390/coatings14081066 - 20 Aug 2024
Viewed by 725
Abstract
Vanadis 8 steel is a tool steel manufactured by powder metallurgic processing. Its main alloy elements are V, Cr and Mo. By implementing an experimental design with five factors—all of them are related to the thermal processing of this steel and with ionic [...] Read more.
Vanadis 8 steel is a tool steel manufactured by powder metallurgic processing. Its main alloy elements are V, Cr and Mo. By implementing an experimental design with five factors—all of them are related to the thermal processing of this steel and with ionic nitriding—the effects of said factors on adhesive wear resistance and corrosion resistance were studied. For this purpose, Pin-on-Disc wear tests and lineal polarization resistance tests were carried out using an aqueous solution with 3.5% NaCl by weight. The main aim was to increase this steel use in more aggressive environmental conditions, such as in coastal environments. By means of XRD, the percentage of retained austenite was determined, and by SEM-EDX, the microstructure was revealed. The conclusion is that adhesive wear resistance is improved if thermal processing parameters are at such levels that increase austenite destabilization and reduce retained austenite content. This means to destabilize austenite at 1180 °C during 1 h, with oil quenching, tempering at 520 °C during 2 h and ionic nitriding at 520 °C during 2 h. Corrosion resistance is highly improved with ionic nitriding. At the same time, to compensate for the negative effect on corrosion resistance of a high density of primary and secondary carbides, it is essential to carry out the ionic nitriding treatment. The harmful effect of electrochemical microcells that appear in the carbide/matrix interface is compensated by the passivating effect generated by the nitrided surface. Full article
(This article belongs to the Special Issue Heat Treatment and Surface Engineering of Tools and Dies)
Show Figures

Figure 1

Figure 1
<p>Microstructure of Vanadis 8 steel after thermal treatments. (<b>a</b>) Experiment 4 (non-nitrided sample); (<b>b</b>) experiment 7 (non-nitrided sample); (<b>c</b>) thickness of nitrided layer in experiment 6; (<b>d</b>) nitrided layer in experiment 8; (<b>e</b>) nitrided layer in experiment 1.</p>
Full article ">Figure 2
<p>Graphic representation of effects on hardness: (<b>a</b>) representation on Pareto chart; (<b>b</b>) representation on normal probability plot.</p>
Full article ">Figure 3
<p>Graphical representation of the effects on retained austenite: (<b>a</b>) representation on Pareto chart; (<b>b</b>) representation on normal probability plot.</p>
Full article ">Figure 4
<p>Graphic representation of the effects on the adhesive wear resistance: (<b>a</b>) representation on the Pareto chart; (<b>b</b>) representation on the normal probability plot. In the latter, factors with significant effects are underlined.</p>
Full article ">Figure 5
<p>Polarization curves. Curves with a higher percentage of corrosion correspond to the nitrided samples.</p>
Full article ">Figure 6
<p>Graphical representation of the effects on corrosion potential (Ecorr), polarization resistance (Rp) and corrosion intensity (Icorr). (<b>a</b>,<b>c</b>,<b>e</b>) show the representation on Pareto charts; (<b>b</b>,<b>d</b>,<b>f</b>) show the representation on normal probability plots. In the last diagram, factors with significant effects are underlined.</p>
Full article ">Figure 7
<p>Surface after 4 h of being in touch with electrolyte. (<b>a</b>) Experiment 7 (non-nitrided), ×10,000; (<b>b</b>) experiment 8 (nitrided), ×10,000.</p>
Full article ">
22 pages, 10423 KiB  
Article
Process-Integrated Component Microtexturing for Tribologically Optimized Contacts Using the Example of the Cam Tappet—Numerical Design, Manufacturing, DLC-Coating and Experimental Analysis
by Christian Orgeldinger, Manuel Reck, Armin Seynstahl, Tobias Rosnitschek, Marion Merklein and Stephan Tremmel
Lubricants 2024, 12(8), 291; https://doi.org/10.3390/lubricants12080291 - 16 Aug 2024
Viewed by 774
Abstract
To meet the demand for energy-efficient and, at the same time, durable, functional components, the improvement of tribological behavior is playing an increasingly important role. One approach to reducing friction in lubricated tribological systems is the microtexturing of the surfaces tailored to the [...] Read more.
To meet the demand for energy-efficient and, at the same time, durable, functional components, the improvement of tribological behavior is playing an increasingly important role. One approach to reducing friction in lubricated tribological systems is the microtexturing of the surfaces tailored to the application, but in most cases, this leads to increased manufacturing costs and thus often makes their use in industry more difficult. In this work, we, therefore, present an approach for an efficient design and fully integrated production process using a cam tappet as an example. For the used cam tappet contact, we first determined the optimal texture geometries using two differently complex EHL (elastohydrodynamic lubrication) simulation models. Based on these, textured tappets were manufactured in a combined manner using sheet-bulk metal-forming and deposition with a diamond-like-carbon (DLC) coating for additional wear protection without further post-processing of the coating. We show that the simulation approach used has a rather subordinate influence on the optimization result. The combined forming of components with textured surfaces is limited by the local material flow, the resulting texture distortion, and tool wear. However, a targeted process design can help to exploit the potential of single-stage forming. The applied DLC coating has good adhesion and can completely prevent wear in subsequent reciprocal pin-on-disc tests, while the friction in the run-in behavior is initially higher due to the soothing effects of the coating. The experiments also show a tendency for shallow textures to exhibit lower friction compared to deeper ones, which corresponds to the expectations from the simulation. Full article
(This article belongs to the Special Issue Tribology in Germany: Latest Research and Development)
Show Figures

Figure 1

Figure 1
<p>Tool setup of the combined forming process for the manufacturing of microtextured cam tappets.</p>
Full article ">Figure 2
<p>Process kinematics for the single-stage forming.</p>
Full article ">Figure 3
<p>Used pin-on-disc tribometer (<b>a</b>) with linear reciprocal kinematics with excentric clamping (<b>b</b>).</p>
Full article ">Figure 4
<p>Resulting relative mean frictional force as a function of the selected design parameters for all test points (<span class="html-italic">n</span> = 40) and as a function of the simulation approach (2D and 3D) for the cam tip contact at 500 rpm.</p>
Full article ">Figure 5
<p>Average friction forces predicted by the GPR models for load cases (<b>a</b>–<b>f</b>).</p>
Full article ">Figure 6
<p>Distortion of the textures depending on the positions shown in the SEM. The rectangle texture corresponds to the texture used for the manufacturing tests.</p>
Full article ">Figure 7
<p>Texture height with respect to the position on the punch (<b>a</b>) and the corresponding signs of wear (<b>b</b>).</p>
Full article ">Figure 8
<p>Light microscopic and laser scanning microscopic images of the texture patterns applied to the tappets: global textures (<b>a</b>,<b>b</b>), distribution of the local texture patterns (<b>c</b>) and their characteristics (<b>d</b>–<b>g</b>). The different height scales must be considered.</p>
Full article ">Figure 9
<p>Coating design in the FIB cut (<b>a</b>) and the corresponding EDS mapping (<b>b</b>).</p>
Full article ">Figure 10
<p>Coating in the textures T_g (<b>a</b>–<b>c</b>) and T_l2 (<b>d</b>–<b>f</b>) as well as exemplary enlargement in the area of the texture edges (<b>g</b>).</p>
Full article ">Figure 11
<p>Mean friction values (<b>a</b>) and diameter of the wear calotte of the counter body (<b>b</b>) of the tested variants (<span class="html-italic">n</span> = 3 uncoated and <span class="html-italic">n</span> = 2 with DLC) with selected significance levels (<span class="html-italic">t</span>-test with <span class="html-italic">p</span> &lt; 0.05 (*) and <span class="html-italic">p</span> &lt; 0.01 (**)). Please note the axes cut off for reasons of presentation.</p>
Full article ">Figure 12
<p>Wear of the coated (<b>a</b>) and uncoated (<b>b</b>) textures (two examples) with the corresponding LSM images and the friction curves (<b>c</b>) in the long-term tests with texture T_l2. Note the axis cut off for display reasons. In addition, the signs of wear on the unfavorable texture patterns T_g (<b>d</b>) and R_l (<b>e</b>), which occurred sporadically even after the short test running times.</p>
Full article ">
16 pages, 6842 KiB  
Article
Tribological Analysis of Fused Filament Fabrication PETG Parts Coated with IGUS
by Moises Batista, Delia Tenorio, Irene Del Sol and Juan Manuel Vazquez-Martinez
Appl. Sci. 2024, 14(16), 7161; https://doi.org/10.3390/app14167161 - 15 Aug 2024
Viewed by 635
Abstract
This paper studied the tribological behaviour of parts manufactured using fused filament fabrication (FFF) technology with PETG (polyethylene terephthalate glycol) coated with IGUS tribological filaments. The research focuses on analysing how these multi-material parts behave under different loads. The objective of this study [...] Read more.
This paper studied the tribological behaviour of parts manufactured using fused filament fabrication (FFF) technology with PETG (polyethylene terephthalate glycol) coated with IGUS tribological filaments. The research focuses on analysing how these multi-material parts behave under different loads. The objective of this study is to evaluate the wear resistance and friction coefficient of parts coated with different thicknesses of IGUS material. The methodology employs pin-on-disc (PoD) tribological tests to measure behaviour under various load conditions and coating thicknesses. The results indicate that increasing the coating thickness improves surface stability and reduces roughness, although it does not significantly affect the average friction coefficient. This research concludes that coating thickness has a moderate impact on surface quality and that the applied load significantly influences the depth and width of the wear groove. This contribution is valuable for the field of additive manufacturing as it provides a better understanding of how to optimise the tribological properties of parts manufactured using FFF, which is crucial for industrial applications where wear and friction are critical factors. The practical application includes the potential improvement of components in the automotive and aerospace industries. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of specimens and manufacturing strategies.</p>
Full article ">Figure 2
<p>Images of PETG specimens coated with IGUS manufactured with FFF: (<b>a</b>) with 0.3 mm coating thickness; (<b>b</b>) with 0.6 mm coating thickness; (<b>c</b>) with 0.9 mm coating thickness.</p>
Full article ">Figure 3
<p>Evolution of surface quality in terms of (<b>a</b>) Sa, (<b>b</b>) Sz, and (<b>c</b>) Sdc.</p>
Full article ">Figure 4
<p>Evolution of tests conducted on specimens coated with 0.3 mm coating.</p>
Full article ">Figure 5
<p>Evolution of tests conducted on specimens coated with 0.6 mm coating.</p>
Full article ">Figure 6
<p>Evolution of tests conducted on specimens coated with 0.9 mm coating.</p>
Full article ">Figure 7
<p>Evolution of the average coefficient of friction.</p>
Full article ">Figure 8
<p>Particles detached during the tribological tests.</p>
Full article ">Figure 9
<p>Evolution of the wear groove width.</p>
Full article ">Figure 10
<p>Evolution of the wear groove depth.</p>
Full article ">
17 pages, 6345 KiB  
Article
Enhancing the Tribological Properties of Bearing Surfaces in Hip Arthroplasty by Shot-Peening the Metal Surface
by Chavarat Jarungvittayakon, Anak Khantachawana and Paphon Sa-ngasoongsong
Lubricants 2024, 12(8), 278; https://doi.org/10.3390/lubricants12080278 - 3 Aug 2024
Viewed by 714
Abstract
Total hip arthroplasty (THA) is a surgical procedure for patients with pain and difficulty walking due to hip osteoarthritis. In primary THA, the acetabulum and femoral head are replaced by a prosthesis where the modular femoral head and inner liner of the acetabulum [...] Read more.
Total hip arthroplasty (THA) is a surgical procedure for patients with pain and difficulty walking due to hip osteoarthritis. In primary THA, the acetabulum and femoral head are replaced by a prosthesis where the modular femoral head and inner liner of the acetabulum form the bearing surface. The most popular bearing surface used in the United States, metal-on-polyethylene, consists of a cobalt–chromium molybdenum (CoCrMo) alloy femoral head that articulates with a polyethylene acetabular liner, typically made of highly cross-linked polyethylene. While successful in most cases, THA sometimes fails, commonly from aseptic loosening due to the wear debris of polyethylene. Fine-particle shot peening (FPSP) is a simple method for enhancing the mechanical properties and surface properties of metal, including reducing friction and enhancing the lubrication properties of the metal surface. In this study, we applied FPSP to the CoCr in the femoral head of a hip prosthesis to improve its surface properties and conducted experiments with pin-on-disc tribometers using CoCr as a pin and highly cross-linked polyethylene as a disc to mimic the THA implant. The results show that FPSP significantly enhances the tribological properties of the CoCr surface, including lubrication; decreases the friction coefficient; and decreases the polyethylene wear volume. Full article
(This article belongs to the Special Issue Biomechanics and Tribology)
Show Figures

Figure 1

Figure 1
<p>Workpiece for evaluation in the tribology test, including the friction coefficient and polyethylene wear by tribometer (pin-on-disc type).</p>
Full article ">Figure 2
<p>Workpiece for evaluating the wettability of the surface.</p>
Full article ">Figure 3
<p>The process of measuring the contact angle on the surface material.</p>
Full article ">Figure 4
<p>Tribology test (pin-on-disc type) using CoCr as a pin and polyethylene as a disc.</p>
Full article ">Figure 5
<p>Chamber used for containing the lubricant between the motion surfaces.</p>
Full article ">Figure 6
<p>(<b>A</b>) The wear track on polyethylene after the tribological test. (<b>B</b>) The wear track was evaluated by a surface roughness tester. (<b>C</b>) The wear volume was evaluated by area under the surface material.</p>
Full article ">Figure 6 Cont.
<p>(<b>A</b>) The wear track on polyethylene after the tribological test. (<b>B</b>) The wear track was evaluated by a surface roughness tester. (<b>C</b>) The wear volume was evaluated by area under the surface material.</p>
Full article ">Figure 7
<p>The association of wettability and the contact angle.</p>
Full article ">Figure 8
<p>(<b>A</b>) The contact angle on the un-peened surface. (<b>B</b>) The contact angle on the surface peened with ceramic particles. (<b>C</b>) The contact angle on the surface peened with silica particles.</p>
Full article ">Figure 9
<p>Contact angles of un-peened surface vs. peened surfaces of various types (ceramic and silica) and particle sizes.</p>
Full article ">Figure 10
<p>The coefficient of friction was evaluated by a tribometer.</p>
Full article ">Figure 11
<p>The result of the average coefficient of friction compared between the un-peened surface and surfaces peened with various particle types.</p>
Full article ">Figure 12
<p>The result of wear volume compared between the un-peened surface and peened surfaces using various particle types.</p>
Full article ">
12 pages, 5511 KiB  
Article
Wear Performance Evaluation of Polymer Overlays on Engine Bearings
by Ismail Ozdemir, Bahattin Bulbul, Ugur Kiracbedel, Thomas Grund and Thomas Lampke
Materials 2024, 17(15), 3802; https://doi.org/10.3390/ma17153802 - 1 Aug 2024
Viewed by 733
Abstract
Modern engine bearing materials encounter the challenge of functioning under conditions of mixed lubrication, low viscosity oils, downsizing, start–stop engines, potentially leading to metal-to-metal contact and, subsequently, premature bearing failure. In this work, two types of polymer overlays were applied to the bearing [...] Read more.
Modern engine bearing materials encounter the challenge of functioning under conditions of mixed lubrication, low viscosity oils, downsizing, start–stop engines, potentially leading to metal-to-metal contact and, subsequently, premature bearing failure. In this work, two types of polymer overlays were applied to the bearing surface to compensate for extreme conditions, such as excessive loads and mixed lubrication. Two different polymer overlays, created through a curing process on a conventional engine bearing surface with an approximate thickness of 13 µm, were investigated for their friction and wear resistances under a 30 N load using a pin-on-disc setup. The results indicate that the newly developed polymer overlay (NDP, PAI-based coating) surface has a coefficient of friction (COF) of 0.155 and a wear volume loss of 0.010 cm3. In contrast, the currently used polymer overlay (CPO) in this field shows higher values with a COF of 0.378 and a wear volume loss of 0.024 cm3, which is significantly greater than that of the NDP. It was found that, in addition to accurately selecting the ratios of solid lubricants, polymer resins, and wear-resistant hard particle additives (metal powders, metal oxides, carbides, etc.) within the polymer coating, the effective presence of a transfer film providing low friction on the counter surface also played a crucial role. Full article
(This article belongs to the Special Issue Corrosion and Tribological Behaviour of Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the bearing structure of a conventional electroplated coated bearing and (<b>b</b>) the investigated type of polymer-coated bearing.</p>
Full article ">Figure 2
<p>Copper-Based Sintering Bimetal production system.</p>
Full article ">Figure 3
<p>SEM images of the polymer overlay microstructure on the bronze bearing (<b>a</b>) at low (500×) and (<b>b</b>) high (2500×) magnifications.</p>
Full article ">Figure 4
<p>Wear test setup for a polymer-coated bearing segment.</p>
Full article ">Figure 5
<p>Macroscopic view of the tested polymer overlays: (<b>a</b>) the reference material (commercial), and (<b>b</b>) the newly developed material sprayed onto the bronze bearing surface.</p>
Full article ">Figure 6
<p>The variation in the coefficient of friction for the reference coating (<b>a</b>) and the newly developed polymer overlay coating (<b>b</b>) under a 20 N applied wear load for a longer sliding distance.</p>
Full article ">Figure 7
<p>The variation in the coefficient of friction and ECR measurement (indicated by the red arrow) for the reference coating (<b>a</b>) and the NDP overlay (<b>b</b>) material tested under 30 N load.</p>
Full article ">Figure 8
<p>The macroscopic appearance of the worn surfaces of the reference polymer (<b>a</b>) and the NDP overlay (<b>b</b>) after the wear test under a 30 N load.</p>
Full article ">Figure 9
<p>Three-dimensional wear track images of the reference polymer (<b>a</b>) and the NDP overlay (<b>b</b>) after wear tests under 30 N also displaying the depth of worn surfaces at (<b>c</b>) and notable breaks in the side walls of the wear path (red arrows) (<b>d</b>), respectively.</p>
Full article ">Figure 10
<p>The variation in the COF values (<b>a</b>) and linear wear loss (<b>b</b>) tested under 30 N.</p>
Full article ">Figure 11
<p>The change in the COF as a function of wear distance when the reference polymer coating initially comes into contact with a ball under a 30 N load (1st zone), depicting the region where it functions in harmony with the ball (2nd zone) and the point at which it completely separates from the wear surface (3rd zone).</p>
Full article ">
17 pages, 12044 KiB  
Article
Study on the Tribological Properties of DIN 16MnCr5 Steel after Duplex Gas-Nitriding and Pack Boriding
by Rafael Carrera Espinoza, Melvyn Alvarez Vera, Marc Wettlaufer, Manuel Kerl, Stefan Barth, Pablo Moreno Garibaldi, Juan Carlos Díaz Guillen, Héctor Manuel Hernández García, Rita Muñoz Arroyo and Javier A. Ortega
Materials 2024, 17(13), 3057; https://doi.org/10.3390/ma17133057 - 21 Jun 2024
Viewed by 848
Abstract
DIN 16MnCr5 is commonly used in mechanical engineering contact applications such as gears, joint parts, shafts, gear wheels, camshafts, bolts, pins, and cardan joints, among others. This study examined the microstructural and mechanical properties and tribological behavior of different surface treatments applied to [...] Read more.
DIN 16MnCr5 is commonly used in mechanical engineering contact applications such as gears, joint parts, shafts, gear wheels, camshafts, bolts, pins, and cardan joints, among others. This study examined the microstructural and mechanical properties and tribological behavior of different surface treatments applied to DIN 16MnCr5 steel. The samples were hardened at 870 °C for 15 min and then quenched in water. The surface conditions evaluated were as follows: quenched and tempered DIN 16MnCr5 steel samples without surface treatments (control group), quenched and tempered DIN 16MnCr5 steel samples with gas-nitriding at 560 °C for 6 h, quenched and tempered DIN 16MnCr5 steel samples with pack boriding at 950 °C for 4 h, and quenched and tempered DIN 16MnCr5 steel samples with duplex gas-nitriding and pack boriding. Microstructure characterization was carried out using metallographic techniques, optical microscopy, scanning electron microscopy with energy-dispersive spectroscopy, and X-ray diffraction. The mechanical properties were assessed through microhardness and elastic modulus tests using nanoindentation. The tribological behavior was evaluated using pin-on-disc tests following the ASTM G99-17 standard procedure under dry sliding conditions. The results indicated that the surface treated with duplex gas-nitriding and pack boriding exhibited the highest wear resistance and a reduced coefficient of friction due to improved mechanical properties, leading to increased hardness and elastic modulus. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of duplex surface treatment of (<b>a</b>) gas-nitriding and (<b>b</b>) pack boriding thermochemical treatment.</p>
Full article ">Figure 2
<p>SEM micrographs with secondary electron detector and chemical composition mapping of layers for (<b>a</b>) gas-nitriding, (<b>b</b>) pack boriding, and (<b>c</b>) duplex nitriding and boriding.</p>
Full article ">Figure 3
<p>Diffraction XRD patterns for quenched and tempered DIN 16MnCr5 and surface treatments with nitriding and boriding.</p>
Full article ">Figure 4
<p>Microhardness profile as a function of a distance from the surface for quenched and tempered DIN 16MnCr5 and surface treatments with nitriding and boriding.</p>
Full article ">Figure 5
<p>Nanoindentation load–unload curves for quenched and tempered DIN 16MnCr5 and surface treatments with nitriding and boriding.</p>
Full article ">Figure 6
<p>SEM micrographs with secondary electron detector wear tracks for Q&amp;T DIN 16MnCr5 steel (<b>a</b>,<b>b</b>) with 6 N load, (<b>c</b>,<b>d</b>) with 4 N load, and (<b>e</b>,<b>f</b>) with 2 N load.</p>
Full article ">Figure 7
<p>SEM micrographs with secondary electron detector wear tracks for Q&amp;T+N DIN 16MnCr5 steel (<b>a</b>,<b>b</b>) with 6 N load, (<b>c</b>,<b>d</b>) with 4 N load, and (<b>e</b>,<b>f</b>) with 2 N load.</p>
Full article ">Figure 8
<p>SEM micrographs with secondary electron detector wear tracks for Q&amp;T+B DIN 16MnCr5 steel (<b>a</b>,<b>b</b>) with 6 N laod, (<b>c</b>,<b>d</b>) with 4 N load, and (<b>e</b>,<b>f</b>) with 2 N load.</p>
Full article ">Figure 9
<p>SEM micrographs with secondary electron detector wear tracks for Q&amp;T+N+B DIN 16MnCr5 steel (<b>a</b>,<b>b</b>) with 6 N load, (<b>c</b>,<b>d</b>) with 4 N load, and (<b>e</b>,<b>f</b>) with 2 N load.</p>
Full article ">Figure 10
<p>Worn surfaces of (<b>a</b>) Q&amp;T DIN 16MnCr5 steel and the (<b>b</b>) WC pin with a 6 N load.</p>
Full article ">Figure 11
<p>Chemical composition mapping of worn surface Q&amp;T DIN 16MnCr5 steel (<b>a</b>) and energy-dispersive spectroscopy EDS (<b>b</b>) with 6 N load.</p>
Full article ">Figure 12
<p>Worn surface of (<b>a</b>) Q&amp;T+N+B DIN 16MnCr5 steel and (<b>b</b>) WC pin with 6 N load.</p>
Full article ">Figure 13
<p>Chemical composition mapping of worn surface Q&amp;T+N+B DIN 16MnCr5 steel (<b>a</b>) and energy-dispersive spectroscopy (EDS) (<b>b</b>) with 6 N load.</p>
Full article ">Figure 14
<p>Cross-sectional profilometry of the worn surface for (<b>a</b>) Q&amp;T, (<b>b</b>) Q&amp;T+N, (<b>c</b>) Q&amp;+B, and (<b>d</b>) Q&amp;T+N+B.</p>
Full article ">Figure 15
<p>Volume loss results for Q&amp;T, Q&amp;T+N, Q&amp;T+B, and Q&amp;T+N+B.</p>
Full article ">Figure 16
<p>Coefficient of friction for quenched and tempered DIN 16MnCr5 and surface treatments with nitriding and boriding.</p>
Full article ">
Back to TopTop