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

Journals

Article Types

Countries / Regions

Search Results (10)

Search Parameters:
Keywords = wear rate (WR)

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 6414 KiB  
Article
Experimental Investigation and Machine Learning Modeling of Tribological Characteristics of AZ31/B4C/GNPs Hybrid Composites
by Dhanunjay Kumar Ammisetti, Bharat Kumar Chigilipalli, Baburao Gaddala, Ravi Kumar Kottala, Radhamanohar Aepuru, T. Srinivasa Rao, Seepana Praveenkumar and Ravinder Kumar
Crystals 2024, 14(12), 1007; https://doi.org/10.3390/cryst14121007 - 21 Nov 2024
Viewed by 350
Abstract
In this study, the AZ31 hybrid composites reinforced with boron carbide (B4C) and graphene nano-platelets (GNPs) are prepared by the stir casting method. The main aim of the study is to study the effect of various wear parameters (reinforcement percentage (R), [...] Read more.
In this study, the AZ31 hybrid composites reinforced with boron carbide (B4C) and graphene nano-platelets (GNPs) are prepared by the stir casting method. The main aim of the study is to study the effect of various wear parameters (reinforcement percentage (R), applied load (L), sliding distance (D), and velocity (V)) on the wear characteristics (wear rate (WR)) of the AZ91/B4C/GNP composites. Experiments are designed using the Taguchi technique, and it was determined that load (L) is the most significant parameter affecting WR, followed by D, R, and V. The wear mechanisms under conditions of maximum and minimum wear rates are examined using SEM analysis of the worn-out surfaces of the specimens. From the result analysis on the WR, the ideal conditions for achieving the lowest WR are R = 4 wt.%, L = 15 N, V = 3 m/s, and D = 500 m. Machine learning (ML) models, including linear regression (LR), polynomial regression (PR), random forest (RF), and Gaussian process regression (GPR), are implemented to develop a reliable prediction model that forecasts output responses in accordance with input variables. A total of 90% of the experimental data points were used to train and 10% to evaluate the models. The PR model exceeded the accuracy of other models in predicting WR, with R2 = 0.953, MSE = 0.011, RMSE = 0.103, and COF with R2 = 0.937, MSE = 0.013, and RMSE = 0.114, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart.</p>
Full article ">Figure 2
<p>SEM image of (<b>a</b>) GNPs and (<b>b</b>) B<sub>4</sub>C; EDS image of (<b>c</b>) GNPs and (<b>d</b>) B<sub>4</sub>C; XRD image of (<b>e</b>) GNPs and (<b>f</b>) B<sub>4</sub>C.</p>
Full article ">Figure 2 Cont.
<p>SEM image of (<b>a</b>) GNPs and (<b>b</b>) B<sub>4</sub>C; EDS image of (<b>c</b>) GNPs and (<b>d</b>) B<sub>4</sub>C; XRD image of (<b>e</b>) GNPs and (<b>f</b>) B<sub>4</sub>C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Wear testing machine. (<b>b</b>) Experimental setup.</p>
Full article ">Figure 4
<p>SEM microstructures of (<b>a</b>) AZ31 + 1 wt.% graphene + 1 wt.% B<sub>4</sub>C; (<b>b</b>) AZ31 + 1 wt.% graphene + 2 wt.% B<sub>4</sub>C; and (<b>c</b>) AZ31 + 1 wt.% graphene + 3 wt.% B<sub>4</sub>C.</p>
Full article ">Figure 5
<p>Effect of various factors on WR (means data).</p>
Full article ">Figure 6
<p>Effect of various factors on WR (S/N ratios data).</p>
Full article ">Figure 7
<p>Interaction plot for means.</p>
Full article ">Figure 8
<p>Residual plots for WR.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) High worn surfaces. (<b>c</b>,<b>d</b>) Low worn out surfaces.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>,<b>b</b>) High worn surfaces. (<b>c</b>,<b>d</b>) Low worn out surfaces.</p>
Full article ">Figure 10
<p>Regression plots for WR data with (<b>a</b>) LR, (<b>b</b>) PR, (<b>c</b>) RF, and (<b>d</b>) GPR. (<b>e</b>) Comparison plot for training and testing of LR, PR, RF, and GPR techniques.</p>
Full article ">Figure 11
<p>Regression plots for COF data with (<b>a</b>) LR, (<b>b</b>) PR, (<b>c</b>) RF, and (<b>d</b>) GPR. (<b>e</b>) Comparison plot for training and testing of LR, PR, RF, and GPR techniques.</p>
Full article ">
19 pages, 6088 KiB  
Article
Tribological Behaviour of Hypereutectic Al-Si Composites: A Multi-Response Optimisation Approach with ANN and Taguchi Grey Method
by Slavica Miladinović, Sandra Gajević, Slobodan Savić, Ivan Miletić, Blaža Stojanović and Aleksandar Vencl
Lubricants 2024, 12(2), 61; https://doi.org/10.3390/lubricants12020061 - 17 Feb 2024
Cited by 6 | Viewed by 1761
Abstract
An optimisation model for small datasets was applied to thixocasted/compocasted composites and hybrid composites with hypereutectic Al-18Si base alloys. Composites were produced with the addition of Al2O3 (36 µm/25 nm) or SiC (40 µm) particles. Based on the design of [...] Read more.
An optimisation model for small datasets was applied to thixocasted/compocasted composites and hybrid composites with hypereutectic Al-18Si base alloys. Composites were produced with the addition of Al2O3 (36 µm/25 nm) or SiC (40 µm) particles. Based on the design of experiment, tribological tests were performed on the tribometer with block-on-disc contact geometry for normal loads of 100 and 200 N, a sliding speed of 0.5 m/s, and a sliding distance of 1000 m. For the prediction of the tribological behaviour of composites, artificial neural networks (ANNs) were used. Three inputs were considered for ANN training: type of reinforcement (base alloy, Al2O3 and SiC), amount of Al2O3 nano-reinforcement (0 and 0.5 wt.%), and load (100 and 200 N). Various ANNs were applied, and the best ANN for wear rate (WR), with an overall regression coefficient of 0.99484, was a network with architecture 3-15-1 and a logsig (logarithmic sigmoid) transfer function. For coefficient of friction (CoF), the best ANN was the one with architecture 3-6-1 and a tansig (hyperbolic tangent sigmoid) transfer function and had an overall regression coefficient of 0.93096. To investigate the potential of ANN for the prediction of two outputs simultaneously, an ANN was trained, and the best results were from network 3-5-2 with a logsig transfer function and overall regression coefficient of 0.99776, but the predicted values for CoF in this case did not show good correlation with experimental results. After the selection of the best ANNs, the Taguchi grey multi-response optimisation of WR and CoF was performed for the same combination of factors as the ANNs. For optimal WR and CoF, the combination of factors was as follows: composite with 3 wt.% Al2O3 micro-reinforcement, 0.5 wt.% Al2O3 nano-reinforcement, and a load of 100 N. The results show that developed ANN, the Taguchi method, and the Taguchi grey method can, with high reliability, be used for the optimisation of wear rate and coefficient of friction of hypereutectic Al-Si composites. Microstructural investigations of worn surfaces were performed, and the wear mechanism for all tested materials was light abrasion and adhesion. The findings from this research can contribute to the future development of hypereutectic Al-Si composites. Full article
(This article belongs to the Special Issue Wear Behavior of Aluminum Matrix Composite)
Show Figures

Figure 1

Figure 1
<p>Microstructures of produced composites: (<b>a</b>) matrix material Al-18Si and (<b>b</b>) composite with 0.5 wt.% Al<sub>2</sub>O<sub>3</sub> (25 nm) and 3 wt.% Al<sub>2</sub>O<sub>3</sub> (36 µm).</p>
Full article ">Figure 2
<p>Schematics of contact pair.</p>
Full article ">Figure 3
<p>Schematic model of simplified ANN.</p>
Full article ">Figure 4
<p>Regression coefficients for the best ANN for (<b>a</b>) wear rate (WR), (<b>b</b>) coefficient of friction (CoF), and (<b>c</b>) both outputs.</p>
Full article ">Figure 5
<p>Diagrams of S/N ratio: (<b>a</b>) main effect for wear rate, (<b>b</b>) interaction plot for wear rate, (<b>c</b>) main effect for CoF, and (<b>d</b>) interaction plot for CoF.</p>
Full article ">Figure 6
<p>Comparative display of the experimental results and ANN predictions for (<b>a</b>) wear rate and (<b>b</b>) coefficient of friction.</p>
Full article ">Figure 7
<p>SEM micrograph of worn surfaces: (<b>a</b>) matrix alloy and (<b>b</b>) Al-18Si with 0.5 wt.% Al<sub>2</sub>O<sub>3</sub> nanoparticles and 3 wt.% Al<sub>2</sub>O<sub>3</sub> microparticles.</p>
Full article ">Figure 8
<p>EDS analysis of Al-18Si with 0.5 wt.% Al<sub>2</sub>O<sub>3</sub> nanoparticles and 3 wt.% Al<sub>2</sub>O<sub>3</sub> microparticles tested under 100 N normal load: (<b>a</b>) spectrum positions, (<b>b</b>) Spectrum 1, and (<b>c</b>) Spectrum 2.</p>
Full article ">
30 pages, 70370 KiB  
Article
Influence of the Matrix Material and Tribological Contact Type on the Antifriction Properties of Hybrid Reinforced Polyimide-Based Nano- and Microcomposites
by Dmitry G. Buslovich, Sergey V. Panin, Jiangkun Luo, Ksenya N. Pogosyan, Vladislav O. Alexenko and Lyudmila A. Kornienko
Polymers 2023, 15(15), 3266; https://doi.org/10.3390/polym15153266 - 31 Jul 2023
Viewed by 1219
Abstract
This paper addresses peculiarities in the formation and adherence of a tribofilm on the wear track surface of antifriction PI- and PEI-based composites, as well as a transfer film (TF) on a steel counterface. It is shown that during hot pressing, PTFE nanoparticles [...] Read more.
This paper addresses peculiarities in the formation and adherence of a tribofilm on the wear track surface of antifriction PI- and PEI-based composites, as well as a transfer film (TF) on a steel counterface. It is shown that during hot pressing, PTFE nanoparticles melted and coalesced into micron-sized porous inclusions. In the PEI matrix, their dimensions were much larger (up to 30 µm) compared to those in the PI matrix (up to 6 µm). The phenomenon eliminated their role as effective uniformly distributed nanofillers, and the content of 5 wt.% was not always sufficient for the formation of a tribofilm or a significant decrease in the WR values. At the loaded content, the role of MoS2 and graphite (Gr) microparticles was similar, although filling with MoS2 microparticles more successfully solved the problem of adhering to a PTFE-containing tribofilm in the point tribological contact. This differed under the linear tribological contact. The higher roughness of the steel counterpart, as well as the larger area of its sliding surface with the same PTFE content in the three-component PI- and PEI-based composites, did not allow for a strong adherence of either the stable PTFE-containing tribofilm on the wear track surface or the TF on the steel counterpart. For the PEI-based composites, the inability to shield the steel counterpart from the more reactive polymer matrix, especially under the conditions of PTFE deficiency, was accompanied by multiple increases in the WR values, which were several times greater than that of neat PEI. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>The SEM micrographs of the loaded fillers: µPTFE (<b>a</b>), nanoPTFE (<b>b</b>), µMoS<sub>2</sub> (<b>c</b>), and µGr (<b>d</b>).</p>
Full article ">Figure 2
<p>SEM micrographs showing the structure of the PI/10µPTFE/0.5µMoS<sub>2</sub> (<b>a</b>), PI/10µPTFE/0.5µGr (<b>b</b>), PI/5nanoPTFE (<b>c</b>), PI/5nanoPTFE/0.5µMoS<sub>2</sub> (<b>d</b>), PI/5nanoPTFE/0.5µGr (<b>e</b>), PEI/10µPTFE/0.5µMoS<sub>2</sub> (<b>f</b>), PEI/10µPTFE/0.5µGr (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PEI/5nanoPTFE/0.5µMoS<sub>2</sub> (<b>i</b>), and PEI/5nanoPTFE/0.5µGr (<b>j</b>) composites.</p>
Full article ">Figure 2 Cont.
<p>SEM micrographs showing the structure of the PI/10µPTFE/0.5µMoS<sub>2</sub> (<b>a</b>), PI/10µPTFE/0.5µGr (<b>b</b>), PI/5nanoPTFE (<b>c</b>), PI/5nanoPTFE/0.5µMoS<sub>2</sub> (<b>d</b>), PI/5nanoPTFE/0.5µGr (<b>e</b>), PEI/10µPTFE/0.5µMoS<sub>2</sub> (<b>f</b>), PEI/10µPTFE/0.5µGr (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PEI/5nanoPTFE/0.5µMoS<sub>2</sub> (<b>i</b>), and PEI/5nanoPTFE/0.5µGr (<b>j</b>) composites.</p>
Full article ">Figure 3
<p>The density (<b>a</b>), Shore D hardness (<b>b</b>), elastic modulus (<b>c</b>), and ultimate tensile strength (<b>d</b>) radar diagrams.</p>
Full article ">Figure 4
<p>The CoF (<b>a</b>) and WR (<b>b</b>) radar diagrams; the “B-o-D” scheme.</p>
Full article ">Figure 5
<p>The CoF vs. distance dependencies for the PI- (<b>a</b>) and PEI-based (<b>b</b>) composites under the “B-o-D” scheme.</p>
Full article ">Figure 6
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>) and wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, as well as the wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after the tribological tests for the PI/10µPTFE (<b>a</b>–<b>c</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 6 Cont.
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>) and wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, as well as the wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after the tribological tests for the PI/10µPTFE (<b>a</b>–<b>c</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 7
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>) and wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces and the wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PEI/10PTFE (<b>a</b>–<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PEI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PEI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 7 Cont.
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>) and wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces and the wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PEI/10PTFE (<b>a</b>–<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PEI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PEI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 8
<p>CoF (<b>a</b>,<b>c</b>) and WR (<b>b</b>,<b>d</b>) radar diagrams under the “B-o-R” scheme, <span class="html-italic">p</span> = 60 (<b>a</b>,<b>b</b>) and 180 N (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 9
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE (<b>a</b>), PEI/10µPTFE (<b>b</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>), PI/10µPTFE/µGr (<b>e</b>), PEI/10µPTFE/µGr (<b>f</b>), PI/5nanoPTFE (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>i</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>j</b>), PI/5nanoPTFE/µGr (<b>k</b>), and PEI/5nanoPTFE/µGr (<b>l</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 9 Cont.
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE (<b>a</b>), PEI/10µPTFE (<b>b</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>), PI/10µPTFE/µGr (<b>e</b>), PEI/10µPTFE/µGr (<b>f</b>), PI/5nanoPTFE (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>i</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>j</b>), PI/5nanoPTFE/µGr (<b>k</b>), and PEI/5nanoPTFE/µGr (<b>l</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 10
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE (<b>a</b>), PEI/10µPTFE (<b>b</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>), PI/10µPTFE/µGr (<b>e</b>), PEI/10µPTFE/µGr (<b>f</b>), PI/5nanoPTFE (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>i</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>j</b>), PI/5nanoPTFE/µGr (<b>k</b>), and PEI/5nanoPTFE/µGr (<b>l</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 180 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 10 Cont.
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE (<b>a</b>), PEI/10µPTFE (<b>b</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>), PI/10µPTFE/µGr (<b>e</b>), PEI/10µPTFE/µGr (<b>f</b>), PI/5nanoPTFE (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>i</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>j</b>), PI/5nanoPTFE/µGr (<b>k</b>), and PEI/5nanoPTFE/µGr (<b>l</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 180 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 10 Cont.
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE (<b>a</b>), PEI/10µPTFE (<b>b</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>), PI/10µPTFE/µGr (<b>e</b>), PEI/10µPTFE/µGr (<b>f</b>), PI/5nanoPTFE (<b>g</b>), PEI/5nanoPTFE (<b>h</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>i</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>j</b>), PI/5nanoPTFE/µGr (<b>k</b>), and PEI/5nanoPTFE/µGr (<b>l</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 180 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 11
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>), wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, and wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PI/10µPTFE (<b>a</b>–<b>c</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 11 Cont.
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>), wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, and wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PI/10µPTFE (<b>a</b>–<b>c</b>), PI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PI/5nanoPTFE (<b>j</b>–<b>l</b>), PI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 12
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>),wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, and wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PEI/10µPTFE (<b>a</b>–<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PEI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PEI/5nanoPTFE (<b>j</b>–<b>l</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PEI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 12 Cont.
<p>Optical photographs showing the steel counterface (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>),wear track (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>) surfaces, and wear track profiles (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>) after tribological tests of the PEI/10µPTFE (<b>a</b>–<b>c</b>), PEI/10µPTFE/µMoS<sub>2</sub> (<b>d</b>–<b>f</b>), PEI/10µPTFE/µGr (<b>g</b>–<b>i</b>), PEI/5nanoPTFE (<b>j</b>–<b>l</b>), PEI/5nanoPTFE/µMoS<sub>2</sub> (<b>m</b>–<b>o</b>), and PEI/5nanoPTFE/µGr (<b>p</b>–<b>r</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 13
<p>The dependencies vs. distance of the CoF values for the PI/10µPTFE/µMoS<sub>2</sub> (<b>a</b>) and PEI/10µPTFE/µMoS<sub>2</sub> (<b>b</b>) composites and their wear track surfaces (<b>c</b>,<b>d</b>) under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s, the ceramic counterpart.</p>
Full article ">Figure 14
<p>The time dependencies of the CoF values for the PI/10µPTFE/µMoS<sub>2</sub> (<b>a</b>) ad PEI/10µPTFE/µMoS<sub>2</sub> (<b>b</b>) composites and their wear track surfaces (<b>c</b>,<b>d</b>) under the “B-o-D” scheme; <span class="html-italic">p</span> = 5 N, <span class="html-italic">V</span> = 0.3 m/s; the ceramic counterpart, <span class="html-italic">p</span> = 60 N, <span class="html-italic">V</span> = 0.3 m/s under the “B-o-R” scheme.</p>
Full article ">Figure 15
<p>Photographs showing the steel counterpart after tribological tests against the PI/10µPTFE (<b>a</b>) and PEI/10µPTFE (<b>b</b>) composites under the “B-o-R” scheme; <span class="html-italic">p</span> = 60 and 180 N, <span class="html-italic">V</span> = 0.3 m/s.</p>
Full article ">Figure 16
<p>The Raman spectra of the steel counterface: outside (1) and inside (2) the TF after the tribological test of the PEI/10µPTFE/µMoS<sub>2</sub> composite under the “B-o-D” scheme.</p>
Full article ">Figure 17
<p>The Raman spectra of the neat PEI (<b>a</b>) and the wear track surface of the PEI-based composites after the tribological test (<b>b</b>) under the “B-o-R” scheme; <span class="html-italic">p</span> = 60, <span class="html-italic">V</span> = 0.3 m/s, the steel counterpart.</p>
Full article ">Figure 18
<p>Materials selection approach [<a href="#B34-polymers-15-03266" class="html-bibr">34</a>] applied to the qualification of the PI/PEI-based composites for tribological applications; the point (<b>a</b>) and linear (<b>b</b>,<b>c</b>) tribological contact, <span class="html-italic">p</span> = 5 N (<b>a</b>), 60 N (<b>b</b>), and 180 N (<b>c</b>); V = 0.3 m/s (<b>a</b>–<b>c</b>).</p>
Full article ">
17 pages, 6045 KiB  
Article
Effect of MoO3 Content on Ni3Al-Ag-MoO3 Composite Coating Microstructure and Tribological Properties
by Xiangjuan Fan, Wensheng Li, Jun Yang, Shengyu Zhu, Shuai Cui, Bo Cheng and Haimin Zhai
Coatings 2023, 13(3), 624; https://doi.org/10.3390/coatings13030624 - 15 Mar 2023
Cited by 3 | Viewed by 1386
Abstract
In order to provide effective solid lubrication to Ni3Al coating, 10 wt.% Ag and different amounts of MoO3 solid lubricant were mechanically mixed with the SHSed Ni3Al powder and sprayed HVOF. Microstructure, mechanical properties, and tribological behavior from [...] Read more.
In order to provide effective solid lubrication to Ni3Al coating, 10 wt.% Ag and different amounts of MoO3 solid lubricant were mechanically mixed with the SHSed Ni3Al powder and sprayed HVOF. Microstructure, mechanical properties, and tribological behavior from 25 °C to 800 °C of the coatings were studied, and the basic wear mechanisms were explored and discussed as well. Results show that the hardness and adhesive bonding strength of the coatings are slightly decreased, while there is little effect on the microstructure and mechanical properties of the Ni3Al-based composite coating when the content of MoO3 additive in the feedstock powder is lower than 15 wt.%. The composite coating formed by feedstock powder containing 15 wt.% MoO3 additive also presents excellent anti-friction and anti-wear performance from 25 °C to 800 °C, especially at 800 °C, where a complete, smooth, and thicker lubricating film comprised of NiO, Al2O3, MoO3, and Ag2MoO4 was formed, which reduced the friction coefficient (COF) and wear rate (WR) to the lowest value of 0.36 and 6.03 × 10−5 mm3/(Nm), respectively. An excessive amount of MoO3 in the feedstock powder (20 wt.%) results in inferior interlayer bonding of the formed coating, and the coating is more prone to delamination and abrasive wear above 200 °C. Full article
(This article belongs to the Special Issue Friction, Wear, Lubrication and Mechanics of Surfaces and Interfaces)
Show Figures

Figure 1

Figure 1
<p>XRD results of the NAM15 feedstock powder (<b>a</b>) and corresponding as-sprayed NAM15 coating (<b>b</b>).</p>
Full article ">Figure 2
<p>The polished surface and sectional SEM morphologies of (<b>a1</b>,<b>a2</b>) NAM0, (<b>b1</b>,<b>b2</b>) NAM10, (<b>c1</b>,<b>c2</b>) NAM15, and(<b>d1</b>,<b>d2</b>) NAM20 coatings. (<b>c2′</b>) Local magnified morphology of (<b>c2</b>) and corresponding EDS elemental distributions.</p>
Full article ">Figure 3
<p>COFs (<b>a</b>) and WRs (<b>b</b>) of the NAM coatings at various temperatures.</p>
Full article ">Figure 4
<p>SEM images of the worn tracks of NAM0 composite coating: (<b>a</b>) 25 °C, (<b>b</b>) 200 °C, (<b>c</b>) 400 °C, (<b>d</b>) 600 °C, (<b>e</b>) 800 °C.</p>
Full article ">Figure 5
<p>SEM images of the worn tracks: (<b>a1</b>–<b>e1</b>) NAM10 coating, (<b>a2</b>–<b>e2</b>) NAM15 coating, (<b>a3</b>–<b>e3</b>) and NAM20 coating. (<b>a1</b>–<b>a3</b>) 25 °C, (<b>b1</b>–<b>b3</b>) 200 °C, (<b>c1</b>–<b>c3</b>) 400 °C, (<b>d1</b>–<b>d3</b>) 600 °C, (<b>e1</b>–<b>e3</b>) 800 °C.</p>
Full article ">Figure 6
<p>Raman results of the worn tracks of NAM0 and NAM15 coatings at various temperatures: (<b>a</b>) NAM0 coating and (<b>b</b>) NAM15 coating.</p>
Full article ">Figure 7
<p>XPS spectra of the worn tracks of NAM15 coating: (<b>a</b>) Ag at various temperatures, (<b>b</b>) Mo at 600 °C, (<b>c</b>) Mo at 800 °C.</p>
Full article ">Figure 8
<p>EDS results of the worn tracks (<a href="#coatings-13-00624-f004" class="html-fig">Figure 4</a> and <a href="#coatings-13-00624-f005" class="html-fig">Figure 5</a>) at various temperatures: (<b>a</b>) Ag content and (<b>b</b>) O content.</p>
Full article ">Figure 9
<p>Three-dimensional images of the worn tracks, SEM images, and Ag distribution on the corresponding Si<sub>3</sub>N<sub>4</sub> counterpart balls corresponding to the coatings at 25 °C: (<b>a1</b>,<b>a2</b>) NAM0 coating, (<b>b1</b>,<b>b2</b>) NAM10 coating, (<b>c1</b>,<b>c2</b>) NAM15 coating, (<b>d1</b>,<b>d2</b>) NAM20 coating.</p>
Full article ">Figure 10
<p>Raman results of the NAM10 coating tested at 800 °C: (<b>a</b>) within the worn track and (<b>b</b>) outside the worn track.</p>
Full article ">Figure 11
<p>SEM morphologies and elemental distribution of the Si<sub>3</sub>N<sub>4</sub> counterpart balls against (<b>a</b>) NAM10 coating, (<b>b</b>) NAM15 coating, and (<b>c</b>) NAM20 coating at 800 °C.</p>
Full article ">Figure 12
<p>SEM morphologies of the wear debris at 800 °C: (<b>a</b>) NAM10 coating, (<b>b</b>) NAM15 coating, (<b>c</b>) NAM20 coating.</p>
Full article ">
16 pages, 5092 KiB  
Article
Tribological Behavior of Reduced Graphene Oxide–Al2O3 Nanofluid: Interaction among Testing Force, Rotational Speed and Nanoparticle Concentration
by Chenglong Wang, Jianlin Sun, Linghui Kong and Jiaqi He
Materials 2022, 15(15), 5177; https://doi.org/10.3390/ma15155177 - 26 Jul 2022
Cited by 5 | Viewed by 1344
Abstract
The tribological properties of nanofluids are influenced by multiple factors, and the interrelationships among the factors are deserving of further attention. In this paper, response surface methodology (RSM) was used to study the tribological behavior of reduced graphene oxide–Al2O3 (rGO-Al [...] Read more.
The tribological properties of nanofluids are influenced by multiple factors, and the interrelationships among the factors are deserving of further attention. In this paper, response surface methodology (RSM) was used to study the tribological behavior of reduced graphene oxide–Al2O3 (rGO-Al2O3) nanofluid. The interaction effects of testing force, rotational speed and nanoparticle concentration on the friction coefficient (μ), wear rate (Wr) and surface roughness (Ra) of steel disks were investigated via the analysis of variance. It was confirmed that all the three input variables were significant for μ and Wr values, while testing force, nanoparticle concentration and its interaction with testing force and rotational speed were identified as significant parameters for Ra value. According to regression quadratic models, the optimized response values were 0.088, 2.35 × 10−7 mm3·N−1·m−1 and 0.832 μm for μ, Wr and Ra, which were in good agreement with the actual validation experiment values. The tribological results show that 0.20% was the optimum mass concentration which exhibited excellent lubrication performance. Compared to the base fluid, μ, Wr and Ra values had a reduction of approximately 45.6%, 90.3% and 56.0%. Tribochemical reactions occurred during the friction process, and a tribofilm with a thickness of approximately 20 nm was generated on the worn surface, consisting of nanoparticle fragments (rGO and Al2O3) and metal oxides (Fe2O3 and FeO) with self-lubrication properties. Full article
(This article belongs to the Section Thin Films and Interfaces)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the disk-on-disk tribotester and the photo of the pin and disk.</p>
Full article ">Figure 2
<p>TEM images of (<b>a</b>) rGO, (<b>b</b>) Al<sub>2</sub>O<sub>3</sub> and (<b>c</b>) rGO-Al<sub>2</sub>O<sub>3</sub> nanoparticles.</p>
Full article ">Figure 3
<p>(<b>a</b>) C 1s, (<b>b</b>) O 1s, (<b>c</b>) Al 2p XPS spectra of the rGO-Al<sub>2</sub>O<sub>3</sub> nanocomposite, and (<b>d</b>) XRD patterns of different nanoparticles.</p>
Full article ">Figure 4
<p>Normal probability plot of residual for (<b>a</b>) <span class="html-italic">μ</span>, (<b>b</b>) <span class="html-italic">W</span><sub>r</sub> and (<b>c</b>) <span class="html-italic">R</span><sub>a</sub>.</p>
Full article ">Figure 5
<p>The parity plot illustrating the correlation between actual and predicted values for (<b>a</b>) <span class="html-italic">μ</span>, (<b>b</b>) <span class="html-italic">W</span><sub>r</sub> and (<b>c</b>) <span class="html-italic">R</span><sub>a</sub>.</p>
Full article ">Figure 6
<p>3D response surface plots showing the interaction effects of different variables on (<b>a</b>–<b>c</b>) <span class="html-italic">μ</span>, (<b>d</b>–<b>f</b>) <span class="html-italic">W</span><sub>r</sub> and (<b>g</b>–<b>i</b>) <span class="html-italic">R</span><sub>a</sub> values.</p>
Full article ">Figure 7
<p>Tribological behavior of base fluid and rGO-Al<sub>2</sub>O<sub>3</sub> nanofluid: (<b>a</b>) friction coefficient–time curves and (<b>b</b>) wear rate and average friction coefficient.</p>
Full article ">Figure 8
<p>Surface topography and cross-section depth profile of wear track on the disk lubricated by (<b>a</b>) base fluid and (<b>b</b>) rGO-Al<sub>2</sub>O<sub>3</sub> nanofluid.</p>
Full article ">Figure 9
<p>SEM micrograph of worn disk surface lubricated with 0.20 wt.% rGO-Al<sub>2</sub>O<sub>3</sub> nanofluid and EDS map scanning results.</p>
Full article ">Figure 10
<p>(<b>a</b>) TEM image of tribofilm on the worn disk surface, XPS resolved fitting curves of (<b>b</b>) Fe 2p, (<b>c</b>) O 1s and (<b>d</b>) Al 2p spectra of the worn surface lubricated by 0.20 wt.% rGO-Al<sub>2</sub>O<sub>3</sub> nanofluid.</p>
Full article ">
14 pages, 2513 KiB  
Article
Tribological Analysis of Jute/Coir Polyester Composites Filled with Eggshell Powder (ESP) or Nanoclay (NC) Using Grey Rational Method
by Ganesan Karuppiah, Kailasanathan Chidambara Kuttalam, Nadir Ayrilmis, Rajini Nagarajan, M. P. Indira Devi, Sivasubramanian Palanisamy and Carlo Santulli
Fibers 2022, 10(7), 60; https://doi.org/10.3390/fib10070060 - 12 Jul 2022
Cited by 12 | Viewed by 2296
Abstract
The wear performance of jute/coir unsaturated polyester composites, filled with eggshell powder (ESP) and nanoclay (NC), were examined, concentrating on two measured parameters, coefficient of friction (COF) and wear rate (WR). To assess the possibilities of this material, a Taguchi study, based on [...] Read more.
The wear performance of jute/coir unsaturated polyester composites, filled with eggshell powder (ESP) and nanoclay (NC), were examined, concentrating on two measured parameters, coefficient of friction (COF) and wear rate (WR). To assess the possibilities of this material, a Taguchi study, based on grey relational analysis (GRA), was carried out, based on three testing parameters of the wear performance, load (10, 20, and 30 N), speed (100, 150, and 200 rpm), and sliding distance (30, 40, and 50 m). The material showed promising characteristics especially at high load, low speed, and high sliding distance. When comparing the respective influence of the three different parameters, the speed proved to be the most critical, this suggested the possible application of the biocomposite only for very low values of it. On the other hand, it was also elucidated that the presence and interfacial adhesion of the two fillers considerably hindered the formation of ploughing during wear test, despite the fact that degradation might be continuous and critical as far as loading progresses. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the laminates lay-up before pressure consolidation.</p>
Full article ">Figure 2
<p>Experimental setup for pin-on-disc wear tester.</p>
Full article ">Figure 3
<p>Grey rational grades.</p>
Full article ">Figure 4
<p>Sample with low COF and high wear rate (sample n.1): (<b>a</b>) Before wear test; (<b>b</b>) After wear test.</p>
Full article ">Figure 5
<p>Sample with high COF and high wear rate (sample n.6): (<b>a</b>) Before wear test; (<b>b</b>) After wear test.</p>
Full article ">Figure 6
<p>Sample with optimal conditions (sample n. 7): (<b>a</b>) Before wear test; (<b>b</b>) After wear test.</p>
Full article ">
20 pages, 1447 KiB  
Article
A Conceptual Comparison of Six Nature-Inspired Metaheuristic Algorithms in Process Optimization
by Shankar Rajendran, Ganesh N., Robert Čep, Narayanan R. C., Subham Pal and Kanak Kalita
Processes 2022, 10(2), 197; https://doi.org/10.3390/pr10020197 - 20 Jan 2022
Cited by 22 | Viewed by 3182
Abstract
In recent years, several high-performance nature-inspired metaheuristic algorithms have been proposed. It is important to study and compare the convergence, computational burden and statistical significance of these metaheuristics to aid future developments. This study focuses on six recent metaheuristics, namely, ant lion optimization [...] Read more.
In recent years, several high-performance nature-inspired metaheuristic algorithms have been proposed. It is important to study and compare the convergence, computational burden and statistical significance of these metaheuristics to aid future developments. This study focuses on six recent metaheuristics, namely, ant lion optimization (ALO), arithmetic optimization algorithm (AOA), dragonfly algorithm (DA), grey wolf optimizer (GWO), salp swarm algorithm (SSA) and whale optimization algorithm (WOA). Optimization of an industrial machining application is tackled in this paper. The optimal machining parameters (peak current, duty factor, wire tension and water pressure) of WEDM are predicted using the six aforementioned metaheuristics. The objective functions of the optimization study are to maximize the material removal rate (MRR) and minimize the wear ratio (WR) and surface roughness (SR). All of the current algorithms have been seen to surpass existing results, thereby indicating their superiority over conventional optimization algorithms. Full article
Show Figures

Figure 1

Figure 1
<p>Convergence of optimum <span class="html-italic">MRR</span> with respect to iterations.</p>
Full article ">Figure 2
<p>Box plots denoting the spread of all function evaluations during <span class="html-italic">MRR</span> optimization in a typical independent run. The horizontal blue line (and numeric value) and the red line indicate the median and mean of all the function evaluations.</p>
Full article ">Figure 3
<p>Convergence of optimum <span class="html-italic">WR</span> with respect to iterations.</p>
Full article ">Figure 4
<p>Box plots denoting the spread of all the function evaluations during <span class="html-italic">WR</span> optimization in a typical independent run. The horizontal blue line (and numeric value) and the red line indicate the median and mean of all the function evaluations.</p>
Full article ">Figure 5
<p>Average time taken by the algorithms in the optimization of the responses. The bars and whiskers denote the mean and standard deviation of 10 runs (3 responses X 10 independent runs/response).</p>
Full article ">
15 pages, 3405 KiB  
Article
Tribological Characterization of Ni-Free Duplex Stainless Steel Alloys Using the Taguchi Methodology
by Hammam Daraghma, Mohammed Abdul Samad, Ihsan ul Haq Toor, Farid M. Abdallah and Faheemuddin Patel
Metals 2020, 10(3), 339; https://doi.org/10.3390/met10030339 - 3 Mar 2020
Cited by 4 | Viewed by 2906
Abstract
Duplex stainless steels (DSSs) exhibit excellent corrosion resistance and are being used in a variety of industrial applications. Reducing/eliminating the amount of nickel in such alloys will contribute significantly to its economic viability. Moreover, a well-established wear behavior for these alloys is also [...] Read more.
Duplex stainless steels (DSSs) exhibit excellent corrosion resistance and are being used in a variety of industrial applications. Reducing/eliminating the amount of nickel in such alloys will contribute significantly to its economic viability. Moreover, a well-established wear behavior for these alloys is also an essential development in most of their applications. Hence, in this work, the Taguchi technique was effectively implemented to investigate the effect of operating factors such as sliding speed and applied load on the wear behavior of different compositions of nickel-free DSSs. It was observed that the composition had a higher contribution of 33.66% to the wear rate (WR) and the contribution of the sliding speed to the coefficient of friction (COF) was found to be 68.17%. With a good agreement, a regression model was also developed to predict the WR and COF within a certain range of factors. Wear tests have also shown that the developed nickel-free DSS is a promising candidate in terms of wear resistance as compared to austenitic stainless steels (ASS). Full article
Show Figures

Figure 1

Figure 1
<p>Representative images of the wear track from the optical profilometer for one of the experimental runs: (<b>a</b>) 3D profile; (<b>b</b>) surface profile; and (<b>c</b>) 2D profile.</p>
Full article ">Figure 2
<p>Main effect plot for the mean specific wear rate.</p>
Full article ">Figure 3
<p>Plot for the corresponding S/N ratios.</p>
Full article ">Figure 4
<p>Typical graph for coefficient of friction (COF) vs. time for a wear test conducted on the DSS1 sample at a sliding speed of 0.3 m/s and at a load of 30N (Run #3 in <a href="#metals-10-00339-t008" class="html-table">Table 8</a>).</p>
Full article ">Figure 5
<p>Main effect plot for the mean COF.</p>
Full article ">Figure 6
<p>Plot for the corresponding S/N ratios.</p>
Full article ">Figure 7
<p>Microscopic images for the counterface ball (<b>a</b>) before and after for (<b>b</b>) DSS1, 0.2 m/s, 25 N; (<b>c</b>) DSS2, 0.1 m/s, 25 N; and (<b>d</b>) DSS3, 0.3 m/s, 25 N.</p>
Full article ">Figure 8
<p>SEM images with different magnifications for the DSS1 tested sample with 0.1 m/s speed and 25 N load. (<b>a</b>) 60×, (<b>b</b>) 100× and (<b>c</b>) 500×.</p>
Full article ">Figure 9
<p>SEM images with different magnifications for the DSS2 tested sample with 0.1 m/s speed and 25 N load. (<b>a</b>) 60×, (<b>b</b>) 100× and (<b>c</b>) 500×.</p>
Full article ">Figure 10
<p>SEM images with different magnifications for the DSS3 tested sample with 0.3 m/s speed and 25 N load. (<b>a</b>) 60×, (<b>b</b>) 100× and (<b>c</b>) 500×.</p>
Full article ">Figure 11
<p>SEM image of the wear track after the wear test along with the EDS mapping showing the presence of oxygen indicative of oxidative wear.</p>
Full article ">
17 pages, 6371 KiB  
Article
Synthesis and Characterization of Novel Ti3SiC2 Reinforced Ni-Matrix Multilayered Composite-Based Solid Lubricants
by Quan Tran, Matt Fuka, Maharshi Dey and Surojit Gupta
Lubricants 2019, 7(12), 110; https://doi.org/10.3390/lubricants7120110 - 9 Dec 2019
Cited by 3 | Viewed by 3704
Abstract
We report the synthesis and characterization of two different types of Ni-based laminated composites (Types I and II). In Type-I composites, layers of Ni and Ti3SiC2 (Ni–Ti3SiC2) were interleaved with Ni, whereas in Type-II composites, Ni–Ti [...] Read more.
We report the synthesis and characterization of two different types of Ni-based laminated composites (Types I and II). In Type-I composites, layers of Ni and Ti3SiC2 (Ni–Ti3SiC2) were interleaved with Ni, whereas in Type-II composites, Ni–Ti3SiC2 layers were interleaved with Al and Ni. The laminate thickness and Ti3SiC2 content in the individual Ni–Ti3SiC2 layers were systematically varied in both the composites. Detailed SEM studies showed that Ti3SiC2 particulates are well distributed in the Ni-matrix with little or no interfacial reactions with interparticle porosity. However, there were interfacial reactions between Ni and Al in Type II composites. In general, Type I multilayered composites had higher ultimate compressive strength (UCS) in parallel orientation as compared to perpendicular orientation (layers are aligned parallel or perpendicular to the wear surface then it will be referred to as parallel or perpendicular orientation). Comparatively, in Type II composites, the UCS was greater in perpendicular orientation as compared to parallel due to the presence of Al layers as bonding layers. Both the composite designs showed triboactive behavior against alumina disks and sensitivity to laminate thickness and orientation. In Type-I composites, the decrease in µ and wear rate (WR) with laminate thickness was more pronounced in the perpendicular orientation as compared to the parallel orientation. Comparatively, Ni–Ti3SiC2/Al/Ni composites showed that the parallel orientation was more effective in enhancing the triboactive performance. SEM analysis of tribosurfaces showed signs of triboxidation and abrasion, which led to the formation of O-rich tribofilms. Full article
Show Figures

Figure 1

Figure 1
<p>Schematics of, (<b>a</b>) Ni–Ti<sub>3</sub>SiC<sub>2</sub>/Ni, and (<b>b</b>) Ni–Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni multilayered composites.</p>
Full article ">Figure 2
<p>Concentration of Ti<sub>3</sub>SiC<sub>2</sub> in: (<b>a</b>) isotropic Ni–Ti<sub>3</sub>SiC<sub>2</sub> composites, (<b>b</b>) multilayered Ni–Ti<sub>3</sub>SiC<sub>2</sub> composites.</p>
Full article ">Figure 3
<p>Schematics of (<b>a</b>) parallel and (<b>b</b>) perpendicular orientation during mechanical and tribological testing.</p>
Full article ">Figure 4
<p>SEM microstructure of (<b>a</b>) Ni-10%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) in backscattered electron (BSE) image, (<b>b</b>) BSE image at higher magnifications, (<b>c</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) in secondary electron (SE) image, (<b>d</b>) BSE of the same region, (<b>e</b>) Ni-40%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) in BSE, and (<b>f</b>) Ni-40%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) in BSE at higher magnifications.</p>
Full article ">Figure 5
<p>SEM microstructures of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (20 µm) in BSE, (<b>b</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (20 µm) in BSE at higher magnifications, (<b>c</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (200 µm) in SE, and (<b>d</b>) BSE of the same region.</p>
Full article ">Figure 6
<p>SEM microstructure of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni (20 µm) in BSE, (<b>b</b>) higher magnification in BSE, (<b>c</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni (100 µm) in BSE, and (<b>d</b>) higher magnification in BSE.</p>
Full article ">Figure 7
<p>Plot of compressive stress versus displacement of (<b>a</b>) Ni-10%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (thickness of laminate is 100 µm), (<b>b</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (thickness of laminate is 100 µm), (<b>c</b>) Ni-40%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (thickness of laminate is 100 µm), (<b>d</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (thickness of laminate is 20 µm), (<b>e</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (thickness of laminate is 200 µm), and (<b>f</b>) Ni-10%Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni composites with different thicknesses.</p>
Full article ">Figure 8
<p>Plot of ultimate compressive strength (UCS) versus Ti<sub>3</sub>SiC<sub>2</sub> content in (<b>a</b>)Ni–Ti<sub>3</sub>SiC<sub>2</sub> (100 µm), (<b>b</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni, and (<b>c</b>) Ni–Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni multilayered composites.</p>
Full article ">Figure 9
<p>Plot of (<b>a</b>) friction coefficient, and (<b>b</b>) wear rate (<span class="html-italic">WR</span>) of Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni multilayered composites as a function of laminate thickness.</p>
Full article ">Figure 10
<p>Plot of (<b>a</b>) friction coefficient, and (<b>b</b>) wear rate as a function of Ti<sub>3</sub>SiC<sub>2</sub> content in Ni–Ti<sub>3</sub>SiC<sub>2</sub>/Ni composites (laminate thickness in all composites was 100 µm) [<a href="#B20-lubricants-07-00110" class="html-bibr">20</a>,<a href="#B21-lubricants-07-00110" class="html-bibr">21</a>].</p>
Full article ">Figure 11
<p>Comparative plot of (<b>a</b>) friction coefficient, and (<b>b</b>) <span class="html-italic">WR</span> of Ni–Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni multilayered composites.</p>
Full article ">Figure 12
<p>SEM micrographs of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) surface in SE, (<b>b</b>) BSE of the same region, (<b>c</b>) alumina surface in SE, and (<b>d</b>) BSE of the same region after tribological testing in parallel orientation.</p>
Full article ">Figure 13
<p>SEM micrographs of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Ni (100 µm) surface (perpendicular to the casting direction) in SE, (<b>b</b>) BSE of the same region, (<b>c</b>) alumina surface in SE, and (<b>d</b>) BSE of the same region after tribological testing in perpendicular orientation.</p>
Full article ">Figure 14
<p>SEM micrographs of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni (100 µm) surface (parallel orientation) in SE, (<b>b</b>) BSE of the same region, (<b>c</b>) alumina surface in SE, and (<b>d</b>) BSE of the same region after tribological testing.</p>
Full article ">Figure 15
<p>SEM micrographs of (<b>a</b>) Ni-20%Ti<sub>3</sub>SiC<sub>2</sub>/Al/Ni (100 µm) surface (perpendicular to the casting direction) in SE, (<b>b</b>) BSE of the same region, (<b>c</b>) alumina surface in SE, and (<b>d</b>) BSE of the same region after tribological testing.</p>
Full article ">
7989 KiB  
Article
Synthesis and Tribological Behavior of Ultra High Molecular Weight Polyethylene (UHMWPE)-Lignin Composites
by Surojit Gupta, M. F. Riyad and Yun Ji
Lubricants 2016, 4(3), 31; https://doi.org/10.3390/lubricants4030031 - 31 Aug 2016
Cited by 3 | Viewed by 6100
Abstract
In this paper, we report the synthesis and characterization of ultra-high molecular weight polyethylene (UHMWPE)-lignin composites. During this study four different compositions, namely UHMWPE, UHMWPE-13 wt. % lignin, UHMWPE-25 wt. % lignin and UHMWPE-42.5 wt. % lignin were fabricated by hot pressing. Detailed [...] Read more.
In this paper, we report the synthesis and characterization of ultra-high molecular weight polyethylene (UHMWPE)-lignin composites. During this study four different compositions, namely UHMWPE, UHMWPE-13 wt. % lignin, UHMWPE-25 wt. % lignin and UHMWPE-42.5 wt. % lignin were fabricated by hot pressing. Detailed microstructural studies by scanning electron microscopy (SEM) showed that UHMWPE and UHMWPE-13 wt. % lignin had a uniform microstructure, whereas UHMWPE-25 wt. % lignin and UHMWPE-42.5 wt. % lignin samples were riddled with pores. UHMWPE and UHMWPE-13% lignin showed comparable flexural strengths of ~32.2 MPa and ~32.4 MPa, respectively. However, the flexural strength dropped drastically in UHMWPE-25 wt. % lignin and UHMWPE-42.5 wt. % samples to ~13 MPa and ~8 MPa, respectively. The tribology of UHMWPE-lignin composites is governed by the tribofilm formation. All the compositions showed similar µmean values and the specific wear rates (WR) decreased gradually as the concentration of lignin in UHMWPE was increased. Full article
(This article belongs to the Special Issue Green Tribology)
Show Figures

Figure 1

Figure 1
<p>SE SEM micrographs of polished surfaces of (<b>a</b>) UHMWPE; (<b>b</b>) UHMWPE-13% lignin; (<b>c</b>) UHMWPE-25% lignin surface machined by a blade and; (<b>d</b>) fractured surface of UHMWPE-25% lignin; (<b>e</b>) UHMWPE-42.5% lignin; and (<b>f</b>) fractured surface of UHMWPE-42.5% lignin.</p>
Full article ">Figure 2
<p>Plot of flexural stress versus displacement of different UHMWPE-lignin composites.</p>
Full article ">Figure 3
<p>Plot of flexural strength versus lignin addition (wt. %) in the UHMWPE matrix.</p>
Full article ">Figure 4
<p>Variation of friction coefficient versus distance of different lignin-UHMWPE composites.</p>
Full article ">Figure 5
<p>Plot of friction coefficient (<span class="html-italic">Y</span>1) and <span class="html-italic">WR</span> (<span class="html-italic">Y</span>2) versus lignin additions (wt. %).</p>
Full article ">Figure 6
<p>Digital pictures of alumina discs after tribological testing against, (<b>a</b>) UHMWPE and (<b>b</b>) UHMWPE-42.5 wt. % and SE SEM micrographs of (<b>c</b>) UHMWPE; (<b>d</b>) alumina surface; (<b>e</b>) UHMWPE-42.5 wt. % and (<b>f</b>) alumina surface (inset shows the morphology of polymer wear debris) after tribological testing.</p>
Full article ">Figure 7
<p>Schematics of tribofilm formation in UHMWPE-lignin and alumina tribocouple: (<b>a</b>) tribocontact development; and (<b>b</b>) formation of discontinuous tribofilm formation by abrasive wear of UHMWPE-lignin surface.</p>
Full article ">
Back to TopTop