Next Article in Journal
Non-Periodic Quantized Model Predictive Control Method for Underwater Dynamic Docking
Previous Article in Journal
Navicular Drop Height Asymmetry as an Intrinsic Risk Factor for Lower Limb Injury in Male Recreational Athletes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unraveling the Subsurface Damage and Material Removal Mechanism of Multi-Principal-Element Alloy FeCrNi Coatings During the Scratching Process

1
Hunan Province Key Laboratory of Materials Surface/Interface Science Technology, Central South University of Forestry Technology, Changsha 410004, China
2
State Key Laboratory of Powder Metallurgy, Central South University, Changsha 410083, China
3
State Key Laboratory of Advanced Design and Manufacturing for Vehicle Body, College of Mechanical and Vehicle Engineering, Hunan University, Changsha 410082, China
*
Authors to whom correspondence should be addressed.
Symmetry 2024, 16(10), 1391; https://doi.org/10.3390/sym16101391
Submission received: 2 September 2024 / Revised: 14 October 2024 / Accepted: 15 October 2024 / Published: 18 October 2024
(This article belongs to the Section Engineering and Materials)

Abstract

:
Multi-principal-element alloys (MPEAs) exhibit superior strength and good ductility. However, tribological properties of FeCrNi MPEAs remain unknown at nanoscale and complex environments. Here, we investigate the effects of scratching speed, depth, and temperature on microstructural and tribological characteristics of FeCrNi using molecular dynamics simulations combined with an elevated temperature tribological experiment. The scratching force experiences the increase stage, the undulated stage, and the stable stage due to chip formation. Compared to traditional alloy coatings, low force enhances the useful life. With increased speed, the friction coefficient decreases, agreeing with previous work. High speed impacting includes severe local plastic deformation, from dislocation to amorphization. As the scratching depth increases, the average scratch force and friction coefficient increases owing to material accumulation in front of the abrasive particles. The surface morphology and dislocation behavior are significantly different during the scratching process. In addition, we revealed a temperature-dependent friction mechanism. FeCrNi MPEAs have excellent wear resistance at an intermediate temperature, which is attributed to the high Cr content promoting the formation of the compact oxide layer. This work provides atomic-scale mechanistic insights into the tribological behavior of FeCrNi, and would be applied to the design of MPEAs with high performance.

1. Introduction

The multi-principal-element alloys (MPEAs) exhibit an exceptional combination of high tensile ductility and good strength [1,2,3,4,5], and have promising application prospects in aerospace, medicine, national defense, and other fields. In fact, MPEA with three or more base elements, as a new alloy development philosophy, is extremely different from conventional alloys that have one principal element [1,2]. This new alloying strategy not only vastly improves the number of possible alloy systems, but also provides a rich composition and largely unknown phase space.
Recently, the face-centred-cubic (FCC) FeCrNi MPEAs have shown attractive properties, such as balanced strength and ductility, as well as outstanding corrosion resistance [6,7,8]. For example, the FeCrNi MPEA manufactured by selective laser melting shows an excellent strength-ductility combination due to hierarchical microstructures [9]. Compared with the 316 L steels, the FeCrNi MPEA prepared by powder metallurgy exhibits a high yield strength, good elongation, and outstanding corrosion resistance [10]. In addition, the oxide-dispersion-strengthened (ODS) FeCrNi MPEA fabricated through mechanical alloying and hot extrusion has a significantly enhanced strength, and suppresses defect growth and irradiation-induced segregation, resulting in the excellent irradiation resistance [11]. Therefore, considering the excellent strength, toughness, and corrosion resistance exhibited by the FeCrNi MPEA, studying its tribological characteristics, especially its wear mechanisms under high temperature environments, is an interesting topic. However, the tribological characteristics of FeCrNi MPEA should be explored at the atomic scale, and would be difficult to study through experiments.
Molecular dynamics (MD) simulations play a critical role in revealing the atomic-scale kinetic microstructural and tribological characteristics in metals and alloys [12,13,14]. The nanoindentation deformation mechanisms of TiN along the [001] and [111] crystal orientation has been investigated by atomic simulation, and the coating withstood high stresses to effectively protect the substrate from damage [15]. The nanoscale wear behavior of SiC-strengthened aluminum alloys was also studied, and strong dislocation strengthening occurs during nanoscratching [16]. The roles of the alloy composition on the wear behavior, dislocation density, and von-Mises stress have also been investigated using MD simulation [17]. It was found that the high alloy depth results in a large number of defects in the deformed region, and the high Cu content induces high dislocation density. In addition, the friction behavior of MPEAs as coatings has been partially studied. For example, by adjusting the friction speed, it was found that high velocity reduces the friction coefficient, owing to the inhibition behavior of the dislocation nucleation in the FeNiCrCoCu high entropy alloy coating [18]. The doping of Mn in FeCoCrNiMn induces a phase transition and improves friction reduction and wear resistance [19]. Meanwhile, the effect of grain size and scratching depth on the tribological behavior of CoCrNi MPEA are investigated. Friction force is reduced due to grain boundary sliding and dislocation release caused by scratching [20]. In addition, more mechanical properties and atomic scale deformation mechanisms of MPEAs have been explored by MD simulation [21,22,23,24,25]. However, the tribological performance of FeCrNi MPEA at nanoscale has not been rigorously examined by MD simulation under complex environments, such as the high temperature.
In the current work, the scratching processes of FeCrNi MPEA are investigated by MD simulations and experiments. The effect of scratching speed, depth, and temperature on the tribological characteristics in FeCrNi MPEAs are systematically discussed. Combined with frictional experiments, the effects of operating temperatures from room temperature to 1100 K on its microstructural and tribological properties are studied, and a temperature-dependent friction mechanism is revealed. The current results provide a reference for the possibility of FeCrNi MPEA coatings being applied in the high temperature environments.

2. Methods

2.1. Experiment

The FeCrNi MPEA was fabricated by a powder metallurgy method, and the chemical composition is detected to be Fe33.7Cr33.7Ni32.5 (at.%) by inductively coupled plasma mass spectrometry analysis. Through an X-ray diffractometer (XRD, Advance D8, Bruker, Karlsruhe, Germany) equipped with a Cu Ka radiation, phase identification of FeCrNi MPEA was studied. Using a scanning electron microscope (SEM, Helios G3 UC, FEI, Brno, Czech Republic) equipped with an energy dispersive spectrometer (EDS) instrument, microstructure characterization of FeCrNi MPEA was examined. The tribological test was carried out on a HT-1000 ball-on-disk tribometer at 300 K, 900 K, and 1100 K, respectively. Si3N4 balls (d = 3 mm) were used as counterpart materials. The applied load, sliding speed and test time were 10 N, 0.084 m/s, and 30 min, respectively. To ensure accuracy, each test was repeated at least three times. The wear surfaces were examined by a three-dimensional (3D) optical profilometer (VHX-5000, Keyence, Osaka, Japan), SEM and EDS.

2.2. Molecular Dynamics Simulation

In the current work, the simulation model of the FeCrNi MPEA coating on nickel substrate was built using the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) [26,27], as shown in Figure 1. Based on our experiment, the FeCrNi MPEA has an FCC lattice structure, and the lattice constant is 3.525 Å. The size of the FeCrNi MPEA coating is 388 Å × 180 Å × 36 Å, and it contains a total of 229,986 atoms. The size of Ni substrate is 388 Å × 180 Å × 144 Å, and it contains a total of 875,964 atoms. A spherical tool with 47,262 atoms has a radius of 4.0 nm. The crystal orientations of the FeCrNi MPEA coating and the Ni substrate are [100] along the X axis, [010] along the Y axis, and [001] along the Z axis. The time step is 1.0 fs. Figure 1 shows the workpiece composed of three layers (boundary layer, thermostat layer, and Newtonian layer) [28,29]. The thickness of both the boundary layer and thermostat layer is set to 5 nm. The thermostat layer keeps a constant temperature based on the velocity scaling method. The Newtonian layer meets the classic Newton’s second law [28,29,30]. The y direction is set to the periodic boundary condition. Free boundary conditions are used to set the directions of x and z. Before scratching, the conjugate gradient algorithm was applied to the sample to minimize system energy. During the scratching process, the scratching tool has a constant speed along the x direction (Figure 1). The OVITO software (version Ovito 2.9.0.exe) with the dislocation extraction algorithm (DXA) was used to obtain dislocation characteristics [31,32,33]. The nanoscratching parameters of FeCrNi are summarized in Table 1.
The nanoscratching model consists of the FeCrNi sample, Ni substrate, and diamond tool. There are Fe, Cr, Ni, and C atom types. Atomic interactions between Fe-Fe, Cr-Cr, Ni-Ni, Fe-Cr, Fe-Ni, Cr-Ni, C-Fe, C-Cr, C-Ni, and C-C are considered. A mixture of Morse potential and embedded atom (EAM) potential are used, and the C-C interatomic interactions are not considered owing to the diamond tool being treated as a rigid body. The interactions of Fe-Fe, Cr-Cr, Ni-Ni, Fe-Cr, Fe-Ni, and Cr-Ni are described by EAM potential applicable to FeCrNi [34], which is suitable for general performance analysis compared with experimental results [34,35,36]. The interactions between C-Fe, C-Cr, and C-Ni atoms are described by Morse potential, and the parameters are listed in Table 2 [37].

3. Results and Discussions

3.1. Influence of Scratching Speed

Figure 2a shows the relationship between scratching force and moving distance for different speeds. Here, the three regions are divided, including the increase stage, undulated stage, and stable stage [36,37,38,39,40]. At the initial scratching stage, a few scratching chips formed at the front of the tool. With increasing distance, chips increased, up to the high value [39,40]. For the stable stage, the local chips could be left at the front of the tool. On the other hand, the mechanical properties from the complex elements always changed along the scratching direction, thus leading to the force curve. It can be noted that the changed degree relies on the speed. To study the speed effects on the friction force, the relationship between average friction force and speed is presented in Figure 2b. As the speed goes up, the force first increases, and then keeps a constant value. Interestingly, the obvious transited region only occurs in the relative small speed range. These findings agree well with previous work [38,39,40,41], where the speed controlled force exits during the nanoscale scratching. Compared to the classic alloy [37,38,39,40], the MPEAs show low force, thus enhancing the useful life.
The surface morphology is presented in Figure 3, where the atoms are colored based on their height. The higher speed causes the more chips at the front of the tool. This means that a low speed has a better ability to withstand the scratching process [39,40,41,42]. High speed impacting would include severe local plastic deformation, where the mechanism ranges from dislocation to amorphization. This trend significantly affects the scratching behavior.
Figure 4a shows the friction coefficient-moving distance curves of FeCrNi MPEA under different speeds during the nanoscratching process. In order to study the effect of speed on the friction coefficient, the average friction coefficient is defined as the ratio of the average tangential force Fx to the average normal force Fz during the scratching process [38,39,40], which is widely used in the study of the interaction between spherical abrasive particles and samples. According to the above definition, the average friction coefficient of FeCrNi MPEA under different speeds was calculated. The calculation results are shown in Figure 4b. With the increase of scratching velocity, the friction coefficient decreases. This result agrees with the previous work, which reports the relationships among the wear, strain rate hardening, and thermal softening in the alloys [41,42]. The high scratching velocity results in increasing temperature to soften the scratching material in front of the tool. As a result, the tangential force changes slightly.
The plastic deformation is completely realized by dislocation nucleation and dislocation slip. Figure 5 shows the dislocation evolution in FeCrNi MPEA at different scratching speeds during the nanoscratching process [40]. It can be seen from Figure 5a–d that the dislocation distribution changes dramatically with the increase of scratching speed. The corresponding microstructural evolution is observed in Figure 6. In order to accurately compare the difference between dislocation length and speed, Figure 5e shows the trend of the dislocation length. In the case of high density dislocation entanglement, tangential forces increase appropriately, resulting in a strengthening effect. As can be seen from Figure 5e, the number of the dislocations reduces markedly at high scratching velocity [36,37,38]. When the scratching velocity is 100 m/s, the length of the dislocation decreases significantly.

3.2. Effect of Scratching Depth

During the scratching process, Figure 7a shows the scratching force for different depths. As the moving distance increases, the force value increases, and then tends to a constant. At the beginning of the scratching stage, the scratching force gradually goes up with the increasing contact between the tool and workpiece. When the scratching distance is larger than 0.5 nm, the fluctuation of the scratching force tends to be stable. Figure 7b shows the change curve of the average scratching force in the FeCrNi MPEA under different scratch depths. It can be observed in Figure 7 that the average scratching force also increases with the increase of scratching depth, which is related to the forward extension of dislocation generated in front of the abrasive particles [37,38,39,40,41]. This trend results in material accumulation in front of the abrasive particle. As the scratching process progresses, many atoms accumulate in front of the abrasive particle, resulting in an increase in the scratching force.
Figure 8 shows the scratching surface morphology of the substrate at different scratching depths during the scratching process. The atoms in the diagram are colored according to their heights in the z direction. It can be seen that the scratching morphology is different at different scratching depths, and the stacking structure appears as an asymmetric shape after scratching on the surface of the workpiece. As the scratching depth and distance increase, the side and front of the scratching tool produce accumulations [38,39,40], resulting in an increase in the contact atoms between the workpiece and the tool during the scratching process. The result shows an increase in total wear. The number of wear atoms increases with the increase of scratching length, especially at large scratching depths. At the larger scratching depth, a sharp increase in the number of worn atoms in the scratch causes a high material removal rate.
Figure 9a describes the evolution of the friction coefficient of FeCrNi MPEA coatings with scratching distance when the scratching depth is 0.5, 1, 1.5, and 2 nm. With the increase in the scratching distance, the friction coefficient first increases rapidly due to the increase of the number of atoms around the particle, and then fluctuates around a constant value. The fluctuation of the friction coefficient is caused by the complex dislocation motion during the nanoscratching process of the FeCrNi MPEA coating. The dislocations nucleate and expand from the scratched area, and eventually accumulate at the interface of the FeCrNi MPEA coating [38,39,40,41,42], resulting in an increase in the friction coefficient. Dislocations pass through the FeCrNi MPEA coating interface or nucleate and emit at the interface, releasing strain energy and resulting in a lower friction coefficient. Figure 9b shows that when the scratching depth increases, more atoms around the abrasive particles are squeezed out and deposited on the surface of the FeCrNi MPEA coating, resulting in more serious wear [36,38]. Because the larger scratching depth induces more lattice defects, the hardening behavior of the FeCrNi MPEA coating is more significantly impacted at the larger scratching depth. Thus, a larger coefficient of friction is required for the scratching process to maintain a constant speed.
Figure 10a–d clearly shows the dislocation distribution and dislocation density inside the FeCrNi MPEA at different scratching depths during the nanoscratch process. Here, the Shockley, Perfect, Stair-rod lines are presented in green, blue, and pink, respectively. In order to observe the dislocation evolution, HCP structure atoms and other structure atoms are removed. As shown in Figure 10a–d, at a scratching depth of 0.5 nm, a large number of Shockley dislocations accumulate in the subsurface and then slip and annihilate on the free surface (Figure 11). At a scratching depth of 1.0 nm, it can be found that the dislocation around the particle is reduced. The dislocation distribution results in fewer atoms deposited before abrasive particles (Figure 11). Obviously, as the scratching depth continues to increase, many structural defects form [36,37,38,39]. Therefore, in the process of scratching, the scratching depth has an important effect on the internal defect structure of the matrix. To better analyze the dislocation distribution during scratching, Figure 10e shows the evolution of dislocation length at different scratching depths. The dislocation length does not increase significantly in the first scratching stage, but oscillates horizontally or decreases slightly in the later scratching stage. As the scratching depth continues to increase, the dislocation length increases significantly [39,40]. This trend indicates that subsurface damage and internal defects increase with the increase of depth.

3.3. Role of Operating Temperature

Figure 12a shows the scratching force-displacement curve of FeCrNi MPEA during the nanoscratching process at different temperatures. As can be seen from Figure 12, with the increase in scratching distance, due to the increase in the atom number around the abrasive particle, the scratching force increases rapidly, and then fluctuates around a constant value [37,38,39,40]. The scratching force oscillates horizontally within a certain range after a scratching distance of 1 nm. In addition, Figure 12b shows the variation trend of the scratching force of FeCrNi MPEA coating at different temperatures. The high temperature causes a decrease in the scratching force.
In order to study the effect of temperature on the surface quality of the scratched material, Figure 13 shows the surface morphology of FeCrNi MPEA at different temperatures after nanoscratching. As can be seen from Figure 13, the scratching-induced grooves are formed on the surface of the sample after nanoscratching, accompanied by surface accumulation and uniform wear at high temperature. In addition, the wear trajectory is wider and deeper [38,39,40]. The cyclic friction can cause the temperature of the wear surface to rise, which leads to surface oxidation and cracking. Thus, the wear surface appears to have lamellar morphology and is accompanied by local micro-cracks. With the change of test temperature, there is no obvious difference in the morphology of wear marks. The amount of surface accumulation is basically the same at different temperatures. The accumulation can significantly increase the resistance of abrasive particles.
Figure 14a describes the evolution of the friction coefficient of the FeCrNi MPEA coating with scratching distance when the scratching temperature is 300 K, 500 K, 700 K, 900 K, and 1100 K. At the initial stage, the friction coefficient decreases sharply with the increase of scratching displacement. When the scratching distance reaches 6 nm, the average friction coefficient of FeCrNi MPEA increases slightly [39,40]. As can be seen from Figure 14b, the friction coefficient at increasing temperature is negatively correlated with the temperature, and rapidly decreases to 0.8 at 900 K. The reason why the friction coefficient decreases with the increase of temperature may be because the wear surface forms a dense oxide layer and the binding force of the contact interface is low [43,44]. The oxide layer leads to the large contribution of a weak van der Waals force compared with the strong interatomic binding force, and causes a decrease in the friction coefficient [44,45]. The high temperature can control the adhesion behavior and reduce the adhesion [46], which is also a reason why the high temperature reduced the friction coefficient.
Figure 15a–d shows the dislocation evolution and the dislocation distribution in the sample at different scratching temperatures. When the abrasive particles are pressed into the sample, plastic deformation occurs, and a large number of dislocations are released (Figure 16). In the scratching stage, the dislocation nucleates around the abrasive particle. As the temperature increases, the dislocation moves toward the interior of the sample along the scratching direction [35,36,37,38,39,40]. The temperature causes the dislocation to spread later (Figure 16), which is the process of scratching repair. When the dislocation spreads in the scratched area, the recovery effect of the material is significantly improved, and the dislocation density in the post-scratched area is significantly reduced. Figure 15e shows that the length of the dislocation changes dramatically with increasing temperatures after the scratching stage. High density dislocation entanglement can produce a strengthening effect [38,39,40,41], resulting in a moderate increase in tangential forces. When the temperature rises from 300 K to 700 K, the length of the dislocation first increases slowly and then rapidly. When the temperature reaches 700 K, the dislocation length reaches the maximum. When the temperature continues to rise, the dislocation length is obviously smaller than other conditions, which reduces the tangential force. Because the length of the dislocation formed by plastic deformation of the sample after the end of the scratching stage is obviously the least, the tangential force here is also the least.

3.4. Experimental Verification

3.4.1. Phase Constitution and Microstructure

The XRD pattern shows that the FeCrNi MPEA has the FCC structure (Figure 17a). The hot extrusion does not cause the phase transformation of the FeCrNi MPEA. This means that FeCrNi has excellent thermal stability [14,15,16], especially at high temperature conditions. The result of inverse pole figure (IPF) maps shows that the coarse grain without obvious textures is uniformly distributed in FeCrNi MPEA (Figure 17b). The equiaxed grain has an average diameter of 12 µm. The annealing twins are formed after hot extrusion in the grain interior, due to the low stacking fault energy. The EDS result of Fe, Cr and Ni elements is presented in Figure 17c. It reveals the uniform element distribution of FeCrNi MPEA, which is beneficial for suppressing high-temperature segregation

3.4.2. Tribological Performance

For a temperature range between 300 K and 1100 K, Figure 18a shows the relationship between friction coefficient and time in FeCrNi MPEA. The friction coefficient is calculated from the experimental data. At a temperature of 300 K, the friction coefficient of FeCrNi MPEA gradually increases until it reaches a steady state, owing to the fracture of the wear surface and the settlement of the tool. A high friction coefficient is about 0.67. With the increase in temperature, the friction coefficient of FeCrNi MPEA greatly reduces with the increasing test temperature (Figure 18b), which is consistent with the MD simulation results (Figure 14b).
Figure 19 shows the friction morphology of FeCrNi MPEA wear tracks at test temperatures of 300 K, 900 K, and 1100 K. The FeCrNi MPEA wears uniformly at 300 K, and the wear track is wide (Figure 19a). Under the effects of cyclic stresses, a large number of micro cracks, wear debris, and delamination can be found on the wear surface, and abrasive wear is the dominant wear mechanism (Figure 19b). With the increase in temperature, the wear resistance of FeCrNi MPEA gradually improved [47,48]. It can be seen from Figure 19c,e that the width of the wear track reduces significantly at 900 K and 1100 K, which is consistent with the changing trend of the friction coefficient. Figure 19d,f shows the wear tracks at the high temperatures of 900 K and 1100 K. Different from that at 300 K, the wear surface at high temperature is flooded with scattered lamellar oxides, which means that the wear mechanism gradually changes from abrasive wear at room temperature to oxidation and delamination wear at high temperature. Since the oxide layer has low shear strength and high hardness, the oxide layer on the high-temperature wear surface can prevent direct contact between friction pairs, helping to reduce the wear on the sample surface and improve the wear performance.

4. Conclusions

In the current work, the microstructural and tribological characteristics of FeCrNi MPEAs were investigated by MD simulations combined with high temperature tribological experiments from nanoscale and microscale perspectives. Using MD simulations, the effects of scratching speed, depth, and temperature on the tribological properties of FeCrNi MPEAs were systematically investigated, which determined the quantitative relationships of the average scratching force (friction coefficient) and scratching speed (depth). Combined with experimental results, the temperature dependent wear mechanism in FeCrNi multicomponent alloys was revealed, and the origin of high-temperature wear resistance was elucidated. The conclusions are as follows:
(1)
The results reveal that the high scratching speed leads to larger scratching force because more atomic bonds have to be broken at the same time. The high scratching depth causes an increase in the friction coefficient due to the larger resistance to tool movement. The thermal softening results in a slightly increased tangential force at high speed, and strong strain rate hardening induces an increase in normal force. This trend reduces the friction coefficient. The friction coefficient decreases with the increase in scratch speed, and increases with the increase of scratch depth.
(2)
The effect of temperature on the tribological behavior of MPEA, the quantitative relationship between friction coefficient and temperature, and the mechanism of oxidation behavior on friction coefficient and atomic adhesion is investigated. The friction coefficient monotonously decreases with the increasing test temperatures. High Cr content results in the formation of a compact oxide layer at intermediate temperatures to improve tribological performance due to the oxidative delamination wear. The experiment shows that the oxide layer with low shear strength and high hardness prevents direct contact between friction pairs, reduces the wear, and improves the wear performance. These findings give an atomistic insight into scratching-induced damage mechanisms, and beneficially develop new MPEA coatings with superior wear performance.

Author Contributions

Conceptualization, X.L.; methodology, J.P.; software, J.P. and Y.C.; validation, J.P. and A.F.; formal analysis, Y.C.; investigation, Y.C. and A.F.; data curation, Y.C. and A.F.; writing—original draft preparation, Y.C. and A.F.; writing—review and editing, X.L. and J.P.; visualization, J.P.; supervision, X.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (52075559), Hunan Provincial Key Research Development Program (2022GK2030), and Hunan Provincial Natural Science Foundation (2021JJ31161).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Pan, Q.; Yang, M.; Feng, R.; Chuang, A.C.; An, K.; Liaw, P.K.; Wu, X.; Tao, N.; Lu, L. Atomic faulting induced exceptional cryogenic strain hardening in gradient cell–structured alloy. Science 2023, 382, 185–190. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, Y.; Chen, T.; Tan, L.; Poplawsky, J.D.; An, K.; Wang, Y.; George, E.P. Bifunctional nanoprecipitates strengthen and ductilize a medium-entropy alloy. Nature 2021, 595, 245–249. [Google Scholar] [CrossRef]
  3. Krishna, S.A.; Noble, N.; Radhika, N.; Saleh, B. A comprehensive review on advances in high entropy alloys: Fabrication and surface modification methods, properties, applications, and future prospects. J. Manuf. Process. 2024, 109, 583–606. [Google Scholar] [CrossRef]
  4. Zhang, P.; Li, Z.; Liu, H.; Zhang, Y.; Li, H.; Shi, C.; Liu, P. Recent progress on the microstructure and properties of high entropy alloy coatings prepared by laser processing technology: A review. J. Manuf. Process. 2022, 76, 397–411. [Google Scholar] [CrossRef]
  5. Peng, J.; Xie, B.; Zeng, X.; Fang, Q.; Liu, B.; Liaw, P.K.; Li, J. Vacancy dependent mechanical behaviors of high-entropy alloy. Int. J. Mech. Sci. 2022, 218, 107065. [Google Scholar] [CrossRef]
  6. Li, J.; Fu, X.; Feng, H.; Liu, B.; Liaw, P.K.; Fang, Q. Evaluating the solid solution, local chemical ordering, and precipitation strengthening contributions in multi-principal-element alloys. J. Alloys Compd. 2023, 938, 168521. [Google Scholar] [CrossRef]
  7. Schneider, M.; Laplanche, G. Effects of temperature on mechanical properties and deformation mechanisms of the equiatomic CrFeNi medium-entropy alloy. Acta Mater. 2021, 204, 116470. [Google Scholar] [CrossRef]
  8. Fu, A.; Liu, B.; Lu, W.; Liu, B.; Li, J. A novel supersaturated medium entropy alloy with superior tensile properties and corrosion resistance. Scr. Mater. 2020, 186, 381–386. [Google Scholar] [CrossRef]
  9. Duan, H.; Liu, B.; Fu, A.; Liu, Y. Segregation enabled outstanding combination of mechanical and corrosion properties in a FeCrNi medium entropy alloy manufactured by selective laser melting. J. Mater. Sci. Tech. 2022, 99, 207–214. [Google Scholar] [CrossRef]
  10. Tang, L.; Jiang, F.Q.; Wróbel, J.S.; Liu, B.; Kabra, S.; Duan, R.X.; Cai, B. In situ neutron diffraction unravels deformation mechanisms of a strong and ductile FeCrNi medium entropy alloy. J. Mater. Sci. Tech. 2022, 116, 103–120. [Google Scholar] [CrossRef]
  11. Fu, A.; Liu, B.; Cao, Y.; Wang, J.; Liao, T.; Liu, Y. A novel cobalt-free oxide dispersion strengthened medium-entropy alloy with outstanding mechanical properties and irradiation resistance. J. Mater. Sci. Tech. 2023, 152, 190–200. [Google Scholar] [CrossRef]
  12. Fan, P.; Katiyar, N.K.; Goel, S.; He, Y.; Geng, Y.; Yan, Y. Oblique nanomachining of gallium arsenide explained using AFM experiments and MD simulations. J. Manuf. Process. 2023, 90, 125–138. [Google Scholar] [CrossRef]
  13. Yang, L.Y.; Wu, C.Y.; Peng, D.K.; Shen, Y.J.; Fan, J.L.; Gong, H.R. Effect of Cu on strain-stress relationship of W bicrystal models with grain boundary from molecular dynamics simulation. Vacuum 2023, 208, 111712. [Google Scholar] [CrossRef]
  14. Sato, Y.; Shinzato, S.; Ohmura, T.; Hatano, T.; Ogata, S. Unique universal scaling in nanoindentation pop-ins. Nat. Commun. 2020, 11, 4177. [Google Scholar] [CrossRef] [PubMed]
  15. Li, J.; Yang, S.; Dong, L.; Zhang, J.; Zheng, Z.; Liu, J. Effect of crystal orientation on the nanoindentation deformation behavior of TiN coating based on molecular dynamics. Surf. Coat. Technol. 2023, 467, 129721. [Google Scholar] [CrossRef]
  16. Yin, Z.; Zhu, P.; Li, B. Study of nanoscale wear of SiC/Al nanocomposites using molecular dynamics simulations. Tribolo. Lett. 2021, 69, 1–17. [Google Scholar] [CrossRef]
  17. Zhou, A.; Liu, X.B.; Wang, Q.; Zhang, S.Y.; Meng, Y. Investigation of nano-tribological behaviors and deformation mechanisms of Cu-Ni alloy by molecular dynamics simulation. Tribo. Int. 2023, 180, 108258. [Google Scholar] [CrossRef]
  18. Li, J.; Dong, L.; Dong, X.; Zhao, W.; Liu, J.; Xiong, J. Study on wear behavior of FeNiCrCoCu high entropy alloy coating on Cu substrate based on molecular dynamics. Appl. Sur. Sci. 2021, 570, 151236. [Google Scholar] [CrossRef]
  19. Yang, C.M.; Liu, X.B.; Zhu, Z.X.; Zhou, A.; Zhou, H.B.; Zhang, S.H. Study of tribological property of laser-cladded FeCoCrNiMnx high-entropy alloy coatings via experiment and molecular dynamics simulation. Tribol. Int. 2024, 191, 109106. [Google Scholar] [CrossRef]
  20. Zhang, J.; Deng, G.; Li, W.; Li, S.; Yan, Y.; Liu, X.; Gao, L. Grain size and scratching depth dependent tribological characteristics of CrCoNi medium-entropy alloy coatings: A molecular dynamics simulation study. Surf. Coat. Technol. 2023, 468, 129772. [Google Scholar] [CrossRef]
  21. Nguyen, H.G.; Fang, T.H.; Doan, D.Q. Cyclic plasticity and deformation mechanism of AlCrCuFeNi high entropy alloy. J. Alloys Compd. 2023, 940, 168838. [Google Scholar] [CrossRef]
  22. Doan, D.Q.; Nguyen, V.H.; Tran, T.V.; Hoang, M.T. Atomic-scale analysis of mechanical and wear characteristics of AlCoCrFeNi high entropy alloy coating on Ni substrate. J. Manuf. Process. 2023, 85, 1010–1023. [Google Scholar] [CrossRef]
  23. Feng, H.; Cui, S.; Chen, H.; Song, X.; Fang, Q.; Li, J. A molecular dynamics investigation into deformation mechanism of nanotwinned Cu/high entropy alloy FeCoCrNi nanolaminates. Surf. Coat. Technol. 2020, 401, 126325. [Google Scholar] [CrossRef]
  24. Gao, S.; Wang, H.; Huang, H.; Kang, R. Molecular simulation of the plastic deformation and crack formation in single grit grinding of 4H-SiC single crystal. Int. J. Mech. Sci. 2023, 247, 108147. [Google Scholar] [CrossRef]
  25. Wu, Z.; Dong, B.; Peng, G.; Luo, W.; Shan, X.; Nie, L. Novel insight into the self-lubricating mechanism of Cu–Pb bearing alloy during sliding friction with the guidance of molecular dynamics. J. Mater. Res. Technol. 2023, 24, 3157–3163. [Google Scholar] [CrossRef]
  26. Seki, T.; Futazuka, T.; Morishige, N.; Matsubara, R.; Ikuhara, Y.; Shibata, N. Incommensurate grain-boundary atomic structure. Nat. Commun. 2023, 14, 7806. [Google Scholar] [CrossRef] [PubMed]
  27. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef]
  28. Qiu, Y.; Qi, Y.; Zheng, H.; He, T.; Feng, M. Atomistic simulation of nanoindentation response of dual-phase nanocrystalline CoCrFeMnNi high-entropy alloy. J. Appl. Phys. 2021, 130, 12. [Google Scholar] [CrossRef]
  29. Lee, S.; Vaid, A.; Im, J.; Kim, B.; Prakash, A.; Guénolé, J.; Oh, S.H. In-situ observation of the initiation of plasticity by nucleation of prismatic dislocation loops. Nat. Commun. 2020, 11, 2367. [Google Scholar] [CrossRef]
  30. Sato, Y.; Shinzato, S.; Ohmura, T.; Ogata, S. Atomistic prediction of the temperature-and loading-rate-dependent first pop-in load in nanoindentation. Int. J. Plast. 2019, 121, 280–292. [Google Scholar] [CrossRef]
  31. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO–the open visualization tool. Model. Simul. Mater. Sci. Eng. 2010, 18, 015012. [Google Scholar] [CrossRef]
  32. Stukowski, A.; Bulatov, V.V.; Arsenlis, A. Automated identification and indexing of dislocations in crystal interfaces. Model. Simul. Mater. Sci. Eng. 2012, 20, 085007. [Google Scholar] [CrossRef]
  33. Stukowski, A.; Albe, K. Extracting dislocations and non-dislocation crystal defects from atomistic simulation data. Model. Simul. Mater. Sci. Eng. 2010, 18, 085001. [Google Scholar] [CrossRef]
  34. Farkas, D.; Caro, A. Model interatomic potentials for Fe–Ni–Cr–Co–Al high-entropy alloys. J. Mater. Res. 2020, 35, 3031–3040. [Google Scholar] [CrossRef]
  35. Tan, F.; Li, J.; Feng, H.; Fang, Q.; Jiang, C.; Liu, Y.; Liaw, P.K. Entropy-induced transition on grain-boundary migration in multi-principal element alloys. Scr. Mater. 2021, 194, 113668. [Google Scholar] [CrossRef]
  36. Lu, Y.S.; Chang, M.P.; Fang, T.H. Phase transformation and microstructure evolution of nanoimprinted NiCoCr medium entropy alloys. J. Alloys Compd. 2022, 892, 162138. [Google Scholar] [CrossRef]
  37. Hao, Z.P.; Cui, R.R.; Fan, Y.H.; Lin, J.Q. Diffusion mechanism of tools and simulation in nanoscale cutting the Ni–Fe–Cr series of Nickel-based superalloy. Int. J. Mech. Sci. 2019, 150, 625–636. [Google Scholar] [CrossRef]
  38. Doan, D.Q.; Fang, T.H.; Chen, T.H. Influences of grain size and temperature on tribological characteristics of CuAlNi alloys under nanoindentation and nanoscratch. Int. J. Mech. Sci. 2020, 185, 105865. [Google Scholar] [CrossRef]
  39. Stephan, S.; Schmitt, S.; Hasse, H.; Urbassek, H.M. Molecular dynamics simulation of the stribeck curve: Boundary lubrication, mixed lubrication, and hydrodynamic lubrication on the atomistic level. Friction 2023, 11, 2342–2366. [Google Scholar] [CrossRef]
  40. Li, J.; Liu, B.; Luo, H.; Fang, Q.; Liu, Y.; Liu, Y. A molecular dynamics investigation into plastic deformation mechanism of nanocrystalline copper for different nanoscratching rates. Comput. Mater. Sci. 2016, 118, 66–76. [Google Scholar] [CrossRef]
  41. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  42. Sneddon, I.N. The relation between load and penetration in the axisymmetric Boussinesq problem for a punch of arbitrary profile. Int. J. Eng. Sci. 1965, 3, 47–57. [Google Scholar] [CrossRef]
  43. Geng, Y.; Chen, J.; Tan, H.; Cheng, J.; Zhu, S.; Yang, J. Tribological performances of CoCrFeNiAl high entropy alloy matrix solid-lubricating composites over a wide temperature range. Tribol. Int. 2021, 157, 106912. [Google Scholar] [CrossRef]
  44. Du, L.M.; Lan, L.W.; Zhu, S.; Yang, H.J.; Shi, X.H.; Liaw, P.K.; Qiao, J.W. Effects of temperature on the tribological behavior of Al0. 25CoCrFeNi high-entropy alloy. J. Mater. Sci. Technol. 2019, 35, 917–925. [Google Scholar] [CrossRef]
  45. Wang, Y.; Yang, Y.; Yang, H.; Zhang, M.; Ma, S.; Qiao, J. Microstructure and wear properties of nitrided AlCoCrFeNi high-entropy alloy. Mater. Chem. Phys. 2018, 210, 233–239. [Google Scholar] [CrossRef]
  46. Gåård, A.; Krakhmalev, P.; Bergström, J.; Grytzelius, J.H.; Zhang, H.M. Experimental study of the relationship between temperature and adhesive forces for low-alloyed steel, stainless steel, and titanium using atomic force microscopy in ultrahigh vacuum. J. Appl. Phys. 2008, 103, 12. [Google Scholar] [CrossRef]
  47. Joseph, J.; Haghdadi, N.; Annasamy, M.; Kada, S.; Hodgson, P.D.; Barnett, M.R.; Fabijanic, D.M. On the enhanced wear resistance of CoCrFeMnNi high entropy alloy at intermediate temperature. Scr. Mater. 2020, 186, 230–235. [Google Scholar] [CrossRef]
  48. Tian, X.; Zhang, Y.; Li, J. Investigation on tribological behavior of advanced high strength steels: Influence of hot stamping process parameters. Trib. Lett. 2012, 45, 489–495. [Google Scholar] [CrossRef]
Figure 1. The atomic model of the nanoscratching process in the FeCrNi MPEA coating on the Ni substrate. Symmetry 16 01391 i001 C, Symmetry 16 01391 i002 Fe, Symmetry 16 01391 i003 Cr, and Symmetry 16 01391 i004 Ni.
Figure 1. The atomic model of the nanoscratching process in the FeCrNi MPEA coating on the Ni substrate. Symmetry 16 01391 i001 C, Symmetry 16 01391 i002 Fe, Symmetry 16 01391 i003 Cr, and Symmetry 16 01391 i004 Ni.
Symmetry 16 01391 g001
Figure 2. The relationship between friction force and distance (a), average force and speed (b).
Figure 2. The relationship between friction force and distance (a), average force and speed (b).
Symmetry 16 01391 g002
Figure 3. Variation of surface morphology at scratching distance of 35 nm for different speeds: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d) in the FeCrNi MPEA coating.
Figure 3. Variation of surface morphology at scratching distance of 35 nm for different speeds: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d) in the FeCrNi MPEA coating.
Symmetry 16 01391 g003
Figure 4. The relationship between friction coefficient and distance (a), friction coefficient and speed (b).
Figure 4. The relationship between friction coefficient and distance (a), friction coefficient and speed (b).
Symmetry 16 01391 g004
Figure 5. Evaluation of dislocation structure with the increase speed at scratching distance of 35 nm: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d). The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The relationship between dislocation length and speed.
Figure 5. Evaluation of dislocation structure with the increase speed at scratching distance of 35 nm: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d). The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The relationship between dislocation length and speed.
Symmetry 16 01391 g005
Figure 6. Evaluation of the microstructure with increased speed at a scratching distance of 35 nm: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d). Symmetry 16 01391 i005 FCC structure, Symmetry 16 01391 i006 dislocation core structure, and Symmetry 16 01391 i007 BCC structure.
Figure 6. Evaluation of the microstructure with increased speed at a scratching distance of 35 nm: 25 m/s (a), 50 m/s (b), 100 m/s (c), and 200 m/s (d). Symmetry 16 01391 i005 FCC structure, Symmetry 16 01391 i006 dislocation core structure, and Symmetry 16 01391 i007 BCC structure.
Symmetry 16 01391 g006
Figure 7. The relationship between friction force and distance (a), average force and depth (b).
Figure 7. The relationship between friction force and distance (a), average force and depth (b).
Symmetry 16 01391 g007
Figure 8. Variation of surface morphology at a scratching distance of 35 nm for different depths: 0.5 nm (a), 1.0 nm (b), 1.5 nm (c), and 2 nm (d) in the FeCrNi MPEA coating.
Figure 8. Variation of surface morphology at a scratching distance of 35 nm for different depths: 0.5 nm (a), 1.0 nm (b), 1.5 nm (c), and 2 nm (d) in the FeCrNi MPEA coating.
Symmetry 16 01391 g008
Figure 9. The relationship between friction coefficient and distance (a), average friction coefficient and depth (b).
Figure 9. The relationship between friction coefficient and distance (a), average friction coefficient and depth (b).
Symmetry 16 01391 g009
Figure 10. Evaluation of dislocation structure at the distance of 35 nm for different depths: 0.5 nm (a), 1.0 nm (b), 1.5 nm (c), and 2 nm (d) The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The dislocation length with the increase in depth.
Figure 10. Evaluation of dislocation structure at the distance of 35 nm for different depths: 0.5 nm (a), 1.0 nm (b), 1.5 nm (c), and 2 nm (d) The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The dislocation length with the increase in depth.
Symmetry 16 01391 g010aSymmetry 16 01391 g010b
Figure 11. Evaluation of the microstructure at a scratching distance of 35 nm for different depths: 0.5 nm (a), 1.5 nm (b), and 2 nm (c). Symmetry 16 01391 i005 FCC structure, Symmetry 16 01391 i006 dislocation core structure, and Symmetry 16 01391 i007 BCC structure.
Figure 11. Evaluation of the microstructure at a scratching distance of 35 nm for different depths: 0.5 nm (a), 1.5 nm (b), and 2 nm (c). Symmetry 16 01391 i005 FCC structure, Symmetry 16 01391 i006 dislocation core structure, and Symmetry 16 01391 i007 BCC structure.
Symmetry 16 01391 g011
Figure 12. The relationship between friction force and distance (a), average force and temperature (b).
Figure 12. The relationship between friction force and distance (a), average force and temperature (b).
Symmetry 16 01391 g012
Figure 13. Variation of surface morphology at a scratching distance of 35 nm for different temperatures 300 K (a), 500 K (b), 900 K (c), and 1100 K (d) in the FeCrNi MPEA coating.
Figure 13. Variation of surface morphology at a scratching distance of 35 nm for different temperatures 300 K (a), 500 K (b), 900 K (c), and 1100 K (d) in the FeCrNi MPEA coating.
Symmetry 16 01391 g013
Figure 14. The relationship between friction coefficient and distance (a), average friction coefficient and temperature (b).
Figure 14. The relationship between friction coefficient and distance (a), average friction coefficient and temperature (b).
Symmetry 16 01391 g014
Figure 15. Evaluation of subsurface structure with the increase in temperature of 300 K (a), 500 K (b), 900 K (c), and 1100 K (d) at a scratching distance of 35 nm.The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The dislocation length with the increase in temperature.
Figure 15. Evaluation of subsurface structure with the increase in temperature of 300 K (a), 500 K (b), 900 K (c), and 1100 K (d) at a scratching distance of 35 nm.The blue line is perfect dislocation, the green line is Shockley partial dislocation, the red line is other dislocation, the sky-blue line is Frank partial dislocation, the pink line is stair-rod dislocation, and the yellow line is Hirth dislocation. (e) The dislocation length with the increase in temperature.
Symmetry 16 01391 g015aSymmetry 16 01391 g015b
Figure 16. Evaluation of microstructure with the increase in temperature at a scratching distance of 35 nm: 900 K (a), and 1100 K (b).
Figure 16. Evaluation of microstructure with the increase in temperature at a scratching distance of 35 nm: 900 K (a), and 1100 K (b).
Symmetry 16 01391 g016
Figure 17. Initial microstructure of the specimen. (a) X-ray diffractometer pattern, (b) inverse pole figure map, and (c) elemental distribution of the FeCrNi.
Figure 17. Initial microstructure of the specimen. (a) X-ray diffractometer pattern, (b) inverse pole figure map, and (c) elemental distribution of the FeCrNi.
Symmetry 16 01391 g017
Figure 18. (a) Friction coefficient with increasing time for different test temperatures in FeCrNi MPEA. (b) The average friction coefficient with increasing temperatures.
Figure 18. (a) Friction coefficient with increasing time for different test temperatures in FeCrNi MPEA. (b) The average friction coefficient with increasing temperatures.
Symmetry 16 01391 g018
Figure 19. Wear surface in the FeCrNi MPEA at temperatures of (a) 300 K, (c) 900 K and (e) 1100 K. The SE images for the corresponding surface damage are plotted at temperatures of (b) 300 K, (d) 900 K and (f) 1100 K.
Figure 19. Wear surface in the FeCrNi MPEA at temperatures of (a) 300 K, (c) 900 K and (e) 1100 K. The SE images for the corresponding surface damage are plotted at temperatures of (b) 300 K, (d) 900 K and (f) 1100 K.
Symmetry 16 01391 g019
Table 1. Computational parameters used in the MD simulations.
Table 1. Computational parameters used in the MD simulations.
MaterialFeCrNi CoatingNi SubstrateTool
Dimension388 Å × 180 Å × 36 Å388 Å × 180 Å × 144 ÅRadius 40 Å
Atom count229,986875,96447,262
Time step1 fs
Initial temperature300 K, 500 K, 700 K, 900 K, and 1100 K
Cutting depth0.5 nm, 1.0 nm, 1.5 nm, and 2 nm
Scratching velocity25 m/s, 50 m/s, 100 m/s, and 200 m/s
Scratching distance35 nm
Scratching direction[100] direction on (1 0 0) surface
Table 2. Morse potential parameters among different atoms.
Table 2. Morse potential parameters among different atoms.
ParameterD (eV) α   ( 1 / A n ) r 0   ( Å )
C-Fe1.00571.97182.6493
C-Cr1.03422.06362.6176
C-Ni1.00391.98752.6199
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, Y.; Liu, X.; Fu, A.; Peng, J. Unraveling the Subsurface Damage and Material Removal Mechanism of Multi-Principal-Element Alloy FeCrNi Coatings During the Scratching Process. Symmetry 2024, 16, 1391. https://doi.org/10.3390/sym16101391

AMA Style

Chen Y, Liu X, Fu A, Peng J. Unraveling the Subsurface Damage and Material Removal Mechanism of Multi-Principal-Element Alloy FeCrNi Coatings During the Scratching Process. Symmetry. 2024; 16(10):1391. https://doi.org/10.3390/sym16101391

Chicago/Turabian Style

Chen, Yuan, Xiubo Liu, Ao Fu, and Jing Peng. 2024. "Unraveling the Subsurface Damage and Material Removal Mechanism of Multi-Principal-Element Alloy FeCrNi Coatings During the Scratching Process" Symmetry 16, no. 10: 1391. https://doi.org/10.3390/sym16101391

APA Style

Chen, Y., Liu, X., Fu, A., & Peng, J. (2024). Unraveling the Subsurface Damage and Material Removal Mechanism of Multi-Principal-Element Alloy FeCrNi Coatings During the Scratching Process. Symmetry, 16(10), 1391. https://doi.org/10.3390/sym16101391

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop
  NODES
3d 1
Association 2
coding 2
Experiments 6
games 2
games 2
Idea 2
idea 2
innovation 2
Interesting 4
Intern 32
iOS 5
Javascript 2
languages 2
mac 26
Note 10
OOP 1
os 175
text 6
twitter 1
visual 3
web 2