Next Article in Journal
Evolution of Microstructure, Mechanical Properties and Residual Stress of a Cold Rolled Invar Sheet Due to Heat Treatment
Next Article in Special Issue
Regularities of Friction Stir Processing Hardening of Aluminum Alloy Products Made by Wire-Feed Electron Beam Additive Manufacturing
Previous Article in Journal
Coupled Modeling Approach for Laser Shock Peening of AA2198-T3: From Plasma and Shock Wave Simulation to Residual Stress Prediction
Previous Article in Special Issue
Friction Stir Processing of Additively Manufactured Ti-6Al-4V Alloy: Structure Modification and Mechanical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructural Evolution of AA5154 Layers Intermixed with Mo Powder during Electron Beam Wire-Feed Additive Manufacturing (EBAM)

Institute of Strength Physics and Materials Science, Siberian Branch of Russian Academy of Sciences, 634055 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Submission received: 2 December 2021 / Revised: 20 December 2021 / Accepted: 29 December 2021 / Published: 6 January 2022
(This article belongs to the Special Issue Light Metals and Their Composites)

Abstract

:
AA5154 aluminum alloy wall was built using EBAM where the wall’s top layers were alloyed by depositing and then remelting a Mo powder-bed with simultaneous transfer of aluminum alloy from the AA5154 wire. The powder-beds with different concentrations of Mo such as 0.3, 0.6, 0.9 and 1.2 g/layer were used to obtain composite AA5154/Mo samples. All samples were characterized by inhomogeneous structures composed of as-deposited AA5154 matrix with coarse unreacted Mo articles and intermetallic compounds (IMC) such as Al12Mo, Al5Mo, Al8Mo3, Al18Mg3Mo2 which formed in the vicinity of these Mo particles. The IMC content increased with the Mo powder-bed concentrations. The AA5154 matrix grains away from the Mo particles contained Al-Fe grain boundary precipitates. Mo-rich regions in the 0.3, 0.6, 0.9 and 1.2 g/layer Mo samples had maximum microhardness at the level of 2300, 2600, 11,500 and 9000 GPa, respectively. Sliding pin-on-steel disk test showed that wear of A5154/Mo composite reduced as compared to that of as-deposited AA5154 due to composite structure, higher microhardness as a well as tribooxidation of Al/Mo IMCs and generation of mechanically mixed layers containing low shear strength Mo8O23 and Al2(MoO4)3 oxides.

1. Introduction

The development of new high-strength and temperature-resistant metallic materials is an urgent task faced by specialists working in space, aircraft, and automobile industries. Nowadays, nickel-based, high-temperature strength alloys are widely applied for fabricating gas and steam turbine blades or exhaustive gas heat shields [1]. However, their working temperatures are less than those of refractory metal aluminides [2,3,4,5]. Molibdenum aluminides are attracting more and more attention of the researchers since they possess a combination of high melting temperature and thermal stability with high mechanical characteristics [6,7,8]. Their resistance to oxidizing is also very high because of the formation of passivation Al2O3 film [9].
Due to its low solubility in aluminum, molybdenum forms a number of hard intermetallic AlxMoy compounds (IMCs) that might have been used for reinforcing the aluminum alloy matrix if properly dispersed and distributed in it. Such a composite structure can be used for improving both mechanical and functional characteristics of the alloy.
Despite there has always been some constant interest in obtaining and studying the AlxMoy IMCs, their mechanical characteristics are still understudied and there are discrepancies in the experimental results reported by different researchers. For instance, Yihan Bian et al. [10] showed that MoAl12 not only had a high melting point but also retained its strength under high-temperature loading. It was reported also [11] that MoAl12, MoAl5, MoAl4, Mo3Al8 and Mo3Al IMCs possess thermodynamic and dynamic stabilities, which are depended on the Mo content so that Mo3Al8 is the most stable IMC as compared to other ones, and bulk modulus of Al-Mo alloys tends to grow with the Mo content [11]. Mingliang Wang et al. [7] established that the calculated bulk elasticity and shear moduli of MoAl5 were 115.7 GPa and 95.0 GPa, respectively. Eumann M. et al. [6] demonstrated that Mo3Al had a yield stress of 1600 MPa and high brittleness. Rikiya Nino et al. [8] showed that coarse lamellar Mo3Al–Mo3Al8 served to suppress crack nucleation in Vickers indentation at loads as high as 9.8 N, while fine lamellar structures fractured during indentation at 2.94 N load. Such a result seems to be surprising since common opinion is that the finer the structure the higher is the fracture toughness.
The processes used for synthesizing the AlxMoy IMCs mainly are as follows: laser powder-bed fusion [12], hot dipping of Mo rods into a commercial Al-12 wt.% Si alloy melt [13], stir casting [14,15], reactive hot-pressing [16], magnetron sputtering [17], friction stir processing [18,19], etc.
Additive manufacturing (AM) is an alternative to traditionally used methods of fabricating materials and articles. Small-scale AM machines equipped with laser, plasma arc, or electron beam energy sources are produced commercially to enable production of machine components directly at place of use from powders or wires. AM has potential for fabricating figurine-shaped components from in-situ formed metal matrix composites, functionally graded or polymetallic materials as depending on the source materials used. Progress achieved in the field of additive manufacturing and materials science allowed using the AM in machine building, aircraft or space applications industries. One of the most important examples is using the AM for fabricating a jet engine collector for Airbus [20].
Using the AM methods for in-situ preparation of metallic matrix composites or functionally graded materials requires resolving problems with respect to materials compatibility, reactive diffusion zone characteristics, improving as-cast structure characteristics, avoiding porosity, shrinkage pores, etc. Some of the above problems may be resolved with properly selecting the AM method.
Additive manufacturing methods used for making metallic components include wire arc additive manufacturing (WAAM) [21,22,23,24,25], selective laser melting (SLM) [26,27,28] and electron beam additive manufacturing (EBAM) [29,30,31,32]. These methods have their own advantages and disadvantages but the majority of experiments on making aluminum alloy matrix composites were carried out using these three.
For instance, Gasper et al. [33] used a combined powder/wire feed direct energy deposition for in-situ preparation of titanium aluminides. SLM on AlSi10Mg/SiC allowed obtaining AlSi10Mg/SiC composites [33] characterized by reduced friction and wear. Intermixing TiC particles with aluminum alloy during the WAAM resulted in grain refining and improved mechanical characteristics of the composites obtained [34,35]. Applicability of inoculating particles of oxides, carbides, metallic for improving the microstructure and mechanical characteristics of alloys was discussed in a number of review articles as follows: [36,37,38,39].
A combination of powder-bed and wire-feed EBAM processes allowed fabricating boron carbide/aluminum bronze composite [40]. Such a process can also be used for preparation of an Al/Mo composite reinforced by in-situ reactive diffusion formed IMCs. Such an approach was never used earlier in this system and can be of interest from the viewpoint of materials science.
The objective of this work is, therefore, to prepare Al/Mo composite layers using EBAM and study their microstructural and mechanical characteristics depending on the Mo concentration.

2. Materials and Methods

Electron beam deposition was carried out using an aluminum alloy of ⌀1.2 mm wire of composition, wt.%: 3.8 ± 0.4 Mg, 0.2 ± 0.01 Fe, 0.1 ± 0.02 Mn, 0.1 ± 0.01 Ti, 0.02 ± 0.003 Cu, 0.01 ± 0.002 Zn, and Al—balance, that corresponded to that of AA5154. A molybdenum powder bed was formed by spreading 98.7% purity Mo particles on the as-deposited AA51514. The Mo powder was composed of deagglomerated ~2 µm particles and 3–12 μm agglomerates.
EBAM parameters used for deposition of first eight AA5154 layers on a AA5154 substrate (Figure 1a, pos. 3), are shown in Table 1. These parameters were the same as those used earlier for growing the quality and defectless AA5356 walls [41].
On depositing these 8 layers, a powder feeder (Figure 1b, pos. 4) was used to form a molybdenum powder bed of the total weight of 0.3 g on the as-deposited metal. The next step was remelting the powder bed with simultaneous deposition of AA5154 from the wire. Such a cycle was then repeated twice so that the total number of the remelted powder beds was 3. Similar procedures were carried out for powder beds amounting to 0.6, 0.9, and 1.2 g/per a layer. The resulting walls had the dimensions as follows: 60 × 20 × 30 mm3.
The distribution of Mo in the as-deposited samples was studied using an X-ray tomograph Cheetah X-ray Inspection System EVO (YXLON International GmbH, Hamburg, Cermany). Metallographic examination of samples was carried out using optical microscope Altami Met 1S (LCC, Altami, St. Petersburg, Russia) and scanning electron microscope Tescan MIRA 3 LMU (TESCAN ORSAY HOLDING, Brno, Czech Republic), equipped with Oxford Instruments Ultim Max 40 EDS detector (Oxford Instruments, High Wycombe, UK). TEM examinations were performed using TEM instrument JEOL-2100 (JEOL Ltd., Akishima, Japan) on foils thinned by a focused ion beam.
X-ray diffraction (XRD) phase analysis was carried out by means of an X-ray diffractometer DRON-7 (Innovation Center “Burevestnik”, St. Petersburg, Russia) using CoKα radiation at 0.05° increments and an exposure time of 25 s/step that allowed detecting phases such as Al and Mo as well as their intermetallic compounds (IMCs). Grazing-incidence X-ray diffraction (GIAXD) at beam incidence angle of 10° was applied for studying phases formed in a thin subsurface layer by friction against the steel disk. The XRD peak identification and treatment was performed using Crystal Impact software “Match!” (version 3.9, Crystal Impact, Bonn, Germany). The phase contents were obtained from the XRD data according to Equation (1), where Vphase is the phase percentage, Iphase is the phase’s integral phase peak intensity –, and Iphase is the integral intensity of any other reflection.
V p h a s e = ( Σ ( I p h a s e ) Σ ( I s u m m ) ) × 100 %
Microhardness of the phases was measured using a Duramin5 (Struers A/S, Ballerup, Denmark) machine. A tensile test machine UTS-110M-100 (Testsystems, Ivanovo, Russia) and tribometer TRIBOtechnic (Tribotechnic, Clichy, France) were used for determining the sample strength and tribological behavior, respectively.
Pin-on-disk sliding was applied for measuring the friction and wear characteristics of samples. Samples in the form of 3 × 3 × 10 mm3 pins were cut off the AA5154/Mo zone, while Ø35 mm, 5 mm thickness counterbody disks were machined from 40 × 13 quenched steel. Rotation rate, normal load and sliding path length were 94 RPM, 15 N and 5600 m, respectively.

3. Results

Macrostructural cross section views of as-deposited AA5154 (zone 1) and AA5154/Mo (zone 2) demonstrate many semicircular fusion boundaries left after each track deposition according to the strategy selected, as shown in Figure 2.
The microstructure of as-deposited AA5154 (zone 1) is the same for all samples and represented by almost non-etched α-Al dendrites with numerous interdendrite precipitates. Zone 2 of all samples demonstrates both bright and dark-etched areas that allow suggesting inhomogeneous distribution of Mo powder in them (Figure 2). Coarse light particles and discontinuities can be observed in zone 2 of all samples, which could result from sintering the Mo powder agglomerates and diffusion reaction with Al.
Cross-section views of each deposition track show semicircular dark-etched bands with intermetallic compounds formed during the solidification of the melted pool.
X-ray tomography images in Figure 3 allow observing distribution of Mo in the cross section views of as-grown AA5154/Mo samples. The dark contrasting particles can be seen owing to their higher density as compared to the as-deposited A5154 layers and therefore reveal distribution of both unreacted Mo and Al-Mo IMCs formed from the Mo powder-bed. Samples obtained with 0.3 g/layer content of Mo show the most uniform dark-contrast phase distribution over the cross section view.
Typical needle-like IMC microstructures were found near the unreacted Mo-rich particles (Figure 4a,b) that precipitated from the solid solution during solidification. Areas darker than those of unreacted Mo ones can be observed in Figure 4b that were formed by dissolution of Mo particle in aluminum alloy α-Al with ensuing precipitation of Al/Mo IMCs.

3.1. Phase Detection with XRD

XRD patterns were obtained from cross section views of the AA5154/Mo samples (Figure 5a) that allowed detecting peaks of α-Al and unreacted Mo as well as those of Al/Mo IMCs identified as Al12Mo, Al5Mo, Al8Mo3, Al18Mg3Mo2 (Figure 5a). Total amount of IMCs increased with the total Mo content (Figure 5b). The unreacted Mo content is reduced starting from 0.6 g/layer load when more IMCs are formed.

3.2. Microstructure of AA5154/Mo Composite

The zone 2 microstructure of samples intermixed with 0.3 wt.% Mo shows large 0.5–1 mm sized particles of Mo with about 5 at.% Al (Figure 6a,b,f, Table 2, spectrum 1). Transition areas can be seen between these Mo-rich particles and Al matrices that are filled with intermetallic dendritic structures composed of ~30 at.% Mo and ~70 at.% Al (Figure 6b). The atomic concentration Al/Mo ratio is close to that of the Al8Mo3 IMC that, according to Al-Mo diagram [42], exists within a wide concentration range 27–71 at.% Al and 27–77 at.% Mo.
Smaller particles (Figure 6b, Table 2, spectrums 3–5) contain more aluminum and less molybdenum so that their concentration ratio allows identifying them as Al8Mo3 and Al5Mo. In addition to transition areas filled with dendritic IMCs, there are sharp interfaces between Mo-rich and α-Al-rich areas (Figure 6e,f). Almost homogeneously distributed needle-like and faceted 2–5 μm IMCs can be observed in the α-Al areas (Figure 6c,d). EDS shows that these precipitates contain 87–89 at.% Al, 8–11 at.% Mo and Mg, Mn, Ti, Fe as impurities (Figure 6d, Table 2, spectrums 6–8).
It should be noted that some Al-Mo precipitates in zone 2, α-Al, are characterized by the presence of Mg (Figure 7d), which allows suggesting the formation of Al18Mg3Mo2 IMCs. The existence of these IMCs can be supported by TEM images and corresponding SAED patterns in Figure 7g as well as EDS (Figure 7, Table 3, spectrum 1). These faceted Al18Mg3Mo2 particles are about 0.8 μm in size and localize on the α-Al grain boundaries together with the Fe-rich ones (Figure 7g–j, Table 3 spectrums 2, 3, 5, 6). The total concentration of Mo in the α-Al is not higher than 0.4 at.% (Figure 6d, Table 2, spectrum 9, Figure 7h, Table 3, spectrums 4, 7). The Fe-rich grain boundary precipitates are inherent in the Al-Mg aluminum alloys obtained using EBAM [26].
The layers deposited with 0.6 g/layer Mo show higher amounts of both ~1 μm and <100 nm Mo-rich particles as well as Al-Mo IMCs (Figure 8) in accordance to Figure 6b. EDS spectra 1 and 2 (Table 4) show that Mo-rich particles contain up to 2.3 at.% Al and are encompassed by IMC needle-like and faceted precipitates (Figure 8a,b). The α-Al grains contain mostly faceted IMCs that correspond to Al12Mo according to EDS data (Figure 8b–d, Table 4, spectrums 3–7, 10).
TEM shows that these precipitates are really Al12Mo (Figure 9a–e,g). Another finding was that irregularly shaped IMCs were found in the α-Al regions which contained up to ~7 at.% Mo and <1 at% of impurities such as Fe, Ti, Mn, Mg (Table 4, spectrums 8, 9). The concentration of Mg in the α-Al regions is at the level of 0.3 at.% (Table 4, spectrums 11, 12). According to the TEM data there are also precipitates such as Al18Mg3Mo2 and as well as the grain boundary precipitates containing up to 11–15 at.% Fe (Figure 9f,h).
For samples obtained with Mo loads of 0.9 and 1.2 g/layer, it becomes observable that the volume content of Mo-rich particles is reduced (Figure 10 and Figure 11) with simultaneous increasing of the areas occupied by Al-Mo IMCs (Figure 10d and Figure 11c,e). Mo-rich particles containing 1.6–7.6 at.% Al were detected (Figure 10d and Figure 11c, Table 5, Table 6 and Table spectrum 1) that were encompassed with the light-grey Al8Mo3 shells (~46–62 at.% Al, ~38–54 at.% Mo) (Figure 10d and Figure 11c, Table 5, Table 6 and Table spectrums 2, 3). According to the Al-Mo diagram, this phase may exist in a wide concentration range so that increasing Al concentration by 20 at.% would not change the stoichiometry but only provide darker contrast in SEM BSE images. Al17Mo4 and Al12Mo shells are formed in regions closer to α-Al ones (Figure 10d, Table 5, spectrums 4, 5).
Greater numbers of needle-like and faceted Al12Mo IMCs are formed in samples charged with 0.9 g/layer Mo powder (Figure 10). Additionally, these particles are larger than those formed in samples with 0.3 and 0.6 g/layer.
Similar structures were observed in samples obtained with even higher concentration of Mo (1.2 g/layer) (Figure 11c,e, Table 6). Along with the Mo-rich multi-shell Al-Mo regions, there are large regions composed of eutectic-like Al8Mo3 structures (Figure 11f, Table 6). Coarse platelet- or needle-like Al5Mo precipitates can be found near these IMC shells (Figure 11e, Table 6). The α-Al regions contain particles with up to 2 at.% of Mo as well as faceted Al12Mo ones (Table 5, spectrums 8, 9).
Partial sedimentation of Mo particles on the melted pool bottom was observed during the deposition of AA5154 layer on the Mo powder bed so that the zone 1/zone 2 IMC boundary became clearly seen in Figure 11a. This IMC boundary contains 75.1–97.7 at.% Mo, 2.3–20 at.% Al, 11.1–14.42 at.% Mg and 0.7–0.9 at.% Fe and shows some cracking.
TEM images show the presence of Al12Mo IMCs in samples obtained with Mo powder concentrations 0.9 and 1.2 g/layer (Figure 12a,b). Their composition coincides with those obtained from EDS on SEM images (Table 5 and Table 6). Grain boundary Fe-rich particles were also identified in the TEM images (Figure 12c).

3.3. Microhardness of AA5154/Mo

Microhardness number vertical profiles were obtained on the cross-section views, as shown in Figure 13a–d, at a distance of 100 μm between the lines. The microhardness number distribution along the corresponding lines revealed sharp changes at the matrix/IMC interfaces. Mo-rich regions in the 0.3, 0.6, 0.9 and 1.2 g/layer Mo samples had maximum microhardness at the level of 2300, 2600, 11,500 and 9000 GPa, respectively (Figure 13a–d). The microhardness of these Mo-rich regions in the last two samples was much greater than those in the first and second ones (Figure 13c,d). Such a result may be explained by indenting those wide Al-Mo IMC shells formed around the Mo-rich particles. The microhardness of α-Al solid solution in as-deposited AA5154 (zone 1) was higher than that of zone 2 with IMCs due to depletion by Mg which diffused into Al18Mg3Mo2 IMCs.

3.4. Sliding Wear A5154/Mo

All samples had their coefficient of friction (CoF) vs. time dependencies as shown in Figure 14. All CoFs mean values lie within the 0.45–0.55 range including that of the base AA5154. Somewhat lower CoF values are demonstrated by samples obtained with the maximal 1.2 g/layer input of Mo powder.
The maximum wear was detected in sliding the base alloy against the steel disk (Figure 15). The in situ formed Al/Mo IMCs allowed reducing the wear so that its minimum value was shown by samples obtained with the maximal 1.2 g/layer input of Mo powder.
The worn surface EDS analysis shows the presence of Mo-rich areas as well as Fe-rich areas in addition to those containing aluminum. Intense tribooxidation of all samples occurred in all samples during sliding wear against a steel disk.
Mo-rich areas (Figure 16) are generated by deformation, adhesion transfer and smearing of Mo-rich IMCs during sliding, while Fe-rich ones appeared because of reverse transfer from the steel counterbody (Figure 17).
Oxide phases such as Fe3O4, non-stoichiometric Mo8O23 and mixed Al2(MoO4)3 have been detected on the worn surfaces of samples by means of GIAXD (Figure 18) to the exclusion of samples with an initial Mo concentration of 0.6 g/layer where no Al2(MoO4)3 was present. Such a finding may be the result of selective wear of the tribooxidized aluminum matrix (Figure 19a). Semi-quantitative analysis of GIAXD peaks of Fe3O4, Mo8O23 and Al2(MoO4)3 oxides showed that their amount increased from 15 to 24 vol.% in a 10–15 μm thickness subsurface layer for samples with initial Mo concentrations of 0.3–1.2 g/layer.
It can be observed from Figure 18 that free molybdenum was found only on the worn surface of the 0.9 g/layer sample, which also showed the maximum wear among all the AA5154/Mo samples (Figure 15). The worn surface of this sample demonstrates large, damaged light-grey areas formed by the removal of a mechanically mixed layer (Figure 19b).
It looks obvious that coarse unoxidized Mo particles underlying the mechanically mixed oxide layer make their way to the surface, thus becoming easily detected with the GIAXD. Fewer damaged areas are observed on the worn surfaces of all other AA5154 g/layer samples which are covered by mechanically mixed layer areas composed of Al/Mo IMCs tribooxidation products such as Mo8O23 and Al2(MoO4)3.
The slight increase in wear for samples with initial concentrations of Mo 0.3, 0.6 and 0.9 g/layer can be explained by the insufficient amount of Mo8O23 and Al2(MoO4)3 oxides formed but an increasing amount of brittle Al/Mo IMCs (Figure 5b) [43].

4. Discussion

Molybdenum has low diffusion factor, rather low solubility in aluminum and is capable of forming a row of hard AlxMoy IMCs. The results of this work allowed showing that changing the amount of Mo charged into the aluminum matrix at a step as low as 0.3 g resulted in a notable increase in the volume fraction of IMCs in the aluminum alloy matrix. Both Mo-rich and Al-rich IMCs were formed depending on the distribution of the Mo powder particles in the melted pool. The latter ones, such as Al12Mo and Al18Mg3Mo2, were detected using XRD, EDS and partially TEM. Let us note that many IMCs were pulled out of the matrix during sample preparation procedures and therefore had never been analyzed. The morphology of Al12Mo IMCs changed depending on the concentration differences and solidification front atomic morphology [44] that is determined according to the expression as follows:
α = Δ S f R
where α is the fusion entropy, ΔSf is the specific fusion entropy per mole, R is the universal gas constant. The unfaceted crystals grow at α < 2, while values higher than α > 2 facilitate the formation of crystalline faces [44].
The IMC phase formation is determined by Gibbs energy change during the diffusion reaction between Al and Mo as follows: ( Δ G 0 ) [45]:
Δ G 0 = Δ H 0 T Δ S 0
where Δ H 0 is the change in enthalpy; T is the reaction temperature; Δ S 0 is the change in entropy that can be neglected in this solid state case to allow rewriting Equation (3) as follows: Δ G 0 Δ H 0 .
For the Al-Mo system, the probability of an IMC formation by diffusion reaction can be explained using several adopted models [45,46]. Pretorius R. et al. [45] proposed a model for determining the effective heat of formation ( Δ H ) on the binary metal/aluminum phase interface:
Δ H i = Δ H i 0 · C e C c
where Δ H i is the effective heat of formation of the i-phase, Ce is the effective concentration of the limiting element at the interface, and Cc is the concentration of the limiting element in the compound. Assuming that Δ G 0 Δ H 0 , the effective Gibbs energy change in the case of forming the i-phase in the Al/Mo layer may be determined as follows:
Δ G i = Δ G i 0 · C e C c
Table 7 contains the effective Gibbs free energy changes calculated for all Al/Mo phases.
It follows from Table 7 that all AlxMoy phases have negative values of the Gibbs free energy changes and therefore can be formed in our experimental conditions. The minimal values belong to the Al-rich ones detected in the experiment. Al-Mo solid solutions may also form in the Al-rich areas
Inhomogeneous distribution of Mo particles and brittle Al/Mo IMCs, as well as the presence of defects such as voids, resulted in non-uniform distribution of the microhardness numbers. The friction and wear behavior of samples showed reduced friction and wear for all the AA5154/Mo samples as compared to that of AA5154. The IMCs allowed increasing the worn subsurface hardness and generating mechanically mixed layers (MML) consisting of Fe3O4, Mo8O23 and Al2(MoO4)3 oxides, which formed due to tribooxidation of AlxMoy IMCs and adhesive transfer of iron from the steel counterbody. These oxides may be related to easy shear compounds that provide a lubrication effect and improve the tribological characteristics of composites [47]. It is also known that [48] introducing Mo into ceramic coatings served not only for their enhanced hardness and wear resistance, but also reduced friction due to the formation of MoO3 on the worn surfaces.
However, there was another factor that served for some minor increase in wear for samples containing from 0.3 to 0.9 g/layer of Mo. This factor was the bearing ability of the MML depending on the number of brittle IMCs in the subsurface of the samples, which were pulled out and abraded the worn surface. However, this did not occur with the 1.2 g/layer sample where the maximal amount of self-lubrication oxides was formed to reduce the abrasion intensity.

5. Conclusions

Combination of wire-feed and powder-bed electron beam additive manufacturing was used for intermixing Mo powder into melted AA5154. Different amounts of Mo powder were used to form the powder-beds that were remelted together with previously deposited AA5154 layers to result in inhomogeneous distribution of coarse Mo particles in the AA5154 matrix. Intermetallic Al/Mo compounds were found near these particles in the form of needle-like precipitates. The intermixed zone structure was composed of AA5154 matrix with unreacted Mo particles and Al/Mo intermetallic precipitates whose amount depended on the amount of Mo powder introduced. The resulting structures possessed high microhardness and reduced sliding wear. The worn surface analysis showed the presence of mechanically mixed layers composed of wear particles and Fe3O4, Mo8O23 and Al2(MoO4)3 oxides which could be the reason for reduced wear of the composite materials as compared to that of the as-deposited AA5154.

Author Contributions

Conceptualization, A.Z. and S.T.; Data curation, A.Z. and A.C.; Formal analysis, A.Z., A.C., E.M., D.G. and N.S. (Nickolai Savchenko); Funding acquisition, A.Z.; Investigation, A.Z., A.C., A.V., N.S. (Nickolay Shamarin), E.M., D.G. and N.S. (Nickolai Savchenko); Methodology, A.Z., A.V. and S.T.; Resources, A.Z.; Software, N.S. (Nickolay Shamarin), A.P. and E.K. (Evgeny Knyazhev); Supervision, E.K. (Evgeny Kolubaev); Validation, A.P. and E.K. (Evgeny Knyazhev); Visualization, A.V.; Writing—original draft, A.Z. and S.T.; Writing—review & editing, A.Z., E.K. (Evgeny Kolubaev) and S.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Federation Government under research assignment for ISPMS SB RAS, project no. FWRW-2021-0012.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yener, T. Low temperature aluminising of Fe-Cr-Ni super alloy by pack cementation. Vacuum 2019, 162, 114–120. [Google Scholar] [CrossRef]
  2. Ding, Y.; You, G.; Wen, H.; Li, P.; Tong, X.; Zhou, Y. Microstructure and mechanical properties of inertia friction welded joints between alloy steel 42CrMo and cast Ni-based superalloy K418. J. Alloys Compd. 2019, 803, 176–184. [Google Scholar] [CrossRef]
  3. Wang, C.; Wu, Y.; Guo, Y.; Guo, J.; Zhou, L. Precipitation behavior of sigma phase and its influence on mechanical properties of a Ni-Fe based alloy. J. Alloys Compd. 2019, 784, 266–275. [Google Scholar] [CrossRef]
  4. Huang, Y.; Liu, Y.; Li, C.; Ma, Z.; Yu, L.; Li, H. Microstructure evolution and phase transformations in Ti-22Al-25Nb alloys tailored by super-transus solution treatment. Vacuum 2019, 161, 209–219. [Google Scholar] [CrossRef]
  5. Pan, Y.; Wen, M. Ab-initio calculations of mechanical and thermodynamic properties of TM (transition metal: 3d and 4d)-doped Pt3Al. Vacuum 2018, 156, 419–426. [Google Scholar] [CrossRef]
  6. Eumann, M.; Sauthoff, G.; Palm, M. Re-evaluation of phase equilibria in the Al–Mo system. Int. J. Mater. Res. 2006, 97, 1502–1511. [Google Scholar] [CrossRef]
  7. Wang, M.; Chen, Z.; Xia, C.; Wu, Y.; Chen, D. Theoretical study of elastic and electronic properties of Al5Mo and Al5W intermetallics under pressure. Mater. Chem. Phys. 2017, 197, 145–153. [Google Scholar] [CrossRef]
  8. Nino, R.; Miura, S.; Mohri, T. The behavior of cracking in Mo3Al–Mo3Al8 two-phase intermetallics with lamellar structures. Intermetallics 2001, 9, 113–118. [Google Scholar] [CrossRef]
  9. Kiryukhantsev-Korneev, P.; Iatsyuk, I.; Shvindina, N.; Levashov, E.; Shtansky, D. Comparative investigation of structure, mechanical properties, and oxidation resistance of Mo-Si-B and Mo-Al-Si-B coatings. Corros. Sci. 2017, 123, 319–327. [Google Scholar] [CrossRef]
  10. Bian, Y.; Gao, T.; Li, Z.; Sun, Q.; Ma, X.; Liu, X. In-situ synthesis of an Al composite reinforced with multi–scale Al12Mo, (Al, Zr, Si) and Al2O3 particles through a multi–stage reaction. Mater. Sci. Eng. A 2019, 762, 138069. [Google Scholar] [CrossRef]
  11. Pan, Y.; Pu, D.; Liu, G. Influence of Mo concentration on the structure, mechanical and thermodynamic properties of Mo–Al compounds from first-principles calculations. Vacuum 2020, 175, 109291. [Google Scholar] [CrossRef]
  12. Minasyan, T.; Ivanov, R.; Toyserkani, E.; Hussainova, I. Laser powder-bed fusion of Mo(Si,Al)2—Based composite for elevated temperature applications. J. Alloys Compd. 2021, 884, 161034. [Google Scholar] [CrossRef]
  13. Sharifitabar, M.; Sadeq, F.O.; Afarani, M.S. Synthesis and kinetic study of Mo(Si,Al)2 coatings on the surface of molybdenum through hot dipping into a commercial Al-12 wt.%Si alloy melt. Surf. Interfaces 2021, 24, 101044. [Google Scholar] [CrossRef]
  14. Kumar, J.; Singh, D.; Kalsi, N.S.; Sharma, S.; Mia, M.; Singh, J.; Rahman, M.A.; Khan, A.M.; Rao, K.V. Investigation on the mechanical, tribological, morphological and machinability behavior of stir-casted Al/SiC/Mo reinforced MMCs. J. Mater. Res. Technol. 2021, 12, 930–946. [Google Scholar] [CrossRef]
  15. Pan, S.; Guan, Z.; Yao, G.; Yuan, J.; Li, X. Mo-enhanced chemical stability of TiC nanoparticles in molten Al. J. Alloys Compd. 2021, 856, 158169. [Google Scholar] [CrossRef]
  16. Richardson, P.J.; Keast, V.J.; Cuskelly, D.T.; Kisi, E.H. Theoretical and experimental investigation of the W-Al-B and Mo-Al-B systems to approach bulk WAlB synthesis. J. Eur. Ceram. Soc. 2021, 41, 1859–1868. [Google Scholar] [CrossRef]
  17. Cordill, M.; Kreiml, P.; Putz, B.; Mitterer, C.; Thiaudière, D.; Mocuta, C.; Renault, P.-O.; Faurie, D. Role of layer order on the equi-biaxial behavior of Al/Mo bilayers. Scr. Mater. 2021, 194, 113656. [Google Scholar] [CrossRef]
  18. Lee, I.; Kao, P.; Chang, C.; Ho, N. Formation of Al–Mo intermetallic particle-strengthened aluminum alloys by friction stir processing. Intermetallics 2013, 35, 9–14. [Google Scholar] [CrossRef]
  19. Mahesh, V.P.; Alphonsa, J.; Arora, A. Electrochemical Behavior of Aluminum-Molybdenum Surface Composites Developed by Friction Stir Processing. J. Mater. Eng. Perform. 2021, 30, 8663–8676. [Google Scholar] [CrossRef]
  20. Khorasani, M.; Ghasemi, A.; Rolfe, B.; Gibson, I. Additive manufacturing a powerful tool for the aerospace industry. Rapid Prototyp. J. 2021. ahead of print. [Google Scholar] [CrossRef]
  21. Köhler, M.; Fiebig, S.; Hensel, J.; Dilger, K. Wire and Arc Additive Manufacturing of Aluminum Components. Metals 2019, 9, 608. [Google Scholar] [CrossRef] [Green Version]
  22. Huang, W.; Chen, S.; Xiao, J.; Jiang, X.; Jia, Y. Laser wire-feed metal additive manufacturing of the Al alloy. Opt. Laser Technol. 2021, 134, 106627. [Google Scholar] [CrossRef]
  23. Klein, T.; Schnall, M.; Gomes, B.; Warczok, P.; Fleischhacker, D.; Morais, P.J. Wire-arc additive manufacturing of a novel high-performance Al-Zn-Mg-Cu alloy: Processing, characterization and feasibility demonstration. Addit. Manuf. 2021, 37, 101663. [Google Scholar] [CrossRef]
  24. Tang, S.; Wang, G.; Song, H.; Li, R.; Zhang, H. A novel method of bead modeling and control for wire and arc additive manufacturing. Rapid Prototyp. J. 2021, 27, 311–320. [Google Scholar] [CrossRef]
  25. Kulkarni, J.D.; Goka, S.B.; Parchuri, P.K.; Yamamoto, H.; Ito, K.; Simhambhatla, S. Microstructure evolution along build direction for thin-wall components fabricated with wire-direct energy deposition. Rapid Prototyp. J. 2021, 27. [Google Scholar] [CrossRef]
  26. Yang, Y.; Geng, K.; Li, S.; Bermingham, M.; Misra, R. Highly ductile hypereutectic Al-Si alloys fabricated by selective laser melting. J. Mater. Sci. Technol. 2021, 110, 84–95. [Google Scholar] [CrossRef]
  27. Sun, S.; Liu, P.; Hu, J.; Hong, C.; Qiao, X.; Liu, S.; Zhang, R.; Wu, C. Effect of solid solution plus double aging on microstructural characterization of 7075 Al alloys fabricated by selective laser melting (SLM). Opt. Laser Technol. 2019, 114, 158–163. [Google Scholar] [CrossRef]
  28. Ma, R.-L.; Peng, C.-Q.; Cai, Z.-Y.; Wang, R.-C.; Zhou, Z.-H.; Li, X.-G.; Cao, X.-Y. Finite element analysis of temperature and stress fields during selective laser melting process of Al–Mg–Sc–Zr alloy. Trans. Nonferrous Met. Soc. China 2021, 31, 2922–2938. [Google Scholar] [CrossRef]
  29. Utyaganova, V.; Filippov, A.; Tarasov, S.; Shamarin, N.; Gurianov, D.; Vorontsov, A.; Chumaevskii, A.; Fortuna, S.; Savchenko, N.; Rubtsov, V.; et al. Characterization of AA7075/AA5356 gradient transition zone in an electron beam wire-feed additive manufactured sample. Mater. Charact. 2021, 172, 110867. [Google Scholar] [CrossRef]
  30. Domack, M.S.; Taminger, K.M.; Begley, M. Metallurgical Mechanisms Controlling Mechanical Properties of Aluminium Alloy 2219 Produced by Electron Beam Freeform Fabrication. Mater. Sci. Forum 2006, 519–521, 1291–1296. [Google Scholar] [CrossRef]
  31. Taminger, K.M.; Hafley, R.A.; Domack, M.S. Evolution and Control of 2219 Aluminium Microstructural Features through Electron Beam Freeform Fabrication. Mater. Sci. Forum 2006, 519–521, 1297–1302. [Google Scholar] [CrossRef]
  32. Brice, C.A.; Dennis, N. Cooling Rate Determination in Additively Manufactured Aluminum Alloy 2219. Met. Mater. Trans. A 2015, 46, 2304–2308. [Google Scholar] [CrossRef]
  33. Gasper, A.; Catchpole-Smith, S.; Clare, A. In-situ synthesis of titanium aluminides by direct metal deposition. J. Mater. Process. Technol. 2017, 239, 230–239. [Google Scholar] [CrossRef]
  34. Langelandsvik, G.; Eriksson, M.; Akselsen, O.M.; Roven, H.J. Wire arc additive manufacturing of AA5183 with TiC nanoparticles. Int. J. Adv. Manuf. Technol. 2021, 1–12. [Google Scholar] [CrossRef]
  35. Jin, P.; Liu, Y.; Li, F.; Sun, Q. Realization of synergistic enhancement for fracture strength and ductility by adding TiC particles in wire and arc additive manufacturing 2219 aluminium alloy. Compos. Part B Eng. 2021, 219, 108921. [Google Scholar] [CrossRef]
  36. Behera, M.P.; Dougherty, T.; Singamneni, S. Conventional and Additive Manufacturing with Metal Matrix Composites: A Perspective. Procedia Manuf. 2019, 30, 159–166. [Google Scholar] [CrossRef]
  37. Mostafaei, A.; Heidarzadeh, A.; Brabazon, D. Production of Metal Matrix Composites Via Additive Manufacturing. In Encyclopedia of Materials: Composites; Elsevier BV: Amsterdam, The Netherlands, 2020; pp. 605–614. [Google Scholar]
  38. Seetharaman, S.; Gupta, M. Additive Manufacturing of Metal Matrix Composites. In Encyclopedia of Materials: Composites; Elsevier BV: Amsterdam, The Netherlands, 2021; pp. 209–229. [Google Scholar]
  39. Parsons, E.M.; Shaik, S.Z. Additive manufacturing of aluminum metal matrix composites: Mechanical alloying of composite powders and single track consolidation with laser powder bed fusion. Addit. Manuf. 2021, 50, 102450. [Google Scholar] [CrossRef]
  40. Filippov, A.; Khoroshko, E.; Shamarin, N.; Savchenko, N.; Moskvichev, E.; Utyaganova, V.; Kolubaev, E.; Smolin, A.; Tarasov, S. Characterization of gradient CuAl–B4C composites additively manufactured using a combination of wire-feed and powder-bed electron beam deposition methods. J. Alloys Compd. 2021, 859, 157824. [Google Scholar] [CrossRef]
  41. Utyaganova, V.R.; Filippov, A.V.; Shamarin, N.N.; Vorontsov, A.V.; Savchenko, N.L.; Fortuna, S.V.; Gurianov, D.A.; Chumaevskii, A.V.; Rubtsov, V.E.; Tarasov, S.Y. Controlling the porosity using exponential decay heat input regimes during electron beam wire-feed additive manufacturing of Al-Mg alloy. Int. J. Adv. Manuf. Technol. 2020, 108, 2823–2838. [Google Scholar] [CrossRef]
  42. Schuster, J.C. Materials Science International Team, MSIT® The Assessed Phase Diagram of the Al-Mo System: Datasheet from MSI Eureka in Springer Materials. Available online: https://0-materials-springer-com.brum.beds.ac.uk/msi/phase-diagram/docs/sm_msi_r_20_012123_01_full_LnkDia0 (accessed on 27 November 2021).
  43. Wu, Q.; Yang, C.; Xue, F.; Sun, Y. Effect of Mo addition on the microstructure and wear resistance of in situ TiC/Al composite. Mater. Des. 2011, 32, 4999–5003. [Google Scholar] [CrossRef]
  44. Almeida, A.; Carvalho, F.; Carvalho, P.; Vilar, R. Laser developed Al–Mo surface alloys: Microstructure, mechanical and wear behaviour. Surf. Coat. Technol. 2006, 200, 4782–4790. [Google Scholar] [CrossRef]
  45. Pretorius, R.; Vredenberg, A.M.; Saris, F.W.; De Reus, R. Prediction of phase formation sequence and phase stability in binary metal-aluminum thin-film systems using the effective heat of formation rule. J. Appl. Phys. 1991, 70, 3636–3646. [Google Scholar] [CrossRef]
  46. Saunders, N.; Thermotech Ltd. The Al-Mo system (aluminum-molybdenum). J. Phase Equilibria Diffus. 1997, 18, 370–378. [Google Scholar] [CrossRef]
  47. Voevodin, A.; Muratore, C.; Aouadi, S. Hard coatings with high temperature adaptive lubrication and contact thermal management: Review. Surf. Coat. Technol. 2014, 257, 247–265. [Google Scholar] [CrossRef]
  48. Grigoriev, S.; Vereschaka, A.; Milovich, F.; Sitnikov, N.; Andreev, N.; Bublikov, J.; Kutina, N. Investigation of the properties of the Cr,Mo-(Cr,Mo,Zr,Nb)N-(Cr,Mo,Zr,Nb,Al)N multilayer composite multicomponent coating with nanostructured wear-resistant layer. Wear 2020, 468–469, 203597. [Google Scholar] [CrossRef]
Figure 1. EBAM process stages such as electron beam deposition of AA5154 layers (a) spreading a Mo powder-bed (b) remelting the Mo powder-bed with deposition of AA5154 (c) as-built AA5154/Mo composite zones (d). 1—electron beam, 2—AA5154 filament, 3—AA5154 layers, 4—powder feeder, 5—Mo powder, 6—samples for SEM and TEM.
Figure 1. EBAM process stages such as electron beam deposition of AA5154 layers (a) spreading a Mo powder-bed (b) remelting the Mo powder-bed with deposition of AA5154 (c) as-built AA5154/Mo composite zones (d). 1—electron beam, 2—AA5154 filament, 3—AA5154 layers, 4—powder feeder, 5—Mo powder, 6—samples for SEM and TEM.
Metals 12 00109 g001
Figure 2. Optical cross section images of AA5154/Mo track zones.
Figure 2. Optical cross section images of AA5154/Mo track zones.
Metals 12 00109 g002
Figure 3. X-ray tomography cross section positive images of the Zone 2 structures as-viewed from the direction parallel to that of layer deposition. Dark and light-grey areas reveal Mo-rich areas and pores, respectively.
Figure 3. X-ray tomography cross section positive images of the Zone 2 structures as-viewed from the direction parallel to that of layer deposition. Dark and light-grey areas reveal Mo-rich areas and pores, respectively.
Metals 12 00109 g003
Figure 4. AA5154/Mo IMCs formed near the Mo-rich particles in 0.9 g/layer (a) and 1.2 g/layer (b) samples.
Figure 4. AA5154/Mo IMCs formed near the Mo-rich particles in 0.9 g/layer (a) and 1.2 g/layer (b) samples.
Metals 12 00109 g004
Figure 5. XRD diffractograms (a) and volume fractions of structural components (b) as depending on the Mo powder bed input.
Figure 5. XRD diffractograms (a) and volume fractions of structural components (b) as depending on the Mo powder bed input.
Metals 12 00109 g005
Figure 6. SEM BSE images of zone 2 AA5154/Mo coarse Mo particles (a,c,f),and enlarged views of IMCs formed around the coarse particles (b,d,e) in the 0.3 g/layer Mo sample. 1–9 denote points selected for EDS (Table 2).
Figure 6. SEM BSE images of zone 2 AA5154/Mo coarse Mo particles (a,c,f),and enlarged views of IMCs formed around the coarse particles (b,d,e) in the 0.3 g/layer Mo sample. 1–9 denote points selected for EDS (Table 2).
Metals 12 00109 g006
Figure 7. SEM BSE image and EDS maps of zone 2 AA5154/Mo microstructures obtained with 0.3 g/layer Mo (af). TEM images (gj) and SAED pattern showing the IMCs formed. Numbers 1–7 denote points selected for EDS (Table 3).
Figure 7. SEM BSE image and EDS maps of zone 2 AA5154/Mo microstructures obtained with 0.3 g/layer Mo (af). TEM images (gj) and SAED pattern showing the IMCs formed. Numbers 1–7 denote points selected for EDS (Table 3).
Metals 12 00109 g007
Figure 8. SEM BSE images of zone 2 AA5154/Mo microstructures obtained with 0.6 g/layer Mo (ad). Numbers 1–11 denote point selected for EDS (Table 4).
Figure 8. SEM BSE images of zone 2 AA5154/Mo microstructures obtained with 0.6 g/layer Mo (ad). Numbers 1–11 denote point selected for EDS (Table 4).
Metals 12 00109 g008
Figure 9. Bright-field TEM image (a) and EDS maps (bd) of zone 2 microstructures obtained with 0.6 g/layer Mo. EDS profile along the line in Figure 8a (e), and TEM images of IMCs (fh).
Figure 9. Bright-field TEM image (a) and EDS maps (bd) of zone 2 microstructures obtained with 0.6 g/layer Mo. EDS profile along the line in Figure 8a (e), and TEM images of IMCs (fh).
Metals 12 00109 g009
Figure 10. SEM SE (a,c) and BSE (b,df) images of zone 2 AA5154/Mo microstructures obtained with 0.9 g/layer Mo. Numbers 1–9 denote point selected for EDS (Table 5).
Figure 10. SEM SE (a,c) and BSE (b,df) images of zone 2 AA5154/Mo microstructures obtained with 0.9 g/layer Mo. Numbers 1–9 denote point selected for EDS (Table 5).
Metals 12 00109 g010
Figure 11. SEM SE (a,c) and BSE (b,df) images of zone 2 AA5154/Mo microstructures obtained with 1.2 g/layer Mo. 1–10 denote points selected for EDS (Table 6).
Figure 11. SEM SE (a,c) and BSE (b,df) images of zone 2 AA5154/Mo microstructures obtained with 1.2 g/layer Mo. 1–10 denote points selected for EDS (Table 6).
Metals 12 00109 g011
Figure 12. TEM bright-field images of zone 2 AA5154/Mo microstructures obtained with 0.9 g/layer (a) and 1.2 g/layer (b,c) Mo.
Figure 12. TEM bright-field images of zone 2 AA5154/Mo microstructures obtained with 0.9 g/layer (a) and 1.2 g/layer (b,c) Mo.
Metals 12 00109 g012
Figure 13. Microhardness profiles obtained along the lines shown in the insets 0.3 g/layer Mo (a), 0.6 g/layer Mo (b), 0.9 g/layer Mo (c) and 1.2 g/layer Mo (d).
Figure 13. Microhardness profiles obtained along the lines shown in the insets 0.3 g/layer Mo (a), 0.6 g/layer Mo (b), 0.9 g/layer Mo (c) and 1.2 g/layer Mo (d).
Metals 12 00109 g013
Figure 14. Coefficient of friction vs. time dependencies of samples.
Figure 14. Coefficient of friction vs. time dependencies of samples.
Metals 12 00109 g014
Figure 15. Wear of the composite layers obtained.
Figure 15. Wear of the composite layers obtained.
Metals 12 00109 g015
Figure 16. Mo-rich areas on the worn surface of 0.9 g/layer sample.
Figure 16. Mo-rich areas on the worn surface of 0.9 g/layer sample.
Metals 12 00109 g016
Figure 17. Fe-rich areas on the worn surface of 1.2 g/layer sample.
Figure 17. Fe-rich areas on the worn surface of 1.2 g/layer sample.
Metals 12 00109 g017
Figure 18. The glancing X-ray diffractograms of worn surface of AA5154/Mo with incidence angle 10°.
Figure 18. The glancing X-ray diffractograms of worn surface of AA5154/Mo with incidence angle 10°.
Metals 12 00109 g018
Figure 19. The worn surfaces of: (a) 0.6 g/layer sample; (b) 0.9 g/layer sample.
Figure 19. The worn surfaces of: (a) 0.6 g/layer sample; (b) 0.9 g/layer sample.
Metals 12 00109 g019
Table 1. The additive process parameters.
Table 1. The additive process parameters.
Acceleration Voltage, kVBeam Current, mAWire Feed Rate, mm/minTable Speed, mm/min
30401590380
Table 2. EDS analysis of the AA5154/Mo with 0.3 g/layer was obtained by using SEM (Figure 6b,d).
Table 2. EDS analysis of the AA5154/Mo with 0.3 g/layer was obtained by using SEM (Figure 6b,d).
SpectrumConcentration of Elements, at.%Calculated Phase
AlMoMgTiFeMn
1595----Mo(Al)
269.330.7----Al8Mo3
335.664.4----Al8Mo3
482.217.8----Al5Mo
566.233.8----Al8Mo3
689.68.80.40.20.10.4Al12Mo
787.710.10.40.20.5-Al12Mo
887.810.90.40.2-0.2Al12Mo
998.90.30.70.1--α-Al
Table 3. EDS analysis of the AA5154/Mo with 0.3 g/layer was obtained by using TEM (Figure 7g–j).
Table 3. EDS analysis of the AA5154/Mo with 0.3 g/layer was obtained by using TEM (Figure 7g–j).
SpectrumConcentration of Elements, at.%Calculated Phase
AlMoMgTiFeMn
176.88.314.9---Al18Mg3Mo2
287.6-0.81.49.90.3Fe-containing phase
384.1-0.61.213.80.3Fe-containing phase
498.2-1.8---α-Al
586.84.58.1-0.50.1Al18Mg3Mo2
683.94.07.8-3.90.2Fe-containing phase
798.50.41.1--0.3α-Al
Table 4. EDS analysis of the AA5154/Mo with 0.6 g/layer was obtained by using SEM (Figure 8b–d).
Table 4. EDS analysis of the AA5154/Mo with 0.6 g/layer was obtained by using SEM (Figure 8b–d).
SpectrumConcentration of Elements, at.%Calculated Phase
AlMoMgTiFeMn
12.597.5----Mo(Al)
22.597.5----Mo(Al)
383.816.70.30.10.1-Al12Mo
482.117.60.20.20.1-Al12Mo
587.311.80.20.30.10.3Al12Mo
686.712.6-0.30.10.3Al12Mo
786.912.4-0.30.10.3Al12Mo
892.56.70.10.10.20.4α-Al(Mo)
993.26.5-0.10.2-α-Al(Mo)
1086.813-0.10.1-Al12Mo
1199.70.20.3---α-Al
1299.40.20.30.1--α-Al
Table 5. EDS analysis of the AA5154/Mo with 0.9 g/layer was obtained by using SEM (Figure 10d,f).
Table 5. EDS analysis of the AA5154/Mo with 0.9 g/layer was obtained by using SEM (Figure 10d,f).
SpectrumConcentration of Elements, at.%Calculated Phase
AlMoMgTiFeMn
17.692.4----Mo(Al)
249.550.5----Al8Mo3
361.938.1----Al8Mo3
472.627.4----Al17Mo4
586.712.1----Al12Mo
692.65.40.30.10.30.6Al12Mo
787.810.00.50.30.10.4Al12Mo
898.5 1.5---α-Al
998.20.11.6--0.1α-Al
Table 6. EDS analysis of the AA5154/Mo with 1.2 g/layer was obtained by using SEM (Figure 11e,f).
Table 6. EDS analysis of the AA5154/Mo with 1.2 g/layer was obtained by using SEM (Figure 11e,f).
SpectrumConcentration of Elements, at.%Calculated Phase
AlMo
11.698.4Mo(Al)
252.347.7Al8Mo3
361.938.1Al8Mo3
41.498.6Mo(Al)
546.253.8Al8Mo3
671.628.4Al17Mo4
752.347.7Al8Mo3
881.818.2Al5Mo
943.156.9Al8Mo3
1065.533.5Al8Mo3
Table 7. Effective Gibbs free energy changes calculated for all Al/Mo phases in AA5154/Mo at 2073 K.
Table 7. Effective Gibbs free energy changes calculated for all Al/Mo phases in AA5154/Mo at 2073 K.
CompoundComposition Δ G i ,   J / mol Δ G i   for   2073   K ,   kJ / mol Δ G 0.3 ,   kJ / mol Δ G 0.6 ,   kJ / mol Δ G 0.9 ,   kJ / mol Δ G 1.2 ,   kJ / mol
Al12MoAl0.923Mo0.077−10,700  +  1.865 T–6.8338–5.97–3.70–5.72–6.83
Al5MoAl0.833 Mo0.167−23,000  +  4.381 T–13.9182–6.02–6.41–6.41–5.88
Al17Mo4Al0.810 Mo0.190−27,200  +  4.8 T–16.8350–6.62–6.62–4.84–4.67
Al8Mo3Al0.728 Mo0.272−35,470  +  7.12 T–20.7102–3.71–3.93–4.19–2.96
AlMo3Al0.25Mo0.75−25,250  + 3 T–19.031–1.95–1.95–1.95–1.95
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zykova, A.; Chumaevskii, A.; Vorontsov, A.; Shamarin, N.; Panfilov, A.; Knyazhev, E.; Moskvichev, E.; Gurianov, D.; Savchenko, N.; Kolubaev, E.; et al. Microstructural Evolution of AA5154 Layers Intermixed with Mo Powder during Electron Beam Wire-Feed Additive Manufacturing (EBAM). Metals 2022, 12, 109. https://0-doi-org.brum.beds.ac.uk/10.3390/met12010109

AMA Style

Zykova A, Chumaevskii A, Vorontsov A, Shamarin N, Panfilov A, Knyazhev E, Moskvichev E, Gurianov D, Savchenko N, Kolubaev E, et al. Microstructural Evolution of AA5154 Layers Intermixed with Mo Powder during Electron Beam Wire-Feed Additive Manufacturing (EBAM). Metals. 2022; 12(1):109. https://0-doi-org.brum.beds.ac.uk/10.3390/met12010109

Chicago/Turabian Style

Zykova, Anna, Andrey Chumaevskii, Andrey Vorontsov, Nickolay Shamarin, Aleksandr Panfilov, Evgeny Knyazhev, Evgeny Moskvichev, Denis Gurianov, Nickolai Savchenko, Evgeny Kolubaev, and et al. 2022. "Microstructural Evolution of AA5154 Layers Intermixed with Mo Powder during Electron Beam Wire-Feed Additive Manufacturing (EBAM)" Metals 12, no. 1: 109. https://0-doi-org.brum.beds.ac.uk/10.3390/met12010109

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