Next Article in Journal
Production of Closed-Cell Foams Out of Aluminum Chip Waste: Mathematical Modeling and Optimization
Previous Article in Journal
Effects of Diatomite Contents on Microstructure, Microhardness, Bioactivity and Biocompatibility of Gradient Bioceramic Coating Prepared by Laser Cladding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Mechanical Properties of Low-Density, B2-Ordered AlNbZrTix Multi-Principal Element Alloys

1
School of Mechanical, Electrical & Information Engineering, Putian University, Putian 351100, China
2
College of Mechanical and Electrical Engineering, Fujian Agriculture and Forestry University, Fuzhou 350002, China
3
College of Materials Science and Engineering, Fuzhou University, Fuzhou 350108, China
4
School of Materials Science and Engineering, Fujian University of Technology, Fuzhou 350108, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(6), 932; https://doi.org/10.3390/met12060932
Submission received: 3 May 2022 / Revised: 23 May 2022 / Accepted: 25 May 2022 / Published: 28 May 2022

Abstract

:
Low-density multi-principal element alloys (MPEAs) combining a high specific strength and considerable ductility have remained a research hotspot, due to their promising prospects for energy-saving industrial applications. Light Ti-containing AlNbZrTix (x = 1−3) MPEAs were designed and prepared by induction melting and annealing. As the Ti content increases, the microstructure of these MPEAs evolves from dual phase (B2-ordered and Zr5Al3-type structure) into a single-phase B2-ordered structure, while the density reduces by ~8.7%, from ~5.85 g·cm−3 (x = 1) to ~5.34 g·cm−3 (x = 3). Unexpectedly, the AlNbZrTix (x = 1, 2, 3) alloys possess high specific yield strengths of ~270 kPa·m3·kg−1, ~221 kPa·m3·kg−1, >208 kPa·m3·kg−1, along with excellent fracture strains of ~17.8%, 21.8%, and >50%, respectively. These combined compressive properties are superior to the reported data of most BCC/B2-dominant MPEAs. The deformation mechanism of the B2-ordered structure is explained as a dislocation-based mechanism, accompanied by antiphase domains. Here, the effect of Ti on the microstructure and compressive properties of AlNbZrTix MPEAs was investigated, providing scientific support for the development of advanced low-density materials.

1. Introduction

Multi-principal element alloys (MPEAs), also known as high-entropy alloys because of their high mixing entropy of four or more principal elements, have received extensive attention from the scientific community, due to their unique microstructure and unexpected mechanical properties [1,2,3]. To date, various preparation approaches have produced single-phase solid-solution MPEAs, e.g., face-centered cubic (FCC) MPEAs with excellent room-temperature ductility and cryogenic damage tolerance [4]; body-centered cubic (BCC) MPEAs with superior strength [5,6]; and hexagonal-close-packed (HCP) MPEAs combined with rare-earth elements [7]. Unfortunately, most high-strength BCC MPEAs so far have been affected by a low ductility at room temperature, suggesting a BCC strength–ductility trade-off dilemma [8,9]. For example, Senkov et al. reported that systematic BCC Al-Mo-Nb-Ta-Ti-Zr MPEAs exhibited a room-temperature fracture strain of ≤4.1%; although, among which, Al0.5Mo0.5NbTa0.5TiZr alloy possessed a highest compressive strength of 2350 MPa [5]. Research efforts into improving the room-temperature ductility of BCC-based MPEAs should be urgent, for further industrial applications.
Low-density (e.g., lightweight) designs of MPEAs with high performance and energy efficiency have been extensively reported recently, promoting the development of second-generation structural materials [10]. In addition to the previously published AlLiMg-based FCC MPEAs with low density (<3 g·cm3) [11], another type of low-density MPEAs with BCC structure has been designed through selecting elements with a light weight (e.g., Al and Ti) and high melting points (Cr, Nb, V, and Zr) [12,13]. For example, Yurchenko et al. revealed that ordered BCC (B2) AlNbTiVZrx (x = 0−1.5) alloys possessed a low density of 5.53–5.87 g·cm3 and high room-temperature yield strength of 1000–1535 MPa [6]. Senkov et al. proposed a Cr-Nb-Ti-V-Zr alloy system that had an excellent yield stress, even at 1000 °C [12,14]. Among these Cr-Nb-Ti-V-Zr alloys, BCC NbTiVZr and NbTiV2Zr could be plastically deformed up to 50% of strain without fracture, but two Cr-added CrNbTiZr and CrNbTiVZr alloys had limited fracture strains of 6% and 3%, respectively [14]. This is because the Cr element added to the two BCC alloys promotes the formation of Laves secondary phase, thereby deteriorating the ductility. Low-density MPEAs with a single-phase BCC structure usually possess a better ductility than those with multiple phases [8,15]. Fortunately, light Al elements, differently from the heavy Cr element, can stabilize the BCC structures of MPEAs and simultaneously lighten their weight [16]. Thus, our previous publication, replacing Cr by Al in CrNbTiZr alloy, found that the density of a B2-based AlNbTiZr with mirror Zr5Al3-type phase was reduced to ~5.85 g·cm3 and the Laves secondary phase disappeared [17]. Meanwhile, Qiu et al. reported that the B2 phase is more stable than the disordered BCC phase [13]. Moreover, a light Ti element was demonstrated to be a BCC/B2 stabilizer [18,19]. To overcome the poor ductility in most lightweight Al-containing MPEAs, an increment of Ti content, to stabilize single-phase B2 structure, may provide an effective pathway for a better strength–ductility combination at room temperature. Here, the effect of light Ti content on the microstructure and mechanical properties of AlNbZrTix MPEAs was investigated, aiming to simultaneously obtain a low density, high specific yield strength (strength-to-density ratio, SYS), and superior ductility at room temperature.

2. Materials and Methods

The AlNbZrTix (x = 1, 1.5, 2, 3) samples were prepared by vacuum induction melting. Each ingot was melted at least five times with pure Al, Nb, Zr, and Ti (≥99.9 wt.%). The produced samples were annealed at 1473 K for 5 h in a vacuumed quartz tube, and subsequently furnace cooled to room temperature for further investigation [6,12,17]. The crystal structure of the annealed samples was characterized using X-ray diffraction (XRD, Bruker D8 Advance, Karlsruhe, Germany) with Cu-Kα radiation. XRD measurement was conducted from 20° to 100° at a scanning rate of 4°/min. The microstructure of the annealed samples was studied using transmission electron microscopy (TEM, FEI Titan Themis, Hillsboro, OR, USA), scanning transmission electron microscopy (STEM), and scanning electron microscopy (SEM, Zeiss Sigma 350, Jena, Germany). Thin samples for TEM observations were prepared by ion milling using 3 keV Ar ions and cooled by liquid N2 during ion milling. The SEM was accompanied by backscatter electron (BSE) technology, energy-dispersive spectrometer (EDS) technology, and electron backscatter diffraction (EBSD, Oxford Aztec, Abingdon, UK) technology. The samples for SEM-BSE observation were treated by mechanical polishing, and those for EBSD measurement were electro-polished in a solution of 10% perchloric acid and 90% alcohol at sub-zero temperatures [17]. The EBSD test was performed at scanning steps of 5 μm.
The density of the annealed samples was measured using the hydrostatic weighting method, and three samples of each alloy were tested to obtain the average density value. The samples for compression tests were cut from annealed ingots with dimensions of ~4.7 × 4.7 × 7.7 mm3. Compression tests of at least three specimens were carried out at room temperature, using a strain rate of 1 × 10−4 s−1. The microstructure measurement of the 20%-strained AlNbZrTi3 sample was conducted on the longitudinal section.

3. Results and Discussion

3.1. Phase Formation

The phase formation of alloys is dependent on the Gibbs free energy (∆Gmix) formula, as presented in Equation (1) [20]. The negative ∆Gmix is associated with reducing the mixing enthalpy (∆Hmix, Equation (2)) and increasing the mixing entropy (∆Smix, Equation (3)), which favors the formation of a single-phase solid solution. Related abbreviations/nomenclature is shown in Table A1. Owing to the component complexity of MPEAs, this explanation has its limitations. In order to predict the phase formation for MPEA design, three parameters, δ, Ω and VEC, were proposed [21,22]. The above parameters were defined as follows:
Δ G mix = Δ H mix T m Δ S mix ,
Δ H mix = i = 1 , i j n 4 CiCj Δ H mix ij ,
Δ S mix = R i = 1 n c i lnc i ,
T m = i = 1 n c i T i ,
δ = ( i = 1 n c i ( 1 r i / i = 1 n c i r i ) 2 ) 1 / 2 ,
Ω = T m Δ S mix / | Δ H mix | ,
VEC = i = 1 n c i ( VEC ) i ,
where Tm is the average calculated melting point, ci and cj represent atomic fraction of the ith and jth constituents, Δ H m i x i j denotes ith-jth-constituent mixing enthalpy, R is gas constant, Ti is melting point of ith pure metal, δ denotes atomic misfit, ri is ith-constituent radius, Ω represents a multi-component solid solution rule, and VEC is the valence electron concentration. When Ω ≥ 1.1 and δ ≤ 6.6%, a single-phase solid-solution structure becomes easier to form [22,23,24]. Additionally, VEC is helpful for predicting a solid–solution structure. Guo et al. [25] revealed that the range of VEC < 6.87, 6.87 ≤ VEC ≤ 8, and VEC ≥ 8 favors the stabilization of BCC and FCC, and the coexistence of BCC and FCC phases, respectively.
The ith-jth-constituent mixing enthalpy in the AlNbZrTix alloys is listed in Table 1. The parameters calculated according to Equations (2) to (7) are listed in Table 2. With the increment of the Ti in these MPEAs, the δ critical value below of 6.6% decreased from 3.85% (x = 1) to 3.34% (x = 3), by 13.2%, and the Ω increased from 1.04 (x = 1), to exceed the critical value of 1.1 (x = 1.5, 2, 3). This suggests the easier formation of a single-phase solid–solution structure. Additionally, the calculated VEC value (4) of these alloys belongs to the VEC range (<6.87), which favors the stabilization of the BCC structure. Thus, this means that the Ti element stabilized the single-phase BCC structure in the AlNbZrTix alloys.
Interestingly, a specialty of a Ω value (1.04) below the critical value (1.1) appears in AlNbZrTi alloy, which promotes the ability to form a multi-phase structure. Our previous report confirmed that the AlNbZrTi alloy exhibited dual phases (B2 and Zr5Al3-type structure) [17]. It is speculated that the lower mixing enthalpy of Al-Zr components (Table 1) is responsible for the Zr5Al3-type phase. Meanwhile, the ordering of the BCC structure observed in Al-Ti-containing MPEAs, i.e., AlNbTiVZrx [6] and AlTiVCr [13], was ascribed to the negative mixing enthalpy of Al with other elements (Table 1). This becomes a dominant factor in facilitating the formation of B2. Based on the site occupations in ordered Ti alloys [6,27], B2 AlNbZrTix MPEAs should have one sublattice preferably occupied by Ti and the other sublattice preferably occupied by Al, while Nb and Zr are present on both sublattices.

3.2. Microstructural Characterization

The XRD patterns of the AlNbZrTix alloys are shown in Figure 1. It can be seen that AlNbZrTi alloy contains a combination of a dominant BCC and minor Zr5Al3-type structure, as also identified in our previous report [17]. The superlattice peak at ~27° demonstrates the ordering of the BCC structure (CsCl-type B2, Pm3m). As the Ti content increases in AlNbZrTix (x = 1.5−3), the XRD intensity of the Zr5Al3-type phase decreases (x = 1.5) and finally disappears (x = 2, 3). It can be speculated that the Ti increment in these MPEAs results in the formation of a single-phase B2 structure. Additionally, the {110}B2 peak at ~38° reveals a right shift and suggests a reduction in the lattice parameters of the B2 phase, which were calculated to be ~0.3336 nm, ~0.3325 nm, ~0.3317 nm, and ~0.3309 nm in the AlNbZrTix (x = 1, 1.5, 2, 3) alloys, respectively. This reduction of B2 lattice parameters is attributed to a weakened lattice distortion induced by a reduced δ value (Table 2).
To understand the phase variations, SEM-BSE images of AlNbZrTix (x = 1–3) alloys are shown in Figure 2. As seen in Figure 2a, the AlNbZrTi alloy possesses an equiaxed-grain B2 matrix (Region I), accompanied b polygonal (Region II) and rod-like (Region III) phases. The polygonal phase is at the matrix interior, while the rod-like phase is distributed discontinuously along the B2 grain boundaries. In AlNbZrTi1.5 alloy, the polygonal and rod-like phase becomes irregular and forms a continuous net along the B2 grain boundaries, respectively (Figure 2b). Additionally, the amount of secondary phase decreases. In AlNbZrTi2 and AlNbZrTi3 alloys, the B2 matrix remains equiaxed and the secondary phase disappears. This trend in the amount of secondary phase is consistent with the XRD peak intensity of Zr5Al3-type structure in Figure 1. Meanwhile, the EDS results, as listed in Table 3, reveal that the matrix composition was close to the nominal composition; where the deviation between the matrix and nominal composition is less than ~7.6%. Both the polygonal and rod-like phases are enriched in Zr and Al, and the content of Zr is greater than that of Al. This Zr-Al-rich phenomenon confirms the Zr5Al3-type structure of the secondary phase shown in Figure 2. This Zr5Al3-type phase is also observed in AlNbTiVZrx (x = 0.1–1.5) alloys, because the enthalpy of the binary Zr5Al3-type phase formation was very negative (−51.5 kJ·mol−1) [6,28].
Al-Ti-containing MPEAs, such as Al0.5NbTa0.8Ti1.5V0.2Zr [29] and Al0.25NbTaTiZr [5], have been reported to usually possess a nanoscale BCC/B2 mixture with similar lattice parameters, which is difficult to identify by XRD analysis. To distinguish disordered BCC and/or B2, an aberration-corrected STEM high-angle annular dark-field (HAADF) image of AlNbZrTi3 is present in Figure 3. The 1/2{100} lattice fringes with lightly dark contrast and the corresponding Fourier transformation (FFT) inset show a weak {100} superlattice spot, confirming a single B2 structure rather than a nanoscale BCC/B2 mixture. The spacing of (100) and ( 01 1 ¯ ) planes in Figure 3 indicates a lattice parameter of ~0.34 nm, in approximate agreement with the XRD results. The alternately dark and bright distribution of the 1/2{100} lattice fringes confirms the preferable entrance of light Al atoms into one sublattice of the CsCl-type B2 structure. In summary, when the Ti content is increased in AlNbZrTix (x = 1~3), the microstructure transforms from a mixture of B2 matrix and mirror Zr5Al3-type phase, to a single-phase B2 structure, confirming that the Ti element stabilizes the B2 structure rather than the B2/BCC mixture.

3.3. Mechanical Property: Compressive Stress–Strain Analysis

Compressive engineering stress–strain curves of the AlNbZrTix (x = 1−3) alloys at room temperature are presented in Figure 4a. As the Ti content increases, the yield strength continuously declines, from ~1579 MPa (x = 1) to ~1111 MPa (x = 3). The corresponding mechanical properties are also listed in Table 4. The downward trend in the yield strength of AlNbZrTix alloys is explained by two aspects. First, the strengthening contribution of the brittle Zr5Al3-type phase to the yield strength becomes less efficient, because the amount of Zr5Al3-type hard phase decreases as the Ti content increases in the MPEAs (Figure 2). Second, a drop in ∆Hmix (Table 2) suggests a lower chemical affinity of the multiple elements in the B2 matrix, which may be responsible for the decreased yield strength [26].
As shown in Figure 4a and Table 4, the fracture strain of the AlNbZrTix (x = 1, 2, 3) alloys were ~17.8%, 21.8%, and >50%, respectively, except for AlNbZrTi1.5, which had a poor fracture strain of ~7.8%. Note that no macroscopic fracture was observed in the AlNbZrTi3 alloy before 50% strain. In these MPEAs consisting of a B2 matrix and Zr5Al3-type phase (Figure 2), the former is relatively softer and provides a larger contribution to the ductility than the latter. It is well known that a high δ in the MPEAs is a dominant factor in causing severe lattice distortion, which would deteriorate the ductility [10]. Thus, as the Ti content increases in the AlNbZrTix (x = 1, 2, 3) alloys, the improvement of the fracture strain can be attributed to the reduced δ (Table 2) and gradual disappearance of the Zr5Al3-type phase. Formation of a single-phase B2 structure is beneficial for improving the ductility in these MPEAs. Exceptionally, compared to AlNbZrTi alloy, AlNbZrTi1.5 alloy possesses a lower δ, but a poor fracture strain. The morphology and distribution transition of a hard Zr5Al3-type phase (Figure 2) provide a reasonable explanation. When Ti was added to AlNbZrTix alloys from x = 1 to x = 1.5, rod-like Zr5Al3-type phase becomes a continuous net distributed along the B2 grain boundaries and is a cause of stress-induced intergranular fracture; while irregular Zr5Al3-type phase at the matrix interior has some sharp tips, where the deformation-induced stress is concentrated. Consequently, the fracture strain becomes poor only in the AlNbZrTi1.5 alloy.
The actual densities of the AlNbZrTix (x = 1, 1.5, 2, 3) MPEAs were measured as ~5.85 g·cm−3, ~5.72 g·cm−3, ~5.56 g·cm−3, and ~5.34 g·cm−3, respectively. This suggests a lowered density of those MPEAs through the addition of the light-weight Ti element. These AlNbZrTix MPEAs are lighter than most reported MPEAs (≥6 g·cm−3) [6,8]. Regarding yield strength and actual density, the SYS versus compressive fracture strain of the AlNbZrTix and previously reported BCC/B2-dominant MPEAs are displayed in Figure 4b. As seen, the fracture strain of most single-phase BCCs (~5–20%) is better than most multi-phase MPEAs (≤ ~10%). The SYS of the AlNbZrTix (x = 1, 1.5, 2, 3) MPEAs were calculated as ~270 kPa·m3·kg−1, ~238 kPa·m3·kg−1, ~221 kPa·m3·kg−1, >208 kPa·m3·kg−1, respectively. Except for AlNbZrTi1.5 with a limited ductility of ~7.8%, the AlNbZrTix (x = 1, 2, 3) have, so far, a better combination of the SYS and fracture strain than any of the other BCC/B2-dominant alloys; with a fracture strain of <50% [30,31,32,33]. This suggests a combination of high strength, low density, and excellent ductility for the AlNbZrTix (x = 1, 2, 3) MPEAs.
Compared to binary B2 materials with inherent brittleness, e.g., CoTi, CoAl, and NiAl [34], the single-phase B2 AlNbZrTi3 alloy reported here has exceptional room-temperature ductility. To uncover the underlying compressive deformation mechanisms of the single-phase B2 structure, the microstructure of 20%-strained AlNbZrTi3 was examined using EBSD, TEM, and STEM-HAADF, as present in Figure 5. The EBSD inverse pole figure (IPF) image in Figure 5a shows elongated grains, in which the color gradients suggest compression-induced deformation bands. Notably, two typical deformation band zones, denoted by A and B in Figure 5a, have misorientation relations of 20°~25°<110> and 30°~40°<123>, which are inconsistent with potential twin relations in B2 materials [35]. This suggests the deformation bands are not twins. Similarly, Yamaguch et al. [34] also reported that superdislocation should exhibit a twinning sense. In Figure 5b, the EBSD kernel average misorientation (KAM) image shows local misorientations, mainly ranging from 0.2° to 1.8°, and delimits some lenticular-like structures spreading from the grain boundaries into the grain interior. The low-magnification TEM image in Figure 5c displays the two dislocation slip bands responsible for the local misorientations and for the lenticular-like structures, while the high-magnification TEM micrograph in the lower-left inset shows the dislocation morphology. This dislocation-mediated lenticular-like structure was also observed in the deformation of a previous BCC TiHfZrTaNb alloy [36]. The inset of the [011]B2 selected area electron diffraction (SAED) pattern in the upper-right corner in Figure 5c indicates that the phase structure is B2 or a B2/BCC mixture. Although the atomic resolution STEM-HAADF technique can only image the edge component of dislocations, but is not able to see screw dislocations, it can easily trace the antiphase domains caused by the motion of dislocations. As massive atomic resolution STEM-HAADF observations revealed few edge- or mixed-type dislocations, but antiphase domains were frequently observed, it is reasonable to suggest the existence of extensive screw dislocations. An example of an antiphase domain is shown in the [011]B2 STEM-HAADF micrograph in Figure 5d, in which the antiphase boundary is marked by a red line. Meanwhile, unlike the results of the phase structure indicated in the upper-right corner in Figure 5c, Figure 5d confirms a single-phase B2 structure with dark 1/2{100} lattice fringes in the compressed sample; indicating that the deformation process did not induce the B2/BCC transition, differently from the B2 AlNbTiVZrx alloys [6]. In summary, these results show that the remarkable compressive properties of the low-density B2 AlNbZrTix are closely related to a dislocation-based ductility mechanism accompanied by antiphase domains, as also demonstrated in other Al-Ti containing B2 MPEAs by Feuerbacher et al. [37].

4. Conclusions

The effect of Ti content on the microstructure and room-temperature mechanical properties of AlNbZrTix (x = 1−3) MPEAs was investigated here. The following conclusions were drawn:
(1)
As Ti content increases, the microstructure of the MPEAs transforms, from a mixture of B2 matrix and Zr5Al3-type phase (x = 1, 1.5) to a single-phase B2 structure (x = 2, 3). When x = 1, the Zr5Al3-type phase exhibits polygonal and rod-like shapes. When x = 1.5, the polygonal particles become irregular and the rod-like phase evolves into a continuous net along the B2 grain boundaries.
(2)
The Ti element in these alloys can stabilize a single-phase B2 structure, rather than a nanoscale BCC/B2 mixture, which can also be predicted by three previously-proposed parameters (Ω ≥ 1.1, δ ≤ 6.6%, and VEC < 6.87). These MPEAs exhibited a steady decrease in B2 lattice parameters, from ~0.3336 nm (x = 1) to ~0.3309 nm (x = 3), which is attributed to weaken lattice distortion caused by a reduced δ value.
(3)
The actual densities of the AlNbZrTix (x = 1, 1.5, 2, 3) MPEAs reduced by ~8.7%, from ~5.85 g·cm−3 (x = 1) to ~5.34 g·cm−3 (x = 3), and were lower than that of most reported MPEAs (≥6 g·cm−3). Thus, the specific yield strengths (SYS) of the AlNbZrTix (x = 1, 1.5, 2, 3) MPEAs were calculated to be ~270 kPa·m3·kg−1, ~238 kPa·m3·kg−1, ~221 kPa·m3·kg−1, >208 kPa·m3·kg−1, respectively.
(4)
The excellent fracture strain of AlNbZrTix (x = 1, 2, 3) alloys were ~17.8%, 21.8%, and >50%, respectively. This Ti-addition-induced ductility improvement was attributed to the reduced δ and gradual disappearance of the Zr5Al3-type phase. An exception was AlNbZrTi1.5 with a limited ductility of ~7.8%, which was explained through the distinct morphology and distribution of hard Zr5Al3-type phase.
(5)
The combined compressive properties (SYS and fracture strain) of AlNbZrTix (x = 1, 2, 3) were superior to the reported data of most BCC/B2-dominant MPEAs. The deformation mechanism of the B2 structure is closely related to a dislocation-based ductility mechanism, accompanied by antiphase domains. Our results provide an explanation for the high strength and exceptional room-temperature ductility of low-density B2 MPEAs.

Author Contributions

Conceptualization, Q.T.; methodology, Q.T.; software, H.S.; validation, S.P.; formal analysis, Q.T. and P.D.; investigation, W.C.; resources, Q.T.; data curation, W.C.; writing—original draft preparation, Q.T., S.P., and P.D.; writing—review and editing, H.S., S.P., and W.C.; supervision, Q.T.; project administration, Q.T.; funding acquisition, Q.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Fujian Province, China, grant number 2019J01812 and Scientific Research Project of Putian University, China, grant number 2021073.

Data Availability Statement

The data sets used or analyzed in the current study are available from the corresponding author upon reasonable request.

Acknowledgments

Xiaoming Zhang of ZKKF (Beijing) Science & Technology Co., Ltd. is acknowledged for TEM and STEM-HAADF observations.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Abbreviations/Nomenclature.
Table A1. Abbreviations/Nomenclature.
AbbreviationsNomenclatureDefinition
MPEA Multi-principal element alloy
FCC Face-centered cubic
BCC Body-centered cubic
HCP Hexagonal-close-packed
B2 Ordered body-centered cubic
SYS Specific yield strength (strength-to-density ratio)
XRD X-ray diffraction
TEM Transmission electron microscopy
STEM Scanning transmission electron microscopy
SEM Scanning electron microscopy
BSE Backscatter electron
EDS Energy-dispersive spectrometer
EBSD Electron backscatter diffraction
HAADF High-angle annular dark-field
FFT Fourier transformation
IPF Inverse pole figure
KAM Kernel average misorientation
SAED Selected area electron diffraction
∆GmixGibbs free energy
∆HmixMixing enthalpy
∆SmixMixing entropy
δAtomic misfit
ΩMulti-component solid solution rule
VECValence electron concentration
TmAverage calculated melting point
TiMelting point of ith pure metal
ciAtomic fraction of the ith constituent
cjAtomic fraction of the jth constituent
Δ H m i x i j ith-jth-constituent mixing enthalpy
RGas constant
riith-constituent radius
(VEC)iValence electron concentration of the ith constituent

References

  1. Wang, Z.; Wu, H.; Wu, Y.; Huang, H.; Zhu, X.; Zhang, Y.; Zhu, H.; Yuan, X.; Chen, Q.; Wang, S. Solving oxygen embrittlement of refractory high-entropy alloy via grain boundary engineering. Mater. Today 2022, 54, 83–89. [Google Scholar] [CrossRef]
  2. Lei, Z.; Liu, X.; Wu, Y.; Wang, H.; Jiang, S.; Wang, S.; Hui, X.; Wu, Y.; Gault, B.; Kontis, P.; et al. Enhanced strength and ductility in a high-entropy alloy via ordered oxygen complexes. Nature 2018, 563, 546–550. [Google Scholar] [CrossRef]
  3. Liu, F.; Liaw, P.K.; Zhang, Y. Recent Progress with BCC-Structured High-Entropy Alloys. Metals 2022, 12, 501. [Google Scholar] [CrossRef]
  4. Gludovatz, B.; Hohenwarter, A.; Catoor, D.; Chang, E.H.; George, E.P.; Ritchie, R.O. A fracture-resistant high-entropy alloy for cryogenic applications. Science 2014, 345, 1153–1158. [Google Scholar] [CrossRef] [Green Version]
  5. Senkov, O.N.; Jensen, J.K.; Pilchak, A.L.; Miracle, D.B.; Fraser, H.L. Compositional variation effects on the microstructure and properties of a refractory high-entropy superalloy AlMo0.5NbTa0.5TiZr. Mater. Des. 2017, 139, 498–511. [Google Scholar] [CrossRef]
  6. Yurchenko, N.Y.; Stepanov, N.D.; Zherebtsov, S.V.; Tikhonovsky, M.A.; Salishchev, G.A. Structure and mechanical properties of B2 ordered refractory AlNbTiVZrx (x = 0–1.5) high-entropy alloys. Mater. Sci. Eng. A 2017, 704, 82–90. [Google Scholar] [CrossRef] [Green Version]
  7. Ter-Isahakyan, A.; Rau, J.S.; Balk, T.J. High entropy alloys with hexagonal close-packed structure derived from thin film combinatorial approach. J. Alloys Compd. 2021, 893, 162293. [Google Scholar] [CrossRef]
  8. Miracle, D.B.; Senkov, O.N. A critical review of high entropy alloys and related concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef] [Green Version]
  9. Zýka, J.; Málek, J.; Veselý, J.; Lukáč, F.; Čížek, J.; Kuriplach, J.; Melikhova, O. Microstructure and Room Temperature Mechanical Properties of Different 3 and 4 Element Medium Entropy Alloys from HfNbTaTiZr System. Entropy 2019, 21, 114. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Y.; Liaw, P.K.; Zhang, Y. Microstructures and Properties of the Low-Density Al15Zr40Ti28Nb12M(Cr, Mo, Si)5 High-Entropy Alloys. Metals 2022, 12, 496. [Google Scholar] [CrossRef]
  11. Youssef, K.M.; Zaddach, A.J.; Niu, C.; Irving, D.L.; Koch, C.C. A Novel Low-Density, High-Hardness, High-entropy Alloy with Close-packed Single-phase Nanocrystalline Structures. Mater. Res. Lett. 2014, 3, 95–99. [Google Scholar] [CrossRef] [Green Version]
  12. Senkov, O.N.; Senkova, S.V.; Woodward, C.; Miracle, D.B. Low-density, refractory multi-principal element alloys of the Cr-Nb-Ti-V-Zr system: Microstructure and phase analysis. Acta Mater. 2013, 61, 1545–1557. [Google Scholar] [CrossRef]
  13. Qiu, Y.; Hu, Y.J.; Taylor, A.; Styles, M.J.; Marceau, R.K.W.; Ceguerra, A.V.; Gibson, M.A.; Liu, Z.K.; Fraser, H.L.; Birbilis, N. A lightweight single-phase AlTiVCr compositionally complex alloy. Acta Mater. 2017, 123, 115–124. [Google Scholar] [CrossRef] [Green Version]
  14. Senkov, O.N.; Senkova, S.V.; Miracle, D.B.; Woodward, C. Mechanical properties of low-density, refractory multi-principal element alloys of the Cr-Nb-Ti-V-Zr system. Mater. Sci. Eng. A 2013, 565, 51–62. [Google Scholar] [CrossRef]
  15. Kumar, A.; Gupta, M. An Insight into Evolution of Light Weight High Entropy Alloys: A Review. Metals 2016, 6, 199. [Google Scholar] [CrossRef] [Green Version]
  16. Senkov, O.N.; Senkova, S.V.; Woodward, C. Effect of aluminum on the microstructure and properties of two refractory high-entropy alloys. Acta Mater. 2014, 68, 214–228. [Google Scholar] [CrossRef]
  17. Chen, W.; Tang, Q.H.; Wang, H.; Xie, Y.C.; Yan, X.H.; Dai, P.Q. Microstructure and mechanical properties of a novel refractory AlNbTiZr high-entropy alloy. Mater. Sci. Technol. 2018, 34, 1309–1315. [Google Scholar] [CrossRef]
  18. Chen, Y.; Liu, W.; Wang, H.; Xie, J.; Zhang, T.; Yin, L.; Huang, Y. Effect of Ti Content on the Microstructure and Properties of CoCrFeNiMnTix High Entropy Alloy. Entropy 2022, 24, 241. [Google Scholar] [CrossRef]
  19. Qiu, X.W.; Zhang, Y.P.; Liu, C.G. Effect of Ti content on structure and properties of Al2CrFeNiCoCuTix high-entropy alloy coatings. J. Alloys Compd. 2014, 585, 282–286. [Google Scholar] [CrossRef]
  20. Hwl, A.; Yang, S.A.; Jfl, B.; Wlm, A.; Zdh, A.; Cs, C.; Kfy, A. Phase stabilities of high entropy alloys. Scr. Mater. 2020, 179, 40–44. [Google Scholar]
  21. Zhang, Y.; Zhou, Y.J.; Lin, J.P.; Chen, G.L.; Liaw, P.K. Solid-Solution Phase Formation Rules for Multi-component Alloys. Adv. Eng. Mater. 2008, 10, 534–538. [Google Scholar] [CrossRef]
  22. Yang, X.; Chen, S.Y.; Cotton, J.D.; Zhang, Y. Phase Stability of Low-Density, Multiprincipal Component Alloys Containing Aluminum, Magnesium, and Lithium. JOM 2014, 66, 2009–2020. [Google Scholar] [CrossRef]
  23. Yang, X.; Zhang, Y. Prediction of high-entropy stabilized solid-solution in multi-component alloys. Mater. Chem. Phys. 2012, 132, 233–238. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Yang, X.; Liaw, P.K. Alloy Design and Properties Optimization of High-Entropy Alloys. JOM 2012, 64, 830–838. [Google Scholar] [CrossRef]
  25. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of valence electron concentration on stability of fcc or bcc phase in high entropy alloys. J. Appl. Phys. 2011, 109, 103505. [Google Scholar] [CrossRef] [Green Version]
  26. Takeuchi, A.; Inoue, A. Mixing enthalpy of liquid phase calculated by miedema’s scheme and approximated with sub-regular solution model for assessing forming ability of amorphous and glassy alloys. Intermetallics 2010, 18, 1779–1789. [Google Scholar] [CrossRef]
  27. Kazantseva, N.V.; Sazonova, V.A.; Lyzhina, G.A. Effect of the annealing temperature on the long-range order in the B2 phase of the Ti-Al-Nb(Zr, Mo) alloy. Phys. Met. Metallogr. 2006, 102, 288–294. [Google Scholar] [CrossRef]
  28. Tao, W.; Jin, Z.; Zhao, J.C. Thermodynamic assessment of the Al-Zr binary system. J. Phase Equilib. Diff. 2001, 22, 544–551. [Google Scholar]
  29. Soni, V.; Senkov, O.N.; Gwalani, B.; Miracle, D.B.; Banerjee, R. Microstructural Design for Improving Ductility of An Initially Brittle Refractory High Entropy Alloy. Sci. Rep. 2018, 8, 8816. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, Y.; Liu, Y.; Li, Y.; Chen, X.; Zhang, H. Microstructure and mechanical properties of a refractory HfNbTiVSi0.5 high-entropy alloy composite. Mater. Lett. 2016, 174, 82–85. [Google Scholar] [CrossRef]
  31. Yao, H.W.; Qiao, J.W.; Gao, M.C.; Hawk, J.A.; Ma, S.G.; Zhou, H.F.; Zhang, Y. NbTaV-(Ti,W) refractory high-entropy alloys: Experiments and modeling. Mater. Sci. Eng. A 2016, 674, 203–211. [Google Scholar] [CrossRef]
  32. Senkov, O.N.; Senkova, S.V.; Meisenkothen, F.; Miracle, D.B.; Woodward, C.F. Microstructure and elevated temperature properties of a refractory TaNbHfZrTi alloy. J. Mater. Sci. 2012, 47, 4062–4074. [Google Scholar] [CrossRef]
  33. Juan, C.C.; Tsai, M.H.; Tsai, C.W.; Lin, C.M.; Wang, W.R.; Yang, C.C.; Chen, S.K.; Lin, S.J.; Yeh, J.W. Enhanced mechanical properties of HfMoTaTiZr and HfMoNbTaTiZr refractory high-entropy alloys. Intermetallics 2015, 62, 76–83. [Google Scholar] [CrossRef]
  34. Yamaguchi, M.; Umakoshi, Y. The deformation behaviour of intermetallic superlattice compounds. Prog. Mater. Sci. 1990, 34, 1–148. [Google Scholar] [CrossRef]
  35. Goo, E.; Duerig, T.; Melton, K.; Sinclair, R. Mechanical twinning in Ti50Ni47Fe3 and Ti49Ni51 alloys. Acta Metall. 1985, 33, 1725–1733. [Google Scholar] [CrossRef]
  36. Dirras, G.; Lilensten, L.; Djemia, P.; Laurent-Brocq, M.; Tingaud, D.; Couzinié, J.P.; Perrière, L.; Chauveau, T.; Guillot, I. Elastic and plastic properties of as-cast equimolar TiHfZrTaNb high-entropy alloy. Mater. Sci. Eng. A 2016, 654, 30–38. [Google Scholar] [CrossRef]
  37. Feuerbacher, M. Dislocations and deformation microstructure in a B2-ordered Al28Co20Cr11Fe15Ni26 high-entropy alloy. Sci. Rep. 2016, 6, 29700. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of the AlNbZrTix alloys.
Figure 1. XRD patterns of the AlNbZrTix alloys.
Metals 12 00932 g001
Figure 2. SEM-BSE images of the AlNbZrTix alloys. (a) AlNbZrTi, (b) AlNbZrTi1.5, (c) AlNbZrTi2, and (d) AlNbZrTi3.
Figure 2. SEM-BSE images of the AlNbZrTix alloys. (a) AlNbZrTi, (b) AlNbZrTi1.5, (c) AlNbZrTi2, and (d) AlNbZrTi3.
Metals 12 00932 g002
Figure 3. [011]B2 STEM-HAADF image of AlNbZrTi3 alloy with an inset FFT pattern.
Figure 3. [011]B2 STEM-HAADF image of AlNbZrTi3 alloy with an inset FFT pattern.
Metals 12 00932 g003
Figure 4. Compressive properties of the AlNbZrTix alloys at room temperature, (a) Engineering stress–strain curve; (b) Comparison of our SYS versus fracture strain to those of previously reported BCC/B2-dominant MPEAs.
Figure 4. Compressive properties of the AlNbZrTix alloys at room temperature, (a) Engineering stress–strain curve; (b) Comparison of our SYS versus fracture strain to those of previously reported BCC/B2-dominant MPEAs.
Metals 12 00932 g004
Figure 5. Microstructure of the 20%-strained AlNbZrTi3 alloy, (a) EBSD IPF image; (b) EBSD KAM image; (c) Low-magnification TEM image with an inset high-magnification TEM image at the lower-left corner and an inset SAED pattern at the upper-right corner; (d) [011]B2 STEM-HAADF image.
Figure 5. Microstructure of the 20%-strained AlNbZrTi3 alloy, (a) EBSD IPF image; (b) EBSD KAM image; (c) Low-magnification TEM image with an inset high-magnification TEM image at the lower-left corner and an inset SAED pattern at the upper-right corner; (d) [011]B2 STEM-HAADF image.
Metals 12 00932 g005
Table 1. The ith-jth-constituent mixing enthalpy in the AlNbZrTix alloys [26].
Table 1. The ith-jth-constituent mixing enthalpy in the AlNbZrTix alloys [26].
H m i x i j AlNbZrTi
Al-−18.2−43.7−29.5
Nb--3.92
Zr---−0.2
Ti----
Table 2. The calculated parameters of the AlNbZrTix alloys.
Table 2. The calculated parameters of the AlNbZrTix alloys.
AlloysTm (K)∆Hmix (kJ·mol−1)∆Smix (J·K−1·mol−1)δ (%)ΩVEC
AlNbZrTi1935.74−21.425011.533.851.044
AlNbZrTi1.51936.56−19.664211.383.701.124
AlNbZrTi21937.22−18.144011.083.571.184
AlNbZrTi31938.21−15.677810.333.341.284
Table 3. EDS results of the AlNbZrTix alloys.
Table 3. EDS results of the AlNbZrTix alloys.
AlloysRegionChemical Compositions (at.%)
AlNbZrTi
AlNbZrTiNominal25252525
I23.124.725.227
II34.212.941.111.8
III35.112.841.210.9
AlNbZrTi1.5Nominal22.222.222.233.4
I21.921.522.733.9
II33.512.938.315.3
III28.712.940.917.5
AlNbZrTi2Nominal20202040
I19.919.320.540.3
AlNbZrTi3Nominal16.716.716.749.9
I16.317.016.949.8
Table 4. The actual density and mechanical properties of the AlNbZrTix alloys.
Table 4. The actual density and mechanical properties of the AlNbZrTix alloys.
AlloysActual Density (g·cm−3)Yield Strength (MPa)Fracture Strain (%)Strength-to-Density Ratio (SYS, kPa·m3·kg−1)
AlNbTiZr5.85157917.8270
AlNbTiZr1.55.7213647.8238
AlNbTiZr25.56122721.8221
AlNbTiZr35.341111>50>208
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tang, Q.; Su, H.; Peng, S.; Chen, W.; Dai, P. Microstructure and Mechanical Properties of Low-Density, B2-Ordered AlNbZrTix Multi-Principal Element Alloys. Metals 2022, 12, 932. https://doi.org/10.3390/met12060932

AMA Style

Tang Q, Su H, Peng S, Chen W, Dai P. Microstructure and Mechanical Properties of Low-Density, B2-Ordered AlNbZrTix Multi-Principal Element Alloys. Metals. 2022; 12(6):932. https://doi.org/10.3390/met12060932

Chicago/Turabian Style

Tang, Qunhua, Honghong Su, Shilong Peng, Wei Chen, and Pinqiang Dai. 2022. "Microstructure and Mechanical Properties of Low-Density, B2-Ordered AlNbZrTix Multi-Principal Element Alloys" Metals 12, no. 6: 932. https://doi.org/10.3390/met12060932

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