Next Article in Journal
Process Temperature Control for Low Dishing in CMP
Previous Article in Journal
Incorporation of Nitinol (NiTi) Shape Memory Alloy (SMA) in Concrete: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Mechanical Properties of Hot-Rolled ZrAl14Ti3 and ZrAl14Ti9 Alloys

College of Physical Science and Technology, Hebei Normal University of Science & Technology, Qinhuangdao 066004, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(19), 4459; https://doi.org/10.3390/ma18194459
Submission received: 24 August 2025 / Revised: 17 September 2025 / Accepted: 19 September 2025 / Published: 24 September 2025
(This article belongs to the Section Metals and Alloys)

Abstract

This study systematically investigated the microstructure and mechanical properties of hot-rolled and quenched ZrAl14Ti3 and ZrAl14Ti9 (at.%) alloys. Microstructural analysis revealed that both alloys consisted of equiaxed α-Zr and Zr3Al grains. Increasing Ti content lowered the dissolution temperature of Zr3Al in α-Zr, enhancing the solubility of Al in α-Zr under identical thermal conditions and decreasing the Zr3Al phase fraction. Moreover, higher Ti content in the ZrAl14Ti9 alloy significantly promoted Zr3Al recrystallization and α-Zr globularization, leading to grain refinement and complete elimination of the α-Zr basal texture. Mechanical property evaluation showed that the ZrAl14Ti3 alloy exhibited offset yield and tensile strengths of 888 ± 12 MPa and 1056 ± 19 MPa, respectively, with a fracture elongation of 23 ± 1%. The ZrAl14Ti9 alloy displayed enhanced strength without compromising ductility, achieving a 110 MPa increase in offset yield strength (998 ± 6 MPa) while maintaining the same fracture elongation (23 ± 2%). The strengthening effects observed in the ZrAl14Ti9 alloy stemmed from multiple synergistic mechanisms: solid-solution strengthening due to increased Ti content in α-Zr, refinement of both Zr3Al and α-Zr grains, a higher proportion of the harder α-Zr phase, and orientation hardening resulting from the elimination of the α-Zr basal texture.

1. Introduction

Zirconium (Zr) alloys are widely used as structural materials in nuclear reactors due to their exceptional corrosion resistance, low thermal neutron absorption cross-section, and excellent high-temperature mechanical properties [1]. Additionally, their low density and thermal expansion coefficient render them promising candidates for aerospace applications [2]. However, currently commercialized Zr alloys, primarily designed for nuclear applications with strict constraints on thermal neutron economy and corrosion resistance, are limited in the variety and concentration of alloying elements, with the hexagonal close-packed (HCP) α-Zr dominating their microstructure [1,2,3,4,5]. The α-Zr phase exhibits a high critical resolved shear stress (CRSS) for activating <a + c>-type slip and twinning at room temperature, which accommodate strain along the c-axis [6]. This intrinsic characteristic results in insufficient slip and twinning system activation during deformation, failing to meet the von Mises criterion (requiring five independent slip systems for homogeneous deformation), and consequently leads to inferior plasticity compared to high-symmetry phases such as face-centered cubic (FCC) and body-centered cubic (BCC) structures. Therefore, achieving a balanced strength–ductility synergy in α-Zr alloys remains challenging. Coarse-grained pure Zr exhibits a tensile elongation of up to 24%, but its tensile strength is limited to approximately 380 MPa [7]. Solid-solution strengthening, grain refinement, and interface strengthening can increase the strength of α-Zr alloys to 800 MPa; however, this is typically accompanied by a significant reduction in ductility (<10%) [7]. For non-nuclear applications, where alloying constraints are less stringent than in the nuclear industry, simultaneous enhancement of strength and ductility may be achieved by introducing ductile secondary phases into α-Zr matrix. Inspired by titanium (Ti) alloy design principles, incorporating BCC-structured β-Zr phase into α-Zr alloys was initially considered a viable strategy. Although Zr and Ti belong to the same group with identical outer electron configurations, their alloying behaviors differ significantly [8]. While β-stabilizing elements can effectively retain the metastable β-phase in Ti alloys through quenching, their introduction in Zr alloys typically induces the formation of intermetallic compounds or the ω-phase, even under rapid cooling rates (>100 °C/s) [8]. This fundamental discrepancy suggests that β-Zr may not serve as an ideal ductile phase in Zr-based alloys.
The Zr3Al phase, possessing an L12 structure, differs from other intermetallics by exhibiting an increase in electrical resistivity with rising temperature, indicative of a metallic nature [9]. Schulson conducted systematic studies on Zr3Al-based alloys, revealing that, in addition to good corrosion resistance and low neutron absorption, Zr3Al deforms at room temperature via a/3<211> partial dislocations, accompanied by stacking faults, demonstrating reasonable plastic deformability, albeit with limited strength [10]. Therefore, the Zr3Al phase could potentially serve as a ductile phase in α-Zr alloys to enhance plasticity. Moreover, aluminum (Al), as an economical industrial metal, not only significantly strengthens α-Zr but also reduces alloy density and raw material costs [11]. Chen et al. demonstrated that adding 14 at.% Al to a ZrSn1.5 (at.%) alloy introduced a substantial Zr3Al phase fraction, achieving simultaneous enhancement of strength and ductility [12].
Ti exhibits complete solid solubility in Zr at room temperature, providing effective solid-solution strengthening while reducing material density and cost [8]. Based on this background, the present study investigated the microstructure and mechanical properties of hot-rolled ZrAl14Ti3 (at.%) and ZrAl14Ti9 (at.%) alloys, focusing on the effects of Ti content variation on their strengthening mechanisms.

2. Experimental Procedure

Both alloys were fabricated from industrial-grade raw materials, including commercially pure Zr (Zr + Hf ≥ 99.8 wt.%), Al (99.9 wt.%), and Ti (99.9 wt.%). Each alloy was synthesized by non-consumable arc melting under an argon atmosphere, followed by six remelting cycles to ensure chemical homogeneity. The as-cast ingots were homogenized at 800 °C for 30 min in a muffle furnace, followed by multi-pass rolling with 5 min interpass holding times. A total thickness reduction of 60% was achieved prior to water quenching. Phase identification of the hot-rolled and water-quenched alloys was performed using X-ray diffraction (XRD, Rigaku SmartLab, Tokyo, Japan). Crystallographic orientation mapping was conducted using a Zeiss GeminiSEM 450 field-emission scanning electron microscope (FEG-SEM, Oberkochen, Germany) equipped with an EDAX Velocity electron backscatter diffraction (EBSD, Mahwah, NJ, USA) detector. Identical acquisition parameters were employed for comparative analysis: an accelerating voltage of 20 kV, a step size of 0.5 μm, and a scan area of 60 μm × 50 μm. The EBSD datasets for both alloys were reindexed using the spherical indexing algorithm, with over 97% of the points in each dataset exhibiting confidence index (CI) values above 0.1. All crystallographic analyses were performed using OIM Analysis™ 9.1 software. The EBSD specimens were prepared from the transverse direction (TD plane) of the hot-rolled alloys by mechanical polishing, followed by electrochemical polishing in a solution of 10 vol% perchloric acid in methanol at 20 V and −20 °C for 120 s. Electron channeling contrast imaging (ECCI) was carried out using a Hitachi SU7000 FEG-SEM (Hino, Japan). The ECCI specimen (TD plane) was prepared by mechanical polishing. Quasi-static tensile tests at room temperature were conducted following the ASTM E8/E8M standard [13] guidelines using an Instron 5982 universal testing machine (Norwood, MA, USA) with dog-bone specimens having gauge dimensions of 20 × 4 × 2 mm3. The tests were performed at an initial strain rate of 0.5 × 10−3 s−1, with strain measured using a mechanical extensometer with a 10 mm gauge length.

3. Results and Discussion

3.1. XRD Results

Figure 1a displays the XRD patterns of the hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys, while Figure 1b presents a magnified view of the 2θ range from 30° to 40°. Phase analysis confirmed the presence of α-Zr and Zr3Al phases in both alloys, with no detectable diffraction peaks corresponding to the β-phase (BCC) or other metastable phases. Furthermore, a distinct shift in the α-Zr diffraction peaks toward higher angles was observed in the magnified patterns with increasing Ti content. This peak shift is attributed to changes in the lattice parameters of the solid-solution phase caused by the incorporation of solute atoms. The atomic radii of Ti and Al are smaller than that of Zr. Their incorporation into the α-Zr matrix reduces the lattice parameters, resulting in the observed peak shift. The extent of the peak shift correlates with the solute content: a greater displacement toward higher angles indicates higher concentrations of Ti or Al atoms dissolved in the α-Zr lattice.

3.2. Microstructure Evolution

Figure 2(a1) shows the inverse pole figure (IPF) map of the hot-rolled ZrAl14Ti3 alloy, and Figure 2(a2) presents the corresponding phase map overlaid with CI values. The alloy exhibits a dual-phase microstructure consisting of equiaxed α-Zr and Zr3Al grains, with phase fractions of 66.0 vol.% α-Zr and 34.0 vol.% Zr3Al. A high density of 60°/<111> twin boundaries was observed within the Zr3Al phase. According to Gagné, these twin boundaries originate from stacking faults introduced during recrystallization, confirming that the Zr3Al phase underwent dynamic recrystallization during hot rolling [14]. Moreover, the Zr3Al grains exhibit two distinct morphologies: (i) elongated coarse grains (>10 μm in length) and (ii) fine equiaxed grains ranging from sub-micron size to a few micrometers. Figure 2(b1,b2) present the IPF map and the CI-overlaid phase map of the hot-rolled ZrAl14Ti9 (at.%) alloy, respectively. Similarly to the ZrAl14Ti3 alloy, the ZrAl14Ti9 alloy also exhibits an equiaxed α-Zr + Zr3Al dual-phase microstructure, but with a lower Zr3Al fraction of 28.9 vol.%. The increase in Ti content suppresses the precipitation temperature of Zr3Al from α-Zr, thereby increasing the solid solubility of Al in the α-Zr matrix at the rolling temperature and reducing the equilibrium Zr3Al fraction. In the ZrAl14Ti9 alloy, the Zr3Al grains are more uniformly distributed, with fewer coarse elongated grains compared to those in the ZrAl14Ti3 alloy. In addition, 60°/<111> twin boundaries remain prevalent in the Zr3Al phase. Quantitative analysis shows that the Zr3Al phase in the ZrAl14Ti3 alloy has a twin boundary density of 0.509 μm−1 (518.80 μm boundary length/1020 μm2 area), whereas the ZrAl14Ti9 alloy exhibits a higher value of 0.609 μm−1 (528.3 μm/867 μm2). The increased twin boundary density indicates a higher degree of recrystallization in the Zr3Al phase of the ZrAl14Ti9 alloy.
Figure 3 presents the kernel average misorientation (KAM) maps of the Zr3Al and α-Zr phases in the ZrAl14Ti3 and ZrAl14Ti9 alloys. In the ZrAl14Ti3 alloy, coarse and elongated Zr3Al grains (Figure 3(a1)) exhibit higher KAM values with an inhomogeneous distribution; regions of significantly elevated KAM (indicated by red arrows) alternate with areas of relatively low KAM (marked by green arrows). Fine and equiaxed Zr3Al grains consistently exhibit lower KAM values. The Zr3Al phase in the ZrAl14Ti9 alloy shows similar KAM distribution characteristics (Figure 3(b1)): coarse grains retain high and unevenly distributed KAM values, whereas fine grains exhibit uniformly low values. These results indicate that dynamic recrystallization has extensively occurred in the fine Zr3Al grains of both alloys, whereas coarse grains have undergone partial or no recrystallization. In the ZrAl14Ti3 alloy (Figure 3(a2)), the α-Zr contains a high density of subgrain boundaries (highlighted by yellow arrows), whereas the ZrAl14Ti9 alloy (Figure 3(b2)) exhibits significantly fewer such boundaries. Globularization of α-Zr in Ti and Zr alloys is a strain-induced geometric transformation fundamentally different from the conventional nucleation–growth mechanism of dynamic recrystallization. It proceeds through three stages: (i) kinking and shear-band formation within α-lamellae, (ii) subdivision of lamellae by low-angle dislocation walls, and (iii) globularization driven by minimization of interfacial energy [15,16,17,18,19]. During the second stage, low-angle boundaries (LABs) form as geometrically necessary dislocations accumulate into walls that subdivide the lamellae. High-angle grain boundaries (HABs) develop through progressive misorientation accumulation within these walls. Therefore, it is concluded that the α-Zr in the ZrAl14Ti9 alloy has experienced a more advanced stage of globularization compared with that in the ZrAl14Ti3 alloy. Quantitative KAM analysis (Figure 4) shows a marked decrease in the average KAM value of the Zr3Al phase, from 0.17° in the ZrAl14Ti3 alloy to 0.07° in the ZrAl14Ti9 alloy. In contrast, the average KAM values of the α-Zr decrease only slightly, from 0.08° to 0.05°. Although subgrain boundaries exhibit high KAM values, the low KAM within α-Zr grains in the ZrAl14Ti3 alloy results in only a marginal increase in the overall average KAM.
Figure 5(a1,b1) presents the misorientation angle distributions of the Zr3Al phase in the ZrAl14Ti3 and ZrAl14Ti9 alloys, respectively. Both alloys exhibit a pronounced peak near 60°, corresponding to the high density of 60°/<111> twin boundaries. The ZrAl14Ti9 alloy exhibits a markedly greater fraction of 60° misorientations, whereas the ZrAl14Ti3 alloy contains a higher proportion of subgrain boundaries (<5°). This indicates that the Zr3Al phase in the ZrAl14Ti3 alloy underwent a lower degree of recrystallization, preserving numerous subgrain boundaries generated during thermomechanical processing. For the α-Zr phase (Figure 5(a2,b2)), the ZrAl14Ti3 alloy exhibits a higher fraction of low-angle subgrain boundaries, whereas the ZrAl14Ti9 alloy is dominated by high-angle grain boundaries. This finding is consistent with the KAM results. Figure 6 presents the grain size statistics for the α-Zr and Zr3Al phases in both alloys. In the ZrAl14Ti3 alloy, the α-Zr exhibits an average grain size of 3.1 ± 1.6 µm, while the Zr3Al phase has an average grain size of 2.1 ± 1.3 µm. In the ZrAl14Ti9 alloy, the corresponding grain sizes are 2.6 ± 1.2 µm for α-Zr and 1.5 ± 0.7 µm for Zr3Al. The enhanced extent of recrystallization and globularization contributes to this grain refinement.
Figure 7a,b presents the pole figures of the Zr3Al phase in the ZrAl14Ti3 and ZrAl14Ti9 alloys, respectively. In both alloys, the Zr3Al exhibits a uniform orientation distribution, indicating the absence of pronounced texture. Figure 8a,b presents the pole figures of the α-Zr phase in the ZrAl14Ti3 and ZrAl14Ti9 alloys. In the ZrAl14Ti3 alloy, the α-Zr exhibits a strong basal texture, with a peak multiple of random distribution (MRD) value of 11.45. In contrast, the ZrAl14Ti9 alloy exhibits a much weaker texture, with a peak MRD of only 2.34, indicating no significant crystallographic preference. The basal texture in α-Zr alloys during hot rolling develops through an initial stage dominated by {1 0 −1 2} tensile twinning, which reorients the c-axis towards the normal direction (ND) and generates high-stored-energy twin boundaries that serve as preferential nucleation sites, followed by recrystallization-driven selective growth of ND-oriented grains, resulting in texture sharpening [19,20]. In the ZrAl14Ti9 alloy, subgrain boundaries within the α-Zr evolved into high-angle boundaries through dislocation absorption, resulting in the elimination of the basal texture. Figure 9(a1,b1) presents the Schmid factor maps for prismatic slip of the α-Zr phase in the ZrAl14Ti3 and ZrAl14Ti9 alloys, respectively, while Figure 9(a2,b2) shows the corresponding statistical distributions. The average Schmid factor for the ZrAl14Ti3 alloy is 0.43, significantly higher than the 0.32 measured for the ZrAl14Ti9 alloy. The lower Schmid factor of the ZrAl14Ti9 alloy indicates orientation hardening, attributed to the elimination of the basal texture.
Figure 10 and Figure 11 show the ECCI images of the ZrAl14Ti3 and ZrAl14Ti9 alloys at different magnifications. The low-magnification images reveal that both alloys contain coarse elongated and fine equiaxed Zr3Al grains. Higher-magnification images show that the α-Zr grains in both alloys exhibit a low dislocation density. Within the coarse Zr3Al grains of both alloys, some regions remain unrecrystallized (indicated by red arrows), containing a high density of dislocations, whereas other regions are recrystallized (indicated by green arrows) and contain fewer dislocations. These recrystallized and unrecrystallized regions within the coarse Zr3Al grains correspond, respectively, to the low- and high-KAM regions observed in the KAM maps. Furthermore, stacking faults (indicated by white arrows) and twin boundaries (indicated by blue arrows) are observed within the recrystallized Zr3Al grains in both alloys.
The influence of Ti content on the microstructure of ZrAl14-based alloys is primarily attributed to its effect on the initial martensitic transformation behavior. According to previous studies on Zr-Ti martensitic transformations [21,22], increasing Ti content induces a significant transition in the as-quenched microstructure. At lower Ti concentrations, the structure consists predominantly of dislocated lath martensite characterized by coarse packets of similarly oriented laths separated by low-angle boundaries. In contrast, higher Ti content promotes the formation of twinned plate martensite comprising fine, acicular plates belonging to multiple crystallographic variants separated by high-angle boundaries. This fundamental difference in initial microstructure profoundly affects subsequent thermomechanical processing. The fine, crystallographically diverse plate structure provides numerous high-energy nucleation sites for recrystallization, facilitating more complete recrystallization while inhibiting the development of strong texture and coarse macrozones. Consequently, alloys with higher Ti content exhibit enhanced recrystallization kinetics, finer equiaxed grains, and more randomized texture compared to their low-Ti counterparts.

3.3. Mechanical Properties

Figure 12 presents the engineering stress–strain curves of the hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys. The ZrAl14Ti9 alloy exhibited a yield plateau, indicative of discontinuous yielding, while the ZrAl14Ti3 alloy showed continuous yielding. The ZrAl14Ti3 alloy exhibits an offset yield strength (σ0.2) of 888 ± 12 MPa, an ultimate tensile strength (σb) of 1056 ± 19 MPa, and a fracture elongation (εf) of 23 ± 1%. Compared with ZrAl14Ti3, the ZrAl14Ti9 alloy shows an increase in offset yield strength by 110 MPa (to 998 ± 6 MPa) and in tensile strength by 13 MPa (to 1069 ± 7 MPa), while maintaining the same fracture elongation (23 ± 2%). These properties are highly competitive when compared to other Zr-based alloys reported in the literature [7,11,23,24,25,26], as visually summarized in the comparative plot provided in Figure 13. Microstructural characterization indicates that these variations in strength and ductility are associated with changes in α-Zr solid-solution composition, dislocation density, grain size, texture, and phase fraction. The EBSD results reveal that, relative to ZrAl14Ti3, the ZrAl14Ti9 alloy exhibits a higher degree of recrystallization in both α-Zr and Zr3Al phases, leading to lower geometrically necessary dislocation densities (Zr3Al: 5.28 × 1013 m−2 vs. 8.45 × 1013 m−2; α-Zr: 3.44 × 1014 m−2 vs. 3.79 × 1014 m−2). This dislocation reduction tends to decrease offset yield strength but enhances ductility. The Zr3Al phase fractions in ZrAl14Ti3 and ZrAl14Ti9 are 34.0% and 28.9%, respectively; given that the Zr3Al phase contains ~25 at.% Al and the total alloy Al content is 14 at.%, the nominal Al contents in the α-Zr matrix are calculated as 8.3 at.% and 9.5 at.%, respectively. The slight difference in Al content suggests that the pronounced shift in α-Zr diffraction peaks toward higher angles in XRD primarily results from increased Ti solubility, confirming the solid-solution strengthening effect of Ti. Another strengthening factor in ZrAl14Ti9 is grain refinement induced by enhanced recrystallization: the average grain sizes of α-Zr and Zr3Al phases decrease from 3.1 µm and 2.1 µm in ZrAl14Ti3 to 2.6 µm and 1.5 µm, respectively. Using Hall–Petch slopes of 280 MPa·µm1/2 [27] for α-Zr and 760 MPa·µm1/2 [28] for Zr3Al, the grain size changes contribute 19.2 MPa and 135.9 MPa, respectively, to the offset yield strength; considering the phase fractions, grain refinement contributes a net 52.9 MPa to the offset yield strength of ZrAl14Ti9. Furthermore, the α-Zr phase fraction increases from 66.0% to 71.1% in ZrAl14Ti9, providing additional strengthening since α-Zr is harder than Zr3Al. The lower average Schmid factor for basal slip in the α-Zr of ZrAl14Ti9 also implies increased orientation hardening. Overall, the offset yield strength improvement in ZrAl14Ti9 arises from solid-solution strengthening, grain refinement, an increased fraction of the harder α-Zr phase, and orientation hardening. Regarding ductility, although solid-solution strengthening and reduced Zr3Al content tend to lower fracture elongation, these effects are offset by grain refinement and the ductility improvement associated with reduced dislocation density.

4. Conclusions

This study systematically investigated the microstructure and mechanical properties of hot-rolled ZrAl14Ti3 and ZrAl14Ti9 (at.%) alloys. Key conclusions are summarized as follows:
(1)
The addition of Ti reduces the dissolution temperature of Zr3Al in the α-Zr matrix, enhancing Al solubility in α-Zr and decreasing the equilibrium Zr3Al phase fraction.
(2)
The addition of Ti promotes the recrystallization of the Zr3Al phase and the globularization of the α-Zr phase, refines the grains, and eliminates the basal texture in the α-Zr phase.
(3)
The ZrAl14Ti9 alloy exhibits an offset yield strength of 998 ± 6 MPa, tensile strength of 1069 ± 7 MPa, and fracture elongation of 23% ± 2. Compared with the ZrAl14Ti3 alloy, the offset yield strength is increased by 110 MPa, the tensile strength is increased by 13 MPa, while the fracture elongation remains unchanged. The main strengthening mechanisms include solid-solution strengthening, grain refinement, an increased fraction of the harder α-Zr phase, and orientation hardening.

Author Contributions

Conceptualization, Y.Y.; Methodology, Y.Y.; Software, M.Y.; Investigation, X.Z. and W.L.; Data curation, X.Z.; Writing—original draft, X.Z.; Writing—review & editing, W.L.; Visualization, M.Y.; Supervision, Z.L.; Funding acquisition, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was financially supported by the National Natural Science Foundation of China (No. 52201117).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jarugula, R.; Halodová, P.; Zimina, M.; Sundararaman, M.; Malá, M.; Klouzal, J.; Ševeček, M.; Běláč, J.; Řeháček, R.; Linhart, S. Effect of neutron irradiation on microstructural evolution and deformation behavior of Zirconium (Zr-1Nb) alloy. J. Nucl. Mater. 2025, 605, 155596. [Google Scholar] [CrossRef]
  2. Zhang, N.; Xia, C.; Qin, J.; Li, Q.; Zhang, X.; Liu, R. Research progress of novel zirconium alloys with high strength and toughness. J. Met. Mater. Miner. 2022, 32, 23–36. [Google Scholar] [CrossRef]
  3. Kautz, E.; Gwalani, B.; Yu, Z.F.; Varga, T.; Geelhood, K.; Devaraj, A.; Senor, D. Investigating zirconium alloy corrosion with advanced experimental techniques: A review. J. Nucl. Mater. 2023, 585, 154586. [Google Scholar] [CrossRef]
  4. King, D.J.M.; Knowles, A.J.; Bowden, D.; Wenman, M.; Capp, S.; Gorley, M.; Shimwell, J.; Packer, L.; Gilbert, M.; Harte, A. High temperature zirconium alloys for fusion energy. J. Nucl. Mater. 2022, 559, 153431. [Google Scholar] [CrossRef]
  5. Slobodyan, M. High-energy surface processing of zirconium alloys for fuel claddings of water-cooled nuclear reactors. Nucl. Eng. Des. 2021, 382, 111364. [Google Scholar] [CrossRef]
  6. Lecomte, J.S.; Crépin, J.; Barberis, P. Critical Resolved Shear Stress and Work Hardening Determination in HCP Metals: Application to Zr Single Crystals. Metals 2024, 14, 1101. [Google Scholar] [CrossRef]
  7. Ma, Z.; Liu, C.; Wu, J.; Yang, Z.; Liu, H.; Jia, X.; Liu, R.; Zhang, Y. Achieving high strength and plasticity in zirconium alloy by introducing bimodal microstructure composed of lamellar microstructure and recrystallized grains. Mater. Sci. Eng. A 2025, 924, 147794. [Google Scholar] [CrossRef]
  8. Banerjee, S.; Mukhopadhyay, P. Phase Transformations: Examples from Titanium and Zirconium Alloys; Elsevier: Oxford, UK, 2010. [Google Scholar]
  9. Schulson, E.M. The tensile and corrosion behaviour of ordered Zr3Al-based alloys. J. Nucl. Mater. 1974, 50, 127–138. [Google Scholar] [CrossRef]
  10. Schulson, E.M.; Roy, J.A. Work-hardening of the L12 phase Zr3Al. Acta Metall. 1978, 26, 15–28. [Google Scholar] [CrossRef]
  11. Zhang, X.; Zhang, B.; Liu, S.G.; Xia, C.; Zhang, X.; Ma, M.; Liu, R. Microstructures and mechanical properties of Zr–Al binary alloys processed by hot-rolling. Mater. Sci. Eng. A 2020, 773, 138723. [Google Scholar] [CrossRef]
  12. Chen, B.H.; Liu, S.G.; Zhang, X.Y.; Ma, M.; Liu, R. Effect of Al change on microstructure and mechanical properties of hot-rolled ZrSnAl alloy. Mater. Lett. 2020, 267, 127552. [Google Scholar] [CrossRef]
  13. ASTM E8/E8M-24; Standard Test Methods for Tension Testing of Metallic Materials. ASTM International: West Conshohocken, PA, USA, 2024.
  14. Gagne, M.; Schulson, E.M. The effects of annealing on cold-worked Zr3Al. Metall. Trans. A 1976, 7, 1775–1783. [Google Scholar] [CrossRef]
  15. Li, L.; Li, M.Q.; Luo, J. Mechanism in the β phase evolution during hot deformation of Ti–5Al–2Sn–2Zr–4Mo–4Cr with a transformed microstructure. Acta Mater. 2015, 94, 36–45. [Google Scholar] [CrossRef]
  16. Semiatin, S.L.; Seetharaman, V.; Weiss, I. Flow behavior and globularization kinetics during hot working of Ti–6Al–4V with a colony alpha microstructure. Mater. Sci. Eng. A 1999, 263, 257–271. [Google Scholar] [CrossRef]
  17. Semiatin, S.L.; Bieler, T.R. The effect of alpha platelet thickness on plastic flow during hot working of TI–6Al–4V with a transformed microstructure. Acta Mater. 2001, 49, 3565–3573. [Google Scholar] [CrossRef]
  18. Zherebtsov, S.; Murzinova, M.; Salishchev, G.; Semiatin, S. Spheroidization of the lamellar microstructure in Ti–6Al–4V alloy during warm deformation and annealing. Acta Mater. 2011, 59, 4138–4150. [Google Scholar] [CrossRef]
  19. Zhu, K.Y.; Bacroix, B.; Chauveau, T.; Chaubet, D.; Castelnau, O. Texture evolution and associated nucleation and growth mechanisms during annealing of a Zr alloy. Metall. Mater. Trans. A 2009, 40, 2423–2434. [Google Scholar] [CrossRef]
  20. Prakash, D.G.L.; Preuss, M.; Dahlbäck, M.; da Fonseca, J.Q. Microstructure and texture evolution during thermomechanical processing of β-quenched Zr. Acta Mater. 2015, 88, 389–401. [Google Scholar] [CrossRef]
  21. Banerjee, S.; Vijayakar, S.J.; Krishnan, R. Strength of zirconium-titanium martensites and deformation behaviour. Acta Metall. 1978, 26, 1815–1831. [Google Scholar] [CrossRef]
  22. Banerjee, S.; Krishnan, R. Martensitic transformation in Zr− Ti alloys. Metall. Trans. 1973, 4, 1811–1819. [Google Scholar] [CrossRef]
  23. Yang, H.; Shen, J.; Matsukawa, Y.; Satoh, Y.; Kano, S.; Zhao, Z.; Li, Y.; Li, F.; Abe, H. Effects of alloying elements (Sn, Nb, Cr, and Mo) on the microstructure and mechanical properties of zirconium alloys. J. Nucl. Sci. Technol. 2015, 52, 1162–1173. [Google Scholar] [CrossRef]
  24. Feng, Z.H.; Xia, C.Q.; Jing, R.; Jiang, X.; Zhou, Y.; Zhong, H.; Zhang, X.; Ma, M.; Liu, R. Microstructure and mechanical properties of ZrBe alloys processed by hot rolling. Mater. Sci. Eng. A 2016, 667, 286–292. [Google Scholar] [CrossRef]
  25. Xia, C.Q.; Jiang, X.J.; Wang, X.Y.; Zhou, Y.; Feng, Z.; Liang, S.; Tan, C.; Ma, M.; Liu, R. Microstructure and mechanical properties of hot-rolled ZrB alloys. Mater. Sci. Eng. A 2015, 628, 168–175. [Google Scholar]
  26. Zhang, Z.G.; Zhou, Y.K.; Jiang, X.J.; Feng, Z.; Xia, C.; Zhang, X.; Ma, M.; Liu, R. A novel Zr-based alloy microstructure with high strength and excellent ductility. Mater. Sci. Eng. A 2016, 651, 370–375. [Google Scholar] [CrossRef]
  27. Cordero, Z.C.; Knight, B.E.; Schuh, C.A. Six decades of the Hall–Petch effect–a survey of grain-size strengthening studies on pure metals. Int. Mater. Rev. 2016, 61, 495–512. [Google Scholar]
  28. Schulson, E.M.; Roy, J.A. The yield strength of the L12 phase Zr3Al. Acta Metall. 1978, 26, 29–38. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) magnified view (2θ = 30–40°) for hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys.
Figure 1. (a) XRD patterns and (b) magnified view (2θ = 30–40°) for hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys.
Materials 18 04459 g001
Figure 2. (a1,a2) Inverse pole figure map and confidence index-overlaid phase map of hot-rolled ZrAl14Ti3 alloy, and (b1,b2) corresponding maps of ZrAl14Ti9 alloy.
Figure 2. (a1,a2) Inverse pole figure map and confidence index-overlaid phase map of hot-rolled ZrAl14Ti3 alloy, and (b1,b2) corresponding maps of ZrAl14Ti9 alloy.
Materials 18 04459 g002
Figure 3. (a1,a2) Kernel average misorientation (KAM) maps of Zr3Al and α-Zr phases in hot-rolled ZrAl14Ti3 alloy, and (b1,b2) corresponding KAM maps in ZrAl14Ti9 alloy.
Figure 3. (a1,a2) Kernel average misorientation (KAM) maps of Zr3Al and α-Zr phases in hot-rolled ZrAl14Ti3 alloy, and (b1,b2) corresponding KAM maps in ZrAl14Ti9 alloy.
Materials 18 04459 g003
Figure 4. Statistical distribution charts of kernel average misorientation (KAM) values: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Figure 4. Statistical distribution charts of kernel average misorientation (KAM) values: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Materials 18 04459 g004
Figure 5. Misorientation angle distribution charts: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Figure 5. Misorientation angle distribution charts: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Materials 18 04459 g005
Figure 6. Grain size statistical charts: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Figure 6. Grain size statistical charts: (a1) Zr3Al phase in ZrAl14Ti3 alloy, (b1) Zr3Al phase in ZrAl14Ti9 alloy, (a2) α-Zr phase in ZrAl14Ti3 alloy, (b2) α-Zr phase in ZrAl14Ti9 alloy.
Materials 18 04459 g006
Figure 7. {001}, {111} and {110} pole figures of Zr3Al phase: (a) ZrAl14Ti3 alloy, (b) ZrAl14Ti9 alloy.
Figure 7. {001}, {111} and {110} pole figures of Zr3Al phase: (a) ZrAl14Ti3 alloy, (b) ZrAl14Ti9 alloy.
Materials 18 04459 g007
Figure 8. {0001}, {10 1 ¯ 0} and {11 2 ¯ 0} pole figures of α-Zr phase: (a) ZrAl14Ti3 alloy, (b) ZrAl14Ti9 alloy.
Figure 8. {0001}, {10 1 ¯ 0} and {11 2 ¯ 0} pole figures of α-Zr phase: (a) ZrAl14Ti3 alloy, (b) ZrAl14Ti9 alloy.
Materials 18 04459 g008
Figure 9. Schmid factor analysis for prismatic slip in α-Zr phase: (a1) Map of ZrAl14Ti3 alloy, (a2) Statistical distribution chart of ZrAl14Ti3 alloy, (b1) Map of ZrAl14Ti9 alloy, (b2) Statistical distribution chart of ZrAl14Ti9 alloy.
Figure 9. Schmid factor analysis for prismatic slip in α-Zr phase: (a1) Map of ZrAl14Ti3 alloy, (a2) Statistical distribution chart of ZrAl14Ti3 alloy, (b1) Map of ZrAl14Ti9 alloy, (b2) Statistical distribution chart of ZrAl14Ti9 alloy.
Materials 18 04459 g009
Figure 10. ECCI images of ZrAl14Ti3 alloy: (a) 1000× magnification, (b) 10,000× magnification, (c) 10,000× magnification, (d) 40,000× magnification.
Figure 10. ECCI images of ZrAl14Ti3 alloy: (a) 1000× magnification, (b) 10,000× magnification, (c) 10,000× magnification, (d) 40,000× magnification.
Materials 18 04459 g010
Figure 11. ECCI images of ZrAl14Ti9 alloy: (a) 1000× magnification, (b) 10,000× magnification, (c) 20,000× magnification, (d) 50,000× magnification.
Figure 11. ECCI images of ZrAl14Ti9 alloy: (a) 1000× magnification, (b) 10,000× magnification, (c) 20,000× magnification, (d) 50,000× magnification.
Materials 18 04459 g011
Figure 12. Engineering stress–strain curves of hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys.
Figure 12. Engineering stress–strain curves of hot-rolled ZrAl14Ti3 and ZrAl14Ti9 alloys.
Materials 18 04459 g012
Figure 13. A comparative plot of yield strength versus elongation for present alloys and other representative Zr-based alloys from the literature [7,11,23,24,25,26].
Figure 13. A comparative plot of yield strength versus elongation for present alloys and other representative Zr-based alloys from the literature [7,11,23,24,25,26].
Materials 18 04459 g013
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

Zhang, X.; Yang, Y.; Yang, M.; Li, W.; Li, Z. Microstructure and Mechanical Properties of Hot-Rolled ZrAl14Ti3 and ZrAl14Ti9 Alloys. Materials 2025, 18, 4459. https://doi.org/10.3390/ma18194459

AMA Style

Zhang X, Yang Y, Yang M, Li W, Li Z. Microstructure and Mechanical Properties of Hot-Rolled ZrAl14Ti3 and ZrAl14Ti9 Alloys. Materials. 2025; 18(19):4459. https://doi.org/10.3390/ma18194459

Chicago/Turabian Style

Zhang, Xing, Yujing Yang, Mingchao Yang, Wang Li, and Zhixin Li. 2025. "Microstructure and Mechanical Properties of Hot-Rolled ZrAl14Ti3 and ZrAl14Ti9 Alloys" Materials 18, no. 19: 4459. https://doi.org/10.3390/ma18194459

APA Style

Zhang, X., Yang, Y., Yang, M., Li, W., & Li, Z. (2025). Microstructure and Mechanical Properties of Hot-Rolled ZrAl14Ti3 and ZrAl14Ti9 Alloys. Materials, 18(19), 4459. https://doi.org/10.3390/ma18194459

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