logo
孪生和异质变形诱导强化作用协同提升Ti-6554合金的损伤容限性

孪生和异质变形诱导强化作用协同提升Ti-6554合金的损伤容限性

明珠
斯韫
文茜
书贤
一夔
继雄
平辉
会群
素平
300

异质结构材料具有优越的力学性能,而当前的研究对其疲劳损伤性能的评估尚且缺乏。本研究采用高通量梯度热处理方法在单个试样上快速制备了在757~857 ℃较宽温度范围的Ti-6Cr-5Mo-5V-4Al(Ti-6554)合金固溶组织,随后进行了500 ℃/4 h时效处理。利用扫描电子显微镜(SEM)、电子背散射衍射(EBSD)、透射电子显微镜(TEM)等对比分析了三种不同再结晶程度(0%,40%,100%)典型微观组织的拉伸变形及疲劳裂纹扩展行为。结果表明:β异质结构的屈服强度(σYS)高达1403 MPa,同时均匀伸长率(UE)显著提升至6.5%。相较于均质结构,σYS和UE分别提升了6.7%和109.7%。异质结构的引入不仅克服了强度-韧性平衡问题,还明显增强了疲劳裂纹扩展(FCP)性能。在FCP过程中,β异质结构通过异质变形诱导(HDI)强化效应,显著促进了粗αS相内几何必需位错(GNDs)的积累,从而加速达到孪晶临界分切应力并增加孪晶密度。这一机制有助于裂纹尖端区域的应力释放和塑性变形能力的提升,使得临界快速断裂阈值由30.4提升至36.0 MPa·m1/2,扩大了稳态扩展区。本研究为钛合金通过异质结构调整以提高疲劳损伤容限性提供了新的见解。

Ti-6554合金疲劳裂纹扩展(FCP)异质变形诱导HDI强化变形诱导纳米孪晶

J.Cent.South Univ.(2025) 32: 744-759

1 Introduction

The metastable β titanium alloy exhibits exceptional toughness, high fatigue performance, and excellent corrosion resistance, making it highly promising for aerospace applications [1-3]. However, increasing fatigue failure in aerospace components has necessitated research on the fatigue damage mechanism of metastable β titanium alloys [4-6].

Various mechanisms have been proposed to explain the fatigue deformation behavior of titanium alloys, including slip [7-10], deformation twinning [11, 12], stacking fault [13], and phase transformation induced crack deflection [14]. BRIFFOD et al [15] investigated that fatigue crack initiation occurred in the primary α phase (αP) and lamellar colonies, particularly in prismatic slip planes with high Schmid factors in the Ti-6Al-4V alloy. HE et al [16] also discovered that the crack path was related to the ease of slip transfer. When the Burgers orientation relationship (BOR) is fully or partially followed, slip can easily transfer via the β phase to the adjacent lamellar α phase. ZHANG et al [17] observed that numerous {pic} type mechanical twins were induced in the lamellar and bimodal microstructure of the Ti-5Al-5Mo-5V-3Cr-1Zr alloy. SONG et al [18] observed that crack deflection within the β grain of Ti-24Nb-4Zr-7.9Sn alloy was closely related to martensitic transformation during high cycle fatigue. Conventional homogeneous structures achieve excellent damage tolerance performance by precisely regulating microstructure characteristics, such as grain size [19, 20], phase composition [21], morphology [22, 23], proportion [7] and texture [24, 25]. LEI et al [26] found that bimodal microstructure has superior strength and ductility but lower crack propagation resistance compared to those of Widmanstätten microstructure. These different microstructures demonstrate excellent fatigue performance; however, their mechanical properties are not always matched. Therefore, an effective structure control strategy should be determined to balance the fatigue performance and mechanical properties of titanium alloys.

Recently, heterogeneous structure has attracted considerable research attention due to their unique strain-hardening effect (hetero-deformation induced strengthening, HDI) and outstanding strength and ductility [27-29]. ZHOU et al [27] fabricated a bimodal microstructure with multiscale α phases by hot rolling and aging treatment of Ti-6Al-2Sn-4Zr-2Mo-0.1Si alloy. The cooperation between the multiscale α phases generated an apparent HDI stress to simultaneously enhance strength and ductility. WU et al [29] reported a heterogeneous lamella structure in Ti, produced by asymmetric rolling and partial static recrystallization (SRX) that combines ultrafine-grained metal strength with conventional coarse-grained metal ductility. They attributed the high strength to enhanced HDI stress from heterogeneous yielding, and the exceptional ductility to HDI and dislocation hardening. GU et al [30] found that the HDI causes extra <c+a> GNDs to accumulate in constrained micro-grains, providing enough strain hardening to maintain or slightly enhance the ductility of CP Ti. Using phase field simulations, WU et al [31] developed a Ti-5Al-5Mo-5V-3Cr-1Zr heterogeneous structure with coarse and ultra-fine lamellar α phases by adjusting the element concentration in the β matrix. The ultimate tensile strength (σUTS) and engineering strain of the heterogeneous structure were 1496 MPa and 5.8%, respectively. YU et al [32] achieved a heterogeneous structure of the Ti-6Cr-5Mo-5V-4Al (Ti-6554) alloy by adjusting the degree of SRX. The heterogeneous structure exhibited σUTS and ductility values of ~1670 MPa and 5%, respectively, outperforming the homogeneous structure that demonstrated complete macroscopic brittleness at the same strength level. Heterogeneous structures facilitate the simultaneous increment of strength and ductility. However, the severe non-uniform plastic deformation in multiphase microstructures may deteriorate the fatigue resistance [33, 34]. Therefore, evaluating the fatigue damage tolerance of heterogeneous structure is crucial in assessing their application.

This study employed a high-throughput gradient heat treatment method [35, 36] to rapidly screen the following three microstructures in the Ti-6554 alloy: β-recovery, β-hetero, and β-fully SRX. The relationship between tensile properties and damage tolerance (fatigue crack propagation, FCP) induced by the three microstructures would be investigated in details and the related mechanism would be also discussed. This study would unveil the contribution of heterogeneous structure to crack propagation resistance, and provide substantial strategy to tune microstructure for titanium alloys used for fatigue resistance component.

2 Material and experiments

2.1 Material

The Ti-6554 alloy used in this study was a 145 mm diameter hot forged bar supplied by Baoti Group Co., Ltd. with the following composition: 5.01Cr, 4.72Mo, 4.92V, 4.73Al, balance Ti (wt%). The β-transus temperature (Tα+β→β) was identified to be 810 ℃ by the metallographic observations in our previous work [37]. The specific experimental process is depicted in Figure S1 (Supplementary material).

A high-throughput gradient heat treatment method was employed to rapidly obtain the solution microstructure evolution laws [35, 36]. As illustrated in Figure 1(a), a tubular furnace was used to establish the temperature gradient distribution. A round bar of Ti-6554 alloy with a diameter of 10 mm was processed by wire-cutting, and holes with a diameter of 1 mm and a depth of 8 mm were drilled on it every 10 mm using EDM. K-type thermocouples were fixed in the holes using high-temperature conductive adhesive, and the temperature was recorded in real time using a YP5016G multi-channel temperature tester. Figure 1(b) depicts the schematic diagram of the heat treatment of a single sample at different locations, including solution annealing at 757, 774, 792, 810, 825, 840 and 857 ℃ for 2 h, followed by air cooling (AC) and subsequent aging at 500 ℃ for 4 h.

Figure 1
(a) Temperature gradient distribution of tubular furnace and (b) schematic of heat treatment
pic
2.2 Mechanical testing

Tensile, cyclic loading and unloading, and FCP tests were performed on Ti-6554 alloy using the MTS Landmark testing machine. The tensile performance test was conducted following the ISO 6892-1: 2019 standard, with a loading rate of 1 mm/min. The standard sample dimensions are shown in Figure 2(a). The cyclic loading and unloading test sample size was the same as the tensile sample. The loading strain rate was 1×10-3 s-1 and then unloaded to 50 N at a rate of 1000 N/min. The dimensions of the compact tension (CT) samples were 40 mm × 38.4 mm × 3 mm (Figure 2(b)). Prior to the FCP experiment, both surfaces of the samples were polished to eliminate the influence of surface roughness. A 2 mm pre-crack was formed on the specimens based on the ISO 12108:2018 standard. The FCP specimens were tested under sinusoidal stress control with a frequency of 10 Hz and a sinusoidal load ratio (R) of 0.1.

Figure 2
Test sample geometry (unit: mm): (a) Tensile, cyclic loading and unloading sample; (b) FCP sample
pic
2.3 Microstructural characterization

The microstructure of the sample was characterized using optical microscope (OM, Leica DM ILM), backscattered electron mode scanning electron microscope (SEM, JSM-7900F), electron backscattered diffraction (EBSD, Regulus8230), and transmission electron microscope (TEM, Tecnai G2F20, Talos F200X). Prior to the EBSD examination, the samples were mechanically ground followed by electrolytic polishing in a solution of 90% CH3COOH and 10% HClO4 (vol%). The electrolytic polishing voltage was set at 30 V, and the temperature was maintained at -30 ℃. The step size was 5 μm and the operating voltage was 20 kV. The EBSD data were post-processed using the Atex software [38]. The TEM sample was thinned to a thickness of 70 μm and then cut into 3 mm diameter circular discs. These discs were further thinned by automatic dual-jet electrolytic polishing in a solution of 5% HClO4, 35% C4H9OH, and 60% CH3OH at 20 V and -30 ℃.

3 Results

3.1 Soluted and aged microstructure

Figure 3 represents the microstructure of the Ti-6554 alloy after undergoing various solution and aging treatments. As shown in Figures 3(a1)-(c1), the volume fraction of αP gradually decreases from 9% at 757 ℃ to 2% at 792 ℃ with increasing solution temperature. Furthermore, the bright regions in Figures 3(b1) and (c1) indicate partial SRX grains in the solution microstructure at 774 and 792 ℃, respectively. When the solution temperature attains the β-transus temperature of 810 ℃, all α phases transform into the β matrix (Figures 3(d1)-(g1)). According to the statistical data in Table 1, the β grain size increases with increasing solution temperature.

Figure 3
SEM and OM microstructures of the Ti-6554 alloy after (a1-g1) gradient solution treatment for 2 h and (a2-g2) aging at 500 ℃ for 4 h
pic
Table 1
Statistical microstructure parameters of the Ti-6554 alloy after solution and aging treatments
SampleαP volume fraction/%β grain size/μmαS thickness/nm
757STA940.1
774STA554.3
792STA272.2
810STA179.678.9
825STA261.879.3
840STA297.380.6
857STA316.284.3
展开更多

The microstructures after aging at 500 ℃ for 4 h are depicted in Figures 3(a2)-(g2). It is evident that a substantial amount of ultra-fine αS phases have precipitated in the microstructures of the 757 solution-aging (STA) and 774STA samples. Conversely, the higher solution temperature samples exhibit a noticeable increase in the length and thickness of the coarse αS phases formed after aging (Figures 3(c2)-(g2)).

Based on the above observations, three typical microstructures of 757STA, 792STA and 825STA with different degrees of SRX were selected for detailed characterization and property analysis. Figure 4 presents the microstructures of the longitudinal sections of Ti-6554 bars after undergoing the selected three different heat treatments. The orientation map in Figure 4(a) shows that the microstructure of the 757STA sample contains numerous β-deformed grains in a fibrous morphology (βf). This fibrous deformation structure is inherited from the forging microstructure. The βf grains present slight color variations, indicating the presence of localized small misorientations. Figure 4(d) is the grain boundary map corresponding to Figure 4(a), where low-angle grain boundaries (LAGBs) and high-angle grain boundaries (HAGBs) are distinguished based on whether the misorientation between adjacent grains is 5°-15° or >15°. The presence of small misorientation variations within the βf grains, where the fraction of LAGBs (FLAGBs) is 61%, is further demonstrated in Figure 4(d).

Figure 4
Orientation maps of Ti-6554 alloy samples of (a) 757STA, (b) 792STA and (c) 825STA; The grain boundary maps of (d) 757STA, (e) 792STA and (f) 825STA
pic

In Figure 4(b), it can be seen that the 792STA sample undergoes approximately 40% partial SRX. The recrystallized β grains are present in an equiaxed shape (βe) with an average grain size (picβe) of 98.7 μm, significantly smaller than the βf grains. Notably, these βe grains exhibit random crystal orientation without a specific texture. Figure 4(e) illustrates that the majority of the βe grains are separated by HAGBs, while numerous LAGBs still remain in the βf grains. Clearly, the obtained microstructure containing both βe and βf grains exhibits structural heterogeneity due to the partial SRX effect. The 825STA sample undergoes sufficient SRX (Figure 4(c)), with a picβe of 261.8 μm. In conjunction with Figure 4(f), only 18% of LAGBs remain; the rest disappear. For convenience, the three types of STA samples are referred to as ‘β-recovery samples’, ‘β-hetero samples’, and ‘β-fully SRX samples’ in the following sections.

Another important feature to consider is the geometrically necessary dislocation (GND) density of the three samples, as GND plays a paramount role in maintaining the continuity of plastic flow during material deformation [39]. A high GND density not only enhances the material strength but also improves its ductility [40]. Table 2 presents the results obtained from the EBSD data of the β-recovery, β-hetero, and β-fully SRX samples. The GND density of βe grains (ρβe) is significantly lower than that of the βf grains (ρβf). Figure 5 illustrates the GND distribution of the three specimens. The β-recovery samples exhibit the highest dislocation density due to their low solution temperature, which effectively preserves dislocations within βf grains (Figure 5(a)). The β-fully SRX samples demonstrate the lowest GND owing to the extensive consumption of dislocations, leading to the formation of new βe grains (Figure 5(c)). Interestingly, as shown in Figure 5(b), the β-hetero samples comprise hard domain βf grains with high GND density and soft domain βe grains with low GND density.

Table 2
Statistical microstructure features from the EBSD data of the β-recovery, β-hetero, and β-fully SRX samples, including the LAGBs fraction (FLAGBs), volume fraction of βf (f) and βe (e) grains, average grain size of βf (picβf) and βe (picβe) grains, and GND density of βf (ρβf) and βe (ρβe) grains
SampleFLAGBs/%f/%e/%picβf/μmpicβe/μmρβf/m-2ρβe/m-2
β-recovery61100468.21.75×1013
β-hetero54.86040424.998.71.77×10135.94×1012
β-fully SRX18100261.81.87×1012
展开更多
Figure 5
GND distribution of the Ti-6554 alloy: (a) β-recovery; (b) β-hetero; (c) β-fully SRX sample
pic

The αS-precipitation morphologies of the three samples were observed using TEM and the thickness distribution was calculated, as shown in Figure 6. A common characteristic of three samples is the presence of three variants of αS, intersecting at an angle of ~60° [41]. The αS-precipitates in the β-recovery sample (Figures 6(a) and (d)), with an average thickness of 40.1 nm, were smaller and denser than those within other samples. Interestingly, Figures 6(b) and (c) present that the αS-precipitates in β-hetero and β-fully SRX samples exhibit significant size variations. These αS-precipitates consist of both coarse αS and ultra-fine αS, with increased length and thickness. The thickness distribution plot clearly shows two peaks (Figures 6(e) and (f)), indicating that αS thickness is characterized by an ultra-fine and coarse dual-scale. The reason for such differences is that the βf grains are rich in dislocations, providing abundant nucleation sites that facilitate the uniform formation of the ultra-fine αS phase. Contrastingly, GNDs within the βe grains are consumed during the SRX process, forming larger and unevenly distributed αS phases during the subsequent aging treatment.

Figure 6
HAADF-STEM images of the aged Ti-6554 alloy samples of (a) β-recovery, (b) β-hetero and (c) β-fully SRX samples; the αS thickness of (d) β-recovery, (e) β-hetero and (f) β-fully SRX samples
pic
3.2 Tensile properties and fatigue propagation

Figure 7(a) displays the tensile engineering stress-strain curves, and Table 3 presents the typical mechanical property data. The ultimate tensile strength (σUTS) of the β-recovery and β-fully SRX samples is 1367 MPa and 1391 MPa, respectively, while the corresponding uniform elongation (UE) is 4.6% and 3.1%, respectively. Evidently, both samples exhibit high strength but low ductility, thus experiencing a strength-ductility trade off. For the β-hetero sample, the yield strength (σYS) is 1403 MPa while the UE remains at 6.5%. Excitingly, compared to the β-recovery and β-fully SRX samples, the σYS of the β-hetero sample increased by 6.4% and 6.7%, respectively, while the corresponding UE improved by 41.3% and 109.7%. The tensile fracture characteristics of the samples are shown in Figure S2. During deformation, the soft domain βe enters the plastic deformation stage first, leading to a significant accumulation of GNDs owing to the difficulty encountered by GNDs in surpassing the βf /βe interface [42]. Consequently, the dislocation accumulation generates HDI stresses [43], resulting in an enhancement of the overall σYS. After yielding, GNDs in the βe phase contribute to HDI hardening and delay necking during tensile testing, thereby improving UE [28].

Figure 7
(a) The tensile engineering stress and strain curves, and (b) FCP curves of β-recovery, β-hetero and β-fully SRX samples
pic
Table 3
Tensile and FCP test data of β-recovery, β-hetero and β-fully SRX samples, including yield strength (σYS), ultimate tensile strength (σUTS), uniform elongation (UE), total elongation (TE), stress concentration factor (pic), and critical value of pic of fracture (pic)
SampleσYS/MPaσUTS/MPaUE/%TE/%(da/dN)/(mm·cycle-1) at pic=24.9 picpic/pic
β-recovery131913674.66.1pic30.4
β-hetero140314626.57.7pic36.0
β-fully SRX131513913.14.3pic32.4
展开更多

Figure 7(b) depicts the FCP curves for the three samples, showing varying rates of crack propagation (da/dN). In comparison to the β-recovery specimen, the FCP curves of the β-hetero and β-fully SRX specimens both exhibit a rightward shift, indicating a greater critical fast fracture threshold (pic). The pic value of the β-hetero sample is 18.4% and 11.1% greater than those of the β-recovery and β-fully SRX samples respectively, indicating an enlarged steady state propagation region. At the stress concentration factor of 24.9 pic, the β-hetero sample demonstrates the lowest FCP rate of pic mm/cycle. The successful design of such a heterogeneous structure overcomes the problem of tensile brittleness in high strength metastable β titanium alloys while maintaining excellent FCP resistance.

To quantitatively measure the HDI strengthening effect, we performed cyclic loading and unloading tests (Figure 8(a)). Calculate the HDI stress (σHDI) [42] according to the diagram in Figure 8(b): σHDI=(σr+σu)/2 where σr and σu are the stresses at which the stress-strain curves of reloading and unloading deviate from the linear relationship, respectively. As shown in Figure 8(c), the σHDI values of β-hetero are significantly higher than those of β-recovery and β-fully SRX samples, and the rate of increment in HDI values for the former is also slightly greater. These results suggest that the β-hetero microstructure does indeed provide extra HDI strengthening due to heterogeneous yielding.

Figure 8
(a) Cyclic loading and unloading curves, (b) schematic of calculating HDI stress and (c) HDI stress
pic
3.3 Fatigue crack propagation path

Figure 9(a) presents the FCP path profile of the β-recovery sample, indicating the transgranular FCP mechanism. In Figures 9(b)-(d), numerous crack deflections and branching are observed, and αS phases are distorted in the vicinity of the crack due to large stress concentration. Additionally, voids and microcracks bridging are observed at the αP/β interface (Figures 9(e) and (f)). WANG et al [44] proposed that microcracks propagate and bridge along the αP/β interface, facilitating easier fracture and accelerating the FCP rate, which is consistent with our results.

Figure 9
SEM images of the FCP path in the Ti-6554 alloy for the β-recovery sample
pic

Figure 10(a) illustrates the tortuous FCP path of the β-fully SRX sample, exhibiting a mixed mode of intergranular and transgranular fracture. Figures 10(a)-(c) reveal abundant branching and closure of cracks, which helps hinder crack propagation. In Figure 10(d) the bridging of voids and microcracks at grain boundaries can be found, accompanied by multiple deflections. The grain boundaries are considered significant barrier that inhibits the rate of FCP [45]. The crack deflection is attributed to the presence of coarse αS phases oriented in various directions, as evidenced in Figure 10(e). In addition, numerous secondary cracks surround the main crack, and the crack tip of αS-precipitates is deformed (Figure 10(f)). The propagation behavior of these cracks decreases the da/dN, improving the crack propagation resistance of the β-fully SRX sample despite its lowest UE value.

Figure 10
SEM images of the FCP path in the Ti-6554 alloy for the β-fully SRX sample
pic

Figure 11 shows the FCP profile of the β-hetero sample, where the fatigue cracks alternately pass through the βf and βe grains. The crack propagation path is relatively tortuous (Figures 11(b) and (c)), accompanied by the wide distribution of secondary cracks surrounding the main crack (Figure 11(f)). Similarly, Figure 11(d) displays the voids and microcrack bridging. In particular, there is a pronounced contrast in αS-precipitation within the βf and βe grains (Figure 11(e)), which aligns with the statistical results presented in Figure 6. This crack propagation behavior consumes more energy, resulting in a lower da/dN and higher crack propagation resistance for the β-hetero sample.

Figure 11
SEM images of the FCP path in the Ti-6554 alloy for the β-hetero sample
pic

4 Discussion

4.1 Mechanism for improving damage tolerance during FCP

To further elucidate the source of the superior FCP resistance of the β-hetero specimen, the GND distribution maps near the crack surfaces of the three samples were plotted as shown in Figure 12. The β-recovery microstructure exhibits variations in GND density, which is influenced by the uneven forging deformation (Figure 12(a)). In fact, such differences do not result in significant heterogeneity effects, as supported by the density increment of GNDs in βf grains (Δρβf) of only 0.60×1012 m-2 presented in Table 4. In contrast, the β-fully SRX sample comprises new βe grains that demonstrate a higher capacity for dislocation accumulation and a greater work-hardening rate. Consequently, this microstructure exhibits the highest density increment of GNDs in βe grains (Δρβe) of 5.19× 1012 m-2, as shown in Figure12(c).

Figure 12
GND distribution of FCP for the (a) β-recovery, (b) β-hetero, and (c) β-fully SRX samples, (d) enlarged image of the yellow box in (b), and (e) GND density distribution curve along the loading direction in (d)
pic
Table 4
Statistical results of EBSD data of the β-recovery, β-hetero and β-fully SRX samples after FCP deformation, including the density of GNDs in βf (ρβf) and βe grains (ρβe), and the density increment of GNDs in βfρβf) and βe grains (Δρβe) before and after deformation
Sampleρβf /m-2ρβe/m-2Δρβf/m-2Δρβe/m-2
β-recovery1.81×10130.60×1012
β-hetero1.85×10139.47×10120.80×10123.53×1012
β-fully SRX7.06×10125.19×1012
展开更多

In the β-hetero sample, the GND value of the hard domain βf grain is substantially higher than that of the soft domain βe grain before FCP (Figure 5). It should be noted that inducing the HDI strengthening effect in the βe grain results in a gradient enhancement of GND density along the grain interior towards the grain boundary, accompanied by a rapid increase in the GND density [43]. To substantiate this, the GND density distribution (Figure 12(d)) and its curve along the loading direction were plotted (Figure 12(e)). The GND density is higher at the βf /βe interface, and the gradient increases from the interior to the grain boundary. The cumulative incremental variation in the GNDs density of the three microstructures was counted as shown in Table 4. The Δρβe of the β-hetero sample (3.53×1012 m–2) is higher than the density increment of GNDs in βf grains (Δρβf) of 0.80×1012 m-2, which validates the above view. Therefore, the β-hetero sample exhibits HDI strengthening effect, and the gradient structure of high-density GNDs has a significant and beneficial effect on FCP resistance.

4.2 Deformation inducing nano-scale twins during FCP

Figure 13 depicts the TEM images of the microstructure near the fatigue crack region in the β-fully SRX specimen. A large number of dislocations accumulate at the αS/β interface, and obvious dislocations appear in the αS phase (Figures 13(a) and (b)). Figure 13(c) corresponds to the SAED pattern taken along the picβ zone axis. The classical BOR is maintained between α and the surrounding β grains, i.e., (0001)α//(110)β, <pic>α//<111>β [46]. After traversing the β matrix, the dislocations leave behind parallelly arranged debris that form dislocation walls (Figure 13(e)). Furthermore, trace analysis indicates that the dislocation walls lie on the picβ plane, which suggests that {pic}β <111>β slip system is activated. The presence of dense dislocation tangles indicates the occurrence of sufficient dislocation activity, as shown in Figure 13(f).

Figure 13
TEM microstructure near the fatigue crack of the β-fully SRX sample after FCP test: (a, b) Dislocation interacting with the αS and β matrix; (c) SAED pattern along the [pic]β zone axis; (d) Band structures in the coarse αS phase; (e) Dislocation slip arranged along the picβ plane; (f) Dislocation tangled in the β matrix; (g) HADDF-STEM image of band structures in the coarse αS phase; (h) SAED pattern of the red region in (g)
pic

Interestingly, as shown in Figure 13(d), band structures are found within the coarse αS phase, which is absent in the β-recovery deformation microstructure (Figure S3). To further reveal the band structure characteristics, two sets of diffraction spots for the band structure interface are clearly visible in Figure 13(h). By calibrating diffraction patterns, the zone axis relationship between the band structures and αS is obtained as [0001]//pic, and the relationship of overlapping spots is (pic) band structures//(pic)αS. To determine the orientation relationship between the band structures and αS, the zone axes index and overlap spots are considered as the crystal direction index (u v w) and crystal plane (h k l) of the Mille index respectively [47]. Hence, the orientation relationship can be expressed as (pic) [0001] band structures// (pic) [pic]αS, such that the axis angle relationship is <pic>/32°. Therefore, the band structure is a {pic} type nano-scale twin formed by the approximately 32° rotation around the pic axis. This deformation-induced twinning has been observed in other titanium alloys. CHEN et al [48] studied the deformation behavior of the Ti-55511 alloy and demonstrated that the deformation induces nano-scale twins in the lamellar α phase. LIU et al [49] pointed out that multiple deformation nano-scale twins can pile up and accommodate incoming strain, favoring stress relief by dislocation transfer and twinning-induced plasticity effect. Therefore, the plastic deformation of the coarse αS phase is enhanced by nano-scale twins.

Figure 14 shows the typical TEM microstructure near the fatigue crack region of the β-hetero sample. The TEM bright-field image (Figure 14(a)) reveals a significant accumulation of dislocations at the αS/β phase interface, whereas a few dislocations exist in the αS phase. To elucidate the dislocation types, dislocation analysis was performed on the same region under the two-beam TEM condition. Based on the invisibility criterion of g·b=0 [50, 51], the dislocation indicated by the blue arrow in Figure 14(b) is invisible in g=pic (Figure 14(f)). Hence, it is a type of pic dislocation. In Figure 14(c), yellow arrows represent dislocations that are invisible in g=pic, indicating that they are either pic or pic type dislocations. It is interesting to note that in the heterogeneous structure we also find the same nano-scale twins described earlier, and a common feature is that all these nano-scale twins exist only within the coarse αS phase and not in the ultra-fine αS phase, i.e. only within the recrystallized βe grains. Due to the HDI strengthening effect, the coarse αS phase in the β-hetero deformed microstructure induces more nano-scale twins compared to that in the β-fully SRX microstructure, promoting the accumulation of GNDs within the soft domain βe grains. Hence, the soft domain is strengthened such that the twinning critical shear stresses are achieved more easily. Finally, the {pic}<pic26> twinning systems are further activated within the coarse αS phase.

Figure 14
TEM microstructure near the fatigue crack of the β-hetero specimen after FCP test: (a) TEM bright-field image under the [011]β zone axis; (d) SAED pattern along the [011]β zone axis; TEM bright-field and dark-field images observed under two-beam condition with diffraction vectors (b, e) g=pic and (c, f) g=pic
pic

SCHMIDT et al [52] believed that a significant ductility loss may deteriorate the FCP resistance of the titanium alloy. In this work, the β-fully SRX sample has a UE of only 3.1%, but shows a higher FCP resistance compared to the β-recovery sample with a UE of 4.6%. According to the results in Figure 7(b), the β-hetero and the β-fully SRX samples demonstrate better FCP resistance, with the deformation-induced nanoscale twins being the common deformation mechanism. In comparison, no deformation mechanism other than dislocation slip is observed in the ultra-fine αS phase of the β-recovery microstructure. Based on this, we can infer that there is a close correlation between the formation of the nano-scale twins and the FCP resistance. Despite the occurrence of nano-scale twins within the homogeneous structure (the β-fully SRX), the HDI strengthening effect of the β-hetero microstructure could further promote the accumulation of GNDs, accelerating the achievement of the critical shear stress and increasing the nanoscale twin density or probability, thereby obviously alleviating stress concentration and improving plastic deformation in the crack tip zone. Consequently, the unique heterogeneous structure exhibits superior FCP resistance.

5 Conclusions

In this study, a high-throughput gradient thermal treatment method was employed to obtain the evolution law of microstructures in the Ti-6554 alloy, thereby constructing the β-hetero microstructure. The mechanical properties and damage tolerance of the β-recovery, β-hetero and β-fully SRX specimens were investigated. The main conclusions are as follows:

1) The β-hetero microstructure was formed by partial SRX annealing, comprising 40% equiaxed (βe) and 60% fibrous (βf) grains. The morphology of the β grains affects the precipitation of the αS phase during subsequent aging. The βf grains, which were abundant in dislocations, facilitated the uniform formation of ultra-fine αS precipitates. Conversely, coarse and ultrafine dual-scale were present in the αS precipitation within βe grains.

2) Compared to the β-recovery and β-fully

SRX samples, the β-hetero sample achieved a 6.7% and 109.7% increase in yield strength and elongation, respectively, without compromising the alloy ductility. This improvement is mainly attributed to the high density of GNDs and the HDI strengthening effect.

3) The FCP path of the β-hetero microstructure was tortuous, with fatigue cracks alternating through the βf and βe grains accompanied by widely distributed secondary cracks around the primary crack. The β-hetero microstructure demonstrated excellent FCP resistance, characterized by lower

da/dN and increased pic. This is ascribed to the HDI strengthening effect accelerating the accumulation of GNDs within the soft domain βe grains, which further promotes the deformation induced {pic} type nano-scale twins in coarse αS phase. This alleviated the stress concentration and improved the plastic deformation in crack tip zone.

References
1BANERJEE D, WILLIAMS J C.

Perspectives on titanium science and technology

[J]. Acta Materialia, 2013, 61(3): 844-879. DOI: 10.1016/j.actamat.2012.10.043.
百度学术谷歌学术
2PAN Su-ping, LIU Hui-qun, CHEN Yu-qiang, et al.

Lamellar α fencing effect for improving stress relaxation resistance in Ti-55511 alloy

[J]. Materials Science and Engineering A, 2021, 808: 140945. DOI: 10.1016/j.msea.2021.140945.
百度学术谷歌学术
3HUANG Hua-long, LI Dan, CHEN Chao, et al.

Selective laser melted near-beta titanium alloy Ti-5Al-5Mo-5V-1Cr-1Fe: Microstructure and mechanical properties

[J]. Journal of Central South University, 2021, 28(6): 1601-1614. DOI: 10.1007/s11771-021-4720-z.
百度学术谷歌学术
4PAN Su-ping, LIU Hui-qun, CHEN Yu-qiang, et al.

αs dissolving induced mechanical properties decay in Ti-55511 alloy during uniaxial fatigue

[J]. International Journal of Fatigue, 2020, 132: 105372. DOI: 10.1016/j.ijfatigue.2019. 105372.
百度学术谷歌学术
5LI Zhi-ming, PRADEEP K G, DENG Yun, et al.

Metastable high-entropy dual-phase alloys overcome the strength-ductility trade-off

[J]. Nature, 2016, 534(7606): 227-230. DOI: 10.1038/nature17981.
百度学术谷歌学术
6XU Sheng-hang, HAN Meng, SHEN Kai-jie, et al.

Fatigue properties of binary Ti-Ta metal-metal composite with lamellar microstructure

[J]. Journal of Central South University, 2023, 30(9): 2878-2889. DOI: 10.1007/s11771-023-5433-2.
百度学术谷歌学术
7WU Zhi-hong, KOU Hong-chao, CHEN Na-na, et al.

Microstructural influences on the high cycle fatigue life dispersion and damage mechanism in a metastable β titanium alloy

[J]. Journal of Materials Science & Technology, 2021, 70: 12-23. DOI: 10.1016/j.jmst.2020.07.018.
百度学术谷歌学术
8BRIDIER F, VILLECHAISE P, MENDEZ J.

Slip and fatigue crack formation processes in an α/β titanium alloy in relation to crystallographic texture on different scales

[J]. Acta Materialia, 2008, 56(15): 3951-3962. DOI: 10.1016/j.actamat.2008.04.036.
百度学术谷歌学术
9CHEN Wei, YU Guo-xiang, LI Ke-er, et al.

Plastic instability in Ti-6Cr-5Mo-5V-4Al metastable β-Ti alloy containing the β-spinodal decomposition structures

[J]. Materials Science and Engineering A, 2021, 811: 141052. DOI: 10.1016/j.msea. 2021.141052.
百度学术谷歌学术
10ZHENG Ze-bang, BALINT D S, DUNNE F P E.

Dwell fatigue in two Ti alloys: An integrated crystal plasticity and discrete dislocation study

[J]. Journal of the Mechanics and Physics of Solids, 2016, 96: 411-427. DOI: 10.1016/j.jmps. 2016.08.008.
百度学术谷歌学术
11ZHENG Shao-kai, SHEN Jun, WANG Wei, et al.

Multi-twinned deformation and fracture characteristics of directional solidified Ti-45.5Al-5Nb-0.5Ta alloys during high-temperature rotary-bending fatigue process

[J]. Materials Science and Engineering A, 2023, 876: 145157. DOI: 10.1016/j.msea.2023.145157.
百度学术谷歌学术
12BOSH N, MÜLLER C, MOZAFFARI-JOVEIN H.

Deformation twinning in cp-Ti and its effect on fatigue cracking

[J]. Materials Characterization, 2019, 155: 109810. DOI: 10.1016/j.matchar.2019.109810.
百度学术谷歌学术
13XU Zi-lu, HUANG Chao-wen, WAN Ming-pan, et al.

Influence of microstructure on strain controlled low cycle fatigue crack initiation and propagation of Ti-55531 alloy

[J]. International Journal of Fatigue, 2022, 156: 106678. DOI: 10.1016/j.ijfatigue.2021.106678.
百度学术谷歌学术
14WANG Xiao-gang, LIU Cheng-huan, SUN Bin-han, et al.

The dual role of martensitic transformation in fatigue crack growth

[J]. Proceedings of the National Academy of Sciences of the United States of America, 2022, 119(9): e2110139119. DOI: 10.1073/pnas.2110139119.
百度学术谷歌学术
15BRIFFOD F, BLEUSET A, SHIRAIWA T, et al.

Effect of crystallographic orientation and geometrical compatibility on fatigue crack initiation and propagation in rolled Ti-6Al-4V alloy

[J]. Acta Materialia, 2019, 177: 56-67. DOI: 10.1016/j.actamat.2019.07.025.
百度学术谷歌学术
16HE Dong, ZHU Jing-chuan, ZAEFFERER S, et al.

Effect of retained beta layer on slip transmission in Ti-6Al-2Zr-1Mo-1V near alpha titanium alloy during tensile deformation at room temperature

[J]. Materials & Design, 2014, 56: 937-942. DOI: 10.1016/j.matdes.2013.12.018.
百度学术谷歌学术
17ZHANG Zhong, HUANG Chao-wen, WEN Xin, et al.

Synergistic influence mechanism of microstructure type and loading mode on the long crack propagation in Ti-55531 alloy

[J]. Engineering Fracture Mechanics, 2022, 266: 108404. DOI: 10.1016/j.engfracmech.2022.108404.
百度学术谷歌学术
18SONG M, HE S Y, DU K, et al.

Transformation induced crack deflection in a metastable titanium alloy and implications on transformation toughening

[J]. Acta Materialia, 2016, 118: 120-128. DOI: 10.1016/j.actamat. 2016.07.041.
百度学术谷歌学术
19JUNET A, MESSAGER A, WECK A, et al.

Internal fatigue crack propagation in a Ti-6Al-4V alloy: An in situ study

[J]. International Journal of Fatigue, 2023, 168: 107450. DOI: 10.1016/j.ijfatigue.2022.107450.
百度学术谷歌学术
20CHEN Wei, LI Chao, FENG Kang-tun, et al.

Strengthening of a near β-Ti alloy through β grain refinement and stress-induced α precipitation

[J]. Materials, 2020, 13(19): 4255. DOI: 10.3390/ma13194255.
百度学术谷歌学术
21ZHU Cheng-peng, ZHANG Xiao-yong, LI Chao, et al.

A strengthening strategy for metastable β titanium alloys: Synergy effect of primary α phase and β phase stability

[J]. Materials Science and Engineering A, 2022, 852: 143736. DOI: 10.1016/j.msea.2022.143736.
百度学术谷歌学术
22TARIK HASIB M, OSTERGAARD H E, LI Xiao-peng, et al.

Fatigue crack growth behavior of laser powder bed fusion additive manufactured Ti-6Al-4V: Roles of post heat treatment and build orientation

[J]. International Journal of Fatigue, 2021, 142: 105955. DOI: 10.1016/j.ijfatigue.2020. 105955.
百度学术谷歌学术
23PAN Su-ping, FU Ming-zhu, LIU Hui-qun, et al.

In situ observation of the tensile deformation and fracture behavior of Ti-5Al-5Mo-5V-1Cr-1Fe alloy with different microstructures

[J]. Materials, 2021, 14(19): 5794. DOI: 10. 3390/ma14195794.
百度学术谷歌学术
24UMEZAWA O, LI Wei-bo.

Effects of crystallographic texture on subsurface fatigue crack generation in Ti-Fe-O alloy at low temperature

[J]. ISIJ International, 2022, 62(3): 593-601. DOI: 10.2355/isijinternational.isijint-2021-381.
百度学术谷歌学术
25YE Han, LE Fang-bin, WEI Chao, et al.

Fatigue crack growth behavior of Ti-6Al-4V alloy fabricated via laser metal deposition: Effects of building orientation and heat treatment

[J]. Journal of Alloys and Compounds, 2021, 868: 159023. DOI: 10.1016/j.jallcom.2021.159023.
百度学术谷歌学术
26LEI Z N, GAO P F, LI H W, et al.

Comparative analyses of the tensile and damage tolerance properties of tri-modal microstructure to widmanstätten and bimodal microstructures of TA15 titanium alloy

[J]. Journal of Alloys and Compounds, 2019, 788: 831-841. DOI: 10.1016/j.jallcom. 2019.02.300.
百度学术谷歌学术
27ZHOU Yu, WANG Ke, WEN Xin, et al.

Achieving synergy enhancement of strength and ductility in Ti-6Al-2Sn-4Zr-2Mo-0.1Si alloy by fabricating a new multiscale microstructure

[J]. Scripta Materialia, 2023, 226: 115233. DOI: 10.1016/j.scriptamat.2022.115233.
百度学术谷歌学术
28WU Xiao-lei, ZHU Yun-tian.

Heterogeneous materials: A new class of materials with unprecedented mechanical properties

[J]. Materials Research Letters, 2017, 5(8): 527-532. DOI: 10.1080/21663831.2017.1343208.
百度学术谷歌学术
29WU Xiao-lei, YANG Mu-xin, YUAN Fu-ping, et al.

Heterogeneous lamella structure unites ultrafine-grain strength with coarse-grain ductility

[J]. Proceedings of the National Academy of Sciences of the United States of America, 2015, 112(47): 14501-14505. DOI: 10.1073/pnas. 1517193112.
百度学术谷歌学术
30GU Lei, MENG Ao, CHEN Xiang, et al.

Simultaneously enhancing strength and ductility of HCP titanium via multi-modal grain induced extra dislocation hardening

[J]. Acta Materialia, 2023, 252: 118949. DOI: 10.1016/j.actamat. 2023.118949.
百度学术谷歌学术
31WU Di, HAO Meng-yuan, ZHANG Tian-long, et al.

Heterostructures enhance simultaneously strength and ductility of a commercial titanium alloy

[J]. Acta Materialia, 2023, 257: 119182. DOI: 10.1016/j.actamat.2023.119182.
百度学术谷歌学术
32YU Guo-xiang, ZHAO Ding-xuan, LI Ke-er, et al.

Homogeneous α-precipitation and enhanced plasticity in Ti-6Cr-5Mo-5V-4Al high-strength metastable titanium alloy with heterogenous β-structure

[J]. Materials Science and Engineering A, 2022, 858: 144180. DOI: 10.1016/j.msea. 2022.144180.
百度学术谷歌学术
33ZHANG Ming-da, CAO Jing-xia, HUANG Xu.

Local softening behavior accompanied by dislocation multiplication accelerates the failure of Ti-6Al-2Sn-4Zr-2Mo-0.1Si alloy under dwell fatigue load

[J]. Scripta Materialia, 2020, 186: 33-38. DOI: 10.1016/j.scriptamat.2020.03.054.
百度学术谷歌学术
34TANG Xue-feng, WANG Zhi-zhou, WANG Xin-yun, et al.

Unraveling size-affected plastic heterogeneity and asymmetry during micro-scaled deformation of CP-Ti by non-local crystal plasticity modeling

[J]. International Journal of Plasticity, 2023, 170: 103733. DOI: 10.1016/j.ijplas.2023.103733.
百度学术谷歌学术
35WANG Chang, ZHOU Wei, LI Si-yun, et al.

Achieving high strength and ductility in Ti-6.8Mo-3.9Al-2.8Cr-2Nb-1.2V-1Zr-1Sn alloy by rapid optimizing microstructure through gradient heat treatment

[J]. Journal of Central South University, 2023, 30(2): 387-399. DOI: 10.1007/s11771-023-5248-1.
百度学术谷歌学术
36ZHU Cheng-peng, LI Chao, WU Di, et al.

A titanium alloys design method based on high-throughput experiments and machine learning

[J]. Journal of Materials Research and Technology, 2021, 11: 2336-2353. DOI: 10.1016/j.jmrt. 2021.02.055.
百度学术谷歌学术
37ZHOU Wei, WANG Chang, LIU Ji-xiong, et al.

Ageing precipitation sequence and effect of ω and secondary α phases on tensile properties of metastable β Ti-6Cr-5Mo-5V-4Al alloy

[J]. Transactions of Nonferrous Metals Society of China, 2023, 33(6): 1742-1754.
百度学术谷歌学术
38BEAUSIR J B.

FUNDENBERGER J

. ATEX software [CP]. 2017: www.atex-software.eu.
百度学术谷歌学术
39MA Ming-yang, LAI Rui-lin, QIN Jin, et al.

Achieving exceptionally tensile properties and damage tolerance of 5083 aluminum alloy by friction stir processing assisted by ultrasonic and liquid nitrogen field

[J]. Materials Science and Engineering A, 2021, 806: 140824. DOI: 10.1016/j.msea. 2021.140824.
百度学术谷歌学术
40JIANG J, BRITTON T B, WILKINSON A J.

Measurement of geometrically necessary dislocation density with high resolution electron backscatter diffraction: Effects of detector binning and step size

[J]. Ultramicroscopy, 2013, 125: 1-9. DOI: 10.1016/j.ultramic.2012.11.003.
百度学术谷歌学术
41LIN Cheng, YIN Gui-li, ZHANG Ai-min, et al.

Simple models to account for the formation and decomposition of athermal ω phase in titanium alloys

[J]. Scripta Materialia, 2016, 117: 28-31. DOI: 10.1016/j.scriptamat.2016.01.042.
百度学术谷歌学术
42ZHU Yun-tian, WU Xiao-lei.

Heterostructured materials

[J]. Progress in Materials Science, 2023, 131: 101019. DOI: 10.1016/j.pmatsci.2022.101019.
百度学术谷歌学术
43ZHU Yun-tian, WU Xiao-lei.

Perspective on hetero-deformation induced (HDI) hardening and back stress

[J]. Materials Research Letters, 2019, 7(10): 393-398. DOI: 10. 1080/21663831.2019.1616331.
百度学术谷歌学术
44WANG Huan, ZHAO Qin-yang, XIN She-wei, et al.

Fatigue crack propagation behaviors in Ti-5Al-3Mo-3V-2Zr-2Cr-1Nb-1Fe alloy with STA and BASCA heat treatments

[J]. International Journal of Fatigue, 2021, 151: 106348. DOI: 10.1016/j.ijfatigue.2021.106348.
百度学术谷歌学术
45ZHAI T, JIANG X P, LI J X, et al.

The grain boundary geometry for optimum resistance to growth of short fatigue cracks in high strength Al-alloys

[J]. International Journal of Fatigue, 2005, 27(10-12): 1202-1209. DOI: 10.1016/j.ijfatigue.2005.06.021.
百度学术谷歌学术
46SHI R, DIXIT V, VISWANATHAN G B, et al.

Experimental assessment of variant selection rules for grain boundary α in titanium alloys

[J]. Acta Materialia, 2016, 102: 197-211. DOI: 10.1016/j.actamat.2015.09.021.
百度学术谷歌学术
47LUO Wei, XUE Teng, ZUO Ding, et al.

Formation and strengthening mechanism of kink bands in an ultra-coarse-grained Fe-Cr-Al alloy

[J]. Journal of Materials Science & Technology, 2024, 186: 1-14. DOI: 10.1016/j.jmst.2023. 11.017.
百度学术谷歌学术
48CHEN Wei, LI Chao, ZHANG Xiao-yong, et al.

Deformation-induced variations in microstructure evolution and mechanical properties of bi-modal Ti-55511 titanium alloy

[J]. Journal of Alloys and Compounds, 2019, 783: 709-717. DOI: 10.1016/j.jallcom.2018.12.262.
百度学术谷歌学术
49LIU Chang, CHEN Jia-nan, WANG Yi-fan, et al.

Strong and ductile nanoscale Ti-1Fe dual-phase alloy via deformation twinning

[J]. Scripta Materialia, 2023, 237: 115720. DOI: 10.1016/j.scriptamat.2023.115720.
百度学术谷歌学术
50WEI Kang, HU Rong, YIN Dong-di, et al.

Grain size effect on tensile properties and slip systems of pure magnesium

[J]. Acta Materialia, 2021, 206: 116604. DOI: 10.1016/j.actamat. 2020.116604.
百度学术谷歌学术
51YAO Kai, MIN Xiao-hua.

Abnormal strain rate strengthening and strain hardening with constitutive modeling in body-centered cubic{332}TWIP titanium alloy

[J]. Acta Materialia, 2022, 226: 117641. DOI: 10.1016/j.actamat. 2022.117641.
百度学术谷歌学术
52SCHMIDT P, EL-CHAIKH A, CHRIST H J.

Effect of duplex aging on the initiation and propagation of fatigue cracks in the solute-rich metastable β titanium alloy Ti38-644

[J]. Metallurgical and Materials Transactions A, 2011, 42(9): 2652-2667. DOI: 10.1007/s11661-011-0662-7.
百度学术谷歌学术
注释

FU Ming-zhu, LUO Wei, LI Si-yun, YAO Wen-xi, PENG Shu-xian, LIU Yi-kui, LIU Ji-xiong, ZHANG Ping-hui, LIU Hui-qun, and PAN Su-ping declare that they have no conflict of interest.

FU Ming-zhu, LUO Wei, LI Si-yun, YAO Wen-xi, PENG Shu-xian, LIU Yi-kui, LIU Ji-xiong, ZHANG Ping-hui, LIU Hui-qun, PAN Su-ping. Enhancement of damage tolerance in Ti-6554 alloy through twinning and hetero-deformation induced strengthening synergy [J]. Journal of Central South University, 2025, 32(3): 744-759. DOI: https://doi.org/10.1007/s11771-025-5919-1.

伏明珠,罗威,李斯韫等.孪生和异质变形诱导强化作用协同提升Ti-6554合金的损伤容限性[J].中南大学学报(英文版),2025,32(3):744-759.