Journal of Materials Science & Technology, 2020, 50(0): 204-214 DOI: 10.1016/j.jmst.2020.01.060

Research Article

Duality of the fatigue behavior and failure mechanism in notched specimens of Ti-7Mo-3Nb-3Cr-3Al alloy

Zhihong Wu, Hongchao Kou,*, Nana Chen, Mengqi Zhang, Ke Hua, Jiangkun Fan, Bin Tang, Jinshan Li

State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi'an 710072, China

Corresponding authors: * E-mail adresshchkou@nwpu.edu.cn(H. Kou)

Received: 2019-12-11   Accepted: 2020-01-27   Online: 2020-08-1

Abstract

An interesting phenomenon of dual S-N fatigue behavior is investigated in a metastable β titanium alloy, Ti-7Mo-3Nb-3Cr-3Al notched cylindrical specimens with an elastic stress concentration factor of Kt = 3. Fractographic studies revealed all specimens, and irrespective of lifetime, failed from the specimen surface because of stress concentration occurs at the notch root. Typically, the short-life-distribution is usually associated with surface-failure-without-facets and the long-life-distribution generally occurs due to surface-failure-with-facets. This competing failure leads to increasing the variability in fatigue lifetime and further facilitates the difficulty in prediction of fatigue lifetime. Crack-initiation area characterization was conducted by using mechanical grinding, focused ion beam milling and subsequent electron back-scattered diffraction (EBSD) analysis of the 2D section across faceted grains. Results show that the αp particles (especially the elongated αp particles) well-oriented for basal slip activation is a preferential fatigue-critical microstructural configuration. Additionally, the β+αs matrix has a higher KAM value than the αp particles in fatigued microstructures and significant dislocation activity in the form of dislocation tangles is observed in αp boundaries.

Keywords: High cycle fatigue ; Metastable β titanium alloys; ; Notch effects ; S-N curves ; Surface crack initiation

PDF (5886KB) Metadata Metrics Related articles Export EndNote| Ris| Bibtex  Favorite

Cite this article

Zhihong Wu, Hongchao Kou, Nana Chen, Mengqi Zhang, Ke Hua, Jiangkun Fan, Bin Tang, Jinshan Li. Duality of the fatigue behavior and failure mechanism in notched specimens of Ti-7Mo-3Nb-3Cr-3Al alloy. Journal of Materials Science & Technology[J], 2020, 50(0): 204-214 DOI:10.1016/j.jmst.2020.01.060

1. Introduction

The metastable β titanium alloys are considered as attractive materials for aerospace industries due to its high specific strength, excellent hardenability, and high resistance to fatigue crack-initiation (FCI) [1,2]. This excellent strength is associated with a large volume fraction of fine α precipitates in the grains of transformed-β (βt) [3,4]. However, this microstructure design often leads to failure initiation due to interface cracking between α phase and the β matrix, especially undergoing cyclic loading. The metastable β titanium alloys exhibit higher fatigue performance, causing increased interest in studying the microstructure-fatigue performance relationship, nevertheless, the underlying fatigue damage mechanisms are still lags behind [5,6].

Recent studies indicate that the conventional representation in terms of single mechanism is perhaps an oversimplification as the fatigue data converge on two curves. This phenomenon is the duality in fatigue behavior and the two-stage S-N behavior where the fatigue lifetime data grouped into two separate S-N curves, especially in high-strength alloys including titanium alloys [[7], [8], [9], [10], [11], [12]], high strength steels [13,14], nickel-base superalloys [15], and γ-TiAl based alloy [16]. Typically, the short-life-distribution is generally associated with the surface-crack-initiated failure and the long-life-distribution usually occurs due to the internal-crack-initiated failure. When considered in relation to some recent studies [[17], [18], [19], [20], [21]], the variability in fatigue response often tends to separate into bimodal distributions representing life-limiting behavior and a mean-lifetime dominated behavior. The duality of the S-N curve is not an artifact but perhaps an unusual fatigue phenomenon caused by the material microstructure, such as the volume fraction of primary-α (αp) phase [7]. The nature of dual fatigue behavior can be termed as “competing failures” as both surface- and internally-initiated cracks can occurred with nearly equal probability. Chandran and Jha [8] reported that these two mechanisms compete to cause fatigue failure of Ti-10V-2Fe-3Al with α + β microstructures. This competing failure phenomenon leads to the unpredictability of how and when a specimen or a component will fail under a given test condition, leading an increased dispersion of fatigue lifetime data. In addition, researches show that the depth of crack-initiation sites show a stress ratio dependence and shifts from surface to interior with increasing of stress ratio [[22], [23], [24]]. In some cases, the apparent shorter- and longer-life-distribution in terms of the S-N curve would not be observed, although surface- and internal-failures are evident [25]. To date, there is no convincing explanation for the presence of dual S-N curves. It is worth noting that some researchers have suggested that competing failures will not occur in notched specimens because the stress is concentrated in the root of the notch [26]. Reportedly, the behavior of the fatigue data of a nickel-based superalloy notched fatigue specimens showed an effect of competing modes, but the fatigue lifetime could be adequately represented by a single S-N curve [27]. Also, the stress state at a notch is multiaxial even under uniaxial loading, leading to large scatter in fatigue lifetime for notched specimens. The failure competition in notched fatigue behavior is still ambiguous and to be explored further.

FCI in metallic materials without significant metallurgical defects occurs due to the cyclic damage accumulation of irreversibly slipping dislocations. Plenty of studies on near-α and α + β titanium alloys evidenced that FCI is accompanied by facet formation across αp particles [[28], [29], [30], [31], [32], [33], [34]]. Besides, many studies have suggested that the FCI facets form on or very near basal planes [[28], [29], [30], [31], [32],35,36], although FCI by prismatic facet formation has also been reported [28]. In terms of the faceting process, three kinds of mechanisms have been proposed and summarized in Refs. [17,18]. The pure slip is suggested as the most dominant factor in facet formation [30].

In the present work, we have discovered the occurrence of a clear duality of the S-N fatigue curve in the Ti-7Mo-3Nb-3Cr-3Al (Ti-7333 here afterwards) notched specimens. The two competing failure mechanisms are discussed based on microstructure observation and fractographic analysis. Furthermore, high cycle fatigue (HCF) damage mechanism was analyzed by electron back-scattered diffraction (EBSD) in a focused-ion-beam cross-section (FIB-CS) across the faceted grains, emphasizing the slip modes of faceted grains.

2. Materials and experimental procedures

2.1. Materials

The material of interest in the present study is a metastable β titanium alloy Ti-7333. Its chemical composition (wt%) is given in Table 1. The β transus temperature (Tβ) was measured to be approximately 850 ℃ by using the metallographic observation method. More specifically, the block specimens extracted from the Ti-7333 forged stocks were solution treated at 860 ℃, 855 ℃, 850 ℃, 845 ℃, 840 ℃ and 830 ℃ for 40 min with subsequent water quenched, then, these specimens were characterized using scanning electron microscope to detect the α phase percentage. Three kinds of forging stocks after different thermo-mechanical processing (TMP) routes were used to evaluate fatigue behavior. On one hand, the Ti-7333 square billets were forged at α + β region (∼790 ℃) and near β region (∼835 ℃), then, the same duplex annealing heat treatment process was conducted on fatigue specimen blanks machined from the forged plates by electric-discharge machining. To be precise, these specimens were solution treated at 820 ℃ for 1 h, furnace cooled, then annealed at 765 ℃ for 1 h, air cooled, and aged at 520 ℃ for 6 h, air cooled. The corresponding microstructures denoted by Ti7333-A and Ti7333-B, respectively. On the other hand, the fatigue specimen blanks machined from Ti-7333 forged bar were solution treated at 820 ℃ for 50 min, air cooled, and then aged at 540 ℃ for 6 h, air cooled, hereafter referred to as Ti7333-C.

Table 1   Chemical compositions of Ti-7333 alloy used in this study (wt%).

TiMoNbCrAlFeONC
Bal.7.143.003.103.040.050.120.0090.018

New window| CSV


2.2. High cycle fatigue experiments

The notched cylindrical fatigue specimens with an elastic stress concentration factor of Kt = 3 were machined using electrical discharge machining (EDM) followed by lathe turning. The shape and dimension of the fatigue specimens is shown in Fig. 1. HCF experiments were performed in tension-tension fatigue (R = 0.1) using an electromagnetic resonance high frequency fatigue testing machine (QBG-100B high-frequency fatigue tester) in accordance with the HB 5287 “Metal Material Axial Loading Fatigue Test Method”. The tests were run in stress-controlled at room temperature and a lab-air environment, using a sinusoidal waveform with a frequency of around 120 Hz. Tests were stopped until the specimen failed completely or fatigue lifetime reached 107 cycles (run-out). It should be pointed out that the lifetime data of unbroken specimens do not participate in S-N curve fitting but still show in the S-N scatter plot. The fatigue strength at 107 cycles was determined by staircase test method using 13, 12 and 10 specimens for Ti7333-A, Ti7333-B and Ti7333-C, respectively.

Fig. 1.

Fig. 1.   Picture (a) and schematic representation (b) (in mm, not to scale) of a cylindrical dog-bone-shaped notched specimen with Kt = 3. The blue dashed box shows the enlarged view of gage section of specimen with a circumferential V-notch.


2.3. Characterization methods

The initial microstructures were characterized by using scanning electron microscope (SEM, Carl Zeiss Microscopy GmbH) and transmission electron microscope (TEM, FEI Themis Z). The samples for SEM were prepared following standard grinding procedures and subsequent electropolished (Struers LectroPol-5) in a solution of 5% perchloric acid in ethanol at 35 V for 40 s. TEM foils were cut perpendicular to the cyclic loading direction by an EDM and were subsequently mechanically ground to ∼60 μm followed by electrolytic polishing in a solution of 5% perchloric acid in methanol at a voltage of 45 V, using a Struers Tenupol-5 twin-jet electropolisher. Post-mortem fractographic analysis were performed using a SEM (TESCAN MIRA3 SEM) in order to determine fracture surface details.

The characterization of FCI site involved measurement of the spatial angles of the faceted planes with respect to the loading direction (LD) and an analysis of crystallographic orientation of crack-initiation grains and their neighbors. The measurement of spatial angle of facets was accomplished by using a quantitative tilt fractography (QTF) technique [30,37]. Stereo-image pairs of facets were acquired in the SEM at tilt angles of 0° and 30°. Crystallographic information of faceted grains and their neighbors was obtained via EBSD (OXFORD NordlysNano) analysis on a 2D FIB-CS across the faceted grains. A FEI Helios G4 CX FIB equipment was used for machining this section. A focused beam of Ga+ ions was used at an accelerating voltage of 30 kV and ion beam current of 65 nA, 9.3 nA and 0.23 nA are used in sequence. TEM foil at crack-initiation site was lifted out by FIB. A low Ga+ current of 80 pA was used for the final thinning. EBSD analysis was carried out under a voltage of 20 kV and a current of 6.0 nA. The Channel 5 software was used to obtain the EBSD measurements. Following EBSD data acquisition, the slip trace maps of β grain were generated using raw EBSD orientation data [38,39]. The misorientation between a data point and its neighbors was analyzed using the Kernel average misorientation (KAM) map to approximate measure the geometrically necessary dislocation (GND) density [40,41].

3. Results

3.1. Characterization of initial microstructures

As-received microstructures for fatigue tests are shown in Fig. 2 and the corresponding microstructure parameters are listed in Table 2. Apparently, the microstructures consists of different volume fractions of αp particles distributed in a matrix of βt structure, and the βt further consisted of acicular secondary-α (αs) precipitates. Besides, the globular αp particles outnumber the elongated αp particles. In this paper, the αp particles with an aspect ratio greater than 2.5 are defined as the elongated αp particles. It should be noted that an almost continuous precipitation of grain boundary α (GBα) phase is observed in Ti7333-C (Fig. 2(i)). The tensile mechanical properties of the three kinds of microstructures are shown in Fig. 3. It shows that the Ti7333-C has the highest strength and give an ultimate tensile strength (UTS) of 1434 MPa, yield strength (YS) of 1385 MPa and the Ti7333-B exhibits the best plasticity with an elongation to fracture of 12.1%. Fig. 4 shows the EBSD results of the initial microstructure of Ti7333-B. As shown in phase map and the corresponding inverse pole figure (IPF) map, the different αp variants are distributed in the β grain, and three αp variants prefer to precipitate. Fig. 4(c) represents the pole figures of ten selected α particles and β grain labeled in Fig. 4b. The objective here is not to prove that all of the 12 α variants exist but to verify the Burgers orientation relationship (BOR) between β phase and selected α particles after TMPs. From Fig. 4(c), except for α6 and α9, other eight α particles exhibit strict BOR relation with β grain. This indicates that a majority of intragranular α particles maintain a strict BOR relation with its β phase.

Fig. 2.

Fig. 2.   SEM/SE and TEM micrographs of Ti7333-A (a-c), Ti7333-B (d-f), and Ti7333-C (g-i). SEM/SE images (a and b, d and e, g and h) show the αp particles and transformed-β matrix (β+αs). TEM bright-field images (c, f and i) show the αs precipitates. The cross sections parallel to the loading direction were used for SEM microstructure observations.


Table 2   Thermo-mechanical processing routes and corresponding microstructure descriptions of Ti-7333 alloy.

IDTMP routesVf of αp (%)D of αp (μm)
Ti7333-A790 ℃ forging+820 ℃/1 h/FC → 765 ℃/1 h/AC + 520 ℃/6 h/AC205
Ti7333-B835 ℃ forging+820 ℃/1 h/FC → 765 ℃/1 h/AC + 520 ℃/6 h/AC234
Ti7333-C820 ℃/50 min/AC + 540 ℃/6 h/AC72

Vf : volume fraction; D: average diameter; AC: air cooling; FC: furnace cooling.

New window| CSV


Fig. 3.

Fig. 3.   Tensile mechanical properties and fatigue strength of Ti-7333 alloy (YS: yield strength; UTS: ultimate tensile strength; EL: elongation; RA: reduction of area; FS: fatigue strength at 107 cycles and Kt = 3).


Fig. 4.

Fig. 4.   (a) EBSD-derived phase map of the initial microstructure of Ti7333-B, (b) the corresponding inverse pole figure (IPF) map and (c) pole figures of selected α particles and β grain, showing that a majority of intragranular α particles obey the Burgers orientation relationship (BOR) with β phase. The cross-section parallel to the loading direction was used for EBSD testing.


3.2. Duality of the S-N fatigue curve

The S-N fatigue curves of three kinds of microstructures are presented in Fig. 5. An interesting phenomenon, called the duality of the S-N fatigue behavior, is observed, in which the data points grouped into two separate curves. Fatigue variability behavior are marked by an increase in the lifetime variability to almost two orders in magnitude as the stress level is decreased, thereby promoting increased fatigue variability. The fatigue data of short-life-distribution and long-life-distribution overlapped each other, indicating a superposition of two mechanisms. Besides, it can be seen that the complete separation of the two failure distributions in terms of fatigue life (Fig. 5(b) and (c)) or they can dominate at high and low stress ranges with step in the mid-stress-range (Fig. 5(a)). It is worth noting that there is very little fatigue data point at the life cycles from 105 to 106. The fatigue lives decrease with steep slope in the short-life-distribution of Ti7333-C (Fig. 5(c)), indicating that the fatigue life decrease rapidly with increasing of stress level. The fatigue strength of 50% survival rate is shown in Fig. 3. It shows that fatigue strength values of three kinds of microstructures are comparable, and the Ti7333-C exhibits the highest fatigue strength, 335 MPa.

Fig. 5.

Fig. 5.   Two-stage S-N fatigue behavior in Ti7333-A (a), the dual S-N fatigue behavior in Ti7333-B (b) and Ti7333-C (c). Arrows denote that the tests were terminated when the specimen survived by 107 cycles.


3.3. Fractographic analysis

The careful fractographic investigation in four representative specimens is presented in Fig. 6. As expected, surface crack-initiation was found in all failed specimens due to the stress concentration at the location of notch root. As shown in Fig. 6(a), (d), (g) and (i), the macroscopic fracture surfaces are relatively flat. For the short-lifetime failed specimen (shown in Fig. 6(a) and (b)), the crack origin can trace back to a surface crack-initiation site without the presence of facets, with deformation accumulation rapidly for crack-initiation in only a few cycles, and the corresponding fatigue lifetime was controlled by the crack-growth process. For the long-lifetime failed specimens (shown in Fig. 6(c-k)), however, the crack origin can be traced back to a surface crack-initiation site with the presence of facets, and the corresponding fatigue lifetime was dominated by the crack-initiation process. These facets were formed across αp particles, which were confirmed by examining both the fractured specimen and the FIB-CS characterization. It can be seen that an elongated αp facet (Fig. 6(e) and (k)) or an equiaxed αp facet (Fig. 6(h) and (k)) locates on or just slightly below the specimen surface. The presence of the elongated αp facets indicates that the crack-initiation favor in elongated αp particles although the number of equiaxed αp particles is much more than the elongated αp particles. This indicates that the combination of crystallography and morphology of αp particles plays a role in generating the critical conditions necessary for facet formation. Additionally, the equiaxed dimples and tear ridges (Fig. 6(c)) and sometimes rough fatigue striations (Fig. 6(f)) can be seen in fatigue crack-growth (FCG) region due to the poor plasticity of this alloy.

Fig. 6.

Fig. 6.   Typical fractographies of four notched specimens showing surface crack-initiation with- and without-facets. The overall fracture surface morphologies of Ti7333-A alloy specimen (σmax =400 MPa, Nf = 57900 cycles a, Ti7333-B alloy specimen σmax =330 MPa, Nf = 4.66 × 106 cycles) (d), Ti7333-C alloy specimen (σmax =340 MPa, Nf = 8.04 × 106 cycles) (g) and Ti7333-C alloy specimen (σmax =400 MPa, Nf = 3.408 × 106 cycles) (i); while magnified views of crack-initiation sites are shown in (b, e, h and k), and magnified views of crack-growth regions are shown in (c and f).


3.4. Characterization of fatigue crack-initiation site

As mentioned above, all long-lifetime failures occurred by crack-initiation in the specimen surface with the presence of αp facets. The FCI site in a representative specimen is presented in Fig. 7. This specimen was tested at 330 MPa, and the lifetime was 4,660,000 cycles. The careful fractographic investigation is presented in Fig. 6(e). The inclination of the facet-normal with respect to the LD was about 44°. This suggests a strong role of pure shear deformation along the facet plane in the formation of the αp facet.

Fig. 7.

Fig. 7.   FIB cross-section and EBSD characterization of a long-lifetime failure specimen microstructural arrangement (σmax =330 MPa, Nf = 4.66 × 106 cycles): (a) faceted αp particles and the neighboring grains shown on the FIB-CS; (b) EBSD-derived IPF//LD map of the crack-initiation neighborhood showing contiguous αp facets; Corresponding phase map are shown in (c); (d) the corresponding Kernel average misorientation (KAM) map showing the heterogeneous distribution of KAM values between αp phase and β phase. {110} slip traces and 3D-view of crystal are superimposed on maps. Schmid factor values for each slip system in β grain are listed here.


In order to explore the crystallographic plane of the αp facet and the deformation mode of the faceted grain and the β grain, this specimen was first mechanically grinding to get close to the FCI site, then a 2D section was machined across faceted grains and parallel to the LD by FIB milling, as shown in the secondary electron image of Fig. 7(a). The specimen tilt in this image is 52°. The αp particles and β+αs matrix underlying the crack-initiation facet were revealed on this section. It is evident from the figure that the facets (designated as F1, F2 and F3) were formed across three adjacent αp particles, resulting in an elongated facet. Also labeled are the neighbor αp particles, designated as N1, N2, N3, N4 and N5. The EBSD-derived IPF//LD map is presented in Fig. 7(b). The corresponding phase map is shown in Fig. 7(c). The 3D-view of crystal of the faceted αp particles indicates that the edges of the facet planes were close to the basal plane. In order to characterize this phenomenon more accurately, the angles of the basal plane normal of three facets respect to the LD are determined with Euler angles of these faceted grains, which are presented in Table 3 and are about 43°, 38° and 44° for F1, F2 and F3, respectively. This demonstrates that the crystallographic planes of the facets are less than 10° inclination from the basal plane. Schmid factor can be used as a parameter (but not the sole parameter) to determine whether the slip system activation or not in a given grain during fatigue loading [17,33]. The Schmid factors, along the LD, calculated for five slip systems of faceted αp particles and neighbor αp particles are shown in Table 4. It can be seen that the basal <a> slip is the dominant deformation mode in each of the faceted αp particles where the Schmid factor values for basal slip are 0.49, 0.48 and 0.48, respectively, for the three faceted αp particles, although the facet-normal has an inclination of about 44° to the LD. These results strongly suggest that these facets were formed by almost pure shear deformation.

Table 3   A comparative analysis of the angles of base normal respective to loading direction (LD) and spatial angle of facet plane.

IDEuler anglesThe angles of base normal respective to LDSpatial angle of facet
φ1Фφ2
F121.1°125.2°26.6°43°44°
F217.7°125.1°15.8°38°
F319.4°122.0°31.7°44°

New window| CSV


Table 4   Schmid factors for αp particles (the highest Schmid factor values are shown in bold text).

IDBasal {0002}<11-20>Prismatic {10-10}<11-20>Pyramidal {10-11}<11-20>First-order pyramidal {10-11}<11-23>Second-order pyramidal {11-22}<11-23>
N10.340.400.430.370.38
F10.490.200.370.410.27
F20.480.200.390.420.30
F30.480.150.330.420.29
N20.470.310.460.320.17

New window| CSV


It can be inferred that the facet formation process ought to be influenced by the neighboring αp particles and β grain. The slip modes in the first-nearest-neighbor αp particles (designated as N1 and N2) and β grain were examined with the help of Schmid factor. It can be seen that the N2 was oriented for basal slip, and no clear trend in deformation modes for N1 particle, as shown in Table 4. For a better understanding of the deformation modes of the β grain, the angles between the all possible {110} slip plane traces and the LD and the corresponding Schmid factors values were determined using Euler angles. As shown in Fig. 7(b), the {110} slip traces are superimposed on IPF map, and corresponding Schmid factors are also presented. The crack-initiation site show that the rugged β fracture is roughly parallel to the slip plane with highest Schmid factor, which is (-101)[-11-1] in this specimen, among all slip systems available. It should be noted that the β fracture is very rough due to there is no a dominant slip plane, i.e. the Schmid factor of 0.44 in.(101)[1-1-1], 0.39 in.(110)[1-11] and 0.39 in.(-110)[-1-1-1]. Additionally, there is no clear β facet appear in the notched specimen may be due to the unique stress distribution at the notch root.

The EBSD-derived pole figures of β phase and αp particles from the Fig. 7(b) IPF map are displayed in Fig. 8 to analyze the orientation relationship between α-β phases. The pole figures illustrate that there is a strict Burgers’ relationship between N3, N5 and β phase, which obeys the Burgers orientation relationship i.e., {110}β//{0001}α, <-11-1>β//<11-20>α [42,43]. Besides, the faceted αp particles (F1, F2 and F3) have a near-Burgers orientation relationship with respect to the β grain. However, N1, N2 and N4 exhibit no Burgers’ relationship between α-β phases due to the stress [44].

Fig. 8.

Fig. 8.   EBSD-derived pole figures of the faceted αp particles, neighbor αp particles and β grain in crack-initiation area.


4. Discussion

4.1. Effect of microstructure on the fatigue behavior

The TMP routes will affect the microstructure features and further affect the fatigue strength and fatigue S-N behavior. As shown in Fig. 3, three kinds of microstructures exhibit the comparable fatigue strength values, although the tensile strengths are obviously different. Generally, the fatigue strength of the alloys is positively correlated with the yield strength (YS). The differences in the static strength of these three alloys are associated with the αs phase characteristics, because of the strengthening effect of the metastable β titanium alloys is mainly dependent on the αs phase characteristics. A majority of researches show that the FCI by facet formation across αp particles, and that includes the Ti-7333 alloy. The Ti7333-A and Ti7333-B offer very similar αp grain sizes, shape and distribution, resulting in comparable fatigue strength. Nevertheless, the Ti7333-C also shows the comparable fatigue strength. It can be speculated that the notch effect weakens the effect of microstructure factor, such as αp phase characteristic, on crack-initiation. These indicate that the fatigue strength of notched specimens at HCF regime is closely related to the notch effect, i.e., the stress state and stress distribution in the vicinity of the notch root. As shown in Fig. 5, compared with the Ti7333-A and Ti7333-B, the Ti7333-C shows a smaller scatter in fatigue data of the short-lifetime failed specimens. Given that the initial microstructure of Ti7333-C contains a smaller αp phase percentage, the surface crack-initiation in this microstructure is closely associated with the β+αs matrix due to a small number of αp particles in the notch root. This weakens the effect of αp particles, which is closely related to the formation of the facet, on FCI process. Further, the αp particle characteristics, i.e., the crystallographic orientation, size and the morphology, contribute significantly to variability of fatigue life and minimum fatigue life.

4.2. Competing fatigue failures in notched specimen

As revealed by fatigue fracture surface and the S-N curves analysis, the dual fatigue behavior is found to result from variability in fatigue life due to a probabilistic selection between two competing fracture mechanisms. Unlike the smooth cylindrical specimen failures where short- and long-life-distributions of S-N curve result from surface- and internal-crack-initiation, respectively [8], the duality of the S-N fatigue behavior of Ti-7333 notched specimens is associated with the surface-failure-without- and with-facets. Moreover, the long-lifetime failed specimens appear to produce fatigue failure after a significant delay (about two orders of magnitude at the lowest applied stresses) with respect to the short-lifetime failed specimens. At a given cyclic stress range, whether a specimen would fail with- or without-facets, it is often not predictable before the test. Given the identical test conditions and the test environment for each of microstructure of Ti-7333, as well as there are no inclusions or pores in the microstructures, except for special forming process such as powder metallurgy [45], the duality of the fatigue behavior reported here is not an artifact but does seem to be an unusual fatigue phenomenon mainly originating from the unique microstructure level, such as αp and αs characteristic, and the notch effects. As expected, the stress state and stress distribution in the vicinity of the notch root can significantly affect the process of formation of fatigue micro-cracks. The reduced fatigue lives in surface-without-facet specimens are due to the acceleration of surface crack-initiation process in the lab-air environment as compared to the quasi-vacuum conditions in the interior of the specimens.

4.3. Phenomenological model for high cycle fatigue crack-initiation

For multi-phase alloys, the phase boundaries are the preferential crack-initiation site, resulting in a decline in the resistance of crack-initiation during cyclic loading. This is due to the discontinuity of slip transfer and the difference of elastic modulus between α and β phases. As discussed above, fatigue micro-cracks favorably initiated from an area of microstructural weakness where αp particles favorably oriented for basal <a> slip. It can be inferred that edges and the sharper corners of αp particles are the geometrically weak sites for micro-cracks initiation, as observed the elongated αp facets in FCI site (Fig. 6(e) and (k)). The strain gradient distribution in FCI site was evaluated in terms of KAM map. As shown in Fig. 7(d), the heterogeneously distributed KAM values are observed in αp particles and β+αs matrix. The KAM value in β+αs matrix is higher than that observed in αp particles. Moreover, the KAM value in faceted αp particles is slightly higher than that in αp particles beneath the fracture surface. Fig. 9 shows the EBSD results (50 nm step size) in crack-initiation region of Ti7333-A specimen. The KAM map (Fig. 9(c)) and the corresponding phase map (Fig. 9(b)) also show the heterogeneously distributed KAM values and the KAM value in β phase is higher than that of α phase, resulting in the strain localization within β phase. Interestingly, a high KAM value is observed at the subgrain boundaries, as shown in Fig. 9(a) and (c). This indicates that dislocation slip occurs within the αp particles during cyclic loading. On the other hand, a TEM foil at the crack-initiation site of the Ti7333-A shows that the significant dislocation activity in the form of dislocation tangles in β+αs matrix (Fig. 10(a)). Additionally, the faceted αp particle reveals the presence of the subgrain boundary, as shown in HAADF image (Fig. 10(b)) and the corresponding SAED patterns (Fig. 10(c) and (d)).

Fig. 9.

Fig. 9.   (a) EBSD-derived IPF//LD map of the fatigued microstructure from Ti7333-A showing the crystallographic orientation of α and β phases. (b) The corresponding phase map showing the distribution of α phase and β phase. (c) The corresponding KAM map with step size of 50 nm showing the heterogeneous distribution of KAM values between αp phase and βt matrix. The insets of local misorientation angle distribution shows that the β phase has a larger percentage of high misorientation angle distribution.


Fig. 10.

Fig. 10.   TEM micrographs showing the dislocation structures of the fatigued microstructure in the crack-initiation region: (a) bright field image showing significant dislocation tangles in the β+αs matrix; (b) HAADF image showing the subgrain boundary in the faceted αp particle. The SAED patterns taken from (b) are shown in (c, d and e).


The small fatigue cracks revealed by FIB milling were found to form on basal plane of αp particle. Schematic for phenomenological model of the HCF crack-initiation in Ti-7333 alloy with bimodal microstructure is shown in Fig. 11. Under fatigue loading, cyclic microplastic deformation first occurs in β+αs matrix result in a mass of GNDs and statistically stored dislocations (SSDs). Besides, αp particles can act as a barrier to dislocation motion, resulting in the dislocation pileup around αp particles. Only the basal and prism slip systems most suitably orienting for slip will be activated due to that the stress levels are far below the YS of the alloy. The <c+a> slip are generally inactivated for the stress levels that produce fatigue life greater than 106 cycle [30]. Therefore, the basal dislocation pileup begins to form at the αp particle boundaries, especially in elongated αp particles and multiple adjacent αp particles well orientated for basal slip, until a dislocation through phase boundary. This results in the formation of a residual boundary dislocation. Once the residual boundary dislocations continue to accumulate to a critical value, a fatigue crack nucleus is formed and then propagates back along the basal slip band under the tension stress, leading to the presence of the αp facet. Then, fatigue crack propagates along (110) slip plane with the high Schmid factor value to form a β facet or not, depending on the crystallographic orientation of β grain. Subsequent crack propagation is no longer sensitive to microstructure and shows the typical equiaxed dimples and fatigue striations in FCG region (Fig. 6(c) and (f)). It should be noted that the size of αp facet in Fig. 6(e) and (k) is slightly larger than that in Fig. 2. This is because the larger the grain size, the larger the dislocation slip path, the faster the dislocation pileup and the easier the fatigue crack nucleation. In general, fatigue strength depends on the lattice strength and the dislocation slip path of the alloys. Therefore, the large or elongated αp particles orientating for basal slip often act as the FCI sites. This provides a deep understanding of the fatigue damage mechanisms by correlating crystallographic orientation and morphology of αp particles. The faceting process and early stages of crack growth are the particularly important points due to the majority of the total fatigue life are typically spent in this regime. However, for the short-lifetime failed specimens, FCI occurred at the specimen surface without the presence of facets. This is because that the fatigue damages accumulation rapidly for surface crack-initiation in only a few cycles, leading to a shorter fatigue lifetime. As shown in Fig. 2, typical microstructure consists of an equiaxed grained β matrix and two populations of α precipitates, including larger equiaxed αp particles and acicular αs precipitations. These fine acicular αs precipitations contribute a lot to the static strength improvement. The fracture observation and FIB-CS analysis suggest that the FCI is strongly related to the αp particles more than the αs precipitations. Nevertheless, FCI process is inevitably associated with αs precipitations due to their significantly strengthening effect and the αp particles are surrounded by the β+αs matrix, but unfortunately, we fail to study this αs effect because they are too small to be characterized and no αs facets were observed in the FCI sites.

Fig. 11.

Fig. 11.   Schematics for phenomenological models for fatigue crack-initiation: (a) the dislocation pileup is formed at αp phase boundaries at the beginning of fatigue loading; (b) the basal dislocation pile up is formed within αp particle until a dislocation through phase boundary, leaving a residual boundary dislocation. The residual boundary dislocations continue to accumulate to a critical value, leading to facet formation along basal plane of αp particles under the tension stress, followed by short crack propagates along (110) slip plane with the highest Schmid factor value.


5. Conclusions

In this work, the duality of the S-N curve in notched specimens of Ti-7333 is reported, and the fatigue damage mechanism was revealed by FIB-CS characterization across the crack-initiation site and surrounding neighborhoods and TEM characterization of crack-initiation site. The following conclusions can be drawn:

(1) The S-N fatigue curves of notched Ti-7333 specimens exhibit duality. The short- and long-life-distribution occurs due to surface failure without and with facets. Fatigue strength at high cycle fatigue regime is associated with static strength and microstructure factor, and strongly dependent on the notch effect.

(2) For long-lifetime failures, crack-initiation occurred by facet formation in αp particles, and the facet plane was inclined closely to the maximum shear orientation with respect to the loading direction (45°). Faceted αp particles were found to be well-oriented for basal <a> slip activation, and no apparent β facet appeared in the notched specimen. Besides, the αp facet planes almost corresponded to the basal plane, and the faceting process was largely controlled by the shear deformation mode.

(3) FCI is strongly related to the αp particles more than the αs precipitations. The morphological and crystallographic effects of αp particles are significantly responsible for facet formation. Elongated αp particles favorably oriented for basal slip is a preferential fatigue-critical microstructural configuration. FCI process is inevitably associated with αs characteristic because of its significantly strengthening effect.

(4) The β+αs matrix has a higher KAM value than the αp particles in fatigued microstructures. Significant dislocation activity in the form of dislocation tangles is observed in αp boundaries and leads to crack-initiation along basal plane of αp particles.

(5) The duality of the fatigue behavior of notched Ti-7333 is found to be resulted from the competing fatigue mechanisms, and mainly originates from the unique microstructure level and the stress state and stress distribution in the vicinity of the notch root.

Acknowledgements

This work was financially supported by the National Key Research and Development Program of China (No. 2016YFB0701303), the National Natural Science Foundation of China (No. 51801156) and the Natural Science Basic Research Plan in Shaanxi Province of China (No. 2019JM-584). The authors also appreciate the help from Dr. Ying Ding at Analytical & Testing Center of NPU for FIB milling.

Reference

Y.F. Zheng, R.E.A. Williams, D. Wang, R. Shi, S. Nag, P. Kami, J.M. Sosa, R. Banerjee, Y. Wang, H.L. Fraser, Acta Mater. 103 2016 850-858.

DOI      URL     [Cited within: 9]

A.F. Gerday, M.B. Bettaieb, L. Duchene, N. Clement, H. Diarra, A.M. Habraken, Acta Mater. 57 2009 5186-5195.

DOI      URL     [Cited within: 1]

N. Jones, R. Dashwood, D. Dye, M. Jackson, Mater. Sci. Eng. A 490 (2008) 369-377.

DOI      URL     [Cited within: 1]

C.M. Liu, X.J. Tian, H. Tang, H.M. Wang, J. Alloys. Compd. 572 2013 17-24.

DOI      URL     [Cited within: 1]

C.W. Huang, Y.Q. Zhao, S.W. Xin, C.S. Tan, W. Zhou, Q. Li, W.D. Zeng, Int. J. Fatigue 94 (2017) 30-40.

DOI      URL     [Cited within: 1]

X.H. Shi, W.D. Zeng, S.K. Xue, Z.Q. Jia, J. Alloys. Compd. 631 2015 340-349.

DOI      URL     [Cited within: 1]

S.K. Jha, K.S.R. Chandran, Scr. Mater. 48 2003 1207-1212.

DOI      URL     [Cited within: 2]

K.S.R. Chandran, S.K. Jha, Acta Mater. 53 2005 1867-1881.

DOI      URL     [Cited within: 3]

AbstractThe mechanism of the duality of the S–N fatigue data (S is stress and N is cycles to failure) or the grouping of data into two distinct S–N curves as a result of two competing failure modes, occurring during fatigue of some α + β microstructures of the metastable beta-titanium alloy Ti–10V–2Fe–3Al, is quantitatively assessed in this research. The competing failure modes result in two separate fatigue curves in this material: one for failures from surface-nucleated cracks and the other for the subsurface-nucleated. The consequence of this behavior is that one can not predict exactly how a specimen or component would fail, regardless of the number of samples tested. This work discusses in detail the microstructural conditions that trigger this phenomenon in the titanium alloy studied. Based on Poisson defect statistics and Monte-Carlo simulations, it is shown that this phenomenon can arise from a Poisson spatial pattern of low populations of microscale defects. An interesting finding is that there should be a combination of average defect density and specimen area at which this phenomenon is observable in any material having sparsely populated and Poisson-distributed defects. The complete analysis presented here also helps to explain the duality of fatigue failures caused by competing failure modes observed in steels and superalloys.]]>

K.S.R. Chandran, P. Chang, G.T. Cashman, Int. J. Fatigue 32 (2010) 482-491.

DOI      URL     [Cited within: 1]

H. Yokoyama, O. Umezawa, K. Nagai, T. Suzuki, ISIJ Int. 37 1997 1237-1244.

DOI      URL     [Cited within: 1]

O. Umezawa, K. Nagai, K. Ishikawa, Mater. Sci. Eng. A 129 (1990) 217-221.

DOI      URL     [Cited within: 3]

K. Tokaji, K. Ohya, K. Kariya, Fatigue Fract. Eng. Mater. Struct. 23 2000 759.

DOI      URL     [Cited within: 1]

Q.Y. Wang, J.Y. Berard, A. Dubarre, G. Baudry, S. Rathery, C. Bathias, Fatigue Fract. Eng. Mater. Struct. 22 1999 667.

DOI      URL     [Cited within: 1]

A. Atrens, W. Hoffelner, T.W. Duerig, J.E. Allison, Scr. Metall. 17 1983 601.

DOI      URL     [Cited within: 1]

J.M. Hyzak, I.M. Bernstein, Metall. Trans. A 13 (1982) 33-43.

DOI      URL     [Cited within: 1]

S.K. Jha, J.M. Larsen, A.H. Rosenberger, Acta Mater. 53 2005 1293-1304.

DOI      URL     [Cited within: 1]

AbstractThe influence of competing mechanisms on the fatigue life behavior of a γ-TiAl based alloy was studied at room temperature and 600 °C. The alloy exhibited almost flat stress–life (S–N) behavior at both temperatures. This type of behavior was found to result from variability in life due to a probabilistic selection between two mechanisms, viz. failure by surface crack nucleation and that by subsurface crack nucleation. It is suggested that the S–N behavior of the alloy is actually composed of dual curves, over the same range of stresses, corresponding to each mechanism. This improved description of the S–N behavior (in terms of dual curves) of the alloy led to a significant reduction in the apparent uncertainty in life.]]>

S.K. Jha, C.J. Szczepanski, R. John, J.M. Larsen, Acta Mater. 82 2015 378-395.

DOI      URL     [Cited within: 3]

S.K. Jha, C.J. Szczepanski, P.J. Golden, W.J. Porter III, R. John, Int. J. Fatigue 42 (2012) 248-257.

DOI      URL     [Cited within: 2]

J.M. Larsen, S.K. Jha, C.J. Szczepanski, M.J. Caton, R. John, A.H. Rosenberger, D.J. Buchanan, P.J. Golden, J.R. Jira, Int. J. Fatigue 57 (2013) 103-112.

DOI      URL     [Cited within: 1]

S.K. Jha, M.J. Caton, J.M. Larsen, Mater. Sci. Eng.A 468-470 2007 23-32.

[Cited within: 1]

S.K. Jha, J.M. Larsen, A.H. Rosenberger, Eng. Fract. Mech. 76 2009 681-694.

DOI      URL     [Cited within: 1]

W. Li, H.Q. Zhao, A. Nehila, Z.Y. Zhang, T. Sakai, Int. J. Fatigue 104 (2017) 342-354.

DOI      URL     [Cited within: 1]

Y. Wu, J.R. Liu, H. Wang, S.X. Guan, R. Yang, H.F. Xiang, J. Mater. Sci. Technol. 34 2018 1189-1195.

DOI      URL     [Cited within: 1]

X.L. Liu, C.Q. Sun, Y.S. Hong, Int. J. Fatigue 92 (2016) 434-441.

DOI      URL     [Cited within: 1]

K. Yang, C. He, Q. Huang, Z.Y. Huang, C. Wang, Q.Y. Wang, Y.J. Liu, B. Zhong, Int.J. Fatigue 99 (2017) 35-43.

[Cited within: 1]

G.T. Cashman, Int. J. Fatigue 32 (2010) 492-496.

DOI      URL     [Cited within: 1]

G.T. Cashman, A Perspective on Competing Modes Fatigue Behavior, in: Proc. Material Science and Technology (MS&T), Pittsburgh, PA, 5-8 October, 2008.

[Cited within: 1]

F. Bridier, P. Villechaise, J. Mendez, Acta Mater. 56 2008 3951-3962.

DOI      URL     [Cited within: 3]

AbstractThe purpose of this experimental study is to investigate the micromechanical fatigue behavior, in terms of slip nature and preferential cracking sites, of a commercial α/β-forged Ti–6Al–4V alloy. Electron backscattering diffraction is extensively used to identify the deformation (prismatic, basal, pyramidal slip) and crack formation modes activated by fatigue at the surface of several hundred primary α nodules. Some fatal crack formation sites are also characterized. Cracking in basal planes is identified as the most critical damage mode leading to fracture. An explanation is proposed which involves the resolved shear stress, taking into account the Schmid factor and the normal stress in relation to the elastic anisotropy of the α-phase. Finally, the spatial distribution of the secondary cracks is analyzed according to the crystallographic textures (macrozones) present on a mesoscopic scale in the Ti–6Al–4V alloy.]]>

M.R. Bache, W.J. Evans, H.M. Davies, J. Mater. Sci. 32 1997 3435-3442.

DOI      URL     [Cited within: 2]

A.L. Pilchak, R.E.A. Williams, J.C. Williams, Metall. Mater. Trans. A 41 (2010) 106-124.

DOI      URL     [Cited within: 5]

I. Bantounas, D. Dye, T.C. Lindley, Acta Mater. 58 2010 3908-3918.

DOI      URL     [Cited within: 2]

AbstractMicrotextured regions (or macrozones) are commonly reported in titanium alloys and are believed to be related to fatigue life. Here, fractographic investigations are conducted on bimodal Ti–6Al–4V plate, including transmission electron microscopy to determine the mechanism of fatigue facet formation and electron backscattered diffraction to examine the underlying macrozone structures. It is found that macrozones oriented with their c-axis close to the loading direction are responsible for facet formation, and that the facets are associated with basal slip. Microtextured regions with c-axis orientations near-perpendicular to the loading direction are believed to act as barriers to faceted crack growth, based on the change in crack morphology as the crack crosses a macrozone boundary. The variant selection occurring during the transformation of the retained beta appears to favour a common orientation with the surrounding primary alpha grains, contributing to the size of the macrozones and therefore to the extent of the observed faceted regions.]]>

I. Bantounas, D. Dye, T.C. Lindley, Acta Mater. 57 2009 3584-3595.

DOI      URL     [Cited within: 2]

AbstractThree different product forms of Ti–6Al–4V, unidirectionally rolled and cross-rolled plate and forged bar, have been cyclically loaded within the high-cycle fatigue regime to investigate fatigue crack initiation. The fracture surfaces of fatigued specimens contained large regions of neighbouring facets. The majority of facets examined had a near-basal fracture plane. It was shown that grains favouring crack initiation were primarily those with misorientations between 15° and 40° from the loading direction, whereas other orientations served as crack growth paths. This implies that the formation of a facet requires a combination of a moderately high Schmid factor for basal slip coupled with a tensile component perpendicular to the basal plane. The large regions of neighbouring facets on the fracture surface were found to be a consequence of cracking within a macrozone unfavourably oriented for slip found within the microstructure, i.e. one with its c-axis near to the loading direction.]]>

F. Bridier, P. Villechaise, J. Mendez, Acta Mater. 53 2005 555-567.

DOI      URL     [Cited within: 2]

AbstractThe gliding modes of a duplex Ti–6Al–4V titanium alloy were investigated through in situ (scanning electron microscopy) tensile tests. A method based on electron back-scattering diffraction (EBSD) measurements was used to identify activated slip systems. The approach applied to a large number of grains allowed a statistical analysis of the nature (basal, prismatic, pyramidal) and distribution of the slip systems according to the crystallographic texture. A discussion concerning the pertinence of Schmid’s law to explain the occurrence and succession of slip events is then proposed. The domain in favor of each type of slip system is finally presented by using inverse pole figures mapped with Schmid’s factor iso-curves.]]>

F.P.E. Dunne, D. Rugg, Fatigue Fract.Eng. Mater. Struct. 31 2008 949-958.

[Cited within: 1]

S.K. Jha, J.M. Larsen, VHCF-4, Random Heterogeneity Scales and Probabilistic Description of the Long-Lfetime Regime of Fatigue, Proceedings of the Fourth International Conference on Very High Cycle Fatigue (2007) 385-396.

C.J. Szczepanski, S.K. Jha, J.M. Larsen, J.W. Jones, Metall. Mater. Trans. A 39 (2008) 2841-2851.

DOI      URL    

V. Sinha, M.J. Mills, J.C. Williams, J. Mater. Sci. 42 2007 8334-8341.

DOI      URL     [Cited within: 1]

S. Birosca, F. Di Gioacchino, S. Stekovic, M. Hardy, Acta Mater. 74 2014 110-124.

DOI      URL     [Cited within: 1]

//loading direction) and near cube orientations (phi(1), Phi, phi(2): 0, 0-15, 0) behaved as "soft" grains with a high Schmid factor and contained low geometrically necessary dislocation (GND) density as a result of low strain hardening at the early stage of deformation for such grains. The near cube orientation (typically phi(1), Phi, phi(2): 0, 9, 0) was the softest orientation among the cube family. While the brass grains (//loading direction) acted as "hard" grains that have the lowest Schmid factor with the highest Taylor factor and GND density compared with other oriented grains. A high GND content was found in the vicinity of the grain boundaries in the soft grains and on slip plane traces within the hard grains. It is concluded that GND absolute value for each grain can vary, as it is interrelated with deformation degree, but the GND locations within the grains give indications of the strain hardening state and deformation stages in hard and soft grains. Furthermore, the areas with random local texture contained high strain incompatibilities between neighbouring grains, and generated microtexture clusters that prompt preferential secondary crack initiation and propagation. (C) 2014 Acta Materialia Inc. Published by Elsevier Ltd.]]>

F.P.E. Dunne, A.J. Wilkinson, R. Allen, Int. J. Plast. 23 2007 273-295.

DOI      URL     [Cited within: 1]

S.I. Wright, M.M. Nowell, D.P. Field, Microsc. Microanal. 17 2011 316-329.

DOI      URL     [Cited within: 1]

Since the automation of the electron backscatter diffraction (EBSD) technique, EBSD systems have become commonplace in microscopy facilities within materials science and geology research laboratories around the world. The acceptance of the technique is primarily due to the capability of EBSD to aid the research scientist in understanding the crystallographic aspects of microstructure. There has been considerable interest in using EBSD to quantify strain at the submicron scale. To apply EBSD to the characterization of strain, it is important to understand what is practically possible and the underlying assumptions and limitations. This work reviews the current state of technology in terms of strain analysis using EBSD. First, the effects of both elastic and plastic strain on individual EBSD patterns will be considered. Second, the use of EBSD maps for characterizing plastic strain will be explored. Both the potential of the technique and its limitations will be discussed along with the sensitivity of various calculation and mapping parameters.

M. Calcagnotto, D. Ponge, E. Demir, D. Raabe, Mater. Sci. Eng. A 527 (2010) 2738-2746.

DOI      URL     [Cited within: 1]

D. Banerjee, J.C. Williams, Acta Mater. 61 2013 844-879.

DOI      URL     [Cited within: 1]

The basic framework and - conceptual understanding of the metallurgy of Ti alloys is strong and this has enabled the use of titanium and its alloys in safety-critical structures such as those in aircraft and aircraft engines. Nevertheless, a focus on cost-effectiveness and the compression of product development time by effectively integrating design with manufacturing in these applications, as well as those emerging in bioengineering, has driven research in recent decades towards a greater predictive capability through the use of computational materials engineering tools. Therefore this paper focuses on the complexity and variety of fundamental phenomena in this material system with a focus on phase transformations and mechanical behaviour in order to delineate the challenges that lie ahead in achieving these goals.

T. Furuhara, S. Takagi, H. Watanabe, T. Maki, Metall. Mater. Trans. A 27 (1996) 1635-1646.

DOI      URL     [Cited within: 1]

K. Hua, Y.D. Zhang, H.C. Kou, J.S. Li, W.M. Gan, J.J. Fundenberger, C. Esling, Acta Mater. 132 2017 307-326.

DOI      URL     [Cited within: 1]

F. Cao, P. Kumar, M. Koopman, C. Lin, Z. Zak Fang, K.S.R. Chandran, Mater. Sci. Eng. A 630 (2015) 139-145.

DOI      URL     [Cited within: 1]

/