Next Article in Journal
Surface Modification with Phosphate and Hydroxyapatite of Porous Magnesium Scaffolds Fabricated by Binder Jet Additive Manufacturing
Next Article in Special Issue
Effects of Compositional Inhomogeneity on the Microstructures and Mechanical Properties of a Low Carbon Steel Processed by Quenching-Partitioning-Tempering Treatment
Previous Article in Journal
Correction: Li et al. Characterizations of Single-Crystal Lithium Niobate Thin Films. Crystals 2022, 12, 667
Previous Article in Special Issue
Effect of Short-Range Ordering on the Grain Boundary Character Distribution Optimization of FCC Metals with High Stacking Fault Energy: A Case Study on Ni-Cr Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Secondary α Phase on Stress Corrosion Cracking of a Novel Metastable β Titanium Alloy in 3.5% NaCl Solution

1
School of Materials Science and Engineering, Shenyang University of Technology, Shenyang 110870, China
2
College of Light Industry, Liaoning University, Shenyang 110036, China
3
ZS Advanced Materials Co., Ltd., Donggang 118305, China
4
Light and High Strength Materials, Institute of Metal Research, Shenyang 110016, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(12), 1849; https://doi.org/10.3390/cryst12121849
Submission received: 30 November 2022 / Revised: 11 December 2022 / Accepted: 16 December 2022 / Published: 19 December 2022
(This article belongs to the Special Issue Advances in High Strength Steels)

Abstract

:
The effect of the secondary α phase on stress corrosion cracking of a novel metastable β titanium alloy, Ti-6Mo-5V-3Al-2Fe, in 3.5% NaCl solution was investigated by slow strain rate testing. Fine acicular secondary α phase was obtained by aging at the low temperature of 520 °C, and coarsened rod-like secondary α phase was obtained by aging at the high temperature of 680 °C. The electrochemical measurement results and slow strain rate testing results show that the microstructure contained with fine acicular secondary α phase exhibits better corrosion resistance and less stress corrosion cracking susceptibility. The fracture morphology exhibits a mixed fracture characteristic with shallow and small dimples, as well as tear ridges and flat facets with undulating surfaces. The combination of Absorption Induced Dislocation Emission and Hydrogen Enhanced Localized Plasticity is the main mechanism for stress corrosion cracking. Fine acicular secondary α phase with narrow spacing leads to less accumulated dislocations and smaller localized stress, so that has a beneficial effect on stress corrosion performance.

1. Introduction

Titanium alloys exhibit high specific strength, sufficient fatigue resistance, and good toughness and are widely used for structural applications in many engineering fields [1,2,3]. Particularly, the superior corrosion resistance and excellent strength of lead titanium alloys are ideal for marine engineering applications [4]. For instance, titanium alloys have been used to make oil country tubular goods in the offshore petroleum exploration industry and cooling systems in seawater-cooled plants [5,6]. Seawater contains various kinds of solutes and microorganisms and gives a corrosion environment for metals [7]. A protective oxide film forms on the surface of titanium alloys because of spontaneous passivation, resulting in the corrosion resistance of seawater. However, the oxide film can be ruptured under the ongoing mechanical load and lead to the unprotected metal matrix being corroded by seawater. Therefore, titanium alloys for marine engineering structural applications may be suffered from stress corrosion cracking (SCC).
Many efforts have been made to investigate the SCC of titanium alloys. Pazhanivel et al. [8] investigated the SCC of Ti-6Al-4V alloy in an aqueous 3.5% NaCl environment by performing a slow strain rate test. They demonstrated that the loss in elongation of the specimen in an aqueous 3.5% NaCl environment is attributed to corrosion susceptibility and hydride formation at the α/β interface. Leon et al. [9] proposed that the decreased amount of α phase and increased amount of β phase has a beneficial effect on the structural integrity of the passive protective film. This is one of the reasons why the stress corrosion performance of Ti-6Al-4V alloy in 3.5% NaCl aqueous solution is improved. Several findings show that corrosive attack mainly occurs at the α/β interface because of the micro-galvanic effect between α and β phases so that the SCC tends to take place at such place in a 3.5% NaCl environment [10,11,12]. Li et al. [13] found that the plastic deformation behaviors are different between α phase and β phase of TC2 alloy, as well as affect the hydrogen embrittlement. The SCC susceptibility of TC2 alloy is accelerated in simulated seawater due to the enhanced anodic dissolution and hydrogen embrittlement affected by deformations. Cao et al. [14] investigated the SCC properties of Ti-8Al-1Mo-1V alloy and proposed that α2 precipitates promote localized planar slip of SCC facets so that SCC accelerates. Dong et al. [15] investigated the stress corrosion cracking of TC4 ELI alloy with different microstructures in a 3.5% NaCl solution. The TC4 ELI alloy with lamellar microstructure exhibited significantly higher stress corrosion sensitivity than the alloy with equiaxed microstructure. The research work of Gao et al. [16] confirmed that the titanium alloy Ti6321 has stress corrosion sensitivity in a 3.5% NaCl solution. Junior et al. [17] investigated the SCC behavior of Ti-6Al-4V alloy at different aging stages and found that the SCC crack propagation was preferential through prior β grains. Ahn et al. [18] found that the change in cooling rate can lead to the random orientation of acicular α platelets, resulting in different SCC susceptibility.
The hexagonal close-packed structural α phase and body-centered cubic structural β phase exist together for titanium alloys. The contrasting crystal structure between α phase and β phase leads to different deformation behaviors under continuous stress loading [19]. Furthermore, it has been reported that the different volume fractions and morphology of α phase and β phase leads to different corrosion behaviors in titanium alloys [20]. Therefore, the α phase can greatly influence the SCC behaviors of titanium alloys. Some findings mentioned above [8,9,13,14,15,18] also indirectly confirmed this deduction. For metastable β titanium alloy, the precipitation of the secondary α phase is the main factor for alloy strengthening. The secondary α phase can be regulated by heat treatment and strongly affect the strength and ductility [21]. However, less attention has been paid to the effect of the secondary α phase on SCC behavior in metastable β titanium alloy. Ti-6Mo-5V-3Al-2Fe alloy is a novel metastable β titanium alloy, which shows potential for high strength [22]. However, limited available information on the relationship between the secondary α phase and SCC behavior exists at present. Therefore, this work is to investigate the effect of the secondary α phase on stress corrosion cracking of the Ti-6Mo-5V-3Al-2Fe alloy in a 3.5%NaCl solution. The objective of the present work is to improve the understanding of the SCC mechanism in the alloy, as well as to provide meaningful references for its marine engineering application.

2. Materials and Methods

The experimental materials are the Ti-6Mo-5V-3Al-2Fe (wt.%) alloy. The alloy ingot was produced by triple vacuum arc remelting. An alloy ingot with 120 mm in diameter and 200 mm in height was prepared. Then, β forging was carried out to obtain the alloy plates. The final forging temperature was 20 °C above β-transus (870 °C). The alloy plates with 90 mm in width and 32 mm in thickness were obtained. Subsequently, the as-forged alloys were heat treated to precipitate the secondary α phase. Two aging treatments were adopted to obtain the secondary α phase with large characteristic differences. According to the literature [22], the fine secondary α phase would precipitate during aging at the temperature range of 450–550 °C. When the aging temperature is larger than 600 °C, the secondary α phase would be significantly coarsened. Thus, the specimens were divided into two sets during heat treatment. One set was aged at a low temperature of 520 °C for 8 h, and another set was aged at a high temperature of 680 °C for 8 h. The purpose of adopting different aging temperatures was to obtain a secondary α phase with the difference in microstructural characteristics.
The chemical composition of the alloys before and after heat treatments were analyzed by ARL PERFORM X4200 X-ray fluorescence spectrometer. The microstructure was characterized by ZEISS GeminiSEM300 scanning electron microscope (SEM, Carl Zeiss AG, Oberkochen, Germany) and electron backscattered diffraction (EBSD, Oxford Instruments, Oxford, Britain). The specimens for SEM detection were cut from the aged alloy plates by electrical discharging machining and ground on metallographic sandpaper. Subsequently, the specimens were electropolished at~−15 °C in a solution of 6% perchloric acid, 35% n-butyl alcohol and 59% methanol.
The corrosion resistance of the aged alloy was gauged by potentiodynamic polarization and electrochemical impedance spectroscopy (EIS, Bio-Logic, Grenoble, France). Electrochemical measurements were performed at VSP-300 electrochemical experimental system in 3.5 wt.% NaCl aqueous solution. A conventional three-electrode cell was used in electrochemical tests. The specimen was used as the working electrode, a saturated calomel electrode was used as the reference electrode, and a carbon rod was used as the counter electrode. The open-circuit potential was measured at first. Potentiodynamic polarization curves were measured from the cathodic and anodic sides of corrosion potential, respectively. The scanning rate of the potentiodynamic polarization curves was 0.5 mV/s. EIS measurements were performed using a frequency range from 100 kHz and 10 mHz at a 10 mV amplitude over the open-circuit potential.
Slow strain rate testing (SSRT) of the aged alloys was carried out to investigate the SCC behaviors. SSRT was conducted at a WDML-10 slow strain rate stress corrosion testing machine with a strain rate of 1 × 10−6 s−1 in the air, as well as in 3.5% NaCl aqueous solution, respectively. Before tests, the specimens were immersed for 12 h to ensure a stable surface condition. Except for the standard distance area, the rest of the area was sealed with epoxy resin, installed in the corrosion solution tank and further sealed with silicone rubber. The fracture morphology of the SSRT specimens was observed by using an S-3400N scanning electron microscope (Hitachi, Tokyo, Japan).

3. Results

3.1. Chemical Composition and Microstructure

The chemical composition of the alloys before and after heat treatments are shown in Table 1. It can be seen that the chemical composition did not change significantly after the aging treatment. The microstructure of the aged alloy is shown in Figure 1. After aging treatment, a large amount of dispersed secondary α phase precipitate on the β matrix. When being aged at a relatively lower temperature of 520 °C, the precipitated secondary α phase is acicular with a high aspect ratio and has a large quantity (Figure 1a). When a relatively high aging temperature of 680 °C is adopted, the precipitated secondary α phase is coarsened and presents a rod-like shape with a low aspect ratio, and its quantity is reduced (Figure 1b). Similar aging response phenomena have been found in other metastable β titanium alloys, such as Ti-5Al-4Zr-8Mo-7V alloy [23], Ti-3.5Al-5Mo-6V-3Cr-2Sn-0.5Fe alloy [24], and Ti-5Al-3Mo-3V-2Cr-2Zr-1Nb-1Fe alloy [25]. The microstructures shown in Figure 1a and Figure 1b are referred to as “AM1” and “AM2” in the following parts, respectively. The morphology and volume fraction of the secondary α phase identified by the electron backscattered diffraction (EBSD) are listed in Table 2. It can be found that the secondary α phase volume fraction of the AM2 alloy is higher than that of the AM1 alloy. Additionally, the spacing of coarsened rod-like secondary α phase is significantly wider than that of the fine acicular secondary α phase.

3.2. Polarization Curves and EIS Analysis

The potentiodynamic polarization curves of the AM1 and AM2 alloys in a 3.5% NaCl aqueous solution are shown in Figure 2. According to Figure 2, the corrosion potential (Ecorr) and corrosion current density (Icorr) are identified by Tafel extrapolation and are summarized in Table 3. The Ecorr of the AM1 alloys exhibits a higher value of −356 mV, as compared to the Ecorr of −412 mV of the AM2 alloys. Moreover, the Icorr of the AM1 alloys is 0.445 μA/cm2, which is lower than that of the AM2 alloys with 1.73 μA/cm2. The relatively low current density values for corrosion demonstrate the relatively better corrosion resistance of the AM1 alloys compared to the AM2 alloy.
The Nyquist and Bode plots for the AM1 and AM2 alloys in a 3.5% NaCl aqueous solution are shown in Figure 3. According to Figure 3a, the capacitive loop radius for the AM2 alloys is smaller than that for the AM1 alloys. From Figure 3b, the impedance modulus |Z| at the frequency of 0.01 Hz for the AM2 alloys, which can reflect the characterization of the corrosion protection of the oxide film on the alloys, is much less than that for the AM1 alloys, although the two phase-angle plots are partially superimposed. All of these EIS results reflect that the compact property of the oxide film growing on the surface of the alloys is worse at AM2 alloys than that at AM1 alloys. According to the microstructural divergence between AM1 and AM2 alloys, the abovementioned difference in passivation and corrosion resistance may be related to the difference in microstructural characteristics of the secondary α phase.

3.3. SSRT Results

The stress–strain curves of SSRT, which were tested in air and in a 3.5% NaCl aqueous solution, are shown in Figure 4. It can be found that the strength and elongation of the specimens tested in a corrosive environment obviously decreased compared with the specimens tested without a corrosive environment.
In order to quantitatively evaluate the stress corrosive performance of the AM1 and AM2 alloys, an SCC susceptibility index (Iscc) is calculated according to the following formula [26].
Iscc = [(EairEsol)/Eair] × 100%
where Eair is the elongation of SSRT in air, and Esol is the elongation of SSRT in the corrosive environment. The SCC susceptibility index of the AM1 and AM2 alloys is 3.5% and 8.8%, respectively. The increased Iscc values clearly indicate that the SCC susceptibility of the AM2 alloys is more than that of the AM1 alloys. It can be speculated that the coarsened and widely spaced secondary α phase leads the alloy to be more susceptible to SCC.

3.4. Fracture Morphologies

The fracture morphologies of the AM1 alloys after SSRT are shown in Figure 5. The dimples on the fracture surface of the AM1 alloys tested in the air are deep and large, as shown in Figure 5a. For the AM1 alloys tested in 3.5% NaCl aqueous solution, the fracture morphology exhibits a mixed characteristic of ductile fracture and brittle fracture, as shown in Figure 5b. It is noteworthy that the dimples located in the ductile region present a shallow and small morphology under the influence of the corrosion environment, as shown in Figure 5c. The fracture morphologies of the AM2 alloys after SSRT tested in a 3.5% NaCl aqueous solution are shown in Figure 6. Similar to the AM1 alloys tested in a 3.5% NaCl aqueous solution, a mixed fracture characteristic of ductile and brittle can be observed in the AM2 alloys, as shown in Figure 6a. In addition to the shallow and small dimples, tear ridges (indicated by the white arrow) can be observed in the ductile region, as shown in Figure 6b. Moreover, a flat facet (circled by the dashed line) with undulating surfaces surrounded by dimples can be observed, as shown in Figure 6c. These special fracture morphologies provide threads for SCC mechanism analysis of the alloys.

4. Discussion

The precipitation behavior of the secondary α phase is dominated by the transformation driving force in β titanium alloy [27]. Lower aging temperature provided a higher undercooling degree as well as brought about a larger transformation driving force and higher nucleation rate. Moreover, lower aging temperature leads to a slow elements diffusion rate, resulting in a slow growth rate of the secondary α phase. Thus, the relatively lower aging temperature of 520 °C leads to fine acicular α precipitates. On the contrary, the larger driving force for the growth of the secondary α phase would be provided by a higher aging temperature. Meanwhile, the nucleation driving force could be reduced because of the lower undercooling. Then, coarsened secondary α phase precipitate under a relatively higher aging temperature of 680 °C. Furthermore, a region with enriched β-stabilizer forms in the vicinity accompanied by the precipitation of the secondary α phase due to the low tolerance of β-stabilizers. Such a region is so stable that hard to precipitate. Thus, coarsened precipitates lead to wider interspacing of the secondary α phase in AM2 alloys, compared with that in AM1 alloys.
Many research results [14,28] show that the SCC mechanism for titanium alloys is related to absorbed hydrogen. For the titanium alloys in 3.5% NaCl aqueous solution, two electrochemical actions take place simultaneously, i.e., the formation of an oxide film and the dissolution of the bare titanium alloy matrix [29]. The hydrogen for external absorption was produced by the chemical reaction that occurred on the alloy surface during the formation of the oxide film and the dissolution of bare metal. The occurred reactions during the repassivation of titanium alloys are shown in Equations (2)–(4), the reactions during the anodic dissolution are shown in Equations (5)–(7), and the production of hydrogen is shown in Equation (8) [30].
Ti → Ti3+ + 3e
Ti3+ + 2H2O → TiO2+ + 2H+ + e
Ti + 2H2O → TiO2 + 4H+ + 4e
Ti → Ti2+ + 2e
Ti2+ + H+ → [Ti3+ H]ads
[Ti3+ H]ads → Ti3+ + H+ + e
H+ + e → H
As shown in the area marked by the dotted box of polarization curves (Figure 2), the anodic current densities of both AM1 and AM2 alloys slowly increase with the increase in potential. It can be inferred that the slow dissolution of the metal occurs instead of maintaining passivation, even if the oxide films are grown on the metal surfaces. The same phenomenon also exists on the potentiodynamic polarization curves of Ti-6Al-4V alloy tested in 3.5 wt.% NaCl aqueous solution [30]. Therefore, titanium dissolution can be promoted. The anodic current density of the AM2 alloys increases faster than that of the AM1 alloys (the dotted box of Figure 2). Moreover, the corrosion potential of the AM2 alloys is more negative. Thus, the reactions during the anodic dissolution that occurred in the AM2 alloys would be more intense, resulting in a higher hydrogen concentration.
The fracture morphologies of the alloys tested in 3.5% NaCl aqueous solution present two typical characteristics:
(1) On the fracture surfaces of the alloys tested in 3.5% NaCl aqueous solution, shallow and small dimples could be observed (Figure 5c and Figure 6b). The formation of dimples with such characteristics has been clarified by an Absorption Induced Dislocation Emission (AIDE) mechanism [31]. It is proposed that intermetallic bond is weakened by absorbed hydrogen resulting in acceleration of dislocation emissions from crack tips [32]. The increasing dislocation density around the crack tip leads to a resharpening of the crack tip so that shallow and small dimples are formed.
(2) Tear ridges, as well as flat facets with undulating surfaces, could be observed on the fracture surfaces of the alloys tested in 3.5% NaCl aqueous solution (Figure 6b,c). It can be inferred that high-degree dislocation movement and localized plastic deformation occurred around cracks [14]. A Hydrogen Enhanced Localized Plasticity (HELP) mechanism can be used to explain such a phenomenon. The adsorbed hydrogen atom spread quickly to the front of crack tips through a continuous β phase, resulting in an increase in solute hydrogen concentration around crack tips [30]. The interaction between hydrogen and dislocation stimulates dislocation activation and then leads to localized plastic deformation [30].
In summary, the mixed fracture mode with shallow dimples and localized plasticity mainly contributed to the function of absorbed hydrogen, which is dominated by AIDE and HELP mechanisms. Moreover, these two mechanisms work together to trigger the SCC of the alloy in a 3.5% NaCl aqueous solution.
For metastable β titanium alloy during deformation, dislocations tend to activate within β matrix rather than within slender α precipitates. In addition, the interfaces between α phase and β matrix can act as an obstacle for dislocations slip because of a considerable lattice mismatch [33]. Then, A pile-up of dislocations can form in front of the interfaces between α phase and β matrix. The localized stress resulting from a pile-up of dislocations τ is as Equation (9) [34].
τ = sbT
where N is the number of accumulated dislocations, τs is the applied stress, and bT is the Burgers vector of the alloy.
The number of accumulated dislocations N is as Equation (10) [23].
N = π τ s ( 1 v T ) / G T b T · λ / 2
where vT is Poisson’s ratio of the alloy, GT is shear modulus of the alloy, λ/2 is the length of the pile-up zone of dislocations, and λ is the length of α phase spacing.
According to Equations (9) and (10), there is a positive correlation between τ, N, and λ. Thus, the increase in α phase spacing leads to the increase in accumulated dislocations quantity, which results in the increase in localized stress. Compared with the AM1 alloys, the AM2 alloys have wider secondary α phase spacing (Figure 1). The localized stress resulting from dislocations pile-up of the AM2 alloys is higher than that of the AM1 alloys. This stress concentration resulting from dislocation pile-up is also the initiation of cracks. Moreover, because of the action of absorbed hydrogen, the interatomic band is weakened, and the emission and motion of dislocations are promoted by AIDE and HELP mechanisms. Therefore, the AM2 alloys are more prone to crack initiation. In addition, worse corrosion resistance leads to a higher hydrogen concentration around the AM2 alloys. As a result, the AM2 alloys are more tend to occur SCC. Thus, the influence mechanism of the secondary α phase on SCC for the alloy in a 3.5% NaCl aqueous solution is proposed, shown in Figure 7. It can be found that compared with the coarsened rod-like secondary α with wide spacing, fine acicular secondary α with narrow spacing has a beneficial effect on the stress corrosion performance of the alloy.

5. Conclusions

In this work, the effect of the secondary α phase on stress corrosion cracking in 3.5% NaCl aqueous solution of a novel metastable β titanium alloy, Ti-6Mo-5V-3Al-2Fe, was investigated. The following conclusions can be drawn from this work;
(1) Compared with the alloy containing coarsened and widely spaced secondary α phase, the alloy containing fine and narrow spaced secondary α phase exhibits better corrosion resistance and less SCC susceptibility.
(2) The SCC mechanism of the alloy is a combination of Absorption Induced Dislocation Emission and Hydrogen Enhanced Localized Plasticity.
(3) Fine acicular secondary α phase with narrow spacing has a beneficial effect on stress corrosion performance. An influence mechanism of the secondary α phase on SCC for the alloy in a 3.5% NaCl aqueous solution is proposed.

Author Contributions

Conceptualization, H.Z.; methodology, C.W. and S.Z.; validation, X.Y. and Z.Z.; investigation, H.Z. and G.Z.; writing—original draft preparation, H.Z.; writing—review and editing, L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “National Natural Science Foundation of China, grant number U21A20117” and the “Technological Tacking Project of Liaoning Province, grant number 2021JH1/10400069”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shan, D.; Wang, L.; Fang, C.Q.; Shao, H.; Fan, D.; Wang, K.X.; Yan, G. The impact of aging time on fracture mode of a metastable β alloy with network-like microstructure. Eng. Fail. Anal. 2021, 129, 105731. [Google Scholar] [CrossRef]
  2. Souza, P.M.; Sivaswamy, G.; Andreu, A.; Rahimi, S. A novel cyclic thermal treatment for enhanced globularisation kinetics in Ti-6Al-4V alloy: Experimental, constitutive and FE based analyse. J. Alloys Compd. 2022, 898, 162859. [Google Scholar] [CrossRef]
  3. Prakash Kolli, R.; Devaraj, A. A review of metastable beta titanium alloys. Metals 2018, 8, 506. [Google Scholar] [CrossRef] [Green Version]
  4. Yan, S.K.; Song, G.L.; Li, Z.X.; Wang, H.N.; Zheng, D.J.; Cao, F.Y.; Horynova, M.; Dargusch, M.S.; Zhou, L. A state-of-the-art review on passivation and biofouling of Ti and its alloys in marine environments. J. Mater. Sci. Technol. 2018, 34, 421–435. [Google Scholar] [CrossRef]
  5. Liu, Q.; Li, N.; Shen, Z.X.; Zhao, M.F.; Xie, J.F.; Zhu, G.C.; Xu, X.; Yin, C.X. Calculation model and experimental study of the collapse strength of titanium alloy tubing and casing. Sci. Rep. 2022, 12, 4526. [Google Scholar] [CrossRef]
  6. Wake, H.; Takahashi, H.; Takimoto, T.; Takayanagi, H.; Ozawa, K.; Kadoi, H.; Okochi, M.; Matsunaga, T. Development of an electrochemical antifouling system for seawater cooling pipelines of power plants using titanium. Biotechnol. Bioeng. 2006, 95, 468–473. [Google Scholar] [CrossRef]
  7. Xu, Y.L.; Li, L.T.; Yin, Y.C.; Zhang, B.B.; Jiang, P. Stress corrosion cracking and fracture behaviors of gaseous hydrogenated titanium alloy Ti-6321 during slow strain rate tests. Mater. Res. Express 2021, 8, 046531. [Google Scholar] [CrossRef]
  8. Pazhanivel, B.; Sathiya, P.; Muthuraman, K.; Sozhan, G. Influence of NaCl environment on stress corrosion cracking of additive manufactured Ti-6Al-4V alloy. Eng. Failure Anal. 2021, 127, 105515. [Google Scholar] [CrossRef]
  9. Leon, A.; Levy, G.K.; Ron, T.; Shirizly, A.; Aghion, E. The effect of strain rate on stress corrosion performance of Ti6Al4V alloy produced by additive manufacturing process. J. Mater. Res. Technol. 2020, 9, 4097–4105. [Google Scholar] [CrossRef]
  10. Ciszak, C.; Popa, I.; Brossard, J.M.; Monceau, D.; Chevalier, S. NaCl induced corrosion of Ti-6Al-4V alloy at high temperature. Corros. Sci. 2016, 110, 91–104. [Google Scholar] [CrossRef]
  11. Atapour, M.; Pilchak, A.L.; Shamanian, M.; Fathi, M.H. Corrosion behavior of Ti-8Al-1Mo-1V alloy compared to Ti-6Al-4V. Mater. Des. 2011, 32, 1692–1696. [Google Scholar] [CrossRef]
  12. Zhang, S.H.; Deng, L.; Che, L.Z. An integrated model of rolling force for extra-thick plate by combining theoretical model and neural network model. J. Manuf. Processes. 2022, 75, 100–109. [Google Scholar] [CrossRef]
  13. Li, Y.; Pei, Z.B.; Zaman, B.; Zhang, Y.X.; Yuan, H.Z.; Cao, B. Effects of plastic deformations on the electrochemical and stress corrosion cracking behaviors of TC2 titanium alloy in simulated seawater. Mater. Res. Express 2018, 5, 116516. [Google Scholar] [CrossRef]
  14. Cao, S.; Samuel Lim, C.V.; Hinton, B.; Wu, X.H. Effects of microtexture and Ti3Al (α2) precipitates on stress-corrosion cracking properties of a Ti-8Al-1Mo-1V alloy. Corros. Sci. 2017, 116, 22–33. [Google Scholar] [CrossRef]
  15. Dong, Y.C.; Huang, S.; Wang, Y.Y.; Zhang, B.; Alexandrov, I.V.; Chang, H.; Dan, Z.H.; Ma, L.; Zhou, L. Stress corrosion cracking of TC4 ELI alloy with different microstructure in 3.5% NaCl solution. Mater. Charact. 2022, 194, 112357. [Google Scholar] [CrossRef]
  16. Gao, F.Y.; Sun, Z.J.; Yang, S.L.; Jiang, P.; Liao, Z.Q. Stress corrosion characteristics of electron beam welded titanium alloys joints in NaCl solution. Mater. Charact. 2022, 192, 112126. [Google Scholar] [CrossRef]
  17. Junior, P.C.P.S.; Correa, E.O. Effect of duplex aging heat treatment on the stress corrosion cracking behavior of Ti-6Al-4V α+β titanium alloy in methanol. Mater. Res. 2021, 24, e20200456. [Google Scholar] [CrossRef]
  18. Ahn, S.; Park, J.; Jeong, D.; Sung, H.; Kwon, Y.; Kim, S. Effect of cooling rate on SCC susceptibility of β-processed Ti–6Al–4V alloy in 0.6M NaCl solution. Met. Mater. Int. 2018, 24, 327–336. [Google Scholar] [CrossRef]
  19. Zhu, W.G.; Tan, C.S.; Xiao, R.Y.; Sun, Q.Y.; Sun, J. Slip behavior of Bi-modal structure in a metastable β titanium alloy during tensile deformation. J. Mater. Sci. Technol. 2020, 57, 188–196. [Google Scholar] [CrossRef]
  20. Ji, P.F.; Li, B.; Liu, S.G.; Zhang, X.; Chen, B.H.; Zhang, X.Y.; Ma, M.Z.; Liu, R.P. Controlling the corrosion behavior of Ti-Zr alloy by tuning the α/β phase volume fraction and morphology of β phase. J. Alloys. Compd. 2020, 825, 154153. [Google Scholar] [CrossRef]
  21. Wang, Y.L.; Hao, M.Y.; Li, D.; Li, P.; Liang, Q.L.; Wang, D.; Zheng, Y.F.; Sun, Q.Y.; Wang, Y.Z. Enhanced mechanical properties of Ti-5Al-5Mo-5V-3Cr-1Zr by bimodal lamellar precipitate microstructures via two-step aging. Mater. Sci. Eng. A 2022, 829, 142117. [Google Scholar] [CrossRef]
  22. Zhang, H.Y.; Li, X.H.; Lin, L.; Zhang, S.Q.; Wang, C.; Chen, L.J. Effect of aging temperature on microstructural evolution and mechanical properties of a novel β titanium alloy. Rare Met. Mater. Eng. 2019, 48, 3812–3818. [Google Scholar]
  23. Zhu, W.G.; Lei, J.; Zhang, Z.X.; Sun, Q.Y.; Chen, W.; Xiao, L.; Sun, J. Microstructural dependence of strength and ductility in a novel high strength β titanium alloy with Bi-modal structure. Mater. Sci. Eng. A 2019, 762, 138086. [Google Scholar] [CrossRef]
  24. Chen, Y.Y.; Du, Z.X.; Xiao, S.L.; Xu, L.J.; Tian, J. Effect of aging heat treatment on microstructure and tensile properties of a new β high strength titanium alloy. J. Alloys. Compd. 2014, 586, 588–592. [Google Scholar] [CrossRef]
  25. Ren, L.; Xiao, W.L.; Chang, H.; Zhao, Y.Q.; Ma, C.L.; Zhou, L. Microstructural tailoring and mechanical properties of a multi-alloyed near β titanium alloy Ti-5321 with various heat treatment. Mater. Sci. Eng. A 2018, 711, 553–561. [Google Scholar] [CrossRef]
  26. Tsao, L.C. Stress-corrosion cracking susceptibility of AZ31 alloy after varied heat treatment in 3.5 wt.% NaCl solution. Int. J. Mater. Res. 2010, 101, 1166–1171. [Google Scholar] [CrossRef]
  27. Zhang, X.; Kou, H.C.; Li, J.S.; Zhang, F.S.; Zhou, L. Evolution of the secondary α phase morphologies during isothermal heat treatment in Ti-7333 alloy. J. Alloys Compd. 2013, 577, 516–522. [Google Scholar] [CrossRef]
  28. Shih, D.S.; Robertson, I.M.; Birnbaum, H.K. Hydrogen embrittlement of α titanium: In situ TEM studies. Acta Mater. 1988, 36, 111–124. [Google Scholar] [CrossRef]
  29. Yu, S. Corrosion Resistance of Titanium Alloys. In ASM Handbook; Cramer, S.D., Covino, B.S., Eds.; ASM International: Geauga, OH, USA, 2003. [Google Scholar]
  30. Liu, R.; Cui, Y.; Liu, L.; Zhang, B.; Wang, F.H. A primary study of the effect of hydrostatic pressure on stress corrosion cracking of Ti-6Al-4V alloy in 3.5% NaCl solution. Corros. Sci. 2020, 165, 108402. [Google Scholar] [CrossRef]
  31. Lynch, S.P. Environmentally assisted cracking: Overview of evidence for an adsorption-induced localised-slip process. Acta Metall. 1988, 36, 2639–2661. [Google Scholar] [CrossRef]
  32. Cao, S.; Zhu, S.M.; Lim, C.V.S.; Zhou, X.G.; Chen, X.B.; Hinton, B.R.W.; Boyer, R.R.; Williams, J.C.; Wu, X. The mechanism of aqueous stress-corrosion cracking of α + β titanium alloys. Corros. Sci. 2017, 125, 29–39. [Google Scholar] [CrossRef]
  33. Zhang, H.Y.; Wang, C.; Zhou, G.; Zhang, S.Q.; Chen, L.J. Dependence of strength and ductility on secondary α phase in a novel metastable-β titanium alloy. J. Mater. Res. Technol. 2022, 18, 5257–5266. [Google Scholar] [CrossRef]
  34. Scudino, S.; Liu, G.; Sakaliyska, M.; Surreddi, K.B.; Eckert, J. Powder metallurgy of Al-based metal matrix composites reinforced with β-Al3Mg2 intermetallic particles: Analysis and modeling of mechanical properties. Acta Mater. 2009, 57, 4529–4538. [Google Scholar] [CrossRef]
Figure 1. SEM and EBSD images of the alloy: (a) SEM, aged at 520 °C for 8 h; (b) SEM, aged at 680 °C for 8 h; (c) EBSD, aged at 520 °C for 8 h; (d) EBSD, aged at 680 °C for 8 h.
Figure 1. SEM and EBSD images of the alloy: (a) SEM, aged at 520 °C for 8 h; (b) SEM, aged at 680 °C for 8 h; (c) EBSD, aged at 520 °C for 8 h; (d) EBSD, aged at 680 °C for 8 h.
Crystals 12 01849 g001
Figure 2. Potentiodynamic polarization curves.
Figure 2. Potentiodynamic polarization curves.
Crystals 12 01849 g002
Figure 3. (a) The Nyquist plots measurement; (b) Bode plots measurement of the AM1 and AM2 alloys.
Figure 3. (a) The Nyquist plots measurement; (b) Bode plots measurement of the AM1 and AM2 alloys.
Crystals 12 01849 g003
Figure 4. The stress–strain curves of SSRT, (a) AM1 alloys, (b) AM2 alloys.
Figure 4. The stress–strain curves of SSRT, (a) AM1 alloys, (b) AM2 alloys.
Crystals 12 01849 g004
Figure 5. Fracture morphologies after SSRT of the AM1 alloys: (a) the image for dimples of the specimen tested in air; (b) macro morphology of the specimen tested in 3.5% NaCl aqueous solution; (c) the image for dimples of the specimen tested in 3.5% NaCl aqueous solution.
Figure 5. Fracture morphologies after SSRT of the AM1 alloys: (a) the image for dimples of the specimen tested in air; (b) macro morphology of the specimen tested in 3.5% NaCl aqueous solution; (c) the image for dimples of the specimen tested in 3.5% NaCl aqueous solution.
Crystals 12 01849 g005
Figure 6. Fracture morphology after SSRT of the AM2 alloy tested in 3.5% NaCl aqueous solution: (a) macro morphology; (b) the image of tear ridges; (c) the image of flat facets.
Figure 6. Fracture morphology after SSRT of the AM2 alloy tested in 3.5% NaCl aqueous solution: (a) macro morphology; (b) the image of tear ridges; (c) the image of flat facets.
Crystals 12 01849 g006
Figure 7. Schematic diagram of effect of secondary α phase on SCC for Ti-6Mo-5V-3Al-2Fe alloy in 3.5% NaCl aqueous solution: (a) AM1 alloy; (b) AM2 alloy.
Figure 7. Schematic diagram of effect of secondary α phase on SCC for Ti-6Mo-5V-3Al-2Fe alloy in 3.5% NaCl aqueous solution: (a) AM1 alloy; (b) AM2 alloy.
Crystals 12 01849 g007
Table 1. Chemical composition of the alloys before and after heat treatments.
Table 1. Chemical composition of the alloys before and after heat treatments.
Heat TreatmentMo (wt.%)V (wt.%)Al (wt.%)Fe (wt.%)Ti (wt.%)
Before heat treatment6.245.192.982.14Bal.
aged at 520 °C for 8 h6.135.122.992.22Bal.
aged at 680 °C for 8 h6.225.152.982.19Bal.
Table 2. Microstructural characteristic of α phase in the aged alloys.
Table 2. Microstructural characteristic of α phase in the aged alloys.
No.Aging Treatment ProcessCharacteristic of α PhaseVolume Fraction of α Phase/%
AM1520 °C for 8 hFine and narrow spaced33.2
AM2680 °C for 8 hCoarsened and widely spaced38.7
Table 3. The corrosion potential, corrosion current density and passive current density.
Table 3. The corrosion potential, corrosion current density and passive current density.
AlloysEcorr (mV)Icorr (μA/cm2)
AM1−3560.445
AM2−4121.730
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, H.; Wang, C.; Zhang, S.; Yu, X.; Zhou, G.; Zhang, Z.; Chen, L. Effect of Secondary α Phase on Stress Corrosion Cracking of a Novel Metastable β Titanium Alloy in 3.5% NaCl Solution. Crystals 2022, 12, 1849. https://doi.org/10.3390/cryst12121849

AMA Style

Zhang H, Wang C, Zhang S, Yu X, Zhou G, Zhang Z, Chen L. Effect of Secondary α Phase on Stress Corrosion Cracking of a Novel Metastable β Titanium Alloy in 3.5% NaCl Solution. Crystals. 2022; 12(12):1849. https://doi.org/10.3390/cryst12121849

Chicago/Turabian Style

Zhang, Haoyu, Chuan Wang, Shuai Zhang, Xiaoling Yu, Ge Zhou, Zhiqiang Zhang, and Lijia Chen. 2022. "Effect of Secondary α Phase on Stress Corrosion Cracking of a Novel Metastable β Titanium Alloy in 3.5% NaCl Solution" Crystals 12, no. 12: 1849. https://doi.org/10.3390/cryst12121849

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop