Next Article in Journal
Influence of Minor Alloying Element Additions on the Crack Susceptibility of a Nickel Based Superalloy Manufactured by LPBF
Previous Article in Journal
Pilot Demonstration of Hot Sheet Metal Forming Using 3D Printed Dies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Thermo-Mechanical Treatment on the Microstructure and Tensile Properties of the Fe-22Cr-5Al-0.1Y Alloy

1
Advanced Technology & Materials Limited Company, China Iron & Steel Research Institute Group, Area C, No. 10 Yongcheng North Road, Haidian District, Beijing 100081, China
2
State Key Laboratory of Advanced Processing and Recycling of Nonferrous Metals, Lanzhou University of Technology, Lanzhou 730050, China
3
School of Materials Science and Engineering, Lanzhou University of Technology, Lanzhou 730050, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(19), 5696; https://doi.org/10.3390/ma14195696
Submission received: 6 September 2021 / Revised: 26 September 2021 / Accepted: 28 September 2021 / Published: 30 September 2021

Abstract

:
Oxide dispersion strengthened ferritic steel is considered an important structural material in fusion reactors due to its excellent resistance to radiation and oxidation. Fine and dispersed oxides can be introduced into the matrix via the powder metallurgy process. In the present study, large grain sizes and prior particle boundaries (PPBs) formed in the FeCrAlY alloy prepared via powder metallurgy. Thermo-mechanical treatment was conducted on the FeCrAlY alloy. Results showed that microstructure was optimized: the average grain diameter decreased, the PPBs disappeared, and the distribution of oxides dispersed. Both ultimate tensile strength and elongation improved, especially the average elongation increased from 0.5% to 23%.

1. Introduction

Oxide dispersion strengthened (ODS) alloys are candidate structural materials for the nuclear power industry owing to their excellent performance at high temperature and corrosive environments [1,2,3,4]. The excellent abilities to withstand high temperatures and resistance to oxidation were obtained by the addition of rare earth elements and thermal stable oxides [5,6,7,8], such as Y2O3 [9] and Al2O3 [10]. The thermodynamically stable oxide particles dispersed in the matrix can improve the strength, creep, and fatigue resistance via hindering the movement of dislocations and grain boundaries [2,9,11,12,13,14].
The Ti addition was expected to react with Y and O to form Y-Ti-O oxides within alloys (Y2Ti2O7, Y2TiO5) [15]. The addition of Al in the ODS steel was not only expected to form Y-Al-O oxides [16,17,18,19,20] but also to improve the corrosion resistance by forming dense Al2O3 scales on the surface of the alloy, which limited further corrosion of the matrix under the oxide layer. In addition, the presence of Al can suppress embrittlement at 475 °C by suppressing the phase separation of the ferritic phase into the two phases, Fe-rich ferrite and Cr-rich ferrite in high Cr steel [21,22,23]. However, Y-Al-O oxides tended to coarsen and decrease the strength of the alloy at high temperatures [15,24,25]. The addition of Zr can improve it by forming more smaller and more stable Y-Zr-O (Y4Zr3O12, Y6ZrO11, Y2Zr2O7) oxides [26].
Traditionally, the introduction of dispersed oxides mainly relies on the mechanical alloying (MA) process [27,28]. Therefore, the MA parameters played a vital role in the process of preparing ODS alloys [27,29,30]. However, the extended delivery time of the MA process as well as the impurities introduced during the ball milling procedure present a challenge to the industrial production of large-sized structural parts [31,32,33]. Consequently, an alternative powder metallurgy route without the MA process is essential for the production of ODS steel.
Gas atomization reaction synthesis (GARS) [34] and surface treatment of gas atomized powder followed by reactive synthesis (STARS) [33] are two optional processing routes. During the GARS and STARS processes, the metastable Fe- and Cr-rich oxide scale is formed on the surface of gas atomized powders. Additionally, the metastable oxides are transformed into stable and dispersed oxides during hot isostatic pressing (HIP) [32,34,35]. Without the MA process, the thermally stable alumina oxides on the surface of gas atomized powders formed prior particle boundaries (PPBs) that resulted in a coarse grain size during the HIP process [33] and were detrimental to the mechanical properties of ODS alloys. A suitable thermo-mechanical treatment (TMT) process can eliminate the effect of PPBs via refining the grain size and optimizing the distribution of the precipitates [36,37,38].
In the current study, the Fe-22Cr-5Al-0.15Y alloy was prepared through argon-gas-atomization and HIP without the MA process. Afterward, the TMT process was conducted on the as-HIPed alloy. Moreover, the microstructure, tensile test, and fracture surfaces of the as-HIPed and as-TMTed alloys were characterized and analyzed in detail. This was performed to examine the effect of the TMT process on the microstructure and properties of the FeCrAlY alloy.

2. Materials and Methods

2.1. Materials

Fe-22Cr-5Al-0.15Y powders were prepared through gas atomization reactive synthesis (GARS). The powders ranged from 50 μm to 350 μm. As shown in Figure 1a, the height of the powder can was 120 mm, its inner diameter was 60 mm, and the wall thickness was 2.5 mm. The vacuum within the can was 10−2 Pa before HIP. Subsequently, the powders were consolidated by HIP at 1120 °C under the pressure of 120 MPa for 2 h. The chemical composition of the as-HIPed alloy is shown in Table 1. After HIP, the as-HIPed alloy, with a diameter of about 50 mm, underwent a TMT, as shown in Figure 1b. The as-HIPed ingot was hot-rolled into plates with a thickness of 11 mm by multi-pass rolling. The ingot was hot-rolled at an initial rolling temperature of 1050 °C and a final rolling temperature of 800 °C, followed by annealing at 1000 °C for 2 h, then air-cooled to room temperature.

2.2. Microstructure Characterization

The microstructure and precipitates in the FeCrAlY alloys were characterized via scanning electron microscopy (SEM, Quanta 450FEG, FEI, Hillsboro, OR, USA) equipped with an energy dispersive X-ray spectroscopy (EDS) (Oxford, London, UK) and transmission electron microscope (TEM, JEM-2100F, JEOL, Tokyo, Japan). Samples investigated by SEM were etched electrolytically with a solution of 5% HClO4 + 95% alcohol at a voltage of 22 V for 30–40 s. On the other hand, samples observed through TEM were prepared using the TenuPol-5 double jet electro-polisher (Struers, Ballerup, Danmark) at −10 °C with a solution of 10% HClO4 + 90% alcohol.

2.3. Tensile Test

The tensile tests were carried out at 25 °C and at a strain rate of 0.002 s−1. As shown in Figure 2, the samples with a gauge length of 15 mm and a gauge diameter of 3 mm were used for the testing on an AGS-XS300 machine (Shimadzu, Kyoto, Japan). Three samples of the as-HIPed alloy and the as-TMTed alloy were tested, respectively. Moreover, the fracture surfaces of the broken samples were investigated through SEM to reveal the fracture mode and the influence of precipitates on their mechanical properties.

3. Results

3.1. Microstructure

Figure 3 shows the secondary electron microstructure images of the as-HIPed alloys at different magnifications along with the grain diameter statistical histogram. Figure 3a shows the macro-microstructure view of the as-HIPed alloy with a mean grain diameter of about 127 μm (as shown in Figure 3h). The grain size of the alloy was calculated from the average lengths of the long axes and the short axes of the grains. Approximately five hundred grains were calculated. Powder prior boundaries were found along the grain boundaries in Figure 3b,c. EDS maps (Figure 3d–g) revealed that the stripe precipitates mainly were Al-containing oxides, Fe, and Cr elements distributed evenly within the grains. In addition, oxides particles were found within the grains.
TEM was used to analyze the oxides. Figure 4 shows the TEM images and EDS spectra of the precipitates in the as-HIPed sample as well as the diameter statistical histogram. Continuous and large-sized precipitates, with a mean size of about 121 nm, were found along the grain boundary and within the grain as shown in Figure 4a,c. The continuous precipitates along the grain boundary were Al-oxides (Figure 3 and Figure 4a,b). The large-sized precipitates within the grain boundary had a size of about 500 nm and were enriched in Al and O. Notably, these small oxides within the grains (with a size of about 50~100 nm) were Y-rich, as shown in Figure 4c,d.
After the HIP, the TMT process was conducted on the alloy to optimize the microstructure and eliminate the adverse effect of the continuous Al-containing precipitates. The results of microstructure analysis for the as-TMTed sample were shown in Figure 5 and Figure 6.
Figure 5a,b shows the macro and micro view of the microstructure of the as-TMTed alloy. The mean grain diameter of the as-TMPed alloy was about 22 μm (Figure 5c). Precipitates were dispersed and no continuous precipitates were found at the grain boundaries.
Figure 6a,c shows that there were dispersed precipitates within the grains and at the grain boundaries, with an average size of about 112 nm (as shown in Figure 6e). The EDS results of precipitates (pointed by the red arrows in Figure 6a,c) are shown in Figure 6b,d, respectively. Not only aluminum oxide particles (P1), but also Y-Al-O particles (P2,P3,P4) were found at the boundaries. Notably, the large and continuous aluminum oxide disappeared, which correspond to Figure 5a,b. In addition, extensive dislocations (pointed by yellow arrows) were observed around the grain boundaries as shown in Figure 6c.

3.2. Tensile Properties and Fracture Surface

Figure 7 shows the yield strength (YS), ultimate tensile strength (UTS), and elongation of the as-HIPed and as-TMTed samples tested at 25 °C. The average UTS of the as-HIPed and as-TMTed alloys was 522 MPa and 658 MPa, respectively. The average YS of the as-HIPed and as-TMTed alloys were 515 MPa and 442 MPa. In addition, the average fracture elongation of the as-HIPed and as-TMTed were 0.5% and 23%. The error bars of YS and UTS in the as-TMTed alloy were smaller than those of the as-HIPed alloys as shown in Figure 7.
Afterward, the SEM was used to examine the fracture surfaces of the two tensile samples to understand the effect of precipitates on the tensile deformation behavior of the as-HIPed and as-TMTed alloys with different distribution of precipitates at room temperature. The fractography of the as-HIPed and as-TMTed alloys after the tensile test are shown in Figure 8 and Figure 9.
Figure 8 shows the fractography of the as-HIPed alloy. The macro-morphology of the as-HIPed alloy was shown in Figure 8a. The fracture surface was mainly an intergranular fracture, with the partial area having a cleavage fracture (marked by a yellow dotted line in Figure 8a). The tensile test revealed that there was a brittle fracture in the as-HIPed alloy. Figure 8b is the magnified image of the intergranular fracture area marked by a red dashed frame in Figure 8a. Moreover, Figure 8c shows that there existed precipitates on the surface of the intergranular fracture in the as-HIPed sample, which was the main reason for the occurrence of the partial intergranular fracture in the as-HIPed alloy.
Additionally, the macro-morphology of the fracture for the as-TMTed sample in Figure 9a shows that there was an obvious necking phenomenon and ductile fracture. Moreover, as the micro-morphology of the fracture shown in Figure 9b, there were regular and small-sized dimples on the fracture surface, which was consistent with the high strength and elongation of the as-TMTed alloy at room temperature.

4. Discussion

4.1. Formation of Al-Containing Precipitates

During the preparation process of the FeCrAl alloy powders, the atomized powders inevitably came into contact with air and were subsequently oxidized [39]. As a result, Fe- and Cr-metastable oxides would form on the surface [31,32,33,35]. Afterward, these metastable oxides would decompose during the HIP process, the oxygen from the decomposition would diffuse into the grains and then react with Y and Al to form more stable oxides [31,32,33,34,35].
In this study, the Al content in the pre-alloy was higher than that in other types of Fe-based ODS steel [40]. Thus, the Al-rich oxides easily formed on the surface of the powders [41,42]. In addition, the Al-containing oxides can exist stably [10] and formed the large and continuous Al-containing oxides along the grain boundaries, as shown in Figure 3a–d. The continuous precipitates formed at the grain boundaries during HIP, the so-called PPBs, were very common among the powder metallurgy (PM) alloys [43,44]. Similarly, PPBs were found in the ODS 12Cr-2W prepared by E. Gil [31]. The presence of PPBs can hinder the element diffuse and reduce bonding strength between powders, so as to form the continuous stripe of Al2O3 at the grain boundaries [24,25,45,46].
As E. Gil reported, PPBs can be dissolved by enhancing HIP temperature [31] so the adverse effect of PPBs can be eliminated. However, the melting point of Al2O3 is higher than that of FeCrAl, so that it is not wise to eliminate Al2O3-PPBs via enhancing HIP temperature. Thus, we attempted to eliminate the PPBs by the TMT process.

4.2. Effect of TMT Process on Microstructure

After the TMT, the microstructure was optimized, as Figure 3, Figure 4, Figure 5 and Figure 6 show. The average grain diameter decreased from 127 μm to 22 μm, and the distribution of Al-containing oxide dispersed. Normally, equiaxed grains are elongated into ellipse along the rolling direction, and the subsequent heat treatment process activates the recovery and recrystallization [47,48,49]. As can be seen in Figure 5, the as-TMTed alloy exhibited early equiaxed, which indicate that full recrystallization occurred for the as-TMTed alloy. Nucleation of the new grains replaced the original deformed grains, which led to grain refinement [48,49]. PPBs were eliminated during the TMT process, and there are no PPBs along the grain boundaries. The most likely explanation was that the stripe Al2O3 along the grain boundary was broken and divided into oxide particles during the TMT process.
Notably, the precipitates at the grain boundaries were mainly Y-Al-O ternary oxides of as-TMTed alloys compared with that of the as-HIPed alloys. This indicated that the TMT may promote the formation of Y-Al-O oxides. The possible reaction mechanisms were as follows. During the TMT process, Al2O3 reacted with Y2O3 or the Y element solid solution in the matrix to form the Y-Al-O particles (YAlO3, Y4Al2O9, Y3Al5O12) through reactions Equations (1)–(3) [18]:
Y2O3 + Al2O3 → 2YAlO3
3Y2O3 + 5Al2O3 → 2Y3Al5O12
3Y2O3 + Al2O3 → Y4Al2O9
In addition, these Y-Al-O oxides were smaller than Al2O3 oxides [10,50] and more useful in improving the mechanical properties of alloys. These Y-Al-O oxides were very common among the commercial ODS alloys, such as PM2000 [51,52,53].

4.3. Possible Strengthening Mechanism

Strength-ductility trade-off is a typical phenomenon for high-strength materials [54,55]. The strength of the steel was increased by introducing fine and dispersed particles into the matrix. This was accompanied by a sacrifice of ductility. The particles can generate stress concentration at the grain boundaries and tend to localize and crack near those particles [56]. In our study, the stripe Al2O3 oxides [57] at the grain boundaries of the as-HIPed alloys not only worsened the elongation deeply, leading to the partial intergranular fracture, but also deteriorated the strength.
TMT can improve plasticity effectively while maintaining the strength of the alloy by optimizing the microstructure [33,58,59]. Here, before and after TMT, the average diameter of the oxides was similar. The distribution of oxides became more dispersed, and the huge PPBs disappeared, which contributed to the improvement of the alloys. The disappearance of continuous Al2O3, which contributed to decreasing the stress concentration at the grain boundaries, and the formation of Y-Al-O within the grains, which contributes to blocking and storing of dislocation [56], contributed to delay the cracks that occurred at the grain boundaries and the alloy underwent more deformation before the fracture. During the tensile deformation, a large number of dislocations generated within the grains, the existence of Y-Al-O oxides can pin the dislocations, which limit the movement of dislocations; thus, a large number of dislocations were stored within the grains. So, it is not difficult to understand the elongation of the as-TMTed alloy improved significantly.
Optimizing the distribution of second phase particles to improve the ductility of an alloy is an effective strategy. This strategy was applied to nanostructured alloys [56], such as molybdenum alloys, and age-hardenable alloys such as UFG Al alloys [60,61,62]. The refinement of the grain size is another important method to improve elongation. The number of the grain boundaries increases and the grain size decreases, which increases the difficulty of crack propagation within alloys because more energy should be consumed.
Refining the grain size can improve the ductility and the strength of the alloy at the same time [58,63]. According to the Hall–Petch relation [59,64,65], which can be written as follows (4):
σy = σ0 + kd−1/2
where k is a material constant, as well as normal stress σ0, and d represents the grain size. Obviously, alloys with smaller grain sizes should possess higher yield strength according to this Hall–Petch relation. However, only UTS was improved while the YS decreased after TMT, which was beyond the expectation. The YS of the alloy is affected by many factors, not just the grain size. The following Equation (5) can be used to describe the influence of each factor on YS quantitatively [66,67]:
σ y = σ 0 + σ ss + σ gb + ( σ dis 2 + σ p 2 ) 0
where σy is the yield strength of ODS steel, σ0 is the friction stress of pure iron, σss and σgb is the solid solution strengthening and grain boundary strengthening, σdis and σp represent the dislocation strengthening and oxide dispersion strengthening, respectively. After TMT, the grain size becomes smaller, more dislocations were introduced into the matrix and the precipitates became more dispersed, so contributions of the grain boundary strengthening, dislocation strengthening and oxide dispersion strengthening to the strength is positive [7,50,65,68,69,70]. Contributions of solution strengthening to the YS decreased. The possible reasons are as follows: During the gas atomization pulverization stage, due to the rapid cooling of the metal droplets to room temperature, the solid solution strengthening elements such as Cr and Al failed to precipitate from the Fe crystal lattice in time to form a supersaturated solid solution [34,35]. In the subsequent hot isostatic pressing stage and thermal deformation stage, the supersaturated Cr, Al, and other elements can be precipitated, which decreased the lattice distortion, and thus, decreased the σss [71]. The formation of Y-Al-O oxides within the grains would consume part of the Al element solid in the matrix, which can decrease the σss to a certain extent.

5. Conclusions

In this study, Fe-22Cr-5Al-0.1Y alloys were produced through hot isostatic pressing, then a thermal-mechanical treatment was applied. In addition, the relationship between the microstructure and tensile property of the as-HIPed and as-TMTed FeCrAlY alloys was investigated. Consequently, the following conclusions were drawn:
  • Both continuous and dispersed Al2O3 particles formed at the grain boundaries within as-HIPed alloys, which led to low strength and intergranular fracture of the as-HIPed alloy at 25 °C;
  • Microstructure of the as-HIPed alloy was optimized after TMT. The mean grain diameter decreased from 127 μm to 22 μm. In addition, large and continuous Al2O3 particles were refined, and the distribution of the oxide became dispersed. Part of the Al2O3 reacted with Y2O3 to form more stable Y-Al-O oxides at the grain boundaries and within the grains;
  • The optimization of the microstructure promoted the improvement in mechanical properties. The UTS increased from 522 MPa to 658 MPa, and the elongation increased from 0.5% to 23%. The fracture of alloy transformed from a cleavage/intergranular fracture into a ductile fracture.

Author Contributions

Conceptualization, Y.Y.; methodology, Y.Y. and R.C.; software, Y.C.; formal analysis, Y.Y.; investigation, H.C. and W.Q.; resources, W.Q.; data curation, H.C. and Y.C. and R.C.; writing-original draft preparation, H.C., Y.Z. and Y.Y.; writing-review and editing, Y.Z.; project administration, T.W.; funding acquisition, H.C. and T.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (51961024, 52071170, 51761027).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also forms part of an ongoing study.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The authors declare that conflicts of interest do not exist.

References

  1. Odette, G.R.; Alinger, M.J.; Wirth, B.D. Recent Developments in Irradiation-Resistant Steels. Annu. Rev. Mater. Res. 2008, 38, 471–503. [Google Scholar] [CrossRef]
  2. Hu, Y.; Zhao, Y.; Peng, Y.; Yang, W.; Ma, X.; Wang, B. High-strength joint of nuclear-grade FeCrAl alloys achieved by friction stir welding and its strengthening mechanism. J. Manuf. Process. 2021, 65, 1–11. [Google Scholar] [CrossRef]
  3. Engkvist, J.; Canovic, S.; Hellström, K.; Järdnäs, A.; Svensson, J.E.; Johansson, L.G.; Olsson, M.; Halvarsson, M. Alumina Scale Formation on a Powder Metallurgical FeCrAl Alloy (Kanthal APMT) at 900–1,100 °C in Dry O2 and in O2 + H2O. Oxid. Met. 2009, 73, 233–253. [Google Scholar] [CrossRef]
  4. Aydogan, E.; El-Atwani, O.; Takajo, S.; Vogel, S.C.; Maloy, S.A. High temperature microstructural stability and recrystallization mechanisms in 14YWT alloys. Acta Mater. 2018, 148, 467–481. [Google Scholar] [CrossRef]
  5. Zhang, L.; Yu, L.; Liu, Y.; Liu, C.; Li, H.; Wu, J. Influence of Zr addition on the microstructures and mechanical properties of 14Cr ODS steels. Mater. Sci. Eng. A 2017, 695, 66–73. [Google Scholar] [CrossRef]
  6. Siska, F.; Stratil, L.; Hadraba, H.; Fintova, S.; Kubena, I.; Hornik, V.; Husak, R.; Bartkova, D.; Zalezak, T. Strengthening mechanisms of different oxide particles in 9Cr ODS steel at high temperatures. Mater. Sci. Eng. A 2018, 732, 112–119. [Google Scholar] [CrossRef]
  7. Shen, J.; Li, Y.; Li, F.; Yang, H.; Zhao, Z.; Kano, S.; Matsukawa, Y.; Satoh, Y.; Abe, H. Microstructural characterization and strengthening mechanisms of a 12Cr-ODS steel. Mater. Sci. Eng. A 2016, 673, 624–632. [Google Scholar] [CrossRef]
  8. Kumar, D.; Prakash, U.; Dabhade, V.V.; Laha, K.; Sakthivel, T. High yttria ferritic ODS steels through powder forging. J. Nucl. Mater. 2017, 488, 75–82. [Google Scholar] [CrossRef]
  9. Li, Z.; Lu, Z.; Xie, R.; Lu, C.; Shi, Y.; Liu, C. Effects of Y2O3, La2O3 and CeO2 additions on microstructure and mechanical properties of 14Cr-ODS ferrite alloys produced by spark plasma sintering. Fusion Eng. Des. 2017, 121, 159–166. [Google Scholar] [CrossRef]
  10. Zhao, Q.; Yu, L.; Liu, Y.; Huang, Y.; Guo, Q.; Li, H.; Wu, J. Evolution of Al-containing phases in ODS steel by hot pressing and annealing. Powder Technol. 2017, 311, 449–455. [Google Scholar] [CrossRef]
  11. Zhao, Q.; Yu, L.; Liu, Y.; Huang, Y.; Ma, Z.; Li, H.; Wu, J. Microstructure and tensile properties of a 14Cr ODS ferritic steel. Mater. Sci. Eng. A 2017, 680, 347–350. [Google Scholar] [CrossRef]
  12. Chauhan, A.; Walter, M.; Aktaa, J. Towards improved ODS steels: A comparative high-temperature low-cycle fatigue study. Fatigue Fract. Eng. Mater. Struct. 2017, 40, 2128–2140. [Google Scholar] [CrossRef]
  13. Ukai, S.; Ohtsuka, S. Low cycle fatigue properties of ODS ferritic–martensitic steels at high temperature. J. Nucl. Mater. 2007, 367–370, 234–238. [Google Scholar] [CrossRef]
  14. Chauhan, A.; Hoffmann, J.; Litvinov, D.; Aktaa, J. High-temperature low-cycle fatigue behavior of a 9Cr-ODS steel: Part 1—Pure fatigue, microstructure evolution and damage characteristics. Mater. Sci. Eng. A 2017, 707, 207–220. [Google Scholar] [CrossRef]
  15. Xu, H.; Lu, Z.; Wang, D.; Zhang, Z.; Han, Y.; Liu, C. Structure and composition of oxides in FeCrAl ODS alloy with Zr addition. Mater. Sci. Technol. 2017, 33, 1790–1795. [Google Scholar] [CrossRef]
  16. Xu, S.; Zhou, Z.; Long, F.; Jia, H.; Guo, N.; Yao, Z.; Daymond, M.R. Combination of back stress strengthening and Orowan strengthening in bimodal structured Fe–9Cr–Al ODS steel with high Al addition. Mater. Sci. Eng. A 2019, 739, 45–52. [Google Scholar] [CrossRef]
  17. Xu, S.; Zhou, Z.; Jia, H.; Yao, Z. Microstructure Characterization and Mechanical Properties of Al Alloyed 9Cr ODS Steels with Different Al Contents. Steel Res. Int. 2019, 90, 1800594. [Google Scholar] [CrossRef]
  18. Zhang, G.; Zhou, Z.; Mo, K.; Wang, P.; Miao, Y.; Li, S.; Wang, M.; Liu, X.; Gong, M.; Almer, J.; et al. The microstructure and mechanical properties of Al-containing 9Cr ODS ferritic alloy. J. Alloys Compd. 2015, 648, 223–228. [Google Scholar] [CrossRef] [Green Version]
  19. Mohan, S.; Kaur, G.; Panigrahi, B.K.; David, C.; Amarendra, G. Effect of Zr and Al addition on nanocluster formation in oxide dispersion strengthened steel—An ab initio study. J. Alloys Compd. 2018, 767, 122–130. [Google Scholar] [CrossRef]
  20. Zhao, Q.; Yu, L.-M.; Liu, Y.-C.; Huang, Y.; Ma, Z.-Q.; Li, H.-J. Effects of aluminum and titanium on the microstructure of ODS steels fabricated by hot pressing. Int. J. Miner. Metall. Mater. 2018, 25, 1156–1165. [Google Scholar] [CrossRef]
  21. Kobayashi, S.; Takasugi, T. Mapping of 475 °C embrittlement in ferritic Fe–Cr–Al alloys. Scr. Mater. 2010, 63, 1104–1107. [Google Scholar] [CrossRef]
  22. Li, W.; Lu, S.; Hu, Q.-M.; Mao, H.; Johansson, B.; Vitos, L. The effect of Al on the 475 °C embrittlement of Fe–Cr alloys. Comput. Mater. Sci. 2013, 74, 101–106. [Google Scholar] [CrossRef]
  23. Száraz, Z.; Hähner, P.; Stráská, J.; Ripplinger, S. Effect of phase separation on tensile and Charpy impact properties of MA956 ODS steel. Mater. Sci. Eng. A 2017, 700, 425–437. [Google Scholar] [CrossRef]
  24. Kimura, A.; Kasada, R.; Iwata, N.; Kishimoto, H.; Zhang, C.H.; Isselin, J.; Dou, P.; Lee, J.H.; Muthukumar, N.; Okuda, T.; et al. Development of Al added high-Cr ODS steels for fuel cladding of next generation nuclear systems. J. Nucl. Mater. 2011, 417, 176–179. [Google Scholar] [CrossRef]
  25. Xia, Y.P.; Wang, X.P.; Zhuang, Z.; Sun, Q.X.; Zhang, T.; Fang, Q.F.; Hao, T.; Liu, C.S. Microstructure and oxidation properties of 16Cr–5Al–ODS steel prepared by sol–gel and spark plasma sintering methods. J. Nucl. Mater. 2013, 432, 198–204. [Google Scholar] [CrossRef]
  26. Gao, R.; Zhang, T.; Wang, X.P.; Fang, Q.F.; Liu, C.S. Effect of zirconium addition on the microstructure and mechanical properties of ODS ferritic steels containing aluminum. J. Nucl. Mater. 2014, 444, 462–468. [Google Scholar] [CrossRef]
  27. Rahmanifard, R.; Farhangi, H.; Novinrooz, A.J. Optimization of mechanical alloying parameters in 12YWT ferritic steel nanocomposite. Mater. Sci. Eng. A 2010, 527, 6853–6857. [Google Scholar] [CrossRef]
  28. Williams, C.A.; Unifantowicz, P.; Baluc, N.; Smith, G.D.W.; Marquis, E.A. The formation and evolution of oxide particles in oxide-dispersion-strengthened ferritic steels during processing. Acta Mater. 2013, 61, 2219–2235. [Google Scholar] [CrossRef]
  29. Waghmode, S.P.; Dabade, U.A. Optimization of process parameters during turning of Inconel 625. Mater. Today Proc. 2019, 19, 823–826. [Google Scholar] [CrossRef]
  30. Alinger, M.J.; Odette, G.R.; Hoelzer, D.T. On the role of alloy composition and processing parameters in nanocluster formation and dispersion strengthening in nanostuctured ferritic alloys. Acta Mater. 2009, 57, 392–406. [Google Scholar] [CrossRef]
  31. Gil, E.; Ordás, N.; García-Rosales, C.; Iturriza, I. ODS ferritic steels produced by an alternative route (STARS): Microstructural characterisation after atomisation, HIPping and heat treatments. Powder Metall. 2016, 59, 359–369. [Google Scholar] [CrossRef]
  32. Ordás, N.; Gil, E.; Cintins, A.; de Castro, V.; Leguey, T.; Iturriza, I.; Purans, J.; Anspoks, A.; Kuzmin, A.; Kalinko, A. The role of yttrium and titanium during the development of ODS ferritic steels obtained through the STARS route: TEM and XAS study. J. Nucl. Mater. 2018, 504, 8–22. [Google Scholar] [CrossRef]
  33. Pazos, D.; Cintins, A.; de Castro, V.; Fernández, P.; Hoffmann, J.; Vargas, W.G.; Leguey, T.; Purans, J.; Anspoks, A.; Kuzmin, A.; et al. ODS ferritic steels obtained from gas atomized powders through the STARS processing route: Reactive synthesis as an alternative to mechanical alloying. Nucl. Mater. Energy 2018, 17, 1–8. [Google Scholar] [CrossRef]
  34. Rieken, J.R.; Anderson, I.E.; Kramer, M.J.; Odette, G.R.; Stergar, E.; Haney, E. Reactive gas atomization processing for Fe-based ODS alloys. J. Nucl. Mater. 2012, 428, 65–75. [Google Scholar] [CrossRef]
  35. Gil, E.; Cortés, J.; Iturriza, I.; Ordás, N. XPS and SEM analysis of the surface of gas atomized powder precursor of ODS ferritic steels obtained through the STARS route. Appl. Surf. Sci. 2018, 427, 182–191. [Google Scholar] [CrossRef]
  36. Zhou, Z.; Sun, S.; Zou, L.; Schneider, Y.; Schmauder, S.; Wang, M. Enhanced strength and high temperature resistance of 25Cr20Ni ODS austenitic alloy through thermo-mechanical treatment and addition of Mo. Fusion Eng. Des. 2019, 138, 175–182. [Google Scholar] [CrossRef]
  37. Song, M.; Sun, C.; Fan, Z.; Chen, Y.; Zhu, R.; Yu, K.Y.; Hartwig, K.T.; Wang, H.; Zhang, X. A roadmap for tailoring the strength and ductility of ferritic/martensitic T91 steel via thermo-mechanical treatment. Acta Mater. 2016, 112, 361–377. [Google Scholar] [CrossRef] [Green Version]
  38. Huang, X.; Wang, H.; Qiu, S.; Zhang, Y.; He, K.; Wu, B. Cold-rolling & annealing process for nuclear grade wrought FeCrAl cladding alloy to enhance the strength and ductility. J. Mater. Process. Technol. 2020, 277, 116434. [Google Scholar]
  39. Karlsson, H.; Nyborg, L.; Berg, S. Surface chemical analysis of prealloyed water atomised steel powder. Powder Metall. 2005, 48, 51–58. [Google Scholar] [CrossRef]
  40. Lim, J.; Hwang, I.S.; Kim, J.H. Design of alumina forming FeCrAl steels for lead or lead–bismuth cooled fast reactors. J. Nucl. Mater. 2013, 441, 650–660. [Google Scholar] [CrossRef]
  41. Wu, Y.; Haney, E.M.; Cunningham, N.J.; Odette, G.R. Transmission electron microscopy characterization of the nanofeatures in nanostructured ferritic alloy MA957. Acta Mater. 2012, 60, 3456–3468. [Google Scholar] [CrossRef]
  42. Hirata, A.; Fujita, T.; Liu, C.T.; Chen, M.W. Characterization of oxide nanoprecipitates in an oxide dispersion strengthened 14YWT steel using aberration-corrected STEM. Acta Mater. 2012, 60, 5686–5696. [Google Scholar] [CrossRef]
  43. Bai, Q.; Lin, J.; Tian, G.; Zou, J.; Dean, T.A. Review and Analysis of Powder Prior Boundary (PPB) Formation in Powder Metallurgy Processes for Nickel-based Super Alloys. J. Powder Metall. Min. 2015, 4, 2. [Google Scholar]
  44. Hou, J.; Dong, J.X.; Yao, Z.H.; Jiang, H.; Zhang, M.C. Influences of PPB, PPB affect zone, grain boundary and phase boundary on crack propagation path for a P/M superalloy FGH4096. Mater. Sci. Eng. A 2018, 724, 17–28. [Google Scholar] [CrossRef]
  45. Li, S.; Zhou, Z.; Li, M.; Wang, M.; Zhang, G. Microstructure characterization and tensile properties of 18Cr–4Al-oxide dispersion strengthened ferritic steel. J. Alloys Compd. 2015, 648, 39–45. [Google Scholar] [CrossRef]
  46. Kasada, R.; Toda, N.; Yutani, K.; Cho, H.S.; Kishimoto, H.; Kimura, A. Pre- and post-deformation microstructures of oxide dispersion strengthened ferritic steels. J. Nucl. Mater. 2007, 367–370, 222–228. [Google Scholar] [CrossRef]
  47. Sun, Z.; Edmondson, P.D.; Yamamoto, Y. Effects of Laves phase particles on recovery and recrystallization behaviors of Nb-containing FeCrAl alloys. Acta Mater. 2018, 144, 716–727. [Google Scholar] [CrossRef]
  48. Martínez-de-Guerenu, A.; Jorge-Badiola, D.; Gutiérrez, I. Assessing the recovery and recrystallization kinetics of cold rolled microalloyed steel through coercive field measurements. Mater. Sci. Eng. A 2017, 691, 42–50. [Google Scholar] [CrossRef]
  49. Guan, D.; Rainforth, W.M.; Gao, J.; Sharp, J.; Wynne, B.; Ma, L. Individual effect of recrystallisation nucleation sites on texture weakening in a magnesium alloy: Part 1- double twins. Acta Mater. 2017, 135, 14–24. [Google Scholar] [CrossRef]
  50. Dadé, M.; Malaplate, J.; Garnier, J.; De Geuser, F.; Barcelo, F.; Wident, P.; Deschamps, A. Influence of microstructural parameters on the mechanical properties of oxide dispersion strengthened Fe-14Cr steels. Acta Mater. 2017, 127, 165–177. [Google Scholar] [CrossRef] [Green Version]
  51. Boegelein, T.; Louvis, E.; Dawson, K.; Tatlock, G.J.; Jones, A.R. Characterisation of a complex thin walled structure fabricated by selective laser melting using a ferritic oxide dispersion strengthened steel. Mater. Charact. 2016, 112, 30–40. [Google Scholar] [CrossRef] [Green Version]
  52. Morris, D.G.; Muñoz-Morris, M.A. Nanoprecipitation of oxide particles and related high strength in oxide-dispersion-strengthened iron–aluminium–chromium intermetallics. Acta Mater. 2013, 61, 4636–4647. [Google Scholar] [CrossRef] [Green Version]
  53. Shen, Y.Z.; Zou, T.T.; Zhang, S.; Sheng, L.Z. Identification of Oxide Phases in Oxide Dispersion Strengthened PM2000 Steel. ISIJ Int. 2013, 53, 304–310. [Google Scholar] [CrossRef] [Green Version]
  54. Wei, Y.; Li, Y.; Zhu, L.; Liu, Y.; Lei, X.; Wang, G.; Wu, Y.; Mi, Z.; Liu, J.; Wang, H.; et al. Evading the strength-ductility trade-off dilemma in steel through gradient hierarchical nanotwins. Nat. Commun. 2014, 5, 3580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Li, Z.; Pradeep, K.G.; Deng, Y.; Raabe, D.; Tasan, C.C. Metastable high-entropy dual-phase alloys overcome the strength-ductility trade-off. Nature 2016, 534, 227–230. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, G.; Zhang, G.J.; Jiang, F.; Ding, X.D.; Sun, Y.J.; Sun, J.; Ma, E. Nanostructured high-strength molybdenum alloys with unprecedented tensile ductility. Nat. Mater. 2013, 12, 344–350. [Google Scholar] [CrossRef]
  57. Zerbst, U.; Madia, M.; Klinger, C.; Bettge, D.; Murakami, Y. Defects as a root cause of fatigue failure of metallic components. II: Non-metallic inclusions. Eng. Fail. Anal. 2019, 98, 228–239. [Google Scholar] [CrossRef]
  58. Yamakov, V.; Wolf, D.; Phillpot, S.R.; Mukherjee, A.K.; Gleiter, H. Deformation-mechanism map for nanocrystalline metals by molecular-dynamics simulation. Nat. Mater. 2004, 3, 43–47. [Google Scholar] [CrossRef] [PubMed]
  59. Zhou, X.; Liu, Y.; Yu, L.; Ma, Z.; Guo, Q.; Huang, Y.; Li, H. Microstructure characteristic and mechanical property of transformable 9Cr-ODS steel fabricated by spark plasma sintering. Mater. Des. 2017, 132, 158–169. [Google Scholar] [CrossRef]
  60. Zhao, Y.H.; Liao, X.Z.; Cheng, S.; Ma, E.; Zhu, Y.T.J.A.M. Simultaneously Increasing the Ductility and Strength of Nanostructured Alloys. Adv. Mater. 2006, 18, 2280–2283. [Google Scholar] [CrossRef]
  61. Sha, G.; Wang, Y.B.; Liao, X.Z.; Duan, Z.C.; Ringer, S.P.; Langdon, T.G. Influence of equal-channel angular pressing on precipitation in an Al–Zn–Mg–Cu alloy. Acta Mater. 2009, 57, 3123–3132. [Google Scholar] [CrossRef]
  62. Rao, P.N.; Singh, D.; Brokmeier, H.-G.; Jayaganthan, R. Effect of ageing on tensile behavior of ultrafine grained Al 6061 alloy. Mater. Sci. Eng. A 2015, 641, 391–401. [Google Scholar] [CrossRef]
  63. Ritchie, R.O. The conflicts between strength and toughness. Nat. Mater. 2011, 10, 817–822. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, Y.Y.; Liu, P.Z.; Li, J.J.; Liaw, P.K.; Spieckermann, F.; Kiener, D.; Qiao, J.W.; Eckert, J. Universally scaling Hall-Petch-like relationship in metallic glass matrix composites. Int. J. Plast. 2018, 105, 225–238. [Google Scholar] [CrossRef]
  65. Armstrong, R.; Codd, I.; Douthwaite, R.M.; Petch, N.J. The plastic deformation of polycrystalline aggregates. Philos. Mag. A J. Theor. Exp. Appl. Phys. 1962, 7, 45–58. [Google Scholar] [CrossRef]
  66. Yu, H.; Xin, Y.; Wang, M.; Liu, Q. Hall-Petch relationship in Mg alloys: A review. J. Mater. Sci. Technol. 2018, 34, 248–256. [Google Scholar] [CrossRef]
  67. Kim, J.H.; Byun, T.S.; Hoelzer, D.T.; Kim, S.-W.; Lee, B.H. Temperature dependence of strengthening mechanisms in the nanostructured ferritic alloy 14YWT: Part I—Mechanical and microstructural observations. Mater. Sci. Eng. A 2013, 559, 101–110. [Google Scholar] [CrossRef]
  68. Chauhan, A.; Bergner, F.; Etienne, A.; Aktaa, J.; de Carlan, Y.; Heintze, C.; Litvinov, D.; Hernandez-Mayoral, M.; Oñorbe, E.; Radiguet, B.; et al. Microstructure characterization and strengthening mechanisms of oxide dispersion strengthened (ODS) Fe-9%Cr and Fe-14%Cr extruded bars. J. Nucl. Mater. 2017, 495, 6–19. [Google Scholar] [CrossRef]
  69. Hall, E.O. The Deformation and Ageing of Mild Steel: II Characteristics of the L ders Deformation. Proc. Phys. Soc. Sect. B 1951, 64, 742–747. [Google Scholar] [CrossRef]
  70. Hansen, N. Hall–Petch relation and boundary strengthening. Scr. Mater. 2004, 51, 801–806. [Google Scholar] [CrossRef]
  71. Ukai, S.; Yano, Y.; Inoue, T.; Sowa, T. Solid-solution strengthening by Al and Cr in FeCrAl oxide-dispersion-strengthened alloys. Mater. Sci. Eng. A 2021, 812, 141076. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagrams of powder can (in mm) and (b) schematic diagrams of thermo-mechanical treatment process.
Figure 1. (a) Schematic diagrams of powder can (in mm) and (b) schematic diagrams of thermo-mechanical treatment process.
Materials 14 05696 g001
Figure 2. Schematic diagrams of tensile samples (in mm).
Figure 2. Schematic diagrams of tensile samples (in mm).
Materials 14 05696 g002
Figure 3. SEM secondary electron images of (a) the macro and (b,c) continuous precipitates for the as-HIPed sample; (dg) EDS map of continuous Al2O3 precipitates along the grain boundaries in (c) for the as-HIPed alloy; (h) distribution of grain diameter of the steel.
Figure 3. SEM secondary electron images of (a) the macro and (b,c) continuous precipitates for the as-HIPed sample; (dg) EDS map of continuous Al2O3 precipitates along the grain boundaries in (c) for the as-HIPed alloy; (h) distribution of grain diameter of the steel.
Materials 14 05696 g003
Figure 4. TEM images of (a) continuous precipitates at grain boundaries and continuous precipitates within grains; (b,d) EDS spectra of the arrowed precipitates in (a,c) for the as-HIPed sample, (e) the diameter statistical histogram of oxides in the as-HIPed sample.
Figure 4. TEM images of (a) continuous precipitates at grain boundaries and continuous precipitates within grains; (b,d) EDS spectra of the arrowed precipitates in (a,c) for the as-HIPed sample, (e) the diameter statistical histogram of oxides in the as-HIPed sample.
Materials 14 05696 g004
Figure 5. SEM secondary electron images of (a) macrostructure, (b) microstructure and (c) grain diameter statistical histogram of the as-TMTed sample.
Figure 5. SEM secondary electron images of (a) macrostructure, (b) microstructure and (c) grain diameter statistical histogram of the as-TMTed sample.
Materials 14 05696 g005
Figure 6. TEM images of (a) precipitates within the grains and (c) at grain boundaries in the as-TMTed sample; (b,d) EDS spectra of the arrowed precipitates in (a,c) for the as-TMTed sample, respectively, (e) the diameter statistical histogram of oxides in the as-TMTed sample.
Figure 6. TEM images of (a) precipitates within the grains and (c) at grain boundaries in the as-TMTed sample; (b,d) EDS spectra of the arrowed precipitates in (a,c) for the as-TMTed sample, respectively, (e) the diameter statistical histogram of oxides in the as-TMTed sample.
Materials 14 05696 g006
Figure 7. The ultimate tensile strength (UTS), yield strength (YS) and percentage elongation after rupture (A) for the as-HIPed and as-TMTed samples tested at 25 °C.
Figure 7. The ultimate tensile strength (UTS), yield strength (YS) and percentage elongation after rupture (A) for the as-HIPed and as-TMTed samples tested at 25 °C.
Materials 14 05696 g007
Figure 8. SEM images of the fracture surface of the tensile tested samples: (a) macro-morphology and (b,c) higher magnification micrograph of the intergranular fracture surface of the as-HIPed sample after tensile tests at 25 °C.
Figure 8. SEM images of the fracture surface of the tensile tested samples: (a) macro-morphology and (b,c) higher magnification micrograph of the intergranular fracture surface of the as-HIPed sample after tensile tests at 25 °C.
Materials 14 05696 g008
Figure 9. SEM images of the fracture surface of the tensile tested samples: (a) macro-morphology and (b) micro-morphology of fracture in the as-TMTed alloy after tensile tests at 25 °C.
Figure 9. SEM images of the fracture surface of the tensile tested samples: (a) macro-morphology and (b) micro-morphology of fracture in the as-TMTed alloy after tensile tests at 25 °C.
Materials 14 05696 g009
Table 1. Chemical composition of the FeCrAlY alloy (wt%).
Table 1. Chemical composition of the FeCrAlY alloy (wt%).
MaterialsCrAlYSiTiFe
FeCrAl alloy22.05.00.150.20.1Bal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Che, H.; Zhai, Y.; Yan, Y.; Chen, Y.; Qin, W.; Wang, T.; Cao, R. Effect of Thermo-Mechanical Treatment on the Microstructure and Tensile Properties of the Fe-22Cr-5Al-0.1Y Alloy. Materials 2021, 14, 5696. https://doi.org/10.3390/ma14195696

AMA Style

Che H, Zhai Y, Yan Y, Chen Y, Qin W, Wang T, Cao R. Effect of Thermo-Mechanical Treatment on the Microstructure and Tensile Properties of the Fe-22Cr-5Al-0.1Y Alloy. Materials. 2021; 14(19):5696. https://doi.org/10.3390/ma14195696

Chicago/Turabian Style

Che, Hongyan, Yazhong Zhai, Yingjie Yan, Yongqing Chen, Wei Qin, Tiejun Wang, and Rui Cao. 2021. "Effect of Thermo-Mechanical Treatment on the Microstructure and Tensile Properties of the Fe-22Cr-5Al-0.1Y Alloy" Materials 14, no. 19: 5696. https://doi.org/10.3390/ma14195696

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