Invited Review
Portevin-Le Châtelier effect in wrought Ni-based superalloys: Experiments and mechanisms

https://doi.org/10.1016/j.jmst.2020.03.023Get rights and content

Abstract

The Portevin-Le Châtelier (PLC) effect is a plastic instability in alloys at certain strain rates and deformation temperatures. This plastic instability exhibits serrated yielding in the temporal domain and strain localization in the spatial domain. Wrought Ni-based superalloys often exhibit the PLC effect. To guarantee the safe and stable operation of equipment, it is important to study the PLC effect in wrought Ni-based superalloys. This paper provides a review of various experimental phenomena and micromechanisms related to the PLC effect in wrought Ni-based superalloys, which have been reported in various publications in recent years and include work from our own group. The influences of stacking fault energy and γ′ precipitates on the PLC effect in wrought Ni-based superalloys are also discussed in detail. Additionally, several suggestions for the future study of the PLC effect in wrought Ni-based superalloys are provided.

Introduction

Wrought Ni-based superalloys play a dominant role in aero engines and industrial gas turbines based on several advantages [[1], [2], [3]]. First, the austenitic structure of nickel can withstand a wide range of temperatures and bond with numerous alloying elements, which is beneficial for material strength. Second, such alloys have excellent oxidation resistance. Third, desired mechanical properties can be obtained by optimizing the morphology, distribution, and volume fraction of γ′ precipitates. Generally, wrought Ni-based superalloys have service temperatures based on a high degree of alloying to intensify solution or precipitation strength [[4], [5], [6]]. For example, a Ni‒Co-based superalloy, which contains Co and Ti, can increase the service temperature by nearly 50 °C [7]. However, a high degree of alloying makes plastic deformational behavior (i.e. the Portevin-Le Châtelier (PLC) effect) more complex.

Plastic deformation behavior is the most fundamental and important mechanical behavior of alloys. The stress‒strain curves of uniform deformation are smooth before necking occurs in tensile tests. However, most alloys exhibit plastic instability with yield points [8] or serrated flow [9] phenomena. The yield point phenomenon generates an abrupt drop in the stress‒strain curve at the end of the elastic region. This drop in stress is followed by fluctuations in the curve, which correspond to local deformation bands (Lüders bands) in a tensile sample. The serrated flow phenomenon in the plastic region is also known as the PLC effect, which is common in most Ni-based superalloys [[10], [11], [12]]. The PLC effect often generates serrations in stress‒time curves and step changes in strain‒time curve. Additionally, tensile samples exhibit the propagation of strain localization, which is often referred to as PLC banding. Deformation bands, including Lüders bands and PLC bands, have a negative impact on the surface quality and further processing of alloys. Lüders bands only appear at the beginning of plastic deformation and can be eliminated via predeformation, while PLC bands can appear at any time during plastic deformation and cannot be removed effectively. Therefore, it is necessary to study the effects of deformation conditions on PLC and avoid the deformation region in which the PLC effect occurs.

Portevin and Châtelier first conducted detailed experimental studies on the plastic instability of aluminum alloys and proposed the concept of “serrated yielding” [13,14]. This plastic instability phenomenon was named as the PLC effect by Cottrell to honor their efforts [15]. Dynamic strain aging (DSA) has been widely accepted as the main mechanism of the PLC effect [[16], [17], [18]]. A serrated flow and negative strain rate sensitivity are the predominant characteristics in the DSA regime [19]. The critical strain of a serrated flow, which depends on the strain rates and temperatures, is a crucial factor in the PLC effect. Based on variations in critical strain, DSA can be divided into the normal DSA and inverse DSA regimes [20]. DSA is based on interactions between solute atoms and mobile dislocations. Therefore, all factors (including stacking fault (SF) energy and γ′ precipitates) that affect these interactions have an influence on the PLC effect [21,22].

Additionally, significant attention has been paid to the PLC effect in Ni-based superalloys based on its influence on mechanical properties. Valsan et al. [23] determined that DSA influences the deformation and fracture behaviors of the Nimonic PE 16 superalloy in a temperature range of 450–550 °C at low strain rates. Gopinath et al. [24] identified dislocations pinned by solutes because DSA can offset softening during the low-cycle fatigue testing of the U720Li alloy. Rezende et al. [25] reported that DSA and precipitates are responsible for the high strength of the Inconel 718 superalloy at elevated temperatures. Pu et al. [26] inferred that the intermediate-temperature embrittlement of the Ni-based UNS N10276 superalloy may be associated with a strong PLC effect. Therefore, the PLC effect can be utilized to enhance mechanical strength, but should be avoided during hot deformation processes. Understanding the PLC effect has great significance for designing and applying wrought Ni-based superalloys.

This paper presents a review of PLC effects in wrought Ni-based superalloys. Our main focuses can be divided into the experimental phenomena and micromechanisms of PLC effects in wrought Ni-based superalloys. Additionally, shortcomings and prospects are discussed in the final section.

Section snippets

Serrated flow

Serrated flow behavior is the main external manifestation of the PLC effect in wrought Ni-based superalloys. This behavior depends on the tensile temperature and strain rate [27]. Generally, the PLC effect can be divided into types A, B, and C. Fig. 1 presents the tensile stress‒strain curves of wrought Ni-based superalloys under different deformation conditions [28,29]. Serration shape changes with variations in tensile temperature are presented in Fig. 1(a). Serrations of types A, B, and C in

Effects of SF energy on the PLC effect

Cui et al. [59] determined that both normal and inverse PLC effects occur at temperatures ranging from 300 to 500 °C for a Ni‒Co-based superalloy, as shown in Fig. 13(a). The defining tensile characteristic for the normal PLC effect is dislocation movement (Fig. 13(b)), while a high density of SFs is a characteristic of the inverse PLC effect (Fig. 13(c)). Therefore, it can be concluded that the inverse PLC effect has a relationship with SFs. The influence of SFs on the PLC effect has also been

Influence of γ′ precipitates on the PLC effect

For precipitation-strengthened wrought Ni-based superalloys, γ′ precipitates make a significant contribution to strength improvement by hindering dislocation movement during plastic deformation [65]. Pink et al. [66] hypothesized that the serrations on the stress‒strain curves of Al‒Li alloys are related to the shearing of precipitates. Sun et al. [67] stated that precipitates should not be ignored when considering the PLC effect in the A2024 aluminum alloy and solute clouds. However, the

Mechanisms of the PLC effect

DSA is a generally accepted micromechanism of the PLC effect [[75], [76], [77]]. However, specific microprocess descriptions (e.g., the dispersal mode of solutes and interactions between solutes and dislocations) are still under discussion based on a lack of direct observational evidence.

Cottrell and Jaswon [78] first presented DSA to explain the PLC effect based on the relationship between the velocity of solute diffusion and the velocity of dislocation movement. If the velocity of solute

Conclusions

This paper reviewed the PLC effect in wrought Ni-based superalloys, which have been featured in numerous publications, including work from our group, in recent years. The main conclusions of our research are as follows:

  • (1)

    The PLC effect in Co-rich wrought Ni-based superalloys is caused by unpinning, which can be explained by interactions of mobile dislocations with substitutional solute (Cr) atmospheres. Increasing Co content (or decreasing SFE) causes the PLC region to move toward higher

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (Nos. 51671189 and 51271174) and the Ministry of Science and Technology of China (Nos. 2017YFA0700703 and 2019YFA0705304).

References (103)

  • L. Tan et al.

    J. Alloys Compd.

    (2019)
  • M. Detrois et al.

    J. Alloys Compd.

    (2019)
  • T.D. Reynolds et al.

    Acta Mater.

    (2019)
  • X. Zhang et al.

    Comput. Mater. Sci.

    (2013)
  • H. Long et al.

    J. Alloys Compd.

    (2018)
  • R. Eriş et al.

    Intermetallics

    (2019)
  • R. Schwab et al.

    Acta Mater.

    (2013)
  • K. Yu et al.

    J. Mater. Sci. Technol.

    (2017)
  • H. Aboulfadl et al.

    Acta Mater.

    (2015)
  • J.B. Seol et al.

    Acta Mater.

    (2017)
  • C. Meng et al.

    Acta Mater.

    (2019)
  • C. Tian et al.

    J. Mater. Sci. Technol.

    (2013)
  • Y. Cai et al.

    Mater. Sci. Eng. A

    (2015)
  • K. Gopinath et al.

    Acta Mater.

    (2009)
  • M.C. Rezende et al.

    J. Alloys Compd.

    (2015)
  • E. Pu et al.

    Mater. Sci. Eng. A

    (2017)
  • C.L. Hale et al.

    Mater. Sci. Eng. A

    (2001)
  • Y. Cao et al.

    Mater. Sci. Eng. A

    (2018)
  • M. Lebyodkin et al.

    Acta Mater.

    (2000)
  • M.A. Lebyodkin et al.

    Acta Mater.

    (2005)
  • P. Maj et al.

    Mater. Sci. Eng. A

    (2014)
  • K. Gopinath et al.

    Acta Mater.

    (2009)
  • A. Nagesha et al.

    Mater. Sci. Eng. A

    (2012)
  • C.V. Rao et al.

    Mater. Sci. Eng. A

    (2019)
  • P. Penning

    Acta Metall.

    (1972)
  • P.G. Mccormick

    Acta Metall.

    (1988)
  • P.G. Mccormick

    Scr. Metall.

    (1978)
  • H.J. Harun et al.

    Acta Metall.

    (1979)
  • G.G. Saha et al.

    Mater. Sci. Eng.

    (1984)
  • K. Chihab et al.

    Scr. Metall.

    (1987)
  • A. Ziegenbein et al.

    Comput. Mater. Sci.

    (2000)
  • H. Louche et al.

    Mater. Sci. Eng. A

    (2005)
  • Q. Zhang et al.

    Int. J. Plast.

    (2005)
  • H. Halim et al.

    Acta Mater.

    (2007)
  • G.F. Xiang et al.

    Scr. Mater.

    (2007)
  • H. Jiang et al.

    Acta Mater.

    (2007)
  • Y.L. Cai et al.

    J. Mater. Sci. Technol.

    (2017)
  • J. Min et al.

    Mech. Mater.

    (2014)
  • P.D. Zavattieri et al.

    Int. J. Plast.

    (2009)
  • K. Renard et al.

    Mater. Sci. Eng. A

    (2010)
  • C.Y. Cui et al.

    Scr. Mater.

    (2011)
  • S. Lee et al.

    Scr. Mater.

    (2011)
  • S.J. Lee et al.

    Acta Mater.

    (2011)
  • S. Venkadesan et al.

    Acta Metall. Mater.

    (1992)
  • E. Pink et al.

    Mater. Sci. Eng. A

    (2000)
  • Y. Cai et al.

    J. Alloys Compd.

    (2017)
  • W.H. Wang et al.

    J. Mater. Sci. Technol.

    (2018)
  • X. Wang et al.

    Metall. Mater. Trans. A

    (2016)
  • X. Wang et al.

    J. Mater. Sci. Technol.

    (2019)
  • B.D. Fu et al.

    Mater. Lett.

    (2015)
  • Cited by (35)

    • The Portevin-Le Chatelier effect in nickel-base superalloys: Origins, consequences and comparison to strain ageing in other alloy systems

      2023, Progress in Materials Science
      Citation Excerpt :

      Concerns have been raised in some alloys such as Al-Mg [20], Twin-Induced Plasticity (TWIP) steels [21] and superalloys [22] as to whether the temperatures at which serrations are observed are sufficient for long range atomic diffusion to occur. There is wide agreement that the diffusion of solute atoms remains responsible for locking at intermediate temperatures in FCC alloys [23,19], but alternate microscopic mechanisms for solute–defect locking interactions have been proposed (Section 3.3), the details of which lack an accepted explanation. While others use the term DSA more generally e.g. [21,24,20] to encompass such solute locking mechanisms of serrated flow, these are distinguished here by taking DSA to refer only to the classical mechanisms of strain ageing, involving long range diffusion of interstitial or substitutional solute to dislocation cores and elastic locking (Section 3.2).

    View all citing articles on Scopus
    View full text