Skip to main content
Top
Published in: Metallurgical and Materials Transactions A 6/2023

Open Access 10-03-2023 | Original Research Article

Influence of Grain Size and Volume Fraction of η/δ Precipitates on the Dwell Fatigue Crack Propagation Rate and Creep Resistance of the Nickel-Base Superalloy ATI 718Plus

Authors: A. Kirchmayer, M. Pröbstle, D. Huenert, S. Neumeier, M. Göken

Published in: Metallurgical and Materials Transactions A | Issue 6/2023

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

Recently, it was found that the orientation of the η/δ plates in the Ni-base superalloys ATI 718Plus has a huge impact on the fatigue crack propagation behavior (Huenert et al. in Superalloys pp.781–792, 2016). However, the targeted alignment of η/δ precipitates in components with complex geometry can be challenging in certain applications. Therefore, the effect of other microstructural features of ATI 718Plus, which are easier to manipulate, is tested with respect to their fatigue crack propagation behavior and creep strength. The grain size, the γ′ precipitate size, as well as the γ′ and η/δ volume fraction are varied by applying different heat treatments. An increase in the η/δ volume fraction as well as the γ′ volume fraction and precipitate size enhanced the crack growth resistance, but worsened the creep properties. On the contrary, increasing the grain size leads to better creep properties and better fatigue crack growth resistance. The impact of the different microstructures investigated in this study on the mechanical properties is smaller compared to microstructures with strongly aligned η/δ plates. These results enable to design the desired microstructures that meet the particular demands of the given application regarding crack propagation and creep resistance.
Notes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Critical materials for applications in jet engines or stationary gas turbines have to show outstanding mechanical properties, such as high temperature stability, creep resistance, and crack resistance at extreme conditions. One alloy that meets these requirements by simultaneous good processability is the polycrystalline Ni-base superalloy ATI718Plus® (hence called 718Plus) which can developed on the basis of the commonly used Ni-base superalloy Inconel 718.[28] 718Plus has two principle precipitate phases: γ′ as the main hardening phase and δ/η phase to control grain growth during processing. It has a higher thermal stability than Inconel 718 due to the utilization of the γ′ precipitate phase instead of the less stable γ″ phase in Inconel 718. Depending on the processing and heat treatment procedure, either the δ or η phase forms as main grain boundary precipitate phase. The δ phase can precipitasuste inside η particles or vice versa.[912]
Investigations by Huenert et al. have shown that the orientation of the plate-like η/δ phase can be adjusted by optimizing the material flow.[1] The orientation of the η/δ plates has a high impact on crack propagation properties. In Reference 1, a “good” orientation, where the η/δ precipitates are aligned perpendicular to the direction of crack propagation, is distinguished from a “bad” orientation with aligned η/δ precipitates that are parallel to the growing crack. The good orientation slowed the crack propagation enormously in comparison with the bad orientation.[1] Since the targeted alignment of η/δ precipitates in components with complex geometry can be challenging in certain applications, the knowledge about the impact of other microstructural features is also of great importance.
In this study, the grain size, the γ′ precipitate size, and the η/δ volume fraction are varied systematically in different samples with randomly oriented η/δ plates. Since an improved crack propagation resistance is typically accompanied with a reduced creep strength, the creep properties of the microstructure variants are also investigated. These findings are finally compared to the orientation effect of the η/δ plates as measured by Huenert et al.[1] as well as the commercially used Ni-base superalloys such as IN718, UDIMET720Li, and WASPALOY in order to classify the different microstructural parameters regarding their impact on crack propagation.

2 Experimental Methods

The material used in this study was provided by Rolls-Royce Deutschland Ltd & Co Kg (RRD, Deutschland), which was processed by Allegheny Technologies Incorporated (ATI, USA) into billets and hot formed at Otto Fuchs KG (Germany) into a near turbine disk form. The nominal composition of the alloy is given in Table I.
Table I
Nominal Composition of 718Plus in Wt Pct According to Ref. 1
Ni
Co
Cr
Nb
Al
Ti
Mo
Fe
W
Bal
8.0 to 10.0
17.0 to 21.0
5.2 to 5.8
1.2 to 1.7
0.5 to 1.0
2.5 to 3.1
8.0 to 10.0
0.8 to 1.4
Specimens were cut from the provided material and further heat treated to create different microstructures. The standard heat treatment is based on the recommendations of ATI[13]: (1) pre-solution heat treatment (PSHT) between 843 and 871 °C for 16 h, (2) air cooling or raising to the solution heat treatment (SHT) between 954 and 982 °C up to 1 h, air cooling or faster cooling, and (3) γ′ age hardening at 788 °C for 2–8 h, furnace cooling + 704 °C for 8 h air cooling. The variations of the different heat treatment steps to generate different microstructures are shown in Table II. The IDs of the investigated conditions describe the properties of the targeted microstructure. Different grain sizes, different η/δ volume fractions, as well as one additional condition with an increased γ′ diameter were realized. Each sample is labeled with an ID for its grain size (small grain size: SG; medium grain size: MG; large grain size: LG) and its η/δ volume fraction. It is important to note that the η/δ precipitates show no preferred alignment in all conditions. Additionally, the same texture in all conditions is guaranteed as the different conditions were generated from the same initial material. Further, no difference in the twin density was confirmed by analyzing the BSD images of the different conditions. Accordingly, the mechanical properties are solely affected by the microstructural features mentioned in Table III.
Table II
Heat Treatment Conditions Applied to 718Plus to Adjust Different Microstructures, See the Different IDs
ID
PSHT
SHT
Two-step γ′ age hardening
SG-0 Pct
980 °C/15 min
*
*
SG-5 Pct
*
*
*
*
SG-10 Pct
900 °C/16 h
*
*
*
SG-17 Pct
900 °C/80 h
*
*
MG-0 Pct
1000 °C/1 h
*
*
LG-0 Pct
1050 °C/3 h
*
*
SG-7 Pct
870 °C/16 h
*
*
*
Only the differences compared to the individual standard heat treatment steps are shown.
*Standard heat treatment steps.
Table III
Stereological Parameters of All Microstructural Conditions
ID
Median Grain Size (µm)
η/δ Volume Fraction (Pct)
Average γ′ size (nm)
γ′ Volume Fraction (Pct)
SG-0 pct
8.1
0.1
28
25.4
SG-5 pct
8.6
5.2
23
23.2
SG-10 pct
9.1
10.3
26
19.0
SG-17 pct
8.2
16.6
23
14.8
SG-7 pct
9.8
7.3
46
21.9
MG-0 pct
15.5
0.2
25
*
LG-0 pct
126.7
0.1
21
*
The IDs of the different conditions relate to their total volume fractions of η and δ phase.
*The γ′ volume fraction of the conditions MG and LG are not relevant for mechanical investigation and are therefore not determined in this study.
To investigate the different microstructures, the specimens were mechanically polished and examined using backscattered electrons (BSE) in a Zeiss Crossbeam 1540 scanning electron microscope (SEM) with an acceleration voltage of 20 kV at a working distance of 7.5 mm. With the same equipment, energy-dispersive X-ray spectroscopy (EDS) was conducted to determine the chemical composition of the γ / γ′ compound in the grain interior. To measure the γ′ precipitate diameter, dark field transmission electron (TEM) images were made using a Philips CM200. The dark field images were taken at 200 kV using the installed 2 k × 2 k CMOS camera F216 (TVIPS). For this, thin foils were mechanically polished and further thinned by electropolishing with an electrolyte solution containing 8 pct perchloric acid and an applied voltage of 40 V at a temperature of − 30 °C.
The grain size as well as the γ′ precipitate diameter was determined using the linear intercept method and the freeware image tool ImageJ.[14] The η/δ area fraction was determined using the free ImageJ plugin “Weka Trainable Segmentation.”[15]
To calculate the γ′ volume fraction, the lever-rule (1) was used according to Blavette et al.[16]:
$$ c_{n} \; = \;f_{{\gamma ^{\prime}}} + \left( {1 - f_{{\gamma ^{\prime}}} } \right)c_{\gamma } $$
(1)
where f γ′ accounts for the γ′ volume fraction and cn is the chemical composition of the γ/γ′ compound in the grain interior without η/δ phase. For the chemical compositions of the γ and γ′ phase (cγ and cγ′ accordingly), atom probe tomography (APT) data from an earlier study on 718Plus were used.[17] Since the microstructure also contains η/δ precipitates, the determined fγ′ must be corrected by the η/δ volume fraction
$$ f_{{\gamma ^{\prime}\left( {{\text{corrected}}} \right)}} \; = \;f_{{\gamma ^{\prime}}} \left( {1 - f_{{\eta /\delta }} } \right), $$
(2)
with the assumption, that the η/δ area fraction equals the η/δ volume fraction fη/δ.
To characterize the mechanical properties of the different microstructures, fatigue crack propagation tests and compressional creep experiments were conducted. Compact tension (CT) specimens according to ASTM E 647-13[18] were used to determine the fatigue crack propagation rate. The samples were heated to 650 °C in air for the actual crack propagation test. A trapezoid signal with 1-120-1-1 was used with 120 seconds dwell time at maximum stress till the ratio of crack length a to sample width w reached 0.6. The crack length was measured using the potential drop method. The specimens for the compression creep experiments were cylindrical with a height of 6 mm and a diameter of 4 mm. The samples were tested in a pneumatic testing machine at 650 °C and different applied constant stresses ranging from 800 to 1100 MPa.

3 Results and Discussion

3.1 Microstructures

3.1.1 Conditions with different η/δ fractions

The microstructures with median grain sizes between 8.1 and 9.8 µm (SG), but different η/δ volume fractions are shown in Figure 1. The condition SG-0 pct possesses nearly no η/δ, see also the overview about all stereological parameters of the different microstructures given in Table III.
Figure 1 shows different conditions with similar grain sizes but different η/δ volume fractions. No alignment of the η/δ phase is visible which mainly precipitates at grain boundaries. While the change in the η/δ volume fractions can be clearly seen in the SE images (a) through (e), the γ′ precipitate size stays virtually the same. The only exception to this is the condition SG-7 pct shown in Figures 1(e) and (f). Here, an increase in the secondary γ′ precipitate size distribution compared to the other SG conditions is revealed. This is also seen in Table III. For the explanation of the formation of the different types of ordered phases in A718Plus, the reader is referred to Reference 12.
Beside the homogenously distributed secondary γ′ precipitates, the two conditions SG-10 pct and SG-17 pct reveal coarse primary γ′ precipitates [Figures 2(a) and (b)]. Figure 2 also shows a depletion of γ′ precipitates around the η/δ precipitates. This is due to a depletion of γ′ building elements in the vicinity of the η/δ precipitates.

3.1.2 Conditions with different grain sizes

Figure 3 shows the difference between the grain sizes of the investigated conditions SG, MG, and LG, see also Table III. Although the grain size strongly increases from SG to LG by a factor of 15, the γ′ precipitate size stays nearly the same with the differences of only a few nm. Additionally, the η/δ volume fraction is about 0 pct in all three conditions. This shows that the grain size is the only microstructural parameter which changed due to the different heat treatments.

3.1.3 Correlation between γ′ and η/δ phase fraction

The γ′ volume fractions that were determined by the lever-rule as well as the η/δ volume fractions of the different heat treatment conditions are shown in Figure 4. The γ′ volume fraction declines linearly with increasing η/δ volume fraction. This is due to the similar chemical composition of γ′ and η/δ[7,8,10,19,20] which leads to a depletion of one phase if the phase-forming elements, mainly Al, Nb, and Ti, are consumed by the other phase. Therefore, a depletion of γ′ precipitates adjacent to grain boundary phases occur, as visible in Figure 2(d). The slope of the fitted line in Figure 4 is significantly lower than 1, which means that a decrease in γ′ volume fraction does not lead to an equal increase of the η/δ volume fraction and vice versa. This indicates that γ′ and η/δ-forming elements like Al and Nb partition more strongly toward the γ′ phase, i.e., more Al and/or Nb are necessary to form γ′ in comparison with η/δ as well as more η/δ-forming elements are available, if the γ′ content is reduced.

3.1.4 Fatigue crack propagation resistance

To determine the influence of the different microstructures on crack propagation and crack growth resistance, dwell fatigue crack propagation tests were conducted. In Figure 5, the measured crack propagation rates of the different conditions are portrayed as a function of the stress intensity factor ΔK. For the SG conditions, the crack growth resistance increases with increasing f η/δ, see Figure 5(a) from 0 to 17pct. Especially, at low values of ΔK from 20 to 30 MPa m−2, the impact of the η/δ volume fraction is significant whereas at higher values of ΔK from 40 to 50 MPa m−2, the impact is less pronounced. An increased average γ′ size of 46 nm of SG-7 pct in comparison with the average γ′ precipitate size of 23 nm of SG-5 pct could have an effect on the higher crack growth resistance. Figure 5(b) shows that an increasing grain size leads to a higher crack growth resistance, as the crack propagation rate of LG with a median grain size of about 127 µm is significantly lower than that of SG and MG with median grain sizes of about 8.1 and 15 µm accordingly. Interestingly, while the change in grain size from 8.1 to 15.5 µm does not lead to a significantly different dwell fatigue crack propagation rate, especially at higher ∆K values, an increase of the grain size to 126.7 µm leads to a much smaller dwell fatigue crack propagation rate.

3.2 Creep Strength

The plot in Figure 6 shows the minimum creep rate as a function of applied stress. Increasing the η/δ volume fraction while keeping the grain size constant clearly leads to a decrease in the creep resistance [see Figure 6(a)]. An increase in γ′ precipitate size reduces the creep resistance further, as the condition SG-7 pct with fγ′ of 21.9 pct and a γ′ precipitate size of 46 nm performs worse than the condition SG-10 pct with fγ′ of 19.0 pct but an average γ′ precipitate size of 26 nm. Figure 6(b) shows the influence of different grain sizes on the minimum creep rate. Increasing the grain size improves the creep properties slightly at the tested conditions. This can be seen in the higher minimum creep rates of SG-0 pct and MG-0 pct, with relatively small grain sizes compared to LG-0 pct. A larger difference in the grain size leads to a slight decrease of the minimum creep rates. This leads to the conclusion that the difference in the grain size has a much smaller influence than a change in the η/δ volume fraction on the creep properties at the tested temperature and stresses.

4 Discussion

An overview of the different influencing factors on the crack resistance is shown in Figure 7. The results from this study are compared to the effect of the orientation of η/δ plates, as studied by Huenert et al.[1] as well as to the commercial Ni-base Superalloys IN718[2124] UDIMET720Li (hence called U720 Li) [25] and WASPALOY.[5]
The condition SG-0 pct with an median grain size of 8.1 µm shows a similar crack propagation resistance as IN718.[2124] The fatigue crack propagation rate can be reduced by about one order of magnitude by strongly increasing the η/δ volume fraction or grain size, as can be seen by the crack propagation rates of SG-17 pct and LG-0 pct. With the precipitation of the η/δ phases at the grain boundaries, the crack does not propagate exclusively along grain boundaries but also along the η/δ plates. Accordingly, the grain boundary η/δ precipitates lead to crack blocking and deflection and thus an increased crack length, which varies strongly with the orientation of the η/δ precipitates compared to the crack propagation direction. Therefore, a higher η/δ volume fraction leads to a higher fatigue crack propagation resistance. Additionally, the crack driving force decreases with higher fractions of η/δ precipitates due to a more pronounced stress relaxation in the microstructure, according to Reference 26.
The grain size seems to have no significant effect on the dwell fatigue crack propagation in the range between 8.1 and 15.5 µm. Only after significantly increasing the grain size to 126.7 µm, a better dwell fatigue crack propagation resistance is obtained. This fits to the results from Chang et al.,[27] where no pronounced difference in the crack growth rate of IN718 at elevated temperatures between a grain size of 10 and 50 µm could be measured as well as other studies, which showed that an strongly increased grain size enhances crack propagation resistance.[28,29] An increase in grain size reduces the density of grain boundaries and therefore reduces possible favorable pathways for the crack to propagate, leading to a deflection of the crack along the grain boundaries. This effect seems to only appear in air, which hints to an embrittlement of the grain boundaries due to high temperature oxidation.[28]
The combination of a higher η/δ volume fraction and larger γ′ precipitate size has a stronger positive effect on the crack propagation resistance than an increase of the volume fraction alone as seen in Figure 5(a), the condition SG-7 pct shows better fatigue crack propagation resistance than the condition SG-10 pct with a higher η/δ content but with smaller γ′ precipitates. A similar effect can be assumed by comparing Figures 5(a) and (b), which shows that a combination of a bigger grain size and higher η/δ volume fraction most likely increases the crack propagation resistance more than solely increasing the η/δ volume fraction or solely increasing the grain size. A positive effect of a combination of different microstructural parameters like η/δ precipitates at grain boundaries in combination with medium grain sizes was also stated by References 26 and 30 where twins played a role, too. Obviously, the increase in grain size, η/δ volume fraction, and γ′ size has a positive impact on the crack propagation resistance of 718Plus. However, in comparison with the effect of the η/δ plate orientation, these improvements are much smaller, which is clearly demonstrated by the blue lines which represent the conditions with microstructures of aligned η/δ plates. The condition labeled as “bad” (upper blue dashed line) corresponds to a η/δ plate orientation of 0° in relation to the crack propagation direction. The condition labeled as “good” shows the results of microstructures with η/δ plates having an orientation of 90° in relation to the crack propagation direction. More detailed information about the different microstructures with aligned η/δ plates can be found in Reference 12 These two groups of conditions with differently aligned η/δ plates demonstrate that beneficially oriented η/δ precipitates result in a strong decrease of the crack propagation rate by at least two orders of magnitude. However, the improvement based on the η/δ plate orientation depends strongly on the crack propagation direction. On the contrary, the improvements due to grain size, η/δ volume fraction, and the γ′ size in conditions with unaligned η/δ plates are independent of the crack propagation direction and therefore should have an advantage under certain conditions.
The creep resistance can be improved by a combination of different microstructural properties, too. Figure 6 shows that a combination of big grains with small γ′ precipitates and a small amount of η/δ phase, which goes along with a high fraction of strengthening γ′ precipitates, enhances the creep resistance at 650 °C. However, as stated before, a decrease of η/δ volume fraction worsens the fatigue crack growth resistance[3135] as shown in Figure 8. Apparently, there seems to be a correlation between an increased fatigue crack propagation resistance and a lower creep resistance and vice versa. This could be attributed to a stress relaxation in front of the crack tip by creep deformation during the dwell time. A lower creep resistance and therefore easier plastic deformation lead to a stronger blunting of the crack tip, which reduces the crack propagation per cycle.[31,33] Besides, the reason for the higher creep strength at a lower fatigue crack growth resistance could be that a lower η/δ volume fraction results in a higher fraction of the strengthening γ′ phase (see Figure 4) and thus higher creep strength, but also in a reduced crack deflection by a lower content of the η/δ plates.

5 Summary and Conclusion

The impact of different microstructural parameters on the dwell fatigue crack properties as well as the creep properties of the Ni-base superalloy 718Plus was studied at 650 °C. Therefore, various microstructures were generated by proper adjustments of the heat treatments that lead to variations in the η/δ volume fraction from 0 to 17 pct, in the grain size from 8 to 127 µm, and in γ′ size from 21 to 46 nm. All η/δ containing microstructures showed a random orientation of the η/δ plates. The following two main conclusions on the correlation between microstructure and mechanical properties can be drawn:
  • An increase of the η/δ volume fraction as well as an increase in the γ′ precipitate size enhances the crack growth resistance but worsens the creep properties. An increase in grain size benefits the crack properties, as well as increasing the creep strength.
  • In comparison with the beneficial effect of well-oriented η/δ plates, the variation of the η/δ volume fraction, grain size, or γ′ size reveals a less significant impact on the fatigue crack propagation resistance. A change of the η/δ volume fraction, grain size, and γ′ precipitate size results in an order of magnitude lower crack propagation rate at the most, while a microstructure with well-oriented η/δ plates improves the crack propagation rate by more than two orders of magnitude.
These results allow to design the desired microstructures that meet the particular demands of the given application regarding crack propagation and creep resistance.

Acknowledgments

The work was conducted as a part of the German-funded research program RoKoTec (LuFoIV) and supported by the Bundesministerium für Wirtschaft und Technologie (BMWi) under Grant Number 20T0813.

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literature
1.
go back to reference D. Huenert, M. Proebstle, A. Casanova, R. Schluetter, R. Schluetter, R. Krakow, R. Krakow, M. Buescher, P. Randelzhofer, A. Evans, K. Loehnert, T. Witulski, S. Neumeier, and C. Rae: in Superalloys 2016, Mark Hardy, Eric Huron, Uwe Glatzel, Brian Griffin, Beth Lewis, Cathie Rae, Venkat Seetharaman, and Sammy Tin, eds., John Wiley & Sons, Inc., Hoboken, NJ, USA, 2016, pp. 781–92. D. Huenert, M. Proebstle, A. Casanova, R. Schluetter, R. Schluetter, R. Krakow, R. Krakow, M. Buescher, P. Randelzhofer, A. Evans, K. Loehnert, T. Witulski, S. Neumeier, and C. Rae: in Superalloys 2016, Mark Hardy, Eric Huron, Uwe Glatzel, Brian Griffin, Beth Lewis, Cathie Rae, Venkat Seetharaman, and Sammy Tin, eds., John Wiley & Sons, Inc., Hoboken, NJ, USA, 2016, pp. 781–92.
2.
go back to reference L. Viskari, Y. Cao, M. Norell, G. Sjöberg, and K. Stiller: Mater. Sci. Eng., A, 2011, vol. 528, pp. 2570–580.CrossRef L. Viskari, Y. Cao, M. Norell, G. Sjöberg, and K. Stiller: Mater. Sci. Eng., A, 2011, vol. 528, pp. 2570–580.CrossRef
3.
go back to reference K. Löhnert and F. Pyczak,: Superalloys, 2010. K. Löhnert and F. Pyczak,: Superalloys, 2010.
4.
go back to reference G.A. Zickler, R. Schnitzer, R. Radis, R. Hochfellner, R. Schweins, M. Stockinger, and H. Leitner: Mater. Sci. Eng. A, 2009, vol. 523, pp. 295–303.CrossRef G.A. Zickler, R. Schnitzer, R. Radis, R. Hochfellner, R. Schweins, M. Stockinger, and H. Leitner: Mater. Sci. Eng. A, 2009, vol. 523, pp. 295–303.CrossRef
5.
go back to reference R.M. Kearsey, J. Tsang, S. Oppenheimer, and E. McDevitt: JOM, 2012, vol. 64, pp. 241–51.CrossRef R.M. Kearsey, J. Tsang, S. Oppenheimer, and E. McDevitt: JOM, 2012, vol. 64, pp. 241–51.CrossRef
6.
go back to reference W.-D. Cao: Superalloys, 2005, vol. 718, pp. 166–77. W.-D. Cao: Superalloys, 2005, vol. 718, pp. 166–77.
7.
go back to reference A. Kruk, G. Cempura, S. Lech, and A. Czyrska -Filemonowicz: Archives of Metallurgy and Materials, 2016, vol. 61, pp. 535–42. A. Kruk, G. Cempura, S. Lech, and A. Czyrska -Filemonowicz: Archives of Metallurgy and Materials, 2016, vol. 61, pp. 535–42.
8.
go back to reference A. Casanova, N. Martín-Piris, M. Hardy, and C. Rae: MATEC Web of Conferences, 2014, vol. 14, p. 09003.CrossRef A. Casanova, N. Martín-Piris, M. Hardy, and C. Rae: MATEC Web of Conferences, 2014, vol. 14, p. 09003.CrossRef
9.
go back to reference B. Hassan and J. Corney: Mater. Sci. Technol., 2017, vol. 33, pp. 1879–889.CrossRef B. Hassan and J. Corney: Mater. Sci. Technol., 2017, vol. 33, pp. 1879–889.CrossRef
10.
go back to reference E.J. Pickering, H. Mathur, A. Bhowmik, O.M.D.M. Messé, J.S. Barnard, M.C. Hardy, R. Krakow, K. Loehnert, H.J. Stone, and C.M.F. Rae: Acta Mater., 2012, vol. 60, pp. 2757–769.CrossRef E.J. Pickering, H. Mathur, A. Bhowmik, O.M.D.M. Messé, J.S. Barnard, M.C. Hardy, R. Krakow, K. Loehnert, H.J. Stone, and C.M.F. Rae: Acta Mater., 2012, vol. 60, pp. 2757–769.CrossRef
11.
go back to reference Robert Krakow, Duncan N. Johnstone, Alexander S. Eggeman, Daniela Hünert, Mark C. Hardy, Catherine M. F. Rae, and Paul A. Midgley: Acta Materialia, 2017, vol. 130, pp. 271–80. Robert Krakow, Duncan N. Johnstone, Alexander S. Eggeman, Daniela Hünert, Mark C. Hardy, Catherine M. F. Rae, and Paul A. Midgley: Acta Materialia, 2017, vol. 130, pp. 271–80.
12.
go back to reference Daniel Hausmann, Andreas Förner, Martin Pröbstle, Daniela Hünert, Peter Felfer, Mathias Göken, and Steffen Neumeier: Advanced Engineering Materials, 2021, p. 2100558. Daniel Hausmann, Andreas Förner, Martin Pröbstle, Daniela Hünert, Peter Felfer, Mathias Göken, and Steffen Neumeier: Advanced Engineering Materials, 2021, p. 2100558.
13.
go back to reference Wei-Di Cao and Richard L. Kennedy: ATI Allvac, 2006. Wei-Di Cao and Richard L. Kennedy: ATI Allvac, 2006.
14.
go back to reference C.A. Schneider, W.S. Rasband, and K.W. Eliceiri: Nat Methods, 2012, vol. 9, pp. 671–75.CrossRef C.A. Schneider, W.S. Rasband, and K.W. Eliceiri: Nat Methods, 2012, vol. 9, pp. 671–75.CrossRef
15.
go back to reference I. Arganda-Carreras, V. Kaynig, C. Rueden, K.W. Eliceiri, J. Schindelin, A. Cardona, and H. Sebastian Seung: Bioinformatics, 2017, vol. 33, pp. 2424–26. I. Arganda-Carreras, V. Kaynig, C. Rueden, K.W. Eliceiri, J. Schindelin, A. Cardona, and H. Sebastian Seung: Bioinformatics, 2017, vol. 33, pp. 2424–26.
16.
go back to reference D. Blavette, P. Caron, and T. Khan: in Superalloys: Sixth International Symposium. TMS, 1988, vol. 1988, pp. 305–14. D. Blavette, P. Caron, and T. Khan: in Superalloys: Sixth International Symposium. TMS, 1988, vol. 1988, pp. 305–14.
17.
go back to reference Andreas Kirchmayer, Hao Lyu, Martin Pröbstle, Frédéric Houllé, Andreas Förner, Daniela Huenert, Mathias Göken, Peter Johann Felfer, Erik Bitzek, and Steffen Neumeier: Advanced Engineering Materials, 2020, p. 2000149. Andreas Kirchmayer, Hao Lyu, Martin Pröbstle, Frédéric Houllé, Andreas Förner, Daniela Huenert, Mathias Göken, Peter Johann Felfer, Erik Bitzek, and Steffen Neumeier: Advanced Engineering Materials, 2020, p. 2000149.
18.
go back to reference ASTM E647-13a: ASTM Int., 2014, vol. 13, pp. 1–50. ASTM E647-13a: ASTM Int., 2014, vol. 13, pp. 1–50.
19.
20.
go back to reference F. Long, Y.S. Yoo, C.Y. Jo, S.M. Seo, Y.S. Song, T. Jin, and Z.Q. Hu: Mater. Sci. Eng. A, 2009, vol. 527, pp. 361–69.CrossRef F. Long, Y.S. Yoo, C.Y. Jo, S.M. Seo, Y.S. Song, T. Jin, and Z.Q. Hu: Mater. Sci. Eng. A, 2009, vol. 527, pp. 361–69.CrossRef
21.
go back to reference X. Liu, S. Rangararan, E. Barbero, K.-M. Chang, W.-D. Cao, R. Kennedy, and T. Cameiro: Superalloys 718, 625, 706 and Various Derivatives. TMS, 2001, vol. 2001, pp. 283–90. X. Liu, S. Rangararan, E. Barbero, K.-M. Chang, W.-D. Cao, R. Kennedy, and T. Cameiro: Superalloys 718, 625, 706 and Various Derivatives. TMS, 2001, vol. 2001, pp. 283–90.
22.
go back to reference S. Li, J. Yang, J. Zhuang, Q. Deng, J. Du, X. Xie, B. Li, Z. Xu, Z. Cao, Z. Su, and C. Jiang: in Superalloys 718, 625, 706 and Various Derivatives. TMS, 1994, vol. 1994, pp. 545–55. S. Li, J. Yang, J. Zhuang, Q. Deng, J. Du, X. Xie, B. Li, Z. Xu, Z. Cao, Z. Su, and C. Jiang: in Superalloys 718, 625, 706 and Various Derivatives. TMS, 1994, vol. 1994, pp. 545–55.
23.
go back to reference K. Sadananda and P. Shahinian: Metall. Trans. A, 1977, vol. 8, pp. 439–49.CrossRef K. Sadananda and P. Shahinian: Metall. Trans. A, 1977, vol. 8, pp. 439–49.CrossRef
24.
go back to reference M. Hörnqvist, T. Månsson, and D. Gustafsson: Procedia Eng., 2011, vol. 10, pp. 147–52.CrossRef M. Hörnqvist, T. Månsson, and D. Gustafsson: Procedia Eng., 2011, vol. 10, pp. 147–52.CrossRef
25.
go back to reference W.J. Evans, J.P. Jones, and S. Williams: Int. J. Fatigue, 2005, vol. 27, pp. 1473–484.CrossRef W.J. Evans, J.P. Jones, and S. Williams: Int. J. Fatigue, 2005, vol. 27, pp. 1473–484.CrossRef
26.
go back to reference Timothy P Gabb, Jack Telesman, Anita Garg, Peter Lin, Virgil Provenzano, Robert Heard, and Herbert M Miller: in 7th International Symposium on Superalloy 718 and Derivatives, 2010, p. 26. Timothy P Gabb, Jack Telesman, Anita Garg, Peter Lin, Virgil Provenzano, Robert Heard, and Herbert M Miller: in 7th International Symposium on Superalloy 718 and Derivatives, 2010, p. 26.
27.
28.
29.
go back to reference Lee A James: Eng. Fract. Mech., 1986, vol. 25, pp. 305–14. Lee A James: Eng. Fract. Mech., 1986, vol. 25, pp. 305–14.
30.
go back to reference P Lin, V Provenzano, G Palumbo, T Gabb, and J Telesman: 7th International Symposium on Superalloy 718 and Derivates, 2010, pp. 243–53. P Lin, V Provenzano, G Palumbo, T Gabb, and J Telesman: 7th International Symposium on Superalloy 718 and Derivates, 2010, pp. 243–53.
31.
go back to reference X. Liu, B. Kang, and K.-M. Chang: Mater. Sci. Eng. A, 2003, vol. 340, pp. 8–14.CrossRef X. Liu, B. Kang, and K.-M. Chang: Mater. Sci. Eng. A, 2003, vol. 340, pp. 8–14.CrossRef
32.
go back to reference H.Y. Li, J.F. Sun, M.C. Hardy, H.E. Evans, S.J. Williams, T.J.A. Doel, and P. Bowen: Acta Mater., 2015, vol. 90, pp. 355–69.CrossRef H.Y. Li, J.F. Sun, M.C. Hardy, H.E. Evans, S.J. Williams, T.J.A. Doel, and P. Bowen: Acta Mater., 2015, vol. 90, pp. 355–69.CrossRef
33.
go back to reference Y.J. Oh, J.H. Hong, and S.W. Nam: Metall. Mater. Trans. A, 2000, vol. 31A, pp. 1761–75.CrossRef Y.J. Oh, J.H. Hong, and S.W. Nam: Metall. Mater. Trans. A, 2000, vol. 31A, pp. 1761–75.CrossRef
34.
go back to reference R.C. Reed: The Superalloys: Fundamentals and Applications, 1st ed. Cambridge University Press, Cambridge, 2006.CrossRef R.C. Reed: The Superalloys: Fundamentals and Applications, 1st ed. Cambridge University Press, Cambridge, 2006.CrossRef
35.
go back to reference I.S. Kim, B.G. Choi, H.U. Hong, J. Do, and C.Y. Jo: Mater. Sci. Eng. A, 2014, vol. 593, pp. 55–63.CrossRef I.S. Kim, B.G. Choi, H.U. Hong, J. Do, and C.Y. Jo: Mater. Sci. Eng. A, 2014, vol. 593, pp. 55–63.CrossRef
Metadata
Title
Influence of Grain Size and Volume Fraction of η/δ Precipitates on the Dwell Fatigue Crack Propagation Rate and Creep Resistance of the Nickel-Base Superalloy ATI 718Plus
Authors
A. Kirchmayer
M. Pröbstle
D. Huenert
S. Neumeier
M. Göken
Publication date
10-03-2023
Publisher
Springer US
Published in
Metallurgical and Materials Transactions A / Issue 6/2023
Print ISSN: 1073-5623
Electronic ISSN: 1543-1940
DOI
https://doi.org/10.1007/s11661-023-07001-3

Other articles of this Issue 6/2023

Metallurgical and Materials Transactions A 6/2023 Go to the issue

Premium Partners