Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-19T20:16:55.391Z Has data issue: false hasContentIssue false

The Microstructure and Creep Behavior of Cold Rolled Udimet 188 Sheet

Published online by Cambridge University Press:  23 December 2010

C.J. Boehlert*
Affiliation:
Department of Chemical Engineering and Materials Science, Michigan State University, East Lansing, MI 48824, USA
S.C. Longanbach
Affiliation:
Department of Chemical Engineering and Materials Science, Michigan State University, East Lansing, MI 48824, USA
*
Corresponding author. E-mail: boehlert@egr.msu.edu
Get access

Abstract

Udimet 188 was subjected to thermomechanical processing (TMP) in an attempt to understand the effects of cold-rolling deformation on the microstructure and tensile-creep behavior. Commercially available sheet was cold rolled to varying amounts of deformation (between 5–35% reduction in sheet thickness) followed by a solution treatment at 1,464 K (1,191°C) for 1 h and subsequent air cooling. This sequence was repeated four times to induce a high-volume fraction of low-energy grain boundaries. The resultant microstructure was characterized using electron backscattered diffraction. The effect of the TMP treatment on the high-temperature [1,033–1,088 K (760–815°C)] creep behavior was evaluated. The measured creep stress exponents (6.0–6.8) suggested that dislocation creep was dominant at 1,033 K (760°C) for stresses ranging between 100–220 MPa. For stresses ranging between 25–100 MPa at 1,033 K (760°C), the stress exponents (2.3–2.8) suggested grain boundary sliding was dominant. A significant amount of grain boundary cracking was observed both on the surface and subsurface of deformed samples. To assess the mechanisms of crack nucleation, in situ scanning electron microscopy was performed during the elevated-temperature tensile-creep deformation. Cracking occurred preferentially along general high-angle grain boundaries (GHAB) and less than 25% of the cracks were found on low-angle grain boundaries (LAB) and coincident site lattice boundaries (CSLB). Creep rupture experiments were performed at T = 1,088 K (815°C) and σ = 165 MPa and the greatest average time-to-rupture was exhibited by the TMP sheet with the greatest fraction of LAB+CSLB. However, a clear correlation was not exhibited between the grain boundary character distribution and the minimum creep rates. The findings of this work suggest that although grain boundary engineering may be possible for this alloy, simply relating the fraction of grain boundary types to the creep resistance is not sufficient.

Type
Electron Backscatter Diffraction Special Section
Copyright
Copyright © Microscopy Society of America 2011

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Alexandreanu, B., Capell, B.M. & Was, G. (2001). Combined effect of special grain boundaries and grain boundary carbides on IGSCC. Mater Sci Eng A 300, 94104.CrossRefGoogle Scholar
ASTM Designation E112-96e3. (2004). Standard Test Methods for Determining Average Grain Size. West Conshohocken, PA: American Society for Testing and Materials.Google Scholar
Boehlert, C.J., Dickmann, D.S. & Eisinger, N.C. (2006). The effect of sheet processing on the microstructure, tensile and creep behavior of INCONEL® alloy 718. Metall Mater Trans A 37, 2740.CrossRefGoogle Scholar
Boehlert, C.J., Longanbach, S.C. & Bieler, T.R. (2008a). The effect of thermomechanical processing on the creep behavior of Udimet Alloy 188. Phil Mag 88, 641664.CrossRefGoogle Scholar
Boehlert, C.J., Longanbach, S.C., Nowell, M. & Wright, S. (2008b). The evolution of grain-boundary cracking evaluated through in-situ tensile creep testing of Udimet alloy 188. J Mater Res 23, 500506.CrossRefGoogle Scholar
Brandon, D.G. (1966). On a more restrictive geometric criterion for “special” CSL grain boundaries. Acta Metall 14, 14791484.CrossRefGoogle Scholar
Chen, L.J., Laiw, P.K., He, Y.H., Benson, M.L., Blust, J.W., Browing, P.F., Seeley, R.R. & Klarstrom, D.L. (2001). Tensile hold low-cycle fatigue behavior of cobalt-based Haynes 188 superalloy. Scripta Mater 44, 859865.CrossRefGoogle Scholar
Cheung, C., Erb, U. & Palumbo, G. (1994). Applications of grain boundary engineering concepts to alleviate intergranular cracking in alloys 600 and 690. Mater Sci Eng A 185, 3943.CrossRefGoogle Scholar
Crossman, F.W. & Ashby, M.F. (1975). The non-uniform flow of polycrystals by grain-boundary sliding accommodated by power-law creep. Acta Metall 23, 425440.CrossRefGoogle Scholar
Dave, V.R., Cola, M.J., Kumar, M., Schwartz, A.J. & Hussen, G.N.A. (2004). Grain boundary character in alloy 690 and ductility-dip cracking susceptibility. Welding J 83, 1-S5-S.Google Scholar
Evans, R.W. & Wilshire, B. (1985). Creep of Metals and Alloys, 1st ed., pp. 914. New York: The Institute of Metals.Google Scholar
Herchenroeder, R.B., Matthews, S.J., Tackett, J.W. & Wlodek, S.T. (1972). Haynes alloy No. 188. Cobalt 54, 313.Google Scholar
Hertzberg, R.W. (1996). Deformation and Fracture Mechanics of Engineering Materials, 4th ed.New York: John Wiley and Sons.Google Scholar
Hilliard, J.E. (1964). Estimating grain size by the intercept method. Met Prog 78, 99102.Google Scholar
King, W.E. & Schwartz, A.J. (1998). Towards optimization of the grain boundary character distribution in OFE copper. Scripta Mater 38, 449455.CrossRefGoogle Scholar
Klarstrom, D.L. (1980). Thermomechanical processing of Haynes Alloy No. 188 sheet to improve creep strength. In Superalloys 1980, Tien, J.K., Wlodek, S.T., Morrow, H. III, Gell, M. & Maurer, G.E. (Eds.), pp. 131140. Warrendale, PA: The Materials Society.Google Scholar
Krupp, U., Kane, W.M., Laird, C. & McMahon, C.J. (2004). Brittle intergranular fracture of a Ni-base superalloy at high temperatures by dynamic embrittlement. Mater Sci Eng A 387389, 409413.CrossRefGoogle Scholar
Krupp, U., Kane, W.M., Liu, X., Dueber, O., Laird, C. & McMahon, C.J. (2003). The effect of grain-boundary-engineering type processing on oxygen-induced cracking of IN718. Mater Sci Eng A 349, 213217.CrossRefGoogle Scholar
Langdon, T.G. (1970). Grain boundary sliding as a deformation mechanism during creep. Phil Mag 22, 689700.CrossRefGoogle Scholar
Lee, T.C., Robertson, I.M. & Birnbaum, H.K. (1990). TEM in situ deformation study of the interaction of lattice dislocations with grain boundaries in metals. Phil Mag A 62, 131153.CrossRefGoogle Scholar
Lehockey, E.M. & Palumbo, G. (1997). On the creep behaviour of grain boundary engineered nickel. Mater Sci Eng A 237, 168172.CrossRefGoogle Scholar
Lehockey, E.M., Palumbo, G. & Lin, P. (1998). Improving the weldability and service performance of nickel- and iron-based suparalloys by grain boundary engineering. Metall Mater Trans A 29, 30693079.CrossRefGoogle Scholar
Lehockey, E.M., Palumbo, G., Lin, P. & Brennenstuhl, A.M. (1997). On the relationship between grain boundary character distribution and intergranular corrosion. Scripta Mater 36, 12111218.CrossRefGoogle Scholar
Lin, P., Palumbo, G., Erb, U. & Aust, K.T. (1995). Influence of grain boundary character distribution on sensitization and intergranular corrosion of alloy 600. Scripta Mater 33, 13871392.CrossRefGoogle Scholar
Lissenden, C.J., Colaiuta, J.F. & Lerch, B.A. (2004). Hardening behavior of three metallic alloys under combined stresses at elevated temperature. Acta Mech 169, 5377.CrossRefGoogle Scholar
McGarrity, E.S., Duxbury, P.M. & Holm, E.A. (2005). Statistical physics of grain boundary engineering. Phys Rev E 71, 26102.CrossRefGoogle ScholarPubMed
Palumbo, G., (1998). Metal alloys having improved resistance to intergranular stress corrosion cracking. U.S. Patent 5,817,193.Google Scholar
Palumbo, G. & Aust, K.T. (1989). Special properties of Σ grain boundaries. In Materials Interfaces: Atomic Level Structure and Properties, 1st ed., Wolf, D. & Yip, S. (Eds.), pp. 190211. New York: Chapman and Hall.Google Scholar
Palumbo, G. & Aust, K.T. (1990). Structure-dependence of intergranular corrosion in high purity nickel. Acta Metall Mater 38, 23432352.CrossRefGoogle Scholar
Palumbo, G., Lehockey, E.M. & Lin, P. (1998). Applications for grain boundary engineered materials. J Met 50, 4043.Google Scholar
Thaveeprungsriporn, V. & Was, G.S. (1997). The role of coincidence-site-lattice boundaries in creep of Ni-16Cr-9Fe at 360°C. Metall Mater Trans A 28, 21012112.CrossRefGoogle Scholar
Was, G.S., Thaveeprungsriporn, V. & Crawford, D.C. (1998). Grain boundary misorientation effects on creep and cracking in Ni-based alloys. J Met 50, 4449.Google Scholar
Watanabe, T. (1984). An approach to grain boundary design for strong and ductile polycrystals. Res Mech 11, 4784.Google Scholar
Whittenberger, J.D. (1992). Mechanical properties of Haynes alloy 188 after exposure to LiF-22CaF2, air, and vacuum at 1093K for periods up to 10,000 hours. J Mater Eng Perf 1, 469482.CrossRefGoogle Scholar
Whittenberger, J.D. (1994). Mechanical properties of Haynes alloy 188 after 22,500 hours of exposure to LiF-22CaF2 and vacuum at 1093K. J Mater Eng Perf 3, 754762.CrossRefGoogle Scholar
Zhu, D., Fox, D.S. & Miller, R.A. (2002). Oxidation- and creep-enhanced fatigue of Haynes 188 alloy-oxide scale system under simulated pulse detonation engine conditions. Ceram Eng Sci Proc 23.4, 547553.CrossRefGoogle Scholar