Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-06-23T06:53:55.587Z Has data issue: false hasContentIssue false

Anisotropic Yield Behavior of Lotus-Type Porous Iron: Measurements and Micromechanical Mean-Field Analysis

Published online by Cambridge University Press:  03 March 2011

M. Tane*
Affiliation:
The Institute of Scientific and Industrial Research, Osaka University, Ibaraki 567-0047, Japan
T. Ichitsubo
Affiliation:
Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan
S.K. Hyun
Affiliation:
The Institute of Scientific and Industrial Research, Osaka University, Ibaraki 567-0047, Japan
H. Nakajima
Affiliation:
The Institute of Scientific and Industrial Research, Osaka University, Ibaraki 567-0047, Japan
*
a)Address all correspondence to this author. e-mail: mtane@sanken.osaka-u.ac.jp
Get access

Abstract

Anisotropic yield behavior of lotus-type porous iron fabricated using the continuous-zone-melting method in a pressurized nitrogen and hydrogen atmosphere has been investigated. The 0.2% offset strength (compressive yield stress) in the loading direction parallel to the longitudinal axis of pores decreases linearly with increasing porosity, while the perpendicular strength decreases steeply. The strength versus porosity curves can be expressed using a well-known power law formula. In addition, the 0.2% offset strength of lotus iron prepared in a nitrogen and hydrogen atmosphere is found to be larger than that of lotus iron prepared in a hydrogen and helium atmosphere, which is attributed to the solid solution hardening by the solute nitrogen. Furthermore, we compare the experimental results to calculations obtained by means of the extended Qiu-Weng’s mean-field method, and the comparison suggests that local stresses deviated from the average stress are dominant in the macroscopic yield behavior of lotus metals.

Type
Articles
Copyright
Copyright © Materials Research Society 2004

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

1Gibson, L.J. and Ashby, M.F.: Cellular Solids, 2nd ed. (Cambridge University Press, Cambridge, U.K., 1997).CrossRefGoogle Scholar
2Banhart, J. and Weaire, D.: On the road again: Metal foams find favor. Phys. Today 55, 37 (2002).Google Scholar
3Nakajima, H.: Fabrication of lotus-structured porous metals. Prod. Tech. 51, 60 (1999).Google Scholar
4Nakajima, H.: Pore-aligned porous metals. J. High. Temp. Soc. 26, 9 (2000).Google Scholar
5Nakajima, H.: Light and strong porous metals. Funct. Mater. 20, 27 (2000).Google Scholar
6Nakajima, H., Hyun, S.K., Ohashi, K., Ota, K. and Murakami, K.: Fabrication of porous copper by unidirectional solidification under hydrogen and its properties. Colloids Surfaces A 179, 209 (2001).Google Scholar
7Hyun, S.K., Murakami, K. and Nakajima, H.: Anisotropic mechanical properties of porous copper fabricated by unidirectional solidification. Mater. Sci. Eng. A 299, 241 (2001).Google Scholar
8Ichitsubo, T., Tane, M., Ogi, H., Hirao, M., Ikeda, T. and Nakajima, H.: Anisotropic elastic constants of lotus-type porous copper: measurements and micromechanics modeling. Acta Mater. 50, 4105 (2002).CrossRefGoogle Scholar
9Tane, M., Ichitsubo, T., Hirao, M., Ikeda, T. and Nakajima, H.: Elastic constants of lotus-type porous magnesium: Comparison with effective-mean-field theory. J. Appl. Phys. 96, 3696 (2004).Google Scholar
10Tane, M., Ichitsubo, T., Hyun, S.K., Nakajima, H. and Hirao, M.: Elastic properties of lotus-type porous iron: Acoustic measurement and extended effective-mean-field theory. Acta Mater. 52, 5195 (2004).CrossRefGoogle Scholar
11Qiu, Y.P. and Weng, G.J.: A theory of plasticity for porous materials and particle-reinforced composites. J. Appl. Mech. Trans. ASME 59, 261 (1992).Google Scholar
12Tane, M., Ichitsubo, T., Hirao, M. and Nakajima, H.: Extended mean-field method for predicting yield behaviors of porous materials. Submitted to Mech. Mater.Google Scholar
13Nakajima, H., Ikeda, T. and Hyun, S.K.: Fabrication of lotus-type porous metals and their physical properties. Adv. Eng. Mater. 6, 377 (2004).Google Scholar
14Gurson, A.L.: Continuum theory of ductile rupture by void nucleation and growth. 1. Yield criteria and flow rules for porous ductile media. J. Eng. Mater. Technol. ASME 99, 2 (1977).CrossRefGoogle Scholar
15Tvergaad, V.: Influence of void nucleation on ductile shear fracture at a free-surface. J. Mech. Phys. Solids 30, 399 (1982).Google Scholar
16Castaneda, P.P.: The effective mechanical-properties of nonlinear isotropic composites. J. Mech. Phys. Solids 39, 45 (1991).CrossRefGoogle Scholar
17Qiu, Y.P. and Weng, G.J.: An energy approach to the plasticity of a two-phase composite containing aligned inclusions. J. Appl. Mech. Trans. ASME 62, 1039 (1995).Google Scholar
18Qiu, Y.P. and Weng, G.J.: Plastic potential and yield function of porous materials with aligned and randomly oriented spheroidal voids. J. Plast. 9, 271 (1993).Google Scholar
19Dunn, M.L. and Ledbetter, H.: Elastic-plastic behavior of textured short-fiber composites. Acta Mater. 45, 3327 (1997).CrossRefGoogle Scholar
20Mori, T. and Tanaka, K.: Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metall. 21, 573 (1973).CrossRefGoogle Scholar
21Eshelby, J.D.: The determination of the elastic field of an ellipsoidal inclusion. Proc. R. Soc. London A 241, 376 (1957).Google Scholar
22Benveniste, Y.: A new approach to the application of mori-tanaka theory in composite-materials. Mech. Mater. 6, 147 (1987).CrossRefGoogle Scholar