Log in

A closed-form yield criterion for porous materials with Mises–Schleicher–Burzyński matrix containing cylindrical voids

  • Original Paper
  • Published:
Acta Mechanica Aims and scope Submit manuscript

Abstract

This work develops a closed-form yield criterion applicable to porous materials with pressure-dependent matrix presenting tension–compression asymmetry (Mises–Schleicher–Burzyński material) containing parallel cylindrical voids. To develop the strength criterion, the stress-based variational homogenization approach due to Cheng et al. (Int J Plast 55:133–151, 2014) is extended to the case of a hollow cylinder under generalized plane strain conditions subjected to axisymmetric loading. Adopting a strictly statically admissible trial stress field, the homogenization procedure results in an approximate yield locus depending on the current material porosity, tension–compression material asymmetry, the mean lateral stress, and an equivalent shear stress. The analytical criterion provides exact solutions for purely hydrostatic loading. Theoretical results are compared with finite element (FE) simulations considering cylindrical unit-cells with distinct porosity levels, different values of the tension–compression asymmetry, and a wide range of stress triaxialities. Based on comparisons, the theoretical results are found to be in good agreement with FE simulations for most of the loading conditions and material features considered in this study. More accurate theoretical predictions are provided when higher material porosities and/or lower tension–compression asymmetries are considered. Overall, the main outcome of this work is a closed-form yield function proving fairly accurate predictions to engineering applications, in which pressure-dependent and tension–compression asymmetric porous materials with cylindrical voids are dealt with. This can be the case of honeycomb structures or additively manufactured materials, in which metal matrix composites are employed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Canada)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  1. ABAQUS/Standard: Simulia, User’s Manual. Dassault Systémes, Providence, USA. version 6.19 edition (2019)

  2. Benzerga, A.A., Leblond, J.B.: Ductile fracture by void growth to coalescence. In: Aref, H., van der Giessen, E. (eds.) Advances in Applied Mechanics, vol. 44, pp. 169–305. Elsevier, Amsterdam (2010)

    Google Scholar 

  3. Burzyński, W.: Ueber die Anstrengungshypothesen. Schweiz. Bauzeitung 94, 259–262 (1929)

    Google Scholar 

  4. Castañeda, P.P.: Nonlinear Composite Materials: Effective Constitutive Behavior and Microstructure Evolution, pp. 131–195. Springer, Vienna (1997)

    MATH  Google Scholar 

  5. Cazacu, O., Plunkett, B., Barlat, F.: Orthotropic yield criterion for hexagonal closed packed metals. Int. J. Plast. 22, 1171–1194 (2006)

    MATH  Google Scholar 

  6. Cazacu, O., Stewart, J.B.: Analytic plastic potential for porous aggregates with matrix exhibiting tension-compression asymmetry. J. Mech. Phys. Solids 57, 325–341 (2009)

    MathSciNet  MATH  Google Scholar 

  7. Cheng, L., Guo, T.: Void interaction and coalescence in polymeric materials. Int. J. Solids Struct. 44, 1787–1808 (2007)

    MATH  Google Scholar 

  8. Cheng, L., de Saxcé, G., Kondo, D.: A stress-based variational model for ductile porous materials. Int. J. Plast. 55, 133–151 (2014)

    Google Scholar 

  9. Coussy, O.: Poromechanics. Wiley, Hoboken (2004)

    MATH  Google Scholar 

  10. Coussy, O.: Mechanics and Physics of Porous Solids. Wiley, Hoboken (2011)

    Google Scholar 

  11. Dormieux, L., Lemarchand, E., Kondo, D., Brach, S.: Strength criterion of porous media: application of homogenization techniques. J. Rock Mech. Geotech. Eng. 9, 62–73 (2017)

    Google Scholar 

  12. Durban, D., Cohen, T., Hollander, Y.: Plastic response of porous solids with pressure sensitive matrix. Mech. Res. Commun. 37, 636–641 (2010)

    MATH  Google Scholar 

  13. Eve, R.A., Reddy, B.D., Rockafellar, R.T.: An internal variable theory of elastoplasticity based on the maximum plastic work inequality. Q. Appl. Math. 48, 59–83 (1990)

    MathSciNet  MATH  Google Scholar 

  14. Fritzen, F., Forest, S., Kondo, D., Böhlke, T.: Computational homogenization of porous materials of Green type. Comput. Mech. 52, 121–134 (2013)

    MathSciNet  MATH  Google Scholar 

  15. Gologanu, M., Leblond, J.B., Devaux, J.: Approximate models for ductile metals containing non-spherical voids-case of axisymmetric prolate ellipsoidal cavities. J. Mech. Phys. Solids 41, 1723–1754 (1993)

    MATH  Google Scholar 

  16. Guo, T., Faleskog, J., Shih, C.: Continuum modeling of a porous solid with pressure-sensitive dilatant matrix. J. Mech. Phys. Solids 56, 2188–2212 (2008)

    MATH  Google Scholar 

  17. Gurson, A.: Continuum theory of ductile rupture by void nucleation and growth. Part I: yield criteria and flow rules for porous ductile media. ASME J. Eng. Mater. Technol. 99, 2–15 (1977)

    Google Scholar 

  18. Halphen, B., Son Nguyen, Q.: Sur les matériaux standard généralisés. J. de Mécanique 14, 39–63 (1975)

    MathSciNet  MATH  Google Scholar 

  19. Han, B., Shen, W., **e, S., Shao, J.: Plastic modeling of porous rocks in drained and undrained conditions. Comput. Geotech. 117, 103277 (2020)

    Google Scholar 

  20. Hill, R.: A theory of the yielding and plastic flow of anisotropic metals. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 193, 281–297 (1948)

    MathSciNet  MATH  Google Scholar 

  21. Keralavarma, S., Benzerga, A.: A constitutive model for plastically anisotropic solids with non-spherical voids. J. Mech. Phys. Solids 58, 874–901 (2010)

    MathSciNet  MATH  Google Scholar 

  22. Keralavarma, S., Benzerga, A.: Numerical assessment of an anisotropic porous metal plasticity model. Mechanics of Materials 90, 212–228. Proceedings of the IUTAM Symposium on Micromechanics of Defects in Solids (2015)

  23. Leblond, J., Perrin, G., Suquet, P.: Exact results and approximate models for porous viscoplastic solids. Int. J. Plast. 10, 213–235 (1994)

    MATH  Google Scholar 

  24. Lee, J., Oung, J.: Yield functions and flow rules for porous pressure-dependent strain-hardening polymeric materials. J. Appl. Mech. 67, 288–297 (2000)

    MATH  Google Scholar 

  25. Martin, J.H., Yahata, B.D., Clough, E.C., Mayer, J.A., Hundley, J.M., Schaedler, T.A.: Additive manufacturing of metal matrix composites via nanofunctionalization. MRS Commun. 8, 297–302 (2018)

    Google Scholar 

  26. McClintock, F.A.: A criterion for ductile fracture by the growth of holes. J. Appl. Mech. 35, 363–371 (1968)

    Google Scholar 

  27. Molinari, A., Mercier, S.: Micromechanical modelling of porous materials under dynamic loading. J. Mech. Phys. Solids 49, 1497–1516 (2001)

    MATH  Google Scholar 

  28. Monchiet, V., Cazacu, O., Charkaluk, E., Kondo, D.: Macroscopic yield criteria for plastic anisotropic materials containing spheroidal voids. Int. J. Plast. 24, 1158–1189 (2008)

    MATH  Google Scholar 

  29. Monchiet, V., Charkaluk, E., Kondo, D.: Macroscopic yield criteria for ductile materials containing spheroidal voids: an Eshelby-like velocity fields approach. Mech. Mater. 72, 1–18 (2014)

    Google Scholar 

  30. Monchiet, V., Kondo, D.: Exact solution of a plastic hollow sphere with a Mises–Schleicher matrix. Int. J. Eng. Sci. 51, 168–178 (2012)

    MathSciNet  MATH  Google Scholar 

  31. Moreau, J.: Application of convex analysis to the treatment of elastoplastic systems, in: Germain, P., Nayroles, B. (Eds.), Applications of Methods of Functional Analysis to Problems in Mechanics. Springer Berlin / Heidelberg. volume 503 of Lecture Notes in Mathematics, pp. 56–89. https://doi.org/10.1007/BFb0088746 (1976)

  32. Pastor, F., Anoukou, K., Pastor, J., Kondo, D.: Limit analysis and homogenization of porous materials with Mohr-Coulomb matrix. Part ii: Numerical bounds and assessment of the theoretical model. J. Mech. Phys. Solids 91, 14–27 (2016)

    MathSciNet  Google Scholar 

  33. Pastor, F., Kondo, D., Pastor, J.: 3d-fem formulations of limit analysis methods for porous pressure-sensitive materials. Int. J. Numer. Methods Eng. 95, 847–870 (2013)

    MathSciNet  MATH  Google Scholar 

  34. Pastor, F., Kondo, D., Pastor, J.: Limit analysis and computational modeling of the hollow sphere model with a Mises–Schleicher matrix. Int. J. Eng. Sci. 66–67, 60–73 (2013)

    MathSciNet  MATH  Google Scholar 

  35. Rice, J., Tracey, D.: On the ductile enlargement of voids in triaxial stress fields. J. Mech. Phys. Solids 17, 201–217 (1969)

    Google Scholar 

  36. Schleicher, F.: Der Spannungszustand an der Fließgrenze (Plastizitätsbedingung). ZAMM J. Appl. Math. Mech. Z. für Angew. Math. und Mech. 6, 199–216 (1926). https://doi.org/10.1002/zamm.19260060303

    Article  MATH  Google Scholar 

  37. Shen, W., Oueslati, A., De Saxcé, G.: Macroscopic criterion for ductile porous materials based on a statically admissible microscopic stress field. Int. J. Plast. 70, 60–76 (2015)

    Google Scholar 

  38. Shen, W., Shao, J.: Some micromechanical models of elastoplastic behaviors of porous geomaterials. J. Rock Mech. Geotech. Eng. 9, 1–17 (2017)

    Google Scholar 

  39. Shen, W., Shao, J.: A micro-mechanics-based elastic-plastic model for porous rocks: applications to sandstone and chalk. Acta Geotech. 13, 329–340 (2018)

    Google Scholar 

  40. Shen, W., Shao, J., Dormieux, L., Kondo, D.: Approximate criteria for ductile porous materials having a Green type matrix: application to double porous media. Comput. Mater. Sci. 62, 189–194 (2012)

    Google Scholar 

  41. Shen, W., Shao, J., Kondo, D., de Saxcé, G.: A new macroscopic criterion of porous materials with a Mises–Schleicher compressible matrix. Eur. J. Mech. A/Solids 49, 531–538 (2015)

    Google Scholar 

  42. Shen, W.Q., Shao, J.F., Liu, Z.B., Oueslati, A., De Saxcé, G.: Evaluation and improvement of macroscopic yield criteria of porous media having a Drucker-Prager matrix. Int. J. Plast. 126, 102609 (2020). https://doi.org/10.1016/j.ijplas.2019.09.015

    Article  Google Scholar 

  43. Shen, W., Shao, J., Oueslati, A., de Saxcé, G., Zhang, J.: An approximate strength criterion of porous materials with a pressure sensitive and tension-compression asymmetry matrix. Int. J. Eng. Sci. 132, 1–15 (2018)

    MathSciNet  MATH  Google Scholar 

  44. Shen, W., Zhang, J., Shao, J., Kondo, D.: Approximate macroscopic yield criteria for Drucker-Prager type solids with spheroidal voids. Int. J. Plast. 99, 221–247 (2017)

    Google Scholar 

  45. Subramani, M., Czarnota, C., Mercier, S., Molinari, A.: Dynamic response of ductile materials containing cylindrical voids. Int. J. Fract. 122, 197–218 (2020). https://doi.org/10.1007/s10704-020-00441-7

    Article  Google Scholar 

  46. Suquet, P.: Effective Properties of Nonlinear Composites, pp. 197–264. Springer, Vienna (1997)

    MATH  Google Scholar 

  47. Suquet, P.M.: Elements of homogenization for inelastic solid mechanics, homogenization techniques for composite media. Lect. Notes Phys. 272, 193 (1985)

    Google Scholar 

  48. Trillat, M., Pastor, J.: Limit analysis and Gurson’s model. Eur. J. Mech. A/Solids 24, 800–819 (2005)

    MATH  Google Scholar 

  49. Tvergaard, V.: Influence of voids on shear band instabilities under plane strain conditions. Int. J. Fract. 17, 389–407 (1981)

    Google Scholar 

  50. Tvergaard, V., Needleman, A.: Analysis of the cup-cone fracture in a round tensile bar. Acta Metall. 32, 157–169 (1984)

    Google Scholar 

  51. Vadillo, G., Fernández-Sáez, J.: An analysis of Gurson model with parameters dependent on triaxiality based on unitary cells. Eur. J. Mech. A/Solids 28, 417–427 (2009)

    MATH  Google Scholar 

  52. Vadillo, G., Fernández-Sáez, J., Pȩcherski, R.: Some applications of Burzyński yield condition in metal plasticity. Mater. Des. 32, 628–635 (2011)

    Google Scholar 

  53. Yi, S., Duo, W.: A lower bound approach to the yield loci of porous materials. Acta Mech. Sin. 5, 237–243 (1989)

    Google Scholar 

  54. Zhang, H., Ramesh, K., Chin, E.: A multi-axial constitutive model for metal matrix composites. J. Mech. Phys. Solids 56, 2972–2983 (2008)

    MATH  Google Scholar 

  55. Zhang, J., Shen, W., Oueslati, A., de Saxcé, G.: Shakedown of porous materials. Int. J. Plast. 95, 123–141 (2017)

    Google Scholar 

Download references

Acknowledgements

TdS wishes to acknowledge the financial support of FAPERGS, Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul, grant agreement 19/2551– 0001054– 0. GV want to acknowledge the financial support provided by the Spanish Ministry of Science and Innovation under Project references DPI2017-88608-P (Proyectos I+D Excelencia 2017) and EIN2019-103276 (Acciones de Dinamización Europa Investigación).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tiago dos Santos.

Additional information

Publisher's Note

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

Appendices

Appendix A: Development of the first stress field \(\varvec{\sigma }_{1}\)

To develop the first trial stress field \(\varvec{\sigma }_{1}\), the analytical development presented by Monchiet and Kondo [30] is considered. They have proposed an exact solution for a plastic hollow sphere with a Mises–Schleicher [36] matrix material. For a plastic loading, the yield criterion has to satisfy the condition (see Eq. (3)):

$$\begin{aligned} \sigma _{e}^{2}+3\alpha \sigma _{0}\sigma _{h}-\sigma _{0}^{2}=0 \end{aligned}$$
(A.1)

where, for a cylindrically symmetric problem, the von Mises equivalent stress and the hydrostatic stress are given, respectively, by:

$$\begin{aligned} \sigma _{e}=\sqrt{\frac{3}{4}\left( \sigma _{\theta }-\sigma _{r}\right) ^{2}+\frac{9}{4}\left( \sigma _{z}-\sigma _{h}\right) ^{2}}\,\,\text { and }\,\,\sigma _{h}=\frac{1}{3}\left( \sigma _{r}+\sigma _{\theta }+\sigma _{z}\right), \end{aligned}$$
(A.2)

being \(\sigma _{r}\), \(\sigma _{\theta }\), and \(\sigma _{z}\) the radial, circumferential, and longitudinal stresses. For the first trial stress field \(\varvec{\sigma }_{1}\) it will be assumed that \(\sigma _{z}=\sigma _{h}\), thus \(\sigma _{e}\) and \(\sigma _{h}\) become:

$$\begin{aligned} \sigma _{e}=\sqrt{\frac{3}{4}}\left( \sigma _{\theta }-\sigma _{r}\right) \epsilon \,\,\text { and }\,\,\sigma _{h}=\frac{1}{2}\left( \sigma _{r}+\sigma _{\theta }\right) \end{aligned}$$
(A.3.1,2)

where \(\epsilon =\mathrm {sign}\left( \sigma _{\theta }-\sigma _{r}\right)\) is the sign of \(\left( \sigma _{\theta }-\sigma _{r}\right)\). It is worth mentioning that, in addition to simplify the expression of the equivalent stress \(\sigma _{e}\), assuming that \(\sigma _{z}=\sigma _{h}\) results in a purely hydrostatic macroscopic stress field (see Eq. (28)).

Following the development of Monchiet and Kondo [30], a positive function \(G\left( r\right)\), depending on the radial coordinate r, is introduced in a manner that:

$$\begin{aligned} \sigma _{e}=\sigma _{0}G\left( r\right). \end{aligned}$$
(A.4)

Therefore, using Eq. (A.1), the hydrostatic stress reads:

$$\begin{aligned} \sigma _{h}=\frac{\sigma _{0}}{3\alpha }\left[ 1-G^{2}\left( r\right) \right]. \end{aligned}$$
(A.5)

Moreover, combining Eqs. (A.3), (A.4), (A.5), the solution in terms of \(\sigma _{r}\) and \(\sigma _{\theta }\) results:

$$\begin{aligned}&\sigma _{r}=\frac{\sigma _{0}}{3\alpha }\left[ 1-G^{2}\left( r\right) \right] -\sqrt{\frac{1}{3}}\sigma _{0}G\left( r\right) \epsilon, \end{aligned}$$
(A.6)
$$\begin{aligned}&\sigma _{\theta }=\frac{\sigma _{0}}{3\alpha }\left[ 1-G^{2}\left( r\right) \right] +\sqrt{\frac{1}{3}}\sigma _{0}G\left( r\right) \epsilon. \end{aligned}$$
(A.7)

Introducing the last two equations into Eq. (9) yields:

$$\begin{aligned} G\left( r\right) \frac{\mathrm{d}G\left( r\right) }{\mathrm{d}r}+\frac{A}{2}\frac{dG\left( r\right) }{\mathrm{d}r}+\frac{A}{r}G\left( r\right) =0 \end{aligned}$$
(A.8)

where \(A=\sqrt{3}\alpha \epsilon\). The solution of the differential equation (A.8) is:

$$\begin{aligned} G\left( r\right) =\frac{A}{2}W\left( \frac{2\exp \left( \frac{C_{1}}{A}\right) }{Ar^{2}}\right) \end{aligned}$$

or

$$\begin{aligned} G\left( r\right) =\frac{\sqrt{3}\alpha \epsilon }{2}W\left( p\frac{a^{2}}{r^{2}}\right) \end{aligned}$$
(A.9)

where \(W\left( x\right)\) is the Lambert W function, which is the inverse of \(x=W\exp \left( W\right)\). Function p is defined as:

$$\begin{aligned} p=\frac{2\exp \left( \frac{C_{1}}{\sqrt{3}\alpha \epsilon }\right) }{\sqrt{3}\alpha \epsilon a^{2}}. \end{aligned}$$
(A.10)

Thus, parameter p has both positive \(\left( p_{+}\right)\) and negative \(\left( p_{-}\right)\) branches:

$$\begin{aligned} p={\left\{ \begin{array}{ll} p_{+}=\frac{2\exp \left( \frac{C_{1}}{\sqrt{3}\alpha }\right) }{\sqrt{3}\alpha a^{2}}\\ p_{-}=-\frac{2\exp \left( -\frac{C_{1}}{\sqrt{3}\alpha }\right) }{\sqrt{3}\alpha a^{2}} \end{array}\right. }. \end{aligned}$$
(A.11)

Since \(G\left( r\right) \ge 0\), from Eqs. (A.9) and (A.11), it is concluded that \(\mathrm {sign}\left( W\right) =\mathrm {sign}\left( p\right)\) and thus \(\epsilon =\mathrm {sign}\left( p\right)\).

Given the solution (A.9), the stress components become (see Eqs. (A.6) and (A.7)):

$$\begin{aligned}&\sigma _{r}=\frac{\sigma _{0}}{3\alpha }\left[ 1-\frac{3\alpha ^{2}}{4}W^{2}\left( p\frac{a^{2}}{r^{2}}\right) -\frac{3\alpha ^{2}}{2}W\left( p\frac{a^{2}}{r^{2}}\right) \right], \end{aligned}$$
(A.12)
$$\begin{aligned}&\sigma _{\theta }=\frac{\sigma _{0}}{3\alpha }\left[ 1-\frac{3\alpha ^{2}}{4}W^{2}\left( p\frac{a^{2}}{r^{2}}\right) +\frac{3\alpha ^{2}}{2}W\left( p\frac{a^{2}}{r^{2}}\right) \right]. \end{aligned}$$
(A.13)

From Eq. (A.3.2), the hydrostatic and axial stresses are then calculated:

$$\begin{aligned} \sigma _{h}=\sigma _{z}=\frac{\sigma _{0}}{3\alpha }\left[ 1-\frac{3\alpha ^{2}}{4}W^{2}\left( p\frac{a^{2}}{r^{2}}\right) \right] \end{aligned}$$
(A.14)

Coefficient p can be determined from the boundary condition \(\sigma _{r}\left( r=a\right) =0\). Thus, from Eq. (A.12):

$$\begin{aligned} \alpha ^{2}W^{2}\left( p\right) +2\alpha ^{2}W\left( p\right) -\frac{4}{3}=0. \end{aligned}$$
(A.15)

The corresponding roots of the last equation are:

$$\begin{aligned} W\left( p\right) =\frac{-\alpha \pm \sqrt{\alpha ^{2}+\frac{4}{3}}}{\alpha }. \end{aligned}$$

Thus, both the positive \(\left( p_{+}\right)\) and negative \(\left( p_{-}\right)\) branches of p are determined:

$$\begin{aligned} {\left\{ \begin{array}{ll} p=p_{+}=z_{+}\exp \left( z_{+}\right) , &{} z_{+}=\frac{-\alpha +\sqrt{\alpha ^{2}+\frac{4}{3}}}{\alpha }\\ \mathrm {or}\\ p=p_{-}=z_{-}\exp \left( z_{-}\right) , &{} z_{-}=\frac{-\alpha -\sqrt{\alpha ^{2}+\frac{4}{3}}}{\alpha } \end{array}\right. } . \end{aligned}$$
(A.16)

Appendix B: Development of the second stress field \(\varvec{\sigma }_{2}\)

To develop the second trial stress field \(\varvec{\sigma }_{2}\), a homogeneous longitudinal stress state is considered:

$$\begin{aligned} \varvec{\sigma }_{2}=\sigma _{z}\varvec{e}_{z}\otimes \varvec{e}_{z}. \end{aligned}$$
(B.1)

Since \(\sigma _{z}\) is constant, the stress tensor \(\varvec{\sigma }_{2}\) readily satisfies the equilibrium equation \(\mathrm {div}\varvec{\sigma }_{2}=\varvec{0}\). For this particular stress tensor, the von Mises equivalent stress and the hydrostatic stress become, respectively:

$$\begin{aligned} \sigma _{e}=\sigma _{z}\mathrm {sign}\left( \sigma _{z}\right) \,\,\text { and }\,\,\sigma _{h}=\frac{1}{3}\sigma _{z}. \end{aligned}$$
(B.2)

Therefore, the yield condition yields (see Eq. (3)):

$$\begin{aligned} \sigma _{z}^{2}+\alpha \sigma _{0}\sigma _{z}-\sigma _{0}^{2}=0. \end{aligned}$$
(B.3)

Thus, solving the previous equation in terms of the longitudinal stress, we obtain:

$$\begin{aligned} \sigma _{z}=\left( -\alpha \pm \sqrt{\alpha ^{2}+4}\right) \frac{\sigma _{0}}{2}. \end{aligned}$$
(B.4)

Since \(\sigma _{0}>0\), the term \(\left( -\alpha \pm \sqrt{\alpha ^{2}+4}\right)\) defines the sign of the longitudinal stress \(\sigma _{z}\). Notice in Eq. (B.4) that the solution depends on the tension–compression asymmetry by means of parameter \(\alpha\) (or k, see Eq. (2)).

Appendix C: Development of the macroscopic yield function \(\Phi \left( \varvec{\Sigma }\right)\)

This Section is intended to present the development leading to the macroscopic yield function \(\Phi \left( \varvec{\Sigma }\right)\). Starting from condition (19), having in mind the matrix yield function (3), the following relation is obtained:

$$\begin{aligned} \Phi \left( \varvec{\Sigma }\right) =\frac{1}{\left| \Omega \right| }\int _{a}^{b}\int _{-\frac{H}{2}}^{\frac{H}{2}}\int _{0}^{2\pi }\left( \sigma _{e}^{2}+3\alpha \sigma _{0}\sigma _{h}-\sigma _{0}^{2}\right) rd\theta \mathrm{d}z\mathrm{d}r. \end{aligned}$$
(C.1)

In view of the stress superposition (21) and the trial stress fields given in Eqs. (22)–(24) and (30), the first term on the right-hand side of Eq. (C.1) can be integrated as follows:

$$\begin{aligned} \frac{1}{\left| \Omega \right| }\int _{-\frac{H}{2}}^{\frac{H}{2}}\int _{0}^{2\pi }\int _{a}^{b}\sigma _{e}^{2}r\mathrm{d}r\mathrm{d}\theta \mathrm{d}z&=\frac{1}{\left| \Omega \right| }\int _{-\frac{H}{2}}^{\frac{H}{2}}\int _{0}^{2\pi }\int _{a}^{b}\left[ A_{1}^{2}\frac{3\alpha ^{2}}{4}W^{2}\left( p\frac{a^{2}}{r^{2}}\right) +\frac{9}{4}\left( \frac{2}{3}A_{2}\right) ^{2}\right] r\mathrm{d}r\mathrm{d}\theta \mathrm{d}z\nonumber \\&=-\left( f+\frac{3\alpha ^{2}}{4}\Upsilon \right) A_{1}^{2}+A_{2}^{2}\left( 1-f\right) \end{aligned}$$
(C.2)

in which relations \(\left| \Omega \right| =H\pi b^{2}\), \(f=\frac{a^{2}}{b^{2}}\) and condition (A.15) have been employed. Moreover, parameter \(\Upsilon\) is calculated using Eq. (29).

Moreover, also using Eqs. (21), (22)–(24), and (30), the second term on the right-hand side of Eq. (C.1) is integrated:

$$\begin{aligned} \frac{3\alpha \sigma _{0}}{\left| \Omega \right| }\int _{-\frac{H}{2}}^{\frac{H}{2}}\int _{0}^{2\pi }\int _{a}^{b}\sigma _{h}r\mathrm{d}r\mathrm{d}\theta \mathrm{d}z&=\frac{2\pi H\sigma _{0}A_{1}}{\Omega }\int _{a}^{b}\left[ 1-\frac{3\alpha ^{2}}{4}W^{2}\left( p\frac{a^{2}}{r^{2}}\right) \right] r\mathrm{d}r+\frac{2\pi H\alpha \sigma _{0}A_{2}}{\Omega }\int _{a}^{b}rdr\nonumber \\&=\sigma _{0}\left( 1+\frac{3}{4}\alpha ^{2}\Upsilon \right) A_{1}+\left( 1-f\right) \alpha \sigma _{0}A_{2} \end{aligned}$$
(C.3)

where relations \(\left| \Omega \right| =H\pi b^{2}\), \(f=\frac{a^{2}}{b^{2}}\), and Eq. (A.15) have been used again.

Finally, the last term on the right hand side of Eq. (C.1) can be easily integrated:

$$\begin{aligned} \frac{1}{\left| \Omega \right| }\int _{-\frac{H}{2}}^{\frac{H}{2}}\int _{0}^{2\pi }\int _{a}^{b}\sigma _{0}^{2}rdrd\theta \mathrm{d}z=\left( 1-f\right) \sigma _{0}^{2}. \end{aligned}$$
(C.4)

Therefore, using Eqs. (C.2)–(C.4), Eq. (C.1) becomes:

$$\begin{aligned} \Phi \left( \varvec{\Sigma }\right) =-\left( f+\frac{3\alpha ^{2}}{4}\Upsilon \right) A_{1}^{2}+A_{2}^{2}\left( 1-f\right) +\sigma _{0}\left( 1+\frac{3}{4}\alpha ^{2}\Upsilon \right) A_{1}+\left( 1-f\right) \alpha \sigma _{0}A_{2}-\left( 1-f\right) \sigma _{0}^{2}=0. \end{aligned}$$
(C.5)

Appendix D: Reference criteria for porous materials with cylindrical voids

This Section aims at summarizing yield criteria that have been proposed in the literature to porous materials with cylindrical voids. Those criteria will be considered in this work for comparison purposes. The first one is the well-known Gurson [17] criterion:

$$\begin{aligned} \Phi _{G}=\frac{\Sigma _{e}^{2}}{\sigma _{0}^{2}}+2f\cosh \left( \frac{\sqrt{3}\Sigma _{m}}{\sigma _{0}}\right) -\left( 1+f^{2}\right) =0 \end{aligned}$$
(D.1)

where \(\Sigma _{e}\) is the von Mises equivalent stress and \(\Sigma _{m}=\nicefrac {\left( \Sigma _{1}+\Sigma _{2}\right) }{2}\) denotes the mean lateral stress. In the specific axisymmetric case of the cylinder shown in Fig. 1c, we have \(\Sigma _{m}=\Sigma _{1}=\Sigma _{2}\) and \(\Sigma _{e}=\left| \Sigma _{3}-\Sigma _{1}\right|\), that is, the von Mises equivalent stress has the absolute value of the equivalent shear stress \(\Sigma _{sh}\) defined in Eq. (32). For a purely hydrostatic loading \(\left( \Sigma _{e}=\Sigma _{sh}=0\right)\), Eq. (D.1) provides the well-known exact solution:

$$\begin{aligned} \left| \Sigma _{m}\right| =\frac{\sigma _{0}}{\sqrt{3}}\ln \left( f\right) \end{aligned}$$
(D.2)

where the identity \(\mathrm {arcosh}\left( x\right) =\ln \left( x+\sqrt{x^{2}-1}\right)\), for \(x\ge 1\), has been used. Notice that, for a purely hydrostatic case, we have \(\Sigma _{m}=\Sigma _{h}\), being \(\Sigma _{h}\) the hydrostatic stress.

The second one is the upper bound that has been developed by Lee and Oung [24] employing Gurson’s kinematic approach to porous materials with Mises–Schleicher matrix material:

$$\begin{aligned} \Phi _{LO}^{up}=\frac{\Sigma _{sh}^{2}}{\sigma _{0}^{2}}+3f\frac{\Sigma _{m}^{2}}{\sigma _{0}^{2}}+\alpha \left( 1-f\right) \frac{\left( 3\Sigma _{m}+\Sigma _{sh}\right) }{\sigma _{0}}-\left( 1-f\right) ^{2}=0 \end{aligned}$$
(D.3)

where relation \(\Sigma _{e}=\pm \Sigma _{sh}\), between the equivalent von Mises stress \(\Sigma _{e}\) and the equivalent shear stress \(\Sigma _{sh}\) (see Eq. (32)), has been employed in order to have the same stress components as shown in Eq. (35). For a purely hydrostatic stress state \(\left( \Sigma _{sh}=0\right)\), the previous criterion becomes:

$$\begin{aligned} 3f\frac{\Sigma _{m}^{2}}{\sigma _{0}^{2}}+3\alpha \left( 1-f\right) \frac{\Sigma _{m}}{\sigma _{0}}-\left( 1-f\right) ^{2}=0. \end{aligned}$$
(D.4)

In addition, when the special case with \(\alpha \rightarrow 0\) is considered, it results in:

$$\begin{aligned} \left| \Sigma _{m}\right| =\frac{\sigma _{0}}{\sqrt{3}}\frac{\left( 1-f\right) }{\sqrt{f}} \end{aligned}$$
(D.5)

which differs from the exact solution (D.2). In order to recover Gurson’s model when \(\alpha \rightarrow 0\) (von Mises matrix material) and also to provide better predictions for high triaxialities, the previous criterion (Eq. (D.3)) has been heuristically modified by Lee and Oung [24]. Their improved approximate criterion reads:

$$\begin{aligned} \Phi _{LO}^{app}=\frac{\Sigma _{sh}^{2}}{\sigma _{0}^{2}}+2f\cosh \left( \frac{\sqrt{3}\Sigma _{m}}{\sigma _{0}}\right) +\alpha \left( 1-f\right) \frac{\left( 3\Sigma _{m}+\Sigma _{sh}\right) }{\sigma _{0}}-\left( 1+f^{2}\right) =0 \end{aligned}$$
(D.6)

where relation \(\Sigma _{e}=\pm \Sigma _{sh}\) has been used. Considering a purely hydrostatic loading \(\left( \Sigma _{sh}=0\right)\), the improved criterion yields:

$$\begin{aligned} 2f\cosh \left( \frac{\sqrt{3}\Sigma _{m}}{\sigma _{0}}\right) +3\alpha \left( 1-f\right) \frac{\Sigma _{m}}{\sigma _{0}}-\left( 1+f^{2}\right) =0 \end{aligned}$$
(D.7)

which clearly recovers Eq. (D.2) when \(\alpha \rightarrow 0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

dos Santos, T., Vadillo, G. A closed-form yield criterion for porous materials with Mises–Schleicher–Burzyński matrix containing cylindrical voids. Acta Mech 232, 1285–1306 (2021). https://doi.org/10.1007/s00707-020-02925-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00707-020-02925-y

Navigation