Skip to main content
Log in

An analysis of deformation and failure in rectangular tensile bars accounting for void shape changes

  • Original Paper
  • Published:
International Journal of Fracture Aims and scope Submit manuscript

Abstract

A three-dimensional finite deformation study of necking and failure in rectangular tensile bars is carried out using a constitutive relation for porous material plasticity. The fully dynamic formulation accounts for void nucleation and growth along with thermal and rate effects, but here focus is on quasi-static response with a specified initial void volume fraction. The constitutive relation takes into account void shape changes and associated void rotations for three-dimensional voids. The constitutive update is carried out using a generalized rate tangent scheme for an elastic-viscoplastic solid. The sensitivity of necking and failure patterns to the aspect ratio of the rectangular bar is investigated with focus on the plane strain limit and a square tensile bar. The calculations predict the well-known slant fracture in plane strain tension and the emergence of a cup-cone like failure region for a square cross-section. Details are provided for the development of porosity in the bar with a square cross-section, including void shape changes and void rotations. The numerical examples show the capability of a constitutive relation for porous plasticity that can model details of void evolution, thus paving the way for advanced analyses of ductile failure under arbitrary loadings.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

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
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  • Agoras M, Ponte Castañeda P (2013) Iterated linear comparison bounds for viscoplastic porous materials with ellipsoidal microstructures. J Mech Phys Solids 61:701–725

    Article  Google Scholar 

  • Aravas N, Papadioti I (2021) A non-local plasticity model for porous metals with deformation-induced anisotropy: mathematical and computational issues. J Mech Phys Solids 146:104190

    Article  Google Scholar 

  • Aravas N, Ponte Castaneda P (2004) Numerical methods for porous metals with deformation-induced anisotropy. Comput Methods Appl Mech Eng 193:3767–3805

    Article  Google Scholar 

  • Bao Y, Wierzbicki T (2004) On fracture locus in the equivalent strain and stress triaxiality space. Int J Mech Sci 46:81–98

    Article  Google Scholar 

  • Barrioz PO, Hure J, Tanguy B (2019) On void shape and distribution effects on void coalescence. J Appl Mech 86:0110061

    Article  Google Scholar 

  • Barsoum I, Faleskog J (2007a) Rupture mechanisms in combined tension and shear—experiments. Int J Solids Struct 44:1768–1786

    Article  CAS  Google Scholar 

  • Barsoum I, Faleskog J (2007b) Rupture mechanisms in combined tension and shear—micromechanics. Int J Solids Struct 44:5481–5498

    Article  CAS  Google Scholar 

  • Bazant Z, Pijaudier-Cabot G (1988) Non local continuum damage. localization, instability and convergence. J Appl Mech 55:287–294

    Article  Google Scholar 

  • Becker R, Needleman A (1986) Effect of yield surface curvature on necking and failure in porous solids. J Appl Mech 53:491–499

    Article  Google Scholar 

  • Becker R, Smelser RE, Richmond O (1989) The effect of void shape on the development of damage and fracture in plane-strain tension. J Mech Phys Solids 37(1):111–129

    Article  Google Scholar 

  • Belytschko T, Chiapetta RL, Bartel HD (1976) Efficient large scale non-linear transient analysis by finite elements. Int J Numer Methods Eng 10:579–596

    Article  Google Scholar 

  • Benzerga AA (2002) Micromechanics of coalescence in ductile fracture. J Mech Phys Solids 50:1331–1362

    Article  Google Scholar 

  • Benzerga AA, Leblond JB (2010) Ductile fracture by void growth to coalescence. Adv Appl Mech 44:169–305

    Article  Google Scholar 

  • Benzerga AA, Leblond JB (2014) Effective yield criterion accounting for microvoid coalescence. J Appl Mech 81(3):031009

    Article  Google Scholar 

  • Benzerga AA, Besson J, Pineau A (1999) Coalescence-controlled anisotropic ductile fracture. J Eng Mater Technol 121:221–229

    Article  CAS  Google Scholar 

  • Benzerga AA, Besson J, Batisse R, Pineau A (2002a) Synergistic effects of plastic anisotropy and void coalescence on fracture mode in plane strain. Model Simul Mater Sci Eng 10:73–102

    Article  Google Scholar 

  • Benzerga AA, Tvergaard V, Needleman A (2002b) Size effects in the Charpy V-notch test. Int J Fract 116:275–296

    Article  CAS  Google Scholar 

  • Benzerga AA, Besson J, Pineau A (2004a) Anisotropic ductile fracture. Part I: experiments. Acta Mater 52:4623–4638

    Article  CAS  Google Scholar 

  • Benzerga AA, Besson J, Pineau A (2004b) Anisotropic ductile fracture. Part II: theory. Acta Mater 52:4639–4650

    Article  CAS  Google Scholar 

  • Benzerga AA, Leblond JB, Needleman A, Tvergaard V (2016) Ductile failure modeling. Int J Fract 201:29–80

    Article  Google Scholar 

  • Benzerga AA, Thomas N, Herrington JS (2019) Plastic flow anisotropy drives shear fracture. Sci Rep 9: Art. No. 1425

  • Beremin FM, Pineau A, Mudry F, Devaux J, D’Escatha Y, Ledermann P (1981) Cavity formation from inclusions in ductile fracture of A508 steel. Metall Trans A 12A:723–731

    Article  Google Scholar 

  • Besson J, Steglich D, Brocks W (2001) Modeling of crack growth in round bars and plane strain specimens. Int J Solids Struct 38:8259–8284

    Article  Google Scholar 

  • Besson J, Steglich D, Brocks W (2003) Modeling of plane strain ductile rupture. Int J Plast 19:1517–1541

    Article  Google Scholar 

  • Burke MA, Nix WD (1979) Numerical study of necking in the plane tension test. Int J Solids Struct 15:379–393

    Article  Google Scholar 

  • Cazacu O, Rodríguez-Martínez J (2019) Effects of plastic anisotropy on localization in orthotropic materials: new explicit expressions for the orientation of localization bands in flat specimens subjected to uniaxial tension. J Mech Phys Solids 126:272–284

    Article  Google Scholar 

  • Chu C, Needleman A (1980) Void nucleation effects in biaxially stretched sheets. J Eng Mater Technol 102:249–256

    Article  Google Scholar 

  • Danas K, Aravas N (2012) Numerical modeling of elasto-plastic porous materials with void shape effects at finite deformations. Composites B 43:2544–2559

    Article  Google Scholar 

  • Danas K, Ponte Castañeda P (2009a) A finite-strain model for anisotropic viscoplastic porous media: I. Theory. Eur J Mech 28:387–401

    Article  Google Scholar 

  • Danas K, Ponte Castañeda P (2009b) A finite-strain model for anisotropic viscoplastic porous media: II. Applications. Eur J Mech 28:402–416

    Article  Google Scholar 

  • Dunand M, Mohr D (2011) On the predictive capabilities of the shear modified Gurson and the modified Mohr–Coulomb fracture models over a wide range of stress triaxialities and lode angles. J Mech Phys Solids 59:1374–1394

    Article  CAS  Google Scholar 

  • Eshelby JD (1957) The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc R Soc Lond A 241:376–396

    Article  Google Scholar 

  • Gologanu M, Leblond JB, Devaux J (1993) Approximate models for ductile metals containing non-spherical voids—case of axisymmetric prolate ellipsoidal cavities. J Mech Phys Solids 41(11):1723–1754

    Article  Google Scholar 

  • Gologanu M, Leblond JB, Devaux J (1994) Approximate models for ductile metals containing non-spherical voids—case of axisymmetric oblate ellipsoidal cavities. J Eng Mater Technol 116:290–297

    Article  Google Scholar 

  • Gologanu M, Leblond JB, Perrin G, Devaux J (1997) Recent extensions of Gurson’s model for porous ductile metals. In: Suquet P (ed) Continuum micromechanics. CISM lectures series. Springer, New York, pp 61–130

    Chapter  Google Scholar 

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

    Article  Google Scholar 

  • Hadamard J (1903) Chapter 6. In: Herrmann A (ed) Leçons sur la propagation des ondes et les équations de l’hydrodynamique. Librairie Scientifique, Paris

    Google Scholar 

  • Haddag B, Abed-Meraim F, Balan T (2009) Strain localization analysis using a large deformation anisotropic elastic–plastic model coupled with damage. Int J Plast 25:1970–1996

    Article  CAS  Google Scholar 

  • Hancock JW, MacKenzie AC (1976) On the mechanisms of ductile failure in high-strength steels subjected to multi-axial stress states. J Mech Phys Solids 24:147–169

    Article  Google Scholar 

  • Hill R (1958) A general theory of uniqueness and stability in elastic–plastic solids. J Mech Phys Solids 6:236–249

    Article  Google Scholar 

  • Hill R (1962) Acceleration waves in solids. J Mech Phys Solids 10:1–16

    Article  Google Scholar 

  • Hill R, Hutchinson J (1975) Bifurcation phenomena in the plane tension test. J Mech Phys Solids 23:239–264

    Article  Google Scholar 

  • Huespe A, Needleman A, Oliver J, Sánchez P (2012) A finite strain, finite band method for modeling ductile fracture. Int J Plast 28:53–69

    Article  Google Scholar 

  • Hutchinson J, Miles J (1974) Bifurcation analysis of the onset of necking in an elastic/plastic cylinder under uniaxial tension. J Mech Phys Solids 22:61–71

    Article  Google Scholar 

  • Jablokov V, Goto DM, Koss DA (2001) Damage accumulation and failure of HY-100 Steel. Metall Mater Trans A 32A:2985–2994

    Article  CAS  Google Scholar 

  • Kailasam M, Ponte Castaneda P (1998) A general constitutive theory for linear and nonlinear particulate media with microstructure evolution. J Mech Phys Solids 46(3):427–465

    Article  CAS  Google Scholar 

  • Keralavarma SM, Chockalingam S (2016) A criterion for void coalescence in anisotropic ductile materials. Int J Plast 82:159–176

    Article  Google Scholar 

  • Krieg RO, Key SW (1973) Transient shell response by numerical time integration. Int J Numer Methods Eng 7:273–286

    Article  Google Scholar 

  • Leblond JB, Perrin G, Devaux J (1994a) Bifurcation effects in ductile metals with nonlocal damage. J Appl Mech 61:236–242

    Article  Google Scholar 

  • Leblond JB, Perrin G, Suquet P (1994b) Exact results and approximate models for porous viscoplastic solids. Intl J Plast 10:213–225

    Article  Google Scholar 

  • Madou K, Leblond JB (2012a) A Gurson-type criterion for porous ductile solids containing arbitrary ellipsoidal voids—I: limit-analysis of some representative cell. J Mech Phys Solids 60:1020–1036

    Article  Google Scholar 

  • Madou K, Leblond JB (2012b) A Gurson-type criterion for porous ductile solids containing arbitrary ellipsoidal voids—II: determination of yield criterion parameters. J Mech Phys Solids 60:1037–1058

    Article  Google Scholar 

  • Madou K, Leblond JB (2013) Numerical studies of porous ductile materials containing arbitrary ellipsoidal voids—I: yield surfaces of representative cells. Eur J Mech 42:480–489

    Article  Google Scholar 

  • Madou K, Leblond JB, Morin L (2013) Numerical studies of porous ductile materials containing arbitrary ellipsoidal voids—II: evolution of the length and orientation of the void axes. Eur J Mech 42:490–507

    Article  Google Scholar 

  • Mansouri LZ, Chalal H, Abed-Meraim F (2014) Ductility limit prediction using a GTN damage model coupled with localization bifurcation analysis. Mech Mater 76:64–92

    Article  Google Scholar 

  • Morin L, Leblond JB, Benzerga AA (2015) Coalescence of voids by internal necking: theoretical estimates and numerical results. J Mech Phys Solids 75:140–158

    Article  Google Scholar 

  • Morin L, Leblond JB, Tvergaard V (2016) Application of a model of plastic porous materials including void shape effects to the prediction of ductile failure under shear-dominated loadings. J Mech Phys Solids 94:148–166

    Article  Google Scholar 

  • Morin L, Leblond JB, Mohr D, Kondo D (2017) Prediction of shear-dominated ductile fracture in a butterfly specimen using a model of plastic porous solids including void shape effects. Eur J Mech 61:433–442

    Article  Google Scholar 

  • Nahshon K, Hutchinson JW (2008) Modification of the Gurson Model for shear failure. Eur J Mech 27:1–17

    Article  Google Scholar 

  • Needleman A (1972) A numerical study of necking in circular cylindrical bars. J Mech Phys Solids 20:111–127

    Article  Google Scholar 

  • Needleman A (1982) Finite elements for finite strain plasticity problems. In: Lee E, Mallett R (eds) Plasticity of metals at finite strain: theory, computations and experiments, p 387

  • Needleman A (1988) Material rate dependence and mesh sensitivity in localization problems. Comput Methods Appl Mech Eng 67:69–85

    Article  Google Scholar 

  • Needleman A (2018) Effect of size on necking of dynamically loaded notched bars. Mech Mater 116:180–188

    Article  Google Scholar 

  • Needleman A, Rice JR (1978) Limits to ductility set by plastic flow localization. In: Koistinen DP, Wang NM (eds) Mechanics of sheet metal forming. Plenum Press, New York, pp 237–267

    Chapter  Google Scholar 

  • Nemcko MJ, Li J, Wilkinson DS (2016) Effects of void band orientation and crystallographic anisotropy on void growth and coalescence. J Mech Phys Solids 95:270–283

    Article  CAS  Google Scholar 

  • Norris D, Moran B, Scudder J, Quinones D (1978) A computer simulation of the tension test. J Mech Phys Solids 26:1–19

    Article  Google Scholar 

  • Ottosen NS, Runesson K (1991) Properties of discontinuous bifurcation solutions in elasto-plasticity. Int J Solids Struct 27:401–421

    Article  Google Scholar 

  • Pan J, Saje M, Needleman A (1983) Localization of deformation in rate sensitive porous plastic solids. Int J Fract 21:261–278

    Article  Google Scholar 

  • Pardoen T (2006) Numerical simulation of low stress triaxiality ductile fracture. Comput Struct 84:1641–1650

    Article  Google Scholar 

  • Pardoen T, Hutchinson JW (2000) An extended model for void growth and coalescence. J Mech Phys Solids 48:2467–2512

    Article  Google Scholar 

  • Peirce D, Shih CF, Needleman A (1984) A tangent modulus method for rate dependent solids. Comput Struct 18:875–887

    Article  Google Scholar 

  • Ponte Castañeda P, Zaidman M (1994) Constitutive models for porous materials with evolving microstructure. J Mech Phys Solids 42:1459–1495

    Article  Google Scholar 

  • Rice JR (1976) The localization of plastic deformation. In: Koiter WT (ed) 14th International congress on theoretical and applied mechanics. North-Holland, Amsterdam, pp 207–220

  • Rousselier G (1987) Ductile fracture models and their potential in local approach of fracture. Nucl Eng Des 105:97–111

    Article  CAS  Google Scholar 

  • Rudnicki JW, Rice JR (1975) Conditions for the localization of deformation in pressure-sensitive dilatant materials. J Mech Phys Solids 23:371–394

  • Runesson K, Ottosen NS, Dunja P (1991) Discontinuous bifurcations of elastic–plastic solutions at plane stress and plane strain. Int J Plast 7:99–121

    Article  Google Scholar 

  • Speich GR, Spitzig WA (1982) Effect of volume fraction and shape of sulfide inclusions on through-thickness ductility and impact energy of high-strength 4340 plate steels. Metall Trans 13A:2239–2258

    Article  Google Scholar 

  • Srivastava A, Ponson L, Osovski S, Bouchaud E, Tvergaard V, Needleman A (2014) Effect of inclusion density on ductile fracture toughness and roughness. J Mech Phys Solids 63:62–79

    Article  Google Scholar 

  • Torki ME, Benzerga AA (2018) A mechanism of failure in shear bands. Extreme Mech Lett 23:67–71

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Tvergaard V (1982) On localization in ductile materials containing spherical voids. Int J Fract 18:237–252

    Article  Google Scholar 

  • Tvergaard V (1993) Necking in tensile bars with rectangular cross-section. Comput Methods Appl Mech Eng 103:273–290

    Article  Google Scholar 

  • Tvergaard V (2009) Behaviour of voids in a shear field. Int J Fract 158:41–49

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Tvergaard V, Needleman A (1995) Effects of nonlocal damage in porous plastic solids. Int J Solids Struct 32(8/9):1063–1077

    Article  Google Scholar 

  • Tvergaard V, Needleman A, Lo KK (1981) Flow localization in the plane strain tensile test. J Mech Phys Solids 29:115–142

    Article  Google Scholar 

  • Willis JR (1977) Bounds and self-consistent estimates for the overall properties of anisotropic composites. J Mech Phys Solids 25:185–202

    Article  Google Scholar 

  • Zhang Y, Lorentz E, Besson J (2018) Ductile damage modelling with locking-free regularised GTN model. Int J Numer Methods Eng 113:1871–1903

    Article  Google Scholar 

Download references

Acknowledgements

AAB acknowledges funding from the Lawrence Livermore National Laboratory under Master Task Agreements No. B599687 and B602391. We are grateful for the high performance research computing resources provided by Texas A&M University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. A. Benzerga.

Additional information

Publisher's Note

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

Appendices

Appendix A: Specification of terms in the Madou–Leblond potential

The components of the quadratic form \({{\mathcal {Q}}}({\varvec{\sigma }})\) and linear form \({{\mathcal {H}}}({\varvec{\sigma }})\) in Eqs. (15) and (16), respectively, are expressed on the orthonormal basis (\(\mathbf {n}_1, \mathbf {n}_2, \mathbf {n}_3\)) associated with the voids and components of tensors are given in Cartesian tensor notation. The general form of the expressions was obtained by homogenization but additional heuristics was included, based on a large set of numerical (limit-analysis) simulations in (Madou and Leblond 2012b). This involved some interpolation from the previously studied cases of prolate and oblate voids using geometric parameters, such as void and cell eccentricities as parameters. Hence, some familiarity with the geometry of the underlying unit cell is needed.

The semi-axes of the void are denoted \(a \ge b \ge c\). Those of the cell, taken to be confocal with the void, are denoted \(A \ge B \ge C\). Thus,

$$\begin{aligned} A=\sqrt{a^{2}+\varLambda }, \quad B=\sqrt{b^{2}+\varLambda }, \quad C=\sqrt{c^{2}+\varLambda }\nonumber \\ \end{aligned}$$
(A-1)

where \(\varLambda \) is the unique positive root of:

$$\begin{aligned} \left( a^{2}+\varLambda \right) \left( b^{2}+\varLambda \right) \left( c^{2}+\varLambda \right) -\frac{a^{2}b{^2}c^{2}}{f^{2}}=0\nonumber \\ \end{aligned}$$
(A-2)

A special case of interest is the completely flat confocal ellipsoid having semi-axes \({{\bar{a}}} \ge {{\bar{b}}} \ge {{\bar{c}}} =0\) and aspect ratio k:

$$\begin{aligned} {\bar{a}}=\sqrt{a^{2}-c^{2}}, \quad {\bar{b}}=\sqrt{b^{2}-c^{2}}, \quad k=\frac{{\bar{b}}}{{\bar{a}}} \end{aligned}$$
(A-3)

It is used to define the ‘second porosity’ g and related quantities through:

$$\begin{aligned} g=\frac{{\bar{a}}{\bar{b}}^{2}}{ABC}, \quad g_{1}=\frac{g}{1+g}, \quad g_{f}=\frac{g}{f+g} \end{aligned}$$
(A-4)

Physically, g has the meaning of void volume fraction of some fictitious prolate void obtained by rotating the completely flat ellipsoid (of semi-axes \(\bar{a}, \bar{b}, \bar{c} = 0 \)) about its major axis (Madou and Leblond 2012b).

Finally, the eccentricities of the inner and outer ellipsoids are given by:

$$\begin{aligned} e_{xz}=\frac{{\bar{a}}}{a}, \quad E_{xz}=\frac{{\bar{a}}}{A}, \quad \quad E_{yz}=\frac{{\bar{b}}}{B} \end{aligned}$$
(A-5)

The tensor \({\varvec{H}}\) in Eq. (16) has a unit trace (\(H_{kk} = 1\)) and is diagonal in the ‘void’ basis:

$$\begin{aligned} {\varvec{H}} = \sum _{i=1,3} H_i \, \mathbf {n}_i\otimes \mathbf {n}_i\end{aligned}$$
(A-6)

with

$$\begin{aligned} \begin{aligned} \left\{ \begin{array}{l l} \displaystyle {H_1= \left( 1-k^{2}\right) H_1^\mathrm{prol}+k^{2}H_1^\mathrm{obl}, } \\ \displaystyle {H_2= \left( 1-k\right) H_2^\mathrm{prol}+kH_2^\mathrm{obl}+} \\ \displaystyle { \qquad \frac{1}{2}\left( 1-k\right) \frac{\alpha ^{2}+\beta ^{2}}{\alpha }\frac{E_{xz}^{3/2} \left( 1-\alpha -E_{xz} \right) }{\left( 1-\alpha -E_{xz} \right) ^{2}+\beta ^{2}},} \\ \displaystyle {\alpha = \frac{4k^{2}}{1+9k^{2}}, \quad \beta =\frac{3k^{2}}{1+30k^{2}} } \end{array} \right. \end{aligned} \end{aligned}$$
(A-7)

where

$$\begin{aligned} \begin{aligned} H_1^\mathrm{prol}=1-2H_2^\mathrm{prol}, \quad H_2^\mathrm{prol}=\frac{1}{3}\left( 1+E_{xz}^{2}-\frac{E_{xz}^{4}}{2} \right) , \\ H_1^\mathrm{obl} = H_2^\mathrm{obl} = \frac{1}{3}\frac{\left( 2-7E_{xz}^{2}+5E_{xz}^{4}\right) }{\left( 2-7E_{xz}^{2}+10E_{xz}^{4}\right) } \end{aligned} \end{aligned}$$
(A-8)

A key parameter entering Eq. (13) is

$$\begin{aligned} \kappa = \frac{3}{2 {{\bar{F}}}} \end{aligned}$$
(A-9)

with, in general,

$$\begin{aligned} \begin{aligned}&{\bar{F}}=1+\frac{1}{\ln \left( g_{f}/g_{1} \right) } \Biggl [ -\left( 1-k\right) \left( 1-\frac{\sqrt{3}}{2} \right) \\&\quad \ln \frac{11k^{2}+5g_{f}}{11k^{2}+5g_{1}}+ \\&\quad \frac{3}{5}\left( 1-k\right) ^{2}\ln \frac{8-5g_{1}}{8-5g_{f}} +\frac{13}{10}k\left( g_{f}-g_{1} \right) -\frac{3}{10}k\left( g_{f}^{5}-g_{1}^{5} \right) \Biggr ] \end{aligned} \end{aligned}$$
(A-10)

For a prolate spheroidal void, the following expression is used:

$$\begin{aligned} {\bar{F}}=1+\frac{1-\sqrt{3}/2}{\ln f} \ln \frac{11+5e_{xz}^{3}/\left( 1-e_{xz}^{2} \right) }{11+5fe_{xz}^{3}/\left( 1-e_{xz}^{2} \right) } \end{aligned}$$
(A-11)

In the case of spherical voids, \(H_1 = H_2 = H_3= 1/3\) and \({{\bar{F}}} = 1\) so that \({{\mathcal {H}}}({\varvec{\sigma }})\) reduces to the mean normal stress \(\sigma _{kk}/3\) and \(\kappa = 3/2\) as in the GT constitutive relation.

The quadratic form in Eq. (15) is based on the Willis bound. It may be written as:

$$\begin{aligned}&{\mathcal {Q}}\left( {\varvec{\sigma }}\right) = {\mathcal {Q}}^{W}\left( {\varvec{\sigma }}\right) - (1+g)(f+g)\kappa ^{2} {{{\mathcal {H}}}}^2 \end{aligned}$$
(A-12)

where it is convenient to formally isolate diagonal stress components as follows:

$$\begin{aligned}&{\mathcal {Q}}^{W}\left( {\varvec{\sigma }}\right) = {\varvec{\sigma }}^{\texttt {dg}} \cdot {\varvec{Q}}^{\texttt {dg}} \cdot {\varvec{\sigma }}^{\texttt {dg}}\nonumber \\&+ {\varvec{\sigma }}^{\texttt {offdg}} \cdot {\varvec{Q}}^{\texttt {offdg}} \cdot {\varvec{\sigma }}^{\texttt {offdg}} \end{aligned}$$
(A-13)

with

$$\begin{aligned} {\varvec{\sigma }}^{\texttt {dg}} \equiv \left( \begin{array}{l l} \displaystyle {\sigma _{xx}} \\ \displaystyle {\sigma _{yy}} \\ \displaystyle {\sigma _{zz}} \\ \end{array} \right) , \quad {\varvec{\sigma }}^{\texttt {offdg}} \equiv \left( \begin{array}{l l} \displaystyle {\sigma _{xy}} \\ \displaystyle {\sigma _{yz}} \\ \displaystyle {\sigma _{xz}} \\ \end{array} \right) \end{aligned}$$
(A-14)

Here and in what follows, we identify the x-, y- and z-axes with the void axes, for convenience. Then \({\varvec{Q}}^{\texttt {dg}}\) and \({\varvec{Q}}^{\texttt {offdg}}\) are symmetric second-rank tensors defined by

$$\begin{aligned}&{\varvec{Q}}^{\texttt {dg}} \equiv \left( 1-f \right) \begin{bmatrix} 1 &{} -\frac{1}{2} &{} -\frac{1}{2} \\ -\frac{1}{2} &{} 1 &{} -\frac{1}{2} \\ -\frac{1}{2} &{} -\frac{1}{2} &{} 1 \end{bmatrix}\nonumber \\&+ \frac{3f}{2} \begin{bmatrix} T_{\texttt {xxxx}} &{} T_{\texttt {xxyy}} &{} T_{\texttt {xxzz}} \\ T_{\texttt {yyxx}} &{} T_{\texttt {yyyy}} &{} T_{\texttt {yyzz}} \\ T_{\texttt {zzxx}} &{} T_{\texttt {zzyy}} &{} T_{\texttt {zzzz}} \end{bmatrix}^{-1} \end{aligned}$$
(A-15)
$$\begin{aligned}&{\varvec{Q}}^{\texttt {offdg}} \equiv 3\left( 1-f \right) \begin{bmatrix} 1 &{} 0 &{} 0 \\ 0 &{} 1 &{} 0 \\ 0 &{} 0 &{} 1 \end{bmatrix}\nonumber \\&+ \frac{3f}{2} \begin{bmatrix} T_{\texttt {xyxy}} &{} 0 &{} 0 \\ 0 &{} T_{\texttt {yzyz}} &{} 0\\ 0 &{} 0 &{} T_{\texttt {zxzx}} \end{bmatrix}^{-1} \end{aligned}$$
(A-16)

where \(\mathbb {T}\) is a fourth-order tensor connected to the Eshelby tensor \(\mathbb {S}(\nu )\) of an ellipsoidal void of semi-axes a,b,c embedded in a fictitious infinite elastic medium with shear modulus \(\mu \) and Poisson ratio \(\nu \), through the relation

$$\begin{aligned} \mathbb {T} = \lim _{\nu \rightarrow 1/2} \frac{1}{2\mu } \mathbb {L}\left( \mu ,\nu \right) :\left[ \mathbb {I}-\mathbb {S}\left( \nu \right) \right] \end{aligned}$$
(A-17)

where \(\mathbb {L}\left( \mu ,\nu \right) \) denotes the stiffness tensor of the medium and \(\mathbb {I}\) the identity tensor. The components of tensor \(\mathbb {T}\) involve implicit integrals and obey the symmetry relations \(T_{ijkl}\)= \(T_{jikl}\)=\(T_{ijlk}\)=\(T_{klij}\). They are listed in Appendix C of Madou and Leblond (2012b).

The parameters involved in the linear form \({{{\mathcal {H}}}}({\varvec{\sigma }})\) were assessed against a large dataset of numerical limit analyses (Madou and Leblond 2012b). Also, such calculations are typically carried out using a “frozen’ microstructure” and the quadratic form \({{{\mathcal {Q}}}}({\varvec{\sigma }})\) is based on a tight bound, which remains nevertheless a bound. To account for the effects of deviation from the assumptions underlying the limit analyses (including, for example, the effects of the evolution of the normal to the flow potential surface, strain hardening and strain-rate hardening) Tvergaard’s parameters \(q_1\) and \(q_2\) (multiplying \(\kappa \)) have been kept. In the results shown in the paper, constant values of \(q_1\) and \(q_2\) have been specified. Morin et al. (2016) have proposed a more complex formula for \(q_1\), which is not used here.

Appendix B: Specification of the concentration tensors

As in Appendix A component expressions, when given, are expressed in Cartesian tensor notation. For an elastic porous matrix, strain concentration is determined in some average sense by

$$\begin{aligned} {\varvec{D}}^{v} = \mathbb {C}^\mathrm{e}: \mathbf{d}, \quad \mathbb {C}^\mathrm{e} = \left[ \mathbb {I} -\left( 1-f\right) \mathbb {S} \right] ^{-1} \end{aligned}$$
(B-1)

where \(\mathbb {S}\) is Eshelby (1957)’s first tensor as above; see (Ponte Castañeda and Zaidman 1994).

For a plastic porous matrix, using the above linear estimate leads to poor predictions, as expected (Madou et al. 2013). Using the following definition of a plastic strain concentration tensor:

$$\begin{aligned} {\varvec{D}}^{v} = \mathbb {C}: \mathbf{d}^p \end{aligned}$$
(B-2)

Madou and Leblond introduced a well-calibrated tensor \(\mathbb {C}\) based on a correction of \(\mathbb {C}^\mathrm{e}\) as follows. For convenience, a change of variable is made for strain-rate tensors, e.g. \(\mathbf{d}\):

$$\begin{aligned} \left\{ \begin{array}{l llll} \displaystyle {{\bar{d}}_{1}}= d_{m} \\ \displaystyle {{\bar{d}}_{2}}= d_{xx}-d_{m} \\ \displaystyle {{\bar{d}}_{3}}= d_{yy}-d_{zz} \\ \displaystyle {{\bar{d}}_{4}}= d_{xy}\\ \displaystyle {{\bar{d}}_{5}}= d_{xz}\\ \displaystyle {{\bar{d}}_{6}}= d_{yz} \end{array} \right. \end{aligned}$$
(B-3)

Similar transformations are made for \({\varvec{D}}^v\) and \(\mathbf{d}^p\). This is somewhat like a Voigt transformation. Also, note that we have kept the x-, y- and z indices for components in the void basis (\(\mathbf {n}_1, \mathbf {n}_2, \mathbf {n}_3\)) as in presenting the quadratic form in Appendix A. In matrix form, Eqs. (B-1) and (B-2) now write using barred notation:

$$\begin{aligned} {\bar{D}}^v = {\bar{C}}^\mathrm{e} \cdot {\bar{d}}, \qquad {\bar{D}}^v = {\bar{C}} \cdot {\bar{d}}^p \end{aligned}$$
(B-4)

We thus have:

$$\begin{aligned} {\bar{C}} = \begin{pmatrix} {\bar{C}}_{11} &{} {\bar{C}}_{12} &{} {\bar{C}}_{13} &{} 0 &{} 0 &{} 0 \\ {\bar{C}}_{21} &{} {\bar{C}}_{22} &{} {\bar{C}}_{23} &{} 0 &{} 0 &{} 0 \\ {\bar{C}}_{31} &{} {\bar{C}}_{32} &{} {\bar{C}}_{33} &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} {\bar{C}}_{44} &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} {\bar{C}}_{55} &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} {\bar{C}}_{66} \\ \end{pmatrix} \end{aligned}$$
(B-5)

where the following relations hold in view of the notation condensation inherent to Eq. (B-3):

$$\begin{aligned} \left\{ \begin{array}{l lllll} \displaystyle {{\bar{C}}_{11}}= 1/f; \quad {\bar{C}}_{12}={\bar{C}}_{13}=0 \\ \displaystyle {{\bar{C}}_{21}= C_{xxxx}+C_{xxyy}+C_{xxzz}-1/f} \\ \displaystyle {{\bar{C}}_{22}}= \left( 2C_{xxxx}-C_{xxyy}-C_{xxzz}\right) /2 \\ \displaystyle {{\bar{C}}_{23}}= \left( C_{xxyy}-C_{xxzz}\right) /2 \\ \displaystyle {{\bar{C}}_{31}}= C_{yyxx}+C_{yyyy}+C_{yyzz}\\ -C_{zzxx}-C_{zzyy}+C_{zzzz} \\ \displaystyle {{\bar{C}}_{32}}= \Big ( 2C_{yyxx}-C_{yyyy}-C_{yyzz}\\ -2C_{zzxx}+C_{zzyy}+C_{zzzz}\Big )/2 \\ \displaystyle {{\bar{C}}_{33}}= \left( C_{yyyy}-C_{yyzz}-C_{zzyy}+C_{zzzz}\right) /2 \\ \displaystyle {{\bar{C}}_{44}}= 2C_{xyxy}; \quad {\bar{C}}_{55}=2C_{xzxz}; \quad {\bar{C}}_{66}=2C_{yzyz} \\ \end{array} \right. \end{aligned}$$
(B-6)

The problem reduces to specifying the components of \( {\bar{C}}\) in terms of the known components of \( {\bar{C}}^\mathrm{e}\):

$$\begin{aligned} {\bar{C}}_{\alpha \beta } = h_{\alpha \beta }{\bar{C}}_{\alpha \beta }^\mathrm{e} \qquad \left( \text {no} \text{ sum } \text{ on } \alpha \text{ or } \beta \right) \end{aligned}$$
(B-7)

where \(h_{\alpha \beta }\) are correction factors taken to depend on void volume fraction f, void shape, as well as the stress state (via dimensionless ratios of stress invariants; see below). In fact, only three coefficients are needed since (Madou et al. 2013):

$$\begin{aligned} {\mathbf {h}} = \begin{pmatrix} 1 &{} 1 &{} 1 &{} 0 &{} 0 &{} 0 \\ 1 &{} h_{22} &{} h_{22} &{} 0 &{} 0 &{} 0 \\ 1 &{} h_{22} &{} h_{33} &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} h_{44} &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} h_{44} &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} h_{33} \\ \end{pmatrix} \end{aligned}$$
(B-8)

The stress state dependence, which reflects strongly nonlinear effects, is taken into account via the stress triaxiality ratio, T, the Lode parameter L, the lateral triaxiality \(T_h\) and a deviatoric transverse invariant X defined through:

$$\begin{aligned} T \equiv \frac{\sigma _{\mathrm{m}}}{\sigma _{\mathrm{eq}}}; \qquad L \equiv \cos \left( 3\phi \right) \equiv \frac{27}{2} \texttt {det} \frac{{\varvec{\sigma }}'}{\sigma _{\mathrm{eq}}} \end{aligned}$$
(B-9)

where \(0 \le \phi \le \pi /3\), and

$$\begin{aligned} T_{h} \equiv \frac{\frac{1}{2}\left( \sigma _{yy}+\sigma _{zz} \right) }{\sigma _{\mathrm{eq}}}; \qquad X \equiv \frac{\sigma _{xx}-\frac{1}{2}\left( \sigma _{yy}+\sigma _{zz} \right) }{\sigma _{\mathrm{eq}}}\nonumber \\ \end{aligned}$$
(B-10)

The void shape dependence of the \(h_{\alpha \beta }\)’s enters via factor k and eccentricity \(e_{xz}\), defined in Eqs. (A-3) and (A-5), respectively.

With the above definitions, the corrections proceed by interpolating the general case between the prolate and oblate special cases, which themselves are interpolated between a sphere and a cylinder. Thus,

$$\begin{aligned} \left\{ \begin{array}{l llll} \displaystyle {{h}_{22}\left( f,k,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-k \right) h_{22}^\mathrm{prol} \left( f,e_{xz},{\varvec{\sigma }}\right) \\ +k h^\mathrm{obl} \left( f,e_{xz},{\varvec{\sigma }}\right) \\ \displaystyle {{h}_{33}\left( f,k,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-k \right) h_{33}^\mathrm{prol} \left( f,e_{xz},{\varvec{\sigma }}\right) \\ +k h^\mathrm{obl} \left( f,e_{xz},{\varvec{\sigma }}\right) \\ \displaystyle {{h}_{44}\left( f,k,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-k \right) h_{44}^\mathrm{prol} \left( f,e_{xz},{\varvec{\sigma }}\right) \\ +k h^\mathrm{obl} \left( f,e_{xz},{\varvec{\sigma }}\right) \end{array} \right. \end{aligned}$$
(B-11)

where \({\varvec{\sigma }}\) indicates stress-state dependence and \(h_{22}^\mathrm{prol}\), \(h_{33}^\mathrm{prol}\), \(h_{44}^\mathrm{prol}\) and \(h^\mathrm{obl}\) are given by:

$$\begin{aligned}&\left\{ \begin{array}{l llll} \displaystyle {{h}_{22}^\mathrm{prol}\left( f,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-e_{xz}^{30} \right) h^\mathrm{sph} \left( f,{\varvec{\sigma }}\right) + e_{xz}^{30}\\ \displaystyle {{h}_{33}^\mathrm{prol}\left( f,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-e_{xz}^{30} \right) h^\mathrm{sph} \left( f,{\varvec{\sigma }}\right) + e_{xz}^{30}h_{33}^\mathrm{cyl}\left( f,{\varvec{\sigma }}\right) \\ \displaystyle {{h}_{44}^\mathrm{prol}\left( f,e_{xz},{\varvec{\sigma }}\right) }= \left( 1-e_{xz}^{30} \right) h^\mathrm{sph} \left( f,{\varvec{\sigma }}\right) + e_{xz}^{30}h_{44}^\mathrm{cyl}\left( f,{\varvec{\sigma }}\right) \end{array} \right. \nonumber \\\end{aligned}$$
(B-12)
$$\begin{aligned}&{h}^\mathrm{obl}\left( f,e_{xz},{\varvec{\sigma }}\right) = \left( 1-e_{xz}^{50} \right) h^\mathrm{sph} \left( f,{\varvec{\sigma }}\right) + e_{xz}^{50}\nonumber \\ \end{aligned}$$
(B-13)

in which

$$\begin{aligned} h^\mathrm{sph}= & {} \frac{1}{7}\left( 9-10\sqrt{f}+8f \right) - \nonumber \\&\left( 1-\sqrt{f} \right) ^{8}\frac{11T^{4}}{5\left( 11+|T|^{3} \right) } \left[ 1+\frac{L}{3}{} \texttt {sgn}\left( T\right) \right] \end{aligned}$$
(B-14)
$$\begin{aligned}&h_{33}^\mathrm{cyl}= \frac{1}{2}\left( 3-5\sqrt{f}+4f \right) -\left( 1-\sqrt{f} \right) ^{5}\frac{9T_{h}^{4}}{20+T_{h}^{4}} \nonumber \\\end{aligned}$$
(B-15)
$$\begin{aligned}&h_{44}^\mathrm{cyl}= 1+\frac{11}{2}\left( 1-\sqrt{f} \right) ^{11} -\left( 1-\sqrt{f} \right) ^{11}\frac{84X^{2}}{1+20X^{2}}\nonumber \\ \end{aligned}$$
(B-16)

All of the above expressions were validated against direct numerical simulations by limit analysis (Madou et al. 2013). Large values of exponents indicate the strong nonlinearity of some effects.

The procedure for calculating the void rotation rate is quite similar to that use for the void strain rate and the same numerical results were used by Madou and Leblond (Madou et al. 2013) to obtain corrections for mean void rotations.

Define void rotation-rate concentration tensors \(\mathbb {R}^\mathrm{e}\) and \(\mathbb {R}\) through:

$$\begin{aligned} {\varvec{\varOmega }}^{v}= & {} {\varvec{\varOmega }}+ \mathbb {R}^\mathrm{e}: \mathbf{d}, \quad \mathbb {R}^\mathrm{e} \equiv \left( 1-f\right) \mathbb {P}: \mathbb {C}^\mathrm{e} \end{aligned}$$
(B-17)
$$\begin{aligned} {\varvec{\varOmega }}^{v}= & {} {\varvec{\varOmega }}+ \mathbb {R} : \mathbf{d}^p , \end{aligned}$$
(B-18)

for the elastic and plastic cases, respectively. Here, \(\mathbb {P}\) is Eshelby’s second tensor and \(\mathbb {C}^\mathrm{e}\) the elastic strain-rate concentration tensor, as above. Eq. (B-17B-18) was derived using homogenization by Kailasam and Ponte Castaneda (1998).

Simpler corrections are introduced for the non-zero components of \(\mathbb {R}\) on the basis of those, known, of \(\mathbb {R}^\mathrm{e}\):

$$\begin{aligned}&R_{xyxy}= \rho R_{xyxy}^\mathrm{e}, \quad R_{xzxz}= \rho R_{xzxz}^\mathrm{e},\nonumber \\&\quad R_{yzyz}= R_{yzyz}^\mathrm{e} \end{aligned}$$
(B-19)

which involve only one coefficient, \(\rho \), given by:

$$\begin{aligned} \rho \left( f,k,e_{xz},{\varvec{\sigma }}\right) = \left( 1-k\right) \rho ^\mathrm{prol}\left( f,e_{xz},{\varvec{\sigma }}\right) +k\nonumber \\ \end{aligned}$$
(B-20)

with

$$\begin{aligned} \rho ^\mathrm{prol}\left( f,e_{xz},{\varvec{\sigma }}\right) = 1-e_{xz}^{30}+e_{xz}^{30}\rho ^\mathrm{cyl} \left( f,{\varvec{\sigma }}\right) \end{aligned}$$
(B-21)

and

$$\begin{aligned} \rho ^\mathrm{cyl}= 1+11\left( 1-\sqrt{f} \right) ^{11} -\left( 1-\sqrt{f} \right) ^{11}\frac{170X^{2}}{1+20X^{2}}\nonumber \\ \end{aligned}$$
(B-22)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khan, I.A., Srivastava, A., Needleman, A. et al. An analysis of deformation and failure in rectangular tensile bars accounting for void shape changes. Int J Fract 230, 133–156 (2021). https://doi.org/10.1007/s10704-021-00532-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10704-021-00532-z

Keywords

Navigation