Skip to main content
Log in

A stabilized, symmetric Nitsche method for spatially localized plasticity

  • Original Paper
  • Published:
Computational Mechanics Aims and scope Submit manuscript

Abstract

A heterogeneous interface method is developed for combining primal displacement and mixed displacement-pressure formulations across nonconforming finite element meshes to treat volume-preserving plastic flow. When the zone of inelastic response is localized within a larger domain, significant computational savings can be achieved by confining the mixed formulation solely to the localized region. The method’s distinguishing feature is that the coupling terms for joining dissimilar element types are derived from a time-discrete free energy functional, which is based on a Lagrange multiplier formulation of the interface constraints. Incorporating residual-based stabilizing terms at the interface enables the condensation of the multiplier field, leading to a symmetric Nitsche formulation in which the interface operators respect the differing character of the governing equations in each region. In a series of numerical problems, the heterogeneous interface method achieved comparable results on coarser meshes as those obtained from applying the mixed formulation throughout the domain.

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
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23

Similar content being viewed by others

References

  1. Masud A, Scovazzi G (2011) A heterogeneous multiscale modeling framework for hierarchical systems of partial differential equations. Int J Numer Methods Fluids 65(1–3):28–42

    Article  MathSciNet  MATH  Google Scholar 

  2. Simo JC, Hughes TJR (1998) Computational inelasticity, 7th edn. Springer, New York

    MATH  Google Scholar 

  3. Simo JC, Taylor RL, Pister KS (1985) Variational and projection methods for the volume constraint in finite deformation elasto-plasticity. Comput Methods Appl Mech Eng 51(1–3):177–208

    Article  MathSciNet  MATH  Google Scholar 

  4. Cervera M, Chiumenti M, Valverde Q, de Saracibar CA (2003) Mixed linear/linear simplicial elements for incompressible elasticity and plasticity. Comput Methods Appl Mech Eng 192(49–50):5249–5263

    Article  MATH  Google Scholar 

  5. Masud A, Xia K (2006) A variational multiscale method for inelasticity: application to superelasticity in shape memory alloys. Comput Methods Appl Mech Eng 195(33–36):4512–4531

    Article  MathSciNet  MATH  Google Scholar 

  6. Le Tallec P, Sassi T (1995) Domain decomposition with nonmatching grids: augmented Lagrangian approach. Math Comput 64(212):1367–1396

    Article  MATH  Google Scholar 

  7. Brezzi F, Fortin M (1991) Mixed and hybrid finite elelement methods, vol 15. Springer, New York

    Book  Google Scholar 

  8. Gerstenberger A, Wall WA (2010) An embedded Dirichlet formulation for 3D continua. Int J Numer Methods Eng 82(5):537–563

    MathSciNet  MATH  Google Scholar 

  9. Barbosa HJC, Hughes TJR (1991) The finite element method with Lagrange multipliers on the boundary: circumventing the Babuška-Brezzi condition. Comput Methods Appl Mech Eng 85(1):109–128

    Article  MathSciNet  MATH  Google Scholar 

  10. Nitsche J (1971) Über ein Variationsprinzip zur Lösung von Dirichlet-Problemen bei Verwendung von Teilräumen, die keinen Randbedingungen unterworfen sind. Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg 36(1):9–15

    Article  MathSciNet  MATH  Google Scholar 

  11. Arnold DN, Brezzi F, Cockburn B, Marini LD (2002) Unified analysis of discontinuous Galerkin methods for elliptic problems. SIAM J Numer Anal 39(5):1749–1779

    Article  MathSciNet  MATH  Google Scholar 

  12. Annavarapu C, Hautefeuille M, Dolbow JE (2012) A robust Nitsche’s formulation for interface problems. Comput Methods Appl Mech Eng 225:44–54

    Article  MathSciNet  Google Scholar 

  13. Hautefeuille M, Annavarapu C, Dolbow JE (2012) Robust imposition of Dirichlet boundary conditions on embedded surfaces. Int J Numer Methods Eng 90(1):40–64

    Article  MathSciNet  MATH  Google Scholar 

  14. Sanders JD, Laursen TA, Puso MA (2012) A Nitsche embedded mesh method. Comput Mech 49(2):243–257

    Article  MathSciNet  MATH  Google Scholar 

  15. Djoko JK, Ebobisse F, McBride AT, Reddy BD (2007) A discontinuous Galerkin formulation for classical and gradient plasticity—Part 1: formulation and analysis. Comput Methods Appl Mech Eng 196(37–40):3881–3897

    Article  MathSciNet  MATH  Google Scholar 

  16. Djoko JK, Ebobisse F, McBride AT, Reddy BD (2007) A discontinuous Galerkin formulation for classical and gradient plasticity. Part 2: algorithms and numerical analysis. Comput Methods Appl Mech Eng 197(1–4):1–21

    Article  MathSciNet  MATH  Google Scholar 

  17. Hansbo P (2010) A discontinuous finite element method for elasto-plasticity. Int J Numer Methods Biomed Eng 26(6):780–789

    MathSciNet  MATH  Google Scholar 

  18. Liu R, Wheeler MF, Dawson CN, Dean RH (2010) A fast convergent rate preserving discontinuous Galerkin framework for rate-independent plasticity problems. Comput Methods Appl Mech Eng 199(49):3213–3226

    Article  MathSciNet  MATH  Google Scholar 

  19. Liu R, Wheeler MF, Yotov I (2013) On the spatial formulation of discontinuous Galerkin methods for finite elastoplasticity. Comput Methods Appl Mech Eng 253:219–236

    Article  MathSciNet  MATH  Google Scholar 

  20. Noels L, Radovitzky R (2008) An explicit discontinuous Galerkin method for non-linear solid dynamics: formulation, parallel implementation and scalability properties. Int J Numer Methods Eng 74(9):1393–1420

    Article  MathSciNet  MATH  Google Scholar 

  21. Truster TJ, Masud A (2014) Primal interface formulation for coupling multiple PDEs: a consistent derivation via the variational multiscale method. Comput Methods Appl Mech Eng 268:194–224

    Article  MathSciNet  MATH  Google Scholar 

  22. Hlepas G, Truster T, Masud A (2014) A heterogeneous modeling method for porous media flows. Int J Numer Methods Fluids 75(7):487–518

    Article  MathSciNet  Google Scholar 

  23. Truster T, Chen P, Masud A (2015) On the algorithmic and implementational aspects of a discontinuous Galerkin method at finite strains. Comput Math Appl 70(6):1266–1289

    Article  MathSciNet  Google Scholar 

  24. Truster TJ, Chen P, Masud A (2015) Finite strain primal interface formulation with consistently evolving stabilization. Int J Numer Methods Eng 102(3–4):278–315

    Article  MathSciNet  Google Scholar 

  25. Masud A, Truster TJ (2013) A framework for residual-based stabilization of incompressible finite elasticity: stabilized formulations and methods for linear triangles and tetrahedra. Comput Methods Appl Mech Eng 267:359–399

    Article  MathSciNet  MATH  Google Scholar 

  26. Masud A, Truster TJ, Bergman LA (2011) A variational multiscale a posteriori error estimation method for mixed form of nearly incompressible elasticity. Comput Methods Appl Mech Eng 200(47–48):3453–3481

    Article  MathSciNet  MATH  Google Scholar 

  27. Masud A, Truster TJ, Bergman LA (2012) A unified formulation for interface coupling and frictional contact modeling with embedded error estimation. Int J Numer Methods Eng 92(2):141–177

    Article  MathSciNet  Google Scholar 

  28. Stenberg R (1995) On some techniques for approximating boundary conditions in the finite element method. J Comput Appl Math 63(1–3):139–148

    Article  MathSciNet  MATH  Google Scholar 

  29. Kiran R, Khandelwal K (2014) Complex step derivative approximation for numerical evaluation of tangent moduli. Comput Struct 140:1–13

    Article  Google Scholar 

  30. Miehe C (1996) Numerical computation of algorithmic (consistent) tangent moduli in large-strain computational inelasticity. Comput Methods Appl Mech Eng 134(3–4):223–240

    Article  MathSciNet  MATH  Google Scholar 

  31. Radovitzky R, Seagraves A, Tupek M, Noels L (2011) A scalable 3D fracture and fragmentation algorithm based on a hybrid, discontinuous Galerkin, cohesive element method. Comput Methods Appl Mech Eng 200(1–4):326–344

    Article  MathSciNet  Google Scholar 

  32. Truster TJ, Masud A (2013) A Discontinuous/continuous Galerkin method for modeling of interphase damage in fibrous composite systems. Comput Mech 52(3):499–514

    Article  MathSciNet  MATH  Google Scholar 

  33. Eyck AT, Celiker F, Lew A (2008) Adaptive stabilization of discontinuous Galerkin methods for nonlinear elasticity: analytical estimates. Comput Methods Appl Mech Eng 197(33–40):2989–3000

    Article  MathSciNet  MATH  Google Scholar 

  34. Hughes TJR (1995) Multiscale phenomena: Green’s functions, the Dirichlet-to-Neumann formulation, subgrid scale models, bubbles and the origins of stabilized methods. Comput Methods Appl Mech Eng 127(1–4):387–401

    Article  MATH  Google Scholar 

  35. Calderer R, Masud A (2012) Residual-based variational multiscale turbulence models for unstructured tetrahedral meshes. Comput Methods Appl Mec Eng 254:238–253

    Article  MathSciNet  Google Scholar 

  36. Oden JT, Babuška I, Nobile F, Feng Y, Tempone R (2005) Theory and methodology for estimation and control of errors due to modeling, approximation, and uncertainty. Comput Methods Appl Mech Eng 194(2–5):195–204

    Article  MATH  Google Scholar 

  37. Oden JT, Prudhomme S, Romkes A, Bauman P (2006) MultiScale modeling of physical phenomena: adaptive control of models. SIAM J Sci Comput 28(6):2359–2389

    Article  MathSciNet  MATH  Google Scholar 

  38. Prudhomme S, Chamoin L, Dhia HB, Bauman PT (2009) An adaptive strategy for the control of modeling error in two-dimensional atomic-to-continuum coupling simulations. Comput Methods Appl Mech Eng 198(21–26):1887–1901

    Article  MATH  Google Scholar 

  39. Lubliner J (1990) Plasticity theory. Macmillan Publishing Company, New York

    MATH  Google Scholar 

Download references

Acknowledgments

T. Truster was supported by faculty startup funds from the University of Tennessee.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Timothy J. Truster.

Appendices

Appendix 1: Stationary conditions

A sketch of a proof is provided to show that the stationary conditions of the discrete total free energy functional (25) are equivalent to the standard Galerkin weak form (19)–(21) combined with the constitutive laws (22)–(24). Note that the non-smooth nature of the consistency condition (9) suggests that a rigorous proof requires techniques for variational inequalities or constrained optimization; see [2].

These stationary conditions are defined through the variation of \(\widehat{\mathcal {P}}_n \) with respect to each primary field:

$$\begin{aligned} \hbox {D}_u \widehat{\mathcal {P}}_n \cdot {\varvec{\eta }} ^{h}+\hbox {D}_p \widehat{\mathcal {P}}_n \cdot q^{h}+\hbox {D}_\lambda \widehat{\mathcal {P}}_n \cdot {\varvec{\mu }} ^{h}=0 \end{aligned}$$
(65)

The variations of (26)–(29) with respect to the displacement field are obtained as follows, accounting for the symmetry of \(\overline{\mathbf{C}}\) and \(\mathbf{D}\):

$$\begin{aligned} \hbox {D}_u \mathcal {P}_{\mathrm{int}} \cdot {\varvec{\eta }}^{h}= & {} \int _\Omega \left[ {{\varvec{\varepsilon }}\left( {{\varvec{\eta }}^{h}} \right) -\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right) \cdot {\varvec{\eta }} ^{h}} \right] :\nonumber \\&\overline{\mathbf{C}}:\left[ {{\varvec{\varepsilon }} \left( {{\varvec{u}}_{n+1}^h } \right) -{\varvec{\varepsilon }}_n^p -\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right] \;\hbox {d}V\nonumber \\&+\int _\Omega {{\varvec{\varepsilon }} \left( {{\varvec{\eta }} ^{h}} \right) :p_{n+1}^h {\varvec{1}}\;\hbox {d}V} \nonumber \\&+\int _\Omega {\left[ {-\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \cdot \mathbf{D}^{-1} }\nonumber \\&\ \cdot \left[ {{\varvec{q}}_n -\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right] \;\hbox {d}V \end{aligned}$$
(66)
$$\begin{aligned} \hbox {D}_u {\mathcal {P}}_{\mathrm{ext}} \cdot {\varvec{\eta }}^{h}=-\int _\Omega {{\varvec{\eta }} ^{h}\cdot {\varvec{b}}_{n+1}^h \;\hbox {d}V} -\int _{\Gamma _h } {{\varvec{\eta }} ^{h}\cdot {\varvec{t}}_{n+1}^h \;\hbox {d}V} \end{aligned}$$
(67)
$$\begin{aligned} \hbox {D}_u \mathcal {P}_I \cdot {\varvec{\eta }} ^{h}= & {} -\int _{\Gamma _I } {{\varvec{\lambda }} _{n+1}^h \cdot \llbracket {{\varvec{\eta }}^{h}} \rrbracket \;\hbox {d}A} \end{aligned}$$
(68)
$$\begin{aligned}&\hbox {D}_u \Delta \mathcal{L}^{p}\cdot {\varvec{\eta }}^{h}\nonumber \\&\quad =\int _\Omega \left[ {\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] :\overline{\mathbf{C}}:\nonumber \\&\qquad \left[ {{\varvec{\varepsilon }} \left( {{\varvec{u}}_{n+1}^h } \right) -{\varvec{\varepsilon }}_n^p -\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right] \;\hbox {d}V\nonumber \\&\qquad +\int _\Omega \left[ {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right] :\overline{\mathbf{C}}:\nonumber \\&\qquad \left[ {{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) -\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \;\hbox {d}V \nonumber \\&\qquad -\int _\Omega \left[ {-\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \cdot \mathbf{D}^{-1} \nonumber \\&\qquad \cdot \left[ {-\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right] \;\hbox {d}V-\int _\Omega \left[ {{\varvec{q}}_n -\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right] \cdot \mathbf{D}^{-1}\nonumber \\&\qquad \cdot \left[ -\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h} \right] \;\hbox {d}V \nonumber \\&\qquad -\int _\Omega {\Delta \gamma _{n+1} \left( {\hbox {D}_u f_{n+1} \cdot {\varvec{\eta }}^{h}} \right) \;\hbox {d}V}\nonumber \\&\qquad -\int _\Omega {\left( {\hbox {D}_u \Delta \gamma _{n+1} \cdot {\varvec{\eta }}^{h}} \right) f_{n+1} \;\hbox {d}V} \end{aligned}$$
(69)

The variation of (26) with respect to the pressure field is given by:

$$\begin{aligned} \hbox {D}_p \mathcal {P}_{\mathrm{int}} \cdot q^{h}=\int _\Omega {{\varvec{\varepsilon }} \left( {{\varvec{u}}_{n+1}^h } \right) :q^{h}{\varvec{1}}\;\hbox {d}V} -\int _\Omega {{q^{h}p_{n+1}^h }/K\;\hbox {d}V}\nonumber \\ \end{aligned}$$
(70)

Lastly, the variation of (28) with respect to the Lagrange multiplier field follows as:

$$\begin{aligned} \hbox {D}_\lambda \mathcal {P}_I \cdot {\varvec{\mu }}^{h}=-\int _{\Gamma _I } {{\varvec{\mu }}^{h}\cdot \llbracket {{\varvec{u}}_{n+1}^h } \rrbracket \;\hbox {d}A} \end{aligned}$$
(71)

Focusing on the variation of the time-discrete yield function \(f_{n+1} \) in (69), applying the chain rule to definition (9) and accounting for the flow rule and hardening law (22)–(24) along with the associate plasticity conditions (10)–(11) results in the expression below:

$$\begin{aligned} \hbox {D}_u f_{n+1} \cdot {\varvec{\eta }} ^{h}= & {} \frac{\partial f_{n+1} }{\partial {\varvec{\sigma }} _{n+1} }:\hbox {D}_u {\varvec{\sigma }}_{n+1} \cdot {\varvec{\eta }} ^{h}\!+\!\frac{\partial f_{n+1} }{\partial {\varvec{q}}_{n+1} }\cdot \hbox {D}_u {\varvec{q}}_{n+1} \cdot {\varvec{\eta }}^{h} \nonumber \\= & {} {\varvec{r}}_{n+1} :\overline{\mathbf{C}}:\left[ {{\varvec{\varepsilon }} \left( {{\varvec{\eta }} ^{h}} \right) -\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \nonumber \\&-\,{\varvec{h}}_{n+1} \cdot \mathbf{D}^{-1}\cdot \left[ {\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \end{aligned}$$
(72)

Substituting (72) into (69) and cancelling like terms yields:

$$\begin{aligned} \hbox {D}_u \Delta \mathcal{L}^{p}\cdot {\varvec{\eta }}^{h}= & {} \int _\Omega \left[ \hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}\right] :\overline{\mathbf{C}}:\nonumber \\&\left[ {{\varvec{\varepsilon }} \left( {{\varvec{u}}_{n+1}^h } \right) -{\varvec{\varepsilon }}_n^p -\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right] \;\hbox {d}V \nonumber \\&-\int _\Omega \left[ {{\varvec{q}}_n -\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right] \cdot \mathbf{D}^{-1}\nonumber \\&\cdot \left[ {-\hbox {D}_u \left( {\Delta \gamma _{n+1} {\varvec{h}}_{n+1} } \right) \cdot {\varvec{\eta }}^{h}} \right] \;\hbox {d}V \nonumber \\&-\int _\Omega {\left( {\hbox {D}_u \Delta \gamma _{n+1} \cdot {\varvec{\eta }} ^{h}} \right) f_{n+1} \;\hbox {d}V} \end{aligned}$$
(73)

Combining (73) with (66)–(68) provides the variation of \(\widehat{\mathcal {P}}_n \) with respect \({\varvec{u}}_{n+1}^h\):

$$\begin{aligned}&\hbox {D}_u \left( {\mathcal {P}_{\mathrm{int}} +\mathcal {P}_{\mathrm{ext}} +\mathcal {P}_I +\Delta \mathcal{L}^{p}} \right) \cdot {\varvec{\eta }}^{h}=\int _\Omega \left[ {{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) } \right] :\nonumber \\&\qquad \overline{\mathbf{C}}:\left[ {{\varvec{\varepsilon }} \left( {{\varvec{u}}_{n+1}^h } \right) -{\varvec{\varepsilon }}_n^p -\Delta \gamma _{n+1} {\varvec{r}}_{n+1} } \right] \;\hbox {d}V \nonumber \\&\quad +\int _\Omega {{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) :p_{n+1}^h {\varvec{1}}\;\hbox {d}V} -\int _\Omega {{\varvec{\eta }}^{h}\cdot {\varvec{b}}_{n+1}^h \;\hbox {d}V} \nonumber \\&\quad -\int _{\Gamma _h } {{\varvec{\eta }}^{h}\cdot {\varvec{t}}_{n+1}^h \;\hbox {d}V} -\int _{\Gamma _I } {{\varvec{\lambda }}_{n+1}^h \cdot \llbracket {{\varvec{\eta }}^{h}} \rrbracket \;\hbox {d}A} \nonumber \\&\quad -\int _\Omega {\left( {\hbox {D}_u \Delta \gamma _{n+1} \cdot {\varvec{\eta }}^{h}} \right) f_{n+1} \;\hbox {d}V} \end{aligned}$$
(74)

Finally, taking note that either \(f_{n+1} =0\) or \(\Delta \gamma _{n+1} =\hbox {D}_u \Delta \gamma _{n+1} \cdot {\varvec{\eta }}^{h}=0\) according to the consistency condition (9), the last term in (74) vanishes identically.

Hence, the stationary conditions of \(\widehat{\mathcal {P}}_n \) subject to the constraints (22)–(24) are given by (74), (70), and (71), which are identical to the weak form (19), (20), and (21), thereby completing the proof.

Appendix 2: Consistent linearization

We briefly summarize the consistent linearization of the elasto-plastic interface method (39)–(40) for use in nonlinear solution strategies such as the Newton–Raphson algorithm. Taking the variational derivative of (39) yields the following expression:

$$\begin{aligned}&K_{uu} \left( {{\varvec{\eta }} ^{h},\Delta {\varvec{u}}^{h};{\varvec{u}}_{n+1}^h } \right) :=\hbox {D}_u \left[ {R_u \left( {{\varvec{\eta }}^{h};{\varvec{u}}_{n+1}^h ,p_{n+1}^h } \right) } \right] \cdot \Delta {\varvec{u}}^{h} \nonumber \\&\quad =\sum _{\alpha =1}^2 {\int _{\Omega ^{(\alpha )}} {{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{(\alpha )}} \right) : \overline{\mathbf{C}}_{n+1}^{(\alpha )} : {\varvec{\varepsilon }} \left( {\Delta {\varvec{u}}^{(\alpha )}} \right) \;\hbox {d}V} } \nonumber \\&\qquad -\sum _{\alpha =1}^2 {\int _{\Gamma _I } {\llbracket {{\varvec{\eta }}^{h}} \rrbracket \cdot \hbox {D}_u \left[ {\left\{ {{\varvec{\sigma }}_{n+1} \cdot {\varvec{n}}} \right\} } \right] \cdot \Delta {\varvec{u}}^{(\alpha )}\;\hbox {d}A} } \nonumber \\&\qquad -\sum _{\alpha =1}^2 \int _{\Gamma _I } \left\{ {\left( {\overline{\mathbf{C}}_{n+1} : {\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) } \right) \cdot {\varvec{n}}} \right\} \cdot \hbox {D}_u \left[ {\llbracket {{\varvec{u}}_{n+1}^h } \rrbracket } \right] \nonumber \\&\qquad \cdot \, \Delta {\varvec{u}}^{(\alpha )}\;\hbox {d}A \nonumber \\&\qquad -\sum _{\alpha =1}^2 \int _{\Gamma _I } \left\{ {\left( {\left( {\hbox {D}_u \left[ {\overline{\mathbf{C}}_{n+1} } \right] \cdot \Delta {\varvec{u}}^{(\alpha )}} \right) : {\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) } \right) \cdot {\varvec{n}}} \right\} \nonumber \\&\qquad \cdot \, \llbracket {{\varvec{u}}_{n+1}^h } \rrbracket \;\hbox {d}A \nonumber \\&\qquad +\sum _{\alpha =1}^2 {\int _{\Gamma _I } {\llbracket {{\varvec{\eta }}^{h}} \rrbracket \cdot {\varvec{\tau }}_s \cdot \hbox {D}_u \left[ {\llbracket {{\varvec{u}}_{n+1}^h } \rrbracket } \right] \cdot \Delta {\varvec{u}}^{(\alpha )}\;\hbox {d}A} } \end{aligned}$$
(75)

The linearization of the average stress terms mirrors (36) by replacing \({\varvec{\eta }}^{h}\) with \(\Delta {\varvec{u}}^{h}\), and the jump term follows simply by replacing \({\varvec{u}}_{n+1}^h \) with \(\Delta {\varvec{u}}^{h}\) due to linearity. However, the last term in (75) involves the linearization of the algorithmic tangent moduli \(\overline{\mathbf{C}}_{n+1} \). Proceeding as in [24], this contribution yields the 6th order tensor of algorithmic tangent moduli denoted as \({\overline{\varvec{\Xi }}}_{n+1}\):

$$\begin{aligned}&\hbox {D}_u \left[ {\overline{\mathbf{C}}_{n+1}^{(\alpha )} } \right] \cdot \Delta {\varvec{u}}^{(\alpha )}={\overline{\varvec{\Xi }}}_{n+1}^{(\alpha )} : {\varvec{\varepsilon }} \left( {\Delta {\varvec{u}}^{(\alpha )}} \right) \end{aligned}$$
(76)
$$\begin{aligned}&{\overline{\varvec{\Xi }}}_{n+1}^{(\alpha )} =\frac{d^{2}{\varvec{\sigma }} _{n+1}^{(\alpha )} }{d{\varvec{\varepsilon }} _{n+1}^{(\alpha )} d{\varvec{\varepsilon }} _{n+1}^{(\alpha )} } \end{aligned}$$
(77)

This tensor in general possesses minor symmetries as well as major symmetry within the last two pairs of indices. For associative flow rules, the other possible major symmetries are also retained. Since this algorithmic tensor can be difficult to derive in closed-form, a numerical approximation procedure based on finite differences [30] is provided in Appendix 3. Recently, other approaches have been proposed [29] for evaluating tangent matrices that use complex variables to avoid numerical cancellation errors.

Making these substitutions into (75) yields the final linearized form:

$$\begin{aligned}&K_{uu} \left( {{\varvec{\eta }}^{h},\Delta {\varvec{u}}^{h};{\varvec{u}}_{n+1}^h } \right) =\int _\Omega {{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) :\overline{\mathbf{C}}_{n+1} :{\varvec{\varepsilon }} \left( {\Delta {\varvec{u}}^{h}} \right) \;\hbox {d}V} \nonumber \\&\quad -\int _{\Gamma _I } {\llbracket {{\varvec{\eta }}^{h}} \rrbracket \cdot \left\{ {\left[ {\overline{\mathbf{C}}_{n+1} :{\varvec{\varepsilon }} \left( {\Delta {\varvec{u}}^{h}} \right) } \right] \cdot {\varvec{n}}} \right\} \;\hbox {d}A} \nonumber \\&\quad -\int _{\Gamma _I } {\left\{ {\left[ {\overline{\mathbf{C}}_{n+1} :{\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) } \right] \cdot {\varvec{n}}} \right\} \cdot \llbracket {\Delta {\varvec{u}}^{h}} \rrbracket \;\hbox {d}A} \nonumber \\&\quad -\int _{\Gamma _I } {\left\{ {\left[ {\left( {{\overline{\varvec{\Xi }}}_{n+1} : {\varvec{\varepsilon }} \left( {\Delta {\varvec{u}}^{h}} \right) } \right) : {\varvec{\varepsilon }} \left( {{\varvec{\eta }}^{h}} \right) } \right] \cdot {\varvec{n}}} \right\} \cdot \llbracket {{\varvec{u}}_{n+1}^h } \rrbracket \;\hbox {d}A} \nonumber \\&\quad +\int _{\Gamma _I } {\llbracket {{\varvec{\eta }}^{h}} \rrbracket \cdot {\varvec{\tau }}_s \cdot \llbracket {\Delta {\varvec{u}}^{h}} \rrbracket \;\hbox {d}A} \end{aligned}$$
(78)

Upon inspection, the bilinear form (78) is symmetric in the arguments \({\varvec{\eta }}^{h}\) and \(\Delta {\varvec{u}}^{h}\) by taking account of the assumed associative flow rule. As remarked in [23], the higher-order moduli term can typically be neglected without significant impact to the Newton–Raphson convergence rate due to the presence of the displacement jump \(\llbracket {{\varvec{u}}_{n+1}^h} \rrbracket \approx \mathbf{0}\).

The contributions from the pressure field are also straightforward due to linearity:

(79)
(80)
(81)

In conclusion, the consistent linearization (78)–(81) is symmetric, which is a direct consequence of the existence of a potential functional (30) and implies that the VMDG formulation (39)–(40) possesses variational and adjoint consistency. These consistency properties are numerically verified in Sect. 5.1; the correctness of (78)–(81) is attested by the quadratic rate of convergence of the Newton–Raphson algorithm in Sect. 5.2.

Appendix 3: Numerical evaluation of \({\overline{\varvec{\Xi }}}_{n+1}\)

Herein, we outline a numerical procedure for evaluating the 6th order tensor of material moduli \({\overline{\varvec{\Xi }}}_{n+1}\). Similarly, computational methods have been proposed for evaluating the algorithmic tangent moduli \(\overline{\mathbf{C}}_{n+1} \) as an alternative to analytical methods; see for example [29, 30]. In the former, a numerical differentiation scheme is advocated that consists of multiple evaluations of the stress update routine with perturbed strain increments. Namely, each “column” of \(\overline{\mathbf{C}}_{n+1} \) is computed from differencing the algorithmic stress resulting from the actual strain \({\varvec{\varepsilon }}\) and the perturbed strain \(\tilde{{\varvec{\varepsilon }} }\). We extend the concept to compute each generalized “column” of \({\overline{\varvec{\Xi }}}_{n+1}\) by differencing the perturbed algorithmic tangent moduli \(\overline{\mathbf{C}}_{n+1} \). For three dimensional analyses, the cost of this procedure is effectively six additional function-calls to the stress-update subroutine \(\mathbf{C}^{\mathrm{algo}}\). The algorithm, presented in Table 6, is quite general and can be extended to the geometrically nonlinear regime using ideas from [30]. As demonstrated in Sect. 5.2, the numerical version of \({\overline{\varvec{\Xi }}}_{n+1}\) only slightly reduces the quadratic rate of the Newton–Raphson method. Herein, the perturbation size \(\epsilon =10^{-8}\varepsilon _{\left( {mn} \right) } \) as recommended by [30].

Table 6 Numerical computation of 6th order material moduli tensor

Appendix 4: Constitutive tensors for von Mises plasticity

We provide a complete characterization of the material tensors \({\varvec{\sigma }}_{n+1} \), \(\overline{\mathbf{C}}_{n+1} \), \({\overline{\varvec{\Xi }}}_{n+1}\) corresponding to the von Mises model with kinematic hardening from Sect. 5.2. Isotropic elastic response is assumed according to the model given in Sect. 2. The elastic predictor–plastic corrector scheme is adopted from [2]. The trial yield function and flow stress at step \(n+1\) are as follows:

$$\begin{aligned} f_{n+1}^{tr}= & {} \left\| {{\varvec{\eta }}_{n+1}^{tr} } \right\| -\sqrt{\frac{2}{3}}\sigma _Y , \nonumber \\ {\varvec{\eta }} _{n+1}^{tr}= & {} \hbox {dev}\left[ {{\varvec{\sigma }}_n } \right] +2G\left( {\hbox {dev}\left[ {{\varvec{\varepsilon }}_{n+1} -{\varvec{\varepsilon }}_n } \right] } \right) -\overline{\varvec{\beta }}_n \end{aligned}$$
(82)

The unit normal tensor to the yield surface is given by:

$$\begin{aligned} {\varvec{N}}={{\varvec{\eta }} _{n+1}^{tr} }\big /{\left\| {{\varvec{\eta }}_{n+1}^{tr} } \right\| } \end{aligned}$$
(83)

Assuming plastic flow \(f_{n+1}^{tr} >0\), the incremental consistency parameter follows from [2] as:

$$\begin{aligned} \Delta \gamma =\frac{f_{n+1}^{tr} }{2G\left( {1+H/{3G}} \right) } \end{aligned}$$
(84)

Thus, the algorithmic stress at step \(n+1\) takes the following form:

$$\begin{aligned} {\varvec{\sigma }}_{n+1}= & {} p_{n+1} {\varvec{I}}+\hbox {dev}\left[ {{\varvec{\sigma }}_n } \right] \nonumber \\&+\,2G\left( {\hbox {dev}\left[ {{\varvec{\varepsilon }}_{n+1} -{\varvec{\varepsilon }} _n } \right] } \right) -2G\Delta \gamma {\varvec{N}} \end{aligned}$$
(85)

We record the results from [2] for the derivatives of the unit normal (83) and consistency parameter (84):

$$\begin{aligned} \frac{\partial N_{ij} }{\partial \varepsilon _{kl} }= & {} \frac{\partial N_{ij} }{\partial \eta _{mn}^{tr} }\frac{\partial \eta _{mn}^{tr} }{\partial \varepsilon _{kl} }={2G}/{\left\| {{\varvec{\eta }}_{n+1}^{tr} } \right\| } \nonumber \\&\left[ {\frac{1}{2}\left( {\delta _{ik} \delta _{jl} +\delta _{il} \delta _{jk} } \right) -\frac{1}{3}\delta _{ij} \delta _{kl} -N_{ij} N_{kl} } \right] \end{aligned}$$
(86)
$$\begin{aligned} \frac{\partial \Delta \gamma }{\partial \varepsilon _{ij} }= & {} \left( {1+H/{3G}} \right) ^{-1}N_{ij} \end{aligned}$$
(87)

The algorithmic tangent moduli is obtained by directly differentiating the algorithmic stress (85) accounting for (86) and (87):

$$\begin{aligned} \frac{d\sigma _{ij} }{d\varepsilon _{kl} }\,:=\,\overline{\mathrm{C}}_{ijkl}= & {} 2G\left[ {\frac{1}{2}\left( {\delta _{ik} \delta _{jl} +\delta _{il} \delta _{jk} } \right) -\frac{1}{3}\delta _{ij} \delta _{kl} } \right] \nonumber \\&-\,2G\left[ {\frac{\partial \Delta \gamma }{\partial \varepsilon _{kl} }N_{ij} +\Delta \gamma \frac{\partial N_{ij} }{\partial \varepsilon _{kl} }} \right] \end{aligned}$$
(88)

Similarly, the 6th order material tensor \({\overline{\varvec{\Xi }}}_{n+1}\) is derived by differentiating (88). To do so, the second derivatives of the unit normal tensor and consistency parameter are required. Following directly from (86) and (87):

$$\begin{aligned} \frac{\partial ^{2}N_{ij} }{\partial \varepsilon _{kl} \partial \varepsilon _{mn} }= & {} -2G\left[ {{N_{ij} }/{\left\| {{\varvec{\eta }}_{n+1}^{tr} } \right\| }\frac{\partial N_{kl} }{\partial \varepsilon _{mn} }+{N_{kl} }/{\left\| {{\varvec{\eta }} _{n+1}^{tr} } \right\| }\frac{\partial N_{ij} }{\partial \varepsilon _{mn} } } \right. \nonumber \\&\left. +\,{N_{mn} }/{\left\| {{\varvec{\eta }} _{n+1}^{tr} } \right\| } \frac{\partial N_{ij} }{\partial \varepsilon _{kl} } \right] \end{aligned}$$
(89)
$$\begin{aligned} \frac{\partial ^{2}\Delta \gamma }{\partial \varepsilon _{ij} \partial \varepsilon _{kl} }= & {} \left( {1+H/{3G}} \right) ^{-1}\frac{\partial N_{ij} }{\partial \varepsilon _{kl} } \end{aligned}$$
(90)

Observe that expressions (86), (87), and (90) imply that \(N_{ij} \frac{\partial ^{2}\Delta \gamma }{\partial \varepsilon _{kl} \partial \varepsilon _{mn} }=\frac{\partial N_{kl} }{\partial \varepsilon _{mn} }\frac{\partial \Delta \gamma }{\partial \varepsilon _{ij} }\). Thus, we employ the results from (89) and (90) to obtain the definition for the 6th order tensor by differentiating (88):

$$\begin{aligned} \frac{\partial ^{2}\sigma _{ij} }{\partial \varepsilon _{kl} \partial \varepsilon _{mn} }\equiv & {} \overline{{\Xi }}_{ijklmn} \nonumber \\= & {} -2G\left[ \Delta \gamma \frac{\partial ^{2}N_{ij} }{\partial \varepsilon _{kl} \partial \varepsilon _{mn} }+\frac{\partial N_{kl} }{\partial \varepsilon _{mn} }\frac{\partial \Delta \gamma }{\partial \varepsilon _{ij} }+\frac{\partial N_{ij} }{\partial \varepsilon _{mn} }\frac{\partial \Delta \gamma }{\partial \varepsilon _{kl} }\right. \nonumber \\&\left. +\frac{\partial N_{ij} }{\partial \varepsilon _{mn} }\frac{\partial \Delta \gamma }{\partial \varepsilon _{kl} } +\frac{\partial N_{ij} }{\partial \varepsilon _{kl} }\frac{\partial \Delta \gamma }{\partial \varepsilon _{mn} } \right] \end{aligned}$$
(91)

Note that \({\overline{\varvec{\Xi }}}_{n+1}\) possesses all possible major symmetries due to the associative nature of the flow rule and minor symmetries due to the symmetry of the strain tensor.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Truster, T.J. A stabilized, symmetric Nitsche method for spatially localized plasticity. Comput Mech 57, 75–103 (2016). https://doi.org/10.1007/s00466-015-1222-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-015-1222-6

Keywords

Navigation