Skip to main content
Log in

Numerical solution for plasticity models using consistency bisection and a transformed-space closest-point return: a nongradient solution method

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

Abstract

A new approach is presented for computing the return in numerical solutions for computational plasticity models that ensures convergence through bisection of the plastic consistency parameter, while using a transformed-space closest-point return based on a geometric search that eliminates the need to compute gradients of the yield function or a consistent tangent operator. Numerical solution of the governing equations for computational plasticity is highly-nontrivial for complex constitutive laws. In particular for geomaterials, a predictive model may account for nonlinear elasticity, shear strength that depends nonlinearly on pressure and Lode angle, and nonlinear evolution models for internal variables such as porosity or pore pressure. Traditional gradient-based integration methods may perform poorly when the hardening laws are highly nonlinear or when the yield function has an ill-defined or cumbersome gradient because of high curvature, vertices, or complicated functional form. The application of this new approach to geomaterial modeling is described, along with verification benchmarks.

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

Similar content being viewed by others

Notes

  1. This assumes no elastic–plastic coupling, the effect of which is discussed in Sect. 2.2.7.

  2. The Arenisca user’s guide [19] provides input parameters and a description of the model features. The recent publication [20] shows the model development and application. Both of these related references omit the details of the numerical solution approach, which are the substance of the current manuscript.

References

  1. Anandarajah A (2010) Computational methods in elasticity and plasticity: solids and porous media. Springer, New York

    Book  Google Scholar 

  2. Armero F, Pérez-Foguet A (2002) On the formulation of closest-point projection algorithms in elastoplasticity part I: the variational structure. Int J Num Methods Eng 53(2):297–329

    Article  MATH  Google Scholar 

  3. Brannon R Plasticity verification suite. Unpublished

  4. Brannon R, Fossum A, Strack O (2009) Kayenta: theory and user’s guide. Sandia Rep 2009(2282):393–397

    Google Scholar 

  5. Brannon R, Leelavanichkul S (2010) A multi-stage return algorithm for solving the classical damage component of constitutive models for rocks, ceramics, and other rock-like media. Int J Fract 163(1–2):133–149

    Article  MATH  Google Scholar 

  6. Brannon RM (2007) Elements of phenomenological plasticity: geometrical insight, computational algorithms, and topics in shock physics. ShockWave science and technology reference library. Springer, BErlin, pp 225–274

    Chapter  Google Scholar 

  7. Burdon RL, Faires JD (2000) Numerical analysis, 9th edn. Richard Stratton, Boston

    Google Scholar 

  8. Burghardt JA (2011) Nonlocal plasticity, instability, and well-posedness of the elastic-plastic initial-boundary value problem. Ph.D. thesis, The University of Utah

  9. Chau K (2012) Analytic methods in geomechanics. CRC Press, Boca Raton

    Google Scholar 

  10. Collins I (2002) Associated and non-associated aspects of the constitutive laws for coupled elastic/plastic materials. Int J Geomech 2(2):259–267

    Article  Google Scholar 

  11. Coombs WM, Crouch RS (2011) Non-associated reuleaux plasticity: analytical stress integration and consistent tangent for finite deformation mechanics. Comput Methods Appl Mech Eng 200(9):1021–1037

    Article  MATH  MathSciNet  Google Scholar 

  12. Crouch RS, Askes H, Li T (2009) Analytical cpp in energy-mapped stress space: application to a modified drucker-prager yield surface. Comput Methods Appl Mech Eng 198(5):853–859

    Article  MATH  Google Scholar 

  13. De Borst R (1987) Integration of plasticity equations for singular yield functions. Comput Struct 26(5):823–829

    Article  MATH  Google Scholar 

  14. Fuller T, Brannon R (2011) On the thermodynamic requirement of elastic stiffness anisotropy in isotropic materials. Int J Eng Sci 49(4):311–321

    Article  MATH  MathSciNet  Google Scholar 

  15. Guilkey J, Harman T, Luitjens J, Schmidt J, Thornock J, Davison de St. Germain J, Shankar S, Peterson J, Brownlee C (2009) Uintah user guide version 1.1. SCI Institute, University of Utah UUSCI-2009-007

  16. Gurson AL (1975) Continuum theory of ductile rupture by void nucleation and growth. Part I. yield criteria and flow rules for porous ductile media. Tech. rep., Div. of Engineering, Brown University, Providence, RI (USA)

  17. Hill R (1998) The mathematical theory of plasticity, vol 11. Oxford University Press, Oxford

    MATH  Google Scholar 

  18. Hjiaj M, Fortin J, De Saxcé G (2003) A complete stress update algorithm for the non-associated drucker-prager model including treatment of the apex. Int J Eng Sci 41(10):1109–1143

    Article  Google Scholar 

  19. Homel MA, Brannon RM, Colovos JW, Guilkey JE. Arenisca: theory guide and user’s manual. University of Utah, Mechanical Engineering, Computational Solid Mechanics, 50 S. Central Campus Dr

  20. Homel MA, Guilkey J, Brannon RM (2015) Continuum effective-stress approach for high-rate plastic deformation of fluid-saturated geomaterials with application to shaped-charge jet penetration. Acta Mech. doi:10.1007/s00707-015-1407-2

  21. Jeffrey A, Dai HH (2008) Handbook of mathematical formulas and integrals. Academic Press, New York

    MATH  Google Scholar 

  22. Kachanov LM (1971) Foundations of the theory of plasticity, vol 12. North-Holland, Amsterdam

    MATH  Google Scholar 

  23. Koiter W (1953) Stress-strain relations, uniqueness and variational theorems for elastic-plastic materials with a singular yield surface. Q Appl Math 11(3):350–354

    MATH  MathSciNet  Google Scholar 

  24. Koiter WT (1960) General theorems for elastic-plastic solids. North-Holland, Amsterdam

    Google Scholar 

  25. Lagarias JC, Reeds JA, Wright MH, Wright PE (1998) Convergence properties of the nelder-mead simplex method in low dimensions. SIAM J Optim 9(1):112–147

    Article  MATH  MathSciNet  Google Scholar 

  26. Lewis RM, Torczon V, Trosset MW (2000) Direct search methods: then and now. J Comput Appl Math 124(1):191–207

    Article  MATH  MathSciNet  Google Scholar 

  27. Mandel J (1965) Generalisation de la theorie de plasticite de W. T. Koiter. Int J Solids Struct 1(3):273–295

    Article  Google Scholar 

  28. Pan J, Rice JR (1983) Rate sensitivity of plastic flow and implications for yield-surface vertices. Int J Solids Struct 19(11):973–987

    Article  MATH  Google Scholar 

  29. Rashid M (1993) Incremental kinematics for finite element applications. Int J Num Methods Eng 36(23):3937–3956

    Article  MATH  MathSciNet  Google Scholar 

  30. Simo J, Hughes T (1998) Computational Inelasticity. Interdisciplinary applied mathematics: mechanics and materials. Springer, New York

    Google Scholar 

  31. Simo JC (1988) A framework for finite strain elastoplasticity based on maximum plastic dissipation and the multiplicative decomposition. Part II: computational aspects. Comput Methods Appl Mech Eng 68(1):1–31

    Article  MATH  Google Scholar 

  32. Simo JC, Taylor RL (1985) Consistent tangent operators for rate-independent elastoplasticity. Comput Methods Appl Mech Eng 48(1):101–118

    Article  MATH  MathSciNet  Google Scholar 

  33. Sloan SW, Abbo AJ, Sheng D (2001) Refined explicit integration of elastoplastic models with automatic error control. Eng Comput 18(1/2):121–194

    Article  Google Scholar 

  34. Sloan SW, Booker JR (1986) Removal of singularities in Tresca and Mohr-Coulomb yield functions. Commun Appl Num Methods 2(2):173–179

    Article  Google Scholar 

  35. Stupkiewicz S, Denzer R, Piccolroaz A, Bigoni D (2014) Implicit yield function formulation for granular and rock-like materials. arXiv preprint arXiv:1405.0125

  36. Terzaghi K (1936) The shearing resistance of saturated soils and the angle between planes of shear. In: First international conference on soil mechanics and foundation engineering, Harvard University, vol. I, pp. 54–56

  37. Vorobiev OY, Liu B, Lomov I, Antoun T (2007) Simulation of penetration into porous geologic media. Int J Impact Eng 34(4):721–731

    Article  Google Scholar 

  38. Weng MC (2014) A generalized plasticity-based model for sandstone considering time-dependent behavior and wetting deterioration. Rock Mech Rock Eng 47:1197–1209

    Article  Google Scholar 

Download references

Acknowledgments

The authors are grateful for the support received from the Perforating Research Group within Schlumberger Technology, Ltd, Harvey Williams, Director, for computational resources provided by the University of Utah Center for High Performance Computing (CHPC), and for c++ coding expertise provided by James Colovos that aided in the implementation of the solution method.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael A. Homel.

Appendices

Appendix 1: Proof of the transformed-space closest-point return

From Simo and Hughes [30], “For an associative flow rule, \(\varvec{\sigma }^{\text {n+1}}\) is the closest point projection onto the yield surface of the trial elastic stress \(\varvec{\sigma }^{\text {trial}}\) in the inner product induced by the compliance tensor \({\mathbb {C}}^{-1}\)”, where the energy norm is defined

$$\begin{aligned} \Vert \sigma \Vert _{{\mathbb {C}}^{-1}}:=\sqrt{\sigma :{\mathbb {C}}^{-1}:\sigma }. \end{aligned}$$
(33)

This is equated to an \(L_2\) norm of some transformed tensor \(\sigma ^\star ={\mathbb {C}}^{-1/2}:\sigma \) so that

$$\begin{aligned} \Vert \sigma \Vert _{{\mathbb {C}}^{-1}}^2=\sigma :{\mathbb {C}}^{-1}:\sigma =\sigma ^\star :\sigma ^\star \end{aligned}$$
(34)

Writing \({\mathbb {C}}^{-1/2}\) in terms of its spectral decomposition, the transformed tensor is defined in terms of eigenvalues \(\lambda _k\) and eigenprojectors \({\mathbb {P}}_k\)

$$\begin{aligned} \sigma ^\star = \sum \limits _{k=1}^{m}\frac{1}{\sqrt{\lambda _k}}{\mathbb {P}}_k:\sigma , \end{aligned}$$
(35)

where m is the number of distinct unique eigenprojectors.

Under certain restrictions [14], an isotropic material has an isotropic elastic tangent stiffness, implying that it can be additively decomposed

$$\begin{aligned} {\mathbb {C}}^{-1} = \frac{1}{3K}{\mathbb {P}}^\text {iso} + \frac{1}{2G}{\mathbb {P}}^\text {symdev} \end{aligned}$$
(36)

so that

$$\begin{aligned} {\mathbb {C}}^{-1/2} = \frac{1}{\sqrt{3K}}{\mathbb {P}}^\text {iso} + \frac{1}{\sqrt{2G}}{\mathbb {P}}^\text {symdev} \end{aligned}$$
(37)

where \({\mathbb {P}}^\text {iso}\) and \({\mathbb {P}}^\text {symdev}\) are the isotropic and symmetric-deviatoric fourth-order eigenprojectors. Specifically,

$$\begin{aligned}&{\mathbb {P}}_{ijkl}^\text {iso} = \frac{1}{3}(\delta _{ij}\delta _{kl}) \end{aligned}$$
(38a)
$$\begin{aligned}&{\mathbb {P}}_{ijkl}^\text {symdev} = \frac{1}{2}(\delta _{ik} \delta _{jl}+\delta _{il}\delta _{jk})-\frac{1}{3}\delta _{ij}\delta _{kl} \end{aligned}$$
(38b)

The stress tensor may always be decomposed as

$$\begin{aligned} \sigma = r \mathbf {E}_r + z \mathbf {E}_z, \end{aligned}$$
(39)

where \({\mathbf {E}}_z = \mathbf {I}/\sqrt{3}\) and \(\mathbf {E}_r = \sigma ^\text {dev}/r\). Therefore, the transformed tensor is

$$\begin{aligned} \sigma ^\star&= {\mathbb {C}}^{-1/2}:\sigma \nonumber \\&= \left( \frac{1}{\sqrt{3K}}{\mathbb {P}}^\text {iso} + \frac{1}{\sqrt{2G}}{\mathbb {P}}^\text {symdev}\right) :(r \mathbf {E}_r + z \mathbf {E}_z) \end{aligned}$$
(40)

Using the identities \({\mathbb {P}}^\text {iso}:\mathbf {E}_r = \mathbf {0},\,{\mathbb {P}}^\text {iso}:\mathbf {E}_z = \mathbf {E}_z,\,{\mathbb {P}}^\text {symdev}:\mathbf {E}_z = \mathbf {0}\), and \({\mathbb {P}}^\text {symdev}:\mathbf {E}_r = \mathbf {E}_r\), the transformed stress can be written

$$\begin{aligned} \sigma ^\star&= \frac{z}{\sqrt{3K}}{\mathbb {P}}^\text {iso}:\mathbf {E}_z + \frac{r}{\sqrt{2G}}{\mathbb {P}}^\text {symdev}:\mathbf {E}_r \nonumber \\&= \frac{z}{\sqrt{3K}}\mathbf {E}_z + \frac{r}{\sqrt{2G}}\mathbf {E}_r. \end{aligned}$$
(41)

Equating this to \(\sigma ^\star = z^\star \mathbf {E}_z + r^\star \mathbf {E}_r\), the scaling of the axis is then \(r^\star = \frac{1}{\sqrt{2G}}r\) and \(z^\star = \frac{1}{\sqrt{3K}}z\).

To obtain the same change in angles, it is sufficient to transform a single axis to produce the same stretch ratio, giving

$$\begin{aligned} r'=\sqrt{\frac{3K}{2G}}r, \text { and } z'=z \end{aligned}$$
(42)

or

$$\begin{aligned} r'=r, \text { and } z'=\sqrt{\frac{2G}{3K}}z \end{aligned}$$
(43)

Throughout this manuscript, the former relationship Eq. (42) has been applied to scale the r invariant for a 2-D solution, or to scale the stress deviator for the 3-D and full-stress solutions, though it may be more efficient to use Eq. (43), as it would result in fewer multiplication operators to scale the isotropic stress rather than the stress deviator.

Appendix 2: Formulation of the simplified yield surface

This section gives the formulation for the simplified constitutive model described in Sect. 2.2.2 and Fig. 5. With this construction, stress space is divided into three regions (I) where the closest point is the vertex, (II) where the closest point is on the linear Drucker–Prager surface, (III) where the closest point is on the elliptical cap.

The ellipse is centered at \((z^a,0)\), and passes though \((z^x,0)\) with principal axis aligned with the \(r'\) and \(z'\) axis, aspect ratio \(\mathcal {R} =\frac{a}{b}\), and tangent to the linear Drucker–Prager surface at the point \(\left\{ z^{\kappa },r^{\kappa }\right\} \). The values of the vertex position \(\left( z^v\right) \) and cap position \(\left( z^x\right) \) are known, as is the slope of the linear surface \((\beta )\) and the aspect ratio of the ellipse, \(\mathcal {R} \).

The axis of the ellipse \((a \text { and } b)\), the coordinates of the branch point \(\left( z^{\kappa },r^{\kappa }\right) \), and definitions for the three return regions, can be expressed in terms of the known values.

The coordinates of the branch point:

$$\begin{aligned} z^{\kappa }= & {} \frac{\beta z^v}{\sqrt{\beta ^2+\mathcal {R} ^2}}-\frac{\beta z^x}{\sqrt{\beta ^2+\mathcal {R} ^2}}+z^x \end{aligned}$$
(44a)
$$\begin{aligned} r^{\kappa }= & {} \frac{\beta \left( \sqrt{\beta ^2+\mathcal {R} ^2}-\beta \right) \left( z^v-z^x\right) }{\sqrt{\beta ^2+\mathcal {R} ^2}} \end{aligned}$$
(44b)

The z-coordinate of the apex:

$$\begin{aligned} z^a=\frac{\beta \left( \sqrt{\beta ^2+\mathcal {R} ^2}-\beta \right) \left( z^v-z^x\right) +\mathcal {R} ^2 z^x}{{\mathcal {R}}^2} \end{aligned}$$
(45)

The principal axis of the ellipse:

$$\begin{aligned} a= & {} \frac{\beta \left( \sqrt{\beta ^2+\mathcal {R} ^2}-\beta \right) \left( z^v-z^x\right) }{\mathcal {R} ^2} \end{aligned}$$
(46a)
$$\begin{aligned} b= & {} \frac{\beta \left( \sqrt{\beta ^2+\mathcal {R} ^2}-\beta \right) \left( z^v-z^x\right) }{\mathcal {R}} \end{aligned}$$
(46b)

A piecewise function \(g\left( r',z',z^X,z^{v},\beta '\right) \) is defined at every point to be the distance in transformed and shifted stress space to the closest point on the yield surface.

In region I this is simply the distance to the vertex:

$$\begin{aligned} g^I=\sqrt{\left( z-z^v\right) ^2+r^2} \end{aligned}$$
(47)

In region II this is the distance from a point to a line.

$$\begin{aligned} g^{\text {II}}=\frac{r+\beta \left( z-z^v\right) }{\sqrt{1+\beta ^2}} \end{aligned}$$
(48)

In region III this is the distance from a point to an ellipse. This can be solved for analytically, but doing so requires evaluating the solution for a quartic polynomial, which is subject to significant numerical error. The error can be imaginary, which is problematic for implementation in a numerical simulation. Instead a simple numerical algorithm computes the closest point on the elliptical curve. For simplicity we present the algorithm in the coordinate system of the ellipse and consider a return from a point \((x_0,y_0)\) in the first quadrant.

The point on the ellipse is represented parametrically in terms of the angle \(\theta \) as \(x=a\cos (\theta )\) and \(y=b\sin (\theta )\), and we seek to find the value \(\theta \) to minimize the length \(l(\theta )=\sqrt{\left( x_0-x(\theta )\right) {}^2+\left( y_0-y(\theta )\right) {}^2}\). In the first quadrant, the solution will lie in the range \(0\le \theta \le \pi /2\), and can be efficiently computed with a Newton’s method.

Figure 5 shows the ideal flow function contours for the simplified yield surface, which allow a steepest descent to the closest point on the yield surface. Since the value of the flow function is the distance to the closest point, the updated stress for a nonhardening return can be evaluated exactly in a single step as

$$\begin{aligned} \varvec{\sigma }^{\text {n+1}} =\varvec{\sigma }^{\text {trial}}-\nabla g\left( \varvec{\sigma }^{\text {trial}} \right) g\left( \varvec{\sigma }^{\text {trial}} \right) . \end{aligned}$$
(49)

Appendix 3: Pseudocode for the computing the updated stress and plastic strain for a generalized 2-D nonhardening return

Inputs

  • \(\varvec{\sigma }^{\text {trial}}\), Trial stress tensor

  • \(\varvec{\sigma }^\text {n}\), Stress tensor at start of substep

  • \(\delta \varvec{\varepsilon }\), Increment in total strain (tensor)

  • \(X\), Hydrostatic compressive strength

  • \(\zeta \), Trace of isotropic backstress

  • \(K\), Elastic tangent bulk modulus

  • \(G\), Elastic tangent shear modulus

Outputs

  • \(\varvec{\sigma }^{\text {n+1}}\), Stress tensor from nonhardening return

  • \(\delta \varvec{\varepsilon }^\text {p,n+1}\), Increment in plastic strain (tensor) for the nonhardening return

Pseudocode

  1. 1.

    Compute invariants, for the trial stress \(\left( I_1^\text {trial},J_2^\text {trial},\mathbf {S}^\text {trial} \right) \), as well as for an interior point, which we will take as \(\left\{ I_1^0=\zeta +\frac{1}{2} \left( X + I_1^\mathbf{peak } \right) ,J_2^0=0\right\} \)

  2. 2.

    Transform the trial and interior points as follows, where \(\beta \) defines the degree of non-associativity.

    • \(z^{\text {trial}\star }=\frac{I_1^\text {trial}}{\sqrt{3}}\)

    • \(r^{\text {trial},\star } = \sqrt{2J_2^\text {trial}}\beta \sqrt{\frac{3K}{2G}}\)

    • \(z^{0 \star } = \frac{I_1^0}{\sqrt{3}}\)

    • \(r^{0,\star } = \sqrt{2J_2^0}\beta \sqrt{\frac{3K}{2G}}\)

  3. 3.

    Perform bisection in transformed space between \((z^{0\star },r^{0\star })\) and \((z^{\text {trial}\star },r^{\text {trial}\star })\) to find the new point on the yield surface \((z^{\text {new}\star },r^{\text {new}\star })\)

  4. 4.

    Perform a rotation of \(\left\{ z^{\text {new}\star },r^{\text {new}\star }\right\} \) about \(\left\{ z^{\text {Trial}\star },r^{\text {Trial}\star }\right\} \) until a new internal point is found, set this as \(\left\{ z^{0 \star },r^{0 \star }\right\} \)

    1. (a)

      \(n=0\)

    2. (b)

      \(n=n+1\)

    3. (c)

      \(\theta =\frac{\pi }{2}(-1)^{n+1} \left( \frac{1}{2}\right) ^{\text {Floor}\left( \frac{n+1}{2}\right) }\)

    4. (d)

      \(\left( \begin{array}{c} z^{\text {test}\star } \\ z^{\text {test}\star } \\ \end{array} \right) \!=\![Q]\cdot \left[ \left( \begin{array}{c} z^{\text {new}\star } \\ r^{\text {new}\star } \\ \end{array} \right) \!-\!\left( \begin{array}{c} z^{\text {trial}\star } \\ r^{\text {trial}\star } \\ \end{array} \right) \right] +\left( \begin{array}{c} z^{\text {trial}\star } \\ r^{\text {trial}\star } \\ \end{array} \right) \), where \([Q]=\left( \begin{array}{cc} \cos \theta &{} \quad -\sin \theta \\ \sin \theta &{} \quad \cos \theta \\ \end{array} \right) \)

      • IF the transformed yield function is plastic when evaluated at \((z^{\text {test}\star },r^{\text {test}\star })\), GOTO[4b]

      • ELSE, Set \(\left\{ z^{0 \star },r^{0 \star }\right\} =\left\{ z^{\text {new}\star },r^{\text {new}\star }\right\} \), GOTO[5]

  5. 5.

    Test for convergence:

    • IF \(\theta \le \text {TOL},\,\hbox {Solution is converged}\), GOTO[6]

    • ELSE \(\theta >\text {TOL},\,\hbox {Not converged}\), GOTO[3]

  6. 6.

    Solution converged, compute untransformed updated stress:

    • \(I_1^{\text {new}}=\sqrt{3} z^{\text {new}\star }\)

    • \(\sqrt{J_2^{\text {new}}}=\frac{(2 G) r^{\text {new}\star }}{\sqrt{2} \beta (3 K)}\)

    • \(\varvec{\sigma }^{\text {n+1}} =\frac{1}{3}I_1^\text {trial} \mathbf {I} + \frac{\sqrt{J_2^{\text {new}}}}{\sqrt{J_2^{\text {trial}}}}\mathbf {S}^\text {trial} \)

  7. 7.

    Compute increment in plastic strain for return: \(\delta \varvec{\varepsilon }^p =\delta \varvec{\varepsilon }- \left( \mathbb {C} ^{-1}\right) :\left( \varvec{\sigma }^{\text {n+1}}-\varvec{\sigma }^\text {n} \right) \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Homel, M.A., Guilkey, J.E. & Brannon, R.M. Numerical solution for plasticity models using consistency bisection and a transformed-space closest-point return: a nongradient solution method. Comput Mech 56, 565–584 (2015). https://doi.org/10.1007/s00466-015-1187-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-015-1187-5

Keywords

Navigation