Skip to main content
Log in

Shape gradients for the failure probability of a mechanic component under cyclic loading: a discrete adjoint approach

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

Abstract

This work provides a numerical calculation of shape gradients of failure probabilities for mechanical components using a first discretize, then adjoin approach. While deterministic life prediction models for failure mechanisms are not (shape) differentiable, this changes in the case of probabilistic life prediction. The probabilistic, or reliability based, approach thus opens the way for efficient adjoin methods in the design for mechanical integrity. In this work we propose, implement and verify a method for the numerical calculation of the shape gradients of failure probabilities for the failure mechanism low cycle fatigue, which applies to polycrystalline metal. Numerical examples range from a bended rod to a complex geometry from a turbo charger in 3D.

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

Similar content being viewed by others

Notes

  1. www.darwin.swri.org/.

  2. The slip directions s mentioned in the introduction are perpendicular to the normal \(\nu \) of the plane of densest packing, thus \(s\cdot \sigma \nu =0\) if \(\sigma =\sigma _0\mathcal{I}\), \(\sigma _0\in {\mathbb {R}}\), and no force acts on the slip system s [22].

  3. For further information see www.calculix.de.

References

  1. Agmon S, Douglis A, Nirenberg L (1964) Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions II. Commun Pure Appl Math XVII:35–92

    Article  MathSciNet  MATH  Google Scholar 

  2. Allaire G (2007) Numerical analysis and optimization. Oxford University Press, Oxford

    MATH  Google Scholar 

  3. Allaire G, Bonneter E, Francfort G, Jouve F (1997) Shape optimization by the homogenization method. Numer Math 76:27–68

    Article  MathSciNet  MATH  Google Scholar 

  4. Atkinson HV, Davies S (2000) Fundamental aspects of hot isostatic pressing: an overview. Metall Mater Trans A 31(12):2981–3000

    Article  Google Scholar 

  5. Babuška I, Sawlan Z, Scavino M, Szabó B, Tempone R (2018) Spatial Poisson processes for crack initiation. arXiv:1805.03433

  6. Bangerth W, Rannacher R (2003) Adaptive finite element methods for differential equations. Birkhäuser, Basel

    Book  MATH  Google Scholar 

  7. Bank RE, Xu J-C, Zheng B (2007) Superconvergent derivative recovery for Lagrange triangular elements of degree p on unstructured grids. SIAM J Numer Anal 45:2032–2046

    Article  MathSciNet  MATH  Google Scholar 

  8. Bäker M, Harders H, Rösler J (2007) Mechanical behaviour of engineering materials: metals, ceramics, polymers, and composites, German edition published by Teubner Verlag (Wiesbaden, 2006). Springer, Berlin, Heidelberg

    Google Scholar 

  9. Batdorf SB, Crosse JG (1974) A statistical theory for the fracture of brittle structures subject to nonuniform polyaxial stress. J Appl Mech 41:459–465

    Article  MATH  Google Scholar 

  10. Bolten M, Hahn C, Gottschalk H, Saadi M (2017) Numerical shape optimization to decrease the failure probability of ceramic structures. arXiv:1705.05776

  11. Bolten M, Gottschalk H, Schmitz S (2015) Minimal failure probability for ceramic design via shape control. J Optim Theory Appl 166:983–1001

    Article  MathSciNet  MATH  Google Scholar 

  12. Borrego LP, Abreu LM, Costa JM, Ferreira JM (2004) Analysis of low cycle fatigue in AlMgSi aluminium alloys. Eng Fail Anal 11:715–725

    Article  Google Scholar 

  13. Borzi A, Schulz V (2012) Computational optimization of systems governed by partial differential equations. SIAM series on computational engineering. SIAM, Philadelphia

    MATH  Google Scholar 

  14. Braess D (2007) Finite Elemente—Theorie, schnelle Löser und Anwendungen in der Elastizitätstheorie, 4th edn. Springer, Berlin

    MATH  Google Scholar 

  15. Ciarlet P (1988) Mathematical elasticity—volume I: three-dimensional elasticity, vol 20. Studies in mathematics and its applications. North-Holland, Amsterdam

    MATH  Google Scholar 

  16. Ern A, Guermond J-L (2004) Theory and practice of finite elements. Springer, New York

    Book  MATH  Google Scholar 

  17. Escobar LA, Meeker WQ (1998) Statistical methods for reliability data. Wiley, New York

    MATH  Google Scholar 

  18. Munz D, Fett T (2012) Ceramics, engl. edn. Springer, Berlin

    Google Scholar 

  19. Gottschalk H, Saadi M, Doganay OT, Klamroth K, Schmitz S (2018) Adjoint method to calculate shape gradients of failure probabilaties for turbomachinery components. In: ASME-turbo-expo GT2018-75759

  20. Gottschalk H, Schmitz S (2014) Optimal reliability in design for fatigue life I: existence of optimal shapes. SIAM J Control Optim 52(5):2727–2752

    Article  MathSciNet  MATH  Google Scholar 

  21. Gottschalk H, Schmitz S, Seibel T, Krause R, Rollmann G, Beck T (2015) Probabilistic schmid factors and scatter of LCF life. Mater Sci Eng 46(2):156–164

    Google Scholar 

  22. Gottstein G (2004) Physical foundations of material science. Springer, Berlin

    Book  Google Scholar 

  23. Haslinger J (2003) Introduction to shape optimization—theory, approximation and computation. SIAM—advances in design and control. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  24. Hertel O, Vormwald M (2012) Statistical and geometrical size effects in notched members based on weakest-link and short-crack modelling. Eng Fract Mech 95:72–83

    Article  Google Scholar 

  25. Hoffmann M, Seeger T (1985) A generalized method for estimating elastic–plastic notch stresses and strains, part 1: theory. J Eng Mater Technol 107:250–254

    Article  Google Scholar 

  26. Kallenberg O (1983) Random measures. Akademie, Berlin

    MATH  Google Scholar 

  27. Knop M, Jones R, Molent L, Wang L (2000) On Glinka and Neuber methods for calculating notch tip strains under cyclic load spectra. Int J Fatigue 22:743–755

    Article  Google Scholar 

  28. Liu WB, Neittaanmäki P, Tiba D (2003) Existence for shape optimization problems in arbitrary dimension. SIAM J Control Optim 41:1440–1454

    Article  MathSciNet  MATH  Google Scholar 

  29. Leverant GR, Littlefield DL, McClung RC, Millwater HR, Wu JY (1997) A probabilistic approach to aircraft turbine rotor material design. In: Paper 97-GT-22, ASME turbo expo ’97, Orlando

  30. Mäde L, Schmitz S, Gottschalk H, Beck T (2018) Combined notch and size effect modeling in a local probabilistic approach for LCF. Comput Mater Sci 142:377388

    Article  Google Scholar 

  31. Neuber H (1961) Theory of stress concentration for shear-strained prismatical bodies with arbitrary nonlinear stress–strain law. J Appl Mech 26:544

    Article  MathSciNet  MATH  Google Scholar 

  32. Radaj D, Vormwald M (2007) Ermüdungsfestigkeit, 3rd edn. Springer, Berlin, Heidelberg

    Google Scholar 

  33. Ramberg W, Osgood WR (1943) Description of stress–strain curves by three parameters. Technical notes—National Advisory Committee for Aeronautics, No. 902, Washington

  34. Riesch-Oppermann H, Brückner-Foit A, Ziegler C (2000) STAU—a general purpose tool for probabilistic reliability assessment of ceramic components under multi axial loading. In: Proceedings of the 13th international conference on ECF 13, San Sebastian

  35. Schmitz S (2014) A probabilistic local model for low cycle fatigue—new aspects of structurel mechanics. Dissertation Lugano and Wuppertal 2014, appeared in Hartung-Gorre Verlag

  36. Schmitz S, Rollmann G, Gottschalk H, Krause R, Risk estimation for LCF crack initiation. In: Proceedings of ASME turbo expo 2013, GT2013-94899. arXiv:1302.2909v1

  37. Schmitz S, Rollmann G, Gottschalk H, Krause R (2013) Probabilistic analysis of the LCF crack initiation life for a turbine blade under thermo-mechanical loading. In: Proceedings of international conference on LCF 7

  38. Schmitz S, Seibel T, Beck T, Rollmann G, Krause R, Gottschalk H (2013) A probabilistic model for LCF. Comput Mater Sci 79:584–590

    Article  Google Scholar 

  39. Sokolowski J, Zolesio J-P (1992) Introduction to shape optimization—shape sensivity analysis, 1st edn. Springer, Berlin, Heidelberg

    Book  MATH  Google Scholar 

  40. Sornette D, Magnin T, Brechet Y (1992) The physical origin of the Coffin–Manson law in low-cycle fatigue. Europhys Lett 20:433–438

    Article  Google Scholar 

  41. Tröltzsch F (2010) Optimal control of partial differential equations (in German). Vieweg + Teubner, Braunschweig

    MATH  Google Scholar 

  42. Watanabe S (1964) On discontinuous additive functionals and Lévy measures of a Markov process. Jpn J Math 34:53–70

    Article  MATH  Google Scholar 

  43. Weibull EW (1939) A statistical theory of the strength of materials. Ingeniors Vetenskaps Akad Handl 151:1–45

    Google Scholar 

  44. Wittig K (1993) Construction of a gas turbine for model air planes (in German), Munich. www.calculix.de

Download references

Acknowledgements

We thank the Siemens Gas Turbine Engineering Department for Probabilistic Design for valuable discussions and support, in Dr. Georg Rollmann and Dr. Sebastian Schmitz in particular. We also thank Dr. T. Seibel (FZ Jülich), Prof. Dr. T. Beck (Technical University of Kaiserslautern) and R. Krause (ICS Lugano) for prior joint work on probabilistic LCF, on which this paper is based. This work has been made possible by financial support under the AG Turbo Grant 4.1.13 funded by Siemens Power and Gas, the German Federal Ministry of Economic Affairs (BMWi) under the Grant No. 03ET7041J and by the BMBF project GIVEN under the Grant No. 05M2018. Let us also thank both anonymous referees for their help in improving this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hanno Gottschalk.

Additional information

Publisher's Note

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

Detailed calculations

Detailed calculations

1.1 Notation for finite elements

To fix the notation for later purposes, we introduce the standard discretization of the PDE (1) with Lagrange finite elements [14, 16].

Let the finite element be given by \(\{K,P(K),{\varSigma }(K)\}\) where \(K\subseteq {\mathbb {R}}^3\) is a compact, connected Lipschitz set, P is a finite vector space polynomials \(p:K\rightarrow {\mathbb {R}}\). The local degrees of freedom are linear functionals on P(K) defined by \(\varphi _{K,j}(p)=p(X_j)\) where there are nodes \(X_{1}^K,\ldots ,X_{n_{\mathrm{sh}}}^K\in K\) such that the mapping from P(K) to its local degrees of freedom is bijective. The dual basis of P(K) with respect to the local degrees of freedom are the local shape functions \(\theta _{K,k}\in P(K)\) and \(\varphi _{K,j}(\theta _{K,k})=\delta _{j,k}\), \(j,k\in \{1,\ldots ,n_\mathrm{sh}\}\).

Let \({\mathcal {T}}_h\) be a finite element mesh on \({\varOmega }\) with \(N_{\mathrm{el}}\) elements. As usually we assume that there is a reference element \(\{{\widehat{K}},{\widehat{P}}, {\widehat{{\varSigma }}})\) with a finite dimensional linear space of reference polynomials \({\widehat{P}}\). For each element \(K\in {\mathcal {T}}_h \) in the mesh, we assume that there is a bijective transformation \(T_K:{{\widehat{K}}}\rightarrow K\) such that \({\widehat{P}}=P\circ T_K\) , \({{\widehat{\theta }}}_{j}=\theta _{j}\circ T_K\) and \({{\widehat{\varphi }}}_{j}(p\circ T_K)=\varphi _{j}(p)\), \(j\in \{1,\ldots ,n_{\mathrm{sh}}\}\). In many cases, we have

$$\begin{aligned} T_K({\hat{\xi }})=T_K({\hat{\xi }},X)=\sum _{j=1}^{n_\mathrm{sh}}{{\widehat{\theta }}}_j({\widehat{\xi }})X_{j}^K,~~{{\widehat{\xi }}}\in {\widehat{K}}. \end{aligned}$$
(26)

Let \(X=\{X_1,...,X_N\}=\bigcup _{K\in {\mathcal {T}}_h}\{X_{1}^K,...,X_{n_{\mathrm{sh}}}^K\}\) be the set of all the Lagrange nodes. For \(K\in {\mathcal {T}}_h\) and \(m\in \{1,...,n_\mathrm{sh}\}\), let \({\widehat{j}}(K,m)\in \{1,...,N\}\) be the corresponding index of the Lagrange node. The mapping \({\widehat{j}}(\cdot ,\cdot ):{\mathcal {T}}_h\times \{1,\ldots ,n_\mathrm{sh}\}\rightarrow \{1,\ldots ,N\}\) is also called the connectivity.

Let \(\{\theta _{j}:j\in \{1,\ldots ,N\}\}\), be the set of global shape functions \(\theta _{j}:{\varOmega }\rightarrow {\mathbb {R}}\). When these functions are restricted to \(K\in {\mathcal {T}}_h\), they belong to P(K) and fulfill

(27)

The global finite element space \(H_h^1({\varOmega },{\mathbb {R}})\) is the linear span of \(\{\theta _{j}:j\in \{1,\ldots ,N\}\}\). Let \(H^1_h({\varOmega },{\mathbb {R}}^3)=H_h^1({\varOmega },{\mathbb {R}})^ {\times 3}\). We have \(H_h^1({\varOmega },{\mathbb {R}}^3)\subseteq H^1({\varOmega },{\mathbb {R}}^3)\) and let \(H^1_{D,h}({\varOmega },{\mathbb {R}}^3)\) be the subspace of \(H^1_h({\varOmega },{\mathbb {R}}^3)\) with functions \(u\in H_h^1({\varOmega },{\mathbb {R}}^3)\) vanishing on all boundary nodes \( \overline{\partial {\varOmega }_D}\cap \{X_1,\ldots ,X_n\}\). If vanishing of u on the nodes in \(\partial {\varOmega }_D\) implies vanishing of u on \(\partial {\varOmega }_D\), we have \(H_{D,h}^1({\varOmega },{\mathbb {R}}^3)=H_h^1({\varOmega },{\mathbb {R}}^3)\cap H^1_D({\varOmega },{\mathbb {R}}^3)\). A solution \(u\in H_{D,h}^1({\varOmega },{\mathbb {R}}^3)\) to the discretized elasticity PDE then fulfills

$$\begin{aligned}&B(u,v)=\int \limits _{{\varOmega }}f\cdot v\, dx+\int \limits _{\partial {\varOmega }_N}g\cdot v \,dA , \nonumber \\&\quad \forall v\in H_{D,h}^1({\varOmega },{\mathbb {R}}^3). \end{aligned}$$
(28)

Furthermore, such a solution always exists by the coercivity of the bilinear form B(uv), which also holds on \(H^1_{D,h}({\varOmega },{\mathbb {R}}^3)\).

In the derivation of the adjoint equation in Sect. 4 we need explicit expressions for both sides of (28) in terms of the global degrees of freedom \(U=(u_{j})_{j\in \{1,\ldots ,N\}}\), \(u_j\in {\mathbb {R}}^3\), and the node coordinates X, where it is understood that \(u_{j}=0\) if \(X_j\in \partial {\varOmega }_D\). We rewrite (28) in terms of global variables U via

$$\begin{aligned} B(X)U&=F(X),~~B(X)_{(j,r),(k,s)}=B(e_r\theta _j,e_s\theta _k) \text{ and } \nonumber \\ F_{(j,r)}&=\int \limits _{{\varOmega }}f\cdot e_r\theta _j\, dx+\int \limits _{\partial {\varOmega }_N}g\cdot e_r\theta _j \,dA, \end{aligned}$$
(29)

with \(e_r\), \(r=1,2,3\) the standard Basis on \({\mathbb {R}}^ 3\). The linear equation (29) is understood in the sense that the (ks) indices of the stiffness matrix B(X) are contracted with the related indices of U. The solution of this equation is denoted by U(X).

We have emphasised the dependency of the stiffness matrix B(X) and the load vector F(X) from the global node coordinate \(N\times 3\)-matrix X. The explicit calculations, which are completely standard, are given in “Appendix A.2” in order to prepare the ground for the calculation of shape sensitivities.

In some situations, also the surface and volume force densities \(f=f(X)\) and \(g=g(X)\) are allowed to have a differentiable dependency on X. For f this makes sense e.g. in the case of centrifugal loads and we might want to adapt a surface force density g under geometry changes, in order to keep the total force acting on a part of the surface fixed, even if the surface volume changes. For notational simplicity, we will only introduce this when needed for a physically correct description.

1.2 Discretization of the PDE

The integrals on both sides of (28) are implemented using a numerical quadrature with quadrature points \({\widehat{\xi }}_l\) and weights \({\widehat{\omega }}_l\), \(l=1,\ldots ,l_q\), on the reference element \({\widehat{K}}\). We start with the discretization of the left hand side of (28):

$$\begin{aligned}&B(u,v)\nonumber \\&\quad =\lambda \sum _{K\in {\mathcal {T}}_h}\int \limits _{K} \nabla \cdot u\nabla \cdot v dx+2\mu \sum _{K\in {\mathcal {T}}_h}\int \limits _{K}\varepsilon (u):\varepsilon (v) dx \nonumber \\&\quad =\lambda \sum _{K\in {\mathcal {T}}_h}\int \limits _{{\widehat{K}}} \nabla \cdot u(T_K({\widehat{\xi }}))\nabla \cdot v(T_K({\widehat{\xi }}))\det ({\widehat{\nabla }}T_K({\widehat{\xi }})) d{\widehat{\xi }} \nonumber \\&\qquad +2\mu \sum _{K\in {\mathcal {T}}_h}\int \limits _{{\widehat{K}}} \varepsilon \bigl (u(T_K({\widehat{\xi }}))\bigr ):\varepsilon \bigl (v(T_K({\widehat{\xi }}))\bigr )\det ({\widehat{\nabla }}T_K({\widehat{\xi }})) d{\widehat{\xi }} \nonumber \\&\quad =\lambda \sum _{K\in {\mathcal {T}}_h}\sum _{l=1}^{l_q} \widehat{\omega _l} \det (\widehat{\nabla }T_K(\widehat{\xi _l}))\nabla \cdot u(T_K(\widehat{\xi _l}))\nabla \cdot v(T_K(\widehat{\xi _l})) \nonumber \\&\qquad +2\mu \sum _{K\in {\mathcal {T}}_h}\sum _{l=1}^{l_q}\widehat{\omega _l} \det ({\widehat{\nabla }}T_K(\widehat{\xi _l}))\varepsilon \bigl (u(T_K(\widehat{\xi _l}))\bigr ):\varepsilon \bigl (v(T_K(\widehat{\xi _l}))\bigr ) \end{aligned}$$
(30)

After setting \(\omega _{lK}=\widehat{\omega _l} \det \left( {\widehat{\nabla }} T_K(\widehat{\xi _l})\right) \) and \(\xi _l^K=T_K(\widehat{\xi _l})\), this can be written as

$$\begin{aligned} B(u,v)&=\underbrace{\lambda \sum _{K\in {\mathcal {T}}_h}\sum _{l=1}^{l_q}\omega _{lK}\nabla \cdot u(\xi _l^K)\nabla \cdot v(\xi _l^K)}_{B_1(u,v)} \nonumber \\&\quad +\underbrace{2\mu \sum _{K\in {\mathcal {T}}_h}\sum _{l=1}^{l_q}\omega _{lK}\varepsilon \bigl (u(\xi _l^K)\bigr ):\varepsilon \bigl (v(\xi _l^K)\bigr )}_{B_2(u,v)} \end{aligned}$$
(31)

If we apply (16) and (17) to u and v and insert the result into (31), we obtain the discretization of the left hand side.

For the volume integral on the right hand side of (28) we obtain by a similar argument

$$\begin{aligned} \int \limits _{{\varOmega }} f\cdot v\, dx=\sum _{K\in {\mathcal {T}}_h}\sum _{l=1}^{l_q} \omega _{lK} f(\xi _l^K)\cdot v(\xi _l^K). \end{aligned}$$
(32)

And finally we get the following expression for the discretized surface integral

$$\begin{aligned}&\int \limits _{\partial {\varOmega }}g\cdot v\,dA=\sum _{F\in {\mathcal {N}}_h}\sum _{l=1}^{l_q^F}\omega _{lF}g(\xi _{lF})\cdot v(\xi _{lF}),\text {where} \nonumber \\&\omega _{lF} ={\widehat{\omega }}^F_l\sqrt{\det g_F({\widehat{\xi }}_l^F)}~\text {and}~\xi _{l}^F=T_{K(F)} ({\widehat{\xi }}_l^F). \end{aligned}$$
(33)

Here \({\widehat{\xi }}_l^F\) and \({\widehat{\omega }}_l^F\) are quadrature points on a reference face \({{\widehat{F}}}\) of the reference element \({\widehat{K}}\) and \({\mathcal {N}}_h\) is the collection of all finite element faces that lie in \(\partial {\varOmega }\). is the Gram matrix on \({\widehat{F}}\), where \(T_K:{\widehat{F}}\rightarrow F\).

If the selected quadratures are not exact, the above identities have to be understood in the sense of approximations.

1.3 Computing \(\frac{\partial J}{\partial U}\)

In the following, we use some conventions related to the connectivity mapping \({\widehat{j}}:{\mathcal {T}}_h\times \{1,\ldots ,n_\mathrm{sh})\rightarrow \{1,\ldots ,N\}\). For \(K\in {\mathcal {T}}_h\), we denote by \({\widehat{j}}_K:\{1,\ldots ,n_{\mathrm{sh}}\}\rightarrow \{1,\ldots ,N\}\) the restriction of the connectivity mapping \({\widehat{j}}\) to the set \(\{(K,1),\ldots ,(K,n_k)\}\), where we identify \(1,\ldots ,n_{\mathrm{sh}}\) with \((K,1),\ldots ,(K,{n_{\mathrm{sh}}})\}\).

For a fixed global index \(j\in \{1,2,...,N\}\) we have \({\widehat{j}}^{-1}(j)=\{(K_1,m_1),...,(K_f,m_f)\}\), where \(K_1,...,K_f \in {\mathcal {T}}_h\) and \(m_1,...,m_f\in \{1,2,...,n_{\mathrm{sh}}\}\). With \({\widehat{j}}^{-1}(j)_1=\{K_1,\ldots ,K_f\}\) we denote the set projection to the first component.

For \(k=1,2,3\) let \(u_{jk}\) be the xyz coordinate of the global degree of freedom \(u_j\). We obtain for the partial derivative of J with respect to this variable

$$\begin{aligned} \begin{aligned} \frac{\partial J(X,U)}{\partial u_{jk}}&= \sum _{\begin{array}{c} F\in {\mathcal {N}}_h \\ j\in {\widehat{j}}_{K(F)}(\{1,\ldots ,n_{\mathrm{sh}}\}) \end{array}}\sum _{l=1}^{l_q^F}\omega _{lF}\\&\quad \times \frac{\partial }{\partial u_{jk}}\left[ \mathrm{CMB}^{-1}\left( \varepsilon _{a}^{\mathrm{el-pl}}\biggl (\sum _{m=1}^{n_\mathrm{sh}}u_{{\widehat{j}}(K(F),m)}\otimes \right. \right. \\&\qquad \qquad \left. \left. (({\widehat{\nabla }} T_{K(F)}^{T}({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_m({\widehat{\xi }}_l^F))\biggr ) \right) \right] ^{-{\bar{m}}}. \end{aligned} \end{aligned}$$

We will frequently use the abbreviation

$$\begin{aligned} q&=q_{lF}=q_{lF}(U,X)=\nabla u(\xi _l^F) \nonumber \\&=\sum _{m=1}^{n_{\mathrm{sh}}}u_{j_{(K(F),m)}}\otimes ({\widehat{\nabla }} T_{K(F)}^{T}({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_m(\widehat{\xi _l})), ~l=1,\ldots ,l_q^ F, \end{aligned}$$
(34)

with \(\xi _l^F=T_{K_F}({\hat{\xi }}_l)\). We thus have

$$\begin{aligned}&\frac{\partial }{\partial u_{jk}}\left[ \mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF}))\right] ^{-{\bar{m}}} \nonumber \\&\quad = -{\bar{m}}\left[ \mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF}))\right] ^{-{\bar{m}}-1}\nonumber \\&\qquad \,\, \times \frac{\partial }{\partial u_{jk}} \left[ \mathrm{CMB}^{-1}(\varepsilon _{a}^{\mathrm{el-pl}}(q_{lF}))\right] . \end{aligned}$$
(35)

In the next step, we compute

$$\begin{aligned} \frac{\partial }{\partial u_{jk}}[\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF}))]&= \frac{\frac{\partial }{\partial u_{jk}}\bigl (\varepsilon _{a}^\mathrm{el-pl}(q_{lF})\bigr )}{\frac{\partial \mathrm{CMB}}{\partial \varepsilon _a^{\mathrm{el-pl}}}(\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF})))} \end{aligned}$$
(36)

We now calculate

$$\begin{aligned} \frac{\partial }{\partial u_{jk}}\bigl (\varepsilon _{a}^\mathrm{el-pl}(q_{lF})\bigr )&= \frac{\partial \varepsilon _{a}^\mathrm{el-pl}(q_{lF})}{\partial q_{lF}}:\frac{\partial q_{lF}}{\partial u_{jk}} \nonumber \\&= \,\text {tr}\,\biggl (\Bigl (\frac{\partial \varepsilon _{a}^\mathrm{el-pl}(q_{lF})}{\partial q_{lF}} \Bigr )^T\frac{\partial q_{lF}}{\partial u_{jk}}\biggr ), \end{aligned}$$
(37)

where for \(s,n=1,2,3\),

$$\begin{aligned}&\frac{\partial (q_{lF})_{sn}}{\partial u_{kj}} \nonumber \\&\quad =\frac{\partial }{\partial u_{jk}}\left( \sum _{m=1}^{n_\mathrm{sh}}u_{{\widehat{j}}(K(F),m)s}\left( ({\widehat{\nabla }}T_{K(F)}^{T} ({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_m({\widehat{\xi }}_l^F)\right) _n\right) \nonumber \\&\quad =\sum _{m=1}^{n_{\mathrm{sh}}}\frac{\partial u_{{\widehat{j}}(K(F),m)s}}{\partial u_{jk}}\left( ({\widehat{\nabla }}T_{K(F)}^{T} ({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_m({\widehat{\xi }}_l^F)\right) _n \nonumber \\&\quad ={\left\{ \begin{array}{ll} \delta _{sk}\left( ({\widehat{\nabla }}T_{K(F)}^{T} ({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_{{\widehat{j}}^{-1}_{K(F)}(j)} ({\widehat{\xi }}_l^F)\right) _n&{} { \text{ if } j\in {\widehat{j}}_{K(F)}}\\ 0&{} \text{ otherwise }\end{array}\right. }, \end{aligned}$$
(38)

with \(\delta _{sk}\) Kronecker’s symbol and \({\widehat{j}}_{K(F)} = {\widehat{j}}_{K(F)}(\{1,\ldots ,n_{\mathrm{sh}}\})\). In order to simplify the calculation, we square (5) and divide by E. The resulting relation between \(\varepsilon _{a}^{\mathrm{el-pl}}\) and \(\frac{\sigma _a^2}{E}\) then is denoted by \(\varepsilon _{a}^\mathrm{el-pl}=\bar{\mathrm{SD}}\left( \frac{\sigma _a^2}{E}\right) \). The partial derivative of \(\varepsilon _{a}^{\mathrm{el-pl}}\) with respect to \(q_{sm}\) is now:

$$\begin{aligned} \frac{\partial \varepsilon _{a}^{\mathrm{el-pl}}}{\partial q_{sm}}=\frac{d \mathrm{RO}\circ \bar{\mathrm{SD}}^{-1}\left( \frac{(\sigma _a)^2}{E}\right) }{d \left( \frac{(\sigma _a)^2}{E}\right) }\cdot \frac{\partial }{\partial q_{sm}}\biggl (\frac{(\sigma _a)^2}{E}\biggr ). \end{aligned}$$
(39)

The derivative of the function \(\mathrm{RO}\circ \bar{\mathrm{SD}}^{-1}(\cdot )\) is easily calculated from (3) and (5). We further calculate

$$\begin{aligned} \frac{\partial }{\partial q_{sm}} \biggl (\frac{(\sigma _a)^2}{E}\biggr ) = \frac{3\mu ^2}{4E}\sum _{i,j=1}^{3}A_{ij}\frac{\partial A_{ij}}{\partial q_{sm}} \end{aligned}$$
(40)

using \(A_{ij}=\biggl (-\frac{2}{3}\bigl (\frac{1}{2}(q_{ij}+q_{ji})\bigr )\delta _{ij}+q_{ij}+q_{ji}\biggr )\). We thus obtain

$$\begin{aligned} \frac{\partial A_{ij}}{\partial q_{s,m}}=-\frac{2}{3} \delta _{ij}\delta _{sm}+ \delta _{is}\delta _{jm}+\delta _{js}\delta _{im}. \end{aligned}$$
(41)

We thus get for the left hand side of (40)

$$\begin{aligned}&\frac{3\mu ^2}{4E}\sum _{i,j=1}^{3}\biggl (-\frac{2}{3} \bigl (\frac{1}{2}(q_{ij}+q_{ji})\bigr )\delta _{ij}+q_{ij}+q_{ji}\biggr ) \nonumber \\&\quad \bigl (-\frac{2}{3}\delta _{ij}\delta _{sm}+\delta _{is}\delta _{jm}+ \delta _{js}\delta _{im}\bigr ). \end{aligned}$$
(42)

1.4 Computing \(\frac{\partial J}{\partial X}\)

As in Sect. A.3, We use the abbreviation \(q_{lF}\) for \(\nabla u(\xi _l^F)\), see (34). The partial derivative of J(XU) w.r.t. the global jth mesh node i-coordinate \(X_{ji}\), \(i=1,2,3\) and \(j=1,\ldots ,N\), is

$$\begin{aligned}&\frac{\partial J(X,U)}{\partial X_{ji}} \nonumber \\&\quad =\sum _{\begin{array}{c} F\in {\mathcal {N}}_h \\ j\in {\widehat{j}}_{K(F)}(\{1,\ldots ,n_\mathrm{sh}\}) \end{array}}\sum _{l=1}^{l_q^F}\frac{\partial }{\partial X_{ji}}\Biggl (\omega _{lF}\left( \mathrm{CMB}^{-1}\left( \varepsilon _{a}^\mathrm{el-pl}(q_{lF})\right) \right) ^{-{\bar{m}}}\Biggr ) \nonumber \\&\quad =\sum _{\begin{array}{c} F\in {\mathcal {N}}_h \\ j\in {\widehat{j}}_{K(F)}(\{1,\ldots ,n_\mathrm{sh}\}) \end{array}}\sum _{l=1}^{l^F_q}\Biggl [\frac{\partial \omega _{lF}}{\partial X_{ji}}\Bigl (\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF})))\Bigr )^{-{\bar{m}}} \nonumber \\&\qquad +\, \omega _{lF}\frac{\partial }{\partial X_{ji}}\Bigl (\Bigl (\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF}))\Bigr )^{-{\bar{m}}}\Bigr )\Biggr ]. \end{aligned}$$
(43)

We compute first the partial derivative of \(\frac{\partial \omega _{lF}}{\partial X_{ji}}\) as

$$\begin{aligned} \frac{\partial \omega _{lF}}{\partial X_{ji}}&= \frac{\partial }{\partial X_{ji}}\left( \widehat{\omega _l} \sqrt{\det (g_F({\widehat{\xi }}_l^F))}\right) \nonumber \\&=\widehat{\omega _l}\frac{1}{2}\bigl (\det (g_F({\widehat{\xi }}_l^F)) \bigr )^{-1/2}\frac{\partial }{\partial X_{ji}}(\det (g_F({\widehat{\xi }}^F_l))). \end{aligned}$$
(44)

The derivative of the determinant is

$$\begin{aligned}&\frac{\partial }{\partial X_{ji}}\biggl (\det \Bigl (g_F({\widehat{\xi }}_l^F)\Bigr )\biggr ) \nonumber \\&\quad =\det \Bigl (g_F({\widehat{\xi }}^F_l)\Bigr ) \,\text {tr}\,\biggl (\Bigl (g_F({\widehat{\xi }}^F_l)\Bigr )^{-1}\frac{\partial g_F({\widehat{\xi }}^F_l)}{\partial X_{ji}}\biggr ), \end{aligned}$$
(45)

where \(g_F({\widehat{\xi }}_l^F)=J_F({\widehat{\xi }}_l^F)^T J_F({\widehat{\xi }}_l^F)\) and

$$\begin{aligned} J_F({\widehat{\xi }}_l^F)&=\frac{\partial T_K({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}^F} \nonumber \\&=\sum _{r=1}^{n_{\mathrm{sh}}} \left( \begin{array}{c@{\quad }c} X_{r1}^{K(F)}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}_1^F} &{} X_{1r}^{K(F)}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}_2^F} \\ X_{r2}^{K(F)}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}_1^F} &{} X_{r2}^{K(F)}\frac{\partial \theta _r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}^F_2} \\ X_{r3}^{K(F)}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}_1^F} &{} X_{r3}^{K(F)}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial {\widehat{X}}^F_2} \\ \end{array} \right) . \end{aligned}$$
(46)

Here \({\widehat{X}}^F_{i}\), \(i=1,2\), are the coordinates on the two dimensional reference face \({\widehat{F}}\) corresponding to F in \({\widehat{K}}\).

The derivative \(\frac{\partial g_F({\widehat{\xi }}_l^F)}{\partial X_{ji}}\) is thus given by

$$\begin{aligned} \frac{\partial g_F({\widehat{\xi }}_l^F)}{\partial X_{ji}}=&\frac{\partial }{\partial X_{ji}}\bigl (J_F({\widehat{\xi }}_l^F)\bigr )^T J_F({\widehat{\xi }}_l^F) \nonumber \\&+\bigl (J_F({\widehat{\xi }}_l^F)\bigr )^T\frac{\partial }{\partial X_{ji}}\bigl (J_F({\widehat{\xi }}_l^F)\bigr ). \end{aligned}$$
(47)

Furthermore, for \(s=1,2,3\) and \(k=1,2\) we have

$$\begin{aligned}&\frac{\partial }{\partial X_{ji}}\bigl (J_F({\widehat{\xi }}_l^F)_{sk}\bigr )= \sum _{r=1}^{n_\mathrm{sh}}\frac{\partial }{\partial X_{ji}}\left( X_{rs}^{K(F)}\frac{\partial \theta _r({\widehat{\xi }}_l^F)}{\partial \widehat{X_k}} \right) \nonumber \\&\quad ={\left\{ \begin{array}{ll} \delta _{is}\frac{\partial {\widehat{\theta }}_{{\widehat{j}}^{-1}_{K(F)}(j)}({\widehat{\xi }}_l^F)}{\partial \widehat{X_k}}&{}\quad \text{ if } j\in \widehat{j}_{K(F)}(\{1,\ldots ,n_{\mathrm{sh}}\})\\ 0&{}\quad \text{ otherwise } \end{array}\right. }. \end{aligned}$$
(48)

This finishes the computation of the first term on the right hand side of (43).

To compute the second term in (43), we take the partial derivative

$$\begin{aligned} \begin{aligned}&\frac{\partial }{\partial X_{ji}}\Bigl [\mathrm{CMB}^{-1}(\varepsilon _{a}^{\mathrm{el-pl}}(q_{lF}))\Bigr ]^{-{\bar{m}}} \\&\quad =\frac{-{\bar{m}}[\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF}))]^{-{\bar{m}}-1}}{\frac{\partial \mathrm{CMB}}{\partial \varepsilon _{a}^{\mathrm{el-pl}}}(\mathrm{CMB}^{-1}(\varepsilon _{a}^\mathrm{el-pl}(q_{lF})))} \frac{\partial }{\partial X_{ji}}\bigl (\varepsilon _{a}^{\mathrm{el-pl}}(q_{lF})\bigr ), \end{aligned} \end{aligned}$$

with

$$\begin{aligned} \frac{\partial }{\partial X_{ji}}\bigl (\varepsilon _{a}^\mathrm{el-pl}(q_{lF})\bigr )&= \frac{\partial \varepsilon _{a}^\mathrm{el-pl}(q_{lF})}{\partial q_{lF}}:\frac{\partial q_{lF}}{\partial X_{ji}}. \end{aligned}$$
(49)

In the above equation,  :  stands for the contraction of both q indices. Next we have to compute

$$\begin{aligned}&\frac{\partial q_{lF}}{\partial X_{ji}}(X) \nonumber \\&\quad =\frac{\partial }{\partial X_{ji}}\biggl (\sum _{m=1}^{n_\mathrm{sh}}u_{{\widehat{j}}(K(F),m)}\otimes (({\widehat{\nabla }}T_{K(F)}^T ({\widehat{\xi }}_l^F))^{-1}{\widehat{\nabla }} {\widehat{\theta }}_m({\widehat{\xi }}_l^F))\biggr ) \nonumber \\&\quad =\sum _{m=1}^{n_\mathrm{sh}}u_{j(K(F),m)}\otimes \Bigl (\frac{\partial }{\partial X_{ji}}({\widehat{\nabla }}T_{K(F)}^T({\widehat{\xi }}_l^F))^{-1} \Bigr ){\widehat{\nabla }} {\widehat{\theta }}_m({\widehat{\xi }}_l^F), \end{aligned}$$
(50)

where

$$\begin{aligned}&\frac{\partial }{\partial X_{ji}}\bigl ({\widehat{\nabla }} T_{K}^{T}({\widehat{\xi }}_l^ F)\bigr )^{-1} \nonumber \\&\quad = -\bigl ({\widehat{\nabla }} T_{K}^T({\widehat{\xi }}_l^ F)\bigr )^{-1}\frac{\partial \bigl ({\widehat{\nabla }} T_{K}^T({\widehat{\xi }}_l^ F)\bigr )}{\partial X_{ji}}\bigl ({\widehat{\nabla }} T_{K}^T({\widehat{\xi }}_l^ F)\bigr )^{-1}. \end{aligned}$$
(51)

The Jacobian matrix has the form

$$\begin{aligned} {\widehat{\nabla }} T_{K}({\widehat{\xi }}_l^ F)&= \left( \begin{array}{c@{\quad }c@{\quad }c} \frac{\partial X_{1}^{K}}{\partial \widehat{X_1}} &{} \frac{\partial X_{1}^{K}}{\partial \widehat{X_2}} &{} \frac{\partial X_{1}^{K}}{\partial \widehat{X_3}} \\ \frac{\partial X_{2}^{K}}{\partial \widehat{X_1}} &{} \frac{\partial X_{2}^{K}}{\partial \widehat{X_2}} &{} \frac{\partial X_{2}^{K}}{\partial \widehat{X_3}}\\ \frac{\partial X_{3}^{K}}{\partial \widehat{X_1}} &{} \frac{\partial X_{3}^{K}}{\partial \widehat{X_2}} &{} \frac{\partial X_{3}^{K}}{\partial \widehat{X_3}}\\ \end{array} \right) \nonumber \\&=\sum _{r=1}^{n_{\mathrm{sh}}} \left( \begin{array}{c@{\quad }c@{\quad }c} X_{r1}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_1}} &{} X_{r1}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_2}} &{} X_{r1}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_3}} \\ X_{r2}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_1}} &{} X_{r2}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_2}} &{} X_{r2}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_3}}\\ X_{r3}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_1}} &{} X_{r3}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_2}} &{} X_{r3}^{K}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^ F)}{\partial \widehat{X_3}}\\ \end{array} \right) . \end{aligned}$$
(52)

Finally, we get

$$\begin{aligned}&\frac{\partial \bigl ({\widehat{\nabla }} T_{K}^T({\widehat{\xi }}_l^ F)_{sk}\bigr )}{\partial X_{ji}}=\sum _{r=1}^{n_\mathrm{sh}}\frac{\partial X_{rk}^{K}}{\partial X_{ji}}\frac{\partial {\widehat{\theta }}_r({\widehat{\xi }}_l^F)}{\partial \widehat{X_s}} \nonumber \\&\quad ={\left\{ \begin{array}{ll}\delta _{ik} \frac{\partial {\widehat{\theta }}_{{\widehat{j}}^{-1}_{K}(j)}({\widehat{\xi }}_l^ F)}{\partial \widehat{X_s}},&{} \text{ if } j\in {\widehat{j}}_{K}(\{1,\ldots ,n_{\mathrm{sh}}\})\\ 0&{} \text{ otherwise }\end{array}\right. }. \end{aligned}$$
(53)

This finishes the computation of the second term.

1.5 Computing \(\frac{\partial B}{\partial X}\)

As in (30) we have

$$\begin{aligned} B(X)_{(q,r),(k,s)}&=B(e_r\theta _q,e_s\theta _k) \nonumber \\&=B_1(e_r\theta _q,e_s\theta _k)+B_2(e_r\theta _q,e_s\theta _k), \end{aligned}$$

then \(\frac{\partial B(X)_{(q,r),(k,s)}}{\partial X}=\frac{\partial B_1(e_r\theta _q,e_s\theta _k)}{\partial X_{ji}}+\frac{\partial B_2(e_r\theta _q,e_s\theta _k)}{\partial X_{ji}}\). For the first partial derivative, one obtains

$$\begin{aligned} \frac{\partial B_1(e_r\theta _q,e_s\theta _k)}{\partial X_{ji}}&=\lambda \sum _{K\in {\widehat{j}}^{-1}(j)_1\cap {\widehat{j}}^{-1}(q)_1\cap {\widehat{j}}^{-1}(k)_1}\sum _{l=1}^{l_q} \nonumber \\&\Bigg \{\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr )\nabla _r \theta _q(\xi _l^{K})\nabla _s \theta _k(\xi _l^{K}) \nonumber \\&\quad +\omega _{lK}\frac{\partial }{\partial X_{ji}}\bigl (\nabla _r \theta _q(\xi _l^{K})\bigr )\nabla _s \theta _k(\xi _l^{K}) \nonumber \\&\quad +\omega _{lK}\nabla _r \theta _q(\xi _l^{K})\frac{\partial }{\partial X_{ji}}\bigl (\nabla _s \theta _k(\xi _l^{K})\bigr )\Bigg \}. \end{aligned}$$
(54)

Here we use the convention that \({\widehat{j}}^{-1}(j)_1\) is the projection of the set \({\widehat{j}}^{-1}(j)\) to the first component (the index of the element).

We thus have to compute the three partial derivatives \(\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr )\), \(\frac{\partial }{\partial X_{ji}}\bigl (\nabla _r\theta _q(\xi _l^K)\bigr )\) and \(\frac{\partial }{\partial X_{ji}}\bigl (\nabla _s\theta _k(\xi _l^K)\bigr )\). We start with the first partial derivative and we obtain

$$\begin{aligned} \frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr )&=\frac{\partial }{\partial X_{ji}}\bigl (\widehat{\omega _{l}}\det ({\widehat{\nabla }} T_K(\widehat{\xi _l})\bigr ) \nonumber \\&=\widehat{\omega _l}\frac{\partial }{\partial X_{ji}}\bigl (\det ( {\widehat{\nabla }}T_K(\widehat{\xi _l})\bigr ). \end{aligned}$$
(55)

For an invertible matrix A, we have

$$\begin{aligned} \frac{d}{d\alpha }\det (A)=\det (A)\,\text {tr}\,(A^{-1}\frac{dA}{d\alpha }), \end{aligned}$$

so we get

$$\begin{aligned}&\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr ) \nonumber \\&\quad =\widehat{\omega _l}\det \left( {\widehat{\nabla }} T_K(\widehat{\xi _l})\right) \,\text {tr}\,\left( ({\widehat{\nabla }}T_K({\widehat{\xi }}_l))^{-1}\frac{\partial {\widehat{\nabla }}T_K({\widehat{\xi }}_l)}{\partial X_{ji}}\right) . \end{aligned}$$
(56)

The partial \(X_{ji}\) derivative of the matrix \({\widehat{\nabla }}T_K({\widehat{\xi }}_l)\) has been calculated in (53), where we have to replace \({\widehat{\xi }}_l^F\) with \({\widehat{\xi }}_l\).

The second partial derivative in \(\frac{\partial B_1}{\partial X_{ji}}\) is easily calculated

$$\begin{aligned} \frac{\partial \nabla _r \theta _q}{\partial X_{ji}}(\xi _l^K)&= \frac{\partial }{\partial X_{ji}}\left[ \left( {\widehat{\nabla }}T_K({\hat{\xi }}_l)^T \right) ^{-1}{\widehat{\nabla }} {\widehat{\theta }}_{{\widehat{j}}_{K}^{-1}(q)}({\hat{\xi }}_l)\right] _r. \end{aligned}$$
(57)

We now refer to Eqs. (51)–(53) in in Sect. A.4 to conclude the computation. Here, of course, the surface quadrature point \({\hat{\xi }}^F_l\) has to be replaced by the volume quadrature point \({\hat{\xi }}_l\). The third partial derivative in (54) is completely analogous to the second.

For the partial derivative \(\frac{\partial B_2}{\partial X}\) we obtain

$$\begin{aligned}&\frac{\partial B_2(e_r\theta _q,e_s\theta _k)}{\partial X_{ji}} \nonumber \\&\quad =2\mu \sum _{K\in {\widehat{j}}^{-1}(j)_1\cap {\widehat{j}}^{-1}(q)_1\cap {\widehat{j}}^{-1}(k)_1}\nonumber \\&\qquad \sum _{l=1}^{l_q} \Bigg \{\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr )\varepsilon \left( e_r \theta _q(\xi _l^{K}))\right) :\varepsilon \left( e_s \theta _k(\xi _l^{K})\right) \nonumber \\&\quad \qquad +\omega _{lK}\frac{\partial }{\partial X_{ji}}\bigl (\varepsilon \left( e_r \theta _q(\xi _l^{K}))\right) \bigr ):\varepsilon \left( e_s \theta _k(\xi _l^{K})\right) \nonumber \\&\quad \qquad +\omega _{lK}\varepsilon \left( e_r \theta _q(\xi _l^{K}))\right) :\frac{\partial }{\partial X_{ji}}\bigl (\varepsilon \left( e_s \theta _k(\xi _l^{K})\right) \bigr )\Bigg \}. \end{aligned}$$
(58)

The first term is calculated in (55). For the second term, we observe that the linear elastic strain tensor field is given by

$$\begin{aligned} \varepsilon (e_r\theta _q(\xi _l^ K))_{ab}= \frac{1}{2}\left( \delta _{rb}\nabla _a \theta _q(\xi _l^ K))+\delta _{ra}\nabla _b \theta _q(\xi _l^ K)\right) ,\nonumber \\ \end{aligned}$$
(59)

and we refer to the argument following Eq. (57) to conclude the computation of (58).

1.6 Computing \(\frac{\partial F}{\partial X}\)

We have \(\frac{\partial F_{(q,r)}}{\partial X_{ji}}=\frac{\partial F_{(q,r)}^{\mathrm{vol}}}{\partial X_{ji}}+\frac{\partial F_{(q,r)}^\mathrm{surf}}{\partial X_{ji}}\). Starting with the volume term, we obtain

$$\begin{aligned} \frac{\partial F_{(q,r)}^{\mathrm{vol}}}{\partial X_{ji}}&=\sum _{K\in {\widehat{j}}^{-1}(j)_1\cap {\widehat{j}}^{-1}(q)_1} \nonumber \\&\quad \sum _{l=1}^{l_q}\Bigg \{\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lK}\bigr )f_r(\xi _l^K)\theta _q(\xi _l^K) \nonumber \\&\quad \qquad +\,\omega _{lK}\frac{\partial }{\partial X_{ji}}\bigl (f_r(\xi _l^K)\bigr )\theta _q(\xi _l^K) \nonumber \\&\quad \qquad +\,\omega _{lK}f_r(\xi _l^K)\frac{\partial }{\partial X_{ji}}\bigl (\theta _q(\xi _l^K)\bigr )\bigg \} \end{aligned}$$
(60)

The partial derivative of the volume weight has been calculated in (55). The third term in (60) vanishes, as \(\theta _q(\theta _l)={\hat{\theta }}_{{\widehat{j}}^{-1}_K(q)}({\hat{\xi }}_l)\) does not depend on \(X_{ji}\). Unless the volume force density f does depend explicitly on the position (like in the case of centrifugal loads) this term vanishes as the third one.

Finally we have to calculate the partial derivative of the surface loads

$$\begin{aligned} \frac{\partial F_{(q,r)}^{\mathrm{surf}}}{\partial X_{ji}}&= \sum _{\begin{array}{c} F\in {\mathcal {N}}_h \\ K(F)\in {\widehat{j}}^{-1}(j)_1 \cap {\widehat{j}}^{-1} (q)_1 \end{array}} \nonumber \\&\quad \sum _{l=1}^{l_q^F}\Bigg \{\frac{\partial }{\partial X_{ji}}\bigl (\omega _{lF}\bigr )f_r(\xi _l^F)\theta _q(\xi _l^F) \nonumber \\&\quad \qquad +\,\omega _{lF}\frac{\partial }{\partial X_{ji}}\bigl (f_r(\xi _l^F)\bigr )\theta _q(\xi _l^F) \nonumber \\&\quad \qquad +\,\omega _{lF}f_r(\xi _l^F)\frac{\partial }{\partial X_{ji}}\bigl (\theta _q(\xi _l^F)\bigr )\Bigg \}. \end{aligned}$$
(61)

The first term can be calculated with the aid of (44). For the second and third term, the same reasoning applies as for partial derivative of the volume force.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gottschalk, H., Saadi, M. Shape gradients for the failure probability of a mechanic component under cyclic loading: a discrete adjoint approach. Comput Mech 64, 895–915 (2019). https://doi.org/10.1007/s00466-019-01686-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-019-01686-3

Keywords

Mathematics Subject Classification

Navigation