Skip to main content
Log in

Convergence analysis of fixed stress split iterative scheme for anisotropic poroelasticity with tensor Biot parameter

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

Abstract

We perform a convergence analysis of the fixed stress split iterative scheme for the Biot system modeling coupled flow and deformation in anisotropic poroelastic media with tensor Biot parameter. The fixed stress split iterative scheme solves the flow subproblem with all components of the stress tensor frozen using a multipoint flux mixed finite element method, followed by the poromechanics subproblem using a conforming Galerkin method in every coupling iteration at each time step. The coupling iterations are repeated until convergence and Backward Euler is employed for time marching. The convergence analysis is based on studying the equations satisfied by the difference of iterates to show that the fixed stress split iterative scheme for anisotropic poroelasticity with Biot tensor is contractive. We also demonstrate that the scheme is numerically convergent using the classical Mandel’s problem solution for transverse isotropy.

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.

Institutional subscriptions

Similar content being viewed by others

References

  1. Abousleiman, Y., Cheng, A.H.D., Cui, L., Detournay, E., Roegiers, J.C.: Mandel’s problem revisited. Géotechnique 46(2), 187–195 (1996)

    Article  Google Scholar 

  2. Almani, T., Kumar, K., Dogru, A., Singh, G., Wheeler, M.F.: Convergence analysis of multirate fixed-stress split iterative schemes for coupling flow with geomechanics. Comput. Methods Appl. Mech. Eng. 311, 180–207 (2016)

    Article  Google Scholar 

  3. Aoki, T., Tan, C.P., Bamford, W.E.: Effects of deformation and strength anisotropy on borehole failures in saturated shales. Int. J. Rock Mech. Min. Sci. Geomech. Abstracts 30(7), 1031–1034 (1993)

    Article  Google Scholar 

  4. Armero, F., Simo, J.C.: A new unconditionally stable fractional step method for non-linear coupled thermomechanical problems. Int. J. Numer. Methods Eng. 35(4), 737–766 (1992)

    Article  Google Scholar 

  5. Biot, M.A.: General theory of three dimensional consolidation. J. Appl. Phys. 12, 155–164 (1941)

    Article  Google Scholar 

  6. Biot, M.A.: Theory of elasticity and consolidation for a porous anisotropic solid. J. Appl. Phys. 26(2), 182–185 (1955)

    Article  Google Scholar 

  7. Biot, M.A.: Mechanics of deformation and acoustic propagation in porous media. J. Appl. Phys. 33(4), 1482–1498 (1962)

    Article  Google Scholar 

  8. Biot, M.A., Willis, D.G.: The elastic coefficients of the theory of consolidation. J. Appl. Mech. 24, 594–601 (1957)

    Google Scholar 

  9. Boresi, A.P., Chong, K.P., Lee, J.D.: Elasticity in Engineering Mechanics, 3rd edn. Wiley (2010)

  10. Brezzi, F., Douglas, J., Durán, R., Fortin, M.: Mixed finite elements for second order elliptic problems in three variables. Numer. Math. 51(2), 237–250 (1987)

    Article  Google Scholar 

  11. Carroll, M.M.: An effective stress law for anisotropic elastic deformation. J. Geophys. Res. 84, B13 (1979)

    Article  Google Scholar 

  12. Carroll, M.M., Katsube, N.: The role of terzaghi effective stress in linearly elastic deformation. J. Energy Resour. Technol. 105(4), 509–511 (1983)

    Article  Google Scholar 

  13. Castelletto, N., White, J.A., Tchelepi, H.A.: Accuracy and convergence properties of the fixed-stress iterative solution of two-way coupled poromechanics. Int. J. Numer. Anal. Methods Geomech. 39(14), 1593–1618 (2015)

    Article  Google Scholar 

  14. Cheng, A.H.D.: Material coefficients of anisotropic poroelasticity. Int. J. Rock Mech. Min. Sci. 34, 199–205 (1997)

    Article  Google Scholar 

  15. Cui, L., Cheng, A.H.D., Kaliakin, V.N., Abousleiman, Y., Roegiers, J.C.: Finite element analyses of anisotropic poroelasticity: A generalized mandel’s problem and an inclined borehole problem. Int. J. Numer. Anal. Methods Geomech. 20(6), 381–401 (1996)

    Article  Google Scholar 

  16. Felippa, C.A., Park, K.C., Farhat, C.: Partitioned analysis of coupled mechanical systems. Comput. Methods Appl. Mech. Eng. 190(24), 3247–3270 (2001)

    Article  Google Scholar 

  17. Geertsma, J.: The effect of fluid pressure decline on volumetric changes of porous rocks. SPE 210, 331–340 (1957)

    Google Scholar 

  18. Gurtin, M.E., Fried, E., Anand, L: The Mechanics and Thermodynamics of Continua, 1st edn. Cambridge University Press (2010)

  19. Ingram, R., Wheeler, M.F., Yotov, I.: A multipoint flux mixed finite element method on hexahedra. SIAM J. Numer. Anal. 48(4), 1281–1312 (2010)

    Article  Google Scholar 

  20. Katsube, N.: The constitutive theory for fluid-filled porous materials. J. Appl. Mech. 52(1), 185–189 (1985)

    Article  Google Scholar 

  21. Kim, J., Tchelepi, H.A., Juanes, R.: Stability, accuracy and efficiency of sequential methods for coupled flow and geomechanics. SPE J. 16(2), 249–262 (2011)

    Article  Google Scholar 

  22. Mikelić, A., Wheeler, M.F.: Convergence of iterative coupling for coupled flow and geomechanics. Comput. Geosci. 17(3), 455–461 (2013)

    Article  Google Scholar 

  23. Mikelić, A., Wang, B., Wheeler, M.F.: Numerical convergence study of iterative coupling for coupled flow and geomechanics. Comput. Geosci. 18(3), 325–341 (2014)

    Article  Google Scholar 

  24. Nur, A., Byerlee, J.D.: An exact effective stress law for elastic deformation of rock with fluids. J. Geophys. Res. 76(26), 6414–6419 (1971)

    Article  Google Scholar 

  25. Phillips, P.J., Wheeler, M.F.: A coupling of mixed and continuous galerkin finite element methods for poroelasticity ii: the discrete-in-time case. Comput. Geosci. 11(2), 145–158 (2007)

    Article  Google Scholar 

  26. Schrefler, B.A., Simoni, L., Turska, E.: Standard staggered and staggered newton schemes in thermo-hydro-mechanical problems. Comput. Methods Appl. Mech. Eng. 144(1-2), 93–109 (1997)

    Article  Google Scholar 

  27. Skempton, A.W.: The pore-pressure coefficients a and b. Géotechnique 4(4), 143–147 (1954)

    Article  Google Scholar 

  28. Skempton, A.W., Peek, R.B.: Significance of terzaghi’s concept of effective stress (terzaghi’s discovery of effective stress). In: Bjerrum, L., Casagrande, A., Skempton, A.W. (eds.) From Theory to Practice in Soil Mechanics, p. 1960. Wiley, New York (1960)

  29. Steele, J.M.: The Cauchy-Schwarz Master Class: An Introduction to the Art of Mathematical Inequalities. Maa Problem Books Series, Cambridge University Press (2004)

  30. Thompson, M., Willis, J.R.: A reformation of the equations of anisotropic poroelasticity. J. Appl. Mech. 58(3), 612–616 (1991)

    Article  Google Scholar 

  31. Ting, T.C.T.: Anisotropic Elasticity: Theory and Applications. The Oxford Engineering Science Series, Oxford University Press (1996)

  32. Turska, E., Schrefler, B.A.: On convergence conditions of partitioned solution procedures for consolidation problems. Comput. Methods Appl. Mech. Eng. 106(1–2), 51–63 (1993)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Saumik Dana.

Appendices

Appendix A: finite element mapping

Let ri, i = 1,.., 8 be the vertices of E. Now, consider a reference cube \(\hat {E}\) with vertices \(\hat {\mathbf {r}}_{1}=[0\,\,0\,\,0]^{T}\), \(\hat {\mathbf {r}}_{2}=[1\,\,0\,\,0]^{T}\), \(\hat {\mathbf {r}}_{3}=[1\,\,1\,\,0]^{T}\), \(\hat {\mathbf {r}}_{4}=[0\,\,1\,\,0]^{T}\), \(\hat {\mathbf {r}}_{5}=[0\,\,0\,\,1]^{T}\), \(\hat {\mathbf {r}}_{6}=[1\,\,0\,\,1]^{T}\), \(\hat {\mathbf {r}}_{7}=[1\,\,1\,\,1]^{T}\) and \(\hat {\mathbf {r}}_{8}=[0\,\,1\,\,1]^{T}\) as shown in Fig. 3. Let \(\hat {\mathbf {x}}=(\hat {x},\hat {y},\hat {z})\in \hat {E}\) and x = (x,y,z) ∈ E. The function \(F_{E}(\hat {\mathbf {x}}):\hat {E}\rightarrow E\) is

$$\begin{array}{@{}rcl@{}} F_{E}(\hat{\mathbf{x}})&=&\mathbf{r}_{1}(1\,-\,\hat{x})(1\,-\,\hat{y})(1\,-\,\hat{z})+\mathbf{r}_{2}\hat{x}(1\,-\,\hat{y})(1\,-\,\hat{z})+\mathbf{r}_{3}\hat{x}\hat{y}(1\,-\,\hat{z})\\ &&+\mathbf{r}_{4}(1\,-\,\hat{x})\hat{y}(1\,-\,\hat{z})+\mathbf{r}_{5}(1\,-\,\hat{x})(1\,-\,\hat{y})\hat{z}+\mathbf{r}_{6}\hat{x}(1\,-\,\hat{y})\hat{z}\\ && +\mathbf{r}_{7}\hat{x}\hat{y}\hat{z}+ \mathbf{r}_{8}(1\,-\,\hat{x})\hat{y}\hat{z} \end{array} $$
Fig. 3
figure 3

Trilinear mapping \(F_{E}:\hat {E}\rightarrow E\) for 8 noded distorted hexahedral elements. The faces of E can be non-planar

Denote Jacobian matrix by DFE and let JE = det(DFE). Defining rijrirj, we have

$$\scriptstyle DF_{E}(\hat{\mathbf{x}}) = \left[\begin{array}{ccccccc} \!\mathbf{r}_{21}\,+\,(\mathbf{r}_{34}\,-\,\mathbf{r}_{21})\hat{y} +(\mathbf{r}_{65}\,-\,\mathbf{r}_{21})\hat{z} \,+\,((\mathbf{r}_{21}\,-\,\mathbf{r}_{34}) -(\mathbf{r}_{65}\,-\,\mathbf{r}_{78}))\hat{y}\hat{z};\\ \!\mathbf{r}_{41}\,+\,(\mathbf{r}_{34}\,-\,\mathbf{r}_{21})\hat{x} +(\mathbf{r}_{85}\,-\,\mathbf{r}_{41})\hat{z} \,+\,((\mathbf{r}_{21}\,-\,\mathbf{r}_{34}) -(\mathbf{r}_{65}\,-\,\mathbf{r}_{78}))\hat{x}\hat{z};\\ \!\mathbf{r}_{51}\,+\,(\mathbf{r}_{65}\,-\,\mathbf{r}_{21})\hat{x} +(\mathbf{r}_{85}\,-\,\mathbf{r}_{41})\hat{y} \,+\,((\mathbf{r}_{21}\,-\,\mathbf{r}_{34}) -(\mathbf{r}_{65}\,-\,\mathbf{r}_{78}))\hat{x}\hat{y} \end{array}\!\right] $$

Denote inverse mapping by \(F_{E}^{-1}\), its Jacobian matrix by \(DF_{E}^{-1}\) and let \(J_{F_{E}^{-1}}=det(DF_{E}^{-1})\) such that

$$DF_{E}^{-1}(\mathbf{x})=(DF_{E})^{-1}(\hat{\mathbf{x}});\qquad J_{F_{E}^{-1}}(\mathbf{x})=(J_{E})^{-1}(\hat{\mathbf{x}}) $$

Let ϕ(x) be any function defined on E and \(\hat {\phi }(\hat {\mathbf {x}})\) be its corresponding definition on \(\hat {E}\). Then we have

$$ \nabla \phi = (DF_{E}^{-1})^{T}(\mathbf{x})\,\hat{\nabla} \hat{\phi} = (DF_{E})^{-T}(\hat{\mathbf{x}})\,\hat{\nabla} \hat{\phi} $$
(39)

Appendix B: enhanced BDDF1 spaces

For the sake of clarity, we provide a brief description of the mixed finite element spaces used in the flow model. Let \(\mathbf {V}^{*}_{h}\times W_{h}\) be the lowest order BDDF1 MFE spaces on hexahedra (see Brezzi et al. [10]). On the reference unit cube, these spaces are defined as

$$\begin{array}{@{}rcl@{}} \hat{\mathbf{V}}^{*}(\hat{E})\!&=&\!P_{1}(\hat{E}) \,+\,r_{0}\,curl(0,0,\hat{x}\hat{y}\hat{z})^{T} \,+\,r_{1}\,curl(0,0,\hat{x}\hat{y}^{2})^{T}\\ &&+s_{0}\,curl(\hat{x}\hat{y}\hat{z},0,0)^{T}+s_{1}\,curl(\hat{y}\hat{z}^{2},0,0)^{T}\\&& +t_{0}\,curl(0,\hat{x}\hat{y}\hat{z},0)^{T}+t_{1}\,curl(0,\hat{x}^{2}\hat{z},0)^{T}\\ \hat{W}(\hat{E})\!&=&\!\mathbb{P}_{0}(\hat{E}) \end{array} $$

with the following properties

$$\hat{\nabla} \cdot \hat{\mathbf{V}}^{*}(\hat{E})=\hat{W}(\hat{E}), \quad \text{and} \quad \forall \hat{\mathbf{v}}\in \hat{\mathbf{V}}^{*}(\hat{E}),\,\forall \hat{e}\subset \partial \hat{E},\,\,\hat{\mathbf{v}}\cdot \hat{\mathbf{n}}_{\hat{e}}\in P_{1}(\hat{e}) $$

The multipoint flux approximation procedure requires on each face one velocity degree of freedom to be associated with each vertex. Since the BDDF1 space \(\mathbf {V}^{*}_{h}\) has only three degrees of freedom per face, it is augmented with six more degrees of freedom (one extra degree of freedom per face). Since the constant divergence, the linear independence of the shape functions and the continuity of the normal component across the element faces are to be preserved, six curl terms are added [19]. Let Vh × Wh be the enhanced BDDF1 spaces on hexahedra. On the reference unit cube, these spaces are

$$\begin{array}{@{}rcl@{}} \hat{\mathbf{V}}(\hat{E})\!&=&\!\hat{\mathbf{V}}^{*}(\hat{E}) \,+\,r_{2}\,curl({\kern-.3pt}0{\kern-.3pt},0{\kern-.3pt},\hat{x}^{2}\hat{z})^{T} \,+\,r_{3}\,curl(0,0,\hat{x}^{2}\hat{y}\hat{z})^{T} \\ &&+s_{2}\,curl(\hat{x}\hat{y}^{2},0,0)^{T}+s_{3}\,curl(\hat{x}\hat{y}^{2}\hat{z}^{2},0,0)^{T}\\&& +t_{2}\,curl(0,\hat{y}\hat{z}^{2},0)^{T} +t_{3}\,curl(0,\hat{x}\hat{y}\hat{z}^{2},0)^{T}\\ \hat{W}(\hat{E})\!&=&\!\mathbb{P}_{0}(\hat{E}) \end{array} $$

with the following properties

$$\hat{\nabla} \cdot \hat{\mathbf{V}}(\hat{E})=\hat{W}(\hat{E}), \quad \text{and} \quad \forall \hat{\mathbf{v}}\in \hat{\mathbf{V}}(\hat{E}),\,\,\forall \hat{e}\subset \partial \hat{E},\,\,\hat{\mathbf{v}}\cdot \hat{\mathbf{n}}_{\hat{e}}\in \mathbb{Q}_{1}(\hat{e}) $$

where \(\mathbb {Q}_{1}\) is the space of bilinear functions. Since \(dim\,\mathbb {Q}_{1}(\hat {e})= 4\), the dimension of \(\hat {\mathbf {V}}(\hat {E})\) is 24 as shown in Fig. 4.

Fig. 4
figure 4

Degrees of freedom and basis functions for the enhanced BDDF1 velocity space on hexahedra

Appendix C: discrete variational statements for the flow subproblem in terms of coupling iteration differences

Before arriving at the discrete variational statement of the flow model, we impose the fixed stress constraint on the strong form of the mass conservation (1). In lieu of Eq. 13, we write (1) as

$$\begin{array}{@{}rcl@{}} &&\frac{\partial}{\partial t}(Cp+\frac{C}{3}\mathbf{B}:\boldsymbol{\sigma})+\nabla \cdot \mathbf{z}=q\\ &&C\frac{\partial p}{\partial t} +\nabla \cdot \mathbf{z}=q-\frac{C}{3}\mathbf{B}:\frac{\partial \boldsymbol{\sigma}}{\partial t} \end{array} $$
(40)

Using backward Euler in time, the discrete in time form of Eq. 40 for the mth coupling iteration in the (n + 1)th time step is written as

$$C\frac{1}{{\Delta} t}(p^{m,n + 1}-p^{n}) +\nabla \cdot \mathbf{z}^{m,n + 1}=q^{n + 1}-\frac{1}{{\Delta} t}\frac{C}{3}\mathbf{B}:(\boldsymbol{\sigma}^{m,n + 1}-\boldsymbol{\sigma}^{n}) $$

where Δt is the time step and the source term as well as the terms evaluated at the previous time level n do not depend on the coupling iteration count as they are known quantities. The fixed stress constraint implies that σm,n+ 1 gets replaced by σm− 1,n+ 1 i.e. the computation of pm,n+ 1 and zm,n+ 1 is based on the value of stress updated after the poromechanics solve from the previous coupling iteration m − 1 at the current time level n + 1. The modified equation is written as

$$C(p^{m,n + 1}-p^{n})+{\Delta} t\nabla \cdot \mathbf{z}^{m,n + 1}={\Delta} t q^{n + 1}-\frac{C}{3}\mathbf{B}:(\boldsymbol{\sigma}^{m,n + 1}-\boldsymbol{\sigma}^{n}) $$

As a result, the discrete weak form of Eq. 1 is given by

$$\begin{array}{@{}rcl@{}} &&C(p_{h}^{m,n + 1}-{p_{h}^{n}},\theta_{h})_{{\Omega}}+{\Delta} t(\nabla \cdot \mathbf{z}_{h}^{m,n + 1},\theta_{h})_{{\Omega}}\\ &=&{\Delta} t(q^{n + 1},\theta_{h})_{{\Omega}}-\frac{C}{3}(\mathbf{B}:(\boldsymbol{\sigma}^{m-1,n + 1}-\boldsymbol{\sigma}^{n}),\theta_{h})_{{\Omega}} \end{array} $$

Replacing m by m + 1 and subtracting the two equations, we get

$$C(\delta^{(m)}p_{h},\theta_{h})_{{\Omega}}+{\Delta} t(\nabla \cdot \delta^{(m)}\mathbf{z}_{h},\theta_{h})_{{\Omega}}=-\frac{C}{3}(\mathbf{B}:\delta^{(m-1)}\boldsymbol{\sigma},\theta_{h})_{{\Omega}} $$

The weak form of the Darcy law (2) for the mth coupling iteration in the (n + 1)th time step is given by

$$ (\boldsymbol{\kappa}^{-1}\mathbf{z}^{m,n + 1},\mathbf{v})_{{\Omega}}\,=\,-(\nabla p^{m,n + 1},\mathbf{v})_{{\Omega}}+(\rho_{0} \mathbf{g},\mathbf{v})_{{\Omega}}\,\, \!\forall\,\,\mathbf{v}\!\in\! \mathbf{V}({\Omega}) $$
(41)

where V(Ω) is given by

$$\mathbf{V}({\Omega})\equiv \mathbf{H}(div,{\Omega})\cap \big\{\mathbf{v}:\mathbf{v}\cdot \mathbf{n}= 0\,\,\text{on}\,\,{{\Gamma}_{N}^{f}}\big\} $$

and H(div,Ω) is given by

$$\mathbf{H}(div,{\Omega})\equiv\big\{\mathbf{v}:\mathbf{v}\in (L^{2}({\Omega}))^{3},\nabla \cdot \mathbf{v}\in L^{2}({\Omega}) \big\} $$

We use the divergence theorem to evaluate the first term on RHS of Eq. 41 as follows

$$\begin{array}{@{}rcl@{}} (\nabla p^{m,n + 1},\mathbf{v})_{{\Omega}}&=&(\nabla,p^{m,n + 1}\mathbf{v})_{{\Omega}}-(p^{m,n + 1},\nabla \cdot \mathbf{v})_{{\Omega}}\\ &=&(p^{m,n + 1},\mathbf{v}\cdot \mathbf{n})_{\partial {\Omega}} -(p^{m,n + 1},\nabla \cdot \mathbf{v})_{{\Omega}}\\ &=&(g,\mathbf{v}\cdot \mathbf{n})_{{{\Gamma}_{D}^{f}}} -(p^{m,n + 1},\nabla \cdot \mathbf{v})_{{\Omega}} \end{array} $$
(42)

where we invoke vn = 0 on \({{\Gamma }_{N}^{f}}\). In lieu of Eqs. 41 and 42, we get

$$\begin{array}{@{}rcl@{}} (\boldsymbol{\kappa}^{-1}\mathbf{z}^{m,n + 1},\mathbf{v})_{{\Omega}}=-(g,\mathbf{v}\cdot \mathbf{n})_{{{\Gamma}_{D}^{f}}}+(p^{m,n + 1},\nabla \cdot \mathbf{v})_{{\Omega}}+(\rho_{0} \mathbf{g},\mathbf{v})_{{\Omega}} \end{array} $$

Replacing m by m + 1 and subtracting the two equations, we get

$$ (\boldsymbol{\kappa}^{-1}\delta^{(m)} \mathbf{z}_{h}, \mathbf{v}_{h})_{{\Omega}}=(\delta^{(m)} p_{h},\nabla \cdot \mathbf{v}_{h})_{{\Omega}} $$

Appendix D: discrete variational statement for the poromechanics subproblem in terms of coupling iteration differences

The weak form of the linear momentum balance (6) is given by

$$ (\nabla \cdot \boldsymbol{\sigma},\mathbf{q})_{{\Omega}}+(\mathbf{f}\cdot \mathbf{q})_{{\Omega}}= 0\qquad (\forall\,\,\mathbf{q}\in \mathbf{U}({\Omega})) $$
(43)

where U(Ω) is given by

$$\mathbf{U}({\Omega})\equiv \big\{\mathbf{q}=(u,v,w):u,v,w\in H^{1}({\Omega}),\mathbf{q}=\mathbf{0}\,\,\text{on}\,\,{{\Gamma}_{D}^{p}}\big\} $$

where Hm(Ω) is defined, in general, for any integer m ≥ 0 as

$$H^{m}({\Omega})\equiv\big\{w:D^{\alpha}w\in L^{2}({\Omega})\,\,\forall |\alpha| \leq m \big\}, $$

where the derivatives are taken in the sense of distributions and given by

$$D^{\alpha}w=\frac{\partial^{|\alpha|}w}{\partial x_{1}^{\alpha_{1}}..\partial x_{n}^{\alpha_{n}}},\,\,|\alpha|=\alpha_{1}+\cdots+\alpha_{n}, $$

We know from tensor calculus that

$$ (\nabla \cdot \boldsymbol{\sigma},\mathbf{q})_{{\Omega}}\equiv (\nabla ,\boldsymbol{\sigma}\mathbf{q})_{{\Omega}}-(\boldsymbol{\sigma}:\nabla \mathbf{q})_{{\Omega}} $$
(44)

Further, using the divergence theorem and the symmetry of σ, we arrive at

$$ (\nabla ,\boldsymbol{\sigma}\mathbf{q})_{{\Omega}}\equiv (\mathbf{q},\boldsymbol{\sigma}\mathbf{n})_{\partial {\Omega}} $$
(45)

We decompose ∇q into a symmetric part \((\nabla \mathbf {q})_{s}\equiv \frac {1}{2}\big (\nabla \mathbf {q}+(\nabla \mathbf {q})^{T}\big )\equiv \boldsymbol {\epsilon }(\mathbf {q})\) and skew-symmetric part (∇q)ss and note that the contraction between a symmetric and skew-symmetric tensor is zero to obtain

figure a

From Eqs. 434445 and 46, we get

$$(\boldsymbol{\sigma}\mathbf{n},\mathbf{q})_{\partial {\Omega}} - (\boldsymbol{\sigma}:\boldsymbol{\epsilon}(\mathbf{q}))_{{\Omega}} + (\mathbf{f},\mathbf{q})_{{\Omega}}= 0 $$

which, after invoking the traction boundary condition, results in the discrete weak form for the mth coupling iteration as

$$ (\mathbf{t}^{n + 1},\mathbf{q}_{h})_{{{\Gamma}_{N}^{p}}} - (\boldsymbol{\sigma}^{m,n + 1}:\boldsymbol{\epsilon}(\mathbf{q}_{h}))_{{\Omega}} + (\mathbf{f}^{n + 1},\mathbf{q}_{h})_{{\Omega}}= 0 $$

Replacing m by m + 1 and subtracting the two equations, we get

$$ (\delta^{(m)} \boldsymbol{\sigma}:\boldsymbol{\epsilon}(\mathbf{q}_{h}))_{{\Omega}}= 0 $$

Appendix E: contracted notation

The generalized Hooke’s law (11) written in indicial notation

$$\sigma_{ij}=\sigma_{0_{ij}}+\mathbb{M}_{ijkl}\epsilon_{kl}-\alpha_{ij}(p-p_{0}) $$

is rewritten in contracted notation as

$$\sigma_{\beta}=\sigma_{0_{\beta}}+\mathbb{M}_{\beta\gamma}\epsilon_{\gamma}-\alpha_{\beta}(p-p_{0}) $$

where the transformation is accomplished by replacing the subscripts ij (or kl) by β (or γ) using the following rules

$$\begin{array}{ccc} ij\,\,(\text{or}\,\,kl) & \longleftrightarrow & \beta\,\,(\text{or}\,\,\gamma)\\ 11 & \longleftrightarrow & 1\\ 22 & \longleftrightarrow & 2\\ 33 & \longleftrightarrow & 3\\ 23\,\,(\text{or}\,\,32) & \longleftrightarrow & 4\\ 31\,\,(\text{or}\,\,13) & \longleftrightarrow & 5\\ 12\,\,(\text{or}\,\,21) & \longleftrightarrow & 6 \end{array} $$

In other words, the stress and strain tensors are represented as

$$\begin{array}{@{}rcl@{}} &&\boldsymbol{\sigma}= \left\{\begin{array}{cccccc} \sigma_{xx} & \sigma_{yy} & \sigma_{zz} & \sigma_{yz} & \sigma_{xz} & \sigma_{xy} \end{array}\right\}^{T},\\ &&\boldsymbol{\epsilon}= \left\{\begin{array}{cccccc} \epsilon_{xx} & \epsilon_{yy} & \epsilon_{zz} & \epsilon_{yz} & \epsilon_{xz} & \epsilon_{xy} \end{array}\right\}^{T} \end{array} $$

Appendix F: material symmetry

A material is said to possess a symmetry with respect to an orthogonal transformation χ if the elasticity tensor \(\mathbb {M}\) is invariant under the orthogonal transformation χ as follows (see Ting [31])

$$\mathbb{M}^{\prime}=\mathbf{K}\mathbb{M}\mathbf{K}^{T}\equiv \mathbb{M} $$

where \(\mathbb {M}^{\prime }\) is the transformed elasticity tensor and K is given by

$$\begin{array}{@{}rcl@{}} \mathbf{K}&=& \left[\begin{array}{ccccccc} \chi_{11}^{2} & \chi_{12}^{2} & \chi_{13}^{2} & 2\chi_{12}\chi_{13} & 2\chi_{13}\chi_{11} & 2\chi_{11}\chi_{12}\\ \chi_{21}^{2} & \chi_{22}^{2} & \chi_{23}^{2} & 2\chi_{22}\chi_{23} & 2\chi_{23}\chi_{21} & 2\chi_{21}\chi_{22}\\ \chi_{31}^{2} & \chi_{32}^{2} & \chi_{33}^{2} & 2\chi_{32}\chi_{33} & 2\chi_{33}\chi_{31} & 2\chi_{31}\chi_{32}\\ \chi_{21}\chi_{31} & \chi_{22}\chi_{32} & \chi_{23}\chi_{33}\\ \chi_{31}\chi_{11} & \chi_{32}\chi_{12} & \chi_{33}\chi_{13} & & \mathbf{K}^{\prime} & \\ \chi_{11}\chi_{21} & \chi_{12}\chi_{22} & \chi_{13}\chi_{23} \end{array}\right]\\ \mathbf{K}^{\prime}&=& \left[\begin{array}{ccccccc} \chi_{22}\chi_{33}+\chi_{23}\chi_{32} & \chi_{23}\chi_{31}+\chi_{21}\chi_{33} & \chi_{21}\chi_{32}+\chi_{22}\chi_{31}\\ \chi_{32}\chi_{13}+\chi_{33}\chi_{12} & \chi_{33}\chi_{11}+\chi_{31}\chi_{13} & \chi_{31}\chi_{12}+\chi_{32}\chi_{11}\\ \chi_{12}\chi_{23}+\chi_{13}\chi_{22} & \chi_{13}\chi_{21}+\chi_{11}\chi_{23} & \chi_{11}\chi_{22}+\chi_{12}\chi_{21} \end{array}\right] \end{array} $$

where χ given by

$$\boldsymbol{\chi}=\mathbf{I}-2\mathbf{n}\mathbf{n}^{T} $$

is a reflection with respect to a plane whose normal is n. The plane is also referred to as the plane of material symmetry.

1.1 F.1 Transverse isotropy

The symmetry planes can be one of the following three possibilities

  1. 1.

    x = 0 plane and any plane that contains the x-axis

  2. 2.

    y = 0 plane and any plane that contains the y-axis

  3. 3.

    z = 0 plane and any plane that contains the z-axis

For the Mandel’s problem (see Fig. 2), the material symmetry planes are the second possibility (y = 0 plane and any plane that contains the y-axis). In lieu of that, the elasticity tensor \(\mathbb {M}\) reduces to

$$ \left[\begin{array}{ccccccc} \mathbb{M}_{11} & \mathbb{M}_{12} & \mathbb{M}_{13} & 0 & 0 & 0\\ \mathbb{M}_{12} & \mathbb{M}_{22} & \mathbb{M}_{12} & 0 & 0 & 0\\ \mathbb{M}_{13} & \mathbb{M}_{12} & \mathbb{M}_{11} & 0 & 0 & 0\\ 0 & 0 & 0 & 2\mathbb{M}_{44} & 0 & 0\\ 0 & 0 & 0 & 0 & (\mathbb{M}_{11}-\mathbb{M}_{13}) & 0\\ 0 & 0 & 0 & 0 & 0 & 2\mathbb{M}_{44} \end{array}\right] $$
(47)

with five independent components \(\mathbb {M}_{11}\), \(\mathbb {M}_{12}\), \(\mathbb {M}_{13}\), \(\mathbb {M}_{22}\) and \(\mathbb {M}_{44}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dana, S., Wheeler, M.F. Convergence analysis of fixed stress split iterative scheme for anisotropic poroelasticity with tensor Biot parameter. Comput Geosci 22, 1219–1230 (2018). https://doi.org/10.1007/s10596-018-9748-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10596-018-9748-2

Keywords

Navigation