Staggered explicit-implicit time-discretization for elastodynamics with dissipative internal variables
Abstract
An extension of the two-step staggered time discretization of linear elastodynamics in stress-velocity form to systems involving internal variables subjected to a possibly non-linear dissipative evolution is proposed. The original scheme is thus enhanced by another step for the internal variables which, in general, is implicit, although even this step might be explicit if no spatial gradients of the internal variables are involved. Using an abstract Banach-space formulation, a-priori estimates and convergence are proved under a CFL condition. The developed three-step scheme finds applications in various problems of continuum mechanics at small strain. Here, we consider in particular plasticity, viscoelasticity (creep), diffusion in poroelastic media, and damage.
AMS Subject Classification 65M12, 65P10, 65Z05, 74C10, 74F10, 74H15, 74R20, 74S05, 76S05.
Keywords Elastodynamics, explicit discretization, fractional steps, mixed finite-element method, plasticity, creep, poro-elasticity, damage.
1 Introduction
In computational mechanics one can distinguish two main classes of time-dependent problems, quasistatic and dynamic. Focusing on the latter one, one can further distinguish two other cases: (i) Low-frequency regimes, which are typically related with vibrations of structures and where the energy is not dominantly transmitted through space. In this case implicit time-discretization is relatively efficient, even though large systems of algebraic equations are to be solved at each time step. (ii) High-frequency regimes which arise typically within wave propagation and for which only explicit time-discretizations are reasonably efficient, in particular in three-dimensions. These explicit methods can essentially be used in hyperbolic problems, as mere elastodynamics (treated here) or the elasto-magnetic Maxwell system or some conservation laws. Yet, many applications need to combine the convervative hyperbolic problems with various dissipative processes of a parabolic character, involving typically some internal variables. For parabolic problems, however, explicit methods are known to be problematic due to severe time-step restrictions. Therefore, we propose and analyse an explicit/implicit scheme, using the fractional-step (also called staggered) technique. In fact, the proposed method becomes completely explicit in the absence of gradients of internal variables, as it can be in plasticicity or viscoelasticity as in Sect. 5.1, or an interfacial variant of damage models (so-called delamination) as in [41].
Our starting point is the linear elastodynamic problem: Find the displacement satisfying
(1.1a) | |||||
(1.1b) | |||||
(1.1c) |
for a fixed time horizon. Here is a bounded Lipschitz domain, or , is its boundary, and the unit outward normal. The dot-notation stands for the time derivative. In (1.1), denotes the mass density, is the elasticity tensor which is symmetric and positive definite, denotes the small-strain tensor defined as , and is a symmetric positive semidefinite 2nd-order tensor determining the elastic support on the boundary. The terms appearing in the second member of (1.1) are, the bulk force and the surface loading . In (1.1c), denotes the initial displacement and the initial velocity. For a more compact notation, we write the initial-boundary-value problem (1.1) in the following abstract form
(1.2) |
Here is the kinetic energy, is the stored energy, and is the external force, while denotes the Gâteaux derivative. In the context of (1.1), we have
and
Thus is a linear functional, let us denote it shortly by .
For high-frequency wave propagation problems, implicit time discretizations are computationally expensive, especially in the three dimensional setting. Therefore, we focus our attention on explicit methods. The simplest explicit scheme is the following second-order finite difference scheme
(1.3) |
Here denotes the time step, and (resp. ) denote some discrete approximations of the respective continuous functionals obtained by a suitable finite-element method (FEM) with mesh size . For simplicity, we assume the mass density constant (or at least piecewise constant in space) so that the kinetic energy does not need any numerical approximation. In particular, a numerical approximation leading to a diagonal the mass matrix in (1.3), typically referred to as mass lumping, is an important ingredient so as to obtain efficient explicit methods. Here we will consider that is discretized in an element-wise constant way so that leads to a diagonal matrix even without any approximation. Multiplying (1.3) by and using the scheme, it is easy to show that the energy preserved is
This is an energy, i.e., a positive quantity under the following Courant-Fridrichs-Lewy (CFL) condition [20]
(1.4) |
for any from the respective finite-dimensional subspace. The CFL typically bounds the time discretization step with the smallest element size on a FEM discretization. This method has frequently been used and analysed from various aspects, including comparison with implicit time discretizations, cf. e.g. [34, 33]. However, the form of the discrete stored energy makes this discretization less suitable for the problem that we wish to consider in this paper where the stored energy is enhanced by some internal variables and (possibly) nonlinear processes on them.
Therefore, we use another explicit discretization scheme, the so-called leap-frog scheme. To this end, we first write the velocity/stress formulation of (1.1a). Introducing the velocity, and the stress tensor , we get
(1.5a) | |||||
(1.5b) | |||||
(1.5c) |
In the abstract form (1.2), when writing with denoting the linear operator and with denoting the trace of on the boundary , this reads as
(1.6a) | ||||||
(1.6b) |
where is the adjoint operator to . The stored energy governing (1.5) is
while the external loading is now split into two parts acting differently, namely
Let us note that (1.6) involves the equation on , as well as, the equation on . Thus is to be understood as the functional on , that is trivial on since no inertia is considered on the )-dimensional boundary . In particular, the “generalized” stress contains, besides the bulk stress tensor, also the traction stress vector. Relying on the linearity of , we have with , as used in (1.6b). Let us note that the adjoint operator in (1.6a) with the traction force determines a bulk force as a functional on test displacements by
which clarifies the force in (1.6a). The leap-frog time discretization of (1.6) then reads as
(1.7) |
where and are suitable FEM discretizations of and , cf. (3.1a) and (3.1) below, and
(1.8) |
As mentioned before, we assume here that is discretized in an element-wise constant way so that leads to a diagonal matrix. In this case we do not need to employ numerical integration to approximate the mass matrix. For higher-order discretizations, however, mass lumping is necessary so as to obtain explicit discretization schemes. We refer to [10, 31, 50] for details in the case . The proposed FEM leads to a block diagonal matrix for , which means that the resulting scheme does not require the solution of a big linear system at each iteration in time. The spatial FEM discretization exploits regularity available in linear elastodynamics, in particular that and in (1.5a) live in -spaces. Moreover, the equations in (1.7) are decoupled in the sense that, first, is calculated from the former equation and, second, is calculated from the latter equation assuming, that is known from the previous time step. For , it starts from and from a half time step . For the space discretization, the lower order finite element is obtained for and in this case the velocity is discretized as piecewise constant on rectangular or cubic elements while the stress is discretized by piecewise bi-linear functions with some continuities. Namely the normal component of the stress is continuous across edges of adjacent elements while the tangential component is allowed to be discontinuous. For more details about the space discretization we refer the interested reader to [10]. Alternative discretizations for the linear elasticity problem have been proposed by D. Arnold and his collaborators who designed mixed finite elements for general rectangular and triangular grids [6, 4, 5]. For tetrahedral leap-frog discretization of the elastodynamics see [21]. In general, the leap-frog scheme has been frequently used in geophysics to calculate seismic wave propagation with the finite differences method, cf. e.g. [13, 25, 51].
When taking the average (i.e. the sum with the weights and ) of the second equation in (1.7) in the level and tested by and summing it with the first equation in (1.7) tested by , we obtain
Summing it up, we get that the following discrete energy is conserved
(1.9) |
Note that is the discrete stored energy expressed in terms of the generalized stress. In contrast to (1.3), this formulation allows for enhancement of the discrete stored energy by some internal variables. The energy (1.9) is shown to be a positive quantity under the following CFL condition
(1.10) |
for any from the respective finite-dimensional subspace. Moreover, is often considered, which makes the a-priori estimation easier. Let us also note that the adjective “leap-frog” is sometimes used also for the time-discretization (1.3) if written as a two-step scheme, cf. e.g. [19, Sect. 7.1.1.1].
The plan of this article is as follows: In Section 2, we complement the abstract system (1.6) by another equation for some internal variable and cast its weak formulation without relying on any regularity. Then, in Section 3, we extend the two-step leap-frog discrete scheme (1.7) to a suitable three-step scheme, and study the energy properties of the proposed scheme. Then, in Section 4, we prove the numerical stability of the 3-step staggered approximation scheme and its convergence under the CFL condition (4.1). Such an abstract scheme is then illustrated in Section 5 on several examples from continuum mechanics, in particular on models of plasticity, creep, diffusion, and damage. For illustration of computational efficiency, we refer to [43] where this scheme was implemented for another problem, namely a delamination (i.e. interfacial damage).
It should be emphasized that, to the best of our knowledge, a rigorously justified (as far as numerical stability and convergence) combination of the explicit staggered discretization with nonlinear dissipative processes on some internal variables is new, although occasionally some dissipative nonlinear phenomena can be found in literature as in [45] for a unilateral contact, in [13] for a Maxwell viscoelastic rheology, in [47] for electroactive polymers, in [23] for an aeroelastic system, or in [24] for general thermomechanical systems, but without any numerical stability (a-priori estimates) and convergence guaranteed.
2 Internal variables and their dissipative evolution.
The concept of internal variables has a long tradition, cf. [36], and opens wide options for material modelling, cf. e.g. [35, 37] and references therein. Typically, internal variables are governed by a parabolic-type 1st-order evolution. The abstract system (1.2) is thus generalized to
(2.1a) | |||||
(2.1b) |
The inclusion in (2.1b) refers to a possibility that the convex (pseudo)potential of dissipative forces may be nonsmooth and then its subdifferential can be multivalued.
Combination of the 2nd-order evolution (1.2) with such 1st-order evolution is to be handled carefully. In contrast to the implicit staggered schemes, cf. [42], the constitutive equation is differentiated in time, cf. (1.5a), and it seems necessary to use the staggered scheme so that the internal-variable flow rule can be used without being differentiated in time, even for a quadratic stored energy .
Moreover, to imitate the leap-frog scheme, it seems suitable (or maybe even necessary) that the stored energy may be expressed in terms of the generalized stress as
(2.2) |
where stands for a “generalized” elasticity tensor and is an abstract gradient-type operator. Typically or simply are considered here in the context of continuum mechanics at small strains, cf. the examples in Sect. 5. Here, may not directly enter the balance of forces and is thus to be called rather as some “proto-stress”, while the actual generalized stress will be denoted by . For a relaxation of the last requirement of (2.2) see Remark 4.4 below.
Then, likewise (1.6), we can write the system (2.1) in the velocity/proto-stress formulation as
(2.3a) | |||||
(2.3b) | |||||
(2.3c) | |||||
(2.3d) |
Here is a “generalized” strain and, when multiplied by , it becomes a generalized stress. Actually, (1.6) is a special case of (2.3) when and while , so that indeed .
The energy properties of this system can be revealed by testing the particular equations/inclusions in (2.3) by , , and . Thus, at least formally, we obtain
(2.4a) | |||
(2.4b) | |||
(2.4c) |
The functional is a dissipative rate and the “inf” in it refers to the fact that the dissipative potential can be nonsmooth and thus the subdifferential can be multivalued even at , otherwise an equality in (2.4c) holds. Summing it up and using the calculus and , we obtain the following inequality for the energy,
(2.17) |
Let us now formulate some abstract functional setting of the system (2.3). For some Banach spaces , , and and for a Hilbert space , let be smooth and coercive, be quadratic and coercive, and let be convex, lower semicontinuous, and coercive on , cf. (4.2) below. Intentionally, we do not want to rely on any regularity which is usually at disposal in linear problems but might be restrictive in some nonlinear problems. For this reason, we reconstruct the abstract “displacement” and use (2.3a) integrated in time, i.e.
(2.18) |
Moreover, we still need another Banach space and define the Banach space equipped with the standard graph norm. Then, by definition, we have the continuous embedding and the continuous linear operator . We assume that is embedded into densely, so that and that is identified with its dual , so that we have the so-called Gelfand triple
We further consider the abstract elasticity tensor as a linear continuous operator . Therefore provided so that the equation (2.18) is meant in and one needs . Let us note that , , , and , so that provided and also and . In particular, the equation (2.3b) can be meant in if integrated in time, and one needs valued in .
For a Banach space , we will use the standard notation for Bochner spaces of the Bochner measurable functions whose norm is integrable with the power or essentially bounded if , and the space of functions from whose distributional time derivative is also in . Also, will denote the space of functions whose -derivative is continuous, and will denote the space of weakly continuous functions . Later, we will also use , denoting the space of linear bounded operators normed by the usual sup-norm.
A weak formulation of (2.3b) can be obtained after by-part integration over the time interval when tested by a smooth function. It is often useful to consider
(2.19) |
and to use integration by-parts for the term . We thus arrive to the following definition.
Definition 2.1 (Weak solution to (2.3).)
Let us note that the remaining initial condition is contained in (2.20a). Definition 2.1 works successfully for , i.e. for rate-dependent evolution of the abstract internal variable , so that . For the rate-dependent evolution when , we would need to modify it. Here, we restrict ourselves to , because of the a-priori estimates in Proposition 4.1.
3 A three-step staggered time discretization
We derive in this section the leap-frog discretization of (2.3a,b) combined with an implicit discretization for (2.3c), using a fractional-step split (called also a staggered scheme) with a mid-point formula for (2.3c). Instead of a two-step scheme (1.7), we obtain a three-step scheme and therefore, from now on, we abandon the convention of a half-step notation as used in (1.7) and write instead of .
To this aim, we consider sequences of nested finite-dimensional subspaces , , and where the values of the respective discrete variables , , and will be, assuming that their unions are dense in the respective Banach spaces. We will use an interpolation operator and the embedding operator ; it is important that the collection is uniformly bounded and, since is dense in , the sequence converges to the identity on strongly. We consider . Let us note that we allow for a “non-conformal” approximation of , i.e., is not necessarily a subspace of . This is in agreement with discretizations of the velocity as in [10, 9, 11, 18, 50].
Considering that we know from previous step , then the proposed discretization scheme is
1) calculate : | (3.1a) | |||||
2) calculate : | (3.1b) | |||||
3) calculate : | ||||||
and : | (3.1c) |
where and are from (1.8). It seems important in the non-linear case to compute the variables in the order given above. We note however, that for the linear viscoelastic problem with Maxwell rheology a scheme with a different ordering has been proposed in [27, Part I, Sect.2]. The potentials , , and are considered restricted on , , and , so that their corresponding (sub)differentials , , , and are valued in , , and , respectively. The particular equations/inclusion in (3.1) are thus to be understood in , , , , and , respectively. The only implicit equation is (3.1b). Note however that even this equation becomes explicit if there are no spatial gradients in and . In view of the definition of the convex subdifferential, (3.1b) means the variational inequality
(3.2) |
for any . In any case, the equation (3.1b) posseses a potential
(3.3) |
which is to be minimized on . Therefore the existence of a solution to this inclusion (3.1b), or equivalently of the variational inequality (3.2), can be shown by a direct method, cf. also [42]. The scheme (3.1) is thus to be solved recurrently for , starting from the initial conditions (2.3d) assumed, for simplicity, to live in the respective finite-dimensional spaces.
The energy properties of this scheme can be obtained by imitating (2.4)–(2.17). More specifically, we proceed as follows: we test (3.1a) by , then test (3.1b) by , and eventually test the average of (3.1) at the level and by . Using that and are quadratic as assumed in (2.2), we have
(3.4a) | ||||
where we used also (3.1a), and | ||||
(3.4b) |
Therefore, using again the particular equations/inclusion in (3.1), we get respectively
(3.5a) | |||
(3.5b) | |||
(3.5c) |
Let us also note that, if is assumed, the substitution into the inequality (3.2) gives instead of the dissipation rate in (3.5b), which is a suboptimal estimate except if is degree-1 positively homogeneous.
Summing (3.5) up, we benefit from the cancellation of the terms , which is the usual goal and attribute of well-designed fractional-split schemes. Thus, using also the simple algebra
(3.6) |
we obtain the analog of (2.17), namely
(3.7) |
If is smooth except possibly at zero, there is even equality in (3.7).
Considering some approximate values of the variable with , we define the piecewise-constant and the piecewise affine interpolants respectively by
(3.8a) | |||||
(3.8b) |
Similar meaning is implied for , , , , , etc. The discrete scheme (3.1) can be written in a “compact” form as
(3.9a) | ||||
(3.9b) | ||||
(3.9c) |
to be valid a.e. on the time interval .
4 Numerical stability and convergence
Because the energy (1.9) involves now also the internal variable, the CFL condition has to be modified. More specifically, we assume that
(4.1) |
where and are considered from the corresponding finite-dimensional subspaces. Let us still introduce the Banach space . We further assume invertible and that the collection of the interpolation operators is bounded in .
Proposition 4.1 (Numerical stability.)
Let be constant in time, valued in , , so that , , , the functionals , , and be coercive and be Lipschitz continuous uniformly for in the sense
(4.2a) | ||||
(4.2b) | ||||
(4.2c) |
Let also the CFL condition (4.1) hold with sufficiently small (in order to make the discrete Gronwall inequality effective). Then the following a-priori estimates hold:
(4.3a) | ||||
(4.3b) | ||||
(4.3c) | ||||
(4.3d) |
Proof. The energy imbalance that we have here is (3.7) which can be re-written as
(4.4) |
with an analog of the energy (1.9), namely
(4.5) |
We need to show that is indeed a sum of the kinetic and the stored energies at least up to some positive coefficients. To do so, like e.g. [45, Lemma 4.2] or [50, Sect. 6.1.6], let us write
(4.6) |
where all the duality pairings are between and ; here also (3.1) has been used. Thus, using also , we can write the energy (4.5) as
(4.7) |
and with . The energy yields a-priori estimates if the coefficient is non-negative, which is just ensured by our CFL condition (4.1) used for and . Here in (4.7) is just from (4.1).
Altogether, summing (4.4) for and using (4.7), we obtain the estimate
(4.8) |
where , , and come from (4.2) and (4.7). Here we also have used that the collection is bounded. Using (4.2b), we estimate and , and then use the summability of needed for the discrete Gronwall inequality; here the assumption is needed. The last term in (4.8) is to be estimated by the Young inequality as
with some depending on , , and . Here we needed ; note that this is related with the specific explicit time discretization due to the last term in (3.7) but not with the problem itself. Then we use the discrete Gronwall inequality to obtain the former estimates in (4.3b,c) and the estimates (4.3a,d). Using the discrete Gronwall inequality is a bit tricky because of the term on the left-hand side of (4.8) while there is instead of on the right-hand side of (4.8). To cope with it, we have to rely on being constant (as assumed). We prove the estimate for , then we sum up (4.8) for and to get also on the right-hand side. Note also that, in view of (3.6) for , we have obtained the term on the right-hand side of (4.8) which, however, can simply be ignored if taking the “fictitious” velocity at level as .
The equation gives the latter estimate in (4.3b) by estimating
(4.9) |
for and using also the already proved boundedness of in and the assumed boundedness of uniform in ; here we used also that .
Eventually, the already obtained estimates (4.2b) give bounded in . Therefore is bounded in , hence is bounded in , so that gives the latter estimate in (4.3c).
We will now need the approximation properties for :
(4.10a) | ||||
(4.10b) |
Proposition 4.2 (Convergence.)
Let (2.19) and (4.10a) hold, all the involved Banach spaces be separable, and the assumptions of Proposition 4.1 hold. Moreover, let
(4.11a) | |||
(4.11b) | |||
(4.11c) |
for some Banach space into which is embedded compactly, where and are from (2.19). Then there is a selected subsequence, again denoted converging weakly* in the topologies indicated in the estimates (4.3) to some . Moreover, any obtained as such a limit is a weak solution according Definition 2.1.
Proof. By the Banach selection principle, we can select the weakly* converging subsequence as claimed; here the separability of the involved Banach spaces is used.
Referring to the compact embedding used in the former option in (4.11a,b) and relying on a generalization the Aubin-Lions compact-embedding theorem with being bounded in the space of the -valued measures on , cf. [39, Corollary 7.9], we have strongly in for any .
Further, we realize that the approximate solution satisfy identities/inequality analogous to what is used in Definition 2.1. In view of (2.20a), the equations (3.9c) now means
(4.12a) | ||||
for any valued in and with . Like in (2.20b), the inclusion (3.9b) means | ||||
(4.12b) |
for all . This is completed by (3.9a).
It is further important that the equations in (3.9a) and the first equation in (3.9c) are linear, so that the weak convergence is sufficient for the limit passage there. In particular, we use (4.10a) and the Lebesgue dominated-convergence theorem.
As to the weak convergence of (3.9a) integrated in time towards (3.1a) integrated in time, i.e. towards as used in Definition 2.1, we need to prove that
(4.13) |
for any . By (4.10a), we have also in for any , in particular for . Thus certainly in strongly. Using the weak* convergence in , we obtain (4.13). Moreover, in the limit so that .
For the limit passage in (4.12a), we also use weakly* in because is continuous in the (weakstrong,weak)-mode, cf. (4.11a), and because of the mentioned strong convergence of .
Furthermore, we need to show the convergence . For this, we use again the mentioned generalized Aubin-Lions theorem to have the strong convergence in for any and then the continuity of in the (weakstrong,weak)-mode, cf. the former option in (4.11b). The limit passage of (4.12b) towards (2.20b) then uses also the weak lower semicontinuity of and the weak convergence in ; here for this pointwise convergence in all time instants and in particular in , we also used that we have some information about , cf. (4.3d).
So far, we have relied on the former options in (4.11a,b) and the Aubin-Lions compactness argument as far as the -variable is concerned. If is quadratic (as e.g. in the examples in Sects. 5.1–5.2 below), we can use the latter options in (4.11a,b) and simplify the above arguments, relying merely on the weak convergence and .
Remark 4.3 (Alternative weak formulation)
Here, we used the weak formulation of (2.3c) containing the term which often does not have a good meaning since may not be regular enough in some applications. This term is thus eliminated by substituting it, after integration over the time interval, by or even rather by . This however, would bring even more difficulties because we would need to prove a strong convergence of , or of , or in our explicit-discretization scheme, which seems not easy.
Remark 4.4 (Nonquadratic )
Remark 4.5 (State-dependent dissipation)
The generalization of dependent also on or even on is easy. Then is to be replaced by the partial subdifferential and (3.1b) should use instead of .
Remark 4.6 (Spatial numerical approximation)
From the coercivity of the stored energy , we have for any and thus, from (3.1a), so that , although the limit cannot be assumed valued in in general. Similarly, from (3.1), one can read that although this cannot be expected in the limit in general. Anyhow, on the time-discrete level, one can use the FEM discretization similarly as in the linear elastodynamics where regularity can be employed, cf. [10, 9, 11, 50] for a mixed finite-element method and [18] for the more recently developed staggered discontinuous Galerkin method for elastodynamics.
Remark 4.7 (Other explicit-implicit schemes)
Combination of explicit and implicit time discretization might not only be due to parabolic evolution of internal variables but also due to geometrical reasons, e.g. transmission through a thin layer, that lead to a very restrictive CFL condition, cf. [14].
5 Particular examples
We present three examples from continuum mechanics of deformable bodies at small strains of different characters to illustrate applicability of the ansatz (2.2) and the above discretization scheme. Various combinations of these examples are possible, too, covering thus a relatively wide variety of models.
We use a standard notation concerning function spaces. Beside the Lebesgue -spaces, we denote by the Sobolev space of functions whose distributional derivatives are from .
5.1 Plasticity or creep
The simplest example with quadratic stored energy and local dissipation potential is the model of plasticity or creep. The internal variable is then the plastic strain , valued in the set of symmetric trace-free matrices . For simplicity, we consider only homogeneous Neumann or Dirichlet boundary conditions, so that simply and . The stored energy in terms of strain is
(5.1) |
which is actually a function of the elastic strain . The additive decomposition is referred to as Green-Naghdi’s [26] decomposition. This energy leads to
(5.2) |
Let us note that , i.e. the elastic strain , and that the proto-stress is indeed different from the actual stress .
The dissipation potential is standardly chosen as
(5.3) |
with a prescribed yield stress and a positive semidefinite viscosity tensor. The dissipation rate is then . For and , we obtain mere creep model or, in other words, the linear viscoelastic model in the Maxwell rheology. For both and , we obtain viscoplasticity. For and , we would obtain the rate-independent (perfect) plasticity but our Proposition 4.1 does not cover this case (i.e. is not admitted).
The functional setting is , where denotes symmetric -matrices. Thus by Korn’s inequality.
A modification of the stored energy models an isotropic hardening, enhancing (5.1) as
(5.4) |
so that the energy from (5.2) is modified as
(5.5) |
In the pure creep variant , this is actually the standard linear solid (in a so-called Zener form), considered together with the leap-frog time discretization in [7]. The isochoric constraint can then be avoided, assuming that is positive definite.
All these models lead to a flow rule which is localized on each element when an element-wise constant approximation of is used, and no large system of algebraic equations need to be solved so that the combination with the explicit discretization of the other equations leads to a very fast computational procedure.
Another modification for gradient plasticity by adding terms into the stored energy is easily possible, too. This modification uses and (2.19) with and makes, however, the flow rule nonlocal but at least one can benefit from that the usual space discretization of the proto-stress uses the continuous piecewise smooth elements which allows for handling gradients if used consistently also for . For the quasistatic variant of this model, we refer to the classical monographs [28, 49], while the dynamical model with is e.g. in [37, Sect.5.2].
Noteworthy, all these models bear time regularity if the loading is smooth and initial conditions regular enough, which can be advantageously reflected in space FEM approximation, too.
The Maxwell visco-elastodynamics was also studied by J.-P. Groby [27, Part I, Sect.2] using a slightly modified time discretization scheme, namely the order of (3.1a) and (3.1) was exchanged.
The CFL condition (4.1) here is actually the same as the standard one (1.4). This is because the internal variable actually does not influence the elasticity response and, likewise, the inertia is independent of the internal variable, so the wave speed is not influenced either. The CFL is thus of the form where is the maximal eigenvalue of .
5.2 Poroelasticity in isotropic materials
Another example with quadratic stored energy but less trivial dissipation potential is a saturated Darcy or Fick flow of a diffusant in porous media, e.g. water in porous elastic rocks or in concrete, or a solvent in elastic polymers. The most simple model is the classical Biot model [12], capturing effects as swelling or seepage. In a one-component flow, the internal variable is then the scalar-valued diffusant content (or concentration) denoted by .
As in the previous Section 5.1 , we consider only Neumann or Dirichlet boundary conditions, so that . Here we use the orthogonal decomposition with the spherical (volumetric) part and the deviatoric part and confine ourselves to isotropic materials where the elastic-moduli tensor with the bulk modulus and the shear modulus (= the second Lamé constant), which is the standard notation hopefully without any confusion with the notation used in (1.6). Then the proto-stress . In particular, so that .
Adopting the gradient theory for this internal variable , the stored energy in terms of strain is considered
which, in terms of the (here partial) stress , reads as . Here and are so-called Biot modulus and coefficient, respectively, is a capillarity coefficient, and is a given equilibrium content. From (5.2), we arrive at the overall stored energy as:
(5.9) |
Let us note that , i.e. the elastic strain, and that the proto-stress indeed differs from an actual stress by the spherical pressure part .
The driving force for the diffusion is the chemical potential , i.e. here
(5.10a) | |||
The diffusion equation is | |||
(5.10b) |
with denoting the diffusivity tensor. The system (5.10) is called the Cahn-Hilliard equation, here combined with elasticity so that the flow of the diffusant is driven both by the gradient of concentration (Fick’s law) and the gradient of the mechanical pressure (Darcy’s law). The dissipation potential in terms of , let us denote it by behind this system, is
(5.11) |
For the analysis cf. e.g. [35, Sect. 7.6].
One would expect the dissipation potential as a function of the rate of internal variables, as in (2.3c). In fact, the system (5.10) turns into the form (2.3c) if one takes the dissipation potential as
(5.12) |
with denoting the convex conjugate of . Now, is nonlocal. The functional setting is as in the previous example but now and . For a discretization of the type (3.1b), see [40].
Often, the diffusivity is considered dependent on . Or even one can think about . Then the modification in Remark 4.5 is to be applied. In particular, and .
For this Biot model in the dynamical variant, the reader is also referred to the books [1, 16, 17, 48] or also [35, 37]. In any case, the diffusion involves gradients and in the implicit discretization it leads to large systems of algebraic equations, which inevitably slows down the fast explicit discretization of the mechanical part itself.
For this case also the CFL condition (4.1) is the same as the standard one and leads to a restriction of the form where denotes the maximal speed with which waves propagate in the medium. We note that the pressure velocity which is the maximal speed of propagation in isotropic solids is enhanced in the Biot model. The stability analysis of the discrete scheme is quite technical and does not always lead to a practical CFL condition. A. Ezziani in his Ph.D thesis [22] studied a discretization of Biot’s model similar to (5.2) but the stability analysis of the discrete scheme is very nontrivial, cf. [22, Formula (7.4.11)] and, as he points out, cannot be translated into a practical condition. Therefore, in practice he proposes to use where is the speed of the fast wave and is a constant depending on the order of the particular space discretization used. The attenuation caused by diffusion causes also some dispersion of wave velocities which stay however bounded from above by a high-frequency limit, cf. also [22, Fig. 5.2.1], so the CFL condition expectedly holds uniformly like for the pure elastodynamics.
5.3 Damage
The simplest examples of nonconvex stored energy are models of damage. The most typical models use as an internal variable the scalar-valued bulk damage having the interpretation as a phenomenological volume fraction of microcracks or microvoids manifested macroscopically as a certain weakening of the elastic response. This concept was invented by L.M. Kachanov [32] and Yu.N. Rabotnov [38].
Considering gradient theories, the stored energy in terms of the strain and damage is here considered as
(5.13) |
where is an energy of damage which gives rise to an activation threshold for damage evolution and may also lead to healing (if allowed). The last term is mainly to facilitate the mathematics towards convergence and existence of a weak solution in such purely elastic materials without involving any viscosity, cf. [35, Sect. 7.5.3]. This regularization can also control dispersion of elastic waves. More specifically, the 4th-order term resulted in the momentum-equilibrium equation from the -term in (5.13) causes an anomalous dispersion, i.e. waves with shorter wavelength propagate faster than longer wavelength ones, cf. e.g. [35, Remark 6.3.6]. The -term also facilitates the analysis and controls the internal length-scale of damage profiles.
Let us consider the “generalized” elasticity tensor independent of . As in the previous examples, and . According (2.3a), the proto-stress , denoted by , now looks as ; in damage mechanics, the proto-stress is also called an effective stress with a specific mechanical interpretation, cf. [38]. In terms of , the stored energy is then
(5.17) |
Then and the true stress is then provided is constant and symmetric. The damage driving force (energy) is . When and , then always also in the discrete scheme if .
The other ingredient is the dissipation potential. To comply with the coercivity on with as needed in Proposition 4.1, one can consider either
(5.18) |
with some (presumably small) coefficient . The former option corresponds to a unidirectional (i.e. irreversible) damage not allowing any healing (as used in engineering) while the latter option allows for (presumably slow) healing as used in geophysical models on large time scales.
Since appears nonlinearly in , the strong convergence in is needed. For this, the strain-gradient term with is needed and the Aubin-Lions compact embedding theorem is used. This gives the strong convergence even in the norm of for arbitrarily small provided also is bounded in some norm, which can be shown by using and the Green formula
with depending on . Cf. also the abstract estimation (4.9).
When or are not quadratic but continuously differentiable, one can use the abstract difference quotient (4.14) defined, in the classical form, as
(5.19) |
Of course, rigorously, the -operator in (5.19) is to be understood in the weak form when using it in (3.1b).
Due to the gradient -term in (5.17), the implicit incremental problem (3.1b) leads to an algebraic problem with a full matrix, which may substantially slow down the otherwise fast explicit scheme. Like in the previous model the capillarity, now this gradient theory controls the length-scale of the damage profile and also serves as a regularization to facilitate mathematical analysis. Sometimes, a nonlocal “fractional” gradient can facilitate the analysis, too. Then, some wavelet equivalent norm can be considered to accelerate the calculations, cf. also [3]. As far as the stress-gradient term, it is important that the discretization of the proto-stress in the usual implementation of the leap-frog method is continuous piecewise smooth, so that has a good sense in the discretization without need to use higher-order elements. Here we use that the latter relation in (3.1) is to be understood in the weak form, namely for , which means
for any from the corresponding finite-dimensional subspace of . Thus we indeed do not need higher-order elements, and also we do not need to specify explicitly homogeneous boundary conditions in this boundary-value problem.
The functional setting is , , , and . Then , and is understood as an operator , and is understood as an a operator from to itself.
A particular case of this model is a so-called phase-field fracture, taking as a basic choice
(5.20) |
with denoting the energy of fracture and with controlling a “characteristic” width of the phase-field fracture. The physical dimension of as well as of is m (meters) while the physical dimension of is J/m2. This is known as the so-called Ambrosio-Tortorelli functional [2]. In the dynamical context, only various implicit discretization schemes seems to be used so far, cf. [15, 29, 44, 46]. There are a lot of improvements of this basic model, approximating a mode-sensitive fracture, or -insensitive models (with referring to (5.20)), or ductile fracture, cf. [40]. This last variant combines this model with the plasticity as in Sect. 5.1.
As mentioned above in this case we have anomalous dispersion, i.e. the high frequencies propagate faster, cf. e.g. [35, Remark 6.3.6]. The resulting CFL condition is a combination of the usual CFL (1.4) for the 2nd-order elastodynamic model with the CFL condition for 4th-order plate as in [8]. More specifically, the speed of elastic waves in such combined model is like with the speed in the elastodynamic case (i.e. ) and with the wavelength, cf. [35, Remark 6.3.6] for a one-dimensional analysis. For particular space discretisations, implementable wavelengths are bounded from below just by . This yields to a CFL condition of the type
(5.21) |
Asymptotically, for we can see that is to be small as . For fixed , this is actually very restrictive like in the explicit discretization of the heat equation where it practically prevents from efficient usage of explicit discretizations. However, here the role of is primarily to facilitate rigorous existence of weak solutions of this model and can be assumed to be small. Then the influence of this 4th-order term and this restrictive asymptotics is presumably small, and the usual CFL condition resulting from (5.21) with will dominate except on very fine space discretizations.
Let us eventually remark that better asymptotics of the type for can be obtained by replacing the 4th-order term by a nonlocal term of the order for some small, which would even allow for more general dispersion [30] and simultaneously make the analytically desired regularization of the damage model, cf. [35, Remarks 6.3.7 and 7.5.29]
Acknowledgments
This research has been partly supported from the grants 17-04301S (especially as far as the focus on the dissipative evolution of internal variables concerns) and 19-04956S (especially as far as the focus on the dynamic and nonlinear behaviour concerns) of the Czech Sci. Foundation, and by the institutional support RVO: 61388998 (ČR). T.R. also acknowledges the hospitality of FORTH in Heraklion, Crete.
The authors are deeply thankful to Christos G. Panagiotopoulos for fruitful discussions. Also the careful and critical reading of the original version and valuable suggestions by two anonymous referees are acknowledged.
References
- [1] Y.N. Abousleiman, A.H.-D. Cheng, and F.-J. Ulm, editors. Poromechanics III: Biot Centennial (1905–2005), London, 2005. Taylor & Francis.
- [2] L. Ambrosio and V.M. Tortorelli. Approximation of functional depending on jumps via by elliptic functionals via -convergence. Comm. Pure Appl. Math., 43:999–1036, 1990.
- [3] M. Arndt, M. Griebel, and T.Roubíček. Modelling and numerical simulation of martensitic transformation in shape memory alloys. Continuum Mech. Thermodyn., 15:463–485, 2003.
- [4] D.N. Arnold and G. Awanou. Rectangular mixed finite elements for elasticity. Math. Models Methods Appl. Sci., 15:1417–1429, 2005.
- [5] D.N. Arnold, G. Awanou, and R. Winther. Finite elements for symmetric tensors in three dimensions. Math. Comput., 77:1229–1251, 2008.
- [6] D.N. Arnold and R. Winther. Mixed finite elements for elasticity. Numerische Mathematik, 92:401–419, 2002.
- [7] E. Bécache, A. Ezziani, and P. Joly. A mixed finite element approach for viscoelastic wave propagation. Computational Geosciences, 8:255–299, 2004.
- [8] E. Bécache, G.Derveaux, and P. Joly. An efficient numerical method for the resolution of the Kirchhoff-Love dynamic plate equation. Numer. Meth. P. D. E., 21:323–348, 2005.
- [9] E. Bécache, P. Joly, and C. Tsogka. Fictitious domains, mixed finite elements and perfectly matched layers for 2D elastic wave propagation. J. of Comput. Acoustics, 9(3):1175–1202, 2001.
- [10] E. Bécache, P. Joly, and C. Tsogka. A new family of mixed finite elements for the linear elastodynamic problem. SIAM J. Numer. Anal., 39:2109–2132, 2002.
- [11] E. Bécache, J. Rodríguez, and C. Tsogka. Convergence results of the fictitious domain method for a mixed formulation of the wave equation with a Neumann boundary condition. ESAIM: Math. Model. Numer. Anal., 43:377–398, 2009.
- [12] M.A. Biot. General theory of three-dimensional consolidation. J. Appl. Phys., 12:155–164, 1941.
- [13] T. Bohlen. Parallel 3-D viscoelastic finite difference seismic modelling. Computers & Geosciences, 28:887–899, 2002.
- [14] M. Bonnet, A. Burel, M. Duruflé, and P. Joly. Effective transmission conditions for thin-layer transmission problems in elastodynamics. The case of a planar layer model. ESAIM: Math. Model. Numer. Anal., 50:43–75, 2016.
- [15] M.J. Borden, C.V. Verhoosel, M.A. Scott, T.J.R. Hughes, and C.M. Landis. A phase-field description of dynamic brittle fracture. Comput. Meth. Appl. Mech. Engr., 217—-220:77–95, 2012.
- [16] J.M. Carcione. Wave Fields in Real Media, Wave Propagation in Anisotropic, Anelastic, Porous and Electromagnetic Media. Elsevier, Amsterdam, 2015.
- [17] A.H.-D. Cheng. Poroelasticity. Springer, Switzerland, 2016.
- [18] E.T. Chung, C.Y. Lam, and J. Qian. A staggered discontinuous Galerkin method for the simulation of seismic waves with surface topography. Geophysics, 80:T119–T135, 2015.
- [19] G. Cohen and S. Pernet. Finite Element and Discontinuous Galerkin Methods for Transient Wave Equations. Springer, Dordrecht, 2017.
- [20] R. Courant, K. Friedrichs, and H. Lewy. Über die partiellen Differenzengleichungen der mathematischen Physik. Math. Annalen, 100:32–74, 1928.
- [21] S. Delcourte and N. Glinsky. Analysis of a high-order space and time discontinuous Galerkin method for elastodynamic equations. Application to 3D wave propagation. ESAIM: Math. Model. Numer. Anal., 49:1085–1126, 2015.
- [22] A. Ezziani. Modélisation mathématique et numérique de la propagation d’ondes dans les milieux viscoélastiques et poroélastiques. PhD thesis, Univ. Paris IX Dauphine, 2005.
- [23] C. Farhat, M. Lesoinne, and N. Maman. Mixed explicit/implicit time integration of coupled aeroelastic problems: Three‐field formulation, geometric conservation and distributed solution. Intl. J. Numer. Meth. Fluids, 21:807–835, 1995.
- [24] C.A. Felippa, K.C. Park, and C. Farhat. Partitioned analysis of coupled mechanical systems. Comput. Methods Appl. Mech. Engrg., 190:3247–3270, 2001.
- [25] R.W. Graves. Simulating seismic wave propagation in 3D elastic media using staggered-grid finite differences. Bull. Seismological Soc. Amer., 86:1091–1106, 1996.
- [26] A.E. Green and P.M. Naghdi. A general theory of an elastic-plastic continuum. Arch. Rational Mech. Anal., 18:251–281, 1965.
- [27] J.-P. Groby. Modélisation de la propagation des ondes élastiques générées par un séisme proche ou éloigné à l’intérieur d’une ville. PhD thesis, Université de la Méditerranée - Aix-Marseille II, 2005.
- [28] W. Han and B.D. Reddy. Plasticity. Springer, New York, 1999.
- [29] M. Hofacker and C. Miehe. Continuum phase field modeling of dynamic fracture: variational principles and staggered FE implementation. Intl. J. Fract., 178:113–129, 2012.
- [30] M. Jirásek. Nonlocal theories in continuum mechanics. Acta Polytechnica, 44:16–34, 2004.
- [31] P. Joly and C. Tsogka. Finite Element Methods with Discontinuous Displacement, chapter 11. Chapman & Hall/CRC, Boca Raton, FL, 2008.
- [32] L.M. Kachanov. Time of rupture process under creep conditions. Izv. Akad. Nauk SSSR, 8:26, 1958.
- [33] R. Kolman, J. Plešek, J. Červ, and D. Gabriel. Grid dispersion analysis of plane square biquadratic serendipity finite elements in transient elastodynamics. Int. J. Numer. Meth. Engng., 96:1–28, 2013.
- [34] R. Kolman, J. Plešek, J. Červ, M. Okrouhlík, and P. Pařík. Temporal-spatial dispersion and stability analysis of finite element method in explicit elastodynamics. Intl. J. Numer. Meth. Engr., 106:113–128, 2016.
- [35] M. Kružík and T. Roubíček. Mathematical Methods in Continuum Mechanics of Solids. Springer, Switzeland, 2019.
- [36] G.A. Maugin. The saga of internal variables of state in continuum thermo-mechanics (1893-2013). Mechanics Research Communications, 69:79–86, 2015.
- [37] A. Mielke and T. Roubíček. Rate-Independent Systems – Theory and Application. Springer, New York, 2015.
- [38] Yu.N. Rabotnov. Creep Problems in Structural Members. North-Holland, Amsterdam, 1969.
- [39] T. Roubíček. Nonlinear Partial Differential Equations with Applications. Birkhäuser, Basel, 2nd edition, 2013.
- [40] T. Roubíček. An energy-conserving time-discretisation scheme for poroelastic media with phase-field fracture emitting waves and heat. Disc. Cont. Dynam. Syst. S, 10:867–893, 2017.
- [41] T. Roubíček, M. Kružík, V. Mantič, C.G. Panagiotopoulos, R. Vodička, and J. Zeman. Delamination and adhesive contacts, their mathematical modeling and numerical treatment. In V.Mantič, editor, Math. Methods and Models in Composites, chapter 11. Imperial College Press, 2nd edition.
- [42] T. Roubíček and C.G. Panagiotopoulos. Energy-conserving time-discretisation of abstract dynamical problems with applications in continuum mechanics of solids. Numer. Funct. Anal. Optim., 38:1143–1172, 2017.
- [43] T. Roubíček, C.G. Panagiotopoulos, and C. Tsogka. Explicit time-discretisation of elastodynamics with some inelastic processes at small strains. Preprint arXiv no.1903.11654, 2019.
- [44] T. Roubíček and R. Vodička. A monolithic model for phase-field fracture and waves in solid-fluid media towards earthquakes. Intl. J. Fracture, 219, 2019.
- [45] G. Scarella. Etude théorique et numérique de la propagation d’ondes en présence de contact unilatéral dans un milieu fissuré. PhD thesis, Univ. Paris Dauphine, 2004.
- [46] A. Schlüter, A. Willenbücher, C. Kuhn, and R. Müller. Phase field approximation of dynamic brittle fracture. Comput Mech, 54:1141–1161, 2014.
- [47] S. Seifi, K.C. Park, and H.S. Park. A staggered explicit-implicit finite element formulation forelectroactive polymers. Comput. Methods Appl. Mech. Engrg., 337:150–164, 2018.
- [48] B. Straughan. Stability and Wave Motion in Porous Media. Springer, New York, 2008.
- [49] R. Temam. Mathematical Problems in Plasticity. Gauthier-Villars, Paris, 1985. (French original in 1983).
- [50] C. Tsogka. Modelisation mathématique et numérique de la propagation des ondes élastiques tridimensionnelles dans des milieux fissurés. PhD thesis, Univ. Paris IX Dauphine, 1999.
- [51] J. Virieux. SH-wave propagation in heterogeneous media: Velocity-stress finite-difference method. Geophysics, 49:1933–1957, 1984.