Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Finite Elements in Analysis and Design 113 (2016) 14–29

Contents lists available at ScienceDirect

Finite Elements in Analysis and Design


journal homepage: www.elsevier.com/locate/finel

A fracture-controlled path-following technique for phase-field


modeling of brittle fracture
N. Singh a, C.V. Verhoosel a,n, R. de Borst b, E.H. van Brummelen a
a
Department of Mechanical Engineering, Eindhoven University of Technology, 5600 MB Eindhoven, The Netherlands
b
Department of Civil and Structural Engineering, University of Sheffield, Sheffield, UK

art ic l e i nf o a b s t r a c t

Article history: In the phase-field description of brittle fracture, the fracture-surface area can be expressed as a func-
Received 16 June 2015 tional of the phase field (or damage field). In this work we study the applicability of this explicit
Received in revised form expression as a (non-linear) path-following constraint to robustly track the equilibrium path in quasi-
17 November 2015
static fracture propagation simulations, which can include snap-back phenomena. Moreover, we derive a
Accepted 8 December 2015
fracture-controlled staggered solution procedure by systematic decoupling of the path-following con-
trolled elasticity and phase-field problems. The fracture-controlled monolithic and staggered solution
Keywords: procedures are studied for a series of numerical test cases. The numerical results demonstrate the
Brittle fracture robustness of the new approach, and provide insight in the advantages and disadvantages of the
Phase-field modeling
monolithic and staggered procedures.
Path-following methods
& 2015 Elsevier B.V. All rights reserved.
Staggered solution procedures

1. Introduction developed in the context of snap-back behavior caused by geometrical


non-linearities, over the past decades various enhancements to the
In many problems in brittle-fracture mechanics, phenomena such original path-following procedures have been proposed in order to
as nucleation, propagation, branching and merging occur. Complex increase their versatility and computational efficiency.
crack patterns appear as a consequence of e.g. the presence of multiple A particularly interesting application of path-following techniques
cracks, anisotropy and heterogeneity. Using discrete fracture models it is their use to track snap-back behavior as a result of material non-
is generally difficult to capture such topologically complex crack pat- linearities, especially localized failure phenomena. In such situations
terns, which has led to the development of smeared or continuum the original path-following constraints have proven to lack robustness
crack models, including phase-field models [1,2]. In phase-field models by the fact that they fail to account for the localized nature of the
the crack surface is regularized by a smeared damage (or phase-field) source of non-linearity. Various modified techniques have been pro-
function, which avoids the need for the explicit tracking of fracture posed to account for this localized behavior, among which are a series
surfaces. Over the past years phase-field modeling of fracture has been of (semi-)automatic procedures for selecting degrees of freedom that
applied to a wide range of problems, including dynamic fracturing contribute to the nonlinear behavior of the system [13,14]. Our work
[3,4], large deformation fracturing [5], fracturing of electromechanical builds on the idea that an appropriate path-following technique can be
materials [6], cohesive fracturing [7], and fluid-driven fracture propa- obtained by selecting a physically motivated constraint equation. In
gation [8]. this regard the crack mouth opening displacement (CMOD) and crack
In this work we consider the quasi-static evolution of brittle frac- mouth sliding displacement (CMSD) control equations proposed by De
tures in an elastic solid, where fractures are driven by gradual incre- Borst [15] can be considered as pioneering works. Inspired by these
mentation of the loading conditions. Since softening and snap-back control equations energy-release rate path-following control was
behavior are frequently encountered in such situations, path-following developed for the simulation of localized failure phenomena, including
control is required to adequately track the complete equilibrium path discrete cracking, smeared damage and softening plasticity [16,17]. The
[9]. Path-following techniques have been an indispensable tool in non- versatility of the energy-release rate control has been demonstrated
linear solid mechanics since the pioneering works of Riks [10], Cris- for a variety of applications, including cases in which geometrical and
field [11] and Ramm [12]. While these path-following techniques were material nonlinearities are competing [18].
When applied in the context of discrete fracture simulations, the
n
energy release-rate path-following technique has the ability to indir-
Corresponding author. Tel.: þ 31 40 247 2382.
E-mail addresses: n.singh@tue.nl (N. Singh),
ectly control the rate at which a fracture propagates by proper selection
c.v.verhoosel@tue.nl (C.V. Verhoosel), r.deborst@sheffield.ac.uk (R. de Borst), of the energy dissipation increment. In Ref. [19] it has been shown
e.h.v.brummelen@tue.nl (E.H. van Brummelen). that the energy-release rate control can be successfully applied to

http://dx.doi.org/10.1016/j.finel.2015.12.005
0168-874X/& 2015 Elsevier B.V. All rights reserved.
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 15

phase-field simulations, where the dissipation increment is related to


the fracture-surface area increment through the critical energy release
rate. In the case of phase-field simulations the relation between the
path-following constraint and the fracture-surface area increase can be
made explicit, i.e. the fracture-surface area can be expressed as a
functional of the phase-field solution. This allows for direct prescription
of the surface-area increments. This explicit dependence allows for the
selection of the path-following parameter increment based on a cri-
terion that relates the crack surface growth to the size of the employed
(finite element) mesh, which provides a natural way of controlling the
accuracy of the path-following scheme. In this work we formulate and
study such a fracture-based path-following technique, which – if used
in combination with a monolithic incremental-iterative path-following
Fig. 1. Schematic representation of a domain Ω with regularized fracture surface
procedure – allows for the parametrization of the equilibrium path by
Γ lc ðdÞ representing a fractured solid medium.
specified fracture-surface area increments.
Quasi-static phase-field simulations of brittle fracture phe-
Under the above conditions, the strong form for the displace-
nomena have mostly relied on the use of a staggered solution
ment and phase field is given by:
strategy, in which the elasticity problem and phase-field problems
are decoupled [2]. This staggered solution strategy has been pro- 8
>
> ∇σ¼0 in Ω ð1aÞ
>
ven to be computationally efficient. A drawback of this solution >
>
> Gc  
> l d  lc Δd ¼ 2ð1 dÞH in Ω
2
> ð1bÞ
strategy is that the step sizes need to be selected appropriately in < c
order to control the accuracy of the procedure. The currently ðSÞ σ  n ¼ t on Γ N ð1cÞ ð1Þ
>
>
available staggered schemes are not capable of representing snap- >
>
>
>
> u¼u on Γ D ð1dÞ
back behavior. In this work a staggered fracture-based path-fol- >
: ∇d  n ¼ 0 on Γ ð1eÞ
lowing method is derived from the monolithic scheme, which has
the possibility of reducing the computational effort of the mono-
In this strong form, Gc is the Griffith type critical energy-release
lithic scheme at the cost of only satisfying the path-following
rate and lc is the length scale associated with the phase-field
increments in an approximate sense. This fracture-controlled
regularization of the fracture surface1 (i.e. the width of the cracks)
staggered scheme does, however, inherit the property of the
[1,2]. In order to restrict the fracturing process to tensile stress
underlying monolithic scheme that the fracture propagation
states, the Cauchy stress tensor in the above problem is defined as
increments can directly be controlled (albeit in an approximate
sense). This simplifies the selection of the step size compared to
σ ðε; dÞ ¼ gðdÞσ 0þ ðεÞ þ σ 0 ðεÞ; ð2Þ
e.g. the displacement-based staggered scheme.
In Section 2 we introduce the phase-field formulation for
where gðdÞ ¼ ð1  dÞ2 is the degradation function, σ 0þ and σ 0 are
brittle fracture and its discretization using the finite element
the tensile and compressive parts of the virgin (d ¼0) Cauchy
method. In Section 3 we derive the fracture-based path-following
stress tensor [2], and ε ¼ ∇s u is the infinitesimal strain tensor.
constraint. In this section we also discuss various aspects of the
From the above stress definition it evidently follows that this
corresponding incremental-iterative path-following procedure. In
Section 4 we systematically derive the staggered path-following degradation function must satisfy the conditions gð0Þ ¼ 1 and
scheme, after which the monolithic scheme and staggered scheme gð1Þ ¼ 0. The property that g 0 ð1Þ ¼ 0 ensures that the thermo-
are studied in detail in terms of computational effort and accuracy dynamic driving force for the phase-field model (i.e. the right-
in Section 5. In this section we also study the nature of the snap- hand-side of the phase-field equation) vanishes once a fracture
back behavior encountered in phase-field simulations for brittle has completely evolved.
fracture. Finally, conclusions are drawn in Section 6. Irreversibility, i.e. the notion that the fracture surface can only
extend (Γ_ lc Z 0), is enforced in the strong form (1) by means of the
history field H : Ω-R þ . This history field satisfies the Kuhn–
Tucker conditions for loading and unloading, defined as
2. Phase-field formulation for brittle fracture
ψ 0þ  H r 0; _ Z0;
H _ ψ þ HÞ ¼ 0;
Hð ð3Þ
0
2.1. Problem formulation
with ψ 0þ the tensile part of the virgin elastic energy density.
We consider the evolution of a regularized fracture surface,
Γ lc ðdÞ, in an ndim dimensional elastic medium, Ω  Rndim , under
quasi-static loading (see Fig. 1). The outward-pointing unit normal 2.2. Finite element discretization
vector to the surface of the domain Ω is denoted by n : Γ -Rndim .
Small deformations and deformation gradients are assumed, and To compute an approximate solution to the strong form (1)
the deformation of the medium is described by the displacement using the finite element method, the weak form is derived. Using
field u : Ω-Rndim . The fracture surface, Γ lc ðdÞ, is represented by the function spaces V u ¼ fu A H 1 ðΩÞ∣u ¼ u on Γ D g and V d ¼ H 1 ðΩÞ
the phase field d : Ω-½0; 1, which approaches 1 inside a reg- for the trial functions, and V u0 ¼ fu A H 1 ðΩÞ∣u ¼ 0 on Γ D g and V d for
ularized crack and vanishes far away from the fracture surface.
External tractions, t , are applied along the Neumann boundary Γ N
and prescribed displacements, u, are considered at the Dirichlet 1
Here the length scale lc is defined as in Ref. [2]. We note that in the literature
boundary Γ D . sometimes use is made of the alternative length scale definition ε ¼ lc =2, e.g. [1].
16 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

the test functions, we obtain: 3. Fracture-controlled monolithic solution procedure


8
>
> Find ðu; dÞ A V u  V d such that : Commonly, the solution to the quasi-static nonlinear problem (6) is
>
< R σ : ∇s w dV ¼ R t  w dS
>
8 w A V u0 ð4aÞ computed through an incremental-iterative solution procedure. In such
Ω ΓN
ðWÞ    k
a procedure either the external loading (f ext ) or the boundary dis-
>
> R Gc
>
> Ω þ 2H d  2H e þ Gc lc ∇d  ∇e dV ¼ 0 8 e A Vd ð4bÞ
: lc placement (akp ) is prescribed in a stepwise incremental fashion, with
index k ¼ 1; …; nsteps . In every step, the corresponding solution incre-
ð4Þ
ment, Δak ¼ ak  ak  1 , is computed using Newton–Raphson iterations.
The tangent stiffness matrices required by the Newton solution pro-
This weak form is discretized using a (Bubnov–) Galerkin finite
cedure can be found in Appendix A.
element discretization, for which the displacement field and phase
Evidently, force-controlled and displacement-controlled solu-
field are interpolated by
tion procedures break down in the case of softening or snap-back
behavior, respectively; see e.g. [9]. Therefore, in order to track the
X
nu X
nd
uðxÞ ¼ NuI ðxÞauI ; dðxÞ ¼ N dI ðxÞadI ; ð5Þ complete equilibrium path we need to supplement the system of
I¼1 I¼1 equations with a path-following control.
In this contribution we restrict ourselves to the case of pro-
where nu and nd denote the number of displacement and phase portional loading, i.e. we assume that the external force vector f ext
k

field degrees of freedom, respectively. The vector-valued shape ^


can be written as a load level λ times a “unit” load vector f in the
k

functions NuI ðxÞ : Ω-Rndim and scalar-valued shape functions N dI : case of force loading, and the boundary displacements are
Ω-R span subsets of H 1 ðΩÞ and H1 ðΩÞ, respectively. The nodal expressed as akp ¼ λ a^ in the case of displacement loading.
k

displacement components and phase-field values are respectively The monolithic path-following procedure is outlined in the
u d
represented by au A Rn and ad A Rn . The degrees of freedom are pseudo-code Algorithm 1. This standard algorithm is here pre-
T
assembled in a single vector of coefficients: aT ¼ ½au ; ad . The
T
sented to place some specific algorithmic aspects of the current
Dirichlet boundary conditions are enforced strongly by means of a work in the proper perspective (see Section 3.3). Moreover, this
constraints matrix C, such that a ¼ Caf þ ap , with af the free algorithm will serve as the basis for the novel staggered path-
degrees of freedom and ap the prescribed degrees of freedom. following procedure to be derived in the next section.

Algorithm 1. Monolithic incremental-iterative path-following procedure.

Input: (a0 ; λ ), H0 # State vector, load level & history field


0

Output: ða ; λ Þ, ða2 ; λ Þ; …; ðansteps ; λ steps Þ # Discrete eq. path


1 1 2 n

# Initialization
control¼‘displacement’
# Load steps
for k ¼ 1; …; nsteps :

 ak ¼ ak  1 ; λk ¼ λk  1 #Initialization of Newton iterations
 0 0

 #Newton iterations

 for m ¼ 1; …; m
 max :

  f ; K ¼ assemble_augmented_systemðak
m  1 ; λm  1 ; H
k k1

 int Þ

  ζ ; h; q ¼ assemble_control_equationðak
m  1 ; λm  1 ; controlÞ
k


  ak ; λk ¼ solve_augmented_systemðK; f ; ζ ; h; qÞ
 m m int

  converged ¼ check_convergenceðak ; λk Þ
 m m

  if converged : break


 end

 if converged :

 k
  a ¼ ak ; λk ¼ λk #Update state vector and load level
 m m
 k
  H ¼ update_history_fieldðak ; Hk  1 Þ

  control ¼ select_control_equationðak ; Hk Þ


 else :


 j restart_newton_iterationsðÞ

 end
end

Using the finite element discretization (5), the weak form (4)
can be written as a non-linear system of equations
f int ðaÞ ¼ f ext ; ð6Þ 3.1. The path-following constraint
T
T uT d T uT
with f int ¼ ½f int ; f int  and f ext ¼ ½f ext ; 0T , see Appendix A for the Using a path-following technique, the equilibrium path is
expressions of these force vectors. defined as the set of all points fðaðtÞ; λðtÞÞj t A ½0; Tg which are a
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 17

solution to the non-linear system of equations (6), where t The choice for a particular path-following constraint is dictated
represents the time parameter ranging from 0 to the final time T. by the existence of solutions to the non-linear system of equations
In practice the equilibrium path is represented by a finite sequence (6) upon the incrementation of the path parameter τ. For example,
nsteps
of equilibrium points fðak ; λ Þgk ¼ 0 , computed through the above-
k
force control will be unable to represent softening behavior, while
mentioned incremental-iterative solution procedure. In order to displacement control will fail when snap-back occurs. In fracture
compute the discrete equilibrium points, the general idea of path- mechanics problems various path-following constraints have been
following techniques is to supplement the system of equations (6) found to be very effective. The CMOD (or CMSD) control proposed
with a path-following constraint of the form in [15] has successfully been applied in many cases. Over the past
decade the use of dissipation-based control has been studied
ζ ðak ; Δak ; λk ; Δλk ; ΔτÞ ¼ 0; ð7Þ
extensively and was found to be very reliable for problems in
where Δτ 4 0 is the positive increment of the path-following which severe non-linear behavior is expected [17]. The rationale
parameter τ, which can be regarded as a pseudo-time parameter. behind the dissipation-based control is that, from a physical per-
By the incremental-iterative solution of the non-linear system of spective, dissipation has to be non-negative as a consequence of
equations (6) in combination with this constraint equation, a dis- the irreversibility of fracture propagation. When fracture propa-
crete parametrization of the equilibrium path in terms of the path- gation is the dominant source of dissipation, this control is very
nsteps
following parameter τ is obtained: faðτk Þ; λðτk Þgk ¼ 0 . We note that effective in simulating the evolution of fractures.
the case of force control, i.e. λ ¼ λ
k k1 _
þ λΔτ, is in fact the most Inspired by the idea of dissipation-control, in this contribution we
simple case of path-following control possible: ζ ¼ Δλ  λΔτ
k _ ¼ 0, propose a path-following constraint directly based on the fracture-
_
where λ represents a prescribed loading rate. surface area (or fracture length in 2D). This has become tractable only

Fig. 2. Problem setup and finite element mesh for the single edge notched tension test. (a) Problem setup. (b) Finite element mesh.

Fig. 3. Mesh convergence study for the single edge notched tension test solved with the displacement-controlled staggered solution algorithm. (a) Force vs. displacement.
(b) Crack length vs. displacement.
18 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

with the introduction of phase-field models for brittle fracture, due to and nodal phase-field coefficients, which yields:
the availability of an explicit functional expression for the fracture- Z
u ∂ζ d ∂ζ 1 2
surface area. We note that in the case of Griffith's theory of fracture, h ¼ u ¼0 h ¼ d ¼ dNd þ lc ∇d  ∇Nd dV ð10Þ
∂a ∂a lc
there is a direct relation between the fracture-surface area and the Ω
amount of dissipation, and hence, under specific assumptions, the
with Nd the column vector of phase-field shape functions. In the
control equation developed herein is identical to that developed in T uT dT
remainder we will consider the combined vector h ¼ ½h ; h .
[16] (see Appendix B for details).
Since the constraint (9) does not depend on the load level expli-
citly, it follows that
3.2. Fracture-based path-following constraint
∂ζ
q¼ ¼ 0: ð11Þ
∂λ
In the phase-field formulation for brittle fracture, the fracture
surface area is expressed by The fracture-based path-following constraint (9) has two major
Z benefits. First, it is evident that the path-following parameter is
1 2 2
Γ lc ðdÞ ¼ d þ lc ∇d dV: non-decreasing in time, and hence this constraint choice is
2
ð8Þ
2lc Ω anticipated to yield robust results, also in the case of material
softening and/or snapback. The second advantage is that this
In this work we prescribe the rate of fracture propagation, Γ_ lc , by
choice for the constraint provides an intuitive way of selecting the
means of the path-following constraint
appropriate step size. By requiring that the fracture surface should
ζ ¼ Γ lc ðdk Þ  Γ_ lc τk not propagate across multiple elements within in a single step, the
irreverisibility condition can be adequately imposed. We will fur-
Þ  Γ_ lc Δτ
k1
¼ Γ lc ðd Þ  Γ lc ðd ther study the choice of the step size in Section 5.
k
ð9Þ

Note that this path-following constraint is a non-linear equation of 3.3. Algorithmic aspects
the phase field. Since we apply this constraint in a Newton–
Raphson solution procedure, it is required to compute the deri- In this section we discuss three algorithmic aspects that are specific
vative of this constraint with respect to the nodal displacements, to the current work: (i) the solution of the augmented system of

Fig. 4. Step size study for the single edge notched tension test solved with the displacement-controlled staggered solution algorithm. (a) Force vs. displacement. (b) Crack
length vs. displacement. (c) Maximum phase field value vs. displacement.
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 19

equations within each Newton–Raphson iteration; (ii) the convergence sensitivity of the solution vector. One particular situation in which
criterion employed for the phase-field model; and (iii) the initialization this occurs is when there is no damage present at all, and hence
and selection procedure for the control equation and the restarting J h J ¼ 0. This is, however, not the only situation in which pro-
procedure for the Newton–Raphson iterations. blems occur. Also in the case that the phase field is rather insen-
sitive to the load level (J ∂ad =∂λ J  0), the constraint equation fails.
3.3.1. Solving the augmented system of equations This situation is encountered in the case that elastic behavior
The solution-vector increment and load level increment in step occurs, which happens particularly in the cases of initial loading
k1
k, Δak ¼ ak  ak  1 and Δλ ¼ λ  λ
k k
, are computed using New- and unloading. For this reason, initially displacement control is
ton–Raphson iterations. As a starting vector and load level for used. The switch to the fracture-surface area constraint is made
these iterations the solution to the previous step is used: ak0 ¼ ak  1 after a significant amount of fracture-surface area has been
k1
and λ0 ¼ λ
k formed.
. Subsequently, the solution vector increment is
iteratively updated by Δakm ¼ Δakm  1 þ δakm and Δλm ¼ Δλm  1
k k Depending on the number of Newton–Raphson iterations the
path-parameter increment is adjusted [16]. To this end a target
þ δλm , where m ¼ 1; …; mmax is the Newton–Raphson iteration
k
number of Newton–Raphson iterations, mtarg , is specified. The
counter. For the computation of the update vector δakm and update
path-parameter for the next increment is then scaled with a factor
load level δλm we distinguish between the cases of force loading
k
mtarg =m with a maximum of Δτmax . Evidently, when the path-
and displacement loading. Note that for notational brevity we omit
parameter increment is chosen too large, it can occur that the
the step number k and iteration number m in the following
Newton–Raphson iterations do not converge within mmax itera-
paragraphs (δakm ¼ δa and δλm ¼ δλ). All matrices and vectors are
k
tions. In that case the Newton–Raphson procedure for the same
evaluated at the state ðam  1 ; λm  1 Þ ¼ ða; λÞ, i.e. the solution com-
k k
step is repeated with the path-parameter increment scaled by
puted after m  1 iterations. mtarg =mmax .
Force loading: The external force vector in the discrete equili-
brium equations (6) is then given by f ext ¼ λf^ and the constraints
are imposed by a ¼ Caf þap , where both the matrix C and the 4. Fracture-controlled staggered solution procedure
vector ap are constant throughout the simulation. The solution
update is then computed through
Taking the monolithic path-following procedure in Algorithm 1
! " T #  10 T h ^ i1
for the case of displacement loading (ap ¼ λa)^ as a starting point,
δaf C KðaÞC  CT f^ C λf  f int ðaÞ
¼ @ A ð12Þ we derive a staggered path-following procedure. The most notable
δλ T
h ðaÞC q  ζ ðaÞ difference of this staggered algorithm compared to the monolithic
Algorithm 1 is that no Newton–Raphson iterations are conducted
and δa ¼ Cδaf . We compute the solution to this augmented system
by solving through the Sherman–Morrison procedure discussed in and that the associated convergence criterion is omitted. A con-
e.g. Ref. [17]. The two linear systems of equations encountered in sequence of this is that an additional source of error is introduced
this procedure are solved using a GMRES solver with sparse ILU in the staggered scheme, which, in practice, needs to be com-
pre-conditioning. Since both systems have the same left-hand- pensated for by using smaller load step sizes. By virtue of the fact
side, the pre-conditioner needs to be computed only once per that no (Newton) iterations are performed within a single time
Newton iteration. step, this staggered approach is, however, considerably faster per
Displacement loading: In this case the external force vector in load step than the Newton procedure.
Eq. (6) is equal to zero, and the constraints depend on the load
level: a ¼ Caf þap þ λa.
^ Note that the vector ap accounts for
Dirichlet constraints that are not dependent on the load level λ. The staggered procedure developed herein is outlined in
The solution update is then obtained by Algorithm 2. In the following sections we study the stagger-
! " T #1 ! ed_solution_update procedure. In Section 4.1 we show how
δaf C KðaÞC CT KðaÞa^  CT f int ðaÞ
¼ ð13Þ
δλ h ðaÞC h ðaÞa^ þ q
T T
 ζ ðaÞ

and δa ¼ Cδaf þ δλa.


^ As for the case of force loading we apply a
Sherman–Morrison procedure to solve this augmented system of
equations.

3.3.2. The convergence criterion


After each Newton–Raphson iteration, convergence is checked
based on the residual of the displacement field solution and
phase-field solution, i.e. the solution is accepted when
J ru ðakm Þ J r ϵu J ru ðak1 Þ J and J rd ðakm Þ J r ϵd J rd ðak1 Þ J ; ð14Þ
where ϵu and ϵ are tolerances for the displacement residual ru
d

and phase-field residual rd , respectively.

3.3.3. The control selection procedure and restarting procedure


Consider the sensitivity of the load level to the path-following
parameter:
   
d 1
∂λ _ ∂a  1 d ∂a
¼ Γ lc h  ¼ Γ_ lc h  : ð15Þ
∂τ ∂λ ∂λ
Fig. 5. Phase field at fracture length Γ lc  0:4 mm for the single edge notched
From this expression it is evident that the path-following con- tension test. Note that the plotted grid is merely a visual aid, and is not related to
straint will fail if the vector h in Eq. (10) is orthogonal to the the finite element mesh. (a) Δu n ¼ 4  10  5 mm. (b) Δu n ¼ 0:5  10  5 mm.
20 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

the displacement-controlled staggered procedure as proposed by end, the updates after a single iteration are accepted as the solu-
Miehe et al. [2] follows as a simplification of the incremental- tion increments, i.e. Δa ¼ δa and Δλ ¼ δλ. An approximate solu-
iterative procedure in the previous section. This procedure is tion to the system is then obtained in three steps. In Step 1 the

Algorithm 2. Staggered path-following procedure.

Input: (a0 ; λ ), H0 # State vector, load level & history field


0

Output: ða1 ; λ Þ, ða2 ; λ Þ, … , ðansteps ; λ steps Þ # Discrete eq. path


1 2 n

# Initialization
control¼‘displacement’
# Load steps
for k ¼ 1; …; nsteps :
 k  1 !

 ak ; λk ¼ staggered_solution_update ak  1 ; λk  1 ; Hk  1 ; ∂H  ; control

 ∂λ 

 k
 k ∂H 
H ;  ¼ update_history_fieldðak ; Hk  1 Þ
 ∂λ 


 control ¼ select_control_equationðak ; Hk Þ
end

employed in the initial stage of loading, when fracture propagation phase-field sub-problem is solved with the load level, displace-
does not yet occur. In Section 4.2 the staggered fracture-controlled ment field and history field resulting from the previous load step.
path-following procedure is derived as a simplification of the In Step 2 the load level is updated, and finally in Step 3 the dis-
monolithic fracture-controlled procedure outlined in the previous placement sub-problem is solved with the phase field as com-
section. puted in Step 1 and the load level as determined in Step 2. These
Note that in Algorithm 2 the sensitivity of the history field with three sub-problems can be written in total form as:
respect to the load parameter is evaluated along with the history d
Kdd ðH0 Þad ¼  f int ðad0 ; H0 Þ þ Kdd ðH0 Þad0 ad ¼ Cd adf þ adp ð19aÞ
field itself at the end of each load step. Evaluation of this sensi-
tivity is required at the end of step k for the staggered fracture-
_
λ ¼ λ0 þ λΔτ ð19bÞ
controlled procedure and is given by
8  u u
 > ∂ψ eþ k Kuu ðau0 ; ad Þau ¼  f int ðau0 ; ad Þ þ Kuu ðau0 ; ad Þau0 au ¼ Cu auf þ aup þ λa^
∂Hk <  : ε^ ψ þ ðεk Þ ZHk  1
 ¼ ∂ε  e ; ð16Þ ð19cÞ
∂λ >
:
0 otherwise
P u  Further simplification using
where ε^ ¼ ∇s n u ∂auI Z
I ¼ 1 NI ðxÞ ∂λ is the strain field sensitivity to the d d
f int ð0; H0 Þ ¼ f int ðad0 ; H0 Þ  Kdd ðH0 Þad0 ¼  2 H0 Nd dV ð20Þ
load level λ, with: Ω
∂af u
∂a ∂a
^ and f int ðau0 ; ad Þ ¼ Kuu ðau0 ; ad Þau0 finally results in the control¼ ¼‘displace-
K ¼0 and constraint ¼C þ a: ð17Þ
∂λ ∂λ ∂λ ment’ conditional block in the staggered_solution_update function
In order to simplify notation, in the following sections we drop the shown in Algorithm 3. We note that this algorithm is equivalent to
superscript k indicating the load step. Instead, the initial state for a the staggered algorithm presented in Ref. [2].
given load step is indicated by a subscript 0 (following the nota-
tion for the Newton–Raphson initial estimate) and the updated 4.2. Fracture-controlled staggered procedure
state is represented without sub- or superscripts, i.e. a ¼ a0 þ Δa
and λ ¼ λ0 þ Δλ. Using the fracture control equation (9), the monolithic aug-
mented system of equations (13) for the Newton–Raphson itera-
4.1. Displacement-controlled staggered procedure tions can be written as
2 uu 30 u 1 0  f u 1
K Kud 0 δaf int
A displacement-controlled simulation can be cast into the form 6 du 7B d C B C
4K Kdd 0 5B δ a
@ f A @
C ¼ B  f dint C;
A ð21Þ
of a path-following procedure by using the control equation ζ ¼
_ T d T
δλ _
Γ Δτ
Δλ  λΔτ (and ap ¼ λa),
^ from which it follows that h ¼ 0 and q ¼1. 0 h 0 lc

The monolithic augmented system of equations (13) for the u


with the constraints δau ¼ Cu δauf þ δλa^ and δad ¼ Cd δadf , and with
Newton–Raphson iterations is then given by d
2 uu 30 u 1 0 u
1 h as defined in Eq. (10).
K Kud 0 δa  f int Following the same procedure as for the case of staggered
6 du 7B C B C
Kdd 0 5@ δad A ¼ B
d C
4K @  f int A; ð18Þ displacement control, the updates of the state vector and load
0T 0T 1 δλ _
λΔτ level after a single Newton–Raphson iteration are accepted as the
solution increments, i.e. Δa ¼ δa and Δλ ¼ δλ. In contrast to the
u
with the constraints δau ¼ Cu δauf þ δλa^ and δad ¼ Cd δadf , where case of displacement control discussed in the previous section, in
the constraints matrix C has been decomposed in a displacement this case there is no natural decoupling of the phase-field and the
part Cu and a phase-field part Cd . load level. An approximate solution to the system (21) is therefore
The system (18) can serve as the starting point for the deriva- obtained in only two steps. In Step 1 the combined phase field and
tion of a displacement-controlled staggered procedure. To this load level system is solved, with the displacement field and history
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 21

field following from the previous load step. In Step 2 the dis- 5. Numerical simulations
placement sub-problem is solved with the phase field and load
level as computed in Step 1. In total form, this results in the fol- In this section the performance of the numerical algorithms
lowing sub-problems: outlined in the previous sections is studied. We will investigate the
2   3 ! proposed numerical algorithms using two standard benchmark
d d ∂H
Kdd ðH0 Þ  f^ a0 ; ∂λ 0 d
4 5 a simulations: the single edge notched tension test (Section 5.1) and
d
h ðadÞ
0
T
0 λ the single edge notched pure shear test (Section 5.2). Moreover,
0  
d  1 we will study the performance of the monolithic and staggered
 f int 0; H0 Þ  λ0 f^ ðad0 ; ∂H
d
B ∂λ 0 C schemes for a tension test with multiple pre-existing fractures
¼@ A ð22aÞ
_ (Section 5.3). In contrast to the two benchmark tests, this simu-
Γ lc ðτ  τ0 Þ þ h ða0 Þ a0
d d T d

lation demonstrates the performance of the algorithms in the case


u of fracture interactions.
Kuu au ¼  f int þ Kuu au0 ð22bÞ
For all simulations we assume plane strain conditions. The first
with a d
¼ Cd adf þ adp
and a u
¼ Cu auf þ aup þ
u
λa^ and where the unit Lamé parameter is taken as λ ¼ 121:15 kN=mm2 , while the second
driving force is defined as Lamé parameter (or shear modulus) is μ ¼ 80:77 kN=mm2 . The
      Z    critical energy release rate equals Gc ¼ 2:7  10  3 kN=mm.
d ∂H d ∂H
 d ^ ∂H

f^ ad0 ;
u
¼  K du
a ; ^
a ¼ 2ð1  d ÞN ε : dV
∂λ 0 ∂ε 0 ε 
0 0
Ω ∂ 0
Table 1
Z 
∂H Dependence of various solution characteristics on the step size for the single edge
¼ 2ð1  d0 ÞNd  dV: ð23Þ notched tension test with the displacement-controlled staggered scheme.
Ω ∂λ 0
Δu n (mm) F peak (N) u n;peak (mm) Γ ult (mm)
Using the fact that h ad0 ¼ 2Γ_ lc τ0 this results in the con-
T
d

trol¼ ¼ ‘fracture’ conditional block in the procedure shown in 4  10  5 758.0 0.00628 0.848
Algorithm 3. The augmented system of equations (22a) is solved 2  10  5 741.3 0.00608 0.780
using the Sherman–Morrison procedure. 1  10  5 731.4 0.00597 0.742
0:5  10  5 725.6 0.00589 0.723

Algorithm 3. Staggered solution update procedure.


22 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

Fig. 6. Fracture-controlled Newton–Raphson solutions for the single edge notched tension test. (a) Fixed step size. (b) Adaptive step size.

5.1. Single edge notched tension test

We consider a two-dimensional square specimen of size 1 


1 mm2 with a horizontal notch starting at the left boundary and
ending in the middle of the specimen (Fig. 2a). The bottom
boundary is constrained in the vertical direction and is free to
move in the horizontal direction. In order to eliminate rigid body
motions, the bottom-left corner point is also constrained in hor-
izontal direction. The top boundary is stretched in vertical direc-
tion, and free to move horizontally. For all simulations in this
section the phase-field length scale is taken as lc ¼ 0:015 mm.
Linear triangular meshes for both the displacement field and
phase field have been used, with local mesh refinement along the
anticipated crack path. The number of elements for the various
meshes is reported in Fig. 2b, where h is the characteristic element
size in the refinement region.
Below we will study the performance of three solution algo-
Fig. 7. Comparison of the equilibrium path computed using the monolithic frac-
rithms for this test case: the displacement-controlled staggered ture-controlled scheme with and without tip enrichment.
scheme as employed in e.g. Ref. [2], the fracture-controlled New-
ton–Raphson scheme as outlined in Section 3, and the fracture-
controlled staggered scheme proposed in Section 4. field reaches a value of 1 at a later moment than for a smaller step
size (Fig. 4c). On the other hand we observe that the overall crack
5.1.1. Displacement-controlled staggered scheme length is considerably overestimated for too large step sizes
In this subsection we consider the solution obtained by the (Fig. 4b). In Fig. 5 we illustrate the primary reason for this over-
displacement-based staggered solution procedure as proposed by estimation by considering the phase field at Γ lc  0:4 mm for step
Miehe et al. [2]. The motivation for considering this solution pro- sizes of Δu n ¼ 4  10  5 mm and Δu n ¼ 0:5  10  5 mm. As can be
cedure is to study the dependence of the solution on the selected seen, the delay in the update of the phase field due to the use of
displacement increment size and to enable direct comparison with the staggered solution procedure causes the crack to widen, and
fracture-controlled schemes. This study provides insight in the hence the total fracture length (Γ lc at u n ¼ 0:009 mm) to be
performance of staggered solution procedures compared to the overestimated. In Table 1 we report the computed peak force
monolithic scheme considered in this work. values (F peak ) and its corresponding displacement (u n;peak ), as well
In Fig. 3 we study the influence of the mesh size by con- as the crack length at u n ¼ 0:009 mm (Γ ult ). From the results in
sideration of meshes with characteristic element sizes of h ¼ lc =2, Table 1 it can be inferred that all reported quantities converge
lc =4, lc =6 and lc =8 in the region where the crack is anticipated to linearly under step size refinement.
propagate. For all simulations a relatively large step size of Δu n ¼ Evidently, using uniform step sizes is not optimal in terms of
1  10  5 mm is used. Figures 3a and 3b depict the dependence of computational effort versus step size error. For example, relatively
the response on the selected mesh size. As observed, the measured large step sizes can be used in the elastic regime. As we will see in
response converges upon mesh refinement. Based on these Section 5.1.3, the fracture-based scheme provides a natural adap-
observations, in the remainder of this section we will employ a tive refinement strategy.
fixed mesh size with a characteristic element size of h ¼ lc =6 in the
refinement region. 5.1.2. Fracture-controlled monolithic scheme
In Fig. 4 we study the dependence of the response on the In this section we study the monolithic fracture-controlled
selected step size, Δu n . From the force–displacement curves we scheme (Algorithm 1) with and without adaptive crack size
observe that the overall dissipation is overestimated when using increments. As outlined in Section 3.3.3, we use displacement
too large step sizes. On one hand this is explained by the fact that a control to initiate the solution procedure. In this case Δu n ¼ 1 
too large step size delays the instance of propagation, i.e. the phase 10  4 mm is used. When the crack length increment exceeds
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 23

Fig. 8. Fracture-controlled staggered solutions for the single edge notched tension specimen with various crack length increments.

Table 2
Comparison of the monolithic and staggered path-following schemes for the tension simulation. For the monolithic scheme the mean value and standard deviation (in
brackets) are given when applicable.

Scheme ΔΓ (  h) nsteps niter nsolve F peak (N) u n;peak (mm) u n;snap (mm) Γ ult (mm)

Monolithic 0.52(0.64) 488 2.48 (1.56) 1210 716.8 0.00579 0.00525 0.639
Staggered 2 150 1 150 782.3 0.00651 0.00654 0.717
Staggered 1 275 1 275 754.9 0.00618 0.00597 0.676
Staggered 1
2
529 1 529 738.0 0.00599 0.00565 0.657
Staggered 1
4
1235 1 1235 728.8 0.00590 0.00530 0.647

Newton procedure. In line with this is the observation that


decreasing the step size increases – albeit moderately – the extend
to which the equilibrium path can be computed. A notable dif-
ference between these monolithic results and the staggered
results presented above is that the errors due to the staggered
steps are completely eliminated, i.e. virtually the same result is
obtained independent of the selected crack length increment.
In order to track the complete equilibrium path we have
employed the adaptive scheme as discussed in Section 3.3.3 with
mtarg ¼ 4. The results are presented in Fig. 6b. With this adaptive
step size increment, the same equilibrium path is recovered
regardless of the maximum allowable increment. In fact, the
maximum step size increment is ineffective as a result of the
limitation imposed by the target number of Newton iterations. We
observe that the monolithic solution procedure is very effective in
capturing the peak load. The computed value of F peak ¼ 715:26 N is
Fig. 9. Variation of the displacement increment over the fracture-controlled stag-
free of the step size errors introduced by the staggered procedure
gered iterations for first 80 steps.
in Section 5.1.1, and can be obtained in relatively few steps. In
ΔΓ switch ¼ 1  10  5 mm, the switch is made to fracture control. addition, the fracture-controlled Newton scheme is capable of
This choice for ΔΓ switch is based on the fact that it should be tracking the snap-back part of the equilibrium path.
considerably larger than the machine precision, and considerably The origin of this snap-back behavior is that at the crack tip a
smaller than the representative element size h, since the switch to phase field needs to nucleate. In the case that we enrich the tip of
fracture control should be made well before the fracture starts to the pre-existing fracture with a phase field (see Appendix
propagate. The obtained solution was observed to be insensitive to Appendix C) – thereby regularizing the stress field around the tip –
variations in this switching value. this snap-back feature vanishes (Fig. 7).
In Fig. 6a we show the force–displacement curves for the case
in which the fracture surface increment ΔΓ is kept fixed. We 5.1.3. Fracture-controlled staggered scheme
observe that for all simulations the Newton–Raphson procedure As for the monolithic scheme discussed above, for the fracture-
with a tolerance of 1  10  5 fails to converge at some point in the controlled staggered scheme we use an initial displacement
incrementation process after softening and/or snapback has step size of Δu n ¼ 1  10  4 mm and switch to the fracture-controlled
occurred. This is caused by the fact that the initial estimate for the scheme when the crack length increment exceeds
Newton procedure is outside the radius of convergence of the ΔΓ switch ¼ 1  10  5 mm. In Fig. 8 the results are shown for various
24 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

Fig. 10. Problem setup and finite element mesh for the single edge notched pure shear test.

Fig. 11. Step size study for the single edge notched pure shear test solved with the Fig. 12. Comparison of the equilibrium path computed using the monolithic frac-
displacement-controlled staggered solution algorithm. ture-controlled scheme with and without tip enrichment.

crack length increments, and some solution characteristics are col- step size, ΔΓ , comparable to the staggered scheme with ΔΓ ¼ 12h.
lected in Table 2. Also the number of steps to track the shown equilibrium path is
It is observed that as for the displacement-controlled staggered similar (488 for the monolithic scheme vs. 529 for the staggered
solution procedure, an error is introduced by this staggered scheme. scheme), but evidently the number of linear system solves for the
For the quantities in Table 2 this error is observed to decrease at least monolithic scheme is considerably higher (1210 for the monolithic
linearly with the selected step size. Compared to the displacement- scheme vs. 529 for the staggered scheme) and in addition each
based scheme, the fracture-controlled staggered scheme has two system solve in the monolithic scheme is computationally more
advantages. First, under step-size refinement it converges to the expensive. The error related to the staggered procedure remains
Newton–Raphson solution, including snap-back behavior. Such con- limited to a few percent for both the peak load and the overall
vergence is not observed for the displacement-based staggered crack length. When comparing with the displacement-controlled
scheme of Section 5.1.1. A second advantage is that the step size for the staggered scheme with Δu n ¼ 2  10  5 mm (Table 1), for which a
fracture controlled simulation can be selected conveniently by relating similar number of system solves is required (500), we observe that
it to the representative element size (h). This permits us to allow for the fracture-controlled staggered scheme provides a better
the gradual motion of a crack through the mesh, i.e. the crack is not approximation of the peak load and total crack length than the
permitted to propagate through multiple elements in a single step displacement-controlled scheme. For the peak load the obtained
when ΔΓ is limited by the element size. As indicated above, the improvement is moderate, and can be attributed to the fact that
fracture-controlled procedure serves as an automatic displacement the fracture-controlled scheme automatically provides displace-
step size adjuster. This is shown in Fig. 9 where the displacement step ment step size adjustments. A significant improvement is obtained
size is plotted versus the step size number. As can be seen, the stag- for the total crack length, which is a consequence of the fact that
gered scheme automatically accounts for a smaller (or even negative) the displacement-controlled scheme fails to account for the snap-
displacement increment when crack propagation occurs. back behavior. For the staggered scheme with ΔΓ ¼ 14h a similar
In Table 2 we also compare the monolithic scheme with number of system solves is required as for the monolithic scheme.
adaptive step size with the staggered scheme for various step In this case errors of less than 2% in the peak load and crack length
sizes. We observe that the monolithic scheme on average has a are obtained.
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 25

Fig. 13. Fracture-controlled staggered solutions for the single edge notched shear specimen with various crack length increments.

Table 3 least at a linear rate. As for the tensile test we observe that the
Comparison of the monolithic and staggered path-following schemes for the pure crack path is predicted appropriately by the staggered scheme for
shear simulation. For the monolithic scheme the mean value and standard devia- relatively large step sizes. For ΔΓ ¼ h=2 we observe errors of a few
tion (in brackets) are given when applicable.
percent, while the involved number of system solves is con-
Scheme ΔΓ (  h) nsteps niter nsolve F peak (N) Γ ult (mm) siderably smaller than for the monolithic scheme.

Monolithic 1.17(0.58) 207 3.69(1.06) 764 557.2 0.93 5.3. Multiple inclusion test
Staggered 2 162 1 162 684.6 1.05
Staggered 1 281 1 281 629.6 0.98
Staggered 1 522 1 522 593.6 0.94
We finally study the performance of the fracture-based path-
2
Staggered 1 1061 1 1061 575.3 0.93 following schemes for a test case with complex fracture surface
4
evolution. To this end we consider a 1  1 mm2 tensile test with
six, randomly distributed, pre-existing cracks (Fig. 14). The dis-
5.2. Single edge notched pure shear test cretized displacement field is discontinuous over the pre-existing
cracks, which is established by aligning the elements of the bulk
In this section we investigate the setup represented in Fig. 10a. material with the pre-existing cracks and duplicating the nodes on
The geometry is identical to that considered for the tension the cracks. An irregular triangular finite element mesh with 28 826
simulation discussed above, but pure shear boundary conditions equal-sized linear elements and 14 700 nodes is used to discretize
are used. This means that the vertical displacement component is the bulk material. The element length along the boundaries and
constrained on all four sides of the domain. Moreover, the bottom pre-existing cracks is h ¼ 0:01 mm. The same material parameters
boundary is constrained horizontally, and a prescribed horizontal as for the test cases discussed above have been used. The crack
displacement, u s , is applied to the top boundary. The same length scale is equal to lc ¼ 0:025 mm.
material parameters are used as for the tension simulation. The In Fig. 15 we show the solutions obtained by the monolithic
fracture length scale is equal to lc ¼ 0:015 mm. In order to accu- scheme, with and without tip enrichment. As for the above
rately capture the phase-field evolution, the mesh is refined along experiments we observe overshoots in the response curve in the
the anticipated crack path (Fig. 10b). The characteristic element case that the pre-existing tips are not regularized by a phase field.
size in this refinement region is h ¼ lc =4 ¼ 0:00375 mm, which This effect is here more pronounced due to the fact that the ele-
results in a mesh with 26 472 elements. ments around the tips are relatively coarse (the same element size
In Fig. 11 we study the convergence of the displacement-based is used throughout the complete domain). Evidently, due to the
staggered solution procedure under step size refinement. We iteration-based step size adjustment strategy, the monolithic
observe very close agreement with the results reported in litera- scheme is capable of tracking the snap-back paths.
ture [2]. Using the fracture-controlled Newton–Raphson procedure In Fig. 16 we show six snapshots of the fracture evolution
with adaptive step size and ΔΓ max ¼ 2h (Fig. 12) we observe that pattern. The labels (a)–(f) are reflected in the force–displacement
the bump in the force–displacement curve at crack nucleation is diagram in Fig. 15b. Initially, the specimen is loaded elastically (a),
related to the occurrence of snap-back, a phenomenon not cap- until pre-existing crack 2 propagates toward the right edge of the
tured by the displacement-based staggered scheme. By compar- specimen (b). When this happens, the specimen unloads, after
ison with the results with phase-field tip enrichment, we observe which a secondary crack propagates from the bottom tip of pre-
that this snap-back behavior is closely related to the nucleation of existing crack 2 (c) and merges with pre-existing crack 1 (d). After
the phase-field fracture at the tip of the pre-existing fracture. another unloading stage, finally pre-existing crack 1 propagates
In Fig. 13 we study the influence of the crack-length increment toward the left edge of the specimen (e) until it reaches the left
size for the fracture-controlled staggered solution procedure. This edge and the specimen lost all its load-carrying capacity (f).
figure conveys that the staggered procedure converges to the In Fig. 17 we show the force–displacement curves computed
monolithic result as the step size decreases. In Table 3 we compare using the staggered path-following scheme with ΔΓ ¼ h, h2 and h4.
the monolithic and staggered scheme for various quantities of We observe that already with a step size of h, the correct fracture
interest. The total crack length Γ ult is measured at u s ¼ 0:016 mm. pattern is predicted. The effect that the energy dissipation is
The peak load and total crack length are observed to converge at increased is also observed here. As the step size decreases, the
26 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

Fig. 14. Schematic representation of a 1  1 mm2 tensile specimen with six, randomly generated, pre-existing cracks.

Fig. 15. Comparison of the equilibrium path for the multiple pre-existing crack case computed using the fracture-controlled staggered scheme with and without tip
enrichment.

force–displacement curve converges toward the Newton–Raphson dissipation-based constraint proposed in [16], but is formulated in
case. It is important to note here that since the fractures evolve in terms of the phase field instead of stresses and strains (and rates
stages, during the fracture process there is always one dominant thereof). Formulation of this constraint in terms of the phase field
fracture. This allows for the interpretation that the crack extends invokes a natural decomposition of the phase-field problem and
by approximately a single element in the case that ΔΓ ¼ h is used. the elasticity problem. Based on this decomposition, we developed
In the case that the evolution of a secondary crack is non-negli- a fracture-controlled staggered solution procedure. The derivation
gible, effectively a smaller crack incrementation length (per crack) of this staggered procedure proceeds in essentially the same
is used. In this sense, the choice of the crack length increment is a manner as the derivation of the commonly used displacement-
conservative choice, which permits its usage also in the case of based staggered scheme [2] from a displacement-controlled
complex fracture evolutions as considered here. monolithic solution procedure.
In Table 4 the monolithic scheme and staggered schemes with There are two advantages to the use of a fracture-controlled
various fracture surface increments are compared in terms of the
path-following constraint. First, this constraint permits for the
predicted fracture strength and total crack length. For all simula-
simulation of snap-back phenomena. In the studied numerical
tions regularized pre-existing crack tips are considered. We
examples we have observed that snap-back is typically encoun-
observe errors of a few percent for the staggered scheme with a
tered when a phase-field crack nucleates from a sharp crack tip. By
step size of h=4, which is in agreement with the observations of
enriching the tips of pre-existing fractures with a phase field, this
the benchmark simulations discussed above. In terms of the
number of system solves, this staggered simulation requires snap-back behavior vanishes. The fracture-controlled constraint
approximately half the number of solves of the monolithic opens the doors to a systematic study of this snap-back behavior,
scheme. but this study is considered beyond the scope of this manuscript.
The second advantage of fracture-control is that it provides a
natural way to select the step size increments, this in contrast to
6. Conclusions the displacement-controlled staggered procedure. By requiring
that the cracks propagate gradually through the mesh, the step
In this contribution we studied the application of a fracture- size can be related to the characteristic element size. We have
based path-following constraint for the simulation of phase-field demonstrated that also in the case of multiple cracks this way of
cracks. The employed constraint is closely related to the selecting the fracture step sizes renders meaningful results.
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 27

Fig. 16. Six snapshots of the phase field for the tensile test with pre-existing cracks.

(up to the precision of the Newton–Raphson process). In contrast,


the fracture-based staggered scheme introduces an additional
source of errors by not resolving the non-linearities in every step.
The advantage of the staggered scheme is, however, that it is
computationally cheaper per load step by virtue of the fact that
only a single decoupled elasticity problem and phase-field pro-
blem is solved. In the studied numerical examples we have found
that the staggered scheme is capable of computing typical quan-
tities of interest such as the peak load and total crack length with
errors of a few percent when step sizes of half the representative
element length scale are used. Overall it can be concluded that
when a high accuracy is required, the monolithic scheme is pre-
ferred. When minor inaccuracies are acceptable, the staggered
procedure can be expected to outperform the monolithic scheme
in terms of computational effort.
We note that herein we have considered a staggered solution
Fig. 17. Force-displacement diagrams for the monolithic and staggered path-fol- procedure with a single iteration per load step. It is possible to
lowing schemes for the specimen with multiple pre-existing cracks.
improve the accuracy of this staggered procedure by using mul-
tiple sub-iterations per load step. This will provide the opportunity
to make a trade-off between computational effort and solution
Table 4
Comparison of the monolithic and staggered path-following schemes for the spe- accuracy. A detailed study of a staggered solution procedure with
cimen with multiple pre-existing cracks. For the monolithic scheme the mean value sub-iterations is a topic of further study.
and standard deviation (in brackets) are given when applicable.

Scheme ΔΓ (  h) nsteps niter nsolve F peak (N) Γ ult (mm)


Acknowledgments
Monolithic 0.62 (0.81) 195 3.10 (1.29) 605 1122 1.532
Staggered 1 209 1 209 1178 1.656 This work is part of the Industrial Partnership Programme (IPP)
Staggered 1 327 1 327 1156 1.594
2 ‘Computational sciences for energy research’ of the Foundation for
Staggered 1 540 1 540 1143 1.563
4 Fundamental Research on Matter (FOM), which is part of the
Netherlands Organisation for Scientific Research (NWO).
This research programme is co-financed by Shell Global Solutions
We have studied the performance of the fracture-controlled International B.V. The research of C.V. Verhoosel was funded by
monolithic and staggered solution procedures. Evidently, an the NWO under the VENI scheme. All simulations in this work
advantage of the monolithic scheme is that in every step the non- were performed using the open source software package Nutils
linear system to compute the solution updates is solved exactly (www.nutils.org).
28 N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29

Appendix A. Internal force vectors and tangent stiffness Tucker conditions (3), this expression can be rewritten as
matrices Z Z Z
2 2 1
Γ_ lc ¼ ð1 dÞHd_ dV ¼ ð1  dÞψ 0þ d_ dV ¼  g_ ψ 0þ dV;
Gc Ω Gc Ω Gc Ω
For the phase-field fracture formulation introduced in Section 2
ðB:2Þ
the internal and external force vectors follow directly from sub-
stitution of the finite element basis function (5) as test functions in with degradation function gðdÞ ¼ ð1  dÞ2 . The rate of dissipation,
the weak form problem (4). For the momentum equation, this defined as the external power minus the rate of elastic energy, can
yields: be written as:
Z Z  Z  Z
_ ¼ P W _ ¼ d 1 1
u
f int;I ¼ σ : ∇s NuI dV I ¼ 1; …; nu ðA:1aÞ D t  u_ dS  σ : ε dV ¼ ½σ : ε_  σ_ : ε dV
ΓN dt 2 Ω 2 Ω
Ω Z
1 
Z ¼ ½C : ε : ε_  gC _ 0þ : ε þ C : ε_ : ε dV
u 2 Ω
f ext;I ¼ t  NuI dS I ¼ 1; …; nu ðA:1bÞ Z Z
1
ΓN ¼ g_ σ 0þ : ε dV ¼  g_ ψ 0þ dV: ðB:3Þ
2 Ω Ω
For the phase-field equation the following discrete equations are
_ ¼ Gc Γ_ l .
Combining with equation (B.2) shows that indeed D
obtained: c

Z   
d Gc
f int;I ¼ þ 2H d  2H N dI þ Gc lc
Ω lc Appendix C. Phase-field tip enrichment
∇d  ∇N dI dV I ¼ 1; …; nd ðA:2aÞ
When simulating the fracture process in specimens with pre-
d d
existing cracks, in principle the tip-stresses will be singular when
f ext;I ¼0 I ¼ 1; …; n ðA:2bÞ these cracks are modeled as strong discontinuities. Evidently, in a
The corresponding tangent stiffness matrices follow by differ- finite element context, finite stresses are obtained due to the
entiation of these forces with respect to the nodal solution vectors regularizing effect of the interpolation functions. However, in
as: principle, this regularizing effect is merely a discretization error. In
Z relation to phase-field modeling, the tip stress does influence the
u
f int;I value of the phase field at the tip [4], which causes a significant
K uu
IJ ¼ u ¼ ∇s NuI : C : ∇s NuJ dV ðA:3aÞ
∂aJ Ω grid size dependence of the phase-field nucleation at the tip. In
order to moderate this mesh dependence, in this work we enrich
u Z the fracture tips of pre-existing cracks with a phase field, thereby
f int;I
K ud
IJ ¼ ¼ 2ðd  1ÞN dJ ∇s NuI : σ 0þ dV ðA:3bÞ regularizing the stress field at these tips.
∂adJ Ω
In order to enrich the tips of pre-existing cracks, we compute
d Z   the history field prior to loading, H0 : Ω-R. In order to obtain this
∂f int;I s u ∂H field, we first solve the weak form problem for the phase field
K du
IJ ¼ ¼ 2ðd  1ÞN d
∇ N : dV ðA:3cÞ
∂auJ Ω
I J
∂ε d : Ω-R:
0

8
d Z   < Find d0 A V dtip such that :
∂f int;I Gc R ðC:1Þ
K dd
IJ ¼ ¼ þ 2H N dI N dJ þGc lc ∇NdI  ∇N dJ dV ðA:3dÞ : d0 eþ l2c ∇d0  ∇e dV ¼ 0 8 eA V dtip;0
∂adJ Ω lc Ω

with V dtip ¼ fd A H 1 ðΩÞj d ¼ 1 on Γ tip g, with Γ tip  Ω the set of


0 0
where C ¼ ∂σ =∂ε is the material tangent, and σ 0þ is the tensile part
of the virgin Cauchy stress tensor. Note that the tangent stiffness crack tip points. Subsequently, we determine the corresponding
matrix is generally not symmetric, since history field, H0 : Ω-R:
8
( < Find H0 A H 1 ðΩÞsuch that :
∂H σ 0þ H_ Z 0 R R  
¼ : ðA:4Þ : Ω 2lc ð1  dÞH0 J dV ¼ Ω Gc d J þ lc ∇d  ∇J dV 8 J A H 1 ðΩÞ
0 2 0
∂ε 0 H_ o0

ðC:2Þ
For the discretization of both weak form problems we employ
Appendix B. Equivalence of fracture control with energy linear finite element spaces.
release-rate control We note that an alternative approach to this tip-enrichment
strategy is to model the pre-existing fractures completely by
Since in Griffith's theory for fracture the rate of dissipation is phase-field fractures. An advantage of this approach is that there is
defined as the fracture toughness (Gc ) times the rate at which new no need to create sharp discontinuities in the mesh. However, the
fracture surface is created, the constraint equation derived in creation of such cracks is generally non-trivial when they do not
Section 3 relies on the same assumptions as the energy release align with the finite element grid.
rate path-following control in [16,17]. In this appendix the relation
between the path-following constraint developed in this work and
the constraint of [16,17] is examined. References
Assuming infinitely small path-parameter increments, Δτ -0,
the constraint equation (9) can be used to obtain [1] B. Bourdin, G.A. Francfort, J. Marigo, The variational approach to fracture,
J. Elast. 91 (2008) 5–148.
Z [2] C. Miehe, M. Hofacker, F. Welschinger, A phase field model for rate-
_ ¼ 1 dd_ þ l2 ∇d  ∇d_ dV;
Γ_ lc ðd; dÞ ðB:1Þ independent crack propagation: robust algorithmic implementation based
c
lc Ω on operator splits, Comput. Methods Appl. Mech. Eng. 199 (45–48) (2010)
2765–2778.
which corresponds to the time derivative of the fracture surface [3] B. Bourdin, C.J. Larsen, C. Richardson, A time-discrete model for dynamic
area (8). Using the weak form (4) in combination with the Kuhn– fracture based on crack regularization, Int. J. Fract. 168 (2011) 133–143.
N. Singh et al. / Finite Elements in Analysis and Design 113 (2016) 14–29 29

[4] M.J. Borden, C.V. Verhoosel, M.A. Scott, T.J.R. Hughes, C.M. Landis, A phase-field [12] E. Ramm, Strategies for Tracing the Nonlinear Response Near Limit Points,
description of dynamic brittle fracture, Comput. Methods Appl. Mech. Eng. Springer, Heidelberg, 1981.
217–220 (2012) 77–95. [13] M.G.D. Geers, Enhanced solution control for physically and geometrically non-
[5] C. Miehe, L.M. Schänzel, Phase field modeling of fracture in rubbery polymers. linear problems. Part I—the subplane control approach, Int. J. Numer. Methods
Part I: finite elasticity coupled with brittle failure, J. Mech. Phys. Solids 65 Eng. 46 (2) (1999) 177–204.
(2013) 93–113. [14] M.G.D. Geers, Enhanced solution control for physically and geometrically non-
[6] C. Miehe, F. Welschinger, M. Hofacker, A phase field model of electro- linear problems. Part II—comparative performance analysis, Int. J. Numer.
mechanical fracture, J. Mech. Phys. Solids 58 (10) (2010) 1716–1740. Methods Eng. 46 (2) (1999) 205–230.
[7] C.V. Verhoosel, R. de Borst, A phase-field model for cohesive fracture, Int. [15] R. de Borst, Computation of post-bifurcation and post-failure behavior of
J. Numer. Methods Eng. 96 (1) (2013) 43–62. strain-softening solids, Comput. Struct. 25 (2) (1987) 211–224.
[8] M.F. Wheeler, T. Wick, W. Wollner, An augmented-Lagrangian method for the [16] M.A. Gutiérrez, Energy release control for numerical simulations of failure in
phase-field approach for pressurized fractures, Comput. Methods Appl. Mech. quasi-brittle solids, Commun. Numer. Methods Eng. 20 (1) (2004) 19–29.
Eng. 271 (2014) 69–85. [17] C.V. Verhoosel, J.J.C. Remmers, M.A. Gutiérrez, A dissipation-based arc-length
[9] R. de Borst, M.A. Crisfield, J.J.C. Remmers, C.V. Verhoosel, Non-Linear Finite method for robust simulation of brittle and ductile failure, Int. J. Numer.
Element Analysis of Solids and Structures, 2nd edition, Wiley, Chichester, Methods Eng. 77 (9) (2009) 1290–1321.
2012. [18] F.P. van der Meer, L.J. Sluys, A phantom node formulation with mixed mode
[10] E. Riks, An incremental approach to the solution of snapping and buckling cohesive law for splitting in laminates, Int. J. Fract. 158 (2) (2009) 107–124.
problems, Int. J. Solids Struct. 15 (7) (1979) 529–551. [19] J. Vignollet, S. May, R. de Borst, C.V. Verhoosel, Phase-field models for brittle
[11] M.A. Crisfield, Accelerated solution techniques and concrete cracking, Comput. and cohesive fracture, Meccanica 49 (11) (2014) 2587–2601.
Methods Appl. Mech. Eng. 33 (1) (1982) 585–607.

You might also like