Weinberg 2016
Weinberg 2016
201610004
Key words phase field, brittle fracture, NURBS, crack propagation, thermomigration, void
growth, spallation
Phase field methods allow for convenient and efficient moving interface simulations. In this
paper phase field approaches of different order are presented, and applied to simulate damage
in solids of temperature dependent and non-linear elastic materials. The numerical frame-
work provides a NURBS based finite element method which minimizes the numerical and
computational effort without impairing the smoothness required by the problem.
In order to demonstrate the possibilities of such general phase field approaches a series of
different models from material science and fracture mechanics is investigated. Specifically,
a priori unknown crack propagation in different fracture modes is studied, simulations of
thermomigration in a technical alloy and of void growth are presented and an inverse analysis
of a dynamic fracture experiment is performed. The examples show the versatility of the
presented low-order and high-order phase field approaches.
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
1 Introduction
One of the main challenges in computational mechanics is to predict damage, cracks and
fragmentation patterns. Besides the high demands on the modeling side, the complicated
structure and non-regular behavior of cracks turn numerical simulations of such problems
into a difficult task. A promising tool to overcome such difficulties are phase field methods.
The main idea behind it is to mark the material’s different states or phases by continuous
order parameter field s(x, t), and to let them evolve in space and time. Since the physical
properties within the phases are all in all given, the evolving structure is fully described by the
position and motion of the phase interfaces. However, an order parameter -or phase field- is by
definition a continuous field and thus, the moving boundaries are ’smeared’ over a small but
finite length. Therefore, phase field models constitute so-called diffuse-interface formulations.
Originally derived for diffusion problems, phase field models are meanwhile used for a
variety of interface problems like decomposition, phase transformations or aging of a mi-
crostructure. The core of every model is a Landau free energy functional. For two phases it
∗ Corresponding author E-mail: kerstin.weinberg@uni-siegen.de, Phone: +49 271 740 4641, Fax:
+49 271 740 2241
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
56 K. Weinberg et al.: Phase field approach to damage and crack growth
where Ψcon (s) denotes a configurational energy density which controls the decomposition of
the phases -it often has a double-well shape with minima at s = 0 and s = 1- and Ψsur (∇s)
is the interfacial or surface free energy density, cf. Fig. 1. Additional fields may contribute to
the phase evolution but are omitted here for brevity.
Phase field models of fracture have gained attention only recently, see, e.g., [5, 12, 16,
19, 29]. The main idea is here to regularize the local discontinuities of brittle cracks and to
approximate them by diffuse crack zones. The phase field indicates the state of the material
which may be solid (s = 1) or, if cracked, empty (s = 0). The field s(x, t) is controlled
by an additional differential equation which results in a coupled field problem but completely
avoids the resolution of discontinuities.
For purpose of illustration let us consider a deforming solid with domain B ⊂ R3 and
boundary ∂B ≡ Γ ⊂ R2 . Crack growth corresponds to the creation of new boundaries Γ(t).
Hence the total potential energy of a homogenous but cracking solid is composed of its bulk
energy with free Helmholtz energy density Ψbulk and of surface energy contributions from
growing crack boundaries.
E= Ψbulk dV + Gc dS (2)
B Γ(t)
The fracture-energy density Gc quantifies the material’s resistance to cracking, for brittle frac-
ture it corresponds to Griffith’s critical energy release rate. However, the energy functional (2)
cannot be optimized in general and even an incremental approach [20] is challenging because
of the moving boundaries Γ(t). Highly sophisticated discretization techniques have been de-
veloped to solve such problems, e.g. cohesive zone models [21, 24, 32], the extended finite
element method [18, 27], eroded finite elements or recently developed eigenfracture strate-
gies [23, 26]. In a phase field approach to fracture the set of evolving crack boundaries is
instead replaced by a surface-density functional γ(t) = γ(s(x, t)) and an approximation of
the form
dS ≈ γ(t) dV , (3)
Γ(t) B
which allows to re-write the total potential energy of a cracking solid and to formulate the
optimization problem locally.
bulk
E= Ψ + Gc γ dV → optimum (4)
B
In potential (4) the material’s energy is again composed of two terms, a bulk energy density
Ψbulk and a surface energy contribution Gc γ. By definition γ is only different from zero along
cracks.
Optimization of the potentials (1) or (4) leads to evolution equations for the phase field
s(x, t). For a simple ordering type of phase field, the variation of energy leads the wanted
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 57
ṡ = c1 φ(s) + c2 s (5)
In this formulation φ(s) denotes the reaction term, e.g. φ(s) = 2s3 − s. If the phase field
variable is a conserved quantity like mass concentration or volume fraction, its evolution has
additionally to account for the continuity equation which leads to an evolution equation of
Cahn-Hilliard type.
Obviously, there is no general phase field evolution equation but instead the specific for-
mulation has to map the physics of the underlying problem. In this paper we apply different
phase field models to simulate damage in solids of temperature dependent and non-linear elas-
tic materials. The necessary fundamental equations are stated in Section 2 and the NURBS
based finite element framework used for numerical simulation is shortly outlined in Section 3.
The use of NURBS minimizes the numerical and computational effort without impairing the
smoothness required by the problem. Example problems in Section 4 demonstrate the versa-
tility of the phase field approach for crack growth and damage simulations.
1.5
Ψ
0.5
0
0 0.5 1
s
Fig. 1 (online colour at: www.gamm-mitteilungen.org) Energy density as a function of phase field s:
classical double well functional of the configurational energy density with two minima and simplified
quadratic energy functional employed for phase field fracture.
We start by defining the basic fields. A material point of a solid in its reference configu-
ration B0 is labeled by X = (X1 , X2 , X3 )T and deforms during a time interval [0, t̄] with a
mapping
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
58 K. Weinberg et al.: Phase field approach to damage and crack growth
We denote x = χ(X, t) with the gradient of the deformation F : B0 × [0, t̄] → Rn×n .
∂χ
F = ∇X χ = (8)
∂X
The temperature field
effects the free energy potential of the solid. Additionally, the material state is characterized
by a phase field function
All fields are assumed to be sufficiently smooth. For the phase field we define s ∈ [0, 1].
where P is the first Piola-Kirchhoff stress tensor; ρ0 denotes the mass density, B̄, T̄ the
prescribed body force and traction, and v the material velocity. The boundary of the solid
is subdivided into displacement and traction boundaries ∂B0u , ∂B0σ with ∂B0 = ∂B0u ∪ ∂B0σ ,
∂B0u ∩ ∂B0σ = ∅ and
Let Ψbulk (F , T, s, . . . ) be the local energy density of the bulk material. There may be
additional dependencies of Ψbulk , e.g. on other phase fields or on internal variables, but we
will restrict ourselves here to an isotropic non-linear elastic material with one or two phases.
This material will develop damage and/or cracks. From physics we know that fracture requires
a local state of tension whereas the compressive part of the deformation does not contribute to
crack growth. Therefore, we decompose the deformation gradient (8) into elastic and inelastic
components, F = F e F i . Moreover it is expressed in principal stretches λa , a = 1, 2, 3,
which will be decomposed into tensile λ+ − + −
a and compressive components λa via λa = λa λa
and with
1
λ±
a = [(λa − 1) ± |λa − 1|] + 1 . (14)
2
Then, the fracture insensitive part of the deformation gradient F i reads
3
F =
i
(λ+
a)
(1−s) −
λa na ⊗ N a , (15)
a=1
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 59
where na and N a denote the principal directions of the left and right stretch tensors, respec-
tively. In the following we abbreviate λia = (λ+a)
(1−s) −
λa and write, in a slight misuse of no-
tation, the Helmholtz free energy density equivalently as Ψ := Ψ(F , T, s) ≡ Ψ(λa , T, s) ≡
Ψ(λia , T ).
The first Piola-Kirchhoff stress tensor then follows as
3 3
∂Ψ ∂Ψ ∂λia
P = = i
na ⊗ N a = Pa na ⊗ N a , (16)
∂F a=1
∂λa ∂λa a=1
where q denotes the heat generation rate per unit volume. The corresponding boundary con-
ditions are
where ∂B0l and ∂B0h denote the low and high temperature boundary of the body B0 and it
holds ∂B0 = ∂B0l ∪ ∂B0h ∪ ∂B0n and ∂B0l ∩ ∂B0h ∩ ∂B0n = ∅. We assume the material to be
composed of two components A and B and choose the mole fraction as an order parameter.
Then, the specific heat capacity cp , the mass density ρ and the thermal conductivity k are
set as a convex combination of the parameter cA A A B B B
p , ρ , k and cp , ρ , k of the individual
components.
cp := sA cA B B
p + s cp , ρ := sA ρA + sB ρB , k := sA k A + sB k B (20)
λ = γ̄lc (22)
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
60 K. Weinberg et al.: Phase field approach to damage and crack growth
is determined by surface energy density γ̄. Length lc is the ’interface width’, i.e. the width
of the transition region between the two phases. Now, with a chemical mobility Mc and
j s = −Mc ∇μ follows the phase field evolution equation
ṡ = ∇ · (Mc ∇μ) in B0 × [0, t̄], (23)
which is clearly of Cahn-Hilliard type (6), with source term Mc ∂s Ψcon and c4 = Mc λ for con-
stant mobility. If the phase evolution is additionally driven by a temperature field, the evolu-
tion equation (23) needs to be extended. A detailed, thermodynamically consistent derivation
of such a thermal diffusion model is given in [3]. The resulting equation reads
where M denotes a kinematic mobility [1/sec] and Y (x, s) summarizes all (dimensionless)
driving forces which typically represent a competition of bulk and surface forces.
In phase field fracture such driving force results from a release of stored elastic energy of
the body into the formation of free surfaces. Typically it is derived from an energy potential
and we reformulate potential (4) as
Gc e lc Ψ
E= (Ψ̄ + lc γ ) dV with Ψ̄e = . (27)
B lc Gc
Ψ̄
For normalization we introduce here a potential Ψ̄ which summarizes elastic and fracture
energy contributions, Ψ̄ = Ψ̄e + lc γ, and a characteristic length lc which corresponds to half
of the diffuse ’crack width’, i.e. the transition zone between intact and broken material. The
surface-density functional γ may be understood as a wavenumber of the moving disturbance;
it characterizes the shape of the diffuse zone. By definition, function γ has a small support
and is symmetric to the ’real’ crack path. In the following it is defined as a function of the
phase field parameter s solely.
1
γ= f (s) (28)
lc
Then, a general ansatz of the form
f (s) = c0 |1 − s|2 + c1 lc2 |∇s|2 + c2 lc4 |s|2 + c3 lc6 |∇3 s|2 + . . . (29)
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 61
phase field s
2nd order
4th order
0
−2lc 0 2lc x
Fig. 2 (online colour at: www.gamm-mitteilungen.org) Uniaxial model with a crack at x = 0 and with
a continuous phase field s ∈ [0, 1]; phase field approximation for a second order and a fourth order
approximation of γ.
can be made. Inserting it into (3) and minimizing the corresponding potential (27) ana-
lytically, leads for the simplest uniaxial case to an exponential solution of the form s =
1 − exp(−|x|/lc ), see Fig. 2. Now we determine the constants c0 , c1 , c2 , c3 , . . . in such a
way, that this disturbance is approximated properly. In consequence we obtain for the surface-
density function a second order approximation of the form
1
γ= (1 − s)2 + lc2 (∇s)2 . (30)
2lc
The consideration of higher order terms gives an overall continuous analytical solution. Ap-
proximating the corresponding disturbance s = 1 − exp(−|x|/lc ) · (1 + |x|/lc ), the fourth
order crack-density functional reads
1
γ= (1 − s)2 + 2lc2 (∇s)2 + lc4 (s)2 . (31)
4lc
Note that an approximation (29) with the first term only describes a sharp transition and would
result in the typical difficulties of moving discontinuities. The gradient term (∇s)2 regularizes
the crack zone and renders the method non-local. The additional Laplacian in (31) affects the
curvature of the diffuse interface approximation and smoothes the transition. We would like
to emphasize that gradient terms are known from continuum damage mechanics. However,
in opposite to a damage variable here the material’s state is well defined only for phase field
parameter s = 1 (intact) and s = 0 (broken). The transition zone is a consequence of the
regularized model and an intermediate value 0 < s < 1 state has no physical meaning.
where Y e summarizes the normalized crack driving force and lc δs γ represents a kinematic
fracture resistance. It evolves for the second order crack-density approximation (30) and for
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
62 K. Weinberg et al.: Phase field approach to damage and crack growth
This function g(s) is such that in regions where the material is broken (s = 0), the contribution
to the elastic energy is zero, while in the intact regions the elastic energy contribution recovers
the one prescribed by the material’s energy density. Preferably it is stationary at both limits.
This leads to a weight function of the general form
Often a = 3 is set, i.e. g(s) = s3 (4 − 3s), and together with a double well term in the dimen-
sionless free energy (see red curve in Fig. 1), this approach has been used in the early works
of phase field fracturing, e.g. by Karma et al. [16] and Henry et al. [8,12]. Its major drawback
is the third order term in the variation (32) which impedes finite element approximations.
In order to simplify the numerical solution, a quadratic weight can be employed which,
however, cannot result in stationarity at both limits, s = 0 and s = 1. Typically, instead of
(35) with a = 1 the function g(s) = s2 is chosen, see blue curve in Fig. 1. The corresponding
variational functional of brittle fracture was first introduced by Francfort, Marigo [11] and
Bourdin [6] and with slight modifications this ansatz has become very popular since, see, e.g.,
Kuhn & Müller [17], Miehe et al. [19] and Abdollahi & Arias [1], Borden et al. [5]. Here the
local energy function is of the form
1 lc
Y = δs Ψ̄ with Ψ̄ = s2 Ψ̄e0 + (1 − s)2 + |∇s|2 , (36)
2lc 2
which corresponds to (32) with a second order crack-density approximation (30). Please note
that for a finite deformation approach to fracture the elastic energy density always needs to be
consistently coupled via Ψe (λia ) as outlined above, [14].
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 63
where, by taking the positiv part of · only, we additionally account for irreversibility of
cracking, i.e., we do not allow the crack to close.
In a more general way we may define a Mohr-Coulomb failure criterion that describes a
maximum principal stress criterion with an isotropic critical failure stress. Since such a defi-
nition allows an easy distinction between tensile and compression regions, the decomposition
of the deformation gradient (15) is not necessary here, cf. [4].
∇s · n = 0 on ∂B0 (39)
for the second order approach (30). For the fourth order ansatz (31) the boundary conditions
are
The initial conditions are given with s(X, 0) = 1 on B0 , describing an unbroken state.
3 Numerical approximation
The set of equations (11), (18), (24) and (38) forms -with corresponding boundary and initial
conditions- the coupled phase field boundary value problem. For finite element approximation
we now deduce the weak forms by taking the variations with respect to x, T and s.
For the mechanical problem (11-12) we obtain
ρ0 v̇ · δx dV + S : F T δF dV = B̄ · δx dV + T̄ · δx dA, ∀δx ∈ V x . (41)
B0 B0 B0 ∂B0σ
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
64 K. Weinberg et al.: Phase field approach to damage and crack growth
where the functional space of admissible test functions of the second-order phase field is
V1s = {δs ∈ H1 (B0 )|δs = 0 on Γ(t)}. For a fourth order approximation (31) the weak form
reads
1 l4
δs ∂s Ψ̄ − (1 − s) −lc2 ∇(δs)·∇(s)− c (δs)(s) dV = 0 ∀δs ∈ V1s , (45)
2 2
B0
where the functional space is V2s = {δs ∈ H2 (B0 )|δs = 0, ∇δs · n = 0 on Γ(t)}.
⊗
0 1 2 3 0 1 2 3
Clearly the spaces V0s and V2s require at least C 1 -continuous approximation functions,
whereas for V x , V T and V1s the C 0 -continuity of classical finite element basis functions is
sufficient. In order to meet these continuity requirements within one finite element frame-
work we employ NURBS (Non-uniform rational B-Splines) as finite element basis, Fig. 3.
A multivariate B-spline basis of degree p = [p1 , . . . , pd ] and dimension d ∈ N is defined
by the tensor product Θ1 ⊗ ... ⊗ Θd of knot vectors, built by a sequence of knots Θl =
[ξ1l ≤ ξ2l ≤ . . . ≤ ξnl l +pl +1 ], l ∈ {1, . . . , d}. In the absence of repeated knots, the partition
[ξi11 , ξi11 +1 ] × . . . × [ξidd , ξidd +1 ] forms an element of the mesh in the parametric domain. A
single multivariate B-spline B A , A ∈ [1, . . . , n], n := n1 ...nd , is then defined by
d
i (ξ) = B i (ξ 1 , ..., ξ d ) =
B A = Bp Nil ,pl (ξ l ),
p (46)
l=1
with multi-index i = [i1 , . . . , id ] and supp(B A ) = [ξi11 , ξi11 +p1 +1 ] × . . . × [ξidd , ξidd +pd +1 ].
It provides the necessary support for the required continuity. The recursive definition of a
univariate B-spline is given as follows
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 65
Linear independence as a fundamental property of finite element basis as well as local sup-
port are provided by a NURBS basis. Smoothness is related to knot multiplicity, i.e., the num-
ber of repetitions in Θ at node i. Unfortunately, the tensor product structure in (46) impedes
standard local refinement strategies which motivated us to introduce a specific hierarchical
refinement strategy in [13].
For the time integration we divide the considered time interval [0, t̄] into nt pairwise dis-
joint equidistant subintervals In = [tn , tn+1 ] with time step Δt := tn+1 − tn and employ an
implicit Crank-Nicolson scheme, known to be second-order accurate.
4 Numerical simulations
4.1 Crack growth in different fracture modes
In order to illustrate the versatility of the phase field approach for the computation of crack
growth and damage we begin with sample simulations of phase field fracture. Here, the three-
dimensional blocks are loaded in tension and shear. The size of the blocks is 1 × 0.2 × 1 mm;
the finite element mesh consists of 20 × 4 × 20 hexahedral elements with p = 2 before
further refinement. The material is assumed to be uniform, non-linear elastic and temperature
independent. Its energy density (2) is of Neo-Hookean type,
λ μ
Ψ= (ln(J))2 − μ ln(J) + (F : F − 3) (50)
2 2
with Lamé coefficients typical for steel, μ = 80769 N/mm2 and λ = 121154 N/mm2 . The
critical energy release rate is chosen to be Gc = 2.7 N/mm.
At first we consider a symmetric tension test with the specific boundary conditions shown
in Fig. 4. The load is applied as prescribed displacement at the upper boundary using constant
displacement increments of u = 5 · 10−5 mm. The mesh is a priori refined locally around
the expected crack path and consists now of 46288 elements which corresponds to a total of
305856 degrees of freedom for both, the mechanical and the phase field. In the refined domain
the elements have a size of hmin = 0.0063 mm which allows to choose a small critical length
parameter, lc = 0.01575 mm. The growing crack is computed with a fourth-order phase
field model (31). In the diagram of Fig. 4 the load-displacement curves are shown, now for
plane-strain simulations of a second-order and a fourth-order phase field model, cf. [30]. All
parameters, i.e., Lamé coefficients, Griffith’s energy release rate, critical length lc , applied
displacement steps as well as the mesh using quadratic NURBS shape functions are exactly
identical for both calculations. Obviously, a higher continuity in the phase field leads in the
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
66 K. Weinberg et al.: Phase field approach to damage and crack growth
8
4th order PF
2nd order PF
6
load [kN]
4
critical range to a more compliant response. This has been confirmed by further tests. The
effect, however, is not necessarily contributed to the order of the phase field only. Additionally,
the choice of critical length parameter lc and the compliance of the discretization (quadratic
NURBS vs. linear finite elements) play a role.
At next we simulate an in-plane shear test with boundary conditions shown in Fig. 5. Anal-
ogous to the tension test the load is applied in constant displacement increments, here with
u = 1 · 10−3 mm. The refined mesh is composed of 21396 elements that leads to 112192
degrees of freedom in this case. We get an element size of hmin = 0.0125 mm and choose the
length-scale parameter as lc = 0.03125 mm. The snapshot of the phase field is presented in
Fig. 5 (left).
Very similar, with a mesh of 21480 elements, a total of 113632 degrees of freedom, and
same element size and length-scale parameter lc we performed the out-of-plane shear test
Fig. 5 (online colour at: www.gamm-mitteilungen.org) Boundary conditions and computed phase field
for mode II and mode III fracture
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 67
Fig. 6 (online colour at: www.gamm-mitteilungen.org) Stress distributions in the current configuration
for the three fracture modes I,II and III
Solder joints, made of fusible metal alloys with a low melting point, are used to connect metal-
lic surfaces. They cover a broad range of technical applications. In particular in electronics
soldering plays an important role, where solder balls provide mechanical as well as electri-
cal connections between different components and, consequently, the reliability of the joints
determines the life expectation of the whole electronic device.
Here we study eutectic Sn-Pb solder balls subjected to an inhomogeneous temperature
field, see Fig. 7. When such initially homogeneous binary alloy is annealed, its microstructure
becomes inhomogeneous, e.g. [7,15]. Under a temperature gradient heat conduction, spinodal
decomposition and mass diffusion couple and cause an irreversible thermomigration.
In the corresponding system of equations (18-25) we set the mole fraction of tin, as our
phase field variable s := cSn = 1 − cPb but omit the subscript Sn. In order to obtain a tem-
perature and mole fraction dependent free energy function, we constructed energy functions
for a discrete set of temperatures in a first step. Then we used the least square method and the
pure metal’s free enthalpy coefficients g̃ik to determine function
Fk
gk (T ) = Ak + Bk T + Ck T ln(T ) + Dk T 2 + Ek T 3 + , k = 1, ..., 4,
T
3
g5 (T ) = A5 + B5 T in J/m .
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
68 K. Weinberg et al.: Phase field approach to damage and crack growth
Fig. 7 (online colour at: www.gamm-mitteilungen.org) Solder bump of Sn-Pb with an eutectic mole
fraction of 0.63 ; distribution of the initial mole fraction field (left), and initial temperature profile (right).
As a result the system’s configurational free energy density can be described by a smooth
function in temperature and mole fraction,
Ψcon (c, T ) = g1 c+g2 (1−c)+g3 c ln(c)+g4 (1−c) ln(1−c)+g5 c(1−c), in J/m3 . (51)
The fitted values are listed in table 1, see [2] for details of the method.
Table 1 Fit values for the coefficients gi , i = 1, ..., 5, of the free energy function Ψcon . The discrete set
of temperatures is Tk = 250 + hk ∈ (250 K, 505 K), h := 85/8334, k = 1, ...25000.
MJ
Ai m 3 Bi mMJ
3K Ci mMJ3K Di m3JK2 Ei m3JK3 Fi GJKm3
i=1 −355.77 3.9628 −0.96695 −1146 0.1898 −3.7692
i=2 −405.42 5.5101 −1.3286 −191.5847 −0.0153 0.027891
i=3 275.05 1.338 −0.18453 389.8139 −0.1331 1.9559
i=4 −85.964 1.3153 −0.13691 244.7001 −0.0888 1.274
i=5 1224.9 −0.49161 - - - -
For realistic values of the chemical mobility Mc we adapted a work of Ubachs et al. [28]
and determined the mobility for a Sn-Pb alloy by comparisons between a two-dimensional
simulation and an ageing experiment. Since they derived the mobility only at a homogeneous
temperature of 423.12 K and, to the best knowledge of the authors, comparable results at
different temperatures are not available in the literature, we follow Dreyer & Müller [10] in
order to relate the mobility to the diffusion coefficients. To this end they considered diffusion
of the Fickean type and stated the relations
−1
∂ 2 Ψcon cβ/α (T ), T
∗
MSn/Pb (T ) = γ DSn/Pb (T ) in m5 /(Js), (52)
∂c2
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 69
∗ ∗
where cα/β are the equilibrium mole fractions and MSn /DSn , MPb /DPb are the mobility/diffusivity
of lead in tin and tin in lead, respectively, given by an Arrhenius relationship
94400 61370
DSn (T ) = 4.1 · 10 −5
exp − , DPb (T ) = 3.533 · 10 −6
exp − in m2 /s,
RT RT
(53)
with the universal gas constant R = 8.31451 Jmol−1K−1 . The parameter γ will be used later
∗
on to fit MSn/Pb such that the magnitude of MC at T = 423.15 K corresponds to the mobility
employed in [28]; for now we set γ = 1. In order to model the equilibrium mole fractions
as functions in temperature, we now apply the common tangent construction to the configu-
rational free energy density Ψcon (c, Tk ) at temperatures Tk = 250 + hk ∈ (250 K, 505 K),
h := 85/8334, k = 1, ...25000 and obtain a corresponding set cα/β (Tk ) of equilibrium mole
fractions and subsequently polynomial functions
In this context, Vm is the molar volume, Mi the atomic mobility and Q∗i the molar heat
∗
of transport of element i ∈ {A, B}. Comparison of the units yields the relation MSn/Pb =
∗ −1
Vm MSn/Pb ⇔ MSn/Pb = MSn/Pb Vm and consequently the temperature and mole fraction
dependent mobilities for the Sn-Pb alloy.
∗
Mc (c, T ) = c(1 − c) [cMPb ∗
(T ) + (1 − c)MSn (T )] in m5 /(Js), (58)
∗ 2
Mq (c, T ) = c(1 − c) [MPb (T )Q∗Pb − ∗
MSn (T )Q∗Sn ] T −1 Vm−1 in m /(sK) (59)
The remaining material parameter for a Sn-Pb alloy are listed in table 2.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
70 K. Weinberg et al.: Phase field approach to damage and crack growth
Fig. 8 (online colour at: www.gamm-mitteilungen.org) Microstructural evolution of the Sn-Pb solder
bump subjected to a temperature gradient.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 71
For numerical simulation a Sn-Pb solder bump with a diameter of 0.5 μm and a height
of 0.2692 μm is subjected to a temperature gradient with Tmin = 423.1212 K and Tmax =
423.1788 K. The dimensions and the temperature difference between the lower and the upper
surface of the bump are scaled such that they correspond to the data from the experiments
performed by Hsiao & Chen [15]. Fig. 8.
Referring to physical inhomogeneities in alloys the initial setting for the simulation is ar-
ranged in such a manner that the alloy has a constant mass fraction of c0 = 0.63 with randomly
generated perturbations as shown in Fig. 7. Here and in Fig. 8 the reddish areas denote the
tin-rich phase whereas the blue domains represent the lead-rich phase. Similar to the experi-
mental observations, the typical spinodal decomposition of phases occurring in the isothermal
case, [2], does not take place here. Instead, a migration of the lead rich phase to the cold end
of the bump and one of the tin rich phase to the hot end of the bump can be observed. The
solder degradation is illustrated for different aging times in Fig. 8.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
72 K. Weinberg et al.: Phase field approach to damage and crack growth
minima for c = 0, s = 1 (no vacancies, no void damage) and c = 1, s = 0 (empty void full
of vacancies). The states at c = 1, s = 1 and c = 0, s = 0 which describe physically unlikely
states have a high energy potential.
Ψcon α = const.
α = α(c)
s 1
1
0 c Δt
Fig. 9 (online colour at: www.gamm-mitteilungen.org) Void growth in a homogeneous continuum:
Landau free-energy potential for a combined phase field model (left), two-dimensional simulations of
voids growing by vacancy concentration within 200 seconds (middle), and for different surface coeffi-
cients αc in potential (61) void volume vs. time (right).
First results of void growth are also shown in Fig. 9. Because we do not consider vacancy
generation at this point, the voids grow slowly. In this computation we use a mesh of 128×128
NURBS based finite elements over a domain of 128 nm2 and an initial concentration of 10−6 .
In our computation we assume the penalty parameter αc to be constant. This, however, is not
realistic because it favors unlimited void growth - until all vacancies are absorbed. In practise
we typically have an equilibrium concentration of vacancies in every material. Therefore, we
express the coefficient αc as a positive function of concentration c with an ansatz
The exponential form allows a simple derivation ∂c α = α ∂c k. For the equilibrium state
the chemical potential assumed to be constant. Then, using a Legendre transformation of the
(squared) concentration gradient (∂c/∂x)2 = c2 , and the consequence, that the square of the
concentration gradient divided by its second derivative is proportional to the concentration,
we obtain an equation for the exponent in (62).
∂c Ψcon
k(c) = ln (63)
A(2c − c2 ∂c k)
After a lengthy calculation the simplified solution leads to an approximation αc = 1 − c
which, in turn, renders the variation of the surface energy term in (61) non-linear. The effect
of a concentration dependent αc is visualized in the diagram of Fig. 9. We observe a saturation
of growth at a specific vacancy concentration here.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 73
Im
material, fracture occurs. We used this experimental setup in order to determine the fracture
properties of an Ultra-High Performance Concrete (UHPC), see Fig. 10.
For finite element analysis we work with an axisymmetric model of the spallation exper-
iment. The mechanical problem is described with the balance of momentum (11) in domain
B0 = [0, rs ] × [0, l], where rs = 10 mm and l = 200 mm denote radius and length of the
UHPC specimen. Stress boundary conditions are applied on B0σ = {(r, z) ∈ B0 : z = l},
i.e. at the left side of the cylindrical specimen. The incoming stress pulse has a rectangular
shape with maximum σmax . Its length and value are adapted to the experimentally measured
data. For the specimen we use the material parameters of UHPC, E = 59 GPa, ν = 0.2,
ρ = 2370 kg/m3 .
The specimen has the typical properties of a brittle ceramic and thus, we employ a small
strain approach. The linear-elastic material is presumed to follow Hooke’s law with elastic
free energy density
1
Ψ (u) = ε(u) : C : ε(u), (64)
2
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
74 K. Weinberg et al.: Phase field approach to damage and crack growth
where strain ε(u(x)) denotes the symmetric gradient of displacement u(x) and C is the
Hookean tensor. Also, the split into tensile and compressive components (14) is linearized,
ε = [ε+ ] + [ε− ], where [ε+ ] denote the positive and tensile components of the principal
strains, and [ε− ] the remaining compressive parts. This leads to an elastic energy density
function which only accounts for tension:
1
Ψe (u) = [ε+ (u)] : C∗ : [ε+ (u)] (65)
2
Here the tensor C∗ = g(s)C is multiplied with a weight function (34), specifically we set
g(s) = s2 + η. The choice of η 1 but η = 0 avoids numerical problems and results in a
substitute-material approach for the damaged zones. In dynamical computations η can be set
to zero.
The resolution of the diffuse crack zone is determined by the critical length lc which de-
pends on the finite element mesh size. We used here a uniform mesh with 5760 triangular
elements and lc = 2.5 mm. A finer mesh would lead to improved results in the sense of a
narrower crack zone, see Fig. 4. However, also with such a rather coarse mesh a quantita-
tive agreement to cohesive element simulations was obtained, cf. [9]. The mobility parameter
in equation (38) is set to M = 106 /s. It needs to be chosen in such a way that the phase
field is able to decrease locally to zero within a time range of few microseconds, once the
energy-release rate exceeds the specific fracture energy.
We performed our numerical simulations with the aim to validate experimental measure-
ments of Griffith’s critical energy release rate Gc in a dynamic regime. Out of the experimental
data Gc is calculated from an energy balance,
1 1 1 1
Gcexp = mall (v(t1 ))2 − m1 (v1 (t2 ))2 − m2 (v2 (t2 ))2 , (66)
Ac 2 2 2
as the energy decreases between t1 immediately before crack initiation and t2 immediately
after fragmentation of the specimen; v(t1 ) denotes the specimen’s velocity, v1 (t2 ), v2 (t2 ) the
velocities of the fragments and mall , m1 , m2 the corresponding masses. The specific fracture
energy then follows by division with the fractured area Ac .
This value is now compared to the input value Gc of the phase-field model. In Fig. 11 we
illustrate qualitatively in the crack formation for different values of Gc which -by definition-
Gc [N/m]
200
180
160
140
120
100
80
60
40
20
0
0 20 40 60 80 100 120 140 160 180 200
input-Gc [N/m]
Fig. 12 Comparison between the input Gc -value of the phase-field model and the Gc -value calculated
by means of (66) with the variables v, v1 , v2 , m1 , m2 observed in the finite element simulation.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 75
m
v s
8
0
0 0.2 0.4 0.6 0.8 1
Fig. 13 (online colour at: www.gamm-mitteilungen.org) Computed velocity of the specimen be-
fore crack initiation and velocity of the two fragments after crack initiation with Gc = 90N/m,
lc = 20μm, M = 106 /s, σmax = 17.3MPa. The experimental velocity-time diagram of the spalla-
tion experiment is shown on the right. The red dots represent the velocity vs,1 of the first fragment
(average of 31 control points), the blue squares represent the velocity vs,2 of the second fragment (aver-
age of 8 control points) and the green diamonds represent the extrapolated initial center of mass velocity
vi .
have a significant effect on the crack evolution. If Gc is set to a high value, Griffith’s criterion
cannot be fulfilled anymore and the phase field parameter s remains at the constant value 1
during the whole simulation. A too small value leads to ’distributed’ crack, i.e., a wide range
of s < 1. Also, Gc influences the crack position of the damaged zone, a reduction of Gc moves
the crack position closer to the free end of the specimen because the driving force ∂Ψe /∂s
exceeds the fracture resistance at earlier time.
Additionally the specific fracture energy of UHPC is determined in the sense of an inverse
analysis. In Fig. 12 the values of Gc , with eq. (66) inversely calculated out of the fragments
velocities, are plotted over the input values of second-order phase field fracture calculations.
They show a good agreement, all points are located near the bisector that declares the ’exact’
values.
Finally, Fig. 13 illustrates the velocity of the specimen and of the two fragments. We
assume an initial velocity of v0 = 5 m/s for the specimen due to the acceleration of the incident
bar and a specific fracture energy of Gc = 90 N/m. A permanent crack is defined if the phase
field parameter s is below a value of 0.3 over the whole width of the specimen. As can be
seen from Fig. 13, the phase field fracture simulation agrees well with the fragments’ velocity
measured in the experiments. Immediately after the cracking this difference is approximately
1 m/s.
5 Summary
In our paper we address the problem of computing material failure within a finite element
framework. We chose phase fields of different type and approximation order to resolve
and regularize the discontinuities typically arising between intact and damaged material. By
means of a phase field ansatz such sharp boundaries are replaced by diffuse interfaces which
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
76 K. Weinberg et al.: Phase field approach to damage and crack growth
avoid a difficult explicit tracking of the boundaries. Here we discuss the theoretical frame-
work of phase diffusion and phase field fracture methods and suggest a numerical solution of
the resulting multi-field problems with spline-based finite elements. These NURBS functions
are an efficient tool to provide the continuity required for different phase field approaches.
Computations of several multi-field problems such as quasi-static crack propagation in a hy-
perelastic material, thermomigration and void growth in solder alloys, and crack initiation by
wave reflection have been modeled and solved numerically.
Acknowledgment
The authors gratefully acknowledge the support of the Deutsche Forschungsgemeinschaft
(DFG) under the grant WE2525/4-1, WE2525/5-1, HE4593/3-2 and WE2525/8-1.
References
[1] A. Abdollahi and I. Arias. Phase-field modeling of crack propagation in piezoelectric and ferro-
electric materials with different electromechanical crack conditions. Journal of the Mechanics and
Physics of Solids, 60(12):2100 – 2126, 2012.
[2] D. Anders, C. Hesch and K. Weinberg. Computational modeling of phase separation and coarsen-
ing in solder alloys. Int. J. Solids Structures 49, 1557–1572, 2012
[3] D. Anders and K. Weinberg. Thermophoresis in binary blends. Mechanics of Materials, 47:33–50,
2012.
[4] C. Bilgen, T. Reppel and K. Weinberg. Determining fracture properties of polyurea by an inverse
analysis using phase field fracture. subm. to Experimental Mechanics, 2016.
[5] M.J. Borden, C.V. Verhoosel, M.A. Scott, T.J.R. Hughes, and C.M. Landis. A phase-field descrip-
tion of dynamic brittle fracture. Comput. Methods Appl. Mech. Engrg., 217–220:77–95, 2012.
[6] B. Bourdin. The variational formulation of brittle fracture: Numerical implementation and exten-
sions., Volume 5 of IUTAM Symposium on Discretization Methods for Evolving Discontinuities,
IUTAM Bookseries, Chapter 22, pages 381–393. Springer Netherlands, 2007.
[7] C. Chen, H.-Y. Hsiao, Y.-W. Chang, F. Ouyang, K.N. Tu. Thermomigration in solder joints.
Materials Science and Engineering R 73, 85–100, 2012
[8] F. Corson, M. Adda-Bedia, H. Henry, and E. Katzav. Thermal fracture as a framework for quasi-
static propagation. International Journal of Fracture, 158:1–14, 2009.
[9] T. Dally and K. Weinberg. The phase-field approach as a tool for experimental validations in
fracture mechanics. Continuum Mech. Thermodyn., 10.1007/s00161-015-0443-4, 2015
[10] W. Dreyer and W.H. Müller. Modeling diffusional coarsening in eutectic tin/lead solders: a quan-
titative approach. Int. J. Solids Structures 38, 1433–1458, 2001
[11] G.A. Francfort and J.-J. Marigo. Revisiting brittle fracture as an energy minimization problem.
Journal of the Mechanics and Physics of Solids, 46:1319–1342, 1998.
[12] H. Henry and H. Levine. Dynamic instabilities of fracture under biaxial strain using a phase field
model. Physics Review Letters, 93:105505, 2004.
[13] C. Hesch, S. Schuß, M. Dittmann, M. Franke and K. Weinberg. Isogeometric analysis and hi-
erarchical refinement for higher-order phase-field models. Compt. Methods Appl. Mech. Engrg.,
2016
[14] C. Hesch and K. Weinberg. Thermodynamically consistent algorithms for a finite-deformation
phase-field approach to fracture. Int. J. Numer. Meth. Engng. 99, 906–924, 2014
[15] H.-Y. Hsiao and C. Chen. Thermomigration in flip-chip SnPb solder joints under alternating
current stressing. Applied Physics Letters 90, 152105, 2007
[16] A. Karma, D.A. Kessler, and H. Levine. Phase-field model of mode iii dynamic fracture. Physical
Review Letter, 81:045501, 2001.
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
GAMM-Mitt. 39, No. 1 (2016) 77
[17] C. Kuhn and R. Müller. A continuum phase field model for fracture. Engineering Fracture Me-
chanics, 77:3625–3634, 2010.
[18] S. Mariani and U. Perego. Extended finite element method for quasi-brittle fracture. Int. J. Numer.
Meth. Engng. 58, 103–126, 2003
[19] C. Miehe, M. Hofacker, and F. Welschinger. A phase field model for rate-independent crack
propagation: Robust algorithmic implementation based on operator splits. Comput. Methods Appl.
Mech. Engrg., 199:2765–2778, 2010.
[20] M. Ortiz and L. Stainier. The variational formulation of viscoplastic constitutive updates. Com-
puter Methods in Applied Mechanics and Engineering, 171(3-4):419–444, 1999.
[21] M. Ortiz and A. Pandolfi. Finite-deformation irreversible cohesive elements for three-dimensional
crack-propagation analysis. Int. J. Numer. Meth. Engng. 44, 1267–1282, 1999
[22] F.Y. Ouyang, K.N. Tu, Y.S. Lai and M. Gusak. Effect of entropy production on microstructure
change in eutectic snpb flip chip solder joints by thermomigration Applied Physics Letters 89,
2006
[23] A. Pandolfi and M. Ortiz. An eigenerosion approach to brittle fracture. Int. J. Numer. Meth.
Engng., 92:694–714, 2012.
[24] K.L. Roe and T. Siegmund. An irreversible cohesive zone model for interface fatigue crack growth
simulation. Eng. Fract. Mech. 70, 209–232, 2003
[25] S. Rokkam, Anter El-Azab, P. Millett and D. Wolf. Phase field modeling of void nucleation and
growth in irradiated metals. Modelling and Simulation in Materials Science and Engineering, 17
064002, 2009.
[26] B. Schmidt, F. Fraternali and M. Ortiz. Eigenfracture: An eigendeformation approach to varia-
tional fracture. Multisclae Model. Simul. 7,1237–1266, 2009
[27] N. Sukumar, D.J. Srolovitz, T.J. Baker and J.-H. Prevost. Brittle fracture in polycrystalline mi-
crostructures with the extended finite element method. Int. J. Numer. Meth. Engng. 56,2015–2037,
2003
[28] R.L.J.M Ubachs, P.J.G Scheurs, and M.G.D Geers. A nonlocal diffuse interface model for mi-
crostructural evolution of tin-lead solder. J. Mech. Phys. Solids 52, 1763–1792, 2004
[29] C.V. Verhoosel and R. de Borst. A phase-field model for cohesive fracture. Int. J. Numer. Methods
Eng., page in press, 2013.
[30] K. Weinberg and C. Hesch. A high-order finite-deformation phase-field approach to fracture.
Continuum Mech. Thermodyn., DOI 10.1007/s00161-015-0440-7, 2015
[31] K. Weinberg, T. Böhme, and W. H. Müller. Kirkendall voids in the intermetallic layers of solder
joints in MEMS. Computational Materials Science, 45(3):827–831, 2009.
[32] X.P. Xu and A. Needleman. Numerical simulations of fast crack growth in brittle solids. J. Mech.
Phys. Solids 42,1397–1434, 1994
www.gamm-mitteilungen.org
c 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim