Corrosion CM

You might also like

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

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272590809

Numerical modeling of corrosion pit propagation using the combined


extended finite element and level set method

Article  in  Computational Mechanics · September 2014


DOI: 10.1007/s00466-014-1010-8

CITATIONS READS

76 1,052

1 author:

Ravindra Duddu
Vanderbilt University
58 PUBLICATIONS   1,180 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Phase field fracture modelling of crevasses in grounded glaciers and floating ice shelves View project

Exponential integrator based methods and applications View project

All content following this page was uploaded by Ravindra Duddu on 26 February 2015.

The user has requested enhancement of the downloaded file.


Comput Mech (2014) 54:613–627
DOI 10.1007/s00466-014-1010-8

ORIGINAL PAPER

Numerical modeling of corrosion pit propagation using the


combined extended finite element and level set method
Ravindra Duddu

Received: 7 November 2013 / Accepted: 19 March 2014 / Published online: 19 April 2014
© Springer-Verlag Berlin Heidelberg 2014

Abstract A sharp-interface Eulerian formulation for mation of pits on the metal surface. A major problem with
modeling the propagation of localized pitting corrosion is this localized form of corrosion is that it is difficult to detect
presented. This formulation allows for an accurate repre- and often leads to sudden and catastrophic failures by induc-
sentation of the corrosion front independent of the underly- ing fatigue cracks or stress corrosion cracks near the pits
ing finite element mesh and handles complex morphological [1]. Thus, pitting corrosion poses a formidable threat to the
transitions such as pit merging without requiring remeshing durability and integrity of industrial, civil infrastructure and
or mesh-moving procedures. First, the governing equations military systems [2]. Computational modeling of corrosion,
of the moving interface problem associated with pit growth together with accelerated laboratory experiments, can yield
are derived for the two-phase (metal-solution) system. Next, quantitative tools for scientists and engineers to better under-
the implementation of the combined extended finite element stand and predict corrosion, and develop efficient strategies
and level set method and the procedure to enforce interface for mitigating corrosion. Therefore, corrosion modeling has
conditions are discussed. Finally, the method is validated by been an active area of research, and for a detailed literature
conducting several benchmark numerical studies, and com- review of the various empirical, stochastic and deterministic
paring the results with published experimental data and exist- models the reader is referred to [2]. The purpose and scope of
ing numerical studies. Simulation studies indicate that: dur- this article is to present a novel extended finite element and
ing diffusion-controlled corrosion an isolated pit grows as level set formulation for modeling pitting corrosion propa-
a semi-circular shape, whereas closely spaced pits merge gation based on continuum mechanics and electrochemical
and grow into an elongated elliptical shape; and only during theory.
activation-controlled corrosion the initial pit morphology is Pitting corrosion, or simply pitting, progresses primar-
preserved. ily by the anodic dissolution of metal atoms as metal ions
(loss of electrons or oxidation) into the pit solution (elec-
Keywords Pitting corrosion · XFEM · Level sets · trolyte) [3]. Isaacs [4] distinguished three basic stages of
Moving interface · Interface constraints pitting corrosion of stainless steels: nucleation, metastable
growth and stable pit growth. In this article, only the stable
growth (or propagation) stage is studied, when the pit inter-
1 Introduction face (separating the solid metal phase from the liquid pit
solution phase) acts as a sharp moving electrode boundary
Alloys of steel, nickel and aluminum are commonly used as across which there are jumps or discontinuities in both the
structural materials and are susceptible to pitting corrosion, atom/ion concentrations and their normal derivatives. Thus,
a form of extremely localized corrosion that leads to the for- the pit propagation problem belongs to the class of moving
free boundary problems. It is well known from experimental
R. Duddu (B) studies that the rate of pit propagation is not only dependent
Department of Civil and Environmental Engineering,
on the compositional variables (e.g. alloy, inter-metallic par-
Vanderbilt University, 400 24th Avenue South, 274 Jacobs Hall,
Nashville, TN 37212, USA ticles composition) but also dependent on the environmental
e-mail: rduddu@gmail.com; ravindra.duddu@vanderbilt.edu variables (e.g. pH, concentration of ionic and neutral species

123
614 Comput Mech (2014) 54:613–627

in the pit solution) [5]. The need for tracking the concen- using published experimental data on stainless steel alloys.
trations of multiple ionic (anions and cations) and neutral In addition, an improved numerical implementation of the
species in the pit solution using electro-diffusion equations, XFE-LSM is presented. Unlike the original implementation
arising from either Nernst-Planck or Stefan-Maxwell theo- described in [41], herein Heaviside functions are employed
ries [6], tremendously increases the computational burden. to enrich the finite element basis and Lagrange multipliers
Additionally, corrosion pits can evolve into different com- are used to enforce the Dirichlet interface condition on the
plex morphologies over time that can range from narrow- evolving interface during diffusion-controlled corrosion.The
deep to wide-shallow shapes owing to the different pit sur- generalized Heaviside enrichment or step enrichment is cho-
face and environments and local microstructure features, as sen as it allows for the jumps in both the concentration field
identified by ASTM [7]. Thus, the challenges associated with and its normal derivative at the pit interface. The rest of the
numerical simulation of pitting corrosion are significant due article is organized as follows: in Sect. 2 the strong form of
to the presence of a sharp moving interface, discontinuous the elliptic interface problem governing metal dissolution is
concentrations and gradients across the interface, complex derived using the continuum theory of two-phase interfaces
pit geometries, and heterogeneous microstructure. Conse- described by Gurtin et al. [46,47], and the corresponding
quently, specialized methods for handling moving bound- weak or variational formulation is presented; in Sect. 3 the
aries and for solving elliptic/parabolic partial differential XFE-LSM discretization and strategies for enforcing interfa-
equations (PDEs) on irregularly shaped domains are needed cial constraints and for evaluating the interface propagation
for producing accurate simulations. speed are detailed; in Sect. 4 numerical results are presented
Early efforts on numerical modeling of localized corro- that demonstrate the viability of this approach for studying
sion focussed on one-dimensional pit growth using either localized pitting corrosion in two-dimensions; and finally,
the finite difference method [8,9] or the finite element Sect. 5 presents some concluding remarks.
method [10]. Recent modeling studies simulated pit growth
in higher dimensions (mostly in 2D) using either the finite
volume method [11,12] or the finite element method [13– 2 Model formulation
16]. Scheiner and Hellmich [11,12] simulated pit propaga-
tion in two-dimensions using the finite volume method, but 2.1 Governing equations
the pit interface was not consistently tracked; consequently,
the reported pit shapes were zigzagged. The finite element 2.1.1 Domain description
method (FEM) was employed to study pitting in higher
dimensions; however, simulating pit propagation using the Let us consider a regular domain  ∈ R3 bounded by a
FEM can be cumbersome as it relies either on remeshing smooth boundary ∂ =  that consists of two phases:
[16], moving mesh techniques [17], or arbitrary Lagrangian- the solid metal domain s (t) and the pit solution domain
Eulerian techniques (ALE) [14]. To overcome the limitations p (t). The solid and the solution domains are separated by
of the FEM, herein an efficient and accurate method based on a sharp moving interface  int (t) at all times t, as shown
the extended finite element method (XFEM) [18,19] and the in Fig. 1. Herein, the general 3D problem is reduced to
level set method (LSM) [20–24] is presented. In the literature, 2D by considering uniformity in the out of plane direc-
many investigations employed the XFEM to study a variety tion. The two phases are assumed exist in closed regions
of embedded interface problems related to: crack propaga- with their union  = s (t) ∪ p (t) and their intersection
tion [25,26], phase transformation [27–29], fluid structure  int (t) = s (t) ∩ p (t) is a smoothly propagating inter-
interaction [30], hydrogels [31–33], biological cells [34–37], face. The interface orientation is defined by the outward unit
finite deformations with arbitrary inclusions or voids [38,39], normal n(x, t) pointing away from the pit solution domain
and nano-composites [40]. The above cited work only con- p (t), where x denotes the spatial Cartesian system of coor-
sidered the level set representation of the embedded inter- dinates. Let V (x, t) denote the normal interface speed of
face, but did not employ the full LSM along with the fast  int (t) in the direction n at any spatial location x and time
marching scheme to consistently simulate its time evolution. t. The Dirichlet boundary at pit mouth and the Neumann
With the goal of consistently simulating complex interface boundary enclosing the metal solid domain are denoted by
p p
morphology evolution and shape transitions, the author and hs (t) and d (t), respectively. Note that  = hs ∪ d at all
co-workers developed the combined extended finite element times.
and level set method (XFE-LSM) to study microbial biofilm
growth [41,42] and precipitate evolution in Ni-Al alloys [43– 2.1.2 Continuum mechanics of phase interfaces
45].
In this article, a model for pitting corrosion propagation is Following the two-phase continuum theory and the notation
proposed within the XFE-LSM framework and is validated introduced by Gurtin et al. [46,47], let us derive the govern-

123
Comput Mech (2014) 54:613–627 615

Fig. 1 A schematic diagram of the pitting corrosion problem. The the pit solution domain. The outward normal to the domain boundary
interface  int (blue) partitions the sub-domains s (light grey) and p is denoted by n0 . A control volume R ∈  with outward unit normal
p
(white). The Dirichlet boundary d (red) and the Neumann boundary n R on ∂ R is considered to derive the governing equations. (Color figure
hs (black) are shown. The outward interface normal n points away from online)

 
ing equations in three-dimensions in the bulk domains and J · n R da → [[ J]] · n da. (5)
at the interfaces. Let c(x, t) denote the metal atom/ion con-
∂R r (t)
centration of a particular species in  that is function of the
spatial coordinates x and time t, and let c denote its mate-
rial time derivative. For this bulk field, let cp and cs denote 2.1.3 Mass balance
the limits of c when the interface is approached from the pit
solution and solid domains, respectively, and [[c]] denotes the The balance of metal atom/ion mass for all control volumes
jump in c across the interface: for x ∈  int (t) R is given by,
cp (x, t) = lim c( y, t), cs (x, t) = lim c( y, t), (1) ⎧ ⎫
y→x y→x
y∈p (t) y∈s (t) ⎨ ⎬ 
[[c]] = c − c .
s p
(2) Mc dv =− M J · n R da, (6)
⎩ ⎭
R ∂R
Now, consider a control volume R ∈  with n R the out-
ward unit normal on ∂ R, as shown in Fig. 1. Let R contain
a portion of the interface r such that r (t) = R ∩  int (t). An where M is the mass of one mole of metal atoms/ions. By
important transport identity [47], valid for the bulk field c is, applying the above equation to control volumes that exclude
the interface and using the divergence theorem, the standard
⎧ ⎫
⎨ ⎬   bulk relation can written as,
c dv = c dv − [[c]]V da, (3)
⎩ ⎭
R R r (t) c = −∇ · J (7)

where V is the normal interface speed, dv is an infinitesimally


small volume element in  and da is an infinitesimally small where ∇· is the divergence operator. The above relation must
area element on the moving interface. Note that the above be satisfied in each of the bulk regions s (t) and p (t) at
identity can be expressed in 2D by replacing volume integrals all times. Shrinking the control volume R (with r = ∅) to
by area integrals and area integrals by line integrals. As R the interface in (6) and using the identities in (4) and (5), the
shrinks to the interface the above Eq. (3) yields the limiting interfacial mass balance can be written as,
relation,
⎧ ⎫ [[c]]V = [[ J]] · n, (8)
⎨ ⎬ 
c dv →− [[c]]V da. (4)
⎩ ⎭
R r (t) which must be satisfied on the interface  int (t) at all times.
Note that the above equation is the Rankine-Hugoniot jump
As ∂ R approaches the interface from the solution or the condition in Eulerian description [48]. In the case of multiple
solid phase, the normal n R has the limit −n or n, respectively. species, bulk mass balance for each of the species in solution
Thus, for the bulk molar flux vector field J the following and interface mass balance for all the dissolving solid species
relation is obtained, needs to be considered.

123
616 Comput Mech (2014) 54:613–627

2.1.4 Diffusive flux Note that i 0 is dependent on the temperature T and the pH of
the pit solution (electrolyte). During the activation-controlled
The driving force for diffusion is the gradient of the chemical corrosion i A can be related to the molar flux of metal atoms
potential μ and the total atomic/ionic flux associated with c J m dissolving into the pit solution or to the normal interface
is given by, speed V according to Faraday’s laws of electrolysis as,

Dc i A = z F J m · n = z F V cs . (13)
J =− ∇μ, μ = μ0 + kB T ln (γ c) + z F E, (9)
kB T
Combining Eqs. (8), (10) and (13), the interface condition
where μ0 is the chemical potential at a reference state, kB for activation-controlled corrosion can be written as,
is the Boltzmann’s constant, T is the absolute temperature,
D is the diffusion coefficient or diffusivity, γ is the activ-
iA
ity coefficient, z is the number of electrons involved in the [[D∇c]] · n + [[c]] = 0. (14)
z Fcs
reaction, F is the Faraday constant, E is the electric poten-
tial and ∇ denotes the gradient operator. In this study, D is
assumed to vary linearly with temperature T as given by the 2.2 Modeling assumptions
Einstein’s relation. For dilute solutions (i.e. γ ≈ 1), assum-
ing zero thermal gradients (i.e. ∇T = 0) and zero electric The following assumptions were made for simplifying the
potential gradient in the domain (i.e. ∇ E = 0), except for a pit growth problem:
sharp potential drop across the pit interface  int , the diffusive
flux for the bulk field c is simply given by Fick’s first law, 1. The motion of the pit interface is due to the dissolution
and diffusion of a single independent chemical compo-
J = −D∇c, (10) nent in the pit solution [11,12]. Since only one charged
species (Fe2+ in aqueous solution) is considered, electro-
migration effects can be modeled by simply choosing an
2.1.5 Electrochemical kinetics
appropriate diffusion coefficient. Also, the aqueous pit
solution is assumed to be under stagnant conditions, so
Metal corrosion usually occurs through electrochemical reac-
mass transport is purely via diffusion. In a more gen-
tions at the pit interface,  int , which acts as an anodic site.
eral case scenario, one should track the concentrations
When an anodic activation overpotential η A is applied, gener-
of all the corrosion rate limiting species by considering
ally referred to as polarizing the metal, a current is established
the flux contribution from electric potential gradient and
and corrosion progresses due to the dissolution of the metal
advection in addition to that from concentration gradient
at the anode. The overpotential is related to applied potential
[15,16]. Since the focus here is on developing the XFE-
E according to the relation η A = E − E corr , where E corr
LSM for modeling pit propagation only Fickian diffusion
is a material parameter called the corrosion or open-circuit
of one rate-limiting species is considered.
potential. The magnitude of the anodic current density i A is
2. The solid domain is homogeneous and local microstruc-
given by the Bulter-Volmer equation, but when η A is not too
ture features such as precipitates are ignored. The con-
small, i A may be approximated by the Tafel equation [3],
centration of metal atoms is therefore assumed a constant

in the solid domain:
αzη A F
i A = i 0 exp (11)
kB T
c = csolid in s . (15)
where i 0 is the exchange current density and α is the dimen-
sionless anodic charge transfer coefficient. The exchange cur- Thus, there is no need to solve for mass balance in s
rent density i 0 is a parameter that represents the current in the and the interface concentration cs = csolid at all times.
absence of net electrolysis and at zero overpotential. Math- 3. The diffusion rate of ions in pit solution is relatively much
ematically, i 0 can be thought of as a background current to faster compared to corrosion (dissolution) rate; so it is
which the observed current i A at various overpotentials η A reasonable to assume that the concentration of ions in
is normalized. According to [12], i 0 can be related to the p at any time t is at a steady state, that is, c = 0. Using
dissolution affinity Adiss as, this in Eq. (7) and combining with Eq. (10) gives the
steady state diffusion equation for Fe2+ ions in solution:

z F E corr
i 0 = z F Adiss exp (12) ∇ · J = ∇ · (D∇c) = 0 (16)
kB T

123
Comput Mech (2014) 54:613–627 617

For corrosion of stainless steels, it can be shown that rior Laplace problem. For activation controlled corrosion,
the steady-state assumption holds, using the analytical the variational or weak form can be written as follows:
solution of the 1D moving interface problem [11,49]. Find c ∈ C for all v ∈ V such that,
4. The temperature T and the applied overpotential η A  
are spatially constant at the pit interface  int , so the iA
− D ∇v · ∇c d + v d
molar flux J m is a constant along the interface during zF
 p
 int
activation-controlled corrosion. In addition, the precipi- 
i Ac
tation reaction products are assumed to be infinitely sol- = v d if cp < csat , (20)
z Fcsolid
uble in the solution, so the formation of a salt layer on  int
the pit interface is neglected. The formation of hydrogen
gas and changes in pH are also ignored. For this simple where C and V are spaces of sufficiently smooth functions for
case, the current density i A during activation-controlled the bulk field and their variations. For diffusion controlled
corrosion is spatially constant along the interface. corrosion, the use of Lagrange multipliers to enforce the
interfacial Dirichlet constraint results in the following mixed
2.3 Strong form of the free boundary problem variational form:
Find c ∈ C and λ ∈ L for all v ∈ V and u ∈ L such that,
The strong form of the boundary value problem for the metal   ⎫
 
concentration in the pit solution is given by, − D∇v · ∇c d+ λ c−csat d = 0,⎪ ⎬
⎫  p   int
∇ · (D∇c) = 0 in p , ⎬ u c d = u csat d, ⎪

p
c=0 on d , (17)  int  int

D∇c · n0 = 0 on h .s if cp = csat , (21)
When corrosion is activation-controlled, the normal inter- where L defines an admissible space for the Lagrange mul-
face speed V is directly known from the current density i A , as tipliers.
given by Eq. (13). When corrosion is diffusion-controlled, the
concentration at pit interface is given by the saturation limit
csat . This gives a Robin type condition (activation-controlled
process) or a Dirichlet type interface condition (diffusion- 3 Numerical implementation
controlled process) on the moving boundary  int depending
upon the local interface concentration cp as follows: In this section, the details of the XFE-LSM and the algo-
rithm are provided. The numerical implementation, herein,
⎫ is different from [41] in following ways: (1) The step enrich-
iA iA ⎪
if cP < csat −D∇cp · n + cp = ,⎪⎬ ment function is used instead of the discontinuous deriva-
z Fcsolid zF
on  int (t)

⎪ tive enrichment function; (2) the Lagrange multiplier method

else cP = csat , is employed for enforcing the Dirichlet interface condition,
(18) instead of the variationally inconsistent penalty method; (3)
a nonlocal domain integral scheme is used for evaluating the
By solving the above two Eqs. (17) and (18), the con- normal interface speed, instead of a direct evaluation scheme
centration field c(x, t) can be established in the pit solution based on the local concentration gradient; and (4) a bilinear
domain p (t). The normal interface speed under activation- rectangular finite element (FE) mesh is employed instead of
controlled cp < csat ) and diffusion-controlled corrosion a triangular FE mesh. The current XFEM-LSM implemen-
regimes (cp = csat ) is given by: tation uses a structured finite difference (FD) grid for the
⎧ level set evolution, whose nodes coincide with the nodes of
⎪ iA

⎪ if c p < csat ,
the structured bilinear FE (four-noded quadrilateral) mesh;
⎨ z Fcsolid
V = (19) therefore interpolation of the level set function φ between the

⎪ p·n

⎩− D∇c FD grid and the FE mesh is avoided at every time step. Since
if cp = csat , φ is continuous across the interface standard FE shape func-
c solid −c sat
tions are used to interpolate φ within finite elements. If the
2.4 Weak form FE mesh is unstructured or is composed of different order
elements (e.g. triangles or nine-noded quadrilaterals) or if
The above described boundary value problem along with the the FD grid is of finer resolution, then a bicubic interpolation
interface conditions belongs to the class of elliptic inter- scheme can be used to project φ from the FD grid to the FE
face problems. When D is constant it reduces to an inte- mesh.

123
618 Comput Mech (2014) 54:613–627

3.1 The level set method tracking method (e.g. marker particle method [51]) is that
the LSM can handle arbitrary topological transitions such as
The level set method (LSM) for moving interfaces was orig- corrosion pit merging (see Fig. 8c). This is important as corro-
inally introduced by Osher and Sethian [20,21], wherein the sion pit morphologies can evolve into complex shapes due to
interface location at any time t is given by the zero contour undercutting, subsurface pitting, and vertical and horizontal
of the scalar level set function, φ(x, t), that is, grain attack [7]. The LSM is relatively expensive compared to
the particle methods due to the solution of hyperbolic level set
 int = {x ∈  | φ(x, t) = 0}. (22) equation and the CFL condition (27) for stability; however,
one needs to pay this extra cost for a more robust handling
Following this approach, the interface is implicitly repre- of interface topology. An alternative numerical implementa-
sented by φ, such that φ = 0 on the pit interface  int , φ < 0 tion of the LSM could utilize a Stream-line Upwind Petrov-
in the pit solution domain p , and φ > 0 in the solid domain Galerkin (SUPG) finite element scheme. Although the SUPG
s . As corrosion progresses,  int (t) moves in the normal finite element level set scheme completely avoids the need
direction with speed V and the level set function φ satisfies for interpolation of φ and is second order accurate [52], it will
the level set evolution equation in the whole domain, be computationally much more expensive than the LSM.

∂φ 3.2 The extended finite element method (XFEM)


+ Vext
∇φ
= 0 in . (23)
∂t
The XFEM is used to discretize the weak form of the interior
where Vext is the extended normal interface speed. The above
Laplace problem defined by (20) and (21). Note that the con-
level set evolution equation is a first order hyperbolic PDE
centration field c(x, t) and its normal derivative ∇c · n(x, t)
and is subjected to the following initial and boundary condi-
are discontinuous across the pit interface  int (t) due to the
tions:
different metal ion concentrations and fluxes in the solid and
φ(x, 0) = ± min ||x − x  || in , (24) solution phases. Therefore, the enriched approximation is
x  ∈ int (0) defined as,
∇φ · n = 0 on . (25)

n e (t)
n
The above equations specify that φ is initialized using the
c(x, t) = N (x)c (t) +
A A
S B (x, t)a B (t), (28)
signed distance function to the interface and that the normal
A=1 B=1
flux is zero on the exterior boundary . The normal speed
is defined in the entire computational domain using velocity where N A (x) are the standard finite element shape functions,
extensions as, [22–24], c A are nodal degrees of freedom (DOFs), n is the total number

∇Vext · ∇φ = 0 in , of nodes in the mesh, a B are the nodal enrichment DOFs
(26) (which are additional DOFs at the enriched nodes), n e (t) is
Vext = V on  int .
the total number of enriched nodes in the mesh at time step t,
The velocity extension procedure is necessary because the and the step enrichment function corresponding to enriched
normal interface speed V is known only at points that lie on node B, is defined as,
the interface; however, the LSM requires that V be defined
at all nodes in the FD grid. Note that Eq. (26) is derived such  
S B (x, t) = N B (x) H(φ(x, t)) − H(−φ(x, t)) , (29)
that it preserves the signed distance property of the level set
function as it is evolved [23]; therefore, reinitialization of the
level set function is seldom required. where H(φ) is the Heaviside step function given by,

The level set evolution equation is solved using an upwind 1 i f φ > 0,
finite difference scheme and the fast marching method [24]. H(φ) = (30)
0 i f φ < 0.
The explicit forward Euler scheme is employed to update
φ using a stable adaptive time step given by the Courant- Unlike the discontinuous derivative enrichment function
Friedrichs-Lewy (CFL) condition [50], used in [41], S incorporates a discontinuity in the value of
the unknown field across the interface; however, S itself does
h min not enforce the jump in value or normal derivative across

t ≤ , (27) the interface, so interface conditions need to be enforced


Vmax
for the problem to be well-posed. In the pitting corrosion
where h min is the minimum FD grid size and the maximum problem, since the concentration is only changing in p ,
interface speed Vmax = max{V (x  , t)} for all x  ∈  int (t). during diffusion-controlled corrosion the jump in concentra-
The motivation for using the LSM rather than an explicit tion is captured by enforcing a one-sided Dirichlet constraint

123
Comput Mech (2014) 54:613–627 619

cp = csat on the moving interface using Lagrange multipli- where N̂ A are the standard 1D FE shape functions, λ A are
ers as described in Sect. 3.3; and during activation-controlled LM nodal DOFs and m(t) is the total number of LM nodes
corrosion the jump in the normal derivative is captured by at time step t. This strategy for generating the LM mesh
adding interface line source terms corresponding to a one- [54] has many similarities with the mortared finite element
sided normal flux (i.e. f c and f a ) to the right hand side of the method presented in [55].
system matrix, defined in Sect. 3.4. A more detailed account A key challenge in using Lagrange multipliers (LMs) for
on incorporating intra-element strong and weak discontinu- enforcing Dirichlet constraints on embedded interfaces in
ities is given in [53]. conjunction with the XFEM lies in the construction of a sta-
ble discrete LM space L. A poor choice of L often leads to
3.3 Imposition of the interfacial constraints spurious oscillations in the multipliers referred to as bound-
ary locking effect, which has been reported in several of the
In order for the pitting corrosion problem to be well-posed, studies cited below. To resolve this issue, the vital-vertex
boundary conditions must be enforced on the moving inter- algorithm was proposed in [56–58], which constructs a sta-
face. Previously in [41], the penalty method was used to ble reduced LM space that passes the numerical inf-sup test;
enforce Dirichlet conditions on the moving interface, but however, its implementation is relatively complex compared
this approach is variationally inconsistent. For the sake of to our implementation. Although some oscillations in the LM
accuracy, one must choose the penalty parameter to be sev- values were noticed in the current implementation, the mag-
eral orders of magnitude larger than the diffusion coefficient; nitude of these oscillations was relatively small; and by using
however, choosing a large penalty parameter often leads to ill- a smoothing procedure based on the domain integral scheme
conditioning of the system matrix which could lead to erro- [31] interface fluxes can be recovered accurately, as described
neous results or even singular systems. Instead, herein the in sect. 3.5. An alternative approach for imposing the inter-
variationally consistent Lagrange multiplier (LM) method face Dirichlet condition involves the Nitsche’s formulation
is employed to enforce the Dirichlet conditions on the pit [59], which could be computationally more efficient and be
solution side of the interface under diffusion-controlled cor- readily extensible to three-dimensions; moreover, it can lead
rosion. In the current scheme, a 2D mesh consisting of four- to optimal convergence [60–65], and possibly, eliminate the
noded (bilinear) square finite elements is used to discretize need for the smoothing procedure.
the bulk concentration field. Next, a 1D interface mesh that is
independent of the underlying 2D mesh is constructed such 3.4 Discrete equations
that there is only one LM node per interface segment con-
tained in a bulk enriched element, as shown in Fig. 2. If the Upon introducing the XFE and FE approximations into the
length of the interface segment in an enriched element is weak form (20) and (21), the following linear algebraic sys-
small then LM node is not added for this segment, so as to tem is obtained: for activation-controlled corrosion,
improve stability. On this interface mesh, the LMs are inter-


polated using piecewise linear functions as [54], Kcc Kca c fc


= , (32)
Kac Kaa a fa

m(t)
λ= N̂ A λ A , (31) and for diffusion-controlled corrosion,
A=1
⎡ ⎤⎡ ⎤ ⎡ ⎤
Kcc Kca Gc c 0
⎣ Kac Kaa Ga ⎦ ⎣ a⎦ = ⎣ 0 ⎦ , (33)
(Gc )T (Ga )T 0 λ g

where c, a, and λ are the arrays containing the stan-


dard, enriched and Lagrange multiplier DOFs, respectively;
Kcc , Kca , Kac and Kaa are the standard and enriched tangent
matrix terms; f c and f a are the standard and enriched vec-
tors of the line source terms, respectively; Gc , Ga , and g are
the matrix and vector terms corresponding to the Lagrange
multipliers. The domain integrals corresponding to all these
Fig. 2 The 2D XFE mesh and the corresponding 1D Lagrange multi-
matrices and vectors are given in Appendix A. The concentra-
plier mesh used for the initially circular pit interface are shown. Notice
that the XFE-LSM bulk mesh is independent of the embedded interface, tion field c(x, t) is obtained by solving the above linear sys-
so there is no need to remesh or move the mesh as the pit grows, which tem. The interface speed can then be computed as described
leads to computational efficiency in the next section.

123
620 Comput Mech (2014) 54:613–627

3.5 Domain integral evaluation of the normal interface where  A denotes the support of the FE node A and  int A is
speed the interface segment contained in an element e that belongs
to the support  A . Finally, the interface speed at point x ∗ is
As described in Sect.2, for diffusion-controlled corrosion calculated according to Eq. (19) as,
(cp = csat ), the normal interface speed V is given in terms of 
the normal interface flux D∇cp · n on the pit solution side, ∗
D∇cp · n x ∗
V (x , t) = − . (37)
according to Eq. (19). However, a direct evaluation of the flux csolid − csat
using the local concentration gradient at a point on the inter-
face, as previously proposed in [41], can give rise to errors The above procedure is a nonlocal scheme for calculating
in V and will yield a poor quality interface topology [66]. V and yields a better approximation compared to the local
Although the LM nodal values can be interpreted as interface scheme implemented in [41].
fluxes, they do not provide a good estimate because the 1D
interface mesh is chosen to be sufficiently coarse for stability 3.6 Algorithm
reasons. Also, there could be minor oscillations in the LM
nodal values as discussed Sect. 3.3. Instead a domain integral The strategy here is to adopt a staggered explicit time scheme,
scheme proposed in [31] gives a better approximation of the and the algorithm for evolving the corrosion front at time t
normal interface flux, which is used herein to calculate V . is:
Let us now briefly review the procedure to approximate the
flux D∇cp · n at a point x ∗ on the pit interface  int using the 1. Project the level set function φ from the FD grid onto the
domain integral scheme. Consider a section of the interface XFE mesh;
∗int containing the point x ∗ , within compact support ∗ of a 2. Assemble the linear systems in Eq. (32) or (33) using the
scalar weight function w∗ . Upon multiplying the first equa- XFEM and solve for the concentration field c(x, t) in p ;
tion in (17) by w ∗ , integrating over ∗ by parts and then 3. Compute the normal interface speed V according to Eq.
applying the divergence theorem, one can obtain the follow- (37) using the domain integral scheme, as described in
ing relation: Sect. 3.5;
  4. Solve the interface evolution Eq. (23) using the LSM and
D ∇w ∗ · ∇c d = w ∗ D∇cp · n d. (34) the explicit forward Euler method, as described in Sect.
3.1;
∗ ∗int 5. If t < tfinal then go to step 1, else end computation.
Next, let us adopt a first order approximation by assum-
ing the normal derivative D∇cp · n to be constant over ∗int .
4 Numerical results
Thus, a simple approximation for the normal component of
the concentration flux at x ∗ ∈ ∗int , when the interface is
In this section, two types of numerical experiments are simu-
approached from p , can be written as,
lated. First, the pencil electrode corrosion tests are simulated
 with the aim of validating the approach by comparing our
 ∗ ∇w ∗ D · ∇c d
D∇c · n x ∗ =
p 

. (35) numerical results with those from published experimental
 int w d
∗ [67] and numerical studies [12]. Second, accelerated foil cor-
rosion tests are simulated to predict corrosion morphologies
As noted in [31], the above result is independent of the par- with the goal of demonstrating the viability of the approach
ticular choice of the weight function w ∗ , so there are several
 in two-dimensions.
ways to construct the approximation. Herein, D∇cp · n x ∗ is
evaluated by: (1) identifying the nodes of the finite element 4.1 Pencil electrode corrosion test
containing x ∗ ; (2) evaluating the contribution of those nodes
to the normal flux on the pit solution side; and (3) interpolat- 4.1.1 Experimental validation
ing the value at x ∗ using these nodal contributions according
to the standard finite element approximation as, Let us first simulate the 1D pencil electrode experiments
  described in [67], wherein a stainless steel (304 SS) wire
NA ∇w ∗ D · ∇c d of 50 μm diameter is mounted in epoxy resin and immersed
 A  in electrolyte solution so that only one end of the wire was
D∇cp · n x ∗ =  
A
(36) exposed to the solution. In this case, the concentration gradi-
w ∗ d ent is only along the longitudinal axis of the wire (i.e. a 1D
e∈ A int problem), so it is reasonable to model this as a flat interface
A

123
Comput Mech (2014) 54:613–627 621

Table 1 Model parameters and constants used in this study


Parameter Value Units Source

Adiss 4.0 mol/cm2 /s [12]


csat 5.1 mol/l [68]
csolid 143 mol/l [9]
D(T = 25 ◦ C) 0.824 × 10−5 cm2 /s [9]
E corr −0.24 mV [69]
F 96485.53 C/mol -
kB 8.314 J/mol./K -
z 2.19 - [11]
α 0.65 - [12]
Fig. 3 A schematic diagram of the pencil electrode corrosion test. The
side walls are made of non-corrosive epoxy resin and so they serve as
zero flux boundaries. The pit depth d is the distance between the pit
mouth and the metal surface to the anodic metal dissolution, the anodic current density i A
provides a good estimate of the rate of corrosion. Figure 4d
shows that pit depth increases linearly with the inverse of the
in a rectangular domain with gradients only in the x2 direc-
current density i A , which is also consistent with experiments.
tion. Figure 3 shows the schematic diagram of the test set-up
along with the domain–boundary definitions. The numeri-
cal results are obtained using a 2D computational domain 4.1.2 Numerical validation
that is 0.16 mm (length x1 ) × 0.16 mm (height x2 ) with
a flat initial interface of height x2 = 0.157 mm and a flat Let us next consider three different values of overpotential
pit mouth boundary at x2 = 0.16 mm. The initial pit depth η A =100 mV, 140 mV, 200 mV and compare our numerical
d = 3 μm is the distance between the pit mouth and the metal results with those of [12] obtained from a finite volume model
surface. Thus, the 2D numerical problem is constrained to (FVM). For this study, D = 0.85 × 10−5 cm2 /s, according to
mimic 1D experimental set-up. The experiments were con- [12]. The same domain and mesh discretization as described
ducted at a constant applied potential of E = +600 mV above is employed. The scatter (diamonds, circles, squares)
versus the reference Saturated Calomel Electrode (SCE) at a of the numerical curves shown in Fig. 5 is obtained from Fig.
temperature of T = 15 ◦ C so that diffusion-controlled sta- 6 in [12]. For the low overpotential η A = 100 mV, corrosion
ble pitting occurs. The diffusion coefficient at T = 25 ◦ C is activation-controlled, so the current density i A is constant
is reported to be D25 = 0.824 × 10−5 cm2 /s by Gaudet et throughout the simulation. Therefore, the pit depth versus
al. [9], and using Einstein’s relation (kinetic theory) one can time plot is a straight line as shown in Fig. 5a and b. For
obtain D15 = 0.796 × 10−5 cm2 /s. The concentration on the intermediate overpotential η A = 140 mV, corrosion is
p initially activation controlled, although after 40 seconds the
the pit mouth d = {x ∈  | x2 = 0.16 mm} is fixed at
zero and the concentration at the pit interface  int is fixed process becomes diffusion-controlled. This is evident from
at csat = 5.1 mol/l [68]. All the model parameter values are the kink in the pit growth rate and current density curves
listed in Table 1. In all the simulations, the finite difference in Fig. 5. For the high overpotential η A = 200 mV, corro-
(FD) grid for the level set method is chosen so that the spa- sion process is diffusion-controlled right from the beginning,
tial coordinates of the FD nodes coincide with those of XFE and the current density is inversely is proportional to time.
nodes. For this problem, an XFE mesh consisting of 200 × As expected the XFE-LSM results match exactly with those
200 elements is employed, so the FD grid consists of 201 × from the FVM, which validates the proposed approach. The
201 nodes. Under the steady state assumption, the average main advantage of the XFE-LSM over the FVM is its ability
ion concentration field in the pit solution varies linearly with to accurately evolve realistic corrosion pit morphology.
depth in p as shown in Fig. 4a. The numerical results of
real-time pit growth and current density are compared with 4.2 Accelerated foil corrosion test
the experimental data of [67] in Fig. 4b and c. The scatter of
the experimental data points is obtained from Figs. 2 and 5 in 4.2.1 Growth of an isolated pit
[67], and the data was adjusted by taking time t = 25 seconds
as t = 0 so as to only consider the regime when η A is truly Let us next simulate the growth of a single isolated pit
constant in the experiment. As expected, during diffusion- similar to the experiment conducted by Ernst and Newman
controlled corrosion the pit depth increases linearly with the [67], wherein a thin stainless steel (SS) foil was sandwiched
square root of time (see Fig. 4b). Since pit formation is due between two optical glass plates so that only one edge is

123
622 Comput Mech (2014) 54:613–627

(a) (b)

(c) (d)
Fig. 4 Comparison of our numerical results (lines) with the experi- depth increases linearly with the square root of time and with the inverse
ments results (scatter) reported in Figs. 2 and 5 in [67]. The excellent of current density as expected for a diffusion-controlled corrosion
agreement of the results demonstrates the validity of the model. The pit

80
10

60 8

6
40
4

20
2

0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) (b)

Fig. 5 Comparison of our numerical results obtained using the circles, squares) in subfigure (a) is obtained from Fig. 6 in [12] and
extended finite element and level set method (XFE-LSM) with the the lines show our numerical results. The current density is constant
numerical results of Scheiner and Hellmich [12] obtained using the with time during activation-controlled corrosion and is inversely pro-
finite volume method (FVM). The scatter of the data points (diamonds, portional to time during diffusion-controlled corrosion

exposed to the solution. The exposed surface of the foil 2D idealization of the problem is reasonable as experimen-
was coated with lacquer leaving only a small square area tal observations support the assumption of uniformity in the
of 250 μm2 in contact with the pit solution. In this case, a x3 (thickness) direction. The test set-up can be described

123
Comput Mech (2014) 54:613–627 623

200

150

100

50

0
0 250 500 750 1000

(a) (b)
0.6

0.55

0.5

0.45
0.2 0.3 0.4

(c) (d)

Fig. 6 Comparison of semi-circular shape pit growth with experiments of Ernst and Newman [67]. The pit is grown from the edge of a 304 stainless
steel (SS) foil where the edge was covered with lacquer except for a small hole that is ≈ 250 μm2

by the schematic diagram in Fig. 1, except that the inter- × 200 elements. The results from these 3 meshes overlap
face shape is semi-circular. The experiment was conducted with each other demonstrating good convergence; however,
at a constant applied potential of E = +600 mV versus the simulation predicted smaller pit depth at t = 720 seconds
the reference Saturated Calomel Electrode (SCE) at temper- compared to the experiment, as indicated by data point (red
ature T = 20 ◦ C so that diffusion-controlled stable pitting square) in Fig. 6b. This discrepancy could be arising because
occurs. Using Einstein’s relation, the diffusion coefficient at of: (1) uncertainty in the value of diffusion coefficient at dif-
T = 20 ◦ C is calculated to be D20 = 0.8102 × 10−5 cm2 /s. ferent temperatures; and (2) factors such as electro-migration
The 2D computational domain is taken to be 0.6 mm (length and activity of ions in the saturated or super-saturated pit
x1 ) × 0.6 mm (height x2 ) with an initial pit of diameter 0.016 solution. To address this discrepancy, the model may need to
mm so as to mimic the physical experimental set-up of [67]. be extended by incorporating the Nernst-Planck-Poisson for
As corrosion progresses the pit grows wider and deeper, but electro-diffusion as opposed to Fickian diffusion, in conjunc-
the pit mouth is assumed to remain 0.016 mm wide and the tion with more details experiments. Finally, from Fig. 6c it
rest of the top boundary is a zero flux boundary because it is evident that an isolated pit grows as a semi-circular shape
is coated with lacquer. The concentration at the pit mouth throughout the diffusion-controlled corrosion process, which
p
d = {x ∈  | 0.292 mm ≤ x1 ≤ 0.308 mm, x2 = 0.6 mm} is qualitatively consistent with the observations of Ernst and
is fixed at zero and the concentration at the pit interface  int Newman [67] shown in Fig. 6d.
is fixed at csat = 5.1 mol/l using Lagrange multipliers. The
numerical results were obtained using an XFE mesh consist- 4.2.2 Pit shape evolution
ing of 600 × 600 elements. A typical field plot of the average
ion concentration in the solution domain is shown in Fig. 6a. Let us next examine the morphology evolution of an iso-
The maximum pit depth is plotted as a function of time in Fig. lated semi-elliptical (crack-like) pit during activation- and
6b, which plateaus as the corrosion rate decreases with time diffusion- controlled corrosion processes. The same domain
under diffusion-control. Next, a numerical convergence study and XFEM mesh as above is employed and initialized with a
of the XFE-LSM is conducted by considering 3 different XFE semi-elliptical pit with a major axis length 0.032 mm and a
mesh resolutions consisting of 600 × 600, 400 × 400, 200 minor axis length 0.008 mm. For activation-controlled cor-

123
624 Comput Mech (2014) 54:613–627

0.6 0.6

0.58 0.58

0.56 0.56

0.54 0.54

0.25 0.3 0.35 0.25 0.3 0.35

(a) (b)

0.6 0.6

0.59 0.59

0.58 0.58

0.57 0.57

0.28 0.3 0.32 0.28 0.3 0.32

(c) (d)

Fig. 7 Comparison of pit morphologies under activation- and process (b, d) smoothens the pit into a semi-circular shape. The nor-
diffusion-controlled corrosion processes. The numerical results of pit mal interface speed is spatially constant during the activation-controlled
interface evolution show that the activation-controlled process (a, c) pre- corrosion
serves the semi-elliptical shape of the pit, whereas diffusion-controlled

rosion η A = 100 mV and for diffusion-controlled corrosion grow as elongated elliptical shapes with lacy metal cover
csat = 5.1 mol/l. The numerical results of pit interface evo- formation on the top boundary. The mechanisms behind the
lution, in Fig. 7a and b, show that the semi-elliptical shape formation of the lacy cover are complex; however, the perfo-
of the pit is preserved during activation-controlled corro- rated cover leads to the evolution of elongated pit shapes. To
sion, whereas the pit morphology transitions into a semi- illustrate the role of the perforated metal cover, a 2D com-
circular shape during diffusion-controlled corrosion. This putational domain that is 0.36 mm (length x1 ) × 0.36 mm
happens because the corrosion rate at the interface during (height x2 ) is initialized with several small semi-circular pits,
an activation-controlled process is spatially constant for a each of diameter 0.004 mm, along the top boundary of the
given T and η A . Therefore, V is constant along the interface, domain. The domain is discretized using an XFE mesh con-
as indicated by length of the arrows in Fig. 7c. On the other sisting of 200 × 200 elements. In this case, since there was
hand, in a diffusion-controlled process the corrosion rate at a lacy cover formation on the exposed surface due to numerous
point on the interface is inversely proportional to its distance small pits, the assumption of uniformity in x3 direction is not
to the pit mouth, so V is larger at points on the interface that strictly valid, and so the 2D idealization of the problem is a
are closer to the pit mouth, as shown in Fig. 7d. Thus, the simplifying approximation. Another assumption here is that
pit shape is preserved only during an activation-controlled there is no change in the protective layer on the metal surface
process, which was also pointed out in [12,70]. as corrosion progresses via metal dissolution inside the pit
beneath the lacy cover. Therefore, the pit mouth of each pit
4.2.3 Growth of multiple pits remains 0.004 mm wide even as the pits grow and coalesce,
which can be clearly seen from the field plot of the concen-
Let us next investigate the morphology evolution of multi- tration variable as shown in Fig. 8a. The pit is evolved under
ple closely spaced pits during diffusion-controlled corrosion. a diffusion-controlled corrosion process at T = 15 ◦ C and
Ernst and Newman [67] experimentally observed pits that E = +600 mV (SCE) for about 300 seconds and the corre-

123
Comput Mech (2014) 54:613–627 625

150

100

50

0
0 50 100 150 200 250 300

(a) (b)

0.36

0.3

0.24
0 0.06 0.12 0.18 0.24 0.3 0.36

(c) (d)

Fig. 8 Multiple pit growth in SS foils with partially broken passive layer on the top surface. The coalescence of nearby pits leads to the formation
of elongated elliptical pits. The numerical pit shapes are in qualitative agreement with the experimental shapes

sponding maximum pit depth is plotted as a function of time normal interface speed in conjunction with the Lagrange mul-
in Fig. 8b. The model seems to predict a faster rate of corro- tiplier method. The numerical results from the pencil elec-
sion as compared to the experiment at t = 240 seconds, as trode test are in excellent quantitative agreement with both
indicated the data point (red square) in Fig. 8b. This discrep- experimental and numerical results in the literature, demon-
ancy may be attributed to the 2D idealization and arbitrary strating the validity of the model for simpler pit geometries.
choice of the initial pit locations and sizes. Figure 8c shows The foil corrosion test results are qualitatively consistent with
that the individual pits initially grow as circular shapes but experiments, illustrating that the model is able to capture the
soon coalesce to form an elongated elliptical shape. Thus, mechanisms of pit propagation under diffusion-controlled
this simulation study provides an explanation for the mecha- corrosion. Only one experimental data point in time is avail-
nism behind the formation of elongated pits experimentally able from the foil corrosion tests reported in [67], and the
observed by Ernst and Newman [67] shown in Fig. 8d. predicted pit growth rates in the case of an isolated pit or
multiple pits do not agree quantitatively with the data. This
discrepancy could be arising because of the 2D idealization
5 Conclusion and because the model does not consider complex phenom-
ena associated with activity of ions in concentrated solutions,
A simple yet robust sharp interface formulation for model- electro-migration, reactions, local microstructural features
ing corrosion pit propagation in alloys is presented, within and the effects of chloride ion concentration in pit solution.
the framework of the XFE-LSM. In contrast with existing Also, there is uncertainty associated with the model parame-
approaches based on the standard FEM or the FVM, the pro- ters and experimental data that could be another source of
posed approach allows for an accurate representation of the the discrepancy. Clearly, this work is only a first step toward
geometry of the corrosion front independent of the underly- developing a comprehensive extended finite element model
ing finite element mesh. The method is convergent and can for localized corrosion, and future work will focus on improv-
capture complex morphological transitions such as pit merg- ing the model in conjunction with more detailed experimen-
ing without requiring remeshing. The generalized Heaviside tation.
or step function allows for the jumps in both the concentration
field and its normal derivative across the pit interface. The Acknowledgments This work was supported by a start-up grant from
domain integral scheme yields a good approximation of the Vanderbilt University. The author is grateful to Dr. Chandrasekhar

123
626 Comput Mech (2014) 54:613–627

Annavarapu at Lawrence Livermore National Laboratory, USA, for the 4. Isaacs HS (1973) The behavior of resistive layers in the localized
discussions on the XFEM implementation corrosion of stainless steel. J Electrochem Soc 120(11):1456–1462
5. Sharland SM (1987) A review of the theoretical modelling of
crevice and pitting corrosion. Corros Sci 27(3):289–323
6. Kontturi K, Murtomäki L, Manzanares JA (2008) Manzanares.
Appendix: Element matrices and vectors Lonic transport processes: in electrochemistry and membrane sci-
ence. OUP Oxford, Oxford
7. ASTM Standard (2013) G46, guide for examination and evalua-
For the sake of clarity, the domain integral terms associated tion of pitting corrosion. American Society for Testing Materials
with the element tangent matrix and element force vector are International, West Conshohocken. doi:10.1520/G0046.
provided below: 8. Alkire R, Siitari D (1979) The location of cathodic reaction during
 localized corrosion. J Electrochem Soc 126(1):15–22
9. Gaudet GT, Mo WT, Hatton TA, Tester JW, Tilly J, Isaacs HS, New-
AB
K cc = − D ∇ N A · ∇ N B d, (38) man RC (1986) Mass transfer and electrochemical kinetic interac-
e tions in localized pitting corrosion. AIChE J 32(6):949–958
 10. Sharland SM, Jackson CP, Diver AJ (1989) A finite-element model
AB
K ca = BA
K ac =− D ∇ N A · ∇S B d, (39) of the propagation of corrosion crevices and pits. Corros Sci
29(9):1149–1166
e
 11. Scheiner S, Hellmich C (2007) Stable pitting corrosion of stain-
AB
K aa =− D ∇S A · ∇S B d, (40) less steel as diffusion-controlled dissolution process with a sharp
moving electrode boundary. Corros Sci 49(2):319–346
e 12. Scheiner S, Hellmich C (2009) Finite volume model for diffusion-

and activation-controlled pitting corrosion of stainless steel. Com-
G cAB = N A N̂ B d, (41) put Method Appl Mech Eng 198(37–40):2898–2910
I 13. Turnbull A, Horner DA, Connolly BJ (2009) Challenges in mod-
 eling the evolution of stress corrosion cracks from pits. Eng Fract
G aAB = SpA (0) N̂ B d, (42) Mech 76(5):633–640
14. Oltra R, Malki B, Rechou F (2010) Influence of aeration on
I the localized trenching on aluminium alloys. Electrochim Acta
 55(15):4536–4542
iA
f cA = N A
d, (43) 15. Xiao J, Chaudhuri S (2011) Predictive modeling of localized
zF corrosion: an application to aluminum alloys. Electrochim Acta
I
 56(16):5630–5641
iA 16. Sarkar S, Warner JE, Aquino W (2012) A numerical framework
f aA = SpA (0) d, (44) for the modeling of corrosive dissolution. Corros Sci 65:502–511
zF
I 17. Zhu Z, Tajallipour N Teevens PJ (2011) A mechanistic model for
 predicting localized pitting corrosion in a brine water-CO2 system.
g =
A
N̂ A csat d (45) Corrosion
18. Belytschko T, Black T (1999) Elastic crack growth in finite ele-
I ments with minimal remeshing. Int J Numer Method Eng 45:601–
620
where superscripts A, B denote the nodal indices; N A is the 19. Moës N, Dolbow J, Belytschko T (1999) A finite element method
standard bilinear (2D) FE shape functions corresponding to for crack growth without remeshing. Int J Numer Method Eng
node A; N̂ B is the standard linear (1D) FE shape functions 46(1):131–150
20. Sethian JA (1999) Level set methods & fast marching methods:
corresponding to the Lagrange multiplier node B; Sp (0) indi- evolving interfaces in computational geometry, fluid mechanics,
cates the enrichment function evaluated on the pit solution computer vision, and materials science. Cambridge University
side of the interface; e indicates the domain correspond- Press, Cambridge
ing to enriched element e and  I is the interface segment 21. Osher S, Sethian JA (November 1988) Fronts propagating with
curvature-dependent speed: algorithms based on Hamilton–Jacobi
contained by the enriched element e. For more details on the formulations. J Comput Phys 79(1):12–49
numerical integration the reader is referred to [54]. 22. Sethian JA (1996) A marching level set method for monotonically
advancing fronts. Proc Nat Acad Sci USA 93(4):1591–1595
23. Adalsteinsson D, Sethian JA (1999) The fast construction of
extension velocities in level set methods. J Comput Phys 148:2–22
References 24. Chopp David L (2001) Some improvements of the fast marching
method. SIAM J Sci Comput 23(1):230–244
1. Hoeppner DW, Taylor AMH (2011) AVT-140 Corrosion fatigue 25. Holl M, Rogge T, Loehnert S, Wriggers P, Rolfes R (2013) 3D
and environmentally assisted cracking in aging military vehicles, multiscale crack propagation using the XFEM applied to a gas
chapter 13. Modelling pitting corrosion fatigue: Pit growth and turbine blade. Computational Mechanics, pp 1–16.
Pit/Crack transition issues. NATO, RTO, NATO, France 26. Amiri F, Anitescu C, Arroyo M, Bordas SPA, Rabczuk T (2013)
2. Macdonald DD, Engelhardt GR (2010) Schreir’s corrosion, chapter XLME interpolants, a seamless bridge between XFEM and
predictive modeling of corrosion. Elsevier, Amsterdam, pp 1630– enriched meshless methods. Computational Mechanics, pp 1–13.
1679 27. Chessa J, Smolinski P, Belytschko T (2002) The extended finite
3. Scully JC (1990) The fundamentals of corrosion, 3rd edn. Perga- element method (XFEM) for solidification problems. Int J Numer
mon Press, New York Method Eng 53(8):1959–1977

123
Comput Mech (2014) 54:613–627 627

28. Merle R, Dolbow J (2002) Solving thermal and phase change 49. Mainguy M, Coussy O (2000) Propagation fronts during calcium
problems with the extended finite element method. Comput Mech leaching and chloride penetration. J Eng Mech 126(3):250–257
28(5):339–350 50. Courant R, Friedrichs K, Lewy H (1928) Über die partiellen
29. Ji H, Chopp D, Dolbow JE (2002) A hybrid extended finite element differenzengleichungen der mathematischen Physik. Math Ann
/ level set method for modeling phase transformations. Int J Numer 100:32–74
Method Eng 54(8):1209–1233 51. Du J, Fix B, Glimm J, Jia X, Li X, Li Y, Wu L (2006) A simple
30. Legay A, Chessa J, Belytschko T (2006) An Eulerian-Lagrangian package for front tracking. J Comput Phys 213(2):613–628
method for fluid-structure interaction based on level sets. Comput 52. Shepel SV, Paolucci S, Smith BL (2005) Implementation of a level
Method Appl Mech Eng 195(17–18):2070–2087 set interface tracking method in the FIDAP and CFX-4 codes. J
31. Ji H, Dolbow JE (2004) On strategies for enforcing interfacial con- Fluids Eng 127(4):674–686
straints and evaluating jump conditions with the extended finite 53. Hansbo A, Hansbo P (2004) A finite element method for the simula-
element method. Int J Numer Method Eng 61(14):2508–2535 tion of strong and weak discontinuities in solid mechanics. Comput
32. Dolbow JE, Fried E, Ji H (2005) A numerical strategy for investigat- Method Appl Mech Eng 193(33–35):3523–3540
ing the kinetic response of stimulus-responsive hydrogels. Comput 54. Vaughan BL, Smith BG, Chopp DL (2006) A comparison of
Method Appl Mech Eng 194(42–44):4447–4480 the extended finite element method with the immersed interface
33. Ji H, Mourad H, Fried E, Dolbow JE (2006) Kinetics of thermally method for elliptic equations with discontinuous coefficients and
induced swelling of hydrogels. Intl J Solids Struct 43(7–8):1878– singular sources. Communications in Applied Mathematics. and
1907 Computational. Science 1(1):207–228
34. Farsad M, Vernerey FJ, Park HS (2010) An extended finite ele- 55. Kim T-Y, Dolbow J, Laursen T (2007) A mortared finite element
ment/level set method to study surface effects on the mechanical method for frictional contact on arbitrary interfaces. Comput Mech
behavior and properties of nanomaterials. Int J Numer Method Eng 39(3):223–235
84(12):1466–1489 56. Moës N, Béchet E, Tourbier M (2006) Imposing Dirichlet bound-
35. Vernerey FJ, Farsad M (2011) An Eulerian/XFEM formulation for ary conditions in the extended finite element method. Int J Numer
the large deformation of cortical cell membrane. Comput Method Method Eng 67(12):1641–1669
Biomech Biomed Eng 14(5):433–445 57. Béchet E, Moës N, Wohlmuth B (2009) A stable Lagrange multi-
36. Farsad M, Vernerey FJ (2012) An XFEM-based numerical strategy plier space for stiff interface conditions within the extended finite
to model mechanical interactions between biological cells and a element method. Int J Numer Method Eng 78(8):931–954
deformable substrate. Int J Numer Method Eng 92(3):238–267 58. Nistor I, Guiton MLE, Massin P, Moës N, Géniaut S (2009) An
37. Vernerey FJ, Farsad M (2014) A mathematical model of the coupled X-FEM approach for large sliding contact along discontinuities.
mechanisms of cell adhesion, contraction and spreading. J Math Int J Numer Method Eng 78(12):1407–1435
Biol 68(4):989–1022 59. Nitsche JA (1970–1971) Über ein variationsprinzip zur lösung von
38. Hiriyur B, Waisman H, Deodatis G (2011) Uncertainty quantifica- dirichlet-problemen bei verwendung von teilräumen, die keinen
tion in homogenization of heterogeneous microstructures modeled randbedingungen unterworfen sind. Abhandlungen aus dem Math-
by XFEM. Int J Numer Method Eng 88(3):257–278 ematischen Seminar der Universität Hamburg, 36:9–15.
39. Benowitz BA, Waisman H (2013) A spline-based enrichment func- 60. Dolbow J, Harari I (2009) An efficient finite element method for
tion for arbitrary inclusions in extended finite element method embedded interface problems. Int J Numer Method Eng 78(2):229–
with applications to finite deformations. Int J Numer Method Eng 252
95(5):361–386 61. Harari I, Dolbow J (2010) Analysis of an efficient finite ele-
40. Yvonnet J, Le Quang H, He Q-C (2008) An XFEM/level set ment method for embedded interface problems. Comput Mech
approach to modelling surface/interface effects and to computing 46(1):205–211
the size-dependent effective properties of nanocomposites. Com- 62. Embar A, Dolbow J, Harari I (2010) Imposing Dirichlet boundary
put Mech 42(1):119–131 conditions with Nitsche’s method and spline-based finite elements.
41. Duddu R, Bordas S, Chopp DL, Moran B (2008) A combined Int J Numer Method Eng 83(7):877–898
extended finite element and level set method for biofilm growth. 63. Hautefeuille M, Annavarapu C, Dolbow JE (2012) Robust imposi-
Int J Numer Method Eng 74(5):848–870 tion of Dirichlet boundary conditions on embedded surfaces. Int J
42. Duddu R, Chopp DL, Moran B (2009) A two-dimensional contin- Numer Method Eng 90(1):40–64
uum model of biofilm growth incorporating fluid flow and shear 64. Annavarapu C, Hautefeuille M, Dolbow JE (2012) A robust
stress based detachment. Biotechnol Bioeng 103(1):92–104 Nitsche’s formulation for interface problems. Comput Method
43. Duddu R, Chopp DL, Voorhees PW, Moran B (2009) Diffusional Appl Mech Eng 225–228:44–54
evolution of precipitates in elastic media using the extended finite 65. Annavarapu C, Hautefeuille M, Dolbow JE (2012) Stable impo-
element and the level set methods. J Comput Phys 230(4):1249– sition of stiff constraints in explicit dynamics for embedded finite
1264 element methods. Int J Numer Method Eng 92(2):206–228
44. Zhao X, Duddu R, Bordas SPA, Qu J (2013) Effects of elastic 66. Smith BG, Vaughan BL, Chopp DL (2007) The extended finite
strain energy and interfacial stress on the equilibrium morphology element method for boundary layer problems in biofilm growth.
of misfit particles in heterogeneous solids. J Mech Phys Solids Comm Appl Math Comp Sci 2(1):35–56
61(6):1433–1445 67. Ernst P, Newman RC (2002) Pit growth studies in stainless steel
45. Zhao X, Bordas SPA, Qu J (2013) A hybrid smoothed extended foils. I. Introduction and pit growth kinetics. Corros Sci 44(5):927–
finite element/level set method for modeling equilibrium shapes of 941
nano-inhomogeneities. Computational Mechanics, pp 1–12. 68. Laycock NJ, Newman RC (1998) Temperature dependence of pit-
46. Gurtin ME, Ian MA (1975) A continuum theory of elastic material ting potentials for austenitic stainless steels above their critical
surfaces. Arch Rational Mech Anal 57(4):291–323 pitting temperature. Corros Sci 40(6):887–902
47. Gurtin ME, Voorhees PW (1993) The continuum mechanics of 69. Chaudhari S, Sainkar SR, Patil PP (2007) Poly(o-ethylaniline) coat-
coherent two-phase elastic solids with mass transport. Proc Royal ings for stainless steel protection. Prog Org Coat 58(1):54–63
Soc Lond A 440(1909):323–343 70. Sato N (1995) The stability of localized corrosion. Corros Sci
48. Coussy O (2004) Poromechanics. Wiley, New York 37(12):1947–1967

123

View publication stats

You might also like