arxiv:2004.14379v1 [math.ap] 29 apr 2020

31
ON A NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES OLENA BURKOVSKA 1 AND MAX GUNZBURGER 2 Abstract. A nonlocal Cahn–Hilliard model with a nonsmooth potential of double-well obstacle type that promotes sharp interfaces in the solution is presented. To capture long-range interactions between particles, a nonlocal Ginzburg-Landau energy functional is defined which recovers the classical (lo- cal) model as the extent of nonlocal interactions vanish. In contrast to the local Cahn–Hilliard problem that always leads to diffuse interfaces, the pro- posed nonlocal model can lead to a strict separation into pure phases of the substance. Here, the lack of smoothness of the potential is essential to guar- antee the aforementioned sharp-interface property. Mathematically, this in- troduces additional inequality constraints that, in a weak formulation, lead to a coupled system of variational inequalities which at each time instance can be restated as a constrained optimization problem. We prove the well posed- ness and regularity of the semi-discrete and continuous in time weak solutions, and derive the conditions under which pure phases are admitted. Moreover, we develop discretizations of the problem based on finite element methods and implicit-explicit time stepping methods that can be realized efficiently. Finally, we illustrate our theoretical findings through several numerical experiments in one and two spatial dimensions that highlight the differences in features of local and nonlocal solutions and also the sharp interface properties of the nonlocal model. 1. Introduction The Cahn-Hilliard model was proposed in [14] as the model to describe phase separation of a binary alloy. Since then, the model and its variants have been widely used in different areas of science such as, e.g., tumor growth, image segmentation and copolymer melts, cf. [7, 41, 45, 58, 59, 65]. 1 Computer Science and Mathematics Division, Oak Ridge National Laboratory, One Bethel Valley Road, TN 37831, USA 2 Department of Scientific Computing, Florida State University, 400 Dirac Science Library, Tallahasse, FL 32306-4120, USA, and the Oden Institute for Computer Engi- neering and Sciences, University of Texas at Austin, Austin, TX 78712, USA E-mail addresses: [email protected], [email protected]. 2010 Mathematics Subject Classification. 45K05; 35K55; 35B65; 49J40; 65M60; 65K15. Key words and phrases. nonlocal Cahn-Hilliard model, variational inequality, well-posedness, regularity, finite elements. This material is based upon work supported by the U.S. Department of Energy, Office of Advanced Scientific Computing Research, Applied Mathematics Program under the award num- bers ERKJ345 and ERKJE45; and was performed at the Oak Ridge National Laboratory, which is managed by UT-Battelle, LLC under Contract No. De-AC05-00OR22725. U.S. Government retains a non-exclusive, royalty-free license to publish or reproduce the published form of this con- tribution, or allow others to do so, for U.S. Government purposes. MG was also supported by U.S. Department of Energy grant DE-SC0021077 as part of the AEOLUS Multifaceted Mathematics Integrated Capability Center. 1 arXiv:2004.14379v2 [math.AP] 9 Sep 2021

Upload: others

Post on 06-Apr-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

ON A NONLOCAL CAHN-HILLIARD MODEL

PERMITTING SHARP INTERFACES

OLENA BURKOVSKA1 AND MAX GUNZBURGER2

Abstract. A nonlocal Cahn–Hilliard model with a nonsmooth potential ofdouble-well obstacle type that promotes sharp interfaces in the solution is

presented. To capture long-range interactions between particles, a nonlocal

Ginzburg-Landau energy functional is defined which recovers the classical (lo-cal) model as the extent of nonlocal interactions vanish. In contrast to the

local Cahn–Hilliard problem that always leads to diffuse interfaces, the pro-

posed nonlocal model can lead to a strict separation into pure phases of thesubstance. Here, the lack of smoothness of the potential is essential to guar-

antee the aforementioned sharp-interface property. Mathematically, this in-

troduces additional inequality constraints that, in a weak formulation, lead toa coupled system of variational inequalities which at each time instance can

be restated as a constrained optimization problem. We prove the well posed-

ness and regularity of the semi-discrete and continuous in time weak solutions,and derive the conditions under which pure phases are admitted. Moreover,

we develop discretizations of the problem based on finite element methods andimplicit-explicit time stepping methods that can be realized efficiently. Finally,

we illustrate our theoretical findings through several numerical experiments in

one and two spatial dimensions that highlight the differences in features of localand nonlocal solutions and also the sharp interface properties of the nonlocal

model.

1. Introduction

The Cahn-Hilliard model was proposed in [14] as the model to describe phaseseparation of a binary alloy. Since then, the model and its variants have been widelyused in different areas of science such as, e.g., tumor growth, image segmentationand copolymer melts, cf. [7, 41, 45, 58, 59, 65].

1 Computer Science and Mathematics Division, Oak Ridge National Laboratory, OneBethel Valley Road, TN 37831, USA

2 Department of Scientific Computing, Florida State University, 400 Dirac Science

Library, Tallahasse, FL 32306-4120, USA, and the Oden Institute for Computer Engi-neering and Sciences, University of Texas at Austin, Austin, TX 78712, USA

E-mail addresses: [email protected], [email protected].

2010 Mathematics Subject Classification. 45K05; 35K55; 35B65; 49J40; 65M60; 65K15.Key words and phrases. nonlocal Cahn-Hilliard model, variational inequality, well-posedness,

regularity, finite elements.This material is based upon work supported by the U.S. Department of Energy, Office of

Advanced Scientific Computing Research, Applied Mathematics Program under the award num-bers ERKJ345 and ERKJE45; and was performed at the Oak Ridge National Laboratory, which

is managed by UT-Battelle, LLC under Contract No. De-AC05-00OR22725. U.S. Governmentretains a non-exclusive, royalty-free license to publish or reproduce the published form of this con-tribution, or allow others to do so, for U.S. Government purposes. MG was also supported by U.S.Department of Energy grant DE-SC0021077 as part of the AEOLUS Multifaceted MathematicsIntegrated Capability Center.

1

arX

iv:2

004.

1437

9v2

[m

ath.

AP]

9 S

ep 2

021

2 O. BURKOVSKA AND M. GUNZBURGER

Given a bounded domain Ω ⊂ Rn, n ≤ 3, and a fixed final time T > 0, themodel is described by the coupled system of equations

∂tu−∇ · (σ(u)∇w) = 0,

w = −ε2∆u+ F ′(u),in (0, T )×Ω, (1.1)

together with appropriate boundary conditions. Here, u is an order parametertaking the values in [−1, 1] and is related to the concentration of a substance, w isa chemical potential, F (u) is a double-well potential, σ(·) is the mobility, and ε > 0is the interface parameter, which is proportional to the thickness of the interface.The double-well potential promotes pure phases and attains its minimum close topure phases, i.e., when u = ±1. Typically, a regular potential is employed in (1.1),given by the fourth-order polynomial

F (u) :=cF4

(u2 − 1)2, cF > 0. (1.2)

Being the simplest choice, the regular potential (1.2) is not physically realistic, sinceit may provide a non-feasible solution |u| > 1. In practice it serves as an approx-imation of the more complex, but physically more relevant, logarithmic potentialfor u ∈ (−1, 1)

F (u) :=θ

2((1 + u) log(1 + u) + (1− u) log(1− u))− cF

2u2, 0 < θ < cF , (1.3)

or the obstacle potential

F (u) :=

F0(u) if |u| ≤ 1

+∞ if |u| > 1

= F0(u) + I[−1,1](u), (1.4)

where I[−1,1] is the convex indicator function of the admissible range [−1, 1] and

F0(u) = cF /2(1− u2), cF > 0. In contrast to (1.2), the logarithmic potential (1.3)and obstacle potential (1.4) always provide a solution within an admissible range|u| ≤ 1. However, the logarithmic potential does not allow u to attain pure phases,i.e., u ∈ (−1, 1), whereas the obstacle potential promotes pure states in the model,i.e., u = 1 or u = −1.

Mathematically, the Cahn-Hilliard model (1.1) can be derived as the H−1-gradient flow of the Ginzburg-Landau energy

E(u) =ε2

2

∫Ω

|∇u|2 dx+

∫Ω

F (u) dx. (1.5)

The gradient square term in (1.5) represents short-range interactions between par-ticles. To account for long-term interactions, one can consider a nonlocal variantof the Cahn-Hilliard problem (1.1):

∂tu−∇ · (σ(u)∇w) = 0,

w = Bu+ F ′(u),in Ω × (0, T ), (1.6)

where B is a nonlocal diffusion operator. In contrast to the derivation of the localcounterpart (1.1), Giacomin and Lebowitz in [46] (see also [47]) provide a rigorousmicroscopic derivation of the nonlocal model (1.6) in the case when Ω ≡ Tn is setto be the n-dimensional torus. More specifically, the macroscopic continuum modelis obtained via a hydrodynamic limit of particle models that are dynamic versionsof lattice gases undergoing long interactions. In this case the nonlocal operatorB = BT is set to

BTu(x) =

∫Tn

(u(x)− u(y))γ(x− y) dy = cγu(x)− (γ ∗ u)(x), ∀x ∈ Tn,

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 3

where γ : Tn → R is an integrable convolution kernel that sets up the law forthe nonlocal interactions. The kernel is assumed to depend only on the distance|x−y|Tn on the torus, the convolution is defined as (γ ∗u)(x) =

∫Tn u(y)γ(x−y) dy,

and cγ =∫Tn γ(y) dy = (γ ∗ 1)(x), which is independent of x. The corresponding

nonlocal free energy functional has the form

E(u) =1

4

∫Tn

∫Tn

(u(x)− u(y))2γ(x− y) dxdy +

∫TnF (u) dx.

We note that the local model can be considered as an approximation of the nonlocalmodel for vanishing nonlocal interactions (cf., e.g., [24, 31] and references citedtherein).

cγu− γ ∗ u ≈ −ε2∆u with ε2 =1

2n

∫Rnγ(ξ)|ξ|2 dξ. (1.7)

An adoption of the Giacomin-Lebowitz model to the case when Ω ⊂ Rn is abounded domain and not a torus is possible by setting B = BR in (1.6), where

BRu(x) =

∫Ω

(u(x)− u(y))γ(x− y) dy = cγ(x)u(x)− (γ ∗ u)(x), ∀x ∈ Ω. (1.8)

Here, the kernel γ : Rn → R is no longer defined on the torus (which corresponds toa periodic function on Rn), but often depends on the distance in Rn, which impliesthat cγ(x) = (γ ∗ 1)(x) =

∫Ωγ(x − y) dy is no longer constant. The operator BR

is often referred to as the regional nonlocal operator, since it restricts the nonlocalinteractions to the region Ω, and has been one of the most frequently analyzed inthe context of nonlocal Cahn-Hilliard models; see, e.g., [6, 18, 20, 24, 42, 43, 44]and the references cited therein. While most of these works address integrablekernels, which is also a subject of the current work, singular kernels have also beeninvestigated; see, e.g., [19, 23]. For an overview of early and recent references andextensions of the local and nonlocal Cahn-Hilliard model, we refer interested readersto [3, 57, 29]. In terms of the notation, the present modification has only minordifferences to (1.6). However, it has direct implications on the properties of thesolutions, as will become apparent shortly.

In this work, we are interested in the model (1.6) posed on a bounded domain Ωwith cγ being constant throughout Ω. For this, we consider B = BN to be thefollowing nonlocal operator:

BNu(x) =

∫Ω∪ΩI

(u(x)− u(y))γ(x− y) dy = cγu(x)− (γ ∗ u)(x), ∀x ∈ Ω, (1.9)

where cγ =∫Ω∪ΩI γ(x− y) dy for all x ∈ Ω and (γ ∗u)(x) =

∫Ω∪ΩI u(y)γ(x− y) dy.

Here, ΩI is to as the nonlocal interaction domain, that represents the extent ofnonlocal interactions outside of Ω, and is defined as

ΩI := y ∈ Rn \Ω : γ(x, y) 6= 0, x ∈ Ω.To define (1.9), one must know the values of u also on the interaction domain ΩI .Usually, the local Cahn–Hilliard model (1.1) is complemented with the homoge-neous Neumann boundary condition (to guarantee the mass-conservation property),and in the nonlocal setting considered here we impose a homogeneous nonlocal fluxcondition for u on ΩI , which is analogous of the Neumann boundary conditions inthe local setting:

Nu(x) =

∫Ω∪ΩI

(u(x)− u(y))γ(x− y) dy = 0 ∀x ∈ ΩI . (1.10)

The type of the nonlocal operator BN (1.9) has gained significant attention recentlyin the context of various applications, cf. [38, 30, 28] and references cited therein.In contrast to (1.8), the nonlocal interactions are not limited to Ω and occur in all

4 O. BURKOVSKA AND M. GUNZBURGER

of Ω ∪ ΩI . If the kernel has infinite support then Ω ∪ ΩI ≡ Rn, and the nonlocalinteractions occur on the whole of Rn.

1.1. Model setting. Finally, we present the model that encompasses both formu-

lations. We denote Ω to be either Ω or Ω ∪ ΩI , and we employ a non-smoothobstacle potential (1.4), and, for simplicity, consider a constant mobility σ(u) ≡ 1.While the problem formulation (1.6) is appropriate for smooth potentials, the non-smooth obstacle potential (1.4) requires us to introduce the concept of subdifferen-tials to define the derivative of I[−1,1]. In this case, F ′(u) should be replaced by ageneralized differential of F , ∂F (u) = −cFu + ∂I[−1,1](u), where ∂I[−1,1](u) is thesubdifferential of the indicator function,

∂I[−1,1](u) =

(−∞, 0] if u = −1,

0 for u ∈ (−1, 1),

[0,+∞) if u = 1.

(1.11)

Finally, the nonlocal model we are interested to study becomes∂tu−∆w = 0,

w = ξu− γ ∗ u+ λ, λ ∈ ∂I[−1,1](u),(1.12)

where ξ(x) :=∫Ωγ(x − y) dy − cF and γ ∗ u =

∫Ωu(y)γ(x − y) dy. Here, we

will already impose that ξ(x) ≥ 0 on Ω, which is required to obtain a well-posedproblem. We will show that ξ is an appropriate interface parameter in the nonlocalmodel. We complement (1.12) with the local and nonlocal flux conditions

∂w

∂n= 0 on ∂Ω, and Nu(x) =

∫Ω

(u(x)− u(y))γ(x− y) dy = 0 on Ω \Ω.

(1.13)We note that, in the variational form the problem (1.12) can be restated as acoupled system of equations involving variational inequalities. The correspondingnonlocal free energy is given as

E(u) =1

4

∫Ω

∫Ω

(u(x)− u(y))2γ(x− y) dxdy +

∫Ω

F (u) dx (1.14)

where F is defined in (1.4).

Sharp interfaces. The choice of the obstacle potential in (1.12) is not only physicallymotivated, but brings an interesting insight into the mathematical properties of thesolution, such as the occurrence of sharp interfaces. In particular, as one of the maincontributions of this work, we demonstrate that the solution of (1.12) is discontinu-ous and can admit sharp interfaces for some non-trivial and non-vanishing nonlocalinteractions. On the contrary, the local problem (1.1), which is a diffuse interfacemodel, does not allow the appearance of sharp interfaces, other than in the limitingcase ε → 0 which corresponds to vanishing local interactions in (1.5). In Figure 1we illustrate the appearance of sharp interfaces for the local and nonlocal solutionsof (1.1) and (1.12) with the obstacle potential (1.4). Here, we compare the solutionsobtained with a nonlocal model (1.12) for ξ ≈ 0 to the local model (1.1), wherewe have replaced BN with the Neumann Laplacian −ε2∆ according to (1.7). Wepoint out that, in contrast to the local model, we do not need to perform the limitto obtain sharp interfaces in the nonlocal solution. Moreover, the sharp interfacecase ξ = 0 for the nonlocal model is well-posed and can be solved numerically withthe same time-dsicretization as for ξ > 0, as we will discuss below.

Mathematically, the appearance of sharp interfaces in the nonlocal model (1.12)is explained by realizing that the solution of (1.12) for ξ > 0 can be obtained as a

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 5

(a) comparison in 1D (b) local solution (2D) (c) nonlocal solution (2D)

Figure 1. Comparison of the local and nonlocal solutions in one-and two-dimensions. The local and nonlocal solutions are initial-ized with the same initial condition, cF , and plotted at the sametime T . For the nonlocal model we consider the nonlocal opera-tor (1.9) with Gaussian type kernel (5.3). The interface parameterε is given by (1.7), and with this choice the local model correspondsto the local limit of the nonlocal model for vanishing radius of non-local interactions.

pointwise projection of g = w + γ ∗ u onto the admissible set,

u(t) = P[−1,1]

1

ξg(t)

in Ω × (0, T ), ξ > 0.

Note that, for ξ = 0, the above turns into (see also Theorem 4.3)

u(t) ∈ signg(t)=

1 if g(t) > 0,

−1 if g(t) < 0,

[−1, 1] if g(t) = 0,

in Ω × (0, T ).

If the level set of g = 0 is of zero measure, this implies that the solution can admitonly pure phases u = 1 or u = −1. It becomes apparent that ξ is not constant forthe case of the regional operator (1.8), since cγ is variable in Ω. Hence, in generalthe corresponding solution does not posses sharp interfaces, unless the obstaclepotential is equipped with a spatially-dependent cF (x).

1.2. Contribution and related works. One of the contributions of this work isto propose a well-posed nonlocal model (1.12) that admits discontinuous solutionswith only pure phases, as described in the previous section. To the best of ourknowledge this is the first result in this direction. However, we note that the jump-discontinuous behaviour of the solution closely resembles the bang-bang controlprinciple, where almost all values of the control function lie on the boundary of theadmissible set; see, e.g., [64]. Indeed, a single time-step of the model (1.12) can bealso understood as a variational inequality, which serves as a necessary optimalitycondition for a specific bang-bang optimal control problem.

Moreover, discontinuous solutions for related Allen-Cahn phase transition modelshave already been addressed in, e.g., [5, 40, 4, 15, 17]; see also [39] and the referencescited therein. They are based on the same Ginzburg-Landau energy as in (1.14)and described by the evolution equation

∂tu+ cγu− γ ∗ u+ F ′(u) = 0. (1.15)

In more recent works [35, 32, 36], the authors also investigated numerically andtheoretically the relation between kernel parameters with finite range of nonlocalinteractions and the appearance of discontinuities in the steady state solution of the

6 O. BURKOVSKA AND M. GUNZBURGER

Allen-Cahn equation. In [35] it was reported that the condition under which thesesolutions may admit discontinuities is related to the same parameter ξ = cγ − cF ,defined for a smooth potential F as in (1.2) with cF = −minu∈[−1,1] F

′′(u) = 1.More specifically, for a regular potential these discontinuities can occur for ξ < 0,which is only possible for the Allen-Cahn equation. However, even though thesolution contains a discontinuity, it does not admit only pure phases. That is, udoes not take values only in the set −1, 1 and it does not jump from 1 to −1 ata single point. On the other hand, using an obstacle potential in the Allen-Cahnsetting can permit exclusively pure phases in the steady-state solutions, u = ±1.However, the evolution law (1.15) still does not allow for the discontinuities tomigrate over time; cf. [40, 39]. This is due to the fact that the time derivative of utakes values in L2(Ω) for the evolution law (1.15). In contrast, we prove that thesolution of the nonlocal Cahn-Hilliard model with the obstacle potential (1.12) canadmit jump-discontinuities with only pure phases in the solution profile not onlyat the steady-state but also during the whole time evolution.

In addition to studying the properties of the solution, we provide a detailedanalysis of the nonlocal problem (1.12). First, we establish general well-posednessresults for an appropriate time-discrete formulation of (1.12) (see Theorem 3.4–3.5), where only integrability is required for the kernel. Moreover, we show thatunder certain conditions the solution is given by an L2-projection on the admissibleset. This representation allows us to deduce improved regularity properties ofthe solution (see Theorem 3.7) and the appearance of sharp-interfaces for ξ = 0(see Theorem 3.8). We derive a convergence result towards a continuous in timeformulation. As a consequence we obtain well-posedness of the time continuousproblem (see Theorem 4.1–4.2), including an existence result for ξ = 0. The latteris noteworthy since most of the available results focus on the settings equivalent toξ(x) ≥ ξmin > 0, which does not cover the more delicate case of sharp interfaces,cf. [24, 43]. One of the works we are aware of that covers the case ξ = 0 is [20], wherea model similar to (1.12) is studied in an abstract setting. However, there well-posedness is derived in the time-continuous framework by performing a limit of anappropriate regularized problem. In contrast, we do not regularize the non-smoothpotential, but exploit its limited regularity to characterize the fine properties of thesolution, including sharp interfaces. For the analysis, we rely solely on the timediscretization, which also serves as a basis for the numerical realization introducedhere, including the case ξ = 0.

In addition to the complete analysis of the model, we also discuss efficientspace and time approximations of the nonlocal solution based on the finite elementmethod which we complement with corresponding numerical results. Numericalanalysis of the local Cahn-Hilliard problem has been an active subject of research;see, e.g., [8, 9, 10, 11, 12]. The approximation of the nonlocal diffuse interfaceproblem has been addressed in various works, e.g., [1, 2, 31, 32]. A design of ap-propriate time inegration schemes have been investigated in [52, 51, 50, 31, 32].The aforementioned works focused mainly on smooth potentials. An inclusion ofthe obstacle potential in the model leads to a non-smooth and nonlinear systeminvolving variational inequalities. The solvability of such a system can be realizedby the semi-smooth Newton method [53] which has been addressed in the localCahn-Hilliard setting in, e.g., [8, 54], and which we adopt here in the nonlocal set-ting. Lastly, due to the complex structure of the Cahn-Hilliard system, in general,there is great demand for the development of efficient approximation techniquesbased on, e.g., Krylov solvers [11] or model order reduction methods [48, 49].

The paper is organized as follows. In Section 2 we introduce appropriate func-tional settings and the variational formulation of the problem we consider. Section 3

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 7

is devoted to the well-posedness analysis of the semi-discrete problem, improved reg-ularity, and the derivation of the sharp-interface condition for the correspondingsolution. The respective results for the time-continuous problem formulation aregiven in Section 4. Then, in Section 5 we discuss finite-element approximationstogether with time-marching schemes for the nonlocal problem and provide variousnumerical examples. Finally, we conclude the paper in Section 6.

2. Preliminaries

We denote by Lp(Ω), p ∈ [1,∞], the usual Lebesgue spaces and by W k,p(Ω),k ∈ N, 1 ≤ p ≤ ∞, we mean the Sobolev space of all functions in Lp(Ω) having alldistributional derivatives up to order k and endowed with the norm ‖u‖pWp,k(Ω) =∑|α|≤k ‖Dαu‖pLp(Ω). We denote by Lp(0, T ;X), 1 ≤ p ≤ ∞ the Bochner space of

all measurable functions u : [0, T ]→ X, for which the norms

‖u‖Lp(0,T ;X) =

(∫ T

0

‖u(t)‖pX dt

)1/p

, p <∞, and ‖u‖L∞(0,T :X) = ess supt∈(0,T )

‖u(t)‖X

are finite. If not specified, we often use the relation a ≤ Cb, with some genericconstant C > 0 that may be different in every instance, but is independent of aand b.

2.1. Nonlocal operators. We introduce a kernel function γ : Rn → R+, which isradial, has finite support and integrable, i.e.,

γ(x) = γ(|x|), γ ∈ L1(Rn),

with supp(γ) ⊂ (0, δ], δ > 0,

and there exists σ > 0, such that (0, σ) ⊂ supp(γ).

(H1)

While most of the results in this paper are valid under the above conditions, incertain cases, we will need to require a higher regularity of the kernel, i.e.,

γ ∈W 1,1(Rn). (H2)

In the following we also employ (by a slight abuse of notation) the two-point versionof γ given by

γ(x, y) = γ(x− y), ∀x, y ∈ Rn,

which will be used as an integration kernel to define the nonlocal operator B.We define the nonlocal operator B that encompasses the definitions of the “Neu-

mann” (Case 1) (1.9) and “regional” (Case 2) (1.8) type of nonlocal operators

Bu(x) =

∫Ω

(u(x)− u(y))γ(x, y) dy = cγ(x)u(x)− (γ ∗ u)(x), (2.1)

where Ω = Ω (“regional”) or Ω = Ω ∪ ΩI (“Neumann”). For convenience of

notation we suppose that u is extended by zero outside of Ω, and (γ ∗u)(x) denotes

the convolution on Ω of γ with u, and we define

cγ(x) :=

∫Ω

γ(x, y) dy, Cγ :=

∫Rnγ(x, y) dy, and Cγ := ‖∇γ‖L1(Rn), (2.2)

and note that 0 ≤ cγ(x) ≤ Cγ < ∞ for all x ∈ Ω. Since γ has finite support, Ω is

a bounded domain, and for Ω = Ω ∪ΩI , cγ is a constant for all x ∈ Ω and can be

computed explicitly as cγ = 2πn/2

Γ (n/2)

∫ δ0|ξ|n−1γ(|ξ|) dξ.

8 O. BURKOVSKA AND M. GUNZBURGER

Nonlocal Neumann-type boundary conditions. The problem with the “Neu-

mann” type nonlocal operator for which Ω = Ω ∪ ΩI and B = BN (see (1.9))implies that there is a nonlocal flux from ΩI to Ω prescribed in the model. Froma probabilistic perspective in the context of particle-based models such boundaryconditions provide information about what happens to a particle upon leaving adomain Ω, whereas u(x, t) in this case represents the probability density functionof the position of a particle subject to a jump-diffusion process. Such boundaryconditions have been proposed in [30]. A characterization of this condition in termsof Levy flights for unsteady fractional Laplace equation has been studied in [26].Additional discussions about Neumann type boundary conditions and their varia-tion in the nonlocal setting can be found in, e.g., [21, 22, 56, 27, 61, 25, 66]; seealso [38] and the references therein. In contrast, for the “regional” nonlocal oper-

ator B = BR (1.8) we have that Ω = Ω and Ω \ Ω = ∅, which implies that thereis no nonlocal interactions between Ω and Rn \ Ω, and, hence, no nonlocal fluxcondition such as (1.10) needs to be imposed. In this case, from the probabilisticpoint of view the random process of the movement of the particle occurs solelywhithin Ω, and while in the “Neumann” case the particle is not allowed to leaveΩ ∪ ΩI , here, the particle is not allowed to jump out of Ω; this is referred to as acensored process.

Function spaces. We set VA := H1(Ω), which is endowed with the usual H1-

norm, ‖v‖2VA := ‖v‖2H1(Ω) = |v|2H1(Ω)+‖v‖2L2(Ω), and |v|VA := |v|H1(Ω) = ‖∇v‖L2(Ω).

We also define the space of mean-free functions

VA0:= v ∈ VA : (v, 1)L2(Ω) = 0,

endowed with the same semi-norm as on VA. Thanks to the local Poincare inequality

‖u‖L2(Ω) ≤ CP |u|VA , ∀u ∈ VA0 , CP > 0, (2.3)

the seminorm on VA0also defines a norm. We denote by V ′A = (H1(Ω))′ the dual

space of VA, and by 〈·, ·〉V ′A×VA a corresponding duality pairing, and for f ∈ V ′A, we

denote

‖f‖V ′A := supv∈VA‖v‖VA 6=0

〈f, v〉V ′A×VA‖v‖VA

.

We introduce the dual space of VA0, denoted by V ′A0

= (VA0)′, which consists of

functionals with mean-value zero:

V ′A0:=f ∈ V ′A : 〈f, 1〉V ′A×VA = 0

.

We note that for any f ∈ V ′A0, we have that ‖f‖V ′A0

= ‖f‖V ′A . Indeed, for v ∈ VA,

such that ‖v‖VA 6= 0, we obtain

‖f‖V ′A = supv∈VA

|〈f, v〉V ′A×VA |‖v‖VA

= supv∈VA0

, M

|〈f, v +M〉V ′A×VA |

(‖v‖2VA +M2)1/2= supv∈VA0

|〈f, v〉V ′A×VA |‖v‖VA

.

Let G denote a Green’s operator for the inverse of the Laplacian with zero Neu-mann boundary conditions. That is, given f ∈ V ′A, we define Gf ∈ VA0

as theunique solution of

(∇Gf,∇v)L2(Ω) = 〈f, v〉V ′A×VA , ∀v ∈ VA0. (2.4)

We note that, if f ∈ V ′A0, then (2.4) also holds true for all v ∈ VA:

(∇Gf,∇v)L2(Ω) = 〈f, v〉V ′A×VA , ∀v ∈ VA.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 9

The existence and uniqueness of Gf follows from the Poincare inequality and theLax-Milgram lemma, and it holds that

‖f‖2V ′A = |Gf |2VA = ‖∇Gf‖2L2(Ω) = 〈f,Gf〉V ′A×VA . (2.5)

If f ∈ V ′A ∩ L2(Ω), then ‖f‖2V ′A = (Gf, f)L2(Ω). Moreover, owing to (2.5), it also

holds that

〈f ′(t),Gf(t)〉V ′A×VA =1

2

d

dt‖f(t)‖2V ′A , for a.e. t ∈ (0, T ), ∀f ∈ H1(0, T ;V ′A).

(2.6)For the nonlocal space VB we distinguish between two cases:

VB :=

v ∈ L2(Ω ∪ΩI) : N v = 0 on ΩI

, for Case 1 ,

L2(Ω), for Case 2 ,

where an inclusion of the nonlocal flux conditions in the function space setting inCase 1 is important as it will become apparent shortly. We recall the generalizednonlocal Green’s first identity [30]:

(Bu, v)L2(Ω) =

∫Ω

v(x)

∫Ω

(u(x)− u(y))γ(x, y) dy

=1

2

∫Ω

∫Ω

(u(x)− u(y))(v(x)− v(y))γ(x, y) dy dx+

∫Ω\Ω

v(x)Nu(x) dx. (2.7)

Next, we define the inner product and semi-norm on VB :

(u, v)VB :=1

2

∫Ω

∫Ω

(u(x)− u(y))(v(x)− v(y))γ(x, y) dxdy, |v|2VB := (v, v)VB ,

and endow VB with the norm ‖v‖2VB := |v|2VB + ‖v‖2L2(Ω). We also define a bilinear

form:b(u, v) := (u, v)VB . (2.8)

Proposition 2.1. For all u, v ∈ VB we obtain that

b(u, v) = (cγu, v)L2(Ω) − (γ ∗ u, v)L2(Ω), (2.9)

where cγ is defined in (2.2).

Proof. The proof follows directly from (2.9), (2.1) and an identity (2.7).

We recall the Young’s inequality for products:

ab ≤ 1

2εa2 +

ε

2b2, ∀a, b ≥ 0, ε > 0, (2.10)

and also make use of the Young’s inequality for the convolution: letting f ∈ Lp(Rn)and g ∈ Lq(Rn) with 1 ≤ p ≤ ∞, 1 ≤ q ≤ ∞, and 1/r = 1/p + 1/q − 1 ≥ 0, thenf ∗ g ∈ Lr(Rn) and

‖f ∗ g‖Lr(Rn) ≤ ‖f‖Lp(Rn)‖g‖Lq(Rn). (2.11)

2.2. Exterior problem. We note that for Case 1 the nonlocal Neumann typevolume constraints could be considered as an “exterior” nonlocal problem posedon ΩI with non-homogeneous Dirichlet-type volume constraints imposed on Ω. Inparticular, we consider the problem

Nu(x) = 0 on ΩI ,

u = g on Ω.(2.12)

We define a nonlocal space V 0B incorporating homogeneous volume constraints as

V 0B :=

v ∈ L2(Ω ∪ΩI) : b(v, v) <∞, v = 0 on Ω

, |v|2V 0

B:= b(v, v),

10 O. BURKOVSKA AND M. GUNZBURGER

where b(·, ·) is defined as in (2.8) with Ω = Ω ∪ ΩI . For u := u − g ∈ V 0B , where

g is extended by zero outside of Ω, and, by a slight abuse of notation, denotedby the same symbol, we consider the corresponding weak formulation of the aboveproblem:

b(u, v) = −(N g, v)L2(ΩI) ∀v ∈ V 0B . (2.13)

We recall a nonlocal Poincare-type inequality (see, e.g., [55, Proposition 1]):

‖u‖L2(ΩI)≤ CP |u|V 0

B, ∀u ∈ V 0

B , (2.14)

with CP = CP (γ, Ω), and note that b(·, ·) is continuous and coercive on V 0B

b(u, u) ≥ C−2P ‖u‖2L2(ΩI)

, |b(u, v)| ≤ 4Cγ‖u‖L2(ΩI)‖v‖L2(ΩI)

, ∀u, v ∈ V 0B .

Then, by the Lax-Milgram argument there exist a unique solution of (2.13) with

‖u‖L2(ΩI)= ‖u‖L2(ΩI)

≤ C2PCγ‖g‖L2(Ω), (2.15)

where the last inequality above has been obtained by using the definition ofN , (1.13),Young’s inequality and the fact that g = 0 on Rn \Ω. Next, we show that, for bothcases Case 1 and Case 2, VB ∼= L2(Ω).

Proposition 2.2. The space(VB , ‖·‖VB

)is a Hilbert space that is equivalent to L2(Ω).

Proof. To show that VB is equivalent to L2(Ω) (which immediately implies thatVB is a Hilbert space) it is suffices to show the norm equivalence between thesespaces. First, it is clear that ‖v‖L2(Ω) ≤ ‖v‖VB . On the other hand, for all u ∈ VBusing (2.9) and Young’s inequality (2.11) we obtain

‖u‖2VB = |u|VB + ‖u‖L2(Ω) = (cγu, u)L2(Ω) − (γ ∗ u, u)L2(Ω) + ‖u‖2L2(Ω)

≤ (Cγ + 1)‖u‖2L2(Ω) + Cγ‖u‖L2(Ω)‖u‖L2(Ω)

≤(

(2Cγ + 1)‖u‖L2(Ω) + Cγ‖u‖L2(Ω\Ω)

)‖u‖L2(Ω) ≤ C‖u‖VB‖u‖L2(Ω),

where in the last inequality we exploit the fact that for Case 2, Ω \Ω = ∅, and forCase 1, u ∈ VB is a solution of the exterior problem (2.12) and (2.15) holds true.Then, dividing by ‖u‖VB concludes the proof.

We denote by V ′B the dual space of VB , V ′B = (VB)′. Clearly, with the previousresult at hand, we obtain that the embedding

VB → L2(Ω) → V ′B

forms a Gelfand triple, i.e., it is continuous and dense.

2.3. Variational formulation. Now, we are ready to present a weak formula-tion of the problem (1.12). To incorporate inequality constraints |u| ≤ 1 in thevariational formulation, we introduce the set

K := v ∈ VB : |v| ≤ 1 a. e. in Ω, (2.16)

and the setKm := v ∈ K : (v, 1)L2(Ω) = m of functions with massm ∈ (−|Ω|, |Ω|),where |Ω| is the volume of Ω. We also define a positive cone

M = v ∈ V ′B : v ≥ 0 a. e. on Ω.

We introduce space-time cylinders Q := Ω × (0, T ) and Q := Ω × (0, T ), T > 0.Then, we consider the following problem: Find u(t) ∈ K, w(t) ∈ VA such that⟨

∂u(t)

∂t, φ

⟩V ′A×VA

+ (∇w(t),∇φ)L2(Ω) = 0, ∀φ ∈ VA,

b(u(t), ψ − u(t)) + (F ′0(u(t))− w(t), ψ − u(t))L2(Ω) ≥ 0, ∀ψ ∈ K,(P)

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 11

subject to u(0) = u0 ∈ Km. In many instances, it is convenient to work withan equivalent formulation, where we eliminate the first equation in (P) by set-ting w(t) = −G(∂tu(t)) − µ(t), where G is a Green’s operator defined in (2.4)and µ(t) = − 1

|Ω|∫Ωw(t) dx ∈ R is the negative mean-value of w(t). In addition,

to incorporate the inequality constraints (2.16), we introduce additional Lagrangemultipliers λ±(t) ∈ L2(Ω). Then, we arrive at the following problem having a sad-dle point structure: Find u(t) ∈ VB , λ(t) ∈ V ′B , λ(t) := λ+(t)− λ−(t), λ±(t) ∈M ,µ(t) ∈ R such that

(G(∂tu(t)) +Bu(t) + F ′0(u(t)) + λ(t) + µ(t), ψ)L2(Ω) = 0, ∀ψ ∈ VB ,(η − λ±(t), 1∓ u)L2(Ω) ≥ 0, ∀η ∈M.

(2.17)

We note that the inequalities in (2.17), which refer to complementarity conditions,can be simplified further to

(η, u(t)± 1)L2(Ω) ≥ 0, (λ±(t), u(t)∓ 1)L2(Ω) = 0, ∀η ∈M. (2.18)

The above equations are obtained by simply taking η = 2λ±(t) and η = 0 in (2.17).We note that the problem (2.17) is also equivalent to its time-integrated version:

Find u ∈ L2(0, T ;K) such that

(G(∂tu) +Bu+ F ′0(u) + λ+ µ, ψ)L2(Q) = 0, ∀ψ ∈ L2(0, T ;VB),

(η − λ±, 1∓ u)L2(Q) ≤ 0, ∀η ∈ L2(0, T ;M),(2.19)

where L2(0, T ;M) := v ∈ L2(0, T ;V ′B) : v ≥ 0 a. e. in Q and L2(0, T ;K) := v ∈L2(0, T ;VB) : |v| ≤ 1 a. e. in Q.

3. Semi-discrete problem: Existence, uniqueness and a sharpinterface condition

In this section we present the semi-discretization of the problem (P) and establishexistence, uniqueness and the main properties of the solution. In the course of ouranalysis we introduce and make use of different equivalent characterizations of theproblem.

For T > 0 and K ∈ N, we define τ = T/K and uk := u(tk), wk := w(tk) fork = 1, . . . ,K. Following [10] we consider an implicit time stepping scheme. Givenu0 ∈ Km, we consider for k = 1, . . . ,K the following problem: Find (uk, wk) ∈K × VA such that

1

τ(uk − uk−1, φ)L2(Ω) + (∇wk,∇φ)L2(Ω) = 0, ∀φ ∈ VA,

b(uk, ψ − uk) + (F ′0(uk)− wk, ψ − uk)L2(Ω) ≥ 0, ∀ψ ∈ K.(Pk)

We point out that the above scheme is mass conserving. Indeed, taking φ = 1in (Pk) we obtain that (uk−1, 1) = (uk, 1) = m.

If uk ∈ Km is a solution of (Pk), then setting wk = −G(1/τ(uk − uk−1)) − µk,µk ∈ R, and restricting the test functions to ψ ∈ Km, we can restate the coupledsystem (Pk) as a single variational inequality: Given u0 ∈ Km, find uk ∈ Km suchthat for k = 1, . . . ,K it holds that

1

τ

(G(uk − uk−1

), ψ − uk

)L2(Ω)

+b(uk, ψ−uk)+(F ′0(uk), ψ−uk)L2(Ω) ≥ 0, ∀ψ ∈ Km.(Qk)

Furthermore, the above problem (Qk) can be equivalently posed as a system ofcomplementarity conditions: For a given u0 ∈ Km, find (uk, λk±, µ

k) ∈ VB ×M ×R

12 O. BURKOVSKA AND M. GUNZBURGER

such that for k = 1, . . . ,K, it holds that(1

τG(uk − uk−1) +Buk + F ′0(uk) + λk + µk, ψ

)L2(Ω)

= 0, ∀ψ ∈ VB ,

(η − λk±, 1∓ uk)L2(Ω) ≥ 0, ∀η ∈M.

(3.1)

The variational inequality has a direct connection to a minimization problem.

Minimization problem. For u ∈ VB we define the functional

Jk(u) :=1

2|u|2VB +

∫Ω

F0(u) dx+1

∥∥∇G(u− uk−1)∥∥2L2(Ω)

=1

2(ξu, u)L2(Ω) −

1

2(γ ∗ u, u)L2(Ω) +

cF2|Ω|+ 1

∥∥u− uk−1∥∥2V ′A,

where ξ(x) := cγ(x)−cF . Then, we consider the following constrained minimizationproblem:

minu∈K

Jk(u) subject to

∫Ω

udx = m, m ∈ (−|Ω|, |Ω|). (J k)

Proposition 3.1. If uk ∈ Km is a solution of (J k), then it solves the variationalinequality (Qk). Conversely, if uk ∈ Km is a solution of (Qk) and the conditionson the time step τ given in Theorem 3.5 hold true, then uk ∈ Km is also a solutionto (J k).

Proof. Since Jk is convex and Gateaux differentiable in VB (see Theorem 3.5), itfollows that the variational inequality (Qk) is a necessary and sufficient first-orderoptimality condition for the minimization problem (J k); see [64, Lemma 2.21].

Next, we state a useful property for the function belonging to the set K, whichwill be used to prove the existence result in the next section.

Proposition 3.2 (Maximum principle). Let u ∈ K, then it follows that |u| ≤ 1 in

Ω, and hence u ∈ L∞(Ω).

Proof. The proof follows the same lines as in the proof of Theorem 4.4 in [13].

3.1. Existence and uniqueness. To show the existence and uniqueness of thesemi-discrete problem (Pk), we first discuss the existence and uniqueness of the so-lution of (Qk), and then demonstrate the equivalence of problem formulations (Pk)and (Qk).

We collect in Proposition 3.3 some properties of the kernel which will be usefullater.

Proposition 3.3. For the kernel γ satisfying (H1) the following properties holdtrue.

(i) For η > 0 there exists a family of functions γη : Rn → R+ satisfying (H1)–(H2), such that

‖∇γη‖L1(Rn) ≤ Cη, and ‖γ − γη‖L1(Rn) ≤ η, Cη > 0. (3.2)

(ii) The sequence γη ∗ u → γ ∗ u converges uniformly in L∞(Ω) for any u ∈L∞(Ω), and the limiting function is continuous

γ ∗ u ∈ C(Ω), ∀u ∈ L∞(Ω). (3.3)

(iii) If, in addition, γ satisfies (H2), then

γ ∗ u ∈W 1,∞(Rn), ∀u ∈ L∞(Rn). (3.4)

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 13

Proof. The conditions (i) and (iii) follow directly by density argument and Young’sinequality (2.11), respectively. Then, using the properties (i), (iii), and the factthat γ ∗ u = γη ∗ u − (γ − γη) ∗ u, we obtain that γη ∗ u ∈ W 1,∞(Rn) for anyu ∈ L∞(Rn), and

‖γ ∗ u− γη ∗ u‖L∞(Ω) ≤ ‖γ − γη‖L1(Rn)‖u‖L∞(Ω) ≤ η‖u‖L∞(Ω).

Now, letting η → 0 we obtain (ii) and conclude the proof.

Next, we establish the existence and uniqueness of the solution of the semi-discrete problem (Qk).

Theorem 3.4 (Existence). Let γ satisfy (H1) and let ξ(x) := cγ(x) − cF ≥ 0 forall x ∈ Ω, then there exists a solution of (Qk).

Proof. To show the existence of the solution of (Qk), it is sufficient to show anexistence result for the solution of (J k), which can also be equivalently writtenas minu∈Km Jk(u). Since u ∈ Km, and Km 6= ∅, it follows that u ∈ L∞(Ω) and

‖u‖2L2(Ω) ≤ |Ω|. Then, we have that Jk(u) ≥ 0 for all u ∈ Km, and there exists

Jk ≥ 0 that realizes Jk = infu∈K Jk(u). Hence, there exists a minimizing sequenceunn∈N ⊂ Km for Jk(u) such that

Jk(un)→ Jk := infu∈Km

Jk(u), n→∞.

Since un ⊂ Km, it follows by Proposition 3.2 that un ∈ L∞(Ω). Then, by theBanach-Alaoglu theorem, we can extract a weak*-convergent subsequence unjnj∈N ⊂L∞(Ω):

unj∗ u, j →∞.

Since Km is convex and closed, it is weakly closed and u ∈ Km. Next, we are goingto pass to the limit in Jk(unj ). Since ξ ≥ 0 and using expression (2.9), we can write

Jk(unj ) =1

2‖√ξunj‖2L2(Ω) −

1

2(γ ∗ unj , unj )L2(Ω) +

cF2|Ω|+ 1

∥∥unj − uk−1∥∥2V ′A ,(3.5)

where the first and the last terms in the expression above are continuous and convex,and, hence, weakly lower semi-continuous, i.e.,

‖√ξu‖2L2(Ω) ≤ lim

j→∞inf ‖

√ξunj‖2L2(Ω),

∥∥u− uk−1∥∥2V ′A≤ limj→∞

inf∥∥unj − uk−1∥∥2V ′A .

Next, we consider the second term in (3.5). Let φnj := γ ∗ unj and φ := γ ∗ u.

Since, u, unj ∈ L∞(Ω), by Proposition 3.3 it follows that φnj , φ ∈ C(Ω). Then,

using the fact that unj∗ u, j →∞, and γ ∈ L1(Rn), we obtain that∫

Ω

unj (y)γ(|x− y|) dy →∫Ω

u(y)γ(|x− y|) dy, j →∞,

that is, φnj (x) → φ(x), ∀x ∈ Ω. Then, by the Lebesgue dominated convergencetheorem, we obtain∫

Ω

unjφnj dx =

∫Ω

unjφ dx+

∫Ω

unj (φnj − φ) dx→∫Ω

uφ dx, j →∞.

Finally, taking the limit in Jk(unj ) we obtain

limj→∞

inf Jk(unj ) ≥1

2‖√ξu‖2L2(Ω)−(u, γ ∗ u)L2(Ω) +

cF2|Ω|+

∥∥u− uk−1∥∥2V ′A

= Jk(u).

That is,

Jk(u) ≤ limj→∞

inf Jk(unj ) = limn→∞

Jk(un) = infu∈Km

Jk(u) = Jk.

14 O. BURKOVSKA AND M. GUNZBURGER

On the other hand, by definition of Jk, it follows that Jk ≤ Jk(u), and, hence

Jk = Jk(u).

Theorem 3.5 (Uniqueness). Let γ satisfy (H1) and ξ(x) := cγ(x) − cF > 0 for

all x ∈ Ω. Then, the solution of (Qk) is unique for τ < 4C−2η (ξ/(1 + C4PC

2γ) −

η) (Case 1), and τ < 4(ξ − η)/C2η (Case 2), where Cη, 0 < η < ξ are as in

Proposition 3.3.

Proof. We prove it by contradiction. Let uk1 , uk2 ∈ Km be two solutions of (Qk),and let θk := uk1 − uk2 ∈ K0. Taking an addition of (Qk) tested with ψ = uk2 , whenuk1 is a solution, and vice-versa, and using (2.9) we obtain

0 ≥ b(θk, θk)− cF ‖θk‖2L2(Ω) +1

τ(G(θk), θk)L2(Ω)

= (ξθk, θk)L2(Ω) − (γ ∗ θk, θk)L2(Ω) +1

τ

∥∥θk∥∥2V ′A0

.

Invoking Proposition 3.3 and (2.11), we can estimate the second term above∣∣(γ ∗ θk, θk)L2(Ω)

∣∣ ≤ ∣∣∣⟨γη ∗ θk, θk⟩VA×V ′A∣∣∣+∣∣∣((γ − γη) ∗ θk, θk

)L2(Ω)

∣∣∣≤∥∥∇γη ∗ θk∥∥L2(Ω)

∥∥θk∥∥V ′A

+ ‖γ − γη‖L1(Rn)∥∥θk∥∥2

L2(Ω)

≤ Cη∥∥θk∥∥

L2(Ω)

∥∥θk∥∥V ′A

+ η∥∥θk∥∥2

L2(Ω)

≤(C2ητ/4 + η

) ∥∥θk∥∥2L2(Ω)

+1

τ

∥∥θk∥∥2V ′A,

where the last estimate is obtained by using Young’s inequality (2.10) with ε =2/(Cητ). Since θk ∈ K0, we have that

∥∥θk∥∥V ′A

=∥∥θk∥∥

V ′A0

. By combining the

previous estimates we obtain

(ξθk, θk)L2(Ω) − (C2ητ/4 + η)

∥∥θk∥∥2L2(Ω)

≤ 0. (3.6)

For Case 2, the above simplifies to ((ξ − C2ητ/4 − η)θk, θk)L2(Ω) ≤ 0, and, since

τ < 4(ξ− η)/C2η , it follows that θk = 0, and hence uk1 = uk2 . For Case 1, we denote

θkΩ := θk|Ω , and using the fact that θk ∈ VB , we obtain that θk solves the exteriorproblem (2.13) on ΩI with g = θkΩ on Ω, and from (2.15), we get∥∥θk∥∥

L2(ΩI)≤ C2

PCγ∥∥θkΩ∥∥L2(Ω)

= C2PCγ

∥∥θk∥∥L2(Ω)

.

Then, the above estimate together with (3.6) brings us to((ξ − (C2

ητ/4 + η)(1 + C4PC

2γ))θk, θk

)L2(Ω)

≤ 0,

which implies for τ < 4C−2η (ξ/(1+C4PC

2γ)−η), that θk = 0, and hence uk1 = uk2 .

Corollary 3.1. If under the assumptions of Theorem 3.5, the kernel γ additionallysatisfies (H2), then η = 0, Cη = Cγ , and the solution of (Qk) is unique if τ <

4ξ/(C2γ(1 + C4

PC2γ)) (Case 1), and τ < 4ξ/C2

γ (Case 2).

We have shown existence and uniqueness for (Qk). Next, we show that (Qk)and (Pk) are equivalent.

Proposition 3.6. Under conditions of Theorem 3.5, problems (Qk) and (Pk) ad-mit unique solutions and are equivalent.

Proof. Since the problems (Qk) and (J k) are equivalent, it suffices to show that (J k)and (Pk) are equivalent. First, we show the existence of the Lagrange multiplierµk ∈ R by verifying a constraint qualification condition [64, Theorem 6.3]. Foru ∈ K and m ∈ (|Ω|, |Ω|), we define the equality constraint G(u) = (u, 1) −m. If

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 15

uk ∈ Km is a unique minimizer of (J k), then G(uk) = 0, and we define the tangentcone

C(uk) = α(u− uk) : α ≥ 0, u ∈ K.

Taking into account that G′(u) = 1, we derive that

S := G′(uk)C(uk) =

α

∫Ω

(u− uk) dx, α ≥ 0, u ∈ K.

Since u ∈ K and uk ∈ Km, it follows that S ≡ R, and a constraint qualification isfulfilled. By [64, Theorem 6.3], there exist µk ∈ R such that

(∇uL(uk, µk), v − uk)L2(Ω) ≥ 0, ∀v ∈ K,

where L(u, µ) := Jk(u) + µG(u) is the Lagrange function, and

∇uL(uk, µ) = ∇Jk(uk) + µ∇G(uk) = Buk + F ′0(uk) +1

τG(uk − uk−1) + µk.

Setting wk := − 1τ G(uk−uk−1)−µk, and noting that 1

τ G(uk−uk−1) ∈ VA0 , µk ∈ R,

we obtain that wk ∈ VA, and (uk, wk) ∈ K × VA is a solution of (Pk).

3.2. Properties of the solution. In this section we prove that, under certainconditions, the solution of the nonlocal Cahn-Hilliard problem (Pk) at each timestep admits discontinuous solutions that imply sharp interfaces in the model.

From (3.1) and (2.9) we obtain that the following relation holds for k = 1, . . . ,Ka.e. in Ω:

wk = Buk + F ′0(uk) + λk = ξuk − γ ∗ uk + λk, (3.7)

where, we recall ξ(x) = cγ(x)− cF . We define for k = 1, . . . ,K

gk := wk + γ ∗ uk = −1

τG(uk − uk−1)− µk + γ ∗ uk,

and from (3.7) it follows that gk = ξuk + λk. Hence, for ξ(x) > 0, x ∈ Ω, sinceλk ∈ ∂I[−1,1](uk), where we recall

λk ∈ ∂I[−1,1](uk) =

(−∞, 0] if uk = −1,

0 for uk ∈ (−1, 1),

[0,+∞) if uk = 1,

(3.8)

the solution uk at each time step tk can be calculated as a pointwise projection ofgk/ξ onto [−1, 1]:

uk = P[−1,1]

(1

ξgk)

=

1, if gk ≥ ξ,−1, if gk ≤ −ξ,gk/ξ, if gk ∈ (−ξ, ξ).

(3.9)

The projection formula (3.9) provides a crucial insight into the properties of thesolution, such as the regularity stated below.

Theorem 3.7 (Improved regularity). Let (uk, wk) ∈ K × VA be the solution pairof (Pk), and the kernel γ satisfies (H1)–(H2). If ξ(x) > 0 for all x ∈ Ω, thenuk ∈ H1(Ω), for all K = 1, . . . ,K, and∥∥∇uk∥∥

L2(Ω)≤∥∥∇(ξ−1gk)

∥∥L2(Ω)

≤ C∥∥gk∥∥

H1(Ω), (3.10)

where C > 0 depends only on ξ and γ.

16 O. BURKOVSKA AND M. GUNZBURGER

Proof. Since γ ∈W 1,1(Rn) and uk ∈ L∞(Rn) (by Proposition 3.2), invoking Propo-sition 3.3 it follows that γ ∗ uk ∈ W 1,∞(Rn). Since wk ∈ H1(Ω), we obtain thatgk = wk + γ ∗ uk ∈ H1(Ω). Invoking the projection formula (3.9) and using thestability of the L2-projection in H1, we obtain the following estimate∥∥∇uk∥∥

L2(Ω)≤∥∥∇(ξ−1gk)

∥∥L2(Ω)

. (3.11)

We note that for Case 1, ξ(x) is constant for all x ∈ Ω. Therefore, from the aboveestimate we immediately deduce that

∣∣uk∣∣H1(Ω)

≤ ξ−1∣∣gk∣∣

H1(Ω)≤ C, and, hence,

u ∈ H1(Ω). For Case 2, since γ ∈ W 1,1(Rn), we have that ξ, ξ−1 ∈ W 1,∞(Ω).Indeed, since ξ(x) > 0 in Ω, it follows that there exists ξmin > 0, such that ξ(x) ≥ξmin > 0, and∥∥∇(ξ−1)

∥∥L∞(Ω)

≤ ξ−2min‖∇ξ‖L∞(Ω) ≤ ξ−2min‖∇γ‖L1(Rn) <∞. (3.12)

Then, using the product rule, the estimate (3.11) reduces to∣∣uk∣∣H1(Ω)

≤∣∣ξ−1gk∣∣

H1(Ω)≤ Cγ

∣∣gk∣∣H1(Ω)

+∥∥∇(ξ−1)

∥∥L∞(Ω)

∥∥gk∥∥L2(Ω)

≤ C∥∥gk∥∥

H1(Ω),

and, thus, uk ∈ H1(Ω).

Corollary 3.2. Let γ satisfy (H1)–(H2). Then, for ξ(x) ≥ 0 for all x ∈ Ω, theLagrange multiplier in (3.1) fulfills λk ∈ H1(Ω), k = 1, . . . ,K.

In contrast to the previous result, we show next that if ξ = 0, the solutionuk is generally not smoother than uk ∈ L∞(Ω), and furthermore, it can possesjump-discontinuities, which allow that the solution can admit only pure phases.

Theorem 3.8 (Sharp interfaces). Let (uk, wk) ∈ K × VA be the solution pairof (Pk), and the kernel γ satisfies (H1). If ξ(x) := cγ(x) − cF = 0, ∀x ∈ Ω,then it holds for all k = 1, . . . ,K a.e. in Ω:

uk ∈

1, if gk > 0,

−1, if gk < 0,

[−1, 1], if gk = 0,

(3.13)

where gk = wk + γ ∗ uk. Thus, in the case that the function gk assumes thevalues zero only on a set of measure zero, i.e., |gk = 0| = 0, the variable uk

assumes only the extreme values −1 and 1 almost everywhere. That is, the solutionuk, k = 1, . . . ,K is discontinuous and consists only of pure phases uk = 1 anduk = −1.

Proof. We recall that from (3.7), gk = ξuk + λk, where λk = λk+ − λk−, λk± ≥ 0.

Then, if ξ = 0, we obtain that gk = λk, where we recall that for k = 1, . . . ,Kλk ∈ ∂I[−1,1](u

k) (3.8) holds true. Hence, if gk > 0 we get λk > 0 and uk = 1.

Similarly, if gk < 0, then λk < 0 and uk = −1. Finally, for gk = 0 we obtain thatλk = 0 and uk ∈ [−1, 1], and we conclude the proof.

Remark 3.1. There are settings for which we can guarantee that |gk = 0| = 0.In particular, if we consider the steady state problem in Ω ⊂ R1, i.e., for K →∞,wk → −µ and uk → u, where u and µ are the steady state solution and mean-value, respectively. Then, taking, e.g., an analytic convolution kernel γ we obtainthat g := γ ∗ u − µ is analytic and not constantly equal to zero, unless u ≡ meverywhere. Hence, if u 6≡ m, |g = 0| = 0 always holds true and the solutionadmits only pure phases.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 17

4. Continuous problem: Existence, uniqueness and a sharp interfacecondition

In this section, we analyze the continuous in time problem (P). To derive thecorresponding existence and uniqueness results, we analyze the semidiscrete prob-lem defined in the previous section by taking the limit as τ → 0. We demonstratethat the projection formula (3.9) also remains valid for the continuous case.

4.1. Existence of a solution. Let X be either VB , L2(Ω) or VA, and recall that

Q = (0, T ) × Ω and Q = (0, T ) × Ω. Then, for K ∈ N, τ = T/K, T > 0, anda given sequence of functions zkKk=1 ⊂ X, we introduce piecewise constant andpiecewise linear interpolants:

zτ (t) := zk, zτ (t) := zk−1, t ∈ (tk−1, tk], k = 1, . . . ,K,

zτ (t) :=t− tk−1

τzk +

tk − tτ

zk−1, t ∈ [tk−1, tk], k = 1, . . . ,K.(4.1)

In terms of the above notations, it follows that ∂tzτ = (zk − zk−1)/τ , t ∈ [tk−1, tk],k = 1, . . . ,K. Then by setting wτ (t) := −G(∂tzτ (t)) − µτ (t), the semi-discreteproblem (3.1) can be recast for a.e. t ∈ (0, T ) as follows:

(G(∂tuτ (t)) + ξuτ (t)− γ ∗ uτ (t) + λτ (t) + µτ (t), ψ)L2(Ω) = 0, ∀ψ ∈ VB , (4.2a)

(η − λ±,τ (t), 1∓ uτ (t))L2(Ω) ≥ 0, ∀η ∈M. (4.2b)

Next, we establish the existence result, and invoking the projection formula (3.9)derive an improved regularity of the solution.

Theorem 4.1 (Existence and improved regularity). Let γ satisfy (H1)–(H2), andξ(x) := cγ(x) − cF ≥ 0 for all x ∈ Ω, then there exists a solution pair (u(t), w(t))of (P), such that

u ∈ L∞(Q), ∂tu ∈ L2(0, T ;V ′A), and w ∈ L2(0, T ;VA).

Moreover, if ξ(x) > 0 for all x ∈ Ω, then

u ∈W (0, T ) ∩ L∞(Q) and w ∈ L2(0, T ;VA),

where W (0, T ) :=v ∈ L2(0, T ;VA) : ∂tv ∈ L2(0, T ;V ′A)

.

Proof. A priori estimates. Since |uτ (t)| ≤ 1 for t ∈ (0, T ) a.e., it immediately

follows that u ∈ L∞(Q). Next, we show that ∂tuτ ∈ L2(0, T ;V ′A). From theenergy minimization (J k) and Proposition 3.1 it follows that Jk(uk) ≤ Jk(uk−1),k = 1, . . . ,K. That is,

1

∥∥uk − uk−1∥∥2V ′A≤ 1

2

∣∣uk−1∣∣2VB− 1

2

∣∣uk∣∣2VB− cF

2

∥∥uk−1∥∥2L2(Ω)

+cF2

∥∥uk∥∥2L2(Ω)

.

Then, utilizing the above expression we deduce

‖∂tuτ‖2L2(0,T ;V ′A) =

∫ T

0

‖∂tuτ‖2V ′A dt =

K∑k=1

∫ tk

tk−1

1

τ2∥∥uk − uk−1∥∥2

V ′Adt

≤∣∣u0∣∣2

VB+ cF

∥∥uK∥∥2L2(Ω)

≤ C. (4.3)

Case ξ > 0. We consider (4.2) with ψ = m − u(t) ∈ VB , where m =1|Ω|∫Ωu(x) dx = m/|Ω|, m ∈ (−1, 1), and taking into account that (µ(t),m −

u(t))L2(Ω) = 0 for t ∈ (0, T ) a.e., we arrive at

−(G(∂tuτ (t)), uτ (t))L2(Ω) +(ξuτ (t)− γ ∗ uτ (t) + λτ (t),m− uτ (t)

)L2(Ω)

= 0.

18 O. BURKOVSKA AND M. GUNZBURGER

From the complementarity conditions (4.2b) it follows that uτ (t) = ±1 if λ±,τ (t) >0. Then, using repeatedly Cauchy and Young’s inequalities, and (2.3), we obtain

0 ≤ (λ+,τ , 1−m)L2(Ω) + (λ−,τ (t), 1 +m)L2(Ω) = −(λτ (t),m− uτ (t))L2(Ω)

≤∣∣∣〈G(∂tuτ (t)), uτ (t)〉VA×V ′A

∣∣∣+ ‖ξuτ (t)− γ ∗ uτ (t)‖L2(Ω)‖m− uτ (t)‖L2(Ω)

≤ ‖G(∂tuτ (t))‖VA‖uτ (t)‖V ′A + C(‖ξuτ (t)‖L2(Ω) + ‖uτ (t)‖L2(Ω)

)≤ ‖G(∂tuτ (t))‖2VA + C

(‖uτ (t)‖2V ′A + ‖ξuτ (t)‖L2(Ω) + ‖uτ (t)‖L2(Ω)

)≤ C

(|G(∂tuτ (t))|2VA + 1

).

Now integrating the last expression from (0, T ) and invoking the estimate (4.3), weobtain that∫ T

0

|λτ (t)|dt ≤∫ T

0

(λ+,τ (t) + λ−,τ (t)) dt ≤ C

(∫ T

0

‖∂tuτ (t)‖2V ′A dt+ T

)≤ C,

and, hence λτ ∈ L2(0, T ;L1(Ω)). Therefore, from (4.2a) it follows that

µτ = −G(∂tuτ )− ξuτ + γ ∗ uτ − λτ ∈ L2(0, T ;R),

and, hence, −G(∂tuτ ) − µτ =: wτ ∈ L2(0, T ;VA), with the uniform bounds

‖µτ‖2L2(0,T ;R) ≤ C and ‖wτ‖2L2(0,T ;VA) ≤ C. From the derived regularity esti-

mates, and invoking again (4.2a), we immediately conclude that λτ ∈ L2(Q), and∥∥λτ∥∥L2(Q)≤ C.

Next, we are going to invoke the higher regularity results from Theorem 3.7. Weintroduce an interpolant of gk, defined in Section 3.2, gτ (t) = wτ (t) + γ ∗ uτ (t).From the projection formula (3.9), we obtain that uτ = P[−1,1](ξ

−1gτ (t)) holds a.e.for t ∈ (0, T ), and from (3.10) it holds that

‖∇uτ (t)‖2L2(Ω) ≤ C‖gτ (t)‖2H1(Ω).

Integrating the above inequality from (0, T ), and invoking (4.3), (2.3), and (2.11),leads to∫ T

0

‖∇uτ (t)‖2L2(Ω) dt ≤ C∫ T

0

‖gτ (t)‖2H1(Ω) dt

≤ C

(∫ T

0

|G(∂tuτ (t))|2VA dt+

∫ T

0

‖uτ (t)‖2L2(Ω) dt+ ‖µτ‖2L2(0,T ;R)

)≤ C

(‖∂tuτ‖2L2(0,T ;V ′A) + ‖uτ‖2L∞(Q) + ‖µτ‖

2L2(0,T ;R)

)≤ C,

and, hence, uτ ∈ L2(0, T ;VA). Since, ξ ∈ W 1,∞(Ω) (3.12), it follows that ξuτ ∈L2(0, T ;VA), and, hence, from (4.2) we immediately deduce that

λτ = −G(∂tuτ )− ξuτ + γ ∗ uτ − µτ ∈ L2(0, T ;VA).

Collecting the above estimates, we arrive at the following a priori energy bound

‖∂tuτ‖2L2(0,T ;V ′A) + ‖uτ‖2L2(0,T ;VA) +∥∥λτ∥∥2L2(0,T ;VA)

+ ‖µτ‖2L2(0,T ;R) ≤ C.

Case ξ = 0. Following the same steps as above we obtain that λτ ∈ L2(Q), µτ ∈L2(0, T ;R), and wτ ∈ L2(0, T ;VA) with the corresponding uniform norm bounds.However, in contrast to the the previous case, here the projection formula (3.9)does not hold, and, in general, we can not expect a solution to have an improvedregularity. Nevertheless, we still can derive an improved regularity for the Lagrangemultiplier λτ . Indeed, using that γ ∗ uτ ∈ L2(0, T ;VA) for γ ∈ W 1,1(Rn), uτ ∈

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 19

L∞(Q), and G(∂tuτ ) ∈ L2(0, T ;VA), µτ ∈ L2(0, T ;R), we conclude from (4.2a),that λτ ∈ L2(0, T ;VA), and we have the following estimate:

‖∂tuτ‖2L2(0,T ;V ′A) + ‖uτ‖2L2(0,T ;VB) +∥∥λτ∥∥2L2(0,T ;VA)

+ ‖µτ‖2L2(0,T ;R) ≤ C.

Limit. From the a priori estimates and the Banach-Alaouglu theorem, thereexist functions u, λ and µ such that

(ξ ≥ 0) uτ u weakly-* in L∞(Q),

(ξ > 0) uτ u weakly-* in L∞(0, T ;VA),

∂tuτ ∂tu weakly in L2(0, T ;V ′A),

µτ µ weakly in L2(0, T ;R),

λτ λ weakly in L2(0, T ;VA),

as τ → 0 (or equivalently, as K →∞), where by a slight abuse of notation we keepthe same subscript for convergent sub-sequences. Next, we are going to pass to thelimit τ → 0 in an equivalent time-integrated variant of (4.2). That is, we consider

(G(∂tuτ ) + ξuτ − γ ∗ uτ + λτ + µτ , ψ)L2(Q) = 0, ∀ψ ∈ L2(0, T ;VB),

(η − λ±,τ , 1∓ uτ )L2(Q) ≥ 0, ∀η ∈ L2(0, T ;M),(4.4)

where now, letting τ → 0 and using the linearity of the first equation in (4.4), weobtain that the limiting functions u, λ and µ satisfy the first equation in (2.19).To be able to pass to the limit in the variational inequality constraints (4.2b),and to deal with the induced nonlinearity, we need to upgrade a weak convergenceto a strong one. Using the Aubin-Lions compactness lemma we obtain that theembedding L2(0, T ;L2(Ω)) ∩H1(0, T ;V ′A) → L2(0, T ;V ′A) is compact, and hence,we obtain that

uτ → u strongly in L2(0, T ;V ′A).

Next, we show we show that uτ → u strongly in L2(0, T ;V ′A). First, we notethat the following estimate holds true for piecewise linear and piecewise constantinterpolants, see, e.g., [19, Proposition 3.9]:

‖uτ − uτ‖L2(0,T ;V ′A) ≤ Cτ‖∂tuτ‖L2(0,T ;V ′A).

Then, using the above estimate and a triangle inequality we arrive at

‖uτ − u‖L2(0,T ;V ′A) ≤ ‖uτ − uτ‖L2(0,T ;V ′A) + ‖uτ − u‖L2(0,T ;V ′A) → 0,

as τ → 0 (or equivalently, as K → ∞). Similarly as in (2.18) we can decomposethe complementarity conditions (4.4) as∫ T

0

(η(t), 1∓ uτ (t))L2(Ω) dt ≥ 0 and

∫ T

0

(λ±,τ (t), 1∓ uτ (t))L2(Ω) dt = 0.

Now passing to the limit τ → 0 in the above expressions and using the fact thatthe inner product of strong and weak convergences converges, we obtain

0 ≥ (η, uτ ∓ 1)L2(Q) → (η, u∓ 1)L2(Q),

0 =⟨λ±,τ , 1∓ uτ

⟩L2(0,T ;VA)×L2(0,T ;V ′A)

→ 〈λ±, 1∓ u〉L2(0,T ;VA)×L2(0,T ;V ′A),

that leads that the limiting functions λ = λ+−λ− and u satisfy the complementarityconditions in (2.19), and since the problems (2.19) and (P) are equivalent, thisconcludes the proof.

20 O. BURKOVSKA AND M. GUNZBURGER

4.2. Uniqueness and continuous dependence result.

Theorem 4.2. Let γ satisfy (H1)–(H2), and ξ(x) := cγ(x)− cF > 0 for all x ∈ Ω,

then the solution u ∈ W (0, T ) ∩ L∞(Q) of the problem (P) is unique. Moreover,if u1 and u2 are two solutions of (P) with the initial conditions u01 = u1(0) andu02 = u2(0), respectively, then we have the following continuous dependence result

‖u1(t)− u2(t)‖2V ′A ≤ eCT∥∥u01 − u02∥∥2V ′A , for a.e. t ∈ (0, T ), (4.5)

where a constant C > 0 is independent of the initial condition.

Proof. Let (u1, w1) and (u2, w2) be two solutions of (P) with the correspondinginitial conditions u01 = u1(0) and u02 = u2(0). We define u(t) := u1(t) − u2(t),w(t) := w1(t)−w2(t) and u(0) = u1(0)−u2(0). Next, we consider (P) for solutions(u1, w1) and (u2, w2). Taking a difference and choosing the test function φ = Gu(t),noting that u(t) ∈ VB ⊂ V ′A, and exploiting (2.6), we arrive at

d

dt‖u(t)‖2V ′A + (w(t), u(t))L2(Ω) = 0, a.e. for t ∈ (0, T ). (4.6)

We consider the second inequality in the problem (P) for (u1, w1) and (u2, w2)tested with φ = u2 and φ = u1, respectively. Then, taking the difference of the twoinequalities and accounting for (2.9) we obtain that

(ξu(t), u(t))L2(Ω) − (γ ∗ u(t), u(t))L2(Ω) − (w(t), u(t))L2(Ω) ≤ 0. (4.7)

Using similar techniques as in the proof of Theorem 3.5 we can estimate the secondterm in the above inequality:∣∣(γ ∗ u(t), u(t))L2(Ω)

∣∣ =∣∣∣〈γ ∗ u(t), u(t)〉VA×V ′A

∣∣∣ ≤ ‖∇γ ∗ u(t)‖L2(Ω)‖u(t)‖V ′A

≤ Cγ‖u(t)‖L2(Ω)‖u(t)‖V ′A ≤ Cγ(ε

2‖u(t)‖2L2(Ω) +

1

2ε‖u(t)‖2V ′A

)≤ Cγε

2(1 + C4

PC2γ)‖u(t)‖2L2(Ω) +

Cγ2ε‖u(t)‖2V ′A , (4.8)

with ε > 0. Since, ξ(x) ≥ 0 in Ω, we obtain that there exists ξmin > 0, such thatξ(x) ≥ ξmin > 0. Then, combining (4.7) and (4.8) we obtain(

ξmin −Cγε

2(1 + C4

PC2γ)

)‖u(t)‖2L2(Ω) −

Cγ2ε‖u(t)‖2V ′A − (w(t), u(t))L2(Ω) ≤ 0.

Choosing ε = ξminC−1γ (1 +C4

PC2γ)−1 and adding the above inequality to (4.6) leads

tod

dt‖u(t)‖2VA′ +

ξmin

2‖u(t)‖2L2(Ω) ≤ C‖u(t)‖2V ′A .

Now, applying the Gronwall lemma we obtain the desired estimate and concludethe proof.

4.3. Properties of the solution. Similarly as in the time discrete case discussedin Section 3.2, for ξ(x) > 0, ∀x ∈ Ω, we obtain the projection formula that holdsa.e. in (0, T )×Ω:

u(t) = P[−1,1]

(1

ξg(t)

)=

1, if g(t) ≥ ξ,−1, if g(t) ≤ −ξ,(1/ξ)g(t), if g(t) ∈ (−ξ, ξ),

(4.9)

where, we recall, g(t) := w(t) + γ ∗ u(t). Analogously, for ξ(x) = 0, x ∈ Ω, thesolution can admit sharp interfaces.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 21

Theorem 4.3 (Sharp interfaces). Let (u(t), w(t)) be the solution pair of (P), andγ satisfy (H1). If ξ(x) := cγ(x)− cF = 0, ∀x ∈ Ω, then it holds a.e. in (0, T )×Ω:

u(t) ∈

1, if g(t) > 0,

−1, if g(t) < 0,

[−1, 1], if g(t) = 0,

(4.10)

where g(t) = w(t) + γ ∗ u(t), and, if |g(t) = 0| = 0 a.e. t ∈ (0, T ), then u(t)assumes only phases −1 and 1 a.e. in (0, T )×Ω.

5. Numerical results

In this section we present numerical experiments for one and two-dimensionalcases for the nonlocal operator (2.1) defined for Case 1 and Case 2.

5.1. Discretization. Let Ω ⊂ Rn, d ≥ 1, be a polygonal domain. We partitionΩ ∪ ΩI into a shape regular quasi-uniform triangulation Thh such that thereexists a proper triangulation T ′hh ⊂ Thh of Ω, that respects the boundaryof Ω. We denote by h the maximum diameter of the elements K ∈ Th and set

Ωh ∪ΩhI = ∪K∈ThK and Ωh = ∪K∈T ′hK.

We employ the implicit Euler time stepping scheme (Pk) together with piecewise-linear continuous finite elements for spatial discretization. We define

Sh = vh ∈ C0(Ω) : vh|K ∈ P1(K), ∀K ∈ Th,

Sh = vh ∈ C0(Ω) : vh|K ∈ P1(K) ∀K ∈ Th.

Let JΩh denote the set of nodes corresponding to the triangulation of Ω (including

the boundary nodes), J Ih the set of nodes in Ω∪ΩI \Ω, and J Ωh the set of all nodes.

Also, we set pj ∈ J ∗h , ∗ ∈ Ω, I, Ω to be the coordinates of the corresponding

nodes. Then, we can represent Sh = spanφp, p ∈ J Ωh and Sh = spanφp, p ∈J Ωh , where φi are the nodal Lagrange basis functions, φi(xj) = δi,j .

Similarly as in [8] we employ a mass-lumping approach, where by (·, ·)h and (·, ·)hwe denote a mass-lumped L2(Ω) and L2(Ω \ Ω) inner products, respectively. Toadapt the mass-lumping procedure for the nonlocal term, we employ the trapezoidalquadrature rule to assemble b(·, ·). In this way, the resulting discretized bilinearform preserves the mass-lumping property. Indeed, let Ixh [·] be the nodal interpolantfor the variable x (respectively for y, Iyh) and let chγ(x) :=

∫ΩIyh [γ(x, y)] dy. For any

φ, ψ ∈ Sh we introduce the discrete convolution

(γ ~ φ)(x) :=

∫Ω

Iyh [γ(x, y)φ(y)] dy =∑k∈J Ωh

mkγ(x, yk)φ(yk), mk =

∫Ω

φk(x) dx

and obtain the mass-lumped bilinear form

bh(φ, ψ) :=1

2

∫Ω

∫Ω

IxhIyh [(φ(x)− φ(y))(ψ(x)− ψ(y))] dxdy

=

∫Ω

Ixh[chγφ(x)ψ(x)

]dx−

∫Ω

Ixh [ψ(x)(γ ~ φ)(x)] dx+

∫Ω\Ω

Ixh [ψ(x)Nhφ(x)] dx

= (chγφ, ψ)h − (γ ~ φ, ψ)h + (Nhφ, ψ)h,

where

Nhφ(x) :=

∫Ω

Iyh [(φ(x)− φ(y))γ(x, y)] dy = chγφ(x)− (γ ~ φ)(x).

22 O. BURKOVSKA AND M. GUNZBURGER

From the above expression we see that mass-lumping for the nonlocal term holds

on the discrete level. Now, letting ψj ∈ Sh to be the Lagrangian basis functions,we obtain that bh(φ, ψj) = (chγφ, ψj)h− (γ~φ, ψj)h for all j ∈ JΩh and bh(φ, ψj) =

(Nhφ, ψj)h for all j ∈ J Ih .

Then, the fully discrete problem reads as follows: For a given u0h ∈ Sh we seek

(wkh, ukh, λ

kh) ∈ Sh × Sh × Sh, k = 1, . . . ,K, such that the following holds

1

τ(ukh − uk−1h , φ)h + (∇wkh,∇φ)L2(Ω) = 0, ∀φ ∈ Sh,

(wkh, ψ)h − bh(ukh, ψ) + cF (ukh, ψ)h − (λkh, ψ)h = 0, ∀ψ ∈ Sh,

λkh = λkh,+ − λkh,−, λkh,+ ≥ 0, λkh,− ≥ 0, |ukh| ≤ 1,

λkh,±(pj)(ukh(pj)∓ 1) = 0, ∀pj ∈ JΩh .

(5.1)

Thanks to the mass-lumping property of bh(·, ·), we can also obtain the projection

formula (4.9), for the fully discrete solution ukh ∈ Sh of (5.1):

ukh(pj) = P[−1,1]

(1

ξh(pj)gkh(pj)

), pj ∈ JΩh , k = 1, . . . ,K,

where ξh(pj) = chγ(pj) − cF > 0 and gkh := wkh + γh ~ ukh. The availability of theabove formula is useful, as it provides an insight into the stability properties of thediscrete solution.

We note that the discrete system (5.1) is computationally demanding due tothe inclusion of the nonlocal term bh(·, ·) on the left-hand side of the equation.This leads to inverting a large and dense matrix at each time step in (5.1). Toovercome it, one can consider instead an implicit-explicit time stepping scheme,where we discretize the local part of bh(·, ·) implicitly, and the remaining nonlocalpart explicitly, i.e., we replace bh(ukh, ψ) in (5.1) with the approximation

bh(ukh, ψ) ≈ (chγukh, ψ)h − (γh ~ uk−1h , ψ)h. (5.2)

The discrete systems (5.1) can be solved by the primal-dual-active set strategy(PDAS), which has been already successfully employed for the local Cahn-Hilliardvariational inequality problem; see, e.g., [8, 54]. Under suitable conditions, PDASis equivalent to the semismooth Newton method, which converges locally superlin-early; for more details, see [53].

We note that, while the present work focuses on solutions that may admit jumpdiscontinuitites, an adoption of the continuous finite element method with the mass-lumping approach allows the avoidance of instabilities associated with the approx-imation of sharp interfaces. Alternatively, one could also consider a discontinuousGalerkin finite element method, which has been recently explored in nonlocal set-tings [16, 62, 33, 37, 34, 60].

Next we present several numerical examples which illustrate the behaviour ofthe nonlocal solution of (5.1) and a comparative study of the nonlocal vs. localsolutions of the Cahn-Hilliard variational inequality. By “local” we mean the solu-tion of (1.1) with the constant mobility and with the same obstacle potential F asdefined in (1.4).

5.2. Numerical examples. The domain Ω is set to be Ω = (0, 1)d, d ∈ 1, 2and Ω is discretized with a quasi-uniform mesh with a number of nodes N that willbe specified in each instance. For the choice of the kernel γ, we consider a scaledGaussian kernel, similar to the one defined in [31]:

γ(x, y) =4ε2

πn/2(δ/3)n+2e−|x−y|2

(δ/3)2 , x ∈ Rn, δ > 0. (5.3)

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 23

For δ → 0 the nonlocal operator B recovers the Laplace operator −ε2∆, see,e.g., [31, 63]. We additionally truncate the kernel at |x − y| > δ to ensure thatwe have a finite extent of nonlocal interactions defined by δ. Due to the exponen-tial decay in the kernel (5.3), this truncation introduces a negligible error in thecorresponding solution. For Case 1 the constant cγ in (2.2) is approximately equalcγ ≈ 36ε2/δ2. The parameters ε, cF and δ are specified for each case separately.

Example 1a. Let Ω = (0, 1) and we consider the “Neumann” type nonlocal op-erator B, defined as in Case 1. We set T = 2, τ = 2 · 10−4, δ = 0.25, andξ = cγ − cF = 0.008, that corresponds to ε2 = 0.00175 and cF = 1. The initial con-dition is chosen as u0 = 0.1(sin(2πx)+sin(3πx)). In Figure 2 we depict the nonlocaland local solutions of the Cahn-Hilliard variational inequality at different time in-stances. We can clearly observe that the nonlocal solution has jump-discontinuities

t = 0.02 t = 0.03

t = 0.06 t = 2

Figure 2. Evolution of the nonlocal (Case 1) and local solutionsof the Cahn-Hilliard variational inequality at different time in-stances for Example 1a.

(up to the resolution of the employed computational grid) and admits mostly purestates, while the local solution also takes values in (−1, 1) even for large times.These observations are in agreement with our theoretical results, where we knowthat for ξ = 0 the solution can admit pure phases throughout the domain.

Example 1b. Next, we are going to investigate the behaviour of the solutioncorresponding to the “regional” type nonlocal operator, defined in Case 2. Wekeep the same settings as in the previous example, apart from ξ, which is no longerconstant in the present case. We set ξmin = minξ(x) = 0.008, x ∈ Ω, whichcorresponds to ε2 = 0.00175 and cF = 0.4960. In Figure 3 we depict the evolutionof the solution at different time instances. In contrast to the previous example,

24 O. BURKOVSKA AND M. GUNZBURGER

we are no longer observing sharp interfaces of the solution. This is explained bythe fact that ξ = ξ(x) is spatially dependent and does not vanish for all ξ ∈ Ω.Furthermore, we observe that the phase transitions that occur near the boundaryof Ω have steeper gradients compared to the phase transitions appearing in theinterior of the domain. This is explained by the fact that ξ(x) is the smallest andequal to ξmin in the boundary nodes and according to the theory this leads tosharper interfaces in the solution.

t = 0.02 t = 0.03

t = 0.06 t = 2

Figure 3. Evolution of the nonlocal (Case 2) and local solutionsof the Cahn-Hilliard variational inequality at different time in-stances for Example 1b.

Example 1c. Now, we again consider the “regional” type nonlocal operator, how-ever set cF = cF (x) to be a spatially dependent coefficient such that ξ(x) =cγ(x) − cF (x) is constant throughout Ω. In particular, taking ε2 = 0.00175 andusing the same settings as in Example 1a, we consider cF (x) = cγ(x) − 0.008.This leads to ξ = 0.008 being constant in Ω. While such modification allows usto get constant values of ξ uniformly in Ω, this changes a potential, i.e., F0(u) =(cf (x)/2)(1 − u2(x)), which, in turn, also leads to a modification of the underly-ing problem. In Figure 4 we depict the snapshots of the solution at different timeinstances. Here, in contrast to the previous example, we are able to get sharpinterfaces in the solution similar as in the case of the “Neumann” type nonlocaloperator. However, in contrast to the “Neumann” type problem, this comes withthe need to use a non-typical double-well potential.

Example 1d. Now, using the“Neumann” type nonlocal operator we investigatethe variation of the solution with respect to the nonlocal interface parameter ξ. Wekeep the same settings as in Example 1a, and to obtain different values of ξ we

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 25

t = 0.02 t = 0.03

t = 0.06 t = 2

Figure 4. Evolution of the nonlocal (Case 2) and local solutionsof the Cahn-Hilliard variational inequality at different time in-stances and with cF = cF (x) for Example 1c.

vary either the coefficient cF or a nonlocal interaction radius δ. The correspondingsnapshots are depicted in Figure 5. Here, we can observe that the width of theinterface is changing with respect to ξ. In particular, for ξ = 0 that correspondseither to cF = 1.56 or δ = 0.25, we can see that the solution admits sharp interfaces,which is also in agreement with the previous examples, and for a larger ξ theinterface becomes more diffuse and the nonlocal solution conforms more closelyto the corresponding local solution. We also observe that for a smaller extent ofnonlocal interactions δ the nonlocal solutions are close to the local one and theinterface becomes more diffuse, which is an expected behavior here. Similar resultshave been also reported in the Cahn-Hilliard case with the regular potential [31].

Overall, we could see that the parameter ξ plays a role of an interface parameterin the nonlocal model, similar to the interface parameter ε in the local setting.Changing the support of the kernel and the scaling of the double-well potentialhave a great affect on ξ and, hence, on the width of the interface.

Next, we conduct a comparative study for two-dimensional examples.

Example 2a. Now, let Ω = (0, 1)2 and consider the “Neumann” type nonlocaloperator B, defined as in Case 1, where we set T = 1, K = 2000, N = 39009,δ = 0.25, ξ = 0, cF = 1, ε2 = 0.0017. The initial condition u0 is chosen as

u0(x) = 2(e−(6x−2.1)

2−(6y−3)2 + e−(7x−4.9)2−(7y−3.5)2

)− 1.

In Figure 6 we plot the snapshots of the local and nonlocal solutions at different timesteps. We can observe that local and nonlocal solutions look alike quantitatively.However, whereas the nonlocal model can describe perfectly sharp interfaces up to

26 O. BURKOVSKA AND M. GUNZBURGER

Figure 5. Snapshots of the nonlocal (Case 2) and local solutionsof the Cahn-Hilliard variational inequality for different values ofcF (left) and δ (right).

the resolution of the discretization mesh, the interfaces for local solution are diffuse.

Figure 6. Evolution of the nonlocal (Case 1) (bottom) and lo-cal (top) solutions of the Cahn-Hilliard variational inequality atdifferent time instances and for Example 2a. From left to right:t = 0.004, 0.009, 0.0225, 1.

Example 2b. Now, let Ω = (0, 1)2 and consider the “Neumann” type nonlocaloperator B, defined as in Case 1, where we set T = 2, K = 1000, N = 43073,δ = 0.1, ξ = 0.07, cF = 1, ε2 = 0.0003. The initial condition u0 is chosen asu0(x) = τ(x), where τ(x) is drawn from a uniform random distribution on [−1, 1]at each grid point. In Figure 7 we plot the snapshots of the local and nonlocalsolutions at different time-steps. From these plots, we observe that similarly asin the previous example, the proposed nonlocal model for Case 1 delivers sharpinterfaces in the solution whereas the local model results in diffuse interfaces. Wealso notice that due the random initial condition, the patterns in the local andnonlocal solutions in the intermediate time steps differ significantly in contrast toExample 2a for which a deterministic initial condition has been used. This is dueto the random and non-smooth initial condition, which has the effect that changesin the model can lead to large changes in the final solution.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 27

t = 0.01 t = 0.1 t = 0.4

t = 0.01 t = 0.1 t = 0.4

Figure 7. Evolution of the nonlocal (Case 1) (bottom) and lo-cal (top) solutions of the Cahn-Hilliard variational inequality atdifferent time instances for Example 2.

Example 3. Next, similarly as in Example 1c, we investigate the case of the “re-gional” nonolocal operator, given as in Case 2. Again, we consider cF = cF (x) isa spatially dependent coefficient, such that ξ(x) = cγ(x) − cF (x) is close to zerothroughout a whole domain Ω. In particular, we chose cF (x) := 0.9cγ(x) and obtainξ(x) = 0.1cγ(x), which is positive but close to zero in Ω.

We set Ω = (0, 1)2, N = 4225, T = 1, K = 100, δ = 0.3, and ε2 = 0.004. Theinitial condition u0 is chosen as u0(x) = τ(x), where τ(x) is drawn from a uniformrandom distribution on [−1, 1] at each grid point. In Figure 8 we plot the nonlocaland local solutions at different time instances. In this case, similarly as before, bymeans of modifying the double-well potential we could achieve sharper interfacesin the nonlocal solution for the “regional” nonlocal operator compared to the localcase. This corresponds to the fact that ξ(x) is very small, much smaller than cγ(x).

In summary, we can conclude that the numerical results reported are in agree-ment with the theoretical developments. Furthermore, the numerical investigationsalso reveals how the choice of the boundary conditions, model, and kernel param-eters affect the sharpness of the interface in the solution. Moreover, we have alsodemonstrated that the nonlocal solution with sharp interfaces can be stably sim-ulated on a given mesh for arbitrary interface parameter ξ ≥ 0 without tying thediscretization parameters to the interface width (including the case ξ = 0, whichcorresponds to a sharp interface on an analytic level).

6. Concluding remarks

In this work we have presented a nonlocal Cahn-Hilliard model that permitssolutions to achieve only pure phases. We have performed a detailed analyses ofthe well-posedness of the problem as well as for the regularity of solutions. Wehave also provided an efficient discretization scheme, based on finite elements andimplicit/explicit time-stepping schemes, that are used in several numerical experi-ments that illustrate the theoretical results.

28 O. BURKOVSKA AND M. GUNZBURGER

t = 0.02 t = 0.5 t = 1

t = 0.02 t = 0.5 t = 1

Figure 8. Evolution of the nonlocal (Case 2) (bottom) and local(top) solutions of the Cahn-Hilliard variational inequality at dif-ferent time instances and with cF (x) = 0.9ε2cγ(x) for Example 3.

Further study of efficient discretization and approximation techniques for themodel based on, e.g., explicit-implicit, cf. (5.2), or higher-order time-marchingschemes, or even model order reduction approaches, are interesting questions forfuture investigation.

We also note that local phase-field models are often used as a substitute of thesharp-interface models in applications such as, e.g., solidification. The obviousadvantage of using a phase-field model is that it avoids explicit tracking of theinterface. However, the interface width in the corresponding local discrete modelsis limited by the mesh discretization, which can lead to requiring excessively finemeshes in implementations and thus to a high computational cost. The nonlocalinterface model, when compared to the local one in this context, could offer muchgreater flexibility and a promising alternative. Here, we do not need to compromiseon the sharpness of the interface, which can be chosen independently of the meshwidth, and allow for higher fidelity simulations on coarser grids.

References

[1] M. Ainsworth and Z. Mao, Analysis and Approximation of a Fractional Cahn–Hilliard

Equation, SIAM Journal on Numerical Analysis, 55 (2017), pp. 1689–1718.

[2] G. Akagi, G. Schimperna, and A. Segatti, Fractional Cahn-Hilliard, Allen-Cahn andporous medium equations, Journal of Differential Equations, 261 (2016), pp. 2935 – 2985.

[3] P. W. Bates, On some nonlocal evolution equations arising in materials science, Nonlineardynamics and evolution equations, 48 (2006), pp. 13–52.

[4] P. W. Bates and A. Chmaj, An Integrodifferential Model for Phase Transitions: Stationary

Solutions in Higher Space Dimensions, Journal of Statistical Physics, 95 (1999), pp. 1119–1139.

[5] P. W. Bates, P. C. Fife, X. Ren, and X. Wang, Traveling waves in a convolution model

for phase transitions, Archive for Rational Mechanics and Analysis, 138 (1997), pp. 105–136.[6] P. W. Bates and J. Han, The Neumann boundary problem for a nonlocal Cahn-Hilliard

equation, Journal of Differential Equations, 212 (2005), pp. 235–277.

[7] A. L. Bertozzi, S. Esedoglu, and A. Gillette, Inpainting of Binary Images Using theCahn–Hilliard Equation, IEEE Transactions on Image Processing, 16 (2007), pp. 285–291.

[8] L. Blank, M. Butz, and H. Garcke, Solving the Cahn-Hilliard variational inequality with

a semi-smooth Newton method, ESAIM: COCV, 17 (2011), pp. 931–954.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 29

[9] J. F. Blowey and C. M. Elliott, The Cahn-Hilliard gradient theory for phase separation

with non-smooth free energy Part I: Mathematical analysis, European Journal of Applied

Mathematics, 2 (1991), p. 233–280.[10] , The Cahn-Hilliard gradient theory for phase separation with non-smooth free energy

Part II: Numerical analysis, European Journal of Applied Mathematics, 3 (1992), p. 147–179.[11] J. Bosch, M. Stoll, and P. Benner, Fast solution of Cahn–Hilliard variational inequalities

using implicit time discretization and finite elements, Journal of Computational Physics, 262

(2014), pp. 38 – 57.[12] S. C. Brenner, A. E. Diegel, and L. Y. Sung, A Robust Solver for a Mixed Finite Element

Method for the Cahn–Hilliard Equation, Journal of Scientific Computing, 77 (2018), pp. 1234–

1249.[13] O. Burkovska and M. Gunzburger, Regularity analyses and approximation of nonlocal

variational equality and inequality problems, Journal of Mathematical Analysis and Applica-

tions, 478 (2019), pp. 1027–1048.[14] J. W. Cahn and J. E. Hilliard, Free energy of a nonuniform system. I. Interfacial free

energy, The Journal of chemical physics, 28 (1958), pp. 258–267.

[15] X. Chen, Existence, uniqueness, and asymptotic stability of traveling waves in nonlocalevolution equations, Adv. Differential Equations, 2 (1997), pp. 125–160.

[16] X. Chen and M. Gunzburger, Continuous and discontinuous finite element methods for aperidynamics model of mechanics, Computer Methods in Applied Mechanics and Engineering,

200 (2011), pp. 1237–1250.

[17] A. Chmaj and X. Ren, Homoclinic solutions of an integral equation: Existence and stability,Journal of Differential Equations, 155 (1999), pp. 17–43.

[18] P. Colli, S. Frigeri, and M. Grasselli, Global existence of weak solutions to a nonlocal

Cahn-Hilliard-Navier-Stokes system, Journal of Mathematical Analysis and Applications, 386(2012), pp. 428–444.

[19] P. Colli, G. Gilardi, and J. Sprekels, Well-posedness and regularity for a generalized

fractional Cahn-Hilliard system, arXiv e-prints, (2018).[20] P. Colli, P. Krejcı, E. Rocca, and J. Sprekels, Nonlinear evolution inclusions arising

from phase change models, Czechoslovak Mathematical Journal, 57 (2007), pp. 1067–1098.

[21] C. Cortazar, M. Elgueta, J. D. Rossi, and N. Wolanski, Boundary fluxes for nonlocaldiffusion, Journal of Differential Equations, 234 (2007), pp. 360–390.

[22] C. Cortazar, M. Elgueta, J. D. Rossi, and N. Wolanski, How to approximate the heatequation with Neumann boundary conditions by nonlocal diffusion problems, Archive for

Rational Mechanics and Analysis, 187 (2008), pp. 137–156.

[23] E. Davoli, H. Ranetbauer, L. Scarpa, and L. Trussardi, Degenerate nonlocal Cahn-Hilliard equations: Well-posedness, regularity and local asymptotics, Annales de l’Institut

Henri Poincare (C) Analyse Non Lineaire, (2019), pp. 1–28.

[24] E. Davoli, L. Scarpa, and L. Trussardi, Local asymptotics for nonlocal convective Cahn-Hilliard equations with W 1,1 kernel and singular potential, Journal of Differential Equations,

289 (2021), pp. 35–58.

[25] M. D’Elia, X. Tian, and Y. Yu, A physically consistent, flexible, and efficient strategyto convert local boundary conditions into nonlocal volume constraints, SIAM Journal on

Scientific Computing, 42 (2020), pp. A1935–A1949.

[26] W. Deng, B. Li, W. Tian, and P. Zhang, Boundary Problems for the Fractional andTempered Fractional Operators, Multiscale Modeling & Simulation, 16 (2018), pp. 125–149.

[27] S. Dipierro, X. Ros-Oton, and E. Valdinoci, Nonlocal Problems with Neumann BoundaryConditions, Revista Mathematica Iberoamericana, 33 (2017), pp. 377–416.

[28] Q. Du, Nonlocal Modeling, Analysis, and Computation, Society for Industrial and Applied

Mathematics, Philadelphia, PA, 2019.[29] Q. Du and X. Feng, Chapter 5 - The phase field method for geometric moving interfaces

and their numerical approximations, in Geometric Partial Differential Equations - Part I,A. Bonito and R. H. Nochetto, eds., vol. 21 of Handbook of Numerical Analysis, Elsevier,2020, pp. 425–508.

[30] Q. Du, M. Gunzburger, R. B. Lehoucq, and K. Zhou, Analysis and approximation of

nonlocal diffusion problems with volume constraints, SIAM Rev., 54 (2012), pp. 667–696.[31] Q. Du, L. Ju, X. Li, and Z. Qiao, Stabilized linear semi-implicit schemes for the nonlocal

Cahn–Hilliard equation, Journal of Computational Physics, 363 (2018), pp. 39 – 54.[32] Q. Du, L. Ju, X. Li, and Z. Qiao, Maximum Principle Preserving Exponential Time Differ-

encing Schemes for the Nonlocal Allen–Cahn Equation, SIAM Journal on Numerical Analysis,

57 (2019), pp. 875–898.[33] Q. Du, L. Ju, and J. Lu, A discontinuous Galerkin method for one-dimensional time-

dependent nonlocal diffusion problems, Mathematics of Computation, 88 (2019), pp. 123–147.

30 O. BURKOVSKA AND M. GUNZBURGER

[34] Q. Du, L. Ju, J. Lu, and X. Tian, A discontinuous Galerkin method with penalty for

one-dimensional nonlocal diffusion problems, Communications on Applied Mathematics and

Computation, 2 (2020), pp. 31–55.[35] Q. Du and J. Yang, Asymptotically Compatible Fourier Spectral Approximations of Nonlocal

Allen–Cahn Equations, SIAM Journal on Numerical Analysis, 54 (2016), pp. 1899–1919.[36] Q. Du and J. Yang, Fast and accurate implementation of Fourier spectral approximations

of nonlocal diffusion operators and its applications, Journal of Computational Physics, 332

(2017), pp. 118–134.[37] Q. Du and X. Yin, A conforming DG method for linear nonlocal models with integrable

kernels, Journal of Scientific Computing, 80 (2019), pp. 1913–1935.

[38] M. D’Elia, Q. Du, C. Glusa, M. Gunzburger, X. Tian, and Z. Zhou, Numerical methodsfor nonlocal and fractional models, Acta Numerica, 29 (2020), p. 1–124.

[39] P. Fife, Some nonclassical trends in parabolic and parabolic-like evolutions, in Trends in

Nonlinear Analysis, M. Kirkilionis, S. Kromker, R. Rannacher, and F. Tomi, eds., Berlin,Heidelberg, 2003, Springer Berlin Heidelberg, pp. 153–191.

[40] P. C. Fife, Travelling waves for a nonlocal double-obstacle problem, European Journal of

Applied Mathematics, 8 (1997), p. 581–594.[41] M. Fritz, E. A. B. F. Lima, J. Tinsley Oden, and B. Wohlmuth, On the unsteady

Darcy–Forchheimer–Brinkman equation in local and nonlocal tumor growth models, Mathe-matical Models and Methods in Applied Sciences, 29 (2019), pp. 1691–1731.

[42] H. Gajewski and K. Zacharias, On a nonlocal phase separation model, Journal of Mathe-

matical Analysis and Applications, 286 (2003), pp. 11–31.[43] C. G. Gal, A. Giorgini, and M. Grasselli, The nonlocal Cahn-Hilliard equation with

singular potential: Well-posedness, regularity and strict separation property, Journal of Dif-

ferential Equations, 263 (2017), pp. 5253–5297.[44] C. G. Gal and M. Grasselli, Longtime behavior of nonlocal Cahn-Hilliard equations, Dis-

crete and Continuous Dynamical Systems- Series A, 34 (2014), pp. 145–149.

[45] H. Garcke, K. F. Lam, E. Sitka, and V. Styles, A Cahn–Hilliard–Darcy model for tumourgrowth with chemotaxis and active transport, Mathematical Models and Methods in Applied

Sciences, 26 (2016), pp. 1095–1148.

[46] G. Giacomin and J. L. Lebowitz, Phase segregation dynamics in particle systems with longrange interactions. I. Macroscopic limits, Journal of Statistical Physics, 87 (1997), pp. 37–61.

[47] G. Giacomin and J. L. Lebowitz, Phase segregation dynamics in particle systems with long

range interactions II: Interface motion, SIAM Journal on Applied Mathematics, 58 (1998),pp. 1707–1729.

[48] C. Graßle and M. Hinze, POD reduced-order modeling for evolution equations utilizingarbitrary finite element discretizations, Advances in Computational Mathematics, 44 (2018),

pp. 1941–1978.

[49] C. Graßle, M. Hintermuller, M. Hinze, and T. Keil, Simulation and control of a non-smooth Cahn-Hilliard Navier-Stokes system, arXiv:1907.04285, (2019).

[50] Z. Guan, J. Lowengrub, and C. Wang, Convergence analysis for second-order accurate

schemes for the periodic nonlocal Allen-Cahn and Cahn-Hilliard equations, MathematicalMethods in the Applied Sciences, 40 (2017), pp. 6836–6863.

[51] Z. Guan, J. S. Lowengrub, C. Wang, and S. M. Wise, Second order convex splitting

schemes for periodic nonlocal Cahn-Hilliard and Allen-Cahn equations, Journal of Compu-tational Physics, 277 (2014), pp. 48–71.

[52] Z. Guan, C. Wang, and S. M. Wise, A convergent convex splitting scheme for the periodic

nonlocal Cahn-Hilliard equation, Numerische Mathematik, 128 (2014), pp. 377–406.[53] M. Hintermuller, K. Ito, and K. Kunisch, The primal-dual active set strategy as a semis-

mooth Newton method, SIAM Journal on Optimization, 13 (2002), pp. 865–888.[54] M. Hintermuller, M. Hinze, and M. H. Tber, An adaptive finite-element Moreau–Yosida-

based solver for a non-smooth Cahn–Hilliard problem, Optimization Methods and Software,26 (2011), pp. 777–811.

[55] T. Mengesha and Q. Du, Analysis of a scalar peridynamic model with a sign changing

kernel, Discrete Contin. Dynam. Systems B, 18 (2013), pp. 1415–1437.

[56] T. Mengesha and Q. Du, On the variational limit of a class of nonlocal functionals relatedto peridynamics, Nonlinearity, 28 (2015), pp. 3999–4035.

[57] A. Miranville, The Cahn-Hilliard equation and some of its variants, AIMS Mathematics,2 (2017), pp. 479–544.

[58] J. T. Oden, A. Hawkins, and S. Prudhomme, General diffuse-interface theories and an

approach to predictive tumor growth modeling, Mathematical Models and Methods in AppliedSciences, 20 (2010), pp. 477–517.

NONLOCAL CAHN-HILLIARD MODEL PERMITTING SHARP INTERFACES 31

[59] T. Ohta and K. Kawasaki, Equilibrium morphology of block copolymer melts, Macro-

molecules, 19 (1986), pp. 2621–2632.

[60] B. Ren, C. Wu, and E. Askari, A 3D discontinuous Galerkin finite element method withthe bond-based peridynamics model for dynamic brittle failure analysis, International Journal

of Impact Engineering, 99 (2017), pp. 14–25.[61] Y. Tao, X. Tian, and Q. Du, Nonlocal diffusion and peridynamic models with Neumann type

constraints and their numerical approximations, Applied Mathematics and Computation, 305

(2017), pp. 282–298.[62] X. Tian and Q. Du, Nonconforming Discontinuous Galerkin Methods for Nonlocal Varia-

tional Problems, SIAM Journal on Numerical Analysis, 53 (2015), pp. 762–781.

[63] , Asymptotically compatible schemes for robust discretization of parametrized problemswith applications to nonlocal models, SIAM Review, 62 (2020), pp. 199–227.

[64] F. Troltzsch, Optimal control of partial differential equations: theory, methods, and appli-

cations, vol. 112, American Mathematical Soc., 2010.[65] S. Wise, J. Lowengrub, H. Frieboes, and V. Cristini, Three-dimensional multispecies

nonlinear tumor growth—I: Model and numerical method, Journal of Theoretical Biology,

253 (2008), pp. 524 – 543.[66] You, Huaiqian, Lu, Xin Yang, Trask, Nathaniel, and Yu, Yue, An asymptotically

compatible approach for Neumann-type boundary condition on nonlocal problems, ESAIM:M2AN, 55 (2021), pp. S811–S851.