ArticlePDF Available

An upwind DG scheme preserving the maximum principle for the convective Cahn-Hilliard model

Authors:

Abstract and Figures

The design of numerical approximations of the Cahn-Hilliard model preserving the maximum principle is a challenging problem, even more if considering additional transport terms. In this work, we present a new upwind discontinuous Galerkin scheme for the convective Cahn-Hilliard model with degenerate mobility which preserves the maximum principle and prevents non-physical spurious oscillations. Furthermore, we show some numerical experiments in agreement with the previous theoretical results. Finally, numerical comparisons with other schemes found in the literature are also carried out
This content is subject to copyright. Terms and conditions apply.
https://doi.org/10.1007/s11075-022-01355-2
ORIGINAL PAPER
An upwind DG scheme preserving the maximum
principle for the convective Cahn-Hilliard model
Daniel Acosta-Soba1,2 ·Francisco Guill´
en-Gonz´
alez3·
J. Rafael Rodr´
ıguez-Galv´
an1
Received:16 November 2021 / Accepted:13 June 2022 /
©The Author(s) 2022
Abstract
The design of numerical approximations of the Cahn-Hilliard model preserving the
maximum principle is a challenging problem, even more if considering additional
transport terms. In this work, we present a new upwind discontinuous Galerkin
scheme for the convective Cahn-Hilliard model with degenerate mobility which
preserves the maximum principle and prevents non-physical spurious oscillations.
Furthermore, we show some numerical experiments in agreement with the previous
theoretical results. Finally, numerical comparisons with other schemes found in the
literature are also carried out.
Keywords Discontinuous Galerkin ·Upwind scheme ·Diffuse interface ·
Convection ·Degenerate mobility
Daniel Acosta-Soba, Francisco Guill´
en-Gonz´
alez and J. Rafael Rodr´
ıguez-Galv´
an contributed
equally to this work.
J. Rafael Rodr´
ıguez-Galv´
an
rafael.rodriguez@uca.es
Daniel Acosta-Soba
daniel.acosta@uca.es
Francisco Guill´
en-Gonz´
alez
guillen@us.es
1Departamento de Matem´
aticas, Universidad de C´
adiz, Campus Universitario de Puerto Real
S/N, Puerto Real, 11510, C´
adiz, Spain
2Department of Mathematics, University of Tennessee at Chattanooga, 615 McCallie Ave,
Chattanooga, 37403, TN, USA
3Departamento de Ecuaciones Diferenciales y An´
alisis Num´
erico & IMUS, Universidad de
Sevilla, Campus de Reina Mercedes S/N, Sevilla, 41012, Spain
Published online: 12 August 2022
Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1 Introduction
This paper is concerned with the development of discontinuous Galerkin (DG)
numerical schemes for the following convective Cahn-Hilliard (CCH) problem (writ-
ten as a second-order system): Given an incompressible velocity field vC()d
with ∇·v=0in, such that v·n=0on∂, find two real valued functions, the
phase uand the chemical potential μ, defined in ×[0,T]such that:
tu=1
Pe ∇·(M(u)μ)−∇·(uv)in ×(0,T), (1a)
μ=F(u) ε2u in ×(0,T), (1b)
u·n=(M(u)μuv)·n=0on∂ ×(0,T), (1c)
u(0)=u0in . (1d)
The phase field variable ulocalizes the two different phases at the values u=0and
u=1, while the interface occurs when 0 <u<1. Here is a bounded polygonal
domain in Rd(d=2 or 3 in practice), with boundary ∂ whose outward-pointing
unit normal is denoted n. Parameters are Pe >0theP
´
eclet number of the flow and
ε>0 related to the width interface. For simplicity, we take Pe =1. The given initial
phase is denoted by u0. The classical Cahn–Hilliard equation (CH) corresponds to
the case without convection, i.e. v=0.
We consider the double-well Ginzburg-Landau potential
F (u) =1
4u2(1u)2
and the degenerate mobility
M(u) =u(1u).(2)
This type of degenerate mobility, vanishing at the pure phases u=0andu=1,
implies the following maximum principle property (see [16] for the CH equation): if
0u01inthen 0 u(·,t) 1infor t(0,T). This property does not
hold for constant mobility, and in general for fourth-order parabolic equations. We
are concerned in numerical schemes maintaining this property at the discrete level.
Cahn-Hilliard equation was originally introduced in [8,9] as a phenomenologi-
cal model of phase separation in a binary alloy and it has been successfully applied
in different contexts as a model which characterizes phase segregation and inter-
face dynamics. Applications include tumor tissues [34,35], image processing [6]and
multi-phase fluid systems (see, e.g. the review [26]). The dynamics of this equation
comprises a first stage in which a rapid separation process takes place, leading to
the creation of interfaces, followed by a second stage where aggregation and devel-
opment of bulk phases separated by thin diffuse interfaces take place. These two
phenomena (separation and aggregation) are characterized by different temporal and
spatial scales which, together with the non-linear and fourth-order nature of CH,
makes efficiently solving this equation an interesting computational challenge.
The interface is represented as a layer of small thickness of order εand the auxil-
iary function u(the phase-field function) is used to localize the (pure) phases u=0
and u=1. The chemical potential μis the variational derivative of the Helmholtz
1590 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
free energy E(u) with respect to u,μ=uE(u),where
E(u) =ε2
2|∇u|2+F (u)dx.(3)
Then, the dynamics of CH correspond to the physical energy dissipation of ut=
−∇ · J (u),whereJ (u) =−M(u)μ=−M (u)uE(u). For existence of solution
for CH equation with a degenerate mobility of the type (2), besides bounds for the
phase over time, we can refer to [16]. A review and many variants can be read in [30].
Numerical approximation of CH is a research topic of great interest due to the wide
spectrum of its applications, as well as a source of interesting mathematical chal-
lenges due to the computational issues referred above. Time approximations include
both linear schemes [21] and non-linear ones, where a Newton method is usually
employed. Although some authors (see, e.g. [15]) consider the fourth-order equation
obtained by substitution in the equation (1a) of the expression of μgivenin(1b),
the spatial discretization is usually based on the mixed phase field/chemical poten-
tial formulation presented in (1). These equations can be approximated by classical
numerical methods, including (a) finite differences (see, e.g. [12,20] for constant
mobility M(u) 1and[25] for degenerate mobility similar to (2), see also [32]for
applications in crystallography); (b) finite volumes (see, e.g. [4,13] for degenerate
mobility schemes); (c) and, above all, finite elements (see the precursory papers [15]
(constant mobility) and [5] (degenerate mobility)).
In recent years, an increasing number of advances has been published explor-
ing the use of discontinuous Galerkin (DG) methods both for constant [2,24]and
degenerate mobility [18,28,33,36]. Although DG methods lead in general to more
complex algorithms and to bigger amounts of degrees of freedom, they exhibit some
benefits of which one can take advantage also in CH equations, for instance doing
mesh refining and adaptivity [2]. In this paper, we are interested in the follow-
ing relevant point: the possibility of designing conservative schemes preserving the
maximum-principle 0 u(t , x) 1 at the discrete level, even for the CCH variant
of CH equations, namely where a convective term models the convection or transport
of the phase-field.
The idea of introducing a convection term in the CH equations modeling a phase-
field driven by a flow, arriving at the CCH problem (1), arouses great interest,
specially in the case where the phase field is coupled with fluid equations.
To design adequate numerical approximations of these CCH equations is an
extremely challenging problem because it adds the hyperbolic nature of convective
terms to the inherent difficulties of the CH equation. There are not many numeri-
cal methods in the literature on this topic, although some interesting contributions
can be found. The first of them [3] is worth mentioning for the application of high-
resolution spectral Fourier schemes. We also underline the papers [7,27], based on
finite volume approximations, and also [11,22], where finite difference techniques
are applied for Navier-Stokes CH equations.
Some authors have recently worked in DG methods for spatial discretization
of the CCH problem. Using DG schemes in this context is natural because they
1591Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
are well-suited in convection-dominated problems, even when the P´
eclet number is
substantially large. For instance, the work of [24] is focused on construction and
convergence analysis of a DG method for the CCH equations with constant mobil-
ity, applying an interior penalty technique to the second-order terms and an upwind
operator for discretization of the convection term. Authors of [18] consider CCH
with degenerate mobility applying also an interior penalty to second-order terms in
the mixed form (1) and a more elaborated upwinding technique, based on a sigmoid
function, to the convective term.
These previous works show that DG methods are well suited for the approximation
of CCH equations, obtaining optimal convergence order and maintaining most of
the properties of the continuous model. But they have room for improvement in one
specific question: getting a maximum principle for the phase in the discrete case.
Although there are some works in which the maximum principle for the CH model
is preserved at the discrete level using finite volumes [4] and flux limiting for DG
[19], to the best knowledge of the authors, no scheme has been published in which
an upwind DG technique is used to obtain a discrete maximum principle property.
Our main contribution in this paper is made in this regard: the development of
numerical schemes which guarantee punctual estimates of the phase at the discrete
level, in addition to maintaining the rest of continuous properties of (1). The main
idea is to introduce an upwind DG discretization associated to the transport of the
phase by the convective velocity v, but also associated to the degenerate mobility
M(u). The main result in this work is Theorem 3.11, where we show that, for a piece-
wise constant approximation of the phase, our DG scheme preserves the maximum
principle, that is the discrete phase is also bounded in [0,1].
Our scheme is specifically designed for the CH equation with non-singular (poly-
nomial) chemical potential and degenerate mobility. For the CH equation with
the logarithmic Flory-Huggins potential, a strict maximum principle is satisfied
without the need of degenerate mobility. In this case, designing maximum principle-
preserving numerical schemes is a different issue. See, for instance, the paper [10]
where a finite difference numerical is proposed in this regard.
The paper is organized as follows: In Section 2, we fix notation and review DG
techniques for discretization of conservative laws. We consider the linear convec-
tion given by a velocity field and show that the usual linear upwind DG numerical
scheme with P0approximation preserves positivity in general, and the maximum
principle in the divergence-free case. In Section 3, we consider the CCH problem
(1). Using a truncated potential and a convex-splitting time discretization, we obtain
a scheme satisfying an energy law, which is decreasing for the CH case (v=0).
Then, we introduce our fully discrete scheme (29) providing a maximum principle
for the discrete phase variable. Finally, in Section 4, we present several numerical
tests, comparing our DG scheme with two different space discretizations found in the
literature: classical continuous P1finite elements and the SIP+upwind sigmoid DG
approximation proposed in [18]. We show error order tests and also we present qual-
itative comparisons where the maximum principle of our scheme is confirmed (and
it is not conserved by the other ones).
1592 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2 DG discretization of conservative laws
Throughout this section, we are going to study the problem of approximating linear
and non-linear conservative laws preserving the positivity of the continuous models.
To this purpose, we introduce an upwinding DG scheme which follows the ideas of
the paper [23] based on an upwinding finite volume method.
2.1 Notation
Firstly, we are going to set the notation that will be used from now on. Let Rd
be a bounded polygonal domain. Then, we consider a shape-regular triangular mesh
Th={K}KThof size hover , and we denote as Ehthe set of the edges of Th(faces
if d=3) distinguishing between the interior edges Ei
hand the boundary edges Eb
h.
Then, Eh=Ei
hEb
h.
AccordingtoFig.1, we will consider for an interior edge eEi
hshared by the
elements Kand L,i.e.e=∂K ∂L, that the associated unit normal vector neis
exterior to Kpointing to L. Moreover, for the boundary edges eEb
h, the unit normal
vector newill be pointing outwards of the domain .
Therefore, we can define the average {{·}} and the jump [[· ]] of a scalar function v
on an edge eEhas follows:
{{v}} := vK+vL
2if e=∂K ∂LEi
h
vKif e=∂KEb
h
,
[[ v]] := vKvLif e=∂K ∂LEi
h
vKif e=∂KEb
h
.
We will denote by Pdisc
k(Th)and Pcont
k(Th)the spaces of discontinuous and
continuous functions, respectively, whose restriction to the elements Kof Thare
polynomials of degree k0.
Moreover, we will take an equispaced partition 0 =t0<t
1<···<t
N=Tof
the time domain [0,T]with t =tm+1tmthe time step.
Fig. 1 Orientation of the unit
normal vector ne
1593Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Finally, we set the following notation for the positive and negative parts of a scalar
function v:
v:= |v|+v
2=max {v, 0},v
:= |v|v
2=−min {v, 0},v=vv.
2.2 Conservative laws
We consider the following non-linear conservative problem for v:
vt+∇·F(v) =0en×(0,T), (4)
where the flux F(·)is a vectorial continuous function.
Let Sh=Pdisc
k(Th). Multiplying by any vShand using the Green Formula in
each element KTh:
tv v
KThK
F(v) ·∇v+
KTh∂K F(v) ·nKv=0,
where nKis the normal vector to ∂K pointing outwards K.Ifvis a strong solution
of (4), such that vis continuous over eEh,weget:
tv v
KThK
F(v) ·∇v+
eEheF(v) ·ne[[ v]] =0. (5)
But, if we look for functions vShwhich are discontinuous over eEh, then the
term F(v) ·nein the last integral of (5) is not well defined and we need to approach it.
Hence, the concept of numerical flux is used, taking expressions like (vK,v
L,ne)
such that: eF(v) ·ne[[ v]] e
(vK,v
L,ne)[[ v]] .
From now on, we will consider fluxes of the form
F(v) =M(v)β,
with M=M(v) Rand β=β(x, t ) Rd. Since in this paper we are interested in
studying conservative problems defined over isolated domains, we impose from now
on that the transport vector βsatisfies the slip condition
β·n=0on∂.(6)
Remark 2.1 We will have to take into consideration the sign of M(v),i.e.ifM(v)
is non-increasing or non-decreasing, in order to work out a well-suited upwinding
scheme. This is due to
∇·F(v) =∇·(M(v)β)=M(v)β·∇v+M(v)∇·β.(7)
where
The first term M(v)β·∇vis a transport of vin the direction of M(v)β.
The second term M(v)∇·βcan be seen as a reaction term with coefficient ∇·β.
It may be specially interesting the case β=−v, where the whole term ∇·
(M(v)β)=−∇·(M(v)v) is a non-linear diffusion term.
1594 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2.3 Linear convection and positivity
Taking M(v) =v, we arrive at the linear conservative problem:
vt+∇·(v β)=0in×(0,T), (8a)
v(0)=v0in , (8b)
where β:Rdis continuous.
Remark 2.2 The problem (8) is well-posed without imposing boundary conditions
due to the slip condition (6) (it can be derived writing an integral expression of solu-
tion in terms of the characteristic curves, see for example [31]). In particular, this
integral expression implies the positivity of the solution, namely, if v00in,then
the solution vof (8) satisfies v0in×(0,T).
For the linear problem (8a), we introduce the upwind numerical flux
(vK,v
L,ne):= (β·ne)vK(β·ne)vL,(9)
which leads to the following discrete scheme for (8a): Given vmSh,findvm+1
Shsolving
δtvm+1v+aupw
h(β;vm+1
,v) =0,vSh,(10)
where
aupw
h(β;v, v) :=
KThK
v(β·∇v)
+
eEi
h,e=KLe
((β·ne)vK(β·ne)vL)[[v]] (11)
and δtvm+1=(vm+1vm)/t denotes a discrete time derivative.
Notice that we have truncated vby its positive part v, taking into account that
the solution of the continuous model (8a) is positive.
Remark 2.3 The numerical flux (vK,v
L,ne)giving in (9) can be rewritten as
follows:
(vK,v
L,ne)=(β·ne){{v}} +1
2|β·ne|[[ v]] ,
where (β·ne){{v}} is a centered-flux term and 1
2|β·ne|[[ v]] is the upwind term.
Now, we will prove that if we set Sh=Pdisc
0(Th), the scheme (10) preserves the
positivity of the continuous model. In this case, since KThK(β·∇v)v =0, the
upwind term (11) reduces to
aupw
h(β;v, v) =
eEi
h,e=KLe
((β·ne)vK(β·ne)vL)[[ v]] . (12)
1595Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Theorem 2.4 (DG scheme (10) preserves positivity) If Sh=Pdisc
0(Th), then the
scheme (10)preserves positivity, that is, for any vmShwith vm0in , then any
solution vm+1of (10)satisfies vm+10in .
Proof Taking the following test function
v=(vm+1
K)in K
0 out of K, v Sh,
where Kis an element of Thsuch that vm+1
K=minKThvm+1
K, the scheme (10)
becomes Kδtvm+1
K(vm+1
K)=−aupw
h(β;vm+1
,(v
m+1
K)). (13)
By applying for all LTh,vm+1
Lvm+1
Khence in particular (vm+1
L)
(vm+1
K)(the positive part is a non-decreasing function), we deduce
aupw
h(β;vm+1
,(v
m+1
K))=
=
eEi
h,e=KLe(β·ne)(vm+1
K)(β·ne)(vm+1
L)(vm+1
K)
eEi
h,e=KLe(β·ne)(vm+1
K)(β·ne)(vm+1
K)(vm+1
K).
Since (vm+1
K)(vm+1
K)=0then
aupw
h(β;vm+1
,(v
m+1
K))0.
Therefore, from (13),
|K|δtvm+1
K(vm+1
K)0.
On the other hand,
0≤|K|δtvm+1
K(vm+1
K)=−|K|
t (vm+1
K)2
+vm
K(vm+1
K)0;
hence, (vm+1
K)=0. Thanks to the choice of K, we can assure that vm+10.
Theorem 2.5 (Existence of DG scheme (10)) Assume that Sh=Pdisc
0(Th). For any
vmSh, there is at least one solution vm+1Shof the scheme (10).
Proof Consider the following well-known theorem:
Theorem 2.6 (Leray-Schauder fixed point theorem) Let Xbe a Banach space
and let T:X−→ Xbe a continuous and compact operator. If the set
{xX:x=αT(x) for some 0α1}
is bounded (with respect to α), then Thas at least one fixed point.
1596 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Given a function wPdisc
0(Th)with w0, we are going to define the operator
T:Pdisc
0(Th)−→ Pdisc
0(Th)such that T(v) =vwhere vis the unique solution of the
linear scheme:
vPdisc
0(Th),
vw
t v=−aupw
h(β;v,v), vPdisc
0(Th). (14)
The idea is to use the Leray-Schauder fixed point Theorem 2.6 to the operator T.
First of all, Tis well defined.
Secondly, we will check that Tis continuous. Let {vj}jNPdisc
0(Th)be a
sequence such that limj→∞vj=v. Taking into account that all norms are equiv-
alent in Pdisc
0(Th)since it is a finite-dimensional space, the convergence vjvis
equivalent to the convergence elementwise (vj)KvKfor every KTh(this may
be seen, for instance, by using the norm ·
L()). Taking limits when j→∞in
the scheme (14) (with v:= vjand v:= T(vj)), and using the notion of convergence
elementwise, we get that
lim
j→∞ T(vj)=T(v) =Tlim
j→∞vj;
hence, Tis continuous.
In particular, as Tis a continuous operator defined over Pdisc
0(Th), which is finite-
dimensional, Tis also compact.
Finally, let us prove that the set
B={vPdisc
0(Th):v=αT(v) for some 0 α1}
is bounded. The case α=0 is trivial so we will assume that α(0,1].IfvB,
then vis the solution of the scheme
vPdisc
0(Th),
vαw
t v=−αa
upw
h(β;v,v), vPdisc
0(Th). (15)
Now, testing (15)byv=1, we get that
v=α
w,
and, since w0, and since it can be proved that v0 using the same arguments
than in Theorem 2.4, we get that
vL1() wL1().
Hence, since Pdisc
0(Th)is a finite-dimensional space where the norms are equiva-
lent,wehaveprovedthatBis bounded. Thus, using the Leray-Schauder fixed point
Theorem 2.6, there is a solution of the scheme (10).
Let us focus on the the following linear scheme (without truncating vby its pos-
itive part, v): Given vmPdisc
0(Th),findvm+1Pdisc
0(Th)such that, for every
vPdisc
0(Th):
δtvm+1v+aupw
h(β;vm+1,v) =0. (16)
1597Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
It is well-known (see, e.g. [14] and references therein) that (16) has a unique solution.
Since we have shown positivity of the non-linear truncated scheme (10), then any
solution of (10) is the unique solution of (16). This argument implies uniqueness of
(10) and positivity of (16).
Corollary 2.7 (Linear DG scheme (16) preserves positivity) For any vmPdisc
0(Th)
with vm0in , the unique solution vm+1of the linear scheme (16)is positive, i.e.
vm+10in .
2.4 Linear convection with incompressible velocity and maximum principle
Now, let us focus on the particular case of the conservation problem (8)where
β:Rdis continuous and incompressible,i.e.∇·β=0in. In this case,
the solution of the problem (8), satisfies the following maximum principle (it can be
proved, for instance, using the characteristics method as in Remark 2.2):
min
v0vmax
v0in ×(0,T).
We are going to show that the solution of the linear scheme (16) (without trun-
cating v) preserves this maximum principle. The proof is based on the fact that, as
consequence of the divergence theorem, one has
∂K
β·nK=0,KTh. (17)
Proposition 2.8 (Linear DG (16) preserves the maximum principle) For any vm
Pdisc
0(Th), the solution vm+1Pdisc
0(Th)of the scheme (16)satisfies the maximum
principle, that is
min
vmvm+1max
vmin .
Proof First, we will prove that vm+1minvm. Let us denote vmi n =minvm.
Taking the following test function
v=(vm+1
Kvmin )in K
0 out of K,
where Kis an element of Thsuch that the value vm+1
K=minKThvm+1
K, the scheme
(16) becomes
|K|δtvm+1
K(vm+1
Kvmin )=−aupw
h(β;vm+1,(v
m+1
Kvmin )).
1598 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Now, as we have chosen Kwe can assure that vm+1
Lvm+1
Kfor all LTh. Hence,
using (17), one has
aupw
h(β;vm+1,(v
m+1
Kvmin ))=
=
eEi
h,e=KLe(β·ne)vm+1
K(β·ne)vm+1
L(vm+1
Kvmin )
eEi
h,e=KLe(β·ne)vm+1
K(β·ne)vm+1
K(vm+1
Kvmin )
=vm+1
K(vm+1
Kvmin )
eEi
h,e=KLe
(β·ne)=0.
Therefore,
|K|δtvm+1
K(vm+1
Kvmin )0.
Moreover,
0≤|K|tvm+1
K)(vm+1
Kvmin )
=|K|
t (vm+1
Kvmin )+(vmin vm
K)(vm+1
Kvmin )
=|K|
t (vm+1
Kvmin )2
+(vmin vm
K)(vm+1
Kvmin )0,
then we have proved that (vm+1
Kvmin )=0. Hence, from the choice of K,we
can assure vm+1minvm.
Now, we will prove that vm+1maxvm. Let us denote vmax =maxvm.
Taking the following test function
v=(vm+1
Kvmax )in K
0 out of K,
where Kis an element of Thsuch that the value vm+1
K=maxKThvm+1
Kand using
similar arguments to those above, we arrive at
|K|δtvm+1
K(vm+1
Kvmax )0.
Moreover,
0|K|
t (vm+1
Kvmax )2
+(vmax vm
K)(vm+1
Kvmax )
=|K|
t (vm+1
Kvmax )+(vmax vm
K)(vm+1
Kvmax )
=|K|δtvm+1
K(vm+1
Kvmax )0,
then we have proved that (vm+1
Kvmax )=0. From the choice of K, we can assure
that vm+1maxvm.
1599Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3 Cahn-Hilliard with degenerate mobility and incompressible
convection
At this point, given v:×(0,T) −→ Rda continuous incompressible velocity field
satisfying the slip condition (6), we are in position to consider the CCH problem (1).
Remark 3.1 Any smooth enough solution (u, μ) of the CCH model (1) satisfies the
maximum principle 0 u1in×(0,T)whenever 0 u01in. The proof
of this statement is a straightforward consequence of Remark 2.2.
To prove that u0, it is enough to notice that uis the solution of (8) with
β:= (1u)μ+vandtouseRemark2.2.
To check that u1 we make the change of variables w:= 1uand, using that
∇·v=0, notice that wis the solution of (8) with β:= (1w)(μ+v). Then,
we just use Remark 2.2.
Owing to this maximum principle, the following C2(R)truncated potential is
considered
F (u) := 1
4
u2u<0,
u2(1u)2u∈[0,1],
(u 1)2u>1.
(18)
This truncated potential will allow us to define a linear time discrete convex-splitting
scheme satisfying an energy law (where the energy is decreasing in the case v=0)
(see Section 3.1).
Assume 0 u01in. The weak formulation of problem (1) consists of
finding (u, μ) such that, u(t) H1(),μ(t ) H1() with tu(t ) H1(),
M(u(t))μ(t ) L2() a.e. t(0,T), and satisfying the following variational
problem a.e. t(0,T)for every μ, uH1():
tu(t), μ=(M(u(t ))μ(t) u(t)v(t), μ)(L2())d,(19a)
(μ(t ), u)L2() =ε2(u(t), u)(L2())d+(F (u(t )), u)L2(),(19b)
with the initial condition u(0)=u0in .
Remark 3.2 By taking μ=1in(19a), any solution uof (19) conserves the mass,
because
d
dt
u(x, t )dx =0.
Remark 3.3 By taking μ=μ(t) and u=tu(t ) in (19), and adding the resulting
expressions, one has that any solution (u, μ) of (19) satisfies the following energy
law
d
dt E(u(t )) +
M(u(x, t ))|∇μ(x, t )|2dx =
u(x, t ) v(x, t ) ·∇μ(x, t )dx, (20)
1600 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
where E:H1() −→ Ris the Helmholtz free energy
E(u) := ε2
2|∇u|2+F (u)dx. (21)
Indeed, by applying the chain rule we get
d
dt E(u(t )) =δE
δu (u(t)), tu(t)=
μ(x, t )∂tu(x, t )dx
=−
M(u(x, t )) |μ(x, t )|2dx +
u(x, t ) v(x, t ) ·∇μ(x, t )dx.
In particular, in the CH case (v=0), the energy E(u(t)) is dissipative. Contrarily,
to the best knowledge of the authors, for the CCH problem with v= 0, there is no
evidence of the existence of a dissipative energy.
3.1 Convex splitting time discretization
Now, we are ready to focus on a convex splitting time discretization (of Eyre’s type
[17])of(19). Specially, we decompose the truncated potential (18) as follows:
F (u) =Fi(u) +Fe(u),
where
Fi(u) := 3
8u2,F
e(u) := 1
4
1
2u2u<0,
u42u31
2u2u∈[0,1],
12u1
2u2u>1.
It can be easily proved that Fi(u) is a convex operator, which will be treated implicitly
whereas Fe(u) is a concave operator that will be treated explicitly. Then, we consider
the following convex Ei(u(t)) and concave Ee(u(t)) energy terms:
Ei(u) := ε2
2|∇u(x )|2dx +
Fi(u(x))d x,
Ee(u) :=
Fe(u(x))d x,
such that the free energy (21) is split as E(u) =Ei(u) +Ee(u).
Finally, we define the following time discretization of (19): find um+1H1()
and μm+1H1() such that, for every μ, uH1():
tum+1,μ)L2() =−(M(um+1)μm+1um+1v(tm+1), μ)(L2())d,(22a)
m+1,u)L2() =ε2(um+1,u)(L2())d+(f (um+1,u
m), u)L2(),(22b)
1601Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
where we denote the convex-implicit and concave-explicit linear approximation of
the potential as follows
f(u
m+1,u
m):= F
i(um+1)+F
e(um)
=3
4um+1+1
4
um,u
m(−∞,0),
4(um)36(um)2um,u
m∈[0,1],
(um+2),u
m(1,+∞).
(23)
Notice that the positive part of the mobilitity has been taken in (22a), regarding the
Remark 4, in order to prevent possible overshoots of the solution um+1beyond the
interval [0,1].
3.1.1 Discrete energy law
By adding (22a)and(22b)forμ=μm+1and u=δtum+1in (22), we get:
M(um+1)|∇μm+1|2+ε2(um+1δtum+1)(L2())d
+(f (um+1,u
m), δtum+1)L2() =
um+1v(·,t
m+1)·∇μm+1. (24)
Taking into account that
ε2(um+1
tum+1)(L2())d=ε2
2δtum+12
+2
2δtum+12
,
and by adding and substracting δtF(u
m+1), we get the following equality
δtE(um+1)+
M(um+1)|∇μm+1|2+2
2|δtum+1|2
+(f (um+1,u
m), δtum+1)L2() δt
F(u
m+1)=
um+1v(·,t
m+1)·∇μm+1,
where E(u) is defined in (21).
Then, using the Taylor theorem, we get
Fi(um)=Fi(um+1)+F
i(um+1)(umum+1)+F
i(um+ξ)
2(umum+1)2,
Fe(um+1)=Fe(um)+F
e(um)(um+1um)+F
e(um+η)
2(um+1um)2,
for certain ξ,η (0,1)with um+ξ=ξum+1+(1ξ)um,um+η=ηum+1+(1η)um.
Hence, adding these expressions and taking into consideration that F (u) =Fi(u) +
Fe(u) for every uR, we arrive at
F(u
m+1)F(u
m)=
=f(u
m+1,u
m)(um+1um)F
i(um+ξ)F
e(um+η)
2(um+1um)2.
1602 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Furthermore, as
F
i(um+ξ)F
e(um+η)
2k=
=
1
2k,u
m+η(−∞,0)(1,+∞),
312(um+η)2+12(um+η)+1
8k,u
m+η∈[0,1],
=1
2k0,u
m+η(−∞,0)(1,+∞),
1
2k13um+η(um+η1)0,u
m+η∈[0,1],
we have
F
i(um+ξ)F
e(um+η)
2k0,
and finally
(f (um+1,u
m), δtum+1)L2()
δtF(u
m+1)0.
Therefore, we arrive at the following result:
Theorem 3.4 Any solution of the scheme (22)satisfies the following discrete energy
law
δtE(um+1)+
M(um+1)|∇μm+1|2+2
2|δtum+1|2
um+1v(·,t
m+1)·∇μm+1. (25)
In particular, if v=0the time-discrete scheme is unconditionally energy stable,
because E(um+1)E(um).
3.2 Fully discrete scheme
At this point, we are going to introduce the key idea for the spatial approximation:
to treat (1a) as a conservative problem where there are two different fluxes (linear
and non-linear). In this sense, we propose an upwind DG scheme approximating the
non-linear flux F(u) =M(u)μproperly.
To this aim, we take for values vRthe increasing and decreasing part of M(v),
denoted respectively by M(v) and M(v), as follows:
M(v) =v
0
(s(M(s)))ds =min(v,1)
0
M(s)ds =min(v ,1)
0
(12s)ds,
M(v) =−
v
0
(s(M(s)))ds =−min(v,1)
0
M(s)ds
=−
min(v,1)
0
(12s)ds.
1603Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Therefore,
M(v) =M(v)if v1
2
M1
2if v>1
2
,M
(v) =0ifv1
2
M(v)M1
2if v>1
2
. (26)
Notice that M(v) +M(v) =M(v). We define the following generalized
upwind bilinear form to be applied for the non-linear flux F(u) =M(u)βwhere
now βcan be discontinuous over Ei
h(in fact we will take β=∇μ):
aupw
h(β;M(v),v) :=
(β·∇v)M(v)
+
eEi
h,e=KLe({{β}} ·ne)(M (vK)+M(vL))
({{β}} ·ne)(M (vL)+M(vK))[[ v]] . (27)
Remark 3.5 We refer to (27) as a generalized bilinear form since it generalizes
the definition of (11) considering the case where βmay be discontinuous. If βis
continuous both definitions are equivalent.
Then, we propose the following fully discrete DG+Eyre scheme (named DG-
UPW) for the model (1):
Find um+1Pdisc
0(Th), with μm+1,w
m+1Pcont
1(Th), solving
tum+1,u)L2()
+aupw
h(−∇μm+1;M(um+1),u) +aupw
h(v(tm+1);um+1,u) =0,(28a)
m+1,μ)L2() =ε2(wm+1,μ)L2() +(f (um+1,u
m), μ)L2(),(28b)
(wm+1,w)L2() =(um+1,w)L2(),(28c)
for all uPdisc
0(Th)and μ, wPdisc
1(Th). Following the notation of the Section 2.3,
aupw
h(v;u, u) =
eEi
h,e=KLe
((v·ne)uK(v·ne)uL)[[ u]] .
In this scheme, we have introduced a truncation of the function M(u) taking its
positive part M(u), which is consistent as the solution of the continuous model (1)
satisfies 0 u1.
Notice that we have introduced a new continuous variable wPcont
1(Th)in (28c).
It can be seen as a regularization of the variable uPcont
0(Th), which is used in the
diffusion term in (28b) (which corresponds to the philic term in the energy of the
model (21)). In fact, both variables wm+1and um+1are approximations of u(tm+1).
Remark 3.6 We are using the same notation in the fully discrete scheme (28)than
the one we have used in the time-discrete scheme (3.1) given in the Section 3.1.1,
satisfying an energy law.
1604 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Nevertheless, in this case, we are changing the meaning of the equations since we
are treating (28a)astheu-equation and (28b)astheμ-equation, contrary to computa-
tions done to reach the energy law. This has been done for the purpose of preserving
the maximum principle in the (28a) and adequately approximating the laplacian term
of the (28b).
Remark 3.7 The boundary condition wm+1·n=0on∂ ×(0,T) is imposed
implicitly by the term (wm+1,μ)L2() in (28b).
Remark 3.8 Since f(·,u
m)is linear, we have the following equality of the potential
term of (28b):
(f (wm+1,u
m), μ)L2() =(f (um+1,u
m), μ)L2().
Remark 3.9 The scheme (28) is non-linear; hence, we will have to use an iterative
procedure, the Newton’s method, to approach its solution.
Proposition 3.10 The scheme (28)conserves the mass of both um+1and wm+1
variables:
um+1=
umand
wm+1=
wm.
Proof Just need to take u=1in(28a)andw=1in(28c).
Theorem 3.11 (DG (28) preserves the maximum principle) For any umPdisc
0(Th)
with 0um1in , then any solution um+1of (28)satisfies 0um+11in .
Proof Firstly, we prove that um+10. Taking the following Pdisc
0(Th)test function
u=(um+1
K)in K
0 out of K,
where Kis an element of Thsuch that um+1
K=minKThum+1
K,(28a) becomes
|K|δtum+1
K(um+1
K)=
=−aupw
h−∇μm+1;M(um+1), uaupw
hv(tm+1);um+1, u. (29)
Now, since um+1
Lum+1
Kfor all LTh, we can assure that
M(um+1
L)M(um+1
K)and M(um+1
L)M(um+1
K).
1605Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Then, we can bound as follows:
aupw
h(−∇μm+1;M(um+1),u) =
=
eEi
h,e=KLe(−∇μm+1·ne)(M (um+1
K)+M(um+1
L))
(−∇μm+1·ne)(M(um+1
L)+M(um+1
K))(um+1
K)
eEi
h,e=KLe(−∇μm+1·ne)(M (um+1
K)+M(um+1
K))
(−∇μm+1·ne)(M(um+1
K)+M(um+1
K))(um+1
K)
=
eEi
h,e=KLe
(−∇μm+1·ne)M(um+1
K)(um+1
K)=0.
On the other hand, applying the incompressibility of vand proceeding as in
Section 2.4, one has that
aupw
h(v(tm+1);um+1,u) 0.
Therefore, from (29)
|K|δtum+1
K(um+1
K)0.
Consequently, it is satisfied that
0≤|K|tum+1
K)(um+1
K)=−|K|
t (um+1
K)2
+um
K(um+1
K)0;
hence, since um
K0, we prove that (um+1
K)=0. Hence, um+10.
Secondly, we prove that um+11. Taking the following test function
u=(um+1
K1)in K
0 out of K,
where Kis an element of Thsuch that um+1
K=maxKThum+1
Kand using similar
arguments than above, we arrive at
|K|δtum+1
K(um+1
K1)0.
Besides, it is satisfied that
0≥|K|δtum+1
K(um+1
K1)=|K|
t (um+1
K1)+(1um
K)(um+1
K1)
=|K|
t (um+1
K1)2
+(1um
K)(um+1
K1)0;
hence, we deduce that (um+1
K1)=0 and, therefore, um+11.
The following result is a direct consequence of Theorem 3.11.
Corollary 3.12 If we use mass-lumping to compute wm+1in (28c),then0
wm+11in for m0.
1606 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Theorem 3.13 There is at least one solution of the scheme (28).
Proof Given a function zPdisc
0(Th)with 0 z1, we define the map
T:Pdisc
0(Th)×Pcont
1(Th)×Pcont
1(Th)−→ Pdisc
0(Th)×Pcont
1(Th)×Pcont
1(Th)
such that T(
u, μ, w) =(u, μ, w) Pdisc
0(Th)×Pcont
1(Th)×Pcont
1(Th)is the unique
solution, for every uPdisc
0(Th),μ, wPcont
1(Th), of the linear (and decoupled)
scheme:
1
t (uz, u)L2() +aupw
h(v;u, u) =−aupw
h(−∇μ;M(
u),u), (30a)
(μ, μ)L2() =ε2(w, μ)L2() +(f (u, z), μ)L2(),(30b)
(w, w)L2() =(u, w)L2(),(30c)
To check that Tis well defined, one may use the following steps. First, it is easy
to prove that there is a unique solution uof (30a) which implies that there is a unique
solution wof (30c) using, for instance, the Lax-Milgram theorem. Then, it is straight-
forward to see that the solution μof (30b) is unique, which implies its existence as
Pcont
1(Th)is a finite-dimensional space.
It can be proved, using the notion of convergence elementwise, as it was done in
Theorem 2.5 and taking into consideration that μPdisc
0(Th)d, that the operator
Tis continuous, and, therefore, it is compact since Pdisc
0(Th)and Pcont
1(Th)have finite
dimension.
Finally, let us prove that the set
B={(u, μ, w) Pdisc
0(Th)×Pcont
1(Th)×Pcont
1(Th):
(u, μ, w) =αT (u, μ, w) for some 0 α1}
is bounded (independent of α). The case α=0 is trivial so we will assume that
α(0,1].
If (u, μ, w) B,thenuPdisc
0(Th)is the solution, for every uPdisc
0(Th),of
1
t (u αz, u)L2() +aupw
h(v;u, u) =−αa
upw
h(−∇μ;M (u),u). (31)
Now, testing (31)byu=1, we get that
u=α
z,
and, since 0 z1, and since it can be proved that 0 u1usingthesame
arguments than in Theorem 3.11, we get that
uL1() zL1() .
Moreover, wPcont
1(Th)is the solution of the equation
(w, w)L2() =(u, w)L2(),wPcont
1(Th). (32)
1607Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Testing with w=wand using that u1 and that the norms are equivalent in
Pcont
1(Th), we obtain
w2
L2() =(u, w)L2() wL1() ≤||1/2wL2() ;
hence, wL2() ≤||1/2holds.
Finally, we will check that μis bounded. Regarding that, μPcont
1(Th)is the
solution of
(μ, μ)L2() =ε2(w, μ)L2() +(f (u, z), μ)L2(),μPcont
1(Th), (33)
by testing by μ=μwe get that
μ2
L2() ε2w(L2()) μ(L2()) +f (u, z)L2() μL2()
ε2wH1() μH1() +f (u, z)L2() μL2() .
The norms are equivalent in the finite-dimensional space Pcont
1(Th); therefore, there
is Ccont 0 such that
μL2() ε2Ccont wL2() +f (u, z)L2() .
Hence, as 0 u, z 1, we know that f (u, z)L2() is bounded, and therefore
μL2() is bounded.
Since Pdisc
0(Th)and Pcont
1(Th)are finite-dimensional spaces where all the norms
are equivalent, we have proved that Bis bounded.
Thus, using the Leray-Schauder fixed point Theorem 2.6, there is a solution
(u, μ, w) of the Scheme (28).
Corollary 3.14 There is at least one solution of the following (non-truncated)
scheme:
Find um+1Pdisc
0(Th)with 0um+11and μm+1,w
m+1Pcont
1(Th)with
0wm+11, solving
tum+1,u)L2()
+aupw
h(−∇μm+1;M(um+1), u) +aupw
h(v(tm+1);um+1,u) =0,(34a)
μm+1, μL2() =ε2wm+1,μL2() +f(u
m+1,u
m), μL2() ,(34b)
wm+1, wL2() =um+1, wL2() . (34c)
for all uPdisc
0(Th)and μ, wPcont
1(Th). Here, we have considered M(um+1)
instead of M(um+1).
Proof By Theorems 3.11 and 3.13 we know that there is a solution of the scheme
(28) such that 0 um1infor every m0. Hence, M(um)=M(um)for
every m0, and therefore the solution of (28) is also a solution of (34), which
moreover satisfies the discrete maximum principle.
1608 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 2 Left: initial condition in the case v=0. Right: initial condition in the case v=100(y, x)
4 Numerical experiments
We now present several numerical tests in which we explore the behaviour of the
new upwind DG scheme presented in this work (DG-UPW)(28) and compare it
with two other space semidiscretizations found in the literature: firstly a classical
FEM discretization (FEM) and secondly the DG scheme proposed in [18], based on
an SIP + (sigmoid upwind) technique, that we call (DG-SIP).WeuseP1piecewise
polynomials for both schemes unless otherwise specified.
For the DG-UPW scheme, we use mass lumping to compute wm+1in (28c)so
that wm+1∈[0,1]by the Corollary 3.12. This wm+1, which is a regularization of
the primal variable um+1, is considered as the main phase-field variable, which is
used when showing the results of the numerical experiments. Moreover, we consider
Pe =1 unless another value is specified.
4.1 Qualitative tests and comparisons
Our first numerical tests are devoted to qualitative experiments about our DG-UPW
scheme in rectangular and circular domains with different kinds of velocity fields.
We also inspect the discrete energy and the maximum principle property, confirming
Fig. 3 Aggregation of circular regions at T=0.001 without convection, 3D view (height represents the
value of the phase variable on each point of the squared domain)
1609Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 4 Aggregation of circular phases. On the left, maximum (top) and minimum (bottom) of the phase
field variable over time without convection (v=0). On the right, energy over time
that the latter one holds for our scheme but not for the two aforementioned ones,
FEM and DG-SIP.
4.1.1 Agreggation of circular regions without convection
First, we consider the Cahn–Hilliard equation without convection (v=0) in the
squared domain =(0,1)2and the following initial condition (two small circles of
radius 0.2, see Fig. 2,left):
u0=1
2
tanh 0.2 (x x1)2+(y y1)2
2ε+1
+1
2
tanh 0.2 (x x2)2+(y y2)2
2ε+1
,(35)
with centers (x1,y
1)=(0.3,0.5)and (x2,y
2)=(0.7,0.5). We take a structured
mesh with h2.8284 ·102and run time iterations with t =106. Each itera-
tion consists of solving a non-linear system for computing (un+1
n+1,w
n+1),for
which we use Newton’s method iterations, programmed on the FEniCS finite ele-
ment library [1,29]. For linear systems, we used a MPI parallel solver (GMRES) in
the computing cluster of the Universidad de C´
adiz.
In Fig. 3, we show a 3D view of the phase field function at the time step T=
0.001, when the aggregation process has started. It is interesting to notice that for our
upwind DG scheme (28) there are no spurious oscillations meanwhile for the FEM
and DG-SIP schemes we obtain several numerical issues (vertical fluctuations in the
3D graphics).
Moreover, in Fig. 4, we can clearly observe how the maximum principle is pre-
served by DG-UPW scheme (and not by the two other ones). Regarding the energy,
we obtain a non-increasing behaviour as expected from the continuous model. An
analytical proof of this property in the discrete case is left as future work.
1610 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 5 Aggregation of circular regions over time with a strong convection (v=100(y, x))
4.1.2 Agreggation of circular regions with convection
Second, we define as the unit ball in R2and, again, the initial condition (35)(two
small circles of radius 0.2), with centers (x1,y
1)=(0.2,0)and (x2,y
2)=(0.2,0)
(see Fig. 2(right)). Moreover, for testing the effect of convection in our scheme, we
take ε=0.001 and v=100(y, x),sothatv·n=0on∂. We take an unstructured
mesh with h4·102and run time iterations with t =103. Figure 5shows
the values of the phase field function at different time steps. We can observe that,
despite of our election of a highly significant convection term, the results of scheme
DG-UPW are qualitatively correct. Nevertheless, we can observe how the spurious
oscillations become more important and the solution begins to have an unexpected
behaviour when using both the FEM and DG-SIP schemes.
1611Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 6 Aggregation of circular phases with strong convection (v=100(y, x)). On the left, maximum
and minimum of the phase field variable over time. On the right, we plot
um+1um
L()
umL() to observe the
dynamics of the approximations
Concerning the maximum principle, in Fig. 6, we can see how this property is
preserved by the DG-UPW scheme (28), while the phase field variable reaches non-
physical values, very far from [0,1], when using the other aforementioned schemes.
Moreover, it is interesting to observe the approximation of the steady state of the
schemes, represented by the quantity
um+1um
L()
umL() , which tends to 0 in the case
DG-UPW. This fact indicates that the solution converges to a stationary state, while
for the other schemes it remains in an oscillatory state.
The computational time spent to obtain the results (computed sequentially) with
each of these schemes is 2:32min using DG-UPW, 1:10min using FE and 3:24min
using DG-SIP.
For the fairness of comparisons, we redo the tests using both FE and DG-SIP
reducing the step size of the mesh, on the one hand, and using higher order polyno-
mials, on the other hand. First, if we reduce the mesh size to h/22·102and we
Fig. 7 Aggregation of circular phases with strong convection (v=100(y, x)) using the DG-SIP scheme.
On the left, the result obtained with h/22·102and Pdisc
1(Th). On the right, the result obtained with
h4·102and Pdisc
2(Th)
1612 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 8 Random initial
perturbation for the spinoidal
decomposition test
use P1polynomials, the FEM scheme does not converge (the linear solver, GMRES,
fails to converge) while the DG-SIP scheme gives us the results shown in Fig. 7
(left), requiring 35:20min to complete the computations sequentially. Second, if we
keep the mesh size h4·102and we use P2polynomials, the FEM scheme does
not converge either (Newton’s method does not converge) while the DG-SIP scheme
gives us the results shown in Fig. 7(right), requiring 11min to complete the compu-
tations sequentially. Therefore, the FEM scheme does not even converge if we try to
improve the results above and, while the DG-SIP scheme does converge, the results
still show spurious oscillations and require a much longer computational time to be
completed than those shown in Fig. 5(right) using our DG-UPW scheme.
4.1.3 Spinoidal decomposition driven by Stokes cavity flow
We show the results from a spinoidal decomposition test, in which the initial con-
dition is a small uniformly distributed random perturbation around 0.5, u0(x)
[0.49,0.51]for xasshowninFig.8.
As convection vector v, we take the flow resulting from solving a cavity test for
the Stokes equations in the domain =[0,2]×[0,1]with Dirichlet boundary
conditions given by a parabolic profile
v(x, y ) =(x (2x),0), (x , y) top ={(x, 1)R2:x∈[0,2]}.
We used this parabolic profile in order to avoid the discontinuities of the Stokes
velocity for a standard boundary condition v=1ontop, which produces a non-
vanishing divergence in the corners where the discontinuities arise. In fact, we have
checked that, in this particular case where v=1ontop, the scheme DG-UPW does
not preserve the upper bound um1, although it does preserve the lower bound
0um(recall that, for compressible velocity, positivity is the only property of the
solution, see Remark 2.2).
We set ε=0.005, t =0.001, h0.07 and, in this case, we take Pe =10 in
order to emphasize the convection effect. We can observe in Figs. 9and 10 how the
1613Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 9 Spinoidal decomposition over time with convection vector obtained from a cavity test
maximum principle is preserved by our DG-UPW scheme (28) while for the other
schemes the solution takes values out of the interval [0,1](to notice that, we must
take into account the scale of the values shown on the right-hand side of each picture).
Furthermore, in Fig. 10, we can also notice that the approximation obtained using
the DG-UPW scheme converges to a stationary state, while it remains in an non-
physical oscillatory state when we use the other schemes.
1614 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 1 Errors and convergence orders at T=0.001 without convection (v=0)
Scheme Norm h2.8284 ·102h/21.4142 ·102h/39.428 ·103h/47.071 ·103
Error Error Order Error Order Error Order
DG-UPW ·L28.5268 ·1033.0933 ·1031.46 1.7645 ·1031.38 1.2134 ·1031.30
·H18.0000 ·1014.0199 ·1010.99 2.6081 ·1011.07 1.8849 ·1011.13
FEM ·L25.3224 ·1031.5679 ·1031.76 6.9944 ·1041.99 4.0191 ·1041.93
·H18.9963 ·1014.1080 ·1011.13 2.5252 ·1011.2 1.7799 ·1011.22
DG-SIP ·L24.6466 ·1031.3023 ·1031.84 5.8945 ·1041.96 3.2710 ·1042.05
·H11.1784 5.8331 ·1011.01 3.6254 ·1011.17 2.6024 ·1011.15
1615Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 2 Errors and convergence orders at T=0.001 with convection (v=(y, x))
Scheme Norm h4·102h/22·102h/31.3333 ·102h/41·102
Error Error Order Error Order Error Order
DG-UPW ·L21.7288 ·1026.9446 ·1031.32 3.3102 ·1031.83 2.0578 ·1031.65
·H11.4549 6.0305 ·1011.27 3.0204 ·1011.71 2.0315 ·1011.38
FEM ·L26.8347 ·1032.1213 ·1031.69 9.7749 ·1041.91 5.3883 ·1042.07
·H18.3104 ·1013.8060 ·1011.13 2.1887 ·1011.36 1.4991 ·1011.32
DG-SIP ·L26.5242 ·1031.9557 ·1031.74 8.9471 ·1041.93 5.0257 ·1042.00
·H11.1980 6.1624 ·1010.96 3.8451 ·1011.16 2.7439 ·1011.17
1616 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 10 Spinoidal decomposition with convection vector obtained from a cavity test. On the left, maximum
and minimum of the phase field variable over time. On the right, we plot
um+1um
L()
umL() to observe the
dynamics of the approximations
4.2 Error order test
We now introduce the results of a numerical test in which we study the convergence
order of our numerical scheme DG-UPW, verifying experimentally that the expected
order is obtained. For the sake of completeness, we also compare the convergence
orders of the two others aforementioned space semidiscretizations: FEM, DG-SIP.
We consider again the same initial conditions than in Section 4.1.1 with ε=0.01
(see Fig. 2).
First, in the non-convective case, errors and convergence order are compared with
respect to an approximate solution which is computed using the FEM scheme in a
very fine mesh of size h=1.414 ·103and a time step t =106(which is taken
as the “exact solution”). In this case, we have used conforming structured meshes for
the space discretization. The results for the DG-UPW scheme, which are shown in
the first row of Table 1, confirm order 1 (in fact, slightly over 1) in norm ·L2.These
results match our expectations for the P0approximation of um+1, with the upwind
discretization of the non-linear second-order term. It is interesting to emphasize that,
unexpectedly, the scheme produces kind results in ·
H1, reaching order 1. On the
other hand, the FEM and DG-SIP schemes reach order 2 in L2and order 1 in H1
norms, as expected (see also Table 1).
Next, we focus on the case with convection, where we take v=(y, x) in the unit
ball. The resulting errors and convergence orders computed using the three different
schemes over a conforming unstructured mesh are shown in Table 2. In this case, the
errors are computed with respect to the solution obtained for the FEM scheme in a
mesh of size h=4·103.
In this case, it is interesting to notice that the error order in ·L2of the DG-UPW
scheme is improved and it approaches the order 2 of the other schemes, while order
in ·
H1slightly beats the other schemes.
1617Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
As a technical comment, notice that, in order to compute the errors, we projected
on a Pcont
1space both the exact and the DG solution obtained when using the DG-SIP.
In the case of the DG-UPW scheme, we have taken was the continuous solution.
Funding Open Access funding provided thanks to the CRUE-CSIC agreement with Springer Nature.
The first author has been supported by UCA FPU contract UCA/REC14VPCT/2020 funded by Univer-
sidad de adiz andbyaGraduate Scholarship funded by the University of Tennessee at Chattanooga.
The second and third authors have been supported by Grant PGC2018-098308-B-I00 by MCI N/AEI/
10.13039/501100011033 and by ERDF a way of making Europe.
Data availability Data sharing not applicable to this article as no datasets were generated or analyzed
during the current study.
Declarations
Conflictofinterest The authors declare no competing interests.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License,
which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as
you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
licence, and indicate if changes were made. The images or other third party material in this article are
included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
1. Alnæs, M., Blechta, J., Hake, J., et al.: The FEniCS project version 1.5. Arch. Numer. Softw 3(100),
9–23 (2015)
2. Aristotelous, AC., Karakashian, OA., Wise, SM.: Adaptive, second-order in time, primitive-variable
discontinuous Galerkin schemes for a Cahn–Hilliard equation with a mass source. IMA J. Numer.
Anal 35(3), 1167–1198 (2015)
3. Badalassi, V., Ceniceros, H., Banerjee, S.: Computation of multiphase systems with phase field
models. J. Comput. Phys 190(2), 371–397 (2003)
4. Bailo, R., Carrillo, JA., Kalliadasis, S., et al.: Unconditional bound-preserving and energy-dissipating
finite-volume schemes for the Cahn-Hilliard equation. arXiv:210505351 (2021)
5. Barrett, JW., Blowey, JF., Garcke, H.: Finite element approximation of the Cahn–Hilliard equation
with degenerate mobility. SIAM J. Numer. Anal 37(1), 286–318 (1999)
6. Bertozzi, AL., Esedoglu, S., Gillette, A.: Inpainting of binary images using the Cahn–Hilliard
equation. IEEE Trans. Image Process 16(1), 285–291 (2007)
7. Boyer, F., Nabet, F.: A DDFV method for a Cahn-Hilliard/Stokes phase field model with dynamic
boundary conditions. ESAIM Math. Model. Numer. Anal 51(5), 1691–1731 (2017)
8. Cahn, JW.: On spinodal decomposition. Acta Metall 9(9), 795–801 (1961)
9. Cahn, JW., Hilliard, JE.: Free energy of a nonuniform system. I. Interfacial free energy. J. Chem. Phys
28(2), 258–267 (1958)
10. Chen, W., Wang, C., Wang, X., et al.: Positivity-preserving, energy stable numerical schemes for the
Cahn-Hilliard equation with logarithmic potential. J. Comput. Phys. X 3, 100,031 (2019)
11. Chen, Y., Shen, J.: Efficient, adaptive energy stable schemes for the incompressible Cahn–Hilliard
Navier–Stokes phase-field models. J. Comput. Phys 308, 40–56 (2016)
12. Cheng, K., Feng, W., Wang, C., et al.: An energy stable fourth order finite difference scheme for the
Cahn–Hilliard equation. J. Comput. Appl. Math 362, 574–595 (2019)
1618 Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
13. Cueto-Felgueroso, L., Peraire, J.: A time-adaptive finite volume method for the Cahn–Hilliard and
Kuramoto–Sivashinsky equations. J. Comput. Phys 227(24), 9985–10,017 (2008)
14. Di Pietro, DA., Ern, A.: Mathematical aspects of discontinuous Galerkin methods. Springer, New
York (2012)
15. Elliott, CM., French, DA.: A nonconforming finite-element method for the two-dimensional Cahn–
Hilliard equation. SIAM J. Numer. Anal 26(4), 884–903 (1989)
16. Elliott, CM., Garcke, H.: On the Cahn–Hilliard equation with degenerate mobility. SIAM J. Math.
Anal 27(2), 404–423 (1996)
17. Eyre, DJ.: An unconditionally stable one-step scheme for gradient systems. Unpublished article (1998)
18. Frank, F., Liu, C., Alpak, FO., et al.: A finite volume / discontinuous Galerkin method for the advec-
tive Cahn–Hilliard equation with degenerate mobility on porous domains stemming from micro-CT
imaging. Comput Geosci 22(2), 543–563 (2018)
19. Frank, F., Rupp, A., Kuzmin, D..: Bound-preserving flux limiting schemes for DG discretizations of
conservation laws with applications to the Cahn–Hilliard equation. Comput. Methods Appl. Mech.
Eng 359, 112–665 (2020)
20. Furihata, D.: A stable and conservative finite difference scheme for the Cahn–Hilliard equation.
Numer. Math 87(4), 675–699 (2001)
21. Guill´
en-Gonz´
alez, F., Tierra, G.: On linear schemes for a Cahn–Hilliard diffuse interface model.
JComput. Phys 234, 140–171 (2013)
22. Guo, Z., Lin, P., Lowengrub, J., et al.: Mass conservative and energy stable finite difference methods
for the quasi-incompressible Navier–Stokes–Cahn–Hilliard system: primitive variable and projection-
type schemes. ComputMethods Appl. Mech. Eng 326, 144–174 (2017)
23. Ibrahim, M., Saad, M.: On the efficacy of a control volume finite element method for the capture of
patterns for a volume-filling chemotaxis model. Comput. Math. Appl 68(9), 1032–1051 (2014)
24. Kay, D., Styles, V., uli, E.: Discontinuous Galerkin finite element approximation of the Cahn–
Hilliard equation with convection. SIAM J. Numer. Anal 47(4), 2660–2685 (2009)
25. Kim, J.: A numerical method for the Cahn–Hilliard equation with a variable mobility. Commun.
Nonlinear Sci. Numer. Simul 12(8), 1560–1571 (2007)
26. Kim, J.: Phase-field models for multi-component fluid flows. Commun. Comput. Phys 12(03), 613–
661 (2012)
27. Li, R., Gao, Y., Chen, J., et al.: Discontinuous finite volume element method for a coupled Navier-
Stokes-Cahn-Hilliard, phase field model. Adv. Comput. Math 46(2), 1–35 (2020)
28. Liu, H., Yin, P.: Unconditionally energy stable discontinuous Galerkin schemes for the Cahn–Hilliard
equation. JComput. Appl. Math 390, 113,375 (2021)
29. Logg, A., Mardal, KA., Wells, GN., et al.: Automated solution of differential equations by the finite
element method. Springer Berlin, Heidelberg (2012)
30. Miranville, A.: The Cahn–Hilliard equation and some of its variants. AIMS Math 2(3), 479–544
(2017)
31. Pironneau, O.: Finite element methods for fluids. Wiley (1989)
32. Wang, C., Wise, SM.: An energy stable and convergent finite-difference scheme for the modified
phase field crystal equation. SIAM J. Numer. Anal 49(3), 945–969 (2011)
33. Wells, GN., Kuhl, E., Garikipati, K.: A discontinuous Galerkin method for the Cahn–Hilliard equation.
J. Comput. Phys 218(2), 860–877 (2006)
34. Wise, S., Lowengrub, J., Frieboes, H., et al.: Three-dimensional multispecies nonlinear tumor
growth—I. J. Theor. Biol 253(3), 524–543 (2008)
35. Wu, X., Zwieten, GJ., Zee, KG.: Stabilized second-order convex splitting schemes for Cahn–Hilliard
models with application to diffuse-interface tumor-growth models. Int. J. Numer. Methods Biomed
Eng 30(2), 180–203 (2014)
36. Xia, Y., Xu, Y., Shu, CW.: Local discontinuous Galerkin methods for the Cahn–Hilliard type
equations. J. Comput Phys 227(1), 472–491 (2007)
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published
maps and institutional affiliations.
1619Numerical Algorithms (2023) 92:1589–1619
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center
GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers
and authorised users (“Users”), for small-scale personal, non-commercial use provided that all
copyright, trade and service marks and other proprietary notices are maintained. By accessing,
sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of
use (“Terms”). For these purposes, Springer Nature considers academic use (by researchers and
students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and
conditions, a relevant site licence or a personal subscription. These Terms will prevail over any
conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription (to
the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of
the Creative Commons license used will apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may
also use these personal data internally within ResearchGate and Springer Nature and as agreed share
it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not otherwise
disclose your personal data outside the ResearchGate or the Springer Nature group of companies
unless we have your permission as detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial
use, it is important to note that Users may not:
use such content for the purpose of providing other users with access on a regular or large scale
basis or as a means to circumvent access control;
use such content where to do so would be considered a criminal or statutory offence in any
jurisdiction, or gives rise to civil liability, or is otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association
unless explicitly agreed to by Springer Nature in writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a
systematic database of Springer Nature journal content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a
product or service that creates revenue, royalties, rent or income from our content or its inclusion as
part of a paid for service or for other commercial gain. Springer Nature journal content cannot be
used for inter-library loans and librarians may not upload Springer Nature journal content on a large
scale into their, or any other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not
obligated to publish any information or content on this website and may remove it or features or
functionality at our sole discretion, at any time with or without notice. Springer Nature may revoke
this licence to you at any time and remove access to any copies of the Springer Nature journal content
which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or
guarantees to Users, either express or implied with respect to the Springer nature journal content and
all parties disclaim and waive any implied warranties or warranties imposed by law, including
merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published
by Springer Nature that may be licensed from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a
regular basis or in any other manner not expressly permitted by these Terms, please contact Springer
Nature at
onlineservice@springernature.com
... In the last decades, a lot of papers have been published dealing with these kinds of theoretical issues both for the classical model (1) and for other models based on some extensions or generalizations. In general, they start from the Keller-Segel classical equations and modify them with the purpose of avoiding the non-physical blow up of solutions, producing solutions which are closer to the "real chemotaxis" phenomena observed in biology. ...
... Moreover, if u m ≥ 0 and v m ≥ 0 (when τ > 0) it follows from the Eq. (13b) that u m+1 ≥ 0 using the same arguments that are shown to prove the positivity result in [1] since the proof given in [1] is independent of the flux ∇μ m+1 . ...
... Moreover, if u m ≥ 0 and v m ≥ 0 (when τ > 0) it follows from the Eq. (13b) that u m+1 ≥ 0 using the same arguments that are shown to prove the positivity result in [1] since the proof given in [1] is independent of the flux ∇μ m+1 . ...
Article
Full-text available
The well-suited discretization of the Keller–Segel equations for chemotaxis has become a very challenging problem due to the convective nature inherent to them. This paper aims to introduce a new upwind, mass-conservative, positive and energy-dissipative discontinuous Galerkin scheme for the Keller–Segel model. This approach is based on the gradient-flow structure of the equations. In addition, we show some numerical experiments in accordance with the aforementioned properties of the discretization. The numerical results obtained emphasize the really good behaviour of the approximation in the case of chemotactic collapse, where very steep gradients appear.
... However, while this model is mass-conservative and energy-dissipative, it does not implicitly impose the necessary pointwise bounds on the volume fraction variables. Therefore, we propose a modification of the aforementioned tumor model, see (2) below, in accordance with its physical interpretation. As a result of this modification, we obtain pointwise bounds on the tumor and the nutrient volume fractions (0 ≤ u, n ≤ 1) which are consistent with the physical meaning of the variables. ...
... Among the currently available structure-preserving schemes we can find some schemes based on finite volumes, [6,27], and on finite elements, [25]. Moreover, the authors have published a recent work, [2], where the pointwise bounds of the CH model in the case with convection are preserved using an upwind discontinuous Galerkin (DG) approximation. To the best knowledge of the authors, no previous work has been published defining a fully discrete DG scheme preserving the mass conservation, pointwise bounds and energy stability of the CH model with degenerate mobility. ...
... In this sense, we introduce a well-suited convex-splitting DG scheme of the proposed model (2), based on the previous works [1,2,28]. This approximation preserves the physical properties of the phase-field tumor model (mass conservation, pointwise bounds and energy stability) and prevent numerical spurious oscillations. ...
Preprint
Full-text available
In this work, we present a modification of the phase-field tumor growth model given in [26] that leads to bounded, more physically meaningful, volume fraction variables. In addition, we develop an upwind discontinuous Galerkin (DG) scheme preserving the mass conservation, pointwise bounds and energy stability of the continuous model. Finally, some computational tests in accordance with the theoretical results are introduced. In the first test, we compare our DG scheme with the finite element (FE) scheme related to the same time approximation. The DG scheme shows a well-behavior even for strong cross-diffusion effects in contrast with FE where numerical spurious oscillations appear. Moreover, the second test exhibits the behavior of the tumor-growth model under different choices of parameters and also of mobility and proliferation functions.
... It is well known that computing of numerical solutions to this kind of hyperbolic PDE equations is not an easy task [EG10,QSS10]. Fortunately we can successfully apply the upwind Discontinuous Galerkin (DG) techniques published in [ASGGRG22] to obtain quite appropriate results. The PDE is supplemented with some extra reaction terms described below, modelling the born of neuroblastas and their disapearence due to evolution in mature neurons. ...
... [EG10,QSS10]) that FEM methods are not well suited for approximations to hyperbolic equations like (1), leading to the loss of its maximum principle property, spurious oscillations and wrong results. Here we will use an upwind discontinuous Galerkin method introduced in [ASGGRG22]. Before getting into the evolution problem, it will be essential to compute appropriate estimates to the olfactory bulb attraction function O and to the initial condition u 0 . ...
... where O ∈ P cont 1 (T h ) is the solution to (7) for the parameter σ and a upw h (· ; ·, ·) is the upwind bilinear form discussed in [ASGGRG22] and defined as ...
Preprint
Full-text available
This article is devoted to the mathematical modeling of migration of neuroblasts, precursor cells of neurons, along the pathway they usually follow before maturing. This pathway is determined mainly by chemotaxis and the heterogeneous mobility of neuroblasts in different regions of the brain. In numerical simulations, the application of novel discontinuous Galerkin methods allows to maintain the properties of the continuous model such as the maximum principle. We present some successful computer tests including parameter adjust to fit real data from rodent brains.
... The main advantage is that it can simultaneously ensure energy stability, mass conservation and boundedness, but it is usually useless for degenerate mobility and potential where convex-concave decomposition is inaccessible. -Upwind approach [1,7] applies upwind scheme to deal with the flux term, leading to a scheme suitable for degenerate PDEs and can guarantee mass conservation, but usually has first-order accuracy in time. Recently, this method was further combined with the convex splitting strategy to achieve the unconditional energy stability [1,4]. ...
... -Upwind approach [1,7] applies upwind scheme to deal with the flux term, leading to a scheme suitable for degenerate PDEs and can guarantee mass conservation, but usually has first-order accuracy in time. Recently, this method was further combined with the convex splitting strategy to achieve the unconditional energy stability [1,4]. -Lagrange multiplier approach [17,18] introduces a space-time Lagrange multiplier and uses the Karush-Kuhn-Tucker conditions as a constraint to achieve bound/positivity preserving, which, however, does not guarantee the energy stability, although the mass conservation can be ensured. ...
Article
Full-text available
This paper establishes a structure-preserving numerical scheme for the Cahn–Hilliard equation with degenerate mobility. First, by applying a finite volume method with upwind numerical fluxes to the degenerate Cahn–Hilliard equation rewritten by the scalar auxiliary variable (SAV) approach, we creatively obtain an unconditionally bound-preserving, energy-stable and fully-discrete scheme, which, for the first time, addresses the boundedness of the classical SAV approach under H-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$H^{-1}$$\end{document}-gradient flow. Then, a dimensional-splitting technique is introduced in high-dimensional cases, which greatly reduces the computational complexity while preserves original structural properties. Numerical experiments are presented to verify the bound-preserving and energy-stable properties of the proposed scheme. Finally, by applying the proposed structure-preserving scheme, we numerically demonstrate that surface diffusion is approximated by the Cahn–Hilliard equation with degenerate mobility and Flory–Huggins potential, when the absolute temperature is sufficiently low, which agrees well with the theoretical result by using formal asymptotic analysis.
... The case of logarithmic potentials together with degenerate mobilities were studied using mixed finite element approximations in [5,9], while in [7] the authors focus on the same problem but using finite differences. Discontinuous Galerkin methods have also become very popular to approximate phase field systems with degenerate mobilities [2,3,24,25,28]. ...
Preprint
In this work we present two new numerical schemes to approximate the Navier-Stokes-Cahn-Hilliard system with degenerate mobility using finite differences in time and finite elements in space. The proposed schemes are conservative, energy-stable and preserve the maximum principle approximately (the amount of the phase variable being outside of the interval [0,1] goes to zero in terms of a truncation parameter). Additionally, we present several numerical results to illustrate the accuracy and the well behavior of the proposed schemes, as well as a comparison with the behavior of the Navier-Stokes-Cahn-Hilliard model with constant mobility.
... Applying the maximum principle, they investigated the generalized time-fractional variable-order Caputo diffusion equations and fractional variable-order Riesz-Caputo diffusion equations. For other new developments of the maximum principle, the reader can refer to [8][9][10][11][12][13][14][15][16] and the references therein. ...
Article
Full-text available
In this paper, we investigate properties of solutions to a space-time fractional variable-order conformable nonlinear differential equation with a generalized tempered fractional Laplace operatorby using the maximum principle. We first establish some new important fractional various-order conformable inequalities. With these inequalities, we prove a new maximum principle with space-time fractional variable-order conformable derivatives and a generalized tempered fractional Laplace operator. Moreover, we discuss some results about comparison principles and properties of solutions for a family of space-time fractional variable-order conformable nonlinear differential equations with a generalized tempered fractional Laplace operator by maximum principle.
... Mixed finite element approximations using logarithmic potentials and degenerate mobilities were studied in [3,7]. More recently, several discontinuous Galerkin methods have been considered for the problem with and without convection [1,16,17,25]. ...
Preprint
Two new numerical schemes to approximate the Cahn-Hilliard equation with degenerate mobility (between stable values 0 and 1) are presented, by using two different non-centered approximation of the mobility. We prove that both schemes are energy stable and preserve the maximum principle approximately, i.e. the amount of the solution being outside of the interval [0,1] goes to zero in terms of a truncation parameter. Additionally, we present several numerical results in order to show the accuracy and the well behavior of the proposed schemes, comparing both schemes and the corresponding centered scheme.
Article
Full-text available
The stabilization approach has been known to permit large time-step sizes while maintaining stability. However, it may “slow down the convergence rate” or cause “delayed convergence” if the time-step rescaling is not well resolved. By considering a fourth-order-in-space viscous Cahn–Hilliard (VCH) equation, we propose a class of up to the fourth-order single-step methods that are able to capture the correct physical behaviors with high-order accuracy and without time delay. By reformulating the VCH as a system consisting of a second-order diffusion term and a nonlinear term involving the operator \(({I} - \nu \Delta )^{-1}\), we first develop a general approach to estimate the maximum bound for the VCH equation equipped with either the Ginzburg–Landau or Flory–Huggins potential. Then, by taking advantage of new recursive approximations and adopting a time-step-dependent stabilization, we propose a class of stabilization Runge–Kutta methods that preserve the maximum principle for any time-step size without harming the convergence. Finally, we transform the stabilization method into a parametric Runge–Kutta formulation, estimate the rescaled time-step, and remove the time delay by means of a relaxation technique. When the stabilization parameter is chosen suitably, the proposed parametric relaxation integrators are rigorously proven to be mass-conserving, maximum-principle-preserving, and the convergence in the \(l^\infty \)-norm is estimated with pth-order accuracy under mild regularity assumption. Numerical experiments on multi-dimensional benchmark problems are carried out to demonstrate the stability, accuracy, and structure-preserving properties of the proposed schemes.
Article
Full-text available
In this paper, we propose a discontinuous finite volume element method to solve a phase field model for two immiscible incompressible fluids. In this finite volume element scheme, discontinuous linear finite element basis functions are used to approximate the velocity, phase function, and chemical potential while piecewise constants are used to approximate the pressure. This numerical method is efficient, optimally convergent, conserving the mass, convenient to implement, flexible for mesh refinement, and easy to handle complex geometries with different types of boundary conditions. We rigorously prove the mass conservation property and the discrete energy dissipation for the proposed fully discrete discontinuous finite volume element scheme. Using numerical tests, we verify the accuracy, confirm the mass conservation and the energy law, test the influence of surface tension and small density variations, and simulate the driven cavity, the Rayleigh-Taylor instability.
Article
Full-text available
In this paper we present and analyze a finite difference numerical scheme for the Allen Cahn/Cahn-Hilliard equation with a logarithmic Flory Huggins energy potential. In more details, we treat the nonlinear logarithmic terms and the surface diffusion term implicitly, and update the linear expansive term and the mobility explicitly. We provide a theoretical justification that, this numerical algorithm has a unique solution such that the positivity is always preserved for the logarithmic arguments, i.e., the phase variable is always between $-1$ and 1, at a point-wise level. In particular, our analysis reveals a subtle fact: the singular nature of the logarithmic term around the values of $-1$ and 1 prevents the numerical solution reaching these singular values, so that the numerical scheme is always well-defined as long as the numerical solution stays similarly bounded at the previous time step. Furthermore, an unconditional energy stability of the numerical scheme is derived, without any restriction for the time step size. As a result of the positivity-preserving property, unique solvability, and energy stability, we are able to derive an optimal rate convergence analysis for the numerical scheme, which gives an $O (\dt + h^2)$ error estimate. The case with a non-constant mobility is analyzed as well. We also describe a practical and efficient multigrid solver for the proposed numerical scheme, and present some numerical results, which demonstrate the robustness of the numerical scheme.
Article
Full-text available
Our aim in this article is to review and discuss the Cahn-Hilliard equation, as well as some of its variants. Such variants have applications in, e.g., biology and image inpainting.
Article
Full-text available
A numerical method is formulated for the solution of the advective Cahn-Hilliard (CH) equation with constant and degenerate mobility in three-dimensional porous media with non-vanishing velocity on the exterior boundary. The CH equation describes phase separation of an immiscible binary mixture at constant temperature in the presence of a mass constraint and dissipation of free energy. Porous media/pore-scale problems specifically entail high-resolution images of rocks in which the solid matrix and pore spaces are fully resolved. The interior penalty discontinuous Galerkin method is used for the spatial discretization of the CH equation in mixed form, while a semi-implicit convex-concave splitting is utilized for temporal discretization. The spatial approximation order is arbitrary, while it reduces to a finite volume scheme for the choice of elementwise constants. The resulting nonlinear systems of equations are reduced using the Schur complement and solved via Newton's method. The numerical scheme is first validated using numerical convergence tests and then applied to a number of fundamental problems for validation and numerical experimentation purposes including the case of degenerate mobility. First-order physical applicability and robustness of the numerical method are shown in a breakthrough scenario on a voxel set obtained from a micro-CT scan of a real sandstone rock sample.
Article
Full-text available
p>The FEniCS Project is a collaborative project for the development of innovative concepts and tools for automated scientific computing, with a particular focus on the solution of differential equations by finite element methods. The FEniCS Projects software consists of a collection of interoperable software components, including DOLFIN, FFC, FIAT, Instant, UFC, UFL, and mshr. This note describes the new features and changes introduced in the release of FEniCS version 1.5.</p
Article
In this paper, we introduce novel discontinuous Galerkin (DG) schemes for the Cahn-Hilliard equation, which arises in many applications. The method is designed by integrating the mixed DG method for the spatial discretization with the Invariant Energy Quadratization (IEQ) approach for the time discretization. Coupled with a spatial projection, the resulting IEQ-DG schemes are shown to be unconditionally energy dissipative, and can be efficiently solved without resorting to any iteration method. Both one and two dimensional numerical examples are provided to verify the theoretical results, and demonstrate the good performance of IEQ-DG in terms of efficiency, accuracy, and preservation of the desired solution properties.
Article
Many mathematical models of computational fluid dynamics involve transport of conserved quantities, which must lie in a certain range to be physically meaningful. The analytical or numerical solution u of a scalar conservation law is said to be bound-preserving if global bounds u∗ and u∗ exist such that u∗≤u≤u∗ holds in the domain of definition. These bounds must be known a priori. To enforce such inequality constraints at least for element averages in the context of discontinuous Galerkin (DG) methods, the numerical fluxes must be defined and constrained in an appropriate manner. In this work, we introduce a general framework for calculating fluxes that produce non-oscillatory DG approximations and preserve relevant global bounds for element averages even if the analytical solution of the PDE violates them due to modeling errors. The proposed methodology is based on a combination of flux and slope limiting. The (optional) slope limiter adjusts the gradients to impose local bounds on pointwise values of the high-order DG solution, which is used to calculate the fluxes. The flux limiter constrains changes of element averages so as to prevent violations of global bounds. Since manipulations of the target flux may introduce a consistency error, it is essential to guarantee that physically admissible fluxes remain unchanged. We propose two kinds of flux limiters, which meet this requirement. The first one is of monolithic type and its time-implicit version requires the iterative solution of a nonlinear problem. Only a fully converged solution is provably bound-preserving. The time-explicit version of this limiter is subject to a time step restriction, which we derive in this article. The second limiter is an iterative version of the multidimensional flux-corrected transport (FCT) algorithm and works as postprocessed correction scheme. This fractional step limiting approach guarantees that each iterate is bound-preserving but avoidable consistency errors may occur if the iterative process is terminated too early. Each iterate depends only on local information of the previous iterate. This concept of limiting the numerical fluxes is also applicable to finite volume methods. Practical applicability of the proposed flux limiters as well as the benefits of using an optional slope limiter are demonstrated by numerical studies for the advection equation (hyperbolic, linear) and the Cahn–Hilliard equation (parabolic, nonlinear) for first-order polynomials. While both flux limiters work for arbitrary order polynomials, we discuss the construction of bound-preserving slope limiters, and show numerical studies only for first-order polynomials.
Article
In this paper we propose and analyze an energy stable numerical scheme for the Cahn-Hilliard equation, with second order accuracy in time and the fourth order finite difference approximation in space. In particular, the truncation error for the long stencil fourth order finite difference approximation, over a uniform numerical grid with a periodic boundary condition, is analyzed, via the help of discrete Fourier analysis instead of the the standard Taylor expansion. This in turn results in a reduced regularity requirement for the test function. In the temporal approximation, we apply a second order BDF stencil, combined with a second order extrapolation formula applied to the concave diffusion term, as well as a second order artificial Douglas-Dupont regularization term, for the sake of energy stability. As a result, the unique solvability, energy stability are established for the proposed numerical scheme, and an optimal rate convergence analysis is derived in the $\ell^\infty (0,T; \ell^2) \cap \ell^2 (0,T; H_h^2)$ norm. A few numerical experiments are presented, which confirm the robustness and accuracy of the proposed scheme.
Article
In this paper we describe two fully mass conservative, energy stable, finite difference methods on a staggered grid for the quasi-incompressible Navier-Stokes-Cahn-Hilliard (q-NSCH) system governing a binary incompressible fluid flow with variable density and viscosity. Both methods, namely the primitive method (finite difference method in the primitive variable formulation) and the projection method (finite difference method in a projection-type formulation), are so designed that the mass of the binary fluid is preserved, and the energy of the system equations is always non-increasing in time at the fully discrete level. We also present an efficient, practical nonlinear multigrid method - comprised of a standard FAS method for the Cahn-Hilliard equation, and a method based on the Vanka-type smoothing strategy for the Navier-Stokes equation - for solving these equations. We test the scheme in the context of Capillary Waves, rising droplets and Rayleigh-Taylor instability. Quantitative comparisons are made with existing analytical solutions or previous numerical results that validate the accuracy of our numerical schemes. Moreover, in all cases, mass of the single component and the binary fluid was conserved up to 10 to -8 and energy decreases in time.
Article
In this paper we propose a "Discrete Duality Finite Volume" method (DDFV for short) for the diffuse interface modelling of incompressible flows. This numerical method is, conservative, robust and is able to handle general geometries and meshes. The model we study couples the Cahn-Hilliard equation and the unsteady Stokes equation and is endowed with particular nonlinear boundary conditions called dynamic boundary conditions. To implement the scheme for this model we have to define new discrete consistent DDFV operators that allows an energy stable coupling between both discrete equations. We are thus able to obtain the existence of a family of solutions satisfying a suitable energy inequality, even in the case where a first order time-splitting method between the two subsystems is used. We illustrate various properties of such a model with some numerical results.