• Tidak ada hasil yang ditemukan

Existence and weak–strong uniqueness for Maxwell–Stefan–Cahn–Hilliard systems

N/A
N/A
Protected

Academic year: 2023

Membagikan "Existence and weak–strong uniqueness for Maxwell–Stefan–Cahn–Hilliard systems"

Copied!
50
0
0

Teks penuh

(1)

Existence and weak–strong uniqueness for Maxwell–Stefan–Cahn–Hilliard systems

Item Type Article

Authors Huo, Xiaokai; Jüngel, Ansgar; Tzavaras, Athanasios

Citation Huo, X., Jüngel, A., & Tzavaras, A. E. (2023). Existence and weak–

strong uniqueness for Maxwell–Stefan–Cahn–Hilliard systems.

Annales de l’Institut Henri Poincaré C, Analyse Non Linéaire.

https://doi.org/10.4171/aihpc/89 Eprint version Post-print

DOI 10.4171/aihpc/89

Publisher European Mathematical Society - EMS - Publishing House GmbH Journal Annales de l'Institut Henri Poincaré C, Analyse non linéaire Rights Archived with thanks to Annales de l'Institut Henri Poincaré C,

Analyse non linéaire under a Creative Commons license, details at: https://creativecommons.org/licenses/by/4.0/

Download date 21/06/2023 07:09:49

Link to Item http://hdl.handle.net/10754/677939

(2)

MAXWELL–STEFAN–CAHN–HILLIARD SYSTEMS

XIAOKAI HUO, ANSGAR J ¨UNGEL, AND ATHANASIOS E. TZAVARAS

Abstract. A Maxwell–Stefan system for fluid mixtures with driving forces depending on Cahn–Hilliard-type chemical potentials is analyzed. The corresponding parabolic cross- diffusion equations contain fourth-order derivatives and are considered in a bounded do- main with no-flux boundary conditions. The nonconvex part of the energy is assumed to have a bounded Hessian. The main difficulty of the analysis is the degeneracy of the diffusion matrix, which is overcome by proving the positive definiteness of the matrix on a subspace and using the Bott–Duffin matrix inverse. The global existence of weak solutions and a weak-strong uniqueness property are shown by a careful combination of (relative) energy and entropy estimates, yieldingH2(Ω) bounds for the densities, which cannot be obtained from the energy or entropy inequalities alone.

1. Introduction

The evolution of fluid mixtures is important in many scientific fields like biology and nanotechnology to understand the diffusion-driven transport of the species. The transport can be modeled by the Maxwell–Stefan equations [33, 35], which consist of the mass bal- ance equations and the relations between the driving forces and the fluxes. The driving forces involve the chemical potentials of the species, which in turn are determined by the (free) energy. When the fluid is immiscible, the energy can be assumed to consist of the thermodynamic entropy and the phase separation energy, given by a density gradient [6].

The gradient energetically penalizes the formation of an interface and restrains the segre- gation. This leads to a system of cross-diffusion equations with fourth-order derivatives.

The aim of this paper is to provide a global existence and weak-strong uniqueness analysis for multicomponent systems of Maxwell–Stefan–Cahn–Hilliard-type.

Date: February 26, 2023.

2000Mathematics Subject Classification. 35A02, 35G20, 35G31, 35K51, 35K55, 35Q35.

Key words and phrases. Cross-diffusion systems, global existence, weak-strong uniqueness, relative en- tropy, relative free energy, parabolic fourth-order equations, Maxwell–Stefan equations, Cahn–Hilliard equations.

XH and AJ acknowledge partial support from the Austrian Science Fund (FWF), grants P33010, W1245, and F65. AET acknowledges support from baseline funds of the King Abdullah University of Science and Technology (KAUST). This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme, ERC Advanced Grant no. 101018153.

1

(3)

1.1. Model equations and state of the art. The equations for the partial densities ci and partial velocitiesui are given by

tci+ div(ciui) = 0, i= 1, . . . , n, (1)

ci∇µi− ci Pn

k=1ck

n

X

j=1

cj∇µj =−

n

X

j=1

Kij(c)cjuj, (2)

n

X

j=1

cjuj = 0, (3)

supplemented by the initial and boundary conditions

(4) c(·,0) = c0 in Ω, ciui·ν =∇ci·ν= 0 on ∂Ω, t >0, i= 1, . . . , n,

where Ω⊂Rd(d= 1,2,3) is a bounded domain,ν is the exterior unit normal vector on the boundary∂Ω,c= (c1, . . . , cn) is the density vector, andKij(c) are the friction coefficients.

The left-hand side of (2) can be interpreted as the driving forces of the thermodynamic system, and the right-hand side is the sum of the friction forces. The chemical potentials

(5) µi = δE

δci = logci−∆ci, i= 1, . . . , n, are the variational derivatives of the (free) energy

(6) E(c) =H(c) + 1 2

n

X

i=1

Z

|∇ci|2dx, H(c) =

n

X

i=1

Z

ci(logci−1) + 1 dx,

and H(c) is the thermodynamic entropy. Note that this energy is convex; we show in Section 5 that our results still hold if the energy contains a nonconvex part with bounded Hessian (like the potential in [13]). We assume that Pn

i=1Kij(c) = 0 for j = 1, . . . , n, meaning that the linear system in∇µj is invertible only on a subspace, and thatPn

i=1c0i = 1 in Ω, which implies that Pn

i=1ci(t) = 1 in Ω for all time t > 0. This means that the mixture is saturated andci can be interpreted as volume fraction. For simplicity, we have normalized all physical constants.

Model (1)–(5) has been derived rigorously in [24] in the high-friction limit from a multi- component Euler–Korteweg system for a general convex energy functional depending onc and∇c. A thermodynamics-based derivation can be found in [34]. When the energy equals E(c) = H(c), the model reduces to the classical Maxwell–Stefan equations, analyzed first in [4, 21, 22] for local-in-time smooth solutions and later in [30] for global-in-time weak solu- tions. In the single-species case, model (1)–(5) becomes a fourth-order Cahn–Hilliard-type equation with convex potentialφ(c) =c(logc−1). Such a model, additionally including a nonconvex potential, was analyzed in, e.g., [13, 27]. Convergence from the Euler-Korteweg in the high-friction limit to the Cahn-Hilliard equation with nonconvex potential is pro- vided in [20]. Only few works are concerned with the multi-species situation, and all of them require additional conditions. The mobility matrix in [5, 32] is assumed to be diago- nal and that one in [31] has constant entries, while the works [11, 14] suppose a particular

(4)

(but nondiagonal) structure of the mobility matrix. We also mention the works [2, 3] on related models with free energies of the typeH.

The proof of the uniqueness of solutions to cross-diffusion or fourth-order systems is quite delicate due to the lack of a maximum principle and regularity of the solutions.

The uniqueness of strong solutions to Maxwell–Stefan systems has been shown in [22, 26], and uniqueness results for weak solutions in a very special case can be found in [8]. A weak-strong uniqueness result was proved for reaction-diffusion systems in [18] and for Maxwell–Stefan systems in [25]. Concerning uniqueness results for fourth-order equations, we refer to [9] for single-species Cahn–Hilliard equations, [28] for single-species thin-film equations, and [17] for the quantum drift-diffusion equations. Up to our knowledge, there are no uniqueness results for multicomponent Cahn–Hilliard-type systems. In this paper, we analyze these equations in a general setting for the first time.

1.2. Key ideas of the analysis. Before stating the main results, we explain the math- ematical ideas needed to analyze model (1)–(5). First, we rewrite (2) by introducing the matrix D(c)∈Rn×n with entries

Dij(c) = 1

√ci

Kij(c)√ cj

in the unknowns (√

c1u1, . . . ,√ cnun):

(7)

√ci∇µi

√ci Pn

k=1ck

n

X

j=1

cj∇µj =−

n

X

j=1

Dij(c)√ cjuj,

n

X

i=1

√ci √ ciui

= 0.

We show in Lemma 3 that this linear system has a unique solution in the space L(c) :=

{z ∈Rn:Pn i=1

√cizi = 0}, and the solution reads as

√ciui =−

n

X

j=1

DijBD(c)√

cj∇µj,

where DBD(c) is the so-called Bott–Duffin matrix inverse; see Lemmas 3 and 4 for the definition and some properties. Then, defining the matrix B(c)∈Rn×n with elements

(8) Bij(c) =√

ciDijBD(c)√

cj, i, j = 1, . . . , n, system (1)–(2) can be formulated as (see Section 2.1 for details)

tci = div

n

X

j=1

Bij(c)∇µj, i= 1, . . . , n.

The matrix B(c) is often called Onsager or mobility matrix in the literature. The major difficulty of the analysis consists in the fact that the matrixB(c) is singular and degenerates

(5)

when ci →0 for some i∈ {1, . . . , n}. Computing formally the energy identity dE

dt(c) +

n

X

i,j=1

Z

Bij(c)∇µi· ∇µjdx = 0,

the degeneracy atci = 0 prevents uniform estimates for∇µi inL2(Ω). In some works, this issue has been compensated. For instance, there exists an entropy equality for the model of [14] yielding an L2(Ω) bound for ∆ci, and the decoupled mobilities in [7, 32] allow for decoupled entropy estimates. In our model, the energy identity does not provide a gradient estimate for the full vector (∇µ1, . . . ,∇µn) but only for a projection:

dE

dt(c) +C1

n

X

i=1

Z

n

X

j=1

ij −√ cicj)√

cj∇µj

2

dx≤0,

where δij is the Kronecker delta; see Lemma 5. (The constant C1 > 0 and all constants that follow do not depend on c.) To address the degeneracy issue, we compute the time derivative of the entropy:

dH dt (c) +

n

X

i,j=1

Z

Bij(c)∇logci· ∇µjdx= 0.

This does not provide a uniform estimate for ∆ci, but we show (see Lemma 5) that dH

dt (c) +C2

n

X

i=1

Z

(∆ci)2dx≤C3

n

X

i=1

Z

n

X

j=1

ij −√ cicj)√

cj∇µj

2

dx.

Combining the energy and entropy inequalities in a suitable way, the last integral cancels:

(9) d

dt

H(c) + C3 C1E(c)

+C2

n

X

i=1

Z

(∆ci)2dx≤0.

This provides the desired H2(Ω) bound for ci. Note that the energy or entropy inequality alone does not give estimates for ci. The combined energy-entropy inequality is the key idea of the paper for both the existence and weak-strong uniqueness analysis. Observe that the termH(c) + (C3/C1)E(c) can also be written as (1 +C3/C1)H(c) +12Pn

i=1

R

|∇ci|2dx.

1.3. Main results. We make the following assumptions:

(A1) Domain: Ω ⊂ Rd with d ≤ 3 is a bounded domain. We set QT = Ω×(0, T) for T >0.

(A2) Initial data: c0i ∈H1(Ω) satisfies c0i ≥0 in Ω,i= 1, . . . , n, and Pn

i=1c0i = 1 in Ω.

The assumptiond≤3 is made for convenience, it can be relaxed for higher space dimen- sion, by choosing another regularization in the existence proof; see (88). The constraint Pn

i=1c0i = 1 expresses the saturation of the mixture and it propagates to the solution. We introduce the matrix Dij(c) = (1/√

ci)Kij(c)√

cj for i, j = 1, . . . , n and set (10) L(c) = {x∈Rn :√

c·x= 0}, L(c) = span{√ c},

(6)

where √

c = (√

c1, . . . ,√

cn). The projections PL(c), PL(c) ∈ Rn×n on L(c), L(c), respectively, are given by

(11) PL(c)ijij −√

cicj, PL(c)ij =√

cicj for i, j = 1, . . . , n.

We impose for any given c∈[0,1]n the following assumptions onD(c) = (Dij(c))∈Rn×n: (B1) D(c) is symmetric and ranD(c) = L(c), ker(D(c)PL(c)) =L(c).

(B2) For all i, j = 1, . . . , n, Dij ∈C1([0,1]n) is bounded.

(B3) The matrix D(c) is positive semidefinite, and there exists ρ > 0 such that all eigenvaluesλ 6= 0 of D(c) satisfy λ≥ρ.

(B4) For all i, j = 1, . . . , n, Kij(c) = √

ciDij(c)/√

cj is bounded in [0,1]n.

Examples of matricesD(c) satisfying these assumptions are presented in Section 6. Our first main result is the global existence of weak solutions.

Theorem 1 (Global existence). Let Assumptions (A1)–(A2) and (B1)–(B4) hold. Then there exists a weak solution c to (1)–(5) satisfying 0≤ci ≤1, Pn

i=1ci = 1 in Ω×(0,∞), ci ∈Lloc(0,∞;H1(Ω))∩L2loc(0,∞;H2(Ω)), ∂tci ∈L2loc(0,∞;H1(Ω)0),

the initial condition in (4) is satisfied in the sense of H1(Ω)0, and for all φi ∈ C0(Ω× (0,∞)),

0 =− Z

0

Z

citφidxdt+

n

X

j=1

Z 0

Z

Bij(c)∇logci· ∇φidxdt (12)

+

n

X

j=1

Z 0

Z

div(Bij(c)∇φi)∆cjdxdt, where Bij(c) is defined in (8). Furthermore,

H(c(·, T)) +C1E(c(·, T)) +C2

Z T 0

Z

(|∇√

c|2+|∆c|2)dxdt (13)

+C2 Z T

0

Z

|ζ|2dxdt≤ H(c0) +C1E(c0),

where C1 >0 depends on ρ, n, kD(c)kF and C2 >0 depends on n, kD(c)kF (k · kF is the Frobenius matrix norm and ρ is introduced in Assumption (B3)). Moreover, ζ is the weak L2(Ω) limit of an approximating sequence of Pn

j=1PL(c)ij

cj∇µj.

Some comments are in order. First, by Assumption (B2), the elements of the matrixD(c) are bounded for anyc∈[0,1]n and therefore, the quantitykD(c)kF is bounded uniformly inc. Second, the weak formulation (12) makes sence sinceBij(c)∇logci ∈L2(QT). Indeed, by the definition of B(c), we have

Bij(c)∇logcj =√

ciDBDij (c) 1

√cj∇cj,

(7)

and the matrix √

ciDijBD(c)/√

cj is bounded for all c ∈ [0,1]n; see Lemma 4 (iii) below.

However, note that the expressionPn

j=1Bij(c)∇µj is generally not an element ofL2(QT).

In particular, we cannot expect that ∇∆ci ∈ L2(QT). Third, we have not been able to identify the weak limit ζ because of low regularity. However, if Pn

j=1PL(c)ij

cj∇µj ∈ L2loc(0,∞;L2(Ω)) holds for alli= 1, . . . , n, then we can identifyζi =Pn

j=1PL(c)ij

cj∇µj; see Lemma 9.

To prove Theorem 1, we first introduce a truncation with parameterδ∈(0,1) as in [14]

to avoid the degeneracy. Then we reduce the cross-diffusion system ton−1 equations by replacing cn by 1−Pn−1

i=1 ci. The advantage is that the diffusion matrix of the reduced system is positive definite (with a lower bound depending onδ). The existence of solutions cδi to the truncated, reduced system is proved by an approximation as in [29] and the Leray–Schauder fixed-point theorem; see Section 3.1. An approximate version of the free energy estimate (13) (proved in Lemma 8 in Section 3.2) provides suitable uniform bounds that allow us to perform the limit δ → 0. The approximate densities cδi may be negative but, by exploiting the entropy bound for cδi, its limitci turns out to be nonnegative. The limitδ →0 is then performed in Section 3.3, using the uniform estimates and compactness arguments.

Our second main result is concerned with the weak-strong uniqueness. For this, we define the relative entropy and free energy in the spirit of [19] by, respectively,

H(c|¯c) :=H(c)− H(¯c)− ∂H

∂c (¯c)·(c−c) =¯

n

X

i=1

Z

cilogci

¯

ci −(ci−¯ci)

dx, (14)

E(c|¯c) :=E(c)− E(¯c)− ∂E

∂c(¯c)·(c−c) =¯ H(c|¯c) + 1 2

n

X

i=1

Z

|∇(ci−¯ci)|2dx.

(15)

Theorem 2 (Weak-strong uniqueness). Let Assumptions (A1)–(A2), (B1)–(B4) hold, let c be a weak solution to (1)–(5) with initial datum c0, and let c¯ be a strong solution to (1)–(5) with initial datum c¯0. We assume that the weak solution c satisfies

(16)

n

X

j=1

PL(c)ij

cj∇µj ∈L2loc(0,∞;L2(Ω)) for i, j = 1, . . . , n

(see (11) for the definition of PL(c)) and for all T >0 the energy and entropy inequalities E(c(T)) +

n

X

i,j=1

Z T 0

Z

Bij(c)∇µi· ∇µjdxdt≤ E(c0), (17)

H(c(T)) +

n

X

i,j=1

Z T 0

Z

Bij(c)∇logci· ∇µjdxdt≤ H(c0).

(18)

(8)

The strong solution c¯ is supposed to be strictly positive, i.e., there exists m >0 such that

¯

ci ≥m in Ω, t >0, and satisfies the regularity

¯

ci ∈Lloc(0,∞;W3,∞(Ω)), ∇div 1

¯

ciBij(¯c)∇¯µj

∈Lloc(0,∞;L(Ω))

for i= 1, . . . , n, as well as for any T > 0 the energy and entropy conservation identities E(¯c(T)) +

n

X

i,j=1

Z T 0

Z

Bij(¯c)∇¯µi· ∇¯µjdxdt=E(¯c0), (19)

H(¯c(T)) +

n

X

i,j=1

Z T 0

Z

Bij(¯c)∇log ¯ci· ∇¯µjdxdt=H(¯c0), (20)

where µi = logci−∆ci and µ¯i = log ¯ci−∆¯ci. Then, for any T >0, there exist constants C1, only depending on kD(c)kF, n, ρ, and C2(T) >0, only depending on T, meas(Ω), n, ρ, such that

(21) H(c(T)|¯c(T)) +C1E(c(T)|¯c(T))≤C2(T) H(c0|¯c0) +C1E(c0|c¯0) . In particular, if c0 = ¯c0 then the weak and strong solutions coincide.

Observe that we need stronger assumptions on the weak solutions than those obtained in Theorem 1. Assumption (16) guarantees that the flux Pn

j=1Bij(c)∇µj lies in L2(QT).

Indeed, we prove in Lemma 4 (i) in Section 2 that DBDij (c) is bounded for c ∈ [0,1]n. Therefore, sinceDBD(c) = DBD(c)PL(c), assumption (16) and ci ∈L(QT) imply that (22)

n

X

j=1

Bij(c)∇µj =√ ci

n

X

j,k=1

DBDik (c)PL(c)kj

√cj∇µj ∈L2(QT).

By the way, it follows fromPn

j=1PL(c)ij

cj∇logcj = 2∇√

ci ∈L2(QT) that (23)

n

X

j=1

PL(c)ij

cj∇∆cj =

n

X

j=1

PL(c)ij

cj∇(logcj −µj)∈L2(QT).

Since ∇∆ci may be not in L2(QT), we interpret (23) in the sense of distributions, i.e. for all Φ∈C0(Ω;Rd),

n

X

j=1

PL(c)ij

cj∇∆cj

=−

n

X

j=1

Z

∇(PL(c)ij

cj)·Φ +PL(c)ij

cjdiv Φ

∆cjdx.

For the proof of Theorem 2, we estimate first the time derivative of the relative entropy (14):

dH

dt (c|¯c) +C1

n

X

i=1

Z

n

X

j=1

PL(c)ij

cj∇logcj

¯ cj

2

dx+C1

n

X

i=1

Z

(∆(ci−¯ci))2dx

(9)

≤C2

n

X

i=1

Z

n

X

j=1

PL(c)ij

cj∇(µj −µ¯j)

2

dx+C3 Z

E(c|¯c)dx,

where Ci > 0 are some constants depending only on the data. The first term on the right-hand side can be handled by estimating the time derivative of the relative energy (15):

dE

dt(c|c) +¯ C4

n

X

i=1

Z

n

X

j=1

PL(c)ij

cj∇(µj −µ¯j)

2

dx

≤θ

n

X

i=1

Z

n

X

j=1

PL(c)ij

cj∇logcj

¯ cj

2

dx+θ

n

X

i=1

Z

(∆(ci −c¯i))2dx +C5(θ)

Z

E(c|¯c)dx,

where θ > 0 can be arbitrarily small. Choosing θ = C1C4/C2, we can combine both estimates leading to

d dt

H(c|¯c) + C2

C4E(c|¯c)

C3+C2C5

C4

E(c|¯c),

and the theorem follows after applying Gronwall’s lemma. As the computations are quite involved, we compute first in Section 4.1 the time derivative of the relative entropy and energy for smooth solutions. The rigorous proof of the combined relative entropy-energy inequality for weak solutions c and strong solutions ¯c is then performed in Section 4.2.

The paper is organized as follows. The Bott–Duffin matrix inverse is introduced in Section 2, some properties of the mobility matrix B(c) are proved, and the combined energy-entropy inequality (9) is derived for smooth solutions. The global existence of solutions (Theorem 1) is shown in Section 3, while Section 4 is concerned with the proof of the weak-strong uniqueness property (Theorem 2). The case of nonconvex energies is investigated in Section 5, showing that Theorems 1 and 2 still hold if the convex part of the energy equals the Boltzmann entropy (see (6)) and the Hessian of the nonconvex part is bounded. Finally, we present some examples verifying Assumptions (B1)–(B4) in Section 6.

Notation. Elements of the matrix A∈ Rn×n are denoted byAij, i, j = 1, . . . , n, and the elements of a vectorc∈Rnarec1, . . . , cn. We use the notationf(c) = (f(c1), . . . , f(cn)) for c ∈ Rn and a function f : R → R. The expression |∇f(c)|2 is defined by Pn

i=1|∇f(ci)|2 and | · | is the usual Euclidean norm. The matrix R(c) ∈ Rn×n is the diagonal matrix with elements √

c1, . . . ,√

cn, i.e. Rij(c) = √

ciδij for i, j = 1, . . . , n, where δij denotes the Kronecker delta. We understand by ∇µ the matrix with entries ∂xiµj. Furthermore, C >0, Ci >0 are generic constants with values changing from line to line.

(10)

2. Properties of the mobility matrix and a priori estimates

We wish to express the fluxesciuias a linear combination of the gradients of the chemical potentials. SinceK(c) has a nontrivial kernel, we need to use a generalized matrix inverse, the Bott–Duffin inverse. This inverse and its properties are studied in Section 2.1. The properties allow us to derive in Section 2.2 some a priori estimates for the Maxwell–Stefan–

Cahn–Hilliard system.

2.1. The Bott–Duffin inverse. We wish to invert (2) or, equivalently, (7). We recall definition (11) of the projection matrices PL(c) ∈ Rn×n on L(c) and PL(c) ∈ Rn×n on L(c), where L(c) and L(c) are defined in (10). Then (7) is equivalent to the problem:

(24) Solve D(c)z =−PL(c)R(c)∇µ in the spacez ∈L(c), where zi =√

ciui, recalling that R(c) = diag(√ c).

Lemma 3 (Solution of (24)). Suppose that D(c) satisfies Assumption (B1). The Bott–

Duffin inverse

DBD(c) =PL(c) D(c)PL(c) +PL(c)−1

is well-defined, symmetric, and satisfies kerDBD(c) = L(c). Furthermore, for any y ∈ L(c), the linear problem D(c)z = y for z ∈ L(c) has a unique solution given by z = DBD(c)y.

We refer to [25, Lemma 17] for the proof. The property for the kernel follows from kerDBD(c) = kerPL(c) = L(c). Since PL(c)R(c)∇µ ∈ L(c) (this follows from the definition of PL(c) and Pn

i=1ci = 1), we infer from Lemma 3 that (24) has the unique solutionz =−DBD(c)PL(c)R(c)∇µ∈L(c) or, componentwise,

ciui =√

cizi =−

n

X

j=1

√ci DBD(c)PL(c)

ij

√cj∇µj =−

n

X

j=1

√ciDBD(c)ij√ cj∇µj

for i = 1, . . . , n, where the last equality follows from DBD(c)PL(c) = DBD(c); see [25, (81)]. Then we can formulate equation (1) as

(25) ∂tci = div

n

X

j=1

Bij(c)∇µj, where Bij(c) = √

ciDijBD(c)√

cj, i, j = 1, . . . , n.

The boundary conditions ciui·ν = 0 on∂Ω yield (26)

n

X

j=1

Bij(c)∇µj ·ν= 0 on ∂Ω, t >0, i= 1, . . . , n.

We recall some properties of the Bott–Duffin inverse.

Lemma 4(Properties ofDBD(c)). Suppose thatD(c)∈Rn×n satisfies Assumptions (B1)–

(B4). Then:

(i) The coefficients DBDij ∈C1([0,1]n) are bounded for i, j = 1, . . . , n.

(11)

(ii) Let λ(c) be an eigenvalue of (D(c)PL(c) +PL(c))−1. Then λm ≤ λ(c) ≤ λM, where

λm = (1 +nkD(c)kF)−1, λM = max{1, ρ−1},

k · kF is the Frobenius matrix norm, and ρ >0 is a lower bound for the eigenvalues of D(c); see Assumption (B3).

(iii) The functions c7→√

ciDBDij (c)/√

cj are bounded in [0,1]n for i, j = 1, . . . , n.

A consequence of (ii) are the inequalities

(27) λm|PL(c)z|2 ≤zTDBD(c)z ≤λM|PL(c)z|2 for z ∈Rn.

Note that the Frobenius norm of D(c) is bounded uniformly in c ∈ [0,1]n, since Dij is bounded by Assumption (B1).

Proof. The points (i) and (ii) are proved in [25, Lemma 11] in an interval [m,1]n for some m >0. In fact, we can conclude (i)–(ii) in the full interval [0,1]n, since our Assumptions (B2)–(B3) are stronger than those in [25].

For the proof of (iii), dropping the argument c and observing that RDR−1 = K, we obtain

RDBDR−1 =RPL(DPL+PL)−1R−1 =RPL(R−1R)(DPL+PL)−1R−1

=RPLR−1 R(DPL+PL)R−1−1

=RPLR−1 RDR−1RPLR−1+RPLR−1−1

=RPLR−1 KRPLR−1+RPLR−1−1

. The determinant of the expression in the brackets equals

det R(DPL+PL)R−1

= det(DPL+PL).

Therefore, denoting by “adj” the adjugate matrix, it follows that (28) RDBDR−1 = RPLR−1adj(KRPLR−1+RPLR−1)

det(DPL+PL) .

By Assumption (B3), the eigenvalues of D are not smaller than ρ > 0. The proof of [25, Lemma 11] shows that the eigenvalues ofDPL+PL are not smaller thanρ >0, too. This implies that det(DPL+PL)≥ρn−1 >0. The coefficients

(RPLR−1)ijij −ci, (RPLR−1)ij =ci

are bounded forc∈[0,1]nand, by Assumption (B4), the coefficients ofK are also bounded.

Therefore, all elements of adj(KRPLR−1+RPLR−1) are bounded. We conclude from (28) that the entries ofRDBDR−1 are bounded in [0,1]n, i.e., point (iii) holds.

The most important property is the positive definiteness of DBD(c) on L(c); see (27).

This property implies the a priori estimates proved in the following subsection.

(12)

2.2. A priori estimates. We show an energy inequality for smooth solutions.

Lemma 5 (Free energy inequality). Let c ∈ C(Ω×(0,∞);Rn) be a positive, bounded, smooth solution to (1)–(5). Then, for any 0< λ < λm,

d dt

H(c) + (λM −λ)2 λmλ E(c)

+ 2λ

Z

|∇√

c|2dx+λ Z

|∆c|2dx

+ (λM −λ)2

Z

|PL(c)R(c)∇µ|2dx≤0.

where the entropy H(c) and the free energy E(c) are given by (6) and λm, λM are defined in Lemma 4.

Proof. We derive first the energy inequality. To this end, we multiply equation (25) for ci

by µi = (∂E/∂ci)(c), integrate over Ω, integrate by parts (using the boundary conditions (26)), and take into account the lower bound (27) for DBD(c):

dE dt(c) =

n

X

i=1

Z

∂E

∂ci(c)∂tcidx=−

n

X

i,j=1

Z

Bij(c)∇µi· ∇µjdx (29)

=−

n

X

i,j=1

DijBD(c)(√

ci∇µi)·(√

cj∇µj)dx≤ −λm Z

|PL(c)R(c)∇µ|2dx.

The entropy inequality is derived by multiplying (25) by logci, integrating over Ω, and integrating by parts (using the boundary conditions (26)):

dH dt (c) =

n

X

i=1

Z

(logci)∂tcidx=−

n

X

i,j=1

Z

Bij(c)∇logci· ∇µjdx.

To estimate the right-hand side, we set G=RPLR (omitting the argument c) and M :=

B−λG for λ∈(0, λm). Then (30) dH

dt (c) =−

n

X

i,j=1

Z

Mij∇logci· ∇µjdx−λ

n

X

i,j=1

Z

Gij∇logci· ∇µjdx=:I1+I2. Before estimating the integrals I1 and I2, we start with some preparations. We use Lemma 4 (ii) and PLTPL=PL to obtain

zTBz = (Rz)TDBDRz ≥λm|PLRz|2m(PLRz)T(PLRz) =λmzTGz for z ∈Rn. The matrix M is positive semidefinite since for any z ∈Rn,

(31) zTMz=zTBz−λzTGz ≥(λm−λ)zTGz = (λm−λ)|PLRz|2. Furthermore, by Lemma 4 (ii) again, we have the upper bound

(32) zTMz =zT(B−λG)z ≤(λM −λ)zTGz = (λM −λ)|PLRz|2.

(13)

We are now in the position to estimate the integral I1, using Young’s inequality for any θ >0:

I1 ≤ θ 2

n

X

i,j=1

Z

Mij∇logci· ∇logcjdx+ 1 2θ

n

X

i,j=1

Z

Mij∇µi· ∇µjdx (33)

≤ θ

2(λM −λ) Z

|PLR∇logc|2dx+λM −λ 2θ

Z

|PLR∇µ|2dx

= 2θ(λM −λ) Z

|∇√

c|2dx+λM −λ 2θ

Z

|PLR∇µ|2dx,

where the last step follows from Pn

j=1(PL)ijRj∇logcj = 2∇√

ci, which is a consequence of Pn

j=1∇cj = 0. For the integral I2, we use the definitions Gij = ciδij −cicj and µj = logcj −∆cj:

I2 =−λ

n

X

i,j=1

Z

(ciδij −cicj)∇ci

ci · ∇(logcj−∆cj)dx

=−λ

n

X

i=1

Z

∇ci· ∇(logci −∆ci)dx+λ Z

n

X

i=1

∇ci·

n

X

j=1

cj∇(logcj−∆cj)dx

=−λ

n

X

i=1

Z

∇ci· ∇(logci −∆ci)dx=−λ Z

4|∇√

c|2+|∆c|2 dx,

where we integrated by parts in the last step.

Inserting the estimates for I1 and I2 into (30) yields dH

dt (c) + 4λ Z

|∇√

c|2dx+λ Z

|∆c|2dx

≤2θ(λM −λ) Z

|∇√

c|2dx+λM −λ 2θ

Z

|PLR∇µ|2dx.

We set θ=λ/(λM −λ) to conclude that

(34) dH

dt (c) + 2λ Z

|∇√

c|2dx+λ Z

|∆c|2dx≤ (λM −λ)2

Z

|PLR∇µ|2dx.

The right-hand side can be absorbed by the corresponding term in (29). Indeed, adding the previous inequality to (29) times (λM −λ)2/(λmλ) finishes the proof.

Note that the energy inequality (29) or the entropy inequality (34) alone are not sufficient to control the derivatives of c but only a suitable linear combination. We will prove these inequalities rigorously in the following section for weak solutions; see Lemma 8.

(14)

3. Proof of Theorem 1

We prove the existence of global weak solutions to (1)–(4). For this, we construct an approximate system depending on a parameter δ > 0, similarly as in [14], and then pass to the limit δ→0.

3.1. An approximate system. In order to deal with the degeneracy of the matrixB(c) when a component of c vanishes, we introduce the cutoff function χδ :Rn→Rn by

δc)i :=

δ for ci < δ,

ci for δ≤ci ≤1−δ, 1−δ for ci >1−δ, and define the approximate matrix

(35) Bδ(c) :=R(χδc)DBDδc)R(χδc), recalling that R(χδc) = diag(√

χδc). We wish to solve the approximate problem

tcδi = div

n

X

j=1

Bijδ(cδ)∇µδj, µδj = ∂Eδ

∂cj(cδ) in Ω, t >0, (36)

cδi(·,0) =c0i in Ω,

n

X

j=1

Bijδ(cδ)∇µδj ·ν = 0, ∇cδi ·ν = 0 on ∂Ω, (37)

where i= 1, . . . , n, Pn

i=1c0i = 1 and the approximate energy is defined by Eδ(c) :=Hδ(c) + 1

2

n

X

i=1

Z

|∇ci|2dx, Hδ(c) :=

n

X

i=1

Z

hδi(ci)dx,

hδi(r) =

rlogδ−δ/2 +r2/(2δ) for r < δ,

rlogr for δ≤r≤1−δ,

rlog(1−δ)−(1−δ)/2 +r2/(2(1−δ)) for r >1−δ.

(38)

Observe that the solutionscδi may be negative. We will show below that cδi converges to a nonnegative function as δ →0. The approximate entropy density is chosen in such a way that hδi ∈C2(R). Indeed, we obtain

(hδi)0(ci) =

logδ+ci/δ for ci < δ,

logci+ 1 for δ < ci <1−δ, log(1−δ) +ci/(1−δ) for ci >1−δ,

(hδi)00(ci) = 1 (χδc)i. With these definitions, we obtain µδi = (hδi)0(cδi)−∆cδi for i= 1, . . . , n.

Theorem 6 (Existence for the approximate system). Let Assumptions (A1)–(A2) and (B1)–(B4) hold and let δ > 0. Then there exists a weak solution (cδδ) to (36)–(37) satisfying Pn

i=1cδi(t) = 1 in Ω, t >0,

cδi ∈Lloc(0,∞;H1(Ω))∩L2loc(0,∞;H2(Ω)),

tci ∈L2loc(0,∞;H2(Ω)0), µδi ∈L2loc(0,∞;H1(Ω)), i= 1, . . . , n,

(15)

and the first equation in (36) as well as the initial condition in (37) are satisfied in the sense of L2loc(0,∞;H2(Ω)0).

The proof of this theorem is deferred to Appendix A, since it is technical and involves well-established techniques. We show some properties of the matrix Bδ(c). We introduce the matrices PLδc), PLδc)∈Rn×n with entries

PLδc)ijij

p(χδc)iδc)j Pn

k=1δc)k , PLδc)ij =

p(χδc)iδc)j Pn

k=1δc)k , i, j = 1, . . . , n.

Lemma 7 (Properties of Bδ(c)). Suppose that D(c) satisfies Assumptions (B1)–(B4).

Then Lemmas 3 and 4 hold withPL(c), PL(c), andDBD(c)replaced byPLδc), PLδc), and DBDδc). As a consequence, the matrix Bδ(c), defined in (35), satisfies

(39) zTBδ(c)z≥λm|PLδc)R(χδc)z|2 for any z,c∈Rn,

and the first (n−1)×(n−1) submatrix Beδ(c) of Bδ(c) is positive definite and satisfies for η(δ) =λmδ2/n,

(40) zeTBeδ(c)ze≥η(δ)|z|e2 for any ze∈Rn−1.

Proof. It can be verified that Assumptions (B1)–(B2) hold for D(χδc), so Lemmas 3 and 4 still hold for the matrix D(χδc). Inequality (39) is a direct consequence of Lemma 4 (ii).

It remains to prove (40). We define for given ze∈Rn−1 the vector z ∈Rn with zi =zei for i= 1, . . . , n−1 and zn= 0. Then (39) becomes

(41) zeTBeδ(c)ze≥λm

ePLδc)R(χe δc)ze

2m R(χe δc)zeT

PeLδc) R(χe δc)ze ,

where Ae denotes the first (n−1)×(n −1) submatrix of a given matrix A ∈ Rn×n. It follows from the Cauchy–Schwarz inequality that for anyζ ∈Rn−1,

ζTPeLδc)ζ =

n−1

X

i=1

ζi2

n−1

X

j=1

s

δc)j

Pn

k=1δc)kζj

!2

≥ |ζ|2

n−1

X

j=1

δc)j

Pn

k=1δc)k|ζ|2

= (χδc)n Pn

k=1δc)k|ζ|2 ≥ δ n|ζ|2. Therefore, (41) becomes

zeTBeδ(c)ze≥ λmδ n

n−1

X

i=1

p(χδc)izei

2 = λmδ n

n−1

X

i=1

δc)i ezi

2 ≥ λmδ2 n |z|e2,

which proves (40).

3.2. Uniform estimates. We derive energy and entropy estimates for the solutions to (36), being uniform in δ.

(16)

Lemma 8 (Energy and entropy inequalities). Let cδ be a weak solution to (36)–(37), constructed in Theorem 6. Then the following inequalities hold for any T >0,

Eδ(cδ(·, T)) +

n

X

i,j=1

Z T 0

Z

Bδij(cδ)∇µδi · ∇µδjdxdt≤ Eδ(c0), (42)

Hδ(cδ(·, T)) +

n

X

i,j=1

Z T 0

Z

Bijδ(cδ)∇(hδi)0(cδi)· ∇µδjdxdt≤ Hδ(c0), (43)

Hδ(cδ(·, T)) + (λM −λ)2

mλ Eδ(cδ(·, T)) +λ

n

X

i=1

Z T 0

Z

|∇cδi|2δcδ)idxdt (44)

n

X

i=1

Z T 0

Z

(∆cδi)2dxdt+(λM −λ)2

Z T 0

Z

PLδcδ)R(χδcδ)∇µδ

2dxdt

≤ Hδ(c0) + (λM −λ)2

mλ Eδ(c0),

where 0< λ < λm, λm, λM are introduced in Lemma 4, and R(χδcδ) = diag(p χδcδ).

Proof. Summing (94) with σ = 1 overk = 1, . . . , N, we find that Eeδ(ce(τ)(·, T)) +

n−1

X

i,j=1

Z T 0

Z

Beδij(ce(τ))∇wi(τ)· ∇w(τ)j dxdt +ε

n

X

i=1

Z T 0

Z

(∆w(τ)i )2+ (w(τ)i )2

dxdt≤Eeδ(ce0).

We know from (97) and the construction of χδ that (w(τ)) is bounded in L2(0, T;H1(Ω)) and (Beijδ(ec)) is bounded in L(QT) with respect to (ε, τ). Therefore, we can pass to the limit (ε, τ) → 0 in the previous inequality, and weak lower semicontinuity of the integral functionals leads to (42).

To show (43), we use (hδi)0(cδi)−(hδi)0(cδn) as a test function in the weak formulation of (87) and sum over i= 1, . . . , n−1:

Hδ(c(·, T)) +

n−1

X

i,j=1

Z T 0

Z

Beijδ(ecδ)∇ (hδi)0(cδi)−(hδi)0(cδn)

· ∇wδjdxdt≤ Hδ(c0).

This inequality can be rewritten as (43) using wδi = µδi −µδn. Finally, we derive (44) by combining (43) and (42) and proceeding as in the proof of Lemma 5.

3.3. Proof of Theorem 1. We perform the limit δ → 0 to finish the proof of Theorem 1. It follows from [15, Lemma 2.1] that for sufficiently small δ > 0, there exists C > 0 (independent of δ) such that for allr1, . . . , rn∈R satisfying Pn

i=1ri = 1, (45)

n

X

i=1

hδi(ri)≥ −C.

Referensi

Dokumen terkait

The spindle cells demonstrated strong nuclear and cytoplasmic staining to S-100 protein while the granular cells demonstrated only weak cytoplasmic staining Fig.. Reactivity of the