This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Invariant cones for linear elliptic systems
with gradient coupling

I. Capuzzo Dolcetta a{}^{\hbox{\small{ a}}}, L. Rossi a,b {}^{\hbox{\small{ a,b }}} and A. Vitolo c{}^{\hbox{\small{ c}}}

{}^{\hbox{a }}Dipartimento di Matematica, Sapienza Università di Roma, Italy
{}^{\hbox{b }}CAMS, CNRS-EHESS, Paris, France
{}^{\hbox{c }}Dipartimento di Ingegneria Civile, Università di Salerno, Italy


Abstract. We discuss counterexamples to the validity of the weak Maximum Principle for linear elliptic systems with zero and first order couplings and prove, through a suitable reduction to a nonlinear scalar equation, a quite general result showing that some algebraic condition on the structure of gradient couplings and a cooperativity condition on the matrix of zero order couplings guarantee the existence of invariant cones in the sense of Weinberger [21].

MSC 2010 Numbers: 35J47, 35J70, 35B50, 35P30, 35D40 111This work has been partially supported by GNAMPA-INdAM
[email protected]
[email protected]
[email protected]

Keywords and phrases: elliptic differential inequalities, weak Maximum Principle, invariant cones, Bellman operators, principal eigenvalue.

1 Introduction

We consider smooth vector-valued functions u=(u1,,um)u=(u_{1},\dots,u_{m}) of the variable xx in a bounded open subset Ωn\Omega\subset{\mathbb{R}}^{n} satisfying linear systems of partial differential inequalities of the following form

Au+i=1nB(i)Diu+Cu0inΩAu+\sum_{i=1}^{n}B^{(i)}D_{i}u+Cu\geq 0\;\;\hbox{in}\;\;\Omega (1.1)

where AA is the second order operator

Au=(Δu1...Δum)Au=\left(\begin{array}[]{c}\Delta u_{1}\\ .\\ .\\ .\\ \Delta u_{m}\end{array}\right) (1.2)

B(i)B^{(i)} and CC are m×mm\times m real matrices and with constant coefficients, and for i=1,,ni=1,\dots,n,

Diu=(u1xi...umxi)D_{i}u=\left(\begin{array}[]{c}\frac{\partial u_{1}}{\partial x_{i}}\\ .\\ .\\ .\\ \frac{\partial u_{m}}{\partial x_{i}}\end{array}\right) (1.3)

denotes the ithi-th column of the Jacobian matrix of the vector function uu.
Note that the above defined structure of the systems allows coupling between the uju_{j} and their gradients but not at the level of second derivatives.
Specific assumptions on the B(i)B^{(i)} and CC will be made later on.

Systems of this kind naturally arise in several different contexts such as modeling of simultaneous diffusions of mm substances which decay spontaneously or in the case of systems describing switching diffusion processes in probability theory. In the latter case the homogeneous Dirichlet problem for system (1.1) describes discounted exit times from Ω\Omega, see for example [11].

We are interested here in investigating the validity of the weak Maximum Principle, wMP in short, that is the sign propagation property from the boundary to the interior for solutions u=(u1,,um)u=(u_{1},\dots,u_{m}) of the differential inequalities (1.1), i.e.,

𝐰𝐌𝐏:u0onΩu0inΩ.\mathbf{wMP}:\quad u\leq 0\ \mbox{on}\ \partial\Omega\ \ \implies\ \ u\leq 0\ \mbox{in}\ \Omega. (1.4)

The vector function uu will be always assumed to belong to [C2(Ω)]m[C0(Ω¯)]m[C^{2}(\Omega)]^{m}\cap[C^{0}(\overline{\Omega})]^{m} and we will adopt the standard notation u0u\leq 0 if uj0u_{j}\leq 0 for each j=1,,mj=1,\dots,m. We adopt the same notation for real-valued matrices, namely for a matrix AA, A0A\geq 0 means that all its entries are nonnegative.
The validity of wMP is well-understood in the scalar case m=1m=1 even for general degenerate elliptic fully nonlinear partial differential inequalities such as

F(x,u,u,2u)0F(x,u,\nabla u,\nabla^{2}u)\geq 0

in a bounded Ω\Omega and also in some unbounded domain of n{\mathbb{R}}^{n}, see [8, 9] for recent results in this direction. Let us point out that the wMP property in the scalar case is related, and in fact equivalent, to the positivity of the principal eigenvalue (may be a pseudo one, if degeneracy occur in the dependence of FF with respect to Hessian matrix 2u\nabla^{2}u) of the Dirichlet problem for FF, see [2],[3].

The case m>1m>1 has been the object of several papers mainly in the case of diagonal weakly coupled systems, that is when the matrices B(i)B^{(i)} are diagonal and couplings between the functions uju_{j} only occur at the level of zero-order terms, described by a matrix C=(cjk)j,kC=(c_{jk})_{j,k}, satisfying the cooperativity condition

cjk0for jk,k=1mcjk0for j=1,,mc_{jk}\geq 0\ \ \mbox{for }j\neq k\;,\qquad\sum_{k=1}^{m}c_{jk}\leq 0\ \ \mbox{for }\,j=1,\dots,m (1.5)

Referring to the aforementioned exit time model, condition (1.5) requires the discount factor for the jj-th process to dominate the sum of the interactions coefficients with all the other processes.

In the framework of purely weak cooperative couplings, let us mention the results in Section 8 of the book by Protter and Weinberger [18] and the references therein. For generalizations of those results in some semilinear cases see [19],[5],[1], while [6] contains results in the same direction concerning fully nonlinear uniformly elliptic operators F=F(x,u,u,2u)F=F(x,u,\nabla u,\nabla^{2}u).
The recent paper [10] extends the validity of some of the results in [6] concerning wMP to a large class of fully nonlinear degenerate elliptic operators.
In particular, for the case of linear systems as (1.1) with no coupling in first derivatives (i.e. when each B(i)B^{(i)} is diagonal), the main result in [10] is that wMP holds true for system (1.1) provided CC is cooperative.
Let us also point out that the main result in [10] holds even in the more general case where the Laplace operator is replaced by more general expressions Tr(Aj2uj)\mathrm{Tr}(A^{j}\nabla^{2}u_{j}) satisfying Tr(Aj)>0\mathrm{Tr}(A^{j})>0 for j=1,,mj=1,\dots,m.

When coupling in first order terms occurs in (1.1), simple examples as the following one taken from [6] show that the wMP property (1.4) may indeed fail:

Example 1.

The vector u(x1,x2)=(1x12x22,13x13+4x220)u(x_{1},x_{2})=(1-x_{1}^{2}-x_{2}^{2},\frac{1}{3}x_{1}^{3}+4x_{2}-20) is a solution of

{Δu1+u2x2=0Δu2+u1x1=0\begin{cases}\displaystyle\Delta u_{1}+\frac{\partial u_{2}}{\partial x_{2}}=0\\ \displaystyle\Delta u_{2}+\frac{\partial u_{1}}{\partial x_{1}}=0\par\end{cases}

in the unit ball Ω2\Omega\subset{\mathbb{R}}^{2}, u1=0u_{1}=0, u2<0u_{2}<0 on Ω\partial\Omega but u1>0u_{1}>0 in Ω\Omega. Observe that the zero-order matrix is C0C\equiv 0 in this example, so that (1.5) is fulfilled.

As a matter of fact, even a first-order coupling of arbitrarily small size in the system can be responsible of the loss of wMP, as the following example shows:

Example 2.

The system

{Δuεvx10Δvεux10\begin{cases}\displaystyle\Delta u-\varepsilon\frac{\partial v}{\partial x_{1}}\geq 0\\ \displaystyle\Delta v-\varepsilon^{\prime}\frac{\partial u}{\partial x_{1}}\geq 0\end{cases}

in a bounded domain Ωn\Omega\subset{\mathbb{R}}^{n}, fulfills wMP if and only if ε=ε=0\varepsilon=\varepsilon^{\prime}=0.
Indeed the validity of wMP when ε=ε=0\varepsilon=\varepsilon^{\prime}=0 is classical. Conversely, if, say, ε0\varepsilon^{\prime}\neq 0, then wMP is violated by the pair

u(x)=δ|xx¯|2,v(x)=v(x1)=C(ex1H),u(x)=\delta-|x-\bar{x}|^{2},\qquad v(x)=v(x_{1})=C(e^{-x_{1}}-H),

where x¯Ω\bar{x}\in\Omega and δ>0\delta>0 is small enough to have u<0u<0 on Ω\partial\Omega, and C1C\gg 1, H1H\gg 1.

Example 2 enlightens an instability property of wMP for cooperative systems with respect to first order perturbations. This is in striking contrast with the scalar case. Indeed, for a uniformly elliptic scalar inequality, not only the presence of a first order term does not affect the validity of wMP when the zero-order term is nonpositive, but in addition wMP is stable with respect to perturbations of the coefficients, in the LL^{\infty} norm. This can be seen as a consequence of the fact that wMP is characterized by the positivity of the associated principal eigenvalue, and the latter depends continuously on the coefficients of the operator, see e.g. [18, 3] and also [5, 6] where such characterization in terms of the same notion of principal eigenvalue as in [3] is extended to cooperative systems without first-order coupling. Example 2 reveals either that such notion does not exist when there is a first-order coupling, or that it is not continuous with respect to the coefficients.

According to the above considerations, two perspectives can be adopted in order to investigate the sign-propagation properties for coupled systems such as (1.1). The first one consists in strengthening the hypotheses on the coefficients of the operators, namely the cooperativity condition (1.5). The second one is to replace wMP by some different kind of propagation property which reflects in some way the geometry of the coupling terms. We will explore both directions.

Observe that the systems in Examples 1 and 2 fulfill the cooperativity condition (1.5) in the “border case”, that is, when all inequalities are replaced by equalities. A natural question is then whether it is possible, for the wMP to hold, to allow some coupling in first-order terms in the system, at least when the cooperativity conditions (1.5) hold with strict inequalities, i.e.,

cjkKfor jk,k=1mcjkKfor j=1,,m,c_{jk}\geq K\ \ \mbox{for }j\neq k\;,\qquad\sum_{k=1}^{m}c_{jk}\leq-K\ \ \mbox{for }\,j=1,\dots,m, (1.6)

with KK possibly very large. The next result shows that this is not possible.

Proposition 1.

Let ε>0\varepsilon>0, α,c~0\alpha,\tilde{c}\geq 0. Then the following system with m=2m=2 and n=1n=1

{u′′±εvcu+αv0v′′c~v0xIρ=(0,ρ)\left\{\begin{array}[]{lll}u^{\prime\prime}\pm\varepsilon v^{\prime}-cu+\alpha v\geq 0\\ v^{\prime\prime}-\tilde{c}v\geq 0\quad\quad\quad\quad\quad\quad x\in I_{\rho}=(0,\rho)\end{array}\right. (1.7)

where uu,vv are scalar functions of xx\in{\mathbb{R}} does not satisfy wMP, provided that

ζ(ρc)c>αε where ζ(τ):=coshτ1sinhττ.\zeta(\rho\sqrt{c})\sqrt{c}>\frac{\alpha}{\varepsilon}\quad\mbox{ where }\ \zeta(\tau):=\frac{\cosh\tau-1}{\sinh\tau-\tau}\;. (1.8)
Remark 3.

Since ζ(0+)=+\zeta(0^{+})=+\infty and ζ(+)=1\zeta(+\infty)=1, this proposition entails that, for every ε,K>0\varepsilon,K>0, there exists a system of the type (1.1), with B(i)B^{(i)} satisfying |Bjk(i)|ε|B^{(i)}_{jk}|\leq\varepsilon and C=(cjk)j,kC=(c_{jk})_{j,k} satisfying (1.6), for which wMP fails. Namely, even an arbitrary small amount of coupling at the level of first derivatives can prevent the validity of wMP although the zero order matrix is, so to say, “very strongly cooperative”. It also shows that, for any ε,c>0\varepsilon,c>0 and α,c~0\alpha,\tilde{c}\geq 0, wMP fails for (1.7) in a small enough interval IρI_{\rho}. The fact that wMP fails when the diagonal zero-order term cc is sufficiently large or when the size ρ\rho of the interval is sufficiently small can be surprising, if one has in mind the picture for the scalar equation (where both having a large –negative– zero-order term and a small domain help the validity of the maximum principle).
This phenomenon could be related to a non-monotonic structure of the system when a first-order coupling is in force.

Remark 4.

A few more comments are in order here. We are considering a system with coupled gradients (ε>0\varepsilon>0). The first part of Proposition 1 says that wMP cannot be satisfied in all bounded domains as soon as ε>0\varepsilon>0, whatever the amount of cooperativity (α>0\alpha>0) is. The second part means that in a fixed interval wMP fails for cc large enough. In cooperative systems under consideration (0αc0\leq\alpha\leq c) an excess of coercivity with respect to the coupling (cc large compared with α/ε\alpha/\varepsilon) seems to be responsible for invalidating wMP. In particular this is the case in any interval IρI_{\rho} when α=0\alpha=0.

We exhibit in Proposition 2 below that the same qualitative phenomenon occurs for a larger class of systems. The proofs of Propositions 1 and 2 are detailed in Section 2.

Proposition 2.

For every ε0\varepsilon\neq 0, c~>0\tilde{c}>0 and ε~,α,β\tilde{\varepsilon},\alpha,\beta\in{\mathbb{R}}, there exists c>0c>0 large enough such that the system

{u′′εvcu+αv0v′′ε~uc~v+βu0in (0,1)\left\{\begin{array}[]{lll}u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v\geq 0\\ v^{\prime\prime}-\tilde{\varepsilon}u^{\prime}-\tilde{c}v+\beta u\geq 0\end{array}\right.\quad\text{in }(0,1) (1.9)

violates the wMP.
If, instead, c>0c>0 is also fixed in the system (1.9), there exists an interval I(0,1)I\subset(0,1) in which such system does not fulfill the
wMP.

Let us turn now to the positive results. The study of sufficient conditions for the validity of the weak Maximum Principle in the form wMP in the case where coupling occurs also at the level of first or second order derivatives is apparently less explored in literature, see however [13],[14],[15] and also [17],[16] for the related issue of maximum norm estimates of the form supxΩ|u(x)|CsupxΩ|f(x)|\sup_{x\in\Omega}|u(x)|\leq C\,\sup_{x\in\Omega}|f(x)| for solutions uu of non-homogeneous systems of equations involving higher order couplings.

The wMP property (1.4) can be understood in the framework of the general theory of invariant sets introduced by H. F. Weinberger in [21] in the context of elliptic and parabolic weakly coupled systems. We refer to the recent paper by G. Kresin and V. Mazya [15] where the notion of invariance is thoroughly developed for general systems with couplings at the first and the second order in the case C0C\equiv 0.
According to the notion introduced in [21], a set SmS\subseteq{\mathbb{R}}^{m} is invariant for system (1.1) if the following property holds

𝐈𝐍𝐕:u(x)Sfor allxΩu(x)Sfor allxΩ\mathbf{INV}:\qquad u(x)\in S\;\;\hbox{for all}\;x\in\partial\Omega\ \ \implies\ \ u(x)\in S\;\;\hbox{for all}\;x\in\Omega (1.10)

The sign propagation property (1.4) can then be rephrased as the property of the negative orthant m={u=(u1,um):uj0,j=1,,m}{\mathbb{R}}^{m}_{-}=\{u=(u_{1},\dots u_{m}):u_{j}\leq 0\,\,,\,j=1,\dots,m\} being an invariant set for system (1.1) of partial differential inequalities.
In [21] it is proved in particular that wMP holds for weakly coupled uniformly elliptic systems such as

Tr(Aj2uj)+bjuj+f(u)=0,j=1,,m\mathrm{Tr}(A^{j}\nabla^{2}u_{j})+b^{j}\cdot\nabla u_{j}+f(u)=0\;,\;j=1,\dots,m (1.11)

under the condition that the vector field ff satisfies the property that for any pp belonging to the outward normal cone to m{\mathbb{R}}^{m}_{-} at a point uu on the boundary of m{\mathbb{R}}^{m}_{-} the inequality

pf(u)0p\cdot f(u)\leq 0 (1.12)

holds. For f(u)=Cuf(u)=Cu, this geometric condition turns out to be the cooperativity property (1.5) of matrix CC. Note also that this condition implies that m{\mathbb{R}}^{m}_{-} is invariant under the flow du/dt=Cu,t>0du/dt=Cu\,,t>0.

We recall that Proposition 1 entails that m{\mathbb{R}}^{m}_{-} may fail to be an invariant set even when the coupling of the first order terms is very small. As a matter of fact, the first order matrix of system (1.7) is

(0ε00)\left(\begin{array}[]{cc}0&-\varepsilon\\ 0&0\end{array}\right)

which is not diagonalizable. This is indeed consistent with results in [15]. It is in fact shown in that paper, see in particular results in Section 3, that the sufficient conditions involving the relations between the geometry of a closed convex set SS and the matrices B(i)B^{(i)} which imply the invariance of SS, necessarily require, in the case S=mS={\mathbb{R}}^{m}_{-}, the diagonal structure of the first order couplings.

On the account of the example (1.7) exhibited in Proposition 1 we are forced to investigate the validity of a weaker form of the sign propagation property or, in other words, to single out an appropriate invariant set for system (1.1) when first order couplings occur.

It turns out that under some algebraic conditions, including notably the simultaneous diagonalizability of the matrices B(i)B^{(i)}, a cone propagation type result holds:

Theorem 3.

Let Ω\Omega be a bounded open subset of n{\mathbb{R}}^{n}. Assume that there exists an invertible m×mm\times m matrix QQ such that, for all i=1,,ni=1,\dots,n,

Q1B(i)Q=Diag(β1(i),,βm(i))for someβj(i),(j=1,,m)Q^{-1}B^{(i)}Q=\mathrm{Diag}\big{(}\beta_{1}^{(i)},\dots,\beta_{m}^{(i)}\big{)}\quad\hbox{for some}\;\;\beta^{(i)}_{j}\in{\mathbb{R}},\ \ \ (j=1,\dots,m) (1.13)
Q10Q^{-1}\geq 0 (1.14)

and, moreover,

Q1CQfulfills the cooperativity condition (1.5)Q^{-1}CQ\ \ \hbox{fulfills the cooperativity condition \eqref{coop}} (1.15)

Then the convex cone S={um:Q1u0}S=\{u\in{\mathbb{R}}^{m}:Q^{-1}u\leq 0\} is invariant for system (1.1).

Remark 5.

Concerning the linear algebraic conditions of Theorem 3, observe first that a matrix QQ simultaneously satisfying (1.11) for i=1,,ni=1,\dots,n exists if the B(i)B^{(i)}’s have a common basis of eigenvectors. This is the case when the matrices B(i)B^{(i)} commute each other for all i=1,,n.i=1,\dots,n. Observe also that if QQ is an invertible M-matrix, that is Q=sIXQ=sI-X where X0X\geq 0 and ss is strictly greater than the spectral radius of XX, then QQ fulfills condition (1.14), see [4].
Next, it is worth to point out that conditions (1.14) and (1.15) are compatible.
For example, Q=(2112)Q=\left(\begin{array}[]{cc}2&-1\\ -1&2\end{array}\right) is an invertible M-matrix, C=(3212)C=\left(\begin{array}[]{cc}-3&2\\ 1&-2\end{array}\right) is cooperative and Q1CQ=(4301)Q^{-1}CQ=\left(\begin{array}[]{cc}-4&3\\ 0&-1\end{array}\right) is cooperative as well.
If no coupling occurs in first derivatives, so that Q=Q1=IQ=Q^{-1}=I, the above result reproduces the one in [10].

Remark 6.

A related remark is that permutation matrices satisfies both Q10Q^{-1}\geq 0 and Q0Q\geq 0, so that in this case the conclusion of Theorem 3 is in fact that the negative orthant RmR^{m}_{-} is invariant. However, it is easy to check that in this situation condition (1.13) implies that each B(i)B^{(i)} is diagonal and the results of [10] apply.
A further remark is that one cannot expect in general the invariance of the negative orthant RmR^{m}_{-}. This is indeed coherent with results in [15]; Lemma 2 there states in fact that the geometric sufficient condition on the matrices B(i)B^{(i)} guaranteeing the invariance of RmR^{m}_{-} implies their diagonal structure.

Remark 7.

Theorem 3 can in fact be extended (with a completely analogous proof) to a second order matrix operator

Au=(Tr(A2u1)...Tr(A2um))Au=\left(\begin{array}[]{c}\mathrm{Tr}(A\nabla^{2}u_{1})\\ .\\ .\\ .\\ \mathrm{Tr}(A\nabla^{2}u_{m})\end{array}\right) (1.16)

where AA is a positive semidefinite matrix such that Aννλ>0A\nu\cdot\nu\geq\lambda>0 for some direction νn\nu\in{\mathbb{R}}^{n}. For some applications of this notion of directional uniform ellipticity condition see [7],[8],[9],[20].

A key role in the proof of this result, which is postponed to the next section, is based on a reduction to a suitable fully nonlinear scalar differential inequality governed by the elliptic convex Bellman-type operator FF defined, on scalar functions ψ:Ω\psi:\Omega\to{\mathbb{R}}, as

F[ψ]=Δψ+maxj=1,,mi=1nβj(i)ψxi=Δψ+maxj=1,,mbjψF[\psi]=\Delta\psi+\max_{j=1,\dots,m}\sum_{i=1}^{n}\beta^{(i)}_{j}\frac{\partial\psi}{\partial x_{i}}=\Delta\psi+\max_{j=1,\dots,m}b^{j}\cdot\nabla\psi (1.17)

where βj(i)\beta^{(i)}_{j} are as in (1.13) and bj:=(βj(1),,βj(m))b^{j}:=(\beta^{(1)}_{j},\dots,\beta^{(m)}_{j}). The main ingredients in the proof are results in [10], see in particular Theorems 1.1 and 1.3, and the notion of generalized principal eigenvalue for scalar fully nonlinear degenerate elliptic operators and its relations with the validity of wMP, see [2].

The next example provides a simple illustration of the result of Theorem 3:

Example 8.

Let u=(u1,u2)u=(u_{1},u_{2}) be a solution of

{Δu1+6u1x1+u2x1u10Δu28u1x1u20\begin{cases}\displaystyle\Delta u_{1}+6\frac{\partial u_{1}}{\partial x_{1}}+\frac{\partial u_{2}}{\partial x_{1}}-u_{1}\geq 0\\ \displaystyle\Delta u_{2}-8\frac{\partial u_{1}}{\partial x_{1}}-u_{2}\geq 0\end{cases}

in a bounded domain Ωn\Omega\subset{\mathbb{R}}^{n}. In this case B(1)=(6180)B^{(1)}=\left(\begin{array}[]{cc}6&1\\ -8&0\end{array}\right), B(2)=0B^{(2)}=0, C=(1001)C=\left(\begin{array}[]{cc}-1&0\\ 0&-1\end{array}\right) and Theorem 3 applies with

Q=(11/241)Q1=(11/241)Q=\left(\begin{array}[]{cc}-1&1/2\\ 4&-1\end{array}\right)\quad Q^{-1}=\left(\begin{array}[]{cc}1&1/2\\ 4&1\end{array}\right)

yielding that inequality u2min(2u1;4u1)u_{2}\leq\min(-2u_{1};-4u_{1}) propagates from Ω\partial\Omega to the whole Ω\Omega.

The result of Theorem 3 can be somewhat refined by a suitable weakening of the assumptions there. Firstly, observe that B(i)B^{(i)} is not necessarily diagonalizable. A suitable change of basis generally generally leads to an upper triangular matrix, which yields a real Jordan canonical form of B(i)B^{(i)}.
Suppose that the B(i)B^{(i)} ’s have a common eigenspace of dimension kmk\leq m and consider a basis of m\mathbb{R}^{m} where the first kk vectors are linearly independent (common) eigenvectors of B(i)B^{(i)}, then we can find an m×mm\times m invertible real matrix Q^\hat{Q} that produces a real Jordan canonical form J(i)=Q^1B(i)Q^J^{(i)}=\hat{Q}^{-1}B^{(i)}\hat{Q}, where the k×mk\times m sub-matrix with the first kk rows is made up by a k×kk\times k diagonal block Λ\Lambda plus the k×(mk)k\times(m-k) zero matrix.

In this setting we have the following:

Theorem 4.

Assume in addition to the above that the k×mk\times m sub-matrix containing the first kk rows of C^=Q^1CQ^\hat{C}=\hat{Q}^{-1}C\hat{Q} is made by a cooperative k×kk\times k block plus the k×(mk)k\times(m-k) zero matrix. Let pijp_{ij} be the entries of the matrix P^:=Q^1\hat{P}:=\hat{Q}^{-1}.
If the k×mk\times m sub-matrix with the first kk rows of P^\hat{P} is positive, then the closed convex set

S^={um:j=1mp1juj0,,j=1mpkjuj0}\hat{S}=\left\{u\in{\mathbb{R}}^{m}:\sum_{j=1}^{m}p_{1j}u_{j}\leq 0,\dots,\sum_{j=1}^{m}p_{kj}u_{j}\leq 0\right\}

is invariant for the system (1.1).

An illustrative example is provided next:

Example 9.

Consider the 2×22\times 2 system

{Δu1+u2x1u10Δu2+u1x1u20\begin{cases}\displaystyle\Delta u_{1}+\frac{\partial u_{2}}{\partial x_{1}}-u_{1}\geq 0\\ \displaystyle\Delta u_{2}+\frac{\partial u_{1}}{\partial x_{1}}-u_{2}\geq 0\end{cases}

in a domain Ω\Omega of 2{\mathbb{R}}^{2}. In this case B(1)=(0110)B^{(1)}=\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right), B(2)=0B^{(2)}=0, C=(1001)C=\left(\begin{array}[]{cc}-1&0\\ 0&-1\end{array}\right) and Theorem 4 applies with

Q^=(1111)Q^1=(1/21/21/21/2)\hat{Q}=\left(\begin{array}[]{cc}1&1\\ 1&-1\end{array}\right)\quad\hat{Q}^{-1}=\left(\begin{array}[]{cc}1/2&1/2\\ 1/2&-1/2\end{array}\right)

Since the first row of P^=Q^1\hat{P}=\hat{Q}^{-1} is nonnegative the above result yields the invariance of the convex set S={u=(u1,u2):u1+u20}S=\{u=(u^{1},u^{2}):u^{1}+u^{2}\leq 0\}.
Note that wMP, that is the invariance of 2{\mathbb{R}}^{2}_{-}, does not hold true in this example. Indeed, the vector u=((x1x12)(x2x22)3,(x12+2x14)(x2x22))u=\big{(}(x_{1}-x_{1}^{2})(x_{2}-x_{2}^{2})^{3},(x_{1}^{2}+2x_{1}-4)(x_{2}-x_{2}^{2})\big{)} is a solution in the square Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1] taking non positive values on Ω\partial\Omega with u20u_{2}\leq 0, u10u_{1}\geq 0 in Ω\Omega.

2 Proofs of the results

The first part of this section is dedicated to the proofs of Proposition 1 and 2.

Proof of Proposition 1.

We restrict to the case ε-\varepsilon, with ε>0\varepsilon>0. In fact, we can reduce to it by using the change of coordinate xρxx\to\rho-x. We also observe that the argument is not affected by c~0\tilde{c}\geq 0, so that we will omit to mention it when discussing on the parameters.

Firstly, we observe that v(x)=xv(x)=-x obviously satisfies the second equation with v0v\leq 0 in [0,ρ][0,\rho].
Next, we introduce the sequence of functions

uk(x)=εc{1ec/k2sinh(c/k)ecx+ec/k12sinh(c/k)ecx1αε(sinh(cx)ksinh(c/k)x)}.u_{k}(x)=-\frac{\varepsilon}{c}\left\{\frac{1-e^{-\sqrt{c}/k}}{2\sinh(\sqrt{c}/k)}\,e^{\sqrt{c}x}+\frac{e^{\sqrt{c}/k}-1}{2\sinh(\sqrt{c}/k)}\,e^{-\sqrt{c}x}-1-\frac{\alpha}{\varepsilon}\left(\frac{\sinh(\sqrt{c}x)}{k\sinh(\sqrt{c}/k)}-x\right)\right\}. (2.1)

Then, a direct computation, shows that for all kk\in\mathbb{N},

uk′′εvcuk+αv=0inIρ(0,ρ)u_{k}^{\prime\prime}-\varepsilon v^{\prime}-cu_{k}+\alpha v=0\ \ \hbox{\rm in}\ I_{\rho}\equiv(0,\rho) (2.2)

and uk(0)=0u_{k}(0)=0.

The case c=0c=0 is ruled out either by taking the limit as k+k\to+\infty or directly by putting v=xv=-x and c=0c=0 in the above equation.

Note also that for kk\to\infty

uk(x)u0(x)=εc{coshcx1αε(sinhcxcx)}\begin{split}u_{k}(x)\to u_{0}(x)&=-\frac{\varepsilon}{c}\left\{\cosh\sqrt{c}x-1-\frac{\alpha}{\varepsilon}\left(\frac{\sinh\sqrt{c}x}{\sqrt{c}}-x\right)\right\}\end{split} (2.3)

Let the parameters ε\varepsilon, cc, α\alpha and ρ\rho be fixed. For large kk\in\mathbb{N}:

uk(0)=εc{1cosh(c/k)sinh(c/k)/cαε(c/ksinh(c/k)1)}=ε2k+o(1/k)\begin{split}u_{k}^{\prime}(0)&=-\frac{\varepsilon}{c}\left\{\frac{1-\cosh(\sqrt{c}/k)}{\sinh(\sqrt{c}/k)/\sqrt{c}}-\frac{\alpha}{\varepsilon}\left(\frac{\sqrt{c}/k}{\sinh(\sqrt{c}/k)}-1\right)\right\}\\ &=\frac{\varepsilon}{2k}+o(1/k)\end{split} (2.4)

so that uk(0)>0u^{\prime}_{k}(0)>0 for kk large enough.

Since uk(0)=0u_{k}(0)=0, we also have uk(x)>0u_{k}(x)>0 for some xIρx\in I_{\rho} for such kk\in\mathbb{N}. So wMP will be violated if uk(ρ)0u_{k}(\rho)\leq 0.

Next, computing (2.3) for x=ρx=\rho,

uk(ρ)=εcsinhcρcρ{coshcρ1αε(sinhcρcρ)}+uk(ρ)u0(ρ)=εcsinhcρcρ{ζ(cρ)cαε}+uk(ρ)u0(ρ)\begin{split}u_{k}(\rho)&=-\frac{\varepsilon\sqrt{c}}{\sinh\sqrt{c}\rho-\sqrt{c}\rho}\left\{\cosh\sqrt{c}\rho-1-\frac{\alpha}{\varepsilon}\left(\frac{\sinh\sqrt{c}\rho}{\sqrt{c}}-\rho\right)\right\}+u_{k}(\rho)-u_{0}(\rho)\\ &=-\frac{\varepsilon\sqrt{c}}{\sinh\sqrt{c}\rho-\sqrt{c}\rho}\left\{\zeta(\sqrt{c}\rho)\sqrt{c}-\frac{\alpha}{\varepsilon}\right\}+u_{k}(\rho)-u_{0}(\rho)\end{split} (2.5)

where

ζ(τ)=coshτ1sinhττ.\begin{split}\zeta(\tau)&=\frac{\cosh\tau-1}{\sinh\tau-\tau}.\end{split} (2.6)

Therefore condition ζ(cρ)c>αε\zeta(\sqrt{c}\rho)\sqrt{c}>\frac{\alpha}{\varepsilon}, see (1.8), yields uk(ρ)0u_{k}(\rho)\leq 0, for large kk\in\mathbb{N}, so that wMP is not satisfied. Once established this fact, we search for condition (1.8) to prove that wMP fails.
A straightforward calculation shows that ζ(τ)\zeta(\tau)\to\infty as τ0+\tau\to 0^{+} and ζ(τ)1\zeta(\tau)\to 1 as τ\tau\to\infty. Therefore there exists ρ0=ρ0(c;αε)\rho_{0}=\rho_{0}(c;\frac{\alpha}{\varepsilon}) such that condition (1.8) holds for ρ<ρ0\rho<\rho_{0}, and wMP is not satisfied, thereby proving that as soon as ε>0\varepsilon>0 there are intervals IρI_{\rho}, small enough, where wMP fails, whatever cc and α\alpha are.

On the other hand, let ρ>0\rho>0 be fixed. The function ζ(cρ)c\zeta(\sqrt{c}\rho)\sqrt{c} is increasing with respect to cc, and ζ(cρ)1\zeta(\sqrt{c}\rho)\to 1 as cc\to\infty, so that

limcζ(cρ)c=.\lim_{c\to\infty}\zeta(\sqrt{c}\rho)\sqrt{c}=\infty. (2.7)

Hence there exists c0=c0(ρ;αε)c_{0}=c_{0}(\rho;\frac{\alpha}{\varepsilon}) such that condition (1.8) holds for c>c0c>c_{0}, and wMP is not satisfied, thereby proving that as soon as ε>0\varepsilon>0 then wMP fails in any interval IρI_{\rho} and for any α0\alpha\geq 0 when a sufficiently large cc is taken.

By the increasing monotonicity of the function cζ(cρ)cc\to\zeta(\sqrt{c}\rho)\sqrt{c} we get

infc>0ζ(cρ)c=limc0+ζ(cρ)c=3ρ.\inf_{c>0}\zeta(\sqrt{c}\rho)\sqrt{c}=\lim_{c\to 0^{+}}\zeta(\sqrt{c}\rho)\sqrt{c}=\frac{3}{\rho}\,.

It follows that, if αε<3ρ\frac{\alpha}{\varepsilon}<\frac{3}{\rho}, then condition (1.8) is satisfied for all c>0c>0. This means that in this case we can choose c0(ρ;αε)=0c_{0}(\rho;\frac{\alpha}{\varepsilon})=0.

Finally, recalling that ζ(cρ)1\zeta(\sqrt{c}\rho)\to 1 as cc\to\infty, then ζ(cρ)cc\zeta(\sqrt{c}\rho)\sqrt{c}\cong\sqrt{c} for cc1(ρ)c\geq c_{1}(\rho). Hence condition (1.8) is equivalent to

c>αε.\sqrt{c}>\frac{\alpha}{\varepsilon}. (2.8)

It follows that, if αε>c1\frac{\alpha}{\varepsilon}>c_{1}, then we can choose c0(ρ;αε)=(αε)2c_{0}(\rho;\frac{\alpha}{\varepsilon})=\left(\frac{\alpha}{\varepsilon}\right)^{2}. ∎

Refer to caption
Figure 1: The function ζ(cρ)c\zeta(\sqrt{c}\rho)\sqrt{c}

A picture of the function cζ(cρ)cc\to\zeta(\sqrt{c}\rho)\sqrt{c} for different values of ρ>0\rho>0 is shown in Figure 1, where condition (1.8) with the threshold c0c_{0} is graphically exhibited on the track ρ=12\rho=\frac{1}{2} for different values of αε\frac{\alpha}{\varepsilon}.

Proof of Proposition 2.

Up to replacing u(x),v(x)u(x),v(x) with u(x),v(x)u(-x),v(-x), it is not restrictive to assume that ε>0\varepsilon>0. We claim that, for cc sufficiently large, there exists a pair (u,v)(u,v) satisfying (1.9) in a strict sense, namely

{inf(0,1)(u′′εvcu+αv)>0inf(0,1)(v′′ε~uc~v+βu)>0\left\{\begin{array}[]{lll}\inf_{(0,1)}\Big{(}u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v\Big{)}>0\\ \inf_{(0,1)}\Big{(}v^{\prime\prime}-\tilde{\varepsilon}u^{\prime}-\tilde{c}v+\beta u\Big{)}>0\end{array}\right.

such that u(0)=0>u(1)u(0)=0>u(1), v(0)0v(0)\leq 0, v(1)0v(1)\leq 0 and u(0)=0u^{\prime}(0)=0. Then, for δ>0\delta>0 sufficiently small, the pair of functions (u(x)+δx,v(x))(u(x)+\delta x,v(x)) still satisfies the system (1.9) and both functions are 0\leq 0 on the boundary of (0,1)(0,1), but u(x)>0u(x)>0 for x>0x>0 small, hence the wMP is violated.

Let us construct the pair of strict subsolutions (u,v)(u,v). They are defined as follows:

v(x)=x2x,u(x)=σχ(x),v(x)=x^{2}-x,\qquad u(x)=\sigma\chi(x),

where χ\chi is a smooth, non-increasing function satisfying

χ(0)=χ(0)=0,χ(x)=1for xmin{14,ε4|α|+1}\chi(0)=\chi^{\prime}(0)=0,\qquad\chi(x)=-1\ \ \text{for }\,x\geq\min\Big{\{}\,\frac{1}{4}\,,\,\frac{\varepsilon}{4|\alpha|+1}\,\Big{\}}

and σ\sigma is a positive constant that will be chosen later. We compute, for x(0,1)x\in(0,1),

v′′ε~uc~v+βu2σ(|β|+ε~χL((0,1))),v^{\prime\prime}-\tilde{\varepsilon}u^{\prime}-\tilde{c}v+\beta u\geq 2-\sigma\big{(}|\beta|+\tilde{\varepsilon}\|\chi^{\prime}\|_{L^{\infty}((0,1))}\big{)},

which is larger than 11 for σσ1:=1/(|β|+ε~χL((0,1))+1)\sigma\leq\sigma_{1}:=1/(|\beta|+\tilde{\varepsilon}\|\chi^{\prime}\|_{L^{\infty}((0,1))}+1). Next, for x(0,1)x\in(0,1), we have that

u′′εvcu+αv\displaystyle u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v ε(12x)|α|xσ|χ′′|cσχ.\displaystyle\geq\varepsilon(1-2x)-|\alpha|x-\sigma|\chi^{\prime\prime}|-c\sigma\chi.

We estimate the right-hand considering first 0<xmin{1/4,ε/(4|α|+1)}0<x\leq\min\{1/4,\varepsilon/(4|\alpha|+1)\}, where we have

u′′εvcu+αvε4σχ′′L((0,1)),u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v\geq\frac{\varepsilon}{4}-\sigma\|\chi^{\prime\prime}\|_{L^{\infty}((0,1))},

which is larger than ε/8\varepsilon/8 for σσ2:=ε/(8χ′′L((0,1)))\sigma\leq\sigma_{2}:=\varepsilon/(8\|\chi^{\prime\prime}\|_{L^{\infty}((0,1))}). While, for min(1/4,|ε|/(4|α|+1))<x<1\min(1/4,|\varepsilon|/(4|\alpha|+1))\!<\!x\!<\!1, we see that

u′′εvcu+αvε|α|+cσ,u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v\geq-\varepsilon-|\alpha|+c\sigma,

which is positive for c>(ε+|α|)/σc>(\varepsilon+|\alpha|)/\sigma. Summing up, taking σ=min{σ1,σ2}\sigma=\min\{\sigma_{1},\sigma_{2}\} and then c>(ε+|α|)/σc>(\varepsilon+|\alpha|)/\sigma, we have that (u,v)(u,v) satisfies (1.9) in a strict sense. The first statement of the proposition is thereby proved.

Let us turn to the second statement. We have seen above that wMP fails for (1.9) provided cc is larger than some c¯>0\bar{c}>0, and more precisely that it is violated by a pair (u,v)(u,v) with v<0v<0 on (0,1)(0,1) and u>0u>0 somewhere. Consider the pair (u,v)(u,v) associated with c=c¯c=\bar{c} and let II be a connected component of the set where u>0u>0 in (0,1)(0,1), hence u=0u=0 on I\partial I. For cc¯c\leq\bar{c} there holds in II,

u′′εvcu+αvu′′εvc¯u+αv0.u^{\prime\prime}-\varepsilon v^{\prime}-cu+\alpha v\geq u^{\prime\prime}-\varepsilon v^{\prime}-\bar{c}u+\alpha v\geq 0.

This means that the wMP fails in II and then concludes the proof. ∎

Let us go now to the proof of Theorem 3.

Proof of Theorem 3.

Assume that u[C2(Ω)]m[C0(Ω¯)]mu\in[C^{2}(\Omega)]^{m}\cap[C^{0}(\overline{\Omega})]^{m} satisfies (1.1) and that u0u\leq 0 on Ω\partial\Omega. Set

B^(i):=Q1B(i)QandC^:=Q1CQ.\hat{B}^{(i)}:=Q^{-1}B^{(i)}Q\ \ \ \text{and}\ \ \ \hat{C}:=Q^{-1}CQ.

Observe that the change of unknown u=Qvu=Qv gives, on the account of assumptions (1.13), (1.14), that vv satisfies

Av+i=1nB^(i)Div+C^v0inΩandv0onΩAv+\sum_{i=1}^{n}\hat{B}^{(i)}D_{i}v+\hat{C}v\geq 0\;\;\hbox{in}\;\;\Omega\;\;\hbox{and}\;\;v\leq 0\;\hbox{on}\;\;\partial\Omega (2.9)

that is, componentwise,

{Δv1+b1v1+C^1v0Δvm+bmvm+C^mv0\left\{\begin{array}[]{lll}\Delta v_{1}+b^{1}\cdot\nabla v_{1}+\hat{C}_{1}v\,\,\,\;\,\,\geq 0\\ \qquad\qquad\cdots\\ \Delta v_{m}+b^{m}\cdot\nabla v_{m}+\hat{C}_{m}v\geq 0\\ \end{array}\right. (2.10)

where bj=(βj(1),,βj(m))b^{j}=(\beta^{(1)}_{j},\dots,\beta^{(m)}_{j}) and C^j\hat{C}_{j} is the jj-th row of C^\hat{C}, for j=1,,mj=1,\dots,m.
We now employ the argument of the proof of Theorem 1 in [10] which reduces the above system to a scalar inequality governed by the uniformly elliptic (nonlinear) Bellman operator FF in (1.17). By viscosity calculus results based on the cooperativity condition (1.15), see [10, 6], since v=(v1,,vm)v=(v_{1},\dots,v_{m}) is a classical solution of (2.9) then the scalar function

v(x):=maxj=1,,m(vj)+(x),v^{*}(x):=\max_{j=1,\dots,m}(v_{j})^{+}(x)\,,

where “+” denotes the positive part, is a continuous weak solution in the viscosity sense, see [12], of

F[v]0inΩandv=0onΩ.F[v^{*}]\geq 0\;\;\hbox{in}\;\;\Omega\;\;\hbox{and}\;\;v^{*}=0\;\hbox{on}\;\;\partial\Omega\,. (2.11)

Suppose indeed that a smooth function φ\varphi touches from above vv^{*} at some point in Ω\Omega. If at that point v=0v^{*}=0 then clearly F[φ]0F[\varphi]\geq 0 there. Otherwise φ\varphi touches from above the component vjv_{j} realizing the positive maximum vv^{*} at that point and thus there holds

Δφ+bjφ+C^jv0.\Delta\varphi+b^{j}\cdot\nabla\varphi+\hat{C}_{j}v\geq 0.

But then recalling that C^j\hat{C}_{j} fulfills the cooperativity condition (1.5), one infers that

C^jvvjkC^jk0,\hat{C}_{j}v\leq v_{j}\sum_{k}\hat{C}_{jk}\leq 0,

whence again F[φ]0F[\varphi]\geq 0.

In order to apply the general result of [2] we need to show that the generalized principal eigenvalue, see [2], of FF is positive, which amounts to finding a strict supersolution which is strictly positive in Ω¯\overline{\Omega}. The latter is simply provided by ψ(x)=ψ(x1,,xm)=1δeγx1\psi(x)=\psi(x_{1},\dots,x_{m})=1-\delta e^{\gamma x_{1}}. Indeed, this function satisfies

F[ψ]=δγeγx1(γ+minj=1,,mβj(1)),F[\psi]=-\delta\gamma e^{\gamma x_{1}}\Big{(}\gamma+\min_{j=1,\dots,m}\beta_{j}^{(1)}\Big{)},

which is strictly negative in n{\mathbb{R}}^{n} provided γ>|minj=1,,mβj(1)|\gamma>|\min_{j=1,\dots,m}\beta_{j}^{(1)}|. We then choose δ\delta small enough, depending on γ\gamma and Ω\Omega, so that ψ>0\psi>0 in Ω¯\overline{\Omega}. Summing up, ψ\psi is positive in Ω¯\overline{\Omega} and satisfies there F[ψ]<0F[\psi]<0, hence also F[ψ]+λψ<0F[\psi]+\lambda\psi<0 for λ>0\lambda>0 suitably small. This implies that the numerical index μ1(F,Ω)\mu_{1}(F,\Omega) defined by

μ1(F,Ω)=sup{λ:ψC(Ω¯),ψ>0,F[ψ]+λψ0inΩ}\mu_{1}(F,\Omega)=\sup\{\lambda\in{\mathbb{R}}:\psi\in C(\overline{\Omega}),\ \psi>0,\ F[\psi]+\lambda\psi\leq 0\;\mbox{in}\;\Omega\} (2.12)

is strictly positive. Therefore, according to [2], the weak Maximum Principle for the scalar problem (2.11) holds, that is v0v^{*}\leq 0 in Ω\Omega.

This means that Q1u=v0Q^{-1}u=v\leq 0 in Ω\Omega and the proof is complete. ∎

We conclude the section with the proof of Theorem 4

Proof of Theorem 4.

Following the same lines of the proof of Theorem 3, we set u=Q^vu=\hat{Q}v. When multiplying by P^=Q^1\hat{P}=\hat{Q}^{-1}, this time we keep, by assumption, the positivity for the first kk equations, which again by the assupmtions made are decoupled in the gradient variables. So, letting C~\tilde{C} be the diagonal part of C^\hat{C} and v~=(v1,,vk)\tilde{v}=(v_{1},\dots,v_{k}), we get

{Δv1+b1vk+C~1v~0Δvk+bmvk+C~kv~0\left\{\begin{array}[]{lll}\Delta v_{1}+b^{1}\cdot\nabla v_{k}+\tilde{C}_{1}\tilde{v}\geq 0\\ \qquad\qquad\cdots\\ \Delta v_{k}+b^{m}\cdot\nabla v_{k}+\tilde{C}_{k}\tilde{v}\geq 0\\ \end{array}\right. (2.13)

where C~j\tilde{C}_{j} is the jj-th row of C^\hat{C}. The conclusion follows as in the proof of Theorem 2 with kk instead of mm. ∎

References

  • [1] H. Amann, Maximum principles and principal eigenvalues. (English summary) Ten mathematical essays on approximation in analysis and topology. 1–60, Elsevier B.V., Amsterdam, 2005. E.F. Beckenbach, R. Bellman, Inequalities.   Springer-Verlag (1961).
  • [2] H. Berestycki, I. Capuzzo Dolcetta, A. Porretta, L. Rossi, Maximum Principle and generalized principal eigenvalue for degenerate elliptic operators. J. Math. Pures Appl. (9) 103 (2015), 1276–1293.
  • [3] H. Berestycki, L. Nirenberg, S.N. Varadhan, The principal eigenvalue and maximum principle for second-order elliptic operators in general domains. Comm. Pure Appl. Math. 47 (1994), no. 1, 47–92.
  • [4] A. Berman, R.J. Plemmons, Nonnegative Matrices in the Mathematical Sciences. Classics in Applied Mathematics, SIAM, Philadelfia (1994).
  • [5] I. Birindelli, E. Mitidieri, G. Sweers, Existence of the principal eigenfunction for cooperative elliptic systems in a general domain. Differential Equations 35 (3) (1999).
  • [6] J. Busca, B. Sirakov, Harnack type estimates for nonlinear elliptic systems and applications, Ann. I. H. Poincaré 21 (2004) 543–590.
  • [7] L. A. Caffarelli, Y.Y. Li and L. Nirenberg, Some remarks on singular solutions of nonlinear elliptic equations III: viscosity solutions including parabolic operators. Comm. Pure Appl. Math. 66 (2013), no. 1, 109–143.
  • [8] I. Capuzzo Dolcetta, A. Vitolo, The weak maximum principle for degenerate elliptic operators in unbounded domains. Int. Math. Res. Not. IMRN 2018, no. 2, 412–431.
  • [9] I. Capuzzo Dolcetta, A. Vitolo, Directional ellipticity on special domains: weak maximum and Phragmén-Lindelöf principles. Nonlinear Anal. 184 (2019), 69–82.
  • [10] I. Capuzzo Dolcetta, A. Vitolo, Weak Maximum Principle for Cooperative Systems: the Degenerate Elliptic Case, Journal of Convex Analysis Volume 28 (2021), No. 2
  • [11] Z. Q. Chen, Z. Zhao, Potential theory for elliptic systems, The Annals of Probability 1996, Vol. 24, No. 1, 293-319
  • [12] M.G. Crandall, H. Ishii, P.L. Lions, User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc., New Ser. 27 (1992), pp. 1–67.
  • [13] G.I. Kresin and V.G. Maz’ya, On the maximum principle with respect to smooth norms for linear strongly coupled parabolic systems, Functional Differential Equations, 5:3-4 (1998), 349–376.
  • [14] G. Kresin and V. Maz’ya, Maximum Principles and Sharp Constants for Solutions of Elliptic and Parabolic Systems, Math. Surveys and Monographs, 183, Amer. Math. Soc., Providence, Rhode Island, 2012.
  • [15] G. Kresin, V. Maz’ya, Invariant convex bodies for strongly elliptic systems, J. Anal. Math. 135 (2018), no. 1, 203–224.
  • [16] Xu Liu, Xu Zhang, The weak maximum principle for a class of strongly coupled elliptic differential systems. Journal of Functional Analysis 263 (2012) 1862–1886.
  • [17] M. Miranda, Sul teorema del massimo modulo per una classe di sistemi ellittici di equazioni del secondo ordine e per le equazioni a coefficienti complessi. Istituto Lombardo (Rend.Sc.) A 104 736-745 (1970).
  • [18] M.H. Protter, H.F. Weinberger. Maximum Principles in Differential Equations. Springer-Verlag, New York, (1984).
  • [19] G. Sweers, Strong positivity in C(Ω¯)C(\overline{\Omega}) for elliptic systems. Math. Z. 209 (1992) 251–271.
  • [20] A. Vitolo, Singular elliptic equations with directional diffusion, Math.Eng. 3 (2021) no. 3 pp.1–16.
  • [21] H.F. Weinberger, Invariant sets for weakly coupled parabolic and elliptic systems. Rendiconti di Matematica 1 (1975), Vol. 8, Serie VI.