
We show that there are no stable stationary nonconstant solutions of the evolution problem (1) for fully autonomous reaction-diffusion-equations on the edges of a finite metric graph
$(1) \ \ \ \ \ \ \ \ \ \ {u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+f(uj)on the edges kj,(K) N∑j=1dijcij∂juj(vi,t)=0at the vertices vi.
Citation: Joachim von Below, José A. Lubary. Stability implies constancy for fully autonomous reaction-diffusion-equations on finite metric graphs[J]. Networks and Heterogeneous Media, 2018, 13(4): 691-717. doi: 10.3934/nhm.2018031
[1] | Joachim von Below, José A. Lubary . Stability implies constancy for fully autonomous reaction-diffusion-equations on finite metric graphs. Networks and Heterogeneous Media, 2018, 13(4): 691-717. doi: 10.3934/nhm.2018031 |
[2] | Matthias Erbar, Dominik Forkert, Jan Maas, Delio Mugnolo . Gradient flow formulation of diffusion equations in the Wasserstein space over a Metric graph. Networks and Heterogeneous Media, 2022, 17(5): 687-717. doi: 10.3934/nhm.2022023 |
[3] | Robert Carlson . Dirichlet to Neumann maps for infinite quantum graphs. Networks and Heterogeneous Media, 2012, 7(3): 483-501. doi: 10.3934/nhm.2012.7.483 |
[4] | Mirela Domijan, Markus Kirkilionis . Graph theory and qualitative analysis of reaction networks. Networks and Heterogeneous Media, 2008, 3(2): 295-322. doi: 10.3934/nhm.2008.3.295 |
[5] | Robert Carlson . Myopic models of population dynamics on infinite networks. Networks and Heterogeneous Media, 2014, 9(3): 477-499. doi: 10.3934/nhm.2014.9.477 |
[6] | Andrea Corli, Lorenzo di Ruvo, Luisa Malaguti, Massimiliano D. Rosini . Traveling waves for degenerate diffusive equations on networks. Networks and Heterogeneous Media, 2017, 12(3): 339-370. doi: 10.3934/nhm.2017015 |
[7] | Riccardo Bonetto, Hildeberto Jardón Kojakhmetov . Nonlinear diffusion on networks: Perturbations and consensus dynamics. Networks and Heterogeneous Media, 2024, 19(3): 1344-1380. doi: 10.3934/nhm.2024058 |
[8] | Dilip Sarkar, Shridhar Kumar, Pratibhamoy Das, Higinio Ramos . Higher-order convergence analysis for interior and boundary layers in a semi-linear reaction-diffusion system networked by a $ k $-star graph with non-smooth source terms. Networks and Heterogeneous Media, 2024, 19(3): 1085-1115. doi: 10.3934/nhm.2024048 |
[9] | Sergei Avdonin, Julian Edward . An inverse problem for quantum trees with observations at interior vertices. Networks and Heterogeneous Media, 2021, 16(2): 317-339. doi: 10.3934/nhm.2021008 |
[10] | Aslam Khan, Abdul Ghafoor, Emel Khan, Kamal Shah, Thabet Abdeljawad . Solving scalar reaction diffusion equations with cubic non-linearity having time-dependent coefficients by the wavelet method of lines. Networks and Heterogeneous Media, 2024, 19(2): 634-654. doi: 10.3934/nhm.2024028 |
We show that there are no stable stationary nonconstant solutions of the evolution problem (1) for fully autonomous reaction-diffusion-equations on the edges of a finite metric graph
$(1) \ \ \ \ \ \ \ \ \ \ {u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+f(uj)on the edges kj,(K) N∑j=1dijcij∂juj(vi,t)=0at the vertices vi.
In a fundamental paper of 1979 H. Matano [18] showed that autonomous reaction-diffusion equations involving the Laplacian under Neumann boundary conditions in a convex
$
{u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,
$
|
(1) |
where
$ \label{gkc} \sum\limits_{j = 1}^N d_{ij} c_{ij}\partial_ju_{j}(v_i, \cdot) = 0 \ \ \ \ \ \text{for}\ \ \ \ \ 1\leq i\leq n $ | (2) |
at each vertex
$ \label{1004} \forall v_i\in V_{\rm r}\, :\, \, k_j\cap k_s = \{v_i\}\;\Longrightarrow \;u_j(v_i) = u_s(v_i), $ | (3) |
that clearly is contained in the condition
$ f\in \mathcal{C}^1(\mathbb{R} ),\ \ \ \ c_{ij} > 0. $ | (4) |
Now we can state the
Main Theorem 1. On any finite connected metric graph all stable stationary solutions of Problem (1) are constant.
In 2001 E. Yanagida [21] published a list of five exceptional graphs that do not allow stable nonconstant stationary solutions in the consistent case of (1), see Section 5. Moreover, he established some fundamental instability tools, as the instability criterion in the presence of two different critical points in one edge. In 2015 the authors [9] showed that the assertion of the Main Theorem holds for any metric graph with sufficiently small edge lengths, as well as for any metric graph for the cubic balanced case
At first glance, the main result might seem to be surprising in view of the existence of stable nonconstant stationary solutions on non convex domains in higher dimensions, see [13,18] and the references therein. But, as it stands, with respect to the autonomous semilinear parabolic flows defined by (1), finite metric graphs behave like one-dimensional objects, i.e. like intervals of the real line. Clearly, Problem (1) can be regarded as an abstract interaction problem on a suitable interval in the sense of Gramsch and Ali Mehmeti, see e.g. [3,7], where the node transition conditions can be read as equivalent Cauchy conditions of order
The present paper is organized as follows. After some stability prerequisites and graph theoretical preliminaries in Section 2, Section 3 presents the exclusion of stable nonconstant stationary solutions on paths and circuits. The basic cutting technique that is a crucial tool in showing the main theorem, is established in Section 4 for more general reaction-diffusion-problems and states, for short, that if the metric graph
For any graph
$ n = \# V,\ \ \ \ \ \ \ N = \# E.$ |
The vertices will be numbered by
Moreover, we consider each graph as a topological graph in
$E(\Gamma ) = \{ \pi_j:[0, \ell_j]\to \mathbb{R} ^m| 1\leq j \leq N\}$ |
with the following properties: Each support
$G = \bigcup\limits_{j = 1}^N\pi_j\left([0, \ell_j]\right), $ |
especially each edge
$L = L(\Gamma) = \sum\limits_{j = 1}^N\ell_j.$ |
The orientation of the graph
$
\label{orifa} d_{ij} = {1if πj(ℓj)=vi,−1if πj(0)=vi,0otherwise.
$
|
For a function
$u_j(v_i): = u_j(\pi_j^{-1}(v_i)), \ \ \ \ \partial_j u_j(v_i): = \frac{\partial}{\partial x_j}u_j(x_j)\Bigr|_{\pi_j^{-1}(v_i)} \text{etc. and}$ |
$\int_G u\, dx: = \sum\limits_{j = 1}^N\int_{0}^{\ell_j}u_j(x_j)\, dx_j. $ |
Endowed with a usual product norm we set
$ L^p(G) = \prod\limits_{j = 1}^N L^p(0, \ell_j)\ \ \ \ \text{and}\ \ \ \ H^k(G) = \prod\limits_{j = 1}^N H^k(0, \ell_j) $ |
for
$\mathcal{C}_{K}^m(G) = \{u\in \mathcal{C}(G)\left|\, \forall j\in \left\{1, \ldots, N\right\}:\, u_j\in \mathcal{C}^{m}([0, \ell_j]), \, u\text{ satisfies (2)}\right. \}, $ |
and for functions depending also on a time variable that ranges in an interval
$\forall j\in \left\{1, \ldots, N\right\}:\, u_j\in \mathcal{C}^{m, r}([0, \ell_j]\times I).$ |
Closing this section we recall Lyapunov's notion of stability associated to the following reaction - diffusion - problem
$
{u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂j(aj(xj)∂juj)+fj(uj) on kj for 1≤j≤N,(K)N∑j=1dijcij∂juj(vi,⋅)=0 for 1≤i≤n,
$
|
(5) |
under the hypotheses
$ f_j\in \mathcal{C}^1(\mathbb{R} ), \ \ \ \ a_j\in \mathcal{C}^1 \left([0, \ell_j]\right), \ \ \ \ a_j > 0, \ \ \ \ c_{ij} > 0. $ | (6) |
A stationary solution
$ \forall t > 0 \ :\left\| u(\cdot, t)-w \right\|_{\infty, G} < \epsilon. $ |
Several instability criteria have been established in [9], among which we cite the following.
Lemma 2.1. If a stationary solution
$\int_G f'(u) dx > 0\ \ \ or\ \ \ \int_G f'(u)\, u^2 dx > \int_G f(u)\, u dx, $ |
then
Moreover, let us cite the following elementary result from [9].
Lemma 2.2. A stationary solution
$\int_G f(u) dx = 0\ \ \ \ and\ \ \ \ \int_G f(u)\, u^{2k+1}\, dx\geq 0 $ |
for all
On all finite metric graphs it can happen that there are no stationary solutions at all. E.g. for
In this section we establish the exclusion of stable nonconstant stationary solutions on paths and circuits. Let
$
{u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,(K) N∑j=1dij∂juj(vi,⋅)=0 for 1≤i≤n.
$
|
(7) |
Recall that the Rayleigh quotient of the linearized elliptic operator at
$ R(\varphi ;u) = \frac{\displaystyle\int_{G} \left(\partial \varphi \right)^2-f'(u)\varphi ^2 \ dx} {\displaystyle\int_G \varphi ^2 \ dx}, $ |
and that its admissible functions
$ \partial_j^2 \psi_{j}+f'(u_j) \psi_{j} = 0\ \ \ \ \text{in} \ \ \ \ [0, \ell_j] $ | (8) |
by standard regularity arguments. Now let
$
d_{ij} = {−1if i=j,1if i=j+1,0otherwise.
$
|
(9) |
Lemma 3.1. Suppose that
$\lambda_0(u) = \min \left\{R(\varphi ;u)| \varphi \in \mathcal{C}(\Pi)\cap H^1(\Pi)\right\} < 0.$ |
Proof. Here, in fact,
$0=N∑j=1∫ℓj0(∂2jψj+f′(uj)ψj)ψjdxj=−N∑j=1∫ℓj0(∂jψj)2−f′(uj)ψ2j dxj+N∑j=1[∂jψj ψj]ℓj0=−N(ψ), $
|
since (3) and
$ \sum\limits^N_{j = 1} \bigl[\partial_j \psi_{j}\ \psi_{j}\bigr]^{\ell_j}_0 = -\sum\limits^n_{i = 1} f(u(v_i)) \sum\limits^N_{j = 1} d_{ij} \psi_{j} (v_i) = 0. $ | (10) |
Thus,
$\psi (v_1) = \psi(v_{N+1}) = \partial_1 \psi _{1} (v_1) = \partial_N \psi(v_{N+1}) = 0. $ |
But as a solution of the second order linear ODE
Now we consider the general Kirchhoff law (2) on the path. As on all trees due to its homogeneous character, it can be written in the form
$ \label{gkctrees} \sum\limits_{j = 1}^N d_{ij}\ c_{j}\ \partial_ju_{j}(v_i) = 0 \ \ \ \ \ \text{for}\ \ \ \ \ 1\leq i\leq n $ | (11) |
with positive constants
$ R(\varphi ;u;c) = \frac{\displaystyle \sum\limits_{j = 1}^N c_{j}\int_{0}^{\ell_j} \left(\partial \varphi _j\right)^2-f'(u_j)\varphi _j^2 \ dx_j} {\displaystyle\sum\limits_{j = 1}^N c_{j}\int_{0}^{\ell_j} \varphi _j^2 \ dx_j} $ | (12) |
with
$ c_{1} = 1,\ \ \ c_{j+1} = \frac{c_{j+1, j+1}c_{j} }{c_{j+1, j}} \ \ \ \ \ \text{for}\ \ \ \ \ 2\leq j\leq N-1. $ | (13) |
The same technique applies on arbitrary trees and shows that for equilibria the linearized stability under (2) and (11) are equivalent. Set
$\lambda_0(u;c) = \min \left\{R(\varphi ;u;c)| 0\ne\varphi\in H^1(G)\cap \mathcal{C}(G)\right\}. $ |
As
$\lambda_0(u_0;c) = -f'(u_0) $ |
for all positive coefficients
Theorem 3.2. Let
$
\label{balprobpath}
{u∈C(Π×[0,∞))∩C2,1K(Π×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,(K): N∑j=1dijcij∂juj(vi,⋅)=0for 1≤i≤n,
$
|
(14) |
is unstable.
Proof. The assertion is true on an interval by Matano's result [18,21]. Thus, we can assume that
If
If
Thus, we are lead to the final case
$\partial_j u_j(x_j) = 0 \iff \pi_j(x_j)\in V_{\rm b} = \left\{v_1, v_{N+1}\right\}.$ |
Write
$\psi\in\mathcal{C}^1(\Pi) \ \ \ \ \text{with}\ \ \ \ \psi(v_1) = \psi_1(0) = \psi_N(\ell_N) = \psi(v_{N+1}) = 0.$ |
If
$ \partial_1\psi_1(0)\ne 0\ \ \ \ \text{and}\ \ \ \ \partial_N\psi_N(\ell_N)\ne0. $ | (15) |
Denoting the numerator of
$0=N∑j=11cj∫ℓj0(∂2jψj+f′(uj)ψj)ψjdxj=−N∑j=11cj∫ℓj0(∂jψj)2−f′(uj)ψ2j dxj+N∑j=1cj−1[∂jψj ψj]ℓj0=−N(ψ;u;(cj−1)), $
|
since (3) and
$\label{numc} \sum\limits^N_{j = 1} {c_j}^{-1}\bigl[\partial_j \psi_{j}\ \psi_{j}\bigr]^{\ell_j}_0 = -\sum\limits^n_{i = 1} f(u(v_i)) \sum\limits^N_{j = 1} d_{ij} \psi_{j} (v_i) = 0. $ |
Thus,
$R\left(\psi;u; \left({c_j}^{-1}\right)\right) = 0.$ |
We conclude that the edge Schrödinger operators
$ \sum\limits^N_{j = 1} d_{ij}{c_j}^{-1} \partial_{j} \varphi _j(v_i) = 0. $ | (16) |
lead to a Rayleigh quotient satisfying
$ \lambda_0\left(u; \left({c_j}^{-1}\right)\right) = \min\limits_{\varphi \in H^1(\Pi)\cap \mathcal{C}(\Pi)}R\left(\varphi ;u; \left({c_j}^{-1}\right)\right)\leq 0. $ | (17) |
Clearly,
$ R\left(\eta;u; \left({c_j}^{-1}\right)\right) = \lambda_0 < 0\ \ \ \ \text{and}\ \ \ \ \eta > 0 \text{ in } \Pi. $ | (18) |
This means that the zero solution is unstable for the edge operators
$\left\| \varphi\right\|_{c, L^\infty(\Pi)} = \sum\limits^N_{j = 1} {c_j} \left\| \varphi_j\right\|_{L^\infty[0, \ell_j]} $ |
are equivalent in
On a circuit
Corollary 1. Let
$
\label{balprobzeta}
{u∈C(ζ×[0,∞))∩C2,1K(ζ×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,(K): N∑j=1dijcij∂juj(vi,⋅)=0for 1≤i≤n.
$
|
(19) |
Then
Proof. Suppose that
$\partial_j u_j(x_j) = 0.$ |
Cut the circuit at
$\tilde{u}\in \mathcal{F}: = \left\{w\in \mathcal{C}(\tilde{\Pi})| \tilde{w}\left(\tilde{v}_1\right) = \tilde{w}\left(\tilde{v}_{N+1}\right) \right\}.$ |
Clearly,
$ \left\| \tilde{w}(\cdot, t_0)-\tilde{u}\right\|_{\infty, \tilde{\Pi}}\geq \widetilde{\varepsilon } _0 $ |
for some
For each
$ \forall t > 0 \ :\left\| z(\cdot, t)-u \right\|_{\infty, \zeta} < \epsilon.$ |
But for each
$ \forall t > 0 \ :\left\| \tilde{z}(\cdot, t)-\tilde{u}\right\|_{\infty, \tilde{\Pi}} < \varepsilon < \widetilde{\varepsilon } _0.$ |
In this section we establish some basic facts and results about stationary stable or unstable classical solutions of the more general reaction - diffusion - problem
$
(5)\ \ \ \ \ \ \ \ \ \ \ \ \ \ {u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂j(aj(xj)∂juj)+fj(uj) on kj for 1≤j≤N,(K) N∑j=1dijcij∂juj(vi,⋅)=0 for 1≤i≤n,
$
|
under the hypotheses (6). In particular we compare the stability on a metric graph
Theorem 4.1. [6,7] Suppose
$\exists b_1, b_2\geq 0\, \forall j\in\{1, ..., N\}\, \forall z\in\mathbb{R} :\, zf_j(z)\leq b_1z^2+b_2.$ |
Then, the following estimate holds:
$\left\| u \right\|_{\infty, G\times [0, T]}\leq\inf\limits_{\lambda > b_1} \left(e^{\lambda T}\max\left\{ \left\| u(\cdot, 0) \right\|_{\infty, G}, \sqrt{\frac{b_2}{\lambda-b_1}}\right\}\right).$ |
The basic reduction tool for establishing instability is the following surgery technique.
Theorem 4.2. Suppose that
Then
Proof. First, we consider the case where there is some constant
$ \label{2701} \sup\limits_j f'_j \leq b_1. $ | (20) |
The cutting of the edge
$F = \left\{\varphi \in\mathcal{C}_K^1(G)| \partial_j\varphi _{j} (p) = 0\right\}, $ |
$\tilde{C} = \left\{{\psi} \in \mathcal{C}(\tilde{G})| {\psi}(v_{n+1}) = {\psi}(v_{a}) \right\}, $ |
and for
$ \widetilde{\varphi } (x) = {φ(x)if x∈G∖{vn+1,va},φ(p)if x∈{vn+1,va}.
$
|
Then
$\iota = \left(\varphi \mapsto \widetilde{\varphi } \right) $ |
is an isometric isomorphism with respect to both
Now, suppose that
$0 < \delta\leq \frac{\varepsilon _0}{2}e^{-b_1t_0}.$ |
If
Next, suppose that
Now we claim
$(\ast)\ \ \ \ \ \ \ \ \ \ \left\| \tilde{z}(\cdot, t_0)- \tilde{w}(\cdot, t_0)\right\|_{\infty, \tilde{G}} = \left\| {z}(\cdot, t_0)- {w}(\cdot, t_0)\right\|_{\infty, G} < \frac14 \widetilde{\varepsilon } _0. $ |
Proof.
$\partial_t d_j = a_j\, \partial_j^2d_{j} +f_j(z_j) - f_j(w_j) = a_j\, \partial_j^2d_{j}+f'_j\left(\lambda z_j+ (1-\lambda)w_j\right)d_{j} $ |
with a function
$\left\| d(\cdot, t_0)\right\|_{\infty, \tilde{T}} \leq e^{b_1t_0} \max\limits_{\tilde{T}} |d(\cdot, 0)| = e^{b_1t_0}\left\| z_0-w_0\right\|_{\infty, G} < \frac14 \widetilde{\varepsilon } _0. $ |
This shows the claim
$ \left\| z(\cdot, t_0)-u\right\|_{\infty, G} = \left\| \tilde{z}(\cdot, t_0)-\tilde{u}\right\|_{\infty, \tilde{G}} \\\geq \left\| \tilde{u}-\tilde{w}(\cdot, t_0)\right\|_{\infty, \tilde{G}} - \left\| \tilde{w}(\cdot, t_0)-\tilde{z}(\cdot, t_0)\right\|_{\infty, \tilde{G}} \geq \widetilde{\varepsilon } _0 - \frac14\widetilde{\varepsilon } _0 > \varepsilon _0. $ |
Thus,
Finally, in order to get rid of (20), i.e. for arbitrary nonlinearities
$
\widetilde{f}_j(z) = {fj(z)for z∈[−2M,2M],fj(2M)+(z−2M)f′j(2M)for z≥2M,fj(−2M)+(z+2M)f′j(−2M)for z≤−2M.
$
|
(21) |
Thus,
As a first application of the surgery theorem 4.2 we generalize Yanagida's Two Points Lemma [9,21] known in the constant coefficients case under consistent Kirchhoff conditions to the more general problem (5).
Lemma 4.3. Suppose that
Proof. In the first case, we cut the edge
As already pointed out in [9], the hereditary properties of the stability notion with respect to subgraphs are very bad. However, the edge doubling is a simple extension technique that permits to conclude for stability from a graph containing the original one. Choose any edge
$\tilde{c}_{ms} = {12cisif m∈{i,h} and s∈{j,N+1},cmjotherwise,
$
|
and lead to the Kirchhoff conditions denoted by
Lemma 4.4. Suppose that
Proof. Each
$ \mathcal{C}_{K}^1(G)\hookrightarrow \mathcal{C}_{\tilde{K}}^1(\tilde{G}). $ |
Thus, identifying
Conversely, suppose that
$ \left\| \tilde{w}(\cdot, t_0)-\tilde{u}\right\|_{\infty, \tilde{G}}\geq \widetilde{\varepsilon } _0 $ |
for some
$ \forall t > 0 \ :\left\| z(\cdot, t)-u \right\|_{\infty, G} < \varepsilon.$ |
But for each
$ \forall t > 0 \ :\left\| \tilde{z}(\cdot, t)-\tilde{u}\right\|_{\infty, \tilde{G}} < \varepsilon < \widetilde{\varepsilon } _0.$ |
A non trivial application of the edge doubling is given by the following example.
Example 4.1. Suppose that
Suppose that
If
$\partial_1 u_1 \left(p\right) = \partial_2 u_2 \left(p\right) = 0\ \ \ \ \text{and}\ \ \ \ u_1 \left(p\right) = u_2 \left(p\right). $ |
By unique solvability of the corresponding Cauchy problems,
$ u_1 \left(v_1\right) = u_2 \left(v_1\right). $ |
Thus, if
$\ell_1 = \ell_2 = \frac{\ell}{2}$ |
and that
$ \textstyle \partial_1 u_1 \left(v_1\right) = \partial_2 u_2 \left(v_1\right). $ |
Choose
$ 2 c_{11}\partial_1 u_1 \left(v_1\right) -c_{13}\partial_3 u_3 \left(v_1\right) = 0. $ | (22) |
Cutting
Recall the following result by E. Yanagida from 2001.
Theorem 5.1. ([21]) If
$
\label{eqyan}
{u∈C2(G×[0,∞))∂tuj=∂2juj+f(uj) 1≤j≤N,N∑j=1dij dj ∂juj(vi,⋅)=0 for 1≤i≤n,
$
|
(23) |
has no stable stationary nonconstant solution on
An important ingredient of Yanagida's proof was the self-adjoint character of the associated eigenvalue problem of the linearized problem. Note that Example 4.1 restricted to Problem (23), but without restriction on the
$ \label{eqyangen}
{u∈C2,1(G×[0,∞))∂tuj=∂2juj+f(uj) 1≤j≤N,N∑j=1dij ∂juj(vi,⋅)=0 for 1≤i≤n,
$
|
(24) |
does not admit any stable stationary nonconstant solution, if the graph
(3) either
(4) or
(5) or
Proof. Throughout, let us suppose that
$ \text{Case}\ (3) $ |
By hypothesis,
Exactly as in Example 4.1 and using the notations given there, cut
$∂1u1(p)=∂2u2(p)=0, u1(p)=u2(p),u1(v1)=u2(v1), ∂1u1(v1)=∂2u2(v1), $
|
and
$ 2 \partial_1 u_1 \left(v_1\right) -\partial_3 u_3 \left(v_1\right) = 0. $ | (25) |
Cutting
$ \iota:\mathcal{C}_{K_c}^1(\Pi)\hookrightarrow \left\{w\in\mathcal{C}_K^1(\Sigma)| w_1 \left(p\right) = w_2 \left(p\right)\right\} \hookrightarrow \mathcal{C}_K^1(G), $ | (26) |
since each
$w = \bar{w}|_\Pi\ \ \ \ \text{and}\ \ \ \ \bar{w} = \iota(w). $ |
Secondly, for
$ \left\|\tilde{u}-w(\cdot, t)\right\|_{\infty, \Pi} = \left\|u-\bar{w}(\cdot, t)\right\|_{\infty, \Gamma} < \varepsilon.$ |
Thus, the restriction of
$ \text{ Case}\ (4) $ |
By Lemma 4.3, on each loop or circuit in
$ \text{ Case }\ (5) $ |
Again, dealing with
If
Figure 5 presents some graphs fulfilling the conditions of Theorem 5.2, that are not in Yanagida's list. Note that the embedding (26) yields compatibility of solutions for the extension-reduction-procedure related to appropriate subgraphs. However, in general graphs, such embeddings compatible with the different involved flows do not seem to be available. They would strongly simplify subgraph reduction techniques in showing instability. Therefore, the surgery techniques from Section 4 will be applied for the general case, rather than the ones above.
In this section we shall use the notation
$ \Delta_{ij} = d_{ij} \partial_ju_{j}(v_i) $ | (27) |
for the outer normal derivative of a function
$
\label{balprobtree}
{u∈C(T×[0,∞))∩C2,1K(T×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,(K): N∑j=1dijcij∂juj(vi,⋅)=0for 1≤i≤n,
$
|
(28) |
under Condition (4). Theorem 3.2 settles the case of a tree without essential ramification nodes and forms part of the following
Theorem 6.1. On any finite metric tree
Proof. Let
$ \label{77099} f' \leq b_1 $ | (29) |
with some constant
If
If
Thus, we conclude that
Choose any essential ramification node
$ \Delta_{i1}\Delta_{i2} = d_{i1}\partial_1 u_1(v_i)\, d_{i2}\partial_2 u_2(v_i) > 0. $ |
Remove
$\tilde{c}_{mj} = {ci2+ci1Δi1Δi2if m=i and j=2,cmjotherwise.
$
|
Then the restriction of
$ (\widetilde{K})\ \ \ \ \ \ \ \ \ \ \ \ \ \ \displaystyle \sum\limits_{j} d_{ij} \tilde{c}_{ij}\partial_ju_{j}(v_i) = 0 \ \ \ ( 1\leq i\leq n(\widetilde{T}) ) $ |
and constitutes a stationary nonconstant solution of (28) on
Next, we modify and extend
$ \left\| {w}_0-u\right\|_{\infty, \tilde{T}} < \delta, \ \ \ \ \left\| {w}_0-\tilde{w}_0\right\|_{\infty, \tilde{T}} < \frac{\varepsilon _0}{2}e^{-b_1t_0}. $ |
Then define
$ \left\| {w}_0-u\right\|_{\infty, {T}} = \left\| {w}_0-u\right\|_{\infty, \tilde{T}} < \delta .$ |
Moreover, for the solution
$ (\ast)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left\| w(\cdot, t_0)- \tilde{w}(\cdot, t_0)\right\|_{\infty, \tilde{T}} < \frac{\varepsilon _0}{2}. $ |
Proof.
$\partial_t z_j = \partial_j^2z_{j} +f_j(w_j) - f_j(\tilde{w}_j) = \partial_j^2z_{j}+f'_j\left(\lambda w_j+ (1-\lambda)\tilde{w}_j\right)z_{j} $ |
with a function
$\left\| z(\cdot, t_0)\right\|_{\infty, \tilde{T}} \leq e^{b_1t_0} \max\limits_{\tilde{T}} |z(\cdot, 0)| = e^{b_1t_0}\left\| {w}_0-\tilde{w}_0\right\|_{\infty, \tilde{T}} < \frac{\varepsilon _0}{2}. $ |
This shows the claim
By stability of
$ \left\| w(\cdot, t)-u\right\|_{\infty, T} < \frac{\varepsilon _0}{2} $ |
for all
$ε0≤‖˜w(⋅,t0)−u‖∞,˜T≤‖˜w(⋅,t0)−w(⋅,t0)‖∞,˜T+‖w(⋅,t0)−u‖∞,˜T≤‖˜w(⋅,t0)−w(⋅,t0)‖∞,˜T+‖w(⋅,t0)−u‖∞,T<ε0, $
|
which is absurd.
In order to achieve the proof of the general case we need a technical combinatorial lemma for graphs with circuits.
Lemma 6.2. Let
(a)
(b)
(c)
(d) Each circuit
$\Delta_{ij} \Delta_{is} > 0. $ |
Then there exists an edge
$\Delta_{ij} \Delta_{is} > 0\ \ \ and\ \ \ \Delta_{hj} \Delta_{hr} > 0. $ |
Proof. Note first that, by (c) and (d), each circuit
Let
$\Delta_{11} \Delta_{1 m} > 0. $ |
W.l.o.g. assume that
$\gamma(v_i;\Gamma)\geq 3. $ |
By (c),
$\Delta_{22} > 0,\ \ \ \ \Delta_{2j} > 0, \ \ \ \ \text{and}\ \ \ \ \Delta_{32} < 0. $ |
It follows recursively, that for
$\Delta_{ii} > 0,\ \ \ \ \Delta_{ij} > 0, \ \ \ \ \text{and}\ \ \ \ \Delta_{i+1, i} < 0 $ |
with some incident edge
$\Delta_{m-1, m-1} > 0, \ \ \ \ \Delta_{m-1, j} > 0, \ \ \ \ \Delta_{m-1, m} < 0 \ \ \ \ \text{and}\ \ \ \ \Delta_{m, m} < 0$ |
with some incident edge
Note that the assertion of Lemma 6.2 does not hold on trees as simple examples readily display. Now, we can show the exclusion result in the general case.
Theorem 6.3. There are no stable nonconstant stationary solutions of Problem (1) on any finite connected metric graph
Proof. W.l.o.g. by introducing artificial nodes with Kirchhoff conditions leading to continuous differentiability, we can assume that
$ {\bf{\Pi }}(\Gamma) = \big\langle c\in \ker\mathcal{D}(\Gamma)| \text{supp}(c)\text{ is a circuit in } \Gamma\big\rangle $ |
and satisfies
$ \label{2012} d = \text{corank}(\Gamma) = \dim\ker\mathcal{D}(\Gamma) = \dim{\bf{\Pi }}(\Gamma), $ | (30) |
that amounts to
$d\geq 1. $ |
Let
$ p\not\in V_{\rm b}. $ |
1. If
2. If
$ d(\tilde{\Gamma}_1) < d \ \ \ \ \text{and}\ \ \ \ d(\tilde{\Gamma}_2) < d.$ |
By recurrence, the restrictions of
3. If
Thus, it remains to show the assertion in the case
$\label{casepvb} \partial_j u_j(x_j) = 0 \, \iff\, \pi_j(x_j)\in V_{\rm b}. $ |
Use notation (27). At the endpoints
$\Delta_{ij} \Delta_{hj} < 0$ |
or
$ \Delta_{i1} \Delta_{ir} > 0\ \ \ \ \text{and}\ \ \ \ \Delta_{h1} \Delta_{hs} > 0, $ | (31) |
respectively. Then introduce the modified conductivities by
$\tilde{c}_{mj} = {cmj+cm1Δm1Δmjif (m,j)=(i,r) or (m,j)=(h,s),cmjotherwise.
$
|
Finally, omit the edge
The stationary case of the fully autonomous consistent parabolic problem
$
{u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+f(uj) on kj for 1≤j≤N,(K) N∑j=1dij∂juj(vi,⋅)=0 for 1≤i≤n.
$
|
(32) |
leads to the same first order system (33) defined by
$
{u′j=vj,v′j=−f(uj).
$
|
(33) |
On each edge (33) is Hamiltonian with respect to the function
$H(u, v) = \frac{v^2}{2}+\int_0^u f(\eta)d\eta.$ |
Thus, a stationary solution of (32) corresponds to
Proposition 1. Let
$ \left\|u\right\|_{H_0^1(G)}^2 = \int_G f(u)u\, dx. $ | (34) |
Moreover, if
$2\left(H_j-H_s\right) = \left(\partial_j u_{j}\right)^2(v_i) -\left(\partial_s u_{s}\right)^2 (v_i).$ |
Formula (34) applies e.g. in order to exclude stationary nonconstant solutions between
Lemma 7.1. Let
$ f(z) > 0 \ \ for\ \ z < z_{\min} \ \ \ \ and\ \ \ \ f(z) < 0\ \ for\ \ z > z_{\max}. $ | (35) |
Then
$z_{\min}\leq u \leq z_{\max}.$ |
If, in addition,
$z_{\min} < u < z_{\max}.$ |
In particular,
Proof. At a point
The lemma applies in particular to the case of the balanced cubic
Remark 7.1. Without (35) the assertion of the lemma is no longer true. Clearly, for
$ u(G)\cap \left[z_{\min}, z_{\max}\right]\ne \emptyset.$ |
For a Lyapunov-energy-calculus we introduce
$ \mathcal{E}(u) = \int_G \frac{v^2}{2} -F(u)\, dx\ \ \ \ \text{with}\ \ \ \ F(s) = \int_0^s f(\eta)\, d\eta, $ |
$ \mathcal{H}(u) = \sum\limits_{j = 1}^N \int_0^{\ell_j} H \left(u_j, \partial_j u_j\right)\, dx. $ |
Lemma 7.2. Let
(a) Along
$\dot{\mathcal{E}}(u): = \frac{d}{dt} \mathcal{E}(u) = - \int_G \left(\partial_t u\right)^2 dx. $ |
(b) If
$\mathcal{E}(u) = \int_G \frac12 f(u)u\, dx -\sum\limits_{j = 1}^N\ell_j H_j. $ |
In particular, if
Proof. As for (a), we can follow a standard density argument using the Kirchhoff and the continuity condition:
$˙E(u)=∫G∂xtu∂xu−f(u)∂tudx=∑j[∂tu∂xu]ℓj0⏟=0−∫G∂tu(∂2xu−f(u))⏟∂tudx=−∫G(∂tu)2dx. $
|
As for (b), both assertions follow readily with the definitions and from Proposition 1.
In order to apply Lasalle's principle, we have to impose an additional condition to
$ \left\{z\in\mathbb{R} | F(z)\geq 0\right\} \text{ bounded}, $ | (36) |
we obtain with
$\mathcal{E}(u) = \int_G \frac{v^2}{2} -F(u)\, dx \geq -\int_G F(u)\, dx \geq - ML. $ |
This enables the application of Lasalle's Principle [4] in order to conclude the following
Corollary 2. Under Condition (36) the solutions
We apply the preceding results to the attractivity properties of the equilibria for a nonlinearity
$
{f∈C1(R), −∞<A<B<C<∞, f(A)=f(B)=f(C)=0,f′(A)<0, f′(C)<0, f′(B)>0,f>0 in (−∞,A)∪(B,C), f<0 in (A,B)∪(C,∞).
$
|
(37) |
They clearly include the case of a cubic
$ M: = \max\limits_\mathbb{R} F^+ = \max \left\{F(A), F(C)\right\}, $ |
and Corollary 2 applies. Now we can state the following results about the flow defined by Problem (32) subject to Condition (37).
Theorem 7.3. (a) The equilibrium
$\displaystyle\mathcal{D}(A) = \left\{u_0\in\mathcal{C}(G)| \lim\limits_{t\to\infty} \int_G u(\cdot, t)dx = LA \right\} \supset \mathcal{C} \left(G;(-\infty, B]\right)\backslash\left\{B\right\}. $ |
(b) The equilibrium
$\displaystyle\mathcal{D}(C) = \left\{u_0\in\mathcal{C}(G)| \lim\limits_{t\to\infty} \int_G u(\cdot, t)dx = LC \right\}\supset \mathcal{C} \left(G;[B, \infty)\right)\backslash\left\{B\right\}. $ |
(c) Any nonconstant stationary solution
(d) A stationary solution
Proof. The claimed inclusions in (a) and (b) follow with [9,Theorem 4.1]. By continuity, Corollary 2 ensures that the solutions
$-LA < \int_G w dx < LC. $ |
Thus, for
The assertion (c) follows readily by Lemma 7.1 and (a) and (b). As for (d),
For the balanced cubic
$\lim\limits_{t\to\infty} \mathcal{E}(u(\cdot, t)) = 0, $ |
see [9,Prop. 6.9]. Theorem 7.3 applies in particular to the balanced cubic
The smallest example of the existence of a stable nonconstant stationary solution in presence of reaction terms depending on the edges is the following one.
Example 8.1. Let
$w_1(x) = \frac12 - \frac{x^2}{2},\ \ \ \ w_2(x) = \frac{x^2}{2} -x. $ |
Then
$ \left\|u(\cdot, t)-w\right\|_{\infty, G} \leq \text{const. } \left\|u(\cdot, 0)-w\right\|_{\infty, G}. $ |
Example 8.2. If we admit homogeneous Dirichlet boundary conditions at some vertices, stable nonconstant stationary solution occur already on an interval. On
$
{u∈C([0,ℓ]×[0,∞))∩C2,1([0,ℓ]×(0,∞))∂tu=u"+1 on [0,ℓ]u(0,⋅)=0=u′(ℓ,⋅)
$
|
(38) |
Then
$w(x) = \ell x - \frac{x^2}{2} $ |
defines a global attractor for the solutions of (38) in
$\lambda_1 = \frac{\pi^2}{4\ell^2}$ |
and leads via eigenfunction expansion and Dirichlet's Theorem to the conclusion
$ \left\|u(\cdot, t)-w\right\|_{\infty, G} \leq \text{const. }e^{-\lambda_1 t } \left\|u(\cdot, 0)-w\right\|_{\infty, G}. $ |
Example 8.3. If the nonlinearity depends on the edges and on
$
{u∈C(G×[0,∞))∩C2,1K(G×(0,∞)),∂tuj=∂2juj+fj(xj) on kj for 1≤j≤N,(K) N∑j=1dij∂juj(vi,t)=0 for 1≤i≤n.
$
|
(39) |
If
$\label{fx1001} 0 = \sum\limits_{j = 1}^N\int_0^{\ell_j}\partial_j^2w_{j}(x_j)\ dx_j = - \sum\limits_{j = 1}^N\int_0^{\ell_j}f_{j}(x_j)\ dx_j = \int_G f(x)\ dx. $ |
But if
$F_j(x_j) = \int_0^{x_j}f_{j}(\xi)\ d\xi $ |
Using
$w_j(x_j) = c_j +b_j x_j - \int_0^{x_j} F_j(s_j)\ d s_j $ |
is the unique stationary solution of (39). As in Example 8.1,
Example 8.4. If the diffusion coefficients are allowed to be different, then again stable nonconstant stationary solutions can occur. The example follows a refinement of Matano's type of counterexamples for non convex domains established by Cónsul and Solá-Morales [13]. Consider the path
$\ell_1 = \ell_3 = 1,\ \ \ \ell_2 = \delta > 0,\ \ \ a_1 = a_3 = 1,\ \ \ a_2 = \varepsilon > 0 $ |
with
$E(u) = \sum\limits_{j = 1}^{3} \int_{0}^{\ell_j} \frac{a_j}{2} \left(\partial_j u_j \right)^2 +G(u) dx_j, $ |
it can be shown that for
$ \left\{ u\in\mathcal{C}(\Pi)\cap H^1(\Pi)| \int_{0}^{\ell_1}u_1\ dx_1 \leq 0, \, \int_{0}^{\ell_3}u_3\ dx_3 \geq 0, \, -1\leq u \leq 1\right\} $ |
that is close to
Remark 8.1. The anti-Kirchhoff condition: If we replace the transition conditions (2) and (3) by their orthogonal condition, the famous anti-Kirchhoff condition
$ \sum\limits_{v_i\in k_j} u_{j}(v_i, t) = 0 \ \ \ \ \ \text{for}\ \ \ \ \ 1\leq i\leq n, $ | (40) |
$ k_j\cap k_s = \{v_i\}\ \Longrightarrow \ d_{ij}\partial_ju_{j}(v_i, t) = d_{is}\partial_s u_{s}(v_i, t) \, \, \ \ \text{for}\ \ 1\leq i\leq n, $ | (41) |
then the stability results change dramatically. E.g., there are no exceptional graphs, since for any finite metric graph, there is a suitable nonlinearity
Remark 8.2. As already pointed out above, node transition conditions different from the ones given by (2) and (3) can allow stable nonconstant stationary solutions. In particular, consider the limit problem of a parabolic problem on
$ \sum\limits_{j} d_{ij} c_{ij} \partial_j u_j (v_i, \cdot)+ \sigma \partial_t u(v_i, \cdot) = 0, $ |
see [11]. Letting
As the stationary solutions of the parabolic and the corresponding hyperbolic problem are the same when keeping the same node transition conditions, the interesting question arises, whether stability properties can be carried over from one case to the other. More generally, if the hyperbolic problem is approximated by parabolic ones with established stability criteria, e.g. by relaxation, is it possible to conclude stability criteria for the hyperbolic case from those in the parabolic case? A precise answer to this question is certainly of big interest and could apply e.g. to gas networks considered in [14], where the existence of nonconstant stationary solutions has been established for certain metric graphs.
Joachim von Below is grateful for several invitations to the UPC Barcelona between 2014 and 2016 within the research group GREDPA. José A. Lubary is grateful for the stays as invited professor to ULCO in Calais between 2014 and 2017 within the LMPA Joseph Liouville. The authors are indebted to Professor Xavier Cabré for pointing out Example 8.4 and to the anonymous referee for valuable remarks.
[1] | F. Ali Mehmeti, Lokale und globale Löungen linearer und nichtlinearer hyperbolischer Evolutionsgleichungen mit Transmission, Ph.D. thesis Johannes Gutenberg-Universität, Mainz, 1987. |
[2] |
Regular solutions of transmission and interaction problems for wave equations. Math. Meth. Appl. Sci. (1989) 11: 665-685. ![]() |
[3] |
Nonlinear interaction problems. Nonlinear Analysis, Theory, Methods & Applications (1993) 20: 27-61. ![]() |
[4] |
H. Amann,
Ordinary Differential Equations, de Gruyter, Berlin, 1990. doi: 10.1515/9783110853698
![]() |
[5] |
Classical solvability of linear parabolic equations on networks. J. Differential Equ. (1988) 72: 316-337. ![]() |
[6] | J. v. Below, A maximum principle for semilinear parabolic network equations, in Differential Equations with Applications in Biology, Physics, and Engineering (eds. J. A. Goldstein, F. Kappel, et W. Schappacher), Lect. Not. Pure and Appl. Math., 133 (1991), 37-45. |
[7] | J. v. Below, Parabolic Network Equations, 2nd ed. Tübingen Universitätsverlag 1994. |
[8] | Eigenvalue asymptotics for second order elliptic operators on networks. Asymptotic Analysis (2012) 77: 147-167. |
[9] |
Instability of stationary solutions of reaction-diffusion-equations on graphs. Results in Math. (2015) 68: 171-201. ![]() |
[10] | J. v. Below and J. A. Lubary, Stability properties of stationary solutions of reaction-diffusion-equations on metric graphs under the anti-Kirchhoff node condition, submitted. |
[11] |
J. v. Below and B. Vasseur, Instability of stationary solutions of evolution equations on graphs under dynamical node transition, Mathematical Technology of Networks, (ed. by Delio Mugnolo), Springer Proceedings in Mathematics & Statistics 128 (2015), 13-26. doi: 10.1007/978-3-319-16619-3_2
![]() |
[12] | N. L. Biggs, Algebraic Graph Theory, Cambridge Tracts Math. 67, Cambridge University Press, Cambridge UK, 1967. |
[13] |
Stability of local minima and stable nonconstant equilibria. J. Differential Equ (1999) 157: 61-81. ![]() |
[14] |
Stationary states in gas networks. Networks and Heterogeneous Media (2015) 10: 295-320. ![]() |
[15] |
Boundary feedback stabilization of the Schlöl system. Automatica (2015) 51: 192-199. ![]() |
[16] |
Multiplicity of solutions of second order linear differential equations on networks. Lin. Alg. Appl. (1998) 274: 301-315. ![]() |
[17] | J. A. Lubary, On the geometric and algebraic multiplicities for eigenvalue problems on graphs, in Partial Differential Equations on Multistructures (eds. F. Ali Mehmeti, J. v. Below and S. Nicaise) Lecture Notes in Pure and Applied Mathematics 219, Marcel Dekker Inc. New York, (2000), 135-146. |
[18] |
Asymptotic behavior and stability of solutions of semilinear diffusion equations. Publ. Res. Inst. Math. Sci. Kyoto Univ., (1979) 15: 401-451. ![]() |
[19] | S. Nicaise, Diffusion sur les espaces ramifié, Ph.D. thesis Université Mons, Belgium, 1986. |
[20] | R. J. Wilson, Introduction to Graph Theory, Oliver & Boyd Edinburgh UK, 1972. |
[21] |
Stability of nonconstant steady states in reaction-diffusion systems on graphs. Japan J. Indust. Appl. Math. (2001) 18: 25-42. ![]() |
1. | Joachim von Below, José A. Lubary, 2020, Chapter 1, 978-3-030-44096-1, 1, 10.1007/978-3-030-44097-8_1 | |
2. | Yuta Ishii, Multi-spike Patterns for the Gierer-Meinhardt Model with Heterogeneity on Y-shaped Metric Graph, 2022, 1040-7294, 10.1007/s10884-022-10157-y |