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

Comparison of two aspects of a PDE model for biological network formation

Clarissa Astuto King Abdullah University of Science and Technology (KAUST), 4700, Thuwal, Saudi Arabia Daniele Boffi King Abdullah University of Science and Technology (KAUST), 4700, Thuwal, Saudi Arabia Jan Haskovec King Abdullah University of Science and Technology (KAUST), 4700, Thuwal, Saudi Arabia Peter Markowich King Abdullah University of Science and Technology (KAUST), 4700, Thuwal, Saudi Arabia Giovanni Russo Department of Mathematics and Computer Science, University of Catania, Viale Andrea Doria 6, 95125, Catania, Italy
Abstract

We compare the solutions of two systems of partial differential equations (PDE), seen as two different interpretations of the same model that describes formation of complex biological networks. Both approaches take into account the time evolution of the medium flowing through the network, and we compute the solution of an elliptic-parabolic PDE system for the conductivity vector mm, the conductivity tensor \mathbb{C} and the pressure pp. We use finite differences schemes in a uniform Cartesian grid in the spatially two-dimensional setting to solve the two systems, where the parabolic equation is solved by a semi-implicit scheme in time. Since the conductivity vector and tensor appear also in the Poisson equation for the pressure pp, the elliptic equation depends implicitly on time. For this reason we compute the solution of three linear systems in the case of the conductivity vector m2m\in\mathbb{R}^{2}, and four linear systems in the case of the symmetric conductivity tensor 2×2\mathbb{C}\in\mathbb{R}^{2\times 2}, at each time step. To accelerate the simulations, we make use of the Alternating Direction Implicit (ADI) method.

The role of the parameters is important for obtaining detailed solutions. We provide numerous tests with various values of the parameters involved, to see the differences in the solutions of the two systems.

1 Introduction

We study two elliptic-parabolic systems of partial differential equations (PDE) describing the formation of biological network structures. Both systems are derived as gradient flows of an energy functional, consisting of a diffusive term, activation term and metabolic cost term. The energy functional can be seen as a continuum version of its discrete counterpart introduced in [1]. Here the authors consider a total energy consumption function for a general class of biological transport networks (seen also in [2]), including a material cost function of the network. In the papers [1, 3], the authors adapt the dynamics of local information for biological transport network and their relation with the optimization principle, which is known to be very common in nature.

Assuming the validity of Darcy’s law for slow flow in porous media (see, for instance,  [4, 5]), the energy functional is constrained by a Poisson equation for the fluid pressure. We use two modes of description of the network conductivity: first, in terms of a conductance vector mm, and, second, in terms of a symmetric positive definite conductance tensor \mathbb{C}. Taking the L2L^{2}-gradient flow with respect to mm (see, for instance, [6, 7, 8, 9, 10]) and, resp., with respect to \mathbb{C} (see [11]), leads to two structurally similar elliptic-parabolic PDE systems.

In the parabolic equation, a reaction term appears, representing the metabolic cost of the network, while the diffusion term describes the randomness in the material structure. The third term, called activation term, describes the tendency of the network to align with the principal direction of the material flow. The elliptic Poisson equation describes local mass conservation and is equipped with a right-hand side describing the distribution of sources and sinks of the material. This distribution is supplied as a datum and is supposed to be time independent.

The aim of this paper is to present several numerical simulations of both the vector-valued mm-model and the tensor-valued \mathbb{C}-model, with the goal of comparing their solutions in various parameter settings. The initial datum is chosen such that =mm\mathbb{C}=m\otimes m, i.e., initially, the principal direction (eigenvector corresponding to the largest eigenvalue) of \mathbb{C} is aligned with mm. Note that, in the spatially two-dimensional setting, the second eigenvalue of mmm\otimes m is zero, with eigenspace orthogonal to mm. We shall observe that the two PDE systems develop structurally and qualitatively similar solutions, however, they differ in quantitative details, for instance, the number and location of branches.

2 Mathematical model

We introduce the energy functional tens\mathcal{E}_{\mathrm{tens}} for the tensor-valued model,

tens[]:=ΩD22||2+c2p[][]p[]+M(||)dx,\mathcal{E}_{\mathrm{tens}}[\mathbb{C}]:=\int_{\Omega}\frac{D^{2}}{2}|\nabla\mathbb{C}|^{2}+c^{2}\nabla p[\mathbb{C}]\cdot\mathbb{P}[\mathbb{C}]\nabla p[\mathbb{C}]+M(|\mathbb{C}|)\mathrm{d}x, (1)

where =(x)2×2\mathbb{C}=\mathbb{C}(x)\in\mathbb{R}^{2\times 2} is the conductivity tensor, the diffusivity parameter DD\in\mathbb{R} measures the effect of random fluctuations in the network, and the activation parameter c2>0c^{2}>0 controls the strength of the network formation feedback loop. The total permeability tensor is of the form []:=r𝕀+\mathbb{P}[\mathbb{C}]:=r\mathbb{I}+\mathbb{C}, where the scalar function r=r(x)r0>0r=r(x)\geq r_{0}>0 describes the isotropic background permeability of the medium. The scalar pressure p=p[]p=p[\mathbb{C}] of the fluid transported within the network is the unique solution (up to an additive constant) of the Poisson equation

([]p)=S,-\nabla\cdot\left(\mathbb{P}[\mathbb{C}]\nabla p\right)=S, (2)

subject to homogeneous Neumann boundary condition on Ω\partial\Omega. The source/sink distribution S=S(x)S=S(x) in the mass conservation Eq. (2) is to be supplemented as an input datum and is assumed to be independent of time. The metabolic cost function M:++M:\mathbb{R}^{+}\to\mathbb{R}^{+} describes the dependence of the metabolic expenditure of maintaining the network on its transportation capacity, see [1]. The expression |||\mathbb{C}| denotes the Frobenius norm ||:=i=12j=12Cij2|\mathbb{C}|:=\sqrt{\sum_{i=1}^{2}\sum_{j=1}^{2}{C}_{ij}^{2}} and ||:=i=12j=12k=12(Cij/xk)2|\nabla\mathbb{C}|:=\sqrt{\sum_{i=1}^{2}\sum_{j=1}^{2}\sum_{k=1}^{2}(\partial{C}_{ij}/\partial x_{k})^{2}}.

Taking the L2L^{2}-gradient flow of the energy (1) constrained by (2), we obtain the parabolic-elliptic system

((r𝕀+)p)\displaystyle-\nabla\cdot\left(\left(r\mathbb{I}+\mathbb{C}\right)\nabla p\right) =S\displaystyle=S (3)
tD2Δc2pp+M(||)||\displaystyle\frac{\partial\mathbb{C}}{\partial t}-D^{2}\Delta\mathbb{C}-c^{2}\nabla p\otimes\nabla p+\frac{M^{\prime}\left(|\mathbb{C}|\right)}{\left|\mathbb{C}\right|}\mathbb{C} =0,\displaystyle=0, (4)

see [11] for details of the derivation. In our paper, we shall make the generic choice for the metabolic cost function (see, e.g., [9, 6, 7, 8]),

M(s):=αγsγfor s0,M(s):=\frac{\alpha}{\gamma}s^{\gamma}\qquad\mbox{for }s\geq 0, (5)

where α>0\alpha>0 is the metabolic constant and γ>0\gamma>0 the metabolic exponent. For instance, to model leaf venation in plants, one chooses 1/2<γ<11/2<\gamma<1, see [1]. Then Eq. (4) becomes

tD2Δc2pp+α||γ2=0\frac{\partial\mathbb{C}}{\partial t}-D^{2}\Delta\mathbb{C}-c^{2}\nabla p\otimes\nabla p+\alpha|\mathbb{C}|^{\gamma-2}\mathbb{C}=0 (6)

To derive the vector-valued model, we make the ansatz :=mm\mathbb{C}:=m\otimes m. Inserting this into (1) with D=0D=0 and noting that ||=|m|2|\mathbb{C}|=|m|^{2}, we obtain

=Ωc2p[m][m]p[m]+M(|m|2)dx\mathcal{E}=\int_{\Omega}c^{2}\nabla p[m]\cdot\mathbb{P}[m]\nabla p[m]+M\left(|m|^{2}\right)\mathrm{d}x

with [m]=r𝕀+mm\mathbb{P}[m]=r\mathbb{I}+m\otimes m and, with a slight abuse of notation, we now denote by p=p[m]p=p[m] the solution of the Poisson equation (2) with []\mathbb{P}[\mathbb{C}] replaced by [mm]\mathbb{P}[m\otimes m]. Re-introducing the Dirichlet integral D2Ω|m|2dxD^{2}\int_{\Omega}|\nabla m|^{2}\mathrm{d}x, we arrive at the energy functional

vect[m]:=ΩD2|m|2+c2p[m][mm]p[m]+M(|m|2)dx.\mathcal{E}_{\mathrm{vect}}[m]:=\int_{\Omega}D^{2}|\nabla m|^{2}+c^{2}\nabla p[m]\cdot\mathbb{P}[m\otimes m]\nabla p[m]+M(|m|^{2})\mathrm{d}x. (7)

Note that this functional is different from the one obtained by replacing C=mmC=m\otimes m in Eq. (1), the two functionals mainly differencing in the contribution of the diffusion term. As we shall see, we are mainly interested in studying the behavior of the system for very small diffusion coefficients, so that we expect such difference not being so crucial.

Taking the L2L^{2}-gradient flow with respect to the vector variable mm and using the explicit form (5) for the metabolic function, we obtain the system

((r𝕀+mm)p)\displaystyle-\nabla\cdot\left(\left(r\mathbb{I}+m\otimes m\right)\nabla p\right) =S\displaystyle=S (8)
mtD2Δmc2(pp)m+α|m|2(γ1)m\displaystyle\frac{\partial{m}}{\partial t}-D^{2}\Delta m-c^{2}(\nabla p\otimes\nabla p)m+{\alpha}|m|^{2(\gamma-1)}m =0.\displaystyle=0. (9)

From now on, we shall denote system (3)–(6) as the \mathbb{C}-system, while we call (8)–(9) the mm-system. Let us now shortly discuss the differences between the two systems.First, the activation term c2ppc^{2}\nabla p\otimes\nabla p in (4) is quadratic and depends on \mathbb{C} only through the solution p=p[]p=p[\mathbb{C}] of the Poisson equation (3). In contrast, the activation term c2(pp)mc^{2}(\nabla p\otimes\nabla p)m in (9) is cubic. Moreover, the metabolic term α||γ2\alpha|\mathbb{C}|^{\gamma-2}\mathbb{C} in (4) becomes singular at =0\mathbb{C}=0 if (and only if) γ<1\gamma<1. This obviously causes difficulties for both the analytical treatment of the system (see [11] for details) and its numerical resolution. We shall discuss in Section 3.2.1 how the numerical difficulties can be overcome. In contrast, the metabolic term α|m|2(γ1)m{\alpha}|m|^{2(\gamma-1)}m in (9) only becomes singular at m=0m=0 if γ<1/2\gamma<1/2. Taking into account that for typical applications in biology only the parameter range γ1/2\gamma\geq 1/2 is relevant (see [1]), the metabolic term in (9)) does not require any special treatment for the mm-system.

We define the equations on a domain Ω2\Omega\subset\mathbb{R}^{2}, and we define the boundary conditions on Ω\partial\Omega for the pressure and the conductivity. We choose homogeneous Dirichlet boundary conditions for mm and \mathbb{C}, and homogeneous Neumann conditions for pp.

m(t,x)=0,(t,x)=0,[]p(t,x)ν=0,xΩ,t0m(t,\vec{x})=0,\quad\mathbb{C}(t,\vec{x})=0,\quad\mathbb{P}[\mathbb{C}]\nabla p(t,\vec{x})\cdot\nu=0,\quad\vec{x}\in\partial\Omega,\,t\geq 0 (10)

where ν\nu is the outgoing normal vector to Ω\partial\Omega. In all our numerical test, the numerical support of \mathbb{C} or mm, after a long time, is well within Ω\Omega, therefore the solution should not be particularly sensitive to the boundary conditions on the two variables. In any case, different boundary conditions for mm and \mathbb{C} are currently under investigation.

To close the system, we prescribe an initial condition for the conductivity vector and tensor

m(t=0,x)=m0(x),(t=0,x)=0(x), in Ω.m(t=0,\vec{x})=m^{0}(\vec{x}),\quad\mathbb{C}(t=0,\vec{x})=\mathbb{C}^{0}(\vec{x}),\quad\text{ in }\Omega. (11)

A direct consequence of the boundary condition defined for the pressure, that is also a necessary condition for the solvability of the Poisson equation, is that the source function has to be vanishing mean, i.e., ΩS(x)𝑑Ω=0\displaystyle\int_{\Omega}S(\vec{x})d\Omega=0.

3 Numerical schemes

In this section we define the numerical schemes used to discretize in space and time the Eqs. (8 - 9) and (3-4). We adopt semi-implicit second order schemes.

3.1 Space discretization

In space we consider the square domain Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1], that we discretize by a uniform Cartesian mesh with spatial step h:=Δx=Δyh:=\Delta x=\Delta y. We call Ωh\Omega_{h} the discrete computational domain. The two variables conductivity and the pressure are ij(xi,yj),mijm(xi,yj){\mathbb{C}_{ij}\approx\mathbb{C}(x_{i},y_{j})},m_{ij}\approx m(x_{i},y_{j}) and pijp(xi,yj){p_{ij}\approx p(x_{i},y_{j})}, defined at the center of the cell (i,j)(i,j), therefore the set of grid points is xi=(i1/2)h,yj=(j1/2)h,(i,j){1,,N}2x_{i}=(i-1/2)h,\,y_{j}=(j-1/2)h,\,(i,j)\in\{1,\dots,N\}^{2}, hN=1hN=1.

In order to obtain second order accuracy in space, we use central differences for the computation of the space derivatives [12]. Discretizing in space Eq. (9), we have:

m(1)t\displaystyle\displaystyle\frac{\partial m^{(1)}}{\partial t} =\displaystyle= D2m(1)+c2(𝒟xp)2m(1)+c2𝒟xp𝒟ypm(2)α|m|2(γ1)m(1)\displaystyle D^{2}\mathcal{L}\,m^{(1)}+c^{2}\left(\mathcal{D}_{x}p\right)^{2}\,m^{(1)}+c^{2}\mathcal{D}_{x}p\mathcal{D}_{y}p\,m^{(2)}-\alpha|m|^{2(\gamma-1)}m^{(1)} (12)
m(2)t\displaystyle\displaystyle\frac{\partial m^{(2)}}{\partial t} =\displaystyle= D2m(2)+c2(𝒟yp)2m(2)+c2𝒟xp𝒟ypm(1)α|m|2(γ1)m(2)\displaystyle D^{2}\mathcal{L}\,m^{(2)}+c^{2}\left(\mathcal{D}_{y}p\right)^{2}\,m^{(2)}+c^{2}\mathcal{D}_{x}p\mathcal{D}_{y}p\,m^{(1)}-\alpha|m|^{2(\gamma-1)}m^{(2)} (13)

while the semi-discrete version of Eq. (4) is

C(1,1)t\displaystyle\displaystyle\frac{\partial{C}^{(1,1)}}{\partial t} =\displaystyle= D2C(1,1)+c2(𝒟xp)2γ||γ2C(1,1)\displaystyle D^{2}\mathcal{L}\,{C}^{(1,1)}+c^{2}\left(\mathcal{D}_{x}p\right)^{2}-\gamma|\mathbb{C}|^{\gamma-2}{C}^{(1,1)} (14)
C(1,2)t\displaystyle\displaystyle\frac{\partial{C}^{(1,2)}}{\partial t} =\displaystyle= D2C(1,2)+c2𝒟xp𝒟ypγ||γ2C(1,2)\displaystyle D^{2}\mathcal{L}\,{C}^{(1,2)}+c^{2}\mathcal{D}_{x}p\mathcal{D}_{y}p-\gamma|\mathbb{C}|^{\gamma-2}{C}^{(1,2)} (15)
C(2,2)t\displaystyle\displaystyle\frac{\partial{C}^{(2,2)}}{\partial t} =\displaystyle= D2C(2,2)+c2(𝒟yp)2γ||γ2C(2,2)\displaystyle D^{2}\mathcal{L}\,{C}^{(2,2)}+c^{2}\left(\mathcal{D}_{y}p\right)^{2}-\gamma|\mathbb{C}|^{\gamma-2}{C}^{(2,2)} (16)

where \mathcal{L} is the discrete Laplacian operator and 𝒟x\mathcal{D}_{x} and 𝒟y\mathcal{D}_{y} are, respectively, the discrete xx and yy first derivative operators, both using central difference approximation. The norm |||\cdot| we use is the Frobenius one. In this case we have three equations instead of four, because \mathbb{C} is symmetric.

In order to have a fully implicit scheme, we define a compact form of the semi-discrete Eqs. (12-13) and Eqs. (14-16), as follows:

mcompt\displaystyle\displaystyle\frac{\partial{m}_{\rm comp}}{\partial t} =\displaystyle= D2mcomp+c2𝒫mcompα𝒬m(m)mcomp\displaystyle D^{2}\mathcal{L}\,{m}_{\rm comp}+c^{2}\mathcal{P}\,{m}_{\rm comp}-\alpha\mathcal{Q}^{m}(m){m}_{\rm comp} (17)
compt\displaystyle\displaystyle\frac{\partial{\mathbb{C}}_{\rm comp}}{\partial t} =\displaystyle= D2comp+c2𝒫α𝒬c()comp\displaystyle D^{2}\mathcal{L}\,{\mathbb{C}}_{\rm comp}+c^{2}\mathcal{P}-\alpha\mathcal{Q}^{c}(\mathbb{C}){\mathbb{C}}_{\rm comp} (18)

where mcomp=[m(1),m(2)]T{m}_{\rm comp}=[m^{(1)},m^{(2)}]^{T}, comp=[C(1,1),C(1,2),C(2,2)]T{\mathbb{C}}_{\rm comp}=[C^{(1,1)},C^{(1,2)},C^{(2,2)}]^{T} and 𝒫=𝒟βp𝒟ηp\mathcal{P}=\mathcal{D}_{\beta}p\,\mathcal{D}_{\eta}p, with the suitable choice of β,η{x,y}\beta,\eta\in\{x,y\}. For the metabolic terms, we have

𝒬m(m)\displaystyle\mathcal{Q}^{m}(m) =\displaystyle= |m|2(γ1)\displaystyle|m|^{2(\gamma-1)} (19)
𝒬c()\displaystyle\mathcal{Q}^{c}(\mathbb{C}) =\displaystyle= ||γ2.\displaystyle|\mathbb{C}|^{\gamma-2}. (20)

Here, we discretize the Poisson equation for the pressure. To obtain a conservative scheme, we consider the following discretization: first, we extend the formula in Eq. (3), that becomes

x((r+C(1,1))xp)+x(C(1,2)yp)+y(C(1,2)xp)+y((r+C(2,2))yp)=S\displaystyle\partial_{x}\left(\left(r+C^{(1,1)}\right)\partial_{x}p\right)+\partial_{x}\left(C^{(1,2)}\partial_{y}p\right)+\partial_{y}\left(C^{(1,2)}\partial_{x}p\right)+\partial_{y}\left(\left(r+C^{(2,2)}\right)\partial_{y}p\right)=-S (21)

where we use the symmetry C(1,2)=C(2,1)C^{(1,2)}=C^{(2,1)}.

Now, we discretize the components of the formula, one by one, since we use different discretizations. For simplicity we pose 𝒞1,1=r+C(1,1)\mathcal{C}^{1,1}=r+C^{(1,1)}:

x(𝒞1,1xp)i,j\displaystyle\partial_{x}\left(\mathcal{C}^{1,1}\partial_{x}\,p\right)_{i,j}\approx 1Δx(𝒞i+1/2,j1,1xpi+1/2,j𝒞i1/2,j1,1xpi1/2,j)\displaystyle\frac{1}{\Delta x}\left(\mathcal{C}^{1,1}_{i+1/2,j}\partial_{x}\,p_{i+1/2,j}-\mathcal{C}^{1,1}_{i-1/2,j}\partial_{x}\,p_{i-1/2,j}\right)
=\displaystyle= 12Δx2((𝒞i+1,j1,1+𝒞i,j1,1)pi+1,j+(𝒞i1,j1,1+𝒞i,j1,1)pi1,j)\displaystyle\frac{1}{2\Delta x^{2}}\left(\left(\mathcal{C}^{1,1}_{i+1,j}+\mathcal{C}^{1,1}_{i,j}\right)p_{i+1,j}+\left(\mathcal{C}^{1,1}_{i-1,j}+\mathcal{C}^{1,1}_{i,j}\right)p_{i-1,j}\right)
12Δx2(𝒞i+1,j1,1+𝒞i1,j1,1+2𝒞i,j1,1)pi,j\displaystyle-\frac{1}{2\Delta x^{2}}\left(\mathcal{C}^{1,1}_{i+1,j}+\mathcal{C}^{1,1}_{i-1,j}+2\mathcal{C}^{1,1}_{i,j}\right)p_{i,j} (22)

where in the last line we consider the following approximations:

xpi+1/2,jpi+1,jpi,jΔx,𝒞i+1/2,j1,1𝒞i+1,j1,1+𝒞i,j1,12.\partial_{x}\,p_{i+1/2,j}\approx\frac{p_{i+1,j}-p_{i,j}}{\Delta x},\quad\mathcal{C}^{1,1}_{i+1/2,j}\approx\frac{\mathcal{C}^{1,1}_{i+1,j}+\mathcal{C}^{1,1}_{i,j}}{2}.

We omit the term with both yy-derivatives because it is analogue to the one with xx-derivatives. Now we discretize the term with mix derivatives.

x(C(1,2)yp)i,j\displaystyle\partial_{x}\left(C^{(1,2)}\partial_{y}\,p\right)_{i,j}\approx 1Δx(Ci+1/2,j(1,2)ypi+1/2,jCi1/2,j(1,2)ypi1/2,j)\displaystyle\frac{1}{\Delta x}\left(C^{(1,2)}_{i+1/2,j}\partial_{y}p_{i+1/2,j}-C^{(1,2)}_{i-1/2,j}\partial_{y}p_{i-1/2,j}\right) (23)
=\displaystyle= 18Δx2(Ci+1,j(1,2)+Ci,j(1,2))(pi+1,j+1pi+1,j1)\displaystyle\frac{1}{8\Delta x^{2}}\left(C^{(1,2)}_{i+1,j}+C^{(1,2)}_{i,j}\right)(p_{i+1,j+1}-p_{i+1,j-1})
18Δx2(Ci1,j(1,2)+Ci,j(1,2))(pi1,j+1pi1,j1)\displaystyle-\frac{1}{8\Delta x^{2}}\left(C^{(1,2)}_{i-1,j}+C^{(1,2)}_{i,j}\right)(p_{i-1,j+1}-p_{i-1,j-1}) (24)
+18Δx2(Ci+1,j(1,2)Ci1,j(1,2))(pi,j+1pi,j1)\displaystyle+\frac{1}{8\Delta x^{2}}\left(C^{(1,2)}_{i+1,j}-C^{(1,2)}_{i-1,j}\right)(p_{i,j+1}-p_{i,j-1}) (25)

again, we omit the term with y,xy,x-derivatives because it is analogue to the one with x,yx,y-derivatives.

3.2 Time discretization: symmetric-ADI method

At this stage, we describe the time discretization that we apply to the model. Since systems (8-6) and (3-4) are stiff in all their components, the choice of the time discretization is crucial for the efficiency.

We also need a high performing scheme in time, since at each time step we compute the solution of seven linear systems (two for the conductivity vector mm, three for the conductivity tensor \mathbb{C} and two Poisson equations for the pressure pp).

As we shall see, for some values of the parameters, the well-posedness of the problem becomes weaker, which reflects the bad conditioning of the numerical problem. For such a reason, we adopt a symmetric scheme, which better preserves possible symmetries of the solution. In particular, we adopt a symmetric-ADI scheme for both the conductivity variables, which guarantees efficiency, second order accuracy and spatial symmetry.

Anyway, the scheme is not strictly second order accurate in time for two reasons. First, the pressure is computed at time nn rather than at an intermediate time n+1/2n+1/2. Second, the metabolic term is treated partially explicitly and partially implicitly, thus destroying second order accuracy. Improvements of the order of accuracy in time are currently under investigation.

3.2.1 Time discretization for the conductivity vector

Here we focus on the time discretization for the Eqs. (8-9).

Given mnm(tn)m^{n}\approx m(t^{n}), we compute pnp^{n} by solving the Poisson equation

(mnmn)pn=S,-\mathcal{L}\left(m^{n}\otimes m^{n}\right)\,p^{n}=S, (26)

where (mnmn)N2×N2\mathcal{L}\left(m^{n}\otimes m^{n}\right)\in\mathbb{R}^{N^{2}\times N^{2}} is the discrete elliptic operator, in both directions (xx and yy), with variable coefficients and corresponding to zero Neumann conditions.

The symmetric-ADI method to solve the Eq. (17) works as follows. We start with the yy-direction implicit and xx-direction explicit, and then we consider the opposite order in the second step of the ADI scheme

(yimpl,xexpl)m~\displaystyle(y-{\rm impl},x-{\rm expl})\qquad\tilde{m} =\displaystyle= mn+Δt2ym~+Δt2xmn+Δtc2𝒫xyn(mn)\displaystyle m^{n}+\frac{\Delta t}{2}\mathcal{L}_{y}\tilde{m}+\frac{\Delta t}{2}\mathcal{L}_{x}m^{n}+\Delta t\,c^{2}\mathcal{P}_{xy}^{n}(m^{n})

and we solve for m~\tilde{m},

(ximpl,yexpl)myn+1\displaystyle(x-{\rm impl},y-{\rm expl})\qquad m^{n+1}_{y} =\displaystyle= m~+Δt2xmyn+1+Δt2ym~Δtα𝒬m(mn)myn+1\displaystyle\tilde{m}+\frac{\Delta t}{2}\mathcal{L}_{x}{m^{n+1}_{y}}+\frac{\Delta t}{2}\mathcal{L}_{y}\tilde{m}-\Delta t\,\alpha\mathcal{Q}^{m}\left({m^{n}}\right)m^{n+1}_{y}
+Δtc2𝒫ynmyn+1\displaystyle+\Delta t\,c^{2}\mathcal{P}_{y}^{n}\,m_{y}^{n+1}

and we solve for myn+1m^{n+1}_{y}.

The second time we apply the ADI method, we first consider the xx-direction implicit and yy explicit, and then we exchange the order. Thus we have

(ximpl,yexpl)m^\displaystyle(x-{\rm impl},y-{\rm expl})\qquad\hat{m} =\displaystyle= mn+Δt2xm^+Δt2ymn+Δtc2𝒫xyn(mn)\displaystyle m^{n}+\frac{\Delta t}{2}\mathcal{L}_{x}\hat{m}+\frac{\Delta t}{2}\mathcal{L}_{y}{m}^{n}+\Delta t\,c^{2}\mathcal{P}_{xy}^{n}(m^{n})

here we solve for m^\hat{m},

(yimpl,xexpl)mxn+1\displaystyle(y-{\rm impl},x-{\rm expl})\qquad m^{n+1}_{x} =\displaystyle= m^+Δt2ymxn+1+Δt2xm^Δtα𝒬m(mn)mxn+1\displaystyle\hat{m}+\frac{\Delta t}{2}\mathcal{L}_{y}{m^{n+1}_{x}}+\frac{\Delta t}{2}\mathcal{L}_{x}\hat{m}-\Delta t\,\alpha\mathcal{Q}^{m}\left({m^{n}}\right)m_{x}^{n+1}
+Δtc2𝒫xnmxn+1\displaystyle+\Delta t\,c^{2}\mathcal{P}_{x}^{n}\,m_{x}^{n+1}

and now we solve for mxn+1m^{n+1}_{x}, where 𝒬m(mn)=|mn|2(γ1)\mathcal{Q}^{m}\left(m^{n}\right)=|m^{n}|^{2(\gamma-1)} and β\displaystyle\mathcal{L}_{\beta}, with β=x,y\beta=x,y, are the discrete operators for the second derivatives in xx and yy directions, respectively, with βN×N\displaystyle\mathcal{L}_{\beta}\in\mathbb{R}^{N\times N}. For the pressure term we have 𝒫xn=(𝒟xpn)2\mathcal{P}_{x}^{n}=(\mathcal{D}_{x}p^{n})^{2} for the first component of the vector m(1)m^{(1)}, and 𝒫yn=(𝒟ypn)2\mathcal{P}_{y}^{n}=(\mathcal{D}_{y}p^{n})^{2} for the second component m(2)m^{(2)}. While the force term is 𝒫xyn(mn)=[𝒟xpn𝒟ypnm(2)n,𝒟xpn𝒟ypnm(1)n]T\mathcal{P}^{n}_{xy}(m^{n})=[\mathcal{D}_{x}p^{n}\mathcal{D}_{y}p^{n}\,m^{(2)n},\mathcal{D}_{x}p^{n}\mathcal{D}_{y}p^{n}\,m^{(1)n}]^{T}.

At the end, we calculate the average of the two solutions myn+1m^{n+1}_{y} and mxn+1m^{n+1}_{x} to obtain the conductivity vector at time tn+1t^{n+1}

mn+1\displaystyle m^{n+1} =\displaystyle= 12mxn+1+12myn+1.\displaystyle\frac{1}{2}{m^{n+1}_{x}}+\frac{1}{2}{m^{n+1}_{y}}. (27)

3.2.2 Time discretization for the conductivity tensor

As we said in the description of the two systems, for the conductivity tensor \mathbb{C}, the reaction term is very stiff because of the exponent γ\gamma that belongs to the interval (0.5,1)(0.5,1). After a finite time, we are basically dividing by zero at each time step. For this reason we introduce a small regularizing parameter ε\varepsilon in the equation, as follows

𝒬c()=|+ε|γ2\mathcal{Q}^{c}(\mathbb{C})=|\mathbb{C}+\varepsilon|^{\gamma-2}\mathbb{C} (28)

and we study the behaviour of the system as ε\varepsilon becomes smaller and smaller.

Given the pressure at time tnt^{n} from Eq. (26), we apply the symmetric-ADI method to solve the Eq. (18). As explained before, we first choose the yy-direction implicit and then we exchange the two directions. The scheme reads

(yimpl,xexpl)~1\displaystyle(y-{\rm impl},x-{\rm expl})\qquad\tilde{\mathbb{C}}_{1} =\displaystyle= n+Δt2y~1+Δt2xn+Δt𝒫n\displaystyle\mathbb{C}^{n}+\frac{\Delta t}{2}\mathcal{L}_{y}\tilde{\mathbb{C}}_{1}+\frac{\Delta t}{2}\mathcal{L}_{x}\mathbb{C}^{n}+\Delta t\,\mathcal{P}^{n}

and we solve for ~1\tilde{\mathbb{C}}_{1},

(ximpl,yexpl)yn+1\displaystyle(x-{\rm impl},y-{\rm expl})\qquad\mathbb{C}^{n+1}_{y} =\displaystyle= ~1+Δt2xyn+1+Δt2y~1Δtα𝒬c(n)yn+1\displaystyle\tilde{\mathbb{C}}_{1}+\frac{\Delta t}{2}\mathcal{L}_{x}{\mathbb{C}^{n+1}_{y}}+\frac{\Delta t}{2}\mathcal{L}_{y}\tilde{\mathbb{C}}_{1}-\Delta t\,\alpha\mathcal{Q}^{c}\left({\mathbb{C}^{n}}\right)\mathbb{C}^{n+1}_{y}

and we solve for yn+1\mathbb{C}^{n+1}_{y}.

Now, as before, we apply the ADI method for the second time for \mathbb{C} starting with the xx-direction implicit, and we have

(ximpl,yexpl)~2\displaystyle(x-{\rm impl},y-{\rm expl})\qquad\tilde{\mathbb{C}}_{2} =\displaystyle= n+Δt2x~2+Δt2yn+Δt𝒫n\displaystyle\mathbb{C}^{n}+\frac{\Delta t}{2}\mathcal{L}_{x}\tilde{\mathbb{C}}_{2}+\frac{\Delta t}{2}\mathcal{L}_{y}{\mathbb{C}}^{n}+\Delta t\,\mathcal{P}^{n}

here we solve for ~2\tilde{\mathbb{C}}_{2},

(yimpl,xexpl)xn+1\displaystyle(y-{\rm impl},x-{\rm expl})\qquad\mathbb{C}^{n+1}_{x} =\displaystyle= ~2+Δt2yxn+1+Δt2x~2Δtα𝒬c(n)xn+1\displaystyle\tilde{\mathbb{C}}_{2}+\frac{\Delta t}{2}\mathcal{L}_{y}{\mathbb{C}^{n+1}_{x}}+\frac{\Delta t}{2}\mathcal{L}_{x}\tilde{\mathbb{C}}_{2}-\Delta t\,\alpha\mathcal{Q}^{c}\left({\mathbb{C}^{n}}\right)\mathbb{C}_{x}^{n+1}

and here we solve for xn+1\mathbb{C}^{n+1}_{x}.

Finally we calculate n+1\mathbb{C}^{n+1} from the two solutions yn+1\mathbb{C}^{n+1}_{y} and xn+1\mathbb{C}^{n+1}_{x}

n+1\displaystyle\mathbb{C}^{n+1} =\displaystyle= 12xn+1+12yn+1.\displaystyle\frac{1}{2}{\mathbb{C}^{n+1}_{x}}+\frac{1}{2}{\mathbb{C}^{n+1}_{y}}.

The quantities above have the following expressions: 𝒬c(n)=|n+ε|γ2\mathcal{Q}^{c}\left(\mathbb{C}^{n}\right)=|\mathbb{C}^{n}+\varepsilon|^{\gamma-2}, and for the pressure, 𝒫n=𝒟βpn𝒟ηpn\mathcal{P}^{n}=\mathcal{D}_{\beta}p^{n}\,\mathcal{D}_{\eta}p^{n}, with the suitable choice of β,η{x,y}\beta,\eta\in\{x,y\} for the four components of the conductivity tensor.

4 Numerical results

In this section we perform several simulations with the aim of studying the effect of the various parameters. In particular, we check the agreement of the two models for the mm-system in Eqs. (8-9) and for the \mathbb{C}-system in Eqs. (3-6).

4.1 Accuracy tests and qualitative agreements

In Table 1 we define the tests we want to show in this paper, varying the parameters of the systems. This choice of parameters, a typical time scale is of the order of unit, while after time 15, the solution reached the steady state.

First, we check the accuracy of the schemes adopted. In Table 2 and Table 3 we see the error for the conductivity variables, calculated with Richardson extrapolation (see, e.g., [13]). We show the error for the module of the vector and of the tensor, and the parameters chosen are defined in TestA, TestB and TestC.

α\alpha cc DD ε\varepsilon γ\gamma rr tfint_{\rm fin}
Accuracy mm TestA: 0.5 1 0.01 - 0.75 0.1 1
Accuracy \mathbb{C} TestB: 1 1 0.01 0.1 1.75 0.1 1
Accuracy mm TestC: 0.5 5 0.01 - 0.75 0.01 1
D=0.05D=0.05 TestG: 0.75 5 0.05 10310^{-3} 0.75 0.005 15
D=0.01D=0.01 TestD: 0.75 5 0.01 10310^{-3} 0.75 0.005 15
D=0.001D=0.001 TestE: 0.75 5 0.001 10310^{-3} 0.75 0.005 15
γ=1\gamma=1 TestH: 0.75 5 0.01 10310^{-3} 1 0.005 15
γ=0.75\gamma=0.75 TestD: 0.75 5 0.01 10310^{-3} 0.75 0.005 15
γ=0.5\gamma=0.5 TestF: 0.75 5 0.01 10310^{-3} 0.5 0.005 15
ε=102\varepsilon=10^{-2} TestI: 0.75 5 0.01 10210^{-2} 0.75 0.005 15
ε=103\varepsilon=10^{-3} TestD: 0.75 5 0.01 10310^{-3} 0.75 0.005 15.
ε=104\varepsilon=10^{-4} TestL: 0.75 5 0.01 10410^{-4} 0.75 0.005 15
Table 1: In this table we define all the tests that we show in Section4.1. The first three rows show the parameters for the accuracy tests for mm and \mathbb{C}, and the results are summarized in Table 2-3. The second three rows define the parameters that we use in Fig. 2, where we compare the results changing the diffusivity. The third three rows are the tests showed in Fig. 3, varying gammagamma and the results of the last three rows are in Fig. 4, where we change the stabilization parameter ε\varepsilon. For the accuracy tests, the number of points of the discretization is specified in Table 2-3, while, for all the other tests, the number of points is fixed and it is N=600N=600.

Here we define the initial conditions mcomp0(x)m^{0}_{\rm comp}(\vec{x}) and comp0(x)\mathbb{C}^{0}_{\rm comp}(\vec{x}), and the source function S(x)S(\vec{x})

mcomp0(x)=[1,1]T,comp0(x)=[1,0,1]T,S(x)=EE¯\displaystyle m^{0}_{\rm comp}(\vec{x})=[1,1]^{T},\qquad\mathbb{C}^{0}_{\rm comp}(\vec{x})=[1,0,1]^{T},\qquad S(\vec{x})=E-\bar{E} (29)
E=exp(σ(xx0)2),σ=1000,x0=(0.1,0.1)\displaystyle E=\exp(-\sigma(\vec{x}-\vec{x}_{0})^{2}),\,\sigma=1000,\,\vec{x}_{0}=(0.1,0.1) (30)

where 𝕀\mathbb{I} is the identity matrix and E¯=mean(E)\bar{E}={\rm mean}(E).

N error order
20 - -
40 0.036030 -
80 0.0492860 -0.4520
160 0.01454106 1.7610
320 0.00690830 1.0737
640 0.001529779 2.1750
N error order
20 - -
40 0.036012 -
80 0.0493010 -0.4531
160 0.01456192 1.7594
320 0.00691103 1.0752
640 0.001528055 2.1772
Table 2: Accuracy test of the mm-system (8-9): we show the L2L^{2}-norm of the relative error for |m||m|, with the parameters defined in TestA (left) and TestC (right).
NN error2 order
25 - -
50 9.066×1029.066\times 10^{-2} -
100 4.625×1024.625\times 10^{-2} 0.97
200 1.571×1021.571\times 10^{-2} 1.56
400 4.149×1034.149\times 10^{-3} 1.92
800 7.347×1047.347\times 10^{-4} 2.50
Table 3: Accuracy test of the \mathbb{C}-system (3-4): we show the L2L^{2}-norm of the relative error for |||\mathbb{C}|, with the parameters defined in TestB.

In Fig. 1 we show three different quantities of the TestD: the module of the variables at final time (first column), the two components of the flux |p||\mathbb{C}\nabla p| at final time (second column) and the energy as a function of time (third column). In the first row we have the results for the variable mm and in the second row the results for the variable \mathbb{C}. As expected, the energy decays in time for both variables, and it is very small at final time, which indicates that we are close to the steady state of the systems. The main difference between the two variables are the shape of the network, with a YY-shape for the conductivity vector and a VV-shape for the tensor.

In Fig. 2 we show the results obtained when varying the parameter DD in Eqs. (8-9) and in Eqs. (3-6). The tests we consider are: TestG (first column), TestD (second column) and TestE (third column), with D{0.05,0.01,0.001}D\in\{0.05,0.01,0.001\}, for the variable mm in the first row and for \mathbb{C} in the second row. The ramifications become more evident when decreasing the diffusivity, and they get thinner and thinner. For the first parameter chosen, D=0.05D=0.05, we are not able to see those ramifications for the vector mm because the time-scale associated with the diffusion is too fast to capture the details.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1: In this figure we show three different quantities of the same computations, with the parameters defined in TestD: the module of the variables at final time (left panels), the flux also at final time (central panels) and the energy as a function of time (right panels). The first row is about the variable mm and the second one is for the variable \mathbb{C}.

In Fig. 3 we observe the dependence on the relaxation exponent γ\gamma. In the first column we report the results of TestH, in the second, those corresponding to TestD and in the third one, those corresponding to TestF. Again, in the first row we show the results for the variable mm and in the second row those for the variable \mathbb{C}. If γ=1\gamma=1 the results do not show the details of the network, and it seems that γ=0.75\gamma=0.75 is the parameter that better represents the leaf network.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: In this figure we show the difference of the results on varying the diffusivity DD. In the first column we have the results of the TestG (D = 0.05), in the second column we choose the parameters of the TestD (D = 0.01) and in the third one the TestE (D = 0.001). The first row shows the results of the variable mm and the second row those of the variable \mathbb{C}.

In Fig. 4 we show the behaviour of the solution \mathbb{C}, when ε0\varepsilon\to 0. We see the results for ε{102(left),103(center),104(right)}\varepsilon\in\{10^{-2}(\rm left),10^{-3}(\rm center),10^{-4}(\rm right)\}, and again we notice that for the largest value of ε\varepsilon we are not able to see any ramification. While we see that for ε\varepsilon smaller than 10310^{-3} we are close to the asymptotic behaviour.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: In this figure we show the difference of the results on varying the relaxation exponent γ\gamma. In the first column we have the results of the TestH (γ\gamma = 1), in the second column we choose the parameters of the TestD (γ=0.75\gamma=0.75) and in the third one the TestF (γ=0.5\gamma=0.5). The first row shows the results of the variable mm and the second row of the variable \mathbb{C}.
Refer to caption
Refer to caption
Refer to caption
Figure 4: In this figure we show the results for the variable \mathbb{C} varying the parameter ε\varepsilon. On the left we have the results of the TestI (ε=102\varepsilon=10^{-2}) in the central panel TestD (ε=103)(\varepsilon=10^{-3}) and on the right TestL (ε=104)(\varepsilon=10^{-4}).

4.2 Quantitative agreement

In this section we show some quantitative comparison between the two models. For this reason we consider well prepared initial data and we look for compatible parameters.

The goal of this part is to choose a convenient set of parameters in order to compare the two systems, trying to make them as close as possible. Now we distinguish the parameters (Dl,cl,αl,γl)(D_{l},c_{l},\alpha_{l},\gamma_{l}) with l=1l=1, for the \mathbb{C}-model, and l=2l=2, for the mm-model.

For simplicity, the choice of parameter is performed by comparing the two models in one space dimension. In 1D the systems (3,6) and (8-9) read

CtD12Cxxc12px2+α1|C|γ12C\displaystyle C_{t}-D_{1}^{2}\,C_{xx}-c_{1}^{2}\,p_{x}^{2}+\alpha_{1}|C|^{\gamma_{1}-2}C =0\displaystyle=0 (31)
mtD22mxxc22px2m+α2|m|2(γ21)m\displaystyle m_{t}-D_{2}^{2}\,m_{xx}-c_{2}^{2}\,p_{x}^{2}\,m+\alpha_{2}|m|^{2(\gamma_{2}-1)}m =0.\displaystyle=0. (32)

Now we suppose that CC has the following form

C=m2+B,C=m^{2}+B, (33)

where BB is a measure of the discrepancy between the two models, and we set the initial conditions so that B(t=0)=0B(t=0)=0. If we substitute Eq. (33) in Eq. (31), we have

(m2)tD12(m2)xxc12px2+α1|m2+B|γ2m2=Bt+D12Bxxα1|m2+B|γ2B.\displaystyle(m^{2})_{t}-D_{1}^{2}(m^{2})_{xx}-c_{1}^{2}p_{x}^{2}+\alpha_{1}|m^{2}+B|^{\gamma-2}m^{2}=-B_{t}+D_{1}^{2}B_{xx}-\alpha_{1}|m^{2}+B|^{\gamma-2}B. (34)

At this point we multiply Eq. (32) by a factor (2m)(2m), and we obtain

2mmt2D22mmxx2c22px2m2+2α2|m|2(γ21)m2=0.2m\,m_{t}-2D_{2}^{2}m\,m_{xx}-2\,c_{2}^{2}\,p_{x}^{2}\,m^{2}+2\alpha_{2}|m|^{2(\gamma_{2}-1)}m^{2}=0. (35)

After some manipulation, Eqs. (34,35) become:

2mmtD12(m2)xxc12px2+α1(m2)γ1=Bt+D12Bxxα1|m2|γ2B:=\displaystyle 2m\,m_{t}-D_{1}^{2}(m^{2})_{xx}-c_{1}^{2}p_{x}^{2}+\alpha_{1}\,(m^{2})^{\gamma-1}=\underbrace{-B_{t}+D_{1}^{2}B_{xx}-\alpha_{1}|m^{2}|^{\gamma-2}B}_{:=\mathcal{R}} (36)
2mmtD22(m2)xx+2D22(mx)22c22px2m2+2α2(m2)γ2=0\displaystyle 2m\,m_{t}-D_{2}^{2}(m^{2})_{xx}+2D_{2}^{2}(m_{x})^{2}-2\,c_{2}^{2}\,p_{x}^{2}\,m^{2}+2\alpha_{2}(m^{2})^{\gamma_{2}}=0 (37)

where \mathcal{R} is the residual. We made use of the following identity in the equation for mm

(m)xx2=2mmxx+2((mx)2).(m)^{2}_{xx}=2m\,m_{xx}+2\left((m_{x})^{2}\right). (38)

Now we consider the difference between Eq. (36) and Eq. (37), and we obtain

(D22D12)(m2)xx2D22(mx)2+(2c22m2c12)px2+α1(m2)γ12α2(m2)γ2=(D_{2}^{2}-D_{1}^{2})(m^{2})_{xx}-2D_{2}^{2}(m_{x})^{2}+(2\,c_{2}^{2}m^{2}-c_{1}^{2})p_{x}^{2}+\alpha_{1}\,(m^{2})^{\gamma-1}-2\alpha_{2}(m^{2})^{\gamma_{2}}=\mathcal{R} (39)

Since we want the residual to be small, in absolute value, a convenient choice for the sets of variables is the following

D1=D2,α1=2α2,γ1=γ2+1,c1=2c2|m|D_{1}=D_{2},\quad\alpha_{1}=2\alpha_{2},\quad\gamma_{1}=\gamma_{2}+1,\quad c_{1}=\sqrt{2}\,c_{2}|m| (40)

and for the initial conditions we choose m0m^{0} and C0C^{0} such that, initially, we have

C0=(m0)2=const.C^{0}=(m^{0})^{2}={\rm const}.

In this way the first derivative in space mxm_{x} is also equal to 0 after one time step.

In order to show some results in 2D, we need to define the initial conditions for mcomp0m^{0}_{\rm comp} and comp0\mathbb{C}^{0}_{\rm comp} such that, at initial time, we have again

0=m0m0,\mathbb{C}^{0}=m^{0}\otimes m^{0},

with mcomp0=[2/2,2/2]Tm^{0}_{\rm comp}=[\sqrt{2}/2,\sqrt{2}/2]^{T} and comp0=[0.5,0.5,0.5]T\mathbb{C}^{0}_{\rm comp}=[0.5,0.5,0.5]^{T}, while the values of the parameters are defined in TestM. In 2D, the equivalent expression of (33) is

=mm+𝔹.\mathbb{C}=m\otimes m+\mathbb{B}.

In this subsection we comment on the solutions of the following tests

α1,α2\alpha_{1},\alpha_{2} c1,c2c_{1},c_{2} D1=D2D_{1}=D_{2} ε\varepsilon γ1,γ2\gamma_{1},\gamma_{2} rr NN
set of parameters TestM: 1, 0.5 2\sqrt{2}, 1 0.1 10110^{-1} 1.75, 0.75 0.1 600
Table 4: In this table we define the two set of parameters in Eq. (40).
α\alpha cc DD ε\varepsilon γ\gamma rr tfint_{\rm fin} NN
D=0D=0 TestN: 0.75 5 0 10310^{-3} 0.75 0.005 15 600
D=105D=10^{-5} TestO: 0.75 5 10510^{-5} 10310^{-3} 0.75 0.005 15 600
Table 5: In this table we define the two tests showed in Figs.5-6, with D=0D=0 (TestN) and D=105D=10^{-5} (TestO).

In Table 6 we show the time evolution of the norm of the difference between \mathbb{C} and mmm\otimes m, to see how they move away from each other when we increase the time. In this table we see that the two solutions are very different, even after few time steps, and for this reason, they are difficult to be compared. The definition of 𝔹||\mathbb{B}|| is the following

𝔹:=|||mm||mm|||\mathbb{B}||:=\frac{\Big{|}\Big{|}|\mathbb{C}|-|m\otimes m|\Big{|}\Big{|}}{\Big{|}\Big{|}|m\otimes m|\Big{|}\Big{|}}

after ntn_{t} time steps, with time=ntΔt{\rm time}=n_{t}\Delta t, such that nt=6(2k),k=0,1,2,3,4,5,6n_{t}=6(2^{k}),\,k=0,1,2,3,4,5,6, where ||:=112+2122+222|\mathbb{C}|:=\sqrt{\mathbb{C}_{11}^{2}+2\,\mathbb{C}_{12}^{2}+\mathbb{C}_{22}^{2}} and
|mm|:=(m12)2+2(m1m2)2+(m22)2|m\otimes m|:=\sqrt{(m_{1}^{2})^{2}+2\,(m_{1}\,m_{2})^{2}+(m_{2}^{2})^{2}}.

time 0.01 0.02 0.04 0.08 0.16 0.32 0.64
𝔹||\mathbb{B}|| 0.0348 0.0538 0.0697 0.0982 0.1509 0.2611 0.5320
Table 6: Here we show the values of 𝔹\mathbb{B}, after ntn_{t} time steps, where nt=2k,k=0,1,2,3,4,5,6n_{t}=2^{k},k=0,1,2,3,4,5,6.

As it appears from the table, the two models move quickly far apart from each other, suggesting an intrinsically different behaviour.

Now we are interested in showing the solution of the \mathbb{C}-model, in the case of zero-diffusivity. Since the randomness of the network is common in nature but is also very effective in stabilizing the equations, we want to see if there is some analogy in considering the cases D=0D=0 and D1D\ll 1. In Fig. 5 we show the agreement of the two solutions, with D=0D=0 (left panel) and D=105D=10^{-5} (right panel). The other parameters are defined in TestN and TestO, while the initial condition is defined in Eq. (29) for Fig. 5.

In Fig. 6 we illustrate the long time solution for the \mathbb{C}-model obtained with the following space dependent initial condition:

0=f(x,y)𝕀,f(x,y)=(2|X+Y|)exp(10(|XY|))\mathbb{C}^{0}=f(x,y)\mathbb{I},\quad f(x,y)=(2-|X+Y|)\exp(-10(|X-Y|)) (41)
Refer to caption
Refer to caption
Figure 5: Comparison between TestN and TestO, with initial condition defined in Eq. (29). The main difference is that, on the left we have D=0D=0, while, on the right, D=105D=10^{-5}.
Refer to caption
Refer to caption
Figure 6: Comparison between TestN and TestO. Here the initial condition is defined in Eq. (41), with D=0D=0 on the left and D=105D=10^{-5} on the right.

We also note that, if we set the diffusion coefficient equal to zero, the model for mm reduces to a reaction equation for the conductivity. This means that if γ>1/2\gamma>1/2 the support of the unknown mm remains unchanged. In particular, it cannot extend, while in some regions the numerical support (i.e. the region in which |m||m| becomes lower than a given small threshold) may shrink.

Alternative boundary conditions

Furthermore, in the case of zero-diffusivity for the mm-model, we observe an anomalous behavior of the solution near the boundaries. In order to overcome such a problem we propose an ad hoc boundary condition as illustrated below.

Let us consider the equations for mm with D=0D=0 and in the limit of steady state. We have

c2ppmα|m|2(γ1)m\displaystyle c^{2}\nabla p\otimes\nabla p\,m-\alpha|m|^{2(\gamma-1)}m =0inΩ\displaystyle=0\quad{\rm in}\,\partial\Omega

that we can write as follows

(mp)p=αc2|m|2(γ1)minΩ\displaystyle(m\cdot\nabla p)\nabla p=\frac{\alpha}{c^{2}}|m|^{2(\gamma-1)}m\quad{\rm in}\,\partial\Omega (42)

Thus, we can deduce that mpm\propto\nabla p. This means that there exists a constant β\beta, such that,

m=βp.m=\beta\nabla p. (43)

If we substitute the Eq. (43) in Eq. (42), we obtain

β|p|2p=αc2β2(γ1)|p|2(γ1)βp\beta|\nabla p|^{2}\nabla p=\frac{\alpha}{c^{2}}\beta^{2(\gamma-1)}|\nabla p|^{2(\gamma-1)}\beta\nabla p (44)

and if we solve it for β\beta, we have the boundary condition for mm

m|Ω=(c2α|p|42γ)12(γ1)βp.\displaystyle\left.m\right|_{\partial\Omega}=\underbrace{\left(\frac{c^{2}}{\alpha}|\nabla p|^{4-2\gamma}\right)^{\frac{1}{2(\gamma-1)}}}_{\beta}\nabla p. (45)

Analogously, we can find also a boundary condition for the vector \mathbb{C}. Again, starting with zero-diffusivity and at steady state, we have

pp=αc2||γ2inΩ\nabla p\otimes\nabla p=\frac{\alpha}{c^{2}}|\mathbb{C}|^{\gamma-2}\mathbb{C}\quad{\rm in}\,\partial\Omega (46)

that means that the conductivity is proportional to the tensor product of the pressure gradient, i.e. pp\mathbb{C}\propto\nabla p\otimes\nabla p. Again, we look for a constant β~\tilde{\beta}, such that,

=β~pp.\mathbb{C}=\tilde{\beta}\nabla p\otimes\nabla p. (47)

Now we substitute the Eq. (47) in Eq. (46), and we solve for β~\tilde{\beta}. In this way, as before, we find the expression for the conductivity tensor at the boundary

|Ω=(c2α|p|2(γ2))1γ1β~pp.\left.\mathbb{C}\right|_{\partial\Omega}=\underbrace{\left(\frac{c^{2}}{\alpha}|\nabla p|^{-2(\gamma-2)}\right)^{\frac{1}{\gamma-1}}}_{\tilde{\beta}}\nabla p\otimes\nabla p. (48)

Conditions (45,48) might be a reasonable choice in the case of zero diffusivity. This treatment has the drawback of introducing additional non-linearity to the system. Alternative boundary conditions are currently under investigation.

Another aspect we want to focus on, is the steady state for the mm-model. In two different cases: γ<1\gamma<1 and γ>1\gamma>1 (as the authors show in [7]). For this reason we define different initial conditions for the vector mm, such that

m10,1=1,m20,1=2;m10,2=5,m20,2=5;\displaystyle m^{0,1}_{1}=1,\quad m^{0,1}_{2}=\sqrt{2};\quad m^{0,2}_{1}=5,\quad m^{0,2}_{2}=5;
m10,3=(2|X+Y|)exp(10|XY|),m20,3=m10,3;\displaystyle m^{0,3}_{1}=(2-|X+Y|)\exp{(-10|X-Y|)},\quad m^{0,3}_{2}=m^{0,3}_{1};

and in Fig. 7 we plot the following quantity

diff=mβmηmη{\rm diff}=\frac{||m^{\beta}-m^{\eta}||}{||m^{\eta}||} (49)

where ||||||\cdot|| is the Frobenius norm, and β,η{1,2,3}\beta,\eta\in\{1,2,3\}. In this way we see the difference between the solutions (with initial condition m0,1m^{0,1} and m0,2m^{0,2} in the left panel and with m0,1m^{0,1} and m0,3m^{0,3} in the right panel), as function of time. In this way, we support with numerical evidences that, for the mm-model and for γ>1\gamma>1, the steady state is unique and it does not depend on the initial conditions, as expected (see e.g. [7]).

In Fig. 8, that is the case for γ<1\gamma<1, we see that we reach two different steady states, when choosing two different initial conditions, suggesting that the steady state solution is not unique when γ<1\gamma<1.

Refer to caption
Refer to caption
Figure 7: In this figure we show the difference between two different solutions choosing, as initial condition, m0,1,m0,2m^{0,1},m^{0,2} (on the left) and m0,1,m0,3m^{0,1},m^{0,3} (on the right). We plot the expression defined in Eq. (49), as a function of time, with γ=1.75>1\gamma=1.75>1, and the others parameters are defined in TestD.
Refer to caption
Refer to caption
Figure 8: In this figure we show the steady states when γ=0.75<1\gamma=0.75<1. On the left the initial condition is m0=m0,1=1m^{0}=m^{0,1}=1 while, on the right, the initial condition is a function of space, m0=m0,3m^{0}=m^{0,3}, and the parameters are defined in TestD.

5 Conclusions

In this paper we use an elliptic-parabolic model to study the formation of biological network, and, in particular, of leaf venation networks. Throughout the paper we compare the solutions of two different systems, that derive from the Cai-Hu model, one for the conductivity vector and one for the conductivity tensor, and we explore the dependence of the solution on the parameters of the two models.

As we said before, all the components of the two systems are very stiff. In particularly, the \mathbb{C}-system is more challenging because of the negative exponent in the reaction term. For this reason, we add a regularization parameter ε\varepsilon, and we compute numerical solutions for smaller and smaller values of ε\varepsilon. This parameter prevents the instability coming from the division by zero in the reaction term.

We make use of finite differences scheme to compute the two solutions, with central differences for the space discretization and a symmetric-ADI method in time. The convergence rate is calculated numerically and denotes the second order accuracy.

At the end we added some quantitative comparison between the two systems, choosing more suitable sets of parameters. We see that the two solutions differ significantly, even after few time steps. This aspect makes it problematic any kind of direct comparison between the two systems.

Then, we showed some results in the case of zero diffusivity for the conductivity tensor. The last tests we consider are in agreement with the results achieved in [7], where the authors prove that there is a unique steady state for the mm-system when the metabolic exponent γ\gamma is greater than 1, and provide evidence that this is not true when γ<1\gamma<1.

Acknowledgments

G.R. thanks ITN-ETN Horizon 2020 Project ModCompShock, Modeling and Computation on Shocks and Interfaces, Project Reference 642768, and the Italian Ministry of Instruction, University and Research (MIUR) to support this research with funds coming from PRIN Project 2017 (No.2017KKJP4X entitled ”Innovative numerical methods for evolutionary partial differential equations and applications”.

References

  • [1] Dan Hu and David Cai. Adaptation and optimization of biological transport networks. Physical review letters, 111(13):138701, 2013.
  • [2] Eleni Katifori, Gergely J Szöllősi, and Marcelo O Magnasco. Damage and fluctuations induce loops in optimal transport networks. Physical review letters, 104(4):048704, 2010.
  • [3] Dan Hu. Optimization, adaptation, and initialization of biological transport networks. Notes from lecture, 1:3–1, 2013.
  • [4] Henry Darcy. Les fontaines publiques de la ville de Dijon: exposition et application… Victor Dalmont, 1856.
  • [5] Shlomo P Neuman. Theoretical derivation of darcy’s law. Acta mechanica, 25(3):153–170, 1977.
  • [6] Di Fang, Shi Jin, Peter Markowich, and Benoit Perthame. Implicit and Semi-implicit Numerical Schemes for the Gradient Flow of the Formation of Biological Transport Networks. The SMAI journal of computational mathematics, 5:229–249, 2019.
  • [7] Jan Haskovec, Peter Markowich, and Benoit Perthame. Mathematical analysis of a pde system for biological network formation. Communications in Partial Differential Equations, 40(5):918–956, 2015.
  • [8] Jan Haskovec, Peter Markowich, Benoît Perthame, and Matthias Schlottbom. Notes on a pde system for biological network formation. Nonlinear Analysis, 138:127–155, 2016. Nonlinear Partial Differential Equations, in honor of Juan Luis Vázquez for his 70th birthday.
  • [9] Giacomo Albi, Marco Artina, Massimo Foransier, and Peter A. Markowich. Biological transportation networks: Modeling and simulation. Analysis and Applications, 14(01):185–206, 2016.
  • [10] Jan Haskovec, Peter Markowich, and Simone Portaro. Emergence of biological transportation networks as a self-regulated process, 2022.
  • [11] Jan Haskovec, Peter Markowich, and Giulia Pilli. Tensor pde model of biological network formation. arXiv preprint arXiv:2111.03889, 2021.
  • [12] P. Wesseling. Principles of computational fluid dynamics. Springer, 2001.
  • [13] Lewis Fry Richardson. Ix. the approximate arithmetical solution by finite differences of physical problems involving differential equations, with an application to the stresses in a masonry dam. Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical or Physical Character, 210(459-470):307–357, 1911.